Loading [MathJax]/jax/element/mml/optable/SuppMathOperators.js
Research article Special Issues

Equivalence of solutions for non-homogeneous p(x)-Laplace equations

  • We establish the equivalence between weak and viscosity solutions for non-homogeneous p(x)-Laplace equations with a right-hand side term depending on the spatial variable, the unknown, and its gradient. We employ inf- and sup-convolution techniques to state that viscosity solutions are also weak solutions, and comparison principles to prove the converse. The new aspects of the p(x)-Laplacian compared to the constant case are the presence of log-terms and the lack of the invariance under translations.

    Citation: María Medina, Pablo Ochoa. Equivalence of solutions for non-homogeneous p(x)-Laplace equations[J]. Mathematics in Engineering, 2023, 5(2): 1-19. doi: 10.3934/mine.2023044

    Related Papers:

    [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] 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
    [3] 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
    [4] 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
    [5] 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
    [6] 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
    [7] Prashanta Garain, Kaj Nyström . On regularity and existence of weak solutions to nonlinear Kolmogorov-Fokker-Planck type equations with rough coefficients. Mathematics in Engineering, 2023, 5(2): 1-37. doi: 10.3934/mine.2023043
    [8] Juan-Carlos Felipe-Navarro, Tomás Sanz-Perela . Semilinear integro-differential equations, Ⅱ: one-dimensional and saddle-shaped solutions to the Allen-Cahn equation. Mathematics in Engineering, 2021, 3(5): 1-36. doi: 10.3934/mine.2021037
    [9] 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
    [10] Feida Jiang . Weak solutions of generated Jacobian equations. Mathematics in Engineering, 2023, 5(3): 1-20. doi: 10.3934/mine.2023064
  • We establish the equivalence between weak and viscosity solutions for non-homogeneous p(x)-Laplace equations with a right-hand side term depending on the spatial variable, the unknown, and its gradient. We employ inf- and sup-convolution techniques to state that viscosity solutions are also weak solutions, and comparison principles to prove the converse. The new aspects of the p(x)-Laplacian compared to the constant case are the presence of log-terms and the lack of the invariance under translations.



    To the memory of Ireneo Peral, mentor and friend. Te echamos de menos.

    Let Ω be a bounded domain of Rn. We will consider non-homogeneous p(x)-Laplace equations of the form

    Δp(x)u=f(x,u,Du)in Ω, (1.1)

    where, given a function p:Ω(1,), Δp(x) is the p(x)-Laplace operator defined as

    Δp(x)u:=div(|Du|p(x)2Du). (1.2)

    For a smooth function φ with Dφ0, we can expand the expression above and write

    Δp(x)φ(x)=|Dφ|p(x)2(Δφ+(p(x)2)|Dφ|2Δφ)|Dφ|p(x)2Dp(x)Dφlog|Dφ|,

    where

    Δφ:=D2φDφDφ

    is the -Laplacian. Consequently, two fundamental differences exist between Δp(x) and the p-Laplacian, with p constant: the fact that this operator is not invariant under translations in x and the presence of log-terms.

    Recently, the study of partial differential equations with variable exponents has been motivated by the description of models in electrorheological and thermorheological fluids, image processing [3], or robotics. Moreover, classical references for existence and regularity of solution for p(x)-Laplacian Dirichlet problems are [7,8,11], among others.

    In this work we are interested in analyzing the equivalence between weak and viscosity solutions (see Section 2.2 for the precise definitions of these notions) of the problem (1.1) under certain conditions on f. The relation among different types of solutions for different operators has been studied by several authors in the last decades. For linear problems, the equivalence between distributional and viscosity solutions was obtained by Ishii in [13]. Later on, for the homogeneous p-Laplace operator (i.e., (1.1) with f0), the equivalence between weak and viscosity solutions was first obtained in [16], and later on in [14] with a different proof. For a source term like the one in (1.1), depending on all the lower-order terms, this equivalence for the p-Laplace equation was given in [19], following some ideas from [14]. Similar studies have been recently made for non-local operators, see [1,18].

    In the case of the variable p(x)-Laplacian, the equivalence for the homogeneous equation was proved in [16]. Related results appear in [20] between solutions of homogeneous equations involving the strong and the normalized p(x)-Laplacian. Up to our knowledge, no results are available in the case f0. Indeed, combining techniques from [16,20] to deal with the operator, and from [19] to deal with the function f, the goal of this work is to prove the equivalence of weak and viscosity solutions for the general problem (1.1).

    Indeed, let us assume from now on that the exponent p satisfies

    pC1(¯Ω),1<pp+<,wherep:=minx¯Ωp(x),p+:=maxx¯Ωp(x). (1.3)

    Hence, our first result is the following:

    Theorem 1.1. Let p satisfy (1.3). Assume that f=f(x,t,η) is uniformly continuous in Ω×R×Rn, non increasing in t, Lipschitz continuous in η, and satisfies the growth condition

    |f(x,t,η)|γ(|t|)|η|p(x)1+ϕ(x), (1.4)

    where γ0 is continuous and ϕLloc(Ω). Thus, if u is a locally Lipschitz viscosity supersolution of (1.1) then it is a weak supersolution of the problem.

    It is worth to point out that the regularity assumption on u derives from a technical restriction on p (see Remark 3.2).

    The proof of Theorem 1.1 relies on the approximation by the so called inf-convolutions (see Section 2.3). Roughly speaking, we will regularize u by some functions uε, that will satisfy a related problem in weak sense, and we will pass to the limit here. This idea was first used in [14] for the constant p-Laplacian in the homogeneous case, and then in more general settings in [1,19,20].

    The reverse statement, weak solutions being viscosity, is strongly connected with comparison arguments, and a new class of functions needs to be considered.

    Definition 1.2. Let u be a weak supersolution to (1.1) in DΩ. We say that (u,f) satisfies the comparison principle property (CPP) in D if for every weak subsolution v of (1.1) such that uv a.e. in D we have uv a.e. in D.

    In particular we will see that, for those functions satisfying this property, weak solutions are indeed viscosity solutions.

    Theorem 1.3. Let p satisfy (1.3). Assume u is a continuous weak supersolution of (1.1) and f=f(x,t,η) is continuous in Ω×R×Rn and Lipschitz continuous in η. If (CPP) holds then u is a viscosity supersolution of (1.1).

    In Section 4, apart from this theorem, we prove a comparison principle for the general equation (1.1), which has interest in itself.

    Applications of the equivalence between viscosity and weak solutions can be found in [15,20] to removability of sets and Radó type theorems. Also, the equivalence has been recently used in free-boundary problems (see [2,12]).

    The paper is organized as follows: in Section 2 we give an introduction into the theory of Sobolev spaces with variable exponents, we introduce the notions of viscosity and weak solutions in this context, and the definition and main properties of the infimal convolutions. Section 3 is devoted to the proof of Theorem 1.1, that is, to see that viscosity solutions of (1.1) are also weak solutions. In Section 4 we prove Theorem 1.3 (that weak solutions are viscosity solutions) and a general comparison principle for Eq (1.1).

    In this section we introduce basic definitions and preliminary results concerning the spaces of variable exponent and the related theory of differential equations. Let

    C+(¯Ω):={pC(¯Ω):p(x)>1for anyx¯Ω}

    and

    p:=min¯Ωp(),p+:=max¯Ωp().

