Processing math: 100%
Research article Special Issues

The ergodic limit for weak solutions of elliptic equations with Neumann boundary condition

  • We consider the so-called it ergodic problem for weak solutions of elliptic equations in divergence form, complemented with Neumann boundary conditions. The simplest example reads as the following boundary value problem in a bounded domain of RN: {div(A(x)u)+λ=H(x,u)inΩ,A(x)un=0onΩ, where A(x) is a coercive matrix with bounded coefficients, and H(x,u) has Lipschitz growth in the gradient and measurable x-dependence with suitable growth in some Lebesgue space (typically, |H(x,u)|b(x)|u|+f(x) for functions b(x)∈ LN(Ω) and f (x) ∈ Lm(Ω), m1). We prove that there exists a unique real value λ for which the problem is solvable in Sobolev spaces and the solution is unique up to addition of a constant. We also characterize the ergodic limit, say the singular limit obtained by adding a vanishing zero order term in the equation. Our results extend to weak solutions and to data in Lebesgue spaces LN(Ω) (or in the dual space (H1(Ω))'), previous results which were proved in the literature for bounded solutions and possibly classical or viscosity formulations.

    Citation: François Murat, Alessio Porretta. The ergodic limit for weak solutions of elliptic equations with Neumann boundary condition[J]. Mathematics in Engineering, 2021, 3(4): 1-20. doi: 10.3934/mine.2021031

    Related Papers:

    [1] Mouhamed Moustapha Fall, Veronica Felli, Alberto Ferrero, Alassane Niang . Asymptotic expansions and unique continuation at Dirichlet-Neumann boundary junctions for planar elliptic equations. Mathematics in Engineering, 2019, 1(1): 84-117. doi: 10.3934/Mine.2018.1.84
    [2] David Cruz-Uribe, Michael Penrod, Scott Rodney . Poincaré inequalities and Neumann problems for the variable exponent setting. Mathematics in Engineering, 2022, 4(5): 1-22. doi: 10.3934/mine.2022036
    [3] 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
    [4] 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
    [5] Huyuan Chen, Laurent Véron . Weak solutions of semilinear elliptic equations with Leray-Hardy potentials and measure data. Mathematics in Engineering, 2019, 1(3): 391-418. doi: 10.3934/mine.2019.3.391
    [6] Antonio Vitolo . Singular elliptic equations with directional diffusion. Mathematics in Engineering, 2021, 3(3): 1-16. doi: 10.3934/mine.2021027
    [7] Feida Jiang . Weak solutions of generated Jacobian equations. Mathematics in Engineering, 2023, 5(3): 1-20. doi: 10.3934/mine.2023064
    [8] Mónica Clapp, Filomena Pacella . Existence of nonradial positive and nodal solutions to a critical Neumann problem in a cone. Mathematics in Engineering, 2021, 3(3): 1-15. doi: 10.3934/mine.2021022
    [9] Italo Capuzzo Dolcetta . The weak maximum principle for degenerate elliptic equations: unbounded domains and systems. Mathematics in Engineering, 2020, 2(4): 772-786. doi: 10.3934/mine.2020036
    [10] Andrea Manzoni, Alfio Quarteroni, Sandro Salsa . A saddle point approach to an optimal boundary control problem for steady Navier-Stokes equations. Mathematics in Engineering, 2019, 1(2): 252-280. doi: 10.3934/mine.2019.2.252
  • We consider the so-called it ergodic problem for weak solutions of elliptic equations in divergence form, complemented with Neumann boundary conditions. The simplest example reads as the following boundary value problem in a bounded domain of RN: {div(A(x)u)+λ=H(x,u)inΩ,A(x)un=0onΩ, where A(x) is a coercive matrix with bounded coefficients, and H(x,u) has Lipschitz growth in the gradient and measurable x-dependence with suitable growth in some Lebesgue space (typically, |H(x,u)|b(x)|u|+f(x) for functions b(x)∈ LN(Ω) and f (x) ∈ Lm(Ω), m1). We prove that there exists a unique real value λ for which the problem is solvable in Sobolev spaces and the solution is unique up to addition of a constant. We also characterize the ergodic limit, say the singular limit obtained by adding a vanishing zero order term in the equation. Our results extend to weak solutions and to data in Lebesgue spaces LN(Ω) (or in the dual space (H1(Ω))'), previous results which were proved in the literature for bounded solutions and possibly classical or viscosity formulations.


    dedicato a Italo, con stima e amicizia.

    Let Ω be a bounded, sufficiently regular, connected domain in RN, N1, and let n denote the outward normal unit vector on the boundary Ω. It is well known (see e.g., [12,15]) that, if H(x,p) is a Lipschitz function for (x,p)Ω×RN, then there is a unique real number λ such that the elliptic problem

    {Δu+H(x,u)+λ=0in Ω,un=0on Ω (1.1)

    admits a solution, and this solution is unique up to a constant. The simplest example of this type of problems occurs in the linear case, when H(x,u)=b(x)uf(x). If b(x) is a Lipschitz continuous, divergence free vector field, then λ is the average value of f; if b(x) is not divergence free, the uniqueness of λ is a consequence of Fredholm theory for linear operators (see e.g., [10]), in which case λ=Ωf(x)φ1dx, where φ1 is the first eigenfunction of the adjoint problem (normalized so that Ωφ1=1).

    If H is a nonlinear function, the existence and uniqueness of λ was proved in [15] assuming that H(x,p) satisfies fairly general structure conditions with respect to p and enough regularity with respect to x.

    The real number λ appearing in this kind of problems is sometimes called additive eigenvalue and is definitively a critical value which plays a role in many different contexts. If H(x,p) is convex in p, then λ can be interpreted as the optimal value of an ergodic stochastic control problem; we refer the reader to [2,3,12,15] and especially to [1] for an extensive presentation of the ergodic stochastic control setting. In that framework, it is natural to obtain λ as the limit of εuε, where uε solves the approximating coercive problem

    {Δu+H(x,u)+εu=0in Ω,un=0on Ω. (1.2)

    Indeed, from Bellman's principle, (1.2) is the equation solved by the value function of an infinite horizon stochastic control problem, where ε is the discount factor. The vanishing discount limit leads, through time averaging, to the ergodic control problem represented by (1.1), and λ=limε0εuε. This interpretation awarded to λ the name of ergodic constant, and to the singular limit of solutions of (1.2), as ε0, the name of ergodic limit.

