Abstract

We establish, in dimension two, a regularity result for nonnegative solutions to an adjoint elliptic equation, generalizing a previous result of Escauriaza (1994). We consider elliptic equations with coefficients π‘Žπ‘–π‘—(π‘₯1,π‘₯2) which are measurable with respect to one variable and VMO with respect to the other.

1. Introduction

Let us consider a planar elliptic operator of nondivergence form:ℳ=π‘–π‘—π‘Žπ‘–π‘—πœ•(π‘₯)2πœ•π‘₯π‘–πœ•π‘₯𝑗,(1.1) where π‘Žπ‘–π‘—=π‘Žπ‘—π‘– for 𝑖,𝑗=1,2 are measurable and the symmetric matrixξ‚€π‘Žπ΄(π‘₯)=11(π‘₯)π‘Ž12π‘Ž(π‘₯)12(π‘₯)π‘Ž22(π‘₯)(1.2) is uniformly elliptic, that is, ||πœ‰||2βˆšπΎβˆšβ‰€βŸ¨π΄(π‘₯)πœ‰,πœ‰βŸ©β‰€πΎ||πœ‰||2,(1.3) for all πœ‰βˆˆβ„2 and a.e. π‘₯=(π‘₯1,π‘₯2)∈Ω, a bounded open subset of ℝ2. Here the ratio √√𝐾/1/𝐾=𝐾 is the ellipticity constant.

The study of weak solutions 𝑣 to the adjoint equation (adjoint solutions, for short)β„³βˆ—[𝑣]=ξ“π‘–π‘—πœ•2πœ•π‘₯π‘–πœ•π‘₯π‘—ξ€·π‘Žπ‘–π‘—ξ€Έ(π‘₯)𝑣(π‘₯)=0(1.4) often occurs in the literature (see Section 3, and for a very recent paper, see [1]).

We say that the function π‘£βˆˆπΏ1loc(Ξ©) is a weak solution to (1.4) if ξ€œΞ©[πœ‘]𝑣ℳ=0βˆ€πœ‘βˆˆπΆβˆž0(Ξ©).(1.5) In this paper we make the assumption that the coefficients π‘Žπ‘–π‘—(π‘₯1,π‘₯2) are VMO with respect to one of the two variables (see Section 2). This kind of assumption has been recently considered mainly for divergence (𝐿[𝑒]=div(𝐴(π‘₯)βˆ‡π‘’)=0) or nondivergence (𝑀[𝑀]=Tr(𝐴(π‘₯)𝐷2𝑀)=0) elliptic equations.

On the other hand, in [2] Escauriaza gave a regularity result for nonnegative solutions to adjoint equation with VMO coefficients.

Here, in case 𝑛=2, we give a generalized form of Theorem  1.2 in which he proves that, in particular,ξ™π‘£βˆˆπ‘ž>1πΊπ‘ž,(1.6) where πΊπ‘ž is the Gehring class, as defined in Section 2.

2. Definitions and Notations

In order to describe the results of the present paper, it is necessary to introduce some definitions. We start recalling basic definitions of the πΊπ‘ž classes, introduced by Gehring [3], in connection with local integrability properties of the gradient of quasiconformal mappings.

Let us assume that 𝑣 is a weight, that is, a nonnegative locally integrable function on ℝ2 and consider cubes π‘„βŠ‚β„2 with sides parallel to the coordinate axes. We will set𝑣𝑄=Q1𝑣(π‘₯)𝑑π‘₯=||𝑄||ξ€œπ‘„π‘£(π‘₯)𝑑π‘₯(2.1) to denote the mean value of 𝑣 over 𝑄, where |𝑄| denotes the 2-dimensional Lebesgue measure of a subset 𝑄 of ℝ2.

Definition 2.1. A weight 𝑣 satisfies the πΊπ‘ž-condition if there exists a constant 𝐺β‰₯1 such that, for all cubes π‘„βŠ‚β„2 as above, one has ξ‚€β¨π‘„π‘£π‘žξ‚(π‘₯)𝑑π‘₯1/π‘žβ¨π‘„π‘£(π‘₯)𝑑π‘₯≀𝐺,(2.2) and one refers to (2.2) as a β€œreverse” HΓΆlder inequality.

In the following, we will consider elliptic differential equations with coefficients π‘Žπ‘–π‘—(π‘₯) of the matrix 𝐴 measurable with respect to one variable and vanishing mean oscillation (VMO) with respect to the other (we say partially-VMO, for short). We recall that the space VMO, introduced by Sarason [4], is a subspace of the functions in the John-Nirenberg space BMO. More precisely, VMO is defined as the closure in BMO of the subspace of uniformly continuous functions.

Definition 2.2. A locally integrable function π‘“βˆΆβ„2→ℝ is in VMO if limπ‘Ÿβ†’0ξ€§π΅π‘Ÿ||𝑓(𝑦)βˆ’π‘“π΅π‘Ÿ||𝑑𝑦=0,(2.3) where π΅π‘Ÿ=𝐡(π‘₯,π‘Ÿ) denotes a ball centered at π‘₯βˆˆβ„2, with radius π‘Ÿ. One will also assume that 𝑓 is defined at ∞ in the following average sense: 𝑓(∞)=limπ‘Ÿβ†’βˆž1πœ‹π‘Ÿ2ξ€œπ΅π‘Ÿ(0)𝑓(𝑦)𝑑𝑦,(2.4) (see [5]).

3. Examples

In the present section we collect a certain number of examples where solutions 𝑣 to the adjoint equationβ„³βˆ—[𝑣]=ξ“π‘–π‘—πœ•2πœ•π‘₯π‘–πœ•π‘₯π‘—ξ€·π‘Žπ‘–π‘—ξ€Έ(π‘₯)𝑣(π‘₯)=0(3.1) occur. The first example deals with adjoint solutions which are partial derivatives β„Žπ‘₯1 and β„Žπ‘₯2 of a very weak solution β„Žβˆˆπ‘Š1,2loc to a particular diagonal divergence type equation, and an interesting relation comes out between regularity results.

Example 3.1. For √1/βˆšπΎβ‰€π›½β‰€πΎ, we consider the following elliptic operators in ℝ2: πœ•β„³=2πœ•π‘₯21πœ•+𝛽2πœ•π‘₯22,πœ•β„’=2πœ•π‘₯21+πœ•2πœ•π‘₯2ξ‚΅π›½πœ•πœ•π‘₯2ξ‚Ά.(3.2)
Fortuitous relations occur between adjoint solutions to β„³ and solutions β„Ž to β„’[β„Ž]=0, as the following Lemma reveals (see [6]).

Lemma 3.2 (see [6]). Let β„Žβˆˆπ‘Š1,2loc(π΅π‘Ÿ), where π΅π‘Ÿ denotes the open ball in ℝ2 centered at 0 with radius π‘Ÿ, such that β„’[β„Ž]=0.(3.3) Set 𝑀=πœ•β„Ž/πœ•π‘₯1, 𝑣=πœ•β„Ž/πœ•π‘₯2 Then β„³βˆ—[𝑀]=0,β„³βˆ—[𝑣]=0.(3.4)

Proof. We proceed similarly as in [6]. If πœ™βˆˆπΆβˆž0(π΅π‘Ÿ), we have ξ€œπ΅π‘Ÿξ€œπ‘€β‹…β„³πœ™π‘‘π‘₯=π΅π‘Ÿπœ•β„Žπœ•π‘₯1ξƒ©πœ•2πœ™πœ•π‘₯21πœ•+𝛽2πœ™πœ•π‘₯22ξƒͺ𝑑π‘₯1𝑑π‘₯2=ξ€œπ΅π‘Ÿξƒ©πœ•β„Žπœ•π‘₯1β‹…πœ•2πœ™πœ•π‘₯21+π›½πœ•β„Žπœ•π‘₯2β‹…πœ•2πœ™πœ•π‘₯1πœ•π‘₯2ξƒͺ𝑑π‘₯1𝑑π‘₯2=0.(3.5) Thus β„³βˆ—[𝑀]=0. In analogous way one checks that β„³βˆ—[𝑣]=0.

Corollary 3.3. Let β„Žβˆˆπ‘Š1,2loc such that β„’[β„Ž]=0. If πœ•β„Ž/πœ•π‘₯1β‰₯0 and πœ•β„Ž/πœ•π‘₯2β‰₯0, then for any ball π΅π‘ŸβŠ‚π΅2π‘ŸβŠ‚β„2, one has ξ‚΅ξ€§π΅π‘Ÿ|βˆ‡β„Ž|𝑝𝑑π‘₯1/𝑝≀𝑐(𝐾,𝑝)π΅π‘Ÿ||||βˆ‡β„Žπ‘‘π‘₯,(3.6) where 2≀𝑝<2𝐾/(πΎβˆ’1).

Proof. See [7, Theorem  3.1].

Compare this with the following well-known result of Astala [8] (see also Leonetti-Nesi [9]).

Theorem 3.4 (see [8, 9]). Let β„Žβˆˆπ‘Š1,2loc(Ξ©) be a local solution to the equation div(𝐴(π‘₯)βˆ‡β„Ž(π‘₯))=0,π‘₯∈Ω,(3.7) where 𝐴 is a real symmetric matrix satisfying the ellipticity bounds, ||πœ‰||2βˆšπΎβˆšβ‰€βŸ¨π΄(π‘₯)πœ‰,πœ‰βŸ©β‰€πΎ||πœ‰||2,βˆ€πœ‰βˆˆβ„2,fora.e.π‘₯∈Ω.(3.8) Then, for any ball 𝐡2π‘ŸβŠ‚Ξ© one has ξ‚΅ξ€§π΅π‘Ÿ|βˆ‡β„Ž|𝑠𝑑π‘₯1/𝑠≀𝑐(𝐾,𝑠)𝐡2π‘Ÿ||||βˆ‡β„Žπ‘‘π‘₯,(3.9) where √2≀𝑠<2√𝐾/(πΎβˆ’1).

Notice that, while the exponent in the left-hand side of the reverse inequality (3.9) may be greater than the exponent in the reverse inequality (3.6), this one is stronger in another sense, because it involves the same support π΅π‘Ÿ at both sides.

Example 3.5. In [10] (see also [11]) the Jacobian 𝑣=detπ·π‘ˆ, where π‘ˆβˆΆΞ©βŠ‚β„2→ℝ2 is a locally univalent 𝐴-harmonic mapping; that is, its components are π‘Š1,2loc solution to (3.7), is shown to be solution to an adjoint equation for the elliptic operator β„³[𝑣]=πœ•2π‘£πœ•π‘₯21πœ•+𝑐2π‘£πœ•π‘₯22,(3.10) where 1/𝐾≀𝑐≀𝐾.

Example 3.6. Very recently [12, 13], the reduced Beltrami differential equation πœ•π‘“πœ•π‘§ξ‚΅=πœ†(𝑧)β„π‘šπœ•π‘“ξ‚Ά,||||πœ•π‘§πœ†(𝑧)β‰€π‘˜<1,π‘˜=πΎβˆ’1𝐾+1(3.11) has been introduced and studied because it naturally arises in different contexts in the theory of quasiconformal mappings. It turns out that the partial derivatives of the components 𝑒,𝑣 of 𝑓(𝑧)=𝑒(𝑧)+𝑖𝑣(𝑧) solution to (3.11) satisfy the equation 𝑒π‘₯2=𝑏(𝑧)βˆ’1𝑣𝑏(𝑧)+1π‘₯1(3.12) with 𝑏(𝑧)=β„π‘šπœ†(𝑧). As a consequence of (3.12) in [13], it is proved that 𝑒π‘₯2β‰ 0 a.e, and it is an adjoint solution for a suitable elliptic operator βˆ‘β„³=𝑖𝑗𝑏𝑖𝑗(𝑧)(πœ•2/πœ•π‘₯π‘–πœ•π‘₯𝑗).

Namely, it has been proved [13] that 𝑒 is a solution to an elliptic equation of divergence formdiv𝐴(𝑧)βˆ‡π‘’=0,(3.13) where 𝐴(𝑧) is of the type𝐴(𝑧)=1π‘Ž12(𝑧)0π‘Ž22(𝑧).(3.14) As a consequence, the function 𝑣=𝑒π‘₯2 is a solution to the adjoint equationβ„³βˆ—[𝑣]=0,(3.15) whereπœ•β„³=2πœ•π‘₯21+π‘Ž12πœ•2πœ•π‘₯1πœ•π‘₯2+π‘Ž22πœ•2πœ•π‘₯22.(3.16) Note that the matrix 𝐴(𝑧) is not symmetric; however, the operator β„³ can also be represented by the symmetric and uniformly elliptic matrixπ΅βŽ›βŽœβŽœβŽ1π‘Ž(π‘₯)=12(𝑧)2π‘Ž12(𝑧)2π‘Ž22⎞⎟⎟⎠(𝑧).(3.17) Notice also that 𝑣=𝑒π‘₯2>0 a.e. (see [13]), and moreover, by general properties of nonnegative adjoint solutions, 𝑣 satisfies a reverse HΓΆlder inequality [7, 14, 15]ξ‚΅ξ€§π΅π‘Ÿπ‘£(𝑧)2𝑑𝑧1/2≀𝑐(𝐾)π΅π‘Ÿπ‘£(𝑧)𝑑𝑧,(3.18) in every ball π΅π‘ŸβŠ‚Ξ© such that 𝐡2π‘ŸβŠ‚Ξ©. Hence 𝑣 is identically zero or 𝑣>0 a.e. [13].

Example 3.7. The properties of the adjoint solutions are also very useful for studing the 𝐺-convergence of non divergence operators, as shown, for example, in a paper of D’Onofrio and Greco [16]. In that paper the authors consider elliptic operators β„³ of non divergence type, defined by β„³[𝑒]ξ€·=Tr𝐴𝐷2𝑒,forπ‘’βˆˆπ‘Š2,2(Ξ©),Ξ©βŠ‚β„2,(3.19) where 𝐴=(π‘Žπ‘–π‘—)βˆˆπ•„, the set of all symmetric 2Γ—2 real matrices and satisfy the ellipticity condition (1.3).
The adjoint to the operator β„³ is given by β„³βˆ—[𝑣]=(π‘Ž11𝑣)π‘₯1π‘₯2+2(π‘Ž12𝑣)π‘₯1π‘₯2+(π‘Ž22𝑣)π‘₯2π‘₯2 and reveals useful behaviour with respect to 𝐺-convergence of sequence of operators of the form (3.19).

Proposition 3.8 (see[16]). Let β„³π‘˜, π‘˜=1,2,…,β„³ be operators whose coefficient matrices π΄π‘˜, π΄βˆˆπ•„ and satisfying (1.3). Assume that π‘£π‘˜βˆˆπΏ2(Ξ©) are solutions to the adjoint equations β„³βˆ—π‘˜[π‘£π‘˜]=0 and verify that π‘£π‘˜β‡€π‘£in𝐿2𝑣(Ξ©),π‘˜π΄π‘˜β‡€π‘£π΄in𝐿2(Ξ©;𝕄),(3.20) where 𝑣(π‘₯)>0 a.e. in Ξ©. Then, one has β„³π‘˜Gβˆ’β†’β„³.

In order to prove Proposition 3.8, the following lemma is crucial.

Lemma 3.9 (see [16]). Let β„³π‘˜, π‘˜=1,2,…,β„³ be operators with coefficient matrices π΄π‘˜, π΄βˆˆπ•„ and satisfy (1.3), π‘£π‘˜βˆˆπΏ2(Ξ©) satisfying β„³βˆ—π‘˜[π‘£π‘˜]=0, and let π‘£π‘˜βˆˆπ‘Š2,2loc(Ξ©) be given. If π‘’π‘˜β‡€π‘’inπ‘Š2,2loc𝑣(Ξ©),π‘˜π΄π‘˜β‡€π‘£π΄in𝐿2loc(Ξ©;𝕄),(3.21) then 𝑣Trπ‘˜π΄π‘˜π·2π‘’π‘˜ξ€Έξ€·βŸΆTr𝑣𝐴𝐷2𝑒(3.22) in the sense of distributions.

Moreover, if we consider the Hessian matrix of any π‘€βˆˆπ‘Š2,2(Ξ©),𝐷2𝑀𝑀=π‘₯1π‘₯1𝑀π‘₯1π‘₯2𝑀π‘₯1π‘₯2𝑀π‘₯2π‘₯2ξ‚Ά.(3.23) In [17] it is proved that 𝑀 is a solution toβ„³[𝑀]ξ€·=Tr𝐡𝐷2𝑀=0,(3.24) where 𝐡 is a suitable coefficient matrix, if and only if𝑀2π‘₯1π‘₯2βˆ’π‘€π‘₯1π‘₯1𝑀π‘₯2π‘₯2ξ€Έξ‚€1𝐾+𝐾β‰₯𝑀2π‘₯1π‘₯1+2𝑀2π‘₯1π‘₯2+𝑀2π‘₯2π‘₯2,(3.25) where √𝐾, 𝐾≀1, is the elliptic constant. In the case where the Hessian matrix is diagonal, that is, 𝑀π‘₯1π‘₯2=0, it is easy to see that a solution of β„³βˆ—[𝑣]=0 is the positive function βˆšπ‘£=βˆ’det𝐷2𝑀=|𝑀π‘₯1π‘₯1𝑀π‘₯2π‘₯2|.

4. The Coefficients Measurable with Respect to One Variable and VMO with Respect to the Other

It is well known that, for linear elliptic operators in nondivergence form with continuous coefficients, the π‘Š2,𝑝 estimates hold for all 𝑝>1. It was shown that these estimates still hold in the same range when the coefficients are in VMO [18] or partially in VMO [19]. Our aim here is to generalize a regularity result of Escauriaza (Theorem  1.2, [2]) for the nonnegative adjoint solutions 𝑣 toβ„³βˆ—[𝑣]=0inΞ©βŠ‚β„2,(4.1) as defined in (1.4), withξ€·π‘Žπ΄=𝑖𝑗=𝑑𝐴,||πœ‰||2βˆšπΎβˆšβ‰€βŸ¨π΄(π‘₯)πœ‰,πœ‰βŸ©β‰€πΎ||πœ‰||2,(4.2) for a.e. π‘₯=(π‘₯1,π‘₯2)∈Ω and for πœ‰βˆˆβ„2.

Theorem 4.1. If π‘£βˆˆπΏ1(Ξ©) is a nonnegative solution to (4.1) and the coefficient matrix 𝐴(π‘₯) satisfies (4.2), and moreover 𝐴π‘₯1ξ€Έ,β‹…βˆˆVMO,(4.3) then ξ™π‘£βˆˆπ‘ž>1πΊπ‘ž.(4.4)

Let us begin with the following 𝐿𝑝-global regularity (for all 𝑝β‰₯2) result for the complex Beltrami equation 𝐹𝑧+πœ‡(𝑧)𝐹𝑧+πœ‡(𝑧)𝐹𝑧=𝐻(𝑧)π‘§βˆˆβ„2,(4.5) under a partially-VMO assumption on the Beltrami coefficients πœ‡, as defined in Section 2.

Proposition 4.2. Let π΅π‘Ÿ=𝐡(0,π‘Ÿ)βŠ‚β„2, and let πœ‡(𝑧)=πœ‡(π‘₯1,π‘₯2) be measurable, such that |πœ‡(𝑧)|+|πœ‡(𝑧)|=2|πœ‡(𝑧)|β‰€π‘˜<1 with π‘˜=(πΎβˆ’1)/(𝐾+1) and πœ‡(𝑧)=0 for |𝑧|β‰₯π‘Ÿ>0. Moreover, assume that πœ‡(π‘₯1,β‹…)∈VMO(ℝ,ℝ) for a.e. π‘₯1βˆˆβ„. Then for any 𝑝β‰₯2 and for π»βˆˆπΏπ‘(ℝ2)(𝐻(𝑧)=0 for |𝑧|>π‘Ÿ), there exists a unique solution 𝐹 to the Beltrami equation (4.5) such that πΉπ‘§βˆˆπΏπ‘ and ‖‖𝐹𝑧‖‖𝐿𝑝(ℝ2)≀𝑐(𝑝,π‘˜)‖𝐻‖𝐿𝑝(ℝ2).(4.6)

Remark 4.3. We note that in general, elliptic Beltrami operator πΌβˆ’πœ‡π‘‡βˆ’πœˆπ‘‡,(4.7) where 𝑇 is the Beurling transform defined via the relation 𝑇𝐹𝑧=𝐹𝑧(4.8) under the assumption ξ‚€ξβ„‚ξ‚πœ‡,𝜈∈VMO,(4.9) is invertible in all 𝐿𝑝(β„‚) spaces, 𝑝>1. The proof is much the same [5, 20], considering the complex Beltrami equation πΉπ‘§βˆ’πœ‡(𝑧)πΉπ‘§βˆ’πœˆ(𝑧)𝐹𝑧=β„Ž,β„ŽβˆˆπΏπ‘(β„‚).(4.10)
The meaning of the condition (4.9) is that πœ‡ and 𝜈 have vanishing mean oscillation in the usual sense, that is, belong to the closure of 𝐢∞0(β„‚) in BMO(β„‚) and that πœ‡ and 𝜈 are defined at infinity in the following average sense: πœ‡(∞)=limπ‘Ÿβ†’βˆžξ€§π΅π‘Ÿ(0)||||πœ‡(𝑧)𝑑𝑧2,𝜈(∞)=limπ‘Ÿβ†’βˆžξ€§π΅π‘Ÿ(0)||||𝜈(𝑧)𝑑𝑧2.(4.11)
The following example, due to T. Iwaniec, shows that without such condition the result fails.

Example 4.4. There exists a function π‘“βˆˆVMO(ℝ𝑛), 0≀𝑓(π‘₯)≀1 everywhere, such that 0=liminf|𝐡|β†’βˆžξ€§π΅π‘“(π‘₯)𝑑π‘₯<limsup||𝐡||β†’βˆžξ€§π΅π‘“(π‘₯)𝑑π‘₯=1,(4.12) where 𝐡 stands for a ball centered at the origin.

Preliminaries
Let πΏβˆΆβ„β†’[0,1] be a Lipschitz function given by 1𝐿(𝑑)=2ξ€·||||ξ€Έ.1+|𝑑|βˆ’π‘‘βˆ’1(4.13) The Lipschitz constant of 𝐿 equals 1, and, therefore, for each πœ‘βˆˆBMO(ℝ𝑛), we have β€–πΏβˆ˜πœ‘β€–BMO≀2β€–πœ‘β€–BMO.(4.14) Next, denote by 𝐢(𝑛) the BMO-norm of the function π‘₯β†’log|π‘₯|. We will truncate this function to make building blocks to our construction.

The Building Blocks
For a nonnegative integer π‘˜, we set πœ‘π‘˜ξ€·(π‘₯)=𝐿7βˆ’2βˆ’π‘˜ξ€Έ,πœ“log|π‘₯|π‘˜ξ€·(π‘₯)=𝐿5βˆ’2βˆ’π‘˜ξ€Έ.log|π‘₯|(4.15) We define the building block as π‘“π‘˜=πœ‘π‘˜βˆ’πœ“π‘˜. Note that each π‘“π‘˜ is continuous and supported in the ball |π‘₯|≀𝑒7β‹…2π‘˜, whereas π‘“π‘˜+1 vanishes on this ball. The BMO-norm of π‘“π‘˜ can be estimated as β€–β€–π‘“π‘˜β€–β€–BMOβ‰€β€–β€–πœ‘π‘˜β€–β€–BMO+β€–β€–πœ“π‘˜β€–β€–BMO≀2β‹…2βˆ’π‘˜πΆ(𝑛)+2β‹…2βˆ’π‘˜πΆ(𝑛)=2βˆ’π‘˜+2𝐢(𝑛).(4.16) Thus the infinite series 𝑓=βˆžξ“π‘˜=0π‘“π‘˜(4.17) represents a VMO function.

Computation of 𝐿1-Averages
Given any positive integer 𝑁, we consider concentric balls π΅π‘ŸβŠ‚π΅π‘… centered at the origin and with radii π‘Ÿ=𝑒4β‹…2𝑁<𝑒6β‹…2𝑁=𝑅. Elementary geometric observation reveals that 𝐡𝑅1𝑓β‰₯||𝐡𝑅||ξ€œπ΅π‘…π‘“π‘β‰₯1||𝐡𝑅||ξ€œπ‘’π‘5β‹…2≀|π‘₯|≀𝑒𝑁6β‹…2𝑓𝑁=𝑒6𝑛2π‘βˆ’π‘’5𝑛2𝑁𝑒6𝑛2𝑁=1βˆ’π‘’βˆ’π‘›2π‘βŸΆ1,asπ‘βŸΆβˆž.(4.18) On the other hand 𝐡𝑅1𝑓=||𝐡𝑅||π‘βˆ’1ξ“π‘˜=1ξ€œ|π‘₯|≀𝑒𝑁4β‹…2π‘“π‘˜1(π‘₯)𝑑π‘₯≀||𝐡𝑅||π‘βˆ’1ξ“π‘˜=1ξ€œ|π‘₯|β‰€π‘’π‘˜7β‹…2≀𝑑π‘₯π‘βˆ’1ξ“π‘˜=1𝑒7𝑛2π‘˜π‘’4𝑛2𝑁≀𝑒(π‘βˆ’1)7𝑛2π‘βˆ’1𝑒4𝑛2𝑁=(π‘βˆ’1)π‘’βˆ’π‘›2π‘βˆ’1⟢0,asπ‘βŸΆβˆž,(4.19) as desired.

Proof of Proposition 4.2. If we set 𝐾=(1+π‘˜)/(1βˆ’π‘˜), then there exists a symmetric matrix 𝐴(𝑧)=π‘Žπ‘–π‘—(𝑧)(4.20) such that πΌβˆšπΎβˆšβ‰€π΄(𝑧)≀1𝐾𝐼,πœ‡(𝑧)=2ξ‚Έπ‘Ž11(𝑧)βˆ’π‘Ž22(𝑧)+2π‘Ž12(𝑧)π‘–π‘Ž11(𝑧)+π‘Ž22(ξ‚Ή.𝑧)(4.21) Moreover π‘Žπ‘–π‘—(𝑧)=0 for |𝑧|β‰₯π‘Ÿ.
We may assume the following familiar normalization: Tr𝐴(𝑧)=π‘Ž11(𝑧)+π‘Ž22(𝑧)=1.(4.22) With the previous prescriptions we easily check that for π‘€βˆˆπ‘Š2,1loc(π΅π‘Ÿ) if we define the complex gradient of 𝑀 as 𝐹(𝑧)=𝑀𝑧=12𝑀π‘₯1βˆ’π‘–π‘€π‘₯2ξ€Έ,(4.23) we have ξ€·Tr𝐴(𝑧)𝐷2𝑀𝐹=2𝑧+πœ‡πΉπ‘§+πœ‡πΉπ‘§ξ‚(4.24) (see [5]). Hence (4.5) is equivalent to ξ€·Tr𝐴(𝑧)𝐷2𝑀=𝐻inℝ2(4.25) with coefficient matrix 𝐴(𝑧)=𝐴(π‘₯1,π‘₯2) allowed to be only measurable with respect to π‘₯1 and VMO with respect to π‘₯2βˆˆβ„, (thanks to (4.22)).
Under these assumptions, in [19, Theorem 2.4], the existence of a unique solution π‘€βˆˆπ‘Š2,𝑝 to (4.25) for π»βˆˆπΏπ‘ has been established (𝑝β‰₯2), together with the estimate ‖‖𝐷2𝑀‖‖𝐿𝑝(π΅π‘Ÿ)≀𝑐(𝑝,π‘˜)‖𝐻‖𝐿𝑝(π΅π‘Ÿ).(4.26) Hence (4.6) follows.

Let us now give the following sharp version of the Alexandrov-Bakelman-Pucci maximum principle for non divergence elliptic operatorsβ„³[𝑀]=ξ“π‘Žπ‘–π‘—πœ•2π‘€πœ•π‘₯π‘–πœ•π‘₯𝑗(4.27) with partially VMO coefficients.

Lemma 4.5. Under the assumptions (4.2), (4.3), (4.22) on 𝐴, suppose π΅π‘Ÿ=𝐡(0,π‘Ÿ)βŠ‚β„2 and that π‘€βˆˆπ‘Š2,1loc(π΅π‘Ÿ)∩𝐢0(π΅π‘Ÿ) satisfies, for β„ŽβˆˆπΏπ‘(π΅π‘Ÿ), 𝑝>1, β„³[𝑀]=β„Žπ‘–π‘›π΅π‘Ÿ,𝑀=0π‘œπ‘›πœ•π΅π‘Ÿ.(4.28) Then one has β€–π‘€β€–πΏβˆž(π΅π‘Ÿ)≀𝑐(𝐾,𝑝)π‘Ÿ2βˆ’2/π‘β€–β„Žβ€–πΏπ‘(π΅π‘Ÿ).(4.29)

Proof. In view of [19, Theorem  2.4], we know that the Dirichlet problem (4.28) always has a unique solution π‘€βˆˆπ‘Š2,𝑝(π΅π‘Ÿ)βˆ©π‘Š01,2(π΅π‘Ÿ) for every β„ŽβˆˆπΏπ‘(π΅π‘Ÿ), 𝑝>1.
Define β„Ž(𝑧)=0 for 𝑧=(π‘₯1,π‘₯2)βˆˆβ„2β§΅π΅π‘Ÿ and 1πœ‡(𝑧)=2ξ€Ίπ‘Ž11(𝑧)βˆ’π‘Ž22(𝑧)+2π‘Ž12ξ€»,(𝑧)𝑖(4.30) and set πœ‡(𝑧)=0 for π‘§βˆˆβ„2β§΅π΅π‘Ÿ. According to Proposition 4.2, the equation 𝐹𝑧+πœ‡πΉπ‘§+πœ‡πΉπ‘§=β„Ž2(4.31) has unique solution πΉβˆˆπ‘Š1,𝑝loc(π΅π‘Ÿ) such that ‖‖𝐹𝑧‖‖𝐿𝑝(ℝ2)≀𝑐(𝑝,𝐾)β€–β„Žβ€–πΏπ‘(ℝ2).(4.32)
Now, let us see that 𝐹|π΅π‘Ÿβˆˆπ‘Š1,2locξ€·π΅π‘Ÿξ€Έ.(4.33) Define 𝑓=𝑀𝑧=(1/2)(𝑀π‘₯1βˆ’π‘–π‘€π‘₯2). Then (π‘“βˆ’πΉ)𝑧+πœ‡(π‘“βˆ’πΉ)𝑧+πœ‡(π‘“βˆ’πΉ)𝑧=0inπ΅π‘Ÿ.(4.34) This means that the mapping 𝑔=π‘“βˆ’πΉ(4.35) is weakly 𝐾-quasiregular, and since πΉβˆˆπ‘Š1,𝑝loc and π‘“βˆˆπ‘Š1,2loc, we deduce π‘”βˆˆπ‘Š1,𝑝loc and actually 𝑔 is 𝐾-quasiregular and in particular π‘”βˆˆπ‘Š1,2locξ€·π΅π‘Ÿξ€Έ.(4.36) Then (4.33) follows.
Now, let us introduce the solution π‘ˆ to the problem Ξ”π‘ˆ=4𝐹𝑧,π‘ˆ(0)=0.(4.37) We have Ξ”π‘ˆβˆˆπΏπ‘(ℝ2)∩𝐿2loc(π΅π‘Ÿ) and β„³[π‘ˆ]=β„Ž a.e. π‘§βˆˆβ„2. Moreover, classically β€–Ξ”π‘ˆβ€–πΏπ‘(ℝ2)≀𝑐(𝐾,𝑝)β€–β„Žβ€–πΏπ‘.(4.38) (Notice that 𝑐(𝐾,𝑝) is optimal in Talenti [21].)
Finally, let us introduce 𝑒=π‘ˆβˆ’π‘€.(4.39) Then 𝑒 is continuous in π΅π‘Ÿ by the Sobolev imbedding, and it is the solution to the Dirichlet problem β„³[𝑒]=0a.e.π‘§βˆˆπ΅π‘Ÿ,𝑒/πœ•π΅π‘Ÿ=π‘ˆ,(4.40) and π‘’βˆˆπ‘Š2,2loc(π΅π‘Ÿ)∩𝐢0(π΅π‘Ÿ). By the classical maximum principle [22], β€–π‘ˆβˆ’π‘€β€–πΏβˆž(π΅π‘Ÿ)=β€–π‘ˆβ€–πΏβˆž(πœ•π΅π‘Ÿ).(4.41) Hence, we use Sobolev and the condition π‘ˆ(0)=0 to conclude β€–π‘€β€–πΏβˆž(π΅π‘Ÿ)≀2β€–π‘ˆβ€–πΏβˆž(π΅π‘Ÿ)β‰€π‘π‘Ÿ2βˆ’2/π‘β€–Ξ”π‘ˆβ€–πΏπ‘(π΅π‘Ÿ)β‰€π‘π‘Ÿ2βˆ’2/π‘β€–β„Žβ€–πΏπ‘(π΅π‘Ÿ).(4.42)

Proof of Theorem 4.1. Let us fix π‘ž>1, set 𝑝=π‘ž/(π‘žβˆ’1), and fix a ball π΅π‘Ÿ such that 𝐡2π‘ŸβŠ‚Ξ©. As in [7, 15], we make use of the dual formulation of the πΏπ‘ž-norm ξ‚΅ξ€œπ΅π‘Ÿπ‘£π‘žξ‚Ά1/π‘žξ‚»ξ€œ=supπ΅π‘Ÿπ‘£β„Ž:β„Žβ‰₯0,β„ŽβˆˆπΆ10ξ€·π΅π‘Ÿξ€Έ,β€–β„Žβ€–πΏπ‘(ℝ2)≀1.(4.43) Fix β„ŽβˆˆπΆ10(π΅π‘Ÿ),β€–β„Žβ€–πΏπ‘(ℝ2)≀1,β„Žβ‰₯0, and solve the Dirichlet problem β„³[𝑀]=ξ“π‘Žπ‘–π‘—πœ•2π‘€πœ•π‘₯π‘–πœ•π‘₯𝑗=β„Žin𝐡2π‘Ÿ,𝑀=0onπœ•π΅2π‘Ÿ.(4.44)
Since 𝐴(π‘₯1,β‹…)∈VMO, the problem has a unique solution π‘€βˆˆπ‘Š2,𝑝(𝐡2π‘Ÿ) vanishing on πœ•π΅2π‘Ÿ, satisfying the estimate ‖‖𝐷2𝑀‖‖𝐿𝑝(𝐡2π‘Ÿ)β‰€π‘β€–β„Žβ€–πΏπ‘(𝐡2π‘Ÿ),(4.45)𝑐=𝑐(𝐾,𝑝,‖𝐴‖VMO).
Fix a nonnegative function πœ‘π‘ŸβˆˆπΆ10(𝐡3π‘Ÿ/2) such that πœ‘π‘Ÿ=1 on π΅π‘Ÿ and |πœ•π›Όπœ‘π‘Ÿ/πœ•π‘₯𝛼|≀𝐢(𝛼)/π‘Ÿ|𝛼|.
Then, we have ξ€œπ΅π‘Ÿξ€œπ‘£β„Žβ‰€π΅2π‘Ÿ[𝑀]πœ‘π‘£β„³π‘Ÿξ€œ=βˆ’π΅2π‘Ÿξ€Ίπœ‘π‘£π‘€β„³π‘Ÿξ€»ξ€œβˆ’2𝐡2π‘Ÿπ‘£βŸ¨π΄βˆ‡π‘€,βˆ‡πœ‘π‘ŸβŸ©β‰€π‘π‘Ÿ2β€–π‘€β€–πΏβˆžξ€·π΅2π‘Ÿξ€Έξ€œπ΅3π‘Ÿ/2π‘βˆšπ‘£+πΎπ‘Ÿξ€œπ΅3π‘Ÿ/2𝑣||||.βˆ‡π‘€(4.46) By (4.29) β€–π‘€β€–πΏβˆž(𝐡2π‘Ÿ)≀𝑐(𝐾,𝑝)π‘Ÿ2βˆ’2/π‘β€–β„Žβ€–πΏπ‘(𝐡2π‘Ÿ)β‰€π‘π‘Ÿ2/π‘ž; hence, (4.46) implies ξ€œπ΅π‘Ÿcπ‘£β„Žβ‰€π‘Ÿ2π‘Ÿ2/π‘žξ€œπ΅3π‘Ÿ/2𝑐𝑣+π‘Ÿξ‚΅ξ€œπ΅3π‘Ÿ/2𝑣1/2ξ‚΅ξ€œπ΅2π‘Ÿπ‘£||||βˆ‡π‘€2ξ‚Ά1/2.(4.47)
Now, we estimate the last integral in the right-hand side. By (1.3), one has ξ€œπ΅2π‘Ÿπ‘£||||βˆ‡π‘€2β‰€βˆšπΎξ€œπ΅2π‘Ÿβˆšπ‘£βŸ¨π΄βˆ‡π‘€,βˆ‡π‘€βŸ©=πΎξ€œπ΅2π‘Ÿπ‘£ξƒ¬π‘€ξ€Ίπ‘€2ξ€»2ξƒ­βˆ’π‘€β„Ž.(4.48) Since 𝑀2=0 and βˆ‡(𝑀2)=0 on πœ•π΅2π‘Ÿ, then we deduce ξ€œπ΅2π‘Ÿξ€Ίπ‘€π‘£β„³2ξ€»=0asβ„³βˆ—[𝑣]=0.(4.49) Using again (4.29) yields ξ€œπ΅2π‘Ÿπ‘£||||βˆ‡π‘€2βˆšβ‰€2πΎξ€œπ΅2π‘Ÿβˆšπ‘£|𝑀|β„Žβ‰€2πΎβ€–π‘€β€–πΏβˆž(𝐡2π‘Ÿ)ξ€œπ΅π‘Ÿβˆšπ‘£β„Žβ‰€2πΎπ‘π‘Ÿ2/π‘žξ€œπ΅π‘Ÿπ‘£β„Ž.(4.50)
By (4.47) and (4.50), it follows that ξ€œπ΅π‘Ÿπ‘π‘£β„Žβ‰€π‘Ÿ2(1βˆ’1/π‘ž)ξ€œπ΅3π‘Ÿ/2𝑐𝑣+π‘Ÿ(1βˆ’1/π‘ž)ξ‚΅ξ€œπ΅3π‘Ÿ/2𝑣1/2ξ‚΅ξ€œπ΅π‘Ÿξ‚Άπ‘£β„Ž1/2.(4.51) By elementary inequality βˆšπ‘Žβˆšπ‘β‰€π‘Ž/2+𝑏/2, we obtain ξ€œπ΅π‘Ÿπ‘π‘£β„Žβ‰€π‘Ÿ2(1βˆ’1/π‘ž)ξ€œπ΅3π‘Ÿ/2𝑐𝑣+π‘Ÿ2(1βˆ’1/π‘ž)ξ€œπ΅3π‘Ÿ/21𝑣+2ξ€œπ΅π‘Ÿπ‘£β„Ž.(4.52)
Rearranging yields ξ€œπ΅π‘Ÿπ‘π‘£β„Žβ‰€π‘Ÿ2(1βˆ’(1/π‘ž))ξ€œπ΅3π‘Ÿ/2𝑣.(4.53)
Since β„Ž is arbitrary, by (4.43) and (4.53), we obtain ξ‚΅ξ€§π΅π‘Ÿπ‘£π‘ξ‚Ά1/𝑝≀𝑐𝐡3π‘Ÿ/2𝑣,(4.54) with 𝑐=𝑐(𝐾,𝑝). An application of [14, Lemma  2.0] concludes the proof.

Acknowledgment

The author wish to thank Professor T. Iwaniec for letting her include in this paper Example 4.4, which is a simplified variant of a construction by P. Jones.