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

Square function and non-tangential maximal function estimates for elliptic operators in 1-sided NTA domains satisfying the capacity density condition

Murat Akman Murat Akman
Department of Mathematical Sciences
University of Essex
Colchester CO4 3SQ, United Kingdom
murat.akman@essex.ac.uk
Steve Hofmann Steve Hofmann
Department of Mathematics
University of Missouri
Columbia, MO 65211, USA
hofmanns@missouri.edu
José María Martell José María Martell
Instituto de Ciencias Matemáticas CSIC-UAM-UC3M-UCM
Consejo Superior de Investigaciones Científicas
C/ Nicolás Cabrera, 13-15
E-28049 Madrid, Spain
chema.martell@icmat.es
 and  Tatiana Toro Tatiana Toro
University of Washington
Department of Mathematics
Box 354350
Seattle, WA 98195-4350
toro@uw.edu
Abstract.

Let Ωn+1\Omega\subset\mathbb{R}^{n+1}, n2n\geq 2, be a 1-sided non-tangentially accessible domain (aka uniform domain), that is, Ω\Omega satisfies the interior Corkscrew and Harnack chain conditions, which are respectively scale-invariant/quantitative versions of openness and path-connectedness. Let us assume also that Ω\Omega satisfies the so-called capacity density condition, a quantitative version of the fact that all boundary points are Wiener regular. Consider L0u=div(A0u)L_{0}u=-\mathrm{div}(A_{0}\nabla u), Lu=div(Au)Lu=-\mathrm{div}(A\nabla u), two real (non-necessarily symmetric) uniformly elliptic operators in Ω\Omega, and write ωL0\omega_{L_{0}}, ωL\omega_{L} for the respective associated elliptic measures. The goal of this program is to find sufficient conditions guaranteeing that ωL\omega_{L} satisfies an AA_{\infty}-condition or a RHqRH_{q}-condition with respect to ωL0\omega_{L_{0}}. In this paper we are interested in obtaining square function and non-tangential estimates for solutions of operators as before. We establish that bounded weak null-solutions satisfy Carleson measure estimates, with respect to the associated elliptic measure. We also show that for every weak null-solution, the associated square function can be controlled by the non-tangential maximal function in any Lebesgue space with respect to the associated elliptic measure. These results extend previous work of Dahlberg-Jerison-Kenig and are fundamental for the proof of the perturbation results in [2].

Key words and phrases:
Uniformly elliptic operators, elliptic measure, the Green function, 1-sided non-tangentially accessible domains, 1-sided chord-arc domains, capacity density condition, Ahlfors-regularity, AA_{\infty} Muckenhoupt weights, Reverse Hölder, Carleson measures, square function estimates, non-tangential maximal function estimates, dyadic analysis, sawtooth domains, perturbation
2010 Mathematics Subject Classification:
31B05, 35J08, 35J25, 42B37, 42B25, 42B99
The second author was partially supported by NSF grants DMS-1664047 and DMS-2000048. The third author acknowledges financial support from the Spanish Ministry of Science and Innovation, through the “Severo Ochoa Programme for Centres of Excellence in R&D” (CEX2019-000904-S) and through the grant MTM PID2019-107914GB-I00. The third author also acknowledges that the research leading to these results has received funding from the European Research Council under the European Union’s Seventh Framework Programme (FP7/2007-2013)/ ERC agreement no. 615112 HAPDEGMT. The fourth author was partially supported by the Craig McKibben & Sarah Merner Professor in Mathematics, by NSF grant number DMS-1664867 and DMS-1954545, and by the Simons Foundation Fellowship 614610.
This material is based upon work supported by the National Science Foundation under Grant No. DMS-1440140 while the authors were in residence at the Mathematical Sciences Research Institute in Berkeley, California, during the Spring 2017 semester.

1. Introduction and Main results

The purpose of this program is to study some perturbation problems for second order divergence form real elliptic operators with bounded measurable coefficients in domains with rough boundaries. Let Ωn+1\Omega\subset\mathbb{R}^{n+1}, n2n\geq 2, be an open set and let Lu=div(Au)Lu=-\mathop{\operatorname{div}}\nolimits(A\nabla u) be a second order divergence form real elliptic operator defined in Ω\Omega. Here the coefficient matrix A=(ai,j())i,j=1n+1A=(a_{i,j}(\cdot))_{i,j=1}^{n+1} is real (not necessarily symmetric) and uniformly elliptic, with ai,jL(Ω)a_{i,j}\in L^{\infty}(\Omega), that is, there exists a constant Λ1\Lambda\geq 1 such that

(1.1) Λ1|ξ|2A(X)ξξ,|A(X)ξη|Λ|ξ||η|\displaystyle\Lambda^{-1}|\xi|^{2}\leq A(X)\xi\cdot\xi,\qquad\qquad|A(X)\xi\cdot\eta|\leq\Lambda|\xi|\,|\eta|

for all ξ,ηn+1\xi,\eta\in\mathbb{R}^{n+1} and for almost every XΩX\in\Omega. Associated with LL one can construct a family of positive Borel measures {ωLX}XΩ\{\omega_{L}^{X}\}_{X\in\Omega}, defined on Ω\partial\Omega with ωX(Ω)1\omega^{X}(\partial\Omega)\leq 1 for every XΩX\in\Omega, so that for each fCc(Ω)f\in C_{c}(\partial\Omega) one can define its associated weak-solution

(1.2) u(X)=Ωf(z)𝑑ωLX(z),wheneverXΩ,u(X)=\int_{\partial\Omega}f(z)d\omega_{L}^{X}(z),\quad\mbox{whenever}\,\,X\in\Omega,

which satisfies Lu=0Lu=0 in Ω\Omega in the weak sense. In principle, unless we assume some further condition, uu need not be continuous all the way to the boundary, but still we think of uu as the solution to the continuous Dirichlet problem with boundary data ff. We call ωLX\omega_{L}^{X} the elliptic measure of Ω\Omega associated with the operator LL with pole at XΩX\in\Omega. For convenience, we will sometimes write ωL\omega_{L} and call it simply the elliptic measure, dropping the dependence on the pole.

Given two such operators L0u=div(A0u)L_{0}u=-\mathop{\operatorname{div}}\nolimits(A_{0}\nabla u) and Lu=div(Au)Lu=-\mathop{\operatorname{div}}\nolimits(A\nabla u), one may wonder whether one can find conditions on the matrices A0A_{0} and AA so that some “good estimates” for the Dirichlet problem or for the elliptic measure for L0L_{0} might be transferred to the operator LL. Similarly, one may try to see whether AA being “close” to A0A_{0} in some sense gives some relationship between ωL\omega_{L} and ωL0\omega_{L_{0}}. In this direction, a celebrated result of Littman, Stampacchia, and Weinberger in [27] states that the continuous Dirichlet problem for the Laplace operator L0=ΔL_{0}=\Delta, (i.e., A0A_{0} is the identity) is solvable if and only if it is solvable for any real elliptic operator LL. By solvability here we mean that the elliptic measure solutions as in (1.2) are indeed continuous in Ω¯\overline{\Omega}. It is well known that solvability in this sense is in fact equivalent to the fact that all boundary points are regular in the sense of Wiener, a condition which entails some capacitary thickness of the complement of Ω\Omega. Note that, for this result, one does not need to know that LL is “close” to the Laplacian in any sense (other than the fact that both operators are uniformly elliptic).

On the other hand, if Ω=+2\Omega=\mathbb{R}^{2}_{+} is the upper-half plane and L0=ΔL_{0}=\Delta, then the harmonic measure associated with Δ\Delta is mutually absolutely continuous with respect to the surface measure on the boundary, and its Radon-Nykodym derivative is the classical Poisson kernel. However, Caffarelli, Fabes, and Kenig in [3] constructed a uniformly real elliptic operator LL in the plane (the pullback of the Laplacian via a quasiconformal mapping of the upper half plane to itself) for which the associated elliptic measure ωL\omega_{L} is not even absolutely continuous with respect to the surface measure (see also [29] for another example). Hence, in principle the “good behavior” of harmonic measure does not always transfer to any elliptic measure even in a nice domain such as the upper-half plane. Consequently, it is natural to see if those good properties can be transferred by assuming some conditions reflecting the fact that LL is “close” to L0L_{0} or, in other words, by imposing some conditions on the disagreement of AA and A0A_{0}.

The goal of this program is to solve some perturbation problems that go beyond [9, 11, 12, 28, 5, 6]. Our setting is that of 1-sided NTA domains satisfying the so called capacity density condition (CDC for short), see Section 2 for the precise definitions. The latter is a quantitative version of the well-known Wiener criterion and it is weaker than the Ahlfors regularity of the boundary or the existence of exterior Corkscrews (see Definition 2.1). This setting guarantees among other things that any elliptic measure is doubling in some appropriate sense, hence one can see that a suitable portion of the boundary of the domain endowed with the Euclidean distance and with a given elliptic measure ωL0\omega_{L_{0}} is a space of homogeneous type. In particular, classes like A(ωL0)A_{\infty}(\omega_{L_{0}}) or RHp(ωL0)RH_{p}(\omega_{L_{0}}) have the same good features of the corresponding ones in the Euclidean setting. However, our assumptions do not guarantee that the surface measure has any good behavior and it could even be locally infinite. In one of our main results, we consider the case in which a certain disagreement condition, originating in [12], holds either with small or large constant. The small constant case can be seen as an extension of [12, 28] to a setting in which surface measure is not a good object. The large constant case is new even in nice domains such as balls, upper-half spaces, Lipschitz domains or chord-arc domains. To the best of our knowledge, our work is the first to establish perturbation results on sets with bad surface measures, and our large perturbation results are the first of their type. Finally, we do not require the operators to be symmetric. The precise results, along with its context in the historical developments, will be stated in the sequel to the present paper [2].

In the present article we develop some of the needed tools, and present some other results which are of independent interest. Key to our argument is the construction of certain sawtooth domains adapted to a dyadic grid on the boundary and to the Whitney decomposition of the domain. These domains are shown to inherit the main geometrical/topological features of the original domain (see Proposition 2.37). With this in hand we obtain a discrete sawtooth lemma for projections improving [10, Main Lemma], see Lemma 3.5 and Lemma 3.19. These ingredients are crucial for the main results of the papers which we state next. First we establish that bounded weak-solutions satisfy Carleson measure estimates adapted to the elliptic measure.

Theorem 1.3.

Let Ωn+1\Omega\subset\mathbb{R}^{n+1}, n2n\geq 2, be a 1-sided NTA domain (cf. Definition 2.5) satisfying the capacity density condition (cf. Definition 2.10). Let Lu=div(Au)Lu=-\mathop{\operatorname{div}}\nolimits(A\nabla u) be a real (not necessarily symmetric) elliptic operator and write ωL\omega_{L} and GLG_{L} to denote, respectively, the associated elliptic measure and the Green function. There exists CC depending only on dimension nn, the 1-sided NTA constants, the CDC constant, and the ellipticity constant of LL, such that for every uWloc1,2(Ω)L(Ω)u\in W^{1,2}_{\rm loc}(\Omega)\cap L^{\infty}(\Omega) with Lu=0Lu=0 in the weak-sense in Ω\Omega there holds

(1.4) supBsupB1ωLXΔ(Δ)BΩ|u(X)|2GL(XΔ,X)𝑑XCuL(Ω)2,\sup_{B}\sup_{B^{\prime}}\frac{1}{\omega_{L}^{X_{\Delta}}(\Delta^{\prime})}\iint_{B^{\prime}\cap\Omega}|\nabla u(X)|^{2}\,G_{L}(X_{\Delta},X)\,dX\leq\,C\,\|u\|_{L^{\infty}(\Omega)}^{2},

where Δ=BΩ\Delta=B\cap\partial\Omega, Δ=BΩ\Delta^{\prime}=B^{\prime}\cap\partial\Omega, XΔX_{\Delta} is a corkscrew point relative to Δ\Delta (cf. Definition 2.1), and the sups are taken respectively over all balls B=B(x,r)B=B(x,r) with xΩx\in\partial\Omega and 0<r<diam(Ω)0<r<\operatorname{diam}(\partial\Omega), and B=B(x,r)B^{\prime}=B(x^{\prime},r^{\prime}) with x2Δx^{\prime}\in 2\Delta and 0<r<rc0/40<r^{\prime}<rc_{0}/4, and c0c_{0} is the Corkscrew constant.

This result is in turn the main ingredient to obtain that the conical square function can be locally controlled by the non-tangential maximal function in norm with respect to the elliptic measure, allowing us to extend some estimates from [10] to our general setting.

Theorem 1.5.

Let Ωn+1\Omega\subset\mathbb{R}^{n+1}, n2n\geq 2, be a 1-sided NTA domain (cf. Definition 2.5) satisfying the capacity density condition (cf. Definition 2.10). Let Lu=div(Au)Lu=-\mathop{\operatorname{div}}\nolimits(A\nabla u) be a real (non-necessarily symmetric) elliptic operator and write ωL\omega_{L} to denote the associated elliptic measure and the Green function. For every 0<q<0<q<\infty, there exists CqC_{q} depending only on dimension nn, the 1-sided NTA constants, the CDC constant, the ellipticity constant of LL, and qq, such that for every uWloc1,2(Ω)u\in W^{1,2}_{\rm loc}(\Omega) with Lu=0Lu=0 in the weak-sense in Ω\Omega, for every Q0𝔻Q_{0}\in\mathbb{D}, there holds

(1.6) 𝒮Q0uLq(Q0,ωLXQ0)Cq𝒩Q0uLq(Q0,ωLXQ0),\|\mathcal{S}_{Q_{0}}u\|_{L^{q}(Q_{0},\omega_{L}^{X_{Q_{0}}})}\leq C_{q}\,\|\mathcal{N}_{Q_{0}}u\|_{L^{q}(Q_{0},\omega_{L}^{X_{Q_{0}}})},

where 𝒮Q0\mathcal{S}_{Q_{0}} and 𝒩Q0\mathcal{N}_{Q_{0}} are the localized dyadic conical square function and non-tangential maximal function respectively (cf. (2.28) and (2.27)), and XQ0X_{Q_{0}} is a corkscrew point relative to Q0Q_{0} (see Section 2.4).

We note that the estimate (1.6) is written for the localized dyadic conical square function and non-tangential maximal function. It is not difficult to see that, as a consequence, one can obtain a similar estimate for the regular localized (or truncated) conical square function and non-tangential maximal function with arbitrary apertures (see [4, Lemma 4.8]), precise statements are left to the interested reader.

The plan of this paper is as follows. Section 2 presents some of the preliminaries, definitions, and tools which will be used throughout the paper. Section 3 contains a dyadic version of the main lemma of [10]. In Section 4 we prove our main results, Theorem 1.3 and Theorem 1.5.

We would like to mention that after an initial version of this work was posted on arXiv [1], Feneuil and Poggi in [13] obtained results related to ours, compare for instance Theorem 1.3 with [13, Theorem 1.27]. Also, the recent work [4] complements this paper and its companion [2], see for instance [4, Corollary 1.4].

2. Preliminaries

2.1. Notation and conventions

  • \bullet

    We use the letters c,Cc,C to denote harmless positive constants, not necessarily the same at each occurrence, which depend only on dimension and the constants appearing in the hypotheses of the theorems (which we refer to as the “allowable parameters”). We shall also sometimes write aba\lesssim b and aba\approx b to mean, respectively, that aCba\leq Cb and 0<ca/bC0<c\leq a/b\leq C, where the constants cc and CC are as above, unless explicitly noted to the contrary. Unless otherwise specified upper case constants are greater than 11 and lower case constants are smaller than 11. In some occasions it is important to keep track of the dependence on a given parameter γ\gamma, in that case we write aγba\lesssim_{\gamma}b or aγba\approx_{\gamma}b to emphasize that the implicit constants in the inequalities depend on γ\gamma.

  • \bullet

    Our ambient space is n+1\mathbb{R}^{n+1}, n2n\geq 2.

  • \bullet

    Given En+1E\subset\mathbb{R}^{n+1} we write diam(E)=supx,yE|xy|\operatorname{diam}(E)=\sup_{x,y\in E}|x-y| to denote its diameter.

  • \bullet

    Given a domain Ωn+1\Omega\subset\mathbb{R}^{n+1}, we shall use lower case letters x,y,zx,y,z, etc., to denote points on Ω\partial\Omega, and capital letters X,Y,ZX,Y,Z, etc., to denote generic points in n+1\mathbb{R}^{n+1} (especially those in n+1Ω\mathbb{R}^{n+1}\setminus\partial\Omega).

  • \bullet

    The open (n+1)(n+1)-dimensional Euclidean ball of radius rr will be denoted B(x,r)B(x,r) when the center xx lies on Ω\partial\Omega, or B(X,r)B(X,r) when the center Xn+1ΩX\in\mathbb{R}^{n+1}\setminus\partial\Omega. A surface ball is denoted Δ(x,r):=B(x,r)Ω\Delta(x,r):=B(x,r)\cap\partial\Omega, and unless otherwise specified it is implicitly assumed that xΩx\in\partial\Omega.

  • \bullet

    If Ω\partial\Omega is bounded, it is always understood (unless otherwise specified) that all surface balls have radii controlled by the diameter of Ω\partial\Omega, that is, if Δ=Δ(x,r)\Delta=\Delta(x,r) then rdiam(Ω)r\lesssim\operatorname{diam}(\partial\Omega). Note that in this way Δ=Ω\Delta=\partial\Omega if diam(Ω)<rdiam(Ω)\operatorname{diam}(\partial\Omega)<r\lesssim\operatorname{diam}(\partial\Omega).

  • \bullet

    For Xn+1X\in\mathbb{R}^{n+1}, we set δ(X):=dist(X,Ω)\delta(X):=\operatorname{dist}(X,\partial\Omega).

  • \bullet

    We let n\mathcal{H}^{n} denote the nn-dimensional Hausdorff measure.

  • \bullet

    For a Borel set An+1A\subset\mathbb{R}^{n+1}, we let 𝟏A\mathbf{1}_{A} denote the usual indicator function of AA, i.e. 𝟏A(X)=1\mathbf{1}_{A}(X)=1 if XAX\in A, and 𝟏A(X)=0\mathbf{1}_{A}(X)=0 if XAX\notin A.

  • \bullet

    We shall use the letter II (and sometimes JJ) to denote a closed (n+1)(n+1)-dimensional Euclidean cube with sides parallel to the coordinate axes, and we let (I)\ell(I) denote the side length of II. We use QQ to denote dyadic “cubes” on EE or Ω\partial\Omega. The latter exist as a consequence of Lemma 2.13 below.

2.2. Some definitions

Definition 2.1 (Corkscrew condition).

Following [25], we say that an open set Ωn+1\Omega\subset\mathbb{R}^{n+1} satisfies the Corkscrew condition if for some uniform constant 0<c0<10<c_{0}<1 and for every xΩx\in\partial\Omega and 0<r<diam(Ω)0<r<\operatorname{diam}(\partial\Omega), if we write Δ:=Δ(x,r)\Delta:=\Delta(x,r), there is a ball B(XΔ,c0r)B(x,r)ΩB(X_{\Delta},c_{0}r)\subset B(x,r)\cap\Omega. The point XΔΩX_{\Delta}\subset\Omega is called a Corkscrew point relative to Δ\Delta (or, relative to BB). We note that we may allow r<Cdiam(Ω)r<C\operatorname{diam}(\partial\Omega) for any fixed CC, simply by adjusting the constant c0c_{0}. We say that Ω\Omega satisfies the exterior Corkscrew condition if Ωext:=n+1Ω¯\Omega_{ext}:=\mathbb{R}^{n+1}\setminus\overline{\Omega} satisfies the Corkscrew condition.

Definition 2.2 (Harnack Chain condition).

Again following [25], we say that Ω\Omega satisfies the Harnack Chain condition if there are uniform constants C1,C2>1C_{1},C_{2}>1 such that for every pair of points X,XΩX,X^{\prime}\in\Omega there is a chain of balls B1,B2,,BNΩB_{1},B_{2},\dots,B_{N}\subset\Omega with NC1(2+log2+Π)N\leq C_{1}(2+\log_{2}^{+}\Pi) where

(2.3) Π:=|XX|min{δ(X),δ(X)}.\Pi:=\frac{|X-X^{\prime}|}{\min\{\delta(X),\delta(X^{\prime})\}}.

such that XB1X\in B_{1}, XBNX^{\prime}\in B_{N}, BkBk+1ØB_{k}\cap B_{k+1}\neq\mbox{{\O}} and for every 1kN1\leq k\leq N

(2.4) C21diam(Bk)dist(Bk,Ω)C2diam(Bk).C_{2}^{-1}\operatorname{diam}(B_{k})\leq\operatorname{dist}(B_{k},\partial\Omega)\leq C_{2}\operatorname{diam}(B_{k}).

The chain of balls is called a Harnack Chain.

We note that in the context of the previous definition if Π1\Pi\leq 1 we can trivially form the Harnack chain B1=B(X,3δ(X)/5)B_{1}=B(X,3\delta(X)/5) and B2=B(X,3δ(X)/5)B_{2}=B(X^{\prime},3\delta(X^{\prime})/5) where (2.4) holds with C2=3C_{2}=3. Hence the Harnack chain condition is non-trivial only when Π>1\Pi>1.

Definition 2.5 (1-sided NTA and NTA).

We say that a domain Ω\Omega is a 1-sided non-tangentially accessible domain (1-sided NTA) if it satisfies both the Corkscrew and Harnack Chain conditions. Furthermore, we say that Ω\Omega is a non-tangentially accessible domain (NTA domain) if it is a 1-sided NTA domain and if, in addition, Ωext:=n+1Ω¯\Omega_{\rm ext}:=\mathbb{R}^{n+1}\setminus\overline{\Omega} also satisfies the Corkscrew condition.

Remark 2.6.

In the literature, 1-sided NTA domains are also called uniform domains. We remark that the 1-sided NTA condition is a quantitative form of path connectedness.

Definition 2.7 (Ahlfors regular).

We say that a closed set En+1E\subset\mathbb{R}^{n+1} is nn-dimensional Ahlfors regular (AR for short) if there is some uniform constant C1>1C_{1}>1 such that

(2.8) C11rnn(EB(x,r))C1rn,xE,0<r<diam(E).C_{1}^{-1}\,r^{n}\leq\mathcal{H}^{n}(E\cap B(x,r))\leq C_{1}\,r^{n},\qquad x\in E,\quad 0<r<\operatorname{diam}(E).
Definition 2.9 (1-sided CAD and CAD).

A 1-sided chord-arc domain (1-sided CAD) is a 1-sided NTA domain with AR boundary. A chord-arc domain (CAD) is an NTA domain with AR boundary.

We next recall the definition of the capacity of a set. Given an open set Dn+1D\subset\mathbb{R}^{n+1} (where we recall that we always assume that n2n\geq 2) and a compact set KDK\subset D we define the capacity of KK relative to DD as

Cap2(K,D)=inf{D|v(X)|2dX:vC0(D),v(x)1 in K}.\mathop{\operatorname{Cap}_{2}}\nolimits(K,D)=\inf\left\{\iint_{D}|\nabla v(X)|^{2}dX:\,\,v\in C^{\infty}_{0}(D),\,v(x)\geq 1\mbox{ in }K\right\}.
Definition 2.10 (Capacity density condition).

An open set Ω\Omega is said to satisfy the capacity density condition (CDC for short) if there exists a uniform constant c1>0c_{1}>0 such that

(2.11) Cap2(B(x,r)¯Ω,B(x,2r))Cap2(B(x,r)¯,B(x,2r))c1\frac{\mathop{\operatorname{Cap}_{2}}\nolimits(\overline{B(x,r)}\setminus\Omega,B(x,2r))}{\mathop{\operatorname{Cap}_{2}}\nolimits(\overline{B(x,r)},B(x,2r))}\geq c_{1}

for all xΩx\in\partial\Omega and 0<r<diam(Ω)0<r<\operatorname{diam}(\partial\Omega).

The CDC is also known as the uniform 2-fatness as studied by Lewis in [26]. Using [15, Example 2.12] one has that

(2.12) Cap2(B(x,r)¯,B(x,2r))rn1,for all xn+1 and r>0,\mathop{\operatorname{Cap}_{2}}\nolimits(\overline{B(x,r)},B(x,2r))\approx r^{n-1},\qquad\mbox{for all $x\in\mathbb{R}^{n+1}$ and $r>0$},

and hence the CDC is a quantitative version of the Wiener regularity, in particular every xΩx\in\partial\Omega is Wiener regular. It is easy to see that the exterior Corkscrew condition implies CDC. Also, it was proved in [30, Section 3] and [16, Lemma 3.27] that a set with Ahlfors regular boundary satisfies the capacity density condition with constant c1c_{1} depending only on nn and the Ahlfors regular constant.

2.3. Existence of a dyadic grid

In this section we introduce a dyadic grid along the lines of that obtained in [7]. More precisely, we will use the dyadic structure from [23, 24], with a modification from [22, Proof of Proposition 2.12]:

Lemma 2.13 (Existence and properties of the “dyadic grid”).

Let En+1E\subset\mathbb{R}^{n+1} be a closed set. Then there exists a constant C1C\geq 1 depending just on nn such that for each kk\in\mathbb{Z} there is a collection of Borel sets (called “cubes”)

𝔻k:={QjkE:j𝔍k},\mathbb{D}_{k}:=\big{\{}Q_{j}^{k}\subset E:\ j\in\mathfrak{J}_{k}\big{\}},

where 𝔍k\mathfrak{J}_{k} denotes some (possibly finite) index set depending on kk satisfying:

  • (a)(a)

    E=j𝔍kQjkE=\bigcup_{j\in\mathfrak{J}_{k}}Q_{j}^{k} for each kk\in\mathbb{Z}.

  • (b)(b)

    If mkm\leq k then either QjkQimQ_{j}^{k}\subset Q_{i}^{m} or QimQjk=ØQ_{i}^{m}\cap Q_{j}^{k}=\mbox{{\O}}.

  • (c)(c)

    For each kk\in\mathbb{Z}, j𝔍kj\in\mathfrak{J}_{k}, and m<km<k, there is a unique i𝔍mi\in\mathfrak{J}_{m} such that QjkQimQ_{j}^{k}\subset Q_{i}^{m}.

  • (d)(d)

    For each kk\in\mathbb{Z}, j𝔍kj\in\mathfrak{J}_{k} there is xjkEx_{j}^{k}\in E such that

    B(xjk,C12k)EQjkB(xjk,C2k)E.B(x_{j}^{k},C^{-1}2^{-k})\cap E\subset Q_{j}^{k}\subset B(x_{j}^{k},C2^{-k})\cap E.
Proof.

We first note that EE is geometric doubling. That is, there exists NN depending just on nn such that for every xEx\in E and r>0r>0 one can cover the surface ball B(x,r)EB(x,r)\cap E with at most NN surface balls of the form B(xi,r/2)EB(x_{i},r/2)\cap E with xiEx_{i}\in E —observe that geometric doubling for EE is inherited from the corresponding property on n+1\mathbb{R}^{n+1} and that is why NN depends only on nn and it is independent of EE. Besides, letting η=116\eta=\frac{1}{16}, for every kk\in{\mathbb{Z}} it is easy to find a countable collection {xjk}j𝔍kE\{x_{j}^{k}\}_{j\in\mathfrak{J}_{k}}\subset E such that

|xjkxjk|ηk,j,j𝔍k,jj;minj𝔍k|xxj|<ηk,xE.|x_{j}^{k}-x_{j^{\prime}}^{k}|\geq\eta^{k},\qquad j,j^{\prime}\in\mathfrak{J}_{k},\ j\neq j^{\prime};\qquad\min_{j\in\mathfrak{J}_{k}}|x-x_{j}|<\eta^{k},\qquad\forall\,x\in E.

Invoking then [23, 24] on EE with the Euclidean distance and c0=C0=1c_{0}=C_{0}=1 one can construct a family of dyadic cubes associated with these families of points, say 𝔇k\mathfrak{D}_{k} for kk\in{\mathbb{Z}}. These satisfy (a)(a)(d)(d) in the statement with the only difference that we have to replace 2k2^{-k} by ηk\eta^{k} in (d)(d).

At this point we follow the argument in [22, Proof of Proposition 2.12] with η=116\eta=\frac{1}{16}. For any kk\in{\mathbb{Z}} we set 𝔻j=𝔇k\mathbb{D}_{j}=\mathfrak{D}_{k} for every 4kj<4(k+1)4k\leq j<4(k+1). It is straightforward to show that properties (a)(a), (b)(b) and (c)(c) for the families 𝔻k\mathbb{D}_{k} follow at once from those for the families 𝔇k\mathfrak{D}_{k}. Regarding (d)(d), let Qi𝔻jQ^{i}\in\mathbb{D}_{j} and let kk\in{\mathbb{Z}} such that 4kj<4(k+1)4k\leq j<4(k+1) so that Qi𝔻j=𝔇kQ^{i}\in\mathbb{D}_{j}=\mathfrak{D}_{k}. Writing xiEx^{i}\in E for the corresponding point associated with Qi𝔇kQ^{i}\in\mathfrak{D}_{k} and invoking (d)(d) for 𝔇k\mathfrak{D}_{k} we conclude

B(xi,C12j)EB(xi,C1ηk)EQiB(xi,Cηk)EB(xi,16C2j)E,B(x^{i},{C}^{-1}2^{-j})\cap E\subset B(x^{i},C^{-1}\eta^{k})\cap E\subset Q^{i}\subset B(x^{i},C\eta^{k})\cap E\subset B(x^{i},16C2^{-j})\cap E,

hence (d)(d) holds. ∎

A few remarks are in order concerning this lemma. Note that by construction, within the same generation (that is, within each 𝔻k\mathbb{D}_{k}) the cubes are pairwise disjoint (hence, there are no repetitions). On the other hand, repetitions are allowed in the different generations, that is, one could have that Q𝔻kQ\in\mathbb{D}_{k} and Q𝔻k1Q^{\prime}\in\mathbb{D}_{k-1} agree. Then, although QQ and QQ^{\prime} are the same set, as cubes we understand that they are different. In short, it is then understood that 𝔻\mathbb{D} is an indexed collection of sets where repetitions of sets are allowed in the different generations but not within the same generation. With this in mind, we can give a proper definition of the “length” of a cube (this concept has no geometric meaning for the moment). For every Q𝔻kQ\in\mathbb{D}_{k}, we set (Q)=2k\ell(Q)=2^{-k}, which is called the “length” of QQ. Note that the “length” is well defined when considered on 𝔻\mathbb{D}, but it is not well-defined on the family of sets induced by 𝔻\mathbb{D}. It is important to observe that the “length” refers to the way the cubes are organized in the dyadic grid and in general may not have a geometrical meaning. It is clear from (d)(d) that diam(Q)(Q)\operatorname{diam}(Q)\lesssim\ell(Q) (we will see below that in our setting the converse hold, see Remark 2.56).

Let us observe that the generations run for all kk\in{\mathbb{Z}}. However, as we are about to see, sometimes it is natural to truncate the generations. If EE is bounded and kk\in{\mathbb{Z}} is such that diam(E)<C12k\operatorname{diam}(E)<C^{-1}2^{-k}, then there cannot be two distinct cubes in 𝔻k\mathbb{D}_{k}. Thus, 𝔻k={Qk}\mathbb{D}_{k}=\{Q^{k}\} with Qk=EQ^{k}=E. Therefore, we are going to ignore those kk\in\mathbb{Z} such that 2kdiam(E)2^{-k}\gtrsim\operatorname{diam}(E). Hence, we shall denote by 𝔻(E)\mathbb{D}(E) the collection of all relevant QjkQ_{j}^{k}, i.e., 𝔻(E):=k𝔻k\mathbb{D}(E):=\bigcup_{k}\mathbb{D}_{k}, where, if diam(E)\operatorname{diam}(E) is finite, the union runs over those kk\in\mathbb{Z} such that 2kdiam(E)2^{-k}\lesssim\operatorname{diam}(E).

In what follows given B=B(x,r)B=B(x,r) with xEx\in E we will denote Δ=Δ(x,r)=BE\Delta=\Delta(x,r)=B\cap E. We write Ξ=2C2\Xi=2C^{2}, with CC being the constant in Lemma 2.13, which is a purely dimensional. For Q𝔻(E)Q\in\mathbb{D}(E) we will set k(Q)=kk(Q)=k if Q𝔻kQ\in\mathbb{D}_{k}. Property (d)(d) implies that for each cube Q𝔻Q\in\mathbb{D}, there exist xQEx_{Q}\in E and rQr_{Q}, with Ξ1(Q)rQ(Q)\Xi^{-1}\ell(Q)\leq r_{Q}\leq\ell(Q) (indeed rQ=(2C)1(Q)r_{Q}=(2C)^{-1}\ell(Q)), such that

(2.14) Δ(xQ,2rQ)QΔ(xQ,ΞrQ).\Delta(x_{Q},2r_{Q})\subset Q\subset\Delta(x_{Q},\Xi r_{Q}).

We shall denote these balls and surface balls by

(2.15) BQ:=B(xQ,rQ),ΔQ:=Δ(xQ,rQ),B_{Q}:=B(x_{Q},r_{Q}),\qquad\Delta_{Q}:=\Delta(x_{Q},r_{Q}),
(2.16) B~Q:=B(xQ,ΞrQ),Δ~Q:=Δ(xQ,ΞrQ),\widetilde{B}_{Q}:=B(x_{Q},\Xi r_{Q}),\qquad\widetilde{\Delta}_{Q}:=\Delta(x_{Q},\Xi r_{Q}),

and we shall refer to the point xQx_{Q} as the “center” of QQ.

Let Q𝔻kQ\in\mathbb{D}_{k} and consider the family of its dyadic children {Q𝔻k+1:QQ}\{Q^{\prime}\in\mathbb{D}_{k+1}:Q^{\prime}\subset Q\}. Note that for any two distinct children Q,Q′′Q^{\prime},Q^{\prime\prime}, one has |xQxQ′′|rQ=rQ′′=rQ/2|x_{Q^{\prime}}-x_{Q^{\prime\prime}}|\geq r_{Q^{\prime}}=r_{Q^{\prime\prime}}=r_{Q}/2, otherwise xQ′′Q′′ΔQQ′′Qx_{Q^{\prime\prime}}\in Q^{\prime\prime}\cap\Delta_{Q^{\prime}}\subset Q^{\prime\prime}\cap Q^{\prime}, contradicting the fact that QQ^{\prime} and Q′′Q^{\prime\prime} are disjoint. Also xQ,xQ′′QΔ(xQ,rQ)x_{Q^{\prime}},x_{Q^{\prime\prime}}\in Q\subset\Delta(x_{Q},r_{Q}), hence by the geometric doubling property we have a purely dimensional bound for the number of such xQx_{Q^{\prime}} and hence the number of dyadic children of a given dyadic cube is uniformly bounded.

Lemma 2.17.

Let En+1E\subset\mathbb{R}^{n+1} be a closed set and let 𝔻(E)\mathbb{D}(E) be the dyadic grid as in Lemma 2.13. Assume that there is a Borel measure μ\mu which is doubling, that is, there exists Cμ1C_{\mu}\geq 1 such that μ(Δ(x,2r))Cμμ(Δ(x,r))\mu(\Delta(x,2r))\leq C_{\mu}\mu(\Delta(x,r)) for every xEx\in E and r>0r>0. Then μ(Q)=0\mu(\partial Q)=0 for every Q𝔻(E)Q\in\mathbb{D}(E). Moreover, there exist 0<τ0<10<\tau_{0}<1, C1C_{1}, and η>0\eta>0 depending only on dimension and CμC_{\mu} such that for every τ(0,τ0)\tau\in(0,\tau_{0}) and Q𝔻(E)Q\in\mathbb{D}(E)

(2.18) μ({xQ:dist(x,EQ)τ(Q)})C1τημ(Q).\mu\big{(}\big{\{}x\in Q:\,\operatorname{dist}(x,E\setminus Q)\leq\tau\ell(Q)\big{\}}\big{)}\leq C_{1}\tau^{\eta}\mu(Q).
Proof.

The argument is a refinement of that in [17, Proposition 6.3] (see also [14, p. 403] where the Euclidean case was treated). Fix an integer kk, a cube Q𝔻kQ\in\mathbb{D}_{k}, and a positive integer mm to be chosen. Fix τ>0\tau>0 small enough to be chosen and write

Στ={xQ¯:dist(x,EQ)<τ(Q)}.\Sigma_{\tau}=\big{\{}x\in\overline{Q}:\operatorname{dist}(x,E\setminus Q)<\tau\ell(Q)\big{\}}.

We set

{Qi1}:=𝔻1:=𝔻Q𝔻k+m,\{Q^{1}_{i}\}:=\mathbb{D}^{1}:=\mathbb{D}_{Q}\cap\mathbb{D}_{k+m}\,,

and make the disjoint decomposition Q=Qi1.Q=\bigcup Q^{1}_{i}. We then split 𝔻1=𝔻1,1𝔻1,2\mathbb{D}^{1}=\mathbb{D}^{1,1}\cup\mathbb{D}^{1,2}, where Qi1𝔻1,1Q^{1}_{i}\in\mathbb{D}^{1,1} if Qi1¯\overline{Q_{i}^{1}} meets Στ\Sigma_{\tau}, and Qi1𝔻1,2Q^{1}_{i}\in\mathbb{D}^{1,2} otherwise. We then write Q¯=R1,1R1,2\overline{Q}=R^{1,1}\cup R^{1,2}, where

R1,1:=𝔻1,1Q^i1,R1,2:=𝔻1,2Qi1,R^{1,1}:=\bigcup_{\mathbb{D}^{1,1}}\widehat{Q}^{1}_{i}\,,\qquad R^{1,2}:=\bigcup_{\mathbb{D}^{1,2}}Q^{1}_{i},

and for each cube Qi1𝔻1,1Q^{1}_{i}\in\mathbb{D}^{1,1}, we construct Q^i1\widehat{Q}^{1}_{i} as follows. We enumerate the elements in 𝔻1,1\mathbb{D}^{1,1} as Qi11,Qi21,,QiN1Q^{1}_{i_{1}},Q^{1}_{i_{2}},\dots,Q^{1}_{i_{N}}, and then set (Qi1)=Qi1(Qi1Q)(Q^{1}_{i})^{*}=Q^{1}_{i}\cup(\partial Q^{1}_{i}\cap\partial Q) and

Q^i11:=(Qi11),Q^i21:=(Qi21)(Qi11),Q^i31:=(Qi31)((Qi11)(Qi21)),\widehat{Q}^{1}_{i_{1}}:=(Q^{1}_{i_{1}})^{*},\quad\widehat{Q}^{1}_{i_{2}}:=(Q^{1}_{i_{2}})^{*}\setminus(Q^{1}_{i_{1}})^{*},\quad\widehat{Q}^{1}_{i_{3}}:=(Q^{1}_{i_{3}})^{*}\setminus((Q^{1}_{i_{1}})^{*}\cup(Q^{1}_{i_{2}})^{*}),\ \dots

so that R1,1R^{1,1} covers Στ\Sigma_{\tau} and the modified cubes Q^i1\widehat{Q}^{1}_{i} are pairwise disjoint.

We also note from (2.14) that if 2m<Ξ2/42^{-m}<\Xi^{-2}/4 then

dist(ΔQ,EQ)rQΞ1(Q),diam(Qi1)2ΞrQi12Ξ(Qi1)<Ξ12(Q).\operatorname{dist}\big{(}\Delta_{Q},E\setminus Q\Big{)}\geq r_{Q}\geq\Xi^{-1}\ell(Q),\qquad\operatorname{diam}(Q^{1}_{i})\leq 2\Xi r_{Q^{1}_{i}}\leq 2\Xi\ell(Q^{1}_{i})<\frac{\Xi^{-1}}{2}\ell(Q).

Then R1,1R^{1,1} misses ΔQ\Delta_{Q} provided τ<Ξ1/2\tau<\Xi^{-1}/2. Otherwise, we can find xQi1¯ΔQx\in\overline{Q^{1}_{i}}\cap\Delta_{Q} with Qi1𝔻1,1Q^{1}_{i}\in\mathbb{D}^{1,1}. The latter implies that there is yQi1¯Στy\in\overline{Q^{1}_{i}}\cap\Sigma_{\tau}. All these yield a contradiction:

Ξ1(Q)dist(ΔQ,EQ)|xy|+dist(y,EQ)diam(Qi1¯)+τ(Q)<Ξ1(Q).\displaystyle\Xi^{-1}\ell(Q)\leq\operatorname{dist}\big{(}\Delta_{Q},E\setminus Q\Big{)}\leq|x-y|+\operatorname{dist}\big{(}y,E\setminus Q\big{)}\leq\operatorname{diam}(\overline{Q^{1}_{i}})+\tau\ell(Q)<\Xi^{-1}\ell(Q).

Consequently, by the doubling property,

μ(Q¯)μ(2Δ~Q)Cμμ(ΔQ)Cμμ(R1,2).\mu(\overline{Q})\leq\mu(2\widetilde{\Delta}_{Q})\leq C_{\mu}^{\prime}\,\mu(\Delta_{Q})\leq C_{\mu}^{\prime}\,\mu(R^{1,2}).

Since R1,1R^{1,1} and R1,2R^{1,2} are disjoint, the latter estimate yields

μ(R1,1)(11Cμ)μ(Q¯)=:θμ(Q¯),\mu(R^{1,1})\leq\Big{(}1-\frac{1}{C_{\mu}^{\prime}}\Big{)}\,\mu(\overline{Q})=:\theta\,\mu(\overline{Q}),

where we note that 0<θ<10<\theta<1.

Let us now repeat this procedure, decomposing Q^i1\widehat{Q}^{1}_{i} for each Qi1𝔻1,1Q_{i}^{1}\in\mathbb{D}^{1,1}. We set 𝔻2(Qi1)=𝔻Qi1𝔻k+2m\mathbb{D}^{2}(Q^{1}_{i})=\mathbb{D}_{Q^{1}_{i}}\cap\mathbb{D}_{k+2m} and split it into 𝔻2,1(Qi1)\mathbb{D}^{2,1}(Q^{1}_{i}) and 𝔻2,2(Qi1)\mathbb{D}^{2,2}(Q^{1}_{i}) where Q𝔻2,1(Qi1)Q^{\prime}\in\mathbb{D}^{2,1}(Q^{1}_{i}) if Q¯\overline{Q^{\prime}} meets Στ\Sigma_{\tau}. Associated to any Q𝔻2,1(Qi1)Q^{\prime}\in\mathbb{D}^{2,1}(Q^{1}_{i}) we set (Q)=(QQ^i1)(Q(QQ^i1))(Q^{\prime})^{*}=(Q^{\prime}\cap\widehat{Q}^{1}_{i})\cup(\partial Q^{\prime}\cap(\partial Q\cap\widehat{Q}^{1}_{i})). Then we make these sets disjoint as before and we have that R2,1(Qi1)R^{2,1}(Q^{1}_{i}) is defined as the disjoint union of the corresponding Q^\widehat{Q^{\prime}}. Note that Q^i1=R2,1(Qi1)R2,2(Qi1)\widehat{Q}^{1}_{i}=R^{2,1}(Q^{1}_{i})\cup R^{2,2}(Q^{1}_{i}) and this is a disjoint union. As before, R2,1(Qi1)R^{2,1}(Q^{1}_{i}) misses ΔQi1\Delta_{Q_{i}^{1}} provided τ<2mΞ1/2\tau<2^{-m}\Xi^{-1}/2 so that by the doubling property

μ(Q^i1)μ(2Δ~Qi1)Cμμ(ΔQi1)Cμμ(R2,2(Qi1))\mu(\widehat{Q}^{1}_{i})\leq\mu(2\widetilde{\Delta}_{Q^{1}_{i}})\leq C_{\mu}^{\prime}\,\mu(\Delta_{Q_{i}^{1}})\leq C_{\mu}^{\prime}\,\mu(R^{2,2}(Q^{1}_{i}))

and then μ(R2,1(Qi1))θμ(Q^i1).\mu(R^{2,1}(Q^{1}_{i}))\leq\theta\,\mu(\widehat{Q}^{1}_{i}). Next we set R2,1R^{2,1} and R2,2R^{2,2} as the union of the corresponding R2,1(Qi1)R^{2,1}(Q^{1}_{i}) and R2,2(Qi1)R^{2,2}(Q^{1}_{i}) with Qi1𝔻1,1Q_{i}^{1}\in\mathbb{D}^{1,1}. Then,

μ(R2,1):=μ(Qi1𝔻1,1R2,1(Qi1))=Qi1𝔻1,1μ(R2,1(Qi1))θQi1𝔻1,1μ(Q^i1)=θμ(R1,1)θ2μ(Q¯).\mu(R^{2,1}):=\mu\Big{(}\bigcup_{Q_{i}^{1}\in\mathbb{D}^{1,1}}R^{2,1}(Q^{1}_{i})\Big{)}=\sum_{Q_{i}^{1}\in\mathbb{D}^{1,1}}\mu\big{(}R^{2,1}(Q^{1}_{i})\big{)}\\ \leq\theta\sum_{Q_{i}^{1}\in\mathbb{D}^{1,1}}\mu(\widehat{Q}^{1}_{i})=\theta\,\mu(R^{1,1})\leq\theta^{2}\,\mu(\overline{Q}).

Iterating this procedure we obtain that for every k=0,1,k=0,1,\dots, if τ<2kmΞ1/2\tau<2^{-km}\Xi^{-1}/2 then μ(Rk+1,1)θk+1μ(Q¯)\mu(R^{k+1,1})\leq\theta^{k+1}\mu(\overline{Q}). Let us see that this leads to the desired estimates. Fix τ<Ξ1/2\tau<\Xi^{-1}/2 and find k0k\geq 0 such that 2(k+1)mΞ1/2τ<2kmΞ1/22^{-(k+1)m}\Xi^{-1}/2\leq\tau<2^{-km}\Xi^{-1}/2. By construction ΣτRk+1,1\Sigma_{\tau}\subset R^{k+1,1} and then

μ(Στ)μ(Rk+1,1)θk+1μ(Q¯)(2Ξ)log2θ1mτlog2θ1mμ(Q¯),\mu(\Sigma_{\tau})\leq\mu(R^{k+1,1})\leq\theta^{k+1}\mu(\overline{Q})\leq(2\Xi)^{\frac{\log_{2}\theta^{-1}}{m}}\,\tau^{\frac{\log_{2}\theta^{-1}}{m}}\mu(\overline{Q}),

which easily gives (2.18) with C1=(2Ξ)log2θ1mC_{1}=(2\Xi)^{\frac{\log_{2}\theta^{-1}}{m}} and η=log2θ1m\eta={\frac{\log_{2}\theta^{-1}}{m}}. On the other hand, note that

Qj:2j<Ξ1/2Σ2j,\partial Q\subset\bigcap_{j:2^{-j}<\Xi^{-1}/2}\Sigma_{2^{-j}},

also Σ2(j+1)Σ2j\Sigma_{2^{-(j+1)}}\subset\Sigma_{2^{-j}}. Thus clearly,

0μ(Q)limjμ(Σ2j)limjC12jημ(Q)=0,0\leq\mu(\partial Q)\leq\lim_{j\to\infty}\mu(\Sigma_{2^{-j}})\leq\lim_{j\to\infty}C_{1}2^{-j\eta}\mu(Q)=0,

yielding that μ(Q)=0\mu(\partial Q)=0. ∎

Remark 2.19.

Note that the previous argument is local in the sense that if we just want to obtain the desired estimates for a fixed Q0Q_{0} we would only need to assume that μ\mu is doubling in 2Δ~Q02\widetilde{\Delta}_{Q_{0}}. Indeed we would just need to know that μ(Δ(x,2r))Cμ(Δ(x,r))\mu(\Delta(x,2r))\leq C\,\mu(\Delta(x,r)) for every xQ0x\in Q_{0} and 0<r<Ξ(Q0)0<r<\Xi\ell(Q_{0}), and the involved constants in the resulting estimates will depend only on dimension and CμC_{\mu}. Further details are left to the interested reader.

We next introduce the “discretized Carleson region” relative to QQ, 𝔻Q={Q𝔻:QQ}\mathbb{D}_{Q}=\{Q^{\prime}\in\mathbb{D}:Q^{\prime}\subset Q\}. Let ={Qi}𝔻\mathcal{F}=\{Q_{i}\}\subset\mathbb{D} be a family of pairwise disjoint cubes. The “global discretized sawtooth” relative to \mathcal{F} is the collection of cubes Q𝔻Q\in\mathbb{D} that are not contained in any QiQ_{i}\in\mathcal{F}, that is,

𝔻:=𝔻Qi𝔻Qi.\mathbb{D}_{\mathcal{F}}:=\mathbb{D}\setminus\bigcup_{Q_{i}\in\mathcal{F}}\mathbb{D}_{Q_{i}}.

For a given Q𝔻Q\in\mathbb{D}, the “local discretized sawtooth” relative to \mathcal{F} is the collection of cubes in 𝔻Q\mathbb{D}_{Q} that are not contained in any QiQ_{i}\in\mathcal{F} or, equivalently,

𝔻,Q:=𝔻QQi𝔻Qi=𝔻𝔻Q.\mathbb{D}_{\mathcal{F},Q}:=\mathbb{D}_{Q}\setminus\bigcup_{Q_{i}\in\mathcal{F}}\mathbb{D}_{Q_{i}}=\mathbb{D}_{\mathcal{F}}\cap\mathbb{D}_{Q}.

We also allow \mathcal{F} to be the null set in which case 𝔻Ø=𝔻\mathbb{D}_{\mbox{\tiny{\O}}}=\mathbb{D} and 𝔻Ø,Q=𝔻Q\mathbb{D}_{\mbox{\tiny{\O}},Q}=\mathbb{D}_{Q}.

With a slight abuse of notation, let Q0Q^{0} be either EE, and in that case 𝔻Q0:=𝔻\mathbb{D}_{Q^{0}}:=\mathbb{D}, or a fixed cube in 𝔻\mathbb{D}, hence 𝔻Q0\mathbb{D}_{Q^{0}} is the family of dyadic subcubes of Q0Q^{0}. Let μ\mu be a non-negative Borel measure on Q0Q^{0} so that 0<μ(Q)<0<\mu(Q)<\infty for every Q𝔻Q0Q\in\mathbb{D}_{Q^{0}}. For the rest of the section we will be working with μ\mu which is dyadically doubling in Q0Q^{0}. This means that there exists CμC_{\mu} such that μ(Q)Cμμ(Q)\mu(Q)\leq C_{\mu}\mu(Q^{\prime}) for every Q,Q𝔻Q0Q,Q^{\prime}\in\mathbb{D}_{Q^{0}} with (Q)=2(Q)\ell(Q)=2\ell(Q^{\prime}).

Definition 2.20 (AdyadicA_{\infty}^{\rm dyadic}).

Given Q0Q^{0} and μ\mu, a non-negative dyadically doubling measure in Q0Q^{0}, a non-negative Borel measure ν\nu defined on Q0Q^{0} is said to belong to Adyadic(Q0,μ)A_{\infty}^{\rm dyadic}(Q^{0},\mu) if there exist constants 0<α,β<10<\alpha,\beta<1 such that for every Q𝔻Q0Q\in\mathbb{D}_{Q^{0}} and for every Borel set FQF\subset Q, we have that

(2.21) μ(F)μ(Q)>αν(F)ν(Q)>β.\frac{\mu(F)}{\mu(Q)}>\alpha\qquad\implies\qquad\frac{\nu(F)}{\nu(Q)}>\beta.

It is well known (see [8, 14]) that since μ\mu is a dyadically doubling measure in Q0Q^{0}, νAdyadic(Q0,μ)\nu\in A_{\infty}^{\rm dyadic}(Q^{0},\mu) if and only if νμ\nu\ll\mu in Q0Q^{0} and there exists 1<p<1<p<\infty such that νRHpdyadic(Q0,μ)\nu\in RH_{p}^{\rm dyadic}(Q^{0},\mu), that is, there is a constant C1C\geq 1 such that

(Qk(x)p𝑑μ(x))1pCQk(x)𝑑μ(x)=Cν(Q)μ(Q),\bigg{(}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{Q}k(x)^{p}\,d\mu(x)\bigg{)}^{\frac{1}{p}}\leq C\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{Q}k(x)\,d\mu(x)=C\,\frac{\nu(Q)}{\mu(Q)},

for every Q𝔻Q0Q\in\mathbb{D}_{Q^{0}}, and where k=dν/dμk=d\nu/d\mu is the Radon-Nikodym derivative.

For each ={Qi}𝔻Q0\mathcal{F}=\{Q_{i}\}\subset\mathbb{D}_{Q^{0}}, a family of pairwise disjoint dyadic cubes, and each fLloc1(μ)f\in L^{1}_{\rm loc}(\mu), we define the projection operator

𝒫μf(x)=f(x)𝟏E(QiQi)(x)+Qi(Qif(y)𝑑μ(y))𝟏Qi(x).\mathcal{P}_{\mathcal{F}}^{\mu}f(x)=f(x)\mathbf{1}_{E\setminus(\bigcup_{Q_{i}\in\mathcal{F}}Q_{i})}(x)+\sum_{Q_{i}\in\mathcal{F}}\Big{(}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{Q_{i}}f(y)\,d\mu(y)\Big{)}\mathbf{1}_{Q_{i}}(x).

If ν\nu is a non-negative Borel measure on Q0Q^{0}, we may naturally then define the measure 𝒫μν\mathcal{P}_{\mathcal{F}}^{\mu}\nu as 𝒫μν(F)=E𝒫μ𝟏F𝑑ν\mathcal{P}_{\mathcal{F}}^{\mu}\nu(F)=\int_{E}\mathcal{P}_{\mathcal{F}}^{\mu}\mathbf{1}_{F}\,d\nu, that is,

(2.22) 𝒫μν(F)=ν(FQiQi)+Qiμ(FQi)μ(Qi)ν(Qi),\mathcal{P}_{\mathcal{F}}^{\mu}\nu(F)=\nu\Big{(}F\setminus\bigcup_{Q_{i}\in\mathcal{F}}Q_{i}\Big{)}+\sum_{Q_{i}\in\mathcal{F}}\frac{\mu(F\cap Q_{i})}{\mu(Q_{i})}\nu(Q_{i}),

for each Borel set FQ0F\subset Q^{0}.

2.4. Sawtooth domains

In the sequel, Ωn+1\Omega\subset\mathbb{R}^{n+1}, n2n\geq 2, will be a 1-sided NTA domain satisfying the CDC. Write 𝔻=𝔻(Ω)\mathbb{D}=\mathbb{D}(\partial\Omega) for the dyadic grid obtained from Lemma 2.13 with E=ΩE=\partial\Omega. In Remark 2.56 below we shall show that under the present assumptions one has that diam(Δ)rΔ\operatorname{diam}(\Delta)\approx r_{\Delta} for every surface ball Δ\Delta. In particular diam(Q)(Q)\operatorname{diam}(Q)\approx\ell(Q) for every Q𝔻Q\in\mathbb{D} in view of (2.14). Given Q𝔻Q\in\mathbb{D} we define the “Corkscrew point relative to QQ” as XQ:=XΔQX_{Q}:=X_{\Delta_{Q}}. We note that

δ(XQ)dist(XQ,Q)diam(Q).\delta(X_{Q})\approx\operatorname{dist}(X_{Q},Q)\approx\operatorname{diam}(Q).

As done above, given Q𝔻Q\in\mathbb{D} and \mathcal{F} a possibly empty family of pairwise disjoint dyadic cubes, we can define 𝔻Q\mathbb{D}_{Q}, the “discretized Carleson region”; 𝔻\mathbb{D}_{\mathcal{F}}, the “global discretized sawtooth” relative to \mathcal{F}; and 𝔻,Q\mathbb{D}_{\mathcal{F},Q}, the “local discretized sawtooth” relative to \mathcal{F}. Note that if \mathcal{F} to be the null set in which case 𝔻Ø=𝔻\mathbb{D}_{\mbox{\tiny{\O}}}=\mathbb{D} and 𝔻Ø,Q=𝔻Q\mathbb{D}_{\mbox{\tiny{\O}},Q}=\mathbb{D}_{Q}. We would like to introduce the “geometric” Carleson regions and sawtooths.

Let 𝒲=𝒲(Ω)\mathcal{W}=\mathcal{W}(\Omega) denote a collection of (closed) dyadic Whitney cubes of Ωn+1\Omega\subset\mathbb{R}^{n+1}, so that the cubes in 𝒲\mathcal{W} form a covering of Ω\Omega with non-overlapping interiors, and satisfy

(2.23) 4diam(I)dist(4I,Ω)dist(I,Ω)40diam(I),I𝒲,4\operatorname{diam}(I)\leq\operatorname{dist}(4I,\partial\Omega)\leq\operatorname{dist}(I,\partial\Omega)\leq 40\operatorname{diam}(I),\qquad\forall I\in\mathcal{W},

and

diam(I1)diam(I2), whenever I1 and I2 touch.\operatorname{diam}(I_{1})\approx\operatorname{diam}(I_{2}),\,\text{ whenever }I_{1}\text{ and }I_{2}\text{ touch}.

Let X(I)X(I) denote the center of II, let (I)\ell(I) denote the side length of II, and write k=kIk=k_{I} if (I)=2k\ell(I)=2^{-k}.

Given 0<λ<10<\lambda<1 and I𝒲I\in\mathcal{W} we write I=(1+λ)II^{*}=(1+\lambda)I for the “fattening” of II. By taking λ\lambda small enough, we can arrange matters, so that, first, dist(I,J)dist(I,J)\operatorname{dist}(I^{*},J^{*})\approx\operatorname{dist}(I,J) for every I,J𝒲I,J\in\mathcal{W}. Secondly, II^{*} meets JJ^{*} if and only if I\partial I meets J\partial J (the fattening thus ensures overlap of II^{*} and JJ^{*} for any pair I,J𝒲I,J\in\mathcal{W} whose boundaries touch, so that the Harnack Chain property then holds locally in IJI^{*}\cup J^{*}, with constants depending upon λ\lambda). By picking λ\lambda sufficiently small, say 0<λ<λ00<\lambda<\lambda_{0}, we may also suppose that there is τ(12,1)\tau\in(\frac{1}{2},1) such that for distinct I,J𝒲I,J\in\mathcal{W}, we have that τJI=Ø\tau J\cap I^{*}=\mbox{{\O}}. In what follows we will need to work with dilations I=(1+2λ)II^{**}=(1+2\lambda)I or I=(1+4λ)II^{***}=(1+4\lambda)I, and in order to ensure that the same properties hold we further assume that 0<λ<λ0/40<\lambda<\lambda_{0}/4.

For every Q𝔻Q\in\mathbb{D} we can construct a family 𝒲Q𝒲(Ω)\mathcal{W}_{Q}^{*}\subset\mathcal{W}(\Omega), and define

UQ:=I𝒲QI,U_{Q}:=\bigcup_{I\in\mathcal{W}_{Q}^{*}}I^{*},

satisfying the following properties: XQUQX_{Q}\in U_{Q} and there are uniform constants kk^{*} and K0K_{0} such that

(2.24) k(Q)kkIk(Q)+k,I𝒲Q,X(I)UQXQ,I𝒲Q,dist(I,Q)K02k(Q),I𝒲Q.\displaystyle\begin{split}k(Q)-k^{*}\leq k_{I}\leq k(Q)+k^{*},\quad\forall I\in\mathcal{W}_{Q}^{*},\\[4.0pt] X(I)\rightarrow_{U_{Q}}X_{Q},\quad\forall I\in\mathcal{W}_{Q}^{*},\\[4.0pt] \operatorname{dist}(I,Q)\leq K_{0}2^{-k(Q)},\quad\forall I\in\mathcal{W}_{Q}^{*}.\end{split}

Here, X(I)UQXQX(I)\rightarrow_{U_{Q}}X_{Q} means that the interior of UQU_{Q} contains all balls in a Harnack Chain (in Ω\Omega) connecting X(I)X(I) to XQX_{Q}, and moreover, for any point ZZ contained in any ball in the Harnack Chain, we have dist(Z,Ω)dist(Z,ΩUQ)\operatorname{dist}(Z,\partial\Omega)\approx\operatorname{dist}(Z,\Omega\setminus U_{Q}) with uniform control of the implicit constants. The constants k,K0k^{*},K_{0} and the implicit constants in the condition X(I)UQXQX(I)\rightarrow_{U_{Q}}X_{Q}, depend on the allowable parameters and on λ\lambda. Moreover, given I𝒲(Ω)I\in\mathcal{W}(\Omega) we have that I𝒲QII\in\mathcal{W}_{Q_{I}}^{*}, where QI𝔻Q_{I}\in\mathbb{D} satisfies (QI)=(I)\ell(Q_{I})=\ell(I), and contains any fixed y^Ω\widehat{y}\in\partial\Omega such that dist(I,Ω)=dist(I,y^)\operatorname{dist}(I,\partial\Omega)=\operatorname{dist}(I,\widehat{y}). The reader is referred to [17, 20] for full details.

For a given Q𝔻Q\in\mathbb{D}, the “Carleson box” relative to QQ is defined by

TQ:=int(Q𝔻QUQ).T_{Q}:=\operatorname{int}\bigg{(}\bigcup_{Q^{\prime}\in\mathbb{D}_{Q}}U_{Q^{\prime}}\bigg{)}.

For a given family ={Qi}𝔻\mathcal{F}=\{Q_{i}\}\subset\mathbb{D} of pairwise disjoint cubes and a given Q𝔻Q\in\mathbb{D}, we define the “local sawtooth region” relative to \mathcal{F} by

(2.25) Ω,Q=int(Q𝔻,QUQ)=int(I𝒲,QI),\Omega_{\mathcal{F},Q}=\operatorname{int}\bigg{(}\bigcup_{Q^{\prime}\in\mathbb{D}_{\mathcal{F},Q}}U_{Q^{\prime}}\bigg{)}=\operatorname{int}\bigg{(}\bigcup_{I\in\mathcal{W}_{\mathcal{F},Q}}I^{*}\bigg{)},

where 𝒲,Q:=Q𝔻,Q𝒲Q\mathcal{W}_{\mathcal{F},Q}:=\bigcup_{Q^{\prime}\in\mathbb{D}_{\mathcal{F},Q}}\mathcal{W}_{Q}^{*}. Note that in the previous definition we may allow \mathcal{F} to be empty in which case clearly ΩØ,Q=TQ\Omega_{\mbox{\tiny{\O}},Q}=T_{Q}. Similarly, the “global sawtooth region” relative to \mathcal{F} is defined as

(2.26) Ω=int(Q𝔻UQ)=int(I𝒲I),\Omega_{\mathcal{F}}=\operatorname{int}\bigg{(}\bigcup_{Q^{\prime}\in\mathbb{D}_{\mathcal{F}}}U_{Q^{\prime}}\bigg{)}=\operatorname{int}\bigg{(}\bigcup_{I\in\mathcal{W}_{\mathcal{F}}}I^{*}\bigg{)},

where 𝒲:=Q𝔻𝒲Q\mathcal{W}_{\mathcal{F}}:=\bigcup_{Q^{\prime}\in\mathbb{D}_{\mathcal{F}}}\mathcal{W}_{Q}^{*}. If \mathcal{F} is the empty set clearly ΩØ=Ω\Omega_{\mbox{\tiny{\O}}}=\Omega. For a given Q𝔻Q\in\mathbb{D} and xΩx\in\partial\Omega let us introduce the “truncated dyadic cone”

ΓQ(x):=xQ𝔻QUQ,\Gamma_{Q}(x):=\bigcup_{x\in Q^{\prime}\in\mathbb{D}_{Q}}U_{Q^{\prime}},

where it is understood that ΓQ(x)=Ø\Gamma_{Q}(x)=\mbox{{\O}} if xQx\notin Q. Analogously, we can slightly fatten the Whitney boxes and use II^{**} to define new fattened Whitney regions and sawtooth domains. More precisely, for every Q𝔻Q\in\mathbb{D},

TQ:=int(Q𝔻QUQ),Ω,Q:=int(Q𝔻,QUQ),ΓQ(x):=xQ𝔻Q0UQT_{Q}^{*}:=\operatorname{int}\bigg{(}\bigcup_{Q^{\prime}\in\mathbb{D}_{Q}}U_{Q^{\prime}}^{*}\bigg{)},\quad\Omega^{*}_{\mathcal{F},Q}:=\operatorname{int}\bigg{(}\bigcup_{Q^{\prime}\in\mathbb{D}_{\mathcal{F},Q}}U_{Q^{\prime}}^{*}\bigg{)},\quad\Gamma^{*}_{Q}(x):=\bigcup_{x\in Q^{\prime}\in\mathbb{D}_{Q_{0}}}U_{Q^{\prime}}^{*}

where UQ:=I𝒲QIU_{Q}^{*}:=\bigcup_{I\in\mathcal{W}_{Q}^{*}}I^{**}. Similarly, we can define TQT_{Q}^{**}, Ω,Q\Omega^{**}_{\mathcal{F},Q}, ΓQ(x)\Gamma_{Q}^{**}(x), and UQU^{**}_{Q} by using II^{***} in place of II^{**}.

Given QQ we next define the “localized dyadic non-tangential maximal function”

(2.27) 𝒩Qu(x):=supYΓQ(x)|u(Y)|,xΩ,\mathcal{N}_{Q}u(x):=\sup_{Y\in\Gamma^{*}_{Q}(x)}|u(Y)|,\qquad x\in\partial\Omega,

for every uC(TQ)u\in C(T_{Q}^{*}), where it is understood that 𝒩Qu(x)=0\mathcal{N}_{Q}u(x)=0 for every xΩQx\in\partial\Omega\setminus Q (since ΓQ(x)=Ø\Gamma_{Q}^{*}(x)=\mbox{{\O}} in such a case). Finally, let us introduce the “localized dyadic conical square function”

(2.28) 𝒮Qu(x):=(ΓQ(x)|u(Y)|2δ(Y)1n𝑑Y)12,xΩ,\mathcal{S}_{Q}u(x):=\bigg{(}\iint_{\Gamma_{Q}(x)}|\nabla u(Y)|^{2}\delta(Y)^{1-n}\,dY\bigg{)}^{\frac{1}{2}},\qquad x\in\partial\Omega,

for every uWloc1,2(TQ0)u\in W^{1,2}_{\rm loc}(T_{Q_{0}}). Note that again 𝒮Qu(x)=0\mathcal{S}_{Q}u(x)=0 for every xΩQx\in\partial\Omega\setminus Q.

To define the “Carleson box” TΔT_{\Delta} associated with a surface ball Δ=Δ(x,r)\Delta=\Delta(x,r), let k(Δ)k(\Delta) denote the unique kk\in\mathbb{Z} such that 2k1<200r2k2^{-k-1}<200r\leq 2^{-k}, and set

(2.29) 𝔻Δ:={Q𝔻k(Δ):Q2ΔØ}.\mathbb{D}^{\Delta}:=\big{\{}Q\in\mathbb{D}_{k(\Delta)}:\>Q\cap 2\Delta\neq\mbox{{\O}}\big{\}}.

We then define

(2.30) TΔ:=int(Q𝔻ΔTQ¯).T_{\Delta}:=\operatorname{int}\bigg{(}\bigcup_{Q\in\mathbb{D}^{\Delta}}\overline{T_{Q}}\bigg{)}.

We can also consider fattened versions of TΔT_{\Delta} given by

TΔ:=int(Q𝔻ΔTQ¯),TΔ:=int(Q𝔻ΔTQ¯).T_{\Delta}^{*}:=\operatorname{int}\bigg{(}\bigcup_{Q\in\mathbb{D}^{\Delta}}\overline{T_{Q}^{*}}\bigg{)},\qquad T_{\Delta}^{**}:=\operatorname{int}\bigg{(}\bigcup_{Q\in\mathbb{D}^{\Delta}}\overline{T_{Q}^{**}}\bigg{)}.

Following [17, 20], one can easily see that there exist constants 0<κ1<10<\kappa_{1}<1 and κ016Ξ\kappa_{0}\geq 16\Xi (with Ξ\Xi the constant in (2.14)), depending only on the allowable parameters, so that

(2.31) κ1BQΩTQTQTQTQ¯κ0BQΩ¯=:12BQΩ¯,\displaystyle\kappa_{1}B_{Q}\cap\Omega\subset T_{Q}\subset T_{Q}^{*}\subset T_{Q}^{**}\subset\overline{T_{Q}^{**}}\subset\kappa_{0}B_{Q}\cap\overline{\Omega}=:\tfrac{1}{2}B_{Q}^{*}\cap\overline{\Omega},
(2.32) 54BΔΩTΔTΔTΔTΔ¯κ0BΔΩ¯=:12BΔΩ¯,\displaystyle\tfrac{5}{4}B_{\Delta}\cap\Omega\subset T_{\Delta}\subset T_{\Delta}^{*}\subset T_{\Delta}^{**}\subset\overline{T_{\Delta}^{**}}\subset\kappa_{0}B_{\Delta}\cap\overline{\Omega}=:\tfrac{1}{2}B_{\Delta}^{*}\cap\overline{\Omega},

and also

(2.33) Qκ0BΔΩ=12BΔΩ=:12Δ,Q𝔻Δ,Q\subset\kappa_{0}B_{\Delta}\cap\partial\Omega=\tfrac{1}{2}B_{\Delta}^{*}\cap\partial\Omega=:\tfrac{1}{2}\Delta^{*},\qquad\forall\,Q\in\mathbb{D}^{\Delta},

where BQB_{Q} is defined as in (2.15), Δ=Δ(x,r)\Delta=\Delta(x,r) with xΩx\in\partial\Omega, 0<r<diam(Ω)0<r<\operatorname{diam}(\partial\Omega), and BΔ=B(x,r)B_{\Delta}=B(x,r) is so that Δ=BΔΩ\Delta=B_{\Delta}\cap\partial\Omega. From our choice of the parameters one also has that BQBQB_{Q}^{*}\subset B_{Q^{\prime}}^{*} whenever QQQ\subset Q^{\prime}.

In the remainder of this section we show that if Ω\Omega is a 1-sided NTA domain satisfying the CDC then Carleson boxes and local and global sawtooth domains are also 1-sided NTA domains satisfying the CDC. We next present some of the properties of the capacity which will be used in our proofs. From the definition of capacity one can easily see that given a ball BB and compact sets F1F2B¯F_{1}\subset F_{2}\subset\overline{B} then

(2.34) Cap2(F1,2B)Cap2(F2,2B).\mathop{\operatorname{Cap}_{2}}\nolimits(F_{1},2B)\leq\mathop{\operatorname{Cap}_{2}}\nolimits(F_{2},2B).

Also, given two balls B1B2B_{1}\subset B_{2} and a compact set FB1¯F\subset\overline{B_{1}} then

(2.35) Cap2(F,2B2)Cap2(F,2B1).\mathop{\operatorname{Cap}_{2}}\nolimits(F,2B_{2})\leq\mathop{\operatorname{Cap}_{2}}\nolimits(F,2B_{1}).

On the other hand, [15, Lemma 2.16] gives that if FF is a compact with FB¯F\subset\overline{B} then there is a dimensional constant CnC_{n} such that

(2.36) Cn1Cap2(F,2B)Cap2(F,4B)Cap2(F,2B).C_{n}^{-1}\mathop{\operatorname{Cap}_{2}}\nolimits(F,2B)\leq\mathop{\operatorname{Cap}_{2}}\nolimits(F,4B)\leq\mathop{\operatorname{Cap}_{2}}\nolimits(F,2B).
Proposition 2.37.

Let Ωn+1\Omega\subset\mathbb{R}^{n+1}, n2n\geq 2, be a 1-sided NTA domain satisfying the CDC. Then all of its Carleson boxes TQT_{Q} and TΔT_{\Delta}, and sawtooth regions Ω\Omega_{\mathcal{F}}, and Ω,Q\Omega_{\mathcal{F},Q} are 1-sided NTA domains and satisfy the CDC with uniform implicit constants depending only on dimension and on the corresponding constants for Ω\Omega.

Proof.

A careful examination of the proofs in [17, Appendices A.1-A.2] reveals that if Ω\Omega is a 1-sided NTA domain then all Carleson boxes TQT_{Q} and TΔT_{\Delta}, and local and global sawtooth domains Ω,Q\Omega_{\mathcal{F},Q} and Ω\Omega_{\mathcal{F}} inherit the interior Corkscrew and Harnack chain conditions, hence they are also 1-sided NTA domains. Therefore, we only need to prove the CDC. We are going to consider only the case Ω,Q\Omega_{\mathcal{F},Q} (which in particular gives the desired property for TQT_{Q} by allowing \mathcal{F} to be the null set). The other proofs require minimal changes which are left to the interested reader. To this end, fix Q𝔻Q\in\mathbb{D} and 𝔻Q\mathcal{F}\subset\mathbb{D}_{Q} a (possibly empty) family of pairwise disjoint dyadic cubes. Let xΩ,Qx\in\partial\Omega_{\mathcal{F},Q} and 0<r<diam(Ω,Q)(Q)0<r<\operatorname{diam}(\Omega_{\mathcal{F},Q})\approx\ell(Q).

Case 1: δ(x)=0\delta(x)=0. In that case we have that xΩx\in\partial\Omega and we can use that Ω\Omega satisfies the CDC with constant c1c_{1}, (2.34) and the fact that Ω,QΩ\Omega_{\mathcal{F},Q}\subset\Omega to obtain the desired estimate

c1rn1Cap2(B(x,r)¯Ω,B(x,2r))Cap2(B(x,r)¯Ω,Q,B(x,2r)).c_{1}r^{n-1}\lesssim\mathop{\operatorname{Cap}_{2}}\nolimits(\overline{B(x,r)}\setminus\Omega,B(x,2r))\leq\mathop{\operatorname{Cap}_{2}}\nolimits(\overline{B(x,r)}\setminus\Omega_{\mathcal{F},Q},B(x,2r)).

Case 2: 0<δ(x)<r/M0<\delta(x)<r/M with MM large enough to be chosen. In this case xΩΩ,Qx\in\Omega\cap\partial\Omega_{\mathcal{F},Q} and hence there exist Q𝔻,QQ^{\prime}\in\mathbb{D}_{\mathcal{F},Q} and I𝒲QI\in\mathcal{W}^{*}_{Q^{\prime}} such that xIx\in\partial I^{*}. Note that by (2.24)

|xxQ|diam(I)+dist(I,Q)+diam(Q)(Q)(I)δ(x)rM.|x-x_{Q^{\prime}}|\leq\operatorname{diam}(I^{*})+\operatorname{dist}(I,Q^{\prime})+\operatorname{diam}(Q^{\prime})\lesssim\ell(Q^{\prime})\approx\ell(I)\approx\delta(x)\lesssim\frac{r}{M}.

Let Q′′𝔻QQ^{\prime\prime}\in\mathbb{D}_{Q} be such that xQQ′′x_{Q^{\prime}}\in Q^{\prime\prime} and r2M(Q′′)<rM<(Q)\frac{r}{2M}\leq\ell(Q^{\prime\prime})<\frac{r}{M}<\ell(Q) provided that MM is taken large enough. If ZBQ′′Z\in B_{Q^{\prime\prime}} then taking MM large enough

|Zx||ZxQ′′|+|xQ′′xQ|+|xQx|(Q′′)+rMrM<r|Z-x|\leq|Z-x_{Q^{\prime\prime}}|+|x_{Q^{\prime\prime}}-x_{Q^{\prime}}|+|x_{Q^{\prime}}-x|\lesssim\ell(Q^{\prime\prime})+\frac{r}{M}\lesssim\frac{r}{M}<r

and BQ′′B(x,r)B_{Q^{\prime\prime}}\subset B(x,r). On the other hand, if ZB(x,2r)Z\in B(x,2r), we analogously have provided MM is large enough

|ZxQ′′||Zx|+|xxQ|+|xQxQ′′|<2r+CrM+ΞrQ′′<6MΞrQ′′|Z-x_{Q^{\prime\prime}}|\leq|Z-x|+|x-x_{Q^{\prime}}|+|x_{Q^{\prime}}-x_{Q^{\prime\prime}}|<2r+C\frac{r}{M}+\Xi r_{Q^{\prime\prime}}<6M\Xi r_{Q^{\prime\prime}}

and thus B(x,2r)6MΞBQ′′B(x,2r)\subset 6M\Xi B_{Q^{\prime\prime}}. Once MM has been fixed so that the previous estimates hold, we use them in conjunction with the fact that Ω\Omega satisfies the CDC with constant c1c_{1}, (2.34)–(2.36), and that Ω,QΩ\Omega_{\mathcal{F},Q}\subset\Omega to obtain

c1(2MΞ)n1rn1c1rQ′′n1Cap2(BQ′′¯Ω,2BQ′′)Cap2(BQ′′¯Ω,6MΞBQ′′)Cap2(BQ′′¯Ω,B(x,2r))Cap2(B(x,r)¯Ω,Q,B(x,2r)),\frac{c_{1}}{(2M\Xi)^{n-1}}r^{n-1}\leq c_{1}r_{Q^{\prime\prime}}^{n-1}\lesssim\mathop{\operatorname{Cap}_{2}}\nolimits(\overline{B_{Q^{\prime\prime}}}\setminus\Omega,2B_{Q^{\prime\prime}})\lesssim\mathop{\operatorname{Cap}_{2}}\nolimits(\overline{B_{Q^{\prime\prime}}}\setminus\Omega,6M\Xi B_{Q^{\prime\prime}})\\ \leq\mathop{\operatorname{Cap}_{2}}\nolimits(\overline{B_{Q^{\prime\prime}}}\setminus\Omega,B(x,2r))\leq\mathop{\operatorname{Cap}_{2}}\nolimits(\overline{B(x,r)}\setminus\Omega_{\mathcal{F},Q},B(x,2r)),

which gives us the desired lower bound in the present case.

Ω\partial\OmegaQQTQT_{Q}xxΩ\partial\OmegaQQTQT_{Q}xxBQ′′B_{Q^{\prime\prime}}
Figure 1. Case 1 and Case 2 for TQT_{Q}.

Case 3: δ(x)>r/M\delta(x)>r/M. In this case xΩΩ,Qx\in\Omega\cap\partial\Omega_{\mathcal{F},Q} and hence there exists Q𝔻,QQ^{\prime}\in\mathbb{D}_{\mathcal{F},Q} and I𝒲QI\in\mathcal{W}^{*}_{Q^{\prime}} such that xIx\in\partial I^{*} and int(I)Ω,Q\operatorname{int}(I^{*})\subset\Omega_{\mathcal{F},Q}. Also there exists J𝒲J\in\mathcal{W}, with JxJ\ni x such that J𝒲Q′′J\notin\mathcal{W}^{*}_{Q^{\prime\prime}} for any Q′′𝔻,QQ^{\prime\prime}\in\mathbb{D}_{\mathcal{F},Q} which implies that τJΩΩ,Q\tau J\subset\Omega\setminus\Omega_{\mathcal{F},Q} for some τ(12,1)\tau\in(\frac{1}{2},1) (see Section 2.4). Note that (I)(J)δ(x)r\ell(I)\approx\ell(J)\approx\delta(x)\gtrsim r, and more precisely r/M<δ(x)<41diam(J)r/M<\delta(x)<41\operatorname{diam}(J) by (2.23).

Ω\partial\OmegaQQTQT_{Q}BB^{\prime}xx
Figure 2. Case 3 for TQT_{Q}.

Let B=B(x,s)B^{\prime}=B(x^{\prime},s) with s=r/(300M)s=r/(300M) and xx^{\prime} being the point in the segment joining xx and the center of JJ at distance 2s2s from xx. It is easy to see that BB(x,r)B(x,2r)1000MBB^{\prime}\subset B(x,r)\subset B(x,2r)\subset 1000MB^{\prime} and also B¯int(J)Ω,Q\overline{B^{\prime}}\subset\operatorname{int}(J)\setminus\Omega_{\mathcal{F},Q}. We can then use (2.12) and (2.34)–(2.36) to obtain the desired estimate:

1(300M)n1rn1=sn1Cap2(B¯,2B)Cap2(B¯,1000MB)Cap2(B¯,B(x,2r))Cap2(B(x,r)¯Ω,Q,B(x,2r)).\frac{1}{(300M)^{n-1}}r^{n-1}=s^{n-1}\approx\mathop{\operatorname{Cap}_{2}}\nolimits(\overline{B^{\prime}},2B^{\prime})\lesssim\mathop{\operatorname{Cap}_{2}}\nolimits(\overline{B^{\prime}},1000MB^{\prime})\\ \leq\mathop{\operatorname{Cap}_{2}}\nolimits(\overline{B^{\prime}},B(x,2r))\leq\mathop{\operatorname{Cap}_{2}}\nolimits(\overline{B(x,r)}\setminus\Omega_{\mathcal{F},Q},B(x,2r)).

Collecting the 3 cases and using (2.12) we have been able to show that

(2.38) Cap2(B(x,r)¯Ω,Q,B(x,2r))Cap2(B(x,r)¯,B(x,2r))1,xΩ,Q, 0<r<diam(Ω,Q),\frac{\mathop{\operatorname{Cap}_{2}}\nolimits(\overline{B(x,r)}\setminus\Omega_{\mathcal{F},Q},B(x,2r))}{\mathop{\operatorname{Cap}_{2}}\nolimits(\overline{B(x,r)},B(x,2r))}\gtrsim 1,\qquad\forall\,x\in\partial\Omega_{\mathcal{F},Q},\ 0<r<\operatorname{diam}(\Omega_{\mathcal{F},Q}),

which eventually gives that Ω,Q\Omega_{\mathcal{F},Q} satisfies the CDC. This completes the proof. ∎

Our next auxiliary result adapts [21, Lemma 4.44] to our current setting:

Lemma 2.39.

Let Ωn+1\Omega\subset\mathbb{R}^{n+1} be a 1-sided NTA domain satisfying the CDC. Given Q0𝔻Q_{0}\in\mathbb{D} and N4N\geq 4 consider the family of pairwise disjoint cubes N={Q𝔻Q0:(Q)=2N(Q0)}\mathcal{F}_{N}=\{Q\in\mathbb{D}_{Q_{0}}:\ell(Q)=2^{-N}\,\ell(Q_{0})\} and let ΩN:=ΩN,Q0\Omega_{N}:=\Omega_{\mathcal{F}_{N},Q_{0}} and ΩN:=ΩN,Q0\Omega_{N}^{*}:=\Omega_{\mathcal{F}_{N},Q_{0}}^{*}. There exists ΨNCc(n+1)\Psi_{N}\in C_{c}^{\infty}(\mathbb{R}^{n+1}) and a constant C1C\geq 1 depending only on dimension nn, the 1-sided NTA constants, the CDC constant, and independent of NN and Q0Q_{0} such that the following hold:

  • (i)(i)

    C1 1ΩNΨN𝟏ΩNC^{-1}\,\mathbf{1}_{\Omega_{N}}\leq\Psi_{N}\leq\mathbf{1}_{\Omega_{N}^{*}}.

  • (ii)(ii)

    supXΩ|ΨN(X)|δ(X)C\sup_{X\in\Omega}|\nabla\Psi_{N}(X)|\,\delta(X)\leq C.

  • (iii)(iii)

    Setting

    (2.40) 𝒲N:=Q𝔻N,Q0𝒲Q,𝒲NΣ:={I𝒲N:J𝒲𝒲NwithIJØ}.\mathcal{W}_{N}:=\bigcup_{Q\in\mathbb{D}_{\mathcal{F}_{N},Q_{0}}}\mathcal{W}_{Q}^{*},\quad\mathcal{W}_{N}^{\Sigma}:=\big{\{}I\in\mathcal{W}_{N}:\,\exists\,J\in\mathcal{W}\setminus\mathcal{W}_{N}\ \mbox{with}\ \partial I\cap\partial J\neq\mbox{{\O}}\big{\}}.

one has

(2.41) ΨN0inI𝒲N𝒲NΣI\nabla\Psi_{N}\equiv 0\quad\mbox{in}\quad\bigcup_{I\in\mathcal{W}_{N}\setminus\mathcal{W}_{N}^{\Sigma}}I^{**}

and there exists a family {Q^I}I𝒲NΣ\{\widehat{Q}_{I}\}_{I\in\mathcal{W}_{N}^{\Sigma}} so that

(2.42) C1(I)(Q^I)C(I),dist(I,Q^I)C(I),I𝒲NΣ𝟏Q^IC.C^{-1}\,\ell(I)\leq\ell(\widehat{Q}_{I})\leq C\,\ell(I),\qquad\operatorname{dist}(I,\widehat{Q}_{I})\leq C\,\ell(I),\qquad\sum_{I\in\mathcal{W}_{N}^{\Sigma}}\mathbf{1}_{\widehat{Q}_{I}}\leq C.
Proof.

We proceed as in [21, Lemma 4.44]. Recall that given II, any closed dyadic cube in n+1\mathbb{R}^{n+1}, we set I=(1+λ)II^{*}=(1+\lambda)I and I=(1+2λ)II^{**}=(1+2\,\lambda)I. Let us introduce I~=(1+32λ)I\widetilde{I^{*}}=(1+\frac{3}{2}\,\lambda)I so that

(2.43) Iint(I~)I~int(I).I^{*}\subsetneq\operatorname{int}(\widetilde{I^{*}})\subsetneq\widetilde{I^{*}}\subset\operatorname{int}(I^{**}).

Given I0:=[12,12]n+1n+1I_{0}:=[-\frac{1}{2},\frac{1}{2}]^{n+1}\subset\mathbb{R}^{n+1}, fix ϕ0Cc(n+1)\phi_{0}\in C_{c}^{\infty}(\mathbb{R}^{n+1}) such that 1I0ϕ01I0~1_{I_{0}^{*}}\leq\phi_{0}\leq 1_{\widetilde{I_{0}^{*}}} and |ϕ0|1|\nabla\phi_{0}|\lesssim 1 (the implicit constant depends on the parameter λ\lambda). For every I𝒲=𝒲(Ω)I\in\mathcal{W}=\mathcal{W}(\Omega) we set ϕI()=ϕ0(X(I)(I))\phi_{I}(\cdot)=\phi_{0}\big{(}\frac{\,\cdot\,-X(I)}{\ell(I)}\big{)} so that ϕIC(n+1)\phi_{I}\in C^{\infty}(\mathbb{R}^{n+1}), 1IϕI1I~1_{I^{*}}\leq\phi_{I}\leq 1_{\widetilde{I^{*}}} and |ϕI|(I)1|\nabla\phi_{I}|\lesssim\ell(I)^{-1} (with implicit constant depending only on nn and λ\lambda).

For every XΩX\in\Omega, we let Φ(X):=I𝒲ϕI(X)\Phi(X):=\sum_{I\in\mathcal{W}}\phi_{I}(X). It then follows that ΦCloc(Ω)\Phi\in C_{\rm loc}^{\infty}(\Omega) since for every compact subset of Ω\Omega, the previous sum has finitely many non-vanishing terms. Also, 1Φ(X)Cλ1\leq\Phi(X)\leq C_{\lambda} for every XΩX\in\Omega since the family {I~}I𝒲\{\widetilde{I^{*}}\}_{I\in\mathcal{W}} has bounded overlap by our choice of λ\lambda. Hence we can set ΦI=ϕI/Φ\Phi_{I}=\phi_{I}/\Phi and one can easily see that ΦICc(n+1)\Phi_{I}\in C_{c}^{\infty}(\mathbb{R}^{n+1}), Cλ11IΦI1I~C_{\lambda}^{-1}1_{I^{*}}\leq\Phi_{I}\leq 1_{\widetilde{I^{*}}} and |ΦI|(I)1|\nabla\Phi_{I}|\lesssim\ell(I)^{-1}. With this in hand set

ΨN(X):=I𝒲NΦI(X)=I𝒲NϕI(X)I𝒲ϕI(X),XΩ.\Psi_{N}(X):=\sum_{I\in\mathcal{W}_{N}}\Phi_{I}(X)=\frac{\sum\limits_{I\in\mathcal{W}_{N}}\phi_{I}(X)}{\sum\limits_{I\in\mathcal{W}}\phi_{I}(X)},\qquad X\in\Omega.

We first note that the number of terms in the sum defining ΨN\Psi_{N} is bounded depending on NN. Indeed, if Q𝔻N,Q0Q\in\mathbb{D}_{\mathcal{F}_{N},Q_{0}} then Q𝔻Q0Q\in\mathbb{D}_{Q_{0}} and 2N(Q0)<(Q)(Q0)2^{-N}\ell(Q_{0})<\ell(Q)\leq\ell(Q_{0}) which implies that 𝔻N,Q0\mathbb{D}_{\mathcal{F}_{N},Q_{0}} has finite cardinality with bounds depending only on dimension and NN (here we recall that the number of dyadic children of a given cube is uniformly controlled). Also, by construction 𝒲Q\mathcal{W}_{Q}^{*} has cardinality depending only on the allowable parameters. Hence, #𝒲NCN<\#\mathcal{W}_{N}\lesssim C_{N}<\infty. This and the fact that each ΦICc(n+1)\Phi_{I}\in C_{c}^{\infty}(\mathbb{R}^{n+1}) yield that ΨNCc(n+1)\Psi_{N}\in C_{c}^{\infty}(\mathbb{R}^{n+1}). Note also that (2.43) and the definition of 𝒲N\mathcal{W}_{N} give

suppΨII𝒲NI~=Q𝔻N,Q0I𝒲QI~int(Q𝔻N,Q0I𝒲QI)=int(Q𝔻N,Q0UQ)=ΩN\operatorname{supp}\Psi_{I}\subset\bigcup_{I\in\mathcal{W}_{N}}\widetilde{I^{*}}=\bigcup_{Q\in\mathbb{D}_{\mathcal{F}_{N},{Q}_{0}}}\bigcup_{I\in\mathcal{W}_{Q}^{*}}\widetilde{I^{*}}\subset\operatorname{int}\Big{(}\bigcup_{Q\in\mathbb{D}_{\mathcal{F}_{N},{Q}_{0}}}\bigcup_{I\in\mathcal{W}_{Q}^{*}}I^{**}\Big{)}\\ =\operatorname{int}\Big{(}\bigcup_{Q\in\mathbb{D}_{\mathcal{F}_{N},{Q}_{0}}}U_{Q}^{*}\Big{)}=\Omega_{N}^{*}

This, the fact that 𝒲N𝒲\mathcal{W}_{N}\subset\mathcal{W}, and the definition of ΨN\Psi_{N} immediately give that ΨN𝟏ΩN\Psi_{N}\leq\mathbf{1}_{\Omega_{N}^{*}}. On the other hand if XΩN=ΩN,Q0X\in\Omega_{N}=\Omega_{\mathcal{F}_{N},Q_{0}} then the exists I𝒲NI\in\mathcal{W}_{N} such that XIX\in I^{*} in which case ΨN(X)ΦI(X)Cλ1\Psi_{N}(X)\geq\Phi_{I}(X)\geq C_{\lambda}^{-1}. All these imply (i)(i). Note that (ii)(ii) follows by observing that for every XΩX\in\Omega

|ΨN(X)|I𝒲N|ΦI(X)|I𝒲(I)1 1I~(X)δ(X)1|\nabla\Psi_{N}(X)|\leq\sum_{I\in\mathcal{W}_{N}}|\nabla\Phi_{I}(X)|\lesssim\sum_{I\in\mathcal{W}}\ell(I)^{-1}\,1_{\widetilde{I^{*}}}(X)\lesssim\delta(X)^{-1}

where we have used that if XI~X\in\widetilde{I^{*}} then δ(X)(I)\delta(X)\approx\ell(I) and also that the family {I~}I𝒲\{\widetilde{I^{*}}\}_{I\in\mathcal{W}} has bounded overlap.

To see (iii)(iii) fix I𝒲N𝒲NΣI\in\mathcal{W}_{N}\setminus\mathcal{W}^{\Sigma}_{N} and XIX\in I^{**}, and set 𝒲X:={J𝒲:ϕJ(X)0}\mathcal{W}_{X}:=\{J\in\mathcal{W}:\phi_{J}(X)\neq 0\} so that I𝒲XI\in\mathcal{W}_{X}. We first note that 𝒲X𝒲N\mathcal{W}_{X}\subset\mathcal{W}_{N}. Indeed, if ϕJ(X)0\phi_{J}(X)\neq 0 then XJ~X\in\widetilde{J^{*}}. Hence XIJX\in I^{**}\cap J^{**} and our choice of λ\lambda gives that I\partial I meets J\partial J, this in turn implies that J𝒲NJ\in\mathcal{W}_{N} since I𝒲N𝒲NΣI\in\mathcal{W}_{N}\setminus\mathcal{W}^{\Sigma}_{N}. All these yield

ΨN(X)=J𝒲NϕJ(X)J𝒲ϕJ(X)=J𝒲N𝒲XϕJ(X)J𝒲XϕJ(X)=J𝒲N𝒲XϕJ(X)J𝒲N𝒲XϕJ(X)=1.\Psi_{N}(X)=\frac{\sum\limits_{J\in\mathcal{W}_{N}}\phi_{J}(X)}{\sum\limits_{J\in\mathcal{W}}\phi_{J}(X)}=\frac{\sum\limits_{J\in\mathcal{W}_{N}\cap\mathcal{W}_{X}}\phi_{J}(X)}{\sum\limits_{J\in\mathcal{W}_{X}}\phi_{J}(X)}=\frac{\sum\limits_{J\in\mathcal{W}_{N}\cap\mathcal{W}_{X}}\phi_{J}(X)}{\sum\limits_{J\in\mathcal{W}_{N}\cap\mathcal{W}_{X}}\phi_{J}(X)}=1.

Hence ΨN|I1\Psi_{N}\big{|}_{I^{**}}\equiv 1 for every I𝒲N𝒲NΣI\in\mathcal{W}_{N}\setminus\mathcal{W}^{\Sigma}_{N}. This and the fact that ΨNCc(n+1)\Psi_{N}\in C_{c}^{\infty}(\mathbb{R}^{n+1}) immediately give that ΨN0\nabla\Psi_{N}\equiv 0 in I𝒲N𝒲NΣI\bigcup_{I\in\mathcal{W}_{N}\setminus\mathcal{W}_{N}^{\Sigma}}I^{**}.

We are left with showing the last part of (iv)(iv) and for that we borrow some ideas from [18, Appendix A.2]. Fix I𝒲NΣI\in\mathcal{W}_{N}^{\Sigma} and let JJ be so that J𝒲𝒲NJ\in\mathcal{W}\setminus\mathcal{W}_{N} with IJØ\partial I\cap\partial J\neq\mbox{{\O}}, in particular (I)(J)\ell(I)\approx\ell(J). Since I𝒲NΣI\in\mathcal{W}_{N}^{\Sigma} there exists QI𝔻N,Q0Q_{I}\in\mathbb{D}_{\mathcal{F}_{N},Q_{0}} (that is, QIQ0Q_{I}\subset Q_{0} with 2N(Q0)<(QI)(Q0)2^{-N}\,\ell(Q_{0})<\ell(Q_{I})\leq\ell(Q_{0}) so that I𝒲QII\in\mathcal{W}_{Q_{I}}^{*}). Pick QJ𝔻Q_{J}\in\mathbb{D} so that (QJ)=(J)\ell(Q_{J})=\ell(J) and it contains any fixed y^Ω\widehat{y}\in\partial\Omega such that dist(J,Ω)=dist(J,y^)\operatorname{dist}(J,\partial\Omega)=\operatorname{dist}(J,\widehat{y}). Then, as observed in Section 2.4, one has J𝒲QJJ\in\mathcal{W}_{Q_{J}}^{*}. But, since J𝒲𝒲NJ\in\mathcal{W}\setminus\mathcal{W}_{N}, we necessarily have QJ𝔻N,Q0=𝔻N𝔻Q0Q_{J}\notin\mathbb{D}_{\mathcal{F}_{N},Q_{0}}=\mathbb{D}_{\mathcal{F}_{N}}\cap\mathbb{D}_{Q_{0}}. Hence, 𝒲NΣ=𝒲NΣ,1𝒲NΣ,2𝒲NΣ,3\mathcal{W}_{N}^{\Sigma}=\mathcal{W}_{N}^{\Sigma,1}\cup\mathcal{W}_{N}^{\Sigma,2}\cup\mathcal{W}_{N}^{\Sigma,3} where

𝒲NΣ,1:\displaystyle\mathcal{W}_{N}^{\Sigma,1}: ={I𝒲NΣ:Q0QJ},\displaystyle=\{I\in\mathcal{W}_{N}^{\Sigma}:Q_{0}\subset Q_{J}\},
𝒲NΣ,2:\displaystyle\mathcal{W}_{N}^{\Sigma,2}: ={I𝒲NΣ:QJQ0,(QJ)2N(Q0)},\displaystyle=\{I\in\mathcal{W}_{N}^{\Sigma}:Q_{J}\subset Q_{0},\ \ell(Q_{J})\leq 2^{-N}\,\ell(Q_{0})\},
𝒲NΣ,3:\displaystyle\mathcal{W}_{N}^{\Sigma,3}: ={I𝒲NΣ:QJQ0=Ø}.\displaystyle=\{I\in\mathcal{W}_{N}^{\Sigma}:Q_{J}\cap Q_{0}=\mbox{{\O}}\}.

For later use it is convenient to observe that

(2.44) dist(QJ,I)dist(QJ,J)+diam(J)+diam(I)(J)+(I)(I).\operatorname{dist}(Q_{J},I)\leq\operatorname{dist}(Q_{J},J)+\operatorname{diam}(J)+\operatorname{diam}(I)\approx\ell(J)+\ell(I)\approx\ell(I).

Let us first consider 𝒲NΣ,1\mathcal{W}_{N}^{\Sigma,1}. If I𝒲NΣ,1I\in\mathcal{W}_{N}^{\Sigma,1} we clearly have

(Q0)(QJ)=(J)(I)(QI)(Q0)\ell(Q_{0})\leq\ell(Q_{J})=\ell(J)\approx\ell(I)\approx\ell(Q_{I})\leq\ell(Q_{0})

and since QI𝔻Q0Q_{I}\in\mathbb{D}_{Q_{0}}

dist(I,xQ0)dist(I,QI)+diam(QI)(I).\operatorname{dist}(I,x_{Q_{0}})\leq\operatorname{dist}(I,Q_{I})+\operatorname{diam}(Q_{I})\approx\ell(I).

In particular, #𝒲NΣ,11\#\mathcal{W}_{N}^{\Sigma,1}\lesssim 1. Thus if we set Q^I:=QJ\widehat{Q}_{I}:=Q_{J} it follows from (2.44) that the two first conditions in (2.42) hold and also I𝒲NΣ,1𝟏Q^I#𝒲NΣ,11\sum_{I\in\mathcal{W}_{N}^{\Sigma,1}}\mathbf{1}_{\widehat{Q}_{I}}\leq\#\mathcal{W}_{N}^{\Sigma,1}\lesssim 1.

Consider next 𝒲NΣ,2\mathcal{W}_{N}^{\Sigma,2}. For any I𝒲NΣ,2I\in\mathcal{W}_{N}^{\Sigma,2} we also set Q^I:=QJ\widehat{Q}_{I}:=Q_{J} so that from (2.44) we clearly see that the two first conditions in (2.42) hold. It then remains to estimate the overlap. With this goal in mind we first note that if I𝒲NΣ,2I\in\mathcal{W}_{N}^{\Sigma,2}, the fact that QI𝔻N,Q0Q_{I}\in\mathbb{D}_{\mathcal{F}_{N},Q_{0}} yields

2N(Q0)<(QI)(I)(J)(QJ)2N(Q0),2^{-N}\,\ell(Q_{0})<\ell(Q_{I})\approx\ell(I)\approx\ell(J)\approx\ell(Q_{J})\leq 2^{-N}\,\ell(Q_{0}),

hence (I)2N(Q0)\ell(I)\approx 2^{-N}\,\ell(Q_{0}). Suppose next that QJQJ=Q^IQ^IØQ_{J}\cap Q_{J}^{\prime}=\widehat{Q}_{I}\cap\widehat{Q}_{I}\neq\mbox{{\O}} for I,I𝒲NΣ,2I,I^{\prime}\in\mathcal{W}_{N}^{\Sigma,2}. Then since II touches JJ and II^{\prime} touches JJ^{\prime}

dist(I,I)diam(J)+dist(J,QJ)+diam(QJ)+diam(QJ)+diam(J)(J)+(J)2N(Q0).\operatorname{dist}(I,I^{\prime})\leq\operatorname{diam}(J)+\operatorname{dist}(J,Q_{J})+\operatorname{diam}(Q_{J})+\operatorname{diam}(Q_{J}^{\prime})+\operatorname{diam}(J^{\prime})\\ \approx\ell(J)+\ell(J^{\prime})\approx 2^{-N}\,\ell(Q_{0}).

Hence fixed I𝒲NΣ,2I\in\mathcal{W}_{N}^{\Sigma,2} there is a uniformly bounded number of I𝒲NΣ,2I^{\prime}\in\mathcal{W}_{N}^{\Sigma,2} with Q^IQ^IØ\widehat{Q}_{I}\cap\widehat{Q}_{I^{\prime}}\neq\mbox{{\O}}, and, in particular, I𝒲NΣ,2𝟏Q^I1\sum_{I\in\mathcal{W}_{N}^{\Sigma,2}}\mathbf{1}_{\widehat{Q}_{I}}\lesssim 1.

We finally take into consideration the most delicate collection 𝒲NΣ,3\mathcal{W}_{N}^{\Sigma,3}. In this case for every I𝒲NΣ,3I\in\mathcal{W}_{N}^{\Sigma,3} we pick Q^I𝔻\widehat{Q}_{I}\in\mathbb{D} so that Q^IxQJ\widehat{Q}_{I}\ni x_{Q_{J}} and (Q^I)=2M(QJ)\ell(\widehat{Q}_{I})=2^{-M^{\prime}}\,\ell(Q_{J}) with M3M^{\prime}\geq 3 large enough so that 2M2Ξ22^{M^{\prime}}\geq 2\Xi^{2} (cf. (2.14)). Note that since M3M^{\prime}\geq 3 we have that Q^IQJ\widehat{Q}_{I}\subset Q_{J} which, together with (2.44), implies

dist(I,Q^I)dist(I,QJ)+diam(QJ)(I).\operatorname{dist}(I,\widehat{Q}_{I})\leq\operatorname{dist}(I,Q_{J})+\operatorname{diam}(Q_{J})\lesssim\ell(I).

Hence the first two conditions in (2.42) hold in the current situation.

On the other hand, the choice of MM^{\prime} and (2.14) guarantee that

(2.45) diam(Q^I)2ΞrQ^I2Ξ(Q^I)=2M+1Ξ(QJ)Ξ1(QJ).\operatorname{diam}(\widehat{Q}_{I})\leq 2\,\Xi\,r_{\widehat{Q}_{I}}\leq 2\,\Xi\,\ell(\widehat{Q}_{I})=2^{-M^{\prime}+1}\,\Xi\,\ell(Q_{J})\leq\Xi^{-1}\,\ell(Q_{J}).

Also, since 2ΔQJQJ2\Delta_{Q_{J}}\subset Q_{J}, it follows that Q02ΔQJ=ØQ_{0}\cap 2\Delta_{Q_{J}}=\mbox{{\O}} and therefore 2Ξ1(QJ)dist(xQJ,Q0)2\Xi^{-1}\,\ell(Q_{J})\leq\operatorname{dist}(x_{Q_{J}},Q_{0}). Besides, since QIQ0Q_{I}\subset Q_{0}

dist(xQJ,Q0)diam(QJ)+dist(QJ,J)+diam(J)+diam(I)+dist(I,QI)+diam(QI)(J)(I).\operatorname{dist}(x_{Q_{J}},Q_{0})\leq\operatorname{diam}(Q_{J})+\operatorname{dist}(Q_{J},J)+\operatorname{diam}(J)\\ +\operatorname{diam}(I)+\operatorname{dist}(I,Q_{I})+\operatorname{diam}(Q_{I})\approx\ell(J)\approx\ell(I).

Thus, 2Ξ1(QJ)dist(xQJ,Q0)C(J)2\,\Xi^{-1}\,\ell(Q_{J})\leq\operatorname{dist}(x_{Q_{J}},Q_{0})\leq C\,\ell(J). Suppose next that I,I𝒲NΣ,3I,I^{\prime}\in\mathcal{W}_{N}^{\Sigma,3} are so that Q^IQ^IØ\widehat{Q}_{I}\cap\widehat{Q}_{I^{\prime}}\neq\mbox{{\O}} and assume without loss of generality that Q^IQ^I\widehat{Q}_{I^{\prime}}\subset\widehat{Q}_{I}, hence (J)(J)\ell(J^{\prime})\leq\ell(J). Then, since xQJQ^Ix_{Q_{J}}\in\widehat{Q}_{I} and xQJQ^IQ^Ix_{Q_{J^{\prime}}}\in\widehat{Q}_{I^{\prime}}\subset\widehat{Q}_{I} we get from (2.45)

2Ξ1(QJ)dist(xQJ,Q0)|xQJxQJ|+dist(xQJ,Q0)diam(Q^I)+C(J)Ξ1(QJ)+C(J)2\,\Xi^{-1}\,\ell(Q_{J})\leq\operatorname{dist}(x_{Q_{J}},Q_{0})\leq|x_{Q_{J}}-x_{Q_{J^{\prime}}}|+\operatorname{dist}(x_{Q_{J^{\prime}}},Q_{0})\\ \leq\operatorname{diam}(\widehat{Q}_{I})+C\ell(J^{\prime})\leq\Xi^{-1}\,\ell(Q_{J})+C\ell(J^{\prime})

and therefore Ξ1(QJ)C(J)\Xi^{-1}\,\ell(Q_{J})\leq C\,\ell(J) which in turn gives (I)(J)(J)(I)\ell(I)\approx\ell(J)\approx\ell(J^{\prime})\approx\ell(I^{\prime}). Note also that since II touches JJ, II^{\prime} touches JJ^{\prime}, and Q^IQ^IØ\widehat{Q}_{I}\cap\widehat{Q}_{I^{\prime}}\neq\mbox{{\O}} we obtain

dist(I,I)diam(J)+dist(J,QJ)+diam(QJ)+diam(QJ)+dist(QJ,J)+diam(J)(J)+(J)(I).\operatorname{dist}(I,I^{\prime})\leq\operatorname{diam}(J)+\operatorname{dist}(J,Q_{J})+\operatorname{diam}(Q_{J})+\operatorname{diam}(Q_{J^{\prime}})\\ +\operatorname{dist}(Q_{J^{\prime}},J^{\prime})+\operatorname{diam}(J^{\prime})\approx\ell(J)+\ell(J^{\prime})\approx\ell(I).

Consequently, fixed I𝒲NΣ,3I\in\mathcal{W}_{N}^{\Sigma,3} there is a uniformly bounded number of I𝒲NΣ,3I^{\prime}\in\mathcal{W}_{N}^{\Sigma,3} with Q^IQ^IØ\widehat{Q}_{I}\cap\widehat{Q}_{I^{\prime}}\neq\mbox{{\O}}. As a result, I𝒲NΣ,3𝟏Q^I1\sum_{I\in\mathcal{W}_{N}^{\Sigma,3}}\mathbf{1}_{\widehat{Q}_{I}}\lesssim 1. This clearly completes the proof of (iii)(iii) and hence that of Lemma 2.39. ∎

2.5. Uniformly elliptic operators, elliptic measure and the Green function

Next, we recall several facts concerning elliptic measure and the Green functions. To set the stage let Ωn+1\Omega\subset\mathbb{R}^{n+1} be an open set. Throughout we consider elliptic operators LL of the form Lu=div(Au)Lu=-\mathop{\operatorname{div}}\nolimits(A\nabla u) with A(X)=(ai,j(X))i,j=1n+1A(X)=(a_{i,j}(X))_{i,j=1}^{n+1} being a real (non-necessarily symmetric) matrix such that ai,jL(Ω)a_{i,j}\in L^{\infty}(\Omega) and there exists Λ1\Lambda\geq 1 such that the following uniform ellipticity condition holds

(2.46) Λ1|ξ|2A(X)ξξ,|A(X)ξη|Λ|ξ||η|\displaystyle\Lambda^{-1}|\xi|^{2}\leq A(X)\xi\cdot\xi,\qquad\qquad|A(X)\xi\cdot\eta|\leq\Lambda|\xi|\,|\eta|

for all ξ,ηn+1\xi,\eta\in\mathbb{R}^{n+1} and for almost every XΩX\in\Omega. We write LL^{\top} to denote the transpose of LL, or, in other words, Lu=div(Au)L^{\top}u=-\mathop{\operatorname{div}}\nolimits(A^{\top}\nabla u) with AA^{\top} being the transpose matrix of AA.

We say that uu is a weak solution to Lu=0Lu=0 in Ω\Omega provided that uWloc1,2(Ω)u\in W_{\rm loc}^{1,2}(\Omega) satisfies

A(X)u(X)ϕ(X)𝑑X=0wheneverϕC0(Ω).\iint A(X)\nabla u(X)\cdot\nabla\phi(X)dX=0\quad\mbox{whenever}\,\,\phi\in C^{\infty}_{0}(\Omega).

Associated with LL one can construct an elliptic measure {ωLX}XΩ\{\omega_{L}^{X}\}_{X\in\Omega} and a Green function GLG_{L} (see [20] for full details). Sometimes, in order to emphasize the dependence on Ω\Omega, we will write ωL,Ω\omega_{L,\Omega} and GL,ΩG_{L,\Omega}. If Ω\Omega satisfies the CDC then it follows that all boundary points are Wiener regular and hence for a given fCc(Ω)f\in C_{c}(\partial\Omega) we can define

u(X)=Ωf(z)𝑑ωLX(z),wheneverXΩ,u(X)=\int_{\partial\Omega}f(z)d\omega^{X}_{L}(z),\quad\mbox{whenever}\,\,X\in\Omega,

so that uWloc1,2(Ω)C(Ω¯)u\in W^{1,2}_{\rm loc}(\Omega)\cap C(\overline{\Omega}) satisfies u=fu=f on Ω\partial\Omega and Lu=0Lu=0 in the weak sense in Ω\Omega. Moreover, if Ω\Omega is bounded and fLip(Ω)f\in\operatorname*{Lip}(\Omega) then uW1,2(Ω)u\in W^{1,2}(\Omega). In the same context the Green function satisfies the following properties which will be used along the paper:

(2.47) 0GL(X,Y)C|XY|1n,X,YΩ,XY;\displaystyle 0\leq G_{L}(X,Y)\leq C|X-Y|^{1-n},\quad\forall X,Y\in\Omega,\quad X\neq Y;
(2.48) GL(,Y)Wloc1,2(Ω{Y})C(Ω¯{Y})andGL(,Y)|Ω0YΩ;\displaystyle G_{L}(\cdot,Y)\in W_{\rm loc}^{1,2}(\Omega\setminus\{Y\})\cap C\big{(}\overline{\Omega}\setminus\{Y\}\big{)}\quad\text{and}\quad G_{L}(\cdot,Y)|_{\partial\Omega}\equiv 0\quad\forall Y\in\Omega;
(2.49) GL(X,Y)=GL(Y,X),X,YΩ,XY;\displaystyle G_{L}(X,Y)=G_{L^{\top}}(Y,X),\quad\forall X,Y\in\Omega,\quad X\neq Y;
(2.50) ΩA(X)XGL(X,Y)φ(X)𝑑X=φ(Y),φCc(Ω).\displaystyle\iint_{\Omega}A(X)\nabla_{X}G_{L}(X,Y)\cdot\nabla\varphi(X)\,dX=\varphi(Y),\qquad\forall\,\varphi\in C_{c}^{\infty}(\Omega).

We first define the reverse Hölder class and the AA_{\infty} classes with respect to fixed elliptic measure in Ω\Omega. One reason we take this approach is that we do not know whether n|Ω\mathcal{H}^{n}|_{\partial\Omega} is well-defined since we do not assume any Ahlfors regularity. Hence we have to develop these notions in terms of elliptic measures. To this end, let Ω\Omega satisfy the CDC and let L0L_{0} and LL be two real (non-necessarily symmetric) elliptic operators associated with L0u=div(A0u)L_{0}u=-\mathop{\operatorname{div}}\nolimits(A_{0}\nabla u) and Lu=div(Au)Lu=-\mathop{\operatorname{div}}\nolimits(A\nabla u) where AA and A0A_{0} satisfy (2.46). Let ωL0X\omega^{X}_{L_{0}} and ωLX\omega_{L}^{X} be the elliptic measures of Ω\Omega associated with the operators L0L_{0} and LL respectively with pole at XΩX\in\Omega. Note that if we further assume that Ω\Omega is connected then ωLXωLY\omega_{L}^{X}\ll\omega_{L}^{Y} on Ω\partial\Omega for every X,YΩX,Y\in\Omega. Hence if ωLX0ωL0Y0\omega_{L}^{X_{0}}\ll\omega_{L_{0}}^{Y_{0}} on Ω\partial\Omega for some X0,Y0ΩX_{0},Y_{0}\in\Omega then ωLXωL0Y\omega_{L}^{X}\ll\omega_{L_{0}}^{Y} on Ω\partial\Omega for every X,YΩX,Y\in\Omega and thus we can simply write ωLωL0\omega_{L}\ll\omega_{L_{0}} on Ω\partial\Omega. In the latter case we will use the notation

(2.51) h(;L,L0,X)=dωLXdωL0Xh(\cdot\,;L,L_{0},X)=\frac{d\omega_{L}^{X}}{d\omega_{L_{0}}^{X}}

to denote the Radon-Nikodym derivative of ωLX\omega_{L}^{X} with respect to ωL0X\omega_{L_{0}}^{X}, which is a well-defined function ωL0X\omega_{L_{0}}^{X}-almost everywhere on Ω\partial\Omega.

Definition 2.52 (Reverse Hölder and AA_{\infty} classes).

Fix Δ0=B0Ω\Delta_{0}=B_{0}\cap\partial\Omega where B0=B(x0,r0)B_{0}=B(x_{0},r_{0}) with x0Ωx_{0}\in\partial\Omega and 0<r0<diam(Ω)0<r_{0}<\operatorname{diam}(\partial\Omega). Given pp, 1<p<1<p<\infty, we say that ωLRHp(Δ0,ωL0)\omega_{L}\in RH_{p}(\Delta_{0},\omega_{L_{0}}), provided that ωLωL0\omega_{L}\ll\omega_{L_{0}} on Δ0\Delta_{0}, and there exists C1C\geq 1 such that

(Δh(y;L,L0,XΔ0)p𝑑ωL0XΔ0(y))1pCΔh(y;L,L0,XΔ0)𝑑ωL0XΔ0(y)=CωLXΔ0(Δ)ωL0XΔ0(Δ),\left(\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{\Delta}h(y;L,L_{0},X_{\Delta_{0}})^{p}d\omega_{L_{0}}^{X_{\Delta_{0}}}(y)\right)^{\frac{1}{p}}\leq C\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{\Delta}h(y;L,L_{0},X_{\Delta_{0}})d\omega_{L_{0}}^{X_{\Delta_{0}}}(y)=C\frac{\omega_{L}^{X_{\Delta_{0}}}(\Delta)}{\omega_{L_{0}}^{X_{\Delta_{0}}}(\Delta)},

for every Δ=BΩ\Delta=B\cap\partial\Omega where BB(x0,r0)B\subset B(x_{0},r_{0}), B=B(x,r)B=B(x,r) with xΩx\in\partial\Omega, 0<r<diam(Ω)0<r<\operatorname{diam}(\partial\Omega). The infimum of the constants CC as above is denoted by [ωL]RHp(Δ0,ωL0)[\omega_{L}]_{RH_{p}(\Delta_{0},\omega_{L_{0}})}.

Similarly, we say that ωLRHp(Ω,ωL0)\omega_{L}\in RH_{p}(\partial\Omega,\omega_{L_{0}}) provided that for every Δ0=Δ(x0,r0)\Delta_{0}=\Delta(x_{0},r_{0}) with x0Ωx_{0}\in\partial\Omega and 0<r0<diam(Ω)0<r_{0}<\operatorname{diam}(\partial\Omega) one has ωLRHp(Δ0,ωL0)\omega_{L}\in RH_{p}(\Delta_{0},\omega_{L_{0}}) uniformly on Δ0\Delta_{0}, that is,

[ωL]RHp(Ω,ωL0):=supΔ0[ωL]RHp(Δ0,ωL0)<.[\omega_{L}]_{RH_{p}(\partial\Omega,\omega_{L_{0}})}:=\sup_{\Delta_{0}}[\omega_{L}]_{RH_{p}(\Delta_{0},\omega_{L_{0}})}<\infty.

Finally,

A(Δ0,ωL0)=p>1RHp(Δ0,ωL0)andA(Ω,ωL0)=p>1RHp(Ω,ωL0).A_{\infty}(\Delta_{0},\omega_{L_{0}})=\bigcup_{p>1}RH_{p}(\Delta_{0},\omega_{L_{0}})\quad\mbox{and}\quad A_{\infty}(\partial\Omega,\omega_{L_{0}})=\bigcup_{p>1}RH_{p}(\partial\Omega,\omega_{L_{0}}).

The following result lists a number of properties which will be used throughout the paper, proofs may be found in [20]:

Lemma 2.53.

Suppose that Ωn+1\Omega\subset\mathbb{R}^{n+1}, n2n\geq 2, is a 1-sided NTA domain satisfying the CDC. Let L0=div(A0)L_{0}=-\mathop{\operatorname{div}}\nolimits(A_{0}\nabla) and L=div(A)L=-\mathop{\operatorname{div}}\nolimits(A\nabla) be two real (non-necessarily symmetric) elliptic operators, there exist C11C_{1}\geq 1, ρ(0,1)\rho\in(0,1) (depending only on dimension, the 1-sided NTA constants, the CDC constant, and the ellipticity of LL) and C21C_{2}\geq 1 (depending on the same parameters and on the ellipticity of L0L_{0}), such that for every B0=B(x0,r0)B_{0}=B(x_{0},r_{0}) with x0Ωx_{0}\in\partial\Omega, 0<r0<diam(Ω)0<r_{0}<\operatorname{diam}(\partial\Omega), and Δ0=B0Ω\Delta_{0}=B_{0}\cap\partial\Omega we have the following properties:

  • (a)(a)

    ωLY(Δ0)C11\omega_{L}^{Y}(\Delta_{0})\geq C_{1}^{-1} for every YC11B0ΩY\in C_{1}^{-1}B_{0}\cap\Omega and ωLXΔ0(Δ0)C11\omega_{L}^{X_{\Delta_{0}}}(\Delta_{0})\geq C_{1}^{-1}.

  • (b)(b)

    If B=B(x,r)B=B(x,r) with xΩx\in\partial\Omega and Δ=BΩ\Delta=B\cap\partial\Omega is such that 2BB02B\subset B_{0}, then for all XΩB0X\in\Omega\setminus B_{0} we have that C11ωLX(Δ)rn1GL(X,XΔ)C1ωLX(Δ){C_{1}^{-1}}\omega_{L}^{X}(\Delta)\leq r^{n-1}G_{L}(X,X_{\Delta})\leq C_{1}\omega_{L}^{X}(\Delta).

  • (c)(c)

    If XΩ4B0X\in\Omega\setminus 4B_{0}, then ωLX(2Δ0)C1ωLX(Δ0)\omega_{L}^{X}(2\Delta_{0})\leq C_{1}\omega_{L}^{X}(\Delta_{0}).

  • (d)(d)

    For every XΩ2κ0B0X\in\Omega\setminus 2\kappa_{0}B_{0} with κ0\kappa_{0} as in (2.32), we have that

    1C11ωLX(Δ0)dωLXΔ0dωLX(y)C11ωLX(Δ0),for ωLX-a.e. yΔ0.\frac{1}{C}_{1}\frac{1}{\omega_{L}^{X}(\Delta_{0})}\leq\frac{d\omega_{L}^{{X_{\Delta_{0}}}}}{d\omega_{L}^{X}}(y)\leq C_{1}\frac{1}{\omega_{L}^{X}(\Delta_{0})},\qquad\mbox{for $\omega_{L}^{X}$-a.e. $y\in\Delta_{0}$}.
  • (e)(e)

    For every XB0ΩX\in B_{0}\cap\Omega and for any j1j\geq 1

    dωLXdωLX2jΔ0(y)C1(δ(X)2jr0)ρ,for ωLX-a.e. yΩ2jΔ0.\frac{d\omega_{L}^{X}}{d\omega_{L}^{X_{2^{j}\Delta_{0}}}}(y)\leq C_{1}\,\bigg{(}\frac{\delta(X)}{2^{j}\,r_{0}}\bigg{)}^{\rho},\qquad\mbox{for $\omega_{L}^{X}$-a.e. $y\in\partial\Omega\setminus 2^{j}\,\Delta_{0}$}.
Remark 2.54.

We note that from (d)(d) in the previous result, Harnack’s inequality, and (2.14) one can easily see that

(2.55) dωLXQdωLXQ′′(y)1ωLXQ′′(Q), for ωLXQ′′-a.e. yQ,whenever QQ′′𝔻.\frac{d\omega_{L}^{X_{Q^{\prime}}}}{d\omega_{L}^{X_{Q^{\prime\prime}}}}(y)\approx\frac{1}{\omega_{L}^{X_{Q^{\prime\prime}}}(Q^{\prime})},\qquad\mbox{ for $\omega_{L}^{X_{Q^{\prime\prime}}}$-a.e. }y\in Q^{\prime},\mbox{whenever }Q^{\prime}\subset Q^{\prime\prime}\in\mathbb{D}.

Observe that since ωLXQ′′ωLXQ\omega_{L}^{X_{Q^{\prime\prime}}}\ll\omega_{L}^{X_{Q^{\prime}}} an analogous inequality for the reciprocal of the Radon-Nikodym derivative follows immediately.

Remark 2.56.

Given Ω\Omega, a 1-sided NTA domain satisfying the CDC, we claim that if Δ=Δ(x,r)\Delta=\Delta(x,r) with xΩx\in\partial\Omega and 0<r<diam(Ω)0<r<\operatorname{diam}(\partial\Omega) then diam(Δ)r\operatorname{diam}(\Delta)\approx r. To see this we first observe that diam(Δ)2r\operatorname{diam}(\Delta)\leq 2r. If diam(Δ)c0r/4\operatorname{diam}(\Delta)\geq c_{0}r/4c0c_{0} is the Corkscrew constant— then clearly diam(Δ)r\operatorname{diam}(\Delta)\approx r. Hence, we may assume that diam(Δ)<c0r/4\operatorname{diam}(\Delta)<c_{0}r/4. Let s=2diam(Δ)s=2\operatorname{diam}(\Delta) so that diam(Δ)<s<r\operatorname{diam}(\Delta)<s<r and note that one can easily see that Δ=Δ:=Δ(x,s)\Delta=\Delta^{\prime}:=\Delta(x,s). Associated with Δ\Delta and Δ\Delta^{\prime} we can consider XΔX_{\Delta} and XΔX_{\Delta^{\prime}} the corresponding Corkscrew points. These are different, despite the fact that Δ=Δ(x,r)\Delta=\Delta(x,r). Indeed,

c0rδ(XΔ)|XΔXΔ|+|XΔx||XΔXΔ|+s<|XΔXΔ|+c02rc_{0}r\leq\delta(X_{\Delta})\leq|X_{\Delta}-X_{\Delta^{\prime}}|+|X_{\Delta^{\prime}}-x|\leq|X_{\Delta}-X_{\Delta^{\prime}}|+s<|X_{\Delta}-X_{\Delta^{\prime}}|+\frac{c_{0}}{2}r

which yields that |XΔXΔ|c02r|X_{\Delta}-X_{\Delta^{\prime}}|\geq\frac{c_{0}}{2}r. Note that XΔ2B:=B(x,2s)X_{\Delta}\notin 2B^{\prime}:=B(x,2s) since otherwise we would get a contradiction: c0rδ(XΔ)|XΔx|<2s<c0rc_{0}r\leq\delta(X_{\Delta})\leq|X_{\Delta}-x|<2s<c_{0}r. Hence we can invoke Lemma 2.53 parts (a)(a) and (b)(b) and (2.47) to see that

1ωLXΔ(Δ)=ωLXΔ(Δ)sn1GL(XΔ,XΔ)sn1|XΔXΔ|1n(s/r)n1.1\approx\omega_{L}^{X_{\Delta}}(\Delta)=\omega_{L}^{X_{\Delta}}(\Delta^{\prime})\approx s^{n-1}G_{L}(X_{\Delta},X_{\Delta^{\prime}})\lesssim s^{n-1}|X_{\Delta}-X_{\Delta^{\prime}}|^{1-n}\lesssim(s/r)^{n-1}.

This and the fact that n2n\geq 2 easily yields that rsr\lesssim s as desired.

We close this section by establishing an estimate for the non-tangential maximal function for elliptic-measure solutions.

Proposition 2.57.

Let Ωn+1\Omega\subset\mathbb{R}^{n+1} be a 1-sided NTA domain satisfying the CDC. Given Q0𝔻Q_{0}\in\mathbb{D} and fC(Ω)f\in C(\partial\Omega) with suppf2Δ~Q0\operatorname{supp}f\subset 2\widetilde{\Delta}_{Q_{0}} let

u(X)=Ωf(y)𝑑ωLX(y),XΩ.u(X)=\int_{\partial\Omega}f(y)\,d\omega_{L}^{X}(y),\qquad X\in\partial\Omega.

Then for every xQ0x\in Q_{0},

(2.58) 𝒩Q0u(x)supΔx0<rΔ<4ΞrQ0Δ|f(y)|𝑑ωLXQ0(y),\mathcal{N}_{Q_{0}}u(x)\lesssim\sup_{\begin{subarray}{c}\Delta\ni x\\ 0<r_{\Delta}<4\Xi r_{Q_{0}}\end{subarray}}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{\Delta}|f(y)|\,d\omega_{L}^{X_{Q_{0}}}(y),

and, as a consequence, for every 1<q1<q\leq\infty

(2.59) 𝒩Q0uLq(Q0,ωLXQ0)fLq(2Δ~Q0,ωLXQ0).\|\mathcal{N}_{Q_{0}}u\|_{L^{q}(Q_{0},\omega_{L}^{X_{Q_{0}}})}\lesssim\|f\|_{L^{q}(2\widetilde{\Delta}_{Q_{0}},\omega_{L}^{X_{Q_{0}}})}.

Moreover, the implicit constants depend just on dimension nn, the 1-sided NTA constants, the CDC constant, and the ellipticity constant of LL and on qq in (2.59).

Proof.

By decomposing ff into its positive and negative parts we may assume that ff is non-negative with suppf2Δ~Q0\operatorname{supp}f\subset 2\widetilde{\Delta}_{Q_{0}} and construct the associated uu as in the statement which is non-negative. Fix xQ0x\in Q_{0} and let XΓQ0(x)X\in\Gamma_{Q_{0}}^{*}(x). Then, by definition there are Q𝔻Q0Q\in\mathbb{D}_{Q_{0}} and I𝒲QI\in\mathcal{W}_{Q}^{*} such that xQx\in Q and XIX\in I^{**}. Hence using Harnack’s inequality and the notation introduced in (2.14)–(2.16)

u(X)=Ωf(y)𝑑ωLX(y)Ωf(y)𝑑ωLXQ(y)4Δ~Qf(y)dωLXQ(y)+j=32jΔ~Q2j1Δ~Qf(y)dωLXQ(y)=:j=2j.u(X)=\int_{\partial\Omega}f(y)\,d\omega_{L}^{X}(y)\approx\int_{\partial\Omega}f(y)\,d\omega_{L}^{X_{Q}}(y)\\ \leq\int_{4\,\widetilde{\Delta}_{Q}}f(y)\,d\omega_{L}^{X_{Q}}(y)+\sum_{j=3}^{\infty}\int_{2^{j}\,\widetilde{\Delta}_{Q}\setminus 2^{j-1}\,\widetilde{\Delta}_{Q}}f(y)\,d\omega_{L}^{X_{Q}}(y)=:\sum_{j=2}^{\infty}\mathcal{I}_{j}.

Let k00k_{0}\geq 0 be such that (Q)=2k0(Q0)\ell(Q)=2^{-k_{0}}\ell(Q_{0}). Observe that for every jk0+3j\geq k_{0}+3 one has that 2Δ~Q02j1Δ~Q=Ø2\widetilde{\Delta}_{Q_{0}}\setminus 2^{j-1}\widetilde{\Delta}_{Q}=\mbox{{\O}}. Otherwise there is z2Δ~Q02j1Δ~Qz\in 2\widetilde{\Delta}_{Q_{0}}\setminus 2^{j-1}\widetilde{\Delta}_{Q} and hence we get a contradiction:

4ΞrQ02j1k0ΞrQ0=2j1ΞrQ|zxQ||zxQ0|+|xQxQ0|3ΞrQ0.4\,\Xi\,r_{Q_{0}}\leq 2^{j-1-k_{0}}\,\Xi\,r_{Q_{0}}=2^{j-1}\,\Xi\,r_{Q}\leq|z-x_{Q}|\leq|z-x_{Q_{0}}|+|x_{Q}-x_{Q_{0}}|\leq 3\,\Xi\,r_{Q_{0}}.

With this in hand, and since suppf2Δ~Q0\operatorname{supp}f\subset 2\widetilde{\Delta}_{Q_{0}}, we clearly see that j=0\mathcal{I}_{j}=0 for jk0+3j\geq k_{0}+3.

In order to estimate the j\mathcal{I}_{j}’s we need some preparatives. Note that for every 2jk0+22\leq j\leq k_{0}+2 one has 2jB~Q5B~Q02^{j}\widetilde{B}_{Q}\subset 5\widetilde{B}_{Q_{0}}. We claim that

(2.60) dωLX2jΔ~QdωLXQ0(y)1ωLXQ0(2jΔ~Q),for ωLXQ0-a.e. y2jΔ~Q2jk0+2.\frac{d\omega_{L}^{X_{2^{j}\widetilde{\Delta}_{Q}}}}{d\omega_{L}^{X_{Q_{0}}}}(y)\lesssim\frac{1}{\omega_{L}^{X_{Q_{0}}}(2^{j}\widetilde{\Delta}_{Q})},\qquad\mbox{for $\omega_{L}^{X_{Q_{0}}}$-a.e. $y\in 2^{j}\,\widetilde{\Delta}_{Q}$, $2\leq j\leq k_{0}+2$.}

Indeed, this estimate follows from Harnack’s inequality and Lemma 2.53 part (a)(a) when jk0j\approx k_{0} since 2j(Q)(Q0)2^{j}\,\ell(Q)\approx\ell(Q_{0}), and from Lemma 2.53 part (d)(d) whenever jk0j\ll k_{0}. We also observe that Lemma 2.53 part (a)(a) and Harnack’s inequality readily give that

(2.61) ωLX2jΔ~Q(2jΔ~Q)1,for every 2jk0+2.\omega_{L}^{X_{2^{j}\widetilde{\Delta}_{Q}}}(2^{j}\widetilde{\Delta}_{Q})\approx 1,\qquad\text{for every $2\leq j\leq k_{0}+2$.}

Finally, by Lemma 2.53 part (e)(e) and Harnack’s inequality it follows that

(2.62) dωLXQdωLX2j1Δ~Q(y)2jρ,for ωLXQ-a.e. yΩ2j1Δ~Q,j3.\frac{d\omega_{L}^{X_{Q}}}{d\omega_{L}^{X_{2^{j-1}\widetilde{\Delta}_{Q}}}}(y)\lesssim 2^{-j\,\rho},\qquad\mbox{for $\omega_{L}^{X_{Q}}$-a.e. $y\in\partial\Omega\setminus 2^{j-1}\,\widetilde{\Delta}_{Q}$},\ j\geq 3.

Let us start estimating 2\mathcal{I}_{2}. Use Harnack’s inequality and (2.61), (2.60) with j=2j=2, to conclude that

24Δ~Qf(y)𝑑ωLXΔ~Q(y)4Δ~Qf(y)𝑑ωLXQ0(y).\mathcal{I}_{2}\approx\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{4\,\widetilde{\Delta}_{Q}}f(y)\,d\omega_{L}^{X_{\widetilde{\Delta}_{Q}}}(y)\approx\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{4\,\widetilde{\Delta}_{Q}}f(y)\,d\omega_{L}^{X_{Q_{0}}}(y).

On the other hand, for 3jk0+23\leq j\leq k_{0}+2, we employ (2.62), Harnack’s inequality, (2.61), and (2.60)

j2jρ2jΔ~Q2j1Δ~Qf(y)𝑑ωLX2j1Δ~Q(y)2jρ2jΔ~Qf(y)𝑑ωLX2jΔ~Q(y)2jρ2jΔ~Qf(y)𝑑ωLXQ0(y).\mathcal{I}_{j}\lesssim 2^{-j\,\rho}\int_{2^{j}\,\widetilde{\Delta}_{Q}\setminus 2^{j-1}\,\widetilde{\Delta}_{Q}}f(y)\,d\omega_{L}^{X_{2^{j-1}\,\widetilde{\Delta}_{Q}}}(y)\lesssim 2^{-j\,\rho}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{2^{j}\,\widetilde{\Delta}_{Q}}f(y)\,d\omega_{L}^{X_{2^{j}\,\widetilde{\Delta}_{Q}}}(y)\\ \approx 2^{-j\,\rho}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{2^{j}\,\widetilde{\Delta}_{Q}}f(y)\,d\omega_{L}^{X_{Q_{0}}}(y).

If we now collect all the obtained estimates we conclude as desired (2.58):

u(X)j=2k0+2jj=2k0+22jρ2jΔ~Qf(y)𝑑ωLXQ0(y)supΔx0<rΔ<8ΞrQ0Δ|f(y)|𝑑ωLXQ0(y)j=22jρsupΔx0<rΔ<4ΞrQ0Δ|f(y)|𝑑ωLXQ0(y).u(X)\lesssim\sum_{j=2}^{k_{0}+2}\mathcal{I}_{j}\lesssim\sum_{j=2}^{k_{0}+2}2^{-j\,\rho}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{2^{j}\,\widetilde{\Delta}_{Q}}f(y)\,d\omega_{L}^{X_{Q_{0}}}(y)\\ \leq\sup_{\begin{subarray}{c}\Delta\ni x\\ 0<r_{\Delta}<8\Xi r_{Q_{0}}\end{subarray}}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{\Delta}|f(y)|\,d\omega_{L}^{X_{Q_{0}}}(y)\,\sum_{j=2}^{\infty}2^{-j\,\rho}\lesssim\sup_{\begin{subarray}{c}\Delta\ni x\\ 0<r_{\Delta}<4\Xi r_{Q_{0}}\end{subarray}}\mathchoice{{\vbox{\hbox{$\textstyle-$}}\kern-4.86108pt}}{{\vbox{\hbox{$\scriptstyle-$}}\kern-3.25pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-2.29166pt}}{{\vbox{\hbox{$\scriptscriptstyle-$}}\kern-1.875pt}}\!\int_{\Delta}|f(y)|\,d\omega_{L}^{X_{Q_{0}}}(y).

To complete the proof we just need to obtain (2.59) but this follows at once upon using (2.58) and observing that the local Hardy-Littlewood maximal function on its right hand side is bounded on Lq(20Δ~Q0,ωLXQ0){L^{q}(20\,\widetilde{\Delta}_{Q_{0}},\omega_{L}^{X_{Q_{0}}})} since ωLXQ0\omega_{L}^{X_{Q_{0}}} is a doubling measure in 20Δ~Q020\,\widetilde{\Delta}_{Q_{0}} by Lemma 2.53 parts (a)(a) and (c)(c). ∎

3. Dyadic sawtooth lemma for projections

In this section, we shall prove two dyadic versions of the main lemma in [10]. To set the stage we sate a result which is partially proved in [17, Proposition 6.7] under the further assumption that Ω\partial\Omega is Ahlfors regular

Proposition 3.1.

Let Ωn+1\Omega\subset\mathbb{R}^{n+1}, n2n\geq 2, be a 1-sided NTA domain satisfying the CDC. Fix Q0𝔻Q_{0}\in\mathbb{D} and let ={Qk}k𝔻Q0\mathcal{F}=\{Q_{k}\}_{k}\subset\mathbb{D}_{Q_{0}} be a family of pairwise disjoint dyadic cubes. There exists YQ0ΩΩ,Q0Ω,Q0Y_{Q_{0}}\in\Omega\cap\Omega_{\mathcal{F},Q_{0}}\cap\Omega_{\mathcal{F},Q_{0}}^{*} so that

(3.2) dist(YQ0,Ω)dist(YQ0,Ω,Q0)dist(YQ0,Ω,Q0)(Q0),\operatorname{dist}(Y_{Q_{0}},\partial\Omega)\approx\operatorname{dist}(Y_{Q_{0}},\partial\Omega_{\mathcal{F},Q_{0}})\approx\operatorname{dist}(Y_{Q_{0}},\partial\Omega_{\mathcal{F},Q_{0}}^{*})\approx\ell(Q_{0}),

where the implicit constants depend only on dimension, the 1-sided NTA constants, the CDC constant, and is independent of Q0Q_{0} and \mathcal{F}. Additionally, for each QjQ_{j}\in\mathcal{F}, there is an nn-dimensional cube PjΩ,Q0P_{j}\subset\partial\Omega_{\mathcal{F},Q_{0}}, which is contained in a face of II^{*} for some I𝒲I\in\mathcal{W}, and which satisfies

(3.3) (Pj)dist(Pj,Qj)dist(Pj,Ω)(I)(Qj),\ell(P_{j})\approx\operatorname{dist}(P_{j},Q_{j})\approx\operatorname{dist}(P_{j},\partial\Omega)\approx\ell(I)\approx\ell(Q_{j}),

and

(3.4) j1Pj1,\sum_{j}1_{P_{j}}\lesssim 1,

where the implicit constants depend on allowable parameters.

Proof.

Note first that Ω,Q0\Omega_{\mathcal{F},Q_{0}} is a 1-sided NTA domain satisfying the CDC (see Proposition 2.37). Pick an arbitrary x0Ω,Q0x_{0}\in\partial\Omega_{\mathcal{F},Q_{0}} and let Y0Y_{0} be a Corkscrew point relative to the surface ball B(x0,diam(Ω,Q0)/2)Ω,Q0B(x_{0},\operatorname{diam}(\partial\Omega_{\mathcal{F},Q_{0}})/2)\cap\partial\Omega_{\mathcal{F},Q_{0}} for the bounded domain Ω,Q0\Omega_{\mathcal{F},Q_{0}} (recall that one has diam(Ω,Q0)(Q0)<\operatorname{diam}(\partial\Omega_{\mathcal{F},Q_{0}})\approx\ell(Q_{0})<\infty by (2.31)). Note that Y0Ω,Q0ΩY_{0}\in\Omega_{\mathcal{F},Q_{0}}\subset\Omega, which is comprised of fattened Whitney boxes, then Y0IY_{0}\in I^{**} for some I𝒲I\in\mathcal{W}, with int(I)Ω,Q0\operatorname{int}(I^{**})\subset\Omega_{\mathcal{F},Q_{0}}. Let YQ0=X(I)Y_{Q_{0}}=X(I) be the center of II so that δ(Y0)(I)δ(YQ0)\delta(Y_{0})\approx\ell(I)\approx\delta(Y_{Q_{0}}). Then,

(Q0)diam(Ω,Q0)dist(Y0,Ω,Q0)dist(Y0,Ω,Q0)δ(Y0)δ(YQ0)(I)diam(Ω,Q0)=diam(Ω,Q0)(Q0).\ell(Q_{0})\approx\operatorname{diam}(\partial\Omega_{\mathcal{F},Q_{0}})\approx\operatorname{dist}(Y_{0},\partial\Omega_{\mathcal{F},Q_{0}})\leq\operatorname{dist}(Y_{0},\partial\Omega_{\mathcal{F},Q_{0}^{*}})\leq\delta(Y_{0})\\ \approx\delta(Y_{Q_{0}})\approx\ell(I)\leq\operatorname{diam}(\Omega_{\mathcal{F},Q_{0}})=\operatorname{diam}(\partial\Omega_{\mathcal{F},Q_{0}})\approx\ell(Q_{0}).

To continue we note that the existence of the family {Pj}j\{P_{j}\}_{j} so that (3.3) holds has been proved in [17, Proposition 6.7] under the further assumption that Ω\partial\Omega is Ahlfors regular. However, a careful examination of the proof shows that the same argument applies in our scenario. We are left with showing (3.4). To see this, observe that as in [17, Remark 6.9] if PjPkØP_{j}\cap P_{k}\neq\mbox{{\O}} then (Qj)(Qk)\ell(Q_{j})\approx\ell(Q_{k}). Indeed from the previous result PjIjP_{j}\subset I_{j}^{*} and PkIkP_{k}\subset I_{k}^{*} for some Ij,Ik𝒲I_{j},I_{k}\in\mathcal{W}. Thus IjI_{j}^{*} meets IkI_{k}^{*} and by construction IjI_{j} and IkI_{k} meet. Using (3.3) and the nature of the Whitney cubes we see that (Qj)(Ij)(Ik)(Qk)\ell(Q_{j})\approx\ell(I_{j})\approx\ell(I_{k})\approx\ell(Q_{k}). Using this and (3.3) one can also see that dist(Qj,Qk)(Qj)(Qk)\operatorname{dist}(Q_{j},Q_{k})\lesssim\ell(Q_{j})\approx\ell(Q_{k}). Hence, fixing Pj0P_{j_{0}} and xPj0x\in P_{j_{0}} we have some constant k01k_{0}\geq 1 (depending on the allowable parameters) such that

j1Pj(x)#{Pk:PkPj0Ø}#{Qk: 2k0(Qk)(Qj0)2k0,dist(Qk,Qj0)2k0(Qj0)}=k=k0k0#{Qk:(Qk)=2k(Qj0),dist(Qk,Qj0)2k0(Qj0)}=:k=k0k0Nk.\sum_{j}1_{P_{j}}(x)\leq\#\{P_{k}:\ P_{k}\cap P_{j_{0}}\neq\mbox{{\O}}\}\\ \leq\#\big{\{}Q_{k}:\ 2^{-k_{0}}\leq\tfrac{\ell(Q_{k})}{\ell(Q_{j_{0}})}\leq 2^{k_{0}},\ \operatorname{dist}(Q_{k},Q_{j_{0}})\leq 2^{k_{0}}\ell(Q_{j_{0}})\big{\}}\\ =\sum_{k=-k_{0}}^{k_{0}}\#\big{\{}Q_{k}:\ \ell(Q_{k})=2^{k}\ell(Q_{j_{0}}),\ \operatorname{dist}(Q_{k},Q_{j_{0}})\leq 2^{k_{0}}\ell(Q_{j_{0}})\big{\}}=:\sum_{k=-k_{0}}^{k_{0}}N_{k}.

To estimate each of the terms in the last sum fix kk and note that since the cubes belong to the same generation then QkQ_{k}’s involved are disjoint and hence so they are the corresponding ΔQk\Delta_{Q_{k}}’s which all have radius (2C)12k(Qj0)(2C)^{-1}2^{k}\ell(Q_{j_{0}}). In particular, |xQkxQk|2k(Qj0)2k0(Qj0)|x_{Q_{k}}-x_{Q_{k}^{\prime}}|\gtrsim 2^{k}\ell(Q_{j_{0}})\geq 2^{-k_{0}}\ell(Q_{j_{0}}) for any such cubes QkQ_{k} and QkQ_{k^{\prime}}. Moreover,

|xQkxQj0|diam(Qk)+dist(Qk,Qj0)+diam(Qj0)2k0(Qj0).|x_{Q_{k}}-x_{Q_{j_{0}}}|\leq\operatorname{diam}(Q_{k})+\operatorname{dist}(Q_{k},Q_{j_{0}})+\operatorname{diam}(Q_{j_{0}})\lesssim 2^{k_{0}}\ell(Q_{j_{0}}).

Thus it is easy to see (since n+1\mathbb{R}^{n+1} is geometric doubling) that Nk22k0(n+1)N_{k}\lesssim 2^{2k_{0}(n+1)}. All these together gives us desired (3.4) —we note in passing that the argument in [17, Remark 6.9] used the fact there Ω\partial\Omega is AR to estimate each NkN_{k}, while here we are invoking the geometric doubling property of the ambient space n+1\mathbb{R}^{n+1}. ∎

We are now ready to state the first main result of this chapter which is a version of [17, Lemma 6.15] (see also [10]) valid in our setting:

Lemma 3.5 (Discrete sawtooth lemma for projections).

Suppose that Ωn+1\Omega\subset\mathbb{R}^{n+1}, n2n\geq 2, is a bounded 1-sided NTA domain satisfying the CDC. Let Q0𝔻Q_{0}\in\mathbb{D}, let ={Qi}𝔻Q0\mathcal{F}=\{Q_{i}\}\subset\mathbb{D}_{Q_{0}} be a family of pairwise disjoint dyadic cubes, and let μ\mu be a dyadically doubling measure in Q0Q_{0}. Given two real (non-necessarily symmetric) elliptic L0L_{0}, LL, we write ω0YQ0=ωL0,ΩYQ0\omega_{0}^{Y_{Q_{0}}}=\omega_{L_{0},\Omega}^{Y_{Q_{0}}}, ωLYQ0=ωL,ΩYQ0\omega_{L}^{Y_{Q_{0}}}=\omega_{L,\Omega}^{Y_{Q_{0}}} for the elliptic measures associated with L0L_{0} and LL for the domain Ω\Omega with fixed pole at YQ0Ω,Q0ΩY_{Q_{0}}\in\Omega_{\mathcal{F},Q_{0}}\cap\Omega (cf. Lemma 3.1). Let ωL,YQ0=ωL,Ω,Q0YQ0\omega_{L,*}^{Y_{Q_{0}}}=\omega_{L,\Omega_{\mathcal{F},Q_{0}}}^{Y_{Q_{0}}} be the elliptic measure associated with LL for the domain Ω,Q0\Omega_{\mathcal{F},Q_{0}} with fixed pole at YQ0Ω,Q0ΩY_{Q_{0}}\in\Omega_{\mathcal{F},Q_{0}}\cap\Omega. Consider νLYQ0\nu_{L}^{Y_{Q_{0}}} the measure defined by

(3.6) νLYQ0(F)=ωL,YQ0(FQiQi)+QiωLYQ0(FQi)ωLYQ0(Qi)ωL,YQ0(Pi),FQ0,\nu_{L}^{Y_{Q_{0}}}(F)=\omega_{L,*}^{Y_{Q_{0}}}\Big{(}F\setminus\bigcup_{Q_{i}\in\mathcal{F}}Q_{i}\Big{)}+\sum_{Q_{i}\in\mathcal{F}}\frac{\omega_{L}^{Y_{Q_{0}}}(F\cap Q_{i})}{\omega_{L}^{Y_{Q_{0}}}(Q_{i})}\omega_{L,*}^{Y_{Q_{0}}}(P_{i}),\qquad F\subset Q_{0},

where PiP_{i} is the cube produced in Proposition 3.1. Then 𝒫μνLYQ0\mathcal{P}_{\mathcal{F}}^{\mu}\nu_{L}^{Y_{Q_{0}}} (see (2.22)) depends only on ω0YQ0\omega_{0}^{Y_{Q_{0}}} and ωL,YQ0\omega_{L,*}^{Y_{Q_{0}}}, but not on ωLYQ0\omega_{L}^{Y_{Q_{0}}}. More precisely,

(3.7) 𝒫μνLYQ0(F)=ωL,YQ0(FQiQi)+Qiμ(FQi)μ(Qi)ωL,YQ0(Pi),FQ0.\mathcal{P}_{\mathcal{F}}^{\mu}\nu_{L}^{Y_{Q_{0}}}(F)=\omega_{L,*}^{Y_{Q_{0}}}\Big{(}F\setminus\bigcup_{Q_{i}\in\mathcal{F}}Q_{i}\Big{)}+\sum_{Q_{i}\in\mathcal{F}}\frac{\mu(F\cap Q_{i})}{\mu(Q_{i})}\omega_{L,*}^{Y_{Q_{0}}}(P_{i}),\qquad F\subset Q_{0}.

Moreover, there exists θ>0\theta>0 such that for all Q𝔻Q0Q\in\mathbb{D}_{Q_{0}} and all FQF\subset Q, we have

(3.8) (𝒫μωLYQ0(F)𝒫μωLYQ0(Q))θ𝒫μνLYQ0(F)𝒫μνLYQ0(Q)𝒫μωLYQ0(F)𝒫μωLYQ0(Q).\bigg{(}\frac{\mathcal{P}_{\mathcal{F}}^{\mu}\omega_{L}^{Y_{Q_{0}}}(F)}{\mathcal{P}_{\mathcal{F}}^{\mu}\omega_{L}^{Y_{Q_{0}}}(Q)}\bigg{)}^{\theta}\lesssim\frac{\mathcal{P}_{\mathcal{F}}^{\mu}\nu_{L}^{Y_{Q_{0}}}(F)}{\mathcal{P}_{\mathcal{F}}^{\mu}\nu_{L}^{Y_{Q_{0}}}(Q)}\lesssim\frac{\mathcal{P}_{\mathcal{F}}^{\mu}\omega_{L}^{Y_{Q_{0}}}(F)}{\mathcal{P}_{\mathcal{F}}^{\mu}\omega_{L}^{Y_{Q_{0}}}(Q)}.
Proof.

Our argument follows the ideas from [17, Lemma 6.15] and we use several auxiliary technical results from [17, Section 6] which were proved under the additional assumption that Ω\partial\Omega is AR. However, as we will indicate along the proof, most of them can be adapted to our setting. Those arguments that require new ideas will be explained in detail.

We first observe that (3.7) readily follows from the definitions of 𝒫μ\mathcal{P}_{\mathcal{F}}^{\mu} and νLYQ0\nu_{L}^{Y_{Q_{0}}}. We first establish the second estimate in (3.8). With this goal in mind let us fix Q𝔻Q0Q\in\mathbb{D}_{Q_{0}} and FQ0F\subset Q_{0}.

Case 1: There exists QiQ_{i}\in\mathcal{F} such that QQiQ\subset Q_{i}. By (3.7) we have

𝒫μνLYQ0(F)𝒫μνLYQ0(Q)=μ(FQi)μ(Qi)ωL,YQ0(Pi)μ(QQi)μ(Qi)ωL,YQ0(Pi)=μ(F)μ(Q)=μ(FQi)μ(Qi)ωLYQ0(Qi)μ(QQi)μ(Qi)ωLYQ0(Qi)=𝒫μωLYQ0(F)𝒫μωLYQ0(Q).\frac{\mathcal{P}_{\mathcal{F}}^{\mu}\nu_{L}^{Y_{Q_{0}}}(F)}{\mathcal{P}_{\mathcal{F}}^{\mu}\nu_{L}^{Y_{Q_{0}}}(Q)}=\frac{\frac{\mu(F\cap Q_{i})}{\mu(Q_{i})}\omega_{L,*}^{Y_{Q_{0}}}(P_{i})}{\frac{\mu(Q\cap Q_{i})}{\mu(Q_{i})}\omega_{L,*}^{Y_{Q_{0}}}(P_{i})}=\frac{\mu(F)}{\mu(Q)}=\frac{\frac{\mu(F\cap Q_{i})}{\mu(Q_{i})}\omega_{L}^{Y_{Q_{0}}}(Q_{i})}{\frac{\mu(Q\cap Q_{i})}{\mu(Q_{i})}\omega_{L}^{Y_{Q_{0}}}(Q_{i})}=\frac{\mathcal{P}_{\mathcal{F}}^{\mu}\omega_{L}^{Y_{Q_{0}}}(F)}{\mathcal{P}_{\mathcal{F}}^{\mu}\omega_{L}^{Y_{Q_{0}}}(Q)}.

Case 2: QQiQ\not\subset Q_{i} for any QiQ_{i}\in\mathcal{F}, that is, Q𝔻,Q0Q\in\mathbb{D}_{\mathcal{F},Q_{0}}. In particular if QQiØQ\cap Q_{i}\neq\mbox{{\O}} with QiQ_{i}\in\mathcal{F} then necessarily QiQQ_{i}\subsetneq Q. Let xix_{i}^{\star} denote the center of PiP_{i} and pick ri(Qi)(Pi)r_{i}\approx\ell(Q_{i})\approx\ell(P_{i}) so that PiΔ(xi,ri):=B(xi,ri)Ω,Q0P_{i}\subset\Delta_{\star}(x_{i}^{\star},r_{i}):=B(x_{i}^{\star},r_{i})\cap\partial\Omega_{\mathcal{F},Q_{0}}. Note that by Proposition 2.37, Harnack’s inequality and Lemma 2.53 parts (a)(a) and (c)(c) we have that ωL,YQ0(Pi)ωL,YQ0(Δ(xi,ri))\omega_{L,*}^{Y_{Q_{0}}}(P_{i})\approx\omega_{L,*}^{Y_{Q_{0}}}(\Delta_{\star}(x_{i}^{\star},r_{i})). On the other hand as in [17, Proposition 6.12] one can see that

(3.9) ΔQ:=B(xQ,tQ)Ω,Q0(QQiQi)(Qi:QiQΔ(xi,ri))\Delta_{\star}^{Q}:=B(x_{Q}^{\star},t_{Q})\cap\partial\Omega_{\mathcal{F},Q_{0}}\subset\Big{(}Q\setminus\bigcup_{Q_{i}\in\mathcal{F}}Q_{i}\Big{)}\bigcup\Big{(}\bigcup_{Q_{i}\in\mathcal{F}:Q_{i}\subsetneq Q}\Delta_{\star}(x_{i}^{\star},r_{i})\Big{)}

with tQ(Q)t_{Q}\approx\ell(Q), xQΩ,Q0x_{Q}^{\star}\in\partial\Omega_{\mathcal{F},Q_{0}} and dist(Q,ΔQ)(Q)\operatorname{dist}(Q,\Delta_{\star}^{Q})\lesssim\ell(Q) with implicit constants depending on the allowable parameters. We note that the last expression is slightly different to that in [17, Proposition 6.2], nonetheless the one stated here follows from the proof in account of [17, (6.14) and Proposition 6.1] as Qi\partial Q_{i} is contained in TQi¯\overline{T_{Q_{i}}}. Besides, Proposition 3.1 easily yields

(3.10) (QQiQi)(Qi:QiQPi)(QQiQi)(Qi:QiQΔ(xi,ri))CΔQ,\Big{(}Q\setminus\bigcup_{Q_{i}\in\mathcal{F}}Q_{i}\Big{)}\bigcup\Big{(}\bigcup_{Q_{i}\in\mathcal{F}:Q_{i}\subsetneq Q}P_{i}\Big{)}\subset\Big{(}Q\setminus\bigcup_{Q_{i}\in\mathcal{F}}Q_{i}\Big{)}\bigcup\Big{(}\bigcup_{Q_{i}\in\mathcal{F}:Q_{i}\subsetneq Q}\Delta_{\star}(x_{i}^{\star},r_{i})\Big{)}\subset C\Delta_{\star}^{Q},

hence

(3.11) ωL,YQ0((QQiQi)(Qi:QiQΔ(xi,ri)))ωL,YQ0(ΔQ).\omega_{L,*}^{Y_{Q_{0}}}\Big{(}\Big{(}Q\setminus\bigcup_{Q_{i}\in\mathcal{F}}Q_{i}\Big{)}\bigcup\Big{(}\bigcup_{Q_{i}\in\mathcal{F}:Q_{i}\subsetneq Q}\Delta_{\star}(x_{i}^{\star},r_{i})\Big{)}\Big{)}\lesssim\omega_{L,*}^{Y_{Q_{0}}}(\Delta_{\star}^{Q}).

Writing E0=Q0QiQiΩΩ,QE_{0}=Q_{0}\setminus\bigcup_{Q_{i}\in\mathcal{F}}Q_{i}\subset\partial\Omega\cap\partial\Omega_{\mathcal{F},Q} (see [17, Proposition 6.1]) we have

(3.12) ωL,YQ0(ΔQ)ωL,YQ0(QE0)+Qi:QiQωL,YQ0(Δ(xi,ri))ωL,YQ0(QE0)+Qi:QiQωL,YQ0(Pi)=𝒫μνLYQ0(Q)\omega_{L,*}^{Y_{Q_{0}}}(\Delta_{\star}^{Q})\leq\omega_{L,*}^{Y_{Q_{0}}}(Q\cap E_{0})+\sum_{Q_{i}\in\mathcal{F}:Q_{i}\subsetneq Q}\omega_{L,*}^{Y_{Q_{0}}}(\Delta_{\star}(x_{i}^{\star},r_{i}))\\ \lesssim\omega_{L,*}^{Y_{Q_{0}}}(Q\cap E_{0})+\sum_{Q_{i}\in\mathcal{F}:Q_{i}\subsetneq Q}\omega_{L,*}^{Y_{Q_{0}}}(P_{i})=\mathcal{P}_{\mathcal{F}}^{\mu}\nu_{L}^{Y_{Q_{0}}}(Q)

and, by (3.4),

(3.13) 𝒫μνLYQ0(Q)=ωL,YQ0(QE0)+Qi:QiQμ(QQi)μ(Qi)ωL,YQ0(Pi)=ωL,YQ0(QE0)+Qi:QiQωL,YQ0(Pi)ωL,YQ0((QE0)(Qi:QiQPi))ωL,YQ0(ΔQ).\mathcal{P}_{\mathcal{F}}^{\mu}\nu_{L}^{Y_{Q_{0}}}(Q)=\omega_{L,*}^{Y_{Q_{0}}}(Q\cap E_{0})+\sum_{Q_{i}\in\mathcal{F}:Q_{i}\subsetneq Q}\frac{\mu(Q\cap Q_{i})}{\mu(Q_{i})}\omega_{L,*}^{Y_{Q_{0}}}(P_{i})\\ =\omega_{L,*}^{Y_{Q_{0}}}(Q\cap E_{0})+\sum_{Q_{i}\in\mathcal{F}:Q_{i}\subsetneq Q}\omega_{L,*}^{Y_{Q_{0}}}(P_{i})\\ \lesssim\omega_{L,*}^{Y_{Q_{0}}}\Big{(}(Q\cap E_{0})\bigcup\Big{(}\bigcup_{Q_{i}\in\mathcal{F}:Q_{i}\subsetneq Q}P_{i}\Big{)}\Big{)}\lesssim\omega_{L,*}^{Y_{Q_{0}}}(\Delta_{\star}^{Q}).

Since Q𝔻,Q0Q\in\mathbb{D}_{\mathcal{F},Q_{0}} we can invoke [17, Proposition 6.4] (which also holds in the current setting) to find YQΩ,Q0Y_{Q}\in\Omega_{\mathcal{F},Q_{0}} which serves as a Corkscrew point simultaneously for Ω,Q0\Omega_{\mathcal{F},Q_{0}} with respect to the surface ball Δ(yQ,sQ)\Delta_{\star}(y_{Q},s_{Q}) for some yQΩ,Qy_{Q}\in\Omega_{\mathcal{F},Q} and some sQ(Q)s_{Q}\approx\ell(Q), and for Ω\Omega with respect to each surface ball Δ(x,sQ)\Delta(x,s_{Q}), for every xQx\in Q. Applying (2.55) and Harnack’s inequality to join YQY_{Q} with XQX_{Q} and YQ0Y_{Q_{0}} with YQY_{Q} we have

(3.14) dωLYQdωLYQ01ωLYQ0(Q),ωLYQ0-a.e. in Q.\frac{d\omega_{L}^{Y_{Q}}}{d\omega_{L}^{Y_{Q_{0}}}}\approx\frac{1}{\omega_{L}^{Y_{Q_{0}}}(Q)},\qquad\mbox{$\omega_{L}^{Y_{Q_{0}}}$-a.e. in }Q.

On the other hand one can see that

(3.15) B~Q(Qi:QiQB(xi,ri))B(yQ,s^Q),\widetilde{B}_{Q}\bigcup\Big{(}\bigcup_{Q_{i}\in\mathcal{F}:Q_{i}\subsetneq Q}B(x_{i}^{\star},r_{i})\Big{)}\subset B(y_{Q},\widehat{s}_{Q}),

for some s^QsQ\widehat{s}_{Q}\approx s_{Q}. Invoking then Proposition 2.37, and Lemma 2.53 parts (c)(c) and (d)(d) in the domain Ω,Q0\Omega_{\mathcal{F},Q_{0}} we can analogously see

(3.16) dωL,YQdωL,YQ01ωL,YQ0(Δ(yQ,s^Q))1ωL,YQ0(ΔQ),ωL,YQ0-a.e. in Δ(yQ,s^Q).\frac{d\omega_{L,*}^{Y_{Q}}}{d\omega_{L,*}^{Y_{Q_{0}}}}\approx\frac{1}{\omega_{L,*}^{Y_{Q_{0}}}(\Delta(y_{Q},\widehat{s}_{Q}))}\approx\frac{1}{\omega_{L,*}^{Y_{Q_{0}}}(\Delta^{Q}_{\star})},\qquad\mbox{$\omega_{L,*}^{Y_{Q_{0}}}$-a.e. in }\Delta(y_{Q},\widehat{s}_{Q}).

Next we invoke (3.12), (3.15), and (3.14) to obtain

(3.17) 𝒫μνLYQ0(F)𝒫μνLYQ0(Q)ωL,YQ0(FE0)ωL,YQ0(ΔQ)+Qi:QiQμ(FQi)μ(Qi)ωL,YQ0(Pi)ωL,YQ0(ΔQ)ωL,YQ(FE0)+Qi:QiQμ(FQi)μ(Qi)ωL,YQ(Pi).\frac{\mathcal{P}_{\mathcal{F}}^{\mu}\nu_{L}^{Y_{Q_{0}}}(F)}{\mathcal{P}_{\mathcal{F}}^{\mu}\nu_{L}^{Y_{Q_{0}}}(Q)}\approx\frac{\omega_{L,*}^{Y_{Q_{0}}}(F\cap E_{0})}{\omega_{L,*}^{Y_{Q_{0}}}(\Delta_{\star}^{Q})}+\sum_{Q_{i}\in\mathcal{F}:Q_{i}\subsetneq Q}\frac{\mu(F\cap Q_{i})}{\mu(Q_{i})}\frac{\omega_{L,*}^{Y_{Q_{0}}}(P_{i})}{\omega_{L,*}^{Y_{Q_{0}}}(\Delta_{\star}^{Q})}\\ \approx\omega_{L,*}^{Y_{Q}}(F\cap E_{0})+\sum_{Q_{i}\in\mathcal{F}:Q_{i}\subsetneq Q}\frac{\mu(F\cap Q_{i})}{\mu(Q_{i})}\omega_{L,*}^{Y_{Q}}(P_{i}).

We claim the following estimates hold

(3.18) ωL,YQ(FE0)ωLYQ(FE0),ωL,YQ(Pi)ωLYQ(Qi).\omega_{L,*}^{Y_{Q}}(F\cap E_{0})\lesssim\omega_{L}^{Y_{Q}}(F\cap E_{0}),\qquad\omega_{L,*}^{Y_{Q}}(P_{i})\lesssim\omega_{L}^{Y_{Q}}(Q_{i}).

The first estimate follows easily from the maximum principle since Ω,Q0Ω\Omega_{\mathcal{F},Q_{0}}\subset\Omega and FE0ΩΩ,Q0F\cap E_{0}\subset\partial\Omega\cap\partial\Omega_{\mathcal{F},Q_{0}}. For the second one, by the maximum principle we just need to see that ωLX(Qi)1\omega_{L}^{X}(Q_{i})\gtrsim 1 for XPiX\in P_{i}, but this follows from Lemma 2.53 part (a)(a), (2.14), Harnack’s inequality, and (3.3).

With the previous estimates at our disposal we can the continue with our estimate (3.17):

𝒫μνLYQ0(F)𝒫μνLYQ0(Q)ωLYQ(FE0)+Qi:QiQμ(FQi)μ(Qi)ωLYQ(Qi)ωLYQ0(FE0)ωLYQ0(Q)+Qi:QiQμ(FQi)μ(Qi)ωLYQ0(Qi)ωLYQ0(Q)=𝒫μωLYQ0(F)ωLYQ0(Q)=𝒫μωLYQ0(F)𝒫μωLYQ0(Q),\frac{\mathcal{P}_{\mathcal{F}}^{\mu}\nu_{L}^{Y_{Q_{0}}}(F)}{\mathcal{P}_{\mathcal{F}}^{\mu}\nu_{L}^{Y_{Q_{0}}}(Q)}\lesssim\omega_{L}^{Y_{Q}}(F\cap E_{0})+\sum_{Q_{i}\in\mathcal{F}:Q_{i}\subsetneq Q}\frac{\mu(F\cap Q_{i})}{\mu(Q_{i})}\omega_{L}^{Y_{Q}}(Q_{i})\\ \approx\frac{\omega_{L}^{Y_{Q_{0}}}(F\cap E_{0})}{\omega_{L}^{Y_{Q_{0}}}(Q)}+\sum_{Q_{i}\in\mathcal{F}:Q_{i}\subsetneq Q}\frac{\mu(F\cap Q_{i})}{\mu(Q_{i})}\frac{\omega_{L}^{Y_{Q_{0}}}(Q_{i})}{\omega_{L}^{Y_{Q_{0}}}(Q)}\\ =\frac{\mathcal{P}_{\mathcal{F}}^{\mu}\omega_{L}^{Y_{Q_{0}}}(F)}{\omega_{L}^{Y_{Q_{0}}}(Q)}=\frac{\mathcal{P}_{\mathcal{F}}^{\mu}\omega_{L}^{Y_{Q_{0}}}(F)}{\mathcal{P}_{\mathcal{F}}^{\mu}\omega_{L}^{Y_{Q_{0}}}(Q)},

where we have used (3.15) and that 𝒫μωLYQ0(Q)=ωLYQ0(Q){\mathcal{P}_{\mathcal{F}}^{\mu}\omega_{L}^{Y_{Q_{0}}}(Q)}=\omega_{L}^{Y_{Q_{0}}}(Q). This proves the second estimate in (3.8) in the current case.

Once we have shown the second estimate in (3.8) we can invoke [17, Lemma B.7] (which is a purely dyadic result and hence applies in our setting) along with Lemma 3.22 below to eventually obtain the first estimate in (3.8). ∎

As a consequence of the previous result we can easily obtain a dyadic analog of the main lemma in [10].

Lemma 3.19 (Discrete sawtooth lemma).

Suppose that Ωn+1\Omega\subset\mathbb{R}^{n+1}, n2n\geq 2, is a bounded 1-sided NTA domain satisfying the CDC. Let Q0𝔻Q_{0}\in\mathbb{D} and let ={Qi}𝔻Q0\mathcal{F}=\{Q_{i}\}\subset\mathbb{D}_{Q_{0}} be a family of pairwise disjoint dyadic cubes. Given two real (non-necessarily symmetric) elliptic L0L_{0}, LL, we write ω0YQ0=ωL0,ΩYQ0\omega_{0}^{Y_{Q_{0}}}=\omega_{L_{0},\Omega}^{Y_{Q_{0}}}, ωLYQ0=ωL,ΩYQ0\omega_{L}^{Y_{Q_{0}}}=\omega_{L,\Omega}^{Y_{Q_{0}}} for the elliptic measures associated with L0L_{0} and LL for the domain Ω\Omega with fixed pole at YQ0Ω,Q0ΩY_{Q_{0}}\in\Omega_{\mathcal{F},Q_{0}}\cap\Omega (cf. Lemma 3.1). Let ωL,YQ0=ωL,Ω,Q0YQ0\omega_{L,*}^{Y_{Q_{0}}}=\omega_{L,\Omega_{\mathcal{F},Q_{0}}}^{Y_{Q_{0}}} be the elliptic measure associated with LL for the domain Ω,Q0\Omega_{\mathcal{F},Q_{0}} with fixed pole at YQ0Ω,Q0ΩY_{Q_{0}}\in\Omega_{\mathcal{F},Q_{0}}\cap\Omega. Consider νLYQ0\nu_{L}^{Y_{Q_{0}}} the measure defined by (3.6). Then, there exists θ>0\theta>0 such that for all Q𝔻Q0Q\in\mathbb{D}_{Q_{0}} and all FQF\subset Q, we have

(3.20) (ωLYQ0(F)ωLYQ0(Q))θνLYQ0(F)νLYQ0(Q)ωLYQ0(F)ωLYQ0(Q).\bigg{(}\frac{\omega_{L}^{Y_{Q_{0}}}(F)}{\omega_{L}^{Y_{Q_{0}}}(Q)}\bigg{)}^{\theta}\lesssim\frac{\nu_{L}^{Y_{Q_{0}}}(F)}{\nu_{L}^{Y_{Q_{0}}}(Q)}\lesssim\frac{\omega_{L}^{Y_{Q_{0}}}(F)}{\omega_{L}^{Y_{Q_{0}}}(Q)}.

In particular, if FQQiQiF\subset Q\setminus\bigcup_{Q_{i}\in\mathcal{F}}Q_{i}, we have

(3.21) (ωLYQ0(F)ωLYQ0(Q))θωL,YQ0(F)ωL,YQ0(ΔQ)ωLYQ0(F)ωLYQ0(Q),\bigg{(}\frac{\omega_{L}^{Y_{Q_{0}}}(F)}{\omega_{L}^{Y_{Q_{0}}}(Q)}\bigg{)}^{\theta}\lesssim\frac{\omega_{L,*}^{Y_{Q_{0}}}(F)}{\omega_{L,*}^{Y_{Q_{0}}}(\Delta_{\star}^{Q})}\lesssim\frac{\omega_{L}^{Y_{Q_{0}}}(F)}{\omega_{L}^{Y_{Q_{0}}}(Q)},

where ΔQ:=B(xQ,tQ)Ω,Q0\Delta_{\star}^{Q}:=B(x_{Q}^{\star},t_{Q})\cap\partial\Omega_{\mathcal{F},Q_{0}} with tQ(Q)t_{Q}\approx\ell(Q), xQΩ,Q0x_{Q}^{\star}\in\partial\Omega_{\mathcal{F},Q_{0}}, and dist(Q,ΔQ)(Q)\operatorname{dist}(Q,\Delta_{\star}^{Q})\lesssim\ell(Q) with implicit constants depending on the allowable parameters (cf. [17, Proposition 6.12]).

Proof.

Letting μ=ωLYQ0\mu=\omega_{L}^{Y_{Q_{0}}}, which is dyadically doubling in Q0Q_{0}, one easily has 𝒫μωLYQ0=ωLYQ0\mathcal{P}_{\mathcal{F}}^{\mu}\omega_{L}^{Y_{Q_{0}}}=\omega_{L}^{Y_{Q_{0}}} and 𝒫μνLYQ0=νLYQ0\mathcal{P}_{\mathcal{F}}^{\mu}\nu_{L}^{Y_{Q_{0}}}=\nu_{L}^{Y_{Q_{0}}}. Thus (3.8) in Lemma 3.5 readily yields (3.20). Next, to obtain (3.21) we may assume that FF is non-empty. Observe that if FQQiQiF\subset Q\setminus\bigcup_{Q_{i}\in\mathcal{F}}Q_{i}, then νLYQ0(F)=ωL,YQ0(F)\nu_{L}^{Y_{Q_{0}}}(F)=\omega_{L,*}^{Y_{Q_{0}}}(F). On the other hand, if FQQiQiF\subset Q\setminus\bigcup_{Q_{i}\in\mathcal{F}}Q_{i} we must be in Case 2 of the proof of Lemma 3.5, hence (3.12) and (3.13) hold. With all these we readily obtain (3.21). ∎

Our last result in this section establishes that both νLYQ0\nu_{L}^{Y_{Q_{0}}} and 𝒫μνLYQ0\mathcal{P}_{\mathcal{F}}^{\mu}\nu_{L}^{Y_{Q_{0}}} are dyadically doubling on Q0Q_{0}.

Lemma 3.22.

Under the assumptions of Lemma 3.5, νLYQ0\nu_{L}^{Y_{Q_{0}}} and 𝒫μνLYQ0\mathcal{P}_{\mathcal{F}}^{\mu}\nu_{L}^{Y_{Q_{0}}} are dyadically doubling on Q0Q_{0}.

Proof.

We follow the ideas in [17, Lemma B.2]. We shall first see νLYQ0\nu_{L}^{Y_{Q_{0}}} is dyadically doubling. To this end, let Q𝔻Q0Q\in\mathbb{D}_{Q_{0}} be fixed and let QQ^{\prime} be one of its dyadic children. We consider three cases:

Case 1: There exists QiQ_{i}\in\mathcal{F} such that QQiQ\subset Q_{i}. In this case we have

νLYQ0(Q)=ωLYQ0(Q)ωLYQ0(Qi)ωL,YQ0(Pi)ωLYQ0(Q)ωLYQ0(Qi)ωL,YQ0(Pi)=νLYQ0(Q)\displaystyle\nu_{L}^{Y_{Q_{0}}}(Q)=\frac{\omega_{L}^{Y_{Q_{0}}}(Q)}{\omega_{L}^{Y_{Q_{0}}}(Q_{i})}\omega_{L,*}^{Y_{Q_{0}}}(P_{i})\lesssim\frac{\omega_{L}^{Y_{Q_{0}}}(Q^{\prime})}{\omega_{L}^{Y_{Q_{0}}}(Q_{i})}\omega_{L,*}^{Y_{Q_{0}}}(P_{i})=\nu_{L}^{Y_{Q_{0}}}(Q^{\prime})

where we have used Harnack’s inequality and Lemma 2.53 parts and (a)(a) and (c)(c).

Case 2: QQ^{\prime}\in\mathcal{F}. For simplicity say Q=Q1Q^{\prime}=Q_{1}\in\mathcal{F} and in this case νLYQ0(Q)=ωL,YQ0(P1)\nu_{L}^{Y_{Q_{0}}}(Q^{\prime})=\omega_{L,*}^{Y_{Q_{0}}}(P_{1}). Note that then Q𝔻,Q0Q\in\mathbb{D}_{\mathcal{F},Q_{0}} and we let 1\mathcal{F}_{1} be the family of cubes QiQ_{i}\in\mathcal{F} with QiQØQ_{i}\cap Q\neq\mbox{{\O}} and observe that if Qi1Q_{i}\in\mathcal{F}_{1} then QiQQ_{i}\subsetneq Q. Then by (3.4)

(3.23) νLYQ0(Q)=ωL,YQ0(QQiQi)+Qi1ωLYQ0(QQi)ωLYQ0(Qi)ωL,YQ0(Pi)=ωL,YQ0(QQiQi)+Qi1ωL,YQ0(Pi)ωL,YQ0((QQiQi)(Qi1Pi)).\nu_{L}^{Y_{Q_{0}}}(Q)=\omega_{L,*}^{Y_{Q_{0}}}\Big{(}Q\setminus\bigcup_{Q_{i}\in\mathcal{F}}Q_{i}\Big{)}+\sum_{Q_{i}\in\mathcal{F}_{1}}\frac{\omega_{L}^{Y_{Q_{0}}}(Q\cap Q_{i})}{\omega_{L}^{Y_{Q_{0}}}(Q_{i})}\omega_{L,*}^{Y_{Q_{0}}}(P_{i})\\ =\omega_{L,*}^{Y_{Q_{0}}}\Big{(}Q\setminus\bigcup_{Q_{i}\in\mathcal{F}}Q_{i}\Big{)}+\sum_{Q_{i}\in\mathcal{F}_{1}}\omega_{L,*}^{Y_{Q_{0}}}(P_{i})\lesssim\omega_{L,*}^{Y_{Q_{0}}}\Big{(}\Big{(}Q\setminus\bigcup_{Q_{i}\in\mathcal{F}}Q_{i}\Big{)}\bigcup\Big{(}\bigcup_{Q_{i}\in\mathcal{F}_{1}}P_{i}\Big{)}\Big{)}.

Recall that in Case 2 in the proof of Lemma 3.5 we mentioned that P1Δ(x1,r1)P_{1}\subset\Delta_{\star}(x_{1}^{\star},r_{1}) with x1x_{1}^{\star} being the center of P1P_{1} and r1(P1)(Q1)(Q)r_{1}\approx\ell(P_{1})\approx\ell(Q_{1})\approx\ell(Q) since QQ is the dyadic parent of Q1Q_{1}. Note that since Qi1Q_{i}\in\mathcal{F}_{1} by (3.3)

(Pi)dist(Pi,Q)(Qi)(Q)=2(Q1)(P1)dist(Q1,P1)r1.\ell(P_{i})\approx\operatorname{dist}(P_{i},Q)\approx\ell(Q_{i})\lesssim\ell(Q)=2\ell(Q_{1})\approx\ell(P_{1})\approx\operatorname{dist}(Q_{1},P_{1})\approx r_{1}.

Thus

(QQiQi)(Qi1Pi)Δ(x1,Cr1),\Big{(}Q\setminus\bigcup_{Q_{i}\in\mathcal{F}}Q_{i}\Big{)}\bigcup\Big{(}\bigcup_{Q_{i}\in\mathcal{F}_{1}}P_{i}\Big{)}\subset\Delta_{\star}(x_{1}^{\star},Cr_{1}),

where we here and below we use the notation Δ\Delta_{\star} for the surface balls with respect to Ω,Q0\partial\Omega_{\mathcal{F},Q_{0}}. Using this, (3.23), and Lemma 2.53 parts (a)(a) and (c)(c) and Harnack’s inequality we derive

νLYQ0(Q)ωL,YQ0(Δ(x1,Cr1))ωL,YQ0(Δ(x1,r1))ωL,YQ0(P1)=νLYQ0(Q).\nu_{L}^{Y_{Q_{0}}}(Q)\lesssim\omega_{L,*}^{Y_{Q_{0}}}(\Delta_{\star}(x_{1}^{\star},Cr_{1}))\lesssim\omega_{L,*}^{Y_{Q_{0}}}(\Delta_{\star}(x_{1}^{\star},r_{1}))\lesssim\omega_{L,*}^{Y_{Q_{0}}}(P_{1})=\nu_{L}^{Y_{Q_{0}}}(Q^{\prime}).

Case 3: None of the conditions in the previous cases happen, and necessarily Q,Q𝔻,Q0Q,Q^{\prime}\in\mathbb{D}_{\mathcal{F},Q_{0}}. We take the same set 1\mathcal{F}_{1} as in the previous case and again if Qi1Q_{i}\in\mathcal{F}_{1} then QiQQ_{i}\subsetneq Q (otherwise we are driven to Case 1). Introduce 2\mathcal{F}_{2}, the family of cubes QiQ_{i}\in\mathcal{F} with QiQØQ_{i}\cap Q^{\prime}\neq\mbox{{\O}}. Again, if Qi2Q_{i}\in\mathcal{F}_{2} we have QiQQ_{i}\subsetneq Q^{\prime}; otherwise either Q=QiQ^{\prime}=Q_{i} which is Case 2, or QQiQ^{\prime}\subsetneq Q_{i} which implies QQiQ\subset Q_{i} and we are back to Case 1.

Note that since QQ is the dyadic parent of QQ^{\prime}, using the same notation as in (3.9) applied to Q𝔻,Q0Q^{\prime}\in\mathbb{D}_{\mathcal{F},Q_{0}} we have that

dist(xQ,Q)dist(xQ,Q)(Q)(Q)tQ.\operatorname{dist}(x^{\star}_{Q^{\prime}},Q)\leq\operatorname{dist}(x^{\star}_{Q^{\prime}},Q^{\prime})\lesssim\ell(Q^{\prime})\approx\ell(Q)\approx t_{Q^{\prime}}.

Also by (3.3)

dist(xQ,Pi)dist(xQ,Q)+(Q)+dist(Q,Pi)(Q)+dist(Qi,Pi)(Q)tQ.\operatorname{dist}(x^{\star}_{Q^{\prime}},P_{i})\lesssim\operatorname{dist}(x^{\star}_{Q^{\prime}},Q)+\ell(Q)+\operatorname{dist}(Q,P_{i})\lesssim\ell(Q)+\operatorname{dist}(Q_{i},P_{i})\lesssim\ell(Q)\approx t_{Q^{\prime}}.

These readily give

(QQiQi)(Qi1Pi)Δ(xQ,CtQ).\Big{(}Q\setminus\bigcup_{Q_{i}\in\mathcal{F}}Q_{i}\Big{)}\bigcup\Big{(}\bigcup_{Q_{i}\in\mathcal{F}_{1}}P_{i}\Big{)}\subset\Delta_{\star}(x_{Q^{\prime}}^{\star},Ct_{Q^{\prime}}).

We can then proceed as in the previous case (see (3.23)) to obtain

νLYQ0(Q)ωL,YQ0((QQiQi)(Qi1Pi))ωL,YQ0(Δ(xQ,CtQ))ωL,YQ0(ΔQ)\displaystyle\nu_{L}^{Y_{Q_{0}}}(Q)\lesssim\omega_{L,*}^{Y_{Q_{0}}}\Big{(}\Big{(}Q\setminus\bigcup_{Q_{i}\in\mathcal{F}}Q_{i}\Big{)}\bigcup\Big{(}\bigcup_{Q_{i}\in\mathcal{F}_{1}}P_{i}\Big{)}\Big{)}\lesssim\omega_{L,*}^{Y_{Q_{0}}}(\Delta_{\star}(x_{Q^{\prime}}^{\star},Ct_{Q^{\prime}}))\lesssim\omega_{L,*}^{Y_{Q_{0}}}(\Delta_{\star}^{Q^{\prime}})

where ΔQ=B(xQ,tQ)Ω,Q0\Delta_{\star}^{Q^{\prime}}=B(x_{Q^{\prime}}^{\star},t_{Q^{\prime}})\cap\partial\Omega_{\mathcal{F},Q_{0}} (see (3.9)) and we have used Lemma 2.53 parts (a)(a) and (c)(c) and Harnack’s inequality. On the other hand, proceeding as in (3.12) with QQ^{\prime} in place of QQ since Q𝔻,Q0Q^{\prime}\in\mathbb{D}_{\mathcal{F},Q_{0}}:

ωL,YQ0(ΔQ)ωL,YQ0(QE0)+Qi2ωL,YQ0(Δ(xi,ri))ωL,YQ0(QE0)+Qi2ωL,YQ0(Pi)=ωL,YQ0(QE0)+Qi2ωLYQ0(QQi)ωLYQ0(Qi)ωL,YQ0(Pi)=νLYQ0(Q).\omega_{L,*}^{Y_{Q_{0}}}(\Delta_{\star}^{Q^{\prime}})\leq\omega_{L,*}^{Y_{Q_{0}}}(Q^{\prime}\cap E_{0})+\sum_{Q_{i}\in\mathcal{F}_{2}}\omega_{L,*}^{Y_{Q_{0}}}(\Delta_{\star}(x_{i}^{\star},r_{i}))\\ \lesssim\omega_{L,*}^{Y_{Q_{0}}}(Q^{\prime}\cap E_{0})+\sum_{Q_{i}\in\mathcal{F}_{2}}\omega_{L,*}^{Y_{Q_{0}}}(P_{i})\\ =\omega_{L,*}^{Y_{Q_{0}}}(Q^{\prime}\cap E_{0})+\sum_{Q_{i}\in\mathcal{F}_{2}}\frac{\omega_{L}^{Y_{Q_{0}}}(Q^{\prime}\cap Q_{i})}{\omega_{L}^{Y_{Q_{0}}}(Q_{i})}\omega_{L,*}^{Y_{Q_{0}}}(P_{i})=\nu_{L}^{Y_{Q_{0}}}(Q^{\prime}).

Eventually we obtain that νLYQ0(Q)νLYQ0(Q)\nu_{L}^{Y_{Q_{0}}}(Q)\lesssim\nu_{L}^{Y_{Q_{0}}}(Q^{\prime}), completing the proof of the dyadic doubling property of νLYQ0\nu_{L}^{Y_{Q_{0}}}.

We next deal with 𝒫μνLYQ0\mathcal{P}_{\mathcal{F}}^{\mu}\nu_{L}^{Y_{Q_{0}}}. We can simply follow the previous argument replacing ωLYQ0\omega_{L}^{Y_{Q_{0}}} by 𝒫μνLYQ0\mathcal{P}_{\mathcal{F}}^{\mu}\nu_{L}^{Y_{Q_{0}}} to see that in Cases 2 and 3 we have that 𝒫μνLYQ0(Q)=νLYQ0(Q)\mathcal{P}_{\mathcal{F}}^{\mu}\nu_{L}^{Y_{Q_{0}}}(Q)=\nu_{L}^{Y_{Q_{0}}}(Q) and 𝒫μνLYQ0(Q)=νLYQ0(Q)\mathcal{P}_{\mathcal{F}}^{\mu}\nu_{L}^{Y_{Q_{0}}}(Q^{\prime})=\nu_{L}^{Y_{Q_{0}}}(Q^{\prime}), hence the doubling condition follows from the previous calculations and the constant depend on that of ωL,YQ0\omega_{L,*}^{Y_{Q_{0}}}. With regard to Cases 1, on which QQiQ\subset Q_{i} for some QiQ_{i}\in\mathcal{F}, one can easily see that

𝒫μνLYQ0(Q)=μ(Q)μ(Qi)ωL,YQ0(Pi)μ(Q)μ(Qi)ωL,YQ0(Pi)=𝒫μνLYQ0(Q),\displaystyle\mathcal{P}_{\mathcal{F}}^{\mu}\nu_{L}^{Y_{Q_{0}}}(Q)=\frac{\mu(Q)}{\mu(Q_{i})}\omega_{L,*}^{Y_{Q_{0}}}(P_{i})\lesssim\frac{\mu(Q^{\prime})}{\mu(Q_{i})}\omega_{L,*}^{Y_{Q_{0}}}(P_{i})=\mathcal{P}_{\mathcal{F}}^{\mu}\nu_{L}^{Y_{Q_{0}}}(Q^{\prime}),

which uses that μ\mu is dyadically doubling in Q0Q_{0}. Eventually we have seen that doubling constant depend on that of ωL,YQ0\omega_{L,*}^{Y_{Q_{0}}} and μ\mu as desired. This completes the proof. ∎

4. Proof of the main results

4.1. Proof of Theorem 1.3

By renormalization we may assume without loss of generality that uL(Ω)=1\|u\|_{L^{\infty}(\Omega)}=1. We will first prove a dyadic version of (1.4). Let 𝔻=𝔻(Ω)\mathbb{D}=\mathbb{D}(\partial\Omega) the dyadic grid from Lemma 2.13 with E=ΩE=\partial\Omega. Our goal is to show that

(4.1) M0:=supQ0𝔻supQ0𝔻Q0(Q0)(Q0)M1ωLXQ0(Q0)TQ0|u(X)|2GL(XQ0,X)𝑑X1M_{0}:=\sup_{Q^{0}\in\mathbb{D}}\sup_{\begin{subarray}{c}Q_{0}\in\mathbb{D}_{Q^{0}}\\ \ell(Q_{0})\leq\frac{\ell(Q^{0})}{M}\end{subarray}}\frac{1}{\omega_{L}^{X_{Q^{0}}}(Q_{0})}\iint_{T_{Q_{0}}}|\nabla u(X)|^{2}\,G_{L}(X_{Q^{0}},X)\,dX\lesssim 1

with M4M\geq 4 large enough. Assuming this momentarily let us see how to derive (1.4). Fix BB and BB^{\prime} as in the suprema in (1.4). Let k,kk,k^{\prime}\in{\mathbb{Z}} be so that 2k1<r2k2^{k-1}<r\leq 2^{k} and 2k1<r2k2^{k^{\prime}-1}<r^{\prime}\leq 2^{k^{\prime}}, and define k′′:=min{k,k10kM}k^{\prime\prime}:=\min\{k^{\prime},k-10k_{M}\} where kM1k_{M}\geq 1 is large enough to be chosen depending on MM and the allowable parameters. Set

𝒲:={I𝒲:IBØ,(I)<2k′′}{I𝒲:IBØ,(I)2k′′}=:𝒲1𝒲2.\mathcal{W}^{\prime}:=\{I\in\mathcal{W}:I\cap B^{\prime}\neq\mbox{{\O}},\ell(I)<2^{k^{\prime\prime}}\}\bigcup\{I\in\mathcal{W}:I\cap B^{\prime}\neq\mbox{{\O}},\ell(I)\geq 2^{k^{\prime\prime}}\}=:\mathcal{W}^{\prime}_{1}\cup\mathcal{W}^{\prime}_{2}.

Note that for every I𝒲I\in\mathcal{W} with IBØI\cap B^{\prime}\neq\mbox{{\O}} we have

(I)<diam(I)dist(I,Ω)4<r42k2.\ell(I)<\operatorname{diam}(I)\leq\frac{\operatorname{dist}(I,\partial\Omega)}{4}<\frac{r^{\prime}}{4}\leq 2^{k^{\prime}-2}.

As a consequence, if 𝒲2Ø\mathcal{W}_{2}^{\prime}\neq\mbox{{\O}}, then k′′=k10kMk^{\prime\prime}=k-10\,k_{M}, and picking I𝒲2ØI\in\mathcal{W}_{2}^{\prime}\neq\mbox{{\O}} one has

r2kM2k′′(I)2k2rr.r\approx 2^{k}\approx_{M}2^{k^{\prime\prime}}\leq\ell(I)\leq 2^{k^{\prime}-2}\approx r^{\prime}\lesssim r.

This gives rMrr^{\prime}\approx_{M}r and #𝒲2M1\#\mathcal{W}_{2}^{\prime}\lesssim_{M}1.

To proceed, let us write

BΩ|u(X)|2GL(XΔ,X)𝑑XI𝒲1I|u(X)|2GL(XΔ,X)𝑑X+I𝒲2I|u(X)|2GL(XΔ,X)dX=:+,\iint_{B^{\prime}\cap\Omega}|\nabla u(X)|^{2}\,G_{L}(X_{\Delta},X)\,dX\leq\iint_{\bigcup\limits_{I\in\mathcal{W}^{\prime}_{1}}I}|\nabla u(X)|^{2}\,G_{L}(X_{\Delta},X)\,dX\\ +\sum_{I\in\mathcal{W}^{\prime}_{2}}\iint_{I}|\nabla u(X)|^{2}\,G_{L}(X_{\Delta},X)\,dX=:\mathcal{I}+\mathcal{II},

and we estimate each term in turn.

To estimate \mathcal{II} we may assume that 𝒲2Ø\mathcal{W}^{\prime}_{2}\neq\mbox{{\O}}, hence k′′=k10kMk^{\prime\prime}=k-10\,k_{M}, rrr^{\prime}\approx r and #𝒲21\#\mathcal{W}_{2}^{\prime}\lesssim 1. Then Lemma 2.53, the fact that ω(Ω)1\omega(\partial\Omega)\leq 1, Caccioppoli’s inequality, the normalization uL(Ω)=1\|u\|_{L^{\infty}(\Omega)}=1, and Harnack’s inequality give

=I𝒲2I|u(X)|2GL(XΔ,X)𝑑XI𝒲2(I)1nI|u(X)|2𝑑X#𝒲21ωLXΔ(Δ).\mathcal{II}=\sum_{I\in\mathcal{W}^{\prime}_{2}}\iint_{I}|\nabla u(X)|^{2}\,G_{L}(X_{\Delta},X)\,dX\lesssim\sum_{I\in\mathcal{W}^{\prime}_{2}}\ell(I)^{1-n}\iint_{I}|\nabla u(X)|^{2}\,dX\\ \lesssim\#\mathcal{W}^{\prime}_{2}\lesssim 1\approx\omega_{L}^{X_{\Delta}}(\Delta^{\prime}).

Next we deal with \mathcal{I}. Introduce the disjoint family ={Q𝔻:(Q)=2k′′1,Q3BØ}\mathcal{F}^{\prime}=\{Q\in\mathbb{D}:\ell(Q)=2^{k^{\prime\prime}-1},Q\cap 3\,B^{\prime}\neq\mbox{{\O}}\}. Given I𝒲1I\in\mathcal{W}^{\prime}_{1}, let XIBIX_{I}\in B^{\prime}\cap I, and QI𝔻Q_{I}\in\mathbb{D} be so that (QI)=(I)\ell(Q_{I})=\ell(I) and it contains some fixed yIΩy_{I}\in\partial\Omega such that dist(I,Ω)=dist(I,yI)\operatorname{dist}(I,\partial\Omega)=\operatorname{dist}(I,y_{I}). Then, as observed in Section 2.4, one has I𝒲QII\in\mathcal{W}_{Q_{I}}^{*}. Note that

|yIx|dist(yI,I)+diam(I)+|XIx|54dist(I,Ω)+|XIx|94|XIx|<3r,|y_{I}-x^{\prime}|\leq\operatorname{dist}(y_{I},I)+\operatorname{diam}(I)+|X_{I}-x^{\prime}|\leq\frac{5}{4}\operatorname{dist}(I,\partial\Omega)+|X_{I}-x^{\prime}|\leq\frac{9}{4}|X_{I}-x^{\prime}|<3\,r^{\prime},

hence yIQI3Δy_{I}\in Q_{I}\cap 3\,\Delta^{\prime}. This and the fact that, as observed before, (QI)=(I)<2k′′\ell(Q_{I})=\ell(I)<2^{k^{\prime\prime}} imply that QIQQ_{I}\subset Q for some QQ\in\mathcal{F}^{\prime}. Hence, I(1+λ)IUQITQ¯I\subset(1+\lambda)\,I\subset U_{Q_{I}}\subset\overline{T_{Q}} for some QQ\in\mathcal{F}^{\prime}. This eventually show that I𝒲1IQTQ\bigcup_{I\in\mathcal{W}^{\prime}_{1}}I\subset\bigcup_{Q\in\mathcal{F}^{\prime}}T_{Q} and therefore

QTQ|u(X)|2GL(XΔ,X)𝑑X.\mathcal{I}\leq\sum_{Q\in\mathcal{F}^{\prime}}\iint_{T_{Q}}|\nabla u(X)|^{2}\,G_{L}(X_{\Delta},X)\,dX.

For any QQ\in\mathcal{F}^{\prime} pick the unique (ancestor) Q^𝔻\widehat{Q}\in\mathbb{D} with (Q^)=2k1\ell(\widehat{Q})=2^{k-1} and QQ^Q\subset\widehat{Q}. Note that δ(XΔ)r\delta(X_{\Delta})\approx r, δ(XQ^)(Q^)=2k1r\delta(X_{\widehat{Q}})\approx\ell(\widehat{Q})=2^{k-1}\approx r. Also,

|XΔXQ^||XΔx|+|xx|+|xxQ|+|xQxQ^|+|xQ^XQ^|<3r+3r+diam(Q)+diam(Q^)+(Q^)r+2k′′+2kr.|X_{\Delta}-X_{\widehat{Q}}|\leq|X_{\Delta}-x|+|x-x^{\prime}|+|x^{\prime}-x_{Q}|+|x_{Q}-x_{\widehat{Q}}|+|x_{\widehat{Q}}-X_{\widehat{Q}}|\\ <3\,r+3\,r^{\prime}+\operatorname{diam}(Q)+\operatorname{diam}(\widehat{Q})+\ell(\widehat{Q})\lesssim r+2^{k^{\prime\prime}}+2^{k}\lesssim r.

Hence by the Harnack chain condition one obtains that GL(XΔ,X)GL(XQ^,X)G_{L}(X_{\Delta},X)\approx G_{L}(X_{\widehat{Q}},X) for every XTQX\in T_{Q} (in doing that we need to make sure that kMk_{M} is large enough so that the Harnack chain joining XΔX_{\Delta} and XQ^X_{\widehat{Q}}, which is crc\,r-away from Ω\partial\Omega, does not get near TQT_{Q}, which is κ0(Q)\kappa_{0}\,\ell(Q)-close to Ω\partial\Omega). Note also that (Q)(Q^)=2k′′k2kM<M1\frac{\ell(Q)}{\ell(\widehat{Q})}=2^{k^{\prime\prime}-k}\leq 2^{-k_{M}}<M^{-1}, provided kMk_{M} is large enough depending on MM. All theres and (4.1) yield

QTQ|u(X)|2GL(XQ^,X)𝑑XM0QωLXQ^(Q)M0QωLXΔ(Q)M0ωLXΔ(QQ)M0ωLXΔ(CΔ)M0ωLXΔ(Δ),\mathcal{I}\lesssim\sum_{Q\in\mathcal{F}^{\prime}}\iint_{T_{Q}}|\nabla u(X)|^{2}\,G_{L}(X_{\widehat{Q}},X)\,dX\lesssim M_{0}\,\sum_{Q\in\mathcal{F}^{\prime}}\omega_{L}^{X_{\widehat{Q}}}(Q)\\ \lesssim M_{0}\,\sum_{Q\in\mathcal{F}^{\prime}}\omega_{L}^{X_{\Delta}}(Q)\leq M_{0}\,\omega_{L}^{X_{\Delta}}\Big{(}\bigcup_{Q\in\mathcal{F}^{\prime}}Q\Big{)}\leq M_{0}\,\omega_{L}^{X_{\Delta}}(C\Delta^{\prime})\lesssim M_{0}\,\omega_{L}^{X_{\Delta}}(\Delta^{\prime}),

where we have used Lemma 2.53. This completes the prof of the fact that (4.1) implies (1.4).

We next focus on showing (4.1). With this goal in mind we fix Q0𝔻=𝔻(Ω)Q^{0}\in\mathbb{D}=\mathbb{D}(\partial\Omega) and let Q0𝔻Q0Q_{0}\in\mathbb{D}_{Q^{0}} with (Q0)(Q0)/M\ell(Q_{0})\leq\ell(Q^{0})/M with MM large enough so that XQ04BQX_{Q^{0}}\notin 4\,B_{Q}^{*} (cf. (2.31)). Write ωL=ωLXQ0\omega_{L}=\omega_{L}^{X_{Q^{0}}} and 𝒢L=GL(XQ0,)\mathcal{G}_{L}=G_{L}(X_{Q^{0}},\cdot) and note that our choice of MM, (2.49), and (2.50) guarantee that L𝒢L=LGL(,XQ0)=0L^{\top}\mathcal{G}_{L}=L^{\top}G_{L^{\top}}(\cdot,X_{Q^{0}})=0 in the weak sense in 4BQ4\,B_{Q}^{*}.

Fix N1N\gg 1 and consider the family of pairwise disjoint cubes N={Q𝔻Q0:(Q)=2N(Q0)}\mathcal{F}_{N}=\{Q\in\mathbb{D}_{Q_{0}}:\ell(Q)=2^{-N}\,\ell(Q_{0})\} and let ΩN=ΩN,Q0\Omega_{N}=\Omega_{\mathcal{F}_{N},Q_{0}} (cf. (2.25)). Note that by construction ΩNTQ0\Omega_{N}\subset T_{Q_{0}} is an increasing sequence of sets converging to TQ0T_{Q_{0}}. Our goal is to show that for every N1N\gg 1 there holds

(4.2) ΩN|u(X)|2𝒢L(X)𝑑XM0ωL(Q0),\iint_{\Omega_{N}}|\nabla u(X)|^{2}\,\mathcal{G}_{L}(X)\,dX\leq M_{0}\,\omega_{L}(Q_{0}),

with M0M_{0} independent of Q0Q^{0}, Q0Q_{0}, and NN. Hence the monotone convergence theorem yields

TQ0|u(X)|2𝒢L(X)𝑑X=limNΩN|u(X)|2𝒢L(X)𝑑XM0ωL(Q0),\iint_{T_{Q_{0}}}|\nabla u(X)|^{2}\,\mathcal{G}_{L}(X)\,dX=\lim_{N\to\infty}\iint_{\Omega_{N}}|\nabla u(X)|^{2}\,\mathcal{G}_{L}(X)\,dX\leq M_{0}\,\omega_{L}(Q_{0}),

which is (4.1).

Let us next start estimating (4.2). Using ΨN\Psi_{N} from Lemma 2.39 and the ellipticity of the matrix AA we have

ΩN|u(X)|2𝒢L(X)𝑑Xn+1|u(X)|2𝒢L(X)ΨN(X)𝑑X\displaystyle\iint_{\Omega_{N}}|\nabla u(X)|^{2}\,\mathcal{G}_{L}(X)\,dX\lesssim\iint_{\mathbb{R}^{n+1}}|\nabla u(X)|^{2}\,\mathcal{G}_{L}(X)\,\Psi_{N}(X)\,dX
n+1A(X)u(X)u(X)𝒢L(X)ΨN(X)𝑑X\displaystyle\qquad\lesssim\iint_{\mathbb{R}^{n+1}}A(X)\nabla u(X)\cdot\nabla u(X)\,\mathcal{G}_{L}(X)\,\Psi_{N}(X)\,dX
=n+1A(X)u(X)(u𝒢LΨN)(X)𝑑X\displaystyle\qquad=\iint_{\mathbb{R}^{n+1}}A(X)\nabla u(X)\cdot\nabla(u\,\mathcal{G}_{L}\,\Psi_{N})(X)\,dX
12n+1A(X)(u2ΨN)(X)𝒢L(X)𝑑X\displaystyle\qquad\qquad-\frac{1}{2}\iint_{\mathbb{R}^{n+1}}A(X)\nabla(u^{2}\,\Psi_{N})(X)\cdot\nabla\mathcal{G}_{L}(X)\,dX
12n+1A(X)(u2)(X)ΨN(X)𝒢L(X)𝑑X\displaystyle\qquad\qquad-\frac{1}{2}\iint_{\mathbb{R}^{n+1}}A(X)\nabla(u^{2})(X)\cdot\nabla\Psi_{N}(X)\,\mathcal{G}_{L}(X)\,dX
+12n+1A(X)ΨN(X)𝒢L(X)u(X)2𝑑X\displaystyle\qquad\qquad+\frac{1}{2}\iint_{\mathbb{R}^{n+1}}A(X)\nabla\Psi_{N}(X)\cdot\nabla\mathcal{G}_{L}(X)\,u(X)^{2}\,dX
=:1+2+3+4.\displaystyle\qquad=:\mathcal{I}_{1}+\mathcal{I}_{2}+\mathcal{I}_{3}+\mathcal{I}_{4}.

We observe that u𝒢LΨNu\,\mathcal{G}_{L}\,\Psi_{N} and u2ΨNu^{2}\,\Psi_{N} belong to W1,2(Ω)W^{1,2}(\Omega) since uWloc1,2(Ω)L(Ω)u\in W^{1,2}_{\rm loc}(\Omega)\cap L^{\infty}(\Omega), suppΨNΩN\operatorname{supp}\Psi_{N}\subset\Omega_{N}^{*}, δ(X)2N(Q0)\delta(X)\gtrsim 2^{-N}\,\ell(Q_{0}) for every XΩNX\in\Omega_{N}^{*}, the properties of GLG_{L}, and the fact that XQ0X_{Q^{0}} is away from ΩN\Omega_{N}^{*}δ(XQ0)(Q0)\delta(X_{Q^{0}})\approx\ell(Q^{0}) and by (2.31) one has δ(X)(Q0)(Q0)/Mδ(XQ0)/2\delta(X)\lesssim\ell(Q_{0})\leq\ell(Q^{0})/M\leq\delta(X_{Q^{0}})/2 for every XΩNX\in{\Omega}_{N}^{*} and provided MM is large enough. Using all these one can easily see via a limiting argument that the fact that Lu=0Lu=0 in the weak sense in Ω\Omega implies that 1=0\mathcal{I}_{1}=0. Likewise, one can easily show that 2=0\mathcal{I}_{2}=0 by recalling that suppΨNΩN12BQΩ\operatorname{supp}\Psi_{N}\subset\Omega_{N}^{*}\subset\frac{1}{2}\,{B_{Q}}^{*}\cap\Omega (see (2.31)) and that as mentioned above L𝒢L=0L^{\top}\mathcal{G}_{L}=0 in the weak sense in 4BQ4\,B_{Q}^{*}. Thus we are left with estimating the terms 3\mathcal{I}_{3} and 4\mathcal{I}_{4}. By (iii)(iii) in Lemma 2.39 and the fact that uL(Ω)=1\|u\|_{L^{\infty}(\Omega)}=1 we obtain

|3|+|4|\displaystyle|\mathcal{I}_{3}|+|\mathcal{I}_{4}| I𝒲NΣI(|u|𝒢L+|𝒢L|)δ()1𝑑X\displaystyle\lesssim\iint_{\bigcup_{I\in\mathcal{W}_{N}^{\Sigma}}I^{**}}\big{(}|\nabla u|\,\mathcal{G}_{L}+|\nabla\mathcal{G}_{L}|\big{)}\,\delta(\cdot)^{-1}\,dX
I𝒲NΣ(I)n12((I|u|2𝑑X)12𝒢L(X(I))+(I|𝒢L|2𝑑X)12)\displaystyle\lesssim\sum_{I\in\mathcal{W}_{N}^{\Sigma}}\ell(I)^{\frac{n-1}{2}}\,\Big{(}\Big{(}\iint_{I^{**}}|\nabla u|^{2}\,dX\Big{)}^{\frac{1}{2}}\,\mathcal{G}_{L}(X(I))+\Big{(}\iint_{I^{**}}|\nabla\mathcal{G}_{L}|^{2}\,dX\Big{)}^{\frac{1}{2}}\Big{)}
I𝒲NΣ(I)n32((I|u|2𝑑X)12+(I)n+12)𝒢L(X(I))\displaystyle\lesssim\sum_{I\in\mathcal{W}_{N}^{\Sigma}}\ell(I)^{\frac{n-3}{2}}\,\Big{(}\Big{(}\int_{I^{***}}|u|^{2}\,dX\Big{)}^{\frac{1}{2}}+\ell(I)^{\frac{n+1}{2}}\,\Big{)}\mathcal{G}_{L}(X(I))
I𝒲NΣ(I)n1𝒢L(X(I)),\displaystyle\lesssim\sum_{I\in\mathcal{W}_{N}^{\Sigma}}\ell(I)^{n-1}\,\mathcal{G}_{L}(X(I)),

where X(I)X(I) denotes the center of II, and we have used Harnack’s and Caccioppoli’s inequalities, that L𝒢L=0L^{\top}\mathcal{G}_{L}=0 and Lu=0Lu=0 in the weak sense in I12BQΩI^{***}\subset\frac{1}{2}\,B_{Q}^{*}\cap\Omega (see (2.31)). Invoking Lemmas 2.53 and Lemma 2.39 one can see that (I)n1𝒢L(X(I))ωL(Q^I)\ell(I)^{n-1}\,\mathcal{G}_{L}(X(I))\lesssim\omega_{L}(\widehat{Q}_{I}) for every I𝒲NΣI\in\mathcal{W}_{N}^{\Sigma}. This together with Lemma 2.39 allows us to conclude

|3|+|4|I𝒲NΣωL(Q^I)ωL(I𝒲NΣQ^I).|\mathcal{I}_{3}|+|\mathcal{I}_{4}|\lesssim\sum_{I\in\mathcal{W}_{N}^{\Sigma}}\omega_{L}(\widehat{Q}_{I})\lesssim\omega_{L}\Big{(}\bigcup_{I\in\mathcal{W}_{N}^{\Sigma}}\widehat{Q}_{I}\Big{)}.

Note that if yQ^Iy\in\widehat{Q}_{I} with I𝒲NΣI\in\mathcal{W}_{N}^{\Sigma} one has

|yxQ0|diam(Q^I)+dist(Q^I,I)+diam(I)+dist(I,xQ0)(I)+(Q0)(Q0)|y-x_{Q_{0}}|\leq\operatorname{diam}(\widehat{Q}_{I})+\operatorname{dist}(\widehat{Q}_{I},I)+\operatorname{diam}(I)+\operatorname{dist}(I,x_{Q_{0}})\lesssim\ell(I)+\ell(Q_{0})\lesssim\ell(Q_{0})

where we have used (2.42) and (2.31). Thus, Lemma 2.53 gives

|3|+|4|ωL(CΔQ0)ωL(Q0).|\mathcal{I}_{3}|+|\mathcal{I}_{4}|\lesssim\omega_{L}(C\,\Delta_{Q_{0}})\lesssim\omega_{L}(Q_{0}).

This allows us to complete the proof of Theorem 1.3. ∎

4.2. Proof of Theorem 1.5

We borrow some ideas from [19]. Given kk\in\mathbb{N} introduce the truncated localized conical square function: for every Q𝔻Q0Q\in\mathbb{D}_{Q_{0}} and xQx\in Q, let

𝒮Qku(x):=(ΓQk(x)|u(Y)|2δ(Y)1n𝑑Y)12,whereΓQk(x):=xQ𝔻Q(Q)2k(Q0)UQ,\mathcal{S}_{Q}^{k}u(x):=\bigg{(}\iint_{\Gamma_{Q}^{k}(x)}|\nabla u(Y)|^{2}\delta(Y)^{1-n}\,dY\bigg{)}^{\frac{1}{2}},\quad\text{where}\ \Gamma_{Q}^{k}(x):=\bigcup_{\begin{subarray}{c}x\in Q^{\prime}\in\mathbb{D}_{Q}\\ \ell(Q^{\prime})\geq 2^{-k}\,\ell(Q_{0})\end{subarray}}U_{Q^{\prime}},

where if (Q)<2k(Q0)\ell(Q)<2^{-k}\,\ell(Q_{0}) it is understood that ΓQk(x)=Ø\Gamma_{Q}^{k}(x)=\mbox{{\O}} and 𝒮Qku(x)=0\mathcal{S}_{Q}^{k}u(x)=0. Note that by the monotone convergence theorem 𝒮Qku(x)𝒮Qu(x)\mathcal{S}_{Q}^{k}u(x)\nearrow\mathcal{S}_{Q}u(x) as kk\to\infty for every xQx\in Q.

Fixed k0k_{0} large enough (eventually, k0k_{0}\to\infty), our goal is to show that we can find ϑ>0\vartheta>0 (independent of k0k_{0}) such that for every β,γ,λ>0\beta,\gamma,\lambda>0 we have

(4.3) ωLXQ0({xQ0:𝒮Q0k0u(x)>(1+β)λ,𝒩Q0u(x)γλ})(γβ)ϑωLXQ0({xQ0:𝒮Q0k0u(x)>βλ}),\omega_{L}^{X_{Q_{0}}}\big{(}\big{\{}x\in Q_{0}:\mathcal{S}_{Q_{0}}^{k_{0}}u(x)>(1+\beta)\,\lambda,\ \mathcal{N}_{Q_{0}}u(x)\leq\gamma\,\lambda\big{\}}\big{)}\\ \lesssim\Big{(}\frac{\gamma}{\beta}\Big{)}^{\vartheta}\,\omega_{L}^{X_{Q_{0}}}\big{(}\big{\{}x\in Q_{0}:\mathcal{S}_{Q_{0}}^{k_{0}}u(x)>\beta\,\lambda\big{\}}\big{)},

where the implicit constant depend on the allowable parameters and it is independent of k0k_{0}. To prove this we fix β,γ,λ>0\beta,\gamma,\lambda>0 and set

Eλ:={xQ0:𝒮Q0k0u(x)>λ}.E_{\lambda}:=\big{\{}x\in Q_{0}:\mathcal{S}_{Q_{0}}^{k_{0}}u(x)>\lambda\big{\}}.

Consider first the case EλQ0E_{\lambda}\subsetneq Q_{0}. Note that if xEλx\in E_{\lambda}, by definition 𝒮Q0k0u(x)>λ\mathcal{S}_{Q_{0}}^{k_{0}}u(x)>\lambda. Let Qx𝔻Q0Q_{x}\in\mathbb{D}_{Q_{0}} be the unique dyadic cube such that QxxQ_{x}\ni x and (Qx)=2k0(Q0)\ell(Q_{x})=2^{-k_{0}}\ell(Q_{0}). Then it is clear from construction that for every yQxy\in Q_{x} one has

ΓQ0k0(x)=QxQQ0UQ=ΓQ0k0(y)andλ<𝒮Q0k0u(x)=𝒮Q0k0u(y).\Gamma_{Q_{0}}^{k_{0}}(x)=\bigcup_{Q_{x}\subset Q\subset Q_{0}}U_{Q}=\Gamma_{Q_{0}}^{k_{0}}(y)\qquad\mbox{and}\qquad\lambda<\mathcal{S}_{Q_{0}}^{k_{0}}u(x)=\mathcal{S}_{Q_{0}}^{k_{0}}u(y).

Hence, QxEλQ_{x}\subset E_{\lambda} and we have shown that for every xEλx\in E_{\lambda} there exists Qx𝔻Q0Q_{x}\in\mathbb{D}_{Q_{0}} such that QxxQ_{x}\ni x and QxEλQ_{x}\subset E_{\lambda}. We then take the ancestors of QxQ_{x}, and look for the one with maximal side length QxmaxQxQ_{x}^{\rm max}\supset Q_{x} which is contained in EλE_{\lambda}. That is, QEλQ\subset E_{\lambda} for every QxQQxmaxQ_{x}\subset Q\subset Q_{x}^{\rm max} and Q^xmaxQ0EλØ\widehat{Q}_{x}^{\rm max}\cap Q_{0}\setminus E_{\lambda}\neq\mbox{{\O}} where Q^xmax\widehat{Q}_{x}^{\rm max} is the dyadic parent of QxmaxQ_{x}^{\rm max} (during this proof we will use Q^\widehat{Q} to denote the dyadic parent of QQ, that is, the only dyadic cube containing it with double side length). Note that the assumption EλQ0E_{\lambda}\subsetneq Q_{0} guarantees that Qxmax𝔻Q0{Q0}Q_{x}^{\rm max}\in\mathbb{D}_{Q_{0}}\setminus\{Q_{0}\}. Let 0={Qj}j\mathcal{F}_{0}=\{Q_{j}\}_{j} be the collection of such maximal cubes as xx runs in EλE_{\lambda} and we clearly have that the family is pairwise disjoint and also Eλ=Qj0QjE_{\lambda}=\bigcup_{Q_{j}\in\mathcal{F}_{0}}Q_{j}. Also, by construction (Qj)2k0(Q0)\ell(Q_{j})\geq 2^{-k_{0}}\ell(Q_{0}) and by the maximality of each QjQ_{j} we can select xjQ^jEλx_{j}\in\widehat{Q}_{j}\setminus E_{\lambda}.

On the other hand, for any xQjx\in Q_{j} we have, using that xjQ^jEλx_{j}\in\widehat{Q}_{j}\setminus E_{\lambda},

ΓQ0k0(x)=xQ𝔻Q0(Q)2k0(Q0)UQ=ΓQjk0(x)(QjQQ0UQ)ΓQjk0(x)ΓQ0k0(xj)\Gamma_{Q_{0}}^{k_{0}}(x)=\bigcup_{\begin{subarray}{c}x\in Q\in\mathbb{D}_{Q_{0}}\\ \ell(Q)\geq 2^{-k_{0}}\,\ell(Q_{0})\end{subarray}}U_{Q}=\Gamma_{Q_{j}}^{k_{0}}(x)\bigcup\Big{(}\bigcup_{Q_{j}\subsetneq Q\subset Q_{0}}U_{Q}\Big{)}\subset\Gamma_{Q_{j}}^{k_{0}}(x)\bigcup\Gamma_{Q_{0}}^{k_{0}}(x_{j})

and therefore

𝒮Q0k0u(x)𝒮Qjk0u(x)+𝒮Q0k0u(xj)𝒮Qjk0u(x)+λ.\mathcal{S}_{Q_{0}}^{k_{0}}u(x)\leq\mathcal{S}_{Q_{j}}^{k_{0}}u(x)+\mathcal{S}_{Q_{0}}^{k_{0}}u(x_{j})\leq\mathcal{S}_{Q_{j}}^{k_{0}}u(x)+\lambda.

As a consequence,

{xQj:𝒮Q0k0u(x)>(1+β)λ}{xQj:𝒮Qjk0u(x)>βλ}\big{\{}x\in Q_{j}:\mathcal{S}_{Q_{0}}^{k_{0}}u(x)>(1+\beta)\,\lambda\big{\}}\subset\big{\{}x\in Q_{j}:\mathcal{S}_{Q_{j}}^{k_{0}}u(x)>\beta\lambda\big{\}}

and

{xQ0:𝒮Q0k0u(x)>(1+β)λ}={xQ0:𝒮Q0k0u(x)>(1+β)λ}Eλ=Qj0{xQj:𝒮Q0k0u(x)>(1+β)λ}Qj0{xQj:𝒮Qjk0u(x)>βλ}.\big{\{}x\in Q_{0}:\mathcal{S}_{Q_{0}}^{k_{0}}u(x)>(1+\beta)\,\lambda\big{\}}=\big{\{}x\in Q_{0}:\mathcal{S}_{Q_{0}}^{k_{0}}u(x)>(1+\beta)\,\lambda\big{\}}\cap E_{\lambda}\\ =\bigcup_{Q_{j}\in\mathcal{F}_{0}}\big{\{}x\in Q_{j}:\mathcal{S}_{Q_{0}}^{k_{0}}u(x)>(1+\beta)\,\lambda\big{\}}\subset\bigcup_{Q_{j}\in\mathcal{F}_{0}}\big{\{}x\in Q_{j}:\mathcal{S}_{Q_{j}}^{k_{0}}u(x)>\beta\lambda\big{\}}.

This has been done under the assumption that EλQ0E_{\lambda}\subsetneq Q_{0}. In the case Eλ=Q0E_{\lambda}=Q_{0} we set 0={Q0}\mathcal{F}_{0}=\{Q_{0}\}. Then in both cases we obtain

(4.4) {xQ0:𝒮Q0k0u(x)>(1+β)λ}Qj0{xQj:𝒮Qjk0u(x)>βλ}.\displaystyle\big{\{}x\in Q_{0}:\mathcal{S}_{Q_{0}}^{k_{0}}u(x)>(1+\beta)\,\lambda\big{\}}\subset\bigcup_{Q_{j}\in\mathcal{F}_{0}}\big{\{}x\in Q_{j}:\mathcal{S}_{Q_{j}}^{k_{0}}u(x)>\beta\lambda\big{\}}.

Thus, to obtain (4.3) it suffices to see that for every Qj0Q_{j}\in\mathcal{F}_{0}

(4.5) ωLXQ0({xQj:𝒮Qjk0u(x)>βλ,𝒩Q0u(x)γλ})(γβ)ϑωLXQ0(Qj).\displaystyle\omega_{L}^{X_{Q_{0}}}\big{(}\big{\{}x\in Q_{j}:\mathcal{S}_{Q_{j}}^{k_{0}}u(x)>\beta\,\lambda,\ \mathcal{N}_{Q_{0}}u(x)\leq\gamma\,\lambda\big{\}}\big{)}\lesssim\Big{(}\frac{\gamma}{\beta}\Big{)}^{\vartheta}\,\omega_{L}^{X_{Q_{0}}}(Q_{j}).

From this we just need to sum in Qj0Q_{j}\in\mathcal{F}_{0} to see that (4.4) together with the previous facts yield the desired estimate (4.3):

ωLXQ0({xQ0:𝒮Q0k0u(x)>(1+β)λ,𝒩Q0u(x)γλ})Qj0ωLXQ0({xQj:𝒮Qjk0u(x)>βλ,𝒩Q0u(x)γλ})(γβ)ϑQj0ωLXQ0(Qj)=(γβ)ϑωLXQ0(Qj0Qj)=(γβ)ϑωLXQ0(Eλ).\omega_{L}^{X_{Q_{0}}}\big{(}\big{\{}x\in Q_{0}:\mathcal{S}_{Q_{0}}^{k_{0}}u(x)>(1+\beta)\,\lambda,\ \mathcal{N}_{Q_{0}}u(x)\leq\gamma\,\lambda\big{\}}\big{)}\\ \leq\sum_{Q_{j}\in\mathcal{F}_{0}}\omega_{L}^{X_{Q_{0}}}\big{(}\big{\{}x\in Q_{j}:\mathcal{S}_{Q_{j}}^{k_{0}}u(x)>\beta\,\lambda,\ \mathcal{N}_{Q_{0}}u(x)\leq\gamma\,\lambda\big{\}}\big{)}\\ \lesssim\Big{(}\frac{\gamma}{\beta}\Big{)}^{\vartheta}\,\sum_{Q_{j}\in\mathcal{F}_{0}}\omega_{L}^{X_{Q_{0}}}(Q_{j})=\Big{(}\frac{\gamma}{\beta}\Big{)}^{\vartheta}\,\omega_{L}^{X_{Q_{0}}}\Big{(}\bigcup_{Q_{j}\in\mathcal{F}_{0}}Q_{j}\Big{)}=\Big{(}\frac{\gamma}{\beta}\Big{)}^{\vartheta}\,\omega_{L}^{X_{Q_{0}}}(E_{\lambda}).

Let us then obtain (4.5). Fix Qj0Q_{j}\in\mathcal{F}_{0} and to ease the notation write P0=QjP_{0}=Q_{j}. Set

(4.6) E~λ={xP0:𝒮P0k0u(x)>βλ},Fλ={xP0:𝒩Q0u(x)γλ}.\widetilde{E}_{\lambda}=\big{\{}x\in P_{0}:\mathcal{S}_{P_{0}}^{k_{0}}u(x)>\beta\,\lambda\big{\}},\qquad F_{\lambda}=\big{\{}x\in P_{0}:\mathcal{N}_{Q_{0}}u(x)\leq\gamma\,\lambda\big{\}}.

If ωLXQ0(Fλ)=0\omega_{L}^{X_{Q_{0}}}(F_{\lambda})=0 then (4.5) is trivial, hence we may assume that ωLXQ0(Fλ)>0\omega_{L}^{X_{Q_{0}}}(F_{\lambda})>0 so that P0Fλ=FλØP_{0}\cap F_{\lambda}=F_{\lambda}\neq\mbox{{\O}}. We subdivide P0P_{0} dyadically and stop the first time that QFλ=ØQ\cap F_{\lambda}=\mbox{{\O}}. If one never stops we write P0={Ø}\mathcal{F}_{P_{0}}^{*}=\{\mbox{{\O}}\}, otherwise P0={Pj}j𝔻P0{P0}\mathcal{F}_{P_{0}}^{*}=\{P_{j}\}_{j}\subset\mathbb{D}_{P_{0}}\setminus\{P_{0}\} is the family of stopping cubes which is maximal (hence pairwise disjoint) with respect to the property FλQ=ØF_{\lambda}\cap Q=\mbox{{\O}}. In particular, FλP0(𝔻P0,P0Pj)F_{\lambda}\subset P_{0}\setminus(\cup_{\mathbb{D}_{\mathcal{F}_{P_{0}}^{*},P_{0}}}P_{j}).

Next we claim that

(4.7) xFλΓP0k0(x)Q𝔻P0,P0(Q)2k0(Q0)UQint(Q𝔻P0,P0UQ)=ΩP0,P0=:Ω\bigcup_{x\in F_{\lambda}}\Gamma_{P_{0}}^{k_{0}}(x)\subset\bigcup_{\begin{subarray}{c}Q\in\mathbb{D}_{\mathcal{F}_{P_{0}}^{*},P_{0}}\\ \ell(Q)\geq 2^{-k_{0}}\,\ell(Q_{0})\end{subarray}}U_{Q}\subset{\rm int}\bigg{(}\bigcup_{Q\in\mathbb{D}_{\mathcal{F}_{P_{0}}^{*},P_{0}}}U_{Q}^{*}\bigg{)}=\Omega_{\mathcal{F}_{P_{0}}^{*},P_{0}}^{*}=:\Omega_{*}

To verify the first inclusion, we fix YΓk0P0(x)Y\in\Gamma^{k_{0}}_{P_{0}}(x) with xFλx\in F_{\lambda}. Then, YUQY\in U_{Q} where xQ𝔻P0x\in Q\in\mathbb{D}_{P_{0}}. Since xFλx\in F_{\lambda} we must have Q𝔻P0Q\in\mathbb{D}_{\mathcal{F}_{P_{0}}^{*}} (otherwise QPjQ\subset P_{j} for some PjP0P_{j}\in\mathcal{F}_{P_{0}}^{*} and this would imply that xPjFλ=Øx\in P_{j}\cap F_{\lambda}=\mbox{{\O}}) and therefore Q𝔻P0,P0Q\in\mathbb{D}_{\mathcal{F}_{P_{0}}^{*},P_{0}} which gives the first inclusion. The second inclusion in (4.7) is trivial (since UQint(UQ)U_{Q}\subset{\rm int}(U_{Q}^{*})).

To continue we see that

(4.8) |u(Y)|γλ,for all YΩ.|u(Y)|\leq\gamma\,\lambda,\qquad\mbox{for all }Y\in\Omega_{*}.

Fix such a YY so that YUQY\in U_{Q}^{*} for some Q𝔻P0,P0Q\in\mathbb{D}_{\mathcal{F}_{P_{0}}^{*},P_{0}}. If QFλ=ØQ\cap F_{\lambda}=\mbox{{\O}}, by maximality of the cubes in P0\mathcal{F}_{P_{0}}^{*}, it follows that QPjQ\subset P_{j} for some PjP0P_{j}\in\mathcal{F}_{P_{0}}^{*}, which contradicts the fact Q𝔻P0,P0Q\in\mathbb{D}_{\mathcal{F}_{P_{0}}^{*},P_{0}}. Thus, QFλØQ\cap F_{\lambda}\neq\mbox{{\O}} and we can select xQFλx\in Q\cap F_{\lambda} so that by definition |u(Y)|𝒩Q0u(x)γλ|u(Y)|\leq\mathcal{N}_{Q_{0}}u(x)\leq\gamma\,\lambda since YUQΓQ0(x)Y\in U_{Q}^{*}\subset\Gamma^{*}_{Q_{0}}(x).

Apply Lemma 3.1 to find X:=YP0ΩΩX_{*}:=Y_{P_{0}}\in\Omega_{*}\cap\Omega so that

(4.9) (P0)dist(X,Ω)δ(X).\displaystyle\ell(P_{0})\approx\operatorname{dist}(X_{*},\partial\Omega_{*})\approx\delta(X_{*}).

Let ωL:=ωL,ΩX\omega_{L}^{*}:=\omega_{L,\Omega_{*}}^{X_{*}} be the elliptic measure associated with LL relative to Ω\Omega_{*} with pole at XX_{*} and write δ=dist(,Ω)\delta_{*}=\operatorname{dist}(\cdot,\partial\Omega_{*}). Given YΩY\in\Omega_{*}, we choose yYΩy_{Y}\in\partial\Omega_{*} such that |YyY|=δ(Y)|Y-y_{Y}|=\delta_{*}(Y). By definition, for xFλx\in F_{\lambda} and YΓP0(x)Y\in\Gamma_{P_{0}}(x), there is a Q𝔻P0Q\in\mathbb{D}_{P_{0}} such that YUQY\in U_{Q} and xQx\in Q. Thus, by the triangle inequality, and the definition of UQU_{Q}, we have that for YΓP0(x)Y\in\Gamma_{P_{0}}(x),

(4.10) |xyY||xY|+δ(Y)δ(Y)+δ(Y)δ(Y)|x-y_{Y}|\leq|x-Y|+\delta_{*}(Y)\approx\delta(Y)+\delta_{*}(Y)\approx\delta_{*}(Y)

where in the last step we have used that

(4.11) δ(Y)δ(Y)forYQ𝔻P0,P0UQ.\delta(Y)\approx\delta_{*}(Y)\qquad\text{for}\quad Y\in\bigcup_{Q\in\mathbb{D}_{\mathcal{F}_{P_{0}}^{*},P_{0}}}U_{Q}.

On the other hand, as observed above FλP0(Qj)ΩΩF_{\lambda}\subset P_{0}\setminus(\cup_{\mathcal{F}}Q_{j})\subset\partial\Omega\cap\partial\Omega_{*}, see [17, Proposition 6.1]. Using this and the fact that if QFλØQ\cap F_{\lambda}\neq\mbox{{\O}} then Q𝔻P0,P0Q\in\mathbb{D}_{\mathcal{F}_{P_{0}}^{*},P_{0}} we have

(4.12) Fλ𝒮P0k0u(x)2dωL(x)\displaystyle\int_{F_{\lambda}}\mathcal{S}_{P_{0}}^{k_{0}}u(x)^{2}\,d\omega_{L}^{*}(x) =FλΓP0k0(x)|u(Y)|2δ(Y)1ndYdωL(x)\displaystyle=\int_{F_{\lambda}}\iint_{\Gamma_{P_{0}}^{k_{0}}(x)}|\nabla u(Y)|^{2}\,\delta(Y)^{1-n}\,dY\,d\omega_{L}^{*}(x)
FλxQ𝔻P0(Q)2k0(Q0)UQ|u(Y)|2δ(Y)1ndYdωL(x)\displaystyle\leq\int_{F_{\lambda}}\sum_{\begin{subarray}{c}x\in Q\in\mathbb{D}_{P_{0}}\\ \ell(Q)\geq 2^{-k_{0}}\,\ell(Q_{0})\end{subarray}}\iint_{U_{Q}}|\nabla u(Y)|^{2}\,\delta(Y)^{1-n}\,dY\,d\omega_{L}^{*}(x)
Q𝔻P0,P0(UQ|u(Y)|2dY)(Q)1nωL(QFλ)\displaystyle\lesssim\sum_{Q\in\mathbb{D}_{\mathcal{F}_{P_{0}}^{*},P_{0}}}\Big{(}\iint_{U_{Q}}|\nabla u(Y)|^{2}\,dY\Big{)}\,\ell(Q)^{1-n}\,\omega_{L}^{*}(Q\cap F_{\lambda})
Q𝔻P0,P0(Q)M1(P0)+Q𝔻P0,P0(Q)<M1(P0)\displaystyle\lesssim\sum_{\begin{subarray}{c}Q\in\mathbb{D}_{\mathcal{F}_{P_{0}}^{*},P_{0}}\\ \ell(Q)\geq M^{-1}\ell(P_{0})\end{subarray}}\dots+\sum_{\begin{subarray}{c}Q\in\mathbb{D}_{\mathcal{F}_{P_{0}}^{*},P_{0}}\\ \ell(Q)<M^{-1}\ell(P_{0})\end{subarray}}\dots
=:Σ1+Σ2,\displaystyle=:\Sigma_{1}+\Sigma_{2},

where MM is a large constant to be chosen.

We start estimating Σ1\Sigma_{1}. Note first that #{Q:𝔻P0:(Q)M1(P0)}CM\#\{Q:\in\mathbb{D}_{P_{0}}:\ell(Q)\geq M^{-1}\ell(P_{0})\}\leq C_{M}, thus

Σ1\displaystyle\Sigma_{1} Q𝔻P0,P0(Q)M1(P0)(Q)1nI𝒲QI|u(Y)|2dY\displaystyle\lesssim\sum_{\begin{subarray}{c}Q\in\mathbb{D}_{\mathcal{F}_{P_{0}}^{*},P_{0}}\\ \ell(Q)\geq M^{-1}\ell(P_{0})\end{subarray}}\ell(Q)^{1-n}\sum_{I^{\in}\mathcal{W}_{Q}^{*}}\iint_{I^{*}}|\nabla u(Y)|^{2}\,dY
Q𝔻P0,P0(Q)M1(P0)(Q)1nI𝒲Q(I)2I|u(Y)|2dY\displaystyle\lesssim\sum_{\begin{subarray}{c}Q\in\mathbb{D}_{\mathcal{F}_{P_{0}}^{*},P_{0}}\\ \ell(Q)\geq M^{-1}\ell(P_{0})\end{subarray}}\ell(Q)^{1-n}\sum_{I\in\mathcal{W}_{Q}^{*}}\ell(I)^{-2}\iint_{I^{**}}|u(Y)|^{2}\,dY
(γλ)2Q𝔻P0(Q)M1(P0)(Q)1nI𝒲Q(I)n1\displaystyle\lesssim(\gamma\,\lambda)^{2}\sum_{\begin{subarray}{c}Q\in\mathbb{D}_{P_{0}}\\ \ell(Q)\geq M^{-1}\ell(P_{0})\end{subarray}}\ell(Q)^{1-n}\sum_{I\in\mathcal{W}_{Q}^{*}}\ell(I)^{n-1}
M(γλ)2,\displaystyle\lesssim_{M}(\gamma\,\lambda)^{2},

where we have used (4.8), along with the fact that int(I)int(UQ)Ω\operatorname{int}(I^{**})\subset\operatorname{int}(U_{Q}^{*})\subset\Omega_{*} for any I𝒲QI\in\mathcal{W}_{Q}^{*} with Q𝔻P0,P0Q\in\mathbb{D}_{\mathcal{F}_{P_{0}}^{*},P_{0}}, and the fact that 𝒲Q\mathcal{W}_{Q}^{*} has uniformly bounded cardinality. To estimate Σ2\Sigma_{2} we note that picking yQQFλy_{Q}\in Q\cap F_{\lambda} we have that QFλB(yQ,2diam(Q))Ω=:ΔQQ\cap F_{\lambda}\subset B(y_{Q},2\,\operatorname{diam}(Q))\cap\partial\Omega_{*}=:\Delta_{Q}^{*}. Write XQX_{Q}^{*} for Corkscrew relative to ΔQ\Delta_{Q}^{*} with respect to Ω\Omega_{*} so that δ(XQ)diam(Q)M1(P0)\delta_{*}(X_{Q}^{*})\approx\operatorname{diam}(Q)\lesssim M^{-1}\ell(P_{0}). Note that by (4.9), we clearly have XΩB(yQ,4diam(Q))X_{*}\in\Omega\setminus B(y_{Q},4\,\operatorname{diam}(Q)) provided MM is sufficiently large. Hence, by Lemma 2.53 part (b)(b) applied in Ω\Omega_{*}, which is a 1-sided NTA domain satisfying the CDC by Proposition 2.37, we obtain for every YUQY\in U_{Q}

(4.13) (Q)1nωL(QFλ)diam(Q)1nωL(ΔQ)GL,(X,XQ)GL,(X,Y),\ell(Q)^{1-n}\,\omega_{L}^{*}(Q\cap F_{\lambda})\lesssim\operatorname{diam}(Q)^{1-n}\omega_{L}^{*}(\Delta_{Q}^{*})\lesssim G_{L,*}(X_{*},X_{Q}^{*})\approx G_{L,*}(X_{*},Y),

where GL,G_{L,*} is the Green function for the operator LL relative to the domain Ω\Omega_{*}. Above the last estimate uses Harnack’s inequality (we may need to tale MM slightly larger) and the fact that by (4.11), one has δ(Y)(Q)diam(Q)δ(XQ)\delta_{*}(Y)\approx\ell(Q)\approx\operatorname{diam}(Q)\approx\delta_{*}(X_{Q}^{*}) (see Remark 2.56) and that if IYI\ni Y with I𝒲QI\in\mathcal{W}_{Q}^{*}

|YX|diam(I)+dist(I,Q)+diam(Q)+|yQX|diam(Q).|Y-X_{*}|\leq\operatorname{diam}(I)+\operatorname{dist}(I,Q)+\operatorname{diam}(Q)+|y_{Q}-X^{*}|\lesssim\operatorname{diam}(Q).

Write {P0i}i𝔻P0\{P_{0}^{i}\}_{i}\subset\mathbb{D}_{P_{0}} for the collection of dyadic cubes with M(P0)(P0i)<2M(P0)M\,\ell(P_{0})\leq\ell(P_{0}^{i})<2\,M\ell(P_{0}) which has uniformly bounded cardinality depending on MM. Note that

{Q𝔻P0,P0:(Q)<M1(P0)}i𝔻P0,P0i.\{Q\in\mathbb{D}_{\mathcal{F}_{P_{0}}^{*},P_{0}}:\ell(Q)<M^{-1}\ell(P_{0})\}\subset\bigcup_{i}\mathbb{D}_{\mathcal{F}_{P_{0}}^{*},P_{0}^{i}}.

For each ii, if 𝔻P0,P0iØ\mathbb{D}_{\mathcal{F}_{P_{0}}^{*},P_{0}^{i}}\neq\mbox{{\O}} then P0i𝔻P0,P0P_{0}^{i}\in\mathbb{D}_{\mathcal{F}_{P_{0}}^{*},P_{0}} and hence P0iFλØP_{0}^{i}\cap F_{\lambda}\neq\mbox{{\O}}. Pick then yiP0iFλy_{i}\in P_{0}^{i}\cap F_{\lambda} and note that for every Q𝔻P0,P0iQ\in\mathbb{D}_{\mathcal{F}_{P_{0}}^{*},P_{0}^{i}} by (2.31) it follows that

UQTP0iΩBP0iΩB(yi,Cκ0(P0i))Ω=:BiΩ.U_{Q}\subset T_{P_{0}^{i}}\cap\Omega_{*}\subset B_{P_{0}^{i}}^{*}\cap\Omega_{*}\subset B(y_{i},C\,\kappa_{0}\,\ell(P_{0}^{i}))\cap\Omega_{*}=:B_{i}\cap\Omega_{*}.

Using then (4.13) we have

Σ2\displaystyle\Sigma_{2} Q𝔻P0,P0(Q)<M1(P0)UQ|u(Y)|2GL,(X,Y)dY\displaystyle\lesssim\sum_{\begin{subarray}{c}Q\in\mathbb{D}_{\mathcal{F}_{P_{0}}^{*},P_{0}}\\ \ell(Q)<M^{-1}\ell(P_{0})\end{subarray}}\iint_{U_{Q}}|\nabla u(Y)|^{2}\,G_{L,*}(X_{*},Y)\,dY
iQ𝔻P0,P0i(Q)<M1(P0)UQ|u(Y)|2GL,(X,Y)dY\displaystyle\lesssim\sum_{i}\sum_{\begin{subarray}{c}Q\in\mathbb{D}_{\mathcal{F}_{P_{0}}^{*},P_{0}^{i}}\\ \ell(Q)<M^{-1}\ell(P_{0})\end{subarray}}\iint_{U_{Q}}|\nabla u(Y)|^{2}\,G_{L,*}(X_{*},Y)\,dY
iBiΩ|u(Y)|2GL,(X,Y)dY\displaystyle\lesssim\sum_{i}\iint_{B_{i}\cap\Omega_{*}}|\nabla u(Y)|^{2}\,G_{L,*}(X_{*},Y)\,dY
uL(Ω)2iωL(BiΩ)\displaystyle\lesssim\|u\|_{L^{\infty}(\Omega_{*})}^{2}\sum_{i}\omega_{L}^{*}(B_{i}\cap\partial\Omega_{*})
(γλ)2,\displaystyle\lesssim(\gamma\,\lambda)^{2},

where we have invoked Theorem 1.3 applied in Ω\Omega_{*}, which is a 1-sided NTA domain satisfying the CDC by Proposition 2.37, and we may need to take MM slightly larger and use Harnack’s inequality; (4.8); and the fact that {P0i}i𝔻P0\{P_{0}^{i}\}_{i}\subset\mathbb{D}_{P_{0}} has uniformly bounded cardinality.

Using Chebyshev’s inequality, (4.12), and collecting the estimates for Σ1\Sigma_{1} and Σ2\Sigma_{2} we conclude that

ωL(E~λFλ)1(βλ)2E~λFλ(𝒮P0k0u)2dωL1(βλ)2Fλ𝒮P0k0u(x)2dωL(x)(γβ)2.\displaystyle\omega_{L}^{*}(\widetilde{E}_{\lambda}\cap F_{\lambda})\leq\frac{1}{(\beta\,\lambda)^{2}}\,\int_{\widetilde{E}_{\lambda}\cap F_{\lambda}}(\mathcal{S}_{P_{0}}^{k_{0}}u)^{2}\,d\omega_{L}^{*}\leq\frac{1}{(\beta\,\lambda)^{2}}\,\int_{F_{\lambda}}\mathcal{S}_{P_{0}}^{k_{0}}u(x)^{2}\,d\omega_{L}^{*}(x)\lesssim\Big{(}\frac{\gamma}{\beta}\Big{)}^{2}.

At this point we invoke Lemma 3.19 in P0P_{0} with P0\mathcal{F}_{P_{0}}^{*} —we warn the reader that P0P_{0} and P0={Pj}j\mathcal{F}_{P_{0}}^{*}=\{P_{j}\}_{j} play the role of Q0Q_{0} and {Qj}j\{Q_{j}\}_{j} and that associated to each PjP_{j} one finds P~j\widetilde{P}_{j} as in Proposition 3.1, which now plays the role of PjP_{j} in that result, and μ=ωLX\mu=\omega_{L}^{X_{*}} (recall that X=YP0X_{*}=Y_{P_{0}}) and observe that the fact that FλP0(𝔻P0,P0Pj)F_{\lambda}\subset P_{0}\setminus(\cup_{\mathbb{D}_{\mathcal{F}_{P_{0}}^{*},P_{0}}}P_{j}) implies on account of (3.21) that for some ϑ>0\vartheta>0 we have

ωLX(E~λFλ)ωLX(P0)(ωL(E~λFλ)ωL(ΔP0))ϑ2(γβ)ϑ,\displaystyle\frac{\omega_{L}^{X_{*}}(\widetilde{E}_{\lambda}\cap F_{\lambda})}{\omega_{L}^{X_{*}}(P_{0})}\lesssim\bigg{(}\frac{\omega_{L}^{*}(\widetilde{E}_{\lambda}\cap F_{\lambda})}{\omega_{L}^{*}(\Delta_{\star}^{P_{0}})}\bigg{)}^{\frac{\vartheta}{2}}\lesssim\Big{(}\frac{\gamma}{\beta}\Big{)}^{\vartheta},

where we have used that ωL(ΔP0)1\omega_{L}^{*}(\Delta_{\star}^{P_{0}})\approx 1 since ΔP0:=B(xP0,tP0)Ω\Delta_{\star}^{P_{0}}:=B(x_{P_{0}}^{\star},t_{P_{0}})\cap\partial\Omega_{*} with tP0(P0)diam(Ω)t_{P_{0}}\approx\ell({P_{0}})\approx\operatorname{diam}(\partial\Omega_{*}), xP0Ωx_{P_{0}}^{\star}\in\partial\Omega_{*}, (4.9), Harnack’s inequality, and Lemma 2.53 part (a)(a). We can then use Remark 2.54, Harnack’s inequality, and (4.9), to conclude that

ωLXQ0(E~λFλ)ωLXQ0(P0)ωLXP0(E~λFλ)ωLXP0(P0)ωLX(E~λFλ)ωLX(P0)(γβ)ϑ.\frac{\omega_{L}^{X_{Q_{0}}}(\widetilde{E}_{\lambda}\cap F_{\lambda})}{\omega_{L}^{X_{Q_{0}}}(P_{0})}\approx\frac{\omega_{L}^{X_{P_{0}}}(\widetilde{E}_{\lambda}\cap F_{\lambda})}{\omega_{L}^{X_{P_{0}}}(P_{0})}\approx\frac{\omega_{L}^{X_{*}}(\widetilde{E}_{\lambda}\cap F_{\lambda})}{\omega_{L}^{X_{*}}(P_{0})}\lesssim\Big{(}\frac{\gamma}{\beta}\Big{)}^{\vartheta}.

Recalling that P0=Qj0P_{0}=Q_{j}\in\mathcal{F}_{0}, and the definitions of E~λ\widetilde{E}_{\lambda} and FλF_{\lambda} in (4.6) the previous estimates readily lead to (4.5).

To conclude we need to see how (4.3) yields (1.6). With this goal in mind we first observe that for every xQ0x\in Q_{0} and YΓQ0k0(x)Y\in\Gamma_{Q_{0}}^{k_{0}}(x) one has that YBQ0¯ΩY\in\overline{B_{Q_{0}}^{*}}\cap\Omega (see (2.31)) and also δ(Y)2k0(Q0)\delta(Y)\gtrsim 2^{-k_{0}}\,\ell(Q_{0}). Hence, since uW1,2loc(Ω)u\in W^{1,2}_{\rm loc}(\Omega), one has

(4.14) supxQ0𝒮Q0k0u(x)=supxQ0(ΓQ0k0(x)|u(Y)|2δ(Y)1ndY)12(2k0(Q0))1n2(BQ0{YΩ:δ(Y)2k0(Q0)}|u(Y)|2dY)12<.\sup_{x\in Q_{0}}\mathcal{S}_{Q_{0}}^{k_{0}}u(x)=\sup_{x\in Q_{0}}\Big{(}\iint_{\Gamma_{Q_{0}}^{k_{0}}(x)}|\nabla u(Y)|^{2}\delta(Y)^{1-n}\,dY\Big{)}^{\frac{1}{2}}\\ \lesssim(2^{-k_{0}}\,\ell(Q_{0}))^{\frac{1-n}{2}}\Big{(}\iint_{B_{Q_{0}}^{*}\cap\{Y\in\Omega:\delta(Y)\gtrsim 2^{-k_{0}}\,\ell(Q_{0})\}}|\nabla u(Y)|^{2}\,dY\Big{)}^{\frac{1}{2}}<\infty.

On the other hand, given 0<q<0<q<\infty, we can use (4.3)

(1+β)q𝒮Q0k0uLq(Q0,ωLXQ0)q\displaystyle(1+\beta)^{-q}\,\|\mathcal{S}_{Q_{0}}^{k_{0}}u\|_{L^{q}(Q_{0},\omega_{L}^{X_{Q_{0}}})}^{q}
=0qλqωLXQ0({xQ0:𝒮Q0k0u(x)>(1+β)λ})dλλ\displaystyle\qquad=\int_{0}^{\infty}q\,\lambda^{q}\,\omega_{L}^{X_{Q_{0}}}\big{(}\big{\{}x\in Q_{0}:\mathcal{S}_{Q_{0}}^{k_{0}}u(x)>(1+\beta)\,\lambda\big{\}}\big{)}\,\frac{d\lambda}{\lambda}
0qλqωLXQ0({xQ0:𝒮Q0k0u(x)>(1+β)λ,𝒩Q0u(x)γλ})dλλ\displaystyle\qquad\leq\int_{0}^{\infty}q\,\lambda^{q}\,\omega_{L}^{X_{Q_{0}}}\big{(}\big{\{}x\in Q_{0}:\mathcal{S}_{Q_{0}}^{k_{0}}u(x)>(1+\beta)\,\lambda,\ \mathcal{N}_{Q_{0}}u(x)\leq\gamma\,\lambda\big{\}}\big{)}\,\frac{d\lambda}{\lambda}
+0qλqωLXQ0({xQ0:𝒩Q0u(x)>γλ})dλλ\displaystyle\qquad\qquad\qquad+\int_{0}^{\infty}q\,\lambda^{q}\,\omega_{L}^{X_{Q_{0}}}\big{(}\big{\{}x\in Q_{0}:\mathcal{N}_{Q_{0}}u(x)>\gamma\,\lambda\big{\}}\big{)}\,\frac{d\lambda}{\lambda}
(γβ)ϑ0qλqωLXQ0({xQ0:𝒮Q0k0u(x)>βλ})dλλ\displaystyle\qquad\lesssim\Big{(}\frac{\gamma}{\beta}\Big{)}^{\vartheta}\,\int_{0}^{\infty}q\,\lambda^{q}\,\omega_{L}^{X_{Q_{0}}}\big{(}\big{\{}x\in Q_{0}:\mathcal{S}_{Q_{0}}^{k_{0}}u(x)>\beta\,\lambda\big{\}}\big{)}\,\frac{d\lambda}{\lambda}
+γq𝒩Q0uLq(Q0,ωLXQ0)q\displaystyle\qquad\qquad\qquad+\gamma^{-q}\,\|\mathcal{N}_{Q_{0}}u\|_{L^{q}(Q_{0},\omega_{L}^{X_{Q_{0}}})}^{q}
(γβ)ϑβq𝒮Q0k0uLq(Q0,ωLXQ0)q+γq𝒩Q0uLq(Q0,ωLXQ0)q.\displaystyle\qquad\lesssim\Big{(}\frac{\gamma}{\beta}\Big{)}^{\vartheta}\,\beta^{-q}\,\|\mathcal{S}_{Q_{0}}^{k_{0}}u\|_{L^{q}(Q_{0},\omega_{L}^{X_{Q_{0}}})}^{q}+\gamma^{-q}\,\|\mathcal{N}_{Q_{0}}u\|_{L^{q}(Q_{0},\omega_{L}^{X_{Q_{0}}})}^{q}.

We can then choose γ\gamma small enough so that we can hide the first term in the right hand side of the last quantity (which is finite by (4.14)) and eventually conclude that

𝒮Q0k0uLq(Q0,ωLXQ0)q𝒩Q0uLq(Q0,ωLXQ0)q.\|\mathcal{S}_{Q_{0}}^{k_{0}}u\|_{L^{q}(Q_{0},\omega_{L}^{X_{Q_{0}}})}^{q}\lesssim\|\mathcal{N}_{Q_{0}}u\|_{L^{q}(Q_{0},\omega_{L}^{X_{Q_{0}}})}^{q}.

Since the implicit constant does not depend on k0k_{0} and 𝒮Q0ku(x)𝒮Q0u(x)\mathcal{S}_{Q_{0}}^{k}u(x)\nearrow\mathcal{S}_{Q_{0}}u(x) as kk\to\infty for every xQx\in Q, the monotone convergence theorem yields at once (1.6) and the proof Theorem 1.5 is complete.

References

  • [1] M. Akman, S. Hofmann, J. M. Martell, and T. Toro. Perturbation of elliptic operators in 1-sided NTA domains satisfying the capacity density condition. https://arxiv.org/abs/1901.08261v2. Preprint, 2019.
  • [2] M. Akman, S. Hofmann, J. M. Martell, and T. Toro. Perturbation of elliptic operators in 1-sided NTA domains satisfying the capacity density condition. https://arxiv.org/abs/1901.08261v3. Preprint, 2021.
  • [3] L. A. Caffarelli, E. B. Fabes, and C. E. Kenig. Completely singular elliptic-harmonic measures. Indiana Univ. Math. J., 30(6):917–924, 1981.
  • [4] M. Cao, Ó. Dominguéz, J. M. Martell, and P. Tradacete. On the A{A}_{\infty} condition for elliptic operators in 1-sided NTA domains satisfying the capacity density condition. https://arxiv.org/abs/2101.06064. Preprint, 2021.
  • [5] J. Cavero, S. Hofmann, and J. M. Martell. Perturbations of elliptic operators in 1-sided chord-arc domains. Part I: Small and large perturbation for symmetric operators. Trans. Amer. Math. Soc., 371(4):2797–2835, 2019.
  • [6] J. Cavero, S. Hofmann, J. M. Martell, and T. Toro. Perturbations of elliptic operators in 1-sided chord-arc domains. Part II: non-symmetric operators and Carleson measure estimates. Trans. Amer. Math. Soc., 373(11):7901–7935, 2020.
  • [7] M. Christ. A T(b)T(b) theorem with remarks on analytic capacity and the Cauchy integral. Colloq. Math., 60/61(2):601–628, 1990.
  • [8] R. R. Coifman and C. Fefferman. Weighted norm inequalities for maximal functions and singular integrals. Studia Math., 51:241–250, 1974.
  • [9] B. E. J. Dahlberg. On the absolute continuity of elliptic measures. Amer. J. Math., 108(5):1119–1138, 1986.
  • [10] B. E. J. Dahlberg, D. S. Jerison, and C. E. Kenig. Area integral estimates for elliptic differential operators with nonsmooth coefficients. Ark. Mat., 22(1):97–108, 1984.
  • [11] R. Fefferman. A criterion for the absolute continuity of the harmonic measure associated with an elliptic operator. J. Amer. Math. Soc., 2(1):127–135, 1989.
  • [12] R. A. Fefferman, C. E. Kenig, and J. Pipher. The theory of weights and the Dirichlet problem for elliptic equations. Ann. of Math. (2), 134(1):65–124, 1991.
  • [13] J. Feneuil and B. Poggi. Generalized Carleson perturbations of elliptic operators and applications. https://arxiv.org/abs/2011.06574. Preprint, 2021.
  • [14] J. García-Cuerva and J. L. Rubio de Francia. Weighted norm inequalities and related topics, volume 116 of North-Holland Mathematics Studies. North-Holland Publishing Co., Amsterdam, 1985. Notas de Matemática [Mathematical Notes], 104.
  • [15] J. Heinonen, T. Kilpeläinen, and O. Martio. Nonlinear potential theory of degenerate elliptic equations. Dover Publications, Inc., Mineola, NY, 2006. Unabridged republication of the 1993 original.
  • [16] S. Hofmann, P. Le, J. M. Martell, and K. Nyström. The weak-AA_{\infty} property of harmonic and pp-harmonic measures implies uniform rectifiability. Anal. PDE, 10(3):513–558, 2017.
  • [17] S. Hofmann and J. M. Martell. Uniform rectifiability and harmonic measure I: Uniform rectifiability implies Poisson kernels in LpL^{p}. Ann. Sci. Éc. Norm. Supér. (4), 47(3):577–654, 2014.
  • [18] S. Hofmann, J. M. Martell, and S. Mayboroda. Uniform rectifiability, Carleson measure estimates, and approximation of harmonic functions. Duke Math. J., 165(12):2331–2389, 2016.
  • [19] S. Hofmann, J. M. Martell, and S. Mayboroda. Transference of scale-invariant estimates from Lipschitz to Non-tangentially accessible to Uniformly rectifiable domains. https://arxiv.org/abs/1904.13116. Preprint, 2019.
  • [20] S. Hofmann, J. M. Martell, and T. Toro. General divergence form elliptic operators on domains with ADR boundaries, and on 1-sided NTA domains. Work in Progress, 2014.
  • [21] S. Hofmann, J. M. Martell, and T. Toro. AA_{\infty} implies NTA for a class of variable coefficient elliptic operators. J. Differential Equations, 263(10):6147–6188, 2017.
  • [22] S. Hofmann, D. Mitrea, M. Mitrea, and A. J. Morris. LpL^{p}-square function estimates on spaces of homogeneous type and on uniformly rectifiable sets. Mem. Amer. Math. Soc., 245(1159):v+108, 2017.
  • [23] T. Hytönen and A. Kairema. Systems of dyadic cubes in a doubling metric space. Colloq. Math., 126(1):1–33, 2012.
  • [24] T. Hytönen and A. Kairema. What is a cube? Ann. Acad. Sci. Fenn. Math., 38(2):405–412, 2013.
  • [25] D. S. Jerison and C. E. Kenig. Boundary behavior of harmonic functions in nontangentially accessible domains. Adv. in Math., 46(1):80–147, 1982.
  • [26] J. L. Lewis. Uniformly fat sets. Trans. Amer. Math. Soc., 308(1):177–196, 1988.
  • [27] W. Littman, G. Stampacchia, and H. F. Weinberger. Regular points for elliptic equations with discontinuous coefficients. Ann. Scuola Norm. Sup. Pisa (3), 17:43–77, 1963.
  • [28] E. Milakis, J. Pipher, and T. Toro. Harmonic analysis on chord arc domains. J. Geom. Anal., 23(4):2091–2157, 2013.
  • [29] L. Modica and S. Mortola. Construction of a singular elliptic-harmonic measure. Manuscripta Math., 33(1):81–98, 1980/81.
  • [30] Z. Zhao. BMO solvability and AA_{\infty} condition of the elliptic measures in uniform domains. J. Geom. Anal., 28(2):866–908, 2018.