Equivalence between validity of the -Poincaré inequality and finiteness of the strict -capacitary inradius.
Abstract.
It is shown that the -Poincaré inequality holds on an open set in if and only if the strict -capacitary inradius of 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:
-Poincaré inequality, capacitary inradius, -Laplacian1991 Mathematics Subject Classification:
35P30, 35J70, 31C451. Introduction
Let be an open set in , and the -norm on for . The classical -Poincaré inequality is said to hold on if there exists a constant such that
(1.1) |
where denotes the space of smooth functions with compact support in . That is, (1.1) holds if and only if the infimum of the so-called nonlinear Rayleigh quotient
is positive. If is positive and attained at some non-zero function in , i.e., the closure of with respect to the -Sobolev 1-norm, then it is the smallest generalized eigenvalue for the -Laplacian, with Dirichlet boundary condition, on in the distributional sense.
The validity of (1.1), or equivalently the positivity of , 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 , 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 -capacitary inradius of , originally introduced in the case of in [7] for and in [6] for , and derive the equivalence between its finiteness and the positivity of . In the definition of the strict -capacitary inradius, we use the notion of Sobolev -capacity, , which, for a compact set , is defined as
where denotes the space of Schwartz functions. The strict -capacitary inradius of is defined as
(1.2) |
where denotes the open ball in of radius with center . Roughly speaking, finiteness of means that the complement of is somewhat evenly und uniformly distributed in . In fact, if is finite and , then there exists a such that within -units of any point in , we may find a set in the complement of whose Sobolov -capacity is at least .
Finiteness of is a necessary condition for (1.1) to hold on an open set . Indeed, our first result yields a sharp upper bound for in terms of .
Theorem 1.3.
The heart of the matter of the proof of Theorem 1.3 is a continuity result for . That is, suppose is a sequence of compact sets contained in the closure of some bounded domain with smooth boundary. If tends to zero as and, for each , is an open set with smooth boundary, then tends to . This was orginally proved in [7] for the case , , , and the logarithmic capacity in place of . Regularity results for the Dirichlet problem and for first eigenfunctions of the -Laplacian, , on open, bounded sets with smooth boundary as well as the theory of -harmonic functions allow this continuity result to be extended to the case of the Sobolev -capacity with , see Lemma 3.1.
To state the sufficiency of the finiteness of , we define the scalar
(1.5) |
for any . Note that is positive whenever is finite and .
Theorem 1.6.
Let be an open set, . Suppose and .
-
(i)
Then,
(1.7) for any bounded, linear extension operator satisfying for .
- (ii)
Part (i) of Theorem 1.6 is a direct consequence of a Poincaré-type inequality for any function in which has a representative that vanishes on a set of positive -Sobolev capacity in , where 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 as a union of closed cubes with mutually disjoint interiors and of side length larger than the strict -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 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 be an open set, . Then,
It is well-known that the finiteness of the inradius, , of an open set , i.e., the supremum of the radii of all balls contained in , 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 , that this condition is also sufficient as long as is a domain which satisfies a uniform exterior cone condition; in the case of 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 . 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 -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 -capacity. However, for any pair with , there exists a set of Lebesgue measure zero, such that while , 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 and . 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 and , estimate (1.4) is an improvement over the upper bounded for 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 and is a domain with finite inradius, , then . Note that if , then singletons have positive -Sobolev capacity, so that , 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 , in potential-theoretic terms in [7]. Originally, we intended to investigate which potential-theoretic conditions yield the -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 . Moreover, we showed that Corollary 1.9 holds for and that (1.1) is equivalent to the existence of a smooth, bounded function on such that its Laplacian has a positive lower bound on . These results were later shown to hold for , with the Newtonian capacity in place of in [6] by the author of this note. We show in Lemma 2.11 that the strict -capacitary inradius defined in (1.2) does not depend on the choice of -capacity as long as the sets of zero -capacity are the same as the sets of for which 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 , see the paragraph subsequent to the proof of Lemma 2.11.
This note is structured as follows. The notions of capacity, strict -capacitary inradius, and 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 -capacity
Definition 2.1.
