We study partial Hölder regularity for nonlinear elliptic systems in divergence form with double-phase growth, modeling double-phase non-Newtonian fluids in the stationary case.
Citation: Giovanni Scilla, Bianca Stroffolini. Partial regularity for steady double phase fluids[J]. Mathematics in Engineering, 2023, 5(5): 1-47. doi: 10.3934/mine.2023088
[1] | Lucio Boccardo, Giuseppa Rita Cirmi . Regularizing effect in some Mingione’s double phase problems with very singular data. Mathematics in Engineering, 2023, 5(3): 1-15. doi: 10.3934/mine.2023069 |
[2] | Yangyang Qiao, Qing Li, Steinar Evje . On the numerical discretization of a tumor progression model driven by competing migration mechanisms. Mathematics in Engineering, 2022, 4(6): 1-24. doi: 10.3934/mine.2022046 |
[3] | Donatella Danielli, Rohit Jain . Regularity results for a penalized boundary obstacle problem. Mathematics in Engineering, 2021, 3(1): 1-23. doi: 10.3934/mine.2021007 |
[4] | Francesca Tedeschi, Giulio G. Giusteri, Leonid Yelash, Mária Lukáčová-Medvid'ová . A multi-scale method for complex flows of non-Newtonian fluids. Mathematics in Engineering, 2022, 4(6): 1-22. doi: 10.3934/mine.2022050 |
[5] | Mirco Piccinini . A limiting case in partial regularity for quasiconvex functionals. Mathematics in Engineering, 2024, 6(1): 1-27. doi: 10.3934/mine.2024001 |
[6] | Catharine W. K. Lo, José Francisco Rodrigues . On an anisotropic fractional Stefan-type problem with Dirichlet boundary conditions. Mathematics in Engineering, 2023, 5(3): 1-38. doi: 10.3934/mine.2023047 |
[7] | Pawan Kumar, Christina Surulescu, Anna Zhigun . Multiphase modelling of glioma pseudopalisading under acidosis. Mathematics in Engineering, 2022, 4(6): 1-28. doi: 10.3934/mine.2022049 |
[8] | Zeljko Kereta, Valeriya Naumova . On an unsupervised method for parameter selection for the elastic net. Mathematics in Engineering, 2022, 4(6): 1-36. doi: 10.3934/mine.2022053 |
[9] | Gennaro Ciampa, Gianluca Crippa, Stefano Spirito . Propagation of logarithmic regularity and inviscid limit for the 2D Euler equations. Mathematics in Engineering, 2024, 6(4): 494-509. doi: 10.3934/mine.2024020 |
[10] | Nicolai Krylov . On parabolic Adams's, the Chiarenza-Frasca theorems, and some other results related to parabolic Morrey spaces. Mathematics in Engineering, 2023, 5(2): 1-20. doi: 10.3934/mine.2023038 |
We study partial Hölder regularity for nonlinear elliptic systems in divergence form with double-phase growth, modeling double-phase non-Newtonian fluids in the stationary case.
Dedicated to Rosario Mingione for his 50th birthday.
Our research was inspired by two main contributions of Rosario: the regularity for non-Newtonian fluids [2] and the double phase problems, [4,13].
This article deals with nonlinear elliptic systems in divergence form, modeling double-phase non-Newtonian fluids in the stationary case:
divu=0,diva(x,Eu)+Dπ=u[Du]+f in Ω. | (1.1) |
Here Ω⊆Rn denotes an open, bounded set, n≥3, the vector-valued map u:Ω→Rn can be interpreted as the stationary velocity field of a fluid, and the scalar function π:Ω→R plays the role of the pressure. The stress tensor will have double phase growth, as a function of the symmetric gradient Eu.
For the nonlinear diffusion term a:Ω×Rn×nsym→Rn×nsym we are considering a double phase growth condition of the type
H(x,t):=tp+μ(x)tq, | (1.2) |
where
2≤p<q≤p+αpn,0<α≤1, | (1.3) |
and the modulating coefficient μ is non negative, bounded and Hölder continuous with exponent α.
Notice that in the region {μ=0}, we have H(x,t)=tp so that H has p-phase, while in the region {μ>0} the function H has q-phase. Moreover, for each x∈Ω such that μ(x)>0, the function t→H(x,t) is an N-function (see Section 2.2) complying with (2.1) where g1=p and g2=q. The precise assumptions are:
{|a(x,ξ)|+|Dξa(x,ξ)|(1+|ξ|)≤LH′(x,1+|ξ|)⟨Dξa(x,ξ)λ,λ⟩≥νH″(x,1+|ξ|)|λ|2. | (1.4) |
for every x∈Ω and ξ∈Rn×nsym, λ∈Rn×n and for some 0<ν≤L, where H′(x,t) and H″(x,t) denote the first and second derivative of t→H(x,t), respectively. Note that the second inequality in (1.4) implies
(a(x,ξ1)−a(x,ξ2)):(ξ1−ξ2)≥˜νH″(x,1+|ξ1|+|ξ2|)|ξ1−ξ2|2 | (1.5) |
for every x∈Ω and ξ1,ξ2∈Rn×nsym. We further assume the existence of a nondecreasing and concave function ω:[0,∞)→[0,1] with ω(0)=0 such that
|a(x1,ξ)−a(x2,ξ)|≤L|μ(x1)−μ(x2)|(1+|ξ|)q−1, | (1.6) |
and
|Dξa(x,ξ1)−Dξa(x,ξ2)|≤Lω(|ξ1−ξ2|1+|ξ1|+|ξ2|)H″(x,1+|ξ1|+|ξ2|), | (1.7) |
for every x,x1,x2∈Ω and ξ,ξ1,ξ2∈Rn×nsym.
For the force term f we require that
f∈Ln(1+β)loc(Ω;Rn) | (1.8) |
for some β>0.
A pair (u,π)∈W1,1(Ω;Rn)×W1,1(Ω,R) with H(⋅,|Eu|)∈L1loc(Ω) is a weak solution to (1.1) if and only if divu=0 in Ω in the sense of distributions and
∫Ω[⟨a(x,Eu),Eφ⟩+πdivφ]dx=∫Ω[u[Du]+f]⋅φdx | (1.9) |
holds for all φ∈C∞0(Ω,Rn). If we test with divergence free vector fields
φ∈C∞0,div(Ω,Rn):={ψ∈C∞0(Ω,Rn):divψ=0}, |
the pressure term in (1.9) vanishes and the system reduces to divu=0 in Ω in the sense of distributions and
∫Ω⟨a(x,Eu),Eφ⟩dx=∫Ω[u[Du]+f]⋅φdx | (1.10) |
whenever φ∈C∞0,div(Ω,Rn). Within this setting, we say that u∈W1,1(Ω;Rn) with H(⋅,|Eu|)∈L1loc(Ω) is a weak solution to (1.1) if and only if divu=0 in Ω in the sense of distributions and (1.10) holds.
The main result of the paper is a partial Morrey regularity result:
Theorem 1.1. Let H:Ω×[0,+∞)→[0,+∞) be defined as in (1.2), with (1.3). Assume that the vector field a:Ω×Rn×nsym→Rn×nsym complies with (1.4), (1.6), (1.7), and that (1.8) holds. Let (u,π)∈W1,1(Ω;Rn)×W1,1(Ω,R) with H(⋅,|Eu|)∈L1loc(Ω) be a weak solution to (1.1). Then there exists an open subset Ω0⊂Ω such that
u∈C0,βloc(Ω0,Rn),|Ω∖Ω0|=0 |
for every β∈(0,1). Moreover, Ω∖Ω0⊂Σ1∪Σ2 where
Σ1:={x0∈Ω:liminfϱ↘0−∫Bϱ(x0)|Du−(Du)x0,ϱ|dx>0},Σ2:={x0∈Ω:limsupϱ↘0[−∫Bϱ(x0)H(x,1+|Eu|)dx+(|Du|p)x0,ϱ+|(u)x0,ϱ|]=+∞}. |
The study of PDEs/functionals with (p,q) growth, corresponding to the case μ(x)≡1, was initiated by Marcellini in the 90's [31] and has been generalized in many directions, see [11,15,21,32,33,34] and references therein. The main feature of these problems is a gap between coercitivity and growth condition, that can eventually lead even to unbounded minimizers/solutions. In order to prevent such irregular behaviour, one has to impose a bound on the ratio qp<1+1n. This bound has been improved recently by Bella and Schäffner to qp<1+2n−1, [5]. More recently, in [14,16], some tools from nonlinear potential theory have been employed to infer sharp partial regularity results for relaxed minimizers of degenerate/singular, nonuniformly elliptic quasiconvex functionals of (p,q)-growth.
As for the double phase problems, the ellipticity could eventually change from p-phase to q-phase, since the modulating coefficient could annihilate. Pioneer were Colombo and Mingione, Baroni, Colombo and Mingione who established Morrey estimates for local minimizers under the condition qp<1+αn, where α is the modulus of continuity of the modulating coefficient, and also they considered borderline cases (of logarithmic type) and a unified approach to variable exponent, [4,13]. Their method relies on a refined alternative between p-phase and (p,q)-phase at every scale combined with an exit time argument. Their intrinsic approach was pushed further by Ok, who was able to find an easier proof in the superquadratic case, [36] and [37].
In the literature, nonlinear elliptic systems of (1.1) when the symmetric gradient has p(x) growth are known as electrorheological fluids. The model was introduced by Rajagopal and Růžička, [39], while the existence theory was studied by Růžička, [40], see also [22]. Acerbi and Mingione, [2], first established regularity for solutions of such systems. They proved that the gradient is partially Hölder continuous on a set of full measure under suitable continuity assumptions on the exponent p(x).
Our aim is to study regularity properties of solutions of the system (1.1) where the symmetric gradient presents a double phase growth. Featuring the double phase case considered by Colombo and Mingione, we cannot expect more than a partial Morrey regularity result. It is worth mentioning the paper [7] where the authors considered electrorheological fluids with discontinuous coefficients and proved Morrey partial regularity.
Recently, the uniqueness of small solutions for steady double phase fluids has been investigated in [1] for 65<p<2<q.
We briefly explain the strategy of the proof of the main result. First, we stress the fact that our argument relies on two technical major ingredients which, to the best of our knowledge, appear for the first time in this paper.
The first one is the A-Stokes approximation lemma for A-Stokes systems with general growth (Theorem 2.10), which extends an analogous result for Stokes systems with p-growth provided by [10, Theorem 4.2]. The second one, contained in Lemma 2.11, is a higher integrability result for weak solutions to double phase Stokes systems (1.1) with lower order terms. Namely, we can find an exponent s>1 depending on the data of the problem, such that
(−∫Br(x0)[H(x,1+|Eu|)]sdx)1s≤c−∫B2r(x0)H(x,1+|Eu|)dx+c−∫B2r(x0)(rp2|Du|p+|u|p∗+1)dx. |
Analogous results were obtained for Stokes systems with p(⋅)-growth, see [2, Theorem 4.2] and [7, Lemma 2.10].
When addressing the problem of partial regularity, the leading quantity is the Campanato-type excess functional
Φ(x0,ϱ):=−∫Bϱ(x0)H1+|(Eu)x0,ϱ|(x0,|Eu−(Eu)x0,ϱ|)dx |
(see (3.1)). We are dealing with the non-degenerate setting; i.e., when
Φ(x0,ϱ)≤H(x0,1+|(Eu)x0,ϱ|), |
so that we can linearize the problem, via the A-Stokes approximation, Theorem 2.10. The linearization procedure of Lemma 3.3 provides a suitable rescaling w of the solution u which is shown to be approximately A-Stokes on a ball Bϱ(x0), and the deviation from being A-Stokes is measured in terms of the "hybrid" excess
Φ(x0,ϱ)+ϱ˜γH(x0,1+|(Eu)x0,ϱ|). | (1.11) |
Then, if (1.11) is small enough, Theorem 2.10 applies and ensures the existence of a smooth A-Stokes function h which is close in an integral sense to w and for which classical decay estimates are available, see [23]. Scaling back from w to u, this allows to prove an excess-decay estimate, which, in turn, permits the iteration of the rescaled excess Φ(x0,ϱ)H(x0,1+|(Eu)x0,ϱ|) and of a "Morrey-type" excess
Θ(x0,ϱ):=ϱ12[H(x0,⋅)]−1(−∫BϱH(x0,1+|Du|)dx) |
at each scale.
Namely, there exists ϑ∈(0,1) such that, if the smallness conditions
Φ(x0,ϱ)H(x0,1+|(Eu)x0,ϱ|)≤ε∗,Θ(x0,ϱ)≤δ∗,|(u)x0,ϱ|≤12M |
hold on some ball Bϱ(x0), then
Φ(x0,ϑmϱ)H(x0,1+|(Eu)x0,ϑmϱ|)≤ε∗,Θ(x0,ϑmϱ)≤δ∗,|(u)x0,ϑmϱ|≤M |
hold for every m=0,1,….. Notice that, if on one hand |(Du)x0,ϱ| might blow up in the iteration since we cannot expect C1-regularity; on the other hand, the Morrey excess Θ(x0,ϑkϱ) stays bounded, exactly as it should be for a C0,α-regularity result. The smallness of Θ at any level ensure Hölder continuity of u at x0 provided the excess functionals Φ and Θ are small at some initial radius ϱ (actually, this holds in a neighborhood of x0, since these smallness conditions are open). Finally, it is then proven that such a smallness condition on the excesses is indeed satisfied on the complement of the set Σ1∪Σ2 of Theorem 1.1. Notice that we were already using the Morrey type excess in [27], where we faced the partial regularity for discontinuous quasiconvex integrals with general growth and [38] where we considered the boundary partial regularity.
Outline of the paper. The paper is organized as follows. Section 2 collects some basic definitions and auxiliary results useful throughout the paper. Specifically, in Section 2.1 we fix some notation, Section 2.2 contains basic facts about Orlicz functions, while Sections 2.3–2.5 concern with typical auxiliary results in the study of systems depending on the symmetrized gradient, as a lemma of Bogowskiǐ and Sobolev-Korn-type inequalities. In Section 2.6 we formulate the A-Stokes approximation lemma, which together with the self improving properties of weak solutions (Section 2.7), is a key tool in the linearization procedure. The problem of the partial regularity of weak solutions starts with Section 3. Namely, in Section 3.1 we prove the necessary Caccioppoli-type estimates, in Section 3.2 we show that a suitable scaling of the weak solution is almost A-Stokes, while in Section 3.3 we get some excess decay estimates. Finally, Section 4 is entirely devoted to the proof of the main result, Theorem 1.1.
We denote by Ω an open bounded domain of Rn. For x0∈Rn and r>0, Br(x0) is the open ball of radius r centred at x0. In the case x0=0, we will often use the shorthand Br in place of Br(x0). If f∈L1(Br(x0)), we denote the average of f by
(f)x0,r:=−∫Br(x0)fdx, |
and we use the abbreviate notation (f)r for (f)0,r. We denote by Rn×nsym the set of all symmetric n×n matrices. For x,y∈Rn, we denote their tensor product by x⊗y:={xiyj}i,j∈Rn2 ad their tensor symmetric product by x⊙y:=12(x⊗y+y⊗x)∈Rn×nsym. For a function u=(ui)∈L1(Ω), we denote by Eu and Wu its symmetric and skew-symmetric distributional derivative, respectively:
Eu≡(Eu)i,j:=∂jui+∂iuj2,Wu≡(Wu)i,j:=∂jui−∂iuj2. |
If p>1, then p′:=pp−1 denotes the conjugate exponent of p. If 1<p<n, the number p∗:=npn−p stands for the Sobolev conjugate exponent of p, whereas p∗ is any real number if p≥n.
We recall here some elementary definitions and basic results about Orlicz functions. The following definitions and results can be found, e.g., in [3,6,28,30].
A real-valued function G:[0,∞)→[0,∞) is said to be an N-function if it is convex and satisfies the following conditions: G(0)=0, G admits the derivative G′ and this derivative is right continuous, non-decreasing and satisfies G′(0)=0, G′(t)>0 for t>0, and limt→∞G′(t)=∞.
We assume that G also satisfies
g1≤inft>0tG′(t)G(t)≤supt>0tG′(t)G(t)≤g2, | (2.1) |
for some 1<g1≤g2<∞. For instance, G(t):=tp, 1<p<∞, is an N-function complying with (2.1) where g1=g2=p. Also G(t):=H(x0,t), where H is defined in (1.2), is an N-function satisfying (2.1) with g1=p and g2=q. Moreover, it can be seen that if G∈C2(0,∞) and satisfies
0<g1−1≤inft>0tG″(t)G′(t)≤supt>0tG″(t)G′(t)≤g2−1, | (2.2) |
then G is an N-function and (2.1) holds.
G∗ denotes the Young-Fenchel-Yosida conjugate function of G, given by G∗(t)=sups≥0(st−G(s)). It is again an N-function; it satisfies (2.1) with g2g2−1 and g1g1−1 in place of g1 and g2, respectively. Also, (G∗)∗=G.
Proposition 1. Let G:[0,∞)→[0,∞) be an N-function complying with (2.1). Then
(i) the mappings
t∈(0,+∞)→G′(t)tg1−1,G(t)tg1andt∈(0,+∞)→G′(t)tg2−1,G(t)tg2 |
are increasing and decreasing, respectively. In particular,
G(at)≤ag1G(t),G′(at)≤ag1−1G′(t),0<a<1,G(bt)≤bg2G(t),G′(bt)≤bg2−1G′(t),b>1. | (2.3) |
Moreover,
G∗(at)≤ag2g2−1G(t),G∗(bt)≤ag1g1−1G(t). |
(ii)
G(s+t)≤2g2−1(G(s)+G(t)); |
(iii) (Young's inequality) for any λ∈(0,1] it holds that
st≤λ−g2+1G(s)+λG∗(t),st≤λG(s)+λ−1g1−1G∗(t). | (2.4) |
(iv) there exists a constant c=c(g1,g2)>1 such that
c−1G(t)≤G∗(t−1G(t))≤cG(t). | (2.5) |
We say that G satisfies the Δ2-condition if there exists c>0 such that for all t≥0 holds G(2t)≤cG(t). We denote the smallest possible such constant by Δ2(G). Since G(t)≤G(2t), the Δ2-condition is equivalent to G(2t)∼G(t), where "∼" indicates the equivalence between N-functions.
In particular, from (2.3) it follows that both G and G∗ satisfy the Δ2-condition with constants Δ2(G) and Δ2(G∗) determined by g1 and g2. We will denote by Δ2(G,G∗) constants depending on Δ2(G) and Δ2(G∗). Moreover, for t>0 we have
G(t)∼G′(t)t,G(t)∼G″(t)t2. | (2.6) |
The following inequalities hold for the inverse function G−1:
a1g1G−1(t)≤G−1(at)≤a1g2G−1(t) | (2.7) |
for every t≥0 with 0<a≤1. The same result holds also for a≥1 by exchanging the role of g1 and g2.
Another important set of tools are the shifted N-functions {Ga}a≥0 (see [17]). We define for t≥0
Ga(t):=∫t0G′a(s)dswith G′a(t):=G′(a+t)ta+t. | (2.8) |
We have the following relations:
Ga(t)∼G′a(t)t;Ga(t)∼G″(a+t)t2∼G(a+t)(a+t)2t2∼G′(a+t)a+tt2, | (2.9) |
G(a+t)∼[Ga(t)+G(a)],G(a+t)≤cGa(t)for t≥a. | (2.10) |
The families {Ga}a≥0 and {(Ga)∗}a≥0 satisfy the Δ2-condition uniformly in a≥0.
We recall also that, by virtue of [17, Lemma 30], uniformly in λ∈[0,1] and a≥0 holds
G∗a(λG′(a))∼λ2G(a). | (2.11) |
The following lemma (see x[19, Corollary 26]) deals with the change of shift for N-functions.
Lemma 2.1 (change of shift). Let G be an N-function with Δ2(G),Δ2(G∗)<∞. Then for any η>0 there exists cη>0, depending only on η and Δ2(G), such that for all a,b∈Rd and t≥0
G|a|(t)≤cηG|b|(t)+ηG|a|(|a−b|). | (2.12) |
Let P0,P1∈RN×n, θ∈[0,1] and define Pθ:=(1−θ)P0+θP1. Then the following result holds (see [17, Lemma 20]).
Lemma 2.2. Let G be a N-function with Δ2(G,G∗)<∞. Then uniformly for all P0,P1∈RN×n with |P0|+|P1|>0 holds
∫10G′(|Pθ|)|Pθ|dθ∼G′(|P0|+|P1|)|P0|+|P1| |
where the constants only depend on Δ2(G,G∗).
In view of the previous considerations, the same proposition holds true for the shifted functions, uniformly in a≥0.
By LG and W1,G we denote the classical Orlicz and Orlicz-Sobolev spaces, i.e., f∈LG iff ∫G(|f|)dx<∞ and f∈W1,G iff f,Df∈LG. The space W1,G0(Ω) will denote the closure of C∞0(Ω) in W1,G(Ω). The following version of Sobolev-Poincaré inequality can be found in [17, Lemma 7].
Theorem 2.3 (Sobolev-Poincaré inequality). Let G be an N-function with Δ2(G,G∗)<+∞. Then there exist numbers θ=θ(n,Δ2(G,G∗))∈(0,1) and K=K(n,N,Δ2(G,G∗))>0 such that the following holds. If B⊂Rn is any ball with radius R and w∈W1,G(B,RN), then
−∫BG(|w−(w)B|R)dx≤K(−∫BGθ(|Dw|)dx)1θ, |
where (w)B:=−∫Bw(x)dx. Moreover, if w∈W1,G0(B,RN), then
−∫BG(|w|R)dx≤K(−∫BGθ(|Dw|)dx)1θ, |
where K and θ have the same dependencies as before.
We conclude this section with the following useful lemma about an almost concave condition.
Lemma 2.4. Let Ψ:[0,∞)→[0,∞) be non-decreasing and such that t→Ψ(t)t be non-increasing. Then there exists a concave function ˜Ψ:[0,∞)→[0,∞) such that
12˜Ψ(t)≤Ψ(t)≤˜Ψ(t)for all t≥0. |
Proof. See [36, Lemma 2.2].
The following lemma is a key tool to deal with the constraint of divergence free vector fields, as we have to construct testing functions in divergence free form.
Lemma 2.5. Let Br(x0) be a ball in Rn and f∈Lγ(Br(x0)) with (f)x0,r=0, where 1<γ1≤γ≤γ2<+∞. Then there exists w∈W1,γ0(Br(x0),Rn), weak solution to divw=f in Br(x0), such that
∫Br(x0)|Dw|tdx≤c(n,γ1,γ2)∫Br(x0)|f|tdx | (2.13) |
for every t∈[γ1,γ].
Proof. See [8] and [24, Chapter 3,Section 3].
Corollary 1. Let Br(x0) be a ball in Rn, and G be an N-function such that Δ2(G)<+∞ and Δ2(G∗)<+∞. Let G(|f|)∈L1(Br(x0)) with (f)x0,r=0. Then there exists w∈W1,G0(Br(x0),Rn), solution to divw=f a.e. in Br(x0), such that
∫Br(x0)G(|Dw|)dx≤cG∫Br(x0)G(|f|)dx. | (2.14) |
Proof. See, e.g., [29, Corollary 4.2].
We define the space of traceless affine functions
T(Rn):={ℓ:Rn→Rn: ℓ is affine and tr(Dℓ)=0}. |
In particular, each ℓ∈T satisfies divℓ=0.
The set of rigid motions in Rn is defined as
R:={c+Sx:c∈Rn,S∈Rn×n,TS=−S}, |
the set of affine functions with skew-symmetric gradient. Let x0∈Ω and r>0. We define the affine function
ℓx0,r(x):=(u)x0,r+(Du)x0,r(x−x0). | (2.15) |
Note that ℓx0,r∈T(Rn) and
(u−ℓx0,r)x0,r=0,Eℓx0,r=(Eu)x0,r,(W(u−ℓx0,r))x0,r=0. |
Let Ω be a bounded open subset of Rn and let x0 be its centroid; i.e.,
x0:=1|Ω|∫Ωxdx. |
For every u∈L1(Ω;Rn) and x∈Ω we set
(PΩu)i(x):=ci+Sij(x−x0)j, | (2.16) |
where
ci:=(u)i=−∫Ωui(x)dx,Sij:=−∫Ω{[u−(u)]i(x−x0)j−[u−(u)]j(x−x0)i}dx−∫Ω[|(x−x0)i|2+|(x−x0)j|2]dx. |
As shown in [2, Proposition 2.6], PΩu∈R for all u; this implies, in particular, that E(PΩu)=0. Moreover, if u∈L2(Ω;Rn) and Ω is a ball, then
‖u−PΩu‖L2(Ω;Rn)=min{‖u−r‖L2(Ω;Rn):r∈R}. |
Finally, we recall the following Korn-type inequality [2, Eq (2.21)]:
1t(−∫Br|u−PBrur|tdx)1t≤c(−∫Br|Eu|pdx)1p | (2.17) |
for every t∈[1,p∗].
A standard tool in order to obtain local bounds of Du in terms of Eu on the scale of Lp spaces is the following Sobolev-Korn inequality (see, e.g., [35]).
Lemma 2.6. Let 1<p≤r≤q and u∈Lp(Bϱ(x0),Rn) be such that Eu∈Lr(Bϱ(x0),Rn×nsym). Then Du∈Lr(Bϱ(x0),Rn2) and for some constant c=c(n,p,q) it holds that
−∫Bϱ(x0)|Du|rdx≤c−∫Bϱ(x0)|Eu|rdx+c(−∫Bϱ(x0)|u−(u)x0,ϱϱ|dx)r. | (2.18) |
If, in addition, u=0 on ∂Bϱ(x0), then
−∫Bϱ(x0)|Du|rdx≤c−∫Bϱ(x0)|Eu|rdx. | (2.19) |
We also need the following version of the Korn's inequality in Orlicz spaces (see [18, Lemma 3.3] and [9]).
Lemma 2.7. Let B⊂Rn be a ball. Let G be an N-function such that both G and G∗ satisfy the Δ2-condition. Then for all u∈W1,G0(B,Rn) (with (Wu)B=0) the inequality
∫BG(|Du|)dx≤cKorn∫BG(|Eu|)dx |
holds, and cKorn=cKorn(Δ2(G,G∗)).
The following lemma is useful to derive reverse Hölder estimates. It is a variant of the results by Gehring [25] and Giaquinta-Modica [26, Theorem 6.6].
Lemma 2.8. Let B0⊂Rn be a ball, f∈L1(B0), and g∈Lσ0(B0) for some σ0>1. Assume that for some θ∈(0,1), c1>0 and all balls B with 2B⊂B0
−∫B|f|dx≤c1(−∫2B|f|θdx)1/θ+−∫2B|g|dx. |
Then there exist σ1>1 and c2>1 such that g∈Lσ1loc(B) and for all σ2∈[1,σ1]
(−∫B|f|σ2dx)1/σ2≤c2−∫2B|f|dx+c2(−∫2B|g|σ2dx)1/σ2. |
Let A be a bilinear form on Rn×nsym. We say that A is strongly elliptic in the sense of Legendre-Hadamard if for all η,ζ∈Rn×nsym it holds that
κA|η||ζ|≤A(η,ζ)≤LA|η||ζ| | (2.20) |
for some LA≥κA>0. The biggest possible constant κA is called the ellipticity constant of A. By |A| we denote the Euclidean norm of A. We say that a Sobolev function w on a ball Bϱ(x0) is A-Stokes on Bϱ(x0) if it satisfies −div(AEw)=0 in the sense of distributions; i.e.,
∫Bϱ(x0)A(Ew,Eψ)dx=0, for all ψ∈C∞0,div(Bϱ(x0),Rn). |
Now, we address the problem of the A-Stokes approximation: given a Sobolev function v on a ball B, we want to find an A-Stokes function h which is "close" the function v. It will be the A-Stokes function with the same boundary values as v; i.e., a Sobolev function h which satisfies
{−div(AEh)=0on Bh=von ∂B | (2.21) |
in the sense of distributions. Setting w:=h−v, then (2.21) is equivalent to finding a Sobolev function w which satisfies
{−div(AEw)=−div(AEv)on Bw=0on ∂B | (2.22) |
in the sense of distributions. We formulate the A-Stokes approximation result for Stokes systems with general growth. The proof for Stokes systems with p-growth can be found in [10, Theorem 4.2], whence the case of general growth can be inferred by using a duality argument as in [20], based on the following Lemma.
Lemma 2.9. Let G be an N-function with Δ2(G,G∗)<∞. Let B⊂Rn be a ball, and let A be strongly elliptic in the sense of Legendre-Hadarmard. Then the following variational estimates hold for all v∈W1,G0,dive(B)
∫BG(|Ev|)dx≲supξ∈(C∞0.dive(B))N[∫BA(Ev,Eξ)dx−∫BG∗(|Dξ|)dx]. | (2.23) |
The implicit constants in (2.23) only depend on n, κA, ||A|| and Δ2(G,G∗).
Proof. This result can be deduced from [10, Lemma 4.1] along the lines of the proof of [20, Lemma 20]. We then omit the details.
The A-Stokes approximation can be stated as follows.
Theorem 2.10. Let B⊂Rn be a ball with radius rB and let ˜B⊂Rn denote either B or 2B. Let G be an N-function with Δ2(G,G∗)<∞ and let s>1. Then for every κ>0, there exists δ>0 only depending on n, κA, |A|, Δ2(G,G∗) and s such that the following holds. If v∈W1,Gdive(˜B) is almost A-Stokes on B in the sense that
|−∫BA(Ev,Eξ)dx|≤δ−∫˜B|Ev|dx||Dξ||∞ | (2.24) |
for all ξ∈C∞0,dive(B). Then the unique solution w∈(W1,Gs0,dive(B))n of (2.22) satisfies
−∫BG(|w|rB)dx+−∫BG(|Dw|)dx≤κ(−∫˜B(G(|Dv|))sdx)1s. | (2.25) |
It holds κ=κ(G,s,δ) and limδ→0κ(G,s,δ)=0. The function h=v−w is called the A-Stokes approximation of v.
Remark 1. We will exploit the previous approximation result in a slightly modified version. Indeed, under the additional assumption (Wv)˜B=0 and
−∫˜BG(|Ev|)dx≤(−∫˜B[G(|Ev|)]sdx)1s≤G(ϵ) |
for some exponent s>1 and for a constant ϵ>0, and (2.24) replaced by
|−∫BA(Ev,Eξ)dx|≤δϵ||Dξ||L∞(B), |
by using Korn's inequality (Lemma 2.7) and following [12, Lemma 2.7], it can be seen that the unique solution w∈(W1,Gs0,dive(B))n of (2.22) satisfies
−∫BG(|w|rB)dx+−∫BG(|Dw|)dx≤κG(ϵ). | (2.26) |
We prove the self improving properties for double phase Stokes systems (1.1) with lower order terms. These are some higher integrability results which will play a crucial role in the sequel. They can be compared with analogous results obtained for Stokes systems with p(⋅)-growth, see [2, Theorem 4.2] and [7, Lemma 2.10]. In this direction, our main difficulty is in handling the lower order terms without any "closeness" condition between the extreme exponents p and q, which, on the contrary, can be ensured in the variable exponent setting by requiring some continuity assumption on p(⋅). For instance, in the main estimate (2.28), we are forced to control the average of the term rp2|Du|p instead of the more natural rp|Du|p appearing in [7, Lemma 2.10]. Nevertheless, these issues can be overcome when dealing with the terms involving H(x,t), just playing with the different behavior of H according to the magnitude of μ inside a small ball.
Lemma 2.11. Let H:Ω×[0,∞)→[0,∞) be defined as in (1.2), and satisfying (1.3). Let a:Ω×Rn×nsym→Rn×nsym be such that
|a(x,ξ)|≤LH′(x,1+|ξ|)anda(x,ξ):ξ≥νH(x,|ξ|)−ν0H(x,1) | (2.27) |
for all x∈Ω and ξ∈Rn×nsym, and for some 0<ν≤L<∞, ν0∈[0,1]. Let u∈W1,1(Ω;Rn) with H(⋅,|Eu|)∈L1(Ω) be a weak solution to (1.1). Then there exists s0>1, depending on n,p,q,[μ]Cα,ν,L,ν0,‖Eu‖Lp(Ω), such that H(⋅,|Eu|)∈Ls0loc(Ω) and, for every s∈(1,s0] and every B2r(x0)⊂⊂Ω, r<1,
(−∫Br(x0)[H(x,1+|Eu|)]sdx)1s≤c−∫B2r(x0)H(x,1+|Eu|)dx+c−∫B2r(x0)(rp2|Du|p+|u|p∗+1)dx | (2.28) |
for some c=c(n,p,q,[μ]Cα,ν,L,ν0,‖Eu‖Lp(Ω))>0.
Moreover, for each t∈[0,1] we have
(−∫Br(x0)[H(x,1+|Eu|)]sdx)1s≤ct(−∫B2r(x0)[H(x,1+|Eu|)]tdx)1t+c−∫B2r(x0)(rp2|Du|p+|u|p∗+1)dx | (2.29) |
for some constant ct=ct(n,p,q,[μ]Cα,ν,L,ν0,‖Eu‖Lp(Ω),t)>0.
Proof. Our argument partly follows that of [36, Theorem 3.4], devised for homogeneous double-phase systems depending on the gradient Du, while the estimates for the right hand side's terms in (1.10) are essentially taken from Acerbi and Mingione [2, Theorem 4.2].
Let η∈C∞0(B2r(x0)) be a cut-off function such that η≡1 on Br(x0), 0≤η≤1 and |Dη|≤c2r, and let P:=PB2r(x0) be the rigid displacement introduced in (2.16). We consider
ψ:=ηq(u−P)+w, |
where the function w is defined through Lemma 2.5 as a solution to
divw=−div(ηq(u−P))=−qηq−1Dη⋅(u−P). |
Such a w exists since divu=0 and
∫B2r(x0)(u−P)D(ηq)dx=0, |
and by (1.3) and the summability properties of u it holds that w∈W1,q0(B2r(x0);Rn). We recall also that, by (2.13), we have the estimate
−∫B2r(x0)|Dw|tdx≤c−∫B2r(x0)|u−P2r|tdx | (2.30) |
for every exponent t for which the right hand side is finite. Taking ψ as a test function in (1.1) we get
J1:=−∫B2rηqa(x,Eu):Eudx=−q−∫B2rηq−1a(x,Eu):(Dη⊙(u−P))dx−−∫B2ra(x,Eu):Ewdx+−∫B2rηqu[Du](u−P)dx+−∫B2ru[Du]wdx+−∫B2rηqf(u−P)dx+−∫B2rfwdx=:J2+J3+J4+J5+J6+J7. |
By the second condition in (2.27) we have
J1≥ν−∫B2rηqH(x,|Eu|)dx−ν0−∫B2rηqH(x,1)dx, |
while from the first one we get
J2≤c−∫B2rηqH′(x,1+|Eu|)|u−P|2rdx. |
Now, by using Young's inequality, for any κ∈(0,1) we obtain
−∫B2rηq−1|Eu|p−1|u−P|2rdx≤κ−∫B2rηq|Eu|pdx+cκ−∫B2r(|u−P|2r)pdx,−∫B2rηq−1μ(x)|Eu|q−1|u−P|2rdx≤κ−∫B2rηqμ(x)|Eu|qdx+cκ−∫B2rμ(x)(|u−P|2r)qdx, |
whence
−∫B2rηqH′(x,1+|Eu|)|u−P|2rdx≤κ−∫B2rηqH(x,1+|Eu|)dx+cκ−∫B2rH(x,|u−P|2r)dx. |
As for J3, using again the first condition in (2.27) and Young's inequality, and Corollary 1, we have
J3≤c−∫B2rH′(x,1+|Eu|)|Ew|dx≤κ−∫B2rH(x,1+|Eu|)dx+cκ−∫B2rH(x,|u−P|2r)dx. |
We can write
J4+J6≤−∫B2r|u||Du||u−P|dx+−∫B2r|f||u−P|dxJ5+J7≤−∫B2r|u||Du||w|dx+−∫B2r|f||w|dx. |
Let 0<β<1n+2. We set
γ:=1+β2(n+2n),σ:=[12(pγ)∗]′, |
and note that
1≤pγ<n,1<σ<σγ≤p,(pγ)∗≤p∗γ, | (2.31) |
since
2β+3nn+2≤p<n<nγ. |
By using Hölder's inequality, the Poincaré inequality for u−P and (2.31) we have
−∫B2r|u||Du||ηq(u−P)|dx≤2r(−∫B2r|u|(pγ)∗dx)1(pγ)∗(−∫B2r|Du|σdx)1σ(−∫B2r|u−P2r|(pγ)∗dx)1(pγ)∗≤cr(−∫B2r|u|(pγ)∗dx)1(pγ)∗(−∫B2r|Du|σdx)1σ(−∫B2r|Du|pγdx)γp≤cr(−∫B2r|u|p∗γdx)γp∗(−∫B2r|Du|pγdx)2γp |
whence, with Young's inequality,
−∫B2r|u||Du||ηq(u−P)|dx≤c−∫B2r|u|p∗γdx+crnpnp−nγ+pγ(−∫B2r|Du|pγdx)2nγnp−nγ+pγ≤c(−∫B2r|u|p∗γdx+rp2γ−∫B2r|Du|pγdx+1). |
Again using Hölder's and Young's inequalities, (2.17) and the fact that (2.31) implies 1∗≤(pγ)∗, we obtain
−∫B2r|f||u−P|dx≤c(−∫B2r|f|ndx)1n(−∫B2r|u−P|nn−1dx)n−1n≤c‖f‖Ln(Ω)(−∫B2r|u−P2r|p∗dx)1p∗≤c‖f‖Ln(Ω)(−∫B2r|Eu|pdx)1p≤κ−∫B2r|Eu|pdx+cκ≤κ−∫B2rH(x,|Eu|)dx+cκ |
for every κ∈(0,1). Using Hölder's inequality, recalling that w∈W1,q0(B2r;Rn), using Poincaré inequality, (2.30) and then Young's inequality we obtain the estimate
−∫B2r|u||Du||w|dx≤(−∫B2r|u|(pγ)∗dx)1(pγ)∗(−∫B2r|Du|σdx)1σ(−∫B2r|w|(pγ)∗dx)1(pγ)∗≤cr(−∫B2r|u|(pγ)∗dx)1(pγ)∗(−∫B2r|Du|σdx)1σ(−∫B2r|Dw|pγdx)γp≤cr(−∫B2r|u|(pγ)∗dx)1(pγ)∗(−∫B2r|Du|σdx)1σ(−∫B2r|u−P2r|(pγ)∗dx)1(pγ)∗≤cr(−∫B2r|u|p∗γdx)γp∗(−∫B2r|Du|pγdx)2γp≤c(−∫B2r|u|p∗γdx+rp2γ−∫B2r|Du|pγdx+1). |
Arguing as above and with Korn's inequality we get
−∫B2r|f||w|dx≤c(−∫B2r|f|ndx)1n(−∫B2r|w|p∗dx)1p∗≤cr‖f‖Ln(Ω)(−∫B2r|Dw|pdx)1p≤c‖f‖Ln(Ω)(−∫B2r|u−P2r|pdx)1p≤c‖f‖Ln(Ω)(−∫B2r|Eu|pdx)1p≤κ−∫B2r|Eu|pdx+cκ≤κ−∫B2rH(x,|Eu|)dx+cκ, |
where in the last inequality we applied Young's inequality for some κ∈(0,1). Collecting the previous estimates and taking into account (2.31), we finally have
−∫BrH(x,1+|Eu|)dx≤κ−∫B2rH(x,1+|Eu|)dx+cκ−∫B2rH(x,|u−P|2r)dx+cκ−∫B2r(x0)(rp2γ|Du|pγ+|u|p∗γ+1)dx, | (2.32) |
for every κ∈(0,1). Note that setting
g:=rp2γ|Du|pγ+|u|p∗γ+1 |
we have g∈Lγ(B2r), since by Korn's inequality and the Sobolev embedding theorem it holds that |Du|p+|u|p∗∈L1(Ω). Now we claim that there exists θ=θ(n,p,q)∈(0,1) such that
−∫B2rH(x,|u−P|2r)dx≤c(−∫B2r[H(x,|Eu|)]θdx)1θ, | (2.33) |
where c=c(q,p,μ,‖Eu‖Lp(Ω))>0. We first note that, setting q∗:=nqn+q, we have q∗<p by (1.3). Then, by Korn's inequality it holds that
−∫B2r|u−P2r|qdx≤c(−∫B2r|Eu|q∗dx)qq∗. | (2.34) |
We distinguish between two cases. If supB2rμ(⋅)≤4[μ]Cαrα, using (2.34) and Hölder's inequality we then find
−∫B2rμ(x)|u−P2r|qdx≤c4[μ]Cαrα(−∫B2r|Eu|q∗dx)qq∗≤c4[μ]Cαrα(−∫B2r|Eu|pdx)q−pp(−∫B2r|Eu|q∗dx)pq∗≤c[μ]Cα‖Eu‖q−pLp(Ω)rα−n(q−p)p(−∫B2r|Eu|q∗dx)pq∗, |
whence (2.33) follows by choosing θ:=q∗p<1 and noting that, by Korn's inequality,
−∫B2r|u−P2r|pdx≤c(−∫B2r|Eu|p∗dx)pp∗≤c(−∫B2r|Eu|pθdx)1θ, |
where in the latter we used p∗≤q∗. If instead infB2rμ(⋅)>3[μ]Cαrα, then we can freeze μ at x=x0 and apply Theorem 2.3 and Lemma 2.7 with G(t):=H(x0,t), since in this case 34H(x,t)≤H(x0,t)≤43H(x,t) for every x∈B2r(x0). This concludes the proof of the claim.
Now, combining (2.32) and (2.33) we can write
−∫BrH(x,1+|Eu|)dx≤κ−∫B2rH(x,1+|Eu|)dx+cκ(−∫B2r[H(x,1+|Eu|)]θdx)1θ+cκ−∫B2r(x0)gdx. |
This estimate holds for every κ∈(0,1) and every ball B2r⊂⊂Ω, and the constants cκ only depend on the data. Thus, by virtue of a variant of Gehring's lemma (Lemma 2.8) we get (2.28). This concludes the proof.
We define
ε0:=ε0(n,p,q,[μ]Cα,ν,L,‖Eu‖Lp(Ω)):=min{α2,n(q−p)(s0−1)2ps0}, | (2.35) |
where s0 is that of Lemma 2.11. We now state an higher integrability estimate near the p-phase.
Lemma 2.12. Let ε0 be as in (2.35). For B2r(x0)⊂⊂Ω with r≤1, if
infB2rμ(⋅)≤[μ]Cα(2r)α−ε0, | (2.36) |
then
(−∫Br(x0)[H(x,1+|Eu|)]s0dx)1s0≤c(1+2αrε0)−∫B2r(x0)(1+|Eu|)pdx+c−∫B2r(x0)(rp2|Du|p+|u|p∗+1)dx | (2.37) |
for some c=c(n,p,q,[μ]Cα,ν,L,ν0,‖Eu‖Lp(Ω),‖1+|Eu|‖Lps0(B2r(x0))).
If, in addition,
rp2(|Du|p)x0,2r+|(u)x0,2r|≤M, | (2.38) |
then
(−∫Br(x0)[H(x,1+|Eu|)]s0dx)1s0≤c(1+2αrε0)−∫B2r(x0)(1+|Eu|)pdx. | (2.39) |
Proof. The argument is similar to that of [37, Lemma 3.6]. Thus, we only sketch the proof.
We first note that under assumption (2.36), for every x\in B_{2r}(x_0) we have \mu(x)\leq 3[\mu]_{C^\alpha}r^{\alpha-\varepsilon_0} . This fact together with (2.29), applied with t = \frac{t_0}{2}\in(0, 1) where
\begin{equation} t_0: = \min\left\{\frac{2p}{q}, \frac{p}{q-p}\right\}\, , \end{equation} | (2.40) |
implies
\begin{equation} \begin{split} \mathop {{\rlap{-} \smallint }}\limits_{B_r(x_0)} \mu(x)(1+| \mathcal{E}{\bf u}|)^q \, \mathrm{d}x & \leq c \left(\mathop {{\rlap{-} \smallint }}\limits_{B_{2r}(x_0)}(1+| \mathcal{E}{\bf u}|)^\frac{pt_0}{2}\right)^\frac{2}{t_0} + c r^{\alpha-\varepsilon_0} \left(\mathop {{\rlap{-} \smallint }}\limits_{B_{2r}(x_0)}(1+| \mathcal{E}{\bf u}|)^\frac{qt_0}{2}\, \mathrm{d}x\right)^\frac{2}{t_0} \\ & \, \, \, \, + c\mathop {{\rlap{-} \smallint }}\limits_{B_{2r}(x_0)}(r^\frac{p}{2}|D{\bf u}|^p+|{\bf u}|^{p^*}+1) \, \mathrm{d}x \, . \end{split} \end{equation} | (2.41) |
In (2.41), the second integral is finite since (qt_0)/2\leq p by (2.40). Moreover, with Hölder's inequality and the fact that (q-p)t_0\leq p again by (2.40), we have
\begin{equation} \left(\mathop {{\rlap{-} \smallint }}\limits_{B_{2r}(x_0)}(1+| \mathcal{E}{\bf u}|)^\frac{qt_0}{2}\right)^\frac{2}{t_0} \leq c r^{-\frac{(q-p)n}{ps_0}} \|1+| \mathcal{E}{\bf u}|\|_{L^{ps_0}(B_{2r}(x_0))}^{q-p} \mathop {{\rlap{-} \smallint }}\limits_{B_{2r}(x_0)}(1+| \mathcal{E}{\bf u}|)^p\, \mathrm{d}x \, . \end{equation} | (2.42) |
Therefore, plugging (2.41) into (2.42) and using r\leq1 , (1.3) and the definition of \varepsilon_0 , we get
\begin{equation*} \begin{split} \mathop {{\rlap{-} \smallint }}\limits_{B_r(x_0)} \mu(x)(1+| \mathcal{E}{\bf u}|)^q \, \mathrm{d}x & \leq c \left(1+ 2^\alpha r^{\alpha-\varepsilon_0-\frac{(q-p)n}{p}-\frac{(q-p)n(s_0-1)}{ps_0}}\right)\mathop {{\rlap{-} \smallint }}\limits_{B_{2r}(x_0)}(1+| \mathcal{E}{\bf u}|)^p \\ & \, \, \, \, + c\mathop {{\rlap{-} \smallint }}\limits_{B_{2r}(x_0)}(r^\frac{p}{2}|D{\bf u}|^p+|{\bf u}|^{p^*}+1) \, \mathrm{d}x \\ & \leq c \left(1+ 2^\alpha r^{\varepsilon_0}\right)\mathop {{\rlap{-} \smallint }}\limits_{B_{2r}(x_0)}(1+| \mathcal{E}{\bf u}|)^p + c\mathop {{\rlap{-} \smallint }}\limits_{B_{2r}(x_0)}(r^\frac{p}{2}|D{\bf u}|^p+|{\bf u}|^{p^*}+1) \, \mathrm{d}x \, . \end{split} \label{eq:3.14oknla} \end{equation*} |
This concludes the proof of (2.37). In order to prove (2.39), we only have to check that under assumption (2.38), we have
\begin{equation} \mathop {{\rlap{-} \smallint }}\limits_{B_{2r}(x_0)}|{\bf u}|^{p^*} \, \mathrm{d}x \leq c \mathop {{\rlap{-} \smallint }}\limits_{B_{2r}(x_0)}(1+| \mathcal{E}{\bf u}|)^p \, \mathrm{d}x \, . \end{equation} | (2.43) |
Indeed, using the Poincaré inequality and the Sobolev-Korn inequality (2.18), we get
\begin{equation*} \begin{split} \mathop {{\rlap{-} \smallint }}\limits_{B_{2r}(x_0)}|{\bf u}|^{p^*} \, \mathrm{d}x & \leq c\mathop {{\rlap{-} \smallint }}\limits_{B_{2r}(x_0)}|{\bf u}-({\bf u})_{x_0, 2r}|^{p^*} \, \mathrm{d}x + c |({\bf u})_{x_0, 2r}|^{p^*} \\ & \leq c \left(r^{p}\mathop {{\rlap{-} \smallint }}\limits_{B_{2r}(x_0)}|D{\bf u}|^{p} \, \mathrm{d}x\right)^\frac{p^*}{p} + M^{p^*} \\ & \leq c \left( r^{p}\mathop {{\rlap{-} \smallint }}\limits_{B_{2r}(x_0)}| \mathcal{E}{\bf u}|^{p}\, \mathrm{d}x + \left (\mathop {{\rlap{-} \smallint }}\limits_{B_{2r}(x_0)}|{{\bf u} - ({\bf u})_{x_0, 2r}}| \, \mathrm{d}x\right)^p\right)^\frac{p^*}{p} + M^{p^*} \\ & \leq c\left( r^{p} \mathop {{\rlap{-} \smallint }}\limits_{B_{2r}(x_0)}| \mathcal{E}{\bf u}|^{p}\, \mathrm{d}x + 1 \right)^\frac{p^*}{p}+ M^{p^*}\\ & \leq c r^{p^*-n(\frac{p^*}{p}-1)} \|1+| \mathcal{E}{\bf u}|\|_{L^p(\Omega)}^{{p^*}-p} \mathop {{\rlap{-} \smallint }}\limits_{B_{2r}(x_0)}(1+| \mathcal{E}{\bf u}|^{p})\, \mathrm{d}x +M^{p^*}\\ & \leq c(n, p, \|1+| \mathcal{E}{\bf u}|\|_{L^p(\Omega)}, M) \mathop {{\rlap{-} \smallint }}\limits_{B_{2r}(x_0)}(1+| \mathcal{E}{\bf u}|^{p})\, \mathrm{d}x\, . \end{split} \end{equation*} |
The proof is concluded.
Let x_0\in\Omega , B_\varrho(x_0)\subset\Omega and {\mathit{\boldsymbol{\ell}}}_{x_0, \varrho} be the traceless affine function defined in (2.15). We introduce the following Campanato-type excess functionals, measuring the oscillations of \mathcal{E}{\bf u} :
\begin{equation} \begin{split} {\mathit{\Phi}}(x_0, \varrho): & = \mathop {{\rlap{-} \smallint }}\limits_{B_\varrho(x_0)} H_{1+|( \mathcal{E}{\bf u})_{x_0, \varrho}|} (x_0, | \mathcal{E}{\bf u}- \mathcal{E}{\mathit{\boldsymbol{\ell}}}_{x_0, \varrho}|)\, \mathrm{d}x \\ & = \mathop {{\rlap{-} \smallint }}\limits_{B_\varrho(x_0)} H_{1+|( \mathcal{E}{\bf u})_{x_0, \varrho}|} (x_0, | \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_{x_0, \varrho}|)\, \mathrm{d}x \, , \end{split} \end{equation} | (3.1) |
where H_{1+|(\mathcal{E}{\bf u})_{x_0, \varrho}|} (x_0, t) denotes the shifted N function H_a with shift a: = 1+|(\mathcal{E}{\bf u})_{x_0, \varrho}| .
The first key tool is the following "conditioned" Caccioppoli type estimate for {\bf u}-{\mathit{\boldsymbol{\ell}}}_{x_0, r} , which holds under suitable smallness assumptions, see (3.2) and (3.3) below.
Lemma 3.1. Let B_{\varrho}(x_0)\subset\subset\Omega , with \varrho\leq1 . Assume that
\begin{equation} \varrho^\frac{p}{2} (|D{\bf u}|^p)_{x_0, \varrho} + |({\bf u})_{x_0, \varrho}| \leq M\, , \end{equation} | (3.2) |
and
\begin{equation} {\mathit{\Phi}}(x_0, \varrho) \leq c H(x_0, 1+|( \mathcal{E}{\bf u})_{x_0, \varrho}|)\, . \end{equation} | (3.3) |
Then we have
\begin{equation} \begin{split} & \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho/2}(x_0)} H_{1+|( \mathcal{E}{\bf u})_{x_0, \varrho}|}(x_0, | \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_{x_0, \varrho}|)\, \mathrm{d}x \\ & \, \, \, \, \, \, \, \, \, \, \, \, \, \, \, \, \, \, \leq c \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}(x_0)} H_{1+|( \mathcal{E}{\bf u})_{x_0, \varrho}|}\left(x_0, \frac{|{\bf u}-{\mathit{\boldsymbol{\ell}}}_{x_0, \varrho}|}{\varrho}\right)\, \mathrm{d}x + c \varrho^{\bar{\gamma}}\, H(x_0, 1+|( \mathcal{E}{\bf u})_{x_0, \varrho}|)\, \end{split} \end{equation} | (3.4) |
for some constant c > 0 depending on data, and \bar{\gamma}: = \min\{\varepsilon_0, 1-\frac{1}{\beta+1}\} .
Proof. We follow the argument of [36, Lemma 4.1]. We use the shorthands {\mathit{\boldsymbol{\ell}}}_{\varrho} and (\mathcal{E}{\bf u})_{\varrho} for {\mathit{\boldsymbol{\ell}}}_{x_0, \varrho} and (\mathcal{E}{\bf u})_{x_0, \varrho} , respectively. We consider a cut-off function \eta\in C_0^\infty(B_{\varrho}; [0, 1]) such that \eta\equiv1 on B_{\varrho/2} and |D\eta|\leq \frac{c(n)}{\varrho} . Correspondingly, we define the function {\mathit{\boldsymbol{\psi}}}: = \eta^q({\bf u} - {\mathit{\boldsymbol{\ell}}}_{\varrho})\in W^{1, p}(B_{\varrho}; {\mathbb{R}}^n) . Since
\begin{equation*} {\rm div}\, {\mathit{\boldsymbol{\psi}}} = D(\eta^q)\cdot ({\bf u} - {\mathit{\boldsymbol{\ell}}}_{\varrho}) + \eta^q {\rm div}\, ({\bf u} - {\mathit{\boldsymbol{\ell}}}_{\varrho}) = q\eta^{q-1}D\eta\cdot ({\bf u} - {\mathit{\boldsymbol{\ell}}}_{\varrho})\, , \end{equation*} |
where we used that {\rm div}\, {\bf u} = 0 and {\mathit{\boldsymbol{\ell}}}_{\varrho}\in\mathcal{T}({\mathbb{R}}^n) , we conclude that {\mathit{\boldsymbol{\psi}}} is not a divergence-free vector field. By virtue of Lemma 2.5 we can find {{\bf w}}\in W^{1, q}_0(B_{\varrho}; {\mathbb{R}}^n) such that, setting
\begin{equation*} {\mathit{\boldsymbol{\varphi}}}: = {\mathit{\boldsymbol{\psi}}}-{{\bf w}} \, , \end{equation*} |
we have {\rm div}\, {\mathit{\boldsymbol{\varphi}}} = 0 . Taking {\mathit{\boldsymbol{\varphi}}} as a test function in (1.1) we get
\begin{equation} \begin{split} \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} \eta^q {\bf a}(x, \mathcal{E}{\bf u}): \mathcal{E}({\bf u} - {\mathit{\boldsymbol{\ell}}}_{\varrho})\, \mathrm{d}x & = -q\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}}\eta^{q-1} {\bf a}(x, \mathcal{E}{\bf u}):(D \eta\odot({\bf u} - {\mathit{\boldsymbol{\ell}}}_{\varrho})- \mathcal{E}{{\bf w}})\, \mathrm{d}x \\ & \, \, \, \, \, \, \, \, +\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} {\bf u}[D {\bf u}][\eta^q({\bf u} - {\mathit{\boldsymbol{\ell}}}_{\varrho})-{{\bf w}}]\, \mathrm{d}x \\ & \, \, \, \, \, \, \, \, + \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} f[\eta^q({\bf u} - {\mathit{\boldsymbol{\ell}}}_{\varrho})-{{\bf w}}]\, \mathrm{d}x\, . \end{split} \end{equation} | (3.5) |
This and the identity
\begin{equation} \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}}{\bf a}(x_0, ( \mathcal{E}{\bf u})_{\varrho}): \mathcal{E}{\mathit{\boldsymbol{\varphi}}}\, \mathrm{d}x = 0 \end{equation} | (3.6) |
imply that
\begin{equation} \begin{split} J_1:& = \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} \eta^q({\bf a}(x_0, \mathcal{E}{\bf u})-{\bf a}(x_0, ( \mathcal{E}{\bf u})_{\varrho})):( \mathcal{E}{\bf u} -( \mathcal{E}{\bf u})_{\varrho})\, \mathrm{d}x \\ & = \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}}({\bf a}(x_0, \mathcal{E}{\bf u})-{\bf a}(x, \mathcal{E}{\bf u})): \mathcal{E}{{\mathit{\boldsymbol{\varphi}}}}\, \mathrm{d}x \\ & \, \, \, \, \, \, - q\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}}\eta^{q-1}({\bf a}(x_0, \mathcal{E}{\bf u})-{\bf a}(x_0, ( \mathcal{E}{\bf u})_{\varrho})):(D\eta\odot({\bf u} - {\mathit{\boldsymbol{\ell}}}_{\varrho})- \mathcal{E}{{\bf w}})\, \mathrm{d}x \\ & \, \, \, \, \, \, +\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}}{\bf u}[D {\bf u}][\eta^q({\bf u} - {\mathit{\boldsymbol{\ell}}}_{\varrho})-{{\bf w}}]\, \mathrm{d}x + \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} f[\eta^q({\bf u} - {\mathit{\boldsymbol{\ell}}}_{\varrho})-{{\bf w}}]\, \mathrm{d}x \\ & = - \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}}({\bf a}(x_0, ( \mathcal{E}{\bf u})_{\varrho})-{\bf a}(x, ( \mathcal{E}{\bf u})_{\varrho})): \mathcal{E}{{\mathit{\boldsymbol{\varphi}}}}\, \mathrm{d}x \\ & \, \, \, \, \, \, - q \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}}\eta^{q-1}({\bf a}(x, \mathcal{E}{\bf u})-{\bf a}(x, ( \mathcal{E}{\bf u})_{\varrho})):(({\bf u} - {\mathit{\boldsymbol{\ell}}}_{\varrho})\odot D\eta- \mathcal{E}{{\bf w}})\, \mathrm{d}x \\ & \, \, \, \, \, \, +\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} {\bf u}[D {\bf u}][\eta^q({\bf u} - {\mathit{\boldsymbol{\ell}}}_{\varrho})-{{\bf w}}]\, \mathrm{d}x+ \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} f[\eta^q({\bf u} - {\mathit{\boldsymbol{\ell}}}_{\varrho})-{{\bf w}}]\, \mathrm{d}x \\ & = : J_2+J_3+J_4+ J_5\, . \end{split} \end{equation} | (3.7) |
Now, we proceed to estimate each term above separately. With (1.5) and (2.9) we get
\begin{equation} \begin{split} J_1 & \geq c \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} \eta^q H''(x_0, 1+| \mathcal{E}{\bf u}|+|( \mathcal{E}{\bf u})_{\varrho}|)| \mathcal{E}{\bf u} - ( \mathcal{E}{\bf u})_{\varrho}|^2\, \mathrm{d}x \\ & \geq c \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} \eta^q H_{1+|( \mathcal{E}{\bf u})_{\varrho}|}(x_0, | \mathcal{E}{\bf u} - ( \mathcal{E}{\bf u})_{\varrho}|)\, \mathrm{d}x\, . \end{split} \end{equation} | (3.8) |
To estimate J_3 we use the second inequality in (1.4), (2.13) with G(t) = H_{1+|(\mathcal{E}{\bf u})_{\varrho}|}(x_0, t) , Young's inequality for G, G^* , (2.5) and we get
\begin{equation} \begin{split} |J_3| & \leq c \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}}\left(\int_0^1|D_\xi {\bf a}(x_0, \sigma( \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_{\varrho}) + ( \mathcal{E}{\bf u})_{\varrho})|\, \mathrm{d}\sigma\right)| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_{\varrho}|\left(\frac{|{\bf u} - {\mathit{\boldsymbol{\ell}}}_{\varrho}|}{\varrho}+| \mathcal{E}{{\bf w}}|\right)\, \mathrm{d}x \\ & \leq c \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}}\frac{H(x_0, 1+|( \mathcal{E}{\bf u})_{\varrho}|+| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_{\varrho}|)}{(1+|( \mathcal{E}{\bf u})_{\varrho}|+| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_{\varrho}|)^2}| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_{\varrho}|\left(\frac{|{\bf u} - {\mathit{\boldsymbol{\ell}}}_{\varrho}|}{\varrho}+| \mathcal{E}{{\bf w}}|\right)\, \mathrm{d}x \\ & \leq c \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}}{H'_{1+|( \mathcal{E}{\bf u})_{\varrho}|}(x_0, | \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_{\varrho}|)}\left(\frac{|{\bf u} - {\mathit{\boldsymbol{\ell}}}_{\varrho}|}{\varrho}+| \mathcal{E}{{\bf w}}|\right)\, \mathrm{d}x \\ & \leq c \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}}\left(H^*_{1+|( \mathcal{E}{\bf u})_{\varrho}|}\left(x_0, {H'_{1+|( \mathcal{E}{\bf u})_{\varrho}|}(x_0, | \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_{\varrho}|)}\right)+ H_{1+|( \mathcal{E}{\bf u})_{\varrho}|}\left(x_0, \frac{|{\bf u} - {\mathit{\boldsymbol{\ell}}}_{\varrho}|}{\varrho}+| \mathcal{E}{{\bf w}}|\right)\right)\, \mathrm{d}x \\ & \leq c \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}}\left(H_{1+|( \mathcal{E}{\bf u})_{\varrho}|}(x_0, | \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_{\varrho}|)+H_{1+|( \mathcal{E}{\bf u})_{\varrho}|}\left(x_0, \frac{|{\bf u} - {\mathit{\boldsymbol{\ell}}}_{\varrho}|}{\varrho}\right)+H_{1+|( \mathcal{E}{\bf u})_{\varrho}|}(x_0, | \mathcal{E}{{\bf w}}|)\right)\, \mathrm{d}x \\ & \leq c \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}}H_{1+|( \mathcal{E}{\bf u})_{\varrho}|}(x_0, | \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_{\varrho}|)\, \mathrm{d}x+ c \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} H_{1+|( \mathcal{E}{\bf u})_{\varrho}|}\left(x_0, \frac{|{\bf u} - {\mathit{\boldsymbol{\ell}}}_{\varrho}|}{\varrho}\right)\, \mathrm{d}x\, . \end{split} \end{equation} | (3.9) |
We now treat separately the two cases where the inequality
\begin{equation} \inf\limits_{x\in B_{\varrho}} \mu(x) \leq [\mu]_{C^\alpha}\varrho^{\alpha-\varepsilon_0} \end{equation} | (3.10) |
holds true or not. Note that from (3.10), for every x\in B_{\varrho} we get
\begin{equation} \mu(x) \leq [\mu]_{C^\alpha} \varrho^\alpha + [\mu]_{C^\alpha}\varrho^{\alpha-\varepsilon_0} \leq 2[\mu]_{C^\alpha} \varrho^{\alpha-\varepsilon_0}\, . \end{equation} | (3.11) |
Step 1: p -phase. Let \mu comply with (3.10). Then, using Hölder's inequality, (1.3), the definition of \epsilon_0 and Lemma 2.12 we get
\begin{equation} \begin{split} \varrho^{\alpha-\epsilon_0}(1+|( \mathcal{E}{\bf u})_{\varrho}|)^{q-p} & \leq \varrho^{\alpha-\epsilon_0} \left(\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}}[1+| \mathcal{E}{\bf u}|]^{ps_0}\, \mathrm{d}x\right)^{\frac{q-p}{ps_0}} \\ & \leq \varrho^{\alpha-\frac{(q-p)n}{p}+\varepsilon_0} \|1+| \mathcal{E}{\bf u}|\|_{L^{ps_0}(B_{\varrho})}^{q-p} \leq c \varrho^{\varepsilon_0}\, . \end{split} \end{equation} | (3.12) |
From this and (3.11) we deduce that
\begin{equation} \begin{split} H(x_0, 1+|( \mathcal{E}{\bf u})_{\varrho}|) & = (1+\mu(x_0)(1+|( \mathcal{E}{\bf u})_{\varrho}|)^{q-p})(1+|( \mathcal{E}{\bf u})_{\varrho}|)^p \\ & \leq c(1+\varrho^{\alpha-\varepsilon_0}(1+|( \mathcal{E}{\bf u})_{\varrho}|)^{q-p})(1+|( \mathcal{E}{\bf u})_{\varrho}|)^p \\ & \leq c(1+\varrho^{\varepsilon_0})(1+|( \mathcal{E}{\bf u})_{\varrho}|)^p \leq c(1+|( \mathcal{E}{\bf u})_{\varrho}|)^p\, . \end{split} \end{equation} | (3.13) |
As for J_2 , from (1.6), (3.12), (3.13), and using Young's inequality (2.4) for H_{1+|(\mathcal{E}{\bf u})_{\varrho}|}(x_0, t) and its conjugate, (2.11), and (2.14),
we obtain
\begin{equation} \begin{split} |J_2| & \leq \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} c\varrho^\alpha (1+|( \mathcal{E}{\bf u})_{\varrho}|)^{q-1}| \mathcal{E}{{\mathit{\boldsymbol{\varphi}}}}|\, \mathrm{d}x \\ & \leq c \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}}\varrho^{\alpha-\epsilon_0}(1+|( \mathcal{E}{\bf u})_{\varrho}|)^{q-p}(1+|( \mathcal{E}{\bf u})_{\varrho}|)^{p-1}| \mathcal{E}{{\mathit{\boldsymbol{\varphi}}}}|\, \mathrm{d}x \\ & \leq c \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}}\varrho^{\varepsilon_0} H'(x_0, 1+|( \mathcal{E}{\bf u})_{\varrho}|)| \mathcal{E}{{\mathit{\boldsymbol{\varphi}}}}|\, \mathrm{d}x \\ & \leq \frac{1}{4}\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}}\eta^q H_{1+|( \mathcal{E}{\bf u})_{\varrho}|}(x_0, | \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_{\varrho}|)\, \mathrm{d}x + c \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}}H_{1+|( \mathcal{E}{\bf u})_{\varrho}|}\left(x_0, \frac{|{\bf u}-{{\mathit{\boldsymbol{\ell}}}}_{\varrho}|}{\varrho}\right)\, \mathrm{d}x \\ &\, \, \, \, \, \, + c \varrho^{\varepsilon_0} H(x_0, 1+|( \mathcal{E}{\bf u})_{\varrho}|)\, . \end{split} \end{equation} | (3.14) |
Plugging the estimates (3.8), (3.14) and (3.9) into (3.7) and reabsorbing some terms we obtain (3.4).
