We study local boundedness for subsolutions of nonlinear nonuniformly elliptic equations whose prototype is given by ∇⋅(λ|∇u|p−2∇u)=0, where the variable coefficient 0≤λ and its inverse λ−1 are allowed to be unbounded. Assuming certain integrability conditions on λ and λ−1 depending on p and the dimension, we show local boundedness. Moreover, we provide counterexamples to regularity showing that the integrability conditions are optimal for every p>1.
Citation: Peter Bella, Mathias Schäffner. Local boundedness for p-Laplacian with degenerate coefficients[J]. Mathematics in Engineering, 2023, 5(5): 1-20. doi: 10.3934/mine.2023081
[1] | Giovanni Cupini, Paolo Marcellini, Elvira Mascolo . Local boundedness of weak solutions to elliptic equations with $ p, q- $growth. Mathematics in Engineering, 2023, 5(3): 1-28. doi: 10.3934/mine.2023065 |
[2] | David Cruz-Uribe, Michael Penrod, Scott Rodney . Poincaré inequalities and Neumann problems for the variable exponent setting. Mathematics in Engineering, 2022, 4(5): 1-22. doi: 10.3934/mine.2022036 |
[3] | Italo Capuzzo Dolcetta . The weak maximum principle for degenerate elliptic equations: unbounded domains and systems. Mathematics in Engineering, 2020, 2(4): 772-786. doi: 10.3934/mine.2020036 |
[4] | Lucio Boccardo . A "nonlinear duality" approach to $ W_0^{1, 1} $ solutions in elliptic systems related to the Keller-Segel model. Mathematics in Engineering, 2023, 5(5): 1-11. doi: 10.3934/mine.2023085 |
[5] | Youchan Kim, Seungjin Ryu, Pilsoo Shin . Approximation of elliptic and parabolic equations with Dirichlet boundary conditions. Mathematics in Engineering, 2023, 5(4): 1-43. doi: 10.3934/mine.2023079 |
[6] | Isabeau Birindelli, Giulio Galise . Allen-Cahn equation for the truncated Laplacian: Unusual phenomena. Mathematics in Engineering, 2020, 2(4): 722-733. doi: 10.3934/mine.2020034 |
[7] | Edgard A. Pimentel, Miguel Walker . Potential estimates for fully nonlinear elliptic equations with bounded ingredients. Mathematics in Engineering, 2023, 5(3): 1-16. doi: 10.3934/mine.2023063 |
[8] | Aleksandr Dzhugan, Fausto Ferrari . Domain variation solutions for degenerate two phase free boundary problems. Mathematics in Engineering, 2021, 3(6): 1-29. doi: 10.3934/mine.2021043 |
[9] | Antonio Greco, Francesco Pisanu . Improvements on overdetermined problems associated to the $ p $-Laplacian. Mathematics in Engineering, 2022, 4(3): 1-14. doi: 10.3934/mine.2022017 |
[10] | François Murat, Alessio Porretta . The ergodic limit for weak solutions of elliptic equations with Neumann boundary condition. Mathematics in Engineering, 2021, 3(4): 1-20. doi: 10.3934/mine.2021031 |
We study local boundedness for subsolutions of nonlinear nonuniformly elliptic equations whose prototype is given by ∇⋅(λ|∇u|p−2∇u)=0, where the variable coefficient 0≤λ and its inverse λ−1 are allowed to be unbounded. Assuming certain integrability conditions on λ and λ−1 depending on p and the dimension, we show local boundedness. Moreover, we provide counterexamples to regularity showing that the integrability conditions are optimal for every p>1.
In this note, we study local boundedness of weak (sub)solutions of non-uniformly elliptic quasi-linear equations of the form
∇⋅a(x,∇u)=0inΩ, | (1.1) |
where Ω⊂Rd with d≥2 and a:Ω×Rd→Rd is a Caratheodory function. The main example that we have in mind are p-Laplace type operators with variable coefficients, that is, there exist p>1 and A:Ω→Rd×d such that a(x,ξ)=A(x)|ξ|p−2ξ for all x∈Ω and ξ∈Rd. In order to measure the ellipticity of a, we introduce for fixed p>1
λ(x):=infξ∈Rd∖{0}a(x,ξ)⋅ξ|ξ|pμ(x):=supξ∈Rd∖{0}|a(x,ξ)|p(a(x,ξ)⋅ξ)p−1 | (1.2) |
and suppose that λ and μ are nonnegative. In the uniformly elliptic setting, that is that there exists 0<m≤M<∞ such that m≤λ≤μ≤M in Ω, solution to (1.1) are locally bounded, Hölder continuous and even satisfy Harnack inequality, see e.g., classical results of Ladyzhenskaya & Ural'tseva, Serrin and Trudinger [34,41,42].
In this contribution, we are interested in a nonuniformly elliptic setting and assume that λ−1∈Lt(Ω) and μ∈Ls(Ω) for some integrability exponents s and t. In [7], we studied this in the case of linear nonuniformly elliptic equations, that is a(x,ξ)=A(x)ξ corresponding to the case p=2, and showed local boundedness and Harnack inequality for weak solutions of (1.1) provided it holds 1s+1t<2d−1. The results of [7] improved classical findings of Trudinger [43,44] (see also [39]) from the 1970s and are optimal in view of counterexamples constructed by Franchi et al. in [27]. In this manuscript we extend these results to the more general situation of quasilinear elliptic equation with p-growth as described above. More precisely, we show
Theorem 1. Let d≥2,p>1, and let Ω⊂Rd. Moreover, let s∈[1,∞] and t∈(1/(p−1),∞] satisfy
1s+1t<pd−1. | (1.3) |
Let a:Ω×Rd→Rd be a Caratheodory function with a(⋅,0)≡0 such that λ and μ defined in (1.2) satisfy μ∈Ls(Ω) and 1λ∈Lt(Ω). Then any weak subsolution of (1.1) is locally bounded from above in Ω.
Remark 1. Note that Theorem 1, restricted to the case p=2 recovers the local boundedness part of [7, Theorem 1.1].
Remark 2. In [15], Cupini, Marcellini and Mascolo studied local boundedness of local minimizer of nonuniformly elliptic variational integrals of the form ∫Ωf(x,∇v)dx where f satisfies
λ(x)|ξ|p≤f(x,ξ)≤μ(x)+μ(x)|ξ|qwithλ−1∈Lt(Ω)andμ∈Ls(Ω). | (1.4) |
They proved local boundedness under the relation 1pt+1qs+1p−1q<1d (see also [11] for related results). Considering the specific case f(x,ξ)=λ(x)|ξ|p, the result of [15] implies local boundedness of solutions to ∇⋅(λ(x)|∇u|p−2∇u)=0 provided λ−1∈Lt(Ω) and λ∈Ls(Ω) with 1s+1t<pd, which is more restrictive compared to assumption (1.3) in Theorem 1. It would be interesting to investigate if the methods of the present paper can be combined with the ones of [15] to obtain local boundedness for minimizer of functionals satisfying (1.4) assuming 1pt+1qs+1p−1q<1d−1. Note that in the specific case s=t=∞, this follows from [32].
Remark 3. We emphasize that we only impose global integrability conditions on λ−1 and μ. Assuming additional local conditions on the coefficients in the form λ∼μ and μ is in some Muckenhoupt class, local boundedness is proven under weaker integrability conditions in the seminal work [25] in the case p=2 (see also [31] for the case p>1); for further recent results on higher regularity for nonlinear elliptic equations with Muckenhoupt coefficients we refer to [4,5,13,17].
The proof of Theorem 1 is presented in Section 2 and follows a variation of the well-known Moser-iteration method. The main new ingredient compared to earlier works [15,43] lies in an optimized choice of certain cut-off functions – an idea that we first used in [7] for linear nonuniformly elliptic equations (see also [1,10,45] for recent applications to linear parabolic equations).
