Home Continuity properties of solutions to the p-Laplace system
Article Publicly Available

Continuity properties of solutions to the p-Laplace system

  • Angela Alberico , Andrea Cianchi and Carlo Sbordone EMAIL logo
Published/Copyright: November 17, 2015

Abstract

A sharp integrability condition on the right-hand side of the p-Laplace system for all its solutions to be continuous is exhibited. Their uniform continuity is also analyzed and estimates for their modulus of continuity are provided. The relevant estimates are shown to be optimal as the right-hand side ranges in classes of rearrangement-invariant spaces, such as Lebesgue, Lorentz, Lorentz–Zygmund, and Marcinkiewicz spaces, as well as some customary Orlicz spaces.

MSC 2010: 35B65; 35J60; 46E30

1 Introduction

The present paper deals with continuity properties of local solutions to the p-Laplacian system in open subsets of the Euclidean space n with 2pn. Our main results provide minimal integrability conditions on the right-hand side of the relevant system for every solution to be continuous or uniformly continuous. Moreover, they yield sharp information on their modulus of continuity. These results, in their general formulation, are stated in Section 3.

In this section, we enucleate some special cases of our conclusions for the n-Laplace system, which exhibits certain peculiar interesting features due to its borderline character. In recent years, the regularity properties of its solutions have been the subject of various contributions, most of which confined to the case of a single equation, such as [1, 20, 21, 23, 24, 25, 31]. The study of optimal continuity properties of its solutions was, in fact, our original motivation for the present research, which, in particular, complements and improves earlier results in [20, 21].

Let Ω be an open subset of n and let 𝒇:ΩN. Since we are concerned with local solutions to the n-Laplace system

(1.1)-div(|𝒖|n-2𝒖)=𝒇(x)in Ω,

we may suppose that Ω is bounded, with Lebesgue measure |Ω|=1. Our first result tells us that membership of 𝒇 to the Lorentz space L(1,1n-1)(Ω,N) is an optimal condition for every solution 𝒖 to (1.1) to be continuous.

Recall that L(1,1n-1)(Ω,N) is the space of all measurable functions 𝒈:ΩN whose quasi-norm

𝒈L(1,1n-1)(Ω,N)=(01|𝒈|**(s)1n-1s-1+1n-1𝑑s)n-1

is finite. Here, |𝒈|* denotes the decreasing rearrangement of the length |𝒈| of 𝒈 and

|𝒈|**(s)=1s0s|𝒈|*(r)𝑑rfor s>0.

When N=1, the abridged notation L(1,1n-1)(Ω) will be employed for L(1,1n-1)(Ω,). An analogous simplified notation will be adopted for other function spaces.

Theorem 1.1

Let n2. Then, L(1,1n-1)(Ω,RN) is the largest rearrangement-invariant quasi-normed space X(Ω,RN) such that any local solution to (1.1) is continuous for every 𝐟X(Ω,RN).

In the scalar case (N=1), the fact that every solution to (1.1) is continuous under the assumption that 𝒇L(1,1n-1)(Ω) is proved in [21, Theorem 1.1] via a different approach. Theorem 1.1 above not only includes systems, but also characterizes the space L(1,1n-1)(Ω,N) as the optimal rearrangement-invariant quasi-normed space enjoying this property. One important trait of Theorem 1.1 is thus to reduce the question of the continuity of all solutions 𝒖 to (1.1), as 𝒇 ranges in some rearrangement-invariant quasi-normed space X(Ω,N), to the merely functional-analytic problem of the inclusion

X(Ω,N)L(1,1n-1)(Ω,N).

As a consequence, minimal conditions on X(Ω,N) in classes of function spaces for any solution to (1.1) to be continuous for every 𝒇X(Ω,N) can be derived.

The following result deals with the situation when X(Ω,N) is an Orlicz spaceLA(Ω,N) associated with a Young function A.

Theorem 1.2

Let A be a Young function. Assume that

{lim supttlog(1+t)A(t)<if n=2,(ttA(t)-A(t))1n-2dtt<if n3.

If 𝐟LA(Ω,RN), then any local solution 𝐮 to (1.1) is continuous.

Let us now focus on the case when X(Ω,N) is a so-called classical Lorentz spaceΛ1(ν)(Ω,N) associated with a weight ν. Unless otherwise stated, we shall assume that the weight function ν:(0,1)[0,) is continuous. The space Λ1(ν)(Ω,N) consists of all measurable functions 𝒈 on Ω for which the quantity

𝒈Λ1(ν)(Ω,N)=01|𝒈|*(s)ν(s)𝑑s

is finite.

Theorem 1.3

Let ν be a non-decreasing weight. Then, any local solution 𝐮 to (1.1) is continuous for every 𝐟Λ1(ν)(Ω,RN) if and only if

(1.2){lim supt0+tlog1t0tν(s)𝑑s<if n=2,0(tlog1t0tν(s)𝑑s)1n-2dtt<if n3.

Theorem 1.3 should be compared with [21, Corollary 4.1], where the scalar case (N=1) is considered and a more implicit, just sufficient condition on the weight ν is given for any solution to (1.1) to be continuous when fΛ1(ν)(Ω).

Remark 1.4

Theorem 1.3 is a special case of Theorem 3.4 in Section 3, where not only solutions to the p-Laplace system with 2pn are considered, but weights ν which need not be monotone are also allowed.

Example 1.5

Assume that n3, and fΛ1((1+log1t)n-1ω(t))(Ω,N) for some slowly varying non-increasing function ω:(0,1)(0,) in the sense of Karamata, namely, such that limt0+ω(λt)ω(t)=1 for every λ>0. As a consequence of [5, Theorems 1.5.11 and 1.6.1], the function ω is slowly varying if and only if

(1.3)0t(1+log1s)n-1ω(s)𝑑s=t(1+log1t)n-1ω(t)+higher order termsas t0+.

Note that, in particular, (1.3) holds if ω is locally absolutely continuous and the (essential) limit of tω(t)ω(t) as t0+ equals 0.

By Theorem 1.3 and Remark 1.4, any local solution u to (1.1) is continuous if and only if

0dttlog1tω(t)1n-2<.

For instance, the choice

(1.4)ω(t)=2(1t)n-2k-1(1t)n-2k(1t)n-2+εfor t(0,1)

and for any k2 and ε>0 is admissible, where k is inductively defined for s>0 as

(1.5)1(s)=max{1,logs},k(s)=max{1,log(k-1(s))}if k2.

On the other hand, discontinuous (in fact, locally unbounded) solutions to (1.1) may exist if ω is defined as in (1.4) with ε=0. This recovers, and extends to systems, a result of [21].

Let us emphasize that discontinuous solutions may exist when ε=0 even if an infinite product extended to all k2 appears on the right-hand side of (1.4), namely, if

ω(t)=k=2k(1t)n-2for t>0.

Indeed, such a function ω can be shown to fulfil (1.3), whereas (1.2) fails. This last example is closely related to a question raised in [21, Remark 5.2] about the case when the right-hand side of (1.1) belongs to an Orlicz space associated with a Young function defined in terms of an infinite product of logarithms.

Let us now turn to the question of the modulus of continuity of solutions to (1.1). Let φ:(0,)(0,) be a function equivalent near 0 (up to multiplicative constants) to a non-decreasing function. We denote by C0,φ(Ω,N) the space of all functions 𝒖 for which the semi-norm

𝒖C0,φ(Ω,N)=supx,yΩxy|𝒖(x)-𝒖(y)|φ(|x-y|)

is finite. Clearly, functions φ which are equivalent (up to multiplicative constants) near 0 yield the same spaces (up to equivalent semi-norms). The space C0,φ(Ω,N) is nontrivial if and only if lim sups0+sφ(s)<. Furthermore, it consists of uniformly continuous functions in Ω, with modulus of continuity not exceeding φ, provided that lims0+φ(s)=0.

The Hölder space C0,α(Ω,N) with α(0,1] agrees with C0,φ(Ω,N) with the choice φ(s)=sα. In particular, C0,1(Ω,N)=Lip(Ω,N), the space of Lipschitz continuous functions in Ω. We denote by Cloc0,φ(Ω,N) the space of those functions which belong to C0,φ(Ω,N) for every open set ΩΩ.

Let us emphasize that no uniform modulus of continuity, depending only on n, is guaranteed for all solutions 𝒖 to (1.1) under the mere assumption that 𝒇L(1,1n-1)(Ω,N), see Proposition 3.7 in Section 3. However, if right-hand sides 𝒇 from function spaces essentially smaller than L(1,1n-1)(Ω,N) are considered, every solution to (1.1) does admit a modulus of continuity depending only on the relevant function space. The following result is paradigmatic in this connection, in that it involves certain borderline Orlicz spaces of Zygmund type and yields the best possible modulus of continuity. In particular, it improves earlier results for Orlicz spaces from the same family, namely, Orlicz spaces of the form L(logL)α(Ω,N), associated with a Young function equivalent to t1(t)α near infinity. In what follows, we denote by BR a ball in n of radius R>0. The notation BR(x) will be used to specify that the ball is centered at the point xn.

Theorem 1.6

Let 𝐮 be any local solution to (1.1) with 𝐟L(logL)n-1+ε(Ω,RN). Then, 𝐮Cloc0,φ(Ω,RN), where

φ(s)=(log1s)-εn-1near 0,

and there exists a constant C such that

(1.6)𝒖C0,φ(BR,N)C(𝒇L(logL)n-1+ε(B2R,N)1n-1+𝒖L1(B2R,N×n))

for every ball B2RΩ.

