1,118
Views
1
CrossRef citations to date
0
Altmetric
Articles

A variational approach to first order kinetic mean field games with local couplings

&
Pages 1945-2022 | Received 23 Dec 2021, Accepted 09 Jul 2022, Published online: 12 Aug 2022

Abstract

First order kinetic mean field games formally describe the Nash equilibria of deterministic differential games where agents control their acceleration, asymptotically in the limit as the number of agents tends to infinity. The known results for the well-posedness theory of mean field games with control on the acceleration assume either that the running and final costs are regularizing functionals of the density variable, or the presence of noise, i.e. a second-order system. In this article we construct global in time weak solutions to a first order mean field games system involving kinetic transport operators, where the costs are local (hence non-regularizing) functions of the density variable with polynomial growth. We show the uniqueness of these solutions on the support of the agent density. This is achieved by characterizing solutions through two convex optimization problems in duality. As part of our approach, we develop tools for the analysis of mean field games on a non-compact domain by variational methods. We introduce a notion of ‘reachable set’, built from the initial measure, that allows us to work with initial measures with or without compact support. In this way we are able to obtain crucial estimates on minimizing sequences for merely bounded and continuous initial measures. These are then carefully combined with L1-type averaging lemmas from kinetic theory to obtain pre-compactness for the minimizing sequence. Finally, under stronger convexity and monotonicity assumptions on the data, we prove higher order Sobolev estimates of the solutions.

2020 MATHEMATICS SUBJECT CLASSIFICATION:

1. Introduction

The aim of the theory of mean field games (MFG for short) is to characterize limits of Nash equilibria of stochastic or deterministic differential games when the number of agents tends to infinity. Such models were first proposed about 15 years ago, simultaneously by Lasry–Lions [Citation1–3] and Huang-Malhamé-Caines [Citation4].

This theory turned out to be extremely rich in applications and it provided excellent mathematical questions. Its literature has witnessed a huge increase in the last decade. From the theoretical viewpoint, there are two main approaches to the study of MFG. One is based on analytical and PDE techniques, while the other is a probabilistic approach. The first approach goes back to the original works of Lasry-Lions and has been extended in a great variety of directions in the subsequent years by many authors. If a non-degenerate idiosyncratic noise is present in the models, this typically yields a parabolic structure for the corresponding PDEs and one can expect (strong) classical solutions or a suitable regularity for weak solutions to the corresponding PDE systems, even when the corresponding Lagrangians are local functions of the density variable. For a non-exhaustive list of works in this direction we refer the reader to [Citation5–12]. The probabilistic approach proved to be equally successful for problems involving Lagrangians that are nonlocal functions of the measure variable. This approach seems to be very powerful for handling different kinds of noises in combination with the non-degenerate idiosyncratic one, such as the common noise. For a non-exhaustive collection of works in this direction we refer to [Citation13–16].

When the model lacks a non-degenerate idiosyncratic noise, this clearly poses technical difficulties in the analysis. Typically, it means that additional structural assumptions need to be imposed on the data to be able to hope for (weak) solutions. Such conditions are, for instance, suitable notions of convexity/monotonicity (cf. [Citation17,Citation18]), or the presence of a suitable variational structure, as in the case of potential games ([Citation19–24]). In the case of local couplings, it was pointed out by Lions in [Citation25] that the MFG system (including the planning problem) can be transformed into a degenerate elliptic system in space-time with oblique boundary conditions. Relying on this idea, in a quite general setting, under suitable assumptions on the data (such as strict monotonicity and strong convexity of the Hamiltonians in the measure and momentum variables, respectively; regularity and positivity conditions on the initial data), it has been proven recently in [Citation26,Citation27] that the corresponding first order MFG systems have smooth classical solutions.

For an excellent, relatively complete account on the subject and a summary of results to date we refer the reader to the collection [Citation17].

In this article we study a class of first order kinetic MFG systems, involving Lagrangians that are local functions of the density variable and that possess a variational structure, in the sense of [Citation19–21].

In our setting, the MFG system can be formally written as (1.1) {tu(t,x,v)v·Dxu(t,x,v)+H(x,v,Dvu(t,x,v))=f(x,v,m),in(0,T)×M×Rd,tm(t,x,v)+v·Dxm(t,x,v)divv(mDpH(x,v,Dvu(t,x,v)))=0,in(0,T)×M×Rd,m(0,x,v)=m0(x,v),u(T,x,v)=g(x,v,mT),inM×Rd.(1.1)

Here M denotes either that d-dimensional flat torus Td or the whole d-dimensional Euclidean space Rd and is the physical space for the position x of the agents, while the velocity vector v of the agents lies in Rd. T > 0 is an arbitrary time horizon, H:M×Rd×RdR is the Hamiltonian function, while f,g:M×Rd×RR stand for the running and final costs of the agents, respectively.

Under suitable assumptions on the data, we obtain the global in time existence, uniqueness and Sobolev regularity of weak solutions to Equation(1.1), relying on two convex optimization problems in duality. One of these problems can be seen as an optimal control problem for the Hamilton-Jacobi equation, while its dual is an optimal control problem for the continuity equation (cf. [Citation19–21]).

Review of the literature in connection to our work

MFG systems of type Equation(1.1) have been introduced in the context of models when agents control their acceleration. It seems that such a model can be traced back to the work [Citation28] (in the engineering community), where the authors proposed a MFG model where agents control their acceleration. In the mathematical community, the first works in this framework seem to be the ones [Citation29–31]. These works consider Hamiltonians (with our notation Hf) and final cost functions that are nonlocal regularizing functions in the measure variable. Moreover, the Hamiltonians need to be either purely quadratic or have quadratic growth in the momentum variable. In addition, in [Citation29,Citation30] further conditions on the initial measure m0 are also imposed. In [Citation29] m0 is taken to be compactly supported and Hölder continuous, while in [Citation30] m0 is taken to be compactly supported. These two works construct weak solutions to the corresponding MFG system in the sense that the Hamilton-Jacobi equation has to be understood in the viscosity sense, while the continuity equation is understood in the sense of distributions. In [Citation31] the initial measure m0 can be quite general and the corresponding Hamiltonian does not need to have the so-called ‘separable structure’ which was assumed in [Citation29,Citation30] and is also assumed in this article. These more general hypotheses come at the price of obtaining a weaker notion of solution to the MFG system: the so-called mild solutions. However, the authors show that, under the additional separability assumption on the Hamiltonian, mild solutions become more standard weak solutions in the sense described above.

Several interesting new works are built on the models introduced in [Citation29–31]. In [Citation32] the authors study the ergodic behavior of MFG systems, for the case of Hamiltonians that are purely quadratic in the momentum variable and nonlocal regularizing coupling functions f, g, with additional growth assumption on f in the v variable. In [Citation33] the authors obtain mild solutions to MFG under acceleration control and state constraints, under assumptions similar to the ones in [Citation29] on the Hamiltonians, with the possibility to consider Hamiltonians that are power-like functions in the momentum variable. Lastly, in [Citation34] the author studies a perturbation problem associated to MFG under acceleration control, where the (Lagrangian) cost associated to the acceleration vanishes.

MFG models with degenerate diffusion share some common features with kinetic type problems. In this context we can mention several works. In [Citation35] and [Citation36] the authors study time independent MFG systems with purely quadratic Hamiltonians and nonlocal regularizing coupling functions, where the diffusion operator is hypoelliptic or satisfies a suitable Hörmander condition. It is also worth mentioning that our system Equation(1.1) shares some similarities with MFG models where agents interact also through their velocities. In this direction we refer to the works [Citation37–41].

Finally, a second order MFG system of type Equation(1.1) has been recently studied in [Citation42]. In this work the author obtains weak and renormalized solutions (in the spirit of [Citation12]) to a MFG system that involves a non-degenerate diffusion in the v direction. This seems to be the only work in the context of kinetic type MFG models where the coupling functions f and g are taken to be local functions of the density variable m. Here the Hamiltonian H is assumed to depend only on the momentum variable and either to be globally Lipschitz continuous or to have quadratic/sub-quadratic growth. There are several summability properties and moment bounds imposed on the initial density m0. In the case of Lipschitz continuous Hamiltonians, the coupling functions f, g are supposed to fulfill several further assumptions: a strong uniform increasing property in the m variable and their derivatives in the (x, v) variable must have a linear growth condition in the m variable.

In [Citation42] the presence of the diffusion in the v direction allows the author to use suitable De Giorgi type arguments to show that the solution to the Fokker-Planck equation is bounded and has fractional Sobolev regularity. These estimates seem to be instrumental to set up a fixed point scheme and to show that the MFG system has a weak solution. Furthermore, the presence of this diffusion allows to obtain second order Sobolev estimates for the MFG system.

Description of our results

As highlighted above, in this work we are inspired by [Citation19–21] and we obtain existence and uniqueness of weak solutions to Equation(1.1) (in the sense of Definition 2.3) via two convex optimization problems in duality (Problem 3.1 and Problem 3.3). Compared to these works, several major differences arise which require new ideas. A first obvious difference is that in our setting (in contrast to the compact setting of the flat torus which is considered in the mentioned references) the velocity variable v lives in the non-compact space Rd. This clearly introduces technical issues in the analysis.

To prove our main results, the general outline of our programme is the same as the one of [Citation19–21]: prove the duality for Problem 3.1 and Problem 3.3; suitably relax Problem 3.1 (this will be Problem 3.8) and show that the value of this is the same as the original one; show existence of optimizers for the relaxed problem and apply the duality result again to obtain existence of solutions in a suitable weak sense. In this article H is supposed to have a superlinear growth in the momentum variable, and f and g are supposed to have polynomial growth in their last variables. The growth of f, g may be taken independently of the growth of the Hamiltonian (we refer to the next section for the precise assumptions).

To show that the value of the relaxed problem is the same as the original one, a standard approach used in [Citation19–21] is to test the Hamilton-Jacobi inequality of any competitor by competitors of the dual problem (i.e. solutions to the continuity equation). To justify this computation a mollification argument was applied for solutions to the continuity equation. In our case, this mollification alone is not enough because of the non-compact setting. Therefore a delicate cutoff argument has to be also implemented.

The most delicate part, however, is to obtain existence of optimizers to the relaxed problem and in particular to obtain proper compactness results for the minimizing sequences. First, in our case the time trace of the solutions to the Hamilton–Jacobi inequality constraint in Problem 3.8 is quite weak: u(t,·) has to be understood as a locally finite signed Radon measure. Since in this work m0 may have non-compact support, it takes additional effort to give a meaning to M×Rdm0u0(dxdv) (a term that appears in the objective functional present in Problem 3.8). Our construction, although completely different, has some similarities in spirit with the one in [Citation43], to define similar time boundary traces.

In order to obtain suitable estimates for the minimizing sequence of the relaxed problem, in [Citation19–21] a typical trick was to test the Hamilton-Jacobi inequality constraint by the initial measure m0. For this reason, it was necessary to impose enough regularity, and more importantly a uniform positive lower bound of this density everywhere. Because of this, estimates on the quantity Tdm0u0dx, would readily yield summability estimates on u0 solely. We emphasize that in this article we assume that m0 is merely a bounded and continuous probability density and so we take a completely different route when obtaining such estimates. We introduce the reachable set Um0, a set of points in time, space and velocity that can be reached from spt(m0) with arbitrary smooth admissible controls (cf. Definition 2.2). In fact, by the controllability of the underlying ODE system, which satisfies the Kalman rank condition, we have Um0=({0}×spt(m0))((0,T)×M×Rd). In order to obtain our crucial estimates on the corresponding minimizing sequence we use well chosen test functions that are supported in Um0. This construction seems to be new in the literature on variational MFG and we believe that it could be instrumental also in other settings, to possibly relax regularity, positivity or compact support assumptions on m0.

As there is no Hopf–Lax type representation formula available for solutions to our Hamilton-Jacobi equations (which was the case in [Citation19,Citation20]), first, we obtain estimates on truncations of the solutions. These are similar in flavor to the corresponding estimates in [Citation21], and such ideas date back to [Citation44]. As our terminal data typically have merely local summability, this will be the source of additional technical issues (in contrast to [Citation21], where the terminal data was taken to be regular enough).

Let us underline that the ideas and constructions that we have described so far allow us to obtain summability estimates on u and Dvu, using the structure of the problem. This is not sufficient to yield weak precompactness for minimizing sequences due to the lack of regularity estimates in x. To recover the necessary compactness we make use of averaging lemmas available in kinetic theory. Averaging lemmas go back to the works [Citation45,Citation46] and provide improved regularity and compactness properties for velocity averages of solutions of kinetic transport equations (see Subsection Citation6.Citation1 for the precise definitions). For more details and a survey of results we refer the reader to the review [Citation47] and the references cited therein. When regularity with respect to v is additionally available, similar properties can be deduced for the full density function: we refer for instance to [Citation48] for regularity results in the Lp case for 1<p<+. We carefully tailor this approach to our setting, combining our estimates on Dvu with L1 averaging lemmas [Citation49–51] to deduce precompactness for minimizing sequences. In this way we prove Theorem 6.8 on the existence of a minimizer of Problem 3.8. This in turn implies Theorem 2.4, that system Equation(1.1) has a (unique) weak solution. As was similarly obtained in [Citation19–21], we show the uniqueness of m and the uniqueness of u on {m>0}.

A natural question that arises in the context of variational MFG is whether the variational structure and further strong monotonicity and convexity assumptions on the data would yield higher order Sobolev estimates on weak solutions. Such estimates were recently obtained in more classical frameworks in [Citation23,Citation24, Citation39, Citation52–54]. In this article we pursue similar Sobolev estimates, implied by taking stronger assumptions on the data. In comparison with the works [Citation23,Citation24], in our setting we need to work with a considerably weaker notion of time trace of u, which is not stable under perturbations of the initial measure m0. Therefore, our Sobolev estimates remain local in time on (0,T]. Another delicate difference is due to the presence of the kinetic transport term. Because of this, a careful choice of perturbations need to be used, which take into account the kinetic nature of the problem. As a result of this, interestingly, first we obtain estimates on differential operators of the form (tDx+Dv) applied to m and Dvu. For the precise results in this direction we refer to Theorem 8.2, Corollary 8.4 and Corollary 8.5.

The structure of the article is as follows. In Section 2 we state our standing assumptions and main results. In Section 3 we present the two variational problems in duality along with the relaxed problem of the primal problem. In Section 4 we have collected some preliminary estimates on weak solutions of the Hamilton-Jacobi inequality obtained on the reachable set Um0. In Section 5 we show that the relaxed problem has the same value as the primal problem and hence the duality result holds. Section 6 contains the existence result of a solution to the relaxed problem. Here we rely on the combination of the estimates derived in the previous sections and suitably tailored averaging lemmas from kinetic theory, applied in our context for distributional subsolutions to kinetic Hamilton-Jacobi equations. In Section 7 we show that optimizers of the variational problems in duality provide weak solutions to the MFG system and, conversely, weak solutions are also optimizers of the variational problems. Furthermore, strong convexity yields (partial) uniqueness of these solutions. Section 8 is devoted to the derivation of higher order Sobolev estimates for the weak solutions. These require further assumptions on the data.

We end the paper with two appendix sections. In Appendix A we discuss the time regularity of distributional subsolutions to kinetic Hamilton-Jacobi equations which allow us to construct suitable notions of time traces. Finally, in Appendix B we show that truncations and maxima of distributional subsolutions to kinetic Hamilton-Jacobi equations remain distributional subsolutions to suitably modified equations.

2. Standing assumptions and main results

In this section we state our main results on the existence, uniqueness and Sobolev regularity of solutions to the MFG system.

We define F and G to be the anti-derivatives of the coupling functions f and g with respect to m: F(x,v,m)=0mf(x,v,m)dmG(x,v,m)=0mg(x,v,m)dm.

Throughout, we make the following assumptions on the Hamiltonian and coupling functions.

Assumption 1.

  • (H1) The Hamiltonian H is continuous in all variables, and convex and differentiable with respect to p. Furthermore, for some r > 1, H satisfies bounds of the form

(2.1) 1cr|p|rCHH(x,v,p)cr|p|r+CH,(2.1)

for all (x,v,p)M×Rd×Rd and some constants c>0 and CH0. Finally, the function H0(x,v):=H(x,v,0) has positive part (H0)+C0(M×Rd), where C0(M×Rd) denotes the closure of the space Cc(M×Rd) with respect to the uniform norm.

  • (H2) F is continuous in all variables and strictly convex and differentiable with respect to m for m > 0. Moreover, it satisfies the growth condition

(2.2) 1cqmqCF(x,v)F(x,v,m)cqmq+CF(x,v),m0(2.2)

where q > 1 and the function CFL1(M×Rd). For m < 0, we set F(x,v,m)=+.

  • (H3) G is continuous and strictly convex. Moreover, it satisfies the growth condition

(2.3) 1cmsCG(x,v)G(x,v,m)cms+CG(x,v),m0,(2.3)

for some CGL1(M×Rd) and 1<sq. For m < 0, we set G(x,v,m)=+.

  • (H4) The initial datum m0Cb(M×Rd) is a probability density.

We note that since m0 is imposed to be a bounded probability density, by interpolation, it is uniformly bounded in Lα(M×Rd), for any α[1,+]. We emphasize that here we impose growth conditions on F,G rather than on f, g.

Example 2.1.

For any q > 1 and continuous bounded function c such that cc0 with c0>0 a strictly positive constant, the function F(x,v,m)={c(x,v)mqm0+m<0, satisfies the given assumptions.

Definition 2.2

(Reachable set). It will be useful to define the set Um0[0,T]×M×Rd to be the set of points potentially reachable by a collection of agents initially distributed according to m0 and evolving according to the control system ẋ=v,v̇=a, for some control aC([0,T];Rd). Observe that the previous control system satisfies the classical Kalman rank condition, and so we have Um0={0}×{m0>0}(0,T]×M×Rd.

Under these standing assumptions, we define the following notion of weak solution to the MFG system.

Definition 2.3.