    The constant λ is also a critical value for the long time behavior of the evolution problem, since it represents the asymptotic speed of time-dependent solutions: typically, a solution of the evolution problem vtΔv+H(x,v)=0 (with Neumann conditions) satisfies v(x,t)tu(x)+λt for some stationary solution u of (1.1). Again, this is consistent with the ergodicity property of the underlying controlled stochastic trajectory, but of course the long time convergence itself does not need the convexity of H, at least for Lipschitz nonlinearities. Finally, λ also plays a crucial role in homogenization problems (in which context problem (1.1) is referred to as the cell problem), see [4].

    A huge literature has been devoted so far to the existence and characterization of ergodic constants, as well as to the study of ergodic limits and of the long time behavior of evolution problems, at the point that it is impossible here to recall such a long list of contributions. Most papers concerned with the above issues treat the problem in the framework of viscosity solutions' theory, for both second and first order Hamilton-Jacobi-Bellman equations. This explains why this kind of results were proved under many different structure conditions on the "Hamiltonian" H(x,u) as well as on the second order operator, but mostly assuming a regular dependence with respect to x. This regularity is often required for verification theorems, whenever the application to stochastic control is the main motivation.

    The purpose of this note is to give a prototype result of existence, uniqueness of the ergodic constant and a characterization of the ergodic limit under natural assumptions for elliptic operators in divergence form, replacing the L framework (and most times continuity, needed for viscosity solutions) with the L2-setting which is natural for weak solutions in the Sobolev space H1(Ω). To be precise, we consider the elliptic problem

    {div(A(x)u)+λ=H(x,u)+χin Ω,A(x)un=0on Ω, (1.3)

    where χ(H1(Ω)) (the dual space of H1(Ω)), A(x) is a measurable matrix such that

    A(x)L(Ω)N×N,A(x)ξξα|ξ|2, (1.4)

    for some α>0, and where H(x,p) is a Carathéodory function (measurable in x, continuous in p) satisfying the following linear growth condition (this is for N3)

    |H(x,p)|b(x)|p|+f(x),for some b(x)LN(Ω)f(x)L2NN+2(Ω), (1.5)

    for almost every xΩ and for every pRN.

    Let us stress that f(x) (and therefore H(x,0)) is not assumed to be bounded, and only belongs to L(2)(Ω), where 2=2NN2 is the Sobolev exponent for N3. The case N=1,2 is mentioned later, see Remark 2.4. Due to Sobolev embedding, the condition bLN(Ω) is the usual threshold for Lebesgue summability of drift terms in elliptic equations, see e.g., [11]; in fact b(x)|u|L2NN+2(Ω) whenever uL2(Ω) and b(x)LN(Ω).

    Here and below, we assume that Ω is a Lipschitz bounded and connected domain in RN; the Lipschitz regularity being just one possible condition which ensures that the Sobolev embedding (and the Poincaré-Wirtinger inequality) hold true.

    Eventually, for the purpose of uniqueness, we will also assume the following Lipschitz condition upon H, namely that

    |H(x,p)H(x,q)|b(x)|pq|,b(x)LN(Ω), (1.6)

    for almost every xΩ and every p,qRN.

    The first main result that we prove in this note is the following.

    Theorem 1.1. Let N3. Assume that A(x) satisfies (1.4) and that H(x,p) satisfies (1.5). Then there exist a constant λR and a function uH1(Ω) which satisfy the elliptic problem (1.3) in the weak sense, i.e.,

    ΩA(x)uφ+λΩφ=ΩH(x,u)φ+χ,φφH1(Ω).

    In addition, if H satisfies (1.6), problem (1.3) is solvable for a unique constant λ and the corresponding weak solution u is unique up to addition of a constant.

    Even if the result of Theorem 1.1 is quite simple, it seems new to the best of our knowledge, except for the linear case, which was treated in [10] through Fredholm theory. As usual, the existence of the constant λ is proved by considering the singular ergodic limit, as ε0, of solutions uε of

    {div(A(x)uε)+εuε=H(x,uε)+χin Ω,A(x)uεn=0on Ω. (1.7)

    Here the main difference, compared to the classical case ([15]), is that the uniform estimate of εuε, usually given by the maximum principle, is not available because of the more singular x-dependence of the Hamiltonian. In fact, we directly estimate uεL2(Ω) as a first, and then crucial, step; this is done with a similar strategy as suggested in [20] for the Dirichlet problem.

    As a consequence of Theorem 1.1, and of our structure conditions, we eventually give a complete description of the limit of uε, assuming further that H(x,p) is differentiable with respect to p. In that case, the limit of uε can be fully characterized in terms of the additive eigenvalue of the linearized problem: this is the (non homogeneous) linear problem

    {div(A(x)w)+θ=Hp(x,ˉu)wˉuin Ω,A(x)wn=0on Ω, (1.8)

    where ˉu is the unique solution of (1.3) such that Ωˉu=0.

    Theorem 1.2. Assume that H(x,p) satisfies (1.5), (1.6) and H is differentiable with respect to p with Hp(x,p):=H(x,p)p being continuous in p, for a.e. xΩ. Let uε be the unique solution of (1.7) and (λ,ˉu) be the unique solution of (1.3) such that Ωˉu=0. Then we have

    limε0(uελεˉu)=θ

    where the limit is in H1(Ω), and θ is the unique constant for which problem (1.8) is solvable. Moreover, we also have (in L2(Ω))

    uε=ˉu+εw+o(ε)asε0,

    where w is the unique solution of (1.8) with zero average.

    The equivalent of this result for much more general diffusion operators and Hamiltonians, but with a smooth dependence on x, is proved in many recent papers through viscosity solutions' methods, see e.g., [13,17] for second order problems, where this is called the selection problem, since the constant θ selects the limit of uε among all possible solutions of the ergodic problem (1.3). Of course, this is much simpler for elliptic equations rather than for first order (or degenerate elliptic) problems, as treated e.g., in [9] or in the pioneering paper [8].

    The proofs of Theorem 1.1 and Theorem 1.2 are given in the next Section. Later we briefly address some extension of our results to nonlinear divergence form operators (Theorem 2.5) and to more singular x-dependence, including the case of data f(x) in Lebesgue spaces Lm(Ω), m1 (see Theorem 2.9 and Theorem 2.10).

    Let us recall that, in the following, Ω is a bounded connected domain in RN, N3, with Lipschitz boundary, and we denote by n the outward normal unit vector to the boundary Ω. The Sobolev space is denoted by H1(Ω) and its dual by (H1(Ω)). We start with a characterization of all possible weak subsolutions of a Neumann elliptic problem.

    Lemma 2.1. Let wH1(Ω) and λR satisfy

    {div(A(x)w)=B(x)+λinΩ,A(x)wn=0onΩ, (2.1)

    where the matrix A(x) satisfies (1.4) and the function B(x)L2NN+2(Ω) satisfies

    |B(x)|b(x)|w(x)|forsomebLN(Ω). (2.2)

    Then, we have λ=0 and the function w(x) is constant in Ω.

    Proof. We divide the proof in three steps.

    Step 1. We prove that wL(Ω).

    This is standard, but we sketch the argument for the reader's convenience, following [21]. For k>0 we use Gk(w):=(wk)+ as test function in problem (2.1). Then, using the ellipticity of A(x) and condition (2.2), we get

    αΩ|Gk(w)|2dxΩ[b(x)|w|+λ]Gk(w)dx (2.3)

    Let us set Ak:={xΩ:w(x)>k}. Since the integral in the right-hand side is restricted in the set Ak, we deduce, using Hölder inequality,

    αΩ|Gk(w)|2dx(Ak{|w|0}|b|N)1NGk(w)L2(Ω)Gk(w)L2(Ω)+|λ|Gk(w)L2(Ω)|Ak|112.

    This readily implies

    Ω|Gk(w)|2dxc(Ak{|w|0}|b|N)2NGk(w)2L2(Ω)+c|λ|Gk(w)L2(Ω)|Ak|112, (2.4)

    where, here and below, c denotes possibly different numbers only depending on α,N,Ω. By Poincaré-Wirtinger inequality we deduce

    Gk(w)ΩGk(w)2L2(Ω)c(Ak|b|N)2NGk(w)2L2(Ω)+c|λ|Gk(w)L2(Ω)|Ak|112

    which implies

    Gk(w)2L2(Ω)c(Ak|b|N)2NGk(w)2L2(Ω)+c|λ|Gk(w)L2(Ω)|Ak|112+c|ΩGk(w)|2. (2.5)

    We estimate last term as

    |ΩGk(w)|2Gk(w)2L2(Ω)|Ak|222Gk(w)2L2(Ω)(wL1(Ω)k)222.

    Using this estimate in (2.5), we obtain that

    Gk(w)2L2(Ω)[1c(Ak|b|N)2Nc(wL1(Ω)k)222]c|λ|Gk(w)L2(Ω)|Ak|112.

    Let us take k0 sufficiently large such that, for every kk0,

    1c(Ak|b|N)2Nc(wL1(Ω)k)22212.

    Then we have

    Gk(w)L2(Ω)2c|λ||Ak|112kk0.

    Hence

    ΩGk(w)Gk(w)L2(Ω)|Ak|1122c|λ||Ak|222.

    Recall that 222=1+2N, and, for a.e. k, we have ddkΩGk(w)=|Ak|. This means that the function φ(k):=ΩGk(w) is a non increasing function which satisfies φ2c|λ|(φ)1+2N for all kk0. It follows that φ(k1)=0 for some k1>k0. Hence w(x)k1 a.e. in Ω. Repeating the argument for w, we conclude that wL(Ω).

    Step 2. We prove that λ=0.

    To this purpose, we reason by contradiction. Suppose that λ<0. Since wL(Ω), for ε sufficiently small we have λ+εw0. Hence w satisfies

    {div(A(x)w)+εwb(x)|w|,in Ω,A(x)wn=0on Ω.

    This implies (with the same proof as e.g., [20, Proposition 2.1]) that w0. Since w+c is still a solution of (2.1), whatever is the constant cR, we easily get a contradiction. Of course the same argument applies if λ>0. We conclude that λ=0.

    Step 3. We now prove that w is a constant. To this purpose, we recall that the median of a function uH1(Ω) is defined as

    med(u):=sup{kR:meas({u>k})|Ω|2}.

    As for the average of u, a Poincaré type inequality holds for umed(u), see e.g., [22]. Namely there exists a constant C (only depending on N,Ω) such that

    umed(u)L2(Ω)CuL2(Ω)uH1(Ω)

    and therefore, by Sobolev inequality, we have, for a possibly different constant C,

    umed(u)L2(Ω)CuL2(Ω)uH1(Ω). (2.6)

    We now normalize our solution w so that

    med(w)=0.

    This implies that med(ψ(w))=0 for every nondecreasing Lipschitz function ψ:RR such that ψ(0)=0. In particular, we have that med(Gk(w))=0 for all k>0, where now Gk(s)=(|s|k)+sign(s). Defining Ak={x:|w(x)|>k}, we obtain as in (2.4) (with the additional information now that λ=0):

    Ω|Gk(w)|2c(Ak{|w|0}|b|N)2NGk(w)2L2(Ω)

    and so by (2.6) we deduce

    Gk(w)2L2(Ω)c(Ak{|w|0}|b|N)2NGk(w)2L2(Ω).

    This inequality now yields that w=0. In fact, assume by contradiction that M:=sup|w|>0 and take a sequence of k<M, kM. Since k<M, we have Gk(w)2L2(Ω)0 (otherwise this contradicts the definition of M), so the previous inequality implies

    1c(Ak{|w|0}|b|N)2N.

    But, as kM, we have |Ak{|w|0}|0, because w=0 a.e. in the set {|w|=M}. Hence we reach a contradiction. This implies that w=0, so w coincides with its median, and it is a constant.

    Remark 2.2. We notice that this lemma remains true for a nonlinear divergence form operator div(a(x,u)) which satisfies

    a(x,p)pα|p|2forsomeα>0. (2.7)

    We now analyze the limit, as ε0, of the elliptic problem (1.7). For the purposes of Theorem 1.2, it is convenient to state this kind of result in a slightly more general form, where the Hamiltonian H(x,p) may possibly depend on ε as well.

    Lemma 2.3. Assume that A(x) satisfies (1.4), and that Hε(x,p) is a sequence of Carathéodory functions satisfying

    |Hε(x,p)|b(x)|p|+f(x),forsomeb(x)LN(Ω),$f(x)L2NN+2(Ω)$, (2.8)

    for almost every xΩ and for every pRN.

    For χ(H1(Ω)), and ε>0, let uε be solutions of the elliptic problem

    {div(A(x)uε)+εuε=Hε(x,uε)+χinΩ,A(x)uεn=0onΩ, (2.9)

    and assume that there exists a function H(x,p) such that

    Hε(x,p)H(x,p)foreverypRN,andalmosteveryxΩ.

    Then there exist a constant λR and a function uH1(Ω) such that, up to a subsequence,

    εuελanduεΩuεu

    where the limits are in the (strong) topology of H1(Ω). Moreover, (λ,u) solve problem (1.3).

    Proof. We first claim that

    K>0(independent of ε):uεL2(Ω)Kε>0. (2.10)

    We proceed by contradiction and suppose that (2.10) is not true. This implies the existence of a subsequence of uε, not relabeled, such that uεL2(Ω). We set

    wε:=uεΩuεuεL2(Ω).

    Since wε has zero average and wεL2(Ω)=1, we deduce that wε is weakly relatively compact in H1(Ω) and strongly in L2(Ω) by Rellich's theorem. We observe that wε satisfies

    div(A(x)wε)+εwε=1uεL2(Ω)[Hε(x,uε)+χ]ε|Ω|ΩuεuεL2(Ω)in Ω. (2.11)

    Last term is a sequence of real numbers that we estimate, integrating (2.9), as

    εΩuεuεL2(Ω)=1uεL2(Ω)ΩHε(x,uε)+1uεL2(Ω)χ,11uεL2(Ω)Ω[b(x)|uε|+f(x)]+χ(H1(Ω))|Ω|12uεL2(Ω)1uεL2(Ω)[bL2(Ω)uεL2(Ω)+fL1(Ω)+χ(H1(Ω))|Ω|12],

    where we used (2.8) as well. Since bLN(Ω) (N2) and fL1(Ω), and since uεL2(Ω) diverges, the right-hand side is bounded. Hence ε|Ω|ΩuεuεL2(Ω) is a bounded sequence and there exists a real value λR such that, up to subsequences (not relabeled here), we have

    ε|Ω|ΩuεuεL2(Ω)λ. (2.12)

    Now we use once more (2.8) to estimate

    1uεL2(Ω)|Hε(x,uε)|b(x)|wε|+f(x)uεL2(Ω). (2.13)

    Since bLN(Ω) and wε is bounded in H1(Ω), the product b(x)|wε| is bounded in L2NN+2(Ω). In addition, it is also equi-integrable: indeed, for any set E, one has

    E|b(x)wε|2NN+2wε2NN+2L2(Ω)(Eb(x)N)2N+2C(Eb(x)N)2N+2

    and last term (independent of ε) goes to zero as |E|0. Thus, in view of (2.13), the term 1uεL2(Ω)Hε(x,uε) is bounded and equi-integrable in L2NN+2(Ω). Therefore, we have

    {wε is bounded in H1(Ω),div(A(x)wε)+εwε=hε+χuεL2(Ω) (2.14)

    for some hε bounded and equi-integrable in L2NN+2(Ω).

    This implies that wε is actually strongly compact in H1(Ω). We recall the argument for the reader's convenience. The key point is that, if w is a weak limit of wε, then

    Ωhε(wεw)ε00. (2.15)

    In fact, we have

    Ωhε(wεw)=ΩhεTk(wεw)+ΩhεGk(wεw)

    where Tk(s)=max(min(s,k),k) is the truncation function and Gk(s) is the difference sTk(s). By Hölder inequality, since wεw is bounded in L2(Ω) one has, by definition of Gk():

    |ΩhεGk(wεw)|C({|wεw|>k}|hε|2NN+2)N+22N

    where last term vanishes as k, uniformly with respect to ε, since hε is equi-integrable in L2NN+2(Ω). Therefore one has

    Ωhε(wεw)=ΩhεTk(wεw)+rk

    where rkk0 uniformly with respect to ε. But hε is bounded in L2NN+2(Ω) while Tk(wεw)0 strongly in Lp(Ω) for any p< due to Rellich theorem; hence the first integral in the right-hand side converges to zero as ε0. Therefore, by letting ε0 and then k, we deduce that (2.15) holds true. With (2.15) in hands, it is now easy to deduce from (2.14) that

    wεw2L2(Ω)1αΩA(x)(wεw)(wεw)ε00.

    We have obtained so far that there exists a function wH1(Ω) such that, up to subsequences,

    wεwstrongly in H1(Ω).

    In particular, this implies, up to subsequences, that wεw almost everywhere in Ω. Therefore, as a consequence of (2.13), the term 1uεL2(Ω)Hε(x,uε) weakly converges in L2NN+2(Ω) towards some function B(x) which satisfies |B(x)|b(x)|w(x)| a.e. in Ω.

    Finally, due to (2.11), (2.12), (2.13), w satisfies (in weak sense)

    {div(A(x)w)b(x)|w|+λin Ω,A(x)wn=0on Ω. (2.16)

    By Lemma 2.1 we deduce that w is constant, and since Ωw=0 this means w(x)0. But this is a contradiction with the fact that wε strongly converges to w in H1(Ω) and wεL2(Ω)=1 for every ε. The contradiction proves that the assertion (2.10) is true.

    Thanks to (2.10), now we integrate (2.9) and we get

    ε|Ωuε|=|ΩHε(x,uε)+χ,1|Ωb(x)|uε|+f(x)+|χ,1|C,

    due to the growth condition (2.8). Therefore, εΩuε is a bounded sequence, and there exists a real value λR such that, up to subsequences,

    εΩuελ.

    We now set ^uε:uεΩuε. Then ^uε is bounded in H1(Ω) and solves

    div(A(x)^uε)+ε^uε=Hε(x,^uε)+χεΩuεin Ω. (2.17)

    With the same arguments used before for wε, we can show that ^uε is strongly compact in H1(Ω), so there exists a function uH1(Ω) such that ^uεu in H1(Ω). We may also assume that ^uεu almost everywhere in Ω, and therefore almost uniformly as well; since Hε(x,^uε) is equi-integrable and Hε(x,p)H(x,p), we can deduce that Hε(x,^uε) converges to H(x,u) (which belongs to L2NN+2(Ω) due to (2.8)). Finally, passing to the limit in (2.17), we obtain that (u,λ) is a solution to problem (1.3).

    The proof of Theorem 1.1 immediately follows from the above two results.

    Proof of Theorem 1.1. For ε>0, let uε be the solution of the elliptic problem (1.7). Applying Lemma 2.3, we have that, up to subsequences, εuελ and uεΩuεu, where (λ,u) gives a solution to problem (1.3).

    Now we assume further the condition (1.6) and we prove uniqueness of λ and u (up to a constant). This is a straightforward consequence of Lemma 2.1. Indeed, let (u1,λ1) and (u2,λ2) be solutions to problem (1.3). Then w:=u1u2 is a solution to

    {div(A(x)w)=H(x,u1)H(x,u2)+λ1λ2in Ω,A(x)wn=0on Ω,

    where we have, due to (1.6),

    |H(x,u1)H(x,u2)|b(x)|w|.

    From Lemma 2.1 we deduce that λ1=λ2 and that u1u2 is a constant.

    Given the solution uε of (1.7), we now investigate the behavior of uελεˉu where (λ,ˉu) is a solution of the ergodic problem (1.3). In order to fix a reference solution, we normalize ˉu so that Ωˉu=0. So (λ,ˉu) is uniquely defined from Theorem 1.1, provided (1.6) holds. The proof of Theorem 1.2 now follows as a Corollary of the previous results.

    Proof of Theorem 1.2. Let us define

    vε:=1ε(uελεˉu).

    One can check that vε solves

    {div(A(x)vε)+εvε=1ε[H(x,uε)H(x,ˉu)]ˉuin Ω,A(x)vεn=0on Ω. (2.18)

    We notice that, thanks to (1.6), the function

    ˜Hε(x,p):=1ε[H(x,εp+ˉu)H(x,ˉu)]

    satisfies

    |˜Hε(x,p)|b(x)|p|,|˜Hε(x,p)˜Hε(x,q)|b(x)|pq|

    and, from the differentiability of H, we have

    ˜Hε(x,p)Hp(x,ˉu)pfor every pRN and a.e. xΩ.

    Therefore Lemma 2.3 applies and we deduce that there exists a constant θR and a function wH1(Ω) such that

    εvεθ,vεΩvεw

    where the limits are in the strong topology of H1(Ω) and (θ,w) is the unique couple which satisfies the linear elliptic problem (1.8) with the normalized condition Ωw=0. We stress that the uniqueness of the limit couple implies that the whole sequence vε converges.

    Coming back from vε to uε, this means that

    limε0(uελεˉu)=θ

    where the limit is meant in H1(Ω), and in addition

    1ε(uεΩuεˉu)w.

    This latter convergence yields the first order expansion for the gradient in L2(Ω):

    uε=ˉu+εw+o(ε)as ε0.

    Remark 2.4. The case N=2 can be dealt with in exactly the same way as before, except that the threshold summability of the drift term should be adapted to the Sobolev embedding of dimension two. Since for N=2 the space H1(Ω) is embedded in Lp(Ω) for every p<, but not in L(Ω), one needs here b(x)Lq(Ω) for some q>2 in order that the product b(x)|u| belongs to the dual space (H1(Ω)). Therefore, conditions (1.5) and (1.6) should be changed into

    |H(x,p)|b(x)|p|+f(x),for some b(x)Lq(Ω)q>2, and f(x)Lm(Ω)m>1, (2.19)

    and respectively,

    |H(x,p)H(x,q)|b(x)|pq|,b(x)Lq(Ω)q>2, (2.20)

    for almost every xΩ and for every p,qRN.

    Replacing assumptions (1.5) and (1.6) with, respectively, (2.19) and (2.20), the results stated in Theorem 1.1 and Theorem 1.2 are true for the dimension N=2, and the proof remains the same up to the obvious modifications in the Lebesgue spaces which are involved.

    A similar remark holds true for the case of dimension N=1; in that case it is enough to assume b(x)L2(Ω) and f(x)L1(Ω).

    In this subsection we give a short extension of the result of Theorem 1.1 to the case of nonlinear operators. To this purpose, we introduce a function a(x,p):Ω×RNRN which is assumed to be measurable with respect to x, for every pRN, and continuous with respect to p, for a.e. xΩ. We assume that a(x,p) satisfies the following monotonicity and growth conditions:

    (a(x,p)a(x,q))(pq)α|pq|2,for some αRα>0, (2.21)

    and

    |a(x,p)|β(|p|+k(x))for some βRk(x)L2(Ω), (2.22)

    for almost every xΩ and every p,qRN.

    Notice that (2.21)–(2.22) imply also the coercivity condition

    a(x,p)pα|p|2˜k(x),where αRα>0˜k(x)L1(Ω). (2.23)

    We also assume that a(x,p) satisfies the following asymptotic condition for |p|:

    for a.e. xΩ, every pRNlimta(x,tp)t. (2.24)

    Of course, the case a(x,p)=A(x)p+K(x), for a matrix A(x) satisfying (1.4) and a vector field KL2(Ω), is the simplest example where conditions (2.21)–(2.24) are satisfied.

    Thanks to (2.21)–(2.24), we can extend Theorem 1.1 to a nonlinear setting. Notice that additional terms in the equation, belonging to (H1(Ω)), can be included here in the vector field a(x,p) or in the local function H(x,p).

    Theorem 2.5. Assume that a(x,p) satisfies (2.21)–(2.24), and that H(x,p) satisfies (1.5) (or it satisfies (2.19) if N=2). Then there exist a constant λR and a function uH1(Ω) which solve, in the weak sense, the elliptic problem

    {div(a(x,u))+λ=H(x,u)inΩ,a(x,u)n=0onΩ. (2.25)

    In addition, if (1.6) holds true (respectively, (2.20) if N=2), then λ is unique and u is unique up to addition of a constant.

    Proof. For ε>0, we consider the approximating problem

    {div(a(x,uε))+εuε=H(x,uε)in Ω,a(x,uε)n=0on Ω. (2.26)

    Then we aim at showing that the a priori estimate (2.10) holds true and we proceed, as in Lemma 2.3, by contradiction. This allows us to build a sequence wε such that wεL2(Ω)=1, wε has zero average and satisfies

    div(a(x,uεL2(Ω)wε)uεL2(Ω))+εwε=H(x,uε)uεL2(Ω)ε|Ω|ΩuεuεL2(Ω)in Ω, (2.27)

    where uεL2(Ω). As in Lemma 2.3, last term is a relatively compact sequence of real numbers, and we observe that the right-hand side is bounded and equi-integrable in L2NN+2(Ω), so that wε satisfies

    {wε is bounded in H1(Ω),div(ˆaε(x,wε))+εwε=hε (2.28)

    for some hε bounded and equi-integrable in L2NN+2(Ω), where

    ˆaε(x,p):=a(x,uεL2(Ω)p)uεL2(Ω).

    Now, let w be a weak limit of wε in H1(Ω), and a strong limit in L2(Ω). Notice that we have that

    Ωˆaε(x,w)(wεw)0

    because ˆaε(x,w) is strongly convergent in L2(Ω) due to (2.24) and (2.22), by Lebesgue's theorem. Therefore, using (2.28) and (2.15) as well, we deduce that

    limε0Ω(ˆaε(x,wε)ˆaε(x,w))(wεw)=0. (2.29)

    Notice that ˆaε satisfies (2.21) for the same α>0, and for every ε>0. Then (2.29) implies that wεw strongly in H1(Ω) (and, up to a subsequence, wεw almost everywhere in Ω). Now, for every φH1(Ω), we set

    ˆa(x,φ(x)):=limε0ˆaε(x,φ(x)) (2.30)

    which exists after (2.24) and belongs to L2(Ω) due to (2.22). It is a consequence of (2.29) and a standard monotonicity argument (sometimes known as Minty's argument) that ˆaε(x,wε(x)) converges to ˆa(x,w(x)) weakly in L2(Ω). Finally, passing to the limit in (2.27), we obtain that w is a weak solution of

    {div(ˆa(x,w(x)))+λ=B(x)in Ω,ˆa(x,w(x))n=0on Ω

    for some λR and some B(x)L2NN+2(Ω) satisfying condition (2.2). Since ˆa(x,p) satisfies (2.7) (because ˆa(x,0)=0 and (2.21) holds), we deduce from Lemma 2.1, Remark 2.2, that λ=0 and that w is constant (hence w=0 because it has zero average). We get a contradiction with wL2(Ω)=1. This proves that the a priori estimate (2.10) holds true, that is uεL2(Ω) is uniformly bounded. Then the compactness of uε in H1(Ω) follows in a similar way as before, using (2.21), and we conclude as in Theorem 1.1 the existence of λR, uH1(Ω) which solve problem (2.25).

    By using (1.6) and (2.21), the uniqueness of λ and the uniqueness of u, up to a constant, are proved exactly as in Theorem 1.1. Even if the problem is nonlinear, and Lemma 2.1 cannot be literally applied, the arguments are exactly the same as used in Lemma 2.1, now applied to w:=u1u2; the nonlinearity of the operator is readily handled with assumption (2.21).

    Remark 2.6. Condition (2.24) is assumed here in order that we can follow the same approach used before in the proof of Theorem 1.1. We believe this condition to be unnecessary for a similar result to hold for general nonlinear operators, however removing this condition would need a substantial change in the method of proof (e.g. using symmetrization methods), which is beyond the scope of this note.

    Remark 2.7. The result remains true for the case of dimension N=1, up to requiring bL2(Ω) and fL1(Ω) in assumptions (1.5) and (1.6).

    The same approach as before can be used in case of more singular dependence with respect to x. We have in mind here that the assumption (1.5) is replaced by

    |H(x,p)|b(x)|p|+f(x),for some b(x)LN(Ω)f(x)Lm(Ω), (2.31)

    for almost every xΩ and for every pRN, where m1, so that data can belong to Lebesgue spaces of any order.

    Of course, if m2NN+2, assumption (2.31) implies (1.5), so there is nothing new to be proved. By contrast, if m<2NN+2, we cannot expect the solutions to belong to H1(Ω) anymore; however it is still possible to obtain similar results in a setting of generalized solutions. We recall below the notion of renormalized solutions. The truncation function is denoted, as before, by Tk(s):=max(min(s,k),k). In what follows we suppose that N>1.

    Definition 2.8. A function u, belonging to W1,q(Ω) for every q<NN1, is a renormalized solution of problem (2.25) if:

    (i) Tk(u)H1(Ω) for every k>0, and H(x,u)L1(Ω).

    (ii) For every function h:RR which is C1 with compact support, it holds

    Ωa(x,u)(h(u)φ)+λΩh(u)φ=ΩH(x,u)h(u)φφH1(Ω)L(Ω). (2.32)

    (iii)

    limn1n{n<|u|<2n}|u|2=0. (2.33)

    We recall that this notion of solution, introduced by P. L. Lions and F. Murat (in a joint paper unpublished, whose content can be found in [16,18]) is nowadays currently used as a formulation of elliptic problems with L1-data, or even Radon measures. For similar problems with first order terms having linear growth, we refer e.g., to [5].

    It is to be noted that, in the renormalized formulation, the truncations of the solution belong to the energy space H1(Ω), so that (2.32) makes sense for any h with compact support. Condition (2.33), in turn, implies that renormalized solutions are also distributional solutions. To this purpose one can take h=S(rn) where S(t) is a piecewise linear function, supported in [2,2], such that S(t)1 for t[1,1]; then S(rn)1 as n and the distributional formulation (with test functions φC1(¯Ω)) is recovered thanks to (2.33).

    Up to replacing the (H1(Ω)) formulation with the renormalized setting, Theorem 2.5 admits a natural extension which is the following one. We recall (see [7]) that elliptic equations (including possibly nonlinear divergence form operators with discontinuous coefficients) with source terms in Lm(Ω), m<2NN+2, admit solutions which belong to W1,m(Ω), where m is the Sobolev exponent (Nm)/(Nm).

    Theorem 2.9. Let N3. Assume that a(x,p) satisfies (2.21)–(2.24) and that H(x,p) satisfies (2.31) for some 1<m<2NN+2. Then there exists a constant λR and a function uW1,m(Ω) which is a renormalized solution of (2.25). In addition, if (1.6) holds true, λ is unique and u is unique up to addition of a constant.

    Proof. We only sketch the main steps, and the main differences with the arguments of Theorem 2.5. We start with the solutions uε of the approximating problem

    {div(a(x,uε))+εuε=Hε(x,uε)in Ω,a(x,uε)n=0on Ω, (2.34)

    where Hε(x,p)=T1ε(H(x,p)). Here the truncation of H is only needed if one wants to work, at fixed ε, with the more comfortable setting of finite energy solutions. Then we claim that there exists K>0 independent of ε, such that

    uεLm(Ω)Kε>0. (2.35)

    The proof of (2.35) is done, as before, by contradiction. In this case we build a subsequence, not relabeled, such that uεLm(Ω) and we define

    wε=uεΩuεuεLm(Ω)

    which solves

    div(ˆa(x,wε))+εwε=Hε(x,uε)uεLm(Ω)ε|Ω|ΩuεuεLm(Ω)in Ω, (2.36)

    where ˆaε(x,p) is defined now as

    ˆaε(x,p):=a(x,uεLm(Ω)p)uεLm(Ω).

    Compared to Theorem 2.5, we now have that the right-hand side is bounded and equi-integrable in Lm(Ω), hence wε satisfies

    {wε is bounded in W1,m(Ω),div(ˆaε(x,wε))+εwε=hε (2.37)

    for some hε which is bounded and equi-integrable in Lm(Ω). One can prove now that wε is actually strongly compact in W1,m(Ω). The argument needs a modification of what is done in Lemma 2.3: first of all, by only using that hε is weakly converging in L1(Ω), one can prove (see e.g., [18,14]) that

    limε0Ω(ˆaε(x,Tk(wε))ˆaε(x,Tk(w)))(Tk(wε)Tk(w))=0

    which yields, thanks to (2.21),

    Tk(wε)Tk(w)in H1(Ω), for every k>0. (2.38)

    To estimate Gk(wε)=wεTk(wε), we take

    [(σ+|Gk(wε)|)γσγ]sign(wε)

    as test function in (2.37), for σ,γ>0. We get

    γΩˆaε(x,wε)Gk(wε)(σ+|Gk(wε)|)γ1Ωhε[(σ+|Gk(wε)|)γσγ]sign(wε)

    which implies, using (2.23) and the definition of ˆa,

    γαΩ|Gk(wε)|2(σ+|Gk(wε)|)γ1Ωhε|(σ+|Gk(wε)|)γσγ|+γuε2Lm(Ω)Ω˜k(x)(σ+|Gk(wε)|)γ1. (2.39)

    We choose γ=N(m1)N2m, so that |wε|γ=|wε|mm is bounded in Lm(Ω) due to the bound of wε in W1,m(Ω) and Sobolev embedding. Moreover, we have γ<1 since m<2NN+2. Taking for instance σ=1, the previous inequality implies

    γαΩ|Gk(wε)|2(1+|Gk(wε)|)γ12{|wε|>k}|hε|(1+|wε|)γ+γuε2Lm(Ω)˜kL1(Ω)C({|wε|>k}|hε|m)1m+Cuε2Lm(Ω).

    Due to the equi-integrability of hε in Lm(Ω) we deduce

    Ω|Gk(wε)|2(1+|Gk(wε)|)γ1δk+Cuε2Lm(Ω) (2.40)

    where δk denotes a quantity which vanishes as k uniformly with respect to ε. From Hölder inequality, since m<2 we have

    Ω|Gk(wε)|m(Ω|Gk(wε)|2(1+|Gk(wε)|)γ1)m2(Ω(1+|Gk(wε)|)(1γ)m2m)1m2.

    The precise value of γ yields (1γ)m2m=m, so last term is bounded and from (2.40) we deduce

    Ω|Gk(wε)|mCδm2k+CuεmLm(Ω).

    Putting together this information with (2.38), it follows that wεw strongly in W1,m(Ω). In particular, up to a subsequence, wεw almost everywhere in Ω. This implies that Hε(x,uε)uεLm(Ω) weakly converges to some function B(x) in Lm(Ω), and, because of (2.31) and the a.e. convergence of wε, it follows that B satisfies (2.2). If the vector filed ˆa(x,p) is defined as in (2.30), then by passing to the limit we obtain a function w such that

    {wW1,m(Ω),div(ˆa(x,w))+λ=B(x)with B(x)Lm(Ω) satisfying (2.2).

    Now we get a contradiction by showing that w=0. This needs a slight refinement of the argument of Lemma 2.1; indeed, w does not belong a priori to H1(Ω). However, from (2.39) one obtains, letting first ε0 and then σ0, the inequality

    γαΩ|Gk(w)|2|Gk(w)|γ1Ω[b(x)|w|+|λ|]|Gk(w)|γ. (2.41)

    This replaces now the starting inequality (2.3) of Lemma 2.1; using that

    Ω|Gk(w)|m(Ω|Gk(w)|2|Gk(w)|γ1)m2(Ω|Gk(w)|(1γ)m2m)1m2C(Ω[b(x)|Gk(w)|+|λ|]|Gk(w)|γ)m2(Ω|Gk(w)|(1γ)m2m)1m2

    and the equalities γm=m and (1γ)m2m=m, one can prove with similar steps as in Lemma 2.1 that wL(Ω). Once w is proved to be bounded, then it has also finite energy and the conclusion of Lemma 2.1 applies. This completes the contradiction argument and concludes the proof of the a priori estimate (2.35). Once uε is proved to be bounded in W1,m(Ω), we repeat the same argument to deduce that uεΩuε is relatively compact and converges, up to subsequences, to some uW1,m(Ω). Due to (2.31), this implies that H(x,uε)H(x,u) in Lm(Ω). By well-known results (see e.g., [5,18]), the limit function u is a weak and a renormalized solution of the limiting problem (2.25).

    Let us now sketch the uniqueness part. Assume that λ1, u1 and λ2,u2 are (renormalized) solutions of the problem. Then it is possible to prove, using renormalization arguments (e.g., the arguments used for the obtention of inequality (3.14) in [19, Theorem 3.1]), that u1,u2 satisfy

    γΩ(a(x,u1)a(x,u2))Gk(u1u2)(σ+|Gk(w)|)γ1Ω|H(x,u1)H(x,u2)||(σ+|Gk(w)|)γσγ|+|λ1λ2|Ω|(σ+|Gk(w)|)γσγ|

    where w=u1u2 and γ=N(m1)N2m as before. The reader may check that all terms here are well defined because of the precise value of γ and since u1,u2W1,m(Ω). Using (2.21) and (1.6) it follows that

    γαΩ|Gk(w)|2(σ+|Gk(w)|)γ1Ωb(x)|w||(σ+|Gk(w)|)γσγ|+|λ1λ2|Ω|(σ+|Gk(w)|)γσγ|.

    By letting σ0 one gets inequality (2.41) with λ=λ1λ2. This inequality allows us to show that wL(Ω) as we said before. Now, let us assume for instance that λ1>λ2; since u1u2L(Ω), then ε(u1u2)<λ1λ2 for ε sufficiently small. Therefore, we get that u1, u2 are, respectively, a renormalized sub and super solution of the same equation

    div(a(x,u))+εu=H(x,u)+εu2λ2.

    This implies (e.g., as in [19, Theorem 3.1], or in [6]) that u1u2, which yields a contradiction, because u1,u2 are solutions up to addition of any constant. We notice here that, if u is a renormalized solution of (2.25), then it can be easily proved that u+c is also a renormalized solution, whenever cR. The contradiction obtained proves that λ1=λ2. Then, we replace u1 with u1med(u1u2); now again from inequality (2.41) (obtained before for w=u1u2med(u1u2) and λ=λ1λ2=0) we can deduce that w=0, i.e., u1u2=med(u1u2). Hence the two solutions differ by a constant.

    Finally, we conclude by observing that a similar result can be proved in the limiting case m=1, up to requiring that b(x)Lq(Ω) for some q>N. More precisely, we have the following result, whose proof can be done with similar arguments as indicated above.

    Theorem 2.10. Let N2. Assume that a(x,p) satisfies (2.21)–(2.24) and that H(x,p) satisfies (2.31) with m=1 and b(x)Lq(Ω) for some q>N. Then there exists a constant λR and a renormalized solution u of (2.25). In addition, if (1.6) holds true with b(x)Lq(Ω) for some q>N, then λ is unique and u is unique up to addition of a constant.

    Alessio Porretta aknowledges the support of Fondation Sciences Matématiques de Paris and Istituto Nazionale di Alta Matematica (GNAMPA).

    The authors declare no conflict of interest.



    [1] Arisawa M, Lions PL (1998) On ergodic stochastic control. Commun Part Diff Eq 23: 2187-2217.
    [2] Bensoussan A, Frehse J (1987) On Bellman equation of ergodic type with quadratic growth Hamiltonian, In: Contributions to Modern Calculus of Variations, Wiley, 13-26.
    [3] Bardi M, Capuzzo-Dolcetta I (1997) Optimal Control and Viscosity Solutions of Hamilton-JacobiBellman Equations, Boston: Birkhäuser.
    [4] Bensoussan A, Lions JL, Papanicolaou G (1978) Asymptotic Analysis for Periodic Structures, Amsterdam: North-Holland.
    [5] Betta F, Mercaldo A, Murat F, et al. (2003) Existence of renormalized solutions to nonlinear elliptic equations with a lower-order term and right-hand side a measure. J Math Pure Appl 82: 90-124.
    [6] Betta F, Mercaldo A, Murat F, et al. (2005) Uniqueness results for nonlinear elliptic equations with a lower order term. Nonlinear Anal 63: 153-170.
    [7] Boccardo L, Gallouët T (1992) Nonlinear elliptic equations with right hand side measures. Commun Part Diff Eq 17: 641-655.
    [8] Capuzzo Dolcetta I, Menaldi JL (1988) Asymptotic behavior of the first order obstacle problem. J Differ Equations 75: 303-328.
    [9] Davini A, Fathi A, Iturriaga R, et al. (2016) Convergence of the solutions of the discounted equation. Invent Math 206: 29-55.
    [10] Droniou J, Vazquez JL (2009) Noncoercive convection-diffusion elliptic problems with Neumann boundary conditions. Calc Var 34: 413-434.
    [11] Gilbarg D, Trudinger N (1983) Partial Differential Equations of Second Order, 2 Eds., BerlinNew-York: Springer-Verlag.
    [12] Gimbert F (1985) Problèmes de Neumann quasilineaires. J Funct Anal 68: 65-72.
    [13] Ishii H, Mitake H, Tran HV (2017) The vanishing discount problem and viscosity Mather measures. Part 1: the problem on a torus. J Math Pure Appl 108: 125-149.
    [14] Leone C, Porretta A (1998) Entropy solutions for nonlinear elliptic equations in L1. Nonlinear Anal Theor 32: 325-334.
    [15] Lions PL (1985) Quelques remarques sur les problèmes elliptiques quasilinéaires du second ordre. J Anal Math 45: 234-254.
    [16] Lions PL (1996) Mathematical Topics in Fluid Mechanics. Vol. 1. Incompressible Models, New York: The Clarendon Press.
    [17] Mitake H, Tran HV (2017) Selection problems for a discounted degenerate viscous HamiltonJacobi equation. Adv Math 306: 684-703.
    [18] Murat F, Soluciones renormalizadas de EDP elipticas no lineales, Preprint 93023 of Laboratoire d'Analyse Numérique, Université Paris VI, 1993. Available from: http://archive.schools.cimpa.info/anciensite/NotesCours/PDF/2009/Alexandrie_Murat_2.pdf.
    [19] Porretta A (2008) On the comparison principle for p-Laplace type operators with first order terms, In: On the Notion of Solutions to Nonlinear Elliptic Problems: Results and Developments, Quaderni di Matematica 23, Dept. Math. Seconda Univ. Napoli, Caserta, 459-497.
    [20] Porretta A, Elliptic equations with first-order terms, notes of the CIMPA School, Alexandria, 2009. Available from: http://archive.schools.cimpa.info/anciensite/NotesCours/PDF/2009/Alexandrie_Porretta.pdf.
    [21] Stampacchia G (1965) Le problème de Dirichlet pour les equations élliptiques du second ordre à coefficients discontinus, Ann I Fourier 15: 189-258.
    [22] Ziemer WP (1989) Weakly Differentiable Functions, New York: Springer-Verlag.
  • This article has been cited by:

    1. Son N.T. Tu, Vanishing discount problem and the additive eigenvalues on changing domains, 2022, 317, 00220396, 32, 10.1016/j.jde.2022.01.055
  • Reader Comments
  • © 2021 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(4600) PDF downloads(617) Cited by(1)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog