Square function and non-tangential maximal function estimates for elliptic operators in 1-sided NTA domains satisfying the capacity density condition
Abstract.
Let , , be a 1-sided non-tangentially accessible domain (aka uniform domain), that is, satisfies the interior Corkscrew and Harnack chain conditions, which are respectively scale-invariant/quantitative versions of openness and path-connectedness. Let us assume also that satisfies the so-called capacity density condition, a quantitative version of the fact that all boundary points are Wiener regular. Consider , , two real (non-necessarily symmetric) uniformly elliptic operators in , and write , for the respective associated elliptic measures. The goal of this program is to find sufficient conditions guaranteeing that satisfies an -condition or a -condition with respect to . In this paper we are interested in obtaining square function and non-tangential estimates for solutions of operators as before. We establish that bounded weak null-solutions satisfy Carleson measure estimates, with respect to the associated elliptic measure. We also show that for every weak null-solution, the associated square function can be controlled by the non-tangential maximal function in any Lebesgue space with respect to the associated elliptic measure. These results extend previous work of Dahlberg-Jerison-Kenig and are fundamental for the proof of the perturbation results in [2].
Key words and phrases:
Uniformly elliptic operators, elliptic measure, the Green function, 1-sided non-tangentially accessible domains, 1-sided chord-arc domains, capacity density condition, Ahlfors-regularity, Muckenhoupt weights, Reverse Hölder, Carleson measures, square function estimates, non-tangential maximal function estimates, dyadic analysis, sawtooth domains, perturbation2010 Mathematics Subject Classification:
31B05, 35J08, 35J25, 42B37, 42B25, 42B991. Introduction and Main results
The purpose of this program is to study some perturbation problems for second order divergence form real elliptic operators with bounded measurable coefficients in domains with rough boundaries. Let , , be an open set and let be a second order divergence form real elliptic operator defined in . Here the coefficient matrix is real (not necessarily symmetric) and uniformly elliptic, with , that is, there exists a constant such that
(1.1) |
for all and for almost every . Associated with one can construct a family of positive Borel measures , defined on with for every , so that for each one can define its associated weak-solution
(1.2) |
which satisfies in in the weak sense. In principle, unless we assume some further condition, need not be continuous all the way to the boundary, but still we think of as the solution to the continuous Dirichlet problem with boundary data . We call the elliptic measure of associated with the operator with pole at . For convenience, we will sometimes write and call it simply the elliptic measure, dropping the dependence on the pole.
Given two such operators and , one may wonder whether one can find conditions on the matrices and so that some “good estimates” for the Dirichlet problem or for the elliptic measure for might be transferred to the operator . Similarly, one may try to see whether being “close” to in some sense gives some relationship between and . In this direction, a celebrated result of Littman, Stampacchia, and Weinberger in [27] states that the continuous Dirichlet problem for the Laplace operator , (i.e., is the identity) is solvable if and only if it is solvable for any real elliptic operator . By solvability here we mean that the elliptic measure solutions as in (1.2) are indeed continuous in . It is well known that solvability in this sense is in fact equivalent to the fact that all boundary points are regular in the sense of Wiener, a condition which entails some capacitary thickness of the complement of . Note that, for this result, one does not need to know that is “close” to the Laplacian in any sense (other than the fact that both operators are uniformly elliptic).
On the other hand, if is the upper-half plane and , then the harmonic measure associated with is mutually absolutely continuous with respect to the surface measure on the boundary, and its Radon-Nykodym derivative is the classical Poisson kernel. However, Caffarelli, Fabes, and Kenig in [3] constructed a uniformly real elliptic operator in the plane (the pullback of the Laplacian via a quasiconformal mapping of the upper half plane to itself) for which the associated elliptic measure is not even absolutely continuous with respect to the surface measure (see also [29] for another example). Hence, in principle the “good behavior” of harmonic measure does not always transfer to any elliptic measure even in a nice domain such as the upper-half plane. Consequently, it is natural to see if those good properties can be transferred by assuming some conditions reflecting the fact that is “close” to or, in other words, by imposing some conditions on the disagreement of and .
The goal of this program is to solve some perturbation problems that go beyond [9, 11, 12, 28, 5, 6]. Our setting is that of 1-sided NTA domains satisfying the so called capacity density condition (CDC for short), see Section 2 for the precise definitions. The latter is a quantitative version of the well-known Wiener criterion and it is weaker than the Ahlfors regularity of the boundary or the existence of exterior Corkscrews (see Definition 2.1). This setting guarantees among other things that any elliptic measure is doubling in some appropriate sense, hence one can see that a suitable portion of the boundary of the domain endowed with the Euclidean distance and with a given elliptic measure is a space of homogeneous type. In particular, classes like or have the same good features of the corresponding ones in the Euclidean setting. However, our assumptions do not guarantee that the surface measure has any good behavior and it could even be locally infinite. In one of our main results, we consider the case in which a certain disagreement condition, originating in [12], holds either with small or large constant. The small constant case can be seen as an extension of [12, 28] to a setting in which surface measure is not a good object. The large constant case is new even in nice domains such as balls, upper-half spaces, Lipschitz domains or chord-arc domains. To the best of our knowledge, our work is the first to establish perturbation results on sets with bad surface measures, and our large perturbation results are the first of their type. Finally, we do not require the operators to be symmetric. The precise results, along with its context in the historical developments, will be stated in the sequel to the present paper [2].
In the present article we develop some of the needed tools, and present some other results which are of independent interest. Key to our argument is the construction of certain sawtooth domains adapted to a dyadic grid on the boundary and to the Whitney decomposition of the domain. These domains are shown to inherit the main geometrical/topological features of the original domain (see Proposition 2.37). With this in hand we obtain a discrete sawtooth lemma for projections improving [10, Main Lemma], see Lemma 3.5 and Lemma 3.19. These ingredients are crucial for the main results of the papers which we state next. First we establish that bounded weak-solutions satisfy Carleson measure estimates adapted to the elliptic measure.
Theorem 1.3.
Let , , be a 1-sided NTA domain (cf. Definition 2.5) satisfying the capacity density condition (cf. Definition 2.10). Let be a real (not necessarily symmetric) elliptic operator and write and to denote, respectively, the associated elliptic measure and the Green function. There exists depending only on dimension , the 1-sided NTA constants, the CDC constant, and the ellipticity constant of , such that for every with in the weak-sense in there holds
(1.4) |
where , , is a corkscrew point relative to (cf. Definition 2.1), and the sups are taken respectively over all balls with and , and with and , and is the Corkscrew constant.