We say that (u, m) is a weak solution to Equation(1.1), if the following are fulfilled:

  1. uLloc1(Um0),DvuLlocr(Um0) and m|Dvu|rL1((0,T)×M×Rd);

  2. mLq((0,T)×M×Rd) and mTLs(M×Rd);

  3. (u0)+(L+Lq)(M×Rd) and (u0) is a locally finite Radon measure supported in {m0>0}.

  4. {tuv·Dxu+H(x,v,Dvu)f(x,v,m), in D((0,T)×M×Rd)uTg(·,·,mT),in D(M×Rd).

  5. The continuity equation from Equation(1.1) holds in D((0,T)×M×Rd).

  6. M×Rdm0u0(dxdv) is finite.

  7. The following energy equality holds:(2.4) M×Rdm0u0(dxdv)M×Rdg(x,v,mT)mTdxdv=0TM×Rdf(x,v,m)mdxdvdt+0TM×Rd[DpvH(x,v,Dvu)·DvuH(x,v,Dvu)]mdxdvdt.(2.4)

2.1. Existence and uniqueness

The first of our main results is the existence and uniqueness of these weak solutions.

Theorem 2.4.

Let Assumption 1 hold. Then there exists a weak solution (u, m) of the mean field game system Equation(1.1) in the sense of Definition 2.3. This solution is unique, in the sense that if (u1, m1) and (u2, m2) are both weak solutions in the sense of Definition 2.3, then m1 = m2 almost everywhere and u1 = u2 almost everywhere on the set {m1>0}.

2.2. Regularity

Our second main result is Sobolev regularity for weak solutions of the mean field games system Equation(1.1). For this result we assume quadratic growth of the Hamiltonian (r = 2) and stronger convexity and regularity hypotheses on the data, as follows.

Assumption 2.

  • (H5) (Conditions on the coupling functions) There exists C > 0 such that the functions f, g satisfy

(2.5) |f(x1,v1,m)f(x2,v2,m)|C(mq1+1)(|x1x2|+|v1v2|)(x1,v1),(x2,v2)M×Rd,m0.(2.5) and(2.6) |g(x1,v1,m)g(x2,v2,m)|C(ms1+1)(|x1x2|+|v1v2|)(x1,v1),(x2,v2)M×Rd,m0.(2.6)

Moreover, there exists cf,cg>0 such that(2.7) (f(x,v,m˜)f(x,v,m))(m˜m)cfmin{m˜q2,mq2}|m˜m|2m˜,m0,m˜m.(2.7) (2.8) (g(x,v,m˜)g(x,v,m))(m˜m)cgmin{m˜s2,ms2}|m˜m|2m˜,m0,m˜m.(2.8)

In the above assumptions, if q < 2 or s < 2 one should interpret 0q2 and 0s2 as +. In this way, when m˜=0, for instance, Equation(2.7) reduces to f(x,v,m)mcfmq, as in the more regular case q2. Similar comments can be made for Equation(2.8).

  • (H6) (Quadratic growth and strong coercivity assumption on H) Suppose that r=2andthere exist j1,j2:RdRd and cH>0 such that

(2.10) H(x,v,P)+L(x,v,W)P·WcH|j1(P)j2(W)|2.(2.10)

In particular, and in light of our restriction Equation(2.1), we assume that j1 and j2 have linear growth.

  • (H7) L(·,·,W)C2(M×Rd)and|Dxx2L(x,v,W)|,|Dxv2L(x,v,W)|,|Dvv2L(x,v,W)|C0|W|2+C0,)(x,v,W)M×Rd×Rd.

Under these additional assumptions, we prove the following result. The proof is carried out in Section 8.

Theorem 2.5.

Suppose that (u, m) is a weak solution to Equation(1.1) in the sense of Definition 2.3 and that (H5), (H6), (H7) hold.

Then, there exists C¯>0 such thatmq21Dx,vmLloc2((0,T]×M×Rd)C¯,m1/2Dx,vDvuLloc2((0,T]×M×Rd)C¯andmTs21Dx,vmTL2(M×Rd)C¯.

Remark 2.6.

The estimates appearing in this statement are informal; we in fact obtain uniform L2-type summability of differential quotients (see estimate Equation(8.8) below). The corresponding Sobolev estimates, however, are more delicate to obtain, because these would need to be understood in the sense of weighted Sobolev spaces or more generally in the sense of Sobolev spaces with respect to measures. Their precise versions would need to involve tangent spaces with respect to the measure m, but these are beyond the scope of the current article. We refer to [Citation55] on this topic.

3. Variational problems in duality

We will prove existence of a solution to the MFG system Equation(1.1) through a variational characterization. In this section we set up the variational problems used to obtain solutions. We recall that here and throughout the rest of the article, we will work under Assumption 1.

3.1. Optimal control of the Hamilton-Jacobi equation: smooth setting

We define the Fenchel conjugates of F and G respectively by F(x,v,β):=supm0{βmF(x,v,m)}G(x,v,u):=supm0{umG(x,v,m)}.

Under our assumptions on F, we have the bounds (3.1) {c|β|qCF(x,v)F(x,v,β)c1|β|q+CF(x,v)β>0,F(x,v,0)=0F(x,v,β)infm0F(x,v,m)CF(x,v)β0,(3.1)

where q=q/(q1) denotes the Hölder conjugate exponent of q. Note also that F is non-decreasing. Similar observations hold for G.

Using this, we define the following functional: for uCb1([0,T]×M×Rd), let A(u):=0TM×RdF*(x,v,tuv·Dxu+H(x,v,Dvu))dxdvdtM×Rdu(0,x,v)m0(x,v)dxdv+M×RdG*(x,v,u(T,x,v))dxdv, whenever the integrals are meaningful, and set A(u)=+ otherwise. We define a first variational problem associated to this problem.

Problem 3.1.

Minimize A(u) over uE0, where E0 denotes the space (3.2) E0:={uCb1([0,T]×M×Rd):|v||Dxu|L([0,T]×M×Rd)}.(3.2)

Remark 3.2.

E0 is a Banach space when equipped with the norm uE0:=uL+DvuL+(1+|v|)DxuL

3.2. Optimal control of the continuity equation

To state the dual problem we define the Lagrangian L:M×R2dR, which is the Fenchel conjugate of the Hamiltonian H in the last variable. In other words, for any (x,v,α)M×R2d, we define L(x,v,α):=suppRd{α·pH(x,v,p)}.

Note that L then satisfies upper and lower bounds of the form 1cL|α|rCLL(x,v,α)cL|α|r+CL, where r=r/(r1) denotes the Hölder conjugate exponent of r.

For pairs (m,w)L1([0,T]×M×Rd)×L1([0,T]×M×Rd), we define the functional B(m,w):=0TM×RdF(x,v,m)dxdvdt+0TM×RdL(x,v,wm)mdxdvdt+M×RdG(x,v,mT(x,v))dxdv,

with the convention that L(x,v,wm)m={0m=w=0,+m=0,w0.

We then define a second variational problem, (formally) dual to the first.

Problem 3.3.

Minimize B(m,w) over the set KB of pairs (m,w)L1([0,T]×M×Rd)×(L1([0,T]×M×Rd))d with m0, subject to (m,w) satisfying the following continuity equation: (3.3) tm+v·Dxm+divvw=0,in D((0,T)×M×Rd)(3.3) and m|t=0=m0 in the sense of a weak trace.

Remark 3.4.

Let us comment on the weak trace of m with respect to the time variable. Since we are interested in competitors (m, w) for which B(m,w) is finite, there must exist a vector-valued measurable function VLr(mdxdvdt), that is, for which 0TM×Rd|V|rmdxdvdt<+, such that w=Vm (i.e. V is the density of w with respect to m). So, we notice that the previous equation can be written as tm+divx(vm)+divv(Vm)=0.

Since Vm=wL1([0,T]×M×Rd), we have |V|mL1([0,T]×M×Rd). We are then able to prove that m has a narrowly continuous representative [0,T]tmtP(M×Rd), so that in particular m|t=0 and m|t=T are meaningful. This is essentially a consequence of [Citation56, Lemma 8.1.2], with minor modifications to account for the fact that vm is only locally integrable; we sketch this in the appendix in EquationLemma A.12.

3.3. Duality

Lemma 3.5.

We have the following duality:infuE0A(u)=min(m,w)KBB(m,w).

Proof.

This is an application of the classical Fenchel–Rockafeller duality theorem. Recall that we defined the Banach space E0 above in Equation(3.2). Then let E1 be defined by E1:=Cb0([0,T]×M×Rd;R)×Cb0([0,T]×M×Rd;Rd); we will express elements of E1 as pairs (ϕ,ψ) of continuous bounded functions, where ϕ is real-valued and ψ is vector-valued. E1 is a Banach space with respect to the uniform norm. On these spaces we define the respective functionals A0(u):=M×Rdu(0,x,v)m0(x,v)dxdv+M×RdG*(x,v,u(T,x,v))dxdv and A1(ϕ,ψ):=0TM×RdF*(x,v,ϕ+H(x,v,ψ))dxdvdt.

Note that these functionals are convex. We also define the bounded linear map Λ:E0E1 by Λu:=(tu+v·Dxu,Dvu).

Then A(u)=A0(u)+A1(Λu).

We wish to apply Fenchel–Rockafeller duality. In order to do this we must verify the existence of uE0 such that A0(u),A1(Λu)<+ and A1 is continuous at Λu. For example, we may take u to be of the form u(t,x,v):=ζ(xvt,v)+2CH(tT), where CH denotes the constant from the bounds on the Hamiltonian Equation(2.1). We then take ζCb1(M×Rd) non-negative to have sufficiently strong decay at infinity so that (3.4) ζLs(M×Rd),Dx,vζLrq(M×Rd).(3.4)

Explicitly, for the case M=Td we may take for example ζ(x,v)=ζ(v)=(1+|v|2)k/2;k>max{ds,drq}, in which case |v||Dxu|=0 and therefore uE0.

For the case M=Rd we may take ζ(x,v)=ζ(v)=(1+|x|2+|v|2)k/2;k>max{2ds,2drq.}.

In this case, Dxu=Dxζ(xvt,v)=k(xvt)(1+|xvt|2+|v|2)1+k/2, and so |v||Dxu|=k|v||xvt|(1+|xvt|2+|v|2)1+k/2k2(1+|xvt|2+|v|2)k/2k2, which implies that uE0.

Then, in either case, tuv·Dxu+H(x,v,Dvu)=2CH+H(x,v,Dvu)c|Dvu|rCH

It follows that the positive part satisfies [tuv·Dxu+H(x,v,Dvu)]+c|Dvu|r1{c|Dvu|r>CH}Lq([0,T]×M×Rd). and thus by the bounds on F Equation(3.1) we obtain 0TM×RdF*(x,v,tuv·Dxu+H(x,v,Dvu))dxdvdt<+.

That is, A1(Λu) is finite.

Moreover, uTLs(M×Rd) and thus M×RdG*(x,v,u(T,x,v))dxdv<+.

Finally, since u0=ζ2CHT and m0 is a probability density, M×Rdu(0,x,v)m0(x,v)dxdv<+.

Thus A0(u) is finite.

Now we verify that A1 is continuous at Λu with respect to convergence in E1. Consider the sequence of pairs (ϕn,ψn)E1,nN, such that ϕn=tu+v·Dxu+δn,ψn=Dvu+εn, where (δn,εn)E1 satisfy (δn,εn)L2n. Then ϕn+H(x,v,ψn)=2CHδn+H(x,v,Dvu+εn).

Using the bounds Equation(2.1) on the Hamiltonian, we obtain ϕn+H(x,v,ψn)2CHδn+CH+cr|Dvu+εn|rCHδn+2r1cr|Dvu|r+2r1cr|εn|rCH+C(2n+2rn)+C|Dvu|r, for some constant C > 0. Therefore, for all n large enough that C(2n+2rn)<CH, for the positive part we have [ϕn+H(x,v,ψn)]+C|Dvu|r.

Then the bounds Equation(3.1) on F imply that F(ϕn+H(x,v,ψn))2CF+C|Dvu|rqL1([0,T]×M×Rd),

(where the constant C > 0 has changed line to line). The right hand side is in L1 because we constructed ζ to satisfy Equation(3.4). We may therefore use it as a dominating function: since (δn,εn) certainly converges to zero pointwise (in fact in uniform norm), and F is continuous with respect to the variable β, by dominated convergence we may conclude that A1[(ϕn,ψn)]=limn+0TM×RdF*(x,v,ϕn+H(x,v,ψn))dxdvdt=0TM×RdF*(x,v,tuv·Dxu+H(x,v,Dvu))dxdvdt=A1(Λu).

Thus A1 is indeed continuous at Λu.

It remains to check that A is bounded below on E0. Let uE0 and set β:=tuv·Dxu+H(x,v,Dvu). Then, using the growth assumptions on F* and G*, similarly to the inequality Equation(4.2) below, we have A(u)β+Lqq+u(T,,)+LssM×Rd[TCF(x,v)+CG(x,v)]dxdvM×Rdu(0,x,v)+m0(x,v)dxdvβ+Lqq+u(T,·,·)+LssC((uT)+Ls+CHT+T1/qβ+Lq)(m0L1+m0Lq)infa,b0{aqc0a+bsc0bc0}>, where c0 was set to be a large positive constant depending only on m0,T,CH,CF,CG.

Therefore, we are in position to apply the Fenchel–Rockafeller duality theorem (cf. [Citation57, Chapter Citation3, Theorem 4.1]), to conclude infuE0A(u)=max(m,w)E1{A0(Λ(m,w))A1((m,w))}.

Here E1 denotes the dual space of E1. By [Citation58, IV.6] the dual space of Cb0 may be identified with the space of bounded, regular, finitely additive set functions. Thus E1 is the space of pairs (m, w), where m is a real-valued regular finitely additive set function, and w is a Rd-valued regular finitely additive set function.

It remains to identify max(m,w)E1{A0(Λ(m,w))A1((m,w))}.

In what follows, we are going to show that the above maximization problem actually admits solutions in a better space than E1. So, we have max(m,w)E1{A0(Λ(m,w))A1((m,w))}=max(m,w)E˜1{A0(Λ(m,w))A1((m,w))}, where the set E˜1 stands for pairs (m, w) such that m is a finite Radon measure on [0,T]×M×Rd and w is a finite vector-valued Radon measure on [0,T]×M×Rd taking values in Rd. The proof of this is postponed to Lemma 3.6 below.

Then, by arguing as in [Citation19, Section 3.3], we may identify that max(m,w)E1{A0(Λ(m,w))A1((m,w))}=max(m,w)E1B(m,w) where the maximum is taken over (m,w)E1 such that (m,w)L1([0,T]×M×Rd)×L1([0,T]×M×Rd;Rd) and m0 almost everywhere, such that tm+v·Dxm+divvw=0in D((0,T)×M×Rd),m|t=0=m0.

Thus infuE0A(u)=min(m,w)KBB(m,w).

Lemma 3.6.

Using the notations and assumptions from Lemma 3.5, we havemax(m,w)E1{A0(Λ(m,w))A1((m,w))}=max(m,w)E˜1{A0(Λ(m,w))A1((m,w))},

Proof.

Observe that any pair (m,w)E1 induces functionals on Cc0 and (Cc0)d. Therefore, there exist a signed Radon measure m˜ with finite total variation and a finite vector-valued measure w˜ which coincide with, respectively, m and w on (the closure with respect to the uniform norm of) Cc0 and (Cc0)d. Then A1((m,w))=sup(ϕ,ψ)E1{m,ϕ+w,ψ0tM×RdF(x,v,ϕ+H(x,v,ψ))dxdv}

By considering functions of the form ϕ=lχ+H0, for H0(x,v):=H(x,v,0) (note that our assumptions on H imply in particular that H0Cb) and any non-negative χCb0 and l > 0, and ψ = 0, we find that A1((m,w))=+ unless m is a positive functional. Indeed, note that 0tM×RdF(x,v,lχ)dxdv0tM×Rdsupβ0F(x,v,β)dxdv<+, and supl>0m,lχ=+ if m,χ<0.

Next, by taking the supremum over the smaller set (ϕ,ψ)Cc0×(Cc0)d we have A1((m,w))sup(ϕ,ψ)Cc0×(Cc0)d{m,ϕ+H0+w,ψ0tM×RdF(x,v,ϕH0+H(x,v,ψ))dxdvdt}=sup(ϕ,ψ)Cc0×(Cc0)d{m˜,ϕ+H0+w˜,ψ0tM×RdF(x,v,ϕH0+H(x,v,ψ))dxdvdt}mm˜,H0.

Let us underline that the assumption on H0 plays a crucial role, otherwise the integral of F might not be finite for compactly supported test functions.

Since H is convex, for any χRCb0 such that 0χR1, H(x,v,χrψ)χRH(x,v,ψ)+(1χR)H0(x,v).

Thus ϕχRH0+H(x,v,χRψ)χR(ϕH0+H(x,v,ψ)), and in particular we can compare the positive parts: (ϕχRH0+H(x,v,χRψ))+(χR(ϕH0+H(x,v,ψ)))+.

Since F is non-decreasing, F(x,v,ϕχRH0+H(x,v,χRψ))supβ<0F(x,v,β)+F(x,v,ϕH0+H(x,v,ψ))L1.

Hence, for all ϕCb0,ψ(Cb0)d such that 0TM×RdF(x,v,ϕH0+H(x,v,ψ))dxdvdt<+,

by dominated convergence we have 0tM×RdF(x,v,ϕH0+H(x,v,ψ))dxdvdt=limR+0tM×RdF(x,v,ϕRH0+H(x,v,ψR))dxdvdt, where ϕR=ϕχR,ψR=ψχR for some continuous 0χR1 converging pointwise to the constant function 1 as R tends to positive infinity. We conclude that, for any m˜,w˜ (respectively signed, vector-valued) Radon measures with finite total variation, sup(ϕ,ψ)Cc0×(Cc0)d{m˜,ϕ+H0+w˜,ψ0tM×RdF(x,v,ϕH0+H(x,v,ψ))dxdvdt}=sup(ϕ,ψ)Cb0×(Cb0)d{m˜,ϕ+w˜,ψ0tM×RdF(x,v,ϕ+H(x,v,ψ))dxdvdt},

where we have used that H0 is also a Cb0 function in order to relabel ϕ. We have thus proved that A1((m,w))A1((m˜,w˜))mm˜,H0.

