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

Regularity of Almost-Minimizers of Hölder-Coefficient
Surface Energies

David A. Simmons Department of Mathematics, University of Washington, Seattle, WA [email protected]
Abstract.

We study almost-minimizers of anisotropic surface energies defined by a Hölder continuous matrix of coefficients acting on the unit normal direction to the surface. In this generalization of the Plateau problem, we prove almost-minimizers are locally Hölder continuously differentiable at regular points and give dimension estimates for the size of the singular set. We work in the framework of sets of locally finite perimeter and our proof follows an excess-decay type argument.

The author was partially supported by NSF FRG 1853993.

1. Introduction

The Plateau problem is a classical geometric variational problem. It consists in minimizing surface area among all surfaces with a certain prescribed boundary. The analogous physical phenomenon occurs in soap films as they seek to minimize surface tension, an equivalent to minimizing surface area. The existence and regularity of solutions to the Plateau problem has been the subject of study in a variety of settings and continues to be a centerpiece of much mathematical research (to name a few, see [Dou31, Rad30, DG60, Rei60, All72, Tay76a, HP16, DLGM17, KMS19]). A natural generalization of the Plateau problem is to study minimizers of surface energies other than surface area. Anisotropic surface energies are those which depend on the normal direction to the surface and possibly the spatial location of the surface as well. This means that the energy assigned to a surface depends not only on its geometry but also on how and where the surface sits in space. Such anisotropic energies arise in physical phenomena such as the formation of crystals and in crystalline materials.

Almgren was the first to study regularity of minimizers to anisotropic variational problems in his paper [Alm68]. This initial work as well as much of the subsequent work in the area was done in the setting of varifolds and currents with many of the results applying to surfaces of arbitrary codimension but with rather strong regularity assumptions on the integrands of the anisotropic energies.

In this paper we work in the setting of sets of locally finite perimeter and study the existence and regularity of minimizers of anisotropic surface energies of the form

A(E;U)=UEA(x)νE(x),νE(x)1/2dn1(x)\displaystyle\mathscr{F}_{A}(E;U)=\int_{U\cap\partial^{*}\!E}\langle A(x)\nu_{E}(x),\nu_{E}(x)\rangle^{1/2}\operatorname{\,d\mathcal{H}^{n-1}}(x) (1.1)

where A=(aij(x))i,j=1nA=(a_{ij}(x))_{i,j=1}^{n} is a uniformly elliptic, Hölder continuous matrix-valued function, EE is a set of locally finite perimeter in n\mathbb{R}^{n}, and UU is an open set. Here E\partial^{*}\!E denotes the (n1)(n-1)-dimensional reduced boundary of EE and νE\nu_{E} denotes its outward unit normal vector. We note that Hölder continuity is a rather weak regularity assumption and the previously known regularity results for general integrands do not apply (see the discussion below). Our main regularity result applies to almost-minimizers which are sets of locally finite perimeter in n\mathbb{R}^{n} satisfying the minimality condition

A(E;B(x,r))A(F;B(x,r))+κrα+n1\displaystyle\mathscr{F}_{A}(E;\textbf{{B}}(x,r))\leq\mathscr{F}_{A}(F;\textbf{{B}}(x,r))+\kappa r^{\alpha+n-1} (1.2)

whenever EΔFUB(x,r)E\Delta F\subset\subset U\cap\textbf{{B}}(x,r), xUx\in U, and r<r0r<r_{0} (see Section 2 for full definitions and notation).

Theorem 1.1 (Regularity of almost-minimizers).

Let n2n\geq 2 and UU be an open set in n\mathbb{R}^{n}. Suppose A\mathscr{F}_{A} is the anisotropic energy given by (1.1) for a uniformly elliptic, Hölder continuous matrix-valued function A=(aij(x))i,j=1nA=(a_{ij}(x))_{i,j=1}^{n} with Hölder exponent α(0,1)\alpha\in(0,1). If EE is a (κ,α)(\kappa,\alpha)-almost-minimizer of A\mathscr{F}_{A} in UU, that is, it satisfies (1.2), then UEU\cap\partial^{*}\!E is a C1,α/4C^{1,\alpha/4}-hypersurface which is relatively open in UEU\cap\partial E, while the singular set of EE in UU,

Σ(E;U)=U(EE),\displaystyle\Sigma(E;U)=U\cap(\partial E\setminus\partial^{*}\!E), (1.3)

satisfies the following:

  1. (i)

    if 2n72\leq n\leq 7, then Σ(E;U)\Sigma(E;U) is empty;

  2. (ii)

    if n=8n=8, then Σ(E;U)\Sigma(E;U) has no accumulation points in UU;

  3. (iii)

    if n9n\geq 9, then s(Σ(E;U))=0\mathcal{H}^{s}(\Sigma(E;U))=0 for s>n8s>n-8.

A regularity result of the form of Theorem 1.1 was first proved by De Giorgi in [DG60] for minimizers of surface area. De Giorgi worked within the framework of sets of locally finite perimeter which he had introduced and shown to be equivalent to the earlier notion of Caccioppoli sets. Shortly thereafter Reifenberg also proved a similar regularity result for minimizers of surface area in [Rei60, Rei64a, Rei64b]. In [Tam82, Tam84], Tamanini extended De Giorgi’s result to almost-minimizers of perimeter satisfying the minimality condition P(E;B(x,r))P(F;B(x,r))+κrα+n1P(E;\textbf{{B}}(x,r))\leq P(F;\textbf{{B}}(x,r))+\kappa r^{\alpha+n-1}, proving C1,βC^{1,\beta}-regularity at points in the reduced boundary for each β(0,α/2)\beta\in(0,\alpha/2). In fact, his result applies with a more general error term.

The anisotropic surface energies treated by Almgren in [Alm68] are given in terms of the integral of a bounded, continuous, elliptic integrand f=f(x,ξ)f=f(x,\xi) over the surface. Here xx denotes the spatial variable and ξ\xi denotes the directional variable. Almgren proved that if ff is CkC^{k} for some k3k\geq 3, then minimal surfaces with respect to ff are Ck1C^{k-1}-regular almost everywhere. Bombieri extended this to the case k=2k=2 by showing in [Bom82] that if ff is C2C^{2}, then minimal surfaces with respect to ff are C1C^{1}-regular almost everywhere. In [SS82], Schoen and Simon provided an alternate proof of this type of regularity result with weakened hypotheses. They showed that if ff is Lipschitz in the spatial variable xx and C2,βC^{2,\beta} in the directional variable ξ\xi, then minimizers are C1,αC^{1,\alpha}-regular almost everywhere for any α(0,1)\alpha\in(0,1).

A characterization of the singular set for codimension one oriented hypersurfaces as in Theorem 1.1 was proved in the case of the area integrand f1f\equiv 1 in a series of papers by various authors. Miranda proved in [Mir65] that n1\operatorname{\mathcal{H}^{n-1}}-measure of the singular set is zero. The rest of the results deal with the Bernstein problem which asks about the existence of global minimizers of surface area in n\mathbb{R}^{n}. Fleming and Almgren proved some intermediate results of nonexistence in singular minimizing cones in 3\mathbb{R}^{3} and 4\mathbb{R}^{4}, respectively, in [Fle62, Alm66]. The next result was by De Giorgi in [DG65] where he showed that the non-existence of a singular minimal cone in n\mathbb{R}^{n} implies non-existence in n1\mathbb{R}^{n-1}. Simons showed the non-existence of singular minimal cones in dimensions 2n72\leq n\leq 7 in [Sim68] and Bombieri, De Giorgi, and Giusti demonstrated in [BDGG69] that Simons’ cone

Σ={x8:x12+x22+x32+x42=x52+x62+x72+x82}\displaystyle\Sigma=\big{\{}x\in\mathbb{R}^{8}\mathrel{\mathop{\mathchar 58\relax}}x_{1}^{2}+x_{2}^{2}+x_{3}^{2}+x_{4}^{2}=x_{5}^{2}+x_{6}^{2}+x_{7}^{2}+x_{8}^{2}\big{\}} (1.4)

is a singular minimal cone in 8\mathbb{R}^{8} with singular set {0}\{0\}. Federer concluded in [Fed70] by proving the Hausdorff dimension of the singular set is less than or equal to n8n-8. In the anisotropic case, it was shown in [SSA77] that n3\mathcal{H}^{n-3}-measure of the singular set is zero for elliptic integrands which are C3C^{3}.

Surface energies of the particular form of (1.1) first appeared in the paper [Tay76b] by Jean Taylor. This is a follow-up paper to her celebrated paper [Tay76a] in which she proves that the structure of singularities of soap-like minimal surfaces in 3\mathbb{R}^{3} are exactly as conjectured by the experimental physicist Joseph Plateau. In [Tay76b], she proves that minimizers of A\mathscr{F}_{A} in 3\mathbb{R}^{3} are locally C1,αC^{1,\alpha} at regular points and possess a singular set with the same general structure as in the case of surface area minimizers. Taylor worked with varifolds as her notion of surface and only with 22-dimensional surfaces in 3\mathbb{R}^{3}. This enabled her to utilize the classification of 22-dimensional surface area minimizing cones in 3\mathbb{R}^{3}. Such a classification is not known in higher dimensions. Note that the singularities dealt with by Jean Taylor cannot occur within our setting of sets of locally finite perimeter as 22-dimensional minimizing cones in 3\mathbb{R}^{3} come from non-oriented surfaces. This is why, for instance, we do not have singularities when 2n72\leq n\leq 7, even though there are singularities in lower dimensions when working with varifolds.

Allard’s work in [All72] established some important results for the Plateau problem in the setting of varifolds, some of which have been generalized to the anisotropic setting. Allard first proved that a varifold VV with bounded first-variation δV\delta V is rectifiable. He then proved regularity by showing that if there are LpL^{p}-type bounds on the generalized mean-curvature of VV for pp large enough (depending on the dimension of VV), then VV is locally C1,αC^{1,\alpha} for some α(0,1)\alpha\in(0,1) (depending on pp and the dimension of VV) outside a closed singular set of measure zero. A recent breakthrough was made in the setting of anisotropic integrands in [DPDRG18] to prove rectifiability. There, the authors were the first to successfully compute the first-variation δfV\delta_{f}V with respect to an anisotropic integrand ff. Using this they showed that if ff is an elliptic C1C^{1}-integrand satisfying the so-called atomic condition (equivalent to ellipticity in codimension one), then a varifold whose anisotropic first-variation δfV\delta_{f}V is locally bounded is indeed rectifiable. Further regularity is currently not known as the monotonicity formula which is essential in Allard’s regularity arguments does not exist for general integrands as demonstrated in [All74]. Much of the related relevant literature in contained in [DLDRG19, DPDRG16, DR18, DDG20, DRK20].

Another related problem of interest is volume constrained minimization. Regularity is known in the case of volume constrained perimeter minimizers [GMT83] and some results are known in anisotropic settings [Par84, LS20].

Let us briefly describe the organization of this paper. We start in Section 2 by providing the essential definitions pertaining to sets of locally finite perimeter, our anisotropic surface energies, and almost-minimizers. In Section 3 we follow the Direct Method of the Calculus of Variations to establish the existence of minimizers to our formulation of the anisotropic Plateau problem. The rest of the paper is devoted to the study of the regularity of almost-minimizers and to the characterization of the singular set. In Section 4 we cover a key change of variable that allows us to assume A(x0)=IA(x_{0})=I (the identity matrix) at a given point x0x_{0}, as well as prove many important properties of almost-minimizers. These include an almost-monotonicity formula, Theorem 4.7, volume and perimeter bounds, Proposition 4.10, and compactness of the class of almost-minimizers, Proposition 4.13. Next in Section 5 we define the excess, (5.4), an important notion in regularity theory, and recall some of its properties. There we also state the height bound, Proposition 5.9, which allows us to control the height of the boundary of an almost-minimizer given a small excess assumption. Following this we show in Section 6 that a small excess assumption together with the assumption A(x0)=IA(x_{0})=I allows us to find a Lipschitz function that well approximates E\partial E and is ‘almost-harmonic’ with a controlled error, Theorem 6.1. In Section 7 we prove a reverse Poincaré inequality, Theorem 7.1, which in Section 8 we combine with a harmonic approximation of the Lipschitz function from Section 6 to prove a tilt-excess decay result, Theorem 8.3. Finally, in Section 9 we use this and an iteration argument to prove our main regularity result, Theorem 9.2. We conclude the paper in Section 10 by using blow-up analysis and a Federer reduction argument to prove the characterization of singular set, Theorem 10.1.

2. Preliminaries

We will work in n\mathbb{R}^{n} for a fixed n2n\geq 2. The open ball centered at xnx\in\mathbb{R}^{n} of radius r>0r>0 is defined by

B(x,r)={yn:|yx|<r},\displaystyle\textbf{{B}}(x,r)=\{y\in\mathbb{R}^{n}\mathrel{\mathop{\mathchar 58\relax}}|y-x|<r\}, (2.1)

where |||\cdot| denotes the standard Euclidean norm and we write Br\textbf{{B}}_{r} for B(0,r)\textbf{{B}}(0,r). We denote the volume of the nn-dimensional ball by ωn\omega_{n}.

Like De Giorgi and Tamanini, we shall also work with sets of locally finite perimeter following much of the notation and definitions given in the insightful expository book by Maggi [Mag12]. Throughout this paper we will follow a scheme inspired by the one presented there.

A Lebesgue measurable set EnE\subset\mathbb{R}^{n} is said to be of locally finite perimeter if there exists an n\mathbb{R}^{n}-valued Radon measure μE\mu_{E} (called the Gauss-Green measure of EE) such that the Gauss-Green formula

Eφdx=nφ𝑑μE,φCc1(n)\displaystyle\int_{E}\nabla\varphi\>dx=\int_{\mathbb{R}^{n}}\varphi\>d\mu_{E},\qquad\forall\varphi\in C_{c}^{1}(\mathbb{R}^{n}) (2.2)

holds. The induced total-variation measure |μE||\mu_{E}| is called the perimeter measure of EE and is denoted by P(E;)P(E;\>\cdot\>). The set EE is said to be of finite perimeter if P(E)=P(E;n)<P(E)=P(E;\mathbb{R}^{n})<\infty. The set of those |μE||\mu_{E}|-a.e. xsptμEx\in\operatorname{spt}\mu_{E} for which

D|μE|μE(x)=limr0+μE(B(x,r))|μE|(B(x,r)) exists and is in SSn1\displaystyle D_{|\mu_{E}|}\mu_{E}(x)=\lim_{r\to 0^{+}}\frac{\mu_{E}(\textbf{{B}}(x,r))}{|\mu_{E}|(\textbf{{B}}(x,r))}\text{ exists and is in }\SS^{n-1} (2.3)

is called the reduced boundary of EE and is denoted by E\partial^{*}\!E. The measure-theoretic outer unit normal to EE is then defined to be the measurable function νE:ESSn1\nu_{E}\mathrel{\mathop{\mathchar 58\relax}}\partial^{*}\!E\to\SS^{n-1} given by

νE(x)=limr0+μE(B(x,r))|μE|(B(x,r)).\displaystyle\nu_{E}(x)=\lim_{r\to 0^{+}}\frac{\mu_{E}(\textbf{{B}}(x,r))}{|\mu_{E}|(\textbf{{B}}(x,r))}. (2.4)

The De Giorgi structure theorem states that E\partial^{*}\!E is (n1)(n-1)-rectifiable and that μE=νEn1  E\mu_{E}=\nu_{E}\operatorname{\mathcal{H}^{n-1}}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>\>\partial^{*}\!E where n1\operatorname{\mathcal{H}^{n-1}} denotes the (n1)(n-1)-dimensional Hausdorff measure. We may modify a set of locally finite perimeter on and/or up to a set of Lebesgue measure zero without changing its perimeter measure. As a consequence, the topological boundary E\partial E of a generic set of locally finite perimeter may be quite messy and might not be well related to E\partial^{*}\!E. However, we may always modify our set of locally finite perimeter EE so that sptμE=E\operatorname{spt}\mu_{E}=\partial E without changing its perimeter measure, in which case E¯=E\overline{\partial^{*}\!E}=\partial E (see [Mag12, Remark 16.11, Remark 15.3]). When discussing boundary regularity of a set of locally finite perimeter we shall always choose this representative of EE.

2.1. Anisotropic surface energies with Hölder coefficients

Now let’s provide precise definitions for the anisotropic energies and almost-minimizers we will study. Denote by nn\mathbb{R}^{n}\otimes\mathbb{R}^{n} the set of real n×nn\times n-matrices equipped with the operator norm ||||||\cdot||. Let A=(aij(x))i,j=1nA=(a_{ij}(x))_{i,j=1}^{n} be a bounded, measurable function on n\mathbb{R}^{n} that takes values in nn\mathbb{R}^{n}\otimes\mathbb{R}^{n}. We say that AA is symmetric if A(x)=A(x)tA(x)=A(x)^{t} for all xnx\in\mathbb{R}^{n}, where t{\>\cdot\>}^{t} denotes the matrix transpose. We say that AA is uniformly elliptic if there exist constants 0<λΛ<+0<\lambda\leq\Lambda<+\infty such that

λ|ξ|2A(x)ξ,ξΛ|ξ|2\displaystyle\lambda|\xi|^{2}\leq\langle A(x)\xi,\xi\rangle\leq\Lambda|\xi|^{2} (2.5)

for all x,ξnx,\xi\in\mathbb{R}^{n}, where ,\langle\>\cdot\>,\>\cdot\>\rangle denotes the standard Euclidean inner product. We say that AA is Hölder continuous with exponent α(0,1)\alpha\in(0,1) if

ACα=supxyA(x)A(y)|xy|α<\displaystyle||A||_{C^{\alpha}}=\sup_{x\not=y}\frac{||A(x)-A(y)||}{|x-y|^{\alpha}}<\infty (2.6)

and call ACα||A||_{C^{\alpha}} the Hölder seminorm of AA. In particular,

A(x)A(y)ACα|xy|α\displaystyle||A(x)-A(y)||\leq||A||_{C^{\alpha}}|x-y|^{\alpha} (2.7)

holds for all x,ynx,y\in\mathbb{R}^{n}.

Definition 2.1 (A\mathscr{F}_{A}-surface energy).

Let A=(aij(x))i,j=1nA=(a_{ij}(x))_{i,j=1}^{n} be uniformly elliptic, and Hölder continuous. Given a set of locally finite perimeter EE in n\mathbb{R}^{n} and a Borel set FF, we define the A\mathscr{F}_{A}-surface energy of EE in FF by

A(E;F)=FEA(x)νE(x),νE(x)1/2dn1(x)[0,].\displaystyle\mathscr{F}_{A}(E;F)=\int_{F\cap\partial^{*}\!E}\langle A(x)\nu_{E}(x),\nu_{E}(x)\rangle^{1/2}\operatorname{\,d\mathcal{H}^{n-1}}(x)\in[0,\infty]. (2.8)

Note that A(E;)\mathscr{F}_{A}(E;\>\cdot\>) defines a Borel measure on n\mathbb{R}^{n} and we will often denote A(E;n)\mathscr{F}_{A}(E;\mathbb{R}^{n}) by A(E)\mathscr{F}_{A}(E).

Remark 2.1 (Symmetry of AA).

We may assume without loss of generality that AA is symmetric which we do throughout this paper. We may make this assumption as the equality A(x)ξ,ξ=12(A(x)+A(x)t)ξ,ξ\langle A(x)\xi,\xi\rangle=\big{\langle}\frac{1}{2}\big{(}A(x)+A(x)^{t}\big{)}\xi,\xi\big{\rangle} holds for all x,ξnx,\xi\in\mathbb{R}^{n}. Hence we can always symmetrize AA without changing the values of A\mathscr{F}_{A}.

Remark 2.2 (Ellipticity).

The integrand f(x,ξ)=A(x)ξ,ξ1/2f(x,\xi)=\langle A(x)\xi,\xi\rangle^{1/2} is elliptic in the sense of Almgren in [Alm68]. In our setting this means that for every bounded set UU there is a constant c>0c>0 such that for every set of locally finite perimeter EE, half-space HH, and x0Ux_{0}\in U,

A(x0)(E;B(x0,r))A(x0)(H;B(x0,r))c[n1(EB(x0,r))n1(HB(x0,r))]\displaystyle\mathscr{F}_{A(x_{0})}(E;\textbf{{B}}(x_{0},r))-\mathscr{F}_{A(x_{0})}(H;\textbf{{B}}(x_{0},r))\geq c\>[\operatorname{\mathcal{H}^{n-1}}(\partial^{*}\!E\cap\textbf{{B}}(x_{0},r))-\operatorname{\mathcal{H}^{n-1}}(\partial H\cap\textbf{{B}}(x_{0},r))] (2.9)

whenever EΔHUB(x0,r)E\Delta H\subset\subset U\cap\textbf{{B}}(x_{0},r), r>0r>0. Here A(x0)(E;)\mathscr{F}_{A(x_{0})}(E;\>\cdot\>) denotes the energy associated to the frozen integrand fx0(ξ)=A(x0)ξ,ξ1/2f_{x_{0}}(\xi)=\langle A(x_{0})\xi,\xi\rangle^{1/2}. As Almgren notes, this notion is equivalent to uniform convexity in codimension one as is our case by uniform ellipticity of AA and (2.9) holds with c=λc=\lambda. Ellipticity ensures that half-spaces are the unique minimizers when compared with their compactly contained variations.

Remark 2.3 (Hölder continuity of integrand of A\mathscr{F}_{A}).

The integrand f(x,ξ)=A(x)ξ,ξ1/2f(x,\xi)=\langle A(x)\xi,\xi\rangle^{1/2} is Hölder continuous with respect to the spatial variable xx, that is,

|A(x)ξ,ξ1/2A(y)ξ,ξ1/2|12λACα|xy|α.\displaystyle\big{|}\langle A(x)\xi,\xi\rangle^{1/2}-\langle A(y)\xi,\xi\rangle^{1/2}\big{|}\leq\frac{1}{2\lambda}\>||A||_{C^{\alpha}}|x-y|^{\alpha}. (2.10)

for all x,y,ξnx,y,\xi\in\mathbb{R}^{n} with |ξ|=1|\xi|=1. This follows from (2.7) combined with the useful inequality

|A(x)ξ,ξ1/2A(y)ξ,ξ1/2|=|A(x)ξ,ξA(y)ξ,ξ|A(x)ξ,ξ1/2+A(y)ξ,ξ1/212λA(x)A(y)\displaystyle\big{|}\langle A(x)\xi,\xi\rangle^{1/2}-\langle A(y)\xi,\xi\rangle^{1/2}\big{|}=\frac{\big{|}\langle A(x)\xi,\xi\rangle-\langle A(y)\xi,\xi\rangle\big{|}}{\langle A(x)\xi,\xi\rangle^{1/2}+\langle A(y)\xi,\xi\rangle^{1/2}}\leq\frac{1}{2\lambda}||A(x)-A(y)|| (2.11)

for all x,y,ξnx,y,\xi\in\mathbb{R}^{n} with |ξ|=1|\xi|=1. Note our regularity assumption is much weaker than in [Alm68] where he assumes the integrand f=f(x,ξ)f=f(x,\xi) is CkC^{k} for some k3k\geq 3 and weaker than the assumption in [SS82] where they assume the integrand f=f(x,ξ)f=f(x,\xi) is Lipschitz in xx.

Remark 2.4 (Comparability to perimeter).

A(E;)\mathscr{F}_{A}(E;\>\cdot\>) is comparable to P(E;)P(E;\>\cdot\>) since it follows for all Borel sets FF that

λ1/2P(E;F)A(E;F)Λ1/2P(E;F).\displaystyle\lambda^{1/2}P(E;F)\leq\mathscr{F}_{A}(E;F)\leq\Lambda^{1/2}P(E;F). (2.12)

by the uniformly ellipticity of AA. When AA equals the identity matrix II we have the isotropic case A(E;)=P(E;)\mathscr{F}_{A}(E;\>\cdot\>)=P(E;\>\cdot\>).

Remark 2.5.

The complement Ec=nEE^{c}=\mathbb{R}^{n}\setminus E of a set of locally finite perimeter is also a set of locally finite perimeter with μEc=νEn1  E\mu_{E^{c}}=-\nu_{E}\operatorname{\mathcal{H}^{n-1}}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>\partial^{*}\!E and so A(Ec;)=A(E;)\mathscr{F}_{A}(E^{c};\>\cdot\>)=\mathscr{F}_{A}(E;\>\cdot\>).

2.2. Notions of almost-minimizers

We are interested in studying the boundary regularity of those sets of locally finite perimeter which are almost-minimizers of the A\mathscr{F}_{A}-surface energy in an open set when compared to their local compactly contained variations. Recent work addressing regularity of almost-minimizers for other variational problems can be found in [STV18, dQT18, DESVGT19, JPSVG20] and the notions of almost-minimizers we consider are similar.

Fix universal constants n2n\geq 2, 0<λΛ<+0<\lambda\leq\Lambda<+\infty, κ0\kappa\geq 0, α(0,1)\alpha\in(0,1) and r0(0,+)r_{0}\in(0,+\infty), and let A=(aij(x))i,j=1nA=(a_{ij}(x))_{i,j=1}^{n} be a symmetric, uniformly elliptic, and Hölder continuous with respect to λ\lambda, Λ\Lambda, and α\alpha and fix an open set UU in n\mathbb{R}^{n}.

Definition 2.2 ((κ,α)(\kappa,\alpha)-almost-minimizer of A\mathscr{F}_{A}).

We say a set of locally finite perimeter EE in n\mathbb{R}^{n} is a (κ,α)(\kappa,\alpha)-(additive) almost-minimizer of A\mathscr{F}_{A} in UU at scale r0r_{0} if sptμE=E\operatorname{spt}\mu_{E}=\partial E and

A(E;B(x,r))A(F;B(x,r))+κrα+n1\displaystyle\mathscr{F}_{A}(E;\textbf{{B}}(x,r))\leq\mathscr{F}_{A}(F;\textbf{{B}}(x,r))+\kappa r^{\alpha+n-1} (2.13)

whenever EΔFUB(x,r)E\Delta F\subset\subset U\cap\textbf{{B}}(x,r) where FF is a set of locally finite perimeter, xUx\in U, and r<r0r<r_{0}.

When (2.13) holds with κ=0\kappa=0, we say that EE is a local minimizer of A\mathscr{F}_{A} in UU at scale r0r_{0}, and when (2.13) holds for all scales r0(0,+)r_{0}\in(0,+\infty), we say that EE is a minimizer of A\mathscr{F}_{A} in UU. Typically we will omit the descriptor additive when discussing almost-minimizers. However, we will include it when we wish to highlight the difference from the following alternative notion of almost-minimality.

Definition 2.3 ((κ,α)(\kappa,\alpha)-multiplicative almost-minimizer of A\mathscr{F}_{A}).

We say a set of locally finite perimeter EE in n\mathbb{R}^{n} is a (κ,α)(\kappa,\alpha)-multiplicative almost-minimizer of A\mathscr{F}_{A} in UU at scale r0r_{0} if sptμE=E\operatorname{spt}\mu_{E}=\partial E and

A(E;B(x,r))(1+κrα)A(F;B(x,r))\displaystyle\mathscr{F}_{A}(E;\textbf{{B}}(x,r))\leq(1+\kappa r^{\alpha})\mathscr{F}_{A}(F;\textbf{{B}}(x,r)) (2.14)

whenever EΔFUB(x,r)E\Delta F\subset\subset U\cap\textbf{{B}}(x,r) where FF is a set of locally finite perimeter, xUx\in U, and r<r0r<r_{0}.

Note that Taylor worked with this notion of multiplicative almost-minimizer in [Tay76b] but handled a more general error term. We now show that multiplicative almost-minimizers are also additive almost-minimizers. To prove this, we need an upper bound for perimeter bounds of multiplicative almost-minimizers at points in the topological boundary. Whenever we write CC we mean a constant (which may change from line to line) that depends only on the universal constants n,λ,Λ,κ,α,r0n,\lambda,\Lambda,\kappa,\alpha,r_{0} and an upper bound for ACα||A||_{C^{\alpha}}, but does not depend on EE or x0x_{0}. If we wish to specify dependence on fewer constants and write for example, C(n)C(n) for constants that only depend on nn.

Lemma 2.1.

There exists a positive constants C=C(n,λ,Λ,κ,α,r0)C=C(n,\lambda,\Lambda,\kappa,\alpha,r_{0}) with the following property. If EE is a (κ,α)(\kappa,\alpha)-multiplicative almost-minimizer of A\mathscr{F}_{A} in UU at scale r0r_{0}, then for every x0UEx_{0}\in U\cap\partial E with r<d=min{dist(x0,U),r0}<r<d=\min\{\mathrm{dist}(x_{0},\partial U),r_{0}\}<\infty,

P(E;B(x0,r))rn1C\displaystyle\frac{P(E;\textbf{{B}}(x_{0},r))}{r^{n-1}}\leq C (2.15)
Proof.

Since n1  E\operatorname{\mathcal{H}^{n-1}}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>\partial^{*}\!E is Radon, n1(EB(x0,r))=0\operatorname{\mathcal{H}^{n-1}}(\partial^{*}\!E\cap\partial\textbf{{B}}(x_{0},r))=0 for a.e. r(0,d)r\in(0,d). Choose one such radius rr and for s(r,d)s\in(r,d) consider the comparison set F=EB(x0,r)F=E\setminus\textbf{{B}}(x_{0},r) in B(x0,s)\textbf{{B}}(x_{0},s). Then EΔFB(x0,r)B(x0,s)E\Delta F\subset\textbf{{B}}(x_{0},r)\subset\subset\textbf{{B}}(x_{0},s). It follows from comparability to perimeter (2.12) and the multiplicative almost-minimality of EE that

λ1/2P(E;B(x0,s))\displaystyle\lambda^{1/2}P(E;\textbf{{B}}(x_{0},s)) A(E;B(x0,s))\displaystyle\leq\mathscr{F}_{A}(E;\textbf{{B}}(x_{0},s))
(1+κsα)A(EB(x0,r);B(x0,s))\displaystyle\leq(1+\kappa s^{\alpha})\mathscr{F}_{A}(E\setminus\textbf{{B}}(x_{0},r);\textbf{{B}}(x_{0},s))
(1+κr0α)Λ1/2P(EB(x0,r);B(x0,s)).\displaystyle\leq(1+\kappa r_{0}^{\alpha})\Lambda^{1/2}P(E\setminus\textbf{{B}}(x_{0},r);\textbf{{B}}(x_{0},s)). (2.16)

Hence

P(E;B(x0,s))\displaystyle P(E;\textbf{{B}}(x_{0},s)) CP(EB(x0,r);B(x0,s))\displaystyle\leq C\>P(E\setminus\textbf{{B}}(x_{0},r);\textbf{{B}}(x_{0},s))
=C(n1(E(1)B(x0,r))+P(E;B(x0,s)B¯(x0,r)))\displaystyle=C\big{(}\operatorname{\mathcal{H}^{n-1}}(E^{(1)}\cap\partial\textbf{{B}}(x_{0},r))+P(E;\textbf{{B}}(x_{0},s)\setminus\overline{\textbf{{B}}}(x_{0},r))\big{)} (2.17)

since n1(EB(x0,r))=0\operatorname{\mathcal{H}^{n-1}}(\partial^{*}\!E\cap\partial\textbf{{B}}(x_{0},r))=0. Sending sr+s\to r^{+} and noting n1(E(1)B(x0,r))nωnrn1\operatorname{\mathcal{H}^{n-1}}(E^{(1)}\cap\partial\textbf{{B}}(x_{0},r))\leq n\omega_{n}r^{n-1} gives

P(E;B(x0,r))Cn1(E(1)B(x0,r))Crn1.\displaystyle P(E;\textbf{{B}}(x_{0},r))\leq C\operatorname{\mathcal{H}^{n-1}}(E^{(1)}\cap\partial\textbf{{B}}(x_{0},r))\leq Cr^{n-1}. (2.18)

By density of these radii, this holds for all r(0,d)r\in(0,d). ∎

Proposition 2.2 (Multiplicative almost-minimizers are (additive) almost-minimizers).

If EE is a (κ,α)(\kappa,\alpha)-multiplicative almost-minimizer of A\mathscr{F}_{A} in UU at scale r0r_{0}, then for each open set VUV\subset\subset U, there is a constant κ=κ(n,λ,Λ,κ,α,r0)\kappa^{\prime}=\kappa^{\prime}(n,\lambda,\Lambda,\kappa,\alpha,r_{0}) such that EE is a (κ,α)(\kappa^{\prime},\alpha)-(additive) almost-minimizer of A\mathscr{F}_{A} in VV at scale r0=min{(1/2)r0,(1/4)dist(V,Uc)}r_{0}^{\prime}=\min\{(1/2)r_{0},(1/4)\mathrm{dist}(V,U^{c})\}.

Proof.

Let EΔFB(x,r)VE\Delta F\subset\subset\textbf{{B}}(x,r)\cap V, xVx\in V, and r<r0r<r_{0}^{\prime}. Suppose EE is a (κ,α)(\kappa,\alpha)-multiplicative almost-minimizer of A\mathscr{F}_{A} in UU at scale r0r_{0}. The minimality condition is trivially satisfied if A(E;B(x,r))A(F;B(x,r))\mathscr{F}_{A}(E;\textbf{{B}}(x,r))\leq\mathscr{F}_{A}(F;\textbf{{B}}(x,r)) or P(E;B(x,r))=0P(E;\textbf{{B}}(x,r))=0. So suppose A(F;B(x,r))A(E;B(x,r))\mathscr{F}_{A}(F;\textbf{{B}}(x,r))\leq\mathscr{F}_{A}(E;\textbf{{B}}(x,r)) and P(E;B(x,r))>0P(E;\textbf{{B}}(x,r))>0. Then there is yB(x,r)Ey\in\textbf{{B}}(x,r)\cap\partial E. So by Lemma 2.1, which applies since 2r<r02r<r_{0} and B(y,2r)B(x,4r)U\textbf{{B}}(y,2r)\subset\textbf{{B}}(x,4r)\subset U, we have P(E;B(x,r))P(E;B(y,2r))C(2r)n1P(E;\textbf{{B}}(x,r))\leq P(E;\textbf{{B}}(y,2r))\leq C(2r)^{n-1}. Hence by comparability to perimeter (2.12) we have A(F;B(x,r))Λ1/2P(E;B(x,r))Crn1\mathscr{F}_{A}(F;\textbf{{B}}(x,r))\leq\Lambda^{1/2}P(E;\textbf{{B}}(x,r))\leq Cr^{n-1}. It follows that

A(E;B(x,r))A(F;B(x,r))+κrαA(F;B(x,r))A(F;B(x,r))+κrα+n1\displaystyle\mathscr{F}_{A}(E;\textbf{{B}}(x,r))\leq\mathscr{F}_{A}(F;\textbf{{B}}(x,r))+\kappa r^{\alpha}\mathscr{F}_{A}(F;\textbf{{B}}(x,r))\leq\mathscr{F}_{A}(F;\textbf{{B}}(x,r))+\kappa^{\prime}r^{\alpha+n-1} (2.19)

for some κ=κ(n,λ,Λ,κ,α,r0)\kappa^{\prime}=\kappa^{\prime}(n,\lambda,\Lambda,\kappa,\alpha,r_{0}). ∎

Thus Proposition 2.2 implies that any interior regularity results for (additive) almost-minimizers shall also apply to multiplicative almost-minimizers. We shall focus on proving a regularity theorem for (additive) almost-minimizers and shall henceforth only work with (additive) almost-minimizers which we simply refer to as almost-minimizers.

3. Existence of Anisotropic Minimizers

Our first order of business is to establish existence of solutions to the anisotropic Plateau problem for A\mathscr{F}_{A}. The existence of anisotropic minimizers in the setting of varifolds and currents is known in general in the framework of varifolds and currents (see [Fed69, Chapter 5]) which should imply existence of minimizers of A\mathscr{F}_{A} in the framework of sets of locally finite perimeter. However, for completeness, we present our own full proof of this result in our setting. Additionally, the lower semicontinuity result of Proposition 3.4 will prove useful at several places in the regularity portion of our paper.

Let A=(aij(x))i,j=1nA=(a_{ij}(x))_{i,j=1}^{n} be a symmetric, uniformly elliptic, continuous function on n\mathbb{R}^{n} with values in nn\mathbb{R}^{n}\otimes\mathbb{R}^{n} (we do not need Hölder continuity to show existence of minimizers) and consider the A\mathscr{F}_{A}-surface energy. Fix an open bounded set UU and a set of finite perimeter E0E_{0} in n\mathbb{R}^{n}. The anisotropic Plateau problem for A\mathscr{F}_{A} in UU with boundary data E0E_{0} is to show that the infimum

γA(E0,U)=inf{A(E):E a set of finite perimeter in n with EU=E0U}\displaystyle\gamma_{A}(E_{0},U)=\inf\Big{\{}\mathscr{F}_{A}(E)\mathrel{\mathop{\mathchar 58\relax}}E\text{ a set of finite perimeter in }\mathbb{R}^{n}\text{ with }E\setminus U=E_{0}\setminus U\Big{\}} (3.1)

is attained (Cf. [Mag12, (12.29)]). That is, we minimize A\mathscr{F}_{A} in n\mathbb{R}^{n} among those sets of finite perimeter which agree with E0E_{0} outside of UU.

To show that (3.1) is achieved by a set of finite perimeter, we follow the Direct Method of the Calculus of Variations. This consists of (i)(i) taking a sequence {Eh}h\{E_{h}\}_{h\in\mathbb{N}} of competitors such that A(Eh)γA(E0,U)\mathscr{F}_{A}(E_{h})\to\gamma_{A}(E_{0},U), (ii)(ii) using a key compactness result in an appropriate topology to extract a subsequence {Eh(k)}k\{E_{h(k)}\}_{k\in\mathbb{N}} converging to some competitor EE satisfying EU=E0UE\setminus U=E_{0}\setminus U, and (iii)(iii) applying lower semicontinuity of A\mathscr{F}_{A} with respect to the convergence in the chosen topology which shows that A(E)\mathscr{F}_{A}(E) equals the infimum γA(E0,U)\gamma_{A}(E_{0},U) in (3.1).

3.1. Compactness of sets of locally finite perimeter

The first key ingredient of the Direct Method is compactness of our class of admissible competitors. One of the primary reasons that sets of locally finite perimeter provide a suitable setting to work on geometric variational problems is that they possess compactness with respect to local convergence of sets. Let’s recall the definition of this convergence and a known compactness theorem for sets of locally finite perimeter.

We say that a sequence of sets of locally finite perimeter {Eh}h\{E_{h}\}_{h\in\mathbb{N}} in n\mathbb{R}^{n} converges locally to EE (and write EhlocEE_{h}\overset{\text{loc}}{\rightarrow}E) if

|(EhΔE)K|0 as h\displaystyle|(E_{h}\Delta E)\cap K|\to 0\text{ as }h\to\infty (3.2)

for each compact KnK\subset\mathbb{R}^{n}, and say {Eh}h\{E_{h}\}_{h\in\mathbb{N}} converges to EE (and write EhEE_{h}\to E) if

|EhΔE|0 as h.\displaystyle|E_{h}\Delta E|\to 0\text{ as }h\to\infty. (3.3)

Recall that EΔF=(EF)(FE)E\Delta F=(E\setminus F)\cup(F\setminus E) and that |||\cdot| denotes Lebesgue measure on n\mathbb{R}^{n}.

Theorem 3.1 (Compactness from perimeter bounds, [Mag12, Theorem 12.26]).

If R>0R>0 and {Eh}h\{E_{h}\}_{h\in\mathbb{N}} are sets of finite perimeter in n\mathbb{R}^{n}, with

EhBR,h, and suphP(Eh)<,\displaystyle E_{h}\subset\textbf{{B}}_{R},\ \forall h\in\mathbb{N},\qquad\text{ and }\qquad\sup_{h\in\mathbb{N}}P(E_{h})<\infty, (3.4)

then there exist a set EE of finite perimeter in n\mathbb{R}^{n} and indices h(k)h(k)\to\infty as kk\to\infty, with

Eh(k)E,μEh(k)μE, and EBR.\displaystyle E_{h(k)}\to E,\qquad\mu_{E_{h(k)}}\overset{\ast}{\rightharpoonup}\mu_{E},\qquad\text{ and }\qquad E\subset\textbf{{B}}_{R}. (3.5)

3.2. Lower semicontinuity of A\mathscr{F}_{A}

The second key ingredient of the Direct Method is to show lower semicontinuity of the A\mathscr{F}_{A}-surface energy. Here we have some work to do and start with a couple lemmas. The first lemma deals with lower semicontinuity when AA is constant, while the second one is a technical lemma we need in the proof when AA is no longer constant.

Lemma 3.2 (Lower semicontinuity for constant AA).

If AA is a constant, uniformly elliptic matrix, and {Eh}h\{E_{h}\}_{h\in\mathbb{N}} and EE are sets of locally finite perimeter with νEhn1  EhνEn1  E\nu_{E_{h}}\operatorname{\mathcal{H}^{n-1}}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>\>\partial^{*}\!E_{h}\overset{\ast}{\rightharpoonup}\nu_{E}\operatorname{\mathcal{H}^{n-1}}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>\>\partial^{*}\!E, then for any open set UU,

A(E;U)lim infhA(Eh;U).\displaystyle\mathscr{F}_{A}(E;U)\leq\liminf_{h\to\infty}\mathscr{F}_{A}(E_{h};U). (3.6)
Proof.

By Remark 2.1 we may assume AA is symmetric and by uniform ellipticity its eigenvalues are positive. So by the spectral theorem we can write A=VDV1A=VDV^{-1} where DD is a diagonal matrix with the eigenvalues of AA and where VV is the matrix of corresponding orthonormal eigenvectors. Setting A1/2=VD1/2V1A^{1/2}=VD^{1/2}V^{-1}, we have A=A1/2A1/2A=A^{1/2}A^{1/2} with A1/2A^{1/2} symmetric since V1=VtV^{-1}=V^{t}. So Aξ,ξ1/2=|A1/2ξ|\langle A\xi,\xi\rangle^{1/2}=|A^{1/2}\xi|. Define n\mathbb{R}^{n}-valued Radon measures on n\mathbb{R}^{n},

μh=A1/2νEhn1 Eh and μ=A1/2νEn1 E.\displaystyle\mu_{h}=A^{1/2}\nu_{E_{h}}\operatorname{\mathcal{H}^{n-1}}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>\>\partial^{*}\!E_{h}\qquad\text{ and }\qquad\mu=A^{1/2}\nu_{E}\operatorname{\mathcal{H}^{n-1}}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>\>\partial^{*}\!E. (3.7)

It then follows that μhμ\mu_{h}\overset{\ast}{\rightharpoonup}\mu because, given any φCc(n;n)\varphi\in C_{c}(\mathbb{R}^{n};\mathbb{R}^{n}), we have A1/2φCc(n;n)A^{1/2}\varphi\in C_{c}(\mathbb{R}^{n};\mathbb{R}^{n}) and thus

limhφ𝑑μh\displaystyle\lim_{h\to\infty}\int\varphi\cdot d\mu_{h} =limhEhφ,A1/2νEhdn1=limhEhA1/2φ,νEhdn1\displaystyle=\lim_{h\to\infty}\int_{\partial^{*}\!E_{h}}\langle\varphi,A^{1/2}\nu_{E_{h}}\rangle\operatorname{\,d\mathcal{H}^{n-1}}=\lim_{h\to\infty}\int_{\partial^{*}\!E_{h}}\langle A^{1/2}\varphi,\nu_{E_{h}}\rangle\operatorname{\,d\mathcal{H}^{n-1}}
=EA1/2φ,νEdn1=Eφ,A1/2νEdn1=φ𝑑μ.\displaystyle=\int_{\partial^{*}\!E}\langle A^{1/2}\varphi,\nu_{E}\rangle\operatorname{\,d\mathcal{H}^{n-1}}=\int_{\partial^{*}\!E}\langle\varphi,A^{1/2}\nu_{E}\rangle\operatorname{\,d\mathcal{H}^{n-1}}=\int\varphi\cdot d\mu. (3.8)

By lower semicontinuity of the total variation of weak-star convergent vector-valued Radon measures ([Mag12, Proposition 4.19]), we have

A(E;U)\displaystyle\mathscr{F}_{A}(E;U) =UE|A1/2νE|dn1=|μ|(U)lim infh|μh|(U)\displaystyle=\int_{U\cap\partial^{*}\!E}|A^{1/2}\nu_{E}|\operatorname{\,d\mathcal{H}^{n-1}}=|\mu|(U)\leq\liminf_{h\to\infty}|\mu_{h}|(U)
=lim infhUEh|A1/2νEh|dn1=lim infhA(Eh;U)\displaystyle=\liminf_{h\to\infty}\int_{U\cap\partial^{*}\!E_{h}}|A^{1/2}\nu_{E_{h}}|\operatorname{\,d\mathcal{H}^{n-1}}=\liminf_{h\to\infty}\mathscr{F}_{A}(E_{h};U) (3.9)

which concludes the proof. ∎

Lemma 3.3.

Let {Φh}h\{\Phi_{h}\}_{h\in\mathbb{N}} and Φ\Phi be Radon measures on n\mathbb{R}^{n} and φCc(n;[0,))\varphi\in C_{c}(\mathbb{R}^{n};[0,\infty)) such that
lim suphΦh({φ>0})<\displaystyle\limsup_{h\to\infty}\Phi_{h}(\{\varphi>0\})<\infty. Then the following two statements hold:

  1. (i)

    If Φ(U)lim infhΦh(U)\displaystyle\Phi(U)\leq\liminf_{h\to\infty}\Phi_{h}(U) for any open set UU, then

    (φΦ)(U)lim infh(φΦh)(U)\displaystyle(\varphi\Phi)(U)\leq\liminf_{h\to\infty}(\varphi\Phi_{h})(U) (3.10)

    for any open set UU in n\mathbb{R}^{n}.

  2. (ii)

    If lim suphΦh(K)Φ(K)\displaystyle\limsup_{h\to\infty}\Phi_{h}(K)\leq\Phi(K) for any compact set KK, then

    lim suph(φΦh)(K)(φΦ)(K)\displaystyle\limsup_{h\to\infty}(\varphi\Phi_{h})(K)\leq(\varphi\Phi)(K) (3.11)

    for any compact set KK in n\mathbb{R}^{n}.

Proof.

Let ε>0\varepsilon>0 and choose 0=t0<t1<<tN1<supφ<tN0=t_{0}<t_{1}<\cdots<t_{N-1}<\sup\varphi<t_{N} such that tjtj1<εt_{j}-t_{j-1}<\varepsilon and Φ({φ=tj})=0\Phi(\{\varphi=t_{j}\})=0 for j=1,,Nj=1,\dots,N. This is possible since Φ\Phi is Radon and so Φ({φ=t})>0\Phi(\{\varphi=t\})>0 for at most countably many tt. Set

Uj={tj1<φ<tj}andKj=U¯j\displaystyle U_{j}=\{t_{j-1}<\varphi<t_{j}\}\qquad\text{and}\qquad K_{j}=\overline{U}_{j} (3.12)

and note that the UjU_{j}’s are open and the KjK_{j}’s are compact.

Proof of (i)(i): Assume the hypothesis and let UU be an open set. Observe that

(φΦ)(U)=Uφ𝑑Φ=j=1NUUjφ𝑑Φ+j=0N1tjΦ(U{φ=tj})=j=1NUUjφ𝑑Φ\displaystyle(\varphi\Phi)(U)=\int_{U}\varphi\>d\Phi=\sum_{j=1}^{N}\int_{U\cap U_{j}}\varphi\>d\Phi+\sum_{j=0}^{N-1}t_{j}\Phi(U\cap\{\varphi=t_{j}\})=\sum_{j=1}^{N}\int_{U\cap U_{j}}\varphi\>d\Phi (3.13)

since Φ(U{φ=tj})=0\Phi(U\cap\{\varphi=t_{j}\})=0 for j=1,,N1j=1,\dots,N-1. Since φ<tj\varphi<t_{j} on UUjU\cap U_{j} and Φ(UUj)lim infhΦh(UUj)\Phi(U\cap U_{j})\leq\liminf_{h}\Phi_{h}(U\cap U_{j}) for j=1,,Nj=1,\dots,N, we have

(φΦ)(U)j=1NUUjtj𝑑Φj=1Ntjlim infhΦh(UUj)=lim infhj=1NtjΦh(UUj),\displaystyle(\varphi\Phi)(U)\leq\sum_{j=1}^{N}\int_{U\cap U_{j}}t_{j}\>d\Phi\leq\sum_{j=1}^{N}t_{j}\liminf_{h\to\infty}\Phi_{h}(U\cap U_{j})=\liminf_{h\to\infty}\sum_{j=1}^{N}t_{j}\Phi_{h}(U\cap U_{j}), (3.14)

where we used the property that lim infhah+lim infhbhlim infh(ah+bh)\liminf_{h}a_{h}+\liminf_{h}b_{h}\leq\liminf_{h}(a_{h}+b_{h}) for any sequences {ah},{bh}\{a_{h}\},\{b_{h}\}. Note that tj<tj1+ε<φ+εt_{j}<t_{j-1}+\varepsilon<\varphi+\varepsilon on UUjU\cap U_{j} and so

(φΦ)(U)\displaystyle(\varphi\Phi)(U) lim infhj=1NtjΦh(UUj)lim infhj=1NUUj(φ+ε)𝑑Φh\displaystyle\leq\liminf_{h\to\infty}\sum_{j=1}^{N}t_{j}\Phi_{h}(U\cap U_{j})\leq\liminf_{h\to\infty}\sum_{j=1}^{N}\int_{U\cap U_{j}}(\varphi+\varepsilon)\>d\Phi_{h}
lim infh(φΦh)(U)+εlim suphΦh(U{φ>0})\displaystyle\leq\liminf_{h\to\infty}(\varphi\Phi_{h})(U)+\varepsilon\limsup_{h\to\infty}\Phi_{h}(U\cap\{\varphi>0\}) (3.15)

since U{φ>0}=j=1NUUjU\cap\{\varphi>0\}=\bigcup_{j=1}^{N}U\cap U_{j}. Sending ε0+\varepsilon\to 0^{+} completes the proof of (i)(i).

