Strong -extension and -extension domains
Abstract.
We show that a bounded domain in a Euclidean space is a -extension domain if and only if it is a strong -extension domain. In the planar case, bounded and strong -extension domains are shown to be exactly those -extension domains for which the set is purely -unrectifiable, where are the open connected components of .
Key words and phrases:
Sobolev extension, BV-extension2000 Mathematics Subject Classification:
Primary 46E35.1. Introduction
Let be a domain for some . For every , we define the Sobolev space to be
where denotes the distributional gradient of . We equip this space with the non-homogeneous norm
We say that is a -extension domain if there exists an operator and a constant so that
and for every . We denote the minimal constant above by . We point out that by the results from [10, 21], for one can always assume the operator to be linear, and also for the case of bounded simply connected planar domains if by [16]. It is not yet known if this is the case for general domains when .
It is well-known from the works of Calderón and Stein [5, 23] that Lipschitz domains are -extension domains for every . Moreover, Jones showed in [12] that every uniform domain is a -extension domain for all . However, these conditions are not necessary for a domain to be a Sobolev extension domain. For bounded simply connected planar domains a geometric characterization of Sobolev extension domains by means of a curve condition has been given in the works [22, 15, 16]. Namely, for the case we have the following: A bounded planar simply connected domain is a -extension domain if and only if for every there exists a curve connecting and with
(1.1) |
A typical example of a simply connected planar domain which is not a -extension domain for any is the slit disk . However, by the results of [14], knowing that the complement is quasiconvex is enough to ensure that is a -extension domain.
Recall that
is the space of functions of bounded variation where
denotes the total variation of on . We endow this space with the norm Note that is a Radon measure on that is defined for every set as
We say that is a -extension domain if there exists a constant and a (not necessarily linear) extension operator so that and
for all and where is an absolute constant, independent of . Let us point out that being a -extension domain always implies that it is also a -extension domain (see [14, Lemma 2.4]).
Our first main result is the characterization of bounded -extension domains in terms of strong extendability of -functions, or equivalently, in terms of strong extendability of sets of finite perimeter. The equivalence between strong extendability of -functions and strong extendability of sets of finite perimeter is inspired by the work of Mazy’a and Burago [4] (see also [20, Section 9.3]). They showed that for all with finite total variation we may find an extension with , for some constant , if and only if any set of finite perimeter in admits an extension satisfying and where is some constant. Recall that a Lebesgue measurable subset has finite perimeter in if , where denotes the characteristic function of the set . We set and call it the perimeter of in . If a set does not have finite perimeter in we set .
Before stating our characterization, we introduce the terminology of strong extendability, following [11] and [17].
Definition 1.1 (Strong -extension domain).
A domain is called a strong -extension domain if there exists a constant so that for any there exists with , , and .
In the spirit of Definition 1.1, we define the analogous concept for sets of finite perimeter.
Definition 1.2 (Strong extension property for sets of finite perimeter).
A domain is said to have the strong extension property for sets of finite perimeter if there exists a constant so that for any set of finite perimeter in there exists a set such that
-
(PE1)
modulo measure zero sets,
-
(PE2)
, and
-
(PE3)
.
With the above definitions we can state our first main result.
Theorem 1.3.
Let be a bounded domain. Then the following are equivalent:
-
(1)
is a -extension domain.
-
(2)
is a strong -extension domain.
-
(3)
has the strong extension property for sets of finite perimeter.
Our main motivation behind this theorem is to understand better the geometry of -extension domains. From Theorem 1.3 we see that for a bounded -extension domain, except for a purely -unrectifiable set, the boundary consists of points where the domain has density at most . See Section 4 for the proof of this. In the same section we give an example showing that the above density bound is not sufficient to imply that a bounded -extension domain is a -extension domain, even in the plane. Another corollary of Theorem 1.3 is that for a bounded -extension domain, again up to a purely -unrectifiable set, the boundary consists of points that are boundary points also for some component of the interior of the complement of the domain. In Section 4 we provide also an example showing that in this property does not characterize -extension domains among bounded -extension domains. However, our second main result states that in the planar case this is true.
Theorem 1.4.
Let be a bounded -extension domain. Then is a -extension domain if and only if the set
is purely -unrectifiable, where are the connected components of .
Let us mention that Theorem 1.4 recovers partly the theorems in [16]. Namely, it immediately follows that Jordan -extension domains are -extension domains since the set required in Theorem 1.4 to be purely unrectifiable, is indeed empty. The curve characterization (1.1) also follows quite easily from Theorem 1.4 using a small observation recorded in [16]. Let us briefly sketch this. Since a -extension domain is known to be a -extension domain, its complement is quasiconvex. Then, a quasiconvex curve between two points in the complement can be modified to intersect the boundaries of each at most twice (see Lemma 5.3). Theorem 1.4 now says that the rest of the curve intersects in a -measure zero set, giving condition (1.1). Conversely, (1.1) implies quasiconvexity, and hence that is a -extension domain. For a simply connected , we can connect every pair of components and with a curve satisfying (1.1). Since the set
is contained in countably many of such curves by [16, Lemma 4.6], we see that it is purely -unrectifiable.
Let us point out, however, that the extension operator that we construct in Theorem 1.4, is not always linear. One of the main points of [16] was to construct a linear extension operator. At the moment we do not see how our construction could be modified to give a linear extension operator. Still, the general smoothing operator we use for proving Theorem 1.3 (and Theorem 1.4) immediately gives the following.
Corollary 1.5.
Suppose is a bounded strong -extension domains where the extension operator is linear. Then there exists a linear -extension operator from to .
Although not strictly used in our proofs, we include the following result for future use: Every -extension domain satisfies the measure density condition, that is, there exists a constant so that for every and we have One may find this result in Section 2.2. The same conclusion for -extension domains with is also true and was already shown in [10].
2. Preliminaries
When making estimates, we often write the constants as positive real numbers which may vary between appearances, even within a chain of inequalities. These constants normally only depend on the dimension of the underlying space unless otherwise stated.
For any point and radius we denote the open ball by
More generally, for a set we define the open -neighbourhood as
We denote by the -dimensional outer Lebesgue measure of a set . For any Lebesgue measurable subset and any point we then define the upper density of at as
and the lower density of at as
If , we call the common value the density of at and denote it by . The essential interior of is then defined as
the essential closure of as
and the essential boundary of as
As usual, will stand for the -dimensional Hausdorff measure of a set obtained as the limit
where is the -dimensional Hausdorff -content of defined as
We say that a set is -rectifiable, for some , if there exist countably many Lipschitz maps so that . A set will be called purely -unrectifiable if for every Lipschitz map we have
Observe that by Rademacher’s theorem one can deduce that if is Lipschitz, then there are countably many sets on which is bi-Lipschitz and such that .
Moreover, it easily follows that if is not -purely unrectifiable, then there exists a Lipschitz map so that up to a rotation, the set
has positive -measure, where .
By a dyadic cube we refer to for some and . We denote the side-length of such dyadic cube by .
2.1. -functions and sets of finite perimeter
Let us recall some basic results related to -functions and sets of finite perimeter. For a more detailed account, we refer to the books [2, 6, 8].
Differently to this paper, Mazy’a and Burago’ (see [4] and also [20, Section 9.3]) considered the space
equipped with the seminorm . This way they defined -extension domains to be those for which just the total variation of the extension is controlled, that is, whenever . As we already explained in the introduction, they proved that being a -extension domain was equivalent to the fact that any set of finite perimeter in admits an extension satisfying only (PE1) and (PE2) from Definition 1.2. Note however, that thanks to [14, Lemma 2.1] -extension domains are equivalent to extension domains if is bounded.
When working with functions we will make use of the well-known -Poincaré inequality that we now state (see for instance [2, Theorem 3.44] for the proof).
Theorem 2.1.
Let be an open bounded set with Lipschitz boundary. Then there exists a constant depending only on and so that for every we have
In particular, there exists a constant only depending on so that if are two dyadic cubes with and connected, then for every ,
(2.1) |
We are using here the notation of the mean value integral of a function on the set as
Let us record as well the coarea formula for functions. See for example [6, Section 5.5].
Theorem 2.2.
Given a function , the superlevel sets have finite perimeter in for almost every and
for every Borel set . Conversely, if and then .
An important result due to Federer [8, Section 4..5.11] tells us that a set has finite perimeter in if and only if . Moreover, thanks to De Giorgi’s pioneering work [7] we can understand the structure of the boundary of sets of finite perimeter even better. Namely, if has finite perimeter in then for every subset ,
and if has finite perimeter in then
where and are compact subsets of hypersurfaces. Furthermore, for any set with finite perimeter we have
for -almost every . Moreover, , and hence for -almost every we have
Let us finally recall some terminology and results from [1]. A Lebesgue measurable set with is called decomposable if there exist two Lebesgue measurable sets so that , , , and
A set is called indecomposable if it is not decomposable. For example, any connected open set with is indecomposable.
For any set of finite perimeter we can always find a unique countable family of disjoint indecomposable subsets so that , and, moreover,
For a proof of this result we refer to [1, Theorem 1].
2.2. Measure density condition for -extension domains
Nowadays it is a well-known fact that all -extension domains for satisfy the measure density condition (see [10]). Although we do not need this in our proofs, we record here the fact that the same property holds for -extension domains. Let us remark that a measure density condition for planar -extension domains was proven in [14, Lemma 2.10]. However, the proof does not seem to extend to domains in . The method of proof we employ here follows the same lines as [10] and can be adapted for -extension domains as well.
Proposition 2.3.
Let be a -extension or a -extension domain, then there exists a constant , depending only on and on the operator norm, so that for every and we have
Proof.
We will only make the proof for -extension domains. For -extension domains one can use the results from [10], or the fact that -extension domains are -extension domains. A proof of this fact can be found in [14, Lemma 2.4]. The reader will notice that the key point will be to apply the Sobolev embedding theorem, which is both valid for and functions.
Let us denote . By induction, we define for every the radius by the equality
Since , we have that as .
For each , consider the function
Note that these functions belong to the class , in particular they are functions. We can estimate their -norms by
Call and denote by the extension operator. By the Sobolev inequality for BV functions (see [6, Theorem 5.10]) we know that
where depends only on the dimension . Hence we have the following chain of inequalities
We also have
and therefore
Consequently,
By summing up all these quantities we conclude that
This gives the claimed inequality. ∎
3. Equivalence of -extension and strong -extension domains
This section is devoted to the proof of Theorem 1.3. The idea in going from a strong -extension to a -extension is to first extend the -function from the domain as a -function to the whole space and then mollify it in the exterior of the domain. In the mollification process it is important to check that we do not change the function too much near the boundary.
3.1. Whitney smoothing operator
In this subsection we prove existence of a suitable smoothing operator from BV to . For similar constructions we recommend to the reader to have a look at [3, 9, 18].
Theorem 3.1.
Let be open subsets. There exist a constant depending only on the dimension and a linear operator
so that for any we have ,
(3.1) |
and
(3.2) |
where is understood to be defined in the whole via a zero-extension. Moreover, the operator is also bounded when acting from the space into the homogeneous Sobolev space .
Recall that stands for the homogeneous Sobolev space endowed with the seminorm .
Let us briefly explain how the operator is constructed. We first take a Whitney decomposition of the open set and a partition of unity based on it. The operator on a -function is then defined as the sum of restricted to the complement of and the average values of in each Whitney cube of times the associated partition function. This way, we immediately have that the function is left unchanged in the complement of , and that in it is smooth. The inequality (3.1) will follow in a standard way from the Poincaré inequality for -functions, whereas for showing (3.2) we will show that the average difference between and near -almost every boundary point of tends to zero as we get closer to the point.
Let us now give the definition of the operator doing the smoothing part. Suppose is an open set, not equal to the entire space . Let be the standard Whitney decomposition of , by which we mean that it satisfies the following properties:
-
(W1)
Each is a dyadic cube inside .
-
(W2)
and for every we have .
-
(W3)
For every we have ,
-
(W4)
If , we have .
The reader can find a proof of the existence of such a dyadic decomposition of the set in [23, Chapter VI].
For a given set and its Whitney decomposition we take a partition of unity so that for every we have , , , with a constant depending only on , and
With the partition of unity we then define for any a function
(3.3) |
Let us start by showing that maps to boundedly. Even though we could obviously equivalently use the BV norm also on the target, we prefer to write it as the -norm in order to underline the spaces where the operator will be used.
Lemma 3.2.
Let be the operator defined in (3.3). Then for any we have and with a constant depending only on .
Proof.
By (W2) and the fact that for every , we know that implies that . Therefore, any point in has a neighbourhood where is defined as a sum of finitely many -functions. Consequently, . For the -norm of the function we can estimate
For the estimate on the -norm of the gradient we start with an estimate via the -Poincaré inequality (2.1)
which then gives, by summing over all , and noticing that in the final double sum the sets have finite overlap with a constant depending only on ,
(3.4) |
This concludes the proof of the lemma. ∎
The next lemma gives the crucial boundary behaviour that will imply (3.2).
Lemma 3.3.
Proof.
Suppose (3.5) fails on a set with . Without loss of generality, we may assume compact. By going to a subset of if needed, we may further assume that there exists a constant so that
for every .
Let . By the -covering lemma there exists a disjointed countable collection so that , for all ,
(3.6) |
and
Similarly as in the proof of Lemma 3.2, we first estimate in a Whitney cube using the (1,1)-Poincaré inequality (2.1)
(3.7) |
By the property (W3) of the Whitney decomposition, we conclude that if is such that , we have
and hence
Similarly, for the same , if for some , by (W4), we get
and so
Now, using the definition of the Hausdorff content, the inequality (3.6), the estimate (3.7), and the above consideration for the cubes , we get
as . Thus
giving a contradiction and concluding the proof. ∎
With the previous two lemmas we can now prove the main theorem of the section.
Proof of Theorem 3.1.
Let be the operator defined in (3.3) and suppose that is given. We define
Let us introduce the superlevel sets for every , where is defined in the whole via a zero-extension. We want to show that for almost every and the equality (3.2) will follow by a simple application of the coarea formula. We proceed as follows.
In the case that , observe that for every we have , then for -almost all , by (3.8),
This, together with the fact that , means that the set has density at -almost all points .
If we take , for every we have , and then for -almost all , again by (3.8),
This means, using , that the set has density at -almost all points .
From these previous observations we deduce that for all . We therefore obtain (3.2), applying the coarea formula,
3.2. Proof of Theorem 1.3
In this section we will prove Theorem 1.3 with the help of Theorem 3.1. Recall that we are claiming that for a bounded domain the following are equivalent:
-
(1)
is a -extension domain.
-
(2)
is a strong -extension domain.
-
(3)
has the strong extension property for sets of finite perimeter.
We will show the equivalence by showing the implications
Proof of the implication (1) (3).
We start with the assumption that is a bounded -extension domain. In particular, it is known that is also a -extension domain (see [13]). That is, there exists an extension operator with for every . Since is bounded, after multiplying with a suitable Lipschitz cutoff-function we may assume that and still keep the control on the gradient norm.
We claim that has the strong extension property for sets of finite perimeter. Thus, let be a set of finite perimeter in . We need to find a set so that (PE1)–(PE3) of Definition 1.2 hold.
Towards this, let be the operator given by Theorem 3.1. We now define a function by
By truncating the function if needed, we may assume that .
Applying the coarea formula (Theorem 2.2) for the function ,
This gives, in particular, that there exists a set with for which for every we have
(3.9) |
In the penultimate inequality we are using (3.4).
Proof of the implication (3) (2).
By assumption has the strong extension property for sets of finite perimeter, so there exists a constant such that for any set of finite perimeter there exists a set such that (PE1)–(PE3) are satisfied.
Take a function and let be a large enough ball. Without loss of generality, we may assume that . Let us write for the superlevel sets for each . Since , by the coarea formula, for almost every . For these , we select to be a strong perimeter extension of . For convenience, for the remaining we define . Notice that these are not strong perimeter extensions of . This will not pose a problem for us, since we will not use these values of in the construction below.
Before going to the actual proof, let us note that if the strong perimeter extensions could be chosen so that is measurable, by Fubini’s theorem we would obtain
for the function . In order to circumvent the measurability issue, we proceed by defining the extension in a similar way, but as a limit of simple functions .
For every , let us denote by the (half-open) dyadic interval of length containing . For almost every we then have
(3.11) |
For almost every we also have
Since is a compact set, for almost every we then have
Let us write for each
Notice that the sets are not necessarily measurable. Nevertheless, since is a regular outer measure, we have
for every .
We define a sequence inductively as follows. First take so that
Suppose now that has been defined for all . Then we take so that
(3.12) |
for all . Notice that this requirement can be obtained since (3.12) is with a strict inequality and again by outer regularity, for every we have
as .
Now, for we also take for which
(3.13) |
where
The index then gives us the scale at which the simple function is constructed.
Let us now construct the function for a given . For each define
Notice that always since .
We then select
Next, we define
which satisfies and .
For every , let us denote
and
Since is measurable, (3.13) gives , and thus, by (3.12)
Hence, we have
(3.14) |
For the norm of , by the fact that for every , we get the estimate
Hence, there exists a subsequence of , which converges in to a function . For it, we have
Moreover, clearly on .
In order to estimate we observe that, for every , we have, by (3.14),
(3.15) |
where
as by the absolute continuity of the integral. Since the upper bound in (3.15) goes to zero as independently of , we have
Let us assume that the function is extended as zero outside . Recall that we have
In order to conclude the proof we control the -norm of the function in the whole as follows.
Consider a Lipschitz function which takes the value on and has support in . Then one can check that
and using the Poincaré inequality that
Therefore we have , where the constant depends on the constant coming from (PE2), on and on the constant coming from the Poincaré inequality. We then can assure that is an extension operator.
Obviously we still have . Hence is indeed a strong -extension domain. ∎
Proof of the implication (2) (1).
We start with a strong -extension operator
In particular, we know that
(3.16) |
for every .
Let be a Whitney smoothing operator given by Theorem 3.1. We assert that the operator defined by is a -extension operator.
Observe that on and
To conclude we must check that indeed , so that in particular . In order to get this let us show that the Radon measure consists only of its absolutely continuous part, and not of its singular part. Since we already now that and are functions we merely have to prove that . By the special properties of our smoothing operator given by (3.2) and by our assumption (3.16) we have that
and we are done. ∎
4. Further properties of -domains
In this section we prove some corollaries to Theorem 1.3.
Corollary 4.1.
Let be a bounded -extension domain. Then the set of points with is purely -unrectifiable.
Proof.
If the set
is not purely -unrectifiable, there exists a Lipschitz map so that, after a suitable rotation,
Notice that the set consists of two connected components. Select one of the components that has nonempty intersection with (actually, both have) and call its restriction to . Then
and so in particular has finite perimeter in . Let be any set of finite perimeter with . Since
at -almost every point , and for every , we have
Using again the fact that
at -almost every point , and for every , we have
for -almost every point . This means that there exists a set with for which
Consequently,
Hence does not have the strong extension property for sets of finite perimeter ((PE3) fails), and so by Theorem 1.3 it is not a -extension domain. ∎
The next example shows that even in the plane the conclusion of Corollary 4.1 is not sufficient to imply that a bounded -extension domain is a -extension domain.
Example 4.2.
Let us construct a planar -extension domain so that the upper-density of at all except at countably many boundary-points is at most , but the domain is not a -extension domain. We set
where, for every , we define
See Figure 1 for an illustration.

Now, the upper-density of is clearly at most at all the points of the boundary except for the corners of the connected components of , and the points and , which together form only a countable set. (One could remove balls instead of rectangles to get the upper-density bound for all boundary points.)
The domain is a -extension domain because each removed square has a neighbourhood inside from which the -function can be extended to the square with a uniform constant. These neighbourhoods can be taken pairwise disjoint. This will result in an extension operator
The target set clearly admits an extension to .
The domain is not a -extension domain, because the set is not purely -unrectifiable, and this is the set in the following Corollary 4.3.
The next corollary to Theorem 1.3 shows that one direction in Theorem 1.4 holds also in higher dimensions.
Corollary 4.3.
Suppose that is a bounded -extension domain. Let , for , be the connected components of . Then the set
is purely -unrectifiable.
Proof.
Supposing to be a -extension domain, by Theorem 1.3 we know that it has the strong perimeter extension property.
Now, towards a contradiction, suppose that is an -Lipschitz map so that
after a suitable rotation. Let be a component of such that the set
has positive -measure. By the measure density of (Proposition 2.3), at least one of the components must satisfy this. Without loss of generality, we may assume that
Take and let be the strong perimeter extension of . Now, since , the set
has positive -measure. Take . Since , which was bounded by a graph of an -Lipschitz map, the set
does not intersect . If there exists a small radius for which
we conclude that there exists a connected component of for which contradicting the fact that . Hence, there exists a sequence of points such that . Since is -Lipschitz, writing , we have
By the measure density (Proposition 2.3), we have
for all . Thus,
giving that
which contradicts the fact that . ∎
Let us point out that if in addition we require to be planar and simply connected in the previous corollary, we would get the stronger fact that .
In the next section we will show that in the planar case the conclusion of Corollary 4.3 is also a sufficient condition for a bounded -extension domain to be a -extension domain. The following example shows that this is not the case in dimension three.
Example 4.4.
Let us construct a bounded -extension domain which is not a -extension domain so that consists of only one component for which . Consequently, in the statement of Corollary 4.3 we have .
Let be a Cantor set with and let
The fact that consists of only one component for which is immediate from the construction.
Also, with the same arguments as in the previous two corollaries, we see that
does not have a strong perimeter extension.
In order to see that is a -extension domain, take . First notice that since the parts
have Lipschitz boundaries, similarly to [6, Theorem 5.8] we can consider the zero extension of both and to the whole and calling them and respectively, we have with
(4.1) |
for every . Here
for are bounded linear operators, called the traces, which are defined as
for -almost every . Now it is easy to check, following (4.1),
for . To conclude, we just let our extension operator be , which is the zero extension of outside .
In the case where , the study of extension domains is the same as the study of closed removable sets. Notice that by the measure density (Proposition 2.3) the Lebesgue measure of is zero for a Sobolev or -extension domain. We call a set of Lebesgue measure zero a removable set for , if as sets and for every . Similarly, we call removable for , if . We obtain the following equivalence of removability.
Corollary 4.5.
Let be a closed set of Lebesgue measure zero. Then is removable for if and only if is removable for .
Proof.
Suppose is removable for . Then is purely -unrectifiable. Otherwise, similarly as in the proof of Corollary 4.3, we can construct a set of finite perimeter so that . Hence, , contradicting the assumption that is removable for . Now, since is removable for , for every radius , the set is a -extension domain. Since is purely -unrectifiable, trivially has the strong perimeter extension property and is thus a -extension domain by Theorem 1.3. Consequently, is removable for .
Suppose then that is removable for . Let . We only need to check that the function when seeing as a function defined on the whole , satisfies . With the Whitney smoothing operator from Theorem 3.1 we can modify to be a -function on and moreover, by (3.2),
where can be defined as any value on . Since is removable for , we have . Thus because and therefore
and we get that with . ∎
5. Characterization of planar -extension domains
In this section we prove Theorem 1.4 using the higher dimensional result stated in Theorem 1.3. Since the necessity part of Theorem 1.4 holds in the higher-dimensional case by Corollary 4.3, we only need to prove the sufficiency. We first set some notations and definitions.
We say that is a Jordan curve if for some , , and some continuous map , injective on and such that . Accordingly to the famous Jordan curve theorem any Jordan curve splits in exactly two connected components, a bounded one and an unbounded one that we call and respectively. We will often talk about rectifiable Jordan curves , for which we mean that is a Jordan curve and it is -rectifiable. A set whose boundary is a Jordan curve is called a Jordan domain.
For technical reasons we also add to the class of Jordan curves the formal ”Jordan” curves and , whose interiors are and the empty set respectively and for which we set .
We say that a set has a decomposition into other sets up to -measure zero sets if
and for every .
For the particular case of planar sets of finite perimeter we have the following decomposition theorem from [1, Corollary 1].
Theorem 5.1.
Let have finite perimeter. Then, there exists a unique decomposition of into rectifiable Jordan curves , up to -measure zero sets, such that
-
(1)
Given , , , they are either disjoint or one is contained in the other; given , , , they are either disjoint or one is contained in the other. Each is contained in one of the .
-
(2)
.
-
(3)
If , , then there is some rectifiable Jordan curve such that . Similarly, if , , then there is some rectifiable Jordan curve such that .
-
(4)
Setting the sets are pairwise disjoint, indecomposable and .
Since sets of finite perimeter are defined via the total variation of -functions, they are understood modulo -dimensional measure zero sets. In particular, the last equality in (4) of Theorem 5.1 is modulo measure zero sets.
In order to prove the sufficiency part of Theorem 1.4 we will proceed as follows: Starting from a set of finite perimeter we first find an extension to using the fact that is a -extension domain. Then we decompose using Theorem 5.1 and after proving the quasiconvexity of each of the open connected components of , we will be able to perturb the Jordan curves of the decomposition of around each so that we get a final set which will be a strong extension of . An application of Theorem 1.3 will conclude the proof.
We start by presenting a couple of lemmas showing the quasiconvexity of all the connected components of .
Lemma 5.2.
Suppose that is a bounded -extension domain. Then there exists a constant so that for any connected component of , any two points can be connected by a curve with .
Proof.
One can essentially follow step by step the proof of [14, Theorem 1.1], once we have taken into account some facts.
-
(1)
For a given , since is a -extension domain, so is . As an extension operator we can take
where is the extension operator from to . Let us explain more in detail why our resulting function is well-defined as a function in . Observe that the closures of the different components can only intersect between themselves in just one point. That is,
(5.1) Otherwise we would be losing the connectedness of or either and are the same component. This means that is a countable set. Once we are aware of this simple fact it is clear that behaves well around and it belongs to .
Observe that since is a -extension domain there is a constant for which the property of extension of sets of finite perimeter holds. Note that this constant only depends on the norm , which only depends on and which in turn only depends on the constant of the same property but now applied to the -extension domain .
-
(2)
We can assume that is bounded, and hence also a -extension domain thanks to [14, Lemma 2.1]. If was not bounded then had to be bounded and we can take a large enough radius so that
It is clear that changing by does not affect the -extension property.
The proof of [14, Theorem 1.1] is made under the assumptions that a set is a bounded simply connected -extension domain, reaching as a conclusion that is quasiconvex.
In the case was unbounded, will be a bounded simply connected -extension domain and we apply the previous result directly to show the quasiconvexity of .
If was bounded, after the modification mentioned above, will be a bounded -extension domain. To prove the quasiconvexity of in [14, Theorem 1.1] the simply connectedness was just used at the following point: when we take two points and join them with a line-segment , the set consists on the disjoint union of countably many line-segments , with . Now, under the assumption of simply connectedness of one can assert that has two disjoint connected components. However, in our case this is still true because otherwise would not be connected.
The previous facts yield that every set is quasiconvex. A careful reading of the proof [14, Theorem 1.1] also shows that the constant of quasiconvexity of all these sets is uniformly bounded by a constant , independent of . Indeed, the quasiconvexity constant of any set only depends on the constant of the extension property of sets of finite perimeter for the -extension domains , which, as we already noted, depends only on the constant for the -extension domain independently of what we are fixing. ∎
Notice that the previous Lemma 5.2 implies, in particular, that if is a bounded -extension domain, then all open connected components of are Jordan domains.
We record the following general lemma which might be of independent interest. A version of it for quasiconvex sets was proven via conformal maps in [16]. Let us also point out that with the sharp Painleve-length result for a connected set [19] one could quite easily prove a version of the lemma with a multiplicative constant .
Lemma 5.3.
Let be a Jordan domain. For every , every and any rectifiable curve joining to there exists a curve joining to so that my g
Proof.
Without loss of generality, we may assume that minimizes the length of curves joining to in , , , and that has unit speed.
If , we are done. Suppose this is not the case and define
and
If , by minimality the curve is the concatenation of line-segments and . In this case, for small , the curve divides into two parts so that one of them is a subset of . Thus, we may replace part of by an arc of the circle , and we are done.
We are then left with the more substantial case where . Since is a Jordan loop, the set consists of two connected components and .
We will show that can be slightly pushed away from in directions that change in a locally Lipschitz way in . Namely, we assert that there exist functions
so that and are locally Lipschitz continuous and satisfy for all and .
In order to show this, let . If , then with we have for all , and .
Suppose then that . Without loss of generality we may assume that . The concatenation of with forms a closed loop so that one of the components of its complement is contained in , and . Now, let . Then, by minimality of , the set is contained on the boundary of a convex set with non-empty interior. Consequently, there exists a constant so that for any for which the outer normal vectors and to exist at and respectively, there is a Lipschitz map so that , , and for all and .
Write to be the points where a normal direction to exists at . Now, cover with the intervals and then take a subcover that is finite for compact subsets of , and so that every belongs to at most two intervals . Assume the intervals are in order, that is only intersects and . By dividing into smaller intervals if needed, we may also assume that if and , then for or . This allows us to select the normal directions in a way so that they agree for the intervals and at the points . Notice that for for which we have to make a choice between two opposite directions.
We will then have subset with , and an open covering of of multiplicity at most two, where are intervals, so that
-
•
for every there exists a constant so that for every , , there is a Lipschitz map so that for all and ,
-
•
if and have been defined as above, .
For each we will now fix a , and define on . A locally Lipschitz choice for can be given by defining
when .
Let be such that . Then, for any , the function is Lipschitz in and . Hence, if we define
we have for all , and also for every . Now if we let
defining
we get a function such that and for all . Note that the function is continuous as a limit of an absolutely and uniformly convergent series of continuous functions, and it is differentiable except on . Thus, defined by
is a curve joining and , and
finishing the proof. ∎
The next lemma, together with Theorem 5.1, are the key tools for our proof of the sufficiency part of Theorem 1.4, that we will show afterwards.
Lemma 5.4.
Let be a bounded -extension domain and the open connected components of . Suppose that the set is purely -unrectifiable and let be a Jordan domain with rectifiable. Then there exists a set of finite perimeter so that
-
(i)
,
-
(ii)
, and
-
(iii)
,
where the constant is absolute.
Proof.
Consider the at most countably many components of . For each we want to modify the set in to get some with rectifiable so that
(5.2) |
and
(5.3) |
Let us show how to conclude the proof of the lemma after assuming these facts. Since we are not changing the set inside the property (i) is clear. To check (ii) let us first write
We will estimate each of these terms separately. For the first one is clear that . For the second one we use the fact that is rectifiable, that is purely -unrectifiable and (5.2),
(5.4) |
For the third term we use (5.3) to get
All these estimates together yield
Since is at most countable by (5.1), we conclude that
proving (ii). Finally (iii) has already been shown in (5).
If , we may skip this and move to the next. Let us thus assume . Let be a parameterisation of the boundary by a homeomorphism. By the Lebesgue density theorem, for almost every there exists a so that for all
(5.5) |
By the Vitali covering lemma, we then find a disjointed collection so that (5.5) holds for each of the balls and
Now, we define for each and obtain a collection of closed arcs in whose interiors are pairwise disjoint,
and
for every .

For the next argument we have fixed. The set consists of at most countably many open curves . For each for which , we use Lemma 5.2 to find a curve such that where and are the endpoints of . Now, for , Lemma 5.3 provides us with another curve so that
(5.6) |
The curves and enclose a bounded subset that we call . Similarly, if we let be the first, and the last point of we again use Lemmas 5.2 and 5.3 to connect to with a curve so that
(5.7) |
Let be the bounded set enclosed by (from to ) and by . Now, we will modify by considering
See Figure 2 for an illustration of the modification.
Proof of Theorem 1.4.
One direction is proven in Corollary 4.3. Thus we only need to prove the converse. Thus, assume that is a bounded -extension domain and that the set is purely -unrectifiable, where are the open connected components of .
We will show that has the strong extension property for sets of finite perimeter and hence, by Theorem 1.3, will be a -extension domain. Using the fact that is a bounded -extension domain if we let be a set of finite perimeter in then there exists an extension to so that . This extension can be obtained for instance by the Maz’ya and Burago result [20, Section 9.3].
Let now be the rectifiable Jordan curves of Theorem 5.1 for the set . By applying Lemma 5.4, each Jordan domain can be replaced by a set so that , , and . Similarly, each can be replaced by a set so that , , and .
Now,
holds modulo a measure zero set. Thus, the set
is an extension of to , and
Since,
the set is the strong extension of that we had to find. ∎
Acknowledgements
The authors thank Panu Lahti for several comments on an earlier version of this paper.
References
- [1] L. Ambrosio, V. Caselles, S. Masnou and J.-M. Morel, Connected components of sets of finite perimeter and applications to image processing. J. Eur. Math. Soc. 3, 39–92 (2001).
- [2] L. Ambrosio, N. Fusco and D. Pallara, Functions of bounded variation and free discontinuity problems. Oxford Mathematical Monographs. The Clarendon Press, Oxford University Press, New York, 2000.
- [3] B. Bojarski, P. Hajłasz, P. Strzelecki, Improved approximation of higher order Sobolev functions in norm and capacity. Indiana Univ. Mat. J. 51 (2002), 507–540.
- [4] Yu. D. Burago and V. G. Maz’ya, Certain questions of potential theory and function theory for regions with irregular boundaries (Russian), Zap. Nauc̆n. Sem. Leningrad. Otdel. Mat. Inst. Steklov (LOMI) 3 (1967) 152pp.
- [5] A. P. Calderón, Lebesgue spaces of differentiable functions and distributions, in Proc. Symp. Pure Math., Vol. IV, 1961, 33–49.
- [6] L. C. Evans and R. F. Gariepy, Measure theory and fine properties of functions. Revised edition, Textbooks in Mathematics. CRC Press, Boca Raton, FL (2015), xiv+299 pp.
- [7] E. De Giorgi, Nuovi teoremi relativi alle misure -dimensionali in uno spazio a dimensioni, Ricerche Mat. 4 (1955), 95–113.
- [8] H. Federer, Geometric Measure Theory. Berlin, Heidelberg, New York: Springer 1969.
- [9] P. Hajłasz and J. Kinnunen, Holder quasicontinuity of Sobolev functions on metric spaces. Rev. Mat. Iberoamericana, 14 (1998), 601–622.
- [10] P. Hajłasz, P. Koskela, and H. Tuominen, Sobolev embeddings, extensions and measure density condition, J. Funct. Anal. 254 (2008), no. 5, 1217–1234.
- [11] H. Hakkarainen, J. Kinnunen, P. Lahti and P. Lehtelä, Relaxation and Integral Representation for Functionals of Linear Growth on Metric Measure spaces, Anal. Geom. Metr. Spaces 2016 4, 288–313.
- [12] P. W. Jones, Quasiconformal mappings and extendability of Sobolev functions, Acta Math. 47 (1981), 71–88.
- [13] P. Koskela, Capacity extension domains, Ann. Acad. Sci. Fenn. Ser. A I Math. Dissertationes No. 73 (1990), 42 pp.
- [14] P. Koskela, M. Miranda Jr. and N. Shanmugalingam, Geometric properties of planar BV extension domains, in: Around the Research of Prof. Maz’ya I, in: International Mathematical Series, 2010, pp.255–272, Function Spaces; Topics (Springer collection).
- [15] P. Koskela, T. Rajala and Yi Ru-Ya Zhang, A geometric characterization of planar Sobolev extension domains, preprint.
- [16] P. Koskela, T. Rajala and Y. Zhang, Planar -extension domains, preprint.
- [17] P. Lahti, Extensions and traces of functions of bounded variation on metric spaces, J. Math. Anal. Appl. 423(1) (2015) 521–537.
- [18] P. Lahti, X. Li and Z. Wang, Traces of Newton-Sobolev, Hajłasz-Sobolev, and BV functions on metric spaces, Ann. Sc. Norm. Super. Pisa Cl. Sci. (5), accepted.
- [19] D. Lučić, E. Pasqualetto and T. Rajala, Sharp estimate on the inner distance in planar domains, Ark. Mat., 58 (2020), 133–159.
- [20] V. Maz’ya, Sobolev spaces with applications to elliptic partial differential equations. Second, revised and augmented edition. Grundlehren der Mathematischen Wissenschaften, 342. Springer, Heidelberg, 2011. xxviii+866 pp.
- [21] P. Shvartsman, Local approximations and intrinsic characterization of spaces of smooth functions on regular subsets of , Math. Nachr. 279 (2006), no. 11, 1212–1241.
- [22] P. Shvartsman, On Sobolev extension domains in , J. Funct. Anal. 258 (2010), no. 7, 2205–2245.
- [23] E. M. Stein, Singular integrals and differentiability properties of functions, Princeton University Press, Princeton, New Jersey, 1970.