Next, note that if m(Cb0) is a positive functional with Radon measure part m˜, then mm˜ is also a positive functional: given 0ϕCb0, let 0χR1 be a sequence of continuous functions, non-decreasing with R and converging pointwise to the constant function 1 as R tends to positive infinity. Then, since 0ϕχRϕ, by dominated convergence and the positivity of m, m˜,ϕ=limR0m˜,ϕχR=limR0m,ϕχRm,ϕ.

Since (H0)+C0,mm˜,(H0)+=0 and thus mm˜,H00 for all m such that A1((m,w,)) is finite. Then A1((m,w))A1((m˜,w˜)).

We now consider A0. We assume from now on that m(Cb0) is a positive functional, since we only wish to consider (m, w) for which A1((m,w))<+. A0(Λ(m,w))=supuE0{Λ(m,w),uM×Rdu(0,x,v)m0(x,v)dxdvM×RdG*(x,v,u(T,x,v))dxdv}=supuE0{(m,w),ΛuM×Rdu(0,x,v)m0(x,v)dxdvM×RdG*(x,v,u(T,x,v))dxdv}.

Then, taking supremum over the smaller set uCc1, we have A0(Λ(m,w))supuCc1{(m,w),ΛuM×Rdu(0,x,v)m0(x,v)dxdvM×RdG*(x,v,u(T,x,v))dxdv}.

If uCc1, then ΛuCc0. Thus A0(Λ(m,w))supuCc1{(m˜,w˜),ΛuM×Rdu(0,x,v)m0(x,v)dxdvM×RdG*(x,v,u(T,x,v))dxdv}.

We show that the right hand side is in fact equal to A0(Λ(m˜,w˜)): given uE0, let χRCc1(M×Rd) be a sequence of cutoff functions such that 0χR1. We construct χR such that their support is contained in B¯2R(0)M×Rd, the closed ball of radius 2 R, χR=1 on B¯R(0), the closed ball of radius R, and χRCR for some constant C > 0 independent of R. Thus note in particular that χR1 and χR0 pointwise as R+. Let uR:=uχR. Then |ΛuR|=|tuR+v·DxuR|CuE0,|uR(0,x,v)||u(0,x,v)|,[uR(T,x,v)]+[u(T,x,v)]+.

Since m(Cb0), we have m˜,CuE0=CuE0m˜,1<+. Moreover u(0,·) is bounded and therefore integrable with respect to m0. Finally, note that G*(x,v,uR(T,x,v))G*(x,v,u(T,x,v))+supβT<0G(x,v,βT).

Hence, if M×RdG*(x,v,u(T,x,v))dxdv<+,

we may apply the dominated convergence theorem to find that (m˜,w˜),ΛuM×Rdu(0,x,v)m0(x,v)dxdvM×RdG*(x,v,u(T,x,v))dxdv=limR+(m˜,w˜),ΛuRM×RduR(0,x,v)m0(x,v)dxdvM×RdG*(x,v,uR(T,x,v))dxdv.

This completes the proof that the suprema over E0 and Cc1 are equal for the Radon measure parts. We conclude that A0(Λ(m,w))A0(Λ(m˜,w˜)).

Now observe that, since the set E˜1 is contained in E1 (it is precisely the set of Radon measure parts of elements of E1), max(m,w)E1{A0(Λ(m,w))A1((m,w))}sup(m,w)E1{A0(Λ(m˜,w˜))A1((m˜,w˜))}sup(m˜,w˜)E˜1{A0(Λ(m˜,w˜))A1((m˜,w˜))}sup(m,w)E1{A0(Λ(m,w))A1((m,w))}.

All of the above inequalities are therefore equalities. Moreover, since A0(Λ(m,w))A1((m,w))A0(Λ(m˜,w˜))A1((m˜,w˜)), if (m, w) attains the supremum then the same is true of the Radon measure part (m˜,w˜). Thus, without loss of generality, the optimizer is given by some (m,w)E˜1, i.e. a finite measure and a finite Rd-valued measure. □

Remark 3.7.

Let us notice that the minimizer of B(m,w) is unique (by the convexity of F,G and L in their last variables). Moreover, the growth conditions on F,G and L imply that mLq((0,T)×M×Rd),mTLs(M×Rd) and |w|rmr1L1((0,T)×M×Rd). Furthermore, by Hölder’s inequality, wLp((0,T)×M×Rd)), with p:=rqr+q1. These arguments are similar to the ones in [Citation20, Theorem 2.1] and [Citation19, Lemma 2]. Furthermore, the equation satisfied by m conserves mass, so that mLtLx,v1, and in fact mtLx,v1=m0Lx,v1 for all t[0,T].

3.4. The relaxed problem

The third problem we define is a relaxation of Problem 3.1. Consider the functional A˜(u,β,βT):=0TM×RdF*(x,v,β)dxdvdtM×Rdm0(x,v)u0(dxdv)+M×RdG*(βT)dxdv.

Problem 3.8.

Minimize A˜(u,β,βT) over the set KA of triples (u,β,βT)Lloc1(Um0)×Lloc1(Um0)×L1(M×Rd) satisfying

  • The positive part of u satisfies u+Lloc1([0,T]×M×Rd);

  • The positive part of β satisfies β+Lq([0,T]×M×Rd).

  • The positive part of βT satisfies (βT)+Ls(M×Rd);

  • DvuLlocr(Um0),

and subject to Equation(3.5), understood in the sense of Definition 3.9.

Definition 3.9.

We say that a triple (u,β,βT) that belongs to the spaces from Problem 3.8 is a weak distributional solution to (3.5) {tuv·Dxu+H(x,v,Dvu)β,in(0,T)×M×Rd,uTβT,inM×Rd,(3.5) if(3.6) 0TM×Rdu[tϕ+div(vϕ)]+ϕH(x,v,Dvu)dxdvdt0TM×Rdβϕdxdvdt+M×RdβTϕTdxdv,(3.6) for any ϕCc1((0,T]×M×Rd) nonnegative.

Remark 3.10.

  1. Let us emphasize that the weak form Equation(3.6) encodes both inequalities from Equation(3.5), as we show this in Lemma A.6.

  2. u0 is similarly understood as a certain notion of a trace at t = 0 in a weak sense. In particular, the term M×Rdm0(x,v)u0(dxdv)=u0,m0,

which appears in the definition of A˜ is to be understood as in EquationDefinition A.10. Moreover, we underline that this quantity is set to be +, if there exist ϕCc1({m0>0}) nonnegative such that u0,ϕ=.

4. The Hamilton–Jacobi equation

In this section, we analyze the Equationequation (3.5). We take the assumptions appropriate to the minimization problem we will consider. Therefore, we suppose throughout that (u,β,βT)KA is such that A˜(u,β,βT)<+. From the finiteness of the energy we deduce in particular that M×Rdu0m0<+.

4.1. Upper bounds

We prove upper bounds on u. First, we observe that for any constant lR the function (ul)+:=max{ul,0} satisfies (see Lemma B.1) (4.1) {(t+v·Dx)(ul)++H(x,v,Dvu)1{u>l}β1{u>l}, in D((0,T)×M×Rd),(uTl)+(βTl)+,in D(M×Rd).(4.1)

We use the notation L+Lq to denote the set of functions {h=h1+h2:h1L,h2Lq},

which becomes a Banach space when equipped with the norm hL+Lq:=inf{h1L+h2Lq:h=h1+h2}.

We also use the notation L1Lq to denote the intersection of L1 and Lq made into a Banach space under the norm hL1Lq:=max{hL1,hLq}.

Note that the dual space is given by (L1Lq)=L+Lq.

Lemma 4.1.

Let lR be given and let (u,β,βT)KA satisfy Equation(4.1).

  1. Then (ul)+Lt(L+Lq)x,v, with the a priori estimate(ul)+Lt(L+Lq)x,v(βTl)+(L+Lq)(M×Rd)+CHT+T1/qβ+Lq(βTl)+Ls(M×Rd)+CHT+T1/qβ+Lq.

  2. Suppose in addition that A˜(u,β,βT)<CA˜. Then, there existsC=C(CA˜,m0L1Lq,T,CF,CG)>0

such thatβ+Lq((0,T)×M×Rd)+(βT)+Ls(M×Rd)C.

Proof.

First, let us note that, since (βT)+Ls(M×Rd) and sq (by Assumption 1), (βT)+(L+Lq)(M×Rd) and thus also (βTl)+(L+Lq)(M×Rd).

  1. Let t[0,T) be fixed. Let 0ψCc(M×Rd) and consider ζ(τ,x,v):=ψ(x+(tτ)v,v).

Then ζ is smooth and compactly supported and satisfies τζ+v·Dxζ=0.

By using ζ as a test function for (ul)+ over τ[t,T], we obtain M×Rd(ul)+(t,x,v)ψ(x,v)dxdvM×Rd(βTl)+ζTdxdv+tTM×Rdζ[βH(x,v,Dvu)]1{u>l}dxdvdτ.

Recall that when we write (ul)+(t,·,·), we are always referring to the version of u that is weakly right continuous with respect to time (cf. Appendix A, Lemma A.1).

Since HCH, we have M×Rd(ul)+(t,x,v)ψ(x,v)dxdvM×Rd(βTl)+ζTdxdv+tTM×Rdζ[β++CH]dxdvdτ.

Then M×Rd(utl)+ψdxdv(βTl)+L+LqζTL1Lq+ζLq([t,T]×M×Rd)β+Lq([t,T]×M×Rd)+CHζL1([t,T]×M×Rd).

We compute ζL1([t,T]×M×Rd)=tTM×Rdψ(x+(tτ)v,v)dxdvdτ=tTM×Rdψ(x,v)dxdvdτ=(Tt)ψL1(M×Rd).

Similarly ζTL1(M×Rd)=ψL1(M×Rd) and ζLq([t,T]×M×Rd)=(Tt)1/qψLq(M×Rd).

Thus M×Rd(utl)+ψdxdv((βTl)+L+Lq+CHT+T1/qβ+Lq)ψL1Lq.

This extends by density to all non-negative ψ(L1Lq)(M×Rd), and general ψ(L1Lq)(M×Rd) by non-negativity of (ul)+. We conclude by the fact that (βTl)+L+Lq(βTl)+Ls. The result follows.

  1. By the definition of A˜ and the assumptions on the data one has (4.2) A˜(u,β,βT)β+Lqq+(βT)+LssM×Rd[TCF(x,v)+CG(x,v)]dxdvM×Rdu(0,x,v)+m0(x,v)dxdvβ+Lqq+(βT)+LssC((βT)+Ls+CHT+T1/qβ+Lq)m0L1Lq,(4.2)

where in the last inequality we used the estimate from (i). This further yields the claim in (ii). □

Corollary 4.2.

Let (u,β,βT) be as in the statement of Lemma 4.1 such that there exists CA˜>0 with A˜(u,β,βT)<CA˜. Then, there exists C=C(CA˜,m0L1Lq,T,CF,CG)>0 such that the following hold.

  1. (u0)+L+Lq(M×Rd)C;

  2. M×Rdm0(u0)(dxdv)C.

Proof.

We notice that (i) is a simple consequence of Lemma 4.1(i)-(ii), by setting l = 0 and t = 0 (in the sense of weak trace, given in EquationDefinition A.5).

For (ii), we observe M×Rdm0(u0)(dxdv)=M×Rdm0u0(dxdv)+M×Rd(u0)+m0dxdvA˜(u,β,βT)0TM×RdF*(x,v,β)dxdvdtM×RdG*(βT)dxdv+(u0)+L+Lqm0L1Lq.

By the bounds Equation(3.1) on F and the corresponding estimates for G, supβ{F(x,v,β)}CF(x,v),supβT{G(x,v,βT)}CG(x,v).

Hence, using the above bounds and (i), we obtain M×Rdm0(u0)(dxdv)CA˜+CFL1+CGL1+Cm0L1Lq, which completes the proof. □

4.2. Local L1 bounds

Next, we prove bounds on the negative parts of u and β. We will obtain Lloc1(Um0) bounds, by use of a duality argument involving a certain class of test functions which satisfy the continuity equation associated to the control system.

Lemma 4.3.

Let aCb1([0,T]×M×Rd;Rd) be a bounded control. Let ϕ0Cc1(M×Rd) satisfy 0ϕ0m0. Let ϕCc1(Um0) be the solution of the continuity equationtϕ+v·Dxϕ+divv(aϕ)=0,ϕ|t=0=ϕ0.

Then, for any (u,β,βT)KA such that A˜(u,β,βT)<+, the following hold:

  • uLtLx,v1(ϕ), that is,ess supt[0,T]M×Rd|ut|ϕtdxdv<+.

  • The negative part of β satisfies βLt,x,v1(ϕ).

  • The negative part of βT satisfies (βT)Lx,v1(ϕT).

  • The following estimate holds:uLtLx,v1(ϕ)+βLt,x,v1(ϕ)+(βT)Lx,v1(ϕT)+DvuLt,x,vr(ϕ)rC(a,ϕ,m0,H,T)(1+β+Lq+(βT)+L+Lq)M×Rdu0m0dxdv.

Proof.

Note the following properties of ϕ:

  • ϕ is non-negative,

  • ϕ has compact support contained in Um0.

  • ϕ(L1L)t,x,v.

In particular, since ϕt(L1Lq)(M×Rd) for any t[0,T], then M×Rd(ut)+ϕtdxdvϕt(L1Lq)x,vu+Lt(L+Lq)x,v.

By Lemma 4.1, (4.3) u+Lt(L+Lq)x,vC(T,H)(1+(βT)+(L+Lq)(M×Rd)+β+Lt,x,vq)(4.3) and thus for t[0,T], we have M×Rd(ut)+ϕtdxdvC(T,H,ϕ)(1+(βT)+(L+Lq)(M×Rd)+β+Lt,x,vq), where (ut)+ is understood in the sense of weak trace (cf. Lemma A.1, EquationDefinition A.5).

For the negative part we make use of the equation. A density argument shows that ϕ is admissible as a test function in the weak form of the Hamilton-Jacobi inequality satisfied by u. Thus for 0s<tT, M×RdusϕsdxdvM×Rdutϕtdxdv+stM×Rdu(tϕ+v·Dxϕ)dxdvdτ+stM×RdϕH(x,v,Dvu)dxdvdτstM×Rdβϕdxdvdτ.

We apply this in the case s = 0, t(0,T]. Using the fact that ϕ satisfies the continuity equation in a pointwise sense, M×Rdu0ϕ0dxdvM×Rdu(t,x,v)ϕ(t,x,v)dxdv0tM×Rdudivv(aϕ)dxdvdτ+0tM×RdϕH(x,v,Dvu)dxdvdτ0tM×Rdβϕdxdvdτ.

Here, let us notice that we have used the existence of weak traces in the sense of Lemma A.1. In particular the integral M×Rdu0ϕ0dxdv is meaningful and finite, since spt(ϕ0)spt(m0) (EquationDefinition A.10).

Since DvuLlocr(Um0) and aϕC1 has compact support contained in Um0, we may integrate by parts to obtain 0tM×Rdu divv(aϕ)dxdvdτ=0tM×Rda·Dvu ϕdxdvdτ.

Then estimate |0tM×Rda·Dvu ϕdxdvdτ|aLr(ϕ)DvuLr(ϕ),

where, in order to lighten the notation, we have used the shorthand hLp(ϕ):=(0TM×Rd|h|pϕdxdvdt)1/p,p[1,+)

to denote Lp norms with respect to the measure on [0,T]×M×Rd with density ϕ with respect to Lebesgue measure. Thus M×Rdutϕtdxdv+0tM×RdϕH(x,v,Dvu)dxdvdτaLr(ϕ)DvuLr(ϕ)M×Rdu0ϕ0dxdv+0tM×Rdβϕdxdvdτ.

Using the lower bounds on the Hamiltonian H, rearranging terms and using Young’s inequality for products (with a small parameter), we obtain M×Rdutϕtdxdv+0tM×Rdβϕdxdvdτ+c0tM×Rdϕ|Dvu|rdxdvdτCaLr(ϕ)r+CH0tM×RdϕdxdvdτM×Rdu0ϕ0dxdv+0tM×Rdβ+ϕdxdvdτ.

Then (4.4) M×Rdutϕtdxdv+0tM×Rdβϕdxdvdτ+c0tM×Rdϕ|Dvu|rdxdvdτC(a,ϕ)(CH+β+Lq)M×Rdu0m0dxdv+M×Rd(u0)+(m0ϕ0)dxdv.(4.4)

Finally, since 0m0ϕ0m0L1Lq, we use the (L+Lq)x,v bounds on the positive part (u0)+ (EquationEquation (4.3)) to conclude that M×Rd|ut|ϕtdxdv+0tM×Rdβϕdxdvdτ+c0tM×Rdϕ|Dvu|rdxdvdτC(a,ϕ,m0,H,T)(1+β+Lq+(βT)+L+Lq)M×Rdu0m0dxdv.

Notice that by setting t = T, Equation(4.4) and the fact that uTβT (together with the bounds that we already have on (βT)+) readily yield also that (βT)L1(ϕT).

This completes the proof. □

Corollary 4.4.

Let (u,β,βT)KA such that A˜(u,β,βT)<+. Then uLloc1(Um0),βLloc1(Um0),(βT)Lloc1(M×Rd) and DvuLlocr(Um0).

Proof.

First, consider a compact setKaCb1,0ϕ0Cc1({m0>0}){ϕ>0}, where ϕC1 denotes the solution of the continuity equation (4.5) tϕ+v·Dxϕ+divv(aϕ)=0,ϕ|t=0=ϕ0.(4.5)

By compactness of K, there exist finitely many ϕi,i=1,,k such that Ki=1k{ϕi>0}.