The modulus of continuity φ is optimal, in the sense that if (1.6) holds with φ replaced by some other modulus of continuity ψ for every 𝐟L(logL)n-1+ε(Ω,RN) and any local solution 𝐮 to (1.1), then Cloc0,φ(Ω,RN)Cloc0,ψ(Ω,RN).

Theorem 1.6 augments [20, Theorem 1.1], where just the scalar case was considered and the weaker conclusion uCloc0,ψ(Ω) with

ψ(s)=(log1s)-εnnear 0

was established under the same assumption that fL(logL)n-1+ε(Ω). In fact, one can more generally show that if 𝒇L(1L)n-1(2L)n-2(k-1L)n-2(kL)n-2+ε(Ω,N), the Orlicz space built upon a Young function equivalent to t1(t)n-12(t)n-2k-1(t)n-2k(t)n-2+ε near infinity, then 𝒖Cloc0,φ(Ω,N), where

φ(t)=k(1t)-εn-1near 0.

This follows as a special case of Theorem 3.8 in Section 3, via (2.5) and (2.11).

After recalling the necessary background from the theory of function spaces in the next section, our main results for the p-Laplace system for any p[2,n] are stated in Section 3. Specifically, Section 3.1 deals with plain continuity, whereas uniform continuity is the subject of Section 3.2. Proofs of these two sets of results are presented in Sections 4 and 5, respectively. Our approach makes substantial use of precise pointwise estimates for the gradient of solutions to p-Laplacian type elliptic systems which are established in [25] (and refine earlier estimates from [27, 14, 15]), of optimal Sobolev type embeddings into spaces of uniformly continuous functions from [11], and of various Hardy type inequalities and related embeddings between spaces of measurable functions appearing in diverse contributions, including [6, 7, 8, 9, 17, 19, 26].

We note that part of the results of the present paper were announced in [2].

2 Spaces of measurable functions

Let Ω be a measurable subset of n having finite Lebesgue measure, which, without loss of generality, will be assumed to be equal to 1, and let g:Ω be a measurable function. The decreasing rearrangement of g is the function g*:[0,)[0,] defined as

g*(s)=inf{t0:|{xΩ:|g(x)|>t}|s}for s0.

In other words, g* is the (unique) non-increasing, right-continuous function in [0,) equidistributed with g. By g** we denote the maximal type function associated with g* as recalled in Section 1. Notice that

(2.1)g*(s)g**(s)for s>0

and for every measurable function g, since g* is non-increasing.

A quasi-normed function spaceX(Ω) on Ω is a linear space of measurable functions on Ω equipped with a quasi-norm X(Ω) satisfying the following properties.

  1. gX(Ω)>0 if g0;λgX(Ω)=|λ|gX(Ω) for every λ and gX(Ω);g+hX(Ω)c(gX(Ω)+hX(Ω)) for some constant c1 and for every g,hX(Ω).

  2. 0|h||g| a.e. in Ω implies hX(Ω)gX(Ω).

  3. 0gkg a.e. implies gkX(Ω)gX(Ω) as k.

  4. 1X(Ω)<.

  5. There exists a constant C such that Ω|g|𝑑xCgX(Ω) for every gX(Ω).

Given a measurable set GΩ, we define

gX(G)=gχGX(Ω)

for every measurable function g on Ω. Here, χG stands for the characteristic function of G. Moreover, we denote by Xloc(Ω) the space of measurable functions g on Ω such that gX(G)< for every compact set GΩ.

If assumption (i) holds with c=1, then the functional X(Ω) is a norm which makes X(Ω) a Banach space. In this case, X(Ω) will be called a Banach function space.

Given a Banach function space X(Ω) and a quasi-normed space Y(Ω), we have that

(2.2)X(Ω)Y(Ω)if and only ifX(Ω)Y(Ω),

where the arrow “” means that there exists a constant C such that gY(Ω)CgX(Ω) for every gX(Ω). Let us notice that property (2.2) is usually stated under the assumption that both X(Ω) and Y(Ω) are Banach function spaces (see [4, Chapter 1, Theorem 1.8]). An inspection of the proof reveals that it continues to hold even if Y(Ω) is just a quasi-normed space.

A quasi-normed function space (in particular, a Banach function space) X(Ω) is called rearrangement-invariant if

gX(Ω)=hX(Ω)wheneverg*=h*.

If this is the case, the (quasi)-norm X(Ω) is also said to be rearrangement-invariant.

Given a nonnegative functional X(Ω) defined on the space of measurable functions (not necessarily a norm nor a quasi-norm), its associate functional X(Ω) is given by

hX(Ω)=supg0Ω|g(x)h(x)|𝑑xgX(Ω).

If X(Ω) is a rearrangement-invariant norm, then X(Ω) is also a rearrangement-invariant norm.

Let σ,ρ(0,) and let ν be a weight on (0,1). The classical Lorentz spacesΛσ(ν)(Ω) and Γσ(ν)(Ω), and the weak Lorentz spacesΛρ,(ν)(Ω) and Γρ,(ν)(Ω), are the families of those measurable functions g in Ω such that the corresponding functional from

(2.3){gΛσ(ν)(Ω)=(01(g*(t))σν(t)𝑑t)1σ,gΛρ,(ν)(Ω)=sup0<t<1g*(t)𝒱(t)1ρ,gΓσ(ν)(Ω)=(01(g**(t))σν(t)𝑑t)1σ,gΓρ,(ν)(Ω)=sup0<t<1g**(t)𝒱(t)1ρ

is finite. Here, 𝒱:(0,1)[0,) denotes the function defined by

(2.4)𝒱(t)=0tν(s)𝑑sfor t(0,1).

Observe that, owing to (2.1), any functional of type Γ is not weaker than the corresponding functional of type Λ associated with the same weight and powers. Inequalities among these functionals have been extensively studied in the literature. In particular, it is easily seen, via Fubini’s theorem, that

(2.5)Γ1(ν)(Ω)=Λ1(ν~)(Ω),

where ν~(t)=t1ν(s)s𝑑s for t(0,1). Moreover, if ρ(0,), then

{Λρ,(ν)(Ω)=Λ1,(νρ)(Ω),Γρ,(ν)(Ω)=Γ1,(νρ)(Ω),

where νρ(t)=1ρ𝒱(t)1ρ-1ν(t) for t(0,1). The quantities defined in (2.3) are not always norms. For example, if σ1, then Λσ(ν)(Ω) is a norm if and only if ν is non-increasing. If σ(1,), then Λσ(ν)(Ω) is equivalent to a norm if and only if

(2.6)tσt1s-σν(s)𝑑sC0tν(s)𝑑sfor t(0,1)

and for some constant C (see [29, Theorem 4]). A necessary and sufficient condition for Λ1(ν)(Ω) to be equivalent to a norm, and hence for Λ1(ν)(Ω) to be equivalent to a Banach space, is that

(2.7)𝒱(t)tC𝒱(s)sfor 0<st1

and for some constant C (see [6, Theorem 2.3]).

Given ρ(0,), the quasi-norm Λρ,(ν)(Ω) is equivalent to a norm if

1t0tds𝒱(s)1ρC𝒱(t)1ρfor t(0,1)

and for some constant C. The space Γσ(ν)(Ω) is a Banach space if σ1. The space Γρ,(ν)(Ω) is a Banach space for every ρ(0,).

For technical reasons, Lorentz type functionals associated with weights dμ associated with a Borel measure μ, which are not necessarily absolutely continuous with respect to the Lebesgue measure, will come into play in some of our proofs. Their definition parallels (2.3) and reads

{gΛσ(dμ)(Ω)=(01(g*(t))σ)dμ(t))1σ,gΛρ,(dμ)(Ω)=sup0<t<1g*(t)(1t0t𝑑μ(s))1ρ,gΓσ(dμ)(Ω)=(01(g**(t))σ𝑑μ(t))1σ,gΓρ,(dμ)(Ω)=sup0<t<1g**(t)(1t0t𝑑μ(s))1ρ

for a measurable function g in Ω.

Standard instances of classical Lorentz spaces are the customary Lorentz spaces L(r,q)(Ω) with r,q(0,], and, more generally, the Lorentz–Zygmund spacesL(r,q)(logL)α(Ω). Given r,q(0,] and α, they are defined in terms of the quasi-norm

(2.8)gL(r,q)(logL)α(Ω)=s1r-1q(1+log1s)αg**(s)Lq(0,1)

for a measurable function g in Ω. The space L(r,q)(Ω) is defined as L(r,q)(logL)0(Ω). If r,q1 and α, the space L(r,q)(logL)α(Ω) is a rearrangement-invariant space endowed with the norm (2.8). The quantity obtained on replacing g** by g* in (2.8) will be denoted by gLr,q(logL)α(Ω). Accordingly, we set gLr,q(Ω)=gLr,q(logL)0(Ω). If r>1, q(0,), and α, then

(2.9)gL(r,q)(logL)α(Ω)gLr,q(logL)α(Ω)

for every measurable function g in Ω. The notation “” means that the two sides are bounded by each other up to multiplicative constants, which in (2.9) depend on r,q,α. In particular, if r>1, then L(r,r)(logL)α(Ω)=Lr,r(logL)α(Ω)) (up to equivalent norms). On the contrary, L(1,1)(Ω)L1(Ω). Actually, L(1,1)(Ω) is equivalent to L1,1(logL)1(Ω), whereas L1(Ω)=L1,1(logL)0(Ω)=L(1,)(Ω).