    Assume that p belongs to C+(¯Ω) and satisfies the following log-Hölder condition: there exists C>0 so that

    |p(x)p(y)|C1|log|xy||,for allx,yΩ,xy. (2.1)

    We define the Lebesgue variable exponent space as

    Lp()(Ω):={u:ΩR:u is measurable and Ω|u(x)|p(x)dx<},

    and we denote by Lp()(Ω) the conjugate space of Lp()(Ω), where

    1p()+1p()=1.

    Consider also the Luxemburg norm

    uLp():=inf{λ>0:Ω|u(x)λ|p(x)dx1}.

    Then the following results, that can be found in [5], hold.

    Theorem 2.1 (Hölder's inequality). The space (Lp()(Ω),Lp()(Ω)) is a separable, uniform convex Banachspace. Furthermore, if uLp()(Ω) and vLp()(Ω), then

    |Ωuvdx|(1p+1(p))uLp()(Ω)vLp()(Ω).

    The next proposition states the relation between norms and integrals of p(x)-th power.

    Proposition 2.2. Let

    ρ(u):=Ω|u|p(x)dx,uLp()(Ω),

    be the convex modular. Then the following assertions hold:

    (i) uLp()(Ω)<1 (resp. =1,>1) if and only if ρ(u)<1 (resp. =1,>1);

    (ii) uLp()(Ω)>1 implies upLp()(Ω)ρ(u)up+Lp()(Ω), and uLp()(Ω)<1 implies up+Lp()(Ω)ρ(u)upLp()(Ω);

    (iii) uLp()(Ω)0 if and only if ρ(u)0, and uLp()(Ω) if and only if ρ(u).

    The following result allows us to relate the norms of different Lebesgue variable exponent spaces (see [6] for a proof).

    Lemma 2.3. Suppose that p,qC+(¯Ω). Let fLq()p()(Ω). Then

    (i) fp+Lp()q()(Ω)fp()Lq()(Ω)fpLp()q()(Ω) if fLp()q()(Ω)1;

    (ii) fpLp()q()(Ω)fp()Lq()(Ω)fp+Lp()q()(Ω) if fLp()q()(Ω)1.

    Let us denote the distributional gradiente by Du. Then we can define the variable Sobolev space W1,p()(Ω) as

    W1,p()(Ω):={uLp()(Ω):|Du|Lp()(Ω)},

    equipped with the norm

    uW1,p()(Ω):=uLp()(Ω)+DuLp()(Ω),

    and we denote by W1,p()0(Ω) the closure of C0(Ω) in W1,p()(Ω). Notice that, due to the log-Hölder condition (2.1), C0(Ω) is dense in W1,p()(Ω). The following Embedding Theorem can be proved (see for instance [10]).

    Theorem 2.4. If p+<n, then

    0<S(p(),q(),Ω):=infvW1,p()0(Ω)DvLp()(Ω)vLq()(Ω),

    for all

    1q()p()=np()np().

    Remark 2.5. The q() exponent has to be uniformly subcritical, i.e., infΩ(p()q())>0, to enssure that W1,p()0(Ω)Lq()(Ω) is compact.

    Let X=W1,p()0(Ω). Recalling Definition 1.2, the operator Δp(x) can be seen as the weak derivative of the functional J:XR,

    J(u):=Ω1p(x)|Du|p(x)dx,

    in the sense that if L:=J:XX then

    (L(u),v)=Ω|Du|p(x)2DuDvdx,u,vX.

    We also recall the following properties from [11].

    Theorem 2.6. Let X=W1,p()0(Ω). Then:

    (i) L:XX is continuous, bounded and strictly monotone;

    (ii) L is a mapping of type (S+), that is, if unu in X and

    lim supn(L(un)L(u),unu)0

    then unu in X;

    (iii) L is a homeomorphism.

    Considering the variable Sobolev spaces defined before, we can already introduce the notion of weak solution.

    Definition 2.7. We say that uW1,p(x)(Ω) is a weak supersolution of (1.1) if for any non-negative φC0(Ω) there holds

    Ω|Du|p(x)2DuDφdxΩf(x,u,Du)φdx.

    Likewise, we say that uW1,p(x)(Ω) is a weak subsolution of (1.1) if

    Ω|Du|p(x)2DuDφdxΩf(x,u,Du)φdx

    for any non-negative φC0(Ω).

    Finally, uW1,p(x)(Ω) is a weak solution to (1.1) if it is a weak sub- and supersolution.

    Denote by Sn the set of symmetric n×n matrices. In order to introduce the concept of viscosity solution, let us recall the definition of jets.

    Definition 2.8. The superjet J2,u(x) of a function u:ΩR at xΩ is defined as the set of pairs (η,X)(Rn{0})×Sn satisfying

    u(y)u(x)+η(yx)+12X(yx)(yx)+o(|xy|2)

    as yx. The closure of a superjet is denoted by ¯J2,u(x) and it is defined as the set of pairs (η,X)Rn×Sn for which there exists a sequence (ηi,Xi)J2,u(xi), with xiΩ so that

    (xi,ηi,Xi)(x,η,X)as i.

    The subjet J2,+u(x) and its closure ¯J2,+u(x) are defined in a similar fashion.

    Observe that the operator can be written as

    Δp(x)φ(x)=tr(A(x,Dφ(x))D2φ(x))+B(x,Dφ(x)),

    where

    A(x,ξ):=|ξ|p(x)2(I+(p(x)2)ξ|ξ|ξ|ξ|),

    and

    B(x,ξ):=|ξ|p(x)2log|ξ|ξDp(x).

    We can now precise the notion of viscosity solution.

    Definition 2.9. A lower semicontinuous function u:ΩR is a viscosity supersolution of (1.1) if for any (η,X)J2,u(x) there holds

    tr(A(x,X))B(x,η)f(x,u(x),η).

    Similarly, an upper semicontinuous function u:ΩR is a viscosity subsolution of (1.1) if

    tr(A(x,X))B(x,η)f(x,u(x),η),

    for all (η,X)J2,+u(x). Finally, a viscosity solution is a continuous function which is a viscosity sub- and a supersolution.

    Observe that we do not require anything at jets of the form (0,X) or if J2,u(x)=. Moreover, the above definition of viscosity supersolution is equivalently given if we replace the superjet by its closure or if we take for jets (η,X) pairs of the form (Dφ(x),D2φ(x))(Rn{0})×Sn, where φ is smooth and touches u from below at x.

    A standard smoothing operator in the theory of viscosity solutions is the infimal convolution.

    Definition 2.10. Given ε>0 and q2 we define the infimal convolution of a function u:ΩR as

    uε(x):=infyΩ(u(y)+|xy|qqεq1),xΩ.

    The infimal convolution will be one of the main tools to prove that viscosity solutions are weak solutions. For the next result see for instance [14,20] and the references therein.

    Lemma 2.11. Let u be a bounded and lower semicontinuous function in Ω. Then:

    (i) There exists r(ε)>0 such that

    uε(x)=infyBr(ε)(x)(u(y)+|xy|qqεq1),

    where r(ε)0 as ε0.

    (ii) The sequence {uε}ε>0 is increasing as ε0 and uεu pointwise in Ω.

