Paper The following article is Free article

Sobolev $W^{1}_{p}$-spaces on $d$-thick closed subsets of $ {\mathbb {R}}^{n}$

and

© 2020 Russian Academy of Sciences (DoM) and London Mathematical Society
, , Citation S. K. Vodopyanov and A. I. Tyulenev 2020 Sb. Math. 211 786 DOI 10.1070/SM9199

1064-5616/211/6/786

Abstract

Let $S \subset \mathbb{R}^{n}$ be a nonempty closed set such that for some ${d \in [0,n]}$ and $\varepsilon>0$ the $d$-Hausdorff content $\mathscr{H}^{d}_{\infty}(S \cap Q(x,r)) \geq \varepsilon r^{d}$ for all cubes $Q(x,r)$ with centre $x \in S$ and edge length $2r \in (0,2]$. For each $p>\max\{1,n-d\}$ we give an intrinsic characterization of the trace space $W_{p}^{1}(\mathbb{R}^{n})|_{S}$ of the Sobolev space $W_{p}^{1}(\mathbb{R}^{n})$ to the set $S$. Furthermore, we prove the existence of a bounded linear operator $\operatorname{Ext}\colon W_{p}^{1}(\mathbb{R}^{n})|_{S} \,{\to}\, W_{p}^{1}(\mathbb{R}^{n})$ such that $\operatorname{Ext}$ is the right inverse to the standard trace operator. Our results extend those available in the case $p \in (1,n]$ for Ahlfors-regular sets $S$.

Bibliography: 36 titles.

Export citation and abstract BibTeX RIS

The research of S. K. Vodopyanov was carried out within the framework of a state assignment of the Ministry of Education and Science of the Russian Federation for the Institute of Mathematics of the Siberian Branch of the Russian Academy of Sciences (project no. 0314-2019-0006).

§ 1. Introduction

Given $m,n \in \mathbb{N}$ let $C^{m}(\mathbb{R}^{n})$ denote the linear space of all functions on $\mathbb{R}^{n}$ with continuous partial derivatives up to order $m$ equipped with the standard seminorm. The problem considered by Whitney in 1934 in his famous papers [1] and [2] reads as follows.

Classical Whitney Extension Problem.  Let $m,n\in\mathbb{N}$ and let $S$ be an arbitrary nonempty subset of $\mathbb{R}^{n}$. How can we decide whether a given function $f\colon S \to \mathbb{R}$ extends to a $C^{m}(\mathbb{R}^{n})$-function?

Whitney [2] solved this problem completely only in the case $n=1$. Furthermore, he gave a solution of the analogous problem in the context of the Lipschitz spaces $C^{m-1,1}(\mathbb{R}^{n})$, $m,n \in \mathbb{N}$ (see [1]). After [1] and [2], great progress was made by many authors (see [3]–[5] and also the references there). Fefferman gave a complete solution (that is, for all $m,n \in \mathbb{N}$) of the Classical Whitney Extension Problem [6]–[9] only recently.

Recall that, according to the classical Sobolev embedding theorem (for example, see [10], Ch. I, §1.8.2), in the case $m \in \mathbb{N}$, $j \in \{1,\dots,m\}$ and $p>n/(m-j+1)$, for every $F \in W_{p}^{m}(\mathbb{R}^{n})$ there exists a representative $\widehat{F} \in C^{j-1}(\mathbb{R}^{n})$. This fact enables one to identify each element $F \in W_{p}^{m}(\mathbb{R}^{n})$ with its unique continuous representative. This implies that $F$ has a well-defined restriction to any given nonempty subset of $\mathbb{R}^{n}$. As a result, in the case $p>n$ we can consider the analogue of the Classical Whitney Extension Problem, where $C^{m}(\mathbb{R}^{n})$ is replaced by $W_{p}^{m}(\mathbb{R}^{n})$. There is extensive literature devoted to these problems (see [11]–[19]).

In the case when $m,n \in \mathbb{N}$, $n \geq 2$ and $1<mp \leqslant n$, functions in the space $W_{p}^{m}(\mathbb{R}^{n})$ do not (in general) have continuous representatives (see Ch. 5, §6 in [20]). Nevertheless, every $F \in W_{p}^{m}(\mathbb{R}^{n})$ has a 'sufficiently nice' representative $\widehat{F}$ which has a well-defined trace $\widehat{F}|_{S}$ on each set $S \subset \mathbb{R}^{n}$ with positive $C_{m,p}$-capacity. Unfortunately, by contrast with the case $p>n$, in the case $1<p \leqslant n$ Whitney-type problems have been posed and solved only in very particular cases. More precisely, either Ahlfors $d$-regular sets $S$ [21]–[23] or special cusps in $\mathbb{R}^{2}$ [24] have been considered. Under minimal restrictions on $S$ the corresponding problem is very complicated and has never been considered. In [25], for each $d \in [0,n]$ Rychkov introduced so-called $d$-thick sets $S \subset \mathbb{R}^{n}$, which are regular with respect to the $d$-Hausdorff content, that is, $\mathscr{H}^{d}_{\infty}(Q(x,r) \cap S) \approx r^{d}$ for all $x \in S$ and $r \in (0,1]$. For such sets he considered Whitney-type problems in the context of Besov and Lizorkin-Triebel spaces. Given $d \in [0,n]$, the class of Ahlfors $d$-regular sets is strictly contained in the class of $d$-thick sets, but the latter is much wider. For example, every path connected subset of $\mathbb{R}^{n}$ is 1-thick but in general is not Ahlfors 1-regular (see Example 2.1 below).

In this paper we solve the following problem.

Problem A.  Let $d \in [0,n]$, $p \in (\max\{1,n-d\},\infty]$, and let $S\subset \mathbb{R}^{n}$ be a $d$-thick closed set. Given a function $f\colon S \to \mathbb{R}$, how can we decide whether there exists a function $F \in W_{p}^{1}(\mathbb{R}^{n})$ such that $\widehat{F}|_{S}(x)=f(x)$ for $C_{1,p}$-quasi-every $x \in S$? Consider the $W_{p}^{1}(\mathbb{R}^{n})$-norms of all functions $F \in W_{p}^{1}(\mathbb{R}^{n})$ such that $\widehat{F}|_{S}(x)=f(x)$ for $C_{1,p}$-q.e. $x \in S$. How small can these norms be?

Given a set $S \subset \mathbb{R}^{n}$ with $C_{1,p}(S) > 0$, we denote the usual trace space of the space $W_{p}^{1}(\mathbb{R}^{n})$ by $W_{p}^{1}(\mathbb{R}^{n})|_{S}$ and let $\operatorname{Tr}|_{S}$: $W_{p}^{1}(\mathbb{R}^{n}) \to W_{p}^{1}(\mathbb{R}^{n})|_{S}$ denote the corresponding trace operator. In our paper we obtain also a solution to the following problem.

Problem B.  Let $d \in [0,n]$ and $p \in (\max\{1,n-d\},\infty]$, and let $S\subset \mathbb{R}^{n}$ be a closed $d$-thick set. Does there exist a bounded linear operator $\operatorname{Ext}\colon W_{p}^{1}(\mathbb{R}^{n})|_{S} \to W_{p}^{1}(\mathbb{R}^{n})$ such that $\operatorname{Tr}|_{S} \circ \operatorname{Ext}=\operatorname{Id}$ on $W_{p}^{1}(\mathbb{R}^{n})|_{S}$?

In §4 we present solutions to Problems A and B. In §5 we consider simplified versions of these problems, when the set $S$ has a porous boundary. In this case the corresponding criterion (the solution to Problem A) can be simplified. In §6 we show that our main results include the corresponding results concerning $W^{1}_{p}$-spaces obtained in [21], [23] and [24] as particular cases.

Finally we would like to underline that the methods in [25] gave a solution to Problem B only in the case $d>n-1$. To the best of our knowledge, Problem A has never been considered in the literature. We introduce new methods which have never been used before. For example, we introduce the concept of a $d$-regular sequence of measures and generalized Calderón-type maximal functions with respect to such sequences. Such tools help us to capture the smoothness properties of functions in the trace space and enable us to solve Problems A and B for every $d \in [0,n]$.

The authors wish to acknowledge their gratitude to Professor P. Shvartsman, who read early versions of the manuscript and made many valuable comments.

§ 2. Necessary background and statements of the main results

Throughout the paper $C,C_{1},C_{2},\dots$ will be generic positive constants. These constants can change even in a single string of estimates. The dependence of a constant on certain parameters is shown, for instance, by the notation $C=C(n,p,k)$. We write $A \approx B$ if there is a constant $C \geq 1$ such that $A/C \leqslant B \leqslant C A$.

Throughout the paper $x=(x_{1},\dots,x_{n})$ denotes an element of the space $\mathbb{R}^{n}$. The symbols $\alpha$ and $\beta$ will be used to denote multi-indices, that is, elements of the space $\mathbb{N}^{n}_{0}$. Following [23] it will often be convenient to measure distances in $\mathbb{R}^{n}$ in the uniform norm $\|x\|:=\|x\|_{\infty}:=\max\{|x_{i}|\colon i=1,\dots,n\}$, $x \in \mathbb{R}^{n}$. Given two subsets $A$ and $B$ of $ \mathbb{R}^{n}$, set $\operatorname{dist}(A,B):=\inf\{\|a-b\|_{\infty}\colon a \in A,\, b \in B\}$. For any $C \subset \mathbb{R}^{n}$ we also set $\operatorname{diam}C:=\sup\{\|a-a'\|_{\infty}\colon a,a' \in C\}$.

The symbols $B(x,r)$ and $Q(x,r)$ stand for the closed balls with centre $x$ and radius $r>0$ in the standard Euclidean norm $\|\cdot\|_{2}$ and in the uniform norm $\|\cdot\|_{\infty}$, respectively (we will also call $Q=Q(x,r)$ a cube) Given a number $c>0$ we write $cB$ ($cQ$) to denote the ball $B(x,cr)$ (the cube $Q(x,cr)$, respectively). By a dyadic cube we mean an arbitrary half-open cube $\widetilde{Q}_{k,m}:=\prod_{i=1}^{n}[{m_{i}}/{2^{k}}, (m_{i}+1)/{2^{k}})$, $k \in \mathbb{Z}$, $m=(m_{1},\dots,m_{n}) \in \mathbb{Z}^{n}$. Given $k \in \mathbb{Z}$, we let $\mathscr{Q}_{k}$ denote the mesh of all dyadic cubes with edge length $2^{-k}$.

For any set $E \subset \mathbb{R}^{n}$ we let $\overline{E}$ and $\operatorname{int}E$ denote the closure and interior of $E$ in the topology induced by an arbitrary norm in $\mathbb{R}^{n}$ (recall that all norms in $\mathbb{R}^{n}$ are equivalent). For $A \subset \mathbb{R}^{n}$ and $\delta>0$ we define the $\delta$-neighbourhood of $A$ to be $U_{\delta}(A):=\bigcup_{x \in A}\operatorname{int}B_{\delta}(x)$.

Given a Borel measure $\mathfrak{m}$ and a nonempty Borel set $S \subset \mathbb{R}^{n}$, we define the restriction of $\mathfrak{m}$ to $S$. More precisely, we set $\mathfrak{m}\lfloor_S (G) := \mathfrak{m}(G \cap S)$ for every nonempty Borel set $G \subset \mathbb{R}^{n}$.

Let $\mathfrak{m}$ be an arbitrary Borel measure on $\mathbb{R}^{n}$. Given $f \in L^{\mathrm{loc}}_{1}(\mathbb{R}^{n},\mathfrak{m})$, for every Borel set $G \subset \mathbb{R}^{n}$ with $\mathfrak{m}(G)<+\infty$ we set

Equation (2.1)

2.1. Ahlfors $d$-regular sets and $d$-thick sets

Given $S \subset \mathbb{R}^{n}$, $0 \leqslant d \leqslant n$ and $\delta \in (0,+\infty]$ we set

where the infimum is taken over all countable covers of $S$ by cubes $Q(x_{j},r_{j})$ with arbitrary centres $x_{j}$ and radii $ r_{j}<\delta$. We call the quantity $\mathscr{H}^{d}_{\infty} (S)$ the $d$-Hausdorff content of $S$.

We define the Hausdorff $d$-measure of $S$ by $\mathscr{H}^{d}(S):=\lim_{\delta \to 0}\mathscr{H}^{d}_{\delta} (S)$. Note that our definitions of Hausdorff contents and measures are slightly different from the classical ones (cf. §5.1 in [26]). We use covers by balls in the $\|\cdot\|_{\infty}$-norm, that is, by cubes instead of classical balls. Up to some universal constants both approaches return the same values of the corresponding measures. Note also that in our case $\mathscr{H}^{n}$ coincides with the classical Lebesgue measure $\mathscr{L}_{n}$ on the Borel $\sigma$-algebra of $\mathbb{R}^{n}$.

Given $m \in \mathbb{N}$ and $p \in (1,\infty)$, recall the concept of $C_{m,p}$-capacity (see §2.1 of [26]). In what follows we say that some property holds $(m,p)$-quasi-everywhere ($(m,p)$-q.e. for short) if it holds everywhere outside some set $E$ with ${C_{m,p}(E)=0}$. The following property summarizes the connections between $C_{1,p}$-capacity and Hausdorff measures (see Theorems 5.1.9 and 5.1.13 in [26]).

Proposition 2.1.  Let $p \in (1,n]$ and let $E \subset \mathbb{R}^{n}$. If $\mathscr{H}^{n-p}(E)<+\infty$, then $C_{1,p}(E)=0$. Conversely, if $C_{1,p}(E)=0$, then $\mathscr{H}^{d}(E)=0$ for every $d>n-p$.

We have taken the following definition from [22], Ch. 2, §§1.1 and 1.2.

Definition 2.1.  Given $d \in [0,n]$, we say that a closed set $S \subset \mathbb{R}^{n}$ is Ahlfors $d$-regular (or just a $d$-set for short) if there exists a $d$-measure on $S$, that is, a Borel measure $\mathfrak{m}$ with $\operatorname{supp}\mathfrak{m}=S$ such that for all $x \in S$ and $r \in (0,1]$

Equation (2.2)

where the positive constants $c_{1}(\mathfrak{m}), c_{2}(\mathfrak{m})$ depend on $\mathfrak{m}$ but do not depend on $x$ or $r$.

Remark 2.1.  We can show that given a (closed) $d$-set $S$ the restriction $\mathscr{H}^{d}\lfloor_{S}$ is a $d$-measure on $S$ (see [22], §1.2, Theorem 1, for details).

Using Remark 2.1 we introduce the following important notation. In what follows, given an Ahlfors $d$-regular (closed) set $S$ we set $c^{d}_{i}(S):=c_{i}$ ($\mathscr{H}^{d}\lfloor_{S}$), $i=1,2$.

The following proposition will be useful for comparing $d$-sets with $d$-thick sets below.

Lemma 2.1.  Let $d\in[0,n]$ and let $\mathscr{I}$ be a nonempty index set. Let $\{S_{\alpha}\}_{\alpha \in \mathscr{I}}$ be a family of $d_{\alpha}$-sets with $d_{\alpha} \in [d,n]$, $\alpha \in \mathscr{I}$, such that the set $S:=\bigcup_{\alpha \in \mathscr{I}}S_{\alpha}$ is closed and Ahlfors $d$-regular. Then the following holds.

(1) $d_{\alpha}=d$ for every $\alpha \in \mathscr{I}$.

(2) Suppose the family $\{S_{\alpha}\}_{\alpha \in \mathscr{I}}$ is such that $\mathscr{H}^{d}(S_{\alpha} \cap S_{\alpha'}) = 0$ for every ${\alpha, \alpha' \in \mathscr{I}}$, ${\alpha \neq \alpha'}$, $\inf_{\alpha \in \mathscr{I}}c^{d}_{1}(S_{\alpha}) > 0$ and $\bigcap_{\alpha \in \mathscr{I}}S_{\alpha} \neq \varnothing$. Then $\operatorname{card}\mathscr{I} < \infty$.

Proof.  (1) If there exists $\alpha_{0} \in \mathscr{I}$ for which $d_{\alpha_{0}}>d$, then Remark 2.1 and (2.2) with $\mathfrak{m}=\mathscr{H}^{d_{\alpha_{0}}}\lfloor_S$, together with elementary properties of Hausdorff measures, show that $\mathscr{H}^{d}(Q(x,r) \cap S)=+\infty$ for all $x \in S_{\alpha_{0}}$ and all $r \in (0,1]$. Hence the right-hand side of (2.2) must be violated for $\mathfrak{m}=\mathscr{H}^{d}\lfloor_S$, each $x \in S_{\alpha_{0}}$ and $r \in (0,1]$. This contradicts the Ahlfors $d$-regularity of $S$.

