
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
[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
Problem 1. Let
{ϵuϵ+12|Duϵ|2+V(x)=g(mϵ)in Td,ϵmϵ−div(mϵDuϵ)=ϵin Td. | (1.1) |
We say that
Problem 2. With
{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
∫Tdudx=0. |
According to [16] (also see [17]), Problem 2 admits weak solutions under suitable polynomial growth conditions of
For Hamilton-Jacobi equations, the discounted problem corresponds to the following control problem. Let
J(x;v)=∫∞0e−ϵtL(x(t),˙x(t))dt, |
for a given Lagrangian,
uϵ(x)=infvJ(x;v), |
where the infimum is taken over
The Hamiltonian,
H(x,p)=supv∈Rd−p⋅v−L(x,v). |
Under standard coercivity and convexity assumptions on
ϵuϵ+H(x,Duϵ)=0in Td. |
For coercive Hamiltonians, the results in [31] give that
H(x,Du)=˜Hin Td, | (1.3) |
where the unknowns are
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,
ˆJ(x)=∫∞0e−ϵt[12|˙x(t)|2−V(x(t))+g(m(x(t),t))]dt. |
Now, we suppose that the value function,
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.
Assumption 2. There exist constants
g′(z)⩾C1zβ, |
g(z)⩽C2+12zg(z). |
Assumption 3.
For
An example that satisfies the preceding assumptions is the following:
g(m)=mα(α>0),V(x)=csin(2πx)(0<c<1/2), |
where
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
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
uϵ−∫Tduϵdx→uin C2,α(Td),mϵ→min 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
uϵ−ˉH/ϵ∼u+λ+ϵv,mϵ∼m+ϵθ | (1.4) |
for the solution of Problem 1. Using this expansion in (1.1), assuming that
Problem 3. Let
{λ+u+Du⋅Dv=g′(m)θin Td,−div(mDv)−div(θDu)=1−min Td. | (1.5) |
Remark 1.3. The normalization condition
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
limϵ→0‖uϵ−ˉHϵ−u−λ‖∞+‖mϵ−m‖∞=0. |
Remark 1.5. The preceding theorem remains valid if we replace Assumption 2 with the weaker condition that for any
g′(z)>γ(z0) |
for all
In the last section of the paper, Section 7, we investigate the asymptotic behavior of
Assumption 4. There exist positive constants,
c1mα−1⩽g′(m)⩽c2mα−1 |
for all
Remark 1.6. From the preceding hypothesis, we obtain that there exist positive constants
˜c1mα−C⩽g(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
1.
2.
3.
4.
5.
Moreover,
−ϵuϵ−|Duϵ|22−V(x)+g(mϵ)⩾0 | (1.6) |
in the sense of distributions, with
(−ϵuϵ−|Duϵ|22−V(x)+g(mϵ))mϵ=0, | (1.7) |
almost everywhere. Furthermore,
ϵmϵ−div(mϵDuϵ)=ϵ, | (1.8) |
in the sense of distributions and almost everywhere.
We note that in [16], the specific form of Hamiltonian
¯H−|Du|22−V(x)+g(m)⩾0 | (1.9) |
in the sense of distributions, with
(¯H−|Du|22−V(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
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
∫Td(g(mϵ)−g(m))(mϵ−m)dx→0 | (1.12) |
and
∫Td⟨ˉu⟩mdx⩽∫Td⟨u⟩mdx, | (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,
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
{12|Du1|2−12|Du2|2+ˉH1−ˉH2=g(m1)−g(m2)−div(m1Du1)+div(m2Du2)=0. |
Now, we multiply the first equation by
∫Td(m1−m2)(g(m1)−g(m2))+12(m1+m2)|Du1−Du2|2dx=0. | (2.1) |
Accordingly,
Here, we show two distinct regular weak solutions to (1.2). In the example below, the existence of a unique smooth solutions fails and
Let
{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),
m(x)=(πcos(2πx))+. |
The preceding expression vanishes in an interval, as can be seen in Figure 1.