Let be a compact set in , . Then
This definition may be extended to open sets by setting
(2.2) |
It then follows that
which can be proven analogously to [1, Proposition 2.2.3]. The definition of may now be extended to arbitrary sets by setting
(2.3) |
for . A set is called -polar, if . Moreover, two functions are said to equal -quasi everywhere if they equal outside a -polar set.
Next, we present some standard properties of .
Lemma 2.4.
Let . Then
-
(i)
If , then .
-
(ii)
If , then .
-
(iii)
If , then .
-
(iv)
If such that , then
-
(v)
Any Borel set is capacitable, i.e.,
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
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 -capacitary inradius
In a slight deviation from (1.2), we define the strict -capacitary inradius as follows.
Definition 2.5.
Let be an open set, . Then, the strict -capacitary inradius of is defined as
(2.6) |
We show first that this definition of agrees with (1.2), although is not invariant under taking closures.
Lemma 2.7.
Let be an open set, . Then
(2.8) |
Proof.
Let us denote the right hand side of (2.8) by . By the monotonicity of the Sobolev -capacity, it is immediate that . Now, suppose that . Then, for all there exists an such that , and hence
Therefore, for all , which implies . ∎
In the following, we collect some basic properties of the strict -capacitary inradius. To do so, we recall that the inradius, , of an open set is defined as
We also define the -capacitary inradius, , by
Lemma 2.9.
Let be an open set, .
-
(i)
If and , then and .
-
(ii)
If is an open set, then . If additionally, is -polar, then .
-
(iii)
, and equality holds if .
-
(iv)
If is bounded, then . Moreover, there exists an , such that
i.e., is attained.
-
(v)
.
Proof.
The translation invariance of holds because it holds for , see (ii) of Lemma 2.4. To check the linearity under dilations we first note that
Thus, if for a given and there exists an such that
then
by (iii) of Lemma 2.4. It then follows that whenever , and hence, for any . We now may repeat this argument with in place of and in place of to obtain
which yields
hence, the proof of (i) is complete.
The first part of (ii) follows from the definition. The second part follows after observing
by (iv) of Lemma 2.4 and the fact that is -polar.
The set of inequalities in (iii) follows directly from the definitions of the inradii. Equality holds if , because for all . To wit, if , then is non-empty for all so that for all . Hence, , so that follows.
For the proof of (iv), suppose is bounded. Then, by definition of , there exists a sequence in such that is an increasing sequence which converges to , and
Since is a bounded set, it follows that is a bounded sequence, thus, has a convergent subsequence. For ease of notation, let us denote the subsequence by . Write for the limit point. It suffices to prove that
(2.10) |
holds. To prove (2.10), let and choose such that for all . Then, choose such that for all . It follows that
and, therefore,
Letting then yields
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 -capacitary inradius does not depend on the choice of -capacity.
Lemma 2.11.
Let be open, . Let be such that
-
(a)
,
-
(b)
,
-
(c)
all Borel sets are capacitable with respect to .
Suppose iff for all bounded Borel sets . Then
(2.12) |
Proof.
Note first that properties (a)–(c) ensure that (iv) and (v) of Lemma 2.9 hold for the capacitary inradius, , defined by the right hand side of (2.12).
Next, write for for . Then, by (v) and (iv) of Lemma 2.9, it follows that
(2.13) |
where the last step follows from the assumption that iff for all bounded Borel set . Since (iv)-(v) of Lemma 2.9 hold for the (strict) capacitary inradius with respect to , it follows
i.e., is invariant under the choice of . ∎
Both the logarithmic capacity for and the Newtonian capacity for 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 by Theorem 1 (, ) in [13], see also Theorems A and B (, ) in [21]. As a consequence, the strict capacitary inradius with respect to is the same as the one defined in [7] with respect to the logarithmic capacity for and the one defined in [6] with respect to the Newtonian capacity for .
Maz’ya and Shubin used a different notion of inradius, formulated in terms of the Wiener capacity, in their work [14], for and . To wit, they defined the interior capacitary radius, , of an open set for by
where the Wiener capacity , , is defined by
for compact.
This interior capacitary radius relates to , for , as follows.
Lemma 2.14.
Let , , be an open set. Then .