The function maxiϕi is continuous and so 0<δK:=infKmaxiϕi.

Then uL1(K)+βL1(K)+DvuLr(K)δK1i=1kuL1(ϕi)+βL1(ϕi)+DvuLr(ϕi).

By Lemma 4.3, this leads to the estimate uL1(K)+βL1(K)+DvuLr(K)C, where C=C(K,A˜(u,β,βT)). We now claim that Um0aCb1,0ϕ0Cc1({m0>0}){ϕ>0}.

This follows from the controllability of the ODE system (4.6) ẋ=v,v̇=a(4.6) on M×Rd. That is, for any initial datum (x0,v0)M×Rd and target (t,x,v)(0,T]×M×Rd, there exists a control function a such that the solution (x(τ),v(τ)) of the ODE Equation(4.5) with (x(0),v(0))=(x0,v0) satisfies (x(t),v(t))=(x,v).

Next, note that (since m0 is continuous) {m0>0} contains a closed ball B¯r(x0,v0) for some point (x0, v0) and some r > 0. Thus there exists 0ϕ0Cc1({m0>0}) such that ϕ0>0 on B¯r(x0,v0). Consider the solution ϕ of Equation(4.4) for the control a found above and with this choice of ϕ0. It follows that (t,x,v){ϕ>0}.

Finally, we notice that by the structure of the set Um0, we have the bound (βT)Lloc1(M×Rd).

5. Duality for the relaxed problem

Theorem 5.1.

Problems 3.3 and 3.8 are in duality:inf(u,β,βT)KAA˜(u,β,βT)=min(m,w)KBB(m,w).

Proof.

For uCb1([0,T]×M×Rd) such that A(u)<+, the triple (u,tuv·Dxu+H(x,v,Dvu),uT) lies in KA. Thus inf(u,β)KAA˜(u,β,βT)infuCb1A(u).

By the duality result of Lemma 3.5, inf(u,β,βT)KAA˜(u,β,βT)infuCb1A(u)=min(m,w)KBB(m,w).

It therefore remains only to prove the reverse inequality. This follows from Lemma 5.2 below, which states that for all (u,β,βT)KA and (m,w)KB, A˜(u,β,βT)B(m,w).

Taking the infimum over (u,β,βT)KA and supremum over (m,w)KB gives inf(u,β,βT)KAA˜(u,β,βT)min(m,w)KBB(m,w) as required. □

Lemma 5.2.

Let (u,β,βT)KA and (m,w)KB such that A˜(u,β,βT),B(m,w)<+. ThenA˜(u,β,βT)+B(m,w)0.

In the proof of this lemma we require the following observation regarding the commutator between the operator v·Dx and the operator given by convolution with a fixed function.

Lemma 5.3.

Let χ:Rd[0,+) be a function such that (1+|v|)χLv1(Rd). Let hLvp(Rd) for p[1,+]. Thenvχhχ(vh)=(vχ)h.

Proof.

By direct computation, for all vRd, vχh(v)χ(vh)=vRdh(vz)χ(z)dzRd(vz)h(vz)χ(z)dz=Rdzχ(z)h(vz)dz=(vχ)h.

Proof of Lemma 5.2.

The overall idea of the proof is to use m as a test function in the weak form of the inequality tuv·Dxu+H(x,v,Dvu)β

and its terminal condition uTβT.

To make this valid, we must first introduce an approximation procedure.

First, we introduce a lower cutoff on u and β. Let l0 and define ul:=max{u,l}. Similarly, for k0, let βk:=max{β,k}. Then by Lemma B.1 we obtain (5.1) {tulv·Dxul+H(x,v,Dvu)1{u>l}βk1{u>l},(0,T)×M×Rd(uT)l(βT)l,M×Rd,(5.1) in the sense of distributions. By Lemma 4.1, ulLt(L+Lq)x,v. We emphasize that k and l are taken to be possibly independent at this point.

Next, we approximate m by a function in Cc1, which is then an admissible test function for the Hamilton-Jacobi EquationEquation (5.1). We regularize m by convolution with a mollifier. For ease of presentation, it will be convenient to work with the time, space and velocity variables separately. Fix χCc(Rd) and define χε for ε>0 by χε(v):=εdχ(vε).

For the space variable, consider ψCc(Rd) and for δ>0 let ψδ(x):=δdψ(xδ).

For the time variable, fix θCc(R) and for η>0 let θη(t):=η1θ(tη).

We then define the full mollifier φ by φ(t,x,v)=θη(t)ψδ(x)χε(v).

Then define the smooth functions m˜:=φt,x,vmandw˜=φt,x,vw.

Notice that for the convolution in time, (m, w) needs to be extended. We choose the following extensions. We set w(t,·,·)=0 to for t < 0 and t > T. Then, if t < 0, we set m(t,·,·) to be the solution to the problem {tm+v·Dxm=0,in(η,0)×M×Rd,m(0,·,·)=m0,inM×Rd.

Similarly, for t > T we set m(t,·,·) to be the solution to {tm+v·Dxm=0,in(T,T+η)×M×Rd,m(T,·,·)=mT,inM×Rd, where mT is the trace of m in time at t = T.

As the final step in the approximation, we localize m˜. As localizers we consider smooth functions ζRCc(M×Rd) such that (5.2) ζR(x,v)={1|x|R2,|v|R0|x|>2R2,|v|>2R.|DxζR|CR2,|DvζR|CR .(5.2)

We then define m˜(R):=ζRm˜ and w˜(R):=ζRw˜.

Then m˜(R) satisfies the equation tm˜(R)+v·Dxm˜(R)+divvw˜(R)=Eη,δ,ε,R, where the error term is given by (5.3) Eη,δ,ε,R:=ζR [v·Dx,χεv]θηtψδxm+(v·DxζR)m˜+(DvζR)·w˜.(5.3)

Here, we use the standard commutator notation [Λ1,Λ2]f:=Λ1(Λ2f)Λ2(Λ1f), where Λ1,Λ2 are some operators acting on the function f.

5.1. Convergence of the error term

We show that the error term Eη,δ,ε,R defined by Equation(5.3) converges to zero in the space Lt1(L1+Lq)x,v, as R+ and η,ε,δ0, under a certain relationship between these parameters.

For the first term, either for p = 1 or p = q, using the explicit formula for the commutator we estimate ζR [v·Dx,χεv]θηtψδxmLt1Lx,vp[v·Dx,χεv](θηψδ)t,xmLt1Lx,vp[v,χεv]θηt(Dxψδ)xmLt1Lx,vp(vχε)vθηt(Dxψδ)xmLt1Lx,vpvχεL1θηL1DxψδL1mLt1Lx,vpεδ1·CT1/pmLt,x,vp, where we have used Lemma 5.3 in the third inequality. Thus by choosing ε=ε(δ) sufficiently small with respect to δ, we may ensure that limδ0supη,RζR [v·Dx,χε(δ)v]θηtψδxmLt1Lx,vp=0.

For the second term, observe that for all R > 0, |v·DxζR|CR1 and thus (v·DxζR)m˜Lt1(L1Lq)x,vCR1m˜Lt1(L1Lq)x,vCR1mLt1(L1Lq)x,vCTR1m(L1Lq)t,x,v.

It follows that limR+supη,δ,ε(v·DxζR)m˜Lt1(L1Lq)x,v=0.

For the third term, for either p = 1 or p = q we have (DvζR)·w˜Lt1Lx,vpCRw˜Lt1Lx,vpCRφLt1Lx,vpwLt,x,v1CRδd/pεd/pwLt,x,v1.

Taking ε=ε(δ) as above, we can then ensure this term converges to zero by choosing R=R(δ) sufficiently large with respect to δ and ε(δ). Thus, for this choice of ε(δ),R(δ), we have limδ0supη(DvζR)·w˜Lt1Lx,vp=0.

Altogether, we have found that there exists a regime R=R(δ) and ε=ε(δ) such that limδ0supηEη,δ,ε,RLt1Lx,vp=0.

5.2. Testing the equation

Using m˜(R) as a test function in the weak form of the equation for ul, one obtains M×Rdul(0,x,v)m˜(R)(0,x,v)dxdvM×Rd(βT)lm˜(R)(T,x,v)dxdv+0TM×Rdul(tm˜(R)+v·Dxm˜(R))dxdvdt+0TM×Rdm˜(R)H(x,v,Dvu)1{u>l}dxdvdt0TM×Rdβkm˜(R)1{u>l}dxdvdt. Using the equation satisfied by m˜(R), we have (5.4) M×Rdul(0,x,v)m˜(R)(0,x,v)dxdvM×Rd(βT)lm˜(R)(T,x,v)dxdv0TM×Rduldivvw˜(R)dxdvdt+0TM×Rdm˜(R)H(x,v,Dvu)1{u>l}dxdvdt0TM×Rdβkm˜(R)1{u>l}dxdvdt0TM×Rdul Eη,δ,ε,Rdxdvdt.(5.4)

Next, note that DvuLlocr(Um0). By the chain rule for Lipschitz functions composed with Sobolev-regular functions, Dvul=Dvu1{u>l}.

Thus, using the definition of distributional derivative we may integrate by parts to obtain 0TM×Rdul divvw˜(R)dxdvdt=0TM×Rdw˜(R)·Dvu1{u>l}dxdvdt.

Since B(m,w) is finite, w is absolutely continuous with respect to m. It follows that there exists α˜Lr([0,T]×M×Rd;m˜) such that w˜=α˜m˜.

Thus 0TM×Rd(w˜(R)·Dvu+m˜(R)H(x,v,Dvu))1{u>l}dxdvdt=0TM×Rd(α˜·Dvu+H(x,v,Dvu))m˜(R)1{u>l}dxdvdt0TM×RdL(x,v,α˜)m˜(R)1{u>l}dxdvdt.

Substituting this, we obtain (5.5) M×Rdul(0,x,v)m˜(R)(0,x,v)dxdvM×Rd(βT)lm˜(R)(T,x,v)dxdv0TM×RdL(x,v,α˜)m˜(R)1{u>l}dxdvdt0TM×Rdβkm˜(R)1{u>l}dxdvdt0TM×Rdul Eη,δ,ε,Rdxdvdt.(5.5)

We have shown above that there exists a regime R=R(δ) and ε=ε(δ) such that the final term converges to zero uniformly in η as δ tends to zero, since ulLt(L+Lq)x,v. We now discuss the convergence of the other terms.

5.3. Boundary terms

We consider the boundary terms at t=0,T. Note that Lemma 4.1 and Corollary 4.2 yield ul(0,·)Lq+L (since (u0)+L+Lq and ul(0,·) is bounded below), while by Lemma 4.1 and Corollary 4.4 we have (βT)lLs+L.

We first show that m˜0(R) converges to m0 in L1Lq, and m˜T(R) converges to mT in L1Lq, in the limit as δ tends to zero for a certain regime η=η(δ) and R=R(δ),ε=ε(δ) according to the regime already found above.

For t=0,T, we write (5.6) m˜t(R)mt=(m˜t(R)ζRψδxχεvmt)+ζR(ψδxχεvmtmt)+mt(ζR1).(5.6)

We first note that m0(L1Lq)x,v by the assumption that it is a bounded probability density, while mTLx,vs since the energy B(m,w) is finite, and mTLx,v1 since the continuity equation conserves mass.

Then, since |mt(ζR1)|mt, if mtLp (where p{1,q} or p{1,s}) then by dominated convergence limR+mt(ζR1)Lx,vp=0.

Moreover, by continuity of translations in Lp, limδ0ψδxχε(δ)vmtmtLx,vp=0.

Since for all R > 0 ζR(ψδxχεvmtmt)Lx,vpψδxχεvmtmtLx,vp,

it follows that limδ0supRζR(ψδxχε(δ)vmtmt)Lx,vp=0.

Therefore, the latter two terms of Equation(5.6) converge to zero as δ tends to zero with ε=ε(δ) and R=R(δ) as already specified above, in (L1Lq)x,v for t = 0 and in (L1Ls)x,v for t = T.

It remains to estimate the difference ψδxχεvmtm˜t. For any function fLx,vp (p[1,+] to be specified), M×Rd(ψδxχεvmtm˜t)ζRfdxdv=M×Rd(mt(θηtm)t)ψδxχεv(ζRf)dxdv.

We use the notation f˜:=ψδxχεv(ζRf). Writing the time convolution explicitly, we obtain M×Rd(mt(θηtm)t)f˜dxdv=M×Rdθη(τ)(mtmtτ)f˜dxdvdτ.

Next, we use estimates on tm: mtmtτ,f˜=tτttm,f˜sds=tτtm,v·xf˜s+w,vf˜sds.

Then, since m,wL1, |mtmtτ,f˜|(tτtmsLx,v1+wsLx,v1ds)(v·xf˜L+vf˜L).

We estimate f˜: v·xf˜L=v·xψδxχεv(ζRf)L(R+Cε)xψδLpχεLpfLp(R+Cε)δ(1+d/p)εd/pfLp and similarly vf˜L=ψδxvχεv(ζRf)LψδLpvχεLpfLpδd/pε(1+d/p)fLp.

Thus |mtmtτ,f˜|C(δ,ε,R)(tτtmsLx,v1+wsLx,v1ds)fLx,vp.

Finally |M×Rd(mt(θηtm)t)f˜dxdv|C(δ,ε,R)(tηt+ηmsLx,v1+wsLx,v1ds)fLx,vpω(η) C(δ,ε,R) fLx,vp, where limη0ω(η)=0. Thus it is possible to choose η=η(δ) depending on δ,ε(δ),R(δ) in such a way that limδ0ω(η) C(δ,ε,R)=0,limδ0η(δ)=0.

We apply this in the case f = fi for i = 1, 2, where (βT)l=f1+f2orul(0)=f1+f2,f1{Lx,vs if t=TLx,vq if t=0f2Lx,v.

Consequently, for t=0,T, M×Rdm˜0(R)ul(0)dxdvM×Rdm0ul(0)dxdv,

and M×Rdm˜T(R)(βT)ldxdvM×RdmT(βT)ldxdv, as δ0 with η,R,ε chosen to depend on δ in the manner specified.

Finally, we take the limit l. For the term, t = 0, convergence holds by monotonicity, and the limit is finite since M×Rdu0m0dxdv<+

by finiteness of A˜(u,β,βT). For the term t = T, we first note that M×Rd(βT)lmT(x,v)dxdvM×RdG(mT)dxdv+M×RdG((βT)l)dxdv.

The second term on the right hand side converges due to the assumption on G Equation(2.3), since the integrand is dominated by supu<0G(u)CGL1(M×Rd).

5.4. Term involving βm

Since mL1Lq, by standard results on approximation by mollification in Lp spaces we have lim(η,δ,ε)0,R+m˜(R)mL1Lq=0, and thus the same limit holds with η,ε,R chosen to depend on δ as described above. Then, since βkL+Lq, we deduce that lim(δ,ε)0,R+0TM×Rdβkm˜(R)1{u>l}dxdvdt=0TM×Rdβkm1{u>l}dxdvdt.

By the definition of Fenchel conjugate, 0TM×Rdβkm1{u>l}dxdvdt0TM×Rd(F(x,v,βk)+F(x,v,m))1{u>l}dxdvdt.

We then take the limit l. Note F and F are both lower bounded by integrable functions (conditions Equation(2.2) and Equation(3.1)). Then, by monotonicity, liml0TM×Rd(F(x,v,m)infmF(x,v,m))1{u>l}dxdvdt=0TM×Rd(F(x,v,m)infmF(x,v,m))dxdvdt.

Moreover liml0TM×RdinfmF(x,v,m)1{u>l}dxdvdt=0TM×RdinfmF(x,v,m)dxdvdt.

Since the lower bound is integrable and F(·,·,m) has finite integral by finiteness of the energy, both of these limits are finite. Thus liml0TM×RdF(x,v,m)1{u>l}dxdvdt=0TM×RdF(x,v,m)dxdvdt.

A similar argument shows that liml0TM×RdF(x,v,βk)1{u>l}dxdvdt=0TM×RdF(x,v,βk)dxdvdt where the right hand side is finite.

Finally, we consider k. Note that supβ0F(x,v,β)CF(x,v)L1 by assumption (see Equation(3.1)). Since F(β) is a continuous non-decreasing function of β, as k decreases to negative infinity F(·,·,βk) is decreasing and converges almost everywhere to F(·,·,β). Thus we deduce the convergence limk{β0}F(x,v,βk)dxdvdt={β0}F(x,v,β)dxdvdt.

By the bounds Equation(3.1), the right hand side is finite. Moreover, for any k0, {β>0}F(x,v,βk)dxdvdt={β>0}F(x,v,β)dxdvdt.

Thus we conclude that limk0TM×RdF(x,v,βk)dxdvdt=0TM×RdF(x,v,β)dxdvdt.

5.5. Lagrangian term

For the term involving the Lagrangian, we use a similar argument as was used in [Citation20]. This argument is based on the joint convexity of L(x,v,w/m)m as a function of (m, w). In our case we must additionally account for the convergence of the localizer ζR. By convexity, for all (t, x, v), the integrand satisfies the inequality L(x,v,w˜m˜)m˜ζR1{u>l}0TM×Rdθη(tt)ψδ(xx)χε(vv)L(x,v,w(t,x,v)m(t,x,v))m(t,x,v)dtdxdv ζR1{u>l}=θηtψδxχεv[L(x,v,wm)m]ζR1{u>l}. Then, note that θηtψδxχεv[L(x,v,wm)m](t,x,v)=θηtψδxχεv[L(·,wm)m](t,x,v)+θηtψδxχεv[L(x,v,wm)mL(·,·,wm)m](t,x,v).