    (iii) uε is locally Lipschitz and twice differentiable a.e. Actually, for almost every x,yΩ,

    uε(y)=uε(x)+Duε(x)(xy)+12D2uε(x)(xy)2+o(|xy|2).

    (iv) uε is semiconcave, that is, there exists a constant C=C(q,ε,osc(u))>0 such that the function xuε(x)C|x|2 is concave. In particular

    D2uε(x)2CI,a.e.xΩ,

    where I is the identity matrix.

    (v) The set Yε(x):={yBr(ε)(x):uε(x)=u(y)+|xy|qqεq1} is non empty and closed for every xΩ.

    (vi) If xΩr(ε):={xΩ:dist(x,Ω)>r(ε)}, then there exists xεBr(ε) such that

    uε(x)=u(xε)+|xxε|qqεq1.

    (vii) If (η,X)J2,uε(x) with xΩr(ε), then

    η=(xxε)εq1|xxε|q2andXq1ε|η|q2q1I.

    Remark 2.12. For later purposes (see the proof of Lemma 3.3) we will choose

    q2 such that p2+q2q10. (2.2)

    Let us consider the inf-convolution uε given by Definition 2.10. We can summarize the strategy to prove Theorem 1.1 in several steps. Assuming that u is a viscosity supersolution, we will identify what problem is satisfied by uε in a pointwise sense, and later on in a weak sense. We will finish from here by passing to the limit in ε, obtaining the weak problem satisfied by u.

    We thus start by identifying the problem fulfilled by uε.

    Lemma 3.1. Assume p satisfies (1.3). Let u:ΩR locally Lipschitz, and letf=f(x,t,η) be continuous in Ω×R×Rn and non increasing in t. If u is a viscosity supersolution of (1.1) then

    Δp(x)uε(x)fε(x,uε(x),Duε(x))+E(ε)a.e. in Ωr(ε), (3.1)

    where

    fε(x,s,η):=infyBr(ε)(x)f(y,s,η),

    and E(ε)0 as ε0+. Here E(ε) depends only on p, q and ε.

    Notice that, differently from the constant case (see [19,Lemma 3.3]), when we identify the problem satisfied by uε in a pointwise sense, an error term E(ε) arises. We expect it that to disappear when passing to the limit in the final step. To prove this lemma we borrow some computations from [17,proof of Proposition 6.1], where they are used to prove a comparison-type result.

    Proof. Fix xΩr(ε) and let (η,Z)J2,uε(x), with η0. Then, by Lemma 2.11, there is xεBr(ε)(x) such that

    uε(x)=u(xε)+|xεx|qqεq1 and η=(xεx)εq1|xεx|q2. (3.2)

    Let φC2(Rn) such that φ touches uε from below at x and

    Dφ(x)=η,D2φ(x)=Z.

    Then, by definition of uε,

    u(y)φ(z)+|yz|qqεq1uε(z)φ(z)0, (3.3)

    for all y,zΩr(ε). Since by (3.2) we have

    u(xε)=φ(x)|xεx|qqεq1,

    it follows from (3.3) that

    u(y)φ(z)+|yz|qqεq1

    has a minimum at (xε,x). Thus,

    u(y)+φ(z)|yz|qqεq1

    attains its maximum over Ωr(ε)×Ωr(ε) at (xε,x). Let us consider

    Φ(y,z):=|yz|qqεq1.

    By the Maximum Principle for semicontinuous functions (see [4,Theorem 3.2]), there exist symmetric matrices (Y,Z) such that

    (η,Y)¯J2,u(xε),(η,Z)¯J2,+φ(x),

    and

    (Y00Z)D2Φ(xε,x)+εq1(D2Φ(xε,x))2, (3.4)

    with

    D2Φ(xε,x)=ε1q|xεx|q4[|xεx|2(IIII)+(q2)((xεx)(xεx)(xεx)(xεx)(xεx)(xεx)(xεx)(xεx))].

    Inequality (3.4) implies that, for any ξ,ηRn,

    ZξξYηηε1q[(q1)|xεx|q2+2(q1)2|xεx|2(q2)]|ηξ|2. (3.5)

    By the equivalence of the definition of viscosity solutions between tests functions and the closure of jets for continuous operators (recall η0), we deduce

    f(xε,u(xε),η)tr(A(xε,η)Y)B(xε,η)=tr(A(x,η)Z)tr(A(xε,η)Y)tr(A(x,η)Z)+B(x,η)B(xε,η)B(x,η). (3.6)

    Observe that since η0, A(,η) is symmetric and positive definite, and hence the square root A(x,η)1/2 exists and is symmetric. We define

    A(x)1/2:=A(x,η)1/2andA(xε)1/2:=A(xε,η)1/2.

    Now,

    tr(A(x,η)Z)=tr(A(x)1/2A(x)1/2Z)=nk=1ZAk(x)1/2Ak(x)1/2, (3.7)

    where Ak()1/2 is the k-th column of A()1/2. Hence, (3.5), (3.6), and (3.7) give

    f(xε,u(xε),η)nk=1ZAk(x)1/2Ak(x)1/2nk=1YAk(xε)1/2Ak(xε)1/2tr(A(x,η)Z)+B(x,η)B(xε,η)B(x,η)Cε1q|xεx|q2A(x)1/2A(xε)1/222+B(x,η)B(xε,η)tr(A(x,η)Z)B(x,η). (3.8)

    Proceeding as in [17,proof of Proposition 6.1], it can be seen that

    A(x)1/2A(xε)1/222A(x)A(xε)22(λmin(A(x))+λmin(A(xε)))2. (3.9)

    and, using that pC1,

    B(x,η)B(xε,η)|η|p(x)1|log|η|||Dp(x)Dp(xε)|+C|η|s1log2|η||p(x)p(xε)|. (3.10)

    for some s in the interval connecting p(x) and p(xε). Furthermore,

    A(x,η)A(xε,η)2C((p++1)|log|η|||η|s2+|η|p(xε)2)|xxε|, (3.11)

    and

    λmin(A(x)1/2)=(min|ξ|=1A(x,η)ξξ)1/2min{1,p(x)1}|η|p(x)22. (3.12)

    Thus, combining (3.8), (3.9), (3.10), (3.11) and (3.12), we deduce

    f(xε,u(xε),η)C|η|p(x)1|log|η|||xxε|+C|η|s1log2|η||xxε|+Cε1q(|log|η|||η|s2+|η|p(xε)2)2min{1,p1}[|η|p(x)22+|η|p(xε)22]2|xxε|qtr(A(x,η)Z)B(x,η). (3.13)

    By Lemma 2.11 ((ⅰ) and (ⅵ)), and since u is locally Lipschitz,

    |xxε|qqεq1|uε(x)u(xε)|qεq1|u(x)u(xε)|Cεq1|xxε|. (3.14)

    Hence,

    |η|=qε1q|xxε|q1C. (3.15)

    Consequently, the terms

    |η|p(x)1|log|η||and|η|s1log2|η|

    remain bounded, and so the first two terms in (3.13) tend to 0 as ε0+. For the third term, we obtain by (3.15) that

    ε1q(|log|η||η|s2||η|p(x)22+|η|p(xε)22)2|xxε|qClog2|η||η|2sp(x)2|η||xxε|Cεlog2|η||η|2sp(x)1.

    Since 2sp(x)1p(x)1p1>0, the term log2|η||η|2sp(x)1 is uniformly bounded for ε sufficiently small and thus

    ε1q(|log|η||η|s2||η|p(x)22+|η|p(xε)22)2|xxε|q0as ε0+.

    Regarding the term

    ε1q(|η|p(xε)2)2min{1,p1}[|η|p(x)22+|η|p(xε)22]2|xxε|q,

    by (3.15) it may be bounded by

    ε1q|η|p(xε)2|xxε|q=q1|η|p(xε)1|xxε|0

    as ε0+. Therefore, we get from (3.13) that

    f(xε,uε(x),η)=f(xε,u(xε),η)tr(A(x,η)Z)B(x,η)+E(ε),

    with E(ε)0 as ε0+. Thus, uε is a viscosity supersolution of (3.1) and, since it is twice differentiable almost everywhere, (3.1) holds a.e. in Ωr(ε).

    Remark 3.2. Observe that the Lipschitz condition on u was used to prove (3.14). If u is merely uniformly continuous, there exists a modulus of continuity ω so that |u(x)u(y)|ω(xy) for all x,yΩ. Hence, by Lemma 2.11 ((ⅰ) and (ⅵ)), we get

    |xεx|qCεq1ω(r(ε))

    and so

    |η|Cε1q|xεx|q1Cε(q1)/q.

    Consequently, the converges in (3.13) may not hold.

    Let us pass now from the pointwise formulation in Lemma 3.1 to a weak inequality. We will use in the proof some computations from the proof of [20,Lemma 5.5].

    Lemma 3.3. Assume f=f(x,t,η) uniformly continuous inΩ×R×Rn and Lipschitz continuous in η, satisfying (1.4). Suppose in addition that f(x,r,0)=0 for all (x,r)Ω×R. If u is a locally Lipschitz viscosity solution of (1.1), then for any non-negative φC0(Ω) there holds

    Ωr(ε)|Duε|p(x)2DuεDφdxΩr(ε)fε(x,uε,Duε)φdx+E(ε)Ωr(ε){Duε=0}φdx,

    for all ε>0 small enough.

    Proof. Let φC0(Ω), φ. Let \varepsilon be small enough so that \varphi \in \mathcal{C}_0^{\infty}(\Omega_{r(\varepsilon)}) . Since u_\varepsilon is semi-concave, there is a constant C(q, \varepsilon, u) > 0 so that

    \phi(x): = u_\varepsilon(x)-C(q, \varepsilon, u)|x|^{2}

    is concave in \Omega_{r(\varepsilon)} . Hence, by mollification, there is a sequence of smooth concave functions \phi_j so that

    (\phi_j, D\phi_j, D^{2}\phi_j) \to (\phi, D\phi, D^{2}\phi) \quad a.e. \, \, \Omega_{r(\varepsilon)}.

    Define

    u_{\varepsilon, j}(x): = \phi_j (x)+ C(q, \varepsilon, u)|x|^{2}.

    Given \delta > 0 , by integration by parts we obtain

    \begin{equation} \begin{split} \int_{\Omega_{r(\varepsilon)}} -\text{div }\left[ \left(\delta+ |Du_{\varepsilon, j}|^{2}\right)^{\frac{p(x)-2}{2}}Du_{\varepsilon, j}\right] \varphi\, dx = \int_{\Omega_{r(\varepsilon)}} \left(\delta+ |Du_{\varepsilon, j}|^{2}\right)^{\frac{p(x)-2}{2}}Du_{\varepsilon, j} \cdot D\varphi\, dx. \end{split} \end{equation} (3.16)

    Observe that, since u_\varepsilon is locally Lipschitz, there exists a constant M > 0 , independent of j , so that

    \begin{equation} \sup\limits_{j}\|D u_{\varepsilon, j}\|_{L^{\infty}(supp\, \varphi)}, \sup\limits_{j}\|D p_{j}\|_{L^{\infty}(supp\, \varphi)} \leq M. \end{equation} (3.17)

    Hence, by the Dominated Convergence Theorem, the right-hand side of (3.16) converges, as j \to \infty , to

    \int_{\Omega_{r(\varepsilon)}} (\delta+ |Du_\varepsilon|^{2})^{\frac{p(x)-2}{2}}Du_\varepsilon \cdot D\varphi\, dx.

    Let us treat now the left-hand side of (3.16). Observe that

    \begin{equation} \begin{split} &\int_{\Omega_{r(\varepsilon)}} -\text{div }\left[ \left(\delta+ |Du_{\varepsilon, j}|^{2}\right)^{\frac{p_j(x)-2}{2}}Du_{\varepsilon, j}\right] \varphi\, dx \\ & \quad = - \int_{\Omega_{r(\varepsilon)}}\left(\delta+ |Du_{\varepsilon, j}|^{2}\right)^{\frac{p_j(x)-2}{2}}\left( \Delta u_{\varepsilon, j} + \dfrac{p_j(x)-2}{\delta + |Du_{\varepsilon, j}|^{2}}\Delta_\infty u_{\varepsilon, j}\right)\varphi\, dx \\ & \qquad-\frac{1}{2}\int_{\Omega_{r(\varepsilon)}}\left(\delta+ |Du_{\varepsilon, j}|^{2}\right)^{\frac{p_j(x)-2}{2}}\log(\delta + |Du_{\varepsilon, j}|^{2})Du_{\varepsilon, j}\cdot Dp_j \varphi \, dx \\ & \qquad = : I_1 + I_2. \end{split} \end{equation} (3.18)

    By (3.17) and the Dominated Convergence Theorem we obtain that, when j \to \infty ,

    I_2 \to -\frac{1}{2}\int_{\Omega_{r(\varepsilon)}}\left(\delta+ |Du_{\varepsilon}|^{2}\right)^{\frac{p(x)-2}{2}}\log(\delta + |Du_{\varepsilon}|^{2})Du_{\varepsilon}\cdot Dp\, \varphi \, dx.

    For I_1 we will use Fatou's Lemma. Observe that by concavity of \phi_j ,

    D^{2}u_{\varepsilon, j} \leq C(q, \varepsilon, u)I.

    Hence, the integrand in I_1 is bounded from below by a constant independent of j if Du_{\varepsilon, j} = 0 . On the other hand, if Du_{\varepsilon, j} \neq 0 , it can be checked (see [20,Lemma 5.5]) that

    \left(\delta+ |Du_{\varepsilon, j}|^{2}\right)^{\frac{p_j(x)-2}{2}}\left( \Delta u_{\varepsilon, j} + \dfrac{p_j(x)-2}{\delta + |Du_{\varepsilon, j}|^{2}}\Delta_\infty u_{\varepsilon, j}\right) \leq C(\varepsilon, q, u, M, \delta).