As mentioned above, an example constructed in [27] shows that condition (1.3) is optimal for the conclusion of Theorem 1 in the case p=2. In the second main result of this paper, we show – building on the construction of [27] – that condition (1.3) is optimal for the conclusion of Theorem 1 for all p∈(1,∞). More precisely, we have
Theorem 2. Let d≥3, 1+1d−2<p<∞, and let s≥1 and t>1p−1 be such that 1s+1t≥pd−1 and p1+1/t<d−1. Then there exists λ:B(0,1)→(0,∞) satisfying λ∈Ls(B1) and λ−1∈Lt(B1) and an unbounded weak subsolution of
−∇⋅(λ|∇v|p−2∇v)=0 | (1.5) |
in B(0,1). Moreover, the same conclusion is valid for d≥3, 1<p≤1+1d−2 and s≥1 and t>1p−1 satisfying the strict inequalities 1s+1t>pd−1 and tt+1p<d−1.
In particular, we see that condition (1.3) is sharp on the scale of Lebesgue-integrability for the conclusion of Theorem 1. We note that in the particularly interesting case p=2 and d=3 the construction in Theorem 2 fails in the critical case 1s+1t=pd−1, see [1] for counterexamples to local boundedness for related problems in d=3.
Let us now briefly discuss a similar but different instance of non-uniform ellipticity which is one of the many areas within the Calculus of Variations, where G. Mingione made significant contributions. Consider variational integrals
∫ΩF(x,∇u)dx, | (1.6) |
where the integrand F satisfies (p,q) growth conditions of the form
|ξ|p≲F(x,ξ)≲1+|ξ|q1<p≤q<∞, | (1.7) |
which where first systematically studied by Marcellini in [35,36]; see also the recent reviews [37,38]. The focal point in the regularity theory for those functionals is to obtain Lipschitz-bounds on the minimizer. Indeed, once boundedness of |∇u| is proven the unbalanced growth in (1.7) becomes irrelevant and there is a huge literature dedicated to Lipschitz estimates under various assumptions on F, see e.g., the interior estimates [6,8,9,23] in the autonomous case, [2,14,16,18,19,20,22,30] in the non-autonomous case, [12,21] for Lipschitz-bounds at the boundary, and also examples where the regularity of minimizer fail [3,24,26,28,36]. Finally, we explain a link between functionals with (p,q)-growth and (linear) equations with unbounded coefficients. Consider the autonomous case that F(x,ξ)=F(ξ) and let u∈W1,p(Ω) be a local minimizer of (1.6). Linearizing the corresponding Euler-Largrange equation yield (formally)
∇⋅D2F(∇u)∇∂iu=0. |
Assuming (p,q)-growth with p=2 of the form |ζ|2≲D2F(ξ)ζ⋅ζ≲(1+|ξ|)q−2|ζ|2 implies that |D2F(∇u)|∈L2q−2loc(Ω). Hence condition (1.3) with p=2 yield local boundedness of ∂iu if q−22<2d−1, which is the currently best known general bound ensuring Lipschitz-continuity of local minimizer of (1.6) – this reasoning was made rigorous in [8] for p≥2 (see also [9] for the case p∈(1,∞)).
Before we prove Theorem 1, we introduce the notion of solution that we consider here.
Definition 1. Fix a domain Ω⊂Rd and a Caratheodory function a:Ω×Rd→Rd such that for a fixed p∈(1,∞) the functions λ,μ≥0 given in (1.2) satisfy 1λ∈L1p−1(Ω) and μ∈L1(Ω). The spaces H1,p0(Ω,a) and H1,p(Ω,a) are respectively defined as the completion of C1c(Ω) and C1(Ω) with respect to the norm ‖⋅‖H1,p(Ω,a), where
‖u‖H1,p(Ω,a):=(∫Ωλ|∇u|p+μ|u|pdx)1p. |
We call u a weak solution (subsolution, supersolution) of (1.1) in Ω if and only if u∈H1,p(Ω,a) and
∀ϕ∈H1,p0(Ω,a),ϕ≥0:A(u,ϕ)=0(≤0,≥0),whereA(u,ϕ):=∫Ωa(x,∇u)⋅∇ϕdx. | (2.1) |
Moreover, we call u a local weak solution of (1.1) in Ω if and only if u is a weak solution of (1.1) in Ω′ for every bounded open set Ω′⋐Ω. Throughout the paper, we call a solution (subsolution, supersolution) of (1.1) in Ωa-harmonic (a-subharmonic, a-superharmonic) in Ω.
The above definitions generalize the concepts of weak solutions and the spaces H1(Ω,a) and H10(Ω,a) discussed by Trudinger [43,44] in the linear case, that is a(x,ξ)=A(x)ξ. We stress that the condition λ−1∈L1p−1(Ω) and Hölder inequality imply
‖∇u‖L1(Ω)≤‖λ−1‖L1p−1(Ω)(∫Ωλ|∇u|p)1p≤‖λ−1‖L1p−1(Ω)‖u‖H1,p(Ω,a) |
and thus, we have that W1,1(Ω)⊂H1,p(Ω,a), where we use that by the same computation as above it holds ‖u‖L1(Ω)≤‖μ−1‖L1p−1(Ω)‖u‖H1,p(Ω,a) and that by definition we have λ≤μ. From this, we also deduce that the elements of H1,p(Ω,a) are strongly differentiable in the sense of [29]. In particular this implies that there holds a chain rule in the following sense
Remark 4. Let g:R→R be uniformly Lipschitz-continuous with g(0)=0 and consider the composition F:=g(u). Then, u∈H1,p0(Ω,a) (or ∈H1,p(Ω,a)) implies F∈H1,p0(Ω,a) (or ∈H1,p(Ω,a)), and it holds ∇F=g′(u)∇u a.e. (see e.g., [44, Lemma 1.3]). In particular, if u satisfies u∈H1,p(Ω,a) (or ∈H1,p(Ω,a)) then also the truncations
u+:=max{u,0};u−:=−min{u,0} |
satisfy u+,u−∈H1,p(Ω,a) (or ∈H1,p(Ω,a)).
Now we come to the local boundedness from above for weak subsolutions of (1.1). In order to state the estimates in the right dimensionality, we introduce for v∈W1,γ(Ω) with γ≥1 the notation
‖v‖W_1,γ(Ω):=|Ω|−1γ‖v‖Lγ(Ω)+|Ω|1d−1γ‖∇v‖Lγ(Ω), | (2.2) |
where |Ω| denotes the d-dimensional Lebesgue-measure of Ω. Moreover, we denote by ‖v‖W1,γ(Ω) the usual Sobolev-norm given by ‖v‖W1,γ(Ω):=‖v‖Lγ(Ω)+‖∇v‖Lγ(Ω).
Theorem 3. Let d≥3, Ω⊂Rd and p∈(1,∞). Moreover, let s∈[1,∞] and t∈(1p−1,∞] satisfy (1.3). Let a:Ω×Rd→Rd be a Caratheodory function with a(⋅,0)≡0 such that λ and μ defined in (1.2) satisfy μ∈Ls(Ω) and 1λ∈Lt(Ω) and for every measurable set S⊂Ω, we set
Λ(S):=(−∫Sμs)1/s(−∫Sλ−t)1/t. |
Then, there exists c=c(d,p,s,t)∈[1,∞) such that for any weak subsolution u of (1.1) and for any ball BR⊂Ω it holds
supBR/2u≤cΛ(BR)1p1δ‖u+‖W_1,11+1/tp(BR), |
where ‖⋅‖W_1,γ(Br) is defined in (2.2); and δ:=1s∗−(1p−1pt)>0 (see Lemma 1 for the definition of s∗). Moreover, in the case 1+1t<pd−1, there exists c=c(d,p,t)∈[1,∞) such that
supBR/2u≤c‖u+‖W_1,11+1/tp(BR). |
In the two-dimensional case, we have the following
Proposition 1. Let Ω⊂R2 and p∈(1,∞). Let a:Ω×Rd→Rd be a Caratheodory function with a(⋅,0)≡0 such that λ and μ defined in (1.2) satisfy μ∈L1(Ω) and 1λ∈L1p−1(Ω). Then, there exists c=c(d,p)∈[1,∞) such that for any weak subsolution u of (1.1) and for any ball BR⊂Ω it holds
supBR/2u≤c‖u+‖W_1,1(BR). |
Before we proof Theorem 3 and Proposition 1, we show that they imply the claim of Theorem 1.
Proof of Theorem 1. In view of Theorem 3 and Proposition 1 it remains to show that for any weak subsolution u of (1.1) and for any ball BR⊂Ω it holds ‖u+‖W_1,tt+1p(BR)<∞. This is a consequence of Hölder inequality and the concept of weak subsolution, see Definition 1. Indeed, we have
(∫BR(|u|+|∇u|)tpt+1)t+1t≤(∫BRλ−t)1t∫BRλ(|u|+|∇u|)p<∞, |
where the right-hand side is finite since u∈H1,p(Ω,a) (note that λ≤μ by definition).
For the proof of Theorem 3, we need a final bit of preparation, namely the following optimization lemma
Lemma 1 (Radial optimization). Let d≥3, p>1, s>1, and let s∗:=max{1,(1p(1−1s)+1d−1)−1}. For 12≤ρ<σ≤2, let v∈W1,s∗(Bσ) and μ∈Ls(Bσ), μ≥0, be such that μ|v|p∈L1(Bσ). Then there exists c=c(d,p,s) such that
J(ρ,σ,v):=inf{∫Bσμ|v|p|∇η|pdx:η∈C10(Bσ),η≥0,η=1inBρ} |
satisfies
J(ρ,σ,v)≤c(σ−ρ)−pdd−1‖μ‖Ls(Bσ∖Bρ)(‖∇v‖pLs∗(Bσ∖Bρ)+ρ−p‖v‖pLs∗(Bσ∖Bρ)). |
Lemma 1 generalizes [7, Lemma 2.1] from p=2 to p>1 and we provide a proof in the appendix.
Proof of Theorem 3. By standard scaling and translation arguments it suffices to suppose that B1⋐Ω and u is locally bounded in B12. Hence, we suppose from now on that B1⋐Ω. In Steps 1–4 below, we consider the case s>1. We first derive a suitable Caccioppoli-type inequality for powers of u+ (Step 1) and perform a Moser-type iteration (Steps 2–4). In Step 5, we consider the case 1+1t<pd−1 which includes the case s=1.
Step 1. Caccioppoli inequality.
Assuming B⊂Ω, for any cut-off function η∈C10(B), η≥0 and any β≥1, there holds
∫ηpλ(x)uβ−1+|∇u+|p≤(pβ)p∫up+β−1+μ(x)|∇η|p. | (2.3) |
For β≥1, we use the weak formulation (2.1) with ϕ:=ηpuβ+: *
*Rigorously, we are a priori not allowed to test with uβ. Instead, for N≥1 one should modify uβ by replacing uβ with affine αNα−1u−(α−1)Nβ in the set u≥N, obtain the conclusion by testing the weak formulation with this modified function, and subsequently sends N→∞ – for details, see [7, Page 460].
∫a(x,∇u)⋅∇(ηpuβ+)≤0. |
We have ∫(a(x,∇u)−a(x,∇u+))⋅∇(ηpu+)=0, so that we were able to replace u with u+ inside a(x,⋅). Applying Leibniz rule we get from the previous display
β∫ηpuβ−1a(x,∇u)⋅∇u≤−∫pηp−1uβa(x,∇u)⋅∇η, | (2.4) |
where to simplify the notation for the rest of this proof we write u instead of u+. Using definition of μ in (1.2) in form of |a(x,ξ)|≤μ(x)1p(a(x,ξ)⋅ξ)p−1p for any ξ∈Rd (in fact we use (1.2) for ξ≠0 and for ξ=0 the inequality follow from the assumption a(x,0)=0), we can bound the r.h.s. in the last math display from above by
p∫ηp−1uβμ(x)1p(a(x,∇u)⋅∇u)p−1p|∇η|=p∫uβ−(β−1)p−1pμ(x)1p|∇η|(ηpuβ−1a(x,∇u)⋅∇u)p−1p≤p(∫up+β−1μ(x)|∇η|p)1p(∫ηpuβ−1a(x,∇u)⋅∇u))p−1p, |
where in the second step we applied Hölder inequality with exponents p and pp−1, respectively. Observe that the last term on the r.h.s. appears on the l.h.s. in (2.4), so that after absorbing it we get from (2.4)
β(∫ηpuβ−1a(x,∇u)⋅∇u)1p≤p(∫up+β−1μ(x)|∇η|p)1p, |
which after taking the p-th power turns into
∫ηpuβ−1a(x,∇u)⋅∇u≤(pβ)p∫up+β−1μ(x)|∇η|p. |
By definition of λ in (1.2) in form of λ(x)|ξ|p≤a(x,ξ)⋅ξ for any ξ∈Rd, one has λ(x)|∇u|p≤a(x,∇u)⋅∇u, thus implying the claimed Caccioppoli inequality (2.3).
Step 2. Improvement of integrability.
We claim that there exists c=c(d,p,s)∈[1,∞) such that for 12≤ρ<σ≤1 and α≥1 it holds
‖∇(uα)‖Lptt+1(Bρ)≤c(σ−ρ)−dd−1Λ(Bσ)1p‖uα‖W1,s∗(Bσ∖Bρ). | (2.5) |
Let η∈C10(Bσ), η≥0, with η=1 in Bρ. First, we rewrite the Caccioppoli inequality (2.3) from Step 1 as inequality for u1+β−1p:
(pp+β−1)p∫ηpλ(x)|∇(u1+β−1p)|p≤(pβ)p∫μ(x)(u1+β−1p)p|∇η|p. | (2.6) |
Calling v:=u1+β−1p, we can estimate the r.h.s. with the help of Lemma 1, yielding
∫ηpλ(x)|∇v|p≤c(p+β−1β)p(σ−ρ)−pdd−1‖μ‖Ls(Bσ∖Bρ)(‖∇v‖pLs∗(Bσ∖Bρ)+ρ−p‖v‖pLs∗(Bσ∖Bρ)). |
Using Hölder inequality with exponents (t+1t,t+1) and the fact that η=1 in Bρ, we see that
‖∇v‖pLptt+1(Bρ)≤‖λ−1‖Lt(Bρ)‖λ|∇v|p‖L1(Bρ)≤‖λ−1‖Lt(Bρ)∫ηpλ(x)|∇v|p. |
Using that 12≤ρ≤σ≤1, combination of two previous relations yields
‖∇v‖pLptt+1(Bρ)≤c(p+β−1β)p(σ−ρ)−pdd−1Λ(Bσ)‖v‖pW1,s∗(Bσ∖Bρ), |
which after taking p-root turns into
‖∇(uα)‖Lptt+1(Bρ)≤c(σ−ρ)−dd−1Λ(Bσ)1p‖uα‖W1,s∗(Bσ∖Bρ), |
with α:=1+β−1p.
Step 3. One-step improvement.
First, we note that (1.3) and t>1p−1 imply δ:=1s∗−1p(1+1t)>0. In particular it holds s∗<tpt+1. We claim that there exists c=c(d,s,t,p) such that for 12≤ρ<σ≤1 there holds
‖uχα‖1χαW1,s∗(Bρ)≤(cΛ(Bσ)1p(σ−ρ)dd−1)1χα‖uα‖1αW1,s∗(Bσ), | (2.7) |
where χ:=1+δ>1. Using Hölder inequality with exponent pt(t+1)s∗>1 and its dual exponent ptpt−(t+1)s∗=1δs∗ we get
(∫Bρ|∇(u(1+δ)α)|s∗)1s∗=(1+δ)α(∫Bρ|∇u|s∗u(α−1)s∗uαδs∗)1s∗=(1+δ)(∫Bρ|∇(uα)|s∗uαδs∗)1s∗≤(1+δ)(∫Bρ|∇(uα)|ptt+1)t+1pt(∫Bρuα)δ. |
Combining the above estimate with (2.5) from Step 2, we get (recall χ=1+δ)
‖∇(uχα)‖Ls∗(Bρ)≤c(σ−ρ)−dd−1Λ(Bσ)1p‖uα‖χW1,s∗(Bσ), |
where we hided χ=1+δ<dd−1 into c. In order to have full W1,s∗(Bρ)-norm also on the l.h.s., using s∗≥1 as well as χ<dd−1 we can use Sobolev inequality to the effect
‖uχα‖Ls∗(Bρ)≤c‖uα‖W1,s∗(Bρ), |
thus obtaining the claim.
Step 4. Iteration.
We iterate the outcome of Step 3. For ˉα≥1 and n∈N let αn:=ˉαχn−1, ρn:=12+12n+1, σn:=ρn+12n+1=ρn−1. Then (2.8) from Step 4 with α:=αn has the form
‖uαn+1‖1αn+1W1,s∗(Bρn)≤(cΛ(B1)1p4n)1ˉαχn‖uαn‖1αnW1,s∗(Bρn−1). | (2.8) |
Using that Lp approximates L∞ as p→∞, we see that
‖u‖L∞(B1/2)≤(∞∏n=1(cΛ(Bσ)1p4n)1ˉαχn)‖uˉα‖1ˉαW1,s∗(B1)≤cΛ(Bσ)1pˉα1χ−1‖uˉα‖1ˉαW1,s∗(B1), | (2.9) |
which for ˉα=1 yields the desired claim where we use χ=1+δ and s∗≤tpt+1.
Step 5. The remaining case 1+1t<pd−1.
Using Fubini theorem, we can choose a generic radius r0∈(12,1) such that
‖u+‖ptt+1W1,ptt+1(Sr0)≤2‖u+‖ptt+1W1,ptt+1(B1). |
We test the weak formulation of −∇⋅a(x,∇u)≤0 see (2.1) with the non-negative test function ϕ:=(u+−supSr0u+)+, which obviously vanishes on Sr0 and can be therefore trivially extended by zero to the whole domain Ω. This yields
0(2.1)≥∫Br0a(x,∇u)⋅∇ϕ=∫Br0a(x,∇ϕ)⋅∇ϕ(1.2)≥∫Br0λ(x)|∇ϕ|p. |
In particular, we see that ∇ϕ=0 a.e. in Br0, hence ϕ≡0 and thus
‖u+‖L∞(B12)≤‖u+‖L∞(Br0)≤supSr0u+. |
Using that ptt+1>d−1, which follows from 1+1t<pd−1, we have by Sobolev embedding that supSr0u+≤c‖u+‖W1,ptt+1(Sr0) for some c=c(d,p,t)>0 which by the above choice of r0 completes the claim.
Proof of Proposition 1. This follows exactly as in Step 5 of the proof of Theorem 3 using that for d=2 it holds supSr0u+≤c‖u+‖W1,1(Sr0).
We close this section by deriving from Theorem 3 in the case s>1 an L∞−Lγ estimate.
Corollary 1. Let d≥2, Ω⊂Rd and p∈(1,∞). Moreover, let s∈(1,∞] and t∈(1p−1,∞] satisfy (1.3). Let a:Ω×Rd→Rd be a Caratheodory function with a(⋅,0)≡0 such that λ and μ defined in (1.2) satisfy μ∈Ls(Ω) and 1λ∈Lt(Ω). Then, any weak subsolution u of (1.1) and any γ>0 there exists c=c(γ,d,p,s,t)∈[1,∞) such that for any ball BR⊂Ω
supBR/2u≤cΛ(BR)1γss−1(1+1δ)(−∫BRuγ+)1γ. |
Proof. Without loss of generality we consider R=1 and suppose that B1⋐Ω. Caccioppoli inequality (2.6) with β=1+p(α−1) for α≥1 and η∈C1c(B1) with η=1 on B12 and |∇η|≤2 and Hölder inequality yield
‖∇(uα+)‖pLptt+1(B1/2)≤‖λ−1‖Lt(B1)∫B1ηpλ|∇(uα+)|p≤(2p)p‖λ−1‖Lt(B1)∫B1μuαp+≤(2p)p‖λ−1‖Lt(B1)‖μ‖Ls(B1)‖uα+‖Lss−1p(B1). |
The above inequality combined with tpt+1≤p≤sps−1 implies ‖uα+‖1αW1,tpt+1(B1/2)≤cΛ(B1)1αp‖u+‖Lαpss−1(B1) (note that 1≤Λ(Br)) for some c=c(d,p)∈[1,∞). Hence, we have in combination with (2.9) that
‖u+‖L∞(B1/4)≤cΛ(B1)1αp(1+1δ)‖u+‖Lαpss−1(B1), | (2.10) |
where c=c(α,d,p,t,s)∈[1,∞).
From estimate (2.10) the claim follows by routine arguments and we only sketch the idea (see [7, Proof of Theorem 3.3, Step 2] for precise arguments in the case p=2). By scaling and translation, we deduce from (2.10) that for all ρ>0 and x∈B1 such that Bρ(x)⊂B1 it holds for α≥1
‖u+‖L∞(Bρ/4(x))≤cΛ(Bρ(x))1αp(1+1δ)ρ−dp(1−1s)‖u+‖Lαpss−1(Bρ(x)), |
where c is as in (2.10). Combining the above estimate with a simple covering argument, we obtain that there exists c=c(α,d,p,s,t)∈[1,∞) such that for all θ∈(0,1) and r∈(0,1] it holds
‖u+‖L∞(Bθr)≤cΛ(Br)1αp(1+1δ)(1−θ)−κr−ds−1αps‖u+‖Lαpss−1(Br), |
where κ:=dαp((1t+1s)(1+1δ)+1−1s) which is the claim for all γ≥pss−1 (by choosing α=s−1psγ). The claim for γ∈(0,pss−1) follows by a standard interpolation and iteration argument see e.g., the textbook reference [33, p. 75] in the uniformly elliptic case or as mentioned above [7, Proof of Theorem 3.3, Step 2] for a closely related setting.
Proof of Theorem 2. The following construction is very much inspired by a construction in [27] in the linear case, that is p=2, and d=4 (which was already extended to d≥3 in [40]).
Let d≥3. Throughout the proof, we set
x=(x1,…,xd)=(x1,x′)and|x′|=√d∑j=2x2j. |
For any p∈(1,∞) and θ∈[0,1], we define λθ(x):=ωθ(|x′|) where ωθ:(0,1)→R+ is defined as
ωθ(r)={(i+1)(p−1)θ4−piθwhenr∈[124−i,4−i),((i+1)−(p−1)4pi)1−θwhenr∈[144−i,124−i) | (3.1) |
for i∈N. We will construct an explicit subsolution to −∇⋅(λθ|∇v|p−2∇v)=0, which is of the form
v(x)=eαx1ϕ(|x′|) | (3.2) |
for some parameter α=α(d,p)>0 and ϕ:(0,1)→R is defined by
ϕ(r)={i+ηi2Q−1((4ir)−Q−1)whenr∈[124−i,4−i),(i+1)−(1−ηi)(4i+1r−1)2whenr∈[144−i,124−i),withQ={max{d−3,1}ifp≥2d−2p−1−1if1<p<2 | (3.3) |
where ηi∈[0,1] will be specified below. Note that Q>0 and ϕ is continuous by definition. We choose ηi∈(0,1) such that the flux λθ|∇v|p−2∇v is continuous at |x′|=124−i for every i∈N. More precisely, we set ηi to be the largest constant (in [0,1]) satisfying
Fi(ηi)=0, | (3.4) |
where Fi:(0,1]→R is given by
Fi(η):=√(α(i+η)4−i)2+(CQη)2p−2CQη−√(α(i+η)(i+1)−1)2+(8(1−η)4i(i+1)−1)2p−28(1−η)42i(i+1)−1 |
with
CQ=Q2Q+12Q−1. |
Note that ηi is well-defined since Fi:(0,1)→R is continuous with
limη→0Fi(η)=−√(αi)2+(2⋅4i+1)2p−28⋅42i(i+1)−(p−1)<0 |
and
limη→1Fi(η)=√(α(i+1)4−i)2+C2Qp−2CQ>0. |
The definition of ηi is rather implicit and we provide now some explicit bounds on ηi which will be useful for later computations. We distinguish two cases. For p≥2 and α≥CQ, we have that
∃j=j(d,p)≥2 such that∀i≥j:ηi≥1−8−1(4p−2CQ)4−2i(i+1)=:η_i. | (3.5) |
Indeed, let j=j(d,p)≥2 be such that η_i∈(0,1) for all i≥j. By definition of ηi, it suffices to show that Fi(η_i)≤0 for i≥j. We have
Fi(η_i)≤√(α(i+1)4−i)2+C2Qp−2CQ−√(αi/(i+1))2p−2(4p−2CQ)=√((i+1)4−i)2+(CQ/α)2p−2αp−2CQ−αp−2√(i/(i+1))2p−2(4p−2CQ)≤αp−2(2p−2CQ−2−(p−2)(4p−2CQ))=0, |
where we used for the last inequality (i+1)4−i≤1 and i/(i+1)≥12 for i≥1 and α≥CQ.
In the case p∈(1,2), we have for α≥22−pp−1CQ that
∃j=j(α,d,p)≥2 such that∀i≥j:ηi≥1−8−1α4−2i(i+1)=:¯ηi. | (3.6) |
Indeed, this follows as above from
Fi(¯ηi)≤Cp−1Q−√α2+(α4−i)2p−2α≤Cp−1Q−αp−12p−2≤0. |
Step 1. We show that for every α≥max{1,22−pp−1}CQ, the function v defined in (3.2) has finite energy, that is ∫B1λθ(|v|p+|∇v|p)<∞ provided (1−θ)p<d−1.
We show first ∫B1λθ|v|p<∞. For this, we observe that 0≤ϕ(r)≤log(4/r) for all r∈(0,1). Indeed, ϕ≥0 is clear from the definition (3.3) and for r∈[144−i,4−i), we have
ϕ(r)≤i+1=log4(4i+1)≤log4(4r)≤log(4r). |
Similarly, we get
ωθ(r)≤{((2r)plog(4/r)p−1)θwhenr∈[124−i,4−i),(rplog(2/r)p−1)−(1−θ)whenr∈[144−i,124−i). | (3.7) |
Hence, there exists C=C(α,d,p)>0 such that
∫B1λθvpdx≤C∫10r−(1−θ)plog(2/r)p−(1−θ)(p−1)rd−2dr<∞, |
where the last integral is finite since (1−θ)p<d−1.
Next, we show ∫B1λθ|∇v|p<∞. For this we compute the gradient of v:
∇v=(αϕϕ′x′|x′|)eαx1and|∇v|=√α2ϕ2+ϕ′2eαx1. | (3.8) |
Moreover, we compute
ϕ′(r)={−Qηi2Q−1(4ir)−Qr−1whenr∈(124−i,4−i),−2(1−ηi)4i+1(4i+1r−1)whenr∈(144−i,124−i) | (3.9) |
and for later usage
ϕ″(r)={Q(Q+1)ηi2Q−1(4ir)−Qr−2whenr∈(124−i,4−i),−2(1−ηi)42(i+1)whenr∈(144−i,124−i). | (3.10) |
From (3.5) and (3.6), we obtain that there exists C=C(α,d,p)>0 such that 0≤1−ηi≤C4−2i(i+1) for i≥j(α,d,p) and thus in combination with (3.9) there exists C=C(α,d,p)>0 such that
|ϕ′(r)|≤C{r−1whenr∈(124−i,4−i),log(2/r)rwhenr∈(144−i,124−i) |
for all i≥j. Hence, we find C=C(α,d,p)>0 such that
∫B1λθ|∇v|p≤C+C∫10((rplog(2/r)p−1)θr−p+(rplog(2/r)p−1)−(1−θ)(log(2/r)r)p)rd−2dr<∞, |
where we use again (1−θ)p<d−1. Finally, it is easy to check that the sequence (vk)k defined by vk(x)=eαx1ϕk(|x′|) with ϕk(x)=ϕ(x) if |x|>4−k and ϕk(x)=k if |x′|≤4−k is a sequence of Lipschitz functions satisfying limk→∞∫B1λθ(|v−vk|p+|∇v−∇v|p)→0 as k→∞ and a straightforward regularization shows that v in H1,p(B1,a) with a(x,ξ):=λθ(x)|ξ|p−2ξ.
Step 2. We claim that there exist α0=α0(d,p)≥1 such that for every α≥α0 there exists ρ=ρ(α,d,p)∈(0,1] such that v defined in (3.2) is a weak subsolution in {x∈B1:δ<|x′|<ρ} for all δ>0.
For this, we observe first that by (3.8) the nonlinear strain |∇v|p−2∇v of v is given by
|∇v|p−2∇v=√α2ϕ2+ϕ′2p−2(αϕϕ′x′|x′|)eα(p−1)x1. | (3.11) |
Introducing the notation M2i=B1∩{124−i<|x′|<4−i} and M2i+1=B1∩{144−i<|x′|<124−i}, we obtain with help of integrating by parts
∫B1λθ|∇v|p−2∇v⋅∇φ=∑i∈N∫Miωθ|∇v|p−2∇v⋅∇φ=∑i∈N−∫Miωθ∇⋅(|∇v|p−2∇v)φ+∫∂Miωθ|∇v|p−2∇v⋅νφ=∑i∈N−∫Miωθ∇⋅(|∇v|p−2∇v)φ+∫∂Miωθ√α2ϕ2+ϕ′2p−2ϕ′e(p−1)αx1φ, |
where ν denotes the outer unit normal to Mi that is ν=(0,x′/|x′|). Hence, it suffices to show that there exists α0>0 such that for all α≥α0 there exists j=j(α,d,p)≥2 such that
(i) v satisfies ∇⋅(|∇v|p−2∇v)≥0 in the classical sense in each shell Mi for all i≥j;
(ii) the flux has only nonnegative jumps at the interfaces, that is
(ωθ√α2ϕ2+ϕ′2p−2ϕ′)(γ−):=limr→γr<γ(ωθ√α2ϕ2+ϕ′2p−2ϕ′)(r)≤limr→γr>γ(ωθ√α2ϕ2+ϕ′2p−2ϕ′)(r)=:(ωθ|∇v|p−2ϕ′)(γ+) |
for all γ∈⋃i∈N,i≥j{4−i}∪{124−i}.
Substep 2.1. Argument for (i). Let α≥1 be such that
α≥α0(p,d):=max{1,CQ,22−pp−1CQ,2p√CQ(1+d−2p−1),8d−1p−1} | (3.12) |
and let j=j(α,d,p)≥2 be such that the estimates (3.5) and (3.6) are valid.
We show that v with α as above, satisfies ∇⋅(|∇v|p−2∇v)≥0 in the classical sense in each shell Mi for all i≥j. We compute with help of (3.11) on Mi
∇⋅(|∇v|p−2∇v)=(α2(p−1)√α2ϕ2+ϕ′2p−2ϕ+(p−2)√α2ϕ2+ϕ′2p−4|ϕ′|2(α2ϕ+ϕ″)+√α2ϕ2+ϕ′2p−2(ϕ″+(d−2)ϕ′|x′|))eα(p−1)x1=√α2ϕ2+ϕ′2p−4(α2(p−1)(α2ϕ2+ϕ′2)ϕ+(p−2)ϕ′2(α2ϕ+ϕ″)+(α2ϕ2+ϕ′2)(ϕ″+(d−2)ϕ′|x′|))eα(p−1)x1. | (3.13) |
We show that v is a classical subsolution in M2i+1. Note that ϕ>0 and ϕ′,ϕ″<0 on (144−i,124−i).
We consider first the case p≥2. From ϕ>0, ϕ″<0 and ϕ′2≤α2ϕ2+ϕ′2, we deduce
(p−2)ϕ′2(α2ϕ+ϕ″)≥(p−2)(α2ϕ2+ϕ′2)ϕ″ |
and in combination with (3.13) that
∇⋅(|∇v|p−2∇v)≥√α2ϕ2+ϕ′2p−2(α2(p−1)ϕ+(p−1)ϕ″+(d−2)ϕ′|x′|)eα(p−1)x1. |
Hence, ∇⋅(|∇v|p−2∇v)≥0 on M2i+1 is equivalent to
α2(p−1)ϕ(r)+(p−1)ϕ″(r)+(d−2)ϕ′(r)r≥0for allr∈(144−i,124−i), |
which is by (3.3), (3.9) and (3.10) valid if and only if
α2(p−1)(i+1−(1−ηi)(4i+1r−1)2)−2(1−ηi)4i+1((p−1)4i+1+r−1(d−2)(4i+1r−1))≥0 |
for all r∈(144−i,124−i). We estimate with help of ηi∈[0,1],
α2(p−1)((i+1)−(1−ηi)(4i+1r−1)2)−2(1−ηi)4i+1((p−1)4i+1+r−1(d−2)(4i+1r−1))≥α2(p−1)i−2(1−ηi)42(i+1)(p−1+d−2). |
The lower bound on ηi≥η_i, see (3.5), implies 1−ηi≤1−η_i≤8−1(4p−2CQ)4−2i(i+1) and thus
α2(p−1)i−2(1−ηi)42(i+1)(p−1+d−2)≥α2(p−1)i−4p−1CQ(i+1)(p+d−3)≥0, | (3.14) |
where the last inequality is valid since (i+1)/i≤2 for i≥1 and α2≥4p−1CQ2(1+d−2p−1) (which is ensured by α≥α0, see (3.12)).
Next, we consider the case p∈(1,2). We deduce from (3.13) with p−2<0 and ϕ>0, ϕ′,ϕ″<0 that
∇⋅(|∇v|p−2∇v)≥√α2ϕ2+ϕ′2p−4((α2ϕ2+ϕ′2)(α2(p−1)ϕ+ϕ″+(d−2)ϕ′|x′|)−(2−p)ϕ′2ϕ)eα(p−1)x1. | (3.15) |
Similar computations as above yield for all r∈(144−i,124−i) and p∈(1,2)
α2(p−1)ϕ(r)+ϕ″(r)+(d−2)ϕ′(r)r≥α2(p−1)(i+ηi)−2(1−ηi)42(i+1)(d−1)(3.6)≥α2(p−1)i−4α(i+1)(d−1)≥1 |
where the last inequality is valid for all i≥1 and α≥8d−1p−1 (see (3.12)). Inserting this into (3.15), we obtain (using 2−p≤1)
∇⋅(|∇v|p−2∇v)≥√α2ϕ2+ϕ′2p−4(α2ϕ2−ϕ′2ϕ)eα(p−1)x1(3.9),(3.6)≥√α2ϕ2+ϕ′2p−4ϕ(α2ϕ−(α(i+1)4−i)2)eα(p−1)x1≥0, |
where we use in the last inequality that 4−2i(i+1)2≤1 and ϕ≥1 on (144−i,124−i) with i≥1.
Now, we show that v is a classical subsolution in M2i. In view of (3.13) it suffices to show that for all r∈(124−i,4−i) it holds
α4(p−1)ϕ3(r)+α2(2p−3)ϕ(r)ϕ′2(r)+ϕ′2((p−1)ϕ″(r)+d−2rϕ′(r))+α2ϕ2(r)(ϕ″(r)+d−2rϕ′(r))≥0 | (3.16) |
For p≥32, we obviously have
α4(p−1)ϕ3(r)+α2(2p−3)ϕ(r)ϕ′2(r)≥0for allr∈(124−i,4−i). |
Let us first consider p≥2. In the case d≥4, the choice of ϕ ensures
∀r∈(124−i,4−i):ϕ″(r)+d−2rϕ′(r)=0and(p−1)ϕ″(r)+d−2rϕ′(r)=(p−2)ϕ″(r)≥0 |
and similarly for d=3 that ϕ″(r)+d−2rϕ′(r)=12ϕ″(r)≥0 and (p−1)ϕ″(r)+d−2rϕ′(r)≥0. Altogether, we have that (3.16) is valid for all r∈(124−i,4−i) provided p≥2.
Next, we consider the case p∈(1,2). The choice of ϕ ensures
∀r∈(124−i,4−i):(p−1)ϕ″(r)+d−2rϕ′(r)=0andϕ″(r)+d−2rϕ′(r)=(2−p)ϕ″(r)≥0. |
Using the above two identities, we see that (3.16) is equivalent to
α4(p−1)ϕ3(r)+α2(2p−3)ϕ(r)ϕ′2(r)+α2ϕ2(r)(2−p)ϕ″(r)≥0 |
and thus it suffices to show
α2(2p−3)ϕϕ′2+α2ϕ2(2−p)ϕ″≥0. |
For p∈[32,2] the above inequality directly follows from ϕ,ϕ″≥0 and it is left to consider p∈(1,32) in which case the above inequality is equivalent to
3−2p2−pϕ′2ϕ″≤ϕ. |
The above inequality is valid on (124−i,4−i) provided i≥2. Indeed, this follows from ϕ≥i on (124−i,4−i) and
3−2p2−pϕ′2ϕ″≤3−2p2−pQQ+1ηi2Q−12Q≤2QQ+1≤2. |
Substep 2.2. Argument for (ii). Let α≥1 and j=j(α,d,p)≥2 be as in Substep 2.1.
In view of (3.8), we need to show that for all γ∈⋃i∈N,i≥j{4−i}∪{124−i} it holds
(ωθ√α2ϕ2+ϕ′2p−2ϕ′)(γ+)≥(ωθ√α2ϕ2+ϕ′2p−2ϕ′)(γ−). | (3.17) |
For γ∈⋃i∈N{4−i}, we directly observe that
(ωθ√α2ϕ2+ϕ′2p−2ϕ′)(γ+)=0>(ωθ√α2ϕ2+ϕ′2p−2ϕ′)(γ−). |
Moreover, the definition of ηi via (3.4) ensures that (3.17) holds as an equality for all γ∈⋃i∈N,i≥j{124−i} which finishes the argument.
Step 3. Let 1<p<∞ and θ∈[0,1] be such that (1−θ)p<d−1. Let α≥α0 and ρ=ρ(α,d,p)∈(0,1) be as in Step 2. We show that v is a weak subsolution on Ωρ:=B1∩{|x′|<ρ}.
We follow a similar reasoning as in [27]. For k∈N, let ψk∈C1(R;[0,1]) be a cut-off function satisfying
ψk=0on[0,124−k],ψk≡1on[4−k,1],‖ψ′k‖L∞(0,1)≤4k+1 |
and we define φk∈C1(B1) by φk(x)=ψk(|x′|). For every η∈C1c(Ωρ) with η≥0, we have
∫Ωρλθ|∇v|p−2∇v⋅∇ϕdx=∫Ωρλθ|∇v|p−2∇v⋅(∇((1−φk)η)+∇(φkη))dx≤∫Ωρλθ|∇v|p−2∇v⋅∇((1−φk)η)dx, | (3.18) |
where we use that 0≤φkη∈C1c(Ωρ∖Ω4−k−1) and that by Step 2 v is a subsolution on Ωρ∖Ωδ for every δ∈(0,ρ). It remains to show that the integral on the right-hand side in (3.18) vanishes as k→∞. Note that 0≤1−φk≤1 and 1−φk≡0 on Ωρ∖Ω4−k. Hence, with help of the product rule, we obtain
|∫Ωρλθ|∇v|p−2∇v⋅∇((1−φk)η)dx|≤∫Ω4−kλθ|∇v|p−1|∇η|dx+∫Ωρηλθ|∇v|p−2|∇v⋅∇φk|dx. |
By dominated convergence, the first term on the right-hand side converges to zero as k tends to ∞ (recall that we showed in Step 1 that λθ|∇v|p∈L1(B1)). To estimate the remaining integral we use |∇v⋅∇φk|=|ϕ′||∇φk|eαx1≤C4k+1|ϕ′| for some C=C(α)>0 on the set {|x′|∈(124−k,4−k)} and ∇v⋅∇φk=0 otherwise. Hence, we have that |∇v|p−2|∇v⋅∇φk|≤C4k+1|x′|−(p−1) on {|x′|∈(124−k,4−k)} and thus we obtain (using λθ=(k+1)θ(p−1)(2|x′|)pθ on {|x′|∈(124−k,4−k)}, see (3.1))
∫Ωρηλθ|∇v|p−2|∇v⋅∇φk|dx≤C‖η‖L∞(B1)4k+1(k+1)θ(p−1)∫4−k124−kr−(p−1)rpθrd−2dr=C‖η‖L∞(B1)4k+1(k+1)θ(p−1)1d−p(1−θ)4−k(d−p(1−θ))(1−2−(d−p(1−θ))k→∞→0, |
where we use p(1−θ)<d−1 the assumption and thus d−p(1−θ)>1.
Step 4. Conclusion.
Substep 4.1. We consider the case 1+1d−2<p<∞. Let s>1 and t>1p−1 be such that 1s+1t=pd−1 and tt+1p<d−1. We claim that there exist 0≤λ∈Ls(B1) with λ−1∈Lt(B1) and an unbounded weak subsolution to (1.5). We set θ=1td−1p and observe that 1s+1t=pd−1 implies θ∈[0,1] and 1−θ=1sd−1p. Moreover, the restriction tt+1p<d−1 in the form p<(1+1t)(d−1) ensures
(1−θ)p=(1−1td−1p)p=(p−1t(d−1))<d−1. |
Hence, in view of Steps 1–3, there exist the function v defined in (3.2) with α=α0=α0(p,d)≥1 such that v is an unbounded weak subsolution to
−∇⋅(λθ|∇v|p−2∇v)=0inB(0,ρ) withρ=ρ(d,p)∈(0,1], |
where λθ(x)=ωθ(|x′|), cf. (3.1). Appealing to (3.7), we have that there exists C=C(d,p)>0 such that
‖λθ‖Ls(B1)≤C(∫10(r−plog(2/r)−(p−1))d−1prd−2dr)1s=C(∫10r−1log(2/r)−(1−1p)(d−1)dr)1s<∞ |
where we use that p>1+1d−2 implies (1−1p)(d−1)>1. Similarly, we have
‖λ−1θ‖Lt(B1)≤C(∫10r−1log(2/r)−(1−1p)(d−1)dr)1t<∞. |
Finally, we observe that by a simple scaling argument namely considering ˜v(x)=v(x/ρ) and λ(x):=λθ(x/ρ) we find that ˜v is a weak subsolution to (1.5) in B1 and λ satisfies λ∈Ls(B1) and λ−1∈Lt(B1).
Substep 4.2. We consider 1<p≤1+1d−2. Let s and t be as in the statement of the theorem. Clearly, we find ¯s>s and ¯t>t such that 1¯s+1¯t=pd−1. Hence, for λθ with θ=1¯td−1p, we obtain as in Substep 4.1, an unbounded subsolution. It remains to check if λθ∈Ls(B1) and λ−1∈Lt(B1). By construction, we have 1−θ=1¯sd−1p and thus
‖λθ‖Ls(B1)≤C(∫10(r−plog(2/r)−(p−1))d−1ps¯srd−2dr)1s=C(∫10r−(d−1)s¯s+d−2log(2/r)−(1−1p)(d−1)s¯sdr)1s<∞, |
where we use s/¯s<1 and thus −(d−1)s¯s+d−2>−1. A similar argument shows λ−1θ∈Lt(B1) which finishes the argument.
PB was partially supported by the German Science Foundation DFG in context of the Emmy Noether Junior Research Group BE 5922/1-1. PB and MS thank Roberta Marziani for carefully reading parts of the manuscript.
The authors declare no conflict of interest.
Proof of Lemma 1. As a starting point we use [32, Lemma 2.1], which states for any δ∈(0,1]
J(ρ,σ,v)≤(σ−ρ)−(p−1+1δ)(∫σρ(∫Srμ|v|p dHd−1)δdr)1δ. |
With this at hand, we proceed in analogy to the Step 2 of Proof of [7, Lemma 2.1]:
Observe that the assumption s>1 implies s∗∈[1,d−1). To estimate the right-hand side, on each sphere we will use "scale-invariant" Sobolev inequality with α:=s∗ in the form
(∫Sr|ϕ|α∗)1α∗≤c((∫Sr|∇ϕ|α)1α+1r(∫Sr|ϕ|α)1α), |
which holds with c=c(d,α) with 1≤α<d−1, 1α∗=1α−1d−1 and any r>0. Moreover, observe that by Jensen inequality the previous estimate holds also if we change the exponent α∗ on the l.h.s. to a smaller exponent α′∈[1,α∗), while picking up a dimensional factor of |Sr|1α′−1α∗. Since by assumption r∈(ρ,σ)⊂[12,2], we can hide this factor into the constant c on the r.h.s.
The definition of s∗ implies that for α=s∗ holds pss−1≤α∗. Hence, for any δ∈(0,1] we estimate
(∫σρ(∫Srμ|v|p)δdr)1δ≤(∫σρ(∫Srμs)δs(∫Sr|v|pss−1)δs−1sdr)1δ≤c(∫σρ(∫Srμs)δs[(∫Sr|∇v|s∗)pδs∗+1rpδ(∫Sr|v|s∗)pδs∗]dr)1δ, |
with s∗ defined above. To be able to apply Hölder inequality in r to get two bulk integrals, we require δs+pδs∗=1. By choosing δ=(1+pd−1)−1∈(0,1) in the case s∗>1 and δ:=(1s+p)−1 if s∗=1, we obtain
J((ρ,σ,v)≤c(σ−ρ)pdd−1(∫Bσ∖Bρμs)1s[(∫Bσ∖Bρ|∇v|s∗)ps∗+1ρp(∫Bσ∖Bρ|v|s∗)ps∗] |
Observe that in the latter case of s∗=1 and δ=(1s+p) the correct prefactor is actually c(σ−ρ)−(2p−1+1s). Nevertheless, the estimate farther holds thanks to 2p−1+1s≥pdd−1, which in turn is equivalent to 1≤1p(1−1s)+1d−1 – the condition which is exactly fulfilled in this case.
[1] | D. Albritton, H. Dong, Regularity properties of passive scalars with rough divergence-free drifts, arXiv: 2107.12511. |
[2] |
P. Baroni, M. Colombo, G. Mingione, Regularity for general functionals with double phase, Calc. Var., 57 (2018), 62. https://doi.org/10.1007/s00526-018-1332-z doi: 10.1007/s00526-018-1332-z
![]() |
[3] |
A. K. Balci, L. Diening, M. Surnachev, New examples on Lavrentiev gap using fractals, Calc. Var., 59 (2020), 180. https://doi.org/10.1007/s00526-020-01818-1 doi: 10.1007/s00526-020-01818-1
![]() |
[4] |
A. K. Balci, L. Diening, R. Giova, A. Passarelli di Napoli, Elliptic equations with degenerate weights, SIAM J. Math. Anal., 54 (2022), 2373–2412. https://doi.org/10.1137/21M1412529 doi: 10.1137/21M1412529
![]() |
[5] | A. K. Balci, S. S. Byun, L. Diening, H. S. Lee, Global maximal regularity for equations with degenerate weights, arXiv: 2201.03524. |
[6] |
L. Beck, G. Mingione, Lipschitz bounds and non-uniform ellipticity, Commun. Pure Appl. Math., 73 (2020), 944–1034. https://doi.org/10.1002/cpa.21880 doi: 10.1002/cpa.21880
![]() |
[7] |
P. Bella, M. Schäffner, Local boundedness and Harnack inequality for solutions of linear nonuniformly elliptic equations, Commun. Pure Appl. Math., 74 (2021), 453–477. https://doi.org/10.1002/cpa.21876 doi: 10.1002/cpa.21876
![]() |
[8] |
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
![]() |
[9] | P. Bella, M. Schäffner, Lipschitz bounds for integral functionals with (p,q)-growth conditions, Adv. Calc. Var., in press. https://doi.org/10.1515/acv-2022-0016 |
[10] |
P. Bella, M. Schäffner, Non-uniformly parabolic equations and applications to the random conductance model, Probab. Theory Relat. Fields, 182 (2022), 353–397. https://doi.org/10.1007/s00440-021-01081-1 doi: 10.1007/s00440-021-01081-1
![]() |
[11] |
S. Biagi, G. Cupini, E. Mascolo, Regularity of quasi-minimizers for non-uniformly elliptic integrals, J. Math. Anal. Appl., 485 (2020), 123838. https://doi.org/10.1016/j.jmaa.2019.123838 doi: 10.1016/j.jmaa.2019.123838
![]() |
[12] |
V. Bögelein, F. Duzaar, M. Marcellini, C. Scheven, Boundary regularity for elliptic systems with p,q-growth, J. Math. Pure. Appl. (9), 159 (2022), 250–293. https://doi.org/10.1016/j.matpur.2021.12.004 doi: 10.1016/j.matpur.2021.12.004
![]() |
[13] |
D. Cao, T. Mengesha, T. Phan, Weighted-W1,p estimates for weak solutions of degenerate and singular elliptic equations, Indiana Univ. Math. J., 67 (2018), 2225–2277. https://doi.org/10.1512/iumj.2018.67.7533 doi: 10.1512/iumj.2018.67.7533
![]() |
[14] |
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
![]() |
[15] |
G. Cupini, P. Marcellini, E. Mascolo, Nonuniformly elliptic energy integrals with p,q-growth, Nonlinear Anal., 177 (2018), 312–324. https://doi.org/10.1016/j.na.2018.03.018 doi: 10.1016/j.na.2018.03.018
![]() |
[16] | G. Cupini, P. Marcellini, E. Mascolo, A. Passarelli di Napoli, Lipschitz regularity for degenerate elliptic integrals with p,q-growth, Adv. Calc. Var., in press. https://doi.org/10.1515/acv-2020-0120 |
[17] |
D. Cruz-Uribe, K. Moen, V. Naibo, Regularity of solutions to degenerate p-Laplacian equations, J. Math. Anal. Appl., 401 (2013), 458–478. https://doi.org/10.1016/j.jmaa.2012.12.023 doi: 10.1016/j.jmaa.2012.12.023
![]() |
[18] |
C. De Filippis, G. Mingione, On the regularity of minima of non-autonomous functionals, J. Geom. Anal., 30 (2020), 1584–1626. https://doi.org/10.1007/s12220-019-00225-z doi: 10.1007/s12220-019-00225-z
![]() |
[19] |
C. De Filippis, G. Mingione, Lipschitz bounds and nonautonomous integrals, Arch. Rational Mech. Anal., 242 (2021), 973–1057. https://doi.org/10.1007/s00205-021-01698-5 doi: 10.1007/s00205-021-01698-5
![]() |
[20] | C. De Filippis, G. Mingione, Nonuniformly elliptic Schauder theory, arXiv: 2201.07369. |
[21] | C. De Filippis, M. Piccinini, Borderline global regularity for nonuniformly elliptic systems, Int. Math. Res. Notices, 2022, rnac283. https://doi.org/10.1093/imrn/rnac283 |
[22] |
M. Eleuteri, P. Marcellini, E. Mascolo, Regularity for scalar integrals without structure conditions, Adv. Calc. Var., 13 (2020), 279–300. https://doi.org/10.1515/acv-2017-0037 doi: 10.1515/acv-2017-0037
![]() |
[23] |
L. Esposito, F. Leonetti, G. Mingione, Regularity results for minimizers of irregular integrals with (p,q) growth, Forum Math., 14 (2002), 245–272. https://doi.org/10.1515/form.2002.011 doi: 10.1515/form.2002.011
![]() |
[24] |
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
![]() |
[25] |
E. B. Fabes, C. E. Kenig, R. P. Serapioni, The local regularity of solutions to degenerate elliptic equations, Commun. Part. Diff. Eq., 7 (1982), 77–116. https://doi.org/10.1080/03605308208820218 doi: 10.1080/03605308208820218
![]() |
[26] |
I. Fonseca, J. Malý, G. Mingione, Scalar minimizers with fractal singular sets, Arch. Rational Mech. Anal., 172 (2004), 295–307. https://doi.org/10.1007/s00205-003-0301-6 doi: 10.1007/s00205-003-0301-6
![]() |
[27] |
B. Franchi, R. Serapioni, F. S. Cassano, Irregular solutions of linear degenerate elliptic equations, Potential Anal., 9 (1998), 201–216. https://doi.org/10.1023/A:1008684127989 doi: 10.1023/A:1008684127989
![]() |
[28] |
M. Giaquinta, Growth conditions and regularity, a counterexample, Manuscripta Math., 59 (1987), 245–248. https://doi.org/10.1007/BF01158049 doi: 10.1007/BF01158049
![]() |
[29] | D. Gilbarg, N. S. Trudinger, Elliptic partial differential equations of second order, Berlin: Springer, 2001. https://doi.org/10.1007/978-3-642-61798-0 |
[30] | P. Hästö, J. Ok, Maximal regularity for local minimizer of non-autonomous functionals. J. Eur. Math. Soc., 24 (2022), 1285–1334. https://doi.org/10.4171/JEMS/1118 |
[31] | J. Heinonen, T. Kilpeläinen, O. Martio, Nonlinear potential theory of degenerate elliptic equations, Mineola, NY: Dover Publications, Inc., 2006. |
[32] |
J. Hirsch, M. Schäffner, Growth conditions and regularity, an optimal local boundedness result, Commun. Contemp. Math., 23 (2021), 2050029. https://doi.org/10.1142/S0219199720500297 doi: 10.1142/S0219199720500297
![]() |
[33] | Q. Han, F. Lin, Elliptic partial differential equations, New York and Providence: New York University, Courant Institute of Mathematical Sciences and American Mathematical Society, 1997. |
[34] | O. Ladyzhenskaya, N. Ural'tseva, Linear and quasilinear elliptic equations, New York-London: Leon Ehrenpreis Academic Press, 1968. |
[35] |
P. Marcellini, Regularity of minimizers of integrals of the calculus of variations with nonstandard growth conditions, Arch. Rational Mech. Anal., 105 (1989), 267–284. https://doi.org/10.1007/BF00251503 doi: 10.1007/BF00251503
![]() |
[36] |
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
![]() |
[37] |
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
![]() |
[38] |
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
![]() |
[39] |
M. K. V. Murthy, G. Stampacchia, Boundary value problems for some degenerate-elliptic operators, Annali di Matematica Pura ed Applicata, 80 (1968), 1–122. https://doi.org/10.1007/BF02413623 doi: 10.1007/BF02413623
![]() |
[40] | A. Schwarzmann, Optimal boundedness results for degenerate elliptic equations, Thesis TU Dortmund, 2020. |
[41] |
J. Serrin, Local behavior of solutions of quasi-linear equations, Acta Math., 111 (1964), 247–302. https://doi.org/10.1007/BF02391014 doi: 10.1007/BF02391014
![]() |
[42] |
N. S. Trudinger, On Harnack type inequalities and their application to quasilinear elliptic equations, Commun. Pure Appl. Math., 20 (1967), 721–747. https://doi.org/10.1002/cpa.3160200406 doi: 10.1002/cpa.3160200406
![]() |
[43] |
N. S. Trudinger, On the regularity of generalized solutions of linear, non-uniformly elliptic equations, Arch. Rational Mech. Anal., 42 (1971), 50–62. https://doi.org/10.1007/BF00282317 doi: 10.1007/BF00282317
![]() |
[44] | N. S. Trudinger, Linear elliptic operators with measurable coefficients, Ann. Scuola Norm. Sup. Pisa, 27 (1973), 265–308. |
[45] | X. Zhang, Maximum principle for non-uniformly parabolic equations and applications, arXiv: 2012.05026. |