Proof of (ii)(ii): Assume the hypothesis and let KK be a compact set. Recalling Kj=U¯jK_{j}=\overline{U}_{j}, observe that

(φΦ)(K)=Kφ𝑑Φ=j=1NKKjφ𝑑Φj=0N1tjΦ(K{φ=tj})=j=1NKKjφ𝑑Φ,\displaystyle(\varphi\Phi)(K)=\int_{K}\varphi\>d\Phi=\sum_{j=1}^{N}\int_{K\cap K_{j}}\varphi\>d\Phi-\sum_{j=0}^{N-1}t_{j}\Phi(K\cap\{\varphi=t_{j}\})=\sum_{j=1}^{N}\int_{K\cap K_{j}}\varphi\>d\Phi, (3.16)

since Φ(K{φ=tj})=0\Phi(K\cap\{\varphi=t_{j}\})=0 for j=1,,N1j=1,\dots,N-1, and

(φΦh)(K)=Kφ𝑑Φh=j=1NKKjφ𝑑Φhj=0N1tjΦh(K{φ=tj})j=1NKKjφ𝑑Φh.\displaystyle(\varphi\Phi_{h})(K)=\int_{K}\varphi\>d\Phi_{h}=\sum_{j=1}^{N}\int_{K\cap K_{j}}\varphi\>d\Phi_{h}-\sum_{j=0}^{N-1}t_{j}\Phi_{h}(K\cap\{\varphi=t_{j}\})\leq\sum_{j=1}^{N}\int_{K\cap K_{j}}\varphi\>d\Phi_{h}. (3.17)

It follows that

lim suph(φΦh)(K)lim suphj=1NKKjφ𝑑Φhj=1Nlim suphKKjφ𝑑Φh,\displaystyle\limsup_{h\to\infty}(\varphi\Phi_{h})(K)\leq\limsup_{h\to\infty}\sum_{j=1}^{N}\int_{K\cap K_{j}}\varphi\>d\Phi_{h}\leq\sum_{j=1}^{N}\limsup_{h\to\infty}\int_{K\cap K_{j}}\varphi\>d\Phi_{h}, (3.18)

where we used the property that lim suph(ah+bh)lim suphah+lim suphbh\limsup_{h}(a_{h}+b_{h})\leq\limsup_{h}a_{h}+\limsup_{h}b_{h} for any sequences {ah},{bh}\{a_{h}\},\{b_{h}\}. Since φ<tj\varphi<t_{j} on KKjK\cap K_{j} and lim suphΦh(KKj)Φ(KKj)\limsup_{h}\Phi_{h}(K\cap K_{j})\leq\Phi(K\cap K_{j}), we have

lim suph(φΦh)(K)j=1Nlim suphKKjφ𝑑Φhj=1NtjΦ(KKj).\displaystyle\limsup_{h\to\infty}(\varphi\Phi_{h})(K)\leq\sum_{j=1}^{N}\limsup_{h\to\infty}\int_{K\cap K_{j}}\varphi\>d\Phi_{h}\leq\sum_{j=1}^{N}t_{j}\Phi(K\cap K_{j}). (3.19)

Note that tj<tj1+εφ+εt_{j}<t_{j-1}+\varepsilon\leq\varphi+\varepsilon on KKjK\cap K_{j} and so

lim suph(φΦh)(K)j=1N(φ+ε)Φ(KKj)=(φΦ)(K)+εΦ(Ksptφ)\displaystyle\limsup_{h\to\infty}(\varphi\Phi_{h})(K)\leq\sum_{j=1}^{N}(\varphi+\varepsilon)\Phi(K\cap K_{j})=(\varphi\Phi)(K)+\varepsilon\Phi(K\cap\operatorname{spt}\varphi) (3.20)

where we used Φ(K{φ=tj})=0\Phi(K\cap\{\varphi=t_{j}\})=0 for j=1,,N1j=1,\dots,N-1. Sending ε0+\varepsilon\to 0^{+} completes the proof of (ii)(ii). ∎

With these lemmas in hand, we are now ready to state and prove the lower semicontinuity of A\mathscr{F}_{A}.

Proposition 3.4 (Lower semicontinuity of A\mathscr{F}_{A}).

Let A=(aij(x))i,j=1nA=(a_{ij}(x))_{i,j=1}^{n} be a symmetric, uniformly elliptic, continuous function on n\mathbb{R}^{n} with values in nn\mathbb{R}^{n}\otimes\mathbb{R}^{n}. Suppose {Eh}h\{E_{h}\}_{h\in\mathbb{N}} is a sequence of sets of locally finite perimeter in n\mathbb{R}^{n} and EE is Lebesgue measurable, with

EhlocE, and lim suphP(Eh;K)<\displaystyle E_{h}\overset{\text{loc}}{\rightarrow}E,\qquad\text{ and }\qquad\limsup_{h\to\infty}P(E_{h};K)<\infty (3.21)

for every compact set KK in n\mathbb{R}^{n}. Then EE is a set of locally finite perimeter in n\mathbb{R}^{n} with

νEhn1 EhνEn1 E,\displaystyle\nu_{E_{h}}\operatorname{\mathcal{H}^{n-1}}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>\>\partial^{*}\!E_{h}\overset{\ast}{\rightharpoonup}\nu_{E}\operatorname{\mathcal{H}^{n-1}}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>\>\partial^{*}\!E, (3.22)

and for any open set UU in n\mathbb{R}^{n},

A(E;U)lim infhA(Eh;U).\displaystyle\mathscr{F}_{A}(E;U)\leq\liminf_{h\to\infty}\mathscr{F}_{A}(E_{h};U). (3.23)
Proof.

That EE is of locally finite perimeter and νEhn1  EhνEn1  E\nu_{E_{h}}\operatorname{\mathcal{H}^{n-1}}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>\>\partial^{*}\!E_{h}\overset{\ast}{\rightharpoonup}\nu_{E}\operatorname{\mathcal{H}^{n-1}}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>\>\partial^{*}\!E follow from [Mag12, Proposition 12.15]. Thus we need only to prove the lower semicontinuity.

First assume that UU is bounded. By taking a subsequence of {A(Eh;U)}h\{\mathscr{F}_{A}(E_{h};U)\}_{h\in\mathbb{N}}, we may assume up to relabeling that

limhA(Eh;U)=lim infhA(Eh;U)<.\displaystyle\lim_{h\to\infty}\mathscr{F}_{A}(E_{h};U)=\liminf_{h\to\infty}\mathscr{F}_{A}(E_{h};U)<\infty. (3.24)

Note this subsequence depends on UU but this is not an issue. Since lim suphP(Eh;K)<\limsup_{h\to\infty}P(E_{h};K)<\infty for every compact set KK, there is a further subsequence {Eh(k)}k\{E_{h(k)}\}_{k\in\mathbb{N}} and a Radon measure Ψ\Psi such that n1  Eh(k)Ψ\operatorname{\mathcal{H}^{n-1}}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>\>\partial^{*}\!E_{h(k)}\overset{\ast}{\rightharpoonup}\Psi as kk\to\infty (see [Mag12, Remark 4.35]).

Let VUV\subset\subset U be open and fix ε>0\varepsilon>0. Since AA is uniformly continuous on U¯\overline{U}, there exists 0<r<dist(V¯,Uc)0<r<\mathrm{dist}(\overline{V},U^{c}) such that for any x,yU¯x,y\in\overline{U}, we have A(x)A(y)<ε||A(x)-A(y)||<\varepsilon whenever |xy|<r|x-y|<r. Thus by the inequality (2.11), for any x,yU¯x,y\in\overline{U},

A(x)ξ,ξ1/2A(y)ξ,ξ1/2+12λε\displaystyle\langle A(x)\xi,\xi\rangle^{1/2}\leq\langle A(y)\xi,\xi\rangle^{1/2}+\frac{1}{2\lambda}\varepsilon (3.25)

whenever |xy|<r|x-y|<r and |ξ|=1|\xi|=1.

Since VV is compactly contained in UU and sptΦE=E=E¯\operatorname{spt}\Phi_{E}=\partial E=\overline{\partial^{*}\!E}, there exist finitely many balls {B(xj,r)}j=1N\{\textbf{{B}}(x_{j},r)\}_{j=1}^{N} each of radius rr and center xjVEx_{j}\in V\cap\partial^{*}\!E which cover V¯E\overline{V}\cap\partial E. Take a partition of unity {φj}j=1N\{\varphi_{j}\}_{j=1}^{N} with φjCc(B(xj,r),[0,1])\varphi_{j}\in C_{c}(\textbf{{B}}(x_{j},r),[0,1]) such that j=1Nφj=1\sum_{j=1}^{N}\varphi_{j}=1 on VEV\cap\partial^{*}\!E and j=1Nφj1\sum_{j=1}^{N}\varphi_{j}\leq 1 elsewhere. It follows that

lim infhA(Eh;U)\displaystyle\liminf_{h\to\infty}\mathscr{F}_{A}(E_{h};U) =limkA(Eh(k);U)\displaystyle=\lim_{k\to\infty}\mathscr{F}_{A}(E_{h(k)};U)
limkj=1NEh(k)φjA(x)νEh(k),νEh(k)1/2dn1\displaystyle\geq\lim_{k\to\infty}\sum_{j=1}^{N}\int_{\partial^{*}\!E_{h(k)}}\varphi_{j}\langle A(x)\nu_{E_{h(k)}},\nu_{E_{h(k)}}\rangle^{1/2}\operatorname{\,d\mathcal{H}^{n-1}}
limkj=1N[Eh(k)φjA(xj)νEh(k),νEh(k)1/2dn112λεEh(k)φjdn1]\displaystyle\geq\lim_{k\to\infty}\sum_{j=1}^{N}\bigg{[}\int_{\partial^{*}\!E_{h(k)}}\varphi_{j}\langle A(x_{j})\nu_{E_{h(k)}},\nu_{E_{h(k)}}\rangle^{1/2}\operatorname{\,d\mathcal{H}^{n-1}}-\frac{1}{2\lambda}\varepsilon\int_{\partial^{*}\!E_{h(k)}}\varphi_{j}\operatorname{\,d\mathcal{H}^{n-1}}\bigg{]}
j=1Nlim infk[Eh(k)φjA(xj)νEh(k),νEh(k)1/2dn112λεEh(k)φjdn1]\displaystyle\geq\sum_{j=1}^{N}\liminf_{k\to\infty}\bigg{[}\int_{\partial^{*}\!E_{h(k)}}\varphi_{j}\langle A(x_{j})\nu_{E_{h(k)}},\nu_{E_{h(k)}}\rangle^{1/2}\operatorname{\,d\mathcal{H}^{n-1}}-\frac{1}{2\lambda}\varepsilon\int_{\partial^{*}\!E_{h(k)}}\varphi_{j}\operatorname{\,d\mathcal{H}^{n-1}}\bigg{]}
j=1N[EφjA(xj)νE,νE1/2dn112λεlim supkEh(k)φjdn1],\displaystyle\geq\sum_{j=1}^{N}\bigg{[}\int_{\partial^{*}\!E}\varphi_{j}\langle A(x_{j})\nu_{E},\nu_{E}\rangle^{1/2}\operatorname{\,d\mathcal{H}^{n-1}}-\frac{1}{2\lambda}\varepsilon\limsup_{k\to\infty}\int_{\partial^{*}\!E_{h(k)}}\varphi_{j}\operatorname{\,d\mathcal{H}^{n-1}}\bigg{]}, (3.26)

where in the last inequality, for each j=1,,Nj=1,\dots,N, we applied part (i)(i) of Lemma 3.3 to φj\varphi_{j} and the measures dΦk=A(xj)νEh(k),νEh(k)1/2dn1  Eh(k)d\Phi_{k}=\langle A(x_{j})\nu_{E_{h(k)}},\nu_{E_{h(k)}}\rangle^{1/2}\operatorname{\,d\mathcal{H}^{n-1}}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>\ \partial^{*}\!E_{h(k)} and dΦ=A(xj)νE,νE1/2dn1  Ed\Phi=\langle A(x_{j})\nu_{E},\nu_{E}\rangle^{1/2}\operatorname{\,d\mathcal{H}^{n-1}}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>\ \partial^{*}\!E which by Lemma 3.2 satisfies the lower semicontinuity hypothesis. By part (ii)(ii) of Lemma 3.3, applied to n1  Eh(k)Ψ\operatorname{\mathcal{H}^{n-1}}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>\>\partial^{*}\!E_{h(k)}\overset{\ast}{\rightharpoonup}\Psi,

lim supkEh(k)φjdn1sptφjφj𝑑Ψ\displaystyle\limsup_{k\to\infty}\int_{\partial^{*}\!E_{h(k)}}\varphi_{j}\operatorname{\,d\mathcal{H}^{n-1}}\leq\int_{\operatorname{spt}\varphi_{j}}\varphi_{j}\>d\Psi (3.27)

for each j=1,,Nj=1,\dots,N, and so

j=1Nlim supkEh(k)φjdn1j=1Nsptφjφj𝑑ΨΨ(U¯)\displaystyle\sum_{j=1}^{N}\limsup_{k\to\infty}\int_{\partial^{*}\!E_{h(k)}}\varphi_{j}\operatorname{\,d\mathcal{H}^{n-1}}\leq\sum_{j=1}^{N}\int_{\operatorname{spt}\varphi_{j}}\varphi_{j}\>d\Psi\leq\Psi(\overline{U}) (3.28)

since j=1NsptφjU¯\bigcup_{j=1}^{N}\operatorname{spt}\varphi_{j}\subset\overline{U} and j=1Nφj1\sum_{j=1}^{N}\varphi_{j}\leq 1. It follows that

lim infhA(Eh;U)\displaystyle\liminf_{h\to\infty}\mathscr{F}_{A}(E_{h};U) [j=1NEφjA(xj)νE,νE1/2dn1]12λεΨ(U¯)\displaystyle\geq\bigg{[}\sum_{j=1}^{N}\int_{\partial^{*}\!E}\varphi_{j}\langle A(x_{j})\nu_{E},\nu_{E}\rangle^{1/2}\operatorname{\,d\mathcal{H}^{n-1}}\bigg{]}-\frac{1}{2\lambda}\varepsilon\Psi(\overline{U})
j=1N[EφjA(x)νE,νE1/2dn112λεEφjdn1]12λεΨ(U¯)\displaystyle\geq\sum_{j=1}^{N}\bigg{[}\int_{\partial^{*}\!E}\varphi_{j}\langle A(x)\nu_{E},\nu_{E}\rangle^{1/2}\operatorname{\,d\mathcal{H}^{n-1}}-\frac{1}{2\lambda}\varepsilon\int_{\partial^{*}\!E}\varphi_{j}\operatorname{\,d\mathcal{H}^{n-1}}\bigg{]}-\frac{1}{2\lambda}\varepsilon\Psi(\overline{U})
A(E;V)12λε[n1(U¯E)+Ψ(U¯)]\displaystyle\geq\mathscr{F}_{A}(E;V)-\frac{1}{2\lambda}\varepsilon\Big{[}\operatorname{\mathcal{H}^{n-1}}(\overline{U}\cap\partial^{*}\!E)+\Psi(\overline{U})\Big{]} (3.29)

since, as above,

j=1NEφjdn1j=1NsptφjEφjdn1n1(U¯E)\displaystyle\sum_{j=1}^{N}\int_{\partial^{*}\!E}\varphi_{j}\operatorname{\,d\mathcal{H}^{n-1}}\leq\sum_{j=1}^{N}\int_{\operatorname{spt}\varphi_{j}\cap\partial^{*}\!E}\varphi_{j}\operatorname{\,d\mathcal{H}^{n-1}}\leq\operatorname{\mathcal{H}^{n-1}}(\overline{U}\cap\partial^{*}\!E) (3.30)

by j=1NsptφjU¯\bigcup_{j=1}^{N}\operatorname{spt}\varphi_{j}\subset\overline{U} and j=1Nφj1\sum_{j=1}^{N}\varphi_{j}\leq 1. Letting ε0+\varepsilon\to 0^{+}, we obtain A(E;V)lim infhA(Eh;U)\mathscr{F}_{A}(E;V)\leq\liminf_{h}\mathscr{F}_{A}(E_{h};U). Approximating UU by VV from below and using monotone convergence, we obtain A(E;U)lim infhA(Eh;U)\mathscr{F}_{A}(E;U)\leq\liminf_{h}\mathscr{F}_{A}(E_{h};U).

For the case when UU is unbounded, we have A(E;V)lim infhA(Eh;V)lim infhA(Eh;U)\mathscr{F}_{A}(E;V)\leq\liminf_{h\to\infty}\mathscr{F}_{A}(E_{h};V)\leq\liminf_{h\to\infty}\mathscr{F}_{A}(E_{h};U) for every bounded open set VUV\subset U. We conclude by approximating UU from below by bounded open sets VUV\subset U and using monotone convergence. ∎

3.3. Existence theorem of minimizers for A\mathscr{F}_{A}

We now show that the anisotropic Plateau problem for A\mathscr{F}_{A} given by (3.1) has a solution. We follow a similar approach as [Mag12, Theorem 12.29].

Theorem 3.5 (Existence of minimizers for the anisotropic Plateau problem for A\mathscr{F}_{A}).

Let A=(aij(x))i,j=1nA=(a_{ij}(x))_{i,j=1}^{n} be a uniformly elliptic, continuous function on n\mathbb{R}^{n} with values in nn\mathbb{R}^{n}\otimes\mathbb{R}^{n}, let E0E_{0} be a set of finite perimeter in n\mathbb{R}^{n}, and let UU be an open bounded set. There exists a set of finite perimeter EE in n\mathbb{R}^{n} with EU=E0UE\setminus U=E_{0}\setminus U such that A(E)=γA(E0,U)\mathscr{F}_{A}(E)=\gamma_{A}(E_{0},U) from (3.1). In particular, EE is a minimizer of A\mathscr{F}_{A} in UU.

Proof.

Let {Eh}h\{E_{h}\}_{h\in\mathbb{N}} be a sequence of sets of finite perimeter in n\mathbb{R}^{n} with EhU=E0UE_{h}\setminus U=E_{0}\setminus U such that A(Eh)γA(E0;U)\mathscr{F}_{A}(E_{h})\to\gamma_{A}(E_{0};U) as hh\to\infty and A(Eh)A(E0)<\mathscr{F}_{A}(E_{h})\leq\mathscr{F}_{A}(E_{0})<\infty. Consider Mh=EhΔE0UM_{h}=E_{h}\Delta E_{0}\subset U. Noting that by [Mag12, Theorem 16.3], (in particular, by [Mag12, Exercise 16.5]),

P(Mh)P(Eh)+P(E0)2λ1/2A(E0)<.\displaystyle P(M_{h})\leq P(E_{h})+P(E_{0})\leq 2\lambda^{-1/2}\mathscr{F}_{A}(E_{0})<\infty. (3.31)

Hence suphP(Mh)<\sup_{h}P(M_{h})<\infty. Choose R>0R>0 with UBRU\subset\textbf{{B}}_{R} so that MhBRM_{h}\subset\textbf{{B}}_{R}. By Theorem 3.1, there is a set of finite perimeter MBRM\subset\textbf{{B}}_{R} and h(k)h(k)\to\infty as kk\to\infty such that Mh(k)MM_{h(k)}\to M. Up to modifying by a set of measure zero MUM\subset U. Set E=MΔE0E=M\Delta E_{0}. Then EU=E0UE\setminus U=E_{0}\setminus U and note that Eh=MhΔE0E_{h}=M_{h}\Delta E_{0}. Hence Eh(k)EE_{h(k)}\to E since |Eh(k)ΔE|=|Mh(k)ΔM|0|E_{h(k)}\Delta E|=|M_{h(k)}\Delta M|\to 0 as kk\to\infty. Finally, observe that

lim supkP(Eh(k))λ1/2lim supkA(Eh(k))λ1/2A(E0)<.\displaystyle\limsup_{k\to\infty}P(E_{h(k)})\leq\lambda^{-1/2}\limsup_{k\to\infty}\mathscr{F}_{A}(E_{h(k)})\leq\lambda^{-1/2}\mathscr{F}_{A}(E_{0})<\infty. (3.32)

Consequently, by Proposition 3.4,

γA(E0;U)A(E)lim infkA(Eh(k))=γA(E0;U).\displaystyle\gamma_{A}(E_{0};U)\leq\mathscr{F}_{A}(E)\leq\liminf_{k\to\infty}\mathscr{F}_{A}(E_{h(k)})=\gamma_{A}(E_{0};U). (3.33)

Thus A(E)=γA(E0;U)\mathscr{F}_{A}(E)=\gamma_{A}(E_{0};U).

Suppose EΔFUB(x,r)E\Delta F\subset\subset U\cap\textbf{{B}}(x,r), xUx\in U, and r<r0r<r_{0}. Then FU=E0UF\setminus U=E_{0}\setminus U and so A(E)A(F)\mathscr{F}_{A}(E)\leq\mathscr{F}_{A}(F). Since EΔFB(x,r)E\Delta F\subset\subset\textbf{{B}}(x,r), we have A(E;nB(x,r))=A(F;nB(x,r))\mathscr{F}_{A}(E;\mathbb{R}^{n}\setminus\textbf{{B}}(x,r))=\mathscr{F}_{A}(F;\mathbb{R}^{n}\setminus\textbf{{B}}(x,r)). Hence A(E;B(x,r))A(F;B(x,r))\mathscr{F}_{A}(E;\textbf{{B}}(x,r))\leq\mathscr{F}_{A}(F;\textbf{{B}}(x,r)). ∎

4. Basic Properties of Almost-Minimizers

In this section we begin our journey toward proving regularity of almost-minimizers by proving some fundamental properties that almost-minimizers possess and which play a crucial role in our excess-decay argument.

4.1. Invariance under an affine change variable

One of the key ideas that allows us to adapt the standard excess-decay arguments for perimeter minimizers to the setting is a certain change of variable.

If AA is a constant matrix, then by symmetry we can orthogonally diagonalize AA and write A=VDV1A=VDV^{-1}, where DD is a diagonal matrix with the eigenvalues of AA and where VV is the matrix of corresponding orthonormal eigenvectors. By ellipticity the eigenvalues of AA are bounded below and above by the positive constants λ\lambda and Λ\Lambda. Setting A1/2=VD1/2V1A^{1/2}=VD^{1/2}V^{-1}, we have A=A1/2A1/2A=A^{1/2}A^{1/2}. Note that A1/2A^{1/2} and A1/2A^{-1/2} are symmetric since V1=VtV^{-1}=V^{t}. In the coordinate system of VV, the matrix A1/2A^{-1/2} is diagonal and so almost-minimizers of A\mathscr{F}_{A} can be viewed as almost-minimizers of perimeter when deformed by the change of variable y=T(x)=A1/2xy=T(x)=A^{-1/2}x (see Proposition 4.1 below). Of course this change of variable preserves any regularity of almost-minimizers and we know by Tamanini’s work in [Tam84] that almost-minimizers of perimeter are Hölder continuously differentiable.

If A=(aij(x))i,j=1nA=(a_{ij}(x))_{i,j=1}^{n} varies Hölder continuously, then almost-minimizers of A\mathscr{F}_{A} cannot simply be viewed as almost-minimizers of perimeter since deformation varies from point to point. However, philosophically it is reasonable to expect a similar amount of regularity since the deformation varies Hölder continuously. In subsequent sections we will prove decay estimates for the excess at points x0Ex_{0}\in\partial E with small excess on some ball or cylinder. In the proofs of these estimates it will be convenient to be able to assume that A(x0)=IA(x_{0})=I, allowing us to think of A\mathscr{F}_{A} as a perturbation of perimeter at the point x0x_{0}. In order to make this assumption, we shall do the following change of variable which was similarly used in [DESVGT19, JPSVG20] for almost-minimizers of other types of functionals involving coefficients A=(aij(x))i,j=1nA=(a_{ij}(x))_{i,j=1}^{n}. As in the constant case, for each fixed x0nx_{0}\in\mathbb{R}^{n} we can write A(x0)=VDV1A(x_{0})=VDV^{-1}, where DD is a diagonal matrix with the eigenvalues of A(x0)A(x_{0}) and where VV is the matrix of corresponding orthonormal eigenvectors. Setting A1/2(x0)=VD1/2V1A^{1/2}(x_{0})=VD^{1/2}V^{-1}, we have that A1/2(x0)A^{1/2}(x_{0}) and A1/2(x0)A^{-1/2}(x_{0}) are symmetric since V1=VtV^{-1}=V^{t} and satisfy

λ1/2|ξ||A1/2(x0)ξ|Λ1/2|ξ|,Λ1/2|ξ||A1/2(x0)ξ|λ1/2|ξ|.\displaystyle\lambda^{1/2}|\xi|\leq|A^{1/2}(x_{0})\xi|\leq\Lambda^{1/2}|\xi|,\qquad\Lambda^{-1/2}|\xi|\leq|A^{-1/2}(x_{0})\xi|\leq\lambda^{-1/2}|\xi|. (4.1)

In particular, λ1/2A1/2(x0)ξΛ1/2\lambda^{1/2}\leq||A^{1/2}(x_{0})\xi||\leq\Lambda^{1/2} and Λ1/2A1/2(x0)λ1/2\Lambda^{-1/2}\leq||A^{-1/2}(x_{0})||\leq\lambda^{-1/2}.

Define the affine change of variable Tx0T_{x_{0}} at x0Ex_{0}\in\partial E by

Tx0(x)=A1/2(x0)(xx0)+x0,Tx01(y)=A1/2(x0)(yx0)+x0,\displaystyle T_{x_{0}}(x)=A^{-1/2}(x_{0})(x-x_{0})+x_{0},\qquad T_{x_{0}}^{-1}(y)=A^{1/2}(x_{0})(y-x_{0})+x_{0}, (4.2)

and define

Ex0=Tx0(E),Ux0=Tx0(U),Ax0(y)=A1/2(x0)A(Tx01(y))A1/2(x0).\displaystyle E_{x_{0}}=T_{x_{0}}(E),\qquad U_{x_{0}}=T_{x_{0}}(U),\qquad A_{x_{0}}(y)=A^{-1/2}(x_{0})A(T_{x_{0}}^{-1}(y))A^{-1/2}(x_{0}). (4.3)

Note that Tx0(x0)=x0T_{x_{0}}(x_{0})=x_{0}, Ax0(x0)=IA_{x_{0}}(x_{0})=I, while Ax0A_{x_{0}} is symmetric, uniformly elliptic with constants 0<λ/ΛΛ/λ<+0<\lambda/\Lambda\leq\Lambda/\lambda<+\infty and Hölder continuous with exponent α\alpha and Hölder seminorm Ax0Cα(Λα/2/λ)ACα||A_{x_{0}}||_{C^{\alpha}}\leq(\Lambda^{\alpha/2}/\lambda)\>||A||_{C^{\alpha}}. The uniform ellipticity constants follow from

(λ/Λ)|ξ|2λ|A1/2(x0)ξ|2A(Tx01(y))A1/2(x0)ξ,A1/2(x0)ξΛ|A1/2(x0)ξ|2(Λ/λ)|ξ|2\displaystyle(\lambda/\Lambda)|\xi|^{2}\leq\lambda|A^{-1/2}(x_{0})\xi|^{2}\leq\langle A(T_{x_{0}}^{-1}(y))A^{-1/2}(x_{0})\xi,A^{-1/2}(x_{0})\xi\rangle\leq\Lambda|A^{-1/2}(x_{0})\xi|^{2}\leq(\Lambda/\lambda)|\xi|^{2} (4.4)

and the bound on the Hölder norm follows from estimate that for all x,ynx,y\in\mathbb{R}^{n} there holds

Ax0(x)Ax0(y)\displaystyle||A_{x_{0}}(x)-A_{x_{0}}(y)|| =||A1/2(x0)[A(Tx01(x))A(Tx01(y))]A1/2(x0))||\displaystyle=||A^{-1/2}(x_{0})\big{[}A(T_{x_{0}}^{-1}(x))-A(T_{x_{0}}^{-1}(y))\big{]}A^{-1/2}(x_{0}))||
λ1/2A(Tx01(x))A(Tx01(y))λ1/2\displaystyle\leq\lambda^{-1/2}||A(T_{x_{0}}^{-1}(x))-A(T_{x_{0}}^{-1}(y))||\lambda^{-1/2}
λ1ACα|Tx01(x)Tx01(y)|α\displaystyle\leq\lambda^{-1}||A||_{C^{\alpha}}|T_{x_{0}}^{-1}(x)-T_{x_{0}}^{-1}(y)|^{\alpha}
λ1ACαΛα/2|xy|α.\displaystyle\leq\lambda^{-1}||A||_{C^{\alpha}}\Lambda^{\alpha/2}|x-y|^{\alpha}. (4.5)

Thus constants for Ax0A_{x_{0}} depend on the same universal constants as AA.

The ellipsoid at x0nx_{0}\in\mathbb{R}^{n} of radius r>0r>0 is defined by

Wx0(x0,r)=Tx01(B(x0,r)).\displaystyle\textbf{{W}}_{x_{0}}(x_{0},r)=T_{x_{0}}^{-1}(\textbf{{B}}(x_{0},r)). (4.6)

We use Wx0\textbf{{W}}_{x_{0}} for our notation as this is the Wulff shape, introduced in [Wul01], for the integrand f(x0,ξ)=A(x0)ξ,ξ)1/2f(x_{0},\xi)=\langle A(x_{0})\xi,\xi)^{1/2}. The ellipsoid Wx0(x0,r)\textbf{{W}}_{x_{0}}(x_{0},r) has axial directions corresponding to the eigenvectors of A1/2(x0)A^{1/2}(x_{0}) and axial lengths corresponding to the eigenvalues scaled by a factor of rr. Since the eigenvalues of A1/2(x0)A^{1/2}(x_{0}) are bounded between λ1/2\lambda^{1/2} and Λ1/2\Lambda^{1/2}, we have

B(x0,λ1/2r)Wx0(x0,r)B(x0,Λ1/2r).\displaystyle\textbf{{B}}(x_{0},\lambda^{1/2}r)\subset\textbf{{W}}_{x_{0}}(x_{0},r)\subset\textbf{{B}}(x_{0},\Lambda^{1/2}r). (4.7)

We now prove the invariance of almost-minimizers under the change of variable Tx0T_{x_{0}} and refer readers to the change of variable formula given in Proposition A.1 in Appendix A.

Proposition 4.1 (Invariance of almost-minimizers under the change of variable Tx0T_{x_{0}}).

If EE is a (κ,α)(\kappa,\alpha)-almost-minimizer of A\mathscr{F}_{A} in UU at scale r0r_{0}, then Ex0E_{x_{0}} is a (Λ(α+n1)/2λn/2κ,α)(\Lambda^{(\alpha+n-1)/2}\lambda^{-n/2}\kappa,\alpha)-almost-minimizer of Ax0\mathscr{F}_{A_{x_{0}}} in Ux0U_{x_{0}} at scale r0/Λ1/2r_{0}/\Lambda^{1/2}.

Proof.

Suppose Ex0ΔFx0B(z,r)Ux0E_{x_{0}}\Delta F_{x_{0}}\subset\subset\textbf{{B}}(z,r)\cap U_{x_{0}} for some zUx0z\in U_{x_{0}} and r<r0/Λ1/2r<r_{0}/\Lambda^{1/2} (here we write Fx0F_{x_{0}} as an arbitrary competitor for Ex0E_{x_{0}} whose image FF under Tx01T_{x_{0}}^{-1} will be a competitor for EE). Applying Proposition A.1 with y=Tx0(x)y=T_{x_{0}}(x), noting that Jf=JTx0=detA1/2(x0)Jf=JT_{x_{0}}=\det A^{-1/2}(x_{0}) and (gf)t=A1/2(x0)(\nabla g\circ f)^{t}=A^{1/2}(x_{0}) since A1/2(x0)A^{1/2}(x_{0}) is symmetric, we have

Ax0(Ex0;B(z,r))\displaystyle\mathscr{F}_{A_{x_{0}}}(E_{x_{0}};\textbf{{B}}(z,r)) =B(z,r)Ex0Ax0(y)νEx0,νEx01/2dn1(y)\displaystyle=\int_{\textbf{{B}}(z,r)\cap\partial^{*}\!E_{x_{0}}}\langle A_{x_{0}}(y)\nu_{E_{x_{0}}},\nu_{E_{x_{0}}}\rangle^{1/2}\operatorname{\,d\mathcal{H}^{n-1}}(y)
=Tx01(B(z,r))EAx0(Tx0(x))A1/2(x0)νE,A1/2(x0)νE1/2detA1/2(x0)dn1(x)\displaystyle=\int_{T_{x_{0}}^{-1}(\textbf{{B}}(z,r))\cap\partial^{*}\!E}\langle A_{x_{0}}(T_{x_{0}}(x))A^{1/2}(x_{0})\nu_{E},A^{1/2}(x_{0})\nu_{E}\rangle^{1/2}\det A^{-1/2}(x_{0})\operatorname{\,d\mathcal{H}^{n-1}}(x)
=Tx01(B(z,r))EA1/2(x0)Ax0(Tx0(x))A1/2(x0)νE,νE1/2detA1/2(x0)dn1(x).\displaystyle=\int_{T_{x_{0}}^{-1}(\textbf{{B}}(z,r))\cap\partial^{*}\!E}\langle A^{1/2}(x_{0})A_{x_{0}}(T_{x_{0}}(x))A^{1/2}(x_{0})\nu_{E},\nu_{E}\rangle^{1/2}\det A^{-1/2}(x_{0})\operatorname{\,d\mathcal{H}^{n-1}}(x). (4.8)

Note that Ax0(Tx0(x))=A1/2(x0)A(x)A1/2(x0)A_{x_{0}}(T_{x_{0}}(x))=A^{-1/2}(x_{0})A(x)A^{-1/2}(x_{0}) and so A1/2(x0)Ax0(Tx0(x))A1/2(x0)=A(x)A^{1/2}(x_{0})A_{x_{0}}(T_{x_{0}}(x))A^{1/2}(x_{0})=A(x). Hence

Ax0(Ex0;B(z,r))=detA1/2(x0)A(E;Tx01(B(z,r))).\displaystyle\mathscr{F}_{A_{x_{0}}}(E_{x_{0}};\textbf{{B}}(z,r))=\det A^{-1/2}(x_{0})\mathscr{F}_{A}(E;T_{x_{0}}^{-1}(\textbf{{B}}(z,r))). (4.9)

Likewise, Ax0(Fx0;B(z,r))=detA1/2(x0)A(F;Tx01(B(z,r)))\mathscr{F}_{A_{x_{0}}}(F_{x_{0}};\textbf{{B}}(z,r))=\det A^{-1/2}(x_{0})\mathscr{F}_{A}(F;T_{x_{0}}^{-1}(\textbf{{B}}(z,r))). Note that

EΔFTx01(B(z,r))UB(Tx01(z),Λ1/2r)U,Tx01(z)U, and Λ1/2r<r0.\displaystyle E\Delta F\subset\subset T_{x_{0}}^{-1}(\textbf{{B}}(z,r))\cap U\subset\textbf{{B}}(T_{x_{0}}^{-1}(z),\Lambda^{1/2}r)\cap U,\ T_{x_{0}}^{-1}(z)\in U,\text{ and }\Lambda^{1/2}r<r_{0}. (4.10)

Thus A(E;B(Tx01(z),Λ1/2r))A(E;B(Tx01(z),Λ1/2r))+Λ(α+n1)/2κrα+n1\mathscr{F}_{A}(E;\textbf{{B}}(T_{x_{0}}^{-1}(z),\Lambda^{1/2}r))\leq\mathscr{F}_{A}(E;\textbf{{B}}(T_{x_{0}}^{-1}(z),\Lambda^{1/2}r))+\Lambda^{(\alpha+n-1)/2}\kappa\>r^{\alpha+n-1} by the minimality condition. This simplifies to A(E;Tx01(B(z,r))A(E;Tx01(B(z,r))+Λ(α+n1)/2κrα+n1\mathscr{F}_{A}(E;T_{x_{0}}^{-1}(\textbf{{B}}(z,r))\leq\mathscr{F}_{A}(E;T_{x_{0}}^{-1}(\textbf{{B}}(z,r))+\Lambda^{(\alpha+n-1)/2}\kappa\>r^{\alpha+n-1}. It then follows that

Ax0(Ex0;B(z,r))\displaystyle\mathscr{F}_{A_{x_{0}}}(E_{x_{0}};\textbf{{B}}(z,r)) =detA1/2(x0)A(E;Tx01(B(z,r)))\displaystyle=\det A^{-1/2}(x_{0})\mathscr{F}_{A}(E;T_{x_{0}}^{-1}(\textbf{{B}}(z,r)))
detA1/2(x0)A(F;Tx01(B(z,r)))+detA1/2(x0)Λ(α+n1)/2κrα+n1\displaystyle\leq\det A^{-1/2}(x_{0})\mathscr{F}_{A}(F;T_{x_{0}}^{-1}(\textbf{{B}}(z,r)))+\det A^{-1/2}(x_{0})\Lambda^{(\alpha+n-1)/2}\kappa\>r^{\alpha+n-1}
Ax0(Fx0;B(z,r))+Λ(α+n1)/2λn/2κrα+n1\displaystyle\leq\mathscr{F}_{A_{x_{0}}}(F_{x_{0}};\textbf{{B}}(z,r))+\Lambda^{(\alpha+n-1)/2}\lambda^{-n/2}\kappa\>r^{\alpha+n-1} (4.11)

as desired. ∎

Hence any of the properties or estimates we prove for (κ,α)(\kappa,\alpha)-almost-minimizers also hold for the set Ex0E_{x_{0}} (with any bounds or estimates having modified constants but which depend only on the same universal constants). Working with Ex0E_{x_{0}} will allow us to assume A(x0)=IA(x_{0})=I in the proof of many of our estimates and will in turn allow us to prove additional properties and estimates for general (κ,α)(\kappa,\alpha)-almost-minimizers.

As previously mentioned, whenever we write CC we mean a constant (which may change from line to line) that depends only on the universal constants n,λ,Λ,κ,α,r0n,\lambda,\Lambda,\kappa,\alpha,r_{0}, and upper bounds for ACα||A||_{C^{\alpha}}, but does not depend on the set EE or the point x0x_{0}. In cases where we wish to emphasize that a constant depends on fewer constants such as, for example, on the dimension nn only, we write C(n)C(n).

4.2. Scaling of the energy A\mathscr{F}_{A}

In Section 7 we will use the scaling of the energy A\mathscr{F}_{A} to simplify and work at scale 11 instead of of scale rr and in Section 10 we will utilize blow-up analysis to study the singular set almost-minimizers. The blow-ups Ex0,rE_{x_{0},r} of a set EE at a point x0nx_{0}\in\mathbb{R}^{n} and scale r>0r>0 are defined by

Ex0,r=Ex0r=Φx0,r(E)\displaystyle E_{x_{0},r}=\frac{E-x_{0}}{r}=\Phi_{x_{0},r}(E) (4.12)

where Φx0,r:nn\Phi_{x_{0},r}\mathrel{\mathop{\mathchar 58\relax}}\mathbb{R}^{n}\to\mathbb{R}^{n} is the map defined by

Φx0,r(x)=xx0r.\displaystyle\Phi_{x_{0},r}(x)=\frac{x-x_{0}}{r}. (4.13)

We denote the inverse of Φx0,r\Phi_{x_{0},r} by Ψx0,r\Psi_{x_{0},r}, that is, Ψx0,r(y)=ry+x0\Psi_{x_{0},r}(y)=ry+x_{0}. Given a matrix-valued function A=(aij(x))i,j=1nA=(a_{ij}(x))_{i,j=1}^{n}, we denote by Ax0,rA_{x_{0},r} the matrix-valued function

Ax0,r(y)=A(ry+x0)=AΨx0,r(y)\displaystyle A_{x_{0},r}(y)=A(ry+x_{0})=A\circ\Psi_{x_{0},r}(y) (4.14)

(this is not to be confused with Ax0A_{x_{0}} from the previous subsection). Note that Ax0,rCα=rαACα||A_{x_{0},r}||_{C^{\alpha}}=r^{\alpha}||A||_{C^{\alpha}}.

Proposition 4.2 (Scaling of A\mathscr{F}_{A}).

If EE is a set of locally finite perimeter in n\mathbb{R}^{n}, x0nx_{0}\in\mathbb{R}^{n}, r>0r>0, then

Ax0,r(Ex,r;Fx0,r)=A(E;F)rn1.\displaystyle\mathscr{F}_{A_{x_{0},r}}(E_{x,r};F_{{x_{0}},r})=\frac{\mathscr{F}_{A}(E;F)}{r^{n-1}}. (4.15)

for Borel sets FF. In particular, if EE is a (κ,α)(\kappa,\alpha)-almost-minimizer of A\mathscr{F}_{A} in UU at scale r0r_{0}, then Ex0,rE_{{x_{0}},r} is a (κrα,α)(\kappa r^{\alpha},\alpha)-almost-minimizer of Ax0,r\mathscr{F}_{A_{{x_{0}},r}} in Ux0,rU_{{x_{0}},r} at scale r0/rr_{0}/r.

Proof.

We apply Proposition A.1 with the change of variable y=f(x)=Φx0,r(x)=(xx0)/ry=f(x)=\Phi_{{x_{0}},r}(x)=(x-{x_{0}})/r and integrand (x,ξ)Ax0,r(x)ξ,ξ1/2(x,\xi)\mapsto\langle A_{{x_{0}},r}(x)\xi,\xi\rangle^{1/2}. Then g(y)=Ψx0,r(y)=ry+x0g(y)=\Psi_{{x_{0}},r}(y)=ry+{x_{0}}, g=rI\nabla g=rI, and Jf=rnJf=r^{-n}. So |(gf)tνE|=r|(\nabla g\circ f)^{t}\nu_{E}|=r and it follows that

Ax0,r(Ex0,r;Fx0,r)\displaystyle\mathscr{F}_{A_{{x_{0}},r}}(E_{{x_{0}},r};F_{{x_{0}},r}) =Fx0,rEx0,rA(ry+x0)νEx0,r,νEx0,r1/2dn1(y)\displaystyle=\int_{F_{{x_{0}},r}\cap\partial^{*}\!E_{{x_{0}},r}}\langle A(ry+{x_{0}})\nu_{E_{{x_{0}},r}},\nu_{E_{{x_{0}},r}}\rangle^{1/2}\operatorname{\,d\mathcal{H}^{n-1}}(y)
=FEA(x)νE,νE1/2rnrdn1(x)\displaystyle=\int_{F\cap\partial^{*}\!E}\langle A(x)\nu_{E},\nu_{E}\rangle^{1/2}\ r^{-n}r\operatorname{\,d\mathcal{H}^{n-1}}(x)
=A(E;F)rn1\displaystyle=\frac{\mathscr{F}_{A}(E;F)}{r^{n-1}} (4.16)

Now, let FF be a set of locally finite perimeter in n\mathbb{R}^{n} with Ex0,rΔFx0,rB(x,s)Ux0,rE_{{x_{0}},r}\Delta F_{{x_{0}},r}\subset\subset\textbf{{B}}(x,s)\cap U_{{x_{0}},r} for xUx0,rx\in U_{{x_{0}},r} and s<r0/rs<r_{0}/r. Then EΔFΨx0,r(B(x,s))UE\Delta F\subset\subset\Psi_{{x_{0}},r}(\textbf{{B}}(x,s))\cap U. Note Ψx0,r(B(x,s))=B(rx+x0,rs)\Psi_{{x_{0}},r}(\textbf{{B}}(x,s))=\textbf{{B}}(rx+{x_{0}},rs) with rx+x0Urx+{x_{0}}\in U and rs<r0rs<r_{0}. Applying (4.2) to B(x,s)\textbf{{B}}(x,s) and using the almost-minimality of EE in UU at scale r0r_{0}, we have

Ax0,r(Ex0,r;B(x,s))\displaystyle\mathscr{F}_{A_{{x_{0}},r}}(E_{{x_{0}},r};\textbf{{B}}(x,s)) =A(E;B(rx+x0,rs))rn1\displaystyle=\frac{\mathscr{F}_{A}(E;\textbf{{B}}(rx+{x_{0}},rs))}{r^{n-1}}
A(F;B(rx+x0,rs))+κ(rs)α+n1rn1\displaystyle\leq\frac{\mathscr{F}_{A}(F;\textbf{{B}}(rx+{x_{0}},rs))+\kappa(rs)^{\alpha+n-1}}{r^{n-1}}
=Ax0,r(Fx0,r;B(x,s))+κrαsα+n1,\displaystyle=\mathscr{F}_{A_{{x_{0}},r}}(F_{{x_{0}},r};\textbf{{B}}(x,s))+\kappa r^{\alpha}s^{\alpha+n-1}, (4.17)

that is, Ex0,rE_{{x_{0}},r} is an (κrα,α)(\kappa r^{\alpha},\alpha)-almost-minimizer of Ax0,r\mathscr{F}_{A_{{x_{0}},r}} in Ux0,rU_{{x_{0}},r} at scale r0/r>0r_{0}/r>0. ∎

4.3. Comparison sets

To utilize the almost-minimality condition we will often construct competitors by modifying EE inside an open set. The following proposition allows us to do this.

Proposition 4.3 (Comparison sets by replacements).

If EE and FF are sets of locally finite perimeter in n\mathbb{R}^{n} and GG is an open set of finite perimeter in n\mathbb{R}^{n} such that

n1(GE)=n1(GF)=0,\displaystyle\operatorname{\mathcal{H}^{n-1}}(\partial^{*}\!G\cap\partial^{*}\!E)=\operatorname{\mathcal{H}^{n-1}}(\partial^{*}\!G\cap\partial^{*}\!F)=0, (4.18)

then the set defined by

F0=(FG)(EG)\displaystyle F_{0}=\big{(}F\cap G\big{)}\cup\big{(}E\setminus G\big{)} (4.19)

is a set of locally finite perimeter in n\mathbb{R}^{n}. Moreover, if GUG\subset\subset U and UU is open, then

A(F0;U)=A(F;G)+A(E;UG¯)+A(G;E(1)ΔF(1)).\displaystyle\mathscr{F}_{A}(F_{0};U)=\mathscr{F}_{A}(F;G)+\mathscr{F}_{A}(E;U\setminus\overline{G})+\mathscr{F}_{A}(G;E^{(1)}\Delta F^{(1)}). (4.20)
Proof.

In the proof of [Mag12, Theorem 16.16] the decomposition, see (16.35),

μF0=μF G+μG(F(1)E(0))+μE (nG¯)μG (E(1)F(0))\displaystyle\mu_{F_{0}}=\mu_{F}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>G+\mu_{G}(F^{(1)}\cap E^{(0)})+\mu_{E}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>(\mathbb{R}^{n}\setminus\overline{G})-\mu_{G}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>(E^{(1)}\cap F^{(0)}) (4.21)

is proved. Since all of the measures on the right-hand side are concentrated on disjoint sets and since the measures A(Gc;)\mathscr{F}_{A}(G^{c};\>\cdot\>) and A(G;)\mathscr{F}_{A}(G;\>\cdot\>) are equal and μGc=μG\mu_{G^{c}}=-\mu_{G}, we have

A(F0;U)=A(F;G)+A(E;UG¯)+A(G;(F(1)E(0))(E(1)F(0)))\displaystyle\mathscr{F}_{A}(F_{0};U)=\mathscr{F}_{A}(F;G)+\mathscr{F}_{A}(E;U\setminus\overline{G})+\mathscr{F}_{A}(G;(F^{(1)}\cap E^{(0)})\cup(E^{(1)}\cap F^{(0)})) (4.22)

by additivity of A\mathscr{F}_{A}. By n1(GE)=0\operatorname{\mathcal{H}^{n-1}}(\partial^{*}\!G\cap\partial^{*}\!E)=0 and n1(n(E(0)E(1)E))=0\operatorname{\mathcal{H}^{n-1}}(\mathbb{R}^{n}\setminus(E^{(0)}\cup E^{(1)}\cup\partial^{*}\!E))=0,

A(G;F(1)E(0))=A(G;F(1)E(1)).\displaystyle\mathscr{F}_{A}(G;F^{(1)}\cap E^{(0)})=\mathscr{F}_{A}(G;F^{(1)}\setminus E^{(1)}). (4.23)

Likewise, n1(GF)=0\operatorname{\mathcal{H}^{n-1}}(\partial^{*}\!G\cap\partial^{*}\!F)=0 and n1(n(F(0)F(1)F))=0\operatorname{\mathcal{H}^{n-1}}(\mathbb{R}^{n}\setminus(F^{(0)}\cup F^{(1)}\cup\partial^{*}\!F))=0 and so

A(G;E(1)F(0))=A(G;E(1)F(1)).\displaystyle\mathscr{F}_{A}(G;E^{(1)}\cap F^{(0)})=\mathscr{F}_{A}(G;E^{(1)}\setminus F^{(1)}). (4.24)

These along with (4.22) prove (4.20). ∎

4.4. Volume and perimeter bounds and the almost-monotonicity formula

One important property which almost-minimizers of A\mathscr{F}_{A} possess is bounds on both the volume and the perimeter of EE on balls centered at points in their topological boundary. Recall that we require sptμE=E\operatorname{spt}\mu_{E}=\partial E for almost-minimizers. The full set of estimates is given in Proposition 4.10 but we have some work to do to prove this. The first step is showing the upper bound on perimeter.

Define the perimeter density ratio of EE at x0x_{0} by

θ(E,x0,r)=P(E;B(x0,r))rn1.\displaystyle\theta(E,x_{0},r)=\frac{P(E;\textbf{{B}}(x_{0},r))}{r^{n-1}}. (4.25)

and perimeter density of EE at x0x_{0} by

θ(E,x0)=limr0+θ(E,x0,r)\displaystyle\theta(E,x_{0})=\lim_{r\to 0^{+}}\theta(E,x_{0},r) (4.26)

whenever the limit exists.

Lemma 4.4 (Upper perimeter bound).

There exists a positive constant C=C(n,λ,Λ,κ,α,r0)C=C(n,\lambda,\Lambda,\kappa,\alpha,r_{0}) with the following property. If EE is a (κ,α)(\kappa,\alpha)-almost-minimizer of A\mathscr{F}_{A} in UU at scale r0r_{0}, then for every x0UEx_{0}\in U\cap\partial E with r<d=min{dist(x0,U),r0}<r<d=\min\{\mathrm{dist}(x_{0},\partial U),r_{0}\}<\infty,

P(E;B(x0,r))rn1C.\displaystyle\frac{P(E;\textbf{{B}}(x_{0},r))}{r^{n-1}}\leq C. (4.27)
Proof.

Consider the function m:(0,d)m\colon(0,d)\to\mathbb{R} defined by m(r)=|EB(x0,r)|m(r)=|E\cap\textbf{{B}}(x_{0},r)|. Note that mm is increasing, m(r)=n1(E(1)B(x0,r))m^{\prime}(r)=\operatorname{\mathcal{H}^{n-1}}(E^{(1)}\cap\partial\textbf{{B}}(x_{0},r)) for a.e. rr by the coarea formula, and n1(EB(x0,r))=0\operatorname{\mathcal{H}^{n-1}}(\partial^{*}\!E\cap\partial\textbf{{B}}(x_{0},r))=0 for a.e. rr because n1  E\operatorname{\mathcal{H}^{n-1}}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>\>\partial^{*}\!E is a Radon measure. Let r(0,d)r\in(0,d) be one of the a.e. radii that satisfies both m(r)=n1(E(1)B(x0,r))m^{\prime}(r)=\operatorname{\mathcal{H}^{n-1}}(E^{(1)}\cap\partial\textbf{{B}}(x_{0},r)) and n1(EB(x0,r))=0\operatorname{\mathcal{H}^{n-1}}(\partial^{*}\!E\cap\partial\textbf{{B}}(x_{0},r))=0. For s(r,d)s\in(r,d) consider the comparison set F=EB(x0,r)F=E\setminus\textbf{{B}}(x_{0},r) in B(x0,s)\textbf{{B}}(x_{0},s). Then EΔFB(x0,r)B(x0,s)E\Delta F\subset\textbf{{B}}(x_{0},r)\subset\subset\textbf{{B}}(x_{0},s). It follows from comparability to perimeter and the almost-minimality that

λ1/2P(E;B(x0,s))\displaystyle\lambda^{1/2}P(E;\textbf{{B}}(x_{0},s)) A(E;B(x0,s))\displaystyle\leq\mathscr{F}_{A}(E;\textbf{{B}}(x_{0},s))
A(EB(x0,r);B(x0,s))+κsα+n1\displaystyle\leq\mathscr{F}_{A}(E\setminus\textbf{{B}}(x_{0},r);\textbf{{B}}(x_{0},s))+\kappa s^{\alpha+n-1}
Λ1/2P(EB(x0,r);B(x0,s))+κsα+n1\displaystyle\leq\Lambda^{1/2}P(E\setminus\textbf{{B}}(x_{0},r);\textbf{{B}}(x_{0},s))+\kappa s^{\alpha+n-1} (4.28)

and so

P(E;B(x0,s))\displaystyle P(E;\textbf{{B}}(x_{0},s)) C(P(EB(x0,r);B(x0,s))+sα+n1)\displaystyle\leq C\big{(}P(E\setminus\textbf{{B}}(x_{0},r);\textbf{{B}}(x_{0},s))+s^{\alpha+n-1}\big{)}
=C(n1(E(1)B(x0,r))+P(E;B(x0,s)B¯(x0,r))+sα+n1)\displaystyle=C\big{(}\operatorname{\mathcal{H}^{n-1}}(E^{(1)}\cap\partial\textbf{{B}}(x_{0},r))+P(E;\textbf{{B}}(x_{0},s)\setminus\overline{\textbf{{B}}}(x_{0},r))+s^{\alpha+n-1}\big{)} (4.29)

since n1(EB(x0,r))=0\operatorname{\mathcal{H}^{n-1}}(\partial^{*}\!E\cap\partial\textbf{{B}}(x_{0},r))=0. Sending sr+s\to r^{+} yields the inequality

P(E;B(x0,r))C(n1(E(1)B(x0,r))+rα+n1).\displaystyle P(E;\textbf{{B}}(x_{0},r))\leq C\big{(}\operatorname{\mathcal{H}^{n-1}}(E^{(1)}\cap\partial\textbf{{B}}(x_{0},r))+r^{\alpha+n-1}\big{)}. (4.30)

This, together with n1(E(1)B(x0,r))nωnrn1\operatorname{\mathcal{H}^{n-1}}(E^{(1)}\cap\partial\textbf{{B}}(x_{0},r))\leq n\omega_{n}r^{n-1} and r<r0r<r_{0}, gives P(E;B(x0,r))Crn1P(E;\textbf{{B}}(x_{0},r))\leq Cr^{n-1}. ∎

To obtain the lower perimeter bound for almost-minimizers of A\mathscr{F}_{A}, we shall adapt an argument given by Tamanini for almost-minimizers of perimeter in [Tam82, Tam84] which makes use of an almost-monotonicity formula. Monotonicity formulas are often times a valuable tool in regularity theory. For example, the monotonicity of density ratios for minimizers of surface area is heavily relied upon in [All72, Tay76a] as well as in many other papers. By this we mean the fact that if EE is a perimeter minimizer in UU, x0Ux_{0}\in U, then the density ratio

θ(E;x0,r)=P(E;B(x0,r))rn1.\displaystyle\theta(E;x_{0},r)=\frac{P(E;\textbf{{B}}(x_{0},r))}{r^{n-1}}. (4.31)

is monotonically increasing in rr (see, for example, [Mag12, Theorem 17.16]). In [All74], Allard demonstrated for integrands depending solely on the direction variable νE\nu_{E} (and not on the spatial variable xx) that monotonicity formulas exist if and only if the integrand is a linear change of variable from the area integrand. Under the change of variable Tx0T_{x_{0}} we have A(x0)=IA(x_{0})=I so that our sets satisfy the condition for almost-minimality of perimeter when making comparisons on balls centered at x0x_{0} as shown in Lemma 4.5 below. A key observation is that we only need these comparisons to apply the standard cone-competitor argument to obtain an almost-monotonicity formula as we do in Lemma 4.6.

