Citation: Claudia Lederman, Noemi Wolanski. Lipschitz continuity of minimizers in a problem with nonstandard growth[J]. Mathematics in Engineering, 2021, 3(1): 1-39. doi: 10.3934/mine.2021009
[1] | 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 |
[2] | 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 |
[3] | Elena Beretta, M. Cristina Cerutti, Luca Ratti . Lipschitz stable determination of small conductivity inclusions in a semilinear equation from boundary data. Mathematics in Engineering, 2021, 3(1): 1-10. doi: 10.3934/mine.2021003 |
[4] | 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 |
[5] | Antonio Vitolo . Singular elliptic equations with directional diffusion. Mathematics in Engineering, 2021, 3(3): 1-16. doi: 10.3934/mine.2021027 |
[6] | Alberto Farina . Some results about semilinear elliptic problems on half-spaces. Mathematics in Engineering, 2020, 2(4): 709-721. doi: 10.3934/mine.2020033 |
[7] | Francesca G. Alessio, Piero Montecchiari . Gradient Lagrangian systems and semilinear PDE. Mathematics in Engineering, 2021, 3(6): 1-28. doi: 10.3934/mine.2021044 |
[8] | Antonio Iannizzotto, Giovanni Porru . Optimization problems in rearrangement classes for fractional $ p $-Laplacian equations. Mathematics in Engineering, 2025, 7(1): 13-34. doi: 10.3934/mine.2025002 |
[9] | 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 |
[10] | Marco Cirant, Kevin R. Payne . Comparison principles for viscosity solutions of elliptic branches of fully nonlinear equations independent of the gradient. Mathematics in Engineering, 2021, 3(4): 1-45. doi: 10.3934/mine.2021030 |
To our dear friend Sandro Salsa on the occasion of his 70th birthday.
In this paper we study the regularity properties of nonnegative, local minimizers of the functional
J(v)=∫Ω(F(x,v,∇v)+λ(x)χ{v>0})dx, | (1.1) |
under nonstandard growth conditions of the energy function F(x,s,η) and 0<λmin≤λ(x)≤λmax<∞.
There has been a great deal of interest in these type of problems. Their study started with the seminal paper of Alt and Caffarelli [2] where the case F(x,s,η)=12|η|2 was considered. Later on, [3] considered the case F(x,s,η)=G(|η|2) under uniform ellipticity assumptions. The general power case F(x,s,η)=1p|η|p with 1<p<∞ was studied in [8], and F(x,s,η)=G(|η|) with G convex under the assumption that G′ satisfies Lieberman's condition namely, G″(t)∼G′(t)/t, was analyzed in [19]. The linear inhomogeneous case F(x,s,η)=12|η|2+f(x)s was addressed in [12] and [15].
The minimization problem for the functional (1.1) with F(x,s,η)=1p(x)|η|p(x) was first considered in [6] for p(x)≥2 and then, in [16] and [17] in the inhomogeneous case F(x,s,η)=1p(x)|η|p(x)+f(x)s, for 1<p(x)<∞ and f∈L∞(Ω). In [17], among other results, we proved that nonnegative local minimizers u are locally Lipschitz continuous and satisfy
Δp(x)u:=div(|∇u(x)|p(x)−2∇u)=fin {u>0}. |
The operator Δp(x), called the p(x)-Laplacian, extends the Laplacian, where p(x)≡2 and the p-Laplacian, where p(x)≡p. This is a prototype operator with nonstandard growth. The functional setting for the study of this type of operators are the variable exponent Lebesgue and Sobolev spaces Lp(⋅) and W1,p(⋅).
Functionals and PDEs with nonstandard growth have a wide range of applications, such as the modelling of non-Newtonian fluids, as for instance, electrorheological [21] or thermorheological fluids [4]. Other areas of application include non-linear elasticity [24], image reconstruction [1,7], the modelling of electric conductors [25], as well as processes of filtration of gases in non-homogeneous porous media [5].
As far as we know, no result on the minimization of (1.1) with F(x,s,η) a general function with nonstandard growth has been obtained.
The main purpose of our work is to prove the local Lipschitz continuity of nonnegative local minimizers of such an energy. We stress that this is the optimal regularity since it is known from the particular cases refered to above that the gradient of a minimizer u jumps across Ω∩∂{u>0}.
We prove that nonnegative minimizers of (1.1) are solutions to the associated equation in their positivity sets. That is, a local minimizer u≥0 satisfies
divA(x,u,∇u)=B(x,u,∇u) | (1.2) |
in {u>0}, where
A(x,s,η)=∇ηF(x,s,η),B(x,s,η)=Fs(x,s,η). |
Under our assumptions, the governing equation (1.2) is given by A(x,s,η) satisfying
λ0|η|p(x)−2|ξ|2≤∑i,j∂Ai∂ηj(x,s,η)ξiξj≤Λ0|η|p(x)−2|ξ|2, |
and has a right hand side given by B(x,s,η)≢0 of p(x)-type growth in η. This equation is singular in the regions where 1<p(x)<2 and degenerate in the ones where p(x)>2.
Our study thus presents new features, needed in order to overcome the deep technical difficulties arising due to the nonlinear degenerate/singular nature and the x and s dependence of this general operator associated to our energy functional (1.1).
The first part of the paper is devoted to the study of equation (1.2) in a domain Ω, under nonstandard growth conditions of p(x)-type. We prove existence results, a comparison principle, a uniqueness result, a maximum principle and other local L∞ bounds of solutions of this equation. These delicate results are of independent interest.
Some of these results are obtained under the growth assumption (3.14). We remark that this hypothesis on the functions A and B allows to consider very general equations. This condition not only enables us to get the inequality in Proposition 3.3 that is a main tool for all the proofs in the paper, but also it is invariant under rescalings. All these results are included in Section 3.
In the second part of the paper we deal with the minimization problem for the functional (1.1). In fact, in Section 4 we first get an existence result for minimizers. We also prove nonnegativity and boundedness, under suitable assumptions. Then, we prove the local Hölder and Lipschitz continuity of nonnegative local minimizers (Theorems 4.3 and 4.5).
The proofs in Section 4 involve delicate rescalings. One of the main difficulties this problem presents is that it is not invariant under the rescaling u(x)⟼u(tx)k, if t≠k —rescaling that is a crucial tool in dealing with this type of problems. The rescaled functionals lose the uniform properties and nontrivial modifications are needed to get through the proofs. Even after these modifications, there is in general no limit equation for the rescaled problems due to the growth we are allowing to the function B(x,s,η). Novel arguments are used to complete the proof of Theorem 4.4. In fact, we are able to show that, although there is in general no limit equation for the rescaled problems, there is a limit function and it satisfies Harnack's inequality (see (4.58)).
A thorough follow up of the dependence of the bounds found in Section 3 with respect to the structural conditions on F,A and B is of most importance as well.
Let us point out that the results in the paper are new even in the case p(x)≡p constant.
Finally, in Section 5 we present some examples of functionals (1.1) where our results can be used.
Our examples include functionals (1.1) involving energy functions of the form
F(x,s,η)=a(x,s)|η|p(x)p(x)+f(x,s). |
A possible example of admissible functions a(x,s),f(x,s) is given by
a(x,s)=a0(x)(1+s)−q(x),a0(x)>0,0<q(x)≤q0(x), |
for s in the range where the nonnegative local minimizer takes values, q0(x) a function depending on p(x) and
f(x,s)=b(x)|s|τ(x),b(x)≥0,τ(x)≥2, |
with τ(x) satisfying (2.7).
Our results also apply to functionals (1.1) involving energy functions of the form
F(x,s,η)=G(x,η)+f(x,s). |
Some admissible G(x,η),f(x,s) are
G(x,η)=a(x)˜G(|η|p(x))a(x)>0,˜G″≥0, |
G(x,η)=˜A(x)η⋅η|η|p(x)−2˜A(x)∈RN×N uniformly elliptic, |
f(x,s)=g(x)s. |
Also,
F(x,s,η)=a1(x)F1(x,s,η)+a2(x)F2(x,s,η),ai(x)>0, |
is an admissible function if both F1(x,s,η) and F2(x,s,η) are admissible.
We begin our paper with a section where we state the hypotheses on F,A, B, λ and p(x) that will be used throught the article. And we end it with an Appendix where we state some properties of the function spaces Lp(⋅) and W1,p(⋅) where the problem is well posed.
Let p:Ω→[1,∞) be a measurable bounded function, called a variable exponent on Ω and denote pmax=esssupp(x) and pmin=essinfp(x). We define the variable exponent Lebesgue space Lp(⋅)(Ω) to consist of all measurable functions u:Ω→R for which the modular ϱp(⋅)(u)=∫Ω|u(x)|p(x)dx is finite. We define the Luxemburg norm on this space by
‖u‖Lp(⋅)(Ω)=‖u‖p(⋅)=inf{λ>0:ϱp(⋅)(u/λ)≤1}. |
This norm makes Lp(⋅)(Ω) a Banach space.
There holds the following relation between ϱp(⋅)(u) and ‖u‖Lp(⋅):
min{(∫Ω|u|p(x)dx)1/pmin,(∫Ω|u|p(x)dx)1/pmax}≤‖u‖Lp(⋅)(Ω)≤max{(∫Ω|u|p(x)dx)1/pmin,(∫Ω|u|p(x)dx)1/pmax}. |
Moreover, the dual of Lp(⋅)(Ω) is Lp′(⋅)(Ω) with 1p(x)+1p′(x)=1.
Let W1,p(⋅)(Ω) denote the space of measurable functions u such that u and the distributional derivative ∇u are in Lp(⋅)(Ω). The norm
‖u‖1,p(⋅):=‖u‖p(⋅)+‖|∇u|‖p(⋅) |
makes W1,p(⋅)(Ω) a Banach space.
The space W1,p(⋅)0(Ω) is defined as the closure of the C∞0(Ω) in W1,p(⋅)(Ω).
For the sake of completeness we include in an Appendix at the end of the paper some additional results on these spaces that are used throughout the paper.
∙N spatial dimension
∙ |S| N-dimensional Lebesgue measure of the set S
∙ Br(x0) open ball of radius r and center x0
∙ Br open ball of radius r and center 0
∙χS characteristic function of the set S
∙u+= max(u,0), u−= max(−u,0)
∙⟨ξ,η⟩ and ξ⋅η both denote scalar product in RN
In this section we collect all the assumptions that will be made along the paper.
Throughout the paper Ω will denote a C1 bounded domain in RN. In addition, the following assumptions will be made:
We assume that the function p(x) is measurable in Ω and verifies
1<pmin≤p(x)≤pmax<∞,x∈Ω. |
We assume further that p(x) is Lipschitz continuous in Ω and we denote by L the Lipschitz constant of p(x), namely, ‖∇p‖L∞(Ω)≤L.
When we are restricted to a ball Br we use p−r and p+r to denote the infimum and the supremum of p(x) over Br.
We assume that the function λ(x) is measurable in Ω and verifies
0<λmin≤λ(x)≤λmax<∞,x∈Ω. |
We assume that F is measurable in ¯Ω×R×RN, and for every x∈¯Ω, F(x,⋅,⋅)∈C1(R×RN)∩C2(R×RN∖{0}).
We denote A(x,s,η)=∇ηF(x,s,η) and B(x,s,η)=Fs(x,s,η).
We assume that A∈C(¯Ω×R×RN,RN) and for every x∈¯Ω, A(x,⋅,⋅)∈C1(R×RN∖{0},RN). Moreover, there exist positive constants λ0 and Λ0, and β∈(0,1) such that for every x,x1,x2∈¯Ω, s,s1,s2∈R, η∈RN∖{0} and ξ∈RN, the following conditions are satisfied:
A(x,s,0)=0, | (2.1) |
∑i,j∂Ai∂ηj(x,s,η)ξiξj≥λ0|η|p(x)−2|ξ|2, | (2.2) |
∑i,j|∂Ai∂ηj(x,s,η)|≤Λ0|η|p(x)−2, | (2.3) |
|A(x1,s,η)−A(x2,s,η)|≤Λ0|x1−x2|β(|η|p(x1)−1+|η|p(x2)−1)(1+|log|η||), | (2.4) |
|A(x,s1,η)−A(x,s2,η)|≤Λ0|s1−s2||η|p(x)−1. | (2.5) |
We assume that B is measurable in ¯Ω×R×RN and for every x∈¯Ω, B(x,⋅,⋅)∈C1(R×RN), and for every (x,s,η)∈¯Ω×R×RN,
|B(x,s,η)|≤Λ0(1+|η|p(x)+|s|τ(x)), | (2.6) |
where Λ0 is as in the assumptions on A and
τ(x)≥p(x)andτ∈C(¯Ω),τ(x)≤p∗(x)=Np(x)N−p(x) if pmax<N,τ(x) arbitrary if pmin>N,τ(x)=p(x) if pmin≤N≤pmax. | (2.7) |
Remark 2.1. From (2.1) and (2.3) we get
|Ai(x,s,η)|=|Ai(x,s,η)−Ai(x,s,0)|=|∫10∑j∂Ai∂ηj(x,s,tη)ηjdt,|≤ˉα(pmin)Λ0|η|p(x)−1, |
so that
|A(x,s,η)|≤ˉα(pmin)NΛ0|η|p(x)−1. | (2.8) |
From (2.1) and (2.2) we have
A(x,s,η)⋅η=(A(x,s,η)−A(x,s,0))⋅η=∫10∑ij∂Ai∂ηj(x,s,tη)ηjηidt, |
so that
A(x,s,η)⋅η≥α(pmax)λ0|η|p(x). | (2.9) |
In this section we consider A and B as in Section 2 and we prove results for solutions of the equation
divA(x,u,∇u)=B(x,u,∇u) in Ω. | (3.1) |
Namely, existence, comparison principle, uniqueness, maximum principle and bounds of solutions.
Our first result is Proposition 3.1, were we prove existence of a solution to (3.1) with given boundary data. In order to prove the existence of a solution to (3.1) we show that, given u∈W1,p(⋅)(Ω), there exists a minimizer of the functional
JΩ(v)=∫ΩF(x,v,∇v)dx | (3.2) |
in u+W1,p(⋅)0(Ω), where F is as in Section 2, A(x,s,η)=∇ηF(x,s,η) and B(x,s,η)=Fs(x,s,η).
Then, in Proposition 3.2 we get an existence result under a growth assumption on the function F stronger than (3.3) in Proposition 3.1, but without the small oscillation hypothesis there.
