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

Equivalence between validity of the pp-Poincaré inequality and finiteness of the strict pp-capacitary inradius.

A.-K. Gallagher 0000-0001-5269-2879 Gallagher Tool & Instrument LLC, Redmond, WA 98052, USA [email protected]
Abstract.

It is shown that the pp-Poincaré inequality holds on an open set Ω\Omega in n\mathbb{R}^{n} if and only if the strict pp-capacitary inradius of Ω\Omega is finite. To that end, new upper and lower bounds for the infimum of the associated nonlinear Rayleigh quotients are derived.

Key words and phrases:
pp-Poincaré inequality, capacitary inradius, pp-Laplacian
1991 Mathematics Subject Classification:
35P30, 35J70, 31C45

1. Introduction

Let Ω\Omega be an open set in n\mathbb{R}^{n}, and .p,Ω\|\,.\,\|_{p,\Omega} the LpL^{p}-norm on Ω\Omega for 1<p<1<p<\infty. The classical pp-Poincaré inequality is said to hold on Ω\Omega if there exists a constant C>0C>0 such that

(1.1) fp,ΩCfp,Ωfor allf𝒞c(Ω),\displaystyle\|f\|_{p,\Omega}\leq C\|\nabla f\|_{p,\Omega}\;\;\hskip 5.69046pt\text{for all}\;f\in\mathcal{C}^{\infty}_{c}(\Omega),

where 𝒞c(Ω)\mathcal{C}_{c}^{\infty}(\Omega) denotes the space of smooth functions with compact support in Ω\Omega. That is, (1.1) holds if and only if the infimum of the so-called nonlinear Rayleigh quotient

λ1,p(Ω)=inf{fp,Ωp/fp,Ωp:f𝒞c(Ω){0}}\displaystyle\lambda_{1,p}(\Omega)=\inf\left\{\|\nabla f\|_{p,\Omega}^{p}/\|f\|_{p,\Omega}^{p}\;:\;f\in\mathcal{C}^{\infty}_{c}(\Omega)\setminus\{0\}\right\}

is positive. If λ1,p(Ω)\lambda_{1,p}(\Omega) is positive and attained at some non-zero function in W01,p(Ω)W^{1,p}_{0}(\Omega), i.e., the closure of 𝒞c(Ω)\mathcal{C}_{c}^{\infty}(\Omega) with respect to the Lp(Ω)L^{p}(\Omega)-Sobolev 1-norm, then it is the smallest generalized eigenvalue for the pp-Laplacian, with Dirichlet boundary condition, on Ω\Omega in the distributional sense.

The validity of (1.1), or equivalently the positivity of λ1,p(.)\lambda_{1,p}(.), plays a role in establishing existence and uniqueness of solutions to certain quasilinear elliptic equations, see [20, Theorem 1.2] as well as in determining the asymptotic behavior of solutions to some nonlinear parabolic equations, see [18, Sections 3&4]. Furthermore, in the case of n=pn=p, it was recently shown in [5, Theorems 1&2] that a local version of the Moser–Trudinger inequality implies the global Moser–Trudinger inequality if and only (1.1) holds.

In this note, we employ the concept of strict pp-capacitary inradius of Ω\Omega, originally introduced in the case of p=2p=2 in [7] for n=2n=2 and in [6] for n3n\geq 3, and derive the equivalence between its finiteness and the positivity of λ1,p(Ω)\lambda_{1,p}(\Omega). In the definition of the strict pp-capacitary inradius, we use the notion of Sobolev pp-capacity, Cp(K)C_{p}(K), which, for a compact set KnK\subset\mathbb{R}^{n}, is defined as

Cp(K)=inf{up,np+up,np:u𝒮,u1 on K},\displaystyle C_{p}(K)=\inf\left\{\|u\|_{p,\mathbb{R}^{n}}^{p}+\|\nabla u\|_{p,\mathbb{R}^{n}}^{p}\;:\;u\in\mathcal{S},\;u\geq 1\text{ on }K\right\},

where 𝒮\mathcal{S} denotes the space of Schwartz functions. The strict pp-capacitary inradius of Ω\Omega is defined as

(1.2) ρp(Ω)=sup{r>0:ϵ>0xn such that Cp(𝔹r(x)¯Ωc)<ϵ},\displaystyle\rho_{p}(\Omega)=\sup\left\{r>0\;:\;\forall\;\epsilon>0\;\exists\;x\in\mathbb{R}^{n}\text{ such that }C_{p}(\overline{\mathbb{B}_{r}(x)}\cap\Omega^{c})<\epsilon\right\},

where 𝔹r(x)\mathbb{B}_{r}(x) denotes the open ball in n\mathbb{R}^{n} of radius r>0r>0 with center xnx\in\mathbb{R}^{n}. Roughly speaking, finiteness of ρp(Ω)\rho_{p}(\Omega) means that the complement of Ω\Omega is somewhat evenly und uniformly distributed in n\mathbb{R}^{n}. In fact, if ρp(Ω)\rho_{p}(\Omega) is finite and R>ρp(Ω)R>\rho_{p}(\Omega), then there exists a δ>0\delta>0 such that within RR-units of any point in n\mathbb{R}^{n}, we may find a set in the complement of Ω\Omega whose Sobolov pp-capacity is at least δ\delta.

Finiteness of ρp(Ω)\rho_{p}(\Omega) is a necessary condition for (1.1) to hold on an open set Ω\Omega. Indeed, our first result yields a sharp upper bound for ρp(Ω)\rho_{p}(\Omega) in terms of λ1,p(Ω)\lambda_{1,p}(\Omega).

Theorem 1.3.

Let Ωn\Omega\subset\mathbb{R}^{n} be an open set, 1<p<1<p<\infty. Suppose λ1,p(Ω)>0\lambda_{1,p}(\Omega)>0.

  • (i)

    Then

    (1.4) (ρp(Ω))pλ1,p(𝔹1(0))/λ1,p(Ω).\displaystyle\left(\rho_{p}(\Omega)\right)^{p}\leq\lambda_{1,p}(\mathbb{B}_{1}(0))/\lambda_{1,p}(\Omega).
  • (ii)

    If Ω\Omega is connected and bounded, then equality in (1.4) holds if and only if

    Ω=𝔹ρp(Ω)(x)K\displaystyle\Omega=\mathbb{B}_{\rho_{p}(\Omega)}(x)\setminus K

    for some xnx\in\mathbb{R}^{n} and some compact set KK with Cp(K)=0C_{p}(K)=0. Furthermore, if Ω\Omega is connected and unbounded such that ρp(Ω)=ρp(Ω𝔹R(0))\rho_{p}(\Omega)=\rho_{p}(\Omega\cap\mathbb{B}_{R}(0)) for some R>0R>0, then equality in (1.4) cannot hold.

The heart of the matter of the proof of Theorem 1.3 is a continuity result for λ1,p\lambda_{1,p}. That is, suppose {Kj}j\{K_{j}\}_{j\in\mathbb{N}} is a sequence of compact sets contained in the closure of some bounded domain DD with smooth boundary. If Cp(Kj)C_{p}(K_{j}) tends to zero as jj\to\infty and, for each jj\in\mathbb{N}, DKjD\setminus K_{j} is an open set with smooth boundary, then λ1,p(DKj)\lambda_{1,p}(D\setminus K_{j}) tends to λ1,p(D)\lambda_{1,p}(D). This was orginally proved in [7] for the case n=2n=2, p=2p=2, D=𝔹1(0)D=\mathbb{B}_{1}(0), and the logarithmic capacity in place of CpC_{p}. Regularity results for the Dirichlet problem and for first eigenfunctions of the pp-Laplacian, Δp\Delta_{p}, on open, bounded sets with smooth boundary as well as the theory of pp-harmonic functions allow this continuity result to be extended to the case of the Sobolev pp-capacity with 1<p<1<p<\infty, see Lemma 3.1.

To state the sufficiency of the finiteness of ρp(Ω)\rho_{p}(\Omega), we define the scalar

(1.5) δR(Ω):=sup{δ0:Cp(𝔹R(x)¯Ωc)δfor allxn}\displaystyle\delta_{R}(\Omega):=\sup\left\{\delta\geq 0\;:\;C_{p}(\overline{\mathbb{B}_{R}(x)}\cap\Omega^{c})\geq\delta\;\text{for all}\;x\in\mathbb{R}^{n}\right\}

for any R>0R>0. Note that δR(Ω)\delta_{R}(\Omega) is positive whenever ρp(Ω)\rho_{p}(\Omega) is finite and R>ρp(Ω)R>\rho_{p}(\Omega).

Theorem 1.6.

Let Ωn\Omega\subset\mathbb{R}^{n} be an open set, 1<p<1<p<\infty. Suppose ρp(Ω)<\rho_{p}(\Omega)<\infty and R>ρp(Ω)R>\rho_{p}(\Omega).

  • (i)

    Then,

    (1.7) δR(Ω)RnERoppλ1,p(Ω),\displaystyle\frac{\delta_{R}(\Omega)}{R^{n}\cdot\|E_{R}\|_{op}^{p}}\leq\lambda_{1,p}(\Omega),

    for any bounded, linear extension operator ER:W1,p((0,R)n)W1,p(n)E_{R}:W^{1,p}((0,R)^{n})\longrightarrow W^{1,p}(\mathbb{R}^{n}) satisfying (ERf)|(0,R)n=f(E_{R}f)_{|_{(0,R)^{n}}}=f for fW1,p((0,R)n)f\in W^{1,p}((0,R)^{n}).

  • (ii)

    If 1<p<n1<p<n, then there exist constants C=C(n,p)>0C=C(n,p)>0 and γR(Ω)(0,1)\gamma_{R}(\Omega)\in(0,1) such that

    (1.8) CγR(Ω)Rpλ1,p(Ω).\displaystyle C\gamma_{R}(\Omega)\cdot R^{-p}\leq\lambda_{1,p}(\Omega).

    If pnp\geq n, then (1.8) holds with RpR^{-p} replaced by (1+Rp)1(1+R^{p})^{-1}.

Part (i) of Theorem 1.6 is a direct consequence of a Poincaré-type inequality for any function in W1,p(Ω)W^{1,p}(\Omega) which has a representative that vanishes on a set of positive pp-Sobolev capacity in Ω\Omega, where Ω\Omega is a bounded extension domain. This inequality originates in the work of Meyers [17], and was completed by Adams, see Theorem 8.3.3 and the notes in Section 8.3 on pg. 231 in [1]. For the proof of Theorem 1.6, one simply writes n\mathbb{R}^{n} as a union of closed cubes with mutually disjoint interiors and of side length larger than the strict pp-capacitary inradius of the open set under consideration. Then, one applies this Poincaré-type inequality to each cube. This kind of proof is contained in the works of Maz’ya– Shubin [14] and Souplet [18]. Souplet also splits n\mathbb{R}^{n} into cubes and then uses a Poincaré-type inequality for functions which vanish on a set of positive Lebesgue measure. Maz’ya–Shubin derive a Poincaré-type inequality in [14, Lemma 3.1], similar to the one used in this note, but for balls which forces them take the multiplicity of coverings by balls into account in order to obtain a global estimate.

Theorems 1.3 and 1.6 may be summarized in a qualitative manner as follows.

Corollary 1.9.

Let Ωn\Omega\subset\mathbb{R}^{n} be an open set, 1<p<1<p<\infty. Then,

λ1,p(Ω)>0ρp(Ω)<.\displaystyle\lambda_{1,p}(\Omega)>0\;\;\Leftrightarrow\;\;\rho_{p}(\Omega)<\infty.

It is well-known that the finiteness of the inradius, (Ω)\mathfrak{R}(\Omega), of an open set Ω\Omega, i.e., the supremum of the radii of all balls contained in Ω\Omega, is a necessary condition for (1.1). This can be seen by a simple scaling argument, see, e.g., Souplet’s proof of [18, Prop. 2.1 (i)]. Souplet additionally shows in [18, Prop. 2.1] with p[1,)p\in[1,\infty), that this condition is also sufficient as long as Ω\Omega is a domain which satisfies a uniform exterior cone condition; in the case of p=2p=2 this appears to go back to work of Agmon [2]. In [18], Souplet introduces a measure-theoretic inradius, which yields a sufficient condition for the validity of (1.1) without any regularity assumptions on the boundary of the domain Ω\Omega. Souplet’s formulation of inradius actually inspired the notion of the capacitary inradius in (1.2), originally introduced in [7]. A sufficient condition, similar to Souplet’s, was previously obtain by Lieb in [10, Corollary 2]. We note that the assumption of finiteness of either of these conditions is a stronger assumption than the one of finiteness of the strict pp-capacitary inradius defined in (1.2). The reason for that is that these measure-theoretic inradii do not take into account sets in the complement which are of Lebesgue measure zero but of positive Sobolev pp-capacity. However, for any pair (p,p)(p^{\prime},p) with 1<p<pn1<p^{\prime}<p\leq n, there exists a set EnE\subset\mathbb{R}^{n} of Lebesgue measure zero, such that Cp(E)=0C_{p^{\prime}}(E)=0 while Cp(E)>0C_{p}(E)>0, see [1, Theorem 5.5.1].

A complete description in the flavor of Corollary 1.9 was first given by Maz’ya–Shubin in [14] in the case of p=2p=2 and n3n\geq 3. The authors of [14] use different notions of capacity and of capacitary inradius than presented in this note; see Lemma 2.14 on how their capacitary inradius relates to the one defined in (1.2). For p=2p=2 and n3n\geq 3, estimate (1.4) is an improvement over the upper bounded for λ1,2(Ω)\lambda_{1,2}(\Omega) provided in [14] while (1.8) and the lower bound given in [14] are similar. We also point to the work of Vitolo [20] in which he shows that if p>np>n and Ω\Omega is a domain with finite inradius, (Ω)\mathfrak{R}(\Omega), then λ1,p(Ω)>0\lambda_{1,p}(\Omega)>0. Note that if p>np>n, then singletons have positive pp-Sobolev capacity, so that (Ω)=ρp(Ω)\mathfrak{R}(\Omega)=\rho_{p}(\Omega), i.e., Theorem 1.6 rediscovers Vitolo’s result.

Together with Lebl and Ramachandran, we considered the problem of describing the validity of the Poincaré inequality in the case of n=2n=2, p=2p=2 in potential-theoretic terms in [7]. Originally, we intended to investigate which potential-theoretic conditions yield the L2L^{2}-closed range property of (the weak maximal extension of) the Cauchy–Riemann operator which constitutes an open problem in several complex variables. This closed range property turns out to be equivalent to (1.1) on any open set Ω2\Omega\subset\mathbb{R}^{2}. Moreover, we showed that Corollary 1.9 holds for p=2p=2 and that (1.1) is equivalent to the existence of a smooth, bounded function on Ω\Omega such that its Laplacian has a positive lower bound on Ω\Omega. These results were later shown to hold for p=2p=2, n3n\geq 3 with the Newtonian capacity in place of CpC_{p} in [6] by the author of this note. We show in Lemma 2.11 that the strict pp-capacitary inradius defined in (1.2) does not depend on the choice of pp-capacity as long as the sets of zero pp-capacity are the same as the sets of for which CpC_{p} is zero as well. In particular, the strict capacitary inradii defined in [7, 6] are the same as the one defined in (1.2) for p=2p=2, see the paragraph subsequent to the proof of Lemma 2.11.

This note is structured as follows. The notions of capacity, strict pp-capacitary inradius, and λ1,p\lambda_{1,p} and their basic properties are detailed in Section 2. The proofs of Theorem 1.3 and Theorem 1.6 are given in Section 3 and 4, respectively.

Acknowledgement

I am very grateful to Carlo Morpurgo for his insights he shared with me while completing this project.

2. Preliminaries

2.1. Sobolev pp-capacity

Definition 2.1.

Let KK be a compact set in n\mathbb{R}^{n}, 1<p<1<p<\infty. Then

Cp(K):=inf{up,np+up,np:u𝒮,u1 on K}.\displaystyle C_{p}(K):=\inf\left\{\|u\|_{p,\mathbb{R}^{n}}^{p}+\|\nabla u\|_{p,\mathbb{R}^{n}}^{p}\;:\;u\in\mathcal{S},u\geq 1\text{ on }K\right\}.

This definition may be extended to open sets UnU\subset\mathbb{R}^{n} by setting

(2.2) Cp(U)=sup{Cp(K):KU,K compact}.\displaystyle C_{p}(U)=\sup\{C_{p}(K)\;:\;K\subset U,K\text{ compact}\}.

It then follows that

Cp(K)=inf{Cp(U):KU,U open},\displaystyle C_{p}(K)=\inf\{C_{p}(U)\;:\;K\subset U,U\text{ open}\},

which can be proven analogously to [1, Proposition 2.2.3]. The definition of CpC_{p} may now be extended to arbitrary sets by setting

(2.3) Cp(E):=inf{Cp(U):EU,U open}\displaystyle C_{p}(E):=\inf\{C_{p}(U)\,:\,E\subset U,U\text{ open}\}

for EnE\subset\mathbb{R}^{n}. A set EnE\subset\mathbb{R}^{n} is called pp-polar, if Cp(E)=0C_{p}(E)=0. Moreover, two functions are said to equal pp-quasi everywhere if they equal outside a pp-polar set.

Next, we present some standard properties of CpC_{p}.

Lemma 2.4.

Let EnE\subset\mathbb{R}^{n}. Then

  • (i)

    If EEE^{\prime}\subset E, then Cp(E)Cp(E)C_{p}(E^{\prime})\leq C_{p}(E).

  • (ii)

    If xnx\in\mathbb{R}^{n}, then Cp(E+x)=Cp(E)C_{p}(E+x)=C_{p}(E).

  • (iii)

    If s>0s>0, then Cp(sE)snmax{1,sp}Cp(E)C_{p}(sE)\leq s^{n}\max\{1,s^{-p}\}C_{p}(E).

  • (iv)

    If {Ei}in\{E_{i}\}_{i\in\mathbb{N}}\subset\mathbb{R}^{n} such that E=i=1EiE=\bigcup_{i=1}^{\infty}E_{i}, then Cp(E)i=1Cp(Ei).C_{p}(E)\leq\sum_{i=1}^{\infty}C_{p}(E_{i}).

  • (v)

    Any Borel set EE is capacitable, i.e.,

    Cp(E)=sup{Cp(K):KE,K compact}=inf{Cp(U):EU,U open}.C_{p}(E)=\sup\{C_{p}(K)\,:\,K\subset E,K\text{ compact}\}=\inf\{C_{p}(U)\,:\,E\subset U,U\text{ open}\}.
Proof.

The proofs of (i)–(iv) for arbitrary sets follow from (2.2) and (2.3) once (i)–(iii) have been established for compact sets. For compact sets, (i), (ii) and (iv) follow directly from Definition 2.1 while (iii) follows from a change of variable argument yielding

Cp(sK)=inf{snup,Kp+snspup,Kp:u1 on K,u𝒮}.\displaystyle C_{p}(sK)=\inf\left\{s^{n}\|u\|_{p,K}^{p}+\frac{s^{n}}{s^{p}}\|\nabla u\|_{p,K}^{p}\;:\;u\geq 1\text{ on }K,u\in\mathcal{S}\right\}.

For the proof of (v), see Propositions 2.3.12 and 2.3.13 as well as Theorem 2.3.11 and the succeeding remark in [1]. ∎

2.2. The strict pp-capacitary inradius

In a slight deviation from (1.2), we define the strict pp-capacitary inradius as follows.

Definition 2.5.

Let Ωn\Omega\subset\mathbb{R}^{n} be an open set, 1<p<1<p<\infty. Then, the strict pp-capacitary inradius of Ω\Omega is defined as

(2.6) ρp(Ω)=sup{r>0:ϵ>0xn such that Cp(𝔹r(x)Ωc)<ϵ}.\displaystyle\rho_{p}(\Omega)=\sup\left\{r>0\,:\,\forall\;\epsilon>0\;\exists\;x\in\mathbb{R}^{n}\text{ such that }C_{p}(\mathbb{B}_{r}(x)\cap\Omega^{c})<\epsilon\right\}.

We show first that this definition of ρp\rho_{p} agrees with (1.2), although CpC_{p} is not invariant under taking closures.

Lemma 2.7.

Let Ωn\Omega\subset\mathbb{R}^{n} be an open set, 1<p<1<p<\infty. Then

(2.8) ρp(Ω)=sup{r>0:ϵ>0xn such that Cp(𝔹r(x)¯Ωc)<ϵ}.\displaystyle\rho_{p}(\Omega)=\sup\left\{r>0\,:\,\forall\;\epsilon>0\;\exists\;x\in\mathbb{R}^{n}\text{ such that }C_{p}(\overline{\mathbb{B}_{r}(x)}\cap\Omega^{c})<\epsilon\right\}.
Proof.

Let us denote the right hand side of (2.8) by ρ^p(Ω)\hat{\rho}_{p}(\Omega). By the monotonicity of the Sobolev pp-capacity, it is immediate that ρ^p(Ω)ρp(Ω)\hat{\rho}_{p}(\Omega)\leq\rho_{p}(\Omega). Now, suppose that 0<R<ρp(Ω)0<R<\rho_{p}(\Omega). Then, for all ϵ>0\epsilon>0 there exists an xnx\in\mathbb{R}^{n} such that Cp(𝔹R(x)Ωc)<ϵC_{p}(\mathbb{B}_{R}(x)\cap\Omega^{c})<\epsilon, and hence

Cp(𝔹Rδ(x)¯Ωc)<ϵfor allδ(0,R).C_{p}(\overline{\mathbb{B}_{R-\delta}(x)}\cap\Omega^{c})<\epsilon\hskip 5.69046pt\;\;\text{for all}\;\delta\in(0,R).