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 , where
Moreover, the arguments supplied in the proof of (2.8) let us work with this alternative formulation for :
Now, let be given. Let , and choose . Then, by definition of , there exists an such that
Thus, for any given , and hence, .
Next, let . Then, there exists an such that for all . Let . It then follows that
for all . That is, for all . Since is increasing in , we obtain that . Hence, . ∎
Understanding the -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
for and . A proof similar to the one given in Lemma 2.14 then yields the following.
Corollary 2.15.
Let and . Then .
2.3. Infimum of the nonlinear Rayleigh quotient for -Laplacian
Definition 2.16.
Let be an open set, . Then
We collect some elementary properties of in the following lemma.
Lemma 2.17.
Let be an open set. Then
-
(i)
If is an open set, then
-
(ii)
If , then
-
(iii)
If , then
-
(iv)
If is an open set such that is -polar, then
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 is -polar, then , see, for instance, [9, Theorem 2.43]. Hence, by definition, . ∎
If is positive and attained at some , then is a weak solution to
(2.18) |
for . Such a function is called a first eigenfunction of , where
Here, being a weak solution to (2.18) means that
(2.19) |
for all .
It is well-known that if is a smoothly bounded domain, then is positive and attained at some non-zero . Moreover, such may be assumed to be positive in . A comprehensive resource for standard results on the first eigenvalue and eigenfunctions for are the lecture notes by Lindqvist [12]. The regularity result for the first eigenfunctions is due to Gariepy–Ziemer in [8] for ; in the case of , it is known that, under these conditions on , a representative of is in , 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 be a smoothly bounded domain. Let , , be compact sets. Suppose that and is an open set with smooth boundary for all . Then, .
Proof.
Note first that if , then implies that is empty for all sufficiently large, and hence, the conclusion holds trivially.
Since is a bounded domain, , and there exists a weak solution, , to (2.18) which is positive on , see Section 2.3 for references. After rescaling, we may further assume that on . Next, for each , there exists a weak solution to the Dirichlet problem for with boundary data on , i.e.,
see, e.g., [11, Th. 2.16]. It follows from work by Maz’ya [15], see also [11, Th. 2.16], that , and hence on and on . Finally, set on . Then, and holds weakly on . Using (2.19), we compute
where the last step follows from Hölder inequality. Using this estimate, in conjunction with the definition of , yields
and therefore,
Since is bounded on and , so that as well, it follows from the Dominated Convergence Theorem that
Hence, it suffices to show that .
For that, define
for each . Temporarily fix and . Then, we may find a which is a weak solution to
(3.2) |
Note that holds on by the comparison principle for -harmonic functions, see, e.g., [11, Th. 2.15], and the fact that on . Thence,
(3.3) |
By (3.2), . Thus, since , the -Poincaré inequality is applicable to the first term on the right hand side of (3.3) with a uniform constant, i.e.,
with . Therefore, we have
Next, note that (3.2) implies that is a quasi-minimizer, i.e.,
for any with , in particular, for , see, e.g., [11, Th. 2.15]. Thus,
Therefore,
where the last step follows from the fact that . This can be proved the same was as (ii) in §2.2.1 in [16]. This concludes the proof since by hypothesis. ∎
The following is a slight variation of [6, Proposition 6.1].
Lemma 3.4.
Let be a compact set such that . Then, for any , there exists a compact set such that
-
(i)
,
-
(ii)
,
-
(iii)
has smooth boundary.
The above lemma differs from [6, Proposition 6.1] in two ways. On the one hand is a compact subset of instead of a relatively compact subset of , and on the other hand, the Sobolev -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 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 and
hold. Then, we may choose a positive such that
(3.5) |
Let us first consider the case . Then, , and, hence, there exists an such that . Using (3.5) as well as (i) and (iii) of Lemma 2.17, we then obtain
This is a contradiction, and hence, .
For the case of , it follows from the definition of and (3.5) that for all there exists a with
where . For each , define by setting
Then, it follows from parts (ii) and (iii) of Lemma 2.4 that . Now, for each , we may apply Lemma 3.4 to the pair to obtain a compact set such that , is an open set with smooth boundary, and . The latter implies that
It then follows from parts (i)–(iii) of Lemma 2.17
Hence, by (3.5), there is some such that
for all . This is a contradiction to Lemma 3.1, with and , and it follows that
∎
The second part of Theorem 1.3 is proved by using the following lemma.
