• Aucun résultat trouvé

BMO, integrability, Harnack and Caccioppoli inequalities for quasiminimizers

N/A
N/A
Protected

Academic year: 2022

Partager "BMO, integrability, Harnack and Caccioppoli inequalities for quasiminimizers"

Copied!
17
0
0

Texte intégral

(1)

www.elsevier.com/locate/anihpc

BMO, integrability, Harnack and Caccioppoli inequalities for quasiminimizers

Anders Björn

a

, Jana Björn

a

, Niko Marola

b,

aDepartment of Mathematics, Linköpings universitet, SE-581 83 Linköping, Sweden

bDepartment of Mathematics and Statistics, University of Helsinki, P.O. Box 68, FI-00014 University of Helsinki, Finland

Received 10 November 2009; received in revised form 26 August 2010; accepted 20 September 2010 Available online 6 October 2010

Abstract

In this paper we use quasiminimizing properties of radial power-type functions to deduce counterexamples to certain Caccioppoli-type inequalities and weak Harnack inequalities for quasisuperharmonic functions, both of which are well known to hold forp-superharmonic functions. We also obtain new bounds on the local integrability for quasisuperharmonic functions.

Furthermore, we show that the logarithm of a positive quasisuperminimizer has bounded mean oscillation and belongs to a Sobolev type space.

©2010 Elsevier Masson SAS. All rights reserved.

Résumé

Dans cet article nous utilisons les propriétés des fonctions avec puissance radiale afin d’obtenir des contre-exemples à certaines inéquations de type Caccioppoli et Harnack faible pour les fonctions quasisuperharmoniques, lesquelles sont bien connues être valables pour les fonctionsp-superharmoniques. Nous obtenons aussi de nouvelles bornes pour l’intégrabilité locale des fonctions quasisuperharmoniques. De plus nous démontrons que le logarithme d’une fonction positive quasiminimisante est de type BMO, et appartient à un espace de Sobolev.

©2010 Elsevier Masson SAS. All rights reserved.

MSC:primary 49J20; secondary 30L99, 31C45, 31E05, 35J20, 49J27

Keywords:Bounded mean oscillation; Doubling measure; Metric space; Nonlinear;p-harmonic; Poincaré inequality; Potential theory;

Quasiminimizer; Quasisuperharmonic; Quasisuperminimizer; Weak Harnack inequality

1. Introduction

Let 1< p <∞and letΩRnbe a nonempty open set. A functionuWloc1,p(Ω)is aQ-quasiminimizer,Q1, inΩif

ϕ=0

|∇u|pdxQ

ϕ=0

(u+ϕ)pdx (1.1)

* Corresponding author.

E-mail addresses:anbjo@mai.liu.se (A. Björn), jabjo@mai.liu.se (J. Björn), niko.marola@helsinki.fi (N. Marola).

0294-1449/$ – see front matter ©2010 Elsevier Masson SAS. All rights reserved.

doi:10.1016/j.anihpc.2010.09.005

(2)

for allϕW01,p(Ω). Quasiminimizers were introduced by Giaquinta and Giusti [19,20] as a tool for a unified treat- ment of variational integrals, elliptic equations and quasiregular mappings on Rn. They realized that De Giorgi’s method could be extended to quasiminimizers, obtaining, in particular, local Hölder continuity. DiBenedetto and Trudinger [15] proved the Harnack inequality for quasiminimizers, as well as weak Harnack inequalities for quasisub- and quasisuperminimizers. We recall that a functionuWloc1,p(Ω)is aquasisub(super)minimizerif (1.1) holds for all nonpositive (nonnegative)ϕW01,p(Ω).

Compared with the theory ofp-harmonic functions we have no differential equation for quasiminimizers, only the variational inequality can be used. There is also no comparison principle nor uniqueness for the Dirichlet problem.

The following result was recently obtained by Martio [37, Theorem 4.1]. It shows that quasiminimizers are much more flexible under perturbations than solutions of differential equations, which can be useful in applications and in particular shows that results obtained for quasiminimizers are very robust.

Theorem 1.1. Letu be aQ-quasiminimizer inΩ andfWloc1,p(Ω)be such that|∇f|c|∇u|a.e. inΩ, where 0< c < Q1/p. Thenu+f is aQ-quasiminimizer inΩ, whereQ=(1+c)p/(Q1/pc)p.

After the papers by Giaquinta and Giusti [19,20] and DiBenedetto and Trudinger [15], Ziemer [45] gave a Wiener- type criterion sufficient for boundary regularity for quasiminimizers. Tolksdorf [42] obtained a Caccioppoli inequality and a convexity result for quasiminimizers. The results in [15,19,20,45] were extended to metric spaces by Kinnunen and Shanmugalingam [30] and J. Björn [11] in the beginning of this century, see also A. Björn and Marola [9].

Soon afterwards, Kinnunen and Martio [28] showed that quasiminimizers have an interesting potential theory, in particular they introduced quasisuperharmonic functions, which are related to quasisuperminimizers in a similar way as superharmonic functions are related to supersolutions, see Definition 2.4.

The one-dimensional theory was already considered in [19], and has since been further developed by Martio and Sbordone [38], Judin [27], Martio [35] and Uppman [44]. Most aspects of the higher-dimensional theory fit just as well in metric spaces, and this theory, in particular concerning boundary regularity, has recently been developed further in a series of papers by Martio [34–37], A. Björn and Martio [10], A. Björn [1–4] and J. Björn [12].

So far, most of the theory for quasiminimizers has been extending various results known forp-harmonic functions.

This paper goes in the opposite direction: we show that some results are not extendable and the class of quasimini- mizers behaves in a way that was not expected.

Superminimizers, i.e. 1-quasisuperminimizers, are nothing but supersolutions to thep-Laplace equation div

|∇u|p2u

=0.