(2) Fix a point $x_{0} \in \bigcap_{\alpha \in \mathscr{I}}S_{\alpha}$. By (1), $d_{\alpha}=d$ for all $\alpha \in \mathscr{I}$. Assume that $\operatorname{card}\mathscr{I}=\infty$. Then it is easy to see from our assumptions that for every $r \in (0,1]$ there is a countable family of different indices $\{\alpha_{k}\} \subset \mathscr{I}$ such that $\widetilde{S}_{\alpha_{k}}(r):=Q(x_{0},r) \cap (S_{\alpha_{k}} \setminus \bigcup_{j=1}^{k-1}S_{\alpha_{j}}) \neq \varnothing$ for all $k \in \mathbb{N}$. Hence there is a sequence of distinct points $\{x_{k}\}$ such that $x_{k} \in \widetilde{S}_{\alpha_{k}}(r/2)$ for every $k \in \mathbb{N}$. Let $c^{d}(S):=\inf_{\alpha \in \mathscr{I}}c^{d}_{1}(S_{\alpha})$. It is clear from the construction and Remark 2.1 that $\mathscr{H}^{d}(\widetilde{S}_{\alpha_{k}}(r)) \geq \mathscr{H}^{d}(Q(x_{k},r/2)\mkern1mu \cap\mkern1mu \widetilde{S}_{\alpha_{k}}(r))\mkern-1mu \geq \mkern-1mu \mathscr{H}^{d}(Q(x_{k},r/2)\mkern1mu \cap\mkern1mu S_{\alpha_{k}})\mkern-1mu \geq\mkern-1mu c^{d}(S) (r/2)^{d}$ for every ${r\in (0,1]}$. Hence the fact that the sets $\widetilde{S}_{\alpha_{k}}(r)$ are disjoint and measurable, together with the countable additivity of the Hausdorff measures, gives the following estimate, which contradicts (2.2). Namely, for all $r \in (0,1]$

The proof is complete.

To the best of our knowledge the following concept was first introduced in [25].

Definition 2.2.  Let $d \in [0,n]$. A set $S \subset \mathbb{R}^{n}$ is said to be $d$-thick if there exists a constant $c^{d}_{3}(S)>0$ such that for all $x \in S$ and $r \in (0,1]$

Equation (2.3)

The following proposition is an immediate consequence of Definition 2.2. We omit the proof.

Proposition 2.2.  Let $S \subset \mathbb{R}^{n}$ be a $d$-thick set for some $d \in [0,n]$. Then

(1) the closure $\overline{S}$ of $S$ is $d$-thick;

(2) $S$ is $d'$-thick for every $d'$, $0 \leqslant d' \leqslant d$, and $c^{d'}_{3}(S) \geq c^{d}_{3}(S)$.

The following lemma exhibits an important relation between the concepts of Ahlfors $d$-regular sets and $d$-thick sets.

Lemma 2.2.  Let $d \in [0,n]$. Every Ahlfors $d$-regular set $S \subset \mathbb{R}^{n}$ is $d$-thick. Furthermore,

Equation (2.4)

Proof.  Suppose that $S\subset \mathbb{R}^{n}$ is an Ahlfors $d$-regular set. Fix a cube $Q=Q(x,r)$ with $x \in S$ and $0<r \leqslant 1$. Let $\{Q_{j}\}_{j \in \mathbb{N}}=\{Q(x_{j},r_{j})\}_{j \in \mathbb{N}}$ be a covering of $Q \cap S$ such that $\mathscr{H}^{d}_{\infty} (Q \cap S) \geq (1/2)\sum_{j \in \mathbb{N}}(r_{j})^{d}$. Clearly, we can assume that $Q_{j} \cap S \neq \varnothing$ for all $j \in \mathbb{N}$. For every $j \in \mathbb{N}$ fix a point $\widetilde{x}_{j} \in Q_{j} \cap S$. Using Remark 2.1, estimate (2.2), and the subadditivity of $\mathscr{H}^{d}_\infty$ we obtain the required estimate:

Equation (2.5)

The proof is complete.

The next result is a direct consequence of Lemma 2.2, Proposition 2.2, (2), and the monotonicity of $\mathscr{H}^{d}_{\infty}$.

Lemma 2.3.  Let $\mathscr{I}$ be an arbitrary nonempty index set. Let $0 \leqslant d \leqslant d_{\alpha} \leqslant n$ for every $\alpha \in \mathscr{I}$. Let $\{S_{\alpha}\}_{\alpha \in \mathscr{I}}$ be a family of Ahlfors $d_{\alpha}$-regular sets and let $S:=\bigcup_{\alpha \in \mathscr{I}}S_{\alpha}$. Then $S$ is $d$-thick and

Equation (2.6)

Now that we have Lemmas 2.1 and 2.3 at our disposal, we can present useful examples which illustrate the huge difference between Definition 2.1 and Definition 2.2.

Example 2.1.  Let $\Omega$ be a path connected subset of $\mathbb{R}^{n}$. Then $\Omega$ and $\overline{\Omega}$ are $1$-thick. In fact, fix a point $x \in \Omega$. Let $Q=Q(x,r)$ be a cube with edge length $0<2r \leqslant 2$. Consider two cases.

In the first case there is a point $y \in \Omega \setminus Q$. Hence there is a curve $\gamma_{x,y}$ which connects $x$ and $y$. Let $\{Q_{j}\}_{j \in \mathbb{N}}=\{Q(x_{j},r_{j})\}_{j \in \mathbb{N}}$ be an arbitrary covering of $Q \cap \Omega$ for which

Equation (2.7)

We choose an index set $\mathscr{A} \subset \mathbb{N}$ such that $\gamma_{x,y} \cap Q_{j} \neq \varnothing$ for every $j \in \mathscr{A}$ and $\gamma_{x,y} \subset \bigcup_{j \in \mathscr{A}}Q_{j}$. Consider the projections $\gamma^{i}_{x,y}$, $i=1,\dots,n$, of our curve and the projections $Q^{i}_{j}$ of cubes in the cover onto the $i$th coordinate axes. Since we measure distances in the $\|\cdot\|_{\infty}$-norm, there exists $i_{0} \in \{1,\dots,n\}$ for which $\mathscr{H}^{1}(\gamma^{i_{0}}_{x,y}) \geq r$. By construction the family of closed intervals $\{Q^{i_{0}}_{j}\}_{j \in \mathscr{A}}$ covers $\gamma^{i_{0}}_{x,y}$. Hence from (2.7) we derive

Equation (2.8)

In the second case $\Omega \subset Q(x,r)$. Since $r \leqslant 1$, we have

Equation (2.9)

Combining (2.8) and (2.9) shows that $\Omega$ is $1$-thick, and we can take $c^{1}_{3}(\Omega)=\min\{\mathscr{H}^{1}_{\infty}(\Omega), 1/2\}$.

It is obvious that this set cannot be Ahlfors 1-regular for $n \geq 2$. In addition, elementary computations show that for each $s>1$ and $n \geq 2$ the cusp $\Omega_{s}:=\{x=(x',x_{n}) \in \mathbb{R}^{n}\colon x_{n}>0,\, \|x'\|_{\infty} \leqslant x_{n}^{s}\}$ cannot be Ahlfors $d$-regular for $d \in [0,n]$.

Example 2.2.  Let $S := \overline{Q}(0,1) \cup ([1,2] \times \{0\}) \subset \mathbb{R}^{2}$. This set is $1$-thick as a union of a $2$-thick set (a square) and a $1$-thick set (a line interval). From Lemma 2.1, (1), it follows that $S$ is not Ahlfors $1$-regular.

Example 2.3.  Let $E:= \bigcup_{n \in \mathbb{N}}\{(r,\varphi)\colon r \in [0,1], \,\varphi=2^{-n}\} \cup [0,1] \times \{0\} \subset \mathbb{R}^{2}$. From Lemma 2.3 it follows that $E$ is $1$-thick. It is not Ahlfors $1$-regular by Lemma 2.1, (2).

Example 2.4.  We can show that any $(\varepsilon,\delta)$-domain $\Omega$ is $n$-thick. We present only a sketch of the proof. Fix $x \in \Omega$ and $r< \min\{\operatorname{diam}\Omega,\delta\}/4$. Choose an arbitrary $y \in \Omega$ so that $\|x-y\| \geq r$. Then it easily follows from formulae (1.1) and (1.2) in [27] that there exists a curve $\gamma_{x,y}$ and a point $z \in \gamma_{x,y} \cap \partial Q(x,r/3)$ such that $B(z,c(\varepsilon, \delta, n) r ) \subset \Omega$ for some positive constant $c(\varepsilon, \delta, n)$. The case $r \geq \min\{\operatorname{diam}\Omega,\delta\}/4$ can be considered similarly to the second case in Example 2.1.

2.2. Regular sequences of measures and Calderón-type maximal functions

The following concept is one of the cornerstones which enable us to solve Problem A.

Definition 2.3.  Let $S$ be a closed $d$-thick set for some $d \in [0,n]$. Let $\{\mu_{k}\}_{k \in \mathbb{N}_{0}}$ be a sequence of Borel measures such that $\operatorname{supp}\mu_{k}=S$, $k \in \mathbb{N}_{0}$. We say that $\{\mu_{k}\}_{k \in \mathbb{N}_{0}}$ is a $d$-regular sequence of measures on $S$ if and only if for some $C>0$ the following properties hold for every $k \in \mathbb{N}_{0}$:

Equation (2.10)

Equation (2.11)

Equation (none)

Equation (2.12)

Remark 2.2.  It is clear that there exists a largest positive constant $C$ for which (2.11) holds. We denote it by $C_{\{\mu_{k}\}}$.

Remark 2.3.  We show in §3 below that Definition 2.3 is consistent: for every $d$-thick closed set $S$ there exists a $d$-regular sequence of measures on $S$.

Definition 2.4.  Let $\mathfrak{m}$ be an arbitrary nonzero Radon measure. Let $Q=Q(x,r)$ be a closed cube. Given a function $f \in L_{1}^{\operatorname{loc}}(\mathbb{R}^{n},\mathfrak{m})$, the best approximation to $f$ by constants on $Q$, normalized with respect to $\mathfrak{m}$, is defined by .

Remark 2.4.  Elementary computations give

Equation (2.13)

(recall (2.1)).

Here and in the sequel we use the following notation. Given a number ${r \in (0,1]}$ we set $k(r):=\bigl|[\operatorname{log}_{2}r]\bigr|$, so that this is the unique integer such that $r\in [2^{-k(r)}, 2^{-k(r)+1})$.

The following definition is a far-reaching generalization of the classical concept of a maximal function measuring smoothness first introduced by Calderón [28].

Definition 2.5.  Let $S \subset \mathbb{R}^{n}$ be a $d$-thick closed set for some $d \in [0,n]$. Let $\{\mu_{k}\}_{k \in \mathbb{N}_{0}}$ be a $d$-regular sequence of measures on $S$. Let $f \in L^{\mathrm{loc}}_{1}(\mathbb{R}^{n},\mu_{k})$ for every $k \in \mathbb{N}_{0}$. Given $t \in [0,1]$, we consider the Calderón-type maximal function with respect to $\{\mu_{k}\}_{k \in \mathbb{N}_{0}}$. For every $x \in \mathbb{R}^{n}$

Remark 2.5.  We set $f^{\sharp}_{\{\mu_{k}\}}(x):=f^{\sharp}_{\{\mu_{k}\}}(x,0)$ for brevity. If the set $S$ is Ahlfors $n$-regular, we can take $\mu_{k}=\mathscr{H}^{n}\lfloor_{S}$ for every $k \in \mathbb{N}_{0}$. Hence, in this case our maximal function $f^{\sharp}_{\{\mu_{k}\}}$ is similar to that introduced by Shvartsman [23]. In particular, if $S=\mathbb{R}^{n}$, we obtain the Calderón-type maximal function (see [28]).

2.3. Porous sets

Definition 2.6.  Let $S$ be a closed nonempty subset of $\mathbb{R}^{n}$ and let $\lambda \in (0,1)$. For every $j \in \mathbb{N}_{0}$ we define

and call $S_{j}(\lambda)$ the maximal $2^{-j}$-porous subset of $S$. We say that $S$ is porous if there exists $\lambda \in (0,1)$ such that $S_{j}(\lambda)=S$ for every $j \in \mathbb{N}_{0}$.

We gather some useful facts about porous sets. The second (see item (2) below) is a special case of Proposition 9.18 in [29].

Proposition 2.3.  Let $S$ be a closed nonempty subset of $\mathbb{R}^{n}$ and let $\lambda \in (0,1)$. Then

(1) $S_{j}(\lambda)$ is closed for every $j \in \mathbb{N}_{0}$;

(2) if $S$ is Ahlfors $d$-regular for some $d \in [0,n)$, then $S$ is porous.

Example 2.5.  Let $\beta\colon [0,+\infty) \to [0,+\infty)$ be a continuous function that is strictly increasing and such that $\beta(0)=0$ and $\beta(t)>0$, $t>0$. Consider the closed single cusp

It is easy to see that the boundary $\partial G^{\beta}$ of $G^{\beta}$ is porous.

2.4. Trace spaces of Sobolev spaces

Recall that given $p \in [1,\infty]$, $n \in \mathbb{N}$ and an open set $G\subset \mathbb{R}^{n}$, the Sobolev space $W_{p}^{1}(G)$ is the linear space of all (equivalence classes of) real functions $F \in L_{p}(G)$ whose generalized partial derivatives on $G$ $D^{\alpha}F$, $|\alpha| \leqslant 1$, belong to $L_{p}(G)$. This space is equipped with the norm

Equation (2.14)

The next result, which is a very special case of Theorem 6.2.1 in [26], will help us to define the trace of a Sobolev function $F$ consistently on a given 'sufficiently massive' set $S$.

Proposition 2.4.  Let $p \in (1,\infty]$ and $F \in W_{p}^{1}(\mathbb{R}^{n})$. If $p \in (1,n]$, then there exists a set $E_{F} \subset \mathbb{R}^{n}$ with $C_{1,p}(E_{F})=0$ and a representative $\widehat{F}$ of the element $F$ such that every point $x \in \mathbb{R}^{n} \setminus E_{F}$ is a Lebesgue point of $\widehat{F}$. If $p>n$, then there exists a continuous representative $\widehat{F}$ of $F$.

In the sequel we call the representative $\widehat{F}$ constructed in Proposition 2.4 a good representative of $F$. Recall that, given $p \in (1,\infty)$, a property of $F$ is said to hold $(1,p)$-quasi-everywhere ($(1,p)$-q.e. for short) if it holds everywhere except on a set of $C_{1,p}$-capacity zero.

Definition 2.7.  Let $p \in (1,n]$ and $F \in W_{p}^{1}(\mathbb{R}^{n})$. Let $\widehat{F}$ be a good representative of $F$. Given a set $S \subset \mathbb{R}^{n}$ with $C_{1,p}(S)>0$, the trace $F|_{S}$ of $F$ on $S$ is the class of equivalent (modulo sets of $C_{1,p}$-capacity zero) functions $f\colon S \to \mathbb{R}$ such that $\widehat{F}(x)=f(x)$ for $(1,p)$-q.e. $x \in S$.

Definition 2.8.  Let $p \in (n,\infty]$ and $F \in W_{p}^{1}(\mathbb{R}^{n})$. Let $\widehat{F}$ be a continuous representative of $F$. Given a nonempty set $S \subset \mathbb{R}^{n}$ the trace $F|_{S}$ of $F$ on the set $S$ is the pointwise restriction of $\widehat{F}$ to $S$.

Below we identify a function $f\colon S \to \mathbb{R}$ and the class of functions each of which coincides $(1,p)$-quasi-everywhere with $f$ on $S$.

Now using Definitions 2.7 and 2.8 we introduce the following.

Definition 2.9.  Let $p \in (1,\infty]$. Given a nonempty set $S \subset \mathbb{R}^{n}$ with $C_{1,p}(S)> 0$, we define the trace space $W_{p}^{1}(\mathbb{R}^{n})|_{S}$ of the space $W_{p}^{1}(\mathbb{R}^{n})$ as follows:

We equip this space with the usual trace norm

where the infimum is taken over all $F \in W_{p}^{1}(\mathbb{R}^{n})$ such that $F|_{S}=f$. Furthermore, we define the trace operator $\operatorname{Tr}|_{S}\colon W_{p}^{1}(\mathbb{R}^{n}) \to W_{p}^{1}(\mathbb{R}^{n})|_{S}$ which takes $F \in W_{p}^{1}(\mathbb{R}^{n})$ and gives back $F|_{S}$.

Definition 2.10.  Let $p \in (1,\infty]$. Let $S \subset \mathbb{R}^{n}$ be a nonempty set. Assume that $C_{1,p}(S) > 0$ whenever $p \in (1,n]$. We say that a map $\operatorname{Ext}\colon W_{p}^{1}(\mathbb{R}^{n})|_{S} \to W_{p}^{1}(\mathbb{R}^{n})$ is an extension operator if it is the right inverse for the trace operator, so that $\operatorname{Tr}|_{S} \circ \operatorname{Ext}=\operatorname{Id}$ on $W_{p}^{1}(\mathbb{R}^{n})|_{S}$.

Remark 2.6.  Let $d \in [0,n]$, and let $S \subset \mathbb{R}^{n}$ be a $d$-thick set. It is important to underline that Proposition 2.1 and Definition 2.9 clearly imply that for every $p \in (\max\{1,n - d\},\infty)$ the trace space $W^{1}_{p}(\mathbb{R}^{n})|_{S}$ is well defined. Furthermore, our definitions immediately implies that the trace operator $\operatorname{Tr}|_{S}\colon W_{p}^{1}(\mathbb{R}^{n}) \to W_{p}^{1}(\mathbb{R}^{n})|_{S}$ is linear and bounded.

Remark 2.7.  In the case $p=\infty$ the Sobolev space $W_{\infty}^{1}(\mathbb{R}^{n})$ can be identified with the space $\operatorname{LIP}(\mathbb{R}^{n})$ of Lipschitz functions, and it is known that the restriction $\operatorname{LIP}(\mathbb{R}^{n})|_{S}$ of the latter coincides with the space $\operatorname{LIP}(S)$ of Lipschitz continuous functions on $S$ and that, furthermore, the classical Whitney extension operator maps $\operatorname{LIP}(S)$ linearly and continuously into $\operatorname{LIP}(\mathbb{R}^{n})$ (for instance, see [20], Ch. 6). Hence in the sequel we will only deal with the case $1<p <\infty$.

2.5. Statements of the main results

As we said above, without loss of generality we can work with the case $p \neq \infty$.

Given a closed set $S \subset \mathbb{R}^{n}$ and $k \in \mathbb{N}_{0}$, we define

Definition 2.11.  Let $p \in (1,\infty)$, $d \in [0,n]$ and $\lambda \in (0,1)$. Let $S \subset \mathbb{R}^{n}$ be a $d$-thick closed set. Let $\{\mu_{k}\}=\{\mu_{k}\}_{k \in \mathbb{N}_{0}}$ be a $d$-regular sequence of measures on $S$. For every $p \in (1,\infty)$ we define the following nonnegative functionals (with values in $[0,+\infty]$) on the space $W_{p}^{1}(\mathbb{R}^{n})|_{S}$:

Equation (2.15)

Remark 2.8.  From Proposition 2.1 and Lemma 3.6 it follows that all the functionals in (2.15) are well defined on the trace space $W_{p}^{1}(\mathbb{R}^{n})|_{S}$. More precisely their values remain the same after changing the function $f$ on a set of $C_{1,p}$-capacity zero.

Remark 2.9.  The symbols $\mathscr{SN}_{S,p}$, $\widetilde{\mathscr{BN}}_{S,p,\lambda}$ and $\mathscr{BN}_{S,p}$ have not been picked at random. Informally speaking, $\mathscr{SN}_{p}$ is the 'Sobolev part' of the trace norm, while we may regard the functionals $\mathscr{BN}_{S,p}$ and $\widetilde{\mathscr{BN}}_{S,p,\lambda}$ as possible variants for the role of a Besov-type seminorm in the trace space. We clarify this in Examples 6.1 and 6.2, respectively.

Now we are ready to formulate our main result, which solves Problems A and B.

Theorem 2.1.  Let $d \in [0,n]$ and $p \in (\max\{1,n-d\},\infty)$. Let $S \subset \mathbb{R}^{n}$ be a $d$-thick closed set. Let $\{\mu_{k}\}_{k \in \mathbb{N}_{0}}$ be a $d$-regular sequence of measures on $S$. Then a function $f\colon S \to \mathbb{R}$ belongs to the trace space $W_{p}^{1}(\mathbb{R}^{n})|_{S}$ if and only if for $(1,p)$-q.e. $x \in S$

Equation (2.16)

and $\widetilde{\mathscr{N}}_{S,p,\lambda}[f]<\infty$ for some $\lambda \in (0,1)$. Furthermore,

Equation (2.17)

and there exists a bounded linear extension operator $\operatorname{Ext}\colon W_{p}^{1}(\mathbb{R}^{n})|_{S} \to W_{p}^{1}(\mathbb{R}^{n})$.

Remark 2.10.  Recall Example 2.1. Consider a path connected closed set $S \subset \mathbb{R}^{2}$. It is obvious that using Theorem 2.1 we obtain an intrinsic description of the trace space of the Sobolev space $W_{p}^{1}(\mathbb{R}^{n})$ on $S$ in the full range of parameters $p \in (1,\infty)$. We would like to underline that even this particular case of Theorem 2.1 was never considered in the literature.

The results in Theorem 2.1 can be simplified in the case when either $S$ or $\mathbb{R}^{n} \setminus S$ possesses a certain 'plumpness'.

Theorem 2.2.  Let $d \in [0,n]$ and $p \in (\max\{1,n-d\},\infty)$. Let $S \subset \mathbb{R}^{n}$ be a $d$-thick closed set. Let $\{\mu_{k}\}_{k \in \mathbb{N}_{0}}$ be a $d$-regular sequence of measures on $S$. Assume that $\partial S$ is porous. Then a function $f\colon S \to \mathbb{R}$ belongs to the trace space $W_{p}^{1}(\mathbb{R}^{n})|_{S}$ if and only if for $(1,p)$-q.e. $x \in S$

Equation (2.18)

and $\mathscr{N}_{S,p}[f]<\infty$. Furthermore,

Equation (2.19)

and there exists a linear bounded extension operator $\operatorname{Ext}\colon W_{p}^{1}(\mathbb{R}^{n})|_{S} \to W_{p}^{1}(\mathbb{R}^{n})$.

§ 3. Main technical tools

The aim of this section is to bring together all the necessary technical results which are essential to the proofs of Theorems 2.1 and 2.2. As well as some very well-known facts, the section contains some new results. We split the section into several subsections for the reader's convenience.

3.1. Maximal functions and potentials

Let $F \in L^{\mathrm{loc}}_{1}(\mathbb{R}^{n})$ and $\alpha \in [0,n)$. Given $ t, s \in [0,\infty]$, we introduce the fractional maximal operator

We use the notation $\operatorname{M}_{>t}[\cdot,\alpha]:=\operatorname{M}^{< \infty}_{> t}[\cdot,\alpha]$, $\operatorname{M}[\cdot,\alpha]:=\operatorname{M}_{> 0}[\cdot,\alpha]$ and $\operatorname{M}^{<s}_{>t}[F]:=\operatorname{M}^{<s}_{>t}[F,0]$.

Remark 3.1.  Assume that $0<t' \leqslant t<s \leqslant s' \leqslant +\infty$. Then it is easy to see that for every $x \in \mathbb{R}^{n}$ and $y \in Q(x,t)$

Equation (3.1)

The following result is a very particular case of Theorem B in [30].

Theorem A.  Let $d \in [0,n]$, $\alpha \in [0,n)$, $s \in (0,+\infty]$ and $\gamma \in (1,\infty)$. Let $\mathfrak{m}$ be a Radon measure on $\mathbb{R}^{n}$ such that for some (universal) positive constant $C $

Equation (3.2)

If $\gamma\alpha \geq n-d$, then the operator $\operatorname{M}^{< s}[\cdot,\alpha]$ is bounded from $L_{\gamma}(\mathbb{R}^{n})$ into $L_{\gamma}(\mathbb{R}^{n},\mathfrak{m})$.

The following simple fact will be of use in what follows (see [31], §2.4.3, for instance, for the proof).

Proposition 3.1.  Suppose that $d \in [0,n)$. Then given a function $F \in L_{1}^{\operatorname{loc}}(\mathbb{R}^{n})$, there exists a set $E_{F} \subset \mathbb{R}^{n}$ with $\mathscr{H}^{d}(E_{F})=0$ such that for every $x \in \mathbb{R}^{n} \setminus E_{F}$

Let $\alpha \in [0,n)$. Given a function $g \in L^{\operatorname{loc}}_{1}(\mathbb{R}^{n})$, for every cube $Q$ we define the reduced Riesz potential by

Remark 3.2.  It is useful to note a simple relation between Riesz potentials and fractional maximal operators. Suppose that $\alpha \in (0,n)$. Then for every $R> 0$ and $\delta \in (0,\alpha)$

Equation (3.3)

Now we formulate an important beautiful estimate, which is a special case of one of the implications in Theorem 2.1 in [32]. In fact, in [32] classical (nonreduced) Riesz potentials were considered. Nevertheless, the corresponding proof for reduced Riesz potentials is similar (at any rate, in the case of interest to us).

Given a (nonnegative) Radon measure $\mu$ and parameters $q \in (1,\infty)$ and $\alpha \in (0,{n}/{q})$, we define the Wolf potential at the scale $R>0$ by

Equation (3.4)

Theorem B.  Let $R>0$, $q \in (1,\infty)$, $q\alpha \in (0,n)$, and let $\mu$ be a positive Radon measure on $\mathbb{R}^{n}$. Assume that $\mathscr{W}^{2R}_{\alpha,q}[\mu] \in L_{1}(\mathbb{R}^{n},\mu)$. Then there exists a positive constant $C$ (independent of $g$) such that

Equation (3.5)

for every $g \in L_{q}(\mathbb{R}^{n})$. Moreover, the least possible constant $C$ in (3.5) satisfies the inequality

Equation (3.6)

where the positive constant $b$ does not depend on $\mu$.

Recall a classical Poincaré-type inequality (see formula (7.45) in [33]).

Proposition 3.2.  Assume that $F \in W_{1}^{1,\operatorname{loc}}(\mathbb{R}^{n})$. Then for every cube $Q=Q(x,r)$, $r> 0$,

Equation (3.7)

The following estimate is well known.

Proposition 3.3.  Let $p \in (1,\infty)$ and $F \in W^{1}_{p}(\mathbb{R}^{n})$. Then for $(1,p)$-q.e. points $x \in \mathbb{R}^{n}$ (for every point in the case $p>n$) and every cube $Q=Q(y,r) \ni x$

Equation (3.8)

where the positive constant $C$ is independent of $F$, $x$ and $r$.

To prove this we use Propositions 2.4 and 3.2, and then repeat the simple arguments in the proof of Theorem 5.2 in [34] almost verbatim, with minor modifications. We omit the elementary details.

3.2. Overlappings of sets

Given a nonempty family $\{E_{\alpha}\}_{\alpha \in \mathscr{I}}$ of nonempty subsets of $\mathbb{R}^{n}$, we say that the multiplicity of overlapping of the sets $E_{\alpha}$ is finite if there exists $C>0$ such that $\sum_{\alpha \in \mathscr{I}}\chi_{E_{\alpha}}(x) \leqslant C$ for every $x \in \mathbb{R}^{n}$.

Definition 3.1.  Let $E$ be a nonempty set in $\mathbb{R}^{n}$. Let $\varepsilon>0$, and let $\{x_{j}\}_{j \in \mathscr{J}}$, $\mathscr{J} \subset \mathbb{N}$, be a subset of $E$ with the following properties:

(i) $\|x_{i}-x_{j}\|_{\infty} \geq \varepsilon$ for every $i,j \in \mathscr{J}$, $i \neq j$;

(ii) for every $x \in E \setminus \{x_{j}\}_{j \in \mathscr{J}}$ there is a point $x_{j}$ such that $\|x- x_{j}\|_{\infty}<\varepsilon$.

We call the set $\{x_{j}\}_{j \in \mathscr{J}}$ a maximal $\varepsilon$-separated subset of $E$.

The following propositions will be used often in what follows. We omit the elementary proofs.

Proposition 3.4.  Let $\{Q_{j}\}_{j \in \mathscr{J}}$ be a family of pairwise disjoint cubes with the same edge length. Then for every $c \geq 1$ there is a positive constant $C=C(n,c)$ such that the multiplicity of overlapping of the cubes $cQ_{j}$ is finite and bounded above by $C$.

Proposition 3.5.  Let $\mathfrak{m}$ be a finite Borel measure on $\mathbb{R}^{n}$. Let $\{E_{j}\}_{j \in \mathscr{J}}$ be a family of Borel subsets of $\mathbb{R}^{n}$ such that the multiplicity of overlapping of the sets $E_{j}$ is finite and bounded above by some constant $N \in \mathbb{N}$. Then

Equation (3.9)

The following elementary observation is a direct consequence of Definition 3.1.

Lemma 3.1.  Let $E$ be a nonempty subset of $\mathbb{R}^{n}$. Let $\varepsilon > 0$ and let $\{x_{j}\}_{j \in \mathscr{J}}$ be a maximal $\varepsilon$-separated subset of $E$. Then

(1) $E \subset \bigcup_{j \in \mathscr{J}}Q(x_{j},\varepsilon)$;

(2) the family $\{Q(x_{j},\varepsilon/2)\}_{j \in \mathscr{J}}$ is pairwise disjoint;

(3) every point $x \in E$ belongs to at most $3^{n}$ cubes in $\{Q(x_{j},\varepsilon)\}_{j \in \mathscr{J}}$.

3.3. The Whitney decomposition

Recall that we measure the distances in $\mathbb{R}^{n}$ in the uniform norm $\|\cdot\|_{\infty}$. For a cube $Q \subset \mathbb{R}^{n}$ we set $Q^{\ast}:=(9/8)Q$. Recall that, unless otherwise stated, all cubes are assumed to be closed.

The following result is a slight modification of the Classical Whitney Decomposition Lemma. Its proof repeats the proof of Theorem 1 in [29], Ch. 6, with minor changes.

Lemma 3.2.  For each closed nonempty set $S\subset\mathbb{R}^{n}$ there exists a family of closed dyadic cubes $W_{S}=\{Q_{\varkappa}\}_{\varkappa \in I}=\{Q(x_{\varkappa},r_{\varkappa})\}_{\varkappa \in I}$ with the following properties:

(1) $\mathbb{R}^{n} \setminus S= \bigcup_{\varkappa \in I} Q_{\varkappa}$;

(2) for each $\varkappa \in I$

Equation (3.10)

(3) the following inequalities hold:

Equation (3.11)

(4) for each index $\varkappa \in I$ there exist at most $C(n)$ indices $\varkappa'$ such that $Q^{\ast}_{\varkappa} \cap Q^{\ast}_{\varkappa'} \neq \varnothing$;

(5) $\operatorname{int}Q_{\varkappa} \cap \operatorname{int}Q_{\varkappa'}=\varnothing$ for every $\varkappa,\varkappa' \in I$, $\varkappa \neq \varkappa'$, and $Q^{\ast}_{\varkappa} \cap Q^{\ast}_{\varkappa'} \neq \varnothing$ if and only if $Q_{\varkappa} \cap Q_{\varkappa'} \neq \varnothing$.

The family of cubes $W_{S}\!=\!\{Q_{\varkappa}\}_{\varkappa \in I}\!=\!\{Q(x_{\varkappa},r_{\varkappa})\}_{\varkappa \in I}$, constructed in Lemma 3.2 is called a Whitney decomposition of the open set $\mathbb{R}^{n} \setminus S$, and the cubes $Q_{\varkappa}$ are called Whitney cubes. In what follows we also need the part of the Whitney decomposition comprised of the cubes of small edge length. More precisely, we set $\mathscr{W}_{S}=\{Q_{\varkappa}\}_{\varkappa \in \mathscr{I}}$, where $\mathscr{I}:=\{\varkappa \in I \colon r_{\varkappa} \leqslant 1\}$.

The following notation is useful below. Given a closed set $S$, for every $\varkappa \in I$ set

Equation (3.12)

We call a cube $Q_{\varkappa'}$ neighbouring to a cube $Q_{\varkappa}$ if $\varkappa' \in b(Q_{\varkappa})$. Similarly, set $b(x):=\{\varkappa \in I\colon Q^{\ast}_{\varkappa} \ni x\}$ for every $x \in \mathbb{R}^{n} \setminus S$.

To construct our extension operator we use the following (see [20], Ch. 6, §1.3, for details).

Proposition 3.6.  Let $S \subset \mathbb{R}^{n}$ be a closed nonempty set and let $\{Q_{\varkappa}\}_{\varkappa \in I}$ be the Whitney decomposition of the open set $\mathbb{R}^{n} \setminus S$ constructed in Lemma 3.2. Then there exists a family of functions $\{\varphi_{\varkappa}\}_{\varkappa \in I}$ with the following properties:

(1) $\varphi_{\varkappa} \in C^{\infty}_{0}(\mathbb{R}^{n} \setminus S)$ for every $\varkappa \in I$;

(2) $0 \leqslant \varphi_{\varkappa} \leqslant 1$ and $\operatorname{supp}\varphi_{\varkappa} \subset (Q_{\varkappa})^{\ast}:=({9}/{8})Q_{\varkappa}$ for every $\varkappa \in I$;

(3) $\sum_{\varkappa \in I}\varphi_{\varkappa}(x)=1$ for all $x \in \mathbb{R}^{n} \setminus S$;

(4) $\|D^{\alpha} \varphi_{\varkappa}|L_{\infty}(\mathbb{R}^{n})\| \leqslant C(\operatorname{diam}Q_{\varkappa})^{-|\alpha|}$ for every multi-index $\alpha \in \mathbb{N}^{n}_{0}$ and every $\varkappa \in I$, where the positive constant $C$ depends only on $n$.

Definition 3.2.  Given a closed nonempty set $S \subset \mathbb{R}^{n}$ and $x \notin S$, we say that $\widetilde{x}$ is a nearest point to $x$ or a metric projection of $x$ onto $S$ whenever $\operatorname{dist}(x,S)= \operatorname{dist}(x,\widetilde{x})$.

Remark 3.3.  Let $\widetilde{x}$ be a metric projection of $x \in \mathbb{R}^{n} \setminus S$ onto $S$. Consider the line interval

Consider an arbitrary $r \in (0,\|x-\widetilde{x}\|)$ and a point $y_{r}=\partial Q(\widetilde{x},r) \cap [x,\widetilde{x}]$. We show that $\operatorname{dist}(y_{r},S)=\|y_{r}-\widetilde{x}\|_{\infty}=r$.

Clearly, $\operatorname{dist}(y_{r},S) \leqslant r$ because $y_{r} \in \partial Q (\widetilde{x},r)$. Assume that $\operatorname{dist}(y_{r},S)<r$. Then there is a point $y' \in S$ such that $\|y_{r}-y'\|_{\infty}<r= \|y_{r}-\widetilde{x}\|_{\infty}$. Using this and the equality $\|x-\widetilde{x}\|_{\infty}=\|x-y_{r}\|_{\infty}+\|y_{r}-\widetilde{x}\|_{\infty}$ we obtain $\operatorname{dist}(x,S) \leqslant \|x-y'\|_{\infty} \leqslant \|x-y_{r}\|_{\infty}+\|y_{r}-y'\|_{\infty}< \|x-\widetilde{x}\|_{\infty}$. This contradicts the fact that $\|\widetilde{x}-x\|_{\infty}=\operatorname{dist}(x,S)$.

Definition 3.3.  Fix a closed nonempty set $S$. For a cube $Q=Q(x,r) \subset \mathbb{R}^{n}$ with $x \notin S$ we call $\widetilde{Q}=Q(\widetilde{x},r)$ a reflected cube, where $\widetilde{x}$ is a metric projection of $x$ onto $S$.

Remark 3.4.  Clearly, a metric projection onto a closed nonempty set exists. It is not unique in general. We will specify an algorithm for choosing $\widetilde{x}$ only when our constructions require this. Otherwise, given a cube $Q(x,r)$, we fix any metric projection $\widetilde{x}$ and the cube $\widetilde{Q}(\widetilde{x},r)$.

Lemma 3.3.  Let $S \subset \mathbb{R}^{n}$ be a closed nonempty set and let $W_{S}=\{Q_{\varkappa}\}_{\varkappa \in I}$ be a Whitney decomposition of $\mathbb{R}^{n} \setminus S$. Then for every $c>0$ there exists a positive constant $C=C(n,c)$ such that

Proof.  Suppose that $Q(\widetilde{x}_{\varkappa},c r_{\varkappa}) \mkern1mu\cap\mkern1mu Q(\widetilde{x}_{\varkappa'},c r_{\varkappa'})\mkern-1mu \neq \mkern-1mu \varnothing$ for some $\varkappa,\varkappa'\mkern-1mu \in\mkern-1mu I$ with ${r_{\varkappa}\mkern-1mu =\mkern-1mu r_{\varkappa'}}$. In view of (3.10) we have $\operatorname{dist}(Q_{\varkappa}, \widetilde{x}_{\varkappa}) \leqslant 4 \operatorname{diam}(Q_{\varkappa})$ and $\operatorname{dist}(Q_{\varkappa'}, \widetilde{x}_{\varkappa'}) \leqslant 4 \operatorname{diam}(Q_{\varkappa'})$. Hence $\operatorname{dist}(Q_{\varkappa},Q_{\varkappa'}) \leqslant (8+c) \operatorname{diam}(Q_{\varkappa})$. This implies that $Q_{\varkappa'} \subset (19+2c)Q_{\varkappa}$. Then Lemma 3.2, (5), and arguments based on volume estimates give

The proof is complete.

Lemma 3.4.  Let $S\subset \mathbb{R}^{n}$ be a closed nonempty set and let $W_{S}=\{Q_{\varkappa}\}_{\varkappa \in I}$ be a Whitney decomposition of $\mathbb{R}^{n} \setminus S$. Let $\mathfrak{m}$ be a finite Borel measure with $\operatorname{supp}\mathfrak{m} \subset S$. Then for every $c \geq 1$

where the positive constant $C$ depends only on $c$ and $n$.

Proof.  Consider the family of cubes $\{Q(\widetilde{x}_{\varkappa},c)\}_{\varkappa \in \mathscr{I}}$. Using Vitali's covering theorem (see §1.5 of [3] for details) we find an index set $\widehat{\mathscr{I}} \subset \mathscr{I}$ such that all cubes in the family $\{Q(\widetilde{x}_{\varkappa},c)\}_{\varkappa \in \widehat{\mathscr{I}}}$ are mutually disjoint and

Equation (3.13)

Note that if $Q(\widetilde{x}_{\varkappa'},r_{\varkappa'}) \cap Q(\widetilde{x}_{\varkappa},5c) \neq \varnothing$ for some $\varkappa, \varkappa' \in \mathscr{I}$, then

Equation (3.14)

because $c \geq 1$ and $r_{\varkappa'} \leqslant 1$. From this and (3.10) it follows that $Q_{\varkappa'} \subset Q(x_{\varkappa}, 20c)$. Hence, using Lemma 3.2, (5), we obtain

Using this fact, (3.13), (3.14) and Propositions 3.4 and 3.5 we obtain the required estimate

The proof is complete.

Recall Definition 2.1 and Remark 2.1. The following result is a minor modification of Theorem 2.4 in [23] and can be proved analogously.

Theorem C.  Let $S$ be a closed Ahlfors $n$-regular set in $\mathbb{R}^{n}$, and let $W_{S}=\{Q_{\varkappa}\}_{\varkappa \in I}$ be a Whitney decomposition of $\mathbb{R}^{n} \setminus S$. Then there exists a family $\mathfrak{U}:=\{\mathscr{U}_{\varkappa} \colon \varkappa \in \mathscr{I}\}$ of Borel sets with the following properties:

(1) $\mathscr{U}_{\varkappa} \subset (10Q_{\varkappa}) \cap S$ for all $\varkappa \in \mathscr{I}$;

(2) $\mathscr{H}^{n}(Q_{\varkappa}) \leqslant \kappa_{1}\mathscr{H}^{n}(\mathscr{U}_{\varkappa})$ for all $\varkappa \in \mathscr{I}$;

(3) $\sum_{\varkappa \in \mathscr{I}}\chi_{\mathscr{U}_{\varkappa}}(x) \leqslant \kappa_{2}$ for $x \in S$.

The positive constants $\kappa_{1}$ and $\kappa_{2}$ depend only on $n$, and the constants $c^{n}_{1}(S)$ and $c^{n}_{2}(S)$.

3.4. $d$-regular sequences of measures

The following result is a version of Frostman's theorem adapted for our purposes (cf. Theorem 5.1.12 in [26]). For the reader's convenience we present a detailed proof.

Theorem 3.1.  Let $S$ be a closed nonempty subset of $\mathbb{R}^{n}$. Then given $d \in [0,n]$, there exists a sequence of Borel measures $\{\nu_{k}\}_{k \in \mathbb{N}_{0}}$ with $\operatorname{supp}\nu_{k}=S$, $k \in \mathbb{N}_{0}$, such that for every $k \in \mathbb{N}_{0}$ the following properties hold:

(1)

Equation (3.15)

(2) for every finite index set $\mathscr{A} \subset \mathbb{Z}^{n}$ let $V^{k}_{\mathscr{A}}:=\bigcup_{m \in \mathscr{A}} Q_{k,m}$; then

Equation (3.16)

(3) there exists a function $\gamma_{k} \in L_{\infty}(S,\nu_{0})$ such that $\nu_{k}=\gamma_{k}\nu_{0}$ and

Equation (3.17)

Proof.  Fix a nonnegative integer $k$ and let $\nu^{k,0}:=\nu^{k}$ be a measure with constant density that has mass $2^{-kd}$ on each $Q_{k,m}$ that intersects $S$. We now modify $\nu^{k}$ in the following way. If $\nu^{k}(Q_{k-1,m})>2^{-(k-1)d}$ for some $Q_{k-1,m} \in \mathscr{Q}_{k-1}$, we reduce its mass uniformly on $Q_{k-1,m}$ until it becomes $2^{-(k-1)d}$. On the other hand, if $\nu^{k}(Q_{k-1,m}) \leqslant 2^{-(k-1)d}$, we leave $\nu^{k}$ unchanged on $Q_{k-1,m}$. In this way we obtain a new measure $\nu^{k,1}$. Using the fact that every cube $Q_{k-1,m}$ which has nonempty intersection with $S$ contains $\leqslant 2^{n}$ cubes $Q_{k,m'}$ with the property $Q_{k,m'} \cap S \neq \varnothing$ we have

We repeat this procedure for $\nu^{k,1}$, obtaining $\nu^{k,2}$, and after $k$ steps we obtain $\nu^{k,k}$. It follows from this construction that

Equation (3.18)

for every $j=0,1,\dots,k$ and every dyadic cube $Q_{i,m} \in \mathscr{Q}_{i}$, where $i=j,\dots,k$. Furthermore, it is clear that

Equation (3.19)

Using (3.18) it is easy to see that for every $j \in \mathbb{N}_{0}$ the sequence $\{\nu^{k,k-j}(E)\}_{k \geq j}$ is bounded for every compact subset $E$ of $ S$. Then $\{\nu^{k,k-j}\}_{k \geq j}$ has a subsequence that converges weakly to $\nu_{j}$ (see [31], §1.9, Theorem 2), and clearly $\operatorname{supp}\nu_{j} \subset S$ (recall that $S$ is closed).

Fix an arbitrary $j \in \mathbb{N}$ and an arbitrary Borel set $G\subset S$. We compare $\nu_{j}(G)$ and $\nu_{j-1}(G)$. First note that, according to our construction, for every dyadic cube $Q_{k,m}$ we have

Let $C_{c}(\mathbb{R}^{n})$ be the set of continuous functions $f\colon \mathbb{R}^{n} \to \mathbb{R}$ with compact support. For every nonnegative function $f \in C_{c}(\mathbb{R}^{n})$ we have

Equation (3.20)

Fix an arbitrary nonnegative $f\,{\in}\,C_{c}(\mathbb{R}^{n})$. Choosing an appropriate subsequence if necessary and passing to the limit in (3.20) we obtain

Equation (3.21)

Using the Borel regularity of the measures $\nu_{j}$ and the Radon-Nikodym theorem and taking (3.21) into account we obtain (3.17).

We show that $\nu_{j}(Q_{i,m}) \leqslant 3^{n}2^{-id}$ for every $i,j \in \mathbb{N}_{0}$, $i \geq j$, and every dyadic cube $Q_{i,m} \in \mathscr{Q}_{i}$. Indeed, if $f_{i,m} \in C_{0}^{\infty}(\mathbb{R}^{n})$ is such that $\chi_{Q_{i,m}} \leqslant f_{i,m} \leqslant \chi_{3Q_{i,m}}$, then (3.18) yields

Hence, using the fact that every closed cube $Q(x,r)$ with $x \in \mathbb{R}^{n}$ and $r \in (0,2^{-k}]$ has nonempty intersection with $\leq 5^{n}$ dyadic cubes $Q_{k(r),m}$, where $k(r):=|[\log_{2}r]|$, we obtain (3.15).

Fix an arbitrary nonempty index set $\mathscr{A} \subset \mathbb{Z}^{n}$ and $k \in \mathbb{N}_{0}$ and fix an arbitrary $l \in \mathbb{N}$, $l \geq k$. The key observation, which follows directly from our construction, is that every $x \in V^{k}_{\mathscr{A}} \cap S$ belongs to some dyadic cube $Q^{(j)} \in \mathscr{Q}_{n_{j}}$, $k \leqslant n_{j} \leqslant l$, (or several cubes) such that $\nu^{l,l-k}(Q^{(j)})=2^{-n_{j}d}$. We can choose a disjoint covering consisting of maximal dyadic cubes with this property, so that $S \cap V^{k}_{\mathscr{A}} \subset \bigcup_{j}Q^{(j)}$. This gives

where the infimum is taken over all finite or countable coverings of $V^{k}_{\mathscr{A}} \cap S$ with dyadic cubes $Q^{(i)} \in \bigcup_{l \geq k}\mathscr{Q}_{l}$. The right-hand side is independent of $l$. Combining this with the definition of the $d$-Hausdorff content we note that $\overline{V^{k}_{\mathscr{A}}}$ is a compact set. This gives

Equation (3.22)

This completes the proof.

The following result shows that Definition 2.3 is consistent.

Corollary 3.1.  Let $d \in [0,n]$ and let $S \subset \mathbb{R}^{n}$ be a $d$-thick closed set. Then there exists a $d$-regular sequence of measures on $S$.

Proof.  We apply Theorem 3.1 to the set $S$. This gives a sequence of Borel measures $\{\nu_{k}\}_{k \in \mathbb{N}_{0}}$ with $\operatorname{supp}\nu_{k}=S$ satisfying (3.15)–(3.17). We set $\mu_{k}:=15^{-n}\nu_{k}$ for every $k \in \mathbb{N}_{0}$.

It is sufficient to verify (2.11). Fix some $x \in S$, and let $\mathscr{A} \subset \mathbb{Z}^{n}$ be the index set such that $m \in \mathscr{A}$ if and only if $Q_{k+2,m} \cap Q(x,2^{-k-2}) \neq \varnothing$. It is clear that $\overline{V^{k+2}_{\mathscr{A}}}:=\bigcup_{m \in \mathscr{A}}\overline{Q}_{k+2,m} \subset Q(x,2^{-k})$ and $Q(x,2^{-k-2}) \subset V^{k+2}_{\mathscr{A}}$. Hence, using Definition 2.2, estimates (3.16) and (3.17) and the monotonicity of the $\mathscr{H}^{d}_{\infty}$-content we obtain

This completes the verification of the corollary.

The following lemma gives some asymptotic estimates for measures in a fixed $d$-regular sequence. Recall Remark 2.2.

Lemma 3.5.  Let $d \in [0,n]$ and let $S$ be a $d$-thick closed set. Let $\{\mu_{k}\}_{k \in \mathbb{N}_{0}}$ be a $d$-regular sequence of measures on $S$. Then for each $c \geq 1$ and every $k>\operatorname{log}_{2}c$

Equation (3.23)

Proof.  Fix $j \in \mathbb{N}_{0}$ such that $c \in [2^{j},2^{j+1})$. It follows from estimates (2.10)–(2.12) that for every $k>\operatorname{log}_{2}c$ and every $x \in S$

Equation (3.24)

Similarly,

Equation (3.25)

The required estimate (3.23) follows from (3.24) and (3.25), which completes the proof.

Remark 3.5.  Recall that a Borel measure $\mu$ on a metric space $(X,d)$ is called a doubling measure if there exists a constant $C_{\mu} \geq 1$ such that $\mu(B(x,2r)) \leqslant C_{\mu} \mu(B(x,r))$ for all $x \in X$ and $r>0$. It is very important to note that the estimate (3.23) does not imply the doubling property of the measures $\mu_{k}$, $k \in \mathbb{N}_{0}$. Roughly speaking, the point is that, given $k \in \mathbb{N}_{0}$, in Lemma 3.5 we compare $\mu_{k}(Q(x,r))$, where $x \in S$ and $r \in (0,1]$, only with $\mu_{k}(Q(x,2^{-k}))$. If we try to compare $\mu_{k}(Q(x,r))$ and $\mu_{k}(Q(x,r/2))$ for $k \gg |\log_{2}r|$, then we obtain a bad estimate, with the corresponding positive constant $C$ depending heavily on $k$.

Lemma 3.6.  Let $d \in [0,n]$ and let $S$ be a $d$-thick closed set. Let $\{\mu_{k}\}_{k \in \mathbb{N}_{0}}$ be a $d$-regular sequence of measures on $S$. Let $E$ be a Borel subset of $S$. If $\mathscr{H}^{d}(E)=0$, then $\mu_{k}(E)=0$ for every $k \in \mathbb{N}_{0}$. The converse is false.

Proof.  Fix $E\subset S$ with $\mathscr{H}^{d}(E)=0$. Using (2.10) and the definition of Hausdorff measure it is easy to see that $\mu_{k}(E)=0$ for every $k \in \mathbb{N}_{0}$.

To prove that the converse is false we use the construction from Example 6.3 below. More precisely, in Example 6.3 we build a 1-thick path-connected set $S \subset \mathbb{R}^{n}$ with $\operatorname{dim}_{H}S=n$ and a $1$-regular sequence of measures $\{\mu_{k}\}_{k \in \mathbb{N}_{0}}$ on $S$ such that every $\mu_{k}$ is absolutely continuous with respect to $\mathscr{H}^{n}$. Hence for every smooth curve $\gamma \in S$ with $\mathscr{H}^{1}(\gamma)>0$ we obtain $\mu_{k}(\gamma)=0$ for all $k \in \mathbb{N}$. The proof is complete.

Recall that for every $r>0$ we set $k(r):=|[\log_{2}r]|$. The following theorem will be an important technical tool in the sequel. Recall Remark 2.2.

Theorem 3.2.  Let $d \in [0,n]$ and let $S$ be a $d$-thick closed set. Let $\{\mu_{k}\}_{k \in \mathbb{N}_{0}}$ be a $d$-regular sequence of measures on $S$. Then for every $r \in (0,1)$, $x \in S$ and every Borel set $G \subset Q(x,r) \cap S$

Equation (3.26)

The positive constant $C$ depends only on $n$ and $C_{\{\mu_{k}\}}$ in Remark 2.2.

Proof.  Fix $x,y \in S$, and $t$ and $r$, $0<t<r<1$, such that $Q(y,t) \subset Q(x,r)$. It is clear that

Equation (3.27)

On the other hand, using (2.10)–(2.12) (we can use these estimates because $x,y \in S$) we have

Equation (3.28)

Combining (3.27) and (3.28) we obtain

Equation (3.29)

Hence we obtain (3.26) for $G=Q(y,t)$.

Fix $r \in (0,1)$. For every $j \in \mathbb{N}$, $j>j_{0}:= [{2r}/{(1-r)}]+1$, let $\{x^{j}_{i}\}_{i=1}^{N(j)}$ be a maximal $({r}/{j})$-separated subset of $Q(x,r) \cap S$. Clearly, $Q(x,r)\cap S \subset \bigcup_{i}\operatorname{int}Q(x^{j}_{i},{2r}/{j})$ and the cubes $Q(x^{j}_{i},{r}/(2j))$ are pairwise disjoint. For every $j \in \mathbb{N}$ take an arbitrary nonempty set $\mathscr{A}^{j} \subset \{1,\dots,N(j)\}$ and consider the set $U_{j}:=\bigcup_{i \in \mathscr{A}^{j}}\operatorname{int}Q(x^{j}_{i},{2r}/{j})$. It is clear that $U_{j} \subset Q(x,r+{2r}/{j_{0}}) \subset Q(x,1)$ for every $j>j_{0}$. We use this inclusion, (3.29), Propositions 3.4 and 3.5, and (2.12); then for every $j>j_{0}$ we obtain

Equation (3.30)

Fix an arbitrary compact set $K\subset Q(x,r) \cap S$. Now using the $\sigma$-additivity of the measures $\mathscr{H}^{n}$ and $\mu_{k(r)}$ we have

Equation (3.31)

Combining (3.31) and (3.30), we obtain (3.26) for every compact set $K\subset Q(x,r) \cap S$. To establish (3.26) for a general Borel set $G \subset S$ it remains to recall that the measures $\mu_{k}$, $k \in \mathbb{N}$, and $\mathscr{H}^{n}$ are Radon measures. The proof is complete.

Corollary 3.2.  Let $d \in [0,n]$ and let $S$ be a $d$-thick closed set. Let $\{\mu_{k}\}_{k \in \mathbb{N}_{0}}$ be a $d$-regular sequence of measures on $S$. Assume that $f \in L_{1}^{\operatorname{loc}}(S,\mu_{k})$ for some (and hence every) $k \in \mathbb{N}_{0}$. Then there exists $C>0$ (independent of $f$, $x$ and $r$) such that

Equation (3.32)

Proof.  The estimate (3.32) clearly holds for a simple function $f\colon S \to \mathbb{R}$ due to Theorem 3.2. In the general case we have to construct an increasing sequence of simple functions converging to $|f|$ and use the monotone convergence theorem for integrals (see §1.3 of [31]). This completes the verification of the corollary.

Lemma 3.7.  Let $d \in [0,n]$ and let $S$ be a $d$-thick closed set. Let $\{\mu_{k}\}_{k \in \mathbb{N}_{0}}$ be a $d$-regular sequence of measures on $S$. Assume that $g \in L_{1}^{\operatorname{loc}}(\mathbb{R}^{n},\mu_{k})$ for some (and hence every) $k \in \mathbb{N}_{0}$. Let $c \geq 1$, $x \in S$, $x' \in \mathbb{R}^{n}$ and $r \in (0,1)$ satisfy $cr \in (0,1)$ and $Q(x,r) \subset Q(x',cr)$. Then

Equation (3.33)

where the positive constant $C$ does not depend on $x$, $x'$, $r$ or $g$.

Proof.  Clearly, $Q(x,r) \subset Q(x',cr) \subset Q(x,2cr)$. Suppose $cr \in (0,1/2)$. Using this we obtain $\mu_{k(r)}(Q(x',cr)) \leqslant \mu_{k(r)}(Q(x,2cr)) \leqslant C^{-1}_{\{\mu_{k}\}}(4c)^{n} \mu_{k(r)}(Q(x,r))$ by Lemma 3.5, which obviously implies the first inequality in (3.33). When $cr \in (1/2,1)$ the corresponding estimate easily follows from (2.11) and (2.12).

The second estimate in (3.33) follows immediately from (2.10)–(2.12). The proof is complete.

3.5. Calderón-type maximal functions

Recall Definition 2.5 and also that $k(r):=-[\log_{2}r]$ for every $r>0$.

Lemma 3.8.  Let $d \in [0,n]$ and let $S$ be a $d$-thick closed set. Let $\{\mu_{k}\}_{k \in \mathbb{N}_{0}}$ be a $d$-regular sequence of measures on $S$. Assume that $f \in L_{1}^{\rm loc}(\mathbb{R}^{n},\mu_{k})$ for every $k \in \mathbb{N}_{0}$. Let $c \geq 1$, $x \in S$, $x' \in \mathbb{R}^{n}$ and $r \in (0,1)$ be such that $Q(x,r) \subset Q(x',cr)$. Then

Equation (3.34)

The positive constant $C$ in (3.34) depends on the constant $c$ but does not depend on $x,x',r$ or $f$.

Proof.  Fix $t \in (r,1]$. If $tc<1$, then we use the inclusion $Q(x,t) \subset Q(x',ct)$, Lemma 3.7, Remark 2.4 and the monotonicity of $f^{\sharp}_{S}(x',t)$ with respect to $t$. We obtain

Equation (3.35)

Now consider the case $tc \geq 1$. We use Remark 2.4 and Lemma 3.7 and note that $k(t) \leqslant |[\log_{2}c]|+1$. This gives

Equation (3.36)

Now (3.34) follows directly from Definition 2.5 and estimates (3.35) and (3.36). The proof is complete.

Lemma 3.9.  Let $d \in [0,n]$ and let $S$ be a $d$-thick closed set. Let $\{\mu_{k}\}_{k \in \mathbb{N}_{0}}$ be a $d$-regular sequence of measures on $S$. Assume that $f \in L_{1}^{\operatorname{loc}}(\mathbb{R}^{n},\mu_{k})$ for some (and hence every) $k \in \mathbb{N}_{0}$. Let $0<r<t<1$, $x \in S$ and $x' \in \mathbb{R}^{n}$ be such that $Q(x,r) \subset Q(x',t)$ and $R=\|x-x'\|+t<1$. Then

Equation (3.37)

where the positive constant $C>0$ does not depend on $x,x',r,t$ or $f$.

Proof.  Clearly, $Q(x,t)\subset Q(x',R) =: Q'$ and $Q(x',t) \subset Q(x,R) = Q$. Since $Q(x,r) \subset Q(x',t)$ we have $\|x-x'\|+r < t$ and hence $R \leqslant 2t$. Arguing as in (3.35) and using Remark 2.4 we obtain

(note that $R \leqslant 2t$). Clearly, similar inequalities hold true with $Q'$ replaced by $Q$. Hence, we get

Equation (3.38)

Let $j_{0}:=[\operatorname{log}_{2}(t/r)]$. Arguing as in (3.38) and using Remark 2.4 we have

Equation (3.39)

Combining (3.38) and (3.39) we obtain (3.37), which completes the proof.

Theorem 3.3.  Let $d \in [0,n]$ and let $S$ be a $d$-thick closed set. Let $\{\mu_{k}\}_{k \in \mathbb{N}_{0}}$ be a $d$-regular sequence of measures on $S$. Let $f \in L_{1}^{\rm loc}(S,\mu_{k})$ for some (and hence every) $k \in \mathbb{N}_{0}$. Then there exists a positive constant $C$ such that for any $x,x' \in S$, $r,t \in (0,1)$ with $\widetilde{R}:=\max\{r,t\}+\|x-x'\|<1/2$

Equation (3.40)

Proof.  Note that $Q(x,r) \subset Q(x',\widetilde{R})$ and $Q(x',t) \subset Q(x',\widetilde{R})$. Using the triangle inequality, (3.37) and the monotonicity of $f^{\sharp}_{\{\mu_{k}\}}(x,t)$ with respect to $t$ we see that the left-hand side of (3.40) is bounded above by

Equation (3.41)

here we have set $R=\widetilde{R}+\|x-x'\| \leqslant 2\widetilde{R}<1$. This completes the proof.

Corollary 3.3.  Let $d \in [0,n]$ and let $S$ be a $d$-thick closed set. Let $\{\mu_{k}\}_{k \in \mathbb{N}_{0}}$ be a $d$-regular sequence of measures on $S$. Let $f \in L_{1}^{\rm loc}(S,\mu_{k})$ for some (and hence every) $k \in \mathbb{N}_{0}$. Let $x,x' \in S$ be such that $\|x-x'\|<1/2$ and let

Equation (3.42)

Then there exists $C>0$ (independent of $x$, $x'$ and $f$) such that

Equation (3.43)

Proof.  From Theorem 3.3, using the monotonicity of $f^{\sharp}_{\{\mu_{k}\}}(x,t)$ with respect to $t$ we obtain the required estimate

Equation (3.44)

This completes the verification of the corollary.

The following result is crucial in proving the 'direct trace theorem'. Recall Proposition 2.4 and Definitions 2.7 and 2.8.

Theorem 3.4.  Let $d \in [0,n]$ and let $S$ be a $d$-thick closed set. Let $\{\mu_{k}\}_{k \in \mathbb{N}_{0}}$ be a $d$-regular sequence of measures on $S$. Let $q \in (\max\{1,n-d\},\infty)$ and $F \in W_{q}^{1}(\mathbb{R}^{n})$. Then for every cube $Q=Q(x,r)$ with $x \in S$ and $r \in (0,1]$

Equation (3.45)

where the positive constant $C$ does not depend on $x$, $r$ or $F$.

Proof.  Fix a cube $Q=Q(x,r)$ with $x \in S$ and $r \in (0,1]$. We consider several cases.

Case 1. Assume that $q>n$. Using the well-known Sobolev embedding theorem (see [10], §1.8.2) we obtain the estimate required in this case:

Here $\widehat{F}$ is the continuous representative of $F$ and $C>0$ depends only on $n$ and $q$.

Case 2. Now we consider the most complicated case, when $d>0$ and $q \in (\max\{1,n-d\},n)$. We set $g:=\chi_{Q}|\nabla F|$. It is clear that $Q \cap S \subset Q(y,2r)$ for every $y \in Q \cap S$. We can rewrite (3.8) as follows. For $(1,p)$-quasi-every (and hence for $\mu_{k(r)}$-almost every) $y \in Q \cap S$

Equation (3.46)

Set $\mu_{Q}:=\mu_{k(r)}\lfloor_{{Q \cap S}}$. Since $Q=Q(x,r)$ lies in a union of at most $5^{n}$ cubes with edge length $2^{-k(r)}$, it follows from (2.10) that $\mu_{Q}(Q(y,t)) \leqslant 5^{n}t^{d}$ for $t \in (0,4r)$ and $y \in Q$. This gives (recall the definition of the Wolf potential (3.4))

Equation (3.47)

The positive constant $C$ in (3.47) does not depend on $x$ or $r$.

Now we recall Lemma 3.6 and apply Theorem B with measure $\mu_{Q}$ (instead of $\mu$) and with $\alpha=1$. Hence, using (2.11), (3.46) and (3.47) we obtain

Equation (3.48)

Case 3. In the case $d>0$, $q=n$ we choose an arbitrary $\widetilde{q} \in (\max\{1,n-d\},n)$ and use the previous step to obtain (3.45) with $\widetilde{q}$ instead of $q$. To complete the proof it remains to apply Hölder's inequality.

Corollary 3.4.  Let $d \in [0,n]$ and let $S$ be a $d$-thick closed set. Assume that $q \in (\max\{1,n-d\},\infty)$ and $F \in W_{q}^{1}(\mathbb{R}^{n})$ and set $f:=F|_{S}$. Then for every $r \in [0,1]$ and every $x \in S$

Equation (3.49)

Proof.  Consider the case $r<1$. The case $r=1$ is similar. Using Remark 2.4 and Theorem 3.4, for every $x \in S$ we have the required estimate

Equation (3.50)

The verification of the corollary is complete.

3.6. Porous sets

Recall Lemma 3.2 and Definitions 2.6 and 3.2. Also recall that we let $k(\varkappa)$ denote the unique integer such that $r_{\varkappa}= 2^{-k(\varkappa)}$. We continue to measure distances in $\mathbb{R}^{n}$ in the $\|\cdot\|_{\infty}$-norm.

Lemma 3.10.  Let $S$ be a closed nonempty set in $\mathbb{R}^{n}$. Let $Q_{\varkappa}=Q(x_{\varkappa},r_{\varkappa})$ be a Whitney cube in $W_{S}$. Then $\widetilde{x}_{\varkappa} \in S_{j}(\lambda)$ for every $j \geq k(\varkappa)$ and $\lambda \in (0,1)$. Furthermore, $Q(\widetilde{x}_{\varkappa},{r_{\varkappa}(c-1)}/{c}) \cap S \subset S_{k(\varkappa)}(\lambda)$ for every $c >1$ and every ${\lambda \in (0,{1}/{c}]}$.

Proof.  Consider the interval $(x_{\varkappa},\widetilde{x}_{\varkappa}) :=\{x=x_{\varkappa}+t(\widetilde{x}_{\varkappa}-x_{\varkappa})\colon t \in (0,1)\}$. Clearly, $S \cap (x_{\varkappa},\widetilde{x}_{\varkappa})=\varnothing$ because otherwise there exists a point $x' \in S$ such that ${\|x_{\varkappa}-x'\|<\|x_{\varkappa}-\widetilde{x}_{\varkappa}\|=\operatorname{dist}(x_{\varkappa},S)}$. For every $r \in (0,r_{\varkappa}]$ consider the point $y_{r}:=(x_{\varkappa},\widetilde{x}_{\varkappa}) \cap \partial Q(\widetilde{x}_{\varkappa},r)$. It follows from Remark 3.3 that $\operatorname{dist}(y_{r},S)=r$. Hence for every $\lambda \in (0,1)$ the cube $Q(y_{r},\lambda r)$ lies in $\mathbb{R}^{n} \setminus S$. This proves the first claim in the lemma.

Given a number $c >1$ we set $r_{c}:={r_{\varkappa}}/{c}$. Then from Remark 3.3 we conclude that $\operatorname{dist}(y_{r_{c}},S)={r_{\varkappa}}/{c}$. On the other hand it is clear that $y_{r_{c}} \in Q(x,r_{\varkappa})$ for every $x \in Q(\widetilde{x}_{\varkappa},r_{\varkappa}(c-1)/c)$. This proves the second claim.

Lemma 3.11.  Let $S$ be a closed nonempty set in $\mathbb{R}^{n}$. Let $W_{S}=\{Q_{\varkappa}\}_{\varkappa \in I}$ be a Whitney decomposition of $\mathbb{R}^{n} \setminus S$. Let $\lambda \in (0,1)$ and $k \in \mathbb{N}_{0}$. Then for every $x \in S_{k}(\lambda)$ there exists a point $y(x) \in Q(x,2^{-k})$ such that

Equation (3.51)

for every Whitney cube $Q_{\varkappa} \ni y(x)$.

Proof.  By Definition 2.6 there exists a point $y\in Q(x,2^{-k})$ such that $Q(y,\lambda 2^{-k}) \subset \mathbb{R}^{n} \setminus S$. We set $y(x):=y$. Now we prove (3.51). Consider an arbitrary Whitney cube $Q_{\varkappa} \ni y(x)$. From (3.10) we have

On the other hand, using (3.10) again, we see that

Combining the above estimates we complete the proof.

§ 4. The main results

Recall that, given $r>0$, we set $k(r):=-[\log_{2}r]$. For each closed nonempty set $S \subset \mathbb{R}^{n}$ and a Whitney decomposition $\{Q_{\alpha}\}_{\alpha \in I}$ of $\mathbb{R}^{n} \setminus S$ we set $k(\varkappa):=k(r_{\varkappa})$ for each $\varkappa \in I$. Throughout this section, unless otherwise stated, we equip $\mathbb{R}^{n}$ with the uniform norm $\|\cdot\|_{\infty}$.

4.1. The direct trace theorem

Recall Definitions 2.3 and 2.6 and the notation following Lemma 3.2.

Lemma 4.1.  Let $d \in [0,n]$, $p \in (1,\infty)$ and $\lambda \in (0,1)$. Let $S$ be a $d$-thick closed set. Let $\{\mu_{k}\}=\{\mu_{k}\}_{k \in \mathbb{N}_{0}}$ be a $d$-regular sequence of measures on $S$. Assume that $f \in L^{\operatorname{loc}}_{1}(S,\mu_{k})$ for some (and hence every) $k \in \mathbb{N}_{0}$. Then there is a positive constant $C$ independent of $f$ such that

Equation (4.1)

Proof.  Given $k \in \mathbb{N}_{0}$, let $\{x_{k,j}\}_{j \in \mathscr{J}_{k}}$ be a maximal $2^{-k}$-separated subset $S_{k}(\lambda)$ of $S$. Using Lemma 3.1, (1), for every $k \in \mathbb{N}_{0}$ we have

Equation (4.2)

Using Lemma 3.11, for every $k \in \mathbb{N}_{0}$ and $j \in \mathscr{J}_{k}$ we choose a point $y(x_{k,j}) \in Q(x_{k,j},2^{-k})$ and fix an index $\varkappa(k,j) \in \mathscr{I}$ such that $Q_{\varkappa(k,j)} \ni y(x_{k,j})$ and (3.51) holds. We define a map $\Theta\colon \bigcup_{k}\mathscr{J}_{k} \to I$ as follows:

Equation (4.3)

Since $Q_{\varkappa} \cap Q(x_{k,j},2^{-k}) \neq \varnothing$, it follows from (3.51) that $Q_{\varkappa} \subset 3Q(x_{k,j},2^{-k})$. Using this, Lemma 3.1, (2), and Proposition 3.4 for $c=6$, it is easy to see that there exists a positive constant $C(n)$ such that for every fixed $k \in \mathbb{N}_{0}$ and every $\varkappa \in \mathscr{I}$

Equation (4.4)

It follows from (3.51) that the equality $\Theta(k,j)=\varkappa=\Theta(k',j')$ implies that

Equation (4.5)

Combining (3.51), (4.4) and (4.5), for every $\varkappa \in \mathscr{I}$ we obtain

Equation (4.6)

If $\varkappa=\Theta(k,j)$ for some $k \in \mathbb{N}_{0}$, $j \in \mathscr{J}_{k}$, then it follows from (3.10) and (3.51) that

This gives the inclusion

Equation (4.7)

Now we use (2.10) and (4.7) and apply Lemma 3.8 for $c=8$. Then for every $k \in \mathbb{N}$ we obtain

Equation (4.8)

From (4.6) it follows that

Equation (4.9)

Similarly

Equation (4.10)

Combining (4.2), (4.8), (4.9) and (4.10) we complete the proof of Lemma 4.1.

Lemma 4.2.  Let $d \in [0,n]$, $p \in (1,\infty)$ and $\lambda \in (0,1)$. Let $S$ be a $d$-thick closed set. Let $\{\mu_{k}\}=\{\mu_{k}\}_{k \in \mathbb{N}_{0}}$ be a $d$-regular sequence of measures in $S$. Assume that ${f \in L_{p}(S,\mu_{0})}$. Then for each $ c \geq 1$ there is a positive constant $C$ depending only on $d,p,n,c$ and $C_{\{\mu_{k}\}}$ such that

Equation (4.11)

Proof.  We use (2.11) and then apply Lemma 3.4 with $d\mathfrak{m}(y)=|f(y)|^{p}\,d\mu_{0}(y)$. This gives the required estimate

Equation (4.12)

The proof is complete.

Recall the notion of a good representative $\widehat{F}$ of a given element $F \in W_{p}^{1}(\mathbb{R}^{n})$.

Lemma 4.3.  Let $d \in (0,n]$ and $p \in (\max\{1,n-d\},n]$. Let $S \subset \mathbb{R}^{n}$ be a $d$-thick closed set. Let $\{\mu_{k}\}_{k \in \mathbb{N}_{0}}$ be a $d$-regular sequence of measures on $S$. Then for every $F \in W_{p}^{1}(\mathbb{R}^{n})$

Equation (4.13)

Proof.  Let $S' \subset S$ be the intersection of $S$ with the set of all Lebesgue points of the function $\widehat{F}$. It follows from Proposition 2.4 that $C_{1,p}(S \setminus S')=0$. For every $x \in S'$ we have

Equation (4.14)

Clearly, $J_{1}(x,r) \to 0$ as $r \to 0$ according to the construction of $S'$. Combining this fact with (4.14) and Proposition 2.1 we conclude that it is sufficient to show that $J_{2}(x,r) \to 0$ as $r \to 0$ for $\mathscr{H}^{n-p}$-a.e. $x \in S$. In fact, applying Theorem 3.4 (for $q=p$) and Proposition 3.1 gives

Equation (4.15)

The lemma is proved.

Now we are ready to prove the main result of this subsection. Recall Definitions 2.72.9 and 2.11.

Theorem 4.1.  Let $d \in [0,n]$, $p \in (\max\{1,n-d\},\infty)$ and $\lambda \in (0,1)$. Let $S \subset \mathbb{R}^{n}$ be a $d$-thick closed set, and let $\{\mu_{k}\}_{k \in \mathbb{N}_{0}}$ be a $d$-regular sequence of measures on $S$. Then the functional $\widetilde{\mathscr{N}}_{S,p,\lambda}$ is bounded on the trace space $W_{p}^{1}(\mathbb{R}^{n})|_{S}$.

Proof.  It is sufficient to verify that there exists a positive constant $C$ (independent of $F$) such that the inequality

Equation (4.16)

holds for each $F \in W_{p}^{1}(\mathbb{R}^{n})$ with $F|_{S}=f$.

We fix some $F \in W_{p}^{1}(\mathbb{R}^{n})$ with $F|_{S}=f$ throughout the proof.

Step 1. First of all we estimate $\mathscr{SN}_{S,p}[f]$ from above. Fix an arbitrary $q \in (\max\{1,n-d\},p)$. We apply Corollary 3.4, Remark 3.1 and Theorem A for $\mathfrak{m}=\mathscr{H}^{n}$, $\alpha=0$, $d=n$ and exponent ${p}/{q}$ instead of $p$. Then we obtain

Equation (4.17)

Step 2. Fix some $q \in (\max\{1,n-d\},p)$. From Lemma 3.2 it follows that $Q(\widetilde{x}_{\varkappa},r) \subset Q(x,10r)$ for every $\varkappa \in \mathscr{I}$, $x \in Q_{\varkappa}$ and $r \in (r_{\varkappa},1)$. Applying Corollary 3.4, Remark 3.1, Theorem A for $\mathfrak{m}=\mathscr{H}^{n}$, $\alpha=0$ and $\gamma={p}/{q}$, and the fact that the interiors of different Whitney cubes are mutually disjoint gives

Equation (4.18)

Step 3. We estimate $\|f|L_{p}(S,\mu_{0})\|$ from above. Let $\{x_{j}\}_{j \in \mathscr{J}}$ be a maximal $1$-separated subset of $S$. Consider the family of cubes $\{Q_{j}\}_{j \in \mathscr{J}}:=\{Q(x_{j},1)\}_{j \in \mathscr{J}}$. It is clear that $ Q(x,1) \subset 2Q_{j}$ for every $x \in S \cap Q_{j}$. Using this and Lemma 3.1, (1), we derive the following estimate:

Equation (4.19)

Using Hölder's inequality, (2.10), Lemma 3.1, (2), Proposition 3.4 with $c=4$ and Proposition 3.5 with $d\mathfrak{m}(y)=|F(y)|^{p}\,d\mathscr{H}^{n}(y)$, we easily obtain

Equation (4.20)

Recall Proposition 2.1 and Lemma 3.6 (also recall that $d>n-p$). Then applying Proposition 3.3 and Remark 3.2 shows that there exists $C>0$ (independent of $F$) such that for every $\delta \in (0,1)$

Equation (4.21)

Since $d>n-p$, we can choose $\delta$ in such a way that $p(1-\delta)>n-d$. Now we use Lemma 3.1, then Proposition 3.5 with $d\mathfrak{m}(x)=\bigl(\operatorname{M}^{<2}[|\nabla F|,(1-\delta)](x)\bigr)^{p}\,d\mu_{0}(x)$, and finally apply Theorem A with $\gamma=p$ and $\alpha=(1-\delta)$. As a result, using (4.21) we obtain

Equation (4.22)

Combining (4.19), (4.20) and (4.22) yields

Equation (4.23)

Step 4. Combining Lemmas 4.1 and 4.2, and estimates (4.18) and (4.23) we obtain

Equation (4.24)

Now the required estimate (4.16) follows directly from (4.17), (4.23) and (4.24). The proof is complete.

4.2. The reverse trace theorem

The following pointwise characterization of the functions in the space $W_{p}^{1}(\mathbb{R}^{n})$ is given in [35].

Theorem D.  Let $p \in (1,\infty]$ and $F \in L_{p}(\mathbb{R}^{n})$. Then $F \in W_{p}^{1}(\mathbb{R}^{n})$ if and only if there exist a nonnegative function $g \in L_{p}(\mathbb{R}^{n})$, a set $E_{F}$ with $\mathscr{H}^{n}(E_{F})=0$ and a constant $\delta>0$ such that

Equation (4.25)

for every $x,y \in \mathbb{R}^{n} \setminus E_{F}$ with $\|x-y\|< \delta$. Furthermore,

Equation (4.26)

where the positive constant $C$ does not depend on $g$.

Proof.  We only sketch the proof. One implication was established in [35] (see the text before Theorem 1 there); see also [36]. For the reverse implication we have to cover $\mathbb{R}^{n}$ by a countable family of 'sufficiently nicely overlapping' balls $\{B_{i}\}_{i \in \mathbb{N}}$ of diameter $\delta>0$ and for every $i \in \mathbb{N}$ apply Theorem 1 in [35] for $\Omega=\operatorname{int}B_{i}$. Then we observe that, given a function $F \in L_{1}^{\operatorname{loc}}(\mathbb{R}^{n})$, the fact that $F \in \bigcap^{N}_{i=1} W_{p}^{1}( \operatorname{int}B_{i})$ implies $F \in W_{p}^{1}\bigl(\bigcup^{N}_{i=1} \operatorname{int}B_{i}\bigr)$ for any $N \in \mathbb{N}$. This shows that locally integrable weak derivatives $D^{\alpha}F$ exist on $\mathbb{R}^{n}$ and their restrictions to any open ball $\operatorname{int}B_{i}$ coincide with the corresponding weak derivatives of $F|_{\operatorname{int}B_{i}}$. It remains to sum the appropriate analogues of estimate (4.26) for all open balls $\operatorname{int}B_{i}$.

Now we are ready to present our construction of the extension operator.

Definition 4.1.  Let $S \subset \mathbb{R}^{n}$ be a $d$-thick closed set for some $d \in [0,n]$. Let $\{\mu_{k}\}=\{\mu_{k}\}_{k \in \mathbb{N}_{0}}$ be a $d$-regular sequence of measures on $S$. Assume that $f \in L^{\mathrm{loc}}_{1}(S,\mu_{k})$ for every $k \in \mathbb{N}_{0}$. For the same family of functions $\{\varphi_{\varkappa}\}_{\varkappa \in I}$ as in Proposition 3.6 we set

Equation (4.27)

where

Remark 4.1.  In fact, (4.27) defines not a single extension operator, but a whole family of operators. The reason is that the choice of the $d$-regular sequence of measures $\{\mu_{k}\}$ is not unique. Furthermore, generally speaking, the choice of the reflected cubes $\widetilde{Q}_{\varkappa}$ is not unique either.

The following result plays a crucial role. It gives a pointwise estimate of the extension constructed in (4.27). Recall the notation $b(\varkappa)$: see (3.12).

Lemma 4.4.  Let $d \in [0,n]$ and let $S$ be a $d$-thick closed set. Let $\{\mu_{k}\}=\{\mu_{k}\}_{k \in \mathbb{N}_{0}}$ be a $d$-regular sequence of measures on $S$. Let $f \in L_{1}^{\operatorname{loc}}(S,\mu_{k})$ for some (and hence every) $k \in \mathbb{N}_{0}$. Suppose that

Equation (4.28)

Then there exists a positive constant $C$ depending only on $n$, $p$, $d$ and $C_{\{\mu_{k}\}}$ such that for each $\delta \in (0,1/150)$, for $(\mathscr{H}^{n} \times \mathscr{H}^{n})$-a.e. $(x,y) \in \mathbb{R}^{2n}$, $\|x-y\|<\delta$, the function $F:=\operatorname{Ext}[f]\colon \mathbb{R}^{n} \to \mathbb{R}$ defined in (4.27) satisfies

Equation (4.29)

where, for every $x \in \mathbb{R}^{n}$,

Equation (4.30)

Proof.  Fix an arbitrary $\delta \in (0,{1}/{150})$. It is obvious that we have to consider five cases:

(1) $x,y \in S$ and $\|x-y\|<\delta$;

(2) $x \in S$, $y \in \mathbb{R}^{n} \setminus S$ and $\|x-y\|<\delta$;

(3) $y \in S$, $x \in \mathbb{R}^{n} \setminus S$ and $\|x-y\|<\delta$;

(4) $x,y \in U_{{1}/{60}}(S) \setminus S$ and $\|x-y\|<\delta$;

(5) $x,y \in \mathbb{R}^{n} \setminus S$ and, furthermore, $\|x-y\|<\delta$ and either $x \notin U_{{1}/{60}}(S)$ or $y \notin U_{{1}/{60}}(S)$.

By the symmetry of the left-hand side of (4.29) with respect to $x$ and $y$, we can identify cases (2) and (3) up to changes in notation.

Case (1). From (4.28) and Corollary 3.3 it follows that for $\mathscr{H}^{n} \times \mathscr{H}^{n}$-almost every $(x,y) \in S \times S$

Equation (4.31)

Case (2) ($=$ case (3)). Consider the case when $x \in S$, $y \in U_{\delta}(S) \setminus S$ and $\|x-y\| \leqslant \delta$. Since $\delta \in (0,1/150)$, estimate (3.11) and Proposition 3.6, (2), give

Equation (4.32)

Recall the notation $b(y)$ (given after Lemma 3.2). From Proposition 3.6, (2), it follows that $b(y)=\{\varkappa \in \mathscr{I}\colon y \in Q^{\ast}_{\varkappa}\}$ for every $y \in \mathbb{R}^{n} \setminus S$. Hence from (4.27) and (4.32) we have

Equation (4.33)

Fix $\varkappa \in b(y)$ and set $r=2\max\{\|x-\widetilde{x}_{\varkappa}\|, \operatorname{diam}\widetilde{Q}_{\varkappa}\}$. Now, (3.10) and Proposition 3.6, (2), give

Hence, using (3.10) again we obtain

Equation (4.34)

We use (4.34) and the fact that $\|x-y\| \leqslant \delta \in (0,1/120)$. Then we obtain

Equation (4.35)

We note that $r_{\varkappa}$ can be much smaller than $r$. Hence we need to estimate with care. Note that $\widetilde{Q}_{\varkappa}:=Q(\widetilde{x}_{\varkappa},r_{\varkappa}) \subset Q(\widetilde{x}_{\varkappa},r/2) \subset Q(x,r)$. Using the same arguments as in the proof of Corollary 3.3, for every $\varkappa \in b(y)$ and $\mathscr{H}^{n}$-a.e. $x \in S$ we have

Equation (4.36)

we have also used the fact that $2r<1$. As a result, combining (4.33), (4.35) and (4.36), for $\mathscr{H}^{n}$-a.e. $x \in S$ and all $y \in Q(x,\delta) \setminus S$ we deduce that

Equation (4.37)

Case (4). Fix $\delta \in (0,1/150)$ and $x,y \in U_{1/60}(S) \setminus S$ with $\|x-y\| \leqslant \delta$. From (3.10) it follows that $\varkappa_{0},\varkappa_{1} \in \mathscr{I}$ for every $Q^{\ast}_{\varkappa_{0}} \ni x$ and $Q^{\ast}_{\varkappa_{1}} \ni y$. Furthermore, from (3.11) it follows that $b(\varkappa_{0}) \cup b(\varkappa_{1}) \subset \mathscr{I}$ in this case. Hence we have

Equation (4.38)

There are two subcases here. In the first $Q_{\varkappa} \cap Q_{\varkappa'}=\varnothing$ for any $Q_{\varkappa}^{\ast} \ni x$ and $Q_{\varkappa'}^{\ast} \ni y$; in the second $Q_{\varkappa_{0}} \cap Q_{\varkappa_{1}} \neq \varnothing$ for some cubes $Q_{\varkappa_{0}}^{\ast} $ and $Q_{\varkappa_{1}}^{\ast} $ containing $x$ and $y$, respectively.

Consider the first subcase. Using (4.38) and arguing as in (4.33) we have

Equation (4.39)

For fixed $\varkappa \in b(x)$ and $\varkappa' \in b(y)$ we set

It is clear that $\varkappa' \notin b(\varkappa)$ and $\varkappa \notin b(\varkappa')$ in this subcase. Hence we obtain

Equation (4.40)

On the other hand, using (3.10) we have

Equation (4.41)

Combining (4.40) and (4.41) we easily deduce that

Equation (4.42)

It is clear that $\widetilde{Q}_{\varkappa} \subset Q(\widetilde{x}_{\varkappa},r)$ and $\widetilde{Q}_{\varkappa'} \subset Q(\widetilde{x}_{\varkappa},r)$. We take (4.42) into account and apply Theorem 3.3 (which is possible due to (4.42) and the restrictions on $\delta$). Then we obtain

Equation (4.43)

Combining (4.39) and (4.43) and using (4.30), for this choice of $x$ and $y$ we have

Equation (4.44)

Now consider the second subcase. Fix arbitrary $Q_{\varkappa_{0}}^{\ast} \ni x$ and $Q_{\varkappa_{1}}^{\ast} \ni y$ with $Q_{\varkappa_{0}} \cap Q_{\varkappa_{1}} \neq \varnothing$. Let $\gamma_{x,y}\colon [0,1] \to \cup_{\varkappa \in b(\varkappa_{0}) \cup b(\varkappa_{1})}Q_{\varkappa}$ be a smooth curve with $|\dot{\gamma}_{x,y}(t)| \leqslant C \|x-y\|$, $t \in [0,1]$. We use (4.27), Lagrange's mean value inequality, (4.38), Proposition 3.6, (4), and estimates (3.11). As a result, we obtain (recall that $\operatorname{supp}\mu_{k}=S$ for each $k \in \mathbb{N}_{0}$)

Equation (4.45)

From (3.10) and (3.11) it follows that

Similarly, $\|\widetilde{x}_{\varkappa}-\widetilde{x}_{\varkappa_{1}}\| \leqslant 25\min\{\operatorname{diam}Q_{\varkappa},\operatorname{diam}Q_{\varkappa_{1}}\}$ for every $\varkappa \in b(\varkappa_{1})$. As a result, since $\operatorname{diam}Q_{\varkappa_{i}} \leqslant \max\{\operatorname{dist}(x,S),\operatorname{dist}(y,S)\} \leqslant 1/60$ for $i=0,1$ (due to (3.10)) we obtain

Equation (4.46)

Using (4.46) we apply Theorem 3.3 and continue with (4.45). This gives

Equation (4.47)

Combining (4.44) and (4.47) we can establish case (4).

Case (5). Fix $\delta \in (0,1/120)$ and $x,y \in \mathbb{R}^{n}$ such that $\|x-y\|<\delta$. Without loss of generality assume that $x \in \mathbb{R}^{n} \setminus U_{1/60}(S)$. Then $y \in \mathbb{R}^{n} \setminus U_{1/120}(S)$. By (3.10) this implies that for every $Q_{\varkappa} \ni x$ and $Q_{\varkappa'} \ni y$

Equation (4.48)

Consider two subcases by analogy with case (4).

In the first subcase $Q_{\varkappa} \cap Q_{\varkappa'}=\varnothing$ for any $Q_{\varkappa}^{\ast} \ni x$ and $Q_{\varkappa'}^{\ast} \ni y$. Then the same arguments as in (4.40) together with (4.48) give

Equation (4.49)

By (4.27) and (4.30) this implies that

Equation (4.50)

In the second subcase there are $Q_{\varkappa_{0}}^{\ast} \ni x$ and $Q_{\varkappa_{1}}^{\ast} \ni y$ with $Q_{\varkappa_{0}} \cap Q_{\varkappa_{1}} \neq \varnothing$. Arguing as in (4.45)) and taking (4.48) into account we obtain

Equation (4.51)

Combining (4.50) and (4.51) we obtain case (5).

The proof of Lemma 4.4 is complete.

Lemma 4.5.  Let $d \in [0,n]$, $p \in (1,\infty)$ and $\lambda \in (0,1)$. Let $S$ be a $d$-thick closed set. Let $\{\mu_{k}\}=\{\mu_{k}\}_{k\in\mathbb{N}_{0}}$ be a $d$-regular sequence of measures on $S$. Then there exists a positive constant $C$, depending only on $d$, $n$, $p$ and $C_{\{\mu_{k}\}}$, such that

Equation (4.52)

Proof.  Clearly, $Q(\widetilde{x}_{\varkappa},r) \subset Q(x,2r)\subset Q(\widetilde{x}_{\varkappa},3)$ for every $x \in Q(\widetilde{x}_{\varkappa},r_{\varkappa})$ and $r \geq r_{\varkappa}$. Hence, applying Lemma 3.8 (for $c=2$) gives

Equation (4.53)

for every $\varkappa \in \mathscr{I}$. Note that $\mu_{k(r_{\varkappa})}(Q(\widetilde{x}_{\varkappa},(1-\lambda)r_{\varkappa}) \cap S) =\mu_{k(r_{\varkappa})}(Q(\widetilde{x}_{\varkappa},(1-\lambda)r_{\varkappa}))$ because $\operatorname{supp}\mu_{k}=S$. Hence using (2.11) and (3.23) for $c=(1-\lambda)^{-1}$, from (4.53) we derive that

Equation (4.54)

It follows from Lemma 3.3 that for every fixed $k \in \mathbb{N}_{0}$ the multiplicity of overlapping of the sets $Q(\widetilde{x}_{\varkappa},(1-\lambda)r_{\varkappa}) \cap S$ with $r_{\varkappa}=2^{-k}$ is bounded above by some positive constant $C(n)$. Furthermore, from Lemma 3.10 it follows that $Q(\widetilde{x}_{\varkappa},(1-\lambda)r_{\varkappa}) \cap S \subset S_{k}(\lambda)$ for each $k \in \mathbb{N}_{0}$, provided that $r_{\varkappa}=2^{-k}$. This, (4.54) and Lemma 4.2 (with $c=3$) give the required estimate

Equation (4.55)

Combining (4.54) and (4.55) we complete the proof.

Lemma 4.6.  Let $d \in [0,n]$, $p \in (1,\infty)$ and $\lambda \in (0,1)$. Let $S$ be a $d$-thick closed set. Let $\{\mu_{k}\}=\{\mu_{k}\}_{k\in\mathbb{N}_{0}}$ be a $d$-regular sequence of measures on $S$. Let $f\colon S \to \mathbb{R}$ be a Borel function such that $\widetilde{\mathscr{N}}_{S,p,\lambda}[f]<\infty$. Let $F:=\operatorname{Ext}[f]$ be the function constructed in (4.27) and $g$ be the function defined in (4.30). Then

Equation (4.56)

The positive constant $C$ in (4.56) is independent of $f$.

Proof.  From (4.27) and Proposition 3.6 it is clear that $|F(x)| \leqslant \chi_{S}(x)|f(x)|+\chi_{\mathbb{R}^{n}\setminus S}(x)g(x)$. It follows from (4.30) that for some positive constant $C $ independent of $f$ we have

Equation (4.57)

Hence it is sufficient to establish that

Equation (4.58)

with a positive constant $C$ independent of $f$. It is clear that (4.56) follows from (4.57) and (4.58).

Step 1. We establish the first estimate in (4.58). Using Lemma 3.2, (3)–(5), we obtain

Equation (4.59)

From Lemma 4.5 and (2.15) we clearly have

Equation (4.60)

Using (2.13) and arguing as in (3.39) we obtain

Equation (4.61)

Using Hölder's inequality, (4.60) and Lemma 4.2 for $c=1$, from (4.61) we obtain

Equation (4.62)

As a result, combination of (4.59), (4.60) and (4.62) gives the first estimate in (4.58).

Step 2. Let $\{x_{j}\}_{j \in \mathscr J}$ be a maximal $(1/2)$-separated subset in $S$. Using Definition 2.3, Lemma 3.1, Corollary 3.2 and Propositions 3.4 and 3.5, we obtain the second estimate in (4.58):

Equation (4.63)

The lemma is proved.

Recall the definitions of the trace $F|_{S}$ of a given element $F\in W_{p}^{1}(\mathbb{R}^{n})$ on the set $S$ (Definitions 2.7 and 2.8). Also recall Definition 4.1.

Theorem 4.2.  Let $d \in [0,n]$ and $p \in (\max\{1,n-d\},\infty)$. Let $S \subset \mathbb{R}^{n}$ be a $d$-thick closed set and $\{\mu_{k}\}_{k \in \mathbb{N}_{0}}$ a $d$-regular sequence of measures on $S$. If $\widetilde{\mathscr{N}}_{S,p,\lambda}[f]<+\infty$ for some $\lambda>0$ and

Equation (4.64)

then $\operatorname{Ext}[f]|_{S}=f$.

Proof.  Let $S' \subset S$ be the set of all points $x$ where (4.64) holds. Set $F:=\operatorname{Ext}[f]$. We are going to show that $\widehat{F}=F$. Fix a cube $Q(x,r)$ with $x \in S$ and $r \in (0,{1}/{100})$. Using (3.10) and the definition of $Q^{\ast}_{\varkappa}$ we see that

Equation (4.65)

From (4.65) it follows that $r_{\varkappa}<1$ and thus $\varkappa \in \mathscr{I}$. Then it follows from Lemma 3.2, (4), and Proposition 3.6, (2), (3), that

Equation (4.66)

Using Corollary 3.2 and (4.64), for $(1,p)$-q.e. $x \in S$ we have

Equation (4.67)

To estimate $J_{2}(x,r)$ we need some preliminaries.

From (4.65) we deduce the inequality $\|x-x_{\varkappa}\|_{\infty} \leqslant r+9r_{\varkappa}/8\leqslant r+9r/14$. Hence for all $\varkappa \in \mathscr{I}$ such that $Q^{\ast}_{\varkappa} \cap Q(x,r) \neq \varnothing$ we have

Equation (4.68)

Recall that Whitney cubes have disjoint interiors. Hence, using (4.68) we obtain

Equation (4.69)

Now for every $\varkappa \in \mathscr{I}$ with $Q^{\ast}_{\varkappa} \cap Q(x,r) \neq \varnothing$ it follows from (3.10) that

Equation (4.70)

For all such $\varkappa$, from (4.70) it follows that

Equation (4.71)

Taking the inclusions (4.71) into account we use (2.13) and then apply Lemma 3.9 for $R=\|x-\widetilde{x}\|+9r$ (note that $17r<1/2$). Finally, we use (4.70) and the monotonicity of $f^{\sharp}_{\{\mu_{k}\}}(\cdot,r)$ with respect to $r$. This gives

Equation (4.72)

Combining (4.69) and (4.72) we obtain

Equation (4.73)

From (4.69), by Hölder's inequality for sums with exponents $p$ and $p'$ we see that the second term on the right-hand side of (4.73) is bounded above by

Equation (4.74)

Clearly, the first term on the right-hand side of (4.73) tends to zero as $r \to +0$ for $(1,p)$-q.e. points $x \in S$ because of (4.64).

In the case $p>n$ we use (4.59) and (4.60) and obtain

Equation (4.75)

for every $x \in S$ because $\widetilde{\mathscr{N}}_{S,p,\lambda}[f]<\infty$. Hence $J_{2}(x,r) \to 0$ as $r \to 0$ everywhere in this case. This, with (4.66) and (4.67), shows that $(1,p)$-quasi-every point $x \in S$ where (4.64) holds is a Lebesgue point of $F$. This fact together with Definition 2.7 proves the claim in the case $p>n$.

Consider the case $p \in (1,n]$. In view of Proposition 2.1, to show that $J_{2}(x,r) \to 0$ as $r \to +0$ for $(1,p)$-q.e. points $x \in S$ it is sufficient to verify that $K(x,r) \to 0$ as $r \to 0$ for $\mathscr{H}^{n-p}$-q.e. points $x \in S$.

In the case $p=n$ this is easy. Indeed, since $\widetilde{\mathscr{N}}_{S,p,\lambda}[f]<\infty$ and (4.60) holds, we can estimate $K(x,r)$ from above by a remainder part of a convergent series:

Equation (4.76)

Consider now the case $p \in (1,n)$. Fix $j \in \mathbb{N}$ and define

Equation (4.77)

Fix some $\delta \in (0,10^{-3})$. For each point $x \in S^{j}$ we find $\delta_{x} \in (0,{\delta}/{50})$ with ${K(x,\delta_{x})>2^{-j}}$. The family of cubes $\{10Q(x,\delta_{x})\}_{x \in S^{j}}$ covers $S^{j}$. Using Vitali's covering theorem (see [31], Ch. 1, §1.5.1) we find an at most countable family of disjoint cubes $\{10Q^{\delta}_{k}\}=\{Q(x_{k},10\delta_{x_{k}})\}_{k \in \mathscr{J}}$ such that $S^{j} \subset \bigcup_{k \in \mathscr{J}}50Q^{\delta}_{k}$. Since the cubes $10Q^{\delta}_{k}$ are disjoint, the inclusions in (4.68) imply that for each $\varkappa \in \mathscr{I}$ the cube $Q^{\ast}_{\varkappa}$ can have nonempty intersection with at most one cube $Q^{\delta}_{k}$. This observation together with (4.74) and (4.77) yields

Equation (4.78)

Recall that according to our assumption $\widetilde{\mathscr{N}}_{S,p,\lambda}[f]<\infty$. This fact, with estimates (4.59) and (4.60), shows that the right-hand side of (4.78) is a remainder part of a convergent series and hence tends to zero as $\delta \to 0$. This clearly implies that $\mathscr{H}^{n-p}(S^{j})= \lim_{\delta \to 0}\mathscr{H}^{n-p}_{\delta}(S^{j})=0$, and hence $\mathscr{H}^{n-p}(\bigcup_{j}S^{j})=0$. As a result, the arguments above and (4.73) imply that

Equation (4.79)

Combining (4.66), (4.67) and (4.79) we complete the proof in the case $p \in (1,n)$. Theorem 4.2 is proved.

Now we can formulate the reverse trace theorem. Recall our construction of the extension operator (4.27).

Theorem 4.3.  Let $p \in (1,\infty)$, $\lambda \in (0,1)$, $d \in [0,n]$ and $d>n-p$. Let $S \subset \mathbb{R}^{n}$ be a $d$-thick closed set. Let $\{\mu_{k}\}_{k \in \mathbb{N}_{0}}$ be a $d$-regular sequence of measures on $S$. Let $\widetilde{\mathscr{N}}_{S,p,\lambda}[f]< \infty$ and

Equation (4.80)

Then $f \in W_{p}^{1}(\mathbb{R}^{n})|_{S}$ and

Equation (4.81)

where the positive constant $C$ depends only on the parameters $p$, $n$, $\lambda$, $d$ and $C_{\{\mu_{k}\}}$.

Proof.  Assume that $\widetilde{\mathscr{N}}_{S,p,\lambda}[f]<\infty$. Then it is obvious that $f_{\varkappa}<\infty$ for all $\varkappa \in \mathscr{I}$. Consequently, (4.27) yields a well-defined function $F:=\operatorname{Ext}[f] \in C^{\infty}(\mathbb{R}^{n} \setminus S)$, whose pointwise restriction to $S$ coincides with the original function $f$. Applying Theorem 4.2 gives $F|_{S}=f$, and hence $\operatorname{Tr}|_{S} \circ \operatorname{Ext}=\operatorname{Id}$.

From Theorem D and Lemmas 4.4 and 4.6 it follows that $F \in W_{p}^{1}(\mathbb{R}^{n})$ and

Equation (4.82)

where the positive constant $C$ does not depend on $f$. Combining Definition 2.9 and (4.82) we obtain (4.81). The theorem is proved.

4.3. The proof of the main result

Now we are ready to prove the main result in this paper.

Proof of Theorem 2.1. Given $f \in W_{p}^{1}(\mathbb{R}^{n})|_{S}$, from Theorem 4.1 it follows that

Equation (4.83)

where the positive constant $C$ does not depend on $f$. Furthermore, from Lemma 4.3 we deduce that (2.16) holds.

Conversely, assume that $\widetilde{\mathscr{N}}_{S,p,\lambda}[f]<+\infty$. Then we deduce from Theorem 4.3 that $f \in W_{p}^{1}(\mathbb{R}^{n})|_{S}$ and

Equation (4.84)

where the positive constant $C$ does not depend on $f$.

Clearly operator $\operatorname{Ext}$ constructed in (4.27) is linear. Furthermore, it was mentioned in the proof of Theorem 4.3 that $\operatorname{Tr}|_{S} \circ \operatorname{Ext} = \operatorname{Id}$.

Finally, estimates (4.83) and (4.84) obviously imply (2.17) and boundedness of the operator $\operatorname{Ext}$. The proof is complete.

Remark 4.2.  As we noted above, while constructing the extension operator we chose a $d$-regular sequence of measures. It is remarkable, however, that both the proofs of Theorems 2.1 and 4.3 and the constants in these proofs depend only on the constant $C_{\{\mu_{k}\}}$ in Remark 2.2 but are independent of the concrete choice of the $d$-regular sequence of measures.

Remark 4.3.  We would like to draw the reader's attention to the fact that (in general) it is impossible to obtain (2.16) from the condition $\widetilde{\mathscr{N}}_{S,p,\lambda}[f]<\infty$ alone. The point is that, given a set $E \subset \mathbb{R}^{n}$ with $\mu_{k}(E)=0$, $k \in \mathbb{N}$, we cannot claim that $C_{1,p}(E)=0$. Hence, changing the given function $f\colon S \to \mathbb{R}$ on a set $E$ such that $\mu_{k}(E)=0$, $k \in \mathbb{N}$, does not affect the value of $\widetilde{\mathscr{N}}_{S,p,\lambda}[f]$, but can violate (2.16).

§ 5. A simplified criterion for sets with porous boundary

In this section we are going to prove Theorem 2.2, which is a simplified version of Theorem 2.1 in the case of sets with porous boundary. Recall Definition 2.6.

Let $S$ be a closed set in $\mathbb{R}^{n}$ with porous boundary. Given $\lambda>0$, for every $k \in \mathbb{N}_{0}$ we set

Equation (5.1)

From Definition 2.6 it is clear that if $\partial S$ is porous then there exists a number $\lambda>0$ such that for all $k \in \mathbb{N}_{0}$

Equation (5.2)

Definition 5.1.  Let $S$ be a closed nonempty subset of $\mathbb{R}^{n}$. We set

Equation (5.3)

Lemma 5.1.  Let $d \in [0,n]$, $p \in (1,\infty)$ and $p>n-d$. Let $S \subset \mathbb{R}^{n}$ be a $d$-thick closed set and $\{\mu_{k}\}_{k \in \mathbb{N}_{0}}$ a $d$-regular sequence of measures on $S$. Assume that $\partial S$ is porous. Then there exists a positive constant $C$ depending only on $d$, $n$, $p$ and $C_{\{\mu_{k}\}}$ such that for every $F \in W_{p}^{1}(\mathbb{R}^{n})$

Equation (5.4)

Proof.  Given $k \in \mathbb{N}$, let $\{x_{k,j}\}_{j \in \mathscr{J}_{k}}$ be an arbitrary maximal $2^{-k}$-separated subset of $\Sigma_{k}$. For every $k \in \mathbb{N}_{0}$ we set $Q_{k,j}:=Q(x_{k,j},2^{-k})$, $j \in \mathscr{J}_{k}$.

Step 1. Given $k \in \mathbb{N}$ and $j \in \mathscr{J}_{k}$, it is clear that $2Q_{k,j} \supset Q(x,2^{-k})$ for every $x \in Q_{k,j}$. Hence elementary computations (similar to (3.35)) give

Equation (5.5)

Using Lemma 3.1, (1), estimates (2.10) and (5.5) and Remark 2.4 we obtain

Equation (5.6)

Fix some $q \in (\max\{1,n-d\},p)$. Arguing as in (3.50), from (5.6) we derive the following estimate:

Equation (5.7)

Step 2. Fix some $\lambda>0$ such that (5.2) holds. Since all the cubes $Q_{k,j}$ are assumed to be closed, $Q_{k,j} \cap \partial S\neq \varnothing$ for all $k \in \mathbb{N}$ and $j \in \mathscr{J}_{k}$. Let $\mathscr{J}^{1}_{k}$ be the set of all $j \in \mathscr{J}_{k}$ such that $Q_{k,j} \cap \partial S^{+}_{k}(\lambda) \neq \varnothing$. Let $\mathscr{J}^{2}_{k}$ be the set of all $j \in \mathscr{J}_{k}$ such that $Q_{k,j} \cap \partial S^{-}_{k}(\lambda) \neq \varnothing$. It is clear that $\mathscr{J}_{k}=\mathscr{J}^{1}_{k} \cup \mathscr{J}^{2}_{k}$ for every $k \in \mathbb{N}$.

Let $W^{1}$ and $W^{2}$ be Whitney decompositions of $\mathbb{R}^{n} \setminus S$ and $S \setminus \partial S$, respectively. Let $\mathscr{I}^{1}$ and $\mathscr{I}^{2}$ be the sets of indices corresponding to the cubes with edge length $\leq1$ in $W^{1}$ and $W^{2}$, respectively.

For every $k \in \mathbb{N}$ and $j \in \mathscr{J}^{1}_{k}$ we choose a point $x'_{k,j} \in Q_{k,j} \cap \partial S^{+}_{k}(\lambda)$, and for every $j \in \mathscr{J}^{2}_{k}$ we choose a point $x'_{k,j} \in Q_{k,j} \cap \partial S^{-}_{k}(\lambda)$. Since $Q(x'_{k,j},2^{-k}) \subset 2Q_{k,j}$ for every $k$ and $j$, we apply Lemma 3.11 and for each $j \in \mathscr{J}^{1}_{k}$ we find a point $y(x'_{k,j}) \in 2Q_{k,j} \cap \mathbb{R}^{n} \setminus S$ such that

Equation (5.8)

Similarly, for each $j \in \mathscr{J}^{2}_{k}$ we find a point $z(x'_{k,j}) \in \operatorname{int} S \cap 2Q_{k,j}$ such that

Equation (5.9)

Consider a map $\Theta^{1}$ that takes a pair $(k,j)$ ($k \in \mathbb{N}, j \in \mathscr{J}^{1}_{k}$) and returns an arbitrary index $\varkappa=\Theta^{1}(k,j) \in \mathscr{I}^{1}$ such that (5.8) holds. Similarly, we build a map $\Theta^{2}$ that takes $(k,j)$ and returns an arbitrary $\varkappa'=\Theta^{2}(k,j) \in \mathscr{I}^{2}$ such that (5.9) holds. Arguing as in (4.6), from (5.8) and (5.9) we derive the existence of a positive constant $C(n,\lambda)$ such that for every $\varkappa \in \mathscr{I}^{1}$ and $\varkappa' \in \mathscr{I}^{2}$

Equation (5.10)

Step 3. Let $k \in \mathbb{N}$, $j \in \mathscr{J}^{1}_{k}$ and $\varkappa= \Theta^{1}(k,j)$. From (5.8) it follows that $2Q_{k,j} \subset 5Q(x,2^{-k})$ for every $x \in Q_{\varkappa}$. Using this and (3.1) we obtain

Equation (5.11)

Similarly, if $k \in \mathbb{N}$, $j \in \mathscr{J}^{2}_{k}$ and $\varkappa'= \Theta^{2}(k,j)$, then

Equation (5.12)

Combining (5.10) and (5.11) we have

Equation (5.13)

Similarly, from (5.10) and (5.12) we obtain

Equation (5.14)

We combine estimates (5.7), (5.13) and (5.14) and apply Theorem A with $\alpha=0$, $d=n$ and $\mathfrak{m}=\mathscr{H}^{n}$. This gives

Equation (5.15)

Lemma 5.1 is proved.

Lemma 5.2.  Let $d \in [0,n]$, $p \in (1,\infty)$ and $p>n-d$. Let $S \subset \mathbb{R}^{n}$ be a $d$-thick closed set with porous boundary $\partial S$. Let $\{\mu_{k}\}=\{\mu_{k}\}_{k \in \mathbb{N}_{0}}$ be a $d$-regular sequence of measures on $S$. Assume that $f \in L_{1}^{\operatorname{loc}}(\mathbb{R}^{n},\mu_{k})$ for some (and hence every) $k \in \mathbb{N}_{0}$. Then

Equation (5.16)

The positive constant $C$ in (5.16) does not depend on $f$.

Proof.  Fix an arbitrary positive number $k \geq 5$. We set $\mathscr{I}_{k}:=\{\varkappa \in \mathscr{I}\colon r_{\varkappa} \leqslant 2^{-k}\}$. Using Vitali's covering theorem (see [31], §1.5.1) we find an index set $\widehat{\mathscr{I}}_{k} \subset \mathscr{I}_{k}$ such that the cubes belonging to the family $\{Q_{\varkappa}(\widetilde{x}_{\varkappa},2^{-k})\}_{\varkappa \in \widehat{\mathscr{I}}_{k}}$ are mutually disjoint and

Equation (5.17)

Note that if $Q\bigl(\widetilde{x}_{\varkappa'},1/2^{k}\bigr) \cap Q\bigl(\widetilde{x}_{\varkappa},5/2^{k}\bigr) \neq \varnothing$ for some $\varkappa' \in \mathscr{I}_{k}$ and $\varkappa \in \widehat{\mathscr{I}}_{k}$, then $Q\bigl(\widetilde{x}_{\varkappa'},2^{-k}\bigr) \subset Q\bigl(\widetilde{x}_{\varkappa},7/2^{k}\bigr)$. Using Remark 2.4 and reasoning as in (3.35) it is easy to show that for such $\varkappa$ and $\varkappa'$

Equation (5.18)

Since different Whitney cubes have disjoint interiors, we find that for every $\varkappa \in \widehat{\mathscr{I}}_{k}$

Equation (5.19)

Combining (5.17), (5.18) and (5.19) we obtain

Equation (5.20)

It is clear that $Q\bigl(\widetilde{x}_{\varkappa},7/2^{k}\bigr) \subset Q(x,15/2^{k})$ for every $x \in S \cap Q\bigl(\widetilde{x}_{\varkappa},7/2^{k}\bigr)$. Using this observation, Remark 2.4, (2.11) and (2.12) and arguing as in (3.35), for every $\varkappa \in \mathscr{I}_{k}$ we have

Equation (5.21)

It is clear that $Q\bigl(\widetilde{x}_{\varkappa},7/2^{k}\bigr) \cap S \subset \Sigma_{k-4}$. Furthermore, according to our construction of $\widehat{\mathscr{I}}_{k}$ the multiplicity of overlapping of the sets $Q\bigl(\widetilde{x}_{\varkappa},7/2^{k}\bigr) \cap S$, $\varkappa \in \widehat{\mathscr{I}}_{k}$ is finite and independent of $k$. Hence, substituting (5.21) into (5.20) and using Proposition 3.5 we obtain

Equation (5.22)

Now we use Remark 2.4 and Hölder's inequality. Then we use Lemma 3.7 and Lemma 4.2 for $c=1$. This gives

Equation (5.23)

To complete the proof it is sufficient to put (5.22) and (5.23) together.

Lemma 5.3.  Let $\lambda \in (0,1)$, $d \in [0,n]$, $p \in (1,\infty)$ and $p>n-d$. Let $S$ be a $d$-thick closed set with porous boundary $\partial S$. Let $\{\mu_{k}\}=\{\mu_{k}\}_{k \in \mathbb{N}_{0}}$ be a $d$-regular sequence of measures on $S$. Assume that $f \in L_{1}^{\operatorname{loc}}(\mathbb{R}^{n},\mu_{k})$ for every $k \in \mathbb{N}_{0}$. Then for each $\lambda \in (0,1)$

Equation (5.24)

The positive constant $C$ in (5.24) is independent of $f$.

Proof.  It is clear from Lemma 3.5 and Remark 2.4 that for every $\varkappa \in \mathscr{I}$ (recall that $\operatorname{diam}Q_{\varkappa} \leqslant 1$ for such $\varkappa$) we can choose $j_{\varkappa} \in \mathbb{N}_{0}$ so that $0 \leqslant j_{\varkappa}<|\operatorname{log}_{2}r_{\varkappa}|$ and

Equation (5.25)

Given $k \in \mathbb{N}$, let $\mathscr{I}_{k}:=\{\varkappa \in \mathscr{I}\colon j_{\varkappa}=k\}$. Estimate (5.25) together with Lemmas 4.1, 4.2 and 5.2 allow us to deduce that

Equation (5.26)

The proof is complete.

Proof of Theorem 2.2. Let $f \in W_{p}^{1}(\mathbb{R}^{n})|_{S}$. It follows from Lemma 5.1 and (4.16) that

Equation (5.27)

where the positive constant $C$ does not depend on $f$. Furthermore, from Lemma 4.3 we deduce that (2.18) holds.

Conversely, let $\mathscr{N}_{S,p}[f]<+\infty$. Then, from Lemma 5.3 we deduce that $\widetilde{\mathscr{N}}_{S,p,\lambda}[f] \leqslant C\mathscr{N}_{S,p}[f]$ for some $\lambda \in (0,1)$, with a positive constant $C$ independent of $f$. Hence, from Theorem 4.3 we deduce that $f \in W_{p}^{1}(\mathbb{R}^{n})|_{S}$ and

Equation (5.28)

where the positive constant $C$ does not depend on $f$.

Finally, estimates (5.27) and (5.28) obviously imply (2.19). The proof is complete.

§ 6. Example

The aim of this section is to present several useful examples, which show the power of our main results.

Example 6.1.  Let $S$ be an Ahlfors $n$-regular closed subset of $\mathbb{R}^{n}$, and let $p>1$. In this case we can take $\mu_{k}=\mathscr{H}^{n}\lfloor S$ for every $k \in \mathbb{N}_{0}$ to obtain an $n$-regular sequence of measures on $S$. Hence, taking Remark 2.4 into account, for every $t \in [0,1)$ we have

Equation (6.1)

To simplify our notation we set $f^{\sharp}_{S}:=f^{\sharp}_{\{\mu_{k}\}}$ in this case. This notation was used in [23].

We establish the following key estimate:

Equation (6.2)

where the positive constant $C$ does not depend on $f$.

We combine Theorem C, (2), and Lemmas 4.1 and 4.2. This gives

Equation (6.3)

Adapting the arguments in the proof of Lemma 4.5 to this case (and also using the monotonicity of $f^{\sharp}_{S}(\cdot,t)$ with respect to $t$) we easily obtain

Equation (6.4)

Now (6.2) clearly follows from (6.3) and (6.4).

Recall that $\mathscr{H}^{n}$-a.e. points $x \in S$ are Lebesgue points of a function $f \in L_{p}(S)$. If we relax the notion of the trace of a given $F \in W_{p}^{1}(\mathbb{R}^{n})$ and identify $F|_{S}$ with the class of functions equivalent modulo coincidence $\mathscr{H}^{n}$-a.e. on $S$, then from (6.2) we deduce the following simplified version of Theorem 2.1.

Let $S \subset \mathbb{R}^{n}$ be an Ahlfors $n$-regular set. Then a function $f\colon S \to \mathbb{R}$ belongs to the trace space $W_{p}^{1}(\mathbb{R}^{n})|_{S}$ if and only if

Equation (6.5)

Moreover, the operator $\operatorname{Ext}$ constructed in (4.27) is a bounded linear extension operator $\operatorname{Ext}\colon W_{p}^{1}(\mathbb{R}^{n})|_{S} \to W_{p}^{1}(\mathbb{R}^{n})$ and

This result is a slight modification of the corresponding result obtained by Shvartsman [23] in the context of first-order Sobolev spaces.

Example 6.2.  Let $d \in [0,n)$ and $p \in (\max\{1,n-d\},\infty)$. Let $S$ be an Ahlfors $d$-regular subset of $\mathbb{R}^{n}$. In this case there exists a simple $d$-regular sequence of measures on $S$. More precisely, set $\mu_{k}=\mathscr{H}^{d}\lfloor S$ for every $k \in \mathbb{N}_{0}$. Clearly, ${\mathscr{H}^{n}(S)=0}$. Furthermore, $\operatorname{int}S=\varnothing$ and $\partial S$ is porous (see Proposition 2.3).

Note that the measure $\mathscr{H}^{d}\lfloor S$ is Radon. Hence, from Theorem 1 in [31], §1.7.1, we conclude that if $f \in L^{\operatorname{loc}}_{1}(S,\mathscr{H}^{d}\lfloor S)$, then

for $\mathscr{H}^{d}$-almost every $x \in S$.

Now we apply Theorem 2.2 and take the above facts into account. If we relax the notion of the trace of a given element $F \in W_{p}^{1}(\mathbb{R}^{n})$ and identify $F|_{S}$ with the class of functions equivalent modulo coincidence $\mathscr{H}^{d}$-a.e. on $S$, we obtain the following simplified version of Theorem 2.2.

Given an Ahlfors $d$-regular set $S \subset \mathbb{R}^{n}$, for some $d \in [0,n)$, let $p \in (\max\{1, {n-d}\},\infty)$. Then a function $f \colon S \to \mathbb{R}$ belongs to the trace space $W_{p}^{1}(\mathbb{R}^{n})|_{S}$ if and only if

Moreover, the operator $\operatorname{Ext}$ constructed in (4.27) is a bounded linear extension operator from $W_{p}^{1}(\mathbb{R}^{n})|_{S}$ to $W_{p}^{1}(\mathbb{R}^{n})$ and

Note that this result coincides with that obtained in [21] in the context of first-order Sobolev spaces.

In the simplest case when $S=\mathbb{R}^{d} \subset \mathbb{R}^{n}$, $d=1,\dots,n-1$, this is a classical result. Namely, $W_{p}^{1}(\mathbb{R}^{n})|_{\mathbb{R}^{d}}=B^{1-(n-d)/{p}}_{p,p}(\mathbb{R}^{d})$. This fact together with Theorem 2.2 implies that $\mathscr{BN}_{\mathbb{R}^{d},p,\lambda}[f] \approx \|f|B^{1-(n-d)/{p}}_{p,p}(\mathbb{R}^{d})\|$. This equivalence has motivated us to call $\mathscr{BN}_{S,p}[f]$ a 'Besov-type seminorm'.

Example 6.3.  Let $\beta\colon [0,+\infty) \to [0,+\infty)$ be a strictly increasing continuous function such that $\beta(0)=0$ and $\beta(t)>0$ for every $t>0$. Let $\beta^{-1}$ denote the inverse function, so that $\beta^{-1}\circ \beta= \operatorname{id}$ on $[0,+\infty)$. Consider the closed single cusp $G^{\beta}:=\{x=(x',x_{n})\colon \max_{i=1,\dots,n-1}|x_{i}| \leqslant \beta(x_{n})\}$. For each $k \in \mathbb{N}_{0}$ we also consider the sets

Equation (6.6)

Recall Definition 5.1. It is clear that $G^{\beta}_{k}$ coincides with $\Sigma_{k}(G^{\beta})$.

For every $k \in \mathbb{N}_{0}$ consider the measure $d\mu_{k}(x)=w^{\beta}_{k}(x)\,d\mathscr{H}^{n}(x)$, where

Equation (6.7)

It is clear from (6.7) (recall that $\beta$ is strictly increasing) that

Equation (6.8)

Using the monotonicity of $\beta$ and elementary geometric observations, it is easy to see that for every point $x=(x',x_{n}) \in G^{\beta}$ and $r \in (0,1)$

Equation (6.9)

On the other hand, using (6.7) and the monotonicity and continuity properties of $\beta$, it is easy to show that for every $x=(x',x_{n})\in G^{\beta}$

Equation (6.10)

Direct computations give

Equation (6.11)

for every $x=(x',x_{n}) \in G^{\beta}$.

Combining (6.8)–(6.11) we see that the sequence of measures $\{\mu_{k}\}_{k \in \mathbb{N}_{0}}$, possibly after multiplying by a fixed constant (depending on $n$ and $\beta$), becomes $1$-regular on $G^{\beta}$.

Recall Example 2.1, (2), and Example 2.5. Thus we see that the set $G^{\beta}$ is $1$-thick and has a porous boundary. Consider a slightly relaxed definition of the trace of $F\in W_{p}^{1}(\mathbb{R}^{n})$ on the set $G^{\beta}$. Namely, we write $F|_{G^{\beta}}=f$ if $F(x)=f(x)$ for $\mathscr{H}^{n}$-a.e. $x \in G^{\beta}$. Then from Theorem 2.2 we clearly derive the following criterion.

Let $p>n-1$. Then a function $f\colon G^{\beta} \to \mathbb{R}^{n}$ lies in the trace space $W_{p}^{1}(\mathbb{R}^{n})|_{G^{\beta}}$ if and only if

Equation (6.12)

Furthermore, the functional $\mathscr{N}$ gives an equivalent norm in the trace space $W_{p}^{1}(\mathbb{R}^{n})|_{G^{\beta}}$ and the operator $\operatorname{Ext}$ in (4.27) is a bounded linear extension operator from $W_{p}^{1}(\mathbb{R}^{n})|_{G^{\beta}}$ to $W_{p}^{1}(\mathbb{R}^{n})$.

Remark 6.1.  To the best of our knowledge the results in Example 6.3 are new and could not be obtained using the techniques previously known. However, we have to mention [24], where a similar example was considered under certain additional assumptions on $\beta$. More precisely, it was assumed there that $\beta$ is Lipschitz. Furthermore, in [24] precise statements were only formulated in the case $n=2$.

Please wait… references are loading.
10.1070/SM9199