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

Bounded weak solutions to elliptic PDE with data in Orlicz spaces

David Cruz-Uribe, OFS and Scott Rodney David Cruz-Uribe, OFS
Dept. of Mathematics
University of Alabama
Tuscaloosa, AL 35487, USA
[email protected] Scott Rodney
Dept. of Mathematics, Physics and Geology
Cape Breton University
Sydney, NS B1Y3V3, CA
[email protected]
(Date: November 29, 2020)
Abstract.

A classical regularity result is that non-negative solutions to the Dirichlet problem Δu=f\Delta u=f in a bounded domain Ω\Omega, where fLq(Ω)f\in L^{q}(\Omega), q>n2q>\frac{n}{2}, satisfy uL(Ω)CfLq(Ω)\|u\|_{L^{\infty}(\Omega)}\leq C\|f\|_{L^{q}(\Omega)}. We extend this result in three ways: we replace the Laplacian with a degenerate elliptic operator; we show that we can take the data ff in an Orlicz space LA(Ω)L^{A}(\Omega) that lies strictly between Ln2(Ω)L^{\frac{n}{2}}(\Omega) and Lq(Ω)L^{q}(\Omega), q>n2q>\frac{n}{2}; and we show that that we can replace the LAL^{A} norm in the right-hand side by a smaller expression involving the logarithm of the “entropy bump” fLA(Ω)/fLn2(Ω)\|f\|_{L^{A}(\Omega)}/\|f\|_{L^{\frac{n}{2}}(\Omega)}, generalizing a result due to Xu.

Key words and phrases:
Orlicz spaces, degenerate elliptic equations, bounded solutions, a priori estimates
1991 Mathematics Subject Classification:
35B45, 35D30, 35J25, 46E30
D. Cruz-Uribe is supported by  research funds from the Dean of the College of Arts & Sciences, the University of Alabama. S. Rodney is supported by the NSERC Discovery Grant program. The authors would like to thank Andrea Cianchi for calling to our attention several important references.

1. Introduction

In this paper, we consider boundedness properties of weak (sub)solutions to the following Dirichlet problem:

(1.3) {Div(Qu)=fvfor xΩu=0for xΩ\displaystyle\left\{\begin{array}[]{rcll}-\operatorname{Div}\left(Q\nabla u\right)&=&fv&\textrm{for }x\in\Omega\\ u&=&0&\textrm{for }x\in\partial\Omega\end{array}\right.

Throughout this paper, Ω\Omega is a bounded domain (i.e., an open and connected subset) of n{\mathbb{R}^{n}} with n3n\geq 3, Q=Q(x)Q=Q(x) is a non-negative definite, symmetric, measurable matrix with QLloc1(Ω)Q\in L^{1}_{loc}(\Omega), vv is a weight (i.e., a non-negative, measurable function) on Ω\Omega such that vL1(Ω)v\in L^{1}(\Omega), and the data function ff is in Lloc1(Ω)L^{1}_{loc}(\Omega).

When QQ is a uniformly elliptic matrix and v(x)=1v(x)=1, it is a classical result (See Mazya [16, 17], Stampacchia [23] and Trudinger [25, Theorem 4.1][12, Theorem 8.16]) that there is a constant C>0C>0 such that if fLq(Ω)f\in L^{q}(\Omega), then

uL(Ω)CfLq(Ω)\|u\|_{L^{\infty}(\Omega)}\leq C\|f\|_{L^{q}(\Omega)}

for any non-negative weak subsolution uH1(Ω)u\in H^{1}(\Omega) of (1.3) provided q>n2q>\frac{n}{2}. Moreover, a counter-example shows that this bound is sharp even for the Laplacian and we cannot take q=n2q=\frac{n}{2}.

The standard proof of this result uses the classical Sobolev inequality,

ψL2nn2(Ω)CψL2(Ω),\|\psi\|_{L^{\frac{2n}{n-2}}(\Omega)}\leq C\|\nabla\psi\|_{L^{2}(\Omega)},

valid for any ψH01(Ω)\psi\in H^{1}_{0}(\Omega), combined with Moser iteration. The restriction q>n2q>\frac{n}{2} is naturally connected to the classical Sobolev gain factor σ=nn2=(n2)\sigma=\frac{n}{n-2}=\left(\frac{n}{2}\right)^{\prime}.

The goal of this paper is to generalize this result. First we show that this estimate can be improved by replacing the space Lq(Ω)L^{q}(\Omega) by an Orlicz space LA(Ω)L^{A}(\Omega) that lies strictly between Ln2(Ω)L^{\frac{n}{2}}(\Omega) and Lq(Ω)L^{q}(\Omega) for q>n2q>\frac{n}{2}. For brevity, we will defer many definitions to Section 2 below.

Theorem 1.4.

Let QQ be a uniformly elliptic matrix, and let fLA(Ω)f\in L^{A}(\Omega), where A(t)=tn2log(e+t)qA(t)=t^{\frac{n}{2}}\log(e+t)^{q}, q>n2q>\frac{n}{2}. Then there exists a constant C=C(n,q,Q)C=C(n,q,Q) such that, given any non-negative weak subsolution uu of

{Div(Qu)=ffor xΩ,u=0for xΩ,\displaystyle\left\{\begin{array}[]{rcll}-\operatorname{Div}\left(Q\nabla u\right)&=&f&\textrm{for }x\in\Omega,\\ u&=&0&\textrm{for }x\in\partial\Omega,\end{array}\right.

we have the estimate

uL(Ω)CfLA(Ω).\|u\|_{L^{\infty}(\Omega)}\leq C\|f\|_{L^{A}(\Omega)}.
Remark 1.5.

After completing this paper we learned that a somewhat more general version of Theorem 1.4 was proved by Cianchi [4, Theorem 5] using very different methods.

We will prove Theorem 1.4 as a special case of a more general result for solutions of (1.3) that holds for a much larger class of matrices QQ. We can allow QQ to be both degenerate and singular, but must impose some restrictions on the largest and smallest eigenvalues. We encode these restrictions in two assumptions on the integrability of the largest eigenvalue and on the existence of an L2L^{2} Sobolev inequality with gain. We state them together as a general hypothesis.

Hypothesis 1.6.

Given the matrix QQ and the weight vL1(Ω)v\in L^{1}(\Omega), assume that for some constant k>0k>0,

|Q(x)|op=sup{|Q(x)ξ|:ξn,|ξ|=1}kv(x)a.e.|Q(x)|_{op}=\sup\{|Q(x)\xi|:\xi\in{\mathbb{R}^{n}},|\xi|=1\}\leq kv(x)\;\text{a.e.}

Moreover, assume that there exist constants σ=σ(n,Q,v,Ω)>1,C01\sigma=\sigma(n,Q,v,\Omega)>1,~{}C_{0}\geq 1 such that for every ψLip0(Ω)\psi\in Lip_{0}(\Omega)

(1.7) (Ω|ψ(x)|2σv(x)𝑑x)12σC0(Ω|Q(x)ψ(x)|2𝑑x)12.\left(\int_{\Omega}|\psi(x)|^{2\sigma}~{}v(x)dx\right)^{\frac{1}{2\sigma}}\leq C_{0}\left(\int_{\Omega}\left|\sqrt{Q(x)}\nabla\psi(x)\right|^{2}~{}dx\right)^{\frac{1}{2}}.

The first assumption, that |Q|opkv|Q|_{op}\leq kv, in Hypothesis 1.6 is necessary to prove many of the necessary properties of weak derivatives in the corresponding degenerate Sobolev space. The second assumption, that inequality (1.7) holds, in Hypothesis 1.6 reduces to the classical Sobolev inequality if QQ is a uniformly elliptic matrix in Ω\Omega and σ=nn2=(n2)\sigma=\frac{n}{n-2}=(\frac{n}{2})^{\prime}. It allows us to perform the necessary De Giorgi iteration.

These assumptions hold in several important special cases. If vv satisfies the Muckenhoupt A2A_{2} condition,

[v]A2=supB1|B|Bv(x)𝑑x1|B|Bv(x)1𝑑x<,[v]_{A_{2}}=\sup_{B}\frac{1}{|B|}\int_{B}v(x)\,dx\frac{1}{|B|}\int_{B}v(x)^{-1}\,dx<\infty,

where the supremum is taken over all balls BB in n{\mathbb{R}^{n}}, and if QQ satisfies the degenerate ellipticity condition

λv(x)|ξ|2Q(x)ξ,ξΛv(x)|ξ|2,\lambda v(x)|\xi|^{2}\leq\langle Q(x)\xi,\xi\rangle\leq\Lambda v(x)|\xi|^{2},

where 0<λΛ<0<\lambda\leq\Lambda<\infty, then |Q|opΛv|Q|_{op}\leq\Lambda v and (1.7) holds. (See [10].) More generally, suppose that uu and vv are a pair of weights such that u(x)v(x)u(x)\leq v(x) a.e., vv satisfies a doubling condition, uA2u\in A_{2}, and there exists σ>1\sigma>1 such that given any balls B1B2ΩB_{1}\subset B_{2}\subset\Omega,

r(B1)r(B2)(v(B1)v(B2))12σC(u(B1)u(B2))12,\frac{r(B_{1})}{r(B_{2})}\bigg{(}\frac{v(B_{1})}{v(B_{2})}\bigg{)}^{\frac{1}{2\sigma}}\leq C\bigg{(}\frac{u(B_{1})}{u(B_{2})}\bigg{)}^{\frac{1}{2}},

and QQ satisfies the degenerate ellipticity condition

u(x)|ξ|2Q(x)ξ,ξv(x)|ξ|2,u(x)|\xi|^{2}\leq\langle Q(x)\xi,\xi\rangle\leq v(x)|\xi|^{2},

then |Q|opv|Q|_{op}\leq v and we have the Sobolev inequality

(Ω|ψ(x)|2σv(x)𝑑x)12σC0(Ω|ψ(x)|2u(x)𝑑x)12,\left(\int_{\Omega}|\psi(x)|^{2\sigma}~{}v(x)dx\right)^{\frac{1}{2\sigma}}\leq C_{0}\left(\int_{\Omega}\left|\nabla\psi(x)\right|^{2}~{}u(x)dx\right)^{\frac{1}{2}},

so again  (1.7) holds. (See [1].)

Remark 1.8.

In [7], the authors and Rosta proved that when v=1v=1, with minor additional hypotheses the global Sobolev inequality (1.7) follows from a weaker, local Sobolev inequality,

(1|B|B|ψ(x)|2σ𝑑x)12σC[r(B)|B|B|Qψ(x)|2𝑑x+1|B|B|ψ(x)|2𝑑x]12,\left(\frac{1}{|B|}\int_{B}|\psi(x)|^{2\sigma}\,dx\right)^{\frac{1}{2\sigma}}\leq C\bigg{[}\frac{r(B)}{|B|}\int_{B}|\sqrt{Q}\nabla\psi(x)|^{2}\,dx+\frac{1}{|B|}\int_{B}|\psi(x)|^{2}\,dx\bigg{]}^{\frac{1}{2}},

that holds for all (sufficiently small) balls BΩB\subset\Omega.

We can now state our main result, which is a generalization of Theorem 1.4 to degenerate elliptic operators. Again, for precise definitions see Section 2.

Theorem 1.9.

Given a weight vv and the non-negative definite, symmetric matrix QQ, suppose that Hypothesis 1.6 holds for some σ>1\sigma>1. Let A(t)=tσlog(e+t)qA(t)=t^{\sigma^{\prime}}\log(e+t)^{q} where q>σq>\sigma^{\prime}. If fLA(v;Ω)f\in L^{A}(v;\Omega), then any non-negative weak subsolution 𝐮=(u,u)QH01(v;Ω){\bf u}=(u,\nabla u)\in QH^{1}_{0}(v;\Omega) of (1.3) satisfies

(1.10) uL(v;Ω)CfLA(v;Ω),\|u\|_{L^{\infty}(v;\Omega)}\leq C\|f\|_{L^{A}(v;\Omega)},

where CC is independent of both 𝐮{\bf u} and ff.

We originally conjectured that the exponent qq in Theorem 1.9 is sharp in general, but we were not able to prove this or find a counter-example. We then learned that Cianchi [4, Theorem 5], as a consequence of a more general result, showed that in the classical setting when QQ is uniformly elliptic and σ=(n2)\sigma=(\frac{n}{2})^{\prime}, we must have that q>n21q>\frac{n}{2}-1. We round out his result by giving a simple counter-example proving that it is sharp for the Laplacian.

Example 1.11.

Let n3n\geq 3, and let Ω=B(0,1)\Omega=B(0,1). Then there exists a function fLA(Ω)f\in L^{A}(\Omega), where A(t)=tn2log(e+t)qA(t)=t^{\frac{n}{2}}\log(e+t)^{q}, q<n21q<\frac{n}{2}-1, such that the non-negative weak solution of the Poisson equation

{Δu=ffor xΩ,u=0for xΩ,\displaystyle\left\{\begin{array}[]{rcll}-\Delta u&=&f&\textrm{for }x\in\Omega,\\ u&=&0&\textrm{for }x\in\partial\Omega,\end{array}\right.

is unbounded.

Remark 1.12.

We conjecture that the sharp exponent in Theorem 1.9 is q>σ1q>\sigma^{\prime}-1. However, the bound q>σq>\sigma^{\prime} appears to be intrinsic to our proof, so either our proof needs to be refined or another approach is needed. We note that the proof in [4] relies on re-arrangement estimates and so does not readily extend to the case of degenerate operators.

Our second main result shows that inequality (1.10) can be sharpened so that the right-hand side only depends on the logarithm of the LAL^{A} norm.

Theorem 1.13.

Given a weight vv and a non-negative definite, symmetric matrix QQ, suppose that Hypothesis 1.6 holds for some σ>1\sigma>1. Let A(t)=tσlog(e+t)qA(t)=t^{\sigma^{\prime}}\log(e+t)^{q}, where q>σq>\sigma^{\prime}. If fLA(v;Ω)f\in L^{A}(v;\Omega), then any non-negative weak subsolution 𝐮=(u,u)QH01(v;Ω){\bf u}=(u,\nabla u)\in QH^{1}_{0}(v;\Omega) of (1.3) satisfies

(1.14) uL(v;Ω)CfLσ(v;Ω)(1+log(1+fLA(v;Ω)fLσ(v;Ω))),\|u\|_{L^{\infty}(v;\Omega)}\leq C\|f\|_{L^{\sigma^{\prime}}(v;\Omega)}\left(1+\log\left(1+\frac{\|f\|_{L^{A}(v;\Omega)}}{\|f\|_{L^{\sigma^{\prime}}(v;\Omega)}}\right)\right),

where CC is independent of both 𝐮{\bf u} and ff.

Theorem 1.13 generalizes the main result of Xu [26], but we note that there is a mistake in the statement of his main result. Working in the same setting as Theorem 1.4, he claims to show that

(1.15) uL(Ω)CfLn2(Ω)(1+log(1+fLq(Ω))),\|u\|_{L^{\infty}(\Omega)}\leq C\|f\|_{L^{\frac{n}{2}}(\Omega)}\left(1+\log\left(1+{\|f\|_{L^{q}(\Omega)}}\right)\right),

where q>n2q>\frac{n}{2}. However, a close examination of his proof shows that he only proves this when fLn2(Ω)1\|f\|_{L^{\frac{n}{2}}(\Omega)}\geq 1, and in fact what he proves is the analog of Theorem 1.13. It is straightforward to see that (1.15) cannot hold if fLn2(Ω)<1\|f\|_{L^{\frac{n}{2}}(\Omega)}<1; if it did, then if we fix ff and the corresponding solution uu, then we could apply this inequality to f/Nf/N (N>1N>1 large) and u/Nu/N. Then we could take the limit as NN\rightarrow\infty to conclude that uL(Ω)CfLn2(Ω)\|u\|_{L^{\infty}(\Omega)}\leq C\|f\|_{L^{\frac{n}{2}}(\Omega)}, which is false in general.

Remark 1.16.

The ratio fLA(v;Ω)/fLσ(v;Ω){\|f\|_{L^{A}(v;\Omega)}}/{\|f\|_{L^{\sigma^{\prime}}(v;\Omega)}} in Theorem 1.13 measures how much bigger the Orlicz norm is than the associated Lebesgue space norm. It is similar in spirit, though not in detail, to the “entropy bump” conditions introduced in the study of weighted norm inequalities in harmonic analysis [15, 24].

Our two main results are established via De Giorgi iteration on the level sets. De Giorgi’s original arguments are in [8] but more helpful descriptions are found in [2] and in [13], where De Giorgi iteration is applied in an infinitely degenerate elliptic regime. We were unable to adapt Moser iteration to work in the context of Orlicz norms, and it remains an open question whether such an approach is possible in this setting.

The remainder of the paper is organized as follows. In Section 2 we gather some preliminary results. We give a definition of Young functions and the associated Orlicz spaces, and record some useful properties. We then define weak solutions to the Dirichlet problem. This definition has to include the possibility that the matrix QQ can be both degenerate and singular, and we give it in terms of a degenerate Sobolev space, building upon results in [6] and elsewhere. We prove a number of properties of weak derivatives in this setting; we believe these results should be useful tools for other problems. We also prove that bounded, non-negative subsolutions of (1.3) must satisfy an exponential integrability condition. This result is a key lemma for the proof of Theorem 1.13 and is modeled on a similar result due to Xu [26] in the classical setting. For completeness we include the details of the proof. In Section 3 we prove Theorem 1.9; as noted above, the proof uses a version of De Giorgi iteration adapted to the scale of Orlicz spaces. This iteration argument was gotten by a careful adaptation of an argument due Korbenko, et al. [13, Section 4.2]. In Section 4 we prove Theorem 1.13; our proof is a generalization of the argument in [26] and requires us to deal with a number of technical obstacles. Finally, in Section 5 we construct Example 1.11.

2. Preliminaries

In this section we gather some preliminary definitions and results. We begin with some notation. The constant nn will always denote the dimension of the underlying space n{\mathbb{R}^{n}}. By CC, cc, etc. we will mean a constant that may change from appearance to appearance, but whose value depends only on the underlying parameters. If we want to specify this dependence, we will write, for instance, C(n,p)C(n,p), etc. If we write ABA\lesssim B, we mean that there exists a constant cc such that AcBA\leq cB. If ABA\lesssim B and BAB\lesssim A, we write ABA\approx B.

A weight vv will always be a non-negative, measurable function such that vL1(Ω)v\in L^{1}(\Omega). Given a set EΩE\subset\Omega, v(E)=Ev(x)𝑑xv(E)=\int_{E}v(x)\,dx. Given a weight vv, Lp(v;Ω)L^{p}(v;\Omega) is the collection of all those measurable functions g:Ωg:\Omega\rightarrow\mathbb{R} for which

gp=gLp(v;Ω)=(Ω|g(x)|pv(x)𝑑x)1/p<.\|g\|_{p}=\|g\|_{L^{p}(v;\Omega)}=\displaystyle\left(\int_{\Omega}|g(x)|^{p}~{}v(x)dx\right)^{1/p}<\infty.

Orlicz spaces

Our main hypothesis on the data function ff in (1.3) is that it belongs to the Orlicz space LA(v;Ω)L^{A}(v;\Omega). Here we gather some essential results about these spaces but we assume the reader has some familiarity with them. For complete information, we refer to [14, 20]. For a briefer summary, see [5, Chapter 5].

By a Young function we mean a function A:[0,)[0,)A:[0,\infty)\rightarrow[0,\infty) that is continuous, convex, strictly increasing, A(0)=0A(0)=0, and A(t)t\frac{A(t)}{t}\rightarrow\infty as tt\rightarrow\infty. Given a Young function AA, define LA(v;Ω)L^{A}(v;\Omega) to be the Banach space of measurable functions h:Ωh:\Omega\rightarrow\mathbb{R} equipped with the Luxembourg norm,

hA=hLA(v;Ω)=inf{λ>0:ΩA(|f(x)|λ)v(x)𝑑x1}<.\|h\|_{A}=\|h\|_{L^{A}(v;\Omega)}=\inf\bigg{\{}\lambda>0~{}:~{}\int_{\Omega}A\left(\frac{|f(x)|}{\lambda}\right)~{}v(x)dx\leq 1\bigg{\}}<\infty.

Given Young functions A,BA,\,B we can compare the associated norms by appealing to a point-wise estimate. We say that A(t)B(t)A(t)\preceq B(t) if there is a t0>0t_{0}>0 and a constant c1c\geq 1 depending only on A,BA,\,B so that A(t)B(ct)A(t)\leq B(ct) for tt0t\geq t_{0}. For a proof of the following result, see [14, Theorem 13.3] or [20, Section 5.1].

Lemma 2.1.

Given Young functions A,BA,\,B, if ABA\preceq B, then there exists a constant C=C(A,B,v(Ω))C=C(A,B,v(\Omega)) such that for every fLB(v;Ω)f\in L^{B}(v;\Omega),

fLA(v;Ω)CfLB(v;Ω).\|f\|_{L^{A}(v;\Omega)}\leq C\|f\|_{L^{B}(v;\Omega)}.

Given a Young function AA, we define the conjugate Orlicz function, A¯\bar{A}, via the pointwise formula

A¯(t)=sup{stB(s):s>0}.\bar{A}(t)=\sup\{st-B(s):s>0\}.

The pair A,A¯A,\,\bar{A} satisfy a version of Hölder’s inequality in the scale of Orlicz spaces. If fLA(v;Ω)f\in L^{A}(v;\Omega) and gLA¯(v;Ω)g\in L^{\bar{A}}(v;\Omega), then fgL1(v;Ω)fg\in L^{1}(v;\Omega) and

(2.2) Ω|f(x)g(x)|v(x)𝑑x2fAgA¯.\int_{\Omega}|f(x)g(x)|v(x)\,dx\leq 2\|f\|_{A}\|g\|_{\bar{A}}.

In our main results we consider Young functions of the form

(2.3) B(t)=tplog(e+t)q,B(t)=t^{p}\log(e+t)^{q},

where 1<p,q<1<p,\,q<\infty. The inverse and conjugate functions associated with these Young functions are well-known: see, for instance, [5]. We have that

(2.4) B¯(t)tplog(t)qtplog(e+t)q,\displaystyle\bar{B}(t)\approx\displaystyle\frac{t^{p^{\prime}}}{\log(t)^{q}}\approx\displaystyle\frac{t^{p^{\prime}}}{\log(e+t)^{q}},
(2.5) B¯1(t)t1/plog(e+t)qp,\displaystyle\bar{B}^{-1}(t)\approx t^{1/p^{\prime}}\log(e+t)^{\frac{q}{p}},

where the implicit constants depend on p,qp,\,q. As a consequence of Lemma 2.1 we have the following estimate which we will need below; details are straightforward and are omitted.

Lemma 2.6.

Let 1p1p2<1\leq p_{1}\leq p_{2}<\infty, 1q1q2<1\leq q_{1}\leq q_{2}<\infty and define

A(t)=tp1log(e+t)q1,B(t)=tp2log(e+t)q2.A(t)=t^{p_{1}}\log(e+t)^{q_{1}},\qquad B(t)=t^{p_{2}}\log(e+t)^{q_{2}}.

Then, given fLB(v;Ω)f\in L^{B}(v;\Omega),

fLp1(v,Ω)fLA(v;Ω)fLp2(v;Ω)fLB(v;Ω).\|f\|_{L^{p_{1}}(v,\Omega)}\lesssim\|f\|_{L^{A}(v;\Omega)}\lesssim\|f\|_{L^{p_{2}}(v;\Omega)}\lesssim\|f\|_{L^{B}(v;\Omega)}.

The implicit constants depend on pip_{i} and qiq_{i}, i=1, 2i=1,\,2, and v(Ω)v(\Omega).

We conclude this section with an estimate for the LB¯(Ω)L^{\bar{B}}(\Omega) norm of an indicator function 𝟙S\mathbbm{1}_{S} for SΩS\subset\Omega; this quantity plays an essential role in our proofs of Theorems 1.9 and 1.13. This computation is well-known, but to make clear the dependence on the constants we include its short proof.

Lemma 2.7.

Given the Young function BB defined by (2.3) then for any SΩS\subset\Omega, v(S)>0v(S)>0,

(2.8) 𝟙SLB¯(v;Ω)cv(S)1plog(1+v(S)1)qp,\|\mathbbm{1}_{S}\|_{L^{\bar{B}}(v;\Omega)}\leq\displaystyle\frac{cv(S)^{\frac{1}{p^{\prime}}}}{\log(1+v(S)^{-1})^{\frac{q}{p}}},

where c=c(p,q)>0c=c(p,q)>0.

Proof.

Given BB, B¯\bar{B} is defined by (2.4). Set

F={λ>0:ΩB¯(𝟙S(x)λ)v(x)𝑑x1}F=\left\{\lambda>0~{}:~{}\int_{\Omega}\bar{B}\left(\frac{\mathbbm{1}_{S}(x)}{\lambda}\right)v(x)\,dx\leq 1\right\}

and notice that FF\neq\emptyset. For each λF\lambda\in F,

v(S)B¯(1λ)=ΩB¯(𝟙S(x)λ)v(x)𝑑x1.v(S)~{}\bar{B}\left(\frac{1}{\lambda}\right)=\int_{\Omega}\bar{B}\left(\frac{\mathbbm{1}_{S}(x)}{\lambda}\right)v(x)\,dx\leq 1.

Since B¯\bar{B} is invertible and increasing,

λ[B¯1(1v(S))]1=m0>0.\lambda\geq\left[\bar{B}^{-1}\left(\frac{1}{v(S)}\right)\right]^{-1}=m_{0}>0.

Again by the invertibility of B¯\bar{B},

ΩB¯(𝟙S(x)m0)v(x)𝑑x=v(S)B¯(m01)=1.\int_{\Omega}\bar{B}\left(\frac{\mathbbm{1}_{S}(x)}{m_{0}}\right)v(x)\,dx=v(S)\bar{B}(m_{0}^{-1})=1.

Hence, m0Fm_{0}\in F, and it follows that 𝟙SLB¯(Ω)=m0\|\mathbbm{1}_{S}\|_{L^{\bar{B}}(\Omega)}=m_{0}. By inequality (2.5),

m0=B¯1(1v(S))c(p,q)v(S)1plog(e+v(S)1)qpm_{0}=\bar{B}^{-1}\left(\frac{1}{v(S)}\right)\geq c(p,q)v(S)^{-\frac{1}{p^{\prime}}}\log(e+v(S)^{-1})^{\frac{q}{p}}

and (2.8) follows. ∎

Degenerate Sobolev spaces and weak solutions

We now give a precise definition of weak (sub)solutions to the Dirichlet problem (1.3). This question has been explored in a number of papers by ourselves and others: see [3, 6, 7, 12, 18, 19, 21, 22]. Here we sketch the relevant details.