On the other hand, from the first equation in (2.2), we see that
ux≡0a.e.on{0<x<1/4}∪{3/4<x<1}, | (2.3) |
and
|ux|2⩽−2πcos(2πx)a.e.on{1/4<x<3/4}. | (2.4) |
For example, We set the functions
˜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
In this section, we establish preliminary a priori estimates for solutions of Problem 1. To simplify the notation, we denote by
Proposition 3.1. Let
∫Tdmdx=1. | (3.1) |
Proof. First, we show the positivity. Suppose that
ϵm(x0)−Dm(x0)Du(x0)−m(x0)Δu(x0)=ϵ. |
However, the left-hand side is
ϵ∫Tdmdx=ϵ. |
Thus, we get the conclusion.
Next, we get a uniform lower bound for
Proposition 3.2. Suppose that Assumption 1 holds. Then, there exists a constant,
‖ϵu‖L∞(Td)+‖1m‖L∞(Td)⩽C. |
Proof. First, we bound
m(˜x)=ϵϵ−Δu(˜x). |
Since
ϵu(˜x)⩾g(1)−maxx∈TdV(x). | (3.2) |
Next, let
ϵu(ˆx)⩽g(1)+minx∈TdV(x). |
Thus,
Now, we address the lower bound for
g(m(x))=ϵu(x)+12|Du(x)|2+V(x)⩾ϵu(˜x)+minx∈TdV(x)⩾g(1)−oscV. |
Using Assumption 1, we get the lower bound for
In the following Lemma, we give an upper bound for
Lemma 3.3. Suppose that Assumptions 1 and 2 hold. Then, there exists a constant,
∫Td1+m2|Du|2+mβ+2dx⩽C. | (3.3) |
Proof. First, we multiply the first equation in (1.1) by
∫Td1+m2|Du|2+mg(m)dx=∫Td(m−1)V+g(m)dx. | (3.4) |
Using Assumption 2, we get
∫Tdmg(m)dx⩽C. |
On the other hand, in light of Proposition 3.2, there exists
g(t)⩾Cβ+1tβ+1−Cβ+1m0β+1+g(m0). |
Therefore, combining the preceding inequalities with (3.4), we obtain (3.3).
In the next proposition, we establish that
Proposition 3.4. Suppose that Assumptions 1 and 2 hold. Then, there exists a constant
‖Du‖L∞(Td)+‖m‖L∞(Td)⩽C. |
Proof. At first, we show the case
{ϵumxx+12mxxu2x+mxxV=g(m)mxxϵmuxx−uxx(mux)x=ϵuxx. |
Next, we subtract these equations and integrate by parts to get
∫Tmu2xx+g′(m)m2xdx=∫TmxVxdx⩽δ∫Tm2xdx+14δ∫TV2xdx, |
using a weighted Cauchy-Schwarz inequality with
Next, we discuss the case
∫Tdϵmdiv(mpDu) dx=∫Tddiv(mDu)div(mpDu) dx. | (3.5) |
Differentiating the first equation in (1.1), we get
∑juxjuxjxi=g′(m)mxi−ϵuxi−Vxi. | (3.6) |
Next, we rewrite the right-hand side of (3.5) as follows:
∫Tddiv(mDu)div(mpDu)=∑i,j∫Td(muxi)xi(mpuxj)xj dx=∑i,j∫Td(muxi)xj(mpuxj)xi dx=∑i,j∫Tdpmp−1(uximxi)(uxjmxj)+pmpmxiuxjuxixjdx+∑i,j∫Tdmpuximxjuxixj+mp+1u2xixjdx=∑i,j∫Tdpmp−1(uximxi)(uxjmxj)+(p+1)mpmxiuxjuxixj+mp+1u2xixjdx=∫Tdpmp−1|Dm⋅Du|2+(p+1)g′(m)mp|Dm|2dx+∫Tdmp+1∑i,ju2xixj−(p+1)mp∑jmxj(ϵuxj+Vxj)dx, | (3.7) |
using (3.6) in the last line. Combining (3.7) and (3.5), we obtain
∫Tdpmp−1|Dm⋅Du|2+(p+1)g′(m)mp|Dm|2+mp+1∑i,ju2xixjdx=∫Tdϵmdiv(mpDu)+(p+1)mp∑jmxj(ϵuxj+Vxj)dx=∫TdϵpmpDm⋅Du+(p+1)mpDm⋅DVdx⩽∫Tdϵp2(mp+1+mp−1|Dm⋅Du|2)+(p+1)[δ|Dm⋅DV|2mp+β+Cδmp−β]dx, |
where the last inequality follows from a weighted Cauchy inequality with
∫Tdmp+β|Dm|2dx⩽C∫Tdmp+1⩽C∫Tdmp+β+2dx, | (3.8) |
where the last inequality is a consequence of
For
[∫Td(mp+β+22)2∗dx]12∗⩽C[∫Tdmp+β+2+|Dmp+β+22|2dx]1/2⩽C(1+|p+β+2|)[∫Tdmp+β+2dx]1/2. |
Thus, there exists a positive constant
‖m‖L2∗(q+1)2⩽[C(1+q)]2q+1‖m‖Lq+1. |
Next, we take
rnrn+1=αn+1−αn2∗/2. |
By Hölder's inequality and the above estimate with
‖m‖rn+1⩽‖m‖αnrn‖m‖1−αn2∗rn/2⩽‖m‖αnrn{(Crn)2rn‖m‖rn}1−αn=(Crn)2(1−αn)rn‖m‖rn. |
Iterating the prior inequality, we get
‖m‖rn+1⩽‖m‖r0∏ni=0(Cri)2(1−αi)ri. |
The right-hand side is bounded uniformly in
log(∏ni=0(Cri)2(1−αi)ri)⩽∑ni=02ri[C+log(ri)]<+∞. |
Hence,
When
Proposition 3.5. Suppose that Assumption 1 holds. Let
‖ux‖C1,α(T)+‖m‖C1,α(T)⩽C. |
Proof. Differentiating the first equation in (1.1)and multiplying by
ϵmux+uxmuxx+mVx=g′(m)mmx. |
Solving the second equation in (1.1) for
mx=2ϵmux−ϵux+mVx.(u2x+g′(m)m). | (3.9) |
Because
Now, we obtain additional estimates for the solutions of (1.1) in the case
m=g−1(ϵu+12|Du|2+V). |
Next, replacing the resulting expression into the second equation in (1.1), we obtain
div[g−1(ϵu+|Du|22+V)Du]−ϵ[g−1(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
(aijvxj)xi=ϕxk+ψxi, | (4.2) |
where
aij(x)=g−1(ϵu+12|Du|2+V)δij+(g−1)′(ϵu+12|Du|2+V)uxiuxj, |
ϕ(x)=ϵ[g−1(ϵu+12|Du|2+V)−1], |
ψ(x)=(g−1)′(ϵu+12|Du|2+V)(ϵuxiuxk+uxiVxk), |
and
‖ϕ‖∞+‖ψ‖∞⩽C. | (4.3) |
Moreover, using again Propositions 3.2 and 3.4, we see that there exists a constant
λ|ξ|2⩽∑di,j=1aij(x)ξiξj⩽1λ|ξ|2. | (4.4) |
Next, we prove that
Proposition 4.1. Suppose that Assumptions 1 and 2 hold. Let
‖Du‖C1,α(Td)⩽C. |
Proof. Take
(aijzxj)xi=ϕxk+ψxi, | (4.5) |
in
(aijwxj)xi=0, | (4.6) |
in
We begin by establishing the following claim.
Claim 1. For
‖z‖L∞(B2R)⩽CRfor anyR>0. |
Remark 4.2. If
Let
∫B2Raijzxj(z−k)+xidx=∫B2Rϕ(z−k)+xk+ψ(z−k)+xidx. | (4.7) |
Set
A(k)={z>k}∩B2R. |
It suffices to prove that we can choose a constant
∫A(k)aijzxjzxidx=∫A(k)ϕzxk+ψzxidx. |
In view of (4.4) and the bounds in (4.3), we get
∫A(k)|Dz|2dx⩽C|A(k)|. |
Next, using Sobolev's inequality and taking into account that
(h−k)2|A(h)|2/2∗⩽[∫A(h)[(z−k)+]2∗dx]2/2∗⩽[∫B2R[(z−k)+]2∗dx]2/2∗⩽C∫B2R|D(z−k)+|2dx⩽C∫A(k)|Dz|2dx. |
Combining the two preceding estimates, we obtain
|A(h)|⩽C|A(k)|2∗2(h−k)2∗. |
Next, we take a sequence
M=(C|A(0)|2∗2−12(2∗)22∗−2)12∗. |
Using the above estimate, we obtain
|A(kn+1)|⩽C(kn+1−kn)2∗|A(kn)|2∗/2⩽C22∗(n+1)M2∗|A(kn)|2∗/2. |
We now prove by induction that
|A(kn)|⩽|A(0)|2−nμ, |
where
|A(kn+1)|⩽C22∗(n+1)M2∗(|A(0)|2−nμ)2∗/2=|A(0)|2−(n+1)μ, |
using our choice for
Finally, by considering the limit
M=ˆC|A(0)|1/d⩽ˆC|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
Claim 2. We have
osc(R2,w)⩽ηosc(R,w), |
for some constant
osc(R,w):=supBRw−infBRw. |
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
osc(R,v)⩽CRα. |
Set
Mn=sup14n+1⩽R⩽14nosc(R,v)Rα, |
where
Here, we prove by induction that there exist
Mn⩽Mμ−n | (4.10) |
for some sufficiently large
1<μ<min(41−α,14αη), | (4.11) |
and then we choose
[ˆC(μ41−α)n+1+η4αMμ]⩽M. | (4.12) |
The prior choice of
For
Mn+1⩽[ˆC(μ41−α)n+1+η4αMμ]μ−(n+1). |
Using the defining property of
Finally, for
osc(R,v)⩽(4ˆC+η4αM)Rα, |
which establishes the claim.