Then, since L(·,·,wm)mL1((0,T)×M×Rd), θηtψδxχεv[L(·,·,wm)m] converges to L(·,·,wm)m in L1((0,T)×M×Rd) as η,δ,ε tend to zero. Since 0ζR1, limη,δ,ε0supR{θηtψδxχεv[L(·,·,wm)m]L(·,·,wm)m}ζR1{u>l}L1=0.

It follows that if we take the regime R=R(δ),ε=ε(δ),η=η(δ) established above, then limδ0{θηtψδxχεv[L(·,·,wm)m]L(·,·,wm)m}ζR1{u>l}L1=0.

We stay with this regime and consider the remaining term 0TM×Rdθηtψδxχεv{L(x,v,wm)mL(·,·,wm)m}(t,x,v)ζR(x,v)1{u>l}dxdvdt.

The integrand converges to zero almost everywhere: note then that |θηtψδxχεv{L(x,v,wm)mL(·,·,wm)m}ζR1{u>l}|θηtψδxχεv[(C+|w|rmr)m].

The right hand side converges in L1 to Cm+|w|rm1r as δ tends to zero. Thus by dominated convergence we conclude that limδ00TM×RdL(x,v,w˜m˜)m˜ζR1{u>l}dxdvdt0TM×RdL(x,v,wm)m1{u>l}dxdvdt.

The reverse inequality follows from Fatou’s lemma. Thus limδ00TM×RdL(x,v,w˜m˜)m˜ζR1{u>l}dxdvdt=0TM×RdL(x,v,wm)m1{u>l}dxdvdt.

Finally, we take the limit l. Since L(x,v,wm)mL1, in the limit we obtain 0TM×RdL(x,v,wm)mdxdvdt.

5.6. Conclusion

From the discussion above, we have obtained M×Rdu0m0dxdvM×RdG(x,v,mT)dxdvM×RdG(x,v,uT)dxdv0TM×RdL(x,v,wm)mdxdvdt0TM×Rd(F(x,v,β)+F(x,v,m))dxdvdt, where all terms are finite. Rearranging this inequality, we obtain the statement A˜(u,β,βT)+B(m,w)0.

Corollary 5.4.

Let (u,β,βT)KA and (m,w)KB be such that A˜(u,β,βT)<+ and B(m,w)<+.

ThenβmL1((0,T)×M×Rd)and(βT)mTL1(M×Rd).

Moreover, for almost all t[0,T], M×Rd[utmtβTmT]dxdvtTM×RdL(x,v,wm)mdxdvdttTM×Rdβmdxdvdt,and(5.7) M×Rd[u0m0utmt]dxdv0tM×RdL(x,v,wm)mdxdvdt0tM×Rdβmdxdvdt.(5.7)

In particular,(5.8) M×Rd[u0m0βTmT]dxdv0TM×RdL(x,v,wm)mdxdvdt0TM×Rdβmdxdvdt.(5.8)

Proof.

This is a consequence of the proof of Lemma 5.2 and in particular the inequality Equation(5.5). First, let us show the first part of the statement, i.e. that βmL1((0,T)×M×Rd)and(βT)mTL1(M×Rd).

As in the mentioned proof, let us first pass to the limit with η,δ,ε,R in the inequality Equation(5.5). Then, we pass to the limit as l and k the remaining terms.

All the terms, except the ones involving ((βT))lmT and (β)km1{u>l} pass to the limit, as in the proof. After rearranging, we also find that both M×Rd((βT))lmTdxdv and 0TM×Rd(β)km1{u>l}dxdvdt are uniformly bounded, independently of l and k. Therefore, the monotone convergence theorem yields limlM×Rd((βT))lmTdxdv=M×Rd(βT)mTdxdv and limllimk0TM×Rd(β)km1{u>l}dxdvdt=0TM×Rdβmdxdvdt.

The summability results follow, and so does Equation(5.8).

For the case of general t[0,T], we begin by testing the equation to find that, for example, M×Rdul(t,x,v)m˜(R)(t,x,v)dxdvM×Rd(βT)lm˜(R)(T,x,v)dxdvtTM×RdL(x,v,α˜)m˜(R)1{u>l}dxdvdttTM×Rdβkm˜(R)1{u>l}dxdvdttTM×Rdul Eη,δ,ε,Rdxdvdt.

We note that we are referring to the version of ul that is weakly right continuous in time (see Appendix A). The only term that requires attention is M×Rdul(t,x,v)m˜(R)(t,x,v)dxdv.

For almost all t[0,T],ul(t)(Lq+L)(M×Rd) and mtL1Lq. Thus the arguments for the boundary terms in Lemma 5.2 show that limδ0M×Rdul(t,x,v)m˜(R)(t,x,v)dxdv=M×Rdul(t,x,v)m(t,x,v)dxdv.

The limit l is then taken by monotone convergence, noting that (ut)+mtL1(M×Rd). Note that the argument for the case Equation(5.7) shows that limit is not negative infinity, since all other terms have finite limits. □

Corollary 5.5.

Let (u,β,βT)KA and (m,w)KB be such that A˜(u,β,βT)<+ and B(m,w)<+. Then

  1. m|Dvu|r is uniformly bounded in L1((0,T)×M×Rd), by a constant that depends only on the data and A˜(u,β,βT) and B(m,w).

  2. The following estimate holds:0TM×RdmH(x,v,Dvu)+mL(x,v,wm)+Dvu·w dxdvdtM×RdβTmTdxdvM×Rdu0m0dxdv+0TM×Rdβmdxdvdt+0TM×RdmL(x,v,wm)dxdvdt.

Proof.

This is a consequence of Equation(5.4). Using the same notation as in the proof of Lemma 5.2, we rewrite Equation(5.4) as 0TM×Rdm˜(R)H(x,v,Dvu)1{u>l}dxdvdtM×Rd(βT)lm˜(R)(T,x,v)dxdvM×Rdul(0,x,v)m˜(R)(0,x,v)dxdv+0TM×Rdβkm˜(R)1{u>l}dxdvdt0TM×RdDvul·w˜(R)1{u>l}dxdvdt0TM×Rdul Eη,δ,ε,RdxdvdtM×Rd(βT)+m˜(R)(T,x,v)dxdv+M×Rd[ul(0,x,v)]m˜(R)(0,x,v)dxdv+0TM×Rdβ+m˜(R)dxdvdt0TM×RdDvul·w˜(R)1{u>l}dxdvdt0TM×Rdul Eη,δ,ε,Rdxdvdt. Let us observe that for some θ>0 parameter that we choose later, Young’s inequality yields 0TM×RdDvul·w˜(R)1{u>l}dxdvdt=0TM×Rdθ(m˜(R))1rDvul·w˜(R)θ1(m˜(R))1r1{u>l}dxdvdt1rθr0TM×Rdm˜(R)|Dvul|r1{u>l}dxdvdt+1rθr0TM×Rd|w˜(R)|r(m˜(R))rr1{u>l}dxdvdt.

We notice that rr=1r. Thus, by using the growth condition Equation(2.1) on the Hamiltonian and choosing θ appropriately, we can conclude that there exists a constant C > 0 (independent of the parameters l,k,R,η,ε,δ), such that after passing to the limit with R,η,ε,δ, as in the proof of Lemma 5.2, we obtain 0TM×Rdm|Dvul|r1{u>l}dxdvdtC+M×Rd(βT)+m(T,x,v)dxdv+M×Rd[ul(0,x,v)]m0(x,v)dxdv+0TM×Rdβ+mdxdvdt+C0TM×Rd|w|rm1rdxdvdt.

Since the right hand side of this inequality is uniformly bounded, independently of l (by Lemma 4.1(ii), Corollary 4.2(ii) and Remark 3.7), the result follows by Fatou’s lemma by sending l.

Using Equation(5.4), we have 0TM×Rdm˜(R)H(x,v,Dvu)1{u>l}dxdvdt+0TM×Rdm˜(R)L(x,v,w˜m˜)1{u>l}dxdvdt+0TM×RdDvu·w˜(R)1{u>l}dxdvdtM×Rd(βT)lm˜(R)(T,x,v)dxdvM×Rdul(0,x,v)m˜(R)(0,x,v)dxdv+0TM×Rdβkm˜(R)1{u>l}dxdvdt+0TM×Rdm˜(R)L(x,v,w˜m˜)1{u>l}dxdvdt0TM×Rdul Eη,δ,ε,Rdxdvdt.

Passing to the limit with R,η,ε,δ, as in the proof of Lemma 5.2, by Fatou’s lemma we obtain 0TM×RdmH(x,v,Dvu)1{u>l}dxdvdt+0TM×RdmL(x,v,wm)1{u>l}dxdvdt+0TM×RdDvu·w1{u>l}dxdvdtM×Rd(βT)lm(T,x,v)dxdvM×Rdul(0,x,v)m(0,x,v)dxdv+0TM×Rdβkm1{u>l}dxdvdt+0TM×RdmL(x,v,wm)1{u>l}dxdvdt.

Finally, taking the limit l,k as in the proofs of Lemma 5.2 and Corollary 5.4, we obtain 0TM×RdmH(x,v,Dvu)+mL(x,v,wm)+Dvu·w dxdvdtM×RdβTmTdxdvM×Rdu0m0dxdv+0TM×Rdβmdxdvdt+0TM×RdmL(x,v,wm)dxdvdt.

6. Existence of a solution of the relaxed problem

In this section we prove the existence of a solution for the relaxed problem. Consider a minimizing sequence (un,βn,(βT)n)n. We will extract a convergent subsequence, and show that the limit constitutes a minimizer of the objective functional.

6.1. Compactness of the minimizing sequence

The following proposition enables the extraction of a convergent subsequence.

Proposition 6.1.

Let (un)n be a sequence of solutions to the Hamilton-Jacobi equationstunv·Dxun+H(x,v,Dvun)=βn,

Assume that:

  • The family (un)n is uniformly bounded in LtLx,v,loc1(Um0).

  • The family (Dvun)n is uniformly bounded in Llocr(Um0).

  • The family (βn)n is uniformly bounded in Lloc1(Um0).

Then there exists a subsequence (unj)j that is strongly convergent in Lloc1(Um0).

The proof of this result will be a consequence of some intermediate results that we detail below.

Remark 6.2.

Let us notice that the assumptions of Proposition 6.1 hold true by Corollary 4.4 and Lemma 4.1.

Proposition 6.1

is proved by treating the Hamilton-Jacobi equation as a kinetic transport equation with right hand side bounded in Lloc1(Um0): tunv·Dxun=βnH(x,v,Dvun).

A form of compactness for the solutions can be obtained by using an averaging lemma. Averaging lemmas are results in kinetic theory showing that, for Lp-bounded families of solutions to the kinetic transport equation, with Lp-bounded source terms, the velocity averages ρϕ[u](t,x):=Rdu(t,x,v)ϕ(v)dvϕCc(Rd), enjoy additional fractional Sobolev regularity and/or strong Lp-compactness. In our case we are in the setting p = 1, and we use an L1 averaging result from [Citation49]. It is necessary to assume a certain equi-integrability condition on the solution un. This condition is defined below.

Definition 6.3

(Equi-integrability in velocity). Let (uλ)λΛ be a bounded family in Lloc1([0,T]×M×Rd). The family is locally equi-integrable in v if, for all ε>0 and all compact sets K[0,T]×M×Rd, there exists η>0 such that for all measurable families (At,x)(t,x)[0,T]×M of measurable subsets of Rd for which sup(t,x)[0,T]×M|At,x|<η, 0TMAt,x1K|uλ(t,x,v)|dvdxdt<εfor all λΛ.

The required averaging lemma is quoted below. This result was proved in [Citation49] for the stationary case, i.e. the equation v·Dxu=β. The result can be adapted to the time dependent equation by standard techniques; see [Citation50] or [Citation51] for statements in the time dependent setting.

Theorem 6.4.

Let (uλ)λΛ be a bounded family in L([0,T];Lloc1(M×Rd)) satisfying(6.1) tuλ+v·Dxuλ=βλ,(6.1)

where (βλ)λΛ is a bounded family in Lloc1([0,T]×M×Rd). Assume that (uλ)λΛ is equi-integrable in v. Then

  • The family (uλ)λΛ is locally equi-integrable in all variables in [0,T]×M×Rd.

  • For each ϕCc(Rd), the family of averages (ρϕ[uλ])λΛ is relatively (strongly) compact in Lloc1([0,T]×M).

In our setting we expect to have local summability estimates in Um0 rather than [0,T]×M×Rd. To deal with this technicality we make use of a localization procedure: given a compact set K, consider a smooth bump function ζ supported in K. If uλ satisfies the kinetic transport EquationEquation (6.2), then uλζ satisfies (6.2) t(uλζ)+v·Dx(uλζ)=βλζ+(tζ+v·xζ)uλ.(6.2)

The right hand side of the above equation is bounded in L1([0,T]×M×Rd), uniformly in λ, as long as uλ and βλ are bounded in Lloc1(Um0).

We wish to apply this to the solutions of the Hamilton-Jacobi equation. To do this, we verify the equi-integrability condition. To prove equi-integrability, we make use of the Lr estimates available for the v-derivative Dvu.

Lemma 6.5.

Let (uλ)λΛ be a bounded family in LtLx,v,loc1(Um0). Assume that (Dvuλ)λΛ is a bounded family in Llocr(Um0). Then:

  • (uλ)λΛ is bounded in Lt,x1Lv,locα(Um0), where 1/α=1/r1/d if r < d, or any α<+ if rd.

  • (uλ)λΛ is equi-integrable in v, locally in Um0.

Proof.

We obtain higher integrability in the velocity variable by using Sobolev embedding. We first apply a localization procedure. Given a compact set KUm0, let ζK denote a smooth bump function with compact support contained in Um0, such that ζK takes values contained in [0,1] and ζK1 on K. Then Dv(uλζK)=Dvuλ ζK+uλ DvζK.

Let K denote the support of ζK. Then Dv(uλζK)L1C(ζK)(DvuλLr(K)+uλLtLx,v1(K)), thus Dv(uλζK) is bounded in L1, uniformly in λ. Moreover it is compactly supported.

We then apply Sobolev embedding in the v variable. Letting 1:=dd1, we have uλLt,x1Lv1(K)uλζKLt,x1Lv1CdDv(uλζK)L1C(d,ζK)(DvuλLr(K)+uλLtLx,v1(K)).

Thus (uλ)λ is uniformly bounded in Lt,x1Lv,loc1.

We now apply a bootstrap argument: it follows from the above that Dv(uλζK) is bounded in Lt,x1Lv1, and thus (uλ)λ is bounded in Lt,x1Lv,locα for α=(12/d)1. This process can be repeated until we obtain that (uλ)λ is bounded in Lt,x1Lv,locr, for 1/r*=1/r1/d, if r < d; otherwise we may obtain Lt,x1Lv,locα for any α<+.

We now prove local equi-integrability. Let (At,x)(t,x)[0,T]×M be a measurable family of measurable subsets of Rd, such that |At,x|<η for all t, x. Then 0TMAt,x1K|uλ|dvdxdt1At,x1KLt,xLvαuλ1KLt,x1Lvα,

where denotes a Hölder conjugate exponent. From the condition on the measure of At,x, we have 0TMAt,x1K|uλ(t,x,v)|dvdxdtC(K)supt,x|At,x|1/αuλLt,xrLvα(K)C(K)uλLt,x1Lvα(K) η1/α, which proves equi-integrability. □

It follows that, under the assumptions of Proposition 6.1, Theorem 6.4 can be applied to (ζKun)n for any ζKCc(Um0). We deduce strong Lloc1-compactness for the averages (ρϕ[unζK])n. We now use this to prove strong compactness for the full solutions un.

Lemma 6.6.

Assume that the family (un)n satisfies the following:

  • (un)n is uniformly bounded in LtLx,v1.

  • (un)n is equi-integrable in all variables.

  • (un)n share the same compact support K.

  • (Dvun)n is uniformly bounded in Lt,x1Lvr.

  • For each ϕCc(Rd), the family of averages (ρϕ[un])n is relatively (strongly) compact in Lloc1.

Then the family (un)n is relatively (strongly) compact in L1.

Proof.

First, note that the first two assumptions imply the weak L1 sequential compactness of (un)n. We pass (without relabelling) to a weakly convergent subsequence un, and let u denote the weak limit. In the remainder of the proof we improve the mode of convergence of un to u to strong convergence in L1.

  • Step 1: Approximation by smoothing in v. We approximate un by a function that is smooth with respect to the v variable. Fix ϕCc(Rd) and define, for ε>0,

ϕε(v):=εdϕ(vε).

Let un,ε(t,x,v):=Rdun(t,x,v)ϕε(vv)dv.

  • Step 2: Compactness for the approximations. Fix v*Rd. For any ψLt,x with compact support we consider testing the sequence (un)n against the test function

ψ(t,x)ϕε(v*v).

Since un,ψϕε(v*·)u,ψϕε(v*·),n+, we deduce that un,ε(·,·,v*) converges weakly in Lt,x,loc1 to uvϕε(·,·,v*) as n+.

Note moreover for each fixed vRd, un,ε(t,x,v) is a velocity average with respect to the test function ϕε(v·). Therefore, by Theorem 6.4 the convergence in fact holds strongly in Lt,x,loc1 for each vRd.

Furthermore, for fixed ε>0, the family (un,ε)n is equi-continuous in v into LtLx1: indeed |un,ε(t,x,v+h)un,ε(t,x,v)|=|Rdun(t,x,v)(ϕε(vv+h)ϕε(vv))dv||h|ϕεLRd|un(t,x,v)|dv.

Thus un,ε(·,·,v+h)un,ε(·,·,v)LtLx1CεsupmumLtLx,v1 |h|.

By an Arzelà-Ascoli argument the convergence therefore holds locally uniformly in v, with respect to the strong topology on Lt,x,loc1: that is, for all compact sets KvRd and Kt,x[0,T]×Rd, limnsupvKvun,ε(v)uϕε(v)Lt,x1(Kt,x)=0.

Consequently, the convergence holds in Lloc1; in fact, since un,εuϕε is supported for all n in K+BCε(0), the convergence holds in L1.

  • Step 3: Removing the approximation.

The bound on Dvun implies that, for any hRd, un(t,x,·+h)un(t,x,·)Lvr|h|Dvun(t,x,·)Lvr.

It follows that un,εunLt,x1LvrεsupmDvumLt,x1Lvr.

Indeed, by definition of un,ε, [un,εun](t,x,v)=Rd[un(t,x,vh)un(t,x,v)]ϕε(h)dh.

Thus, for any gLvr(Rd), Rd[un,εun](t,x,v)g(v)dv=RdRd[un(t,x,vh)un(t,x,v)]ϕε(h)g(v)dhdvgLvrRdun(t,x,+h)un(t,x,)Lvr|ϕε(h)|dhgLvrDvun(t,x,)LvrRd|h||ϕ1(hε)|εddhCϕεDvun(t,x,)LvrgLvr.

That is, un,ε(t,x,·)un(t,x,·)LvrεDvun(t,x,·)Lvr, then, one integrates in t, x and takes supremum.

Finally, estimate unuLt,x,v1unun,εLt,x,v1+un,εuvϕεLt,x,v1+uvϕεuLt,x,v1CKunun,εLt,x1Lvr+un,εuvϕεLt,x,v1+CKuvϕεuLt,x1LvrCKεsupmDvumLt,x1Lvr+un,εuvϕεLt,x,v1+uvϕεuLt,x1Lvr.

Thus limsupnunuLt,x,v1KεsupmDvumLt,x1Lvr+uvϕεuLt,x1Lvr0 as ε0, which completes the proof. □

Proof of Proposition 6.1.

The proof of this proposition follows by applying Lemma 6.6 to (ζKun)n.

It remains to obtain the necessary convergence of (βn)n and (βT,n)n.

Lemma 6.7.

Let (un,βn,βT,n) be a minimizing sequence for Problem 3.8. There exists a modification (un,β˜n,β˜T,n) of this sequence that is also minimizing such that (β˜n)n is weakly precompact in Lloc1(Um0) and (β˜T,n)n is weakly precompact in Lloc1(M×Rd).

Proof.

We replace βn with some β˜nβn and (βT,n)n with β˜T,nβT,n such that (β˜n,β˜T,n) is uniformly integrable, and (un,β˜n,β˜T,n) is still a minimizing sequence. We do this in a similar manner to [Citation20]: since (βn) is bounded in Lloc1(Um0), using a compact exhaustion of Um0 and a diagonal argument, by [Citation59] it is possible to pass to a subsequence such that the following holds for some JnR. We define β˜n by (β˜n):=(βn)1{(βn)Jn},(β˜n)+=(βn)+.

Then it is possible to choose Jn in such a way that:

  • For each compact set KUm0, the sequence (β˜n)1K is uniformly integrable.

  • The measure of the set {(βn)>Jn}K converges to zero as n tends to infinity.

We use the exact same construction for β˜T,n, and we can get the same properties (now taking KM×Rd).

We notice, that by construction the constraints tunv·Dxun+H(x,v,Dvun)β˜n and uT,nβ˜T,n are still satisfied. Finally, |0TM×RdF(x,v,β˜n)dxdvdt0TM×RdF(x,v,βn)dxdvdt|0TM×Rd|F(x,v,0)F(x,v,βn)|1{βnJn}dxdvdt.

By the estimate Equation(3.1), the integrand on the right hand side is dominated by 2CFL1, and thus the right hand side converges to zero as n tends to infinity.

The exact same arguments apply to G and β˜T,n too. Thus (un,β˜n,β˜T,n) is a minimizing sequence. Moreover, there exists (u,β,βT) such that up to passing to a subsequence, (un)n converges to u strongly in Lloc1(Um0),(β˜n)n converges weakly to β in Lloc1(Um0) and (β˜T,n)n converges to βT weakly in Lloc1(M×Rd).

6.2. Existence of a minimizer of A˜ over KA

In this subsection, we prove that there exists a minimizer (u,β,βT) by passing to the limit in the functional A˜(un,β˜n,β˜T,n):=0TM×RdF*(x,v,β˜n)dxdvdtM×Rdun(0,x,v)m0(x,v)dxdv+M×RdG*(β˜T,n(x,v))dxdv.

Theorem 6.8.

Under our standing assumptions, the functional A˜ admits a minimizer over KA.

Proof.

Let (un,βn,βT,n)nN be a minimizing sequence. Without loss of generality, for example by considering unC1, we may assume equality in the Hamilton-Jacobi equation: tunv·Dxun+H(x,v,Dvun)=βn,un(T,x,v)=βT,n(x,v).

For this minimizing sequence we have, for some constant C > 0, supnA˜(un,βn,βT,n)C.

We have discussed that this implies uniform in n bounds on the following quantities: (βT,n)+Ls(M×Rd),M×Rd(u0,n)m0dxdv,(βT,n)Lx,v,loc1(M×Rd),(u0,n)+(L+Lq)x,v,DvunLlocr(Um0),(βn)+Lt,x,vq,(βn)Lloc1(Um0).

To get the uniform integrability on (βn) and (βT,n), we perform the surgery argument as in Lemma 6.7. So, let (un,β˜n,β˜T,n)nN be the modification of the minimizing sequence (which will still have uniformly bounded energy). By Proposition 6.1 we know that (un)nN is strongly precompact in Lloc1(Um0), while Lemma 6.7 yields that (β˜n)nN and (β˜T,n)nN are weakly precompact in Lloc1(Um0) and Lloc1(M×Rd), respectively. In particular, after passing to a subsequence let us denote by u the strong Lloc1(Um0) limit of (un)n. In what follows, to ease the notation, we drop the tilde symbol, but whenever we write βn and βT,n, we mean the corresponding modified versions.

Passing to further subsequences (that we do not relabel), there exist limit functions so that we may also assume the following weak convergences:

  • (βn)+β+, weakly in Lt,x,vq([0,T]×M×Rd), as n+.

  • βnβ, weakly in Lloc1(Um0), as n+.

  • (βT,n)+(βT)+, weakly in Ls(M×Rd), as n+.

  • βn,TβT, weakly in Lloc1(M×Rd), as n+.

  • DvunDvu, weakly in Llocr(Um0), as n+.

With these convergences in hand, we are ready to pass to the limit in the Hamilton-Jacobi inequality constraint and the functional. Note that the weak form of the inequality (Definition 3.9) implies that, for all n and all test functions ϕCc((0,T]×M×Rd) such that ϕ0, (6.3) 0TM×Rd(tϕ+v·Dxϕ)un+ϕH(x,v,Dvun)dxdvdt0TM×Rdϕβndxdvdt+M×Rdϕ(T,x,v)βT,ndxdv.(6.3)

Note again that ϕ is compactly supported in Um0. By the weak convergence of Dvun in Llocr(U0) and the convexity of H it follows that 0TM×RdϕH(x,v,Dvu)dxdvdt=Um0ϕH(x,v,Dvu)dxdvdtlim infnUm0ϕH(x,v,Dvun)dxdvdt=lim infn0TM×RdϕH(x,v,Dvun)dxdvdt.

All the other convergences stated above are sufficient to guarantee convergence against ϕ. So, we obtain that the limit (u,β,βT) satisfies Equation(3.6).

Next, we consider the convergence in the functional. In addition to the previous convergences, along the previously chosen subsequence, we have

  • (u0,n)+(u¯0)+, weakly-* in (L+Lq)(M×Rd), as n+.

The convergence of the sequence ((u0,n)m0)n requires special attention. The boundedness of this sequence in L1(M×Rd) lets us conclude that there exists a nonnegative Radon measure ν such that after passing to a subsequence (that we do not relabel) (u0,n)m0ν,asn+.

This means in particular that for all ϕCc(M×Rd), we have M×Rdϕ(u0,n)m0dxdvM×Rdϕν(dxdv),asn+.

Since the the sequence ((u0,n)m0)nN is supported in the open set {m0>0}, we get that spt(ν)spt(m0). Now, let us take ϕCc({m0>0}) arbitrary and define ψ:=ϕ/m0. Since m0C(M×Rd), by assumption, we have that ψCc({m0>0}) and so M×Rdψ(u0,n)m0dxdv=M×Rd(ϕ/m0)(u0,n)m0dxdv=M×Rdϕ(u0,n)dxdvM×Rd(ϕ/m0)ν(dxdv),asn+.

Thus, this means that as n+,(un,0) converges weakly-* to the nonnegative Radon measure (u¯0):=1m0·ν, i.e. (u0) has density 1m0 with respect to ν. We notice that this means that (u¯0) is absolutely continuous with respect to ν. In fact, we also have that ν is absolutely continuous with respect to (u¯0), and so we can write ν=m0·(u¯0).

Let us take now ϕCc(Um0), and test the inequalities satisfied by (un,βn,βT,n), similarly to Equation(6.3), to obtain 0TM×Rd(tϕ+v·Dxϕ)un+ϕH(x,v,Dvun)dxdvdt0TM×Rdϕβndxdvdt+M×Rdϕ(T,x,v)βT,ndxdvM×Rdϕ(0,x,v)u0,ndxdv.

Incorporating also the previously described convergence of (u0,n)n, we can pass to the limit along the chosen subsequence and obtain 0TM×Rd(tϕ+v·Dxϕ)u+ϕH(x,v,Dvu)dxdvdtM×RdβTϕTdxdvM×Rdϕ0u¯0(dxdv)+0TM×Rdϕβdxdvdt, where u¯0:=(u¯0)+(u¯0). We notice that u¯0 is a signed Radon measure, supported in spt(m0).

Having in hand this last inequality, we readily check that the assumptions of EquationLemma A.11 are fulfilled with the choice of β0=u¯0 and βT as before. This means in particular that u satisfies {tuv·Dxu+H(x,v,Dvu)β, in D((0,T)×M×Rd)u0u¯0in D({m0>0});uTβTin D(M×Rd), where when writing the traces u0 and uT, we are referring to the right continuous version of u in time. Since by construction, u¯0,m0=M×Rd(u¯0)+m0dxdv(u¯0),m0 is finite, we have that u0,m0 is meaningful and finite, with u0,m0u¯0,m0.

Lower semicontinuity of the term involving M×Rdm0u0(dxdv).

Claim.

{m0>0}m0(u0)(dxdv)M×Rdm0(u¯0)(dxdv)lim infnM×Rd(u0,n)m0dxdv.

Proof of Claim.

First, notice that since u0u¯0 is a positive distribution, it can be represented by a Radon measure. We may therefore write, for some ν0M+(M×Rd), such that spt(ν0)spt(m0) and u0=u¯0+ν0=[(u¯0)++ν0](u¯0).

It follows that the Hahn-Jordan decomposition of u0 satisfies (u0)+(u¯0)++ν0,(u0)(u¯0).

Now consider any compactly supported function ζCc({m0>0}), such that 0ζ1. Then M×Rdζm0(u¯0)(dxdv)=limnM×Rdζ(u0,n)m0dxdvlim infnM×Rd(u0,n)m0dxdv.

Since (u0)(u¯0) as measures, M×Rdζm0(u0)(dxdv)lim infnM×Rd(u0,n)m0dxdv.

Then take a non-decreasing sequence of functions ζk such that ζk converges pointwise to the indicator function of the set {m0>0} as k tends to infinity: consider for example functions such that ζk(x,v)={1ifm0(x,v)>2k0ifm0(x,v)2(k+1).

This is always possible since m0 is continuous. Then, by monotone convergence, we indeed have M×Rdm0(u0)(dxdv)=limk+M×Rdζkm0(u0)(dxdv)lim infnM×Rd(u0,n)m0dxdv, as desired and the claim follows.

By the weak star convergence of (u0,n)+ to (u¯0)+ in (L+Lq)(M×Rd), we also have that, for u¯0, the positive part (u0,n)+m0 converges to (u¯0)+m0 strongly in L1(M×Rd). Since u0u¯0 as signed measures, we deduce that (6.4) M×Rdm0u0(dxdv)M×Rdm0u¯0(dxdv)lim infnM×Rdu0,nm0dxdv ,(6.4) as required.

The term involving G.

For the term involving G, we notice that by convexity (6.5) M×RdG*(x,v,βT)dxdvlim infn+M×RdG*(x,v,βT,n)dxdv.(6.5)

Indeed, by classical results (cf. [Citation57, Proposition I.2.3, Corollary I.2.2]), this is a consequence of the convexity of the integrand in the last variable and Fatou’s lemma that yields the lower semi-continuity with respect to the strong topology on Lloc1(M×Rd).

The term involving F.

First note that, for functions β such that β+Lq([0,T]×M×Rd) and βLloc1([0,T]×M×Rd), by Equation(3.1) the following inequality holds: |F(x,v,β)|c1|β+|q+CF(x,v)L1([0,T]×M×Rd).

Thus (6.6) 0TM×RdF*(x,v,β)dxdvdt=Um0F*(x,v,β)dxdvdt.(6.6)

Indeed, since Um0={0}×{m0>0}(0,T)×M×Rd, for all δ>0 we have |0TM×RdF*(x,v,β)dxdvdtUm0F*(x,v,β)dxdvdt|0δ{m0>0}|F*(x,v,β)|dxdvdt.

The integrand is bounded by the L1 function |F*(x,v,β)| and converges to zero almost everywhere as δ tends to zero. Thus, taking δ0 we obtain Equation(6.6). A similar equality holds for all βn.

Therefore, by the convexity of F* (and by arguments similar to the one for G*), we conclude that (6.7) 0TM×RdF*(x,v,β)dxdvdt=Um0F*(x,v,β)dxdvdtliminfn+Um0F*(x,v,βn)dxdvdt=liminfn+0TM×RdF*(x,v,βn)dxdvdt.(6.7)

Thus, collecting all the previous arguments, one deduces that A˜(u,β,βT)liminfn+A˜(un,βn,wn).

The thesis of the theorem follows. □

Corollary 6.9.

In the setting and notation of the previous theorem, in fact u0=u¯0 on {m0>0}.

Proof.

Since (u,β,βT) is a minimizer, A˜(u,β,βT)M×Rdm0u¯0(dxdv)+lim infn+0TM×RdF*(x,v,βn)dxdvdt+lim infnM×RdG*(x,v,βT,n)dxdvlimn+(M×Rdm0u0,n(dxdv)+0TM×RdF*(x,v,βn)dxdvdt+M×RdG*(βT,n)dxdv)=A˜(u,β,βT), where in the last equality we have used that (un,βn,βT,n) is a minimizing sequence.

All the above inequalities are therefore equalities. From the inequalities Equation(6.4), Equation(6.5) and Equation(6.7) for each of the terms, we deduce that M×Rdm0u0(dxdv)=M×Rdm0u¯0(dxdv).

It follows that u0=u¯0 as signed measures on {m0>0}. Indeed, first note that u0u¯0 as signed measures, or in other words u0u¯0 is a nonnegative measure. For any non-negative test function ϕ0Cc({m0>0}) we have m0ε>0 on the support of ϕ0, for some ε>0. Thus there exists a constant C such that ϕ0Cm0. Thus 0M×Rdϕ0(u0u¯0)(dxdv)CM×Rdm0(u0u¯0)(dxdv)=0.

Thus u0=u¯0 as signed measures on {m0>0}.

7. Existence and uniqueness of a solution to the MFG system

In this section we prove Theorem 2.4. First, we show that the minimizers of Problems 3.3 and 3.8 that we have obtained in the previous sections provide weak solutions (u, m) of the MFG.

Theorem 7.1.

Let (u,β,βT) be a minimizer of A˜ over KA and let (m, w) be a minimizer of B over KB. Then

  1. β(t,x,v,)=f(x,v,m(t,x,v)) for a.e. (t,x,v)(0,T)×M×Rd, βT(x,v)=g(x,v,mT(x,v)) for a.e. (x,v)M×Rd;

  2. w(t,x,v)=m(t,x,v)DpvH(x,v,Dvu(t,x,v)) for a.e. (t,x,v)(0,T)×M×Rd.

As a consequence, (u, m) is a weak solution to Equation(1.1) in the sense of Definition 2.3.

Proof.

By Theorem 5.1, A˜(u,β,βT)+B(m,w)=0.

Substituting the definitions of the functionals, we obtain 0TM×RdF(x,v,m)+F*(x,v,β)dxdvdtM×Rdm0u0(dxdv)+M×RdG(x,v,mT)+G*(x,v,βT)dxdv+0TM×RdL(x,v,wm)mdxdvdt=0.

Fenchel’s inequality then implies that 0TM×Rdβmdxdvdt+M×RdβTmTm0u0dxdv+0TM×RdL(x,v,wm)mdxdvdt0.

By Corollary 5.4, the left hand side is non-negative, and therefore equality holds: (7.1) 0TM×Rdβmdxdvdt+M×RdβTmTm0u0dxdv+0TM×RdL(x,v,wm)mdxdvdt=0.(7.1)

Moreover, equality also holds almost everywhere in the applications of Fenchel’s inequality. Thus the following hold almost everywhere in [0,T]×M×Rd: (7.2) β=f(x,v,m(t,x,v)),βT=g(x,v,m(T,x,v)).(7.2)

By Equation(7.1) and Corollary 5.5, 0TM×RdmH(x,v,Dvu)+mL(x,v,wm)+Dvu·w dxdvdt0.

By Fenchel’s inequality, the integrand on the right hand side is non-negative; we deduce that equality holds in the above estimate and thus the integrand is equal almost everywhere to zero. It follows that wm=DpvH(x,v,Dvu) almost everywhere on the support of m. Moreover, (7.3) mL(x,v,wm)=m(Dvu·DpvH(x,v,Dvu)H(x,v,Dvu)).(7.3) The energy equality then follows from substituting Equation(7.2) and Equation(7.3) into Equation(7.1). □

We show now, conversely, that weak solutions to the MFG system are in fact minimizers in the corresponding variational problems. The proof of this result follows similar ideas as the corresponding ones from [Citation20,Citation21].

Theorem 7.2.

Let (u, m) be a weak solution to Equation(1.1) in the sense of Definition 2.3. Then by setting β:=f(·,·,m),βT:=g(·,·,mT) and w:=mDpvH(·,·,Dvu), we find that (m, w) is a solution of Problem 3.3, while (u,β,βT) is a solution of Problem 3.8.

Proof.

First let us notice that by Fenchel’s equality one has F(·,·,f(·,·,m))=F(·,·,m)mf(·,·,m).

We define the Borel set B:={(t,x,v)[0,T]×M×Rd:f(x,v,m(t,x,v))0}. Restricted to this set, we find CFF(·,·,f(·,·,m))F(·,·,m),a.e.inB, where in the first inequality we have used our assumptions (3.1). Since, CF, F*(·,·,0) and F(·,·,m) are summable, this implies in particular that F(·,·,f(·,·,m)+)L1([0,T]×M×Rd). Using the growth condition on F* we find furthermore that f(·,·,m)+=β+Lq([0,T]×M×Rd).

Now, on Bc, i.e. when f(·,·,m)0, we find 0mf(·,·,m)=mf(·,·,m)=F(·,·,f(·,·,m))F(·,·,m)supβ<0F(·,·,β)F(·,·,m).

Again, the summability of the right hand side, we find that mf(·,·,m)L1([0,T]×M×Rd). Using the exact same arguments for G*, we find similarly that (βT)+Ls(M×Rd) and mT(βT)L1(M×Rd).

Moreover, we have that DvuLlocr(Um0), mL1([0,T]×M×Rd) and wL1([0,T]×M×Rd;Rd), so (m, w) and (u,β,βT) are admissible competitors for the two optimization problems.

Now, take (u¯,β¯,β¯T) as an admissible competitor for the problem involving the functional A˜. By the convexity and differentiability of F* and G* in their last variable we have A˜(u¯,β¯,β¯T)=0TM×RdF*(x,v,β¯)dxdvdtM×Rdm0u¯0(dxdv)+M×RdG*(β¯T)dxdv0TM×RdF*(x,v,β)dxdvdt+0TM×RdβF*(x,v,β)(β¯β)dxdvdtM×Rdm0u0(dxdv)+M×Rdm0(u0u¯0)(dxdv)+M×RdG*(βT)dxdv+M×RdβTG*(βT)(β¯TβT)dxdv=A˜(u,β,βT)+0TM×Rdm(β¯f(·,·,m))dxdvdt+M×Rdm0(u0u¯0)(dxdv)+M×RdmT(β¯Tg(·,·,mT))dxdv where we have used the fact that mf(·,·,m)L1([0,T]×M×Rd) and mTg(·,·,mT)L1(M×Rd) (by the arguments at the beginning of this proof). Moreover, mβ¯L1([0,T]×M×Rd) and mTβ¯TL1(M×Rd) (cf. Corollary 5.4) and βF*(x,v,β)=βF*(x,v,f(x,v,m))=βF*(x,v,mF(x,v,m))=m,βTG*(x,v,βT)=βG*(x,v,g(x,v,mT))=βTG*(x,v,mTG(x,v,mT))=mT.

Now, using Equation(2.4), one obtains A˜(u¯,β¯,β¯T)A˜(u,β,βT)+0TM×Rdmβ¯dxdvdt+M×RdmTβ¯TdxdvM×Rdm0u¯0(dxdv)+0TM×RdL(·,·,w/m)mdxdvdt,

where in the last line we have used DpvH(·,·,Dvu)·DvuH(·,·,Dvu)=L(·,·,DpvH(·,·,Dvu)).

By Corollary 5.4 we conclude that A˜(u¯,β¯,β¯T)A˜(u,β,βT), as desired.

Using the very same ideas and the convexity of F and G, we can conclude similarly that (m, w) must be a minimizer in Problem 3.3. □

Finally, we show that solutions in the sense of Definition 2.3 are unique, again following similar ideas as the corresponding ones from [Citation21]. One major difference, however, is that we develop a suitable comparison principle for the distributional solutions to the corresponding Hamilton-Jacobi inequalities. This completes the proof of Theorem 2.4.

Proof of Theorem 2.4.

The existence of a weak solution (u, m) follows from combining Theorem 6.8 (existence of a minimizer for A˜), Theorem 5.1 (duality, and the fact that the infimum for B˜ is attained) and Theorem 7.1 (minimizers are weak solutions in the sense of Definition 2.3).

For the uniqueness, we first apply Theorem 7.2 to obtain that for i = 1, 2, (ui,f(·,·,mi),g(·,·,mi(T))) are minimizers of A˜ over KA˜ and (mi,miDpH(·,·,Dvui)) are minimizers of B˜ over KB˜. Since the minimizer of B˜ is unique by strict convexity, m1=m2=:m almost everywhere and m1DpH(·,·,Dvu1)=m2DpH(·,·,Dvu2)=:w almost everywhere.

To show that u1 = u2 almost everywhere on the set {m>0}, we first define u=max{u1,u2}. By Lemma B.2, u also satisfies the Hamilton-Jacobi inequality, with β=f(·,·,m) and βT=g(·,·,mT). Since uiu for i = 1, 2, we have M×Rdm0u0(dxdv)M×Rdm0[ui]0(dxdv), and thus A˜(u,β,βT)A˜(ui,β,βT). Since ui is a minimizer, equality holds. By duality, equality then holds in the energy inequalities of Corollary 5.4 for u and m, with β,βT,w as defined previously. Thus, for almost all t[0,T], M×Rdutmtdxdv=tTM×RdL(x,v,wm)mdxdvdt+tTM×Rdβmdxdvdt+M×RdβTmTdxdv.

The same is true replacing u by ui, and so M×Rdutmtdxdv=M×Rd(ui)tmtdxdv,i=1,2.

Thus, since also uiu, we deduce that ui = u almost everywhere on the set {m>0}.

8. Sobolev estimates on the solutions

In this section, we obtain Sobolev estimates on the optimizers of the variational problems, and hence on weak solutions for the MFG system Equation(1.1). The general idea is to “compare” the optimality of the optimizers in the variational problems with their carefully chosen translates. Then using strong convexity of the data one can deduce differential quotient estimates.

These results are inspired by [Citation23,Citation24]. However, because of the kinetic nature of the model we need completely new ideas when we consider perturbations. So, the estimates that we obtain are on suitable kinetic differential operators applied to the solutions. Another crucial difference between our results and the ones in [Citation23,Citation24] is that our Sobolev estimates in the x and v variables are local in time on (0,T]. The main reason behind this is that we have a weaker notion of trace for u0, that we cannot ensure to be stable under perturbations. This imposed further technical complications that require us to work in the case of r = 2.

We emphasize that these estimates are consequences of the stronger convexity and regularity assumptions on the data stated in Assumption 2.

8.1. Local in time Sobolev estimates

Let ζ:[0,T]R be a smooth cutoff function such that ζ(0)=0 and ζ(t)0 for all t > 0. We define η:[0,T]R as η(t):=0tζ(s)ds.

For competitors (m, w) in Problem 3.3, without loss of generality one might assume the representation w = Vm, for a suitable vector field V. Let δRd with |δ|1 and define mδ(t,x,v):=m(t,x+η(t)δ,v+ζ(t)δ)andVδ(t,x,v):=V(t,x+η(t)δ,v+ζ(t)δ)ζ(t)δ.

We use the notation wδ:=Vδmδ.

We notice that by construction, if (m,w)=(m,Vm) is a distributional solution to Equation(3.3), so is (mδ,wδ)=(mδ,Vδmδ) and mδ(0,·,·)=m0.

Similarly, for competitors (u,β,βT) in Problem 3.8 we define uδ(t,x,v):=u(t,x+η(t)δ,v+ζ(t)δ),βδ(t,x,v):=β(t,x+η(t)δ,v+ζ(t)δ),andβTδ(x,v):=βT(x+η(t)δ,v+ζ(t)δ).

Furthermore, we define Hδ(x,v,ξ):=H(x+η(t)δ,v+ζ(t)δ,ξ)+ζ(t)δ·ξ,Fδ(x,v,θ):=F(x+η(t)δ,v+ζ(t)δ,θ),Gδ(x,v,θ):=G(x+η(t)δ,v+ζ(t)δ,θ).

When computing the Legendre transforms of these functions in their last variables we obtain (Hδ)*(x,v,ξ):=H*(x+η(t)δ,v+ζ(t)δ,ξζ(t)δ),(Fδ)*(x,v,θ):=F*(x+η(t)δ,v+ζ(t)δ,θ),(Gδ)*(x,v,θ):=G*(x+η(t)δ,v+ζ(t)δ,θ).

Let us notice that Hδ satisfies in particular the hypotheses imposed in Assumptions 1. Correspondingly, we define the functionals A˜δ and Bδ and the constraint sets KAδ and KBδ, using the shifted versions of the data functions. By construction, as a consequence of a change of variable formula, the proof of the following lemma is immediate.

Lemma 8.1.

(m, w) is an optimizer of B over KB if and only if (mδ,wδ) is an optimizer of Bδ over KBδ. Similarly, (u,β,βT) is an optimizer of A˜ over KA if and only if (uδ,βδ,βTδ) is an optimizer of A˜δ over KAδ.

Proof.

We provide the proof of one of the statements only, the other ones follow similar steps. Suppose that (mδ,wδ) is an optimizer of Bδ over KBδ. This means in particular the minimality of the quantity 0TM×RdFδ(x,v,mδ)dxdvdt+0TM×Rd(Hδ)*(x,v,wδmδ)mdxdvdt+M×RdGδ(x,v,mTδ(x,v))dxdv=0TM×RdF(x+η(t)δ,v+ζ(t)δ,m(t,x+η(t)δ,v+ζ(t)δ))dxdvdt+0TM×RdH*(x+η(t)δ,v+ζ(t)δ,w(t,x+η(t)δ,v+ζ(t)δ)m(t,x+η(t)δ,v+ζ(t)δ))mdxdvdt+M×RdG(x+η(t)δ,v+ζ(t)δ,mT(x+η(t)δ,v+ζ(t)δ))dxdv=0TM×RdF(x,v,m)dxdvdt+0TM×RdH*(x,v,wm)mdxdvdt+M×RdG(x,v,mT(x,v))dxdv, where in the last equality we have used the change of variables (x,v)(xη(t)δ,vζ(t)δ). So, this means that the minimality of (mδ,wδ), after a change of variables, yields the minimality of (m, w). □

Now we are ready to state the main result of this subsection.

Theorem 8.2.

Suppose that (u, m) is a weak solution to Equation(1.1) in the sense of Definition 2.3 and that (H5), (H6), (H7) hold.

Then, there exists C¯>0 such thatmq21(ηDx+ζDv)mL2((0,T]×M×Rd)C¯,m1/2(ηDx+ζDv)DvuL2((0,T]×M×Rd)C¯andmTs21(η(T)Dx+ζ(T)Dv)mTL2(M×Rd)C¯.

Remark 8.3.

As for Theorem 2.5, this is an informal statement: the result we obtain is on suitable difference quotients as in estimate Equation(8.8) below.

Proof of Theorem 8.2.

Let (un,βn,βT,n)nN be a minimizing sequence for Problem 3.8 such that unCc1([0,T]×M×Rd), βn=tunv·Dxun+H(x,v,Dvun),βT,n=u(T,·,·).

Let us recall that after passing to a subsequence, that we do not relabel, as a consequence of Proposition 6.1, Lemma 6.7 and by Claim 2 in the proof of Theorem 7.1, we have that

  • (βn)+β+, weakly in Lq([0,T]×M×Rd), as n+.

  • (βn)β, weakly in Lloc1(Um0), as n+.

  • (βT,n)+(βT)+, weakly in Ls(M×Rd), as n+.

  • (βT,n)(βT), weakly in Lloc1(M×Rd), as n+.

  • (u0,n)+(u˜0)+, weakly-* in (L+Lq)(M×Rd), as n+.

  • DvunDvu, weakly in Lmr([0,T]×M×Rd), as n+.

Notice that the previous arguments imply also that the subsequence can be chosen such that for all M < 0 (8.1) βn1{βnM}β1{βM},weakly* in (L+Lq)([0,T]×M×Rd)as n+(8.1) and βT,n1{βT,nM}βT1{βTM},weakly* in (L+Ls)(M×Rd)as n+.

Furthermore, by Theorem 7.1, we have that β=f(·,·,m) and βT=g(·,·,mT). Let w=mDpvH(·,·,Dvu).

Fix δRd such that |δ|1 and ζ:[0,T]R as described at the beginning of this subsection.

Now, the main idea is to use unδ as a test function in the weak formulation of the equation satisfied by (m, w) and un as test function in the weak form of the equation satisfied by (mδ,wδ). Then we combine these inequalities with the energy equality Equation(2.4) written for (m, w) and (mδ,wδ), respectively, and rely on the strong convexity and regularity properties of the data to deduce a differential quotient estimate.

Following these steps, we obtain M×Rd[βT,nδmTu0,nδm0]dxdv0TM×Rd(Hδ(x,v,Dvunδ)βnδ)m+Dvunδwdxdt.

We combine this with the energy equality Equation(2.4) for (m, w) to get (8.2) M×Rd(βT,nδg(·,·,mT))mT(u0,nδu0)m0dxdv0TM×Rd(Hδ(x,v,Dvunδ)+H*(x,v,w/m)+Dvunδ·w/mβnδ+f(·,·,m))mdxdvdt.(8.2)

Similarly, using un as a test function in the weak form of the equation for (mδ,wδ) and combining with Equation(2.4) for (mδ,wδ), (8.3) M×Rd[(βT,ngδ(x,v,mTδ))mTδ(u0,nu0δ)m0δ]dxdv0TM×Rd(H(x,v,Dvun)+(Hδ)*(x,v,wδ/mδ)+Dvun·wδ/mδβn+fδ(mδ))mδdxdt(8.3)

Adding Equation(8.2) and Equation(8.3), after some changes of variables (translations) and a Taylor expansion of L, we deduce (8.4) 0TM×Rd(H(x+η(t)δ,v+ζ(t)δ,Dvunδ)+H*(x+η(t)δ,v+ζ(t)δ,w/m)+Dvunδ·w/m)mdxdvdt+0TM×Rd(H(xη(t)δ,vζ(t)δ,Dvunδ)+H*(xη(t)δ,vζ(t)δ,w/m)+Dvunδ·w/m)mdxdvdtM×Rd[βT,n(mTδ+mTδ)2g(x,v,mT)mT]dxdvM×Rd2(u0,nu0)m0dxdv+0TM×Rd[βnδ+βnδ2f(m)]mdxdvdt,+0TM×Rd[H*(x,v,w/m)+H*(x+η(t)δ,v+ζ(t)δ,w/m)ζ(t)δ·Dvunδ]mdxdvdt+0TM×Rd[H*(x,v,w/m)+H*(xη(t)δ,vζ(t)δ,w/m)+ζ(t)δ·Dvunδ]mdxdvdt=M×Rd[βT,n(mTδ+mTδ)2g(x,v,mT)mT]dxdvM×Rd2(u0,nu0)m0dxdv+0TM×Rd[βnδ+βnδ2f(m)]mdxdvdt,+0TM×Rd[ζ(t)δ·Dvunδζ(t)δ·Dvunδ]mdxdvdt+0TM×Rd01ss[η2(t)Dxx2H*+ζ2(t)Dvv2H*](x+τη(t)δ,v+τζ(t)δ,w/m)δ,δmdτdsdxdvdt+0TM×Rd01ss2η(t)ζ(t)Dxv2H*(x+τη(t)δ,v+τζ(t)δ,w/m)δ,δmdτdsdxdvdt.(8.4) where we have also used the facts that by the choice of η and ζ, we have u0,nδ=u0,n,u0δ=u0 and m0δ=m0.

Our aim now is to pass to the limit n+ in Equation(8.4) and derive a differential quotient estimate. For this, we consider each of the terms separately.

  • Step 1. First, we notice that by (H7) and by the fact that |w|rmr1L1([0,T]×M×Rd), there exists C > 0 such that

0TM×Rd01ss[η2(t)Dxx2H*+ζ2(t)Dvv2H*](x+τη(t)δ,v+τζ(t)δ,w/m)δ,δmdτdsdxdvdt+0TM×Rd01ss2η(t)ζ(t)Dxv2H*(x+τη(t)δ,v+τζ(t)δ,w/m)δ,δmdτdsdxdvdtC0|δ|20TM×Rd(|w|rmr1+m)dxdvdtC|δ|2.
  • Step 2. Second, let us notice that by the fact that m(L1Lq)([0,T]×M×Rd) and by Equation(8.1), for any M < 0 we have

limn+0TM×Rdβn±δ1{βn±δM}mdxdvdt=0TM×Rdf±δ(m±δ)1{f±δ(m±δ)M}mdxdvdt.

Therefore, limsupn+0TM×Rd[βnδ+βnδ2f(m)]mdxdvdtlimn+0TM×Rd[βnδ1{βnδM}+βnδ1{βnδM}2f(m)]mdxdvdt=0TM×Rd[fδ(mδ)1{fδ(mδ)M}+fδ(mδ)1{fδ(mδ)M}2f(m)]mdxdvdt.

Now, sending M, we conclude that (8.5) limsupn+0TM×Rd[βnδ+βnδ2f(m)]mdxdvdt0TM×Rd[fδ(mδ)+fδ(mδ)2f(m)]mdxdvdt,(8.5)

where we have used the fact that f(m)m,(fδ(mδ))+,(fδ(mδ))+L1 so that the integrand is upper bounded by an L1 function to allow us to apply the monotone convergence theorem. Since the left hand side of inequality Equation(8.4) is bounded from below by zero, it follows the right hand side of Equation(8.5) is not negative infinity.

By the very same arguments one can conclude that limsupn+M×Rd[βT,n(mTδ+mTδ)2g(x,v,mT)mT]dxdvM×Rd[gδ(mTδ)+gδ(mTδ)2g(mT)]mTdxdv.