Step 2: (p, q) -phase. If (3.10) does not hold, then \mu(x_0)\geq\inf_{x\in B_{\varrho}} \mu(x) > [\mu]_{C^\alpha}\varrho^{\alpha-\varepsilon_0} , so that
\begin{equation} \varrho^\alpha \leq c \varrho^{\varepsilon_0}\mu(x_0)\, . \end{equation} | (3.15) |
Now, arguing as for Step 1 and using Young's inequality (2.4) for H_{1+|(\mathcal{E}{\bf u})_{\varrho}|}(x_0, t) and its conjugate, and (2.11), we have
\begin{equation*} \begin{split} |J_2| & \leq c \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}}\varrho^{\varepsilon_0}H'(x_0, 1+|( \mathcal{E}{\bf u})_{\varrho}|)| \mathcal{E}{{\mathit{\boldsymbol{\varphi}}}}|\, \mathrm{d}x \\ & \leq \frac{1}{4}\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}}\eta^q H_{1+|( \mathcal{E}{\bf u})_{\varrho}|}(x_0, | \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_{\varrho}|)\, \mathrm{d}x + c \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}}H_{1+|( \mathcal{E}{\bf u})_{\varrho}|}\left(x_0, \frac{|{\bf u}-{{\mathit{\boldsymbol{\ell}}}}_{\varrho}|}{\varrho}\right)\, \mathrm{d}x \\ &\, \, \, \, \, \, + c \varrho^{\varepsilon_0}H(x_0, 1+|( \mathcal{E}{\bf u})_{\varrho}|)\, . \end{split} \end{equation*} |
Now, we set \gamma: = \left(\frac{p^*}{2}\right)' . Note that if p > \frac{3n}{n+2} , then \gamma < p . Using Hölder's inequality we have
\begin{equation*} \begin{split} |J_4| & \leq \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} |{\bf u}| |D{\bf u}||\eta^q({\bf u} - {\mathit{\boldsymbol{\ell}}}_{\varrho})-{{\bf w}}|\, \mathrm{d}x \\ & \leq \left(\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} |{\bf u}|^{p^*}\, \mathrm{d}x\right)^\frac{1}{p^*}\left(\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} |D{\bf u}|^{\gamma}\, \mathrm{d}x\right)^\frac{1}{\gamma}\left(\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} (|{\bf u} - {\mathit{\boldsymbol{\ell}}}_\varrho|^{p^*}+|{{\bf w}}|^{p^*})\, \mathrm{d}x\right)^\frac{1}{p^*} \\ & = : I_1\cdot I_2\cdot I_3 \, . \end{split} \end{equation*} |
Estimate of I_1 : using Poincaré inequality and (3.2), we get
\begin{equation*} |I_1| \leq \left(\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} |{\bf u}-({\bf u})_\varrho|^{p^*}\, \mathrm{d}x\right)^\frac{1}{p^*} + |({\bf u})_\varrho| \leq \varrho(|D{\bf u}|^p)_\varrho + M \leq 2M\, . \end{equation*} |
Estimate of I_2 : with the Sobolev-Korn's inequality (2.18) and (3.2) we obtain
\begin{equation*} \begin{split} |I_2| & \leq c \left[\left(\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} | \mathcal{E}{\bf u}|^{\gamma}\, \mathrm{d}x\right)^\frac{1}{\gamma} + \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} \left|\frac{{\bf u}-({\bf u})_\varrho}{\varrho}\right|\, \mathrm{d}x\right] \\ & \leq c \left[\left(\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} | \mathcal{E}{\bf u}|^{\gamma}\, \mathrm{d}x\right)^\frac{1}{\gamma} + C_M \right]\, . \end{split} \end{equation*} |
Now, using Lemma 2.1 and (3.3),
\begin{equation*} \begin{split} &\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} H(x_0, 1+| \mathcal{E}{\bf u}|)\, \mathrm{d}x \\ & \leq cc_\eta \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} H_{1+|( \mathcal{E}{\bf u})_\varrho|}(x_0, | \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|)\, \mathrm{d}x + (\eta+c) H(x_0, 1+|( \mathcal{E}{\bf u})_\varrho|) \\ & \leq c H(x_0, 1+|( \mathcal{E}{\bf u})_\varrho|)\, , \end{split} \end{equation*} |
whence, from Jensen's inequality for the function \Psi(t): = t^\frac{p}{\gamma}+\mu(x_0)t^\frac{q}{\gamma} (convex, since \gamma < p\leq q ) we obtain
\begin{equation} \begin{split} \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} (1+| \mathcal{E}{\bf u}|)^{\gamma}\, \mathrm{d}x & \leq \Psi^{-1}\left(\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} H(x_0, 1+| \mathcal{E}{\bf u}|)\, \mathrm{d}x\right) \\ & \leq c \Psi^{-1}\left(H(x_0, 1+|( \mathcal{E}{\bf u})_\varrho|)\right) \\ & \leq c(1+|( \mathcal{E}{\bf u})_\varrho|)^\gamma\, . \end{split} \end{equation} | (3.16) |
Thus,
\begin{equation*} \begin{split} |I_2| \leq c \left[1+|( \mathcal{E}{\bf u})_\varrho|+ C_M \right]\, . \end{split} \end{equation*} |
Estimate of I_3 : we use Poincaré's inequality, Lemma 2.5 and Lemma 2.7, and we get
\begin{equation*} \begin{split} |I_3| \leq \left(\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} (|{\bf u} - {\mathit{\boldsymbol{\ell}}}_\varrho|^{p^*}+\varrho^{p^*}|D {{\bf w}}|^{p})\, \mathrm{d}x\right)^\frac{1}{p^*} & \leq c \left(\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} |{\bf u} - {\mathit{\boldsymbol{\ell}}}_\varrho|^{p^*}\, \mathrm{d}x\right)^\frac{1}{p^*} \\ & \leq c \varrho\left(\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} |D{\bf u} - (D{\bf u})_\varrho|^{p}\, \mathrm{d}x\right)^\frac{1}{p}\\ & \leq c c_{\rm Korn} \varrho\left(\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} | \mathcal{E}{\bf u} - ( \mathcal{E}{\bf u})_\varrho|^{p}\, \mathrm{d}x\right)^\frac{1}{p}\, . \end{split} \end{equation*} |
Now, using the change of shift formula, (3.3) and an analogous argument as for (3.16) with the convex function \widetilde{\Psi}(t): = t +\mu(x_0)t^\frac{q}{p} , we obtain
\begin{equation*} \left(\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} | \mathcal{E}{\bf u} - ( \mathcal{E}{\bf u})_\varrho|^{p}\, \mathrm{d}x\right)^\frac{1}{p} \leq c (1+|( \mathcal{E}{\bf u})_\varrho|)\, . \end{equation*} |
Collecting the previous estimates, we finally get
\begin{equation*} \begin{split} \left|\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} {\bf u}[D{\bf u}][\eta^q({\bf u} - {\mathit{\boldsymbol{\ell}}}_{\varrho})-{{\bf w}}]\, \mathrm{d}x\right| & \leq c \varrho (1+|( \mathcal{E}{\bf u})_\varrho|)^2 \\ & \leq c \varrho H(x_0, 1+|( \mathcal{E}{\bf u})_\varrho|)\, . \end{split} \end{equation*} |
We are left to estimate J_5 . Using Hölder's inequality, p^* > \frac{n}{n-1} and the estimate of I_3 we have
\begin{equation*} \begin{split} |J_5| & \leq c \left(\mathop {{\rlap{-} \smallint }}\limits_{B_\varrho} |f|^n\, \mathrm{d}x\right)^\frac{1}{n} \left(\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} (|{\bf u} - {\mathit{\boldsymbol{\ell}}}_\varrho|^{\frac{n}{n-1}}+|{{\bf w}}|^{\frac{n}{n-1}})\, \mathrm{d}x\right)^\frac{n-1}{n} \\ & \leq c \varrho^{-\frac{1}{\beta+1}}\|f\|_{L^{n(1+\beta)}(\Omega)} \left(\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} (|{\bf u} - {\mathit{\boldsymbol{\ell}}}_\varrho|^{p^*}+|{{\bf w}}|^{p^*})\, \mathrm{d}x\right)^\frac{1}{p^*} \\ & \leq c \varrho^{1-\frac{1}{\beta+1}} (1+|( \mathcal{E}{\bf u})_\varrho|) \\ & \leq c \varrho^{1-\frac{1}{\beta+1}}\, H(x_0, 1+|( \mathcal{E}{\bf u})_\varrho|)\, . \end{split} \end{equation*} |
The proof of (3.4) is then concluded.
The following result can be obtained combining the argument of [36, Theorem 3.2] with Lemma 2.7.
Lemma 3.2. There exist 0 < \sigma_1 < 1 depending on n, p, q such that for any B_\varrho(x_0)\subset\subset\Omega we have
\begin{equation*} \begin{split} &\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}(x_0)} H_{1+|( \mathcal{E}{\bf u})_{x_0, \varrho}|}\left(x_0, \frac{|{\bf u}-{\mathit{\boldsymbol{\ell}}}_{x_0, \varrho}|}{\varrho}\right)\, \mathrm{d}x \\ & \leq c \left(1+ [\mu]_{C^\alpha}\| \mathcal{E}{\bf u}\|^{q-p}_{L^p(\Omega)}\varrho^{\alpha-\frac{(q-p)n}{p}}\right) \left(\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}(x_0)} \left[H_{1+|( \mathcal{E}{\bf u})_{x_0, \varrho}|}(x_0, | \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_{x_0, \varrho}|)\right]^{\sigma_1}\, \mathrm{d}x \right)^{\frac{1}{\sigma_1}} \end{split} \end{equation*} |
for some constant c = c(n, p, q) > 1 .
Proof. Let a\geq0 and denote by H_a(x_0, t) the shifted N -function of H(x_0, t) defined as in (2.8). Note that, by the definition of H and (2.9), we have
\begin{equation*} \begin{split} H_a(x_0, t) & \sim (a+t)^{p-2} t^2 + \mu(x_0) (a+t)^{q-2} t^2 \\ & \, \, \, \, \, \, \, \, = : G_{p, a}(t) + \mu(x_0) G_{q, a}(t)\, , \end{split} \end{equation*} |
where the N -functions G_{p, a} and G_{q, a} are equivalent to the corresponding shifted N -functions of t^p and t^q , respectively. Note that
\begin{equation*} G_{p, a}(t) \sim a^p + t^p \sim (a+t)^p\, , \quad G_{q, a}(t) \sim a^q + t^q \sim (a+t)^q\, . \end{equation*} |
We first assume that
\begin{equation} \sup\limits_{x\in B_\varrho} \mu(x) \leq 4 [\mu]_{C^\alpha} \varrho^\alpha\, . \end{equation} | (3.17) |
To enlighten the notation, we set
\begin{equation*} {\bf f}_{x_0, \varrho}: = {\bf u}-{\mathit{\boldsymbol{\ell}}}_{x_0, \varrho}\, . \end{equation*} |
Note that ({\bf f}_{x_0, \varrho})_{x_0, \varrho} = {\bf 0} .
Using the Sobolev-Poincaré inequality, together with (3.17), we get
\begin{equation} \begin{split} \mathop {{\rlap{-} \smallint }}\limits_{B_\varrho}\left(a+\frac{|{\bf f}_{x_0, \varrho}|}{\varrho}\right)^p \, \mathrm{d}x & \leq c \left(\mathop {{\rlap{-} \smallint }}\limits_{B_\varrho} \left(a+|D{\bf f}_{x_0, \varrho}|\right)^\frac{np}{n+p}\, \mathrm{d}x\right)^\frac{n+p}{n}\, , \\ \mathop {{\rlap{-} \smallint }}\limits_{B_\varrho} \mu(x_0)\left(a+\frac{|{\bf f}_{x_0, \varrho}|}{\varrho}\right)^q \, \mathrm{d}x & \leq c 4 [\mu]_{C^\alpha} \varrho^\alpha \left(\mathop {{\rlap{-} \smallint }}\limits_{B_\varrho} (a+|D{\bf f}_{x_0, \varrho}|)^\frac{nq}{n+q}\, \mathrm{d}x\right)^\frac{n+q}{n}\, . \end{split} \end{equation} | (3.18) |
Now, invoking the Korn's inequality of Lemma 2.7, we have
\begin{equation*} \begin{split} \mathop {{\rlap{-} \smallint }}\limits_{B_\varrho}\left(a+\frac{|{\bf f}_{x_0, \varrho}|}{\varrho}\right)^p \, \mathrm{d}x & \leq c_{\rm Korn} \left(\mathop {{\rlap{-} \smallint }}\limits_{B_\varrho} \left(a+| \mathcal{E}{\bf f}_{x_0, \varrho}|\right)^\frac{np}{n+p}\, \mathrm{d}x\right)^\frac{n+p}{n}\, , \\ \mathop {{\rlap{-} \smallint }}\limits_{B_\varrho} \mu(x_0)\left(a+\frac{|{\bf f}_{x_0, \varrho}|}{\varrho}\right)^q \, \mathrm{d}x & \leq c c_{\rm Korn} 4 [\mu]_{C^\alpha} \varrho^\alpha \left(\mathop {{\rlap{-} \smallint }}\limits_{B_\varrho} (a+| \mathcal{E}{\bf f}_{x_0, \varrho}|)^\frac{nq}{n+q}\, \mathrm{d}x\right)^\frac{n+q}{n}\, . \end{split} \end{equation*} |
By Hölder's inequality, we get
\begin{equation*} \begin{split} \left(\mathop {{\rlap{-} \smallint }}\limits_{B_\varrho} (a+| \mathcal{E}{\bf f}_{x_0, \varrho}|)^\frac{nq}{n+q}\, \mathrm{d}x\right)^\frac{n+q}{n} & \leq \left(\mathop {{\rlap{-} \smallint }}\limits_{B_\varrho} (a+| \mathcal{E}{\bf f}_{x_0, \varrho}|)^p\, \mathrm{d}x\right)^\frac{q-p}{p} \left(\mathop {{\rlap{-} \smallint }}\limits_{B_\varrho} (a+| \mathcal{E}{\bf f}_{x_0, \varrho}|)^\frac{nq}{n+q}\, \mathrm{d}x\right)^\frac{p(n+q)}{n} \\ & \leq \varrho^{-\frac{n(q-p)}{p}} \|a+| \mathcal{E}{\bf u}|\|_{L^p(\Omega)}^{q-p} \left(\mathop {{\rlap{-} \smallint }}\limits_{B_\varrho} (a+| \mathcal{E}{\bf f}_{x_0, \varrho}|)^{p\sigma_1}\, \mathrm{d}x\right)^\frac{1}{\sigma_1}\, , \end{split} \end{equation*} |
where we have set \sigma_1: = \frac{nq}{p(n+q)} < 1 . Combining with (3.18), we then obtain
\begin{equation*} \begin{split} \mathop {{\rlap{-} \smallint }}\limits_{B_\varrho} & \left[\left(a+\frac{|{\bf f}_{x_0, \varrho}|}{\varrho}\right)^p \, \mathrm{d}x+ \mu(x_0)\left(a+\frac{|{\bf f}_{x_0, \varrho}|}{\varrho}\right)^q \right]\, \mathrm{d}x \\ & \leq c \left[1+4 [\mu]_{C^\alpha} \varrho^{\alpha-\frac{n(q-p)}{p}} \|a+| \mathcal{E}{\bf u}|\|_{L^p(\Omega)}^{q-p}\right] \left(\mathop {{\rlap{-} \smallint }}\limits_{B_\varrho} (a+| \mathcal{E}{\bf f}_{x_0, \varrho}|)^{p\sigma_1}\, \mathrm{d}x\right)^\frac{1}{\sigma_1}\, . \end{split} \end{equation*} |
If instead (3.17) does not hold, we can apply Theorem 2.3 and Lemma 2.7 with G(t): = H_a(x_0, t) . The proof is concluded.
Now, we are in position to establish a higher integrability result for H_{1+|(\mathcal{E}{\bf u})_{x_0, \varrho}|}(x_0, | \mathcal{E}{\bf u}-(\mathcal{E}{\bf u})_{x_0, \varrho}|) . The result follows from Lemma 3.2 and Lemma 3.1 as a consequence of Gehring's lemma with increasing supports (Lemma 2.8):
Corollary 2. There exist a constant c_{\rm high} = c_{\rm high}(data) > 0 and \sigma > 1 such that
\begin{equation} \begin{split} &\left(\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho/2}(x_0)} \left[H_{1+|( \mathcal{E}{\bf u})_{x_0, \varrho}|}(x_0, | \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_{x_0, \varrho}|)\right]^\sigma\, \mathrm{d}x \right)^{\frac{1}{\sigma}} \\ &\leq c_{\rm high} \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}(x_0)} H_{1+|( \mathcal{E}{\bf u})_{x_0, \varrho}|}(x_0, | \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_{x_0, \varrho}|)\, \mathrm{d}x + c_{\rm high} \varrho^{\widetilde{\gamma}}\, H(x_0, 1+|( \mathcal{E}{\bf u})_{x_0, \varrho}|)\, . \end{split} \end{equation} | (3.19) |
We can start with the linearization procedure for system (1.1). Let us set
\begin{equation} \mathcal{A}\colon = \frac{ D_\xi{\bf a}(x_0, ( \mathcal{E}{\bf u})_{x_0, \varrho})}{ H''(x_0, 1+|( \mathcal{E}{\bf u})_{x_0, \varrho}|)} \quad \mbox{ and } \quad {{\bf w}}: = {{\bf u} - {\mathit{\boldsymbol{\ell}}}_{x_0, \varrho}}\, . \end{equation} | (3.20) |
Note that the bilinear form \mathcal{A} satisfies the Legendre-Hadamard condition (2.20) by virtue of (1.4).
We aim to prove that {{\bf w}} is approximately \mathcal{A} -Stokes. This fact, together with the higher integrability result (3.19) will allow us to apply the \mathcal{A} -Stokes approximation theorem.
Lemma 3.3. Let B_{\varrho}(x_0)\subset\subset\Omega . Assume that (3.2) and (3.3) hold, and let \varrho\leq1 . Then there exists a constant c_{\rm Stokes} > 0 such that
\begin{equation} \begin{split} & \left|\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho/2}(x_0)}\mathcal{A} \mathcal{E}{{\bf w}}: \mathcal{E}{\mathit{\boldsymbol{\varphi}}}\, \mathrm{d}x\right| \\ & \leq c_{\rm Stokes} (1+|( \mathcal{E}{\bf u})_{x_0, \varrho}|)\mathcal{R}\left(\frac{{\mathit{\Phi}}(x_0, \varrho)}{H(x_0, 1+|( \mathcal{E}{\bf u})_{x_0, \varrho}|)}+\varrho^{\widetilde{\gamma}}\right)\|D{\mathit{\boldsymbol{\varphi}}}\|_{L^\infty(B_{\varrho/2}(x_0))} \end{split} \end{equation} | (3.21) |
for every {\mathit{\boldsymbol{\varphi}}}\in C_{0, {\rm div}}^\infty(B_{\varrho/2}(x_0)) , where \mathcal{R}(t): = [\omega(t^\frac{1}{2}) + t]^\frac{1}{2} \sqrt{t} .
Proof. It will suffice to prove (3.21) for any fixed {\mathit{\boldsymbol{\varphi}}}\in C_0^\infty(B_{\varrho/2}(x_0)) with {\rm div}\, {\mathit{\boldsymbol{\varphi}}} = 0 such that \|D{\mathit{\boldsymbol{\varphi}}}\|_{L^\infty(B_{\varrho/2}(x_0))}\leq1 , since the general case will follow by a standard normalization argument. To enlighten notation, we omit the explicit dependence on x_0 .
From the definitions of \mathcal{A} and {{\bf w}} we have
\begin{equation} \begin{split} & H''(x_0, 1+|( \mathcal{E}{\bf u})_\varrho|) \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho/2}}\mathcal{A} \mathcal{E}{{\bf w}}: \mathcal{E}{\mathit{\boldsymbol{\varphi}}}\, \mathrm{d}x \\ &\, \, = \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho/2}} D_\xi{\bf a}(x_0, ( \mathcal{E}{\bf u})_\varrho) \mathcal{E}({\bf u}-{\mathit{\boldsymbol{\ell}}}_\varrho): \mathcal{E}{\mathit{\boldsymbol{\varphi}}}\, \mathrm{d}x \\ & \, \, = \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho/2}} \int_0^1 \left[D_\xi{\bf a}(x_0, ( \mathcal{E}{\bf u})_\varrho) - D_\xi{\bf a}(x_0, ( \mathcal{E}{\bf u})_\varrho + t \mathcal{E}({\bf u}-{\mathit{\boldsymbol{\ell}}}_\varrho))\right] \mathcal{E}({\bf u}-{\mathit{\boldsymbol{\ell}}}_\varrho) : \mathcal{E}{\mathit{\boldsymbol{\varphi}}}\, \mathrm{d}t\, \mathrm{d}x \\ & \, \, \, \, \, \, + \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho/2}} \int_0^1 \left[D_\xi{\bf a}(x_0, ( \mathcal{E}{\bf u})_\varrho + t \mathcal{E}({\bf u}-{\mathit{\boldsymbol{\ell}}}_\varrho))\right] \mathcal{E}({\bf u}-{\mathit{\boldsymbol{\ell}}}_\varrho) : \mathcal{E}{\mathit{\boldsymbol{\varphi}}}\, \mathrm{d}t\, \mathrm{d}x \\ & \, \, = : J_1 + J_2\, . \end{split} \end{equation} | (3.22) |
We set G(t): = H(x_0, t) . From (1.7), (2.6), (2.2) and Lemma 2.2, we have
\begin{equation*} \begin{split} |J_1| & \leq c \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho/2}} \int_0^1 \omega\left(\frac{t | \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|}{1+|( \mathcal{E}{\bf u})_\varrho|}\right) G''(1+|( \mathcal{E}{\bf u})_\varrho|+|( \mathcal{E}{\bf u})_\varrho+t( \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho)|) \, \mathrm{d}t\, | \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|\, \mathrm{d}x \\ & \leq c \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho/2}} \omega\left(\frac{ | \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|}{1+|( \mathcal{E}{\bf u})_\varrho|}\right) \left[\frac{G'(1+|( \mathcal{E}{\bf u})_\varrho|+| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|)}{1+|( \mathcal{E}{\bf u})_\varrho|+| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|}\right]| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|\, \mathrm{d}x \\ & \leq c \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho/2}} {\mathbb{1}}_E(x) \omega\left(\frac{ | \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|}{1+|( \mathcal{E}{\bf u})_\varrho|}\right) \frac{G'(1+|( \mathcal{E}{\bf u})_\varrho|+| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|)}{1+|( \mathcal{E}{\bf u})_\varrho|}| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|\, \mathrm{d}x \\ & \, \, \, \, \, \, + c \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho/2}} {\mathbb{1}}_F(x) \omega\left(\frac{ | \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|}{1+|( \mathcal{E}{\bf u})_\varrho|}\right) \frac{G'(1+|( \mathcal{E}{\bf u})_\varrho|+| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|)}{1+|( \mathcal{E}{\bf u})_\varrho|}| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|\, \mathrm{d}x \\ & = : J_{1, E} + J_{1, F}\, , \end{split} \end{equation*} |
where E: = \{x\in B_{\varrho/2}:\, \, | \mathcal{E}{\bf u}(x)-(\mathcal{E}{\bf u})_\varrho| > 1+|(\mathcal{E}{\bf u})_\varrho|\} , and F: = B_\varrho\backslash E .
We start with the estimate of J_{1, E} . We have
\begin{equation*} \begin{split} \frac{|J_{1, E}|}{G'(1+|( \mathcal{E}{\bf u})_\varrho|)} & \leq c \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho/2}} {\mathbb{1}}_E(x) \omega\left(\frac{| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|}{1+|( \mathcal{E}{\bf u})_\varrho|}\right) \frac{G'(1+|( \mathcal{E}{\bf u})_\varrho|+| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|)}{G(1+|( \mathcal{E}{\bf u})_\varrho|)}| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|\, \mathrm{d}x\, . \end{split} \end{equation*} |
For a.e. x\in E , we have
\begin{equation*} 1+|( \mathcal{E}{\bf u})_\varrho|+| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho| \leq 2 | \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|\, , \end{equation*} |
and taking into account (2.10), we obtain
\begin{equation*} \begin{split} G'(1+|( \mathcal{E}{\bf u})_\varrho|+| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|) | \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho| & \leq c G(| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|) \\ & \leq c G(1+|( \mathcal{E}{\bf u})_\varrho|+| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|)\\ & \leq c G_{1+|( \mathcal{E}{\bf u})_\varrho|}(| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|)\, . \end{split} \end{equation*} |
Now, using that \omega\leq1 , we finally get
\begin{equation*} {\mathbb{1}}_E(x) \omega\left(\frac{| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|}{1+|( \mathcal{E}{\bf u})_\varrho|}\right) \frac{G'(1+|( \mathcal{E}{\bf u})_\varrho|+| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|)}{G(1+|( \mathcal{E}{\bf u})_\varrho|)}| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho| \leq \frac{G_{1+|( \mathcal{E}{\bf u})_\varrho|}(| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|)} {G(1+|( \mathcal{E}{\bf u})_\varrho|)}\, , \end{equation*} |
whence
\begin{equation*} |J_{1, E}| \leq c G'(1+|( \mathcal{E}{\bf u})_\varrho|)\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho/2}} \frac{G_{1+|( \mathcal{E}{\bf u})_\varrho|}(| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|)} {G(1+|( \mathcal{E}{\bf u})_\varrho|)}\, \mathrm{d}x\, . \end{equation*} |
For what concerns J_{1, F} , arguing as for J_{1, E} we get the preliminary estimate
\begin{equation*} \frac{|J_{1, F}|}{G'(1+|( \mathcal{E}{\bf u})_\varrho|)}\leq c \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho/2}} {\mathbb{1}}_F(x) \omega\left(\frac{| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|}{1+|( \mathcal{E}{\bf u})_\varrho|}\right) \frac{G'(1+|( \mathcal{E}{\bf u})_\varrho|+| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|)}{G(1+|( \mathcal{E}{\bf u})_\varrho|)}| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|\, \mathrm{d}x\, . \end{equation*} |
For a.e. x\in F , we have
\begin{equation*} 1+|( \mathcal{E}{\bf u})_\varrho|+| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho| < 2(1+|( \mathcal{E}{\bf u})_\varrho|)\, , \end{equation*} |
and with (2.6) and since G' is increasing, we can write
\begin{equation*} |J_{1, F}|\leq c G'(1+|( \mathcal{E}{\bf u})_\varrho|)\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho/2}} {\mathbb{1}}_F(x) \omega\left(\frac{| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|}{1+|( \mathcal{E}{\bf u})_\varrho|}\right) \frac{| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|}{1+|( \mathcal{E}{\bf u})_\varrho|}\, \mathrm{d}x\, . \end{equation*} |
Using the definition of F again, the monotonicity of G' and (2.6), for a.e. x\in F we have
\begin{equation*} \begin{split} & \omega\left(\frac{| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|}{1+|( \mathcal{E}{\bf u})_\varrho|}\right) \frac{| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|}{1+|( \mathcal{E}{\bf u})_\varrho|} \\ & \leq \omega\left(\left[\frac{G'(1+|( \mathcal{E}{\bf u})_\varrho|)| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|^2}{G'(1+|( \mathcal{E}{\bf u})_\varrho|)(1+|( \mathcal{E}{\bf u})_\varrho|)^2}\right]^\frac{1}{2}\right) \left[\frac{G'(1+|( \mathcal{E}{\bf u})_\varrho|)| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|^2}{G'(1+|( \mathcal{E}{\bf u})_\varrho|)(1+|( \mathcal{E}{\bf u})_\varrho|)^2} \right]^\frac{1}{2}\\ & \leq \omega\left(\left[\frac{G'(1+|( \mathcal{E}{\bf u})_\varrho|+| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|)| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|^2}{G(1+|( \mathcal{E}{\bf u})_\varrho|)(1+|( \mathcal{E}{\bf u})_\varrho|+| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|)}\right]^\frac{1}{2}\right) \left[\frac{G'(1+|( \mathcal{E}{\bf u})_\varrho|+| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|)| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|^2}{G(1+|( \mathcal{E}{\bf u})_\varrho|)(1+|( \mathcal{E}{\bf u})_\varrho|+| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|)}\right]^\frac{1}{2}\, . \end{split} \end{equation*} |
Using (2.9) we then get
\begin{equation*} \begin{split} & \omega\left(\frac{| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|}{1+|( \mathcal{E}{\bf u})_\varrho|}\right) \frac{| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|}{1+|( \mathcal{E}{\bf u})_\varrho|} \\ & \leq \omega\left(\left[\frac{G_{1+|( \mathcal{E}{\bf u})_\varrho|}(| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|)}{G(1+|( \mathcal{E}{\bf u})_\varrho|)}\right]^\frac{1}{2}\right) \left[\frac{G_{1+|( \mathcal{E}{\bf u})_\varrho|}(| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|)}{G(1+|( \mathcal{E}{\bf u})_\varrho|)}\right]^\frac{1}{2}\, . \end{split} \end{equation*} |
Combining the previous estimates and using Hölder's inequality, the fact that \omega\leq1 and Jensen's inequality for the concave function \omega(t^\frac{1}{2}) , we get
\begin{equation*} \frac{|J_{1, F}|}{G'(1+|( \mathcal{E}{\bf u})_\varrho|)} \leq c \left\{ \left[\omega\left(\left[\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho/2}}\frac{G_{1+|( \mathcal{E}{\bf u})_\varrho|}(| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|)} {G(1+|( \mathcal{E}{\bf u})_\varrho|)}\, \mathrm{d}x \right]^\frac{1}{2}\right) \right]^\frac{1}{2} \left( \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho/2}} \frac{G_{1+|( \mathcal{E}{\bf u})_\varrho|}(| \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|)} {G(1+|( \mathcal{E}{\bf u})_\varrho|)}\, \mathrm{d}x \right)^\frac{1}{2}\right\}\, . \end{equation*} |
Collecting the estimates for J_{1, E} and J_{1, F} , we then infer
\begin{equation} \begin{split} & \frac{|J_1|}{G''(1+|( \mathcal{E}{\bf u})_\varrho|)(1+|( \mathcal{E}{\bf u})_\varrho|)} \\ & \leq c \left[\frac{{\mathit{\Phi}}(\varrho)}{G(1+|( \mathcal{E}{\bf u})_\varrho|)} + \sqrt{\omega\left(\left[\frac{{\mathit{\Phi}}(\varrho)}{G(1+|( \mathcal{E}{\bf u})_\varrho|)}\right]^\frac{1}{2}\right)}\sqrt{\frac{{\mathit{\Phi}}(\varrho)}{G(1+|( \mathcal{E}{\bf u})_\varrho|)}}\right] \, . \end{split} \end{equation} | (3.23) |
In order to estimate J_2 , we use (3.6), the definition of weak solution, (1.6) and we preliminarly obtain
\begin{equation} \begin{split} |J_2| & = \left|\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho/2}} \left[{\bf a}(x_0, \mathcal{E}{\bf u})-{\bf a}(x_0, ( \mathcal{E}{\bf u})_\varrho)\right]: \mathcal{E}{\mathit{\boldsymbol{\varphi}}}\, \mathrm{d}x \right| \\ & = \left|\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho/2}} \left[{\bf a}(x_0, \mathcal{E}{\bf u})-{\bf a}(x, \mathcal{E}{\bf u})\right]: \mathcal{E}{\mathit{\boldsymbol{\varphi}}}\, \mathrm{d}x + \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho/2}} [{\bf u}[D{\bf u}]+f]{\mathit{\boldsymbol{\varphi}}}\, \mathrm{d}x \right| \\ & \leq c \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho/2}}|\mu(x)-\mu(x_0)|(1+| \mathcal{E}{\bf u}|)^{q-1}\, \mathrm{d}x + \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho/2}}|{\bf u}||D{\bf u}||{\mathit{\boldsymbol{\varphi}}}|\, \mathrm{d}x + \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho/2}}|f||{\mathit{\boldsymbol{\varphi}}}|\, \mathrm{d}x \\ & = : J_{2, 1}+J_{2, 2}+J_{2, 3}. \end{split} \end{equation} | (3.24) |
Now, we have to distinguish between two cases, depending on whether condition (3.10) is satisfied or not.
Step 1: p -phase. Under assumption (3.10), we have
\begin{equation} \begin{split} |J_{2, 1}| & \leq c \varrho^{\frac{\alpha-\varepsilon_0}{q}}\left(\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho/2}}|\mu(x)|^{\frac{q-1}{q}}(1+| \mathcal{E}{\bf u}|)^{q-1}\, \mathrm{d}x + \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho/2}}|\mu(x_0)|^{\frac{q-1}{q}}(1+| \mathcal{E}{\bf u}|)^{q-1}\, \mathrm{d}x\right) \\ & \leq c \varrho^{\frac{\alpha-\varepsilon_0}{q}}\left[\left(\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho/2}}\mu(x)(1+| \mathcal{E}{\bf u}|)^{q}\, \mathrm{d}x\right)^\frac{q-1}{q} + \left(\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho/2}}\mu(x_0)(1+| \mathcal{E}{\bf u}|)^{q}\, \mathrm{d}x\right)^\frac{q-1}{q}\right]\, . \end{split} \end{equation} | (3.25) |
An analogous argument as for (3.12) and (3.13) shows that
\begin{equation*} \begin{split} H(x_0, 1+|( \mathcal{E}{\bf u})_\varrho|) & \leq c(1+\varrho^{\alpha-\varepsilon_0}(1+|( \mathcal{E}{\bf u})_\varrho|)^{q-p}) (1+|( \mathcal{E}{\bf u})_\varrho|)^p \\ & \leq c (1+|( \mathcal{E}{\bf u})_\varrho|)^p\, , \end{split} \end{equation*} |
whence, using the change of shift formula (2.12) and (3.3), we obtain
\begin{equation*} \begin{split} \mathop {{\rlap{-} \smallint }}\limits_{B_\varrho} H(x_0, 1+| \mathcal{E}{\bf u}|)\, \mathrm{d}x & \leq c_\eta \mathop {{\rlap{-} \smallint }}\limits_{B_\varrho} H_{1+|( \mathcal{E}{\bf u})_\varrho|}(x_0, | \mathcal{E}{\bf u} - ( \mathcal{E}{\bf u})_\varrho|)\, \mathrm{d}x + \eta H(x_0, 1+|( \mathcal{E}{\bf u})_\varrho|) \\ & \leq c H(x_0, 1+|( \mathcal{E}{\bf u})_\varrho|) \\ & \leq c (1+|( \mathcal{E}{\bf u})_\varrho|)^p\, . \end{split} \end{equation*} |
With this and (2.39) we then get
\begin{equation*} \begin{split} \varrho^{\frac{\alpha-\varepsilon_0}{q}} \left(\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho/2}}\mu(x)(1+| \mathcal{E}{\bf u}|)^q\, \mathrm{d}x\right)^\frac{q-1}{q} & \leq \varrho^{\frac{\alpha-\varepsilon_0}{q}} \left(\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}}(1+| \mathcal{E}{\bf u}|)^p\, \mathrm{d}x\right)^\frac{q-1}{q} \\ & \leq c \varrho^{\frac{\alpha-\varepsilon_0}{q}} \left(\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}}(1+| \mathcal{E}{\bf u}|)^{ps_0}\, \mathrm{d}x\right)^\frac{q-p}{qs_0}\left(\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}}H(x_0, 1+| \mathcal{E}{\bf u}|)^{ps_0}\, \mathrm{d}x\right)^\frac{p-1}{q} \\ & \leq c \varrho^{\frac{1}{q}(\alpha-\varepsilon_0-(q-p)n+2p\varepsilon_0)}\|1+| \mathcal{E}{\bf u}|\|_{L^{ps_0}(B_{\varrho})}(1+|( \mathcal{E}{\bf u})_\varrho|)^\frac{p(p-1)}{q} \\ & \leq c \varrho^{\frac{\varepsilon_0}{q}}(1+|( \mathcal{E}{\bf u})_\varrho|)^{p-1}\, . \end{split} \end{equation*} |
A similar estimate holds for the second summand in the right hand side of (3.25), with \mu(x_0) in place of \mu(x) . Inserting these estimates in (3.25) and using the definition of H_{1}(x_0, 1+|(\mathcal{E}{\bf u})_\varrho|) we finally obtain
\begin{equation} \begin{split} \frac{|J_{2, 1}|}{H''(x_0, 1+|( \mathcal{E}{\bf u})_\varrho|)(1+|( \mathcal{E}{\bf u})_\varrho|)} & \leq c \varrho^{\frac{\varepsilon_0}{q}}\, .\\ \end{split} \end{equation} | (3.26) |
Step 2: (p, q) -phase. If (3.10) does not hold, then (3.15) is in force. Now, using (3.3), the triangle inequality and the change of shift formula, we get
\begin{equation*} \begin{split} \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}}H(x_0, 1+| \mathcal{E}{\bf u}|)\, \mathrm{d}x & \leq c \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}}H(x_0, | \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|)\, \mathrm{d}x + c H(x_0, 1+|( \mathcal{E}{\bf u})_\varrho|) \\ & \leq c_\eta \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}}H_{1+|( \mathcal{E}{\bf u})_\varrho|}(x_0, | \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_\varrho|)\, \mathrm{d}x +(c+\eta)H(x_0, 1+|( \mathcal{E}{\bf u})_\varrho|) \\ & \leq c H(x_0, 1+|( \mathcal{E}{\bf u})_\varrho|)\, . \end{split} \end{equation*} |
With this, Hölder's inequality and Jensen's inequality, we can estimate J_{2, 1} as
\begin{equation*} \begin{split} |J_{2, 1}| & \leq c \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho/2}}\varrho^{\varepsilon_0}H'(x_0, 1+| \mathcal{E}{\bf u}|)\, \mathrm{d}x \\ & \leq c \varrho^{\varepsilon_0}\left(\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}}[H'(x_0, 1+| \mathcal{E}{\bf u}|)]^{\frac{2q-1}{2q-2}}\, \mathrm{d}x\right)^\frac{2q-2}{2q-1} \\ & \leq c \varrho^{\varepsilon_0}(H'(x_0, \cdot)\circ H^{-1}(x_0, \cdot))\left(\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}}H(x_0, 1+| \mathcal{E}{\bf u}|)\, \mathrm{d}x\right) \\ & \leq \varrho^{\varepsilon_0} H'(x_0, 1+|( \mathcal{E}{\bf u})_\varrho|)\, , \end{split} \end{equation*} |
which corresponds to (3.26). In order to estimate J_{2, 2} and J_{2, 3} , we may argue as in the proof of Lemma 3.1, exploiting the smallness assumptions in (3.2), so we briefly sketch the proof.
Setting \gamma: = \left(\frac{p^*}{2}\right)' again, and using Hölder's inequality we have
\begin{equation*} \begin{split} |J_{2, 2}| & \leq \left(\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} |{\bf u}|^{p^*}\, \mathrm{d}x\right)^\frac{1}{p^*}\left(\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} |D{\bf u}|^{\gamma}\, \mathrm{d}x\right)^\frac{1}{\gamma}\left(\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} |{\mathit{\boldsymbol{\varphi}}}|^{p^*}\, \mathrm{d}x\right)^\frac{1}{p^*} \, , \end{split} \end{equation*} |
then the proof is similar. The only difference is the use of the Sobolev-Korn inequality (2.19) for {\mathit{\boldsymbol{\varphi}}} with | \mathcal{E}{\mathit{\boldsymbol{\varphi}}}|\leq1 . We then obtain
\begin{equation*} |J_{2, 2}| \leq c \varrho (1+|( \mathcal{E}{\bf u})_\varrho|) \leq c \varrho H'(x_0, 1+|( \mathcal{E}{\bf u})_\varrho|)\, . \end{equation*} |
For what concerns J_{2, 3} , using Hölder's inequality, the fact that p^* > \frac{n}{n-1} and the Sobolev-Korn inequality for {\mathit{\boldsymbol{\varphi}}} with | \mathcal{E}{\mathit{\boldsymbol{\varphi}}}|\leq1 , we have
\begin{equation*} \begin{split} |J_{2, 3}| & \leq c \left(\mathop {{\rlap{-} \smallint }}\limits_{B_\varrho} |f|^n\, \mathrm{d}x\right)^\frac{1}{n} \left(\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} |{\mathit{\boldsymbol{\varphi}}}|^{\frac{n}{n-1}}\, \mathrm{d}x\right)^\frac{n-1}{n} \\ & \leq c \varrho^{-\frac{1}{\beta+1}}\|f\|_{L^{n(1+\beta)}(\Omega)} \left(\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} |{\mathit{\boldsymbol{\varphi}}}|^{p^*}\, \mathrm{d}x\right)^\frac{1}{p^*} \\ & \leq c \varrho^{1-\frac{1}{\beta+1}}\\ & \leq c \varrho^{1-\frac{1}{\beta+1}}\, H'(x_0, 1+|( \mathcal{E}{\bf u})_\varrho|)\, . \end{split} \end{equation*} |
Collecting the previous estimates, we then obtain
\begin{equation} \frac{|J_2|}{H'(x_0, 1+|( \mathcal{E}{\bf u})_\varrho|)} \leq c \varrho^{\widetilde{\gamma}}\, , \end{equation} | (3.27) |
where \widetilde{\gamma}: = \min\{\varepsilon_0, 1- \frac{1}{\beta+1}\} . The final estimate (3.21) then follows inserting (3.23) and (3.27) into (3.22).
In this section, we prove an excess improvement estimate for weak solutions to (1.1). This will be the content of Lemma 3.5. We start with a technical tool useful in the sequel.
Lemma 3.4. Let \vartheta\in(0, 1) . Assume that
\begin{equation} \frac{{\mathit{\Phi}}(x_0, \varrho)}{H(x_0, 1+|( \mathcal{E}{\bf u})_{x_0, \varrho}|)}\leq \frac{\vartheta^n}{2^{q+1} c_{q}}\, , \end{equation} | (3.28) |
where c_{q} is the constant of the change of shift formula (2.12) with \eta = \frac{1}{2^{q+1}} . Then,
\begin{equation} 1+|( \mathcal{E}{\bf u})_{x_0, \varrho}| \leq 2(1+ |( \mathcal{E}{\bf u})_{x_0, \vartheta\varrho}|)\, . \end{equation} | (3.29) |
Proof. As a consequence of (2.12) for \eta = \frac{1}{2^{q+1}} and with (3.28) we get
\begin{equation*} \begin{split} H(x_0, |( \mathcal{E}{\bf u})_{\varrho}-( \mathcal{E}{\bf u})_{\vartheta\varrho}|) & \leq \mathop {{\rlap{-} \smallint }}\limits_{B_{\vartheta\varrho}} H(x_0, | \mathcal{E}{\bf u} - ( \mathcal{E}{\bf u})_{\varrho}|)\, \mathrm{d}x \\ & \leq c_{q}\vartheta^{-n}{\mathit{\Phi}}(\varrho) + \frac{1}{2^{q+1}}H(x_0, 1+|( \mathcal{E}{\bf u})_{\varrho}|) \\ & \leq \frac{1}{2^{q}} H(x_0, 1+|( \mathcal{E}{\bf u})_{\varrho}|)\, , \end{split} \label{eq:computation} \end{equation*} |
whence, passing to [H(x_0, \cdot)]^{-1} and taking into account (2.7), we obtain
\begin{equation*} |( \mathcal{E}{\bf u})_{\varrho}-( \mathcal{E}{\bf u})_{\vartheta\varrho}| \leq \frac{1}{2}(1+|( \mathcal{E}{\bf u})_{\varrho}|)\, . \end{equation*} |
Now,
\begin{equation*} \begin{split} 1+|( \mathcal{E}{\bf u})_{\varrho}| \leq |( \mathcal{E}{\bf u})_{\varrho}-( \mathcal{E}{\bf u})_{\theta\varrho}|+1+|( \mathcal{E}{\bf u})_{\theta\varrho}| \leq \frac{1}{2}(1+|( \mathcal{E}{\bf u})_{\varrho}|) + 1+ |( \mathcal{E}{\bf u})_{\theta\varrho}|\, , \end{split} \end{equation*} |
whence (3.29) follows by re-absorbing the first term of the right-hand side into the left.
We are now in position to prove the excess improvement estimate.
Lemma 3.5. For any fixed \theta\in(0, \frac{1}{8}) , there exists \varepsilon_1 = \varepsilon_1(n, \nu, L, p, q, [\mu]_{C^\alpha}, \theta)\in(0, 1) such that if
\begin{equation} \frac{{\mathit{\Phi}}(x_0, \varrho)}{H(x_0, 1+|( \mathcal{E}{\bf u})_{x_0, \varrho}|)}+\varrho^{\widetilde{\gamma}}\leq \varepsilon_1\, , \quad \varrho\leq\theta^n\, , \end{equation} | (3.30) |
then
\begin{equation} {\mathit{\Phi}}(x_0, \theta\varrho) \leq c_{\rm dec} \theta^2\left[{\mathit{\Phi}}(x_0, \varrho) + \varrho^{\widetilde{\gamma}}H(x_0, 1+|( \mathcal{E}{\bf u})_{x_0, \varrho}|)\right] \end{equation} | (3.31) |
for some constant c_{\rm dec} = c_{\rm dec}(n, \nu, L, p, q, [\mu]_{C^\alpha}, \| \mathcal{E}{\bf u}\|_{L^p(\Omega)})\geq1 .
Proof. Step 1: By virtue of (3.4) we have
\begin{equation} \begin{split} & \mathop {{\rlap{-} \smallint }}\limits_{B_{\theta\varrho}} H_{1+|( \mathcal{E}{\bf u})_{\theta\varrho}|}(x_0, | \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_{\theta\varrho}|)\, \mathrm{d}x \\ & \, \, \, \, \leq c \mathop {{\rlap{-} \smallint }}\limits_{B_{\theta\varrho}} H_{1+|( \mathcal{E}{\bf u})_{\theta\varrho}|}\left(x_0, \frac{|{\bf u}-{\mathit{\boldsymbol{\ell}}}_{\theta\varrho}|}{\theta\varrho}\right)\, \mathrm{d}x + c (\theta\varrho)^{\widetilde{\gamma}}\, H(x_0, 1+|( \mathcal{E}{\bf u})_{\theta\varrho}|)\, . \end{split} \end{equation} | (3.32) |
Choosing
\begin{equation} \sqrt{\varepsilon_1}\leq \frac{\theta^n}{8}\, , \end{equation} | (3.33) |
we obtain
\begin{equation} |( \mathcal{E}{\bf u})_{\theta\varrho}| \leq \frac{\sqrt{\varepsilon_1}}{8} |( \mathcal{E}{\bf u})_{\varrho}| < |( \mathcal{E}{\bf u})_{\varrho}| \, . \end{equation} | (3.34) |
Therefore, combining (3.30) and (3.33) we finally obtain
\begin{equation*} \begin{split} (\theta\varrho)^{\widetilde{\gamma}} \leq \left[\frac{{\mathit{\Phi}}(\varrho) }{H(x_0, 1+|( \mathcal{E}{\bf u})_{\varrho}|)} + \varrho^{\widetilde{\gamma}}\right]^2 \leq c \theta^2 \left[\frac{{\mathit{\Phi}}(\varrho) }{H(x_0, 1+|( \mathcal{E}{\bf u})_{\varrho}|)} + \varrho^{\widetilde{\gamma}}\right]\, . \end{split} \end{equation*} |
Inserting this estimate in (3.32) and using (3.34), we then obtain
\begin{equation} \begin{split} {\mathit{\Phi}}(\theta\varrho) \leq c \mathop {{\rlap{-} \smallint }}\limits_{B_{\theta\varrho}} H_{1+|( \mathcal{E}{\bf u})_{\theta\varrho}|}\left(x_0, \frac{|{\bf u}-{\mathit{\boldsymbol{\ell}}}_{\theta\varrho}|}{\theta\varrho}\right)\, \mathrm{d}x + c \theta^2 \left[{\mathit{\Phi}}(x_0, \varrho) + \varrho^{\widetilde{\gamma}}H(x_0, 1+|( \mathcal{E}{\bf u})_{x_0, \varrho}|)\right]\, . \end{split} \end{equation} | (3.35) |
Step 2: Following [12, Lemma 4.2], we prove that Theorem 2.10 and the subsequent remark can be applied to function
\begin{equation} {\bf v}: = \frac{{{\bf u} - {\mathit{\boldsymbol{\ell}}}_{\varrho}}}{1+|( \mathcal{E}{\bf u})_{\varrho}|}\, . \end{equation} | (3.36) |
Setting
\begin{equation} G(t): = \frac{H_{1+|( \mathcal{E}{\bf u})_{\varrho}|}(x_0, (1+|( \mathcal{E}{\bf u})_{\varrho}|)t)}{H(x_0, 1+|( \mathcal{E}{\bf u})_{\varrho}|)}\, , \end{equation} | (3.37) |
by using the fact that G is indeed a shifted N -function it can be seen that
\begin{equation} \tilde{c}G(t)\geq t^2\, , \quad t\in[0, 1]\, , \end{equation} | (3.38) |
for some constant \tilde{c} > 0 . Moreover, by Corollary 2 we have that
\begin{equation} \begin{split} \left(\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} \left[G\left(\frac{| \mathcal{E}({\bf u}-{\mathit{\boldsymbol{\ell}}}_{\varrho})|}{1+|( \mathcal{E}{\bf u})_{\varrho}|}\right)\right]^\sigma\, \mathrm{d}x \right)^{\frac{1}{\sigma}} & = \left(\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} \left[\frac{H_{1+|( \mathcal{E}{\bf u})_{\varrho}|}(x_0, | \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_{\varrho}|)}{H(x_0, 1+|( \mathcal{E}{\bf u})_{\varrho}|)}\right]^\sigma\, \mathrm{d}x \right)^{\frac{1}{\sigma}} \\ &\leq c_{\rm high} \frac{1}{H(x_0, 1+|( \mathcal{E}{\bf u})_{\varrho}|)}\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} H_{1+|( \mathcal{E}{\bf u})_{\varrho}|}(x_0, | \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_{\varrho}|)\, \mathrm{d}x \\ &\, \, \, \, \, \, \, \, \, \, + c_{\rm high}\varrho^{\widetilde{\gamma}} \\ & \leq c_{\rm high} \left [\frac{{\mathit{\Phi}}(\varrho)}{H(x_0, 1+|( \mathcal{E}{\bf u})_{\varrho}|)}+\varrho^{\widetilde{\gamma}}\right]\, . \end{split} \end{equation} | (3.39) |
Now, setting
\begin{equation} \varepsilon: = \max\{c_{\rm Stokes}, \sqrt{c_{\rm high} \tilde{c}}\}\left [\frac{{\mathit{\Phi}}(\varrho)}{H(x_0, 1+|( \mathcal{E}{\bf u})_{\varrho}|)}+\varrho^{\widetilde{\gamma}}\right]^{\frac{1}{2}} \leq \max\{c_{\rm Stokes}, \sqrt{c_{\rm high} \tilde{c}}\} \sqrt{\varepsilon_1}\, , \end{equation} | (3.40) |
we can choose \varepsilon_1 such that \varepsilon < 1 . Moreover, combining with (3.39) and (3.38), we get
\begin{equation} \begin{split} \left(\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} \left[G\left(\frac{| \mathcal{E}({\bf u}-{\mathit{\boldsymbol{\ell}}}_{\varrho})|}{1+|( \mathcal{E}{\bf u})_{\varrho}|}\right)\right]^\sigma\, \mathrm{d}x \right)^{\frac{1}{\sigma}} \leq c_{\rm high} \left [\frac{{\mathit{\Phi}}(\varrho)}{H(x_0, 1+|( \mathcal{E}{\bf u})_{\varrho}|)}+\varrho^{\widetilde{\gamma}}\right] & = c_{\rm high}\frac{\varepsilon^2}{\max\{c_{\rm Stokes}, \sqrt{c_{\rm high} \tilde{c}}\}^2} \\ &\leq G(\varepsilon) \, . \end{split} \end{equation} | (3.41) |
Inserting (3.40) and (3.30) into (3.21), written for {\bf v} and for some {\mathit{\boldsymbol{\varphi}}}\in C_{0, {\rm div}}^\infty(B_{\varrho}) with \|D{\mathit{\boldsymbol{\varphi}}}\|_{L^\infty(B_{\varrho})}\leq1 , we obtain
\begin{equation} \begin{split} \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}}\mathcal{A}\left(\frac{ \mathcal{E}({\bf u}-{\mathit{\boldsymbol{\ell}}}_{\varrho})}{1+|( \mathcal{E}{\bf u})_{\varrho}|}\right): \mathcal{E}{\mathit{\boldsymbol{\varphi}}}\, \mathrm{d}x \leq c_{\rm Stokes} \frac{[\omega(\varepsilon_1^\frac{1}{2})+\varepsilon_1]^{\frac{1}{2}}}{\max\{c_{\rm Stokes}, \sqrt{c_{\rm high} \tilde{c}}\}}\varepsilon\, . \end{split} \end{equation} | (3.42) |
Then, choosing \varepsilon_1 small enough, the assumptions of Theorem 2.10 and Remark 1 for function {\bf v} are in force. We denote by {{\bf h}} the \mathcal{A} -Stokes function in B_\varrho such that {{\bf h}} = {\bf v} on \partial B_\varrho . Assertion (2.26) gives
\begin{equation*} \begin{split} \frac{1}{H(x_0, 1+|( \mathcal{E}{\bf u})_{\varrho}|)}\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} H_{1+|( \mathcal{E}{\bf u})_{\varrho}|}(x_0, | \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_{\varrho}- (1+|( \mathcal{E}{\bf u})_{\varrho}|) \mathcal{E}{{\bf h}}|)\, \mathrm{d}x & = \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} G(| \mathcal{E}{\bf v}- \mathcal{E}{{\bf h}}|)\, \mathrm{d}x \\ & \leq \kappa G(\varepsilon)\, . \end{split} \label{eq:4.15celok} \end{equation*} |
Moreover, we also have
\begin{equation*} G(\varepsilon) \leq c \frac{H(x_0, (1+|( \mathcal{E}{\bf u})_{\varrho}|)(1+\varepsilon))}{H(x_0, 1+|( \mathcal{E}{\bf u})_{\varrho}|)(1+\varepsilon)^2}\varepsilon^2 \leq c \varepsilon^2\, , \end{equation*} |
whence
\begin{equation} \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}} H_{1+|( \mathcal{E}{\bf u})_{\varrho}|}(x_0, | \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_{\varrho}- (1+|( \mathcal{E}{\bf u})_{\varrho}|) \mathcal{E}{{\bf h}}|)\, \mathrm{d}x \leq \bar{c}\kappa \left[ {\mathit{\Phi}}(\varrho) + \varrho^{\widetilde{\gamma}} H(x_0, 1+|( \mathcal{E}{\bf u})_{\varrho}|) \right] \end{equation} | (3.43) |
for some constant \bar{c} > 0 .
Step 3: Now, we go back to the estimate of the right hand side of (3.35). As a consequence of Lemma 3.2, with (3.34) and (3.29), we get
\begin{equation*} \begin{split} &\mathop {{\rlap{-} \smallint }}\limits_{B_{\theta\varrho}} H_{1+|( \mathcal{E}{\bf u})_{\theta\varrho}|}\left(x_0, \frac{|{\bf u}-{\mathit{\boldsymbol{\ell}}}_{\theta\varrho}|}{\varrho}\right)\, \mathrm{d}x \\ & \leq c \left(1+ [\mu]_{C^\alpha}\| \mathcal{E}{\bf u}\|^{q-p}_{L^p(\Omega)}\theta^{\alpha-\frac{(q-p)n}{p}}\right)\mathop {{\rlap{-} \smallint }}\limits_{B_{\theta\varrho}} \left[H_{1+|( \mathcal{E}{\bf u})_{\theta\varrho}|}(x_0, | \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_{\theta\varrho}|)\right]\, \mathrm{d}x \\ & \leq c \left(1+ [\mu]_{C^\alpha}\| \mathcal{E}{\bf u}\|^{q-p}_{L^p(\Omega)}\theta^{\alpha-\frac{(q-p)n}{p}}\right)\mathop {{\rlap{-} \smallint }}\limits_{B_{\theta\varrho}} \left[H_{1+|( \mathcal{E}{\bf u})_{\varrho}|}(x_0, | \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_{\theta\varrho}|)\right]\, \mathrm{d}x\, . \end{split} \end{equation*} |
Now, the integral in the right hand side above can be treated exactly as in [12, pp. 24-25] starting from (3.43), to obtain the estimate}
\begin{equation} \mathop {{\rlap{-} \smallint }}\limits_{B_{\theta\varrho}} H_{1+|( \mathcal{E}{\bf u})_{\varrho}|}(x_0, | \mathcal{E}{\bf u}-( \mathcal{E}{\bf u})_{\theta\varrho}|)\, \mathrm{d}x \leq c \theta^2 \left[\frac{\kappa}{\theta^{n+2}}+1\right] \left [\frac{{\mathit{\Phi}}(\varrho)}{H(x_0, 1+|( \mathcal{E}{\bf u})_{\varrho}|)}+\varrho^{\widetilde{\gamma}}\right]\, . \end{equation} | (3.44) |
Thus, we omit further details. This concludes the proof of (3.31).
Now, we prove that the previous excess improvement estimate can be iterated at each scale. For this, we introduce the Morrey-type excess
\begin{equation} \Theta(x_0, \varrho): = \varrho^\frac{1}{2} [H(x_0, \cdot)]^{-1}\left(\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}}H(x_0, 1+|D {\bf u}|)\, \mathrm{d}x\right)\, . \end{equation} | (3.45) |
Lemma 3.6. Let {\mathit{\Phi}}(x_0, \varrho) and \Theta(x_0, \varrho) be defined as in (3.1) and (3.45), respectively. Then there exist constants \delta_* , \varepsilon_*, \varrho_*\in(0, 1] , M\geq1 and \vartheta such that the following holds: if the conditions
\begin{equation} \frac{{\mathit{\Phi}}(x_0, \varrho)}{H(x_0, 1+|( \mathcal{E}{\bf u})_{x_0, \varrho}|)}\leq \varepsilon_*\, , \qquad \Theta(x_0, \varrho)\leq\delta_*\, , \qquad |({\bf u})_{x_0, \varrho}|\leq\frac{1}{2}M \end{equation} | (3.46) |
hold on B_\varrho(x_0)\subseteq\Omega for \varrho\in(0, \varrho_*] , then
\begin{equation} \frac{{\mathit{\Phi}}(x_0, \vartheta^m\varrho)}{H(x_0, 1+|( \mathcal{E}{\bf u})_{x_0, \vartheta^m\varrho}|)}\leq \varepsilon_*\, , \qquad \Theta(x_0, \vartheta^m\varrho)\leq\delta_*\, , \qquad |({\bf u})_{x_0, \vartheta^m\varrho}|\leq M \end{equation} | (3.47) |
for every m = 0, 1, \dots. In particular, this would imply
\begin{equation} (\vartheta^m\varrho)^\frac{p}{2}(|D{\bf u}|^p)_{x_0, \vartheta^m\varrho} \leq \delta_*^p\, , \quad m = 0, 1, \dots\, . \end{equation} | (3.48) |
Moreover, for any \beta\in(0, 1) the following Morrey-type estimate holds:
\begin{equation} \Theta(y, r)\leq c\delta_*\left(\frac{r}{\varrho}\right)^\frac{\beta}{2} \end{equation} | (3.49) |
for all y\in B_{\varrho/2}(x_0) and r\in(0, \varrho/2] .
Proof. As usual, we omit the explicit dependence on x_0 . Let \vartheta\in(0, 1) be such that
\begin{equation} \vartheta\leq \min\left\{(8c_{\rm dec}2^{q-1})^{-\frac{1}{2}}, \frac{1}{2^{2q}}, \frac{1}{2^{\frac{q}{p (1-\beta)}}}\right\}\, , \end{equation} | (3.50) |
where c_{\rm dec} is the constant of Lemma 3.5 depending only on n, \nu, L, p, q, [\mu]_{C^\alpha}, \| \mathcal{E}{\bf u}\|_{L^p(\Omega)} . Correspondingly, let \varepsilon_1 = \varepsilon_1(n, \nu, L, p, q, [\mu]_{C^\alpha}, \| \mathcal{E}{\bf u}\|_{L^p(\Omega)}, \vartheta) be the constant of Lemma 3.5, applied with the choice \varepsilon = \vartheta^{n+2} . We choose \varepsilon_* > 0 such that
\begin{equation} \varepsilon_*\leq\min\left\{\frac{\varepsilon_1}{2}, \frac{\vartheta^n}{\max\{2c_{\frac{1}{2}}, 2^{q+1} c_{q}\}}\right\}\, , \end{equation} | (3.51) |
where c_{\frac{1}{2}} is the constant in the change-shift formula (2.12) with \eta = \frac{1}{2} , while c_q is obtained with \eta = \frac{1}{2^{q+1}} . Moreover, we choose a radius 0 < \varrho_*\leq1 such that
\begin{equation} \varrho_* < \min \left\{\varepsilon_*^\frac{1}{\widetilde{\gamma}}, \left(\frac{\vartheta^n (1-\vartheta^{p^*-\frac{p}{2}})}{2\delta_*^p}\right)^\frac{1}{p^*-\frac{p}{2}}\right\}\, . \end{equation} | (3.52) |
As a consequence, \varepsilon_* and \varrho_* have the same dependencies as \varepsilon_1 . We argue by induction on m . Since (3.47) are trivially true for m = 0 by assumption (3.46), our aim is to show that if (3.47) holds for some m\geq1 , then the corresponding inequalities hold with m+1 in place of m . Setting
\begin{equation*} E(B_{\vartheta^m\varrho}): = \mathop {{\rlap{-} \smallint }}\limits_{B_{\vartheta^m\varrho}}H(x_0, 1+|D {\bf u}|)\, \mathrm{d}x\, , \end{equation*} |
in order to prove the second inequalities in (3.47) it will suffice to show that
\begin{equation} E(B_{\vartheta^m\varrho})\leq H\left(x_0, \frac{\delta_*}{(\vartheta^m\varrho)^\frac{1}{2}}\right)\, . \end{equation} | (3.53) |
With (3.47) at step m , the shift-change formula (2.12) with \eta = \frac{1}{2} , (3.51) and Lemma 2.7, we have the estimate
\begin{equation} \begin{split} E(B_{\vartheta^{m+1}\varrho}) & \leq 2^{q-1}\left(\mathop {{\rlap{-} \smallint }}\limits_{B_{\vartheta^{m+1}\varrho}}H(x_0, |D {\bf u} - (D {\bf u})_{\vartheta^{m}\varrho}|)\, \mathrm{d}x + H(x_0, 1+|(D {\bf u})_{\vartheta^{m}\varrho}|)\right) \\ & \leq 2^{q-1}\left(c_{\frac{1}{2}}\mathop {{\rlap{-} \smallint }}\limits_{B_{\vartheta^{m+1}\varrho}}H_{1+|( \mathcal{E}{\bf u})_{\vartheta^{m}\varrho}|}(x_0, |D {\bf u} - (D {\bf u})_{\vartheta^{m}\varrho}|)\, \mathrm{d}x \right. \\ & \, \, \, \, \, \, \, \, \, \, \, \, \, \, \, \, \, \, \left.\vphantom{c_{\frac{1}{2}}\mathop {{\rlap{-} \smallint }}\limits_{B_{\vartheta^{m+1}\varrho}}H_{1+|( \mathcal{E}{\bf u})_{\vartheta^{m}\varrho}|}(x_0, |D {\bf u} - (D {\bf u})_{\vartheta^{m}\varrho}|)\, \mathrm{d}x}+\frac{1}{2}H(x_0, 1+|( \mathcal{E} {\bf u})_{\vartheta^{m}\varrho}|)+ H(x_0, 1+|(D {\bf u})_{\vartheta^{m}\varrho}|)\right) \\ & \leq 2^{q-1}\left(c_{\frac{1}{2}}\theta^{-n}\mathop {{\rlap{-} \smallint }}\limits_{B_{\vartheta^{m}\varrho}}H_{1+|( \mathcal{E}{\bf u})_{\vartheta^{m}\varrho}|}(x_0, |D {\bf u} - (D {\bf u})_{\vartheta^{m}\varrho}|)\, \mathrm{d}x +\frac{3}{2}E(B_{\vartheta^{m}\varrho}) \right) \\ & \leq 2^{q-1}\left(c_{\frac{1}{2}}\theta^{-n}{\mathit{\Phi}}(\vartheta^m\varrho) +\frac{3}{2}E(B_{\vartheta^{m}\varrho}) \right)\\ & \leq 2^{q-1}\left(c_{\frac{1}{2}}\vartheta^{-n}\varepsilon_*+\frac{3}{2}\right) E(B_{\vartheta^m\varrho})\\ &\leq 2^{q-1}\left(c_{\frac{1}{2}}\vartheta^{-n}\varepsilon_*+\frac{3}{2}\right)\vartheta^\frac{1}{2} H\left(x_0, \frac{\delta_*}{(\vartheta^{m+1}\varrho)^\frac{1}{2}}\right)\\ & \leq H\left(x_0, \frac{\delta_*}{(\vartheta^{m+1}\varrho)^\frac{1}{2}}\right)\, , \end{split} \end{equation} | (3.54) |
which proves (3.47)_2 for m+1 . With this at hand, since the function G(t^\frac{1}{p}): = H(x_0, t^\frac{1}{p}) is convex, from Jensen's inequality we get
\begin{equation*} G\left(\left(\mathop {{\rlap{-} \smallint }}\limits_{B_{\vartheta^i\varrho}}|D{\bf u}|^p \, \mathrm{d}x\right)^\frac{1}{p}\right) \leq \mathop {{\rlap{-} \smallint }}\limits_{B_{\vartheta^mi\varrho}}G(1+|D{\bf u}|)\, \mathrm{d}x \leq G\left(\frac{\delta_*}{(\vartheta^i\varrho)^\frac{1}{2}}\right)\, , \end{equation*} |
whence, passing to G^{-1} , we obtain (3.48). Now, we prove by induction the first inequality in (3.47) for m+1 . From (3.47) at step k and the choice of \varrho_* as in (3.52), we have
\begin{equation*} \begin{split} \frac{{\mathit{\Phi}}(\vartheta^m\varrho)}{H(x_0, 1+|( \mathcal{E}{\bf u})_{\vartheta^m\varrho}|)} & \leq \varepsilon_* < 2\varepsilon_*\leq\varepsilon_1\, , \\ (\vartheta^m\varrho)^{\widetilde{\gamma}} & < 2\varepsilon_*\leq \varepsilon_1\, , \end{split} \end{equation*} |
and
\begin{equation*} \begin{split} {\mathit{\Phi}}(\vartheta^m\varrho) + (\vartheta^m\varrho)^{\widetilde{\gamma}} H(x_0, 1+|( \mathcal{E}{\bf u})_{\vartheta^m\varrho}|) & \leq 4 \epsilon_* H(x_0, 1+|( \mathcal{E}{\bf u})_{\vartheta^m\varrho}|)\, . \end{split} \end{equation*} |
Then, by virtue of Lemma 3.5 and Lemma 3.4 applied with radius \vartheta^m\varrho in place of \varrho , and recalling the choice of \vartheta (3.50), we get
\begin{equation*} \begin{split} {\mathit{\Phi}}(\vartheta^{m+1}\varrho) &\leq 2c_{\rm dec}\vartheta^2[{\mathit{\Phi}}(\vartheta^m\varrho) + (\vartheta^m\varrho)^{\widetilde{\gamma}} H(x_0, 1+|( \mathcal{E}{\bf u})_{\vartheta^m\varrho}|)] \\ &\leq 8c_{\rm dec}\epsilon_*\vartheta^2H(x_0, 1+|( \mathcal{E}{\bf u})_{\vartheta^m\varrho}|)\\ & \leq \epsilon_* H(x_0, 1+|( \mathcal{E}{\bf u})_{\vartheta^{m+1}\varrho}|)\, . \end{split} \end{equation*} |
To conclude the proof, we are left to prove (3.47)_3 and (3.49). We use (3.46)_3 , the Poincaré inequality with (3.47)_2 , \varrho\leq \varrho_* and (3.52), (3.48) for every i = 0, \dots, m and we get
\begin{equation} \begin{split} |({\bf u})_{\vartheta^{m+1}\varrho}| & \leq |({\bf u})_{\varrho}| + \vartheta^{-n} \sum\limits_{i = 0}^m\mathop {{\rlap{-} \smallint }}\limits_{B_{\vartheta^i\varrho}} |{\bf u}-({\bf u})_{\vartheta^i\varrho}|^{p^*} \, \mathrm{d}x \\ & \leq \frac{1}{2} M + \vartheta^{-n} \sum\limits_{i = 0}^m (\vartheta^i\varrho)^{p^*} \mathop {{\rlap{-} \smallint }}\limits_{B_{\vartheta^i\varrho}}|D{\bf u}|^p \, \mathrm{d}x \\ & \leq \frac{1}{2} M + \delta_*^p\vartheta^{-n} \sum\limits_{i = 0}^m (\vartheta^i\varrho)^{p^*-\frac{p}{2}} \\ & = \frac{1}{2} M + \delta_*^{p}\vartheta^{-n} \frac{\varrho^{p^*-\frac{p}{2}}}{1-\vartheta^{p^*-\frac{p}{2}}} \\ & \leq \frac{1}{2} M + \frac{1}{2}\leq M\, . \end{split} \end{equation} | (3.55) |
Finally, setting G(t): = H(x_0, t) , since the iteration starting from m = 0 of the estimate G^{-1}(E(B_{\vartheta^{m+1}\varrho}))\leq 2^{\frac{q}{p}}G^{-1}(E(B_{\vartheta^{m}\varrho})) , obtained by (3.54) and (2.7), with (3.50) yields
\begin{equation*} (\vartheta^m\varrho)^\frac{1-\beta}{2}G^{-1}(E(B_{\vartheta^m\varrho}))\leq \varrho^\frac{1-\beta}{2}G^{-1}(E(B_{\varrho}))\leq \delta_*\varrho^{-\frac{\beta}{2}}\, , \end{equation*} |
and this estimate a fortiori holds if we consider E(B_{\vartheta^m\varrho}(y)) for y\in B_{\varrho/2} in place of E(B_{\vartheta^m\varrho}) , we deduce the Morrey-type estimate
\begin{equation*} r^\frac{1-\beta}{2}G^{-1}(E(B_{r}(y)))\leq c\delta_*\varrho^{-\frac{\beta}{2}} \end{equation*} |
for all y\in B_{\varrho/2} and r\leq\varrho/2 , which is equivalent to (3.49). This concludes the proof.
We are now in position to prove Theorem 1.1.
Proof. Let \delta_* , \varepsilon_*, \varrho_*\in(0, 1] be the constants of Lemma 3.6. We define
\begin{equation*} \Omega_0: = \left\{z_0\in\Omega:\, \, {\bf u}\in C^\beta(U_{z_0}; {\mathbb{R}}^n) \mbox{ for every $\beta\in(0, 1)$ and for some }U_{z_0}\subset\Omega\right\}\, , \end{equation*} |
where U_{z_0} is an open neighborhood of z_0 . Assuming that x_0\in\Omega complies with
\begin{equation} \begin{split} &\lim\limits_{\varrho\searrow0} \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}(x_0)}|D{\bf u}-(D{\bf u})_{x_0, \varrho}|\, \mathrm{d}x = 0\, , \\ M_{x_0}: = &\mathop{\lim\sup}_{\varrho\searrow 0}\left[\mathop {{\rlap{-} \smallint }}\limits_{B_\varrho(x_0)}H(x, 1+| \mathcal{E}{\bf u}|)\, \mathrm{d}x + (|D{\bf u}|^p)_{x_0, \varrho} + |({\bf u})_{x_0, \varrho}|\right] < +\infty\, ; \end{split} \end{equation} | (4.1) |
i.e., x_0\in\Omega\setminus(\Sigma_1\cup\Sigma_2) , we will prove that x_0\in\Omega_0 . We fix \beta\in(0, 1) and choose t\in(0, 1) such that
\begin{equation} \frac{1}{q} = t + \frac{1-t}{qs_0}\, , \end{equation} | (4.2) |
where s_0 > 1 is the exponent of Lemma 2.11. We also set
\begin{equation} \varepsilon\leq\min \left\{\left[\frac{1}{c_*}\left(\frac{\varepsilon_*}{2}\right)^\frac{p}{2}\frac{1}{\bar{c}(M_{x_0}+1)^{\frac{q(1-t)}{p}}}\right]^\frac{1}{pt}, 1 \right\}\, , \end{equation} | (4.3) |
where c_* is the constant c_\eta in the change of shift formula for \eta_*: = \frac{\varepsilon_*}{2^{n+1} (M_{x_0}+1) c_H} . By virtue of (4.1), we can find 0 < \varrho with
\begin{equation} \varrho \leq \min\left\{\frac{\delta_*}{[H(x_0, \cdot)]^{-1}(\tilde{c}^\frac{1}{s_0}(M_{x_0}+1)^{q}) }, \varrho_*\right\}\, . \end{equation} | (4.4) |
and B_{3\varrho}(x_0)\subset\subset\Omega such that
\begin{equation} \begin{split} &\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}(x_0)}|D{\bf u}-(D{\bf u})_{x_0, \varrho}|\, \mathrm{d}x < \varepsilon\, , \\ & \mathop {{\rlap{-} \smallint }}\limits_{B_{2\varrho}(x_0)}H(x, 1+| \mathcal{E}{\bf u}|)\, \mathrm{d}x + (|D{\bf u}|^p)_{x_0, \varrho} + |({\bf u})_{x_0, \varrho}| < M_{x_0}+1\, . \end{split} \end{equation} | (4.5) |
Step 1: We first establish the higher integrability of H(x_0, 1+|D{\bf u}|) on B_\varrho(x_0) . Namely, we show that
\begin{equation} \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}(x_0)}[H(x_0, 1+|D{\bf u}|)]^{s_0}\, \mathrm{d}x \leq \tilde{c} (M_{x_0}+1)^{qs_0} \, , \end{equation} | (4.6) |
where s_0 > 1 is the exponent of Lemma 2.11.
If \mu satisfies (2.36), since (4.5) implies (2.38), from (2.39) we infer
\begin{equation*} \begin{split} \left(\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}(x_0)}[H(x, 1+| \mathcal{E}{\bf u}|)]^{s_0}\, \mathrm{d}x\right)^{\frac{1}{s_0}} & \leq c \mathop {{\rlap{-} \smallint }}\limits_{B_{2\varrho}(x_0)}H(x, 1+| \mathcal{E}{\bf u}|)\, \mathrm{d}x \\ & \leq c (M_{x_0}+1)\, , \end{split} \end{equation*} |
whence
\begin{equation} \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}(x_0)}[H(x, 1+| \mathcal{E}{\bf u}|)]^{s_0}\, \mathrm{d}x \leq c(M_{x_0}+1)^{s_0}\, . \end{equation} | (4.7) |
Now, from Korn's inequality, Poincaré inequality, (4.5), (2.39) and \varrho\leq1 we have
\begin{equation*} \begin{split} \mathop {{\rlap{-} \smallint }}\limits_{B_\varrho(x_0)} (1+|D{\bf u}|)^{ps_0}\, \mathrm{d}x & \leq c \mathop {{\rlap{-} \smallint }}\limits_{B_\varrho(x_0)} (1+| \mathcal{E}{\bf u}|)^{ps_0}\, \mathrm{d}x + c \left(\mathop {{\rlap{-} \smallint }}\limits_{B_\varrho(x_0)} \left |\frac{{\bf u} - ({\bf u})_{x_0, \varrho}}{\varrho}\right|\, \mathrm{d}x\right)^{ps_0} \\ & \leq c \mathop {{\rlap{-} \smallint }}\limits_{B_\varrho(x_0)} (1+| \mathcal{E}{\bf u}|)^{ps_0}\, \mathrm{d}x + c \left(\mathop {{\rlap{-} \smallint }}\limits_{B_\varrho(x_0)} |D{\bf u}|\, \mathrm{d}x\right)^{ps_0} \\ & \leq c \left(\mathop {{\rlap{-} \smallint }}\limits_{B_{2\varrho}(x_0)} (1+| \mathcal{E}{\bf u}|)^{p}\, \mathrm{d}x \right)^{s_0} + c(M_{x_0}+1)^{s_0}\, . \end{split} \end{equation*} |
Analogously, using also Hölder's inequality and \varrho\leq1 , we have
\begin{equation*} \begin{split} \mathop {{\rlap{-} \smallint }}\limits_{B_\varrho(x_0)} [\mu(x_0)(1+|D{\bf u}|)^q]^{s_0}\, \mathrm{d}x & \leq c\varrho^{\alpha s_0}\left [\mathop {{\rlap{-} \smallint }}\limits_{B_\varrho(x_0)} (1+| \mathcal{E}{\bf u}|)^{qs_0}\, \mathrm{d}x + \left(\mathop {{\rlap{-} \smallint }}\limits_{B_\varrho(x_0)} |D{\bf u}|\, \mathrm{d}x\right)^{qs_0} \right] \\ & \leq c\varrho^{\alpha s_0}\left [\left(\mathop {{\rlap{-} \smallint }}\limits_{B_{2\varrho}(x_0)} (1+| \mathcal{E}{\bf u}|)^{p}\, \mathrm{d}x\right)^\frac{qs_0}{p} + \left(\mathop {{\rlap{-} \smallint }}\limits_{B_\varrho(x_0)} |D{\bf u}|\, \mathrm{d}x\right)^{qs_0} \right] \\ & \leq c\varrho^{(\alpha-\frac{(q-p)}{p}) s_0}\|1+| \mathcal{E}{\bf u}|\|_{L^p}^{q-p}\left(\mathop {{\rlap{-} \smallint }}\limits_{B_{2\varrho}(x_0)} (1+| \mathcal{E}{\bf u}|)^{p}\, \mathrm{d}x\right)^{s_0} \\ & \, \, \, \, \, \, + c\varrho^{\alpha s_0} \left(\mathop {{\rlap{-} \smallint }}\limits_{B_\varrho(x_0)} |D{\bf u}|\, \mathrm{d}x\right)^{qs_0} \\ & \leq c \left(\mathop {{\rlap{-} \smallint }}\limits_{B_{2\varrho}(x_0)} (1+| \mathcal{E}{\bf u}|)^{p}\, \mathrm{d}x\right)^{s_0} + c (M_{x_0}+1)^\frac{qs_0}{p} \, . \end{split} \end{equation*} |
Combining the previous estimates, we then obtain
\begin{equation} \begin{split} \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}(x_0)}[H(x_0, 1+|D{\bf u}|)]^{s_0}\, \mathrm{d}x & \leq c \left(\mathop {{\rlap{-} \smallint }}\limits_{B_{2\varrho}(x_0)}H(x, 1+| \mathcal{E}{\bf u}|)\, \mathrm{d}x\right)^{s_0} \\ & \leq c_1 (M_{x_0}+1)^\frac{qs_0}{p} \, . \end{split} \end{equation} | (4.8) |
If (2.36) doesn't hold true, then it can be shown that H(x_0, t)\leq c H(x, t) . In this case, with (4.5) and (2.28), (2.43), we get
\begin{equation} \left(\mathop {{\rlap{-} \smallint }}\limits_{B_{r}(x_0)}[H(x, 1+| \mathcal{E}{\bf u}|)]^{s_0}\, \mathrm{d}x\right)^{\frac{1}{s_0}} \leq c \mathop {{\rlap{-} \smallint }}\limits_{B_{2r}(x_0)}H(x, 1+| \mathcal{E}{\bf u}|)\, \mathrm{d}x\, , \end{equation} | (4.9) |
whence assertion (4.7) can still be inferred. Now, using Korn's inequality, Poincaré inequality, (4.5), (4.9) we have
\begin{equation*} \begin{split} \mathop {{\rlap{-} \smallint }}\limits_{B_\varrho(x_0)} (1+|D{\bf u}|)^{ps_0}\, \mathrm{d}x & \leq c \mathop {{\rlap{-} \smallint }}\limits_{B_\varrho(x_0)} (1+| \mathcal{E}{\bf u}|)^{ps_0}\, \mathrm{d}x + c \left(\mathop {{\rlap{-} \smallint }}\limits_{B_\varrho(x_0)} |D{\bf u}|\, \mathrm{d}x\right)^{ps_0} \\ & \leq c \left(\mathop {{\rlap{-} \smallint }}\limits_{B_{2\varrho}(x_0)} H(x, 1+| \mathcal{E}{\bf u}|)\, \mathrm{d}x \right)^{s_0} + c(M_{x_0}+1)^{s_0}\, . \end{split} \end{equation*} |
In a similar way, using also that \mu\leq1 , we have
\begin{equation*} \begin{split} \mathop {{\rlap{-} \smallint }}\limits_{B_\varrho(x_0)} [\mu(x_0)(1+|D{\bf u}|)^q]^{s_0}\, \mathrm{d}x & \leq c\left [\mathop {{\rlap{-} \smallint }}\limits_{B_\varrho(x_0)} [\mu(x_0)(1+| \mathcal{E}{\bf u}|)^q]^{s_0}\, \mathrm{d}x + \left(\mathop {{\rlap{-} \smallint }}\limits_{B_\varrho(x_0)} |D{\bf u}|\, \mathrm{d}x\right)^{qs_0} \right] \\ & \leq c \left(\mathop {{\rlap{-} \smallint }}\limits_{B_{2\varrho}(x_0)} H(x_0, 1+| \mathcal{E}{\bf u}|)\, \mathrm{d}x\right)^{s_0} + c (M_{x_0}+1)^{qs_0}\\ & \leq c \left(\mathop {{\rlap{-} \smallint }}\limits_{B_{2\varrho}(x_0)} H(x, 1+| \mathcal{E}{\bf u}|)\, \mathrm{d}x\right)^{s_0} + c (M_{x_0}+1)^{qs_0}\, , \end{split} \end{equation*} |
which combined with the previous one gives
\begin{equation} \begin{split} \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}(x_0)}[H(x_0, 1+|D{\bf u}|)]^{s_0}\, \mathrm{d}x & \leq c \left(\mathop {{\rlap{-} \smallint }}\limits_{B_{2\varrho}(x_0)}H(x, 1+| \mathcal{E}{\bf u}|)\, \mathrm{d}x\right)^{s_0} \\ & \leq c_2 (M_{x_0}+1)^{qs_0} \, . \end{split} \end{equation} | (4.10) |
Assertion (4.6) then follows combining (4.8) and (4.10) and choosing \tilde{c}: = \max\{c_1, c_2\} .
Step 2: From Hölder's inequality and the choice of t in (4.2) we get
\begin{equation*} \begin{split} \mathop {{\rlap{-} \smallint }}\limits_{B_\varrho(x_0)} H(x_0, |D{\bf u}-(D{\bf u})_{x_0, \varrho}|)\, \mathrm{d}x & \leq \left(\mathop {{\rlap{-} \smallint }}\limits_{B_\varrho(x_0)} [H(x_0, |D{\bf u}-(D{\bf u})_{x_0, \varrho}|)]^\frac{1}{q}\, \mathrm{d}x\right)^{tq} \\ & \, \, \, \, \times \left(\mathop {{\rlap{-} \smallint }}\limits_{B_\varrho(x_0)} [H(x_0, |D{\bf u}-(D{\bf u})_{x_0, \varrho}|)]^{s_0}\, \mathrm{d}x\right)^\frac{1-t}{s_0}\, . \end{split} \end{equation*} |
Now, using Jensen's inequality for the concave function \tilde{\Psi} such that \frac{1}{2}\tilde{\Psi}(t)\leq \Psi(t) : = [H(x_0, t)]^\frac{1}{q} \leq \tilde{\Psi}(t) (see Lemma 2.4) and (4.5), we have
\begin{equation*} \mathop {{\rlap{-} \smallint }}\limits_{B_\varrho(x_0)} [H(x_0, |D{\bf u}-(D{\bf u})_{x_0, \varrho}|)]^\frac{1}{q}\, \mathrm{d}x \leq 2 [H(x_0, \varepsilon)]^\frac{1}{q} \leq 2[(1+\|\mu\|_\infty)\varepsilon^p]^\frac{1}{q}\, . \end{equation*} |
On the other hand, using Jensen's inequality for the convex map t\to [H(x_0, t)]^{s_0} we obtain
\begin{equation*} \mathop {{\rlap{-} \smallint }}\limits_{B_\varrho(x_0)} [H(x_0, |D{\bf u}-(D{\bf u})_{x_0, \varrho}|)]^{s_0}\, \mathrm{d}x \leq \mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}(x_0)}[H(x_0, 1+|D{\bf u}|)]^{s_0}\, \mathrm{d}x\, . \end{equation*} |
Combining the previous estimates with (4.6) we finally get
\begin{equation*} \mathop {{\rlap{-} \smallint }}\limits_{B_\varrho(x_0)} H(x_0, |D{\bf u}-(D{\bf u})_{x_0, \varrho}|)\, \mathrm{d}x \leq \bar{c}\varepsilon^{tp} (M_{x_0}+1)^{\frac{q(1-t)}{p}}\, , \end{equation*} |
for a constant \bar{c} = \bar{c}(n, \nu, L, p, q, [\mu]_{C^\alpha}, \|\mu\|_{L^\infty}, \|1+| \mathcal{E}{\bf u}|\|_{L^p}) . With the choice of \varepsilon in (4.3), this implies
\begin{equation*} \mathop {{\rlap{-} \smallint }}\limits_{B_\varrho(x_0)} H(x_0, |D{\bf u}-(D{\bf u})_{x_0, \varrho}|)\, \mathrm{d}x \leq \frac{1}{c_*}\left(\frac{\varepsilon_*}{2}\right)^\frac{p}{2}\, . \end{equation*} |
Now, using the change of shift formula with \eta_*: = \frac{\varepsilon_*}{2^{n+1} (M_{x_0}+1) c_H} and H(x_0, 1)\geq1 ,
\begin{equation*} \begin{split} \frac{{\mathit{\Phi}}(x_0, \varrho)}{H(x_0, 1+|( \mathcal{E}{\bf u})_{x_0, \varrho}|)} & \leq \frac{1}{H(x_0, 1)}\mathop {{\rlap{-} \smallint }}\limits_{B_\varrho(x_0)}H_{1+|( \mathcal{E}{\bf u})_{x_0, \varrho}|}(x_0, | \mathcal{E}{\bf u} - ( \mathcal{E}{\bf u})_{x_0, \varrho}|)\, \mathrm{d}x \\ & \leq \mathop {{\rlap{-} \smallint }}\limits_{B_\varrho(x_0)}H_{1+|( \mathcal{E}{\bf u})_{x_0, \varrho}|}(x_0, | \mathcal{E}{\bf u} - ( \mathcal{E}{\bf u})_{x_0, \varrho}|)\, \mathrm{d}x \\ & \leq c_* \mathop {{\rlap{-} \smallint }}\limits_{B_\varrho(x_0)}H(x_0, | \mathcal{E}{\bf u} - ( \mathcal{E}{\bf u})_{x_0, \varrho}|)\, \mathrm{d}x + \eta_* H_{1+|( \mathcal{E}{\bf u})_{x_0, \varrho}|}(x_0, 1+|( \mathcal{E}{\bf u})_{x_0, \varrho}|) \\ & \leq c_* \mathop {{\rlap{-} \smallint }}\limits_{B_\varrho(x_0)}H(x_0, |D{\bf u} - (D{\bf u})_{x_0, \varrho}|)\, \mathrm{d}x + c_H \eta_* H(x_0, 1+|( \mathcal{E}{\bf u})_{x_0, \varrho}|) \\ & \leq \left(\frac{\varepsilon_*}{2}\right)^\frac{p}{2} + c_H \eta_* 2^n\mathop {{\rlap{-} \smallint }}\limits_{B_{2\varrho}(x_0)}H(x_0, 1+| \mathcal{E}{\bf u}|)\, \mathrm{d}x \\ & \leq \left(\frac{\varepsilon_*}{2}\right)^\frac{p}{2} + \left(\frac{\varepsilon_*}{2}\right) \leq \varepsilon_* \, . \end{split} \end{equation*} |
Moreover, with (4.6) and (4.4) we have
\begin{equation*} \begin{split} \Theta(x_0, \varrho) & = \varrho^\frac{1}{2} [H(x_0, \cdot)]^{-1}\left(\mathop {{\rlap{-} \smallint }}\limits_{B_{\varrho}}H(x_0, 1+|D {\bf u}|)\, \mathrm{d}x\right) \\ & \leq \varrho^\frac{1}{2} [H(x_0, \cdot)]^{-1}(\tilde{c}^\frac{1}{s_0}(M_{x_0}+1)^{q}) \\ & \leq \delta_*\, . \end{split} \end{equation*} |
By the absolute continuity of the integral, we can find an open neighborhood U_{x_0} of x_0 such that
\begin{equation*} \frac{{\mathit{\Phi}}(x, \varrho)}{H(x_0, 1+|( \mathcal{E}{\bf u})_{x_0, \varrho}|)} < \varepsilon_*\quad \mbox{ and }\quad \Theta(x, {\varrho}) < \delta_* \end{equation*} |
for every x\in U_{x_0} . Then, taking into account also (4.1), we can apply Lemma 3.6 at each point of U_{x_0} . This provides a Morrey-type estimate as in (3.49) proving that {\bf u}\in C^{0, \beta}(U_{x_0}, {\mathbb{R}}^n) for every \beta\in(0, 1) (see, e.g., [27, Lemma 3.15]). Thus, x_0\in\Omega_0 and the proof is concluded.
The authors are members of Gruppo Nazionale per l'Analisi Matematica, la Probabilità e le loro Applicazioni (GNAMPA) of INdAM. G. Scilla and B. Stroffolini have been supported by the project STAR PLUS 2020 – Linea 1 (21‐UNINA‐EPIG‐172) "New perspectives in the Variational modeling of Continuum Mechanics".
The authors declare no conflict of interest.
[1] |
M. Abdelwahed, L. C. Berselli, N. Chorfi, On the uniqueness for weak solutions of steady double-phase fluids, Adv. Nonlinear Anal., 11 (2022), 454–468. https://doi.org/10.1515/anona-2020-0196 doi: 10.1515/anona-2020-0196
![]() |
[2] |
E. Acerbi, G. Mingione, Regularity results for stationary electro-rheological fluids, Arch. Rational Mech. Anal., 164 (2002), 213–259. https://doi.org/10.1007/s00205-002-0208-7 doi: 10.1007/s00205-002-0208-7
![]() |
[3] | R. A. Adams, Sobolev spaces, New York-London: Academic Press, 1975. |
[4] |
P. Baroni, M. Colombo, G. Mingione, Nonautonomous functionals, borderline cases and related function classes, St. Petersburg Math. J., 27 (2016), 347–379. https://doi.org/10.1090/spmj/1392 doi: 10.1090/spmj/1392
![]() |
[5] |
P. Bella, M. Schäffner, On the regularity of minimizers for scalar integral functionals with (p, q)-growth, Anal. PDE, 13 (2020), 2241–2257. https://doi.org/10.2140/apde.2020.13.2241 doi: 10.2140/apde.2020.13.2241
![]() |
[6] | C. Bennett, R. Sharpley, Interpolation of operators, Boston, MA: Academic Press Inc., 1988. |
[7] |
V. Bögelein, F. Duzaar, J. Habermann, C. Scheven, Stationary electro-rheological fluids: low order regularity for systems with discontinuous coefficients, Adv. Calc. Var., 5 (2012), 1–57. https://doi.org/10.1515/acv.2011.009 doi: 10.1515/acv.2011.009
![]() |
[8] | M. E. Bogovskiǐ, Solutions of some problems of vector analysis, associated with the operators div and grad, (Russian), In: Theory of cubature formulas and the application of functional analysis to problems of mathematical physics, Novosibirsk: Akad. Nauk SSSR Sibirsk. Otdel., Inst. Mat., 1980, 5–40. |
[9] |
D. Breit, L. Diening, Sharp conditions for Korn inequalities in Orlicz spaces, J. Math. Fluid Mech., 14 (2012), 565–573. https://doi.org/10.1007/s00021-011-0082-x doi: 10.1007/s00021-011-0082-x
![]() |
[10] |
D. Breit, L. Diening, M. Fuchs, Soleinodal Lipschitz truncations and application in fluid mechanics, J. Differ. Equations, 253 (2012), 1910–1942. https://doi.org/10.1016/j.jde.2012.05.010 doi: 10.1016/j.jde.2012.05.010
![]() |
[11] |
M. Carozza, J. Kristensen, A. Passarelli di Napoli, Regularity of minimizers of autonomous convex variational integrals, Ann. Sc. Norm. Super. Pisa Cl. Sci. (5), 13 (2014), 1065–1089. https://doi.org/10.2422/2036-2145.201208_005 doi: 10.2422/2036-2145.201208_005
![]() |
[12] |
P. Celada, J. Ok, Partial regularity for non-autonomous degenerate quasi-convex functionals with general growth, Nonlinear Anal., 194 (2020), 111473. https://doi.org/10.1016/j.na.2019.02.026 doi: 10.1016/j.na.2019.02.026
![]() |
[13] |
M. Colombo, G. Mingione, Regularity for double phase variational problems, Arch. Rational Mech. Anal., 215 (2015), 443–496. https://doi.org/10.1007/s00205-014-0785-2 doi: 10.1007/s00205-014-0785-2
![]() |
[14] |
C. De Filippis, Quasiconvexity and partial regularity via nonlinear potentials, J. Math. Pure. Appl., 163 (2022), 11–82. https://doi.org/10.1016/j.matpur.2022.05.001 doi: 10.1016/j.matpur.2022.05.001
![]() |
[15] |
C. De Filippis, F. Leonetti, Uniform ellipticity and (p, q)-growth., J. Math. Anal. Appl., 501 (2021), 124451. https://doi.org/10.1016/j.jmaa.2020.124451 doi: 10.1016/j.jmaa.2020.124451
![]() |
[16] |
C. De Filippis, B. Stroffolini, Singular multiple integrals and nonlinear potentials, J. Funct. Anal., 285 (2023), 109952. https://doi.org/10.1016/j.jfa.2023.109952 doi: 10.1016/j.jfa.2023.109952
![]() |
[17] |
L. Diening, F. Ettwein, Fractional estimates for non-differentiable elliptic systems with general growth, Forum Math., 20 (2008), 523–556. https://doi.org/10.1515/FORUM.2008.027 doi: 10.1515/FORUM.2008.027
![]() |
[18] |
L. Diening, P. Kaplicky, S. Schwarzacher, Campanato estimates for the generalized Stokes system, Annali di Matematica, 193 (2014), 1779–1794. https://doi.org/10.1007/s10231-013-0355-5 doi: 10.1007/s10231-013-0355-5
![]() |
[19] |
L. Diening, C. Kreuzer, Linear convergence of an adaptive finite element method for the p-laplacian equation, SIAM J. Numer. Anal., 46 (2008), 614–638. https://doi.org/10.1137/070681508 doi: 10.1137/070681508
![]() |
[20] |
L. Diening, D. Lengeler, B. Stroffolini, A. Verde, Partial regularity for minimizers of quasiconvex functionals with general growth, SIAM J. Math. Anal., 44 (2012), 3594–3616. https://doi.org/10.1137/120870554 doi: 10.1137/120870554
![]() |
[21] |
L. Esposito, F. Leonetti, G. Mingione, Sharp regularity for functionals with (p, q) growth, J. Differ. Equations, 204 (2004), 5–55. https://doi.org/10.1016/j.jde.2003.11.007 doi: 10.1016/j.jde.2003.11.007
![]() |
[22] |
J. Frehse, J. Málek, M. Steinhauer, On analysis of steady flows of fluids with shear-dependent viscosity based on the Lipschitz truncation method, SIAM J. Math. Anal., 34 (2003), 1064–1083. https://doi.org/10.1137/S0036141002410988 doi: 10.1137/S0036141002410988
![]() |
[23] | M. Fuchs, G. Seregin, Variational methods for problems from plasticity theory and for generalized Newtonian fluids, Berlin, Heidelberg: Spinger, 2000. https://doi.org/10.1007/BFb0103751 |
[24] | G. P. Galdi, An introduction to the mathematical theory of the Navier-Stokes equations, Volume I: Linearised steady problems, New York, NY: Springer, 1994. https://doi.org/10.1007/978-1-4757-3866-7 |
[25] |
F. W. Gehring, The L^p-integrability of the partial derivatives of a quasiconformal mapping, Acta Math., 130 (1973), 265–277. https://doi.org/10.1007/BF02392268 doi: 10.1007/BF02392268
![]() |
[26] | E. Giusti, Direct methods in the calculus of variations, World Scientific, 2003. https://doi.org/10.1142/5002 |
[27] |
C. Goodrich, G. Scilla, B. Stroffolini, Partial regularity for minimizers of discontinuous quasiconvex integrals with general growth, Proc. Roy. Soc. Edinb. A, 152 (2022), 1191–1232. https://doi.org/10.1017/prm.2021.53 doi: 10.1017/prm.2021.53
![]() |
[28] | M. A. Krasnosel'ski\mathop {\rm{i}}\limits^ \vee , Ya. B. Ruticki\mathop {\rm{i}}\limits^ \vee , Convex functions and orlicz spaces, Groningen: P. Noordhoff LTD., 1961. |
[29] |
M. Křepela, M. Růžička, Solenoidal difference quotients and their application to the regularity theory of the p-Stokes system, Calc. Var., 59 (2020), 34. https://doi.org/10.1007/s00526-019-1691-0 doi: 10.1007/s00526-019-1691-0
![]() |
[30] | A. Kufner, O. John, S. Fuc\mathop {\rm{i}}\limits^ \vee k, Function spaces, Dordrecht: Springer, 1977. |
[31] |
P. Marcellini, Regularity and existence of solutions of elliptic equations with p, q-growth conditions, J. Differ. Equations, 90 (1991), 1–30. https://doi.org/10.1016/0022-0396(91)90158-6 doi: 10.1016/0022-0396(91)90158-6
![]() |
[32] |
P. Marcellini, Local Lipschitz continuity for p, q-PDEs with explicit u-dependence, Nonlinear Anal., 226 (2022), 113066. https://doi.org/10.1016/j.na.2022.113066 doi: 10.1016/j.na.2022.113066
![]() |
[33] |
G. Mingione, Regularity of minima: an invitation to the dark side of the calculus of variations, Appl. Math., 51 (2006), 355–426. https://doi.org/10.1007/s10778-006-0110-3 doi: 10.1007/s10778-006-0110-3
![]() |
[34] |
G. Mingione, V. Rǎdulescu, Recent developments in problems with nonstandard growth and nonuniform ellipticity, J. Math. Anal. Appl., 501 (2021), 125197. https://doi.org/10.1016/j.jmaa.2021.125197 doi: 10.1016/j.jmaa.2021.125197
![]() |
[35] |
P. P. Mosolov, V. P. Mjasnikov, On the correctness of boundary value problems in the mechanics of continuous media, Math. USSR Sb., 17 (1972), 257. https://doi.org/10.1070/SM1972v017n02ABEH001503 doi: 10.1070/SM1972v017n02ABEH001503
![]() |
[36] |
J. Ok, Partial Hölder regularity for elliptic systems with non-standard growth, J. Funct. Anal., 274 (2018), 723–768. https://doi.org/10.1016/j.jfa.2017.11.014 doi: 10.1016/j.jfa.2017.11.014
![]() |
[37] |
J. Ok, Partial regularity for general systems of double phase type with continuous coefficients, Nonlinear Anal., 177 (2018), 673–698. https://doi.org/10.1016/j.na.2018.03.021 doi: 10.1016/j.na.2018.03.021
![]() |
[38] |
J. Ok, G. Scilla, B. Stroffolini, Boundary partial regularity for minimizers of discontinuous quasiconvex integrals with general growth, Commun. Pure Appl. Anal., 21 (2022), 4173–4214. https://doi.org/10.3934/cpaa.2022140 doi: 10.3934/cpaa.2022140
![]() |
[39] |
K. R. Rajagopal, M. Růžička, Mathematical modeling of electrorheological materials, Continuum Mech. Thermodyn., 13 (2001), 59–78. https://doi.org/10.1007/s001610100034 doi: 10.1007/s001610100034
![]() |
[40] | M. Růžička, Electrorheological fluids: modeling and mathematical theory, Berlin, Heidelberg: Springer, 2000. https://doi.org/10.1007/BFb0104029 |
1. | Jihoon Ok, Giovanni Scilla, Bianca Stroffolini, Partial regularity for degenerate systems of double phase type, 2025, 432, 00220396, 113207, 10.1016/j.jde.2025.02.078 | |
2. | Nirjan Biswas, Laura Gambera, Umberto Guarnotta, On the linear independence condition for the Bobkov-Tanaka first eigenvalue of the double-phase operator, 2025, 166, 08939659, 109549, 10.1016/j.aml.2025.109549 |