Claim 4.
‖Du‖C1,α(Td)⩽C. |
Due to Claim 3, we get
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
η1(x)={10<x⩽12m00x⩾m0. |
Now, we consider
{ϵuϵ+12|Duϵ|2+V(x)=f(mϵ)in Td,ϵmϵ−div(mϵDuϵ)=ϵin Td, | (5.1) |
where
f(m):=η(m)h(m)+{1−η(m)}g(m). |
Because
div{f−1(ϵu+|Du|22+V)Du}−ϵ{f−1(ϵu+|Du|22+V)−1}=0. | (5.2) |
We check that
f(z)⩽η(z){C2+12zh(z)}+{1−η(z)}{C2+12zg(z)}=C2+12zf(z). |
Finally, since
f′(z)=η′(z){h(z)−g(z)}+η(z)h′(z)+{1−η(z)}g′(z)⩾η(z)h′(z)+{1−η(z)}g′(z)⩾C1zβ. |
Thus, we obtain the estimates in Section 3 and 4 for (5.1) and (5.2). In the case
Remark 5.1. Replacing
Remark 5.2. Without Assumption 3, we need to modify
Here, we show the existence of classical solution to (5.2), using the continuation method. We begin by defining an operator,
J(u,λ)=div[f−1(ϵu+|Du|22+λV)Du]−ϵ[f−1(ϵu+|Du|22+λV)−1]. |
We set
Λ={λ∈[0,1]:∃uλ∈C2,α(Td):J(uλ,λ)=0}. | (5.3) |
We claim that
Proposition 5.3. Consider the setting of Problem 1 and suppose that Assumptions 1-3 hold. Let
Proof. Fix a sequence
Now, we show that
mλ=f−1(ϵuλ+|Duλ|22+λV). |
We consider the linearization of
Lλ(ϕ)=∂J∂μ(uλ+μϕ,λ)|μ=0=div[mλDϕ+(f−1)′(f(mλ))(ϵϕ+Duλ⋅Dϕ)Duλ]−ϵ(f−1)′(f(mλ))(ϵϕ+Duλ⋅Dϕ). | (5.4) |
Lemma 5.4. Consider the setting of Problem 1. Let
Proof. We must prove that for any
div[mλDϕ+(f−1)′(f(mλ))(ϵϕ+Duλ⋅Dϕ)Duλ]−ϵ(f−1)′(f(mλ))(ϵϕ+Duλ⋅Dϕ)=ξ | (5.5) |
has a unique solution,
Bλ[v,w]=∫TdDw[mλDv+(f−1)′(f(mλ))(ϵv+Duλ⋅Dv)Duλ]+ϵ∫Tdw(f−1)′(f(mλ))(ϵv+Duλ⋅Dv)dx. |
Note that if
Claim 1. There exists a constant,
We establish this claim by contradiction. For that, suppose that there exists
Bλ[vn,vn]=(Avn,vn)→0. |
Next, we have
Bλ[vn,vn]=∫Tdmλ|Dvn|2+(f−1)′(f(mλ)){ϵvn+Duλ⋅Dvn}2dx. |
Since
Claim 2. The range of
Take a sequence
Next, to establish the last part of the claim, suppose that
0=(Av,v)=Bλ[v,v]=∫Tdmλ|Dv|2+(f−1)′(f(mλ)){ϵv+Duλ⋅Dv}2dx. |
This contradicts
Claim 3. (5.5) has a unique solution
To prove this last claim, we define a bounded linear functional,
T(w)=∫Tdξwdx. |
The Riesz Representation Theorem guarantees that there exists a unique
Bλ[ϕ,w]=(Aϕ,w)E=(ˆw,w)E=T(w); |
that is,
To finish the proof of Theorem 1.1 we verify that
Proposition 5.5. Consider the setting of Problem 1 and suppose that Assumptions 1-3 hold. Let
Proof. By the preceding lemma, Lemma 5.4, we can apply the implicit function theorem (see [6]) to the operator
By combining the previous results, we prove Theorem 1.1 as follows.
Proof of Theorem 1.1. Since
mϵ(x)⩾f−1(f(1)−oscV)=f−1(g(1)−oscV)⩾g−1(g(1)−oscV)>m0. |
Because
The identity (2.1) gives that (1.1) has a unique classical solution. Indeed, let
{ϵ(uϵ1−uϵ2)+12|Duϵ1|2−12|Duϵ2|2=g(mϵ1)−g(mϵ2)ϵ(mϵ1−mϵ2)−div(mϵ1Duϵ1)+div(mϵ2Duϵ2)=0. |
Now, we multiply the first equation by
∫Td(mϵ1−mϵ2)(g(mϵ1)−g(mϵ2))+12(mϵ1+mϵ2)|Duϵ1−Duϵ2|2dx=0. |
Accordingly,
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
Now, we investigate the asymptotic behavior of
First, to address Problem 3, we consider the linearized discounted problem that we state now.
Problem 4. Let
{ϵvϵ+u+Du⋅Dvϵ=g′(m)θϵ+Ain Td,ϵθϵ−div(mDvϵ)−div(θϵDu)=1−m+div(B)in Td. | (6.1) |
Proposition 6.1. Suppose Assumption 2 holds.
Then, Problem 4 has a unique weak solution
Proof. Because
θϵ=ϵvϵ+u+Du⋅Dvϵ−Ag′(m). | (6.2) |
Using the previous expression for
−div(mDvϵ+ϵvϵ+Du⋅Dvϵg′(m)Du)+ϵ(ϵvϵ+Du⋅Dvϵ)g′(m)=1−m+div(B)+div(u−Ag′(m)Du)−ϵ(u−A)g′(m). | (6.3) |
Therefore, it suffices to show that (6.3) has a weak solution. For that, we define a bilinear form,
K[ϕ1,ϕ2]=∫TdmDϕ1⋅Dϕ2+ϵϕ1+Du⋅Dϕ1g′(m)Du⋅Dϕ2+ϵϕ2(ϵϕ1+Du⋅Dϕ1)g′(m)dx. |
Because
K[ϕ,ϕ]=∫Tdm|Dϕ|2+(Du⋅Dϕ+ϵϕ)2g′(m)dx. |
Hence,
Proposition 6.2.
Let
‖ϵvϵ‖L2(Td)+‖θϵ‖L2(Td)+‖Dvϵ‖L2(Td)⩽C(‖A‖L2(Td)+‖B‖L2(Td)+1). |
Proof. We multiply the first equation in (6.1) by
uθϵ+θϵDu⋅Dvϵ+vϵdiv(mDvϵ)+vϵdiv(θϵDu)=g′(m)|θϵ|2−(1−m)vϵ−vϵdiv(B)+Aθϵ. |
Integrating by parts, we obtain
∫Tdg′(m)|θϵ|2+m|Dvϵ|2 dx=∫Td(u−A)θϵ+(1−m)vϵ+vϵdiv(B) dx. |
Using Poincaré's inequality, we conclude that
∫Tdm|Dvϵ|2 dx⩽∫Td(u−A)θϵ+(1−m)vϵ+vϵdiv(B) dx=∫Td(u−A)θϵ+(1−m)(vϵ−∫vϵ dx)+vϵdiv(B) dx⩽‖u‖L2(Td)‖θϵ‖L2(Td)+‖A‖L2(Td)‖θϵ‖L2(Td)+‖1−m‖L2(Td)‖Dvϵ‖L2(Td)+‖B‖L2(Td)‖Dvϵ‖L2(Td). |
Hence, taking into account that
‖Dvϵ‖2L2(Td)⩽C(‖θϵ‖L2(Td)+‖A‖L2(Td)‖θϵ‖L2(Td)+‖B‖2L2(Td)). | (6.4) |
Arguing analogously, we obtain
∫Tdg′(m)|θϵ|2 dx⩽‖u‖L2(Td)‖θϵ‖L2(Td)+‖A‖L2(Td)‖θϵ‖L2(Td)+‖1−m‖L2(Td)‖Dvϵ‖L2(Td)+‖B‖L2(Td)‖Dvϵ‖L2(Td). |
Hence, by (6.4),
‖θϵ‖2L2(Td)⩽C(‖A‖2L2(Td)+‖B‖2L2(Td)+1). |
Combining the preceding inequality with (6.4), we have the estimate
‖Dvϵ‖2L2(Td)⩽C(‖A‖2L2(Td)+‖B‖2L2(Td)+1). |
Finally, the first equation in (6.1) yields
‖ϵvϵ‖2L2(Td)⩽C(‖A‖2L2(Td)+‖B‖2L2(Td)+1). |
Next, we bootstrap higher regularity for
Proposition 6.3.
Let
‖z‖Hk(Td)⩽Ck(1+‖A‖Hk(Td)+‖B‖Hk(Td)). |
Proof. We begin by rewriting (6.3) as
−div(mDvϵ+Du⋅Dvϵg′(m)Du)=−div(ϵvϵg′(m)Du)−ϵ(ϵvϵ+Du⋅Dvϵ)g′(m)+1−m+div(B)+div(u−Ag′(m)Du)−ϵ(u−A)g′(m). |
Next, we fix
(aijzxj)xi=ϕxh+ψixi, | (6.5) |
where
aij=δijm+uxiuxjg′(m), |
ψi=−mxhvxi−Du⋅Dvϵg′(m)uxixh+Du⋅Dvϵg′(m)2g″(m)uximxh+Dvϵ⋅Duxhg′(m)uxi, |
and
ϕ=−div(ϵvϵg′(m)Du)−ϵ(ϵvϵ+Du⋅Dvϵ)g′(m)+1−m+div(B)+div(u−Ag′(m)Du)−ϵ(u−A)g′(m). |
By the previous proposition, we know that
‖z‖L2(Td)⩽C(‖A‖L2(Td)+‖B‖L2(Td)+1). |
Furthermore, we have the estimates
‖ϕ‖L2(Td)⩽C(‖A‖H1(Td)+‖B‖H1(Td)+1), |
and
‖ψ‖L2(Td)⩽C(‖A‖L2(Td)+‖B‖L2(Td)+1). |
Let
∫Tdaijzxjzxidx=∫Td−ϕzxh+−hzxidx. |
Because of the uniform ellipticity of
∫Td|Dz|2dx⩽C(‖ϕ‖L2(Td)+‖h‖L2(Td))‖Dz‖L2(Td). |
Hence,
‖Dz‖L2(Td)⩽C(‖ϕ‖L2(Td)+‖h‖L2(Td))⩽C(1+‖A‖H1(Td)+‖B‖H1(Td)). |
Therefore,
‖ϕ‖H1(Td)⩽C(‖A‖H2(Td)+‖B‖H2(Td)+1), |
and
‖ψ‖H1(Td)⩽C(‖A‖H1(Td)+‖B‖H1(Td)+1). |
The conclusion follows by iterating this argument for higher derivatives.
Proposition 6.4.
Let
‖ϵvϵ‖Hk(Td)+‖θϵ‖Hk(Td)⩽Ck(1+‖A‖Hk(Td)+‖B‖Hk(Td)). |
In particular,
Proof. Differentiating the first equation in (6.1), we get
‖Dθ‖L2(Td)⩽C(1+‖A‖H1(Td)+‖B‖H1(Td)). |
The above implies
‖ϵvϵ‖H1(Td)⩽C(1+‖A‖H1(Td)+‖B‖H1(Td)). |
Iterating the preceding steps, we get the result.
Proposition 6.5. For
‖A‖L∞(Td)+‖B‖L∞(Td)⩽Cϵ |
for some constant
limϵ→0(‖ϵvϵ−λ‖L∞(Td)+‖θϵ−θ‖Hk(Td)+‖(vϵ−∫Tdvϵdx)−v‖Hk(Td))=0. |
Proof. By the previous estimates on the solutions of Problem 4, we can choose a subsequence such that
Finally, we present the proof of Theorem 1.4.
Proof of Theorem 1.4. Fix
Ek={(v,θ)∈Hk+1(Td)×Hk(Td),‖ϵv‖Hk(Td)+‖Dv‖Hk(Td)+‖θ‖Hk(Td)⩽^Ck}, |
where
A(x)=−ϵ2v−ϵ22|Dv|2+g(m+ϵθ)−g(m)−ϵg′(m)θϵ, |
and
B(x)=ϵ2θDvϵ. |
Because,
‖A‖Hk(Td)+‖B‖Hk(Td)⩽C^Ck2ϵ, |
we obtain
‖ϵv‖Hk(Td)+‖Dv‖Hk(Td)+‖θ‖Hk(Td)⩽Ck(‖A‖Hk(Td)+‖B‖Hk(Td)+1)⩽Ck(1+^Ck2ϵ). |
We can choose
‖uϵ−ˉHϵ−u−λ‖∞=‖ϵvϵ−λ‖∞→0. |
Now, we investigate the behavior of
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
Proof. Properties 1 and 3 in Theorem 1.7 are immediate; that is,
m⩾0,∫Tdmdx=1,‖uϵ−∫Tduϵdx‖W1,2(Td)⩽C. |
From Property 2, we conclude that, through a subequence
‖mα+12‖W1,2(Td)⩽C. |
Next, we examine Property 5 in Theorem 1.7. By compactness,
∫Tdφ(mϵ)α+12Duϵdx→∫Tdφmα+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,
∫Tdϵ(mϵ−1)φdx→0, |
we have
∫TdmϵDφDuϵdx→0. |
Because of Rellich-Kondrachov theorem
∫Td(mϵ)qα+12dx→∫Tdmqα+12dx, |
for any
∫TdmDφDudx=0. |
Next, we select a smooth non-negative function,
∫Td(ϵuϵ−V(x))φdx→∫Td(−¯H−V(x))φdx. |
Moreover, by convexity
lim infϵ→0∫Td|Duϵ|22φdx⩾∫Td|Du|22φdx. |
Finally, we observe that
∫Tdg(mϵ)−g(m)dx=∫Td∫10g′(smϵ+(1−s)m)(mϵ−m)dx. |
For any
‖g′(smϵ+(1−s)m)‖Lp⩽C(‖mϵ‖Lp(α−1)+‖m‖Lp(α−1))⩽C. |
Therefore, since
∫Tdg(mϵ)−g(m)dx→0. |
Proposition 7.2. Suppose that Assumption 4 holds. Let
∫Td|Duϵ|2mϵ+mϵg(mϵ)dx⩽C. | (7.1) |
Proof. Because of properties 1 and 2 in Theorem 1.7, Assumption 4 implies that
∫Tdmϵg(mϵ)dx⩽C. |
Then, (7.1) follows by integrating (1.7).
Remark 7.3. If Assumptions 4 holds, a similar estimate holds for regular weak solutions
∫Td|Du|2m+mg(m)dx⩽C. |
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
{ϵuϵ+12|Duϵ|2+Wϵ(x)=0in Td,ϵmϵ−div(mϵDuϵ)=ϵin Td, |
where
{12|Du|2+W(x)=0in Td,−div(mDu)=0in Td, |
where
Proposition 7.4. Suppose that Assumption 4 holds. Let
∫Td×Rdϕ(x,v)dμϵ=∫Tdϕ(x,−Duϵ)mϵdx |
for all
∫Td×Rd(−ϵφ(x)+vDxφ(x))dμϵ=−ϵ∫Tdφdx | (7.2) |
for all
∫Td×RdLϵ(x,v)dμϵ=ϵ∫Tduϵdx. | (7.3) |
Proof. Because (1.8) holds in the sense of distributions, the discounted holonomy condition for
∫Td|f|2(1+mϵ)dx<∞, |
is a Hilbert space. Moreover, if
ηδ∗f→f |
in
Duϵ∈L21+mϵ(Td). |
Accordingly, because of (7.2), we have
∫Td×Rd−ϵ(ηδ∗uϵ)+vDx(ηδ∗uϵ) dμϵ=−ϵ∫Td(ηδ∗uϵ)dx. |
Taking the limit
∫Td×Rd−ϵuϵ+vDxuϵ dμϵ=−ϵ∫Tduϵdx. | (7.4) |
Taking into account the definition of
∫Td×RdLϵ(x,v) dμϵ=∫Td×Rd12|v|2−Wϵ(x) dμϵ=∫Td(12|Duϵ|2−Wϵ(x))mϵdx+∫Td×RdvDxuϵ−ϵuϵ dμϵ+ϵ∫Tduϵdx=ϵ∫Tduϵdx. |
Remark 7.5. From (7.2) and (7.3), we conclude that
Similarly, for Problem 2, we have the following result.
Proposition 7.6. Suppose that Assumption 4 holds. Let
∫Td×Rdϕ(x,v)dμ=∫Tdϕ(x,−Du)mdx, |
for all
∫Td×RdvDxφ(x)dμ=0 |
for all
∫Td×RdL(x,v)dμ=∫Td(12|Du|2−W(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
Proof of Theorem 1.9. Let
For any
−v⋅p−Lϵ(x,v)⩽12|p|2+Wϵ(x), |
where
−v⋅D(ηδ∗u)−Lϵ(x,v)+W−Wϵ⩽12|D(ηδ∗u)|2+Wϵ+W−Wϵ=12|D(ηδ∗u)|2+W, | (7.5) |
where
∫Td×Rdv⋅D(ηδ∗u)−Lϵ(x,v)+W−Wϵdμϵ=−ϵ∫Td(ηδ∗u)mϵdx+ϵ∫Td(ηδ∗u)dx−ϵ∫Tduϵdx+∫Td(W−Wϵ)mϵdx=−ϵ∫Td(ηδ∗u)mϵdx+ϵ∫Td(ηδ∗u)dx−ϵ∫Tduϵdx+∫Td(W−Wϵ)mϵdx. | (7.6) |
Next, we integrate the right-hand side of (7.5) and use Jensen's inequality to obtain
∫Td×Rd12|D(ηδ∗u)|2+Wdμϵ⩽∫Td×Rd12ηδ∗(|Du|2)+Wdμϵ⩽∫Td×Rd−ηδ∗W+Wdμϵ, | (7.7) |
taking into account (1.9). Because
∫Td×Rd−ηδ∗W+Wdμϵ→0, |
as
−ϵ∫Tdumϵdx+ϵ∫Tdudx−ϵ∫Tduϵdx+∫Td(W−Wϵ)mϵdx⩽0. | (7.8) |
On the other hand, we observe that, for all
ϵuϵ+v⋅D(ηδ∗uϵ)−L(x,v)+Wϵ−W⩽ϵuϵ+12|D(ηδ∗uϵ)|2+W+Wϵ−W. |
Integrating with respect to
ϵ∫Tduϵmdx+∫Td(Wϵ−W)mdx⩽0. | (7.9) |
Next, from (7.8), we gather
−ϵ∫Td⟨u⟩mϵdx−ϵ∫Tduϵdx+∫Td(W−Wϵ)mϵdx⩽0. |
Finally, from (7.9), we get
ϵ∫Td⟨uϵ⟩mdx+ϵ∫Tduϵdx+∫Td(Wϵ−W)mdx⩽0. |
Adding the above two inequalities, we obtain
ϵ(∫Td⟨uϵ⟩mdx−∫Td⟨u⟩mϵdx)+∫Td(W−Wϵ)(mϵ−m)dx⩽0. | (7.10) |
By Sobolev's inequality, we have
∫Td⟨uϵ⟩mdx−∫Td⟨u⟩mϵdx⩽0. |
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
By (1.13), we get the following result:
Proposition 7.7. Let
∫T1⟨ˉu⟩mdx⩽∫T1⟨u⟩mdx. |
Using the above criterion, we can show that
Proposition 7.8. Let
˜ux=√(−2πcos(2πx))+⋅χ{1/4<x<1/2}−√(−2πcos(2πx))+⋅χ{1/2<x<3/4}, |
and
Proof. Let
∫T1⟨u⟩ dm=∫T1um dx−(∫T1u dx)(∫T1m dx)=∫[1/4,3/4]um dx−∫T1u dx=−∫T1u dx. |
Hence, it suffices to discuss the quantities of
Because of (2.4), we can see that
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. ![]() |
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 |