Skip to content
BY 4.0 license Open Access Published by De Gruyter July 2, 2020

Gradient estimate of a variable power for nonlinear elliptic equations with Orlicz growth

  • Shuang Liang and Shenzhou Zheng EMAIL logo

Abstract

In this paper, we prove a global Calderón-Zygmund type estimate in the framework of Lorentz spaces for a variable power of the gradients to the zero-Dirichlet problem of general nonlinear elliptic equations with the nonlinearities satisfying Orlicz growth. It is mainly assumed that the variable exponents p(x) satisfy the log-Hölder continuity, while the nonlinearity and underlying domain (A, Ω) is (δ, R0)-vanishing in xΩ.

MSC 2010: 35J60; 35B65

1 Introduction

Throughout this paper, let Ω ⊂ ℝn for n ≥ 2 be a given bounded domain with its rough boundary specified later. Given a vectorial valued function f = (f1, f2, ⋯, fn) : Ω → ℝn. The aim of this present article is to study a global Calderón-Zygmund type estimate in the framework of Lorentz spaces for a variable power of the gradients of weak solutions to the zero-Dirichlet problem of general nonlinear elliptic equations

div A(x,Du)=div G(x,f)in Ω,u=0on Ω, (1.1)

where the nonlinearity A = A(x, ξ) : ℝn × ℝn → ℝn is measurable for each ξ ∈ ℝn and differentiable for almost every x ∈ ℝn, and there exist constants 0 < ν ≤ 1 ≤ Λ < ∞ such that for all x, ξ, η ∈ ℝn,

DξA(x,ξ)ηηνφ(|ξ|)|η|2,|A(x,ξ)|+|ξ||DξA(x,ξ)|Λφ(|ξ|). (1.2)

Here Dξ denotes the differentiation in ξ, and φ(s) : [0, +∞) → [0, +∞) has the following properties

φ(0)=0if and only ifs=0,φC1(R+),σφsφ(s)φ(s)τφfor everys>0,where0<σφ1τφ<. (1.3)

Moreover, the nonhomogeneous term G(x, ξ) : ℝn × ℝn → ℝn satisfies

|G(x,ξ)|Λφ(|ξ|)for anyx,ξRn. (1.4)

We also define

Φ=Φ(s)=0sφ(r)drfors0. (1.5)

The weak solution of (1.1) is understood in the following usual sense, if for uW1,Φ(Ω) it holds

ΩA(x,Du)Dϕdx=ΩG(x,f)DϕdxϕW01,Φ(Ω),

where W1,Φ is a Orlicz-Sobolev space defined in the following.

Definition 1.1

A Orlicz space LΦ(Ω) is the set of all measurable functions f : Ω → ℝ such that

ΩΦ(|f|)dx<.

Therefore, LΦ(Ω) is a Banach space equipped with the Luxemburg norm

fLΦ(Ω):=inf{μ>0:ΩΦ(|f|μ)dx1}.

A Orlicz-Sobolev space W1,Φ(Ω) is the set of all measurable functions fLΦ(Ω) for which DfLΦ(Ω,ℝn) with the norm ∥fW1,Φ(Ω) = ∥fLΦ(Ω) + ∥DfLΦ(Ω,ℝn). For more details about Orlicz spaces, we refer to [14, 15, 28, 29].

The equation of generalized p-Laplacian type is arising in the fields of fluid dynamics, magnetism and mechanics. Particularly, if φ(s) = sp−1 for 1 < p < ∞, our problem becomes the p-Laplace equation. For this case, there is a great deal of literature concerning the regularity of weak solutions, for instance, see [4, 9, 11, 13, 24]. While φ(s) = sp−1 + asq−1, where 1 < p < q and a is a positive constant, Filippis and Mingione [18] obtained a global estimate in the setting of Lebesgue spaces, which is covered by our problem (1.1). We also refer to [8] for the existence of solutions to the (p, q)-Laplace equations.

In recent years, there have been significant advances of this type of problem in the regularity theory, see for example [12, 14, 34, 35] since starting with a seminal paper of Lieberman [23]. In 2011, Verde [34] studied the elliptic system

div A(x,Du)= div G(x,f),

where A(x, ξ) = G(x, ξ) = φ(|ξ|)|ξ|ξ with φ(s) defined as in (1.3). He proved the global Calderón-Zygmund estimates over whole domain ℝn for the weak solutions. Very recently, Yao and Zhou [35] obtained local Lq-estimates for weak solutions of (1.1), which implies the fact that

Φ(|f|)Llocq(Ω)Φ(|Du|)Llocq(Ω),for anyq1.

Also, Byun and Cho [12] further extended it to a global gradient estimate in the setting of Orlicz spaces under the assumption that the boundary of underlying domain is Reifenberg flat. We would like to point out that Cho [14] also considered the zero-Dirichlet problem of (1.1) for A defined as in (1.2) and G(x, ξ) = φ(|ξ|)|ξ|ξ , who obtained a global Calderón-Zygmund estimate in the setting of Orlicz spaces. We also refer to [6, 7, 17, 36] for a further study of these problems with Orlicz growth. In particular, Baroni and Lindfors [7], and Yao [36] studied parabolic problems with Orlicz growth.

In this paper we are to prove a global Calderón-Zygmund type estimate in the Lorentz spaces for general nonlinear elliptic equations (1.1). As we know, there are mainly three kinds of different arguments to handle the Calderón-Zygmund theory for elliptic and parabolic problems with VMO or small BMO discontinuous coefficients except for a classical technique by using singular integral operators and their commutators. The first one is the so-called geometric method originally traced from Byun and Wang’s work in [11], which is based on the weak compactness, the Hardy-Littlewood maximal operators and the modified Vitali covering. Here, the so-called modified Vitali covering actually refers to the argument as “crawling of ink spots” as in the early papers by Safonov and Krylov [20, 30]. Indeed, this is also a development from Caffarelli and Peral’s paper in [13]. Secondly, Kim and Krylov [19, 21] gave a unified approach of studying Lp solvability for elliptic and parabolic problems due to the Fefferman-Stein theorem, which is mainly based on the sharp functions. In this present paper, we have to highlight the third technique being called large-M-inequality principle originating from Acerbi and Mingione’s work [1, 2], which is directly based on arguing on certain Calderón-Zygmund-type covering instead of the boundedness of a maximal function operator and the so-called good-λ-inequality used by Byun and Wang [11] and Kim and Krylov’s papers [19, 21].

Here we are revising the so-called large-M-inequality principle and geometric method to get an estimate in the Lorentz spaces for the variable power of the gradients to (1.1). Regarding what we consider, we would like to point out that Byun, Ok and Wang [10] studied the zero-Dirichlet problem of linear elliptic system in divergence form under the assumptions that their coefficients are partially BMO and the variable exponent p(x) is log-Hölder continuous. They first obtained a global Calderón-Zygmund estimate in the variable exponent Lebesgue spaces:

FLp(x)(Ω,Rn)DuLp(x)(Ω,Rn).

Also, Tian and Zheng in [32] further extended the above result to the global Calderón-Zygmund type estimate in the Lorentz spaces for a variable power of the gradients of weak solutions. Note that the Lorentz space is a two-parameter scale of the Lebesgue space obtained by refining it in the fashion of a second index, and there are a lot of research activities on Lorentz regularity for partial differential equations. For examples, the first estimates in the Lorentz spaces are obtained by Mingione [27]. Later, Mengesha and Phuc [24] derived the weighted Lorentz estimate to quasilinear p-Laplacian type equations based on the geometric approach. In 2014, Zhang and Zhou [39] extended the above result in [24] to that for a general equation of p(x)-Laplacian also using a geometrical argument. Adimurthil and Phuc [3] proved the global Lorentz and Lorentz-Morrey estimates for quasilinear equations below the natural exponent. Meanwhile, Baroni [4, 5] obtained interior Lorentz estimates to evolutionary p-Laplacian systems and obstacle parabolic p-Laplacian with the given obstacle function Ly,q locally in ΩT, respectively, which means that

F,DψLlocy,q(ΩT)DuLlocy,q(ΩT)

for y > p and q ∈ (0, ∞], where he just used the so-called large-M-inequality principle. Very recently, Tian and Zheng [31] showed a global weighted Lorentz estimate to linear elliptic equations with lower order items under assumptions of partially BMO coefficients in Reifenberg flat domain. Zhang and Zheng [37, 38] also proved Hessian Lorentz estimates for fully nonlinear parabolic and elliptic equations with small BMO nonlinearities, and weighted Hessian Lorentz estimates of strong solutions for nondivergence linear elliptic equations with partially BMO coefficients, respectively.

To this end, we introduce some related notation and basic facts being useful in the article. The Lorentz space Lt,q(U) for open subset U ⊂ ℝn with parameters 1 ≤ t < ∞ and 0 < q < ∞, is the set of all measurable functions g : U → ℝ requiring

gLt,q(U)q:=t0(μt|{ξU:|g(ξ)|>μ}|)qtdμμ<;

while the Lorentz space Lt,∞ for 1 ≤ t < ∞ and q = ∞ is defined by the Marcinkiewicz space 𝓜t(U) as usual, which is the set of all measurable functions g with

gLt,=gMt(U):=supμ>0(μt|{ξU:|g(ξ)|>μ}|)1t<.

The local variant of such spaces is defined as usual way. If t = q, then the Lorentz space Lt,t(U) is nothing but a classical Lebesgue space. Indeed, by Fubini’s theorem it yields

gLt(U)t=t0μt|{ξU:|g(ξ)|>μ}|dμμ=gLt,t(U)t,

which implies Lt(U) = Lt,t(U), see also [4, 5, 24, 37].

Let us denote

Br(y)={xRn:|xy|<r}

for yΩ and radius r > 0, and

Ωr(y)=Br(y)Ω,Br+(y)=Br(y){xn>0},Tr(y)=Br(y){xn=0}

with briefly Br = Br(0), Ωr = Ωr(0) and Tr = Tr(0). For a bounded open set U ⊂ ℝn, we write the integral average of g(x) over U of a locally integrable function g in ℝn by

gU=Ug(x)dx=1|U|Ug(x)dx.

As usual, it is a necessary assumption that the variable exponent p(⋅) is log-Hölder continuous, which ensures that the mollification, the singular integrals and the Hardy-Littlewood maximal operator are all bounded within the framework of generalized Lebesgue space. For this, we are recalling the definition that p(x) is log-Hölder continuous denoted it by p(x) ∈ LH(Ω), if there exist positive constants c0 and δ such that for all x, yΩ with ∣xy∣ < δ it holds

|p(x)p(y)|c0log(|xy|).

In the following context, we assume that p(x) : Ω → ℝ is a log-Hölder continuous function and there exist positive constants y1 and y2 such that

1<y1p(x)y2<,for all xΩ. (1.6)

Without loss of generality, let

|p(x)p(y)|ω(|xy|),for allx,yΩ, (1.7)

where ω: [0, ∞) → [0, ∞) is a modulus of continuity of p(x) such that ω is a nondecreasing continuous function with ω(0) = 0 and limsupr0ω(r)log(1r)<. With the above assumptions in hand, it is clear that p(x) ∈ LH(Ω) yields that there exists a positive number K0 such that

ω(r)log(1r)K0rω(r)eKOfor anyr(0,1). (1.8)

To obtain a global Calderón-Zygmund type estimate for general nonlinear elliptic problem (1.1), it is also necessary to impose some regular assumptions on the nonlinearity A = A(x, ξ) and the rough boundary of underlying domain Ω. Let us set

θ(A,Br(y))(x):=supξRn{0}|A(x,ξ)A¯Br(y)(ξ)|φ(|ξ|)

with

A¯Br(y)(ξ)=Br(y)A(x,ξ)dx.

Assumption 1.2

Let R0 > 0, we say that (A, Ω) is (δ, R0)-vanishing if

  1. sup0<rR0supyRnBr(y)θ(A,Br(y))(x)dxδ. (1.9)
  2. For any y Ω and for every number r ∈ (0, R0], there exists a coordinate system depending only on y and r, such that in this new coordinate system, y is the origin and

    Br{xn>δr}BrΩBr{xn>δr}. (1.10)

Remark 1.3

In this article, we always assume that δ is a small positive constant with 0 < δ < 18 by a scaling transformation. Note that the rough boundary with the so-called (δ, R0)-Reifenberg flatness (1.10) yields that the boundary might be locally very rough between two hyperplanes, which may go beyond the boundaries with C1-smooth or the Lipschitz category with a small Lipschitz constant. However, this is still an A-type domain with the following measure density condition

sup0<rR0supyΩ|Br(y)||Br(y)Ω|21δn167n, (1.11)

ensuring some natural properties in geometric analysis to hold, such as Sobolev embedding theorem and Sobolev extension theorem, see [11, 22, 25, 33].

Finally, we state the main result of this article.

Theorem 1.4

Let u W01,Φ (Ω) be a solution of (1.1) under the assumptions (1.2), (1.3) and (1.4) with the following higher integrability data

(Φ(|f|))p(x)Lt,q(Ω)fort>1,q(0,+].

If p(⋅) ∈ LH(Ω) with (1.6) and (1.7), and there exists a small constant δ = δ(n, ν, Λ, y1, y2, t, q, σφ, τφ, R0, K0, ∣Ω∣) > 0 such that (A, Ω) is (δ, R0)-vanishing with Assumption 1.2. Then we have (Φ(∣Du∣))p(x)Lt,q(Ω) with the estimate

(Φ(|Du|))p(x)Lt,q(Ω)c((Φ(|f|))p(x)Lt,q(Ω)+1)y2y1, (1.12)

where the constant c depends only on n, ν, Λ, y1, y2, t, q, σφ, τφ, R0, K0, ω(⋅) andΩ(except in the case q = ∞, where c depends on n, ν, Λ, y1, y2, t, σφ, τφ, R0, K0, ω(⋅) andΩ).

This article is devoted to a global Calderón-Zygmund type estimate in the framework of Lorentz spaces for a variable power of the gradients of weak solutions to the zero-Dirichlet problem of general nonlinear elliptic equations (1.1) over Reifenberg domains. As already mentioned, our problem and proof are inspired by work of Acerbi and Mingione [1, 2] and Baroni [4, 5], and recent work from Cho [14]. The key ingredient is to make use of the so-called large-M-inequality principle, Calderón-Zygmund-type covering, approximate estimate and an iteration argument to attain the variable Lorentz estimate (1.12).

The rest of the paper is organized as follows. In Section 2, we introduce notation and some useful lemmas. In Section 3, we focus on proving the main theorem.

2 Technical tools

In this section, we introduce some useful lemmas. From now on, we denote c to mean a universal constant that can be computed in terms of given data such as n, ν, Λ, y1, y2, t, q, σφ, τφ, R0, ω(⋅) and ∣Ω∣. First of all, we recall the existence and energy estimate of weak solution to general nonlinear elliptic problem (1.1).

Lemma 2.1

Let Ω ⊂ ℝn be a bounded domain, and fLΦ(Ω). Then there exists a unique solution u W01,Φ (Ω) to (1.1) such that the following energy estimate is valid:

ΩΦ(|Du|)dxcΩΦ(|f|)dx, (2.1)

where c = c(n, ν, Λ, σφ, τφ).

Proof

The existence and uniqueness of weak solution to (1.1) has been proved in [14]. Next, we take a test function ϕ = u W01,Φ (Ω) and use (1.2) and (1.4) to get the energy estimate (2.1).□

In what follows, let us give the scaling invariant property of (1.1), see Lemma 3.1 in [14].

Lemma 2.2

Let u W01,Φ (Ω) be the weak solution of (1.1) under the assumptions (1.2), (1.3) and (1.4). For fixed x0Ω, ρ > 0 and K > 0, we define

A~(x,ξ):=A(ρx,Kξ),G~(x,ξ):=G(ρx,Kξ),u~(x):=u(x0+ρx)Kρ,f~(x):=f(x0+ρx)K,φ~(t):=φ(Kt)

and the set Ω~={zx0ρ:zΩ}, where ξ ∈ ℝn and t ≥ 0. It leads to the following conclusions:

  1. ũ W01,Φ~ (Ω̃) is a weak solution to

    divA~(x,Du~)=divG~(x,f~)inΩ~,u~=0onΩ~,

    where

    Φ~(s)=0sφ~(r)dr=1KΦ(Ks).
  2. Ã and satisfy assumptions (1.2) and (1.4), respectively, with the same constants ν and L.

  3. φ̃ satisfies (1.3).

  4. If (A, Ω) is (δ, R0)-vanishing, then (Ã, Ω̃) is (δ,R0ρ) -vanishing.

Next, we recall a reverse Hölder inequality for Φ(∣ξ∣), see Theorem 9 in [16].

Lemma 2.3

Let u W01,Φ (Ω) be the weak solution of (1.1) under the assumptions (1.2), (1.3) and (1.4). Suppose that (Φ(∣f∣))p(x)Lt for p(x) > y1 > 1 and t > 1. Then there exists σ0 = σ0(n, ν, Λ, σφ, τφ) ∈ (0, ty1 − 1) such that for Br with B2r ⊂ ⊂ Ω and any σ ∈ (0, σ0], it holds

(Br(Φ(|Du|))1+σdx)11+σcB2rΦ(|Du|)dx+c(B2r(Φ(|f|))1+σdx)11+σ,

where c = c(n, ν, Λ, σφ, τφ) > 0.

In addition, the following reverse Hölder inequality on the boundary version is also a self-improving result due to the Reifenberg flatness domain belonging to A-type condition as in (1.11).

Lemma 2.4

Let u W01,Φ (Ω) be the weak solution of (1.1) under the assumptions (1.2), (1.3) and (1.4). Suppose that (Φ(∣f∣))p(x)Lt for p(x) > y1 > 1 and t > 1, and (A, Ω) satisfies (δ, R0)-vanishing with Assumption 1.2. Then there exists σ0 = σ0(n, ν, Λ, σφ, τφ) ∈ (0, ty1 − 1) such that for Ωr = ΩBr(x0) with x0Ω and any σ ∈ (0, σ0], it holds

(Ωr(Φ(|Du|))1+σdx)11+σcΩ2rΦ(|Du|)dx+c(Ω2r(Φ(|f|))1+σdx)11+σ,

where c = c(n, ν, Λ, σφ, τφ) > 0.

Proof

By a similar procedure to Theorem 9 in [16] and using the measure density property (1.11) for Ω, and a zero-extension of u in B2r, the conclusion is clearly true.□

Note that Φ(s) is an N-function from the definition of Φ(s) in (1.5). Besides, we deduce that

min{ασφ+1,ατφ+1}Φ(s)Φ(αs)max{ασφ+1,ατφ+1}Φ(s),for anys,α0, (2.2)

and

Φ(s1+s2)12Φ(2s1)+12Φ(2s2),fors1,s20. (2.3)

We define an auxiliary vector field V : ℝn → ℝn given by

V(ξ):=(φ(|ξ|)|ξ|)1/2ξ,for eachξRn.

Then there is the following relation between V and Φ, see [14]:

Φ(|ξη|)m0(|V(ξ)V(η)|2Φ(η)), (2.4)

for all ξ, η ∈ ℝn and m0 > 0 depending only on n, ν, Λ, σφ and τφ.

Now, we give the comparison estimates with a limiting problem, and show its Lipschitz regularity by following from Proposition 5.5 and 5.11 in [14]. We state the required comparison estimates in the following lemmas.

For the interior case, we assume

B5θ(A,B5)dxδ, (2.5)

and

B5Φ(|Du|)dx1andB5(Φ(|f|))ηdxδy1y2η (2.6)

for η = 1+σ2 ≤ 1+σ with σ as the same of Lemma 2.3, and δ (0,18) determined later.

Lemma 2.5

Let u W01,Φ (Ω) be a weak solution of (1.1) under the assumptions (1.2), (1.3) and (1.4). If for any 0 < ϵ < 1, there exists a constant δ = δ(n, ϵ, ν, Λ, σφ, τφ) > 0 such that (2.5) and (2.6) hold. Then there exists a weak solution v W01,Φ (B3) of

divA¯B3(Dv)=0inB3,v=uonB3,

where ĀB3 = B3 A(x, ξ)dx, such that we have

B1|V(Du)V(Dv)|2dxϵ

and

Φ(|Dv|)L(B1)c1,

where c1 = c1(n, ν, Λ, σφ, τφ) > 1.

For the boundary case, let Ω5 satisfy that

B5+Ω5B5{xRn,xn>5δ}, (2.7)
B5+θ(A,B5+)dxδ, (2.8)

and

Ω5Φ(|Du|)dx1andΩ5(Φ(|f|))ηdxδy1y2η. (2.9)

Lemma 2.6

Let u W01,Φ (Ω) be a weak solution of (1.1) under the assumptions (1.2), (1.3) and (1.4). If for any 0 < ϵ < 1, there exists a constant δ = δ(n, ϵ, ν, Λ, σφ, τφ) > 0 such that (2.7), (2.8) and (2.9) hold. Then there exists a weak solution v W01,Φ(B2+) of

divA¯B2+(Dv)=0inB2+,v=uonT2,

where A¯B2+=B2+A(x,ξ)dx, and we have

Ω1|V(Du)V(Dv¯)|2dxϵ

and

Φ(|Dv¯|)L(Ω1)c2,

where c2 = c2(n, ν, Λ, σφ, τφ) > 1 and is zero extension of v from B1+ to Ω1.

Let us now collect some preliminary results concerning the so-called embedding relations involved in the Lorentz spaces, which will be used in the sequel.

Proposition 2.7

Let U be a bounded measurable subset ofn, then the following relations hold:

  1. If 0 < q ≤ ∞, and 1 ≤ t1 < t2 < ∞, then Lt2,q(U) ⊂ Lt1,q(U) with the estimate

    gLt1,q(U)c|U|1t11t2gLt2,q(U). (2.10)
  2. If 1 ≤ t < ∞, and 0 < q1 < q2 ≤ ∞, then Lt,q1(U) ⊂ Lt,q2(U) ⊂ Lt,∞(U) with the estimate

    gLt,q2(U)c(t,q1,q2)gLt,q1(U). (2.11)
  3. IfgαLt,q(U) for some 0 < α < ∞, then gLαt,αq(U) with the estimate

    |g|αLt,q(U)=gLαt,αq(U)α. (2.12)

    The following two lemmas will play important roles in our main proof, which are the variants of classical Hardy’s inequality and a reverse Hölder inequality, respectively, see Lemma 3.4 and 3.5 in [4].

Lemma 2.8

Let ψ: [0, +∞) → [0, +∞) be a measurable function such that

0ψ(λ)dλ<. (2.13)

Then for any α ≥ 1 and r > 0, there holds

0λrλψ(μ)dμαdλλαrα0λr(λψ(λ))αdλλ.

Lemma 2.9

Let h : [0, +∞) → [0, +∞) be a nonincreasing, measurable function. For α1α2 ≤ ∞ and r > 0, if α2 < ∞, then we have

λμrh(μ)α2dμμ1α2ελrh(λ)+cεα2/(α11)λμrh(μ)α1dμμ1α1 (2.14)

with every ε ∈ (0, 1] and λ ≥ 0. If α2 = ∞, then it holds

supμ>λμrh(μ)cλrh(λ)+cλμrh(μ)α1dμμ1α1, (2.15)

where the constant c depends only on α1,α2 and r; except in the case α2 = ∞ with cc(α1, r).

3 Proof of Theorem 1.4

This section is mainly devoted to proving Theorem 1.4. Our proof consists of 6 steps. In step 1, for given λ0 in (3.7), we show the Calderón-Zygmund type covering on the super-level set E(λ, Ωr1(x0)), and establish a decay estimate of Ωry(y). In step 2, we give various comparison estimates to the reference problems. In step 3, we employ the so-called “crawling of ink spots” approach to show an estimate for the super-level set. In steps 4, we get the estimate of ∥(Φ(∣Du∣))p(x)Lt,q(Ωr1(x0)) for q < ∞. In steps 5 and 6, we deduce our conclusions in the cases of q < ∞ and q = ∞, respectively, under a priori assumption ∥(Φ(∣Du∣))p(x)Lt,q(Ωr2(x0)) < ∞ which is proved in step 5.

Proof

Let σ0 be the same as in Lemma 2.4, and let σ2 = min{σ0,y1 − 1} > 0 due to y1 > 1. For a fixed point yΩ, we take ry < R250 with

RminR02,R0c,1andω(4R)<y1σ2, (3.1)

where c* = c*(n, y1, y2, ν, Λ, σφ, τφ, ω(⋅), ∣Ω∣) ≥ ∣Ω∣ + 1 determined later.

Here, we only consider the boundary case, saying B25ry(y) ⊄ Ω. For this, we can find a boundary point ȳB25ry (y) ∩ Ω and for ry (0,R250), there exists a coordinate system depending only on ȳ and ry, such that in this new coordinate system it holds

z=y,y¯+125δry(0,,0,1)is the origin,B125ry+(0)Ω125ry(0)B125ry(0)zn>250δry. (3.2)

We select 0 < δ < 18 such that Ω5ry(z) ⊂ Ω50ry(0), which implies the fact that

Ω125ry(0)Ω170ry(z). (3.3)

Then we also have

B125ry+(0)θ(A,B125ry+(0))(x)dx4B125ry(0)θ(A,B125ry(0))(x)dx4δ. (3.4)

For a fixed point x0Ω, we set

p:=infΩ2R(x0)p(x),p+:=supΩ2R(x0)p(x);py:=infΩ125ryp(x),py+:=supΩ125ryp(x).

For the weak solution u of original problem, by a scaling argument to the nonhomogeneous terms f and the N-functions φ, Φ, then we write

u¯=u((Φ(|f|))p(x)Lt,q(Ω)+1)1/y1,f¯=f((Φ(|f|))p(x)Lt,q(Ω)+1)1/y1,φ¯(s)=φ(((Φ(|f|))p(x)Lt,q(Ω)+1)1/y1s),Φ¯(s)=Φ(((Φ(|f|))p(x)Lt,q(Ω)+1)1/y1s)((Φ(|f|))p(x)Lt,q(Ω)+1)1/y1. (3.5)

By the assumption (Φ(∣f∣))p(x)Lt,q(Ω), then there holds

(Φ¯(|f¯|))p(x)Lt,q(Ω)=(Φ(|f|)((Φ(|f|))p(x)Lt,q(Ω)+1)1/y1)p(x)Lt,q(Ω)(Φ(|f|))p(x)Lt,q(Ω)(Φ(|f|))p(x)Lt,q(Ω)+11. (3.6)

Hereafter, for the sake of simplicity, we still use u, f, φ and Φ replacing ū, , φ̄ and Φ̄ in the following.

  1. In this step, we give a Calderón-Zygmund type covering on the super-level set E(λ, Ωr1(x0)) as below. Let u be the weak solution of (1.1), we define the quantity for any r1 and r2 with Rr1r2 ≤ 2R

    λ0:=Ωr2(x0)(Φ(|Du|))p(x)pdx+1δ(Ωr2(x0)(Φ(|f|))p(x)pηdx+1)1η, (3.7)

    where δ > 0 and η > 1 will be specified later. We treat the super-level set

    Eλ,Ωr1(x0):={xΩr1(x0),(Φ(|Du|))p(x)p>λ}

    for any λ > 0 ≥ 1 with M=(2320r27(r2r1))n. For yE(λ, Ωr1(x0)) and radii 0 < rr2-r1, we let

    CZ(Ωr(y)):=Ωr(y)(Φ(|Du|))p(x)pdx+1δ(Ωr(y)(Φ(|f|))p(x)pηdx)1η. (3.8)

    If r2r1145 rr2-r1, then we discover that

    CZ(Ωr(y))|Ωr2(x0)||Ωr(y)|Ωr2(x0)(Φ(|Du|))p(x)pdx+(|Ωr2(x0)||Ωr(y)|)1η1δ(Ωr2(x0)(Φ(|f|))p(x)pηdx)1η|Ωr2(x0)||Ωr(y)|(Ωr2(x0)(Φ(|Du|))p(x)pdx+1δ(Ωr2(x0)(Φ(|f|))2p(x)pηdx)1η)|Br2(x0)||Br(y)||Br(y)||Ωr(y)|(Ωr2(x0)(Φ(|Du|))p(x)pdx+1δ(Ωr2(x0)(Φ(|f|))2p(x)pηdx)1η)(r2r)n(167)nλ0(2320r27(r2r1))nλ0<λ,

    which means that while r2r1145 rr2r1 one has CZ(Ωr(y)) < λ. At the same time, by Lebesgue’s differentiation theorem we find that CZ(Ωr(y)) > λ for 0 < r ≪ 1. Therefore, by absolute continuity of the integral with respect to the domain we can pick the maximal radius ry such that

    CZ(Ωry(y))=Ωry(y)(Φ(|Du|))p(x)pdx+1δ(Ωry(y)(Φ(|f|))p(x)pηdx)1η=λ (3.9)

    for each point yE(λ, Ωr1(x0)). Moreover, one has

    CZ(Ωr(y))<λ,for anyr(ry,r2r1]. (3.10)

    From (3.9), we conclude the following alternatives:

    λ2Ωry(y)(Φ(|Du|))p(x)pdxor(δλ2)ηΩry(y)(Φ(|f|))p(x)pηdx. (3.11)

    First, we suppose that the first case of (3.11) is valid to have

    Ωry(y)(Φ(|Du|))p(x)pdx=|Ωry(y)E(λ4,Ωr2(x0))||Ωry(y)|Ωry(y)E(λ4,Ωr2(x0))(Φ(|Du|))p(x)pdx+1|Ωry(y)|Ωry(y)E(λ4,Ωr2(x0))(Φ(|Du|))p(x)pdxλ4+c(|Ωry(y)E(λ4,Ωr2(x0))||Ωry(y)|)111+σ1(Ωry(y)(Φ(|Du|))p(x)p(1+σ1)dx)11+σ1, (3.12)

    where σ1 > 0 is determined later.

    Considering that ω(4R) < y1σ2 in (3.1), it leads to y1(1+σ2)y1+ω(4R)1>0. Now let us take

    0<σ1y1(1+σ2)y1+ω(4R)1,

    which yields that

    py+p(1+σ1)=(1+p+pp)(1+σ1)(1+ω(4R)p)(1+σ1)(1+σ2),

    where py+:=supΩ125ryp(x) with Ω125ryΩ2R(x0), and ω(⋅) is the modulus of continuity for p(x). Then we use the reverse Hölder inequality shown in Lemma 2.4 to obtain that

    (Ωry(y)(Φ(|Du|))p(x)p(1+σ1)dx)11+σ1(Ωry(y)(Φ(|Du|))py+p(1+σ1)dx+1)11+σ1c(Ω2ry(y)(Φ(|Du|))dx)py+p+c(Ω2ry(y)(Φ(|f|))py+p(1+σ1)dx)11+σ1+cc(Ω2ry(y)Φ(|Du|)dx)py+p+c(Ω2ry(y)(Φ(|f|))1+σ2dx)py+p11+σ2+c.

    Taking into account (3.10), we have CZ(Ω2ry(y)) < λ, then by a similar proof of (3.19) in Step 2, we get that

    Ω2ry(y)Φ(|Du|)dxcλppy+,Ω2ry(y)(Φ(|f|))ηdxcλppy+ηδy1y2η.

    Thus, by taking η = 1 + σ2 we have

    (Ωry(y)(Φ(|Du|))p(x)p(1+σ1)dx)11+σ1c(λ+λδy1y2py+p+1)cλ. (3.13)

    Therefore, we combine (3.11), (3.12) and (3.13) to have

    λ4c(|Ωry(y)E(λ4,Ωr2(x0))||Ωry(y)|)111+σ1λ,

    which implies

    |Ωry(y)|c|Ωry(y)E(λ4,Ωr2(x0))| (3.14)

    with the positive constant c depending only on n, ν, Λ, y2, y2, σφ, τφ, R0, K0, and ∣Ω∣.

    If the case of the second estimate in (3.11) is valid, by taking ζ=δ4 and Fubini’s theorem, we get

    (λδ2)ηΩry(y)(Φ(|f|))p(x)pηdx=η|Ωry(y)|0μη|{xΩry(y):(Φ(|f|))p(x)p>μ}|dμμ=η|Ωry(y)|0ζλμη|{xΩry(y):(Φ(|f|))p(x)p>μ}|dμμ+η|Ωry(y)|ζλμη|{xΩry(y):(Φ(|f|))p(x)p>μ}|dμμ(ζλ)η+η|Ωry(y)|ζλμη|{xΩry(y):(Φ(|f|))p(x)p>μ}|dμμ.

    Let δ = 4ζ, we derive that

    |Ωry(y)|η(ζλ)ηζλμη|{xΩry(y):(Φ(|f|))p(x)p>μ}|dμμ. (3.15)

    Now we put (3.14) and (3.15) together to get that

    |Ωry(y)|c|Ωry(y)E(λ4,Ωr2(x0))|+cη(ζλ)ηζλμη|{xΩry(y):(Φ(|f|))p(x)p>μ}|dμμ. (3.16)
  2. This step is devoted to various comparison estimates with the reference problems and the limiting one. Note that Ω is (δ, R0)-Reifenberg flat, then it follows from (3.3) and (3.10) that

    Ω125ry(Φ(|Du(z)|))p(z)pdz+1δ(Ω125ry(Φ(|f|))p(z)pηdz)1η2(3425)n(Ω170ry(y)(Φ(|Du(z)|))p(z)pdz+1δ(Ω170ry(y)(Φ(|f|))p(z)pηdz)1η)2(3425)nλ, (3.17)

    which implies that

    Ω125ry(Φ(|Du|))p(x)pdxcλ,(Ω125ry(Φ(|f|))p(x)pηdx)1ηcλδ, (3.18)

    where we still use variable x for simplicity. Therefore, it suffices to show that

    /Ω125ryΦ(|Du|)dxc3λppy+,Ω125ry(Φ(|f|))ηdxc3λppy+ηδy1y2η (3.19)

    for a constant c3 ≥ 1. We first claim that

    (Ω125ryΦ(|Du|)dx)py+pyc4 (3.20)

    with c4 ≥ 1 a universal constant. In fact, since (Φ(∣f∣))p(x)Lt,q(Ω) for t > 1 and 0 < q ≤ ∞, it deduces that

    ΩΦ(|f|)dxΩ(Φ(|f|)+1)p(x)dxc(Φ(|f|))p(x)Lt,q(Ω)+c|Ω|c(1+|Ω|),

    where we have used (3.6). By the energy estimate of Lemma 2.1, it leads to that

    ΩΦ(|Du|)dxc(1+|Ω|). (3.21)

    By considering py+pyω(250ry), we have

    (Ω125ryΦ(|Du|)dx)py+py=(1|Ω125ry|)py+py(Ω125ryΦ(|Du|)dx)py+pyc(1|B250ry|)py+py(Ω125ryΦ(|Du|)dx)py+pyc(1250ry)nω(250ry)(Ω125ryΦ(|Du|)dx)py+pyc(Ω125ryΦ(|Du|)dx)py+py.

    On the other hand, we use (3.21) and 1250ry1RcR0|Ω|+1 with (3.1) to find

    (Ω125ryΦ(|Du|)dx)py+py(ΩΦ(|Du|)dx)py+pyc(|Ω|+1)py+pyc(1250ry)ω(250ry)c,

    where we have used the so-called log-Hölder continuity (1.8) for p(x) in the last inequality, which yields (3.20). Recalling y1py+ and (3.20) with λ > 1, we obtain

    Ω125ryΦ(|Du|)dx=(Ω125ryΦ(|Du|)dx)py+pypy+(Ω125ryΦ(|Du|)dx)pypy+c41y1(Ω125ryΦ(|Du|)dx)pypy+c(Ω125ry(Φ(|Du|))pypdx)ppy+c(Ω125ry(Φ(|Du|))p(x)pdx+1)ppy+cλppy+.

    Note that (Φ(∣f∣))p(x)Lt,q(Ω) for 1 < y1p(x) ≤ y2 < ∞, and 1 < η ≤ 1+σ with σ as the same of Lemma 2.4, it holds

    Ω(Φ(|f|))ηdxΩ((Φ(|f|))1+σ+2)dxc(1+|Ω|).

    Similarly, recalling δλ0 ≥ 1 and λ0 we find

    Ω125ry(Φ(|f|))ηdxc(Ω125ry(Φ(|f|))p(x)pηdx+1)ppy+c(δλ+1)ppy+ηc(δλ+δλ0)ppy+ηcλppy+ηδy1y2η.

    Next, we define

    A~(x,ξ):=A(25ryx,λppy+ξ),G~(x,ξ):=G(25ryx,λppy+ξ),
    u~(x):=u(25ryx)25ryλppy+,f~(x):=f(25ryx)λppy+,φ~(s):=φ(λppy+s),Φ~(s):=Φ(λppy+s)λppy+.

    In light of (3.2), (3.4), (3.19) and Lemma 2.2, we find that the hypothesis of Lemma 2.6 is valid, which implies that

    Ω25ry|V(Du)V(Dv¯)|2dxϵλppy+,

    and

    Φ(|Dv¯|)L(Ω25ry)c0λppy+,

    where ϵ > 0 is small and c0 = max{c1, c2} ≥ 1.

    For the case of interior, by Lemma 2.5 similarly we have

    B25ry|V(Du)V(Dv)|2dxϵλppy+,

    and

    Φ(|Dv|)L(B25ry)c0λppy+.
  3. We are here to estimate the super level set E(λ, Ωr2(x0)). For any fixed point xΩ, we select a universal constant R with 0 < R min{R02,R0|Ω|+1,1}, and there exists a constant δ = δ(ϵ) > 0 such that Lemma 2.5 and 2.6 hold. Let A=(2τφ+1c0(m0+1))y2y1, for any xE(, Ωr1(x0)) we consider the collection 𝓑λ of all subset Ωry(y). By“ crawling of ink spots” argument, we extract a countable subcollection {Ωri(yi)} ∈ 𝓑λ, such that

    Ωri(yi)Ωrj(yj)=,wheneverij,andE(Aλ,Ωr1(x0))iNΩ5ri(yi)Nλ

    with ∣𝓝λ∣ = 0. Let us denote pi+=pyi+, then we derive that

    |E(Aλ,Ωr1(x0))|=|E((2τφ+1c0(m0+1))y2y1λ,Ωr1(x0)|=|{xΩr1(x0):(Φ(|Du|))p(x)p>(2τφ+1c0(m0+1))y2y1λ}|i1|{xΩ5ri(yi):Φ(|Du|)>2τφ+1c0(m0+1)λpp(x)}|i1|{xΩ5ri(yi):Φ(|Du|)>2τφ+1c0(m0+1)λppi+}|= interior case|{xΩ5ri(yi):Φ(|Du|)>2τφ+1c0(m0+1)λppi+}|+ boundary case|{xΩ5ri(yi):Φ(|Du|)>2τφ+1c0(m0+1)λppi+}|. (3.22)

    For the boundary case, we recall (2.2), (2.3) and (2.4) to find that

    Φ(|Du|)Φ(|DuDv¯|+|Dv¯|)12(Φ(2|DuDv¯|)+Φ(2|Dv¯|))2τφ(Φ(|DuDv¯|)+Φ(|Dv¯|))2τφ(m0+1)(|V(Du)V(Dv¯)|2+Φ(|Dv¯|)),

    which implies that

    |{xΩ5ri(yi):Φ(|Du|)>2τφ+1c0(m0+1)λppi+}||{xΩ50ri:|V(Du)V(Dv¯)|2>c0λppi+}|+|{xΩ50ri:Φ(|Dv¯|)>c0λppi+}|1c0λppi+Ω50ri|V(Du)V(Dv¯)|2dxcϵ|Ω50ri|cϵ|B50ri|=cϵ|Bri(yi)|cϵ|Bri(yi)||Ωri(yi)||Ωri(yi)|cϵ(21δ)n|Ωri(yi)|cϵ|Ωri(yi)|,

    where we used the following weak (1, 1)−type estimate:

    |{xU:g(x)>λ}|1λUg(x)dx.

    Similarly, for the interior case, we discover that

    |{xΩ5ri(yi):Φ(|Du|)>2τφ+1c0(m0+1)λppi+}|=|{xB5ri(yi):Φ(|Du|)>2τφ+1c0(m0+1)λppi+}|cϵ|Bri(yi)|.

    Therefore, it follows from (3.22) that

    |E(Aλ,Ωr1(x0))|cϵi1|Ωri(yi)|. (3.23)

    Using“ crawling of ink spots” argument again and (3.16), we conclude that

    |E(Aλ,Ωr1(x0))|cϵi1|Ωri(yi)E(λ4,Ωr2(x0))|+cϵη(ζλ)ηi1ζλμη|{xΩri(yi):(Φ(|f|))p(x)p>μ}|dμμcϵ|E(λ4,Ωr2(x0))|+cϵη(ζλ)ηζλμη|{xΩr2(x0):(Φ(|f|))p(x)p>μ}|dμμ. (3.24)
  4. This step is devoted to the estimate of ∥(Φ(∣Du∣))p(x)Lt,q(Ωr1(x0)) for 0 < q < ∞. Since t > 1, we multiply the inequality (3.24) by (tp)tq(Aλ)tp, and integrate the resulting expression with a power qt in the measure dλAλ from 0 to ∞, which yields that

    tpMλ0((Aλ)tp|{xΩr1(x0):(Φ(|Du|))p(x)p>Aλ}|)qtdλAλctpϵqt0(λtp|{xΩr2(x0):(Φ(|Du|))p(x)p>λ4}|)qtdλλ+ctpϵqt0λq(pηt)(ζλμη|{xΩr2(x0):(Φ(|f|))p(x)p>μ}|dμμ)qtdλλ=cϵqt(I1+I2), (3.25)

    where c depends on n, ν, Λ, y1, y2, t, q, σφ, τφ, R0, K0,∣Ω∣, and ω(⋅). Thanks to (2.12) in Proposition 2.7, we have

    (Φ(|Du|))p(x)Lt,q(Ωr1(x0))q=(Φ(|Du|))p(x)pLtp,qp(Ωr1(x0))qp=tp0(μtp|{xΩr1(x0):(Φ(|Du|))p(x)p>μ}|)qtdμμ. (3.26)

    By a simple change of variable and (3.26) it leads to that

    I1=c(q)(Φ(|Du|))p(x)Lt,q(Ωr2(x0))q.

    We are now to estimate I2. For this we part it in two cases.

  5. If qt, noticing that (2.13) is satisfied since (Φ(|f|))p(x)pLη(Ωr2). By making the change of variables λ̄ = ζλ and ζ=δ4, then we employ Lemma 2.8 with ψ(μ)=μη1|{xΩr2(x0):(Φ(|f|))p(x)P>μ}|,α=qt1, r=q(pηt)>0 and (3.26) to infer

    I2=ctp0λ¯q(pηt)(λ¯μη|{xΩr2(x0):(Φ(|f|))p(x)p>μ}|dμμ)qtdλ¯λ¯ctp0λ¯qp|{xΩr2(x0):(Φ(|f|))p(x)p>λ¯}|qtdλ¯λ¯=c(Φ(|f|))p(x)Lt,q(Ωr2(x0))q,

    where c = c(y1, y2, t, q).

  6. If 0 < q < t, we use Lemma 2.9 with h(μ)=|{xΩr2(x0):(Φ(|f|))p(x)p>μ}|qt,r=ηqt,α1=1<tq=α2 and ε = 1, which yields the following

    (λμη|{xΩr2(x0):(Φ(|f|))p(x)p>μ}|dμμ)qtληqt|{xΩr2(x0):(Φ(|f|))p(x)p>λ}|qt+cλμηqt|{xΩr2(x0):(Φ(|f|))p(x)p>μ}|qtdμμ.

    After a change variable ζ λλ, (3.26) and Fubini’s theorem we get

    I2ctp0λq(pηt)ληqt|{xΩr2(x0):(Φ(|f|))p(x)p>λ}|qtdλλ+ctp0λq(pηt)λμηqt1|{xΩr2(x0):(Φ(|f|))2p(x)p>μ}|qtdμdλλc(Φ(|f|))p(x)Lt,q(Ωr2(x0))q+ctp0λq(pηt)(λμηqt1|{xΩr2(x0):(Φ(|f|))p(x)p>μ}|qtdμ)dλλc(Φ(|f|))p(x)Lt,q(Ωr2(x0))q,

    where c = c(y1, y2, t, q).

    We are now in a position to put the estimates of I1 and I2 into (3.25), and after simple manipulation, then for t > 1 it follows that

    (Φ(|Du|))p(x)Lt,q(Ωr1(x0))c(tpMλ0((Aλ)tp|{xΩr1(x0):(Φ(|Du|))p(x)p>Aλ}|)qtd(Aλ)Aλ)1q+c(tp0Mλ0((Aλ)tp|{xΩr1(x0):(Φ(|Du|))p(x)p>Aλ}|)qtd(Aλ)Aλ)1qc(tp0Mλ0((Aλ)tp|{xΩr1(x0):(Φ(|Du|))p(x)p>Aλ}|)qtd(Aλ)Aλ)1q+c¯ϵ1t((Φ(|Du|))p(x)Lt,q(Ωr2(x0))+(Φ(|f|))p(x)Lt,q(Ωr2(x0)))cλ0p|Ωr2(x0)|1t+c¯ϵ1t((Φ(|Du|))p(x)Lt,q(Ωr2(x0))+(Φ(|f|))p(x)Lt,q(Ωr2(x0))), (3.27)

    where = (n, ν, Λ, y1, y2, t, q, σφ, τφ, R0, K0, ∣Ω∣, ω(⋅)) > 0. Now we choose ϵ > 0 small enough such that c¯ϵ1t12, then we can find a corresponding positive constant δ = δ(n, ϵ, ν, Λ, y1, y2, t, q, σφ, τφ, R0, K0, ∣Ω∣) such that we deduce

    (Φ(|Du|))p(x)Lt,q(Ωr1(x0))cλ0p|Ωr2(x0)|1t+12(Φ(|Du|))p(x)Lt,q(Ωr2(x0))+c(Φ(|f|))p(x)Lt,q(Ωr2(x0)). (3.28)
  7. This step is devoted to proving that ∥(Φ(∣Du∣))p(x)Lt,q(Ωr2(x0)) < ∞. To this end, we first refine the estimate of (Φ(∣Du∣))p(x) in the scale of Lorentz spaces. Consider the following truncated function

    |(Φ(|Du|))p(x)|k=min{(Φ(|Du|))p(x),k}forxΩandkN[Mλ0,).

    By considering Ek(λ, Ωρ) = {xΩρ : ∣(Φ(∣Du∣))p(x)k > λ} in line with (3.24), we discover that

    Ek(Aλ,Ωr1(x0))cϵ|Ek(λ4,Ωr2(x0))|+cϵη(ζλ)ηζλμη|{xΩr2(x0):(Φ(|f|))p(x)p>μ}|dμμ. (3.29)

    For 0 < k we have Ek(, Ωr1(x0)) = ∅, which implies that the above estimate holds trivially. For k > , the estimate is also valid since

    Ek(Aλ,Ωr1(x0))=E(Aλ,Ωr1(x0))={xΩr1(x0),(Φ(|Du|))p(x)>Aλ}andEk(λ4,Ωr2(x0))=E(λ4,Ωr2(x0)).

    Then working exactly as in Step 4, we get that (3.28) holds with ∣(Φ(∣Du∣))p(x)k in place of (Φ(∣Du∣)p(x).

    Now we let L=cλ0p|Ω2R(x0)|1t+c(Φ(|f|))p(x)Lt,q(Ω2R(x0)) for a fixed small radius R > 0, Θ(ρ) = ∥∣(Φ(∣Du∣))p(x)kLt,q(Ωρ(x0)) for any Rρ < 2R, and take a small ϵ > 0 such that cϵ=12 in (3.28). Then for any Rρ1 < ρ2 < ρ3 < ⋯ < 2R with ρd → 2R, d → ∞,

    Θ(ρ1)12Θ(ρ2)+L,Θ(ρ2)12Θ(ρ3)+L,

    by the iteration we get that

    Θ(R)12dΘ(ρd)+Ll=0d112l.

    Note that

    Θ(2R)=|(Φ(|Du|))p(x)|kLt,q(Ω2R(x0))<,

    we let d → ∞ to get

    |(Φ(|Du|))p(x)|kLt,q(ΩR(x0))cλ0p|Ω2R(x0)|1t+c(Φ(|f|))p(x)Lt,q(Ω2R(x0)).

    In what follows, we use a standard finite covering argument to realize our global estimate. Note that Ω is a bounded domain in ℝn and let us now take x0 as every point in Ω. Then there exist N = N(n, ∣Ω∣) ∈ ℕ and xjΩ for j = 1, 2, ⋯, N, where we replace the point x0 by each xj, such that

    Ω¯j=1NΩR(xj).

    Therefore, we deduce that

    |(Φ(|Du|))p(x)|kLt,q(Ω)j=1N|(Φ(|Du|))p(x)|kLt,q(ΩR(xj))cj=1N(λ0p|Ω2R(xj)|1t+(Φ(|f|))p(x)Lt,q(Ω2R(xj)))cj=1N|Ω2R(xj)|1t(Ωr2(xj)(Φ(|Du|))p(x)pdx+(Ωr2(xj)((Φ(|f|))p(x)pη+1)dx)1η)p+cj=1N(Φ(|f|))p(x)Lt,q(Ω2R(xj))cj=1N|Ω|1t(Ωr2(xj)(Φ(|Du|))p(x)pdx+(Ωr2(xj)((Φ(|f|))p(x)pη+1)dx)1η)p+cN(Φ(|f|))p(x)Lt,q(Ω), (3.30)

    where we use the definition of λ0. By (3.1), we notice that

    p+p=1+p+pp1+ω(4R)y11+σ,

    where σ is the same as in Lemma 2.4. Then, it yields that

    Ωr2(xj)(Φ(|Du|))p(x)pdxΩr2(xj)(Φ(|Du|))p+pdx+1c(Ω2r2(xj)(Φ(|Du|))dx+1)p+p+cΩ2r2(xj)(Φ(|f|))p+pdx, (3.31)

    where we employed the reverse Höder inequality of Lemma 2.4 in the last inequality. Using (2.1) and Höder inequality, we obtain that

    (Ω2r2(xj)Φ(|Du|)dx)p+p(1|Ω2r2(xj)|)p+p(Ω(Φ(|Du|))dx)p+pc(1|Ω2r2(xj)|)p+p(ΩΦ(|f|)dx)p+pc(1|Ω2r2(xj)|)p+p|Ω|1pp+Ω(Φ(|f|))p+pdxc(1|Ω2r2(xj)|)p+pΩ((Φ(|f|))p(x)pp+p+1)dx. (3.32)

    We now combine (3.30), (3.31) and (3.32) to get that

    |(Φ(|Du|))p(x)|kLt,q(Ω)cj=1N((1|Ω2r2(xj)|)p+pΩ((Φ(|f|))p(x)pp+p+1)dx+(1|Ωr2(xj)|)1η(Ω((Φ(|f|))p(x)pη+1)dx)1η)p+cN(Φ(|f|))p(x)Lt,q(Ω). (3.33)

    Using a standard Hardy’s inequality in the Marcinkiewicz spaces (cf. Lemma 2.8 in [26]) and the reverse Hölder inequality of Lemma 2.9, we conclude that

    Ω(Φ(|f|))p(x)pp+pdxt(p)2t(p)2p+|Ω|1p+t(p)2(Φ(|f|))p(x)pMtp(Ω)p+p=t(p)2t(p)2p+|Ω|1p+t(p)2(suph>0(htp|{xΩ:(Φ(|f|))p(x)p>h}|)1tp)p+pc|Ω|1p+t(p)2(Φ(|f|))p(x)pLtp,qp(Ω)p+pc|Ω|1p+t(p)2(Φ(|f|))p(x)Lt,q(Ω)p+(p)2. (3.34)

    Similarly, we also show that

    (Ω(|Φ(|f|))p(x)pηdx)1ηc(y1,y2,t,q)|Ω|1η1tp(Φ(|f|))p(x)pLtp,qp(Ω)c(y1,y2,t,q)|Ω|1η1tp(Φ(|f|))p(x)Lt,q(Ω)1p.

    For the case of q < ∞, from (3.33) we then infer the following relations

    |(Φ(|Du|))p(x)|kLt,q(Ω)cj=1N((1|Ω2r2(xj)|)p+((Φ(|f|))p(x)Lt,q(Ω)p+p+1)+(1|Ωr2(xj)|)pη((Φ(|f|))p(x)Lt,q(Ω)+1))cj=1N(1|B2r2(xj)||B2r2(xj)||Ω2r2(xj)|)p+((Φ(|f|))p(x)Lt,q(Ω)p+p+1)+cj=1N(1|Br2(xj)||Br2(xj)||Ωr2(xj)|)pη((Φ(|f|))p(x)Lt,q(Ω)+1)cj=1N(1|B2R(xj)|(21δ)n)p+((Φ(|f|))p(x)Lt,q(Ω)p+p+1)+cj=1N(1|BR(xj)|(21δ)n)pη((Φ(|f|))p(x)Lt,q(Ω)+1)cN((Φ(|f|))p(x)Lt,q(Ω)+1)y2y1c,

    where we used the uniformly estimate (3.6). Now let us take k → ∞. By the lower semi-continuity of Lorentz quasi-norm we have

    (Φ(|Du|))p(x)Lt,q(Ω)c,

    where c depends only on n, ν, Λ, y1, y2, t, q, σφ, τφ, R0, K0, ω(⋅), and ∣Ω∣. And recalling the definition in (3.5), we get the desired result (1.12)

  8. Finally, for the case of q = ∞, we come back to the second inequality in (3.11) and split it into two parts with a small ι > 0 determined later:

    (λ2)η1δηΩry(y)(Φ(|f|))p(x)pηdx(ιλ)ηδη+1δη|Ωry(y)|{xΩry(y):(Φ(|f|))p(x)p>ιλ}(Φ(|f|))p(x)pηdx.

    We set

    Ψ(ιλ,Ωry(y))={xΩry(y):(Φ(|f|))p(x)p>ιλ},Ψ(μ,Ωry(y))={xΩry(y):(Φ(|f|))p(x)p>μ}.

    Similar to the estimate (3.34), by using Hölder inequality we get

    (λ2)η(ιλδ)η1δη|Ωry(y)|{xΩry(y):(Φ(|f|))p(x)p>ιλ}(Φ(|f|))p(x)pηdxt(tη)δη|Ψ(ιλ,Ωry(y))|1ηt|Ωry(y)|supμ>0μη|{xΨ(ιλ,Ωry(y)):(Φ(|f|))p(x)pημ}|ηtt|Ψ(ιλ,Ωry(y))|1ηt(tη)δη|Ωry(y)|((ιλ)η|Ψ(ιλ,Ωry(y))|ηt+supμ>ιλμη|Ψ(μ,Ωry(y))|ηt)=t(tη)δη((ιλ)η+|Ψ(ιλ,Ωry(y))|1ηt|Ωry(y)|supμ>ιλμη|Ψ(μ,Ωry(y))|ηt).

    Now we choose ι > 0 sufficiently small so as to satisfy

    (λ2)η(ιλδ)ηttη(ιλδ)η=(λ2)η(ιλδ)η(1+ttη)(λ4)η,

    and there exists a positive constant c(t) depending only on t such that ιc(t)δ. Therefore, it follows that

    |Ωry(y)|cttη|Ψ(ιλ,Ωry(y))|1ηt(ιλ)η(supμ>ιλμt|Ψ(μ,Ωry(y))|)ηtct(ιλ)ttη((ιλ)t|Ψ(ιλ,Ωry(y))|)1ηt(supμ>ιλμt|Ψ(μ,Ωry(y))|)ηtct(ιλ)ttηsupμ>ιλμt|Ψ(μ,Ωry(y))|. (3.35)

    Next we put the estimates of two cases in (3.11) together into (3.23), that is, we insert the formulas (3.14) and (3.35) into (3.23) to get

    |E(Aλ,Ωr1(x0))|cϵ|E(λ4,Ωr2(x0))|+cϵ(ιλ)tsupμ>ιλμt|Ψ(μ,Ωr2(x0))|, (3.36)

    then we multiply (3.36) by ()tp, and take the supremum with respect to λ over (0, ∞) to show

    supλ>Mλ0(Aλ)tp|{xΩr2(x0):(Φ(|Du|))p(x)p>Aλ}|cϵAtp(supλ>Mλ0λtp|{xΩr2(x0):(Φ(|Du|))p(x)p>λ4}|+supλ>Mιλ0λtpt(supμ>λμt|Ψ(μ,Ωr2(x0))|))cϵ(supλ>Mλ0λtp|{xΩr2(x0):(Φ(|Du|))p(x)p>λ4}|+supλ>Mιλ0(supμ>λμtp|Ψ(μ,Ωr2(x0))|)).

    Note that

    supλ>Mιλ0supμ>λμtp|Ψ(μ,Ωr2(x0))|(Φ(|f|))p(x)Mt(Ωr2(x0))t.

    If we take ϵ > 0 so small that it ensures cϵ1t12, it follows that

    (Φ(|Du|))p(x)Mt(Ωr1(x0))cϵ1t((Φ(|Du|))p(x)Mt(Ωr2(x0))+(Φ(|f|))p(x)Mt(Ωr2(x0)))+cMλ0p|Ωr2(x0)|1t12(Φ(|Du|))p(x)Mt(Ωr2(x0))+c(Φ(|f|))p(x)Mt(Ωr2(x0))+c|Ωr2(x0)|1t(Ωr2(x0)(Φ(|Du|))p(x)pdx+(Ωr2(x0)((Φ(|f|))p(x)p+1)ηdx)1η)p.

    In the remainder we use a similar way of the argument in Step 5, and it leads to the desired result for the case q = ∞.□

Acknowledgement

We would like to thank the anonymous referee for his valuable comments and suggestions that led to improvement of this paper. This research was supported by NSFC grant 11371050.

References

[1] E. Acerbi and G. Mingione, Gradient estimates for the p(x)-Laplacean system, J. Reine Angew. Math. 584 (2005), 117–148.10.1515/crll.2005.2005.584.117Search in Google Scholar

[2] E. Acerbi and G. Mingione, Gradient estimates for a class of parabolic systems, Duke Math. J. 136 (2007), no. 2, 285–320.10.1215/S0012-7094-07-13623-8Search in Google Scholar

[3] K. Adimurthil and N. C. Phuc, Global Lorentz and Lorentz-Morrey estimates below the natural exponent for quasilinear equations, Calc. Var. Partial Differential Equations 54 (2015), 3107–3139.10.1007/s00526-015-0895-1Search in Google Scholar

[4] P. Baroni, Lorentz estimates for degenerate and singular evolutionary systems, J. Differential Equations 255 (2013), 2927–2951.10.1016/j.jde.2013.07.024Search in Google Scholar

[5] P. Baroni, Lorentz estimates for obstacle parabolic problems, Nonlinear Anal. 96 (2014), 167–188.10.1016/j.na.2013.11.004Search in Google Scholar

[6] P. Baroni, Riesz potential estimates for a general class of quasilinear equations, Calc. Var. Partial Differential Equations 53 (2015), no. 3-4, 803–846.10.1007/s00526-014-0768-zSearch in Google Scholar

[7] P. Baroni and C. Lindfors, The Cauchy-Dirichlet problem for a general class of parabolic equations, Ann. Inst. H. Poincaré Anal. Non Linéaire 34 (2017), no. 3, 593–624.10.1016/j.anihpc.2016.03.003Search in Google Scholar

[8] V. Bobkov and M. Tanaka, On sign-changing solutions for (p, q)-Laplace equations with two parameters, Adv. Nonlinear Anal. 8 (2019), no. 1, 101–129.10.1515/anona-2016-0172Search in Google Scholar

[9] V. Bögelein, F. Duzaar and G. Mingione, Degenerate problems with irregular obstacles, J. Reine Angew. Math. 650 (2011), 107–160.10.1515/crelle.2011.006Search in Google Scholar

[10] S. S. Byun, J. Ok and L. H. Wang, W1,p(x)-Regularity for elliptic equations with measurable coefficients in nonsmooth domains, Commun. Math. Phys. 329 (2014), 937–958.10.1007/s00220-014-1962-8Search in Google Scholar

[11] S. S. Byun and L. H. Wang, Elliptic equations with BMO coefficients in Reifenberg domains, Commun. Pure Appl. Math. 57 (2004), no. 10, 1283–1310.10.1002/cpa.20037Search in Google Scholar

[12] S. S. Byun and Y. Cho, Nonlinear gradient estimates for generalized elliptic equations with nonstandard growth in non-smooth domains, Nonlinear Anal. 140 (2016), 145–165.10.1016/j.na.2016.03.016Search in Google Scholar

[13] L. A. Caffarelli and I. Peral, On W1,p estimates for elliptic equations in divergence form, Commun. Pure Appl. Math. 51 (1998), 1–21.10.1002/(SICI)1097-0312(199801)51:1<1::AID-CPA1>3.0.CO;2-GSearch in Google Scholar

[14] Y. Cho, Global gradient estimates for divergence-type elliptic problems involving general nonlinear operators, J. Differential Equations 264 (2018), 6152–6190.10.1016/j.jde.2018.01.026Search in Google Scholar

[15] E. Correa and A. de Pablo, Remarks on a nonlinear nonlocal operator in Orlicz spaces, Adv. Nonlinear Anal. 9 (2020), no. 1, 305–326.10.1515/anona-2020-0002Search in Google Scholar

[16] L. Diening and F. Ettwein, Fractional estimates for non-differentiable elliptic systems with general growth, Forum Math. 20 (2008), 523–556.10.1515/FORUM.2008.027Search in Google Scholar

[17] L. Esposito, G. Mingione and C. Trombetti, On the Lipschitz regularity for certain elliptic problems, Forum Math. 18 (2006), no. 2, 263–292.10.1515/FORUM.2006.016Search in Google Scholar

[18] C. De Filippis and G. Mingione, A borderline case of Calderón-Zygmund estimates for non-uniformly elliptic problems. St. Petersburg Math. J. (2020) online. http://arxiv.org/abs/1901.05603v110.1090/spmj/1608Search in Google Scholar

[19] D. Kim and N. V. Krylov, Elliptic differenrial equations with coefficients measurable with respact to one variable and VMO with respect to the others, SIAM J. Math. Anal. 32 (2007), no. 9, 489–506.10.1137/050646913Search in Google Scholar

[20] N. V. Krylov and M. V. Safonov, A property of the solutions of parabolic equations with measurable coefficients, Izv. akad. nauk Ssr Ser. mat 26 (1980), no. 4, 345–361.10.1070/IM1981v016n01ABEH001283Search in Google Scholar

[21] N. V. Krylov, Parabolic and elliptic equations with VMO coefficients, Commun. Partial Differential Equations 32 (2007), no. 1-3, 453–475.10.1080/03605300600781626Search in Google Scholar

[22] A. Lemenant and E. Milakis, On the extension property of Reifenberg-flat domains, Ann. Acad. Sci. Fenn. Math. 39 (2014), no. 1, 51–71.10.5186/aasfm.2014.3907Search in Google Scholar

[23] G. M. Lieberman, The natural generalization of the natural conditions of Ladyzhenskaya and Uraľtseva for elliptic equations, Commun. Partial Differential Equations 16 (1991), no. 2-3, 311–361.Search in Google Scholar

[24] T. Mengesha and N. C. Phuc, Global estimates for quasilinear elliptic equations on Reifenberg flat domains, Arch. Rational Mech. Anal. 203 (2012), 189–216.10.1007/s00205-011-0446-7Search in Google Scholar

[25] E. Milakis and T. Toro, Divergence form operators in Reifenberg flat domains, Math. Z. 264 (2010), no. 1, 15–41.10.1007/s00209-008-0450-2Search in Google Scholar

[26] G. Mingione, The Calderón-Zygmund theory for elliptic problems with measure data, Ann. Sc. Norm. Super. Pisa. Cl. Sci. 6 (2007), no. 5, 195–261.10.2422/2036-2145.2007.2.01Search in Google Scholar

[27] G. Mingione, Gradient estimates below the duality exponent, Math. Ann. 346 (2010), no. 3, 571–627.10.1007/s00208-009-0411-zSearch in Google Scholar

[28] J. Musielak, Orlicz Spaces and Modular Spaces, Springer Verlag, Berlin, 1983.10.1007/BFb0072210Search in Google Scholar

[29] V. D. Rădulescu and D. D. Repovš, Partial differential equations with variable exponents. Variational methods and qualitative analysis, Monographs and Research Notes in Mathematics. CRC Press, Boca Raton, FL, 2015.Search in Google Scholar

[30] M. V. Safonov, Harnack’s inequality for elliptic equations and Hölder property of their solutions, J. Soviet Mathematics 21 (1983), no. 5, 851–863.10.1007/BF01094448Search in Google Scholar

[31] H. Tian and S. Z. Zheng, Uniformly nondegenerate elliptic equations with partially BMO coefficients in nonsmooth domains, Nonlinear Anal. 156 (2017), 90–110.10.1016/j.na.2017.02.013Search in Google Scholar

[32] H. Tian and S. Z. Zheng, Lorentz estimates for the gradient of weak solutions to elliptic obstacle problems with partially BMO coefficients, Boundary Value Problems 2017 (2017), 128.10.1186/s13661-017-0859-9Search in Google Scholar

[33] T. Toro, Doubling and flatness: geometry of measures, Notices Amer. Math. Soc. 44 (2000), no. 9, 1087–1094.Search in Google Scholar

[34] A. Verde, Calderón-Zygmund estimates for systems of φ-growth, J. Convex Anal. 18 (2011), no. 1, 67–84.Search in Google Scholar

[35] F. Yao and S. Zhou, Calderón-Zygmund estimates for a class of quasilinear elliptic equations, J. Funct. Anal. 272 (2017), 1524–1552.10.1016/j.jfa.2016.11.008Search in Google Scholar

[36] F. Yao, Calderón-Zygmund estimates for a class of quasilinear parabolic equations, Nonlinear Anal. 176 (2018), 84–99.10.1016/j.na.2018.06.008Search in Google Scholar

[37] J. J. Zhang and S. Z. Zheng, Lorentz estimates for fully nonlinear parabolic and elliptic equations, Nonlinear Anal. 148 (2017), 106–125.10.1016/j.na.2016.09.012Search in Google Scholar

[38] J. J. Zhang and S. Z. Zheng, Weighted Lorentz estimates for nondivergence linear elliptic equations with partially BMO coefficients, Commun. Pure Appl. Anal. 16 (2017), no. 3, 899–914.10.3934/cpaa.2017043Search in Google Scholar

[39] C. Zhang and S. L. Zhou, Global weighted estimates for quasilinear elliptic equations with non-standard growth, J. Funct. Anal. 267 (2014), 605–642.10.1016/j.jfa.2014.03.022Search in Google Scholar

Received: 2019-05-05
Accepted: 2020-04-30
Published Online: 2020-07-02

© 2021 Shuang Liang and Shenzhou Zheng, published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 18.4.2024 from https://www.degruyter.com/document/doi/10.1515/anona-2020-0121/html
Scroll to top button