Bounded weak solutions to elliptic PDE with data in Orlicz spaces
Abstract.
A classical regularity result is that non-negative solutions to the Dirichlet problem in a bounded domain , where , , satisfy . We extend this result in three ways: we replace the Laplacian with a degenerate elliptic operator; we show that we can take the data in an Orlicz space that lies strictly between and , ; and we show that that we can replace the norm in the right-hand side by a smaller expression involving the logarithm of the “entropy bump” , generalizing a result due to Xu.
Key words and phrases:
Orlicz spaces, degenerate elliptic equations, bounded solutions, a priori estimates1991 Mathematics Subject Classification:
35B45, 35D30, 35J25, 46E301. Introduction
In this paper, we consider boundedness properties of weak (sub)solutions to the following Dirichlet problem:
(1.3) |
Throughout this paper, is a bounded domain (i.e., an open and connected subset) of with , is a non-negative definite, symmetric, measurable matrix with , is a weight (i.e., a non-negative, measurable function) on such that , and the data function is in .
When is a uniformly elliptic matrix and , it is a classical result (See Maz′ya [16, 17], Stampacchia [23] and Trudinger [25, Theorem 4.1], [12, Theorem 8.16]) that there is a constant such that if , then
for any non-negative weak subsolution of (1.3) provided . Moreover, a counter-example shows that this bound is sharp even for the Laplacian and we cannot take .
The standard proof of this result uses the classical Sobolev inequality,
valid for any , combined with Moser iteration. The restriction is naturally connected to the classical Sobolev gain factor .
The goal of this paper is to generalize this result. First we show that this estimate can be improved by replacing the space by an Orlicz space that lies strictly between and for . For brevity, we will defer many definitions to Section 2 below.
Theorem 1.4.
Let be a uniformly elliptic matrix, and let , where , . Then there exists a constant such that, given any non-negative weak subsolution of
we have the estimate
Remark 1.5.
We will prove Theorem 1.4 as a special case of a more general result for solutions of (1.3) that holds for a much larger class of matrices . We can allow to be both degenerate and singular, but must impose some restrictions on the largest and smallest eigenvalues. We encode these restrictions in two assumptions on the integrability of the largest eigenvalue and on the existence of an Sobolev inequality with gain. We state them together as a general hypothesis.
Hypothesis 1.6.
Given the matrix and the weight , assume that for some constant ,
Moreover, assume that there exist constants such that for every
(1.7) |
The first assumption, that , in Hypothesis 1.6 is necessary to prove many of the necessary properties of weak derivatives in the corresponding degenerate Sobolev space. The second assumption, that inequality (1.7) holds, in Hypothesis 1.6 reduces to the classical Sobolev inequality if is a uniformly elliptic matrix in and . It allows us to perform the necessary De Giorgi iteration.
These assumptions hold in several important special cases. If satisfies the Muckenhoupt condition,
where the supremum is taken over all balls in , and if satisfies the degenerate ellipticity condition
where , then and (1.7) holds. (See [10].) More generally, suppose that and are a pair of weights such that a.e., satisfies a doubling condition, , and there exists such that given any balls ,
and satisfies the degenerate ellipticity condition
then and we have the Sobolev inequality
Remark 1.8.
We can now state our main result, which is a generalization of Theorem 1.4 to degenerate elliptic operators. Again, for precise definitions see Section 2.
Theorem 1.9.
We originally conjectured that the exponent in Theorem 1.9 is sharp in general, but we were not able to prove this or find a counter-example. We then learned that Cianchi [4, Theorem 5], as a consequence of a more general result, showed that in the classical setting when is uniformly elliptic and , we must have that . We round out his result by giving a simple counter-example proving that it is sharp for the Laplacian.
Example 1.11.
Let , and let . Then there exists a function , where , , such that the non-negative weak solution of the Poisson equation
is unbounded.
Remark 1.12.
We conjecture that the sharp exponent in Theorem 1.9 is . However, the bound appears to be intrinsic to our proof, so either our proof needs to be refined or another approach is needed. We note that the proof in [4] relies on re-arrangement estimates and so does not readily extend to the case of degenerate operators.
Our second main result shows that inequality (1.10) can be sharpened so that the right-hand side only depends on the logarithm of the norm.
Theorem 1.13.
Theorem 1.13 generalizes the main result of Xu [26], but we note that there is a mistake in the statement of his main result. Working in the same setting as Theorem 1.4, he claims to show that
(1.15) |
where . However, a close examination of his proof shows that he only proves this when , and in fact what he proves is the analog of Theorem 1.13. It is straightforward to see that (1.15) cannot hold if ; if it did, then if we fix and the corresponding solution , then we could apply this inequality to ( large) and . Then we could take the limit as to conclude that , which is false in general.
Remark 1.16.
Our two main results are established via De Giorgi iteration on the level sets. De Giorgi’s original arguments are in [8] but more helpful descriptions are found in [2] and in [13], where De Giorgi iteration is applied in an infinitely degenerate elliptic regime. We were unable to adapt Moser iteration to work in the context of Orlicz norms, and it remains an open question whether such an approach is possible in this setting.
The remainder of the paper is organized as follows. In Section 2 we gather some preliminary results. We give a definition of Young functions and the associated Orlicz spaces, and record some useful properties. We then define weak solutions to the Dirichlet problem. This definition has to include the possibility that the matrix can be both degenerate and singular, and we give it in terms of a degenerate Sobolev space, building upon results in [6] and elsewhere. We prove a number of properties of weak derivatives in this setting; we believe these results should be useful tools for other problems. We also prove that bounded, non-negative subsolutions of (1.3) must satisfy an exponential integrability condition. This result is a key lemma for the proof of Theorem 1.13 and is modeled on a similar result due to Xu [26] in the classical setting. For completeness we include the details of the proof. In Section 3 we prove Theorem 1.9; as noted above, the proof uses a version of De Giorgi iteration adapted to the scale of Orlicz spaces. This iteration argument was gotten by a careful adaptation of an argument due Korbenko, et al. [13, Section 4.2]. In Section 4 we prove Theorem 1.13; our proof is a generalization of the argument in [26] and requires us to deal with a number of technical obstacles. Finally, in Section 5 we construct Example 1.11.
2. Preliminaries
In this section we gather some preliminary definitions and results. We begin with some notation. The constant will always denote the dimension of the underlying space . By , , etc. we will mean a constant that may change from appearance to appearance, but whose value depends only on the underlying parameters. If we want to specify this dependence, we will write, for instance, , etc. If we write , we mean that there exists a constant such that . If and , we write .
A weight will always be a non-negative, measurable function such that . Given a set , . Given a weight , is the collection of all those measurable functions for which
Orlicz spaces
Our main hypothesis on the data function in (1.3) is that it belongs to the Orlicz space . Here we gather some essential results about these spaces but we assume the reader has some familiarity with them. For complete information, we refer to [14, 20]. For a briefer summary, see [5, Chapter 5].
By a Young function we mean a function that is continuous, convex, strictly increasing, , and as . Given a Young function , define to be the Banach space of measurable functions equipped with the Luxembourg norm,
Given Young functions we can compare the associated norms by appealing to a point-wise estimate. We say that if there is a and a constant depending only on so that for . For a proof of the following result, see [14, Theorem 13.3] or [20, Section 5.1].
Lemma 2.1.
Given Young functions , if , then there exists a constant such that for every ,
Given a Young function , we define the conjugate Orlicz function, , via the pointwise formula
The pair satisfy a version of Hölder’s inequality in the scale of Orlicz spaces. If and , then and
(2.2) |
In our main results we consider Young functions of the form
(2.3) |
where . The inverse and conjugate functions associated with these Young functions are well-known: see, for instance, [5]. We have that
(2.4) | |||
(2.5) |
where the implicit constants depend on . As a consequence of Lemma 2.1 we have the following estimate which we will need below; details are straightforward and are omitted.
Lemma 2.6.
Let , and define
Then, given ,
The implicit constants depend on and , , and .
We conclude this section with an estimate for the norm of an indicator function for ; this quantity plays an essential role in our proofs of Theorems 1.9 and 1.13. This computation is well-known, but to make clear the dependence on the constants we include its short proof.
Lemma 2.7.
Degenerate Sobolev spaces and weak solutions
We now give a precise definition of weak (sub)solutions to the Dirichlet problem (1.3). This question has been explored in a number of papers by ourselves and others: see [3, 6, 7, 12, 18, 19, 21, 22]. Here we sketch the relevant details.
Given a non-negative definite, symmetric and measurable matrix function on and a weight , the solution space for the Dirichlet problem is the matrix weighted Sobolev space . This space is defined as the abstract completion (in terms of Cauchy sequences) of the space (i.e., Lipschitz functions with compact support in ) with respect to the norm
where is the Banach space of vector-valued functions on that satisfy
This norm is well defined for provided ; in particular, is well defined if the first assumption in Hypothesis 1.6 holds.
With this definition, the Sobolev space is a collection of equivalence classes of Cauchy sequences of functions. However, the spaces and are complete: for a proof that is complete, see [22] or [6] where it was proved that is complete for . Therefore, to each equivalence class in we can associate a unique pair whose norm is given by
Conversely, given a pair we will say that it is in if there exists a sequence such that converges to in .
Hereafter, we will denote by since plays the role of a weak gradient of . However, while we adopt this formal notation we want to stress that the function is not the weak gradient of in the sense of classical Sobolev spaces. For further details, [6] contains the construction of for . Additionally, unweighted constructions of are found in [7, 19] for and [18, 21] for .
We can extend the second assumption in Hypothesis 1.6 to functions in ; this follows from density of functions and we omit the proof.
Lemma 2.9.
We can now define the weak solution to the Dirichlet problem.
Definition 2.11.
A pair is said to be a weak solution of the Dirichlet problem (1.3) if
for every . The pair is said to be a non-negative weak subsolution if -a.e. and
for every non-negative .
Note that if with -a.e., then by a standard limiting argument we may use as our test function in Definition 2.11.
Properties of weak gradients
We now develop some useful properties of functions in the degenerate Sobolev space . All of these properties are well known in the classical case: see, for instance, [12]. In the degenerate case, we stress that the first assumption in Hypothesis 1.6 is critical in proving these results and throughout this subsection we assume that and a.e.
Our first result shows that weak gradients are zero almost everywhere on sets of -measure zero.
Lemma 2.13.
Let and . Then, given any set of -measure zero, we have that:
-
(1)
;
-
(2)
;
-
(3)
a.e. .
Proof.
If , then is defined a.e. in by the Rademacher-Stepanov theorem and is in . Therefore, for a.e. ,
hence,
which proves (1).
Let ; then there exists a sequence such that in . Then by the previous argument,
and so (2) holds.
Finally, (3) follows immediately from (1) and (2). ∎
Our second result shows that non-negative truncations of functions in are again in this space.
Lemma 2.14.
Let and fix . If , then .
Proof.
By the definition of there exists a sequence in such that in and in . If we pass to a subsequence, we may assume that pointwise -a.e. We will first prove that in .
Define ; then -a.e. We will show that converges to in ; this follows from the generalized dominated convergence theorem [11, p. 59] if we show that there exist non-negative functions such that and as . But we have that
Moreover, converges pointwise a.e. to , and since . Finally, since in we get that .
Now define ; then -a.e. Moreover, we have that a.e. [12, Lemma 7.6] and so -a.e. By passing to another subsequence, we assume that pointwise -a.e. We claim that they converge in as well. If this is the case, then we have shown that is Cauchy in , and the desired conclusion follows at once.
To prove convergence, note that
The first term on the right-hand side is less than which goes to as . To estimate the second term, let be the set of where does not converge to . Then , and so by Lemma 2.13,
Since , by the dominated convergence theorem we have that as ,
∎
Our next lemma proves the existence of an approximating sequence of Lipschitz functions with some additional useful properties.
Lemma 2.15.
Let with and -a.e. Then there exists a sequence such that:
-
(1)
in ;
-
(2)
-a.e. and also in ;
-
(3)
in and pointwise a.e.;
-
(4)
for each .
Proof.
By the definition of and by passing twice to a subsequence, there exists a sequence such that:
-
()
both -a.e. and also in
-
()
in and a.e.;
-
()
Now let . Since is non-negative -a.e. in , by the triangle inequality we have that
-a.e. Therefore, we have that converges to both in and pointwise -a.e.
By the Rademacher-Stepanov theorem [9], a.e. Hence, a.e. and so as . Thus a.e. and properties ()–() above hold with replaced by .
We now define the sequence of functions . Set and let be such that , is increasing, if , if , and . Define the by . Then ; moreover, a.e. and so
(2.16) |
We claim that satisfies properties (1)–(4) above. By the definition of , , so property (1) holds. Property (4) follow immediately from (2.16) and property () for the .
It remains to prove properties (2) and (3). By the choice of , for -a.e. . We also have that -a.e. Let be the set of all such that both of these hold. Then . Given there exists such that if , , and so . Thus, pointwise -a.e. Since is bounded and , by the dominated convergence theorem we also have that in . This proves (2).
To prove (3) define the set as above. For each , there exists such that for each there exists a ball where for , ; hence, for . Now let be the set of such that ; by (), . Since is an absolutely continuous measure, . Let . Then on we have that pointwise. But so by Lemma 2.13 we have that
This implies that almost everywhere on . Therefore, we have that pointwise a.e.
The next two lemmas give the product rule and chain rule associated to pairs in . The proofs are adapted from those of similar results in [19].
Lemma 2.17.
Let and let . Then we have that .
Proof.
By the definition of there exists a sequence such that in and in . But then we immediately have that
and so in .
Similarly, since a.e., we have that
Thus, in and so . ∎
Lemma 2.18.
Let with -a.e. and . Then, given any non-negative function such that , the pair .
Proof.
Let be the sequence associated with given by Lemma 2.15. Since is Lipschitz with compact support in and , . Since -a.e., the continuity of implies that -a.e. By the fundamental theorem of calculus,
whenever .
Since by assumption and property (1) of Lemma 2.15, for -a.e. , we have that -a.e.,
Since -a.e. and in , by the generalized Lebesgue dominated convergence theorem we get that in .
To show the convergence of the gradients, first note that a.e. in and so by the continuity of and property (3) in Lemma 2.15 we get that a.e. Moreover,
The right-hand term converges to both pointwise a.e. and in . Therefore, we can again apply the generalized dominated convergence theorem to get that in . We conclude that . ∎
Exponential results
In this section we give two results which are needed to prove Theorem 1.13. The first gives a solution to an auxiliary Dirichlet problem and is an application of the previous two lemmas.
Lemma 2.19.
Fix . If is a non-negative bounded weak subsolution of the Dirichlet problem
(2.22) |
then is a non-negative weak subsolution of the Dirichlet problem
(2.25) |
Proof.
Our second result gives the exponential integrability of bounded solutions to (1.3). A version of this result is proved in [26, Lemma B] for uniformly elliptic operators; a qualitative version appeared previously in [4, Example 4]. Here we adapt the proof from [26] to our more general setting.
Lemma 2.26.
Proof.
Let and be as in the hypotheses. Define and with to be chosen below. Since is bounded, by Lemma 2.18 we have that
are in . Further, we immediately have the identities , , and . If we apply the Sobolev inequality (2.10) and use as a test function in Definition 2.11 we can estimate as follows:
If we now apply Hölder’s inequality with exponents and , and then , we get | ||||
If we now fix , then we can re-arrange terms to get
(2.28) |
Therefore, again by Hölder’s inequality and by (2.28) applied twice, we have that
∎
3. Proof of Theorem 1.9
Fix and let be a non-negative weak subsolution of (1.3). We may assume without loss of generality that (equivalently, that is non-zero on a set with ); otherwise, a standard argument shows that -almost everywhere. (Cf. (3.8) below.) By Lemma 2.6, .
For each define and let . Then by Lemma 2.14, . We now estimate as follows: by the Sobolev inequality (2.10), the definition of a weak subsolution with as the test function, and Hölder’s inequality, we have that
since on . If we divide through by , we get
(3.1) |
In order to estimate the norm of the right-hand side, recall that since , , we can define the Young function
It is immediate that and so by Lemma 2.2, a change of variables in the Luxemburg norm, and Lemmas 2.1 and 2.7 we get
where is independent of and .
We now turn to our iteration argument. For all , and, for , . Hence, if we combine the above two inequalities, we get
(3.2) |
Define with to be chosen below. Our goal is to find sufficiently large so that , as this immediately implies that
which is what we want to prove. To do this, we will use an iteration argument based on De Giorgi iteration. For each set
(3.3) |
where will be chosen below, and let . The sequence increases to and by an estimate using the mean-value theorem we have that for each ,
(3.4) |
If we set , in inequality (3.2), we get
(3.5) |
for each . By the dominated convergence theorem converges to , so to complete the proof we need to prove that .
Let . We will show that as , which is equivalent to the desired limit. To do so, we will show that we can choose and such that and
(3.6) |
for all .
Fix . Since , if we take logarithms and re-arrange terms, inequality (3.5) becomes, for ,
(3.7) |
The first step is to fix by an appropriate choice of . If we argue as we did to prove (3.1) using as the test function in the definition of a weak subsolution, we get
(3.8) |
If we estimate the right-hand side using Hölder’s inequality and Lemma 2.6, we get
where the constant is independent of and . For each we have that , so by Hölder’s inequality and the above two estimates,
If we re-arrange terms, we get
where again the constant is independent of and . Now choose so that
(3.9) |
where is as in (3.2). Note that is independent of and , and the first inequality implies that .
4. Proof of Theorem 1.13
Our proof requires one technical lemma.
Lemma 4.1.
Given , there exist constants , , and such that
(4.2) |
and
(4.3) |
Proof.
We will first show that we can choose and so that (4.3) holds, and then show that we can refine our choice so that (4.2) holds as well.
Set and , where will be determined below. With this restriction on it is immediate that and lie in the specified intervals. Moreover, if we insert these values into the definition of , we get
This gives (4.3).
Remark 4.4.
In the proof of Lemma 4.1, the range of possible values for shrinks as the dimension increases. In the classical case, , and this value is generally a lower bound on in the more degenerate settings.
Proof of Theorem 1.13.
Let be a non-negative weak subsolution of (1.3). By the homogeneity of equation (1.3) and inequality (1.14), to prove this result it will suffice to assume that and prove that
(4.5) |
To prove (4.5) we will apply an iteration argument very similar to that in the proof of Theorem 1.9, but to the solution of an auxiliary equation we which now define. Given that , and since by Theorem 1.9 is bounded in , we can apply Lemma 2.26 and fix such that
(4.6) |
Define (where will be determined below) and let . By Lemma 2.19, is a non-negative weak subsolution of
(4.9) |
For each , let and . By Lemma 2.14, . By Lemma 4.1, there exist , and such that (4.2) holds. We can now argue as we did in the proof of Theorem 1.9 with as a test function, and then apply Hölder’s inequality twice to get
(4.10) |
the last inequality follows since and since by (4.6), with a constant independent of and .
Now define the Young function and note that . Therefore, arguing as before, by Lemma 2.7 and (4) we have that
We can now argue as we did in the proof of Theorem 1.9 to get that for all ,
the last inequality holds by (4.3).
We continue the proof of Theorem 1.9 and define , , , as in (3.3), and to again get the iteration inequality
(4.11) |
We will again prove that we can choose the parameter such that and for every ,
(4.12) |
Assume for the moment that (4.12) holds. Then arguing as before we have that : that is,
which in turn implies that (4.5) holds as desired.
Therefore, to complete the proof we need to show that (4.12) holds. The proof is almost identical to the proof of (3.6): the only difference is in the choice of which we will describe. We first estimate as we did for inequality (4):
where the last inequality holds since with norm bounded by a constant. Furthermore, by Hölder’s inequality and Lemma 2.6,
Since , for every we have . Thus, combining the above inequalities, we get
Hence,
and so we can choose independent of both such that
where is as in (4.11). We may now proceed exactly as in the proof of (3.6) to get that (4.12) holds. This completes our proof. ∎
5. Theorem 1.9 is almost sharp
In this section we construct Example 1.11 that shows that Theorem 1.9 is almost sharp in the case of the Laplacian. Our example is intuitively straightforward. Let our domain , , be the unit ball , and define
Let . We will show that if and only if . Moreover, we claim that, at least formally, if is the solution of on , then . For if we use the well-known fact that the Green’s function for the unit ball is , then
To make this argument rigorous we must justify our use of Green’s formula which requires that the function be continuous on . To overcome this, we give an approximation argument and show that the inequality
cannot hold with a uniform constant. For each , let be a continuous, non-negative, radial function such that if , and if . Define . Each is continuous, and if is the solution to the Dirichlet problem
then at the origin it is given by
It is immediate that as . Since by monotonicity of the norm, , we have that the inequality
cannot hold with a uniform constant if .
Therefore, to complete the proof, it will suffice to show if and only if . By the definition of the Luxemburg norm, it will suffice to show that . But this is straightforward:
where the implicit constant only depends on . Thus, if and only if , or equivalently, .
References
- [1] S Chanillo and R. L. Wheeden. Weighted Poincaré and Sobolev inequalities and estimates for weighted Peano maximal functions. Amer. J. Math., 107(5):1191–1226, 1985.
- [2] M Christ. Hypoellipticity in the infinitely degenerate regime, complex analysis and geometry (Columbus, OH, 1999). Ohio State Univ. Math. Res. Inst. Publ. de Gruyter, Berlin, 9:59–84, 2001.
- [3] S.-K. Chua and R. L. Wheeden. Existence of weak solutions to degenerate -Laplacian equations and integral formulas. J. Differential Equations, 263(12):8186–8228, 2017.
- [4] A. Cianchi. Strong and weak type inequalities for some classical operators in Orlicz spaces. J. London Math. Soc. (2), 60(1):187–202, 1999.
- [5] D. Cruz-Uribe, J. M. Martell, and C. Pérez. Weights, extrapolation and the theory of Rubio de Francia, volume 215 of Operator Theory: Advances and Applications. Birkhäuser/Springer Basel AG, Basel, 2011.
- [6] D. Cruz-Uribe, S. Rodney, and E. Rosta. Poincaré inequalities and Neumann problems for the -Laplacian. Canad. Math. Bull., 61(4):738–753, 2018.
- [7] D. Cruz-Uribe, S. Rodney, and E. Rosta. Global Sobolev inequalities and degenerate -Laplacian equations. J. Differential Equations, 268(10):6189–6210, 2020.
- [8] E. De Giorgi. Sulla differenziabilità e l’analiticità delle estremali degli integrali multipli regolari. Mem. Accad. Sci. Torino. Cl. Sci. Fis. Math. Nat., 3:25–43, 1957.
- [9] L. C. Evans and R. F. Gariepy. Measure theory and fine properties of functions. Textbooks in Mathematics. CRC Press, Boca Raton, FL, revised edition, 2015.
- [10] E. B. Fabes, C. E. Kenig, and R. P. Serapioni. The local regularity of solutions of degenerate elliptic equations. Comm. Partial Differential Equations, 7(1):77–116, 1982.
- [11] G. B. Folland. Real analysis. Pure and Applied Mathematics (New York). John Wiley & Sons, Inc., New York, second edition, 1999. Modern techniques and their applications, A Wiley-Interscience Publication.
- [12] D. Gilbarg and N. S. Trudinger. Elliptic Partial Differential Equations of Second Order. Classics in Mathematics. Springer-Verlag, Berlin, 2001. Reprint of the 1998 edition.
- [13] L. Korobenko, C. Rios, E. T. Sawyer, and R. Shen. Sharp local boundedness and maximum principle in the infinitely degenerate regime via De Giorgi iteration. Memoirs Amer. Math. Soc., to appear. arXiv1608.01630v4.
- [14] M. A. Krasnosel′skiĭ and Ja. B. Rutickiĭ. Convex functions and Orlicz spaces. Translated from the first Russian edition by Leo F. Boron. P. Noordhoff Ltd., Groningen, 1961.
- [15] M. T. Lacey and S. Spencer. On entropy bumps for Calderón-Zygmund operators. Concr. Oper., 2(1):47–52, 2015.
- [16] V. G. Maz′ya. Some estimates of solutions of second-order elliptic equations. Dokl. Akad. Nauk SSSR, 137:1057–1059, 1961.
- [17] V. G. Maz′ya. Weak solutions of the Dirichlet and Neumann problems. Trudy Moskov. Mat. Obšč., 20:137–172, 1969.
- [18] D. D. Monticelli and S. Rodney. Existence and spectral theory for weak solutions of Neumann and Dirichlet problems for linear degenerate elliptic operators with rough coefficients. J. Differential Equations, 259(8):4009–4044, 2015.
- [19] D. D. Monticelli, S. Rodney, and R. L. Wheeden. Boundedness of weak solutions of degenerate quasilinear equations with rough coefficients. Differential and Integral Equations, 25(1-2):143–200, 2012.
- [20] M. M. Rao and Z. D. Ren. Theory of Orlicz spaces, volume 146 of Monographs and Textbooks in Pure and Applied Mathematics. Marcel Dekker Inc., New York, 1991.
- [21] S. Rodney. Existence of weak solutions of linear subelliptic Dirichlet problems with rough coefficients. Canad. J. Math., 64(6):1395–1414, 2012.
- [22] E. T. Sawyer and R. L. Wheeden. Degenerate Sobolev spaces and regularity of subelliptic equations. Trans. Amer. Math. Soc., 362(4):1869–1906, 2010.
- [23] G. Stampacchia. Le problème de Dirichlet pour les équations elliptiques du second ordre à coefficients discontinus. Ann. Inst. Fourier (Grenoble), 15(fasc. 1):189–258, 1965.
- [24] S. Treil and A. Volberg. Entropy conditions in two weight inequalities for singular integral operators. Adv. Math., 301:499–548, 2016.
- [25] N. S. Trudinger. Linear elliptic operators with measurable coefficients. Ann. Scuola Norm. Sup. Pisa Cl. Sci. (3), 27:265–308, 1973.
- [26] X. Xu. Logarithmic up bounds for solutions of elliptic partial differential equations. Proc. Amer. Math. Soc., 139(10):3485–3490, 2011.