Therefore, Rδ<ρ^p(Ω)R-\delta<\hat{\rho}_{p}(\Omega) for all δ(0,R)\delta\in(0,R), which implies ρp(Ω)ρ^p(Ω)\rho_{p}(\Omega)\leq\hat{\rho}_{p}(\Omega). ∎

In the following, we collect some basic properties of the strict pp-capacitary inradius. To do so, we recall that the inradius, (Ω)\mathfrak{R}(\Omega), of an open set Ωn\Omega\subset\mathbb{R}^{n} is defined as

(Ω)=sup{r>0:xn such that 𝔹r(x)Ω}.\mathfrak{R}(\Omega)=\sup\{r>0\,:\,\exists\;x\in\mathbb{R}^{n}\text{ such that }\mathbb{B}_{r}(x)\subset\Omega\}.

We also define the pp-capacitary inradius, 𝔯p(Ω)\mathfrak{r}_{p}(\Omega), by

𝔯p(Ω)=sup{r>0:xn such that Cp(𝔹r(x)Ωc)=0}.\mathfrak{r}_{p}(\Omega)=\sup\left\{r>0\,:\,\exists\;x\in\mathbb{R}^{n}\text{ such that }C_{p}(\mathbb{B}_{r}(x)\cap\Omega^{c})=0\right\}.
Lemma 2.9.

Let Ωn\Omega\subset\mathbb{R}^{n} be an open set, 1<p<1<p<\infty.

  • (i)

    If xnx\in\mathbb{R}^{n} and s>0s>0, then ρp(Ω+x)=ρp(Ω)\rho_{p}(\Omega+x)=\rho_{p}(\Omega) and ρp(sΩ)=sρp(Ω)\rho_{p}(s\Omega)=s\rho_{p}(\Omega).

  • (ii)

    If ΩΩ\Omega\subset\Omega^{\prime} is an open set, then ρp(Ω)ρp(Ω)\rho_{p}(\Omega)\leq\rho_{p}(\Omega^{\prime}). If additionally, ΩΩ\Omega^{\prime}\setminus\Omega is pp-polar, then ρp(Ω)=ρp(Ω)\rho_{p}(\Omega)=\rho_{p}(\Omega^{\prime}).

  • (iii)

    ρp(Ω)𝔯p(Ω)(Ω)\rho_{p}(\Omega)\geq\mathfrak{r}_{p}(\Omega)\geq\mathfrak{R}(\Omega), and equality holds if p>np>n.

  • (iv)

    If Ω\Omega is bounded, then ρp(Ω)=𝔯p(Ω)\rho_{p}(\Omega)=\mathfrak{r}_{p}(\Omega). Moreover, there exists an xnx^{\circ}\in\mathbb{R}^{n}, such that

    Cp(𝔹ρp(Ω)(x)Ωc)=0,C_{p}(\mathbb{B}_{\rho_{p}(\Omega)}(x^{\circ})\cap\Omega^{c})=0,

    i.e., 𝔯p(Ω)\mathfrak{r}_{p}(\Omega) is attained.

  • (v)

    ρp(Ω)=limRρp(Ω𝔹R(0))\rho_{p}(\Omega)=\lim_{R\to\infty}\rho_{p}(\Omega\cap\mathbb{B}_{R}(0)).

Proof.

The translation invariance of ρp\rho_{p} holds because it holds for CpC_{p}, see (ii) of Lemma 2.4. To check the linearity under dilations we first note that

𝔹r(x)Ωc=1s(𝔹sr(xs)(sΩ)c).\mathbb{B}_{r}(x)\cap\Omega^{c}=\frac{1}{s}\left(\mathbb{B}_{sr}(xs)\cap(s\Omega)^{c}\right).

Thus, if for a given r>0r>0 and ϵ>0\epsilon>0 there exists an xnx\in\mathbb{R}^{n} such that

Cp(𝔹sr(xs)(sΩ)c)<ϵ,C_{p}\left(\mathbb{B}_{sr}(xs)\cap(s\Omega)^{c}\right)<\epsilon,

then

Cp(𝔹r(x)Ωc)snmax{1,sp}ϵ,C_{p}(\mathbb{B}_{r}(x)\cap\Omega^{c})\leq s^{-n}\max\{1,s^{p}\}\epsilon,

by (iii) of Lemma 2.4. It then follows that r<ρp(Ω)r<\rho_{p}(\Omega) whenever sr<ρp(sΩ)sr<\rho_{p}(s\Omega), and hence, sρp(Ω)ρp(sΩ)s\rho_{p}(\Omega)\leq\rho_{p}(s\Omega) for any s>0s>0. We now may repeat this argument with t=1st=\frac{1}{s} in place of ss and t1Ωt^{-1}\Omega in place of Ω\Omega to obtain

tρp(t1Ω)ρp(Ω)s1ρp(sΩ)ρp(Ω),\displaystyle t\rho_{p}(t^{-1}\Omega)\leq\rho_{p}(\Omega)\Rightarrow s^{-1}\rho_{p}(s\Omega)\leq\rho_{p}(\Omega),

which yields

sρp(Ω)ρp(sΩ)sρp(Ω),s\rho_{p}(\Omega)\leq\rho_{p}(s\Omega)\leq s\rho_{p}(\Omega),

hence, the proof of (i) is complete.

The first part of (ii) follows from the definition. The second part follows after observing

Cp(𝔹r(x)Ωc)Cp(𝔹r(x)(Ω)c)C_{p}(\mathbb{B}_{r}(x)\cap\Omega^{c})\leq C_{p}(\mathbb{B}_{r}(x)\cap(\Omega^{\prime})^{c})

by (iv) of Lemma 2.4 and the fact that 𝔹r(x)(ΩΩ)\mathbb{B}_{r}(x)\cap(\Omega^{\prime}\setminus\Omega) is pp-polar.

The set of inequalities in (iii) follows directly from the definitions of the inradii. Equality holds if p>np>n, because Cp({x})>0C_{p}(\{x\})>0 for all xnx\in\mathbb{R}^{n}. To wit, if R>(Ω)R>\mathfrak{R}(\Omega), then 𝔹R(x)Ωc\mathbb{B}_{R}(x)\cap\Omega^{c} is non-empty for all xnx\in\mathbb{R}^{n} so that Cp(𝔹R(x)Ωc)>Cp({0})C_{p}(\mathbb{B}_{R}(x)\cap\Omega^{c})>C_{p}(\{0\}) for all xnx\in\mathbb{R}^{n}. Hence, R>ρp(Ω)R>\rho_{p}(\Omega), so that (Ω)=ρp(Ω)\mathfrak{R}(\Omega)=\rho_{p}(\Omega) follows.

For the proof of (iv), suppose Ω\Omega is bounded. Then, by definition of ρp(Ω)\rho_{p}(\Omega), there exists a sequence {(rj,xj)}j\{(r_{j},x_{j})\}_{j\in\mathbb{N}} in 0+×n\mathbb{R}^{+}_{0}\times\mathbb{R}^{n} such that {rj}j\{r_{j}\}_{j\in\mathbb{N}} is an increasing sequence which converges to ρp(Ω)\rho_{p}(\Omega), and

Cp(𝔹rj(xj)Ωc)<1j.\displaystyle C_{p}(\mathbb{B}_{r_{j}}(x_{j})\cap\Omega^{c})<\frac{1}{j}.

Since Ω\Omega is a bounded set, it follows that {xj}\{x_{j}\} is a bounded sequence, thus, has a convergent subsequence. For ease of notation, let us denote the subsequence by {xj}j\{x_{j}\}_{j\in\mathbb{N}}. Write xx^{\circ} for the limit point. It suffices to prove that

(2.10) Cp(𝔹ρp(Ω)(x)Ωc)=0\displaystyle C_{p}\left(\mathbb{B}_{\rho_{p}(\Omega)}(x^{\circ})\cap\Omega^{c}\right)=0

holds. To prove (2.10), let δ(0,ρp(Ω))\delta\in(0,\rho_{p}(\Omega)) and choose j0j_{0}\in\mathbb{N} such that |xxj|<δ2|x^{\circ}-x_{j}|<\frac{\delta}{2} for all jj0j\geq j_{0}. Then, choose j1j0j_{1}\geq j_{0} such that ρp(Ω)<rj+δ2\rho_{p}(\Omega)<r_{j}+\frac{\delta}{2} for all jj1j\geq j_{1}. It follows that

𝔹ρp(Ω)δ(x)¯𝔹rj(xj)for alljj1,\displaystyle\overline{\mathbb{B}_{\rho_{p}(\Omega)-\delta}(x^{\circ})}\subset\mathbb{B}_{r_{j}}(x_{j})\;\;\hskip 5.69046pt\text{for all}\;j\geq j_{1},

and, therefore,

Cp(𝔹ρp(Ω)δ(x)¯Ωc)Cp(𝔹rj(xj)Ωc)<1jfor alljj1.\displaystyle C_{p}\left(\overline{\mathbb{B}_{\rho_{p}(\Omega)-\delta}(x^{\circ})}\cap\Omega^{c}\right)\leq C_{p}\left(\mathbb{B}_{r_{j}}(x_{j})\cap\Omega^{c}\right)<\frac{1}{j}\;\;\hskip 5.69046pt\text{for all}\;j\geq j_{1}.

Letting jj\to\infty then yields

Cp(𝔹ρp(Ω)δ(x)¯Ωc)=0for allδ(0,ρp(Ω)),\displaystyle C_{p}\left(\overline{\mathbb{B}_{\rho_{p}(\Omega)-\delta}(x^{\circ})}\cap\Omega^{c}\right)=0\;\;\hskip 5.69046pt\text{for all}\;\delta\in(0,\rho_{p}(\Omega)),

and, hence, by part (v) of Lemma 2.4, the claimed (2.10) follows.

Part (v) follows directly from the monotonicity property in (ii). ∎

We now can prove that the strict pp-capacitary inradius does not depend on the choice of pp-capacity.

Lemma 2.11.

Let Ωn\Omega\subset\mathbb{R}^{n} be open, 1<p<1<p<\infty. Let Γp:𝒫(n)0+{}\Gamma_{p}:\mathcal{P}(\mathbb{R}^{n})\longrightarrow\mathbb{R}_{0}^{+}\cup\{\infty\} be such that

  • (a)

    Γp()=0\Gamma_{p}(\emptyset)=0,

  • (b)

    EEE\subset E^{\prime} \Rightarrow Γp(E)Γp(E)\Gamma_{p}(E)\leq\Gamma_{p}(E^{\prime}),

  • (c)

    all Borel sets are capacitable with respect to Γp\Gamma_{p}.

Suppose Γp(E)=0\Gamma_{p}(E)=0 iff Cp(E)=0C_{p}(E)=0 for all bounded Borel sets EnE\subset\mathbb{R}^{n}. Then

(2.12) ρp(Ω)=sup{r>0:ϵ>0xn such that Γp(𝔹r(x)Ωc)<ϵ}.\displaystyle\rho_{p}(\Omega)=\sup\left\{r>0\,:\,\forall\;\epsilon>0\;\exists\;x\in\mathbb{R}^{n}\text{ such that }\Gamma_{p}(\mathbb{B}_{r}(x)\cap\Omega^{c})<\epsilon\right\}.
Proof.

Note first that properties (a)–(c) ensure that (iv) and (v) of Lemma 2.9 hold for the capacitary inradius, ρpΓ(Ω)\rho_{p}^{\Gamma}(\Omega), defined by the right hand side of (2.12).

Next, write ΩR\Omega_{R} for Ω𝔹R(0)\Omega\cap\mathbb{B}_{R}(0) for R>0R>0. Then, by (v) and (iv) of Lemma 2.9, it follows that

ρp(Ω)=limRρp(ΩR)\displaystyle\rho_{p}(\Omega)=\lim_{R\to\infty}\rho_{p}(\Omega_{R}) =𝔯p(ΩR)\displaystyle=\mathfrak{r}_{p}(\Omega_{R})
(2.13) =sup{r>0:xn such that Γp(𝔹r(x)(ΩR)c)=0},\displaystyle=\sup\left\{r>0\,:\,\exists\;x\in\mathbb{R}^{n}\text{ such that }\Gamma_{p}(\mathbb{B}_{r}(x)\cap(\Omega_{R})^{c})=0\right\},

where the last step follows from the assumption that Γp(E)=0\Gamma_{p}(E)=0 iff Cp(E)=0C_{p}(E)=0 for all bounded Borel set EnE\subset\mathbb{R}^{n}. Since (iv)-(v) of Lemma 2.9 hold for the (strict) capacitary inradius with respect to Γp\Gamma_{p}, it follows

sup{r>0:xn such that Γp(𝔹r(x)(ΩR)c)=0}=limRρpΓ(ΩR)=ρpΓ(Ω),\displaystyle\sup\left\{r>0\,:\,\exists\;x\in\mathbb{R}^{n}\text{ such that }\Gamma_{p}(\mathbb{B}_{r}(x)\cap(\Omega_{R})^{c})=0\right\}=\lim_{R\to\infty}\rho_{p}^{\Gamma}(\Omega_{R})=\rho_{p}^{\Gamma}(\Omega),

i.e., ρp\rho_{p} is invariant under the choice of Γp\Gamma_{p}. ∎

Both the logarithmic capacity for n=2n=2 and the Newtonian capacity for n3n\geq 3 satisfy (a)–(c) of Lemma 2.11. Moreover, their bounded polar Borel sets are equal to the bounded Borel sets which are polar with respect to C2C_{2} by Theorem 1 (m=1m=1, p=2p=2) in [13], see also Theorems A and B (α=0\alpha=0, m=2m=2) in [21]. As a consequence, the strict capacitary inradius ρ2\rho_{2} with respect to C2C_{2} is the same as the one defined in [7] with respect to the logarithmic capacity for n=2n=2 and the one defined in [6] with respect to the Newtonian capacity for n3n\geq 3.

Maz’ya and Shubin used a different notion of inradius, formulated in terms of the Wiener capacity, in their work [14], for p=2p=2 and n3n\geq 3. To wit, they defined the interior capacitary radius, rΩ,γr_{\Omega,\gamma}, of an open set Ωn\Omega\subset\mathbb{R}^{n} for γ(0,1)\gamma\in(0,1) by

rΩ,γ=sup{r>0:xn such that C2(𝔹r(x)¯Ω)γC2(𝔹r(0)¯)},r_{\Omega,\gamma}=\sup\left\{r>0\;:\;\exists\;x\in\mathbb{R}^{n}\text{ such that }C_{2}^{\prime}\left(\overline{\mathbb{B}_{r}(x)}\setminus\Omega\right)\leq\gamma C_{2}^{\prime}\left(\overline{\mathbb{B}_{r}(0)}\right)\right\},

where the Wiener capacity CpC_{p}^{\prime}, 1<p<1<p<\infty, is defined by

Cp(K)=inf{up,np:u1 on K,u𝒮}C_{p}^{\prime}(K)=\inf\left\{\|\nabla u\|_{p,\mathbb{R}^{n}}^{p}\;:\;u\geq 1\text{ on }K,u\in\mathcal{S}\right\}

for KnK\subset\mathbb{R}^{n} compact.

This interior capacitary radius relates to ρ2\rho_{2}, for n3n\geq 3, as follows.

Lemma 2.14.

Let Ωn\Omega\subset\mathbb{R}^{n}, n3n\geq 3, be an open set. Then ρ2(Ω)=inf{rΩ,γ:γ(0,1)}\rho_{2}(\Omega)=\inf\{r_{\Omega,\gamma}:\gamma\in(0,1)\}.

We note that such a relation was first observed by Carlo Morpurgo, see also (3.7) in [5].

Proof.

Note first that the Wiener capacity is equivalent to the Newtonian capacity, see, e.g., pg. 4 in [14] for a sketch of the proof. That means in particular that the Wiener capacity satisfies the hypotheses of Lemma 2.11, and hence, it suffices to show that ρ2(Ω)=inf{rΩ,γ:γ(0,1)}\rho_{2}^{\prime}(\Omega)=\inf\{r_{\Omega,\gamma}:\gamma\in(0,1)\}, where

ρ2(Ω)=sup{r>0:ϵ>0xn such that C2(𝔹r(x)Ωc)<ϵ}.\rho_{2}^{\prime}(\Omega)=\sup\{r>0\,:\,\forall\,\epsilon>0\,\exists\;x\in\mathbb{R}^{n}\text{ such that }C_{2}^{\prime}(\mathbb{B}_{r}(x)\cap\Omega^{c})<\epsilon\}.

Moreover, the arguments supplied in the proof of (2.8) let us work with this alternative formulation for ρ2(Ω)\rho_{2}^{\prime}(\Omega):

ρ2(Ω)=sup{r>0:ϵ>0xn such that C2(𝔹r(x)¯Ωc)<ϵ}.\rho_{2}^{\prime}(\Omega)=\sup\{r>0\,:\,\forall\,\epsilon>0\,\exists\;x\in\mathbb{R}^{n}\text{ such that }C_{2}^{\prime}(\overline{\mathbb{B}_{r}(x)}\cap\Omega^{c})<\epsilon\}.

Now, let γ(0,1)\gamma\in(0,1) be given. Let R<ρ2(Ω)R<\rho_{2}^{\prime}(\Omega), and choose ϵ=γC2(𝔹R¯(0))\epsilon=\gamma C_{2}^{\prime}(\overline{\mathbb{B}_{R}}(0)). Then, by definition of ρ2(Ω)\rho_{2}^{\prime}(\Omega), there exists an xnx\in\mathbb{R}^{n} such that

C2(𝔹R(x)¯Ωc)<ϵ=γC2(𝔹R¯(0)).C_{2}^{\prime}(\overline{\mathbb{B}_{R}(x)}\cap\Omega^{c})<\epsilon=\gamma C_{2}^{\prime}(\overline{\mathbb{B}_{R}}(0)).

Thus, R<rΩ,γR<r_{\Omega,\gamma} for any given γ(0,1)\gamma\in(0,1), and hence, ρ2(Ω)inf{rΩ,γ:γ(0,1)}\rho_{2}^{\prime}(\Omega)\leq\inf\{r_{\Omega,\gamma}:\gamma\in(0,1)\}.

Next, let R>ρ2(Ω)R>\rho_{2}^{\prime}(\Omega). Then, there exists an ϵ>0\epsilon>0 such that C2(𝔹R(x)¯Ωc)ϵC_{2}^{\prime}(\overline{\mathbb{B}_{R}(x)}\cap\Omega^{c})\geq\epsilon for all xnx\in\mathbb{R}^{n}. Let γ0=min{1,ϵ/C2(𝔹R¯(0))}\gamma_{0}=\min\{1,\epsilon/C_{2}^{\prime}(\overline{\mathbb{B}_{R}}(0))\}. It then follows that

C2(𝔹R(x)¯Ωc)>γC2(𝔹R¯(0))for allγ(0,γ0)C_{2}^{\prime}(\overline{\mathbb{B}_{R}(x)}\cap\Omega^{c})>\gamma C_{2}^{\prime}(\overline{\mathbb{B}_{R}}(0))\hskip 5.69046pt\;\;\text{for all}\;\gamma\in(0,\gamma_{0})

for all xnx\in\mathbb{R}^{n}. That is, R>rΩ,γR>r_{\Omega,\gamma} for all γ(0,γ0)\gamma\in(0,\gamma_{0}). Since rΩ,γr_{\Omega,\gamma} is increasing in γ\gamma, we obtain that R>inf{rΩ,γ:γ(0,1)}R>\inf\{r_{\Omega,\gamma}:\gamma\in(0,1)\}. Hence, ρ2(Ω)inf{rΩ,γ:γ(0,1)}\rho_{2}^{\prime}(\Omega)\geq\inf\{r_{\Omega,\gamma}:\gamma\in(0,1)\}. ∎

Understanding the pp-capacitary inradius as limit of Maz’ya–Shubin-like inradii as in Lemma 2.14 comes in handy in the proof of part (ii) of Theorem 1.6. We define

rp,γ(Ω):=sup{r>0:xn such that Cp(𝔹r(x)Ωc)γCp(𝔹r(0))}r_{p,\gamma}(\Omega):=\sup\{r>0\;:\;\exists\,x\in\mathbb{R}^{n}\text{ such that }C_{p}(\mathbb{B}_{r}(x)\cap\Omega^{c})\leq\gamma C_{p}(\mathbb{B}_{r}(0))\}

for γ(0,1)\gamma\in(0,1) and 1<p<1<p<\infty. A proof similar to the one given in Lemma 2.14 then yields the following.

Corollary 2.15.

Let Ωn\Omega\subset\mathbb{R}^{n} and 1<p<1<p<\infty. Then ρp(Ω)=inf{rp,γ(Ω):γ(0,1)}\rho_{p}(\Omega)=\inf\{r_{p,\gamma}(\Omega)\;:\;\gamma\in(0,1)\}.

2.3. Infimum of the nonlinear Rayleigh quotient for pp-Laplacian

Definition 2.16.

Let Ωn\Omega\subset\mathbb{R}^{n} be an open set, 1<p<1<p<\infty. Then

λ1,p(Ω):=inf{up,Ωpup,Ω:u𝒞c(Ω){0}}.\lambda_{1,p}(\Omega):=\inf\left\{\frac{\|\nabla u\|_{p,\Omega}^{p}}{\|u\|_{p,\Omega}}\;:\;u\in\mathcal{C}_{c}(\Omega)\setminus\{0\}\right\}.

We collect some elementary properties of λ1,p\lambda_{1,p} in the following lemma.

Lemma 2.17.

Let Ωn\Omega\subset\mathbb{R}^{n} be an open set. Then

  • (i)

    If ΩΩ\Omega^{\prime}\subset\Omega is an open set, then λ1,p(Ω)λ1,p(Ω).\lambda_{1,p}(\Omega)\leq\lambda_{1,p}(\Omega^{\prime}).

  • (ii)

    If xnx\in\mathbb{R}^{n}, then λ1,p(Ω+x)=λ1,p(Ω).\lambda_{1,p}(\Omega+x)=\lambda_{1,p}(\Omega).

  • (iii)

    If s>0s>0, then

    λ1,p(sΩ)=λ1,p(Ω)sp.\lambda_{1,p}(s\Omega)=\lambda_{1,p}(\Omega)s^{-p}.
  • (iv)

    If ΩΩ\Omega^{\prime}\subset\Omega is an open set such that ΩΩ\Omega\setminus\Omega^{\prime} is pp-polar, then λ1,p(Ω)=λ1,p(Ω).\lambda_{1,p}(\Omega)=\lambda_{1,p}(\Omega^{\prime}).

Proof.

Parts (i)–(iii) follow from the definition, the translation invariance of the Lebesgue measure, and a change of variable argument, respectively. For part (iv), we note that if ΩΩ\Omega\setminus\Omega^{\prime} is pp-polar, then W01,p(Ω)=W01,p(Ω)W_{0}^{1,p}(\Omega)=W_{0}^{1,p}(\Omega^{\prime}), see, for instance, [9, Theorem 2.43]. Hence, by definition, λ1,p(Ω)=λ1,p(Ω)\lambda_{1,p}(\Omega)=\lambda_{1,p}(\Omega^{\prime}). ∎

If λ1,p(Ω)\lambda_{1,p}(\Omega) is positive and attained at some uW01,p(Ω)u\in W_{0}^{1,p}(\Omega), then uu is a weak solution to

(2.18) {div(|u|p2u)+λ|u|p2u=0 in Ωu=0 on bΩ\begin{cases}\operatorname{div}\left(|\nabla u|^{p-2}\nabla u\right)+\lambda|u|^{p-2}u=0&\text{ in }\Omega\\ u=0&\text{ on }b\Omega\end{cases}

for λ=λ1,p(Ω)\lambda=\lambda_{1,p}(\Omega). Such a function uu is called a first eigenfunction of Δp\Delta_{p}, where

Δpu:=div(|u|p2u).\Delta_{p}u:=\operatorname{div}(|\nabla u|^{p-2}\nabla u).

Here, uu being a weak solution to (2.18) means that

(2.19) Ω|u|p2uχdm=λΩ|u|p2uχ𝑑m\displaystyle\int_{\Omega}|\nabla u|^{p-2}\nabla u\,{\scriptstyle\circ}\,\nabla\chi\;dm=\lambda\int_{\Omega}|u|^{p-2}u\chi\;dm

for all χ𝒞c(Ω)\chi\in\mathcal{C}_{c}^{\infty}(\Omega).

It is well-known that if Ωn\Omega\Subset\mathbb{R}^{n} is a smoothly bounded domain, then λ1,p(Ω)\lambda_{1,p}(\Omega) is positive and attained at some non-zero uW01,p(Ω)𝒞(Ω¯)u\in W^{1,p}_{0}(\Omega)\cap\mathcal{C}(\overline{\Omega}). Moreover, such uu may be assumed to be positive in Ω\Omega. A comprehensive resource for standard results on the first eigenvalue and eigenfunctions for Δp\Delta_{p} are the lecture notes by Lindqvist [12]. The regularity result for the first eigenfunctions is due to Gariepy–Ziemer in [8] for 1<pn1<p\leq n; in the case of p>np>n, it is known that, under these conditions on Ω\Omega, a representative of uW1,p(Ω)u\in W^{1,p}(\Omega) is in 𝒞(Ω¯)\mathcal{C}(\overline{\Omega}), see, e.g., Theorem 5 in §5.6.2 in [4].

3. Proof of Theorem 1.3

The following lemma is crucial for the proof of Theorem 1.3.

Lemma 3.1.

Let DnD\Subset\mathbb{R}^{n} be a smoothly bounded domain. Let KjD¯K_{j}\subset\overline{D}, jj\in\mathbb{N}, be compact sets. Suppose that limjCp(Kj)=0\lim_{j\to\infty}C_{p}(K_{j})=0 and Dj:=DKjD_{j}:=D\setminus K_{j} is an open set with smooth boundary for all jj\in\mathbb{N}. Then, limjλ1,p(Dj)=λ1,p(D)\lim_{j\to\infty}\lambda_{1,p}(D_{j})=\lambda_{1,p}(D).

Proof.

Note first that if p>np>n, then limjCp(Kj)=0\lim_{j\to\infty}C_{p}(K_{j})=0 implies that KjK_{j} is empty for all jj sufficiently large, and hence, the conclusion holds trivially.

Since DD is a bounded domain, λ1,p(D)>0\lambda_{1,p}(D)>0, and there exists a weak solution, φW01,p(D)𝒞(D¯)\varphi\in W_{0}^{1,p}(D)\cap\mathcal{C}(\overline{D}), to (2.18) which is positive on DD, see Section 2.3 for references. After rescaling, we may further assume that 0<φ10<\varphi\leq 1 on DD. Next, for each jj\in\mathbb{N}, there exists a weak solution hjW1,p(Dj)h_{j}\in W^{1,p}(D_{j}) to the Dirichlet problem for Δp\Delta_{p} with boundary data φ\varphi on DjD_{j}, i.e.,

{Δphj=0 in Djhj=φ on bDj,\begin{cases}\Delta_{p}h_{j}=0&\text{ in }D_{j}\\ h_{j}=\varphi&\text{ on }bD_{j}\end{cases},

see, e.g., [11, Th. 2.16]. It follows from work by Maz’ya [15], see also [11, Th. 2.16], that hj𝒞(Dj¯)h_{j}\in\mathcal{C}(\overline{D_{j}}), and hence hj=0h_{j}=0 on bDjbDbD_{j}\cap bD and 0<hj10<h_{j}\leq 1 on bDjbDbD_{j}\setminus bD. Finally, set ψj=φhj\psi_{j}=\varphi-h_{j} on DjD_{j}. Then, ψjW01,p(Dj)𝒞(D¯j)\psi_{j}\in W_{0}^{1,p}(D_{j})\cap\mathcal{C}(\overline{D}_{j}) and Δpψj=Δpφ\Delta_{p}\psi_{j}=\Delta_{p}\varphi holds weakly on DjD_{j}. Using (2.19), we compute

ψjp,Djp\displaystyle\left\|\nabla\psi_{j}\right\|_{p,D_{j}}^{p} =Dj|ψj|p2ψjψjdm\displaystyle=\int_{D_{j}}\left|\nabla\psi_{j}\right|^{p-2}\nabla\psi_{j}\,{\scriptstyle\circ}\,\nabla\psi_{j}\;dm
=Dj|φ|p2φψjdm\displaystyle=\int_{D_{j}}\left|\nabla\varphi\right|^{p-2}\nabla\varphi\,{\scriptstyle\circ}\,\nabla\psi_{j}\;dm
=λ1,p(D)Dj|φ|p2φψj𝑑mλ1,p(D)φp,Djp1ψjp,Dj,\displaystyle=\lambda_{1,p}(D)\int_{D_{j}}\left|\varphi\right|^{p-2}\varphi\cdot\psi_{j}\;dm\leq\lambda_{1,p}(D)\;\left\|\varphi\right\|_{p,D_{j}}^{p-1}\cdot\left\|\psi_{j}\right\|_{p,D_{j}},

where the last step follows from Hölder inequality. Using this estimate, in conjunction with the definition of λ1,p(Dj)\lambda_{1,p}(D_{j}), yields

1λ1,p(Dj)ψjp,Djpψjp,Djpψjp,Djp1λ1,p(D)φp,Djp1,\displaystyle\frac{1}{\lambda_{1,p}(D_{j})}\geq\frac{\|\psi_{j}\|_{p,D_{j}}^{p}}{\left\|\nabla\psi_{j}\right\|_{p,D_{j}}^{p}}\geq\frac{\|\psi_{j}\|_{p,D_{j}}^{p-1}}{\lambda_{1,p}(D)\cdot\|\varphi\|_{p,D_{j}}^{p-1}},

and therefore,

(λ1,p(D)λ1,p(Dj))1p1ψjp,Djφp,Djφp,Djhjp,Djφp,Dj1hjp,Djφp,Dj.\displaystyle\left(\frac{\lambda_{1,p}(D)}{\lambda_{1,p}(D_{j})}\right)^{\frac{1}{p-1}}\geq\frac{\|\psi_{j}\|_{p,D_{j}}}{\|\varphi\|_{p,D_{j}}}\geq\frac{\|\varphi\|_{p,D_{j}}-\|h_{j}\|_{p,D_{j}}}{\|\varphi\|_{p,D_{j}}}\geq 1-\frac{\|h_{j}\|_{p,D_{j}}}{\|\varphi\|_{p,D_{j}}}.

Since φ\varphi is bounded on DD and limjCp(DDj)=0\lim_{j\to\infty}C_{p}(D\setminus D_{j})=0, so that limjm(DDj)=0\lim_{j\to\infty}m(D\setminus D_{j})=0 as well, it follows from the Dominated Convergence Theorem that

limjφp,Dj=φp,D.\lim_{j\to\infty}\|\varphi\|_{p,D_{j}}=\|\varphi\|_{p,D}.

Hence, it suffices to show that limjhjp,Dj=0\lim_{j\to\infty}\|h_{j}\|_{p,D_{j}}=0.

For that, define

Uj={u𝒮:u1 near Kj, 0u1}\displaystyle U_{j}=\{u\in\mathcal{S}:u\equiv 1\text{ near }K_{j},\;0\leq u\leq 1\}

for each jj\in\mathbb{N}. Temporarily fix jj\in\mathbb{N} and uUju\in U_{j}. Then, we may find a gjW1,p(Dj)𝒞(Dj¯)g_{j}\in W^{1,p}(D_{j})\cap\mathcal{C}(\overline{D_{j}}) which is a weak solution to

(3.2) {Δpgj=0 in Djgj=u on bDj.\begin{cases}\Delta_{p}g_{j}=0&\text{ in }D_{j}\\ g_{j}=u&\text{ on }bD_{j}\end{cases}.

Note that hjgjh_{j}\leq g_{j} holds on DjD_{j} by the comparison principle for pp-harmonic functions, see, e.g., [11, Th. 2.15], and the fact that φu\varphi\leq u on bDjbD_{j}. Thence,

(3.3) hjp,Djgjp,Djgjup,Dj+up,Dj.\displaystyle\|h_{j}\|_{p,D_{j}}\leq\|g_{j}\|_{p,D_{j}}\leq\|g_{j}-u\|_{p,D_{j}}+\|u\|_{p,D_{j}}.

By (3.2), gjuW01,p(Dj)g_{j}-u\in W_{0}^{1,p}(D_{j}). Thus, since DjDD_{j}\subset D, the pp-Poincaré inequality is applicable to the first term on the right hand side of (3.3) with a uniform constant, i.e.,

gjup,DjCgjup,Dj\displaystyle\|g_{j}-u\|_{p,D_{j}}\leq C\|\nabla g_{j}-\nabla u\|_{p,D_{j}}

with C=(λ1,p(D))1/pC=\left(\lambda_{1,p}(D)\right)^{-1/p}. Therefore, we have

hjp,DjCgjp,Dj+Cup,Dj+up,Dj.\displaystyle\|h_{j}\|_{p,D_{j}}\leq C\|\nabla g_{j}\|_{p,D_{j}}+C\|\nabla u\|_{p,D_{j}}+\|u\|_{p,D_{j}}.

Next, note that (3.2) implies that gjg_{j} is a quasi-minimizer, i.e.,

gjp,Djfp,Dj\displaystyle\|\nabla g_{j}\|_{p,D_{j}}\leq\|\nabla f\|_{p,D_{j}}

for any fW1,p(Dj)f\in W^{1,p}(D_{j}) with gjfW01,p(Dj)g_{j}-f\in W^{1,p}_{0}(D_{j}), in particular, for f=uf=u, see, e.g., [11, Th. 2.15]. Thus,

hjp,Dj2Cup,Dj+up,Dj.\displaystyle\|h_{j}\|_{p,D_{j}}\leq 2C\|\nabla u\|_{p,D_{j}}+\|u\|_{p,D_{j}}.

Therefore,

hjp,Djp\displaystyle\|h_{j}\|_{p,D_{j}}^{p} 2p1(2C+1)inf{up,Djp+up,Djp:uUj}\displaystyle\leq 2^{p-1}(2C+1)\cdot\inf\left\{\|u\|_{p,D_{j}}^{p}+\|\nabla u\|^{p}_{p,D_{j}}:u\in U_{j}\right\}
=2p1(2C+1)Cp(Kj),\displaystyle=2^{p-1}(2C+1)C_{p}(K_{j}),

where the last step follows from the fact that Cp(Kj)=inf{up,Djp+up,Djp:uUj}C_{p}(K_{j})=\inf\left\{\|u\|_{p,D_{j}}^{p}+\|\nabla u\|^{p}_{p,D_{j}}:u\in U_{j}\right\}. This can be proved the same was as (ii) in §2.2.1 in [16]. This concludes the proof since limjCp(Kj)=0\lim_{j\to\infty}C_{p}(K_{j})=0 by hypothesis. ∎

The following is a slight variation of [6, Proposition 6.1].

Lemma 3.4.

Let KnK\subset\mathbb{R}^{n} be a compact set such that 𝔹1(0)K\mathbb{B}_{1}(0)\cap K\neq\emptyset. Then, for any ϵ>0\epsilon>0, there exists a compact set Kϵ𝔹1(0)¯K_{\epsilon}\subset\overline{\mathbb{B}_{1}(0)} such that

  • (i)

    K𝔹1(0)¯KϵK\cap\overline{\mathbb{B}_{1}(0)}\subset K_{\epsilon},

  • (ii)

    Cp(Kϵ)Cp(K)+ϵC_{p}(K_{\epsilon})\leq C_{p}(K)+\epsilon,

  • (iii)

    𝔹1(0)Kϵ\mathbb{B}_{1}(0)\setminus K_{\epsilon} has smooth boundary.

The above lemma differs from [6, Proposition 6.1] in two ways. On the one hand KϵK_{\epsilon} is a compact subset of 𝔹1(0)¯\overline{\mathbb{B}_{1}(0)} instead of a relatively compact subset of 𝔹1(0)\mathbb{B}_{1}(0), and on the other hand, the Sobolev pp-capacity is used instead of the Newtonian capacity. That the latter change is acceptable is due to the fact that only the properties of monotonicity, countable subadditivity and outer regularity are used. To ensure that the former change is also correct, one only needs to set Kϵ=Ω¯𝔹1(0)¯K_{\epsilon}=\overline{\Omega}\cap\overline{\mathbb{B}_{1}(0)} in the last paragraph of the proof of [6, Proposition 6.1].

Now we are set to prove part (i) of Theorem 1.3.

Proof of part (i) of Theorem 1.3.

Suppose that λ1,p(Ω)>0\lambda_{1,p}(\Omega)>0 and

(ρp(Ω))p>λ1,p(𝔹1(0))/λ1,p(Ω)\displaystyle(\rho_{p}(\Omega))^{p}>\lambda_{1,p}(\mathbb{B}_{1}(0))/\lambda_{1,p}(\Omega)

hold. Then, we may choose a positive R<ρp(D)R<\rho_{p}(D) such that

(3.5) Rp>λ1,p(𝔹1(0))/λ1,p(Ω).\displaystyle R^{p}>\lambda_{1,p}(\mathbb{B}_{1}(0))/\lambda_{1,p}(\Omega).

Let us first consider the case p>np>n. Then, ρp(Ω)=(Ω)\rho_{p}(\Omega)=\mathfrak{R}(\Omega), and, hence, there exists an xnx\in\mathbb{R}^{n} such that 𝔹R(x)Ω\mathbb{B}_{R}(x)\subset\Omega. Using (3.5) as well as (i) and (iii) of Lemma 2.17, we then obtain

λ1,p(𝔹R(x))λ1,p(Ω)>λ1,p(𝔹1(0))Rp=λ1,p(𝔹R(0)).\lambda_{1,p}(\mathbb{B}_{R}(x))\geq\lambda_{1,p}(\Omega)>\lambda_{1,p}(\mathbb{B}_{1}(0))R^{-p}=\lambda_{1,p}(\mathbb{B}_{R}(0)).

This is a contradiction, and hence, (Ω)λ1,p(𝔹1(0))/λ1,p(Ω)\mathfrak{R}(\Omega)\leq\lambda_{1,p}(\mathbb{B}_{1}(0))/\lambda_{1,p}(\Omega).

For the case of p(1,n]p\in(1,n], it follows from the definition of ρp(Ω)\rho_{p}(\Omega) and (3.5) that for all jj\in\mathbb{N} there exists a xjnx_{j}\in\mathbb{R}^{n} with

Cp(𝔹R(xj)¯Ωc)<12jγ,\displaystyle C_{p}(\overline{\mathbb{B}_{R}(x_{j})}\cap\Omega^{c})<\frac{1}{2j\gamma},

where γ:=Rnmax{1,Rp}\gamma:=R^{-n}\max\{1,R^{p}\}. For each jj\in\mathbb{N}, define 𝔄j𝔹1(0)¯\mathfrak{A}_{j}\subset\overline{\mathbb{B}_{1}(0)} by setting

𝔄j={xn:Rx+xj𝔹R(xj)¯Ωc}.\mathfrak{A}_{j}=\{x\in\mathbb{R}^{n}:Rx+x_{j}\in\overline{\mathbb{B}_{R}(x_{j})}\cap\Omega^{c}\}.

Then, it follows from parts (ii) and (iii) of Lemma 2.4 that Cp(𝔄𝔧)<1/(2j)C_{p}(\mathfrak{A_{j}})<1/(2j). Now, for each jj\in\mathbb{N}, we may apply Lemma 3.4 to the pair (𝔄j,1/(2j))(\mathfrak{A}_{j},1/(2j)) to obtain a compact set AjA_{j} such that Cp(Aj)1/jC_{p}(A_{j})\leq 1/j, Ωj:=𝔹1(0)Aj\Omega_{j}:=\mathbb{B}_{1}(0)\setminus A_{j} is an open set with smooth boundary, and 𝔄j𝔹1(0)¯Aj\mathfrak{A}_{j}\cap\overline{\mathbb{B}_{1}(0)}\subset A_{j}. The latter implies that

RΩj+xjΩ.\displaystyle R\Omega_{j}+x_{j}\subset\Omega.

It then follows from parts (i)–(iii) of Lemma 2.17

λ1,p(Ωj)=Rpλ1,p(RΩj+xj)Rpλ1,p(Ω).\displaystyle\lambda_{1,p}(\Omega_{j})=R^{p}\lambda_{1,p}(R\Omega_{j}+x_{j})\geq R^{p}\lambda_{1,p}(\Omega).

Hence, by (3.5), there is some ϵ>0\epsilon>0 such that

λ1,p(Ωj)>λ1,p(𝔹1(0))+ϵ\displaystyle\lambda_{1,p}(\Omega_{j})>\lambda_{1,p}(\mathbb{B}_{1}(0))+\epsilon

for all jj\in\mathbb{N}. This is a contradiction to Lemma 3.1, with D=𝔹1(0)D=\mathbb{B}_{1}(0) and Kj=AjK_{j}=A_{j}, and it follows that

(ρp(Ω))pλ1,p(𝔹1(0))/λ1,p(Ω).\left(\rho_{p}(\Omega)\right)^{p}\leq\lambda_{1,p}(\mathbb{B}_{1}(0))/\lambda_{1,p}(\Omega).

The second part of Theorem 1.3 is proved by using the following lemma.

Lemma 3.6.

Let Ω1Ω2\Omega_{1}\subset\Omega_{2} be open sets in n\mathbb{R}^{n}, Ω2\Omega_{2} connected and bounded. If λ1,p(Ω1)=λ1,p(Ω2)\lambda_{1,p}(\Omega_{1})=\lambda_{1,p}(\Omega_{2}), then Ω2Ω1\Omega_{2}\setminus\Omega_{1} is pp-polar.

The necessity of the connectedness assumption in Lemma 3.6 can be seen by considering the example of Ω1=𝔹1(0)\Omega_{1}=\mathbb{B}_{1}(0) and Ω2=𝔹1(0)𝔹1(x)\Omega_{2}=\mathbb{B}_{1}(0)\cup\mathbb{B}_{1}(x) for some x𝔹¯2(0)x\notin\overline{\mathbb{B}}_{2}(0). In this case λ1,p(Ω1)=λ1,p(Ω2)\lambda_{1,p}(\Omega_{1})=\lambda_{1,p}(\Omega_{2}) while Ω2Ω1\Omega_{2}\setminus\Omega_{1} is pp-polar.

Although, this result is presumably well-known, at least for p=2p=2, we present a proof of Lemma 3.6 here for the sake of completeness.

Proof.

Suppose that λ1,p(Ω1)=λ1,p(Ω2)\lambda_{1,p}(\Omega_{1})=\lambda_{1,p}(\Omega_{2}). Since Ω1\Omega_{1} is bounded, λ1,p(Ω1)\lambda_{1,p}(\Omega_{1}) is positive and there exists a φW01,p(Ω1)\varphi\in W^{1,p}_{0}(\Omega_{1}) which is an eigenfunction of Δp\Delta_{p} with eigenvalue λ1,p(Ω1)\lambda_{1,p}(\Omega_{1}). Then, there exists φ~W01,p(n)\widetilde{\varphi}\in W_{0}^{1,p}(\mathbb{R}^{n}) such that φ~=φ\widetilde{\varphi}=\varphi almost everywhere on Ω1\Omega_{1} and φ~=0\widetilde{\varphi}=0 pp-quasi everywhere on n\mathbb{R}^{n}, see, e.g., [9, Theorem 4.5]. It follows that φ~\widetilde{\varphi} is not zero pp-quasi everywhere on Ω2\Omega_{2} and φ~W01,p(Ω2)\widetilde{\varphi}\in W_{0}^{1,p}(\Omega_{2}) since Ω1Ω2\Omega_{1}\subset\Omega_{2}. Hence, φ~\widetilde{\varphi} is an eigenfunction with eigenvalue λ1,p(Ω1)\lambda_{1,p}(\Omega_{1}) for Δp\Delta_{p} on Ω2\Omega_{2}. The assumption λ1,p(Ω1)=λ1,p(Ω2)\lambda_{1,p}(\Omega_{1})=\lambda_{1,p}(\Omega_{2}) then implies that φ~\widetilde{\varphi} is a first eigenfunction for Δp\Delta_{p} on Ω2\Omega_{2}. Since Ω2\Omega_{2} is connected and bounded, φ~\widetilde{\varphi} is non-zero almost everywhere on Ω2\Omega_{2}. This is a contradiction to φ~=0\widetilde{\varphi}=0 pp-quasi everywhere on Ω2\Omega_{2} unless Ω2Ω1\Omega_{2}\setminus\Omega_{1} is pp-polar. ∎

Proof of part (ii) of Theorem 1.3.

Suppose Ω\Omega is bounded. Then, λ1,p(Ω)>0\lambda_{1,p}(\Omega)>0, hence ρp(Ω)<\rho_{p}(\Omega)<\infty by part (i) of Theorem 1.3. Moreover, by (iv) of Lemma 2.9, there exists an xx^{\circ}such that

Cp(𝔹ρp(Ω)(x)Ωc)=0.\displaystyle C_{p}\left(\mathbb{B}_{\rho_{p}(\Omega)}(x^{\circ})\cap\Omega^{c}\right)=0.

This means that there exists a set KK with Cp(K)=0C_{p}(K)=0, which is relatively closed in 𝔹ρp(Ω)(x)\mathbb{B}_{\rho_{p}(\Omega)}(x^{\circ}), such that

(𝔹ρp(Ω)(x)K)Ω.\left(\mathbb{B}_{\rho_{p}(\Omega)}(x^{\circ})\setminus K\right)\subset\Omega.

By hypothesis,

λ1,p(Ω)=λ1,p(𝔹ρp(Ω)(0))=λ1,p(𝔹ρp(Ω)(x)K).\displaystyle\lambda_{1,p}(\Omega)=\lambda_{1,p}(\mathbb{B}_{\rho_{p}(\Omega)}(0))=\lambda_{1,p}(\mathbb{B}_{\rho_{p}(\Omega)}(x^{\circ})\setminus K).

Lemma 3.6 is now applicable with Ω1=𝔹ρp(Ω)(x)K\Omega_{1}=\mathbb{B}_{\rho_{p}(\Omega)}(x^{\circ})\setminus K and Ω2=Ω\Omega_{2}=\Omega. Thus, we obtain that Ω(𝔹ρp(Ω)(x)K)\Omega\setminus(\mathbb{B}_{\rho_{p}(\Omega)}(x^{\circ})\setminus K) is pp-polar which concludes the proof of (ii) for the bounded case.

Next, suppose Ω\Omega is unbounded such ρp(Ω)=ρp(Ω𝔹R(0))\rho_{p}(\Omega)=\rho_{p}(\Omega\cap\mathbb{B}_{R}(0)) for some R>0R>0, and

λ1,p(𝔹1(0))(ρp(Ω))p=λ1,p(Ω).\lambda_{1,p}(\mathbb{B}_{1}(0))(\rho_{p}(\Omega))^{-p}=\lambda_{1,p}(\Omega).

For ease of notation, write ΩR\Omega_{R} in place of Ω𝔹R(0)\Omega\cap\mathbb{B}_{R}(0). Without loss of generality, we may assume that ρp(ΩR)<R\rho_{p}(\Omega_{R})<R. Part (i) of Lemma 2.17 and (1.4) for ΩR\Omega_{R} yield

λ1,p(Ω)λ1,p(ΩR)λ1,p(𝔹1(0))(ρp(ΩR))p.\lambda_{1,p}(\Omega)\leq\lambda_{1,p}(\Omega_{R})\leq\lambda_{1,p}(\mathbb{B}_{1}(0))\left(\rho_{p}(\Omega_{R})\right)^{-p}.

Thus, it follows from the hypothesis and the two preceeding estimates that

(3.7) λ1,p(ΩR)=λ1,p(𝔹1(0))(ρp(ΩR))p.\displaystyle\lambda_{1,p}(\Omega_{R})=\lambda_{1,p}(\mathbb{B}_{1}(0))\left(\rho_{p}(\Omega_{R})\right)^{-p}.

Denote by {Zj}j\{Z_{j}\}_{j\in\mathbb{N}} the set of connected components of ΩR\Omega_{R}, and note that

(3.8) ρp(ΩR)\displaystyle\rho_{p}(\Omega_{R}) =sup{ρp(Zj):j},\displaystyle=\sup\left\{\rho_{p}(Z_{j})\;:\;j\in\mathbb{N}\right\},
λ1,p(ΩR)\displaystyle\lambda_{1,p}(\Omega_{R}) =inf{λ1,p(Zj):j}.\displaystyle=\inf\left\{\lambda_{1,p}(Z_{j})\;:\;j\in\mathbb{N}\right\}.

Since ΩR\Omega_{R} is bounded and ρp(ΩR)>0\rho_{p}(\Omega_{R})>0, it follows that ρp(ΩR)\rho_{p}(\Omega_{R}) is attained at ρp(Zj0)\rho_{p}(Z_{j_{0}}) for some j0j_{0}\in\mathbb{N}. Then, (1.4) yields

0<λ1,p(ΩR)λ1,p(Zj0)\displaystyle 0<\lambda_{1,p}(\Omega_{R})\leq\lambda_{1,p}(Z_{j_{0}}) λ1,p(𝔹1(0))(ρp(Zj0))p\displaystyle\leq\lambda_{1,p}(\mathbb{B}_{1}(0))\left(\rho_{p}(Z_{j_{0}})\right)^{-p}
=λ1,p(𝔹1(0))(ρp(ΩR))p\displaystyle=\lambda_{1,p}(\mathbb{B}_{1}(0))\left(\rho_{p}(\Omega_{R})\right)^{-p}

Thus, by (3.7) and the choice of j0j_{0}, it follows that λ1,p(ΩR)=λ1,p(Zj0)\lambda_{1,p}(\Omega_{R})=\lambda_{1,p}(Z_{j_{0}}). We now may apply part (ii) of Theorem 1.3 for the bounded case to Zj0Z_{j_{0}}. That is, there exists an xnx\in\mathbb{R}^{n} such that Zj0Z_{j_{0}} is equal to 𝔹ρp(ΩR)(x)\mathbb{B}_{\rho_{p}(\Omega_{R})}(x) modulo a pp-polar set KK. Now, either the boundary of 𝔹ρp(ΩR)(x)\mathbb{B}_{\rho_{p}(\Omega_{R})}(x) meets the boundary of 𝔹R(0)\mathbb{B}_{R}(0) tangentially or the closure of 𝔹ρp(ΩR)(x)\mathbb{B}_{\rho_{p}(\Omega_{R})}(x) is contained in 𝔹R(0)\mathbb{B}_{R}(0). In the latter case, it follows that Zj0Z_{j_{0}} is actually a connected component of Ω\Omega, and hence Ω=Zj0\Omega=Z_{j_{0}} which is a contradiction to Ω\Omega being unbounded. In the former case, it follows from the openess of Ω\Omega and ρp(ΩR)<R\rho_{p}(\Omega_{R})<R that Zj0Z_{j_{0}} is in fact a connected component of Ω\Omega which, again, is a contradiction. Thus, equality in (1.4) cannot hold.

4. Proof of Theorem 1.6

The next theorem is the essence of Theorem 1.6. It is a special case of Theorem 8.3.3 in [1] which is formulated for (1,p)(1,p)-extension domains. A domain Ω\Omega is of this class, if there exists a linear, bounded operator EΩ:W1,p(Ω)W1,p(n)E_{\Omega}:W^{1,p}(\Omega)\longrightarrow W^{1,p}(\mathbb{R}^{n}) such that (EΩf)|Ω=f(E_{\Omega}f)_{|_{\Omega}}=f for fW1,p(Ω)f\in W^{1,p}(\Omega).

Theorem 4.1.

Let Ωn\Omega\subset\mathbb{R}^{n} be a bounded (1,p)(1,p)-extension domain, and suppose that fW1,p(Ω)f\in W^{1,p}(\Omega), 1<p<1<p<\infty. Let KK be a closed subset of Ω\Omega such that Cp(K)>0C_{p}(K)>0. Suppose that f|K=0f_{|_{K}}=0, then

(4.2) fp,Ω(m(Ω))1pEΩopfp,Ω(Cp(K))1/p,\displaystyle\|f\|_{p,\Omega}\leq\left(m(\Omega)\right)^{\frac{1}{p}}\cdot\|E_{\Omega}\|_{\operatorname{op}}\cdot\frac{\|\nabla f\|_{p,\Omega}}{\left(C_{p}(K)\right)^{1/p}},

where m(Ω)m(\Omega) is the Lebesgue measure of Ω\Omega and EΩop\|E_{\Omega}\|_{\operatorname{op}} is the operator norm of a bounded, linear extension operator EΩ:W1,p(Ω)W1,p(n)E_{\Omega}:W^{1,p}(\Omega)\longrightarrow W^{1,p}(\mathbb{R}^{n}) such that (EΩf)|Ω=f(E_{\Omega}f)_{|_{\Omega}}=f for fW1,p(Ω)f\in W^{1,p}(\Omega).

Proof.

This follows straightforwardly from Theorem 8.3.3 and Lemma 8.3.2 in [1]. First, consider Lemma 8.3.2, with m=1m=1, σ=0\sigma=0, so that a0=Ωf𝑑μ0a_{0}=\int_{\Omega}f\;d\mu_{0} and

LfW1,p(Ω)=a0(m(Ω))1/p.\|Lf\|_{W^{1,p}(\Omega)}=a_{0}\cdot(m(\Omega))^{1/p}.

Then, use Theorem 8.3.3, with m=1m=1, β=0\beta=0, and G1μ0p=Cp(K)1/p\|G_{1}*\mu_{0}\|_{p^{\prime}}=C_{p}(K)^{-1/p}. Collecting the constants determining AA in (8.3.7), we get A=(m(Ω))1/pEΩopA=(m(\Omega))^{1/p}\cdot\|E_{\Omega}\|_{\operatorname{op}}. ∎

Proof of Theorem 1.6.

It follows from Calderón’s work in [3, Theorem 12] that any open square in RnR^{n} is a (1,p)(1,p)-extension domain. Hence, for given R>0R>0, there exist bounded, linear operators from W1,p((0,R)n)W^{1,p}((0,R)^{n}) to W1,p(n)W^{1,p}(\mathbb{R}^{n}) which are the identity operator on (0,R)n(0,R)^{n}. Let ERE_{R} be such a operator.

Suppose ρp(Ω)\rho_{p}(\Omega) is finite. Let R>ρp(Ω)R>\rho_{p}(\Omega) and δR(Ω)\delta_{R}(\Omega) be defined as in (1.5). We note that arguments analogous to the ones used in the proof of Lemma 2.7 yield

δR(Ω)=sup{δ>0:Cp(𝔹R(x)Ωcδfor allxn}.\delta_{R}(\Omega)=\sup\{\delta>0\,:\;C_{p}(\mathbb{B}_{R}(x)\cap\Omega^{c}\geq\delta\;\text{for all}\;x\in\mathbb{R}^{n}\}.

Now, for each mnm\in\mathbb{Z}^{n}, set

Qm={xn:xj(Rmj,R(mj+1))for allj{1,,n}}.\displaystyle Q_{m}=\left\{x\in\mathbb{R}^{n}:x_{j}\in\left(Rm_{j},R(m_{j}+1)\right)\;\;\text{for all}\;j\in\{1,\dots,n\}\right\}.

Since

n|.|pdm=mnQm|.|pdm,\displaystyle\int_{\mathbb{R}^{n}}|\,.\,|^{p}\;dm=\sum_{m\in\mathbb{Z}^{n}}\int_{Q_{m}}|\,.\,|^{p}\;dm,

it suffices to show that

fp,QmRn/pERop(δR(Ω))1/pfp,Qmfor allf𝒞c(Ω)\displaystyle\|f\|_{p,Q_{m}}\leq\frac{R^{n/p}\|E_{R}\|_{\operatorname{op}}}{(\delta_{R}(\Omega))^{1/p}}\cdot\|\nabla f\|_{p,Q_{m}}\;\;\;\;\text{for all}\;f\in\mathcal{C}_{c}^{\infty}(\Omega)

holds for each mnm\in\mathbb{Z}^{n}.

For that, let η(0,δR(Ω))\eta\in(0,\delta_{R}(\Omega)) be fixed. Then, by our choice of RR, we may choose a compact set KmQmΩcK_{m}\subset Q_{m}\cap\Omega^{c} such that Cp(Km)ηC_{p}(K_{m})\geq\eta for each mnm\in\mathbb{Z}^{n}. Since f𝒞c(Ω)f\in\mathcal{C}_{c}^{\infty}(\Omega) implies that f𝒞(Qm¯)f\in\mathcal{C}^{\infty}(\overline{Q_{m}}) with f|Km=0f_{|_{K_{m}}}=0 for all mnm\in\mathbb{Z}^{n}, Theorem 4.1 is applicable here and yields

fp,QmRn/pERopη1/pfp,Qmfor allf𝒞c(Ω)\displaystyle\|f\|_{p,Q_{m}}\leq\frac{R^{n/p}\|E_{R}\|_{\operatorname{op}}}{\eta^{1/p}}\|\nabla f\|_{p,Q_{m}}\;\;\text{for all}\;f\in\mathcal{C}_{c}^{\infty}(\Omega)

for all mnm\in\mathbb{Z}^{n}. As this holds for any η(0,δR(Ω))\eta\in(0,\delta_{R}(\Omega)), the proof of part (i) of Theorem 1.6, i.e.,

δR(Ω)RnERoppλ1,p(Ω)\frac{\delta_{R}(\Omega)}{R^{n}\|E_{R}\|_{op}^{p}}\leq\lambda_{1,p}(\Omega)

is complete.

To prove part (ii) of Theorem 1.6, again, suppose ρp(Ω)<\rho_{p}(\Omega)<\infty and R>ρp(Ω)R>\rho_{p}(\Omega). Set γR(Ω)=δR(Ω)/Cp(𝔹R(0))\gamma_{R}(\Omega)=\delta_{R}(\Omega)/C_{p}(\mathbb{B}_{R}(0)), i.e.,

γR(Ω)=sup{γ(0,1):Cp(𝔹R(x)Ωc)γCp(𝔹R(0))for allxn}.\gamma_{R}(\Omega)=\sup\{\gamma\in(0,1)\;:\;C_{p}(\mathbb{B}_{R}(x)\cap\Omega^{c})\geq\gamma C_{p}(\mathbb{B}_{R}(0))\;\;\text{for all}\;x\in\mathbb{R}^{n}\}.

This means that rp,γ(Ω)<Rr_{p,\gamma}(\Omega)<R for all γ(0,γR(Ω))\gamma\in(0,\gamma_{R}(\Omega)), see also Corollary 2.15. Now, if 1<p<n1<p<n, then there exists a constant c=c(n,p)c=c(n,p) such that

(4.3) Cp(𝔹R(0))cRnp,\displaystyle C_{p}(\mathbb{B}_{R}(0))\geq cR^{n-p},

see Example 2.12 in [9]. Hence,

cγR(Ω)RnpδR(Ω).c\gamma_{R}(\Omega)\cdot R^{n-p}\leq\delta_{R}(\Omega).

Next, we may choose an extension operator ERE_{R} for (0,R)n(0,R)^{n} such that ERop\|E_{R}\|_{op} only depends on nn and pp, and not on RR, see Stein’s construction of extension operators in [19, Theorem 5], in particular, note Theorem 5’ and (c) on pg. 190 therein in regards to the operator norm dependencies. It follows that there exists a constant C=C(n,p)>0C=C(n,p)>0 such that

(4.4) CγR(Ω)Rpλ1,p(Ω).\displaystyle C\gamma_{R}(\Omega)\cdot R^{-p}\leq\lambda_{1,p}(\Omega).

If pnp\geq n, we replace (4.3) in the above argument by

Cp(𝔹R(0))cRnp1+Rp,\displaystyle C_{p}(\mathbb{B}_{R}(0))\geq c\frac{R^{n-p}}{1+R^{-p}},

which is obtained from Theorem 2.38 and Example 2.12 in [9]. ∎

Remark.

Estimate (4.4) is in line with the estimate obtained by Maz’ya–Shubin in case of p=2p=2, n3n\geq 3, in (3.19) of [14], since rp,γR(Ω)(Ω)=Rr_{p,\gamma_{R}(\Omega)}(\Omega)=R.

Statements and Declarations

Conflict of interest

The author has no conflict of interest to declare.

Data availability

Not applicable.

References

  • [1] Adams, D. R., and Hedberg, L. I. Function spaces and potential theory, vol. 314 of Grundlehren Math. Wiss. Berlin: Springer-Verlag, 1995.
  • [2] Agmon, S. Lectures on elliptic boundary value problems. Prepared for publication by B. Frank Jones jun. with the assistance of George W. Batten jun. Van Nostrand Mathematical Studies. 2. Princeton, N.J.-Toronto-New York-London: D. van Nostrand Company, Inc. v, 291 p. (1965)., 1965.
  • [3] Calderón, A. P. Lebesgue spaces of differentiable functions and distributions. Proc. Sympos. Pure Math. 4, 33-49 (1961)., 1961.
  • [4] Evans, L. C. Partial differential equations, vol. 19 of Graduate Studies in Mathematics. American Mathematical Society, 1998.
  • [5] Fontana, L., Morpurgo, C., and Qin, L. Moser–Trudinger inequalities: from local to global. To appear in Annali di Matematica Pura ed Applicata (2024). doi:10.1007/s10231-024-01481-9, arXiv:2408.06989.
  • [6] Gallagher, A.-K. On the Poincaré inequality on open sets in n\mathbb{R}^{n}. To appear in Computational Methods and Function Theory. doi:10.1007/s40315-024-00550-7, arXiv:2404.19207.
  • [7] Gallagher, A.-K., Lebl, J., and Ramachandran, K. The closed range property for the ¯\overline{\partial}-operator on planar domains. J. Geom. Anal. 31, 2 (2021), 1646–1670.
  • [8] Gariepy, R., and Ziemer, W. P. A regularity condition at the boundary for solutions of quasilinear elliptic equations. Arch. Ration. Mech. Anal. 67 (1977), 25–39.
  • [9] Heinonen, J., Kilpeläinen, T., and Martio, O. Nonlinear potential theory of degenerate elliptic equations. With a new preface, corrigenda, and epilogue consisting of other new material, reprint of the 2006 edition ed. Dover Books Math. Mineola, NY: Dover Publications, 2018.
  • [10] Lieb, E. H. On the lowest eigenvalue of the Laplacian for the intersection of two domains. Invent. Math. 74 (1983), 441–448.
  • [11] Lindqvist, P. Notes on the pp-Laplace equation, vol. 102 of Rep., Univ. Jyväskylä, Dep. Math. Stat. Jyväskylä: University of Jyväskylä, Department of Mathematics and Statistics, 2006.
  • [12] Lindqvist, P. A nonlinear eigenvalue problem. In Topics in mathematical analysis. Hackensack, NJ: World Scientific, 2008, pp. 175–203.
  • [13] Littman, W. A connection between α\alpha-capacity and mpm-p polarity. Bull. Am. Math. Soc. 73 (1967), 862–866.
  • [14] Maz’ya, V., and Shubin, M. Can one see the fundamental frequency of a drum? Letters in Mathematical Physics 74 (2005), 135–1151.
  • [15] Maz’ya, V. G. On continuity in a boundary point of solutions of quasilinear elliptic equations. Vestn. Leningr. Univ., Mat. Mekh. Astron. 25, 3 (1970), 42–55.
  • [16] Maz’ya, V. G. Sobolev spaces. With applications to elliptic partial differential equations. Transl. from the Russian by T. O. Shaposhnikova, 2nd revised and augmented ed. ed., vol. 342 of Grundlehren Math. Wiss. Berlin: Springer, 2011.
  • [17] Meyers, N. G. Integral inequalities of Poincaré and Wirtinger type. Arch. Ration. Mech. Anal. 68 (1978), 113–120.
  • [18] Souplet, P. Geometry of unbounded domains, Poincaré inequalities and stability in semilinear parabolic equations. Comm. Partial Differential Equations 24, 5-6 (1999), 951–973.
  • [19] Stein, E. M. Singular integrals and differentiability properties of functions, vol. 30 of Princeton Math. Ser. Princeton University Press, Princeton, NJ, 1970.
  • [20] Vitolo, A. H1,pH^{1,p}-eigenvalues and LL^{\infty}-estimates in quasicylindrical domains. Commun. Pure Appl. Anal. 10, 5 (2011), 1315–1329.
  • [21] Wallin, H. A connection between α\alpha-capacity and LpL^{p}-classes of differentiable functions. Ark. Mat. 5 (1965), 331–341.