    Taking \liminf as j \to \infty in (3.18), we obtain

    \begin{equation} \begin{split} &- \int_{\Omega_{r(\varepsilon)}}\left(\delta+ |Du_{\varepsilon}|^{2}\right)^{\frac{p(x)-2}{2}}\left( \Delta u_{\varepsilon} + \dfrac{p(x)-2}{\delta + |Du_{\varepsilon}|^{2}}\Delta_\infty u_{\varepsilon}\right)\varphi\, dx \\ & \qquad -\frac{1}{2}\int_{\Omega_{r(\varepsilon)}}\left(\delta+ |Du_{\varepsilon}|^{2}\right)^{\frac{p(x)-2}{2}}\log(\delta + |Du_{\varepsilon}|^{2})Du_{\varepsilon}\cdot Dp\, \varphi \, dx \\ & \leq \int_{\Omega_{r(\varepsilon)}} (\delta + |Du_\varepsilon|^{2})^{\frac{p(x)-2}{2}}Du_\varepsilon \cdot D\varphi\, dx. \end{split} \end{equation} (3.19)

    By the Dominated Convergence Theorem, as \delta \to 0 we have

    \begin{equation} \begin{split} \int_{\Omega_{r(\varepsilon)}}&\left(\delta+ |Du_{\varepsilon}|^{2}\right)^{\frac{p(x)-2}{2}}\log(\delta + |Du_{\varepsilon}|^{2})Du_{\varepsilon}\cdot Dp\, \varphi \, dx \to 2\int_{\Omega_{r(\varepsilon)}}|Du_{\varepsilon}|^{p(x)-2}\log|Du_{\varepsilon}|Du_{\varepsilon}\cdot Dp\, \varphi \, dx \end{split} \end{equation} (3.20)

    and

    \begin{equation} \int_{\Omega_{r(\varepsilon)}} (\delta+ |Du_\varepsilon|^{2})^{p(x)-2}Du_\varepsilon \cdot D\varphi\, dx \to \int_{\Omega_{r(\varepsilon)}} |Du_\varepsilon|^{p(x)-2}Du_\varepsilon \cdot D\varphi\, dx. \end{equation} (3.21)

    Moreover, using (2.2) and proceeding as in the proof of [20,Lemma 5.5], we can apply Fatou's lemma in the integral

    \int_{\Omega_{r(\varepsilon)} \setminus \left\lbrace Du_\varepsilon = 0 \right\rbrace}\left(\delta+ |Du_{\varepsilon}|^{2}\right)^{\frac{p(x)-2}{2}}\left( \Delta u_{\varepsilon} + \dfrac{p(x)-2}{\delta + |Du_{\varepsilon}|^{2}}\Delta_\infty u_{\varepsilon}\right)\, dx.

    Thus, from (3.19)–(3.21), and Lemma 3.1 we conclude that

    \begin{equation*} \begin{split} & \int_{\Omega_{r(\varepsilon)}} |Du_\varepsilon|^{p(x)-2}Du_\varepsilon \cdot D\varphi\, dx \\ & \qquad \geq \liminf\limits_{\delta \to 0} \int_{\Omega_{r(\varepsilon)} \setminus \left\lbrace Du_\varepsilon = 0 \right\rbrace}-\left(\delta+ |Du_{\varepsilon}|^{2}\right)^{\frac{p(x)-2}{2}}\left( \Delta u_{\varepsilon} + \dfrac{p(x)-2}{\delta + |Du_{\varepsilon}|^{2}}\Delta_\infty u_{\varepsilon}\right)\varphi\, dx \\ & \qquad + \liminf\limits_{\delta \to 0}\int_{\Omega_{r(\varepsilon)} \setminus \left\lbrace Du_\varepsilon = 0 \right\rbrace}- \frac{1}{2}\left(\delta+ |Du_{\varepsilon}|^{2}\right)^{\frac{p(x)-2}{2}}\log(\delta + |Du_{\varepsilon}|^{2})Du_{\varepsilon}\cdot Dp\, \varphi \, dx \\ & \qquad \geq -\int_{\Omega_{r(\varepsilon)} \setminus \left\lbrace Du_\varepsilon = 0 \right\rbrace}\Delta_{p(x)}u_\varepsilon\varphi \, dx \\ & \qquad \geq\int_{\Omega_{r(\varepsilon)} \setminus \left\lbrace Du_\varepsilon = 0 \right\rbrace} f_\varepsilon(x, u_\varepsilon, Du_\varepsilon) \varphi\, dx + E(\varepsilon)\int_{\Omega_{r(\varepsilon)} \setminus \left\lbrace Du_\varepsilon = 0 \right\rbrace} |Du_\varepsilon|^{\max\left\lbrace p(x)-2, 0\right\rbrace}\varphi\, dx \\ & \qquad = \int_{\Omega_{r(\varepsilon)}} f_\varepsilon(x, u_\varepsilon, Du_\varepsilon) \varphi\, dx + E(\varepsilon)\int_{\Omega_{r(\varepsilon)}} \varphi\, dx. \end{split} \end{equation*}

    The proofs of the following lemmas follow the strategy in [20,Lemma 5.6,Lemma 5.7] for the homogeneous case, so we will only highlight the differences coming from the non-homogeneous term.

    Lemma 3.4. Under the assumptions of Lemma 3.3, u \in W^{1, p(x)}_{loc}(\Omega) and, for each \Omega' \Subset \Omega , we have that, up to a subsequence, u_\varepsilon \to u weakly in W^{1, p(x)}(\Omega') as \varepsilon\to 0 .

    Proof. Let \Omega' \Subset \Omega and let \xi \in\mathcal{C}^{\infty}_0(\Omega) be such that 0 \leq \xi \leq 1 and \xi = 1 in \overline{\Omega'} . Assume that

    K: = supp\, \xi \subset \Omega_{r(\varepsilon)},

    and define

    \varphi: = (L-u_\varepsilon)\xi^{p^+}, \quad \mbox{with } L: = \sup\limits_{\varepsilon, \Omega'}|u_\varepsilon(x)|.

    By Lemma 3.3, we have

    \begin{equation} \begin{split} &\int_{\Omega_{r(\varepsilon)}}|Du_\varepsilon|^{p(x)}\xi^{p^+}dx \leq \int_{\Omega_{r(\varepsilon)}}|Du_\varepsilon|^{p(x)-1}\xi^{p^+-1}(L-u_\varepsilon)p^+|D\xi|\, dx \\ & \qquad \qquad+ \int_{\Omega_{r(\varepsilon)}}|f_\varepsilon(x, u_\varepsilon, Du_\varepsilon)\varphi \, dx + |E(\varepsilon)|\int_{\Omega_{r(\varepsilon)}}\varphi\, dx. \end{split} \end{equation} (3.22)

    By using Young's inequality it can be seen that

    \begin{equation} \begin{split} \int_{\Omega_{r(\varepsilon)}}|Du_\varepsilon|^{p(x)-1}\xi^{p^+-1}(L-u_\varepsilon)p^+|D\xi|\, dx &\leq \delta \int_{\Omega_{r(\varepsilon)}}|Du_\varepsilon|^{p(x)}\xi^{p^+}dx + C(\delta, p, , L, D\xi), \end{split} \end{equation} (3.23)

    and, using (1.4),

    \begin{equation} \begin{split} \int_{\Omega_{r(\varepsilon)}}|f_\varepsilon(x, & u_\varepsilon, Du_\varepsilon)|\varphi \, dx \\ &\leq \gamma_\infty \int_{\Omega_{r(\varepsilon)}}|Du_\varepsilon|^{p(x)-1}\xi^{p^+-1}(L-u_\varepsilon)\xi\, dx + \int_{\Omega_{r(\varepsilon)}}\phi(x)(L-u_\varepsilon)\xi^{p^+}\, dx \\ & \leq \gamma_\infty \delta \int_{\Omega_{r(\varepsilon)}}|Du_\varepsilon|^{p(x)}\xi^{p^+}+ \gamma_\infty \int_{\Omega_{r(\varepsilon)}}\left(\frac{2}{\delta}Lp^+\right)^{p(x)}\, dx+C(\phi, L, \Omega) \\ & \leq \gamma_\infty \delta \int_{\Omega_{r(\varepsilon)}}|Du_\varepsilon|^{p(x)}\xi^{p^+}+C(\phi, \delta, p, L, \gamma, \Omega). \end{split} \end{equation} (3.24)

    Finally, it is easy to check that

    \begin{equation} \begin{split} |E(\varepsilon)|\int_{\Omega_{r(\varepsilon)}}\varphi\, dx & \leq |E(\varepsilon)| C(L, p, \Omega). \end{split} \end{equation} (3.25)

    Combining (3.22)–(3.25) and recalling that \xi = 1 in \Omega' , we obtain the uniform boundedness of Du_\varepsilon in L^{p(x)}(\Omega') . Therefore, up to a subsequence, u_\varepsilon \to u weakly in W^{1, p(x)}(\Omega') as \varepsilon\to 0 .

    Lemma 3.5. Under the assumptions of Lemma 3.3, for each \Omega' \Subset \Omega , we have that, up to a subsequence, u_\varepsilon \to u in W^{1, p(x)}(\Omega') as \varepsilon\to 0 .

    Proof. Let \Omega' \Subset \Omega and let \xi \in \mathcal{C}^{\infty}_0(\Omega) be such that 0 \leq \xi \leq 1 and \xi = 1 in \overline{\Omega'} . Consider the test function

    \varphi: = (u-u_\varepsilon)\xi,

    and choose \varepsilon small enough so that K: = \text{supp}\, \varphi \subset \Omega_{r(\varepsilon)} . Observe that \varphi \in W^{1, p(x)}(\Omega) and has compact support. By Lemma 3.3 and (1.4), we have

    \begin{equation} \begin{split} \int_{\Omega_{r(\varepsilon)}} &\left(|Du|^{p(x)-2}Du - |Du_\varepsilon|^{p(x)-2}Du_\varepsilon\right)\cdot (Du - Du_\varepsilon)\xi\, dx \\ & \leq\int_{\Omega_{r(\varepsilon)}} |Du|^{p(x)-2}Du \cdot (Du - Du_\varepsilon)\xi\, dx + \int_{\Omega_{r(\varepsilon)}} |f_\varepsilon(x, u_\varepsilon, Du_\varepsilon)|(u-u_\varepsilon)\xi\, dx \\ & \quad + |E(\varepsilon)|\int_{\Omega_{r(\varepsilon)}} (u-u_\varepsilon)\xi\, dx + \int_{\Omega_{r(\varepsilon)}} |Du_\varepsilon|^{p(x)-2}Du_\varepsilon\cdot D\xi(u-u_\varepsilon)dx \\ & \leq\int_{\Omega_{r(\varepsilon)}} |Du|^{p(x)-2}Du \cdot (Du - Du_\varepsilon)\xi\, dx \\ &\quad + \|u-u_\varepsilon \|_{L^{\infty}(K)}\left(\gamma_\infty \int_K |Du_\varepsilon|^{p(x)-1} dx + \int_K \phi(x)\, dx\right) \\ & \quad +\|u-u_\varepsilon \|_{L^{\infty}(K)}\left(|E(\varepsilon)||K|+\int_K |Du_\varepsilon|^{p(x)-1}D\xi dx\right). \end{split} \end{equation} (3.26)

    Since u_\varepsilon \to u locally uniformly and, by Lemma 3.4, u_\varepsilon \to u weakly in W^{1, p(x)}(K) , the right-hand side of (3.26) converges to 0 , up to a subsequence, when \varepsilon \to 0 . By Theorem 2.6, the operator L: W^{1, p(x)}(\Omega') \to [W^{1, p(x)}(\Omega')]^{*} given by

    \langle L(v), w\rangle = \int_{\Omega'}|Dv|^{p(x)-2}Dv\cdot Dw\, dx, \quad v, w \in W^{1, p(x)}(\Omega')

    is a mapping of type ( S_{+} ). Then, it follows from (3.26) that u_\varepsilon \to u strongly in W^{1, p(x)}(\Omega') as \varepsilon\to 0 .

    We can already prove that viscosity solutions of (1.1) are also weak solutions.

    Proof of Theorem 1.1. Let \varphi \in \mathcal{C}_0^{\infty}(\Omega) and take \Omega' \Subset\Omega such that

    \text{supp }\varphi \subset \Omega'.

    Let us fix \varepsilon_0 > 0 such that 0 < \varepsilon < \varepsilon_0 implies

    \Omega' \subset \Omega_{r(\varepsilon)}.

    In view of Lemma 3.3, to prove the theorem, it will be enough to show the following convergences:

    ({\rm{I}})

    \lim\limits_{\varepsilon \to 0^{+}} \int_{\Omega'}|Du_\varepsilon|^{p(x)-2}Du_\varepsilon \cdot D\varphi\, dx = \int_{\Omega'}|Du|^{p(x)-2}Du\cdot D\varphi\, dx

    ({\rm{II}})

    \lim\limits_{\varepsilon \to 0^{+}} \int_{\Omega_{r(\varepsilon)}}f_\varepsilon(x, u_\varepsilon, Du_\varepsilon)\varphi\, dx = \int_{\Omega'}f(x, u, Du)\varphi \, dx

    ({\rm{III}})

    \lim\limits_{\varepsilon \to 0^{+}} E(\varepsilon)\int_{\Omega_{r(\varepsilon)}}\varphi \, dx = 0.

    Proceeding exactly as in the proof of [20,Theorem 5.8], it can be seen that (Ⅰ) holds, and (Ⅲ) follows in a straightforward way.

    Let us prove (Ⅱ). Let \varepsilon , \varphi and \Omega' as above. By the uniform continuity of f , for every \rho > 0 , there exists \delta > 0 such that

    |f(x, u_\varepsilon(x), D u_\varepsilon(x))-f(y, u_\varepsilon(x), D u_\varepsilon(x))| \leq \rho, \quad y \in B_\delta(x).

    Choose \varepsilon_0 > 0 so that r(\varepsilon) < \delta for every \varepsilon < \varepsilon_0 . Thus, from the previous inequality we get

    f(x, u_\varepsilon(x), D u_\varepsilon(x)) < \rho+f(y, u_\varepsilon(x), D u_\varepsilon(x)),

    for every x\in \Omega' and y\in B_{r(\varepsilon)}(x) . In particular,

    f(x, u_\varepsilon(x), D u_\varepsilon(x)) < \rho+f_\varepsilon(x, u_\varepsilon(x), D u_\varepsilon(x)),

    and therefore

    0 \leq |f(x, u_\varepsilon(x), D u_\varepsilon(x))-f_\varepsilon(x, u_\varepsilon(x), D u_\varepsilon(x))| < \rho.

    Hence,

    \begin{equation} \int_{\Omega'} |f(x, u_\varepsilon, D u_\varepsilon)-f_\varepsilon(x, u_\varepsilon, D u_\varepsilon)|\varphi\, dx \leq \rho \|\varphi\|_{L^{\infty}(\Omega')}|\Omega'|, \end{equation} (3.27)

    for \rho arbitrarily small. Since \|u_\varepsilon\|_{L^{\infty}(\Omega')} \leq \|u\|_{L^{\infty}(\Omega')} for all \varepsilon , it follows

    \max\limits_{[-\|u_\varepsilon\|_{L^{\infty}}, \|u_\varepsilon\|_{L^{\infty}}]}|\gamma(t)| \leq \max\limits_{[-\|u\|_{L^{\infty}}, \|u\|_{L^{\infty}}]}|\gamma(t)|,

    and then by (1.4) we have

    |f(x, u_\varepsilon, D u)| \leq C|D u|^{p(x)-1}+\phi(x) \in L^{p'(x)}(\Omega') \subset L^{1}(\Omega')

    for a constant C independent of \varepsilon . Then, by the Lebesgue Convergence Theorem,

    \begin{equation} \lim\limits_{\varepsilon \to 0}\int_{\Omega'} f(x, u_\varepsilon, D u)\varphi\, dx = \int_{\Omega'}f(x, u, D u)\varphi\, dx. \end{equation} (3.28)

    Moreover, the convergence Du_\varepsilon \to Du in L^{p(x)}(\Omega') and the Lipschitz continuity of f with respect to the third variable imply

    \begin{equation} \begin{split} \int_{\Omega'}|f(x, u_\varepsilon, Du_\varepsilon)-f(x, u_\varepsilon, Du )|\varphi\, dx \leq C\int_{\Omega'}|Du_\varepsilon-Du|\, dx \to 0 \quad \text{ as } \varepsilon \to 0. \end{split} \end{equation} (3.29)

    Therefore, combining (3.27)–(3.29) we obtain (Ⅱ).

    To prove this implication we follow the strategy of [19]. As we said in the Introduction, the argument is strongly connected to the availability of comparison principles. After the proof of the theorem we will state and prove an example of comparison result that applies here.

    Proof of Theorem 1.3. Let u \in \mathcal{C}(\Omega) be a weak supersolution to (1.1). To reach a contradiction, assume that u is not a viscosity supersolution. By assumption, there exist x_0 \in \Omega and \varphi \in \mathcal{C}^{2}(\Omega) so that D \varphi (x_0) \neq 0 ,

    \begin{equation} u(x_0) = \varphi(x_0), \quad u(x) > \varphi(x) \mbox{ for all $x\neq x_0$}, \end{equation} (4.1)

    and

    \begin{equation} -\Delta_{p(x_0)} \varphi(x_0) < f(x_0, u(x_0), D \varphi(x_0)). \end{equation} (4.2)

    Moreover, the mapping

    x \to f(x, u(x), D \varphi(x))

    is continuous in \Omega , and (4.1) yields

    -\Delta_{p(x)} \varphi(x)-f(x, u(x), D\varphi(x)) < 0, \quad \mbox{ for all }x\in B_r(x_0),

    for some r > 0 small eunogh. Hence, there exists r_0 > 0 so that D \varphi(x) \neq 0 for all x \in B_{r_0}(x_0) and

    \begin{equation} -\Delta_{p(x)}\varphi(x)\leq f(x, u(x), D \varphi(x)), \qquad x\in B_{r_0}(x_0). \end{equation} (4.3)

    Let

    m: = \min\limits_{\partial B_{r_0}(x_0)}\left( u-\varphi\right).

    Then by (4.1), m > 0 . Consider

    \tilde{\varphi}(x): = \varphi(x) +m, \quad x \in \Omega.

    By (4.3), \tilde{\varphi} is a weak subsolution to

    \begin{equation} -\Delta_{p(x)} v = \tilde{f}(x, D v), \end{equation} (4.4)

    in B_{r_0}(x_0) , where \tilde{f}(x, \eta): = f(x, u(x), \eta) . Observe that \tilde{f} is locally Lipschitz in \Omega \times \mathbb{R}^{n} . Moreover, in the weak sense, we have

    \begin{equation*} -\Delta_{p(x)} u \geq f(x, u, D u) = \tilde{f}(x, D u), \end{equation*}

    which shows that u is a weak supersolution to (4.4). In addition, observe that u \geq \tilde{\varphi} on \partial B_{r_0}(x_0) , and that (4.5) holds since D \tilde{\varphi} \neq 0 in B_{r_0}(x_0) . Thus, by the (CPP) we conclude that u \geq \tilde{\varphi} in B_{r_0}(x_0) . This contradicts (4.1).

    Some comparison principles may be found in the literature for p(x) -Laplace equations. See for instance [9,17,21]. Here, we provide a comparison principle for a Lipschitz right-hand side depending on all the lower terms.

    Theorem 4.1. Assume that f = f(x, r, \eta) is locally Lipschitz in \Omega \times \mathbb{R}\times \mathbb{R}^{n} . Let u, v \in \mathcal{C}^{1}(\Omega) be weak sub- and supersolutions, respectively, of (1.1) such that

    \begin{equation} |D u(x)|+ |D v(x)| > 0, \end{equation} (4.5)

    for a.e. x satisfying p(x) > 2 . Then, there exists \delta > 0 such that for any domain B , \overline{B} \subset \Omega , with |B| < \delta and u \leq v on \partial B , there holds u \leq v in B .

    Proof. Take (u-v)^{+}\chi_B \in W_0^{1, p(x)}(\Omega) as a test function in (1.1) to get

    \begin{equation} \begin{split} &\int_B \left( |D u|^{p(x)-2}D u - |D v|^{p(x)-2}D v \right)\cdot D (u-v)^{+} \\ & \qquad \leq \int_B \frac{f(x, u, D u)-f(x, v, D v)}{u-v}[(u-v)^{+}]^{2}. \end{split} \end{equation} (4.6)

    Now, using the inequality

    c(|\xi|+|\eta|)^{p(x)-2}|\xi-\eta|^{2}\leq \left( |\xi|^{p(x)-2}\xi-|\eta|^{p(x)-2}\eta\right)\cdot ( \xi-\eta),

    the Lipschitz assumption on f and the boundedness of u and v in \mathcal{C}^{1} , we get

    \begin{equation} \begin{split} &\int_B (|D u|+ |D v|)^{p(x)-2}|D (u-v)^{+}|^{2} \leq C(u, v) \left[\int_B [(u-v)^{+}]^{2}+ \int_B |D u-D v|(u-v)^{+}\right]. \end{split} \end{equation} (4.7)

    By Poincaré and Hölder inequalities, and the assumption (4.5), we obtain

    \begin{equation} \begin{split} C(u, &v)\left[\int_B [(u-v)^{+}]^{2}+ \int_B |D u-D v|(u-v)^{+}\right] \\ & \leq C(B)C(u, v)\int_B |D (u-v)^{+}|^{2} \quad ( \text{here, }C(B) \to 0 \, \, \text{as }|B|\to 0) \\ & = C(B)C(u, v)\int_B (|D u|+ |D v|)^{2-p(x)}(|D u|+ |D v|)^{p(x)-2}|D (u-v)^{+}|^{2} \\ & \leq C(B)C(u, v)\int_B (|D u|+ |D v|)^{p(x)-2}|D (u-v)^{+}|^{2}. \end{split} \end{equation} (4.8)

    Combining (4.7) and (4.8), we obtain

    \int_B (|D u|+ |D v|)^{p(x)-2}|D (u-v)^{+}|^{2} \leq C(B)C(u, v)\int_B (|D u|+ |D v|)^{p(x)-2}|D (u-v)^{+}|^{2}.

