Volume 2011, Issue 1 198606
Research Article
Open Access

On Carlson′s Type Removability Test for the Degenerate Quasilinear Elliptic Equations

Farman I. Mamedov

Corresponding Author

Farman I. Mamedov

Institute of Mathematics and Mechanics, National Academy of Sciences, Baku Az 141, F. Agaev 9, Azerbaijan

Search for more papers by this author
Aslan D. Quliyev

Aslan D. Quliyev

Institute of Mathematics and Mechanics, National Academy of Sciences, Baku Az 141, F. Agaev 9, Azerbaijan

Search for more papers by this author
Mirfaig M. Mirheydarli

Mirfaig M. Mirheydarli

Institute of Mathematics and Mechanics, National Academy of Sciences, Baku Az 141, F. Agaev 9, Azerbaijan

Search for more papers by this author
First published: 02 November 2011
Academic Editor: Xingfu Zou

Abstract

Carlson′s type theorem on removable sets for α-Holder continuous solutions is investigated for the quasilinear elliptic equations div A(x, u, ∇u) = 0, having degeneration ω in the Muckenhoupt class. In partial, when α is sufficiently small and the operator is weighted p-Laplacian, we show that the compact set E is removable if and only if the Hausdorff measure .

1. Introduction

In this paper, we will consider questions of a removable singularity for the class of quasilinear elliptic equations of the form
(1.1)
where A(x, ξ, ζ) = {A1(x, ξ, ζ), A2(x, ξ, ζ), …, An(x, ξ, ζ)} : D × × nn are continuous with respect to ξ, continuously differentiable with respect to ζ functions. For ξ1, ζn, it is assumed that Aj(x, ξ, ζ);   j = 1,2, …n are measurable functions with respect to a variable x in the open domain Dn. Let the following growth conditions be satisfied:
(1.2)
where 1 < p < ,   λ ∈ (0,1). Throughout the paper, ω : n → [0, ] is a measurable function satisfying the doubling condition: for any ball B = B(x, r) with centre at a point x and of radius r > 0, the inequality
(1.3)
is satisfied, where the constant C is positive and does not depend on the ball Bn. For the system of functions Aj(x, ξ, ζ);   j = 1,2, …n, we can write the following expression:
(1.4)
where . Therefore, (1.1) can be written in the form
(1.5)
By virtue of (1.2) and (1.4), the system of functions is bounded and measurable. Moreover, the condition of uniform ellipticity is satisfied: for a.e. xD,   ηn there exist positive constants C1,   C2 such that
(1.6)
Denote by Cα(D),   0 < α ≤ 1, the class of continuous in D functions f : D satisfying the condition
(1.7)
with some K > 0 not depending on the points x, yD. Denote by the space of measurable functions in D, which have the finite norm
(1.8)
where are the derivatives of a function uLpω(D) in the sense of the distribution theory, which belong to the space Lpω(D). The norm of the space Lpω(D) is given in the form for p ≥ 1; for p = . Denote by a subspace of the space , where the class of functions is an everywhere dense set. Denote by the closure of the set of functions C(D) with respect to the norm . The spaces and coincide and are completely reflexive [1, 2] if the conditions 1 < p < and Ap-Muckenhoupt are fulfilled for ω:
(1.9)
where |B| denotes the Lebesgue measure of an arbitrary ball Bn. In the sequel, we will also use the A1-condition:

Definition 1.1. A function is called a solution of (1.1) if the integral identity

(1.10)
is fulfilled for any test function .

Definition 1.2. Let E ⊂ ⊂D be a compact subset of the bounded domain Dn. One will say that the set E is removable for the class of Cα(D) of solutions of (1.1) if any solution of (1.1) in DE from the space belongs to the space throughout the domain D and is extendable inside the compactum E as solution.

Definition 1.3. Let En be a bounded closed subset, h : → (0, ) a continuous function, and h(0) = 0,   μ some Radon measure. A finite system of balls , the radii of which do not exceed δ > 0, covers the set E, that is, E ⊂ ⋃νBν. Assume that , where the lower bound is taken with respect to all the mentioned balls.

Assume that

(1.11)
In the case μ = dx,   h(t) = tp+(p−1)α, the number is a Hausdorff measure of order np + (p − 1)α of the set E. We will sometimes denote it by mesnp+(p−1)α(E). Denote by the term for h(t) = tp+(p−1)α, dμ = ωdx.

By Carlson’s theorem [3], a necessary and sufficient condition for the compact set E to be removable in the class of harmonic outside E functions belonging to the class Cα(D) is expressed in terms of a Hausdorff measure of order n − 2 + α having the form
(1.12)
(For the case α = 1, the same result was proved in [4, 5].) In [6], the corresponding result was proved for a general linear elliptic equation with variable coefficients (see also [7, 8]). In [9], for the case p ≥ 2, a sufficient condition was proved for a solution of the p-Laplace equation (A = |∇u|p−2u) to be removable in the class Cα(D) (0 < α ≤ 1) in the form
(1.13)
Furthermore, a complete analogue of Carlson’s result was proved in [10], where the authors not only proved the necessity of condition (1.13), but also gave another proof of sufficiency that includes also range of exponent 1 < p < 2. Their approach was applied to the case of a metric measure space in [11].
It should be said that in [9] a somewhat general result was in fact obtained for the compact set E to be removable for the class Cα(D) of solutions of the equation
(1.14)
in the form of a sufficient condition
(1.15)
Note, in [9], the case p ≥ 2 was considered and it was required that the function ω to satisfy the doubling condition.

The present paper continues the development of the approach of [9]. We show that condition (1.15) is also the necessary one for the compact set E to be removable. Moreover, imposing some restrictions on the degeneration function, we manage to make the proof embrace a range of the exponent 1 < p < 2.

We will use the following auxiliary statements.

Lemma 1.4 (see [12].)Assume that a function uL1(D) satisfies the inequality

(1.16)
for any ball B(x, r) ⊂ D, where α ∈ (0,1). Then, uCα(D) and for any D ⊂ ⊂D the estimate
(1.17)
where C = C(n, α, D, D), is satisfied.

We also need the following analogue of the well-known Giaquinta’s lemma [13].

Lemma 1.5. Let ϕ(t), ω(t) be nonnegative nondecreasing functions on [0, R]. Assume that s > 0 is such that

(1.18)
for all r > 0 and 0 < λ < 1. Suppose that
(1.19)
for any 0 < ρ < r < R, with A, B, α, β nonnegative constants and β < α. Then, for any γ ∈ (β, α), there exists a constant ε0 = ε0(A, α, β, γ, s) such that if ε < ε0, then one has, far all 0 < ρ < r < R,
(1.20)
where c = c(β, α, A, s, γ) > 0.

Proof. For τ ∈ (0,1) and r < R, we have

(1.21)
Choose τ < 1 in such a way that 2Aτα = τγ and assume ε0ταs ≤ 1. Then, we get, for every r < R,
(1.22)
and therefore, for all integers k > 0,
(1.23)
Choosing k such that τk+2r < ρτk+1r, the last inequality gives
(1.24)
This proves Lemma 1.5.

We did not find the proof of the next inequality in the literature and therefore give here our proof.

Lemma 1.6. Let 1 ≤ p ≤ 2. Let x, yn be arbitrary points. Then, the estimate

(1.25)
is valid.

Proof. Let us introduce the vector function

(1.26)
acting from [0,1] into n. Applying the methods of differential calculus for the vector function, we obtain
(1.27)

The set of points {l(θ) ∈ n : l(θ) = θx + (1 − θ)y;   0 ≤ θ ≤ 1} in n forms a segment of the straight line that connects the point x with the point y. We denote this segment by [x, y]. Let |dl| be a length element of this segment. It is obvious that |dl| = |xy|dθ. Therefore, for the above integral expression, we have the estimate

(1.28)

To proceed with the estimation of this expression, we introduce into consideration the triangle, the base of which is the segment [x, y] and the vertex lies at the point 0. Now, the integration in the preceding estimate will be carried out with respect to the base of the triangle. It is not difficult to verify that the above integral expression takes a maximal value when the point 0 lies in the middle of the segment [x, y], which means that for it we have the estimate

(1.29)

To show that this is true, let us choose a new coordinate system, where the xn-axis is directed along the segment [x, y]. Let (u1, u2, …, un) be the coordinates of the the point 0 in the new coordinate system. Then, the preceding integral expression is equal to

(1.30)

The main result of this paper is contained in the following statements.

Theorem 1.7. Let Dn be a bounded domain, E ⊂ ⊂D be a compact subset. Let 2 < p < and ω be a positive, locally integrable function satisfying condition (1.3) or 1 < p < 2 and let any of the following conditions be fulfilled for the function ω:

  • (1)

    the function ω is integrable along any finite smooth n − 1-dimensional surface and condition (1.9′) is fulfilled for it;

  • (2)

    for any xD and sufficiently small ρ > 0, the condition is fulfilled for some s > np + 1, where the constant C > 0 does not depend on x.

Then, for a compact set E to be removable in the class Cα(D), 0 < α ≤ 1 of solutions of (1.1) in DE, , it is sufficient that condition (1.15) be fulfilled.

Here, we will use also the fact that a solution of generating equations of the form (1.1) is Hölderian. According to [14], when a weight ω belongs to the Muckenhoupt Ap-class, a solution of (1.1) belongs to the class Cκ(Dρ) in any subdomain Dρ = {xD : dist (x, D) > ρ} of the domain D. For solutions, we have the estimate
(1.31)
where κ = κ(n, p, Cp, λ)∈(0,1] and C = C(n, p, Cp, λ). Let κ denote a maximal number κ = κ(n, p, Cp, λ), for which the estimate (1.31) holds for solutions of (1.1). The following statement is valid.

Theorem 1.8. Let ωAp, E ⊂ ⊂D be a compact subset of the domain D. Let 0 < α < κ be some number. In that case, if , then the set E is not removable in the class of solutions of (1.1) which belong to Cα(D).

The foregoing statements give rise to the following corollaries.

Corollary 1.9. Let 0 < α < κ, 2 ≤ p < , ωAp, or 1 < p < 2 and any of the following conditions be fulfilled:

  • (1)

    the function ω satisfies condition (1.9′) and is integrable along any finite smooth n − 1-dimensional surface;

  • (2)

    for any xD and sufficiently small ρ > 0, the condition is fulfilled by some s > np + 1, where the constant C > 0 does not depend on x.

Then, for the compact set E to be removable for the class of solutions of (1.1) belonging to Cα(D), it is necessary and sufficient that condition (1.15) be fulfilled.

Corollary 1.10. Let 0 < α ≤ 1, 1 < p < . Then, for the compact set E to be removable in the class of solutions of the equation

(1.32)
belonging to the class Cα(D) throughout the domain D, it is necessary and sufficient that condition (1.13) be fulfilled.

2. Proof of the Main Results

In [9], the method of proving Theorem 1.7 was based on the application of an analogue of Landis-Gerver’s mean value theorem [15]. The restrictive condition p ≥ 2 used in [9] was necessitated by the proof of Lemma 2.1 below (see also [15, Lemma  1]). Below, we prove a such type lemma for the case 1 < p < 2, ignoring some smoothness of the function f and making some additional assumptions for the function ω.

Lemma 2.1. Let D be a bounded domain. Let 2 ≤ p < and the function ω : n → [0, ] satisfy condition (1.3) or 1 < p < 2, and let any of the following conditions be fulfilled for the function ω:

  • (1)

    condition (1.9′) is fulfilled and ω is integrable along any finite smooth n − 1-dimensional surface;

  • (2)

    for any xD and sufficiently small ρ > 0, the condition , where the constant C > 0 does not depend on x, is fulfilled for some s > np + 1.

Assume that f : D is a sufficiently smooth function (one can also assume the condition f(x) ∈ Cβ(D), where β ≥ min {p, 1}). Then, for any ε > 0, there exist a finite number of balls {Bν}, ν = 1,2, …, N, such that

(2.1)

Proof. We will follow the same reasoning as that used in proving Lemma  1  in [9] (see also [3], Lemma 2.1). The set Of = {xD : ∇f(x) = 0} is divided into two parts ; here, is the set of points where ∇2f(x) ≠ 0, and is the set of points where ∇2f(x) = 0. Let 1 < p < 2. Then, for the set , our reasoning is as follows. By virtue of the implicit function theorem, the set lies on a countable quantity of smooth n − 1-dimensional surfaces {Sj};   j = 1,2, …. Let xSj be a fixed point on the j-th surface. For sufficiently small r > 0, we have

(2.2)
where C1 = sup D |∇2f| and ω(S(x, r)) is integral omega over the n − 1 dimensional surface S(x, r). By virtue of Fubini’s formula,
(2.3)
Let E be the set of points t ∈ (rx, 2rx), for which the following condition is fulfilled:
(2.4)
Then
(2.5)
whence for the one-dimensional Lebesgue measure of the set E we obtain the estimate μ1(E) ≤ rx/2. Hence, by virtue of the doubling condition, there exists a point tx ∈ (rx, 2rx), for which
(2.6)
Then, for sufficiently small tj > 0, for any xSj, it can be assumed that there exists a number ρx ∈ (tj, 2tj), for which
(2.7)
Therefore,
(2.8)
For the surface Sj, from the system of balls {B(x, ρx); xSj}, we can extract, by virtue of Besicovtich theorem [16], a subcovering {B(xν, ρν);   xνSj, ν} with finite intersections:
(2.9)
Therefore and by construction, for xνSj,   ν = 1,2, …, we have ρν ∈ (tj, 2tj). Thus,
(2.10)
Here, we have used the sufficient smallness of tj, condition (1.9′), and the inequality . Choosing now tj, , where ε > 0 is an arbitrary number, we obtain , whence, after summing the inequalities over all surfaces Sj,   j, we find
(2.11)

In the case of the second condition 1 < p < 2, using , we immediately pass from the inequality (2.8) to (2.10) as

(2.12)
Due to the latter inequality, an estimate analogous to (2.10) will have the form
(2.13)
After choosing tj sufficiently small and taking the condition s > np + 1 into account, we can make the right-hand part smaller than ε/2j.

In the case p ≥ 2, the whole reasoning of [9] is applicable. Note that only instead of the inequality (2.10) we will have

(2.14)
where is the tj neighborhood of the surface Sj. After choosing a sufficiently small tj, we can make the right-hand part of this inequality smaller than ε/2j. This is possible because the n-dimensional Lebesgue measure of the surface Sj is equal to zero.

Now, it remains to obtain the covering for the set of points . Let 1 < p < 2. Let us decompose , where is the set of points , for which ∇3f ≠ 0. Here, we repeat the reasoning for . As above, the set is divided into two parts. In one part, we have ∇4f(x) ≠ 0, to which we apply the same reasoning as for . The second part of , where ∇4f(x) = 0, is again divided into two parts. At the k-th step, when k(p − 1) ≥ 1 and t > 0 is sufficiently small, this process yields the estimate

(2.15)
where η > 0 is arbitrary.

Note that, in the case p ≥ 2, the foregoing estimate gives the desired results immediately at the first step (k = 1).

Remark 2.2. It is not difficult to verify that under the assumptions of Lemma 2.1, instead of the condition of sufficient smoothness it sufficed to assume that f(x) ∈ Cβ(D), where β ≥ min {p, 1}.

Applying approaches similar to those in [15, Theorem  2.2, page 128] and [9], we prove the following analogue of Landis-Gerver’s lemma.

Lemma 2.3. Let 2 ≤ p < and the function ω : n → [0, ] satisfy condition (1.3) or 1 < p < 2 and let any of the following conditions be fulfilled for the function ω:

  • (1)

    condition (1.9′) is fulfilled and ω is integrable along any finite smooth n − 1-dimensional surface;

  • (2)

    for any xD and sufficiently small ρ > 0, the condition is fulfilled by some s > np + 1, where the constant C > 0 does not depend on x.

Let D be some domain lying in the spherical layer and having limit points on the surfaces of the spheres S(x0, 2r) and S(x0, r). Let be the quadratic form, the coefficients of which are well defined and continuously differentiable in the domain D and for which the inequalities

(2.16)
are fulfilled for any xD,   ηn for some λ ∈ (0,1). Assume that f : D is a sufficiently smooth function.

Then, there exists a piecewise-smooth surface Σ, separating, in the domain D, the surfaces of the spheres S(x0, r) and S(x0, 2r) and being such that

(2.17)
where is conormal derivative on Σ, n = (n1, n2, …, nn) is unit orthogonal vector to the surface Σ, and the constant K depends on p,  λ, and the dimension n.

Proof. It suffices to consider the case r = 1. Indeed, after the change of variables x = ry, the function f : DR transforms to the function , where . Also, . lies in the spherical layer B(0,2)∖B(0,1). It suffices to show that . Indeed, let a sufficiently small element of the surface Σ satisfy the equation φ(x) = 0 in coordinates x. Then, after the change of variables, this equation takes the form , where . In other words, the normals of the surfaces Σ and are related by . Therefore,

(2.18)
Applying these equalities, from the estimate
(2.19)
we obtain (2.17).

Let us now prove (2.19). Following the notation and reasoning of [15] (see also [9]), we assume that ε = ω(D)(oscf)p−1. For this ε, we find the corresponding balls Q1, Q2, …, QN of Lemma 2.1 and remove them from the domain D. Assume that and intersect D* with the closed layer (1 + 1/4) ≤ |x| ≤ (1 + 3/4). Denote this intersection by D. On the closed set D, we have ∇f ≠ 0. Let us choose some δ-neighborhood with δ < 1/4 so small that in we would have |∇f| > α > 0. We consider on the system of equations

(2.20)
In , there are no stationary points of the system (2.20), and at every point the direction of the field forms with the direction of the gradient an angle different from the straight angle. Let l(x) be the vector of the field at the point x. Then, using , we obtain
(2.21)
From this inequality, it follows that in there are no closed trajectories and all the trajectories have the uniformly bounded length.

Let some surface S be tangential, at each of its points, to the field direction. Then,

(2.22)
since the integrand is identically zero. We will use this fact in constructing the needed surface Σ. The base of Σ consists of ruled surfaces, while the generatrices are the trajectories of the system (2.20). Note that they will add nothing to the integral in which we are interested. These surfaces will have the form of fine tubes which will cover the entire D. Let us insert partitions into some of the tubes. The integral over these partitions will not any longer be equal to zero, but we can make it infinitesimal. The construction of tubes practically repeats that given in [15, pages 129–132].

Denote by E the intersection of D with the sphere . Let N be the set of points xE, where the direction of the field of the system (2.20) is tangential to the sphere . At the points xN, we have f/v = 0, where /v is the derivative with respect to the conormal to the sphere . Cover N by a set G, open on the sphere and being such that

(2.23)
Put E = EG. At the points xE, the direction of the field is transversal to the sphere. Cover E on the sphere by a finite number of uncovered domains with piecewise-smooth boundaries. We will call them cells. We will choose their diameters so small that at the points of the cells the field would be transversal to the sphere and the bundle of trajectories passing through each of the cells would diverge by δ/2n at most. The surface with trajectories lying inside the ball |x| < (1 + 3/4) and passing through the cell boundary will be called a tube. Thus, we obtain a finite number of tubes. We will call a tube a through tube if, without intersecting this tube, we can connect by a broken line a point of its corresponding cell with a point of the sphere within the limits of the intersection of D with a spherical layer 1 + 1/4 − δ < |x| < 1 + 3/4. Such through tubes are denoted by T1, T2, …, Ts. If every through tube is partitioned, then the spheres and are separated in D by the set-theoretic sum of nonthrough tubes, partitions T1, T2, …, Ts, the spheres S1, S2, …, SN, and the set G on the sphere .

Let us now take care to choose partitions in such a way that the integral ∫ω|∇f|p−2|f/v|ds over them would have the value which we need. Denote by Ui the domain bounded by Ti. Choose any trajectory on this tube. Denote it by Li. The length μ1Li of the curve Li satisfies the inequality μ1Li > 1/2. Introduce, on Li, the parameter l which is the length of the arc counted from . Denote by σi(l) the section Ui with a hypersurface which is orthogonal, at the point l, to the trajectory Li. Let the diameter at the beginning of the tube be so small that . Then, the set H of points lLi, where , satisfies the inequality μ1Li < 1/4. Thus, for E = LiH, the inequality μ1Li > 1/4 is valid and

(2.24)
At the points of the curve Li, the derivative f/l preserves the sign and therefore
(2.25)
Hence, using μ1Li > 1/4 and the mean value theorem, we see that there exists a point l0E such that . On the other hand, since, by virtue of (2.21), , we have . This together with (2.24) gives the estimate
(2.26)
Let us now choose a cell diameter so small that
(2.27)
This can be done since the derivatives f/xk,   k = 1,2, …, n, are uniformly continuous. Therefore,
(2.28)
Denote by Σ the set-theoretic sum of all nonthrough tubes, all σi(l0), all spheres Si, and the set G on the sphere . Then, from Lemma 2.1 and (2.22)–(2.28), we obtain
(2.29)

Lemma 2.3 is proved.

In this paper, we give the complete proof of Theorem 1.7. Some part of the proof of sufficiency is in fact identical to the proof given in [9]. The method of proving Theorem 1.8 is analogous to the method [10], where the nonweight case was considered.

Proof of Theorem 1.7 (Approximation). Let and uCα(D); let be a solution of (1.1). Denote by uj a mean value of the function u with smooth kernel ρ with finite support, . Then, it is obvious that u(j)C(D), j = 1,2, …. Moreover, uju uniformly in any subdomain . Also, for any open set EE contained in G, u(j)u in the norm of the space (see [2, 16]). Since, by condition (1.15), we have mesnE = 0, it can be assumed that mesnE < η, where η > 0 is an arbitrary number.

Let ε > 0 be an arbitrary number. Cover the set E by a finite system of balls , such that diam Bν < δ,

(2.30)
Assume that the number δ = δ(ε, η) is so small that the set lies in E.

For every ν, there exists, by virtue of Lemma 2.3 and inequality (1.31), a piecewise-smooth surface , separating the surfaces of the spheres (2Bν) and (4Bν), such that

(2.31)
Denote by the interiority of the surface . Then, . Assume that . Let for some ν. Then, for ν, the inequality (2.31) implies the estimate
(2.32)

It is obvious that the set is a strictly interior subdomain of the domain DE. Thus, we have the identity

(2.33)
for any . From this, by virtue of the convergence as j and the fact that the satisfies (2.16) by some λ > 0, we find
(2.34)
This follows from the fact that the integrand is a system of equi-integrable functions: for any subset gD∖Γ, we have
(2.35)
as mesng → 0. Here and in the sequel, speaking in general, we denote by δj different sequences tending to zero as j.

Green’s Formulae for Approximations Let now be an arbitrary function. Assume that ψ = φξ(d(x)/τ), where 0 ≤ ξ(s) ≤ 1 is an infinitely differentiable function equal to zero for s ≤ 0 and to one for s ≥ 1 and τ > 0 is a parameter, , d(x) = dist (x, Γj). It is obvious that . Then, (2.34) implies, for j = 1,2, …,

(2.36)
By virtue of the majorant Lebesgue theorem, for τ → 0, the first summand in (2.36) tends to the limit . Let us now find the limit of the second summand. Applying the Federer formula, we have
(2.37)
Applying the mean value theorem, for some t0 ∈ (0, τ), we obtain
(2.38)
Taking this limit relation into account, from (2.36), we obtain, as τ → 0, the following equality which is Green’ formula for approximation function uj:
(2.39)
Whence, in view of the inequality |φ(x)| ≤ ∥φC(D), xD, we have
(2.40)

Using the convergence   (j), Lemma 1.6, conditions (1.2), and Hölder inequality, we have the estimate for 1 < p < 2:

(2.41)
An analogous estimate for 2 ≤ p < has the form
(2.42)

The Belongness<!--${ifMathjaxEnabled: 10.1155%2F2011%2F198606}-->  u∈Wpω1(D)<!--${/ifMathjaxEnabled:}--><!--${ifMathjaxDisabled: 10.1155%2F2011%2F198606}--><!--${/ifMathjaxDisabled:}--> Taking into account (2.41) and (2.42), the estimate (2.32), and the uniform convergence u(j)u in G, convergence a.e. , we find

(2.43)
Therefore,
(2.44)
Taking into account the density of the class of functions in the space and the fact that , we also come to the same equality (2.44) for any function . Assuming now that, in (2.44) φ = uξp, where is a positive function equal to one in G, since Γj ⊂ Γ′′ and the integrand is positive, we obtain
(2.45)
Whence, by means of Young’s inequality, we derive
(2.46)
Then, by virtue of the arbitrariness of ε,  η (mesnE = 0), we obtain .

proof of that u(x) is a solution in D. Let us return to relation (2.44), from which, in view of and Γ′′E, we have

(2.47)
for any . By virtue of the arbitrariness of ε,   η, we find
(2.48)
that is, the function is a solution of (1.5) throughout the domain D and thereby of (1.1), too.

Theorem 1.7 is proved.

Proof of Theorem 1.8. Let for some compact set ED. Let us use the recent results for a Frosman type lemma with measure [11, 17] and follow the reasoning of the original paper [3]. We come to the following conclusion. There exists a Radon measure μ with a support on the set E, such that μ(E) > 0 and for any ball B = B(x, r) we have

(2.49)

Let be a solution of the equation

(2.50)
in the domain D = B(0, R), where B is a sufficiently large ball. The solution of (2.50) is understood in the sense as follows: the integral identity
(2.51)
is fulfilled for any test function . Such a solution exists by virtue of . Let us show the latter inclusion.

By virtue of ωAp for q > p and the fact that q is sufficiently close to p, inequality (2.49) implies for 0 < ρ < r:

where . This inequality defines the constant in the Adams inequality
(2.52)
as
(2.53)
Then, by virtue of (37′), we have
Whence, by virtue of Hölder inequality, we obtain
(2.54)
Taking into account inequalities (2.49), (37′′) and the Ap-condition, for any function , we have
whence it follows that .

Let us, following ideas of [10], show that u(x) ∈ Cα(D). Let be a solution of the equation

(2.55)
with the condition . Then, for it, we have the integral identity
(2.56)
where v = uh. The integrand in the left-hand part of (2.56) is positive. Therefore, for ρ < r/2, we have
(2.57)
By virtue of Young’s inequality and conditions (1.2), (1.3), from (2.57), we find
(2.58)
From (2.55), we obtain
(2.59)
whence by virtue of Hölder inequality, we find
For the first summand of (2.58), we have the following estimates. According to [14], there exists a positive number κ = κ(n, p, λ, Cp)∈(0,1) such that for the solution of (1.1) the inequality
(2.60)
where C = C(n, p, Cp, λ, κ), is fulfilled for any r1 < r2. If we take into account the Caccioppoli type estimate (see [14])
(2.61)
then, by virtue of Moser’s inequality, we obtain
(2.62)
From (2.60), we derive the estimate
(2.63)
Indeed,
(2.64)
where is the lower bound of the function h in the ball B(x0, r). Inequality (2.63) is proved.

Using the estimate (2.63) in (2.58), by virtue of (41′), we have for 0 < ρ < r/2

(2.65)

Now, let us derive an estimate for the last summand in (2.65). To this end, we use inequality (38′) to obtain

(2.66)
where for the constant we have the estimate
(2.67)
By virtue of (2.65) and (2.67), we find
(2.68)
whence, by virtue of Young’s inequality, we obtain
(2.69)
Assuming that 0 < α < κ, from (2.69) and Lemma 1.5, we obtain the estimate
(2.70)
This inequality implies
(2.71)
whence, by virtue of the Poincaré inequality, we obtain
(2.72)
where (u)ρ is average of the function u with respect to the ball B(x0, ρ).

By (2.72) and Campanato’s Lemma 1.4, we find uCα.

Theorem 1.8 is proved.

    The full text of this article hosted at iucr.org is unavailable due to technical difficulties.