In Proposition 3.4 and Corollary 3.2 we prove comparison and uniqueness for this problem, assuming that condition (3.14) below holds. In Proposition 3.5 we prove that solutions to (3.1) with bounded boundary data are bounded and in Proposition 3.6 we prove a maximum principle for this problem, under suitable assumptions. In Proposition 3.7 we give another existence result of a bounded solution.
We start with the definition of solution to (3.1).
Definition 3.1. Let p, A and B be as in Section 2. We say that u is a solution to (3.1) if u∈W1,p(⋅)(Ω) and, for every φ∈C∞0(Ω), there holds that
−∫ΩA(x,u,∇u)⋅∇φdx=∫ΩB(x,u,∇u)φdx. |
We are using that, under the conditions in (2.7), the embedding theorem (see Theorem A.5) applies.
Our first existence result is
Proposition 3.1. Let p,F,A,B as in Section 2 and let Ω′⊂Ω be a C1 domain. Let u∈W1,p(⋅)(Ω′) and let us call p+=supΩ′p(x), p−=infΩ′p(x). Assume that there exist μ,c1∈R+, pmin>δ>0 and g∈L1(Ω) such that
F(x,s,η)≥μ|η|p(x)−c1|s|p(x)−δ−g(x)inΩ. | (3.3) |
Assume, moreover that δ>p+−p− and that
F(x,s,η)≤μ−1|η|p(x)+c1|s|τ(x)+g(x)inΩ, | (3.4) |
with τ satisfying (2.7).
Then, there exists a solution v∈u+W1,p(⋅)0(Ω′) to (3.1) in Ω′.
Moreover, ‖v‖W1,p(⋅)(Ω′)≤C, for a constant C depending only ‖u‖W1,p(⋅)(Ω′), ‖g‖L1(Ω′), |Ω′|, diam(Ω′), N, p−, p+, δ, L, μ, c1, ||τ||L∞(Ω′) and the C1 norm of ∂Ω′.
Proof. We will show that there is a minimizer of JΩ′ in u+W1,p(⋅)0(Ω′) where
JΩ′(v)=∫Ω′F(x,v,∇v)dx. |
This minimizer is a solution to the associated Euler-Lagrange equation (3.1) in Ω′.
We will use the embedding theorem (see Theorem A.5) that states that, under the conditions in (2.7), W1,p(⋅)(Ω′)↪Lτ(⋅)(Ω′) continuously.
So, let vn be a minimizing sequence. That is, vn∈u+W1,p(⋅)0(Ω′) and
I=limn→∞JΩ′(vn)=infu+W1,p(⋅)0(Ω′)JΩ′(v)≤∫Ω′F(x,u,∇u)dx. |
Let us show that there is a constant κ>0 such that ‖vn‖Lp(⋅)(Ω′)≤κ. In fact, by (3.3), for n large,
∫Ω′|∇vn|p(x)dx≤1+∫Ω′F(x,u,∇u)dx+c1μ∫Ω′|vn|p(x)−δdx+1μ∫Ω′g(x)dx. |
By Poincare's inequality (Theorem A.4)
‖vn−u‖Lp(⋅)(Ω′)≤CΩ′‖∇(vn−u)‖Lp(⋅)(Ω′). |
Hence, recalling Proposition A.1,
‖vn‖Lp(⋅)(Ω′)≤‖u‖Lp(⋅)(Ω′)+CΩ′[‖∇vn‖Lp(⋅)(Ω′)+‖∇u‖Lp(⋅)(Ω′)]≤C[‖u‖W1,p(⋅)(Ω′)+max{(∫Ω′|∇vn|p(x)dx)1/p−,(∫Ω′|∇vn|p(x)dx)1/p+}]≤ˉC[1+max{(∫Ω′|vn|p(x)−δdx)1/p−,(∫Ω′|vn|p(x)−δdx)1/p+}] |
with ˉC depending on ‖u‖W1,p(⋅)(Ω′),‖g‖L1(Ω′),N,p−, p+, δ, |Ω′|, diam(Ω′), L, ||τ||L∞(Ω′), the C1 norm of ∂Ω′, and the constants in (3.3).
Observe that in case u≡M, there holds that ∫Ω′F(x,u,∇u)dx is bounded by a constant that depends only on M, ‖τ‖L∞(Ω′) and |Ω′|. Hence, in that case ˉC is independent of the regularity of Ω′.
Since we want to find a uniform bound of ‖vn‖Lp(⋅)(Ω′), we may assume that this norm is larger than 1. Let q be the middle point of the interval [p+−δ,p−]. By Young's inequality with r(x)=qp(x)−δ,
∫Ω′|vn|p(x)−δdx≤Cε+ε∫Ω′|vn|qdx, |
for 0<ε<1 with Cε depending only on |Ω′|,ε,p−,p+ and δ. On the other hand, since ‖vn‖Lq(Ω′)≤C‖vn‖Lp(⋅)(Ω′) with C depending only on |Ω′|,p−,p+ and δ,
∫Ω′|vn|qdx≤(C‖vn‖Lp(⋅)(Ω′))q. |
So that
‖vn‖Lp(⋅)(Ω′)≤C[˜Cε+ε1p+(‖vn‖Lp(⋅)(Ω′))qp−]≤C[˜Cε+ε1p+‖vn‖Lp(⋅)(Ω′)]. |
By choosing ε small enough, we find that
‖vn‖Lp(⋅)(Ω′)≤C | (3.5) |
with C depending on |Ω′|, diam(Ω′), ‖u‖W1,p(⋅)(Ω′),p−,p+, N, δ, ‖g‖L1(Ω′), L, ||τ||L∞(Ω′), the C1 norm of ∂Ω′, μ and c1.
From the computations above we find that ∫Ω′|vn|p(x)−δdx≤C1. So that we have that I>−∞ and
‖∇vn‖Lp(⋅)(Ω′)≤C2, | (3.6) |
with C2 depending on |Ω′|, diam(Ω′), ‖u‖W1,p(⋅)(Ω′),p−,p+, N, δ, ‖g‖L1(Ω′), L, ||τ||L∞(Ω′), the C1 norm of ∂Ω′, μ and c1.
From our comment above, we have that in case u≡M in Ω′, the constant C2 is independent of the regularity of ∂Ω′.
Let us proceed with the proof of the existence of a minimizer. By the estimates above, for a subsequence that we still call vn, there holds that there exists v∈u+W1,p(⋅)0(Ω′), such that
vn⇀vinW1,p(⋅)(Ω′),vn→vinLp−(Ω′)and almost everywhere, |
and such that the bounds (3.5) and (3.6) also hold for v.
By Egorov's Theorem, for every ε>0 there exists Ωε such that |Ω′∖Ωε|<ε and vn→v uniformly in Ωε.
On the other hand, if we set ΩK={x∈Ω′/|v|+|∇v|≤K}, there holds that |Ω′∖ΩK|→0 as K→∞.
Let Ωε,K=Ωε∩ΩK. Then, |Ω′∖Ωε,K|→0 as ε→0 and K→∞.
There holds
lim supn→∞∫Ωε,KF(x,vn,∇vn)dx≤I+c1∫Ω′∖Ωε,K|v|p(x)−δdx+∫Ω′∖Ωε,Kgdx. | (3.7) |
Let us prove that
∫Ωε,KF(x,v,∇v)dx≤I+c1∫Ω′∖Ωε,K|v|p(x)−δdx+∫Ω′∖Ωε,Kgdx. |
In fact,
∫Ωε,KF(x,vn,∇vn)dx−∫Ωε,KF(x,v,∇v)dx=∫Ωε,K[F(x,vn,∇vn)−F(x,vn,∇v)]dx+∫Ωε,K[F(x,vn,∇v)−F(x,v,∇v)]dx=A+B. |
On the one hand, B→0 since F(x,vn,∇v)−F(x,v,∇v)→0 uniformly in Ωε,K and it is uniformly bounded. On the other hand, by the convexity assumption on F(x,s,η) with respect to η,
A≥∫Ωε,KA(x,vn,∇v)⋅(∇vn−∇v)dx→0asn→∞ |
since A(x,vn,∇v)→A(x,v,∇v) uniformly in Ωε,K, they are uniformly bounded and ∇vn⇀∇v weakly in Lp(⋅)(Ωε,K).
Hence, for every ε,K,
∫Ωε,KF(x,v,∇v)dx≤I+c1∫Ω′∖Ωε,K|v|p(x)−δdx+∫Ω′∖Ωε,Kgdx. |
Now, by letting ε→0 and K→∞, we get
∫Ω′F(x,v,∇v)dx≤I, |
and therefore, v is a minimizer of JΩ′ in u+W1,p(⋅)0(Ω′) and a solution to (3.1).
As a corollary of Proposition 3.1 we have the following existence result that will be used in the next section.
Corollary 3.1. Let p,F,A,B as in Section 2 and let Ω′⊂Ω be a C1 domain. Let u∈W1,p(⋅)(Ω′) and let us call p+=supΩ′p(x), p−=infΩ′p(x). Assume that there exist μ,c1∈R+ and pmin>δ>0 such that
F(x,s,η)≥μ|η|p(x)−c1(|s|p(x)−δ+1)inΩ. | (3.8) |
Assume, moreover that δ>p+−p− and that
F(x,s,η)≤μ−1|η|p(x)+c1(|s|τ(x)+1)inΩ, | (3.9) |
with τ(x) satisfying (2.7).
Then, there exists a solution v∈u+W1,p(⋅)0(Ω′) to (3.1) in Ω′ and ‖v‖W1,p(⋅)(Ω′)≤C, for a constant C depending only ‖u‖W1,p(⋅)(Ω′), |Ω′|, diam(Ω′), N, p−, p+, δ, L, μ, c1, ||τ||L∞(Ω′) and the C1 norm of ∂Ω′.
With a stronger growth assumption on the s variable for the function F(x,s,η) we get an existence result without the small oscillation assumption of the function p.
Proposition 3.2. Let p,F,A,B as in Section 2 and let Ω′⊂Ω be a C1 domain. Let u∈W1,p(⋅)(Ω′). Assume that there exist μ,c1∈R+, g∈L1(Ω) and 1≤q<pmin such that
F(x,s,η)≥μ|η|p(x)−c1|s|q−g(x)inΩ. | (3.10) |
Assume, moreover that
F(x,s,η)≤μ−1|η|p(x)+c1|s|τ(x)+g(x)inΩ, | (3.11) |
with τ satisfying (2.7).
Then, there exists a solution v∈u+W1,p(⋅)0(Ω′) to (3.1) in Ω′ and ‖v‖W1,p(⋅)(Ω′)≤C, for a constant C depending only ‖u‖W1,p(⋅)(Ω′), ‖g‖L1(Ω′), |Ω′|, diam(Ω′), N, pmin, pmax, q, L, μ, c1, ||τ||L∞(Ω′) and the C1 norm of ∂Ω′.
Proof. We proceed as in the proof of Proposition 3.1 and we prove that a minimizing sequence {vn} satisfies
μ∫Ω′|∇vn|p(x)dx≤∫Ω′F(x,u,∇u)+1+∫Ω′g(x)dx+c1∫Ω′|vn|qdx. | (3.12) |
We want to prove that there is a constant such that ∫Ω′|∇vn|p(x)dx≤C. So, we can assume that ∫Ω′|∇vn|p(x)dx>1.
Thus,
‖vn‖Lq(Ω′)≤C‖vn‖Lp(⋅)(Ω′)≤C[‖u‖W1,p(⋅)(Ω′)+‖∇vn‖Lp(⋅)(Ω′)]≤C[‖u‖W1,p(⋅)(Ω′)+(∫Ω′|∇vn|p(x)dx)1/pmin], |
where C depends on q,pmin,pmax, N, L and |Ω′|, diam(Ω′). Hence, as q<pmin,
∫Ω′|vn|qdx≤C(1+(∫Ω′|∇vn|p(x)dx)q/pmin)≤˜C+ε∫Ω′|∇vn|p(x)dx | (3.13) |
with C depending only on q,pmin,pmax,N,|Ω′|,diam(Ω′),L,‖u‖W1,p(⋅)(Ω′), and ˜C depending on the same constants and also on ε.
Thus, by (3.12) and (3.13),
∫Ω′|∇vn|p(x)dx≤ˆC |
with ˆC depending only on q,pmin,pmax,N,μ,|Ω′|,diam(Ω′),L,∫Ω′g(x)dx, c1, ||τ||L∞(Ω′), the C1 norm of ∂Ω′ and ‖u‖W1,p(⋅)(Ω′).
Now, as in the proof of Proposition 3.1, we get that there exists a subsequence that we still call {vn} and a function v∈u+W1,p(⋅)0(Ω′) such that
vn→vinLpmin(Ω′),vn⇀vweakly in W1,p(⋅)(Ω′). |
Now, the proof follows as that of Proposition 3.1.
We next prove a result valid for solutions of equation (3.1) that will be of use in the proofs of Hölder and Lipschitz continuity of minimizers of the energy functional (1.1)
Proposition 3.3. Let p,F,A and B be as in Section 2. Assume moreover that
2|As(x,s,η)⋅ξw|≤12∑i,j∂Ai∂ηj(x,s,η)ξiξj+Bs(x,s,η)w2, | (3.14) |
for every (x,s,η)∈¯Ω×R×RN∖{0}, ξ∈RN and w∈R.
Let u∈W1,p(⋅)(Ω)∩L∞(Ω) and let v∈W1,p(⋅)(Ω)∩L∞(Ω) be such that
{divA(x,v,∇v)=B(x,v,∇v)inΩ,v=uon∂Ω. | (3.15) |
Then,
∫Ω(F(x,u,∇u)−F(x,v,∇v))dx≥12αλ0(∫Ω∩{p(x)≥2}|∇u−∇v|p(x)dx+∫Ω∩{p(x)<2}(|∇u|+|∇v|)p(x)−2|∇u−∇v|2dx), | (3.16) |
where α=α(pmin,pmax) and λ0 is as in (2.2).