This result is in turn the main ingredient to obtain that the conical square function can be locally controlled by the non-tangential maximal function in norm with respect to the elliptic measure, allowing us to extend some estimates from [10] to our general setting.
Theorem 1.5.
Let , , be a 1-sided NTA domain (cf. Definition 2.5) satisfying the capacity density condition (cf. Definition 2.10). Let be a real (non-necessarily symmetric) elliptic operator and write to denote the associated elliptic measure and the Green function. For every , there exists depending only on dimension , the 1-sided NTA constants, the CDC constant, the ellipticity constant of , and , such that for every with in the weak-sense in , for every , there holds
(1.6) |
where and are the localized dyadic conical square function and non-tangential maximal function respectively (cf. (2.28) and (2.27)), and is a corkscrew point relative to (see Section 2.4).
We note that the estimate (1.6) is written for the localized dyadic conical square function and non-tangential maximal function. It is not difficult to see that, as a consequence, one can obtain a similar estimate for the regular localized (or truncated) conical square function and non-tangential maximal function with arbitrary apertures (see [4, Lemma 4.8]), precise statements are left to the interested reader.
The plan of this paper is as follows. Section 2 presents some of the preliminaries, definitions, and tools which will be used throughout the paper. Section 3 contains a dyadic version of the main lemma of [10]. In Section 4 we prove our main results, Theorem 1.3 and Theorem 1.5.
We would like to mention that after an initial version of this work was posted on arXiv [1], Feneuil and Poggi in [13] obtained results related to ours, compare for instance Theorem 1.3 with [13, Theorem 1.27]. Also, the recent work [4] complements this paper and its companion [2], see for instance [4, Corollary 1.4].
2. Preliminaries
2.1. Notation and conventions
-
We use the letters to denote harmless positive constants, not necessarily the same at each occurrence, which depend only on dimension and the constants appearing in the hypotheses of the theorems (which we refer to as the “allowable parameters”). We shall also sometimes write and to mean, respectively, that and , where the constants and are as above, unless explicitly noted to the contrary. Unless otherwise specified upper case constants are greater than and lower case constants are smaller than . In some occasions it is important to keep track of the dependence on a given parameter , in that case we write or to emphasize that the implicit constants in the inequalities depend on .
-
Our ambient space is , .
-
Given we write to denote its diameter.
-
Given a domain , we shall use lower case letters , etc., to denote points on , and capital letters , etc., to denote generic points in (especially those in ).
-
The open -dimensional Euclidean ball of radius will be denoted when the center lies on , or when the center . A surface ball is denoted , and unless otherwise specified it is implicitly assumed that .
-
If is bounded, it is always understood (unless otherwise specified) that all surface balls have radii controlled by the diameter of , that is, if then . Note that in this way if .
-
For , we set .
-
We let denote the -dimensional Hausdorff measure.
-
For a Borel set , we let denote the usual indicator function of , i.e. if , and if .
-
We shall use the letter (and sometimes ) to denote a closed -dimensional Euclidean cube with sides parallel to the coordinate axes, and we let denote the side length of . We use to denote dyadic “cubes” on or . The latter exist as a consequence of Lemma 2.13 below.
2.2. Some definitions
Definition 2.1 (Corkscrew condition).
Following [25], we say that an open set satisfies the Corkscrew condition if for some uniform constant and for every and , if we write , there is a ball . The point is called a Corkscrew point relative to (or, relative to ). We note that we may allow for any fixed , simply by adjusting the constant . We say that satisfies the exterior Corkscrew condition if satisfies the Corkscrew condition.
Definition 2.2 (Harnack Chain condition).
Again following [25], we say that satisfies the Harnack Chain condition if there are uniform constants such that for every pair of points there is a chain of balls with where
(2.3) |
such that , , and for every
(2.4) |
The chain of balls is called a Harnack Chain.
We note that in the context of the previous definition if we can trivially form the Harnack chain and where (2.4) holds with . Hence the Harnack chain condition is non-trivial only when .
Definition 2.5 (1-sided NTA and NTA).
We say that a domain is a 1-sided non-tangentially accessible domain (1-sided NTA) if it satisfies both the Corkscrew and Harnack Chain conditions. Furthermore, we say that is a non-tangentially accessible domain (NTA domain) if it is a 1-sided NTA domain and if, in addition, also satisfies the Corkscrew condition.
Remark 2.6.
In the literature, 1-sided NTA domains are also called uniform domains. We remark that the 1-sided NTA condition is a quantitative form of path connectedness.
Definition 2.7 (Ahlfors regular).
We say that a closed set is -dimensional Ahlfors regular (AR for short) if there is some uniform constant such that
(2.8) |
Definition 2.9 (1-sided CAD and CAD).
A 1-sided chord-arc domain (1-sided CAD) is a 1-sided NTA domain with AR boundary. A chord-arc domain (CAD) is an NTA domain with AR boundary.
We next recall the definition of the capacity of a set. Given an open set (where we recall that we always assume that ) and a compact set we define the capacity of relative to as
Definition 2.10 (Capacity density condition).
An open set is said to satisfy the capacity density condition (CDC for short) if there exists a uniform constant such that
(2.11) |
for all and .
The CDC is also known as the uniform 2-fatness as studied by Lewis in [26]. Using [15, Example 2.12] one has that
(2.12) |
and hence the CDC is a quantitative version of the Wiener regularity, in particular every is Wiener regular. It is easy to see that the exterior Corkscrew condition implies CDC. Also, it was proved in [30, Section 3] and [16, Lemma 3.27] that a set with Ahlfors regular boundary satisfies the capacity density condition with constant depending only on and the Ahlfors regular constant.
2.3. Existence of a dyadic grid
In this section we introduce a dyadic grid along the lines of that obtained in [7]. More precisely, we will use the dyadic structure from [23, 24], with a modification from [22, Proof of Proposition 2.12]:
Lemma 2.13 (Existence and properties of the “dyadic grid”).
Let be a closed set. Then there exists a constant depending just on such that for each there is a collection of Borel sets (called “cubes”)
where denotes some (possibly finite) index set depending on satisfying:
-
for each .
-
If then either or .
-
For each , , and , there is a unique such that .
-
For each , there is such that
Proof.