Lemma 4.5.

There exists a positive constant C=C(n,λ,Λ,κ,α,r0)C=C(n,\lambda,\Lambda,\kappa,\alpha,r_{0}) with the following property. If EE is a (κ,α)(\kappa,\alpha)-almost-minimizer of A\mathscr{F}_{A} in UU at scale r0>0r_{0}>0, x0UEx_{0}\in U\cap\partial E, and A(x0)=IA(x_{0})=I, then

P(E;B(x0,r))P(F;B(x0,r))+C(κ+ACα)rα+n1\displaystyle P(E;\textbf{{B}}(x_{0},r))\leq P(F;\textbf{{B}}(x_{0},r))+C(\kappa+||A||_{C^{\alpha}})r^{\alpha+n-1} (4.32)

whenever EΔFB(x0,r)E\Delta F\subset\subset\textbf{{B}}(x_{0},r) and r<d=min{r0,dist(x0,U)}r<d=\min\{r_{0},\mathrm{dist}(x_{0},\partial U)\}.

Proof.

Let EΔFB(x0,r)UE\Delta F\subset\subset\textbf{{B}}(x_{0},r)\subset U and r<r0r<r_{0}. If P(E;B(x0,r))P(F;B(x0,r))P(E;\textbf{{B}}(x_{0},r))\leq P(F;\textbf{{B}}(x_{0},r)), then (4.32) trivially holds true. So consider the case when P(F;B(x0,r))P(E;B(x0,r))P(F;\textbf{{B}}(x_{0},r))\leq P(E;\textbf{{B}}(x_{0},r)).

By inequality (2.11) and A(x0)=IA(x_{0})=I, we have

|νE|A(x)νE,νE1/2+12λA(x0)A(x)A(x)νE,νE1/2+12λACα|xx0|α\displaystyle|\nu_{E}|\leq\langle A(x)\nu_{E},\nu_{E}\rangle^{1/2}+\frac{1}{2\lambda}||A(x_{0})-A(x)||\leq\langle A(x)\nu_{E},\nu_{E}\rangle^{1/2}+\frac{1}{2\lambda}||A||_{C^{\alpha}}|x-x_{0}|^{\alpha} (4.33)

and so |νE|A(x)νE,νE1/2+(1/2λ)ACαrα|\nu_{E}|\leq\langle A(x)\nu_{E},\nu_{E}\rangle^{1/2}+(1/2\lambda)||A||_{C^{\alpha}}r^{\alpha} for xB(x0,r)x\in\textbf{{B}}(x_{0},r). Integrating with respect to n1  E\operatorname{\mathcal{H}^{n-1}}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>\>\partial^{*}\!E gives

P(E;B(x0,r))A(E;B(x0,r))+12λACαrαP(E;B(x0,r))\displaystyle P(E;\textbf{{B}}(x_{0},r))\leq\mathscr{F}_{A}(E;\textbf{{B}}(x_{0},r))+\frac{1}{2\lambda}||A||_{C^{\alpha}}r^{\alpha}P(E;\textbf{{B}}(x_{0},r)) (4.34)

Similarly, A(x)νF,νF1/2|νF|+(1/2λ)ACαrα\langle A(x)\nu_{F},\nu_{F}\rangle^{1/2}\leq|\nu_{F}|+(1/2\lambda)||A||_{C^{\alpha}}r^{\alpha} for xB(x0,r)x\in\textbf{{B}}(x_{0},r) and so

A(F;B(x0,r))P(F;B(x0,r))+12λACαrαP(F;B(x0,r))\displaystyle\mathscr{F}_{A}(F;\textbf{{B}}(x_{0},r))\leq P(F;\textbf{{B}}(x_{0},r))+\frac{1}{2\lambda}||A||_{C^{\alpha}}r^{\alpha}P(F;\textbf{{B}}(x_{0},r)) (4.35)

Combining the almost-minimizer inequality with (4.34), (4.35), and P(F;B(x0,r))P(E;B(x0,r))P(F;\textbf{{B}}(x_{0},r))\leq P(E;\textbf{{B}}(x_{0},r)) gives P(E;B(x0,r))P(F;B(x0,r))+κrα+n1+(1/2λ)ACαrαP(E;B(x0,r))P(E;\textbf{{B}}(x_{0},r))\leq P(F;\textbf{{B}}(x_{0},r))+\kappa r^{\alpha+n-1}+(1/2\lambda)||A||_{C^{\alpha}}r^{\alpha}P(E;\textbf{{B}}(x_{0},r)). The upper perimeter bound P(E;B(x0,r))Crn1P(E;\textbf{{B}}(x_{0},r))\leq Cr^{n-1} gives P(E;B(x0,r))P(F;B(x0,r))+C(κ+ACα)rα+n1P(E;\textbf{{B}}(x_{0},r))\leq P(F;\textbf{{B}}(x_{0},r))+C(\kappa+||A||_{C^{\alpha}})r^{\alpha+n-1}. ∎

Lemma 4.6.

There exists a positive constant C=C(n,λ,Λ,κ,α,r0,ACα)C=C(n,\lambda,\Lambda,\kappa,\alpha,r_{0},||A||_{C^{\alpha}}) such that the following holds. If EE is a (κ,α)(\kappa,\alpha)-almost-minimizer of A\mathscr{F}_{A} in UU at scale r0>0r_{0}>0 with x0UEx_{0}\in U\cap\partial E and A(x0)=IA(x_{0})=I, then the function

rP(E;B(x0,r))rn1+Crα\displaystyle r\mapsto\frac{P(E;\textbf{{B}}(x_{0},r))}{r^{n-1}}+Cr^{\alpha} (4.36)

is monotonically increasing on (0,d)(0,d) where d=min{r0,dist(x0,U)}>0d=\min\{r_{0},\mathrm{dist}(x_{0},\partial U)\}>0.

Proof.

Without loss of generality assume x0=0x_{0}=0 and write Br=B(x0,r)\textbf{{B}}_{r}=\textbf{{B}}(x_{0},r). Define the function Φ:(0,d)(0,)\Phi\colon(0,d)\to(0,\infty) by Φ(r)=P(E;Br)\Phi(r)=P(E;\textbf{{B}}_{r}). Φ\Phi is increasing and hence differentiable for a.e. r(0,d)r\in(0,d). Thus it suffices to prove

ddr(Φ(r)rn1+Crα)0 for a.e. r(0,d),\displaystyle\frac{d}{dr}\Big{(}\frac{\Phi(r)}{r^{n-1}}+Cr^{\alpha}\Big{)}\geq 0\qquad\text{ for a.e. }r\in(0,d), (4.37)

which can be rewritten as

Φ(r)rn1Φ(r)+Crα+n1 for a.e. r(0,d).\displaystyle\Phi(r)\leq\frac{r}{n-1}\Phi^{\prime}(r)+Cr^{\alpha+n-1}\qquad\text{ for a.e. }r\in(0,d). (4.38)

The idea of the proof of (4.38) is to construct cone competitors over EBrE\cap\partial\textbf{{B}}_{r} with vertex at 0 for each r>0r>0 to use in the comparison inequality (4.32). To do this we will need to approximate EE by open sets with smooth boundary and construct the cone competitors for the approximating sets.

By [Mag12, Theorem 13.8], there is a sequence {Eh}h\{E_{h}\}_{h\in\mathbb{N}} of open sets with smooth boundary in n\mathbb{R}^{n} such that EhlocEE_{h}\overset{\text{loc}}{\rightarrow}E and |μEh||μE||\mu_{E_{h}}|\overset{\ast}{\rightharpoonup}|\mu_{E}|. For now hold hh\in\mathbb{N} fixed. The set EhBrE_{h}\cap\partial\textbf{{B}}_{r} is relatively open in Br\partial\textbf{{B}}_{r} for every r>0r>0. By Sard’s lemma,

EhBr is a smooth (n2)-dimensional surface for a.e. r>0.\displaystyle\partial E_{h}\cap\partial\textbf{{B}}_{r}\text{ is a smooth $(n-2)$-dimensional surface for a.e. }r>0. (4.39)

Consider the cones with vertex 0 over EhBrE_{h}\cap\partial\textbf{{B}}_{r},

Kh(r)={λxn:λ>0,xEhBr}.\displaystyle K_{h}(r)=\big{\{}\lambda x\in\mathbb{R}^{n}\mathrel{\mathop{\mathchar 58\relax}}\lambda>0,\>x\in E_{h}\cap\partial\textbf{{B}}_{r}\big{\}}. (4.40)

For the a.e. r>0r>0 such that (4.39) holds we have that Kh(r)K_{h}(r) is a set of locally finite perimeter in n\mathbb{R}^{n} with

μKh(r)=νKh(r)n1 Kh(r),andνKh(r)(x)x=0,xKh(r){0}.\displaystyle\mu_{K_{h}(r)}=\nu_{K_{h}(r)}\operatorname{\mathcal{H}^{n-1}}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>\>\partial K_{h}(r),\qquad\text{and}\qquad\nu_{K_{h}(r)}(x)\cdot x=0,\qquad\forall x\in K_{h}(r)\setminus\{0\}. (4.41)

For r>0r>0 such that (4.39) holds, the coarea formula for (n1)(n-1)-dimensional rectifiable sets (see [Mag12, Theorem 18.8]) on Kh(r)\partial K_{h}(r) with u(x)=|x|u(x)=|x| yields

P(Kh(r);Br)=0rn2(Kh(r)Bt)𝑑t\displaystyle P(K_{h}(r);\textbf{{B}}_{r})=\int_{0}^{r}\mathcal{H}^{n-2}(\partial K_{h}(r)\cap\partial\textbf{{B}}_{t})\>dt (4.42)

since |Kh(r)u|=|u|=1|\nabla^{\partial K_{h}(r)}u|=|\nabla u|=1. Note that for t<rt<r we have

Kh(r)Bt=(tr)(Kh(r)Br)=(tr)(EhBr)\displaystyle\partial K_{h}(r)\cap\partial\textbf{{B}}_{t}=\Big{(}\frac{t}{r}\Big{)}\big{(}\partial K_{h}(r)\cap\partial\textbf{{B}}_{r}\big{)}=\Big{(}\frac{t}{r}\Big{)}\big{(}\partial E_{h}\cap\partial\textbf{{B}}_{r}\big{)} (4.43)

and hence for rr such that (4.39) holds we have

P(Kh(r);Br)=0r(tr)n2n2(EhBr)𝑑t=rn1n2(EhBr)\displaystyle P(K_{h}(r);\textbf{{B}}_{r})=\int_{0}^{r}\bigg{(}\frac{t}{r}\bigg{)}^{n-2}\mathcal{H}^{n-2}(\partial E_{h}\cap\partial\textbf{{B}}_{r})\>dt=\frac{r}{n-1}\mathcal{H}^{n-2}(\partial E_{h}\cap\partial\textbf{{B}}_{r}) (4.44)

Consider a radius r>0r>0 such that for all hh\in\mathbb{N} both (4.39) and n1(EBr)=n1(EhBr)=0\operatorname{\mathcal{H}^{n-1}}(\partial^{*}\!E\cap\partial\textbf{{B}}_{r})=\operatorname{\mathcal{H}^{n-1}}(\partial E_{h}\cap\partial\textbf{{B}}_{r})=0 hold (and consequently (4.44) as well). This true for a.e. r(0,d)r\in(0,d) since n1  E\operatorname{\mathcal{H}^{n-1}}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>\>\partial^{*}\!E and n1  Eh\operatorname{\mathcal{H}^{n-1}}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>\>\partial E_{h} are Radon measures and by Sard’s lemma. Consider the comparison sets Fh=(Kh(r)Br)(EBr)F_{h}=(K_{h}(r)\cap\textbf{{B}}_{r})\cup(E\setminus\textbf{{B}}_{r}). Let ss be such that r<s<dr<s<d. By (16.32) of [Mag12] we have

P(Fh;Bs)=P(Kh(r);Br)+P(E;BsB¯r)+n1((E(1)ΔKh(r))Br).\displaystyle P(F_{h};\textbf{{B}}_{s})=P(K_{h}(r);\textbf{{B}}_{r})+P(E;\textbf{{B}}_{s}\setminus\overline{\textbf{{B}}}_{r})+\operatorname{\mathcal{H}^{n-1}}\big{(}\big{(}E^{(1)}\Delta K_{h}(r))\cap\partial\textbf{{B}}_{r}\big{)}. (4.45)

Since EΔFhBrBsE\Delta F_{h}\subset\textbf{{B}}_{r}\subset\subset\textbf{{B}}_{s}, applying Lemma 4.5 gives

P(E;Bs)P(Kh(r);Br)+P(E;BsB¯r)+n1((E(1)ΔKh(r))Br)+Csα+n1\displaystyle P(E;\textbf{{B}}_{s})\leq P(K_{h}(r);\textbf{{B}}_{r})+P(E;\textbf{{B}}_{s}\setminus\overline{\textbf{{B}}}_{r})+\operatorname{\mathcal{H}^{n-1}}\big{(}\big{(}E^{(1)}\Delta K_{h}(r))\cap\partial\textbf{{B}}_{r}\big{)}+Cs^{\alpha+n-1} (4.46)

which by subtracting P(E;BsB¯r)P(E;\textbf{{B}}_{s}\setminus\overline{\textbf{{B}}}_{r}) from each side together with (4.44) simplifies to

P(E;Br)rn1n2(EhBr)+n1((E(1)ΔEh)Br)+Csα+n1.\displaystyle P(E;\textbf{{B}}_{r})\leq\frac{r}{n-1}\mathcal{H}^{n-2}(\partial E_{h}\cap\partial\textbf{{B}}_{r})+\operatorname{\mathcal{H}^{n-1}}\big{(}\big{(}E^{(1)}\Delta E_{h})\cap\partial\textbf{{B}}_{r}\big{)}+Cs^{\alpha+n-1}. (4.47)

Sending sr+s\to r^{+} gives

P(E;Br)rn1n2(EhBr)+n1((E(1)ΔEh)Br)+Crα+n1.\displaystyle P(E;\textbf{{B}}_{r})\leq\frac{r}{n-1}\mathcal{H}^{n-2}(\partial E_{h}\cap\partial\textbf{{B}}_{r})+\operatorname{\mathcal{H}^{n-1}}\big{(}\big{(}E^{(1)}\Delta E_{h})\cap\partial\textbf{{B}}_{r}\big{)}+Cr^{\alpha+n-1}. (4.48)

This inequality holds for a.e. r(0,d)r\in(0,d) and integrating over the interval (s,t)(0,d)(s,t)\subset(0,d) yields

stP(E;Br)𝑑r1n1strn2(EhBr)𝑑r+n1((E(1)ΔEh)Bd)+C(tα+nsα+n).\displaystyle\int_{s}^{t}P(E;\textbf{{B}}_{r})\>dr\leq\frac{1}{n-1}\int_{s}^{t}r\>\mathcal{H}^{n-2}(\partial E_{h}\cap\partial\textbf{{B}}_{r})\>dr+\operatorname{\mathcal{H}^{n-1}}\big{(}\big{(}E^{(1)}\Delta E_{h})\cap\textbf{{B}}_{d}\big{)}+C(t^{\alpha+n}-s^{\alpha+n}). (4.49)

Applying the coarea formula for (n1)(n-1)-dimensional rectifiable sets ([Mag12, Theorem 18.8]) on Eh\partial E_{h} with u(x)=|x|u(x)=|x|, and g=|x|g=|x|, gives

strn2(EhBr)𝑑r=Eh(BtB¯s)|x||Ehu|tP(Eh;B¯tBs)\displaystyle\int_{s}^{t}r\>\mathcal{H}^{n-2}(\partial E_{h}\cap\partial\textbf{{B}}_{r})\>dr=\int_{\partial E_{h}\cap(\textbf{{B}}_{t}\setminus\overline{\textbf{{B}}}_{s})}|x||\nabla^{\partial E_{h}}u|\leq t\>P(E_{h};\overline{\textbf{{B}}}_{t}\setminus\textbf{{B}}_{s}) (4.50)

since |Ehu||u|=1|\nabla^{\partial E_{h}}u|\leq|\nabla u|=1. Thus combining (4.49) and (4.50) gives

stP(E;Br)𝑑rtn1P(Eh;B¯tBs)+n1((E(1)ΔEh)Bd)+C(tα+nsα+n).\displaystyle\int_{s}^{t}P(E;\textbf{{B}}_{r})\>dr\leq\frac{t}{n-1}P(E_{h};\overline{\textbf{{B}}}_{t}\setminus\textbf{{B}}_{s})+\operatorname{\mathcal{H}^{n-1}}\big{(}\big{(}E^{(1)}\Delta E_{h})\cap\textbf{{B}}_{d}\big{)}+C(t^{\alpha+n}-s^{\alpha+n}). (4.51)

By EhlocEE_{h}\overset{\text{loc}}{\rightarrow}E and |μEh||μE||\mu_{E_{h}}|\overset{\ast}{\rightharpoonup}|\mu_{E}|, sending hh\to\infty gives

stP(E;Br)𝑑rtn1P(E;B¯tBs)+C(tα+nsα+n).\displaystyle\int_{s}^{t}P(E;\textbf{{B}}_{r})\>dr\leq\frac{t}{n-1}P(E;\overline{\textbf{{B}}}_{t}\setminus\textbf{{B}}_{s})+C(t^{\alpha+n}-s^{\alpha+n}). (4.52)

Dividing by tst-s and sending ts+t\to s^{+} at points of differentiability of Φ\Phi yields (4.38) as desired. ∎

Now we are able to use a perturbation argument and the change of variable Tx0T_{x_{0}} to obtain an almost-monotonicity formula when A(x0)A(x_{0}) is not assumed to equal II.

Theorem 4.7 (Almost-monotonicity formula).

There exists a positive constant
C=C(n,λ,Λ,κ,α,r0,ACα)C=C(n,\lambda,\Lambda,\kappa,\alpha,r_{0},||A||_{C^{\alpha}}) with the following property. If EE is a (κ,α)(\kappa,\alpha)-almost-minimizer of A\mathscr{F}_{A} in UU at scale r0r_{0}, then for every x0UEx_{0}\in U\cap\partial E, we have

A(E;Wx0(x0,s))sn1A(E;Wx0(x0,r))rn1+Crα\displaystyle\frac{\mathscr{F}_{A}(E;\textbf{{W}}_{x_{0}}(x_{0},s))}{s^{n-1}}\leq\frac{\mathscr{F}_{A}(E;\textbf{{W}}_{x_{0}}(x_{0},r))}{r^{n-1}}+Cr^{\alpha} (4.53)

whenever 0<sr<d0<s\leq r<d where d=Λ1/2min{r0,dist(x0,U)}d=\Lambda^{-1/2}\min\{r_{0},\mathrm{dist}(x_{0},\partial U)\}.

Proof.

Applying Lemma 4.6 to Ex0E_{x_{0}} and Ax0\mathscr{F}_{A_{x_{0}}} gives that rr(n1)P(Ex0;B(x0,r))+Crαr\mapsto r^{-(n-1)}P(E_{x_{0}};\textbf{{B}}(x_{0},r))+Cr^{\alpha} is monotone increasing on (0,dx0)(0,d_{x_{0}}) where dx0=min{Λ1/2r0,dist(x0,Ux0)}d_{x_{0}}=\min\{\Lambda^{-1/2}r_{0},\mathrm{dist}(x_{0},\partial U_{x_{0}})\}. The change of variable y=Tx0(x)y=T_{x_{0}}(x) applied to EE gives P(Ex0;B(x0,r))=detA1/2(x0)A(x0)(E;Wx0(x0,r))P(E_{x_{0}};\textbf{{B}}(x_{0},r))=\det A^{-1/2}(x_{0})\mathscr{F}_{A(x_{0})}(E;\textbf{{W}}_{x_{0}}(x_{0},r)). This and the bound (detA1/2(x0))1Λn/2(\det A^{-1/2}(x_{0}))^{-1}\leq\Lambda^{n/2} imply that

rA(x0)(E;Wx0(x0,r))rn1+Crα\displaystyle r\mapsto\frac{\mathscr{F}_{A(x_{0})}(E;\textbf{{W}}_{x_{0}}(x_{0},r))}{r^{n-1}}+Cr^{\alpha} (4.54)

is monotone increasing on (0,dx0)(0,d_{x_{0}}). Note U=Tx01(Ux0)U=T_{x_{0}}^{-1}(U_{x_{0}}) and LipTx01=A1/2(x0)Λ1/2\operatorname{Lip}T_{x_{0}}^{-1}=||A^{1/2}(x_{0})||\leq\Lambda^{1/2}. Given xUx\in\partial U, setting y=Tx0(x)Ux0y=T_{x_{0}}(x)\in\partial U_{x_{0}}, it follows that

|xx0|=|Tx01(y)Tx01|LipTx01|yx0|Λ1/2dist(x0,Ux0).\displaystyle|x-x_{0}|=|T_{x_{0}}^{-1}(y)-T_{x_{0}}^{-1}|\leq\operatorname{Lip}T_{x_{0}}^{-1}|y-x_{0}|\leq\Lambda^{1/2}\mathrm{dist}(x_{0},\partial U_{x_{0}}). (4.55)

Hence dist(x0,U)Λ1/2dist(x0,Ux0)\mathrm{dist}(x_{0},\partial U)\leq\Lambda^{1/2}\mathrm{dist}(x_{0},\partial U_{x_{0}}) and so d=Λ1/2min{r0,dist(x0,U}dx0d=\Lambda^{-1/2}\min\{r_{0},\mathrm{dist}(x_{0},\partial U\}\leq d_{x_{0}}. Thus (4.54) is monotone increasing on (0,d)(0,d). By (2.11) we have

A(x)νE,νE1/2\displaystyle\langle A(x)\nu_{E},\nu_{E}\rangle^{1/2} A(x0)νE,νE1/2+(A(x)A(x0))νE,νE1/2\displaystyle\leq\langle A(x_{0})\nu_{E},\nu_{E}\rangle^{1/2}+\langle(A(x)-A(x_{0}))\nu_{E},\nu_{E}\rangle^{1/2}
A(x0)νE,νE1/2+12λACα|xx0|α\displaystyle\leq\langle A(x_{0})\nu_{E},\nu_{E}\rangle^{1/2}+\frac{1}{2\lambda}||A||_{C^{\alpha}}|x-x_{0}|^{\alpha} (4.56)

and so A(x)νE,νE1/2A(x0)νE,νE1/2+CACαsα\langle A(x)\nu_{E},\nu_{E}\rangle^{1/2}\leq\langle A(x_{0})\nu_{E},\nu_{E}\rangle^{1/2}+C||A||_{C^{\alpha}}s^{\alpha} for xWx0(x0,s)x\in\textbf{{W}}_{x_{0}}(x_{0},s) by (4.7). It follows that

A(E;Wx0(x0,s))sn1\displaystyle\frac{\mathscr{F}_{A}(E;\textbf{{W}}_{x_{0}}(x_{0},s))}{s^{n-1}} A(x0)(E;Wx0(x0,s))sn1+CsαP(E;Wx0(x0,s))sn1\displaystyle\leq\frac{\mathscr{F}_{A(x_{0})}(E;\textbf{{W}}_{x_{0}}(x_{0},s))}{s^{n-1}}+Cs^{\alpha}\frac{P(E;\textbf{{W}}_{x_{0}}(x_{0},s))}{s^{n-1}}
A(x0)(E;Wx0(x0,s))sn1+Csα\displaystyle\leq\frac{\mathscr{F}_{A(x_{0})}(E;\textbf{{W}}_{x_{0}}(x_{0},s))}{s^{n-1}}+Cs^{\alpha} (4.57)

where we used that P(E;Wx0(x0,s))P(E;B(x0,Λ1/2s))Csn1P(E;\textbf{{W}}_{x_{0}}(x_{0},s))\leq P(E;\textbf{{B}}(x_{0},\Lambda^{1/2}s))\leq Cs^{n-1} by the upper perimeter bound (4.27). Similarly, we have

A(x0)(E;Wx0(x0,r))rn1A(E;Wx0(x0,r))rn1+Crα\displaystyle\frac{\mathscr{F}_{A(x_{0})}(E;\textbf{{W}}_{x_{0}}(x_{0},r))}{r^{n-1}}\leq\frac{\mathscr{F}_{A}(E;\textbf{{W}}_{x_{0}}(x_{0},r))}{r^{n-1}}+Cr^{\alpha} (4.58)

Combining this last inequality and (4.4) with (4.54) and srs\leq r yields

A(E;Wx0(x0,s))sn1A(E;Wx0(x0,r))rn1+Crα\displaystyle\frac{\mathscr{F}_{A}(E;\textbf{{W}}_{x_{0}}(x_{0},s))}{s^{n-1}}\leq\frac{\mathscr{F}_{A}(E;\textbf{{W}}_{x_{0}}(x_{0},r))}{r^{n-1}}+Cr^{\alpha} (4.59)

as desired. ∎

For x0Ex_{0}\in\partial E, define the A\mathscr{F}_{A}-density ratio of EE at x0x_{0} by

θA(E,x0,r)=A(E;Wx0(x0,r))rn1.\displaystyle\theta_{A}(E,x_{0},r)=\frac{\mathscr{F}_{A}(E;\textbf{{W}}_{x_{0}}(x_{0},r))}{r^{n-1}}. (4.60)

and the A\mathscr{F}_{A}-density of EE at x0x_{0} by

θA(E,x0)=limr0+θA(E,x0,r)\displaystyle\theta_{A}(E,x_{0})=\lim_{r\to 0^{+}}\theta_{A}(E,x_{0},r) (4.61)

when the limit exists.

Corollary 4.8 (Existence of densities).

If EE is a (κ,α)(\kappa,\alpha)-almost-minimizer of A\mathscr{F}_{A} in UU at scale r0r_{0}, then for every x0UEx_{0}\in U\cap\partial E the density

θA(E,x0)=limr0+θA(E,x0,r)\displaystyle\theta_{A}(E,x_{0})=\lim_{r\to 0^{+}}\theta_{A}(E,x_{0},r) (4.62)

exists.

Proof.

For every 0<sr<d0<s\leq r<d we have by almost-monotonicity that θA(E,x0,s)θA(E,x0,r)+Crα\theta_{A}(E,x_{0},s)\leq\theta_{A}(E,x_{0},r)+Cr^{\alpha}. Taking the lim sup\limsup as s0+s\to 0^{+} followed by the lim inf\liminf as r0+r\to 0^{+} yields

lim sups0+θA(E,x0,s)lim infr0+θA(E,x0,r)+lim supr0+Crα=lim infr0+θA(E,x0,r)\displaystyle\limsup_{s\to 0^{+}}\theta_{A}(E,x_{0},s)\leq\liminf_{r\to 0^{+}}\theta_{A}(E,x_{0},r)+\limsup_{r\to 0^{+}}Cr^{\alpha}=\liminf_{r\to 0^{+}}\theta_{A}(E,x_{0},r) (4.63)

Hence θA(E,x0)=limr0+θA(E,x0,r)\theta_{A}(E,x_{0})=\lim_{r\to 0^{+}}\theta_{A}(E,x_{0},r) exists. ∎

Using the almost-monotonicity formula, we are now able to control the perimeter density ratios from below.

Proposition 4.9.

There exists a positive constant C=C(n,λ,Λ,κ,α,r0,ACα)C=C(n,\lambda,\Lambda,\kappa,\alpha,r_{0},||A||_{C^{\alpha}}) with the following property. If EE is a (κ,α)(\kappa,\alpha)-almost-minimizer of A\mathscr{F}_{A} in UU at scale r0r_{0}, then for every x0UEx_{0}\in U\cap\partial E, we have

ωn1(λ/Λ)n/2CrαP(E;B(x0,r))rn1\displaystyle\omega_{n-1}(\lambda/\Lambda)^{n/2}-Cr^{\alpha}\leq\frac{P(E;\textbf{{B}}(x_{0},r))}{r^{n-1}} (4.64)

for r<min{r0,dist(x0,U)}r<\min\{r_{0},\mathrm{dist}(x_{0},\partial U)\}.

Proof.

Let r<min{r0,dist(x0,U)}r<\min\{r_{0},\mathrm{dist}(x_{0},\partial U)\}. First consider the case x0Ex_{0}\in\partial^{*}\!E. The limit of perimeter density rations at a point in the reduced boundary converge to ωn1\omega_{n-1} as r0+r\to 0^{+} [Mag12, Corollary 15.8]. Note that Λ1/2r<Λ1/2min{r0,dist(x0,U)}\Lambda^{-1/2}r<\Lambda^{-1/2}\min\{r_{0},\mathrm{dist}(x_{0},\partial U)\} and so for s<rs<r we can apply Theorem 4.7 with Λ1/2sΛ1/2r\Lambda^{-1/2}s\leq\Lambda^{-1/2}r to obtain

A(E;Wx0(x0,Λ1/2s))(Λ1/2s)n1A(E;Wx0(x0,Λ1/2r))(Λ1/2r)n1+Crα\displaystyle\frac{\mathscr{F}_{A}(E;\textbf{{W}}_{x_{0}}(x_{0},\Lambda^{-1/2}s))}{(\Lambda^{-1/2}s)^{n-1}}\leq\frac{\mathscr{F}_{A}(E;\textbf{{W}}_{x_{0}}(x_{0},\Lambda^{-1/2}r))}{(\Lambda^{-1/2}r)^{n-1}}+Cr^{\alpha} (4.65)

By comparability to perimeter and (4.7), we have

ωn1\displaystyle\omega_{n-1} =lims0+P(E;B(x0,(λ/Λ)1/2s))((λ/Λ)1/2s)n1lims0+1λn/2A(E;Wx0(x0,Λ1/2s))(Λ1/2s)n1\displaystyle=\lim_{s\to 0^{+}}\frac{P(E;\textbf{{B}}(x_{0},(\lambda/\Lambda)^{1/2}s))}{((\lambda/\Lambda)^{1/2}s)^{n-1}}\leq\lim_{s\to 0^{+}}\frac{1}{\lambda^{n/2}}\frac{\mathscr{F}_{A}(E;\textbf{{W}}_{x_{0}}(x_{0},\Lambda^{-1/2}s))}{(\Lambda^{-1/2}s)^{n-1}}
1λn/2A(E;Wx0(x0,Λ1/2r))(Λ1/2r)n1+CrαΛ(n1)/2λn/2A(E;B(x0,r))rn1+Crα\displaystyle\leq\frac{1}{\lambda^{n/2}}\frac{\mathscr{F}_{A}(E;\textbf{{W}}_{x_{0}}(x_{0},\Lambda^{-1/2}r))}{(\Lambda^{-1/2}r)^{n-1}}+Cr^{\alpha}\leq\frac{\Lambda^{(n-1)/2}}{\lambda^{n/2}}\frac{\mathscr{F}_{A}(E;\textbf{{B}}(x_{0},r))}{r^{n-1}}+Cr^{\alpha}
(Λλ)n/2P(E;B(x0,r))rn1+Crα.\displaystyle\leq\Big{(}\frac{\Lambda}{\lambda}\Big{)}^{n/2}\frac{P(E;\textbf{{B}}(x_{0},r))}{r^{n-1}}+Cr^{\alpha}. (4.66)

Hence (4.64) holds for x0Ex_{0}\in\partial^{*}\!E.

Now consider the general case x0Ex_{0}\in\partial E (but perhaps not in E\partial^{*}\!E). Given 0<s<r0<s<r, there is y0Ey_{0}\in\partial^{*}\!E with B(y0,s)B(x0,r)\textbf{{B}}(y_{0},s)\subset\textbf{{B}}(x_{0},r) by sptμE=E=E¯\operatorname{spt}\mu_{E}=\partial E=\overline{\partial^{*}\!E}. It follows that

(sr)n1P(E;B(y0,s))sn1P(E;B(x0,r))rn1\displaystyle\Big{(}\frac{s}{r}\Big{)}^{n-1}\frac{P(E;\textbf{{B}}(y_{0},s))}{s^{n-1}}\leq\frac{P(E;\textbf{{B}}(x_{0},r))}{r^{n-1}} (4.67)

and so applying (4.64) at y0Ey_{0}\in\partial^{*}\!E gives

(sr)n1(ωn1(λ/Λ)n/2Csα)P(E;B(x0,r))rn1.\displaystyle\Big{(}\frac{s}{r}\Big{)}^{n-1}\Big{(}\omega_{n-1}(\lambda/\Lambda)^{n/2}-Cs^{\alpha}\Big{)}\leq\frac{P(E;\textbf{{B}}(x_{0},r))}{r^{n-1}}. (4.68)

Sending srs\to r^{-} completes the proof. ∎

Let us recall a definition. For a set of locally finite perimeter of EE, the essential boundary of EE, denoted by eE\partial^{e}E, is the set of points with neither full nor zero volume density, that is,

eE=n(E(0)E(1)).\displaystyle\partial^{e}E=\mathbb{R}^{n}\setminus(E^{(0)}\cup E^{(1)}). (4.69)

Here E(t)E^{(t)} denotes the points of volume density tt, that is,

E(t)={xn:limr0+|EB(x,r)|ωnrn=t}.\displaystyle E^{(t)}=\Big{\{}x\in\mathbb{R}^{n}\mathrel{\mathop{\mathchar 58\relax}}\lim_{r\to 0^{+}}\frac{|E\cap\textbf{{B}}(x,r)|}{\omega_{n}r^{n}}=t\Big{\}}. (4.70)

In general, we always have eEE\partial^{e}E\subset\partial E. Federer’s theorem states that n1(eEE)=0\operatorname{\mathcal{H}^{n-1}}(\partial^{e}E\setminus\partial^{*}\!E)=0 for sets of locally finite perimeter in n\mathbb{R}^{n}.

A consequence of the volume density bounds (4.71) in the following proposition is that the topological boundary of an almost-minimizer EE cannot contain any points of zero or full volume density, that is, the essential boundary eE\partial^{e}E in UU equals the topological boundary E\partial E in UU. This fact precludes the existence of sharp cusps in the topological boundary of EE as well as prevents two sheets of the topological boundary from touching tangentially. The perimeter bounds (4.72) show that the perimeter measure for EE is (n1)(n-1)-Ahlfors regular up to scale r0r_{0}.

Proposition 4.10 (Volume and perimeter bounds for almost-minimizers).

There exist positive constants c=c(n,λ,Λ)(0,1)c=c(n,\lambda,\Lambda)\in(0,1), C=C(n,λ,Λ,κ,α,r0)C=C(n,\lambda,\Lambda,\kappa,\alpha,r_{0}), and ε=ε(n,λ,Λ,κ,α,r0,ACα)\varepsilon=\varepsilon(n,\lambda,\Lambda,\kappa,\alpha,r_{0},||A||_{C^{\alpha}}) with the following property. with the following property. If EE is a (κ,α)(\kappa,\alpha)-almost-minimizer of A\mathscr{F}_{A} in UU at scale r0r_{0}, then for every x0UEx_{0}\in U\cap\partial E with r<d=min{r0,dist(x0,U),ε}<r<d=\min\{r_{0},\mathrm{dist}(x_{0},\partial U),\varepsilon\}<\infty,

c|EB(x0,r)|ωnrn1c,\displaystyle c\leq\frac{|E\cap\textbf{{B}}(x_{0},r)|}{\omega_{n}r^{n}}\leq 1-c, (4.71)

and

cP(E;B(x0,r))rn1C\displaystyle c\leq\frac{P(E;\textbf{{B}}(x_{0},r))}{r^{n-1}}\leq C (4.72)

Moreover, the volume density bounds (4.71) imply EU=eEU\partial E\cap U=\partial^{e}E\cap U and so Federer’s theorem gives

n1(U(EE))=0.\displaystyle\operatorname{\mathcal{H}^{n-1}}(U\cap(\partial E\setminus\partial^{*}\!E))=0. (4.73)
Proof.

The upper bound of (4.72) was proved in Lemma 4.4. For the lower bound of (4.72) take ε>0\varepsilon>0 small enough so that Cεα(1/2)ωn1(λ/Λ)n/2C\varepsilon^{\alpha}\leq(1/2)\omega_{n-1}(\lambda/\Lambda)^{n/2} where CC is the constant in Proposition 4.9.

Recall from the proof of Lemma 4.4 that for m(r)=|EB(x0,r)|m(r)=|E\cap\textbf{{B}}(x_{0},r)| we have m(r)=n1(E(1)B(x0,r))m^{\prime}(r)=\operatorname{\mathcal{H}^{n-1}}(E^{(1)}\cap\textbf{{B}}(x_{0},r)) for a.e. r<dr<d. Then the inequality (4.30) becomes

P(E;B(x0,r))C(m(r)+rα+n1).\displaystyle P(E;\textbf{{B}}(x_{0},r))\leq C(m^{\prime}(r)+r^{\alpha+n-1}). (4.74)

So by the lower bound of (4.72) we have

crn1Cm(r)+Crα+n1.\displaystyle cr^{n-1}\leq Cm^{\prime}(r)+Cr^{\alpha+n-1}. (4.75)

Taking ε\varepsilon small enough so that Cεαc/2C\varepsilon^{\alpha}\leq c/2 and relabeling c/2c/2 to cc gives crn1m(r)cr^{n-1}\leq m^{\prime}(r) for a.e. r<dr<d. Integrating on (0,r)(0,r) and modifying constants gives cωnrnm(r)=|EB(x0,r)|c\omega_{n}r^{n}\leq m(r)=|E\cap\textbf{{B}}(x_{0},r)| which is the lower bound of (4.71). Since EcE^{c} is also a (κ,α)(\kappa,\alpha)-almost-minimizer of A\mathscr{F}_{A}, we can apply this lower bound of (4.71) to get cωnrn|EcB(x0,r)|c\omega_{n}r^{n}\leq|E^{c}\cap\textbf{{B}}(x_{0},r)| which gives the upper bound of (4.71).

Federer’s theorem [Mag12, Theorem 16.2] states n1(eEE)=0\operatorname{\mathcal{H}^{n-1}}(\partial^{e}E\setminus\partial^{*}\!E)=0. The volume density bounds (4.71) imply EU=eEU\partial E\cap U=\partial^{e}E\cap U and hence n1(U(EE))=0\operatorname{\mathcal{H}^{n-1}}(U\cap(\partial E\setminus\partial^{*}\!E))=0

Hence, given any (κ,α)(\kappa,\alpha)-almost-minimizer EE of A\mathscr{F}_{A} in UU at scale r0r_{0}, we can shrink r0r_{0} by a fixed amount, depending only on the universal constants n,λ,Λ,κ,αn,\lambda,\Lambda,\kappa,\alpha, and an upper bound for ACα||A||_{C^{\alpha}}, so that at points x0UEx_{0}\in U\cap\partial E the volume and perimeter bounds hold for all r<min{r0,dist(x0,U)}r<\min\{r_{0},\mathrm{dist}(x_{0},\partial U)\}. Throughout the rest of this paper, we will work at this smaller scale and use the volume and perimeter bounds.

4.5. Compactness for the class of A\mathscr{F}_{A}-energies

In addition to having fixed nn, λ\lambda, Λ\Lambda, α\alpha, fix a positive constants M1M_{1} and M2M_{2}. Define the class of admissible matrix-valued functions 𝒜\mathscr{A} by

𝒜={ACα(n;nn):A is symmetric,ACαM1,A(x)M2, and λ|ξ|2A(x)ξ,ξΛ|ξ|2 for all x,ξn},\displaystyle\mathscr{A}=\Big{\{}A\in C^{\alpha}(\mathbb{R}^{n};\mathbb{R}^{n}\otimes\mathbb{R}^{n})\mathrel{\mathop{\mathchar 58\relax}}\begin{array}[]{lr}A\text{ is symmetric},\ ||A||_{C^{\alpha}}\leq M_{1},\ ||A(x)||\leq M_{2},\text{ and }\\ \lambda|\xi|^{2}\leq\langle A(x)\xi,\xi\rangle\leq\Lambda|\xi|^{2}\text{ for all }x,\xi\in\mathbb{R}^{n}\end{array}\Big{\}}, (4.78)
Lemma 4.11.

𝒜\mathscr{A} is compact in the topology of uniform convergence on compact sets as a subspace of C(n;nn)C(\mathbb{R}^{n};\mathbb{R}^{n}\otimes\mathbb{R}^{n}).

Proof.

Let {Ah}h\{A_{h}\}_{h\in\mathbb{N}} be a sequence in 𝒜\mathscr{A}. We apply Arzelà-Ascoli to {Ah}h\{A_{h}\}_{h\in\mathbb{N}} noting that pointwise-boundedness follows from Ah(x)M2||A_{h}(x)||\leq M_{2} and equicontinuity follows from AhCαM1||A_{h}||_{C^{\alpha}}\leq M_{1}. Hence there is a subsequence {Ah(k)}k\{A_{h(k)}\}_{k\in\mathbb{N}} and AC(n;nn)A\in C(\mathbb{R}^{n};\mathbb{R}^{n}\otimes\mathbb{R}^{n}) such that Ah(k)AA_{h(k)}\to A uniformly on compact sets. It follows that AA is symmetric and A(x)limkAh(k)(x)M2||A(x)||\leq\lim_{k\to\infty}||A_{h(k)}(x)||\leq M_{2} for any xnx\in\mathbb{R}^{n}. For any x,y,ξnx,y,\xi\in\mathbb{R}^{n},

A(x)A(y)\displaystyle||A(x)-A(y)|| A(x)Ah(k)(x)+M1|xy|α+Ah(k)(y)A(y),\displaystyle\leq||A(x)-A_{h(k)}(x)||+M_{1}|x-y|^{\alpha}+||A_{h(k)}(y)-A(y)||,
λ|ξ|2\displaystyle\lambda|\xi|^{2} A(x)ξ,ξ+(Ah(k)(x)A(x))ξ,ξΛ|ξ|2.\displaystyle\leq\langle A(x)\xi,\xi\rangle+\langle(A_{h(k)}(x)-A(x))\xi,\xi\rangle\leq\Lambda|\xi|^{2}. (4.79)

Sending h(k)h(k)\to\infty, gives ACαM1||A||_{C^{\alpha}}\leq M_{1} and λ|ξ|2A(x)ξ,ξΛ|ξ|2\lambda|\xi|^{2}\leq\langle A(x)\xi,\xi\rangle\leq\Lambda|\xi|^{2}. Thus A𝒜A\in\mathscr{A} and so 𝒜\mathscr{A} is compact. ∎

Fix an open set UU and κ0\kappa\geq 0. Define the class \mathscr{M} of almost-minimizers of A\mathscr{F}_{A} in UU for A𝒜A\in\mathscr{A} by

={En:E is a (κ,α)-almost-min. of A in U at scale r0 for some A𝒜 and r0>0}.\displaystyle\mathscr{M}=\Big{\{}E\subset\mathbb{R}^{n}\mathrel{\mathop{\mathchar 58\relax}}E\text{ is a $(\kappa,\alpha)$-almost-min. of }\mathscr{F}_{A}\text{ in }U\text{ at scale }r_{0}\text{ for some }A\in\mathscr{A}\text{ and }r_{0}>0\Big{\}}. (4.80)

We will show that \mathscr{M} is compact by separately proving precompactness and closedness.

Proposition 4.12 (Precompactness of \mathscr{M}).

Suppose that {Eh}h\{E_{h}\}_{h\in\mathbb{N}}\subset\mathscr{M} (that is, EhE_{h} is a (κ,α)(\kappa,\alpha)-almost-minimizer of Ah\mathscr{F}_{A_{h}} in UU at scale rhr_{h} for some Ah𝒜A_{h}\in\mathscr{A} and rh>0r_{h}>0), and that r0=lim infhrh>0r_{0}=\liminf_{h\to\infty}r_{h}>0. For any open VUV\subset\subset U with P(V)<P(V)<\infty, there exist h(k)h(k)\to\infty as kk\to\infty, a set of finite perimeter EVE\subset V, and A𝒜A\in\mathscr{A} such that

VEh(k)E,μVEh(k)μE,\displaystyle V\cap E_{h(k)}\to E,\qquad\mu_{V\cap E_{h(k)}}\overset{\ast}{\rightharpoonup}\mu_{E},
Ah(k)A uniformly on compact sets.\displaystyle A_{h(k)}\to A\text{ uniformly on compact sets}. (4.81)
Proof.

First we choose h(k)h(k)\to\infty as kk\to\infty such that limkrh(k)=r0\lim_{k\to\infty}r_{h(k)}=r_{0}. Let xVx\in V and B(x,4r)U\textbf{{B}}(x,4r)\subset U with 2r<r02r<r_{0}. Let k0k_{0} be such that 2r<rh(k)2r<r_{h(k)} for kk0k\geq k_{0}. If P(Eh(k);B(x,r))>0P(E_{h(k)};\textbf{{B}}(x,r))>0, there is yB(x,r)Eh(k)y\in\textbf{{B}}(x,r)\cap\partial E_{h(k)} and so P(Eh(k);B(x,r))P(Eh(k);B(y,2r))C(2r)n1P(E_{h(k)};\textbf{{B}}(x,r))\leq P(E_{h(k)};\textbf{{B}}(y,2r))\leq C(2r)^{n-1} by upper density bound (4.72) since B(y,2r)B(x,4r)U\textbf{{B}}(y,2r)\subset\textbf{{B}}(x,4r)\subset U. By [Mag12, (16.10)], we have for kk0k\geq k_{0} that

P(Eh(k)B(x,r))P(Eh(k);B(x,r))+P(B(x,r))Cr0n1<.\displaystyle P(E_{h(k)}\cap\textbf{{B}}(x,r))\leq P(E_{h(k)};\textbf{{B}}(x,r))+P(\textbf{{B}}(x,r))\leq Cr_{0}^{n-1}<\infty. (4.82)

and so supkP(Eh(k)B(x,r))<\sup_{k}P(E_{h(k)}\cap\textbf{{B}}(x,r))<\infty. Since VV is open and compactly contained in UU, the balls with centers in VV that are contained in UU form a covering for V¯\overline{V}. Hence we may cover VV by finitely many balls {Bj}j=1N\{\textbf{{B}}_{j}\}_{j=1}^{N} where Bj=B(xj,sj)\textbf{{B}}_{j}=\textbf{{B}}(x_{j},s_{j}) satisfy B(xj,4sj)U\textbf{{B}}(x_{j},4s_{j})\subset U with xjVx_{j}\in V and 2sj<r02s_{j}<r_{0} for 1jN1\leq j\leq N. Choose R>0R>0 so that j=1NBjBR\bigcup_{j=1}^{N}\textbf{{B}}_{j}\subset\textbf{{B}}_{R}. Then

P(Eh(k)(j=1NBj))\displaystyle P\Big{(}E_{h(k)}\cap\big{(}\bigcup_{j=1}^{N}\textbf{{B}}_{j}\big{)}\Big{)} P(Eh(k);j=1NBj)+P(j=1NBj)j=1NP(Eh(k);Bj)+P(j=1NBj)\displaystyle\leq P\Big{(}E_{h(k)};\bigcup_{j=1}^{N}\textbf{{B}}_{j}\Big{)}+P\Big{(}\bigcup_{j=1}^{N}\textbf{{B}}_{j}\Big{)}\leq\sum_{j=1}^{N}P(E_{h(k)};\textbf{{B}}_{j})+P\Big{(}\bigcup_{j=1}^{N}\textbf{{B}}_{j}\Big{)}
Cr0n1N+P(BR)<\displaystyle\leq Cr_{0}^{n-1}N+P(\textbf{{B}}_{R})<\infty (4.83)

and so we may apply Theorem 3.1 to construct a set FBRF\subset\textbf{{B}}_{R} of finite perimeter and a further subsequence indices h(k)h(k) such that Eh(k)(j=1NBj)FE_{h(k)}\cap\big{(}\bigcup_{j=1}^{N}\textbf{{B}}_{j}\big{)}\to F. Setting E=VFE=V\cap F, we have that VEh(k)EV\cap E_{h(k)}\to E and supkP(VEh(k))supkP(Eh(k)(j=1NBj))<\sup_{k}P(V\cap E_{h(k)})\leq\sup_{k}P(E_{h(k)}\cap\big{(}\bigcup_{j=1}^{N}\textbf{{B}}_{j}\big{)})<\infty. Finally, given φCc0(n)\varphi\in C_{c}^{0}(\mathbb{R}^{n}) and ψCc1(n)\psi\in C_{c}^{1}(\mathbb{R}^{n}), we have

|n\displaystyle\bigg{|}\int_{\mathbb{R}^{n}} φdμVEh(k)nφdμE|\displaystyle\varphi\>d\mu_{V\cap E_{h(k)}}-\int_{\mathbb{R}^{n}}\varphi\>d\mu_{E}\bigg{|}
|n(φψ)𝑑μVEh(k)|+|nψ𝑑μVEh(k)nψ𝑑μE|+|n(ψφ)𝑑μE|\displaystyle\leq\bigg{|}\int_{\mathbb{R}^{n}}(\varphi-\psi)\>d\mu_{V\cap E_{h(k)}}\bigg{|}+\bigg{|}\int_{\mathbb{R}^{n}}\psi\>d\mu_{V\cap E_{h(k)}}-\int_{\mathbb{R}^{n}}\psi\>d\mu_{E}\bigg{|}+\bigg{|}\int_{\mathbb{R}^{n}}(\psi-\varphi)\>d\mu_{E}\bigg{|}
φψsupsupk|μVEh(k)|(n)+|VEh(k)ψdxEψdx|+φψsup|μE|(n)\displaystyle\leq||\varphi-\psi||_{\sup}\sup_{k}|\mu_{V\cap E_{h(k)}}|(\mathbb{R}^{n})+\bigg{|}\int_{V\cap E_{h(k)}}\nabla\psi\>dx-\int_{E}\nabla\psi\>dx\bigg{|}+||\varphi-\psi||_{\sup}|\mu_{E}|(\mathbb{R}^{n})
φψsupsupk|μVEh(k)|(n)+ψsup|(VEh(k))ΔE|+φψsup|μE|(n).\displaystyle\leq||\varphi-\psi||_{\sup}\sup_{k}|\mu_{V\cap E_{h(k)}}|(\mathbb{R}^{n})+||\nabla\psi||_{\sup}|(V\cap E_{h(k)})\Delta E|+||\varphi-\psi||_{\sup}|\mu_{E}|(\mathbb{R}^{n}). (4.84)

Since VEh(k)EV\cap E_{h(k)}\to E, this gives

lim supksup|nφ𝑑μVEh(k)nφ𝑑μE|φψsup(supk|μVEh(k)|(n)+|μE|(n)).\displaystyle\limsup_{k\to\sup}\bigg{|}\int_{\mathbb{R}^{n}}\varphi\>d\mu_{V\cap E_{h(k)}}-\int_{\mathbb{R}^{n}}\varphi\>d\mu_{E}\bigg{|}\leq||\varphi-\psi||_{\sup}\big{(}\sup_{k}|\mu_{V\cap E_{h(k)}}|(\mathbb{R}^{n})+|\mu_{E}|(\mathbb{R}^{n})\big{)}. (4.85)

So by density of Cc1(n)C_{c}^{1}(\mathbb{R}^{n}) in Cc0(n)C_{c}^{0}(\mathbb{R}^{n}) in the sup norm, we have μVEh(k)μE\mu_{V\cap E_{h(k)}}\overset{\ast}{\rightharpoonup}\mu_{E}. Finally, by Lemma 4.11 we may extract a further subsequence such that, up to relabeling, we also have Ah(k)AA_{h(k)}\to A uniformly on compact sets. ∎

Proposition 4.13 (Closedness of \mathscr{M}).

