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

Strong BVBV-extension and W1,1W^{1,1}-extension domains

Miguel García-Bravo  and  Tapio Rajala University of Jyvaskyla
Department of Mathematics and Statistics
P.O. Box 35 (MaD)
FI-40014 University of Jyvaskyla
Finland
[email protected] [email protected]
Abstract.

We show that a bounded domain in a Euclidean space is a W1,1W^{1,1}-extension domain if and only if it is a strong BVBV-extension domain. In the planar case, bounded and strong BVBV-extension domains are shown to be exactly those BVBV-extension domains for which the set ΩiΩ¯i\partial\Omega\setminus\bigcup_{i}\overline{\Omega}_{i} is purely 11-unrectifiable, where Ωi\Omega_{i} are the open connected components of 2Ω¯\mathbb{R}^{2}\setminus\overline{\Omega}.

Key words and phrases:
Sobolev extension, BV-extension
2000 Mathematics Subject Classification:
Primary 46E35.
The authors acknowledge the support from the Academy of Finland, grant no. 314789.

1. Introduction

Let Ωn\Omega\subset\mathbb{R}^{n} be a domain for some n2n\geq 2. For every 1p1\leq p\leq\infty, we define the Sobolev space W1,p(Ω)W^{1,p}(\Omega) to be

W1,p(Ω)={uLp(Ω):uLp(Ω;n)},W^{1,p}(\Omega)=\{u\in L^{p}(\Omega):\,\nabla u\in L^{p}(\Omega;\mathbb{R}^{n})\},

where u\nabla u denotes the distributional gradient of uu. We equip this space with the non-homogeneous norm

uW1,p(Ω)=uLp(Ω)+uLp(Ω).\|u\|_{W^{1,p}(\Omega)}=\|u\|_{L^{p}(\Omega)}+\|\nabla u\|_{L^{p}(\Omega)}.

We say that Ω\Omega is a W1,pW^{1,p}-extension domain if there exists an operator T:W1,p(Ω)W1,p(n)T\colon W^{1,p}(\Omega)\to W^{1,p}(\mathbb{R}^{n}) and a constant C>0C>0 so that

TuW1,p(Rn)CuW1,p(Ω)\|Tu\|_{W^{1,p}(R^{n})}\leq C\|u\|_{W^{1,p}(\Omega)}

and Tu|Ω=uTu|_{\Omega}=u for every uW1,p(Ω)u\in W^{1,p}(\Omega). We denote the minimal constant CC above by T\|T\|. We point out that by the results from [10, 21], for p>1p>1 one can always assume the operator TT to be linear, and also for the case of bounded simply connected planar domains if p=1p=1 by [16]. It is not yet known if this is the case for general domains when p=1p=1.

It is well-known from the works of Calderón and Stein [5, 23] that Lipschitz domains are W1,pW^{1,p}-extension domains for every p1p\geq 1. Moreover, Jones showed in [12] that every uniform domain Ωn\Omega\subset\mathbb{R}^{n} is a W1,pW^{1,p}-extension domain for all p1p\geq 1. However, these conditions are not necessary for a domain to be a Sobolev extension domain. For bounded simply connected planar domains a geometric characterization of Sobolev extension domains by means of a curve condition has been given in the works [22, 15, 16]. Namely, for the W1,1W^{1,1} case we have the following: A bounded planar simply connected domain Ω\Omega is a W1,1W^{1,1}-extension domain if and only if for every x,yΩcx,y\in\Omega^{c} there exists a curve γΩc\gamma\subset\Omega^{c} connecting xx and yy with

(γ)C|xy|, and 1(γΩ)=0.\ell(\gamma)\leq C|x-y|,\text{ and }\mathcal{H}^{1}(\gamma\cap\partial\Omega)=0. (1.1)

A typical example of a simply connected planar domain Ω\Omega which is not a W1,pW^{1,p}-extension domain for any p1p\geq 1 is the slit disk D={(x,y)2:x2+y2<1}([0,1)×{0})}D=\{(x,y)\in\mathbb{R}^{2}:\,x^{2}+y^{2}<1\}\setminus([0,1)\times\{0\})\}. However, by the results of [14], knowing that the complement is quasiconvex is enough to ensure that DD is a BVBV-extension domain.

Recall that

BV(Ω)={uL1(Ω):Du(Ω)<}BV(\Omega)=\{u\in L^{1}(\Omega):\,\|Du\|(\Omega)<\infty\}

is the space of functions of bounded variation where

Du(Ω)=sup{Ωudiv(v)𝑑x:vC0(Ω;n),|v|1}\|Du\|(\Omega)=\sup\left\{\int_{\Omega}u\,\text{div}(v)\,dx:\,v\in C^{\infty}_{0}(\Omega;\mathbb{R}^{n}),\,|v|\leq 1\right\}

denotes the total variation of uu on Ω\Omega. We endow this space with the norm uBV(Ω)=uL1(Ω)+Du(Ω).\|u\|_{BV(\Omega)}=\|u\|_{L^{1}(\Omega)}+\|Du\|(\Omega). Note that Du\|Du\| is a Radon measure on Ω\Omega that is defined for every set FΩF\subset\Omega as

Du(F)=inf{Du(U):FUΩ,Uopen}.\|Du\|(F)=\inf\{\|Du\|(U):\,F\subset U\subset\Omega,\;U\;\text{open}\}.

We say that Ω\Omega is a BVBV-extension domain if there exists a constant C>0C>0 and a (not necessarily linear) extension operator T:BV(Ω)BV(n)T\colon BV(\Omega)\to BV(\mathbb{R}^{n}) so that Tu|Ω=uTu|_{\Omega}=u and

TuBV(n)CuBV(Ω)\|Tu\|_{BV(\mathbb{R}^{n})}\leq C\|u\|_{BV(\Omega)}

for all uBV(Ω)u\in BV(\Omega) and where C>0C>0 is an absolute constant, independent of uu. Let us point out that Ω\Omega being a W1,1W^{1,1}-extension domain always implies that it is also a BVBV-extension domain (see [14, Lemma 2.4]).

Our first main result is the characterization of bounded W1,1W^{1,1}-extension domains in terms of strong extendability of BVBV-functions, or equivalently, in terms of strong extendability of sets of finite perimeter. The equivalence between strong extendability of BVBV-functions and strong extendability of sets of finite perimeter is inspired by the work of Mazy’a and Burago [4] (see also [20, Section 9.3]). They showed that for all uLloc1(Ω)u\in L^{1}_{\text{loc}}(\Omega) with finite total variation we may find an extension TuLloc1(n)Tu\in L^{1}_{\text{loc}}(\mathbb{R}^{n}) with D(Tu)(n)CDu(Ω)\|D(Tu)\|(\mathbb{R}^{n})\leq C\|Du\|(\Omega), for some constant C>0C>0, if and only if any set EΩE\subset\Omega of finite perimeter in Ω\Omega admits an extension E~n\widetilde{E}\subset\mathbb{R}^{n} satisfying E~Ω=E\widetilde{E}\cap\Omega=E and P(E~,n)CP(E,Ω)P(\widetilde{E},\mathbb{R}^{n})\leq CP(E,\Omega) where C>0C>0 is some constant. Recall that a Lebesgue measurable subset EnE\subset\mathbb{R}^{n} has finite perimeter in Ω\Omega if χEBV(Ω)\chi_{E}\in BV(\Omega), where χE\chi_{E} denotes the characteristic function of the set EE. We set P(E,Ω)=DχE(Ω)P(E,\Omega)=\|D\chi_{E}\|(\Omega) and call it the perimeter of EE in Ω\Omega. If a set EE does not have finite perimeter in Ω\Omega we set P(E,Ω)=P(E,\Omega)=\infty.

Before stating our characterization, we introduce the terminology of strong extendability, following [11] and [17].

Definition 1.1 (Strong BVBV-extension domain).

A domain Ωn\Omega\subset\mathbb{R}^{n} is called a strong BVBV-extension domain if there exists a constant C>0C>0 so that for any uBV(Ω)u\in BV(\Omega) there exists TuBV(n)Tu\in BV(\mathbb{R}^{n}) with Tu|Ω=uTu|_{\Omega}=u, TuBV(n)CuBV(Ω)\|Tu\|_{BV(\mathbb{R}^{n})}\leq C\|u\|_{BV(\Omega)}, and D(Tu)(Ω)=0\|D(Tu)\|(\partial\Omega)=0.

In the spirit of Definition 1.1, we define the analogous concept for sets of finite perimeter.

Definition 1.2 (Strong extension property for sets of finite perimeter).

A domain Ωn\Omega\subset\mathbb{R}^{n} is said to have the strong extension property for sets of finite perimeter if there exists a constant C>0C>0 so that for any set EΩE\subset\Omega of finite perimeter in Ω\Omega there exists a set E~n\widetilde{E}\subset\mathbb{R}^{n} such that

  1. (PE1)

    E~Ω=E\widetilde{E}\cap\Omega=E modulo measure zero sets,

  2. (PE2)

    P(E~,n)CP(E,Ω)P(\widetilde{E},\mathbb{R}^{n})\leq CP(E,\Omega), and

  3. (PE3)

    n1(ME~Ω)=0\mathcal{H}^{n-1}(\partial^{M}\widetilde{E}\cap\partial\Omega)=0.

With the above definitions we can state our first main result.

Theorem 1.3.

Let Ωn\Omega\subset\mathbb{R}^{n} be a bounded domain. Then the following are equivalent:

  1. (1)

    Ω\Omega is a W1,1W^{1,1}-extension domain.

  2. (2)

    Ω\Omega is a strong BVBV-extension domain.

  3. (3)

    Ω\Omega has the strong extension property for sets of finite perimeter.

Our main motivation behind this theorem is to understand better the geometry of W1,1W^{1,1}-extension domains. From Theorem 1.3 we see that for a bounded W1,1W^{1,1}-extension domain, except for a purely (n1)(n-1)-unrectifiable set, the boundary consists of points where the domain has density at most 1/21/2. See Section 4 for the proof of this. In the same section we give an example showing that the above density bound is not sufficient to imply that a bounded BVBV-extension domain is a W1,1W^{1,1}-extension domain, even in the plane. Another corollary of Theorem 1.3 is that for a bounded W1,1W^{1,1}-extension domain, again up to a purely (n1)(n-1)-unrectifiable set, the boundary consists of points that are boundary points also for some component of the interior of the complement of the domain. In Section 4 we provide also an example showing that in 3\mathbb{R}^{3} this property does not characterize W1,1W^{1,1}-extension domains among bounded BVBV-extension domains. However, our second main result states that in the planar case this is true.

Theorem 1.4.

Let Ω2\Omega\subset\mathbb{R}^{2} be a bounded BVBV-extension domain. Then Ω\Omega is a W1,1W^{1,1}-extension domain if and only if the set

ΩiIΩi¯\partial\Omega\setminus\bigcup_{i\in I}\overline{\Omega_{i}}

is purely 11-unrectifiable, where {Ωi}iI\{\Omega_{i}\}_{i\in I} are the connected components of 2Ω¯\mathbb{R}^{2}\setminus\overline{\Omega}.

Let us mention that Theorem 1.4 recovers partly the theorems in [16]. Namely, it immediately follows that Jordan BVBV-extension domains are W1,1W^{1,1}-extension domains since the set required in Theorem 1.4 to be purely unrectifiable, is indeed empty. The curve characterization (1.1) also follows quite easily from Theorem 1.4 using a small observation recorded in [16]. Let us briefly sketch this. Since a W1,1W^{1,1}-extension domain is known to be a BVBV-extension domain, its complement is quasiconvex. Then, a quasiconvex curve between two points in the complement can be modified to intersect the boundaries of each Ωi\Omega_{i} at most twice (see Lemma 5.3). Theorem 1.4 now says that the rest of the curve intersects Ω\partial\Omega in a 1\mathcal{H}^{1}-measure zero set, giving condition (1.1). Conversely, (1.1) implies quasiconvexity, and hence that Ω\Omega is a BVBV-extension domain. For a simply connected Ω\Omega, we can connect every pair of components Ωi\Omega_{i} and Ωj\Omega_{j} with a curve satisfying (1.1). Since the set

ΩiIΩi¯.\partial\Omega\setminus\bigcup_{i\in I}\overline{\Omega_{i}}.

is contained in countably many of such curves by [16, Lemma 4.6], we see that it is purely 11-unrectifiable.

Let us point out, however, that the extension operator that we construct in Theorem 1.4, is not always linear. One of the main points of [16] was to construct a linear extension operator. At the moment we do not see how our construction could be modified to give a linear extension operator. Still, the general smoothing operator we use for proving Theorem 1.3 (and Theorem 1.4) immediately gives the following.

Corollary 1.5.

Suppose Ωn\Omega\subset\mathbb{R}^{n} is a bounded strong BVBV-extension domains where the extension operator is linear. Then there exists a linear W1,1W^{1,1}-extension operator from W1,1(Ω)W^{1,1}(\Omega) to W1,1(n)W^{1,1}(\mathbb{R}^{n}).

Although not strictly used in our proofs, we include the following result for future use: Every BVBV-extension domain Ωn\Omega\subset\mathbb{R}^{n} satisfies the measure density condition, that is, there exists a constant c>0c>0 so that for every xΩ¯x\in\overline{\Omega} and r(0,1]r\in(0,1] we have |B(x,r)Ω|crn.|B(x,r)\cap\Omega|\geq cr^{n}. One may find this result in Section 2.2. The same conclusion for W1,pW^{1,p} -extension domains with 1p<1\leq p<\infty is also true and was already shown in [10].

2. Preliminaries

When making estimates, we often write the constants as positive real numbers CC which may vary between appearances, even within a chain of inequalities. These constants normally only depend on the dimension of the underlying space n\mathbb{R}^{n} unless otherwise stated.

For any point xnx\in\mathbb{R}^{n} and radius r>0r>0 we denote the open ball by

B(x,r)={yn:|xy|<r}.B(x,r)=\{y\in\mathbb{R}^{n}\,:\,|x-y|<r\}.

More generally, for a set AnA\subset\mathbb{R}^{n} we define the open rr-neighbourhood as

B(A,r)=xAB(x,r).B(A,r)=\bigcup_{x\in A}B(x,r).

We denote by |E||E| the nn-dimensional outer Lebesgue measure of a set EnE\subset\mathbb{R}^{n}. For any Lebesgue measurable subset EnE\subset\mathbb{R}^{n} and any point xnx\in\mathbb{R}^{n} we then define the upper density of EE at xx as

D¯(E,x)=lim supr0|EB(x,r)||B(x,r)|,\overline{D}(E,x)=\limsup_{r\searrow 0}\frac{|E\cap B(x,r)|}{|B(x,r)|},

and the lower density of EE at xx as

D¯(E,x)=lim infr0|EB(x,r)||B(x,r)|.\underline{D}(E,x)=\liminf_{r\searrow 0}\frac{|E\cap B(x,r)|}{|B(x,r)|}.

If D¯(E,x)=D¯(E,x)\overline{D}(E,x)=\underline{D}(E,x), we call the common value the density of EE at xx and denote it by D(E,x)D(E,x). The essential interior of EE is then defined as

E̊M={xn:D(E,x)=1},\mathring{E}^{M}=\{x\in\mathbb{R}^{n}\,:\,D(E,x)=1\},

the essential closure of EE as

E¯M={xn:D¯(E,x)>0},\overline{E}^{M}=\{x\in\mathbb{R}^{n}\,:\,\overline{D}(E,x)>0\},

and the essential boundary of EE as

ME={xn:D¯(E,x)>0 and D¯(nE,x)>0}.\partial^{M}E=\{x\in\mathbb{R}^{n}\,:\,\overline{D}(E,x)>0\text{ and }\overline{D}(\mathbb{R}^{n}\setminus E,x)>0\}.

As usual, s(A)\mathcal{H}^{s}(A) will stand for the ss-dimensional Hausdorff measure of a set AnA\subset\mathbb{R}^{n} obtained as the limit

s(A)=limδ0δs(A),\mathcal{H}^{s}(A)=\lim_{\delta\searrow 0}\mathcal{H}_{\delta}^{s}(A),

where δs(A)\mathcal{H}_{\delta}^{s}(A) is the ss-dimensional Hausdorff δ\delta-content of AA defined as

δs(A)=inf{i=1missingdiam(Ui)s:Ai=1Ui,missingdiam(Ui)δ}.\mathcal{H}_{\delta}^{s}(A)=\inf\left\{\sum_{i=1}^{\infty}{\mathop{\mathrm{missing}}{\,diam\,}}(U_{i})^{s}\,:\,A\subset\bigcup_{i=1}^{\infty}U_{i},{\mathop{\mathrm{missing}}{\,diam\,}}(U_{i})\leq\delta\right\}.

We say that a set HnH\subset\mathbb{R}^{n} is mm-rectifiable, for some m<nm<n, if there exist countably many Lipschitz maps fj:mnf_{j}\colon\mathbb{R}^{m}\to\mathbb{R}^{n} so that m(Hjfj(m))=0\mathcal{H}^{m}(H\setminus\bigcup_{j}f_{j}(\mathbb{R}^{m}))=0. A set HH will be called purely mm-unrectifiable if for every Lipschitz map f:mnf\colon\mathbb{R}^{m}\to\mathbb{R}^{n} we have

m(Hf(m))=0.\mathcal{H}^{m}(H\cap f(\mathbb{R}^{m}))=0.

Observe that by Rademacher’s theorem one can deduce that if f:mnf\colon\mathbb{R}^{m}\to\mathbb{R}^{n} is Lipschitz, then there are countably many sets EimE_{i}\subset\mathbb{R}^{m} on which ff is bi-Lipschitz and such that m(f(miEi))=0\mathcal{H}^{m}(f(\mathbb{R}^{m}\setminus\bigcup_{i}E_{i}))=0.

Moreover, it easily follows that if HnH\subset\mathbb{R}^{n} is not mm-purely unrectifiable, then there exists a Lipschitz map f:mnmf\colon\mathbb{R}^{m}\to\mathbb{R}^{n-m} so that up to a rotation, the set

HGraph(f)H\cap\textrm{Graph}(f)

has positive m\mathcal{H}^{m}-measure, where Graph(f)={(x,f(x)):xm}\textrm{Graph}(f)=\{(x,f(x))\,:\,x\in\mathbb{R}^{m}\}.

By a dyadic cube we refer to Q=[0,2k]n+𝚓nQ=[0,2^{-k}]^{n}+\mathtt{j}\subset\mathbb{R}^{n} for some kk\in\mathbb{Z} and 𝚓2kn\mathtt{j}\in 2^{-k}{\mathbb{Z}}^{n}. We denote the side-length of such dyadic cube QQ by (Q):=2k\ell(Q):=2^{-k}.

2.1. BVBV-functions and sets of finite perimeter

Let us recall some basic results related to BVBV-functions and sets of finite perimeter. For a more detailed account, we refer to the books [2, 6, 8].

Differently to this paper, Mazy’a and Burago’ (see [4] and also [20, Section 9.3]) considered the space

BVl(Ω)={uLloc1(Ω):Du(Ω)<}BV_{l}(\Omega)=\{u\in L^{1}_{loc}(\Omega):\,\|Du\|(\Omega)<\infty\}

equipped with the seminorm Du(Ω)\|Du\|(\Omega). This way they defined BVlBV_{l}-extension domains to be those Ωn\Omega\subset\mathbb{R}^{n} for which just the total variation of the extension is controlled, that is, whenever D(Tu)(n)CDu(Ω)\|D(Tu)\|(\mathbb{R}^{n})\leq C\|Du\|(\Omega). As we already explained in the introduction, they proved that being a BVlBV_{l}-extension domain was equivalent to the fact that any set EΩE\subset\Omega of finite perimeter in Ω\Omega admits an extension E~n\widetilde{E}\subset\mathbb{R}^{n} satisfying only (PE1) and (PE2) from Definition 1.2. Note however, that thanks to [14, Lemma 2.1] BVlBV_{l}-extension domains are equivalent to BVBV extension domains if Ω\Omega is bounded.

When working with BVBV functions we will make use of the well-known (1,1)(1,1)-Poincaré inequality that we now state (see for instance [2, Theorem 3.44] for the proof).

Theorem 2.1.

Let Ωn\Omega\subset\mathbb{R}^{n} be an open bounded set with Lipschitz boundary. Then there exists a constant C>0C>0 depending only on nn and Ω\Omega so that for every uBV(Ω)u\in BV(\Omega) we have

Ω|u(y)uΩ|𝑑yCDu(Ω).\int_{\Omega}|u(y)-u_{\Omega}|\,dy\leq C\|Du\|(\Omega).

In particular, there exists a constant C>0C>0 only depending on nn so that if Q,QnQ,Q^{\prime}\subset\mathbb{R}^{n} are two dyadic cubes with 14(Q)(Q)4(Q)\frac{1}{4}\ell(Q^{\prime})\leq\ell(Q)\leq 4\ell(Q^{\prime}) and Ω=int(QQ)\Omega=\text{int}(Q\cup Q^{\prime}) connected, then for every uBV(Ω)u\in BV(\Omega),

Ω|u(y)uΩ|𝑑yC(Q)Du(Ω).\int_{\Omega}|u(y)-u_{\Omega}|\,dy\leq C\ell(Q)\|Du\|(\Omega). (2.1)

We are using here the notation of the mean value integral of a function uu on the set Ω\Omega as

uΩ=1|Ω|Ωu(y)𝑑y.u_{\Omega}=\frac{1}{|\Omega|}\int_{\Omega}u(y)\,dy.

Let us record as well the coarea formula for BVBV functions. See for example [6, Section 5.5].

Theorem 2.2.

Given a function uBV(Ω)u\in BV(\Omega), the superlevel sets ut={xΩ:u(x)>t}u_{t}=\{x\in\Omega:\,u(x)>t\} have finite perimeter in Ω\Omega for almost every tt\in\mathbb{R} and

Du(F)=P(ut,F)𝑑t\|Du\|(F)=\int^{\infty}_{-\infty}P(u_{t},F)\,dt

for every Borel set FΩF\subset\Omega. Conversely, if uL1(Ω)u\in L^{1}(\Omega) and P(ut,Ω)𝑑t<\int^{\infty}_{-\infty}P(u_{t},\Omega)\,dt<\infty then uBV(Ω)u\in BV(\Omega).

An important result due to Federer [8, Section 4..5.11] tells us that a set EE has finite perimeter in Ω\Omega if and only if n1(MEΩ)<\mathcal{H}^{n-1}(\partial^{M}E\cap\Omega)<\infty. Moreover, thanks to De Giorgi’s pioneering work [7] we can understand the structure of the boundary of sets of finite perimeter even better. Namely, if EE has finite perimeter in Ω\Omega then for every subset AΩA\subset\Omega,

DχE(A)=P(E,A)=n1(MEA)\|D\chi_{E}\|(A)=P(E,A)=\mathcal{H}^{n-1}(\partial^{M}E\cap A)

and if EE has finite perimeter in n\mathbb{R}^{n} then

ME=FnKn\partial^{M}E=F\cup\bigcup_{n\in\mathbb{N}}K_{n}

where n1(F)=0\mathcal{H}^{n-1}(F)=0 and KnK_{n} are compact subsets of C1C^{1} hypersurfaces. Furthermore, for any set EE with finite perimeter we have

D¯(E,x)=D¯(E,x){0,1/2,1}\overline{D}(E,x)=\underline{D}(E,x)\in\{0,1/2,1\}

for n1\mathcal{H}^{n-1}-almost every xnx\in\mathbb{R}^{n}. Moreover, n1(ME{x:D(E,x)=1/2})=0\mathcal{H}^{n-1}(\partial^{M}E\setminus\{x:\,D(E,x)=1/2\})=0, and hence for n1\mathcal{H}^{n-1}-almost every xMEx\in\partial^{M}E we have

D(E,x)=12.D(E,x)=\frac{1}{2}.

Let us finally recall some terminology and results from [1]. A Lebesgue measurable set EnE\subset\mathbb{R}^{n} with |E|>0|E|>0 is called decomposable if there exist two Lebesgue measurable sets F,GnF,G\subset\mathbb{R}^{n} so that |F|,|G|>0|F|,|G|>0, E=FGE=F\cup G, FG=F\cap G=\emptyset, and

P(E,n)=P(F,n)+P(G,n).P(E,\mathbb{R}^{n})=P(F,\mathbb{R}^{n})+P(G,\mathbb{R}^{n}).

A set is called indecomposable if it is not decomposable. For example, any connected open set EnE\subset\mathbb{R}^{n} with n1(ME)<\mathcal{H}^{n-1}(\partial^{M}E)<\infty is indecomposable.

For any set EnE\subset\mathbb{R}^{n} of finite perimeter we can always find a unique countable family of disjoint indecomposable subsets EiEE_{i}\subset E so that |Ei|>0|E_{i}|>0, P(E,n)=iP(Ei,n)P(E,\mathbb{R}^{n})=\sum_{i}P(E_{i},\mathbb{R}^{n}) and, moreover,

n1(E̊MiE̊iM)=0.\mathcal{H}^{n-1}\left(\mathring{E}^{M}\setminus\bigcup_{i}\mathring{E}_{i}^{M}\right)=0.

For a proof of this result we refer to [1, Theorem 1].

In the particular case of 2\mathbb{R}^{2}, thanks to [1, Corollary 1] one can find a decomposition of sets of finite perimeter into indecomposable sets whose boundaries are rectifiable Jordan curves, except for a set of 1\mathcal{H}^{1}-measure zero. We will state this useful result in the last Section 5, in Theorem 5.1.

2.2. Measure density condition for BVBV-extension domains

Nowadays it is a well-known fact that all W1,pW^{1,p}-extension domains for 1p<1\leq p<\infty satisfy the measure density condition (see [10]). Although we do not need this in our proofs, we record here the fact that the same property holds for BVBV-extension domains. Let us remark that a measure density condition for planar BVlBV_{l}-extension domains was proven in [14, Lemma 2.10]. However, the proof does not seem to extend to domains in n\mathbb{R}^{n}. The method of proof we employ here follows the same lines as [10] and can be adapted for BVlBV_{l}-extension domains as well.

Proposition 2.3.

Let Ωn\Omega\subset\mathbb{R}^{n} be a BVBV-extension or a W1,1W^{1,1}-extension domain, then there exists a constant c>0c>0, depending only on nn and on the operator norm, so that for every xΩ¯x\in\overline{\Omega} and r(0,1]r\in(0,1] we have

|B(x,r)Ω|crn.|B(x,r)\cap\Omega|\geq cr^{n}.
Proof.

We will only make the proof for BVBV-extension domains. For W1,1W^{1,1}-extension domains one can use the results from [10], or the fact that W1,1W^{1,1}-extension domains are BVBV-extension domains. A proof of this fact can be found in [14, Lemma 2.4]. The reader will notice that the key point will be to apply the Sobolev embedding theorem, which is both valid for W1,1W^{1,1} and BVBV functions.

Let us denote r0=rr_{0}=r. By induction, we define for every ii\in\mathbb{N} the radius ri(0,ri1)r_{i}\in(0,r_{i-1}) by the equality

|ΩB(x,ri)|=12|ΩB(x,ri1)|=2i|ΩB(x,r0)|.|\Omega\cap B(x,r_{i})|=\frac{1}{2}|\Omega\cap B(x,r_{i-1})|=2^{-i}|\Omega\cap B(x,r_{0})|.

Since xΩ¯x\in\overline{\Omega}, we have that ri0r_{i}\searrow 0 as ii\to\infty.

For each ii\in\mathbb{N}, consider the function fi:Ωf_{i}:\Omega\to\mathbb{R}

fi(y)={1,for yB(x,ri)Ω,ri1|xy|ri1ri,for y(B(x,ri1)B(x,ri))Ω,0,otherwise.f_{i}(y)=\begin{cases}1,&\text{for }y\in B(x,r_{i})\cap\Omega,\\ \frac{r_{i-1}-|x-y|}{r_{i-1}-r_{i}},&\text{for }y\in(B(x,r_{i-1})\setminus B(x,r_{i}))\cap\Omega,\\ 0,&\text{otherwise}.\end{cases}

Note that these functions belong to the class W1,1(Ω)W^{1,1}(\Omega), in particular they are BVBV functions. We can estimate their BVBV-norms by

fiBV(Ω)\displaystyle\|f_{i}\|_{BV(\Omega)} =fiL1(Ω)+Dfi(Ω)=Ω|f|+Ω|fi|\displaystyle=\|f_{i}\|_{L^{1}(\Omega)}+\|D\,f_{i}\|(\Omega)=\int_{\Omega}|f|+\int_{\Omega}|\nabla f_{i}|
|B(x,ri1)Ω|+|riri1|1|(B(x,ri1)B(x,ri))Ω)|\displaystyle\leq|B(x,r_{i-1})\cap\Omega|+|r_{i}-r_{i-1}|^{-1}|(B(x,r_{i-1})\setminus B(x,r_{i}))\cap\Omega)|
C|riri1|12i|ΩB(x,r0)|.\displaystyle\leq C|r_{i}-r_{i-1}|^{-1}2^{-i}|\Omega\cap B(x,r_{0})|.

Call 1=nn11^{*}=\frac{n}{n-1} and denote by T:BV(Ω)BV(n)T\colon BV(\Omega)\to BV(\mathbb{R}^{n}) the extension operator. By the Sobolev inequality for BV functions (see [6, Theorem 5.10]) we know that

TfiL1(n)CD(Tfi)(n),\|Tf_{i}\|_{L^{1^{*}}(\mathbb{R}^{n})}\leq C\|D(Tf_{i})\|(\mathbb{R}^{n}),

where C>0C>0 depends only on the dimension nn. Hence we have the following chain of inequalities

fiL1(Ω)TfiL1(n)CD(Tfi)(n)CTfiBV(Ω).\|f_{i}\|_{L^{1^{*}}(\Omega)}\leq\|Tf_{i}\|_{L^{1*}(\mathbb{R}^{n})}\leq C\|D(Tf_{i})\|(\mathbb{R}^{n})\leq C\|T\|\,\|f_{i}\|_{BV(\Omega)}.

We also have

Ω|fi(y)|1𝑑y\displaystyle\int_{\Omega}|f_{i}(y)|^{1^{*}}\,dy |B(x,ri)Ω|=2i|B(x,r0)Ω|,\displaystyle\geq|B(x,r_{i})\cap\Omega|=2^{-i}|B(x,r_{0})\cap\Omega|,

and therefore

2i|B(x,r0)Ω|\displaystyle 2^{-i}|B(x,r_{0})\cap\Omega| CT1(|riri1|12i|ΩB(x,r0)|)1.\displaystyle\leq C\|T\|^{1^{*}}\left(|r_{i}-r_{i-1}|^{-1}2^{-i}|\Omega\cap B(x,r_{0})|\right)^{1^{*}}.

Consequently,

ri1ri\displaystyle r_{i-1}-r_{i} CT2i(1/11)|ΩB(x,r0)|11/1\displaystyle\leq C\|T\|2^{i(1/1^{*}-1)}|\Omega\cap B(x,r_{0})|^{1-1/1^{*}}
=CT2i/n|ΩB(x,r0)|1/n.\displaystyle=C\|T\|2^{-i/n}|\Omega\cap B(x,r_{0})|^{1/n}.

By summing up all these quantities we conclude that

r=r0=i=1(ri1ri)CTi=12i/n|ΩB(x,r)|1/n=CT21/n1|ΩB(x,r)|1/n.r=r_{0}=\sum_{i=1}^{\infty}(r_{i-1}-r_{i})\leq C\|T\|\sum_{i=1}^{\infty}2^{-i/n}|\Omega\cap B(x,r)|^{1/n}=\frac{C\|T\|}{2^{1/n}-1}|\Omega\cap B(x,r)|^{1/n}.

This gives the claimed inequality. ∎

3. Equivalence of W1,1W^{1,1}-extension and strong BVBV-extension domains

This section is devoted to the proof of Theorem 1.3. The idea in going from a strong BVBV-extension to a W1,1W^{1,1}-extension is to first extend the W1,1W^{1,1}-function from the domain as a BVBV-function to the whole space and then mollify it in the exterior of the domain. In the mollification process it is important to check that we do not change the function too much near the boundary.

3.1. Whitney smoothing operator

In this subsection we prove existence of a suitable smoothing operator from BV to W1,1W^{1,1}. For similar constructions we recommend to the reader to have a look at [3, 9, 18].

Theorem 3.1.

Let ABnA\subset B\subset\mathbb{R}^{n} be open subsets. There exist a constant CC depending only on the dimension nn and a linear operator

SB,A:BV(B){uBV(B):u|AW1,1(A)}S_{B,A}\colon BV(B)\to\left\{u\in BV(B)\,:\,u|_{A}\in W^{1,1}(A)\right\}

so that for any uBV(B)u\in BV(B) we have SB,Au|BA=uS_{B,A}u|_{B\setminus A}=u,

SB,AuBV(B)CuBV(B),\|S_{B,A}u\|_{BV(B)}\leq C\|u\|_{BV(B)}, (3.1)

and

D(SB,Auu)(A)=0,\|D(S_{B,A}u-u)\|(\partial A)=0, (3.2)

where SB,AuuS_{B,A}u-u is understood to be defined in the whole n\mathbb{R}^{n} via a zero-extension. Moreover, the operator SB,AS_{B,A} is also bounded when acting from the space BVl(A)BV_{l}(A) into the homogeneous Sobolev space L1,1(A)L^{1,1}(A).

Recall that L1,1(A)={uLloc1(A):uL1(A)}L^{1,1}(A)=\{u\in L^{1}_{loc}(A):\,\nabla u\in L^{1}(A)\} stands for the homogeneous Sobolev space endowed with the seminorm uL1,1(A)=uL1(A)\|u\|_{L^{1,1}(A)}=\|\nabla u\|_{L^{1}(A)}.

Let us briefly explain how the operator SB,AS_{B,A} is constructed. We first take a Whitney decomposition of the open set AA and a partition of unity based on it. The operator on a BVBV-function uu is then defined as the sum of uu restricted to the complement of AA and the average values of uu in each Whitney cube of AA times the associated partition function. This way, we immediately have that the function SB,AS_{B,A} is left unchanged in the complement of AA, and that in AA it is smooth. The inequality (3.1) will follow in a standard way from the Poincaré inequality for BVBV-functions, whereas for showing (3.2) we will show that the average difference between uu and SB,AuS_{B,A}u near n1\mathcal{H}^{n-1}-almost every boundary point of AA tends to zero as we get closer to the point.

Let us now give the definition of the operator doing the smoothing part. Suppose AnA\subset\mathbb{R}^{n} is an open set, not equal to the entire space n\mathbb{R}^{n}. Let 𝒲={Qi}i=1\mathcal{W}=\{Q_{i}\}_{i=1}^{\infty} be the standard Whitney decomposition of AA, by which we mean that it satisfies the following properties:

  • (W1)

    Each QiQ_{i} is a dyadic cube inside AA.

  • (W2)

    A=iQiA=\bigcup_{i}Q_{i} and for every iji\neq j we have int(Qi)int(Qj)=\text{int}(Q_{i})\cap\text{int}(Q_{j})=\emptyset.

  • (W3)

    For every ii we have (Qi)missingdist(Qi,A)4n(Qi)\ell(Q_{i})\leq{\mathop{\mathrm{missing}}{\,dist\,}}(Q_{i},\partial A)\leq 4\sqrt{n}\ell(Q_{i}),

  • (W4)

    If QiQjQ_{i}\cap Q_{j}\neq\emptyset, we have 14(Qi)(Qj)4(Qi)\frac{1}{4}\ell(Q_{i})\leq\ell(Q_{j})\leq 4\ell(Q_{i}).

The reader can find a proof of the existence of such a dyadic decomposition of the set AA in [23, Chapter VI].

For a given set AA and its Whitney decomposition 𝒲\mathcal{W} we take a partition of unity {ψi}i=1\{\psi_{i}\}_{i=1}^{\infty} so that for every ii we have ψiC(n)\psi_{i}\in C^{\infty}(\mathbb{R}^{n}), spt(ψi)={xn:ψi(x)0}B(Qi,18(Qi))\text{spt}(\psi_{i})=\{x\in\mathbb{R}^{n}:\,\psi_{i}(x)\neq 0\}\subset B(Q_{i},\frac{1}{8}\ell(Q_{i})), ψi0\psi_{i}\geq 0, |ψi|C(Qi)1|\nabla\psi_{i}|\leq C\ell(Q_{i})^{-1} with a constant CC depending only on nn, and

i=1ψi=χA.\sum_{i=1}^{\infty}\psi_{i}=\chi_{A}.

With the partition of unity we then define for any uBV(A)u\in BV(A) a function

S𝒲u=i=1uQiψi.S_{\mathcal{W}}u=\sum_{i=1}^{\infty}u_{Q_{i}}\psi_{i}. (3.3)

Let us start by showing that S𝒲S_{\mathcal{W}} maps to W1,1(A)W^{1,1}(A) boundedly. Even though we could obviously equivalently use the BV norm also on the target, we prefer to write it as the W1,1W^{1,1}-norm in order to underline the spaces where the operator will be used.

Lemma 3.2.

Let S𝒲S_{\mathcal{W}} be the operator defined in (3.3). Then for any uBV(A)u\in BV(A) we have S𝒲uC(A)S_{\mathcal{W}}u\in C^{\infty}(A) and S𝒲uW1,1(A)CuBV(A)\|S_{\mathcal{W}}u\|_{W^{1,1}(A)}\leq C\|u\|_{BV(A)} with a constant CC depending only on nn.

Proof.

By (W2) and the fact that spt(ψi)B(Qi,18(Qi))\text{spt}(\psi_{i})\subset B(Q_{i},\frac{1}{8}\ell(Q_{i})) for every ii, we know that spt(ψi)spt(ψj)\text{spt}(\psi_{i})\cap\text{spt}(\psi_{j})\neq\emptyset implies that QiQjQ_{i}\cap Q_{j}\neq\emptyset. Therefore, any point in AA has a neighbourhood where S𝒲uS_{\mathcal{W}}u is defined as a sum of finitely many CC^{\infty}-functions. Consequently, S𝒲uC(A)S_{\mathcal{W}}u\in C^{\infty}(A). For the L1L^{1}-norm of the function we can estimate

S𝒲uL1(A)i=1uQiψiL1(A)=i=1|uQi|ψiL1(A)i=1|uQi|2n(Qi)n=2nuL1(A).\|S_{\mathcal{W}}u\|_{L^{1}(A)}\leq\sum_{i=1}^{\infty}\|u_{Q_{i}}\psi_{i}\|_{L^{1}(A)}=\sum_{i=1}^{\infty}|u_{Q_{i}}|\|\psi_{i}\|_{L^{1}(A)}\leq\sum_{i=1}^{\infty}|u_{Q_{i}}|2^{n}\ell(Q_{i})^{n}=2^{n}\|u\|_{L^{1}(A)}.

For the estimate on the L1L^{1}-norm of the gradient we start with an estimate via the (1,1)(1,1)-Poincaré inequality (2.1)

(S𝒲u)L1(Qi)\displaystyle\|\nabla(S_{\mathcal{W}}u)\|_{L^{1}(Q_{i})} QjQi|uQiuQj|ψjL1(A)\displaystyle\leq\sum_{Q_{j}\cap Q_{i}\neq\emptyset}|u_{Q_{i}}-u_{Q_{j}}|\|\nabla\psi_{j}\|_{L^{1}(A)}
QjQi|uQiuQj|C(Qj)n1\displaystyle\leq\sum_{Q_{j}\cap Q_{i}\neq\emptyset}|u_{Q_{i}}-u_{Q_{j}}|C\ell(Q_{j})^{n-1}
CQjQi(Qj)1QiQj|uQiu(y)|+|u(y)uQj|dy\displaystyle\leq C\sum_{Q_{j}\cap Q_{i}\neq\emptyset}\ell(Q_{j})^{-1}\int_{Q_{i}\cup Q_{j}}|u_{Q_{i}}-u(y)|+|u(y)-u_{Q_{j}}|\,dy
CQjQi(Qj)1(2QiQj|uQiQju(y)|+2QiQj|uQiQjuQj|𝑑y)\displaystyle\leq C\sum_{Q_{j}\cap Q_{i}\neq\emptyset}\ell(Q_{j})^{-1}\left(2\int_{Q_{i}\cup Q_{j}}|u_{Q_{i}\cup Q_{j}}-u(y)|+2\int_{Q_{i}\cup Q_{j}}|u_{Q_{i}\cup Q_{j}}-u_{Q_{j}}|\,dy\right)
CQjQi(Du(QiQj)),\displaystyle\leq C\sum_{Q_{j}\cap Q_{i}\neq\emptyset}(\|Du\|(Q_{i}\cup Q_{j})),

which then gives, by summing over all ii, and noticing that in the final double sum the sets QiQjQ_{i}\cup Q_{j} have finite overlap with a constant depending only on nn,

(S𝒲u)L1(A)=i=1(S𝒲u)L1(Qi)Ci=1QjQi(Du(QiQj))CDu(A).\begin{split}\|\nabla(S_{\mathcal{W}}u)\|_{L^{1}(A)}&=\sum_{i=1}^{\infty}\|\nabla(S_{\mathcal{W}}u)\|_{L^{1}(Q_{i})}\\ &\leq C\sum_{i=1}^{\infty}\sum_{Q_{j}\cap Q_{i}\neq\emptyset}(\|Du\|(Q_{i}\cup Q_{j}))\\ &\leq C\|Du\|(A).\end{split} (3.4)

This concludes the proof of the lemma. ∎

The next lemma gives the crucial boundary behaviour that will imply (3.2).

Lemma 3.3.

For the operator S𝒲S_{\mathcal{W}} defined in (3.3) and for any uBV(A)u\in BV(A) we have

limr01|B(x,r)|B(x,r)A|S𝒲u(y)u(y)|𝑑y=0\lim_{r\searrow 0}\frac{1}{|B(x,r)|}\int_{B(x,r)\cap A}|S_{\mathcal{W}}u(y)-u(y)|\,dy=0 (3.5)

for n1\mathcal{H}^{n-1}-almost every point xAx\in\partial A.

Proof.

Suppose (3.5) fails on a set FAF\subset\partial A with n1(F)>0\mathcal{H}^{n-1}(F)>0. Without loss of generality, we may assume FF compact. By going to a subset of FF if needed, we may further assume that there exists a constant δ>0\delta>0 so that

lim supr01|B(x,r)|B(x,r)A|S𝒲u(y)u(y)|𝑑y>δ\limsup_{r\searrow 0}\frac{1}{|B(x,r)|}\int_{B(x,r)\cap A}|S_{\mathcal{W}}u(y)-u(y)|\,dy>\delta

for every xFx\in F.

Let ε>0\varepsilon>0. By the 5r5r-covering lemma there exists a disjointed countable collection {B(xi,ri)}iI\{B(x_{i},r_{i})\}_{i\in I} so that xiFx_{i}\in F, ri<εr_{i}<\varepsilon for all ii,

|B(xi,ri)|1δB(xi,ri)A|S𝒲u(y)u(y)|𝑑y|B(x_{i},r_{i})|\leq\frac{1}{\delta}\int_{B(x_{i},r_{i})\cap A}|S_{\mathcal{W}}u(y)-u(y)|\,dy (3.6)

and

FiIB(xi,5ri).F\subset\bigcup_{i\in I}B(x_{i},5r_{i}).

Similarly as in the proof of Lemma 3.2, we first estimate in a Whitney cube Q𝒲Q\in\mathcal{W} using the (1,1)-Poincaré inequality (2.1)

Q|S𝒲u(y)u(y)|𝑑y=Q|QiQ(uQiψi(y)u(y)ψi(y))|𝑑yQiQQ|uQiu(y)|𝑑yQiQQQi|uQiu(y)|𝑑yC(Q)QiQDu(QQi).\begin{split}\int_{Q}|S_{\mathcal{W}}u(y)-u(y)|\,dy&=\int_{Q}\left|\sum_{Q_{i}\cap Q\neq\emptyset}(u_{Q_{i}}\psi_{i}(y)-u(y)\psi_{i}(y))\right|\,dy\\ &\leq\sum_{Q_{i}\cap Q\neq\emptyset}\int_{Q}|u_{Q_{i}}-u(y)|\,dy\\ &\leq\sum_{Q_{i}\cap Q\neq\emptyset}\int_{Q\cup Q_{i}}|u_{Q_{i}}-u(y)|\,dy\\ &\leq C\ell(Q)\sum_{Q_{i}\cap Q\neq\emptyset}\|Du\|(Q\cup Q_{i}).\end{split} (3.7)

By the property (W3) of the Whitney decomposition, we conclude that if Q𝒲Q\in\mathcal{W} is such that QB(xi,ri)Q\cap B(x_{i},r_{i})\neq\emptyset, we have

(Q)missingdist(Q,A)missingdist(Q,xi)<ri,\ell(Q)\leq{\mathop{\mathrm{missing}}{\,dist\,}}(Q,\partial A)\leq{\mathop{\mathrm{missing}}{\,dist\,}}(Q,x_{i})<r_{i},

and hence

QB(xi,(n+1)ri)B(F,(n+1)ε).Q\subset B(x_{i},(\sqrt{n}+1)r_{i})\subset B(F,(\sqrt{n}+1)\varepsilon).

Similarly, for the same QQ, if QiQQ_{i}\cap Q\neq\emptyset for some Qi𝒲Q_{i}\in\mathcal{W}, by (W4), we get

(Qi)4(Q),\ell(Q_{i})\leq 4\ell(Q),

and so

QiB(xi,(5n+1)ri)B(F,(5n+1)ε).Q_{i}\subset B(x_{i},(5\sqrt{n}+1)r_{i})\subset B(F,(5\sqrt{n}+1)\varepsilon).

Now, using the definition of the Hausdorff content, the inequality (3.6), the estimate (3.7), and the above consideration for the cubes QQ, we get

5ϵn1(F)\displaystyle\mathcal{H}_{5\epsilon}^{n-1}(F) CiIrin1CiI|B(xi,ri)|ri\displaystyle\leq C\sum_{i\in I}r_{i}^{n-1}\leq C\sum_{i\in I}\frac{|B(x_{i},r_{i})|}{r_{i}}
CiI1δriB(xi,ri)A|S𝒲u(y)u(y)|𝑑y\displaystyle\leq C\sum_{i\in I}\frac{1}{\delta r_{i}}\int_{B(x_{i},r_{i})\cap A}|S_{\mathcal{W}}u(y)-u(y)|\,dy
CQ𝒲iI1δ(Q)(Q)riQχB(xi,ri)(y)|S𝒲u(y)u(y)|𝑑y\displaystyle\leq C\sum_{Q\in\mathcal{W}}\sum_{i\in I}\frac{1}{\delta\ell(Q)}\frac{\ell(Q)}{r_{i}}\int_{Q}\chi_{B(x_{i},r_{i})}(y)|S_{\mathcal{W}}u(y)-u(y)|\,dy
Cδ(QB(F,(n+1)ε)1(Q)Q|S𝒲u(y)u(y)|𝑑y)\displaystyle\leq\frac{C}{\delta}\left(\sum_{Q\cap B(F,(\sqrt{n}+1)\varepsilon)\neq\emptyset}\frac{1}{\ell(Q)}\int_{Q}|S_{\mathcal{W}}u(y)-u(y)|\,dy\right)
Cδ(QB(F,(n+1)ε)QiQDu(QQi))\displaystyle\leq\frac{C}{\delta}\left(\sum_{Q\cap B(F,(\sqrt{n}+1)\varepsilon)\neq\emptyset}\sum_{Q_{i}\cap Q\neq\emptyset}\|Du\|(Q\cup Q_{i})\right)
CδDu(B(F,(5n+1)ϵ)A)0\displaystyle\leq\frac{C}{\delta}\|Du\|(B(F,(5\sqrt{n}+1)\epsilon)\cap A)\searrow 0

as ε0\varepsilon\searrow 0. Thus

n1(F)=0,\mathcal{H}^{n-1}(F)=0,

giving a contradiction and concluding the proof. ∎

With the previous two lemmas we can now prove the main theorem of the section.

Proof of Theorem 3.1.

Let S𝒲S_{\mathcal{W}} be the operator defined in (3.3) and suppose that uBV(B)u\in BV(B) is given. We define

SB,Au=u|BA+S𝒲u|A.S_{B,A}u=u|_{B\setminus A}+S_{\mathcal{W}}u|_{A}.

Consider SB,AuuL1(B)S_{B,A}u-u\in L^{1}(B) for which, by (3.5), we have

limr01|B(x,r)|B(x,r)A|SB,Au(y)u(y)|𝑑y=0\lim_{r\searrow 0}\frac{1}{|B(x,r)|}\int_{B(x,r)\cap A}|S_{B,A}u(y)-u(y)|\,dy=0 (3.8)

for n1\mathcal{H}^{n-1}-almost every xAx\in\partial A. Observe that SB,Auu=0S_{B,A}u-u=0 on BAB\setminus A.

Let us introduce the superlevel sets Et={yn:SB,Au(y)u(y)>t}E_{t}=\{y\in\mathbb{R}^{n}:\,S_{B,A}u(y)-u(y)>t\} for every tt\in\mathbb{R}, where SB,AuuS_{B,A}u-u is defined in the whole n\mathbb{R}^{n} via a zero-extension. We want to show that n1(MEtA)=0\mathcal{H}^{n-1}(\partial^{M}E_{t}\cap\partial A)=0 for almost every tt\in\mathbb{R} and the equality (3.2) will follow by a simple application of the coarea formula. We proceed as follows.

In the case that t<0t<0, observe that for every yAEty\in A\setminus E_{t} we have |SB,Au(y)u(y)||t||S_{B,A}u(y)-u(y)|\geq|t|, then for n1\mathcal{H}^{n-1}-almost all xAx\in\partial A, by (3.8),

D¯(AEt,x)\displaystyle\overline{D}(A\setminus E_{t},x) =lim supr0|AEtB(x,r)||B(x,r)|\displaystyle=\limsup_{r\searrow 0}\frac{|A\setminus E_{t}\cap B(x,r)|}{|B(x,r)|}
lim supr01|t||B(x,r)|AB(x,r)|SB,Au(y)u(y)|𝑑y=0.\displaystyle\leq\limsup_{r\searrow 0}\frac{1}{|t||B(x,r)|}\int_{A\cap B(x,r)}|S_{B,A}u(y)-u(y)|\,dy=0.

This, together with the fact that BAEtB\setminus A\subset E_{t}, means that the set EtE_{t} has density 11 at n1\mathcal{H}^{n-1}-almost all points xAx\in\partial A.

If we take t>0t>0, for every yEty\in E_{t} we have |SB,Au(y)u(y)|t|S_{B,A}u(y)-u(y)|\geq t, and then for n1\mathcal{H}^{n-1}-almost all xAx\in\partial A, again by (3.8),

D¯(Et,x)\displaystyle\overline{D}(E_{t},x) =lim supr0|EtB(x,r)||B(x,r)|\displaystyle=\limsup_{r\searrow 0}\frac{|E_{t}\cap B(x,r)|}{|B(x,r)|}
lim supr01t|B(x,r)|AB(x,r)|SB,Au(y)u(y)|𝑑y=0.\displaystyle\leq\limsup_{r\searrow 0}\frac{1}{t|B(x,r)|}\int_{A\cap B(x,r)}|S_{B,A}u(y)-u(y)|\,dy=0.

This means, using EtAE_{t}\subset A, that the set EtE_{t} has density 0 at n1\mathcal{H}^{n-1}-almost all points xAx\in\partial A.

From these previous observations we deduce that n1(MEtA)=0\mathcal{H}^{n-1}(\partial^{M}E_{t}\cap\partial A)=0 for all t0t\neq 0. We therefore obtain (3.2), applying the coarea formula,

D(SB,Auu)(A)=n1(MEtA)𝑑t=0.\|D(S_{B,A}u-u)\|(\partial A)=\int^{\infty}_{-\infty}\mathcal{H}^{n-1}(\partial^{M}E_{t}\cap\partial A)\,dt=0.

We now combine this with Lemma 3.2 to obtain (3.1) and hence also that SB,AuBV(B)S_{B,A}u\in BV(B). We get

D(SB,Au)(B)\displaystyle\|D(S_{B,A}u)\|(B) Du(B)+D(SB,Au)u(B)\displaystyle\leq\|Du\|(B)+\|D(S_{B,A}u)-u\|(B)
=Du(B)+D(SB,Au)u(A)\displaystyle=\|Du\|(B)+\|D(S_{B,A}u)-u\|(A)
Du(B)+D(S𝒲u|A)(A)+Du(A)\displaystyle\leq\|Du\|(B)+\|D(S_{\mathcal{W}}u|_{A})\|(A)+\|Du\|(A)
Du(B)+CDu|A(A)+Du(A)\displaystyle\leq\|Du\|(B)+C\|Du|_{A}\|(A)+\|Du\|(A)
(C+2)Du(B)\displaystyle\leq(C+2)\|Du\|(B)

and conclude the proof. ∎

3.2. Proof of Theorem 1.3

In this section we will prove Theorem 1.3 with the help of Theorem 3.1. Recall that we are claiming that for a bounded domain Ωn\Omega\subset\mathbb{R}^{n} the following are equivalent:

  1. (1)

    Ω\Omega is a W1,1W^{1,1}-extension domain.

  2. (2)

    Ω\Omega is a strong BVBV-extension domain.

  3. (3)

    Ω\Omega has the strong extension property for sets of finite perimeter.

We will show the equivalence by showing the implications

(1)(3)(2)(1).\text{(1)}\Longrightarrow\text{(3)}\Longrightarrow\text{(2)}\Longrightarrow\text{(1)}.
Proof of the implication (1) \Longrightarrow (3).

We start with the assumption that Ω\Omega is a bounded W1,1W^{1,1}-extension domain. In particular, it is known that Ω\Omega is also a L1,1L^{1,1}-extension domain (see [13]). That is, there exists an extension operator T:L1,1(Ω)L1,1(n)T\colon L^{1,1}(\Omega)\to L^{1,1}(\mathbb{R}^{n}) with (Tu)L1(n)TuL1(Ω)\|\nabla(Tu)\|_{L^{1}(\mathbb{R}^{n})}\leq\|T\|\|\nabla u\|_{L^{1}(\Omega)} for every uL1,1(Ω)u\in L^{1,1}(\Omega). Since Ω\Omega is bounded, after multiplying with a suitable Lipschitz cutoff-function we may assume that TuL1(n)Tu\in L^{1}(\mathbb{R}^{n}) and still keep the control on the gradient norm.

We claim that Ω\Omega has the strong extension property for sets of finite perimeter. Thus, let EΩE\subset\Omega be a set of finite perimeter in Ω\Omega. We need to find a set E~n\widetilde{E}\subset\mathbb{R}^{n} so that (PE1)–(PE3) of Definition 1.2 hold.

Towards this, let SΩ,Ω:BV(Ω)W1,1(Ω)S_{\Omega,\Omega}\colon BV(\Omega)\to W^{1,1}(\Omega) be the operator given by Theorem 3.1. We now define a function vW1,1(n)v\in W^{1,1}(\mathbb{R}^{n}) by

v=TSΩ,ΩχE.v=TS_{\Omega,\Omega}\chi_{E}.

By truncating the function if needed, we may assume that 0v10\leq v\leq 1.

Applying the coarea formula (Theorem 2.2) for the function vv,

01P({v>t},n)𝑑t=Dv(n)=n|v(y)|𝑑y<.\int_{0}^{1}P(\{v>t\},\mathbb{R}^{n})\,dt=\|Dv\|({\mathbb{R}^{n}})=\int_{\mathbb{R}^{n}}|\nabla v(y)|\,dy<\infty.

This gives, in particular, that there exists a set I[0,1]I\subset[0,1] with 1(I)12\mathcal{H}^{1}(I)\geq\frac{1}{2} for which for every tIt\in I we have

P({v>t},n)2Dv(n)=2n|v(y)|𝑑y2TSΩ,ΩχEL1(Ω)2TCDχE(Ω)=2CTP(E,Ω).\begin{split}P(\{v>t\},\mathbb{R}^{n})&\leq 2\|Dv\|({\mathbb{R}^{n}})=2\int_{\mathbb{R}^{n}}|\nabla v(y)|\,dy\leq 2\|T\|\,\|\nabla S_{\Omega,\Omega}\chi_{E}\|_{L^{1}(\Omega)}\\ &\leq 2\|T\|\,C\|D\chi_{E}\|(\Omega)=2C\|T\|\,P(E,\Omega).\end{split} (3.9)

In the penultimate inequality we are using (3.4).

By the measure density (Proposition 2.3) we have |Ω|=0|\partial\Omega|=0. This together with the fact that vL1(n)\nabla v\in L^{1}(\mathbb{R}^{n}), gives

Dv(Ω)=0,\|Dv\|(\partial\Omega)=0,

and by (3.2)

D(χEvχΩ)(Ω)=D(χESΩ,ΩχE)(Ω)=0,\|D(\chi_{E}-v\chi_{\Omega})\|(\partial\Omega)=\|D(\chi_{E}-S_{\Omega,\Omega}\chi_{E})\|(\partial\Omega)=0,

where χESΩ,ΩχE\chi_{E}-S_{\Omega,\Omega}\chi_{E} is understood to be defined in the whole n\mathbb{R}^{n} via a zero-extension. Hence, again by the coarea formula

01n1(M(E({v>t}Ω))Ω)𝑑t\displaystyle\int_{0}^{1}\mathcal{H}^{n-1}(\partial^{M}(E\cup(\{v>t\}\setminus\Omega))\cap\partial\Omega)\,dt =D(χE+vχnΩ)(Ω)\displaystyle=\|D(\chi_{E}+v\chi_{\mathbb{R}^{n}\setminus\Omega})\|(\partial\Omega)
Dv(Ω)+D(χEvχΩ)(Ω)=0.\displaystyle\leq\|Dv\|(\partial\Omega)+\|D(\chi_{E}-v\chi_{\Omega})\|(\partial\Omega)=0.

This gives that for almost every t[0,1]t\in[0,1] we have

n1(M(E({v>t}Ω))Ω)=0.\mathcal{H}^{n-1}(\partial^{M}(E\cup(\{v>t\}\setminus\Omega))\cap\partial\Omega)=0. (3.10)

Let us pick tI[0,1]t\in I\subset[0,1] so that both (3.9) and (3.10) hold, and define

E~=E({v>t}Ω).\widetilde{E}=E\cup(\{v>t\}\setminus\Omega).

Now, it is straightforward that condition (PE1) holds. The equation (3.10) gives (PE3), and together with (3.9) it also implies

P(E~,n)\displaystyle P(\widetilde{E},\mathbb{R}^{n}) =n1(ME~)\displaystyle=\mathcal{H}^{n-1}(\partial^{M}\widetilde{E})
n1((ME)Ω)+n1((ME~)Ω)+n1(M{v>t})\displaystyle\leq\mathcal{H}^{n-1}((\partial^{M}E)\cap\Omega)+\mathcal{H}^{n-1}((\partial^{M}\widetilde{E})\cap\partial\Omega)+\mathcal{H}^{n-1}(\partial^{M}\{v>t\})
P(E,Ω)+2CTP(E,Ω)\displaystyle\leq P(E,\Omega)+2C\|T\|P(E,\Omega)

proving (PE2). ∎

Proof of the implication (3) \Longrightarrow (2).

By assumption Ω\Omega has the strong extension property for sets of finite perimeter, so there exists a constant C>0C>0 such that for any set EΩE\subset\Omega of finite perimeter there exists a set E~n\widetilde{E}\subset\mathbb{R}^{n} such that (PE1)–(PE3) are satisfied.

Take a function uBV(Ω)u\in BV(\Omega) and let BΩ¯B\supset\overline{\Omega} be a large enough ball. Without loss of generality, we may assume that u:Ω[0,1]u\colon\Omega\to[0,1]. Let us write Et={ut}={yΩ:u(y)t}E_{t}=\{u\geq t\}=\{y\in\Omega:\,u(y)\geq t\} for the superlevel sets for each tt. Since uBV(Ω)u\in BV(\Omega), by the coarea formula, P(Et,Ω)<P(E_{t},\Omega)<\infty for almost every t[0,1]t\in[0,1]. For these tt, we select E~t\tilde{E}_{t} to be a strong perimeter extension of EtE_{t}. For convenience, for the remaining tt we define E~t=Et\tilde{E}_{t}=E_{t}. Notice that these are not strong perimeter extensions of EtE_{t}. This will not pose a problem for us, since we will not use these values of tt in the construction below.

Before going to the actual proof, let us note that if the strong perimeter extensions E~t\tilde{E}_{t} could be chosen so that (t,x)χE~t(x)(t,x)\mapsto\chi_{\tilde{E}_{t}}(x) is measurable, by Fubini’s theorem we would obtain

DTu01P(E~t)𝑑t\|DTu\|\leq\int_{0}^{1}P(\tilde{E}_{t})\,dt

for the function Tu(x)=1{t[0,1]:xE~t}Tu(x)=\mathcal{H}^{1}\{t\in[0,1]\,:\,x\in\tilde{E}_{t}\}. In order to circumvent the measurability issue, we proceed by defining the extension TuTu in a similar way, but as a limit of simple functions umu_{m}.

For every t[0,1]t\in[0,1], let us denote by Ik(t)I_{k}(t) the (half-open) dyadic interval of length 2k2^{-k} containing tt. For almost every t[0,1]t\in[0,1] we then have

P(Et,Ω)lim supk2kIk(t)P(Es,Ω)𝑑s.P(E_{t},\Omega)\leq\limsup_{k\to\infty}2^{k}\int_{I_{k}(t)}P(E_{s},\Omega)\,ds. (3.11)

For almost every t[0,1]t\in[0,1] we also have

P(E~t,Ω)=0.P(\tilde{E}_{t},\partial\Omega)=0.

Since Ω\partial\Omega is a compact set, for almost every tt we then have

limkP(E~t,B(Ω,2k))=0.\lim_{k\to\infty}P(\tilde{E}_{t},B(\partial\Omega,2^{-k}))=0.

Let us write for each k,mk,m\in\mathbb{N}

Ikm={t[0,1]:P(E~t,B(Ω,2k))<2m and (3.11) holds}.I_{k}^{m}=\{t\in[0,1]\,:\,P(\tilde{E}_{t},B(\partial\Omega,2^{-k}))<2^{-m}\text{ and \eqref{eq:tdensity} holds}\}.

Notice that the sets IkmI_{k}^{m} are not necessarily measurable. Nevertheless, since 1\mathcal{H}^{1} is a regular outer measure, we have

1(Ikm)1,as k\mathcal{H}^{1}(I_{k}^{m})\nearrow 1,\qquad\text{as }k\to\infty

for every mm\in\mathbb{N}.

We define a sequence (km)m=1(k_{m})_{m=1}^{\infty}\subset\mathbb{N} inductively as follows. First take k1k_{1}\in\mathbb{N} so that

1(Ik11)>121.\mathcal{H}^{1}(I_{k_{1}}^{1})>1-2^{-1}.

Suppose now that kik_{i} has been defined for all i<mi<m. Then we take kmk_{m}\in\mathbb{N} so that

1(j=imIkjj)>12i\mathcal{H}^{1}\left(\bigcap_{j=i}^{m}I_{k_{j}}^{j}\right)>1-2^{-i} (3.12)

for all imi\leq m. Notice that this requirement can be obtained since (3.12) is with a strict inequality and again by outer regularity, for every i<mi<m we have

1(Ikmj=im1Ikjj)1(j=im1Ikjj)\mathcal{H}^{1}\left(I_{k}^{m}\cap\bigcap_{j=i}^{m-1}I_{k_{j}}^{j}\right)\to\mathcal{H}^{1}\left(\bigcap_{j=i}^{m-1}I_{k_{j}}^{j}\right)

as kk\to\infty.

Now, for mm\in\mathbb{N} we also take lml_{m}\in\mathbb{N} for which

1(Jm)>12m,\mathcal{H}^{1}(J^{m})>1-2^{-m}, (3.13)

where

Jm={t[0,1]:P(Et,Ω)2lm+1Ilm(t)P(Es,Ω)𝑑s}J^{m}=\left\{t\in[0,1]\,:\,P(E_{t},\Omega)\leq 2^{l_{m}+1}\int_{I_{l_{m}}(t)}P(E_{s},\Omega)\,ds\right\}

The index lml_{m} then gives us the scale at which the simple function umu_{m} is constructed.

Let us now construct the function umu_{m} for a given mm\in\mathbb{N}. For each j{1,,2lm}j\in\{1,\dots,2^{l_{m}}\} define

ijm=min{i:[(j1)2lm,j2lm)Jmh=imIkhh}.i_{j}^{m}=\min\left\{i\,:\,[(j-1)2^{-l_{m}},j2^{-l_{m}})\cap J^{m}\cap\bigcap_{h=i}^{m}I_{k_{h}}^{h}\neq\emptyset\right\}.

Notice that always ijmm+1i_{j}^{m}\leq m+1 since [(j1)2lm,j2lm)Jm[(j-1)2^{-l_{m}},j2^{-l_{m}})\cap J^{m}\neq\emptyset.

We then select

tjm[(j1)2lm,j2lm)Jmh=ijmmIkhh.t_{j}^{m}\in[(j-1)2^{-l_{m}},j2^{-l_{m}})\cap J^{m}\cap\bigcap_{h=i_{j}^{m}}^{m}I_{k_{h}}^{h}.

Next, we define

um=j=12lm2lmχE~tjm,u_{m}=\sum_{j=1}^{2^{l_{m}}}2^{-l_{m}}\chi_{\tilde{E}_{t_{j}^{m}}},

which satisfies 0um10\leq u_{m}\leq 1 and umBV(B)u_{m}\in BV(B).

For every i{1,,m}i\in\{1,\dots,m\}, let us denote

Kim={j{1,,2lm}:[(j1)2lm,j2lm)Jmh=imIkhh=}K_{i}^{m}=\left\{j\in\{1,\dots,2^{l_{m}}\}\,:\,[(j-1)2^{-l_{m}},j2^{-l_{m}})\cap J^{m}\cap\bigcap_{h=i}^{m}I_{k_{h}}^{h}=\emptyset\right\}

and

Bim=jKim[(j1)2lm,j2lm).B_{i}^{m}=\bigcup_{j\in K_{i}^{m}}[(j-1)2^{-l_{m}},j2^{-l_{m}}).

Since JmJ^{m} is measurable, (3.13) gives 1([0,1]Jm)<2m\mathcal{H}^{1}\left([0,1]\setminus J_{m}\right)<2^{-m}, and thus, by (3.12)

1(h=imIkhhJm)\displaystyle\mathcal{H}^{1}\left(\bigcap_{h=i}^{m}I_{k_{h}}^{h}\cap J_{m}\right) =1(h=imIkhh)1(h=imIkhhJm)\displaystyle=\mathcal{H}^{1}\left(\bigcap_{h=i}^{m}I_{k_{h}}^{h}\right)-\mathcal{H}^{1}\left(\bigcap_{h=i}^{m}I_{k_{h}}^{h}\setminus J_{m}\right)
1(h=imIkhh)1([0,1]Jm)\displaystyle\geq\mathcal{H}^{1}\left(\bigcap_{h=i}^{m}I_{k_{h}}^{h}\right)-\mathcal{H}^{1}\left([0,1]\setminus J_{m}\right)
>12i2m12i+1.\displaystyle>1-2^{-i}-2^{-m}\geq 1-2^{-i+1}.

Hence, we have

1(Bim)<2i+1.\mathcal{H}^{1}(B_{i}^{m})<2^{-i+1}. (3.14)

For the norm of DumDu_{m}, by the fact that tjmJmt_{j}^{m}\in J^{m} for every jj, we get the estimate

Dum(n)\displaystyle\|Du_{m}\|(\mathbb{R}^{n}) j=12lm2lmP(E~tjm,n)Cj=12lm2lmP(Etjm,Ω)\displaystyle\leq\sum_{j=1}^{2^{l_{m}}}2^{-l_{m}}P(\tilde{E}_{t_{j}^{m}},\mathbb{R}^{n})\leq C\sum_{j=1}^{2^{l_{m}}}2^{-l_{m}}P(E_{t_{j}^{m}},\Omega)
Cj=12lm2[(j1)2lm,j2lm)P(Es,Ω)=2CDu(Ω).\displaystyle\leq C\sum_{j=1}^{2^{l_{m}}}2\int_{[(j-1)2^{-l_{m}},j2^{-l_{m}})}P(E_{s},\Omega)=2C\|Du\|(\Omega).

Hence, there exists a subsequence of (um)m=1(u_{m})_{m=1}^{\infty}, which converges in L1(B)L^{1}(B) to a function vBV(B)v\in BV(B). For it, we have

Dv(B)lim supmDum(n)2CDu(Ω).\|Dv\|(B)\leq\limsup_{m\to\infty}\|Du_{m}\|(\mathbb{R}^{n})\leq 2C\|Du\|(\Omega).

Moreover, clearly v=uv=u on Ω\Omega.

In order to estimate Dv(Ω)\|Dv\|(\partial\Omega) we observe that, for every i{1,,m}i\in\{1,\dots,m\}, we have, by (3.14),

Dum(B(Ω,2ki))j=12lm2lmP(E~tj,B(Ω,2ki))jKim2lmP(E~tj,B(Ω,2ki))+jKim2lmP(E~tj,n)2i+C2BimP(Es,Ω)𝑑s2i+C2δ(2i+1),\begin{split}\|Du_{m}\|(B(\partial\Omega,2^{-k_{i}}))&\leq\sum_{j=1}^{2^{l_{m}}}2^{-l_{m}}P(\tilde{E}_{t_{j}},B(\partial\Omega,2^{-k_{i}}))\\ &\leq\sum_{j\notin K_{i}^{m}}2^{-l_{m}}P(\tilde{E}_{t_{j}},B(\partial\Omega,2^{-k_{i}}))+\sum_{j\in K_{i}^{m}}2^{-l_{m}}P(\tilde{E}_{t_{j}},\mathbb{R}^{n})\\ &\leq 2^{-i}+C2\int_{B_{i}^{m}}P(E_{s},\Omega)\,ds\\ &\leq 2^{-i}+C2\delta(2^{-i+1}),\end{split} (3.15)

where

δ(r)=sup{AP(Es,Ω)𝑑s:A[0,1],1(A)=r}0\delta(r)=\sup\left\{\int_{A}P(E_{s},\Omega)\,ds\,:\,A\subset[0,1],\mathcal{H}^{1}(A)=r\right\}\to 0

as r0r\to 0 by the absolute continuity of the integral. Since the upper bound in (3.15) goes to zero as ii\to\infty independently of mm, we have

Dv(Ω)=0.\|Dv\|(\partial\Omega)=0.

Let us assume that the function vv is extended as zero outside BB. Recall that we have

Dv(B)2CDu(Ω)andDv(Ω)=0.\|Dv\|(B)\leq 2C\|Du\|(\Omega)\;\;\text{and}\;\;\|Dv\|(\partial\Omega)=0.

In order to conclude the proof we control the BVBV-norm of the function vv in the whole n\mathbb{R}^{n} as follows.

Consider a Lipschitz function η\eta which takes the value 11 on Ω\Omega and has support in BB. Then one can check that

D(ηv)(n)CDv(B)\|D(\eta v)\|(\mathbb{R}^{n})\leq C\|Dv\|(B)

and using the Poincaré inequality that

ηvL1(n)=ηvL1(B)CD(ηv)(B)CDv(B).\|\eta v\|_{L^{1}(\mathbb{R}^{n})}=\|\eta v\|_{L^{1}(B)}\leq C\|D(\eta v)\|(B)\leq C\|Dv\|(B).

Therefore we have ηvBV(n)CDuBV(Ω)\|\eta v\|_{BV(\mathbb{R}^{n})}\leq C\|Du\|_{BV(\Omega)}, where the constant CC depends on the constant coming from (PE2), on |Ω||\Omega| and on the constant coming from the Poincaré inequality. We then can assure that T:BV(Ω)BV(n):uηvT\colon BV(\Omega)\to BV(\mathbb{R}^{n})\colon u\mapsto\eta v is an extension operator.

Obviously we still have D(ηv)(Ω)=Dv(Ω)=0\|D(\eta v)\|(\partial\Omega)=\|Dv\|(\partial\Omega)=0. Hence Ω\Omega is indeed a strong BVBV-extension domain. ∎

Proof of the implication (2) \Longrightarrow (1).

We start with a strong BVBV-extension operator

T:BV(Ω)BV(n).T\colon BV(\Omega)\to BV(\mathbb{R}^{n}).

In particular, we know that

D(Tu)(Ω)=0\|D(Tu)\|(\partial\Omega)=0 (3.16)

for every uBV(Ω)u\in BV(\Omega).

Let S=Sn,(nΩ¯)S=S_{\mathbb{R}^{n},(\mathbb{R}^{n}\setminus\overline{\Omega})} be a Whitney smoothing operator given by Theorem 3.1. We assert that the operator R:W1,1(Ω)W1,1(n)R\colon W^{1,1}(\Omega)\to W^{1,1}(\mathbb{R}^{n}) defined by Ru(x)=(ST)(u)(x)Ru(x)=(S\circ T)(u)(x) is a W1,1W^{1,1}-extension operator.

Observe that Ru=uRu=u on Ω\Omega and

RuBV(n)CTuBV(n)CT|uBV(Ω)=CTuW1,1(Ω).\|Ru\|_{BV(\mathbb{R}^{n})}\leq C\|Tu\|_{BV(\mathbb{R}^{n})}\leq C\|T|\|\,\|u\|_{BV(\Omega)}=C\|T\|\,\|u\|_{W^{1,1}(\Omega)}.

To conclude we must check that indeed RuW1,1(n)Ru\in W^{1,1}(\mathbb{R}^{n}), so that in particular RuBV(n)=RuW1,1(n)\|Ru\|_{BV(\mathbb{R}^{n})}=\|Ru\|_{W^{1,1}(\mathbb{R}^{n})}. In order to get this let us show that the Radon measure D(Ru)\|D(Ru)\| consists only of its absolutely continuous part, and not of its singular part. Since we already now that Ru|ΩRu|_{\Omega} and Ru|nΩ¯Ru|_{\mathbb{R}^{n}\setminus\overline{\Omega}} are W1,1W^{1,1} functions we merely have to prove that D(Ru)(Ω)=0\|D(Ru)\|(\partial\Omega)=0. By the special properties of our smoothing operator given by (3.2) and by our assumption (3.16) we have that

D(Ru)(Ω)=D(STuTu+Tu)(Ω)D(STuTu)(Ω)+D(Tu)(Ω)=0\|D(Ru)\|(\partial\Omega)=\|D(STu-Tu+Tu)\|(\partial\Omega)\leq\|D(STu-Tu)\|(\partial\Omega)+\|D(Tu)\|(\partial\Omega)=0

and we are done. ∎

4. Further properties of W1,1W^{1,1}-domains

In this section we prove some corollaries to Theorem 1.3.

Corollary 4.1.

Let Ωn\Omega\subset\mathbb{R}^{n} be a bounded W1,1W^{1,1}-extension domain. Then the set of points xΩx\in\partial\Omega with D¯(Ω,x)>12\overline{D}(\Omega,x)>\frac{1}{2} is purely (n1)(n-1)-unrectifiable.

Proof.

If the set

F={xΩ:D¯(Ω,x)>12}F=\left\{x\in\partial\Omega\,:\,\overline{D}(\Omega,x)>\frac{1}{2}\right\}

is not purely (n1)(n-1)-unrectifiable, there exists a Lipschitz map f:n1f\colon\mathbb{R}^{n-1}\to\mathbb{R} so that, after a suitable rotation,

n1(Graph(f)F)>0.\mathcal{H}^{n-1}(\textrm{Graph}(f)\cap F)>0.

Notice that the set nGraph(f)\mathbb{R}^{n}\setminus\textrm{Graph}(f) consists of two connected components. Select one of the components that has nonempty intersection with Ω\Omega (actually, both have) and call EE its restriction to Ω\Omega. Then

MEΩ=Graph(f)Ω\partial^{M}E\cap\Omega=\textrm{Graph}(f)\cap\Omega

and so in particular EE has finite perimeter in Ω\Omega. Let E~n\widetilde{E}\subset\mathbb{R}^{n} be any set of finite perimeter with E~Ω=E\widetilde{E}\cap\Omega=E. Since

D(E,x)=12D(E,x)=\frac{1}{2}

at n1\mathcal{H}^{n-1}-almost every point xMEx\in\partial^{M}E, and D¯(Ω,x)>12\overline{D}(\Omega,x)>\frac{1}{2} for every xFx\in F, we have

n1(FME)=n1(FGraph(f)).\mathcal{H}^{n-1}(F\cap\partial^{M}E)=\mathcal{H}^{n-1}(F\cap\textrm{Graph}(f)).

Using again the fact that

D(E,x)=12D(E,x)=\frac{1}{2}

at n1\mathcal{H}^{n-1}-almost every point xMEx\in\partial^{M}E, and D¯(nΩ,x)<12\underline{D}(\mathbb{R}^{n}\setminus\Omega,x)<\frac{1}{2} for every xFx\in F, we have

0<D¯(E,x)D¯(E~,x)D¯(nΩ,x)+D¯(E,x)<10<\underline{D}(E,x)\leq\underline{D}(\widetilde{E},x)\leq\underline{D}(\mathbb{R}^{n}\setminus\Omega,x)+\underline{D}(E,x)<1

for n1\mathcal{H}^{n-1}-almost every point xFMEx\in F\cap\partial^{M}E. This means that there exists a set GFMEG\subset F\cap\partial^{M}E with n1(G)=0\mathcal{H}^{n-1}(G)=0 for which

(FME)GFME~.(F\cap\partial^{M}E)\setminus G\subset F\cap\partial^{M}\widetilde{E}.

Consequently,

n1(ΩME~)n1(FME~)n1(FME)=n1(FGraph(f))>0.\mathcal{H}^{n-1}(\partial\Omega\cap\partial^{M}\widetilde{E})\geq\mathcal{H}^{n-1}(F\cap\partial^{M}\widetilde{E})\geq\mathcal{H}^{n-1}(F\cap\partial^{M}E)=\mathcal{H}^{n-1}(F\cap\textrm{Graph}(f))>0.

Hence Ω\Omega does not have the strong extension property for sets of finite perimeter ((PE3) fails), and so by Theorem 1.3 it is not a W1,1W^{1,1}-extension domain. ∎

The next example shows that even in the plane the conclusion of Corollary 4.1 is not sufficient to imply that a bounded BVBV-extension domain is a W1,1W^{1,1}-extension domain.

Example 4.2.

Let us construct a planar BVBV-extension domain Ω\Omega so that the upper-density of Ω\Omega at all except at countably many boundary-points is at most 1/21/2, but the domain is not a W1,1W^{1,1}-extension domain. We set

Ω=(1,1)2(({0}×[1/2,1/2])i=2Ei),\Omega=(-1,1)^{2}\setminus\left((\{0\}\times[-1/2,1/2])\cup\bigcup_{i=2}^{\infty}E_{i}\right),

where, for every i2i\geq 2, we define

Ei=k=02i1\displaystyle E_{i}=\bigcup_{k=0}^{2^{i}-1} ([2i+1,2i2i10][2i+2i10,2i+1]\displaystyle\bigg{(}[-2^{-i+1},-2^{-i}-2^{-i-10}]\cup[2^{-i}+2^{-i-10},2^{-i+1}]
×[21+k2i,21+(k+1)2i2i10]).\displaystyle\times[-2^{-1}+k2^{-i},-2^{-1}+(k+1)2^{-i}-2^{-i-10}]\bigg{)}.

See Figure 1 for an illustration.

Refer to caption
Figure 1. An illustration of the BVBV-extension domain Ω2\Omega\subset\mathbb{R}^{2} in Example 4.2 which is not a W1,1W^{1,1}-extension domain. Components of the complement accumulate on the vertical line segment where the upper-density is less than 12\frac{1}{2} at almost every point.

Now, the upper-density of Ω\Omega is clearly at most 1/21/2 at all the points of the boundary Ω\partial\Omega except for the corners of the connected components of EiE_{i}, and the points (0,1/2)(0,-1/2) and (0,1/2)(0,1/2), which together form only a countable set. (One could remove balls instead of rectangles to get the upper-density bound for all boundary points.)

The domain Ω\Omega is a BVBV-extension domain because each removed square has a neighbourhood inside Ω\Omega from which the BVBV-function can be extended to the square with a uniform constant. These neighbourhoods can be taken pairwise disjoint. This will result in an extension operator

T:BV(Ω)BV((1,1)2{0}×[1/2,1/2]).T\colon BV(\Omega)\to BV((-1,1)^{2}\setminus\{0\}\times[-1/2,1/2]).

The target set clearly admits an extension to BV(2)BV(\mathbb{R}^{2}).

The domain Ω\Omega is not a W1,1W^{1,1}-extension domain, because the set {0}×[1/2,1/2]\{0\}\times[-1/2,1/2] is not purely 11-unrectifiable, and this is the set HH in the following Corollary 4.3.

The next corollary to Theorem 1.3 shows that one direction in Theorem 1.4 holds also in higher dimensions.

Corollary 4.3.

Suppose that Ωn\Omega\subset\mathbb{R}^{n} is a bounded W1,1W^{1,1}-extension domain. Let Ωi\Omega_{i}, for iIi\in I, be the connected components of nΩ¯\mathbb{R}^{n}\setminus\overline{\Omega}. Then the set

H=ΩiIΩi¯H=\partial\Omega\setminus\bigcup_{i\in I}\overline{\Omega_{i}}

is purely (n1)(n-1)-unrectifiable.

Proof.

Supposing Ω\Omega to be a W1,1W^{1,1}-extension domain, by Theorem 1.3 we know that it has the strong perimeter extension property.

Now, towards a contradiction, suppose that f:n1f\colon\mathbb{R}^{n-1}\to\mathbb{R} is an LL-Lipschitz map so that

n1(Graph(f)H)>0,\mathcal{H}^{n-1}(\textrm{Graph}(f)\cap H)>0,

after a suitable rotation. Let AA be a component of nGraph(f)\mathbb{R}^{n}\setminus\textrm{Graph}(f) such that the set

F={xGraph(f)H:D¯(ΩA,x)>0}F=\{x\in\textrm{Graph}(f)\cap H\,:\,\overline{D}(\Omega\cap A,x)>0\}

has positive n1\mathcal{H}^{n-1}-measure. By the measure density of Ω\Omega (Proposition 2.3), at least one of the components must satisfy this. Without loss of generality, we may assume that

A={(y,f(x)):y<f(x)}.A=\{(y,f(x))\,:\,y<f(x)\}.

Take E=ΩAE=\Omega\cap A and let E~\widetilde{E} be the strong perimeter extension of EE. Now, since n1(ME~Ω)=0\mathcal{H}^{n-1}(\partial^{M}\widetilde{E}\cap\partial\Omega)=0, the set

G={xF:D(E~,x)=1}G=\{x\in F\,:\,D(\widetilde{E},x)=1\}

has positive n1\mathcal{H}^{n-1}-measure. Take x=(x1,,xn)Gx=(x_{1},\dots,x_{n})\in G. Since EAE\subset A, which was bounded by a graph of an LL-Lipschitz map, the set

Rx,L={y=(y1,,yn)n:ynxn>L|(x1,,xn1)(y1,,yn1)|}R_{x,L}=\{y=(y_{1},\dots,y_{n})\in\mathbb{R}^{n}\,:\,y_{n}-x_{n}>L|(x_{1},\dots,x_{n-1})-(y_{1},\dots,y_{n-1})|\}

does not intersect EE. If there exists a small radius r>0r>0 for which

Rx,2LB(x,r)Ω=,R_{x,2L}\cap B(x,r)\cap\Omega=\emptyset,

we conclude that there exists a connected component Ωi\Omega_{i} of nΩ¯\mathbb{R}^{n}\setminus\overline{\Omega} for which xΩix\in\partial\Omega_{i} contradicting the fact that xHx\in H. Hence, there exists a sequence of points xiRx,2LΩx^{i}\in R_{x,2L}\cap\Omega such that |xix|0|x^{i}-x|\to 0. Since ff is LL-Lipschitz, writing δ=14(L+1)\delta=\frac{1}{4(L+1)}, we have

B(xi,δ|xix|)Rx,LnE.B\left(x^{i},\delta|x^{i}-x|\right)\subset R_{x,L}\subset\mathbb{R}^{n}\setminus E.

By the measure density (Proposition 2.3), we have

|B(xi,δ|xix|)Ω|>c|B(xi,δ|xix|)||B(x^{i},\delta|x^{i}-x|)\cap\Omega|>c|B(x^{i},\delta|x^{i}-x|)|

for all xix_{i}. Thus,

|B(x,2|xix|)E~||B(x,2|xix|)|1|B(xi,δ|xix|)Ω||B(x,2|xix|)|<1c(δ2)n<1\frac{|B(x,2|x^{i}-x|)\cap\widetilde{E}|}{|B(x,2|x^{i}-x|)|}\leq 1-\frac{|B(x^{i},\delta|x^{i}-x|)\cap\Omega|}{|B(x,2|x^{i}-x|)|}<1-c\left(\frac{\delta}{2}\right)^{n}<1

giving that

D¯(E~,x)<1,\underline{D}(\widetilde{E},x)<1,

which contradicts the fact that xGx\in G. ∎

Let us point out that if in addition we require Ω\Omega to be planar and simply connected in the previous corollary, we would get the stronger fact that 1(ΩiΩ¯i)=0\mathcal{H}^{1}(\partial\Omega\setminus\bigcup_{i}\overline{\Omega}_{i})=0.

In the next section we will show that in the planar case the conclusion of Corollary 4.3 is also a sufficient condition for a bounded BVBV-extension domain to be a W1,1W^{1,1}-extension domain. The following example shows that this is not the case in dimension three.

Example 4.4.

Let us construct a bounded BVBV-extension domain Ω3\Omega\subset\mathbb{R}^{3} which is not a W1,1W^{1,1}-extension domain so that 3Ω¯\mathbb{R}^{3}\setminus\overline{\Omega} consists of only one component Ω0\Omega_{0} for which Ω=Ω0\partial\Omega=\partial\Omega_{0}. Consequently, in the statement of Corollary 4.3 we have H=H=\emptyset.

Let C[0,1]2C\subset[0,1]^{2} be a Cantor set with 2(C)>0\mathcal{H}^{2}(C)>0 and let

Ω=(1,1)3{(x1,x2,x3):|x3|missingdist((x1,x2),C),(x1,x2)[0,1]2}.\Omega=(-1,1)^{3}\setminus\left\{(x_{1},x_{2},x_{3})\,:\,|x_{3}|\leq{\mathop{\mathrm{missing}}{\,dist\,}}((x_{1},x_{2}),C),(x_{1},x_{2})\in[0,1]^{2}\right\}.

The fact that 3Ω¯\mathbb{R}^{3}\setminus\overline{\Omega} consists of only one component Ω0\Omega_{0} for which Ω=Ω0\partial\Omega=\partial\Omega_{0} is immediate from the construction.

Also, with the same arguments as in the previous two corollaries, we see that

E={(x1,x2,x3)Ω:x3<0}E=\{(x_{1},x_{2},x_{3})\in\Omega\,:\,x_{3}<0\}

does not have a strong perimeter extension.

In order to see that Ω\Omega is a BVBV-extension domain, take uBV(Ω)u\in BV(\Omega). First notice that since the parts

Ω1={(x1,x2,x3)Ω:x3>0} and Ω2={(x1,x2,x3)Ω:x3<0}\Omega_{1}=\{(x_{1},x_{2},x_{3})\in\Omega\,:\,x_{3}>0\}\;\text{ and }\;\Omega_{2}=\{(x_{1},x_{2},x_{3})\in\Omega\,:\,x_{3}<0\}

have Lipschitz boundaries, similarly to [6, Theorem 5.8] we can consider the zero extension of both u|Ω1u|_{\Omega_{1}} and u|Ω2u|_{\Omega_{2}} to the whole 3\mathbb{R}^{3} and calling them u~1\tilde{u}_{1} and u~2\tilde{u}_{2} respectively, we have u~1,u~2BV(3)\tilde{u}_{1},\tilde{u}_{2}\in BV(\mathbb{R}^{3}) with

Du~i(3)=Du(Ωi)+Ωi|Tri(u)|𝑑2\|D\tilde{u}_{i}\|(\mathbb{R}^{3})=\|Du\|(\Omega_{i})+\int_{\partial\Omega_{i}}|\text{Tr}_{i}(u)|\,d\mathcal{H}^{2} (4.1)

for every i=1,2i=1,2. Here

Tri:BV(Ωi)L1(Ωi;2)\displaystyle\text{Tr}_{i}\colon BV(\Omega_{i})\to L^{1}\left(\partial\Omega_{i};\mathcal{H}^{2}\right)

for i=1,2i=1,2 are bounded linear operators, called the traces, which are defined as

Tri(u)(x)=limr01|B(x,r)Ωi|B(x,r)Ωiu(y)𝑑y\displaystyle\text{Tr}_{i}(u)(x)=\lim_{r\searrow 0}\frac{1}{|B(x,r)\cap\Omega_{i}|}\int_{B(x,r)\cap\Omega_{i}}u(y)\,dy

for 2\mathcal{H}^{2}-almost every xx. Now it is easy to check, following (4.1),

Du~i(3)\displaystyle\|D\tilde{u}_{i}\|(\mathbb{R}^{3}) Du(Ωi)+Ωi|Tri(u)|𝑑1\displaystyle\leq\|Du\|(\Omega_{i})+\int_{\partial\Omega_{i}}|\text{Tr}_{i}(u)|\,d\mathcal{H}^{1}
uBV(Ωi)+CuBV(Ωi)=(1+C)uBV(Ω),\displaystyle\leq\|u\|_{BV(\Omega_{i})}+C\|u\|_{BV(\Omega_{i})}=(1+C)\|u\|_{BV(\Omega)},

for i=1,2i=1,2. To conclude, we just let our extension operator T:BV(Ω)BV(3)T\colon BV(\Omega)\to BV(\mathbb{R}^{3}) be Tu=u~1+u~2Tu=\tilde{u}_{1}+\tilde{u}_{2}, which is the zero extension of uu outside Ω\Omega.

In the case where Ω¯=n\overline{\Omega}=\mathbb{R}^{n}, the study of extension domains Ω\Omega is the same as the study of closed removable sets. Notice that by the measure density (Proposition 2.3) the Lebesgue measure of Ω\partial\Omega is zero for a Sobolev or BVBV-extension domain. We call a set FnF\subset\mathbb{R}^{n} of Lebesgue measure zero a removable set for BVBV, if BV(nF)=BV(n)BV(\mathbb{R}^{n}\setminus F)=BV(\mathbb{R}^{n}) as sets and Du(n)=Du(nF)\|Du\|(\mathbb{R}^{n})=\|Du\|(\mathbb{R}^{n}\setminus F) for every uBV(n)u\in BV(\mathbb{R}^{n}). Similarly, we call FF removable for W1,1W^{1,1}, if W1,1(nF)=W1,1(n)W^{1,1}(\mathbb{R}^{n}\setminus F)=W^{1,1}(\mathbb{R}^{n}). We obtain the following equivalence of removability.

Corollary 4.5.

Let FnF\subset\mathbb{R}^{n} be a closed set of Lebesgue measure zero. Then FF is removable for BVBV if and only if FF is removable for W1,1W^{1,1}.

Proof.

Suppose FF is removable for BVBV. Then FF is purely (n1)(n-1)-unrectifiable. Otherwise, similarly as in the proof of Corollary 4.3, we can construct a set EE of finite perimeter so that n1(MEF)>0\mathcal{H}^{n-1}(\partial^{M}E\cap F)>0. Hence, P(E,nF)P(E,n)P(E,\mathbb{R}^{n}\setminus F)\neq P(E,\mathbb{R}^{n}), contradicting the assumption that FF is removable for BVBV. Now, since FF is removable for BVBV, for every radius R>0R>0, the set B(0,R)FB(0,R)\setminus F is a BVBV-extension domain. Since FF is purely (n1)(n-1)-unrectifiable, B(0,R)FB(0,R)\setminus F trivially has the strong perimeter extension property and is thus a W1,1W^{1,1}-extension domain by Theorem 1.3. Consequently, FF is removable for W1,1W^{1,1}.

Suppose then that FF is removable for W1,1W^{1,1}. Let uBV(nF)u\in BV(\mathbb{R}^{n}\setminus F). We only need to check that the function uu when seeing as a function defined on the whole n\mathbb{R}^{n}, satisfies Du(F)=0\|Du\|(F)=0. With the Whitney smoothing operator SnF,nFS_{\mathbb{R}^{n}\setminus F,\mathbb{R}^{n}\setminus F} from Theorem 3.1 we can modify uu to be a W1,1W^{1,1}-function u~=SnF,nFu\tilde{u}=S_{\mathbb{R}^{n}\setminus F,\mathbb{R}^{n}\setminus F}u on nF\mathbb{R}^{n}\setminus F and moreover, by (3.2),

D(u~u)(F)=0,\|D(\tilde{u}-u)\|(F)=0,

where u~\tilde{u} can be defined as any value on FF. Since FF is removable for W1,1W^{1,1}, we have u~W1,1(n)\tilde{u}\in W^{1,1}(\mathbb{R}^{n}). Thus Du~(F)=0\|D\tilde{u}\|(F)=0 because |F|=0|F|=0 and therefore

Du(F)D(uu~)(F)+Du~(F)=0,\|Du\|(F)\leq\|D(u-\tilde{u})\|(F)+\|D\tilde{u}\|(F)=0,

and we get that uBV(n)u\in BV(\mathbb{R}^{n}) with Du(n)=Du(nF)\|Du\|(\mathbb{R}^{n})=\|Du\|(\mathbb{R}^{n}\setminus F). ∎

5. Characterization of planar W1,1W^{1,1}-extension domains

In this section we prove Theorem 1.4 using the higher dimensional result stated in Theorem 1.3. Since the necessity part of Theorem 1.4 holds in the higher-dimensional case by Corollary 4.3, we only need to prove the sufficiency. We first set some notations and definitions.

We say that Γ2\Gamma\subset\mathbb{R}^{2} is a Jordan curve if Γ=γ([a,b])\Gamma=\gamma([a,b]) for some a,ba,b\in\mathbb{R}, a<ba<b, and some continuous map γ\gamma, injective on [a,b)[a,b) and such that γ(a)=γ(b)\gamma(a)=\gamma(b). Accordingly to the famous Jordan curve theorem any Jordan curve Γ\Gamma splits 2Γ\mathbb{R}^{2}\setminus\Gamma in exactly two connected components, a bounded one and an unbounded one that we call int(Γ)\text{int}(\Gamma) and ext(Γ)\text{ext}(\Gamma) respectively. We will often talk about rectifiable Jordan curves JJ, for which we mean that JJ is a Jordan curve and it is 11-rectifiable. A set AA whose boundary A\partial A is a Jordan curve is called a Jordan domain.

For technical reasons we also add to the class of Jordan curves the formal ”Jordan” curves J0J_{0} and JJ_{\infty}, whose interiors are 2\mathbb{R}^{2} and the empty set respectively and for which we set 1(J0)=1(J)=0\mathcal{H}^{1}(J_{0})=\mathcal{H}^{1}(J_{\infty})=0.

We say that a set E2E\in\mathbb{R}^{2} has a decomposition into other sets {Ei}i\{E_{i}\}_{i} up to 1\mathcal{H}^{1}-measure zero sets if

1((EiEi)(iEiE))=0\mathcal{H}^{1}\left(\left(E\setminus\bigcup_{i}E_{i}\right)\cup\left(\bigcup_{i}E_{i}\setminus E\right)\right)=0

and 1(EiEj)=0\mathcal{H}^{1}(E_{i}\cap E_{j})=0 for every iji\neq j.

For the particular case of planar sets of finite perimeter we have the following decomposition theorem from [1, Corollary 1].

Theorem 5.1.

Let E2E\subset\mathbb{R}^{2} have finite perimeter. Then, there exists a unique decomposition of ME\partial^{M}E into rectifiable Jordan curves {Ci+,Ck:i,k}\{C_{i}^{+},C_{k}^{-}\,:\,i,k\in\mathbb{N}\}, up to 1\mathcal{H}^{1}-measure zero sets, such that

  1. (1)

    Given int(Ci+)\text{int}(C_{i}^{+}), int(Ck+)\text{int}(C_{k}^{+}), iki\neq k, they are either disjoint or one is contained in the other; given int(Ci)\text{int}(C_{i}^{-}), int(Ck)\text{int}(C_{k}^{-}), iki\neq k, they are either disjoint or one is contained in the other. Each int(Ci)\text{int}(C_{i}^{-}) is contained in one of the int(Ck+)\text{int}(C_{k}^{+}).

  2. (2)

    P(E,2)=i1(Ci+)+k1(Ck)P(E,\mathbb{R}^{2})=\sum_{i}\mathcal{H}^{1}(C_{i}^{+})+\sum_{k}\mathcal{H}^{1}(C_{k}^{-}).

  3. (3)

    If int(Ci+)int(Cj+)\text{int}(C_{i}^{+})\subset\text{int}(C_{j}^{+}), iji\neq j, then there is some rectifiable Jordan curve CkC_{k}^{-} such that int(Ci+)int(Ck)int(Cj+)\text{int}(C_{i}^{+})\subset\text{int}(C_{k}^{-})\subset\text{int}(C_{j}^{+}). Similarly, if int(Ci)int(Cj)\text{int}(C_{i}^{-})\subset\text{int}(C_{j}^{-}), iji\neq j, then there is some rectifiable Jordan curve Ck+C_{k}^{+} such that int(Ci)int(Ck+)int(Cj)\text{int}(C_{i}^{-})\subset\text{int}(C_{k}^{+})\subset\text{int}(C_{j}^{-}).

  4. (4)

    Setting Lj={i:int(Ci)int(Cj+)}L_{j}=\{i\,:\,\text{int}(C_{i}^{-})\subset\text{int}(C_{j}^{+})\} the sets Yj=int(Cj+)iLjint(Ci)Y_{j}=\text{int}(C_{j}^{+})\setminus\bigcup_{i\in L_{j}}\text{int}(C_{i}^{-}) are pairwise disjoint, indecomposable and E=jYjE=\bigcup_{j}Y_{j}.

Since sets of finite perimeter are defined via the total variation of BVBV-functions, they are understood modulo 22-dimensional measure zero sets. In particular, the last equality in (4) of Theorem 5.1 is modulo measure zero sets.

In order to prove the sufficiency part of Theorem 1.4 we will proceed as follows: Starting from a set EΩE\subset\Omega of finite perimeter we first find an extension EE^{\prime} to 2\mathbb{R}^{2} using the fact that Ω\Omega is a BVBV-extension domain. Then we decompose ME\partial^{M}E^{\prime} using Theorem 5.1 and after proving the quasiconvexity of each of the open connected components Ωi\Omega_{i} of 2Ω¯\mathbb{R}^{2}\setminus\overline{\Omega}, we will be able to perturb the Jordan curves of the decomposition of ME\partial^{M}E^{\prime} around each Ωi\partial\Omega_{i} so that we get a final set E~\widetilde{E} which will be a strong extension of EE. An application of Theorem 1.3 will conclude the proof.

We start by presenting a couple of lemmas showing the quasiconvexity of all the connected components of 2Ω¯\mathbb{R}^{2}\setminus\overline{\Omega}.

Lemma 5.2.

Suppose that Ω2\Omega\subset\mathbb{R}^{2} is a bounded BVBV-extension domain. Then there exists a constant C>0C>0 so that for any connected component Ωi\Omega_{i} of 2Ω¯\mathbb{R}^{2}\setminus\overline{\Omega}, any two points z,wΩiz,w\in\partial\Omega_{i} can be connected by a curve βΩi¯\beta\subset\overline{\Omega_{i}} with (β)C|zw|\ell(\beta)\leq C|z-w|.

Proof.

One can essentially follow step by step the proof of [14, Theorem 1.1], once we have taken into account some facts.

  1. (1)

    For a given ii, since Ω\Omega is a BVBV-extension domain, so is Ω=2Ω¯i\Omega^{\prime}=\mathbb{R}^{2}\setminus\overline{\Omega}_{i}. As an extension operator we can take

    T:BV(Ω)BV(2):uT(u|Ω)|Ω¯i+u,T^{\prime}\colon BV(\Omega^{\prime})\to BV(\mathbb{R}^{2})\colon u\mapsto T(u|_{\Omega})|_{\overline{\Omega}_{i}}+u,

    where TT is the extension operator from BV(Ω)BV(\Omega) to BV(2)BV(\mathbb{R}^{2}). Let us explain more in detail why our resulting function TuT^{\prime}u is well-defined as a function in BV(2)BV(\mathbb{R}^{2}). Observe that the closures of the different components Ω¯i\overline{\Omega}_{i} can only intersect between themselves in just one point. That is,

    #{ΩiΩj}1for everyij.\#{\{\partial\Omega_{i}\cap\partial\Omega_{j}\}}\leq 1\;\;\;\text{for every}\;\;i\neq j. (5.1)

    Otherwise we would be losing the connectedness of Ω\Omega or either Ωi\Omega_{i} and Ωj\Omega_{j} are the same component. This means that ΩijiΩ¯j\partial\Omega_{i}\cap\bigcup_{j\neq i}\overline{\Omega}_{j} is a countable set. Once we are aware of this simple fact it is clear that T(u)T^{\prime}(u) behaves well around Ωi\partial\Omega_{i} and it belongs to BV(2)BV(\mathbb{R}^{2}).

    Observe that since Ω\Omega^{\prime} is a BVBV-extension domain there is a constant C>0C^{\prime}>0 for which the property (PE2)(PE2) of extension of sets of finite perimeter holds. Note that this constant CC^{\prime} only depends on the norm T\|T^{\prime}\|, which only depends on T\|T\| and which in turn only depends on the constant C>0C>0 of the same property (PE2)(PE2) but now applied to the BVBV-extension domain Ω\Omega.

  2. (2)

    We can assume that Ω\Omega^{\prime} is bounded, and hence also a BVlBV_{l}-extension domain thanks to [14, Lemma 2.1]. If Ω\Omega^{\prime} was not bounded then Ωi\Omega_{i} had to be bounded and we can take a large enough radius R>0R>0 so that

    ΩiB(0,R)andΩB(0,R)Ω¯i.\Omega_{i}\subset B(0,R)\;\;\text{and}\;\;\Omega\subset B(0,R)\setminus\overline{\Omega}_{i}.

    It is clear that changing Ω\Omega^{\prime} by ΩB(0,R)\Omega^{\prime}\cap B(0,R) does not affect the BVBV-extension property.

The proof of [14, Theorem 1.1] is made under the assumptions that a set Ω\Omega^{\prime} is a bounded simply connected BVlBV_{l}-extension domain, reaching as a conclusion that 2Ω\mathbb{R}^{2}\setminus\Omega^{\prime} is quasiconvex.

In the case Ωi\Omega_{i} was unbounded, Ω\Omega^{\prime} will be a bounded simply connected BVlBV_{l}-extension domain and we apply the previous result directly to show the quasiconvexity of Ω¯i\overline{\Omega}_{i}.

If Ωi\Omega_{i} was bounded, after the modification mentioned above, Ω\Omega^{\prime} will be a bounded BVlBV_{l}-extension domain. To prove the quasiconvexity of Ω¯i\overline{\Omega}_{i} in [14, Theorem 1.1] the simply connectedness was just used at the following point: when we take two points z,wΩiz,w\in\partial\Omega_{i} and join them with a line-segment Lz,wL_{z,w}, the set ΩLz,w\Omega^{\prime}\cap L_{z,w} consists on the disjoint union of countably many line-segments Lzi,wiL_{z_{i},w_{i}}, with zi,wiΩiz_{i},w_{i}\in\partial\Omega_{i}. Now, under the assumption of simply connectedness of Ω\Omega^{\prime} one can assert that ΩLzi,wi\Omega^{\prime}\setminus L_{z_{i},w_{i}} has two disjoint connected components. However, in our case this is still true because otherwise Ωi\Omega_{i} would not be connected.

The previous facts yield that every set Ω¯i\overline{\Omega}_{i} is quasiconvex. A careful reading of the proof [14, Theorem 1.1] also shows that the constant of quasiconvexity of all these sets is uniformly bounded by a constant C>0C>0, independent of ii. Indeed, the quasiconvexity constant of any set Ω¯i\overline{\Omega}_{i} only depends on the constant of the extension property of sets of finite perimeter (PE2)(PE2) for the BVBV-extension domains Ω=2Ω¯i\Omega^{\prime}=\mathbb{R}^{2}\setminus\overline{\Omega}_{i}, which, as we already noted, depends only on the constant for the BVBV-extension domain Ω\Omega independently of what ii we are fixing. ∎

Notice that the previous Lemma 5.2 implies, in particular, that if Ω\Omega is a bounded BVBV-extension domain, then all open connected components of 2Ω¯\mathbb{R}^{2}\setminus\overline{\Omega} are Jordan domains.

We record the following general lemma which might be of independent interest. A version of it for quasiconvex sets was proven via conformal maps in [16]. Let us also point out that with the sharp Painleve-length result for a connected set [19] one could quite easily prove a version of the lemma with a multiplicative constant 22.

Lemma 5.3.

Let Ω\Omega be a Jordan domain. For every x,yΩ¯x,y\in\overline{\Omega}, every ε>0\varepsilon>0 and any rectifiable curve γΩ¯\gamma\subset\overline{\Omega} joining xx to yy there exists a curve σΩ{x,y}\sigma\subset\Omega\cup\{x,y\} joining xx to yy so that my g

(σ)(γ)+ε.\ell(\sigma)\leq\ell(\gamma)+\varepsilon.
Proof.

Without loss of generality, we may assume that γ:[0,(γ)]2\gamma\colon[0,\ell(\gamma)]\to\mathbb{R}^{2} minimizes the length of curves joining xx to yy in Ω¯\overline{\Omega}, γ(0)=x\gamma(0)=x, γ((γ))=y\gamma(\ell(\gamma))=y, and that γ\gamma has unit speed.

If γ((0,(γ)))Ω=\gamma((0,\ell(\gamma)))\cap\partial\Omega=\emptyset, we are done. Suppose this is not the case and define

s1=min{t[0,(γ)]:γ(t)Ω}s_{1}=\min\{t\in[0,\ell(\gamma)]\,:\,\gamma(t)\in\partial\Omega\}

and

s2=max{t[0,(γ)]:γ(t)Ω}s_{2}=\max\{t\in[0,\ell(\gamma)]\,:\,\gamma(t)\in\partial\Omega\}

If s1=s2s_{1}=s_{2}, by minimality the curve γ\gamma is the concatenation of line-segments [x,γ(s1)][x,\gamma(s_{1})] and [γ(s1),y][\gamma(s_{1}),y]. In this case, for small r(0,ε/(2π))r\in(0,\varepsilon/(2\pi)), the curve γ\gamma divides B(γ(s1),r)B(\gamma(s_{1}),r) into two parts so that one of them is a subset of Ω\Omega. Thus, we may replace part of γ\gamma by an arc of the circle S(γ(s1),r)S(\gamma(s_{1}),r), and we are done.

We are then left with the more substantial case where s1<s2s_{1}<s_{2}. Since Ω\partial\Omega is a Jordan loop, the set Ωγ({s1,s2})\partial\Omega\setminus\gamma(\{s_{1},s_{2}\}) consists of two connected components T1T_{1} and T2T_{2}.

We will show that γ\gamma can be slightly pushed away from Ω\partial\Omega in directions that change in a locally Lipschitz way in (0,(γ))(0,\ell(\gamma)). Namely, we assert that there exist functions

ε:(0,(γ))(0,1),\displaystyle\varepsilon\colon(0,\ell(\gamma))\to(0,1),
v:(0,(γ))𝕊1,\displaystyle v\colon(0,\ell(\gamma))\to\mathbb{S}^{1},

so that ε()\varepsilon(\cdot) and v()v(\cdot) are locally Lipschitz continuous and satisfy γ(t)+hv(t)Ω\gamma(t)+hv(t)\in\Omega for all 0<h<ε(t)0<h<\varepsilon(t) and t(0,(γ))t\in(0,\ell(\gamma)).

In order to show this, let t(0,(γ))t\in(0,\ell(\gamma)). If γ(t)Ω\gamma(t)\in\Omega, then with εt=12missingdist(γ(t),Ω)\varepsilon_{t}=\frac{1}{2}{\mathop{\mathrm{missing}}{\,dist\,}}(\gamma(t),\partial\Omega) we have γ(s)+hvΩ\gamma(s)+hv\in\Omega for all s(tεt,t+εt)(0,(γ))s\in(t-\varepsilon_{t},t+\varepsilon_{t})\cap(0,\ell(\gamma)), v𝕊1v\in\mathbb{S}^{1} and 0<h<εt0<h<\varepsilon_{t}.

Suppose then that tγ1(Ω)(0,(γ))t\in\gamma^{-1}(\partial\Omega)\cap(0,\ell(\gamma)). Without loss of generality we may assume that γ(t)T1\gamma(t)\in T_{1}. The concatenation of γ|[s1,s2]\gamma|_{[s_{1},s_{2}]} with T2T_{2} forms a closed loop α\alpha so that one of the components Ω~\tilde{\Omega} of its complement is contained in Ω\Omega, and γ(t)Ω~\gamma(t)\in\partial\tilde{\Omega}. Now, let rt=missingdist(γ(t),T2)r_{t}={\mathop{\mathrm{missing}}{\,dist\,}}(\gamma(t),T_{2}). Then, by minimality of γ\gamma, the set γB(γ(t),rt)\gamma\cap B(\gamma(t),r_{t}) is contained on the boundary of a convex set Kt=B(γ(t),rt)Ω~K_{t}=B(\gamma(t),r_{t})\setminus\tilde{\Omega} with non-empty interior. Consequently, there exists a constant εt>0\varepsilon_{t}>0 so that for any tεt<τ1<τ2<t+εtt-\varepsilon_{t}<\tau_{1}<\tau_{2}<t+\varepsilon_{t} for which the outer normal vectors w1w_{1} and w2w_{2} to KtK_{t} exist at γ(τ1)\gamma(\tau_{1}) and γ(τ2)\gamma(\tau_{2}) respectively, there is a Lipschitz map [τ1,τ2]𝕊1:tvτ1,τ2(t)[\tau_{1},\tau_{2}]\to\mathbb{S}^{1}\colon t\mapsto v_{\tau_{1},\tau_{2}}(t) so that vτ1,τ2(τ1)=w1v_{\tau_{1},\tau_{2}}(\tau_{1})=w_{1}, vτ1,τ2(τ2)=w2v_{\tau_{1},\tau_{2}}(\tau_{2})=w_{2}, and γ(t)+hvτ1,τ2(t)Ω\gamma(t)+hv_{\tau_{1},\tau_{2}}(t)\in\Omega for all 0<h<εt0<h<\varepsilon_{t} and τ1tτ2\tau_{1}\leq t\leq\tau_{2}.

Write I(0,(γ))I\subset(0,\ell(\gamma)) to be the points tt where a normal direction to γ\gamma exists at γ(t)\gamma(t). Now, cover (0,(γ))(0,\ell(\gamma)) with the intervals (tεt,t+εt)(0,(γ))(t-\varepsilon_{t},t+\varepsilon_{t})\cap(0,\ell(\gamma)) and then take a subcover {Ui=(t¯iεt¯i,t¯i+εt¯i)}i\{U_{i}=(\overline{t}_{i}-\varepsilon_{\overline{t}_{i}},\overline{t}_{i}+\varepsilon_{\overline{t}_{i}})\}_{i\in\mathbb{Z}} that is finite for compact subsets of (0,(γ))(0,\ell(\gamma)), and so that every t(0,(γ))t\in(0,\ell(\gamma)) belongs to at most two intervals UiU_{i}. Assume the intervals UiU_{i} are in order, that is UiU_{i} only intersects Ui1U_{i-1} and Ui+1U_{i+1}. By dividing into smaller intervals if needed, we may also assume that if γ(Ui)Ω\gamma(U_{i})\cap\partial\Omega\neq\emptyset and γ(Ui+1)Ω\gamma(U_{i+1})\cap\partial\Omega\neq\emptyset, then γ(UiUi+1)ΩTj\gamma(U_{i}\cup U_{i+1})\cap\partial\Omega\subset T_{j} for j=1j=1 or j=2j=2. This allows us to select the normal directions w𝕊1w\in\mathbb{S}^{1} in a way so that they agree for the intervals UiU_{i} and Ui+1U_{i+1} at the points tIUiUi+1t\in I\cap U_{i}\cap U_{i+1}. Notice that for ii for which γ(Ui)Ω=\gamma(U_{i})\cap\partial\Omega=\emptyset we have to make a choice between two opposite directions.

We will then have subset I(0,(γ))I\subset(0,\ell(\gamma)) with 1((0,(γ))I)=0\mathcal{H}^{1}((0,\ell(\gamma))\setminus I)=0, and an open covering {Ui}i\{U_{i}\}_{i} of (0,(γ))(0,\ell(\gamma)) of multiplicity at most two, where UiU_{i} are intervals, so that

  • for every ii\in\mathbb{Z} there exists a constant εi>0\varepsilon_{i}>0 so that for every τ1,τ2IUi\tau_{1},\tau_{2}\in I\cap U_{i}, τ1<τ2\tau_{1}<\tau_{2}, there is a Lipschitz map [τ1,τ2]𝕊1:tvτ1,τ2(t)[\tau_{1},\tau_{2}]\to\mathbb{S}^{1}\colon t\mapsto v_{\tau_{1},\tau_{2}}(t) so that γ(t)+hvτ1,τ2(t)Ω\gamma(t)+hv_{\tau_{1},\tau_{2}}(t)\in\Omega for all 0<h<εi0<h<\varepsilon_{i} and τ1tτ2\tau_{1}\leq t\leq\tau_{2},

  • if vτ1,τ2v_{\tau_{1},\tau_{2}} and vτ2,τ3v_{\tau_{2},\tau_{3}} have been defined as above, vτ1,τ2(τ2)=vτ2,τ3(τ2)v_{\tau_{1},\tau_{2}}(\tau_{2})=v_{\tau_{2},\tau_{3}}(\tau_{2}).

For each ii\in\mathbb{Z} we will now fix a tiIUiUi+1t_{i}\in I\cap U_{i}\cap U_{i+1}, and define v(t)=vti,ti+1(t)v(t)=v_{t_{i},t_{i+1}}(t) on [ti,ti+1][t_{i},t_{i+1}]. A locally Lipschitz choice for ε\varepsilon can be given by defining

ε(t)=ti+1tti+1timin(εi,εi+1)+ttiti+1timin(εi+1,εi+2)\varepsilon(t)=\frac{t_{i+1}-t}{t_{i+1}-t_{i}}\min(\varepsilon_{i},\varepsilon_{i+1})+\frac{t-t_{i}}{t_{i+1}-t_{i}}\min(\varepsilon_{i+1},\varepsilon_{i+2})

when t[ti,ti+1]t\in[t_{i},t_{i+1}].

Let i0i_{0}\in\mathbb{N} be such that i0>2s2s12(γ)i_{0}>\frac{2}{s_{2}-s_{1}}\geq\frac{2}{\ell(\gamma)}. Then, for any ii0i\geq i_{0}, the function v()v(\cdot) is Lipschitz in [1/i,(γ)1/i][1/i,\ell(\gamma)-1/i] and ηi:=mint[1/i,(γ)1/i]ε(t)>0\eta_{i}:=\min_{t\in[1/i,\ell(\gamma)-1/i]}\varepsilon(t)>0. Hence, if we define

δi(t)={ηimin{|(γ)1/it|,|1/it|}(γ)2/i,if t[1/i,(γ)1/i],0,otherwise,\delta_{i}(t)=\begin{cases}\eta_{i}\frac{\min\{|\ell(\gamma)-1/i-t|,|1/i-t|\}}{\ell(\gamma)-2/i},&\text{if }t\in[1/i,\ell(\gamma)-1/i],\\ 0,&\text{otherwise}\end{cases},

we have |δi(t)|=ηi(γ)2/i|\delta^{\prime}_{i}(t)|=\frac{\eta_{i}}{\ell(\gamma)-2/i} for all t(1/i,(γ)1/i){(γ)/2}t\in(1/i,\ell(\gamma)-1/i)\setminus\{\ell(\gamma)/2\}, and also |δi(t)|ηi/2ε(t)/2|\delta_{i}(t)|\leq\eta_{i}/2\leq\varepsilon(t)/2 for every tt. Now if we let

Li=0(γ)|(δi(t)v(t))|𝑑t<,L_{i}=\int_{0}^{\ell(\gamma)}|(\delta_{i}(t)v(t))^{\prime}|\,dt<\infty,

defining

δ(t)=i=i02i1min(ε,1)1+Liδi(t),\delta(t)=\sum_{i=i_{0}}^{\infty}\frac{2^{-i-1}\min(\varepsilon,1)}{1+L_{i}}\delta_{i}(t),

we get a function δ:[0,(γ)]\delta\colon[0,\ell(\gamma)]\to\mathbb{R} such that δ(0)=δ((γ))=0\delta(0)=\delta(\ell(\gamma))=0 and 0<δ(t)<ε(t)0<\delta(t)<\varepsilon(t) for all t(0,(γ))t\in(0,\ell(\gamma)). Note that the function δ\delta is continuous as a limit of an absolutely and uniformly convergent series of continuous functions, and it is differentiable except on (γ)/2\ell(\gamma)/2. Thus, σ:[0,(γ)]2\sigma\colon[0,\ell(\gamma)]\to\mathbb{R}^{2} defined by

σ(t)=γ(t)+δ(t)v(t)\sigma(t)=\gamma(t)+\delta(t)v(t)

is a curve joining xx and yy, σ((0,(γ)))Ω\sigma((0,\ell(\gamma)))\subset\Omega and

(σ)\displaystyle\ell(\sigma) =0(γ)|σ(t)|𝑑t0(γ)|γ(t)|𝑑t+0(γ)|(δ(t)v(t))|𝑑t\displaystyle=\int_{0}^{\ell(\gamma)}|\sigma^{\prime}(t)|\,dt\leq\int_{0}^{\ell(\gamma)}|\gamma^{\prime}(t)|\,dt+\int_{0}^{\ell(\gamma)}|(\delta(t)v(t))^{\prime}|\,dt
(γ)+i=i02i1ε1+Li0(γ)|(δi(t)v(t)7)|𝑑t<(γ)+ε,\displaystyle\leq\ell(\gamma)+\sum_{i=i_{0}}^{\infty}\frac{2^{-i-1}\varepsilon}{1+L_{i}}\int_{0}^{\ell(\gamma)}|(\delta_{i}(t)v(t)7)^{\prime}|\,dt<\ell(\gamma)+\varepsilon,

finishing the proof. ∎

The next lemma, together with Theorem 5.1, are the key tools for our proof of the sufficiency part of Theorem 1.4, that we will show afterwards.

Lemma 5.4.

Let Ω2\Omega\subset\mathbb{R}^{2} be a bounded BVBV-extension domain and Ωi\Omega_{i} the open connected components of 2Ω¯\mathbb{R}^{2}\setminus\overline{\Omega}. Suppose that the set H=ΩiΩ¯iH=\partial\Omega\setminus\bigcup_{i}\overline{\Omega}_{i} is purely 11-unrectifiable and let E2E\subset\mathbb{R}^{2} be a Jordan domain with E\partial E rectifiable. Then there exists a set E~2\widetilde{E}\subset\mathbb{R}^{2} of finite perimeter so that

  • (i)

    EΩ=E~ΩE\cap\Omega=\widetilde{E}\cap\Omega,

  • (ii)

    1(ME~)C1(ME)\mathcal{H}^{1}(\partial^{M}\widetilde{E})\leq C\mathcal{H}^{1}(\partial^{M}E), and

  • (iii)

    1(ME~Ω)=0\mathcal{H}^{1}(\partial^{M}\widetilde{E}\cap\partial\Omega)=0,

where the constant CC is absolute.

Proof.

Consider the at most countably many components {Ωi}i\{\Omega_{i}\}_{i} of 2Ω¯\mathbb{R}^{2}\setminus\overline{\Omega}. For each ii we want to modify the set EE in Ωi¯\overline{\Omega_{i}} to get some E~2\widetilde{E}\subset\mathbb{R}^{2} with E~\partial\widetilde{E} rectifiable so that

1(ME~Ωi)=0\mathcal{H}^{1}(\partial^{M}\widetilde{E}\cap\partial\Omega_{i})=0 (5.2)

and

1(ME~Ωi)C1(MEΩi¯).\mathcal{H}^{1}(\partial^{M}\widetilde{E}\cap\Omega_{i})\leq C\mathcal{H}^{1}(\partial^{M}E\cap\overline{\Omega_{i}}). (5.3)

Let us show how to conclude the proof of the lemma after assuming these facts. Since we are not changing the set EE inside Ω\Omega the property (i) is clear. To check (ii) let us first write

1(ME~)=1(ME~Ω)+1(ME~Ω)+1(ME~(nΩ¯)).\mathcal{H}^{1}(\partial^{M}\widetilde{E})=\mathcal{H}^{1}(\partial^{M}\widetilde{E}\cap\Omega)+\mathcal{H}^{1}(\partial^{M}\widetilde{E}\cap\partial\Omega)+\mathcal{H}^{1}(\partial^{M}\widetilde{E}\cap(\mathbb{R}^{n}\setminus\overline{\Omega})).

We will estimate each of these terms separately. For the first one is clear that 1(ME~Ω)=1(MEΩ)\mathcal{H}^{1}(\partial^{M}\widetilde{E}\cap\Omega)=\mathcal{H}^{1}(\partial^{M}E\cap\Omega). For the second one we use the fact that E~\partial\widetilde{E} is rectifiable, that ΩiΩ¯i\partial\Omega\setminus\bigcup_{i}\overline{\Omega}_{i} is purely 11-unrectifiable and (5.2),

1(ME~Ω)\displaystyle\mathcal{H}^{1}(\partial^{M}\widetilde{E}\cap\partial\Omega) =1(ME~[ΩiΩ¯i])+1(ME~[ΩiΩ¯i])\displaystyle=\mathcal{H}^{1}\left(\partial^{M}\widetilde{E}\cap\left[\partial\Omega\setminus\bigcup_{i}\overline{\Omega}_{i}\right]\right)+\mathcal{H}^{1}\left(\partial^{M}\widetilde{E}\cap\left[\partial\Omega\cap\bigcup_{i}\overline{\Omega}_{i}\right]\right)
=1(i(ME~ΩΩ¯i))\displaystyle=\mathcal{H}^{1}\left(\bigcup_{i}(\partial^{M}\widetilde{E}\cap\partial\Omega\cap\overline{\Omega}_{i})\right)
i1(ME~Ωi)=0.\displaystyle\leq\sum_{i}\mathcal{H}^{1}(\partial^{M}\widetilde{E}\cap\partial\Omega_{i})=0. (5.4)

For the third term we use (5.3) to get

1(ME~(nΩ¯))i1(ME~Ωi)Ci1(MEΩi¯).\mathcal{H}^{1}(\partial^{M}\widetilde{E}\cap(\mathbb{R}^{n}\setminus\overline{\Omega}))\leq\sum_{i}\mathcal{H}^{1}(\partial^{M}\widetilde{E}\cap\Omega_{i})\leq C\sum_{i}\mathcal{H}^{1}(\partial^{M}E\cap\overline{\Omega_{i}}).

All these estimates together yield

1(ME~)1(MEΩ)+Ci1(MEΩi¯).\mathcal{H}^{1}(\partial^{M}\widetilde{E})\leq\mathcal{H}^{1}(\partial^{M}E\cap\Omega)+C\sum_{i}\mathcal{H}^{1}(\partial^{M}E\cap\overline{\Omega_{i}}).

Since {xΩi:xΩjfor someji}\{x\in\partial\Omega_{i}:\,x\in\partial\Omega_{j}\;\;\text{for some}\;\;j\neq i\} is at most countable by (5.1), we conclude that

1(ME~)C1(ME),\mathcal{H}^{1}(\partial^{M}\widetilde{E})\leq C\mathcal{H}^{1}(\partial^{M}E),

proving (ii). Finally (iii) has already been shown in (5).

We now move to prove how to modify EE inside each set Ω¯i\overline{\Omega}_{i} in order to get (5.2) and (5.3).

If 1(MEΩi)=0\mathcal{H}^{1}(\partial^{M}E\cap\partial\Omega_{i})=0, we may skip this ii and move to the next. Let us thus assume 1(MEΩi)>0\mathcal{H}^{1}(\partial^{M}E\cap\partial\Omega_{i})>0. Let f:𝕊1Ef\colon\mathbb{S}^{1}\to\partial E be a parameterisation of the boundary by a homeomorphism. By the Lebesgue density theorem, for almost every tf1(MEΩi)t\in f^{-1}(\partial^{M}E\cap\partial\Omega_{i}) there exists a rt>0r_{t}>0 so that for all 0<r<rt0<r<r_{t}

1(f(B(t,r))Ωi)121(f(B(t,r))).\mathcal{H}^{1}\left(f(B(t,r))\cap\partial\Omega_{i}\right)\geq\frac{1}{2}\mathcal{H}^{1}(f(B(t,r))). (5.5)

By the Vitali covering lemma, we then find a disjointed collection {B(tj,rj)}j\{B(t_{j},r_{j})\}_{j} so that (5.5) holds for each of the balls and

1((MEΩi)jf(B(tj,rj))=0.\mathcal{H}^{1}\left((\partial^{M}E\cap\partial\Omega_{i})\setminus\bigcup_{j}f(B(t_{j},r_{j})\right)=0.

Now, we define Ii,j=B(tj,rj)¯𝕊1I_{i,j}=\overline{B(t_{j},r_{j})}\cap\mathbb{S}^{1} for each jj and obtain a collection {Ii,j}j\{I_{i,j}\}_{j} of closed arcs in 𝕊1\mathbb{S}^{1} whose interiors are pairwise disjoint,

1((MEΩi)jf(Ii,j))=0\mathcal{H}^{1}\left((\partial^{M}E\cap\partial\Omega_{i})\setminus\bigcup_{j}f(I_{i,j})\right)=0

and

1(f(Ii,j)Ωi)121(f(Ii,j))\mathcal{H}^{1}\left(f(I_{i,j})\cap\partial\Omega_{i}\right)\geq\frac{1}{2}\mathcal{H}^{1}(f(I_{i,j}))

for every jj.

Refer to caption
Figure 2. An illustration of the construction in Lemma 5.4. The boundary E\partial E intersect the boundaries Ω1\partial\Omega_{1} and Ω2\partial\Omega_{2} in a set of positive 1\mathcal{H}^{1}-measure. The modification of EE inside Ω1\Omega_{1} consists of the added set bounded by γ1,1\gamma_{1,1} from which three sets have been removed, bounded by γ1,1,1\gamma_{1,1,1}, γ1,1,3\gamma_{1,1,3}, and γ1,1,4\gamma_{1,1,4}, respectively. The modification inside Ω2\Omega_{2} consists of only one added part bounded by γ2,1\gamma_{2,1}.

For the next argument we have i,ji,j fixed. The set f(Ii,j)Ωif(I_{i,j})\setminus\partial\Omega_{i} consists of at most countably many open curves {αi,j,k}k\{\alpha_{i,j,k}\}_{k}. For each kk for which αi,j,kΩi=\alpha_{i,j,k}\cap\Omega_{i}=\emptyset, we use Lemma 5.2 to find a curve βi,j,kΩi¯\beta_{i,j,k}\subset\overline{\Omega_{i}} such that (βi,j,k)C|zi,j,kwi,j,k|\ell(\beta_{i,j,k})\leq C|z_{i,j,k}-w_{i,j,k}| where zi,j,kz_{i,j,k} and wi,j,kw_{i,j,k} are the endpoints of αi,j,k\alpha_{i,j,k}. Now, for ε=|zi,j,kwi,j,k|\varepsilon=|z_{i,j,k}-w_{i,j,k}|, Lemma 5.3 provides us with another curve γi,j,kΩi{zi,j,k,wi,j,k}\gamma_{i,j,k}\subset\Omega_{i}\cup\{z_{i,j,k},w_{i,j,k}\} so that

(γi,j,k)(βi,j,k)+|zi,j,kwi,j,k|(C+1)|zi,j,kwi,j,k|.\ell(\gamma_{i,j,k})\leq\ell(\beta_{i,j,k})+|z_{i,j,k}-w_{i,j,k}|\leq(C+1)|z_{i,j,k}-w_{i,j,k}|. (5.6)

The curves αi,j,k\alpha_{i,j,k} and γi,j,k\gamma_{i,j,k} enclose a bounded subset that we call Ei,j,k2E_{i,j,k}\subset\mathbb{R}^{2}. Similarly, if we let zi,jz_{i,j} be the first, and wi,jw_{i,j} the last point of f(Ii,j)Ωif(I_{i,j})\cap\partial\Omega_{i} we again use Lemmas 5.2 and 5.3 to connect zi,jz_{i,j} to wi,jw_{i,j} with a curve γi,jΩi{zi,j,wi,j}\gamma_{i,j}\subset\Omega_{i}\cup\{z_{i,j},w_{i,j}\} so that

(γi,j)(C+1)|zi,jwi,j|.\ell(\gamma_{i,j})\leq(C+1)|z_{i,j}-w_{i,j}|. (5.7)

Let Fi,jF_{i,j} be the bounded set enclosed by Ωi\partial\Omega_{i} (from zi,jz_{i,j} to wi,jw_{i,j}) and by γi,j\gamma_{i,j}. Now, we will modify EE by considering

E~i,j=E(Fi,jkEi,j,k).\widetilde{E}_{i,j}=E\cup\left(F_{i,j}\setminus\bigcup_{k}E_{i,j,k}\right).

See Figure 2 for an illustration of the modification.

Repeating this process for all ii with 1(MEΩi)>0\mathcal{H}^{1}(\partial^{M}E\cap\partial\Omega_{i})>0 and all jj we can finally define

E~=i,jE~i,j.\widetilde{E}=\bigcup_{i,j}\widetilde{E}_{i,j}.

Let us check that the properties (5.2)\eqref{eq:boundarytozero} and (5.3)\eqref{eq:boundary.int.controled} hold. Firstly, observing that we did not modified E\partial E outside the arcs f(Ii,j)f(I_{i,j}),

1(ME~Ωi)\displaystyle\mathcal{H}^{1}(\partial^{M}\widetilde{E}\cap\partial\Omega_{i}) =1((ME~Ωi)jf(Ii,j))\displaystyle=\mathcal{H}^{1}\left((\partial^{M}\widetilde{E}\cap\partial\Omega_{i})\setminus\bigcup_{j}f(I_{i,j})\right)
+j1(ME~Ωif(Ii,j))\displaystyle\qquad+\sum_{j}\mathcal{H}^{1}(\partial^{M}\widetilde{E}\cap\partial\Omega_{i}\cap f(I_{i,j}))
=1((MEΩi)jf(Ii,j))\displaystyle=\mathcal{H}^{1}\left((\partial^{M}E\cap\partial\Omega_{i})\setminus\bigcup_{j}f(I_{i,j})\right)
+j(1(MFi,jΩi)+k1(MEi,j,kΩi))\displaystyle\qquad+\sum_{j}\left(\mathcal{H}^{1}(\partial^{M}F_{i,j}\cap\partial\Omega_{i})+\sum_{k}\mathcal{H}^{1}(\partial^{M}E_{i,j,k}\cap\partial\Omega_{i})\right)
=0,\displaystyle=0,

which gives us (5.2). Secondly,

1(ME~Ωi)\displaystyle\mathcal{H}^{1}(\partial^{M}\widetilde{E}\cap\Omega_{i}) 1(MEΩi)+j(1(γi,j)+k1(γi,j,k))\displaystyle\leq\mathcal{H}^{1}(\partial^{M}E\cap\Omega_{i})+\sum_{j}\left(\mathcal{H}^{1}(\gamma_{i,j})+\sum_{k}\mathcal{H}^{1}(\gamma_{i,j,k})\right)
1(MEΩi)+j((C+1)|zi,jwi,j|+k(C+1)|zi,j,kwi,j,k|)\displaystyle\leq\mathcal{H}^{1}(\partial^{M}E\cap\Omega_{i})+\sum_{j}\left((C+1)|z_{i,j}-w_{i,j}|+\sum_{k}(C+1)|z_{i,j,k}-w_{i,j,k}|\right)
1(MEΩi)+j2C1(f(Ii,j))\displaystyle\leq\mathcal{H}^{1}(\partial^{M}E\cap\Omega_{i})+\sum_{j}2C\mathcal{H}^{1}(f(I_{i,j}))
1(MEΩi)+j2(C+1)1(f(Ii,j)Ωi)\displaystyle\leq\mathcal{H}^{1}(\partial^{M}E\cap\Omega_{i})+\sum_{j}2(C+1)\mathcal{H}^{1}(f(I_{i,j})\cap\partial\Omega_{i})
C1(MEΩ¯i)\displaystyle\leq C\mathcal{H}^{1}(\partial^{M}E\cap\overline{\Omega}_{i})

proving (5.3). ∎

Proof of Theorem 1.4.

One direction is proven in Corollary 4.3. Thus we only need to prove the converse. Thus, assume that Ω2\Omega\subset\mathbb{R}^{2} is a bounded BVBV-extension domain and that the set H=ΩiΩ¯iH=\partial\Omega\setminus\bigcup_{i}\overline{\Omega}_{i} is purely 11-unrectifiable, where Ωi\Omega_{i} are the open connected components of 2Ω¯\mathbb{R}^{2}\setminus\overline{\Omega}.

We will show that Ω\Omega has the strong extension property for sets of finite perimeter and hence, by Theorem 1.3, Ω\Omega will be a W1,1W^{1,1}-extension domain. Using the fact that Ω\Omega is a bounded BVBV-extension domain if we let EΩE\subset\Omega be a set of finite perimeter in Ω\Omega then there exists an extension EE^{\prime} to 2\mathbb{R}^{2} so that P(E,2)CP(E,Ω)P(E^{\prime},\mathbb{R}^{2})\leq CP(E,\Omega). This extension can be obtained for instance by the Maz’ya and Burago result [20, Section 9.3].

Let now {Ci+,Ck:i,k}\{C_{i}^{+},C_{k}^{-}\,:\,i,k\in\mathbb{N}\} be the rectifiable Jordan curves of Theorem 5.1 for the set EE^{\prime}. By applying Lemma 5.4, each Jordan domain int(Ci+)\text{int}(C_{i}^{+}) can be replaced by a set E~i+\widetilde{E}_{i}^{+} so that E~i+Ω=int(Ci+)Ω\widetilde{E}_{i}^{+}\cap\Omega=\text{int}(C_{i}^{+})\cap\Omega, 1(ME~i+)C1(Ci+)\mathcal{H}^{1}(\partial^{M}\widetilde{E}_{i}^{+})\leq C\mathcal{H}^{1}(C_{i}^{+}), and 1(ME~i+Ω)=0\mathcal{H}^{1}(\partial^{M}\widetilde{E}_{i}^{+}\cap\partial\Omega)=0. Similarly, each int(Ck)\text{int}(C_{k}^{-}) can be replaced by a set E~k\widetilde{E}_{k}^{-} so that E~kΩ=int(Ck)Ω\widetilde{E}_{k}^{-}\cap\Omega=\text{int}(C_{k}^{-})\cap\Omega, 1(ME~k)C1(Ck)\mathcal{H}^{1}(\partial^{M}\widetilde{E}_{k}^{-})\leq C\mathcal{H}^{1}(C_{k}^{-}), and 1(ME~kΩ)=0\mathcal{H}^{1}(\partial^{M}\widetilde{E}_{k}^{-}\cap\partial\Omega)=0.

Now,

E=EΩ=(iint(Ci+)kint(Ck))Ω=(iE~i+kE~k)Ω,E=E^{\prime}\cap\Omega=\left(\bigcup_{i}\text{int}(C_{i}^{+})\setminus\bigcup_{k}\text{int}(C_{k}^{-})\right)\cap\Omega=\left(\bigcup_{i}\widetilde{E}_{i}^{+}\setminus\bigcup_{k}\widetilde{E}_{k}^{-}\right)\cap\Omega,

holds modulo a measure zero set. Thus, the set

E~=(iE~i+kE~k)\widetilde{E}=\left(\bigcup_{i}\widetilde{E}_{i}^{+}\setminus\bigcup_{k}\widetilde{E}_{k}^{-}\right)

is an extension of EE to 2\mathbb{R}^{2}, and

P(E~,2)\displaystyle P(\widetilde{E},\mathbb{R}^{2}) i1(ME~i+)+k1(ME~k)\displaystyle\leq\sum_{i}\mathcal{H}^{1}(\partial^{M}\widetilde{E}_{i}^{+})+\sum_{k}\mathcal{H}^{1}(\partial^{M}\widetilde{E}_{k}^{-})
iC1(Ci+)+kC1(Ck)\displaystyle\leq\sum_{i}C\mathcal{H}^{1}(C_{i}^{+})+\sum_{k}C\mathcal{H}^{1}(C_{k}^{-})
=CP(E,2)CP(E,Ω).\displaystyle=CP(E^{\prime},\mathbb{R}^{2})\leq CP(E,\Omega).

Since,

1(ME~Ω)i1(ME~i+Ω)+k1(ME~kΩ)=0,\mathcal{H}^{1}(\partial^{M}\widetilde{E}\cap\partial\Omega)\leq\sum_{i}\mathcal{H}^{1}(\partial^{M}\widetilde{E}_{i}^{+}\cap\partial\Omega)+\sum_{k}\mathcal{H}^{1}(\partial^{M}\widetilde{E}_{k}^{-}\cap\partial\Omega)=0,

the set E~\widetilde{E} is the strong extension of EE that we had to find. ∎

Acknowledgements

The authors thank Panu Lahti for several comments on an earlier version of this paper.

References

  • [1] L. Ambrosio, V. Caselles, S. Masnou and J.-M. Morel, Connected components of sets of finite perimeter and applications to image processing. J. Eur. Math. Soc. 3, 39–92 (2001).
  • [2] L. Ambrosio, N. Fusco and D. Pallara, Functions of bounded variation and free discontinuity problems. Oxford Mathematical Monographs. The Clarendon Press, Oxford University Press, New York, 2000.
  • [3] B. Bojarski, P. Hajłasz, P. Strzelecki, Improved Ck,λC^{k,\lambda} approximation of higher order Sobolev functions in norm and capacity. Indiana Univ. Mat. J. 51 (2002), 507–540.
  • [4] Yu. D. Burago and V. G. Maz’ya, Certain questions of potential theory and function theory for regions with irregular boundaries (Russian), Zap. Nauc̆n. Sem. Leningrad. Otdel. Mat. Inst. Steklov (LOMI) 3 (1967) 152pp.
  • [5] A. P. Calderón, Lebesgue spaces of differentiable functions and distributions, in Proc. Symp. Pure Math., Vol. IV, 1961, 33–49.
  • [6] L. C. Evans and R. F. Gariepy, Measure theory and fine properties of functions. Revised edition, Textbooks in Mathematics. CRC Press, Boca Raton, FL (2015), xiv+299 pp.
  • [7] E. De Giorgi, Nuovi teoremi relativi alle misure (r1)(r-1)-dimensionali in uno spazio a rr dimensioni, Ricerche Mat. 4 (1955), 95–113.
  • [8] H. Federer, Geometric Measure Theory. Berlin, Heidelberg, New York: Springer 1969.
  • [9] P. Hajłasz and J. Kinnunen, Holder quasicontinuity of Sobolev functions on metric spaces. Rev. Mat. Iberoamericana, 14 (1998), 601–622.
  • [10] P. Hajłasz, P. Koskela, and H. Tuominen, Sobolev embeddings, extensions and measure density condition, J. Funct. Anal. 254 (2008), no. 5, 1217–1234.
  • [11] H. Hakkarainen, J. Kinnunen, P. Lahti and P. Lehtelä, Relaxation and Integral Representation for Functionals of Linear Growth on Metric Measure spaces, Anal. Geom. Metr. Spaces 2016 4, 288–313.
  • [12] P. W. Jones, Quasiconformal mappings and extendability of Sobolev functions, Acta Math. 47 (1981), 71–88.
  • [13] P. Koskela, Capacity extension domains, Ann. Acad. Sci. Fenn. Ser. A I Math. Dissertationes No. 73 (1990), 42 pp.
  • [14] P. Koskela, M. Miranda Jr. and N. Shanmugalingam, Geometric properties of planar BV extension domains, in: Around the Research of Prof. Maz’ya I, in: International Mathematical Series, 2010, pp.255–272, Function Spaces; Topics (Springer collection).
  • [15] P. Koskela, T. Rajala and Yi Ru-Ya Zhang, A geometric characterization of planar Sobolev extension domains, preprint.
  • [16] P. Koskela, T. Rajala and Y. Zhang, Planar W1, 1W^{1,\,1}-extension domains, preprint.
  • [17] P. Lahti, Extensions and traces of functions of bounded variation on metric spaces, J. Math. Anal. Appl. 423(1) (2015) 521–537.
  • [18] P. Lahti, X. Li and Z. Wang, Traces of Newton-Sobolev, Hajłasz-Sobolev, and BV functions on metric spaces, Ann. Sc. Norm. Super. Pisa Cl. Sci. (5), accepted.
  • [19] D. Lučić, E. Pasqualetto and T. Rajala, Sharp estimate on the inner distance in planar domains, Ark. Mat., 58 (2020), 133–159.
  • [20] V. Maz’ya, Sobolev spaces with applications to elliptic partial differential equations. Second, revised and augmented edition. Grundlehren der Mathematischen Wissenschaften, 342. Springer, Heidelberg, 2011. xxviii+866 pp.
  • [21] P. Shvartsman, Local approximations and intrinsic characterization of spaces of smooth functions on regular subsets of RnR^{n}, Math. Nachr. 279 (2006), no. 11, 1212–1241.
  • [22] P. Shvartsman, On Sobolev extension domains in n\mathbb{R}^{n}, J. Funct. Anal. 258 (2010), no. 7, 2205–2245.
  • [23] E. M. Stein, Singular integrals and differentiability properties of functions, Princeton University Press, Princeton, New Jersey, 1970.