Given a non-negative definite, symmetric and measurable matrix function QQ on Ω\Omega and a weight vLloc1(Ω)v\in L^{1}_{loc}(\Omega), the solution space for the Dirichlet problem is the matrix weighted Sobolev space QH01(v;Ω)QH^{1}_{0}(v;\Omega). This space is defined as the abstract completion (in terms of Cauchy sequences) of the space Lip0(Ω)Lip_{0}(\Omega) (i.e., Lipschitz functions with compact support in Ω\Omega) with respect to the norm

ψQH01(v;Ω)=ψL2(v;Ω)+ψLQ2(Ω),\|\psi\|_{QH^{1}_{0}(v;\Omega)}=\|\psi\|_{L^{2}(v;\Omega)}+\|\nabla\psi\|_{L^{2}_{Q}(\Omega)},

where LQ2(Ω)L^{2}_{Q}(\Omega) is the Banach space of n\mathbb{R}^{n} vector-valued functions 𝐠{\bf g} on Ω\Omega that satisfy

𝐠LQ2(Ω)=(Ω|Q(x)𝐠(x)|2𝑑x)12<.\|{\bf g}\|_{L^{2}_{Q}(\Omega)}=\bigg{(}\int_{\Omega}|\sqrt{Q(x)}{\bf g}(x)|^{2}~{}dx\bigg{)}^{\frac{1}{2}}<\infty.

This norm is well defined for ψLip0(Ω)\psi\in Lip_{0}(\Omega) provided |Q|opLloc1(Ω)|Q|_{op}\in L_{loc}^{1}(\Omega); in particular, QH01(v;Ω)QH^{1}_{0}(v;\Omega) is well defined if the first assumption in Hypothesis 1.6 holds.

With this definition, the Sobolev space QH01(v;Ω)QH^{1}_{0}(v;\Omega) is a collection of equivalence classes of Cauchy sequences of Lip0(Ω)Lip_{0}(\Omega) functions. However, the spaces L2(v;Ω)L^{2}(v;\Omega) and LQ2(Ω)L^{2}_{Q}(\Omega) are complete: for a proof that LQ2(Ω)L^{2}_{Q}(\Omega) is complete, see [22] or [6] where it was proved that LQp(Ω)L^{p}_{Q}(\Omega) is complete for 1p<1\leq p<\infty. Therefore, to each equivalence class [{ψj}][\{\psi_{j}\}] in QH01(v;Ω)QH^{1}_{0}(v;\Omega) we can associate a unique pair 𝐮=(u,𝐠)L2(v;Ω)×LQ2(Ω){\bf u}=(u,{\bf g})\in L^{2}(v;\Omega)\times L^{2}_{Q}(\Omega) whose norm is given by

𝐮QH01(v;Ω)\displaystyle\|{\bf u}\|_{QH^{1}_{0}(v;\Omega)} =uL2(v;Ω)+𝐠LQ2(Ω)\displaystyle=\|u\|_{L^{2}(v;\Omega)}+\|{\bf g}\|_{L^{2}_{Q}(\Omega)}
=limj(ψjL2(v;Ω)+ψjLQ2(Ω)).\displaystyle=\displaystyle\lim_{j\rightarrow\infty}\left(\|\psi_{j}\|_{L^{2}(v;\Omega)}+\|\nabla\psi_{j}\|_{L^{2}_{Q}(\Omega)}\right).

Conversely, given a pair (u,𝐠)(u,{\bf g}) we will say that it is in QH01(v;Ω)QH^{1}_{0}(v;\Omega) if there exists a sequence {uj}jLip0(Ω)\{u_{j}\}_{j}\subset Lip_{0}(\Omega) such that (ujuj)(u_{j}\nabla u_{j}) converges to (u,𝐠)(u,{\bf g}) in L2(v;Ω)×LQ2(Ω)L^{2}(v;\Omega)\times L^{2}_{Q}(\Omega).

Hereafter, we will denote 𝐠{\bf g} by u\nabla u since 𝐠\bf g plays the role of a weak gradient of uu. However, while we adopt this formal notation we want to stress that the function 𝐠{\bf g} is not the weak gradient of uu in the sense of classical Sobolev spaces. For further details,  [6] contains the construction of QH1,p(v;Ω)QH^{1,p}(v;\Omega) for p1p\geq 1. Additionally, unweighted constructions of QH01,p(1;Ω)QH^{1,p}_{0}(1;\Omega) are found in [7, 19] for p1p\geq 1 and [18, 21] for p=2p=2.

We can extend the second assumption in Hypothesis 1.6 to functions in QH01(Ω)QH^{1}_{0}(\Omega); this follows from density of Lip0(Ω)Lip_{0}(\Omega) functions and we omit the proof.

Lemma 2.9.

Suppose the Sobolev inequality of Hypothesis 1.6 holds. Then,

(2.10) (Ω|w(x)|2σv(x)𝑑x)12σC0(Ω|Q(x)w|2𝑑x)12\left(\int_{\Omega}|w(x)|^{2\sigma}v(x)\,dx\right)^{\frac{1}{2\sigma}}\leq C_{0}\left(\int_{\Omega}\left|\sqrt{Q(x)}\nabla w\right|^{2}\,dx\right)^{\frac{1}{2}}

for every 𝐰=(w,w)QH01(v;Ω){\bf w}=(w,\nabla w)\in QH^{1}_{0}(v;\Omega) where C0C_{0} is the same as in Hypothesis 1.6.

We can now define the weak solution to the Dirichlet problem.

Definition 2.11.

A pair 𝐮=(u,u)QH01(Ω){\bf u}=(u,\nabla u)\in QH^{1}_{0}(\Omega) is said to be a weak solution of the Dirichlet problem (1.3) if

Ωψ(x)Q(x)u(x)𝑑x=Ωf(x)ψ(x)v(x)𝑑x\int_{\Omega}\nabla\psi(x)\cdot Q(x)\nabla u(x)\,dx=\int_{\Omega}f(x)\psi(x)v(x)\,dx

for every ψLip0(Ω)\psi\in Lip_{0}(\Omega). The pair is said to be a non-negative weak subsolution if u(x)0u(x)\geq 0 vv-a.e. and

Ωψ(x)Q(x)u(x)𝑑xΩf(x)ψ(x)v(x)𝑑x\int_{\Omega}\nabla\psi(x)\cdot Q(x)\nabla u(x)\,dx\leq\int_{\Omega}f(x)\psi(x)v(x)\,dx

for every non-negative ψLip0(Ω)\psi\in Lip_{0}(\Omega).

Note that if 𝐡=(h,h)QH01(v;Ω){\bf h}=(h,\nabla h)\in QH^{1}_{0}(v;\Omega) with h(x)0h(x)\geq 0 vv-a.e., then by a standard limiting argument we may use hh as our test function in Definition 2.11.

Remark 2.12.

The existence of weak solutions to (1.3) when v=1v=1 was studied in [12, 18, 21], and when v(x)=|Q(x)|opv(x)=|Q(x)|_{op} in [6, 7].

Properties of weak gradients

We now develop some useful properties of functions in the degenerate Sobolev space QH01(v;Ω)QH^{1}_{0}(v;\Omega). All of these properties are well known in the classical case: see, for instance, [12]. In the degenerate case, we stress that the first assumption in Hypothesis 1.6 is critical in proving these results and throughout this subsection we assume that vL1(Ω)v\in L^{1}(\Omega) and |Q|opkv|Q|_{op}\leq kv a.e.

Our first result shows that weak gradients are zero almost everywhere on sets of vv-measure zero.

Lemma 2.13.

Let 𝐮=(u,u)QH01(v;Ω){\bf u}=(u,\nabla u)\in QH^{1}_{0}(v;\Omega) and wLip0(Ω)w\in Lip_{0}(\Omega). Then, given any set EE of vv-measure zero, we have that:

  1. (1)

    wLQ2(E)=0\|\nabla w\|_{L^{2}_{Q}(E)}=0;

  2. (2)

    uLQ2(E)=0\|\nabla u\|_{L^{2}_{Q}(E)}=0;

  3. (3)

    Q(x)u(x)=0=Q(x)w(x)\sqrt{Q(x)}\nabla u(x)=0=\sqrt{Q(x)}\nabla w(x) a.e. xEx\in E.

Proof.

If wLip0(Ω)w\in Lip_{0}(\Omega), then w\nabla w is defined a.e. in Ω\Omega by the Rademacher-Stepanov theorem and is in LL^{\infty}. Therefore, for a.e. xΩx\in\Omega,

|Q(x)w(x)||Q(x)|op12|w(x)|cv(x)12|w(x)|;|\sqrt{Q(x)}\nabla w(x)|\leq|Q(x)|_{op}^{\frac{1}{2}}|\nabla w(x)|\leq cv(x)^{\frac{1}{2}}|\nabla w(x)|;

hence,

w(x)LQ2(E)2w2v(E)=0,\|\nabla w(x)\|_{L^{2}_{Q}(E)}^{2}\leq\|\nabla w\|_{\infty}^{2}v(E)=0,

which proves (1).

Let 𝐮QH01(v;Ω){\bf u}\in QH^{1}_{0}(v;\Omega); then there exists a sequence {wj}jLip0(Ω)\{w_{j}\}_{j}\subset Lip_{0}(\Omega) such that wju\nabla w_{j}\rightarrow\nabla u in LQ2(Ω)L^{2}_{Q}(\Omega). Then by the previous argument,

uLQ2(E)=limjwjLQ2(E)=0,\|\nabla u\|_{L^{2}_{Q}(E)}=\lim_{j\rightarrow\infty}\|\nabla w_{j}\|_{L^{2}_{Q}(E)}=0,

and so (2) holds.

Finally, (3) follows immediately from (1) and (2). ∎

Our second result shows that non-negative truncations of functions in QH01(v;Ω)QH^{1}_{0}(v;\Omega) are again in this space.

Lemma 2.14.

Let 𝐮=(u,u)QH01(v;Ω){\bf u}=(u,\nabla u)\in QH^{1}_{0}(v;\Omega) and fix r>0r>0. If S(r)={xΩ:u(x)>r}S(r)=\{x\in\Omega:u(x)>r\}, then ((ur)+,𝟙S(r)u)QH01(v;Ω)((u-r)_{+},\mathbbm{1}_{S(r)}\nabla u)\in QH^{1}_{0}(v;\Omega).

Proof.

By the definition of QH01(v;Ω)QH^{1}_{0}(v;\Omega) there exists a sequence {uj}j\{u_{j}\}_{j} in Lip0(Ω)Lip_{0}(\Omega) such that ujuu_{j}\rightarrow u in L2(v;Ω)L^{2}(v;\Omega) and uju\nabla u_{j}\rightarrow\nabla u in LQ2(Ω)L^{2}_{Q}(\Omega). If we pass to a subsequence, we may assume that ujuu_{j}\rightarrow u pointwise vv-a.e. We will first prove that (ujr)+(ur)+(u_{j}-r)_{+}\rightarrow(u-r)_{+} in L2(v;Ω)L^{2}(v;\Omega).

Define fj=|(ujr)+(ur)+|2f_{j}=|(u_{j}-r)_{+}-(u-r)_{+}|^{2}; then fj0f_{j}\rightarrow 0 vv-a.e. We will show that fjf_{j} converges to 0 in L1(v;Ω)L^{1}(v;\Omega); this follows from the generalized dominated convergence theorem [11, p. 59] if we show that there exist non-negative functions gj,gL1(v;Ω)g_{j},\,g\in L^{1}(v;\Omega) such that fjgjf_{j}\leq g_{j} and gjL1(v;Ω)gL1(v;Ω)\|g_{j}\|_{L^{1}(v;\Omega)}\rightarrow\|g\|_{L^{1}(v;\Omega)} as jj\rightarrow\infty. But we have that

fj2(ujr)+2+2(ur)+24(|uj|2+r2)+4(|u|2+r2)=gj.f_{j}\leq 2(u_{j}-r)^{2}_{+}+2(u-r)_{+}^{2}\leq 4(|u_{j}|^{2}+r^{2})+4(|u|^{2}+r^{2})=g_{j}.

Moreover, gjg_{j} converges pointwise a.e. to g=8(|u|2+r2)g=8(|u|^{2}+r^{2}), and g,gjL1(v;Ω)g,\,g_{j}\in L^{1}(v;\Omega) since v(Ω)<v(\Omega)<\infty. Finally, since ujuu_{j}\rightarrow u in L2(v;Ω)L^{2}(v;\Omega) we get that gjL1(v;Ω)gL1(v;Ω)\|g_{j}\|_{L^{1}(v;\Omega)}\rightarrow\|g\|_{L^{1}(v;\Omega)}.

Now define Sj={xΩ:uj(x)>r}S_{j}=\{x\in\Omega:u_{j}(x)>r\}; then 𝟙Sj𝟙S\mathbbm{1}_{S_{j}}\rightarrow\mathbbm{1}_{S} vv-a.e. Moreover, we have that (ujr)+=uj𝟙Sj\nabla(u_{j}-r)_{+}=\nabla u_{j}\mathbbm{1}_{S_{j}} a.e. [12, Lemma 7.6] and so vv-a.e. By passing to another subsequence, we assume that uj𝟙Sju𝟙S\nabla u_{j}\mathbbm{1}_{S_{j}}\rightarrow\nabla u\mathbbm{1}_{S} pointwise vv-a.e. We claim that they converge in LQ2(Ω)L^{2}_{Q}(\Omega) as well. If this is the case, then we have shown that ((ujr)+,(ujr)+)((u_{j}-r)_{+},\nabla(u_{j}-r)_{+}) is Cauchy in QH01(v;Ω)QH^{1}_{0}(v;\Omega), and the desired conclusion follows at once.

To prove LQ2(Ω)L^{2}_{Q}(\Omega) convergence, note that

uj𝟙Sju𝟙SLQ2(Ω)uj𝟙Sju𝟙SjLQ2(Ω)+Qu(𝟙Sj𝟙S)L2(Ω).\|\nabla u_{j}\mathbbm{1}_{S_{j}}-\nabla u\mathbbm{1}_{S}\|_{L^{2}_{Q}(\Omega)}\leq\|\nabla u_{j}\mathbbm{1}_{S_{j}}-\nabla u\mathbbm{1}_{S_{j}}\|_{L^{2}_{Q}(\Omega)}+\|\sqrt{Q}\nabla u(\mathbbm{1}_{S_{j}}-\mathbbm{1}_{S})\|_{L^{2}(\Omega)}.

The first term on the right-hand side is less than ujuLQ2(Ω)\|\nabla u_{j}-\nabla u\|_{L^{2}_{Q}(\Omega)} which goes to 0 as jj\rightarrow\infty. To estimate the second term, let EE be the set of xΩx\in\Omega where 𝟙Sj(x)\mathbbm{1}_{S_{j}}(x) does not converge to 𝟙S(x)\mathbbm{1}_{S}(x). Then v(E)=0v(E)=0, and so by Lemma 2.13,

E|Qu(𝟙Sj𝟙S)|2𝑑xE|Qu|2𝑑x=0.\int_{E}|\sqrt{Q}\nabla u(\mathbbm{1}_{S_{j}}-\mathbbm{1}_{S})|^{2}\,dx\leq\int_{E}|\sqrt{Q}\nabla u|^{2}\,dx=0.

Since |Qu(𝟙Sj𝟙S)||Qu|L2(Ω)|\sqrt{Q}\nabla u(\mathbbm{1}_{S_{j}}-\mathbbm{1}_{S})|\leq|\sqrt{Q}\nabla u|\in L^{2}(\Omega), by the dominated convergence theorem we have that as j0j\rightarrow 0,

Qu(𝟙Sj𝟙S)L2(Ω)=Qu(𝟙Sj𝟙S)L2(ΩE)0.\|\sqrt{Q}\nabla u(\mathbbm{1}_{S_{j}}-\mathbbm{1}_{S})\|_{L^{2}(\Omega)}=\|\sqrt{Q}\nabla u(\mathbbm{1}_{S_{j}}-\mathbbm{1}_{S})\|_{L^{2}(\Omega\setminus E)}\rightarrow 0.

Our next lemma proves the existence of an approximating sequence of Lipschitz functions with some additional useful properties.

Lemma 2.15.

Let 𝐮=(u,u)QH01(v;Ω){\bf u}=(u,\nabla u)\in QH^{1}_{0}(v;\Omega) with uL(v;Ω)u\in L^{\infty}(v;\Omega) and u0u\geq 0 vv-a.e. Then there exists a sequence {uj}jLip0(Ω)\{u_{j}\}_{j}\in Lip_{0}(\Omega) such that:

  1. (1)

    0uj(x)uL(v;Ω)+10\leq u_{j}(x)\leq\|u\|_{L^{\infty}(v;\Omega)}+1 in Ω\Omega;

  2. (2)

    ujuu_{j}\rightarrow u vv-a.e. and also in L2(v;Ω)L^{2}(v;\Omega);

  3. (3)

    uju\nabla u_{j}\rightarrow\nabla u in LQ2(Ω)L^{2}_{Q}(\Omega) and |Quj||Qu|\left|\sqrt{Q}\nabla u_{j}\right|\rightarrow\left|\sqrt{Q}\nabla u\right| pointwise a.e.;

  4. (4)

    ujLQ2(Ω)uLQ2(Ω)+1\|\nabla u_{j}\|_{L^{2}_{Q}(\Omega)}\leq\|\nabla u\|_{L^{2}_{Q}(\Omega)}+1 for each jj\in\mathbb{N}.

Proof.

By the definition of QH01(v;Ω)QH^{1}_{0}(v;\Omega) and by passing twice to a subsequence, there exists a sequence {zj}jLip0(Ω)\{z_{j}\}_{j}\subset Lip_{0}(\Omega) such that:

  1. (11^{\prime})

    zjuz_{j}\rightarrow u both vv-a.e. and also in L2(v;Ω)L^{2}(v;\Omega)

  2. (22^{\prime})

    zju\nabla z_{j}\rightarrow\nabla u in LQ2(Ω)L^{2}_{Q}(\Omega) and |Qzj||Qu||\sqrt{Q}\nabla z_{j}|\rightarrow|\sqrt{Q}\nabla u| a.e.;

  3. (33^{\prime})

    zjLQ2(Ω)uLQ2(Ω)+1.\|\nabla z_{j}\|_{L^{2}_{Q}(\Omega)}\leq\|\nabla u\|_{L^{2}_{Q}(\Omega)}+1.

Now let wj=|zj|w_{j}=|z_{j}|. Since uu is non-negative vv-a.e. in Ω\Omega, by the triangle inequality we have that

|wju|=||zj||u|||zju||w_{j}-u|=||z_{j}|-|u||\leq|z_{j}-u|

vv-a.e. Therefore, we have that wjw_{j} converges to uu both in L2(v;Ω)L^{2}(v;\Omega) and pointwise vv-a.e.

By the Rademacher-Stepanov theorem [9], wj(x)=sgn(zj(x))zj(x)\nabla w_{j}(x)=sgn(z_{j}(x))\nabla z_{j}(x) a.e. Hence, |Qwj(x)|=|Qzj(x)||\sqrt{Q}\nabla w_{j}(x)|=|\sqrt{Q}\nabla z_{j}(x)| a.e. and so wjLQ2(Ω)uLQ2(Ω)\|\nabla w_{j}\|_{L^{2}_{Q}(\Omega)}\rightarrow\|\nabla u\|_{L^{2}_{Q}(\Omega)} as jj\rightarrow\infty. Thus wj0w_{j}\geq 0 a.e. and properties (11^{\prime})–(33^{\prime}) above hold with zjz_{j} replaced by wjw_{j}.

We now define the sequence of Lip0(Ω)Lip_{0}(\Omega) functions {uj}j\{u_{j}\}_{j}. Set M=uL(v;Ω)+1M=\|u\|_{L^{\infty}(v;\Omega)}+1 and let ϕ:[0,)[0,)\phi:[0,\infty)\rightarrow[0,\infty) be such that ϕC\phi\in C^{\infty}, ϕ\phi is increasing, ϕ(x)=x\phi(x)=x if 0xM120\leq x\leq M-\frac{1}{2}, ϕ(x)=M\phi(x)=M if xM+1x\geq M+1, and ϕ(x)1\phi^{\prime}(x)\leq 1. Define the uju_{j} by uj(x)=ϕ(wj(x))u_{j}(x)=\phi(w_{j}(x)). Then ϕjLip0(Ω)\phi_{j}\in Lip_{0}(\Omega); moreover, uj(x)=ϕ(wj(x))wj(x)\nabla u_{j}(x)=\phi^{\prime}(w_{j}(x))\nabla w_{j}(x) a.e. and so

(2.16) |Q(x)uj(x)||Q(x)wj(x)|.|\sqrt{Q(x)}\nabla u_{j}(x)|\leq|\sqrt{Q(x)}\nabla w_{j}(x)|.

We claim that {uj}j\{u_{j}\}_{j} satisfies properties (1)–(4) above. By the definition of ϕ\phi, 0ujM0\leq u_{j}\leq M, so property (1) holds. Property (4) follow immediately from (2.16) and property (33^{\prime}) for the wjw_{j}.

It remains to prove properties (2) and (3). By the choice of MM, u(s)M1u(s)\leq M-1 for vv-a.e. sΩs\in\Omega. We also have that wj(s)u(s)w_{j}(s)\rightarrow u(s) vv-a.e. Let FF be the set of all sΩs\in\Omega such that both of these hold. Then v(ΩF)=0v(\Omega\setminus F)=0. Given sFs\in F there exists N>0N>0 such that if jNj\geq N, wj(s)<M12w_{j}(s)<M-\frac{1}{2}, and so uj(s)=wj(s)u_{j}(s)=w_{j}(s). Thus, ujuu_{j}\rightarrow u pointwise vv-a.e. Since uu is bounded and v(Ω)<v(\Omega)<\infty, by the dominated convergence theorem we also have that ujuu_{j}\rightarrow u in L2(v;Ω)L^{2}(v;\Omega). This proves (2).

To prove (3) define the set FF as above. For each sFs\in F, there exists N>0N>0 such that for each jNj\geq N there exists a ball Bj,sB_{j,s} where for xBj,sx\in B_{j,s}, wj(x)<M12w_{j}(x)<M-\frac{1}{2}; hence, uj(s)=wj(s)\nabla u_{j}(s)=\nabla w_{j}(s) for jNj\geq N. Now let GG be the set of sΩs\in\Omega such that |Q(s)wj(s)||Q(s)u(s)||\sqrt{Q(s)}\nabla w_{j}(s)|\rightarrow|\sqrt{Q(s)}\nabla u(s)|; by (22^{\prime}), |ΩG|=0|\Omega\setminus G|=0. Since vdxv\,dx is an absolutely continuous measure, v(ΩG)=0v(\Omega\setminus G)=0. Let H=FGH=F\cap G. Then on HH we have that |Quj||Qu||\sqrt{Q}\nabla u_{j}|\rightarrow|\sqrt{Q}\nabla u| pointwise. But v(ΩH)=0v(\Omega\setminus H)=0 so by Lemma 2.13 we have that

ujLQ2(ΩH)=0=uLQ2(ΩH).\|\nabla u_{j}\|_{L^{2}_{Q}(\Omega\setminus H)}=0=\|\nabla u\|_{L^{2}_{Q}(\Omega\setminus H)}.

This implies that |Quj|=0=|Qu||\sqrt{Q}\nabla u_{j}|=0=|\sqrt{Q}\nabla u| almost everywhere on ΩH\Omega\setminus H. Therefore, we have that |Quj||Qu||\sqrt{Q}\nabla u_{j}|\rightarrow|\sqrt{Q}\nabla u| pointwise a.e.

Finally, to prove that uju\nabla u_{j}\rightarrow\nabla u in LQ2(Ω)L^{2}_{Q}(\Omega) we use the generalized dominated convergence theorem as in the proof of Lemma 2.14. Let fj=|Q(uju)|2f_{j}=|\sqrt{Q}(\nabla u_{j}-\nabla u)|^{2}; then fj0f_{j}\rightarrow 0 a.e. Further, by (2.16)

fj2|Quj|2+2|Qu|22|Qwj|2+2|Qu|2=gj.f_{j}\leq 2|\sqrt{Q}\nabla u_{j}|^{2}+2|\sqrt{Q}\nabla u|^{2}\leq 2|\sqrt{Q}\nabla w_{j}|^{2}+2|\sqrt{Q}\nabla u|^{2}=g_{j}.

Again by (22^{\prime}), gj4|Qu|2=gg_{j}\rightarrow 4|\sqrt{Q}\nabla u|^{2}=g a.e., and since wju\nabla w_{j}\rightarrow\nabla u in LQ2(Ω)L^{2}_{Q}(\Omega), gjgg_{j}\rightarrow g in L1(Ω)L^{1}(\Omega). Therefore, fj0f_{j}\rightarrow 0 in L1(Ω)L^{1}(\Omega), which completes the proof of (3). ∎

The next two lemmas give the product rule and chain rule associated to pairs in QH01(Ω)QH^{1}_{0}(\Omega). The proofs are adapted from those of similar results in [19].

Lemma 2.17.

Let (u,u)QH01(v;Ω)(u,\nabla u)\in QH_{0}^{1}(v;\Omega) and let ψLip0(Ω)\psi\in Lip_{0}(\Omega). Then we have that (uψ,ψu+uψ)QH01(v;Ω)(u\psi,\psi\nabla u+u\nabla\psi)\in QH^{1}_{0}(v;\Omega).

Proof.

By the definition of QH01(v;Ω)QH_{0}^{1}(v;\Omega) there exists a sequence {wj}Lip0(Ω)\{w_{j}\}\subset Lip_{0}(\Omega) such that wjuw_{j}\rightarrow u in L2(v;Ω)L^{2}(v;\Omega) and wju\nabla w_{j}\rightarrow\nabla u in LQ2(Ω)L^{2}_{Q}(\Omega). But then we immediately have that

wjψuψL2(v;Ω)ψwjuL2(v;Ω),\|w_{j}\psi-u\psi\|_{L^{2}(v;\Omega)}\leq\|\psi\|_{\infty}\|w_{j}-u\|_{L^{2}(v;\Omega)},

and so wjψuψw_{j}\psi\rightarrow u\psi in L2(v;Ω)L^{2}(v;\Omega).

Similarly, since |Q|opkv|Q|_{op}\leq kv a.e., we have that

(wjψ)(uψ+ψu)LQ2(Ω)ψwjψuLQ2(Ω)+wjψuψLQ2(Ω)ψwjuLQ2(Ω)+kψwjuL2(v;Ω).\|\nabla(w_{j}\psi)-(u\nabla\psi+\psi\nabla u)\|_{L^{2}_{Q}(\Omega)}\leq\|\psi\nabla w_{j}-\psi\nabla u\|_{L^{2}_{Q}(\Omega)}+\|w_{j}\nabla\psi-u\nabla\psi\|_{L^{2}_{Q}(\Omega)}\\ \leq\|\psi\|_{\infty}\|\nabla w_{j}-\nabla u\|_{L^{2}_{Q}(\Omega)}+k\|\nabla\psi\|_{\infty}\|w_{j}-u\|_{L^{2}(v;\Omega)}.

Thus, (wjψ)uψ+ψu\nabla(w_{j}\psi)\rightarrow u\nabla\psi+\psi\nabla u in LQ2(Ω)L^{2}_{Q}(\Omega) and so (uψ,uψ+ψu)QH01(v;Ω)(u\psi,u\nabla\psi+\psi\nabla u)\in QH^{1}_{0}(v;\Omega). ∎

Lemma 2.18.

Let (u,u)QH01(v;Ω)(u,\nabla u)\in QH^{1}_{0}(v;\Omega) with u0u\geq 0 vv-a.e. and uL(v;Ω)u\in L^{\infty}(v;\Omega). Then, given any non-negative function φC1()\varphi\in C^{1}(\mathbb{R}) such that φ(0)=0\varphi(0)=0, the pair (φ(u),φ(u)u)QH01(v;Ω)(\varphi(u),\varphi^{\prime}(u)\nabla u)\in QH^{1}_{0}(v;\Omega).

Proof.

Let {uj}jLip0(Ω)\{u_{j}\}_{j}\subset Lip_{0}(\Omega) be the sequence associated with (u,u)(u,\nabla u) given by Lemma 2.15. Since uju_{j} is Lipschitz with compact support in Ω\Omega and φ(0)=0\varphi(0)=0, ψj=φ(uj)Lip0(Ω)\psi_{j}=\varphi(u_{j})\in Lip_{0}(\Omega). Since ujuu_{j}\rightarrow u vv-a.e., the continuity of φ\varphi implies that ψjφ(u)=ψ\psi_{j}\rightarrow\varphi(u)=\psi vv-a.e. By the fundamental theorem of calculus,

|φ(t)|=|0tφ(s)𝑑s|φL([0,M])|t|=A0|t||\varphi(t)|=\left|\int_{0}^{t}\varphi^{\prime}(s)ds\right|\leq\|\varphi^{\prime}\|_{L^{\infty}([0,M])}|t|=A_{0}|t|

whenever 0tM=uL(v;Ω)+10\leq t\leq M=\|u\|_{L^{\infty}(v;\Omega)}+1.

Since by assumption and property (1) of Lemma 2.15, 0u(x),uj(x)M0\leq u(x),\,u_{j}(x)\leq M for vv-a.e. xΩx\in\Omega, we have that vv-a.e.,

|ψjψ|22(|ψj|2+|ψ|2)2A02(|uj|2+|u|2).|\psi_{j}-\psi|^{2}\leq 2(|\psi_{j}|^{2}+|\psi|^{2})\leq 2A_{0}^{2}(|u_{j}|^{2}+|u|^{2}).

Since |uj|2+|u|22|u|2|u_{j}|^{2}+|u|^{2}\rightarrow 2|u|^{2} vv-a.e. and in L1(v;Ω)L^{1}(v;\Omega), by the generalized Lebesgue dominated convergence theorem we get that ψjψ\psi_{j}\rightarrow\psi in L2(v;Ω)L^{2}(v;\Omega).

To show the convergence of the gradients, first note that Qψj=φ(uj)Quj\sqrt{Q}\nabla\psi_{j}=\varphi^{\prime}(u_{j})\sqrt{Q}\nabla u_{j} a.e. in Ω\Omega and so by the continuity of φ\varphi^{\prime} and property (3) in Lemma 2.15 we get that Qψjφ(u)Qu\sqrt{Q}\nabla\psi_{j}\rightarrow\varphi^{\prime}(u)\sqrt{Q}\nabla u a.e. Moreover,

|Q(ψj)φ(u)Qu|22|φ(uj)Quj|2+2|φ(u)Qu|22A02(|Quj|2+|Qu|2).|\sqrt{Q}\nabla(\psi_{j})-\varphi^{\prime}(u)\sqrt{Q}\nabla u|^{2}\leq 2|\varphi^{\prime}(u_{j})\sqrt{Q}\nabla u_{j}|^{2}+2|\varphi^{\prime}(u)\sqrt{Q}\nabla u|^{2}\\ \leq 2A_{0}^{2}(|\sqrt{Q}\nabla u_{j}|^{2}+|\sqrt{Q}\nabla u|^{2}).

The right-hand term converges to 4A02|Qu|24A_{0}^{2}|\sqrt{Q}\nabla u|^{2} both pointwise a.e. and in L1(Ω)L^{1}(\Omega). Therefore, we can again apply the generalized dominated convergence theorem to get that ψjφ(u)u\nabla\psi_{j}\rightarrow\varphi^{\prime}(u)\nabla u in LQ2(Ω)L^{2}_{Q}(\Omega). We conclude that (φ(u),φ(u)u)QH01(v;Ω)(\varphi(u),\varphi^{\prime}(u)\nabla u)\in QH^{1}_{0}(v;\Omega). ∎

Exponential results

In this section we give two results which are needed to prove Theorem 1.13. The first gives a solution to an auxiliary Dirichlet problem and is an application of the previous two lemmas.

Lemma 2.19.

Fix α>0\alpha>0. If (u,u)QH01(Ω)(u,\nabla u)\in QH^{1}_{0}(\Omega) is a non-negative bounded weak subsolution of the Dirichlet problem

(2.22) {Div(Qu)=fvfor xΩ,u=0for xΩ,\displaystyle\left\{\begin{array}[]{rcll}-\operatorname{Div}\left(Q\nabla u\right)&=&fv&\textrm{for }x\in\Omega,\\ u&=&0&\textrm{for }x\in\partial\Omega,\end{array}\right.

then (w,w)=(eαu1,αeαuu)QH01(v;Ω)(w,\nabla w)=(e^{\alpha u}-1,\alpha e^{\alpha u}\nabla u)\in QH_{0}^{1}(v;\Omega) is a non-negative weak subsolution of the Dirichlet problem

(2.25) {Div(Qw)=αf(w+1)vfor xΩ,w=0for xΩ.\displaystyle\left\{\begin{array}[]{rcll}-\operatorname{Div}\left(Q\nabla w\right)&=&\alpha f(w+1)v&\textrm{for }x\in\Omega,\\ w&=&0&\textrm{for }x\in\partial\Omega.\end{array}\right.
Proof.

Fix a non-negative ψLip0(Ω)\psi\in Lip_{0}(\Omega). By our assumptions on (u,u)(u,\nabla u) and by Lemmas 2.17 and 2.18 we have that that both (w,w)=(eαu1,αeαuu)(w,\nabla w)=(e^{\alpha u}-1,\alpha e^{\alpha u}\nabla u) and (ψ(w+1),(w+1)ψ+ψw)(\psi(w+1),(w+1)\nabla\psi+\psi\nabla w) are in QH01(Ω)QH^{1}_{0}(\Omega). Since w=α(w+1)u\nabla w=\alpha(w+1)\nabla u and (u,u)(u,\nabla u) is a non-negative weak subsolution of (2.22), we have that

Ωf(w+1)ψv𝑑x\displaystyle\int_{\Omega}f(w+1)\psi~{}vdx Ω(ψ(w+1))Qudx\displaystyle\geq\int_{\Omega}\nabla(\psi(w+1))Q\nabla u~{}dx
=Ω(w+1)ψQudx+Ωψ(w+1)Qwdx\displaystyle=\int_{\Omega}(w+1)\nabla\psi Q\nabla u~{}dx+\int_{\Omega}\psi\nabla(w+1)Q\nabla w~{}dx
=1αΩψQwdx+ΩψwQwdx\displaystyle=\frac{1}{\alpha}\int_{\Omega}\nabla\psi Q\nabla w~{}dx+\int_{\Omega}\psi\nabla wQ\nabla w~{}dx
1αΩψQwdx.\displaystyle\geq\frac{1}{\alpha}\int_{\Omega}\nabla\psi Q\nabla w~{}dx.

Since ψLip0(Ω)\psi\in Lip_{0}(\Omega) is arbitrary, we conclude that ww is a non-negative weak subsolution of (4.9). ∎

Our second result gives the exponential integrability of bounded solutions to (1.3). A version of this result is proved in [26, Lemma B] for uniformly elliptic operators; a qualitative version appeared previously in [4, Example 4]. Here we adapt the proof from [26] to our more general setting.

Lemma 2.26.

Suppose Hypothesis 1.6 holds. Let fLσ(v;Ω)f\in L^{\sigma^{\prime}}(v;\Omega) satisfy fσ;v1\|f\|_{\sigma^{\prime};v}\leq 1, and let (u,u)QH01(Ω)(u,\nabla u)\in QH_{0}^{1}(\Omega) be a bounded, non-negative weak subsolution of  (1.3). Then, for every γ(0,4C02)\gamma\in(0,\frac{4}{C_{0}^{2}}), with C0C_{0} as in (1.7), there M=M(γ,C0,v(Ω))M=M(\gamma,C_{0},v(\Omega)) such that

(2.27) Ωeγu(x)v(x)𝑑xM.\int_{\Omega}e^{\gamma u(x)}v(x)\,dx\leq M.
Proof.

Let ff and (u,u)(u,\nabla u) be as in the hypotheses. Define φ=eγu1\varphi=e^{\gamma u}-1 and ψ=eγu21\psi=e^{\frac{\gamma u}{2}}-1 with γ>0\gamma>0 to be chosen below. Since uu is bounded, by Lemma 2.18 we have that

(φ,φ)=(eγu1,γeγuu),(ψ,ψ)=(eγu21,γ2eγu2u)(\varphi,\nabla\varphi)=(e^{\gamma u}-1,\gamma e^{\gamma u}\nabla u),\quad(\psi,\nabla\psi)=(e^{\frac{\gamma u}{2}}-1,\frac{\gamma}{2}e^{\frac{\gamma u}{2}}\nabla u)

are in QH01(Ω)QH_{0}^{1}(\Omega). Further, we immediately have the identities φ=ψ2+2ψ\varphi=\psi^{2}+2\psi, ψ=γ2eγu2u\nabla\psi=\frac{\gamma}{2}e^{\frac{\gamma u}{2}}\nabla u, and φ=2eγu2ψ\nabla\varphi=2e^{\frac{\gamma u}{2}}\nabla\psi. If we apply the Sobolev inequality (2.10) and use φ\varphi as a test function in Definition 2.11 we can estimate as follows:

ψL2σ(v;Ω)2\displaystyle\|\psi\|_{L^{2\sigma}(v;\Omega)}^{2} C02Ω|Q(x)ψ(x)|2𝑑x\displaystyle\leq C_{0}^{2}\int_{\Omega}\left|\sqrt{Q(x)}\nabla\psi(x)\right|^{2}\,dx
=C02γ4Ωφ(x)Q(x)u(x)𝑑x\displaystyle=\frac{C_{0}^{2}\gamma}{4}\int_{\Omega}\nabla\varphi(x)\cdot Q(x)\nabla u(x)\,dx
C02γ4Ωf(x)φ(x)v(x)𝑑x\displaystyle\leq\frac{C_{0}^{2}\gamma}{4}\int_{\Omega}f(x)\varphi(x)v(x)\,dx
=C02γ4(Ωf(x)ψ(x)2v(x)𝑑x+2Ωf(x)ψ(x)v(x)𝑑x).\displaystyle=\frac{C_{0}^{2}\gamma}{4}\left(\int_{\Omega}f(x)\psi(x)^{2}v(x)\,dx+2\int_{\Omega}f(x)\psi(x)v(x)\,dx\right).
If we now apply Hölder’s inequality with exponents σ\sigma and 2σ2\sigma, and then 22, we get
=C02γ4(fLσ(v;Ω)ψ2Lσ(v;Ω)+2fLσ(v;Ω)ψLσ(v;Ω))\displaystyle=\frac{C_{0}^{2}\gamma}{4}\left(\|f\|_{L^{\sigma^{\prime}}(v;\Omega)}\|\psi^{2}\|_{L^{\sigma}(v;\Omega)}+2\|f\|_{L^{\sigma^{\prime}}(v;\Omega)}\|\psi\|_{L^{\sigma}(v;\Omega)}\right)
C02γ4(ψL2σ(v;Ω)2+2ψL2σ(v;Ω)v(Ω)12σ).\displaystyle\leq\frac{C_{0}^{2}\gamma}{4}\left(\|\psi\|_{L^{2\sigma}(v;\Omega)}^{2}+2\|\psi\|_{L^{2\sigma}(v;\Omega)}v(\Omega)^{\frac{1}{2\sigma}}\right).

If we now fix γ(0,4C02)\gamma\in(0,\frac{4}{C_{0}^{2}}), then we can re-arrange terms to get

(2.28) ψL2σ(v;Ω)C02γ2(1C02γ4)v(Ω)12σ.\|\psi\|_{L^{2\sigma}(v;\Omega)}\leq\frac{C_{0}^{2}\gamma}{2(1-\frac{C_{0}^{2}\gamma}{4})}v(\Omega)^{\frac{1}{2\sigma}}.

Therefore, again by Hölder’s inequality and by (2.28) applied twice, we have that

Ωeγu(x)v(x)𝑑x\displaystyle\int_{\Omega}e^{\gamma u(x)}v(x)\,dx =Ω(ψ(x)2+2ψ(x))v(x)𝑑x+v(Ω)\displaystyle=\int_{\Omega}\big{(}\psi(x)^{2}+2\psi(x)\big{)}v(x)\,dx+v(\Omega)
ψL2σ(v;Ω)2v(Ω)1σ+2ψL2σ(v;Ω)v(Ω)1(2σ)+v(Ω)\displaystyle\leq\|\psi\|_{L^{2\sigma}(v;\Omega)}^{2}v(\Omega)^{\frac{1}{\sigma^{\prime}}}+2\|\psi\|_{L^{2\sigma}(v;\Omega)}v(\Omega)^{\frac{1}{(2\sigma)^{\prime}}}+v(\Omega)
C(γ,C0)v(Ω)\displaystyle\leq C(\gamma,C_{0})v(\Omega)
=M(γ,C0,v(Ω)).\displaystyle=M(\gamma,C_{0},v(\Omega)).

3. Proof of Theorem 1.9

Fix fLA(v;Ω)f\in L^{A}(v;\Omega) and let 𝐮=(u,u)QH01(Ω){\bf u}=(u,\nabla u)\in QH^{1}_{0}(\Omega) be a non-negative weak subsolution of (1.3). We may assume without loss of generality that fLA(v;Ω)>0\|f\|_{L^{A}(v;\Omega)}>0 (equivalently, that ff is non-zero on a set EΩE\subset\Omega with v(E)>0v(E)>0); otherwise, a standard argument shows that u=0u=0 vv-almost everywhere. (Cf. (3.8) below.) By Lemma 2.6, fLσ(v;Ω)f\in L^{\sigma^{\prime}}(v;\Omega).

For each r>0r>0 define φr=(ur)+\varphi_{r}=(u-r)_{+} and let S(r)={xΩ:u(x)>r}S(r)=\{x\in\Omega~{}:~{}u(x)>r\}. Then by Lemma 2.14, (φr,φr)=((ur)+,𝟙S(r)u)QH01(v;Ω)(\varphi_{r},\nabla\varphi_{r})=((u-r)_{+},\mathbbm{1}_{S(r)}\nabla u)\in QH^{1}_{0}(v;\Omega). We now estimate as follows: by the Sobolev inequality (2.10), the definition of a weak subsolution with φr\varphi_{r} as the test function, and Hölder’s inequality, we have that

φrL2σ(v;Ω)2C02S(r)|Qφr|2𝑑x=C02S(r)φrQφrdx=C02S(r)φrQudxC02S(r)fφrv𝑑xC02fL(2σ)(v;S(r))φrL2σ(v;Ω)\|\varphi_{r}\|_{L^{2\sigma}(v;\Omega)}^{2}\leq C_{0}^{2}\int_{S(r)}|\sqrt{Q}\nabla\varphi_{r}|^{2}\,dx=C_{0}^{2}\int_{S(r)}\nabla\varphi_{r}\cdot Q\nabla\varphi_{r}\,dx\\ =C_{0}^{2}\int_{S(r)}\nabla\varphi_{r}\cdot Q\nabla u\,dx\leq C_{0}^{2}\int_{S(r)}f\varphi_{r}~{}vdx\leq C_{0}^{2}\|f\|_{L^{(2\sigma)^{\prime}}(v;S(r))}\|\varphi_{r}\|_{L^{2\sigma}(v;\Omega)}

since u=φr\nabla u=\nabla\varphi_{r} on S(r)S(r). If we divide through by φrL2σ(v;Ω)\|\varphi_{r}\|_{L^{2\sigma}(v;\Omega)}, we get

(3.1) φrL2σ(v;Ω)CfL(2σ)(v;S(r)).\|\varphi_{r}\|_{L^{2\sigma}(v;\Omega)}\leq C\|f\|_{L^{(2\sigma)^{\prime}}(v;S(r))}.

In order to estimate the norm of the right-hand side, recall that since σ>1\sigma>1, (2σ)<σ(2\sigma)^{\prime}<\sigma^{\prime}, we can define the Young function

B(t)=tσ(2σ)log(e+t)q.B(t)=t^{\frac{\sigma^{\prime}}{(2\sigma)^{\prime}}}\log(e+t)^{q}.

It is immediate that Bσ(t)=B(t(2σ))A(t)B_{\sigma}(t)=B(t^{(2\sigma)^{\prime}})\preceq A(t) and so by Lemma 2.2, a change of variables in the Luxemburg norm, and Lemmas 2.1 and 2.7 we get

fL(2σ)(v;S(r))(2σ)\displaystyle\|f\|_{L^{(2\sigma)^{\prime}}(v;S(r))}^{(2\sigma)^{\prime}} =Ω|f|(2σ)𝟙S(r)v𝑑x\displaystyle=\int_{\Omega}|f|^{(2\sigma)^{\prime}}~{}\mathbbm{1}_{S(r)}v\,dx
2f(2σ)LB(v;Ω)𝟙S(r)LB¯(v;Ω)\displaystyle\leq 2\|f^{(2\sigma)^{\prime}}\|_{L^{B}(v;\Omega)}\|\mathbbm{1}_{S(r)}\|_{L^{\bar{B}}(v;\Omega)}
=2fLBσ(v;Ω)(2σ)𝟙S(r)LB¯(v;Ω)\displaystyle=2\|f\|_{L^{B_{\sigma}}(v;\Omega)}^{(2\sigma)^{\prime}}\|\mathbbm{1}_{S(r)}\|_{L^{\bar{B}}(v;\Omega)}
CfLA(v;Ω)(2σ)v(S(r))12σ1log(e+(v(S(r)))1)q((2σ)σ),\displaystyle\leq C\|f\|_{L^{A}(v;\Omega)}^{(2\sigma)^{\prime}}\frac{v(S(r))^{\frac{1}{2\sigma-1}}}{\log(e+(v(S(r)))^{-1})^{q\left(\frac{(2\sigma)^{\prime}}{\sigma^{\prime}}\right)}},

where C=C(σ,q,v(Ω))C=C(\sigma,q,v(\Omega)) is independent of f,φ,f,\varphi, and 𝐮{\bf u}.

We now turn to our iteration argument. For all s>rs>r, S(s)S(r)S(s)\subset S(r) and, for xS(s)x\in S(s), φr(x)>sr>0\varphi_{r}(x)>s-r>0. Hence, if we combine the above two inequalities, we get

(3.2) v(S(s))12σ(sr)φr𝟙S(s)L2σ(v;Ω)CfLA(v;Ω)v(S(r))12σlog(e+(v(S(r)))1)qσ.v(S(s))^{\frac{1}{2\sigma}}(s-r)\leq\|\varphi_{r}\mathbbm{1}_{S(s)}\|{L^{2\sigma}(v;\Omega)}\leq C\|f\|_{L^{A}(v;\Omega)}~{}\displaystyle\frac{v(S(r))^{\frac{1}{2\sigma}}}{\log(e+(v(S(r)))^{-1})^{\frac{q}{\sigma^{\prime}}}}.

Define r0=τ0fLA(v;Ω)r_{0}=\tau_{0}\|f\|_{L^{A}(v;\Omega)} with τ0\tau_{0} to be chosen below. Our goal is to find τ0\tau_{0} sufficiently large so that v(S(r0))=0v(S(r_{0}))=0, as this immediately implies that

uL(v;Ω)τ0fLA(v;Ω),\|u\|_{L^{\infty}(v;\Omega)}\leq\tau_{0}\|f\|_{L^{A}(v;\Omega)},

which is what we want to prove. To do this, we will use an iteration argument based on De Giorgi iteration. For each kk\in\mathbb{N} set

(3.3) Ck=r0(1(k+1)ϵ)C_{k}=r_{0}(1-(k+1)^{-\epsilon})

where ϵ>0\epsilon>0 will be chosen below, and let C0=C1/2C_{0}=C_{1}/2. The sequence {Ck}k=0\{C_{k}\}_{k=0}^{\infty} increases to r0r_{0} and by an estimate using the mean-value theorem we have that for each kk\in\mathbb{N},

(3.4) Ck+1Ckϵr0(k+2)1+ϵ.C_{k+1}-C_{k}\geq\displaystyle\frac{\epsilon~{}r_{0}}{(k+2)^{1+\epsilon}}.

If we set s=Ck+1,r=Cks=C_{k+1},~{}r=C_{k}, μk=v(S(Ck))\mu_{k}=v(S(C_{k})) in inequality (3.2), we get

(3.5) μk+1[C(k+2)1+ϵϵτ0]2σμklog(e+μk1)2qσσ\mu_{k+1}\leq\left[\displaystyle\frac{C(k+2)^{1+\epsilon}}{\epsilon\tau_{0}}\right]^{2\sigma}\displaystyle\frac{\mu_{k}}{\log(e+\mu_{k}^{-1})^{\frac{2q\sigma}{\sigma^{\prime}}}}

for each kk\in\mathbb{N}. By the dominated convergence theorem μk\mu_{k} converges to v(S(r0))v(S(r_{0})), so to complete the proof we need to prove that μk0\mu_{k}\rightarrow 0.

Let mk=log(μk1)m_{k}=\log(\mu_{k}^{-1}). We will show that mkm_{k}\rightarrow\infty as kk\rightarrow\infty, which is equivalent to the desired limit. To do so, we will show that we can choose ϵ\epsilon and τ0\tau_{0} such that m02m_{0}\geq 2 and

(3.6) mkm0+km_{k}\geq m_{0}+k

for all k{0}k\in\mathbb{N}\cup\{0\}.

Fix ϵ=qσ1>0\epsilon=\frac{q}{\sigma^{\prime}}-1>0. Since 2σ(1+ϵ)=2σqσ2\sigma(1+\epsilon)=\frac{2\sigma q}{\sigma^{\prime}}, if we take logarithms and re-arrange terms, inequality (3.5) becomes, for kk\in{\mathbb{N}},

(3.7) mk+12σlog(ϵτ0C)+2σqσlog(mkk+2)+mk.m_{k+1}\geq 2\sigma\log\left(\displaystyle\frac{\epsilon\tau_{0}}{C}\right)+\frac{2\sigma q}{\sigma^{\prime}}\log\left(\displaystyle\frac{m_{k}}{k+2}\right)+m_{k}.

The first step is to fix m0m_{0} by an appropriate choice of τ0>0\tau_{0}>0. If we argue as we did to prove (3.1) using uu as the test function in the definition of a weak subsolution, we get

(3.8) uL2σ(v;Ω)CfL(2σ)(v;Ω).\|u\|_{L^{2\sigma}(v;\Omega)}\leq C\|f\|_{L^{(2\sigma)^{\prime}}(v;\Omega)}.

If we estimate the right-hand side using Hölder’s inequality and Lemma 2.6, we get

fL(2σ)(v;Ω)Cv(Ω)12σfLA(v;Ω),\|f\|_{L^{(2\sigma)^{\prime}}(v;\Omega)}\leq Cv(\Omega)^{\frac{1}{2\sigma}}\|f\|_{L^{A}(v;\Omega)},

where the constant CC is independent of ff and 𝐮{\bf u}. For each xS(C0)x\in S(C_{0}) we have that 2u(x)/C1>12u(x)/C_{1}>1, so by Hölder’s inequality and the above two estimates,

v(S(C0))2C1S(C0)uv𝑑x2C1uL2σ(v;Ω)v(S(C0))1(2σ)2CC1v(Ω)12σfLA(v;Ω)v(S(C0))1(2σ)2Cv(Ω)12στ0(12ϵ)v(S(C0))1(2σ).v(S(C_{0}))\leq\frac{2}{C_{1}}\int_{S(C_{0})}uv\,dx\leq\frac{2}{C_{1}}\|u\|_{L^{2\sigma}(v;\Omega)}v(S(C_{0}))^{\frac{1}{(2\sigma)^{\prime}}}\\ \leq\frac{2C}{C_{1}}v(\Omega)^{\frac{1}{2\sigma}}\|f\|_{L^{A}(v;\Omega)}v(S(C_{0}))^{\frac{1}{(2\sigma)^{\prime}}}\leq\frac{2Cv(\Omega)^{\frac{1}{2\sigma}}}{\tau_{0}(1-2^{-\epsilon})}v(S(C_{0}))^{\frac{1}{(2\sigma)^{\prime}}}.

If we re-arrange terms, we get

v(S(C0))(Cτ0(12ϵ))2σ,v(S(C_{0}))\leq\left(\frac{C}{\tau_{0}(1-2^{-\epsilon})}\right)^{2\sigma},

where again the constant CC is independent of ff and 𝐮\bf u. Now choose τ0>0\tau_{0}>0 so that

(3.9) μ0=v(S(C0))<e2, and τ0max{2ϵ+1eC2ϵ1,eCϵ},\mu_{0}=v(S(C_{0}))<e^{-2},\;\text{ and }\;\tau_{0}\geq\max\bigg{\{}\frac{2^{\epsilon+1}eC}{2^{\epsilon}-1},\frac{eC}{\epsilon}\bigg{\}},

where CC is as in (3.2). Note that τ0\tau_{0} is independent of 𝐮{\bf u} and ff, and the first inequality implies that m02m_{0}\geq 2.