Suppose that {Eh}h\{E_{h}\}_{h\in\mathbb{N}}\subset\mathscr{M} (that is, EhE_{h} is a (κ,α)(\kappa,\alpha)-almost-minimizer of Ah\mathscr{F}_{A_{h}} in UU at scale rhr_{h} for some Ah𝒜A_{h}\in\mathscr{A} and rh>0r_{h}>0), r0=lim infhrh>0r_{0}=\liminf_{h\to\infty}r_{h}>0, VUV\subset\subset U is an open set with P(V)<P(V)<\infty such that VEhEV\cap E_{h}\to E for a set of finite perimeter EE, and AhAA_{h}\to A uniformly on compact sets for some A𝒜A\in\mathscr{A}. Then EE is a (κ,α)(\kappa,\alpha)-almost-minimizer of A\mathscr{F}_{A} in VV at scale r0r_{0}. Moreover,

μVEhμE,\displaystyle\mu_{V\cap E_{h}}\overset{\ast}{\rightharpoonup}\mu_{E}, (4.86)
Ah(Eh;)A(E;) in V\displaystyle\mathscr{F}_{A_{h}}(E_{h};\>\cdot\>)\overset{\ast}{\rightharpoonup}\mathscr{F}_{A}(E;\>\cdot\>)\text{ in }V (4.87)

where we view Ah(Eh;)\mathscr{F}_{A_{h}}(E_{h};\>\cdot\>) and A(E;)\mathscr{F}_{A}(E;\>\cdot\>) as Radon measures. In particular,

  1. (i)

    if xhVEhx_{h}\in V\cap\partial E_{h}, xhxx_{h}\to x, and xVx\in V, then xVEx\in V\cap\partial E;

  2. (ii)

    if xVEx\in V\cap\partial E, then there exists {xh}h\{x_{h}\}_{h\in\mathbb{N}} with xhVEhx_{h}\in V\cap\partial E_{h} such that xhxx_{h}\to x.

Proof.

By the same argument as in the proof of Proposition 4.12 we can show suphP(VEh)<\sup_{h}P(V\cap E_{h})<\infty. The weak convergence μVEhμE\mu_{V\cap E_{h}}\overset{\ast}{\rightharpoonup}\mu_{E} of (4.86) follows from VEhEV\cap E_{h}\to E as also shown in the proof of Proposition 4.12.

To show that EE is a (κ,α)(\kappa,\alpha)-almost-minimizer of A\mathscr{F}_{A} our strategy is as follows. Given a competitor FF for EE, we modify FF to construct competitors FhF_{h} and apply the almost-minimality of EhE_{h} with respect to Ah\mathscr{F}_{A_{h}}. We then pass the minimality inequalities through limits to obtain the desired almost-minimality inequality for EE.

Suppose EΔFVB(x,r)E\Delta F\subset\subset V\cap\textbf{{B}}(x,r) with xVx\in V and r<r0r<r_{0}. For yVy\in V, set d(y)=min{r0,dist(y,V)}>0d(y)=\min\{r_{0},\mathrm{dist}(y,\partial V)\}>0. Since n1  Eh\operatorname{\mathcal{H}^{n-1}}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>\>\partial^{*}\!E_{h} and n1  F\operatorname{\mathcal{H}^{n-1}}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>\>\partial^{*}\!F are Radon measures, we have that for a.e. s(0,d(y))s\in(0,d(y)),

n1(B(y,s)F)=n1(B(y,s)Eh)=0,h.\displaystyle\operatorname{\mathcal{H}^{n-1}}(\partial\textbf{{B}}(y,s)\cap\partial^{*}\!F)=\operatorname{\mathcal{H}^{n-1}}(\partial\textbf{{B}}(y,s)\cap\partial^{*}\!E_{h})=0,\ \forall h\in\mathbb{N}. (4.88)

Note that |(E(1)ΔEh(1))B(y,d(y))|=|(EΔEh)B(y,d(y))||(E^{(1)}\Delta E_{h}^{(1)})\cap\textbf{{B}}(y,d(y))|=|(E\Delta E_{h})\cap\textbf{{B}}(y,d(y))| because a Lebesgue measurable set is equivalent to its set of points of full density. By the coarea formula, VEhEV\cap E_{h}\to E, and B(y,d(y))V\textbf{{B}}(y,d(y))\subset V, it follows that

0d(y)n1((E(1)ΔEh(1))B(y,s))𝑑s=|(E(1)ΔEh(1))B(y,d(y))|0\displaystyle\int_{0}^{d(y)}\operatorname{\mathcal{H}^{n-1}}((E^{(1)}\Delta E_{h}^{(1)})\cap\partial\textbf{{B}}(y,s))\>ds=|(E^{(1)}\Delta E_{h}^{(1)})\cap\textbf{{B}}(y,d(y))|\to 0 (4.89)

as hh\to\infty. Consequently, by Fatou’s lemma,

lim infhn1((E(1)ΔEh(1))B(y,s))=0\displaystyle\liminf_{h\to\infty}\operatorname{\mathcal{H}^{n-1}}((E^{(1)}\Delta E_{h}^{(1)})\cap\partial\textbf{{B}}(y,s))=0 (4.90)

for a.e. s(0,d(y))s\in(0,d(y)). Since EΔFE\Delta F is compactly contained in VB(x,r)V\cap\textbf{{B}}(x,r), we may find finitely many balls {B(yj,sj)}j=1N\{\textbf{{B}}(y_{j},s_{j})\}_{j=1}^{N} with yjVy_{j}\in V and sj(0,d(yj))s_{j}\in(0,d(y_{j})) satisfying (4.88) and (4.90) with y=yjy=y_{j}, s=sjs=s_{j} such that, setting G=j=1NB(yj,sj)G=\bigcup_{j=1}^{N}\textbf{{B}}(y_{j},s_{j}), we have EΔFGVB(x,r)E\Delta F\subset\subset G\subset\subset V\cap\textbf{{B}}(x,r). Now consider the comparison sets FhF_{h} defined by Fh=(FG)(EhG)F_{h}=(F\cap G)\cup(E_{h}\setminus G). Since Gj=1NB(yj,sj)\partial G\subset\bigcup_{j=1}^{N}\partial\textbf{{B}}(y_{j},s_{j}), by (4.88) there holds

n1(GF)=n1(GEh)=0,h.\displaystyle\operatorname{\mathcal{H}^{n-1}}(\partial G\cap\partial^{*}\!F)=\operatorname{\mathcal{H}^{n-1}}(\partial G\cap\partial^{*}\!E_{h})=0,\ \forall h\in\mathbb{N}. (4.91)

Additionally, E(1)G=F(1)GE^{(1)}\cap\partial G=F^{(1)}\cap\partial G since EΔFGE\Delta F\subset\subset G so that by (4.90) there holds

lim infhn1((F(1)ΔEh(1))G)=0.\displaystyle\liminf_{h\to\infty}\operatorname{\mathcal{H}^{n-1}}((F^{(1)}\Delta E_{h}^{(1)})\cap\partial G)=0. (4.92)

Observe that EhΔFhGUB(x,r)E_{h}\Delta F_{h}\subset G\subset\subset U\cap\textbf{{B}}(x,r) with xUx\in U. Since r<r0=lim infhrhr<r_{0}=\liminf_{h\to\infty}r_{h}, there is h0h_{0} such that r<rhr<r_{h} for all hh0h\geq h_{0}. For now fix hh0h\geq h_{0}. By (4.91) we can apply Proposition 4.3 to obtain

Ah(Fh;B(x,r))\displaystyle\mathscr{F}_{A_{h}}(F_{h};\textbf{{B}}(x,r)) =Ah(F;G)+Ah(Eh;B(x,r)G¯)+Ah(G;F(1)ΔEh(1))\displaystyle=\mathscr{F}_{A_{h}}(F;G)+\mathscr{F}_{A_{h}}(E_{h};\textbf{{B}}(x,r)\setminus\overline{G})+\mathscr{F}_{A_{h}}(G;F^{(1)}\Delta E_{h}^{(1)})
Ah(F;G)+Ah(Eh;B(x,r)G)+Λ1/2n1((F(1)ΔEh(1))G).\displaystyle\leq\mathscr{F}_{A_{h}}(F;G)+\mathscr{F}_{A_{h}}(E_{h};\textbf{{B}}(x,r)\setminus G)+\Lambda^{1/2}\operatorname{\mathcal{H}^{n-1}}((F^{(1)}\Delta E_{h}^{(1)})\cap\partial G). (4.93)

Since FhF_{h} is a competitor for the Ah\mathscr{F}_{A_{h}}-almost-minimality of EhE_{h}, we have

Ah(Eh;B(x,r))Ah(F;G)+Ah(Eh;B(x,r)G)+Λ1/2n1((F(1)ΔEh(1))G)+κrα+n1\displaystyle\mathscr{F}_{A_{h}}(E_{h};\textbf{{B}}(x,r))\leq\mathscr{F}_{A_{h}}(F;G)+\mathscr{F}_{A_{h}}(E_{h};\textbf{{B}}(x,r)\setminus G)+\Lambda^{1/2}\operatorname{\mathcal{H}^{n-1}}((F^{(1)}\Delta E_{h}^{(1)})\cap\partial G)+\kappa r^{\alpha+n-1} (4.94)

which simplifies to

Ah(Eh;G)Ah(F;G)+Λ1/2n1((F(1)ΔEh(1))G)+κrα+n1.\displaystyle\mathscr{F}_{A_{h}}(E_{h};G)\leq\mathscr{F}_{A_{h}}(F;G)+\Lambda^{1/2}\operatorname{\mathcal{H}^{n-1}}((F^{(1)}\Delta E_{h}^{(1)})\cap\partial G)+\kappa r^{\alpha+n-1}. (4.95)

Similar to (2.11) we have Ah(y)νF,νF1/2A(y)νF,νF1/2+CAh(y)A(y)\langle A_{h}(y)\nu_{F},\nu_{F}\rangle^{1/2}\leq\langle A(y)\nu_{F},\nu_{F}\rangle^{1/2}+C||A_{h}(y)-A(y)||. Integrating yields

Ah(F;G)\displaystyle\mathscr{F}_{A_{h}}(F;G) A(F;G)+CAhAsupGP(F;G)\displaystyle\leq\mathscr{F}_{A}(F;G)+C||A_{h}-A||_{\sup G}P(F;G) (4.96)

where we set AhAsupG=supyGAh(y)A(y)||A_{h}-A||_{\sup{G}}=\sup_{y\in G}||A_{h}(y)-A(y)||. Taking the lim sup\limsup as hh\to\infty gives

lim suphAh(F;G)\displaystyle\limsup_{h\to\infty}\mathscr{F}_{A_{h}}(F;G) A(F;G)\displaystyle\leq\mathscr{F}_{A}(F;G) (4.97)

because lim suph0AhAsupG=0\limsup_{h\to 0}||A_{h}-A||_{\sup G}=0 by the uniform convergence AhAA_{h}\to\ A on compact sets. Similarly,

A(VEh;G)Ah(VEh;G)+CAAhsupGP(VEh;G).\displaystyle\mathscr{F}_{A}(V\cap E_{h};G)\leq\mathscr{F}_{A_{h}}(V\cap E_{h};G)+C||A-A_{h}||_{\sup{G}}\>P(V\cap E_{h};G). (4.98)

Using the fact that lim inf(ah+bh)lim infah+lim supbh\liminf(a_{h}+b_{h})\leq\liminf a_{h}+\limsup b_{h} for nonnegative sequences {ah},{bh}\{a_{h}\},\{b_{h}\}, we have

lim infhA(VEh;G)lim infhAh(VEh;G)+lim suph(CAAhsupGP(VEh;G)).\displaystyle\liminf_{h\to\infty}\mathscr{F}_{A}(V\cap E_{h};G)\leq\liminf_{h\to\infty}\mathscr{F}_{A_{h}}(V\cap E_{h};G)+\limsup_{h\to\infty}\big{(}C||A-A_{h}||_{\sup G}\>P(V\cap E_{h};G)\big{)}. (4.99)

By suphP(VEh;G)<\sup_{h}P(V\cap E_{h};G)<\infty and lim suph0AhAsupG=0\limsup_{h\to 0}||A_{h}-A||_{\sup G}=0, this becomes

lim infhA(VEh;G)lim infhAh(VEh;G).\displaystyle\liminf_{h\to\infty}\mathscr{F}_{A}(V\cap E_{h};G)\leq\liminf_{h\to\infty}\mathscr{F}_{A_{h}}(V\cap E_{h};G). (4.100)

Noting Ah(VEh;G)=Ah(Eh;G)\mathscr{F}_{A_{h}}(V\cap E_{h};G)=\mathscr{F}_{A_{h}}(E_{h};G) since GVG\subset\subset V and using the lower semicontinuity of A\mathscr{F}_{A} with respect to μVEhμE\mu_{V\cap E_{h}}\overset{\ast}{\rightharpoonup}\mu_{E} by Proposition 3.4, this implies

A(E;G)lim infhA(Eh;G)lim infhAh(Eh;G).\displaystyle\mathscr{F}_{A}(E;G)\leq\liminf_{h\to\infty}\mathscr{F}_{A}(E_{h};G)\leq\liminf_{h\to\infty}\mathscr{F}_{A_{h}}(E_{h};G). (4.101)

Now we combine our estimates, again using that lim infh(ah+bh)lim suphah+lim infhbh\liminf_{h}(a_{h}+b_{h})\leq\limsup_{h}a_{h}+\liminf_{h}b_{h}, and obtain

A(E;G)\displaystyle\mathscr{F}_{A}(E;G) lim infhAh(Eh;G)\displaystyle\leq\liminf_{h\to\infty}\mathscr{F}_{A_{h}}(E_{h};G) by (4.101)\displaystyle\text{ by }\eqref{comp ineq 3}
lim infh(Ah(F;G)+Λ1/2n1((F(1)ΔEh(1))G)+κrα+n1)\displaystyle\leq\liminf_{h\to\infty}\big{(}\mathscr{F}_{A_{h}}(F;G)+\Lambda^{1/2}\operatorname{\mathcal{H}^{n-1}}((F^{(1)}\Delta E_{h}^{(1)})\cap\partial G)+\kappa r^{\alpha+n-1}\big{)} by (4.95)\displaystyle\text{ by }\eqref{comp ineq 1}
lim suphAh(F;G)+Λ1/2lim infhn1((F(1)ΔEh(1))G)+κrα+n1\displaystyle\leq\limsup_{h\to\infty}\mathscr{F}_{A_{h}}(F;G)+\Lambda^{1/2}\liminf_{h\to\infty}\operatorname{\mathcal{H}^{n-1}}((F^{(1)}\Delta E_{h}^{(1)})\cap\partial G)+\kappa r^{\alpha+n-1}
A(F;G)+κrα+n1.\displaystyle\leq\mathscr{F}_{A}(F;G)+\kappa r^{\alpha+n-1}. by (4.97) and (4.92)\displaystyle\text{ by }\eqref{comp ineq 2}\text{ and }\eqref{liminf measure} (4.102)

Since EΔFGE\Delta F\subset\subset G, we can add A(E;B(x,r)G)=A(F;B(x,r)G)\mathscr{F}_{A}(E;\textbf{{B}}(x,r)\setminus G)=\mathscr{F}_{A}(F;\textbf{{B}}(x,r)\setminus G) to obtain

A(E;B(x,r))A(F;B(x,r))+κrα+n1\displaystyle\mathscr{F}_{A}(E;\textbf{{B}}(x,r))\leq\mathscr{F}_{A}(F;\textbf{{B}}(x,r))+\kappa r^{\alpha+n-1} (4.103)

as desired.

Next we prove the weak convergence of energy measures (4.87). Let Φh=Ah(VEh;)\Phi_{h}=\mathscr{F}_{A_{h}}(V\cap E_{h};\>\cdot\>) and Φ=A(E;)\Phi=\mathscr{F}_{A}(E;\>\cdot\>) which are Radon measures on n\mathbb{R}^{n}. It suffices to show the following claim.

Claim.

If Ψ\Psi is a Radon measure and {Φh(k)}k\{\Phi_{h(k)}\}_{k\in\mathbb{N}} is a subsequence such that Φh(k)Ψ\Phi_{h(k)}\overset{\ast}{\rightharpoonup}\Psi, then Φ  V=Ψ  V\Phi\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>V=\Psi\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>V.

Indeed, suppose the claim is true. By sequential compactness of Radon measures (which applies since suphΦh(n)Λ1/2suphP(VEh)<\sup_{h}\Phi_{h}(\mathbb{R}^{n})\leq\Lambda^{1/2}\sup_{h}P(V\cap E_{h})<\infty), for each subsequence of {Φh}h\{\Phi_{h}\}_{h\in\mathbb{N}} there exists a further subsequence that converges weakly to some Radon measure Ψ\Psi. By the claim Ψ  V=Φ  V\Psi\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>V=\Phi\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>V and so Φh  V\Phi_{h}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>V converges weakly to Φ  V\Phi\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>V. Since Φh=Ah(VEh;)  V=Ah(Eh;)  V\Phi_{h}=\mathscr{F}_{A_{h}}(V\cap E_{h};\>\cdot\>)\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>V=\mathscr{F}_{A_{h}}(E_{h};\>\cdot\>)\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>V by the decomposition formula of the Gauss-Green measure for the intersection of two sets of locally finite perimeter (see (16.4) of [Mag12, Theorem 16.3]), this will complete the proof of (4.87).

Now we prove the above claim. Suppose Φh(k)Ψ\Phi_{h(k)}\overset{\ast}{\rightharpoonup}\Psi for some Radon measure Ψ\Psi and subsequence {Φh(k)}k\{\Phi_{h(k)}\}_{k\in\mathbb{N}}. For convenience we will just write the indices as kk instead of h(k)h(k).

Let us show ΦΨ\Phi\leq\Psi on (n)\mathscr{B}(\mathbb{R}^{n}) where (n)\mathscr{B}(\mathbb{R}^{n}) denotes the Borel sets of n\mathbb{R}^{n}. Let WW be an open bounded set and set Wt={xW:dist(x,W)>t}W_{t}=\{x\in W\mathrel{\mathop{\mathchar 58\relax}}\mathrm{dist}(x,\partial W)>t\} for t>0t>0. Choose φCc(W;[0,1])\varphi\in C_{c}(W;[0,1]) with 1Wtφ1_{W_{t}}\leq\varphi. Note that (4.100) holds for any bounded set in place of GG by the same argument. So applying (4.100) with WtW_{t} in conjunction with the lower semicontinuity of A\mathscr{F}_{A} with respect to μVEhμE\mu_{V\cap E_{h}}\overset{\ast}{\rightharpoonup}\mu_{E} by Proposition 3.4 gives

Φ(Wt)\displaystyle\Phi(W_{t}) =A(E;Wt)lim infkAk(VEk;Wt)\displaystyle=\mathscr{F}_{A}(E;W_{t})\leq\liminf_{k\to\infty}\mathscr{F}_{A_{k}}(V\cap E_{k};W_{t})
lim infk(VEk)φ(x)Ak(x)νVEk,νVEk1/2dn1=φ𝑑ΨΨ(W).\displaystyle\leq\liminf_{k\to\infty}\int_{\partial^{*}\!(V\cap E_{k})}\varphi(x)\langle A_{k}(x)\nu_{V\cap E_{k}},\nu_{V\cap E_{k}}\rangle^{1/2}\operatorname{\,d\mathcal{H}^{n-1}}=\int\varphi\>d\Psi\leq\Psi(W). (4.104)

By monotone convergence, taking t0+t\to 0^{+} gives Φ(W)Ψ(W)\Phi(W)\leq\Psi(W). Since WW was an arbitrary open, bounded set, it follows that ΦΨ\Phi\leq\Psi on (n)\mathscr{B}(\mathbb{R}^{n}).

Now let B(x,s0)V\textbf{{B}}(x,s_{0})\subset\subset V with s0<r0s_{0}<r_{0}. Define

Fk=(EB(x,s))(EkB(x,s))\displaystyle F_{k}=\big{(}E\cap\textbf{{B}}(x,s))\cup(E_{k}\setminus\textbf{{B}}(x,s)) (4.105)

for s(0,s0)s\in(0,s_{0}) with

n1(EB(x,s))=n1(EkB(x,s))=0,k\displaystyle\operatorname{\mathcal{H}^{n-1}}(\partial^{*}\!E\cap\partial\textbf{{B}}(x,s))=\operatorname{\mathcal{H}^{n-1}}(\partial^{*}\!E_{k}\cap\partial\textbf{{B}}(x,s))=0,\qquad\forall k\in\mathbb{N}
lim infkn1(B(x,s)(Ek(1)ΔE(1)))=0.\displaystyle\liminf_{k\to\infty}\operatorname{\mathcal{H}^{n-1}}(\partial\textbf{{B}}(x,s)\cap(E_{k}^{(1)}\Delta E^{(1)}))=0. (4.106)

This holds for a.e. s(0,s0)s\in(0,s_{0}). Then EkΔFkB(x,s)UB(x,s0)E_{k}\Delta F_{k}\subset\textbf{{B}}(x,s)\subset\subset U\cap\textbf{{B}}(x,s_{0}) with xUx\in U and s0<rks_{0}<r_{k} for all kk larger than some k0k_{0}. By the same argument as with GG above to prove (4.95), for such kk there holds

Ak(Ek;B(x,s))Ak(E;B(x,s))+Λ1/2n1((E(1)ΔEk(1))B(x,s))+κsα+n1.\displaystyle\mathscr{F}_{A_{k}}(E_{k};\textbf{{B}}(x,s))\leq\mathscr{F}_{A_{k}}(E;\textbf{{B}}(x,s))+\Lambda^{1/2}\operatorname{\mathcal{H}^{n-1}}((E^{(1)}\Delta E_{k}^{(1)})\cap\partial\textbf{{B}}(x,s))+\kappa s^{\alpha+n-1}. (4.107)

Since B(x,s)V\textbf{{B}}(x,s)\subset\subset V,

Φk(B(x,s))=Ak(VEk;B(x,s))=Ak(Ek;B(x,s)).\displaystyle\Phi_{k}(\textbf{{B}}(x,s))=\mathscr{F}_{A_{k}}(V\cap E_{k};\textbf{{B}}(x,s))=\mathscr{F}_{A_{k}}(E_{k};\textbf{{B}}(x,s)). (4.108)

Sending kk\to\infty and using the lower semicontinuity of weak convergent Radon measures, we have by the same reasoning as for (4.5) that

Ψ(B(x,s))\displaystyle\Psi(\textbf{{B}}(x,s)) lim infkΦk(B(x,s))\displaystyle\leq\liminf_{k\to\infty}\Phi_{k}(\textbf{{B}}(x,s))
lim infk(Ak(E;B(x,s))+Λ1/2n1(B(x,s)(Ek(1)ΔE(1)))+κsα+n1)\displaystyle\leq\liminf_{k\to\infty}(\mathscr{F}_{A_{k}}(E;\textbf{{B}}(x,s))+\Lambda^{1/2}\operatorname{\mathcal{H}^{n-1}}(\partial\textbf{{B}}(x,s)\cap(E_{k}^{(1)}\Delta E^{(1)}))+\kappa s^{\alpha+n-1})
=lim supkAk(E;B(x,s))+Λ1/2lim infhn1(B(x,s)(Ek(1)ΔE(1)))+κsα+n1\displaystyle=\limsup_{k\to\infty}\mathscr{F}_{A_{k}}(E;\textbf{{B}}(x,s))+\Lambda^{1/2}\liminf_{h\to\infty}\operatorname{\mathcal{H}^{n-1}}(\partial\textbf{{B}}(x,s)\cap(E_{k}^{(1)}\Delta E^{(1)}))+\kappa s^{\alpha+n-1}
A(E;B(x,s))+Clim supkAkAsupB(x,s)P(E;B(x,s))+κsα+n1\displaystyle\leq\mathscr{F}_{A}(E;\textbf{{B}}(x,s))+C\limsup_{k\to\infty}||A_{k}-A||_{\sup\textbf{{B}}(x,s)}\>P(E;\textbf{{B}}(x,s))+\kappa s^{\alpha+n-1}
=Φ(B(x,s))+κsα+n1.\displaystyle=\Phi(\textbf{{B}}(x,s))+\kappa s^{\alpha+n-1}. (4.109)

The lower perimeter bound (4.72) and comparability to perimeter give csn1Φ(B(x,s))cs^{n-1}\leq\Phi(\textbf{{B}}(x,s)). So by (4.5) and Φ(B(x,s))Ψ(B(x,s))\Phi(\textbf{{B}}(x,s))\leq\Psi(\textbf{{B}}(x,s)), which we know because ΦΨ\Phi\leq\Psi on (n)\mathscr{B}(\mathbb{R}^{n}), it follows that

1CsαΦ(B(x,s))Ψ(B(x,s))1\displaystyle 1-Cs^{\alpha}\leq\frac{\Phi(\textbf{{B}}(x,s))}{\Psi(\textbf{{B}}(x,s))}\leq 1 (4.110)

for a.e. s(0,s0)s\in(0,s_{0}). Sending s0+s\to 0^{+} gives DΨΦ=1D_{\Psi}\Phi=1 for Ψ\Psi-a.e. xVsptΨx\in V\cap\operatorname{spt}\Psi. Since ΦΨ\Phi\ll\Psi, we have that Ψ=Φ\Psi=\Phi on (V)\mathscr{B}(V), the Borel subsets of VV. This completes the proof of our claim.

We finish by showing (i)(i) and (ii)(ii). For (i)(i), suppose xhVEhx_{h}\in V\cap\partial E_{h} and xhxx_{h}\to x for xVx\in V. Let r>0r>0 with B(x,r)V\textbf{{B}}(x,r)\subset\subset V. Then B(xh,r/4)B(x,r/2)\textbf{{B}}(x_{h},r/4)\subset\textbf{{B}}(x,r/2) for large enough hh. So by the weak convergence of the measures Ah(Eh;)A(E;)\mathscr{F}_{A_{h}}(E_{h};\>\cdot\>)\overset{\ast}{\rightharpoonup}\mathscr{F}_{A}(E;\>\cdot\>) in VV, the lower perimeter bound of (4.72), and λ1/2P(Eh;B(xh,r/4))Ah(Eh;B(xh,r/4))\lambda^{1/2}P(E_{h};\textbf{{B}}(x_{h},r/4))\leq\mathscr{F}_{A_{h}}(E_{h};\textbf{{B}}(x_{h},r/4)), we have

0\displaystyle 0 <crn1lim suphAh(Eh;B(xh,r/4))lim suphAh(Eh;B(x,r/2)¯)\displaystyle<c\>r^{n-1}\leq\limsup_{h\to\infty}\mathscr{F}_{A_{h}}(E_{h};{\textbf{{B}}(x_{h},r/4)})\leq\limsup_{h\to\infty}\mathscr{F}_{A_{h}}(E_{h};\overline{\textbf{{B}}(x,r/2)})
A(E;B(x,r/2)¯)Λ1/2P(E;B(x,r)).\displaystyle\leq\mathscr{F}_{A}(E;\overline{\textbf{{B}}(x,r/2)})\leq\Lambda^{1/2}P(E;{\textbf{{B}}(x,r)}). (4.111)

Hence xsptμE=Ex\in\operatorname{spt}\mu_{E}=\partial E. For (ii)(ii), suppose xVEx\in V\cap\partial E and by way of contradiction that there does not exist a sequence {xh}h\{x_{h}\}_{h\in\mathbb{N}} with xhVEhx_{h}\in V\cap\partial E_{h} and xhxx_{h}\to x. Then there is some r>0r>0 and h(k)h(k)\to\infty as kk\to\infty such that B(x,r)V\textbf{{B}}(x,r)\subset\subset V and B(x,r)Eh(k)=\textbf{{B}}(x,r)\cap\partial E_{h(k)}=\varnothing for every kk\in\mathbb{N}. It follows that

P(E;B(x,r))\displaystyle P(E;\textbf{{B}}(x,r)) λ1/2A(E;B(x,r))λ1/2lim infkAh(k)(Eh(k);B(x,r))\displaystyle\leq\lambda^{-1/2}\mathscr{F}_{A}(E;\textbf{{B}}(x,r))\leq\lambda^{-1/2}\liminf_{k\to\infty}\mathscr{F}_{A_{h(k)}}(E_{h(k)};\textbf{{B}}(x,r))
(Λ/λ)1/2lim infkP(Eh(k);B(x,r))=0,\displaystyle\leq(\Lambda/\lambda)^{1/2}\liminf_{k\to\infty}P(E_{h(k)};\textbf{{B}}(x,r))=0, (4.112)

contradicting the fact that xsptμE=Ex\in\operatorname{spt}\mu_{E}=\partial E. ∎

5. The Excess and the Height Bound

The concept of the excess is a common key tool in the study of regularity for minimizers for many geometric variational problems. This quantity measures the average L2L^{2}-oscillation of outward unit normal vector νE\nu_{E} with respect to a fixed direction ν\nu and will eventually allow us to control the average L2L^{2}-oscillation of νE\nu_{E} from its average. Our aim is to show decay estimates for the excess of almost-minimizers. For our variable coefficient surface energies and the change of variable, it will be useful to measure this oscillation over balls, ellipsoids, and cylinders.

5.1. Definition of the excess and basic properties

Given νSSn1\nu\in\SS^{n-1} we decompose n\mathbb{R}^{n} into n1×\mathbb{R}^{n-1}\times\mathbb{R} by identifying n1\mathbb{R}^{n-1} with ν\nu^{\perp} and \mathbb{R} with spanν\text{span}\ \nu. With a slight abuse of notation, we write x=(px,qx)x=(\operatorname{\textnormal{{p}}}x,\operatorname{\textnormal{{q}}}x) where p:nn1\operatorname{\textnormal{{p}}}\mathrel{\mathop{\mathchar 58\relax}}\mathbb{R}^{n}\to\mathbb{R}^{n-1} and q:n\operatorname{\textnormal{{q}}}\mathrel{\mathop{\mathchar 58\relax}}\mathbb{R}^{n}\to\mathbb{R} are the horizontal and vertical projections defined by

px=x(xν)νandqx=xν.\displaystyle\operatorname{\textnormal{{p}}}x=x-(x\cdot\nu)\>\nu\qquad\text{and}\qquad\operatorname{\textnormal{{q}}}x=x\cdot\nu. (5.1)

We define the open cylinder centered at x0nx_{0}\in\mathbb{R}^{n} of radius r>0r>0 in the direction νSSn1\nu\in\SS^{n-1} by

C(x0,r,ν)={xn:|p(xx0)|<r,|q(xx0)|<r}.\displaystyle\textbf{{C}}(x_{0},r,\nu)=\big{\{}x\in\mathbb{R}^{n}\mathrel{\mathop{\mathchar 58\relax}}|\operatorname{\textnormal{{p}}}(x-x_{0})|<r,\ |\operatorname{\textnormal{{q}}}(x-x_{0})|<r\big{\}}. (5.2)

Note that balls and cylinders are comparable as we have

B(x0,r)C(x0,r,ν)B(x0,2r)\displaystyle\textbf{{B}}(x_{0},r)\subset\textbf{{C}}(x_{0},r,\nu)\subset\textbf{{B}}(x_{0},\sqrt{2}\>r) (5.3)

and we have by (4.7) that balls and the ellipsoids Wx0(x0,r)\textbf{{W}}_{x_{0}}(x_{0},r) are comparable. Thus balls, ellipsoids, and cylinders can all be mutually contained in each other by shrinking or enlarging them by fixed scales.

Given a set of locally finite perimeter EE, a point x0Ex_{0}\in\partial E, a radius rr and direction νSSn1\nu\in\SS^{n-1}, we define the spherical excess by

eB(E,x0,r,ν)=1rn1B(x0,r)E|νE(x)ν|22dn1(x),\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{B}}}(E,x_{0},r,\nu)=\frac{1}{r^{n-1}}\int_{\textbf{{B}}(x_{0},r)\cap\partial^{*}\!E}\frac{|\nu_{E}(x)-\nu|^{2}}{2}\operatorname{\,d\mathcal{H}^{n-1}}(x), (5.4)

the ellipsoidal excess by

eW(E,x0,r,ν)=1rn1Wx0(x0,r)E|νE(x)ν|22dn1(x),\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{W}}}(E,x_{0},r,\nu)=\frac{1}{r^{n-1}}\int_{\textbf{{W}}_{x_{0}}(x_{0},r)\cap\partial^{*}\!E}\frac{|\nu_{E}(x)-\nu|^{2}}{2}\operatorname{\,d\mathcal{H}^{n-1}}(x), (5.5)

and the cylindrical excess by

eC(E,x0,r,ν)=1rn1C(x0,r,ν)E|νE(x)ν|22dn1(x).\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},r,\nu)=\frac{1}{r^{n-1}}\int_{\textbf{{C}}(x_{0},r,\nu)\cap\partial^{*}\!E}\frac{|\nu_{E}(x)-\nu|^{2}}{2}\operatorname{\,d\mathcal{H}^{n-1}}(x). (5.6)

Since balls, ellipsoids, and cylinders are comparable, if we can control one of these types of excess, we can control all of them.

As mentioned in Section 4, it will often be convenient to prove estimates at points x0Ex_{0}\in\partial E with the assumption A(x0)=IA(x_{0})=I. To do this, we make the change of variable under the transformation Tx0T_{x_{0}}. The next proposition shows that the excess of the image set under this transformation is comparable to that of the original set.

Proposition 5.1 (Comparability of excess under change of variable Tx0T_{x_{0}}).

There exists a positive constant C=C(n,λ,Λ)C=C(n,\lambda,\Lambda) with the following property. If EE is a set of locally finite perimeter, Ex0=Tx0(E)E_{x_{0}}=T_{x_{0}}(E) for some x0Ex_{0}\in\partial E, then for any r>0r>0 and νSSn1\nu\in\SS^{n-1},

C1eB(Ex0,x0,r,ν~)eW(E,x0,r,ν)CeB(Ex0,x0,r,ν~)\displaystyle C^{-1}\operatorname{\textnormal{{e}}}_{\textbf{{B}}}(E_{x_{0}},x_{0},r,\widetilde{\nu})\leq\operatorname{\textnormal{{e}}}_{\textbf{{W}}}(E,x_{0},r,\nu)\leq C\operatorname{\textnormal{{e}}}_{\textbf{{B}}}(E_{x_{0}},x_{0},r,\widetilde{\nu}) (5.7)

where ν~SSn1\widetilde{\nu}\in\SS^{n-1} is defined by

ν~=A1/2(x0)ν|A1/2(x0)ν|, or equivalently ν=A1/2(x0)ν~|A1/2(x0)ν~|.\displaystyle\widetilde{\nu}=\frac{A^{1/2}(x_{0})\nu}{|A^{1/2}(x_{0})\nu|},\text{ or equivalently }\nu=\frac{A^{-1/2}(x_{0})\widetilde{\nu}}{|A^{-1/2}(x_{0})\widetilde{\nu}|}. (5.8)
Proof.

Without loss of generality assume x0=0x_{0}=0 and to simplify notation write S=A1/2(0)S=A^{1/2}(0), Wr=W0(0,r)\textbf{{W}}_{r}=\textbf{{W}}_{0}(0,r), and Br=B(0,r)\textbf{{B}}_{r}=\textbf{{B}}(0,r). Noting that SS is symmetric, the change of variable y=T0(x)=S1xy=T_{0}(x)=S^{-1}x gives by Proposition A.2 that νE0(T0(x))=SνE(x)/|SνE(x)| for all xE\nu_{E_{0}}(T_{0}(x))=S\nu_{E}(x)/|S\nu_{E}(x)|\text{ for all }x\in\partial^{*}\!E and

BrE0|νE0(y)ν~|2dn1(y)=WrE|SνE(x)|SνE(x)|Sν|Sν||2detS1|SνE(x)|dn1(x).\displaystyle\int_{\textbf{{B}}_{r}\cap\partial^{*}\!E_{0}}|\nu_{E_{0}}(y)-\widetilde{\nu}|^{2}\operatorname{\,d\mathcal{H}^{n-1}}(y)=\int_{\textbf{{W}}_{r}\cap\partial^{*}\!E}\bigg{|}\frac{S\nu_{E}(x)}{|S\nu_{E}(x)|}-\frac{S\nu}{|S\nu|}\bigg{|}^{2}\det S^{-1}\>|S\nu_{E}(x)|\operatorname{\,d\mathcal{H}^{n-1}}(x). (5.9)

Note that detS1=detA1/2(0)λn/2\det S^{-1}=\det A^{-1/2}(0)\leq\lambda^{-n/2}, |SνE(x)|Λ1/2|S\nu_{E}(x)|\leq\Lambda^{1/2}, and

|SνE|SνE|Sν|Sν||22|SνE|SνE|Sν|SνE||2+2|Sν|SνE|Sν|Sν||2.\displaystyle\bigg{|}\frac{S\nu_{E}}{|S\nu_{E}|}-\frac{S\nu}{|S\nu|}\bigg{|}^{2}\leq 2\bigg{|}\frac{S\nu_{E}}{|S\nu_{E}|}-\frac{S\nu}{|S\nu_{E}|}\bigg{|}^{2}+2\bigg{|}\frac{S\nu}{|S\nu_{E}|}-\frac{S\nu}{|S\nu|}\bigg{|}^{2}. (5.10)

For the first term, we have the estimate

|SνE|SνE|Sν|SνE||21|SνE|2|S(νEν)|2Λλ|νEν|2\displaystyle\bigg{|}\frac{S\nu_{E}}{|S\nu_{E}|}-\frac{S\nu}{|S\nu_{E}|}\bigg{|}^{2}\leq\frac{1}{|S\nu_{E}|^{2}}|S(\nu_{E}-\nu)|^{2}\leq\frac{\Lambda}{\lambda}|\nu_{E}-\nu|^{2} (5.11)

since the maximum eigenvalue of SS is bounded by Λ1/2\Lambda^{1/2} and its minimum eigenvalue is bounded by λ1/2\lambda^{1/2}. For the second term, we have the estimate

|Sν|SνE|Sν|Sν||2\displaystyle\bigg{|}\frac{S\nu}{|S\nu_{E}|}-\frac{S\nu}{|S\nu|}\bigg{|}^{2} |Sν|2|1|SνE|1|Sν||2=|Sν|2|SνE|2|Sν|2||Sν||SνE||2\displaystyle\leq|S\nu|^{2}\bigg{|}\frac{1}{|S\nu_{E}|}-\frac{1}{|S\nu|}\bigg{|}^{2}=\frac{|S\nu|^{2}}{|S\nu_{E}|^{2}|S\nu|^{2}}\bigg{|}|S\nu|-|S\nu_{E}|\bigg{|}^{2}
1|SνE|2|S(ννE)|2Λλ|νEν|2\displaystyle\leq\frac{1}{|S\nu_{E}|^{2}}|S(\nu-\nu_{E})|^{2}\leq\frac{\Lambda}{\lambda}|\nu_{E}-\nu|^{2} (5.12)

as above. It follows that

|SνE|SνE|Sν|Sν||24Λλ|νEν|2.\displaystyle\bigg{|}\frac{S\nu_{E}}{|S\nu_{E}|}-\frac{S\nu}{|S\nu|}\bigg{|}^{2}\leq\frac{4\Lambda}{\lambda}|\nu_{E}-\nu|^{2}. (5.13)

Hence

BrE0|νE0ν~|2dn14Λ3/2λn/2+1WrE|νEν|2dn1,\displaystyle\int_{\textbf{{B}}_{r}\cap\partial^{*}\!E_{0}}|\nu_{E_{0}}-\widetilde{\nu}|^{2}\operatorname{\,d\mathcal{H}^{n-1}}\leq\frac{4\Lambda^{3/2}}{\lambda^{n/2+1}}\int_{\textbf{{W}}_{r}\cap\partial^{*}\!E}|\nu_{E}-\nu|^{2}\operatorname{\,d\mathcal{H}^{n-1}}, (5.14)

or equivalently, eB(E0,0,r,ν~)(4Λn/2+1/λ3/2)eW(E,0,r,ν)\operatorname{\textnormal{{e}}}_{\textbf{{B}}}(E_{0},0,r,\widetilde{\nu})\leq(4\Lambda^{n/2+1}/\lambda^{3/2})\operatorname{\textnormal{{e}}}_{\textbf{{W}}}(E,0,r,\nu). The upper bound for (5.7) follows by a symmetric argument. ∎

We now recall several known properties of the excess, referring readers to [Mag12, Chapter 22] for proofs of these facts.

Proposition 5.2 (Scaling of the excess).

If EE is a set of locally finite perimeter in n\mathbb{R}^{n}, x0Ex_{0}\in\partial E, r>0r>0, νSSn1\nu\in\SS^{n-1}, then

eC(E,x0,r,ν)=eC(Ex0,r,0,1,ν)\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},r,\nu)=\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E_{x_{0},r},0,1,\nu) (5.15)

where Ex0,r=(Ex0)/rE_{x_{0},r}=(E-x_{0})/r as in Section 4.

Proposition 5.3 (Zero excess implies being a half-space).

If EE is a set of locally finite perimeter in n\mathbb{R}^{n}, with sptμE=E\operatorname{spt}\mu_{E}=\partial E, x0Ex_{0}\in\partial E, r>0r>0, and νSSn2\nu\in\SS^{n-2}, then eC(E,x0,r,ν)=0\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},r,\nu)=0 if and only if EC(x0,r,ν)E\cap\textbf{{C}}(x_{0},r,\nu) is equivalent to the set {xC(x0,r,ν):(xx0)ν0}\big{\{}x\in\textbf{{C}}(x_{0},r,\nu)\mathrel{\mathop{\mathchar 58\relax}}(x-x_{0})\cdot\nu\leq 0\big{\}}.

Proposition 5.4 (Vanishing of the excess at the reduced boundary).

If EE is a set of locally finite perimeter in n\mathbb{R}^{n} and x0Ex_{0}\in\partial^{*}\!E, then

limr0+infνSSn1eC(E,x0,r,ν)=0.\displaystyle\lim_{r\to 0^{+}}\inf_{\nu\in\SS^{n-1}}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},r,\nu)=0. (5.16)
Proposition 5.5 (Excess at different scales).

If EE is a set of locally finite perimeter in n\mathbb{R}^{n}, x0Ex_{0}\in\partial E, 0<s<r0<s<r, νSSn1\nu\in\SS^{n-1}, then

eC(E,x0,s,ν)(rs)n1eC(E,x0,r,ν).\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},s,\nu)\leq\Big{(}\frac{r}{s}\Big{)}^{n-1}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},r,\nu). (5.17)
Proposition 5.6 (Excess and changes of direction).

For every n2n\geq 2, there exists a constant
C=C(n,λ,Λ,κ,α,r0)C=C(n,\lambda,\Lambda,\kappa,\alpha,r_{0}) with the following property. If EE is a (κ,α)(\kappa,\alpha)-almost-minimizer of A\mathscr{F}_{A} in UU at scale r0r_{0}, then

eC(E,x0,r,ν)C(eC(E,x0,r,ν0)+|νν0|2)\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},r,\nu)\leq C\big{(}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},r,\nu_{0})+|\nu-\nu_{0}|^{2}\big{)} (5.18)

whenever x0UEx_{0}\in U\cap\partial E, B(x,2r)U\textbf{{B}}(x,2r)\subset\subset U, ν,ν0SSn1\nu,\nu_{0}\in\SS^{n-1}.

Proof.

It follows from the proof of [Mag12, Proposition 22.5] using the upper density estimate of Proposition 4.10. ∎

5.2. Small-Excess Position and the Height Bound

We now recall some standard lemmas we will need about the excess and almost-minimizers and recall the height bound. The first lemma states that if the excess of an almost-minimizer in a given cylinder is small enough, then in a smaller cylinder the topological boundary sits within a narrow strip.

Lemma 5.7 (Small-excess position).

Given n2n\geq 2 and t0(0,1)t_{0}\in(0,1), there is a positive constant ω=ω(t0,n,λ,Λ,κ,α,r0)\omega=\omega(t_{0},n,\lambda,\Lambda,\kappa,\alpha,r_{0}) with the following property. If EE is a (κ,α)(\kappa,\alpha)-almost-minimizer of A\mathscr{F}_{A} in C(x0,2r,ν)\textbf{{C}}(x_{0},2r,\nu) with x0Ex_{0}\in\partial E, 2r<r02r<r_{0}, νSSn2\nu\in\SS^{n-2}, and

eC(E,x0,2r,ν)ω,\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},2r,\nu)\leq\omega, (5.19)

then

|q(xx0)|r<t0,xC(x0,r,ν)E,\displaystyle\frac{|\operatorname{\textnormal{{q}}}(x-x_{0})|}{r}<t_{0},\qquad\forall x\in\textbf{{C}}(x_{0},r,\nu)\cap\partial E, (5.20)
|{xC(x0,r,ν)E:q(xx0)r>t0}|=0, and\displaystyle\Big{|}\Big{\{}x\in\textbf{{C}}(x_{0},r,\nu)\cap E\mathrel{\mathop{\mathchar 58\relax}}\frac{\operatorname{\textnormal{{q}}}(x-x_{0})}{r}>t_{0}\Big{\}}\Big{|}=0,\text{ and } (5.21)
|{xC(x0,r,ν)E:q(xx0)r<t0}|=0.\displaystyle\Big{|}\Big{\{}x\in\textbf{{C}}(x_{0},r,\nu)\setminus E\mathrel{\mathop{\mathchar 58\relax}}\frac{\operatorname{\textnormal{{q}}}(x-x_{0})}{r}<-t_{0}\Big{\}}\Big{|}=0. (5.22)
Proof.

[BEG+19, Lemma 3.8] which applies by Proposition 4.10. ∎

We define the open disk in n1\mathbb{R}^{n-1} centered at zn1z\in\mathbb{R}^{n-1} and of radius r>0r>0 by

D(z,r)={wn1:|zw|<r},\displaystyle\textbf{{D}}(z,r)=\big{\{}w\in\mathbb{R}^{n-1}\mathrel{\mathop{\mathchar 58\relax}}|z-w|<r\big{\}}, (5.23)

Thus we may write C(x0,r,ν)=D(px0,r)×(r,r)\textbf{{C}}(x_{0},r,\nu)=\textbf{{D}}(\operatorname{\textnormal{{p}}}x_{0},r)\times(-r,r).

The second lemma states that if the a set of locally finite perimeter satisfies the separation property of Lemma 5.7, then the difference of measure of perimeter sitting above a set GD(px0,r)G\subset\textbf{{D}}(\operatorname{\textnormal{{p}}}x_{0},r) and n1(G)\operatorname{\mathcal{H}^{n-1}}(G) defines a measure which we call the excess measure.

Lemma 5.8 (Excess measure).

If EE is a set of locally finite perimeter in n\mathbb{R}^{n}, with 0E0\in\partial E, and such that, for some t0(0,1)t_{0}\in(0,1),

|q(xx0)|r<t0,xC(x0,r,ν)E,\displaystyle\frac{|\operatorname{\textnormal{{q}}}(x-x_{0})|}{r}<t_{0},\qquad\forall x\in\textbf{{C}}(x_{0},r,\nu)\cap\partial E, (5.24)
|{xC(x0,r,ν)E:q(xx0)r>t0}|=0, and\displaystyle\Big{|}\Big{\{}x\in\textbf{{C}}(x_{0},r,\nu)\cap E\mathrel{\mathop{\mathchar 58\relax}}\frac{\operatorname{\textnormal{{q}}}(x-x_{0})}{r}>t_{0}\Big{\}}\Big{|}=0,\text{ and } (5.25)
|{xC(x0,r,ν)E:q(xx0)r<t0}|=0,\displaystyle\Big{|}\Big{\{}x\in\textbf{{C}}(x_{0},r,\nu)\setminus E\mathrel{\mathop{\mathchar 58\relax}}\frac{\operatorname{\textnormal{{q}}}(x-x_{0})}{r}<-t_{0}\Big{\}}\Big{|}=0, (5.26)

then, setting for brevity M=C(x0,r,ν)EM=\textbf{{C}}(x_{0},r,\nu)\cap\partial^{*}\!E, we have for every Borel set GD(px0,r)G\subset\textbf{{D}}(\operatorname{\textnormal{{p}}}x_{0},r), function φCc(D(px0,r))\varphi\in C_{c}(\textbf{{D}}(\operatorname{\textnormal{{p}}}x_{0},r)), and t(1,1)t\in(-1,1) that

n1(G)n1(Mp1(G)),\displaystyle\operatorname{\mathcal{H}^{n-1}}(G)\leq\operatorname{\mathcal{H}^{n-1}}(M\cap\operatorname{\textnormal{{p}}}^{-1}(G)), (5.27)
n1(G)=Mp1(G)(νEν)dn1(x),\displaystyle\operatorname{\mathcal{H}^{n-1}}(G)=\int_{M\cap\operatorname{\textnormal{{p}}}^{-1}(G)}(\nu_{E}\cdot\nu)\operatorname{\,d\mathcal{H}^{n-1}}(x), (5.28)
Dφ𝑑x=Mφ(px)(νE(x)ν)dn1(x),\displaystyle\int_{\textbf{{D}}}\varphi\>dx=\int_{M}\varphi(\operatorname{\textnormal{{p}}}x)(\nu_{E}(x)\cdot\nu)\operatorname{\,d\mathcal{H}^{n-1}}(x), (5.29)
EtDφ𝑑x=M{qx>t}φ(px)(νE(x)ν)dn1(x),\displaystyle\int_{E_{t}\cap\textbf{{D}}}\varphi\>dx=\int_{M\cap\{\operatorname{\textnormal{{q}}}x>t\}}\varphi(\operatorname{\textnormal{{p}}}x)(\nu_{E}(x)\cdot\nu)\operatorname{\,d\mathcal{H}^{n-1}}(x), (5.30)

where Et={zn1|(z,t)E}E_{t}=\{z\in\mathbb{R}^{n-1}|(z,t)\in E\}. In fact, the set function

ζ(G)\displaystyle\zeta(G) =P(E;C(x0,r,ν)p1(G))n1(G)\displaystyle=P(E;\textbf{{C}}(x_{0},r,\nu)\cap\operatorname{\textnormal{{p}}}^{-1}(G))-\operatorname{\mathcal{H}^{n-1}}(G) (5.31)
=n1(Mp1(G))n1(G)\displaystyle=\operatorname{\mathcal{H}^{n-1}}(M\cap\operatorname{\textnormal{{p}}}^{-1}(G))-\operatorname{\mathcal{H}^{n-1}}(G) (5.32)

defines a Radon measure on n1\mathbb{R}^{n-1}, concentrated on D(px0,r)\textbf{{D}}(\operatorname{\textnormal{{p}}}x_{0},r), called the excess measure of EE over D(px0,r))\textbf{{D}}(\operatorname{\textnormal{{p}}}x_{0},r)).

