1. Introduction
The Mean Field Games (MFG in short) theory concerns the study of differential games with a large number of rational, indistinguishable agents and the characterization of the corresponding Nash equilibria. In the original model introduced in [
1,
2], an agent can typically act on its velocity (or other first order dynamical quantities) via a control variable. Mean Field Games where agents control the acceleration have been recently proposed in [
3,
4,
5].
A prototype of stochastic process involving acceleration is given by the Langevin diffusion process, which can be formally defined as
where
is the second time derivative of the stochastic process
X,
B a Brownian motion and
a positive parameter. The solution of (
1) can be rewritten as a Markov process
solving
The probability density function of the previous process satisfies the kinetic Fokker–Planck equation
The previous equation, in the case
, was first studied by Kolmogorov [
6] who provided an explicit formula for its fundamental solution. Then considered by Hörmander [
7] as motivating example for the general theory of the hypoelliptic operators (see also [
8,
9,
10]).
We consider a Mean Field Games model where the dynamics of the single agent is given by a controlled Langevin diffusion process, i.e.,
for
. In (
2), the control law
, which is a progressively measurable process with respect to a fixed filtered probability space such that
, is chosen to
maximize the functional
where
is the distribution of the agents at time
s. Let
u the value function associated with the previous control problem, i.e.,
where
is the the set of the control laws. Formally, the couple
satisfies the MFG system (see Section 4.1 in [
3] for more details)
for
. The first equation is a backward Hamilton–Jacobi–Bellman equation, degenerate in the
x-variable and with a quadratic Hamiltonian in the
v variable, and the second equation is forward kinetic Fokker–Planck equation. In the standard setting, MFG systems with quadratic Hamiltonians has been extensively considered in literature both as a reference model for the general theory and also since, thanks to the Hopf-Cole change of variable, the nonlinear Hamilton-Jacobi-Bellman equation can be transformed into a linear equation, allowing to use all the tools developed for this type of problem (see for example [
2,
11,
12,
13,
14,
15]). Recently, a similar procedure has been used for ergodic hypoelliptic MFG with quadratic cost in [
16] and for a flocking model involving kinetic equations in Section 4.7.3 of [
17].
We study (
3) by means of a change of variable introduced in [
11,
14] for the standard case. By defining the new unknowns
and
, the system (
3) is transformed into a system of two kinetic Fokker–Planck equations
for
. In the previous problem, the coupling between the two equations is only in the source terms. Following [
14], we prove existence of a weak solution to (
4) by showing the convergence of an iterative scheme defined, starting from
, by solving alternatively the backward problem
and the forward one
We show that the resulting sequence
,
, monotonically converges to the solution of (
4). Hence, by the inverse change of variable (see again [
11,
14] for details)
we obtain a solution of the original problem (
3). We have
Theorem 1. The sequence defined by (5) and (6) converges in and a.e. to a weak solution of (4). Moreover, the couple defined by (7) is a weak solution to (3). The main difficulty in the study of problems (
3) and (
4) is due both in the degeneracy of the second order operator with respect to
x and in the unbounded dependence of the coefficients of the first order terms with respect to
v. To overcome the previous difficulties we rely on the results for linear kinetic Fokker–Planck equations developed in [
18]. We mention that existence of weak solutions for the standard MFG problem, possibly degenerate, has been studied in [
19], but the results in this paper do not cover the present setting. The previous iterative procedure also suggests a monotone numerical method for the approximation of (
4), hence for (
3). Indeed, by approximating (
5) and (
6) by finite differences and solving alternatively the resulting discrete equations, we obtain an approximation of the sequence
. A corresponding procedure for the standard quadratic MFG system was studied in [
14], where the convergence of the method is proved. We plan to study the properties of the previous numerical procedure in a future work.
2. Well Posedness of the Kinetic Fokker–Planck System
In this section, we study the existence of a solution to system (
4). The proof of the result follows the strategy implemented in Section 2 of [
14] for the case of a standard MFG system with quadratic Hamiltonian and relies on the results for linear kinetic Fokker–Planck equations in Appendix A of [
18]. We remark the model here studied does not fit exactly the problem treated in [
18] because of the presence of a zero order term in the Fokker-Planck equation. Hence some technical aspects should be analyzed in more detail, however the present paper is mainly intended to give some idea on the change of variabile for the kinetic MGF.
We fix the assumptions we will assume in the whole paper. The vector field
and the coupling cost
are assumed to satisfy
Moreover, the diffusion coefficient
is positive and the initial and terminal data satisfy
and
Note that (
9) implies that
. We denote with
the scalar product in
and with
the pairing between
and its dual
. We define the following functional space
and we set
. If
, then it admits (continuous) trace values
,
(see [
18], Lemma A.1) and therefore the initial/terminal conditions for (
4) are well defined in
sense. We first prove the well posedness of problems (
5) and (
6).
- (i)
For any , there exists a unique solution to Moreover, and, for any , there exist and such that - (ii)
Let be the map which associates to ψ the unique solution of (10). Then, if , we have .
Proof. We first prove existence of a solution to the nonlinear problem (
10) by a fixed point argument exploiting the results for the corresponding linear problem proved in [
18]. Fixed
, consider the map
from
into itself that associates with
the weak solution
of the linear problem
By Prop. A.2 of [
18],
belongs to
and it coincides with the unique solution of (
12) in this space. Moreover, the following estimate
holds for some constant
C which depends only on
,
and
. Hence
F maps
, the closed ball of radius
C of
, into itself.
To show that the map
F is continuous on
, consider
such that
and set
. Then
, and, by the estimate (
13), we get that, up to a subsequence, there exists
such that
,
in
,
in
. Moreover,
almost everywhere. By the definition of weak solution to (
12), we have that
for any
, the space of infinite differentiable functions with compact support in
. Employing weak convergence for left hand side of (
14) and the Dominated Convergence Theorem for the right hand one, we get for
for any
. Hence
and
for
in
. The compactness of the map
F in
follows by the compactness of the set of the solutions to (
12), see Theorem 1.2 of [
20]. We conclude, by Schauder’s Theorem, that there exists a fixed-point of the map
F in
, hence in
, and therefore a solution to the nonlinear parabolic Equation (
10).
Observe that, if
is a solution of (
10), then
is a solution of
with the corresponding final condition. In the following, we assume that
. To show that
is non-negative, we will exploit the following property (see Lemma A.3 of [
18]): given
and defined
, then
and
Let
be a solution of (
15), multiply the equation by
and integrate. Then, since
is non-negative, by (
16) we get
where it has been exploited that, by integration by parts,
. Since
and therefore
we get
, hence
.
To prove the uniqueness of the solution to (
10), consider two solutions
,
of (
15) and set
. Multiplying the equation for
by
, integrating and using
, we get
and, by the strict monotonicity of
f, we conclude that
.
To prove that
is bounded from above, we observe that the function
, where
as in (
9), is a supersolution of the linear problem (
12) for any
, i.e.,
and
By the Maximum Principle (see Prop. A.3 (i) in [
18]), we get that
, where
is the solution of (
12). Since the previous property holds for any
, we conclude that
, where
is the solution of the nonlinear problem (
10).
A similar argument show that
, where
as in (
9) and
sufficiently large, is a subsolution of (
12) for any
. Indeed, replacing
in the equation, we get that the inequality
is satisfied for
large enough and, moreover,
. Hence
, where
is the solution of the nonlinear problem (
10), and, from this estimate, we deduce (
11).
We finally prove the monotonicity of the map
. Set
,
, and consider the equation satisfied by
, multiply it by
and integrate. Performing a computation similar to (
17), we get
Since, by monotonicity of
f and non-negativity of
, we have
we get
and therefore
. □
We set
where
is defined as in (
11).
Proposition 3. Given , where as in (8), we have - (i)
For any , there exists a unique solution to - (ii)
Let be the map which associates with the unique solution of (18). Then, if , we have .
Proof. First observe that, since
, then
is well defined for
. The proof of the first part of
is very similar to the one of the corresponding result in Proposition 2, hence we only prove the bound (
19). If
is a solution of (
18), then
is a solution of
Let
be a solution of (
20), set
and observe that
. Multiply the equation for
by
and integrate to obtain
Since
and
, we have
and therefore
. Hence the upper bound (
19).
Now we prove
(ii). Set
,
and
. Multiply the equation satisfied by
by
and integrate. Since, by monotonicity and negativity of
f, we have
Hence and therefore . □
Proof of Theorem 1. Given
, consider the sequence
,
, defined in (
5) and (
6). It can rewritten as
where the maps
,
are as in Propositions 2 and, respectively 3. Observe that, by (
11), we have
for
and
for any
k. Hence the sequence
is well defined. We first prove by induction the monotonicity of the components of
. By non-negativity of solutions to (
18), we have
and therefore
. Moreover, by the monotonicity of
,
. Now assume that
. Then
and
therefore the monotonicity of two sequences.
Since
and, by (
19), for
, the sequence
,
converges a.e. and in
to a couple
. Taking into account the estimate (
13), the a.e. convergence of the two sequences and repeating an argument similar to the one employed for the continuity of the map
F in Proposition 2, we get that the couple
satisfies, in weak sense, the first two equations in (
4). The terminal condition for
is obviously satisfied, while the initial condition for
, in
sense, follows by convergence of
to
.
We now consider the couple
given by the change of variable in (
7). We first observe that, by Theorem 1.5 of [
10], we have
,
,
and a corresponding regularity for
. Taking into account the boundedness of
and the estimate in (
11), we have that
u,
,
,
. Hence we can write the equation for
u in weak form, i.e.,
for any
, with final datum in trace sense. In a similar way, since
m,
,
,
and
m is locally bounded, we can rewrite also the equation for
m in weak form, i.e.,
for any
with the initial datum in trace sense. □