We first note that is geometric doubling. That is, there exists depending just on such that for every and one can cover the surface ball with at most surface balls of the form with —observe that geometric doubling for is inherited from the corresponding property on and that is why depends only on and it is independent of . Besides, letting , for every it is easy to find a countable collection such that
Invoking then [23, 24] on with the Euclidean distance and one can construct a family of dyadic cubes associated with these families of points, say for . These satisfy – in the statement with the only difference that we have to replace by in .
At this point we follow the argument in [22, Proof of Proposition 2.12] with . For any we set for every . It is straightforward to show that properties , and for the families follow at once from those for the families . Regarding , let and let such that so that . Writing for the corresponding point associated with and invoking for we conclude
hence holds. ∎
A few remarks are in order concerning this lemma. Note that by construction, within the same generation (that is, within each ) the cubes are pairwise disjoint (hence, there are no repetitions). On the other hand, repetitions are allowed in the different generations, that is, one could have that and agree. Then, although and are the same set, as cubes we understand that they are different. In short, it is then understood that is an indexed collection of sets where repetitions of sets are allowed in the different generations but not within the same generation. With this in mind, we can give a proper definition of the “length” of a cube (this concept has no geometric meaning for the moment). For every , we set , which is called the “length” of . Note that the “length” is well defined when considered on , but it is not well-defined on the family of sets induced by . It is important to observe that the “length” refers to the way the cubes are organized in the dyadic grid and in general may not have a geometrical meaning. It is clear from that (we will see below that in our setting the converse hold, see Remark 2.56).
Let us observe that the generations run for all . However, as we are about to see, sometimes it is natural to truncate the generations. If is bounded and is such that , then there cannot be two distinct cubes in . Thus, with . Therefore, we are going to ignore those such that . Hence, we shall denote by the collection of all relevant , i.e., , where, if is finite, the union runs over those such that .
In what follows given with we will denote . We write , with being the constant in Lemma 2.13, which is a purely dimensional. For we will set if . Property implies that for each cube , there exist and , with (indeed ), such that
(2.14) |
We shall denote these balls and surface balls by
(2.15) |
(2.16) |
and we shall refer to the point as the “center” of .
Let and consider the family of its dyadic children . Note that for any two distinct children , one has , otherwise , contradicting the fact that and are disjoint. Also , hence by the geometric doubling property we have a purely dimensional bound for the number of such and hence the number of dyadic children of a given dyadic cube is uniformly bounded.
Lemma 2.17.
Let be a closed set and let be the dyadic grid as in Lemma 2.13. Assume that there is a Borel measure which is doubling, that is, there exists such that for every and . Then for every . Moreover, there exist , , and depending only on dimension and such that for every and
(2.18) |
Proof.
The argument is a refinement of that in [17, Proposition 6.3] (see also [14, p. 403] where the Euclidean case was treated). Fix an integer , a cube , and a positive integer to be chosen. Fix small enough to be chosen and write
We set
and make the disjoint decomposition We then split , where if meets , and otherwise. We then write , where
and for each cube , we construct as follows. We enumerate the elements in as , and then set and
so that covers and the modified cubes are pairwise disjoint.
We also note from (2.14) that if then
Then misses provided . Otherwise, we can find with . The latter implies that there is . All these yield a contradiction:
Consequently, by the doubling property,
Since and are disjoint, the latter estimate yields
where we note that .
Let us now repeat this procedure, decomposing for each . We set and split it into and where if meets . Associated to any we set . Then we make these sets disjoint as before and we have that is defined as the disjoint union of the corresponding . Note that and this is a disjoint union. As before, misses provided so that by the doubling property
and then Next we set and as the union of the corresponding and with . Then,
Iterating this procedure we obtain that for every , if then . Let us see that this leads to the desired estimates. Fix and find such that . By construction and then
which easily gives (2.18) with and . On the other hand, note that
also . Thus clearly,
yielding that . ∎
Remark 2.19.
Note that the previous argument is local in the sense that if we just want to obtain the desired estimates for a fixed we would only need to assume that is doubling in . Indeed we would just need to know that for every and , and the involved constants in the resulting estimates will depend only on dimension and . Further details are left to the interested reader.
We next introduce the “discretized Carleson region” relative to , . Let be a family of pairwise disjoint cubes. The “global discretized sawtooth” relative to is the collection of cubes that are not contained in any , that is,
For a given , the “local discretized sawtooth” relative to is the collection of cubes in that are not contained in any or, equivalently,
We also allow to be the null set in which case and .
With a slight abuse of notation, let be either , and in that case , or a fixed cube in , hence is the family of dyadic subcubes of . Let be a non-negative Borel measure on so that for every . For the rest of the section we will be working with which is dyadically doubling in . This means that there exists such that for every with .
Definition 2.20 ().
Given and , a non-negative dyadically doubling measure in , a non-negative Borel measure defined on is said to belong to if there exist constants such that for every and for every Borel set , we have that
(2.21) |
It is well known (see [8, 14]) that since is a dyadically doubling measure in , if and only if in and there exists such that , that is, there is a constant such that
for every , and where is the Radon-Nikodym derivative.
For each , a family of pairwise disjoint dyadic cubes, and each , we define the projection operator
If is a non-negative Borel measure on , we may naturally then define the measure as , that is,
(2.22) |
for each Borel set .
2.4. Sawtooth domains
In the sequel, , , will be a 1-sided NTA domain satisfying the CDC. Write for the dyadic grid obtained from Lemma 2.13 with . In Remark 2.56 below we shall show that under the present assumptions one has that for every surface ball . In particular for every in view of (2.14). Given we define the “Corkscrew point relative to ” as . We note that
As done above, given and a possibly empty family of pairwise disjoint dyadic cubes, we can define , the “discretized Carleson region”; , the “global discretized sawtooth” relative to ; and , the “local discretized sawtooth” relative to . Note that if to be the null set in which case and . We would like to introduce the “geometric” Carleson regions and sawtooths.
Let denote a collection of (closed) dyadic Whitney cubes of , so that the cubes in form a covering of with non-overlapping interiors, and satisfy
(2.23) |
and
Let denote the center of , let denote the side length of , and write if .
Given and we write for the “fattening” of . By taking small enough, we can arrange matters, so that, first, for every . Secondly, meets if and only if meets (the fattening thus ensures overlap of and for any pair whose boundaries touch, so that the Harnack Chain property then holds locally in , with constants depending upon ). By picking sufficiently small, say , we may also suppose that there is such that for distinct , we have that . In what follows we will need to work with dilations or , and in order to ensure that the same properties hold we further assume that .
For every we can construct a family , and define
satisfying the following properties: and there are uniform constants and such that
(2.24) |
Here, means that the interior of contains all balls in a Harnack Chain (in ) connecting to , and moreover, for any point contained in any ball in the Harnack Chain, we have with uniform control of the implicit constants. The constants and the implicit constants in the condition , depend on the allowable parameters and on . Moreover, given we have that , where satisfies , and contains any fixed such that . The reader is referred to [17, 20] for full details.
For a given , the “Carleson box” relative to is defined by
For a given family of pairwise disjoint cubes and a given , we define the “local sawtooth region” relative to by
(2.25) |
where . Note that in the previous definition we may allow to be empty in which case clearly . Similarly, the “global sawtooth region” relative to is defined as
(2.26) |
where . If is the empty set clearly . For a given and let us introduce the “truncated dyadic cone”
where it is understood that if . Analogously, we can slightly fatten the Whitney boxes and use to define new fattened Whitney regions and sawtooth domains. More precisely, for every ,
where . Similarly, we can define , , , and by using in place of .
Given we next define the “localized dyadic non-tangential maximal function”
(2.27) |
for every , where it is understood that for every (since in such a case). Finally, let us introduce the “localized dyadic conical square function”
(2.28) |
for every . Note that again for every .
To define the “Carleson box” associated with a surface ball , let denote the unique such that , and set
(2.29) |
We then define
(2.30) |
We can also consider fattened versions of given by
Following [17, 20], one can easily see that there exist constants and (with the constant in (2.14)), depending only on the allowable parameters, so that
(2.31) | |||
(2.32) |
and also
(2.33) |
where is defined as in (2.15), with , , and is so that . From our choice of the parameters one also has that whenever .
In the remainder of this section we show that if is a 1-sided NTA domain satisfying the CDC then Carleson boxes and local and global sawtooth domains are also 1-sided NTA domains satisfying the CDC. We next present some of the properties of the capacity which will be used in our proofs. From the definition of capacity one can easily see that given a ball and compact sets then
(2.34) |
Also, given two balls and a compact set then
(2.35) |
On the other hand, [15, Lemma 2.16] gives that if is a compact with then there is a dimensional constant such that
(2.36) |
Proposition 2.37.
Let , , be a 1-sided NTA domain satisfying the CDC. Then all of its Carleson boxes and , and sawtooth regions , and are 1-sided NTA domains and satisfy the CDC with uniform implicit constants depending only on dimension and on the corresponding constants for .
Proof.
A careful examination of the proofs in [17, Appendices A.1-A.2] reveals that if is a 1-sided NTA domain then all Carleson boxes and , and local and global sawtooth domains and inherit the interior Corkscrew and Harnack chain conditions, hence they are also 1-sided NTA domains. Therefore, we only need to prove the CDC. We are going to consider only the case (which in particular gives the desired property for by allowing to be the null set). The other proofs require minimal changes which are left to the interested reader. To this end, fix and a (possibly empty) family of pairwise disjoint dyadic cubes. Let and .
Case 1: . In that case we have that and we can use that satisfies the CDC with constant , (2.34) and the fact that to obtain the desired estimate
Case 2: with large enough to be chosen. In this case and hence there exist and such that . Note that by (2.24)
Let be such that and provided that is taken large enough. If then taking large enough
and . On the other hand, if , we analogously have provided is large enough
and thus . Once has been fixed so that the previous estimates hold, we use them in conjunction with the fact that satisfies the CDC with constant , (2.34)–(2.36), and that to obtain
which gives us the desired lower bound in the present case.
Case 3: . In this case and hence there exists and such that and . Also there exists , with such that for any which implies that for some (see Section 2.4). Note that , and more precisely by (2.23).
Let with and being the point in the segment joining and the center of at distance from . It is easy to see that and also . We can then use (2.12) and (2.34)–(2.36) to obtain the desired estimate:
Collecting the 3 cases and using (2.12) we have been able to show that
(2.38) |
which eventually gives that satisfies the CDC. This completes the proof. ∎
Our next auxiliary result adapts [21, Lemma 4.44] to our current setting:
Lemma 2.39.
Let be a 1-sided NTA domain satisfying the CDC. Given and consider the family of pairwise disjoint cubes and let and . There exists and a constant depending only on dimension , the 1-sided NTA constants, the CDC constant, and independent of and such that the following hold:
-
.
-
.
-
Setting
(2.40)
one has
(2.41) |
and there exists a family so that
(2.42) |
Proof.
We proceed as in [21, Lemma 4.44]. Recall that given , any closed dyadic cube in , we set and . Let us introduce so that
(2.43) |
Given , fix such that and (the implicit constant depends on the parameter ). For every we set so that , and (with implicit constant depending only on and ).
For every , we let . It then follows that since for every compact subset of , the previous sum has finitely many non-vanishing terms. Also, for every since the family has bounded overlap by our choice of . Hence we can set and one can easily see that , and . With this in hand set
We first note that the number of terms in the sum defining is bounded depending on . Indeed, if then and which implies that has finite cardinality with bounds depending only on dimension and (here we recall that the number of dyadic children of a given cube is uniformly controlled). Also, by construction has cardinality depending only on the allowable parameters. Hence, . This and the fact that each yield that . Note also that (2.43) and the definition of give
This, the fact that , and the definition of immediately give that . On the other hand if then the exists such that in which case . All these imply . Note that follows by observing that for every
where we have used that if then and also that the family has bounded overlap.
To see fix and , and set so that . We first note that . Indeed, if then . Hence and our choice of gives that meets , this in turn implies that since . All these yield
Hence for every . This and the fact that immediately give that in .
We are left with showing the last part of and for that we borrow some ideas from [18, Appendix A.2]. Fix and let be so that with , in particular . Since there exists (that is, with so that ). Pick so that and it contains any fixed such that . Then, as observed in Section 2.4, one has . But, since , we necessarily have . Hence, where
For later use it is convenient to observe that
(2.44) |
Let us first consider . If we clearly have
and since
In particular, . Thus if we set it follows from (2.44) that the two first conditions in (2.42) hold and also .
Consider next . For any we also set so that from (2.44) we clearly see that the two first conditions in (2.42) hold. It then remains to estimate the overlap. With this goal in mind we first note that if , the fact that yields
hence . Suppose next that for . Then since touches and touches
Hence fixed there is a uniformly bounded number of with , and, in particular, .
We finally take into consideration the most delicate collection . In this case for every we pick so that and with large enough so that (cf. (2.14)). Note that since we have that which, together with (2.44), implies
Hence the first two conditions in (2.42) hold in the current situation.
On the other hand, the choice of and (2.14) guarantee that
(2.45) |
Also, since , it follows that and therefore . Besides, since
Thus, . Suppose next that are so that and assume without loss of generality that , hence . Then, since and we get from (2.45)
and therefore which in turn gives . Note also that since touches , touches , and we obtain
Consequently, fixed there is a uniformly bounded number of with . As a result, . This clearly completes the proof of and hence that of Lemma 2.39. ∎
2.5. Uniformly elliptic operators, elliptic measure and the Green function
Next, we recall several facts concerning elliptic measure and the Green functions. To set the stage let be an open set. Throughout we consider elliptic operators of the form with being a real (non-necessarily symmetric) matrix such that and there exists such that the following uniform ellipticity condition holds
(2.46) |
for all and for almost every . We write to denote the transpose of , or, in other words, with being the transpose matrix of .
We say that is a weak solution to in provided that satisfies
Associated with one can construct an elliptic measure and a Green function (see [20] for full details). Sometimes, in order to emphasize the dependence on , we will write and . If satisfies the CDC then it follows that all boundary points are Wiener regular and hence for a given we can define
so that satisfies on and in the weak sense in . Moreover, if is bounded and then . In the same context the Green function satisfies the following properties which will be used along the paper:
(2.47) | |||
(2.48) | |||
(2.49) | |||
(2.50) |
We first define the reverse Hölder class and the classes with respect to fixed elliptic measure in . One reason we take this approach is that we do not know whether is well-defined since we do not assume any Ahlfors regularity. Hence we have to develop these notions in terms of elliptic measures. To this end, let satisfy the CDC and let and be two real (non-necessarily symmetric) elliptic operators associated with and where and satisfy (2.46). Let and be the elliptic measures of associated with the operators and respectively with pole at . Note that if we further assume that is connected then on for every . Hence if on for some then on for every and thus we can simply write on . In the latter case we will use the notation
(2.51) |
to denote the Radon-Nikodym derivative of with respect to , which is a well-defined function -almost everywhere on .
Definition 2.52 (Reverse Hölder and classes).
Fix where with and . Given , , we say that , provided that on , and there exists such that
for every where , with , . The infimum of the constants as above is denoted by .
Similarly, we say that provided that for every with and one has uniformly on , that is,
Finally,
The following result lists a number of properties which will be used throughout the paper, proofs may be found in [20]:
Lemma 2.53.
Suppose that , , is a 1-sided NTA domain satisfying the CDC. Let and be two real (non-necessarily symmetric) elliptic operators, there exist , (depending only on dimension, the 1-sided NTA constants, the CDC constant, and the ellipticity of ) and (depending on the same parameters and on the ellipticity of ), such that for every with , , and we have the following properties:
-
for every and .
-
If with and is such that , then for all we have that .
-
If , then .
-
For every with as in (2.32), we have that
-
For every and for any
Remark 2.54.
We note that from in the previous result, Harnack’s inequality, and (2.14) one can easily see that
(2.55) |
Observe that since an analogous inequality for the reciprocal of the Radon-Nikodym derivative follows immediately.
Remark 2.56.
Given , a 1-sided NTA domain satisfying the CDC, we claim that if with and then . To see this we first observe that . If — is the Corkscrew constant— then clearly . Hence, we may assume that . Let so that and note that one can easily see that . Associated with and we can consider and the corresponding Corkscrew points. These are different, despite the fact that . Indeed,
which yields that . Note that since otherwise we would get a contradiction: . Hence we can invoke Lemma 2.53 parts and and (2.47) to see that
This and the fact that easily yields that as desired.
We close this section by establishing an estimate for the non-tangential maximal function for elliptic-measure solutions.
Proposition 2.57.
Let be a 1-sided NTA domain satisfying the CDC. Given and with let
Then for every ,
(2.58) |
and, as a consequence, for every
(2.59) |
Moreover, the implicit constants depend just on dimension , the 1-sided NTA constants, the CDC constant, and the ellipticity constant of and on in (2.59).
Proof.
By decomposing into its positive and negative parts we may assume that is non-negative with and construct the associated as in the statement which is non-negative. Fix and let . Then, by definition there are and such that and . Hence using Harnack’s inequality and the notation introduced in (2.14)–(2.16)
Let be such that . Observe that for every one has that . Otherwise there is and hence we get a contradiction:
With this in hand, and since , we clearly see that for .
In order to estimate the ’s we need some preparatives. Note that for every one has . We claim that
(2.60) |
Indeed, this estimate follows from Harnack’s inequality and Lemma 2.53 part when since , and from Lemma 2.53 part whenever . We also observe that Lemma 2.53 part and Harnack’s inequality readily give that
(2.61) |
Finally, by Lemma 2.53 part and Harnack’s inequality it follows that
(2.62) |
3. Dyadic sawtooth lemma for projections
In this section, we shall prove two dyadic versions of the main lemma in [10]. To set the stage we sate a result which is partially proved in [17, Proposition 6.7] under the further assumption that is Ahlfors regular
Proposition 3.1.
Let , , be a 1-sided NTA domain satisfying the CDC. Fix and let be a family of pairwise disjoint dyadic cubes. There exists so that
(3.2) |
where the implicit constants depend only on dimension, the 1-sided NTA constants, the CDC constant, and is independent of and . Additionally, for each , there is an -dimensional cube , which is contained in a face of for some , and which satisfies
(3.3) |
and
(3.4) |
where the implicit constants depend on allowable parameters.
Proof.
Note first that is a 1-sided NTA domain satisfying the CDC (see Proposition 2.37). Pick an arbitrary and let be a Corkscrew point relative to the surface ball for the bounded domain (recall that one has by (2.31)). Note that , which is comprised of fattened Whitney boxes, then for some , with . Let be the center of so that . Then,
To continue we note that the existence of the family so that (3.3) holds has been proved in [17, Proposition 6.7] under the further assumption that is Ahlfors regular. However, a careful examination of the proof shows that the same argument applies in our scenario. We are left with showing (3.4). To see this, observe that as in [17, Remark 6.9] if then . Indeed from the previous result and for some . Thus meets and by construction and meet. Using (3.3) and the nature of the Whitney cubes we see that . Using this and (3.3) one can also see that . Hence, fixing and we have some constant (depending on the allowable parameters) such that
To estimate each of the terms in the last sum fix and note that since the cubes belong to the same generation then ’s involved are disjoint and hence so they are the corresponding ’s which all have radius . In particular, for any such cubes and . Moreover,
Thus it is easy to see (since is geometric doubling) that . All these together gives us desired (3.4) —we note in passing that the argument in [17, Remark 6.9] used the fact there is AR to estimate each , while here we are invoking the geometric doubling property of the ambient space . ∎
We are now ready to state the first main result of this chapter which is a version of [17, Lemma 6.15] (see also [10]) valid in our setting:
Lemma 3.5 (Discrete sawtooth lemma for projections).