Proof. For 0≤σ≤1, let uσ=v+σ(u−v). Then, denoting ∇ηF=A and Fs=B, we obtain
∫Ω(F(x,u,∇u)−F(x,v,∇v))dx=∫10∫ΩA(x,uσ,∇uσ)⋅∇(uσ−v)1σdxdσ+∫10∫ΩB(x,uσ,∇uσ)(uσ−v)1σdxdσ=∫10∫Ω(A(x,uσ,∇uσ)−A(x,v,∇v))⋅∇(uσ−v)1σdxdσ+∫10∫Ω(B(x,uσ,∇uσ)−B(x,v,∇v))(uσ−v)1σdxdσ=I+II, | (3.17) |
where we have used (3.15). Moreover,
I=∫10∫10∫|∇v|≥|∇uσ|As(x,uστ,∇uστ)⋅∇(uσ−v)(uσ−v)1σdxdσdτ+∫10∫10∫|∇v|<|∇uσ|As(x,uσ(1−τ),∇uσ(1−τ))⋅∇(uσ−v)(uσ−v)1σdxdσdτ+∫10∫10∫|∇v|≥|∇uσ|∑i,j∂Ai∂ηj(x,uστ,∇uστ)(uσ−v)xi(uσ−v)xj1σdxdσdτ+∫10∫10∫|∇v|<|∇uσ|∑i,j∂Ai∂ηj(x,uσ(1−τ),∇uσ(1−τ))(uσ−v)xi(uσ−v)xj1σdxdσdτ=I1+I2+I3+I4. | (3.18) |
Now, using (2.2), and the inequality
|η′+t(η−η′)|≥14|η−η′|,for|η′|≥|η|, 0≤t≤14, | (3.19) |
we get
I3+I4≥∫10∫10∫|∇v|≥|∇uσ|λ0|∇uστ|p(x)−2|∇(uσ−v)|21σdxdσdτ+∫10∫10∫|∇v|<|∇uσ|λ0|∇uσ(1−τ)|p(x)−2|∇(uσ−v)|21σdxdσdτ≥αλ0(∫{p(x)≥2}|∇u−∇v|p(x)dx+∫{p(x)<2}(|∇u|+|∇v|)p(x)−2|∇u−∇v|2dx), | (3.20) |
where α=α(pmin,pmax) and λ0 is as in (2.2). On the other hand,
II=∫10∫10∫|∇v|≥|∇uσ|Bs(x,uστ,∇uστ)(uσ−v)21σdxdσdτ+∫10∫10∫|∇v|<|∇uσ|Bs(x,uσ(1−τ),∇uσ(1−τ))(uσ−v)21σdxdσdτ+∫10∫10∫|∇v|≥|∇uσ|∇ηB(x,uστ,∇uστ)⋅∇(uσ−v)(uσ−v)1σdxdσdτ+∫10∫10∫|∇v|<|∇uσ|∇ηB(x,uσ(1−τ),∇uσ(1−τ))⋅∇(uσ−v)(uσ−v)1σdxdσdτ. | (3.21) |
Finally, using that As(x,s,η)=∇ηB(x,s,η), the assumption (3.14) and estimates (3.17), (3.18), (3.20) and (3.21), we get (3.16).
We now prove a comparison principle for equation (3.1), which holds under assumption (3.14).
Proposition 3.4. Let p,A and B be as in Section 2. Assume moreover that condition (3.14) holds. Let u,v∈W1,p(⋅)(Ω) be such that
divA(x,u,∇u)≥B(x,u,∇u)inΩ,divA(x,v,∇v)≤B(x,v,∇v)inΩ,u≤von∂Ω. | (3.22) |
Then,
u≤vinΩ. | (3.23) |
Proof. We will use arguments similar to those in Proposition 3.3. In fact, for R>0 we consider the nonnegative function wR∈W1,p(⋅)0(Ω)∩L∞(Ω) given by
wR={0if u−v≤0,u−vif 0<u−v<R,Rif u−v≥R, | (3.24) |
and by (3.22) we have
0≥∫Ω(A(x,u,∇u)−A(x,v,∇v))⋅∇wRdx+∫Ω(B(x,u,∇u)−B(x,v,∇v))wRdx=I+II. | (3.25) |
Then, denoting ΩR=Ω∩{0<u−v<R} and, for 0≤τ≤1, uτ=v+τ(u−v), we get
I=∫10∫ΩR∩{|∇v|≥|∇u|}As(x,uτ,∇uτ)⋅∇(u−v)(u−v)dxdτ+∫10∫ΩR∩{|∇v|<|∇u|}As(x,u(1−τ),∇u(1−τ))⋅∇(u−v)(u−v)dxdτ+∫10∫ΩR∩{|∇v|≥|∇u|}∑i,j∂Ai∂ηj(x,uτ,∇uτ)(u−v)xi(u−v)xjdxdτ+∫10∫ΩR∩{|∇v|<|∇u|}∑i,j∂Ai∂ηj(x,u(1−τ),∇u(1−τ))(u−v)xi(u−v)xjdxdτ=I1+I2+I3+I4. | (3.26) |
Now, proceeding as in Proposition 3.3, we obtain
I3+I4≥∫10∫ΩR∩{|∇v|≥|∇u|}λ0|∇uτ|p(x)−2|∇(u−v)|2dxdτ+∫10∫ΩR∩{|∇v|<|∇u|}λ0|∇u(1−τ)|p(x)−2|∇(u−v)|2dxdτ≥˜αλ0(∫ΩR∩{p(x)≥2}|∇u−∇v|p(x)dx+∫ΩR∩{p(x)<2}(|∇u|+|∇v|)p(x)−2|∇u−∇v|2dx), | (3.27) |
where ˜α=˜α(pmin,pmax) and λ0 is as in (2.2).
On the other hand, we observe that the evaluation of (3.14) in ξ=0 implies that Bs(x,s,η)≥0. Then, we get
II≥∫10∫ΩR∩{|∇v|≥|∇u|}Bs(x,uτ,∇uτ)(u−v)2dxdτ+∫10∫ΩR∩{|∇v|<|∇u|}Bs(x,u(1−τ),∇u(1−τ))(u−v)2dxdτ+∫10∫ΩR∩{|∇v|≥|∇u|}∇ηB(x,uτ,∇uτ)⋅∇(u−v)(u−v)dxdτ+∫10∫ΩR∩{|∇v|<|∇u|}∇ηB(x,u(1−τ),∇u(1−τ))⋅∇(u−v)(u−v)dxdτ+∫10∫{u−v>R}∩{|∇v|≥|∇u|}Bs(x,uτ,∇uτ)w2Rdxdτ+∫10∫{u−v>R}∩{|∇v|<|∇u|}Bs(x,u(1−τ),∇u(1−τ))w2Rdxdτ+∫10∫{u−v>R}∩{|∇v|≥|∇u|}∇ηB(x,uτ,∇uτ)⋅∇(u−v)wRdxdτ+∫10∫{u−v>R}∩{|∇v|<|∇u|}∇ηB(x,u(1−τ),∇u(1−τ))⋅∇(u−v)wRdxdτ. | (3.28) |
Now, using that As(x,s,η)=∇ηB(x,s,η), (2.3), (3.19), assumption (3.14) and estimates (3.25), (3.26), (3.27) and (3.28), we get
0≥12˜αλ0(∫ΩR∩{p(x)≥2}|∇u−∇v|p(x)dx+∫ΩR∩{p(x)<2}(|∇u|+|∇v|)p(x)−2|∇u−∇v|2dx)−ˆαΛ0(∫{u−v>R}∩{p(x)≥2}(|∇u|+|∇v|)p(x)dx+∫{u−v>R}∩{p(x)<2}|∇u−∇v|p(x)dx), | (3.29) |
where ˆα=ˆα(pmin,pmax) and Λ0 is as in (2.3). Since R>0 is arbirtrary, we can use that u,v∈W1,p(⋅)(Ω) and let R→∞ and we obtain
0≥12˜αλ0(∫Ω∩{p(x)≥2}|∇(u−v)+|p(x)dx+∫Ω∩{p(x)<2}(|∇u|+|∇v|)p(x)−2|∇(u−v)+|2dx), | (3.30) |
which implies that ∇(u−v)+=0 in Ω. Since (u−v)+∈W1,p(⋅)0(Ω), Poincare's inequality (Theorem A.4) gives (u−v)+=0 in Ω. That is, (3.23) holds.
As a corollary of Propostion 3.4 we obtain the following uniqueness result
Corollary 3.2. Let p,A and B be as in Section 2. Assume moreover that condition (3.14) holds. Let φ∈W1,p(⋅)(Ω) and let u1,u2∈W1,p(⋅)(Ω) be such that
{divA(x,ui,∇ui)=B(x,ui,∇ui)inΩ,ui=φon∂Ω, | (3.31) |
for i=1,2. Then, u1=u2 in Ω.
We next prove that solutions to (3.1) with bounded boundary data are bounded, under the assumptions of Proposition 3.1.
Proposition 3.5. Let p,A and B be as in Section 2 and let Ω′⊂Ω be a C1 domain. Assume moreover, that conditions (3.3), (3.4) and (3.14) hold in Ω′ for some p+−p−<δ<pmin where p+=supΩ′p and p−=infΩ′p and with τ satisfying (2.7). Let us also assume that there exists a positive constant Λ0 such that the following condition holds:
|B(x,s,η)|≤Λ0(1+|s|p(x)−1+|η|p(x)−1), | (3.32) |
for every (x,s,η)∈¯Ω′×R×RN. Let u∈W1,p(⋅)(Ω′) be such that
{divA(x,u,∇u)=B(x,u,∇u)inΩ′,|u|≤Mon∂Ω′, | (3.33) |
for some positive constant M. Then, there exists C such that |u|≤C in Ω′, where C depends only on M, |Ω′|, diam(Ω′), N,λ0,Λ0,L,p−,p+, δ, ‖g‖L1(Ω′), ||τ||L∞(Ω′), μ and c1.
Proof. Let v+ be the solution to (3.1) with boundary data M. Then, from the proof of Proposition 3.1 it follows that ||v+||W1,p(⋅)(Ω′) depends only on the constants in the structural conditions, on |Ω′|, diam(Ω′) and M. Since (recall Remark 2.1) we are under the assumptions of Theorem 4.1 in [11], then v+∈L∞(Ω′) with bounds depending only on the constants in the structural conditions, on |Ω′|, diam(Ω′) and M. Now, the comparison principle (Proposition 3.4) implies that u≤v+ in Ω′ and the upper bound follows. Proceeding in an analogous way with v− the solution to (3.1) with boundary data −M, we obtain the lower bound, thus concluding the proof.
As a corollary of Propositions 3.1 and 3.5 we get
Corollary 3.3. Let p,F,A and B as in Section 2 and let Ω′⊂Ω be a C1 domain. Assume, moreover that F satisfies (3.8) and (3.9) with τ satisfying (2.7) and A and B satisfy (3.14) and (3.32) in Ω′ for some p+−p−<δ<pmin where p+=supΩ′p and p−=infΩ′p.
Let u∈W1,p(⋅)(Ω′)∩L∞(Ω′). Then, there exists v∈u+W1,p(⋅)0(Ω′) a solution to
divA(x,v,∇v)=B(x,v,∇v)inΩ′. |
Moreover, v∈L∞(Ω′) and ‖v‖L∞(Ω′) is bounded by a constant C that depends only on ‖u‖L∞(Ω′), |Ω′|, diam(Ω′), N,λ0,Λ0,L,p−,p+, δ, ||τ||L∞(Ω′), μ and c1.
We also prove the following maximum principle
Proposition 3.6. Let p,A and B be as in Section 2. Assume moreover that condition (3.14) holds. We also assume that B(x,0,0)≡0 for every x∈¯Ω. Let u∈W1,p(⋅)(Ω) be such that
{divA(x,u,∇u)=B(x,u,∇u)inΩ,−M1≤u≤M2on∂Ω, | (3.34) |
for some nonnegative constants M1,M2. Then, −M1≤u≤M2 in Ω.
Proof. Since condition (3.14) implies that Bs(x,s,η)≥0 in ¯Ω×R×RN∖{0}, we have B(x,M2,0)≥0 and also B(x,−M1,0)≤0, for every x∈Ω. Recalling (2.1), we take v+≡M2 and v−≡−M1 and observe that divA(x,v+,∇v+)≤B(x,v+,∇v+) and divA(x,v−,∇v−)≥B(x,v−,∇v−) in Ω. Then, we can apply the comparison principle (Proposition 3.4) and obtain −M1≡v−≤u≤v+≡M2 in Ω and the conclusion follows.
As a corollary of Propositions 3.1 and 3.6 we get
Corollary 3.4. Let p,F,A and B as in Section 2 and let Ω′⊂Ω be a C1 domain. Assume, moreover that F satisfies (3.8) and (3.9) with τ satisfying (2.7) and A and B satisfy (3.14) in Ω′ for some p+−p−<δ<pmin where p+=supΩ′p and p−=infΩ′p. We also assume that B(x,0,0)≡0 for every x∈¯Ω′.
Let u∈W1,p(⋅)(Ω′)∩L∞(Ω′). Then, there exists v∈u+W1,p(⋅)0(Ω′) a solution to
divA(x,v,∇v)=B(x,v,∇v)inΩ′. |
Moreover, v∈L∞(Ω′) and ‖v‖L∞(Ω′)≤‖u‖L∞(Ω′).
We also have the following existence result of a bounded solution
Proposition 3.7. Let p as in Section 2. Assume that F(x,⋅,⋅) is locally Lipschitz in R×RN for almost every x∈Ω and that F(x,s,⋅)∈C1(RN)∩C2(RN∖{0}) for s∈R and almost every x∈Ω. Let A=∇ηF, B=Fs. Assume that A satisfies (2.2) and (2.5),
|A(x,s,η)|,|B(x,s,η)|≤Λ0(1+|s|τ(x)+|η|p(x))a.e.inΩ×R×RN, |
and F satisfies (3.3) and (3.4), where τ satisfies (2.7). Assume moreover that
F(x,s,η)=G(x,s,η)+f(x,s) withG, fmeasurable functions | (3.35) |
and,
G≥0inΩ×R×RN,G(x,s,η)=0 ⟺η=0, | (3.36) |
f(x,⋅)monotone decreasing in(−∞,0]and monotone increasing in[0,+∞). | (3.37) |
Then, for every Ω′⊂Ω of class C1 there holds that, if p+−p−<δ<pmin where p+=supΩ′p and p−=infΩ′p for δ in (3.3), given u∈W1,p(⋅)(Ω′) such that 0≤u≤M in Ω′ there exists v that minimizes the functional JΩ′(v) in u+W1,p(⋅)0(Ω′). Moreover, 0≤v≤M in Ω′.
In addition, if there exists ε0>0 such that for almost every x∈Ω, F(x,⋅,⋅)∈C1((−ε0,M+ε0)×RN), then there holds that v is a solution to
{divA(x,v,∇v)=B(x,v,∇v)inΩ′,v=uon∂Ω′. | (3.38) |
Proof. To begin with, the existence of a minimizer v follows proceeding as in Proposition 3.1. Let us prove that a minimizer satisfies 0≤v≤M. In fact, both w1=v−(v−M)+ and w2=v+v− are admissible functions. So that on the one hand,
0≤∫Ω′F(x,w1,∇w1)−F(x,v,∇v)=∫v>MF(x,M,0)−F(x,v,∇v)=∫v>Mf(x,M)−f(x,v)−∫v>MG(x,v,∇v)≤−∫v>MG(x,v,∇v)≤0. |
Hence, G(x,v,∇v)=0 in {v>M}. So that, ∇(v−M)+=0 in Ω′. As (v−M)+=0 on ∂Ω′, we deduce that v≤M in Ω′.
On the other hand, proceeding in a similar way with w2,
0≤∫Ω′F(x,w2,∇w2)−F(x,v,∇v)=∫v<0F(x,0,0)−F(x,v,∇v)=∫v<0f(x,0)−f(x,v)−∫v<0G(x,v,∇v)≤−∫v<0G(x,v,∇v)≤0, |
and we deduce as before that v−=0. This is, v≥0 in Ω′.
Now, in order to proceed with the proof we assume further regularity of F for −ε0≤s≤M+ε0. Let 0≤φ∈C∞0(Ω′) and 0<ε<ε0/‖φ‖L∞. Then, w=v+εφ is an admissible function, −ε0<w<M+ε0 and we deduce that
divA(x,v,∇v)≤B(x,v,∇v)in Ω′. |
Replacing φ by −φ we reverse the inequality. So that, v is a solution to (3.38).
In this section we prove properties of nonnegative local minimizers of the energy functional (1.1). We prove that nonnegative local minimizers are locally Hölder continuous (Theorem 4.3) and are solutions to
divA(x,u,∇u)=B(x,u,∇u)in {u>0}, |
where A(x,s,η)=∇ηF(x,s,η) and B(x,s,η)=Fs(x,s,η). In particular we prove our main result which is the local Lipschitz continuity on nonnegative local minimizers (Theorem 4.5).
We start with a definition, some related remarks and an existence result of a minimizer. We also prove nonnegativity and boundedness, under suitable assumptions.
Definition 4.1. Let p,F and λ be as in Section 2. Assume that F satisfies (3.3) and (3.4) with τ satisfying (2.7). We say that u∈W1,p(⋅)(Ω) is a local minimizer in Ω of
J(v)=JΩ(v)=∫Ω(F(x,v,∇v)+λ(x)χ{v>0})dx |
if for every Ω′⊂⊂Ω and for every v∈W1,p(⋅)(Ω) such that v=u in Ω∖Ω′ there holds that J(v)≥J(u).
We point out that the energy J is well defined in W1,p(⋅)(Ω) since, under the conditions in (2.7), the embedding theorem (see Theorem A.5) applies.
Remark 4.1. Let u be as in Definition 4.1. Let Ω′⊂⊂Ω and w−u∈W1,p(⋅)0(Ω′). If we define
ˉw={win Ω′,uin Ω∖Ω′, |
then ˉw∈W1,p(⋅)(Ω) and therefore J(ˉw)≥J(u). If we now let
JΩ′(v)=∫Ω′(F(x,v,∇v)+λ(x)χ{v>0})dx |
it follows that JΩ′(w)≥JΩ′(u).
Remark 4.2. Let J be as in Definition 4.1. If u∈W1,p(⋅)(Ω) is a minimizer of J among the functions v∈u+W1,p(⋅)0(Ω), then u is a local minimizer of J in Ω.
We start with an existence result of a minimizer to (1.1).
Theorem 4.1. Let p,F,A,B and λ be as in Section 2. Let ϕ∈W1,p(⋅)(Ω) and assume moreover that F satisfies (3.10) and (3.11) with τ satisfying (2.7).
Then, there exists a minimizer u∈ϕ+W1,p(⋅)0(Ω) to (1.1) and there holds that ‖u‖W1,p(⋅)(Ω)≤C, for a constant C depending only on ‖ϕ‖W1,p(⋅)(Ω)‖g‖L1(Ω), λmax, |Ω|, diam(Ω), N, pmin, pmax, q, L, μ, c1, ||τ||L∞(Ω) and the C1 norm of ∂Ω.
Proof. The proof is immediate from the computations in the proof of Proposition 3.2.
We also have,
Theorem 4.2. Let p and λ be as in Section 2. Let F,A and B be as in Propostion 3.7, except for the fact that we require that F satisfies (3.10) and (3.11) with τ satisfying (2.7), instead of (3.3) and (3.4), and with no oscillation assumption on p. Let ϕ∈W1,p(⋅)(Ω) such that 0≤ϕ≤M, for some M>0.
Then, there exists a minimizer u∈ϕ+W1,p(⋅)0(Ω) to (1.1) and 0≤u≤M in Ω.
Proof. Proceeding as in the proof of Proposition 3.2 we obtain that there exists a minimizer u∈ϕ+W1,p(⋅)0(Ω) to (1.1). The proof that 0≤u≤M is similar to that of Proposition 3.7. We only have to observe that
{u−(u−M)+>0}={u>0}and{u+u−>0}={u>0}. |
For local minimizers of (1.1) we first have
Lemma 4.1. Let p,F,A,B and λ be as in Section 2. Assume that F satisfies (3.3) and (3.4) with τ satisfying (2.7). Let u∈W1,p(⋅)(Ω) be a local minimizer of
J(v)=∫Ω(F(x,v,∇v)+λ(x)χ{v>0})dx. |
Then
divA(x,u,∇u)≥B(x,u,∇u)inΩ, | (4.1) |
where A(x,s,η)=∇ηF(x,s,η) and B(x,s,η)=Fs(x,s,η).
Proof. In fact, let t>0 and 0≤ξ∈C∞0(Ω). Using the minimality of u we have
0≤1t(J(u−tξ)−J(u))≤1t∫Ω(F(x,u−tξ,∇u−t∇ξ)−F(x,u,∇u))dx |
and if we take t→0, we obtain
0≤−∫Ω∇ηF(x,u,∇u)⋅∇ξdx−∫ΩFs(x,u,∇u)ξdx, | (4.2) |
which gives (4.1).
From now on we will deal with nonnegative, bounded, local minimizers of (1.1). Next we will prove that they are locally Lipschitz continuous.
We first prove that nonnegative, bounded, local minimizers are locally Hölder continuous.
Theorem 4.3. Let p,F,A,B and λ be as in Section 2. Assume that F satisfies (3.3) and (3.4) with τ satisfying (2.7). Let x0∈Ω, ˆr0>0 such that Bˆr0(x0)⊂⊂Ω. Assume that A,B satisfy condition (3.14) in Bˆr0(x0) and either B(x,0,0)≡0 for x∈Bˆr0(x0) or B satisfies (3.32) for x∈Bˆr0(x0). Let u∈W1,p(⋅)(Ω)∩L∞(Ω) be a nonnegative local minimizer of (1.1). Then, there exist 0<γ<1, γ=γ(N,pmin) and 0<ˆρ0<ˆr0, such that u∈Cγ(¯Bˆρ0(x0)). Moreover, ‖u‖Cγ(¯Bˆρ0(x0))≤C with ˆρ0 and C depending only on β,pmax,pmin,N,L,ˆr0,λ0,Λ0, ‖g‖L1(Bˆr0(x0)), μ, c1, λmax, ‖u‖L∞(Bˆr0(x0)), ||τ||L∞(Bˆr0(x0)) and δ.
Proof. We will prove that there exist 0<γ<1 and 0<ρ0<r0<ˆr0 such that, if Br0(y)⊂Bˆr0(x0) and ρ≤ρ0, then
(–∫–Bρ(y)|∇u|p−dx)1/p−≤Cργ−1, | (4.3) |
where p−=inf{p(x),x∈Br0(y)}. Without loss of generality we will assume that y=0.
In fact, let 0<r0≤min{ˆr02,1}, 0<r≤r0 and v the solution of
divA(x,v,∇v)=B(x,v,∇v)in Br,v−u∈W1,p(⋅)0(Br). | (4.4) |
Observe that, under our assumptions we can apply either Proposition 3.1 and Proposition 3.5 or Proposition 3.6 and deduce that such a solution exists and it is bounded in ¯Br if r0 is small enough depending on δ and L=‖∇p‖L∞(Ω). Hence, by Proposition 3.3, we have
∫Br(F(x,u,∇u)−F(x,v,∇v))dx≥12αλ0(∫Br∩{p(x)≥2}|∇u−∇v|p(x)dx+∫Br∩{p(x)<2}(|∇u|+|∇v|)p(x)−2|∇u−∇v|2dx), | (4.5) |
where α=α(pmin,pmax) and λ0 is as in (2.2).
By the minimality of u, we have (if A1=Br∩{p(x)<2} and A2=Br∩{p(x)≥2})
∫A2|∇u−∇v|p(x)dx≤CrN, | (4.6) |
∫A1|∇u−∇v|2(|∇u|+|∇v|)p(x)−2dx≤CrN, | (4.7) |
where C=C(pmin,pmax,N,λmax,λ0).
Let ε>0. Take ρ=r1+ε and suppose that rε≤1/2. Take 0<η<1 to be chosen later. Then, by Young's inequality, the definition of A1 and (4.7), we obtain
∫A1∩Bρ|∇u−∇v|p(x)dx≤Cη2/pmin∫A1∩Br(|∇u|+|∇v|)p(x)−2|∇u−∇v|2dx+Cη∫Bρ∩A1(|∇u|+|∇v|)p(x)dx≤Cη2/pminrN+Cη∫Bρ∩A1(|∇u|+|∇v|)p(x)dx. | (4.8) |
Therefore, by (4.6) and (4.8), we get
∫Bρ|∇u−∇v|p(x)dx≤Cη2/pminrN+Cη∫Bρ∩A1(|∇u|+|∇v|)p(x)dx, | (4.9) |
where C=C(pmin,pmax,N,λmax,λ0).
Since, |∇u|q≤C(|∇u−∇v|q+|∇v|)q), for any q>1, with C=C(q), we have, by (4.9), choosing η small, that
∫Bρ|∇u|p(x)dx≤CrN+C∫Bρ|∇v|p(x)dx, | (4.10) |
where C=C(pmin,pmax,N,λmax,λ0).
Now let M≥1 such that ||v||L∞(Br)≤M and define
w(x)=v(rx)MinB1. |
Observe that M depends only on ‖u‖L∞(Bˆr0(x0)) if B(x,0,0)≡0 or it depends also on the structural conditions on F, A and B, on ˆr0 and on the bound L of ‖∇p‖L∞ if not.
There holds that,
divˉA(x,w,∇w)=ˉB(x,w,∇w)inB1 |
where
ˉA(x,s,η)=A(rx,Ms,Mrη),ˉB(x,s,η)=rB(rx,Ms,Mrη). |
Now, let
˜A(x,s,η)=(rM)p−r−1ˉA(x,s,η),˜B(x,s,η)=(rM)p−r−1ˉB(x,s,η). |
Observe that w∈W1,ˉp(⋅)(B1)∩L∞(B1) satisfies
div˜A(x,w,∇w)=˜B(x,w,∇w)inB1, | (4.11) |
where ˉp(x)=p(rx).
Let us see that (4.11) is under the conditions of Theorem 1.1 in [10].
First, we clearly have ˜A(x,s,0)=0. Moreover, as 1≤rp−r−p+r≤CL<∞ if r≤1 and we have assumed that M≥1,
∑ij∂˜Ai∂ηj(x,s,η)ξiξj=(rM)p−r−1(Mr)∑ij∂Ai∂ηj(rx,Ms,Mrη)ξiξj≥λ0(rM)p−r−1(Mr)p(rx)−1|η|ˉp(x)−2|ξ|2≥λ0|η|ˉp(x)−2|ξ|2. | (4.12) |
On the other hand,
∑ij|∂˜Ai∂ηj(x,s,η)|≤Λ0(rM)p−r−1(Mr)p(rx)−1|η|ˉp(x)−2≤Λ0CLMpmax−pmin|η|ˉp(x)−2. | (4.13) |
Then, assuming without loss of generality that p(rx1)≥p(rx2),
|˜A(x1,s,η)−˜A(x2,s,η)|≤(rM)p−r−1|A(rx1,Ms,Mrη)−A(rx2,Ms,Mrη)|≤(rM)p−r−1Λ0((Mr)p(rx1)−1|η|p(rx1)−1+(Mr)p(rx2)−1|η|p(rx2)−1)(1+|log(Mr|η|)|)rβ|x1−x2|β≤Λ0CLMpmax−pmin(|η|ˉp(x1)−1+|η|ˉp(x2)−1)(1+|log|η||)|x1−x2|β | (4.14) |
if r≤rM,β.
Similarly,
|˜A(x,s1,η)−˜A(x,s2,η)|≤Λ4|s1−s2||η|ˉp(x)−1 |
with Λ4=Λ0CLMpmax−pmin+1.
On the other hand, denoting ˉτ(x)=τ(rx),
|˜B(x,s,η)|≤Λ0r(rM)p−r−1+Λ0CLMpmax−pmin+1|η|ˉp(x)+Λ0r(rM)p−r−1|Ms|ˉτ(x)≤Λ5(1+|η|ˉp(x)+|s|ˉτ(x)) |
with Λ5 depending on Λ0, L, pmax, pmin, M and ||τ||L∞(Bˆr0(x0)).
Since |w|≤1, we may assume that
˜B(x,s,η)≤Λ6(1+|η|ˉp(x)), |
with Λ6 depending on Λ0, L, pmin, pmax, M and ||τ||L∞(Bˆr0(x0)).
From Theorem 1.1 in [10], it follows that w∈C1,αloc(B1) for some 0<α<1 and that
supB1/2|∇w|≤C, |
with C depending only on β,pmax,pmin,N,L,λ0,Λ0, M and ||τ||L∞(Bˆr0(x0)), which implies
supBr/2|∇v|≤CMr. | (4.15) |
Therefore, from (4.10) and (4.15), we deduce that if r is small depending on M and β,
∫Bρ|∇u|p(x)dx≤CrN+CρNr−p+, | (4.16) |
with p+=sup{p(x),x∈Br0} and C depending on β,pmax,pmin,N,L,λ0,Λ0, λmax, M and ||τ||L∞(Bˆr0(x0)).
Then, if we take ε≤pminN, we have by (4.16) and by our election of ρ, that
–∫–Bρ|∇u|p−dx≤–∫–Bρ|∇u|p(x)dx+1|Bρ|∫Bρ∩{|∇u|<1}|∇u|p−dx≤–∫–Bρ|∇u|p(x)dx+1≤1+C(rρ)N+Cr−p+≤1+Cr−εN+Cr−p+≤Cr−p+=Cρ−p+(1+ε). |
Now let r0≤r0(ε,pmin,L) so that
p+p−=p+(Br0)p−(Br0)≤1+ε2, |
and small enough so that, in addition, rε0≤1/2. Then, if ρ≤ρ0=r1+ε0 and moreover, r0 is small depending on M and β,
–∫–Bρ|∇u|p−dx≤Cρ−(1+ε2)(1+ε)p−=Cρ−(1−γ)p−, |
where γ=ε2(1+ε)=γ(N,pmin). That is, if ρ≤ρ0=r1+ε0
(–∫–Bρ|∇u|p−dx)1/p−≤Cργ−1. |
Thus (4.3) holds, with C depending only on β,pmax,pmin,N,L,ˆr0,λ0,Λ0, ‖g‖L1(Bˆr0(x0)), μ, c1, λmax, ‖u‖L∞(Bˆr0(x0)), ||τ||L∞(Bˆr0(x0)) and δ.
Applying Morrey's Theorem, see e.g., [18], Theorem 1.53, we conclude that u∈Cγ(Bρ0(x0)) and ‖u‖Cγ(¯Bρ0/2(x0))≤C for C depending only on β,pmax,pmin,N,L,ˆr0,λ0,Λ0, ‖g‖L1(Bˆr0(x0)), μ, c1, λmax, ‖u‖L∞(Bˆr0(x0)), ||τ||L∞(Bˆr0(x0)) and δ.
As a corollary we obtain
Corollary 4.1. Let p,F,A,B and λ be as in Section 2. Assume that F satisfies (3.8) and (3.9) with τ satisfying (2.7). Assume that A,B satisfy condition (3.14) and either B(x,0,0)≡0 for x∈Ω or B satisfies (3.32) for x∈Ω. Let u∈W1,p(⋅)(Ω)∩L∞(Ω) be a nonnegative local minimizer of (1.1). Then, there exists 0<γ<1, γ=γ(N,pmin) such that u∈Cγ(Ω). Moreover, if Ω′⊂⊂Ω, then ‖u‖Cγ(¯Ω′)≤C with C depending only on dist(Ω′,∂Ω), β, N, pmin, pmax, L, λmax, λ0, Λ0, μ, c1, ‖u‖L∞(Ω), ||τ||L∞(Ω) and δ.
Then, under the assumptions of the previous corollary we have that u is continuous in Ω and therefore, {u>0} is open. We can now prove the following property for nonnegative local minimizers of (1.1).
Lemma 4.2. Let p,F,A,B and λ be as in Corollary 4.1. If u∈W1,p(⋅)(Ω)∩L∞(Ω) is a nonnegative local minimizer of
J(v)=∫Ω(F(x,v,∇v)+λ(x)χ{v>0})dx, |
there holds that,
divA(x,u,∇u)=B(x,u,∇u)in{u>0}, | (4.17) |
where A(x,s,η)=∇ηF(x,s,η) and B(x,s,η)=Fs(x,s,η).
Proof. From Lemma 4.1 we already know that (4.1) holds. In order to obtain the opposite inequality in {u>0}, we let 0≤ξ∈C∞0({u>0}) and consider u−tξ, for t<0, with |t| small.
Using the minimality of u we have
0≥1t(J(u−tξ)−J(u))=1t∫Ω(F(x,u−tξ,∇u−t∇ξ)−F(x,u,∇u))dx |
and if we take t→0, we obtain
0≥−∫Ω∇ηF(x,u,∇u)⋅∇ξdx−∫ΩFs(x,u,∇u)ξdx, |
which gives the desired inequality, so (4.17) follows.
We will next prove the Lipschitz continuity of nonnegative local minimizers of (1.1).
Before getting the Lipschitz continuity we prove the following result
Theorem 4.4. Let p,F,A,B, λ and u be as in Corollary 4.1. Let Ω′⊂⊂Ω. There exist constants C>0, r0>0 such that if x0∈Ω′∩∂{u>0} and r≤r0 then
supBr(x0)u≤Cr. |
The constants depend only on dist(Ω′,∂Ω), β, N, pmin, pmax, L, λmax, λ0, Λ0, μ, c1, ‖u‖L∞(Ω), ||τ||L∞(Ω) and δ.
Proof. Let us suppose by contradiction that there exist a sequence of nonnegative local minimizers uk corresponding to functionals Jk given by
Jk(v)=∫Ω(Fk(x,v,∇v)+λk(x)χ{v>0})dx, |
with uk∈W1,pk(⋅)(Ω)∩L∞(Ω), pmin≤pk(x)≤pmax, ‖∇pk‖L∞≤L, 0≤λk(x)≤λmax, ||uk||L∞(Ω)≤M, for some M≥1, and points ˉxk∈Ω′∩∂{uk>0}, such that
supBrk/4(ˉxk)uk≥krk and rk≤1k. |
We denote Ak(x,s,η)=∇ηFk(x,s,η) and Bk(x,s,η)=(Fk)s(x,s,η) and we also suppose that pk,Fk,Ak,Bk and τk satisfy the assumptions in Section 2 with constants λ0, Λ0 and β, we assume that Ak,Bk satisfy condition (3.14) and Fk satisfy (3.8) and (3.9) with τk satisfying (2.7) and either and Bk(x,0,0)≡0 for x∈Ω or Bk satisfy (3.32) for x∈Ω. All these conditions with exponent pk and constants independent of k and with ||τk||L∞(Ω)≤τ0, for some τ0>0.
Without loss of generality we will assume that ˉxk=0.
Let us define in B1, for k large, ˉuk(x)=1rkuk(rkx), ˉpk(x)=pk(rkx) and ˉλk(x)=λk(rkx). Then pmin≤ˉpk(x)≤pmax, ‖∇ˉpk‖L∞(B1)≤Lrk and 0≤ˉλk(x)≤λmax. Moreover, ˉuk is a nonnegative minimizer in ˉuk+W1,ˉpk(⋅)0(B1) of the functional
ˉJk(v)=∫B1(ˉFk(x,v,∇v)+ˉλk(x)χ{v>0})dx, | (4.18) |
where
ˉFk(x,s,η)=Fk(rkx,rks,η), |
with
ˉuk(0)=0 and max¯B1/4ˉuk(x)>k. |
Let dk(x)=dist(x,{ˉuk=0}) and Ok={x∈B1:dk(x)≤1−|x|3}. Since ˉuk(0)=0 then ¯B1/4⊂Ok, therefore
mk:=supOk(1−|x|)ˉuk(x)≥max¯B1/4(1−|x|)ˉuk(x)≥34max¯B1/4ˉuk(x)>34k. |
For each fixed k, ˉuk is bounded, then (1−|x|)ˉuk(x)→0 when |x|→1 which means that there exists xk∈Ok such that (1−|xk|)ˉuk(xk)=supOk(1−|x|)ˉuk(x), and then
ˉuk(xk)=mk1−|xk|≥mk>34k | (4.19) |
as xk∈Ok. Observe that δk:=dk(xk)≤1−|xk|3. Let yk∈∂{ˉuk>0}∩B1 such that |yk−xk|=δk. Then,
(1) B2δk(yk)⊂B1, since if y∈B2δk(yk)⇒|y|<3δk+|xk|≤1,(2) Bδk2(yk)⊂Ok, since if y∈Bδk2(yk)⇒|y|≤32δk+|xk|≤1−32δk⇒dk(y)≤δk2≤1−|y|3 and (3) if z∈Bδk2(yk)⇒1−|z|≥1−|xk|−|xk−z|≥1−|xk|−32δk≥1−|xk|2. |
By (2) we have
maxOk(1−|x|)ˉuk(x)≥max¯Bδk2(yk)(1−|x|)ˉuk(x)≥max¯Bδk2(yk)(1−|xk|)2ˉuk(x), |
where in the last inequality we are using (3). Then,
2ˉuk(xk)≥max¯Bδk2(yk)ˉuk(x). | (4.20) |
As Bδk(xk)⊂{ˉuk>0}, then Brkδk(rkxk)⊂{uk>0}. Hence, divAk(x,uk,∇uk)=Bk(x,uk,∇uk) in Brkδk(rkxk). Recalling that ||uk||L∞(Brkδk(rkxk))≤M, we can replace |s|τk(x) in (2.6) for Bk by Mτ0. Then we can apply Harnack's inequality (Theorem 3.2 in [23]) and we thus have
max¯B34rkδk(rkxk)uk(x)≤C[min¯B34rkδk(rkxk)uk(x)+rkδk], | (4.21) |
with C a positive constant depending only on N,pmin,pmax,L, M, λ0, Λ0 and τ0.
It follows that
max¯B34δk(xk)ˉuk(x)≤C[min¯B34δk(xk)ˉuk(x)+δk]. | (4.22) |
Recalling (4.19), we get from (4.22), for k large,
min¯B34δk(xk)ˉuk(x)≥cˉuk(xk), | (4.23) |
with c a positive constant depending only on N,pmin,pmax,LM, λ0, Λ0 and τ0. As ¯B34δk(xk)∩¯Bδk4(yk)≠∅ we have by (4.23)
max¯Bδk4(yk)ˉuk(x)≥cˉuk(xk). | (4.24) |
Let wk(x)=ˉuk(yk+δk2x)ˉuk(xk). Then, wk(0)=0 and, by (4.20) and (4.24), we have
max¯B1wk≤2max¯B1/2wk≥c>0. | (4.25) |
Now, recalling that ˉuk is a nonnegative minimizer in ˉuk+W1,ˉpk(⋅)0(B1) of the functional ˉJk in (4.18) and that Bδk2(yk)⊂B1, we see that wk is a nonnegative minimizer of ˆJk in wk+W1,ˉpk(yk+δk2x)0(B1), where
ˆJk(v)=∫B1(ˆFk(x,v,∇v)+ˆλk(x)χ{v>0})dx, |
ˆFk(x,s,η)=ˉFk(yk+δk2x,ˉuk(xk)s,2ˉuk(xk)δkη) and ˆλk(x)=ˉλk(yk+δk2x). |
We let ck=2ˉuk(xk)δk and we notice that ck→∞. So we define ˜pk(x)=ˉpk(yk+δk2x) and divide the functional ˆJk by c˜p−kk, with ˜p−k=infB1˜pk. Then, it follows that wk is a nonnegative minimizer of ˜Jk in wk+W1,˜pk(⋅)0(B1), where
˜Jk(v)=∫B1(˜Fk(x,v,∇v)+˜λk(x)χ{v>0})dx, |
˜Fk(x,s,η)=c−˜p−kkˆFk(x,s,η) and ˜λk(x)=c−˜p−kkˆλk(x). |
We claim that
˜λk→0uniformly in B1, | (4.26) |
c˜pk(x)−˜p−kk→1 uniformly, 1≤c˜pk(x)−˜p−kk≤M1in B1, | (4.27) |
˜pk→p0 uniformlyandpmin≤p0≤pmaxin B1, | (4.28) |
up to a subsequence, for some constants M1 and p0, where M1=M1(M,L).
On the one hand, 0<˜λk(x)≤λmaxc−1k→0 gives (4.26).
In addition, in B1 there holds, for k large, that 1≤c˜pk(x)−˜p−kk≤e2‖∇˜pk‖L∞logck. But we have ‖∇˜pk‖L∞logck≤Lrkδk2log(2Mrkδk)→0, which implies (4.27).
To see (4.28) we observe that pmin≤pk(x)≤pmax and ‖∇pk‖L∞(Ω)≤L and then, for a subsequence, pk→p uniformly on compacts of Ω, so ˜pk(x)=pk(rk(yk+δk2x))→p0=p(0) uniformly in B1.
We define ˜Ak=∇η˜Fk and ˜Bk=(˜Fk)s and we observe that
˜pk(x)=pk(rk(yk+δk2x)),˜τk(x)=τk(rk(yk+δk2x)), |
˜Fk(x,s,η)=c−˜p−kkˆFk(x,s,η)=c−˜p−kkˉFk(yk+δk2x,ˉuk(xk)s,2ˉuk(xk)δkη)=c−˜p−kkFk(rk(yk+δk2x),rkˉuk(xk)s,ckη), |
˜Ak(x,s,η)=c−˜p−kkckAk(rk(yk+δk2x),rkˉuk(xk)s,ckη), |
˜Bk(x,s,η)=c−˜p−kkrkˉuk(xk)Bk(rk(yk+δk2x),rkˉuk(xk)s,ckη). |
There holds that ˜pk, ˜Fk, ˜Ak, ˜Bk and ˜τk are under the assumptions of Section 2, with constants independent of k. In fact, recalling (4.27), we get for k large
pmin≤˜pk(x)≤pmax,‖∇˜pk‖L∞(Ω)≤L,˜pk(x)≤˜τk(x)≤τ0, |
˜Ak(x,s,0)=0, |
∑i,j∂(˜Ak)i∂ηj(x,s,η)ξiξj=c−˜p−kkc2k∑i,j∂(Ak)i∂ηj(rk(yk+δk2x),rkˉuk(xk)s,ckη)ξiξj≥λ0c˜pk(x)−˜p−kk|η|˜pk(x)−2|ξ|2≥λ0|η|˜pk(x)−2|ξ|2, | (4.29) |
∑i,j|∂(˜Ak)i∂ηj(x,s,η)|=c−˜p−kkc2k∑i,j|∂(Ak)i∂ηj(rk(yk+δk2x),rkˉuk(xk)s,ckη)|≤Λ0c˜pk(x)−˜p−kk|η|˜pk(x)−2≤Λ0M1|η|˜pk(x)−2. | (4.30) |
Assuming, without loss of generality, that ˜pk(x1)≥˜pk(x2) and using that (rkδk2)β1logck≤(rkδk2)β1log(2Mrkδk)→0, we get
|˜Ak(x1,s,η)−˜Ak(x2,s,η)|≤c−˜p−kkckΛ0(rkδk2)β1|x1−x2|β(|ckη|˜pk(x1)−1+|ckη|˜pk(x2)−1)(1+|log|ckη||)≤2M1Λ0|x1−x2|β(|η|˜pk(x1)−1+|η|˜pk(x2)−1)(1+|log|η||). | (4.31) |
Finally, recalling that rkˉuk(xk)≤M, we obtain
|˜Ak(x,s1,η)−˜Ak(x,s2,η)|≤c−˜p−kkckΛ0rkˉuk(xk)|s1−s2||ckη|˜pk(x)−1≤Λ0M1M|s1−s2||η|˜pk(x)−1, | (4.32) |
|˜Bk(x,s,η)|≤c−˜p−kkrkˉuk(xk)Λ0(1+|ckη|˜pk(x)+|rkˉuk(xk)s|˜τk(x))≤MΛ0(c−˜p−kk+M1|η|˜pk(x)+c−˜p−kk|Ms|˜τk(x))≤M1MΛ0(1+|η|˜pk(x)+Mτ0|s|˜τk(x)). | (4.33) |
On the other hand, ˜Ak and ˜Bk satisfy condition (3.14). In fact, since Ak and Bk satisfy condition (3.14),
12∑i,j∂(˜Ak)i∂ηj(x,s,η)ξiξj+(˜Bk)s(x,s,η)w2=12c−˜p−kkc2k∑i,j∂(Ak)i∂ηj(rk(yk+δk2x),rkˉuk(xk)s,ckη)ξiξj+c−˜p−kk(rkˉuk(xk))2(Bk)s(rk(yk+δk2x),rkˉuk(xk)s,ckη)w2≥c−˜p−kk2|(Ak)s(rk(yk+δk2x),rkˉuk(xk)s,ckη)⋅(ckξ)(rkˉuk(xk)w)|=2|(˜Ak)s(x,s,η)⋅ξw|. |
Also, since Fk satisfy (3.8) and (3.9) with τk satisfying (2.7), with exponent pk and constants independent of k, then ˜Fk satisfy (3.8) and (3.9) with ˜τk satisfying (2.7), with exponent ˜pk and constants independent of k. In fact,
˜Fk(x,s,η)≥c−˜p−kkμ|ckη|˜pk(x)−c−˜p−kkc1(|rkˉuk(xk)s|˜pk(x)−δ+1)≥μ|η|˜pk(x)−c1Mpmax(|s|˜pk(x)−δ+1). | (4.34) |
Analogously,
˜Fk(x,s,η)≤c−˜p−kkμ−1|ckη|˜pk(x)+c−˜p−kkc1(|rkˉuk(xk)s|˜τk(x)+1)≤M1μ−1|η|˜pk(x)+c1Mτ0(|s|˜τk(x)+1). | (4.35) |
If Bk(x,0,0)≡0 for x∈Ω, then ˜Bk(x,0,0)≡0 for x∈B1.
On the other hand, if Bk satisfy (3.32) for x∈Ω with exponent pk and constant independent of k, then ˜Bk satisfy (3.32) for x∈B1 with exponent ˜pk and constant independent of k. In fact,
|˜Bk(x,s,η)|=c−˜p−kkrkˉuk(xk)|Bk(rk(yk+δk2x),rkˉuk(xk)s,ckη)|≤c−˜p−kkrkˉuk(xk)Λ0(1+|rkˉuk(xk)s|˜pk(x)−1+|ckη|˜pk(x)−1)≤c−1kM1MpmaxΛ0(1+|s|˜pk(x)−1+|η|˜pk(x)−1)≤M1MpmaxΛ0(1+|s|˜pk(x)−1+|η|˜pk(x)−1). | (4.36) |
We now take vk the solution of
div˜Ak(x,vk,∇vk)=˜Bk(x,vk,∇vk)in B3/4,vk−wk∈W1,˜pk(⋅)0(B3/4). | (4.37) |
In fact, from Corollaries 3.3, 3.4 and 3.2 and the upper bound in (4.25), it follows that if k is large enough
||vk||L∞(B3/4)≤ˉC, | (4.38) |
where ˉC depends only on N, pmin, pmax, L, λ0, Λ0, μ, c1, δ, M and τ0. Here we have used that supB3/4˜pk−infB3/4˜pk≤‖∇˜pk‖L∞32≤3Lrkδk4<δ in (3.8), for k large.
Then, by (4.38), we can replace |s|ˉτk(x) in (4.33) by 1+ˉCτ0 and applying Theorem 1.1 in [10] we obtain that, for k large,
||vk||C1,α(¯B1/2)≤ˆC with 0<α<1 | (4.39) |
where ˆC depends only on β, N, pmin, pmax, L, λ0, Λ0, μ, c1, δ, M and τ0. Therefore, there is a function v0∈C1,α(¯B1/2) such that, for a subsequence,
vk→v0and∇vk→∇v0uniformly in ¯B1/2. | (4.40) |
Let us now show that
wk−vk→0in Lpmin(B3/4). | (4.41) |
From the minimality of wk we have
∫B3/4˜Fk(x,wk∇wk)−˜Fk(x,vk∇vk)≤C(N)‖˜λk‖L∞(B3/4), | (4.42) |
which together with Proposition 3.3 gives
∫Ak2|∇wk−∇vk|˜pk(x)dx≤C‖˜λk‖L∞(B3/4), | (4.43) |
∫Ak1|∇wk−∇vk|2(|∇wk|+|∇vk|)˜pk(x)−2dx≤C‖˜λk‖L∞(B3/4), | (4.44) |
where Ak1=B3/4∩{˜pk(x)<2}, Ak2=B3/4∩{˜pk(x)≥2} and C=C(pmin,pmax,N,λ0).
Applying Hölder's inequality (Theorem A.3) with exponents 2˜pk(x) and 22−˜pk(x), we get
∫Ak1|∇wk−∇vk|˜pk(x)dx≤2 ‖Gak‖L2/˜pk(⋅)(Ak1)‖Gbk‖L2/(2−˜pk(⋅))(Ak1), | (4.45) |
where
Gak=|∇wk−∇vk|˜pk(|∇wk|+|∇vk|)(˜pk−2)˜pk/2Gbk=(|∇wk|+|∇vk|)(2−˜pk)˜pk/2. |
Since
∫Ak1|Gak|2/˜pk(x)dx=∫Ak1|∇wk−∇vk|2(|∇wk|+|∇vk|)˜pk(x)−2dx, |
then, from (4.44), (4.26) and Proposition A.1, we get, for k large,
‖Gak‖L2/˜pk(⋅)(Ak1)≤C‖˜λk‖pmin/2L∞(B3/4), | (4.46) |
C=C(pmin,pmax,N,λ0). On the other hand, (4.37) and the bounds (4.34), (4.35) and (4.38) give
C1∫B3/4|∇vk|˜pk(x)≤∫B3/4˜Fk(x,vk∇vk)+C2≤∫B3/4˜Fk(x,wk∇wk)+C2≤C(1+∫B3/4|∇wk|˜pk(x)). |
This implies
∫Ak1|Gbk|2/(2−˜pk(x))dx≤C∫B3/4(|∇wk|˜pk(x)+|∇vk|˜pk(x))dx≤˜C(1+∫B3/4|∇wk|˜pk(x)), | (4.47) |
for some ˜C≥1, depending only on pmin, pmax and the uniform constants and functions in (4.34), (4.35) and (4.38). Now (4.47) and Proposition A.1 give
‖Gbk‖L2/(2−˜pk(⋅))(Ak1)≤˜C(1+∫B3/4|∇wk|˜pk(x)). | (4.48) |
Let us show that the right hand side in (4.48) can be bounded independently of k.
In fact, let ˜vk be the solution of
div˜Ak(x,˜vk,∇˜vk)=˜Bk(x,˜vk,∇˜vk)in B7/8,˜vk−wk∈W1,˜pk(⋅)0(B7/8). | (4.49) |
Then, similar arguments to those leading to (4.38) and (4.39), give, for k large enough,
||˜vk||L∞(B7/8)≤ˉC, | (4.50) |
and
||˜vk||C1,α(¯B3/4)≤ˆC with 0<α<1, | (4.51) |
where ˉC and ˆC depend only β, N, pmin, pmax, L, λ0, Λ0, μ, c1, δ, M and τ0.
Since wk is a nonnegative minimizer of ˜Jk in B1, then we can argue as in the proof of Theorem 4.3 and get estimate (4.10) for u=wk, v=˜vk, p(x)=˜pk(x), λ(x)=˜λk(x), r=7/8 and ρ=3/4. That is,
∫B3/4|∇wk|˜pk(x)dx≤C+C∫B3/4|∇˜vk|˜pk(x)dx, | (4.52) |
where C=C(pmin,pmax,N,λmax,λ0). Therefore (4.52) and (4.51) give, for k large, a uniform bound for the right hand side in (4.48). That is,
‖Gbk‖L2/(2−˜pk(⋅))(Ak1)≤ˉC, | (4.53) |
with ˉC a constant depending only on β, N, pmin, pmax, L, λ0, Λ0, μ, c1, δ, M and τ0.
Now, putting together (4.43), (4.45), (4.46), (4.53) and (4.26), we obtain
∫B3/4|∇wk−∇vk|˜pk(x)→0. | (4.54) |
Thus, using Poincare's inequality (Theorem A.4) and Theorem A.2, we get (4.41).
In order to conclude the proof, we now observe that, since ˜pk, ˜Fk, ˜Ak, ˜Bk, ˜τk, ˜λk and wk fall (uniformly) under the assumption of Corollary 4.1 in B1, there exists 0<γ<1, γ=γ(N,pmin), such that
‖wk‖Cγ(¯B1/2)≤C |
with C depending only on β, N, pmin, pmax, L, λmax, λ0, Λ0, μ, c1, τ0 and δ (recall that ‖wk‖L∞(B1)≤2).
Therefore, there is a function w0∈Cγ(¯B1/2) such that, for a subsequence,
wk→w0uniformly in ¯B1/2. | (4.55) |
In addition, recalling (4.40) and (4.41), we get v0=w0 in ¯B1/2.
We then observe that, since there holds that wk≥0, wk(0)=0 and (4.25), then (4.55) implies
w0≥0,w0(0)=0,max¯B1/2w0≥c>0. |
That is,
v0≥0,v0(0)=0,max¯B1/2v0≥c>0. | (4.56) |
Let us show that (4.56) gives a contradiction. We will divide the proof in two cases.
Case I. Assume that ˜Bk(x,0,0)≡0 for x∈B1.
We first observe that, since wk≥0, from Proposition 3.6 we deduce that vk≥0.
Recalling (4.39), we choose M0>0 such that, for every k,
||vk||L∞(B1/2)≤M0,||∇vk||L∞(B1/2)≤M0, |
and define
˜˜Ak(x,s,η)=a(s,η)˜Ak(x,s,η)+(1−a(s,η))|η|p0−2η,˜˜Bk(x,s,η)=a(s,η)˜Bk(x,s,η), |
where
a(s,η)=χ{|s|≤M0,|η|≤M0}. |
Then,
div˜˜Ak(x,vk,∇vk)=˜˜Bk(x,vk,∇vk)in B1/2. |
From (4.29) and (4.30) (recall Remark 2.1) we deduce
|˜Ak(x,s,η)|≤˜Λ0|η|˜pk(x)−1,˜Ak(x,s,η)⋅η≥˜Λ−10|η|˜pk(x), | (4.57) |
for some constant ˜Λ0>0 independent of k.
Let us now fix ε>0. Then, if k≥k0(ε), (4.57), (4.33) and (4.28) give, for large k,
|˜˜Ak(x,s,η)|≤˜˜Λ0|η|p0−1+cε,˜˜Ak(x,s,η)⋅η≥˜˜Λ−10|η|p0−cε,|˜˜Bk(x,s,η)|≤˜˜Λ0|η|p0−1+cε, |
for some positive constants {\tilde{\tilde{\Lambda}}}_0 and c (independent of {\varepsilon} and k ).
Applying Harnack's inequality (see [22], Theorems 5 and 6 and Section 5), we get for any 0 < r < 1
\begin{equation*} \max\limits_{\overline{B_{r/2}}}v_k\le C_r \big(\min\limits_{\overline{B_{r/2}}} v_k + {{\varepsilon}}^{\frac{1}{p_0}}\big), \end{equation*} |
with C_r a positive constant.
Now, letting k\to\infty first, and then {\varepsilon}\to 0 , we get
\begin{equation} \max\limits_{\overline{B_{r/2}}}v_0\le C_r \min\limits_{\overline{B_{r/2}}} v_0, \end{equation} | (4.58) |
with
\begin{equation} v_0\ge 0, \qquad v_0(0) = 0. \end{equation} | (4.59) |
Since 0 < r < 1 is arbitrary, we get v_0\equiv 0 in B_{1/2} . This is in contradiction with (4.56) and concludes the proof of Case I.
Case II. Assume that \tilde B_k satisfy (3.32) for x\in B_1 with exponent \tilde p_k and constant independent of k . Then, (4.30), (4.31), (4.32) and (4.36) imply that, for a subsequence,
\begin{equation*} \begin{aligned} \tilde A_k \rightarrow \tilde A & \quad\mbox{uniformly on compacts of }{B_{1/2}}\times{\mathbb R}\times{\mathbb R}^N\setminus\{0\} \, \mbox{and pointwise on }{B_{1/2}}\times{\mathbb R}\times{\mathbb R}^N, \\ &\qquad\qquad\tilde B_k\rightarrow 0\quad\mbox {uniformly on compacts of }{B_{1/2}}\times{\mathbb R}\times{\mathbb R}^N, \end{aligned} \end{equation*} |
and from (4.29) and (4.30) (recall Remark 2.1) we deduce
\begin{equation*} \begin{aligned} |\tilde A(x, s, \eta)|&\le \tilde\Lambda_0|\eta|^{p_0-1}, \\ \tilde A(x, s, \eta)\cdot\eta&\ge \tilde\Lambda_0^{-1}|\eta|^{p_0}, \end{aligned} \end{equation*} |
for some constant \tilde\Lambda_0 > 0 . Then, (4.37) and (4.40) imply that
\begin{equation*} {\rm div} \tilde{A}(x, v_0, \nabla v_0) = 0 \quad\mbox{in }B_{1/2}. \end{equation*} |
Applying Harnack's inequality (see [22], Theorems 5 and 6 and Section 5), we get again, that (4.58) and (4.59) holds for any 0 < r < 1 . This contradicts once more (4.56) and concludes the proof.
We can now prove the Lipschitz continuity of nonnegative local minimizers of (1.1)
Theorem 4.5. Let p, F, A, B, \lambda and u be as in Corollary 4.1. Then u is locally Lipschitz continuous in \Omega . Moreover, for any \Omega'\subset\subset \Omega the Lipschitz constant of u in \Omega' can be estimated by a constant C depending only on {\rm dist}(\Omega', \partial\Omega) , \beta , N , p_{\min} , p_{\max} , L , {\lambda_{\max}} , \lambda_0 , \Lambda_0 , \mu , c_1 , \|u\|_{L^{\infty}(\Omega)} , ||\tau||_{L^{\infty}(\Omega)} and \delta .
Proof. The result is a consequence of Corollary 4.1, Lemma 4.2 and Theorem 4.4 above, and Proposition 2.1 in [16]. We point out that, although the proof of Proposition 2.1 in [16] is written for the particular case in which A(x, s, \eta) = |\eta|^{p(x)-2}\eta and B(x, s, \eta) = f(x) , this same proof is valid for general A and B under the present assumptions, without changes.
In this section we present some examples of application of our results.
Theorem 5.1. Let f(x, s) be a measurable function such that f(x, \cdot)\in C^2({\mathbb R}) for every x\in\Omega . Let a(x, s) be a Hölder continuous function with exponent \alpha , a(x, \cdot)\in C^2({\mathbb R}) for every x\in\Omega . Let p , \tau and \lambda as in Section 2 and 0 < \delta < p_{\min} . Assume that there exist positive constants a_0, a_1, a_2, c_1 and \Lambda_0 such that
f1 -c_1(1+|s|^{p(x)-\delta})\le f(x, s)\le c_1(1+|s|^{\tau(x)}) in \Omega\times{\mathbb R} .
f2 f_s(x, 0)\equiv 0 in \Omega .
f3 f_{ss}(x, s)\ge0 in \Omega\times{\mathbb R} .
f4 |f_s(x, s)|\le \Lambda_0(1+|s|^{\tau(x)}) in \Omega\times{\mathbb R} .
And
a1 0 < a_0\le a(x, s)\le a_1 < \infty in \Omega\times{\mathbb R} .
a2 |a_s(x, s)|\le a_2 in \Omega\times{\mathbb R} .
a3 \big(a(x, s)^{1-\gamma(x)}\big)_{ss}\le0 in \Omega\times{\mathbb R} with \gamma(x) = \frac{2p(x)}{\min\{1, p(x)-1\}} > 1 .
Let
F(x, s, \eta) = a(x, s)\frac{|\eta|^{p(x)}}{p(x)}+f(x, s) |
and let u\in W^{1, p(\cdot)}(\Omega)\cap L^\infty(\Omega) a nonnegative, local minimizer of (1.1). Then, u is locally Lipschitz continuous in \Omega .
Proof. We only have to see that F, A, B satisfy the hypotheses of Theorem 4.5.
There holds that
A(x, s, \eta) = a(x, s)|\eta|^{p(x)-2}\eta, \qquad B(x, s, \eta) = a_s(x, s)\frac{|\eta|^{p(x)}}{p(x)}+f_s(x, s). |
And
\frac{a_0}{p_{\max}}|\eta|^{p(x)}-c_1(1+|s|^{p(x)-\delta})\le F(x, s, \eta)\le \frac{a_1}{p_{\min}}|\eta|^{p(x)}+c_1(1+|s|^{\tau(x)}). |
Moreover,
(1). \; A(x, s, 0) = 0 .
(2). \; \sum_{i, j}\frac{\partial A_i}{\partial\eta_j}(x, s, \eta)\xi_i\xi_j\ge\lambda_0|\eta|^{p(x)-2}|\xi|^2 . In fact,
\begin{equation} \begin{aligned}\sum\limits_{i, j}\frac{\partial A_i}{\partial\eta_j}(x, s, \eta)\xi_i\xi_j& = a(x, s)\Big[(p(x)-2)|\eta|^{p(x)-4}\langle\eta, \xi\rangle^2+|\eta|^{p(x)-2}|\xi|^2\Big]\\ &\ge a(x, s)\min\{1, p(x)-1\}|\eta|^{p(x)-2}|\xi|^2 \ge\lambda_0|\eta|^{p(x)-2}|\xi|^2 \end{aligned} \end{equation} | (5.1) |
with \lambda_0 = a_0\min\{1, p_{\min}-1\} .
(3). \; \sum_{i, j}\Big|\frac{\partial A_i}{\partial\eta_j}(x, s, \eta)\Big|\le \Lambda_0|\eta|^{p(x)-2} if \Lambda_0\ge a_1N(p_{\max}+3) .
(4). \; |A(x_1, s, \eta)-A(x_2, s, \eta)|\le \Lambda_0|x_1-x_2|^\alpha\big(|\eta|^{p(x_1)-1}+|\eta|^{p(x_2)-1}\big)\big|\big(1+\big|\log|\eta|\big|\big) for a big enough constant \Lambda_0 . In fact, without loss of generality we may assume that p(x_1)\ge p(x_2) . There holds,
|A(x_1, s, \eta)-A(x_2, s, \eta)|\le a(x_1, s)\big||\eta|^{p(x_1)-1}-|\eta|^{p(x_2)-1}\big|+\big|a(x_1, s)-a(x_2, s)\big||\eta|^{p(x_2)-1}. |
Now, if |\eta|\ge1 ,
\big||\eta|^{p(x_1)-1}-|\eta|^{p(x_2)-1}\big|\le {L} |x_1-x_2| |\eta|^{p(x_1)-1}\big|\log|\eta|\big|\le {L} |x_1-x_2| \Big(|\eta|^{p(x_1)-1}+|\eta|^{p(x_2)-1}\Big)\big|\log|\eta|\big| . |
A similar inequality holds if |\eta|\le1 . So that,
|A(x_1, s, \eta)-A(x_2, s, \eta)|\le a_1 {L} |x_1-x_2| \Big(|\eta|^{p(x_1)-1}+|\eta|^{p(x_2)-1}\Big)\big|\log|\eta|\big|+C_a|x_1-x_2|^\alpha|\eta|^{p(x_2)-1}, |
where C_a is the Holder constant of the function a . And the result follows if \Lambda_0\ge a_1{L} d(\Omega)^{1-\alpha}+C_a with d(\Omega) the diameter of \Omega .
(5). \; |A(x, s_1, \eta)-A(x, s_2, \eta)|\le a_2|\eta|^{p(x)-1}|s_1-s_2| .
We clearly have,
(1). |B(x, s, \eta)|\le \Lambda_0(1+|\eta|^{p(x)}+|s|^{\tau(x)}) (as we may assume, without loss of generality that \Lambda_0\ge \frac{a_2}{p_{\min}} ).
(2). \; B(x, 0, 0) = 0 .
Finally, let us see that
2|A_s(x, s, \eta)\cdot\xi\, w|\le \frac12\sum\limits_{i, j}\frac{\partial A_i}{\partial\eta_j}(x, s, \eta)\xi_i\xi_j+B_s(x, s, \eta)\, w^2. |
In fact, let
\ell(x) = \frac{p(x)-2}{2(p(x)-1)}\qquad {\varepsilon}(x, s) = a(x, s)\min\{1, p(x)-1\}. |
Then,
\begin{aligned} 2|A_s&(x, s, \eta)\cdot\xi\, w|\le\Big(\sqrt{{\varepsilon}(x, s)} |\eta|^{\ell(x)(p(x)-1)} |\xi|\Big)\Big(\frac2{\sqrt{{\varepsilon}(x, s)}}|a_s(x, s)||\eta|^{(1-\ell(x))(p(x)-1)}\, |w|\Big)\\ &\le\frac{{\varepsilon}(x, s)}2|\eta|^{p(x)-2}|\xi|^2+ \frac2{{\varepsilon}(x, s)}a_s(x, s)^2|\eta|^{p(x)}w^2\\ & = \frac12 a(x, s)\min\{1, p(x)-1\} |\eta|^{p(x)-2}|\xi|^2+ \frac{2a_s(x, s)^2}{a(x, s)\min\{1, p(x)-1\}}|\eta|^{p(x)} w^2. \end{aligned} |
By (5.1), we only have to check that
B_s(x, s, \eta)\ge \frac{2a_s(x, s)^2}{a(x, s)\min\{1, p(x)-1\}}|\eta|^{p(x)}. |
Since f_{ss}(x, s)\ge0 it is enough to check that
\begin{equation} {a_{ss}(x, s)}\ge \gamma(x)\frac {a_s(x, s)^2}{a(x, s)}\quad\mbox{ with }\quad\gamma(x) = \frac{2p(x)}{\min\{1, p(x)-1\}} \gt 1. \end{equation} | (5.2) |
And, (5.2) holds by hypothesis a3.
If a(x, s) is smooth in -M_1 < s < M_2 with M_1, M_2 > 0 , condition a3 only holds in 0\le s\le M < M_2 and the local minimizer u satisfies that 0\le u\le M , we can still apply the results in this paper and get that u is locally Lipschitz continuous.
Theorem 5.2. Let f(x, s) be a measurable function such that f(x, \cdot)\in C^2({\mathbb R}) for every x\in\Omega . Let a(x, s) be a Hölder continuous function with exponent \alpha , a(x, \cdot)\in C^2(-M_1, M_2)\cap Lip({\mathbb R}) for almost every x\in\Omega with M_1, M_2 > 0 . Let p , \tau and \lambda as in Section 2 and 0 < \delta < p_{\min} . Assume that there exist positive constants a_0, a_1, a_2, c_1, \Lambda_0 and 0 < M < M_2 such that
f1 -c_1(1+|s|^{p(x)-\delta})\le f(x, s)\le c_1(1+|s|^{\tau(x)}) in \Omega\times{\mathbb R} .
f2 f_s(x, 0)\equiv 0 in \Omega .
f3 f_{ss}(x, s)\ge0 in \Omega\times{\mathbb R} .
f4 |f_s(x, s)|\le \Lambda_0(1+|s|^{\tau(x)}) in \Omega\times{\mathbb R} .
And
a1 0 < a_0\le a(x, s)\le a_1 < \infty in \Omega\times{\mathbb R} .
a2 |a_s(x, s)|\le a_2 \ \ a.e. in \Omega\times{\mathbb R} .
a3' \big(a(x, s)^{1-\gamma(x)}\big)_{ss}\le0 in \Omega\times[0, M] with \gamma(x) = \frac{2p(x)}{\min\{1, p(x)-1\}} > 1 .
Let
F(x, s, \eta) = a(x, s)\frac{|\eta|^{p(x)}}{p(x)}+f(x, s) |
and let u\in W^{1, p(\cdot)}(\Omega)\cap L^\infty(\Omega) be a local minimizer of (1.1) such that 0\le u\le M . Then, u is locally Lipschitz continuous in \Omega .
Proof. By Proposition 3.7 for such a function f and with a satisfying a1 and a2, for every ball B_r(x_0)\subset\Omega with r small enough there exists a solution v\in u+W^{1, p(\cdot)}_0(B_r(x_0)) of (1.2) such that 0\le v\le \|u\|_{L^\infty(B_r(x_0))} . And this result also holds for all the rescaled equations and functions that appear in the proofs of Section 4. Hence, condition (3.14) is only needed for s\in(0, M) and this is a consequence of a3'.
Example 5.1. A possible example of functions a and f satisfying the assumptions of Theorem 5.2 is
a(x, s) = \begin{cases}(1+s)^{-q(x)}&\quad\mbox{if }-1/2\le s\le M_2, \\ 2^{q(x)}&\quad\mbox{if } s\le -1/2, \\ (1+M_2)^{-q(x)}&\quad\mbox{if } s\ge M_2, \end{cases} |
with M_2 > 0 and q\in L^\infty(\Omega) a Hölder continuous function such that 0 < q(x) < \frac1{\gamma(x)-1} and
f(x, s) = b(x)|s|^{\tau(x)} |
with 0\le b\in L^\infty(\Omega) and \tau(x)\ge 2 in \Omega satisfying (2.7).
Another possible choice of f is
\begin{equation} f(x, s) = b(x)\widetilde f(x, s) \end{equation} | (5.3) |
with 0\le b\in L^\infty(\Omega) and
\begin{equation*} \widetilde f(x, s) = \begin{cases} s^2&\quad\mbox{if } |s|\le 1, \\ \widetilde a(x)|s|^{\tau(x)}+\widetilde b(x)|s|+\widetilde c(x)&\quad\mbox{if } |s|\ge 1, \end{cases} \end{equation*} |
where \tau(x) satisfies (2.7) and the functions \widetilde a, \widetilde b, \widetilde c \in L^\infty(\Omega) are chosen in such a way that \widetilde f(x, \cdot)\in C^2({\mathbb R}) for every x\in\Omega .
With this choice of a and f , for every 0 < M < M_2 there holds that any local minimizer u such that 0\le u\le M is locally Lipschitz continuous in \Omega .
Observe that, by Theorem 4.2, if \phi\in W^{1, p(\cdot)}(\Omega) is such that 0\le\phi\le M < M_2 , such a minimizers always exists.
We have another example.
Theorem 5.3. Let f(x, s) be a measurable function such that f(x, \cdot)\in C^2({\mathbb R}) for every x\in\Omega . Let G(x, \eta) be a measurable function such that G(x, \cdot)\in C^1({\mathbb R}^N)\cap C^2({\mathbb R}^N\setminus\{0\}) for every x\in\Omega . Let p and \lambda as in Section 2 and assume that either f satisfies conditions \rm f1, \cdots, f4 in Theorem 5.1 or f satisfies \rm f1, f3 in Theorem 5.1 and
f4' |f_s(x, s)|\le \Lambda_0\big(1+|s|^{p(x)-1}\big) .
On the other hand, G satisfies
G1 \mu\big(|\eta|^{p(x)}- 1\big)\le G(x, \eta)\le \mu^{-1}\big(|\eta|^{p(x)}+ 1\big) with \mu > 0 .
G2 \nabla_\eta G(x, 0)\equiv0 in \Omega .
G3 \sum_{i, j}\frac{\partial^2 G}{\partial\eta_i\partial\eta_j}\xi_i\xi_j\ge \lambda_0 |\eta|^{p(x)-2}|\xi|^2 .
G4 \sum_{i, j}\Big|\frac{\partial^2 G}{\partial\eta_i\partial\eta_j}\Big|\le \Lambda_0 |\eta|^{p(x)-2} .
G5 |\nabla_\eta G(x_1, \eta)-\nabla_\eta G(x_2, \eta)|\le \Lambda_0 |x_1-x_2|^\beta \big(|\eta|^{p(x_1)-1}+|\eta|^{p(x_2)-1}\big)\big(1+\big|\log|\eta|\big|\big) for some 0 < \beta\le1 .
Let
F(x, s, \eta) = G(x, \eta)+f(x, s) |
and let u\in W^{1, p(\cdot)}(\Omega)\cap L^\infty(\Omega) be a nonnegative, local minimizer of (1.1). Then, u is locally Lipschitz continuous in \Omega .
Proof. There holds that
A(x, s, \eta) = \nabla_\eta G(x, \eta), \qquad B(x, s, \eta) = f_s(x, s). |
And it is clear that F, A and B satisfy the assumptions in Theorem 4.5.
Example 5.2. A possible example of function G satisfying the assumptions of Theorem 5.3 is
G(x, \eta) = a(x)\widetilde G\big(|\eta|^{p(x)}\big), |
with p(x) as in Section 2, a(x) a Hölder continuous function such that a_0\le a(x)\le a_1 , with a_0, a_1 positive constants and \widetilde G\in C^2\big([0, \infty)\big) a function satisfying:
\begin{equation*} \begin{aligned} &\qquad\qquad c_0\le \widetilde G'(t)\le C_0, \\ &0\le \widetilde G''(t)\le \frac{C_0}{1+t}\qquad c_0, C_0 \mbox{ positive constants.} \end{aligned} \end{equation*} |
In fact, since c_0\le\widetilde G'(t)\le C_0 , condition G1 in Theorem 5.3 holds. We have \nabla_\eta G(x, \eta) = a(x)\widetilde G'\big(|\eta|^{p(x)}\big)p(x)|\eta|^{p(x)-2}\eta , so we get condition G2. We obtain condition G3 by reasoning as in (5.1), using that in the present case we have \widetilde G''(t)\ge 0 and \widetilde G'(t)\ge c_0 .
We get condition G4 by using in our computations that \widetilde G'(t)\le C_0 and \widetilde G''(t)t\le C_0 .
Finally, applying again that \widetilde G''(t)t\le C_0 , we can obtain the estimate
|\widetilde G'\big(|\eta|^{p(x_1)}\big)-\widetilde G'\big(|\eta|^{p(x_2)}\big)|\le C_0|p(x_1)-p(x_2)||\log |\eta||, |
which combined with computations similar as those in (4) in Theorem 5.1 leads to condition G5.
A possible example of function f satisfying the assumptions of Theorem 5.3 is
f(x, s) = g(x)s, \qquad \mbox{with}\quad g\in L^{\infty}(\Omega). |
In fact, it is immediate that f satisfies conditions f1, f3 and f4'.
On the other hand, f(x, s) = b(x)|s|^{\tau(x)} with b and \tau as in Example 5.1 and f(x, s) as in (5.3) are other possible choices.
Let us present another example
Example 5.3. Another possible example of function G satisfying the assumptions of Theorem 5.3 is
G(x, \eta) = \widetilde A(x)\eta\cdot\eta|\eta|^{p(x)-2}, |
with p(x) as in Section 2 and \widetilde A(x)\in {{\mathbb R}}^{N\times N} , symmetric, Hölder continuous in \Omega and such that
\lambda(x) I\le \widetilde A(x)\le \Lambda(x) I. |
Here \lambda_0\le \lambda(x)\le \Lambda (x)\le \Lambda_0 with \lambda_0, \Lambda_0 positive constants and \Lambda(x)-\lambda(x)\le c_0 , with c_0 a suitable positive constant depending only on N , p_{\min} , p_{\max} and \lambda_0 .
In fact, conditions G1 and G2 in Theorem 5.3 are easy to verify. The computations leading to G4 and G5 are similar to the computations in Theorem 5.1.
In order to verify G3, we observe that, denoting a(x) the smaller eigenvalue of \widetilde A(x) , there holds that
\widetilde A(x) = a(x)I +\widetilde B(x), \qquad \text{with }\quad ||\widetilde B(x)||_{L^\infty(\Omega)}\le ||\Lambda(x)-\lambda(x)||_{L^\infty(\Omega)}. |
Then we can write
\begin{equation*} \begin{aligned} G(x, \eta) = &a(x)|\eta|^{p(x)} + \widetilde B(x)\eta\cdot\eta|\eta|^{p(x)-2}\\ = &G_1(x, \eta)+G_2(x, \eta). \end{aligned} \end{equation*} |
Now, proceeding as in Theorem 5.1, we get
\begin{equation} \sum\limits_{i, j}\frac{\partial^2 G_1}{\partial\eta_i\partial\eta_j}\xi_i\xi_j\ge c_{p_{\min}}\lambda_0 |\eta|^{p(x)-2}|\xi|^2. \end{equation} | (5.4) |
It is not hard to see that
\begin{equation} \sum\limits_{i, j}\Big|\frac{\partial^2 G_2}{\partial\eta_i\partial\eta_j}\Big|\le C ||\Lambda(x)-\lambda(x)||_{L^\infty(\Omega)} |\eta|^{p(x)-2}, \end{equation} | (5.5) |
with C depending only on N , p_{\min} and p_{\max} . Then, combining (5.4) and (5.5) we deduce that G(x, \eta) satisfies condition G3, if we take ||\Lambda(x)-\lambda(x)||_{L^\infty(\Omega)}\le c_0 , with c_0 depending only on \lambda_0 , N , p_{\min} and p_{\max} .
For choices of suitable functions f(x, s) for this G(x, \eta) we refer to Example 5.2.
Remark 5.1. We can present further examples of functions satisfying our assumptions. Let p and \lambda be as in Section 2. Let F_1 and F_2 satisfy the assumptions on Theorem 4.5, with B_i = \partial_s F_i satisfying B_i(x, 0, 0)\equiv 0 for x\in\Omega , i = 1, 2 . Then Theorem 4.5 also applies to the function
F(x, s, \eta) = a_1(x)F_1(x, s, \eta)+a_2(x)F_2(x, s, \eta), |
for any choice of Hölder continuous functions a_1(x), a_2(x) , which are bounded from above and below by positive constants.
The same result holds if F_1 and F_2 satisfy the assumptions on Theorem 4.5, with B_i = \partial_s F_i satisfying (3.32) for x\in \Omega , i = 1, 2 .
Similar consideration applies to functions F_1 and F_2 under the assumptions of Theorem 5.2.
Supported by the Argentine Council of Research CONICET under the project PIP 11220150100032CO 2016-2019, UBACYT 20020150100154BA and ANPCyT PICT 2016-1022.
The authors declare no conflict of interest.
In Section 1 we included some preliminaries on Lebesgue and Sobolev spaces with variable exponent. For the sake of completeness we collect here some additional results on these spaces.
Proposition A.1. There holds
\begin{align*} \min\Big\{\Big(\int_{\Omega} |u|^{p(x)}\, dx\Big) ^{1/{p_{\min}}}, & \Big(\int_{\Omega} |u|^{p(x)}\, dx\Big) ^{1/{p_{\max}}}\Big\}\le\|u\|_{L^{p(\cdot)}(\Omega)}\\ &\leq \max\Big\{\Big(\int_{\Omega} |u|^{p(x)}\, dx\Big) ^{1/{p_{\min}}}, \Big(\int_{\Omega} |u|^{p(x)}\, dx\Big) ^{1/{p_{\max}}}\Big\}. \end{align*} |
Some important results for these spaces are
Theorem A.1. Let p'(x) such that
\frac{1}{p(x)}+\frac{1}{p'(x)} = 1. |
Then L^{p'(\cdot)}(\Omega) is the dual of L^{p(\cdot)}(\Omega) . Moreover, if p_{\min} > 1 , L^{p(\cdot)}(\Omega) and W^{1, p(\cdot)}(\Omega) are reflexive.
Theorem A.2. Let q(x)\leq p(x) . If \Omega has finite measure, then L^{p(\cdot)}(\Omega)\hookrightarrow L^{q(\cdot)}(\Omega) continuously.
We also have the following Hölder's inequality
Theorem A.3. Let p'(x) be as in Theorem A.1. Then there holds
\int_{\Omega}|f||g|\, dx \le 2\|f\|_{p(\cdot)}\|g\|_{p'(\cdot)}, |
for all f\in L^{p(\cdot)}(\Omega) and g\in L^{p'(\cdot)}(\Omega) .
The following version of Poincare's inequality holds
Theorem A.4. Let \Omega be bounded. Assume that p(x) is log-Hölder continuous in \Omega (that is, p has a modulus of continuity \omega(r) = C(\log \frac{1}{r})^{-1} ). For every u\in W_0^{1, p(\cdot)}(\Omega) , the inequality
\|u\|_{L^{p(\cdot)}(\Omega)}\leq C\|\nabla u\|_{L^{p(\cdot)}(\Omega)} |
holds with a constant C depending only on N, \rm{diam}(\Omega) and the log-Hölder modulus of continuity of p(x) .
The following Sobolev embedding holds. We assume for simplicity that the domain is C^1 , but the result holds with weaker assumptions on the smoothness of the boundary.
Theorem A.5. Let \Omega be a C^1 bounded domain. Assume that p(x) is log-Hölder continuous in \Omega and 1 < p_{\min}\le p(x)\le p_{\max} < \infty . Let \tau be such that \tau(x)\ge p(x) and \tau\in C(\overline\Omega) . Assume moreover that \tau(x)\le p^*(x) = \frac{Np(x)}{N-p(x)} if p_{\max} < N , \tau(x) is arbitrary if p_{\min} > N , \tau(x) = p(x) if p_{\min}\le N \le p_{\max} .
Then, W^{1, p(\cdot)}(\Omega)\hookrightarrow L^{\tau(\cdot)}(\Omega) continuously. The embedding constant depends only on N , |\Omega| , the log-Hölder modulus of continuity of p(x) , p_{\min} , p_{\max} , ||\tau||_{L^{\infty}} and the C^1 norm of \partial\Omega .
For the proof of these results and more about these spaces, see [9,13,14,20] and the references therein.
[1] | Aboulaich R, Meskine D, Souissi A (2008) New diffusion models in image processing. Comput Math Appl 56: 874-882. |
[2] | Alt HW, Caffarelli LA (1981) Existence and regularity for a minimum problem with free boundary. J Reine Angew Math 325: 105-144. |
[3] | Alt HW, Caffarelli LA, Friedman A (1984) A free boundary problem for quasilinear elliptic equations. Ann Scuola Norm Sci 11: 1-44. |
[4] | Antontsev SN, Rodrigues JF (2006) On stationary thermo-rheological viscous flows. Ann Univ Ferrara 52: 19-36. |
[5] | Antontsev SN, Shmarev SI (2005) A model porous medium equation with variable exponent of nonlinearity: Existence, uniqueness and localization properties of solutions. Nonlinear Anal 60: 515-545. |
[6] | Bonder JF, Martínez S, Wolanski N (2010) A free boundary problem for the p(x)-Laplacian. Nonlinear Anal 72: 1078-1103. |
[7] | Chen Y, Levine S, Rao M (2006) Variable exponent, linear growth functionals in image restoration. SIAM J Appl Math 66: 1383-1406. |
[8] | Danielli D, Petrosyan A (2005) A minimum problem with free boundary for a degenerate quasilinear operator. Calc Var 23: 97-124. |
[9] | Diening L, Harjulehto P, Hasto P, et al. (2011) Lebesgue and Sobolev Spaces with Variable Exponents, Berlin: Springer. |
[10] | Fan X (2007) Global C1, α regularity for variable exponent elliptic equations in divergence form. J Differ Equations 235: 397-417. |
[11] | Fan X, Zhao D (1999) A class of De Giorgi type and Hölder continuity. Nonlinear Anal 36: 295-318. |
[12] | Gustafsson B, Shahgholian H (1996) Existence and geometric properties of solutions of a free boundary problem in potential theory. J Reine Angew Math 473: 137-179. |
[13] | Harjulehto P, Hästö P (2019) Orlicz Spaces and Generalized Orlicz Spaces, Berlin: Springer. |
[14] | Kováčik O, Rákosník J (1991) On spaces Lp(x) and Wk, p(x). Czechoslovak Math J 41: 592-618. |
[15] | Lederman C (1996) A free boundary problem with a volume penalization. Ann Scuola Norm Sci 23: 249-300. |
[16] | Lederman C, Wolanski N (2017) Weak solutions and regularity of the interface in an inhomogeneous free boundary problem for the p(x)-Laplacian. Interface Free Bound 19: 201-241. |
[17] | Lederman C, Wolanski N (2019) Inhomogeneous minimization problems for the p(x)-Laplacian. J Math Anal Appl 475: 423-463. |
[18] | Maly J, Ziemer WP (1997) Fine Regularity of Solutions of Elliptic Partial Differential Equations, Providence, RI: American Mathematical Society. |
[19] | Martínez S, Wolanski N (2008) A minimum problem with free boundary in Orlicz spaces. Adv Math 218: 1914-1971. |
[20] | Radulescu VD, Repovs DD (2015) Partial Differential Equations with Variable Exponents: Variational Methods and Qualitative Analysis, Boca Raton, FL: Chapman & Hall / CRC Press. |
[21] | Ruzicka M (2000) Electrorheological Fluids: Modeling and Mathematical Theory, Berlin: Springer-Verlag. |
[22] | Serrin J (1964) Local behavior of solutions of quasi-linear equations. Acta Math 111: 247-302. |
[23] | Wolanski N (2015) Local bounds, Harnack inequality and Hölder continuity for divergence type elliptic equations with non-standard growth. Rev Un Mat Argentina 56: 73-105. |
[24] | Zhikov VV (1987) Averaging of functionals of the calculus of variations and elasticity theory. Math USSR Izv 29: 33-66. |
[25] | Zhikov VV (2008) Solvability of the three-dimensional thermistor problem. Proc Steklov Inst Math 261: 98-111. |
1. | João Vitor da Silva, Giane Casari Rampasso, Gleydson Chaves Ricarte, Hernán Agustín Vivas, Free Boundary Regularity for a Class of One-Phase Problems with Non-Homogeneous Degeneracy, 2022, 0021-2172, 10.1007/s11856-022-2392-5 |