Until now, there have been very few examples of quasi(super)minimizers for which the best quasi(super)minimizer constant is known, apart from a few explicit examples ofp-harmonic functions, i.e. withQ=1. In one dimension there are a couple of examples with optimal quasiminimizer constant in Judin [27], Martio [35] and Uppman [44]. As far as we know, the only explicit examples of quasiminimizers with optimal quasiminimizer constantQ >1 in higher dimensions were recently obtained by Björn and Björn [7]. LetB=B(0,1)denote the unit ball inRn.

Theorem 1.2.Let1< p < n,αβ=(pn)/(p−1)andu(x)= |x|α. Thenuis aQ-quasiminimizer inB\ {0}

and aQ-quasisuperharmonic function inB, where Q=

α β

pp+n p+n

is the best quasiminimizer constant in both cases.

Trudinger [43] obtained a sharp weak Harnack inequality for supersolutions (Q=1), whose exponent coincides with the sharp exponent for local integrability of superharmonic functions. As a consequence of Theorem 1.2, we will show that the best exponent in the weak Harnack inequality forQ-quasisuperminimizers must depend onQ, and tends to 0, asQ→ ∞. The same is true for the best exponent of local integrability forQ-quasisuperharmonic functions.

Theorem 1.2 implies upper bounds for these exponents, which forQ=1 coincide with the known sharp bounds. It is therefore natural to expect that these bounds are sharp also forQ >1.

(3)

Similar conclusions can be drawn for Caccioppoli inequalities for quasisuperminimizers: some of the “classical”

Caccioppoli type inequalities cannot be true with exponents independent of the quasiminimizing constantQ. There is a gap between the sharp exponents forQ=1 and the known exponents forQ >1, and Theorem 1.2 implies bounds for these exponents, which forQ=1 coincide with the known sharp bound.

Caccioppoli inequalities and the weak Harnack inequality for quasisuperminimizers are essential for extending the Moser iteration technique in full to quasiminimizers. A. Björn and Marola [9] have shown that the scheme applies to a large extent to this setting, but a logarithmic Caccioppoli inequality for quasisuperminimizers still needs to be proved in order to obtain the full result. It is traditionally used to show that the logarithm of a positive superminimizer belongs to BMO. We show that the logarithm of a positive quasisuperminimizer belongs to BMO simply by exploiting a classical tool from harmonic analysis in our setting. Interesting enough, in light of Theorem 1.2, this qualitative result seems to be the best we can hope for.

The outline of the paper is as follows: In Section 2 we introduce the relevant background on metric spaces and quasiminimizers. Those readers only interested in Euclidean spaces may simply replace the minimal upper gradient guby the modulus|∇u|of the usual gradient and the Newtonian spaceN1,pby the Sobolev spaceW1,pthroughout the paper.

In Section 3 we turn to Caccioppoli inequalities for quasisuperminimizers and deduce bounds on the exponents.

In Section 4 we obtain bounds on the exponents in weak Harnack inequalities for quasisuperminimizers and for the local integrability of quasisuperharmonic functions. In Section 5 we prove some new Caccioppoli inequalities for quasiminimizers. We also show that the logarithm of a positive quasisuperminimizer belongs to BMO, qualitatively, and toWloc1,q for everyq < p.

Finally, in Section 6 we make a digression and discuss the relation between the quasiconvexity constantLand the dilation constantλin the weak Poincaré inequality; both constants play essential roles in Sections 4 and 5.

2. Preliminaries

The theory of quasiminimizers fits naturally into the analysis on metric spaces, as it uses variational integrals rather than partial differential equations. In this case, the metric space(X, d)is assumed to be complete and equipped with a doubling measureμ, i.e. there exists a doubling constantCμsuch that

0< μ(2B)Cμμ(B) <

for every ball B =B(x, r)= {yX: d(x, y) < r}, where λB =B(x, λr). Moreover we require the measure to support a weak(1, p)-Poincaré inequality, see below. Throughout the paper we assume that 1< p <∞.

We follow Heinonen and Koskela [25] in introducing upper gradients as follows (they called them very weak gradients).

Definition 2.1.A nonnegative Borel functiongonXis anupper gradientof an extended real-valued functionf on Xif for all curvesγ: [0, lγ] →X,

f γ (0)

f

γ (lγ)

γ

g ds (2.1)

whenever bothf (γ (0))andf (γ (lγ))are finite, and

γg ds= ∞otherwise.

A nonnegative measurable functiongonXis ap-weak upper gradientoff if (2.1) holds forp-almost every curve in the sense of Definition 2.1 in Shanmugalingam [40]. It is implicitly assumed that

γg dsis defined (with a value in [0,∞]) forp-almost every rectifiable curveγ.

Thep-weak upper gradients were introduced in Koskela and MacManus [31]. They also showed that ifgLp(X) is ap-weak upper gradient off, then one can find a sequence{gj}j=1of upper gradients off such thatgjgin Lp(X). Iff has an upper gradient inLp(X), then it has aminimalp-weak upper gradientgfLp(X)in the sense that for everyp-weak upper gradientgLp(X)off,gf ga.e., see Corollary 3.7 in Shanmugalingam [41]. (The reader may also consult Björn and Björn [8] where proofs of all the facts mentioned in this section are given, apart from those about quasiminimizers.)

(4)

Next we define a version of Sobolev spaces on the metric spaceXdue to Shanmugalingam in [40]. Cheeger [14]

gave an alternative definition which leads to the same space (whenp >1) see [40].

Definition 2.2.WheneveruLp(X), let uN1,p(X)=

X

|u|p+inf

g

X

gp 1/p

,

where the infimum is taken over all upper gradients ofu. TheNewtonian spaceonXis the quotient space N1,p(X)= u: uN1,p(X)<

/,

whereuvif and only ifuvN1,p(X)=0.

The spaceN1,p(X)is a Banach space and a lattice, see Shanmugalingam [40].

A functionu belongs to thelocal Newtonian spaceNloc1,p(Ω)ifuN1,p(V )for all bounded open setsV with VΩ, the latter space being defined by considering V as a metric space with the metric d and the measure μ restricted to it.

Ifu, vNloc1,p(X), thengu=gva.e. in{xX: u(x)=v(x)}, in particulargmin{u,c}=guχu=cforcR. For these and other facts onp-weak upper gradients, see, e.g., Björn and Björn [5, Section 3] (which is not included in Björn and Björn [6]).

Definition 2.3.We say that Xsupports aweak(q, p)-Poincaré inequalityif there exist constantsC >0 andλ1 such that for all ballsBX, all integrable functionsf onXand all upper gradientsgoff,

B

|ffB|q 1/q

Cdiam(B)

λB

gp 1/p

, (2.2)

wherefB:=

Bf dμ:=

Bf dμ/μ(B).

In the definition of Poincaré inequality we can equivalently assume thatg is ap-weak upper gradient—see the comments above.

For more details see any of the papers on metric spaces in the reference list. Note, in particular, thatRn with the Lebesgue measure=dx, as well as weighted Rnwithp-admissible weights are special cases of metric spaces satisfying our assumptions.

In metric spaces,guis the natural substitute for the scalar|∇u|. Note that we have no natural counterpart to the vector∇u. (See however Cheeger [14].) The Newtonian spaceN1,preplaces the Sobolev spaceW1,p. The definition of quasiminimizers on metric spaces is thus as follows. A functionuNloc1,p(Ω)is aQ-quasiminimizer,Q1, inΩ

if

ϕ=0

gpudxQ

ϕ=0

gup+ϕ (2.3)

for allϕ∈Lipc(Ω). Similarly, a functionuNloc1,p(Ω)is aQ-quasisub(super)minimizerif (2.3) holds for all non- positive (nonnegative)ϕ∈Lipc(Ω). Note also that a function is aQ-quasiminimizer inΩ if and only if it is both a Q-quasisubminimizer and aQ-quasisuperminimizer inΩ.

Our definition of quasiminimizers (and quasisub- and quasisuperminimizers) is one of several equivalent possi- bilities, see Proposition 3.2 in A. Björn [1]. In fact it is enough to test (1.1) with (all, nonpositive and nonnegative, respectively)ϕ∈Lipc(Ω), the space of Lipschitz functions with compact support inΩ.

Every quasiminimizer can be modified on a set of measure zero so that it becomes locally Hölder continuous inΩ. This was proved inRnby Giaquinta and Giusti [20, Theorem 4.2], and in metric spaces by Kinnunen and Shanmu- galingam [30, Proposition 3.8 and Corollary 5.5]. AQ-quasiharmonicfunction is a continuousQ-quasiminimizer.

(5)

Kinnunen and Martio [28, Theorem 5.3] showed that ifuis aQ-quasisuperminimizer inΩ, then itslower semi- continuous regularizationu(x)=ess lim infyxu(y)is also aQ-quasisuperminimizer inΩ belonging to the same equivalence class asuinNloc1,p(Ω). Furthermore,u isQ-quasisuperharmonic inΩ. For our purposes we make the following definition.

Definition 2.4.A functionu:Ω(−∞,∞]isQ-quasisuperharmonicinΩifuis not identically∞in any com- ponent of Ω,u is lower semicontinuously regularized, and min{u, k}is aQ-quasisuperminimizer in Ω for every kR.

This definition is equivalent to Definition 7.1 in Kinnunen and Martio [28], see Theorem 7.10 in [28]. (Note that there is a misprint in Definition 7.1 in [28]—the functionsvi are assumed to beQ-quasisuperminimizers.)

A function isp-harmonicif it is 1-quasiharmonic, it issuperharmonicif it is 1-quasisuperharmonic, and it is a sub(super)minimizerif it is a 1-quasisub(super)minimizer.

Unless otherwise stated, the letterC denotes various positive constants whose exact values are unimportant and may vary with each usage.

3. Caccioppoli inequalities for quasisuperminimizers

In this section we discuss Caccioppoli inequalities for quasi(super)minimizers. These inequalities play an important role, e.g., when proving regularity results for quasiminimizers, see, e.g., DiBenedetto and Trudinger [15], Kinnunen and Shanmugalingam [30] and A. Björn and Marola [9]. We will show how some well-known results for sub- and superminimizers (i.e. withQ=1) do not extend to the caseQ >1.

Letγ0=γ0(Q, p)be the largest number (independent ofXandΩ) such that for everyγ < γ0there is a constant Cγ =Cγ(Q, p, X, Ω)such that theCaccioppoli inequality

Ω

up+γgupηpdμCγ

Ω

uγgηp (3.1)

holds for allQ-quasisuperminimizersu0 inΩand all 0η∈Lipc(Ω).

By Proposition 7.3 in A. Björn and Marola [9], (3.1) holds for allγ <0 and thusγ0(Q, p)0, both inRn and in metric spaces. We also know thatγ0(1, p)p−1, see Heinonen, Kilpeläinen and Martio [24, Lemma 3.57] for the Euclidean case and Kinnunen and Martio [29, Lemma 3.1] for metric spaces. It is probably well known that γ0(1, p)=p−1, but it also follows from the following more general proposition, withα=β.

Proposition 3.1.Letn > pandα(pn)/(p−1)=β. Let further Q=

α β

pp+n

p+n and δ(Q, p)=pn

α . (3.2)

Thenγ0(Q, p)δ(Q, p).

Note thatδ is really a function ofQandponly: there is a one-to-one correspondence betweenQ1 andαβ, and thusδ(Q, p)is a function ofQ,pandna priori, and is independent ofnby Proposition 3.3 below.

In view of this result, and the fact thatγ0(1, p)=p−1, it feels natural to make the following conjecture.

Conjecture 3.2.LetQ >1. We conjecture thatγ0(Q, p)=δ(Q, p).

Recall though that we do not know thatγ0(Q, p) is positive nor that (3.1) holds forγ =0, for anyQ >1. See however Lemma 5.8 below.

Proposition 3.3.Letn > p. Thenδ(Q, p)is the unique solution in(0, p−1]of the equation Q= (p−1)p1

δp1(pδ). (3.3)

In particular, the expression forδ(Q, p)in(3.2)is independent ofn > p.

(6)

Proof. Let 0< δp−1 be fixed andα=(pn)/δ(pn)/(p−1)=β <0. Note thatp+n=β. Inserting this into (3.2) gives

Q= p−1

δ

p (pn)/(p−1)

p(pn)/δp+n= (p−1)p1 δp1(pδ).

Note thatQ=1 forδ=p−1, andQ→ ∞, asδ→0. DifferentiatingQwith respect toδgives

∂Q

∂δ =(p−1)p1

1−p

δp(pδ)+ 1 δp1(pδ)2

=p(p−1)p1+1−p) δp(pδ)2 0

with equality only forδ=p−1. Thus,Qis strictly decreasing as a function ofδin the interval(0, p−1]. Conse- quently, for everyQ1, there exists a uniqueδ(0, p−1]satisfying (3.3). 2

It is easy to see that δ(Q,2)=1−√

1−1/Qandα(Q,2, n)=(2n)(Q+

Q2Q). For generalpwe can use (3.3) to obtain the following asymptotic estimates forδ(Q, p)andαin terms ofQandp.

Corollary 3.4.LetQ >1andn > p. Then p−1

(pQ)1/(p1) < δ(Q, p) < p−1

Q1/(p1). (3.4)

Moreover, ifβ:=(pn)/(p−1), then (pQ)1/(p1)β < α < Q1/(p1)β.

Proof. Writeδ:=δ(Q, p). Asα=(pn)/δ=(p−1)β/δ, it suffices to prove (3.4). Sinceδ >0, we have Q= (p−1)p1

δp1(pδ)>(p−1)p1 δp1p ,

proving the first inequality in (3.4). Similarly, as δ < p−1, we see that Q < (p−1)p1p1, and the second inequality in (3.4) follows. 2

Before giving the proof of Proposition 3.1 we show that it is equivalent to study Caccioppoli inequalities for quasisuperharmonic functions, a fact that we will actually use in the proof of Proposition 3.1.

Proposition 3.5.Letγ,Q,C, Ω and0η∈Lipc(Ω)be fixed. Then the Caccioppoli inequality

Ω

up+γgpuηpdμC

Ω

uγgpη (3.5)

holds for allQ-quasisuperminimizersu0inΩif and only if it holds for allQ-quasisuperharmonic functionsu0 inΩ.

Quasisuperharmonic (and also superharmonic) functions are in general too large to belong toNloc1,p(Ω). However the gradient is naturally defined bygu=gmin{u,k}on{x: u(x) < k}, for allk=1,2, . . ., see e.g. p. 150 in Heinonen, Kilpeläinen and Martio [24] for the Euclidean case or Kinnunen and Martio [29] for the metric space case. Here it is important to know that a quasisuperharmonic function is infinite only on a set with zero measure. Kinnunen and Martio [28, Theorem 10.6] showed even more: that it is infinite only on a set of zero capacity.

Proof of Proposition 3.5. Assume first that (3.5) holds for allQ-quasisuperharmonicu0 and letube a nonnegative Q-quasisuperminimizer. Thenu0 isQ-quasisuperharmonic andu=ua.e. Thus

Ω

up+γgpuηp=

Ω

up+γ

gupηpdμC

Ω

uγ

gηp=C

Ω

uγgηpdμ.

(7)

Conversely, letube a nonnegativeQ-quasisuperharmonic function. Thenuk:=min{u, k},k=1,2, . . ., is a non- negativeQ-quasisuperminimizer. Moreover,guk=χ{u<k}gua.e. By monotone convergence we see that

Ω

up+γgupηp= lim

k→∞

Ω

ukp+γgpukηp lim

k→∞C

Ω

uγkgηp=C

Ω

uγgηpdμ,

where we use the boundedness of|uγkuγ|,gηand suppηto establish the last equality ifγ <0. 2

Proof of Proposition 3.1. By Theorem 1.2,u(x)= |x|α isQ-quasisuperharmonic in B. Let δ:=(pn)/α and η(x)=min{(2−3|x|)+,1}. Then

B

up+δ|∇u|pηpdxC 1/3 0

rα(p+δ)r1)prn1dr=C 1/3 0

dr r = ∞. On the other hand,

B

uδ|∇η|pdx=C 2/3 1/3

rαδrn1dr <.

Thus the Caccioppoli inequality (3.5) does not hold for allQ-quasisuperharmonic functions withγ=δ. In view of Proposition 3.5, this shows thatγ0(Q, p)δ. 2

Remark 3.6.Ifp=n, then by Theorem 7.4 in Björn and Björn [7],u(x)=(−log|x|)α is quasisuperharmonic inB for allα1 andQ=αn/(nαn+1)is the best quasisuperminimizer constant. Arguments similar to those in the proofs of Propositions 3.1 and 3.3 then show thatγ0(Q, n)(n−1)/α=δ(Q, n). As in the proof of Corollary 3.4, one then obtains the estimatesQ1/(n1)< α < (nQ)1/(n1)forQ >1.

4. Weak Harnack inequalities and local integrability for quasisuperharmonic functions

(Weak) Harnack inequalities for quasi(super)minimizers inRn were obtained by DiBenedetto and Trudinger [15, Corollaries 1–3]. In metric spaces they were given by Kinnunen and Shanmugalingam [30, Theorem 7.1 and Corol- lary 7.3]. Some necessary modifications of the proofs and results in [30] were provided in Section 10 in A. Björn and Marola [9].

Example 10.1 in [9] or Example 8.19 in Björn and Björn [8] also shows that (weak) Harnack inequalities (both for quasi(super)minimizers and for (super)minimizers) in metric spaces can only hold on ballsBsuch that a sufficiently large blow-up of B (depending onX) lies in Ω. This is usually formulated ascλBΩ, whereλ is the dilation constant in the weak Poincaré inequality andcis an absolute constant, such as 5, 20 or 50, depending on the proof. In (weighted)Rnone can of course takeλ=1.

In this and the next section we will see how one can improve upon the blow-up constant in various Harnack and Caccioppoli inequalities. The price one has to pay is however that the conditions involve the quasiconvexity constantL, see below, instead ofλ. It is therefore relevant to discuss the relationship betweenLandλ. We postpone this discussion to Section 6.

A metric spaceY isL-quasiconvexif for allx, yY, there is a curveγ: [0, lγ] →Y withγ (0)=xandγ (lγ)=y, parameterized by arc length, such thatlγ Ld(x, y). Under our assumptions,Xis quasiconvex, see Section 6 for more details.

In this section, we formulate the (weak) Harnack inequality for quasi(super)minimizers as follows. The proof below also shows that the constant 2Lcan be replaced by(1+ε)Lfor anyε >0. By Kinnunen and Shanmugalingam [30]

and A. Björn and Marola [9], similar inequalities hold withLreplaced byλand the requirement that a larger blow-up of the balls lies inΩ.

Proposition 4.1. Assume that X is L-quasiconvex, ΩX and Q1. Then there exist constants C and s >0, depending only onQ,p,L,Cμ and the constants in the weak(1, p)-Poincaré inequality, such that for all ballsB with2LB⊂Ω,

(8)

(a) wheneveru >0is aQ-quasiminimizer inΩ, the following Harnack inequality holds:

ess sup

B

uCess inf

B u; (4.1)

(b) wheneverv >0is aQ-quasisuperminimizer inΩ, the following weak Harnack inequality holds:

B

vs 1/s

Cess inf

B v. (4.2)

As in Proposition 3.5, it can be shown that (4.2) holds also for quasisuperharmonic functions, in which case ess inf may be replaced by inf as quasisuperharmonic functions are lower semicontinuously regularized.

Proof of Proposition 4.1. By changingu and v on a set of capacity zero, if necessary, we can assume that u is continuous andvis lower semicontinuously regularized inΩ. It is thus possible to replace ess sup and ess inf by sup and inf. LetB=B(x0, r)and assume that 2LB⊂Ω. By Theorem 4.32 in Björn and Björn [8],Xsupports a weak (1, p)-Poincaré inequality with dilation constantλ=L. Thus, (4.1) and (4.2) hold on all ballsBsuch thatcLBΩ for some fixedc >1. See e.g. Kinnunen and Shanmugalingam [30] and A. Björn and Marola [9].

(a) Letε >0 be arbitrary and findxBsuch thatu(x)infBu+ε. By theL-quasiconvexity ofX, there exists a curveγ: [0, lγ] →X, parameterized by arc length, such thatlγLd(x, x0) < Lr,x0=γ (0)andx=γ (lγ).

Letxj=γ (j r/c)and coverγ by the ballsBj=B(xj, r/c),j=0,1, . . . , N, withN being the largest integer such that N r/clγ. ThenBjBj+1 is nonempty forj =0,1, . . . , N−1. Note thatN < cLandcLBj ⊂2LB⊂Ω. Hence, the Harnack inequality for quasiminimizers holds on eachBj and we have for everyj=0,1, . . . , N−1,

infBj

u inf

BjBj+1

usup

Bj+1

uC0 inf

Bj+1

u.

Iterating this estimate, we obtain infB0

uC0Ninf

BN

uC0Nu(x)C0N

infB u+ε

and lettingε→0 yields u(x0)C0inf

B0

uC0N+1inf

B u. (4.3)

Similarly, choosingyB such thatu(y)supBuεgives u(x0)C0N1sup

B

u,

which together with (4.3) proves (4.1).

(b) First, there exist ballsB1, . . . , Bm with centres inBand radiir/3csuch thatBm

j=1Bj and the balls 12Bj are pairwise disjoint. It follows from the doubling property ofμthat the numbermof these balls does not exceed a constant depending only oncandCμ. In particular, the bound formis independent ofx0andr. Moreover, for all j =1,2, . . . , m, we have 3cLBj ⊂2LB⊂Ωandμ(B)/Cμ(Bj)Cμ(B).

LetB=B(x, r/3c)be one of these balls and connectxtox0by a curve of length at mostLd(x, x0). As in (a), choosezB such thatu(z)infBv+εand connectx0tozby a curve of length at mostLd(x0, z). Adding these two curves gives a connecting curveγ fromxtozof length less than 2Lr and such thatγLB. Coverγ by balls Bj,j =0,1, . . . , N6cL, with radiir/3cas in (a) and note that 3cLBj⊂2LB ⊂Ω andBj+1⊂3Bj for each j =0,1, . . . , N. Hence, the weak Harnack inequality for quasisuperminimizers holds on each 3Bj, implying that

infBjv

Bj

vs 1/s

Cμ2

3Bj

vs 1/s

C0inf

3BjvC0 inf

Bj+1v.

Iterating this estimate, we obtain, asB=B0,

(9)

infB v=inf

B0vC0Ninf

BNvC0Nv(z)C0N

infB v+ε and lettingε→0 yields

B

vs C0inf

B

v s

μ B

C0N+1inf

B v s

μ B

.

