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

The AA_{\infty} condition, ε\varepsilon-approximators, and Varopoulos extensions in uniform domains

S. Bortz, B. Poggi, O. Tapiola, and X. Tolsa Simon Bortz, Department of Mathematics
University of Alabama
Tuscaloosa, AL, 35487, USA
sbortz@ua.edu Bruno Poggi, Departament de Matemàtiques
Universitat Autònoma de Barcelona
Barcelona, Catalonia.
bgpoggi.math@gmail.com Olli Tapiola, Departament de Matemàtiques
Universitat Autònoma de Barcelona
Barcelona, Catalonia.
olli.m.tapiola@gmail.com Xavier Tolsa, ICREA
Barcelona
Departament de Matemàtiques
Universitat Autònoma de Barcelona and Centre de Recerca Matemàtica
Barcelona
Catalonia.
xtolsa@mat.uab.cat
Abstract.

Suppose that Ωn+1\Omega\subset\mathbb{R}^{n+1}, n1n\geq 1, is a uniform domain with nn-Ahlfors regular boundary and LL is a (not necessarily symmetric) divergence form elliptic, real, bounded operator in Ω\Omega. We show that the corresponding elliptic measure ωL\omega_{L} is quantitatively absolutely continuous with respect to surface measure of Ω\partial\Omega in the sense that ωLA(σ)\omega_{L}\in A_{\infty}(\sigma) if and only if any bounded solution uu to Lu=0Lu=0 in Ω\Omega is ε\varepsilon-approximable for any ε(0,1)\varepsilon\in(0,1). By ε\varepsilon-approximability of uu we mean that there exists a function Φ=Φε\Phi=\Phi^{\varepsilon} such that uΦL(Ω)εuL(Ω)\|u-\Phi\|_{L^{\infty}(\Omega)}\leq\varepsilon\|u\|_{L^{\infty}(\Omega)} and the measure μ~Φ\widetilde{\mu}_{\Phi} with dμ~=|Φ(Y)|dYd\widetilde{\mu}=|\nabla\Phi(Y)|\,dY is a Carleson measure with LL^{\infty} control over the Carleson norm.

As a consequence of this approximability result, we show that boundary BMO\operatorname{BMO} functions with compact support can have Varopoulos-type extensions even in some sets with unrectifiable boundaries, that is, smooth extensions that converge non-tangentially back to the original data and that satisfy L1L^{1}-type Carleson measure estimates with BMO\operatorname{BMO} control over the Carleson norm. Our result complements the recent work of Hofmann and the third named author who showed the existence of these types of extensions in the presence of a quantitative rectifiability hypothesis.

S.B. was supported by the Simons foundation grant “Travel support for Mathematicians” (grant number 959861). B.P., O.T. and X.T. were supported by the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation programme (grant agreement 101018680). X.T. is also partially supported by MICINN (Spain) under the grant PID2020-114167GB-I00, the María de Maeztu Program for units of excellence (Spain) (CEX2020-001084-M), and 2021-SGR-00071 (Catalonia).

1. Introduction

Carleson measures (see Definition 2.7) and their connections to geometry, harmonic analysis and partial differential equations (PDE) have been studied actively over the previous decades; see [Mat21] for a recent survey related to many key developments and other relevant research. Carleson-type estimates are particularly powerful for measures μF\mu_{F} such that dμF=|F(Y)|2dist(Y,Ω)dYd\mu_{F}=|\nabla F(Y)|^{2}\operatorname{dist}(Y,\partial\Omega)\,dY and FF is a solution to an elliptic PDE in the set Ω\Omega: they can be used to, for example, characterize functions of bounded mean oscillation (BMO) [FS72], quantitative boundary geometry [HMM16, GMT18] and absolute continuity properties of harmonic measure [HL18]. As it is discussed in [Gar07, Chapters VI and VIII], Carleson-type estimates for measures μ~F\widetilde{\mu}_{F} such that dμ~F=|F(Y)|dYd\widetilde{\mu}_{F}=|\nabla F(Y)|\,dY would be very powerful but they fail even for harmonic functions in the unit disk. To circumvent this problem, Varopoulos [Var78] introduced a way to approximate harmonic functions in LL^{\infty} sense by other functions satisfying these estimates. This ε\varepsilon-approximability theory has been studied from many points of view in the past years [Gar07, Dah80, KKPT00, HKMP15, HMM16, HR18, GMT18, AGMT22, HT20, BT19, BH20, Gar22].

The initial motivation behind ε\varepsilon-approximability theory was to prove an extension theorem for BMO functions inspired by Carleson’s Corona Theorem [Car62]. Varopoulos [Var77, Var78] showed that any compactly supported BMO function ff in n{\mathbb{R}}^{n} has a smooth extension VfV_{f} to the upper half-space such that the extension converges non-tangentially (see Definition 2.8) back to ff and |Vf(Y)|dY|\nabla V_{f}(Y)|\,dY defines a Carleson measure. Inspired by the power of these estimates, Hofmann and the third named author [HT21] recently showed that uniform rectifiability (see Definition 2.6) of the boundary is enough to guarantee the existence of these Varopoulos-type extensions. Since Carleson-type estimates for harmonic functions can be used to characterize uniform rectifiability [HMM16, GMT18] or even stronger geometric properties [AHMMT20], it is natural to ask if the existence of Varopoulos-type extensions (which satisfy better Carleson-type estimates than harmonic functions) characterizes some quantitative geometric properties for the boundary.

In this paper, we show that the existence of Varopoulos-type extensions does not characterize uniform rectifiability but they can exist even in sets with unrectifiable boundaries. Our main result is the following ε\varepsilon-approximability result which can be used to build Varopoulos-type extensions. Throughout, set σn|Ω\sigma\coloneqq\mathcal{H}^{n}|_{\partial\Omega}.

Theorem 1.1.

Let Ωn+1\Omega\subset\mathbb{R}^{n+1}, n1n\geq 1, be a uniform domain (see Definition 2.3) with nn-Ahlfors regular boundary (see Definition 2.4). Let L=divAL=-\operatorname{div}A\nabla be a real, not necessarily symmetric, bounded elliptic operator in Ω\Omega such that the corresponding elliptic measure ωL\omega_{L} satisfies ωLA(σ)\omega_{L}\in A_{\infty}(\sigma) (see Definition 2.31), and let ε(0,1)\varepsilon\in(0,1). Then any solution uWloc1,2(Ω)L(Ω)u\in W^{1,2}_{\operatorname{loc}}(\Omega)\cap L^{\infty}(\Omega) to Lu=0Lu=0 in Ω\Omega is ε\varepsilon-approximable: there exists a constant CεC_{\varepsilon} and a function Φ=ΦεC(Ω)\Phi=\Phi^{\varepsilon}\in C^{\infty}(\Omega) such that

  1. i)

    uΦL(Ω)εuL(Ω)\|u-\Phi\|_{L^{\infty}(\Omega)}\leq\varepsilon\|u\|_{L^{\infty}(\Omega)},

  2. ii)

    Φ\Phi satisfies a quantitative L1L^{1}-type Carleson measure estimate

    supxΩ,r>01rnB(x,r)Ω|Φ(Y)|𝑑YCεuL(Ω),\displaystyle\sup_{x\in\partial\Omega,r>0}\frac{1}{r^{n}}\iint_{B(x,r)\cap\Omega}|\nabla\Phi(Y)|\,dY\leq C_{\varepsilon}\|u\|_{L^{\infty}(\Omega)},
  3. iii)

    |Φ(X)|CεuL(Ω)δ(X)|\nabla\Phi(X)|\leq\tfrac{C_{\varepsilon}\|u\|_{L^{\infty}(\Omega)}}{\delta(X)} for every XΩX\in\Omega,

  4. iv)

    if |XY|dist(X,Ω)|X-Y|\ll\operatorname{dist}(X,\partial\Omega), then |Φ(X)Φ(Y)|CεuL(Ω)δ(X)|XY||\Phi(X)-\Phi(Y)|\leq\tfrac{C_{\varepsilon}\|u\|_{L^{\infty}(\Omega)}}{\delta(X)}|X-Y|,

  5. v)

    there exists a function φL(Ω)\varphi\in L^{\infty}(\partial\Omega) such that

    limYx,n.t.Φ(Y)=φ(x) for σ-a.e. xΩ.\displaystyle\lim_{Y\to x,\,\operatorname{n.t.}}\Phi(Y)=\varphi(x)\text{ for }\sigma\text{-a.e. }x\in\partial\Omega.

The notation limYx,n.t.\lim_{Y\to x,\,\operatorname{n.t.}} means non-tangential convergence (see Definition 2.8) and δ()dist(,Ω)\delta(\cdot)\coloneqq\operatorname{dist}(\cdot,\partial\Omega). Here, CεC_{\varepsilon} depends on ε\varepsilon, the structural constants related to Ω\Omega and Ω\partial\Omega, ellipticity and the ωLA(σ)\omega_{L}\in A_{\infty}(\sigma) constants.

The proof of Theorem 1.1 borrows some ideas from the proof of [HMM16, Theorem 1.3] (which is an adaptation of the classical construction in [Gar07, Chapter VIII, Theorem 6.1]), but very quickly our argument must differ significantly. In [HMM16], the authors constructed the approximators for harmonic functions in the presence of a quantitative rectifiability hypothesis. This allowed them to construct approximating chord-arc domains which, in turn, allowed them to use an “NSN\lesssim S” estimate for harmonic functions. This was the most delicate part of their argument and our main challenges are strongly related to overcoming the fact that we cannot use the same tools due to our geometry (our boundary may be purely unrectifiable) and our operator LL (no control on its structure or smoothness).

We also obtain the following converse to Theorem 1.1.

Theorem 1.2.

Let Ωn+1\Omega\subset\mathbb{R}^{n+1}, n1n\geq 1, be a uniform domain (see Definition 2.3) with nn-Ahlfors regular boundary (see Definition 2.4), and let L=divAL=-\operatorname{div}A\nabla be a real, not necessarily symmetric, bounded elliptic operator in Ω\Omega. Suppose also that every solution uWloc1,2(Ω)L(Ω)u\in W^{1,2}_{\operatorname{loc}}(\Omega)\cap L^{\infty}(\Omega) to Lu=0Lu=0 is ε\varepsilon-approximable for every ε(0,1)\varepsilon\in(0,1) in the sense of Theorem 1.1. Then ωLA(σ)\omega_{L}\in A_{\infty}(\sigma) (see Definition 2.31).

In particular, by combining Theorem 1.1 and Theorem 1.2 with [CHMT20, Theorem 1.1], we get a new characterization of the AA_{\infty} property of elliptic measure on uniform domains.

Corollary 1.3.

Let Ωn+1\Omega\subset\mathbb{R}^{n+1}, n1n\geq 1, be a uniform domain (see Definition 2.3) with nn-Ahlfors regular boundary (see Definition 2.4), and let L=divAL=-\operatorname{div}A\nabla be a real, not necessarily symmetric, bounded elliptic operator in Ω\Omega. The following conditions are equivalent:

  1. (a)

    ωLA(σ)\omega_{L}\in A_{\infty}(\sigma),

  2. (b)

    every solution uWloc1,2(Ω)L(Ω)u\in W^{1,2}_{\operatorname{loc}}(\Omega)\cap L^{\infty}(\Omega) to Lu=0Lu=0 in Ω\Omega is ε\varepsilon-approximable for any ε(0,1)\varepsilon\in(0,1) in the sense of Theorem 1.1,

  3. (c)

    every solution uWloc1,2(Ω)L(Ω)u\in W^{1,2}_{\operatorname{loc}}(\Omega)\cap L^{\infty}(\Omega) to Lu=0Lu=0 in Ω\Omega satisfies an L2L^{2}-type Carleson measure estimate with LL^{\infty} control over the Carleson norm: there exists C1C\geq 1 such that

    supxΩr(0,diam(Ω))1rnB(x,r)Ω|u(X)|2dist(X,Ω)𝑑XCuLΩ2.\displaystyle\sup_{\begin{subarray}{c}x\in\partial\Omega\\ r\in(0,\operatorname{diam}(\partial\Omega))\end{subarray}}\frac{1}{r^{n}}\iint_{B(x,r)\cap\Omega}|\nabla u(X)|^{2}\operatorname{dist}(X,\partial\Omega)\,dX\leq C\|u\|_{L^{\infty}{\Omega}}^{2}.

Only the implications “(a) \implies (b)” and “(b) \implies (a)” in Corollary 1.3 are new. The equivalence “(a)\iff (c)” was already shown in [CHMT20] for n2n\geq 2, while the case n=1n=1 follows as a particular case of a more general result in [FP22], with a similar method of proof.

In the setting of a uniform domain with Ahlfors regular boundary, the conditions (a), (b) and (c) in Corollary 1.3 are known to be equivalent with uniform rectifiability of Ω\partial\Omega for the special case L=ΔL=-\Delta [HM14, HMU14, HMM16, GMT18], or for L=divAL=-\operatorname{div}A\nabla where AA is a locally Lipschitz symmetric matrix such that |A||\nabla A| satisfies an L1L^{1}-type Carleson measure condition [CHMT20, HMT17, AGMT22]. Even for the aforementioned operators with nice structure, all the available proofs in the literature of the implications “(a) \implies (b)” or “(c) \implies (b)” in Corollary 1.3 rely on uniform rectifiability techniques in a decisive fashion, and do not extend to domains with rougher boundaries. The main novelty of this manuscript is that we succeed in proving the implication “(a) \implies (b)” without appealing to uniform rectifiability theory. This allows us to establish the equivalence “(a)\iff (b)” for arbitrary elliptic operators in settings that are beyond chord-arc domains (see Definition 2.5). For further characterizations of the ωLA(σ)\omega_{L}\in A_{\infty}(\sigma) property for arbitrary real divergence form elliptic operators, see [CDMT22] and [MPT].

Let us see an immediate corollary of our new characterization of the ωLA(σ)\omega_{L}\in A_{\infty}(\sigma) property. We say that LL is a Dahlberg–Kenig–Pipher operator if ALiploc(Ω)A\in\operatorname{Lip}_{\operatorname{loc}}(\Omega) with |A|dist(,Ω)L(Ω)|\nabla A|\operatorname{dist}(\cdot,\partial\Omega)\in L^{\infty}(\Omega), and the measure μA\mu_{A} such that dμA=|A(X)|2dist(X,Ω)dXd\mu_{A}=|\nabla A(X)|^{2}\operatorname{dist}(X,\partial\Omega)\,dX is a Carleson measure. For these operators, the conditions (a) and (c) in Corollary 1.3 are equivalent with uniform rectifiability of Ω\partial\Omega when Ω\Omega is a uniform domain with Ahlfors regular boundary [HMMTZ21]. Combining the main result of [HMMTZ21] with Corollary 1.3 gives us a new result for Dahlberg–Kenig–Pipher operators:

Corollary 1.4.

Let Ωn+1\Omega\subset\mathbb{R}^{n+1}, n2n\geq 2, be a uniform domain (see Definition 2.3) with nn-Ahlfors regular boundary (see Definition 2.4), and let L=divAL=-\operatorname{div}A\nabla be a not necessarily symmetric Dahlberg–Kenig–Pipher operator in Ω\Omega. The following are equivalent:

  1. (a)

    Ω\partial\Omega is uniformly rectifiable (see Definition 2.6).

  2. (b)

    every solution uWloc1,2(Ω)L(Ω)u\in W^{1,2}_{\operatorname{loc}}(\Omega)\cap L^{\infty}(\Omega) to Lu=0Lu=0 in Ω\Omega is ε\varepsilon-approximable for any ε(0,1)\varepsilon\in(0,1) in the sense of Theorem 1.1.

As a consequence of Theorem 1.1 and the techniques in [HT21], we get the following generalization of the Varopoulos extension theorem [Var77, Var78]:

Theorem 1.5.

Let Ωn+1\Omega\subset\mathbb{R}^{n+1}, n1n\geq 1, be a uniform domain (see Definition 2.3) with Ahflors regular boundary (see Definition 2.4). If there exists a divergence form elliptic operator L=divAL=-\mathop{\operatorname{div}}\nolimits A\nabla such that the corresponding elliptic measure ωL\omega_{L} satisfies ωLA(σ)\omega_{L}\in A_{\infty}(\sigma) (see Definition 2.31), then every fBMOc(Ω)f\in\operatorname{BMO}_{\operatorname{c}}(\partial\Omega) has a Varopoulos extension in Ω\Omega. That is, there exists FF with the following properties:

  1. (1)

    FC(Ω)F\in C^{\infty}(\Omega) and |F(X)|fBMO(Ω)δ(X)|\nabla F(X)|\lesssim\tfrac{\|f\|_{\operatorname{BMO}(\partial\Omega)}}{\delta(X)}, for all XΩX\in\Omega,

  2. (2)

    limYx,n.t.F(Y)=f(x)\lim_{Y\to x,\,\operatorname{n.t.}}F(Y)=f(x) for σ\sigma-a.e. xΩx\in\partial\Omega, and

  3. (3)

    |F(Y)||\nabla F(Y)| is the density of a Carleson measure in the sense that

    supr>0,xΩ1rnB(x,r)Ω|F(Y)|𝑑YCfBMO(Ω).\sup_{r>0,x\in\partial\Omega}\frac{1}{r^{n}}\iint_{B(x,r)\cap\Omega}|\nabla F(Y)|\,dY\leq C\|f\|_{\operatorname{BMO}(\partial\Omega)}.

The notation limYx,n.t.\lim_{Y\to x,\,\operatorname{n.t.}} means non-tangential limit and δ()dist(,Ω)\delta(\cdot)\coloneqq\operatorname{dist}(\cdot,\partial\Omega). Here, CC depends on the structural constants related to Ω\Omega and Ω\partial\Omega, and the ωLA(σ)\omega_{L}\in A_{\infty}(\sigma) constants.

Theorem 1.5 is not (and is not meant to be) a generalization of the main result in [HT21] where an extension theorem of this type was proven in the presence of a quantitative rectifiability hypothesis for the boundary. The novelty of Theorem 1.5 is that its assumptions hold for some sets with very rough boundaries. In particular, recently David and Mayboroda [DM21] showed that the key hypothesis of Theorem 1.5 holds for the exterior of the 44-corner Cantor set:

Theorem 1.6 ([DM21, Section 4]).

Let Ω\Omega be the complement of the 44-corner Cantor set in 2\mathbb{R}^{2} (see Section 7). There exists a divergence form elliptic operator L=divAL=-\mathop{\operatorname{div}}\nolimits A\nabla in Ω\Omega such that ωLA(σ)\omega_{L}\in A_{\infty}(\sigma) (see Definition 2.31). More precisely, one can take the matrix AA to be diagonal and equal to the identity outside a ball of radius 11 concentric with the Cantor set and so that the LL-elliptic measure with pole at \infty equals σ/σ(Ω)\sigma/\sigma(\partial\Omega).

The remarkable thing about Theorem 1.6 is that since the complement of the 44-corner Cantor set is an unbounded uniform domain with unrectifiable 11-Ahlfors regular boundary, we know that harmonic measure for this set cannot satisfy the A(σ)A_{\infty}(\sigma) condition (see, for example, [AHMMT20]). Thus, constructing operators like this is highly non-trivial. In Section 7, we take the David–Mayboroda example and use it to build an example of a similar operator in 3\mathbb{R}^{3} (see Proposition 7.6).

By combining Theorems 1.5 and 1.6 and Proposition 7.6, we get the following:

Corollary 1.7.

There exist uniform domains Ω\Omega with unrectifiable Ahlfors regular boundaries Ω\partial\Omega in 2\mathbb{R}^{2} and 3\mathbb{R}^{3} such that every function fBMO(Ω)f\in\operatorname{BMO}(\partial\Omega) with compact support has a Varopoulos extension in Ω\Omega. In particular, the existence of Varopoulos extensions does not imply rectifiability for the boundary.

By Corollary 1.7, L1L^{1}-type Carleson measure estimates are simultaneously too strong and too weak from the point of view of the David–Semmes theory: harmonic functions fail these estimates even in the unit disk but the existence of non-harmonic extensions that satisfy these estimates does not imply even qualitative rectifiability for the boundary.

Finally, we want to mention an upcoming paper that is closely related to our results. As we were finishing this work, we were informed111Personal communication. about an upcoming manuscript by M. Mourgoglou and T. Zacharopoulos, where the authors construct Varopoulos extensions in corkscrew domains with Ahlfors regular boundaries satisfying a mild quantitative connectivity hypothesis by different methods.

The paper is organized as follows. In Section 2, we discuss basic definitions and consider some key tools from dyadic analysis and elliptic PDE theory. In Section 3, we prove some important preliminary estimates for Theorem 1.1, and in Section 4, we prove Theorem 1.1. In Section 5, we prove Theorem 1.2 (and hence, Corollary 1.3). Finally, in Section 6 we sketch the proof of Theorem 1.5 and in Section 7 we construct a David–Mayboroda-type example in 3\mathbb{R}^{3} (which completes the proof of Corollary 1.7).

2. Preliminaries

Throughout, we let Ωn+1\Omega\subset\mathbb{R}^{n+1} be an open set with n1n\geq 1. We say that Ω\Omega is a domain if it is also connected.

Usually, we use capital letters X,Y,ZX,Y,Z, and so on to denote points in Ω\Omega, and lowercase letters x,y,zx,y,z, and so on to denote points in Ω\partial\Omega. For Xn+1X\in\mathbb{R}^{n+1} and r>0r>0, we let B(X,r)B(X,r) be the Euclidean open ball of radius rr centered at XX. The letters cc and CC and their obvious variations denote constants that depend only on dimension, nn-Ahlfors regularity constant (see Definition 2.4), corkscrew constant (see Definition 2.1), Harnack chain constants, ellipticity constants (see Section 2.5), and so on. We call these kinds of constants structural constants. We write aba\lesssim b if aCba\leq Cb for a structural constant CC and aba\approx b if C1baC2bC_{1}b\leq a\leq C_{2}b for structural constants C1C_{1} and C2C_{2}.

2.1. Uniform domains, chord-arc domains, Ahlfors regularity and uniform rectifiability

Definition 2.1 (Corkscrew condition).

We say that a domain Ωn+1\Omega\subset\mathbb{R}^{n+1} satisfies the corkscrew condition if there exists a constant γ>0\gamma>0 such that for every xΩx\in\partial\Omega and r(0,diam(Ω))r\in(0,\operatorname{diam}(\Omega)) there exists Yx,rY_{x,r} such that

B(Yx,r,γr)B(x,r)Ω.B(Y_{x,r},\gamma r)\subset B(x,r)\cap\Omega.

We call Yx,rY_{x,r} a corkscrew point relative to xx at scale rr.

Definition 2.2 (Harnack chain Condition).

We say a domain Ωn+1\Omega\subset\mathbb{R}^{n+1} satisfies the Harnack chain condition if there exists a uniform constant CC such that for every ρ>0\rho>0 and Λ0\Lambda\geq 0 and X,XΩX,X^{\prime}\in\Omega with dist(X,Ω),dist(X,Ω)ρ\operatorname{dist}(X,\partial\Omega),\operatorname{dist}(X^{\prime},\partial\Omega)\geq\rho and |XX|Λρ|X-X^{\prime}|\leq\Lambda\rho there exists a chain of open balls B1,,BJB_{1},\dots,B_{J} with JN(Λ)J\leq N(\Lambda) with XB1X\in B_{1}, XBJX^{\prime}\in B_{J}, BjBj+1ØB_{j}\cap B_{j+1}\neq\mbox{{\O}} and C1diam(Bj)dist(Bj,Ω)Cdiam(Bj)C^{-1}\operatorname{diam}(B_{j})\leq\operatorname{dist}(B_{j},\partial\Omega)\leq C\operatorname{diam}(B_{j}).

Definition 2.3 (Uniform domain).

We say that a domain Ωn+1\Omega\subset\mathbb{R}^{n+1} is uniform if it satisfies the corkscrew and Harnack chain conditions.

Definition 2.4 (Ahlfors regularity).

We say Σn+1\Sigma\subset\mathbb{R}^{n+1} is nn-Ahlfors regular (or simply Ahlfors regular) if there exists CC such that

C1rnn(B(x,r)Σ)Crn,for each xr(0,diam(Σ)).C^{-1}r^{n}\leq\mathcal{H}^{n}(B(x,r)\cap\Sigma)\leq Cr^{n},\quad\text{for each }x\in r\in(0,\operatorname{diam}(\Sigma)).

Here and below n\mathcal{H}^{n} denotes the nn-dimensional Hausdorff measure.

Definition 2.5 (Chord-arc domain).

We say that a domain Ωn+1\Omega\subset\mathbb{R}^{n+1} is a chord-arc domain if Ω\Omega satisfies the Harnack chain condition, both Ω\Omega and intΩc\text{int}\,\Omega^{\operatorname{c}} satisfy the corkscrew condition, and the boundary Ω\partial\Omega is nn-Ahlfors regular.

Definition 2.6 (Uniform rectifiability).

Following [DS91], we say that an nn-Ahlfors regular set En+1E\subset\mathbb{R}^{n+1} is uniformly rectifiable if it contains “big pieces of Lipschitz images” of n\mathbb{R}^{n}: there exist constants θ,M>0\theta,M>0 such that for every xEx\in E and r(0,diam(E))r\in(0,\operatorname{diam}(E)) there is a Lipschitz mapping ρ=ρx,r:nn+1\rho=\rho_{x,r}\colon\mathbb{R}^{n}\to\mathbb{R}^{n+1}, with Lipschitz norm no larger that MM, such that

n(EB(x,r)ρ({yn:|y|<r}))θrn.\displaystyle\mathcal{H}^{n}\big{(}E\cap B(x,r)\cap\rho(\{y\in\mathbb{R}^{n}\colon|y|<r\})\big{)}\geq\theta r^{n}.

2.2. Carleson measures, non-tangential convergence, BMO and local BV

Given a domain Ωn+1\Omega\subset\mathbb{R}^{n+1}, we set σn|Ω\sigma\coloneqq\mathcal{H}^{n}|_{\partial\Omega} and δ(X)dist(X,Ω)\delta(X)\coloneqq\operatorname{dist}(X,\partial\Omega) for XΩX\in\Omega.

Definition 2.7 (Carleson measures).

We say that a Borel measure μ\mu in Ω\Omega is a Carleson measure (with respect to Ω\partial\Omega) if we have

CμsupxΩ,r>0μ(B(x,r)Ω)rn<.\displaystyle C_{\mu}\coloneqq\sup_{x\in\partial\Omega,r>0}\frac{\mu(B(x,r)\cap\Omega)}{r^{n}}<\infty.

We call CμC_{\mu} the Carleson norm of μ\mu.

Definition 2.8 (Cones and non-tangential convergence).

Suppose that m>1m>1. For every xΩx\in\partial\Omega, the cone of mm-aperture at xx is the set

(2.9) Γ~(x)Γ~m(x){ZΩ:dist(Z,x)<mdist(Z,Ω)}.\displaystyle\widetilde{\Gamma}(x)\coloneqq\widetilde{\Gamma}^{m}(x)\coloneqq\{Z\in\Omega\colon\operatorname{dist}(Z,x)<m\operatorname{dist}(Z,\partial\Omega)\}.

Let GG be a function defined in Ω\Omega and gg be a function defined on Ω\partial\Omega. We say that GG converges non-tangentially to gg at xΩx\in\partial\Omega if there exists m>1m>1 such that we have limkG(Yk)=g(x)\lim_{k\to\infty}G(Y_{k})=g(x) for every sequence (Yk)(Y_{k}) in Γ~m(x)\widetilde{\Gamma}^{m}(x) such that limkYk=x\lim_{k\to\infty}Y_{k}=x. We denote this by limYx,n.t.G(Y)=g(x)\lim_{Y\to x,\,\operatorname{n.t.}}G(Y)=g(x).

Definition 2.10 (Non-tangential maximal operator).

We denote the non-tangential maximal operator by NN_{*}, that is, for a function uL(Ω)u\in L^{\infty}(\Omega), the function Nu:ΩN_{*}u\colon\partial\Omega\to\mathbb{R} is defined as

Nu(x)=supXΓ~(x)|u(X)|.\displaystyle N_{*}u(x)=\sup_{X\in\widetilde{\Gamma}(x)}|u(X)|.

We call NuN_{*}u the non-tangential maximal function of uu.

Remark 2.11.

The aperture constant mm in Definition 2.8 does not play a big role in this paper and therefore we do not analyze it in detail for our results; we simply have that the results hold for some uniform aperture constant. Naturally, the aperture constant affects the values of the non-tangential maximal function but since the LpL^{p} norms of non-tangential maximal functions that are defined with different aperture constants are comparable (with the comparability constant depending on the aperture constants) (see [FS72, Lemma 1] and [HT20, Lemma 1.10]), this is not important for us. In most computations, it is convenient to use dyadic cones (see (2.17)) instead of cones of the previous type.

Definition 2.12 (BMO).

The space BMO(Ω)\operatorname{BMO}(\partial\Omega) (bounded mean oscillation) consists of fLloc1(Ω)f\in L^{1}_{\operatorname{loc}}(\partial\Omega) with

fBMOsupΔΔ|f(y)fΔ|𝑑σ(y)<,\displaystyle\|f\|_{\operatorname{BMO}}\coloneqq\sup_{\Delta}\fint_{\Delta}|f(y)-\langle f\rangle_{\Delta}|\,d\sigma(y)<\infty,

where the supremum is taken over all surface balls Δ=Δ(x,r)B(x,r)Ω\Delta=\Delta(x,r)\coloneqq B(x,r)\cap\partial\Omega. We denote fBMOc(Ω)f\in\operatorname{BMO}_{\operatorname{c}}(\partial\Omega) if ff is a BMO\operatorname{BMO} function with compact support.

Definition 2.13 (Local BV).

We say that locally integrable function ff has locally bounded variation in Ω\Omega (denote fBVloc(Ω)f\in\operatorname{BV}_{\operatorname{loc}}(\Omega)) if for any open relatively compact set ΩΩ\Omega^{\prime}\subset\Omega the total variation over Ω\Omega^{\prime} is finite:

Ω|f(Y)|𝑑YsupΨC01(Ω)ΨL(Ω)1Ωf(Y)divΨ(Y)𝑑Y<,\displaystyle\iint_{\Omega^{\prime}}|\nabla f(Y)|\,dY\coloneqq\sup_{\begin{subarray}{c}\overrightarrow{\Psi}\in C_{0}^{1}(\Omega^{\prime})\\ \|\overrightarrow{\Psi}\|_{L^{\infty}(\Omega^{\prime})\leq 1}\end{subarray}}\iint_{\Omega^{\prime}}f(Y)\,\text{div}\overrightarrow{\Psi}(Y)\,dY<\infty,

where C01(Ω)C_{0}^{1}(\Omega^{\prime}) is the class of compactly supported continuously differentiable vector fields in Ω\Omega^{\prime}.

2.3. Dyadic cubes, Whitney regions and approximating domains

An nn-Ahlfors regular set En+1E\subset\mathbb{R}^{n+1} equipped with the Euclidean distance and surface measure can be viewed as a space of homogeneous type of Coifman and Weiss [CW71], with ambient dimension n+1n+1. All such sets can be decomposed dyadically in the following sense:

Lemma 2.14 ([Chr90, DS91, HK12]).

Assume that En+1E\subset\mathbb{R}^{n+1} is nn-Ahlfors regular. Then EE admits a dyadic decomposition in the sense that there exist constants a1a0>0a_{1}\geq a_{0}>0 such that for each kk\in\mathbb{Z} there exists a collection of Borel sets, 𝔻k\mathbb{D}_{k}, which we will call (dyadic) cubes, such that

𝔻k{QjkE:jk},\mathbb{D}_{k}\coloneqq\{Q_{j}^{k}\subset E\colon j\in\mathfrak{I}_{k}\},

where k\mathfrak{I}_{k} denotes a countable index set depending on kk, satisfying

  1. (i)

    for each fixed kk\in\mathbb{Z}, the sets QjkQ_{j}^{k} are disjoint and E=jQjkE=\cup_{j}Q_{j}^{k}\,\,,

  2. (ii)

    if mkm\geq k then either QimQjkQ_{i}^{m}\subset Q_{j}^{k} or QimQjk=ØQ_{i}^{m}\cap Q_{j}^{k}=\mbox{{\O}},

  3. (iii)

    for each kk\in\mathbb{Z}, jkj\in\mathfrak{I}_{k} and m<km<k, there is a unique imi\in\mathfrak{I}_{m} such that QjkQimQ_{j}^{k}\subset Q_{i}^{m},

  4. (iv)

    diam(Qjk)a12k\operatorname{diam}(Q_{j}^{k})\leq a_{1}2^{-k},

  5. (v)

    for each QjkQ_{j}^{k}, there exists a point zjkQjkz_{j}^{k}\in Q_{j}^{k} such that EB(zjk,a02k)QjkEB(zjk,a12k)E\cap B(z^{k}_{j},a_{0}2^{-k})\subset Q_{j}^{k}\subset E\cap B(z^{k}_{j},a_{1}2^{-k}).

We denote by 𝔻=𝔻(E)\mathbb{D}=\mathbb{D}(E) the collection of all cubes QjkQ^{k}_{j}, that is,

𝔻k𝔻k.\mathbb{D}\coloneqq\cup_{k}\mathbb{D}_{k}.

If EE is bounded, we ignore cubes where 2kdiam(Ω)2^{-k}\gtrsim\operatorname{diam}(\partial\Omega) (in particular, where a02kdiam(Ω)a_{0}2^{-k}\geq\operatorname{diam}(\partial\Omega)). Given a cube Q=Qjk𝔻Q=Q_{j}^{k}\in\mathbb{D}, we define the side-length of QQ as (Q)2k\ell(Q)\coloneqq 2^{-k}. By Ahlfors regularity and property (v) in Lemma 2.14, we know that (Q)diam(Q)\ell(Q)\approx\operatorname{diam}(Q) and n(Q)(Q)n\mathcal{H}^{n}(Q)\approx\ell(Q)^{n}. Given Q𝔻Q\in\mathbb{D} and mm\in\mathbb{Z}, we set

𝔻Q{Q𝔻:QQ},𝔻m,Q{Q𝔻Q:(Q)=2m(Q)}.\mathbb{D}_{Q}\coloneqq\left\{Q^{\prime}\in\mathbb{D}\colon Q^{\prime}\subseteq Q\right\},\qquad\mathbb{D}_{m,Q}\coloneqq\left\{Q^{\prime}\in\mathbb{D}_{Q}\colon\ell(Q^{\prime})=2^{-m}\ell(Q)\right\}.

We call the cubes in the collection 𝔻1,Q\mathbb{D}_{1,Q} the children of QQ. Notice that by Ahlfors regularity and property (v) in Lemma 2.14, each cube has a uniformly bounded number of children.

Given a cube Q=Qjk𝔻Q=Q_{j}^{k}\in\mathbb{D}, we call the point zjkQz_{j}^{k}\in Q in property (v) in Lemma 2.14 the center of QQ, denote xQzjkx_{Q}\coloneqq z_{j}^{k}, and set

ΔQEB(zjk,a0(Q)).\Delta_{Q}\coloneqq E\cap B(z_{j}^{k},a_{0}\ell(Q)).

We use Lemma 2.14 to decompose Ω\partial\Omega, so that 𝔻(E)=𝔻(Ω)𝔻\mathbb{D}(E)=\mathbb{D}(\partial\Omega)\eqqcolon\mathbb{D}. For each Q𝔻Q\in\mathbb{D}, we let XQX_{Q} be the corkscrew point relative to xQx_{Q} at scale 105a0(Q)10^{-5}a_{0}\ell(Q). We have B(XQ,γ105a0(Q))B(xQ,105a0(Q))ΩB(X_{Q},\gamma 10^{-5}a_{0}\ell(Q))\subset B(x_{Q},10^{-5}a_{0}\ell(Q))\cap\Omega, where γ\gamma is the corkscrew constant in Definition 2.1.

For many of our techniques, it is important that we show that some collections of dyadic cubes are quantitatively small in the following sense:

Definition 2.15 (Carleson packing condition).

Let 𝔻\mathbb{D} be a dyadic system on Ω\partial\Omega and let 𝒜𝔻\mathcal{A}\subset\mathbb{D}. We say that 𝒜\mathcal{A} satisfies a Carleson packing condition if there exists a constant C1C\geq 1 such that for any Q0𝔻Q_{0}\in\mathbb{D} we have

Q𝒜,QQ0σ(Q)Cσ(Q0).\displaystyle\sum_{Q\in\mathcal{A},Q\subset Q_{0}}\sigma(Q)\leq C\sigma(Q_{0}).

We denote the smallest such constant CC by 𝒞𝒜\mathcal{C}_{\mathcal{A}}.

Next, we use a standard decomposition of Ω\Omega into Whitney cubes (see e.g. [Ste70, Chapter VI]), and then associate a collection of such Whitney cubes to each boundary cube to construct suitable Whitney-type regions. These Whitney regions are modeled after regions of the type Q×((Q)/2,(Q))Q\times(\ell(Q)/2,\ell(Q)) (that is, the upper halves of Carleson boxes) in the simpler geometry of the upper half-space. For this, we recall the construction found in [HM14] noting that we make some changes to the notation therein (following the notation of more recent papers, e.g. [HMM16]). We let 𝒲={I}I\mathcal{W}=\{I\}_{I} denote a Whitney decomposition of Ω\Omega, with the properties that each II is a closed (n+1)(n+1)-dimensional cube satisfying

4diam(I)dist(4I,Ω)dist(I,Ω)40diam(I),\displaystyle 4\operatorname{diam}(I)\leq\operatorname{dist}(4I,\partial\Omega)\leq\operatorname{dist}(I,\partial\Omega)\leq 40\operatorname{diam}(I),

where 4I4I is the standard concentric Euclidean dilate of a cube; the interiors of the cubes II are disjoint, and for all I1,I2𝒲I_{1},I_{2}\in\mathcal{W} with I1I2ØI_{1}\cap I_{2}\neq\mbox{{\O}} we have

14diam(I1)diam(I2)4diam(I1).\displaystyle\frac{1}{4}\operatorname{diam}(I_{1})\leq\operatorname{diam}(I_{2})\leq 4\operatorname{diam}(I_{1}).

For I𝒲I\in\mathcal{W} we let (I)\ell(I) denote the side length of II.

For each cube Q𝔻Q\in\mathbb{D} and constant KK0K\geq K_{0}, with K0K_{0} to be described momentarily, we associate an initial collection of Whitney cubes

WQ(K){I𝒲:K1(I)(Q)K(I),dist(I,Q)K(Q)}.W_{Q}(K)\coloneqq\{I\in\mathcal{W}\colon K^{-1}\ell(I)\leq\ell(Q)\leq K\ell(I),\operatorname{dist}(I,Q)\leq K\ell(Q)\}.

We choose K0K_{0} depending on the constants in the corkscrew condition and the Ahlfors regularity condition, insisting on two conditions being met:

  1. (1)

    If XΩX\in\Omega with dist(X,Ω)105diam(Ω)\operatorname{dist}(X,\partial\Omega)\leq 10^{5}\operatorname{diam}(\partial\Omega) then XIWQ(K)X\in I\in W_{Q}(K) for some Q𝔻Q\in\mathbb{D}.

  2. (2)

    For any Q𝔻Q\in\mathbb{D}, we have B(XQ,dist(XQ,Ω)/2)IWQ(K)IB(X_{Q},\operatorname{dist}(X_{Q},\partial\Omega)/2)\subseteq\cup_{I\in W_{Q}(K)}I, and if Q𝔻Q^{\prime}\in\mathbb{D} is another cube such that QQQ^{\prime}\subset Q with (Q)=12(Q)\ell(Q^{\prime})=\tfrac{1}{2}\ell(Q), then we also have B(XQ,dist(XQ,Ω)/2)IWQ(K)IB(X_{Q^{\prime}},\operatorname{dist}(X_{Q^{\prime}},\partial\Omega)/2)\subseteq\cup_{I\in W_{Q}(K)}I.

Of course, condition (1) above is automatically satisfied if diam(Ω)=\operatorname{diam}(\partial\Omega)=\infty.

Following [HM14, Section 3], we augment the collection 𝒲Q(K)\mathcal{W}_{Q}(K) as follows. For each IWQ(K)I\in W_{Q}(K), we take a Harnack chain H(I)H(I) from the center of II to the corkscrew point XQX_{Q}, and we let WQ,I(K)W_{Q,I}(K) be be the collection of Whitney cubes in 𝒲\mathcal{W} that meet at least one ball in the chain H(I)H(I). We then set 𝒲Q(K)IWQ(K)WQ,I(K)\mathcal{W}^{*}_{Q}(K)\coloneqq\bigcup_{I\in W_{Q}(K)}W_{Q,I}(K). Finally, for a small dimensional parameter τ\tau (this is the parameter λ\lambda in [HM14, Section 3]), we define the Whitney region relative to QQ as

UQ=UQ(K0)I𝒲Q(K0)(1+τ)I.\displaystyle U_{Q}=U_{Q}(K_{0})\coloneqq\bigcup_{I\in\mathcal{W}^{*}_{Q}(K_{0})}(1+\tau)I.

By construction, we know that if XUQX\in U_{Q}, then

(2.16) K11(Q)dist(X,Ω)dist(X,Q)K1(Q),K_{1}^{-1}\ell(Q)\leq\operatorname{dist}(X,\partial\Omega)\leq\operatorname{dist}(X,Q)\leq K_{1}\ell(Q),

where K1K_{1} depends on K0K_{0}, the dimension and the Harnack chain condition. For κK0\kappa\gg K_{0} to be chosen, we also define the following fattened versions of the Whitney regions:

UQ=UQ(κ)=I𝒲Q(κ)(1+τ)I,UQ=I𝒲Q(κ)(1+2τ)I,U_{Q}^{*}=U_{Q}(\kappa)=\bigcup_{I\in\mathcal{W}^{*}_{Q}(\kappa)}(1+\tau)I,\qquad U_{Q}^{**}=\bigcup_{I\in\mathcal{W}^{*}_{Q}(\kappa)}(1+2\tau)I,

that is, UQU_{Q}^{*} is constructed the same way as UQU_{Q} but we replace the constant K0K_{0} by κ\kappa (similarly for UQU_{Q}^{**}). We describe the reasoning and choice of κ\kappa in the next subsection. We note that for τ\tau small enough the regions UQU_{Q}, UQU^{*}_{Q} and UQU_{Q}^{**} have bounded overlaps, that is, for a collection of dyadic cubes 𝒬\mathscr{Q} and the (n+1)(n+1)-dimensional Lebesgue measure |||\cdot| we have |Q𝒬UQ|Q𝒬|UQ||\bigcup_{Q\in\mathscr{Q}}U_{Q}^{**}|\approx\sum_{Q\in\mathscr{Q}}|U_{Q}^{**}|.

Using the Whitney regions above, we can now define objects like sawtooth regions, Carleson boxes and dyadic cones. Let Q0𝔻Q_{0}\in\mathbb{D} be a fixed cube and 𝔻Q0\mathcal{F}\subset\mathbb{D}_{Q_{0}} a collection of pairwise disjoint cubes. We set

𝔻,Q0𝔻Q0Q𝔻Q.\mathbb{D}_{\mathcal{F},Q_{0}}\coloneqq\mathbb{D}_{Q_{0}}\setminus\cup_{Q\in\mathcal{F}}\mathbb{D}_{Q}.

We then define the local sawtooth relative to \mathcal{F} (and its fattened version) as

Ω,Q0int(Q𝔻,Q0UQ),Ω,Q0int(Q𝔻,Q0UQ).\Omega_{\mathcal{F},Q_{0}}\coloneqq\operatorname{int}\Big{(}\bigcup_{Q\in\mathbb{D}_{\mathcal{F},Q_{0}}}U_{Q}\Big{)},\qquad\Omega^{*}_{\mathcal{F},Q_{0}}\coloneqq\operatorname{int}\Big{(}\bigcup_{Q\in\mathbb{D}_{\mathcal{F},Q_{0}}}U^{*}_{Q}\Big{)}.

In the special case where =Ø\mathcal{F}=\mbox{{\O}}, we write TQ0=Ω,Q0T_{Q_{0}}=\Omega_{\mathcal{F},Q_{0}} and TQ0=Ω,QT^{*}_{Q_{0}}=\Omega^{*}_{\mathcal{F},Q}, that is,

TQ0int(Q𝔻Q0UQ),TQ0int(Q𝔻Q0UQ),TQ0int(Q𝔻Q0UQ)T_{Q_{0}}\coloneqq\operatorname{int}\Big{(}\bigcup_{Q\in\mathbb{D}_{Q_{0}}}U_{Q}\Big{)},\qquad T_{Q_{0}}^{*}\coloneqq\operatorname{int}\Big{(}\bigcup_{Q\in\mathbb{D}_{Q_{0}}}U^{*}_{Q}\Big{)},\qquad T_{Q_{0}}^{**}\coloneqq\operatorname{int}\Big{(}\bigcup_{Q\in\mathbb{D}_{Q_{0}}}U^{**}_{Q}\Big{)}

We call TQ0T_{Q_{0}} the Carleson box relative to Q0Q_{0} and TQ0T_{Q_{0}}^{*} and TQ0T_{Q_{0}}^{**} its fattened versions. Given a cube Q0𝔻Q_{0}\in\mathbb{D} and a point xQ0x\in Q_{0}, we also define the (truncated) dyadic cone at xΩx\in\partial\Omega, Γ(x)\Gamma(x), by setting

(2.17) Γ(x)ΓQ0(x)int(Q𝔻Q0:xQUQ).\displaystyle\Gamma(x)\coloneqq\Gamma_{Q_{0}}(x)\coloneqq\operatorname{int}\Big{(}\bigcup_{Q\in\mathbb{D}_{Q_{0}}\colon x\in Q}U_{Q}\Big{)}.

Notice that Γ(x)=Ωx,Q0\Gamma(x)=\Omega_{\mathcal{F}_{x},Q_{0}}, where x\mathcal{F}_{x} is the collection of maximal222Since the sizes of the cubes in {Q𝔻Q0:xQ}\{Q\in\mathbb{D}_{Q_{0}}\colon x\notin Q\} are bounded from above, we can always choose the maximal cubes. In the case of non-truncated dyadic cones (that is, cones of the form int(Q𝔻:xQUQ)\operatorname{int}\left(\bigcup_{Q\in\mathbb{D}\colon x\in Q}U_{Q}\right)), we can choose the maximal cubes if the dyadic system 𝔻\mathbb{D} forms a tree-like structure. This structure can be achieved by choosing the center points of the dyadic cubes in a suitable way (see, for example, [HT14, Theorem 2.4]). (and hence, disjoint) cubes in the collection {Q𝔻Q0:xQ}\{Q\in\mathbb{D}_{Q_{0}}\colon x\notin Q\}. It is straightforward to verify that there exists uniform constants m1,m2>1m_{1},m_{2}>1 such that Γ(x)\Gamma(x) contains a truncated version of the cone Γ~m1(x)\widetilde{\Gamma}^{m_{1}}(x) and it is contained in a truncated version of the cone Γ~m2(x)\widetilde{\Gamma}^{m_{2}}(x), where Γ~m1(x)\widetilde{\Gamma}^{m_{1}}(x) and Γ~m2(x)\widetilde{\Gamma}^{m_{2}}(x) are cones of the type (2.9). Thus, since the aperture of the cones is not important for us, we mostly use dyadic cones when studying non-tangential convergence.

By the following lemma, the Whitney regions, sawtooth regions, Carleson boxes, and truncated dyadic cones inherit many quantitative geometric properties from Ω\Omega:

Lemma 2.18 ([HM14, Lemma 3.61]).

Suppose that Ωn+1\Omega\subset\mathbb{R}^{n+1} is a uniform domain with nn-Ahlfors regular boundary. Let Q0𝔻(Ω)Q_{0}\in\mathbb{D}(\partial\Omega) be a cube and 𝔻Q0\mathcal{F}\subset\mathbb{D}_{Q_{0}} be a collection of pairwise disjoint cubes. Then Ω,Q\Omega_{\mathcal{F},Q} and Ω,Q\Omega^{*}_{\mathcal{F},Q} are uniform domains with nn-Ahlfors regular boundary whose structural constants depend only on the dimension, the structural constants of Ω\Omega, and the constant κ\kappa. In particular, the Whitney regions UQU_{Q}, UQU_{Q}^{*} and UQU_{Q}^{**}, the Carleson boxes TQT_{Q}, TQT_{Q}^{*} and TQT_{Q}^{**} and the truncated dyadic cones Γ(x)\Gamma(x) are uniform domains with nn-Ahlfors regular boundaries, with structural constants depending only on the dimension and the structural constants of Ω\Omega and the constant κ\kappa.

2.4. The choice of the parameter κ\kappa

In contrast to the setting of the upper half space, we do not define the sawtooths by removing Whitney regions. This is due to the overlaps of the regions UQU_{Q}: we may encounter situations where for Q0𝔻(Ω)Q_{0}\in\mathbb{D}(\partial\Omega) and a collection of pairwise disjoint cubes 𝔻Q0\mathcal{F}\subset\mathbb{D}_{Q_{0}} there exists a cube Q𝔻,Q0Q\in\mathbb{D}_{\mathcal{F},Q_{0}} such that UQ¯\overline{U_{Q}} does not contribute to the boundary of Ω,Q\Omega_{\mathcal{F},Q}. That being said, if κ\kappa is chosen large enough, then the fattened Whitney region UQU_{Q}^{*} meets the boundary of the unfattened region Ω,Q\Omega_{\mathcal{F},Q} on a portion roughly the measure of QQ. We will consider this in Section 4 where it will be convenient for us, but we prove the technical estimates that give this property below.

Let us fix a cube Q𝔻Q\in\mathbb{D}. Recall that XQX_{Q} is a corkscrew point relative to xQQx_{Q}\in Q at scale rQ105a0(Q)r_{Q}\coloneqq 10^{-5}a_{0}\ell(Q) and ΔQ=B(xQ,105rQ)ΩQ\Delta_{Q}=B(x_{Q},10^{5}r_{Q})\cap\partial\Omega\subset Q is the surface ball associated to QQ. We let x^QΩ\hat{x}_{Q}\in\partial\Omega denote a touching point for XQX_{Q}, that is, a point such that |XQx^Q|=δ(XQ)|X_{Q}-\hat{x}_{Q}|=\delta(X_{Q}). By triangle inequality, |xQx^Q|<2rQ|x_{Q}-\hat{x}_{Q}|<2r_{Q}, and thus,

(2.19) Δ^Q=B(x^Q,103rQ)ΩΔQQ.\hat{\Delta}_{Q}=B(\hat{x}_{Q},10^{3}r_{Q})\cap\partial\Omega\subset\Delta_{Q}\subseteq Q.

For every θ(0,1)\theta\in(0,1), we let

PQ(θ)x^Q+θ(XQx^Q)\displaystyle P_{Q}(\theta)\coloneqq\hat{x}_{Q}+\theta(X_{Q}-\hat{x}_{Q})

be the “θ\theta-point” on the directed line segment from x^Q\hat{x}_{Q} to XQX_{Q}. Then, by definitions,

(2.20) γθrQ|PQ(θ)x^Q|=dist(PQ(θ),Ω)θrQ.\gamma\theta r_{Q}\leq|P_{Q}(\theta)-\hat{x}_{Q}|=\operatorname{dist}(P_{Q}(\theta),\partial\Omega)\leq\theta r_{Q}.
Lemma 2.21.

There exists θ0(0,1)\theta_{0}\in(0,1) depending on K0K_{0} and the structural constants such that if for some Q𝔻Q^{\prime}\in\mathbb{D} we have that

B(PQ(θ0),γθ010rQ)UQØ,\displaystyle B\big{(}P_{Q}(\theta_{0}),\tfrac{\gamma\theta_{0}}{10}r_{Q}\big{)}\cap U_{Q^{\prime}}\neq\mbox{{\O}},

then QQQ^{\prime}\subseteq Q and (Q)<(Q)\ell(Q^{\prime})<\ell(Q).

Proof.

Suppose that θ01/4C(K1)2\theta_{0}\leq 1/4C(K_{1})^{2} for a large structural constant C1C\geq 1 and

(2.22) XB(PQ(θ0),γθ010rQ)UQ\displaystyle X\in B\big{(}P_{Q}(\theta_{0}),\tfrac{\gamma\theta_{0}}{10}r_{Q}\big{)}\cap U_{Q^{\prime}}

for some QQ^{\prime}, where K1K_{1} is the constant in (2.16). By (2.16), (2.22) and (2.20), it holds that

(Q)\displaystyle\ell(Q^{\prime}) K1dist(X,Ω)\displaystyle\leq K_{1}\operatorname{dist}(X,\partial\Omega)
(2.23) K1(dist(PQ(θ0),Ω)+γθ010rQ)2K1θ0rQ=2K1θ0105a0(Q).\displaystyle\leq K_{1}\bigg{(}\operatorname{dist}(P_{Q}(\theta_{0}),\partial\Omega)+\frac{\gamma\theta_{0}}{10}r_{Q}\bigg{)}\leq 2K_{1}\theta_{0}r_{Q}=2K_{1}\theta_{0}10^{-5}a_{0}\ell(Q).

In particular, we have (Q)<(Q)\ell(Q^{\prime})<\ell(Q).

To show that QQQ^{\prime}\subset Q, we first notice that we have

|XxQ|CK1(Q)\displaystyle|X-x_{Q^{\prime}}|\leq CK_{1}\ell(Q^{\prime})

for a structural constant C1C\geq 1 by (2.16) and the fact that diam(Q)(Q)\operatorname{diam}(Q^{\prime})\approx\ell(Q^{\prime}). This and (2.23) then give us

|XxQ|2C(K1)2θ0105a0(Q).\displaystyle|X-x_{Q^{\prime}}|\leq 2C(K_{1})^{2}\theta_{0}10^{-5}a_{0}\ell(Q).

Thus, by (2.20) and (2.22), it holds that

|Xx^Q|2θ0rQ=2θ0105a0(Q).\displaystyle|X-\hat{x}_{Q}|\leq 2\theta_{0}r_{Q}=2\theta_{0}10^{-5}a_{0}\ell(Q).

Combining the previous two inequalities then gives us

(2.24) |x^QxQ|4C(K1)2θ0105a0(Q)=4C(K1)2θ0rQ.\displaystyle|\hat{x}_{Q}-x_{Q^{\prime}}|\leq 4C(K_{1})^{2}\theta_{0}10^{-5}a_{0}\ell(Q)=4C(K_{1})^{2}\theta_{0}r_{Q}.

In particular,

|xQxQ||xQx^Q|+|x^QxQ|4C(K1)2θ0rQ+2rQ<3rQ<a0(Q),\displaystyle|x_{Q^{\prime}}-x_{Q}|\leq|x_{Q^{\prime}}-\hat{x}_{Q}|+|\hat{x}_{Q}-x_{Q}|\leq 4C(K_{1})^{2}\theta_{0}r_{Q}+2r_{Q}<3r_{Q}<a_{0}\ell(Q),

by (2.24), the fact that |x^QxQ|<2rQ|\hat{x}_{Q}-x_{Q}|<2r_{Q}, and the choice θ01/4C(K1)2\theta_{0}\leq 1/4C(K_{1})^{2}. Thus, xQΔQQx_{Q^{\prime}}\in\Delta_{Q}\subset Q. Since QQØQ^{\prime}\cap Q\neq\mbox{{\O}} and (Q)<(Q)\ell(Q^{\prime})<\ell(Q), we know that QQQ^{\prime}\subset Q, which is what we wanted. ∎

Let us then fix θ0\theta_{0} so that Lemma 2.21 holds. For Q𝔻Q\in\mathbb{D}, we set

ΞQθ[θ0,1]B(PQ(θ),γθ010rQ),\displaystyle\Xi_{Q}\coloneqq\bigcup_{\theta\in[\theta_{0},1]}B\big{(}P_{Q}(\theta),\tfrac{\gamma\theta_{0}}{10}r_{Q}\big{)},

which is a cylinder-like object. We get the following straightforward lemma:

Lemma 2.25.

Let Q𝔻Q\in\mathbb{D} be a fixed cube and let 𝒬Q\mathscr{Q}_{Q} be the collection of cubes that share the same dyadic parent as QQ, that is,

𝒬Q{P𝔻:P,QQ0 for a cube Q0𝔻 such that (P)=(Q)=12(Q0)}.\displaystyle\mathscr{Q}_{Q}\coloneqq\{P\in\mathbb{D}\colon P,Q\subset Q_{0}\text{ for a cube }Q_{0}\in\mathbb{D}\text{ such that }\ell(P)=\ell(Q)=\tfrac{1}{2}\ell(Q_{0})\}.

Let κmax{K0,(θ0)1}\kappa\gg\max\{K_{0},(\theta_{0})^{-1}\} and XΞPX\in\Xi_{P} for some P𝒬QP\in\mathscr{Q}_{Q}. Then XUQX\in U_{Q}^{*}.

Proof.

Let κmax{K0,(θ0)1}\kappa\gg\max\{K_{0},(\theta_{0})^{-1}\} and XΞPX\in\Xi_{P} for some P𝒬QP\in\mathscr{Q}_{Q}. By the Whitney decomposition, there exists a Whitney cube I𝒲I\in\mathcal{W} such that XIX\in I. By the definition of UQU_{Q}^{*}, it is enough to show that I𝒲Q(κ)I\in\mathcal{W}_{Q}^{*}(\kappa).

By (2.20) and the definitions, we first notice that

(I)dist(X,Ω)γθ0rPθ0(P)=θ0(Q)\displaystyle\ell(I)\approx\operatorname{dist}(X,\partial\Omega)\approx\gamma\theta_{0}r_{P}\approx\theta_{0}\ell(P)=\theta_{0}\ell(Q)

with uniformly bounded implicit constants. In particular, since θ01κ\theta_{0}\gg\tfrac{1}{\kappa} and κ1\kappa\gg 1, we get

1κ(I)(Q)κ(I).\displaystyle\frac{1}{\kappa}\ell(I)\leq\ell(Q)\leq\kappa\ell(I).

On the other hand, since P𝒬QP\in\mathscr{Q}_{Q}, we know that

dist(I,Q)dist(I,P)+(Q),\displaystyle\operatorname{dist}(I,Q)\lesssim\operatorname{dist}(I,P)+\ell(Q),

and by (2.20), (2.19) and the fact that XΞPX\in\Xi_{P}, we know that

dist(I,P)dist(X,P)|Xx^P|γθ010rP+|XPx^P|2rP2(P)=2(Q).\displaystyle\operatorname{dist}(I,P)\leq\operatorname{dist}(X,P)\leq|X-\hat{x}_{P}|\leq\frac{\gamma\theta_{0}}{10}r_{P}+|X_{P}-\hat{x}_{P}|\leq 2r_{P}\leq 2\ell(P)=2\ell(Q).

In particular, since κ1\kappa\gg 1, we have

dist(I,Q)κ(Q).\displaystyle\operatorname{dist}(I,Q)\leq\kappa\ell(Q).

Thus, I𝒲Q(κ)I\in\mathcal{W}_{Q}^{*}(\kappa), which proves the claim. ∎

Let us also record the following simple lemma for future use:

Lemma 2.26.

Let Q𝔻Q\in\mathbb{D} and let XQΩX_{Q}\in\Omega be a corkscrew point, x^QΩ\hat{x}_{Q}\in\partial\Omega be a touching point and rQ=105a0(Q)r_{Q}=10^{-5}a_{0}\ell(Q) as above. If YB(PQ(θ),rQ)Y\in B(P_{Q}(\theta),r_{Q}) for some θ[0,1]\theta\in[0,1] and y^Ω\hat{y}\in\partial\Omega is a point such that |Yy^|=dist(Y,Ω)|Y-\hat{y}|=\operatorname{dist}(Y,\partial\Omega), then y^Δ^QΔQQ\hat{y}\in\hat{\Delta}_{Q}\subseteq\Delta_{Q}\subseteq Q.

Proof.

By definitions and using (2.20) several times, we get

|y^x^Q||y^Y|+|YPQ(θ)|+|PQ(θ)x^Q|<dist(Y,Ω)+2rQ|YPQ(θ)|+dist(PQ(θ),Ω)+2rQ<rQ+rQ+2rQ=4rQ,|\hat{y}-\hat{x}_{Q}|\leq|\hat{y}-Y|+|Y-P_{Q}(\theta)|+|P_{Q}(\theta)-\hat{x}_{Q}|<\operatorname{dist}(Y,\partial\Omega)+2r_{Q}\\ \leq|Y-P_{Q}(\theta)|+\operatorname{dist}(P_{Q}(\theta),\partial\Omega)+2r_{Q}<r_{Q}+r_{Q}+2r_{Q}=4r_{Q},

and thus, y^Δ^QΔQQ\hat{y}\in\hat{\Delta}_{Q}\subset\Delta_{Q}\subset Q by (2.19). ∎

2.5. Elliptic PDE estimates

Here we collect some of the standard estimates for divergence form elliptic operators with real coefficients that will be used throughout the paper. In this section, Ω\Omega always denotes a uniform domain in n+1\mathbb{R}^{n+1}, n1n\geq 1, with nn-Ahlfors regular boundary. We recall that a divergence form elliptic operator is of the form

L()div(A),\displaystyle L(\cdot)\coloneqq-\mathop{\operatorname{div}}\nolimits(A\nabla\cdot),

viewed in the weak sense, where AA is a uniformly elliptic matrix, that is, A=(ai,j)i,j=1n+1A=(a_{i,j})_{i,j=1}^{n+1} is an (n+1)×(n+1)(n+1)\times(n+1) matrix-valued function on n+1\mathbb{R}^{n+1} and there exists a constant Λ\Lambda, the ellipticity parameter, such that

Λ1|ξ|2A(X)ξξ,andai,jL(n+1)Λ,\displaystyle\Lambda^{-1}|\xi|^{2}\leq A(X)\xi\cdot\xi,\qquad\text{and}\qquad\|a_{i,j}\|_{L^{\infty}(\mathbb{R}^{n+1})}\leq\Lambda,

for all ξ,ζn+1\xi,\zeta\in\mathbb{R}^{n+1} and almost every XΩX\in\Omega. We say that a constant depends on ellipticity if it depends on Λ\Lambda. Given an open set 𝒪n+1\mathcal{O}\subset\mathbb{R}^{n+1} we say a function uWloc1,2(𝒪)u\in W^{1,2}_{\operatorname{loc}}(\mathcal{O}) is a solution to Lu=0Lu=0 in 𝒪\mathcal{O} if

𝒪AuφdX=0, for every φCc(𝒪).\displaystyle\iint_{\mathcal{O}}A\nabla u\cdot\nabla\varphi\,dX=0,\quad\text{ for every }\varphi\in C_{\operatorname{c}}^{\infty}(\mathcal{O}).

The most fundamental estimate for solutions to divergence form elliptic equations is the following local energy inequality.

Lemma 2.27 (Caccioppoli Inequality).

Let L=divAL=-\mathop{\operatorname{div}}\nolimits A\nabla be a divergence form elliptic operator and uu a solution to Lu=0Lu=0 in an open set 𝒪\mathcal{O}. If a>0a>0 and BB is a ball such that (1+a)B𝒪(1+a)B\subset\mathcal{O} then

B|u|𝑑Xr2(1+a)Bu2𝑑X,\displaystyle\iint_{B}|\nabla u|\,dX\lesssim r^{-2}\iint_{(1+a)B}u^{2}\,dX,

where the implicit constant depends only on aa, dimension and ellipticity.

Solutions to divergence form elliptic equations are locally Hölder continuous.

Lemma 2.28 (Hölder continuity of solutions, [DG57, Nas58]).

Let L=divAL=-\mathop{\operatorname{div}}\nolimits A\nabla be a divergence form elliptic operator and uu a non-negative solution to Lu=0Lu=0 in an open set 𝒪\mathcal{O}. Suppose that B=B(X0,R)B=B(X_{0},R) is a ball such that λBB(X0,2λR)𝒪\lambda B\coloneqq B(X_{0},2\lambda R)\subset\mathcal{O} for λ>1\lambda>1. Then we have

|u(X)u(Y)|C(|XY|λR)α(2λB|u|2𝑑Y)1/2 for all X,YB,\displaystyle|u(X)-u(Y)|\leq C\left(\frac{|X-Y|}{\lambda R}\right)^{\alpha}\left(\mathchoice{{\vbox{\hbox{$\textstyle\raisebox{-8.61108pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-7.09656pt}}{{\vbox{\hbox{$\scriptstyle\raisebox{-6.02777pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-6.09258pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-4.29826pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-3.54826pt}}\!\iint_{2\lambda B}|u|^{2}\,dY\right)^{1/2}\quad\text{ for all }X,Y\in B,

where α\alpha and CC depend only on dimension and ellipticity.

Another celebrated result is Moser’s Harnack inequality for non-negative solutions.

Lemma 2.29 (Harnack inequality [Mos61]).

Let L=divAL=-\mathop{\operatorname{div}}\nolimits A\nabla be a divergence form elliptic operator and uu a non-negative solution to Lu=0Lu=0 in an open set 𝒪\mathcal{O}. If BB is a ball such that 2B𝒪2B\subset\mathcal{O} then supBuCinfBu\sup_{B}u\leq C\inf_{B}u, where CC depends only on dimension and ellipticity.

We now turn our attention to the elliptic measure, for which we borrow the setting of [AGMT22]. Consider the compactified space ¯n+1=n+1{}\overline{\mathbb{R}}^{n+1}=\mathbb{R}^{n+1}\cup\{\infty\}; following [HKM93, Section 9], we will understand all topological notions with respect to this space. Hence, for instance, if Ω\Omega is unbounded, then Ω\infty\in\partial\Omega, and the functions in the space C(Ω)C(\partial\Omega) are assumed continuous and real-valued, so that all functions in C(Ω)C(\partial\Omega) lie in L(Ω)L^{\infty}(\partial\Omega) even if Ω\partial\Omega is unbounded.

Given a domain Ω\Omega and a divergence form elliptic operator LL, we let ωL,ΩX\omega_{L,\Omega}^{X} denote the elliptic measure with pole at XΩX\in\Omega. That is, by the Riesz representation theorem, for every XΩX\in\Omega there exists a probability measure ωL,ΩX\omega_{L,\Omega}^{X} such that if fCc(Ω)f\in C_{\operatorname{c}}(\partial\Omega), then the solution to Lu=0Lu=0 in Ω\Omega with uC(Ω¯)u\in C(\overline{\Omega}) and u=fu=f on Ω\partial\Omega, constructed via Perron’s method, satisfies

(2.30) u(X)=uf(X)=Ωf(y)𝑑ωL,ΩX(y).u(X)=u_{f}(X)=\int_{\partial\Omega}f(y)\,d\omega^{X}_{L,\Omega}(y).

When the context is clear, we simply denote ωXωL,ΩX\omega^{X}\coloneqq\omega^{X}_{L,\Omega} and, with slight abuse of terminology, call the family of elliptic measures ω=ωL=ωL,Ω{ωX}X\omega=\omega_{L}=\omega_{L,\Omega}\coloneqq\{\omega^{X}\}_{X} just the elliptic measure.

Our main results consider characterizations and implications given by quantitative absolute continuity of elliptic measure in the sense of Muckenhoupt’s AA_{\infty} condition [Muc72, CF74]:

Definition 2.31 (AA_{\infty} for elliptic measure).

Let LL be a divergence form elliptic operator in Ω\Omega. We say that the elliptic measure ω=ωL,Ω\omega=\omega_{L,\Omega} satisfies the AA_{\infty} condition with respect to surface measure (denote ωA(σ)\omega\in A_{\infty}(\sigma)) if there exist constants C1C\geq 1 and s>0s>0 such that if BB(x,r)B\coloneqq B(x,r) with xΩx\in\partial\Omega and r(0,diam(Ω)/4)r\in(0,\operatorname{diam}(\partial\Omega)/4) and AΔBΩA\subset\Delta\coloneqq B\cap\partial\Omega is a Borel set, then

ωY(A)C(σ(A)σ(Δ))sωY(Δ),for every YΩ4B.\displaystyle\omega^{Y}(A)\leq C\left(\frac{\sigma(A)}{\sigma(\Delta)}\right)^{s}\omega^{Y}(\Delta),\qquad\text{for every }Y\in\Omega\setminus 4B.

We refer to CC and ss here together as the “ωLA(σ)\omega_{L}\in A_{\infty}(\sigma) constants”.

Next we discuss the Green’s function and its properties.

Definition 2.32 (Green’s function).

Let L=divAL=-\operatorname{div}A\nabla be a not necessarily symmetric divergence form elliptic operator with bounded measurable coefficients. There exists a unique non-negative function G=GL=GL,Ω:Ω×ΩG=G_{L}=G_{L,\Omega}:\Omega\times\Omega\rightarrow\mathbb{R}, called Green’s function for LL, satisfying the following properties:

  1. (i)

    For each X,YΩX,Y\in\Omega,

    (2.33) 0G(X,Y){|XY|1n,n2,XY,1,n=1,|XY|δ(Y)/10,log(δ(Y)|XY|),n=1,|XY|δ(Y)/2.0\leq G(X,Y)\lesssim\left\{\begin{matrix}|X-Y|^{1-n},&\qquad n\geq 2,&\qquad X\neq Y,\\[2.84526pt] 1,&\qquad n=1,&\qquad|X-Y|\geq\delta(Y)/10,\\[2.84526pt] \log\big{(}\frac{\delta(Y)}{|X-Y|}\big{)},&\qquad n=1,&\qquad|X-Y|\leq\delta(Y)/2.\end{matrix}\right.
  2. (ii)

    For every a(0,1)a\in(0,1) there exists cac_{a} such that

    (2.34) G(X,Y){ca|XY|1n,n2,|XY|aδ(Y),calog(δ(Y)|XY|),n=1,|XY|aδ(Y).G(X,Y)\geq\left\{\begin{matrix}c_{a}|X-Y|^{1-n},&\qquad n\geq 2,&\qquad|X-Y|\leq a\delta(Y),\\[2.84526pt] c_{a}\log\big{(}\frac{\delta(Y)}{|X-Y|}\big{)},&\qquad n=1,&\qquad|X-Y|\leq a\delta(Y).\end{matrix}\right.
  3. (iii)

    For each YΩY\in\Omega, G(,Y)C(Ω¯\{Y})Wloc1,2(Ω\{Y})G(\cdot,Y)\in C(\overline{\Omega}\backslash\{Y\})\cap W^{1,2}_{\operatorname{loc}}(\Omega\backslash\{Y\}) and G(,Y)|Ω0G(\cdot,Y)|_{\partial\Omega}\equiv 0.

  4. (iv)

    For each XΩX\in\Omega, the identity LG(,X)=δXLG(\cdot,X)=\delta_{X} holds in the distributional sense; that is,

    ΩA(Y)YG(Y,X)Φ(Y)𝑑Y=Φ(X),for any ΦCc(Ω).\int_{\Omega}A(Y)\nabla_{Y}G(Y,X)\nabla\Phi(Y)\,dY=\Phi(X),\qquad\text{for any }\Phi\in C_{c}^{\infty}(\Omega).
  5. (v)

    For each X,YΩX,Y\in\Omega with XYX\neq Y, if L=divATL^{*}=-\operatorname{div}A^{T}\nabla, then

    GL(X,Y)=GL(Y,X).G_{L}(X,Y)=G_{L^{*}}(Y,X).

If n2n\geq 2, then it has been known for a long time that a non-negative Green’s function exists for any domain, without any further regularity assumptions on the geometry [GW82, HK07]. If n=1n=1, the situation has been more challenging; for instance, key Sobolev embeddings, available when n2n\geq 2, fail when n=1n=1, and the fundamental solution changes sign when n=1n=1 [KN85]; nevertheless, the paper [DK09] shows the construction of a Green’s function for any domain with either finite volume or finite width, and also, for the domain above a Lipschitz graph, improving on the result of [DM95] (but non-negativity is not shown in these works). For our setting of uniform domains Ωn+1\Omega\subset\mathbb{R}^{n+1}, n1n\geq 1, with nn-Ahlfors regular boundary, the unified (for n=1n=1 and n2n\geq 2) existence of the non-negative Green’s function for arbitrary divergence form elliptic operators LL of merely bounded measurable coefficients with the properties stated above follows from the much more general, recent construction in [DFM, Theorem 14.60 and Lemma 14.78].

The Green’s function and the elliptic measure are related through the following Riesz formula: For every FCc(n+1)F\in C_{\operatorname{c}}^{\infty}(\mathbb{R}^{n+1}), one has that

(2.35) ΩF(y)𝑑ωX(y)=F(X)ΩATYGL(X,Y)YF(Y)𝑑Y.\int_{\partial\Omega}F(y)\,d\omega^{X}(y)=F(X)-\iint_{\Omega}A^{T}\nabla_{Y}G_{L}(X,Y)\cdot\nabla_{Y}F(Y)\,dY.

We need several estimates from the literature for the elliptic measure and Green’s function in our proofs and we list these estimates below. Although these have appeared in several works in the literature [CFMS81, HKM93, AGMT22], we cite [DFM] for their unified consideration of the cases n=1n=1 and n2n\geq 2 and arbitrary elliptic operators on uniform domains with Ahlfors regular boundary. The first estimate is a non-degeneracy estimate for elliptic measure.

Lemma 2.36 (Bourgain’s estimate, [DFM, Lemma 15.1]).

Let xΩx\in\partial\Omega and r(0,diam(Ω)]r\in(0,\operatorname{diam}(\partial\Omega)] and let Yx,rY_{x,r} be a corkscrew point relative to xx at scale rr. We have

ωYx,r(Δ(x,r))c,\displaystyle\omega^{Y_{x,r}}(\Delta(x,r))\geq c,

where cc depends only on dimension, ellipticity, and the Ahlfors regularity constant. Here and below Δ(x,r)=B(x,r)Ω\Delta(x,r)=B(x,r)\cap\partial\Omega.

The elliptic measure is locally doubling in the following sense.

Lemma 2.37 ((Local) doubling property, [DFM, Lemma 15.43]).

Let xΩx\in\partial\Omega and r(0,diam(Ω)]r\in(0,\operatorname{diam}(\Omega)]. If YΩB(x,4r)Y\in\Omega\setminus B(x,4r) then ωY(Δ(x,2r))CωY(Δ(x,r))\omega^{Y}(\Delta(x,2r))\leq C\omega^{Y}(\Delta(x,r)), where CC depends on dimension, ellipticity, Harnack chain, corkscrew and Ahlfors regularity constants.

The following estimate allows us to connect the Green’s function and elliptic measure in a quantitative way:

Lemma 2.38 (CFMS estimate, [DFM, Lemma 15.28]).

If xΩx\in\partial\Omega and r(0,diam(Ω)]r\in(0,\operatorname{diam}(\partial\Omega)] then

GL(X,Yx,r)rωLX(Δ(x,r))rn\displaystyle\frac{G_{L}(X,Y_{x,r})}{r}\approx\frac{\omega^{X}_{L}(\Delta(x,r))}{r^{n}}

for every XΩB(x,2r)X\in\Omega\setminus B(x,2r), where the implicit constants depend on dimension, ellipticity, Harnack chain, corkscrew and Ahlfors regularity constants, and Yx,rY_{x,r} is any corkscrew point relative to xx at scale rr.

Non-negative solutions uu to Lu=0Lu=0 that vanish on an open subset of the boundary of a uniform domain must do so at the same rate:

Lemma 2.39 (Boundary Harnack Principle, [DFM, Theorem 15.64]).

Let xΩx\in\partial\Omega and r(0,diam(Ω)]r\in(0,\operatorname{diam}(\partial\Omega)], and let uu and vv be positive functions such that Lu=Lv=0Lu=Lv=0 in ΩB(x,4r)\Omega\cap B(x,4r) that vanish continuously on ΩB(x,4r)\partial\Omega\cap B(x,4r). Then

u(X)v(X)u(Yx,r)v(Yx,r),for all XB(x,r)Ω,\displaystyle\frac{u(X)}{v(X)}\approx\frac{u(Y_{x,r})}{v(Y_{x,r})},\qquad\text{for all }X\in B(x,r)\cap\Omega,

where Yx,rY_{x,r} is a corkscrew point relative to xx at scale rr. The implicit constants depend on dimension, ellipticity, Harnack chain, corkscrew and Ahlfors regularity constants.

We have the following standard consequence of the boundary Harnack Principle.

Lemma 2.40 (Change of Poles, [DFM, Lemma 15.61]).

Let xΩx\in\partial\Omega, r(0,diam(Ω))r\in(0,\operatorname{diam}(\partial\Omega)), Yx,rY_{x,r} a corkscrew point relative to xx at scale rr, and EΔ(x,r)E\subset\Delta(x,r) a Borel set. Then

ωYx,r(E)ωX(E)ωX(Δ(x,r)),for each XΩ\B(x,2r),\omega^{Y_{x,r}}(E)\approx\frac{\omega^{X}(E)}{\omega^{X}(\Delta(x,r))},\qquad\text{for each }X\in\Omega\backslash B(x,2r),

where the implicit constants depend on dimension, ellipticity, and structural constants of Ω\Omega.

Finally, solutions uu to Lu=0Lu=0 that vanish continuously on the boundary do so at a Hölder rate:

Lemma 2.41 ([DFM, Lemma 11.32]).

Let xΩx\in\partial\Omega and r(0,diam(Ω)]r\in(0,\operatorname{diam}(\partial\Omega)], and let uu be a solution to Lu=0Lu=0 in ΩB(x,4r)\Omega\cap B(x,4r) that vanishes continuously on ΩB(x,4r)\partial\Omega\cap B(x,4r). We have the bound

|u(X)|C(dist(X,Ω)r)αsupYB(x,4r)|u(Y)|\displaystyle|u(X)|\leq C\left(\frac{\operatorname{dist}(X,\partial\Omega)}{r}\right)^{\alpha}\sup_{Y\in B(x,4r)}|u(Y)|

for every XB(x,r)ΩX\in B(x,r)\cap\Omega, where α\alpha and CC depend on dimension, the Harnack chain, corkscrew and Ahlfors regularity constants.

3. Set-up for Theorem 1.1

In this section we provide some of the preliminary estimates and observations required to prove Theorem 1.1. We divide this section into two subsections. The first records Carleson measure estimates and non-tangential convergence in our setting. The second subsection contains a few lemmas, which are roughly adapted from ideas in [DJK84] and play a crucial role in our analysis. Throughout this section, we suppose that the assumptions of Theorem 1.1 hold; that is, Ω\Omega is a uniform domain with Ahlfors regular boundary and LL is a not necessarily symmetric divergence form elliptic operator such that ωLA(σ)\omega_{L}\in A_{\infty}(\sigma). For the sequel, recall that δ()dist(,Ω)\delta(\cdot)\coloneqq\operatorname{dist}(\cdot,\partial\Omega).

3.1. CME and non-tangential convergence

One of the key tools in most of the constructions of ε\varepsilon-approximators (see, for example, [HMM16]) is L2L^{2}-type Carleson measure estimates (CME), that is, Carleson properties (see Definition 2.7) of measures μu\mu_{u} such that dμu=|u|2δ(Y)dYd\mu_{u}=|\nabla u|^{2}\delta(Y)\,dY for a solution uu to Lu=0Lu=0. Under the hypotheses of Theorem 1.1, we have the following “classical” Carleson measure estimate and LpL^{p}-solvability of the Dirichlet problem for LL for some pp [AHMT19, CHMT20, FP22]:

Lemma 3.1 ([FP22, Corollary 1.32]).

Suppose that Ω\Omega is a uniform domain in n+1\mathbb{R}^{n+1}, n1n\geq 1, with nn-Ahlfors regular boundary and LL is a divergence form elliptic operator such that ωLA(σ)\omega_{L}\in A_{\infty}(\sigma). There exists a constant C1C\geq 1 such that if uWloc1,2(Ω)L(Ω)u\in W^{1,2}_{\operatorname{loc}}(\Omega)\cap L^{\infty}(\Omega) is a solution to Lu=0Lu=0 then

supr>0supxΩ1rnB(x,r)Ω|u(Y)|2δ(Y)𝑑YCuL(Ω)2.\displaystyle\sup_{r>0}\sup_{x\in\partial\Omega}\frac{1}{r^{n}}\iint_{B(x,r)\cap\Omega}|\nabla u(Y)|^{2}\delta(Y)\,dY\lesssim C\|u\|_{L^{\infty}(\Omega)}^{2}.

The constant CC depends on the structural constants and the ωLA(σ)\omega_{L}\in A_{\infty}(\sigma) constants.

Lemma 3.2 ([HL18, Theorem 1.3]).

Suppose that Ωn+1\Omega\subset\mathbb{R}^{n+1}, n1n\geq 1, is a uniform domain with Ahlfors regular boundary and LL is a divergence form elliptic operator in Ω\Omega such that ωLA(σ)\omega_{L}\in A_{\infty}(\sigma). There exist 1<p<1<p<\infty and C1C\geq 1 such that for every fLp(Ω)f\in L^{p}(\partial\Omega) there exists a solution uu to the boundary value problem

{Lu=0 in Ω,NuLp(Ω)fLp(Ω),limYx,n.t.u(Y)=f(x), for σ-a.e.xΩ.\displaystyle\begin{cases}Lu=0&\text{ in }\Omega,\\ \|Nu\|_{L^{p}(\partial\Omega)}\lesssim\|f\|_{L^{p}(\partial\Omega)},\\ \lim\limits_{Y\to x,\,\operatorname{n.t.}}u(Y)=f(x),&\text{ for $\sigma$-a.e.}x\in\partial\Omega.\end{cases}

Moreover, the solution is of the form

u(X)=uf(X)=Ωf(y)𝑑ωX(y).\displaystyle u(X)=u_{f}(X)=\int_{\partial\Omega}f(y)\,d\omega^{X}(y).

Using the LpL^{p} result of Lemma 3.2 gives us the following non-tangential convergence result for LL^{\infty} data:

Lemma 3.3.

Suppose that Ωn+1\Omega\subset\mathbb{R}^{n+1}, n1n\geq 1, is a uniform domain with Ahlfors regular boundary and LL is a divergence form elliptic operator in Ω\Omega such that ωLA(σ)\omega_{L}\in A_{\infty}(\sigma). If fL(dσ)f\in L^{\infty}(d\sigma), then the solution

uf(X)=Ωf(y)𝑑ωX(y)\displaystyle u_{f}(X)=\int_{\partial\Omega}f(y)\,d\omega^{X}(y)

converges non-tangentially to ff; that is,

(3.4) limYx,n.t.uf(Y)=f(x), for σ-a.e. xΩ.\lim\limits_{Y\to x,\,\operatorname{n.t.}}u_{f}(Y)=f(x),\quad\text{ for $\sigma$-a.e. }x\in\partial\Omega.

Equivalently,

(3.5) limYx,n.t.uf(Y)=f(x), for ωY-a.e. xΩ.\lim\limits_{Y\to x,\,\operatorname{n.t.}}u_{f}(Y)=f(x),\quad\text{ for $\omega^{Y}$-a.e. }x\in\partial\Omega.
Proof.

Let pp be as in Lemma 3.2. Let us note that ωY\omega^{Y} and ωY\omega^{Y^{\prime}} are mutually absolutely continuous for all Y,YΩY,Y^{\prime}\in\Omega by the Harnack chain condition and the Harnack inequality (see Lemma 2.29). Moreover, ωY\omega^{Y} is mutually absolutely continuous with respect to σ\sigma by the A(σ)A_{\infty}(\sigma) condition [CF74, Lemma 5] and hence, (3.4) and (3.5) are equivalent. In addition, we have fL(Ω,dσ)f\in L^{\infty}(\partial\Omega,d\sigma) if and only if fL(Ω,dωY)f\in L^{\infty}(\partial\Omega,d\omega^{Y}) for all YΩY\in\Omega, with fL(Ω,dσ)=fL(Ω,dωY)\|f\|_{L^{\infty}(\partial\Omega,d\sigma)}=\|f\|_{L^{\infty}(\partial\Omega,d\omega^{Y})}.

If diam(Ω)<\operatorname{diam}(\partial\Omega)<\infty, then fLp(Ω)f\in L^{p}(\partial\Omega) and the claim follows from Lemma 3.2. Thus, we may assume that diam(Ω)=\operatorname{diam}(\partial\Omega)=\infty. We show that the claim holds in any ball BB centered at Ω\partial\Omega. Let us fix x0Ωx_{0}\in\partial\Omega and r>0r>0. We write f=g+hf=g+h for g=f𝟙Δ(x0,100r)g=f\mathbbm{1}_{\Delta(x_{0},100r)}, where Δ(x0,100r)B(x0,100r)Ω\Delta(x_{0},100r)\coloneqq B(x_{0},100r)\cap\partial\Omega. By linearity, we have that uf(X)=ug(X)+uh(X)u_{f}(X)=u_{g}(X)+u_{h}(X). Since fL(Ω)f\in L^{\infty}(\partial\Omega), we know that gLp(Ω)g\in L^{p}(\partial\Omega) and thus, Lemma 3.2 gives us

limYx,n.t.ug(Y)=g(x)=f(x), for σ-a.e. xΩB(x0,r).\displaystyle\lim\limits_{Y\to x,\,\operatorname{n.t.}}u_{g}(Y)=g(x)=f(x),\quad\text{ for $\sigma$-a.e. }x\in\partial\Omega\cap B(x_{0},r).

Therefore it suffices to show

(3.6) limYx,n.t.uh(Y)=0, for σ-a.e. xΩB(x0,r).\lim\limits_{Y\to x,\,\operatorname{n.t.}}u_{h}(Y)=0,\quad\text{ for $\sigma$-a.e. }x\in\partial\Omega\cap B(x_{0},r).

We now have

(3.7) |uh(X)|fL(Ω)ωX(ΩΔ(x0,100r))|u_{h}(X)|\leq\|f\|_{L^{\infty}(\partial\Omega)}\omega^{X}(\partial\Omega\setminus\Delta(x_{0},100r))

for any XΩX\in\Omega. Let ϕCc(Δ(x0,100r))\phi\in C_{\operatorname{c}}(\Delta(x_{0},100r)) be a non-negative function such that ϕ(x)1\phi(x)\leq 1 for every xΩx\in\partial\Omega and ϕ1\phi\equiv 1 on Δ(x0,50r)\Delta(x_{0},50r). Then, since ωX\omega^{X} is a probability measure, we get

(3.8) ωX(ΩΔ(x0,100r))1v(X)1Ωϕ(y)𝑑ωX(y).\displaystyle\omega^{X}(\partial\Omega\setminus\Delta(x_{0},100r))\leq 1-v(X)\coloneqq 1-\int_{\partial\Omega}\phi(y)\,d\omega^{X}(y).

Since ϕ\phi is a compactly supported continuous function on Ω\partial\Omega, we know that vv is a continuous bounded solution to Lv=0Lv=0 in Ω¯\overline{\Omega} such that v=ϕ1v=\phi\equiv 1 on Δ(x0,50r)\Delta(x_{0},50r). In particular, we have

(3.9) limYx,n.t.v(Y)=1\displaystyle\lim_{Y\to x,\,\operatorname{n.t.}}v(Y)=1

for every xΔ(x0,50r)x\in\Delta(x_{0},50r). Thus, combining (3.7) and (3.8) gives us |uh(X)|fL(Ω)(1v(X))|u_{h}(X)|\leq\|f\|_{L^{\infty}(\partial\Omega)}\left(1-v(X)\right) for every XΩX\in\Omega and (3.6) follows then from (3.9). This completes the proof. ∎

3.2. A few important lemmas

In this subsection, we prove some lemmas that will be important for the proof of Theorem 1.1. The first three lemmas were inspired by the ideas in [DJK84].

Lemma 3.10.

Let Ω\Omega be a domain in n+1\mathbb{R}^{n+1}, n1n\geq 1, and let LL be a not necessarily symmetric divergence form elliptic operator. Suppose that uWloc1,2(Ω)u\in W^{1,2}_{\operatorname{loc}}(\Omega) is a weak solution to Lu=0Lu=0 in Ω\Omega, that ΩΩ\Omega^{\prime}\subset\Omega is a Wiener-regular domain which is compactly contained in Ω\Omega, and fix XΩX_{*}\in\Omega^{\prime}. If n=1n=1, assume in addition that Ω\Omega^{\prime} is a uniform domain with nn-Ahlfors regular boundary. Then |u|2GL,Ω(X,)L1(Ω)|\nabla u|^{2}G_{L,\Omega^{\prime}}(X_{*},\cdot)\in L^{1}(\Omega^{\prime}), and

(3.11) Ω|u(Y)|2GL,Ω(X,Y)𝑑YΩ(u(y)u(X))2𝑑ωL,ΩX(y),\iint_{\Omega^{\prime}}|\nabla u(Y)|^{2}G_{L,\Omega^{\prime}}(X_{*},Y)\,dY\approx\int_{\partial\Omega^{\prime}}(u(y)-u(X_{*}))^{2}\,d\omega_{L,\Omega^{\prime}}^{X_{*}}(y),

where the implicit constants depend only on ellipticity.

Proof.

Throughout this argument we fix XΩX_{*}\in\Omega^{\prime}, we let r=dist(X,Ω)/8r=\operatorname{dist}(X_{*},\partial\Omega^{\prime})/8, and we write G(Y)GΩ(X,Y)G(Y)\coloneqq G_{\Omega^{\prime}}(X_{*},Y). First, we show the finiteness of the integral in the left-hand side of (3.11). If n=1n=1, then it is trivial by (2.33) and the Meyers reverse Hölder estimate for gradients of solutions [Mey63], so suppose that n2n\geq 2. We write

Ω|u|2G𝑑Y=Ω\B(X,r)|u|2G𝑑Y+B(X,r)|u|2G𝑑YT1+T2.\iint_{\Omega^{\prime}}|\nabla u|^{2}G\,dY=\iint_{\Omega^{\prime}\backslash B(X_{*},r)}|\nabla u|^{2}G\,dY\\ +\iint_{B(X_{*},r)}|\nabla u|^{2}G\,dY\eqqcolon T_{1}+T_{2}.

Since GL(Ω\B(X,r))G\in L^{\infty}(\Omega^{\prime}\backslash B_{*}(X_{*},r)) and uL2(Ω)\nabla u\in L^{2}(\Omega^{\prime}), it is clear that T1<T_{1}<\infty. As for T2T_{2}, for each k0k\in\mathbb{N}_{0}, let AkB(X,2kr)\B(X,2k1r)A_{k}\coloneqq B(X_{*},2^{-k}r)\backslash B(X_{*},2^{-k-1}r), and using (2.33), Lemma 2.27, and Lemma 2.28, we see that

(3.12) T2=k=0Ak|u|2G𝑑Yk=0(2kr)n+1B(X,2kr)|(u(Y)u(X))|2𝑑Yk=0(2kr)n1B(X,2k+1r)|u(Y)u(X)|2𝑑Yk=022αkuL(B(X,4r))<.T_{2}=\sum_{k=0}^{\infty}\iint_{A_{k}}|\nabla u|^{2}G\,dY\lesssim\sum_{k=0}^{\infty}(2^{-k}r)^{-n+1}\iint_{B(X_{*},2^{-k}r)}|\nabla(u(Y)-u(X_{*}))|^{2}\,dY\\ \lesssim\sum_{k=0}^{\infty}(2^{-k}r)^{-n-1}\iint_{B(X_{*},2^{-k+1}r)}|u(Y)-u(X_{*})|^{2}\,dY\lesssim\sum_{k=0}^{\infty}2^{-2\alpha k}\|u\|_{L^{\infty}(B(X_{*},4r))}<\infty.

We turn to the proof of (3.11). By the ellipticity of AA and the product rule, we see that

(3.13) Ω|u|2G𝑑YΩAu(u)G𝑑Y=ΩAu(uG)dYΩAu(G)u𝑑YT3+T4,\iint_{\Omega^{\prime}}|\nabla u|^{2}G\,dY\approx\iint_{\Omega^{\prime}}A\nabla u\cdot(\nabla u)G\,dY\\ =\iint_{\Omega^{\prime}}A\nabla u\cdot\nabla(uG)\,dY-\iint_{\Omega^{\prime}}A\nabla u\cdot(\nabla G)u\,dY\eqqcolon T_{3}+T_{4},

provided that the last two integrals are finite, which we now show. First, we claim that |u||u||G|L1(Ω)|u||\nabla u||\nabla G|\in L^{1}(\Omega^{\prime}). As before, it is enough to show that |u||u||G|L1(B(X,r))|u||\nabla u||\nabla G|\in L^{1}(B(X_{*},r)). Let AkA_{k} as above, and for each k0k\in\mathbb{N}_{0}, let {Bkj}j=1Jk\{B_{k}^{j}\}_{j=1}^{J_{k}} be a cover of AkA_{k} by balls centered on AkA_{k} of radius 2k4r2^{-k-4}r with uniformly bounded overlap. Then Jk1J_{k}\lesssim 1, and we have that

B(X,r)|u||u||G|dYCk=0maxjBkj|u||G|dY{Ck=02kα,n2,Ck=0k2kα,n=1,\iint_{B(X_{*},r)}|u||\nabla u||\nabla G|\,dY\leq C\sum_{k=0}^{\infty}\max_{j}\iint_{B_{k}^{j}}|\nabla u||\nabla G|\,dY\leq\left\{\begin{matrix}C\sum_{k=0}^{\infty}2^{-k\alpha},&\quad n\geq 2,\\[2.84526pt] C\sum_{k=0}^{\infty}k2^{-k\alpha},&\quad n=1,\end{matrix}\right.

where the last estimate follows from using Lemma 2.27 for both uu and GG, then (2.33) and Lemma 2.28. This proves the claim. By the product rule and the triangle inequality, we have also shown that |u||(uG)|L1(Ω)|\nabla u||\nabla(uG)|\in L^{1}(\Omega^{\prime}). Finally, by boundedness of AA we conclude that Au(uG)A\nabla u\nabla(uG) and Au(G)uA\nabla u(\nabla G)u belong to L1(Ω)L^{1}(\Omega^{\prime}), as desired.

The next step is to show that T3=0T_{3}=0. For each MM\in\mathbb{N} large enough, let ψMC(Ω)\psi_{M}\in C^{\infty}(\Omega) satisfy ψM0\psi_{M}\geq 0, ψM1\psi_{M}\equiv 1 in Ω\B(X,2M)\Omega\backslash B(X_{*},\frac{2}{M}), ψM0\psi_{M}\equiv 0 in B(X,1M)B(X_{*},\frac{1}{M}), and |ψM|M|\nabla\psi_{M}|\lesssim M. We claim that

(3.14) ΩAu(uG)ψMdY0,as M.\iint_{\Omega^{\prime}}A\nabla u\nabla(uG)\psi_{M}\,dY\rightarrow 0,\qquad\text{as }M\rightarrow\infty.

Fix MM\in\mathbb{N} large enough. By the product rule and the fact that uGψMW01,2(Ω)uG\psi_{M}\in W_{0}^{1,2}(\Omega^{\prime}), since Lu=0Lu=0 in Ω\Omega^{\prime}, we have that

(3.15) |ΩAu(uG)ψMdY|=|ΩAu(ψM)uG𝑑Y|CMB(X,2M)\B(X,1M)|u|GdY{CMα,n2,CMαlogM,n=1,\Big{|}\iint_{\Omega^{\prime}}A\nabla u\nabla(uG)\psi_{M}\,dY\Big{|}=\Big{|}\iint_{\Omega^{\prime}}A\nabla u(\nabla\psi_{M})uG\,dY\Big{|}\\ \leq CM\iint_{B(X_{*},\frac{2}{M})\backslash B(X_{*},\frac{1}{M})}|\nabla u|G\,dY\leq\left\{\begin{matrix}CM^{-\alpha},&\quad n\geq 2,\\[2.84526pt] CM^{-\alpha}\log M,&\quad n=1,\end{matrix}\right.

where in the last estimate we once again used (2.33), Lemma 2.27, and Lemma 2.28. Thus we have shown (3.14). Since it is also true that Au(uG)ψMAu(uG)A\nabla u\nabla(uG)\psi_{M}\rightarrow A\nabla u\nabla(uG) pointwise a.e. in Ω\Omega^{\prime} and since we have already proved that |u||(uG)|L1(Ω)|\nabla u||\nabla(uG)|\in L^{1}(\Omega^{\prime}), then by the Lebesgue Dominated Convergence Theorem we conclude that T3=0T_{3}=0.

We proceed with the proof of (3.11) as follows:

T4=12ΩATG(u2)dY=12(Ωu(y)2𝑑ωL,ΩX(y)u(X)2)=12Ω(u(y)u(X))2𝑑ωL,ΩX(y),T_{4}=-\frac{1}{2}\iint_{\Omega^{\prime}}A^{T}\nabla G\nabla(u^{2})\,dY=\frac{1}{2}\Big{(}\int_{\partial\Omega^{\prime}}u(y)^{2}\,d\omega_{L,\Omega^{\prime}}^{X_{*}}(y)-u(X_{*})^{2}\Big{)}\\ =\frac{1}{2}\int_{\partial\Omega^{\prime}}(u(y)-u(X_{*}))^{2}\,d\omega_{L,\Omega^{\prime}}^{X_{*}}(y),

where we used the Riesz formula (2.35), and in the last identity we used that ωL,ΩX\omega_{L,\Omega^{\prime}}^{X_{*}} is a probability measure, and that Lu=0Lu=0 in Ω\Omega^{\prime} so that

Ωu(y)𝑑ωL,ΩX(y)=u(X),\displaystyle\int_{\partial\Omega^{\prime}}u(y)\,d\omega_{L,\Omega^{\prime}}^{X_{*}}(y)=u(X_{*}),

and hence

Ωu(X)u(y)𝑑ωL,ΩX(y)=u(X)Ωu(y)𝑑ωL,ΩX(y)=u(X)2.\displaystyle\int_{\partial\Omega^{\prime}}u(X_{*})u(y)\,d\omega_{L,\Omega^{\prime}}^{X_{*}}(y)=u(X_{*})\int_{\partial\Omega^{\prime}}u(y)\,d\omega_{L,\Omega^{\prime}}^{X_{*}}(y)=u(X_{*})^{2}.

This finishes the proof. ∎

The following result is a direct consequence of the maximum principle and the DeGiorgi–Nash–Moser theory (see for instance the proof of [ADFJM19, Proposition 3.2]).

Lemma 3.16.

Let Ω1\Omega_{1} and Ω2\Omega_{2} be Wiener regular domains such that Ω1Ω2\Omega_{1}\subseteq\Omega_{2}. If Gi(X,Y)G_{i}(X,Y) is the Green function for Ωi\Omega_{i}, i=1,2i=1,2 then

G1(X,Y)G2(X,Y) for every (X,Y)Ω1×Ω1{X=Y}.\displaystyle G_{1}(X,Y)\leq G_{2}(X,Y)\quad\text{ for every }(X,Y)\in\Omega_{1}\times\Omega_{1}\setminus\{X=Y\}.

We will also use a corona-type decomposition of the elliptic measure ωL\omega_{L} [HLMN17, GMT18, AGMT22]. To formulate this decomposition, we recall the coherency and semicoherency of subcollections of 𝔻\mathbb{D}:

Definition 3.17.

We say that a subcollection 𝒮𝔻\mathcal{S}\subset\mathbb{D} is coherent if the following three conditions hold.

  1. (a)

    There exists a maximal element Q(𝒮)𝒮Q(\mathcal{S})\in\mathcal{S} such that QQ(𝒮)Q\subset Q(\mathcal{S}) for every Q𝒮Q\in\mathcal{S}.

  2. (b)

    If Q𝒮Q\in\mathcal{S} and P𝔻P\in\mathbb{D} is a cube such that QPQ(𝒮)Q\subset P\subset Q(\mathcal{S}), then also P𝒮P\in\mathcal{S}.

  3. (c)

    If Q𝒮Q\in\mathcal{S}, then either all children of QQ belong to 𝒮\mathcal{S} or none of them do.

If 𝒮\mathcal{S} satisfies only conditions (a) and (b), then we say that 𝒮\mathcal{S} is semicoherent.

We are ready to present the corona decomposition that we shall use in the sequel.

Lemma 3.18 ([AGMT22, Proposition 3.1]).

Suppose that Ω\Omega is a uniform domain and L=divAL=-\mathop{\operatorname{div}}\nolimits A\nabla is a divergence form elliptic operator such that ωLA(σ)\omega_{L}\in A_{\infty}(\sigma). Then there exist constants, C,M>1C,M>1 and a decomposition of the dyadic system 𝔻=𝔻(Ω)\mathbb{D}=\mathbb{D}(\partial\Omega), with the following properties.

  1. (i)

    The dyadic grid breaks into a disjoint decomposition 𝔻=𝒢\mathbb{D}=\mathcal{G}\cup\mathcal{B}, the good cubes and bad cubes respectively.

  2. (ii)

    The family 𝒢\mathcal{G} has a disjoint decomposition 𝒢=𝒮\mathcal{G}=\cup\mathcal{S} where each 𝒮\mathcal{S} is a coherent stopping time regime with maximal cube Q(𝒮)Q(\mathcal{S}).

  3. (iii)

    The maximal cubes and bad cubes satisfy a Carleson packing condition:

    QQRσ(Q)+𝒮:Q(𝒮)Rσ(Q(𝒮))Cσ(R) for every R𝔻.\displaystyle\sum_{\begin{subarray}{c}Q\in\mathcal{B}\\ Q\subseteq R\end{subarray}}\sigma(Q)+\sum_{\mathcal{S}\colon Q(\mathcal{S})\subseteq R}\sigma(Q(\mathcal{S}))\leq C\sigma(R)\quad\text{ for every }R\in\mathbb{D}.
  4. (iv)

    On each stopping time 𝒮\mathcal{S}, the elliptic measure ‘acts like surface measure’ in the sense that if Q𝒮Q\in\mathcal{S}, then

    M1σ(Q)σ(Q(𝒮))ωXQ(𝒮)(Q)ωXQ(𝒮)(Q(𝒮))Mσ(Q)σ(Q(𝒮)).\displaystyle M^{-1}\frac{\sigma(Q)}{\sigma(Q(\mathcal{S}))}\leq\frac{\omega^{X_{Q(\mathcal{S})}}(Q)}{\omega^{X_{Q(\mathcal{S})}}(Q(\mathcal{S}))}\leq M\frac{\sigma(Q)}{\sigma(Q(\mathcal{S}))}.
Proof.

Since ωLA(σ)\omega_{L}\in A_{\infty}(\sigma), then by Lemma 3.1 we have that the hypothesis (b) of [AGMT22, Proposition 3.1] is satisfied. By [AGMT22, Proposition 3.1], this yields a decomposition 𝔻=𝒢\mathbb{D}=\mathcal{G}\cup\mathcal{B} like the one described above (in fact we have that =\mathcal{B}=\varnothing), but with the caveat that the stopping time regimes need only be semicoherent. Then, by a standard mechanism (see for instance [DS93, pp.56–57], and [CHM, Remark 2.13]), we can modify the stopping time regimes so that they are coherent, while the rest of the properties are still satisfied. ∎

It is straightforward to check that for each semicoherent stopping time regime 𝒮\mathcal{S}, there exists a collection of pairwise disjoint cubes 𝒮\mathcal{F}_{\mathcal{S}} such that 𝒮=𝔻𝒮,Q(𝒮)\mathcal{S}=\mathbb{D}_{\mathcal{F}_{\mathcal{S}},Q(\mathcal{S})}. By Lemma 2.38 and the Harnack inequality, we get the following:

Lemma 3.19.

Suppose that Ω\Omega is a uniform domain in n+1\mathbb{R}^{n+1}, n1n\geq 1, with nn-Ahlfors regular boundary, that L=divAL=-\mathop{\operatorname{div}}\nolimits A\nabla is a divergence form elliptic operator such that ωLA(σ)\omega_{L}\in A_{\infty}(\sigma), and that 𝒮\mathcal{S} is a semicoherent stopping time regime satisfying the property in Lemma 3.18 (iv). Then

GL(XQ,Y)σ(Q)δ(Y) for every Q𝒮 and YΩ𝒮,Q,\displaystyle G_{L}(X_{Q},Y)\sigma(Q)\approx\delta(Y)\qquad\text{ for every }Q\in\mathcal{S}\text{ and }Y\in\Omega^{*}_{\mathcal{F}_{\mathcal{S}},Q},

where the implicit constants depend only on dimension, ellipticity, κ\kappa, Harnack chain, corkscrew, and Ahlfors regularity constants, as well as the constant MM in Lemma 3.18 (iv).

Proof.

Fix Q𝒮Q\in\mathcal{S} and333For relevant notation in this proof, see Section 2.3. YΩ𝒮,QY\in\Omega_{\mathcal{F}_{\mathcal{S}},Q}^{*}. By definition, there exists P𝔻S,QP\in\mathbb{D}_{\mathcal{F}_{S},Q} so that YUPY\in U_{P}^{*}, and thus P𝒮P\in\mathcal{S}, PQP\subseteq Q, and δ(Y)(P)\delta(Y)\approx\ell(P). Let BPB(xP,ηa0(P))B_{P}^{\prime}\coloneqq B(x_{P},\eta a_{0}\ell(P)) with η(0,1)\eta\in(0,1) small, and let XPX_{P}^{\prime} be the corkscrew point for BPB_{P}^{\prime}. Define BQB_{Q}^{\prime} and XQX_{Q}^{\prime} analogously. We may guarantee that XQΩ\2BPX_{Q}\in\Omega\backslash 2B_{P}^{\prime} if η\eta is chosen small enough depending only on the corkscrew constant. Then, since δ(Y)δ(XP)\delta(Y)\approx\delta(X_{P}^{\prime}), and |YXP|δ(Y)|Y-X_{P}^{\prime}|\lesssim\delta(Y), by the Harnack inequality (for LL and LL^{*}) and Harnack chains, Lemma 2.38, the doubling property of elliptic measure, and Ahlfors regularity, we have that

GL(XQ,Y)δ(Y)GL(XQ,XP)(P)ωXQ(2BP)(P)nωXQ(BP)σ(P).\frac{G_{L}(X_{Q},Y)}{\delta(Y)}\approx\frac{G_{L}(X_{Q},X_{P}^{\prime})}{\ell(P)}\approx\frac{\omega^{X_{Q}}(2B_{P}^{\prime})}{\ell(P)^{n}}\approx\frac{\omega^{X_{Q}^{\prime}}(B_{P}^{\prime})}{\sigma(P)}.

Now, since XQ(𝒮)Ω\2BQX_{Q(\mathcal{S})}\in\Omega\backslash 2B_{Q}^{\prime} and P,Q𝒮P,Q\in\mathcal{S}, then by Lemma 2.40, the doubling property of elliptic measure, Harnack chains, Harnack inequality, Bourgain’s estimate, and the property (iv) in Lemma 3.18, it follows that

ωXQ(BP)σ(P)ωXQ(𝒮)(BP)σ(P)1ωXQ(𝒮)(BQ)ωXQ(𝒮)(P)σ(P)1ωXQ(𝒮)(Q)1σ(Q(𝒮))σ(Q(𝒮))σ(Q)=1σ(Q),\frac{\omega^{X_{Q}^{\prime}}(B_{P})}{\sigma(P)}\approx\frac{\omega^{X_{Q(\mathcal{S})}}(B_{P})}{\sigma(P)}\frac{1}{\omega^{X_{Q(\mathcal{S})}}(B_{Q}^{\prime})}\approx\frac{\omega^{X_{Q(\mathcal{S})}}(P)}{\sigma(P)}\frac{1}{\omega^{X_{Q(\mathcal{S})}}(Q)}\\ \approx\frac{1}{\sigma(Q(\mathcal{S}))}\frac{\sigma(Q(\mathcal{S}))}{\sigma(Q)}=\frac{1}{\sigma(Q)},

which completes the proof. ∎

4. Existence of ε\varepsilon-approximators: proof of Theorem 1.1

In this section we prove Theorem 1.1. The key step in the proof is the construction of BVloc\operatorname{BV}_{\operatorname{loc}} approximators since the existence of smooth approximators with more delicate properties follows then by using almost black box regularization arguments. More precisely, the heart of Theorem 1.1 is the following result:

Theorem 4.1.

Let Ωn+1\Omega\subset\mathbb{R}^{n+1}, n1n\geq 1, be a uniform domain with Ahflors regular boundary. Let L=divAL=-\mathop{\operatorname{div}}\nolimits A\nabla be a (not necessarily symmetric) divergence form elliptic operator satisfying that ωLA(σ)\omega_{L}\in A_{\infty}(\sigma). Then, for any ε(0,1)\varepsilon\in(0,1) there exists a constant CεC_{\varepsilon} such that if uWloc1,2(Ω)L(Ω)u\in W^{1,2}_{\operatorname{loc}}(\Omega)\cap L^{\infty}(\Omega) is a solution to Lu=0Lu=0 in Ω\Omega, then there exists Φ=ΦεBVloc(Ω)\Phi=\Phi^{\varepsilon}\in\operatorname{BV}_{\operatorname{loc}}(\Omega) satisfying

uΦL(Ω)εuL(Ω) and supxΩ,r>01rnB(x,r)Ω|Φ(Y)|𝑑YCεuL(Ω),\displaystyle\|u-\Phi\|_{L^{\infty}(\Omega)}\leq\varepsilon\|u\|_{L^{\infty}(\Omega)}\quad\text{ and }\quad\sup_{x\in\partial\Omega,r>0}\frac{1}{r^{n}}\iint_{B(x,r)\cap\Omega}|\nabla\Phi(Y)|\,dY\leq C_{\varepsilon}\|u\|_{L^{\infty}(\Omega)},

where CεC_{\varepsilon} depends on ε\varepsilon, structural constants, ellipticity and the ωLA(σ)\omega_{L}\in A_{\infty}(\sigma) constants.

As mentioned in the introduction (see the paragraph after Theorem 1.1), the proof of Theorem 4.1 is based on the construction of ε\varepsilon-approximators in [HMM16, Theorem 1.3], with important differences. We will point out when we have to diverge from the strategy in [HMM16].

4.1. Set-up for the proof of Theorem 4.1

We start by making some preliminary considerations and fixing some notation and terminology. Let uu be a bounded solution to Lu=0Lu=0, and let ε(0,1)\varepsilon\in(0,1). Without loss of generality, we may assume that uL(Ω)=1\|u\|_{L^{\infty}(\Omega)}=1, since otherwise we have u=0u=0 or we can replace uu with u/uLu/\|u\|_{L^{\infty}}. We fix a stopping time regime 𝒮\mathcal{S} from Lemma 3.18, with maximal cube Q(𝒮)Q(\mathcal{S}). Following [Gar07, HMM16], we label each cube Q𝔻Q\in\mathbb{D} depending on how much the the function uu oscillates in the corresponding fattened Whitney region UQU_{Q}^{*}. To be more precise, we say that

Q𝔻 is red if supX,YUQ|u(X)u(Y)|ε1000,\displaystyle Q\in\mathbb{D}\text{ is {\bf red} if }\quad\sup_{X,Y\in U_{Q}^{*}}|u(X)-u(Y)|\geq\frac{\varepsilon}{1000},

and

Q𝔻 is blue if supX,YUQ|u(X)u(Y)|<ε1000.\displaystyle Q\in\mathbb{D}\text{ is {\bf blue} if }\quad\sup_{X,Y\in U_{Q}^{*}}|u(X)-u(Y)|<\frac{\varepsilon}{1000}.

We note that these conditions differ from the, ones in [HMM16] in two ways: in [HMM16], the oscillation threshold was ε/10\varepsilon/10 and the level of oscillation was measured in the unfattened region UQU_{Q}. With our conditions we have a little bit more control on the blue cubes, which will be useful for us later. In [HMM16], it was enough to use these two labels but for our analysis it is important to take into consideration the minimal cubes of 𝒮\mathcal{S}, which we label as follows:

Q𝒮 is yellow if Q has a child Q such that Q𝒮.\displaystyle Q\in\mathcal{S}\text{ is {\bf yellow} if }\quad Q\text{ has a child $Q^{\prime}$ such that $Q^{\prime}\not\in\mathcal{S}$}.

Recall that QQ^{\prime} is a child of QQ if QQQ^{\prime}\subset Q and (Q)=(Q)/2\ell(Q^{\prime})=\ell(Q)/2. Notice that yellow cubes are not a separate collection from the red and blue cubes but each yellow cube is also red or blue. We denote the collections of red and yellow cubes by

{Q𝔻:Q is red},𝒴=𝒴(𝒮){Q𝒮:Q is yellow}.\mathcal{R}\coloneqq\{Q\in\mathbb{D}\colon Q\text{ is red}\},\qquad\mathcal{Y}=\mathcal{Y}(\mathcal{S})\coloneqq\{Q\in\mathcal{S}\colon Q\text{ is yellow}\}.

The rough idea of the construction of the ε\varepsilon-approximator Φ\Phi of uu is to first construct Φ\Phi inside a Carleson box TQ0T_{Q_{0}} for a fixed cube Q0Q_{0} and then use a “local to global”-type argument to define a global approximator. Working in a Carleson box allows us to reduce many of the challenging estimates to working with just one stopping time regime 𝒮\mathcal{S} given by Lemma 3.18. We then set Φu\Phi\equiv u in the regions UQU_{Q} such that Q𝒴Q\in\mathcal{R}\cup\mathcal{Y} and break up the blue cubes into smaller stopping time regimes where uu does not vary by more than ε/100\varepsilon/100. In these new regimes, we set Φ=u(X0)\Phi=u(X_{0}) in the union of UQU_{Q}, where X0X_{0} is any point in the union. The LL^{\infty} approximation property follows then just from the way we defined Φ\Phi but verifying the L1L^{1}-type Carleson measure estimate for Φ\Phi is more challenging. For this, one of the key steps is to show that the collections of red, yellow and maximal cubes from the new stopping time regimes satisfy Carleson packing conditions. For the first two collections, this follows in a straightforward way:

Lemma 4.2.

The collections \mathcal{R} and 𝒮𝒴(𝒮)\bigcup_{\mathcal{S}}\mathcal{Y}(\mathcal{S}) satisfy the following Carleson packing conditions: for any P𝔻P\in\mathbb{D} we have

Q,QPσ(Q)1ε2σ(P)\displaystyle\sum_{Q\in\mathcal{R},Q\subset P}\sigma(Q)\lesssim\frac{1}{\varepsilon^{2}}\sigma(P)

and

𝒮Q𝒴(𝒮),QPσ(Q)(C+1)σ(P).\displaystyle\sum_{\mathcal{S}}\sum_{Q\in\mathcal{Y}(\mathcal{S}),Q\subset P}\sigma(Q)\leq(C+1)\sigma(P).

The constant CC is the same constant as CC in Lemma 3.18.

Proof.

The packing condition for 𝒮𝒴(𝒮)\bigcup_{\mathcal{S}}\mathcal{Y}(\mathcal{S}) follows from the definition and the facts that the stopping time regimes are coherent and their maximal cubes satisfy a Carleson packing condition. Indeed, each yellow cube Q𝒴(𝒮)Q\in\mathcal{Y}(\mathcal{S}) is contained in Q(𝒮)Q(\mathcal{S}) and by the coherency of 𝒮\mathcal{S}, no yellow cube can contain a smaller yellow cube Q~𝒴(𝒮)\widetilde{Q}\in\mathcal{Y}(\mathcal{S}). Thus, the cubes in 𝒴(𝒮)\mathcal{Y}(\mathcal{S}) are disjoint and we get

𝒮Q𝒴(𝒮),QPσ(Q)𝒮:Q𝒴(𝒮),QP,Q(𝒮)Pσ(Q)+𝒮:Q𝒴(𝒮),QP,P𝒮σ(Q)𝒮:Q(𝒮)Pσ(Q(𝒮))+𝒮:P𝒮σ(P).\sum_{\mathcal{S}}\sum_{Q\in\mathcal{Y}(\mathcal{S}),Q\subset P}\sigma(Q)\leq\sum_{\begin{subarray}{c}\mathcal{S}:Q\in\mathcal{Y}(\mathcal{S}),Q\subset P,\\ Q(\mathcal{S})\subset P\end{subarray}}\sigma(Q)+\sum_{\begin{subarray}{c}\mathcal{S}:Q\in\mathcal{Y}(\mathcal{S}),Q\subset P,\\ P\in\mathcal{S}\end{subarray}}\sigma(Q)\\ \leq\sum_{\begin{subarray}{c}\mathcal{S}:Q(\mathcal{S})\subset P\end{subarray}}\sigma(Q(\mathcal{S}))+\sum_{\begin{subarray}{c}\mathcal{S}:P\in\mathcal{S}\end{subarray}}\sigma(P).

Since the collection {Q(𝒮)}𝒮\{Q(\mathcal{S})\}_{\mathcal{S}} satisfies a Carleson packing condition by Lemma 3.18, we know that the first sum on the right-hand side is bounded by Cσ(P)C\sigma(P). In addition, since the stopping time regimes 𝐒{\bf S} are disjoint, we have P𝒮P\in\mathcal{S} for at most one stopping time regime 𝒮\mathcal{S}. Thus, the second sum on the right-hand side is bounded by σ(P)\sigma(P). The desired bound follows from combining these two estimates.

Let us then prove the Carleson packing condition for \mathcal{R}. If QQ\in\mathcal{R}, then there exist points X1,X2UQX_{1},X_{2}\in U_{Q}^{*} such that |u(X1)u(X2)|>ε|u(X_{1})-u(X_{2})|>\varepsilon. Since X1,X2UQX_{1},X_{2}\in U_{Q}^{*}, there exist Whitney cubes I1,I2UQI_{1},I_{2}\subset U_{Q}^{*} such that X1(1+τ)I1X_{1}\in(1+\tau)I_{1}, X2(1+τ)I2X_{2}\in(1+\tau)I_{2} and (Q)(I1)(I2)δ(X1)δ(X2)\ell(Q)\approx\ell(I_{1})\approx\ell(I_{2})\approx\delta(X_{1})\approx\delta(X_{2}). In particular, we have dist(X1,UQ)dist(X2,UQ)(Q)\operatorname{dist}(X_{1},\partial U_{Q}^{**})\approx\operatorname{dist}(X_{2},\partial U_{Q}^{**})\approx\ell(Q) for the twice-fattened region UQU_{Q}^{**} (see Section 2.3), where the implicit constants depend on the dilation parameter τ\tau. Thus, since UQU_{Q}^{**} satisfies the Harnack chain condition by Lemma 2.18, there exists a uniformly bounded number of balls B1,B2,,BNB_{1},B_{2},\ldots,B_{N} and points Y1,Y2,,YN1Y_{1},Y_{2},\ldots,Y_{N-1} such that

  1. \bullet

    X1B1X_{1}\in B_{1}, X2BNX_{2}\in B_{N} and YiBiBi+1Y_{i}\in B_{i}\cap B_{i+1} for every i=1,2,,N1i=1,2,\ldots,N-1,

  2. \bullet

    for a constant λ>1\lambda>1 (depending on τ\tau), we have 2λBiUQ2\lambda B_{i}\subset U_{Q}^{**} for every i=1,2,,Ni=1,2,\ldots,N, and

  3. \bullet

    |2λBi||UQ|(Q)n+1|2\lambda B_{i}|\approx|U_{Q}^{**}|\approx\ell(Q)^{n+1} for every i=1,2,,Ni=1,2,\ldots,N.

These properties combined with the triangle inequality, the local Hölder continuity (that is, Lemma 2.28) and the Poincaré inequality applied for solution vv, v(X)u(X)UQu(Z)𝑑Zv(X)\coloneqq u(X)-\fint_{U_{Q}^{**}}u(Z)\,dZ, then give us

ε|v(X1)v(X2)||v(X1)v(Y1)|+i=1N2|v(Yi)v(Yi+1)|+|v(YN)v(X2)|2λB1|v|𝑑X+i=1N22λBi+1|v|𝑑X+2λBN|v|𝑑XUQ|v|𝑑X(Q)(UQ|u|2𝑑X)1/2.\varepsilon\leq|v(X_{1})-v(X_{2})|\leq|v(X_{1})-v(Y_{1})|+\sum_{i=1}^{N-2}|v(Y_{i})-v(Y_{i+1})|+|v(Y_{N})-v(X_{2})|\\ \lesssim\mathchoice{{\vbox{\hbox{$\textstyle\raisebox{-8.61108pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-7.09656pt}}{{\vbox{\hbox{$\scriptstyle\raisebox{-6.02777pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-6.09258pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-4.29826pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-3.54826pt}}\!\iint_{2\lambda B_{1}}|v|\,dX+\sum_{i=1}^{N-2}\mathchoice{{\vbox{\hbox{$\textstyle\raisebox{-8.61108pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-7.09656pt}}{{\vbox{\hbox{$\scriptstyle\raisebox{-6.02777pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-6.09258pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-4.29826pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-3.54826pt}}\!\iint_{2\lambda B_{i+1}}|v|\,dX+\mathchoice{{\vbox{\hbox{$\textstyle\raisebox{-8.61108pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-7.09656pt}}{{\vbox{\hbox{$\scriptstyle\raisebox{-6.02777pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-6.09258pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-4.29826pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-3.54826pt}}\!\iint_{2\lambda B_{N}}|v|\,dX\lesssim\mathchoice{{\vbox{\hbox{$\textstyle\raisebox{-8.61108pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-7.09656pt}}{{\vbox{\hbox{$\scriptstyle\raisebox{-6.02777pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-6.09258pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-4.29826pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-3.54826pt}}\!\iint_{U_{Q}^{**}}|v|\,dX\lesssim\ell(Q)\Big{(}\mathchoice{{\vbox{\hbox{$\textstyle\raisebox{-8.61108pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-7.09656pt}}{{\vbox{\hbox{$\scriptstyle\raisebox{-6.02777pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-6.09258pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-4.29826pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-3.54826pt}}\!\iint_{U_{Q}^{**}}|\nabla u|^{2}\,dX\Big{)}^{1/2}.

Thus, since δ(X)(Q)\delta(X)\approx\ell(Q) for every XUQX\in U_{Q}^{**}, we have

ε2σ(Q)(Q)nε2(Q)n(Q)2UQ|u(X)|2𝑑XUQ|u(X)|2(Q)𝑑XUQ|u(X)|2δ(X)𝑑X.\varepsilon^{2}\sigma(Q)\approx\ell(Q)^{n}\varepsilon^{2}\lesssim\ell(Q)^{n}\ell(Q)^{2}\mathchoice{{\vbox{\hbox{$\textstyle\raisebox{-8.61108pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-7.09656pt}}{{\vbox{\hbox{$\scriptstyle\raisebox{-6.02777pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-6.09258pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-4.29826pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-3.54826pt}}\!\iint_{U_{Q}^{**}}|\nabla u(X)|^{2}\,dX\\ \lesssim\iint_{U_{Q}^{**}}|\nabla u(X)|^{2}\ell(Q)\,dX\approx\iint_{U_{Q}^{**}}|\nabla u(X)|^{2}\delta(X)\,dX.

By construction, we know that the regions UPU_{P}^{**} have bounded overlaps and any twice-fattened Carleson box TPT_{P}^{**} satisfies TPB(xP,RP)ΩT_{P}^{**}\subset B(x_{P},R_{P})\cap\Omega the center xPx_{P} of PP and for a radius RP(P)R_{P}\approx\ell(P). Hence, for any Q0𝔻Q_{0}\in\mathbb{D}, these facts, the previous estimate and the Carleson measure estimate (3.1) give us

PPQ0σ(P)ε2PPQ0UP|u(X)|2δ(X)𝑑Xε2TQ0|u(X)|2δ(X)𝑑Xε2(Q0)nε2σ(Q0),\sum_{\begin{subarray}{c}P\in\mathcal{R}\\ P\subset Q_{0}\end{subarray}}\sigma(P)\lesssim\varepsilon^{-2}\sum_{\begin{subarray}{c}P\in\mathcal{R}\\ P\subset Q_{0}\end{subarray}}\iint_{U_{P}^{**}}|\nabla u(X)|^{2}\delta(X)\,dX\\ \lesssim\varepsilon^{-2}\iint_{T^{**}_{Q_{0}}}|\nabla u(X)|^{2}\delta(X)\,dX\lesssim\varepsilon^{-2}\ell(Q_{0})^{n}\approx\varepsilon^{-2}\sigma(Q_{0}),

which proves the claim. ∎

4.2. A stopping time decomposition for the family of blue cubes

We now move to decomposing the collection of blue cubes into more manageable subcollections using a stopping time procedure. A similar idea is utilized in [HMM16, p. 2360] but, due to our geometry, we use different stopping time conditions and different analysis of the subcollections. Set =(𝒮)\mathcal{L}=\mathcal{L}(\mathcal{S}) to be the collection of blue cubes in 𝒮\mathcal{S}. We first take the largest blue cube in 𝒮\mathcal{S} with respect to side length (if there is more than one such cube, we just pick one) and denote this cube by Q(𝐒1)Q({\bf S}_{1}). The cube Q(𝐒1)Q({\bf S}_{1}) will be the maximal cube in our first refined stopping time regime 𝐒1{\bf S}_{1}. We let 𝐒1\mathcal{F}_{{\bf S}_{1}} be the collection of cubes Q𝒮𝔻Q(𝐒1){Q(𝐒1)}Q\in\mathcal{S}\cap\mathbb{D}_{Q({\bf S}_{1})}\setminus\{Q({\bf S}_{1})\} such that QQ is a maximal cube with respect to having one of the following three properties:

  1. (1)

    QQ or one of its siblings444Here QQ^{\prime} is a sibling of QQ if Q,Q𝔻kQ,Q^{\prime}\in\mathbb{D}_{k} and Q,QQ𝔻k1Q,Q^{\prime}\subset Q^{*}\in\mathbb{D}_{k-1}, that is, QQ and QQ^{\prime} have the same parent. For simplicity, we consider QQ to be its own sibling. is red.

  2. (2)

    QQ and all of its siblings are blue, but for some QQ^{\prime} that is either QQ or a sibling of QQ it holds that

    |u(XQ)u(XQ(𝐒1))|>ε/100.|u(X_{Q^{\prime}})-u(X_{Q({\bf S}_{1})})|>\varepsilon/100.
  3. (3)

    QQ is yellow.

Recall that for every P𝔻P\in\mathbb{D}, the point XPX_{P} is a corkscrew point relative to the center xPx_{P} at scale 105a0(P)10^{-5}a_{0}\ell(P), as defined in Section 2.3. We now set 𝐒1𝔻𝐒1,Q(𝐒1){\bf S}_{1}\coloneqq\mathbb{D}_{\mathcal{F}_{{\bf S}_{1}},Q({\bf S}_{1})}. By construction, 𝐒1{\bf S}_{1} is a coherent stopping time regime in the sense of Definition 3.17. Since the cubes in 𝐒1{\bf S}_{1} are blue and none of them satisfy the stopping condition (2) above, we know that

(4.3) |u(X)u(XQ(𝐒1))|ε/50 for every XΩ𝐒1,Q(𝐒1).|u(X)-u(X_{Q({\bf S}_{1})})|\leq\varepsilon/50\quad\text{ for every }X\in\Omega_{\mathcal{F}_{{\bf S}_{1}},Q({\bf S}_{1})}^{*}.

We now express 𝐒1\mathcal{F}_{{\bf S}_{1}} as a union of three collections,

𝐒1=𝐒1R𝐒1SB𝐒1Y,\displaystyle\mathcal{F}_{{\bf S}_{1}}=\mathcal{F}_{{\bf S}_{1}}^{\operatorname{R}}\cup\mathcal{F}_{{\bf S}_{1}}^{\operatorname{SB}}\cup\mathcal{F}_{{\bf S}_{1}}^{\operatorname{Y}},

where 𝐒1R\mathcal{F}_{{\bf S}_{1}}^{\operatorname{R}} contains the cubes for which (1) holds, 𝐒1SB\mathcal{F}_{{\bf S}_{1}}^{\operatorname{SB}} contains the cubes for which (2) holds and 𝐒1Y\mathcal{F}_{{\bf S}_{1}}^{\operatorname{Y}} contains the cubes for which (3) holds. The superscripts stand for “red”, “stopping blue” and “yellow”. We note that the collections 𝐒1R\mathcal{F}_{{\bf S}_{1}}^{\operatorname{R}} and 𝐒1SB\mathcal{F}_{{\bf S}_{1}}^{\operatorname{SB}} are disjoint but 𝐒1Y\mathcal{F}_{{\bf S}_{1}}^{\operatorname{Y}} may overlap with both of them.

We now continue this way: we let Q(𝐒2)Q({\bf S}_{2}) be the largest blue cube in 𝒮𝐒1\mathcal{S}\setminus{\bf S}_{1} with respect to side length, we extract the collection of maximal stopping cubes Q(𝐒2)\mathcal{F}_{Q({\bf S}_{2})} (with an updated stopping condition (2)), we define the coherent stopping regime 𝐒2{\bf S}_{2} and the collections 𝐒2R\mathcal{F}_{{\bf S}_{2}}^{\operatorname{R}}, 𝐒2SB\mathcal{F}_{{\bf S}_{2}}^{\operatorname{SB}} and 𝐒2Y\mathcal{F}_{{\bf S}_{2}}^{\operatorname{Y}}, choose the largest blue cube Q(𝐒3)Q({\bf S}_{3}) in 𝒮𝐒1𝐒2\mathcal{S}\setminus{\bf S}_{1}\cup{\bf S}_{2}, and so on. Since each 𝐒i{\bf S}_{i} contains at least the cube Q(𝐒i)Q({\bf S}_{i}), we know that this procedure exhausts \mathcal{L} and gives us a disjoint decomposition =j𝐒j\mathcal{L}=\cup_{j}{\bf S}_{j} where each 𝐒j{\bf S}_{j} is a coherent stopping time regime. Just like (4.3), we have the oscillation estimate

(4.4) |u(X)u(XQ(𝐒j))|ε/50 for every XΩ𝐒j,Q(𝐒j)|u(X)-u(X_{Q({\bf S}_{j})})|\leq\varepsilon/50\quad\text{ for every }X\in\Omega_{\mathcal{F}_{{\bf S}_{j}},Q({\bf S}_{j})}^{*}

for every jj. We also get the collections 𝐒jR\mathcal{F}_{{\bf S}_{j}}^{\operatorname{R}}, 𝐒jSB\mathcal{F}_{{\bf S}_{j}}^{\operatorname{SB}} and 𝐒jY\mathcal{F}_{{\bf S}_{j}}^{\operatorname{Y}} for each jj.

Our next goal is to show that the maximal cubes {Q(𝐒j)}j\{Q({\bf S}_{j})\}_{j} satisfy a Carleson packing condition. This goal is an analog of [HMM16, Lemma 5.16] but since the proof of this lemma is based on the use of an “NSN\lesssim S” estimate in sawtooth regions (which is possible in the presence of uniform rectifiability of Ω)\partial\Omega), this is the part where we significantly depart from [HMM16]. Following an idea in [DS91], we let λ(0,1010)\lambda\in(0,10^{-10}) be a small parameter (to be chosen) and break the stopping times into four groups. We say that 𝐒j{\bf S}_{j} is of

  1. \bullet

    Type 1 (T1) if σ(Q(𝐒j)Q𝐒jQ)λσ(Q(𝐒j))\sigma(Q({\bf S}_{j})\setminus\cup_{Q\in\mathcal{F}_{{\bf S}_{j}}}Q)\geq\lambda\sigma(Q({\bf S}_{j})).

  2. \bullet

    Type 2 (T2) if σ(Q𝐒jRQ)λσ(Q(𝐒j))\sigma(\cup_{Q\in\mathcal{F}_{{\bf S}_{j}}^{\operatorname{R}}}Q)\geq\lambda\sigma(Q({\bf S}_{j})).

  3. \bullet

    Type 3 (T3) if σ(Q𝐒jYQ)λσ(Q(𝐒j))\sigma(\cup_{Q\in\mathcal{F}_{{\bf S}_{j}}^{\operatorname{Y}}}Q)\geq\lambda\sigma(Q({\bf S}_{j})).

  4. \bullet

    Type 4 (T4) if 𝐒j{\bf S}_{j} is not type 1, 2 or 3.

The Carleson packing condition for the cubes Q(𝐒j)Q({\bf S}_{j}) follows in a straightforward way when 𝐒j{\bf S}_{j} is of Type 1, 2 or 3:

Lemma 4.5.

We have the following Carleson packing conditions for the maximal cubes of the subregimes 𝐒j{\bf S}_{j} in the decomposition (𝒮)=j𝐒j\mathcal{L}(\mathcal{S})=\cup_{j}{\bf S}_{j}: for any P𝔻P\in\mathbb{D}, we have

j:𝐒j is T1Q(𝐒j)Pσ(Q(𝐒j))+j:𝐒j is T3Q(𝐒j)Pσ(Q(𝐒j))\displaystyle\sum_{\begin{subarray}{c}j\colon{\bf S}_{j}\text{ is T1}\\ Q({\bf S}_{j})\subset P\end{subarray}}\sigma(Q({\bf S}_{j}))+\sum_{\begin{subarray}{c}j\colon{\bf S}_{j}\text{ is T3}\\ Q({\bf S}_{j})\subset P\end{subarray}}\sigma(Q({\bf S}_{j})) 1λσ(P),and,\displaystyle\lesssim\frac{1}{\lambda}\sigma(P),\qquad\text{and,}
j:𝐒j is T2Q(𝐒j)Pσ(Q(𝐒j))\displaystyle\sum_{\begin{subarray}{c}j\colon{\bf S}_{j}\text{ is T2}\\ Q({\bf S}_{j})\subset P\end{subarray}}\sigma(Q({\bf S}_{j})) 1λε2σ(P).\displaystyle\lesssim\frac{1}{\lambda\varepsilon^{2}}\sigma(P).
Proof.

For the regimes of Type 1, we first notice that if Q(𝐒i)Q(𝐒j)Q({\bf S}_{i})\subsetneq Q({\bf S}_{j}), then Q(𝐒i)QQ({\bf S}_{i})\subset Q for some Q𝐒jQ\in\mathcal{F}_{{\bf S}_{j}}. In particular, the sets Q(𝐒j)Q𝐒jQQ({\bf S}_{j})\setminus\cup_{Q\in\mathcal{F}_{{\bf S}_{j}}}Q are pairwise disjoint. Thus, by the definition of Type 1, we get

j:𝐒j is T1Q(𝐒j)Pσ(Q(𝐒j))1λj:𝐒j is T1Q(𝐒j)Pσ(Q(𝐒j)Q𝐒jQ)1λσ(P).\displaystyle\sum_{\begin{subarray}{c}j\colon{\bf S}_{j}\text{ is T1}\\ Q({\bf S}_{j})\subset P\end{subarray}}\sigma(Q({\bf S}_{j}))\leq\frac{1}{\lambda}\sum_{\begin{subarray}{c}j\colon{\bf S}_{j}\text{ is T1}\\ Q({\bf S}_{j})\subset P\end{subarray}}\sigma(Q({\bf S}_{j})\setminus\cup_{Q\in\mathcal{F}_{{\bf S}_{j}}}Q)\leq\frac{1}{\lambda}\sigma(P).

For Type 3 regimes, since the cubes Q𝐒jY𝒴(𝒮)Q\in\mathcal{F}_{{\bf S}_{j}}^{\operatorname{Y}}\subset\mathcal{Y}(\mathcal{S}) are yellow cubes in 𝒮\mathcal{S}, they are disjoint. Thus, the claim for the regimes of Type 3 follows immediately from definition.

For the regimes of Type 2, we recall that if Q𝐒jRQ\in\mathcal{F}_{{\bf S}_{j}}^{\operatorname{R}}, then QQ is red or one of its siblings is red. In particular, each Q𝐒jRQ\in\mathcal{F}_{{\bf S}_{j}}^{\operatorname{R}} has approximately the same measure as some red sibling RQQ(𝐒j)R_{Q}\subset Q({\bf S}_{j}) of QQ. If there is more than one red sibling, we just choose one of them for each Q𝐒jRQ\in\mathcal{F}_{{\bf S}_{j}}^{\operatorname{R}}. On the other hand, since each cube has only a uniformly bounded number of siblings, for each Qj𝐒jRQ\in\bigcup_{j}\mathcal{F}_{{\bf S}_{j}}^{\operatorname{R}} we can have RQ=RQR_{Q}=R_{Q^{\prime}} only for a uniformly bounded number of cubes QQ^{\prime}. Thus, we get

j:𝐒j is T2Q(𝐒j)Pσ(Q(𝐒j))1λj:𝐒j is T2Q(𝐒j)PQ𝐒jRσ(Q)1λj:𝐒j is T2Q(𝐒j)PQ𝐒jRσ(RQ)1λR,RPσ(R)1λε2σ(P),\sum_{\begin{subarray}{c}j\colon{\bf S}_{j}\text{ is T2}\\ Q({\bf S}_{j})\subset P\end{subarray}}\sigma(Q({\bf S}_{j}))\leq\frac{1}{\lambda}\sum_{\begin{subarray}{c}j\colon{\bf S}_{j}\text{ is T2}\\ Q({\bf S}_{j})\subset P\end{subarray}}\sum_{Q\in\mathcal{F}_{{\bf S}_{j}}^{\operatorname{R}}}\sigma(Q)\approx\frac{1}{\lambda}\sum_{\begin{subarray}{c}j\colon{\bf S}_{j}\text{ is T2}\\ Q({\bf S}_{j})\subset P\end{subarray}}\sum_{Q\in\mathcal{F}_{{\bf S}_{j}}^{\operatorname{R}}}\sigma(R_{Q})\\ \lesssim\frac{1}{\lambda}\sum_{R\in\mathcal{R},R\subset P}\sigma(R)\lesssim\frac{1}{\lambda\varepsilon^{2}}\sigma(P),

where we used Lemma 4.2 in the final estimate. ∎

By Lemma 4.5, to show the Carleson packing condition for the collection {Q(𝐒j)}j\{Q({\bf S}_{j})\}_{j} it remains only to consider the regimes 𝐒j{\bf S}_{j} of Type 4.

Lemma 4.6.

There exists λ0>0\lambda_{0}>0 depending only on structural constants, ellipticity, and the ωL,ΩA(σ)\omega_{L,\Omega}\in A_{\infty}(\sigma) constants, such that for any λ(0,λ0)\lambda\in(0,\lambda_{0}), there exist constants C1,C21C_{1},C_{2}\geq 1 depending only on structural constants, ellipticity, and the ωL,ΩA(σ)\omega_{L,\Omega}\in A_{\infty}(\sigma) constants (and independent of ε\varepsilon, λ\lambda, and 𝒮\mathcal{S}), so that

j:𝐒j is T4Q(𝐒j)Pσ(Q(𝐒j))\displaystyle\sum_{\begin{subarray}{c}j\colon{\bf S}_{j}\text{ is T4}\\ Q({\bf S}_{j})\subset P\end{subarray}}\sigma(Q({\bf S}_{j})) C1ε2σ(P).\displaystyle\leq\frac{C_{1}}{\varepsilon^{2}}\sigma(P).

for every P𝔻P\in\mathbb{D}. In particular, we have

j:Q(𝐒j)Pσ(Q(𝐒j))\displaystyle\sum_{j\colon Q({\bf S}_{j})\subset P}\sigma(Q({\bf S}_{j})) C2λε2σ(P).\displaystyle\leq\frac{C_{2}}{\lambda\varepsilon^{2}}\sigma(P).

Proving Lemma 4.6 is much more delicate than proving Lemma 4.5 and we do this in several steps. The key idea is to reduce the proof to proving estimates for which we can use lemmas from Section 3.

Let us fix a regime 𝐒j{\bf S}_{j} that is of Type 4. Since 𝐒j{\bf S}_{j} is not of Type 1, 2 or 3, we know that, roughly speaking, at the “bottom” of the sawtooth domain Ω𝐒j,Q(𝐒j)\Omega_{\mathcal{F}_{{\bf S}_{j}},Q({\bf S}_{j})} there is a large region where uu has some (uniform) oscillation from the value u(XQ(𝐒j))u(X_{Q({\bf S}_{j})}). Let us be more precise. Since 𝐒j{\bf S}_{j} is not of Type 1, we know that σ(Q(𝐒j)Q𝐒jQ)<λσ(Q(𝐒j)\sigma(Q({\bf S}_{j})\setminus\cup_{Q\in\mathcal{F}_{{\bf S}_{j}}}Q)<\lambda\sigma(Q({\bf S}_{j}), and since 𝐒j{\bf S}_{j} is not of Type 2 or 3, we have

σ((Q𝐒jRQ)(Q𝐒jYQ))<2λσ(Q(𝐒j)).\displaystyle\sigma\big{(}(\cup_{Q\in\mathcal{F}_{{\bf S}_{j}}^{\operatorname{R}}}Q)\cup(\cup_{Q\in\mathcal{F}_{{\bf S}_{j}}^{\operatorname{Y}}}Q)\big{)}<2\lambda\sigma(Q({\bf S}_{j})).

Thus, since 𝐒j=𝐒jR𝐒jSB𝐒jY\mathcal{F}_{{\bf S}_{j}}=\mathcal{F}_{{\bf S}_{j}}^{\operatorname{R}}\cup\mathcal{F}_{{\bf S}_{j}}^{\operatorname{SB}}\cup\mathcal{F}_{{\bf S}_{j}}^{\operatorname{Y}}, it holds that

σ(Q𝐒jSBQ)(13λ)σ(Q(𝐒j)).\displaystyle\sigma\big{(}\cup_{Q\in\mathcal{F}_{{\bf S}_{j}}^{\operatorname{SB}}}Q\big{)}\geq(1-3\lambda)\sigma(Q({\bf S}_{j})).

Let N=N(j)N=N(j) be so large that the subcollection

N,jSB{Q𝐒jSB:(Q)>2N(Q(𝐒j))}\displaystyle\mathcal{F}_{N,j}^{\operatorname{SB}}\coloneqq\{Q\in\mathcal{F}_{{\bf S}_{j}}^{\operatorname{SB}}\colon\ell(Q)>2^{-N}\ell(Q({\bf S}_{j}))\}

satisfies

(4.7) σ(QN,jSBQ)(14λ)σ(Q(𝐒j)).\sigma\big{(}\cup_{Q\in\mathcal{F}_{N,j}^{\operatorname{SB}}}Q\big{)}\geq(1-4\lambda)\sigma(Q({\bf S}_{j})).

Our estimates will not depend on NN but we work with the subcollection N,jSB\mathcal{F}_{N,j}^{\operatorname{SB}} to avoid dealing with estimates on the boundary. We let N,j\mathcal{F}_{N,j} denote the collection of maximal cubes in the collection 𝐒j𝔻N,Q(𝐒j)\mathcal{F}_{{\bf S}_{j}}\cup\mathbb{D}_{N,Q({\bf S}_{j})}, where 𝔻N,Q={Q𝔻Q:(Q)=2N(Q)}\mathbb{D}_{N,Q}=\{Q^{\prime}\in\mathbb{D}_{Q}\colon\ell(Q^{\prime})=2^{-N}\ell(Q)\} as earlier. We note that ΩN,j,Q(𝐒j)Ω𝐒j,Q(𝐒j)\Omega_{\mathcal{F}_{N,j},Q({\bf S}_{j})}\subseteq\Omega_{\mathcal{F}_{{\bf S}_{j}},Q({\bf S}_{j})}. Then

N,j=N,jSBN,jO, where N,jON,jN,jSB,\displaystyle\mathcal{F}_{N,j}=\mathcal{F}_{N,j}^{\operatorname{SB}}\cup\mathcal{F}_{N,j}^{\operatorname{O}},\ \text{ where }\ \mathcal{F}_{N,j}^{\operatorname{O}}\coloneqq\mathcal{F}_{N,j}\setminus\mathcal{F}_{N,j}^{\operatorname{SB}},

where the superscript O in N,jO\mathcal{F}_{N,j}^{\operatorname{O}} stands for “other” cubes. By (4.7) we have

(4.8) σ(QN,jOQ)=QN,jOσ(Q)(4λ)σ(Q(𝐒j)).\displaystyle\sigma\big{(}\cup_{Q\in\mathcal{F}_{N,j}^{\operatorname{O}}}Q\big{)}=\sum_{Q\in\mathcal{F}_{N,j}^{\operatorname{O}}}\sigma(Q)\leq(4\lambda)\sigma(Q({\bf S}_{j})).

With the notation above, we can formulate our key estimate for the proof of Lemma 4.6:

Lemma 4.9.

Suppose that 𝐒j{\bf S}_{j} is a stopping time regime of Type 4. There exists λ0>0\lambda_{0}>0 depending only on structural constants, ellipticity, and the ωL,ΩA(σ)\omega_{L,\Omega}\in A_{\infty}(\sigma) constants, such that for any λ(0,λ0)\lambda\in(0,\lambda_{0}), there exists a constant C31C_{3}\geq 1 depending only on structural constants, ellipticity, and the ωL,ΩA(σ)\omega_{L,\Omega}\in A_{\infty}(\sigma) constants (and independent of ε\varepsilon, λ\lambda, NN, jj, and 𝒮\mathcal{S}), so that the following estimate holds:

(4.10) σ(Q(𝐒j))C3ε2ΩN,j,Q(𝐒j)|u(Y)|2δ(Y)𝑑Y.\displaystyle\sigma(Q({\bf S}_{j}))\leq\frac{C_{3}}{\varepsilon^{2}}\iint_{\Omega_{\mathcal{F}_{N,j},Q({\bf S}_{j})}}|\nabla u(Y)|^{2}\delta(Y)\,dY.

Taking Lemma 4.9 for granted momentarily, we can prove Lemma 4.6 in a straightforward way:

Proof of Lemma 4.6.

Fix P𝔻P\in\mathbb{D}. Then we have

j:𝐒j is T4Q(𝐒j)Pσ(Q(𝐒j))C3ε2j:𝐒j is T4Q(𝐒j)PΩN,j,Q(𝐒j)|u(Y)|2δ(Y)𝑑YC3ε2TP|u(Y)|2δ(Y)𝑑YC3ε2σ(P),\sum_{\begin{subarray}{c}j\colon{\bf S}_{j}\text{ is T4}\\ Q({\bf S}_{j})\subset P\end{subarray}}\sigma(Q({\bf S}_{j}))\leq\frac{C_{3}}{\varepsilon^{2}}\sum_{\begin{subarray}{c}j\colon{\bf S}_{j}\text{ is T4}\\ Q({\bf S}_{j})\subset P\end{subarray}}\iint_{\Omega_{\mathcal{F}_{N,j},Q({\bf S}_{j})}}|\nabla u(Y)|^{2}\delta(Y)\,dY\\ \lesssim\frac{C_{3}}{\varepsilon^{2}}\iint_{T_{P}}|\nabla u(Y)|^{2}\delta(Y)\,dY\lesssim\frac{C_{3}}{\varepsilon^{2}}\sigma(P),

where we used Lemma 4.9, the fact that the bounded overlap of the regions UQU_{Q} and the disjointness of the collections 𝐒j{\bf S}_{j} imply that the regions ΩN(j),j,Q(𝐒j)\Omega_{\mathcal{F}_{N(j),j},Q({\bf S}_{j})} have bounded overlaps, the fact that TPB(xP,C(P))T_{P}\subset B(x_{P},C\ell(P)) and Lemma 3.1. The rest of the claim follows now from Lemma 4.5. ∎

4.3. Proof of Lemma 4.9: A high oscillation estimate

Let us then start processing the estimate (4.10). Let 𝐒j{\bf S}_{j} be a fixed stopping time regime of Type 4. To relax the notation, we denote

𝐒𝔻N,j,Q(𝐒j)𝐒j,Q(𝐒)Q(𝐒j),XXQ(𝐒j),\displaystyle{\bf S}_{*}\coloneqq\mathbb{D}_{\mathcal{F}_{N,j},Q({\bf S}_{j})}\subseteq{\bf S}_{j},\qquad Q({\bf S}_{*})\coloneqq Q({\bf S}_{j}),\qquad X_{*}\coloneqq X_{Q({\bf S}_{j})},
ΩΩN,j,Q(𝐒j),N,j,SBN,jSB, andON,jO.\displaystyle\Omega_{*}\coloneqq\Omega_{\mathcal{F}_{N,j},Q({\bf S}_{j})},\qquad\mathcal{F}\coloneqq\mathcal{F}_{N,j},\qquad\mathcal{F}^{\operatorname{SB}}\coloneqq\mathcal{F}_{N,j}^{\operatorname{SB}},\quad\text{ and}\quad\mathcal{F}^{\operatorname{O}}\coloneqq\mathcal{F}_{N,j}^{\operatorname{O}}.

Recall that Q(𝐒)𝒮Q({\bf S}_{*})\in\mathcal{S} and that 𝒮\mathcal{S} is a coherent stopping time regime satisfying the property (iv) in Lemma 3.18. Thus, by using Lemma 3.19, Lemma 3.16 and Lemma 3.10 in this order, we get

(4.11) Ω|u(Y)|2δ(Y)𝑑Yσ(Q(𝐒))Ω|u(Y)|2GL,Ω(X,Y)𝑑Yσ(Q(𝐒))Ω|u(Y)|2GL,Ω(X,Y)𝑑Yσ(Q(𝐒))Ω(u(y)u(X))2𝑑ω(y),\iint_{\Omega_{*}}|\nabla u(Y)|^{2}\delta(Y)\,dY\approx\sigma(Q({\bf S}_{*}))\iint_{\Omega_{*}}|\nabla u(Y)|^{2}G_{L,\Omega}(X_{*},Y)\,dY\\ \geq\sigma(Q({\bf S}_{*}))\iint_{\Omega_{*}}|\nabla u(Y)|^{2}G_{L,\Omega_{*}}(X_{*},Y)\,dY\approx\sigma(Q({\bf S}_{*}))\int_{\partial\Omega_{*}}(u(y)-u(X_{*}))^{2}\,d\omega_{*}(y),

where

ω\omega_{*} is the elliptic measure for LL in Ω\Omega_{*} with pole at XX_{*}.

Thus, Lemma 4.9 follows immediately from the following estimate. Recall that λ\lambda is the parameter we used when we defined the Types 1–4 for the stopping time regimes.

Lemma 4.12.

There exists λ0>0\lambda_{0}>0 depending on structural constants, ellipticity, and the ωL,ΩA(σ)\omega_{L,\Omega}\in A_{\infty}(\sigma) constants, such that for any λ(0,λ0)\lambda\in(0,\lambda_{0}), there exists a constant c4>0c_{4}>0 depending only on structural constants, ellipticity, and the ωL,ΩA(σ)\omega_{L,\Omega}\in A_{\infty}(\sigma) constants (and independent of ε\varepsilon, λ\lambda, NN, jj, and 𝒮\mathcal{S}), so that the following estimate holds:

Ω(u(y)u(X))2𝑑ω(y)c4ε2.\int_{\partial\Omega_{*}}(u(y)-u(X_{*}))^{2}\,d\omega_{*}(y)\geq c_{4}\varepsilon^{2}.

For the the proof of Lemma 4.12, we need some auxiliary constructions and estimates. Recall that X=XQ(𝐒)X_{*}=X_{Q({\bf S}_{*})} is a corkscrew point relative to Q(𝐒)Q({\bf S}_{*}) at scale r105a0(Q(𝐒))r_{*}\coloneqq 10^{-5}a_{0}\ell(Q({\bf S}_{*})) (see Section 2.4). Let x^Ω\hat{x}_{*}\in\partial\Omega be a touching point for XX_{*} on Ω\partial\Omega, that is, |x^X|=dist(X,Ω)|\hat{x}_{*}-X_{*}|=\operatorname{dist}(X_{*},\partial\Omega). For ξ[0,1]\xi\in[0,1], consider the points X(ξ)x^+ξ(Xx^)X(\xi)\coloneqq\hat{x}_{*}+\xi(X_{*}-\hat{x}_{*}) which lie on the line segment from x^\hat{x}_{*} to XX_{*}. Since we know that x^Ω¯\hat{x}_{*}\not\in\overline{\Omega}_{*} and XΩX_{*}\in\Omega_{*}, there exists ξ0(0,1)\xi_{0}\in(0,1) such that X(ξ0)ΩX(\xi_{0})\in\partial\Omega_{*}. Now we set X=X(ξ0)X_{**}=X(\xi_{0}), and ΔB(X,r)Ω\Delta_{*}\coloneqq B(X_{**},r_{*})\cap\partial\Omega_{*}. Since we are working with the truncated collection of stopping cubes =N,j\mathcal{F}=\mathcal{F}_{N,j}, we know that ΩΩ=Ø\partial\Omega\cap\partial\Omega_{*}=\mbox{{\O}}. In particular, ΔΩ\Delta_{*}\subset\Omega. By Lemma 2.26, we know that

(4.13) if yΔ, then y^ΔQ(𝐒)Q(𝐒),\displaystyle\text{if }y\in\Delta_{*},\ \ \text{ then }\hat{y}\in\Delta_{Q({\bf S}_{*})}\subseteq Q({\bf S}_{*}),

where y^\hat{y} is a touching point for yy in Ω\Omega, that is, |yy^|=dist(y,Ω)|y-\hat{y}|=\operatorname{dist}(y,\partial\Omega).

By Lemma 2.18, we know that Ω\Omega_{*} is also a uniform domain with Ahlfors regular boundary.555By Lemma 2.18, the structural constants of Ω=ΩN,j,Q(𝐒j)\Omega_{*}=\Omega_{\mathcal{F}_{N,j},Q({\bf S}_{j})} do not depend on jj or the truncation parameter N=N(j)N=N(j). Thus, by Lemma 2.36, we have

(4.14) ωL,ΩX~(Δ)1,\displaystyle\omega^{\widetilde{X}}_{L,\Omega_{*}}(\Delta_{*})\gtrsim 1,

where X~\widetilde{X} is a corkscrew point relative to XX_{**} at scale rr_{*} in the domain Ω\Omega_{*}. Recall that by the construction in Section 2.3, we know that B(X,δ(X)/2)ΩB(X_{*},\delta(X_{*})/2)\subset\Omega_{*} and diam(Ω)(Q(𝐒))\operatorname{diam}(\Omega_{*})\approx\ell(Q({\bf S}_{*})), and we have δ(X)(Q(𝐒))rδ(X~)\delta(X_{*})\approx\ell(Q({\bf S}_{*}))\approx r_{*}\approx\delta(\widetilde{X}) by the definition of XX_{*} and X~\widetilde{X}. Thus, by the Harnack chain property of Ω\Omega_{*}, there exists a Harnack chain of uniformly bounded length from X~\widetilde{X} to XX_{*} inside Ω\Omega_{*}. Thus, by (4.14), formula (2.30) and Lemma 2.29 (that is, Harnack inequality), there exists a constant c>0c_{*}>0 that depends only on structural constants such that

(4.15) ω(Δ)=ωL,ΩX(Δ)>c.\displaystyle\omega_{*}(\Delta_{*})=\omega^{X_{*}}_{L,\Omega_{*}}(\Delta_{*})>c_{*}.

Next, we will construct a cover of Δ\Delta_{*} that consists of dilated surface balls on Ω\partial\Omega_{*} associated to the cubes QQ\in\mathcal{F}. We will construct the cover in such a way that there is oscillation of uu on the balls associated to cubes in SB\mathcal{F}^{\operatorname{SB}} and the balls associated to cubes in O\mathcal{F}^{\operatorname{O}} do not have much ω\omega_{*}-mass, provided λ\lambda is sufficiently small. Given Q𝔻Q\in\mathbb{D}, denote by x^Q\hat{x}_{Q} a touching point for the corkscrew point XQX_{Q}. For θ[0,1]\theta\in[0,1], recall that we denote PQ(θ)=x^Q+θ(XQx^Q)P_{Q}(\theta)=\hat{x}_{Q}+\theta(X_{Q}-\hat{x}_{Q}), and by Lemma 2.21 we showed that there exists θ0(0,1)\theta_{0}\in(0,1) such that if for some Q𝔻Q^{\prime}\in\mathbb{D} we have B(PQ(θ0),γθ010rQ)UQØB(P_{Q}(\theta_{0}),\tfrac{\gamma\theta_{0}}{10}r_{Q})\cap U_{Q^{\prime}}\neq\mbox{{\O}}, then QQQ^{\prime}\subset Q and (Q)<(Q)\ell(Q^{\prime})<\ell(Q), where γ\gamma is the corkscrew constant in Definition 2.1.

Fix QQ\in\mathcal{F}. Then its parent Q~\widetilde{Q} satisfies Q~𝔻,Q(𝐒)\widetilde{Q}\in\mathbb{D}_{\mathcal{F},Q({\bf S}_{*})} and hence, by the construction of the Whitney regions in Section 2.3, we have PQ(1)=XQUQ~ΩP_{Q}(1)=X_{Q}\in U_{\widetilde{Q}}\subset\Omega_{*}. By Lemma 2.21, we also know that PQ(θ0)ΩP_{Q}(\theta_{0})\not\in\Omega_{*}. Indeed, otherwise PQ(θ0)UQP_{Q}(\theta_{0})\in U_{Q^{\prime}} for a cube Q𝔻,Q(𝐒)Q^{\prime}\in\mathbb{D}_{\mathcal{F},Q({\bf S}_{*})}, but by Lemma 2.21 we would then have QQQ^{\prime}\subset Q with (Q)<(Q)\ell(Q^{\prime})<\ell(Q). This is impossible since QQ\in\mathcal{F}. Thus, there exists θ(θ0,1)\theta^{\prime}\in(\theta_{0},1) such that

XQPQ(θ)Ω.\displaystyle X^{*}_{Q}\coloneqq P_{Q}(\theta^{\prime})\in\partial\Omega_{*}.

We set

ΔQB(XQ,λθ02rQ)Ω and MΔQB(XQ,Mλθ02rQ)Ω\displaystyle\Delta^{*}_{Q}\coloneqq B(X^{*}_{Q},\tfrac{\lambda\theta_{0}}{2}r_{Q})\cap\partial\Omega_{*}\quad\text{ and }\quad M\Delta^{*}_{Q}\coloneqq B(X^{*}_{Q},M\tfrac{\lambda\theta_{0}}{2}r_{Q})\cap\partial\Omega_{*}

for a constant M1M\geq 1 to be chosen momentarily.

Let us describe some of the properties of ΔQ\Delta^{*}_{Q}. First, by Lemma 2.25 and definition of ΔQ\Delta_{Q}^{*}, it holds that

(4.16) ΔQΞQUQ\Delta^{*}_{Q}\subseteq\Xi_{Q}\subseteq U_{Q^{\prime}}^{*}

whenever QQ^{\prime} is a sibling of QQ, where

ΞQ=θ[θ0,1]B(PQ(θ),γθ010rQ).\displaystyle\Xi_{Q^{\prime}}=\bigcup_{\theta\in[\theta_{0},1]}B\big{(}P_{Q^{\prime}}(\theta),\tfrac{\gamma\theta_{0}}{10}r_{Q^{\prime}}\big{)}.

Next, let us observe that if QSBQ\in\mathcal{F}^{\operatorname{SB}}, then there exists a sibling QQ^{\prime} of QQ such that QQ^{\prime} is blue and |u(XQ)u(XQ(𝐒))|>ε/100|u(X_{Q^{\prime}})-u(X_{Q({\bf S}_{*})})|>\varepsilon/100. Thus, for every QSBQ\in\mathcal{F}^{\operatorname{SB}}, it holds that

(4.17) |u(X)u(X)||u(XQ)u(XQ(𝐒))||u(XQ)u(X)|ε/100ε/1000ε/200,|u(X)-u(X_{*})|\geq|u(X_{Q^{\prime}})-u(X_{Q({\bf S}_{*})})|-|u(X_{Q^{\prime}})-u(X)|\\ \geq\varepsilon/100-\varepsilon/1000\geq\varepsilon/200,

for every XΔQX\in\Delta^{*}_{Q} since QQ^{\prime} is blue and ΔQUQ\Delta^{*}_{Q}\subset U_{Q^{\prime}}^{*} by (4.16).

Lemma 4.18.

There exists M1M\geq 1, depending only on structural constants, such that

Δ=B(X,r)ΩQMΔQ.\displaystyle\Delta_{*}=B(X_{**},r_{*})\cap\partial\Omega_{*}\subset\bigcup_{Q\in\mathcal{F}}M\Delta^{*}_{Q}.
Proof.

Fix yΔΩy\in\Delta_{*}\subset\partial\Omega_{*} and a touching point y^Ω\hat{y}\in\partial\Omega for yy. Then by (4.13) we have y^ΔQ(𝐒)\hat{y}\in\Delta_{Q({\bf S}_{*})}. Since XX_{**} lies on the line segment from the corkscrew point XX_{*} relative to xQ(𝐒)x_{Q({\bf S}_{*})} at scale rr_{*} to its touching point x^\hat{x}_{*}, we have |x^X|=dist(X,Ω)dist(X,Ω)r=105a0(Q(𝐒))|\hat{x}_{*}-X_{**}|=\operatorname{dist}(X_{**},\partial\Omega)\leq\operatorname{dist}(X_{*},\partial\Omega)\leq r_{*}=10^{-5}a_{0}\ell(Q({\bf S}_{*})). Thus, yB(x^,2r)=B(x^Q(𝐒),2(10)5a0(Q(𝐒)))y\in B(\hat{x}_{*},2r_{*})=B(\hat{x}_{Q({\bf S}_{*})},2(10)^{-5}a_{0}\ell(Q({\bf S}_{*}))). By (4.13), we have

(4.19) |yy^|=dist(y,Ω)=dist(y,Q(𝐒))(Q(𝐒)),|y-\hat{y}|=\operatorname{dist}(y,\partial\Omega)=\operatorname{dist}(y,Q({\bf S}))\leq\ell(Q({\bf S}_{*})),

and by the definition of 𝒲Q(K0)\mathcal{W}_{Q^{\prime}}(K_{0}) (which we used in the construction of the Whitney regions), for any cube QQ^{\prime} it holds that

(4.20) if y^Q and Cτ1K01|yy^|(Q)CτK0|yy^|, then yint(UQ),\displaystyle\text{if }\hat{y}\in Q^{\prime}\text{ and }C_{\tau}^{-1}K_{0}^{-1}|y-\hat{y}|\leq\ell(Q^{\prime})\leq C_{\tau}K_{0}|y-\hat{y}|,\quad\text{ then }y\in\operatorname{int}(U_{Q^{\prime}}),

where τ\tau is the dilation parameter in the definition of UQU_{Q}. Since 𝐒{\bf S}_{*} and \mathcal{F} are the truncated collections, we have Q(𝐒)=QQQ({\bf S}_{*})=\cup_{Q\in\mathcal{F}}Q. In particular, by Lemma 2.26, we have y^Q(𝐒)=QQ\hat{y}\in Q({\bf S}_{*})=\cup_{Q\in\mathcal{F}}Q. Let Qy^Q_{\hat{y}}\in\mathcal{F} be the cube such that y^Qy^\hat{y}\in Q_{\hat{y}}. We now have (Qy^)Cτ1K01|yy^|\ell(Q_{\hat{y}})\geq C_{\tau}^{-1}K_{0}^{-1}|y-\hat{y}| for the same constant CτC_{\tau} as in (4.20), since otherwise there exists a cube QQ^{\prime} such that Qy^QQ(𝐒)Q_{\hat{y}}\subset Q^{\prime}\subseteq Q({\bf S}_{*}) with Cτ1K01|yy^|(Q)CτK0|yy^|C^{-1}_{\tau}K_{0}^{-1}|y-\hat{y}|\leq\ell(Q^{\prime})\leq C_{\tau}K_{0}|y-\hat{y}|. This is not possible since (4.20) and the fact that Qy^Q_{\hat{y}}\in\mathcal{F} would then imply that yint(UQ)Ωy\in\text{int}\,(U_{Q^{\prime}})\subset\Omega_{*}, but we know that yΩy\in\partial\Omega_{*}. Thus, it holds that

(4.21) dist(Qy^,y)=|y^y|CτK0(Qy^).\displaystyle\operatorname{dist}(Q_{\hat{y}},y)=|\hat{y}-y|\leq C_{\tau}K_{0}\ell(Q_{\hat{y}}).

Recall that xQy^x_{Q_{\hat{y}}} is the center of Qy^Q_{\hat{y}} and XQy^X_{Q_{\hat{y}}}^{*} lies on a line segment from a corkscrew point XQy^X_{Q_{\hat{y}}} to its touching point x^Qy^\hat{x}_{Q_{\hat{y}}}. By Lemma 2.26, we know that x^Qy^Qy^\hat{x}_{Q_{\hat{y}}}\in Q_{\hat{y}}. Thus, (4.21), the definitions of the points and the fact that y^,x^Qy^Qy^\hat{y},\hat{x}_{Q_{\hat{y}}}\in Q_{\hat{y}} give us

|yXQy^||yy^|+|y^x^Qy^|+|x^Qy^XQy^|K0(Qy^)+diam(Qy^)+rQy^rQy^.\displaystyle|y-X^{*}_{Q_{\hat{y}}}|\leq|y-\hat{y}|+|\hat{y}-\hat{x}_{Q_{\hat{y}}}|+|\hat{x}_{Q_{\hat{y}}}-X^{*}_{Q_{\hat{y}}}|\lesssim K_{0}\ell(Q_{\hat{y}})+\operatorname{diam}(Q_{\hat{y}})+r_{Q_{\hat{y}}}\approx r_{Q_{\hat{y}}}.

In particular, there exists M1M\geq 1 such that

yMΔQy^QMΔQ,\displaystyle y\in M\Delta^{*}_{Q_{\hat{y}}}\subseteq\bigcup_{Q\in\mathcal{F}}M\Delta^{*}_{Q},

which is what we wanted. ∎

Let us then fix M1M\geq 1 as in Lemma 4.18. By (4.15), it holds that

(4.22) ω(QMΔQ)c.\displaystyle\omega_{*}\Big{(}\bigcup_{Q\in\mathcal{F}}M\Delta^{*}_{Q}\Big{)}\geq c_{*}.

Our next goal is to analyze how much the cubes QOQ\in\mathcal{F}^{\operatorname{O}} contribute to (4.22) and then limit this contribution by choosing λ\lambda in a suitable way. For this, we prove the following bound:

Lemma 4.23.

For any QQ\in\mathcal{F}, we have

(4.24) ω(ΔQ)σ(Q)σ(Q(𝐒))\displaystyle\omega_{*}(\Delta^{*}_{Q})\lesssim\frac{\sigma(Q)}{\sigma(Q({\bf S}_{*}))}

for an implicit constant depending only on structural constants, ellipticity, and the ωL,ΩA(σ)\omega_{L,\Omega}\in A_{\infty}(\sigma) constants (and independent of λ\lambda, ε\varepsilon, NN, jj, and 𝒮\mathcal{S}).

Proof.

Let QQ\in\mathcal{F}. Recall that ΔQ=B(XQ,λθ02rQ)Ω\Delta^{*}_{Q}=B(X^{*}_{Q},\tfrac{\lambda\theta_{0}}{2}r_{Q})\cap\partial\Omega_{*} is a surface ball on Ω\partial\Omega_{*} with XQΩX^{*}_{Q}\in\partial\Omega_{*}. Since Ω\Omega_{*} is a uniform domain, it satisfies the corkscrew condition. Let X~Q\widetilde{X}_{Q} be a corkscrew point in Ω\Omega_{*} relative to XQX^{*}_{Q} at scale approximately rQ=λθ02rQrQr_{Q}^{\prime}=\frac{\lambda\theta_{0}}{2}r_{Q}\approx r_{Q}. By perhaps insisting that θ0\theta_{0} is smaller we have that this corkscrew point X~Q\widetilde{X}_{Q} is far from XX_{*}. Then by connecting X~Q\widetilde{X}_{Q} to XQX_{Q} with a Harnack chain666Here we use that dist(XQ,Ω)dist(XQ,Q)(Q)rQ\operatorname{dist}(X^{*}_{Q},\partial\Omega)\approx\operatorname{dist}(X^{*}_{Q},Q)\approx\ell(Q)\approx r_{Q}; with very crude bounds this can be seen from (4.16). Then we also use that ΩΩ\Omega_{*}\subset\Omega, so that (Q)rQdist(X~Q,Ω)dist(X~Q,Ω)\ell(Q)\approx r_{Q}^{\prime}\approx\operatorname{dist}(\widetilde{X}_{Q},\partial\Omega_{*})\leq\operatorname{dist}(\widetilde{X}_{Q},\partial\Omega). (of uniformly bounded length) in Ω\Omega and using Lemma 2.38,

GL,Ω(X,X~Q)GL,Ω(X,XQ)(Q)ωL,ΩX(Q)(Q)n.G_{L,\Omega}(X_{*},\widetilde{X}_{Q})\approx G_{L,\Omega}(X_{*},X_{Q})\lesssim\ell(Q)\frac{\omega^{X_{*}}_{L,\Omega}(Q)}{\ell(Q)^{n}}.

Now, by 3.16 we have that GL,Ω(X,X~Q)GL,Ω(X,X~Q)G_{L,\Omega_{*}}(X_{*},\widetilde{X}_{Q})\leq G_{L,\Omega}(X_{*},\widetilde{X}_{Q}), and then by Lemma 2.38 in Ω\Omega_{*} we conclude that

ω(ΔQ)(rQ)nrQ=ωL,ΩX(ΔQ)(rQ)nrQGL,Ω(X,X~Q)GL,Ω(X,X~Q).\frac{\omega_{*}(\Delta^{*}_{Q})}{(r^{\prime}_{Q})^{n}}r^{\prime}_{Q}=\frac{\omega^{X_{*}}_{L,\Omega_{*}}(\Delta^{*}_{Q})}{(r^{\prime}_{Q})^{n}}r^{\prime}_{Q}\lesssim G_{L,\Omega_{*}}(X_{*},\widetilde{X}_{Q})\leq G_{L,\Omega}(X_{*},\widetilde{X}_{Q}).

Combining the two previously displayed inequalities and using that (Q)rQ\ell(Q)\approx r^{\prime}_{Q} we have ω(ΔQ)ωL,ΩX(Q)\omega_{*}(\Delta^{*}_{Q})\lesssim\omega^{X_{*}}_{L,\Omega}(Q). By Lemma 2.40 and Lemma 3.18 applied twice for both Q,Q(𝐒)𝒮Q,Q({\bf S}_{*})\in\mathcal{S}, it holds that

ωL,ΩX(Q)=ωL,ΩXQ(𝐒)(Q)ωL,ΩXQ(𝒮)(Q)ωL,ΩXQ(𝒮)(Q(𝐒))σ(Q)σ(Q(𝒮))σ(Q(𝒮))σ(Q(𝐒))=σ(Q)σ(Q(𝐒)),\omega^{X_{*}}_{L,\Omega}(Q)=\omega^{X_{Q({\bf S}_{*})}}_{L,\Omega}(Q)\approx\frac{\omega_{L,\Omega}^{X_{Q(\mathcal{S})}}(Q)}{\omega_{L,\Omega}^{X_{Q(\mathcal{S})}}(Q({\bf S}_{*}))}\approx\frac{\sigma(Q)}{\sigma(Q(\mathcal{S}))}\frac{\sigma(Q(\mathcal{S}))}{\sigma(Q({\bf S}_{*}))}=\frac{\sigma(Q)}{\sigma(Q({\bf S}_{*}))},

which ends the proof of (4.24). ∎

Now we are ready to conclude the proof of Lemma 4.12. Using (4.24) and the doubling property of ω\omega_{*} we find that

(4.25) QOω(MΔQ)CQOω(ΔQ)CQOσ(Q)σ(Q(𝐒))C^λ,\sum_{Q\in\mathcal{F}^{\operatorname{O}}}\omega_{*}\left(M\Delta^{*}_{Q}\right)\leq C\sum_{Q\in\mathcal{F}^{\operatorname{O}}}\omega_{*}\left(\Delta^{*}_{Q}\right)\leq C\sum_{Q\in\mathcal{F}^{\operatorname{O}}}\frac{\sigma(Q)}{\sigma(Q({\bf S}_{*}))}\leq\hat{C}\lambda,

where we used (4.8) in the last inequality. Now we choose λ>0\lambda>0 so that C^λ<c/2\hat{C}\lambda<c_{*}/2 and use (4.25) and (4.22) to deduce

(4.26) ω(QSBMΔQ)c/2,\omega_{*}\Big{(}\bigcup_{Q\in\mathcal{F}^{\operatorname{SB}}}M\Delta^{*}_{Q}\Big{)}\geq c_{*}/2,

where we used =SBO\mathcal{F}=\mathcal{F}^{\operatorname{SB}}\cup\mathcal{F}^{O}. Now we use the 5R5R-covering lemma [Mat95] to produce a countable collection of disjoint surface balls {MΔk}{MΔQk}\{M\Delta^{*}_{k}\}\coloneqq\{M\Delta^{*}_{Q_{k}}\} where each QkQ_{k} is in SB\mathcal{F}^{\operatorname{SB}} and such that

QSBMΔQk5MΔk.\bigcup_{Q\in\mathcal{F}^{\operatorname{SB}}}M\Delta^{*}_{Q}\subset\cup_{k}5M\Delta^{*}_{k}.

Then using (4.26) and the doubling property of ω\omega_{*} it holds

ω(kΔk)=kω(Δk)kω(5MΔk)ω(QSBMΔQ)c/2,\omega_{*}(\cup_{k}\Delta^{*}_{k})=\sum_{k}\omega_{*}(\Delta^{*}_{k})\gtrsim\sum_{k}\omega_{*}(5M\Delta^{*}_{k})\gtrsim\omega_{*}\Big{(}\bigcup_{Q\in\mathcal{F}^{\operatorname{SB}}}M\Delta^{*}_{Q}\Big{)}\geq c_{*}/2,

where we used that Δk\Delta^{*}_{k} are disjoint. To summarize we have produced a sequence of surface balls Δk=ΔQk\Delta^{*}_{k}=\Delta^{*}_{Q_{k}} with QkSBQ_{k}\in\mathcal{F}^{\operatorname{SB}} such that

(4.27) ω(kΔk)c,\omega_{*}(\cup_{k}\Delta^{*}_{k})\geq c_{**},

where cc_{**} depends on dimension, ellipticity, the Ahlfors regularity constant for Ω\partial\Omega, the corkscrew and Harnack Chain constants for Ω\Omega, and the ωLA(σ)\omega_{L}\in A_{\infty}(\sigma) constants. Thus, using (4.17) we have

Ω(u(y)u(X))2𝑑ω(y)ω(kΔk)infykΔk(u(y)u(X))2c(ε/200)2c4ε2,\int_{\partial\Omega_{*}}(u(y)-u(X_{*}))^{2}\,d\omega_{*}(y)\geq\omega_{*}(\cup_{k}\Delta^{*}_{k})\inf\limits_{y\in\cup_{k}\Delta^{*}_{k}}(u(y)-u(X_{*}))^{2}\\ \geq c_{**}(\varepsilon/200)^{2}\eqqcolon c_{4}\varepsilon^{2},

which proves Lemma 4.12.

As we had reduced the proof of the packing of the Type 4 maximal cubes to Lemma 4.12, this completes the proof of Lemma 4.6.

4.4. Construction of ε\varepsilon-approximators

With the help of the previous constructions and estimates, we can prove the existence of BVloc\operatorname{BV}_{\operatorname{loc}} ε\varepsilon-approximators in a similar way as in [HMM16]. For the convenience of the reader, we recall the key steps of the construction below. For some of the details, we follow the construction of LpL^{p}-type approximators in [HT20] which are an adaptation of the arguments in [HMM16]. Recall that we denote the collection of blue cubes in the disjoint stopping time regimes 𝒮\mathcal{S} in Lemma 3.18 by =(𝒮)\mathcal{L}=\mathcal{L}(\mathcal{S}), and each of these collections has a decomposition =j𝐒j\mathcal{L}=\cup_{j}{\bf S}_{j}.

Proof of Theorem 4.1.

Let us fix a dyadic cube Q0𝔻(Ω)Q_{0}\in\mathbb{D}(\partial\Omega) and construct an ε\varepsilon-approximator ΦQ0=ΦQ0ε\Phi_{Q_{0}}=\Phi_{Q_{0}}^{\varepsilon} first in the Carleson box TQ0T_{Q_{0}}. We start by dividing the Carleson box TQ0T_{Q_{0}} into a few types of different regions where we define the approximator differently. Let us choose the largest good (in the sense of the corona decomposition from Lemma 3.18), blue subcube Q1Q0Q_{1}\subset Q_{0} which may be Q0Q_{0} itself; if there are several such cubes with the largest side length, we choose just one of them. Since Q1Q_{1} is a good blue cube, there exists a stopping time regime 𝒮Q1\mathcal{S}_{Q_{1}} in Lemma 3.18 and subregime 𝐒Q1(𝒮Q1){\bf S}_{Q_{1}}\subset\mathcal{L}(\mathcal{S}_{Q_{1}}) such that Q1Q_{1} is the maximal element of the regime 𝐒1𝐒Q1𝔻Q0{\bf S}^{1}\coloneqq{\bf S}_{Q_{1}}\cap\mathbb{D}_{Q_{0}}. We then choose the largest good blue cube Q2Q_{2} from the collection 𝔻Q0𝐒1\mathbb{D}_{Q_{0}}\setminus{\bf S}^{1}. Similarly, Q2Q_{2} is the maximal element of the regime 𝐒2𝐒Q2𝔻Q0{\bf S}^{2}\coloneqq{\bf S}_{Q_{2}}\cap\mathbb{D}_{Q_{0}}. We then choose the largest good blue cube Q3𝔻Q0(𝐒1𝐒2)Q_{3}\in\mathbb{D}_{Q_{0}}\setminus({\bf S}^{1}\cup{\bf S}^{2}), and continue like this. This gives us a sequence of good blue cubes Q1,Q2,Q_{1},Q_{2},\ldots such that (Q1)(Q2)\ell(Q_{1})\geq\ell(Q_{2})\geq\ldots, each cube QiQ_{i} is a maximal element of a regime 𝐒i{\bf S}^{i} and the collection i𝐒i\cup_{i}{\bf S}^{i} contains all the good blue cubes in 𝔻Q0\mathbb{D}_{Q_{0}}. The cubes QiQ_{i} are “mostly” of the form Q(𝐒)Q({\bf S}) as in the decomposition of the collections (𝒮)\mathcal{L}(\mathcal{S}) earlier in the sense that there exists a collection of pairwise disjoint cubes {Pk}k𝔻Q0\{P_{k}\}_{k}\subset\mathbb{D}_{Q_{0}} (that may be empty) such that every cube in the collection {Qi}i{Pk}k\{Q_{i}\}_{i}\setminus\{P_{k}\}_{k} is of the form of Q(𝐒)Q({\bf S}) for some 𝐒{\bf S}. This is because the cube Q0Q_{0} is arbitrary and hence, it may be a bad cube or a red cube. For each ii, we define the “bottom” cubes of 𝐒i{\bf S}^{i} in the obvious way: we set 𝐒i𝐒Qi\mathcal{F}_{{\bf S}^{i}}\coloneqq\mathcal{F}_{{\bf S}_{Q_{i}}}, that is, 𝐒i\mathcal{F}_{{\bf S}^{i}} is the collection of the stopping cubes associated to the unique regime 𝐒Qi{\bf S}_{Q_{i}} that contains QiQ_{i}.

For each ii, we define the regions AiA_{i} recursively the following way:

A1Ω𝐒1,Q1,AiΩ𝐒i,Qik=1i1Ak for i2.\displaystyle A_{1}\coloneqq\Omega_{\mathcal{F}_{{\bf S}^{1}},Q_{1}},\qquad A_{i}\coloneqq\Omega_{\mathcal{F}_{{\bf S}^{i},Q_{i}}}\setminus\bigcup_{k=1}^{i-1}A_{k}\quad\text{ for }i\geq 2.

By construction, the regions AiA_{i} are pairwise disjoint. We also set Ω0iAi\Omega_{0}\coloneqq\bigcup_{i}A_{i}, and we define the function Φ0\Phi_{0} on Ω0\Omega_{0} as

Φ0iu(XQ(𝐒Qi))𝟙Ai,\displaystyle\Phi_{0}\coloneqq\sum_{i}u(X_{Q({\bf S}_{Q_{i}})})\mathbbm{1}_{A_{i}},

where XQ(𝐒Qi)X_{Q({\bf S}_{Q_{i}})} is the corkscrew point we used in the stopping conditions in the definition of 𝐒Qi{\bf S}_{Q_{i}}. In particular, for any XAiX\in A_{i} we have |u(X)u(XQ(𝐒Qi))|ε/100<ε|u(X)-u(X_{Q({\bf S}_{Q_{i}})})|\leq\varepsilon/100<\varepsilon. Furthermore, by the disjointness of the regions AiA_{i}, we have uΦ0L(Ω0)<ε\|u-\Phi_{0}\|_{L^{\infty}(\Omega_{0})}<\varepsilon.

Let us then consider the cubes in 𝔻Q0i𝐒i\mathbb{D}_{Q_{0}}\setminus\cup_{i}{\bf S}^{i}. Let us fix some enumeration {Rj}j\{R_{j}\}_{j} for the cubes 𝔻Q0i𝐒i\mathbb{D}_{Q_{0}}\setminus\cup_{i}{\bf S}^{i}. The cubes RjR_{j} are red cubes or bad blue cubes. For each jj, we define the regions VjV_{j} recursively the following way:

V1UR1,VjURjk=1j1Vk for j2.\displaystyle V_{1}\coloneqq U_{R_{1}},\qquad V_{j}\coloneqq U_{R_{j}}\setminus\bigcup_{k=1}^{j-1}V_{k}\quad\text{ for }j\geq 2.

By construction, the regions VjV_{j} are pairwise disjoint. We also set Ω1jVj\Omega_{1}\coloneqq\bigcup_{j}V_{j}, and we define the function Φ1\Phi_{1} on Ω1\Omega_{1} as

Φ1(X){u(X), if XVk for a red cube Rk,u(Xk), if XVk for a blue cube Rk,\displaystyle\Phi_{1}(X)\coloneqq\left\{\begin{array}[]{cl}u(X),&\text{ if }X\in V_{k}\text{ for a red cube }R_{k},\\ u(X_{k}),&\text{ if }X\in V_{k}\text{ for a blue cube }R_{k},\end{array}\right.

where XkX_{k} is any fixed point on URkU_{R_{k}}. By the definitions, we have uΦ1L(Ω1)<ε/1000<ε\|u-\Phi_{1}\|_{L^{\infty}(\Omega_{1})}<\varepsilon/1000<\varepsilon.

We define the ε\varepsilon-approximator ΦQ0\Phi_{Q_{0}} of uu in the Carleson box TQ0T_{Q_{0}} as

ΦQ0(X){Φ0(X), if XΩ0,Φ1(X), if XTQ0Ω0.\displaystyle\Phi_{Q_{0}}(X)\coloneqq\left\{\begin{array}[]{cl}\Phi_{0}(X),&\text{ if }X\in\Omega_{0},\\ \Phi_{1}(X),&\text{ if }X\in T_{Q_{0}}\setminus\Omega_{0}.\end{array}\right.

By the construction, we have uΦQ0L(TQ0)<ε\|u-\Phi_{Q_{0}}\|_{L^{\infty}(T_{Q_{0}})}<\varepsilon. The L1L^{1}-type Carleson measure estimate for ΦQ0\Phi_{Q_{0}} in TQ0T_{Q_{0}} can be proven as in [HMM16] with small but quite obvious changes. Using a covering argument, the claim can be reduced to proving the estimate on Carleson boxes TQT_{Q^{\prime}}, and since uL(Ω)u\in L^{\infty}(\Omega), the core challenge is to handle the jumps across the boundaries of the sets AiA_{i} and VjV_{j} that contribute to the total variation of ΦQ0\Phi_{Q_{0}} inside TQ0T_{Q_{0}}. Since the boundaries of the sawtooth regions, Whitney regions and Carleson boxes are Ahlfors regular by Lemma 2.18, the estimates reduce to using the Carleson packing conditions in Lemma 3.18, Lemma 4.2, Lemma 4.5 and Lemma 4.6. The Carleson norm of the measure μΦQ0\mu_{\Phi_{Q_{0}}} such that dμΦQ0(Y)=|ΦQ0(Y)|dYd\mu_{\Phi_{Q_{0}}}(Y)=|\nabla\Phi_{Q_{0}}(Y)|\,dY is given (up to a structural constant) by the sizes of the Carleson packing norms in these results. We omit the details.

Using these kinds of local approximators, we build the global approximator of uu. If diam(Ω)<\operatorname{diam}(\Omega)<\infty, it is enough to build a local approximator for a Carleson box that covers the whole space Ω\Omega. Thus, we may assume that diam(Ω)=\operatorname{diam}(\Omega)=\infty. Suppose first that diam(Ω)<\operatorname{diam}(\partial\Omega)<\infty. Then there exists a dyadic cube Q0Q_{0} that covers the whole boundary Ω\partial\Omega. We build the local approximator ΦQ0\Phi_{Q_{0}} on TQ0T_{Q_{0}}, extend it to whole Ω\Omega by setting it to be 0 outside TQ0T_{Q_{0}} and define the global approximator as Φ=𝟙TQ0ΦQ0+𝟙ΩTQ0u\Phi=\mathbbm{1}_{T_{Q_{0}}}\Phi_{Q_{0}}+\mathbbm{1}_{\Omega\setminus T_{Q_{0}}}u. The L1L^{1}-type Carleson measure estimate follows from the same arguments as with the local approximators.

Finally, suppose that diam(Ω)=\operatorname{diam}(\partial\Omega)=\infty. Fix a sequence of dyadic cubes PkP_{k} such that P1P2P_{1}\subset P_{2}\subset\cdots, (P1)<(P2)<\ell(P_{1})<\ell(P_{2})<\cdots and Ω=kPk\partial\Omega=\cup_{k}P_{k}. This type of sequence of cubes does not exist in every dyadic system, but we can always construct a system where it exists (see, for example, [HT14]). We build a local approximator ΦPk\Phi_{P_{k}} in TPkT_{P_{k}} for every kk and extend the approximators to whole Ω\Omega by setting each of them to be 0 outside TPkT_{P_{k}}. We then define the global approximator as Φ=𝟙TP1ΦP1+k=2𝟙TPkTPk1ΦPk\Phi=\mathbbm{1}_{T_{P_{1}}}\Phi_{P_{1}}+\sum_{k=2}^{\infty}\mathbbm{1}_{T_{P_{k}}\setminus T_{P_{k-1}}}\Phi_{P_{k}}. The L1L^{1}-type Carleson measure estimate follows from the Carleson measure estimates of the local approximators and the fact that the collection {Pk}k\{P_{k}\}_{k} satisfies a Carleson packing condition with a uniformly bounded Carleson packing norm depending only on structural constants. Again, we omit the details. ∎

To finish the proof of Theorem 1.1, we regularize the approximators in Theorem 4.1. This regularization makes the constant CεC_{\varepsilon} significantly larger but since the size of this constant is not important for our results, we do not track its size.

Lemma 4.28.

Let ε(0,1)\varepsilon\in(0,1). There exists a unifomly bounded constant C~ε1\widetilde{C}_{\varepsilon}\geq 1 such that we can choose the ε\varepsilon-approximator Φ=Φε\Phi=\Phi^{\varepsilon} for the solution uW1,2(Ω)L(Ω)u\in W^{1,2}(\Omega)\cap L^{\infty}(\Omega) to Lu=0Lu=0 in Theorem 4.1 so that

  1. i)

    uΦL(Ω)2εuL(Ω)\|u-\Phi\|_{L^{\infty}(\Omega)}\leq 2\varepsilon\|u\|_{L^{\infty}(\Omega)},

  2. ii)

    supxΩ,r>01rnB(x,r)Ω|Φ(Y)|𝑑YC~εuL(Ω),\sup_{x\in\partial\Omega,r>0}\frac{1}{r^{n}}\iint_{B(x,r)\cap\Omega}|\nabla\Phi(Y)|\,dY\leq\widetilde{C}_{\varepsilon}\|u\|_{L^{\infty}(\Omega)},

  3. iii)

    ΦC(Ω)\Phi\in C^{\infty}(\Omega),

  4. iv)

    |Φ(Y)|C~εδ(Y)|\nabla\Phi(Y)|\leq\tfrac{\widetilde{C}_{\varepsilon}}{\delta(Y)} for every YΩY\in\Omega,

  5. v)

    if |XY|δ(X)|X-Y|\ll\delta(X), then |Φ(X)Φ(Y)|C~ε|XY|δ(X)|\Phi(X)-\Phi(Y)|\leq\tfrac{\widetilde{C}_{\varepsilon}|X-Y|}{\delta(X)},

  6. vi)

    there exists a function φL(Ω)\varphi\in L^{\infty}(\partial\Omega) such that

    limYx,n.t.Φ(Y)=φ(x) for σ-a.e. xΩ.\displaystyle\lim_{Y\to x,\,\operatorname{n.t.}}\Phi(Y)=\varphi(x)\text{ for }\sigma\text{-a.e. }x\in\partial\Omega.

The constant C~ε\widetilde{C}_{\varepsilon} depends on ε\varepsilon, the structural constants of Ω\Omega, the constant CεC_{\varepsilon} in Theorem 4.1 and the Hölder continuity constants CC and α\alpha in Lemma 2.28.

Proof.

The proof uses tweaked mollifier techniques combined with a regularized distance function. The properties follow mostly from [HT21, Section 3] but for the convenience of the reader, we define the core objects and give some explicit details below.

Let β\beta be a regularized version of the distance function δ=dist(,Ω)\delta=\operatorname{dist}(\cdot,\partial\Omega), that is, a smooth function in Ω\Omega such that βδ\beta\approx\delta (see [Ste70, Theorem 2, p. 171]). Let ζ0\zeta\geq 0 be a smooth non-negative function supported on B(0,1m)B(0,\tfrac{1}{m}) for a suitable constant m>0m>0 (depending on the implicit constants in δβ\delta\approx\beta), satisfying ζ1\zeta\leq 1 and ζ=1\int\zeta=1. For a constant ξε>0\xi_{\varepsilon}>0 to be chosen momentarily, we set

Λξε(X,Y)ζξεβ(X)(XY)=1(ξεβ(X))n+1ζ(XYξεβ(X)).\displaystyle\Lambda_{\xi_{\varepsilon}}(X,Y)\coloneqq\zeta_{\xi_{\varepsilon}\beta(X)}(X-Y)=\frac{1}{(\xi_{\varepsilon}\beta(X))^{n+1}}\zeta\Big{(}\frac{X-Y}{\xi_{\varepsilon}\beta(X)}\Big{)}.

For a suitable choice of mm, we have suppΛ(X,)B(X,ξεδ(X)/2)\text{supp}\,\Lambda(X,\cdot)\subset B(X,\xi_{\varepsilon}\delta(X)/2). Given the non-smooth ε\varepsilon-approximator Φ0\Phi_{0} of the solution uWloc1,2(Ω)L(Ω)u\in W^{1,2}_{\operatorname{loc}}(\Omega)\cap L^{\infty}(\Omega) to Lu=0Lu=0 in Theorem 4.1, we set

Φ(X)Λε(X,Y)Φ0(Y)𝑑Y.\displaystyle\Phi(X)\coloneqq\iint\Lambda_{\varepsilon}(X,Y)\Phi_{0}(Y)\,dY.

The property iii) follows from a standard modification of the case Ω=n+1\Omega=\mathbb{R}^{n+1} (for example, see [EG15, Theorem 1, p. 123]) and properties ii), iv) and v) are formulated explicitly in [HT21, Section 3]. Property i) follows from the local Hölder continuity of uu (that is, Lemma 2.28) and the fact that Φ0\Phi_{0} is an ε\varepsilon-approximator of uu: for almost every XΩX\in\Omega, we get

|Φ(X)u(X)|\displaystyle|\Phi(X)-u(X)| =|Λε(X,Y)(Φ0(Y)u(X))𝑑Y|\displaystyle=\Big{|}\iint\Lambda_{\varepsilon}(X,Y)(\Phi_{0}(Y)-u(X))\,dY\Big{|}
<εuL(Ω)+Λε(X,Y)|u(Y)u(X)|𝑑Y\displaystyle<\varepsilon\|u\|_{L^{\infty}(\Omega)}+\iint\Lambda_{\varepsilon}(X,Y)\left|u(Y)-u(X)\right|\,dY
εuL(Ω)+Λε(X,Y)C(|XY|14δ(X))α(B(X,12δ(X))|u(Z)|2𝑑Z)1/2𝑑Y\displaystyle\leq\varepsilon\|u\|_{L^{\infty}(\Omega)}+\iint\Lambda_{\varepsilon}(X,Y)\,C\Big{(}\frac{|X-Y|}{\tfrac{1}{4}\delta(X)}\Big{)}^{\alpha}\Big{(}\fint_{B(X,\tfrac{1}{2}\delta(X))}|u(Z)|^{2}\,dZ\Big{)}^{1/2}\,dY
εuL(Ω)+C(2ξε)αuL(Ω)Λε(X,Y)𝑑Y2εuL(Ω)\displaystyle\leq\varepsilon\|u\|_{L^{\infty}(\Omega)}+C\left(2\xi_{\varepsilon}\right)^{\alpha}\|u\|_{L^{\infty}(\Omega)}\iint\Lambda_{\varepsilon}(X,Y)\,dY\leq 2\varepsilon\|u\|_{L^{\infty}(\Omega)}

as long as we choose ξε12(εC)1α\xi_{\varepsilon}\leq\tfrac{1}{2}\big{(}\tfrac{\varepsilon}{C}\big{)}^{\frac{1}{\alpha}}, where CC and α\alpha are the Hölder continuity constants in Lemma 2.28.

Property vi) follows from the same argument that is used in the proof of [HT21, Lemma 4.14] after some small additional considerations. The proof of [HT21, Lemma 4.14] is based on showing that almost every cone on a codimension 11 uniformly rectifiable set has locally exactly two components, these local components satisfy the Harnack chain condition and the Harnack chain condition combined with the L1L^{1}-type Carleson measure estimate ii) implies the existence of the a.e. non-tangential trace φ\varphi. We do not assume that Ω\partial\Omega is uniformly rectifiable but by the definition of dyadic cones (2.17) and Lemma 2.18 we know that any truncated cone on Ω\partial\Omega has exactly one component inside Ω\Omega and this component satisfies the Harnack chain condition. Thus, the argument in the proof of [HT21, Lemma 4.14] works also for us. In particular, we can choose Φ\Phi in such a way that all the properties i) – vi) hold. This completes the proofs of Lemma 4.28 and Theorem 1.1. ∎

5. Proof of Theorem 1.2

In this section, we prove Theorem 1.2, that is, we prove that ε\varepsilon-approximability of solutions uu to Lu=0Lu=0 implies that ωLA(σ)\omega_{L}\in A_{\infty}(\sigma). To be more precise, we prove the following seemingly stronger result:

Theorem 5.1.

Let Ωn+1\Omega\subset\mathbb{R}^{n+1}, n1n\geq 1, be a uniform domain with Ahflors regular boundary, and let LL be a divergence form elliptic operator L=divAL=-\mathop{\operatorname{div}}\nolimits A\nabla in Ω\Omega. Suppose also that for every bounded Borel set SΩS\subset\partial\Omega the solution u=uSu=u_{S} to Lu=0Lu=0 such that u(X)=ωLX(S)u(X)=\omega_{L}^{X}(S) is ε\varepsilon-approximable for every ε(0,1)\varepsilon\in(0,1) in the sense of Theorem 1.1 with the ε\varepsilon-approximability constants depending only on structural constants and ε\varepsilon. Then ωLA(σ)\omega_{L}\in A_{\infty}(\sigma).

In particular, by Theorem 1.1 and Theorem 5.1, ε\varepsilon-approximability of the subclass of solutions uu to Lu=0Lu=0 in Theorem 5.1 is equivalent with ε\varepsilon-approximability of all solutions uu to Lu=0Lu=0 (and hence, it is equivalent with the other conditions in Corollary 1.3).

The proof of Theorem 5.1 is based on the proof of [CHMT20, Theorem 1.1], which itself is based on the techniques used in [KKPT00] and [KKPT16]. The key idea is the following. We fix a cube Q0Q_{0} and a Borel set FQ0F\subset Q_{0} and we build a suitable solution u=uFu=u_{F} associated to FF such that uu oscillates a significant amount. We then use this oscillation to control the L1L^{1} norm of the gradient of an ε\varepsilon-approximator of uu from below near Q0Q_{0} for a very small ε\varepsilon. The L1L^{1}-type Carleson measure estimate of the ε\varepsilon-approximator then allows us to verify the ωLA(σ)\omega_{L}\in A_{\infty}(\sigma) condition.

For the proof of Theorem 5.1, we need some definitions and notation from [CHMT20]. For clarity, we adopt most of the notation as it is from [CHMT20]. We define the following fattened version of the Whitney regions UQU_{Q} and a “wider” version of the truncated dyadic cone:

UQ,η3Q𝔻Q(Q)>η3(Q)UQ, and ΓQ0η(x)Q𝔻Q0QxUQ,η3.\displaystyle U_{Q,\eta^{3}}\coloneqq\bigcup_{\begin{subarray}{c}Q^{\prime}\in\mathbb{D}_{Q}\\ \ell(Q^{\prime})>\eta^{3}\ell(Q)\end{subarray}}U_{Q^{\prime}},\qquad\text{ and }\qquad\Gamma_{Q_{0}}^{\eta}(x)\coloneqq\bigcup_{\begin{subarray}{c}Q\in\mathbb{D}_{Q_{0}}\\ Q\ni x\end{subarray}}U_{Q,\eta^{3}}.

The main difference between our approach and the proof of the implication “CME \implies AA_{\infty}” in [CHMT20] is the following lemma which is a modification of [CHMT20, Lemma 3.10]:

Lemma 5.2.

There exist ε(0,1)\varepsilon\in(0,1), 0<η10<\eta\ll 1, depending only on structural constants, and α0(0,1)\alpha_{0}\in(0,1), Cη1C_{\eta}\geq 1, both depending on structural constants and on η\eta, such that for each Q0𝔻Q_{0}\in\mathbb{D}, for every α(0,α0)\alpha\in(0,\alpha_{0}), and for every FQ0F\subset Q_{0} satisfying ωLXQ0(F)αωLXQ0(Q0)\omega_{L}^{X_{Q_{0}}}(F)\leq\alpha\omega_{L}^{X_{Q_{0}}}(Q_{0}), there exists a Borel set SQ0S\subset Q_{0} such that if u(X)=ωLX(S)u(X)=\omega_{L}^{X}(S) and Φ=Φε\Phi=\Phi^{\varepsilon} is an ε\varepsilon-approximator of uu, then

ΓQ0η(y)|Φ(Y)|δ(Y)n𝑑YCη1log(α1),for each yF.\displaystyle\iint_{\Gamma_{Q_{0}}^{\eta}(y)}|\nabla\Phi(Y)|\delta(Y)^{-n}\,dY\geq C_{\eta}^{-1}\log(\alpha^{-1}),\qquad\text{for each }y\in F.

Before proving Lemma 5.2, let us see how it gives us Theorem 5.1.

Proof of Theorem 5.1.

Following [CHMT20]777Although the results of [CHMT20] are stated only for n2n\geq 2, the proof of “(a) \implies (b)” in [CHMT20, Theorem 1.1] works for a large part also for n=1n=1; see [FP22]., we show that for each β(0,1)\beta\in(0,1), there exists α(0,1)\alpha\in(0,1) such that for every Q0𝔻Q_{0}\in\mathbb{D} and every Borel set FQ0F\subset Q_{0}, we have that

(5.3) ωLXQ0(F)ωLXQ0(Q0)ασ(F)σ(Q0)β,\frac{\omega_{L}^{X_{Q_{0}}}(F)}{\omega_{L}^{X_{Q_{0}}}(Q_{0})}\leq\alpha\quad\implies\quad\frac{\sigma(F)}{\sigma(Q_{0})}\leq\beta,

where the constants α\alpha and β\beta are independent of the choice of the dyadic system 𝔻\mathbb{D}. This is a dyadic version of the A(σ)A_{\infty}(\sigma) condition. Although it looks different than Definition 2.31, the conditions are equivalent since we consider doubling measures (see, for example, [GR85, Chapter IV, Theorem 2.11]) and the constants are independent of the system 𝔻\mathbb{D} (see [CHMT20, pp. 16] or use adjacent dyadic techniques [HK12, HT14]).

Fix β(0,1)\beta\in(0,1) and Q0𝔻Q_{0}\in\mathbb{D}. Moreover, fix η(0,1)\eta\in(0,1) small enough, and constants ε\varepsilon, α0\alpha_{0}, and CηC_{\eta} as in Lemma 5.2. Let FQ0F\subset Q_{0} satisfy ωLX0(F)αωLX0(Q0)\omega_{L}^{X_{0}}(F)\leq\alpha\omega_{L}^{X_{0}}(Q_{0}) with α(0,α0)\alpha\in(0,\alpha_{0}). We now use Lemma 5.2 to see that there exists SQ0S\subset Q_{0} such that if Φε\Phi^{\varepsilon} is an ε\varepsilon-approximator of u(X)=ωLX(S)u(X)=\omega_{L}^{X}(S), then

Cη1log(α1)σ(F)FΓQ0η(y)|Φε(Y)|δ(Y)n𝑑Y𝑑σ(y)Cη3nσ(Q0),C_{\eta}^{-1}\log(\alpha^{-1})\sigma(F)\leq\int_{F}\iint_{\Gamma_{Q_{0}}^{\eta}(y)}|\nabla\Phi^{\varepsilon}(Y)|\delta(Y)^{-n}\,dY\,d\sigma(y)\leq C\eta^{-3n}\sigma(Q_{0}),

where the last estimate follows by Fubini’s theorem and property ii) of ε\varepsilon-approximability in Theorem 1.1 (see [CHMT20, pp. 15–16] for more details). We may then choose α\alpha small enough depending on β\beta so that (5.3) holds. ∎

We turn to the proof of Lemma 5.2. For this, we need the following machinery:

Definition 5.4.

Let Q0𝔻Q_{0}\in\mathbb{D} be a dyadic cube, μ\mu a regular Borel measure on Q0Q_{0}, FQ0F\subset Q_{0} be a Borel set, ϵ0>0\epsilon_{0}>0, and kk\in\mathbb{N} be fixed. We say that a collection of nested Borel subsets {𝒪}=1k\{\mathcal{O}_{\ell}\}_{\ell=1}^{k} of Q0Q_{0} is a good ϵ0\epsilon_{0}-cover of FF of length kk for μ\mu if

  1. (a)

    F𝒪k𝒪k1𝒪2𝒪1Q0F\subset\mathcal{O}_{k}\subset\mathcal{O}_{k-1}\subset\cdots\subset\mathcal{O}_{2}\subset\mathcal{O}_{1}\subset Q_{0},

  2. (b)

    𝒪=iQi\mathcal{O}_{\ell}=\bigcup_{i}Q_{i}^{\ell} for disjoint subcubes Qi𝔻Q0Q_{i}^{\ell}\in\mathbb{D}_{Q_{0}},

  3. (c)

    μ(𝒪Qi1)ϵ0μ(Qi1)\mu(\mathcal{O}_{\ell}\cap Q_{i}^{\ell-1})\leq\epsilon_{0}\mu(Q_{i}^{\ell-1}) for every ii and every 2k2\leq\ell\leq k.

Lemma 5.5 ([CHMT20, Lemma 3.5]).

Fix Q0𝔻Q_{0}\in\mathbb{D} and suppose that ϵ0(0,1e)\epsilon_{0}\in(0,\tfrac{1}{e}). If FQ0F\subset Q_{0} is a Borel set such that ωLXQ0(F)αωLXQ0(Q0)\omega_{L}^{X_{Q_{0}}}(F)\leq\alpha\omega_{L}^{X_{Q_{0}}}(Q_{0}) for α(0,ϵ02/C)\alpha\in(0,\epsilon_{0}^{2}/C^{\prime}), then there exists a good ϵ0\epsilon_{0}-cover of FF of length klogα1logϵ01k\approx\tfrac{\log\alpha^{-1}}{\log\epsilon_{0}^{-1}} for ωLXQ0\omega_{L}^{X_{Q_{0}}}. Here CC^{\prime} depends only on the constant CC of Lemma 2.37.

Let Q0𝔻Q_{0}\in\mathbb{D} be a fixed cube and FQ0F\subset Q_{0} be a fixed subset such that ωLX0(F)αωLX0(Q0)\omega_{L}^{X_{0}}(F)\leq\alpha\omega_{L}^{X_{0}}(Q_{0}) for α\alpha small enough. For each Q𝔻Q\in\mathbb{D}, we let Q~𝔻\widetilde{Q}\in\mathbb{D} be the unique dyadic cube such that xQQ~x_{Q}\in\widetilde{Q} and (Q~)=η(Q)\ell(\widetilde{Q})=\eta\ell(Q) for η>0\eta>0 a small enough parameter to be determined later. Fix ϵ0>0\epsilon_{0}>0 small enough. Let {𝒪l}l=1k\{\mathcal{O}_{l}\}_{l=1}^{k} be a good ϵ0\epsilon_{0}-cover of FF given by Lemma 5.5. We set

𝒪~jiQ~il and Sj=2k𝒪~j1𝒪j\displaystyle\widetilde{\mathcal{O}}_{j}\coloneqq\bigcup_{i}\widetilde{Q}_{i}^{l}\quad\text{ and }\quad S\coloneqq\bigcup_{j=2}^{k}\widetilde{\mathcal{O}}_{j-1}\setminus\mathcal{O}_{j}

By the nestedness of the sets 𝒪j\mathcal{O}_{j}, we have 𝟙S=j=2k𝟙𝒪~j1𝒪j\mathbbm{1}_{S}=\sum_{j=2}^{k}\mathbbm{1}_{\widetilde{\mathcal{O}}_{j-1}\setminus\mathcal{O}_{j}}. We define the nonnegative solution uuF:Ωu\coloneqq u_{F}\colon\Omega\to\mathbb{R} to Lu=0Lu=0 as

u(X)=Ω𝟙S(y)𝑑ωLX(y)=ωLX(S)=j=2kωLX(𝒪~j1𝒪j).\displaystyle u(X)=\int_{\partial\Omega}\mathbbm{1}_{S}(y)\,d\omega_{L}^{X}(y)=\omega_{L}^{X}(S)=\sum_{j=2}^{k}\omega_{L}^{X}(\widetilde{\mathcal{O}}_{j-1}\setminus\mathcal{O}_{j}).

We have the following lower oscillation bound:

Lemma 5.6 ([CHMT20, Lemma 3.24]).

There exists a structural constant c0>0c_{0}>0 such that if η\eta and ϵ0=ϵ0(η,c0)\epsilon_{0}=\epsilon_{0}(\eta,c_{0}) are small enough, then for any yFy\in F and any 1lk11\leq l\leq k-1, there exist dyadic cubes QilQ_{i}^{l} and PilP_{i}^{l} such that

|u(X^Q~il)u(X^P~il)|c0,\displaystyle\big{|}u(\hat{X}_{\widetilde{Q}_{i}^{l}})-u(\hat{X}_{\widetilde{P}_{i}^{l}})\big{|}\geq c_{0},

where QilQ_{i}^{l} is the unique cube (in 𝒪l\mathcal{O}_{l}) such that yQily\in Q_{i}^{l}, Pil𝔻QilP_{i}^{l}\in\mathbb{D}_{Q_{i}^{l}} is the unique cube such that yPily\in P_{i}^{l} and (Pil)=η(Qil)\ell(P_{i}^{l})=\eta\ell(Q_{i}^{l}), Q~il\widetilde{Q}_{i}^{l} and P~il\widetilde{P}_{i}^{l} are defined as we did after Lemma 5.5 and X^Q~il\hat{X}_{\widetilde{Q}_{i}^{l}} and X^P~il\hat{X}_{\widetilde{P}_{i}^{l}} are corkscrew points relative to xQ~ilx_{\widetilde{Q}_{i}^{l}} at scale a1(Q~il)a_{1}\ell(\widetilde{Q}_{i}^{l}) and relative to xP~ilx_{\widetilde{P}_{i}^{l}} at scale a1(P~il)a_{1}\ell(\widetilde{P}_{i}^{l}), respectively.

With the help of Lemma 5.6, we can now prove Lemma 5.2:

Proof of Lemma 5.2.

The proof is a straightforward modification of the proofs of [CHMT20, Lemma 3.10] and [HT21, Lemma 4.14]. Let us fix yFy\in F and l{1,2,,k1}l\in\{1,2,\ldots,k-1\}. We borrow the notation from Lemma 5.6, and set ε=c0/4\varepsilon=c_{0}/4. By adjusting the construction parameters for the Whitney regions in Section 2.3, we may assume that there exist Whitney cubes IQ~ilI_{\widetilde{Q}_{i}^{l}} and IP~ilI_{\widetilde{P}_{i}^{l}} such that888In Section 2.3, we constructed the Whitney regions UQU_{Q} in such a way that we know that a corkscrew point relative to xQx_{Q} at scale 105a0(Q)10^{-5}a_{0}\ell(Q) belongs to UQU_{Q}. Corkscrew points of the type X^Q\hat{X}_{Q} are also relative to xQx_{Q} but they are at scale a1(Q)a_{1}\ell(Q).

(5.7) X^Q~ilIQ~il(1+τ)IQ~ilUQ~il and X^P~ilIP~il(1+τ)IP~ilUP~il.\displaystyle\hat{X}_{\widetilde{Q}_{i}^{l}}\in I_{\widetilde{Q}_{i}^{l}}\subset(1+\tau)I_{\widetilde{Q}_{i}^{l}}\subset U_{\widetilde{Q}_{i}^{l}}\quad\text{ and }\quad\hat{X}_{\widetilde{P}_{i}^{l}}\in I_{\widetilde{P}_{i}^{l}}\subset(1+\tau)I_{\widetilde{P}_{i}^{l}}\subset U_{\widetilde{P}_{i}^{l}}.

Since (Q~il)=η(Qil)\ell(\widetilde{Q}_{i}^{l})=\eta\ell(Q_{i}^{l}) and (P~il)=η(Pil)=η2(Qil)\ell(\widetilde{P}_{i}^{l})=\eta\ell(P_{i}^{l})=\eta^{2}\ell(Q_{i}^{l}). In particular, since η<1\eta<1, we have UQ~il,UP~ilUQil,η3U_{\widetilde{Q}_{i}^{l}},U_{\widetilde{P}_{i}^{l}}\subset U_{Q_{i}^{l},\eta^{3}}.

Since Φ\Phi is a c04\tfrac{c_{0}}{4}-approximator of uu, Lemma 5.6 gives us

c0|u(X^Q~il)u(X^P~il)||u(X^Q~il)Φ(X^Q~il)|+|Φ(X^Q~il)Φ(X^P~il)|+|Φ(X^P~il)u(X^P~il)|c02+|Φ(X^Q~il)Φ(X^P~il)|.c_{0}\leq|u(\hat{X}_{\widetilde{Q}_{i}^{l}})-u(\hat{X}_{\widetilde{P}_{i}^{l}})|\leq|u(\hat{X}_{\widetilde{Q}_{i}^{l}})-\Phi(\hat{X}_{\widetilde{Q}_{i}^{l}})|+|\Phi(\hat{X}_{\widetilde{Q}_{i}^{l}})-\Phi(\hat{X}_{\widetilde{P}_{i}^{l}})|+|\Phi(\hat{X}_{\widetilde{P}_{i}^{l}})-u(\hat{X}_{\widetilde{P}_{i}^{l}})|\\ \leq\frac{c_{0}}{2}+|\Phi(\hat{X}_{\widetilde{Q}_{i}^{l}})-\Phi(\hat{X}_{\widetilde{P}_{i}^{l}})|.

By (5.7), we know that the points X^Q~il\hat{X}_{\widetilde{Q}_{i}^{l}} and X^P~il\hat{X}_{\widetilde{P}_{i}^{l}} are well inside UQil,η3U_{Q_{i}^{l},\eta^{3}} in the sense that dist(X^P~il,UQil,η3)dist(X^Q~il,UQil,η3)(Qil)\operatorname{dist}(\hat{X}_{\widetilde{P}_{i}^{l}},\partial U_{Q_{i}^{l},\eta^{3}})\approx\operatorname{dist}(\hat{X}_{\widetilde{Q}_{i}^{l}},\partial U_{Q_{i}^{l},\eta^{3}})\approx\ell(Q_{i}^{l}). Since diam(UQil,η3)η(Qil)\operatorname{diam}(U_{Q_{i}^{l},\eta^{3}})\approx_{\eta}\ell(Q_{i}^{l}) and UQil,η3U_{Q_{i}^{l},\eta^{3}} satisfies the Harnack chain condition by Lemma 2.18, there exists a chain of N=N(η)N=N(\eta) balls B1,B2,,BNB_{1},B_{2},\ldots,B_{N} such that

  1. (i)

    X^Q~ilB1\hat{X}_{\widetilde{Q}_{i}^{l}}\in B_{1}, X^P~ilBN\hat{X}_{\widetilde{P}_{i}^{l}}\in B_{N} and BjBj+1ØB_{j}\cap B_{j+1}\neq\mbox{{\O}} for every j=1,2,,N1j=1,2,\ldots,N-1,

  2. (ii)

    the radii of the balls BjB_{j} are comparable to (Qil)\ell(Q_{i}^{l}), depending on η\eta and c0c_{0},

  3. (iii)

    |Φ(X^Q~il)Φ(X)|<c08|\Phi(\hat{X}_{\widetilde{Q}_{i}^{l}})-\Phi(X)|<\tfrac{c_{0}}{8} for every XB1X\in B_{1} and |Φ(X^P~il)Φ(Y)|<c08|\Phi(\hat{X}_{\widetilde{P}_{i}^{l}})-\Phi(Y)|<\tfrac{c_{0}}{8} for every YBNY\in B_{N},

  4. (iv)

    for each j=1,2,,N1j=1,2,\ldots,N-1, there exists a cylinder j\mathfrak{C}_{j} connecting BjB_{j} to Bj+1B_{j+1} such that BjBj+1jUQil,η3B_{j}\cup B_{j+1}\subset\mathfrak{C}_{j}\subset U_{Q_{i}^{l},\eta^{3}}, the cylinders have bounded overlaps and |j|η(Qil)n+1|\mathfrak{C}_{j}|\approx_{\eta}\ell(Q_{i}^{l})^{n+1},

where the bound (iii) follows from properties of ε\varepsilon-approximators (see property iii) in Theorem 1.1). Then

|Φ(X^Q~il)Φ(X^P~il)|B1|Φ(X^Q~il)Φ(X)|𝑑X+j=1N1|BjΦ(X)𝑑XBj+1Φ(X)𝑑X|+BN|Φ(X)Φ(X^P~il)|𝑑Xc04+j=1N1|BjΦ(X)𝑑XBj+1Φ(X)𝑑X||\Phi(\hat{X}_{\widetilde{Q}_{i}^{l}})-\Phi(\hat{X}_{\widetilde{P}_{i}^{l}})|\\ \leq\mathchoice{{\vbox{\hbox{$\textstyle\raisebox{-8.61108pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-7.09656pt}}{{\vbox{\hbox{$\scriptstyle\raisebox{-6.02777pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-6.09258pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-4.29826pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-3.54826pt}}\!\iint_{B_{1}}|\Phi(\hat{X}_{\widetilde{Q}_{i}^{l}})-\Phi(X)|\,dX+\sum_{j=1}^{N-1}\Big{|}\mathchoice{{\vbox{\hbox{$\textstyle\raisebox{-8.61108pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-7.09656pt}}{{\vbox{\hbox{$\scriptstyle\raisebox{-6.02777pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-6.09258pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-4.29826pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-3.54826pt}}\!\iint_{B_{j}}\Phi(X)\,dX-\mathchoice{{\vbox{\hbox{$\textstyle\raisebox{-8.61108pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-7.09656pt}}{{\vbox{\hbox{$\scriptstyle\raisebox{-6.02777pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-6.09258pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-4.29826pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-3.54826pt}}\!\iint_{B_{j+1}}\Phi(X)\,dX\Big{|}+\mathchoice{{\vbox{\hbox{$\textstyle\raisebox{-8.61108pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-7.09656pt}}{{\vbox{\hbox{$\scriptstyle\raisebox{-6.02777pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-6.09258pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-4.29826pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-3.54826pt}}\!\iint_{B_{N}}|\Phi(X)-\Phi(\hat{X}_{\widetilde{P}_{i}^{l}})|\,dX\\ \leq\frac{c_{0}}{4}+\sum_{j=1}^{N-1}\Big{|}\mathchoice{{\vbox{\hbox{$\textstyle\raisebox{-8.61108pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-7.09656pt}}{{\vbox{\hbox{$\scriptstyle\raisebox{-6.02777pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-6.09258pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-4.29826pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-3.54826pt}}\!\iint_{B_{j}}\Phi(X)\,dX-\mathchoice{{\vbox{\hbox{$\textstyle\raisebox{-8.61108pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-7.09656pt}}{{\vbox{\hbox{$\scriptstyle\raisebox{-6.02777pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-6.09258pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-4.29826pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-3.54826pt}}\!\iint_{B_{j+1}}\Phi(X)\,dX\Big{|}

and furthermore, by Poincaré inequality and properties of the Whitney regions and cylinders j\mathfrak{C}_{j}, we have

|BjΦ(X)𝑑XBj+1Φ(X)𝑑X|=|BjΦ(X)𝑑XjΦ(X)𝑑X+jΦ(X)𝑑XBj+1Φ(X)𝑑X|ηj|Φ(X)jΦ(Y)𝑑Y|𝑑Xη(Qil)|j|j|Φ(X)|𝑑Xηj|Φ(X)|δ(X)n𝑑X\Big{|}\mathchoice{{\vbox{\hbox{$\textstyle\raisebox{-8.61108pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-7.09656pt}}{{\vbox{\hbox{$\scriptstyle\raisebox{-6.02777pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-6.09258pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-4.29826pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-3.54826pt}}\!\iint_{B_{j}}\Phi(X)\,dX-\mathchoice{{\vbox{\hbox{$\textstyle\raisebox{-8.61108pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-7.09656pt}}{{\vbox{\hbox{$\scriptstyle\raisebox{-6.02777pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-6.09258pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-4.29826pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-3.54826pt}}\!\iint_{B_{j+1}}\Phi(X)\,dX\Big{|}=\Big{|}\mathchoice{{\vbox{\hbox{$\textstyle\raisebox{-8.61108pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-7.09656pt}}{{\vbox{\hbox{$\scriptstyle\raisebox{-6.02777pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-6.09258pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-4.29826pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-3.54826pt}}\!\iint_{B_{j}}\Phi(X)\,dX-\mathchoice{{\vbox{\hbox{$\textstyle\raisebox{-8.61108pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-7.09656pt}}{{\vbox{\hbox{$\scriptstyle\raisebox{-6.02777pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-6.09258pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-4.29826pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-3.54826pt}}\!\iint_{\mathfrak{C}_{j}}\Phi(X)\,dX+\mathchoice{{\vbox{\hbox{$\textstyle\raisebox{-8.61108pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-7.09656pt}}{{\vbox{\hbox{$\scriptstyle\raisebox{-6.02777pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-6.09258pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-4.29826pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-3.54826pt}}\!\iint_{\mathfrak{C}_{j}}\Phi(X)\,dX-\mathchoice{{\vbox{\hbox{$\textstyle\raisebox{-8.61108pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-7.09656pt}}{{\vbox{\hbox{$\scriptstyle\raisebox{-6.02777pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-6.09258pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-4.29826pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-3.54826pt}}\!\iint_{B_{j+1}}\Phi(X)\,dX\Big{|}\\ \lesssim_{\eta}\mathchoice{{\vbox{\hbox{$\textstyle\raisebox{-8.61108pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-7.09656pt}}{{\vbox{\hbox{$\scriptstyle\raisebox{-6.02777pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-6.09258pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-4.29826pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-3.54826pt}}\!\iint_{\mathfrak{C}_{j}}\Big{|}\Phi(X)-\mathchoice{{\vbox{\hbox{$\textstyle\raisebox{-8.61108pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-7.09656pt}}{{\vbox{\hbox{$\scriptstyle\raisebox{-6.02777pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-6.09258pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-4.29826pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-3.54826pt}}\!\iint_{\mathfrak{C}_{j}}\Phi(Y)\,dY\Big{|}\,dX\lesssim_{\eta}\frac{\ell(Q_{i}^{l})}{|\mathfrak{C}_{j}|}\iint_{\mathfrak{C}_{j}}|\nabla\Phi(X)|\,dX\lesssim_{\eta}\iint_{\mathfrak{C}_{j}}|\nabla\Phi(X)|\delta(X)^{-n}\,dX

for every j=1,2,,N1j=1,2,\ldots,N-1. Combining the previous estimates and using the bounded overlaps of the cylinders then gives us

c04j=1N1|BjΦ(X)𝑑XBj+1Φ(X)𝑑X|\displaystyle\frac{c_{0}}{4}\leq\sum_{j=1}^{N-1}\Big{|}\mathchoice{{\vbox{\hbox{$\textstyle\raisebox{-8.61108pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-7.09656pt}}{{\vbox{\hbox{$\scriptstyle\raisebox{-6.02777pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-6.09258pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-4.29826pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-3.54826pt}}\!\iint_{B_{j}}\Phi(X)\,dX-\mathchoice{{\vbox{\hbox{$\textstyle\raisebox{-8.61108pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-7.09656pt}}{{\vbox{\hbox{$\scriptstyle\raisebox{-6.02777pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-6.09258pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-4.29826pt}}{{\vbox{\hbox{$\scriptscriptstyle\raisebox{-4.30554pt}{\rotatebox[origin={c}]{20.0}{${-}\mkern-2.0mu{-}$}}$}}\kern-3.54826pt}}\!\iint_{B_{j+1}}\Phi(X)\,dX\Big{|} ηUQil,η3|Φ(X)|δ(X)n𝑑X.\displaystyle\lesssim_{\eta}\iint_{U_{Q_{i}^{l},\eta^{3}}}|\nabla\Phi(X)|\delta(X)^{-n}\,dX.

Finally, we sum over ll and use Lemma 5.5, the bounded overlaps of the regions UQil,η3U_{Q_{i}^{l},\eta^{3}} and the structure of ΓQilη\Gamma_{Q_{i}^{l}}^{\eta} to get

c04logα1logϵ01c04(k1)ηl=1k1UQil,η3|Φ(X)|δ(X)n𝑑XηΓQilη(y)|Φ(X)|δ(X)n𝑑X,\frac{c_{0}}{4}\frac{\log\alpha^{-1}}{\log\epsilon_{0}^{-1}}\approx\frac{c_{0}}{4}(k-1)\lesssim_{\eta}\sum_{l=1}^{k-1}\iint_{U_{Q_{i}^{l},\eta^{3}}}|\nabla\Phi(X)|\delta(X)^{-n}\,dX\lesssim_{\eta}\iint_{\Gamma_{Q_{i}^{l}}^{\eta}(y)}|\nabla\Phi(X)|\delta(X)^{-n}\,dX,

which proves the claim. ∎

6. Proof of Theorem 1.5

With the help of the tools from the previous sections, we can now prove Theorem 1.5. The proof is an adaptation of the corresponding proof from [HT21] (which itself uses some core ideas of Varopoulos [Var77, Var78] and Garnett [Gar07]).

Proof of Theorem 1.5.

Let fBMOc(Ω)f\in\operatorname{BMO}_{\operatorname{c}}(\partial\Omega) and 𝔻\mathbb{D} be a dyadic system on Ω\partial\Omega. By Lemma 10.1 and Remark 10.3 in [HT21], we know that

(6.1) f=f0+g,\displaystyle f=f_{0}+g,

where f0L(Ω)f_{0}\in L^{\infty}(\partial\Omega) with f0L(Ω)fBMO\|f_{0}\|_{L^{\infty}(\partial\Omega)}\lesssim\|f\|_{\operatorname{BMO}} and g=jαj𝟙Qjg=\sum_{j}\alpha_{j}\mathbbm{1}_{Q_{j}}, for a collection of dyadic cubes 𝒜{Qj}j𝔻\mathcal{A}\coloneqq\{Q_{j}\}_{j}\subset\mathbb{D} satisfying a Carleson packing condition with 𝒞𝒜1\mathcal{C}_{\mathcal{A}}\lesssim 1 and coefficients αj\alpha_{j} satisfying supj|αj|fBMO(Ω)\sup_{j}|\alpha_{j}|\lesssim\|f\|_{\operatorname{BMO}(\partial\Omega)}. By [HT21, Proposition 1.3], there exists an extension GG of gg in Ω\Omega that satisfies the corresponding versions of the properties (1) – (3) in Theorem 1.5. Thus, it is enough to construct the extension for the function f0f_{0} in the decomposition (6.1).

Construction of the extension for f0f_{0} follows the proof of [HT21, Theorem 1.2]. We repeatedly use Lemma 3.3 to solve boundary value problems for LL with updated data and Theorem 1.1 to take smooth 12\tfrac{1}{2}-approximators (that is, ε\varepsilon-approximators for ε=12\varepsilon=\tfrac{1}{2}) for the corresponding solutions and to take their a.e. non-tangential boundary traces. The idea is that a suitable approximator is a Varopoulos-type extension plus an error term and we can keep on halving the size of the error term through iteration.

We start by taking the solution u0u_{0} to the boundary value problem with data f0f_{0}, satisfying u0(X)=Ωf0𝑑ωXu_{0}(X)=\int_{\partial\Omega}f_{0}\,d\omega^{X}. We then take the 12\tfrac{1}{2}-approximator Φ0\Phi_{0} to u0u_{0}. This approximator has an a.e. non-tangential boundary trace φ0\varphi_{0}. By Theorem 1.1, these functions satisfy u0Φ0L(Ω)12u0L(Ω)12f0L(Ω)\|u_{0}-\Phi_{0}\|_{L^{\infty}(\Omega)}\leq\tfrac{1}{2}\|u_{0}\|_{L^{\infty}(\Omega)}\leq\tfrac{1}{2}\|f_{0}\|_{L^{\infty}(\partial\Omega)}, and

supxΩ,r>01rnB(x,r)Ω|Φ0(Y)|𝑑YC~u0L(Ω)C~f0L(Ω).\displaystyle\sup_{x\in\partial\Omega,r>0}\frac{1}{r^{n}}\iint_{B(x,r)\cap\Omega}|\nabla\Phi_{0}(Y)|\,dY\leq\widetilde{C}\|u_{0}\|_{L^{\infty}(\Omega)}\leq\widetilde{C}\|f_{0}\|_{L^{\infty}(\partial\Omega)}.

We set f1f0φ0f_{1}\coloneqq f_{0}-\varphi_{0} which is the a.e. non-tangential boundary trace of u0Φ0u_{0}-\Phi_{0}. We take the solution u1u_{1} with the data f1f_{1}, satisfying u1(X)=Ωf1𝑑ωXu_{1}(X)=\int_{\partial\Omega}f_{1}\,d\omega^{X}. This solution satisfies

u1L(Ω)f1L(Ω)u0Φ0L(Ω)12u0L(Ω)12f0L(Ω),\displaystyle\|u_{1}\|_{L^{\infty}(\Omega)}\leq\|f_{1}\|_{L^{\infty}(\partial\Omega)}\leq\|u_{0}-\Phi_{0}\|_{L^{\infty}(\Omega)}\leq\tfrac{1}{2}\|u_{0}\|_{L^{\infty}(\Omega)}\leq\tfrac{1}{2}\|f_{0}\|_{L^{\infty}(\partial\Omega)},

and thus, we can take a 12\tfrac{1}{2}-approximator Φ1\Phi_{1} of u1u_{1}, with boundary trace φ1\varphi_{1}. We get

u1Φ1L(Ω)12u1L(Ω)14u0L(Ω)\displaystyle\|u_{1}-\Phi_{1}\|_{L^{\infty}(\Omega)}\leq\tfrac{1}{2}\|u_{1}\|_{L^{\infty}(\Omega)}\leq\tfrac{1}{4}\|u_{0}\|_{L^{\infty}(\Omega)}

and

supxΩ,r>01rnB(x,r)Ω|Φ1(Y)|𝑑YC~u1L(Ω)C~2f0L(Ω).\displaystyle\sup_{x\in\partial\Omega,r>0}\frac{1}{r^{n}}\iint_{B(x,r)\cap\Omega}|\nabla\Phi_{1}(Y)|\,dY\leq\widetilde{C}\|u_{1}\|_{L^{\infty}(\Omega)}\leq\frac{\widetilde{C}}{2}\|f_{0}\|_{L^{\infty}(\partial\Omega)}.

We set f2f1φ1=f0φ0φ1f_{2}\coloneqq f_{1}-\varphi_{1}=f_{0}-\varphi_{0}-\varphi_{1}, and continue in the previous way. This gives us a sequences of solutions uku_{k} and their 12\tfrac{1}{2}-approximators Φk\Phi_{k}. The functions uku_{k} and Φk\Phi_{k} have a.e. non-tangential boundary traces fkf_{k} and φk\varphi_{k}, respectively, and we have

  1. (i)

    fk+1=f0i=0kφif_{k+1}=f_{0}-\sum_{i=0}^{k}\varphi_{i},

  2. (ii)

    fk+1L(Ω)ukΦkL(Ω)2k1u0L(Ω)2k1f0L(Ω)\|f_{k+1}\|_{L^{\infty}(\partial\Omega)}\leq\|u_{k}-\Phi_{k}\|_{L^{\infty}(\Omega)}\leq 2^{-k-1}\|u_{0}\|_{L^{\infty}(\Omega)}\leq 2^{-k-1}\|f_{0}\|_{L^{\infty}(\partial\Omega)},

  3. (iii)

    ukL(Ω)fkL(Ω)2ku0L(Ω)\|u_{k}\|_{L^{\infty}(\Omega)}\leq\|f_{k}\|_{L^{\infty}(\partial\Omega)}\leq 2^{-k}\|u_{0}\|_{L^{\infty}(\Omega)},

  4. (iv)

    supxΩ,r>01rnB(x,r)Ω|Φk(Y)|𝑑YC~ukL(Ω)C~2ku0L(Ω)\sup_{x\in\partial\Omega,r>0}\frac{1}{r^{n}}\iint_{B(x,r)\cap\Omega}|\nabla\Phi_{k}(Y)|\,dY\leq\widetilde{C}\|u_{k}\|_{L^{\infty}(\Omega)}\leq\widetilde{C}2^{-k}\|u_{0}\|_{L^{\infty}(\Omega)}, and

  5. (v)

    ΦkL(Ω)2ku0L(Ω)\|\Phi_{k}\|_{L^{\infty}(\Omega)}\lesssim 2^{-k}\|u_{0}\|_{L^{\infty}(\Omega)}.

Property (v) follows from the properties (ii) and (iii) combined with the fact that Φk\Phi_{k} is a 12\tfrac{1}{2}-approximator of uku_{k}. By this property, we can define a uniformly convergent series

Φ(X)k=0Φk(X)\displaystyle\Phi(X)\coloneqq\sum_{k=0}^{\infty}\Phi_{k}(X)

for XΩX\in\Omega. Let F0F_{0} be a version of Φ\Phi that has been smoothened using similar convolution techniques as in the proof of Lemma 4.28 (see [HT21, Section 3] for details). By the arguments in the proof of [HT21, Theorem 1.1], F0F_{0} is an extension of f0f_{0} that satisfies the properties (1)–(3) in Theorem 1.5. Thus, by the decomposition (6.1), we may define the extensions FF in Theorem 1.5 by setting FF0+GF\coloneqq F_{0}+G. This completes the proof. ∎

7. An example in 3\mathbb{R}^{3}

In this section we construct a three-dimensional version of the example provided in Corollary 1.7. That is, we show there is a domain Ω\Omega in 3\mathbb{R}^{3} whose boundary is not rectifiable and such that every function fBMO(Ω)f\in\operatorname{BMO}(\partial\Omega) has a Varopoulos extension in Ω\Omega.

To define Ω\Omega, denote by EE the 44-corner Cantor set in 2\mathbb{R}^{2}. That is E=k=0EkE=\bigcap_{k=0}^{\infty}E_{k}, where EkE_{k} equals the union of 4k4^{k} closed squares QikQ_{i}^{k} of side length 4k4^{-k} located in the corners of the squares Qjk1Q_{j}^{k-1} of the previous generation (see [DM21, Section 3] for the precise definition). We assume that the center of E0=Q10E_{0}=Q_{1}^{0} coincides with the origin in 2\mathbb{R}^{2}. We consider the half-plane Π={(x,y)2:y>2}\Pi=\{(x,y)\in\mathbb{R}^{2}:y>-2\} and we set V=ΠEV=\Pi\setminus E. We also write 𝕃=Π={(x,y)2:y=2}\mathbb{L}=\partial\Pi=\{(x,y)\in\mathbb{R}^{2}:y=-2\}. Then we define Ω=V×\Omega=V\times\mathbb{R}. It is straightforward to check that both VV and Ω\Omega are uniform domains.

Notice that V=E𝕃\partial V=E\cup\mathbb{L} is 11-Ahlfors regular, while Ω=(E𝕃)×\partial\Omega=(E\cup\mathbb{L})\times\mathbb{R} is 22-Ahlfors regular. In fact, the purpose of introducing the half plane Π\Pi and the line 𝕃\mathbb{L} in this previous construction is to ensure the 22-Ahlfors regularity of Ω\partial\Omega. It is also clear that E×E\times\mathbb{R} is purely 22-unrectifiable (since this set has no approximate tangent planes at any point).

Let AA be the 2×22\times 2 matrix in the David–Mayboroda example in Theorem 1.6, let L=divAL=-\operatorname{div}A\nabla be the associated elliptic operator in 2\mathbb{R}^{2}, and let

A^=(A001).\hat{A}=\left(\begin{array}[]{ll}A&0\\ 0&1\end{array}\right).

Set L^=divA^\hat{L}=-\mathop{\operatorname{div}}\nolimits\hat{A}\,\nabla in 3\mathbb{R}^{3}. Below we will show that ωL^,ΩA(σ)\omega_{\hat{L},\Omega}\in A_{\infty}(\sigma), where σ\sigma is the surface measure on Ω\partial\Omega. Consequently, by Theorem 1.5, every function fBMO(Ω)f\in\operatorname{BMO}(\partial\Omega) has a Varopoulos extension in Ω\Omega.

First, we prove the following.

Lemma 7.1.

We have that ωL,VA(1|V)\omega_{L,V}\in A_{\infty}(\mathcal{H}^{1}|_{\partial V}). Further, there is a constant C>0C>0 such that for any surface ball ΔV\Delta\subset\partial V and any corkscrew point pVp\in V for Δ\Delta,

(7.2) dωL,Vpd1|V(x)C1(Δ),for each xΔ.\frac{d\omega_{L,V}^{p}}{d\mathcal{H}^{1}|_{\partial V}}(x)\leq\frac{C}{\mathcal{H}^{1}(\Delta)},\qquad\text{for each }x\in\Delta.
Proof.

It is clear that the estimate (7.2) implies the local AA_{\infty} condition of ωLωL,V\omega_{L}\equiv\omega_{L,V}. First we will show that, for any pΠp\in\Pi,

(7.3) dωL,Πpd1|𝕃(x)CdωΔ,Πpd1|𝕃(x),\frac{d\omega_{L,\Pi}^{p}}{d\mathcal{H}^{1}|_{\mathbb{L}}}(x)\leq C\,\frac{d\omega_{-\Delta,\Pi}^{p}}{d\mathcal{H}^{1}|_{\mathbb{L}}}(x),

where ωL,Π\omega_{L,\Pi} stands for the LL-elliptic measure for the domain Π\Pi, and ωΔ,Π\omega_{-\Delta,\Pi} is the harmonic measure (i.e. for the Laplacian) for the domain Π\Pi. To this end, we consider the auxiliary domain U=ΠB1¯U=\Pi\setminus\overline{B_{1}}, where we denoted Br=B(0,r)B_{r}=B(0,r). Observe that AA is the identity matrix on UU, and thus ωL,UXωΔ,UX\omega_{L,U}^{X}\equiv\omega_{-\Delta,U}^{X} for all XUX\in U. Let F𝕃F\subset\mathbb{L} be an arbitrary closed subset, and let qB3/2q\in\partial B_{3/2} be such that

ωL,Πq(F)=maxXB3/2ωL,ΠX(F).\omega_{L,\Pi}^{q}(F)=\max_{X\in\partial B_{3/2}}\omega_{L,\Pi}^{X}(F).

Since ωL,Πx(F)\omega_{L,\Pi}^{x}(F) is a harmonic function of xx in UU, we have

(7.4) ωL,Πq(F)=UωL,ΠX(F)𝑑ωΔ,Uq(X)=𝕃ωL,ΠX(F)𝑑ωΔ,Uq(X)+B1ωL,ΠX(F)𝑑ωΔ,Uq(X)ωΔ,Uq(F)+supXB1ωL,ΠX(F)ωΔ,Uq(B1).\omega_{L,\Pi}^{q}(F)=\int_{\partial U}\omega^{X}_{L,\Pi}(F)\,d\omega^{q}_{-\Delta,U}(X)=\int_{\mathbb{L}}\omega_{L,\Pi}^{X}(F)\,d\omega^{q}_{-\Delta,U}(X)+\int_{\partial B_{1}}\omega_{L,\Pi}^{X}(F)\,d\omega^{q}_{-\Delta,U}(X)\\ \leq\omega_{-\Delta,U}^{q}(F)+\sup_{X\in\partial B_{1}}\omega_{L,\Pi}^{X}(F)\,\omega_{-\Delta,U}^{q}(\partial B_{1}).

By the maximum principle and the definition of qq, we have

supXB1ωL,ΠX(F)supXB3/2ωL,ΠX(F)=ωL,Πq(F).\sup_{X\in\partial B_{1}}\omega_{L,\Pi}^{X}(F)\leq\sup_{X\in\partial B_{3/2}}\omega_{L,\Pi}^{X}(F)=\omega_{L,\Pi}^{q}(F).

Also, it is immediate that cBωΔ,Uq(B1)<1c_{B}\coloneqq\omega_{-\Delta,U}^{q}(\partial B_{1})<1. Hence,

ωL,Πq(F)ωΔ,Uq(F)+cBωL,Πq(F),\omega_{L,\Pi}^{q}(F)\leq\omega_{-\Delta,U}^{q}(F)+c_{B}\,\omega_{L,\Pi}^{q}(F),

or equivalently, ωL,Πq(F)(1cB)1ωΔ,Uq(F)\omega_{L,\Pi}^{q}(F)\leq(1-c_{B})^{-1}\omega_{-\Delta,U}^{q}(F). By a Harnack chain argument we deduce that ωL,Πp(F)ωΔ,Up(F)=ωL,Up(F)\omega_{L,\Pi}^{p}(F)\lesssim\omega_{-\Delta,U}^{p}(F)=\omega_{L,U}^{p}(F) for all pB3/2p\in\partial B_{3/2}, and then by the maximum principle, it follows that the same estimate is valid for all pΠB¯3/2p\in\Pi\setminus\overline{B}_{3/2}. Using again the maximum principle, we get, for all pΠB¯3/2p\in\Pi\setminus\overline{B}_{3/2},

ωL,Πp(F)ωΔ,Up(F)ωΔ,Πp(F).\omega_{L,\Pi}^{p}(F)\lesssim\omega_{-\Delta,U}^{p}(F)\leq\omega_{-\Delta,\Pi}^{p}(F).

Since ωL,Πp(F)\omega_{L,\Pi}^{p}(F) and ωΔ,Up(F)\omega_{-\Delta,U}^{p}(F) are, respectively, elliptic and harmonic in B3/2B_{3/2}, by a Harnack chain argument it follows that the estimate above also holds for all pB3/2p\in B_{3/2}, possibly with a different implicit constant. This is equivalent to (7.3).

To prove (7.2), let BB be a ball centered in V\partial V such that Δ=BΩ\Delta=B\cap\partial\Omega and consider an arbitrary closed set FΔF\subset\Delta. Let pBVp\in B\cap V be a corkscrew point for Δ\Delta. By the maximum principle and by (7.3),

ωL,Vp(F𝕃)ωL,Πp(F𝕃)ωΔ,Πp(F𝕃)1(F𝕃)1(Δ),\omega_{L,V}^{p}(F\cap\mathbb{L})\leq\omega_{L,\Pi}^{p}(F\cap\mathbb{L})\lesssim\omega_{-\Delta,\Pi}^{p}(F\cap\mathbb{L})\lesssim\frac{\mathcal{H}^{1}(F\cap\mathbb{L})}{\mathcal{H}^{1}(\Delta)},

where in the last estimate we used that dωΔ,Πpd1|𝕃(x)σ(Δ)1\frac{d\omega_{-\Delta,\Pi}^{p}}{d\mathcal{H}^{1}|_{\mathbb{L}}}(x)\lesssim\sigma(\Delta)^{-1} for all xΔx\in\Delta. Indeed, if F𝕃=F\cap\mathbb{L}=\varnothing, then there is nothing to show; if BB is centered on 𝕃\mathbb{L}, this follows from Ahlfors regularity of V\partial V and classical properties of the harmonic measure, and if BB is centered on EE and there exists zF𝕃z\in F\cap\mathbb{L}, then if B=B(z,rad(B))B^{\prime}=B(z,\operatorname{rad}(B)) and pp^{\prime} is a corkscrew point for BB^{\prime} in VV, then by Harnack chains we have that ωΔ,Πp(F𝕃)ωΔ,Πp(F𝕃)\omega^{p}_{-\Delta,\Pi}(F\cap\mathbb{L})\approx\omega^{p^{\prime}}_{-\Delta,\Pi}(F\cap\mathbb{L}), and the claim follows as in the previous case.

To estimate ωL,Vp(FE)\omega_{L,V}^{p}(F\cap E), we assume that BEB\cap E\neq\varnothing and we distinguish two cases. Suppose first that r(B)r(B), the radius of BB, satisfies r(B)1r(B)\leq 1. Denote XE=(0,2)X_{E}=(0,2) and let BB^{\prime} be a ball centered in EE containing BB with radius at most 2r(B)2r(B). By the maximum principle, the change of poles formula, and the fact that dωL,EcXEd1|E(x)1\frac{d\omega_{L,E^{c}}^{X_{E}}}{d\mathcal{H}^{1}|_{E}}(x)\approx 1 (by Theorem 1.6), we obtain

ωL,Vp(FE)ωL,Ecp(FE)ωL,EcXE(FE)ωL,EcXE(B)1(FE)1(BE)1(FE)1(Δ).\omega_{L,V}^{p}(F\cap E)\leq\omega_{L,E^{c}}^{p}(F\cap E)\approx\frac{\omega_{L,E^{c}}^{X_{E}}(F\cap E)}{\omega_{L,E^{c}}^{X_{E}}(B^{\prime})}\approx\frac{\mathcal{H}^{1}(F\cap E)}{\mathcal{H}^{1}(B^{\prime}\cap E)}\approx\frac{\mathcal{H}^{1}(F\cap E)}{\mathcal{H}^{1}(\Delta)}.

In the case r(B)>1r(B)>1, recall that B1=B(0,1)B_{1}=B(0,1), and then using the change of poles formula and the maximum principle, we write

ωL,Vp(FE)ωL,VXE(FE)ωL,Vp(B1)ωL,EcXE(FE)ωL,Vp(B1)1(FE)ωL,Vp(B1).\omega_{L,V}^{p}(F\cap E)\approx\omega_{L,V}^{X_{E}}(F\cap E)\,\omega_{L,V}^{p}(B_{1})\leq\omega_{L,E^{c}}^{X_{E}}(F\cap E)\,\omega_{L,V}^{p}(B_{1})\approx\mathcal{H}^{1}(F\cap E)\,\omega_{L,V}^{p}(B_{1}).

To estimate ωL,Vp(B1)\omega_{L,V}^{p}(B_{1}), consider a ball B1B_{1}^{\prime} of radius 11 centered in 𝕃\mathbb{L}, disjoint from EE, and contained in 4B14B_{1}. We claim that

(7.5) ωL,Vp(B1)ωL,Vp(B1).\omega_{L,V}^{p}(B_{1})\approx\omega_{L,V}^{p}(B_{1}^{\prime}).

Indeed, if p8B1p\notin 8B_{1}, then (7.5) follows directly from Lemma 2.37. On the other hand, if p8B1p\in 8B_{1}, then denote p=(0,16)p^{\prime}=(0,16) and note that δ(p)1\delta(p)\gtrsim 1, δ(p)8\delta(p^{\prime})\geq 8, and |pp|24|p-p^{\prime}|\leq 24. Then, by the Harnack inequality, Harnack chains, and Lemma 2.37, the estimate (7.5) follows. Next, by the maximum principle, (7.3), and Ahlfors regularity, we get

ωL,Vp(B1)ωL,Πp(B1)ωΔ,Πp(B1)1(B1𝕃)1(Δ)11(Δ).\omega_{L,V}^{p}(B_{1}^{\prime})\leq\omega_{L,\Pi}^{p}(B_{1}^{\prime})\lesssim\omega_{-\Delta,\Pi}^{p}(B_{1}^{\prime})\lesssim\frac{\mathcal{H}^{1}(B_{1}^{\prime}\cap\mathbb{L})}{\mathcal{H}^{1}(\Delta)}\approx\frac{1}{\mathcal{H}^{1}(\Delta)}.

Therefore, again we derive

ωL,Vp(FE)1(FE)1(Δ).\omega_{L,V}^{p}(F\cap E)\lesssim\frac{\mathcal{H}^{1}(F\cap E)}{\mathcal{H}^{1}(\Delta)}.

Altogether, we deduce that

ωL,Vp(F)=ωL,Vp(F𝕃)+ωL,Vp(FE)1(F)1(Δ)\omega_{L,V}^{p}(F)=\omega_{L,V}^{p}(F\cap\mathbb{L})+\omega_{L,V}^{p}(F\cap E)\lesssim\frac{\mathcal{H}^{1}(F)}{\mathcal{H}^{1}(\Delta)}

for any closed set FΔF\subset\Delta, which is equivalent to the statement in the lemma, by the Lebesgue-Radon-Nykodim Theorem. ∎

Now we are ready to prove the AA_{\infty} property of the elliptic measure ωL^\omega_{\hat{L}} for the three-dimensional domain Ω\Omega:

Proposition 7.6.

The elliptic measure ωL^\omega_{\hat{L}} for the domain Ω3\Omega\subset\mathbb{R}^{3} defined above satisfies the AA_{\infty} condition with respect to 2|Ω\mathcal{H}^{2}|_{\partial\Omega}. Further, there is a constant C>0C>0 such that for any surface ball ΔΩ\Delta\subset\partial\Omega and any corkscrew point pΩp\in\Omega for Δ\Delta,

(7.7) dωL^pd2|Ω(x)C2(Δ),for each xΔ.\frac{d\omega_{\hat{L}}^{p}}{d\mathcal{H}^{2}|_{\partial\Omega}}(x)\leq\frac{C}{\mathcal{H}^{2}(\Delta)},\qquad\text{for each }x\in\Delta.
Proof.

Let BB be a ball with radius r(B)r(B) centered in Ω\partial\Omega such that Δ=BΩ\Delta=B\cap\partial\Omega. Clearly, to prove the AA_{\infty} condition for ωL^\omega_{\hat{L}}, it suffices to show (7.7). In turn, since L^\hat{L} is symmetric, by a direct application of Lemma 2.38 and the Lebesgue-Radon-Nykodim Theorem, it is enough to prove that

(7.8) GL^(X,p)dist(X,Ω)r(B)2 for all XBΩB(p,12dist(p,Ω)),G_{\hat{L}}(X,p)\lesssim\frac{\operatorname{dist}(X,\partial\Omega)}{r(B)^{2}}\quad\mbox{ for all $X\in B\cap\Omega\setminus B(p,\frac{1}{2}\operatorname{dist}(p,\partial\Omega))$,}

where GL^G_{\hat{L}} is the L^\hat{L}-Green function for Ω\Omega.

Denote by PP the orthogonal projection of 3\mathbb{R}^{3} onto 22×{0}\mathbb{R}^{2}\equiv\mathbb{R}^{2}\times\{0\}. Let p0=P(p)p_{0}=P(p) and consider the function u:Ω¯P1({p0})u:\overline{\Omega}\setminus P^{-1}(\{p_{0}\})\to\mathbb{R} defined by

u(X)=GL(P(X),P(p))=GL(P(X),p0),u(X)=G_{L}(P(X),P(p))=G_{L}(P(X),p_{0}),

where GLG_{L} is the Green’s function for LL in the domain V=P(Ω)V=P(\Omega). It is immediate to check that uu is L^\hat{L}-elliptic in ΩP1({p0})\Omega\setminus P^{-1}(\{p_{0}\}), and clearly it can be extended continuously by zero to the whole Ω\partial\Omega. Thus, by the boundary Harnack principle, choosing pB(p,12dist(p,Ω))p^{\prime}\in\partial B(p,\frac{1}{2}\operatorname{dist}(p,\partial\Omega)) such that P(p)B(p0,12dist(p,Ω))2P(p^{\prime})\in\partial B(p_{0},\frac{1}{2}\operatorname{dist}(p,\partial\Omega))\cap\mathbb{R}^{2}, we have

GL^(X,p)GL^(p,p)u(X)u(p)=GL(P(X),p0)GL(P(p),p0) for all XBΩB(p,12dist(p,Ω)).\frac{G_{\hat{L}}(X,p)}{G_{\hat{L}}(p^{\prime},p)}\approx\frac{u(X)}{u(p^{\prime})}=\frac{G_{L}(P(X),p_{0})}{G_{L}(P(p^{\prime}),p_{0})}\quad\mbox{ for all $X\in B\cap\Omega\setminus B(p,\frac{1}{2}\operatorname{dist}(p,\partial\Omega))$.}

Thus, for such points XX and by (2.33) and (2.34) applied both to GL^G_{\hat{L}} and GLG_{L},

GL^(X,p)|pp|1GL(P(X),p0)1.\frac{G_{\hat{L}}(X,p)}{|p^{\prime}-p|^{-1}}\approx\frac{G_{L}(P(X),p_{0})}{1}.

Thus, by Lemma 2.38 applied to GL,VG_{L,V}, Harnack chains, and the Harnack inequality, we see that

(7.9) GL^(X,p)GL,V(P(X),p0)r(B)ωL,Vp0(ΔV,X)r(B),G_{\hat{L}}(X,p)\approx\frac{G_{L,V}(P(X),p_{0})}{r(B)}\approx\frac{\omega_{L,V}^{p_{0}}(\Delta_{V,X})}{r(B)},

where ΔV,X=B(P(X),2dist(P(X),V))\Delta_{V,X}=B(P(X),2\operatorname{dist}(P(X),\partial V)). From (7.2) we infer that

(7.10) ωL,Vp0(ΔV,X)1(ΔV,X)1(P(B)V)dist(X,Ω)r(B).\omega_{L,V}^{p_{0}}(\Delta_{V,X})\lesssim\frac{\mathcal{H}^{1}(\Delta_{V,X})}{\mathcal{H}^{1}(P(B)\cap\partial V)}\approx\frac{\operatorname{dist}(X,\partial\Omega)}{r(B)}.

From (7.9) and (7.10), we deduce (7.8), which concludes the proof. ∎

References

  • [AHMT19] M. Akman, S. Hofmann, J. M. Martell, and T. Toro. Perturbation of elliptic operators in 1-sided NTA domains satisfying the capacity density condition, 2019. Preprint, arxiv.1901.08261.
  • [ADFJM19] D. N. Arnold, G. David, M. Filoche, D. Jerison, and S. Mayboroda. Localization of eigenfunctions via an effective potential. Comm. Partial Differential Equations, 44(11):1186–1216, 2019.
  • [AGMT22] J. Azzam, J. Garnett, M. Mourgoglou, and X. Tolsa. Uniform rectifiability, elliptic measure, square functions, and ε\varepsilon-approximability via an acf monotonicity formula. Int. Math. Res. Not. IMRN, 06 2022. rnab095.
  • [AHMMT20] J. Azzam, S. Hofmann, J. M. Martell, M. Mourgoglou, and X. Tolsa. Harmonic measure and quantitative connectivity: geometric characterization of the LpL^{p}-solvability of the Dirichlet problem. Invent. Math., 222(3):881–993, 2020.
  • [BH20] S. Bortz and S. Hofmann. Quantitative Fatou theorems and uniform rectifiability. Potential Anal., 53(1):329–355, 2020.
  • [BT19] S. Bortz and O. Tapiola. ε\varepsilon-approximability of harmonic functions in LpL^{p} implies uniform rectifiability. Proc. Amer. Math. Soc., 147(5):2107–2121, 2019.
  • [CFMS81] L. Caffarelli, E. Fabes, S. Mortola, and S. Salsa. Boundary behavior of nonnegative solutions of elliptic operators in divergence form. Indiana Univ. Math. J., 30(4):621–640, 1981.
  • [CDMT22] M. Cao, O. Domínguez, J. M. Martell, and P. Tradacete. On the AA_{\infty} condition for elliptic operators in 1-sided nontangentially accessible domains satisfying the capacity density condition. Forum Math. Sigma, 10:Paper No. e59, 57, 2022.
  • [CHM] M. Cao, P. Hidalgo-Palencia, and J. Martell. Carleson measure estimates, corona decompositions, and perturbation of elliptic operators without connectivity. Preprint. arXiv:2202.06363.
  • [Car62] L. Carleson. Interpolations by bounded analytic functions and the corona problem. Ann. of Math. (2), 76:547–559, 1962.
  • [CHMT20] 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.
  • [Chr90] 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.
  • [CF74] R. R. Coifman and C. Fefferman. Weighted norm inequalities for maximal functions and singular integrals. Studia Math., 51:241–250, 1974.
  • [CW71] R. R. Coifman and G. Weiss. Analyse harmonique non-commutative sur certains espaces homogènes. Lecture Notes in Mathematics, Vol. 242. Springer-Verlag, Berlin-New York, 1971. Étude de certaines intégrales singulières.
  • [Dah80] B. E. J. Dahlberg. Approximation of harmonic functions. Ann. Inst. Fourier (Grenoble), 30(2):vi, 97–107, 1980.
  • [DJK84] 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.
  • [DS91] G. David and S. Semmes. Singular integrals and rectifiable sets in 𝐑n{\bf R}^{n}: Beyond Lipschitz graphs. Astérisque, 193:152, 1991.
  • [DFM] G. David, J. Feneuil, and S. Mayboroda. Elliptic theory in domains with boundaries of mixed dimension. Preprint. May 2020. arXiv:2003.09037.
  • [DM21] G. David and S. Mayboroda. Good elliptic operators on Cantor sets. Adv. Math., 383:Paper No. 107687, 21, 2021.
  • [DS93] G. David and S. Semmes. Analysis of and on uniformly rectifiable sets, volume 38 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 1993.
  • [DG57] E. De Giorgi. Sulla differenziabilità e l’analiticità delle estremali degli integrali multipli regolari. Mem. Accad. Sci. Torino. Cl. Sci. Fis. Mat. Nat. (3), 3:25–43, 1957.
  • [DM95] G. Dolzmann and S. Müller. Estimates for Green’s matrices of elliptic systems by LpL^{p} theory. Manuscripta Math., 88(2):261–273, 1995.
  • [DK09] H. Dong and S. Kim. Green’s matrices of second order elliptic systems with measurable coefficients in two dimensional domains. Trans. Amer. Math. Soc., 361(6):3303–3323, 2009.
  • [EG15] L. C. Evans and R. F. Gariepy. Measure theory and fine properties of functions. Textbooks in Mathematics. CRC Press, Boca Raton, FL, revised edition, 2015.
  • [FS72] C. Fefferman and E. M. Stein. HpH^{p} spaces of several variables. Acta Math., 129(3-4):137–193, 1972.
  • [FP22] J. Feneuil and B. Poggi. Generalized Carleson perturbations of elliptic operators and applications. Trans. Amer. Math. Soc., 375(11):7553–7599, 2022.
  • [GR85] 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.
  • [GMT18] J. Garnett, M. Mourgoglou, and X. Tolsa. Uniform rectifiability from Carleson measure estimates and ε\varepsilon-approximability of bounded harmonic functions. Duke Math. J., 167(8):1473–1524, 2018.
  • [Gar07] J. B. Garnett. Bounded analytic functions, volume 236 of Graduate Texts in Mathematics. Springer, New York, first edition, 2007.
  • [Gar22] J. B. Garnett. Carleson measure estimates and ε\varepsilon-approximation for bounded harmonic functions, without Ahlfors regularity assumptions. Rev. Mat. Iberoam., 38(1):323–351, 2022.
  • [GW82] M. Grüter and K.-O. Widman. The Green function for uniformly elliptic equations. Manuscripta Math., 37(3):303–342, 1982.
  • [HKM93] J. Heinonen, T. Kilpeläinen, and O. Martio. Nonlinear potential theory of degenerate elliptic equations. Oxford Mathematical Monographs. The Clarendon Press, Oxford University Press, New York, 1993. Oxford Science Publications.
  • [HKMP15] S. Hofmann, C. Kenig, S. Mayboroda, and J. Pipher. Square function/non-tangential maximal function estimates and the Dirichlet problem for non-symmetric elliptic operators. J. Amer. Math. Soc., 28(2):483–529, 2015.
  • [HK07] S. Hofmann and S. Kim. The Green function estimates for strongly elliptic systems of second order. Manuscripta Math., 124(2):139–172, 2007.
  • [HL18] S. Hofmann and P. Le. BMO solvability and absolute continuity of harmonic measure. J. Geom. Anal., 28(4):3278–3299, 2018.
  • [HLMN17] 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.
  • [HMMTZ21] S. Hofmann, J. M. Martell, , S. Mayboroda, T. Toro, and Z. Zhao. Uniform rectifiability and elliptic operators satisfying a carleson measure condition. Geom. Funct. Anal., 31:325–401, 2021.
  • [HM14] 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.
  • [HMM16] 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.
  • [HMT17] S. Hofmann, J. M. Martell, and T. Toro. AA_{\infty} implies NTA for a class of variable coefficient elliptic operators. J. Differ. Equations, 263(10):6147–6188, 2017.
  • [HMU14] S. Hofmann, J. M. Martell, and I. Uriarte-Tuero. Uniform rectifiability and harmonic measure, II: Poisson kernels in LpL^{p} imply uniform rectifiability. Duke Math. J., 163(8):1601–1654, 2014.
  • [HT20] S. Hofmann and O. Tapiola. Uniform rectifability and ε\varepsilon-approximability of harmonic functions in LpL^{p}. Ann. Inst. Fourier (Grenoble), 70(4):1595–1638, 2020.
  • [HT21] S. Hofmann and O. Tapiola. Uniform rectifiability implies Varopoulos extensions. Adv. Math., 390:Paper No. 107961, 53, 2021.
  • [HK12] T. Hytönen and A. Kairema. Systems of dyadic cubes in a doubling metric space. Colloq. Math., 126(1):1–33, 2012.
  • [HR18] T. Hytönen and A. Rosén. Bounded variation approximation of LpL_{p} dyadic martingales and solutions to elliptic equations. J. Eur. Math. Soc. (JEMS), 20(8):1819–1850, 2018.
  • [HT14] T. Hytönen and O. Tapiola. Almost Lipschitz-continuous wavelets in metric spaces via a new randomization of dyadic cubes. J. Approx. Theory, 185:12–30, 2014.
  • [KKPT16] C. Kenig, B. Kirchheim, J. Pipher, and T. Toro. Square functions and the AA_{\infty} property of elliptic measures. J. Geom. Anal., 26(3):2383–2410, 2016.
  • [KKPT00] C. Kenig, H. Koch, J. Pipher, and T. Toro. A new approach to absolute continuity of elliptic measure, with applications to non-symmetric equations. Adv. Math., 153(2):231–298, 2000.
  • [KN85] C. E. Kenig and W.-M. Ni. On the elliptic equation Luk+Kexp[2u]=0Lu-k+K\,{\rm exp}[2u]=0. Ann. Scuola Norm. Sup. Pisa Cl. Sci. (4), 12(2):191–224, 1985.
  • [Mat95] P. Mattila. Geometry of sets and measures in Euclidean spaces, volume 44 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 1995. Fractals and rectifiability.
  • [Mat21] P. Mattila. Rectifiability; a survey. preprint, arXiv:2112.00540, 2021.
  • [Mey63] N. G. Meyers. An LpL^{p}-estimate for the gradient of solutions of second order elliptic divergence equations. Ann. Scuola Norm. Sup. Pisa Cl. Sci. (3), 17:189–206, 1963.
  • [Mos61] J. Moser. On Harnack’s theorem for elliptic differential equations. Comm. Pure Appl. Math., 14:577–591, 1961.
  • [MPT] M. Mourgoglou, B. Poggi, and X. Tolsa. Solvability of the Poisson-Dirichlet problem with interior data in Lp{L}^{p^{\prime}}-carleson spaces and its applications to the Lp{L}^{p}-regularity problem. Preprint. January 2023 (v2). July 2022 (v1). arXiv: 2207.10554.
  • [Muc72] B. Muckenhoupt. Weighted norm inequalities for the Hardy maximal function. Trans. Amer. Math. Soc., 165:207–226, 1972.
  • [Nas58] J. Nash. Continuity of solutions of parabolic and elliptic equations. Amer. J. Math., 80:931–954, 1958.
  • [Ste70] E. M. Stein. Singular integrals and differentiability properties of functions. Princeton Mathematical Series, No. 30. Princeton University Press, Princeton, N.J., 1970.
  • [Var77] N. T. Varopoulos. BMO functions and the ¯\overline{\partial}-equation. Pacific J. Math., 71(1):221–273, 1977.
  • [Var78] N. T. Varopoulos. A remark on functions of bounded mean oscillation and bounded harmonic functions. Addendum to: “BMO functions and the ¯\overline{\partial}-equation” (Pacific J. Math. 71 (1977), no. 1, 221–273). Pacific J. Math., 74(1):257–259, 1978.