Suppose that , , is a bounded 1-sided NTA domain satisfying the CDC. Let , let be a family of pairwise disjoint dyadic cubes, and let be a dyadically doubling measure in . Given two real (non-necessarily symmetric) elliptic , , we write , for the elliptic measures associated with and for the domain with fixed pole at (cf. Lemma 3.1). Let be the elliptic measure associated with for the domain with fixed pole at . Consider the measure defined by
(3.6) |
where is the cube produced in Proposition 3.1. Then (see (2.22)) depends only on and , but not on . More precisely,
(3.7) |
Moreover, there exists such that for all and all , we have
(3.8) |
Proof.
Our argument follows the ideas from [17, Lemma 6.15] and we use several auxiliary technical results from [17, Section 6] which were proved under the additional assumption that is AR. However, as we will indicate along the proof, most of them can be adapted to our setting. Those arguments that require new ideas will be explained in detail.
We first observe that (3.7) readily follows from the definitions of and . We first establish the second estimate in (3.8). With this goal in mind let us fix and .
Case 1: There exists such that . By (3.7) we have
Case 2: for any , that is, . In particular if with then necessarily . Let denote the center of and pick so that . Note that by Proposition 2.37, Harnack’s inequality and Lemma 2.53 parts and we have that . On the other hand as in [17, Proposition 6.12] one can see that
(3.9) |
with , and with implicit constants depending on the allowable parameters. We note that the last expression is slightly different to that in [17, Proposition 6.2], nonetheless the one stated here follows from the proof in account of [17, (6.14) and Proposition 6.1] as is contained in . Besides, Proposition 3.1 easily yields
(3.10) |
hence
(3.11) |
Since we can invoke [17, Proposition 6.4] (which also holds in the current setting) to find which serves as a Corkscrew point simultaneously for with respect to the surface ball for some and some , and for with respect to each surface ball , for every . Applying (2.55) and Harnack’s inequality to join with and with we have
(3.14) |
On the other hand one can see that
(3.15) |
for some . Invoking then Proposition 2.37, and Lemma 2.53 parts and in the domain we can analogously see
(3.16) |
Next we invoke (3.12), (3.15), and (3.14) to obtain
(3.17) |
We claim the following estimates hold
(3.18) |
The first estimate follows easily from the maximum principle since and . For the second one, by the maximum principle we just need to see that for , but this follows from Lemma 2.53 part , (2.14), Harnack’s inequality, and (3.3).
As a consequence of the previous result we can easily obtain a dyadic analog of the main lemma in [10].
Lemma 3.19 (Discrete sawtooth lemma).
Suppose that , , is a bounded 1-sided NTA domain satisfying the CDC. Let and let be a family of pairwise disjoint dyadic cubes. Given two real (non-necessarily symmetric) elliptic , , we write , for the elliptic measures associated with and for the domain with fixed pole at (cf. Lemma 3.1). Let be the elliptic measure associated with for the domain with fixed pole at . Consider the measure defined by (3.6). Then, there exists such that for all and all , we have
(3.20) |
In particular, if , we have
(3.21) |
where with , , and with implicit constants depending on the allowable parameters (cf. [17, Proposition 6.12]).
Proof.
Letting , which is dyadically doubling in , one easily has and . Thus (3.8) in Lemma 3.5 readily yields (3.20). Next, to obtain (3.21) we may assume that is non-empty. Observe that if , then . On the other hand, if we must be in Case 2 of the proof of Lemma 3.5, hence (3.12) and (3.13) hold. With all these we readily obtain (3.21). ∎
Our last result in this section establishes that both and are dyadically doubling on .
Lemma 3.22.
Under the assumptions of Lemma 3.5, and are dyadically doubling on .
Proof.
We follow the ideas in [17, Lemma B.2]. We shall first see is dyadically doubling. To this end, let be fixed and let be one of its dyadic children. We consider three cases:
Case 1: There exists such that . In this case we have
where we have used Harnack’s inequality and Lemma 2.53 parts and and .
Case 2: . For simplicity say and in this case . Note that then and we let be the family of cubes with and observe that if then . Then by (3.4)
(3.23) |
Recall that in Case 2 in the proof of Lemma 3.5 we mentioned that with being the center of and since is the dyadic parent of . Note that since by (3.3)
Thus
where we here and below we use the notation for the surface balls with respect to . Using this, (3.23), and Lemma 2.53 parts and and Harnack’s inequality we derive
Case 3: None of the conditions in the previous cases happen, and necessarily . We take the same set as in the previous case and again if then (otherwise we are driven to Case 1). Introduce , the family of cubes with . Again, if we have ; otherwise either which is Case 2, or which implies and we are back to Case 1.
Note that since is the dyadic parent of , using the same notation as in (3.9) applied to we have that
Also by (3.3)
These readily give
We can then proceed as in the previous case (see (3.23)) to obtain
where (see (3.9)) and we have used Lemma 2.53 parts and and Harnack’s inequality. On the other hand, proceeding as in (3.12) with in place of since :
Eventually we obtain that , completing the proof of the dyadic doubling property of .
We next deal with . We can simply follow the previous argument replacing by to see that in Cases 2 and 3 we have that and , hence the doubling condition follows from the previous calculations and the constant depend on that of . With regard to Cases 1, on which for some , one can easily see that
which uses that is dyadically doubling in . Eventually we have seen that doubling constant depend on that of and as desired. This completes the proof. ∎
4. Proof of the main results
4.1. Proof of Theorem 1.3
By renormalization we may assume without loss of generality that . We will first prove a dyadic version of (1.4). Let the dyadic grid from Lemma 2.13 with . Our goal is to show that
(4.1) |
with large enough. Assuming this momentarily let us see how to derive (1.4). Fix and as in the suprema in (1.4). Let be so that and , and define where is large enough to be chosen depending on and the allowable parameters. Set
Note that for every with we have
As a consequence, if , then , and picking one has
This gives and .
To proceed, let us write
and we estimate each term in turn.
To estimate we may assume that , hence , and . Then Lemma 2.53, the fact that , Caccioppoli’s inequality, the normalization , and Harnack’s inequality give
Next we deal with . Introduce the disjoint family . Given , let , and be so that and it contains some fixed such that . Then, as observed in Section 2.4, one has . Note that
hence . This and the fact that, as observed before, imply that for some . Hence, for some . This eventually show that and therefore
For any pick the unique (ancestor) with and . Note that , . Also,
Hence by the Harnack chain condition one obtains that for every (in doing that we need to make sure that is large enough so that the Harnack chain joining and , which is -away from , does not get near , which is -close to ). Note also that , provided is large enough depending on . All theres and (4.1) yield
where we have used Lemma 2.53. This completes the prof of the fact that (4.1) implies (1.4).
We next focus on showing (4.1). With this goal in mind we fix and let with with large enough so that (cf. (2.31)). Write and and note that our choice of , (2.49), and (2.50) guarantee that in the weak sense in .
Fix and consider the family of pairwise disjoint cubes and let (cf. (2.25)). Note that by construction is an increasing sequence of sets converging to . Our goal is to show that for every there holds
(4.2) |
with independent of , , and . Hence the monotone convergence theorem yields
which is (4.1).
Let us next start estimating (4.2). Using from Lemma 2.39 and the ellipticity of the matrix we have
We observe that and belong to since , , for every , the properties of , and the fact that is away from — and by (2.31) one has for every and provided is large enough. Using all these one can easily see via a limiting argument that the fact that in the weak sense in implies that . Likewise, one can easily show that by recalling that (see (2.31)) and that as mentioned above in the weak sense in . Thus we are left with estimating the terms and . By in Lemma 2.39 and the fact that we obtain
where denotes the center of , and we have used Harnack’s and Caccioppoli’s inequalities, that and in the weak sense in (see (2.31)). Invoking Lemmas 2.53 and Lemma 2.39 one can see that for every . This together with Lemma 2.39 allows us to conclude
Note that if with one has
where we have used (2.42) and (2.31). Thus, Lemma 2.53 gives
This allows us to complete the proof of Theorem 1.3. ∎
4.2. Proof of Theorem 1.5
We borrow some ideas from [19]. Given introduce the truncated localized conical square function: for every and , let
where if it is understood that and . Note that by the monotone convergence theorem as for every .
Fixed large enough (eventually, ), our goal is to show that we can find (independent of ) such that for every we have
(4.3) |
where the implicit constant depend on the allowable parameters and it is independent of . To prove this we fix and set
Consider first the case . Note that if , by definition . Let be the unique dyadic cube such that and . Then it is clear from construction that for every one has
Hence, and we have shown that for every there exists such that and . We then take the ancestors of , and look for the one with maximal side length which is contained in . That is, for every and where is the dyadic parent of (during this proof we will use to denote the dyadic parent of , that is, the only dyadic cube containing it with double side length). Note that the assumption guarantees that . Let be the collection of such maximal cubes as runs in and we clearly have that the family is pairwise disjoint and also . Also, by construction and by the maximality of each we can select .
On the other hand, for any we have, using that ,
and therefore
As a consequence,
and
This has been done under the assumption that . In the case we set . Then in both cases we obtain
(4.4) |
Thus, to obtain (4.3) it suffices to see that for every
(4.5) |
From this we just need to sum in to see that (4.4) together with the previous facts yield the desired estimate (4.3):
Let us then obtain (4.5). Fix and to ease the notation write . Set
(4.6) |
If then (4.5) is trivial, hence we may assume that so that . We subdivide dyadically and stop the first time that . If one never stops we write , otherwise is the family of stopping cubes which is maximal (hence pairwise disjoint) with respect to the property . In particular, .
Next we claim that
(4.7) |
To verify the first inclusion, we fix with . Then, where . Since we must have (otherwise for some and this would imply that ) and therefore which gives the first inclusion. The second inclusion in (4.7) is trivial (since ).
To continue we see that
(4.8) |
Fix such a so that for some . If , by maximality of the cubes in , it follows that for some , which contradicts the fact . Thus, and we can select so that by definition since .
Apply Lemma 3.1 to find so that
(4.9) |
Let be the elliptic measure associated with relative to with pole at and write . Given , we choose such that . By definition, for and , there is a such that and . Thus, by the triangle inequality, and the definition of , we have that for ,
(4.10) |
where in the last step we have used that
(4.11) |
On the other hand, as observed above , see [17, Proposition 6.1]. Using this and the fact that if then we have
(4.12) | ||||
where is a large constant to be chosen.
We start estimating . Note first that , thus
where we have used (4.8), along with the fact that for any with , and the fact that has uniformly bounded cardinality. To estimate we note that picking we have that . Write for Corkscrew relative to with respect to so that . Note that by (4.9), we clearly have provided is sufficiently large. Hence, by Lemma 2.53 part applied in , which is a 1-sided NTA domain satisfying the CDC by Proposition 2.37, we obtain for every
(4.13) |
where is the Green function for the operator relative to the domain . Above the last estimate uses Harnack’s inequality (we may need to tale slightly larger) and the fact that by (4.11), one has (see Remark 2.56) and that if with
Write for the collection of dyadic cubes with which has uniformly bounded cardinality depending on . Note that
For each , if then and hence . Pick then and note that for every by (2.31) it follows that
Using then (4.13) we have
where we have invoked Theorem 1.3 applied in , which is a 1-sided NTA domain satisfying the CDC by Proposition 2.37, and we may need to take slightly larger and use Harnack’s inequality; (4.8); and the fact that has uniformly bounded cardinality.
Using Chebyshev’s inequality, (4.12), and collecting the estimates for and we conclude that
At this point we invoke Lemma 3.19 in with —we warn the reader that and play the role of and and that associated to each one finds as in Proposition 3.1, which now plays the role of in that result, and (recall that ) and observe that the fact that implies on account of (3.21) that for some we have
where we have used that since with , , (4.9), Harnack’s inequality, and Lemma 2.53 part . We can then use Remark 2.54, Harnack’s inequality, and (4.9), to conclude that
Recalling that , and the definitions of and in (4.6) the previous estimates readily lead to (4.5).
To conclude we need to see how (4.3) yields (1.6). With this goal in mind we first observe that for every and one has that (see (2.31)) and also . Hence, since , one has
(4.14) |
On the other hand, given , we can use (4.3)
We can then choose small enough so that we can hide the first term in the right hand side of the last quantity (which is finite by (4.14)) and eventually conclude that
Since the implicit constant does not depend on and as for every , the monotone convergence theorem yields at once (1.6) and the proof Theorem 1.5 is complete.
References
- [1] M. Akman, S. Hofmann, J. M. Martell, and T. Toro. Perturbation of elliptic operators in 1-sided NTA domains satisfying the capacity density condition. https://arxiv.org/abs/1901.08261v2. Preprint, 2019.
- [2] M. Akman, S. Hofmann, J. M. Martell, and T. Toro. Perturbation of elliptic operators in 1-sided NTA domains satisfying the capacity density condition. https://arxiv.org/abs/1901.08261v3. Preprint, 2021.
- [3] L. A. Caffarelli, E. B. Fabes, and C. E. Kenig. Completely singular elliptic-harmonic measures. Indiana Univ. Math. J., 30(6):917–924, 1981.
- [4] M. Cao, Ó. Dominguéz, J. M. Martell, and P. Tradacete. On the condition for elliptic operators in 1-sided NTA domains satisfying the capacity density condition. https://arxiv.org/abs/2101.06064. Preprint, 2021.
- [5] J. Cavero, S. Hofmann, and J. M. Martell. Perturbations of elliptic operators in 1-sided chord-arc domains. Part I: Small and large perturbation for symmetric operators. Trans. Amer. Math. Soc., 371(4):2797–2835, 2019.
- [6] J. Cavero, S. Hofmann, J. M. Martell, and T. Toro. Perturbations of elliptic operators in 1-sided chord-arc domains. Part II: non-symmetric operators and Carleson measure estimates. Trans. Amer. Math. Soc., 373(11):7901–7935, 2020.
- [7] M. Christ. A theorem with remarks on analytic capacity and the Cauchy integral. Colloq. Math., 60/61(2):601–628, 1990.
- [8] R. R. Coifman and C. Fefferman. Weighted norm inequalities for maximal functions and singular integrals. Studia Math., 51:241–250, 1974.
- [9] B. E. J. Dahlberg. On the absolute continuity of elliptic measures. Amer. J. Math., 108(5):1119–1138, 1986.
- [10] B. E. J. Dahlberg, D. S. Jerison, and C. E. Kenig. Area integral estimates for elliptic differential operators with nonsmooth coefficients. Ark. Mat., 22(1):97–108, 1984.
- [11] R. Fefferman. A criterion for the absolute continuity of the harmonic measure associated with an elliptic operator. J. Amer. Math. Soc., 2(1):127–135, 1989.
- [12] R. A. Fefferman, C. E. Kenig, and J. Pipher. The theory of weights and the Dirichlet problem for elliptic equations. Ann. of Math. (2), 134(1):65–124, 1991.
- [13] J. Feneuil and B. Poggi. Generalized Carleson perturbations of elliptic operators and applications. https://arxiv.org/abs/2011.06574. Preprint, 2021.
- [14] J. García-Cuerva and J. L. Rubio de Francia. Weighted norm inequalities and related topics, volume 116 of North-Holland Mathematics Studies. North-Holland Publishing Co., Amsterdam, 1985. Notas de Matemática [Mathematical Notes], 104.
- [15] J. Heinonen, T. Kilpeläinen, and O. Martio. Nonlinear potential theory of degenerate elliptic equations. Dover Publications, Inc., Mineola, NY, 2006. Unabridged republication of the 1993 original.
- [16] S. Hofmann, P. Le, J. M. Martell, and K. Nyström. The weak- property of harmonic and -harmonic measures implies uniform rectifiability. Anal. PDE, 10(3):513–558, 2017.
- [17] S. Hofmann and J. M. Martell. Uniform rectifiability and harmonic measure I: Uniform rectifiability implies Poisson kernels in . Ann. Sci. Éc. Norm. Supér. (4), 47(3):577–654, 2014.
- [18] S. Hofmann, J. M. Martell, and S. Mayboroda. Uniform rectifiability, Carleson measure estimates, and approximation of harmonic functions. Duke Math. J., 165(12):2331–2389, 2016.
- [19] S. Hofmann, J. M. Martell, and S. Mayboroda. Transference of scale-invariant estimates from Lipschitz to Non-tangentially accessible to Uniformly rectifiable domains. https://arxiv.org/abs/1904.13116. Preprint, 2019.
- [20] S. Hofmann, J. M. Martell, and T. Toro. General divergence form elliptic operators on domains with ADR boundaries, and on 1-sided NTA domains. Work in Progress, 2014.
- [21] S. Hofmann, J. M. Martell, and T. Toro. implies NTA for a class of variable coefficient elliptic operators. J. Differential Equations, 263(10):6147–6188, 2017.
- [22] S. Hofmann, D. Mitrea, M. Mitrea, and A. J. Morris. -square function estimates on spaces of homogeneous type and on uniformly rectifiable sets. Mem. Amer. Math. Soc., 245(1159):v+108, 2017.
- [23] T. Hytönen and A. Kairema. Systems of dyadic cubes in a doubling metric space. Colloq. Math., 126(1):1–33, 2012.
- [24] T. Hytönen and A. Kairema. What is a cube? Ann. Acad. Sci. Fenn. Math., 38(2):405–412, 2013.
- [25] D. S. Jerison and C. E. Kenig. Boundary behavior of harmonic functions in nontangentially accessible domains. Adv. in Math., 46(1):80–147, 1982.
- [26] J. L. Lewis. Uniformly fat sets. Trans. Amer. Math. Soc., 308(1):177–196, 1988.
- [27] W. Littman, G. Stampacchia, and H. F. Weinberger. Regular points for elliptic equations with discontinuous coefficients. Ann. Scuola Norm. Sup. Pisa (3), 17:43–77, 1963.
- [28] E. Milakis, J. Pipher, and T. Toro. Harmonic analysis on chord arc domains. J. Geom. Anal., 23(4):2091–2157, 2013.
- [29] L. Modica and S. Mortola. Construction of a singular elliptic-harmonic measure. Manuscripta Math., 33(1):81–98, 1980/81.
- [30] Z. Zhao. BMO solvability and condition of the elliptic measures in uniform domains. J. Geom. Anal., 28(2):866–908, 2018.