Summing up over all ballsB=Bj coveringBfinishes the proof. 2

A sharp version of the weak Harnack inequality, due to Trudinger [43], is as follows. Assume thatu0 is a superminimizer in an open setΩRn. If 0< s < (p−1), then for every ballB with 6B⊂Ωwe have

B

usdx 1/s

Csess inf

3B u, (4.4)

where=n/(np)if 1< p < n, and= ∞ifpnin unweightedRn. In the metric space case, this is due to Kinnunen and Martio [29] and >1 is chosen so thatX supports a weak(p, p)-Poincaré inequality. The proof is strongly based on the fact that the Caccioppoli inequality (3.1) holds for allγ < p−1 whenQ=1. The initial requirement onBin metric spaces is that 60λB⊂Ω, see A. Björn and Marola [9, Section 10], or Björn and Björn [8]

(the latter gives the condition 150λB⊂Ω), but as in Proposition 4.1, it can be shown that if X isL-quasiconvex, then (4.4) holds for all ballsBwith 6LB⊂Ω (or even(3+ε)LBΩ for every fixedε >0). Example 10.1 in [9]

or Example 8.19 in [8] shows that the dilation constantλor the quasiconvexity constantL are needed in the weak Harnack inequality.

Arguing as in the proof of Proposition 3.5 it is easy to see that all nonnegativeQ-quasisuperminimizers satisfy (4.4) if and only if all nonnegativeQ-quasisuperharmonic functions satisfy the inequality with the same positive constants s andCs. (As quasisuperharmonic functions are lower semicontinuously regularized it is also equivalent to replace ess inf by inf in the quasisuperharmonic case.)

Letζ0=ζ0(Q, p)be the largest number (independent ofXandΩ) such that for every positives < ζ0there is a constantCs=Cs(Q, p, X, Ω)such that (4.4) holds for allQ-quasisuperharmonic functionsu0. We then have the following consequence of Theorem 1.2.

Proposition 4.2.ζ0(Q, p)δ(Q, p), whereδ(Q, p)is as in(3.2).

By (4.4) we know thatζ0(1, p)=δ(1, p)=p−1, and that this is valid also in metric spaces.

If we fix a metric spaceX(e.g.Rn) andQandp, then, by Proposition 4.1, there is somes >0 such that the weak Harnack inequality (4.4) holds. By the Hölder inequality, it holds for all 0< ssand henceζ0(Q, p, X)s >0.

The proof of the weak Harnack inequality shows that the exponentsonly depends onp,Q,Cμand the constants in the weak(1, p)-Poincaré inequality. Proposition 4.2 however suggests that the upper bound forsonly depends onp andQ, and that the only dependence onXis through. It therefore feels natural to make the following conjecture.

Conjecture 4.3.Assume thatXsupports a weak(p, p)-Poincaré inequality. LetQ >1and 0< s < δ(Q, p). Then there is a constantCs such that the weak Harnack inequality(4.4)holds for everyQ-quasisuperharmonic function u0inΩand every ballBsuch that6LB⊂ΩorcλBΩ.

It is worth observing thatXsupports a weak(p, p)-Poincaré inequality if and only if it supports a weak(1, p)- Poincaré inequality and there is a constantCsuch that for all ballsB=B(x, r)XandB=B(x, r), withxB andrr, the estimate

μ(B) μ(B) C

r r

σ

holds withσ=p/(−1), see Hajłasz and Koskela [22] and Franchi, Gutiérrez and Wheeden [17], or Björn and Björn [8].

(10)

Proof of Proposition 4.2. Letsδ(Q, p)and recall that=n/(np)in unweightedRn,n > p. By Theorem 1.2, u(x)= |x|α, withα=(pn)/δ(Q, p), isQ-quasisuperharmonic inBRn. Then, withB=B(0,16),

B

usdx=C 1/6

0

rαsrn1dr= ∞,

as

αs+n−1αδ(Q, p)+n−1= −1.

Thus, the left-hand side in (4.4) is infinite while the right-hand side is finite, showing that (4.4) does not hold fors. 2 It is well known that the sharp weak Harnack inequality implies sharp local integrability results for superharmonic functions: ifuis superharmonic inΩ, thenuLsloc(Ω)for 0< s < (p−1), where the range forsis sharp.

We immediately get that ifu isQ-quasisuperharmonic inΩX thenuLsloc(Ω)for 0< s < ζ0(Q, p, X).

Moreover, by the proof of Proposition 4.2, we see that ifX=Rn andn > p, then there is aQ-quasisuperharmonic u /Lδ(Q,p)loc (Ω).

Conjecture 4.4.Assume thatXsupports a weak(p, p)-Poincaré inequality. LetQ >1and 0< s < δ(Q, p). Then everyQ-quasisuperharmonic function inΩ belongs toLsloc(Ω).

This conjecture follows directly from Conjecture 4.3. In fact Conjecture 4.3 follows from Conjecture 3.2; to show this one essentially needs to repeat the arguments in Kinnunen and Martio [29].

5. Logarithmic Caccioppoli inequality and BMO

The following proposition is the logarithmic Caccioppoli inequality for superminimizers which plays a crucial role in the proof of the (weak) Harnack inequality using the Moser method. In metric spaces, it was originally proved in Kinnunen and Martio [29] and follows easily from (3.1) withγ=0 and a suitable choice of test function.

Proposition 5.1.Assume thatu >0is a superminimizer inΩ which is locally bounded away from0. Then for every ballBwith2B⊂Ωwe have

B

glogup C

diam(B)p. (5.1)

Proposition 5.1 implies, together with a Poincaré inequality, that the logarithm of a positive superminimizer has bounded mean oscillation. This is needed in the Moser iteration to show the weak Harnack inequality for supermini- mizers.

We have not been able to prove the inequality (5.1) for quasisuperminimizers, and it is therefore not clear whether the Moser iteration runs for quasiminimizers. See however Lemma 5.8 below. In view of Proposition 3.1, the possible proof of (5.1) for quasisuperminimizers should be based on some other method than the proof for superminimizers.

In any case, the constantCin the logarithmic Caccioppoli inequality for quasisuperminimizers would have to depend on Qand grow at least asQp/(p1), see Example 5.7 below. We know, however, that the logarithm of a positive quasisuperminimizer belongs both to BMO and toNloc1,q for allq < p, qualitatively, see Theorems 5.5 and 5.9 below.

It is also rather interesting to observe that ifu >0 is a quasiminimizer then (5.1) follows from the Caccioppoli inequality (3.1) forγ <0 (recall that by Proposition 7.3 in A. Björn and Marola [9], it is true for allγ <0) and from the Harnack inequality.

Proposition 5.2. Assume that u >0 is a Q-quasiminimizer in Ω. Then inequality(5.1)holds true for every ball B=B(x, r)with2B⊂Ω.

(11)

The proof below shows that the factor 2 in the condition 2B⊂Ωcan be replaced by 1+εfor anyε >0.

Proof of Proposition 5.2. Letc >1 be such that the Harnack inequality for quasiminimizers holds on all balls B withcλBΩ, whereλis the dilation constant from the weak(1, p)-Poincaré inequality, see the discussion at the beginning of Section 4.

As in the proof of Proposition 4.1, coverBby ballsB1, . . . , Bmwith centres inBand radiir/cλ, so that the number mof these balls does not exceed a constant depending only onc,λandCμ. Moreover,μ(B)/Cμ(Bj)Cμ(B) forj =1,2, . . . , m.

By construction, cλBjΩ, and hence, by the Harnack inequality for quasiminimizers, there is a constantC depending only onQ,Cμand the constants in the weak(1, p)-Poincaré inequality, such that for allj=1,2, . . . , m,

sup

Bj

uCinf

Bj

u.

Fixj ∈ {1, . . . , m}for the moment. It follows thatuhas to be bounded away from 0 inBj. Letη∈Lipc(2Bj)so that 0η1,η=1 onBj, andgη4/diam(Bj). Let alsoγ <0 and recall that (3.1) holds forγ. By the Harnack inequality, and the doubling property ofμ, we have

Bj

glogp u=

Bj

gupuγpηpuγ sup

Bj

u γ

Bj

gupuγpηp

Cγ

Cinf

Bju γ

2Bj

gηpuγdμC

2Bj

gηp=Cμ(Bj) (r/cλ)p.

Summing up over allBjand using the fact thatμ(Bj)is comparable toμ(B)for allj=1,2, . . . , m, gives the desired logarithmic Caccioppoli inequality onB. 2

We want to remark that in the same way, it can be shown that ifu >0 is a quasiminimizer, thenusatisfies the Caccioppoli inequality (3.1) with arbitrary exponentγ.

Proposition 5.3.Assume thatXisL-quasiconvex and letu >0be aQ-quasiminimizer inΩ. Then the Caccioppoli inequality(3.1)holds for allγRand allη∈Lipc(Ω)such thatsuppηB for some ballB⊂2LB⊂Ω, with a constant independent ofu,ηandB.

This follows easily from Proposition 4.1 and the Caccioppoli inequality forγ <0, as in the proof of Proposition 5.2.

The blow-up constant 2Lcan be replaced bycλ, whereλis the dilation constant in the weak Poincaré inequality andc >1 is such that the Harnack inequality (4.1) for quasiminimizers holds on all ballsBwithcλBΩ, see the discussion at the beginning of Section 4. In (weighted)Rnone can clearly takeL=λ=1.

Forγ p, Proposition 5.3 also follows from the Caccioppoli inequality for quasisubminimizers established in A. Björn and Marola [9, Proposition 7.2] in the metric space setting. (Forγ=pthis was earlier obtained by Tolksdorf [42, Theorem 1.4] in unweightedRn and by A. Björn [2, Theorem 4.1] for metric spaces.) For subminimizers, i.e.

whenQ=1, this also follows forγ > p−1, by the Caccioppoli inequality in Marola [33, Lemma 4.1]. In all these cases one can allow for suppηΩ, not only suppηB. We do not know whether this is possible in Proposition 5.3.

Before stating the main result of this section, we recall that a locally integrable functionf :ΩRbelongs to BMO(Ω)or BMOτ-loc(Ω),τ1, if there is a constantCsuch that

B

|ffB|dμC (5.2)

for all ballsBΩ orBτ BΩ, respectively. The smallest boundCfor which this inequality is satisfied is said to be the BMO-norm (resp. BMOτ-loc-norm) off, and is denoted byfBMO(Ω)(resp.fBMOτ-loc(Ω)).

A locally integrable functionw >0 is anA1-weightinΩ, if there is a constantAsuch that

B

w dxAess inf

B w

(12)

for all ballsBΩ. The least boundAis called theA1-constantofw. An essential feature ofA1-weights is that the average oscillation of its magnitude on every ball is uniformly controlled. A precise version of this is the following theorem which can be found in García-Cuerva and Rubio de Francia [18, Theorem 3.3, p. 157] for unweightedRn. The proof therein holds true also in the metric setting. We refer also to [18] and Duoandikoetxea [16] for more properties ofA1-weights and BMO.

Theorem 5.4. Ifwis anA1-weight inΩ, thenlogw∈BMO(Ω)with a norm depending only on the A1-constant forw.

Theorem 5.5.Assume thatXisL-quasiconvex. Letu >0be aQ-quasisuperminimizer or aQ-quasisuperharmonic function inΩ. Thenlogu∈BMOτ-loc(Ω)withτ =4L. Moreover

loguBMOτ-loc(Ω)< C,

whereConly depends onQ,p,Cμand the constants in the weak(1, p)-Poincaré inequality.

The blow-up constantτ =4Lcan be replaced byτ =2cλ, whereλis the dilation constant in the weak Poincaré inequality andc >1 is such that the weak Harnack inequality (4.2) for quasisuperminimizers holds on all ballsB withcλBΩ, see the discussion at the beginning of Section 4.

Proof of Theorem 5.5. LetBΩ be a ball inΩ such that 4LB⊂Ω. By Proposition 4.1 (or its modification for quasisuperharmonic functions), there iss >0 such that

B

usdxCess inf

B us (5.3)

for all ballsB⊂2LBΩ, and in particular for all subballsBofB. (Observe that ifB(x, r)B(x, r)=X, then it can happen thatr < r2rbut it is impossible to haver>2r.) Hence,usis anA1-weight inBwithA1-constantC. Thus, Theorem 5.4 implies that slogu=logus ∈BMO(B), with BMO-constant only depending onC. As this holds, with the same constant, for all balls B such that 4LB ⊂Ω, we find that logu∈BMOτ-loc(Ω), with the BMOτ-loc-norm only depending on C ands, which in turn only depend onQ,p,L,Cμand the constants in the weak(1, p)-Poincaré inequality. 2

Note that in (weighted)Rn, BMO(Ω)=BMOτ-loc(Ω)(with comparable norms), so that we can takeτ =1 in this case. This follows from the proof of Hilfssatz 2 in Reimann and Rychener [39, pp. 4, 13–17] for unweightedRn. This is also true for length metric spacesX (i.e. withL=1+εfor everyε >1), see Maasalo [32, Theorem 2.2]. For general metric spaces, however, the following example shows thatτ is essential in Theorem 5.5, that forL18 one cannot takeτ 19L, and thatτ (L)=4Lin Theorem 5.5 has the right growth asL→ ∞. Moreover, this is so even forQ=1.

Example 5.6.Let 0< θ16and

X=R2\ (x, y): 0< x <1 and 0< y < θ x .

By J. Björn and Shanmugalingam [13, Theorem 4.4], X supports a weak(1,1)-Poincaré inequality. Also,XisL- quasiconvex withL=(1+√

1+θ2)/θ <3/θ.

The ballsBj=B((2j,0),2j/3),j=1,2, . . ., are pairwise disjoint. Let now G=(0,1)×(−1,0], Ω=G

j=1

Bj, u(x, y)=

1, inG,

j, inBj\G, j=1,2, . . . .

As u is constant in every component of Ω it is a quasiminimizer. Assume that τ <1/3θ and let Bj =(1/τ )Bj, j =1,2, . . ., andv=logu. Then

(13)

Bj

|vvB

j|→ ∞, asj→ ∞,

showing thatv /∈BMOτ-loc(Ω)forτ <1/3θ.

Moreover, we have the following example.

Example 5.7.Letp < nandα(pn)/(p−1)=β. Letk >0 and u(x)=min{|x|β, k}. By Theorem 1.2, the function|x|α isQ-quasisuperharmonic inBRnwith the best quasisuperminimizer constantQdepending onα. For everyk >0, the functionv=uα/β=min{|x|α, kα/β}belongs toW1,p(B)and is thus aQ-quasisuperminimizer inB.

We then have logv=

α

βlogk, if|x|k1/β,

αlog|x|, otherwise, and |∇logv| =

0, if|x|k1/β, α/|x|, otherwise.

It follows that forr >2k1/β,

B(0,r)

|∇logv|pdx= C rn

r

k1/β

|α| ρ

p

ρn1

=C|α|p rn

rnpk(np)/β

C|α|p rp ,

where the constantCdepends only onnandp. Corollary 3.4 shows that|α|is comparable toQ1/(p1). Hence, the constant in the logarithmic Caccioppoli inequality for quasisuperminimizers, if it holds true, must depend onQand grow at least asQp/(p1).

At the same time,uis a superminimizer inBand thus logu∈BMO(B). Note that logv=(α/β)loguand hence logvBMO(B)=(α/β)loguBMO(B). As|α|is comparable toQ1/(p1), by Corollary 3.4, this shows that the con- stantCin Theorem 5.5 must depend onQand grow at least asQ1/(p1).

We finish this section by showing that the logarithm of a positive quasisuperminimizer belongs toNloc1,q for every q < p. For superminimizers, this is known even forq=p, and follows directly from Proposition 5.1. We start by proving a weak version of Proposition 5.1 for functions satisfying a Caccioppoli inequality.

Lemma 5.8.LetB=B(x, r)and assume thatu0satisfies the Caccioppoli inequality

B

up+γgupdμCγ rp

2B

uγ (5.4)

for allγ <0. Assume moreover that for someσ >0,

2B

uσ

2B

uσdμC0. (5.5)

Then for allq < p,

B

guq

uqdμCμC10q/pCγq/p

rq ,

whereγ= −σ (pq)/qandCμis the doubling constant ofμ.

Proof. We can assume thatq > p/2. By the Hölder inequality and the doubling property ofμ, we have for allε >0,

B

guq

uq= −

B

uqεguquε

B

u(qε)p/qgup q/p

Cμ

2B

uεp/(pq) 1q/p

. (5.6)

Références

Documents relatifs

Keywords: Poincaré inequality; Concentration of measure; Transportation-cost inequalities; Inf-convolution inequalities; Logarithmic-Sobolev inequalities; Super Poincaré

First we recall that if ( M , ρ) is a complete Riemannian manifold with Ricci curvature bounded from below, then M is a locally doubling metric space with respect to the

LIN, Local behavior of singular positive solutions of semilinear elliptic equations with Sobolev exponent, Duke Math. LIN, On compactness and completeness of conformal

In particular, for every E every A-harmonicfunction is continuous and Harnack inequalities hold for positive 1-t-harmonic function on every domain in X... Keuntje [Keu90]

The final step in obtaining the coarea formula for mappings defined on an H n -rectifiable metric space is the proof of the area formula on level sets:.. Theorem 7 (The Area Formula

In the case of the Scalar curvature equation and in di- mension 2 Shafrir used the isoperimetric inequality of Alexandrov to prove an inequality of type sup + inf with only L

On met- ric measure spaces, for instance, under some curvature lower bound assumptions, Sturm [39], Rajala [36], Erbar, Kuwada and Sturm [19] and Jiang, Li and Zhang [26] studied

We recover, in particular, well-known results on the right invertibility of the divergence in Sobolev spaces when Ω is Lipschitz or, more generally, when Ω is a John domain, and