    Hence, for |B| small enough, we get (u-v)^{+} = 0 in B and hence u \leq v in B .

    Remark 4.2. Keeping track of the proof of Theorem 1.3, it is easy to see that Theorem 4.1 allows us to prove that weak solutions are viscosity. Indeed, it is enough with choosing r_0 sufficiently small so that |B_{r_0}(x_0)| < \delta , with \delta > 0 provided by Theorem 4.1.

    M. Medina has been supported by Project PDI2019-110712GB-100, MICINN, Spain. P. Ochoa has been supported by CONICET and Grant B080, UNCUYO, Argentina.

    The authors declare no conflict of interest.



    [1] B. Barrios, M. Medina, Equivalence of weak and viscosity solutions in fractional non-homogeneous problems, Math. Ann., 381 (2021), 1979–2012. https://doi.org/10.1007/s00208-020-02119-w doi: 10.1007/s00208-020-02119-w
    [2] J. E. M. Braga, R. A. Leitao, J. E. L. Oliveira, Free boundary theory for singular/degenerate nonlinear equations with right hand side: a non-variational approach, Calc. Var., 59 (2020), 86. https://doi.org/10.1007/s00526-020-01733-5 doi: 10.1007/s00526-020-01733-5
    [3] Y. Chen, S. Levine, M. Rao, Variable exponent, linear growth functionals in image restoration, SIAM J. Appl. Math., 66 (2006), 1383–1406. https://doi.org/10.1137/050624522 doi: 10.1137/050624522
    [4] M. G. Crandall, H. Ishii, P.-L. Lions, User's guide to viscosity solutions of second order partial differential equations, Bull. Amer. Math. Soc., 27 (1992), 1–67. https://doi.org/10.1090/S0273-0979-1992-00266-5 doi: 10.1090/S0273-0979-1992-00266-5
    [5] l. Diening, P. Harjulehto, P. Hästö, M. Ružička, Lebesgue and Sobolev Spaces with Variable Exponents, Berlin, Heidelberg: Springer, 2011. https://doi.org/10.1007/978-3-642-18363-8
    [6] D. Edmunds, J. Rákosník, Sobolev embeddings with variable exponent, Stud. Math., 143 (2000), 267–293. https://doi.org/10.4064/sm-143-3-267-293 doi: 10.4064/sm-143-3-267-293
    [7] X. Fan, D. Zhao, A class of De Giorgi type and Hölder continuity, Nonlinear Anal. Theor., 36 (1999), 295–318. https://doi.org/10.1016/S0362-546X(97)00628-7 doi: 10.1016/S0362-546X(97)00628-7
    [8] X. Fan, D. Zhao, The quasi-minimizer of integral functionals with m(x)-growth conditions, Nonlinear Anal. Theor., 39 (2000), 807–816. https://doi.org/10.1016/S0362-546X(98)00239-9 doi: 10.1016/S0362-546X(98)00239-9
    [9] X. Fan, Y. Zhao, Q. Zhang, A strong maximum principle for p(x)-Laplace equations, Chinese Journal of Contemporary Mathematics, 21 (2000), 277–282.
    [10] X. Fan, D. Zhao, On the spaces L^{p(x)}(\Omega) and W^{1, p(x)}(\Omega), J. Math. Anal. Appl., 263 (2001), 424–446. https://doi.org/10.1006/jmaa.2000.7617 doi: 10.1006/jmaa.2000.7617
    [11] X. Fan, Q. Zhang, Existence of solutions for p(x)-Laplacian Dirichlet problem, Nonlinear Anal. Theor., 52 (2003), 1843–1852. https://doi.org/10.1016/S0362-546X(02)00150-5 doi: 10.1016/S0362-546X(02)00150-5
    [12] F. Ferrari, C. Lederman, Regularity of flat free boundaries for a p(x)-Laplacian problem with right hand side, Nonlinear Anal., 212 (2021), 112444. https://doi.org/10.1016/j.na.2021.112444 doi: 10.1016/j.na.2021.112444
    [13] H. Ishii, On the equivalence of two notions of solutions, viscosity solutions and distribution solutions, Funkcislaj Ekvacioj, 38 (1995), 101–120.
    [14] V. Julin, P. Juutinen, A new proof for the equivalence of weak and viscosity solutions for the p-Laplace equation, Commun. Part. Diff. Eq., 37 (2012), 934–946. https://doi.org/10.1080/03605302.2011.615878 doi: 10.1080/03605302.2011.615878
    [15] P. Juutinen, P. Lindqvist, A theorem of Radó type for solutions of a quasi-linear equation, Math. Res. Lett., 11 (2004), 31–34. https://doi.org/10.4310/MRL.2004.v11.n1.a4 doi: 10.4310/MRL.2004.v11.n1.a4
    [16] P. Juutinen, P. Lindqvist, J. J. Manfredi, On the equivalence of viscosity solutions and weak solutions for a quasilinear equation, SIAM J. Math. Anal., 33 (2001), 699–717. https://doi.org/10.1137/S0036141000372179 doi: 10.1137/S0036141000372179
    [17] P. Juutinen, T. Lukkari, M. Parviainen, Equivalence of viscosity solutions and weak solutions for the p(x)-Laplacian, Ann. Ins. H. Poincaré Anal. Non Linéaire, 27 (2010), 1471–1487. https://doi.org/10.1016/j.anihpc.2010.09.004 doi: 10.1016/j.anihpc.2010.09.004
    [18] J. Korvenpää, T. Kuusi, E. Lindgren, Equivalence of solutions to fractional p-Laplace type equations, J. Math. Pure. Appl., 132 (2019), 1–26. https://doi.org/10.1016/j.matpur.2017.10.004 doi: 10.1016/j.matpur.2017.10.004
    [19] M. Medina, P. Ochoa, On viscosity and weak solutions for non-homogeneous p-Laplace equations, Adv. Nonlinear Anal., 8 (2019), 468–481. https://doi.org/10.1515/anona-2017-0005 doi: 10.1515/anona-2017-0005
    [20] J. Siltakoski, Equivalence of viscosity solutions and weak solutions for the normalized p(x)-Laplacian, Calc. Var., 57 (2018), 95. https://doi.org/10.1007/s00526-018-1375-1 doi: 10.1007/s00526-018-1375-1
    [21] P. Takac, J. Giacomoni, A p(x)-Laplacian extension of the Díaz-Saa inequality and some applications, P. Roy. Soc. Edinb. A, 150 (2020), 205–232. https://doi.org/10.1017/prm.2018.91 doi: 10.1017/prm.2018.91
  • This article has been cited by:

    1. Yuzhou Fang, Vicenţiu D. Rădulescu, Chao Zhang, Equivalence of weak and viscosity solutions for the nonhomogeneous double phase equation, 2023, 0025-5831, 10.1007/s00208-023-02593-y
  • Reader Comments
  • © 2023 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(1777) PDF downloads(147) Cited by(1)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog