We prove the necessary and sufficient condition for the removability of the fundamental singularity, and equivalently for the unique solvability of the singular Dirichlet problem for the heat equation. In the measure-theoretical context the criterion determines whether the h-parabolic measure of the singularity point is null or positive. From the probabilistic point of view, the criterion presents an asymptotic law for conditional Brownian motion.

Consider the fundamental solution of the heat equation:
(1)
It is a distributional solution of the Cauchty Problem
where δ is a unit measure with support at x = 0. For any fixed γRN, let
be a fundamental solution with a pole at O(γ,0). The singularity of h at O represents the natural phenomenon of the space-time distribution of the unit energy initially blown up at a single point. The fundamental singularity is non-removable for the heat equation in R+N+1. In particular, the Cauchy Problem for the heat equation in R+N+1 has infinitely many solutions in class O(h).
The goal of this paper is to reveal the criterion for the removability of the fundamental singularity for open subsets of R+N+1. Let ΩR+N+1 be an arbitrary open set and Ω{t=0}={O}. Let g:ΩR be a boundary function such that g/h is a bounded Borel measurable. Consider a singular parabolic Dirichlet problem (PDP):
(2)
Solution of the singular PDP is understood in the Perron’s sense [see Sec. II A, and formulae (26) and (27)]. Furthemore, the expression “prescribing the behavior of u/h at O” is understood in the sense of requiring relations (26) and (27) at the boundary point O [similar convention is made for the singular PDP (7)]. Without prescribing the behavior of u/h at O, either there exists one and only one or infinitely many solutions of PDP [see Sec. II B and formula (31)]. The main goal of this paper is to find a necessary and sufficient condition for open sets Ω for the uniqueness of the solution to the PDP without prescribing u/h at O. The problem of removability vs non-removability of the fundamental singularity is equivalent to the question of the uniqueness of the solution to PDP (2) without prescribing the behavior of u/h at O. Note that we are not requiring existence of the limit of u/h at O. Assume that limzO,zΩ\{O}g/h exists. We prove that in this case the removability of the fundamental singularity at O is equivalent to the existence of a unique solution of the singular PDP (2) such that
Otherwise speaking, for a unique solution u, u/h will pick up the limit value at the singularity point O without being required (see Sec. II C and Definition II.6).
The following procedure provides a key problem in testing the removability of fundamental singularity. Let
and un be a unique solution of the parabolic Dirichlet problem
(3)
From the maximum principle it follows that
(4)
Therefore, there exists a limit function
(5)
which satisfies (2), and
The following is the key problem.

Problem Aγ: Is u ≡ 0 in Ω or u(x,t)0 in Ω? Equivalently, is the fundamental singularity at O removable or nonremovable for Ω?

Note that the Problems Aγ are equivalent for various γ. Indeed, the heat equation is translation invariant so that solving the problem when the pole of the fundamental solutions is (0, 0) is equivalent to solve the problem when the pole is at (γ, 0) for any γRN.

Next, we formulate the equivalent problem in RN+1. In that context, we are going to consider one-point Alexandrov compactification: RN+1RN+1{}. For any fixed finite γRN, consider a function
(6)
It is a positive solution of the heat equation in RN+1. If γ = 0, then h̃1, while in the case when γ ≠ 0, it is an unbounded solution with singularity at . The singularity is not removable for the heat equation in RN+1. We aim to reveal the criterion for the removability of the fundamental singularity for the open subsets of RN+1.
Let Ω̃RN+1 be an arbitrary open set, and g:Ω̃R be a boundary function, such that g/h̃ is a bounded Borel measurable. Consider a singular parabolic Dirichlet problem (PDP):
(7)
Without prescribing the behavior of u/h̃ at , there exists one and only one or infinitely many solutions of (7) [see Sec. II B, formula (38)]. The alternative statement is the question of removability vs non-removability of the fundamental singularity at for Ω̃. In particular, in the case γ = 0 (h̃1), we are addressing the uniqueness of a bounded solution of (7), the problem solved in Ref. 1. Similar to its counterpart (2), the key problem to test the removability of the singularity at is formulated as follows:
Let
and ũn be a unique solution of the parabolic Dirichlet problem
(8)
From the maximum principle it follows that
(9)
Therefore, there exists a limit function
(10)
which satisfies (7), and
The following is the key problem.

Problem Ãγ: Is ũ0 in Ω̃ or ũ(x,t)0 in Ω̃? Equivalently, is a fundamental singularity at removable or nonremovable for Ω̃?

Remark I.1.

Problem Ãγ can be formulated in RN+1 without any change. Indeed, the parabolic Dirichlet problem for the heat equation is uniquely solvable in any open subset R+N+1 in a class O(h̃). Therefore, given arbitrary open set ΩRN+1, the solution of the parabolic Dirichlet problem in Ω can be constructed as a unique continuation of the solution to the parabolic Dirichlet problem in Ω=ΩRN+1. Moreover, the latter is independent of the boundary values assigned on Ω∩{t = 0}, since Ω∩{t = 0} is a parabolic measure null subset of Ω. This implies that the Problem Ãγ is equivalent for Ω and Ω. Otherwise speaking, the fundamental singularity at is removable for ΩRN+1 if and only if it is removable for Ω.

The only problem in the family of formulated problems that is solved is the Problem Ãγ when γ = 0 (or Problem Ã0). The Problem Ã0 was formulated by Kolmogorov in 1928 in the seminar on the probability theory at Moscow State University in the particular case with Ω={|x|<f(t),<t<0}R2, with fC(−, 0] such that f()=+,f+,(t)12f+ as t↓ − . Kolmogorov’s motivation for posing this problem was a connection to the probabilistic problem of finding asymptotic behavior at infinity of the standard Brownian path. Let {ξ(t): t ≥ 0, P} be a standard one-dimensional Brownian motion and P(B) is the probability of the event B as a function of the starting point ξ(0). Consider the event
Blumenthal’s 01 law implies that P0(B) = 0 or 1; f(−t) is said to belong to the lower class if this probability is 1 and to the upper class otherwise. Remarkably, Kolmogorov Problem’s solution u is = 0 or >0 according to as f(−t) is in lower or upper class accordingly. Kolmogorov Problem in a one-dimensional setting was solved by Petrovsky in 1935, and the celebrated result is called the Kolmogorov–Petrovski test in the probabilistic literature2 (see also Ref. 3).

The full solution of the Kolmogorov Problem for arbitrary open sets Ω (or Problem Ã0) is presented in Ref. 1. A new concept of regularity or irregularity of is introduced according to whether the parabolic measure of is null or positive, and the necessary and sufficient condition for the Problem Ã0 is proved in terms of the Wiener-type criterion for the regularity of .

In the probabilistic context, the formulated problems Aγ and Ãγ are generalizations of the Kolmogorov problem to establish asymptotic laws for the h-Brownian processes.4 

For the special case of domains
(11)
where lC(R̄+;R̄+),l(0)=0,l(t)>0 for t > 0, the solution of the Problem Aγ reads:

Theorem I.2.
u ≡ 0 or u > 0, that is to say, the fundamental singularity is removable or non-removable according to whether the following integral diverges or converges
(12)

The result is local. The removability of the fundamental singularity is dictated by the boundary of the domain near the singularity point. Precisely, it is defined by the thinness of the exterior set R+N+1Ω near the singularity point O.

The removability of the singularity is locally order-preserving. Precisely, if for some δ > 0 we have Ω1 ∩ {0 < t < δ} ⊂ Ω2 ∩ {0 < t < δ}, then removability of the fundamental singularity for Ω2 (or non-removability for Ω1) implies the same for Ω1 (or Ω2) (Lemma IV.2, Sec. II C).

An equivalent form of the criterion can be written if we choose l(t)=(4tlogρ(t))12, and consider the domain Ω such that
(13)
such that
(14)
Then the claim of the Theorem I.2 remains valid if the integral (12) is replaced with
(15)
Some examples of functions ρ with divergent integral (15) are as follows:
(16)
On the other side, for ∀ϵ > 0, the integral (15) converges for the corresponding functions
(17)
We adopt the notation
Hence, we have the following law for the removability of the fundamental singularity. For arbitrary integer k ≥ 4, consider a domain
(18)
Then u ≡ 0 or u > 0, that is to say, the fundamental singularity is removable or non-removable according to ϵ = 0 or ϵ > 0.
Probabilistic counterpart: Let {x(t) = (x1(t), …, xN(t)): t ≥ 0, P} be an N-dimensional h-Brownian process, and P(B) is a probability of the event B as a function of the starting point x(τ) with τ > 0.4 The process x(t) is an almost surely continuous process whose sample functions never leave R+N+1 and proceed downward, that is, in the direction of decreasing t. In fact, almost every path starting at x(τ) has a finite lifetime τ and tends to the boundary point O as t↓ 0.4 Let r(t) = |x(t) − γ|: 0 < tτ be a radial part of the h-Brownian path starting at x(τ). Consider the event
Kolmogorov’s 01 law implies that Px(τ)(B) = 0 or 1; l is said to belong to the lower class if the probability is 1 and to the upper class otherwise. The probabilistic analog of Theorem I.2 states that if l ∈ ↑ and if t−1/2l ∈ ↓ for small t > 0, then l belongs to the upper class or to the lower class according as the integral (12) converges or diverges. In particular, for arbitrary integer k ≥ 4
belongs to the upper or lower class according as ϵ > 0 or ϵ ≤ 0.
Next, we describe the solution of the Problem Ãγ for a special class of domains
(19)
where δ ≪ − 1 and
(20)

Theorem I.3.
ũ0 or ũ>0, that is to say the singularity at is removable or non-removable according to the following integral diverges or converges
(21)

Typical examples for the divergence or the convergence of the integral (21) are given by (16) and (17) just by replacing t with |t|. Hence, we have the following law for the removability of singularity at . For arbitrary integer k ≥ 4, consider a domain (19) with
Then u ≡ 0 or u > 0, that is to say the singularity at is removable or non-removable according to ϵ ≤ 0 or ϵ > 0.

Remarkably, in the particular case γ = 0, Theorem I.3 coincides with the celebrated Kolmogorov–Petrovski test.2,3

Probabilistic counterpart: Let {x(t) = (x1(t), …, xN(t)): t < 0, P} be an N-dimensional h̃-Brownian process, and P(B) is a probability of the event B as a function of the starting point x(τ) with τ < 0.4 The process x(t) is an almost surely continuous process whose sample functions never leave RN+1 and proceed downward, that is, in the direction of decreasing t. Almost every path starting at x(τ) tends to the boundary point as t↓ − .4 Let r(t) = |x(t) + 2|: t < 0 be the radial part of the h̃-Brownian path starting at x(τ). Consider the event
Kolmogorov’s 01 law implies that Px(τ)(B) = 0 or 1; l is said to belong to the lower class if the probability is 1 and to the upper class otherwise. The probabilistic analog of Theorem I.3 states that if ρ satisfies (20), then l(t)=2(tlogρ(t))12 belongs to the upper class or the lower class according as the integral (21) converges or diverges. In particular, for arbitrary integer k ≥ 4
belongs to the upper or lower class according as ϵ > 0 or ϵ ≤ 0.

Being a generalization of the Kolmogorov problem, the Problems Aγ, Ãγ, and their solution expressed in Theorems I.2 and I.3 has far-reaching measure-theoretical, topological and probabilistic implications in Analysis, PDEs and Potential theory. The goal of this section is to formulate three outstanding problems equivalent to the Problems Aγ and Ãγ. Since the problems Aγ and Ãγ are equivalent via the Appell transformation, without loss of generality we are going to formulate the problems in the framework of the Problem Aγ. The equivalent formulation can be pursued in the framework of the Problem Ãγ by replacing the triple (R+N+1,Ω,h) with singularity point at O through the triple (RN+1,Ω̃,h̃) with the singularity point at respectively.

Consider a singular parabolic Dirichlet problem(PDP) (2). The solution of the PDP can be constructed by Perron’s method (or the method by Perron, Wiener, Brelot, and Bauer).4,5 Let us introduce some necessary terminology.

We will often write a typical point zRN+1 as z=(x,t),xRN,tR. A smooth solution of the heat equation is called a parabolic function. A bounded open set URN+1 is regular if for each continuous f:UR there exists one (and only one) parabolic function HfU:UR, such that
A function u is called a superparabolic in Ω if it satisfies the following conditions:
  1. < u ≤ +, u < + on a dense subset of Ω;

  2. u is lower semicontinuous (l.s.c.);

  3. for each regular open set U ⊂ Ω and each parabolic function vC(Ū), the inequality uv on ∂U implies uv in U.

A function u is called a subparabolic if −u is a superparabolic. We use the notation S(Ω) for a class of superparabolic functions. Similarly, the class of subparabolic functions is S(Ω).

A function u = v/h is called a h-parabolic, h-superparabolic, or h-subparabolic in Ω if v is parabolic, superparabolic, or subparabolic.4 

We use the notation Sh(Ω) [respectively Sh(Ω)] for a class of all h-superparabolic (respectively h-subparabolic) functions in Ω.

Given boundary function f on Ω, consider a h-parabolic Dirichlet problem (h-PDP): find h-parabolic function u in Ω such that
(22)
It is easy to see that h-parabolic function u=vh is a bounded solution of the h-PDP if and only if v is a solution of the PDP (2).
Assuming for a moment that fC(Ω), the generalized upper (or lower) solution of the h-PDP is defined as
(23)
(24)
The class of functions defined in (23) [or in (24)] is called upper class (or lower class) of the h-PDP. According to classical potential theory,4  f is a h-resolutive boundary function in the sense that
The indicator function of any Borel measurable boundary subset, and equivalently any bounded Borel measurable boundary function is resolutive. Being h-parabolic in Ω, HfΩh is called a generalized solution of the h-PDP for f. The generalized solution is unique by construction. It is essential to note that the construction of the generalized solution is accomplished by prescribing the behavior of the solution at O.

Equivalently, we can define a generalized solution of the PDP (2):

Definition II.1.
Let g:ΩR be a boundary function, such that g/h is a bounded Borel measurable. Then g is called a resolutive boundary function for the PDP (2), if f = g/h (extended to O) is h-resolutive for the h-PDP. The function
(25)
is called a generalized solution of the PDP (2).

In particular, for the resolutive boundary function g we have
where
(26)
(27)

Again, note that the unique solution HgΩ of the PDP (2) is constructed by prescribing the behavior of the ratio HgΩ/h at O.

The elegant theory, while identifying a class of unique solvability, leaves the following questions open:

  • Would a unique solution of the h-PDP still exist if its limit at O were not specified? That is, could it be that the solutions would pick up the “boundary value” f(O) without being required? Equivalently, would a unique solution of the PDP (2) still exist if the limit of the ratio of solution to h at O is not prescribed? In particular, is the fundamental singularity at O removable?

  • What if the boundary datum f (or g/h) on Ω, while being continuous at Ω\{O}, does not have a limit at O, for example, it exhibits bounded oscillations. Is the h-PDP [or PDP (2)] uniquely solvable?

Example II.2.
Let Ω=R+N+1. It is easy to see that the boundary of R+N+1 is h-resolutive and the only possible solutions of the h-PDP in R+N+1 are constants. Precisely, the unique solution of the h-PDP is identical with the constant f(O). Indeed, for arbitrary ϵ > 0, the function
is in the upper class (or lower class) for h-PDP in R+N+1 for f. Hence,
Since ϵ > 0 is arbitrary, the assertion follows. Equivalently, all possible solutions of the PDP (2) in R+N+1 are constant multiples of h, and the unique solution is identified by prescribing the ratio u/h at O.

Example II.2 demonstrates that if Ω=R+N+1, the answer is negative and arbitrary constant C is a solution of the h-PDP, Ch is a solution of the PDP (2), and the fundamental singularity at O is not removable.

The positive answer to these fundamental questions is possible if Ω is not too sparse, or equivalently ΩcR+N+1 is not too thin near O. The principal purpose of this paper is to prove the necessary and sufficient condition for the non-thinness of ΩcR+N+1 near O which is equivalent to the uniqueness of the solution of the h-PDP [or PDP (2)] without specification of the boundary function (or ratio of the boundary function to h) at O.

Furthermore, given bounded Borel measurable function f=g/h:Ω\{O}R, we fix an arbitrary finite real number f̄, and extend a function f as f(O)=f̄. The extended function is a bounded Borel measurable on Ω and there exists a unique solution HfΩh of the h-PDP, and the unique solution of the PDP (2) is given by (25). The major question now becomes:

Problem 1.

How many bounded solutions do we have, or does the constructed solution depend on f̄?

For a given boundary Borel subset AΩ, denote the indicator function of A as 1A. Indicator functions of the Borel measurable subsets of Ω are resolutive.4, h-Parabolic measure of the boundary Borel subset A is defined as4:
where z ∈ Ω is a reference point. It is said that A is an h-parabolic measure null set if μΩ(·, A) vanishes identically in Ω. If this is not the case, A is a set of positive h-parabolic measure. In particular, the h-parabolic measure of {O} is well defined:
The following formula is true for the solution HfΩh of the h-PDP:4 
(28)
Since f is extended to {O} as f(O)=f̄, we have
(29)
This formula implies that the uniqueness of the solution to the h-PDP without prescribing the behavior of the solution at the singularity point O, that is to say, the independence of HfΩh on f̄ is equivalent to whether or not O is an h-parabolic measure null set. Equivalently, according to the formula (25) the following formula is true for the unique solution of the PDP (2):
(30)
Splitting the integral as in (29) we have
(31)
where f̄ is a prescribed limit value of HgΩ/h at O. Similar to its counterpart (29), the formula (31) demonstrates that the uniqueness of the solution u of the PDP (2) without prescribing u/h at O is equivalent to whether or not O is an h-parabolic measure null set.

Hence, the following problem is the measure-theoretical counterpart of the Problem 1:

Problem 2.

Given Ω, is the h-parabolic measure of {O} null or positive ?

From the Example II.2 demonstrated above it follows that in the particular case with Ω=R+N+1, we have
(32)

Example II.3.
For arbitrary c > 0 consider a domain bounded by the level set of h:
where c=(4πc)N2. It is easy to see that the h-parabolic measure of {O} is positive, and we have
(33)

Both Problem 1 and 2 are equivalent to the Problem Aγ formulated in Sec. I. The connection follows from the following formula:
(34)
To establish (34), first note that the h-parabolic function u/h is in the lower class of the Perron’s solution H1{O}Ωh, which imply that
(35)
Moreover, H1{O}Ωh itself is in the lower class of the Perron’s solution un/h of the h-PDP in Ωn with boundary function 1Ωn{t=n1}, where un is a solution of PDP (3). Therefore, we have
(36)
passing to the limit as n, from (5), (35), (36), and (34) follows.

In light of the measure-theoretical counterpart of the removability of the fundamental singularity, we introduce a concept of h-regularity of the boundary point O.

Definition II.4.

O is said to be h-regular for Ω if it is an h-parabolic measure null set. Conversely, O is h-irregular if it has a positive h-parabolic measure.

Hence, Theorem I.2 establishes a criterion for the removability of the fundamental singularity in terms of the necessary and sufficient condition for the h-regularity of O.

Similarly, in the context of the singular PDP (7), and corresponding h̃-PDP, we have the formulae analogous to (29), (31), and (34):
(37)
(38)
(39)
We introduce a concept of h̃-regularity of the boundary point for Ω̃RN+1.

Definition II.5.

is said to be h̃-regular for Ω̃ if it is an h̃-parabolic measure null set. Conversely, is h̃-irregular if it has a positive h̃-parabolic measure.

In fact, in the particular case with γ=0,h̃1, it coincides with the concept of regularity of introduced in Ref. 1. Theorem I.3 presents a criterion for the removability of the fundamental singularity at in terms of the necessary and sufficient condition for the h̃-regularity of .

The notion of the h-regularity of O is, in particular, relates to the notion of continuity of the solution to the h-PDP at O.

Problem 3.
Given Ω, determine whether or not
(40)
for all bounded fC(Ω\{O}).

Note that if f has a limit at O, (40) simply means that the solution HfΩh is continuous at O.

The equivalent problem in the context of the PDP (2) is the following:

Problem 3′.
Given Ω, whether or not
(41)
for all g such that ghC(Ω\{O}) and bounded.

In particular, if g/h has a limit at O, (41) means that the limit of the ratio HgΩ/h at O exists and equal to the limit of g/h.

Definition II.6.

Singular PDP (2) (and corresponding h-PDP) is said to be regular at O if (40) and (41) are satisfied. It is said to be irregular at O otherwise.

Similarly, we introduce the concept of regularity at of the singular PDP (7), and corresponding h̃-PDP if the following conditions are satisfied:
(42)
(43)
Theorem I.2 (or I.3) express the solutions to equivalent Problems 1–3 in terms of the Kolmogorov–Petrovsky-type criterion for the h-regularity of O (or h̃-regularity of ).

We now reformulate the main results of Theorems I.2 and I.3 in a broader context as a solution of the equivalent Problems 1–3.

Theorem III.1.

For arbitrary open set ΩR+N+1 the following conditions are equivalent:

  1. O is h-regular (or h-irregular).

  2. Singular Parabolic Dirichlet Problem (2), and equivalently h-PDP has a unique (or infinitely many) solution(s).

  3. Singular Parabolic Dirichlet Problem (2), and equivalently h-PDP is regular (or irregular) at O.

Theorem III.2.

Let ΩR+N+1 be an open set satisfying (13) and (14). Then O is h-regular if and only if the integral (15) diverges.

Similarly in the context of the h̃-PDP the main results read:

Theorem III.3.

For arbitrary open set Ω̃RN+1 the following conditions are equivalent:

  1. is h̃-regular (or h̃-irregular).

  2. Singular Parabolic Dirichlet Problem (7), and equivalently h̃-PDP has a unique (or infinitely many) solution(s).

  3. Singular Parabolic Dirichlet Problem (7), and equivalently h̃-PDP is regular (or irregular) at .

Theorem III.4.

Let Ω̃RN+1 be an open set satisfying (19) and (20). Then is h̃-regular if and only if the integral (21) diverges.

The major problem in the Analysis of PDEs is understanding the nature of singularities of solutions to the PDEs reflecting the natural phenomena. It would be convenient to make some remarks on the analysis of singularities for the Laplace and heat equations, as well as more general second-order elliptic and parabolic PDEs. The solvability, in some generalized sense, of the classical DP in a bounded open set ERN, with prescribed data on ∂E, is realized within the class of resolutive boundary functions, identified by Perron’s method and its Wiener6,7 and Brelot8 refinements. Such a method is referred to as the PWB method, and the corresponding solutions are PWB solutions. Paralleling the theory of PWB solutions, the DP for the heat equation in an arbitrary open set is solvable within the class of resolutive boundary functions. We refer to Refs. 4 and 5 for an account of the theory. Wiener, in his pioneering works,6,7 proved a necessary and sufficient condition for the finite boundary point xo∂E to be regular in terms of the “thinness” of the complementary set in the neighborhood of xo. If the boundary of the domain is a graph in a neighborhood of x0, the Wiener criterion is entirely geometrical. A key advance made in Wiener’s work was an introduction of the concept of capacity - sub-additive set function dictated by the Laplacian for the accurate measuring of the thinness of the complementary set in the neighborhood of x0 for the boundary regularity of harmonic function. Formalized through the powerful Choquet capacitability theorem,9 the concept of capacity became a standard tool for the characterization of singularities for the elliptic and parabolic equations. The question of removability of isolated singularities for the linear second-order elliptic and parabolic PDEs was settled in Refs. 10–13. The Wiener criterion for the boundary continuity of harmonic functions became a canonical result driving the boundary regularity theory for the elliptic and parabolic PDEs. In 1935, Petrovsky proved a geometric necessary and sufficient condition for the regularity of the characteristic top boundary point for the heat equation in the domain of revolution2 (see also Refs. 14). In the same paper, he also presented an elegant solution of the Kolmogorov problem (see Sec. I, Problem Ã0) for the special domain of revolution (see also Ref. 3). The results formed the so-called Kolmogorov–Petrovsky test for the asymptotic behavior of the standard Brownian path as t↓ 0 and t↑ + , and opened a path for the deep connection between the regularity theory of elliptic and parabolic PDEs and asymptotic properties of the associated Markov processes.15 The geometric iterated logarithm test for the regularity of the boundary point for an arbitrary open set with respect to heat equation is proved in Ref. 16. Paralleling the Wiener regularity theory, Wiener’s criterion for the regularity of the finite boundary point for the heat equation was formulated in Ref. 17 along with the proof of the irregularity assertion. The problem was accomplished in Ref. 18, where the long-awaited regularity assertion was proved. As in its elliptic counterpart, the concept of heat capacity was a key concept to extend the Wiener regularity theory to the case of heat equation.5 However, the major technical difficulty in doing so was connected to the nature of singularities of the fundamental solution of the Laplace and heat equations. The former is an isolated singularity for the spherical level sets of the fundamental solution, while the latter is a non-isolated singularity point for the level sets of the fundamental solution of the heat equation. To complete the Wiener regularity result at finite boundary points for the heat equation, the major technical advance of paper18 was a proof of elegant boundary Harnack estimate near the non-isolated singularity point of the level sets of the fundamental solution to the backward heat equation. The result of18 was extended to the class of linear second-order divergence form parabolic PDEs with C1-Dini continuous coefficients in Refs. 19 and 20.

In Ref. 21 it is proved that the Wiener test for the regularity of finite boundary points concerning second-order divergence form uniformly elliptic operator with bounded measurable coefficients coincides with the classical Wiener test for the boundary regularity of harmonic functions. The Wiener test for the regularity of finite boundary points for linear degenerate elliptic equations is proved in Ref. 22. The Wiener test for the regularity of finite boundary points for quasilinear elliptic equations was settled due to Refs. 23–25. Nonlinear potential theory was developed along the same lines as classical potential theory for the Laplace operator, for which we refer to Ref. 26.

To solve the DP in an unbounded open set, Brelot introduced the idea of compactifying RN into RN{}, where is the point at of RN.27 The PWB-method is extended to the compactified framework, thus providing a powerful existence and uniqueness result for the DP in arbitrary open sets in the class of resolutive boundary functions. The new concept of regularity of was introduced in Ref. 28 for the classical DP, and in Ref. 1 for its parabolic counterpart. The DP with bounded Borel measurable boundary function has one and only one or infinitely many solutions without prescribing the boundary value at . The point at is called a regular if there is a unique solution, and it is called irregular otherwise. Equivalently, in the measure-theoretical context, the new concept of regularity or irregularity of is introduced according to whether the harmonic measure of is null or positive. In Ref. 28 the Wiener criterion for the regularity of in the classical DP for the Laplace equation in an open set ERN with N ≥ 3 is proved. In Ref. 29 it is proved that the Wiener criterion at for the linear second-order divergence form elliptic PDEs with bounded measurable coefficients coincides with the Wiener criterion at for the Laplacian operator. The Wiener criterion at for the heat equation is proved in Ref. 1. Remarkably, the Kolmogorov problem (see Sec. I, Problem Ã0) is a particular case of the problem of uniqueness of the bounded solution of the parabolic Dirichlet problem in arbitrary open set in RN+1 without prescribing the limit of the solution at . Hence, the Wiener criterion at proved in Ref. 1 presents a full solution to the Kolmogorov problem.

The new Wiener criterion at for the elliptic and parabolic PDEs broadly extends the role of the Wiener regularity theory in mathematics. The Wiener test at arises as a global characterization of uniqueness in boundary value problems in arbitrary unbounded open sets. From a topological point of view, the Wiener test at arises as a thinness criterion at in fine topology. In a probabilistic context, the Wiener test at characterizes asymptotic laws for the Markov processes whose generator is a given differential operator. The counterpart of the new Wiener test at a finite boundary point leads to uniqueness in a Dirichlet problem for a class of unbounded functions growing at a certain rate near the boundary point; a criterion for the removability of singularities and/or for unique continuation at the finite boundary point: let ERN,N3 be an open set, and x0E be a finite boundary point. Consider a singular Dirichlet problem for the linear second order uniformly elliptic PDE with bounded measurable coefficients in a class O(|xx0|2−N) as xx0. In Ref. 29 it is proved that the Wiener test at x0 is a necessary and sufficient condition for the unique solvability of the singular Dirichlet problem, and equivalently for the removability of the fundamental singularity at x0. In a recent paper30 an appropriate 2D analog of this result is established. Let ER2 be a Greenian open set, and x0∂E be a boundary point (finite or ). Consider a singular Dirichlet problem for the linear second-order uniformly elliptic operator with bounded measurable coefficients in the class O(log |xx0|) if x0 is finite, and in a class of functions with logarithmic growth, if x0 = . In Ref. 30 it is proved that the Wiener criterion at x0 is a necessary and sufficient condition for the unique solvability of the singular Dirichlet problem, and equivalently for the removability of the logarithmic singularity. Precisely, in Ref. 30 the concept of log-regularity (or log-irregularity) of the boundary point (finite or ) is introduced according as if log-harmonic measure of it is null or positive, and the removability of the logarithmic singularity is expressed in terms of the Wiener criterion for the log-regularity of x0.

The goal of this paper is to establish a necessary and sufficient condition for the removability of the fundamental singularity, and equivalently for the unique solvability of the singular PDP. In this paper, we prove the Kolmogorov–Petrovsky-type test. We address the proof of the Wiener-type criterion in the forthcoming paper.

The equivalence of two problems formulated in R+N+1 and RN+1 is a consequence of the Appell transformation. Consider a homeomorphism A:R+N+1{O}RN+1{} with
(44)
Let P(Ω) be a class of parabolic functions in an open set Ω. Given open set ΩR+N+1, the Appell transformation is a homeomorphism A:P(Ω)P(AΩ) defined as
(45)
The claim follows from the following formula:
(46)
In particular, the Appell transform of h is given by
(47)
as it is defined in (6).

From the formula (46) it follows that The Appell transformation is a homeomorphism between S(Ω)C2(Ω) and S(AΩ)C2(AΩ), where S(Ω) denotes the class of all superparabolic functions in Ω. A simple approximation argument can be used to demonstrate that the hypothesis C2(Ω) can be removed.4 

Appell transformation presents one-to-one mapping between the singular PDPs (2) and (7).

Lemma IV.1.

  1. Function u is h-parabolic (or h-superparabolic) in open set ΩR+N+1 if and only if u(A−1(x, t)) is h̃-parabolic (or h̃-superparabolic) in AΩRN1.

  2. HfΩh is a solution of the h-parabolic Dirichlet problem in ΩR+N+1 if and only if HfΩh(A1()) is a solution of the h̃-parabolic Dirichlet problem in AΩRN+1 with boundary function f(A−1), i.e.
    (48)
  3. O is h-regular for ΩR+N+1 if and only if is h̃-regular for AΩRN+1.

  4. HgΩ is a solution of the singular PDP (2) if and only if its Appell transform is a solution of the singular PDP (7) in AΩ with boundary function Ag(A−1), i.e.
    (49)
    (50)
  5. Problems Aγ|Ω and Ãγ|AΩ are equivalent, i.e. u ≡ 0 if and only if ũ0.

Proof.

  1. Let ΩR+N+1 be an open set, and u be h-parabolic function on Ω, i.e.
    where v is a parabolic function in Ω. Considering the Appel transform of v = uh we have
    which implies that u(A−1(x, t)) is h̃-parabolic function in AΩ. On the other side, let ΩRN+1 be an open set, and u be h̃-parabolic function on Ω, i.e.
    where v is a parabolic function in Ω. Considering the inverse Appel transform of v=uh̃ we have
    which implies that u(A(x, t)) is h-parabolic function in A−1Ω. The presented proof applies to smooth superparabolic functions without any changes. Using the standard smoothing, the proof is extended to h- and h̃-superparabolic functions as well.
  2. According to Claim (1) HfΩh(A1(x,t)) is h̃-parabolic in AΩ, and we only need to verify that the relations (23) and (24) corresponding to h̃-parabolic Dirichlet problem are satisfied. For arbitrary z = (x, t) ∈ AΩ, we have
    and
    which implies (48).
  3. If f=1{O}, (48) implies
    (51)
    which proves the claim.
  4. By using (25) and (48) we have
    (52)
    since
    (53)
    Proof of the symmetric relation (50) is similar.
  5. The claim is a direct consequence of (34) and (51).□

The next lemma expresses the fact that the property of h-regularity of the singularity point is local and order-preserving.

Lemma IV.2.

  1. If Ω1Ω2R+N+1, then

    • H1{O}Ω2h0H1{O}Ω1h0;

    • H1{O}Ω1h0H1{O}Ω2h0;

  2. Let ΩR+N+1, and Ωδ ≔ Ω ∩ {t<δ}, δ>0. Then H1{O}Ωh0 if and only if H1{O}Ωδh0 for some (and equivalently for all) δ > 0.

  3. If Ω̃1Ω̃2RN+1, then

    • H1{}Ω̃2h̃0H1{}Ω̃1h̃0;

    • H1{}Ω̃1h̃0H1{}Ω̃2h̃0;

  4. Let Ω̃RN+1, and Ω̃δΩ̃{t<δ},δ>0. Then H1{}Ω̃h0 if and only if H1{}Ω̃δh0 for some (and equivalently for all) δ > 0.

  5. H1{}Dδh̃0 for Dδ={xRN,t>δ},δ>0.

Proof.
  1. It is easy to see that H1{O}Ω2h=HgΩ1h on Ω1, where

(54)
Since Perron’s solution is order-preserving, it follows that
(55)
which implies the claims (1a) and (1b).
  1. The “only if” claim is trivial. To demonstrate the “if” claim, note that since Ωδ ∩ {t = δ} is a parabolic measure null set for Ωδ, we have

Therefore, by the maximum principle we have
which proves the claim (2).

The proof of (3) and (4) is identical to the proof of (1) and (2). The claim (5) is a consequence of the uniqueness result for the Cauchy problem.31 

We assume that Ω satisfies (13) and (14). Assume that the integral (15) converges. We aim to demonstrate that u > 0, or equivalently, {O} is h-irregular. Without loss of generality, we can assume that ρC1(0, δ). Indeed, otherwise we can select the function ρ1C1(0, δ) which satisfy
(56)
and consider the domain
(57)
From (56) it follows that the integral (15) is convergent for ρ1, and Ω1 ⊂ Ω ∩ {0 < t < δ}. Therefore, h-irregularity of {O} for Ω1 would imply so for Ω.
Consider a function
which is positive in Ω for 0 < δ ≪ 1, vanishes on Ω ∩ {|xγ|2 = 4t log ρ(t)}, and satisfies
(58)
For a function v = uh we have
(59)
Now we construct a function w with the following properties:
(60)
(61)
for some fixed 0 < T < δ and for all sufficiently large n. From (59) and (60) it follows that the function ũ=v+wh is h-subparabolic, and by the maximum principle we have
(62)
If we select the fixed value T sufficiently small, it follows that ũ(γ,T)>1/3. Therefore, within ΩT = Ω ∩ {t > T}, ũ is greater than the function which is h-parabolic in ΩT, takes the value 1/4 in {(x, t): |xγ| ≤ ϵ, t = T} for some 0 < ϵ ≪ 1, and vanishes on the rest of the parabolic boundary of ΩT. Hence, we have a positive lower bound for un in ΩT which is independent of n. Obviously, the same lower bound holds for the limit function u, that is to say the h-parabolic measure of O would be positive. Thus, to complete the proof we need to construct the function w with the properties (60) and (61).
As a function w we select a particular solution of the equation from (60):
(63)
We only need to check that (61) is satisfied. We have
We split the integral n1t=t/2t+n1t/2, and estimate the first one as follows:
(64)
From the convergence of the integral (15) it follows that the right-hand side of (64) converges to zero as t↓ 0. We then have
(65)
where ωN is a volume of the unit ball in RN. From the convergence of the integral (15) it follows that for some fixed value of T, (61) is satisfied for all sufficiently large n. This completes the proof of the h-irregularity of the singularity point O.
Let us prove the h-regularity of the singularity point O by assuming that the integral (15) diverges. Without loss of generality, we assume that ρ satisfies the additional conditions
(66)
(67)
(68)
The proof of the “if” statement is based on the construction of the family of h-superparabolic functions ũn with the following properties:
  1. ũn(x,t)0 in Ωn;

  2. |1ũn(x,n1)|<12;

  3. ∀ϵ > 0 there exists a number T < δ such that ∀t0 < T and for arbitrary sufficiently large n we have ũn(x,t0)<ϵ.

Indeed, the existence of such a family implies that un(x,t0)2ũn(x,t0)<2ϵ, for all large n. passing to the limit n↑ + it follows that u(x, t0) ≤ 2ϵ. From the maximum principle, it follows that u(x, t) ≤ 2ϵ for all t > t0. Since ϵ > 0 is arbitrary, the assertion of the theorem follows.

To construct such a family {ũn}, we need a more precise asymptotic evaluation of wh in Ω as t↓ 0. We have
(69)
We split the integral n1t=n1tμ(t)+tμ(t)tI+J, where μ(t) = k log−2ρ(t) and k > 0 is a small number at our disposal. Next, we find the asymptotics of I as t↓ 0, and prove that it provides a dominating term for the asymptotic behavior of w/h. Since μ(t) → 0 as t↓ 0, we have (t) ≪ t/2, and tτ > t/2 for n−1τ(t), and t sufficiently small. Therefore, we have
where the last inequality follows from the monotonicity of τ log ρ(τ). Indeed, from the assumption (67) it follows that
(70)
From (67) it follows that
which implies that
and therefore,
(71)
Hence, we have
for all sufficiently small t. That is to say, ∀ ϵ > 0 we can find T > 0 such that ∀t < T
and therefore, we have
where ωN is the volume of the unit ball. Since n−1τ(t), we have
Therefore, ∃ T > 0 such that ∀t < T, and for all sufficiently large n we have
Hence, the following asymptotic relation is proved
Since (15) is divergent, it follows that
(72)
provided that the integrals J and
remain bounded as t↓ 0. We split the integral tμ(t)t=tμ(t)θt+θttI2+I3, with 0 < θ < 1 to be selected. For sufficiently small t we have
and we still need to demonstrate the boundedness of I4 as t↓ 0. This will be proved below while proving the boundedness of the integrals I5 and I6.
Next, we estimate wh inside Ωn for small t. As before, we split the time integral as n1t=θtt+tμ(t)θt+n1tμ(t)I5+I6+I7. To estimate I5, we use the identity
and derive
Let us estimate the integral
(73)
Assuming θ<12, from τ < θt it follows that tτ > t/2. Therefore, we have
From (73) we deduce
At this point, we are going to make a precise choice of the number θ. Since for arbitrary γ > 0
we can reduce the boundedness of |I6| to the boundedness of |I4| if we choose θ such that
We fix the value θ=165. Therefore, for sufficiently small t we have
By using assumption (68) we have
(74)
By using a l’Hôpital’s rule and (67) we have
(75)
and therefore, from (74) we deduce that
(76)
Finally, we estimate the integral
Our goal is to demonstrate that its asymptotics as t↓ 0 coincides with the asymptotics of the corresponding integral I in the expression of w(γ,t)h(γ,t). To prove that we need to demonstrate that the term expτ|xγ|22txγ,γξ4t(tτ) is close to 1 for small t. We have
Since tτ>t2, we have
Using (70) and (71), we also deduce that for all sufficiently small t
and the right-hand side will be sufficiently small if the parameter k at our disposal is chosen small enough.
Hence, we proved that for arbitrary ϵ > 0 there exists T > 0 such that for any fixed t < T and for all large n
Otherwise speaking,
Consider a function
where
From (59), (60), and (67) it follows that H(ũnh)0 for small t. Therefore, ũn is h-superparabolic function for sufficiently small t. We can check that it satisfies the required conditions (1)(3).
  1. We have |vh|<1 in Ωn, w1(x, n−1) = 0, and w1(y,t)h(y,t) as n↑ + , t↓ 0. Therefore,

(77)
  1. We have

(78)
(79)
(80)
From these conditions it follows that ∀ϵ > 0 there exists a number T < δ such that 0 < t0 < T and for arbitrary sufficiently large n we have ũn(x,t0)<ϵ.
  1. ũn(x,t)0 in Ωn, since vh0 and w1h0 in Ωn.

Hence we proved that the divergence of the integral (15) implies the h-regularity of O for Ω provided that additional assumptions (66)(68) are satisfied. Note that the assumptions (66)(68) are satisfied for all functions in (16). Therefore, we completed the proof of h-regularity of O and removability of the fundamental singularity for domains (18) with ϵ ≤ 0.

To complete the proof we only need to demonstrate that the assumptions (66)(68) can be removed. Differentiability assumption (66) can be removed as before with the only difference that we select ρ1C1(0, δ) which satisfy
(81)
and consider Ω1 as in (57). From (81) it follows that the integral (15) is divergent for ρ1, and Ω1 contains Ω ∩ {t < δ}. Therefore, h-regularity of O for Ω1 implies the h-regularity of O for Ω.
To remove assumption (68), assume on the contrary that there are arbitrarily small values of t such that ρ(t) ≤ |log t|. Consider a function
Clearly, the h-regularity of O for Ω1 implies the h-regularity of O for Ω. Hence, we only need to demonstrate that the integral (15) is divergent for ρ1. In view of our assumption we can choose the sequence {tk} with the following properties:
(82)
Let us define a function
We have
(83)
By the last relation of (82) we have
which implies that
(84)
From (83) and (84) it follows that
(85)
and therefore, the integral (15) is also divergent for ρ1. Hence, assumption (68) is removable.
Next, we are going to demonstrate that the assumption (67) is also removable. Assume that the given function ρC1(0, δ) has a divergent integral (15), but doesn’t satisfy the condition (67). Consider a one-parameter family of curves
(86)
Note that for any curve ρC the integral (15) is convergent, and since t1logN2t is monotonically decreasing function for small t, the convergence of (15) holds for any function satisfying ρρC near 0. Therefore, there are arbitrarily small values of t such that ρ(t) < ρC(t). Since ρC(C−1) = 0, a graph of the given function ρ with divergent integral (15) must intersect all curves ρC(t) with Cδ−1. For any point [ρ(t), t] on the positive quarter plane there exists a unique value
(87)
such that ρC(t) passes through the point [ρ(t), t]. Clearly, tC(t) → 0 as t → 0; and there exists a sequence tn↓ 0 such that C(tn)↑ + as n → +. We define a set
(88)
where C1(t)=maxtτδC(τ). Denote by M̄c the complement of M̄. Since M̄c is an open set, we have
(89)
From the definition of the set M it follows that C(t) satisfies the following properties:
(90)
Indeed, if we take t′, t″ ∈ M with t′ > t″, then we have
On the other hand, if t,tM̄ the same conclusion follows from the continuity of the function C(t).

To prove the second assertion, first note that since t2n,t2n1M̄, we have C1(t2n) = C(t2n) and C1(t2n−1) = C(t2n−1). Therefore, assuming that C(t2n) ≠ C(t2n−1) would imply that C1(t2n) > C1(t2n−1). Since C1 is a continuous function for some ϵ ∈ (0, t2n−1t2n) we have C1(t2n−1) < C1(t2n + ϵ). Let C1(t2n + ϵ) = C(θ). Obviously, we must have θ ∈ [t2n + ϵ, t2n−1) and C1(θ) = C(θ). However, this is a contradiction with the fact that (t2n,t2n1)M̄c, which proves the second assertion of (90).

To prove the third assertion of (90), recall that there is a sequence tn↓ 0 with C(tn)↑ + . Assume that for some n we have tnM̄. Then there exists kn such that tn(tkn,tkn+1)M̄c. According to the second assertion of (90) we have
Consider a sequence
(91)
Clearly, we have t̃nM̄;t̃n0 and C(t̃n)+ as n↑ + . From the monotonicity of C(t) on M̄ we easily deduce the third assertion of (90).

Let us define now a new function ρ1 with the following properties:

  1. ρ1(t) = ρ(t) for tM̄,

  2. ρ1(t) = |log(C(t2n)t)|3 for t(t2n,t2n1)M̄c.

Note that equivalently we can define ρ1 as
(92)
In fact, function C(t) defined for the function ρ1 via (86) coincides with C1(t), i.e.
The new function is continuous and satisfy ρ1(t) ≥ ρ(t) everywhere, with ρ1(t) ≠ ρ(t) on intervals (t2n, t2n−1). It can be easily seen that the function ρ satisfies the desired property (67) on M̄. Indeed, for a function ρC we have
Since C(t) is monotonically decreasing on M̄, we have
provided that C = C(t) is chosen as in (86). Therefore, we have
(93)
Since C(t)t → 0 as t↓ 0, the right-hand side is arbitrarily small for small t, and clearly ρ satisfies (67) on M̄.
Our next goal is to demonstrate that
(94)
Since ρρ1 it is satisfactory to demonstrate that the integral (94) is convergent for ρ1. By using the property (2) of ρ1 we have
(95)
where we use a notation CnC(t2n) ≡ C(t2n−1). Since Cnt is arbitrarily small for t sufficiently small, we can use the following inequality for all large n:
(96)
Using (96), from (95) we deduce
(97)
Since
and both ρ(t) and |log t|3 are decreasing functions, we deduce that
(98)
Therefore we also have
(99)
Hence, the series (97) is a telescoping series and therefore integrals (95) and (94) are convergent integrals. Since ρ1ρ on M̄c, it follows that the integral (94) is convergent for ρ. On the other side, since the integral (15) is divergent it follows that
(100)
Now we pursue the identical proof given above under the assumptions (66)(68) by replacing the function (63) with the following one:
(101)
where the function ρ̃ in the integrand is chosen with the following properties:
  1. ρ̃(t)=ρ(t) for tM̄;

  2. In the complementary set M̄c the function ρ̃ is chosen as sufficiently large continuous function with the property that all the estimations of the function w/h from (69) in the proof given above remain valid when the function ρ in the integrand is replaced with ρ̃.

  3. ρ̃(t)ρ(t),0<tδ.

To estimate the function w̃/h in Ω as t↓ 0, the time integral in (n−1, t) is split into two parts over M̄(n1,t) and M̄c(n1,t). The estimation of the first one is identical to the presented proof, for the assumptions (66)(68) are satisfied on M̄. Due to property (2) of the function ρ̃ the second integral remains bounded and accordingly does not affect the leading asymptotic of w̃/h given via the divergent integral (100). Precisely, we establish (72) and (77)(80), where the integral term n1t in expressions (78) and (79) is replaced with M̄(n1,t). This completes the proof of the h-regularity of O without assumptions (66)(68).□

Theorems I.3 follow from the Lemma IV.1 (iv) and the mapping (44).

According to the Definitions II.4, II.5, and formulae (34) and(39), Theorem I.2 and I.3 are equivalent to the Theorems III.2 and III.4 respectively.

The equivalence (1) ⇔ (2) in Theorems III.1 and III.3 follows from the formulae (29), (31), (37), and (38).

The equivalence (2) ⇔ (3) in Theorem III.3 with γ = 0 is proved in Ref. 1 (see Lemma 2.3, p. 472). Applying Lemma IV.1 (4), the equivalence (2) ⇔ (3) in Theorem III.1 with γ = 0 follows. Applying the translation xx + γ, the equivalence of (2) ⇔ (3) in Theorem III.1 with γ ≠ 0 easily follows. Applying Lemma IV.1 (iv) again, the equivalence (2) ⇔ (3) in Theorem III.3 with γ ≠ 0 follows.□

The authors have no conflicts to disclose.

Ugur G. Abdulla: Conceptualization (equal); Formal analysis (equal); Investigation (equal); Methodology (equal); Project administration (equal); Validation (equal); Writing – original draft (equal); Writing – review & editing (equal).

Data sharing is not applicable to this article as no new data were created or analyzed in this study.

1.
U. G.
Abdulla
,
Adv. Differ. Equations
13
(5–6),
457
488
(
2008
).
2.
I. G.
Petrovsky
,
Compos. Math.
1
,
383
(
1935
), http://www.numdam.org/item?id=CM_1935__1__383_0.
3.
U. G.
Abdulla
,
Nonlinear Anal.: Theory, Methods Appl.
63
,
712
(
2005
).
4.
J. L.
Doob
,
Classical Potential Theory and its Probabilistic Counterpart
(
Springer
,
1984
).
5.
N. A.
Watson
,
Introduction to Heat Potential Theory
(
American Mathematical Society
,
2012
).
8.
M.
Brelot
,
On Topologies and Boundaries in Potential Theory
,
Lecture Notes in Mathematics, Vol. 175
(
Springer-Verlag
,
1971
).
9.
G.
Choquet
,
Ann. Inst. Fourier
5
,
131
(
1955
).
10.
J.
Serrin
and
H. F.
Weinberger
,
Am. J. Math.
72
,
258
(
1966
).
11.
D. G.
Aronson
,
Arch. Ration. Mech. Anal.
17
,
79
(
1964
).
12.
D. G.
Aronson
,
Arch. Ration. Mech. Anal.
19
,
231
(
1965
).
13.
D. G.
Aronson
,
Ann. Sc. Norm. Super. Pisa
22
,
607
(
1968
), http://www.numdam.org/item/ASNSP_1968_3_22_4_607_0.pdf.
14.
U. G.
Abdulla
,
Bound. Value Probl.
2005
,
289712
.
15.
K.
Ito
and
H. P.
McKean
,
Diffusion Processes and Their Sample Paths
(
Springer
,
1996
).
16.
U. G.
Abdulla
,
SIAM J. Math, Anal.
34
,
1422
(
2003
).
17.
E.
Lanconelli
,
Ann. Math. Pura. Appl.
97
,
83
(
1973
).
18.
L. C.
Evans
and
R.
Gariepy
,
Arch. Ration. Mech. Anal.
78
,
293
(
1982
).
19.
N.
Garofalo
and
E.
Lanconelli
,
Trans. Am. Math. Soc.
308
,
811
(
1988
).
20.
E.
Fabes
,
N.
Garofalo
, and
E.
Lanconelli
,
Duke Math J.
59
,
191
(
1989
).
21.
W.
Littman
,
G.
Stampacchia
, and
H. F.
Weinberger
,
Ann. Sc. Norm. Super. Pisa Cl. Sci.
17
,
43
(
1963
), http://www.numdam.org/item/ASNSP_1963_3_17_1-2_43_0.pdf.
22.
E.
Fabes
,
D.
Jerison
, and
C.
Kenig
,
Ann. Inst. Fourier
32
,
151
(
1982
).
23.
V. G.
Maz’ya
,
Vestn. Leningrad Univ.: Math.
3
,
225
(
1976
).
24.
R.
Gariepy
and
W. P.
Ziemer
,
Arch. Ration. Mech. Anal.
67
,
25
(
1977
).
25.
T.
Kilpeläinen
and
J.
Maly
,
Acta Math.
172
,
137
(
1994
).
26.
J.
Heinonen
,
T.
Kilpeläinen
, and
O.
Martio
,
Nonlinear Potential Theory of Degenerate Elliptic Equations
(
Clarendon Press
,
1993
).
27.
M.
Brelot
,
Ann. Sci. Éc. Norm. Supér.
61
,
301
(
1944
).
28.
U. G.
Abdulla
,
Nonlinear Anal.: Theory, Methods Appl.
67
,
563
(
2007
).
29.
U. G.
Abdulla
,
Discrete Contin. Dyn. Syst.
32
,
3379
(
2012
).
30.
U. G.
Abdulla
,
Calculus Var. Partial Differ. Equations
57
,
157
(
2018
).
31.
A. N.
Tikhonov
,
Math. Sb.
42
,
199
(
1935
).