It is clear that m0m0m_{0}\geq m_{0} but for the sake of clarity we also show that m1>m0+1m_{1}>m_{0}+1. Since k=0k=0 we cannot use (3.7), but instead use (3.2) directly. If we set s=C1s=C_{1} and r=C0r=C_{0} we find

C12μ112σCfLA(v;Ω)μ01(2σ)log(e+μ01)qσ.\frac{C_{1}}{2}\mu_{1}^{\frac{1}{2\sigma}}\leq C\|f\|_{L^{A}(v;\Omega)}\frac{\mu_{0}^{\frac{1}{\ell(2\sigma)^{\prime}}}}{\log(e+\mu_{0}^{-1})^{\frac{q}{\sigma^{\prime}}}}.

If we use the definition of C1C_{1} and recall that mj=log(μj1)m_{j}=\log(\mu_{j}^{-1}), we get

m12σlog((2ϵ1)τ02ϵ+1C)+m0+2q(σ1)log(m0)log((2ϵ1)τ02ϵ+1eC)+m0+1m0+1;m_{1}\geq 2\sigma\log\left(\frac{(2^{\epsilon}-1)\tau_{0}}{2^{\epsilon+1}C}\right)+m_{0}+2q(\sigma-1)\log(m_{0})\\ \geq\log\left(\frac{(2^{\epsilon}-1)\tau_{0}}{2^{\epsilon+1}eC}\right)+m_{0}+1\geq m_{0}+1;

the second inequality follows since m01m_{0}\geq 1, and the third by our choice of τ0\tau_{0}.

Now suppose that mjm0+jm_{j}\geq m_{0}+j for some jj\in\mathbb{N}. Since m02m_{0}\geq 2, (3.7) and (3.9) together show that

mj+12σlog(ϵτ0C)+2σqσlog(2+j2+j)+m0+jlog(ϵτ0eC)+m0+j+1m0+j+1.m_{j+1}\geq 2\sigma\log\left(\frac{\epsilon\tau_{0}}{C}\right)+\frac{2\sigma q}{\sigma^{\prime}}\log\left(\frac{2+j}{2+j}\right)+m_{0}+j\\ \geq\log\left(\frac{\epsilon\tau_{0}}{eC}\right)+m_{0}+j+1\geq m_{0}+j+1.

Hence, by induction we have that inequality (3.6) holds for all kk, and this completes our proof.

4. Proof of Theorem 1.13

Our proof requires one technical lemma.

Lemma 4.1.

Given σ>1\sigma>1, there exist constants b(σ,2σ)b\in(\sigma,2\sigma), b¯((2σ),σ)\bar{b}\in((2\sigma)^{\prime},\sigma^{\prime}), and p>1p>1 such that

(4.2) 1b+1b¯+1p=1,\frac{1}{b}+\frac{1}{\bar{b}}+\frac{1}{p}=1,

and

(4.3) Γ=2σb¯(σb¯σ+2σb2σ)=1.\Gamma=\frac{2\sigma}{\bar{b}}\left(\frac{\sigma^{\prime}-\bar{b}}{\sigma^{\prime}}+\frac{2\sigma-b}{2\sigma}\right)=1.
Proof.

We will first show that we can choose bb and b¯\bar{b} so that (4.3) holds, and then show that we can refine our choice so that (4.2) holds as well.

Set b=2σ(1β)b=2\sigma(1-\beta) and b¯=(1+β)(2σ)\bar{b}=(1+\beta)(2\sigma)^{\prime}, where 0<β<min(12,σ(2σ)(2σ))0<\beta<\min(\frac{1}{2},\frac{\sigma^{\prime}-(2\sigma)^{\prime}}{(2\sigma)^{\prime}}) will be determined below. With this restriction on β\beta it is immediate that bb and b¯\bar{b} lie in the specified intervals. Moreover, if we insert these values into the definition of Γ\Gamma, we get

Γ=2σ(1+β)(2σ)(σ(1+β)(2σ)σ+2σ2σ(1β)2σ)=2σ(1+β)(2σ)((1+β)(1(2σ)σ))=2σ(1(2σ)1σ)=1.\Gamma=\frac{2\sigma}{(1+\beta)(2\sigma)^{\prime}}\bigg{(}\frac{\sigma^{\prime}-(1+\beta)(2\sigma)^{\prime}}{\sigma^{\prime}}+\frac{2\sigma-2\sigma(1-\beta)}{2\sigma}\bigg{)}\\ =\frac{2\sigma}{(1+\beta)(2\sigma)^{\prime}}\bigg{(}(1+\beta)\bigg{(}1-\frac{(2\sigma)^{\prime}}{\sigma^{\prime}}\bigg{)}\bigg{)}=2\sigma\bigg{(}\frac{1}{(2\sigma)^{\prime}}-\frac{1}{\sigma^{\prime}}\bigg{)}=1.

This gives (4.3).

To show that we can choose p>1p>1 and β\beta so that (4.2) holds, note that

1b+1b¯=12σ(1β)+2σ12σ(1+β)=1+β+2σ12βσ+β2σ(1β2)=σβσ+βσ(1β2).\frac{1}{b}+\frac{1}{\bar{b}}=\frac{1}{2\sigma(1-\beta)}+\frac{2\sigma-1}{2\sigma(1+\beta)}=\frac{1+\beta+2\sigma-1-2\beta\sigma+\beta}{2\sigma(1-\beta^{2})}=\frac{\sigma-\beta\sigma+\beta}{\sigma(1-\beta^{2})}.

Thus, 1b+1b¯<1\frac{1}{b}+\frac{1}{\bar{b}}<1 exactly when 0<β<1σ0<\beta<\frac{1}{\sigma^{\prime}}. Hence, if we choose β\beta sufficiently small we can find p>1p>1 such that  (4.2) holds. ∎

Remark 4.4.

In the proof of Lemma 4.1, the range of possible values for β\beta shrinks as the dimension increases. In the classical case, σ=n2\sigma^{\prime}=\frac{n}{2}, and this value is generally a lower bound on σ\sigma^{\prime} in the more degenerate settings.

Proof of Theorem 1.13.

Let 𝐮=(u,u)QH01(v;Ω){\bf u}=(u,\nabla u)\in QH_{0}^{1}(v;\Omega) be a non-negative weak subsolution of (1.3). By the homogeneity of equation (1.3) and inequality (1.14), to prove this result it will suffice to assume that fLσ(v;Ω)=1\|f\|_{L^{\sigma^{\prime}}(v;\Omega)}=1 and prove that

(4.5) uL(v;Ω)C[1+log(1+fLA(v;Ω))].\|u\|_{L^{\infty}(v;\Omega)}\leq C[1+\log(1+\|f\|_{L^{A}(v;\Omega)})].

To prove (4.5) we will apply an iteration argument very similar to that in the proof of Theorem 1.9, but to the solution of an auxiliary equation we which now define. Given that fLσ(v;Ω)=1\|f\|_{L^{\sigma^{\prime}}(v;\Omega)}=1, and since by Theorem 1.9 uu is bounded in Ω\Omega, we can apply Lemma 2.26 and fix γ(0,4C02)\gamma\in(0,\frac{4}{C_{0}^{2}}) such that

(4.6) Ωeγu(x)v(x)𝑑xM(γ,C0,v(Ω))=M.\int_{\Omega}e^{\gamma u(x)}v(x)\,dx\leq M(\gamma,C_{0},v(\Omega))=M.

Define h=eγu/ph=e^{\gamma u/p} (where p>1p>1 will be determined below) and let w=h1w=h-1. By Lemma 2.19, (w,γphu)QH01(v;Ω)(w,\frac{\gamma}{p}h\nabla u)\in QH^{1}_{0}(v;\Omega) is a non-negative weak subsolution of

(4.9) {Div(Qw)=αfhvfor xΩ,w=0for xΩ.\displaystyle\left\{\begin{array}[]{rcll}-\operatorname{Div}\left(Q\nabla w\right)&=&\alpha fhv&\textrm{for }x\in\Omega,\\ w&=&0&\textrm{for }x\in\partial\Omega.\end{array}\right.

For each r>0r>0, let φr=(wr)+\varphi_{r}=(w-r)_{+} and S(r)={xΩ:w(x)>r}S(r)=\{x\in\Omega~{}:~{}w(x)>r\}. By Lemma 2.14, (φr,φr)QH01(v;Ω)(\varphi_{r},\nabla\varphi_{r})\in QH^{1}_{0}(v;\Omega). By Lemma 4.1, there exist b¯((2σ),σ),b(σ,2σ)\bar{b}\in((2\sigma)^{\prime},\sigma^{\prime}),\,b\in(\sigma,2\sigma), and p>1p>1 such that (4.2) holds. We can now argue as we did in the proof of Theorem 1.9 with φr\varphi_{r} as a test function, and then apply Hölder’s inequality twice to get

φrL2σ(v;Ω)2\displaystyle\|\varphi_{r}\|_{L^{2\sigma}(v;\Omega)}^{2} CS(r)φrQφrdx\displaystyle\leq C\int_{S(r)}\nabla\varphi_{r}Q\nabla\varphi_{r}~{}dx
=CS(r)φrQwdx\displaystyle=C\int_{S(r)}\nabla\varphi_{r}Q\nabla w~{}dx
CS(r)fφrhv𝑑x\displaystyle\leq C\int_{S(r)}f\varphi_{r}h~{}vdx
Cf𝟙S(r)Lb¯(v;Ω)φrLb(v;Ω)hLp(v;Ω)\displaystyle\leq C\|f{\mathbbm{1}}_{S(r)}\|_{L^{\bar{b}}(v;\Omega)}\|\varphi_{r}\|_{L^{{b}}(v;\Omega)}\|h\|_{L^{p}(v;\Omega)}
(4.10) Cf𝟙S(r)Lb¯(v;Ω)φrL2σ(v;Ω)v(S(r))2σb2σ;\displaystyle\leq C\|f{\mathbbm{1}}_{S(r)}\|_{L^{\bar{b}}(v;\Omega)}\|\varphi_{r}\|_{L^{2\sigma}(v;\Omega)}v(S(r))^{\frac{2\sigma-b}{2\sigma}};

the last inequality follows since b<2σb<2\sigma and since by (4.6), hLp(v;Ω)h\in L^{p}(v;\Omega) with a constant independent of 𝐮\bf u and ff.

Now define the Young function B(t)=tσb¯log(e+t)qB(t)=t^{\frac{\sigma^{\prime}}{\bar{b}}}\log(e+t)^{q} and note that B(|t|b¯)A(t)B(|t|^{\bar{b}})\preceq A(t). Therefore, arguing as before, by Lemma 2.7 and (4) we have that

φr2σCfAv(S(r))1b¯(σ/b¯)+2σb2σb¯log(e+v(S(r)1))qσ=CfAv(S(r))σb¯b¯σ+2σb2σb¯log(e+v(S(r))1)qσ\|\varphi_{r}\|_{2\sigma}\leq C\|f\|_{A}\frac{v(S(r))^{\frac{1}{\bar{b}(\sigma^{\prime}/\bar{b})^{\prime}}+\frac{2\sigma-b}{2\sigma\bar{b}}}}{\log(e+v(S(r)^{-1}))^{\frac{q}{\sigma^{\prime}}}}=C\|f\|_{A}\frac{v(S(r))^{\frac{\sigma^{\prime}-\bar{b}}{\bar{b}\sigma^{\prime}}+\frac{2\sigma-b}{2\sigma\bar{b}}}}{\log(e+v(S(r))^{-1})^{\frac{q}{\sigma^{\prime}}}}

We can now argue as we did in the proof of Theorem 1.9 to get that for all s>rs>r,

v(S(s))(CfA(sr))2σv(S(r))2σb¯(σb¯σ+2σb2σ)log(e+v(S(r))1)2qσσ=(CfA(sr))2σv(S(r))log(e+v(S(r))1)2qσσ;v(S(s))\leq\left(\frac{C\|f\|_{A}}{(s-r)}\right)^{2\sigma}\frac{v(S(r))^{\frac{2\sigma}{\bar{b}}\left(\frac{\sigma^{\prime}-\bar{b}}{\sigma^{\prime}}+\frac{2\sigma-b}{2\sigma}\right)}}{\log(e+v(S(r))^{-1})^{\frac{2q\sigma}{\sigma^{\prime}}}}=\left(\frac{C\|f\|_{A}}{(s-r)}\right)^{2\sigma}\frac{v(S(r))}{\log(e+v(S(r))^{-1})^{\frac{2q\sigma}{\sigma^{\prime}}}};

the last inequality holds by (4.3).

We continue the proof of Theorem 1.9 and define ϵ=qσ1>0\epsilon=\frac{q}{\sigma^{\prime}}-1>0, CkC_{k}, k0k\geq 0, as in (3.3), and mk=log(v(S(C(k)))m_{k}=-\log(v(S(C(k))) to again get the iteration inequality

(4.11) mk+12σlog(ϵτ0C)+2σqσlog(mkk+2)+mk.m_{k+1}\geq 2\sigma\log\left(\frac{\epsilon\tau_{0}}{C}\right)+\frac{2\sigma q}{\sigma^{\prime}}\log\left(\frac{m_{k}}{k+2}\right)+m_{k}.

We will again prove that we can choose the parameter τ0\tau_{0} such that m0>1m_{0}>1 and for every k{0}k\in\mathbb{N}\cup\{0\},

(4.12) mkm0+km_{k}\geq m_{0}+k

Assume for the moment that (4.12) holds. Then arguing as before we have that wτ0fA\|w\|_{\infty}\leq\tau_{0}\|f\|_{A}: that is,

ecuτ0(fA+1),e^{c||u||_{\infty}}\leq\tau_{0}(\|f\|_{A}+1),

which in turn implies that (4.5) holds as desired.

Therefore, to complete the proof we need to show that (4.12) holds. The proof is almost identical to the proof of (3.6): the only difference is in the choice of m0m_{0} which we will describe. We first estimate as we did for inequality (4):

wLb(v;Ω)2wL2σ(v;Ω)2v(Ω)2σbσCΩfwhv𝑑xv(Ω)2σbσCfLb¯(v;Ω)wLb(v;Ω)hLp(v;Ω)v(Ω)2σbσCfLb¯(v;Ω)wLb(v;Ω)v(Ω)2σbσ,\|w\|^{2}_{L^{b}(v;\Omega)}\leq\|w\|^{2}_{L^{2\sigma}(v;\Omega)}v(\Omega)^{\frac{2\sigma-b}{\sigma}}\leq C\int_{\Omega}fwh\,vdx\;v(\Omega)^{\frac{2\sigma-b}{\sigma}}\\ \leq C\|f\|_{L^{\bar{b}}(v;\Omega)}\|w\|_{L^{b}(v;\Omega)}\|h\|_{L^{p}(v;\Omega)}v(\Omega)^{\frac{2\sigma-b}{\sigma}}\leq C\|f\|_{L^{\bar{b}}(v;\Omega)}\|w\|_{L^{b}(v;\Omega)}v(\Omega)^{\frac{2\sigma-b}{\sigma}},

where the last inequality holds since hLp(v;Ω)h\in L^{p}(v;\Omega) with norm bounded by a constant. Furthermore, by Hölder’s inequality and Lemma 2.6,

fLb¯(v;Ω)b¯fLσ(v;Ω)b¯v(Ω)1(σ/b¯)fLA(v;Ω)b¯v(Ω)1(σ/b¯).\|f\|_{L^{\bar{b}}(v;\Omega)}^{\bar{b}}\leq\|f\|_{L^{\sigma^{\prime}}(v;\Omega)}^{\bar{b}}v(\Omega)^{\frac{1}{(\sigma^{\prime}/\bar{b})^{\prime}}}\leq\|f\|_{L^{A}(v;\Omega)}^{\bar{b}}v(\Omega)^{\frac{1}{(\sigma^{\prime}/\bar{b})^{\prime}}}.

Since C0=C1/2C_{0}=C_{1}/2, for every xS(C0)x\in S(C_{0}) we have 2w(x)C1>1\frac{2w(x)}{C_{1}}>1. Thus, combining the above inequalities, we get

v(S(C0))2C1S(C0)wv𝑑x2C1wLb(v;Ω)v(S(C0))1b¯2CC1fLA(v;Ω)v(S(C0))1b¯v(Ω)1b¯(σ/b¯)+2σbσ=Cτ0(12ϵ)v(S(C0))1b¯.v(S(C_{0}))\leq\frac{2}{C_{1}}\int_{S(C_{0})}w~{}vdx\leq\frac{2}{C_{1}}\|w\|_{L^{b}(v;\Omega)}v(S(C_{0}))^{\frac{1}{\bar{b}}}\\ \leq\frac{2C}{C_{1}}\|f\|_{L^{A}(v;\Omega)}v(S(C_{0}))^{\frac{1}{\bar{b}}}v(\Omega)^{\frac{1}{\bar{b}(\sigma^{\prime}/\bar{b})^{\prime}}+\frac{2\sigma-b}{\sigma}}=\frac{C}{\tau_{0}(1-2^{-\epsilon})}v(S(C_{0}))^{\frac{1}{\bar{b}}}.

Hence,

v(S0)(Cτ0(12ϵ))b,v(S_{0})\leq\left(\frac{C}{\tau_{0}(1-2^{-\epsilon})}\right)^{b},

and so we can choose τ0>0\tau_{0}>0 independent of both 𝐮,f{\bf u},f such that

μ0=v(S(C0))<e2,τ0max{eC12ϵ,eCϵ}\mu_{0}=v(S(C_{0}))<e^{-2},\quad\tau_{0}\geq\max\bigg{\{}\frac{eC}{1-2^{-\epsilon}},\frac{eC}{\epsilon}\bigg{\}}

where CC is as in (4.11). We may now proceed exactly as in the proof of  (3.6) to get that (4.12) holds. This completes our proof. ∎

5. Theorem 1.9 is almost sharp

In this section we construct Example 1.11 that shows that Theorem 1.9 is almost sharp in the case of the Laplacian. Our example is intuitively straightforward. Let our domain Ωn\Omega\subset\mathbb{R}^{n}, n3n\geq 3, be the unit ball B=B(0,1)B=B(0,1), and define

f(x)=|x|2log(e+|x|1)1.f(x)=|x|^{-2}\log(e+|x|^{-1})^{-1}.

Let A(t)=tn2log(e+t)qA(t)=t^{\frac{n}{2}}\log(e+t)^{q}. We will show that fLA(B)f\in L^{A}(B) if and only if q<n21q<\frac{n}{2}-1. Moreover, we claim that, at least formally, if uu is the solution of Δu=f\Delta u=f on BB, then u(0)=u(0)=\infty. For if we use the well-known fact that the Green’s function for the unit ball is cn|x|2nc_{n}|x|^{2-n}, then

u(0)=cnB|x|nlog(e+|x|1)1dx=.u(0)=c_{n}\int_{B}|x|^{-n}\log(e+|x|^{-1})^{-1}\,dx=\infty.

To make this argument rigorous we must justify our use of Green’s formula which requires that the function ff be continuous on BB. To overcome this, we give an approximation argument and show that the inequality

uL(B)CfLA(B)\|u\|_{L^{\infty}(B)}\leq C\|f\|_{L^{A}(B)}

cannot hold with a uniform constant. For each k1k\geq 1, let χk\chi_{k} be a continuous, non-negative, radial function such that χk(x)=0\chi_{k}(x)=0 if |x|2k1|x|\leq 2^{-k-1}, and χk(x)=1\chi_{k}(x)=1 if 2kx<12^{-k}\leq x<1. Define fk=ukf_{k}=u_{k}. Each fkf_{k} is continuous, and if uku_{k} is the solution to the Dirichlet problem

{Δuk=fkxB,uk=0xB,\begin{cases}\Delta u_{k}=f_{k}&x\in B,\\ u_{k}=0&x\in\partial B,\end{cases}

then at the origin it is given by

uk(0)=cnB|x|2nfk(x)dxcn2k|x|<1|x|nlog(e+|x|1)1dx.u_{k}(0)=c_{n}\int_{B}|x|^{2-n}f_{k}(x)\,dx\geq c_{n}\int_{2^{-k}\leq|x|<1}|x|^{-n}\log(e+|x|^{-1})^{-1}\,dx.

It is immediate that uk(0)u_{k}(0)\rightarrow\infty as kk\rightarrow\infty. Since by monotonicity of the norm, fkLA(B)fLA(B)\|f_{k}\|_{L^{A}(B)}\leq\|f\|_{L^{A}(B)}, we have that the inequality

uk(0)ukL(B)CfkLA(B)CfLA(B)u_{k}(0)\leq\|u_{k}\|_{L^{\infty}(B)}\leq C\|f_{k}\|_{L^{A}(B)}\leq C\|f\|_{L^{A}(B)}

cannot hold with a uniform constant if fLA(B)f\in L^{A}(B).

Therefore, to complete the proof, it will suffice to show fLA(B)f\in L^{A}(B) if and only if q<n21q<\frac{n}{2}-1. By the definition of the Luxemburg norm, it will suffice to show that f(A)L1(B)f(A)\in L^{1}(B). But this is straightforward:

A(f(x))\displaystyle A(f(x)) =f(x)n2log(e+f(x))q\displaystyle=f(x)^{\frac{n}{2}}\log(e+f(x))^{q}
=xnlog(e+|x|1)n2log(e+|x|2log(e+|x|1)1)q\displaystyle=x^{-n}\log(e+|x|^{-1})^{-\frac{n}{2}}\log(e+|x|^{-2}\log(e+|x|^{-1})^{-1})^{q}
xnlog(e+|x|1)n2log(e+|x|1)q,\displaystyle\approx x^{-n}\log(e+|x|^{-1})^{-\frac{n}{2}}\log(e+|x|^{-1})^{q},

where the implicit constant only depends on qq. Thus, A(f)L1(B)A(f)\in L^{1}(B) if and only if n2q>1\frac{n}{2}-q>1, or equivalently, q<n21q<\frac{n}{2}-1.

References

  • [1] S Chanillo and R. L. Wheeden. Weighted Poincaré and Sobolev inequalities and estimates for weighted Peano maximal functions. Amer. J. Math., 107(5):1191–1226, 1985.
  • [2] M Christ. Hypoellipticity in the infinitely degenerate regime, complex analysis and geometry (Columbus, OH, 1999). Ohio State Univ. Math. Res. Inst. Publ. de Gruyter, Berlin, 9:59–84, 2001.
  • [3] S.-K. Chua and R. L. Wheeden. Existence of weak solutions to degenerate pp-Laplacian equations and integral formulas. J. Differential Equations, 263(12):8186–8228, 2017.
  • [4] A. Cianchi. Strong and weak type inequalities for some classical operators in Orlicz spaces. J. London Math. Soc. (2), 60(1):187–202, 1999.
  • [5] D. Cruz-Uribe, J. M. Martell, and C. Pérez. Weights, extrapolation and the theory of Rubio de Francia, volume 215 of Operator Theory: Advances and Applications. Birkhäuser/Springer Basel AG, Basel, 2011.
  • [6] D. Cruz-Uribe, S. Rodney, and E. Rosta. Poincaré inequalities and Neumann problems for the pp-Laplacian. Canad. Math. Bull., 61(4):738–753, 2018.
  • [7] D. Cruz-Uribe, S. Rodney, and E. Rosta. Global Sobolev inequalities and degenerate pp-Laplacian equations. J. Differential Equations, 268(10):6189–6210, 2020.
  • [8] E. De Giorgi. Sulla differenziabilità e l’analiticità delle estremali degli integrali multipli regolari. Mem. Accad. Sci. Torino. Cl. Sci. Fis. Math. Nat., 3:25–43, 1957.
  • [9] L. C. Evans and R. F. Gariepy. Measure theory and fine properties of functions. Textbooks in Mathematics. CRC Press, Boca Raton, FL, revised edition, 2015.
  • [10] E. B. Fabes, C. E. Kenig, and R. P. Serapioni. The local regularity of solutions of degenerate elliptic equations. Comm. Partial Differential Equations, 7(1):77–116, 1982.
  • [11] G. B. Folland. Real analysis. Pure and Applied Mathematics (New York). John Wiley & Sons, Inc., New York, second edition, 1999. Modern techniques and their applications, A Wiley-Interscience Publication.
  • [12] D. Gilbarg and N. S. Trudinger. Elliptic Partial Differential Equations of Second Order. Classics in Mathematics. Springer-Verlag, Berlin, 2001. Reprint of the 1998 edition.
  • [13] L. Korobenko, C. Rios, E. T. Sawyer, and R. Shen. Sharp local boundedness and maximum principle in the infinitely degenerate regime via De Giorgi iteration. Memoirs Amer. Math. Soc., to appear. arXiv1608.01630v4.
  • [14] M. A. Krasnoselskiĭ and Ja. B. Rutickiĭ. Convex functions and Orlicz spaces. Translated from the first Russian edition by Leo F. Boron. P. Noordhoff Ltd., Groningen, 1961.
  • [15] M. T. Lacey and S. Spencer. On entropy bumps for Calderón-Zygmund operators. Concr. Oper., 2(1):47–52, 2015.
  • [16] V. G. Mazya. Some estimates of solutions of second-order elliptic equations. Dokl. Akad. Nauk SSSR, 137:1057–1059, 1961.
  • [17] V. G. Mazya. Weak solutions of the Dirichlet and Neumann problems. Trudy Moskov. Mat. Obšč., 20:137–172, 1969.
  • [18] D. D. Monticelli and S. Rodney. Existence and spectral theory for weak solutions of Neumann and Dirichlet problems for linear degenerate elliptic operators with rough coefficients. J. Differential Equations, 259(8):4009–4044, 2015.
  • [19] D. D. Monticelli, S. Rodney, and R. L. Wheeden. Boundedness of weak solutions of degenerate quasilinear equations with rough coefficients. Differential and Integral Equations, 25(1-2):143–200, 2012.
  • [20] M. M. Rao and Z. D. Ren. Theory of Orlicz spaces, volume 146 of Monographs and Textbooks in Pure and Applied Mathematics. Marcel Dekker Inc., New York, 1991.
  • [21] S. Rodney. Existence of weak solutions of linear subelliptic Dirichlet problems with rough coefficients. Canad. J. Math., 64(6):1395–1414, 2012.
  • [22] E. T. Sawyer and R. L. Wheeden. Degenerate Sobolev spaces and regularity of subelliptic equations. Trans. Amer. Math. Soc., 362(4):1869–1906, 2010.
  • [23] G. Stampacchia. Le problème de Dirichlet pour les équations elliptiques du second ordre à coefficients discontinus. Ann. Inst. Fourier (Grenoble), 15(fasc. 1):189–258, 1965.
  • [24] S. Treil and A. Volberg. Entropy conditions in two weight inequalities for singular integral operators. Adv. Math., 301:499–548, 2016.
  • [25] N. S. Trudinger. Linear elliptic operators with measurable coefficients. Ann. Scuola Norm. Sup. Pisa Cl. Sci. (3), 27:265–308, 1973.
  • [26] X. Xu. Logarithmic up bounds for solutions of elliptic partial differential equations. Proc. Amer. Math. Soc., 139(10):3485–3490, 2011.