Loading [MathJax]/jax/output/SVG/jax.js

The selection problem for some first-order stationary Mean-field games

  • Received: 01 August 2019 Revised: 01 June 2020 Published: 26 August 2020
  • 35A01, 49L25, 91A07

  • Here, we study the existence and the convergence of solutions for the vanishing discount MFG problem with a quadratic Hamiltonian. We give conditions under which the discounted problem has a unique classical solution and prove convergence of the vanishing-discount limit to a unique solution up to constants. Then, we establish refined asymptotics for the limit. When those conditions do not hold, the limit problem may not have a unique solution and its solutions may not be smooth, as we illustrate in an elementary example. Finally, we investigate the stability of regular weak solutions and address the selection problem. Using ideas from Aubry-Mather theory, we establish a selection criterion for the limit.

    Citation: Diogo A. Gomes, Hiroyoshi Mitake, Kengo Terai. The selection problem for some first-order stationary Mean-field games[J]. Networks and Heterogeneous Media, 2020, 15(4): 681-710. doi: 10.3934/nhm.2020019

    Related Papers:

    [1] Diogo A. Gomes, Hiroyoshi Mitake, Kengo Terai . The selection problem for some first-order stationary Mean-field games. Networks and Heterogeneous Media, 2020, 15(4): 681-710. doi: 10.3934/nhm.2020019
    [2] Martino Bardi . Explicit solutions of some linear-quadratic mean field games. Networks and Heterogeneous Media, 2012, 7(2): 243-261. doi: 10.3934/nhm.2012.7.243
    [3] Olivier Guéant . New numerical methods for mean field games with quadratic costs. Networks and Heterogeneous Media, 2012, 7(2): 315-336. doi: 10.3934/nhm.2012.7.315
    [4] Filippo Santambrogio . A modest proposal for MFG with density constraints. Networks and Heterogeneous Media, 2012, 7(2): 337-347. doi: 10.3934/nhm.2012.7.337
    [5] Fabio Camilli, Francisco Silva . A semi-discrete approximation for a first order mean field game problem. Networks and Heterogeneous Media, 2012, 7(2): 263-277. doi: 10.3934/nhm.2012.7.263
    [6] Pierre Cardaliaguet, Jean-Michel Lasry, Pierre-Louis Lions, Alessio Porretta . Long time average of mean field games. Networks and Heterogeneous Media, 2012, 7(2): 279-301. doi: 10.3934/nhm.2012.7.279
    [7] Diogo A. Gomes, Gabriel E. Pires, Héctor Sánchez-Morgado . A-priori estimates for stationary mean-field games. Networks and Heterogeneous Media, 2012, 7(2): 303-314. doi: 10.3934/nhm.2012.7.303
    [8] Fabio Camilli, Italo Capuzzo Dolcetta, Maurizio Falcone . Preface. Networks and Heterogeneous Media, 2012, 7(2): i-ii. doi: 10.3934/nhm.2012.7.2i
    [9] Yves Achdou, Manh-Khang Dao, Olivier Ley, Nicoletta Tchou . A class of infinite horizon mean field games on networks. Networks and Heterogeneous Media, 2019, 14(3): 537-566. doi: 10.3934/nhm.2019021
    [10] Yves Achdou, Victor Perez . Iterative strategies for solving linearized discrete mean field games systems. Networks and Heterogeneous Media, 2012, 7(2): 197-217. doi: 10.3934/nhm.2012.7.197
  • Here, we study the existence and the convergence of solutions for the vanishing discount MFG problem with a quadratic Hamiltonian. We give conditions under which the discounted problem has a unique classical solution and prove convergence of the vanishing-discount limit to a unique solution up to constants. Then, we establish refined asymptotics for the limit. When those conditions do not hold, the limit problem may not have a unique solution and its solutions may not be smooth, as we illustrate in an elementary example. Finally, we investigate the stability of regular weak solutions and address the selection problem. Using ideas from Aubry-Mather theory, we establish a selection criterion for the limit.



    Mean-field games (MFG) model systems with many rational noncooperative players, describe the player's optimal strategies and determine the statistical properties of their distribution. These games are often determined by a system of a Hamilton-Jacobi equation coupled with a transport or Fokker-Planck equation. In the study of stationary Hamilton-Jacobi equations, a standard method to obtain a solution is to consider the vanishing discount problem. This was the strategy used originally in [31] in the study of homogenization problems. For second-order MFG, the existence of a solution for the discounted problem was shown, for example, in [20] and [8] and for first-order MFG in [16] in the sense of weak solutions and in [3] using variational methods. In the second-order case, the vanishing discount limit was studied in [4]. In the first-order case, the theory is not as much developed and the vanishing discount limit has not been examined previously. Here, our goal is to study the limit behavior as ϵ0 of the following discounted first-order stationary mean-field game.

    Problem 1. Let Td be the d-dimensional flat torus identified with [0,1]d. Let V:TdR, VC1,α(Td), g:[0,)R{}, {gC1,α((0,+)), with g strictly increasing, and fix a discount rate, ϵ>0. Find uϵ,mϵ:TdR with mϵ(x)0 such that

    {ϵuϵ+12|Duϵ|2+V(x)=g(mϵ)in Td,ϵmϵdiv(mϵDuϵ)=ϵin Td. (1.1)

    We say that (uϵ,mϵ) is a classical solution of the preceding problem if uϵC2,α(Td) and mϵC1,α(Td) with mϵ0. As we show in Proposition 3.1, mϵ cannot vanish, hence, mϵ>0. As in the case of Hamilton-Jacobi equations, we expect that, as ϵ0, the solutions of (1.1) converge, maybe through subsequences after adding a suitable constant to uϵ, to a solution of the following first-order MFG.

    Problem 2. With g, V as in Problem 1, find u,m:TdR with m0 and ¯HR such that

    {12|Du|2+V(x)=g(m)+ˉHin Td,div(mDu)=0in Td,m(x)0,  Tdmdx=1. (1.2)

    Because (1.2) is invariant under addition of constants to u, we can prescribe the additional normalization condition

    Tdudx=0.

    According to [16] (also see [17]), Problem 2 admits weak solutions under suitable polynomial growth conditions of g, see Corollary 6.3 in [16]. Here, in Section 7, under a different set of hypothesis and using a limiting argument, we establish the existence of solutions for Problem 2. A natural question in the analysis of the limit ϵ0 is the selection problem; that is, whether the sequence (uϵ,mϵ) converges (not just whether a subsequence converges) and if so, what is the limit among all possible solutions of (1.2). This matter is our main focus here.

    For Hamilton-Jacobi equations, the discounted problem corresponds to the following control problem. Let x(t)Rd be the state of an agent at the time t. This agent can change its state by choosing a control vL([0,),Rd). Thus, its trajectory, x(t), is determined by ˙x(t)=v(t), with initial condition x(t)=xTd. The agent selects the control to minimize the cost functional

    J(x;v)=0eϵtL(x(t),˙x(t))dt,

    for a given Lagrangian, L:Td×RdR. The value function, uϵ, is given by

    uϵ(x)=infvJ(x;v),

    where the infimum is taken over vL([0,+),Rd).

    The Hamiltonian, H:Td×RdR, corresponding to this control problem is the Legendre transform of L; that is,

    H(x,p)=supvRdpvL(x,v).

    Under standard coercivity and convexity assumptions on L, uϵ is the unique viscosity solution of the discounted Hamilton-Jacobi equation,

    ϵuϵ+H(x,Duϵ)=0in Td.

    For coercive Hamiltonians, the results in [31] give that ϵuϵ is uniformly bounded and that uϵ is equi-Lipschitz for ϵ>0. Thus, uϵminTduϵ uniformly converges to a function, u, along subsequences, as ϵ0. Moreover, ϵuϵ converges to a constant ˜H. By stability of viscosity solutions, (u,˜H) solves the ergodic Hamilton-Jacobi equation

    H(x,Du)=˜Hin Td, (1.3)

    where the unknowns are u:TdR and ˜HR. However, the solution of (1.3) may not be unique. Hence, the solution constructed above could depend on the particular subsequence used to extract the limit. The study of the selection problem was started in [19] using the discounted Mather measures introduced in [2]. The main convergence result was established in [5]. Subsequently, several authors investigated and extended those ideas in [1], [26], [27], and [34]. Recently, the case of non-convex Hamiltonians was addressed in [21].

    In MFGs, we consider a large population of agents where each agent seeks to optimize an objective function. Here, however, the running cost depends on statistical information about the players, encoded in a probability density, m:Td×[0,)R. In the model discussed here, the Lagrangian is ˆL(x,p)=12|p|2V(x)+g(m) and each agent seeks to minimize the functional

    ˆJ(x)=0eϵt[12|˙x(t)|2V(x(t))+g(m(x(t),t))]dt.

    Now, we suppose that the value function, uϵ:=infv^J, is smooth. Then, uϵ solves the first equation in (1.1) and the optimal control is given by v(t)=Duϵ(x(t)). Because the players are rational, they use this optimal control. Here, ϵ represents the rate at which players quit the game, which occurs at independent and memoryless times. Furthermore, new players join the game randomly at a rate ϵ, as can be seen by looking at the right-hand side of the second equation in (1.1). Then, in the stationary configuration, the density, m, is determined by the second equation in (1.1). Without an inflow of players, the only non-negative solution is trivial, m=0.

    The theory for second-order stationary MFG is now well developed and in many cases the existence of smooth solutions can be established, see for example [25], [24], [35], or [8]. For logarithmic nonlinearities, the existence of smooth solutions was proven in [9]. However, this is a special case; as shown in Section 2, for first-order MFG, the existence of smooth solutions may not hold (see also a detailed discussion in [23] and [22]). Thus, in general, we need to consider weak solutions, see [7] or [16] for an approach using monotone operators and [3] for a variational approach.

    One of the difficulties of first-order stationary MFG is the lack of regularizing terms in both the Hamilton-Jacobi equation and in the transport equation. Nonetheless, the MFG system behaves somewhat like an elliptic equation. Here, we explore this effect and obtain conditions under which Problem 1 has classical solutions. These conditions are given in the following two assumptions.

    Assumption 1. g and V satisfy that g1(g(1)oscxTdV(x))>0.

    Assumption 2. There exist constants C1>0, C2>0 and βR such that for all z>0,

    g(z)C1zβ,
    g(z)C2+12zg(z).

    Assumption 3. limz+g(z)=+.

    For β1, Assumption 2 imples the preceding assumption. This is not the case for β<1.

    An example that satisfies the preceding assumptions is the following:

    g(m)=mα(α>0),V(x)=csin(2πx)(0<c<1/2),

    where d=1 and V is extended by periodicity to R. The preceding two assumptions are used to obtain lower bounds on the density and can be interpreted as follows. Because g is increasing agents want to avoid crowded areas and prefer areas with low density. However, if the oscillation of the potential is large, the trade-off between a low-density area with high potential and a high-density area with low potential may not pay-off. Hence, the control of the oscillation of V given in Assumption 1 implies that no point is totally avoided by the agents.

    As we mentioned previously, the two preceding assumptions imply the existence of a classical solution for Problem 1 as stated in the following theorem.

    Theorem 1.1. Suppose that Assumptions 1-3 hold. Then, for each ϵ>0, Problem 1 has a unique classical solution (uϵ,mϵ) with mϵ>0.

    The proof of this theorem is given in Section 5 using a continuation method combined with the a priori estimates in Section 3 and the DeGiorgi-Nash-Moser argument outlined in Section 4. As a corollary of the preceding theorem, we obtain our first convergence result.

    Corollary 1.2. Suppose that Assumptions 1-3 hold. Then, Problem 2 has a unique classical solution (u,m,ˉH), with m>0 and Tdudx=0. Furthermore, let (uϵ,mϵ) solve Problem 1. Then

    uϵTduϵdxuin C2,α(Td),mϵmin C1,α(Td),ϵuϵ¯Huniformly.

    The proof of this corollary is given at the end of Section 5.

    For second-order MFGs, the vanishing discount problem for mean-field games was addressed in [4]. Inspired by the approach there, we consider the following formal asymptotic expansion

    uϵˉH/ϵu+λ+ϵv,mϵm+ϵθ (1.4)

    for the solution of Problem 1. Using this expansion in (1.1), assuming that (u,m,λ) solves Problem 2, and matching powers of ϵ, we obtain the following problem that determines the terms λ, v, and θ in (1.4). To simplify the presentation, we discuss the case of C- solutions.

    Problem 3. Let g be as in Problem 1 with gC and let (u,m) be C- solutions of Problem 2 with m>0 and u=0. Find v,θ:TdR and λR such that

    {λ+u+DuDv=g(m)θin Td,div(mDv)div(θDu)=1min Td. (1.5)

    Remark 1.3. The normalization condition udx=0 is required for the uniqueness of the constant λ. Given a solution of (1.5), by adding a constant κ to u and subtracting κ to λ, we produce another solution.

    The existence of a solution to the preceding problem is established in Proposition 6.5 in Section 6. In that section, we prove the following improved asymptotic rate of convergence.

    Theorem 1.4. Suppose Assumption 2 holds. Let (uϵ,mϵ) and (u,m,ˉH), with m>0 and u=0, be classical solutions of, respectively, Problems 1 and 2. Let (v,θ,λ) be the corresponding classical solution to Problem 3. Then,

    limϵ0uϵˉHϵuλ+mϵm=0.

    Remark 1.5. The preceding theorem remains valid if we replace Assumption 2 with the weaker condition that for any z0>0 there exists γ(z0)>0 such that

    g(z)>γ(z0)

    for all z>z0.

    In the last section of the paper, Section 7, we investigate the asymptotic behavior of (uϵ,mϵ) as ϵ0. Here, we work with weak solutions in the sense of the definition below, and we consider the case where uniqueness of solution for Problem 2 may not hold. In this case, we replace Assumption 1 and 2 the following assumption that still allows the existence of solutions to be established.

    Assumption 4. There exist positive constants, c1,c20, and a positive real number, α, such that

    c1mα1g(m)c2mα1

    for all m>0.

    Remark 1.6. From the preceding hypothesis, we obtain that there exist positive constants ˜c1,˜c2 and C such that

    ˜c1mαCg(m)˜c2mα+C.

    Of course, if Assumption 1 does not hold, we cannot ensure the existence of smooth solutions to Problem 1. Nonetheless, the existence of weak solutions for Problem 1 was proven in [16].

    Closely related existence results are also addressed in [3]. The results in [16] requires less restrictive assumptions, albeit at the price of not having the uniqueness result from [3]. For a comparison between these two notions of weak solutions, we refer the reader to [16]. More precisely, we consider the following result.

    Theorem 1.7 (from [16]). Suppose that Assumption 4 holds and α>d42 if d8. Then, Problem 1 has a weak solution (mϵ,uϵ) as follows. There exists a constant C, independent of ϵ such that

    1. mϵ0 and Tdmϵdx=1,

    2. (mϵ)α+12W1,2(Td)C,

    3. uϵTduϵdxW1,2(Td)C,

    4. |ϵTduϵdx|C,

    5. (mϵ)α+12DuϵBV(Td)C.

    Moreover,

    ϵuϵ|Duϵ|22V(x)+g(mϵ)0 (1.6)

    in the sense of distributions, with

    (ϵuϵ|Duϵ|22V(x)+g(mϵ))mϵ=0, (1.7)

    almost everywhere. Furthermore,

    ϵmϵdiv(mϵDuϵ)=ϵ, (1.8)

    in the sense of distributions and almost everywhere.

    We note that in [16], the specific form of Hamiltonian H(x,p)=12|p|2+V(x) was crucial to get the existence solutions with the properties above. Similar techniques applied to Problem 2 yield the existence of a number ¯H and functions (m,u) satisfying estimates 1-3 and 5 in Theorem 1.7 such that

    ¯H|Du|22V(x)+g(m)0 (1.9)

    in the sense of distributions, with

    (¯H|Du|22V(x)+g(m))m=0, (1.10)

    almost everywhere. Furthermore,

    div(mDu)=0, (1.11)

    in the sense of distributions and almost everywhere.

    When classical solutions are not available, we need to work with regular weak solutions, as defined next.

    Definition 1.8. A pair (mϵ,uϵ) is a regular weak solution of Problem 1 if it satisfies (1.6), (1.7) and (1.8) in the preceding theorem and, in particular, the same estimates 1-5 with the same constants. Similarly, a triple (u,m,¯H) is a regular weak solution of Problem 2 if it satisfies (1.9), (1.10), (1.11) and the estimates 1-5 in the preceding theorem with the same constants.

    In Section 7, Proposition 7.1, we consider a sequence of regular weak solution of Problem 1 and show that, by extracting a subsequence if necessary, it converges to a regular weak solution of Problem 2. In particular, this approach gives the existence a regular weak solution for Problem 2.

    Our selection result for regular weak solutions, proven in Section 7, is the following theorem.

    Theorem 1.9. Suppose that Assumption 4 holds and α>d6d+2 if d>6. Let (uϵ,mϵ) be a regular weak solution of Problem 1. Suppose that uϵˉu in H1(Td) and that mϵˉm weakly in L1(Td). Let (u,m) be a regular weak solution of Problem 2. Then,

    Td(g(mϵ)g(m))(mϵm)dx0 (1.12)

    and ˉm=m. Moreover, we have

    TdˉumdxTdumdx, (1.13)

    where

    f=f(x)Tdfdx.

    The proof of the preceding theorem relies on ideas from Aubry-Mather theory introduced in [19]. The paper ends with a short example that illustrates the preceding result.

    Here, we examine the uniqueness of solutions of (1.2). First, we use the uniqueness method by Lasry-Lions [29] to show that the probability density, m, is unique. Thus, failure of uniqueness for (1.2) requires multiplicity of solutions, u, of the Hamilton-Jacobi equation. Second, we revisit an example from [23], where uniqueness does not hold. This example serves to illustrate the selection principle derived in Section 7.

    The monotonicity argument introduced by Lasry-Lions (see, [28] or the lectures [30]), can be used to prove the uniqueness of solution for MFGs in the time-dependent case and gives the uniqueness of m in the stationary problem. Here, we apply this technique to Problem 2. Let (u1,m2,ˉH1) and (u2,m2,ˉH2) be classical solutions of (1.2). Then,

    {12|Du1|212|Du2|2+ˉH1ˉH2=g(m1)g(m2)div(m1Du1)+div(m2Du2)=0.

    Now, we multiply the first equation by (m1m2) and the second equation by (u1u2). Next, subtracting the resulting identities and integrating by parts, we obtain

    Td(m1m2)(g(m1)g(m2))+12(m1+m2)|Du1Du2|2dx=0. (2.1)

    Accordingly, m1=m2=m on Td because g is strictly increasing. Moreover, Du1=Du2 on m>0. Hence, classical solutions (u,m,ˉH) of (1.2) with m>0 are unique up to an additive term in u. Uniqueness may fail if m vanishes, as we show in Section 2.2. A similar proof gives that (2.1) holds for the solutions of (1.1). By Lemma 1, mϵ is positive. Hence, classical solutions of Problem 1 are unique.

    Here, we show two distinct regular weak solutions to (1.2). In the example below, the existence of a unique smooth solutions fails and m vanishes at an interval.

    Let g(m)=m, d=1, and V(x)=πcos(2πx). Then, (1.2) becomes

    {12|ux|2+πcos(2πx)=m+ˉHin T,(mux)x=0in T,m(x)0,  m=1. (2.2)

    From the second equation in (2.2), mux=c for some constant. We claim that, mux=0. Indeed, if c0, ux=cm. This is not possible because Tuxdx=0. Thus, u is constant on the set m>0. From the first equation in (2.2) and taking into account that Tmdx=1, we have ˉH=0 and, thus,

    m(x)=(πcos(2πx))+.

    The preceding expression vanishes in an interval, as can be seen in Figure 1.

    Figure 1.  Density m for (2.2) which exhibits areas with no agents.
    Figure 2.  Two distinct solutions, ˆu and ˜u, of the Hamilton-Jacobi equation in (2.2). Their gradients differ only when m vanishes.

    On the other hand, from the first equation in (2.2), we see that uH1(T1) is a regular weak solution if u satisfies

    ux0a.e.on{0<x<1/4}{3/4<x<1}, (2.3)

    and

    |ux|22πcos(2πx)a.e.on{1/4<x<3/4}. (2.4)

    For example, We set the functions ˜u and ˆu by

    ˜ux(x)=(2πcos(2πx))+χ{14<x<12}(2πcos(2πx))+χ{12<x<34},

    and

    ˆux(x)=(2πcos(2πx))+χ{14<x<38}{12<x<58}+(2πcos(2πx))+χ{38<x<12}{58<x<34},

    where χ is the characteristic function. Then, we observe that (ˆu,m,0) and (˜u,m,0) satisfy (2.3) and (2.4). Thus, (˜u,m,¯H) and (ˆu,m,¯H) are regular weak solutions.

    In this section, we establish preliminary a priori estimates for solutions of Problem 1. To simplify the notation, we denote by (u,m) a solution of Problem 1, instead of (uϵ,mϵ). Here, we seek to establish bounds for (u,m) that are uniform in ϵ. Accordingly, the bounds in this section depend only on the data, g, V, and d but not on ϵ nor on the particular solution. First, we show that m is a probability; that is, nonnegative and its integral is 1. Next, we establish a lower bound and higher integrability for m. Finally, we prove Lipschitz bounds for u, which give the regularity of the solutions in the one-dimensional case. The higher dimensional case requires further estimates that are addressed in the following section.

    Proposition 3.1. Let (u,m) be a classical solution of Problem 1. Then, for every xTd, m(x)>0 and

    Tdmdx=1. (3.1)

    Proof. First, we show the positivity. Suppose that x0Td is such that m(x0)=minxTdm(x)=0. At this point, the second equation in (1.1) becomes

    ϵm(x0)Dm(x0)Du(x0)m(x0)Δu(x0)=ϵ.

    However, the left-hand side is 0, which is a contradiction. To check (3.1), we integrate the second equation in (1.1) and use integration by parts. Then, we see that

    ϵTdmdx=ϵ.

    Thus, we get the conclusion.

    Next, we get a uniform lower bound for m.

    Proposition 3.2. Suppose that Assumption 1 holds. Then, there exists a constant, C>0, such that for any classical solution, (u,m) of Problem 1, we have

    ϵuL(Td)+1mL(Td)C.

    Proof. First, we bound ϵuL(Td). Let ˜xTd be a minimum point of u. At this point, Du(˜x)=0 and Δu(˜x)0. From the second equation in (1.1), we get

    m(˜x)=ϵϵΔu(˜x).

    Since m is positive, Δu(˜x)<ϵ and, thus, m(˜x)1. Because g is increasing, it follows from the first equation in (1.1) that

    ϵu(˜x)g(1)maxxTdV(x). (3.2)

    Next, let ˆxTd be a maximum point of u. By an analogous argument, we get

    ϵu(ˆx)g(1)+minxTdV(x).

    Thus, ϵuL(Td)C.

    Now, we address the lower bound for m. By the first equation in (1.1) and (3.2), for all xTd, we have

    g(m(x))=ϵu(x)+12|Du(x)|2+V(x)ϵu(˜x)+minxTdV(x)g(1)oscV.

    Using Assumption 1, we get the lower bound for m.

    In the following Lemma, we give an upper bound for m.

    Lemma 3.3. Suppose that Assumptions 1 and 2 hold. Then, there exists a constant, C>0, such that for any classical solution, (u,m), of Problem 1, we have

    Td1+m2|Du|2+mβ+2dxC. (3.3)

    Proof. First, we multiply the first equation in (1.1) by (1m) and the second equation in (1.1) by u. Integrating by parts and adding the resulting identities, we have

    Td1+m2|Du|2+mg(m)dx=Td(m1)V+g(m)dx. (3.4)

    Using Assumption 2, we get

    Tdmg(m)dxC.

    On the other hand, in light of Proposition 3.2, there exists 0<m0<m(x) for all xTd. Furthermore, Assumption 2 guarantees that for all tm0,

    g(t)Cβ+1tβ+1Cβ+1m0β+1+g(m0).

    Therefore, combining the preceding inequalities with (3.4), we obtain (3.3).

    In the next proposition, we establish that u is Lipschitz continuous and get uniform bounds for m using a technique introduced in [9].

    Proposition 3.4. Suppose that Assumptions 1 and 2 hold. Then, there exists a constant C>0 such that for any classical solution, (u,m), of Problem 1, we have

    DuL(Td)+mL(Td)C.

    Proof. At first, we show the case d=1. Multiplying the first equation by mxx and the second by uxx, we obtain

    {ϵumxx+12mxxu2x+mxxV=g(m)mxxϵmuxxuxx(mux)x=ϵuxx.

    Next, we subtract these equations and integrate by parts to get

    Tmu2xx+g(m)m2xdx=TmxVxdxδTm2xdx+14δTV2xdx,

    using a weighted Cauchy-Schwarz inequality with δ>0. Because m is bounded by below, taking δ>0 sufficiently small, mxL2(T) and uxxL2(T) are bounded. In view of (3.3), we get the desired result.

    Next, we discuss the case d2. Take pβ. Multiplying the second equation in (1.1) by div(mpDu), we obtain

    Tdϵmdiv(mpDu) dx=Tddiv(mDu)div(mpDu) dx. (3.5)

    Differentiating the first equation in (1.1), we get

    juxjuxjxi=g(m)mxiϵuxiVxi. (3.6)

    Next, we rewrite the right-hand side of (3.5) as follows:

    Tddiv(mDu)div(mpDu)=i,jTd(muxi)xi(mpuxj)xj dx=i,jTd(muxi)xj(mpuxj)xi dx=i,jTdpmp1(uximxi)(uxjmxj)+pmpmxiuxjuxixjdx+i,jTdmpuximxjuxixj+mp+1u2xixjdx=i,jTdpmp1(uximxi)(uxjmxj)+(p+1)mpmxiuxjuxixj+mp+1u2xixjdx=Tdpmp1|DmDu|2+(p+1)g(m)mp|Dm|2dx+Tdmp+1i,ju2xixj(p+1)mpjmxj(ϵuxj+Vxj)dx, (3.7)

    using (3.6) in the last line. Combining (3.7) and (3.5), we obtain

    Tdpmp1|DmDu|2+(p+1)g(m)mp|Dm|2+mp+1i,ju2xixjdx=Tdϵmdiv(mpDu)+(p+1)mpjmxj(ϵuxj+Vxj)dx=TdϵpmpDmDu+(p+1)mpDmDVdxTdϵp2(mp+1+mp1|DmDu|2)+(p+1)[δ|DmDV|2mp+β+Cδmpβ]dx,

    where the last inequality follows from a weighted Cauchy inequality with δ>0 and β is the exponent in Assumption 2. For δ sufficiently small, there exists C that does not depend on p, such that

    Tdmp+β|Dm|2dxCTdmp+1CTdmp+β+2dx, (3.8)

    where the last inequality is a consequence of Tdm=1 and p+β+2>p+1.

    For d>2, 2=2dd2 is the Sobolev conjugated exponent to 2; if d=2, we use the convention that 2 is an arbitrarily large real number. Using Sobolev's inequality and (3.8), we gather that

    [Td(mp+β+22)2dx]12C[Tdmp+β+2+|Dmp+β+22|2dx]1/2C(1+|p+β+2|)[Tdmp+β+2dx]1/2.

    Thus, there exists a positive constant C>0 such that for all qβ+1,

    mL2(q+1)2[C(1+q)]2q+1mLq+1.

    Next, we take 1<θ<2/2 and define rn=θn+β+1. In view of (3.3), mr0 is bounded. Now we observe that rnrn+1<22. Thus, for each nN, there exists 0<αn<1 satisfying

    rnrn+1=αn+1αn2/2.

    By Hölder's inequality and the above estimate with q+1=rn, we obtain

    mrn+1mαnrnm1αn2rn/2mαnrn{(Crn)2rnmrn}1αn=(Crn)2(1αn)rnmrn.

    Iterating the prior inequality, we get

    mrn+1mr0ni=0(Cri)2(1αi)ri.

    The right-hand side is bounded uniformly in nN because

    log(ni=0(Cri)2(1αi)ri)ni=02ri[C+log(ri)]<+.

    Hence, m is bounded. According to the first equation in (1.1), and using the bound for ϵu in Proposition 3.2, we obtain that Du is also bounded.

    When d=1, we can get additional estimates for Du and m, as shown in the next proposition. The case d2 is discussed in the next section.

    Proposition 3.5. Suppose that Assumption 1 holds. Let d=1. Then, there exists a constant C>0 such that for any classical solution, (u,m), of Problem 1, we have

    uxC1,α(T)+mC1,α(T)C.

    Proof. Differentiating the first equation in (1.1)and multiplying by m, we get

    ϵmux+uxmuxx+mVx=g(m)mmx.

    Solving the second equation in (1.1) for muxx and substituting in the above identity, we have

    mx=2ϵmuxϵux+mVx.(u2x+g(m)m). (3.9)

    Because m is bounded by below, the denominator in the preceding expression does not vanish. Thus, from the previous Proposition, the right-hand side is bounded. Accordingly, mxL(T) is bounded. Returning to the second equation (1.1), we see that uxxL(T) is bounded. Returning to (3.9), we see that mC1,α(T) is bounded. Thus, from the second equation in (1.1), we gather that uxC1,α(T) is bounded.

    Now, we obtain additional estimates for the solutions of (1.1) in the case d2. As in the previous section, to simplify the notation, we omit the ϵ in (uϵ,mϵ) and denote by (u,m) a solution of Problem 1. First, by solving the first equation in (1.1) for m, we get

    m=g1(ϵu+12|Du|2+V).

    Next, replacing the resulting expression into the second equation in (1.1), we obtain

    div[g1(ϵu+|Du|22+V)Du]ϵ[g1(ϵu+|Du|22+V)1]=0. (4.1)

    Here, we apply the DeGiorgi-Nash-Moser regularity method to (4.1) to obtain our estimates.

    We begin by selecting k with 1kd. Differentiating (4.1) with respect to xk, we conclude that v=uxk solves

    (aijvxj)xi=ϕxk+ψxi, (4.2)

    where

    aij(x)=g1(ϵu+12|Du|2+V)δij+(g1)(ϵu+12|Du|2+V)uxiuxj,
    ϕ(x)=ϵ[g1(ϵu+12|Du|2+V)1],
    ψ(x)=(g1)(ϵu+12|Du|2+V)(ϵuxiuxk+uxiVxk),

    and δij=1 if i=j and δij=0 otherwise. Because of Propositions 3.2 and 3.4, there exists a constant, C>0, such that for any classical solution, u, of (4.1), we have ϵu+DuC. Hence, we get

    ϕ+ψC. (4.3)

    Moreover, using again Propositions 3.2 and 3.4, we see that there exists a constant λ>0 such that for all ξRd, we have

    λ|ξ|2di,j=1aij(x)ξiξj1λ|ξ|2. (4.4)

    Next, we prove that v is Hölder-continuous and, thus, get higher regularity for u.

    Proposition 4.1. Suppose that Assumptions 1 and 2 hold. Let d2. Then, there exist constants, C>0 and 0<α<1, such that for any classical solution, u, of (4.1), we have

    DuC1,α(Td)C.

    Proof. Take R>0. Let v solve (4.2). Write v=z+w where z is a solution of

    (aijzxj)xi=ϕxk+ψxi, (4.5)

    in B2R and z=0 on B2R. Therefore, w solves

    (aijwxj)xi=0, (4.6)

    in B2R with w|B2R=v.

    We begin by establishing the following claim.

    Claim 1. For d>2, there exists a constant, C>0, that depends only on the bounds in (4.4) such that

    zL(B2R)CRfor anyR>0.

    Remark 4.2. If d=2, we get zL(B2R)ˉCRκ, for any κ<12. This difference is due to the exponent 2 in dimension 2 being replaced by an arbitrarily large constant. The argument that follows needs to be adapted accordingly, namely the bound in (4.8) below, but the key steps remain unchanged. This case will be omitted.

    Let k0. By multiplying (4.5) by (zk)+ and integrating by parts, we get

    B2Raijzxj(zk)+xidx=B2Rϕ(zk)+xk+ψ(zk)+xidx. (4.7)

    Set

    A(k)={z>k}B2R.

    It suffices to prove that we can choose a constant C>0 satisfying |A(CR)|=0. Because (zk)+xi=0 on A(k)C and (zk)+xi=zxi on A(k), we obtain from (4.7) that

    A(k)aijzxjzxidx=A(k)ϕzxk+ψzxidx.

    In view of (4.4) and the bounds in (4.3), we get

    A(k)|Dz|2dxC|A(k)|.

    Next, using Sobolev's inequality and taking into account that (zk)+=0 on B2R, we conclude that, for any h>k,

    (hk)2|A(h)|2/2[A(h)[(zk)+]2dx]2/2[B2R[(zk)+]2dx]2/2CB2R|D(zk)+|2dxCA(k)|Dz|2dx.

    Combining the two preceding estimates, we obtain

    |A(h)|C|A(k)|22(hk)2.

    Next, we take a sequence kn=M(112n), where

    M=(C|A(0)|2212(2)222)12.

    Using the above estimate, we obtain

    |A(kn+1)|C(kn+1kn)2|A(kn)|2/2C22(n+1)M2|A(kn)|2/2.

    We now prove by induction that

    |A(kn)||A(0)|2nμ,

    where μ=2221. The case n=0 is clear. Assume our claim holds for some n, we have to check that it holds for n+1. We have

    |A(kn+1)|C22(n+1)M2(|A(0)|2nμ)2/2=|A(0)|2(n+1)μ,

    using our choice for M and μ.

    Finally, by considering the limit n, we get |A(M)|=0. If d>2, we have

    M=ˆC|A(0)|1/dˆC|B2R|1/d=ˉCR. (4.8)

    Thus, we get Claim 1.

    Next, using the ellipticity bounds in (4.4), we apply the DeGiorgi-Nash-Moser estimate (see, [18], Theorem 8.22) to (4.6) to establish the following claim.

    Claim 2. We have

    osc(R2,w)ηosc(R,w),

    for some constant 0<η<1, where we denoted

    osc(R,w):=supBRwinfBRw.

    Combining the two preceding claims, we obtain the following estimate:

    osc(R/4,v)CR+osc(R/4,w)ˆCR+ηosc(R,v). (4.9)

    Claim 3. There exist constants C>0 and 0<α<1 such that for all 0<R<1, we have

    osc(R,v)CRα.

    Set

    Mn=sup14n+1R14nosc(R,v)Rα,

    where α satisfies 0<α<logηlog4.

    Here, we prove by induction that there exist μ>1 satisfying

    MnMμn (4.10)

    for some sufficiently large M>0. We choose μ satisfying

    1<μ<min(41α,14αη), (4.11)

    and then we choose M>24αvL(Td) and such that

    [ˆC(μ41α)n+1+η4αMμ]M. (4.12)

    The prior choice of M is possible due to (4.11).

    For n=0, M024αvL(Td)M. Next, we assume that (4.10) holds for some n0 and verify that it also holds for n replaced by n+1. Using (4.9), we have that

    Mn+1[ˆC(μ41α)n+1+η4αMμ]μ(n+1).

    Using the defining property of M, (4.12), we get (4.10).

    Finally, for 0<R<1, combining (4.9) and (4.10), we obtain

    osc(R,v)(4ˆC+η4αM)Rα,

    which establishes the claim.

    Claim 4.

    DuC1,α(Td)C.

    Due to Claim 3, we get DuC0,α(Td)C. Since (4.2) is a uniformly elliptic equation and the Hölder-norm of the coefficients is bounded. Therefore, it follows from Schauder's estimate that DvC0,α(Td)C. Hence, we conclude that DuC1,α(Td)C.

    Here, we prove the existence of a classical solution for (1.1), using the continuation method. In this Section, we suppose Assumption 1-3.

    A key difficulty is that g1 may not always defined on the whole real line. Thus, to perform the continuation method, we modify g as follows. We consider the case where limz0g(z)<. We begin by selecting m0R satisfying 0<m0<min{1,g1(g(1)oscV)}. Then, we define an increasing funcion h:(0,)R satisfying gh, h(z) as z0 and Assumption 2. Let η be a decreasing smooth function satisfying

    η1(x)={10<x12m00xm0.

    Now, we consider

    {ϵuϵ+12|Duϵ|2+V(x)=f(mϵ)in Td,ϵmϵdiv(mϵDuϵ)=ϵin Td, (5.1)

    where

    f(m):=η(m)h(m)+{1η(m)}g(m).

    Because f:(0,)R is increasing and surjective, we can define its inverse f1:R(0,). Moreoever, fg on (0,m0], and f=g on [m0,+). As (1.1) and (4.1) are equivalent, and (5.1) is equivalent to

    div{f1(ϵu+|Du|22+V)Du}ϵ{f1(ϵu+|Du|22+V)1}=0. (5.2)

    We check that f satisfies both Assumption 1 and Assumption 2. Indeed, because f1 is positive, f1(f(1)oscV)>0. Next, we see that for all z>0,

    f(z)η(z){C2+12zh(z)}+{1η(z)}{C2+12zg(z)}=C2+12zf(z).

    Finally, since gh and η is decreasing, we have

    f(z)=η(z){h(z)g(z)}+η(z)h(z)+{1η(z)}g(z)η(z)h(z)+{1η(z)}g(z)C1zβ.

    Thus, we obtain the estimates in Section 3 and 4 for (5.1) and (5.2). In the case limz0g(z)=, we can omit this resetting.

    Remark 5.1. Replacing g by f, we modify the behavior of g(z) as z0, to define its inverse on the whole real line. However, thanks to the lower bound for mϵ in Proposition 3.2, the above does not change the solution. Indeed, any classical solution (uϵ,mϵ) to (5.1) solves (1.1) as shown in the proof of Theorem 1.1.

    Remark 5.2. Without Assumption 3, we need to modify g(z) as z+, in addition to the preceding changes. However, this is more complicated because the upper bound for mϵ is more qualitative than the lower bound. On the other hand, typical examples of g are power or logarithm functions which satisfy Assumption 3.

    Here, we show the existence of classical solution to (5.2), using the continuation method. We begin by defining an operator, J:C2,α(Td)×[0,1]C0,α(Td), by

    J(u,λ)=div[f1(ϵu+|Du|22+λV)Du]ϵ[f1(ϵu+|Du|22+λV)1].

    We set

    Λ={λ[0,1]:uλC2,α(Td):J(uλ,λ)=0}. (5.3)

    We claim that Λ=[0,1]. First, we observe that 0Λ. In fact, for u0ϵ1f(1), we have J(u0,0)=0. Accordingly, Λ is non-empty. Thus, it suffices to check that Λ is relatively open and closed in [0,1], to get Λ=[0,1]. In the next proposition, we verify that Λ is a closed set.

    Proposition 5.3. Consider the setting of Problem 1 and suppose that Assumptions 1-3 hold. Let Λ as in (5.3). Then, Λ is relatively closed in [0,1].

    Proof. Fix a sequence λnΛ converging to λ[0,1] as n. We must show that λΛ. For that, take uλn satisfying J(uλn,λn)=0. Proceeding as in Section 3 and Section 4, replacing g and V by f and λnV respectively, we get bounds for {uλn}nN. In particular, the a priori bounds in Proposition 3.5 (for d=1) or Proposition 4.1 (for d>1) guarantee that there exists a subsequence of {uλn}nN converging to some uC2,α(Td). By passing to the limit, we conclude that J(u,λ)=0. Accordingly, λΛ.

    Now, we show that Λ is relatively open. For each λΛ, let uλC2,α(Td) solve J(uλ,λ)=0 and set

    mλ=f1(ϵuλ+|Duλ|22+λV).

    We consider the linearization of J around this solution and define Lλ:C2,α(Td)C0,α(Td) for ϕC2,α(Td) by

    Lλ(ϕ)=Jμ(uλ+μϕ,λ)|μ=0=div[mλDϕ+(f1)(f(mλ))(ϵϕ+DuλDϕ)Duλ]ϵ(f1)(f(mλ))(ϵϕ+DuλDϕ). (5.4)

    Lemma 5.4. Consider the setting of Problem 1. Let uλC2,α(Td) solve J(uλ,λ)=0 and let Lλ be given by (5.4). Then, Lλ is an isomorphism between C2,α(Td) and C0,α(Td).

    Proof. We must prove that for any ξC0,α(Td), the equation

    div[mλDϕ+(f1)(f(mλ))(ϵϕ+DuλDϕ)Duλ]ϵ(f1)(f(mλ))(ϵϕ+DuλDϕ)=ξ (5.5)

    has a unique solution, ϕC2,α(Td). We define Bλ:H1(Td)×H1(Td)R by

    Bλ[v,w]=TdDw[mλDv+(f1)(f(mλ))(ϵv+DuλDv)Duλ]+ϵTdw(f1)(f(mλ))(ϵv+DuλDv)dx.

    Note that if v and w are smooth, Bλ becomes Bλ[v,w]=(Lλ(v),w)L2. Using Hölder inequality, we see that Bλ is bounded. Now, using Riesz's Representation Theorem, we see that there exists a bounded linear operator A:H1(Td)H1(Td) such that Bλ[v,w]=(Av,w) for all wH1(Td). We divide the rest of the proof in the following three claims.

    Claim 1. There exists a constant, c>0, such that AvH1(Td)cvH1(Td) for all vH1(Td).

    We establish this claim by contradiction. For that, suppose that there exists {vn}nNH1(Td) with vnH1(Td)=1 and Avn0. Then,

    Bλ[vn,vn]=(Avn,vn)0.

    Next, we have

    Bλ[vn,vn]=Tdmλ|Dvn|2+(f1)(f(mλ)){ϵvn+DuλDvn}2dx.

    Since mλ and (f1) are positive, we see that Dvn0 and ϵvn+DuλDvn0 in L2(Td). Hence, we can construct a subsequence {vnk}kN satisfying vnk0 in H1(Td), which contradicts vnH1(Td)=1.

    Claim 2. The range of A, R(A), is closed and R(A)=H1(Td).

    Take a sequence {zn}nNR(A) that converges to zE. To prove the first part of the claim, we begin by showing that zR(A). For that, take wnH1(Td) satisfying zn=Awn. From the preceding claim, it follows that {wn}nN is a Cauchy sequence converging to some wH1(Td). By the continuity of A, we have z=Aw. Thus, zR(A).

    Next, to establish the last part of the claim, suppose that R(A)H1(Td). In this case, there exists a non-zero vector, vR(A). Then, we get

    0=(Av,v)=Bλ[v,v]=Tdmλ|Dv|2+(f1)(f(mλ)){ϵv+DuλDv}2dx.

    This contradicts v0.

    Claim 3. (5.5) has a unique solution ϕC2,α(Td).

    To prove this last claim, we define a bounded linear functional, T:H1(Td)R, by

    T(w)=Tdξwdx.

    The Riesz Representation Theorem guarantees that there exists a unique ˆwH1(Td) satisfying T(w)=(ˆw,w) for all wH1(Td). Taking ϕ=A1ˆw, we get

    Bλ[ϕ,w]=(Aϕ,w)E=(ˆw,w)E=T(w);

    that is, ϕH1(Td) is the unique weak solution of (5.5). Because (5.5) is a uniformly elliptic equation and the coefficients belong to C0,α(Td), we conclude that ϕC2,α(Td).

    To finish the proof of Theorem 1.1 we verify that Λ is relatively open. This is achieved in the next proposition.

    Proposition 5.5. Consider the setting of Problem 1 and suppose that Assumptions 1-3 hold. Let Λ as in (5.3). Then, Λ is relatively open in [0,1].

    Proof. By the preceding lemma, Lemma 5.4, we can apply the implicit function theorem (see [6]) to the operator J to conclude that Λ is open. Therefore, for any λΛ, there exists δ>0 such that for any ˆλ(λδ,λ+δ), we can find ˆuC2,α(Td) such that J(ˆu,ˆλ)=0.

    By combining the previous results, we prove Theorem 1.1 as follows.

    Proof of Theorem 1.1. Since 1Λ, there exists a classical solution uϵ for (5.2). Take mϵ=f1(ϵuϵ+12|Duϵ|2+V). Then, (uϵ,mϵ) solves (5.1). Arguing as in Proposition 3.2, from (5.1) we obtain that mϵf1(f(1)oscV). By the definition of f, noting that m0<1, we obtain that for all xTd

    mϵ(x)f1(f(1)oscV)=f1(g(1)oscV)g1(g(1)oscV)>m0.

    Because g(mϵ)=f(mϵ) on Td, (uϵ,mϵ) solves (1.1).

    The identity (2.1) gives that (1.1) has a unique classical solution. Indeed, let (uϵ1,mϵ1) and (uϵ2,mϵ2) be classical solutions to (1.1). Then,

    {ϵ(uϵ1uϵ2)+12|Duϵ1|212|Duϵ2|2=g(mϵ1)g(mϵ2)ϵ(mϵ1mϵ2)div(mϵ1Duϵ1)+div(mϵ2Duϵ2)=0.

    Now, we multiply the first equation by (mϵ1mϵ2) and the second equation by (uϵ1uϵ2). Next, subtracting the resulting identities and integrating by parts, we obtain

    Td(mϵ1mϵ2)(g(mϵ1)g(mϵ2))+12(mϵ1+mϵ2)|Duϵ1Duϵ2|2dx=0.

    Accordingly, mϵ1=mϵ2 on Td because g is strictly increasing. Moreover, Duϵ1=Duϵ2 on m>0. In view of positivity of mϵ, we see Duϵ1=Duϵ2 on Td. Hence, from the first equation in (1.1), uϵ1=uϵ2 on Td.

    Finally, we show that under Assumptions 1 and 2, we have the convergence of the solutions of (1.1).

    Proof of Corollary 1.2. The estimates in Section 3 and 4 do not depend on ϵ. Therefore, we can extract a subsequence ϵj such that ϵjuϵj converges uniformly to a constant ˉH and (uϵjTduϵjdx,mϵj) converges to some (u,m) in C2,α×C1,α. Therefore, (u,m,ˉH) solves (1.2). By the results in Section 2, m and ˉH are uniquely determined. Accordingly, the limit of ϵjuϵj and mϵj does not depend on the subsequence. Because of Proposition 3.2, we have that m>0. Thus, there exists a unique solution, (u,m,ˉH) of (1.2) satisfying the additional condition Tdudx=0.

    Now, we investigate the asymptotic behavior of {uϵ¯H/ϵ}ϵ>0 and prove Theorem 1.4, thus improving the converge results in Corollary 1.

    First, to address Problem 3, we consider the linearized discounted problem that we state now.

    Problem 4. Let g be as in Problem 1 with gC and let (u,m)C(Td)×C(Td) solve Problem 2 with m>0. Suppose that ϵ>0. Given A,BC(Td), find vϵ,θϵ:TdR such that

    {ϵvϵ+u+DuDvϵ=g(m)θϵ+Ain Td,ϵθϵdiv(mDvϵ)div(θϵDu)=1m+div(B)in Td. (6.1)

    Proposition 6.1. Suppose Assumption 2 holds. Then, Problem 4 has a unique weak solution (vϵ,θϵ)H1(Td)×L2(Td).

    Proof. Because m>0, Assumption 2 (or the alternative assumption in Remark 1.5) gives that g(m) is bounded by below. From the first equation in (6.1), we get

    θϵ=ϵvϵ+u+DuDvϵAg(m). (6.2)

    Using the previous expression for θϵ in the second equation in (6.1), we obtain

    div(mDvϵ+ϵvϵ+DuDvϵg(m)Du)+ϵ(ϵvϵ+DuDvϵ)g(m)=1m+div(B)+div(uAg(m)Du)ϵ(uA)g(m). (6.3)

    Therefore, it suffices to show that (6.3) has a weak solution. For that, we define a bilinear form, K:H1(Td)×H1(Td)R, by

    K[ϕ1,ϕ2]=TdmDϕ1Dϕ2+ϵϕ1+DuDϕ1g(m)DuDϕ2+ϵϕ2(ϵϕ1+DuDϕ1)g(m)dx.

    Because m and u are smooth with g(m) bounded by below, we see that K is a bounded bilinear form. Moreover, for all ϕH1(Td),

    K[ϕ,ϕ]=Tdm|Dϕ|2+(DuDϕ+ϵϕ)2g(m)dx.

    Hence, K is coercive. Thus, applying the Lax-Milgram theorem, we see that (6.3) has a unique weak solution, vϵH1(Td). Then, using (6.2) and taking into account that g(m) is bounded by below, we obtain a weak solution θϵL2(Td).

    Proposition 6.2. Let (vϵ,θϵ)H1(Td)×L2(Td) be a weak solution of Problem 4. Then, there exists a constant C>0 independent of ϵ such that

    ϵvϵL2(Td)+θϵL2(Td)+DvϵL2(Td)C(AL2(Td)+BL2(Td)+1).

    Proof. We multiply the first equation in (6.1) by θϵ and the second one by vϵ. Next, we subtract the resulting expressions to get

    uθϵ+θϵDuDvϵ+vϵdiv(mDvϵ)+vϵdiv(θϵDu)=g(m)|θϵ|2(1m)vϵvϵdiv(B)+Aθϵ.

    Integrating by parts, we obtain

    Tdg(m)|θϵ|2+m|Dvϵ|2 dx=Td(uA)θϵ+(1m)vϵ+vϵdiv(B) dx.

    Using Poincaré's inequality, we conclude that

    Tdm|Dvϵ|2 dxTd(uA)θϵ+(1m)vϵ+vϵdiv(B) dx=Td(uA)θϵ+(1m)(vϵvϵ dx)+vϵdiv(B) dxuL2(Td)θϵL2(Td)+AL2(Td)θϵL2(Td)+1mL2(Td)DvϵL2(Td)+BL2(Td)DvϵL2(Td).

    Hence, taking into account that m is bounded by below,

    Dvϵ2L2(Td)C(θϵL2(Td)+AL2(Td)θϵL2(Td)+B2L2(Td)). (6.4)

    Arguing analogously, we obtain

    Tdg(m)|θϵ|2 dxuL2(Td)θϵL2(Td)+AL2(Td)θϵL2(Td)+1mL2(Td)DvϵL2(Td)+BL2(Td)DvϵL2(Td).

    Hence, by (6.4),

    θϵ2L2(Td)C(A2L2(Td)+B2L2(Td)+1).

    Combining the preceding inequality with (6.4), we have the estimate

    Dvϵ2L2(Td)C(A2L2(Td)+B2L2(Td)+1).

    Finally, the first equation in (6.1) yields

    ϵvϵ2L2(Td)C(A2L2(Td)+B2L2(Td)+1).

    Next, we bootstrap higher regularity for (vϵ,θϵ).

    Proposition 6.3. Let (vϵ,θϵ) be a weak solution of Problem 4. Fix h{1,2,,...,d} and let z=vϵxh. Then, for each kN, there exists a constant Ck>0 such that

    zHk(Td)Ck(1+AHk(Td)+BHk(Td)).

    Proof. We begin by rewriting (6.3) as

    div(mDvϵ+DuDvϵg(m)Du)=div(ϵvϵg(m)Du)ϵ(ϵvϵ+DuDvϵ)g(m)+1m+div(B)+div(uAg(m)Du)ϵ(uA)g(m).

    Next, we fix h{1,2,...,d} and let z=vϵxh. Differentiating the preceding equation with respect to xh, we obtain

    (aijzxj)xi=ϕxh+ψixi, (6.5)

    where

    aij=δijm+uxiuxjg(m),
    ψi=mxhvxiDuDvϵg(m)uxixh+DuDvϵg(m)2g(m)uximxh+DvϵDuxhg(m)uxi,

    and

    ϕ=div(ϵvϵg(m)Du)ϵ(ϵvϵ+DuDvϵ)g(m)+1m+div(B)+div(uAg(m)Du)ϵ(uA)g(m).

    By the previous proposition, we know that

    zL2(Td)C(AL2(Td)+BL2(Td)+1).

    Furthermore, we have the estimates

    ϕL2(Td)C(AH1(Td)+BH1(Td)+1),

    and

    ψL2(Td)C(AL2(Td)+BL2(Td)+1).

    Let k0. Multiplying (6.5) by z and integrating by parts, we get

    Tdaijzxjzxidx=Tdϕzxh+hzxidx.

    Because of the uniform ellipticity of aij, we get

    Td|Dz|2dxC(ϕL2(Td)+hL2(Td))DzL2(Td).

    Hence,

    DzL2(Td)C(ϕL2(Td)+hL2(Td))C(1+AH1(Td)+BH1(Td)).

    Therefore,

    ϕH1(Td)C(AH2(Td)+BH2(Td)+1),

    and

    ψH1(Td)C(AH1(Td)+BH1(Td)+1).

    The conclusion follows by iterating this argument for higher derivatives.

    Proposition 6.4. Let (vϵ,θϵ) be a weak solution of Problem 4. Then, for each kN, there exists a constant Ck>0 such that

    ϵvϵHk(Td)+θϵHk(Td)Ck(1+AHk(Td)+BHk(Td)).

    In particular, (vϵ,θϵ) is a classical solution of (6.1).

    Proof. Differentiating the first equation in (6.1), we get

    DθL2(Td)C(1+AH1(Td)+BH1(Td)).

    The above implies

    ϵvϵH1(Td)C(1+AH1(Td)+BH1(Td)).

    Iterating the preceding steps, we get the result.

    Proposition 6.5. For ϵ>0, let (vϵ,θϵ) be a weak solution of Problem 4 and assume that

    AL(Td)+BL(Td)Cϵ

    for some constant C>0. Then, there exists a solution (v,θ,λ) of Problem 3 and, for each kN,

    limϵ0(ϵvϵλL(Td)+θϵθHk(Td)+(vϵTdvϵdx)vHk(Td))=0.

    Proof. By the previous estimates on the solutions of Problem 4, we can choose a subsequence such that ϵvϵλ, θϵθ and vϵvϵv. Clearly, (v,θ,λ) solves (1.5). Because the solution to (1.5) is unique, the limit is independent of the particular sequence. Therefore, (vϵ,θϵ) converges to (v,θ).

    Finally, we present the proof of Theorem 1.4.

    Proof of Theorem 1.4. Fix kN and set

    Ek={(v,θ)Hk+1(Td)×Hk(Td),ϵvHk(Td)+DvHk(Td)+θHk(Td)^Ck},

    where ^Ck is to be chosen later. For (v,θ)Ek, we find (ˆv,ˆθ) solving (6.1), where

    A(x)=ϵ2vϵ22|Dv|2+g(m+ϵθ)g(m)ϵg(m)θϵ,

    and

    B(x)=ϵ2θDvϵ.

    Because,

    AHk(Td)+BHk(Td)C^Ck2ϵ,

    we obtain

    ϵvHk(Td)+DvHk(Td)+θHk(Td)Ck(AHk(Td)+BHk(Td)+1)Ck(1+^Ck2ϵ).

    We can choose ^Ck such that, for ϵ small enough, the right-hand side is less than ^Ck. Then, it holds that (v,θ)(ˆv,ˆθ) has a fixed point (vϵ,θϵ). We remark that (ϵvϵ+u+ˉHϵ,m+ϵθϵ) solves (1.1) and therefore it is equal to (uϵ,mϵ). Hence, by the previous proposition, for suitably large k, as ϵ0,

    uϵˉHϵuλ=ϵvϵλ0.

    Now, we investigate the behavior of (uϵ,mϵ) as ϵ0 in the case where Problem 2 may have multiple solutions; that is, when Assumptions 1 and 2 do not hold. We are interested in which of the solutions of Problem 2 arise as a limit of solutions of Problem 1. Without Assumptions 1 and 2, smooth solutions may not exist. Therefore, we need to work with weak solutions. For Problem 1, weak solutions were shown to exist in [16]. In Section 7.1, we review those existence results and use them to show the existence of a solution for Problem 2. Then, in Section 7.2, we construct certain measures on phase space that generalize Mather measures. Next, in Section 7.3, we prove our main selection result, Theorem 1.9. We end the paper with a discussion of an explicit example, in Section 7.4.

    We begin this section by proving the following result on the stability of regular weak solutions. In particular, since the estimate of regular weak solutions of Problem 1 was proved in [16], we obtain the existence of regular weak solutions for Problem 2.

    Proposition 7.1. Suppose that Assumption 4 holds and α>d4d if 4<d. Let (mϵ,uϵ) be a regular weak solution of Problem 1. Assume that mϵm weakly in L1(Td), uϵuϵdxu weakly in W1,2(Td), and that ϵuϵdx¯H. Then, (m,u,¯H) is a regular weak solution of Problem 2.

    Proof. Properties 1 and 3 in Theorem 1.7 are immediate; that is,

    m0,Tdmdx=1,uϵTduϵdxW1,2(Td)C.

    From Property 2, we conclude that, through a subequence (mϵ)α+12, converges weakly in W1,2 to a function η(x). Moreover, by Rellich-Kondrachov theorem, (mϵ)α+12η in L2(Td) and extracting a further sequence if necessary also almost everywhere. Therefore, mϵη2α+1=m almost everywhere. Accordingly (mϵ)α+12 converges to mα+12 weakly in W1,2(Td), strongly in L2(Td) and almost everywhere. Consequently

    mα+12W1,2(Td)C.

    Next, we examine Property 5 in Theorem 1.7. By compactness, (mϵ)α+12Duϵ converges weakly in BV, through a subsequence, to a function ψBV with ψBV(Td)C. Because (mϵ)α+12 converges to mα+12 strongly in L2(Td) and Duϵ converges weakly to Du in L2(Td), we have for any φC(Td)

    Tdφ(mϵ)α+12DuϵdxTdφmα+12Dudx.

    Therefore,

    ψ=mα+12Du.

    Finally, we address the limit properties corresponding to (1.6), (1.7) and (1.8). We take a smooth function, φC(Td), multiply (1.8) by φ and integrate. Because

    Tdϵ(mϵ1)φdx0,

    we have

    TdmϵDφDuϵdx0.

    Because of Rellich-Kondrachov theorem

    Td(mϵ)qα+12dxTdmqα+12dx,

    for any q<2. In particular, for α in the range of Assumption 4 this implies mϵm strongly in L2(Td). Using the weak convergence of uϵ in W1,2(Td) we conclude that

    TdmDφDudx=0.

    Next, we select a smooth non-negative function, φ0, multiply (1.6) by φ, and integrate in Td. We have

    Td(ϵuϵV(x))φdxTd(¯HV(x))φdx.

    Moreover, by convexity

    lim infϵ0Td|Duϵ|22φdxTd|Du|22φdx.

    Finally, we observe that

    Tdg(mϵ)g(m)dx=Td10g(smϵ+(1s)m)(mϵm)dx.

    For any α>0, we can select p and p such that 1p+1p+1, p(α1)<22(α+1), and p<22(α+1). Next, we estimate

    g(smϵ+(1s)m)LpC(mϵLp(α1)+mLp(α1))C.

    Therefore, since mϵmLp0, we conclude that

    Tdg(mϵ)g(m)dx0.

    Proposition 7.2. Suppose that Assumption 4 holds. Let (mϵ,uϵ) be a regular weak solution of Problem 1. Then, there exists a constant C>0 independent of ϵ such that

    Td|Duϵ|2mϵ+mϵg(mϵ)dxC. (7.1)

    Proof. Because of properties 1 and 2 in Theorem 1.7, Assumption 4 implies that

    Tdmϵg(mϵ)dxC.

    Then, (7.1) follows by integrating (1.7).

    Remark 7.3. If Assumptions 4 holds, a similar estimate holds for regular weak solutions (m,u,¯H) of Problem 2. Namely, there exists a constant C>0 such that

    Td|Du|2m+mg(m)dxC.

    We begin by introducing a class of phase-space probability measures called Mather measures, see [32] and [33]. These measures were introduced in the context of Lagrangian mechanics and later used to examine the properties of Hamilton–Jacobi equations in [12,13,14,15] and in [10,11]. In the context of the selection problem, generalized Mather measures were first used in [19]. As previously, we suppose that Assumption 4 holds. Accordingly, we work with regular weak solutions of Problems 1 and 2.

    Fix a regular weak solution (uϵ,mϵ) of Problem 1 and a regular weak solution (u,m) of Problem 2. Next, we rewrite (1.1) as

    {ϵuϵ+12|Duϵ|2+Wϵ(x)=0in Td,ϵmϵdiv(mϵDuϵ)=ϵin Td,

    where Wϵ(x)=V(x)g(mϵ) and, assuming without loss of generality that ¯H=0, we rewrite (1.2) as

    {12|Du|2+W(x)=0in Td,div(mDu)=0in Td,

    where W(x)=V(x)g(m).

    Proposition 7.4. Suppose that Assumption 4 holds. Let (uϵ,mϵ) be a regular weak solution of Problem 1. Let Lϵ=12|v|2Wϵ(x) with Wϵ(x)=V(x)g(mϵ). Consider the phase-space measure, μϵ, the ϵ-Mather measure, determined by

    Td×Rdϕ(x,v)dμϵ=Tdϕ(x,Duϵ)mϵdx

    for all ϕC(Td×Rd). Then, μϵ satisfies the discounted holonomy condition

    Td×Rd(ϵφ(x)+vDxφ(x))dμϵ=ϵTdφdx (7.2)

    for all φC1(Td). Moreover, we have

    Td×RdLϵ(x,v)dμϵ=ϵTduϵdx. (7.3)

    Proof. Because (1.8) holds in the sense of distributions, the discounted holonomy condition for μϵ, (7.2), follows immediately. Next, recall that if mϵ is an integrable non-negative function then L21+mϵ(Td), the space of all measurable functions f:TdR that satisfy

    Td|f|2(1+mϵ)dx<,

    is a Hilbert space. Moreover, if ηδ is a standard mollifier, we have

    ηδff

    in L21+mϵ(Td). Due to (1.6) and (1.7), we have

    DuϵL21+mϵ(Td).

    Accordingly, because of (7.2), we have

    Td×Rdϵ(ηδuϵ)+vDx(ηδuϵ) dμϵ=ϵTd(ηδuϵ)dx.

    Taking the limit δ0, we obtain

    Td×Rdϵuϵ+vDxuϵ dμϵ=ϵTduϵdx. (7.4)

    Taking into account the definition of Lϵ and using the identities (1.7) and (7.4), we conclude that

    Td×RdLϵ(x,v) dμϵ=Td×Rd12|v|2Wϵ(x) dμϵ=Td(12|Duϵ|2Wϵ(x))mϵdx+Td×RdvDxuϵϵuϵ dμϵ+ϵTduϵdx=ϵTduϵdx.

    Remark 7.5. From (7.2) and (7.3), we conclude that μϵ is a discounted Mather measure with trace dx in the sense of the definition in [19].

    Similarly, for Problem 2, we have the following result.

    Proposition 7.6. Suppose that Assumption 4 holds. Let (u,m,¯H) be a regular weak solution of Problem 2. Assume without loss of generality that ¯H=0. Let L=12|v|2W with W(x)=V(x)g(m). Consider the phase-space measure, μ, the Mather measure, determined by

    Td×Rdϕ(x,v)dμ=Tdϕ(x,Du)mdx,

    for all ϕC(Td×Rd). Then, μ satisfies the holonomy constraint

    Td×RdvDxφ(x)dμ=0

    for all φC1(Td). Moreover

    Td×RdL(x,v)dμ=Td(12|Du|2W(x))mdx=0.

    Proof. The proof is analogous to the one of Proposition 7.4

    The goal of this section is to prove Theorem 1.9 and, hence, establish a selection criterion for the limit of uϵ and prove the convergence of mϵ. Our proof is inspired by the one in [19].

    Proof of Theorem 1.9. Let (uϵ,mϵ) be a regular weak solution of Problem 1. Suppose that uϵuϵdxˉu in H1(Td) and that mϵˉm weakly in L1(Td). Note that due to the bounds in Theorem 1.7, we have that mϵˉm strongly in Lp for p<22(α+1). Let (u,m) be a regular weak solution of Problem 2. Let μϵ and μ be the Mather measure constructed in the previous section in Propositions 7.4 and 7.6.

    For any vRd and almost every xTd, we have

    vpLϵ(x,v)12|p|2+Wϵ(x),

    where Wϵ is as in Proposition 7.4. Consider a standard mollifier ηδ, and let p=D(ηδu). Then, for vRd and almost every xTd,

    vD(ηδu)Lϵ(x,v)+WWϵ12|D(ηδu)|2+Wϵ+WWϵ=12|D(ηδu)|2+W, (7.5)

    where W is as in Proposition 7.6. Integrating the left-hand side of the preceding expression with respect to μϵ and using the holonomy condition, (7.2), and (7.3), we obtain

    Td×RdvD(ηδu)Lϵ(x,v)+WWϵdμϵ=ϵTd(ηδu)mϵdx+ϵTd(ηδu)dxϵTduϵdx+Td(WWϵ)mϵdx=ϵTd(ηδu)mϵdx+ϵTd(ηδu)dxϵTduϵdx+Td(WWϵ)mϵdx. (7.6)

    Next, we integrate the right-hand side of (7.5) and use Jensen's inequality to obtain

    Td×Rd12|D(ηδu)|2+WdμϵTd×Rd12ηδ(|Du|2)+WdμϵTd×RdηδW+Wdμϵ, (7.7)

    taking into account (1.9). Because WLα+1α(Td), ηδWW in Lα+1α(Td) as δ0. Therefore, taking into account that mϵLα+1(Td), we have

    Td×RdηδW+Wdμϵ0,

    as δ0. Therefore, combining (7.5), (7.6), and (7.7), and by considering the limit δ0, we conclude that

    ϵTdumϵdx+ϵTdudxϵTduϵdx+Td(WWϵ)mϵdx0. (7.8)

    On the other hand, we observe that, for all vRd and almost every xTd,

    ϵuϵ+vD(ηδuϵ)L(x,v)+WϵWϵuϵ+12|D(ηδuϵ)|2+W+WϵW.

    Integrating with respect to μ and proceeding in a similar manner, we obtain

    ϵTduϵmdx+Td(WϵW)mdx0. (7.9)

    Next, from (7.8), we gather

    ϵTdumϵdxϵTduϵdx+Td(WWϵ)mϵdx0.

    Finally, from (7.9), we get

    ϵTduϵmdx+ϵTduϵdx+Td(WϵW)mdx0.

    Adding the above two inequalities, we obtain

    ϵ(TduϵmdxTdumϵdx)+Td(WWϵ)(mϵm)dx0. (7.10)

    By Sobolev's inequality, we have uϵ,uL2(Td), uniformly in ϵ. Moreover, m,mϵL22(α+1)(Td), uniformly in ϵ. In the range of α, we observe m,mϵL(2)(Td), where (2) is the Hölder's conjugate of 2. Therefore, uϵmdx and umϵdx are bounded uniformly in ϵ. Consequently, the first term in the left-hand side of (7.10) converges to 0. Hence, we obtain (1.12). Moreover, because the second term is non-negative, we conclude by the monotonicity of g that

    TduϵmdxTdumϵdx0.

    Hence, (1.13) holds.

    Finally, in this subsection, we present an application of our selection criterion. We consider the following discount problem:

    {ϵuϵ+12|uϵx|2+πcos(2πx)=mϵin T1,ϵmϵ(mϵuϵx)x=ϵin T1, (7.11)

    Thus d=1, V(x)=πcos(2πx), and g(m)=m. The associated limit problems of (7.11) is (2.2) in Section 2.2. By Theorem 1.7, there exists a regular weak solution (uϵ,mϵ)H1(T1)×H1(T1), of (7.11). We note that uϵTduϵdx converges along subsequence weakly in H1(T1). In view of Proposition 7.1, the limit is a regular weak solution of (2.2). However, regular weak solutions for (2.2) are not unique, as we show in Section 2.2.

    By (1.13), we get the following result:

    Proposition 7.7. Let (uϵ,mϵ) be regular weak solution of (7.11). Suppose that uϵuϵdxˉu as ϵ0 weakly in H1(T1). Let (u,m) be any regular weak solution of (2.2). Then,

    T1ˉumdxT1umdx.

    Using the above criterion, we can show that uϵT1uϵdx fully converges weakly in H1(T1) sense and we can detect the unique limit of uϵT1uϵdx, as we show now.

    Proposition 7.8. Let ˜uH1(T1) be determined by

    ˜ux=(2πcos(2πx))+χ{1/4<x<1/2}(2πcos(2πx))+χ{1/2<x<3/4},

    and ˜u(0)=0. Then ˜u is the unique minimizer of T1umdx over all regular weak solutions u of (2.2).

    Proof. Let u be any regular weak solution of (2.2). Because the quantities

    T1um dx is invariant by addition of a constant to u, we can assume u(0)=˜u(0)=0, without loss of generality. Moreover, because of (2.3) and by periodicity, we have u(x)=˜u(x)=0 for x[0,1/4][3/4,1]. Then,

    T1u dm=T1um dx(T1u dx)(T1m dx)=[1/4,3/4]um dxT1u dx=T1u dx.

    Hence, it suffices to discuss the quantities of T1u dx.

    Because of (2.4), we can see that ˜u(x)u(x) in x[1/4,1/2]. On the other hand, it holds that ˜u(x)u(x) in x[1/2,3/4]. Indeed, suppose that there exists x0[1/2,3/4] and solution u0 such that ˜u(x0)<u0(x0). Then, it follows from (2.4) that u0(3/4)>˜u(3/4)=0, which is a contradiction. Thus, ˜u(x)u(x) for xT1 and we see T1˜u dxT1u dx.

    The authors thank anonymous referees for their careful reading of our manuscript and their many insightful comments and suggestions.



    [1] A convergence result for the ergodic problem for Hamilton-Jacobi equations with Neumann-type boundary conditions. Proc. Roy. Soc. Edinburgh Sect. A (2016) 146: 225-242.
    [2] Error estimates for the approximation of the effective Hamiltonian. Appl. Math. Optim. (2008) 57: 30-57.
    [3] Mean field games systems of first order. ESAIM Control Optim. Calc. Var. (2015) 21: 690-722.
    [4] Long time behavior of the master equation in mean field game theory. Anal. PDE (2019) 12: 1397-1453.
    [5] Convergence of the solutions of the discounted Hamilton-Jacobi equation: Convergence of the discounted solutions. Invent. Math. (2016) 206: 29-55.
    [6] (1969) Foundations of Modern Analysis. New York-London: Enlarged and Corrected Printing, Pure and Applied Mathematics, 10-I, Academic Press.
    [7] First-order, stationary mean-field games with congestion. Nonlinear Analysis (2018) 173: 37-74.
    [8] D. Evangelista and D. Gomes, On the existence of solutions for stationary mean-field games with congestion, J. Dyn. Diff. Equ., (2016), 1–24. doi: 10.1007/s10884-017-9615-1
    [9] Some new PDE methods for weak KAM theory. Calculus of Variations and Partial Differential Equations (2003) 17: 159-177.
    [10] Effective Hamiltonians and averaging for Hamiltonian dynamics. I. Arch. Ration. Mech. Anal. (2001) 157: 1-33.
    [11] Effective Hamiltonians and averaging for Hamiltonian dynamics II. Arch. Ration. Mech. Anal. (2002) 161: 271-305.
    [12] Solutions KAM faibles conjuguées et barrières de Peierls. C. R. Acad. Sci. Paris Sér. I Math. (1997) 325: 649-652.
    [13] Théorème KAM faible et théorie de Mather sur les systèmes lagrangiens. C. R. Acad. Sci. Paris Sér. I Math. (1997) 324: 1043-1046.
    [14] Orbite hétéroclines et ensemble de Peierls. C. R. Acad. Sci. Paris Sér. I Math. (1998) 326: 1213-1216.
    [15] Sur la convergence du semi-groupe de Lax-Oleinik. C. R. Acad. Sci. Paris Sér. I Math. (1998) 327: 267-270.
    [16] Existence of weak solutions to stationary mean-field games through variational inequalities. SIAM J. Math. Anal. (2018) 50: 5969-6006.
    [17] R. Ferreira, D. Gomes and T. Tada, Existence of weak solutions to first-order stationary mean-field games with Dirichlet conditions, To appear in Proc. Amer. Math. Society, 2018. doi: 10.1090/proc/14475
    [18] D. Gilbarg and N. S. Trudinger, Elliptic Partial Differential Equations of Second Order, Classics in Mathematics, Springer-Verlag, Berlin, 2001.
    [19] Generalized Mather problem and selection principles for viscosity solutions and Mather measures. Adv. Calc. Var. (2008) 1: 291-307.
    [20] Existence for stationary mean-field games with congestion and quadratic Hamiltonians. NoDEA Nonlinear Differential Equations Appl. (2015) 22: 1897-1910.
    [21] The selection problem for discounted Hamilton-Jacobi equations: Some non-convex cases. J. Math. Soc. Japan (2018) 70: 345-364.
    [22] D. Gomes, L. Nurbekyan and M. Prazeres, Explicit solutions of one-dimensional, first-order, stationary mean-field games with congestion, 2016 IEEE 55th Conference on Decision and Control, CDC 2016, (2016), 4534–4539. doi: 10.1007/s13235-017-0223-9
    [23] D. Gomes, L. Nurbekyan and M. Prazeres, One-dimensional stationary mean-field games with local coupling, Dyn. Games and Applications, (2017). doi: 10.1007/s13235-017-0223-9
    [24] On the existence of classical solutions for stationary extended mean field games. Nonlinear Anal. (2014) 99: 49-79.
    [25] A stochastic Evans-Aronsson problem. Trans. Amer. Math. Soc. (2014) 366: 903-929.
    [26] The vanishing discount problem and viscosity Mather measures. Part 1: The problem on a torus. J. Math. Pures Appl. (9) (2017) 108: 125-149.
    [27] The vanishing discount problem and viscosity Mather measures. Part 2: Boundary value problems. J. Math. Pures Appl. (9) (2017) 108: 261-305.
    [28] Jeux à champ moyen. I. Le cas stationnaire. C. R. Math. Acad. Sci. Paris (2006) 343: 619-625.
    [29] Mean field games. Jpn. J. Math. (2007) 2: 229-260.
    [30] P.-L. Lions, Collège de France course on mean-field games, 2007-2011.
    [31] P.-L. Lions, G. Papanicolao and S. R. S. Varadhan, Homogeneization of Hamilton-Jacobi equations, Preliminary Version, (1988).
    [32] On the minimizing measures of Lagrangian dynamical systems. Nonlinearity (1992) 5: 623-638.
    [33] Action minimizing invariant measure for positive definite Lagrangian systems. Math. Z (1991) 207: 169-207.
    [34] Selection problems for a discount degenerate viscous Hamilton-Jacobi equation. Adv. Math. (2017) 306: 684-703.
    [35] Regularity for second-order stationary mean-field games. Indiana Univ. Math. J. (2017) 66: 1-22.
  • This article has been cited by:

    1. Xiaotian Hu, Kaizhi Wang, Existence of solutions to contact mean-field games of first order, 2022, 22, 2169-0375, 289, 10.1515/ans-2022-0012
    2. Qinbo Chen, Convergence of solutions of Hamilton–Jacobi equations depending nonlinearly on the unknown function, 2023, 16, 1864-8258, 45, 10.1515/acv-2020-0089
    3. Nikiforos Mimikos-Stamatopoulos, Sebastian Munoz, Regularity and Long Time Behavior of One-Dimensional First-Order Mean Field Games and the Planning Problem, 2024, 56, 0036-1410, 43, 10.1137/23M1547779
    4. Martino Bardi, Hicham Kouhkouh, Long-Time Behavior of Deterministic Mean Field Games with Nonmonotone Interactions, 2024, 56, 0036-1410, 5079, 10.1137/23M1608100
    5. Tigran Bakaryan, Giuseppe Di Fazio, Diogo A. Gomes, $$C^{1,\alpha }$$ regularity for stationary mean-field games with logarithmic coupling, 2024, 31, 1021-9722, 10.1007/s00030-024-00976-x
    6. Panrui Ni, Time periodic solutions of first order mean field games from the perspective of Mather theory, 2024, 412, 00220396, 881, 10.1016/j.jde.2024.09.006
    7. Cristian Mendico, Differential N-players game: Nash equilibria and Mather measures, 2023, 40, 0265-0754, 192, 10.1093/imamci/dnad006
    8. Jules Berry, Olivier Ley, Francisco J. Silva, Approximation and perturbations of stable solutions to a stationary mean field game system, 2025, 00217824, 103666, 10.1016/j.matpur.2025.103666
  • Reader Comments
  • © 2020 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(2146) PDF downloads(301) Cited by(8)

Figures and Tables

Figures(2)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog