1 Introduction

1.1 Motivations

In the recent years, there has been a great surge of interest toward Sobolev spaces of fractional order. This is a very classical topic, essentially initiated by the Russian school in the 1950s of the last century, with the main contributions given by Besov, Lizorkin, Nikol’skiĭ, Slobodeckiĭ and their collaborators. Nowadays, we have a lot of monographies at our disposal on the subject. We just mention the books by Adams [1, 2], by Nikol’skiĭ [25] and by Triebel [30,31,32]. We also refer the reader to [31, Chapter 1] for an historical introduction to the subject.

The reason for this revival lies in the fact that fractional Sobolev spaces seem to play a fundamental role in the study and description of a vast amount of phenomena, involving nonlocal effects. Phenomena of this type have a wide range of applications; we refer to [10] for an overview.

There are many ways to introduce fractional derivatives and, consequently, Sobolev spaces of fractional order. Without any attempt of completeness, let us mention the two approaches which are of interest for our purposes:

  • a concrete approach, based on the introduction of explicit norms, which are modeled on the case of Hölder spaces. For example, by using the heuristic

    $$\begin{aligned} \delta _h^s u(x):=\frac{u(x+h)-u(x)}{|h|^s}\sim \hbox { ``derivative of order }{\textit{s}}\hbox {''},\quad \hbox { for }x,h\in {\mathbb {R}}^N, \end{aligned}$$

    a possible choice of norm is

    $$\begin{aligned} \left( \int \left\| \delta _h^s u\right\| _{L^p}^p\,\frac{\mathrm{d}h}{|h|^N}\right) ^\frac{1}{p}, \end{aligned}$$

    and more generally

    $$\begin{aligned} \left( \int \left\| \delta _h^s u\right\| _{L^p}^q\,\frac{\mathrm{d}h}{|h|^N}\right) ^\frac{1}{q},\quad \hbox { for } 1\le q\le \infty . \end{aligned}$$

    Observe that the integral contains the singular kernel \(|h|^{-N}\); thus, functions for which the norm above is finite must be better than just merely s-Hölder regular, in an averaged sense;

  • an abstract approach, based on the so-called interpolation methods. The foundations of these methods were established at the beginning of the 1960s of the last century, by Calderón, Gagliardo, Krejn, Lions and Petree, among others. A comprehensive treatment of this approach can be found for instance in the books [3, 4, 29] and references therein.

    In a nutshell, the idea is to define a scale of “intermediate spaces” between \(L^p\) and the standard Sobolev space \(W^{1,p}\), by means of a general abstract construction. The main advantage of this second approach is that many of the properties of the spaces constructed in this way can be extrapolated in a direct way from those of the two “endpoint” spaces \(L^p\) and \(W^{1,p}\).

As mentioned above, actually other approaches are possible: a possibility is to use the Fourier transform. Another particularly elegant approach consists in taking the convolution with a suitable kernel (for example, heat or Poisson kernels are typical choices) and looking at the rate of blowup of selected \(L^p\) norms with respect to the convolution parameter. However, we will not consider these constructions in the present paper; we refer the reader to [31] for a wide list of definitions of this type.

In spite of the explosion of literature on Calculus of Variations settled in fractional Sobolev spaces of the last years, the abstract approach based on interpolation seems to have been completely neglected or, at least, overlooked. For example, the well-known survey paper [14], which eventually became a standard reference on the field, does not even mention interpolation techniques.

1.2 Aims

The main scope of this paper is to revitalize some interest toward interpolation theory in the context of fractional Sobolev spaces. In doing this, we will resist the temptation of any unnecessary generalization. Rather, we will focus on a particular, yet meaningful, question which can be resumed as follows:

figure a

We can already anticipate the conclusions of the paper and say that this is not always true. Let us now try to enter more in the details of the present paper.

Our concerns involve the so-called homogeneous fractional Sobolev–Slobodeckiĭ spaces\({\mathcal {D}}^{s,p}_0(\varOmega )\). Given an open set \(\varOmega \subset {\mathbb {R}}^N\), an exponent \(1\le p <\infty \) and a parameter \(0<s<1\), this space is defined as the completion of \(C_0^\infty (\varOmega )\) with respect to the norm

$$\begin{aligned} u\mapsto [u]_{W^{s,p}({\mathbb {R}}^N)}:=\left( \iint _{{\mathbb {R}}^N\times {\mathbb {R}}^N} \frac{|u(x)-u(y)|^p}{|x-y|^{N+s\,p}}\,\mathrm{d}x\,\mathrm{d}y\right) ^\frac{1}{p}. \end{aligned}$$

Such a space is the natural fractional counterpart of the homogeneous Sobolev space \({\mathcal {D}}^{1,p}_0(\varOmega )\), defined as the completion of \(C^\infty _0(\varOmega )\) with respect to the norm

$$\begin{aligned} u\mapsto \left( \int _\varOmega |\nabla u|^p\,\mathrm{d}x\right) ^\frac{1}{p}. \end{aligned}$$

The space \({\mathcal {D}}^{1,p}_0(\varOmega )\) has been first studied by Deny and Lions in [13], among others. We recall that \({\mathcal {D}}^{1,p}_0(\varOmega )\) is a natural setting for studying variational problems of the type

$$\begin{aligned} \inf \left\{ \frac{1}{p}\,\int _\varOmega |\nabla u|^p\,\mathrm{d}x-\int _\varOmega f\,u\,\mathrm{d}x \right\} , \end{aligned}$$

supplemented with Dirichlet boundary conditions, in the absence of regularity assumptions on the boundary \(\partial \varOmega \). In the same way, the space \({\mathcal {D}}^{s,p}_0(\varOmega )\) is the natural framework for studying minimization problems containing functionals of the type

$$\begin{aligned} \frac{1}{p}\,\iint _{{\mathbb {R}}^N\times {\mathbb {R}}^N} \frac{|u(x)-u(y)|^p}{|x-y|^{N+s\,p}}\,\mathrm{d}x\,\mathrm{d}y-\int _\varOmega f\,u\,\mathrm{d}x, \end{aligned}$$
(1.1)

in the presence of nonlocal Dirichlet boundary conditions, i.e., the values of u are prescribed on the whole complement \({\mathbb {R}}^N{{\setminus }}\varOmega \). Observe that even if this kind of boundary conditions may look weird, these are the correct ones when dealing with energies (1.1), which take into account interactions “from infinity.”

The connection between the two spaces \({\mathcal {D}}^{1,p}_0(\varOmega )\) and \({\mathcal {D}}^{s,p}_0(\varOmega )\) is better appreciated by recalling that for \(u\in C^\infty _0(\varOmega )\), we have (see [5] and [26, Corollary 1.3])

$$\begin{aligned} \lim _{s\nearrow 1} (1-s)\,\iint _{{\mathbb {R}}^N\times {\mathbb {R}}^N} \frac{|u(x)-u(y)|^p}{|x-y|^{N+s\,p}}\,\mathrm{d}x\,\mathrm{d}y=\alpha _{N,p}\,\int _{\varOmega } |\nabla u|^p\,\mathrm{d}x, \end{aligned}$$

with

$$\begin{aligned} \alpha _{N,p}=\frac{1}{p}\,\int _{\mathbb {S}^{N-1}}|\langle \omega ,{\mathbf {e}}_1\rangle |^p\,d{\mathcal {H}}^{N-1}(\omega ),\quad {\mathbf {e}}_1=(1,0,\ldots ,0). \end{aligned}$$

On the other hand, as \(s\searrow 0\) we have (see [24, Theorem 3])

$$\begin{aligned} \lim _{s\searrow 0} s\,\iint _{{\mathbb {R}}^N\times {\mathbb {R}}^N} \frac{|u(x)-u(y)|^p}{|x-y|^{N+s\,p}}\,\mathrm{d}x\,\mathrm{d}y=\beta _{N,p}\,\int _{\varOmega }|u|^p\,\mathrm{d}x, \end{aligned}$$

with

$$\begin{aligned} \beta _{N,p}=\frac{2\,N\,\omega _N}{p}, \end{aligned}$$

and \(\omega _N\) is the volume of the N-dimensional unit ball. These two results reflect the “interpolative” nature of the space \({\mathcal {D}}^{s,p}_0(\varOmega )\), which will be, however, discussed in more detail in the sequel.

Indeed, one of our goals is to determine whether \({\mathcal {D}}^{s,p}_0(\varOmega )\) coincides or not with the real interpolation space\({\mathcal {X}}^{s,p}_0(\varOmega )\) defined as the completion of \(C_0^\infty (\varOmega )\) with respect to the norm

$$\begin{aligned} \Vert u\Vert _{{\mathcal {X}}^{s,p}_0(\varOmega )}:=\left( \int _0^{+\infty } \left( \frac{K(t,u,L^p(\varOmega ),{\mathcal {D}}^{1,p}_0(\varOmega ))}{t^s}\right) ^p\,\frac{\mathrm{d}t}{t}\right) ^\frac{1}{p}. \end{aligned}$$

Here \(K(t,\cdot ,L^p(\varOmega ),{\mathcal {D}}^{1,p}_0(\varOmega ))\) is the K-functional associated with the spaces \(L^p(\varOmega )\) and \({\mathcal {D}}^{1,p}_0(\varOmega )\), see Sect. 3 for more details.

In particular, we will be focused on obtaining double-sided norm inequalities leading to answer our initial question, i.e., estimates of the form

$$\begin{aligned} \frac{1}{C}\,[u]_{W^{s,p}({\mathbb {R}}^N)}\le \Vert u\Vert _{{\mathcal {X}}^{s,p}_0(\varOmega )}\le C\,[u]_{W^{s,p}({\mathbb {R}}^N)},\quad u\in C^\infty _0(\varOmega ). \end{aligned}$$

Moreover, we compute carefully the dependence on the parameter s of the constant C. Indeed, we will see that C can be taken independent of s.

1.3 Results

We now list the main achievements of our discussion:

  1. 1.

    the space \({\mathcal {D}}^{s,p}_0(\varOmega )\) is always larger than \({\mathcal {X}}^{s,p}_0(\varOmega )\) (see Proposition 4.1) and they do not coincide for general open sets, as we exhibit with an explicit example (see Example 4.4);

  2. 2.

    they actually coincide on a large class of domains, i.e., bounded convex sets (Theorem 4.7), convex cones (Corollary 4.8), Lipschitz sets (Theorem 4.10);

  3. 3.

    the Poincaré constants for the embeddings

    $$\begin{aligned} {\mathcal {D}}^{s,p}_0(\varOmega )\hookrightarrow L^p(\varOmega ) \quad \hbox {and}\quad {\mathcal {D}}^{1,p}_0(\varOmega )\hookrightarrow L^p(\varOmega ), \end{aligned}$$

    are equivalent for the classes of sets at point 2 (Theorem 6.1). More precisely, by setting

    $$\begin{aligned} \lambda ^s_p(\varOmega )=\inf _{u\in C_0^\infty (\varOmega )}\Big \{ [u]^p_{W^{s,p}(\mathbb {R}^N)}\ :\ \Vert u\Vert _{L^p(\varOmega )}=1\Big \},\quad 0<s<1, \end{aligned}$$

    and

    $$\begin{aligned} \lambda ^1_p(\varOmega )=\inf _{u\in C_0^\infty (\varOmega )}\left\{ \int _\varOmega |\nabla u|^p\,\mathrm{d}x\, :\, \Vert u\Vert _{L^p(\varOmega )}=1\right\} , \end{aligned}$$

    we have

    $$\begin{aligned} \frac{1}{C}\,\Big (\lambda ^1_p(\varOmega )\Big )^s\le s\,(1-s)\,\lambda ^s_p(\varOmega )\le C\, \Big (\lambda ^1_p(\varOmega )\Big )^s. \end{aligned}$$

    Moreover, on convex sets the constant \(C>0\) entering in the relevant estimate is universal, i.e., it depends onNandponly. On the other hand, we show that this equivalence fails if we drop any kind of regularity assumptions on the sets (see Remark 6.3).

As a byproduct of our discussion, we also highlight some weird and unnatural behaviors of the interpolation space \({\mathcal {X}}^{s,p}_0(\varOmega )\):

  • the “extension by zero” operator \({\mathcal {X}}^{s,p}_0(\varOmega )\hookrightarrow {\mathcal {X}}_0^{s,p}({\mathbb {R}}^N)\) is not an isometry for general open sets (see Remark 4.6) and the two norms

    $$\begin{aligned} \Vert \,\cdot \,\Vert _{{\mathcal {X}}^{s,p}_0(\varOmega )}\quad \hbox {and}\quad \Vert \,\cdot \,\Vert _{{\mathcal {X}}^{s,p}_0({\mathbb {R}}^N)}, \end{aligned}$$

    may not be equivalent on \(C^\infty _0(\varOmega )\). This is in contrast with what happens for the spaces \(L^p(\varOmega )\), \({\mathcal {D}}^{1,p}_0(\varOmega )\) and \({\mathcal {D}}^{s,p}_0(\varOmega )\);

  • the sharp Poincaré interpolation constant

    $$\begin{aligned} \varLambda ^s_p(\varOmega )=\inf _{u\in C^\infty _0(\varOmega )} \Big \{ \Vert u\Vert _{{\mathcal {X}}^{s,p}_0(\varOmega )}^p\ :\ \Vert u\Vert _{L^p(\varOmega )}=1\Big \},\quad 0<s<1 \end{aligned}$$

    is sensitive to removing sets with zero capacity. In other words, if we remove a compact set \(E\Subset \varOmega \) having zero capacity in the sense of \({\mathcal {X}}^{s,p}_0(\varOmega )\), it may happen that (see Lemma 5.4)

    $$\begin{aligned} \varLambda ^s_p(\varOmega {\setminus } E)>\varLambda ^s_p(\varOmega ). \end{aligned}$$

    Again, this is in contrast with the case of \({\mathcal {D}}^{1,p}_0(\varOmega )\) and \({\mathcal {D}}^{s,p}_0(\varOmega )\).

Remark 1.1

As recalled at the beginning, nowadays there is a huge literature on Sobolev spaces of fractional order. Nevertheless, to the best of our knowledge, a detailed discussion on the space \({\mathcal {D}}^{s,p}_0(\varOmega )\) in connection with interpolation theory seems to be missing. For this reason, we believe that our discussion is of independent interest.

We also point out that for Sobolev spaces of functions not necessarily vanishing at the boundary, there is a very nice paper [11] by Chandler-Wilde, Hewett and Moiola comparing “concrete” constructions with the interpolation one.

1.4 Plan of the paper

In Sect. 2 we present the relevant Sobolev spaces, constructed with the concrete approach based on the so-called Sobolev–Slobodeckiĭ norms. Then in Sect. 3 we introduce the homogeneous interpolation space we want to work with. Essentially, no previous knowledge of interpolation theory is necessary.

The comparison between the concrete space and the interpolation one is contained in Sect. 4. This in turn is divided in three subsections, each one dealing with a different class of open sets. We point out here that we preferred to treat convex sets separately from Lipschitz sets, for two reasons: the first one is that for convex sets the comparison between the two spaces can be done “by hands,” without using any extension theorem. This in turn permits to have a better control on the relevant constants entering in the estimates. The second one is that in proving the result for Lipschitz sets, we actually use the result for convex sets.

In order to complement the comparison between the two spaces, in Sect. 5 we compare the two relevant notions of capacity, naturally associated with the norms of these spaces. Finally, Sect. 6 compares the Poincaré constants.

The paper ends with three appendices: the first one contains the construction of a counterexample used throughout the whole paper; the second one proves a version of the one-dimensional Hardy inequality; and the last one contains a geometric expedient result for convex sets.

2 Preliminaries

2.1 Basic notation

In what follows, we will always denote by N the dimension of the ambient space. For an open set \(\varOmega \subset {\mathbb {R}}^N\), we indicate by \(|\varOmega |\) its N-dimensional Lebesgue measure. The symbol \({\mathcal {H}}^k\) will stand for the k-dimensional Hausdorff measure. Finally, we set

$$\begin{aligned} B_R(x_0)=\left\{ x\in {\mathbb {R}}^N\, :\, |x-x_0|<R\right\} , \end{aligned}$$

and

$$\begin{aligned} \omega _N=|B_1(0)|. \end{aligned}$$

2.2 Sobolev spaces

For \(1\le p<\infty \) and an open set \(\varOmega \subset {\mathbb {R}}^N\), we use the classical definition

$$\begin{aligned} W^{1,p}(\varOmega ):=\left\{ u\in L^p(\varOmega )\,:\, \int _\varOmega |\nabla u|^p\,\mathrm{d}x <+\infty \right\} . \end{aligned}$$

This is a Banach space endowed with the norm

$$\begin{aligned} \Vert u\Vert _{W^{1,p}(\varOmega )}=\left( \Vert u\Vert ^p_{L^p(\varOmega )}+\Vert \nabla u\Vert ^p_{L^p(\varOmega )}\right) ^\frac{1}{p}. \end{aligned}$$

We also denote by \({\mathcal {D}}^{1,p}_0(\varOmega )\) the homogeneous Sobolev space, defined as the completion of \(C^\infty _0(\varOmega )\) with respect to the norm

$$\begin{aligned} u\mapsto \Vert \nabla u\Vert _{L^p(\varOmega )}. \end{aligned}$$

If the open set \(\varOmega \subset {\mathbb {R}}^N\) supports the classical Poincaré inequality

$$\begin{aligned} c\,\int _\varOmega |u|^p\,\mathrm{d}x\le \int _\varOmega |\nabla u|^p\,\mathrm{d}x,\quad \hbox {for every }u\in C^\infty _0(\varOmega ), \end{aligned}$$

then \({\mathcal {D}}^{1,p}_0(\varOmega )\) is indeed a functional space and it coincides with the closure in \(W^{1,p}(\varOmega )\) of \(C^\infty _0(\varOmega )\). We will set

$$\begin{aligned} \lambda ^1_p(\varOmega )=\inf _{u\in C^\infty _0(\varOmega )} \Big \{\Vert \nabla u\Vert ^p_{L^p(\varOmega )}\, :\, \Vert u\Vert _{L^p(\varOmega )}=1\Big \}. \end{aligned}$$

It occurs \(\lambda ^1_p(\varOmega )=0\) whenever \(\varOmega \) does not support such a Poincaré inequality.

Remark 2.1

We remark that one could also consider the space

$$\begin{aligned} W^{1,p}_0(\varOmega ):= \left\{ u\in W^{1,p}({\mathbb {R}}^N):u = 0 \text { a.e. in } {\mathbb {R}}^N{\setminus }\varOmega \right\} . \end{aligned}$$

It is easy to see that \({\mathcal {D}}^{1,p}_0(\varOmega )\subset {W}^{1,p}_0(\varOmega )\), whenever \({\mathcal {D}}^{1,p}_0(\varOmega )\hookrightarrow L^p(\varOmega )\). If in addition \(\partial \varOmega \) is continuous, then both spaces are known to coincide, thanks to the density of \(C^\infty _0(\varOmega )\) in \(W^{1,p}_0(\varOmega )\), see [20, Theorem 1.4.2.2].

2.3 A homogeneous Sobolev–Slobodeckiĭ space

Given \(0<s< 1\) and \(1\le p <\infty \), the fractional Sobolev space \(W^{s,p}({\mathbb {R}}^N)\) is defined as

$$\begin{aligned} W^{s,p}({\mathbb {R}}^N) := \left\{ u\in L^p({\mathbb {R}}^N)\,:\, [u]_{W^{s,p}({\mathbb {R}}^N)}<+\infty \right\} , \end{aligned}$$

where the Sobolev–Slobodeckiĭ seminorm\([\,\cdot \,]_{W^{s,p}({\mathbb {R}}^N)}\) is defined as

$$\begin{aligned}{}[u]_{W^{s,p}({\mathbb {R}}^N)} := \left( \iint _{{\mathbb {R}}^N\times {\mathbb {R}}^N} \frac{|u(x)-u(y)|^p}{|x-y|^{N+s\,p}}\, \mathrm{d}x\,\mathrm{d}y\right) ^\frac{1}{p}. \end{aligned}$$

This is a Banach space endowed with the norm

$$\begin{aligned} \Vert u\Vert _{W^{s,p}({\mathbb {R}}^N)} = \left( \Vert u\Vert _{L^p({\mathbb {R}}^N)}^p + [u]_{W^{s,p}({\mathbb {R}}^N)}^p\right) ^\frac{1}{p}. \end{aligned}$$

In what follows, we need to consider nonlocal homogeneous Dirichlet boundary conditions, outside an open set \(\varOmega \subset {\mathbb {R}}^N\). In this setting, it is customary to consider the homogeneous Sobolev–Slobodeckiĭ space\({\mathcal {D}}^{s,p}_0(\varOmega )\). The latter is defined as the completion of \(C^\infty _0(\varOmega )\) with respect to the norm

$$\begin{aligned} u\mapsto [u]_{W^{s,p}({\mathbb {R}}^N)}. \end{aligned}$$

Observe that the latter is indeed a norm on \(C^\infty _0(\varOmega )\). Whenever the open set \(\varOmega \subset {\mathbb {R}}^N\) admits the following Poincaré inequality

$$\begin{aligned} c\,\int _\varOmega |u|^p\,\mathrm{d}x\le \iint _{{\mathbb {R}}^N\times {\mathbb {R}}^N} \frac{|u(x)-u(y)|^p}{|x-y|^{N+s\,p}}\, \mathrm{d}x\,\mathrm{d}y, \quad \hbox {for every }u\in C_0^\infty (\varOmega ), \end{aligned}$$

we get that \({\mathcal {D}}^{s,p}_0(\varOmega )\) is a functional space continuously embedded in \(L^p(\varOmega )\). In this case, it coincides with the closure in \(W^{s,p}({\mathbb {R}}^N)\) of \(C^\infty _0(\varOmega )\). We endow the space \({\mathcal {D}}^{s,p}_0(\varOmega )\) with the norm

$$\begin{aligned} \Vert u\Vert _{{\mathcal {D}}^{s,p}_0(\varOmega )}:=[u]_{W^{s,p}({\mathbb {R}}^N)}. \end{aligned}$$

We also define

$$\begin{aligned} \lambda ^s_p(\varOmega )=\inf _{u\in C^\infty _0(\varOmega )} \Big \{\Vert u\Vert ^p_{{\mathcal {D}}^{s,p}_0(\varOmega )}\, :\, \Vert u\Vert _{L^p(\varOmega )}=1\Big \}, \end{aligned}$$

i.e., this is the sharp constant in the relevant Poincaré inequality. Some embedding properties of the space \({\mathcal {D}}^{s,p}_0(\varOmega )\) are investigated in [18].

Remark 2.2

As in the local case, one could also consider the space

$$\begin{aligned} W^{s,p}_0(\varOmega ):= \left\{ u\in W^{s,p}({\mathbb {R}}^N):u = 0 \text { a.e. in } {\mathbb {R}}^N{\setminus }\varOmega \right\} . \end{aligned}$$

It is easy to see that \({\mathcal {D}}^{s,p}_0(\varOmega )\subset {W}^{s,p}_0(\varOmega )\), whenever \({\mathcal {D}}^{s,p}_0(\varOmega )\hookrightarrow L^p(\varOmega )\). As before, if \(\partial \varOmega \) is continuous, then both spaces are known to coincide, again thanks to the density of \(C^\infty _0(\varOmega )\) in \(W^{s,p}_0(\varOmega )\), see [20, Theorem 1.4.2.2].

2.4 Another space of functions vanishing at the boundary

Another natural fractional Sobolev space of functions “vanishing at the boundary” is given by the completion of \(C^\infty _0(\varOmega )\) with respect to the localized norm

$$\begin{aligned}{}[u]_{W^{s,p}(\varOmega )}=\left( \iint _{\varOmega \times \varOmega } \frac{|u(x)-u(y)|^p}{|x-y|^{N+s\,p}}\,\mathrm{d}x\,\mathrm{d}y\right) ^\frac{1}{p}. \end{aligned}$$

We will denote this space by \(\mathring{D}^{s,p}(\varOmega )\) and endow it with the norm \([\Vert u\Vert _{\mathring{D}^{s,p}(\varOmega )}:=[u]_{W^{s,p}(\varOmega )}]\). We recall the following

Lemma 2.3

Let \(1<p<\infty \) and \(0<s<1\). For every \(\varOmega \subset {\mathbb {R}}^N\) open bounded Lipschitz set, we have:

  • if \(s\,p>1\), then

    $$\begin{aligned} {\mathcal {D}}^{s,p}_0(\varOmega )=\mathring{D}^{s,p}(\varOmega ); \end{aligned}$$
  • if \(s\,p\le 1\), then there exists a sequence \(\{u_n\}_{n\in {\mathbb {N}}}\subset C^\infty _0(\varOmega )\) such that

    $$\begin{aligned} \lim _{n\rightarrow \infty } \frac{\Vert u_n\Vert _{\mathring{D}^{s,p}(\varOmega )}}{\Vert u_n\Vert _{{\mathcal {D}}^{s,p}_0(\varOmega )}}=0. \end{aligned}$$

Proof

The proof of the first fact is contained in [7, Proposition B.1].

As for the case \(s\,p\le 1\), in [15, Section 2] Dyda constructed a sequence \(\{u_n\}_{n\in {\mathbb {N}}}\subset C^\infty _0(\varOmega )\) such that

$$\begin{aligned} \lim _{n\rightarrow \infty }\Vert u_n\Vert _{\mathring{D}^{s,p}(\varOmega )}=0\quad \hbox {and}\quad \lim _{n\rightarrow \infty }\Vert u_n-1_\varOmega \Vert _{L^p(\varOmega )}=0. \end{aligned}$$

By observing that for such a sequence we have

$$\begin{aligned} \lim _{n\rightarrow \infty }\Vert u_n\Vert _{{\mathcal {D}}^{s,p}_0(\varOmega )}\ge \Big (\lambda ^s_p(\varOmega )\Big )^\frac{1}{p}\,\lim _{n\rightarrow \infty }\Vert u_n\Vert _{L^p(\varOmega )}=\Big (\lambda ^s_p(\varOmega )\,|\varOmega |\Big )^\frac{1}{p}, \end{aligned}$$

we get the desired conclusion, by observing that \(\lambda ^s_p(\varOmega )>0\) for an open bounded set, thanks to [8, Corollary 5.2]. \(\square \)

Remark 2.4

Clearly, we always have

$$\begin{aligned} \Vert u\Vert _{\mathring{D}^{s,p}(\varOmega )}\le \Vert u\Vert _{{\mathcal {D}}^{s,p}_0(\varOmega )},\quad \hbox {for every } u\in C^\infty _0(\varOmega ). \end{aligned}$$

As observed in [16], the reverse inequality

$$\begin{aligned} \Vert u\Vert _{{\mathcal {D}}^{s,p}_0(\varOmega )}\le C\, \Vert u\Vert _{\mathring{D}^{s,p}(\varOmega )},\quad \hbox {for every } u\in C^\infty _0(\varOmega ), \end{aligned}$$
(2.1)

is equivalent to the validity of the Hardy-type inequality

$$\begin{aligned} \int _\varOmega |u(x)|^p\left( \int _{{\mathbb {R}}^N{\setminus } \varOmega } |x-y|^{-N-s\,p}\,y\right) \,\mathrm{d}x\le C\,\iint _{\varOmega \times \varOmega } \frac{|u(x)-u(y)|^p}{|x-y|^{N+s\,p}}\,\mathrm{d}x\,\mathrm{d}y. \end{aligned}$$

A necessary and sufficient condition for this to happen is proved in [16, Proposition 2]. We also observe that the failure of (2.1) implies that in general the “extension by zero” operator

$$\begin{aligned} {\mathcal {T}}_0:\mathring{D}^{s,p}(\varOmega )\rightarrow \mathring{D}^{s,p}({\mathbb {R}}^N), \end{aligned}$$

is not continuous. We refer to [16] for a detailed discussion of this issue.

Remark 2.5

The space \(\mathring{D}^{s,p}(\varOmega )\) is quite problematic in general, especially in the case \(s\,p\le 1\) where it may fail to be a functional space. A more robust variant of this space is

$$\begin{aligned} \widetilde{D}^{s,p}(\varOmega )=\hbox {``closure of }C^\infty _0(\varOmega ) \hbox { in }W^{s,p}(\varOmega )\hbox {''}. \end{aligned}$$

By definition, this is automatically a functional space, continuously contained in \(W^{s,p}(\varOmega )\). It is a classical fact that if \(\varOmega \) is a bounded open set with smooth boundary, then

$$\begin{aligned} \widetilde{D}^{s,p}(\varOmega )=W^{s,p}(\varOmega ),\quad \hbox {for }s\,p< 1, \end{aligned}$$

see [32, Theorem 3.4.3]. Moreover, we also have

$$\begin{aligned} \widetilde{D}^{s,p}(\varOmega )={\mathcal {D}}^{s,p}_0(\varOmega ),\quad \hbox {for }s\,p\not =1, \end{aligned}$$

see, for example, [7, Proposition B.1].

3 An interpolation space

Let \(\varOmega \subset {\mathbb {R}}^N\) be an open set. If \(X(\varOmega )\) and \(Y(\varOmega )\) are two normed vector spaces containing \(C^\infty _0(\varOmega )\) as a dense subspace, we define for every \(t>0\) and \(u\in C^\infty _0(\varOmega )\) the K-functional

$$\begin{aligned} K(t,u,X(\varOmega ),Y(\varOmega )):=\inf _{v\in C^\infty _0(\varOmega )} \Big \{\Vert u-v\Vert _{X(\varOmega )}+t\,\Vert v\Vert _{Y(\varOmega )}\Big \}. \end{aligned}$$
(3.1)

We are interested in the following specific case: let us take \(0<s<1\) and \(1<p<\infty \), we choose

$$\begin{aligned} X(\varOmega )=L^p(\varOmega )\quad \hbox {and}\quad Y(\varOmega )={\mathcal {D}}^{1,p}_0(\varOmega ). \end{aligned}$$

Then we use the notation

$$\begin{aligned} \Vert u\Vert _{{\mathcal {X}}^{s,p}_0(\varOmega )}:=\left( \int _0^{+\infty } \left( \frac{K(t,u,L^p(\varOmega ),{\mathcal {D}}^{1,p}_0(\varOmega ))}{t^s}\right) ^p\,\frac{\mathrm{d}t}{t}\right) ^\frac{1}{p},\quad u\in C^\infty _0(\varOmega ). \end{aligned}$$

It is standard to see that this is a norm on \(C^\infty _0(\varOmega )\), see [4, Section 3.1]. We will indicate by \({\mathcal {X}}^{s,p}_0(\varOmega )\) the completion of \(C^\infty _0(\varOmega )\) with respect to this norm.

The first result is the Poincaré inequality for the interpolation space \({\mathcal {X}}^{s,p}_0(\varOmega )\). The main focus is on the explicit dependence of the constant on the local Poincaré constant \(\lambda ^1_{p}\).

Lemma 3.1

Let \(1<p<\infty \) and \(0<s<1\). Let \(\varOmega \subset {\mathbb {R}}^N\) be an open set. Then for every \(u\in C^\infty _0(\varOmega )\) we have

$$\begin{aligned} \left( \lambda ^1_{p}(\varOmega )\right) ^s\,\Vert u\Vert _{L^p(\varOmega )}^p\le p\,s\,(1-s)\,\Vert u\Vert _{{\mathcal {X}}^{s,p}_0(\varOmega )}^p. \end{aligned}$$
(3.2)

Proof

We proceed in two stages: we first prove that

$$\begin{aligned} \Vert u\Vert _{L^p(\varOmega )}^p\lesssim \int _0^{+\infty } \left( \frac{K(t,u,L^p(\varOmega ),L^p(\varOmega ))}{t^s}\right) ^p\,\frac{\mathrm{d}t}{t}, \end{aligned}$$

and then we show that the last integral is estimated from the above by the norm \({\mathcal {X}}^{s,p}_0(\varOmega )\).

First stage Let us take \(u\in C^\infty _0(\varOmega )\), for every \(t\ge 1\) and \(v\in C^\infty _0(\varOmega )\)

$$\begin{aligned} \Vert u\Vert _{L^p(\varOmega )}\le \Vert u-v\Vert _{L^p(\varOmega )}+t\,\Vert v\Vert _{L^p(\varOmega )}. \end{aligned}$$

By taking the infimum, we thus get

$$\begin{aligned} \begin{aligned} \Vert u\Vert _{L^p(\varOmega )}\le K(t,u,L^p(\varOmega ),L^p(\varOmega )). \end{aligned} \end{aligned}$$

By integrating with respect to the singular measure \(\mathrm{d}t/t\), we then get

$$\begin{aligned} \int _1^{+\infty } \left( \frac{K(t,u,L^p(\varOmega ),L^p(\varOmega ))}{t^s}\right) ^p\,\frac{\mathrm{d}t}{t}\ge \int _1^{+\infty } t^{-s\,p}\,\Vert u\Vert ^p_{L^p(\varOmega )}\,\frac{\mathrm{d}t}{t}=\frac{\Vert u\Vert _{L^p(\varOmega )}^p}{s\,p}. \end{aligned}$$
(3.3)

We now pick \(0<t<1\), by triangle inequality we get for every \(v\in C^\infty _0(\varOmega )\)

$$\begin{aligned} \begin{aligned} t\,\Vert u\Vert _{L^p(\varOmega )}&\le t\,\Vert u-v\Vert _{L^p(\varOmega )}+t\,\Vert v\Vert _{L^p(\varOmega )}\\&\le \Vert u-v\Vert _{L^p(\varOmega )}+t\,\Vert v\Vert _{L^p(\varOmega )}. \end{aligned} \end{aligned}$$

By taking the infimum over \(v\in C^\infty _0(\varOmega )\), we obtain for \(u\in C^\infty _0(\varOmega )\) and \(0<t<1\)

$$\begin{aligned} t\,\Vert u\Vert _{L^p(\varOmega )}\le K(t,u,L^p(\varOmega ),L^p(\varOmega )). \end{aligned}$$

By integrating again, we get this time

$$\begin{aligned} \int _0^{1} \left( \frac{K(t,u,L^p(\varOmega ),L^p(\varOmega ))}{t^s}\right) ^p\,\frac{\mathrm{d}t}{t}\ge \int _0^1 t^{p-s\,p}\,\Vert u\Vert ^p_{L^p(\varOmega )}\,\frac{\mathrm{d}t}{t}=\frac{\Vert u\Vert _{L^p(\varOmega )}^p}{(1-s)\,p}. \end{aligned}$$
(3.4)

By summing up (3.3) and (3.4), we get the estimate

$$\begin{aligned} \Vert u\Vert _{L^p(\varOmega )}^p\le p\,s\,(1-s)\,\int _0^{+\infty } \left( \frac{K(t,u,L^p(\varOmega ),L^p(\varOmega ))}{t^s}\right) ^p\,\frac{\mathrm{d}t}{t}. \end{aligned}$$
(3.5)

Second stage Given \(u\in C^\infty _0(\varOmega )\), we take \(v\in C^\infty _0(\varOmega )\). We can suppose that \(\lambda ^1_p(\varOmega )>0\); otherwise, (3.2) trivially holds. By definition of \(\lambda ^1_{p}(\varOmega )\), we have that

$$\begin{aligned} \Vert u-v\Vert _{L^p(\varOmega )}+t\,\Vert v\Vert _{L^p(\varOmega )} \le \Vert u-v\Vert _{L^p(\varOmega )}+t\, (\lambda _{p}^1(\varOmega ))^{-\frac{1}{p}}\,\Vert \nabla v\Vert _{L^p(\varOmega )}. \end{aligned}$$

If we recall the definition (3.1) of the K-functional, we get

$$\begin{aligned} K(t,u,L^p(\varOmega ),L^p(\varOmega ))^p \le \left( \Vert u-v\Vert _{L^p(\varOmega )}+\frac{t}{ (\lambda _{p}^1(\varOmega ))^\frac{1}{p}}\Vert \nabla v\Vert _{L^p(\varOmega )} \right) ^p, \end{aligned}$$

and by taking infimum over \(v\in C^\infty _0(\varOmega )\) and multiplying by \(t^{-s\,p}\), we get

$$\begin{aligned} t^{-s\,p}K(t,u,L^p(\varOmega ),L^p(\varOmega ))^p \le t^{-s\,p}\, K\left( \frac{t}{ (\lambda _{p}^1(\varOmega ))^\frac{1}{p}},u\,,L^p(\varOmega ),{\mathcal {D}}^{1,p}_0(\varOmega )\right) ^p. \end{aligned}$$

We integrate over \(t>0\), by performing the change of variable \(\tau =t/(\lambda _{p}^1(\varOmega ))^\frac{1}{p}\) we get

$$\begin{aligned} \begin{aligned} \int _0^{+\infty }&\left( \frac{K(t,u,L^p(\varOmega ),L^p(\varOmega ))}{t^s}\right) ^p\,\frac{\mathrm{d}t}{t}\le \frac{1}{(\lambda _{p}^1(\varOmega ))^s } \Vert u\Vert ^p_{{\mathcal {X}}^{s,p}_0(\varOmega )}. \end{aligned} \end{aligned}$$

By using this in (3.5), we prove the desired inequality (3.2). \(\square \)

We will set

$$\begin{aligned} \varLambda ^s_p(\varOmega )=\inf _{u\in C^\infty _0(\varOmega )}\left\{ \Vert u\Vert ^p_{{\mathcal {X}}^{s,p}_0(\varOmega )}\, :\, \Vert u\Vert _{L^p(\varOmega )}=1\right\} , \end{aligned}$$

i.e., this is the sharp constant in the relevant Poincaré inequality. As a consequence of (3.2), we obtain

$$\begin{aligned} \Big (\lambda ^1_{p}(\varOmega )\Big )^s\le p\,s\,(1-s)\,\varLambda ^s_p(\varOmega ). \end{aligned}$$
(3.6)

Proposition 3.2

(Interpolation inequality) Let \(1<p<\infty \) and \(0<s<1\). Let \(\varOmega \subset {\mathbb {R}}^N\) be an open set. For every \(u\in C^\infty _0(\varOmega )\) we have

$$\begin{aligned} p\,s\,(1-s)\,\Vert u\Vert _{{\mathcal {X}}^{s,p}_0(\varOmega )}^p\le \Vert u\Vert _{L^p(\varOmega )}^{p\,(1-s)}\,\Vert \nabla u\Vert _{L^p(\varOmega )}^{s\,p}. \end{aligned}$$
(3.7)

In particular, we also obtain

$$\begin{aligned} p\,s\,(1-s)\,\varLambda ^s_{p}(\varOmega )\le \Big (\lambda ^1_{p}(\varOmega )\Big )^s. \end{aligned}$$
(3.8)

Proof

We can assume that \(u\not \equiv 0\); otherwise, there is nothing prove. In the definition of the K-functional \(K(t,u,L^p(\varOmega ),{\mathcal {D}}^{1,p}_0(\varOmega ))\), we take \(v=\tau \,u\) for \(\tau >0\); thus, we obtain

$$\begin{aligned} \begin{aligned} K(t,u,L^p(\varOmega ),{\mathcal {D}}^{1,p}_0(\varOmega ))&\le \inf _{\tau >0} \Big [|1-\tau |\,\Vert u\Vert _{L^p(\varOmega )}+t\,\tau \,\Vert \nabla u\Vert _{L^p(\varOmega )}\Big ]\\&=\min \Big \{\Vert u\Vert _{L^p(\varOmega )},\,t\,\Vert \nabla u\Vert _{L^p(\varOmega )}\Big \} . \end{aligned} \end{aligned}$$

By raising to the power p and integrating for \(t>0\), we get

$$\begin{aligned} \begin{aligned} \Vert u\Vert _{{\mathcal {X}}^{s,p}_0(\varOmega )}^p&\le \int _0^{+\infty } \frac{\min \Big \{\Vert u\Vert ^p_{L^p(\varOmega )},\,t^p\,\Vert \nabla u\Vert ^p_{L^p(\varOmega )}\Big \}}{t^{s\,p}}\,\frac{\mathrm{d}t}{t}\\&=\Vert \nabla u\Vert _{L^p(\varOmega )}^p\,\int _0^\frac{\Vert u\Vert _{L^p(\varOmega )}}{\Vert \nabla u\Vert _{L^p(\varOmega )}} t^{p\,(1-s)}\,\frac{\mathrm{d}t}{t}\\&\quad +\Vert u\Vert _{L^p(\varOmega )}^p\,\int _\frac{\Vert u\Vert _{L^p(\varOmega )}}{\Vert \nabla u\Vert _{L^p(\varOmega )}}^{+\infty } t^{-s\,p}\,\frac{\mathrm{d}t}{t}\\&=\Vert u\Vert _{L^p(\varOmega )}^{p\,(1-s)}\,\Vert \nabla u\Vert _{L^p(\varOmega )}^{s\,p}\,\left[ \frac{1}{p\,(1-s)}+\frac{1}{s\,p}\right] . \end{aligned} \end{aligned}$$

We thus get the desired conclusion (3.7). The estimate (3.8) easily follows from the definition of Poincaré constant. \(\square \)

From (3.6) and (3.8), we get, in particular, the following

Corollary 3.3

(Equivalence of Poincaré constants) Let \(1<p<\infty \) and \(0<s<1\). For every \(\varOmega \subset {\mathbb {R}}^N\) open set we have

$$\begin{aligned} p\,s\,(1-s)\,\varLambda ^s_p(\varOmega )=\Big (\lambda ^1_p(\varOmega )\Big )^s. \end{aligned}$$

In particular, there holds

$$\begin{aligned} {\mathcal {D}}^{1,p}_0(\varOmega ) \hookrightarrow L^p(\varOmega ) \quad \Longleftrightarrow \quad {\mathcal {X}}^{s,p}_0(\varOmega ) \hookrightarrow L^p(\varOmega ). \end{aligned}$$

Remark 3.4

(Extensions by zero in \({\mathcal {X}}^{s,p}_0\)) We observe that by interpolating the “extension by zero” operators

$$\begin{aligned} {\mathcal {T}}_0 : {\mathcal {D}}^{1,p}_0(\varOmega ) \rightarrow {\mathcal {D}}^{1,p}_0({\mathbb {R}}^N) \quad \hbox {and}\quad {\mathcal {T}}_0:L^p(\varOmega ) \rightarrow L^{p}({\mathbb {R}}^N), \end{aligned}$$

which are both continuous, one obtains the same result for the interpolating spaces. In other words, we have

$$\begin{aligned} \Vert u\Vert ^p_{{\mathcal {X}}^{s,p}_0({\mathbb {R}}^N)}\le \Vert u\Vert ^p_{{\mathcal {X}}^{s,p}_0(\varOmega )},\quad \hbox {for every }u\in C^\infty _0(\varOmega ). \end{aligned}$$

This can be also seen directly: it is sufficient to observe that \(C^\infty _0(\varOmega )\subset C^\infty _0({\mathbb {R}}^N)\); thus, we immediately get

$$\begin{aligned} K(t,u,L^p({\mathbb {R}}^N),{\mathcal {D}}^{1,p}_0({\mathbb {R}}^N))\le K(t,u,L^p(\varOmega ),{\mathcal {D}}^{1,p}_0(\varOmega )), \end{aligned}$$

since in the K-functional on the left-hand side the infimum is performed on a larger class. By integrating, we get the conclusion.

However, differently from the case of \({\mathcal {D}}^{1,p}_0(\varOmega )\), \(L^p(\varOmega )\) and \({\mathcal {D}}^{s,p}_0(\varOmega )\), in general for \(u\in C^\infty _0(\varOmega )\) we have

$$\begin{aligned} \Vert u\Vert ^p_{{\mathcal {X}}^{s,p}_0({\mathbb {R}}^N)}< \Vert u\Vert ^p_{{\mathcal {X}}^{s,p}_0(\varOmega )}. \end{aligned}$$

In other words, even if \(u\equiv 0\) outside \(\varOmega \), passing from \(\varOmega \) to \({\mathbb {R}}^N\) has an impact on the interpolation norm.

Actually, if \(\varOmega \) has not smooth boundary, the situation can be much worse than this. We refer to Remark 4.6.

4 Interpolation versus Sobolev–Slobodeckiĭ

4.1 General sets

We want to compare the norms of \({\mathcal {D}}^{s,p}_0(\varOmega )\) and \({\mathcal {X}}^{s,p}_0(\varOmega )\). We start with the simplest estimate, which is valid for every open set.

Proposition 4.1

(Comparison of norms I) Let \(1<p<\infty \) and \(0<s<1\). Let \(\varOmega \subset {\mathbb {R}}^N\) be an open set, then for every \(u\in C^\infty _0(\varOmega )\) we have

$$\begin{aligned} \frac{1}{2^{p\,(1-s)}\,N\,\omega _N}\,\Vert u\Vert ^p_{{\mathcal {D}}^{s,p}_0(\varOmega )}\le \Vert u\Vert ^p_{{\mathcal {X}}^{s,p}_0(\varOmega )}. \end{aligned}$$
(4.1)

In particular, we have the continuous inclusion \({\mathcal {X}}^{s,p}_0(\varOmega )\subset {\mathcal {D}}^{s,p}_0(\varOmega )\).

Proof

To prove (4.1), we take \(h\in {\mathbb {R}}^N{\setminus }\{0\}\) and \(\varepsilon >0\), then there exists \(v_{h,\varepsilon }\in C^\infty _0(\varOmega )\) such that

$$\begin{aligned} \Vert u-v_{h,\varepsilon }\Vert _{L^p(\varOmega )}+\frac{|h|}{2}\,\Vert \nabla v_{h,\varepsilon }\Vert _{L^p(\varOmega )}\le (1+\varepsilon )\,K\left( \frac{|h|}{2},u,L^p(\varOmega ),{\mathcal {D}}^{1,p}_0(\varOmega )\right) . \end{aligned}$$
(4.2)

Thus, for \(h\not =0\) we getFootnote 1

$$\begin{aligned} \begin{aligned} \left( \int _{{\mathbb {R}}^N} \frac{|u(x+h)-u(x)|^{p}}{|h|^{N+s\,p}}\,\mathrm{d}x\right) ^\frac{1}{p}&\le \left( \int _{{\mathbb {R}}^N} \frac{|u(x+h)-v_{h,\varepsilon }(x+h)-u(x)+v_{h,\varepsilon }(x)|^{p}}{|h|^{N+s\,p}}\,\mathrm{d}x\right) ^\frac{1}{p}\\&\quad +\left( \int _{{\mathbb {R}}^N} \frac{|v_{h,\varepsilon }(x+h)-v_{h,\varepsilon }(x)|^{p}}{|h|^{N+s\,p}}\,\mathrm{d}x\right) ^\frac{1}{p}\\&\le 2\,|h|^{-\frac{N}{p}-s}\,\Vert u-v_{h,\varepsilon }\Vert _{L^p(\varOmega )}\\&\quad +|h|^{1-\frac{N}{p}-s}\,\Vert \nabla v_{h,\varepsilon }\Vert _{L^p(\varOmega )}\\&\le 2\,|h|^{-\frac{N}{p}-s}\,\left( \Vert u-v_{h,\varepsilon }\Vert _{L^{p}(\varOmega )}+\frac{|h|}{2}\,\Vert \nabla v_{h,\varepsilon }\Vert _{L^p(\varOmega )}\right) . \end{aligned} \end{aligned}$$

By using (4.2), we then obtain

$$\begin{aligned} \int _{{\mathbb {R}}^N} \frac{|u(x+h)-u(x)|^{p}}{|h|^{N+s\,p}}\,\mathrm{d}x\le 2^p\,(1+\varepsilon )^p\, \left( \frac{K(|h|/2,u,L^p(\varOmega ),{\mathcal {D}}^{1,p}_0(\varOmega ))}{|h|^s}\right) ^p\,\frac{1}{|h|^N}. \end{aligned}$$

We now integrate with respect to \(h\in {\mathbb {R}}^N\) and use spherical coordinates. This yields

$$\begin{aligned}&\iint _{{\mathbb {R}}^N\times {\mathbb {R}}^N} \frac{|u(x+h)-u(x)|^{p}}{|h|^{N+s\,p}}\,\mathrm{d}x\,\mathrm{d}h\\&\quad \le 2^p\,(1+\varepsilon )^p\, \int _{{\mathbb {R}}^N} \left( \frac{K(|h|/2,u,L^p(\varOmega ),{\mathcal {D}}^{1,p}_0(\varOmega ))}{|h|^s}\right) ^{p}\,\frac{\mathrm{d}h}{|h|^N}\\&\quad =2^p\,(1+\varepsilon )^p\,N\,\omega _N\, \int _0^{+\infty } \left( \frac{K(t/2,u,L^p(\varOmega ),{\mathcal {D}}^{1,p}_0(\varOmega ))}{t^s}\right) ^{p}\,\frac{\mathrm{d}t}{t}. \end{aligned}$$

By making the change of variable \(t/2=\tau \) and exploiting the arbitrariness of \(\varepsilon >0\), we eventually reach the desired estimate. \(\square \)

Corollary 4.2

(Interpolation inequality for \({\mathcal {D}}^{s,p}_0\)) Let \(1<p<\infty \) and \(0<s<1\). Let \(\varOmega \subset {\mathbb {R}}^N\) be an open set. For every \(u\in C^\infty _0(\varOmega )\) we have

$$\begin{aligned} s\,(1-s)\,\Vert u\Vert _{{\mathcal {D}}^{s,p}_0(\varOmega )}^p\le \frac{2^{p\,(1-s)}\,N\,\omega _N}{p}\,\Vert u\Vert _{L^p(\varOmega )}^{p\,(1-s)}\,\Vert \nabla u\Vert _{L^p(\varOmega )}^{s\,p}. \end{aligned}$$
(4.3)

Proof

It is sufficient to combine Propositions 4.1 and 3.2. \(\square \)

Remark 4.3

For \(p\searrow 1\), the previous inequality becomes [7, Proposition 4.2]. In this case, the constant in (4.3) is sharp for \(N=1\).

For a general open set \(\varOmega \subset {\mathbb {R}}^N\), the converse of inequality (4.1) does not hold. This means that in general we have

$$\begin{aligned} {\mathcal {X}}^{s,p}_0(\varOmega )\subset {\mathcal {D}}^{s,p}_0(\varOmega ) \quad \hbox {and}\quad {\mathcal {X}}^{s,p}_0(\varOmega )\not = {\mathcal {D}}^{s,p}_0(\varOmega ), \end{aligned}$$

the inclusion being continuous. We use the construction of “Appendix A”, in order to give a counterexample.

Example 4.4

With the notation of “Appendix A”, let us takeFootnote 2

$$\begin{aligned} E={\mathbb {R}}^N{\setminus } \left( \bigcup _{z\in {\mathbb {Z}}^N} (F+z)\right) ,\quad \hbox {with }F=\left[ -\frac{1}{4},\frac{1}{4}\right] ^{N-1}\times \{0\}. \end{aligned}$$

For every \(\varepsilon >0\), we take \(u_n\in C^\infty _0({\widetilde{\varOmega }}_n)\subset C^\infty _0(E)\) such that

$$\begin{aligned}{}[u_n]^p_{W^{s,p}({\mathbb {R}}^N)}<\lambda ^s_{p}({\widetilde{\varOmega }}_n)+\varepsilon \quad \hbox {and}\quad \int _{E} |u_n|^p\,\mathrm{d}x=1. \end{aligned}$$

Here the set \({\widetilde{\varOmega }}_n\) is defined by

$$\begin{aligned} {\widetilde{\varOmega }}_n=\bigcup _{z\in {\mathbb {Z}}^N_n} \Big (\varOmega +z\Big )=\left[ -n-\frac{1}{2},n+\frac{1}{2}\right] ^N{\setminus } \bigcup _{z\in {\mathbb {Z}}^N_n} (F+z). \end{aligned}$$

On the other hand, by Corollary 3.3 we have

$$\begin{aligned} \begin{aligned} \Vert u_n\Vert ^p_{{\mathcal {X}}^{s,p}_0(E)}&\ge \varLambda _p^s(E)\,\int _E |u_n|^p\,\mathrm{d}x\\&=\frac{\Big (\lambda _{p}^1(E)\Big )^s}{p\,s\,(1-s)}\,\int _{E} |u_n|^p\,\mathrm{d}x=\frac{\Big (\lambda _{p}^1(E)\Big )^s}{p\,s\,(1-s)}:=\frac{1}{C}. \end{aligned} \end{aligned}$$

where we also used (A.4), to infer that \(\lambda ^1_p(E)>0\). By Lemma A.1, we have that \(\lambda ^s_{p}(\varOmega _n)\) converges to 0 for \(s\,p<1\), so that

$$\begin{aligned} \liminf _{n\rightarrow \infty }\Vert u_n\Vert _{{\mathcal {X}}^{s,p}_0(E)}^p\ge \frac{1}{C}\quad \hbox {and}\quad \limsup _{n\rightarrow \infty }\, [u_n]^p_{W^{s,p}({\mathbb {R}}^N)}\le \varepsilon . \end{aligned}$$

Since \(\varepsilon >0\) is arbitrary, we obtain

$$\begin{aligned} \lim _{n\rightarrow \infty } \frac{\displaystyle \Vert u_n\Vert ^p_{{\mathcal {D}}^{s,p}_0(E)}}{\Vert u_n\Vert ^p_{{\mathcal {X}}^{s,p}_0(E)}}=0,\quad \hbox {for } 1<p<\infty \hbox { and } s< \frac{1}{p}. \end{aligned}$$

4.2 Convex sets

We now prove the converse of (4.1), under suitable assumptions on \(\varOmega \). We start with the case of a convex set. The case \(\varOmega ={\mathbb {R}}^N\) is simpler and instructive; thus, we give a separate statement. The proof can be found, for example, in [28, Lemma 35.2]. We reproduce it, for the reader’s convenience. We also single out an explicit determination of the constant.

Proposition 4.5

(Comparison of norms II: \({\mathbb {R}}^N\)) Let \(1<p<\infty \) and \(0<s<1\). For every \(u\in C^\infty _0({\mathbb {R}}^N)\) we have

$$\begin{aligned} \Vert u\Vert ^p_{{\mathcal {X}}^{s,p}_0({\mathbb {R}}^N)}\le \Big (N\,(N+1)\Big )^p\,\frac{2^p}{N\,\omega _N}\,\Vert u\Vert ^p_{{\mathcal {D}}_0^{s,p}({\mathbb {R}}^N)}. \end{aligned}$$

In particular, we have that \({\mathcal {D}}^{s,p}_0({\mathbb {R}}^N)={\mathcal {X}}^{s,p}_0({\mathbb {R}}^N)\).

Proof

Let \(u\in C^\infty _0({\mathbb {R}}^N)\), we set

$$\begin{aligned} U(h)=\left( \int _{{\mathbb {R}}^N} |u(x+h)-u(x)|^p\,\mathrm{d}x\right) ^\frac{1}{p},\quad h\in {\mathbb {R}}^N, \end{aligned}$$

and observe that by construction

$$\begin{aligned} \int _{{\mathbb {R}}^N} \frac{U(h)^p}{|h|^{N+s\,p}}\,\mathrm{d}h=[u]_{W^{s,p}({\mathbb {R}}^N)}^p. \end{aligned}$$

We also define

$$\begin{aligned} \overline{U}(\varrho )=\frac{1}{N\,\omega _N\,\varrho ^{N-1}}\,\int _{\{h\in {\mathbb {R}}^N\, :\, |h|=\varrho \}} U\,d{\mathcal {H}}^{N-1},\quad \varrho >0; \end{aligned}$$

thus, by Jensen’s inequality we have

$$\begin{aligned} \begin{aligned} \int _0^{+\infty } \left( \frac{\overline{U}}{\varrho ^s}\right) ^p\,\frac{d\varrho }{\varrho }&\le \frac{1}{N\,\omega _N}\,\int _{0}^{+\infty } \left( \int _{\{h\in {\mathbb {R}}^N\, :\, |h|=\varrho \}} U^p\,d{\mathcal {H}}^{N-1}\right) \,\frac{d\varrho }{\varrho ^{N+s\,p}}\\&= \frac{1}{N\,\omega _N}\int _{{\mathbb {R}}^N} \frac{U(h)^p}{|h|^{N+s\,p}}\,\mathrm{d}h=\frac{1}{N\,\omega _N}\,[u]_{W^{s,p}({\mathbb {R}}^N)}^p. \end{aligned} \end{aligned}$$
(4.4)

We now take the compactly supported Lipschitz function

$$\begin{aligned} \psi (x)=\frac{N+1}{\omega _N}\,(1-|x|)_+, \end{aligned}$$

where \((\,\cdot \,)_+\) stands for the positive part. Observe that \(\psi \) has unit \(L^1\) norm, by construction. We then define

$$\begin{aligned} \psi _t(x)=\frac{1}{t^N} \,\psi \left( \frac{x}{t}\right) ,\quad \hbox {for }t>0. \end{aligned}$$

By observing that \(\psi _t*u\in C^\infty _0({\mathbb {R}}^N)\), from the definition of the K-functional, we get

$$\begin{aligned} K(t,u,L^p({\mathbb {R}}^N),{\mathcal {D}}^{1,p}_0({\mathbb {R}}^N))\le \Vert u-\psi _t*u\Vert _{L^p({\mathbb {R}}^N)}+t\,\Vert \nabla \psi _t*u\Vert _{L^p({\mathbb {R}}^N)}. \end{aligned}$$

We estimate the two norms in the right-hand side separately: for the first one, by Minkowski inequality and Fubini Theorem we get

$$\begin{aligned} \begin{aligned} \Vert u-\psi _t*u\Vert _{L^p({\mathbb {R}}^N)}&=\left\| \int _{{\mathbb {R}}^N} [u(\cdot )-u(\cdot -y)]\,\psi _t(y)\,\mathrm{d}y\right\| _{L^p({\mathbb {R}}^N)}\\&\le \int _{{\mathbb {R}}^N}\left( \int _{{\mathbb {R}}^N} |u(x)-u(x-y)|^p\,\mathrm{d}x\right) ^\frac{1}{p}\,\psi _t(y)\,\mathrm{d}y\\&=\int _{{\mathbb {R}}^N} U(-y)\,\psi _{t}(y)\,\mathrm{d}y\le \frac{N+1}{\omega _N\,t^N}\,\int _{B_t(0)} U(-y)\,\mathrm{d}y\\&=\frac{N\,(N+1)}{t^N}\,\int _0^t \overline{U}\,\varrho ^{N-1}\,d\varrho \le \frac{N\,(N+1)}{t}\,\int _0^t \overline{U}\,d\varrho . \end{aligned} \end{aligned}$$

For the norm of the gradient, we first observe that

$$\begin{aligned} \int _{{\mathbb {R}}^N} \nabla \psi _t(y)\,\mathrm{d}y=0; \end{aligned}$$

thus, we can write

$$\begin{aligned} \nabla \psi _t*u=(\nabla \psi _t)*u=\int _{{\mathbb {R}}^N} \nabla \psi _t(y)\,[u(x-y)-u(x)]\,\mathrm{d}y. \end{aligned}$$

Consequently, by Minkowski inequality we get

$$\begin{aligned} \begin{aligned} \Vert \nabla \psi _t*u\Vert _{L^p({\mathbb {R}}^N)}&=\left\| \int _{{\mathbb {R}}^N} \nabla \psi _t(y)\,[u(\cdot -y)-u(\cdot )]\,\mathrm{d}y\right\| _{L^p({\mathbb {R}}^N)}\\&\le \int _{{\mathbb {R}}^N}\left( \int _{{\mathbb {R}}^N} |u(x-y)-u(x)|^p\,\mathrm{d}x\right) ^\frac{1}{p}\,|\nabla \psi _t(y)|\,\mathrm{d}y\\&\le \frac{N+1}{\omega _N\,t^{N+1}} \,\int _{B_t(0)} U(-y)\,\mathrm{d}y\le \frac{N\,(N+1)}{t^2}\,\int _0^t \overline{U}\,d\varrho . \end{aligned} \end{aligned}$$

In conclusion, we obtained for every \(t>0\)

$$\begin{aligned} K(t,u,L^p({\mathbb {R}}^N),{\mathcal {D}}^{1,p}_0({\mathbb {R}}^N))\le \frac{2\,N\,(N+1)}{t}\,\int _0^t \overline{U}\,d\varrho . \end{aligned}$$
(4.5)

If we integrate on (0, T), the previous estimate gives

$$\begin{aligned} \int _0^T \left( \frac{K(t,u,L^p({\mathbb {R}}^N),{\mathcal {D}}^{1,p}_0({\mathbb {R}}^N))}{t^s}\right) ^p\,\frac{\mathrm{d}t}{t}\le \Big (2\,N\,(N+1)\Big )^p\,\int _0^T \left( \int _0^t \overline{U}\,d\varrho \right) ^p\,t^{-p-s\,p}\,\frac{\mathrm{d}t}{t}. \end{aligned}$$

If we now use Lemma B.1 with \(\alpha =p+s\,p\) for the function

$$\begin{aligned} t\mapsto \int _0^t \overline{U}\,d\varrho , \end{aligned}$$

we get

$$\begin{aligned} \begin{aligned} \int _0^T \left( \frac{K(t,u,L^p(\varOmega ),{\mathcal {D}}^{1,p}_0(\varOmega ))}{t^s}\right) ^p\,\frac{\mathrm{d}t}{t}&\le \left( \frac{2\,N\,(N+1)}{s+1}\right) ^p\,\int _0^T \left( \frac{\overline{U}}{t^s}\right) ^p\frac{\mathrm{d}t}{t}\\&\le \left( \frac{2\,N\,(N+1)}{s+1}\right) ^p\,\frac{1}{N\,\omega _N}\,[u]^p_{W^{s,p}({\mathbb {R}}^N)}, \end{aligned} \end{aligned}$$

where we used (4.4) in the second inequality. By letting T going to \(+\infty \), we get the desired estimate. \(\square \)

Remark 4.6

(Extensions by zero in \({\mathcal {X}}^{s,p}_0(\varOmega )\)...reprise) We take the set \(E\subset {\mathbb {R}}^N\) and the sequence \(\{u_n\}_{n\in {\mathbb {N}}}\subset C^\infty _0(E)\) as in Example 4.4. We have seen that

$$\begin{aligned} \lim _{n\rightarrow \infty } \frac{\Vert u_n\Vert _{{\mathcal {D}}^{s,p}_0(E)}}{\Vert u_n\Vert _{{\mathcal {X}}^{s,p}_0(E)}}=0,\quad \hbox {for } s<\frac{1}{p}. \end{aligned}$$

By observing that

$$\begin{aligned} \Vert u_n\Vert _{{\mathcal {D}}^{s,p}_0(E)}=\Vert u_n\Vert _{{\mathcal {D}}^{s,p}_0({\mathbb {R}}^N)}, \end{aligned}$$

and using Proposition 4.5, we obtain

$$\begin{aligned} \lim _{n\rightarrow \infty } \frac{\Vert u_n\Vert _{{\mathcal {X}}^{s,p}_0({\mathbb {R}}^N)}}{\Vert u_n\Vert _{{\mathcal {X}}^{s,p}_0(E)}}=0, \end{aligned}$$

as well, still for \(s\,p<1\). This shows that the “extension by zero” operator

$$\begin{aligned} {\mathcal {T}}_0:{\mathcal {X}}^{s,p}_0(E)\rightarrow {\mathcal {X}}^{s,p}_0({\mathbb {R}}^N), \end{aligned}$$

is not an isometry and, even worse, the two norms

$$\begin{aligned} \Vert \,\cdot \, \Vert _{{\mathcal {X}}^{s,p}_0(E)}\quad \hbox {and}\quad \Vert \,\cdot \,\Vert _{{\mathcal {X}}^{s,p}_0({\mathbb {R}}^N)}, \end{aligned}$$

are not equivalent on \(C^\infty _0(E)\). This is in contrast with the case of \(L^p(E)\), \(\mathcal {D}^{1,p}_0(E)\) and \(\mathcal {D}^{s,p}_0(E)\).

We denote by

$$\begin{aligned} R_\varOmega =\sup _{x\in \varOmega } {\mathrm {dist}}(x,\partial \varOmega ), \end{aligned}$$

the inradius of an open set \(\varOmega \subset {\mathbb {R}}^N\). This is the radius of the largest open ball inscribed in \(\varOmega \). We introduce the eccentricity of an open bounded set \(\varOmega \subset {\mathbb {R}}^N\), defined by

$$\begin{aligned} {\mathcal {E}}(\varOmega )=\frac{\mathrm {diam\,}(\varOmega )}{2\,R_\varOmega }. \end{aligned}$$

Observe that this is a scaling invariant quantity. By generalizing the construction used in [9, Lemma A.6] for a ball, we have the following.

Theorem 4.7

(Comparison of norms II: bounded convex sets) Let \(1<p<\infty \) and \(0<s<1\). If \(\varOmega \subset {\mathbb {R}}^N\) is an open bounded convex set, then for every \(u\in C^\infty _0(\varOmega )\) we have

$$\begin{aligned} \Vert u\Vert ^p_{{\mathcal {X}}^{s,p}_0(\varOmega )}\le C\,\Vert u\Vert ^p_{{\mathcal {D}}_0^{s,p}(\varOmega )}, \end{aligned}$$
(4.6)

for a constant \(C=C(N,p,{\mathcal {E}}(\varOmega ))>0\), which blows up as \({\mathcal {E}}(\varOmega )\rightarrow +\infty \). In particular, we have \({\mathcal {X}}^{s,p}_0(\varOmega )={\mathcal {D}}^{s,p}_0(\varOmega )\).

Proof

The proof runs similarly to that of Proposition 4.5 for \({\mathbb {R}}^N\), but now we have to pay attention to boundary issues. Indeed, the function \(\psi _t*u\) is not supported in \(\varOmega \), unless t is sufficiently small, depending on u itself. In order to avoid this, we need to perform a controlled scaling of the function. By keeping the same notation as in the proof of Proposition 4.5, we need the following modification: we take a point \(x_0\in \varOmega \) such that

$$\begin{aligned} {\mathrm {dist}}(x_0,\partial \varOmega )=R_\varOmega . \end{aligned}$$

Without the loss of generality, we can assume that \(x_0=0\). Then we define the rescaled function

$$\begin{aligned} u_t=u\left( \frac{R_\varOmega }{R_\varOmega -t}\,x\right) ,\quad 0<t<\frac{R_\varOmega }{2}. \end{aligned}$$

We observe that

$$\begin{aligned} {\mathrm {support}}(u_t)=\frac{R_\varOmega -t}{R_\varOmega }\,\varOmega , \end{aligned}$$

and by Lemma C.1, we have

$$\begin{aligned} {\mathrm {dist}}\left( \frac{R_\varOmega -t}{R_\varOmega }\,\varOmega ,\partial \varOmega \right) \ge \left( 1-\frac{R_\varOmega -t}{R_\varOmega }\right) \,R_\varOmega =t. \end{aligned}$$

This implies that

$$\begin{aligned} \psi _t*u_t\in C^\infty _0(\varOmega ),\quad \hbox {for every } 0<t<\frac{R_\varOmega }{2}. \end{aligned}$$

We can now estimate the K-functional by using the choice \(v=\psi _t*u_t\), that is

$$\begin{aligned} \begin{aligned} K(t,u,L^p(\varOmega ),{\mathcal {D}}^{1,p}_0(\varOmega ))&\le \Vert u-\psi _t*u_t\Vert _{L^p(\varOmega )}\\&\quad +t\,\Vert \nabla \psi _t*u_t\Vert _{L^p(\varOmega )},\quad \hbox { for every } 0<t<\frac{R_\varOmega }{2}. \end{aligned} \end{aligned}$$

Let us set

$$\begin{aligned} \varOmega _t=\{x\in {\mathbb {R}}^N\, :\, {\mathrm {dist}}(x,\varOmega )<2\,t\}, \end{aligned}$$

then we have that for every \(x\in \varOmega \),

$$\begin{aligned} y\mapsto \psi _t(x-y)\quad \hbox { has support contained in } \varOmega _t. \end{aligned}$$

By using this and Jensen’s inequality, we obtain

$$\begin{aligned} \Vert u-\psi _t*u_t\Vert _{L^p(\varOmega )}^p\le \int _{\varOmega }\int _{\varOmega _t} \left| u(x)-u\left( \frac{R}{R-t}\,y\right) \right| ^p\,\frac{1}{t^N}\,\psi \left( \frac{x-y}{t}\right) \,\mathrm{d}y\,\mathrm{d}x. \end{aligned}$$

Thus, by using a change of variable and Fubini theorem, we get

$$\begin{aligned} \begin{aligned} \int _0^{R_\varOmega /2}&\left( \frac{\Vert u-\psi _t*u_t\Vert _{L^p(\varOmega )}}{t^s}\right) ^{p}\,\frac{\mathrm{d}t}{t}\\&\le \int _0^{R_\varOmega /2} \int _{\varOmega }\int _{\varOmega _t} t^{-s\,p}\,\left| u(x)-u\left( \frac{R_\varOmega }{R_\varOmega -t}\,y\right) \right| ^{p}\,\frac{1}{t^N}\,\psi \left( \frac{x-y}{t}\right) \,\mathrm{d}y\,\mathrm{d}x\,\frac{\mathrm{d}t}{t}\\&=\left( \frac{R_\varOmega -t}{R_\varOmega }\right) ^N\,\int _0^{R_\varOmega /2} \int _{\varOmega }\int _{\frac{R_\varOmega }{R_\varOmega -t}\,\varOmega _t} t^{-s\,p}\,\left| u(x)-u(z)\right| ^{p}\,\frac{1}{t^N}\,\psi \left( \frac{x}{t}-\frac{R_\varOmega -t}{R_\varOmega \,t}z\right) \,\mathrm{d}z\,\mathrm{d}x\,\frac{\mathrm{d}t}{t}\\&\le \int _{\varOmega }\int _{{\widetilde{\varOmega }}}\left| u(x)-u(z)\right| ^{p}\left( \int _0^{R_\varOmega /2} \,t^{-s\,p-N}\,\psi \left( \frac{x-z}{t}+\frac{z}{R_\varOmega }\right) \,\frac{\mathrm{d}t}{t}\right) \,\mathrm{d}z\,\mathrm{d}x, \end{aligned} \end{aligned}$$

where we used that

$$\begin{aligned} \frac{R_\varOmega }{R_\varOmega -t}\,\varOmega _t\subset {\widetilde{\varOmega }}:=2\,\varOmega _{R_\varOmega /2},\quad \hbox {for } 0<t<\frac{R_\varOmega }{2}. \end{aligned}$$

We now observe that

$$\begin{aligned} \psi \left( \frac{x-z}{t}+\frac{z}{R_\varOmega }\right) \not =0\quad \Longleftrightarrow \quad \left| \frac{x-z}{t}+\frac{z}{R_\varOmega }\right| <1; \end{aligned}$$

thus, in particular,

$$\begin{aligned} \hbox { if }\quad \left| \frac{x-z}{t}\right| \ge 1+\left| \frac{z}{R_\varOmega }\right| \quad \hbox {then} \quad \psi \left( \frac{x-z}{t}+\frac{z}{R_\varOmega }\right) =0, \end{aligned}$$

i.e., for every \(x\in \varOmega \) and \(z\in {\widetilde{\varOmega }}\),

$$\begin{aligned} \hbox {if} \quad 0<t\le \frac{|x-z|}{1+\dfrac{|z|}{R_\varOmega }}\quad \hbox {then} \quad \psi \left( \frac{x-z}{t}+\frac{z}{R_\varOmega }\right) =0. \end{aligned}$$

This implies that for \(x\in \varOmega \) and \(z\in {\widetilde{\varOmega }}\) we get

$$\begin{aligned} \begin{aligned} \int _0^{R_\varOmega /2} \,t^{-s\,p-N}\,\psi \left( \frac{x-z}{t}+\frac{z}{R_\varOmega }\right) \,\frac{\mathrm{d}t}{t}&\le \int _0^{+\infty } \,t^{-s\,p-N}\,\psi \left( \frac{x-z}{t}+\frac{z}{R_\varOmega }\right) \,\frac{\mathrm{d}t}{t}\\&=\int _{\frac{|x-z|}{1+\frac{|z|}{R_\varOmega }}}^{+\infty } \,t^{-s\,p-N}\,\psi \left( \frac{x-z}{t}+\frac{z}{R_\varOmega }\right) \,\frac{\mathrm{d}t}{t}\\&\le \frac{N+1}{\omega _N\,(N+s\,p)}\,\left( 1+\frac{|z|}{R_\varOmega }\right) ^{N+s\,p}\,|x-z|^{-N-s\,p}\\&\le \frac{N+1}{\omega _N\,(N+s\,p)}\,\left( 1+\frac{{\mathrm {diam}}({\widetilde{\varOmega }})}{R_\varOmega }\right) ^{N+s\,p}\,|x-z|^{-N-s\,p}. \end{aligned} \end{aligned}$$

Thus, we obtain

$$\begin{aligned} \begin{aligned} \int _0^{R_\varOmega /2}&\left( \frac{\Vert u-\psi _t*u_t\Vert _{L^{p}(\varOmega )}}{t^s}\right) ^{p}\,\frac{\mathrm{d}t}{t}\\&\le \frac{N+1}{\omega _N\,(N+s\,p)}\,\left( 1+\frac{{\mathrm {diam}}({\widetilde{\varOmega }})}{R_\varOmega }\right) ^{N+s\,p}\,\int _\varOmega \int _{{\widetilde{\varOmega }}} \frac{|u(x)-u(z)|^{p}}{|x-z|^{N+s\,p}}\,\mathrm{d}x\,\mathrm{d}z\\&\le \frac{N+1}{\omega _N\,(N+s\,p)}\,\left( 1+\frac{{\mathrm {diam}}({\widetilde{\varOmega }})}{R_\varOmega }\right) ^{N+s\,p}\, \Vert u\Vert ^{p}_{{\mathcal {D}}^{s,p}_0(\varOmega )}. \end{aligned} \end{aligned}$$
(4.7)

Observe that by construction

$$\begin{aligned} {\mathrm {diam}}({\widetilde{\varOmega }})=2\,{\mathrm {diam}}(\varOmega _{R_\varOmega /2})\le 2\,\Big ({\mathrm {diam}}(\varOmega )+2\,R_\varOmega \Big ). \end{aligned}$$

We now need to show that

$$\begin{aligned} \int _0^{R_\varOmega /2} t^{p}\,\left( \frac{\Vert \nabla \psi _t*u_t\Vert _{L^p(\varOmega )}}{t^s}\right) ^p\,\frac{\mathrm{d}t}{t}\le C\, \Vert u\Vert ^{p}_{{\mathcal {D}}^{s,p}_0(\varOmega )}. \end{aligned}$$
(4.8)

We first observe that

$$\begin{aligned} \begin{aligned} \nabla \psi _t*u_t(x)&=\int _{{\mathbb {R}}^N} u\left( \frac{R_\varOmega }{R_\varOmega -t}\,y\right) \,\frac{1}{t^{N+1}}\,\nabla \psi \left( \frac{x-y}{t}\right) \,\mathrm{d}y, \end{aligned} \end{aligned}$$

and by the divergence theorem

$$\begin{aligned} \int _{{\mathbb {R}}^N}\frac{1}{t^{N+1}}\,\nabla \psi \left( \frac{x-y}{t}\right) \,\mathrm{d}y=0. \end{aligned}$$

Thus, we obtain as well

$$\begin{aligned} -\nabla \psi _t*u_t(x)=\int _{{\mathbb {R}}^N} \left[ u\left( \frac{R_\varOmega }{R_\varOmega -t}\,x\right) -u\left( \frac{R_\varOmega }{R_\varOmega -t}\,y\right) \right] \,\frac{1}{t^{N+1}}\,\nabla \psi \left( \frac{x-y}{t}\right) \,\mathrm{d}y, \end{aligned}$$

and by Hölder’s inequality

$$\begin{aligned} \begin{aligned} \Vert \nabla \psi _t*u_t\Vert _{L^p(\varOmega )}^{p}&\le \int _{{\mathbb {R}}^N} \left( \int _{{\mathbb {R}}^N} \left| u\left( \frac{R_\varOmega }{R_\varOmega -t}\,x\right) -u\left( \frac{R_\varOmega }{R_\varOmega -t}\,y\right) \right| ^{p}\,\frac{1}{t^{N+1}}\right. \\&\quad \left. \left| \nabla \psi \left( \frac{x-y}{t}\right) \right| \,\mathrm{d}y\right) \times \left( \int _{{\mathbb {R}}^N}\frac{1}{t^{N+1}}\,\left| \nabla \psi \left( \frac{x-y}{t}\right) \right| \,\mathrm{d}y\right) ^{p-1}\,\mathrm{d}x\\&= \frac{\Vert \nabla \psi \Vert ^{p-1}_{L^1({\mathbb {R}}^N)}}{t^{p-1}}\, \int _{{\mathbb {R}}^N} \int _{{\mathbb {R}}^N} \left| u\left( \frac{R_\varOmega }{R_\varOmega -t}\,x\right) \right. \\&\quad \left. -\,u\left( \frac{R_\varOmega }{R_\varOmega -t}\,y\right) \right| ^{p}\,\,\frac{1}{t^{N+1}}\,\left| \nabla \psi \left( \frac{x-y}{t}\right) \right| \,\mathrm{d}y\,\mathrm{d}x\\&\le \frac{\Vert \nabla \psi \Vert ^{p-1}_{L^1({\mathbb {R}}^N)}}{t^{p-1}}\,\int _{{\mathbb {R}}^N} \int _{{\mathbb {R}}^N} \left| u\left( z\right) -u\left( w\right) \right| ^{p}\,\,\frac{1}{t^{N+1}}\\&\quad \left| \nabla \psi \left( \frac{R_\varOmega -t}{R_\varOmega \,t}\,(z-w)\right) \right| \,\mathrm{d}z\,\mathrm{d}w, \end{aligned} \end{aligned}$$

for every \(0<t<R_\varOmega /2\). This yields

$$\begin{aligned} \begin{aligned} \int _0^{R_\varOmega /2}&t^{p}\,\left( \frac{\Vert \nabla \psi _t*u_t\Vert _{L^p(\varOmega )}}{t^s}\right) ^{p}\,\frac{\mathrm{d}t}{t}\\&\le (N+1)^{p-1}\,\int _0^{R_\varOmega /2} t^{-s\,p}\,\int _{{\mathbb {R}}^N} \int _{{\mathbb {R}}^N} \left| u(z)-u(w)\right| ^{p}\,\,\frac{1}{t^{N}}\\&\quad \left| \nabla \psi \left( \frac{R_\varOmega -t}{R_\varOmega \,t}\,(z-w)\right) \right| \,\mathrm{d}z\,\mathrm{d}w\,\frac{\mathrm{d}t}{t}\\&=(N+1)^{p-1}\,\int _{{\mathbb {R}}^N} \int _{{\mathbb {R}}^N} |u(z)-u(w)|^{p}\\&\quad \left( \int _0^{R_\varOmega /2} t^{-s\,p}\,\frac{1}{t^{N}}\,\left| \nabla \psi \left( \frac{R_\varOmega -t}{R_\varOmega \,t}\,(z-w)\right) \right| \,\frac{\mathrm{d}t}{t}\right) \,\mathrm{d}z\,\mathrm{d}w. \end{aligned} \end{aligned}$$
(4.9)

As above, we now observe that

$$\begin{aligned} \left| \nabla \psi \left( \frac{R_\varOmega -t}{R_\varOmega \,t}\,(z-w)\right) \right| \not =0\quad \Longleftrightarrow \quad \frac{R_\varOmega -t}{R_\varOmega }\,\frac{|z-w|}{t}<1; \end{aligned}$$

thus, in particular, for \(0<t<R_\varOmega /2\) we have

$$\begin{aligned} \frac{1}{2}\,\frac{|z-w|}{t}>1\quad \Longrightarrow \quad \nabla \psi \left( \frac{R_\varOmega -t}{R_\varOmega \,t}\,(z-w)\right) =0. \end{aligned}$$

This implies that for \(z,w\in {\mathbb {R}}^N\) we have

$$\begin{aligned} \begin{aligned}&\int _0^{R_\varOmega /2} t^{-s\,p}\,\frac{1}{t^{N}}\,\left| \nabla \psi \left( \frac{R_\varOmega -t}{R_\varOmega \,t}\,(z-w)\right) \right| \,\frac{\mathrm{d}t}{t}\\&\quad \le \int _{\frac{|z-w|}{2}}^{+\infty } t^{-s\,p}\,\frac{1}{t^{N}}\,\left| \nabla \psi \left( \frac{R_\varOmega -t}{R_\varOmega \,t}\,(z-w)\right) \right| \,\frac{\mathrm{d}t}{t}\\&\quad \le \frac{(N+1)\,2^{N+s\,p}}{\omega _N\,(N+s\,p)}\,|z-w|^{-N-s\,p}. \end{aligned} \end{aligned}$$

By inserting this estimate in (4.9), we now get (4.8). This and (4.7) then give

$$\begin{aligned} \int _0^{\frac{R_\varOmega }{2}} \left( \frac{K(t,u,L^p(\varOmega ),{\mathcal {D}}^{1,p}_0(\varOmega ))}{t^s}\right) ^p\,\frac{\mathrm{d}t}{t}\le C\,\Vert u\Vert _{{\mathcal {D}}^{s,p}_0(\varOmega )}^p, \end{aligned}$$
(4.10)

for a constant \(C=C(N,p,{\mathcal {E}}(\varOmega ))\).

We are left with estimating the integral of the K-functional on \((R_\varOmega /2,+\infty )\): for this, we can use the trivial decomposition

$$\begin{aligned} u=(u-0)+0, \end{aligned}$$

which gives

$$\begin{aligned} \begin{aligned} \int _{\frac{R_\varOmega }{2}}^{+\infty } \left( \frac{K(t,u,L^p(\varOmega ),{\mathcal {D}}^{1,p}_0(\varOmega ))}{t^s}\right) ^p\,\frac{\mathrm{d}t}{t}&\le \int _{\frac{R_\varOmega }{2}}^{+\infty } \frac{\Vert u\Vert _{L^p(\varOmega )}^p}{t^{s\,p}}\,\frac{\mathrm{d}t}{t}\\&=\frac{\Vert u\Vert _{L^p(\varOmega )}^p}{s\,p}\, \left( \frac{R_\varOmega }{2}\right) ^{-s\,p}\\&\le \frac{2^{s\,p}}{s\,p}\,\Vert u\Vert _{{\mathcal {D}}^{s,p}_0(\varOmega )}^p\, \left( \frac{1}{\lambda ^s_p(\varOmega )\,R_\varOmega ^{s\,p}}\right) , \end{aligned} \end{aligned}$$

where we used the Poincaré inequality for \({\mathcal {D}}^{s,p}_0(\varOmega )\). By recalling that for a convex set with finite inradius, we have (see [8, Corollary 5.1])

$$\begin{aligned} \lambda ^s_p(\varOmega )\,R_\varOmega ^{s\,p}\ge \frac{{\mathcal {C}}}{s\,(1-s)}, \end{aligned}$$

for a constant \({\mathcal {C}}={\mathcal {C}}(N,p)>0\), we finally obtain

$$\begin{aligned} \int _{\frac{R_\varOmega }{2}}^{+\infty } \left( \frac{K(t,u,L^p(\varOmega ),{\mathcal {D}}^{1,p}_0(\varOmega ))}{t^s}\right) ^p\,\frac{\mathrm{d}t}{t}\le \frac{2^{s\,p}}{p}\,\Vert u\Vert _{{\mathcal {D}}^{s,p}_0(\varOmega )}^p\, \left( \frac{1-s}{{\mathcal {C}}}\right) . \end{aligned}$$

By using this in conjunction with (4.10), we finally conclude the proof. \(\square \)

For general unbounded convex sets, the previous proof does not work anymore. However, for convex cones the result still holds. We say that a convex set \(\varOmega \subset {\mathbb {R}}^N\) is a convex cone centered at\(x_0\in {\mathbb {R}}^N\) if for every \(x\in \varOmega \) and \(\tau >0\), we have

$$\begin{aligned} x_0+\tau \,(x-x_0)\in \varOmega . \end{aligned}$$

Then we have the following

Corollary 4.8

(Comparison of norms II: convex cones) Let \(1<p<\infty \) and \(0<s<1\). If \(\varOmega \subset {\mathbb {R}}^N\) is an open convex cone centered at \(x_0\in {\mathbb {R}}^N\), then for every \(u\in C^\infty _0(\varOmega )\) we have

$$\begin{aligned} \Vert u\Vert ^p_{{\mathcal {X}}^{s,p}_0(\varOmega )}\le C\,\Vert u\Vert ^p_{{\mathcal {D}}_0^{s,p}(\varOmega )}, \end{aligned}$$

for a constant \(C=C(N,p,{\mathcal {E}}(\varOmega \cap B_1(x_0)))>0\). In particular, we have \({\mathcal {X}}^{s,p}_0(\varOmega )={\mathcal {D}}^{s,p}_0(\varOmega )\).

Proof

We assume for simplicity that \(x_0=0\) and take \(u\in C^\infty _0(\varOmega )\). Since u has compact support, we have that \(u\in C^\infty _0(\varOmega \cap B_R(0))\), for R large enough. From Theorem 4.7, we know that

$$\begin{aligned} \Vert u\Vert ^p_{{\mathcal {X}}^{s,p}_0(\varOmega \cap B_R(0))}\le C\,\Vert u\Vert _{{\mathcal {D}}^{s,p}_0(\varOmega \cap B_R(0))}= C\, \Vert u\Vert _{{\mathcal {D}}^{s,p}_0(\varOmega )}. \end{aligned}$$

We recall that the constant C depends on the eccentricity of \(\varOmega \cap B_R(0)\). However, since \(\varOmega \) is a cone, we easily get

$$\begin{aligned} {\mathcal {E}}(\varOmega \cap B_R(0))={\mathcal {E}}(\varOmega \cap B_1(0)),\quad \hbox {for every } R>0, \end{aligned}$$

i.e., the constant C is independent of R. Finally, by observing that

$$\begin{aligned} \Vert u\Vert ^p_{{\mathcal {X}}^{s,p}_0(\varOmega )}\le \Vert u\Vert ^p_{{\mathcal {X}}^{s,p}_0(\varOmega \cap B_R(0))}, \end{aligned}$$

we get the desired conclusion. \(\square \)

Remark 4.9

(Rotationally symmetric cones) Observe that if \(\varOmega \) is the rotationally symmetric convex cone

$$\begin{aligned} \varOmega =\{x\in {\mathbb {R}}^N\, :\, \langle x-x_0,\omega \rangle >\beta \,|x-x_0|\},\quad \hbox {for some }0\le \beta <1,\, x_0\in {\mathbb {R}}^N\hbox { and }\omega \in \mathbb {S}^{N-1}, \end{aligned}$$

we have

$$\begin{aligned} {\mathcal {E}}(\varOmega \cap B_1(0))=\frac{1}{2}\,\max \left\{ 2\,\sqrt{1-\beta ^2},\, 1\right\} \,\left( 1+\frac{1}{\sqrt{1-\beta ^2}}\right) , \end{aligned}$$

by elementary geometric considerations.

In particular, when \(\varOmega \) is a half-space (i.e., when \(\beta =0\)), then we have \({\mathcal {E}}(\varOmega \cap B_1(0))=2\).

4.3 Lipschitz sets and beyond

In this section, we show that the norms of \({\mathcal {X}}^{s,p}_0\) and \({\mathcal {D}}^{s,p}_0\) are equivalent on open bounded Lipschitz sets. We also make some comments on more general sets, see Remark 4.11.

By generalizing the idea of [22, Theorem 11.6] (see also [6, Theorem 2.1]) for \(p=2\) and smooth sets, we can rely on the powerful extension theorem for Sobolev functions proved by Stein and obtain the following

Theorem 4.10

(Comparison of norms II: Lipschitz sets) Let \(1<p<\infty \) and \(0<s<1\). Let \(\varOmega \subset {\mathbb {R}}^N\) be an open bounded set, with Lipschitz boundary. Then for every \(u\in C^\infty _0(\varOmega )\) we have

$$\begin{aligned} \Vert u\Vert ^p_{{\mathcal {X}}^{s,p}_0(\varOmega )}\le C_1\,\Vert u\Vert ^p_{{\mathcal {D}}_0^{s,p}(\varOmega )}, \end{aligned}$$

for a constant \(C_1>0\) depending on \(N,p, {\mathrm {diam}}(\varOmega )\) and the Lipschitz constant of \(\partial \varOmega \). In particular, we have \({\mathcal {X}}^{s,p}_0(\varOmega )={\mathcal {D}}^{s,p}_0(\varOmega )\) in this case as well.

Proof

We take an open ball \(B\subset {\mathbb {R}}^N\) with radius \({\mathrm {diam}}(\varOmega )\) and such that \(\varOmega \Subset B\). We then take a linear and continuous extension operator

$$\begin{aligned} {\mathcal {T}}: W^{1,p}(B{\setminus }\overline{\varOmega })\rightarrow W^{1,p}(B), \end{aligned}$$

such that

$$\begin{aligned} \left\{ \begin{array}{rcl} \Vert {\mathcal {T}}(u)\Vert _{L^p(B)}&{}\le &{} \mathfrak {e}_\varOmega \,\Vert u\Vert _{L^p(B{\setminus }\overline{\varOmega })},\\ &{}&{}\\ \Vert \nabla {\mathcal {T}}(u)\Vert _{L^p(B)}&{}\le &{} \mathfrak {e}_\varOmega \,\Vert u\Vert _{W^{1,p}(B{\setminus }\overline{\varOmega })}, \end{array} \right. \end{aligned}$$
(4.11)

where \(\mathfrak {e}_\varOmega >0\) depends on \(N,p,{\mathrm {diam}}(\varOmega )\) and the Lipschitz constant of \(\partial \varOmega \). We observe that such an operator exists, thanks to the fact that \(\varOmega \) has a Lipschitz boundary, see [27, Theorem 5, p. 181]. We also observe that the first estimate in (4.11) is not explicitly stated by Stein, but it can be extrapolated by having a closer look at the proof, see [27, p. 192].

For every \(v\in C^\infty _0(B)\), we define the operator

$$\begin{aligned} {\mathcal {R}}(v)=v-{\mathcal {T}}(v\cdot 1_{B{\setminus }\overline{\varOmega }}), \end{aligned}$$

and observe that

$$\begin{aligned} {\mathcal {R}}(v)\equiv 0 \hbox { in }B{\setminus } \overline{\varOmega }\quad \hbox {and}\quad {\mathcal {R}}(v)\in W^{1,p}(B). \end{aligned}$$

Since \(\varOmega \) has continuous boundary, this implies that \({\mathcal {R}}(v)\in {\mathcal {D}}^{1,p}_0(\varOmega )\), see Remark 2.1. We now fix \(u\in C^\infty _0(B)\), for every \(v\in C^\infty _0(B)\) and every \(t,\varepsilon >0\), we take \(\varphi _{\varepsilon ,t}\in C^\infty _0(\varOmega )\) such that

$$\begin{aligned} \Big (\lambda ^1_p(\varOmega )\Big )^\frac{1}{p}\,\Vert \varphi _\varepsilon -{\mathcal {R}}(v)\Vert _{L^p(\varOmega )}\le \Vert \nabla \varphi _{\varepsilon ,t}-\nabla \mathcal {R}(v)\Vert _{L^p(\varOmega )}<\varepsilon \,\left( t+\left( \lambda ^1_p(\varOmega )\right) ^{-\frac{1}{p}}\right) ^{-1}. \end{aligned}$$

This is possible, thanks to the definition of \({\mathcal {D}}^{1,p}_0(\varOmega )\). Then for \(t>0\) we can estimate the relevant K-functional as follows

$$\begin{aligned} \begin{aligned} K(t,{\mathcal {R}}(u),L^p(\varOmega ),{\mathcal {D}}^{1,p}_0(\varOmega ))&\le \Vert {\mathcal {R}}(u)-\varphi _{\varepsilon ,t}\Vert _{L^p(\varOmega )}+t\,\Vert \nabla \varphi _{\varepsilon ,t}\Vert _{L^p(\varOmega )}\\&\le \Vert {\mathcal {R}}(u)-{\mathcal {R}}(v)\Vert _{L^p(\varOmega )}+\Vert \varphi _{\varepsilon ,t}-{\mathcal {R}}(v)\Vert _{L^p(\varOmega )}\\&+t\,\Vert \nabla {\mathcal {R}}(v)\Vert _{L^p(\varOmega )}+t\,\Vert \nabla \varphi _{\varepsilon ,t}-\nabla {\mathcal {R}}(v)\Vert _{L^p(\varOmega )}\\&\le \Vert {\mathcal {R}}(u-v)\Vert _{L^p(\varOmega )}+t\,\Vert \nabla {\mathcal {R}}(v)\Vert _{L^p(\varOmega )}\\&+\left( t+\left( \lambda ^1_p(\varOmega )\right) ^{-\frac{1}{p}}\right) \,\Vert \nabla \varphi _{\varepsilon ,t}-\nabla {\mathcal {R}}(v)\Vert _{L^p(\varOmega )}\\&\le \Vert u-v\Vert _{L^p(\varOmega )}+\Vert {\mathcal {T}}((u-v)\cdot 1_{B{\setminus } \overline{\varOmega }})\Vert _{L^p(\varOmega )}\\&+t\,\left( \Vert \nabla v\Vert _{L^p(\varOmega )}+\Vert \nabla {\mathcal {T}}(v\cdot 1_{B{\setminus }\overline{\varOmega }})\Vert _{L^p(\varOmega )}\right) +\varepsilon . \end{aligned} \end{aligned}$$

By applying (4.11) and using that

$$\begin{aligned} \Vert \nabla {\mathcal {T}}(v\cdot 1_{B{\setminus }\overline{\varOmega }})\Vert _{L^p(B)}\le \mathfrak {e}_\varOmega \,\Vert v\cdot 1_{B{\setminus }\overline{\varOmega }}\Vert _{W^{1,p}(B{\setminus }\overline{\varOmega })}=\mathfrak {e}_\varOmega \,\Vert v\Vert _{W^{1,p}(B{\setminus }\overline{\varOmega })}\le \mathfrak {e}_\varOmega \,\Vert v\Vert _{W^{1,p}(B)}, \end{aligned}$$

we then get

$$\begin{aligned} \begin{aligned} K(t,{\mathcal {R}}(u),L^p(\varOmega ),{\mathcal {D}}^{1,p}_0(\varOmega ))&\le (1+\mathfrak {e}_\varOmega )\,\Vert u-v\Vert _{L^p(B)}+t\,\left( \Vert \nabla v\Vert _{L^p(B)}+\mathfrak {e}_\varOmega \,\Vert v\Vert _{W^{1,p}(B)}\right) +\varepsilon . \end{aligned} \end{aligned}$$

We now use that

$$\begin{aligned} \Vert v\Vert _{W^{1,p}(B)}=\left( \Vert v\Vert ^p_{L^p(B)}+\Vert \nabla v\Vert ^p_{L^p(B)}\right) ^\frac{1}{p}\le \Vert \nabla v\Vert _{L^p(B)}\,\left( 1+\frac{1}{\lambda ^1_p(B)}\right) ^\frac{1}{p}, \end{aligned}$$

thanks to Poincaré inequality. By spending this information in the previous estimate and using the arbitrariness of \(\varepsilon \), we get

$$\begin{aligned} \begin{aligned} K(t,{\mathcal {R}}(u),L^p(\varOmega ),{\mathcal {D}}^{1,p}_0(\varOmega ))&\le (1+\mathfrak {e}_\varOmega )\,\Vert u-v\Vert _{L^p(B)}\\&+t\,\left( 1+\mathfrak {e}_\varOmega \,\left( 1+\frac{1}{\lambda ^1_p(B)}\right) ^\frac{1}{p}\right) \,\Vert \nabla v\Vert _{L^p(B)}. \end{aligned} \end{aligned}$$

We set for simplicity

$$\begin{aligned} \gamma _\varOmega =1+\mathfrak {e}_\varOmega \,\left( 1+\frac{1}{\lambda ^1_p(B)}\right) ^\frac{1}{p}, \end{aligned}$$

then by taking the infimum over \(v\in C^\infty _0(B)\)

$$\begin{aligned} K(t,{\mathcal {R}}(u),L^p(\varOmega ),{\mathcal {D}}^{1,p}_0(\varOmega ))\le \gamma _\varOmega \,K(t,u,L^p(B),{\mathcal {D}}^{1,p}_0(B)). \end{aligned}$$

As usual, we integrate in t, so to get

$$\begin{aligned} \Vert {\mathcal {R}}(u)\Vert ^p_{{\mathcal {X}}^{s,p}_0(\varOmega )}\le \gamma _\varOmega ^p\,\Vert u\Vert ^p_{{\mathcal {X}}^{s,p}_0(B)},\quad \hbox {for } u\in C^\infty _0(B). \end{aligned}$$
(4.12)

We now observe that if \(u\in C^\infty _0(\varOmega )\), then we have \({\mathcal {R}}(u)=u\). Thus, from (4.12) and Theorem 4.7 for the convex set B, we get

$$\begin{aligned} \Vert u\Vert ^p_{{\mathcal {X}}^{s,p}_0(\varOmega )}\le C\,\gamma _\varOmega ^p\,\Vert u\Vert ^p_{{\mathcal {D}}^{s,p}_0({\mathbb {R}}^N)}=C\,\gamma _\varOmega ^p\,\Vert u\Vert ^p_{{\mathcal {D}}^{s,p}_0(\varOmega )},\quad \hbox { for every }u\in C^\infty _0(\varOmega ), \end{aligned}$$

where C only depends on N and p. This concludes the proof. \(\square \)

Remark 4.11

(More general sets) It is not difficult to see that the previous proof works (and thus \({\mathcal {X}}^{s,p}_0(\varOmega )\) and \({\mathcal {D}}^{s,p}_0(\varOmega )\) are equivalent), whenever the set \(\varOmega \) is such that there exists a linear and continuous extension operator

$$\begin{aligned} {\mathcal {T}}: W^{1,p}(B{\setminus }\overline{\varOmega })\rightarrow W^{1,p}(B), \end{aligned}$$

such that (4.11) holds. Observe that there is a vicious subtlety here: the first condition in (4.11) is vital and, in general, it may fail to hold for an extension operator. For example, there is a beautiful extension result by Jones [21, Theorem 1], which is valid for very irregular domains (possibly having a fractal boundary): however, the construction given by Jones does not assure that the first estimate in (4.11) holds true, see the statement of [21, Lemma 3.2].

In order to complement the discussion of Remarks 3.4 and 4.6 on “extensions by zero” in \({\mathcal {X}}^{s,p}_0\), we explicitly state the following consequence of Proposition 4.1 and Theorem 4.10.

Corollary 4.12

Let \(1<p<\infty \) and \(0<s<1\). Let \(\varOmega \subset {\mathbb {R}}^N\) be an open bounded set, with Lipschitz boundary. Then for every \(u\in C^\infty _0(\varOmega )\), we have

$$\begin{aligned} \Vert u\Vert ^p_{{\mathcal {X}}^{s,p}_0(\varOmega )}\le 2^{p\,(1-s)}\,N\,\omega _N\,C_1\, \Vert u\Vert ^p_{{\mathcal {X}}^{s,p}_0({\mathbb {R}}^N)}, \end{aligned}$$

where \(C_1>0\) is the same constant as in Theorem 4.10.

5 Capacities

Let \(1<p<N\), we recall that for every compact set \(F\subset {\mathbb {R}}^N\), its p-capacity is defined by

$$\begin{aligned} {\mathrm {cap}}_p(F)=\inf _{u\in C^\infty _0({\mathbb {R}}^N)}\left\{ \int _{{\mathbb {R}}^N} |\nabla u|^p\,\mathrm{d}x\, :\, u\ge 0\hbox { and } u\ge 1_F\right\} , \end{aligned}$$

see [17, Chapter 4, Section 7].

Similarly, given \(1<p<\infty \) and \(0<s<1\) such thatFootnote 3\(s\,p<N\), we define the (sp)-capacity ofF through

$$\begin{aligned} {\mathrm {cap}}_{s,p}(F)=\inf _{u\in C^\infty _0({\mathbb {R}}^N)}\left\{ [u]^p_{W^{s,p}({\mathbb {R}}^N)}\, :\, u\ge 0\hbox { and } u\ge 1_F\right\} , \end{aligned}$$

and the interpolation (sp)-capacity ofF by

$$\begin{aligned} {{\mathrm {int\,cap}}}_{s,p}(F)=\inf _{u\in C^\infty _0({\mathbb {R}}^N)}\left\{ \Vert u\Vert ^p_{{\mathcal {X}}^{s,p}_0({\mathbb {R}}^N)}\, :\, u\ge 0 \hbox { and }u\ge 1_F\right\} . \end{aligned}$$

As a straightforward consequence of Propositions 4.1 and 4.5, we have the following

Corollary 5.1

(Comparison of capacities) Let \(1<p<\infty \) and \(0<s<1\) be such that \(s\,p<N\). Let \(F\subset {\mathbb {R}}^N\) be a compact set, then we have

$$\begin{aligned} \frac{1}{C}\,{\mathrm {cap}}_{s,p}(F)\le {{\mathrm {int\,cap}}}_{s,p}(F)\le C\,{\mathrm {cap}}_{s,p}(F), \end{aligned}$$

for a constant \(C=C(N,p)>1\). In particular, it holds

$$\begin{aligned} {\mathrm {cap}}_{s,p}(F)=0\quad \hbox {if and only if}\quad {{\mathrm {int\,cap}}}_{s,p}(F)=0. \end{aligned}$$

Proposition 5.2

Let \(1<p<\infty \) and \(0<s<1\) be such that \(s\,p<N\). For every \(E,F\subset {\mathbb {R}}^N\) compact sets, we have

$$\begin{aligned} {\mathrm {cap}}_{s,p}(E\cup F)\le {\mathrm {cap}}_{s,p}(E)+{\mathrm {cap}}_{s,p}(F). \end{aligned}$$

Proof

We fix \(n\in {\mathbb {N}}{\setminus }\{0\}\) and choose two nonnegative functions \(\varphi _n,\psi _n\in C^\infty _0({\mathbb {R}}^N)\) such that

$$\begin{aligned}{}[\varphi _n]^p_{W^{s,p}({\mathbb {R}}^N)}\le {\mathrm {cap}}_{s,p}(E)+\frac{1}{n},\quad \varphi _n\ge 1_E, \end{aligned}$$

and

$$\begin{aligned}{}[\psi _n]^p_{W^{s,p}({\mathbb {R}}^N)}\le {\mathrm {cap}}_{s,p}(F)+\frac{1}{n},\quad \psi _n\ge 1_F. \end{aligned}$$

We then set

$$\begin{aligned} U_{n,\varepsilon }=\Big (\max \{\varphi _n,\psi _n\}\Big )*\varrho _\varepsilon ,\quad 0<\varepsilon \ll 1, \end{aligned}$$

where \(\{\varrho _\varepsilon \}_{\varepsilon >0}\) is a family of standard Friedrichs mollifiers. We observe that for every \(n\in {\mathbb {N}}{\setminus }\{0\}\), it holds that \(U_{n,\varepsilon }\in C^\infty _0({\mathbb {R}}^N)\). Moreover, by construction we have

$$\begin{aligned} U_{n,\varepsilon }\ge 1_{E\cup F}. \end{aligned}$$

By observing that Jensen’s inequality implies

$$\begin{aligned}{}[U_{n,\varepsilon }]_{W^{s,p}({\mathbb {R}}^N)}\le \Big [\max \{\varphi _n,\psi _n\}\Big ]^p_{W^{s,p}({\mathbb {R}}^N)}, \end{aligned}$$

we thus get

$$\begin{aligned} {\mathrm {cap}}_{s,p}(E\cup F)\le [U_{n,\varepsilon }]_{W^{s,p}({\mathbb {R}}^N)}\le \Big [\max \{\varphi _n,\psi _n\}\Big ]^p_{W^{s,p}({\mathbb {R}}^N)}. \end{aligned}$$

By using the sub-modularity of the Sobolev–Slobodeckiĭ seminorm (see [19, Theorem 3.2 & Remark 3.3]), we obtain

$$\begin{aligned} {\mathrm {cap}}_{s,p}(E\cup F)\le [\varphi _n]^p_{W^{s,p}({\mathbb {R}}^N)}+[\psi _n]^p_{W^{s,p}({\mathbb {R}}^N)}. \end{aligned}$$

Finally, thanks to the choice of \(\varphi _n\) and \(\psi _n\), we get the desired conclusion by the arbitrariness of n. \(\square \)

In the next result, we denote by \({\mathcal {H}}^\tau \) the \(\tau \)-dimensional Hausdorff measure.

Proposition 5.3

Let \(1<p<\infty \) and \(0<s<1\) be such that \(s\,p<N\). Let \(\varOmega \subset {\mathbb {R}}^N\) be an open set. We take a compact set \(E\Subset \varOmega \) such that

$$\begin{aligned} {\mathrm {cap}}_{s,p}(E)=0. \end{aligned}$$

Then we have

$$\begin{aligned} {\mathcal {H}}^\tau (E)=0\quad \hbox {for every }\tau >N-s\,p, \end{aligned}$$
(5.1)

and

$$\begin{aligned} \lambda ^s_p(\varOmega {\setminus } E)=\lambda ^s_p(\varOmega ). \end{aligned}$$
(5.2)

Proof

To prove (5.1), we can easily adapt the proof of [17, Theorem 4, p. 156], dealing with the local case.

In order to prove (5.2), we first assume \(\varOmega \) to be bounded. Let \(\varepsilon >0\), we take \(u_\varepsilon \in C^\infty _0(\varOmega )\) such that

$$\begin{aligned} \Vert u_\varepsilon \Vert _{{\mathcal {D}}^{s,p}_0(\varOmega )}^p<(1+\varepsilon )\,\lambda ^s_p(\varOmega )\quad \hbox {and}\quad \int _\varOmega |u_\varepsilon |^p\,\mathrm{d}x=1. \end{aligned}$$

We further observe that the boundedness of \(\varOmega \) implies that

$$\begin{aligned} \lambda ^s_p(\varOmega )=\inf _{u\in C^\infty _0(\varOmega )} \Big \{\Vert u\Vert ^p_{{\mathcal {D}}^{s,p}_0(\varOmega )}\, :\, \Vert u\Vert _{L^p(\varOmega )}=1\Big \} =\min _{u\in {\mathcal {D}}^{s,p}_0(\varOmega )} \Big \{\Vert u\Vert ^p_{{\mathcal {D}}^{s,p}_0(\varOmega )}\, :\, \Vert u\Vert _{L^p(\varOmega )}=1\Big \}, \end{aligned}$$

and that any solution \(u\in {\mathcal {D}}^{s,p}_0(\varOmega )\) has norm \(L^\infty (\varOmega )\) bounded by a constant \(M=M(N,s,p,\varOmega )\), see [7, Theorem 3.3]. Thus, without the loss of generality, we can also assume that

$$\begin{aligned} \Vert u_\varepsilon \Vert _{L^\infty (\varOmega )}\le M+1,\quad \hbox {for }0<\varepsilon \ll 1. \end{aligned}$$

Since E has null (sp)-capacity, there exists \(\varphi _\varepsilon \in C^\infty _0(\varOmega )\) such that

$$\begin{aligned}{}[\varphi _\varepsilon ]_{W^{s,p}({\mathbb {R}}^N)}^p<\varepsilon ,\quad \varphi _\varepsilon \ge 0\quad \hbox {and}\quad \varphi _\varepsilon \ge 1_E. \end{aligned}$$

We set \(\psi _\varepsilon =\varphi _\varepsilon /\Vert \varphi _\varepsilon \Vert _{L^\infty ({\mathbb {R}}^N)}\) and observe that \(\Vert \varphi _\varepsilon \Vert _{L^\infty ({\mathbb {R}}^N)}\ge 1\). The function \(u_\varepsilon \,(1-\psi _\varepsilon )\) is admissible for the variational problem defining \(\lambda ^s_p(\varOmega {\setminus } E)\); then by using the triangle inequality, we have

$$\begin{aligned} \begin{aligned} \Big (\lambda ^s_{p}(\varOmega {\setminus } E)\Big )^\frac{1}{p}\le \frac{[u_\varepsilon \,(1-\psi _\varepsilon )]_{W^{s,p}({\mathbb {R}}^N)}}{\Vert u_\varepsilon \,(1-\psi _\varepsilon )\Vert _{L^p(\varOmega {\setminus } E)}}&\le \frac{[u_\varepsilon ]_{W^{s,p}({\mathbb {R}}^N)}\,\Vert 1-\psi _\varepsilon \Vert _{L^\infty ({\mathbb {R}}^N)}}{\Vert u_\varepsilon \,(1-\psi _\varepsilon )\Vert _{L^p(\varOmega {\setminus } E)}}\\&\quad +\frac{\Vert u_\varepsilon \Vert _{L^\infty }\,[\psi _\varepsilon ]_{W^{s,p}({\mathbb {R}}^N)}}{\Vert u_\varepsilon \,(1-\psi _\varepsilon )\Vert _{L^p(\varOmega {\setminus } E)}}. \end{aligned} \end{aligned}$$

From the first part of the proof, we know that E has N-dimensional Lebesgue measure 0; thus, the \(L^p\) norm over \(\varOmega {\setminus } E\) is the same as that over \(\varOmega \). If we now take the limit as \(\varepsilon \) goes to 0 and use the properties of \(u_\varepsilon \), together withFootnote 4

$$\begin{aligned}{}[\psi _\varepsilon ]^p_{W^{s,p}({\mathbb {R}}^N)}<\frac{\varepsilon ^p}{\Vert \varphi _\varepsilon \Vert _{L^\infty ({\mathbb {R}}^N)}}\le \varepsilon ^p, \end{aligned}$$

and

$$\begin{aligned} \lim _{\varepsilon \rightarrow 0}\Vert 1-\psi _\varepsilon \Vert _{L^\infty ({\mathbb {R}}^N)}=\lim _{\varepsilon \rightarrow 0} \sup _{{\mathbb {R}}^N}\, (1-\psi _\varepsilon )\le 1, \end{aligned}$$

we get

$$\begin{aligned} \Big (\lambda ^s_{p}(\varOmega {\setminus } E)\Big )^\frac{1}{p}\le \Big (\lambda ^s_{p}(\varOmega )\Big )^\frac{1}{p}. \end{aligned}$$

The reverse inequality simply follows from the fact that \(C^\infty _0(\varOmega {\setminus } E)\subset C^\infty _0(\varOmega )\); thus, we get the conclusion when \(\varOmega \) is bounded.

In order to remove the last assumption, we consider the sets \(\varOmega _R=\varOmega \cap B_R(0)\). For R large enough, this is a nonempty open bounded set and \(E\Subset \varOmega _R\) as well. We thus have

$$\begin{aligned} \lambda ^s_p(\varOmega _R{\setminus } E)=\lambda ^s_p(\varOmega _R). \end{aligned}$$

By taking the limitFootnote 5 as R goes to \(+\infty \), we get the desired conclusion in the general case as well. \(\square \)

The previous result giving the link between the Poincaré constant and sets with null capacity does not hold true in the interpolation space \({\mathcal {X}}^{s,p}_0(\varOmega )\). Indeed, we have the following result, which shows that the interpolation Poincaré constant is sensitive to removing sets with null (sp)-capacity.

Lemma 5.4

Let \(1<p<N\) and \(0<s<1\). Let \(\varOmega \subset {\mathbb {R}}^N\) be an open set and \(E\Subset \varOmega \) a compact set such that

$$\begin{aligned} {{\mathrm {int\,cap}}}_{s,p}(E)=0<{\mathrm {cap}}_p(E). \end{aligned}$$

Then we have

$$\begin{aligned} \varLambda ^s_{p}(\varOmega {\setminus } E)>\varLambda ^s_p(\varOmega ). \end{aligned}$$

Proof

By Corollary 3.3, we know that

$$\begin{aligned} p\,s\,(1-s)\varLambda ^s_{p}(\varOmega {\setminus } E)=\Big (\lambda ^1_p(\varOmega {\setminus } E)\Big )^s \quad \hbox {and}\quad p\,s\,(1-s)\,\varLambda ^s_p(\varOmega )=\Big (\lambda ^1_p(\varOmega )\Big )^s. \end{aligned}$$

It is now sufficient to use that \(\lambda ^1_p(\varOmega {\setminus } E)>\lambda ^1_p(\varOmega )\), as a consequence of the fact that E has positive p-capacity. \(\square \)

Remark 5.5

As an explicit example of the previous situation, we can take \(s\,p<1\) and the \((N-1)\)-dimensional set

$$\begin{aligned} F=[-a,a]^{N-1}\times \{0\}. \end{aligned}$$

Observe that \({\mathrm {cap}}_p(F)>0\) by [17, Theorem 4, p. 156]. On the other hand, we have

$$\begin{aligned} {\mathrm {int\,cap}}_{s,p}(F)=0. \end{aligned}$$

Indeed, we set

$$\begin{aligned} F_\varepsilon =\left\{ x\in {\mathbb {R}}^N \, :\, {\mathrm {dist}}(x,F)<\varepsilon \right\} . \end{aligned}$$

We then take the usual sequence of Friedrichs mollifiers \(\{\varrho _\varepsilon \}_{\varepsilon >0}\subset C^\infty _0({\mathbb {R}}^N)\) and define

$$\begin{aligned} \varphi _{\varepsilon }=1_{F_\varepsilon }*\varrho _\varepsilon \in C^\infty _0({\mathbb {R}}^N). \end{aligned}$$

Observe that by construction we have

$$\begin{aligned} \varphi _\varepsilon \equiv 1 \hbox { on } F_\varepsilon \quad \hbox {and}\quad \varphi _\varepsilon \equiv 0\hbox { on } {\mathbb {R}}^N{\setminus } F_{2\,\varepsilon }. \end{aligned}$$

By definition of (sp)-capacity and using the interpolation estimate (4.3), we get

$$\begin{aligned} \begin{aligned} {\mathrm {cap}}_{s,p}(F)&\le [\varphi _\varepsilon ]_{W^{s,p}({\mathbb {R}}^N)}^p\\&\le C\,\left( \int _{{\mathbb {R}}^N} |\varphi _\varepsilon |^p\,\mathrm{d}x\right) ^{1-s}\,\left( \int _{{\mathbb {R}}^N} |\nabla \varphi _\varepsilon |^p\,\mathrm{d}x\right) ^s\\&\le C\, \left( \int _{{\mathbb {R}}^N} |1_{F_\varepsilon }|^p\,\mathrm{d}x\right) ^{1-s}\,\left( \int _{{\mathbb {R}}^N} |1_{F_\varepsilon }|^p\,\mathrm{d}x\right) ^s \left( \int _{{\mathbb {R}}^N} |\nabla \varrho _\varepsilon |\,\mathrm{d}x\right) ^{s\,p}\\&\le C\,|F_\varepsilon |\,\varepsilon ^{-s\,p}\le C\,\varepsilon ^{1-s\,p}. \end{aligned} \end{aligned}$$

We then observe that the last quantity goes to 0 as \(\varepsilon \) goes to 0, thanks to the fact that \(s\,p<1\). By Corollary 5.1, we have

$$\begin{aligned} {\mathrm {int\, cap}}_{s,p}(F)={\mathrm {cap}}_{s,p}(F)=0. \end{aligned}$$

as desired.

6 Double-sided estimates for Poincaré constants

We already observed that for an open set \(\varOmega \subset {\mathbb {R}}^N\) we have

$$\begin{aligned} p\,s\,(1-s)\,\varLambda ^s_{p}(\varOmega )=\Big (\lambda ^1_p(\varOmega )\Big )^s. \end{aligned}$$

We now want to compare \(\lambda ^1_p\) with the sharp Poincaré constant for the embedding \({\mathcal {D}}^{s,p}_0(\varOmega )\hookrightarrow L^p(\varOmega )\).

Theorem 6.1

Let \(1<p<\infty \) and \(0<s<1\). Let \(\varOmega \subset {\mathbb {R}}^N\) be an open set, then

$$\begin{aligned} s\,(1-s)\,\lambda ^s_{p}(\varOmega )\le \frac{2^{p\,(1-s)}\,N\,\omega _N}{p}\,\Big (\lambda ^1_{p}(\varOmega )\Big )^s. \end{aligned}$$
(6.1)

If in addition:

  • \(\varOmega \subset {\mathbb {R}}^N\) is bounded with Lipschitz boundary, then we also have the reverse inequality

    $$\begin{aligned} \frac{1}{p\,C_1}\,\Big (\lambda ^1_{p}(\varOmega )\Big )^s\le s\,(1-s)\,\lambda ^s_{p}(\varOmega ), \end{aligned}$$
    (6.2)

    where \(C_1>0\) is the same constant as in Theorem 4.10;

  • \(\varOmega \subset {\mathbb {R}}^N\) is convex, then we also have the reverse inequality

    $$\begin{aligned} \frac{1}{C_2}\,\Big (\lambda ^1_{p}(\varOmega )\Big )^s\le s\,(1-s)\,\lambda ^s_{p}(\varOmega ), \end{aligned}$$
    (6.3)

    where \(C_2\) is the universal constant given by

    $$\begin{aligned} C_2=\frac{\Big (\lambda ^1_p(B_1(0))\Big )^s}{{\mathcal {C}}}, \end{aligned}$$

    and \({\mathcal {C}}={\mathcal {C}}(N,p)>0\) is the same constant as in the Hardy inequality for \({\mathcal {D}}^{s,p}_0(\varOmega )\) (see [8, Theorem 1.1]).

Proof

The first inequality (6.1) is a direct consequence of the interpolation inequality (4.3). Indeed, by using the definition of \(\lambda ^s_{p}(\varOmega )\), we obtain from this inequality

$$\begin{aligned} s\,(1-s)\,\lambda ^s_{p}(\varOmega )\,\Vert u\Vert ^p_{L^p(\varOmega )}\le \frac{2^{p\,(1-s)}\,N\,\omega _N}{p}\, \Vert u\Vert _{L^p(\varOmega )}^{(1-s)\,p}\,\Vert \nabla u\Vert _{L^p(\varOmega )}^{s\,p}, \end{aligned}$$

for every \(u\in C^\infty _0(\varOmega )\). By simplifying the factor \(\Vert u\Vert ^p_{L^p(\varOmega )}\) on both sides and taking the infimum over \(C^\infty _0(\varOmega )\), we get the claimed inequality.

In order to prove (6.2), for every \(\varepsilon >0\) we take \(\varphi \in C^\infty _0(\varOmega )\) such that

$$\begin{aligned} \frac{\Vert \varphi \Vert _{{\mathcal {D}}^{s,p}_0(\varOmega )}^p}{\Vert \varphi \Vert ^p_{L^p(\varOmega )}}<\lambda ^s_p(\varOmega )+\varepsilon , \end{aligned}$$

then we use Theorem 4.10 to infer

$$\begin{aligned} \frac{1}{C_1}\,\frac{\Vert \varphi \Vert _{{\mathcal {X}}^{s,p}_0(\varOmega )}^p}{\Vert \varphi \Vert ^p_{L^p(\varOmega )}}<\lambda ^s_p(\varOmega )+\varepsilon . \end{aligned}$$

This in turn implies

$$\begin{aligned} \frac{1}{C_1}\,\varLambda ^s_p(\varOmega )\le \lambda ^s_p(\varOmega ), \end{aligned}$$

by arbitrariness of \(\varepsilon >0\). A further application of Corollary 3.3 leads to the desired conclusion.

Finally, if \(\varOmega \subset {\mathbb {R}}^N\) is convex, we can proceed in a different way. We first observe that we can always suppose that the inradius \(R_\varOmega \) is finite; otherwise, both \(\lambda ^1_p(\varOmega )\) and \(\lambda ^s_p(\varOmega )\) vanish, and there is nothing to prove. Then (6.3) comes by joining the simple estimate

$$\begin{aligned} \lambda ^1_{p}(\varOmega )\le \frac{\lambda ^1_p(B_1(0))}{R_\varOmega ^p}, \end{aligned}$$

which follows from the monotonicity and scaling properties of \(\lambda ^1_p\), and the estimate of [8, Corollary 5.1], i.e.,

$$\begin{aligned} s\,(1-s)\,\lambda ^s_p(\varOmega )\ge \frac{{\mathcal {C}}}{R_\varOmega ^{s\,p}}. \end{aligned}$$

The latter is a consequence of the Hardy inequality in convex sets for \({\mathcal {D}}^{s,p}_0\). \(\square \)

Remark 6.2

For \(p=2\), the double-sided estimate of Theorem 6.1 is contained in [12, Theorem 4.5]. The proof in [12] relies on probabilistic techniques, and the result is proved by assuming that \(\varOmega \) verifies a uniform exterior cone condition.

Remark 6.3

Inequality (6.2) cannot hold for a general open set \(\varOmega \subset {\mathbb {R}}^N\), with a constant independent of \(\varOmega \). Indeed, one can construct a sequence \(\{\varOmega _n\}_{n\in {\mathbb {N}}}\subset {\mathbb {R}}^N\) such that

$$\begin{aligned} \lim _{n\rightarrow \infty }\frac{\Big (\lambda ^1_{p}(\varOmega _n)\Big )^s}{\lambda ^s_{p}(\varOmega _n)}=+\infty ,\quad \hbox {for } 1<p<\infty \hbox { and } s< \frac{1}{p}, \end{aligned}$$

see Lemma A.1.