Proof.

[BEG+19, Theorem A.1] which applies by Proposition 4.10. ∎

We now state the main result we need from this section which is a strengthening of Lemma 5.7 to quantitatively control the height of an almost-minimizer in a cylinder by the excess on a larger cylinder.

Proposition 5.9 (Height bound).

Given n2n\geq 2, there exist positive constants ε0=ε0(n,λ,Λ,κ,α,r0)\varepsilon_{0}=\varepsilon_{0}(n,\lambda,\Lambda,\kappa,\alpha,r_{0}) and C0=C0(n,λ,Λ,κ,α,r0)C_{0}=C_{0}(n,\lambda,\Lambda,\kappa,\alpha,r_{0}) with the following property. If EE is a (κ,α)(\kappa,\alpha)-almost-minimizer of A\mathscr{F}_{A} in C(x0,4r,ν)\textbf{{C}}(x_{0},4r,\nu) at scale r0r_{0} with x0Ex_{0}\in\partial E, 4r<r04r<r_{0}, and

eC(E,x0,4r,ν)ε0,\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},4r,\nu)\leq\varepsilon_{0}, (5.33)

then

sup{|q(xx0)|r:xC(x0,r,ν)E}C0eC(E,x0,4r,ν)1/(2(n1)).\displaystyle\sup\Big{\{}\frac{|\operatorname{\textnormal{{q}}}(x-x_{0})|}{r}\mathrel{\mathop{\mathchar 58\relax}}x\in\textbf{{C}}(x_{0},r,\nu)\cap\partial E\Big{\}}\leq C_{0}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},4r,\nu)^{1/(2(n-1))}. (5.34)
Proof.

[BEG+19, Theorem A.2] which applies by Proposition 4.10. ∎

Throughout the course of the proof of our regularity result, we shall keep track of a number of specific constants for which certain estimates hold. The estimate (5.34) with the constants C0C_{0} and ε0\varepsilon_{0} from Lemma 5.9 are the first of these. Subsequent CiC_{i}’s will be chosen to be larger than previous ones, i.e. C0C1C_{0}\leq C_{1}\leq\dots and subsequent εi\varepsilon_{i}’s will be chosen to be smaller than the previous ones, i.e. ε0ε1\varepsilon_{0}\geq\varepsilon_{1}\geq\dots. This way previous estimates will also hold under any smallness of the excess assumptions. We shall also choose ε0ω(1/4,n,λ,Λ,κ,α,r0)\varepsilon_{0}\leq\omega(1/4,n,\lambda,\Lambda,\kappa,\alpha,r_{0}) so that the height of our topological boundary is at most 1/41/4 of the cylinder.

6. The Lipschitz Approximation Theorem

The next step in our proof is to show that, given a small excess assumption of an almost-minimizer in a cylinder, a large portion of the topological boundary can be covered by the graph of a Lipschitz function in a smaller cylinder. Moreover, if we assume A(x0)=IA(x_{0})=I, this Lipschitz function is quantitatively ‘almost-harmonic’ at x0x_{0} with an error controlled in terms of the excess and the scale. Given a direction νSSn1\nu\in\SS^{n-1} which decomposes n\mathbb{R}^{n} into n1×\mathbb{R}^{n-1}\times\mathbb{R}, we denote the gradient in the first n1n-1 directions by \nabla^{\prime}, that is, =(1,,n1)\nabla^{\prime}=(\partial_{1},\dots,\partial_{n-1}).

Theorem 6.1 (Lipschitz approximation theorem).

There exist positive constants ε1=ε1(n,λ,Λ,κ,α,r0)\varepsilon_{1}=\varepsilon_{1}(n,\lambda,\Lambda,\kappa,\alpha,r_{0}), δ0=δ0(n,λ,Λ,κ,α,r0)\delta_{0}=\delta_{0}(n,\lambda,\Lambda,\kappa,\alpha,r_{0}), and C1=C1(n,λ,Λ,κ,α,r0,ACα)C_{1}=C_{1}(n,\lambda,\Lambda,\kappa,\alpha,r_{0},||A||_{C^{\alpha}}) with the following property. If EE is a (κ,α)(\kappa,\alpha)-almost-minimizer of A\mathscr{F}_{A} in C(x0,r0,ν)\textbf{{C}}(x_{0},r_{0},\nu) at scale r0r_{0} with x0Ex_{0}\in\partial E, 13r<r013r<r_{0}, and

eC(E,x0,13r,ν)ε1,\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},13r,\nu)\leq\varepsilon_{1}, (6.1)

then, setting

M=C(x0,r,ν)E and M0={xM:sup0<s<8reC(E,x,s,ν)δ0},\displaystyle M=\textbf{{C}}(x_{0},r,\nu)\cap\partial E\text{ and }M_{0}=\{x\in M\mathrel{\mathop{\mathchar 58\relax}}\sup_{0<s<8r}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x,s,\nu)\leq\delta_{0}\}, (6.2)

there is a Lipschitz function u:n1u\mathrel{\mathop{\mathchar 58\relax}}\mathbb{R}^{n-1}\to\mathbb{R} with Lipu1\operatorname{Lip}u\leq 1 satisfying

supn1|u|rC1eC(E,x0,13r,ν)1/2(n1)\displaystyle\sup_{\mathbb{R}^{n-1}}\frac{|u|}{r}\leq C_{1}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},13r,\nu)^{1/2(n-1)} (6.3)

such that the translation Γ=x0+{(z,u(z)):zDr}\Gamma=x_{0}+\{(z,u(z))\mathrel{\mathop{\mathchar 58\relax}}z\in\textbf{{D}}_{r}\} of the graph of uu over Dr\textbf{{D}}_{r} contains M0M_{0}, that is, M0MΓM_{0}\subset M\cap\Gamma, and covers a large portion of MM in the sense that

n1(MΔΓ)rn1C1eC(E,x0,13r,ν).\displaystyle\frac{\operatorname{\mathcal{H}^{n-1}}(M\Delta\Gamma)}{r^{n-1}}\leq C_{1}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},13r,\nu). (6.4)

Moreover, uu is ‘almost harmonic’ in Dr\textbf{{D}}_{r} in the sense that

1rn1Dr|u|2C1eC(E,x0,13r,ν)\displaystyle\frac{1}{r^{n-1}}\int_{\textbf{{D}}_{r}}|\nabla^{\prime}u|^{2}\leq C_{1}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},13r,\nu) (6.5)

and if A(x0)=IA(x_{0})=I, then

1rn1|Druφ|C1supDr|φ|(eC(E,x0,13r,ν)+rα/2)φCc1(Dr).\displaystyle\frac{1}{r^{n-1}}\bigg{|}\int_{\textbf{{D}}_{r}}\nabla^{\prime}u\cdot\nabla^{\prime}\varphi\bigg{|}\leq C_{1}\sup_{\textbf{{D}}_{r}}|\nabla^{\prime}\varphi|\big{(}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},13r,\nu)+r^{\alpha/2})\qquad\forall\varphi\in C_{c}^{1}(\textbf{{D}}_{r}). (6.6)
Proof.

Without loss of generality we may assume x0=0x_{0}=0 and ν=en\nu=e_{n}. We simplify notation by setting Cr=C(0,r,ν)\textbf{{C}}_{r}=\textbf{{C}}(0,r,\nu). Everything up to and including (6.5) follows from [BEG+19, Theorem A.3] by Proposition 4.10, for an ε1\varepsilon_{1} chosen sufficiently small. We also choose ε1\varepsilon_{1} small enough so that

ε1ε0ω(1/4,n,λ,Λ,κ,α,r0)\displaystyle\boxed{\varepsilon_{1}\leq\varepsilon_{0}\leq\omega(1/4,n,\lambda,\Lambda,\kappa,\alpha,r_{0})} (6.7)

where ω\omega is the constant from Lemma 5.7 with t0=1/4t_{0}=1/4. It follows that

|qx|r<14xCrE,{xCrE:qxr>14}=, and {xCrE:qxr<14}=.\displaystyle\frac{|\operatorname{\textnormal{{q}}}x|}{r}<\frac{1}{4}\ \ \forall x\in\textbf{{C}}_{r}\cap\partial E,\ \Big{\{}x\in\textbf{{C}}_{r}\cap E\mathrel{\mathop{\mathchar 58\relax}}\frac{\operatorname{\textnormal{{q}}}x}{r}>\frac{1}{4}\Big{\}}=\varnothing,\text{ and }\Big{\{}x\in\textbf{{C}}_{r}\setminus E\mathrel{\mathop{\mathchar 58\relax}}\frac{\operatorname{\textnormal{{q}}}x}{r}<-\frac{1}{4}\Big{\}}=\varnothing. (6.8)

Let φCc1(Dr)\varphi\in\textbf{{C}}_{c}^{1}(D_{r}). By considering φ/supDr|φ|\nabla^{\prime}\varphi/\sup_{\textbf{{D}}_{r}}|\nabla^{\prime}\varphi|, we may assume supDr|φ|=1\sup_{\textbf{{D}}_{r}}|\nabla^{\prime}\varphi|=1 and reduce to proving

1rn1|Druφ|C1(eC(E,x0,13r,en)+rα/2).\displaystyle\frac{1}{r^{n-1}}\bigg{|}\int_{\textbf{{D}}_{r}}\nabla^{\prime}u\cdot\nabla^{\prime}\varphi\bigg{|}\leq C_{1}\big{(}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},13r,e_{n})+r^{\alpha/2}). (6.9)

By the Fundamental Theorem of Calculus and the fact that φ=0\varphi=0 on Dr\partial\textbf{{D}}_{r}, we have supDr|φ|r\sup_{\textbf{{D}}_{r}}|\varphi|\leq r. Let ηCc1((3r/4,3r/4))\eta\in C_{c}^{1}((-3r/4,3r/4)) be a cutoff function such that

η=1 on [r/2,r/2],|η|1, and |η|5/r.\displaystyle\eta=1\text{ on }[-r/2,r/2],\ |\eta|\leq 1,\text{ and }|\eta^{\prime}|\leq 5/r. (6.10)

and define T:nnT\mathrel{\mathop{\mathchar 58\relax}}\mathbb{R}^{n}\to\mathbb{R}^{n} by T(x)=η(qx)φ(px)enT(x)=\eta(\operatorname{\textnormal{{q}}}x)\varphi(\operatorname{\textnormal{{p}}}x)e_{n}. Then TCc1(Dr×(3r/4,3r/4);n)T\in C_{c}^{1}(\textbf{{D}}_{r}\times(-3r/4,3r/4);\mathbb{R}^{n}), supCr|T|r\sup_{\textbf{{C}}_{r}}|T|\leq r, and T(x)=η(qx)φ(px)en+η(qx)φ(px)enen\nabla T(x)=\eta(\operatorname{\textnormal{{q}}}x)\>\nabla^{\prime}\varphi(\operatorname{\textnormal{{p}}}x)\otimes e_{n}+\eta^{\prime}(\operatorname{\textnormal{{q}}}x)\varphi(\operatorname{\textnormal{{p}}}x)\>e_{n}\otimes e_{n}. Hence

|T(x)|=|η(qx)|2|φ(px)|2+|η(qx)|2|φ(px)|26\displaystyle|\nabla T(x)|=\sqrt{|\eta(\operatorname{\textnormal{{q}}}x)|^{2}|\nabla^{\prime}\varphi(\operatorname{\textnormal{{p}}}x)|^{2}+|\eta^{\prime}(\operatorname{\textnormal{{q}}}x)|^{2}|\varphi(\operatorname{\textnormal{{p}}}x)|^{2}}\leq 6 (6.11)

and so supCr|T|6\sup_{\textbf{{C}}_{r}}|\nabla T|\leq 6. Consider the family of maps ft:nnf_{t}\colon\mathbb{R}^{n}\to\mathbb{R}^{n} defined by ft(x)=x+tT(x)f_{t}(x)=x+t\>T(x). Then ft=Id+tT\nabla f_{t}=\operatorname{Id}+t\nabla T and so Jft=det(Id+tT)Jf_{t}=\det(\operatorname{Id}+t\nabla T). We have that T(x)|T(x)|6||\nabla T(x)||\leq|\nabla T(x)|\leq 6 where ||||||\cdot|| denotes the operator norm and |||\cdot| denotes the Frobenius norm. It then follows by [Mag12, Lemma 17.4] that there are positive constants ε(n)\varepsilon(n), C(n)C(n) such that

Jft=(1+tdivT)+O(C(n)t2).\displaystyle Jf_{t}=(1+t\operatorname{div}T)+O(C(n)t^{2}). (6.12)

for |t|<ε(n)|t|<\varepsilon(n). Since divT\operatorname{div}T is bounded, we can choose ε(n)\varepsilon(n) so that ftf_{t} is a diffeomorphism for |t|<ε(n)|t|<\varepsilon(n). Letting gt=ft1g_{t}=f_{t}^{-1}, we also have by [Mag12, Lemma 17.4] that gtft=IdtT+O(C(n)t2\nabla g_{t}\circ f_{t}=\operatorname{Id}-t\nabla T+O(C(n)t^{2} for t<ε(n)t<\varepsilon(n). Choosing ε(n)1/8\varepsilon(n)\leq 1/8, we claim that EΔft(E)CrE\Delta f_{t}(E)\subset\subset\textbf{{C}}_{r} for |t|<ε(n)|t|<\varepsilon(n).

To see why this is the case, take yEΔft(E)y\in E\Delta f_{t}(E). Then y=x+tT(x)y=x+tT(x) for some xsptTx\in\operatorname{spt}T. By definition of TT, py=pxsptφ\operatorname{\textnormal{{p}}}y=\operatorname{\textnormal{{p}}}x\in\operatorname{spt}\varphi and qx(3r/4,3r/4)\operatorname{\textnormal{{q}}}x\in(-3r/4,3r/4). So |qy||qx|+|qyqx|(3/4)r+|t|supCr|T|<7r/8|\operatorname{\textnormal{{q}}}y|\leq|\operatorname{\textnormal{{q}}}x|+|\operatorname{\textnormal{{q}}}y-\operatorname{\textnormal{{q}}}x|\leq(3/4)r+|t|\sup_{\textbf{{C}}_{r}}|T|<7r/8 since |t|<ε(n)1/8|t|<\varepsilon(n)\leq 1/8 and supCr|T|r\sup_{\textbf{{C}}_{r}}|T|\leq r. Hence ysptφ×(7r/8,7r/8)Cry\in\operatorname{spt}\varphi\times(-7r/8,7r/8)\subset\subset\textbf{{C}}_{r}.

By [Mag12, Proposition 17.1] we have that

P(ft(E);Cr)=CrE|(gtft)νE|Jftdn1.\displaystyle P(f_{t}(E);\textbf{{C}}_{r})=\int_{\textbf{{C}}_{r}\cap\partial^{*}\!E}|(\nabla g_{t}\circ f_{t})^{*}\nu_{E}|Jf_{t}\operatorname{\,d\mathcal{H}^{n-1}}. (6.13)
Claim.

We can choose ε(n)\varepsilon(n) small enough so that

P(ft(E);Cr)=P(E;Cr)+tCrEdivET(x)dn1+O(C(n)P(E;Cr)t2)\displaystyle P(f_{t}(E);\textbf{{C}}_{r})=P(E;\textbf{{C}}_{r})+t\int_{\textbf{{C}}_{r}\cap\partial^{*}\!E}\operatorname{div}_{E}T(x)\operatorname{\,d\mathcal{H}^{n-1}}+O(C(n)P(E;\textbf{{C}}_{r})t^{2}) (6.14)

for all |t|<ε(n)|t|<\varepsilon(n), where divET=divTνE(T)νE\operatorname{div}_{E}T=\operatorname{div}T-\nu_{E}\cdot(\nabla T)^{*}\nu_{E}.

To prove the claim, observe that |(Id+tT)νE|2=12tνE(T)νE+t2|(T)νE|2\big{|}(\operatorname{Id}+t\nabla T)^{*}\nu_{E}\big{|}^{2}=1-2t\>\nu_{E}\cdot(\nabla T)^{*}\nu_{E}+t^{2}|(\nabla T)^{*}\nu_{E}|^{2} and so, since 1+x=1x/2+O(x2)\sqrt{1+x}=1-x/2+O(x^{2}) for small |x||x| by Taylor’s theorem, shrinking ε(n)\varepsilon(n) as necessary, we have

|(gtft)νE|=|(IdtT)νE|+O(C(n)t2)=1tνE(T)νE+O(C(n)t2)\displaystyle\big{|}(\nabla g_{t}\circ f_{t})^{*}\nu_{E}\big{|}=\big{|}(\operatorname{Id}-t\nabla T)^{*}\nu_{E}\big{|}+O(C(n)t^{2})=1-t\>\nu_{E}\cdot(\nabla T)^{*}\nu_{E}+O(C(n)t^{2}) (6.15)

whenever |t|<ε(n)|t|<\varepsilon(n). Combining this with (6.12) gives

|(gtft)νE|Jft=1+t(divTνE(T)νE)+O(C(n)t2)\displaystyle\big{|}(\nabla g_{t}\circ f_{t})^{*}\nu_{E}\big{|}Jf_{t}=1+t(\operatorname{div}T-\>\nu_{E}\cdot(\nabla T)^{*}\nu_{E})+O(C(n)t^{2}) (6.16)

for |t|<ε(n)|t|<\varepsilon(n). Integrating with respect to n1  (CrE)\operatorname{\mathcal{H}^{n-1}}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>(\textbf{{C}}_{r}\cap\partial^{*}\!E) completes the proof of the claim.

By the claim, Proposition 4.10, and Lemma 4.5, it follows that

|t||CrEdivET(x)dn1|\displaystyle|t|\>\Big{|}\int_{\textbf{{C}}_{r}\cap\partial^{*}\!E}\operatorname{div}_{E}T(x)\operatorname{\,d\mathcal{H}^{n-1}}\Big{|} |P(ft(E;Cr))P(E;Cr))|+C(n)P(E;Cr)|t|2\displaystyle\leq\big{|}P(f_{t}(E;\textbf{{C}}_{r}))-P(E;\textbf{{C}}_{r}))\big{|}+C(n)P(E;\textbf{{C}}_{r})|t|^{2}
Crα+n1+Ct2rn1\displaystyle\leq Cr^{\alpha+n-1}+Ct^{2}r^{n-1} (6.17)

whenever |t|<ε(n)|t|<\varepsilon(n). Choosing t=ε(n)(r/r0)α/2<ε(n)t=\varepsilon(n)(r/r_{0})^{\alpha/2}<\varepsilon(n) gives that

|ECrdivET(x)dn1|Crα/2+n1.\displaystyle\Big{|}\int_{\partial^{*}\!E\cap\textbf{{C}}_{r}}\operatorname{div}_{E}T(x)\operatorname{\,d\mathcal{H}^{n-1}}\Big{|}\leq Cr^{\alpha/2+n-1}. (6.18)

Now, for n1\operatorname{\mathcal{H}^{n-1}}-a.e. xMΓx\in M\cap\Gamma, there is λ(x){1,1}\lambda(x)\in\{-1,1\} such that

νE(x)=λ(x)(u(px),1)1+|u(px)|2.\displaystyle\nu_{E}(x)=\lambda(x)\frac{(-\nabla^{\prime}u(\operatorname{\textnormal{{p}}}x),1)}{\sqrt{1+|\nabla^{\prime}u(\operatorname{\textnormal{{p}}}x)|^{2}}}. (6.19)

By (6.8) and definition of η\eta, we have η(qx)=1\eta(\operatorname{\textnormal{{q}}}x)=1 on a neighborhood MM and so divT(x)=0\operatorname{div}T(x)=0 and T(x)=φ(px)en\nabla T(x)=\nabla^{\prime}\varphi(\operatorname{\textnormal{{p}}}x)\otimes e_{n} for xMx\in M. Hence for n1\operatorname{\mathcal{H}^{n-1}}-a.e. xMΓx\in M\cap\Gamma, there holds

divET(x)=divT(x)νE(x)((T(x))νE(x))=u(px)φ(px)1+|u(px)|2\displaystyle\operatorname{div}_{E}T(x)=\operatorname{div}T(x)-\nu_{E}(x)\cdot((\nabla T(x))^{*}\nu_{E}(x))=\frac{\nabla^{\prime}u(\operatorname{\textnormal{{p}}}x)\cdot\nabla^{\prime}\varphi(\operatorname{\textnormal{{p}}}x)}{1+|\nabla^{\prime}u(\operatorname{\textnormal{{p}}}x)|^{2}} (6.20)

since λ(x)2=1\lambda(x)^{2}=1. Thus

|p(MΓ)uφ1+|u|2|\displaystyle\bigg{|}\int_{\operatorname{\textnormal{{p}}}(M\cap\Gamma)}\frac{\nabla^{\prime}u\cdot\nabla^{\prime}\varphi}{\sqrt{1+|\nabla^{\prime}u|^{2}}}\bigg{|} =|MΓdivETdn1|\displaystyle=\bigg{|}\int_{M\cap\Gamma}\operatorname{div}_{E}T\operatorname{\,d\mathcal{H}^{n-1}}\bigg{|}
|MdivETdn1|+|MΓdivETdn1|\displaystyle\leq\bigg{|}\int_{M}\operatorname{div}_{E}T\operatorname{\,d\mathcal{H}^{n-1}}\bigg{|}+\bigg{|}\int_{M\setminus\Gamma}\operatorname{div}_{E}T\operatorname{\,d\mathcal{H}^{n-1}}\bigg{|}
C(rα/2+n1+n1(MΓ))\displaystyle\leq C(r^{\alpha/2+n-1}+\operatorname{\mathcal{H}^{n-1}}(M\setminus\Gamma)) (6.21)

by (6.18) and since |divET|C(n)|T|C(n)|\operatorname{div}_{E}T|\leq C(n)|\nabla T|\leq C(n). Since Dr=p(Γ)\textbf{{D}}_{r}=\operatorname{\textnormal{{p}}}(\Gamma), it follows that

|Druφ1+|u|2|\displaystyle\bigg{|}\int_{\textbf{{D}}_{r}}\frac{\nabla^{\prime}u\cdot\nabla^{\prime}\varphi}{\sqrt{1+|\nabla^{\prime}u|^{2}}}\bigg{|} |p(MΓ)uφ1+|u|2|+|p(ΓM)uφ1+|u|2|\displaystyle\leq\bigg{|}\int_{\operatorname{\textnormal{{p}}}(M\cap\Gamma)}\frac{\nabla^{\prime}u\cdot\nabla^{\prime}\varphi}{\sqrt{1+|\nabla^{\prime}u|^{2}}}\bigg{|}+\bigg{|}\int_{\operatorname{\textnormal{{p}}}(\Gamma\setminus M)}\frac{\nabla^{\prime}u\cdot\nabla^{\prime}\varphi}{\sqrt{1+|\nabla^{\prime}u|^{2}}}\bigg{|}
C(rα/2+n1+n1(MΓ))+Cn1(ΓM)\displaystyle\leq C(r^{\alpha/2+n-1}+\operatorname{\mathcal{H}^{n-1}}(M\setminus\Gamma))+C\operatorname{\mathcal{H}^{n-1}}(\Gamma\setminus M)
C(n1(MΔΓ)+rα/2+n1)\displaystyle\leq C(\operatorname{\mathcal{H}^{n-1}}(M\Delta\Gamma)+r^{\alpha/2+n-1}) (6.22)

where we used the fact that Lipu1\operatorname{Lip}u\leq 1, |φ|1|\nabla^{\prime}\varphi|\leq 1, and n1(p(ΓM))n1(ΓM)\operatorname{\mathcal{H}^{n-1}}(\operatorname{\textnormal{{p}}}(\Gamma\setminus M))\leq\operatorname{\mathcal{H}^{n-1}}(\Gamma\setminus M). Again, using Lipu1\operatorname{Lip}u\leq 1, |φ|1|\nabla^{\prime}\varphi|\leq 1, we have

Dr|uφuφ1+|u|2|\displaystyle\int_{\textbf{{D}}_{r}}\bigg{|}\nabla^{\prime}u\cdot\nabla^{\prime}\varphi-\frac{\nabla^{\prime}u\cdot\nabla^{\prime}\varphi}{\sqrt{1+|\nabla^{\prime}u|^{2}}}\bigg{|} Dr|u||φ||111+|u|2|\displaystyle\leq\int_{\textbf{{D}}_{r}}|\nabla^{\prime}u|\>|\nabla^{\prime}\varphi|\>\Big{|}1-\frac{1}{\sqrt{1+|\nabla^{\prime}u|^{2}}}\Big{|}
Dr1+|u|211+|u|2\displaystyle\leq\int_{\textbf{{D}}_{r}}\frac{\sqrt{1+|\nabla^{\prime}u|^{2}}-1}{\sqrt{1+|\nabla^{\prime}u|^{2}}}
=Dr|u|21+|u|2(1+|u|2+1)\displaystyle=\int_{\textbf{{D}}_{r}}\frac{|\nabla^{\prime}u|^{2}}{\sqrt{1+|\nabla^{\prime}u|^{2}}(\sqrt{1+|\nabla^{\prime}u|^{2}}+1)}
12Dr|u|2.\displaystyle\leq\frac{1}{2}\int_{\textbf{{D}}_{r}}|\nabla^{\prime}u|^{2}. (6.23)

By (6.4) and (6.5), it follows that

|Druφ|\displaystyle\bigg{|}\int_{\textbf{{D}}_{r}}\nabla^{\prime}u\cdot\nabla^{\prime}\varphi\bigg{|} |Dr(uφuφ1+|u|2)|+|Druφ1+|u|2|\displaystyle\leq\bigg{|}\int_{\textbf{{D}}_{r}}\Big{(}\nabla^{\prime}u\cdot\nabla^{\prime}\varphi-\frac{\nabla^{\prime}u\cdot\nabla^{\prime}\varphi}{\sqrt{1+|\nabla^{\prime}u|^{2}}}\Big{)}\bigg{|}+\bigg{|}\int_{\textbf{{D}}_{r}}\frac{\nabla^{\prime}u\cdot\nabla^{\prime}\varphi}{\sqrt{1+|\nabla^{\prime}u|^{2}}}\bigg{|}
12Dr|u|2+C(n1(MΔΓ)+rα/2+n1)\displaystyle\leq\frac{1}{2}\int_{\textbf{{D}}_{r}}|\nabla^{\prime}u|^{2}+C(\operatorname{\mathcal{H}^{n-1}}(M\Delta\Gamma)+r^{\alpha/2+n-1})
C(eC(E,x0,13r,en)+rα/2)rn1.\displaystyle\leq C(\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},13r,e_{n})+r^{\alpha/2})r^{n-1}. (6.24)

completing the proof. ∎

7. The Reverse Poincaré Inequality

In Section 5 we saw that a small excess controls the height of the topological boundary of an almost-minimizer. In this section we show that given a small excess assumption on a cylinder, the flatness of the topological boundary controls the excess on a smaller cylinder. Recall that the cylindrical flatness of a set of locally finite perimeter EE at a point x0Ex_{0}\in\partial E, radius r>0r>0, in the direction νSSn1\nu\in\SS^{n-1} is defined by

f(E,x0,r,ν)=1rn1infcC(x0,r,ν)E|(xx0)νc|2r2dn1(x).\displaystyle\operatorname{\textnormal{{f}}}(E,x_{0},r,\nu)=\frac{1}{r^{n-1}}\inf_{c\in\mathbb{R}}\int_{\textbf{{C}}(x_{0},r,\nu)\cap\partial^{*}\!E}\frac{|(x-x_{0})\cdot\nu-c|^{2}}{r^{2}}\operatorname{\,d\mathcal{H}^{n-1}}(x). (7.1)

This quantity measures how far in an L2L^{2} sense the boundary of EE is from the best approximating plane with normal ν\nu.

Theorem 7.1 (Reverse Poincaré Inequality).

Given n2n\geq 2, there is a positive constant
C2=C2(n,λ,Λ,κ,α,r0,ACα)C_{2}=C_{2}(n,\lambda,\Lambda,\kappa,\alpha,r_{0},||A||_{C^{\alpha}}) with the following property. If EE is a (κ,α)(\kappa,\alpha)-almost-minimizer in C(x0,4r,ν)\textbf{{C}}(x_{0},4r,\nu) with x0Ex_{0}\in\partial E, A(x0)=IA(x_{0})=I, 4r<r04r<r_{0}, and

eC(E,x0,4r,ν)ω(1/8,n,λ,Λ,κ,α,r0),\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},4r,\nu)\leq\omega(1/8,n,\lambda,\Lambda,\kappa,\alpha,r_{0}), (7.2)

where ω\omega is the constant from Lemma 5.7, then

eC(E,x0,r,ν)C2(f(E,x0,2r,ν)+rα).\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},r,\nu)\leq C_{2}\big{(}\operatorname{\textnormal{{f}}}(E,x_{0},2r,\nu)+r^{\alpha}\big{)}. (7.3)

To prove this we modify the proofs presented in [Mag12, Chapter 24]. First we need several lemmas. Given νSSn1\nu\in\SS^{n-1} and the decomposition of n\mathbb{R}^{n} into n1×\mathbb{R}^{n-1}\times\mathbb{R}, we define the narrow cylinders

K(z,s)=D(z,s)×(1,1)\displaystyle\textbf{{K}}(z,s)=\textbf{{D}}(z,s)\times(-1,1) (7.4)

for zn1z\in\mathbb{R}^{n-1} and s>0s>0.

Lemma 7.2 (Cone-like competitors [Mag12, Lemma 24.8]).

If s>0s>0 and EE is an open set with smooth boundary in n\mathbb{R}^{n} such that

|qx|<14,xK(z,s)E,\displaystyle|\operatorname{\textnormal{{q}}}x|<\frac{1}{4},\qquad\forall x\in\textbf{{K}}(z,s)\cap\partial E,
{xK(z,s):qx<14}K(z,s)E{xK(z,s):qx<14},\displaystyle\Big{\{}x\in\textbf{{K}}(z,s)\mathrel{\mathop{\mathchar 58\relax}}\operatorname{\textnormal{{q}}}x<-\frac{1}{4}\Big{\}}\subset\textbf{{K}}(z,s)\cap E\subset\Big{\{}x\in\textbf{{K}}(z,s)\mathrel{\mathop{\mathchar 58\relax}}\operatorname{\textnormal{{q}}}x<\frac{1}{4}\Big{\}}, (7.5)

then, for every t(0,1/4)t\in(0,1/4) and |c|<1/4|c|<1/4, there exists I(2/3,3/4)I\subset(2/3,3/4) with |I|1/24|I|\geq 1/24 such that for every rIr\in I, there exists an open set FF of locally finite perimeter in n\mathbb{R}^{n}, satisfying,

FK(z,rs)=EK(z,rs),\displaystyle F\cap\partial\textbf{{K}}(z,rs)=E\cap\partial\textbf{{K}}(z,rs), (7.6)
n1(FK(z,rs))=n1(EK(z,rs))=0,\displaystyle\operatorname{\mathcal{H}^{n-1}}(\partial F\cap\partial\textbf{{K}}(z,rs))=\operatorname{\mathcal{H}^{n-1}}(\partial E\cap\partial\textbf{{K}}(z,rs))=0, (7.7)
K(z,s/2)F=D(z,s/2)×{c},\displaystyle\textbf{{K}}(z,s/2)\cap\partial F=\textbf{{D}}(z,s/2)\times\{c\}, (7.8)
P(F;K(z,rs))\displaystyle P(F;\textbf{{K}}(z,rs)) n1(D(z,rs))\displaystyle-\operatorname{\mathcal{H}^{n-1}}(\textbf{{D}}(z,rs))
C(n){t(P(E;K(z,s))n1(D(z,s))+1tK(z,s)E|qxc|2s2dn1(x)}.\displaystyle\leq C(n)\Big{\{}t\big{(}P(E;\textbf{{K}}(z,s))-\operatorname{\mathcal{H}^{n-1}}(\textbf{{D}}(z,s)\big{)}+\frac{1}{t}\int_{\textbf{{K}}(z,s)\cap\partial E}\frac{|\operatorname{\textnormal{{q}}}x-c|^{2}}{s^{2}}\operatorname{\,d\mathcal{H}^{n-1}}(x)\Big{\}}. (7.9)
Proof.

This is proved in [Mag12, Lemma 24.8], though we point out that (7.7) follows by line (24.29) in [Mag12, Lemma 24.8] and the fact that FF is the cone-like extension of EK(z,rs)E\cap\partial\textbf{{K}}(z,rs) over the disk D(z,(1t)rs)×{c}\textbf{{D}}(z,{(1-t)rs})\times\{c\} (see [Mag12, Lemma 24.6]). ∎

Lemma 7.3 (Weak reverse Poincaré inequality).

If EE is a (κ,α)(\kappa,\alpha)-almost-minimizer of A\mathscr{F}_{A} in C4\textbf{{C}}_{4}, A(0)=IA(0)=I, at scale r0>4r_{0}>4, such that

|qx|<18,xC2E,\displaystyle|\operatorname{\textnormal{{q}}}x|<\frac{1}{8},\qquad\forall x\in\textbf{{C}}_{2}\cap\partial E, (7.10)
|{xC2E:qx<18}|=|{xC2E:qx>18}|=0,\displaystyle\Big{|}\Big{\{}x\in\textbf{{C}}_{2}\setminus E\mathrel{\mathop{\mathchar 58\relax}}\operatorname{\textnormal{{q}}}x<-\frac{1}{8}\Big{\}}\Big{|}=\Big{|}\Big{\{}x\in\textbf{{C}}_{2}\cap E\mathrel{\mathop{\mathchar 58\relax}}\operatorname{\textnormal{{q}}}x>\frac{1}{8}\Big{\}}\Big{|}=0, (7.11)

and if zn1z\in\mathbb{R}^{n-1} and s>0s>0 are such that

K(z,s)C2,n1(EK(z,s))=0,\displaystyle\textbf{{K}}(z,s)\subset\textbf{{C}}_{2},\qquad\operatorname{\mathcal{H}^{n-1}}(\partial^{*}\!E\cap\partial\textbf{{K}}(z,s))=0, (7.12)

then, for every |c|<1/4|c|<1/4,

P(E;K(z,s/2))n1(D(z,s/2))\displaystyle P(E;\textbf{{K}}(z,s/2))-\operatorname{\mathcal{H}^{n-1}}(\textbf{{D}}(z,s/2))
C[([P(E;K(z,s))n1(D(z,s))]K(z,s)E(qxc)2s2dn1(x))1/2+κ+ACα]\displaystyle\leq C\bigg{[}\Big{(}\big{[}P(E;\textbf{{K}}(z,s))-\operatorname{\mathcal{H}^{n-1}}(\textbf{{D}}(z,s))\big{]}\int_{\textbf{{K}}(z,s)\cap\partial^{*}\!E}\frac{(\operatorname{\textnormal{{q}}}x-c)^{2}}{s^{2}}\operatorname{\,d\mathcal{H}^{n-1}}(x)\Big{)}^{1/2}+\kappa+||A||_{C^{\alpha}}\bigg{]} (7.13)

where C=C(n,λ,Λ,κ,α,r0)C=C(n,\lambda,\Lambda,\kappa,\alpha,r_{0}).

Proof.

Properties (7.10) and (7.11) imply by the divergence theorem that

ζ(G)=P(E;C2p1(G))n1(G),GD2,\displaystyle\zeta(G)=P(E;\textbf{{C}}_{2}\cap\operatorname{\textnormal{{p}}}^{-1}(G))-\operatorname{\mathcal{H}^{n-1}}(G),\qquad G\subset\textbf{{D}}_{2}, (7.14)

defines a Radon measure on n1\mathbb{R}^{n-1} concentrated on D2\textbf{{D}}_{2} as in Lemma 5.8. By [Mag12, Theorem 13.8], given εh0+\varepsilon_{h}\to 0^{+} there exists a sequence {Eh}h\{E_{h}\}_{h\in\mathbb{N}} of open sets with smooth boundary such that

EhlocE,n1 Ehn1 E,EhIεh(E),\displaystyle E_{h}\overset{\text{loc}}{\rightarrow}E,\qquad\operatorname{\mathcal{H}^{n-1}}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>\>\partial E_{h}\overset{\ast}{\rightharpoonup}\operatorname{\mathcal{H}^{n-1}}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>\>\partial^{*}\!E,\qquad\partial E_{h}\subset I_{\varepsilon_{h}}(\partial E), (7.15)

where Iεh(E)I_{\varepsilon_{h}}(\partial E) denotes the εh\varepsilon_{h}-neighborhood of E\partial E. The coarea formula and Fatou’s lemma give

2/33/4lim infhn1(Krs(E(1)ΔEh))dr\displaystyle\int_{2/3}^{3/4}\liminf_{h\to\infty}\operatorname{\mathcal{H}^{n-1}}(\partial\textbf{{K}}_{rs}\cap(E^{(1)}\Delta E_{h}))dr lim infh2/33/4n1(Krs(E(1)ΔEh))𝑑r\displaystyle\leq\liminf_{h\to\infty}\int_{2/3}^{3/4}\operatorname{\mathcal{H}^{n-1}}(\partial\textbf{{K}}_{rs}\cap(E^{(1)}\Delta E_{h}))dr
limh|(E(1)ΔEh)Bs|=0.\displaystyle\leq\lim_{h\to\infty}|(E^{(1)}\Delta E_{h})\cap B_{s}|=0. (7.16)

So for a.e. r(2/3,3/4)r\in(2/3,3/4), there holds

lim infhn1(Krs(E(1)ΔEh))=0.\displaystyle\liminf_{h\to\infty}\operatorname{\mathcal{H}^{n-1}}(\partial\textbf{{K}}_{rs}\cap(E^{(1)}\Delta E_{h}))=0. (7.17)

Provided that hh is large enough, EhlocEE_{h}\overset{\text{loc}}{\rightarrow}E and EhIεh(E)\partial E_{h}\subset I_{\varepsilon_{h}}(\partial E) imply by (7.10) and (7.11) that

|qx|<14,xC2Eh,\displaystyle|\operatorname{\textnormal{{q}}}x|<\frac{1}{4},\qquad\forall x\in\textbf{{C}}_{2}\cap\partial E_{h}, (7.18)
|{xC2Eh:qx<14}|=|{xC2Eh:qx>14}|=0.\displaystyle\Big{|}\Big{\{}x\in\textbf{{C}}_{2}\setminus E_{h}\mathrel{\mathop{\mathchar 58\relax}}\operatorname{\textnormal{{q}}}x<-\frac{1}{4}\Big{\}}\Big{|}=\Big{|}\Big{\{}x\in\textbf{{C}}_{2}\cap E_{h}\mathrel{\mathop{\mathchar 58\relax}}\operatorname{\textnormal{{q}}}x>\frac{1}{4}\Big{\}}\Big{|}=0. (7.19)

Given t(0,1/4)t\in(0,1/4) and |c|<1/4|c|<1/4 we can apply Lemma 7.2 to each EhE_{h} for zn1z\in\mathbb{R}^{n-1} and s>0s>0 to find the sets Ih(2/3,3/4)I_{h}\subset(2/3,3/4) with |Ih|1/24|I_{h}|\geq 1/24 such that for each rIhr\in I_{h} there exists an open set FhF_{h} satisfying (7.6), (7.7), (7.8), and (7.2). For each hh\in\mathbb{N}, we have the containment khIkkh+1Ik\bigcup_{k\geq h}I_{k}\supset\bigcup_{k\geq h+1}I_{k} and so

|hkhIk|=limh|khIk|124>0.\displaystyle\Big{|}\bigcap_{h\in\mathbb{N}}\bigcup_{k\geq h}I_{k}\Big{|}=\lim_{h\to\infty}\Big{|}\bigcup_{k\geq h}I_{k}\Big{|}\geq\frac{1}{24}>0. (7.20)

It follows that there exists a subsequence h(k)h(k)\to\infty as kk\to\infty and r(2/3,3/4)r\in(2/3,3/4) such that

rkIh(k),limkn1(Krs(E(1)ΔEh(k)))=0, and\displaystyle r\in\bigcap_{k\in\mathbb{N}}I_{h(k)},\qquad\lim_{k\to\infty}\operatorname{\mathcal{H}^{n-1}}(\partial\textbf{{K}}_{rs}\cap(E^{(1)}\Delta E_{h(k)}))=0,\text{ and}
n1(EK(z,rs))=n1(Eh(k)K(z,rs))=0.\displaystyle\operatorname{\mathcal{H}^{n-1}}(\partial^{*}\!E\cap\textbf{{K}}(z,rs))=\operatorname{\mathcal{H}^{n-1}}(\partial E_{h(k)}\cap\textbf{{K}}(z,rs))=0. (7.21)

By Lemma 7.2 there exist a sequence of open sets FkF_{k} of locally finite perimeter in n\mathbb{R}^{n} such that

FkKrs=Eh(k)Krs,\displaystyle F_{k}\cap\partial\textbf{{K}}_{rs}=E_{h(k)}\cap\partial\textbf{{K}}_{rs}, (7.22)
n1(FkK(z,rs))=n1(Eh(k)K(z,rs))=0,\displaystyle\operatorname{\mathcal{H}^{n-1}}(\partial F_{k}\cap\partial\textbf{{K}}(z,rs))=\operatorname{\mathcal{H}^{n-1}}(\partial E_{h(k)}\cap\partial\textbf{{K}}(z,rs))=0, (7.23)

and

P(Fk;K(z,rs))n1(D(z,rs))\displaystyle P(F_{k};\textbf{{K}}(z,rs))-\operatorname{\mathcal{H}^{n-1}}(\textbf{{D}}(z,rs))
C(n){t(P(Eh(k);K(z,s))n1(D(z,s))+1tK(z,s)Eh(k)|qxc|2s2dn1(x)}.\displaystyle\leq C(n)\Big{\{}t\big{(}P(E_{h(k)};\textbf{{K}}(z,s))-\operatorname{\mathcal{H}^{n-1}}(\textbf{{D}}(z,s)\big{)}+\frac{1}{t}\int_{\textbf{{K}}(z,s)\cap\partial E_{h(k)}}\frac{|\operatorname{\textnormal{{q}}}x-c|^{2}}{s^{2}}\operatorname{\,d\mathcal{H}^{n-1}}(x)\Big{\}}. (7.24)

Now consider the comparison sets

Gk=(FkK(z,rs))(EK(z,rs)).\displaystyle G_{k}=(F_{k}\cap\textbf{{K}}(z,{rs}))\cup(E\setminus\textbf{{K}}(z,rs)). (7.25)

Since EΔGkC2E\Delta G_{k}\subset\subset\textbf{{C}}_{2} and

n1(FkK(z,rs))=n1(EK(z,rs))=0,\displaystyle\operatorname{\mathcal{H}^{n-1}}(\partial F_{k}\cap\partial\textbf{{K}}(z,rs))=\operatorname{\mathcal{H}^{n-1}}(\partial^{*}\!E\cap\partial\textbf{{K}}(z,rs))=0, (7.26)

we have that

P(Gk;C2)=P(Fk;K(z,rs))+P(E;C2K(z,rs)¯)+n1(K(z,rs)(E(1)ΔFk)).\displaystyle P(G_{k};\textbf{{C}}_{2})=P(F_{k};\textbf{{K}}(z,rs))+P(E;\textbf{{C}}_{2}\setminus\overline{\textbf{{K}}(z,rs)})+\operatorname{\mathcal{H}^{n-1}}(\partial\textbf{{K}}(z,rs)\cap(E^{(1)}\Delta F_{k})). (7.27)

By Proposition 4.5 we have

P(E;C2)P(Gk;C2)+C(κ+ACα)\displaystyle P(E;\textbf{{C}}_{2})\leq P(G_{k};\textbf{{C}}_{2})+C(\kappa+||A||_{C^{\alpha}}) (7.28)

for some C=C(n,λ,Λ,κ,α,r0)C=C(n,\lambda,\Lambda,\kappa,\alpha,r_{0}). Hence

P(E;K(z,rs))P(Fk;K(z,rs))+n1(K(z,rs)(E(1)ΔFk))+C(κ+ACα).\displaystyle P(E;\textbf{{K}}(z,rs))\leq P(F_{k};\textbf{{K}}(z,rs))+\operatorname{\mathcal{H}^{n-1}}(\partial\textbf{{K}}(z,rs)\cap(E^{(1)}\Delta F_{k}))+C(\kappa+||A||_{C^{\alpha}}). (7.29)

It follows that

P(E;K(z,rs))n1(D(z,rs))\displaystyle P(E;\textbf{{K}}(z,rs))-\operatorname{\mathcal{H}^{n-1}}(\textbf{{D}}(z,rs))
C(n){t(P(Eh(k);K(z,s))n1(D(z,s))+1tK(z,s)Eh(k)|qxc|2s2dn1(x)}\displaystyle\leq C(n)\Big{\{}t\big{(}P(E_{h(k)};\textbf{{K}}(z,s))-\operatorname{\mathcal{H}^{n-1}}(\textbf{{D}}(z,s)\big{)}+\frac{1}{t}\int_{\textbf{{K}}(z,s)\cap\partial E_{h(k)}}\frac{|\operatorname{\textnormal{{q}}}x-c|^{2}}{s^{2}}\operatorname{\,d\mathcal{H}^{n-1}}(x)\Big{\}}
+n1(K(z,rs)(E(1)ΔFk))+C(κ+ACα).\displaystyle\qquad+\operatorname{\mathcal{H}^{n-1}}(\partial\textbf{{K}}(z,rs)\cap(E^{(1)}\Delta F_{k}))+C(\kappa+||A||_{C^{\alpha}}). (7.30)

Taking the limit as kk\to\infty, using the weak convergence of (7.15) since n1(EK(z,s))=0\operatorname{\mathcal{H}^{n-1}}(\partial E\cap\textbf{{K}}(z,s))=0, and limkn1(Krs(E(1)ΔEh(k)))=0\lim_{k\to\infty}\operatorname{\mathcal{H}^{n-1}}(\partial\textbf{{K}}_{rs}\cap(E^{(1)}\Delta E_{h(k)}))=0, we have

P(E;K(z,rs))n1(D(z,rs))\displaystyle P(E;\textbf{{K}}(z,rs))-\operatorname{\mathcal{H}^{n-1}}(\textbf{{D}}(z,rs))
C(n){t(P(E;K(z,s))n1(D(z,s))+1tK(z,s)E|qxc|2s2dn1(x)}\displaystyle\leq C(n)\Big{\{}t\big{(}P(E;\textbf{{K}}(z,s))-\operatorname{\mathcal{H}^{n-1}}(\textbf{{D}}(z,s)\big{)}+\frac{1}{t}\int_{\textbf{{K}}(z,s)\cap\partial^{*}\!E}\frac{|\operatorname{\textnormal{{q}}}x-c|^{2}}{s^{2}}\operatorname{\,d\mathcal{H}^{n-1}}(x)\Big{\}}
+C(κ+ACα).\displaystyle\qquad+C(\kappa+||A||_{C^{\alpha}}). (7.31)

Hence

ζ(K(z,s/2))\displaystyle\zeta(\textbf{{K}}(z,s/2)) ζ(K(z,rs))\displaystyle\leq\zeta(\textbf{{K}}(z,rs))
C{tζ(K(z,s)+1tK(z,s)E|qxc|2s2dn1(x)+κ+||A||Cα}\displaystyle\leq C\Big{\{}t\zeta(\textbf{{K}}(z,s)+\frac{1}{t}\int_{\textbf{{K}}(z,s)\cap\partial^{*}\!E}\frac{|\operatorname{\textnormal{{q}}}x-c|^{2}}{s^{2}}\operatorname{\,d\mathcal{H}^{n-1}}(x)+\kappa+||A||_{C^{\alpha}}\Big{\}} (7.32)

By (7.14), ζ(K(z,s/2))ζ(K(z,s))\zeta(\textbf{{K}}(z,s/2))\leq\zeta(\textbf{{K}}(z,s)) and so this inequality also holds for t>1/4t>1/4 provided we take C=C(n,λ,Λ,κ,α,r0)4C=C(n,\lambda,\Lambda,\kappa,\alpha,r_{0})\geq 4. Hence it holds for all t(0,)t\in(0,\infty). Minimizing the right hand side over all tt yields

ζ(K(z,s/2))C{(ζ(K(z,s))K(z,s)E|qxc|2s2dn1)1/2+κ+||A||Cα}\displaystyle\zeta(\textbf{{K}}(z,s/2))\leq C\Big{\{}\Big{(}\zeta(\textbf{{K}}(z,s))\int_{\textbf{{K}}(z,s)\cap\partial^{*}\!E}\frac{|\operatorname{\textnormal{{q}}}x-c|^{2}}{s^{2}}\operatorname{\,d\mathcal{H}^{n-1}}\Big{)}^{1/2}+\kappa+||A||_{C^{\alpha}}\Big{\}} (7.33)

as desired. ∎

Proof of Theorem 7.1.