Observe that L(r,q)(logL)α(Ω)=Γq(ν)(Ω) if ν(t)=tqr-1(1+log1t)αq and r,q(0,). Furthermore, L(r,)(logL)α(Ω)=Γr,(ν)(Ω) if ν(t)=t(1+log1t)αr and r(0,). Similarly, Lr,q(logL)α(Ω)=Λq(ν)(Ω) and Lr,(logL)α(Ω)=Λr,(ν)(Ω) with the same choices of the weight ν.

A function A:[0,)[0,] is called a Young function if it has the form

(2.10)A(t)=0ta(τ)𝑑τfor t0

and for some non-decreasing, left-continuous function a:[0,)[0,] which is neither identically equal to 0 nor to . Clearly, any convex (nontrivial) function from [0,) into [0,], which is left-continuous and vanishes at 0, is a Young function.

The Orlicz spaceLA(Ω), associated with a Young function A, is the Banach function space of those measurable functions g in Ω for which the Luxemburg norm

gLA(Ω)=inf{λ>0:ΩA(|g(x)|λ)𝑑x1}

is finite.

The Orlicz spaces are rearrangement-invariant spaces, since

gLA(Ω)=g*LA(0,1)

for every measurable function g in Ω. We shall also need to make use of the rearrangement-invariant space L(A)(Ω), associated with a Young function A, consisting of those measurable functions g in Ω such that the norm

gL(A)(Ω)=g**LA(Ω)

is finite.

A function A of the form (2.10) for some function a which is just nonnegative and locally bounded in (0,) (but not necessarily non-decreasing) will be called a generalized Young function. Clearly, a generalized Young function is not necessarily convex. In particular, it may have sublinear growth near infinity.

If A is just a generalized Young function, the set of all measurable functions g in Ω such that the quantity

|g|LA(Ω)=inf{k>0:ΩA(|g(x)|k)𝑑xk}

is finite will be still denoted by LA(Ω) and is a complete Frechét space, in the sense that ||LA(Ω) satisfies the triangle inequality, |g|LA(Ω)=0 if and only if g=0, and it is a complete metric space endowed with the metric induced by ||LA(Ω). One also has that |λg|LA(Ω)0 if λ0. Let us point out that functions from a generalized Orlicz space need not be even locally integrable. Classical examples of generalized Orlicz spaces are the spaces Lp(Ω) with 0<p<1.

The functional ||LA(Ω) is rearrangement-invariant, since

|g|LA(Ω)=|g*|LA(0,1)

for every measurable function g in Ω.

The Frechét space L(A)(Ω), associated with a generalized Young function A, consists of those measurable functions g in Ω such that the functional

gL(A)(Ω)=|g**|LA(Ω)

is finite.

A generalized Young function A is said to dominate another generalized Young function B near infinity if there exist constants C>0 and t00 such that

B(t)A(Ct)for tt0.

The functions A and B are called equivalent near infinity if they dominate each other near infinity. If A dominates B near infinity, then LA(Ω)LB(Ω). Hence, LA(Ω)=LB(Ω) if A and B are equivalent near infinity. Analogous inclusion relations hold for the spaces L(A)(Ω) and L(B)(Ω).

Orlicz spaces associated with a (generalized) Young function equivalent to the function tp1(t)α12(t)α2k(t)αk near infinity, where k is defined as in (1.5), will be denoted by Lp(1L)α1(2L)α2(kL)αk(Ω). The expression 1L will also simply be denoted by logL. One can show that

L(q,q)(logL)α(Ω)=Lq,q(logL)α(Ω)=Lq(logL)αq(Ω),

up to equivalent norms, if q>1 and α. On the other hand, if α0,

L1,1(logL)α(Ω)=L1(logL)α(Ω),

but

L(1,1)(logL)α(Ω)=L1(logL)α+1(Ω),

up to equivalent norms. More generally, analogous relations hold among the Orlicz spaces Lp(1L)α1(2L)α2(kL)αk(Ω) and the Lorenz spaces Λq(ν) and Γq(ν) associated with weights ν of power-multiple-logarithmic type. For instance,

(2.11)Γ1(ν)(Ω)=L(1L)α1+1(2L)α2(kL)αk(Ω),

up to equivalent norms, if ν(s)=1(1s)α12(1s)α2k(1s)αk.

Spaces X(Ω,N) of vector-valued functions 𝒈:ΩN are defined analogously, on replacing g with |𝒈| in the definitions given above. In particular, a space X(Ω,N) will be called rearrangement-invariant if

𝒈X(Ω,N)=𝒉X(Ω,N)whenever|𝒈|*=|𝒉|*.

Spaces X(Ω,N×n) of matrix-valued functions from Ω into N×n are defined accordingly.

3 Main results

A generalized notion of solutions to the p-Laplace system will be adopted, which also applies to the case when the right-hand side is merely an integrable function. In the spirit of [13, 12, 25], the solutions to be considered are required to be approximable, in a suitable sense, by a sequence of solutions to the p-Laplace system corresponding to a sequence of smooth right-hand sides converging to the original one. If the latter enjoys a sufficiently high degree of integrability, then an approximable solution turns out to be a weak solution in the usual sense.

Specifically, let 𝒇L1(Ω,N) and p2. A function 𝒖 in the Sobolev space W1,p-1(Ω,N) is called a local approximable solution to the system

(3.1)-div(|𝒖|p-2𝒖)=𝒇in Ω

if there exists a sequence of smooth functions {𝒇k} and a sequence {𝒖k} of local weak solutions to the systems

-div(|𝒖k|p-2𝒖k)=𝒇kin Ω

such that 𝒇k𝒇 in Lloc1(Ω,N) and 𝒖k𝒖 in Wloc1,p-1(Ω,N) as k.

In particular, any local approximable solution satisfies

Ω|𝒖|p-2𝒖ϕdx=Ω𝒇ϕ𝑑x

for every ϕC0(Ω,N).

3.1 Continuity

The key result of this section is a description of the optimal (largest) rearrangement-invariant quasi-normed space X(Ω,N) such that membership of 𝒇 to X(Ω,N) ensures the continuity of any local solution to (3.1).

Theorem 3.1

Theorem 3.1 (Optimal space for continuity)

Let 2pn. Then, L(np,1p-1)(Ω,RN) is the largest rearrangement-invariant quasi-normed space X(Ω,RN) such that any local solution 𝐮 to (3.1) is continuous for every 𝐟X(Ω,RN).

In the case of a single equation, namely, when N=1, the continuity of any solution to (3.1) under the assumption that fL(np,1p-1)(Ω) was established in [16] for p<n and, as mentioned above, in [21] for p=n. The special case when p=n=2 goes back to [3]. Theorem 3.1 provides us with the optimality of the space L(np,1p-1)(Ω,N) and tells us that the same conclusion holds for systems as well.

Theorem 3.1 can be exploited, in combination with the characterization of embeddings between function spaces, to provide sharp conditions on the right-hand side 𝒇 in (3.1), in classes of function spaces, for the continuity of solutions. Data 𝒇 in classical Lorentz spaces of type Γσ(ν)(Ω,N) are considered in the following result.

Theorem 3.2

Theorem 3.2 (Continuity with fΓσ(ν)(Ω,RN))

Assume that 2pn. Let 0<σ< and let ν be a weight. Let Vσ:(0,1)[0,) be defined as

Vσ(t)=0tν(s)𝑑s+tσt1s-σν(s)𝑑sfor t(0,1).

Assume that

(3.2){lim supt0+tpnVσ(t)1σ<if 0<σ1p-1, 2p<n,lim supt0+t(log1t)n-1Vσ(t)1σ<if 0<σ1n-1, 2p=n,0tpσn[σ(p-1)-1]-1Vσ(t)1σ(n-1)-1𝑑t<if 1p-1<σ1, 2<pn, or σ>1, 2p<n,0t(log1t)1σ-1Vσ(t)1σ-1𝑑t<if σ>1, 2=p=n,0tσσ(n-1)-1-1(log1t)1σ(n-1)-1Vσ(t)1σ(n-1)-1𝑑t<if σ1, 2<p=n.

If 𝐟Γσ(ν)(Ω,RN), then every local solution 𝐮 to (3.1) is continuous.

Conversely, if σ1 and for every 𝐟Γσ(ν)(Ω,RN) any local solution to (3.1) is continuous, then one of the alternatives in (3.2) holds.

Right-hand sides 𝒇 in weak type spaces Γρ,(ν)(Ω,N) are the subject of the next theorem.

Theorem 3.3

Theorem 3.3 (Continuity with fΓρ,(ν)(Ω,RN))

Assume that 2pn. Let 0<ρ< and let ν be a weight. Let V be defined as in (2.4). Assume that

(3.3)0t1p-1(pn-1)-1supst(s-1𝒱(s)1ρ)𝑑t<.

If 𝐟Γρ,(ν)(Ω,RN), then every local solution 𝐮 to (3.1) is continuous.

Conversely, if ρ1 and for every 𝐟Γρ,(ν)(Ω,RN) any local solution to (3.1) is continuous, then (3.3) holds.

In Theorems 3.4 and 3.5 below, functions 𝒇 from the Lorentz spaces Λσ(ν)(Ω,N) and the weak Lorentz spaces Λρ,(ν)(Ω,N), respectively, are taken into account.

Theorem 3.4

Theorem 3.4 (Continuity with fΛσ(ν)(Ω,RN))

Assume that 2pn. Let 0<σ< and let ν be a weight. Let V be defined as in (2.4). Assume that

(3.4){lim supt0+tpn𝒱(t)1σ<if 0<σ1p-1, 2p<n,lim supt0+t(log1t)n-1𝒱(t)1σ<if 0<σ1n-1, 2p=n,0tpσn(σ(p-1)-1)𝒱(t)1σ(p-1)-1dtt<if σ>1p-1, 2p<n,0sup0<st(sσσ(n-1)-1𝒱(s)1σ(n-1)-1)(log1t)1σ(n-1)-1dtt<if 1n-1<σ1, 2<p=n,0(0t(s𝒱(s))1σ-1𝑑s)σ-1σ(n-1)-1(log1t)1σ(n-1)-1dtt<if σ>1, 2p=n.

If 𝐟Λσ(ν)(Ω,RN), then every local solution 𝐮 to (3.1) is continuous.

Conversely, if either σ>1 and (2.6) holds or σ=1 and (2.7) holds, and for every 𝐟Λσ(ν)(Ω,RN) any local solution to (3.1) is continuous, then one of the alternatives in (3.4) holds.

Theorem 3.5

Theorem 3.5 (Continuity with fΛρ,(ν)(Ω,RN))

Assume that 2pn. Let 0<ρ< and let ν be a weight. Let V be defined as in (2.4). Assume that

(3.5)0(1t0t𝒱(s)-1ρ𝑑s)1p-1tpn(p-1)-1𝑑t<.

If 𝐟Λρ,(ν)(Ω,RN), then every local solution 𝐮 to (3.1) is continuous.

Conversely, if (2.6) holds and for every 𝐟Λρ,(ν)(Ω,RN) any local solution to (3.1) is continuous, then (3.5) holds.

We conclude this subsection with the case when 𝒇 belongs to an Orlicz space.

Theorem 3.6

Theorem 3.6 (Continuity with fLA(Ω,RN))

Assume that 2pn. Let A be a Young function. Assume that

{(tA(t))pnp-n-pdttn(p-2)np-n-p<if 2p<n,(ttA(t)-A(t))1n-2dtt<if 2<p=n,lim supttlog(1+t)A(t)<if 2=p=n.

If 𝐟LA(Ω,RN), then any local solution 𝐮 to (3.1) is continuous.

3.2 Modulus of continuity

This subsection is devoted to the modulus of continuity of solutions to (3.1). Loosely speaking, the results to be presented tell us that if a sufficiently strong rearrangement-invariant (quasi)-norm of 𝒇 is finite, then the modulus of continuity of any solution 𝒖 to (3.1) is locally bounded by the relevant norm of 𝒇 and the L1 norm of 𝒖.

In the following proposition, we preliminarily note that no uniform estimate of this kind holds for the modulus of continuity of solutions 𝒖 to the p-Laplace system (3.1) (even for N=1) when 𝒇 just belongs to the space L(np,1p-1)(Ω,N). Recall from Theorem 3.1 that such a space is the largest ensuring the mere continuity of 𝒖.

Proposition 3.7

Proposition 3.7 (Failure of uniform continuity for fL(np,1p-1)(Ω,RN))

Let 2pn. Assume that there exist a non-decreasing function φ and a constant C such that

(3.6)𝒖C0,φ(BR,N)C(𝒇L(np,1p-1)(B2R,N)1p-1+𝒖L1(B2R,N×n))

for every ball B2RΩ, every 𝐟Lloc(np,1p-1)(Ω,RN), and every local solution 𝐮 to (3.1). Then, lims0+φ(s)>0.

Although our approach applies, in principle, to a variety of rearrangement-invariant spaces, we focus on the case when 𝒇 belongs to Lorentz spaces of type Γσ(ν)(Ω,N) or Γσ,(ν)(Ω,N). These include, as special instances, most customary function spaces, such as the usual Lorentz spaces L(r,q)(Ω,N) and certain borderline Orlicz spaces of Zygmund type which are of crucial use in the analysis of the n-Laplace system, as already shown in Theorem 1.1 in Section 1. The modulus of continuity established in these instances can also be shown to be sharp in most situations, see Proposition 3.12 below.

In what follows, given a number σ[1,], we set σ=σσ-1, the Hölder conjugate of σ, with the usual conventions if either σ=1 or σ=.

Theorem 3.8

Theorem 3.8 (Modulus of continuity with fΓσ(ν)(Ω,RN))

Assume that 2pn. Let 𝐟Γσ(ν)(Ω,RN) for some σ[1,). Let 𝐮 be a local solution to (3.1).

Part I. Assume that the function φ, defined as

(3.7)φ(s)={(0snt(σ(p-1))n-1(t1τ-σnν(τ)-1σ-1𝑑τ)σ-1σ(p-1)-1𝑑t)1(σ(p-1))if σ>1, 2pn,(0snt(p-1)n-11(inft<τ<1(τ1nν(τ)))(p-1)-1𝑑t)1(p-1)if σ=1, 2<pn,sup0<t<1min{t12,s}inft<τ<1(τ12ν(τ))if σ=1, 2=p=n,

for s near 0, is finite and lims0+φ(s)=0 (of course, the latter assumption is relevant only in the last case of (3.7)). Then, 𝐮Cloc0,φ(Ω,RN) and there exists a constant C such that

𝒖C0,φ(BR,N)C(𝒇Γσ(ν)(B2R,N)1p-1+𝒖L1(B2R,N×n))

for every ball B2RΩ.

Part II. In particular, if

(3.8)0t-σnν(t)-1σ-1𝑑t<,

then 𝐮Liploc(Ω,RN) and there exists a constant C such that

𝒖Lip(BR,N)C(𝒇Γσ(ν)(B2R,N)1p-1+𝒖L1(B2R,N×n))

for every ball B2RΩ.

Theorem 3.9

Theorem 3.9 (Modulus of continuity with fΓρ,(ν)(Ω,RN))

Assume that 2pn. Let 𝐟Γρ,(ν)(Ω,RN) for some ρ(0,). Let 𝐮 be a local solution to (3.1). Assume that the function φ, defined as

φ(s)=0snt-1n(t1τ-1n(0τν(ϱ)𝑑ϱ)-1ρ𝑑τ)1p-1𝑑t

for s near 0, is finite. Then, 𝐮Cloc0,φ(Ω,RN) and there exists a constant C such that

𝒖C0,φ(BR,N)C(𝒇Γρ,(ν)(B2R,N)1p-1+𝒖L1(B2R,N×n))

for every ball B2RΩ.

Specializing the previous theorems to the case when the weight ν is a power times a logarithm yields the following result.

Theorem 3.10

Theorem 3.10 (Modulus  of  continuity  with  fL(r,q)(logL)α(Ω,RN))

Assume that 2pn and also let 𝐟L(r,q)(logL)α(Ω,RN) with nprn, 1q, and αR. Let 𝐮 be a local solution to (3.1).

Part I. Let φ be the function defined as

(3.9)φ(s)={(log1s)-α+p-1-1qp-1if r=np, 1q,α>p-1-1q, 2pn,srp-nr(p-1)(log1s)-αp-1if np<r<n, 1q,α, 2pn,s(log1s)-α+1-1qp-1if r=n, 1q,α<1q, 2pn,s(log(log1s))1-1qp-1if r=n, 1<q,α=1q, 2pn,

for s near 0. Then, 𝐮Cloc0,φ(Ω,RN) and there exists a constant C such that

(3.10)𝒖C0,φ(BR,N)C(𝒇L(r,q)(logL)α(B2R,N)1p-1+𝒖L1(B2R,N×n))

for every ball B2RΩ.

Part II. In particular, if one of the conditions

{r=n,q=1,α0, 2pn;r=n, 1<q,α>1q, 2pn;r>n, 1q,α, 2pn,

is satisfied, then 𝐮Liploc(Ω,RN) and there exists a constant C such that

𝒖Lip(BR,N)C(𝒇L(r,q)(logL)α(B2R,N)1p-1+𝒖L1(B2R,N×n))

for every ball B2RΩ.

Example 3.11

Theorem 3.8 allows one to deal also with data 𝒇 from unconventional functions spaces. Suppose, for instance, that p=n and 𝒇Γ1(ν)(Ω,N), where ν(t)=eγ(log1t)δ(log1t)δ-1 near 0 for some δ(0,1) and γ>0. The behavior near 0 of this weight is intermediate between a logarithm and a power. If 𝒖 is any local solution to (3.1), then 𝒖Cloc0,φ(Ω,N), where φ(s)=e-γn-1(log1s)δ(log1s)1-δn-1 near 0.

The result of Theorem 3.10 concerning the space Lr,(Ω) with np<r<n and N=1 overlaps with [23, Corollary 8.11] and [31, Theorem 4.2]. In fact, we are also in a position to prove the sharpness of the modulus of continuity provided by Theorem 3.10. This is a special instance of our last result.

Proposition 3.12

Proposition 3.12 (Sharpness of moduli of continuity)

Let 𝐟L(r,q)(logL)α(Ω,RN). Assume that either of the alternatives

(3.11){r=np, 1q,α>p-1-1q, 2pn,n>2;r=1, 1<q,α>1-1q,n=p=2;np<r<n, 1q,α, 2pn;r=n,q=1,α<0, 2pn,

holds. Then, the modulus of continuity φ, defined as in (3.9), is optimal, in the sense that if ψ is another modulus of continuity such that (3.10) holds for every 𝐟L(r,q)(logL)α(Ω,RN) and every local solution 𝐮 to (3.1), then Cloc0,φ(Ω,RN)Cloc0,ψ(Ω,RN).

4 Proofs of the results of Section 3.1

Our point of departure is an estimate, in rearrangement form, for the gradient of local solutions to (3.1). Such an estimate in turn relies upon a pointwise gradient bound from [25].

Proposition 4.1

Assume that 2pn. Let 𝐮 be a local solution to (3.1) for some 𝐟L1(Ω,RN). Assume that |Ω|=1 and Ω is any open set such that ΩΩ. Then, there exists a constant C=C(n,N,p,Ω,Ω) such that

(|𝒖|Ω)*(s)p-1(|𝒖|Ω)**(s)p-11s0s(|𝒖|Ω)*(t)p-1𝑑t
(4.1)Cs1|𝒇|**(t)t-1n𝑑t+C𝒖L1(Ω,N×n)p-1for s(0,1).

Proof.

By [25, Theorem 1.1], there exists a constant C=C(n,N,p) such that

(4.2)|𝒖|(x)p-1C0RBϱ(x)|𝒇(y)|𝑑ydϱϱn+C(BR(x)|𝒖|𝑑y)p-1

for a.e. xΩ and every R>0 such that BR(x)Ω. Fubini’s theorem entails that

0RBϱ(x)|𝒇(y)|𝑑ydϱϱn0Bϱ(x)|𝒇(y)|χBR(x)(y)𝑑ydϱϱn
=n|𝒇(y)|χBR(x)(y)0χBϱ(x)(y)dϱϱn𝑑y
(4.3)=1n-1n|𝒇(y)||x-y|n-1χBR(x)(y)𝑑y

for any such x and R. Owing to (4.2) and (4.3),

(4.4)|𝒖|(x)p-1Cn|𝒇(y)||x-y|n-1χBR(x)(y)𝑑y+C(BR(x)|𝒖|𝑑y)p-1.

Owing to O’Neil’s rearrangement inequality for convolutions [28], we infer from (4.4) that there exist constants C, C, and C′′, depending on n,N,p,Ω, and Ω, such that

1s0s(|𝒖|Ω)*(t)p-1𝑑tC(n|𝒇(y)||-y|n-1χBR(x)(y)𝑑y)*(s)+C𝒖L1(Ω,N×n)p-1
Cs|𝒇|**(s)(1|-y|n-1)**(s)+Cs|𝒇|*(t)(1|-y|n-1)*(t)𝑑t+C𝒖L1(Ω,N×n)p-1
Cs-1n0s|𝒇|*(t)𝑑t+Cs1|𝒇|*(t)t-1n𝑑t+C𝒖L1(Ω,N×n)p-1
(4.5)Cs1′′|𝒇|**(t)t-1n𝑑t+C𝒖L1(Ω,N×n)p-1for s(0,1).

This establishes the last inequality in (4.1). The second inequality is a consequence of Hölder’s inequality, since p2. The first inequality holds by (2.1). The proof is complete. ∎

We are now in a position to prove our first main result.

Proof of Theorem 3.1.

We claim that, if ΩΩ, then there exists a constant C such that

(4.6)𝒖Ln,1(Ω,N×n)C𝒇L(np,1p-1)(Ω,N)1p-1+C𝒖L1(Ω,N×n),

namely,

(4.7)01s-1+1n(|𝒖|Ω)*(s)𝑑sC01(spn|𝒇|**(s))1p-1dss+C𝒖L1(Ω,N×n).

By (4.1), inequality (4.7) will follow if we show that

(4.8)01t-1+1n(t1|𝒇|**(s)s-1n𝑑s)1p-1𝑑tC01(tpn|𝒇|**(t))1p-1dtt

for some constant C and for every measurable function 𝒇 in Ω.

Inequality (4.8) is in turn a consequence of [9, Proposition 2.5]. Having (4.6) at disposal, the continuity of 𝒖 follows, since, by a result of [30], any weakly differentiable function 𝒖 with |𝒖|Llocn,1(Ω,N) is continuous.

In order to prove the optimality of the space L(np,1p-1)(Ω,N), assume, without loss of generality, that N=1, 0Ω, and let R>0 be such that BR(0)Ω. Denote by d the diameter of Ω. Let X(Ω) be a rearrangement-invariant quasi-normed space. Given any nonnegative function hX(Ω) which is radially decreasing about 0 and vanishes outside BR(0), the (radially decreasing) function v:Ω defined as

v(x)=(nωn1n)-pωn|x|nωndn(spnh**(s))1p-1dssfor xΩ

is a local solution to the equation

(4.9)-div(|v|p-2v))=h(x)in Ω.

One has that

(4.10)vL(BR(0))v(0)=chL(np,1p-1)(BR(0))1p-1

for a suitable constant c. If, by contradiction, X(Ω)L(np,1p-1)(Ω), then a function h as above can be chosen in such a way that hX(Ω)L(np,1p-1)(Ω), and hence

hL(np,1p-1)(BR(0))1p-1=.

By (4.10), vLloc(Ω), and, a fortiori, v is not continuous in Ω. ∎

The remaining part of this section is devoted to the proofs of Theorems 3.23.6.

Proof of Theorem 3.2.

By Theorem 3.1, any local solution 𝒖 to (3.1) is continuous provided that

(4.11)Γσ(ν)(Ω,N)L(np,1p-1)(Ω,N).

The embedding (4.11) is equivalent to the inequality

(4.12)(01ϕ**(t)1p-1tpn(p-1)-1𝑑t)p-1C(01ϕ**(t)σν(t)𝑑t)1σ

for every measurable function ϕ on (0,1). The weighted Hardy type inequality for non-increasing functions (4.12) holds if one of the conditions in (3.2) is fulfilled. This follows from [17, Theorem 5.1 (i)] in the first two cases of (3.2) and from [17, Theorem 5.1 (ii)] in the last three cases.

Conversely, assume that σ1 and for every 𝒇Γσ(ν)(Ω,N) any local solution 𝒖 to (3.1) is continuous. Then, by Theorem 3.1,

(4.13)Γσ(ν)(Ω,N)L(np,1p-1)(Ω,N).

Since σ1, Γσ(ν)(Ω,N) is a rearrangement-invariant space, and hence, owing to (2.2), the inclusion (4.13) is equivalent to the embedding (4.11). Thereby, (4.12) holds. The necessity of the conditions of [17, Theorem 5.1] again tell us that either of the conditions in (3.2) must be fulfilled. ∎

Proof of Theorem 3.3.

By Theorem 3.1, any local solution 𝒖 to (3.1) is continuous provided that

(4.14)Γρ,(ν)(Ω,N)L(np,1p-1)(Ω,N)

or, equivalently,

(4.15)(01ϕ**(t)1p-1tpn(p-1)-1𝑑t)p-1Csup0<t<1ϕ**(t)𝒱(t)1ρ

for every measurable function ϕ on (0,1). By [8, Theorem 6.4], (4.15) holds if and only if (3.3) holds.

Conversely, assume that ρ1 and for every 𝒇Γρ,(ν)(Ω,N) any local solution 𝒖 to (3.1) is continuous. Then, by Theorem 3.1, Γρ,(ν)(Ω,N)L(np,1p-1)(Ω,N). Since ρ1, Γρ,(ν)(Ω,N) is a rearrangement-invariant space, and, by (2.2), the last inclusion is in fact equivalent to the embedding (4.14), namely, to (4.15). Owing to (the necessity part of) [8, Theorem 6.4], we conclude that (3.3) must hold. ∎

Proof of Theorem 3.4.

By Theorem 3.1, any local solution 𝒖 to (3.1) is continuous provided that

(4.16)Λσ(ν)(Ω,N)L(np,1p-1)(Ω,N).

If p=n, the embedding (4.16) is equivalent to the inequality

(4.17)(01ϕ**(t)1n-1t1n-1-1𝑑t)n-1C(01ϕ*(t)σν(t)𝑑t)1σ

for every measurable function ϕ on (0,1). On the other hand, if 2p<n, then L(np,1p-1)(Ω,N)=Lnp,1p-1(Ω,N) and the embedding (4.16) is equivalent to

(4.18)(01ϕ*(t)1p-1tpn(p-1)-1𝑑t)p-1C(01ϕ*(t)σν(t)𝑑t)1σ

for every measurable function ϕ on (0,1). We claim that either (4.17) or (4.18) holds if the appropriate condition from (3.4) is satisfied. This can be verified by distinguishing into several cases.

If 0<σ1p-1 and 2p<n, inequality (4.18) is equivalent to the corresponding condition in (3.4) by [8, Theorem 3.1 (i)].

If 0<σ1n-1 and 2p=n, inequality (4.17) is equivalent to the corresponding condition in (3.4) by [8, Theorem 4.1 (ii)].

If σ>1p-1 and 2p<n, inequality (4.18) is equivalent to the corresponding condition in (3.4) by [8, Theorem 3.1 (ii)].

If 1n-1<σ1 and 2p=n, inequality (4.17) is equivalent to the corresponding condition in (3.4) by [7, Theorem 3.1].

If σ>1 and 2=p=n, by [8, Theorem 4.1 (iv)], inequality (4.17) is equivalent to

(4.19)0(t𝒱(t))1σ-1(log1t)σσ-1𝑑t<.

An application of Fubini’s theorem shows that (4.19) is equivalent to the condition on the last line of (3.4).

If σ>1 and 2<p=n, by [8, Theorem 4.1 (iii)], inequality (4.17) holds if and only if

(4.20)(01(tσ0tν(s)𝑑s)1σ(n-1)-1dtt)σ(n-1)-1σ<

and

(4.21)01(0tν(s)sσ(0sν(τ)𝑑τ)σ𝑑s)σ-1σ(n-1)-1(log1t)1σ(n-1)-1dtt<.

An integration by parts in the inner integral on the left-hand side of (4.21) shows that the latter equals

01(σσ-10tsσ-1(0sν(τ)𝑑τ)σ-1𝑑s-tσσ-1(0tν(τ)𝑑τ)1-σ)σ-1σ(n-1)-1(log1t)1σ(n-1)-1dtt.

Thus, the condition appearing on the last line of (3.4), namely,

(4.22)01(0tsσ-1(0sν(τ)𝑑τ)σ-1𝑑s)σ-1σ(n-1)-1(log1t)1σ(n-1)-1dtt<,

implies (4.21). We claim that the reverse implication is also true, and hence (4.21) and (4.22) are actually equivalent. To verify this claim, it suffices to show that there exists a positive constant C such that

01(σσ-10tsσ-1(0sν(τ)𝑑τ)σ-1𝑑s-tσσ-1(0tν(τ)𝑑τ)1-σ)σ-1σ(n-1)-1(log1t)1σ(n-1)-1dtt
(4.23)C01(0tsσ-1(0sν(τ)𝑑τ)σ-1𝑑s)σ-1σ(n-1)-1(log1t)1σ(n-1)-1dtt

for every continuous function ν:(0,1)[0,). On setting

ψ*(s)=(0s1σν(t)𝑑t)1-σandw(s)=sσ-1σ(n-1)-1-1(log1s)1σ(n-1)-1

for s(0,1) and changing variables, (4.23) takes the form

(4.24)01(ψ**(s)-ψ*(s))σ-1σ(n-1)-1w(s)𝑑sC01ψ**(s)σ-1σ(n-1)-1w(s)𝑑s

for some positive constant C. Inequality (4.24), and hence (4.23), follows from [7, Theorem 6.2 (ii)]. The equivalence of (4.22) and (4.21) is fully proved.

Next, observe that, since

01ψ**(s)σ-1σ(n-1)-1w(s)𝑑s01ψ*(s)σ-1σ(n-1)-1w(s)𝑑s,

we have that

(4.25)01(0tsσ-1(0sν(τ)𝑑τ)σ-1ds)σ-1σ(n-1)-1(log1t)1σ(n-1)-1dttC01(tσ(0tν(s)𝑑s))1σ(n-1)-1(log1t)1σ(n-1)-1dtt

for some positive constant C. Thus, (4.22) implies that the right-hand side of (4.25) is finite. On the other hand, the latter condition trivially implies (4.20). Altogether, we have that (4.22) implies both (4.20) and (4.21), and hence (4.17). This concludes the proof of (4.16) and the proof of the embedding is thus complete in all possible cases.

Conversely, assume now that either σ>1 and (2.6) holds or σ=1 and (2.7) holds, and that for every 𝒇Λσ(ν)(Ω,N) any local solution 𝒖 to (3.1) is continuous. Then, by Theorem 3.1,

(4.26)Λσ(ν)(Ω,N)L(np,1p-1)(Ω,N).

Under the present assumptions, Λσ(ν)(Ω,N) is a rearrangement-invariant space, and owing to (2.2), the inclusion (4.26) in fact implies the embedding (4.16), namely (4.17) or (4.18). Distinguishing into the same cases as above, and making use of the necessity of the conditions for (4.17) or (4.18), tells us that the corresponding condition in (3.4) must be in force. ∎

Proof of Theorem 3.5.

By Theorem 3.1, any local solution 𝒖 to (3.1) is continuous provided that

(4.27)Λρ,(ν)(Ω,N)L(np,1p-1)(Ω,N),

namely, if

(4.28)(01ϕ**(t)1p-1tpn(p-1)-1𝑑t)p-1Csup0<t<1ϕ*(t)𝒱(t)1ρ

for every measurable function ϕ on (0,1). By [8, Theorem 4.4], (4.28) holds if (and only if) (3.5) holds.

Conversely, assume that either ρ>1 and (2.6) holds or ρ=1 and (2.7) holds, and that for every 𝒇Λρ,(ν)(Ω,N) any local solution 𝒖 to (3.1) is continuous. Theorem 3.1 then yields the inclusion

Λρ,(ν)(Ω,N)L(np,1p-1)(Ω,N).

Our present assumptions ensure that Λρ,(ν)(Ω,N) is a rearrangement-invariant space, and owing to (2.2) this inclusion in fact implies the embedding (4.27), namely (4.28). Making use of the necessity part of [8, Theorem 4.4] tells us that (3.5) must hold. ∎

In view of the proof of Theorem 3.6, we preliminarily establish a lemma on inclusions between Lorentz and Orlicz type spaces. In fact, just one of the relevant inclusions will be needed in what follows. However, we believe that its full content can be of use for future applications, and we thus provide a complete statement and proof.

Lemma 4.2

Let 0<p,q< and let A be a generalized Young function. Let Ω be a measurable subset of Rn with |Ω|<.

Part I. Assume that

(4.29){lim supttpA(t)<if pq,(tA(t))qp-qtp(q-1)p-q𝑑t<if p>q.

Then,

(4.30)LA(Ω,N)Lp,q(Ω,N)

and

(4.31)L(A)(Ω,N)L(p,q)(Ω,N).

Part II. Assume that

(4.32){lim suptA(t)tp<if pq,(A(t)t)qq-ptq-1𝑑t<if q>p.

Then,

(4.33)Lp,q(Ω,N)LA(Ω,N)

and

(4.34)L(p,q)(Ω,N)L(A)(Ω,N).

Proof.

Let 𝒈:ΩN be a measurable function. An application of Fubini’s theorem tells us that

(4.35)0|Ω||𝒈|*(s)qsqp-1ds=p0|{|𝒈|*>t}|qptq-1dt

and

(4.36)0|Ω|A(|𝒈|*(s))ds=0|{|𝒈|*>t}|a(t)dt,

where a is the nonnegative function appearing in (2.10). Thus, (4.30) holds, provided that the inequality

(4.37)(0ϕ(t)qptq-1𝑑t)pqC0ϕ(t)a(t)𝑑t

holds for some constant C and for every non-increasing function ϕ:(0,)[0,). Inequality (4.37) follows from (4.29), owing to [8, Theorem 3.1]. Here, one also makes use of the fact that, since |Ω|<, the behavior of A near 0 is irrelevant in the definition of the space LA(Ω,N).

Similarly, (4.33) follows from the inequality

(4.38)0ϕ(t)a(t)𝑑tC(0ϕ(t)qptq-1𝑑t)pq

for some constant C and for every non-increasing function ϕ:(0,)[0,). Inequality (4.38) is a consequence of (4.32), thanks to [8, Theorem 3.1] again.

The proof of (4.31) and (4.34) is completely analogous: one has just to replace |𝒈|* with |𝒈|** in (4.35) and (4.36). ∎

Proof of Theorem 3.6.

Assume that 2p<n. By Theorem 3.1, it suffices to show that

LA(Ω,N)L(np,1p-1)(Ω,N).

Since L(np,1p-1)(Ω,N)=Lnp,1p-1(Ω,N) if 2p<n, the conclusion follows via Lemma 4.2.

Assume next that 2<p=n. On replacing, if necessary, A with an equivalent Young function, we may suppose, without loss of generality, that AC2(0,). By Theorem 3.1, it suffices to show that

(4.39)LA(Ω,N)L(1,1n-1)(Ω,N).

Define the generalized Young function B as B(t)=tA(t)-A(t) for t0. Observe that B is non-decreasing, since B(t)=tA′′(t)0 for t0. By [10, 18, 22], LA(Ω,N)L(B)(Ω,N). Hence, (4.39) holds if, in particular, L(B)(Ω,N)L(1,1n-1)(Ω,N), and the latter inclusion follows from (4.31).

Finally, assume that 2=p=n. Now, we have to show that LA(Ω,N)L(1,1)(Ω,N). Since L(1,1)(Ω,N)=LlogL(Ω,N), the conclusion follows, inasmuch as A(t) dominates tlog(1+t) near infinity, owing to our assumption on A. ∎

5 Proofs of the results of Section 3.2

We begin by establishing Theorem 3.8.

Proof of Theorem 3.8.

For the proof of Part I, we distinguish the following cases.

If σ>1, 2pn, we determine a weight m, optimal in a sense, such that

(5.1)𝒖Γσ(p-1)(m)(BR,N×n)C𝒇Γσ(ν)(B2R,N)1(p-1)+C𝒖L1(B2R,N×n)

for every ball B2RΩ. By Proposition (4.1), inequality (5.1) will follow if we show that

(5.2)(01(s1|𝒇|**(t)t-1n𝑑t)σm(s)𝑑s)1σC(01|𝒇|**(s)σν(s)𝑑s)1σ

for every measurable function 𝒇:ΩN. Inequality (5.2) holds, in particular, if

(5.3)(01(s1ϕ(t)𝑑t)σm(s)𝑑s)1σC(01ϕ(s)σs1nν(s)𝑑s)1σ

for every nonnegative measurable function ϕ in (0,1). A classical criterion for weighted Hardy type inequalities [26, Theorem 1.3.2/2] ensures that the weight m, defined by

m(s)=dds(s1(t1nν(t)1σ)-σ𝑑t)1-σif 0<s12

and m(s)=0 if 12<s<1, renders (5.3) true. In conclusion, inequality (5.1) is fulfilled with this choice of m.

By [11, Theorem 1.3],

𝒖C0,φ(BR,N)C𝒖Γσ(p-1)(m)(BR,N×n),

where

φ(s)=t-1nχ(0,sn)(t)(Γσ(p-1)(m))(0,1)for s near 0.

Inasmuch as (Λσ(p-1)(m)(0,1)Γσ(p-1)(m)(0,1), one has that

(Γσ(p-1)(m))(0,1)(Λσ(p-1)(m))(0,1).

Since 1<σ(p-1)<, by [8, Theorem 9.1 (i)],

ϕ(Λσ(p-1)(m))(0,1)(01ϕ**(t)(σ(p-1))t(σ(p-1))m(t)M(t)(σ(p-1))𝑑t)1(σ(p-1))

for every measurable function ϕ in (0,1), where the function M is given by

M(t)=0tm(s)𝑑s=(t1s-σnν(s)-1(σ-1)𝑑s)1-σif 0<t12

and M(t)=M(12) if 12<t<1. Therefore, there exists a constant C such that

φ(s)C{01((()-1nχ(0,sn)())**(t))(σ(p-1))t(σ(p-1))m(t)M(t)(σ(p-1))𝑑t}1(σ(p-1))
{0snt(σ(p-1))nm(t)M(t)(σ(p-1))𝑑t+s(σ(p-1))sn1m(t)M(t)(σ(p-1))𝑑t}1(σ(p-1))
{limt0+t(σ(p-1))nM(t)1-(σ(p-1))(σ(p-1))-1+σ(p-1)n0snt(σ(p-1))n-1M(t)1-(σ(p-1))dt
(5.4)-s(σ(p-1))M(1)1-(σ(p-1))(σ(p-1))-1}1(σ(p-1))for s near 0.

An application of de L’Hôpital’s rule tells us that the second addend in braces on the rightmost side of (5.4) decays to 0 as s0+ more slowly than the third addend. On the other hand,

(5.5)0snt(σ(p-1))n-1M(t)1-(σ(p-1))𝑑t1M(sn)(σ(p-1))-10snt(σ(p-1))n-1𝑑ts(σ(p-1))M(sn)(σ(p-1))-1for s(0,1),

where the first inequality is a consequence of the fact that M(sn)1-(σ(p-1)) is a non-increasing function. Equation (5.5) implies that the limit in (5.4) equals 0. Altogether, we have shown that

φ(s){0snt(σ(p-1))n-1M(t)1-(σ(p-1))𝑑t}1(σ(p-1))
(5.6){0snt(σ(p-1))n-1(t1τ-σnν(τ)-1σ-1𝑑τ)(σ-1)((σ(p-1))-1)𝑑t}1(σ(p-1))for s near 0.

If σ=1, 2<pn, let us determine a measure μ in (0,1) such that

(5.7)𝒖Γp-1(dμ)(BR,N×n)C(𝒇Γ1(ν)(B2R,N)1p-1+𝒖L1(B2R,N×n))

for every B2RΩ. Owing to estimate (4.5), inequality (5.7) is a consequence of the inequality

(5.8)01(t1s-1nϕ(s)𝑑s)𝑑μ(t)c01ϕ(s)ν(s)𝑑s

for every nonnegative measurable function ϕ in (0,1). By [26, Theorem 1.3.2/4], inequality (5.8) holds provided that μ obeys

μ([0,s])=infs<t<1(t1nν(t))if 0<s12

and μ([0,s])=μ([0,12]) if 12<s<1. By [11, Theorem 1.3],

𝒖C0,φ(BR,N)C𝒖Γp-1(dμ)(B2R,N×n),

where

(5.9)φ(s)=t-1nχ(0,sn)(t)(Γp-1(dμ))(0,1)t-1nχ(0,sn)(t)(Λp-1(dμ))(0,1)for s near 0.

Since we are assuming that 1<p-1<,

(5.10)ϕ(Λp-1(dμ))(0,1)(01ϕ**(t)(p-1)t(p-1)dμ(t)M(t)(p-1))1(p-1)

for every measurable function ϕ in (0,1), where now we have set M(t)=μ([0,t]) for t(0,1). Equation (5.10) follows from [8, Theorem 9.1 (i)] in the case when μ is absolutely continuous, but the same proof applies to the present slightly more general situation.

By (5.9) and (5.10), there exists a constant C such that

φ(s)C{01((()-1nχ(0,sn)())**(t))(p-1)t(p-1)dμ(t)M(t)(p-1)}1(p-1)
{0snt(p-1)ndμ(t)M(t)(p-1)+s(p-1)sn1dμ(t)M(t)(p-1)}1(p-1)
{limt0+t(p-1)nM(t)1-(p-1)(p-1)-1+(p-1)n0snt(p-1)n-1M(t)1-(p-1)𝑑t-s(p-1)M(1)1-(p-1)(p-1)-1}1(p-1)
{0snt(p-1)n-1(inft<τ<1(τ1nν(τ)))1-(p-1)𝑑t}1(p-1)for s near 0,

where the last equivalence follows via an argument analogous to that employed in the derivation of (5.6) from (5.4).

If σ=1, 2=p=n, we exhibit a measure μ such that

(5.11)𝒖Γ1(dμ)(BR,N×n)C(𝒇Γ1(ν)(B2R,N)+𝒖L1(B2R,N×n))

for every B2RΩ. By Proposition 4.1, inequality (5.11) holds if the measure μ supports the inequality

(5.12)01(t1s-12ϕ(s)𝑑s)𝑑μ(t)C01ϕ(s)ν(s)𝑑s

for every nonnegative measurable function ϕ in (0,1). Owing to [26, Theorem 1.3.2/4], inequality (5.12) is satisfied, provided that μ obeys

μ([0,s])=infs<t<1(t12ν(t))if 0<s12

and μ([0,s])=μ([0,12]) if 12<s<1. Thanks to [11, Theorem 1.3],

𝒖C0,φ(BR,N)C𝒖Γ1(dμ)(BR,N×n),

where

φ(s)=t-12χ(0,s2)(t)(Γ1(dμ))(0,1)t-12χ(0,s2)(t)(Λ1(dμ))(0,1)for s near 0.

Moreover, by [8, Theorem 9.1 (i)],

ϕ(Λ(dμ))(0,1)sup0<t<1ϕ**(t)tμ([0,t])

for every measurable function ϕ in (0,1). Let us notice that in [11] and [8], absolutely continuous measures μ are considered, but analogous proofs apply to the present slightly more general setting. Thereby, there exists a constant C such that

φ(s)Csup0<t<1((()-12χ(0,s2)())**(t)tμ([0,t]))sup0<t<1(min{t12,s}inft<τ<1(τ12ν(τ)))for s near 0.

For Part II, we assume that (3.8) is in force. We claim that |𝒖|L(BR,N×n), and hence 𝒖Lip(BR,N). In order to prove our claim, it suffices to show that

(5.13)𝒖L(BR,N×n)C(𝒇Γσ(ν)(B2R,N)1p-1+𝒖L1(B2R,N×n))

for every ball B2RΩ. Inequality (5.13) is a consequence of Proposition 4.1 and Hölder’s inequality, since

s1|𝒇|**(t)t-1n𝑑tL(0,1)=01|𝒇|**(t)t-1n𝑑t(01|𝒇|**(t)σν(t)𝑑t)1σ(01(t-1nν(t)-1σ)σ𝑑t)1σ.

Proof of Theorem 3.9.

We follow the outline of the proof of Theorem 3.8 and find a weight m such that

(5.14)𝒖Γρ(p-1),(m)(BR,N×n)C(𝒇Γρ,(ν)(B2R,N)1(p-1)+𝒖L1(B2R,N×n))

for every ball B2RΩ. By Proposition 4.1, inequality (5.14) is implied by the inequality

(5.15)(s1|𝒇|**(t)t-1ndt)M(s)1ρL(0,1)C|𝒇|**(s)𝒱(s)1ρL(0,1)

for every measurable function 𝒇:ΩN, where M(s)=0sm(t)𝑑t and 𝒱(s)=0sν(t)𝑑t for s(0,1). Owing to [26, Theorem 1.3.2/2], the weight m, defined as

m(s)=dds(s1t-1n𝒱(t)-1ρ𝑑t)-ρif 0<s12

and m(s)=0 if 12<s<1, supports (5.15), and hence (5.14).

An application of [11, Theorem 1.3] now tells us that

𝒖C0,φ(BR,N)C𝒖Γρ(p-1),(m)(BR,N×n),

where

φ(s)=t-1nχ(0,sn)(t)(Γρ(p-1),(m))(0,1)t-1nχ(0,sn)(t)(Λρ(p-1),(m))(0,1)for s near 0.

On the other hand, [8, Theorem 9.1 (ii)] ensures that

ϕ(Λρ(p-1),(m))(0,1)01ϕ*(t)M(t)-1ρ(p-1)𝑑t

for every measurable function ϕ in (0,1). As a consequence, there exists a constant C such that

φ(s)C01(()-1nχ(0,sn)())*(t)M(t)-1ρ(p-1)𝑑t=0snt-1n(t1τ-1n𝒱(τ)-1ρ𝑑τ)1p-1𝑑tfor s near 0.

The conclusion follows. ∎

Proof of Theorem 3.10.

The conclusions follow on applying Theorem 3.8 with appropriate choices of the power σ and the weight ν, via standard computations. The details are omitted for brevity. ∎

Proof of Proposition 3.12.

Let BR(0), hL(r,q)(logL)α(Ω), and v be as in the proof of Theorem 3.1, so that v is a (radially decreasing) solution to the p-Laplace equation (4.9). One has that

(5.16)vL1(BR(0),n)=ChL(nnp-p+1,1p-1)(BR(0))ChL(r,q)(logL)α(Ω)

for some constants C,C and for every r,q,α as in the statement. Observe that the inequality holds, by well-known inclusion relations between Lorentz–Zygmund spaces, since r>nnp-p+1 for every r as in (3.11). Now, assume that (3.10) holds with φ replaced with some other modulus of continuity ψ, namely,

(5.17)supx,yBR2(0)|v(x)-v(y)|ψ(|x-y|)C(fL(r,q)(logL)α(BR(0))1p-1+vL1(BR(0))).

Choosing x=0 and y such that |y|=s<R2 in (5.17) and making use of (5.16) yield

(5.18)suphL(r,q)(logL)α(BR(0))sup0<s<R20ωnsn(ϱpnh**(ϱ))1p-1dϱϱψ(s)(hL(r,q)(logL)α(BR(0))1p-1)C.

Exchanging the order of suprema in (5.18) tells us that

(5.19)Cψ(s)suphL(r,q)(logL)α(BR(0))0ωnsn(ϱpnh**(ϱ))1p-1dϱϱ(01(ϱ1r(1+log1ϱ)αh**(ϱ))qdϱϱ)1q(p-1)for 0<s<R2.

We claim that the supremum on the right-hand side of (5.19) is equivalent for s near 0 to the function φ(s) defined by (3.9). This equivalence follows from an application of [17, Theorem 5.1 (i)] if p=2, q=1, and n and α obey (3.11). If instead q= and n,p,α satisfy (3.11), then our claim can be derived via [8, Theorem 6.4]. All the remaining cases can be dealt by [17, Theorem 5.1 (ii)]. The details are omitted for brevity. ∎

Proof of Proposition 3.7.

By the same argument as in the proof of Proposition 3.12, inequality (3.6) implies (5.19) with r=np, q=1p-1, and α=0, namely,

Cφ(s)suphL(np,1p-1)(logL)α(BR(0))0ωnsn(ϱpnh**(ϱ))1p-1dϱϱ0ωnsn(ϱpnh**(ϱ))1p-1dϱϱ=1for 0<s<R2.

Hence, the conclusion follows. ∎


Communicated by Giuseppe Mingione


Funding statement: This research was partly supported by the research project of MIUR (Italian Ministry of Education, University and Research) PRIN 2012, no. 2012TC7588, “Elliptic and Parabolic Partial Differential Equations: Geometric Aspects, Related Inequalities, and Applications”, and by the GNAMPA (National Group for Mathematical Analysis, Probability and their Applications) of the Italian INdAM (National Institute of High Mathematics).

References

[1] Alberico A. and Cianchi A., Optimal summability of solutions to nonlinear elliptic problems, Nonlinear Anal. 67 (2007), no. 6, 1775–1790. 10.1016/j.na.2006.08.023Search in Google Scholar

[2] Alberico A., Cianchi A. and Sbordone C., On the modulus of continuity of solutions to the n-Laplace equation, J. Elliptic Parabolic Equations 1 (2015), 1–11. 10.1007/BF03377364Search in Google Scholar

[3] Alberico A. and Ferone V., Regularity properties of solutions of elliptic equations in 2 in limit cases, Atti Accad. Naz. Lincei Cl. Sci. Fis. Mat. Natur. Rend. Lincei (9) Mat. Appl. 6 (1995), no. 4, 237–250. Search in Google Scholar

[4] Bennett C. and Sharpley R., Interpolation of Operators, Pure Appl. Math. 129, Academic Press, Boston, 1988. Search in Google Scholar

[5] Bingham N. H., Goldie C. M. and Teugels J. L., Regular Variation, Encyclopedia Math. Appl. 27, Cambridge University Press, Cambridge, 1989. Search in Google Scholar

[6] Carro M., García del Amo A. and Soria J., Weak-type weights and normable Lorentz spaces, Proc. Amer. Math. Soc. 124 (1996), no. 3, 849–857. 10.1090/S0002-9939-96-03214-5Search in Google Scholar

[7] Carro M., Gogatishvili A., Martín J. and Pick L., Weighted inequalities involving two Hardy operators with applications to embeddings of function spaces, J. Operator Theory 59 (2008), no. 2, 309–332. Search in Google Scholar

[8] Carro M., Pick L., Soria J. and Stepanov V. D., On embeddings between classical Lorentz spaces, Math. Inequal. Appl. 4 (2001), no. 3, 397–428. 10.7153/mia-04-37Search in Google Scholar

[9] Carro M. and Soria J., Boundedness of some integral operators, Canad. J. Math. 45 (1993), no. 6, 1155–1166. 10.4153/CJM-1993-064-2Search in Google Scholar

[10] Cianchi A., Korn type inequalities in Orlicz spaces, J. Funct. Anal. 267 (2014), no. 7, 2313–2352. 10.1016/j.jfa.2014.07.012Search in Google Scholar

[11] Cianchi A. and Pick L., Sobolev embeddings into spaces of Campanato, Morrey, and Hölder type, J. Math. Anal. Appl. 282 (2003), no. 1, 128–150. 10.1016/S0022-247X(03)00110-0Search in Google Scholar

[12] Dal Maso G. and Malusa A., Some properties of reachable solutions of nonlinear elliptic equations with measure data, Ann. Scuola Norm. Sup. Pisa Cl. Sci. (4) 25 (1997), no. 1–2, 375–396. Search in Google Scholar

[13] Dall’Aglio A., Approximated solutions of equations with L1 data. Application to the H-convergence of quasi-linear parabolic equations, Ann. Mat. Pura Appl. (4) 170 (1996), 207–240. 10.1007/BF01758989Search in Google Scholar

[14] Duzaar F. and Mingione G., Gradient estimates via linear and nonlinear potentials, J. Funct. Anal. 259 (2010), no. 11, 2961–2998. 10.1016/j.jfa.2010.08.006Search in Google Scholar

[15] Duzaar F. and Mingione G., Gradient estimates via non-linear potentials, Amer. J. Math. 133 (2011), no. 4, 1093–1149. 10.1353/ajm.2011.0023Search in Google Scholar

[16] Ferone V. and Fusco N., Continuity properties of minimizers of integral functionals in a limit case, J. Math. Anal. Appl. 202 (1996), no. 1, 27–52. 10.1006/jmaa.1996.0301Search in Google Scholar

[17] Gogatishvili A. and Pick L., Discretization and anti-discretization of rearrangement-invariant norms, Publ. Mat. 47 (2003), no. 2, 311–358. 10.5565/PUBLMAT_47203_02Search in Google Scholar

[18] Greco L., Iwaniec T. and Moscariello G., Limits of the improved integrability of the volume forms, Indiana Univ. Math. J. 44 (1995), no. 2, 305–339. 10.1512/iumj.1995.44.1990Search in Google Scholar

[19] Heinig H. P. and Maligranda L., Weighted inequalities for monotone and concave functions, Studia Math. 116 (1995), no. 2, 133–165. Search in Google Scholar

[20] Iwaniec T. and Onninen J., Continuity estimates for n-harmonic equations, Indiana Univ. Math. J. 56 (2007), no. 2, 805–824. 10.1512/iumj.2007.56.2987Search in Google Scholar

[21] Jiang R., Koskela P. and Yang D., Continuity of solutions to n-harmonic equations, Manuscripta Math. 139 (2012), no. 1–2, 237–248. 10.1007/s00229-011-0514-1Search in Google Scholar

[22] Kita H., On maximal functions in Orlicz spaces, Proc. Amer. Math. Soc. 124 (1996), no. 10, 3019–3025. 10.1090/S0002-9939-96-03807-5Search in Google Scholar

[23] Kuusi T. and Mingione G., Universal potential estimates, J. Funct. Anal. 262 (2012), no. 10, 4205–4269. 10.1016/j.jfa.2012.02.018Search in Google Scholar

[24] Kuusi T. and Mingione G., Linear potentials in nonlinear potential theory, Arch. Ration. Mech. Anal. 207 (2013), no. 1, 215–246. 10.1007/s00205-012-0562-zSearch in Google Scholar

[25] Kuusi T. and Mingione G., Vectorial nonlinear potential theory, J. Eur. Math. Soc., to appear. 10.4171/JEMS/780Search in Google Scholar

[26] Maz’ya V. G., Sobolev Spaces. With Applications to Elliptic Partial Differential Equations, 2nd ed., Grundlehren Math. Wiss. 342, Springer, Berlin, 2011. Search in Google Scholar

[27] Mingione G., Gradient potential estimates, J. Eur. Math. Soc. 13 (2011), no. 2, 459–486. 10.4171/JEMS/258Search in Google Scholar

[28] O’Neil R., Convolution operators and L(p,q) spaces, Duke Math. J. 30 (1963), 129–142. Search in Google Scholar

[29] Sawyer E., Boundedness of classical operators on classical Lorentz spaces, Studia Math. 96 (1990), no. 2, 145–158. 10.4064/sm-96-2-145-158Search in Google Scholar

[30] Stein E. M., Editor’s note: the differentiability of functions in n, Ann. of Math. (2) 113 (1981), no. 2, 383–385. Search in Google Scholar

[31] Teixeira E. V., Sharp regularity for general Poisson equations with borderline sources, J. Math. Pures Appl. (9) 99 (2013), no. 2, 150–164. 10.1016/j.matpur.2012.06.007Search in Google Scholar

Received: 2015-7-22
Accepted: 2015-10-12
Published Online: 2015-11-17
Published in Print: 2017-1-1

© 2017 by De Gruyter

Downloaded on 9.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/acv-2015-0029/html
Scroll to top button