Skip to content
BY 4.0 license Open Access Published by De Gruyter August 7, 2020

Regularity for sub-elliptic systems with VMO-coefficients in the Heisenberg group: the sub-quadratic structure case

  • Jialin Wang EMAIL logo , Maochun Zhu , Shujin Gao and Dongni Liao

Abstract

We consider nonlinear sub-elliptic systems with VMO-coefficients for the case 1 < p < 2 under controllable growth conditions, as well as natural growth conditions, respectively, in the Heisenberg group. On the basis of a generalization of the technique of 𝓐-harmonic approximation introduced by Duzaar-Grotowski-Kronz, and an appropriate Sobolev-Poincaré type inequality established in the Heisenberg group, we prove partial Hölder continuity results for vector-valued solutions of discontinuous sub-elliptic problems. The primary model covered by our analysis is the non-degenerate sub-elliptic p-Laplacian system with VMO-coefficients, involving sub-quadratic growth terms.

MSC 2010: 35H20; 35B65; 32A37

1 Introduction and statements of main results

In this paper, we consider discontinuous sub-elliptic systems with sub-quadratic growth coefficients that belong to the space of functions with vanishing mean oscillation (VMO, for short) in the Heisenberg group ℍn. We establish optimal partial Hölder continuity for vector-valued weak solutions in the sense that the solution is Hölder continuous on an open subset of its domain with full measure. More precisely, let Ω be a bounded domain, and horizontal gradient X = {X1, ⋯ X2n} with the horizontal vector fields Xi : (i = 1, ⋯, 2n) in ℍn, we consider sub-elliptic systems of the type

i=12nXiAiα(ξ,u,Xu)=Bα(ξ,u,Xu),inΩ,α=1,2,,N, (1.1)

where the primary coefficient Aiα VMO and satisfies some standard ellipticity and growth conditions with polynomial growth rate p ∈ (1, 2), and the inhomogeneous term Bα conforms to either controllable growth conditions, or natural growth conditions under an additional smallness assumption on the weak solutions. For the precise statement of the assumptions, and more details about the Heisenberg group, we refer to (H1)-(H4)-(HC) and (HN) below, and Section 2, respectively.

The main new aspect of this paper is the fact that we are able to deal with the inhomogeneity Bα : ℝ2n+1 × ℝN × ℝ2n×N → ℝN that satisfies the sub-quadratic controllable growth conditions, as well as sub-quadratic natural growth conditions, respectively, and the primary coefficient Aiα : ℝ2n+1 × ℝN × ℝ2n×N → ℝ2n×N that satisfies only a VMO-condition in ξ and is continuous in u. More precisely, we assume that the partial mapping ξ Aiα (ξ, u, P)/(1 + |P|)p−1 is VMO uniformly in (u, P), in the sense of (1.5) below and, moreover, u Aiα (ξ, u, P)/(1 + |P|)p−1 is continuous in the sense of (1.3) below. Our tool of choice is the use of an appropriate Sobolev-Poincaré inequality, and the harmonic approximation lemma; see Lemma 3.1, Lemma 3.3 below, respectively. The method of proof employed here will avoid the use of LqLp-estimates for the horizontal gradient and reverse Hölder inequalities. Our results essentially extend those results that the coefficients are continuous with respect to variables ξ and u to the case of the coefficients being VMO in the first variable ξ. We point out that partial Hölder continuity is the best one can expect under such weak assumptions concerning regularity of the structural functions Aiα and Bα in the (ξ, u)-variables.

We now impose the precise structure assumptions for coefficients Aiα and Bα we are dealing with.

(H1). The primary coefficient Aiα satisfies following ellipticity and growth conditions for a growth exponent 1 < p < 2:

DPAiα(ξ,u,P)P0,P0ν(1+|P|)p2|P0|2,|Aiα(ξ,u,P)|+(1+|P|)|DPAiα(ξ,u,P)|L(1+|P|)p1, (1.2)

for any choice of ξΩ, u, u0 ∈ ℝN and P, P0 ∈ ℝ2n×N. Here structure constants ν ≤ 1 ≤ L < ∞.

(H2). The vector field Aiα is continuous with respect to the second variable u. More precisely, there exists a bounded, concave and non-decreasing moduli of continuity ω : [0, ∞ → [0, 1] with lims0 ω(s) = 0 = ω(0) such that

|Aiα(ξ,u,P)Aiα(ξ,u0,P)|Lω|uu0|p(1+|P|)p1,1<p<2. (1.3)

(H3). The vector field Aiα is differentiable in the third variable P with continuous derivatives. This infers the bounded, concave and non-decreasing modulus μ : [0, ∞) → [0, 1] such that μ(t) ≤ t, lims0 μ(s) = 0 = μ(0), and we have

|DPAiα(ξ,u,P)DPAiα(ξ,u,P0)|Lμ|PP0|1+|P|+|P0|(1+|P|+|P0|)p2,1<p<2. (1.4)

With respect to the dependence on the first variable ξ, we do not impose a continuity condition, but we merely assume the following VMO-condition.

(H4). The mapping ξ Aiα (ξ, u, P)/(1 + |P|)p−1 satisfies the following VMO-condition uniformly in u and P:

Aiα(ξ,u,P)Aiα(,u,P)ξ0,rvξ0(ξ,r)(1+|P|)p1,for allξBr(ξ0),1<p<2. (1.5)

where vξ0 : ℝ2n+1 × [0, ρ0] → [0, 2L] are bounded functions satisfying

limρ0V(ρ)=0,whereV(ρ)=supξ0Ωsup0<rρ0Br(ξ0)Ωvξ0(ξ,r)dξ. (1.6)

Here we have used the short-hand notation

Aiα(,u,P)ξ0,r:=Br(ξ0)ΩAiα(ζ,u,P)dζ=Br(ξ0)Ω1Br(ξ0)ΩAiα(ζ,u,P)dζ.

(HC) (Controllable growth condition). The inhomogeneity Bα satisfies sub-quadratic controllable growth condition

|Bα(ξ,u,P)|C1+|u|p1+|P|p(11p),p=pQQp,1<p<Q,any constantpp,pQ, (1.7)

where C is a positive constant. We note that Q ≥ 3 is the homogeneous dimension in non-Abelian Heisenberg groups (see (2.1) below), and the exponent p ∈ (1, 2). So those infer that p < Q, and then, p = pQQp in our setting.

(HN) (Natural growth condition). For |u| ≤ M = supΩ |u|. The term Bα satisfies sub-quadratic natural growth condition

|Bα(ξ,u,P)|a|P|p+b,1<p<2, (1.8)

where a = a(M) and b = b(M) are constants possibly depending on M.

Now we mention some results on elliptic systems. Duzaar and Grotowski [9] prove optimal partial Hölder continuity for nonlinear elliptic systems with quadratic growth p = 2, by a new method so-called 𝓐-harmonic approximation introduced by Duzaar and Steffen [15]. Then, the method was extended to non-quadratic growth cases. Duzaar and Mingione [12, 13] consider systems of p-Laplacian type. Many partial regularity results have been established for more general nonlinear elliptic problems with Hölder, or Dini continuous coefficients; see, for example, [6, 8, 11, 28]. Furthermore, with respect to discontinuous elliptic problems, we refer to Bögelein, Duzaar, Habermann and Scheven [1], Ragusa [23], Zheng [35], Kanazawa [26], Goodrich, Ragusa and Scapellato [20], Polidoro and Ragusa [22], Scapellato [24], and Tan, Wang and Chen [27] and the references therein.

Several regularity results were focused on sub-elliptic systems in Heisenberg groups, or Hörmander vector fields; see Bramanti [2]. Xu and Zuily [34], Capogna and Garofalo [5], and Shores [25] showed partial regularity for quasi-linear sub-elliptic systems with quadratic growth p = 2. Their methods depend on generalization of classical freezing coefficient method. Then, by the generalization of the method of 𝓐-harmonic approximation, Föglein [16] treated homogeneous nonlinear sub-elliptic systems with Hölder continuous coefficients, under super-quadratic growth conditions p ≥ 2 in the Heisenberg group, and established partial Hölder continuity for the horizontal gradient Xu. Later Wang and Liao [30] considered the case of 1 < p < 2 for inhomogeneous systems in Carnot groups. Furthermore, Wang, Liao and Gao [31] weakened assumptions on coefficients Aiα with Hölder continuity in the variables (ξ, u) to the assumptions of Dini continuity, and proved partial regularity result with optimal estimates for the modulus of continuity for the horizontal derivative Xu.

Regularity results for discontinuous sub-elliptic systems with VMO coefficients instead of continuous coefficients have been established in the work [7] by Di Fazio and Fanciullo, and [19] by Gao, Niu and Wang for the case of quadratic growth; [32] by Wang and Manfredi, [14] by Dong and Niu, [36] by Zheng and Feng, and [33] by Wang, Zhang and Yang for non-quadratic growth conditions. We note that the regularity results in [14] and [36] have a limitation of p near 2, and the result in [19] holds only under a strong smallness condition for the dimension. In contrast, our partial Hölder continuity result stated below, is valid for the full range 1 < p < 2 in any dimension.

The typical strategy in partial regularity depends on decay estimates for certain excess functionals, which measure the oscillations of the solution or its gradient in a suitable sense. In this paper, we are working with a combination of a zero-order excess functional Cy and a first-order excess functional Ψ. For the case p ≥ 2, the functional Ψ is defined by

Ψ(ξ0,ρ,l)=Bρ(ξ0)ulρ(1+|Xl|)2+ulρ(1+|Xl|)pdξ,

with the horizontal affine functions l : ℝ2n → ℝN defined in the subsection 2.2 below. It is straightforward to adapt the standard 𝓐-harmonic approximation lemma by utilizing L2-theory combined with the standard Sobolev inequality; see Wang and Manfredi [32] for the super-quadratic natural growth case. However, in the present situation, we treat the case of sub-quadratic controllable growth, and sub-quadratic natural growth, respectively. So one should establish the decay estimate for the following excess functional

Ψ(ξ0,ρ,l)=Bρ(ξ0)Vulρ2dξ, (1.9)

where V(A)=(1+|A|2)4p2A for A ∈ ℝk, k ∈ ℕ+. On the other hand, we define the Campanato type excess functional Cy by

Cy(ξ0,ρ)=ρpyBρ(ξ0)|uuξ0,ρ|pdξ,1<p<2,0<y<1,

which provides a measure of the oscillations in the weak solutions u itself. It is remarkable that the excess functionals defined above involve only u, which simplifies the proofs of our partial regularity results. It is shown that if Ψ is small enough on a ball Bξ0(ρ) ⊂ ⊂ U, then, for some fixed θ ∈ (0, 1), one obtain an excess improvement Ψ (ξ0, θ r, lξ0,θr) ≤ C4θ2Ψ*(ξ0, r, lξ0,r) under smallness condition assumptions; see for example, Lemma 4.3. At this point, one has to assume smallness on the Ψ*-excess. Also we note that such an excess improvement estimate has two different quantities Ψ and Ψ* on the left, and the right hand side, respectively. Therefore, in contrast to the standard proof of partial regularity, the excess improvement cannot be iterated directly to yield an excess-decay estimate for Ψ-excess. In the present situation, however, iteration of the excess improvement yields that the Ψ-excess in (1.9) and also the Cy-excess remain bounded. Finally, the boundedness of the Cy-excess on any scale leads immediately to desired Hölder continuity of weak solutions u via the integral characterization of continuity by Campanato. We point out that the idea of such a combination of two excess functionals has its origin by Foss and Mingione [17] for continuous vector fields and integrands, and then, adapted to discontinuous problems with VMO coefficients for p ≥ 2 by Bögelein- Duzaar-Habermann-Scheven [1]. It is worth mentioning that we obviously do not have access to use L2-theory for functions in the horizontal Sobolev space HW1,p with 1 < p < 2. Therefore, we have to establish the following Sobolev-Poincaré inequality with the function V (see Lemma 3.1 below),

Bρ(ξ0)Vuuξ0,ρρ2QQpdξQp2QCPBρ(ξ0)VXu2dξ12,

with the constant CP dependence only on N, p, Q. This inequality is an essential tool in order to get the regularity result. It is also one technique point where our case differs from the case p ≥ 2 in [32].

Under the previous assumptions(H1)-(H4) and (HC), and (H1)-(H4) and (HN), respectively, we establish the following two partial Hölder continuity results.

Theorem 1.1

Assume that coefficients Aiα (ξ, u, Xu) and Bα(ξ, u, Xu) satisfy the assumptions (H1)-(H4) and (HC). Let uHW1,p(Ω, ℝN) with 1 < p < 2 be weak solutions to the systems (1.1), i.e.,

ΩAiα(ξ,u,Xu)Xφdξ=ΩBα(ξ,u,Xu)φdξ,φC0(Ω,RN). (1.10)

Then, there exists a relatively closed singular set Ω0Ω such that u Cloc0,y (ΩΩ0, ℝN) for every y ∈ (0, 1). Moreover, for any λ ∈ (0, Q) we have Xu Llocp,λ (ΩΩ0, ℝ2n×N) with the Morrey parameter λ = Qp(1 − y). Finally, we have that the singular set satisfies Ω0Σ1Σ2, where

Σ1=ξ0Ω:limr0sup(Xu)ξ0,r=,Σ2=ξ0Ω:limr0infBr(ξ0)V(Xu)V(Xu)ξ0,r2dξ>0

with the functional V defined in (2.3), and the singular set has (2n + 1)-Lebesgue measure zero |Ω0| = 0 and its complement ΩΩ0 is a set of full measure in Ω.

Theorem 1.2

Assume that coefficients Aiα (ξ, u, Xu) and Bα(ξ, u, Xu) satisfy the assumptions (H1)-(H4) and (HN). Let uHW1,p(Ω, ℝN) ∩ L (Ω, ℝN) be weak solutions to the system (1.1). Then, we have the same results that u Cloc0,y (ΩΩ0, ℝN) and Xu Llocp,λ (ΩΩ0, ℝ2n×N) as Theorem 1.

Remark 1.3

It is worth noting that the choice

Aiα(ξ,u,P)=a(ξ)1+|P|2p22Piαfori{1,,2n},α{1,,N}

makes the sub-elliptic p-Laplacian system with VMO-coefficients, involving sub-quadratic growth terms

i=12nXiAiα(ξ)1+Xu2p22Xiuα=Bα(ξ,u,Xu)

just as a special case of (1.1), where Aiα (ξ) ∈ VMO, and 1 < p < 2. So, combining the result for 2 ≤ p < ∞ established by Wang and Manfredi in [32], our partial Hölder continuity results covers the model case of sub-elliptic p-Laplacian system with 1 < p < ∞. It is remarkable that Zheng and Feng [36] showed everywhere regularity for weak solutions of sub-elliptic p-harmonic systems while p is very closed to 2.

The organization of this paper is as follows. In Section 2, we collect some basic notions and facts associated to Heisenberg groups, involving quasi-distance, horizontal Sobolev spaces, and horizontal affine function and some estimates. In Section 3, firstly an appropriate Sobolev-poincaré inequality which plays an important part on proving Hölder regularity is established. Then, an 𝓐-harmonic approximation lemma, and a prior estimate for weak solution hHW1,1 to the constant coefficient homogeneous sub-elliptic systems are given. In Section 4, we prove partial regularity results of Theorem 1.1 under sub-quadratic controllable structure assumptions (H1)-(H4) and (HC) by several steps. Step 1 is to gain a suitable Caccioppoli-type inequality which is an essential tool to get partial regularity. An appropriate linearization strategy is given in the second step. Then, one can achieve that solutions are approximately 𝓐-harmonic by the linearization procedure, and an excess improvement estimate for the functional Ψ is obtained under two smallness condition assumptions, by combining with 𝓐-harmonic approximation lemma in the third steps. Once the excess improvement is established, the iteration for the Ψ-excess and the Cy-excess can be acquired in Step 4. Finally, we show boundedness of the Campanato-type excess which leads immediately to desired Hölder continuity and Morrey regularity of Theorem 1.1. The last section shows the results of Theorem 1.2 under sub-quadratic natural structure assumptions (H1)-(H4) and (HN). In such a case, we establish appropriate estimates just for the natural growth term, and the rest procedure is similar to the proof of Theorem 1.1.

2 Preliminaries

In this section, we will give introduction of the Heisenberg group ℍn and definitions of several function spaces, and some elementary estimates which will be used later.

2.1 Introduction of the Heisenberg group ℍn

The Heisenberg group ℍn is defined as ℝ2n+1 endowed with the following group multiplication:

(ξ¯,t)(η¯,t~)=ξ¯+η¯,t+t~+12i=1nxiy~ix~iyi,

for all ξ = (ξ̄, t) = (x1, x2, ⋯, xn, y1, y2, ⋯, yn, t), ξ̃ = (η̄, ) = (1, 2, ⋯, n, 1, 2, ⋯, n, ). Its neutral element is 0, and its inverse to (ξ̄, t) is given by (− ξ̄, − t).

The basic vector fields corresponding to its Lie algebra can be explicitly calculated, and are given by

XiXi(ξ)=xiyi2t,Xn+iXn+i(ξ)=yi+xi2t,TT(ξ)=t

for i = 1, 2, ⋯, n, and note that the special structure of the commutators:

Xi,Xi+n=Xi+n,Xi=T,elseXi,Xj=0,andT,T=T,Xi=0,

that is, (ℍn, ⋅) is a nilpotent Lie group of step 2. X = (X1, ⋯, X2n) is said to be the horizontal gradient, and T vertical derivative.

The homogeneous norm is defined by ∥(ξ̄, t)∥n = (∥ξ̄4 + 16 t2)1/4, and the metric induced by this homogeneous norm is given by

d(ξ~,ξ)=ξ1ξ~Hn.

The measure used on ℍn is the Haar measure (Lebesgue measure in ℝ2n+1), and the volume of the homogeneous ball BR (ξ0) = {ξ ∈ ℍn : d(ξ0, ξ) < R} is given by |BR (ξ0)|n = R2n+2 |B1 (ξ0)|n =Δ ωn RQ, where the number

Q=2n+2 (2.1)

is called the homogeneous dimension of ℍn, and the quantity ωn is the volume of the homogeneous ball of radius 1.

Let 1 ≤ p ≤ +∞. We denote by

HW1,p(Ω)=uLp(Ω)|XiuLp(Ω),i=1,,k

the horizontal Sobolev space. Then, HW1,p(Ω) is a Banach space under the norm

uHW1,p(Ω)=uLp(Ω)+i=1kXiuLp(Ω).

For uHW1,q(BR (ξ0)), 1 < q < Q and 1 ≤ p qQQq , Lu [29] showed the following Poincaré type inequality associated with Hörmander vector fields, which is naturally valid for ℍn:

BR(ξ0)uuξ0,Rpdξ1pCpRBR(ξ0)Xuqdξ1q. (2.2)

The inequality (2.2) is valid for p = q (≥ 2).

Throughout the paper, we shall use the functions V, W: ℝk → ℝk defined by

V(ς)=(1+ς2)p24ς,W(ς)=ς/(1+ς2p)12 (2.3)

for each ς ∈ ℝk, k ∈ ℕ and p > 1. The functions V and W are locally bi-Lipschitz bijection on ℝk.

The following inequality

1+ς22p21+ς2p2p21+ς22p2,

immediately yields

W(ς)V(ς)2p4W(ς). (2.4)

The purpose of introducing W is the fact that in contrast to V2m, , the function W2m is convex. In fact, firstly by direct computation yields that W2p(t)=t2p(1+t2p)1p is a convex and monotone increasing function on [0, ∞) with W2p (0) = 0; secondly we have

Wς1+ς222pWς1+ς222pWς12p+Wς22p2,ς1,ς2Rn.

The following lemma includes some useful properties of the function V. The proof can be found in Lemma 2.1 of [4]. For simplicity, here, we replace the coefficient 2(p−2)/4 with 12 in the left of the first inequality (1) below, since the fact that 2−1/2 < 2(p−2)/4 for p > 0.

Lemma 2.1

Let p ∈ (1, 2) and V : ℝk → ℝk be the functions defined in (2.3). Then, for any ς1, ς2 ∈ ℝk and t > 0, the following inequalities hold:

  1. 12minς1,ς1p22(p2)/4minς1,ς1p2Vς1minς1,ς1p2;

  2. V(tς1)maxt,tp2V(ς1);

  3. |V(ς1 + ς2)| ⩽ C(p)(|V(ς1)| + |V(ς2)|);

  4. p2ς1ς2V(ς1)V(ς2)/(1+ς12+ς22)p24C(p,k)ς1ς2;

  5. |V(ς1) − V(ς2)| ⩽ C(p, k)|V(ς1ς2)|;

  6. |V(ς1ς2)| ⩽ C(p, M)|V(ς1) − V(ς2)| for all ς2 with |ς2| ≤ M.

2.2 Horizontal affine function and estimates in ℍn

Let uL2(Bρ(ξ0), ℝN), ξ0 ∈ ℝ2n+1, and consider the horizontal components

ξ¯=(x1,...,xn,y1,...,yn)andξ0¯=(x01,...,x0n,y01,...,y0n).

If the function

lξ0,ρ(ξ¯)=lξ0,ρ(ξ0¯)+Xlξ0,ρ(ξ¯ξ0¯),

minimizes the functional

lBρ(ξ0)|ul|2dξ,

among horizontal affine function l : ℝ2n → ℝN, Then, we have

lξ0,ρ(ξ0¯)=uξ0,ρ=Bρ(ξ0)udξ,

and

Xlξ0,ρ=Q2c0QQ+2ρ2Bρ(ξ0)u(ξ¯ξ0¯)dξ, (2.5)

where the vector u ⊗ (ξ̄ ξ0¯ ) has components uα(x1x01,x2x02,...,x2nx02n) with α = 1, 2, …, N, and c0 is a positive constant defined by

c0=0π(sinθ)ndθ0π(sinθ)n1dθ=[(2k2)!!]2(2k1)!!(2k3)!!2π,n=2k1,[(2k1)!!]2(2k)!!(2k2)!!π2,n=2k. (2.6)

Here, we use the notation (2k − 2)!! = (2k − 2)(2k − 4)⋅⋅⋅4 × 2 and (2k − 1)!! = (2k − 1)(2k − 3)⋅⋅⋅3 × 1. The proof of the results above can be found in [32] by Wang and Manfredi. On the basis of this formula, elementary calculations yield the following estimates.

Lemma 2.2

Let uL2(Bρ(ξ0), ℝN), θ ∈ (0, 1).We denote by lξ0,ρ and lξ0,θρ, the horizontal affine function defined as above for the radii ρ and θρ. Then, we have

|Xlξ0,ρXlξ0,θρ|pQ2c0QQ+2θρpBθρ(ξ0)ulξ0,ρpdξ, (2.7)

and, more generally,

Xlξ0,ρXlpQ2c0QQ+2ρpBρ(ξ0)ulpdξ, (2.8)

for every horizontal affine function l : ℝ2n → ℝN.

Proof

By the identity (2.5) and Hölder's inequality, we obtain

Xlξ0,ρXlξ0,θρp=Q2c0QQ+2θρppBθρ(ξ0)ulξ0,ρ(ξ¯0)Xlξ0,ρ(ξ¯ξ¯0)(ξ¯ξ¯0)dξpQ2c0QQ+2θρ2pBθρ(ξ0)ulξ0,ρ(ξ¯0)Xlξ0,ρ(ξ¯ξ¯0)pdξBθρ(ξ0)ξ¯ξ¯0pp1dξp1Q2c0QQ+2θρpBθρ(ξ0)ulξ0,ρpdξ, (2.9)

where we have used the fact that Bθρ(ξ0)ξ¯ξ¯0pp1dξp1(θρ)p.

For (2.8), it follows

Xlξ0,ρXlpQ2c0QQ+2ρ2pBρ(ξ0)ul(ξ¯0)Xl(ξ¯ξ¯0)pdξBρ(ξ0)ξ¯ξ¯0pp1dξp1=Q2c0QQ+2ρpBθρ(ξ0)ulpdξ.

According to the definition of the function lξ0,ρ, the following version of the Poincaré inequality (2.2) is true, that is,

Bρ(ξ0)ulξ0,ρ(ξ¯)pdξ1pCpρBρ(ξ0)XuXlξ0,ρqdξ1q,

where 1 < q < Q, 1 ≤ p qQQq .

3 Sobolev-Poincaré type inequality and 𝓐-harmonic approximation

We know that L2-theory cannot be directly used to obtain appropriate estimates for solutions uHW1,p with 1 < p < 2, so in this section, we first establish a suitable version of Sobolev-Poincaré type inequality with functions V. This inequality is an essential tool in proving partial regularity. Then, we give a prior estimate for 𝓐-harmonic functions hHW1,1, and introduce an 𝓐-harmonic approximation lemma which plays an important part in getting excess improvement estimates.

Lemma 3.1

(Sobolev-Poincaré type inequality). Let p ∈ (1, 2) and uHW1,p(Bρ(ξ0), ℝN) with Bρ (ξ0) ⊂ Ω, Then, it follows

Bρ(ξ0)Vuuξ0,ρρ2ppdξp2pCPBρ(ξ0)VXu2dξ12, (3.1)

with p = pQQp the Sobolev critical exponent of p; here the constant CP depends only on Q, N and p. In particular, the inequality also holds if we substitute 2 for 2pp .

Proof

We introduce the operator of fractional integration on Ω of order 1 as follows

I1(f)(ξ)=Ω|f(η)|d(ξ,η)Bξ,d(ξ,η)dη,ξBρ(ξ0).

Based on Theorem 2.7 in [3] by Capogna, Danielli and Garofalo, we deduce for 1 < p < +∞

Bρ(ξ0)|I1(f)(ξ)|pdξ1p=CρBρ(ξ0)|f(ξ)|pdξ1p, (3.2)

where we denote by p = pQQp the Sobolev critical exponent of p, and the number Q the homogeneous dimension in ℍn.

Lu [21] gave a representation formula for a function on graded nilpotent Lie groups for the left invariant vector fields; see Lemma 3.1 there. One form of the representation states that there exist constants c > 1 and C > 1 such that

u(ξ)uξ0,ρCBcρ(ξ0)|Xu(η)|d(ξ,η)Bξ,d(ξ,η)dη,ξBρ(ξ0).

Noting that W2/p(t) is monotone increasing and convex, we apply W2/p(t) to both sides of the last inequality and have by Jensen’s inequality

W2/pu(ξ)uξ0,ρρCρR2n+1W~2/pXu(η)d(ξ,η)Bξ,d(ξ,η)dη,

and

W~Xu(η)=0,ηBcρ(ξ0),WXu(η),ηBcρ(ξ0).

One can check that W(|Xu(η)|) ∈ Lp(Bρ (ξ0)), which implies (|Xu(η)) ∈ Lp(ℝ2n+1). Then, applying the inequality (3.2) yields

Bρ(ξ0)Wu(ξ)uξ0,ρρ2QQpdξQppQ=Bρ(ξ0)W2/pu(ξ)uξ0,ρρpQQpdξQppQCρBρ(ξ0)R2n+1W~2/pXu(η)d(ξ,η)Bξ,d(ξ,η)dηpQQpdξQppQCρBρ(ξ0)I1W2/p(Xu)(ξ)pQQpdξQppQCBρ(ξ0)W2Xu(ξ)dξ1p,

or

Bρ(ξ0)Wu(ξ)uξ0,ρ2QQpdξQp2QCBρ(ξ0)W2Xu(ξ)dξ12.

We obtain the assertion of the theorem, first for W, and Then, also for V by (2.4). □

Let 𝓐 ∈ Bil(Ω × ℝN × ℝ2n×N, ℝ2n×N) be a bilinear form with constant tensorial coefficients. We say that a map hC(Bρ(ξ0), ℝN) is 𝓐-harmonic if and only if

Bρ(ξ0)A(Xh,Xφ)dξ=0

holds for all testing function φ C0 (Bρ(ξ0), ℝN).

Shores in [25] showed that weak solutions hHW1,2(Ω, ℝN) of the constant coefficient homogeneous sub-elliptic systems belongs to C in the subset Ω0Ω. Then, the following estimate holds for the solution hHW1,2(Ω, ℝN),

supBρ/2(ξ0)Xh2+X2h2C0ρ2Bρ(ξ0)Xh2dξ.

Therefore, we can argue as the proof of Proposition 2.10 in [4] to obtain the same estimate for any 𝓐-harmonic function hHW1,1 (Ω, ℝN).

Lemma 3.2

Let hHW1,1 (Ω, ℝN) be weak solutions of the constant coefficient systems. Then, h is smooth and there exists C0 ≥ 1 such that for any Bρ (ξ0) ⊂ Ω

supBρ/2(ξ0)Xh2+X2h2C0ρ2Bρ(ξ0)Xh2dξ. (3.3)

Similarly to [10], one can establish the following version of 𝓐-harmonic approximation for the case 1 < p < 2 in Heisenberg groups.

Lemma 3.3

Let 0 < νL and 1 < p < 2 be given. For every ε > 0, there is a constant δ = δ(Q, N, p, ν, L, ε) ∈ (0, 1] assume that y ∈ [0, 1] and that 𝓐 is a bilinear form on2n×N with the properties

A(P,P)ν|P|2andA(P,P¯)L|P||P¯|,P,P¯R2n×N.

Furthermore, let wHW1,p(Bρ (ξ0), ℝN) be an approximate 𝓐-harmonic map in the sense that the following estimate holds

Bρ(ξ0)A(Xw,Xφ)dξδysupBρ(ξ0)|Xφ|,φC0(Bρ(ξ0),RN),

and

Bρ(ξ0)|V(Xw)|2dξy2.

Then, there exists an 𝓐-harmonic map hC(Bρ(ξ0), ℝN) which satisfies

Bρ(ξ0)Vwyhρ2dξy2εandBρ(ξ0)|V(Xh)|2dξ1.

4 Partial Hölder continuity for sub-quadratic controllable growth

In this section, we prove the partial regularity result of Theorem 1 under the assumptions of sub-quadratic controllable structure conditions. Now we begin with the following.

4.1 Caccioppoli-type inequality

We know that Caccioppoli-type inequality is a preliminary tool to prove partial regularity for systems. So in this subsection, we shall prove a Caccioppoli-type inequality for weak solutions to the sub-elliptic systems (1.1) with sub-quadratic controllable growth conditions.

Lemma 4.1

(Caccioppoli-type inequality). Let uHW1,p(Ω, ℝN) be weak solutions of the nonlinear sub-elliptic systems (1.1) under the assumptions (H1)-(H4)-(HC). Then, for any ξ0 = (x1, ⋯, xn, y1, ⋯, yn, t) ∈ Ω with Br(ξ0) ⊂ ⊂ Ω, and any horizontal affine functions l : ℝ2n → ℝN with |l(ξ̄0)| + |Xl| ≤ M0, we have the estimate

Br2(ξ0)|V(XuXl)|2dξCcBr(ξ0)Vulr2dξ+ωBr(ξ0)(|ul(ξ0¯)|p)dξ+V(r)+(r2+rp)Br(ξ0)(|Xu|p+|u|p+1)dξp(p),

where Cc is a positive constants depending only on Q, p, ν, L, M0, and the exponent p′ = pp1 , and (p)′ = pp1 with p = pQQp .

Proof

We choose a standard cut-off function ϕ C0 (Br(ξ0), [0, 1]) with ϕ ≡ 1 on Br2 (ξ0) and || ≤ 4r . Then, φ = ϕ2(ul) can be taken as a testing function for sub-elliptic systems (1.1). Hence, we have

Br(ξ0)Aiα(ξ,u,Xu)ϕ2(XuXl)dξ=2Br(ξ0)Aiα(ξ,u,Xu)ϕ(ul)Xϕdξ+Br(ξ0)Bα(ξ,u,Xu)ϕ2(ul)dξ,

where we have used the fact that = ϕ2(XuXl) + 2ϕ(ul).

In view of the identities Br(ξ0)Aiα(,l(ξ0¯),Xl)ξ0,rXφdξ=0, and

Br(ξ0)Aiα(ξ,u,Xl)ϕ2(XuXl)dξ=2Br(ξ0)Aiα(ξ,u,Xl)ϕ(ul)XϕdξBr(ξ0)Aiα(ξ,u,Xl)Xφdξ.

It follows for weak solutions u of systems (1.1) that

I0:=Br(ξ0)[Aiα(ξ,u,Xu)Aiα(ξ,u,Xl)]ϕ2(XuXl)dξ=2Br(ξ0)[Aiα(ξ,u,Xl)Aiα(ξ,u,Xu)]ϕ(ul)Xϕdξ+Br(ξ0)[Aiα(ξ,l(ξ0¯),Xl)Aiα(ξ,u,Xl)]Xφdξ+Br(ξ0)[(Aiα(,l(ξ0¯),Xl))ξ0,rAiα(ξ,l(ξ0¯),Xl)]Xφdξ+Br(ξ0)Bα(ξ,u,Xu)ϕ2(ul)dξ=:2I1+I2+I3+I4, (4.1)

with the obvious labelling for I0I4.

We first estimate the left-hand side of (4.1). By the first inequality of (1.2), Young’s inequality and definition of the function V (2.3), we have

I0=Br(ξ0)01DPAiα(ξ,u,Xl+s(XuXl))(XuXl),(XuXl)ϕ2dsdξBr(ξ0)01ν(1+|Xl+s(XuXl)|)p2|XuXl|2ϕ2dsdξBr(ξ0)ν3(1+|Xl|2+|XuXl|2)p22|XuXl|2ϕ2dξBr(ξ0)ν3(1+M02)p221+|XuXl|2p22|XuXl|2ϕ2dξ=ν3(1+M02)p22Br(ξ0)ϕ2|V(XuXl)|2dξ, (4.2)

where we have used the elementary inequality 1 + |a|2 + |ba|2 ⩽ 3(1 + |a|2 + |b|2), and 1 < p < 2.

Now, we are going to estimate the terms I1I4 on the right-hand side of (4.1). For small positive ϵ < 1 appearing in lines, it will be fixed later.

Estimate for I1. We shall decompose the ball Br(ξ0) into four subsets: Ω1 := Br(ξ0) ∩ {|XuXl| ≤ 1} ∩ ulr1 , Ω2 := Br(ξ0) ∩ {|XuXl| ≤ 1} ∩ ulr1 , Ω3 := Br(ξ0) ∩ {|XuXl| ≥ 1} ∩ ulr1 , and Ω4 := Br(ξ0) ∩ {|XuXl| ≥ 1} ∩ ulr1 .

  1. Using the second inequality of (1.2), |Xϕ| ≤ 4r , Young’s inequality, and Lemma 2.1, we derive the following bound for I1 on the subset Ω1.

    Ω1[Aiα(ξ,u,Xu)Aiα(ξ,u,Xl)]ϕ(ul)XϕdξLΩ101(1+|Xl+s(XuXl)|)p2ds|XuXl||ul||Xϕ|ϕdξ4LΩ1ϕ|2V(XuXl)|ulrdξ2εΩ1ϕ2|V(XuXl)|2dξ+32L2ε1Ω1ulr2dξ, (4.3)

    where we have used the inequality (1 + |Xl + s(Xu - Xl)|)p−2 ≤ 1 for 1 < p < 2.

  2. Similarly to the case 1, there is

    Ω2[Aiα(ξ,u,Xu)Aiα(ξ,u,Xl)]ϕ(ul)Xϕdξ2p2(p1)εΩ2ϕpp1|V(XuXl)|pp1dξ+(4L)pε1pΩ2ulrpdξ2εΩ2ϕ2|V(XuXl)|2dξ+2(4L)pε1pΩ2Vulr2dξ, (4.4)

    where we have used Lemma 2.1, and the inequality |V(XuXl)|pp1 ≤ |V(XuXl)|2 ≤ |XuXl|2 ≤ 1 for 1 < p < 2 on the set Ω2.

  3. By Young’s inequality and Lemma 2.1, it follows that on the subset Ω3,

    Ω3[Aiα(ξ,u,Xu)Aiα(ξ,u,Xl)]ϕ(ul)XϕdξεΩ3ϕp|XuXl|pdξ+(4L)pp1ε11pΩ3ulr2dξ2εΩ3ϕ2|V(XuXl)|2dξ+2(4L)pp1ε11pΩ3Vulr2dξ, (4.5)

    where we have used the fact that ulrpp1ulr2 as pp1 ≥ 2 and ulr ≤ 1 on the subset Ω3.

  4. On the subset Ω4, it holds that by the assumption |l( ξ0¯ )| + |Xl| ≤ M0

    Ω4[Aiα(ξ,u,Xu)Aiα(ξ,u,Xl)]ϕ(ul)Xϕdξ22+1pLΩ4ϕV(XuXl)2pulrdξΩ4εϕpp1V(XuXl)2p1dξ+22p+1LpΩ4ε1pulrpdξεΩ4ϕ2|V(XuXl)|2dξ+C(p,L)ε1pΩ4Vulr2dξ, (4.6)

    here, we have used the smallness assumption Φ(ξ0, r, l) := ⨍Br(ξ0)|V(XuXl)|2 ≤ 1 and ϕpp1 ϕ2.

    From (4.3), (4.4), (4.5) and (4.6), we have the estimate for the term I1 as follows

    I12εBr(ξ0)ϕ2|V(XuXl)|2dξ+C(p,L,M0)ε11pBr(ξ0)Vulr2dξ, (4.7)

    where we have used the inequality ε11pε1ε1p for small positive constant ε < 1.

    Estimate for I2. By the first inequality of (1.3), we get

    I2LBr(ξ0)ω|ul(ξ0¯)|p(1+|Xl|)p1|Xφ|dξC(p,L,M0)Br(ξ0)ω|ul(ξ0¯)|pϕ2|XuXl|dξ+C(p,L,M0)Br(ξ0)ω|ul(ξ0¯)|pϕ|ul||Xϕ|dξ=:I21+I22. (4.8)

    To estimate the term I21, we divide the domain of integration into two parts Ω5 := Br(ξ0) ∩ |XuXl| ≤ 1 and Ω6 := Br(ξ0) ∩ {|XuXl| > 1}.

    1. On the set Ω5 where |XuXl| ≤ 1, it holds

      I21(onΩ5)2εΩ5ϕ4|V(XuXl)|2dξ+C(p,L,M0)ε1Ω5ω2|ul(ξ0¯)|pdξ2εΩ5ϕ2|V(XuXl)|2dξ+C(p,L,M0)ε1ωΩ5|ul(ξ0¯)|pdξ, (4.9)

      where we have used in turn Young’s inequality, ω2ω, the concavity of ω and Jensen’s inequality.

    2. On the part Ω6 where |XuXl| > 1, we find

      I21(onΩ6)εΩ6ϕ2p|XuXl|pdξ+C(p,L,M0)ε11pΩ6ωpp1|ul(ξ0¯)|pdξ2εΩ6ϕ2|V(XuXl)|2dξ+C(p,L,M0)ε11pωΩ6|ul(ξ0¯)|pdξ, (4.10)

      where we have used the inequality ωpp1 ω.

    Combining (4.9) with (4.10) leads to

    I212εBr(ξ0)ϕ2|V(XuXl)|2dξ+C(p,L,M0)ε11pωBr(ξ0)|ul(ξ0¯)|pdξ, (4.11)

    where we have use the fact ε11p ε−1 for 0 < ε < 1.

The term I22 can be estimated similarly as I21 above. Here, we split the ball Br(ξ0) into two subsets Ω7 := Br(ξ0) ∩ {|ulr|1} and Ω8 := Br(ξ0) ∩ {ulr|>1} .

  1. On the subset Ω7, it yields

    I22(onΩ7)2εΩ7ϕ2Vulr2dξ+C(p,L,M0)ε1ωΩ7|ul(ξ0¯)|pdξ. (4.12)
  2. We deduce on Ω8

    I22(onΩ8)2εΩ8ϕpVulr2dξ+C(p,L,M0)ε11pωΩ8|ul(ξ0¯)|pdξ. (4.13)

From (4.12) and (4.13), it follows

I222εBr(ξ0)Vulr2dξ+C(p,L,M0)ε11pωBr(ξ0)|ul(ξ0¯)|pdξ. (4.14)

Joining (4.8), (4.11) and (4.14), we obtain

I22εBr(ξ0)ϕ2|V(XuXl)|2dξ+2εBr(ξ0)Vulr2dξ+C(p,L,M0)ε11pωBr(ξ0)|ul(ξ0¯)|pdξ. (4.15)

We are now in the position to handle the term I3. By VMO-condition (1.5), the term I3 can be estimated as follows

I3Br(ξ0)vξ0(1+|Xl|)p1|Xφ|dξC(p,M0)Br(ξ0)vξ0ϕ2|XuXl|dξ+C(p,M0)Br(ξ0)vξ0|ul|ϕ|Xϕ|dξ=:I31+I32. (4.16)

We can argue the terms I31 and I32 as the same way treating the terms I21 and I22.

  1. On the set Ω5 where |XuXl| ≤ 1, we use vξ0 ≤ 2L and (1.6) to infer the following estimate

    I31(onΩ5)Ω5εϕ42V(XuXl)2dξ+C(p,M0)Ω5ε1vξ02dξ2εΩ5ϕ2|V(XuXl)|2dξ+C(p,L,M0)ε1V(r). (4.17)
  2. On the part Ω6 where |XuXl| > 1, we use (1.6) and the fact that vξ0pp1=vξ0vξ01p1,vξ02L, to deduce

    I31(onΩ6)Ω6εϕ2pXuXlpdξ+C(p,M0)Ω6ε11pvξ0p1pdξ2εΩ6ϕ2|V(XuXl)|2dξ+C(p,L,M0)ε11pV(r). (4.18)

Using (4.17) and (4.18), we get

I312εBr(ξ0)ϕ2|V(XuXl)|2dξ+C(p,L,M0)ε11pV(r). (4.19)

Similarly, the term I32 can be estimated as follows

I322εBr(ξ0)Vulr2dξ+C(p,L,M0)ε11pV(r). (4.20)

Joining (4.16), (4.19) with (4.20), we have

I32εBr(ξ0)ϕ2|V(XuXl)|2dξ+2εBr(ξ0)Vulr2dξ+C(p,L,M0)ε11pV(r). (4.21)

Estimate for I4. Using Hölder inequality, one has

I4CBr(ξ0)Xup+up+111pϕ2(ul)dξCBr(ξ0)Xup+up+1dξ11pBr(ξ0)ϕ2(ul)pdξ1p. (4.22)

To obtain an appropriate estimate for I4, we take the domain Br(ξ0) into four parts as the same way of I1.

  1. For the case of Ω1 = Br(ξ0) ∩ {|XuXl| ≤ 1} ∩ ulr1 , by Sobolev type inequality, Hölder’s inequality, Young’s inequality and Lemma 2.1, it follows that

    Ω1Xup+up+1dξ11pΩ1ϕ2(ul)pdξ1pCpΩ1Xup+up+1dξ11prΩ1ulr+ϕ22V(XuXl)pdξ1pC(Cp,p,ε)r2Ω1Xup+up+1dξ211p+CpεΩ1ulr2+2CpεΩ1ϕ4V(XuXl)2dξC(Cp,p,ε)r2Ω1Xup+up+1dξ211p+2CpεΩ1Vulr2+2CpεΩ1ϕ2VXuXl2dξ. (4.23)
  2. On the part Ω2 = Br(ξ0) ∩ {|XuXl| ≤ 1} ∩ ulr1 , the following estimate holds

    Ω2Xup+up+1dξ11pΩ2ϕ2(ul)pdξ1pC(Cp,p,ε)rpp1Ω2Xup+up+1dξ11ppp1+CpεΩ2ulrp+C(Cp,p,ε)r2Ω2Xup+up+1dξ211p+CpεΩ2ϕ4V(XuXl)2dξC(Cp,p,ε)rpp1+r2Ω2Xup+up+1dξ11ppp1+2CpεΩ2Vulr2+2CpεΩ2ϕ2VXuXl2dξ, (4.24)

    where we have used the fact that ⨍Br(ξ0) (|Xu|p + |u|p + 1) ≥ 1, and ϕ ≤ 1.

  3. On the part Ω3 = Br(ξ0) ∩ {|XuXl| ≥ 1} ∩ ulr1 , it yields

    Ω3Xup+up+1dξ11pΩ3ϕ2(ul)pdξ1pCpΩ3Xup+up+1dξ11prΩ3ulr+ϕ2XuXlpdξ1pC(Cp,p)Ω3Xup+up+1dξ11prΩ3ulr2dξ12+rΩ3ϕ2pXuXlp1pdξC(Cp,p,ε)rpp1+r2Ω3Xup+up+1dξ11ppp1+2CpεΩ3Vulr2+2CpεΩ3ϕ2VXuXl2dξ, (4.25)
  4. For the last case of Ω4 = Br(ξ0) ∩ {|XuXl| ≥ 1} ∩ ulr1 , we get

    Ω4Xup+up+1dξ11pΩ4ϕ2(ul)pdξ1p2p1pCpΩ4Xup+up+1dξ11prΩ4ulrpdξ1p+rΩ4ϕ2pXuXlp1pdξC(Cp,p,ε)rpp1Ω4Xup+up+1dξ11ppp1+2CpεΩ4ulrp+2CpεΩ4ϕ2pXuXlpdξC(Cp,p,ε)rpp1Ω4Xup+up+1dξ11ppp1+4CpεΩ4Vulr2+4CpεΩ4ϕ2VXuXl2dξ. (4.26)

Combining the estimates (4.22)-(4.26), we find that

I44CpεBr(ξ0)ϕ2VXuXl2dξ+C(Cp,p,ε)r2+rpp1Br(ξ0)Xup+up+1dξ11ppp1+4CpεBr(ξ0)Vulr2. (4.27)

Joining the estimates(4.2), (4.7), (4.15), (4.21), (4.27) with (4.1), we arrive at

3(1+M02)p22Br(ξ0)ϕ2|V(XuXl)|2dξ6ε+4CpεBr(ξ0)ϕ2|V(XuXl)|2dξ+C(p,L,M0)ε11p+4ε(1+Cp)Br(ξ0)Vulr2dξ+C(p,L,M0)ε11pωBr(ξ0)|ul(ξ0¯)|pdξ+C(p,L,M0)ε11pV(r)+C(Cp,p,ε)r2+rpp1Br(ξ0)Xup+up+1dξ11ppp1.

Here, choosing ε<3(1+M02)p226+4Cp, we can absorb the first integral of the right-hand side into the left. Keeping in mind the properties of ϕ, we have thus shown

Br2(ξ0)|V(XuXl)|2dξ2QBr(ξ0)|V(XuXl)|2ϕ2dξCcBr(ξ0)Vulr2dξ+ωBr(ξ0)(|ul(ξ0¯)|p)dξ+V(r)+(r2+rp)Br(ξ0)(|Xu|p+|u|p+1)dξp(p),

with a constant Cc = Cc(Q, p, ν, L, M0), p′ = pp1 , and (p)′ = pp1 . This proves the claim. □

For sake of simplicity, we motivated the form of the Caccioppoli inequalities in Lemma 4.1. We set

Φ(ξ0,r,l):=Br(ξ0)|V(XuXl)|2dξ,Ψ(ξ0,r,l):=Br(ξ0)Vulr2dξ,
Ψ(ξ0,r,l):=Ψ(ξ0,r,l)+ωBr(ξ0)|ul(ξ0¯)|pdξ+V(r)+r2+rpBr(ξ0)(|Xu|p+|u|p+1)dξp(p).

In the sequel, when the choice of ξ0 or l is clear, we frequently write Φ(r, l) or Φ(r) respectively, as a replacement of Φ(ξ0, r, l).

4.2 Approximate 𝓐-harmonicity of weak solutions

To apply 𝓐-harmonic approximation lemma, we need to establish the following lemma, which provides a linearization strategy for non-linear sub-elliptic systems (1.1).

Lemma 4.2

Under the assumptions of Theorem 1.1 are satisfied, B2ρ(ξ0) ⊆ Ω with ρρ0 and an arbitrary horizontal function l : ℝ2n → ℝN, we define

A=DPAiα(,l(ξ0¯),Xl)ξ0,ρandw=ul,

Then, w is approximately 𝓐-harmonic in the sense that

Bρ(ξ0)A(Xw,Xφ)dξC1Ψ(2ρ)+μΨ12(2ρ)+μΨ1p(2ρ)supBρ(ξ0)|Xφ|

for all φ C0 (Bρ(ξ0), ℝN), and the positive constant C1 = C(p, M0, L, Cc).

Proof

Without loss of generality, we assume that supBρ(ξ0)|Xφ|1. Noting that w = ul, we compute

Bρ(ξ0)A(Xw,Xφ)dξ=Bρ(ξ0)01DPAiα(,l(ξ0¯),Xl)ξ0,ρDPAiα(,l(ξ0¯),Xl+sXw)ξ0,ρXwdsdξsupBρ(ξ0)|Xφ|+Bρ(ξ0)01DPAiα(,l(ξ0¯),Xl+sXw)ξ0,ρXwdsdξsupBρ(ξ0)|Xφ|=:(J1+J2)supBρ(ξ0)|Xφ|, (4.28)

with obvious labelling of J1 and J2.

In order to get the bound for the first term J1, we first use the inequality (1.4) to obtain

01DPAiα(,l(ξ0¯),Xl)ξ0,ρDPAiα(,l(ξ0¯),Xl+sXw)ξ0,ρds=01Bρ(ξ0)DPAiα(,l(ξ0¯),Xl)DPAiα(,l(ξ0¯),Xl+sXw)dξds01Bρ(ξ0)DPAiα(,l(ξ0¯),Xl)DPAiα(,l(ξ0¯),Xl+sXw)dξdsLBρ(ξ0)μ|XuXl|1+|Xl|(1+2|Xl|)p2dξ.

By the monotonicity of μ and the inequality above, it yields

J1LBρ(ξ0)μ|XuXl|1+|Xl|(1+2|Xl|)p2|XuXl|dξC(p,L,M0)Bρ(ξ0)μ(|XuXl|)|XuXl|dξ.

Here, we decompose the ball Bρ(ξ0) into two parts Ω5 and Ω6.

  1. On the domain Ω5 where |XuXl| ≤ 1, it follows by Lemma 2.1 Young’s inequality, Jensen’s inequality, and Hölder’s inequality in turn

    J1(onΩ5)C(p,L,M0)Ω5μ|V(XuXl)||V(XuXl)|dξC(p,L,M0)μ2Ω5|V(XuXl)|dξ+Ω5|V(XuXl)|2dξC(p,L,M0)μΩ5|V(XuXl)|2dξ12+Ω5|V(XuXl)|2dξ,

    where we have used the inequality μ2μ.

  2. On the set Ω6 where |XuXl| > 1, we have the following bound

    J1(onΩ6)C(p,L,M0)Ω6μ(|XuXl|)|XuXl|dξC(p,L,M0)μpp1Ω6|XuXl|dξ+Ω6|XuXl|pdξC(p,L,M0)μΩ6|XuXl|pdξ1p+Ω6|V(XuXl)|2dξC(p,L,M0)μΩ6|V(XuXl)|2dξ1p+Ω6|V(XuXl)|2dξ,

    where we have used μpp1 μ and Lemma 2.1. Then, we get the following estimate

    J1C(p,L,M0)μ(Φ12(ρ))+μ(Φ1p(ρ))+Φ(ρ). (4.29)

    Based on the following facts

    Bρ(ξ0)Aiα(ξ,u,Xu),XφdξBρ(ξ0)Bα(ξ,u,Xu),φdξ=0andBρ(ξ0)Aiα(,l(ξ0¯),Xl)ξ0,ρ,Xφdξ=0,

    the integral J2 can be rewritten as

    J2=Bρ(ξ0)Aiα(,l(ξ0¯),Xu)ξ0,ρAiα(,l(ξ0¯),Xl)ξ0,ρXφdξ=Bρ(ξ0)Aiα(,l(ξ0¯),Xu)ξ0,ρAiα(ξ,l(ξ0¯),Xu)Xφdξ+Bρ(ξ0)Aiα(ξ,l(ξ0¯),Xu)Aiα(ξ,u,Xu)Xφdξ+Bρ(ξ0)Bα(ξ,u,Xu)φdξ=:J21+J22+J23, (4.30)

    with the obvious meaning of J21 + J22 + J23.

    Using the assumption of |l( ξ0¯ )| + |Xl| ≤ M0 and VMO-condition (1.5), We find that

    J21Bρ(ξ0)vξ0(1+|Xu|)p1dξBρ(ξ0)vξ0(1+|Xl|p1+|XuXl|p1)dξ(1+M0p1)Bρ(ξ0)vξ0+vξ0|XuXl|p1dξ,

    where we have used the inequality 0 < p − 1 < 1 in the second line.

Now, we discuss it on the domain Ω5 and Ω6, respectively.

  1. On the set Ω5 where |XuXl| ≤ 1, the following estimate holds

    J21(onΩ5)(1+M0p1)Ω5vξ0+vξ0|2V(XuXl)|p1dξ(1+M0p1)Ω5vξ0dξ+Ω5vξ022pdξ+2Ω5|V(XuXl)|2dξC(p,L,M0)Ω5vξ0dξ+Ω5|V(XuXl)|2dξ,

    where we have used vξ02p2=vξ0vξ0p2p,vξ02L and Lemma 2.1.

  2. On the part Ω6 where |XuXl| > 1, it yields

    J21(onΩ6)(1+M0)Ω6vξ0dξ+Ω6vξ0pdξ+Ω6|XuXl|pdξC(p,L,M0)Ω6vξ0dξ+Ω6|V(XuXl)|2dξ,

    where we have used the assumption vξ0 ≤ 2L and Lemma 2.1.

    Then, we get the following estimate for J21

    J21C(p,L,M0)V(ρ)+Φ(ρ). (4.31)

    By first inequality of (1.3), the term J22 can be estimated as follows

    J22LBρ(ξ0)ω|ul(ξ0¯)|p(1+|Xu|)p1dξL(1+M0p1)Bρ(ξ0)ω|ul(ξ0¯)|p+ω|ul(ξ0¯)|p|XuXl|p1dξ.

    Similarly, for the case of |XuXl| ≤ 1 on Ω5, applying Young’s inequality, Jensen’s inequality and Lemma 2.1, we deduce that

    J22(onΩ5)C(L,p,M0)ωΩ5|ul(ξ0¯)|pdξ+Ω5|V(XuXl)|2dξ,

    where we have used ω22p ω ≤ 1.

    For the other case of |XuXl| > 1 on Ω6, it follows

    J22(onΩ6)C(L,p,M0)ωΩ6|ul(ξ0¯)|pdξ+Ω6|V(XuXl)|2dξ,

    where we have used ωpω ≤ 1, Jensen’s inequality and Lemma 2.1.

    Thus, we get the following estimate for J22

    J22C(L,p,M0)ωBρ(ξ0)|ul(ξ0¯)|pdξ+Φ(ρ). (4.32)

    Finally, we handle the term J23 by the same as the way for I4 to obtain

    J234CpεBρ(ξ0)ϕ2VXuXl2dξ+4CpεBρ(ξ0)Vulρ2.+C(Cp,ε)ρ2+ρpp1Bρ(ξ0)Xup+up+1dξ11ppp1C(Cp,ε)Φ(ρ)+Ψ(ρ)+ρ2+ρpp1Bρ(ξ0)Xup+up+1dξ11ppp1 (4.33)

    Joining the estimates (4.31)-(4.33) with (4.30), we have

    J2C(p,L,M0,Cp)Φ(ρ)+Ψ(ρ)+ωBρ(ξ0)|ul(ξ0¯)|pdξ+V(ρ)+ρ2+ρpp1Bρ(ξ0)Xup+up+1dξ11ppp1=C(p,L,M0,Cp)Φ(ρ)+Ψ(ρ). (4.34)

    Plugging (4.29) and (4.34) into (4.28), we finally arrive at

    Bρ(ξ0)A(Xw,Xφ)dξC(p,L,M0,Cp)μ(Φ12(ρ))+μ(Φ1p(ρ))+Φ(ρ)+Ψ(ρ)supBρ(ξ0)|Xφ|C(p,L,M0,Cc,Cp)μ(Ψ12(2ρ))+μ(Ψ1p(2ρ))+Ψ(2ρ)supBρ(ξ0)|Xφ|,

    where we have employed the Caccioppoli-type inequality from Lemma 4.1, Ψ(ρ) ≤ C(n, p)Ψ(2ρ) in the last step. This yields the claim. □

4.3 Excess improvement

The strategy of our proof is to approximate the given solution in the sense of L2 by 𝓐-harmonic functions. Now we are in the position to establish the excess improvement.

Lemma 4.3

Suppose that the assumptions of Theorem 1.1 are satisfied and consider a ball Br(ξ0) ⊆ Ω with rρ0. For the constants δ = δ (Q, N, p, L, ν, θQ+4) ∈ (0, 1] and y ∈ (0, 1] from the 𝓐-harmonic approximation Lemma 3.3, we let 0 < θ 14 be arbitrary and also impose the following smallness conditions:

  1. Ψ12(r)<δ2;

  2. y:=Ψ(r)+δ22μΨ12(r)+μΨ1p(r)21.

Then, there holds an excess improvement estimate

Ψ(ξ0,θr,lξ0,θr)C4θ2Ψ(ξ0,r,lξ0,r)

with some constants C4 that depend only on n, N, p, ν, δ and L. Here, lξ0,θr and lξ0,r denote the minimizing affine functions introduced in Lemma 2.2.

Proof

We denote Ψ*(r) = Ψ*(ξ0, r, lξ0,r), and take

w~=ulξ0,rC2

with lξ0,r = uξ0,r+Xlξ0,r(ξ̄ξ̄0) and C2 = max{C1, Cc }. We claim that satisfies the assumptions of 𝓐-harmonic approximation Lemma 3.3.

First note that, for our choice of the bilinear form

A=DPAiα(,lξ0,r(ξ0¯),Xlξ0,r)ξ0,ρ.

Next by Lemma 4.2 with ρ = r2 and l = lξ0,r, and the assumptions (i) and (ii), we find the map is approximately 𝓐-harmonic in the sense that

Br2(ξ0)A(Xw~,Xφ)dξyC1C2Ψ(r)+μΨ12(r)+μΨ1p(r)ysupBr2(ξ0)|Xφ|yΨ12(r)+δ2supBr2(ξ0)|Xφ|yδsupBr2(ξ0)|Xφ| (4.35)

for all φC0(Br2(ξ0),RN), and

Br2(ξ0)|V(Xw~)|2dξ=1C22Br2(ξ0)|V(XuXlξ0,r)|2dξΦ(r/2,lξ0,r)C22CcC22Ψ(r)y2. (4.36)

The estimates (4.35) and (5.2) tell us that the conditions of Lemma 3.3 are satisfied. So, there exists an 𝓐-harmonic hC0(Br2(ξ0),RN) such that

Br2(ξ0)Vw~yhr2dξy2ε,andBr2(ξ0)|V(Xh)|2dξ1.

In order to estimate excess functional

Ψ(ξ0,θr,lξ0,θr)=Bθr(ξ0)Vulξ0,θrθr2dξ,

we now have to handle the integral ⨍Bθr(ξ0) |V(X2h(ξ))|2. Since the function h(ξ) is 𝓐-harmonic, we know that h(ξ) ∈ C(Ω) by Lemma 3.2. Noting that the boundedness |Xh(ξ)| ≤ M in the ball Br2(ξ0) ⊂ ⊂ Ω, and using Hölder’s inequality, we have the estimate for θ ∈ (0, 14 )

Bθr(ξ0)V(X2h(ξ))2dξsupBr4(ξ0)X2h(ξ)2C0r2Br2(ξ0)Xh(ξ)2dξC0Mr2Br2(ξ0)Xh(ξ)2dξ12+Br2(ξ0)Xh(ξ)pdξ1p2C0Mr2Br2(ξ0)V(Xh(ξ))2dξ12+Br2(ξ0)V(Xh(ξ))2dξ1pCr2, (4.37)

where we have used the estimate (3.3) and Lemma 2.1.

We write lh(ξ) = hξ0,θr + (Xh)ξ0,θr(ξ̄ ξ0¯ ). Based on (3.1) and (4.37), it follows that

Bθr(ξ0)Vw~ylhθr2dξCBθr(ξ0)Vw~yhθr2dξ+Bθr(ξ0)yVhhξ0,θr(Xh)ξ0,θr(ξ¯ξ0¯)θr2dξCθ2(2θ)QBr2(ξ0)Vw~yhr2dξ+CCPy2Bθr(ξ0)VXh(Xh)ξ0,θr2dξCθ2(2θ)Qy2ε+CCP2(θr)2y2Bθr(ξ0)V(X2h)2dξC(CP,C0)y2θ2Qε+θ2C(CP,C0)(1+16δ2)θ2Ψ(r),

where we have taken ε = θQ+4.

Scaling back to u, we infer

Bθr(ξ0)Vulξ0,rC2ylhθr2dξC22C(CP,C0)(1+16δ2)θ2Ψ(r).

In view of the defining property of lξ0,θr, we arrive at

Bθr(ξ0)Vulξ0,θrθr2dξC22C(CP,C0)(1+16δ2)θ2Ψ(r)C4θ2Ψ(r),

here, we have denoted C4=C22C(CP,C0)(1+16δ2). Then, it implies excess improvement estimate

Ψ(ξ0,θr,lξ0,θr)C4θ2Ψ(ξ0,r,lξ0,r).

4.4 Iteration

First, let y ∈ (0, 1) be an fixed Hölder exponent. We define the Campanato-type excess

Cy(ξ0,ρ)=ρpyBρ(ξ0)|uuξ0,ρ|pdξ,1<p<2.

Next, we iterate the excess improvement estimate from Lemma 4.3.

Lemma 4.4

Suppose that the assumptions of Theorem 1.1 are satisfied. For every y ∈ (0, 1), there are constant ε*, κ*, ρ* and θ ∈ (0, 18 ], if

Ψ(ξ0,r,lξ0,r)<εandCy(ξ0,r)<κ, (A0)

for r ∈ (0, ρ*) with Br(ξ0) ⊂ Ω, Then,

Ψ(ξ0,θkr,lξ0,θkr)<εandCy(ξ0,θkr)<κ, (Ak)

respectively, for every k ∈ ℕ.

Proof

We begin by choosing the constants. First, we let

θ=minc0Q2p(Q2)(Q+2)11y,12C418, (4.38)

where c0 is defined in (2.6), and C4 is determined in Lemma 4.3, respectively. We note that the choice of θ fixes the constant δ = δ(Q, N, p, ν, L, θQ+4) from Lemma 3.3. Next, we fix an ε* small sufficiently to ensure

εminθQ+py82p,δ23δ2+48andμ(ε)ε. (4.39)

Then, we choose κ* > 0 so small that

ω(κ)ε.

Finally, we fix ρ* > 0 small enough to guarantee

ρminρ0,κ1p(1y),1,V(ρ)ε,andF(ρ)ε, (4.40)

here we have abbreviated F(r)=:(r2+rp)Br(ξ0)(|Xu|p+|u|p+1)dξp(p).

Now, we are in the position to prove the assertion (Ak) by induction. We assume that (Ak) is true for up to some k ∈ ℕ. Then, we prove the first part of the assertion (Ak+1), that is, the one concerning Ψ (θk+1r, lξ0,θk+1r). For this we are going to prove that the small assumptions for the excess improvement in Lemma 4.3 are satisfied. Firstly, by the assumptions of (Ak) and the choices of ε*, κ* and ρ*, we deduce

Ψ(ξ0,θkr,lξ0,θkr)Ψ(ξ0,θkr,lξ0,θkr)+ω(Cy(ξ0,θkr))+V(θkr)+F(θkr)ε+ω(κ)+V(ρ)+F(ρ)4ε.

Now it is easy to check that our choice of ε* implies that the smallness condition assumptions (i)-(ii) in Lemma 4.3 are satisfied on the level θkr, that is, we have

Ψ12(ξ0,θkr,lξ0,θkr)2εδ2,

where we have used εδ216 due to the choice of εδ23δ2+48 in (4.39), and

y(θkr):=Ψ(θkr)+δ22μΨ(θkr)+μΨ(θkr)p23ε+δ223ε+3εp23ε+4δ223ε2=ε3δ2+48δ21.

Consequently, we apply Lemma 4.3 with the radius θk r instead of r, this leads to

Ψ(ξ0,θk+1r,lξ0,θk+1r)C4θ2Ψ(ξ0,θkr,lξ0,θkr)4C4θ2ε<ε,

by the choice of θ in (4.38). This is the result for the first part of (Ak+1).

Now it remains to show the second part, that is, the one concerning Cy(ξ0, θk+1 r). Since lξ0,θkr = uξ0,θkr + Xlξ0,θkr(ξ̄ ξ0¯ ), we can estimate

Cy(ξ0,θk+1r)=(θk+1r)pyBθk+1r(ξ0)|uuξ0,θk+1r|pdξ(θk+1r)pyBθk+1r(ξ0)|uuξ0,θkr|pdξ2p1(θk+1r)pyBθk+1r(ξ0)|ulξ0,θkr|pdξ+|Xlξ0,θkr|p(θk+1r)p2(θk+1r)pyθQBθkr(ξ0)|ulξ0,θkr|pdξ+|Xlξ0,θkr|p(θk+1r)p=2(θkr)p(1y)θQpyBθkr(ξ0)ulξ0,θkrθkrpdξ+|Xlξ0,θkr|pθp(1y).

Now we are going to estimate the term Bθkr(ξ0)ulξ0,θkrθkrpdξ. Similarly, we divide the domain of integration into two subsets Ω9:=Bθkr(ξ0)ulξ0,θkrθkr>1 and Ω10:=Bθkr(ξ0)ulξ0,θkrθkr1.

On the subset Ω9, we get

Ω9ulξ0,θkrθkrpdξ2Ω9Vulξ0,θkrθkr2dξ.

For the case of ulξ0,θkrθkr1 on Ω10, noting the fact of 1 < p < 2, we find

Ω10ulξ0,θkrθkrpdξΩ10ulξ0,θkrθkr2dξp22Ω10Vulξ0,θkrθkr2dξp2.

Therefore, we deduce the following estimate

Bθkr(ξ0)ulξ0,θkrθkrpdξ2Bθkr(ξ0)Vulξ0,θkrθkr2dξp22Ψp2(ξ0,θkr,lξ0,θkr)2εp2.

By means of (2.8) with the choice of luξ0,θkr, and the assumption Ak, we obtain

|Xlξ0,θkr|pQ2c0QQ+2θkrpBθkr(ξ0)|uuξ0,θkr|pdξ(Q2)(Q+2)c0Qp(θkr)p(y1)Cy(ξ0,θkr)(Q2)(Q+2)c0Qp(θkr)p(y1)κ.

Recalling that r ∈ (0, ρ*), we deduce that

Cy(ξ0,θk+1r4ρp(1y)θQpyεp2+(Q2)(Q+2)c0Qpθp(1y)κκ2+κ2κ,

where we have used the choice of ε* in (4.39), the choice of ρ* from (4.40) and θ in (4.38).

This proves the second part of the assertion (Ak+1) and finally complete the proof of the lemma. □

4.5 Proof of Theorem 1

Proof

By Lebesgue’s differentiation theorem, we obtain |Σ1Σ2| = 0. So our aim is to show that every ξ0Ω ∖ (Σ1Σ2) is a regular point. For every 0 < ρ < dist(ξ0, ∂Ω), by Sobolev-Poincaré type inequality, it follows

Ψ(ξ0,ρ,lξ0,ρ)=Bρ(ξ0)Vulξ0,ρρ2dξCP2Bρ(ξ0)VXu(Xu)ξ0,ρ2dξCP2C(p,M)Bρ(ξ0)V(Xu)V(Xu)ξ0,ρ2dξ. (4.41)

Moreover, for any y ∈ (0, 1) and ρ ≤ 1, the following estimate holds

Cy(ξ0,ρ)=ρpyBρ(ξ0)|uuξ0,ρ|pdξρppyBρ(ξ0)ulξ0,ρρpdξ.

If ulξ0,ρρ>1, we have

Bρ(ξ0)ulξ0,ρρpdξ2Bρ(ξ0)Vulξ0,ρρ2dξ.

If ulξ0,ρρ1, we obtain ulξ0,ρρpulξ0,ρρ2+1. Then, it implies

Bρ(ξ0)ulξ0,ρρpdξ2Bρ(ξ0)Vulξ0,ρρ2dξ+1.

So, it yields

Cy(ξ0,ρ)2ρppyBρ(ξ0)Vulξ0,ρρ2dξ+ρppyρppy2Ψ(ξ0,ρ,lξ0,ρ)+1. (4.42)

Keeping in mind the definitions of Σ1, Σ2, from the estimates (4.41) and (4.42), we know that there exists a radius ρ:0 < ρ < min{ρ*, dist(ξ0, ∂Ω)} such that

Ψ(ξ0,ρ,lξ0,ρ)<εandCy(ξ0,ρ)<κ.

Using the absolute continuity of the integral, there exists a neighborhood UΩ of ξ0 with

Ψ(ξ,ρ,lξ0,ρ)<εandCy(ξ,ρ)<κ,ξU.

Applying Lemma 4.4 in any point ξU, Then, we get

Ψ(ξ,θkρ,lξ0,θkρ)<εandCy(ξ,θkρ)<κ,ξU,kN. (4.43)

This together with Campanato’s characterization of Hölder continuous functions imply that

supξU,σ(0,ρ)σpyBσ(ξ)|uuξ,σ|pdη=supξU,σ(0,ρ)Cy(ξ,σ)<κ<.

Hence u Cloc0,y (U, ℝN).

Furthermore, it holds for |XuXlξ,σ| > 1

Bσ(ξ)|XuXlξ,σ|pdξ2Bσ(ξ)|V(XuXlξ,σ)|2dξ, (4.44)

and we have if |XuXlξ,σ| ≤ 1

Bσ(ξ)|XuXlξ,σ|pdξ2Bσ(ξ)|V(XuXlξ,σ)|2dξ+1. (4.45)

Combining (4.44) and (4.45) with (4.43) and (1.6), we get for y ∈ (0, 1)

supξU,σ(0,ρ)σp(1y)Bσ(ξ)|XuXlξ,σ|pdξsupξU,σ(0,ρ)σp(1y)2Bσ(ξ)|V(XuXlξ,σ)|2dξ+1supξU,σ(0,ρ)σp(1y)2CcΨ(ξ,σ,lξ,σ)+ω(Cy(ξ,σ))+V(σ)+F(σ)+1,

with abbreviation of F(r)=:(r2+rp)Br(ξ0)(|Xu|p+|u|p+1)dξp(p).

In view of the well known equivalence of Campanato and Morrey spaces for parameters λ ∈ (0, Q), it yields XuLp,λ(U, ℝ2n×N) with λ = Qp(1 − y). In particular, the parameter λ can be chosen arbitrary chose to Q. This concludes the proof of Theorem 1.1. □

5 Partial Hölder continuity for sub-quadratic Natural growth

In this section, we prove the partial regularity result of Theorem 1.2 under the assumptions of sub-quadratic natural structure conditions (H1)-(H4) and (HN). In this case, we will need to restrict ourselves to bounded solution of (1.1), where the bound M = supΩ |u| satisfies the smallness assumption

2a(M)(M+M0)3(1+M02)2p2<ν

in our present situation with a(M) defined in (1.8). Such a similar smallness condition is necessary for a partial regularity result even in the elliptic case with quadratic growth (p = 2); for example, see [18].

We first introduce an elementary inequality showed by Kanazawa in [26]. It is useful to get suitable estimates for the natural growth term in proving Caccioppoli-type inequality.

Lemma 5.1

Consider fixed a, b ≥ 0, p ≥ 1. Then, for any ε > 0, there exists K = K(p, ε) ≥ 0 satisfying

(a+b)p(1+ε)ap+Kbp. (5.1)

Lemma 5.2

(Caccioppoli-type inequality). Let uHW1,p(Ω, ℝN) ∩ L (Ω, ℝN) be weak solutions of the systems (1.1) under the assumptions (H1)-(H4)-(HN) with ν > 2a(M)(M + M0) 3(1+M02)2p2. Then, for any ξ0 = (x1, ⋯, xn, y1, ⋯, yn, t) ∈ Ω and r ≤ 1 with Br(ξ0) ⊂ ⊂ Ω, and any horizontal affine functions l : ℝ2n → ℝN with |l( ξ0¯ )| + |Xl| ≤ M0, we have the estimate

Br2(ξ0)|V(XuXl)|2dξCcBr(ξ0)Vulr2dξ+ωBr(ξ0)(|ul(ξ0¯)|p)dξ+V(r)+r2+rp,

where Cc is some positive constants depending only on Q, N, p, a, b, L, ν, M, M0, and the exponent p′ = pp1 .

Proof

Let φ = ϕ2(ul) be a testing function for sub-elliptic systems (1.1), where the standard cut-off function ϕ C0 (Br(ξ0), [0, 1]) with ϕ ≡ 1 on Br2(ξ0) and || ≤ 4r . By the same way as the case of controllable growth, we have for weak solutions u of the systems (1.1)

I0:=Br(ξ0)[Aiα(ξ,u,Xu)Aiα(ξ,u,Xl)]ϕ2(XuXl)dξ=2Br(ξ0)[Aiα(ξ,u,Xl)Aiα(ξ,u,Xu)]ϕ(ul)Xϕdξ+Br(ξ0)[Aiα(ξ,l(ξ0¯),Xl)Aiα(ξ,u,Xl)]Xφdξ+Br(ξ0)[(Aiα(,l(ξ0¯),Xl))ξ0,rAiα(ξ,l(ξ0¯),Xl)]Xφdξ+Br(ξ0)Bα(ξ,u,Xu)ϕ2(ul)dξ=:2I1+I2+I3+I4, (5.2)

with the obvious meaning for I0I4 .

With respect to the terms I0I3 , here, we choose the same estimates as (4.2), (4.7), (4.15) and (4.21), that is,

I0ν3(1+M02)p22Br(ξ0)ϕ2|V(XuXl)|2dξ, (5.3)
I12εBr(ξ0)ϕ2|V(XuXl)|2dξ+C(p,L,M0)ε11pBr(ξ0)Vulr2dξ, (5.4)
I22εBr(ξ0)ϕ2|V(XuXl)|2dξ+2εBr(ξ0)Vulr2dξ+C(p,L,M0)ε11pωBr(ξ0)|ul(ξ0¯)|pdξ, (5.5)
I32εBr(ξ0)ϕ2|V(XuXl)|2dξ+2εBr(ξ0)Vulr2dξ+C(p,L,M0)ε11pV(r). (5.6)

Now we are in the position to get an appropriate bound for the term I4 . By (H4), elementary inequality (5.1) and Young’s inequality, it yields

I4Br(ξ0)(a|Xu|p+b)ϕ2|ul|dξaBr(ξ0)|XuXl|+|Xl|pϕ2|ul|dξ+bBr(ξ0)ϕ2|ul|dξaBr(ξ0)(1+ϵ)|XuXl|p+1+K|Xl|pϕ2|ul|dξ+bBr(ξ0)rϕ2ulrdξ. (5.7)

We denote by I41 the first term of the right-hand side of (5.7). If |XuXl| ≥ 1, the following estimate holds

I412a(M0+M)(1+ϵ)Br(ξ0)|V(XuXl)|2ϕ2dξ+a1+K|Xl|pBr(ξ0)ϕrulrdξ.

If |XuXl| ≤ 1, we have |XuXl|p ≤ |XuXl|2 + 1. Then, it follows

I412a(M0+M)(1+ϵ)Br(ξ0)|V(XuXl)|2ϕ2dξ+a1+K(1+|Xl|p)Br(ξ0)ϕrulrdξ.

Combining these estimates above, we have

I42a(M0+M)(1+ϵ)Br(ξ0)|V(XuXl)|2ϕ2dξ+a1+K(1+|Xl|p)+bBr(ξ0)ϕrulrdξ. (5.8)

We denote by I42 the second term of the right-hand side of (5.8). If ulr1, it leads to

I42a1+K(1+|Xl|p)+bpBr(ξ0)ϕpulrpdξ+rpa1+K(1+|Xl|p)+bpBr(ξ0)ϕpVulr2dξ+rp.

If ulr1, it yields

I42a1+K(1+|Xl|p)+b2Br(ξ0)ϕ2ulr2dξ+r2a1+K(1+|Xl|p)+b2Br(ξ0)ϕ2Vulr2dξ+r2.

So, we finally arrive at

I42a(M0+M)(1+ϵ)Br(ξ0)|V(XuXl)|2ϕ2dξ+a1+K(1+M0p)+b2Br(ξ0)Vulr2dξ+r2+rp. (5.9)

Combining (5.3)-(5.6), (5.9) and (5.2), we have

ν3(1+M02)p226ε2a(M0+M)(1+ϵ)Br(ξ0)|V(XuXl)|2ϕ2dξC(p,L,M0)ε11p+4ε+a1+K(1+M0p)+b2Br(ξ0)Vulr2dξ+C(p,L,M0)ε11pωBr(ξ0)|ul(ξ0¯)|pdξ+C(p,L,M0)ε11pV(r)+r2+rp. (5.10)

Noting that the smallness condition 2a(M + M0) 3(1+M02)2p2 < ν, we fix the constant ε > 0 small sufficiently such that the coefficient ν3(1+M02)p226ε2a(M+M0)(1+ϵ)>0. Dividing the inequality (5.10) by the positive constant, finally we deduce

Br2(ξ0)|V(XuXl)|2dξCcBr(ξ0)Vulr2dξ+ωBr(ξ0)|ul(ξ0¯)|pdξ+V(r)+r2+rp,

where Cc = C(Q, p, a, b, L, ν, M0, M). This yields the claim. □

For sake of simplicity, we motivated the form of the Caccioppoli inequalities in Lemma 5.2. We write

Φ¯(ξ0,r,l):=Br(ξ0)|V(XuXl)|2dξ,Ψ¯(ξ0,r,l):=Br(ξ0)Vulr2dξ,Ψ¯(ξ0,r,l):=Ψ(ξ0,r,l)+ωBr(ξ0)|ul(ξ0¯)|pdξ+V(r)+r2+rp.

Lemma 5.3

Under the assumptions of Theorem 1.1 are satisfied, B2ρ(ξ0) ⊆ Ω with ρρ0 and an arbitrary horizontal function l : ℝ2n → ℝN, we define

A=DPAiα(,l(ξ0¯),Xl)ξ0,ρandw=ul,

then, w is approximately 𝓐-harmonic in the sense that

Bρ(ξ0)A(Xw,Xφ)dξC1Ψ¯(2ρ)+μΨ¯12(2ρ)+μΨ¯1p(2ρ)supBρ(ξ0)|Xφ|

for all φ C0 (Bρ(ξ0), ℝN), and the positive constant C1 = C(p, a, b, M0, L, Cc).

Proof

The proof is similar as the case of controllable growth. Here, we just give the different estimate for the natural growth term, that is,

J232a(M0+M)(1+ϵ)Bρ(ξ0)|V(XuXl)|2ϕ2dξ+a1+K(1+M0p)+b2Bρ(ξ0)Vulρ2dξ+ρ2+ρpC(p,a,b,M0,M)Φ¯(ρ)+Ψ¯(ρ)+ρ2+ρpC(p,a,b,M0,M,Cc)Φ¯(ρ)+Ψ¯(ρ),

where we have used the bound for the natural growth term I4 in (5.9). The rest procedure is very similar as the proof in Lemma 4.2, and we omit them. So we obtain the claim. □

Applying Lemma 5.2 and Lemma 5.3, we can establish the improvement estimate for Excess functional Ψ with the same form as Lemma 4.3, that is,

Lemma 5.4

Suppose that the assumptions of Theorem 1.2 are satisfied and consider a ball Br(ξ0) ⊆ Ω with rρ0. For the constants δ = δ (Q, N, p, L, ν, θQ+4) ∈ (0, 1] and y ∈ (0, 1] from the 𝓐-harmonic approximation Lemma 3.3, we let 0 < θ 14 be arbitrary and also impose the following smallness conditions:

  1. Ψ¯12(r)<δ2;

  2. y:=Ψ¯(r)+δ22μΨ¯12(r)+μΨ¯1p(r)21.

Then, the following excess improvement estimate holds

Ψ¯(ξ0,θr,lξ0,θr)C4θ2Ψ¯(ξ0,r,lξ0,r),

where constants C4 depend only on Q, N, p, a, b, ν, δ and L.

By Lemma 5.4, the iteration for the Ψ-excess and the Cy-excess can be obtained as follows,

Lemma 5.5

Suppose that the assumptions of Theorem 1.2 are satisfied. For every y ∈ (0, 1), there are constant ε*, κ*, ρ* and θ ∈ (0, 18 ], if

Ψ¯(ξ0,r,lξ0,r)<εandCy(ξ0,r)<κ, (A0)

for r ∈ (0, ρ*) with Br(ξ0) ⊂ Ω, then,

Ψ¯(ξ0,θkr,lξ0,θkr)<εandCy(ξ0,θkr)<κ, (Ak)

respectively, for every k ∈ ℕ.

Proof of Theorem 1.2

It is enough to use Lemma 5.5, and repeat the procedure for the proof of Theorem 1.1 in the previous Subsection 4.5.


Tel.:+86 0797 8393663; fax: +86 0797 8395933

Acknowledgement

The authors wish to thank the referees for their careful reading of my manuscript and valuable suggestions.

  1. Funding: The research is supported by the National Natural Science Foundation of China (No.11661006), and the Science and Technology Planning Project of Jiangxi Province, China (No. GJJ190741).

References

[1] V. Bögelein, F. Duzaar, J. Habermann, C. Scheven, Partial Hölder continuity for discontinuous elliptic problems with VMO-coefficients, Proc. Lond. Math. Soc. 103 (2011), 371-404.10.1112/plms/pdr009Search in Google Scholar

[2] M. Bramanti, An Invitation to Hypoelliptic Operators and Hörmander’s Vector Fields, Springer, 2014.10.1007/978-3-319-02087-7Search in Google Scholar

[3] L. Capogna, D. Danielli, N. Garofalo, An embedding theorem and the Harnack inequality for nonlinear sub-elliptic equations, Comm. Partial Differential Equations 18 (1993), 1765-1794.10.1080/03605309308820992Search in Google Scholar

[4] M. Carozza, N. Fusco, G. Mingione, Partial regularity of minimizers of quasiconvex integrals with sub-quadratic growth, Ann. Mat. Pura Appl. 175 (1998), 141-164.10.1017/S0308210500002900Search in Google Scholar

[5] L. Capogna, N. Garofalo, Regularity of minimizers of the calculus of variations in Carnot groups via hypoellipticity of systems of Hörmander type, J. European Math. Society 5 (2003), 1-40.10.1007/s100970200043Search in Google Scholar

[6] S. Chen, Z. Tan, The method of A-harmonic approximation and optimal interior partial regularity for nonlinear elliptic systems under the controllable growth condition, J. Math. Anal. Appl. 335 (2007), 20-42.10.1016/j.jmaa.2007.01.042Search in Google Scholar

[7] G. Di Fazio, M. Fanciullo, Gradient estimates for elliptic systems in Carnot-Carathéodory spaces. Comment. Math. Univ. Caroline. 43 (2002), 605-618.Search in Google Scholar

[8] F. Duzaar, A. Gastel, Nonlinear elliptic systems with Dini continuous coefficients, Arch. Math. 78 (2002), 58-73.10.1007/s00013-002-8217-1Search in Google Scholar

[9] F. Duzaar, J. F. Grotowski, Partial regularity for nonlinear elliptic systems: The method of A-harmonic approximation, Manuscripta Math. 103 (2000), 267-298.10.1007/s002290070007Search in Google Scholar

[10] F. Duzaar, J. F. Grotowski, M. Kronz, Regularity of almost minimizers of quasi-convex variational integrals with sub-quadratic growth, Ann. Mat. Pura Appl. 184 (2005), 421-448.10.1007/s10231-004-0117-5Search in Google Scholar

[11] F. Duzaar, A. Gastel, G. Mingione, Elliptic systems, singular sets and Dini continuity, Comm. Partial Differential Equations 29 (2004), 1215-1240.10.1081/PDE-200033734Search in Google Scholar

[12] F. Duzaar, G. Mingione, The p-harmonic approximation and the regularity of p-harmonic maps, Calc. Var. Partial Differention Equations 20 (2004), 235-256.10.1007/s00526-003-0233-xSearch in Google Scholar

[13] F. Duzaar, G. Mingione, Regularity for degenerate elliptic problems via p-harmonic approximation, Ann. Inst. H. Poincaré Anal. Non Linéaire 21 (2004), 735-766.10.1016/j.anihpc.2003.09.003Search in Google Scholar

[14] Y. Dong, P. Niu, Estimates in Morrey spaces and Hölder continuity for weak solutions to degenerate elliptic systems, Manuscripta Math. 138 (2012), 419-437.10.1007/s00229-011-0498-xSearch in Google Scholar

[15] F. Duzaar, K. Steffen, Optimal interior and boundary regularity for almost minimizers to elliptic variational integrals, J. Reine Angew. Math. 546 (2002), 73-138.10.1515/crll.2002.046Search in Google Scholar

[16] A. Föglein, Partial regularity results for sub-elliptic systems in the Heisenberg group, Calc. Var. Partial Differential Equations 32 (2008), 25-51.10.1007/s00526-007-0127-4Search in Google Scholar

[17] M. Foss, G. Mingione, Partial continuity for elliptic problems, Ann. Inst. H. Poincaré Anal. Non Linéaire 25 (2008), 471-503.10.1016/j.anihpc.2007.02.003Search in Google Scholar

[18] M. Giaquinta, Multiple Integrals in the Calculus of Variations and Nonlinear Elliptic Systems, Princeton University Press, Princeton, NJ, 1983.10.1515/9781400881628Search in Google Scholar

[19] D. Gao, P. Niu, J. Wang, Partial regularity for degenerate sub-elliptic systems associated with Hörmander’s vector fields, Nonlinear Anal. 73 (2010), 3209-3223.10.1016/j.na.2010.07.001Search in Google Scholar

[20] C. S. Goodrich, M. A. Ragusa, A. Scapellato, Partial regularity of solutions to p(x)-Laplacian PDEs with discontinuous coefficients, J. Differential Equations 268 (2020), 5440-5468.10.1016/j.jde.2019.11.026Search in Google Scholar

[21] G. Lu, Embedding theorems on Campanato-Morrey space for vector fields on Hörmander type, Approx. Theory Appl. 14 (1998) 69-80.Search in Google Scholar

[22] S. Polidoro, M. A. Ragusa, Harnack inequality for hypoelliptic ultraparabolic equations with a singular lower order term, Revista Matematica Iberoamericana 24 (2008), 1011-1046.10.4171/RMI/565Search in Google Scholar

[23] M. A. Ragusa, A. Tachikawa, Regularity of minimizers of some variational integrals with discontinuity, Z. Anal. Anwend. 27 (2008), 469-482.10.4171/ZAA/1366Search in Google Scholar

[24] A. Scapellato, New perspectives in the theory of some function spaces and their applications, AIP Conference Proceedings 1978, 140002(2018); https://doi.org/10.1063/1.5043782.10.1063/1.5043782Search in Google Scholar

[25] E. Shores, Regularity theory for weak solutions of systems in Carnot groups, Ph. D. Thesis, University of Arkansas. 2005.Search in Google Scholar

[26] T. Kanazawa, Partial regularity for elliptic systems with VMO-coefficients, Riv. Math. Univ. Parma (N.S.) 5 (2014), 311-333.Search in Google Scholar

[27] Z. Tan, Y. Wang, S. Chen, Partial regularity in the interior for discontinuous inhomogeneous elliptic system with VMO-coefficients, Ann. Mat. Pura Appl. 196 (2017), 85-105.10.1007/s10231-016-0564-9Search in Google Scholar

[28] Z. Tan, Y. Wang, S. Chen, Partial regularity up to the boundary for solutions of sub-quadratic elliptic systems, Adv. Nonlinear Anal. 7 (2018), 469-483.10.1515/anona-2016-0054Search in Google Scholar

[29] G. Lu, The sharp Poincaré inequality for free vector fields: an endpoint result, Revista Matematica Iberoamericana 10 (1994), 453-466.10.4171/RMI/158Search in Google Scholar

[30] J. Wang, D. Liao, Optimal partial regularity for sub-elliptic systems with sub-quadratic growth in Carnot groups, Nonlinear Anal. 75 (2012), 2499-2519.10.1016/j.na.2011.10.045Search in Google Scholar

[31] J. Wang, D. Liao, S. Gao, Z. Yu, Optimal partial regularity for sub-elliptic systems with Dini continuous coefficients under the superquadratic natural growth, Nonlinear Anal. 114 (2015), 13-25.10.1016/j.na.2014.10.028Search in Google Scholar

[32] J. Wang, Juan J. Manfredi, Partial Hölder continuity for nonlinear sub-elliptic systems with VMO-coefficients in the Heisenberg group, Adv. Nonlinear Anal. 7 (2018), 96-114.10.1515/anona-2015-0182Search in Google Scholar

[33] J. Wang, S. Zhang, Q. Yang, Partial regularity for discontinuous sub-elliptic systems with sub-quadratic growth in the Heisenberg group, Nonlinear Anal. 195 (2020), 111719; https://doi.org/10.1016/j.na.2019.111719.10.1016/j.na.2019.111719Search in Google Scholar

[34] C. Xu, C. Zuily, Higher interior regularity for quasilinear sub-elliptic systems, Calc. Var. Partial Differential Equations 5 (1997), 323-343.10.1007/s005260050069Search in Google Scholar

[35] S. Zheng, Partial regularity for quasi-linear elliptic systems with VMO coefficients under the natural growth, Chinese Ann. Math. Ser. A 29 (2008), 49-58.Search in Google Scholar

[36] S. Zheng, Z. Feng, Regularity of sub-elliptic p-harmonic systems with subcritical growth in Carnot group, J. Differential Equations 258 (2015), 2471-2494.10.1016/j.jde.2014.12.020Search in Google Scholar

Received: 2020-03-19
Accepted: 2020-06-29
Published Online: 2020-08-07

© 2021 Jialin Wang et al., published by De Gruyter

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

Downloaded on 25.4.2024 from https://www.degruyter.com/document/doi/10.1515/anona-2020-0145/html
Scroll to top button