By the scaling given in Proposition 4.2, we have that Ex0,rE_{x_{0},r} is a (κrα,α)(\kappa r^{\alpha},\alpha)-almost-minimizer of Ax0,r\mathscr{F}_{A_{x_{0},r}} in C4=C(0,4,ν)\textbf{{C}}_{4}=\textbf{{C}}(0,4,\nu) at scale r0/rr_{0}/r with 0Ex0,r0\in\partial E_{x_{0},r}, Ax0,r(0)=IA_{x_{0},r}(0)=I, Ax0,rCα=ACαrα||A_{x_{0},r}||_{C^{\alpha}}=||A||_{C^{\alpha}}r^{\alpha}, and 4<r0/r4<r_{0}/r. Thus to prove (7.3), we may assume eC(Ex0,r,0,4,ν)=eC(E,x0,4r,ν)ω\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E_{x_{0},r},0,4,\nu)=\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},4r,\nu)\leq\omega and show

eC(Ex0,r,0,1,ν)C(f(Ex0,r,0,2,ν)+κrα+Ax0,rCα).\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E_{x_{0},r},0,1,\nu)\leq C\big{(}\operatorname{\textnormal{{f}}}(E_{x_{0},r},0,2,\nu)+\kappa r^{\alpha}+||A_{x_{0},r}||_{C^{\alpha}}\big{)}. (7.34)

By Proposition 5.7 and Proposition 5.8, it follows that

|qx|<18,xC2Ex0,r,\displaystyle|\operatorname{\textnormal{{q}}}x|<\frac{1}{8},\qquad\forall x\in\textbf{{C}}_{2}\cap\partial E_{x_{0},r}, (7.35)
|{xC2Ex0,r:qx<18}|=|{xC2Ex0,r:qx>18}|=0,\displaystyle\Big{|}\Big{\{}x\in\textbf{{C}}_{2}\setminus E_{x_{0},r}\mathrel{\mathop{\mathchar 58\relax}}\operatorname{\textnormal{{q}}}x<-\frac{1}{8}\Big{\}}\Big{|}=\Big{|}\Big{\{}x\in\textbf{{C}}_{2}\cap E_{x_{0},r}\mathrel{\mathop{\mathchar 58\relax}}\operatorname{\textnormal{{q}}}x>\frac{1}{8}\Big{\}}\Big{|}=0, (7.36)

and

n1(G)=C2Ex0,rp1(G)(νEν)dn1,GD2.\displaystyle\operatorname{\mathcal{H}^{n-1}}(G)=\int_{\textbf{{C}}_{2}\cap\partial^{*}\!E_{x_{0},r}\cap\operatorname{\textnormal{{p}}}^{-1}(G)}(\nu_{E}\cdot\nu)\operatorname{\,d\mathcal{H}^{n-1}},\qquad\forall\ G\subset\textbf{{D}}_{2}. (7.37)

Hence eC(Ex0,r,0,1,ν)=P(Ex0,r;C1)n1(D1)\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E_{x_{0},r},0,1,\nu)=P(E_{x_{0},r};\textbf{{C}}_{1})-\operatorname{\mathcal{H}^{n-1}}(\textbf{{D}}_{1}) and so it suffices to show that for every cc\in\mathbb{R},

P(Ex0,r;C1)n1(D1)C{C2Ex0,r|qxc|2dn1+κrα+Ax0,rCα}.\displaystyle P(E_{x_{0},r};\textbf{{C}}_{1})-\operatorname{\mathcal{H}^{n-1}}(\textbf{{D}}_{1})\leq C\Big{\{}\int_{\textbf{{C}}_{2}\cap\partial E_{x_{0},r}}|\operatorname{\textnormal{{q}}}x-c|^{2}\operatorname{\,d\mathcal{H}^{n-1}}+\kappa r^{\alpha}+||A_{x_{0},r}||_{C^{\alpha}}\Big{\}}. (7.38)

If |c|>1/4|c|>1/4, then |qxc|1/8|\operatorname{\textnormal{{q}}}x-c|\geq 1/8 and so

C2Ex0,r|qxc|2dn1P(Ex0,r;C1)82\displaystyle\int_{\textbf{{C}}_{2}\cap\partial E_{x_{0},r}}|\operatorname{\textnormal{{q}}}x-c|^{2}\operatorname{\,d\mathcal{H}^{n-1}}\geq\frac{P(E_{x_{0},r};\textbf{{C}}_{1})}{8^{2}} (7.39)

and we are done provided we take C64C\geq 64. Thus we are left with the case |c|<1/4|c|<1/4. Set

ζ(G)=P(Ex0,r;C2p1(G))n1(G), for GD2,\displaystyle\zeta(G)=P(E_{x_{0},r};\textbf{{C}}_{2}\cap\operatorname{\textnormal{{p}}}^{-1}(G))-\operatorname{\mathcal{H}^{n-1}}(G),\qquad\text{ for }G\subset\textbf{{D}}_{2}, (7.40)

which defines a Radon measure on n1\mathbb{R}^{n-1}, concentrated on D2\textbf{{D}}_{2}. We apply Lemma 7.3 in every cylinder K(z,s)\textbf{{K}}(z,s) with zn1z\in\mathbb{R}^{n-1} and s>0s>0 such that

D(z,2s)D2,n1(Ex0,rK(z,2s))=0,\displaystyle\textbf{{D}}(z,2s)\subset\textbf{{D}}_{2},\qquad\operatorname{\mathcal{H}^{n-1}}(\partial^{*}\!E_{x_{0},r}\cap\partial\textbf{{K}}(z,2s))=0, (7.41)

to find that

ζ(D(z,s))C{(ζ(D(z,2s)infc<1/4K(z,2s)Ex0,r|qxc|2(2s)2dn1)1/2+κrα+||Ax0,r||Cα}.\displaystyle\zeta(\textbf{{D}}(z,s))\leq C\Big{\{}\Big{(}\zeta(\textbf{{D}}(z,2s)\inf_{c<1/4}\int_{\textbf{{K}}(z,2s)\cap\partial^{*}\!E_{x_{0},r}}\frac{|\operatorname{\textnormal{{q}}}x-c|^{2}}{(2s)^{2}}\operatorname{\,d\mathcal{H}^{n-1}}\Big{)}^{1/2}+\kappa r^{\alpha}+||A_{x_{0},r}||_{C^{\alpha}}\Big{\}}. (7.42)

An approximation argument, setting

h=inf|c|<1/4C2Ex0,r|qxc|2dn1,\displaystyle h=\inf_{|c|<1/4}\int_{\textbf{{C}}_{2}\cap\partial^{*}\!E_{x_{0},r}}|\operatorname{\textnormal{{q}}}x-c|^{2}\operatorname{\,d\mathcal{H}^{n-1}}, (7.43)

for brevity, implies by (7.42) that

s2ζ(D(z,s))C(s2ζ(D(z,2s))h+κrα+Ax0,rCα)\displaystyle s^{2}\zeta(\textbf{{D}}(z,s))\leq C\Big{(}\sqrt{s^{2}\zeta(\textbf{{D}}(z,2s))h}+\kappa r^{\alpha}+||A_{x_{0},r}||_{C^{\alpha}}\Big{)} (7.44)

whenever D(z,2s)D2\textbf{{D}}(z,2s)\subset\textbf{{D}}_{2} since s<1s<1. We now use a covering argument to complete the proof. Let

Q=sup{s2ζ(D(z,s)):D(z,2s)D2},\displaystyle Q=\sup\{s^{2}\zeta(\textbf{{D}}(z,s))\mathrel{\mathop{\mathchar 58\relax}}\textbf{{D}}(z,2s)\subset\textbf{{D}}_{2}\}, (7.45)

and notice Q<Q<\infty since for every D(z,2s)D2\textbf{{D}}(z,2s)\subset\textbf{{D}}_{2}, we have

s2ζ(D(z,s))ζ(D2)P(Ex0,r;C2)<.\displaystyle s^{2}\zeta(\textbf{{D}}(z,s))\leq\zeta(\textbf{{D}}_{2})\leq P(E_{x_{0},r};\textbf{{C}}_{2})<\infty. (7.46)

Given D(z,2s)D2\textbf{{D}}(z,2s)\subset\textbf{{D}}_{2}, cover D(z,s)\textbf{{D}}(z,s) by finite many balls {D(zk,s/4)}k=1N\{\textbf{{D}}(z_{k},s/4)\}_{k=1}^{N} with centers zkD(z,s)z_{k}\in\textbf{{D}}(z,s). This can be done with a bounded number of balls depending only on the dimension nn, that is, NN(n)N\leq N(n). So by the subadditivity of the measure ζ\zeta, (7.44), and the definition of QQ, we have

s2ζ(D(z,s))\displaystyle s^{2}\zeta(\textbf{{D}}(z,s)) 16k=1N(s4)2ζ(D(zk,s4))\displaystyle\leq 16\sum_{k=1}^{N}\Big{(}\frac{s}{4}\Big{)}^{2}\zeta\Big{(}\textbf{{D}}\Big{(}z_{k},\frac{s}{4}\Big{)}\Big{)}
Ck=1N((s2)2ζ(D(zk,s2))h+κrα+Ax0,rCα)\displaystyle\leq C\sum_{k=1}^{N}\Big{(}\sqrt{\Big{(}\frac{s}{2}\Big{)}^{2}\zeta\Big{(}\textbf{{D}}\Big{(}z_{k},\frac{s}{2}\Big{)}\Big{)}h}+\kappa r^{\alpha}+||A_{x_{0},r}||_{C^{\alpha}}\Big{)}
CN(n)(Qh+κrα+Ax0,rCα)\displaystyle\leq CN(n)\Big{(}\sqrt{Qh}+\kappa r^{\alpha}+||A_{x_{0},r}||_{C^{\alpha}}\Big{)} (7.47)

where we used that D(zk,s/4)D(z,2s)D2\textbf{{D}}(z_{k},s/4)\subset\textbf{{D}}(z,2s)\subset\textbf{{D}}_{2}. Hence QC(Qh+κrα+Ax0,rCα)Q\leq C(\sqrt{Qh}+\kappa r^{\alpha}+||A_{x_{0},r}||_{C^{\alpha}}) for some C=C(n,λ,Λ,κ,α,r0)C=C(n,\lambda,\Lambda,\kappa,\alpha,r_{0}). By Cauchy-Schwarz, we have CQh=QC2h12Q+12C2hC\sqrt{Qh}=\sqrt{QC^{2}h}\leq\frac{1}{2}Q+\frac{1}{2}C^{2}h. Combining these gives Q12Q+C(h+κrα+Ax0,rCα)Q\leq\frac{1}{2}Q+C(h+\kappa r^{\alpha}+||A_{x_{0},r}||_{C^{\alpha}}). Thus ζ(D1)QC(h+κrα+Ax0,rCα)\zeta(\textbf{{D}}_{1})\leq Q\leq C(h+\kappa r^{\alpha}+||A_{x_{0},r}||_{C^{\alpha}}) for some C=C(n,λ,Λ,κ,α,r0)C=C(n,\lambda,\Lambda,\kappa,\alpha,r_{0}). Recalling the definitions of ζ(D1)\zeta(\textbf{{D}}_{1}) and hh, we see that this completes the proof of (7.38). ∎

8. Tilt-Excess Decay

We showed in Section 6 that almost-minimizers can be approximated by ‘almost-harmonic’ Lipschitz functions at points with small excess. Now we approximate these Lipschitz with harmonic functions which allow us to find new directions for which the excess experiences quadratic decay.

First we recall a couple lemmas about harmonic functions. These are just the rescaled versions of [Mag12, Lemma 25.1, Lemma 25.2]. Note that

Ds=1ωn1sn1Ds\displaystyle\fint_{\textbf{{D}}_{s}}=\frac{1}{\omega_{n-1}s^{n-1}}\int_{\textbf{{D}}_{s}} (8.1)

denotes the integral average.

Lemma 8.1.

There is a positive constant C(n)C(n) with the following property. If v:n1v\colon\mathbb{R}^{n-1}\to\mathbb{R} is harmonic in Dr\textbf{{D}}_{r} and w:n1w\colon\mathbb{R}^{n-1}\to\mathbb{R} is defined by w(z)=v(0)+v(0)zw(z)=v(0)+\nabla v(0)\cdot z, then

supDθr|vw|rC(n)θ2(Dr|v|2)1/2\displaystyle\sup_{\textbf{{D}}_{\theta r}}\frac{|v-w|}{r}\leq C(n)\theta^{2}\bigg{(}\fint_{\textbf{{D}}_{r}}|\nabla^{\prime}v|^{2}\bigg{)}^{1/2} (8.2)

for every θ(0,1/2]\theta\in(0,1/2]. In particular,

Dθr(|vw|θr)2C(n)θ2Dr|v|2.\displaystyle\fint_{\textbf{{D}}_{\theta r}}\bigg{(}\frac{|v-w|}{\theta r}\bigg{)}^{2}\leq C(n)\theta^{2}\fint_{\textbf{{D}}_{r}}|\nabla^{\prime}v|^{2}. (8.3)
Lemma 8.2 (Harmonic approximation).

For every τ>0\tau>0 there exists σ>0\sigma>0 with the following property. If uW1,2(Dr)u\in W^{1,2}(\textbf{{D}}_{r}) is such that

Dr|u|21,|Druφ|supDr|φ|σφCc(Dr),\displaystyle\fint_{\textbf{{D}}_{r}}|\nabla^{\prime}u|^{2}\leq 1,\qquad\bigg{|}\fint_{\textbf{{D}}_{r}}\nabla^{\prime}u\cdot\nabla^{\prime}\varphi\bigg{|}\leq\sup_{\textbf{{D}}_{r}}|\nabla^{\prime}\varphi|\>\sigma\qquad\forall\varphi\in C_{c}^{\infty}(\textbf{{D}}_{r}), (8.4)

then there exists a harmonic function vv on Dr\textbf{{D}}_{r} such that

Dr|v|21, and Dr|vu|2τr2.\displaystyle\fint_{\textbf{{D}}_{r}}|\nabla^{\prime}v|^{2}\leq 1,\qquad\text{ and }\qquad\fint_{\textbf{{D}}_{r}}|v-u|^{2}\leq\tau r^{2}. (8.5)

We now prove the excess improvement by tilting. This states that if the excess is small enough in a given direction, then there is a nearby direction in which the excess at a definite smaller scale sees quadratic decay with the error term seeing α\alphath power decay. Note in the theorem below, the fraction 1/1041/104 comes from the rough bound 13421342=10413\cdot 4\cdot\sqrt{2}\leq 13\cdot 4\cdot 2=104 where the 1313 comes from the Lipschitz approximation theorem, the 44 comes from small excess assumption in the reverse Poincaré inequality, and the 2\sqrt{2} comes from containing one cylinder inside of another cylinder that is tilted in a different direction.

Theorem 8.3 (Excess improvement by tilting).

Given θ(0,1/104]\theta\in(0,1/104], there exist positive constants ε2=ε2(n,λ,Λ,κ,α,r0,ACα,θ)\varepsilon_{2}=\varepsilon_{2}(n,\lambda,\Lambda,\kappa,\alpha,r_{0},||A||_{C^{\alpha}},\theta) and C3=C3(n,λ,Λ,κ,α,r0,ACα)C_{3}=C_{3}(n,\lambda,\Lambda,\kappa,\alpha,r_{0},||A||_{C^{\alpha}}) with the following property. If EE is a (κ,α)(\kappa,\alpha)-almost-minimizer of A\mathscr{F}_{A} in C(x0,r0,ν0)\textbf{{C}}(x_{0},r_{0},\nu_{0}), x0Ex_{0}\in\partial E, A(x0)=IA(x_{0})=I, and r<r0r<r_{0} with

eC(E,x0,r,ν0)+rα/2ε2,\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},r,\nu_{0})+r^{\alpha/2}\leq\varepsilon_{2}, (8.6)

then there exists ν1SSn1\nu_{1}\in\SS^{n-1} such that

eC(E,x0,θr,ν1)C3(θ2eC(E,x0,r,ν0)+θαrα/2)\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},\theta r,\nu_{1})\leq C_{3}(\theta^{2}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},r,\nu_{0})+\theta^{\alpha}r^{\alpha/2}) (8.7)
Proof.

Assuming without loss of generality that x0=0x_{0}=0 and ν0=en\nu_{0}=e_{n}, it suffices to prove that given θ(0,1/8]\theta\in(0,1/8], there exist positive constants ε2=ε2(n,λ,Λ,κ,α,r0,ACα,θ)\varepsilon_{2}=\varepsilon_{2}(n,\lambda,\Lambda,\kappa,\alpha,r_{0},||A||_{C^{\alpha}},\theta) and C3=C3(n,λ,Λ,κ,α,r0,ACα)C_{3}=C_{3}(n,\lambda,\Lambda,\kappa,\alpha,r_{0},||A||_{C^{\alpha}}) with the following property. If EE is a (κ,α)(\kappa,\alpha)-almost-minimizer of A\mathscr{F}_{A} in C(0,r0,en)\textbf{{C}}(0,r_{0},e_{n}), A(0)=IA(0)=I, 0E0\in\partial E, and 13r<r013r<r_{0} with

eC(E,0,13r,en)+rα/2ε2,\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,0,13r,e_{n})+r^{\alpha/2}\leq\varepsilon_{2}, (8.8)

then there exists ν1SSn1\nu_{1}\in\SS^{n-1} such that

eC(E,0,θr,ν1)C3(θ2eC(E,0,13r,en)+θαrα/2)\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,0,\theta r,\nu_{1})\leq C_{3}(\theta^{2}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,0,13r,e_{n})+\theta^{\alpha}r^{\alpha/2}) (8.9)

We set Cs=C(0,s,en)\textbf{{C}}_{s}=\textbf{{C}}(0,s,e_{n}) for brevity.

We shall select a number of criteria for ε2\varepsilon_{2} to satisfy which together give the desired result. We place a box around each of these choices to make it easy for the reader to check that all of these choices are consistent.

Choose ε2\varepsilon_{2} to satisfy

ε2ε1\displaystyle\boxed{\varepsilon_{2}\leq\varepsilon_{1}} (8.10)

where ε1\varepsilon_{1} is from the Lipschitz approximation theorem. Then eC(E,0,13r,en)ε1\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,0,13r,e_{n})\leq\varepsilon_{1} and thus there is a Lipschitz function u:n1u\colon\mathbb{R}^{n-1}\to\mathbb{R} such that Lipu1\operatorname{Lip}u\leq 1 such that

supDr|u(z)|rC1eC(E,0,13r,en)1/(2(n1)),\displaystyle\sup_{\textbf{{D}}_{r}}\frac{|u(z)|}{r}\leq C_{1}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,0,13r,e_{n})^{1/(2(n-1))}, (8.11)
n1(MΔΓ)rn1C1eC(E,0,13r,en),\displaystyle\frac{\operatorname{\mathcal{H}^{n-1}}(M\Delta\Gamma)}{r^{n-1}}\leq C_{1}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,0,13r,e_{n}), (8.12)
Dr|u|2C1eC(E,0,13r,en), and\displaystyle\fint_{D_{r}}|\nabla^{\prime}u|^{2}\leq C_{1}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,0,13r,e_{n}),\text{ and } (8.13)
|Druφ|C1supDr|φ|(eC(E,0,13r,en)+rα/2) for all φCc1(Dr).\displaystyle\bigg{|}\fint_{\textbf{{D}}_{r}}\nabla^{\prime}u\cdot\nabla^{\prime}\varphi\bigg{|}\leq C_{1}\sup_{\textbf{{D}}_{r}}|\nabla^{\prime}\varphi|\big{(}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,0,13r,e_{n})+r^{\alpha/2})\qquad\text{ for all }\varphi\in C_{c}^{1}(\textbf{{D}}_{r}). (8.14)

where C1C_{1} is the constant from the Lipschitz approximation theorem, M=CrEM=\textbf{{C}}_{r}\cap\partial E, and Γ\Gamma is the graph of uu. Choose ε2\varepsilon_{2} to also satisfy

C1ε21.\displaystyle\boxed{C_{1}\varepsilon_{2}\leq 1}. (8.15)

Then C1(eC(E,0,13r,en)+rα/2)1C_{1}(\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,0,13r,e_{n})+r^{\alpha/2})\leq 1 and so setting

β=C1(eC(E,0,13r,en)+rα/2)andu0=u/β,\displaystyle\beta=C_{1}(\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,0,13r,e_{n})+r^{\alpha/2})\qquad\text{and}\qquad u_{0}=u/\sqrt{\beta}, (8.16)

we have

Dr|u0|21and|Dru0φ|supDr|φ|β for all φCc1(Dr).\displaystyle\fint_{\textbf{{D}}_{r}}|\nabla^{\prime}u_{0}|^{2}\leq 1\qquad\text{and}\qquad\bigg{|}\fint_{\textbf{{D}}_{r}}\nabla^{\prime}u_{0}\cdot\nabla^{\prime}\varphi\bigg{|}\leq\sup_{\textbf{{D}}_{r}}|\nabla^{\prime}\varphi|\sqrt{\beta}\qquad\text{ for all }\varphi\in C_{c}^{1}(\textbf{{D}}_{r}). (8.17)

By Lemma 8.2, for every τ>0\tau>0 there is σ(τ)>0\sigma(\tau)>0 such that if

βσ(τ)\displaystyle\sqrt{\beta}\leq\sigma(\tau) (8.18)

then there is v0:n1v_{0}\mathrel{\mathop{\mathchar 58\relax}}\mathbb{R}^{n-1}\to\mathbb{R} which is harmonic in Dr\textbf{{D}}_{r} such that

Dr|u0v0|2τr2andDr|v0|21\displaystyle\fint_{\textbf{{D}}_{r}}|u_{0}-v_{0}|^{2}\leq\tau r^{2}\qquad\text{and}\qquad\fint_{\textbf{{D}}_{r}}|\nabla^{\prime}v_{0}|^{2}\leq 1 (8.19)

Setting v=βv0v=\sqrt{\beta}\>v_{0}, we have that vv is harmonic in Dr\textbf{{D}}_{r} and

Dr|uv|2τr2βandDr|v|2β.\displaystyle\fint_{\textbf{{D}}_{r}}|u-v|^{2}\leq\tau r^{2}\beta\qquad\text{and}\qquad\fint_{\textbf{{D}}_{r}}|\nabla^{\prime}v|^{2}\leq\beta. (8.20)

Since 4θ1/24\theta\leq 1/2, setting w(z)=v(0)+v(0)zw(z)=v(0)+\nabla^{\prime}v(0)\cdot z for zDrz\in\textbf{{D}}_{r}, we see by Lemma 8.1 that

D4θr|vw|2(θr)2C(n)θ2Dr|v|2C(n)θ2β.\displaystyle\fint_{\textbf{{D}}_{4\theta r}}\frac{|v-w|^{2}}{(\theta r)^{2}}\leq C(n)\theta^{2}\fint_{\textbf{{D}}_{r}}|\nabla^{\prime}v|^{2}\leq C(n)\theta^{2}\beta. (8.21)

By (8.20) and D4θrDr\textbf{{D}}_{4\theta r}\subset\textbf{{D}}_{r},

D4θr|uv|2(θr)2Dr|uv|2(θr)2τθ2βrn1\displaystyle\int_{\textbf{{D}}_{4\theta r}}\frac{|u-v|^{2}}{(\theta r)^{2}}\leq\int_{\textbf{{D}}_{r}}\frac{|u-v|^{2}}{(\theta r)^{2}}\leq\frac{\tau}{\theta^{2}}\beta r^{n-1} (8.22)

and so

D4θr|uv|2(θr)2C(n)τθn+1β.\displaystyle\fint_{\textbf{{D}}_{4\theta r}}\frac{|u-v|^{2}}{(\theta r)^{2}}\leq C(n)\frac{\tau}{\theta^{n+1}}\beta. (8.23)

Noting |uw|22|uv|2+2|vw|2|u-w|^{2}\leq 2|u-v|^{2}+2|v-w|^{2}, we have that

D4θr|uw|2(θr)2C(n)(τθn+1+θ2)β.\displaystyle\fint_{\textbf{{D}}_{4\theta r}}\frac{|u-w|^{2}}{(\theta r)^{2}}\leq C(n)\Big{(}\frac{\tau}{\theta^{n+1}}+\theta^{2}\Big{)}\beta. (8.24)

We apply the above with τ=θn+3\tau=\theta^{n+3} and choose ε2\varepsilon_{2} to also satisfy

C1ε2σ(θn+3)\displaystyle\boxed{\sqrt{C_{1}\varepsilon_{2}}\leq\sigma(\theta^{n+3})} (8.25)

with σ()\sigma(\>\cdot\>) as in (8.18). Then βσ(θn+3)\sqrt{\beta}\leq\sigma(\theta^{n+3}) and so

D4θr|uw|2C(n)θn+3βrn+1.\displaystyle\int_{\textbf{{D}}_{4\theta r}}|u-w|^{2}\leq C(n)\theta^{n+3}\beta r^{n+1}. (8.26)

Now set

ν1=(v(0),1)1+|v(0)|2SSn1,c1=v(0)1+|v(0)|2\displaystyle\nu_{1}=\frac{(-\nabla^{\prime}v(0),1)}{\sqrt{1+|\nabla^{\prime}v(0)|^{2}}}\in\SS^{n-1},\qquad c_{1}=-\frac{v(0)}{\sqrt{1+|\nabla^{\prime}v(0)|^{2}}}\in\mathbb{R} (8.27)

and let’s estimate f(E,0,2θr,ν1)\operatorname{\textnormal{{f}}}(E,0,2\theta r,\nu_{1}). Since C(0,2θr,ν1)C4θr\textbf{{C}}(0,2\theta r,\nu_{1})\subset\textbf{{C}}_{4\theta r}, we have that

f(E,0,2θr,ν1)\displaystyle\operatorname{\textnormal{{f}}}(E,0,2\theta r,\nu_{1}) =1(2θr)n+1infcC(0,2θr,ν1)E|xν1c|2dn1(x)\displaystyle=\frac{1}{(2\theta r)^{n+1}}\inf_{c\in\mathbb{R}}\int_{\textbf{{C}}(0,2\theta r,\nu_{1})\cap\partial^{*}\!E}|x\cdot\nu_{1}-c|^{2}\operatorname{\,d\mathcal{H}^{n-1}}(x)
C(n)(θr)n+1C4θrE|xν1c1|2dn1(x).\displaystyle\leq\frac{C(n)}{(\theta r)^{n+1}}\int_{\textbf{{C}}_{4\theta r}\cap\partial^{*}\!E}|x\cdot\nu_{1}-c_{1}|^{2}\operatorname{\,d\mathcal{H}^{n-1}}(x). (8.28)

This last integral we split in terms of MΓM\cap\Gamma and MΓM\setminus\Gamma.

For MΓM\cap\Gamma, by Lip(u)1\operatorname{Lip}(u)\leq 1, (8.27), and (8.26), we have

C4θrMΓ|xν1c1|2dn1(x)\displaystyle\int_{\textbf{{C}}_{4\theta r}\cap M\cap\Gamma}|x\cdot\nu_{1}-c_{1}|^{2}\operatorname{\,d\mathcal{H}^{n-1}}(x) =D4θrp(MΓ)|(z,u(z))ν1c1|21+|u(z)|2𝑑z\displaystyle=\int_{\textbf{{D}}_{4\theta r}\cap\operatorname{\textnormal{{p}}}(M\cap\Gamma)}|(z,u(z))\cdot\nu_{1}-c_{1}|^{2}\sqrt{1+|\nabla^{\prime}u(z)|^{2}}dz
2D4θrp(MΓ)|uw|21+|v(0)|2\displaystyle\leq\sqrt{2}\int_{\textbf{{D}}_{4\theta r}\cap\operatorname{\textnormal{{p}}}(M\cap\Gamma)}\frac{|u-w|^{2}}{1+|\nabla^{\prime}v(0)|^{2}}
2D4θrp(MΓ)|uw|2\displaystyle\leq\sqrt{2}\int_{\textbf{{D}}_{4\theta r}\cap\operatorname{\textnormal{{p}}}(M\cap\Gamma)}|u-w|^{2}
C(n)θn+3βrn+1.\displaystyle\leq C(n)\theta^{n+3}\beta r^{n+1}. (8.29)

For MΓM\setminus\Gamma, observe that

C4θr(MΓ)|xν1c1|2dn1(x)\displaystyle\int_{\textbf{{C}}_{4\theta r}\cap(M\setminus\Gamma)}|x\cdot\nu_{1}-c_{1}|^{2}\operatorname{\,d\mathcal{H}^{n-1}}(x) =C4θr(MΓ)|qx+v(0)pxv(0)|21+|v(0)|2dn1(x)\displaystyle=\int_{\textbf{{C}}_{4\theta r}\cap(M\setminus\Gamma)}\frac{|\operatorname{\textnormal{{q}}}x+v(0)-\operatorname{\textnormal{{p}}}x\cdot\nabla^{\prime}v(0)|^{2}}{1+|\nabla^{\prime}v(0)|^{2}}\operatorname{\,d\mathcal{H}^{n-1}}(x)
C4θr(MΓ)|qx+v(0)pxv(0)|2dn1(x)\displaystyle\leq\int_{\textbf{{C}}_{4\theta r}\cap(M\setminus\Gamma)}|\operatorname{\textnormal{{q}}}x+v(0)-\operatorname{\textnormal{{p}}}x\cdot\nabla^{\prime}v(0)|^{2}\operatorname{\,d\mathcal{H}^{n-1}}(x)
3n1(MΓ)(supxM|qx|2+|v(0)|2+supxM|px|2|v(0)|2).\displaystyle\leq 3\operatorname{\mathcal{H}^{n-1}}(M\setminus\Gamma)(\sup_{x\in M}|\operatorname{\textnormal{{q}}}x|^{2}+|v(0)|^{2}+\sup_{x\in M}|\operatorname{\textnormal{{p}}}x|^{2}|\nabla^{\prime}v(0)|^{2}). (8.30)

By the height bound, we have

supxM|qx|2C02eC(E,0,4r,en)1/(n1)r2C02(13/4)eC(E,0,13r,en)1/(n1)Cβ1/(n1)r2.\displaystyle\sup_{x\in M}|\operatorname{\textnormal{{q}}}x|^{2}\leq C_{0}^{2}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,0,4r,e_{n})^{1/(n-1)}r^{2}\leq C_{0}^{2}(13/4)\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,0,13r,e_{n})^{1/(n-1)}\leq C\beta^{1/(n-1)}r^{2}. (8.31)

Also, supxM|px|2r2\sup_{x\in M}|\operatorname{\textnormal{{p}}}x|^{2}\leq r^{2}. Since vv is harmonic,

|v(0)|2C(n)Dr|v|2and|v(0)|2C(n)r2Dr|v|2.\displaystyle|v(0)|^{2}\leq C(n)\fint_{\textbf{{D}}_{r}}|v|^{2}\qquad\text{and}\qquad|\nabla^{\prime}v(0)|^{2}\leq\frac{C(n)}{r^{2}}\fint_{\textbf{{D}}_{r}}|v|^{2}. (8.32)

By (8.20) and supDr|u|2C12β1/(n1)r2\sup_{\textbf{{D}}_{r}}|u|^{2}\leq C_{1}^{2}\beta^{1/(n-1)}r^{2} from (8.11), it follows that

|v(0)|2+supxM|px|2|v(0)|2\displaystyle|v(0)|^{2}+\sup_{x\in M}|\operatorname{\textnormal{{p}}}x|^{2}|\nabla^{\prime}v(0)|^{2} C(n)Dr|v|2C(n)(Dr|uv|2+Dr|u|2)\displaystyle\leq C(n)\fint_{\textbf{{D}}_{r}}|v|^{2}\leq C(n)\Big{(}\fint_{\textbf{{D}}_{r}}|u-v|^{2}+\fint_{\textbf{{D}}_{r}}|u|^{2}\Big{)}
C(θn+3β+β1/(n1))r2.\displaystyle\leq C(\theta^{n+3}\beta+\beta^{1/(n-1)})r^{2}. (8.33)

Since n1(MΓ)n1(MΔΓ)βrn1\operatorname{\mathcal{H}^{n-1}}(M\setminus\Gamma)\leq\operatorname{\mathcal{H}^{n-1}}(M\Delta\Gamma)\leq\beta r^{n-1}, we have

C4θr(MΓ)|xν1c1|2dn1(x)Cβrn1(θn+3β+β1/(n1))r2.\displaystyle\int_{\textbf{{C}}_{4\theta r}\cap(M\setminus\Gamma)}|x\cdot\nu_{1}-c_{1}|^{2}\operatorname{\,d\mathcal{H}^{n-1}}(x)\leq C\beta r^{n-1}(\theta^{n+3}\beta+\beta^{1/(n-1)})r^{2}. (8.34)

Choose ε2\varepsilon_{2} to also satisfy

ε21/(n1)θn+3.\displaystyle\boxed{\varepsilon_{2}^{1/(n-1)}\leq\theta^{n+3}}. (8.35)

Then

β1/(n1)θn+3.\displaystyle\beta^{1/(n-1)}\leq\theta^{n+3}. (8.36)

which gives

C4θr(MΓ)|xν1c1|2dn1(x)Cθn+3βrn+1.\displaystyle\int_{\textbf{{C}}_{4\theta r}\cap(M\setminus\Gamma)}|x\cdot\nu_{1}-c_{1}|^{2}\operatorname{\,d\mathcal{H}^{n-1}}(x)\leq C\theta^{n+3}\beta r^{n+1}. (8.37)

Combining these estimates we have

f(E,0,2θr,ν1)\displaystyle\operatorname{\textnormal{{f}}}(E,0,2\theta r,\nu_{1}) C(n)(θr)n+1(C4θrMΓ|xν1c1|2dn1+C4θr(MΓ)|xν1c1|2dn1)\displaystyle\leq\frac{C(n)}{(\theta r)^{n+1}}\bigg{(}\int_{\textbf{{C}}_{4\theta r}\cap M\cap\Gamma}|x\cdot\nu_{1}-c_{1}|^{2}\operatorname{\,d\mathcal{H}^{n-1}}+\int_{C_{4\theta r}\cap(M\setminus\Gamma)}|x\cdot\nu_{1}-c_{1}|^{2}\operatorname{\,d\mathcal{H}^{n-1}}\bigg{)}
Cθ2β.\displaystyle\leq C\theta^{2}\beta. (8.38)

Next, we show that provided ε2\varepsilon_{2} is suitably small, then

eC(E,0,4θr,ν1)ω(1/8,n,λ,Λ,κ,α,r0).\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,0,4\theta r,\nu_{1})\leq\omega(1/8,n,\lambda,\Lambda,\kappa,\alpha,r_{0}). (8.39)

By Proposition 5.5 and Proposition 5.6, we have

eC(E,0,4θr,ν1)(13r4θr)n1eC(E,0,13r,ν1)C~(eC(E,0,13r,en)+|enν1|2).\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,0,4\theta r,\nu_{1})\leq\Big{(}\frac{13r}{4\theta r}\Big{)}^{n-1}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,0,13r,\nu_{1})\leq\widetilde{C}\big{(}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,0,13r,e_{n})+|e_{n}-\nu_{1}|^{2}\big{)}. (8.40)

where C~=C~(n,λ,Λ,κ,α,r0,θ)\widetilde{C}=\widetilde{C}(n,\lambda,\Lambda,\kappa,\alpha,r_{0},\theta) Additionally,

|enν1|2\displaystyle|e_{n}-\nu_{1}|^{2} =|(0,1)(v(0),1)1+|v(0)|2|2\displaystyle=\bigg{|}(0,1)-\frac{(-\nabla^{\prime}v(0),1)}{\sqrt{1+|\nabla^{\prime}v(0)|^{2}}}\bigg{|}^{2}
=|v(0)|2+(1+|v(0)|21)21+|v(0)|2\displaystyle=\frac{|\nabla^{\prime}v(0)|^{2}+(\sqrt{1+|\nabla^{\prime}v(0)|^{2}}-1)^{2}}{1+|\nabla^{\prime}v(0)|^{2}}
2|v(0)|2C(n)r2Dr|v|2\displaystyle\leq 2|\nabla^{\prime}v(0)|^{2}\leq\frac{C(n)}{r^{2}}\fint_{\textbf{{D}}_{r}}|v|^{2}
C(n)r2(Dr|uv|2+Dr|u|2)\displaystyle\leq\frac{C(n)}{r^{2}}\Big{(}\fint_{\textbf{{D}}_{r}}|u-v|^{2}+\fint_{\textbf{{D}}_{r}}|u|^{2}\Big{)}
C(θn+3β+β1/(n1))Cβ1/(n1)\displaystyle\leq C(\theta^{n+3}\beta+\beta^{1/(n-1)})\leq C\beta^{1/(n-1)} (8.41)

where the last several inequalities follow as above. Hence

eC(E,0,4θr,ν1)C~β1/(n1).\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,0,4\theta r,\nu_{1})\leq\widetilde{C}\beta^{1/(n-1)}. (8.42)

for some C~=C~(n,λ,Λ,κ,α,r0,θ)\widetilde{C}=\widetilde{C}(n,\lambda,\Lambda,\kappa,\alpha,r_{0},\theta). We choose ε2\varepsilon_{2} to also satisfy

C~ε21/(n1)ω(1/8,n,λ,Λ,κ,α,r0)\displaystyle\boxed{\widetilde{C}\varepsilon_{2}^{1/(n-1)}\leq\omega(1/8,n,\lambda,\Lambda,\kappa,\alpha,r_{0})} (8.43)

so that

eC(E,0,4θr,ν1)ω(1/8,n,λ,Λ,κ,α,r0)\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,0,4\theta r,\nu_{1})\leq\omega(1/8,n,\lambda,\Lambda,\kappa,\alpha,r_{0}) (8.44)

since βε2\beta\leq\varepsilon_{2}. The reverse Poincaré inequality, Theorem 7.1, implies that

eC(E,0,θr,ν1)\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,0,\theta r,\nu_{1}) C2(f(E,0,2θr,ν1)+(θr)α)\displaystyle\leq C_{2}(\operatorname{\textnormal{{f}}}(E,0,2\theta r,\nu_{1})+(\theta r)^{\alpha})
C(θ2β+θαrα)\displaystyle\leq C(\theta^{2}\beta+\theta^{\alpha}r^{\alpha})
C(θ2eC(E,0,13r,en)+θ2rα/2+θαrα)\displaystyle\leq C(\theta^{2}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,0,13r,e_{n})+\theta^{2}r^{\alpha/2}+\theta^{\alpha}r^{\alpha})
C(θ2eC(E,0,13r,en)+θαrα/2)\displaystyle\leq C(\theta^{2}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,0,13r,e_{n})+\theta^{\alpha}r^{\alpha/2}) (8.45)

as desired. ∎

9. Regularity of Almost-Minimizers

We are almost in the position to prove our main regularity result. All we first need is to prove the following lemma which allows us to remove the assumption A(x0)=IA(x_{0})=I and obtain an excess-decay estimate which we will iterate in the proof of our regularity theorem.

Lemma 9.1.

For each β(0,α/4]\beta\in(0,\alpha/4], there exist positive constants θ1=θ1(n,λ,Λ,κ,α,r0,ACα,β)<1\theta_{1}=\theta_{1}(n,\lambda,\Lambda,\kappa,\alpha,r_{0},||A||_{C^{\alpha}},\beta)<1, ε3=ε3(n,λ,Λ,κ,α,r0,ACα,β)\varepsilon_{3}=\varepsilon_{3}(n,\lambda,\Lambda,\kappa,\alpha,r_{0},||A||_{C^{\alpha}},\beta), and C4=C4(n,λ,Λ,κ,α,r0,ACα,β)C_{4}=C_{4}(n,\lambda,\Lambda,\kappa,\alpha,r_{0},||A||_{C^{\alpha}},\beta) with the following property. Let EE be a (κ,α)(\kappa,\alpha)-almost-minimizer of A\mathscr{F}_{A} in C(x0,r,ν0)\textbf{{C}}(x_{0},r,\nu_{0}) with r<r0r<r_{0} and x0Ex_{0}\in\partial E, and set

eC(E,x,s,ν)=max{eC(E,x,s,ν),sα/2θ1n1+2β}, for xn,s>0,νSSn1.\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x,s,\nu)=\max\Big{\{}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x,s,\nu),\frac{s^{\alpha/2}}{\theta_{1}^{n-1+2\beta}}\Big{\}},\qquad\text{ for }x\in\mathbb{R}^{n},\ s>0,\ \nu\in\SS^{n-1}. (9.1)

If r<r0r<r_{0} and

eC(E,x0,r,ν0)ε3,\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x_{0},r,\nu_{0})\leq\varepsilon_{3}, (9.2)

then there exists ν1SSn1\nu_{1}\in\SS^{n-1} such that

eC(E,x0,θ1r,ν1)\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x_{0},\theta_{1}r,\nu_{1}) θ12βeC(E,x0,r,ν0), and\displaystyle\leq\theta_{1}^{2\beta}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x_{0},r,\nu_{0}),\text{ and } (9.3)
|ν1ν0|2\displaystyle|\nu_{1}-\nu_{0}|^{2} C4eC(E,x0,r,ν0).\displaystyle\leq C_{4}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x_{0},r,\nu_{0}). (9.4)
Proof.

We will eventually make choices for positive constants θ~=θ~(n,λ,Λ,κ,α,r0,ACα,β)<1\widetilde{\theta}=\widetilde{\theta}(n,\lambda,\Lambda,\kappa,\alpha,r_{0},||A||_{C^{\alpha}},\beta)<1 and C~=C~(n,λ,Λ,κ,α,r0,ACα,β)\widetilde{C}=\widetilde{C}(n,\lambda,\Lambda,\kappa,\alpha,r_{0},||A||_{C^{\alpha}},\beta) show that (9.3) holds if we set

θ1=(λ4Λ)1/2θ~andε3=C~1ε2\displaystyle\theta_{1}=\Big{(}\frac{\lambda}{4\Lambda}\Big{)}^{1/2}\widetilde{\theta}\qquad\text{and}\qquad\varepsilon_{3}={\widetilde{C}}^{-1}\varepsilon_{2} (9.5)

where ε2\varepsilon_{2} is the constant from Proposition 8.3 applied with θ=θ~\theta=\widetilde{\theta}.

Since 2βα/22\beta\leq\alpha/2 and θ1<1\theta_{1}<1, we have

(θ1r)α/2θ1n1+2βθ1α/2eC(E,x0,r,ν0)θ12βeC(E,x0,r,ν0).\displaystyle\frac{(\theta_{1}r)^{\alpha/2}}{\theta_{1}^{n-1+2\beta}}\leq\theta_{1}^{\alpha/2}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x_{0},r,\nu_{0})\leq\theta_{1}^{2\beta}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x_{0},r,\nu_{0}). (9.6)

Consequently, we only need to show the existence of ν1SSn1\nu_{1}\in\SS^{n-1} such that

eC(E,x0,θ1r,ν1)θ12βeC(E,x0,r,ν0).\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},\theta_{1}r,\nu_{1})\leq\theta_{1}^{2\beta}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x_{0},r,\nu_{0}). (9.7)

If eC(E,x0,r,ν0)rα/2\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},r,\nu_{0})\leq r^{\alpha/2}, then by Proposition 5.5 we have

eC(E,x0,θ1r,ν0)1θ1n1eC(E,x0,r,ν0)θ12βrα/2θ1n1+2βθ12βeC(E,x0,r,ν0)\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},\theta_{1}r,\nu_{0})\leq\frac{1}{\theta_{1}^{n-1}}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},r,\nu_{0})\leq\theta_{1}^{2\beta}\frac{r^{\alpha/2}}{\theta_{1}^{n-1+2\beta}}\leq\theta_{1}^{2\beta}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x_{0},r,\nu_{0}) (9.8)

and so we can take ν1=ν0\nu_{1}=\nu_{0}. Otherwise, eC(E,x0,r,ν0)rα/2\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},r,\nu_{0})\geq r^{\alpha/2}. We will proceed by applying Proposition 8.3, but we need to use the change of variable Tx0T_{x_{0}} since we are not assuming that A(x0)A(x_{0}) equals the II. This enables us to work with the set Ex0E_{x_{0}} which is an almost-minimizer of Ax0\mathscr{F}_{A_{x_{0}}} with Ax0(x0)=IA_{x_{0}}(x_{0})=I. Let ν~0\widetilde{\nu}_{0} denote the image of ν0\nu_{0} under this change of variable, that is,

ν~0=A1/2(x0)ν0|A1/2(x0)ν0|.\displaystyle\widetilde{\nu}_{0}=\frac{A^{1/2}(x_{0})\nu_{0}}{|A^{1/2}(x_{0})\nu_{0}|}. (9.9)

First note that

C(x0,r(2Λ)1/2,ν~0)B(x0,rΛ1/2) and Wx0(x0,rΛ1/2)B(x0,r)C(x0,r,ν0).\displaystyle\textbf{{C}}\big{(}x_{0},\frac{r}{(2\Lambda)^{1/2}},\widetilde{\nu}_{0}\big{)}\subset\textbf{{B}}\big{(}x_{0},\frac{r}{\Lambda^{1/2}}\big{)}\text{\ \ and \ }\textbf{{W}}_{x_{0}}\big{(}x_{0},\frac{r}{\Lambda^{1/2}}\big{)}\subset\textbf{{B}}(x_{0},r)\subset\textbf{{C}}(x_{0},r,\nu_{0}). (9.10)

Then Ex0E_{x_{0}} is an almost-minimizer of Ax0\mathscr{F}_{A_{x_{0}}} in C(x0,r(2Λ)1/2,ν~0)\textbf{{C}}\big{(}x_{0},\frac{r}{(2\Lambda)^{1/2}},\widetilde{\nu}_{0}\big{)} by Proposition 4.1 since

C(x0,r(2Λ)1/2,ν~0)Tx0(Wx0(x0,rΛ1/2))Tx0(C(x0,r,ν0)).\displaystyle\textbf{{C}}\big{(}x_{0},\frac{r}{(2\Lambda)^{1/2}},\widetilde{\nu}_{0}\big{)}\subset T_{x_{0}}\big{(}\textbf{{W}}_{x_{0}}\big{(}x_{0},\frac{r}{\Lambda^{1/2}}\big{)}\big{)}\subset T_{x_{0}}\big{(}\textbf{{C}}(x_{0},r,\nu_{0})\big{)}. (9.11)

It also follows by (9.10) and Proposition 5.7 that

eC(Ex0,x0,r(2Λ)1/2,ν~0)\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}\big{(}E_{x_{0}},x_{0},\frac{r}{(2\Lambda)^{1/2}},\widetilde{\nu}_{0}\big{)} 2(n1)/2eB(Ex0,x0,rΛ1/2,ν~0)\displaystyle\leq 2^{(n-1)/2}\operatorname{\textnormal{{e}}}_{\textbf{{B}}}\big{(}E_{x_{0}},x_{0},\frac{r}{\Lambda^{1/2}},\widetilde{\nu}_{0}\big{)}
2(n1)/2CeW(E,x0,rΛ1/2,ν0)\displaystyle\leq 2^{(n-1)/2}C\operatorname{\textnormal{{e}}}_{\textbf{{W}}}\big{(}E,x_{0},\frac{r}{\Lambda^{1/2}},\nu_{0}\big{)}
CeC(E,x0,r,ν0).\displaystyle\leq C\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},r,\nu_{0}). (9.12)

Hence by our assumption eC(E,x0,r,ν0)rα/2\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},r,\nu_{0})\geq r^{\alpha/2}

eC(Ex0,x0,r(2Λ)1/2,ν~0)+rα/2(2Λ)α/4\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}\big{(}E_{x_{0}},x_{0},\frac{r}{(2\Lambda)^{1/2}},\widetilde{\nu}_{0}\big{)}+\frac{r^{\alpha/2}}{(2\Lambda)^{\alpha/4}} (C+(2Λ)α/4)eC(E,x0,r,ν0)\displaystyle\leq(C+(2\Lambda)^{-\alpha/4})\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},r,\nu_{0})
CeC(E,x0,r,ν0)\displaystyle\leq C\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},r,\nu_{0})
C~ε3ε2\displaystyle\leq\widetilde{C}\varepsilon_{3}\leq\varepsilon_{2} (9.13)

where at this step we make our choice for C~=C\widetilde{C}=C. Thus Proposition 8.3 applies to Ex0E_{x_{0}} with radius r/(2Λ)1/2r/(2\Lambda)^{1/2} and so there is ν~1SSn1\widetilde{\nu}_{1}\in\SS^{n-1} such that

eC(Ex0,x0,θ~r(2Λ)1/2,ν~1)\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}\big{(}E_{x_{0}},x_{0},\frac{\widetilde{\theta}r}{(2\Lambda)^{1/2}},\widetilde{\nu}_{1}\big{)} C3(θ~2eC(Ex0,x0,r(2Λ)1/2,ν~0)+θ~αrα/2(2Λ)α/4)\displaystyle\leq C_{3}\Big{(}\widetilde{\theta}^{2}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E_{x_{0}},x_{0},\frac{r}{(2\Lambda)^{1/2}},\widetilde{\nu}_{0})+\widetilde{\theta}^{\alpha}\frac{r^{\alpha/2}}{(2\Lambda)^{\alpha/4}}\Big{)}
Cθ~αeC(E,x0,r,ν0).\displaystyle\leq C\widetilde{\theta}^{\alpha}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},r,\nu_{0}). (9.14)

Let ν1SSn1\nu_{1}\in\SS^{n-1} denote the preimage of ν~1\widetilde{\nu}_{1} under the change of variable Tx0T_{x_{0}}, that is,

ν1=A1/2(x0)ν~1|A1/2(x0)ν~1|.\displaystyle\nu_{1}=\frac{A^{-1/2}(x_{0})\widetilde{\nu}_{1}}{|A^{-1/2}(x_{0})\widetilde{\nu}_{1}|}. (9.15)

Note that

C(x0,(λ4Λ)1/2θ~r,ν1)B(x0,(λ2Λ)1/2θ~r)Wx0(x0,θ~r(2Λ)1/2).\displaystyle\textbf{{C}}\big{(}x_{0},\Big{(}\frac{\lambda}{4\Lambda}\Big{)}^{1/2}\widetilde{\theta}r,\nu_{1})\subset\textbf{{B}}\big{(}x_{0},\Big{(}\frac{\lambda}{2\Lambda}\Big{)}^{1/2}\widetilde{\theta}r\big{)}\subset\textbf{{W}}_{x_{0}}\big{(}x_{0},\frac{\widetilde{\theta}r}{(2\Lambda)^{1/2}}\big{)}. (9.16)

So by definition of θ1\theta_{1} and Proposition 5.7, we have that

eC(E,x0,θ1r,ν1)\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},\theta_{1}r,\nu_{1}) (2/λ)(n1)/2eW(E,x0,θ~r(2Λ)1/2,ν1)\displaystyle\leq(2/\lambda)^{(n-1)/2}\operatorname{\textnormal{{e}}}_{\textbf{{W}}}\big{(}E,x_{0},\frac{\widetilde{\theta}r}{(2\Lambda)^{1/2}},\nu_{1}\big{)}
(2/λ)(n1)/2CeB(Ex0,x0,θ~r(2Λ)1/2,ν~1)\displaystyle\leq(2/\lambda)^{(n-1)/2}C\operatorname{\textnormal{{e}}}_{\textbf{{B}}}\big{(}E_{x_{0}},x_{0},\frac{\widetilde{\theta}r}{(2\Lambda)^{1/2}},\widetilde{\nu}_{1}\big{)}
CeC(Ex0,x0,θ~r(2Λ)1/2,ν~1).\displaystyle\leq C\operatorname{\textnormal{{e}}}_{\textbf{{C}}}\big{(}E_{x_{0}},x_{0},\frac{\widetilde{\theta}r}{(2\Lambda)^{1/2}},\widetilde{\nu}_{1}\big{)}. (9.17)

Combining this with (9.14) yields

eC(E,x0,θ1r,ν1)\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},\theta_{1}r,\nu_{1}) Cθ~αeC(E,x0,r,ν0)Cθ~α2βθ12βeC(E,x0,r,ν0).\displaystyle\leq C\widetilde{\theta}^{\alpha}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},r,\nu_{0})\leq C\widetilde{\theta}^{\alpha-2\beta}\theta_{1}^{2\beta}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},r,\nu_{0}). (9.18)

Using this CC we now make our choice for θ~\widetilde{\theta} by setting

θ~=min{1104,(1C)1/(α2β)}.\displaystyle\widetilde{\theta}=\min\Big{\{}\frac{1}{104},\Big{(}\frac{1}{C}\Big{)}^{1/(\alpha-2\beta)}\Big{\}}. (9.19)

The condition θ~(0,1/104]\widetilde{\theta}\in(0,1/104] allows us to apply Proposition 8.3 as above and since Cθ~α2β1C\widetilde{\theta}^{\alpha-2\beta}\leq 1, (9.18) implies

eC(E,x0,θ1r,ν1)θ12βeC(E,x0,r,ν0)θ12βeC(E,x0,r,ν0),\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},\theta_{1}r,\nu_{1})\leq\theta_{1}^{2\beta}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},r,\nu_{0})\leq\theta_{1}^{2\beta}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x_{0},r,\nu_{0}), (9.20)

completing the proof of (9.3).

Now we turn to (9.4). Integrating the inequality |ν1ν0|22|νEν1|2+2|νEν0|2|\nu_{1}-\nu_{0}|^{2}\leq 2|\nu_{E}-\nu_{1}|^{2}+2|\nu_{E}-\nu_{0}|^{2} over the set C(x0,θ1r,ν1)\textbf{{C}}(x_{0},\theta_{1}r,\nu_{1}) which is contained in C(x0,r,ν0)\textbf{{C}}(x_{0},r,\nu_{0}) gives

P(E;C(x0,θ1r,ν1))(θ1r)n1|ν1ν0|24eC(E,x0,θ1r,ν1)+4θ1n1eC(E,x0,r,ν0).\displaystyle\frac{P(E;\textbf{{C}}(x_{0},\theta_{1}r,\nu_{1}))}{(\theta_{1}r)^{n-1}}|\nu_{1}-\nu_{0}|^{2}\leq 4\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},\theta_{1}r,\nu_{1})+\frac{4}{\theta_{1}^{n-1}}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},r,\nu_{0}). (9.21)

The lower density estimate of (4.72) along with (9.3) imply

c|ν1ν0|24(1+θ11n)eC(E,x0,r,ν0)\displaystyle c|\nu_{1}-\nu_{0}|^{2}\leq 4(1+\theta_{1}^{1-n})\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},r,\nu_{0}) (9.22)

completing the proof of (9.4). ∎

Now we prove our main theorem. Before we start, let’s briefly describe the structure of the argument. In the Lipschitz approximation theorem, Theorem 6.1, we saw that given a small excess assumption, there is a Lipschitz function u:n1u\mathrel{\mathop{\mathchar 58\relax}}\mathbb{R}^{n-1}\to\mathbb{R} such that, setting

M=C(x0,r,ν)E and M0={xM:sup0<s<8reC(E,x,s,ν)δ0},\displaystyle M=\textbf{{C}}(x_{0},r,\nu)\cap\partial E\text{ and }M_{0}=\{x\in M\mathrel{\mathop{\mathchar 58\relax}}\sup_{0<s<8r}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x,s,\nu)\leq\delta_{0}\}, (9.23)

the translated graph graph Γ=x0+{(z,u(z)):zDr}\Gamma=x_{0}+\{(z,u(z))\mathrel{\mathop{\mathchar 58\relax}}z\in\textbf{{D}}_{r}\} of uu over Dr\textbf{{D}}_{r} contains M0M_{0}. We proceed by iterating (9.3) at points xMx\in M to obtain a sequence of unit vectors νj(x)\nu_{j}(x) for which certain decay estimates of the excess hold, namely (9.35) and (9.36). Using this, we show that xEx\in\partial^{*}\!E and that νj(x)\nu_{j}(x) converges to νE(x)\nu_{E}(x). Moreover, our iteration gives estimates for Hölder continuity of νE\nu_{E}. Lastly, we show M0M_{0} in fact equals MM, that is, C(x0,r,ν)E\textbf{{C}}(x_{0},r,\nu)\cap\partial E equals the graph of uu. Hölder estimates for u\nabla^{\prime}u follow from the ones for νE\nu_{E}.

Theorem 9.2 (C1,α/4C^{1,\alpha/4}-regularity of almost-minimizers of A\mathscr{F}_{A}).

There exist positive constants
ε4=ε4(n,λ,Λ,κ,α,r0,||A||Cα)\varepsilon_{4}=\varepsilon_{4}(n,\lambda,\Lambda,\kappa,\alpha,r_{0},||A||_{C^{\alpha}}) and C5=C5(n,λ,Λ,κ,α,r0,||A||Cα)C_{5}=C_{5}(n,\lambda,\Lambda,\kappa,\alpha,r_{0},||A||_{C^{\alpha}}) with the following property. If EE is a (κ,α)(\kappa,\alpha)-almost-minimizer of A\mathscr{F}_{A} in C(x0,13r,ν0)\textbf{{C}}(x_{0},13r,\nu_{0}) with 13r<r013r<r_{0} and x0Ex_{0}\in\partial E such that

eC(E,x0,13r,ν0)+rα/2ε4,\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},13r,\nu_{0})+r^{\alpha/2}\leq\varepsilon_{4}, (9.24)

then there exists a Lipschitz function u:n1u\colon\mathbb{R}^{n-1}\to\mathbb{R} with Lip(u)1\operatorname{Lip}(u)\leq 1 satisfying

supn1|u|rC5eC(E,x0,13r,ν0)1/(2(n1))\displaystyle\sup_{\mathbb{R}^{n-1}}\frac{|u|}{r}\leq C_{5}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},13r,\nu_{0})^{1/(2(n-1))} (9.25)

such that

C(x0,r,ν0)E\displaystyle\textbf{{C}}(x_{0},r,\nu_{0})\cap\partial E =x0+{(z,u(z)):zDr},\displaystyle=x_{0}+\big{\{}(z,u(z))\mathrel{\mathop{\mathchar 58\relax}}z\in\textbf{{D}}_{r}\big{\}}, (9.26)
C(x0,r,ν0)E\displaystyle\textbf{{C}}(x_{0},r,\nu_{0})\cap E =x0+{(z,t):zDr,r<t<u(z)}\displaystyle=x_{0}+\big{\{}(z,t)\mathrel{\mathop{\mathchar 58\relax}}z\in\textbf{{D}}_{r},\>-r<t<u(z)\big{\}} (9.27)

and uC1,α/4(Dr)u\in C^{1,\alpha/4}(\textbf{{D}}_{r}) with

|u(z)u(w)|\displaystyle|\nabla^{\prime}u(z)-\nabla^{\prime}u(w)| C5(eC(E,x0,13r,ν0)+rα/2)1/2(|zw|r)α/4,\displaystyle\leq C_{5}\big{(}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},13r,\nu_{0})+r^{\alpha/2}\big{)}^{1/2}\Big{(}\frac{|z-w|}{r}\Big{)}^{\alpha/4}, (9.28)
|νE(x)νE(y)|\displaystyle|\nu_{E}(x)-\nu_{E}(y)| C5(eC(E,x0,13r,ν0)+rα/2)1/2(|xy|r)α/4,\displaystyle\leq C_{5}\big{(}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},13r,\nu_{0})+r^{\alpha/2}\big{)}^{1/2}\Big{(}\frac{|x-y|}{r}\Big{)}^{\alpha/4}, (9.29)

for every z,wDrz,w\in\textbf{{D}}_{r} and x,yC(x0,r,ν0)Ex,y\in\textbf{{C}}(x_{0},r,\nu_{0})\cap\partial E.

Proof.

Without loss of generality we may assume x0=0x_{0}=0. Let θ1<1\theta_{1}<1, ε3\varepsilon_{3}, and C4C_{4} denote the constants from Lemma 9.1 with the choice β=α/4\beta=\alpha/4 which hence depend only on n,λ,Λ,κ,α,r0n,\lambda,\Lambda,\kappa,\alpha,r_{0}, and ||A||Cα||A||_{C^{\alpha}}. As mentioned before, we will choose ε4ε3ε2ε1\boxed{\varepsilon_{4}\leq\varepsilon_{3}}\leq\varepsilon_{2}\leq\varepsilon_{1} and apply the Lipschitz approximation theorem, Theorem 6.1. This gives that there is a Lipschitz function u:n1u\mathrel{\mathop{\mathchar 58\relax}}\mathbb{R}^{n-1}\to\mathbb{R} with Lipu1\operatorname{Lip}u\leq 1 satisfying

supn1|u|rC1eC(E,0,13r,ν0)1/2(n1)\displaystyle\sup_{\mathbb{R}^{n-1}}\frac{|u|}{r}\leq C_{1}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,0,13r,\nu_{0})^{1/2(n-1)} (9.30)

and such that, setting

M=C(0,r,ν0)E and M0={xM:sup0<s<8reC(E,x,s,ν0)δ0},\displaystyle M=\textbf{{C}}(0,r,\nu_{0})\cap\partial E\text{ and }M_{0}=\{x\in M\mathrel{\mathop{\mathchar 58\relax}}\sup_{0<s<8r}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x,s,\nu_{0})\leq\delta_{0}\}, (9.31)

the translated graph graph Γ={(z,u(z)):zDr}\Gamma=\{(z,u(z))\mathrel{\mathop{\mathchar 58\relax}}z\in\textbf{{D}}_{r}\} of uu over Dr\textbf{{D}}_{r} contains M0M_{0}, that is, M0MΓM_{0}\subset M\cap\Gamma. As in Lemma 9.1, we define

eC(E,x,s,ν)=max{eC(E,x,s,ν),sα/2θ1n1+α/2}, for xn,s>0,νSSn1.\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x,s,\nu)=\max\Big{\{}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x,s,\nu),\frac{s^{\alpha/2}}{\theta_{1}^{n-1+\alpha/2}}\Big{\}},\qquad\text{ for }x\in\mathbb{R}^{n},\ s>0,\ \nu\in\SS^{n-1}. (9.32)

Let xMx\in M. Then C(x,8r,ν0)C(0,13r,ν0)\textbf{{C}}(x,8r,\nu_{0})\subset\textbf{{C}}(0,13r,\nu_{0}) and so

eC(E,x,8r,ν0)\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x,8r,\nu_{0}) eC(E,x,8r,ν0)+(8r)α/2θ1n1+α/2(138)n1eC(E,0,13r,ν0)+(8r)α/2θ1n1+α/2\displaystyle\leq\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x,8r,\nu_{0})+\frac{(8r)^{\alpha/2}}{\theta_{1}^{n-1+{\alpha/2}}}\leq\Big{(}\frac{13}{8}\Big{)}^{n-1}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,0,13r,\nu_{0})+\frac{(8r)^{\alpha/2}}{\theta_{1}^{n-1+{\alpha/2}}}
(138)n18α/2θ1n1+α/2(eC(E,0,13r,ν0)+rα/2)=C(eC(E,0,13r,ν0)+rα/2).\displaystyle\leq\Big{(}\frac{13}{8}\Big{)}^{n-1}\frac{8^{\alpha/2}}{\theta_{1}^{n-1+{\alpha/2}}}(\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,0,13r,\nu_{0})+r^{\alpha/2})=C(\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,0,13r,\nu_{0})+r^{\alpha/2}). (9.33)

where C=C(n,λ,Λ,κ,α,r0,||A||Cα)C=C(n,\lambda,\Lambda,\kappa,\alpha,r_{0},||A||_{C^{\alpha}}). For this constant CC, choose ε4\varepsilon_{4} to also satisfy

ε4C1ε3\displaystyle\boxed{\varepsilon_{4}\leq C^{-1}\varepsilon_{3}} (9.34)

so that eC(E,x,8r,ν0)ε3\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x,8r,\nu_{0})\leq\varepsilon_{3}.

Claim.

There exists a sequence {νj(x)}j=1SSn1\{\nu_{j}(x)\}_{j=1}^{\infty}\subset\SS^{n-1} and ν(x)SSn1\nu(x)\in\SS^{n-1} with νj(x)ν(x)\nu_{j}(x)\to\nu(x) such that for every j0j\geq 0,

eC(E,x,θ1j8r,νj(x))\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x,\theta_{1}^{j}8r,\nu_{j}(x)) θ1(α/2)jeC(E,x,8r,ν0)\displaystyle\leq\theta_{1}^{(\alpha/2)j}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x,8r,\nu_{0}) (9.35)
|ν(x)νj(x)|2\displaystyle|\nu(x)-\nu_{j}(x)|^{2} Cθ1(α/2)jeC(E,x,8r,ν0)\displaystyle\leq C\>\theta_{1}^{{(\alpha/2)}j}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x,8r,\nu_{0}) (9.36)

for some constant C=C(n,λ,Λ,κ,α,r0,||A||Cα)C=C(n,\lambda,\Lambda,\kappa,\alpha,r_{0},||A||_{C^{\alpha}}).

Proof of claim.

Since eC(E,x,8r,ν0)ε3\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x,8r,\nu_{0})\leq\varepsilon_{3}, we may apply Lemma 9.1 to find ν1(x)SSn1\nu_{1}(x)\in\SS^{n-1} such that

eC(E,x,θ18r,ν1(x))\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x,\theta_{1}8r,\nu_{1}(x)) θ1α/2eC(E,x,8r,ν0),\displaystyle\leq\theta_{1}^{{\alpha/2}}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x,8r,\nu_{0}), (9.37)
|ν1ν0|2\displaystyle|\nu_{1}-\nu_{0}|^{2} C4eC(E,x,8r,ν0).\displaystyle\leq C_{4}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x,8r,\nu_{0}). (9.38)

In particular, since θ1<1\theta_{1}<1,

eC(E,x,θ18r,ν1(x))eC(E,x,8r,ν0)ε3.\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x,\theta_{1}8r,\nu_{1}(x))\leq\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x,8r,\nu_{0})\leq\varepsilon_{3}. (9.39)

Proceeding inductively we find a sequence {νj(x)}j=0SSn1\{\nu_{j}(x)\}_{j=0}^{\infty}\subset\SS^{n-1} such that

eC(E,x,θ1j+18r,νj+1(x))\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x,\theta_{1}^{j+1}8r,\nu_{j+1}(x)) θ1α/2eC(E,x,θ1j8r,νj(x))ε3,\displaystyle\leq\theta_{1}^{{\alpha/2}}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x,\theta_{1}^{j}8r,\nu_{j}(x))\leq\varepsilon_{3}, (9.40)
|νj+1(x)νj(x)|2\displaystyle|\nu_{j+1}(x)-\nu_{j}(x)|^{2} C4eC(E,x,θ1j8r,νj(x)).\displaystyle\leq C_{4}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x,\theta_{1}^{j}8r,\nu_{j}(x)). (9.41)

for j0j\geq 0. Stringing together the inequalities of (9.40) gives (9.35) and stringing together the inequalities of (9.41) gives

|νj+1(x)νj(x)|2\displaystyle|\nu_{j+1}(x)-\nu_{j}(x)|^{2} C4θ1(α/2)jeC(E,x,8r,ν0)\displaystyle\leq C_{4}\theta_{1}^{{(\alpha/2)}j}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x,8r,\nu_{0}) (9.42)

for j0j\geq 0. Given 0j<h0\leq j<h, it follows that

|νh(x)νj(x)|\displaystyle|\nu_{h}(x)-\nu_{j}(x)| k=jh1|νk+1(x)νk(x)|(C4eC(E,x,8r,ν0))1/2k=jh1θ1(α/4)k\displaystyle\leq\sum_{k=j}^{h-1}|\nu_{k+1}(x)-\nu_{k}(x)|\leq\big{(}C_{4}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x,8r,\nu_{0})\big{)}^{1/2}\sum_{k=j}^{h-1}\theta_{1}^{(\alpha/4)k}
(C4eC(E,x,8r,ν0))1/2k=jθ1(α/4)k=(C4eC(E,x,8r,ν0))1/2θ1(α/4)j1θ1α/4\displaystyle\leq\big{(}C_{4}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x,8r,\nu_{0})\big{)}^{1/2}\sum_{k=j}^{\infty}\theta_{1}^{(\alpha/4)k}=\big{(}C_{4}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x,8r,\nu_{0})\big{)}^{1/2}\frac{\theta_{1}^{(\alpha/4)j}}{1-\theta_{1}^{\alpha/4}} (9.43)

and so

|νh(x)νj(x)|2Cθ1(α/2)jeC(E,x,8r,ν0)\displaystyle|\nu_{h}(x)-\nu_{j}(x)|^{2}\leq C\>\theta_{1}^{{(\alpha/2)}j}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x,8r,\nu_{0}) (9.44)

where C=C(n,λ,Λ,κ,α,r0,||A||Cα)C=C(n,\lambda,\Lambda,\kappa,\alpha,r_{0},||A||_{C^{\alpha}}) since θ1\theta_{1} depends only on these constants too. Hence {νj(x)}j=1SSn1\{\nu_{j}(x)\}_{j=1}^{\infty}\subset\SS^{n-1} is Cauchy and so there is ν(x)SSn1\nu(x)\in\SS^{n-1} such that νj(x)ν(x)\nu_{j}(x)\to\nu(x) as jj\to\infty. Sending hh\to\infty in (9.44) gives (9.36) and this first claim is proved.

Claim.

There is a constant C=C(n,λ,Λ,κ,α,r0,||A||Cα)C=C(n,\lambda,\Lambda,\kappa,\alpha,r_{0},||A||_{C^{\alpha}}) such that

eC(E,x,s,ν(x))\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x,s,\nu(x)) C(sr)α/2eC(E,x,8r,ν0)s(0,4r),\displaystyle\leq C\Big{(}\frac{s}{r}\Big{)}^{{\alpha/2}}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x,8r,\nu_{0})\qquad\forall s\in(0,4r), (9.45)
eC(E,x,s,ν0)\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x,s,\nu_{0}) CeC(E,x,8r,ν0)s(0,8r).\displaystyle\leq C\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x,8r,\nu_{0})\qquad\forall s\in(0,8r). (9.46)
Proof of claim.

Given s(0,4r)s\in(0,4r), there is j0j\geq 0 such that

θ1j+18r<2sθ1j8r.\displaystyle\theta_{1}^{j+1}8r<2s\leq\theta_{1}^{j}8r. (9.47)

Integrating |νEν(x)|22|νEνj(x)|2+2|ν(x)νj(x)|2|\nu_{E}-\nu(x)|^{2}\leq 2|\nu_{E}-\nu_{j}(x)|^{2}+2|\nu(x)-\nu_{j}(x)|^{2} with respect to n1  E\operatorname{\mathcal{H}^{n-1}}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>\partial^{*}\!E over C(x,s,ν(x))C(x,2s,νj(x))\textbf{{C}}(x,s,\nu(x))\subset\textbf{{C}}(x,2s,\nu_{j}(x)), and using the the perimeter bound (4.72) and (9.36), it follows that

eC(E,x,s,ν(x))\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x,s,\nu(x)) 2neC(E,x,2s,νj(x))+C|ν(x)νj(x)|2\displaystyle\leq 2^{n}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x,2s,\nu_{j}(x))+C|\nu(x)-\nu_{j}(x)|^{2}
2n(θ1j8r2s)n1eC(E,x,θ1j8r,νj(x))+C|ν(x)νj(x)|2\displaystyle\leq 2^{n}\Big{(}\frac{\theta_{1}^{j}8r}{2s}\Big{)}^{n-1}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x,\theta_{1}^{j}8r,\nu_{j}(x))+C|\nu(x)-\nu_{j}(x)|^{2}
2n(1θ1)n1eC(E,x,θ1j8r,νj(x))+C|ν(x)νj(x)|2\displaystyle\leq 2^{n}\Big{(}\frac{1}{\theta_{1}}\Big{)}^{n-1}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x,\theta_{1}^{j}8r,\nu_{j}(x))+C|\nu(x)-\nu_{j}(x)|^{2}
Cθ1(α/2)jeC(E,x,8r,ν0)Cθ1α/2(θ1j+1)α/2eC(E,x,8r,ν0)\displaystyle\leq C\theta_{1}^{{(\alpha/2)}j}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x,8r,\nu_{0})\leq\frac{C}{\theta_{1}^{{\alpha/2}}}(\theta_{1}^{j+1})^{{\alpha/2}}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x,8r,\nu_{0})
C(sr)α/2eC(E,x,8r,ν0)\displaystyle\leq C\Big{(}\frac{s}{r}\Big{)}^{{\alpha/2}}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x,8r,\nu_{0}) (9.48)

which is (9.45). Now, take s(0,8r)s\in(0,8r). In the case where s(2r,8r)s\in(2r,8r), it follows that

eC(E,x,s,ν0)(8rs)n1eC(E,x,8r,ν0)4n1eC(E,x,8r,ν0).\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x,s,\nu_{0})\leq\Big{(}\frac{8r}{s}\Big{)}^{n-1}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x,8r,\nu_{0})\leq 4^{n-1}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x,8r,\nu_{0}). (9.49)

Otherwise, s(0,2r)s\in(0,2r) and so integrating |νEν0|22|νEν(x)|2+2|ν(x)ν0|2|\nu_{E}-\nu_{0}|^{2}\leq 2|\nu_{E}-\nu(x)|^{2}+2|\nu(x)-\nu_{0}|^{2} with respect to n1  E\operatorname{\mathcal{H}^{n-1}}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>\partial^{*}\!E over C(x,s,ν0)C(x,2s,ν(x))\textbf{{C}}(x,s,\nu_{0})\subset\textbf{{C}}(x,2s,\nu(x)), using (9.45) with 2s(0,4r)2s\in(0,4r) and (9.36) with j=0j=0 gives

eC(E,x,s,ν0)\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x,s,\nu_{0}) 2neC(E,x,2s,ν(x))+C|ν(x)ν0|2\displaystyle\leq 2^{n}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x,2s,\nu(x))+C|\nu(x)-\nu_{0}|^{2}
2nC(2sr)α/2eC(E,x,8r,ν0)+C|ν(x)ν0|2CeC(E,x,8r,ν0).\displaystyle\leq 2^{n}C\Big{(}\frac{2s}{r}\Big{)}^{{\alpha/2}}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x,8r,\nu_{0})+C|\nu(x)-\nu_{0}|^{2}\leq C\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x,8r,\nu_{0}). (9.50)

Hence (9.46) holds. This completes the proof of our second claim.

Suppose x,yM=C(0,r,ν0)Ex,y\in M=\textbf{{C}}(0,r,\nu_{0})\cap\partial E. Then |xy|<2r|x-y|<\sqrt{2}r and so there is some j0j\geq 0 such that

θ1j+12r|xy|<θ1j2r.\displaystyle\theta_{1}^{j+1}\sqrt{2}r\leq|x-y|<\theta_{1}^{j}\sqrt{2}r. (9.51)

Integrating |νj(x)νj(y)|22|νEνj(x)|2+2|νEνj(y)|2|\nu_{j}(x)-\nu_{j}(y)|^{2}\leq 2|\nu_{E}-\nu_{j}(x)|^{2}+2|\nu_{E}-\nu_{j}(y)|^{2} with respect to n1  E\operatorname{\mathcal{H}^{n-1}}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>\partial^{*}\!E over B(x,θ1jr)C(x,θ1jr,νj(x))C(y,θ1j8r,νj(y))\textbf{{B}}(x,\theta_{1}^{j}r)\subset\textbf{{C}}(x,\theta_{1}^{j}r,\nu_{j}(x))\subset\textbf{{C}}(y,\theta_{1}^{j}8r,\nu_{j}(y)) and using the perimeter bounds (4.72) gives

c|νj(x)νj(y)|2\displaystyle c|\nu_{j}(x)-\nu_{j}(y)|^{2} 4eC(E,x,θ1jr,νj(x))+48n1eC(E,y,θ1j8r,νj(y))\displaystyle\leq 4\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x,\theta_{1}^{j}r,\nu_{j}(x))+4\cdot 8^{n-1}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,y,\theta_{1}^{j}8r,\nu_{j}(y))
48n1eC(E,x,θ1j8r,νj(x))+48n1eC(E,y,θ1j8r,νj(y)).\displaystyle\leq 4\cdot 8^{n-1}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x,\theta_{1}^{j}8r,\nu_{j}(x))+4\cdot 8^{n-1}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,y,\theta_{1}^{j}8r,\nu_{j}(y)). (9.52)

Hence by (9.35) and the definition of eC\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*} we have, eC\operatorname{\textnormal{{e}}}_{\textbf{{C}}}

|νj(x)νj(y)|2\displaystyle|\nu_{j}(x)-\nu_{j}(y)|^{2} Cθ1(α/2)j(eC(E,x,8r,ν0)+eC(E,y,8r,ν0))\displaystyle\leq C\theta_{1}^{(\alpha/2)j}(\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x,8r,\nu_{0})+\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,y,8r,\nu_{0})) (9.53)

By this, (9.36), (9.33), and θ1j+12r|xy|\theta_{1}^{j+1}\sqrt{2}r\leq|x-y|, it follows that

|ν(x)ν(y)|2\displaystyle|\nu(x)-\nu(y)|^{2} 3(|ν(x)νj(x)|2+|νj(x)νj(y)|2+|νj(y)ν(y)|2)\displaystyle\leq 3(|\nu(x)-\nu_{j}(x)|^{2}+|\nu_{j}(x)-\nu_{j}(y)|^{2}+|\nu_{j}(y)-\nu(y)|^{2})
Cθ1(α/2)j(eC(E,x,8r,ν0)+eC(E,y,8r,ν0))\displaystyle\leq C\theta_{1}^{(\alpha/2)j}(\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,x,8r,\nu_{0})+\operatorname{\textnormal{{e}}}_{\textbf{{C}}}^{*}(E,y,8r,\nu_{0}))
Cθ1(α/2)j(eC(E,0,13r,ν0)+rα/2)\displaystyle\leq C\theta_{1}^{(\alpha/2)j}(\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,0,13r,\nu_{0})+r^{\alpha/2})
C(eC(E,0,13r,ν0)+rα/2)(|xy|r)α/2\displaystyle\leq C(\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,0,13r,\nu_{0})+r^{\alpha/2})\Big{(}\frac{|x-y|}{r}\Big{)}^{\alpha/2} (9.54)

and so

|ν(x)ν(y)|C(eC(E,0,13r,ν0)+rα/2)1/2(|xy|r)α/4.\displaystyle|\nu(x)-\nu(y)|\leq C\big{(}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,0,13r,\nu_{0})+r^{\alpha/2}\big{)}^{1/2}\Big{(}\frac{|x-y|}{r}\Big{)}^{\alpha/4}. (9.55)

We now prove xEx\in\partial^{*}\!E and νE(x)=ν(x)\nu_{E}(x)=\nu(x) so that (9.55) becomes (9.29), proving the Hölder continuity of the outer normal to EE.

By (9.45), lims0+eB(E,s,r,ν(x))=0\lim_{s\to 0^{+}}\operatorname{\textnormal{{e}}}_{\textbf{{B}}}(E,s,r,\nu(x))=0. So by perimeter bounds (4.72), we

lims0+1P(E;B(x,s))B(x,s)E|νE(z)ν(x)|22dn1(z)=0.\displaystyle\lim_{s\to 0^{+}}\frac{1}{P(E;\textbf{{B}}(x,s))}\int_{\textbf{{B}}(x,s)\cap\partial^{*}\!E}\frac{|\nu_{E}(z)-\nu(x)|^{2}}{2}\operatorname{\,d\mathcal{H}^{n-1}}(z)=0. (9.56)

Expanding |νE(z)ν(x)|2=|νE(z)|22νE(z)ν(x)+|ν(x)|2=22νE(z)ν(x)|\nu_{E}(z)-\nu(x)|^{2}=|\nu_{E}(z)|^{2}-2\nu_{E}(z)\cdot\nu(x)+|\nu(x)|^{2}=2-2\nu_{E}(z)\cdot\nu(x) implies

ν(x)lims0+1P(E;B(x,s))B(x,s)EνE(z)dn1(z)=1.\displaystyle\nu(x)\cdot\lim_{s\to 0^{+}}\frac{1}{P(E;\textbf{{B}}(x,s))}\int_{\textbf{{B}}(x,s)\cap\partial^{*}\!E}\nu_{E}(z)\operatorname{\,d\mathcal{H}^{n-1}}(z)=1. (9.57)

Since |ν(x)|=1|\nu(x)|=1 and

|lims0+1P(E;B(x,s))B(x,s)EνE(z)dn1(z)|1,\displaystyle\Big{|}\lim_{s\to 0^{+}}\frac{1}{P(E;\textbf{{B}}(x,s))}\int_{\textbf{{B}}(x,s)\cap\partial^{*}\!E}\nu_{E}(z)\operatorname{\,d\mathcal{H}^{n-1}}(z)\Big{|}\leq 1, (9.58)

this implies

ν(x)=lims0+1P(E;B(x,s))B(x,s)EνEdn1.\displaystyle\nu(x)=\lim_{s\to 0^{+}}\frac{1}{P(E;\textbf{{B}}(x,s))}\int_{\textbf{{B}}(x,s)\cap\partial^{*}\!E}\nu_{E}\operatorname{\,d\mathcal{H}^{n-1}}. (9.59)

Since ν(x)SSn1\nu(x)\in\SS^{n-1}, this by definition means xEx\in\partial^{*}\!E with νE(x)=ν(x)\nu_{E}(x)=\nu(x) and hence (9.29) holds.

Combining (9.46) with (9.33) gives

eC(E,x,s,ν0)\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x,s,\nu_{0}) C(eC(E,0,13r,ν0)+rα/2),s(0,8r).\displaystyle\leq C(\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,0,13r,\nu_{0})+r^{\alpha/2}),\qquad\forall s\in(0,8r). (9.60)

Lastly, for this constant CC, we choose ε4\varepsilon_{4} to also satisfy

ε4C1δ0\displaystyle\boxed{\varepsilon_{4}\leq C^{-1}\delta_{0}} (9.61)

where δ0\delta_{0} is the constant from the Lipschitz approximation theorem. It follows for xMx\in M that

sup0<s<8reC(E,x,s,ν0)\displaystyle\sup_{0<s<8r}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x,s,\nu_{0}) δ0\displaystyle\leq\delta_{0} (9.62)

and so M=M0ΓM=M_{0}\subset\Gamma. By the Lipschitz graph criterion, [Mag12, Theorem 23.1], the graph of the Lipschitz function uu coincides with E\partial E in C(0,r,ν0)\textbf{{C}}(0,r,\nu_{0}). Moreover,

νE(x)=(u(px),1)1+|u(px)|2\displaystyle\nu_{E}(x)=\frac{(-\nabla^{\prime}u(\operatorname{\textnormal{{p}}}x),1)}{\sqrt{1+|\nabla^{\prime}u(\operatorname{\textnormal{{p}}}x)|^{2}}} (9.63)

for all xC(0,r,ν0)Ex\in\textbf{{C}}(0,r,\nu_{0})\cap\partial E. Since Lipu1\operatorname{Lip}u\leq 1, for x,yC(0,r,ν0)Ex,y\in\textbf{{C}}(0,r,\nu_{0})\cap\partial E, it follows that

|u(px)u(py)|22|u(px)1+|u(px)|2u(py)1+|u(px)|2|2\displaystyle|\nabla^{\prime}u(\operatorname{\textnormal{{p}}}x)-\nabla^{\prime}u(\operatorname{\textnormal{{p}}}y)|^{2}\leq 2\bigg{|}\frac{\nabla^{\prime}u(\operatorname{\textnormal{{p}}}x)}{\sqrt{1+|\nabla^{\prime}u(\operatorname{\textnormal{{p}}}x)|^{2}}}-\frac{\nabla^{\prime}u(\operatorname{\textnormal{{p}}}y)}{\sqrt{1+|\nabla^{\prime}u(\operatorname{\textnormal{{p}}}x)|^{2}}}\bigg{|}^{2}
4|u(px)1+|u(px)|2u(py)1+|u(py)|2|2+4|u(py)1+|u(py)|2u(py)1+|u(px)|2|2\displaystyle\leq 4\bigg{|}\frac{\nabla^{\prime}u(\operatorname{\textnormal{{p}}}x)}{\sqrt{1+|\nabla^{\prime}u(\operatorname{\textnormal{{p}}}x)|^{2}}}-\frac{\nabla^{\prime}u(\operatorname{\textnormal{{p}}}y)}{\sqrt{1+|\nabla^{\prime}u(\operatorname{\textnormal{{p}}}y)|^{2}}}\bigg{|}^{2}+4\bigg{|}\frac{\nabla^{\prime}u(\operatorname{\textnormal{{p}}}y)}{\sqrt{1+|\nabla^{\prime}u(\operatorname{\textnormal{{p}}}y)|^{2}}}-\frac{\nabla^{\prime}u(\operatorname{\textnormal{{p}}}y)}{\sqrt{1+|\nabla^{\prime}u(\operatorname{\textnormal{{p}}}x)|^{2}}}\bigg{|}^{2} (9.64)
4|νE(x)νE(y)|2\displaystyle\leq 4|\nu_{E}(x)-\nu_{E}(y)|^{2} (9.65)

and |xy|2=|pxpy|2+|u(px)u(py)|22|pxpy|2|x-y|^{2}=|\operatorname{\textnormal{{p}}}x-\operatorname{\textnormal{{p}}}y|^{2}+|u(\operatorname{\textnormal{{p}}}x)-u(\operatorname{\textnormal{{p}}}y)|^{2}\leq 2|\operatorname{\textnormal{{p}}}x-\operatorname{\textnormal{{p}}}y|^{2}. So by (9.29) we have uu is C1,α/4C^{1,\alpha/4} with the estimate (9.28). ∎

Theorem 9.3 (Regularity of the reduced boundary and characterization of the singular set).

If UU is an open set in n\mathbb{R}^{n}, n2n\geq 2, and EE is a (κ,α)(\kappa,\alpha)-almost-minimizer of A\mathscr{F}_{A} in UU, then UEU\cap\partial^{*}\!E is a C1,α/4C^{1,\alpha/4}-hypersurface that is relatively open in UEU\cap\partial E, and it is n1\operatorname{\mathcal{H}^{n-1}}-equivalent to UEU\cap\partial E. Hence the singular set Σ(E;U)\Sigma(E;U) of EE in UU,

Σ(E;U)=U(EE),\displaystyle\Sigma(E;U)=U\cap(\partial E\setminus\partial^{*}\!E), (9.66)

is closed. Moreover, Σ(E;U)\Sigma(E;U) is characterized in terms of the excess as follows:

Σ(E;U)={xUE:inf0<13r<r0,B(x,132r)U(infν0SSn1eC(E,x,13r,ν0)+rα/2)ε4}\displaystyle\Sigma(E;U)=\Big{\{}x\in U\cap\partial E\mathrel{\mathop{\mathchar 58\relax}}\inf_{0<13r<r_{0},\textbf{{B}}(x,13\sqrt{2}r)\subset\subset U}\big{(}\inf_{\nu_{0}\in\SS^{n-1}}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x,13r,\nu_{0})+r^{\alpha/2}\big{)}\geq\varepsilon_{4}\Big{\}} (9.67)

where ε4=ε4(n,λ,Λ,κ,α,r0,||A||Cα)\varepsilon_{4}=\varepsilon_{4}(n,\lambda,\Lambda,\kappa,\alpha,r_{0},||A||_{C^{\alpha}}) is the positive constant from Theorem 9.2.

Proof.

The regularity and relative openness of UEU\cap\partial^{*}\!E follows from Theorem 9.2 and the n1\operatorname{\mathcal{H}^{n-1}}-equivalence follows from Proposition 4.10. Consequently, Σ(E;U)\Sigma(E;U) is closed. Hence all we need to show is (9.67). Consider the set defined by

Σ={xUE:inf0<13r<r0,B(x,132r)U(infν0SSn1eC(E,x,13r,ν0)+rα/2)ε4}.\displaystyle\Sigma=\Big{\{}x\in U\cap\partial E\mathrel{\mathop{\mathchar 58\relax}}\inf_{0<13r<r_{0},\textbf{{B}}(x,13\sqrt{2}r)\subset\subset U}\big{(}\inf_{\nu_{0}\in\SS^{n-1}}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x,13r,\nu_{0})+r^{\alpha/2}\big{)}\geq\varepsilon_{4}\Big{\}}. (9.68)

We show Σ=Σ(E;U)\Sigma=\Sigma(E;U).

By Proposition 5.4, for each xUEx\in U\cap\partial^{*}\!E, we have

limr0+(infν0SSn1eC(E,x,13r,ν0)+rα/2)=0\displaystyle\lim_{r\to 0^{+}}\big{(}\inf_{\nu_{0}\in\SS^{n-1}}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x,13r,\nu_{0})+r^{\alpha/2}\big{)}=0 (9.69)

and so x(UE)Σx\in(U\cap\partial E)\setminus\Sigma. Hence UE(UE)ΣU\cap\partial^{*}\!E\subset(U\cap\partial E)\setminus\Sigma.

If x(UE)Σx\in(U\cap\partial E)\setminus\Sigma, then there is 0<13r<r00<13r<r_{0}, ν0SSn1\nu_{0}\in\SS^{n-1}, with C(x,13r,ν0)U\textbf{{C}}(x,13r,\nu_{0})\subset\subset U such that

eC(E,x,13r,ν0)+rα/2<ε5.\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x,13r,\nu_{0})+r^{\alpha/2}<\varepsilon_{5}. (9.70)

By Theorem 9.2, C(x,r,ν0)E\textbf{{C}}(x,r,\nu_{0})\cap\partial E coincides with the graph of a C1,α/4C^{1,\alpha/4}-function and so xEx\in\partial^{*}\!E. Hence (UE)ΣUE(U\cap\partial E)\setminus\Sigma\subset U\cap\partial^{*}\!E. ∎

Now that we have established regularity of almost-minimizers at points in the reduced boundary, we wish to study the singular set which we do in the next section. However, before we move on to that, we prove the convergence of the outer unit normal vectors along sequences of almost-minimizers and points in the reduced boundaries. The contrapositive of this will be a useful tool in showing that the blow-ups at a singular point must converge to a singular point.

We first need the following lemma regarding almost upper semicontinuity of the excess. Recall from Section 4 the class 𝒜\mathscr{A} of uniformly elliptic, Hölder continuous matrices with respect to the given universal constants and the class \mathscr{M} of (κ,α)(\kappa,\alpha)-almost-minimizers of A\mathscr{F}_{A} with A𝒜A\in\mathscr{A}.

Lemma 9.4 (Almost upper semicontinuity of the excess).

Suppose that {Eh}h\{E_{h}\}_{h\in\mathbb{N}}\subset\mathscr{M} is a sequence of (κ,α)(\kappa,\alpha)-almost-minimizers of Ah\mathscr{F}_{A_{h}} in UU at scale rhr_{h}, r0=lim infhrh>0r_{0}=\liminf_{h\to\infty}r_{h}>0, VUV\subset\subset U is an open set with P(V)<P(V)<\infty such that VEhEV\cap E_{h}\to E for a set EE of finite perimeter, and AhAA_{h}\to A uniformly on compact sets for some A𝒜A\in\mathscr{A}. Furthermore suppose x0VEx_{0}\in V\cap\partial E and r<r0r<r_{0} with A(x0)=IA(x_{0})=I, C(x0,r,ν0)V\textbf{{C}}(x_{0},r,\nu_{0})\subset\subset V, and

n1(EC(x0,r,ν0))=0,\displaystyle\operatorname{\mathcal{H}^{n-1}}(\partial^{*}\!E\cap\partial\textbf{{C}}(x_{0},r,\nu_{0}))=0, (9.71)

then

lim supheC(Eh,x0,r,ν0)eC(E,x0,r,ν0)+Crα.\displaystyle\limsup_{h\to\infty}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E_{h},x_{0},r,\nu_{0})\leq\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},r,\nu_{0})+Cr^{\alpha}. (9.72)

for some positive constant C=C(n,λ,Λ,κ,α,r0,||A||Cα)C=C(n,\lambda,\Lambda,\kappa,\alpha,r_{0},||A||_{C^{\alpha}}).

Proof.

By Proposition 4.13, EE is a (κ,α)(\kappa,\alpha)-almost-minimizer of A\mathscr{F}_{A} in VV at scale r0r_{0}, satisfying

μVEhμE,\displaystyle\mu_{V\cap E_{h}}\overset{\ast}{\rightharpoonup}\mu_{E}, (9.73)
Ah(Eh;)A(E;) in V.\displaystyle\mathscr{F}_{A_{h}}(E_{h};\>\cdot\>)\overset{\ast}{\rightharpoonup}\mathscr{F}_{A}(E;\>\cdot\>)\text{ in }V. (9.74)

We write Cr\textbf{{C}}_{r} for C(x0,r,ν)\textbf{{C}}(x_{0},r,\nu) to simplify notation and claim

lim suphP(Eh;Cr)P(E;Cr)+Crα+n1\displaystyle\limsup_{h\to\infty}P(E_{h};\textbf{{C}}_{r})\leq P(E;\textbf{{C}}_{r})+Cr^{\alpha+n-1} (9.75)

for some C=C(n,λ,Λ,κ,α,r0,||A||Cα)C=C(n,\lambda,\Lambda,\kappa,\alpha,r_{0},||A||_{C^{\alpha}}).

To show this, note

|P(Eh;Cr)P(E;Cr)||P(Eh;Cr)Ah(Eh;Cr)|+|Ah(Eh;Cr)P(E;Cr)|.\displaystyle|P(E_{h};\textbf{{C}}_{r})-P(E;\textbf{{C}}_{r})|\leq|P(E_{h};\textbf{{C}}_{r})-\mathscr{F}_{A_{h}}(E_{h};\textbf{{C}}_{r})|+|\mathscr{F}_{A_{h}}(E_{h};\textbf{{C}}_{r})-P(E;\textbf{{C}}_{r})|. (9.76)

Noting |A(x)I|||A||Cα|xx0|α|A(x)-I|\leq||A||_{\textbf{{C}}^{\alpha}}|x-x_{0}|^{\alpha}, we bound the first term of (9.76) by

|P(Eh;Cr)Ah(Eh;Cr)|\displaystyle|P(E_{h};\textbf{{C}}_{r})-\mathscr{F}_{A_{h}}(E_{h};\textbf{{C}}_{r})| |P(Eh;Cr)A(Eh;Cr)|+|A(Eh;Cr)Ah(Eh;Cr)|\displaystyle\leq|P(E_{h};\textbf{{C}}_{r})-\mathscr{F}_{A}(E_{h};\textbf{{C}}_{r})|+|\mathscr{F}_{A}(E_{h};\textbf{{C}}_{r})-\mathscr{F}_{A_{h}}(E_{h};\textbf{{C}}_{r})|
C||A||CαrαP(Eh;Cr)+||AAh||P(Eh;Cr).\displaystyle\leq C||A||_{C^{\alpha}}r^{\alpha}P(E_{h};\textbf{{C}}_{r})+||A-A_{h}||P(E_{h};\textbf{{C}}_{r}). (9.77)

Similarly, for the second term of (9.76), we have

|Ah(Eh;Cr)P(E;Cr)|\displaystyle|\mathscr{F}_{A_{h}}(E_{h};\textbf{{C}}_{r})-P(E;\textbf{{C}}_{r})| |Ah(Eh;Cr)A(E;Cr)|+|A(E;Cr)P(E;Cr)|\displaystyle\leq|\mathscr{F}_{A_{h}}(E_{h};\textbf{{C}}_{r})-\mathscr{F}_{A}(E;\textbf{{C}}_{r})|+|\mathscr{F}_{A}(E;\textbf{{C}}_{r})-P(E;\textbf{{C}}_{r})|
|Ah(Eh;Cr)A(E;Cr)|+C||A||CαrαP(E;Cr).\displaystyle\leq|\mathscr{F}_{A_{h}}(E_{h};\textbf{{C}}_{r})-\mathscr{F}_{A}(E;\textbf{{C}}_{r})|+C||A||_{C^{\alpha}}r^{\alpha}P(E;\textbf{{C}}_{r}). (9.78)

Since x0VEx_{0}\in V\cap\partial^{*}\!E, by Proposition 4.13 there is a sequence xhVEhx_{h}\in V\cap\partial^{*}\!E_{h} such that xhx0x_{h}\to x_{0}. Given r<sr<s, we have Cr=C(x0,r,ν)C(xh,s,ν)U\textbf{{C}}_{r}=\textbf{{C}}(x_{0},r,\nu)\subset\textbf{{C}}(x_{h},s,\nu)\subset\subset U for large hh. So

lim suphP(Eh;Cr)lim suphP(Eh;C(xh,s,ν))Csn1\displaystyle\limsup_{h\to\infty}P(E_{h};\textbf{{C}}_{r})\leq\limsup_{h\to\infty}P(E_{h};\textbf{{C}}(x_{h},s,\nu))\leq Cs^{n-1} (9.79)

by the upper perimeter bound (4.72) for EhE_{h}. Hence lim suphP(Eh;Cr)Crn1\limsup_{h\to\infty}P(E_{h};\textbf{{C}}_{r})\leq Cr^{n-1}. We also have P(E;Cr)Crn1P(E;\textbf{{C}}_{r})\leq Cr^{n-1}. So by (9.76), (9), and (9), we have

|P(Eh;Cr)P(E;Cr)||Ah(Eh;Cr)A(E;Cr)|+C||A||Cαrα+n1+C||AAh||rn1.\displaystyle|P(E_{h};\textbf{{C}}_{r})-P(E;\textbf{{C}}_{r})|\leq|\mathscr{F}_{A_{h}}(E_{h};\textbf{{C}}_{r})-\mathscr{F}_{A}(E;\textbf{{C}}_{r})|+C||A||_{C^{\alpha}}r^{\alpha+n-1}+C||A-A_{h}||r^{n-1}. (9.80)

Note A(VEh;Cr)=A(Eh;Cr)\mathscr{F}_{A}(V\cap E_{h};\textbf{{C}}_{r})=\mathscr{F}_{A}(E_{h};\textbf{{C}}_{r}) because CrV\textbf{{C}}_{r}\subset\subset V and so A(Eh;Cr)A(E;Cr)\mathscr{F}_{A}(E_{h};\textbf{{C}}_{r})\to\mathscr{F}_{A}(E;\textbf{{C}}_{r}) by (9.74) and n1(ECr)=0\operatorname{\mathcal{H}^{n-1}}(\partial^{*}\!E\cap\partial\textbf{{C}}_{r})=0. This and the uniform convergence of AhAA_{h}\to A on Cr\textbf{{C}}_{r} complete the proof of our claim.

Note μVEh(Cr)=μEh(Cr)\mu_{V\cap E_{h}}(\textbf{{C}}_{r})=\mu_{E_{h}}(\textbf{{C}}_{r}) because CrV\textbf{{C}}_{r}\subset\subset V and so

νμEh(Cr)νμE(Cr)\displaystyle\nu\cdot\mu_{E_{h}}(\textbf{{C}}_{r})\to\nu\cdot\mu_{E}(\textbf{{C}}_{r}) (9.81)

by (9.73) and n1(ECr)=0\operatorname{\mathcal{H}^{n-1}}(\partial^{*}\!E\cap\partial\textbf{{C}}_{r})=0. By |ννE|2=2(1(ννE))|\nu-\nu_{E}|^{2}=2(1-(\nu\cdot\nu_{E})) and |ννEh|2=2(1(ννEh))|\nu-\nu_{E_{h}}|^{2}=2(1-(\nu\cdot\nu_{E_{h}})), we have

eC(E,x0,r,ν)=P(E;Cr)νμE(Cr)rn1andeC(Eh,x0,r,ν)=P(Eh;Cr)νμEh(Cr)rn1.\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},r,\nu)=\frac{P(E;\textbf{{C}}_{r})-\nu\cdot\mu_{E}(\textbf{{C}}_{r})}{r^{n-1}}\qquad\text{and}\qquad\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E_{h},x_{0},r,\nu)=\frac{P(E_{h};\textbf{{C}}_{r})-\nu\cdot\mu_{E_{h}}(\textbf{{C}}_{r})}{r^{n-1}}. (9.82)

From this, (9.75), and (9.81), we obtain (9.72). ∎

Theorem 9.5 (Convergence of outer unit normals).

If {Eh}h\{E_{h}\}_{h\in\mathbb{N}} and EE are (κ,α)(\kappa,\alpha)-almost-minimizers of Ah\mathscr{F}_{A_{h}} and A\mathscr{F}_{A}, respectively, in the open set UnU\subset\mathbb{R}^{n} at scale r0r_{0}, and

EhlocE,AhAuniformly on compact sets,xhUEh,x0UE,xhx0,\displaystyle E_{h}\overset{\text{loc}}{\rightarrow}E,\ \ A_{h}\to A\ \text{uniformly on compact sets},\ \ x_{h}\in U\cap\partial E_{h},\ \ x_{0}\in U\cap\partial^{*}\!E,\ \ x_{h}\to x_{0}, (9.83)

then xhUEhx_{h}\in U\cap\partial^{*}\!E_{h} for hh large enough. Moreover,

limhνEh(xh)=νE(x0).\displaystyle\lim_{h\to\infty}\nu_{E_{h}}(x_{h})=\nu_{E}(x_{0}). (9.84)
Proof.

Considering the translated sets Eh+(x0xh)E_{h}+(x_{0}-x_{h}), note that

νEh(xh)=νEh+(x0xh)(x0),\displaystyle\nu_{E_{h}}(x_{h})=\nu_{E_{h}+(x_{0}-x_{h})}(x_{0}), (9.85)

Eh+(x0xh)locEE_{h}+(x_{0}-x_{h})\overset{\text{loc}}{\rightarrow}E, and Ah(+(x0xh))AA_{h}(\>\cdot\>+(x_{0}-x_{h}))\to A uniformly on compact sets. Hence by replacing EhE_{h} with Eh+(x0xh)E_{h}+(x_{0}-x_{h}) and AhA_{h} with Ah(+(x0xh))A_{h}(\>\cdot\>+(x_{0}-x_{h})), and UU with {xU:dist(x,U)>δ}\{x\in U\mathrel{\mathop{\mathchar 58\relax}}\mathrm{dist}(x,\partial U)>\delta\} for some sufficiently small δ>0\delta>0, we may assume that xh=x0x_{h}=x_{0} for every hh.

By applying the change of variable Tx0T_{x_{0}} on EhE_{h}, EE and AhA_{h}, AA, we may assume without loss of generality that A(x0)=IA(x_{0})=I. Choose an open set VUV\subset\subset U with x0Vx_{0}\in V and P(V)<P(V)<\infty. Lemma 9.4 with EhVlocEVE_{h}\cap V\overset{\text{loc}}{\rightarrow}E\cap V implies there is a constant CC for which

lim supheC(Eh,x0,13r,ν0)+rα/2eC(E,x0,13r,ν0)+Crα/2\displaystyle\limsup_{h\to\infty}\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E_{h},x_{0},13r,\nu_{0})+r^{\alpha/2}\leq\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},13r,\nu_{0})+Cr^{\alpha/2} (9.86)

holds for every r>0r>0 such that C(x0,13r,ν0)V\textbf{{C}}(x_{0},13r,\nu_{0})\subset\subset V and n1(EC(x0,13r,ν0))=0\operatorname{\mathcal{H}^{n-1}}(\partial^{*}\!E\cap\partial\textbf{{C}}(x_{0},13r,\nu_{0}))=0. Since x0UEx_{0}\in U\cap\partial^{*}\!E, by Proposition 5.4 there is r>0r>0 and ν0SSn1\nu_{0}\in\SS^{n-1} with 0<13r<r00<13r<r_{0}, C(x0,13r,ν0)V\textbf{{C}}(x_{0},13r,\nu_{0})\subset\subset V, n1(EC(x0,13r,ν0))=0\operatorname{\mathcal{H}^{n-1}}(\partial^{*}\!E\cap\partial\textbf{{C}}(x_{0},13r,\nu_{0}))=0, and

eC(E,x0,13r,ν0)+Crα/2<ε4\displaystyle\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E,x_{0},13r,\nu_{0})+Cr^{\alpha/2}<\varepsilon_{4} (9.87)

where ε4\varepsilon_{4} is the constant from Theorem 9.2. Then eC(Eh,x0,13r,ν0)+rα/2<ε4\operatorname{\textnormal{{e}}}_{\textbf{{C}}}(E_{h},x_{0},13r,\nu_{0})+r^{\alpha/2}<\varepsilon_{4} for large hh and so by Theorem 9.3, x0UEhx_{0}\in U\cap\partial^{*}\!E_{h} for large hh. Moreover, by Theorem 9.2 there exist Lipschitz functions uh,u;Dru_{h},u;\textbf{{D}}_{r}\to\mathbb{R} with Lipuh,Lipu1\operatorname{Lip}u_{h},\operatorname{Lip}u\leq 1 such that

C(x0,r,ν0)Eh\displaystyle\textbf{{C}}(x_{0},r,\nu_{0})\cap E_{h} =x0+{(z,t):zDr,r<t<uh(z)},\displaystyle=x_{0}+\Big{\{}(z,t)\mathrel{\mathop{\mathchar 58\relax}}z\in\textbf{{D}}_{r},\ -r<t<u_{h}(z)\Big{\}}, (9.88)
C(x0,r,ν0)E\displaystyle\textbf{{C}}(x_{0},r,\nu_{0})\cap E =x0+{(z,t):zDr,r<t<u(z)},\displaystyle=x_{0}+\Big{\{}(z,t)\mathrel{\mathop{\mathchar 58\relax}}z\in\textbf{{D}}_{r},\ -r<t<u(z)\Big{\}}, (9.89)

and such that for z,wDrz,w\in\textbf{{D}}_{r},

|uh(z)uh(w)|\displaystyle|\nabla^{\prime}u_{h}(z)-\nabla^{\prime}u_{h}(w)| C(|zw|r)α/4\displaystyle\leq C\Big{(}\frac{|z-w|}{r}\Big{)}^{\alpha/4} (9.90)

where C=C(n,λ,Λ,κ,α,r0,suph||Ah||Cα)C=C(n,\lambda,\Lambda,\kappa,\alpha,r_{0},\sup_{h}||A_{h}||_{C^{\alpha}}). Then

Dr|uhu|=|(EhΔE)C(x0,r,ν0)|0.\displaystyle\int_{\textbf{{D}}_{r}}|u_{h}-u|=|(E_{h}\Delta E)\cap\textbf{{C}}(x_{0},r,\nu_{0})|\to 0. (9.91)

It follows by integration by parts and the density of Cc1(Dr)C_{c}^{1}(\textbf{{D}}_{r}) in Cc(Dr)C_{c}(\textbf{{D}}_{r}) that

DrφuhDrφu\displaystyle\int_{\textbf{{D}}_{r}}\varphi\nabla^{\prime}u_{h}\to\int_{\textbf{{D}}_{r}}\varphi\nabla^{\prime}u (9.92)

for every φCc(Dr)\varphi\in C_{c}(\textbf{{D}}_{r}). By (9.90), {uh}\{\nabla^{\prime}u_{h}\} is equicontinuous and it is bounded by Lipuh1\operatorname{Lip}{u_{h}}\leq 1. Thus by Arzelà-Ascoli it is compact under uniform convergence. By (9.92), u\nabla^{\prime}u is the only possible limit point of {uh}\{\nabla^{\prime}u_{h}\}. Hence uhu\nabla^{\prime}u_{h}\to\nabla^{\prime}u uniformly on Dr\textbf{{D}}_{r}. Consequently, as x0EhEx_{0}\in\partial^{*}\!E_{h}\cap\partial^{*}\!E, it follows that

νEh(x0)=(uh(0),1)1+|uh(0)|2(u(0),1)1+|u(0)|2=νE(x0)\displaystyle\nu_{E_{h}}(x_{0})=\frac{(-\nabla^{\prime}u_{h}(0),1)}{\sqrt{1+|\nabla^{\prime}u_{h}(0)|^{2}}}\to\frac{(-\nabla^{\prime}u(0),1)}{\sqrt{1+|\nabla^{\prime}u(0)|^{2}}}=\nu_{E}(x_{0}) (9.93)

as desired. ∎

10. Analysis of the Singular Set

In this final section, we turn to the portion of Theorem 1.1 which addresses the size of singular set.