Lemma 3.6.
Let be open sets in , connected and bounded. If , then is -polar.
The necessity of the connectedness assumption in Lemma 3.6 can be seen by considering the example of and for some . In this case while is -polar.
Although, this result is presumably well-known, at least for , we present a proof of Lemma 3.6 here for the sake of completeness.
Proof.
Suppose that . Since is bounded, is positive and there exists a which is an eigenfunction of with eigenvalue . Then, there exists such that almost everywhere on and -quasi everywhere on , see, e.g., [9, Theorem 4.5]. It follows that is not zero -quasi everywhere on and since . Hence, is an eigenfunction with eigenvalue for on . The assumption then implies that is a first eigenfunction for on . Since is connected and bounded, is non-zero almost everywhere on . This is a contradiction to -quasi everywhere on unless is -polar. ∎
Proof of part (ii) of Theorem 1.3.
Suppose is bounded. Then, , hence by part (i) of Theorem 1.3. Moreover, by (iv) of Lemma 2.9, there exists an such that
This means that there exists a set with , which is relatively closed in , such that
By hypothesis,
Lemma 3.6 is now applicable with and . Thus, we obtain that is -polar which concludes the proof of (ii) for the bounded case.
Next, suppose is unbounded such for some , and
For ease of notation, write in place of . Without loss of generality, we may assume that . Part (i) of Lemma 2.17 and (1.4) for yield
Thus, it follows from the hypothesis and the two preceeding estimates that
(3.7) |
Denote by the set of connected components of , and note that
(3.8) | ||||
Since is bounded and , it follows that is attained at for some . Then, (1.4) yields
Thus, by (3.7) and the choice of , it follows that . We now may apply part (ii) of Theorem 1.3 for the bounded case to . That is, there exists an such that is equal to modulo a -polar set . Now, either the boundary of meets the boundary of tangentially or the closure of is contained in . In the latter case, it follows that is actually a connected component of , and hence which is a contradiction to being unbounded. In the former case, it follows from the openess of and that is in fact a connected component of 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 -extension domains. A domain is of this class, if there exists a linear, bounded operator such that for .
Theorem 4.1.
Let be a bounded -extension domain, and suppose that , . Let be a closed subset of such that . Suppose that , then
(4.2) |
where is the Lebesgue measure of and is the operator norm of a bounded, linear extension operator such that for .
Proof.
This follows straightforwardly from Theorem 8.3.3 and Lemma 8.3.2 in [1]. First, consider Lemma 8.3.2, with , , so that and
Then, use Theorem 8.3.3, with , , and . Collecting the constants determining in (8.3.7), we get . ∎
Proof of Theorem 1.6.
It follows from Calderón’s work in [3, Theorem 12] that any open square in is a -extension domain. Hence, for given , there exist bounded, linear operators from to which are the identity operator on . Let be such a operator.
Suppose is finite. Let and be defined as in (1.5). We note that arguments analogous to the ones used in the proof of Lemma 2.7 yield
Now, for each , set
Since
it suffices to show that
holds for each .
For that, let be fixed. Then, by our choice of , we may choose a compact set such that for each . Since implies that with for all , Theorem 4.1 is applicable here and yields
for all . As this holds for any , the proof of part (i) of Theorem 1.6, i.e.,
is complete.
To prove part (ii) of Theorem 1.6, again, suppose and . Set , i.e.,
This means that for all , see also Corollary 2.15. Now, if , then there exists a constant such that
(4.3) |
see Example 2.12 in [9]. Hence,
Next, we may choose an extension operator for such that only depends on and , and not on , 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 such that
(4.4) |
If , we replace (4.3) in the above argument by
which is obtained from Theorem 2.38 and Example 2.12 in [9]. ∎
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 . 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 -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 -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 -capacity and 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. -eigenvalues and -estimates in quasicylindrical domains. Commun. Pure Appl. Anal. 10, 5 (2011), 1315–1329.
- [21] Wallin, H. A connection between -capacity and -classes of differentiable functions. Ark. Mat. 5 (1965), 331–341.