Theorem 10.1 (Dimensional estimates of singular sets of (κ,α)(\kappa,\alpha)-almost-minimizers).

If EE is a (κ,α)(\kappa,\alpha)-almost-minimizer of A\mathscr{F}_{A} in the open set UnU\subset\mathbb{R}^{n} at scale r0r_{0}, then the following hold true:

  1. (i)

    if 2n72\leq n\leq 7, then Σ(E;U)\Sigma(E;U) is empty;

  2. (ii)

    if n=8n=8, then Σ(E;U)\Sigma(E;U) has no accumulation points in UU;

  3. (iii)

    if n9n\geq 9, then s(Σ(E;U))=0\mathcal{H}^{s}(\Sigma(E;U))=0 for every s>n8s>n-8.

This result is known to be sharp in the case of perimeter minimizers in the sense that Simons’ cone,

Σ={x8:x12+x22+x32+x42=x52+x62+x72+x82},\displaystyle\Sigma=\Big{\{}x\in\mathbb{R}^{8}\mathrel{\mathop{\mathchar 58\relax}}x_{1}^{2}+x_{2}^{2}+x_{3}^{2}+x_{4}^{2}=x_{5}^{2}+x_{6}^{2}+x_{7}^{2}+x_{8}^{2}\Big{\}}, (10.1)

is a perimeter minimizer in 8\mathbb{R}^{8} with singular set {0}\{0\}, and for n9n\geq 9, Σ×n8\Sigma\times\mathbb{R}^{n-8} is perimeter minimizer in n\mathbb{R}^{n} that gives n8(Σ×n8)>0\mathcal{H}^{n-8}(\Sigma\times\mathbb{R}^{n-8})>0. Since our surface energies A\mathscr{F}_{A} include perimeter when A=IA=I, our theorem is also sharp.

We use blow-up analysis and a standard Federer dimension reduction argument to prove Theorem 10.1. The next theorem shows the convergence of the singular set along sequences of almost-minimizers. Recall again from Section 4 the class 𝒜\mathscr{A} of uniformly elliptic, Hölder continuous matrices with respect to the given universal constants and the class \mathscr{M} of (κ,α)(\kappa,\alpha)-almost-minimizers of A\mathscr{F}_{A} with A𝒜A\in\mathscr{A}.

Theorem 10.2 (Closure and local uniform convergence of singularities).

If {Eh}h\{E_{h}\}_{h\in\mathbb{N}}\subset\mathscr{M} and EE\in\mathscr{M} are (κ,α)(\kappa,\alpha)-almost-minimizers of Ah\mathscr{F}_{A_{h}} and A\mathscr{F}_{A}, respectively, in the open set UU at scale r0r_{0}, and

EhlocE,AhA uniformly on compact sets,xhΣ(Eh;U),x0UE,xhx0,\displaystyle E_{h}\overset{\text{loc}}{\rightarrow}E,\ \ A_{h}\to A\text{ uniformly on compact sets},\ \ x_{h}\in\Sigma(E_{h};U),\ \ x_{0}\in U\cap\partial E,\ \ x_{h}\to x_{0}, (10.2)

then x0Σ(E;U)x_{0}\in\Sigma(E;U). Moreover, given ε>0\varepsilon>0 and HUH\subset U compact, then

Σ(Eh;U)HIε(Σ(E;U)H)\displaystyle\Sigma(E_{h};U)\cap H\subset I_{\varepsilon}(\Sigma(E;U)\cap H) (10.3)

for all large hh where IεI_{\varepsilon} denotes the ε\varepsilon neighborhood of a set.

Proof.

It must be that x0Σ(E;U)x_{0}\in\Sigma(E;U) since x0UEx_{0}\in U\cap\partial^{*}\!E would contradict Theorem 9.5. We prove (10.3) by contradiction. Indeed, assume there exist ε>0\varepsilon>0, HUH\subset U compact, h(k)h(k)\to\infty as kk\to\infty, and ykΣ(Eh(k);U)Hy_{k}\in\Sigma(E_{h(k)};U)\cap H such that dist(yk,Σ(E;U)H)ε\mathrm{dist}(y_{k},\Sigma(E;U)\cap H)\geq\varepsilon. By compactness of HH and reducing to a further subsequence, we may assume yky0y_{k}\to y_{0} for some y0HUy_{0}\in H\subset U. By Proposition 4.13 (i)(i), we have y0UEy_{0}\in U\cap\partial E. By the first part of this theorem, we have y0Σ(E;U)y_{0}\in\Sigma(E;U). This implies ykIε(Σ(E;U)H)y_{k}\in I_{\varepsilon}(\Sigma(E;U)\cap H) for large kk, a contradiction. ∎

10.1. Existence of blow-up limits

We now prove the existence of blow-up limits along subsequences and their convergence to singular minimizing cones. We say that FF is a cone with vertex x0x_{0} if it is invariant under blow-ups at x0x_{0}, that is, if

F=Fx0,r=Fx0r\displaystyle F=F_{x_{0},r}=\frac{F-x_{0}}{r} (10.4)

for all r>0r>0. If FF is a cone which is a (global) minimizer of A\mathscr{F}_{A} in n\mathbb{R}^{n} and Σ(F)=Σ(F;n)\Sigma(F)=\Sigma(F;\mathbb{R}^{n})\not=\varnothing, we say that FF is a singular minimizing cone of A\mathscr{F}_{A}. A singular minimizing cone of perimeter we simply refer to as a singular minimizing cone.

Note that if FF is a singular minimizing cone, then 0Σ(F)0\in\Sigma(F), for otherwise the blow-ups F0,rF_{0,r} would converge to a half-space, implying that F=F0,rF=F_{0,r} is a half-space, in contradiction with Σ(F)\Sigma(F)\not=\varnothing.

Theorem 10.3 (Existence of blow-up limits at singular points).

If EE is a (κ,α)(\kappa,\alpha)-almost-minimizer of A\mathscr{F}_{A} in the open set UU at scale r0r_{0}, x0Σ(E;U)x_{0}\in\Sigma(E;U), and rh0+r_{h}\to 0^{+}, then, setting

Eh=Ex0,rh=Ex0rh,andAh(x)=Ax0,rh(y)=A(rhx+x0),\displaystyle E_{h}=E_{x_{0},r_{h}}=\frac{E-x_{0}}{r_{h}},\qquad\text{and}\qquad A_{h}(x)=A_{x_{0},r_{h}}(y)=A(r_{h}x+x_{0}), (10.5)

there exist h(k)h(k)\to\infty as kk\to\infty and a set of locally finite perimeter FF in n\mathbb{R}^{n} such that

Eh(k)locF,μEh(k)μK,Ah(k)(Eh(k);)A(x0)(F;) on bounded subsets of n\displaystyle E_{h(k)}\overset{\text{loc}}{\rightarrow}F,\qquad\mu_{E_{h(k)}}\overset{\ast}{\rightharpoonup}\mu_{K},\qquad\mathscr{F}_{A_{h(k)}}(E_{h(k)};\>\cdot\>)\overset{\ast}{\rightharpoonup}\mathscr{F}_{A(x_{0})}(F;\>\cdot\>)\text{ on bounded subsets of }\mathbb{R}^{n} (10.6)

and FF is singular minimizing cone of A(x0)\mathscr{F}_{A(x_{0})} in n\mathbb{R}^{n} with vertex 0.

Proof.

The change of variable Tx0T_{x_{0}} allows us to assume without loss of generality that A(x0)=IA(x_{0})=I since the convergence properties and cones are preserved under this affine transformation. For each R>0R>0, BR\textbf{{B}}_{R} is eventually contained in Ux0,rhU_{x_{0},r_{h}} for large hh. Note that EhE_{h} is a (κrhα,α)(\kappa r_{h}^{\alpha},\alpha)-almost-minimizer of Ah\mathscr{F}_{A_{h}} in Ux0,rhU_{x_{0},r_{h}} at scale r0/rhr_{0}/r_{h} by Proposition 4.2 and ||Ah||Cαrhα||A||Cα||A_{h}||_{C^{\alpha}}\leq r_{h}^{\alpha}||A||_{C^{\alpha}}. Hence ||Ah||CαM1||A_{h}||_{C^{\alpha}}\leq M_{1} and ||Ah(x)||||A(rhx+x0)||M2||A_{h}(x)||\leq||A(r_{h}x+x_{0})||\leq M_{2} for some positive constants M1M_{1} and M2M_{2}. Once rhα1r_{h}^{\alpha}\leq 1, we have EhE_{h} is a (κ,α)(\kappa,\alpha)-almost-minimizer of Ah\mathscr{F}_{A_{h}}. Thus we may apply the compactness of Proposition 4.12 and Proposition 4.13 to obtain h(k)h(k)\to\infty as kk\to\infty, a set of locally finite perimeter FF, and a uniformly elliptic, Hölder continuous matrix AA_{\infty} such that BREh(k)F\textbf{{B}}_{R}\cap E_{h(k)}\to F and Ah(k)AA_{h(k)}\to A_{\infty} uniformly on compact sets and FF is a minimizer of A\mathscr{F}_{A_{\infty}} in BR\textbf{{B}}_{R} at scale lim infkr0/rh(k)=\liminf_{k\to\infty}r_{0}/r_{h(k)}=\infty. Note that A(x)=limkA(rh(k)x+x0)=A(x0)=IA_{\infty}(x)=\lim_{k\to\infty}A(r_{h(k)}x+x_{0})=A(x_{0})=I. Hence A=P\mathscr{F}_{A_{\infty}}=P. By Proposition 4.13 and a diagonalization argument, we obtain a subsequence such that up to relabeling

Eh(k)locF,\displaystyle E_{h(k)}\overset{\text{loc}}{\rightarrow}F,
Ah(k)I uniformly on compact sets,\displaystyle A_{h(k)}\to I\text{ uniformly on compact sets},
F is a (global) minimizer of perimeter in n,\displaystyle F\text{ is a (global) minimizer of perimeter in }\mathbb{R}^{n},
Ah(k)(Eh(k);)P(F;) on bounded subsets of n.\displaystyle\mathscr{F}_{A_{h(k)}}(E_{h(k)};\>\cdot\>)\overset{\ast}{\rightharpoonup}P(F;\>\cdot\>)\text{ on bounded subsets of }\mathbb{R}^{n}. (10.7)

By Theorem 10.2 we have 0Σ(F)0\in\Sigma(F) because 0Σ(Eh(k);U)0\in\Sigma(E_{h(k)};U). All that remains is to show that FF is a cone with vertex 0. Choose one of the a.e. r>0r>0 for which we have n1(FBr)=0\operatorname{\mathcal{H}^{n-1}}(\partial^{*}\!F\cap\partial\textbf{{B}}_{r})=0. By (10.1), Proposition 4.2, and Corollary 4.8, it follows that

P(F;Br)\displaystyle P(F;\textbf{{B}}_{r}) =limkAh(k)(Eh(k);Br)\displaystyle=\lim_{k\to\infty}\mathscr{F}_{A_{h(k)}}(E_{h(k)};\textbf{{B}}_{r})
=limkA(E;B(x0,rrh(k)))rh(k)n1=rn1θA(E,x0)\displaystyle=\lim_{k\to\infty}\frac{\mathscr{F}_{A}(E;\textbf{{B}}(x_{0},rr_{h(k)}))}{r_{h(k)}^{n-1}}=r^{n-1}\theta_{A}(E,x_{0}) (10.8)

since Wx0(x0,rrh(k))=B(x0,rrh(k))\textbf{{W}}_{x_{0}}(x_{0},rr_{h(k)})=\textbf{{B}}(x_{0},rr_{h(k)}) as A(x0)=IA(x_{0})=I. Hence

P(F;Br)rn1=θA(E,x0)for a.e. r>0.\displaystyle\frac{P(F;\textbf{{B}}_{r})}{r^{n-1}}=\theta_{A}(E,x_{0})\qquad\text{for a.e. }r>0. (10.9)

The monotonicity formula for perimeter minimizers [Mag12, Theorem 28.9] gives

ddrP(F;Br)rn1=ddrBrF(νF(x)x)2|x|n+1dn1(x)for a.e. r>0.\displaystyle\frac{d}{dr}\frac{P(F;\textbf{{B}}_{r})}{r^{n-1}}=\frac{d}{dr}\int_{\textbf{{B}}_{r}\cap\partial^{*}\!F}\frac{\big{(}\nu_{F}(x)\cdot x\big{)}^{2}}{|x|^{n+1}}\operatorname{\,d\mathcal{H}^{n-1}}(x)\qquad\text{for a.e. }r>0. (10.10)

So (10.9) implies

ddrBrF(νF(x)x)2|x|n+1dn1(x)=0for a.e. r>0.\displaystyle\frac{d}{dr}\int_{\textbf{{B}}_{r}\cap\partial^{*}\!F}\frac{\big{(}\nu_{F}(x)\cdot x\big{)}^{2}}{|x|^{n+1}}\operatorname{\,d\mathcal{H}^{n-1}}(x)=0\qquad\text{for a.e. }r>0. (10.11)

Hence

BsF(νF(x)x)2|x|n+1dn1(x)=BrF(νF(x)x)2|x|n+1dn1(x)\displaystyle\int_{\textbf{{B}}_{s}\cap\partial^{*}\!F}\frac{\big{(}\nu_{F}(x)\cdot x\big{)}^{2}}{|x|^{n+1}}\operatorname{\,d\mathcal{H}^{n-1}}(x)=\int_{\textbf{{B}}_{r}\cap\partial^{*}\!F}\frac{\big{(}\nu_{F}(x)\cdot x\big{)}^{2}}{|x|^{n+1}}\operatorname{\,d\mathcal{H}^{n-1}}(x) (10.12)

for a.e. 0<s<r0<s<r, and thus

(BrB¯s)F(νF(x)x)2|x|n+1dn1(x)=0.\displaystyle\int_{(\textbf{{B}}_{r}\setminus\overline{\textbf{{B}}}_{s})\cap\partial^{*}\!F}\frac{\big{(}\nu_{F}(x)\cdot x\big{)}^{2}}{|x|^{n+1}}\operatorname{\,d\mathcal{H}^{n-1}}(x)=0. (10.13)

This implies νF(x)x=0\nu_{F}(x)\cdot x=0 for n1\operatorname{\mathcal{H}^{n-1}}-a.e. xFx\in\partial^{*}\!F. Thus F(1)F^{(1)} is a cone with vertex 0 by [Mag12, Proposition 28.8]. ∎

10.2. Dimension reduction argument

The next few results we recall from the standard argument for the characterization of the singular set for perimeter minimizers which we use for our adapted proof. We refer readers to [Mag12, Chapter 28] for proofs.

Theorem 10.4.

If n2n\geq 2 and there exists a singular minimizing cone FnF\subset\mathbb{R}^{n} with Σ(F)={0}\Sigma(F)=\{0\}, then n8n\geq 8.

Theorem 10.5 (Dimension reduction theorem).

If FF is a singular minimizing cone in n\mathbb{R}^{n}, x0Σ(F)x_{0}\in\Sigma(F), x00x_{0}\not=0, and if rh0+r_{h}\to 0^{+}, then, up to extracting a subsequence and up to rotation, the blow-ups Fx0,rhF_{x_{0},r_{h}} locally converge to a cylinder G×G\times\mathbb{R}, where GG is a singular minimizing cone in n1\mathbb{R}^{n-1}.

Lemma 10.6 (Half-lines of singular points).

If FF is a singular minimizing cone in n\mathbb{R}^{n}, x0Σ(F)x_{0}\in\Sigma(F), and x00x_{0}\not=0, then {tx0:t>0}Σ(F)\{t\>x_{0}\mathrel{\mathop{\mathchar 58\relax}}t>0\}\subset\Sigma(F) and n3n\geq 3.

Lemma 10.7 (Cylinders of locally finite perimeter).
  1. (i)

    If FF is a set of locally finite perimeter in n1\mathbb{R}^{n-1}, then F×F\times\mathbb{R} is of locally finite perimeter in n\mathbb{R}^{n}, with

    μF×=(νF(px),0)n1 ((F)×).\displaystyle\mu_{F\times\mathbb{R}}=(\nu_{F}(\operatorname{\textnormal{{p}}}x),0)\operatorname{\mathcal{H}^{n-1}}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>\big{(}(\partial^{*}\!F)\times\mathbb{R}\big{)}. (10.14)

    Moreover, if FF is a perimeter minimizing in n1\mathbb{R}^{n-1}, then F×F\times\mathbb{R} is a perimeter minimizer in n\mathbb{R}^{n}.

  2. (ii)

    If EE is a set of locally finite perimeter in n\mathbb{R}^{n} such that

    νE(x)en=0for n1a.e.xE,\displaystyle\nu_{E}(x)\cdot e_{n}=0\qquad\text{for }\operatorname{\mathcal{H}^{n-1}}-a.e.\ x\in\partial^{*}\!E, (10.15)

    then there exists a set of locally finite perimeter FF in n1\mathbb{R}^{n-1} such that EE is equivalent to F×F\times\mathbb{R}. If, moreover, EE is a perimeter minimizer in n\mathbb{R}^{n}, then FF is a perimeter minimizer in n1\mathbb{R}^{n-1}.

Lemma 10.8.
  1. (i)

    If EE is a Borel set such that s(E)<\mathcal{H}^{s}(E)<\infty, s>0s>0, then

    lim supr0+s(EB(x,r))ωsrs12s,for sa.e.xE\displaystyle\limsup_{r\to 0^{+}}\frac{\mathcal{H}_{\infty}^{s}(E\cap\textbf{{B}}(x,r))}{\omega_{s}r^{s}}\geq\frac{1}{2^{s}},\qquad\text{for }\mathcal{H}^{s}-a.e.\ x\in E (10.16)
  2. (ii)

    If EE is an (κ,α)(\kappa,\alpha)-almost-minimizer of A\mathscr{F}_{A} in the open set UnU\subset\mathbb{R}^{n} at scale r0r_{0}, and rh0+r_{h}\to 0^{+}, then, setting

    Eh=Ex0,rh=Ex0rh,andAh(x)=Ax0,rh(x)=A(rhx+x0),\displaystyle E_{h}=E_{x_{0},r_{h}}=\frac{E-x_{0}}{r_{h}},\qquad\text{and}\qquad A_{h}(x)=A_{x_{0},r_{h}}(x)=A(r_{h}x+x_{0}), (10.17)

    we have

    s(Σ(E;U)H)lim suphs(Σ(Eh;U)H)\displaystyle\mathcal{H}_{\infty}^{s}(\Sigma(E;U)\cap H)\geq\limsup_{h\to\infty}\mathcal{H}_{\infty}^{s}(\Sigma(E_{h};U)\cap H) (10.18)

    for every compact set HUH\subset U.

  3. (iii)

    If s0s\geq 0, Fn1F\subset\mathbb{R}^{n-1}, and s(F)=0\mathcal{H}_{\infty}^{s}(F)=0, then s+1(F×)=0\mathcal{H}_{\infty}^{s+1}(F\times\mathbb{R})=0.

Proof.

(i)(i) and (iii)(iii) are proved in [Mag12, Lemma 28.14] and we now adapt the proof of his version of (ii)(ii) to the case of (κ,α)(\kappa,\alpha)-almost-minimizers.

Let \mathcal{F} be a finite covering by open sets of the compact set Σ(E;U)H\Sigma(E;U)\cap H. Then there exists ε>0\varepsilon>0 such that Iε(Σ(E;U)H)FFI_{\varepsilon}(\Sigma(E;U)\cap H)\subset\bigcup_{F\in\mathcal{F}}F. Eventually UUx0,rhU\subset U_{x_{0},r_{h}} and rhα1r_{h}^{\alpha}\leq 1, and so by Proposition 4.2 EhE_{h} is a (κ,α)(\kappa,\alpha)-almost-minimizer of Ah\mathscr{F}_{A_{h}} in UU. Then by Theorem 10.2,

Σ(Eh;U)HIε(Σ(E;U)H)FF\displaystyle\Sigma(E_{h};U)\cap H\subset I_{\varepsilon}(\Sigma(E;U)\cap H)\subset\bigcup_{F\in\mathcal{F}}F (10.19)

for large hh. Hence by definition of s\mathcal{H}_{\infty}^{s} it follows that

lim suphs(Σ(E;U)H)s(FF)ωsF(diam(F)2)s\displaystyle\limsup_{h\to\infty}\mathcal{H}_{\infty}^{s}(\Sigma(E;U)\cap H)\leq\mathcal{H}_{\infty}^{s}(\bigcup_{F\in\mathcal{F}}F)\leq\omega_{s}\sum_{F\in\mathcal{F}}\Big{(}\frac{\operatorname{diam}(F)}{2}\Big{)}^{s} (10.20)

Taking the infimum over all such coverings \mathcal{F} proves the result. ∎

Lastly we recall a technical result [Mat95, Theorem 8.16] before we jump into the proof of Theorem 10.1.

Lemma 10.9.

If EnE\subset\mathbb{R}^{n} and s(E)>0\mathcal{H}^{s}(E)>0, then there exists FEF\subset E with 0<s(F)<0<\mathcal{H}^{s}(F)<\infty.

Proof of Theorem 10.1.

(i)(i) Let EE be a (κ,α)(\kappa,\alpha)-almost-minimizer of A\mathscr{F}_{A} in UU with 2n72\leq n\leq 7. By way of contradiction, suppose there exists x0Σ(E;U)x_{0}\in\Sigma(E;U). As usual we may assume without loss of generality that A(x0)=IA(x_{0})=I by the change of variable Tx0T_{x_{0}}. By Theorem 10.3 there exists a singular minimizing cone FF in n\mathbb{R}^{n}, but this contradicts Simons’ theorem on the nonexistence of singular minimizing cones in dimensions 2n72\leq n\leq 7, see [Mag12, Theorem 28.1 (i)].

(ii)(ii) Let E8E\subset\mathbb{R}^{8} be a (κ,α)(\kappa,\alpha)-almost-minimizer of A\mathscr{F}_{A} in UU. By way of contradiction, suppose x0UEx_{0}\in U\cap\partial E is an accumulation point of Σ(E;U)\Sigma(E;U). Then there is a sequence xhΣ(E;U)x_{h}\in\Sigma(E;U) such that xhx0x_{h}\to x_{0}. Again we may assume without loss of generality that A(x0)=IA(x_{0})=I. Set rh=|xhx0|r_{h}=|x_{h}-x_{0}| and consider the blow-ups Eh=Ex0,rhE_{h}=E_{x_{0},r_{h}}. By Theorem 10.3 there is a subsequence, which upon relabeling as EhE_{h}, converges locally to a singular minimizing cone FF in n\mathbb{R}^{n}. Let yh=(xhx0)/rhy_{h}=(x_{h}-x_{0})/r_{h}. Then yhSSn1y_{h}\in\SS^{n-1} and so by compactness there exists y0SSn1y_{0}\in\SS^{n-1} and a further subsequence so that, up to relabeling, we have yhy0y_{h}\to y_{0}. Note that yhΣ(Eh;Ux0,rh)y_{h}\in\Sigma(E_{h};U_{x_{0},r_{h}}) as xhΣ(E;U)x_{h}\in\Sigma(E;U). So by Theorem 10.2 we have y0Σ(F)y_{0}\in\Sigma(F). Since y00y_{0}\not=0, we have 8(F)>1\mathcal{H}^{8}(F)>1 and so by Theorem 10.5 there exists a singular minimizing cone GG in 7\mathbb{R}^{7}, contradicting (i)(i).

(iii)(iii) Let EnE\subset\mathbb{R}^{n} be a (κ,α)(\kappa,\alpha)-almost-minimizer of A\mathscr{F}_{A} in UU and suppose s(Σ(E;U))>0\mathcal{H}^{s}(\Sigma(E;U))>0 with s>0s>0. Then there exists x0Σ(E;U)x_{0}\in\Sigma(E;U). By the change of variable Tx0T_{x_{0}}, we may assume without loss of generality that A(x0)=IA(x_{0})=I. By [Mat95, Theorem 8.16] and Lemma 10.8 (i)(i), there exists rh0+r_{h}\to 0^{+} such that

s(Σ(E;U)B¯(x0,rh))ωsrhs2s+1\displaystyle\mathcal{H}_{\infty}^{s}(\Sigma(E;U)\cap\overline{\textbf{{B}}}(x_{0},r_{h}))\geq\frac{\omega_{s}r_{h}^{s}}{2^{s+1}} (10.21)

for all hh\in\mathbb{N}. This is equivalently rewritten in terms of the blow-ups Eh=Ex0,rhE_{h}=E_{x_{0},r_{h}} as

s(Σ(Eh;Ux0,rh)B¯1)ωs2s+1\displaystyle\mathcal{H}_{\infty}^{s}(\Sigma(E_{h};U_{x_{0},r_{h}})\cap\overline{\textbf{{B}}}_{1})\geq\frac{\omega_{s}}{2^{s+1}} (10.22)

for all hh\in\mathbb{N}. Eventually B2Ux0,rh\textbf{{B}}_{2}\subset U_{x_{0},r_{h}} when hh is sufficiently and thus

s(Σ(Eh;B2)B¯1)ωs2s+1.\displaystyle\mathcal{H}_{\infty}^{s}(\Sigma(E_{h};\textbf{{B}}_{2})\cap\overline{\textbf{{B}}}_{1})\geq\frac{\omega_{s}}{2^{s+1}}. (10.23)

By Theorem 10.3 there exists a subsequence, which up relabeling as EhE_{h}, converges locally to a singular minimizing cone FF in n\mathbb{R}^{n}. By Lemma 10.8 (ii)(ii) we have

s(Σ(F)B¯1)lim suphs(Σ(Eh;B2)B¯1)ωs2s+1.\displaystyle\mathcal{H}_{\infty}^{s}(\Sigma(F)\cap\overline{\textbf{{B}}}_{1})\geq\limsup_{h\to\infty}\mathcal{H}_{\infty}^{s}(\Sigma(E_{h};\textbf{{B}}_{2})\cap\overline{\textbf{{B}}}_{1})\geq\frac{\omega_{s}}{2^{s+1}}. (10.24)

We may apply the above argument with FF and n\mathbb{R}^{n} in place of EE and UU. By Theorem 10.5 there exists a singular minimizing cone G×G\times\mathbb{R}

s(Σ(G×)B¯1)ωs2s+1.\displaystyle\mathcal{H}_{\infty}^{s}(\Sigma(G\times\mathbb{R})\cap\overline{\textbf{{B}}}_{1})\geq\frac{\omega_{s}}{2^{s+1}}. (10.25)

By Lemma 10.8 (iii)(iii) we must have s1(Σ(G))>0\mathcal{H}^{s-1}(\Sigma(G))>0. If we now assume that n9n\geq 9 and s>n8s>n-8, repeating this argument n8n-8 times gives the existence of a singular minimizing cone GG in 8\mathbb{R}^{8} with s(n8)(Σ(G))>0\mathcal{H}^{s-(n-8)}(\Sigma(G))>0, in contradiction to (ii)(ii). Thus we conclude that sn8s\leq n-8. ∎

A. Appendix A

In this appendix we provide a proof for the change of variable formula for sets of locally finite perimeter with bounded, continuous integrands depending on both xx and νE\nu_{E}. This is a generalization of [Mag12, Proposition 17.1] and [KMS19, Theorem A.1].

Proposition A.1 (Change of variable for sets of locally finite perimeter).

Suppose ff is a diffeomorphism of n\mathbb{R}^{n} and denote g=f1g=f^{-1}. If EE is a set of locally finite perimeter in n\mathbb{R}^{n}, then f(E)f(E) is a set of locally finite perimeter in n\mathbb{R}^{n} such that

f(E)=f(E)andνf(E)(f(x))=(gf)tνE(x)|(gf)tνE(x)| for all xE.\displaystyle\partial^{*}\!f(E)=f(\partial^{*}\!E)\qquad\text{and}\qquad\nu_{f(E)}(f(x))=\frac{(\nabla g\circ f)^{t}\nu_{E}(x)}{|(\nabla g\circ f)^{t}\nu_{E}(x)|}\text{ for all }x\in\partial^{*}\!E. (A.1)

If Φ:n×SSn1[0,)\Phi\colon\mathbb{R}^{n}\times\SS^{n-1}\to[0,\infty) is a bounded and continuous function and UU is an open, bounded set satisfying n1(Uf(E))=0\operatorname{\mathcal{H}^{n-1}}(\partial U\cap\partial^{*}\!f(E))=0, and f\nabla f and g\nabla g are bounded, then the change of variable y=f(x)y=f(x) gives

Uf(E)Φ(y,νf(E))dn1(y)=g(U)EΦ(f(x),(gf)tνE|(gf)tνE|)Jf|(gf)tνE|dn1(x).\displaystyle\int_{U\cap\partial^{*}\!f(E)}\Phi(y,\nu_{f(E)})\operatorname{\,d\mathcal{H}^{n-1}}(y)=\int_{g(U)\cap\partial^{*}\!E}\Phi\Big{(}f(x),\frac{(\nabla g\circ f)^{t}\nu_{E}}{|(\nabla g\circ f)^{t}\nu_{E}|}\Big{)}Jf\>|(\nabla g\circ f)^{t}\nu_{E}|\operatorname{\,d\mathcal{H}^{n-1}}(x). (A.2)

Note that Jf|(gf)tνE|Jf\>|(\nabla g\circ f)^{t}\nu_{E}| is the tangential Jacobian of ff with respect to E\partial^{*}\!E.

Proof.

The fact that f(E)f(E) is a set of locally finite perimeter is shown in [Mag12, Proposition 17.1] and (A.1) is proved in [KMS19, Theorem A.1]. Hence we only need to show (A.2).

Let uε=1Eρεu_{\varepsilon}=1_{E}*\rho_{\varepsilon} where ρε\rho_{\varepsilon} denotes the standard mollifier and let vε=uεgv_{\varepsilon}=u_{\varepsilon}\circ g. Then uε1Eu_{\varepsilon}\to 1_{E} in L1loc(n)L^{1}_{loc}(\mathbb{R}^{n}) and vε1Eg=1f(E)v_{\varepsilon}\to 1_{E}\circ g=1_{f(E)} in L1loc(n)L^{1}_{loc}(\mathbb{R}^{n}) as shown in the proof of [Mag12, Proposition 17.1]. Note that vε=(g)t(uεg)\nabla v_{\varepsilon}=(\nabla g)^{t}(\nabla u_{\varepsilon}\circ g) and so vεf=(gf)tuε\nabla v_{\varepsilon}\circ f=(\nabla g\circ f)^{t}\nabla u_{\varepsilon}. By [Mag12, Remark 8.3], the change of variable y=f(x)y=f(x) gives

UΦ(y,vε|vε|)|vε|dy\displaystyle\int_{U}\Phi\Big{(}y,-\frac{\nabla v_{\varepsilon}}{|\nabla v_{\varepsilon}|}\Big{)}|\nabla v_{\varepsilon}|dy =g(U)Φ(f(x),vεf|vεf|)Jf|vεf|dx\displaystyle=\int_{g(U)}\Phi\Big{(}f(x),-\frac{\nabla v_{\varepsilon}\circ f}{|\nabla v_{\varepsilon}\circ f|}\Big{)}Jf|\nabla v_{\varepsilon}\circ f|dx
=g(U)Φ(f(x),(gf)tuε|(gf)tuε|)Jf|(gf)tuε|dx.\displaystyle=\int_{g(U)}\Phi\Big{(}f(x),-\frac{(\nabla g\circ f)^{t}\nabla u_{\varepsilon}}{|(\nabla g\circ f)^{t}\nabla u_{\varepsilon}|}\Big{)}Jf|(\nabla g\circ f)^{t}\nabla u_{\varepsilon}|dx. (A.3)

We shall show that this equation converges to (A.2) as ε0+\varepsilon\to 0^{+}. To do this, we shall apply the version of Reshetynak’s continuity theorem provided in [Spe11, Theorem 1.3]. Under the hypotheses that Φ\Phi is bounded and continuous and UU is open, this states that

limhUΦ(x,D|μh|μh)d|μh|=UΦ(x,D|μ|μ)d|μ|\displaystyle\lim_{h\to\infty}\int_{U}\Phi(x,D_{|\mu_{h}|}\mu_{h})d|\mu_{h}|=\int_{U}\Phi(x,D_{|\mu|}\mu)d|\mu| (A.4)

whenever μh\mu_{h}, μ\mu are finite n\mathbb{R}^{n}-valued measures satisfying

limhTdμh=Tdμ,TC0(U;n),and|μh|(U)|μ|(U),\displaystyle\lim_{h\to\infty}\int T\cdot d\mu_{h}=\int T\cdot d\mu,\ \forall\>T\in C_{0}(U;\mathbb{R}^{n}),\qquad\text{and}\qquad|\mu_{h}|(U)\to|\mu|(U), (A.5)

where C0(U;n)C_{0}(U;\mathbb{R}^{n}) denotes the completion, with respect to the sup norm, of the compactly supported continuous functions from UU to n\mathbb{R}^{n}.

Starting with the right-hand side of (A), first note that (gf)tuεn(gf)tμE-(\nabla g\circ f)^{t}\nabla u_{\varepsilon}\mathscr{L}^{n}\overset{\ast}{\rightharpoonup}(\nabla g\circ f)^{t}\mu_{E} since (gf)t(\nabla g\circ f)^{t} is continuous. By [Mag12, Theorem 12.20], uεnμE-\nabla u_{\varepsilon}\mathscr{L}^{n}\overset{\ast}{\rightharpoonup}\mu_{E} and |uε|n|μE||\nabla u_{\varepsilon}|\mathscr{L}^{n}\overset{\ast}{\rightharpoonup}|\mu_{E}| where n\mathscr{L}^{n} denotes Lebesgue measure. Since UU is bounded and gg is continuous, g(U)g(U) is bounded. Since g\nabla g is bounded, n1((g(U))E)=n1(g(Uf(E)))Lip(g)n1n1(Uf(E))=0\operatorname{\mathcal{H}^{n-1}}(\partial(g(U))\cap\partial^{*}\!E)=\operatorname{\mathcal{H}^{n-1}}(g(\partial U\cap\partial^{*}\!f(E)))\leq\operatorname{Lip}(g)^{n-1}\operatorname{\mathcal{H}^{n-1}}(\partial U\cap\partial^{*}\!f(E))=0. Hence (|uε|n)((g(U))|μE|((g(U)))(|\nabla u_{\varepsilon}|\mathscr{L}^{n})(\partial(g(U))\to|\mu_{E}|(\partial(g(U))) and

limε0+T(uε)dx=TdμE\displaystyle\lim_{\varepsilon\to 0^{+}}\int T\cdot(-\nabla u_{\varepsilon})dx=\int T\cdot d\mu_{E} (A.6)

for all TC0(g(U);n)T\in C_{0}(g(U);\mathbb{R}^{n}), since TCc(n;n)T\in C_{c}(\mathbb{R}^{n};\mathbb{R}^{n}) as g(U)¯\overline{g(U)} is compact. Thus uεn  g(U)-\nabla u_{\varepsilon}\mathscr{L}^{n}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>g(U), μE  g(U)\mu_{E}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>g(U) are finite n\mathbb{R}^{n}-valued measures which satisfy (A.5) (where we take discrete sequences of εh0+\varepsilon_{h}\to 0^{+}). Hence for each φCc(n)\varphi\in\textbf{{C}}_{c}(\mathbb{R}^{n}), applying [Spe11, Theorem 1.3] to the bounded, continuous function (x,ξ)φ(x)|(gf(x))tξ|(x,\xi)\mapsto\varphi(x)|(\nabla g\circ f(x))^{t}\xi| gives

limε0+g(U)φ|(gf)t(uε)|uε|||uε|dx=g(U)Eφ|(gf)tνE|dn1,\displaystyle\lim_{\varepsilon\to 0^{+}}\int_{g(U)}\varphi\Big{|}\Big{(}\nabla g\circ f\Big{)}^{t}\frac{(-\nabla u_{\varepsilon})}{|\nabla u_{\varepsilon}|}\Big{|}|\nabla u_{\varepsilon}|dx=\int_{g(U)\cap\partial^{*}\!E}\varphi|(\nabla g\circ f)^{t}\nu_{E}|\operatorname{\,d\mathcal{H}^{n-1}}, (A.7)

that is, |(gf)tuε|n  g(U)|(gf)tνE|n1  (g(U)E)|(\nabla g\circ f)^{t}\nabla u_{\varepsilon}|\mathscr{L}^{n}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>g(U)\overset{\ast}{\rightharpoonup}|(\nabla g\circ f)^{t}\nu_{E}|\operatorname{\mathcal{H}^{n-1}}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>\>(g(U)\cap\partial^{*}\!E). In particular, it follows from the fact n1((g(U))E)=0\operatorname{\mathcal{H}^{n-1}}(\partial(g(U))\cap\partial^{*}\!E)=0 that (|(gf)tuε|n)(g(U))(|(gf)tνE|n1  E)(g(U))(|(\nabla g\circ f)^{t}\nabla u_{\varepsilon}|\mathscr{L}^{n})(g(U))\to(|(\nabla g\circ f)^{t}\nu_{E}|\operatorname{\mathcal{H}^{n-1}}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>\>\partial^{*}\!E)(g(U)). Hence (A.5) holds for (gf)tuεn  g(U)(\nabla g\circ f)^{t}\nabla u_{\varepsilon}\mathscr{L}^{n}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>g(U), (gf)tνEn1  (g(U)E)(\nabla g\circ f)^{t}\nu_{E}\operatorname{\mathcal{H}^{n-1}}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>\>(g(U)\cap\partial^{*}\!E) and so we can apply [Spe11, Theorem 1.3] to (x,ξ)Φ(f(x),ξ)Jf(x)(x,\xi)\mapsto\Phi(f(x),\xi)Jf(x) which is bounded continuous since Φ\Phi and f\nabla f are. We obtain

limε0+\displaystyle\lim_{\varepsilon\to 0^{+}} g(U)Φ(f(x),(gf)tuε|(gf)tuε|)Jf|(gf)tuε|dx\displaystyle\int_{g(U)}\Phi\Big{(}f(x),-\frac{(\nabla g\circ f)^{t}\nabla u_{\varepsilon}}{|(\nabla g\circ f)^{t}\nabla u_{\varepsilon}|}\Big{)}Jf|(\nabla g\circ f)^{t}\nabla u_{\varepsilon}|dx
=g(U)EΦ(f(x),(gf)tνE|(gf)tνE|)Jf|(gf)tνE|dn1(x).\displaystyle=\int_{g(U)\cap\partial^{*}\!E}\Phi\Big{(}f(x),\frac{(\nabla g\circ f)^{t}\nu_{E}}{|(\nabla g\circ f)^{t}\nu_{E}|}\Big{)}Jf\>|(\nabla g\circ f)^{t}\nu_{E}|\operatorname{\,d\mathcal{H}^{n-1}}(x). (A.8)

This is the convergence of the right-hand side of (A) to the right-hand side of (A.2).

Now moving on to the left-hand side of (A), note that for all φCc1(n)\varphi\in C_{c}^{1}(\mathbb{R}^{n}),

φ(vε)dy=(φ)vεdyf(E)φdy=φdμf(E)\displaystyle\int\varphi(-\nabla v_{\varepsilon})dy=\int(\nabla\varphi)v_{\varepsilon}dy\to\int_{f(E)}\nabla\varphi\>dy=\int\varphi\>d\mu_{f(E)} (A.9)

since vε1f(E)v_{\varepsilon}\to 1_{f(E)} in L1loc(n)L^{1}_{loc}(\mathbb{R}^{n}). So by density of Cc1(n)C_{c}^{1}(\mathbb{R}^{n}) in Cc(n)C_{c}(\mathbb{R}^{n}), we have vεnμf(E)-\nabla v_{\varepsilon}\mathscr{L}^{n}\overset{\ast}{\rightharpoonup}\mu_{f(E)}. The change of variable y=f(x)y=f(x) ([Mag12, Remark 8.3]) gives

φ|vε|dy\displaystyle\int\varphi|\nabla v_{\varepsilon}|\>dy =(φf)|vεf|Jfdx=(φf)|(gf)tuε|Jfdx\displaystyle=\int(\varphi\circ f)|\nabla v_{\varepsilon}\circ f|Jfdx=\int(\varphi\circ f)|(\nabla g\circ f)^{t}\nabla u_{\varepsilon}|Jfdx
(φf)|(gf)tνE|Jfdx=φd|μf(E)|\displaystyle\to\int(\varphi\circ f)|(\nabla g\circ f)^{t}\nu_{E}|Jfdx=\int\varphi\>d|\mu_{f(E)}| (A.10)

where the last equality is by [Mag12, Proposition 17.1]. Hence |vε|n|μf(E)||\nabla v_{\varepsilon}|\mathscr{L}^{n}\overset{\ast}{\rightharpoonup}|\mu_{f(E)}|. Thus (|vε|n)(U)|μf(E)|(U)(|\nabla v_{\varepsilon}|\mathscr{L}^{n})(U)\to|\mu_{f(E)}|(U) since n1(Uf(E))=0\operatorname{\mathcal{H}^{n-1}}(\partial U\cap\partial^{*}\!f(E))=0 and

limε0+T(vε)dy=Tdμf(E)\displaystyle\lim_{\varepsilon\to 0^{+}}\int T\cdot(-\nabla v_{\varepsilon})dy=\int T\cdot d\mu_{f(E)} (A.11)

for all TC0(U;n)T\in C_{0}(U;\mathbb{R}^{n}), since TCc(n;n)T\in C_{c}(\mathbb{R}^{n};\mathbb{R}^{n}) as U¯\overline{U} is compact. Hence vεn  U,μf(E)  U-\nabla v_{\varepsilon}\mathscr{L}^{n}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>U,\mu_{f(E)}\mathbin{\vrule height=4.30554pt,depth=0.0pt,width=0.43057pt\vrule height=0.43057pt,depth=0.0pt,width=4.30554pt}\>U satisfies (A.5) so by [Spe11, Theorem 1.3], we have

limε0+UΦ(y,vε|vε|)|vε|dy=Uf(E)Φ(y,νf(E))dn1(y).\displaystyle\lim_{\varepsilon\to 0^{+}}\int_{U}\Phi\Big{(}y,-\frac{\nabla v_{\varepsilon}}{|\nabla v_{\varepsilon}|}\Big{)}|\nabla v_{\varepsilon}|dy=\int_{U\cap\partial^{*}\!f(E)}\Phi(y,\nu_{f(E)})\operatorname{\,d\mathcal{H}^{n-1}}(y). (A.12)

Hence left hand side of (A) converges to the left hand side of (A.2) and we are done. ∎

References

  • [All72] W.K. Allard, On the first variation of a varifold, Ann. of Math. 95 (1972), no. 3, 417–491. MR MR0307015
  • [All74] by same author, A characterization of the area integrand, Symposia Mathematica, vol. 14, 1974, pp. 429–444.
  • [Alm66] F.J. Almgren, Some interior regularity theorems for minimal surfaces and an extension of Bernstein’s theorem, Ann. of Math. (2) 84 (1966), 277–292. MR 0200816
  • [Alm68] by same author, Existence and regularity almost everywhere of solutions to elliptic variational problems among surfaces of varying topological type and singularity structure, Annals of Mathematics (1968), 321–391.
  • [BDGG69] E. Bombieri, E. De Giorgi, and E. Giusti, Minimal cones and the Bernstein problem, Ivent. Math. 7 (1969), 243–268.
  • [BEG+19] Simon. Bortz, Max. Engelstein, Max. Goering, Tatiana. Toro, and Zihui. Zhao, Two Phase Free Boundary Problem for Poisson Kernels, 2019.
  • [Bom82] E. Bombieri, Regularity Theory for Almost Minimal Currents, Arch. Rational Mech. Anal. (1982), 99–130.
  • [DDG20] Guido De Philippis, Antonio De Rosa, and Francesco Ghiraldin, Existence Results for Minimizers of Parametric Elliptic Functionals, Journal of Geometric Analysis 30 (2020), no. 2, 1450–1465 (English (US)).
  • [DESVGT19] Guy David, Max Engelstein, Mariana Smit Vega Garcia, and Tatiana Toro, Regularity for almost-minimizers of variable coefficient Bernoulli-type functionals, 2019.
  • [DG60] E. De Giorgi, Frontiere orientate di misura minima, Sem. Mat. Scuola Norm. Sup. Pisa (1960), 1–56.
  • [DG65] by same author, Una estensione del teorema di Bernstein, Annali della Scuola Normale Superiore di Pisa - Classe di Scienze Ser. 3, 19 (1965), no. 1, 79–85. MR 178385
  • [DLDRG19] Camillo De Lellis, Antonio De Rosa, and Francesco Ghiraldin, A direct approach to the anisotropic Plateau problem, Advances in Calculus of Variations 12 (2019), no. 2, 211 – 223.
  • [DLGM17] C. De Lellis, F. Ghiraldin, and F. Maggi, A direct approach to Plateau’s problem, Journal of the European Mathematical Society 19 (2017), no. 8, 2219–2240.
  • [Dou31] Jesse Douglas, Solution of the problem of Plateau, Trans. Amer. Math. Soc. 33 (1931), 263–321.
  • [DPDRG16] G. De Philippis, A. De Rosa, and F. Ghiraldin, A direct approach to Plateau’s problem in any codimension, Advances in Mathematics 288 (2016), 59–80.
  • [DPDRG18] by same author, Rectifiability of varifolds with locally bounded first variation with respect to anisotropic surface energies, Communications on Pure and Applied Mathematics 71 (2018), no. 6, 1123–1148.
  • [dQT18] Olivaine S. de Queiroz and Leandro S. Tavares, Almost minimizers for semilinear free boundary problems with variable coefficients, Mathematische Nachrichten 291 (2018), no. 10, 1486–1501.
  • [DR18] Antonio De Rosa, Minimization of anisotropic energies in classes of rectifiable varifolds, SIAM Journal on Mathematical Analysis 50 (2018), no. 1, 162–181.
  • [DRK20] A. De Rosa and S. Kolasiński, Equivalence of the Ellipticity Conditions for Geometric Variational Problems, Communications on Pure and Applied Mathematics n/a (2020), no. n/a.
  • [Fed69] H. Federer, Geometric measure theory, vol. 1996, Springer New York, 1969.
  • [Fed70] by same author, The singular sets of area minimizing rectifiable currents with codimension one and of area minimizing flat chains modulo two with arbitrary codimension, Bull. Amer. Math. Soc. 76 (1970), 767–771.
  • [Fle62] Wendell H. Fleming, On the oriented plateau problem, Rendiconti del Circolo Matematico di Palermo 11 (1962), no. 1, 69–90.
  • [GMT83] E. Gonzalez, U. Massari, and I. Tamanini, On the Regularity of Boundaries of Sets Minimizing Perimeter with a Volume Constraint, Indiana University Mathematics Journal 32 (1983), no. 1, 25–37.
  • [HP16] Jenny Harrison and Harrison Pugh, Existence and soap film regularity of solutions to Plateau’s problem, Advances in Calculus of Variations 9 (2016), no. 4, 357 – 394.
  • [JPSVG20] Seongmin Jeon, Arshak Petrosyan, and Mariana Smit Vega Garcia, Almost minimizers for the thin obstacle problem with variable coefficients, 2020.
  • [KMS19] Darren King, Francesco Maggi, and Salvatore Stuvard, Plateau’s problem as a singular limit of capillarity problems, 2019.
  • [LS20] Jimmy Lamboley and Pieralberto Sicbaldi, Existence and regularity of Faber-Krahn minimizers in a Riemannian manifold, Journal de Mathématiques Pures et Appliqués (2020).
  • [Mag12] Francesco Maggi, Sets of Finite Perimeter and Geometric Variational Problems: An Introduction to Geometric Measure Theory, Cambridge University Press, 2012.
  • [Mat95] Pertti Mattila, Geometry of sets and measures in Euclidean spaces: fractals and rectifiability, Cambridge studies in advanced mathematics;, vol. 44, Cambridge University Press, United Kingdom, 1995 (English).
  • [Mir65] M. Miranda, Sul minimo dell’integrale del gradiente di una fonzione, Ann. Scuola Norm. Sup. Pisa 19 (1965), no. 3, 626–665.
  • [Par84] Harold R. Parks, Regularity of solutions to elliptic isoperimetric problems., Pacific J. Math. 113 (1984), no. 2, 463–470.
  • [Rad30] Tibor Radó, On Plateau’s Problem, Annals of Mathematics, Second Series 31 (1930), no. 3, 457–469.
  • [Rei60] E.R. Reifenberg, Solution of the Plateau Problem for m-dimensional surfaces of varying topological type, Acta Math. 104 (1960), no. 1-2, 1–92.
  • [Rei64a] by same author, An Epiperimetric Inequality Related to the Analyticity of Minimal Surfaces, Annals of Mathematics, Second Series 80 (1964), no. 1, 1–14.
  • [Rei64b] by same author, On the Analyticity of Minimal Surfaces, Annals of Mathematics, Second Series 80 (1964), no. 1, 15–21.
  • [Sim68] James Simons, Minimal varieties in Riemannian manifolds, Ann. of Math. (2) 88 (1968), 62–105. MR 0233295
  • [Spe11] Daniel Spector, Simple proofs of some results of Reshetnyak, Proc. Amer. Math. Soc. 139 (2011), 1681–1690.
  • [SS82] R. Schoen and L. Simon, A New Proof of the Regularity Theorem for Rectifiable Currents which Minimize Parametric Elliptic Functionals, Indiana University Mathematics Journal 31 (1982), no. 3, 415–434.
  • [SSA77] R. Schoen, L. Simon, and F.J. Almgren, Regularity and singularity estimates on hypersurfaces minimizing parametric elliptic variational integrals, Acta Mathematica 139 (1977), no. 1, 217–265.
  • [STV18] Luca Spolaor, Baptiste Trey, and Bozhidar Velichkov, Free boundary regularity for a multiphase shape optimization problem, 2018.
  • [Tam82] Italo Tamanini, Boundaries of Caccioppoli sets with Hölder-continuois normal vector., Journal für die reine und angewandte Mathematik 334 (1982), 27–39.
  • [Tam84] by same author, Regularity Results for Almost Minimal Oriented Hypersurfaces in RnR^{n}, Quaderni del Dipartimento Matematica dell’Universita de Lecce, 1984.
  • [Tay76a] Jean E. Taylor, The Structure of Singularities in Soap-Bubble-Like and Soap-Film-Like Minimal Surfaces, Annals of Mathematics, Second Series 103 (1976), no. 3, 489–539.
  • [Tay76b] by same author, The Structure of Singularities in Solutions to Ellipsoidal Variational Problems With Constraints in 3\mathbb{R}^{3}, Annals of Mathematics, Second Series 103 (1976), no. 3, 541–546.
  • [Wul01] G. Wulff, Zur Frage der Geschwindigkeit des Wachsturms und der auflösungder Kristall-Flächen, Zeitschrift für Kristallographie. 34 (1901), 449–530.