  • Step 3. By Young’s inequality, we have

0TM×Rd[ζ(t)δ·Dvunδζ(t)δ·Dvunδ]mdxdvdtC|δ|2+c0TM×Rd|DvunδDvunδ|2mdxdvdt, where c > 0 is an arbitrary constant, and C=C(c,T,ζ)>0.
  • Step 4. By the previous steps we can conclude that

0TM×Rd(H(x±η(t)δ,v±ζ(t)δ,Dvun±δ)+H*(x±η(t)δ,v±ζ(t)δ,w/m)+Dvun±δ·w/m)mdxdvdtc0TM×Rd|DvunδDvunδ|2mdxdvdt

is uniformly bounded above, independently of nN. Let us recall that |w|2mL1((0,T)×M×Rd), and so is H*(x±η(t)δ,v±ζ(t)δ,w/m)mL1((0,T)×M×Rd). Using the growth condition on H, by choosing c > 0 small enough in our application of Young’s inequality we deduce that Dvun±δ is uniformly bounded in Lm2((0,T)×M×Rd;Rd). By a change of variable, one can similarly deduce that Dvun is uniformly bounded in Lm±δ2((0,T)×M×Rd;Rd).

Claim.

After passing to a subsequence that we do not relabel, we have Dvun±δDvu±δ weakly in Lm2((0,T)×M×Rd;Rd), as n+.

Proof of Claim.

By the uniform boundedness of the sequence, we know that there exists a subsequence of it (that we do not relabel) and ξLm2((0,T)×M×Rd;Rd), as weak limit, i.e. 0TM×RdDvun±δ·ϕmdxdvdt0TM×Rdξ·ϕmdxdvdt,ϕLm2((0,T)×M×Rd;Rd),asn+.

Thus, we aim to show that ξ=Dvu±δ. As Dvun±δDvu±δ, weakly in Lloc2(Um0), as n+, we can argue similarly as in the proof of Claim 2, in the proof of Theorem 7.1 to deduce the claim.

  • Step 5. By summarizing, Equation(8.4) implies that

0TM×Rd(H(x+η(t)δ,v+ζ(t)δ,Dvunδ)+H*(x+η(t)δ,v+ζ(t)δ,w/m)+Dvunδ·w/m)mdxdvdt+0TM×Rd(H(xη(t)δ,vζ(t)δ,Dvunδ)+H*(xη(t)δ,vζ(t)δ,w/m)+Dvunδ·w/m)mdxdvdtc0TM×Rd|DvunδDvunδ|2mdxdvdtM×Rd[βT,n(mTδ+mTδ)2g(x,v,mT)mT]dxdvM×Rd2(u0,nu0)m0dxdv+0TM×Rd[βnδ+βnδ2f(m)]mdxdvdt+C|δ|2.

Using the additional assumption (2.10) and the inequality |a+b|22(a2+b2), for c > 0 sufficiently small, one can conclude that there exists c0>0 depending only on the data, such that (8.6) c00TM×Rd|DvunδDvunδ|2mdxdvdtM×Rd[βT,n(mTδ+mTδ)2g(x,v,mT)mT]dxdvM×Rd2(u0,nu0)m0dxdv+0TM×Rd[βnδ+βnδ2f(m)]mdxdvdt+C|δ|2.(8.6)

Now, our aim is to pass to the limit with n+ first in Equation(8.6). For this we take lim infn+ of the left hand side and limsupn+ of the right hand side. We notice that the term M×Rd2u0,nm0dxdv needs special attention, since we do not have upper semicontinuity of it. Because of this, we add to both sides of Equation(8.6) the quantity 20TM×RdF*(x,v,βn)dxdvdt+2M×RdG*(βT,n)dxdv

before passing to the limit. Thus we obtain lim infn+c00TM×Rd|DvunδDvunδ|2mdxdvdt+lim infn+20TM×RdF*(x,v,βn)dxdvdt+lim infn+2M×RdG*(βT,n)dxdvlimsupn+M×Rd[βT,n(mTδ+mTδ)2g(x,v,mT)mT]dxdv+2M×Rdu0m0dxdv+limsupn+0TM×Rd[βnδ+βnδ2f(m)]mdxdvdt+limsupn+(2M×Rdu0,nm0dxdv+20TM×RdF*(x,v,βn)dxdvdt+2M×RdG*(βT,n)dxdv)+C|δ|2.

All the arguments in the previous steps allow us to pass to the limit. By the fact that (un,βn,βT,n) is a minimizing sequence, we get that limsupn+(2M×Rdu0,nm0dxdv+20TM×RdF*(x,v,βn)dxdvdt+2M×RdG*(βT,n)dxdv)limn+(2M×Rdu0,nm0dxdv+20TM×RdF*(x,v,βn)dxdvdt+2M×RdG*(βT,n)dxdv)=2A˜(u,β,βt)=2(M×Rdu0m0dxdv+0TM×RdF*(x,v,β)dxdvdt+2M×RdG*(βT)dxdv).

So, after simplification, one obtains (8.7) c00TM×Rd|DvuδDvuδ|2mdxdvdtM×Rd[gδ(mTδ)+gδ(mTδ)2g(x,v,mT]mTdxdv+0TM×Rd[fδ(mδ)+fδ(mδ)2f(m)]mdxdvdt+C|δ|2.(8.7)

Now, using Equation(2.5) and Equation(2.7) the very same arguments as in [Citation23, computation (4.25)] yield M×Rd(fδ(mδ)+fδ(mδ)2f(m))mdxdvC|δ|2(1+M×Rdmin{mδ,m}qdxdv)cf2M×Rdmin{(mδ)q2,mq2}|mδm|2dxdv.

Similarly, Equation(2.6) and Equation(2.8) yield M×Rd(gδ(mTδ)+gδ(mTδ)2g(mT))mTdxdvC|δ|2(1+M×Rdmin{mTδ,mT}sdxdv)cg2M×Rdmin{(mTδ)s2,mTs2}|mTδmT|2dxdv.

Combining these estimates with Equation(8.7), we get (8.8) c00TM×Rd|DvuδDvuδ|2mdxdvdt+cf20TM×Rdmin{(mδ)q2,mq2}|mδm|2dxdvdt+cg2M×Rdmin{(mTδ)s2,mTs2}|mTδmT|2dxdvC|δ|2.(8.8)

Dividing by |δ|2 and letting δ0, we easily obtain the result. □

8.1.1. Recovering estimates on the operator (tDx+Dv) applied to solutions

By choosing a specific structure for the cutoff function ζ, we can recover estimates on more particular differential operators. Suppose that ζ(t)=0 for t[0,t0/2], and ζ(t)=1, for t(t0,T] for some t0(0,T) (to be chosen to be arbitrary), in such a way that also η(t0)=t0. Then in Theorem 8.2, the operator (ηDx+ζDv), for t>t0 becomes (tDx+Dv). So, one can state the following local in time corollary.

Corollary 8.4.

Suppose that the assumptions of Theorem 8.2 take place. Then, there exists C¯>0 such thatmq21(tDx+Dv)mLloc2((0,T]×M×Rd)C¯,m1/2(tDx+Dv)DvuLloc2((0,T]×M×Rd)C¯andmTs21(TDx+Dv)mTL2(M×Rd)C¯.

8.1.2. Recovering estimates on the operator Dx applied to solutions

Now suppose that η(t)=0 for t[0,t0/2] and η(t)=1 for t(t0,T] (where t0(0,T) can be chosen arbitrarily). We still require that ζ:=η. With this choice of cutoff functions η,ζ, we can formulate the following result as a corollary of Theorem 8.2.

Corollary 8.5.

Suppose that the assumptions of Theorem 8.2 take place. Then, there exists C¯>0 such thatmq21DxmLloc2((0,T]×M×Rd)C¯,m1/2DxDvuLloc2((0,T]×M×Rd)C¯andmTs21DxmTL2(M×Rd)C¯.

8.1.3. Proof of Theorem 2.5

Finally, the proof of Theorem 2.5 follows from the previous two corollaries and the inequality |Dvh||(tDx+Dv)h|+T|Dxh|for allt[0,T],

(8.1.3) for any Sobolev function h.

Acknowledgements

We thank Mikaela Iacobelli for useful discussions regarding Lemma A.4. We thank the two anonymous referees for carefully reading our manuscript and for their valuable comments.

Disclosure statement

The authors report that there are no competing interests to declare.

References

  • Lasry, J.-M., Lions, P.-L. (2006). Jeux à champ moyen I. Le cas stationnaire. C. R. Math. Acad. Sci. Paris. 343(9):619–625. DOI: 10.1016/j.crma.2006.09.019.
  • Lasry, J.-M., Lions, P.-L. (2006). Jeux à champ moyen II. Horizon fini et contrôle optimal. C. R. Math. Acad. Sci. Paris. 343(10):679–684. DOI: 10.1016/j.crma.2006.09.018.
  • Lasry, J.-M., Lions, P.-L. (2007). Mean field games. Jpn. J. Math. 2(1):229–260. DOI: 10.1007/s11537-007-0657-8.
  • Huang, M., Malhamé, R. P., Caines, P. E. (2006). Large population stochastic dynamic games: closed-loop McKean-Vlasov systems and the Nash certainty equivalence principle. Commun. Inf. Syst. 6(3):221–251.
  • Ambrose, D. M. (2018). Strong solutions for time-dependent mean field games with non-separable Hamiltonians. J. Math. Pures. Appl. 113:141–154. (9) DOI: 10.1016/j.matpur.2018.03.003.
  • Ambrose, D. M. Existence theory for non-separable mean field games in Sobolev spaces. Indiana U. Math. J.
  • Cirant, M., Goffi, A. (2020). Lipschitz regularity for viscous Hamilton-Jacobi equations with Lp terms. Ann. Inst. H. Poincaré Anal. Non Linéaire. 37(4):757–784. DOI: 10.1016/j.anihpc.2020.01.006.
  • Cirant, M., Goffi, A. (2021). Maximal Lq-regularity for parabolic Hamilton-Jacobi equations and applications to mean field games. Ann. PDE. 7(2). Paper No. 19, 40. DOI: 10.1007/s40818-021-00109-y.
  • Gomes, D. A., Pimentel, E., Sánchez-Morgado, H. (2016). Time-dependent mean-field games in the superquadratic case. ESAIM: COCV. 22(2):562–580. DOI: 10.1051/cocv/2015029.
  • Gomes, D. A., Pimentel, E., Voskanyan, V. (2016). Regularity Theory for Mean-Field Game Systems. SpringerBriefs in Mathematics. Cham: Springer.
  • Gomes, D. A., Pimentel, E. A., Sánchez-Morgado, H. (2015). Time-dependent mean-field games in the subquadratic case. Comm. Partial Diff. Eq. 40(1):40–76. DOI: 10.1080/03605302.2014.903574.
  • Porretta, A. (2015). Weak solutions to Fokker-Planck equations and mean field games. Arch. Rational Mech. Anal. 216(1):1–62. DOI: 10.1007/s00205-014-0799-9.
  • Carmona, R., Delarue, F. (2013). Probabilistic analysis of mean-field games. SIAM J. Control Optim. 51(4):2705–2734. DOI: 10.1137/120883499.
  • Carmona, R., Delarue, F. (2018). Probabilistic Theory of Mean Field Games with Applications. I, Volume 83 of Probability Theory and Stochastic Modelling. Cham: Springer. Mean field FBSDEs, control, and games.
  • Carmona, R., Delarue, F. (2018). Probabilistic Theory of Mean Field Games with Applications. II, Volume 84 of Probability Theory and Stochastic Modelling. Cham: Springer. Mean field games with common noise and master equations.
  • Carmona, R., Delarue, F., Lacker, D. (2016). Mean field games with common noise. Ann. Probab. 44(6):3740–3803. DOI: 10.1214/15-AOP1060.
  • Achdou, Y., Cardaliaguet, P., Delarue, F., Porretta, A., Santambrogio, F. (2020). Mean field games. Lecture Notes in Mathematics, C.I.M.E. Foundation Subseries, Vol. 2281, Springer.
  • Mészáros, A., Mou, C. (2021). Mean field games systems under displacement monotonicity. arXiv:2109.06687.
  • Cardaliaguet, P. (2015). Weak solutions for first order mean field games with local coupling. In Analysis and Geometry in Control Theory and Its Applications, Vol. 11 of Springer INdAM Ser. Cham: Springer, pp. 111–158.
  • Cardaliaguet, P., Graber, P. (2015). Mean field games systems of first order. ESAIM: Cocv. 21(3):690–722. DOI: 10.1051/cocv/2014044.
  • Cardaliaguet, P., Graber, P., Porretta, A., Tonon, D. (2015). Second order mean field games with degenerate diffusion and local coupling. Nonlinear Differ. Equ. Appl. 22(5):1287–1317. DOI: 10.1007/s00030-015-0323-4.
  • Graber, P. J. (2014). Optimal control of first-order Hamilton-Jacobi equations with linearly bounded Hamiltonian. Appl. Math Optim. 70(2):185–224. DOI: 10.1007/s00245-014-9239-3.
  • Graber, P. J., Mészáros, A. R. (2018). Sobolev regularity for first order mean field games. Ann. Inst. H. Poincaré Anal. Non Linéaire. 35(6):1557–1576. DOI: 10.1016/j.anihpc.2018.01.002.
  • Graber, P. J., Mészáros, A. R., Silva, F. J., Tonon, D. (2019). The planning problem in mean field games as regularized mass transport. Calc. Var. 58(3):115–128. Paper No. DOI: 10.1007/s00526-019-1561-9.
  • P.-L, L. Cours au Collège de France. (2007-2011). www.college-de-france.fr.
  • Munoz, S. (2020). Classical and weak solutions to local first order mean field games through elliptic regularity. arXiv:2006.07367
  • Munoz, S. (2021). Classical solutions to local first order extended mean field games. arXiv:2102.13093.
  • Nourian, M., Caines, P. E., Malhamé, R. P. (2011). Mean field analysis of controlled Cucker-Smale type flocking: Linear analysis and perturbation equations. Proceedings of the 18th IFAC WC, Milan, Italy, pp. 4471–4476. DOI: 10.3182/20110828-6-IT-1002.03639.
  • Achdou, Y., Mannucci, P., Marchi, C., Tchou, N. (2020). Deterministic mean field games with control on the acceleration. NoDEA Nonlinear Diff. Equat. Appl. 27(3). Paper No. 33, 32.
  • Bardi, M., Cardaliaguet, P. (2021). Convergence of some mean field games systems to aggregation and flocking models. Nonlinear Anal. 204:112199. Paper No. 112199, 24 DOI: 10.1016/j.na.2020.112199.
  • Cannarsa, P., Mendico, C. (2020). Mild and weak solutions of mean field game problems for linear control systems. Minimax Theory Appl. 5(2):221–250.
  • Cardaliaguet, P., Mendico, C. (2021). Ergodic behavior of control and mean field games problems depending on acceleration. Nonlinear Anal. 203:112185. Paper No. 112185, 40 DOI: 10.1016/j.na.2020.112185.
  • Achdou, Y., Mannucci, P., Marchi, C., Tchou, N. (2021). Deterministic mean field games with control on the acceleration and state constraintsarXiv:2104.07292
  • Mendico, C. (2021). Singular perturbation problem of mean field game of acceleration. arXiv:2107.08479.
  • Dragoni, F., Feleqi, E. (2018). Ergodic mean field games with Hörmander diffusions. Calc. Var. 57(5):116–122. Paper No. DOI: 10.1007/s00526-018-1391-1.
  • Feleqi, E., Gomes, D., Tada, T. (2020). Hypoelliptic mean field games—a case study. Minimax Theory Appl. 5(2):305–326.
  • Achdou, Y., Kobeissi, Z. (2021). Mean field games of controls: finite difference approximations. Math. Eng. 3(3):1–35. Paper No. 024, 35, DOI: 10.3934/mine.2021024.
  • Gomes, D. A., Voskanyan, V. K. (2016). Extended deterministic mean-field games. SIAM J. Control Optim. 54(2):1030–1055. DOI: 10.1137/130944503.
  • Graber, P. J., Mullenix, A., Pfeiffer, L. (2021). Weak solutions for potential mean field games of controls. NoDEA Nonlinear Diff. Eq. Appl. 28(5). Paper No. 50, 34.
  • Kobeissi, Z. (2021). On classical solutions to the mean field game system of controls. Comm. Partial Diff. Eq. to appear,
  • Santambrogio, F., Shim, W. (2021). A Cucker-Smale inspired deterministic mean field game with velocity interactions. SIAM J. Control Optim. 59(6):4155–4187. DOI: 10.1137/20M1368549.
  • Mimikos-Stamatopoulos, N. (2021). Weak and renormalized solutions to a hypoelliptic mean field games system. arXiv:2105.05777.
  • Orrieri, C., Porretta, A., Savaré, G. (2019). A variational approach to the mean field planning problem. J. Funct. Anal. 277(6):1868–1957. DOI: 10.1016/j.jfa.2019.04.011.
  • Stampacchia, G. (1965). Le problème de Dirichlet pour les équations elliptiques du second ordre à coefficients discontinus. Ann. Inst. Fourier (Grenoble). 15(1):189–258. DOI: 10.5802/aif.204.
  • Golse, F., Lions, P.-L., Perthame, B., Sentis, R. (1988). Regularity of the moments of the solution of a transport equation. J. Funct. Anal. 76(1):110–125. DOI: 10.1016/0022-1236(88)90051-1.
  • Golse, F., Perthame, B., Sentis, R. (1985). Un résultat de compacité pour les équations de transport et application au calcul de la limite de la valeur propre principale d’un opérateur de transport. C. R. Acad. Sci. Paris Sér. I Math. 301(7):341–344.
  • Jabin, P.-E. (2009). Averaging lemmas and dispersion estimates for kinetic equations. Riv. Mat. Univ. Parma. 1(8):71–138.
  • Bouchut, F. (2002). Hypoelliptic regularity in kinetic equations. J. Math. Pures Appl. 81(11):1135–1159. (9) DOI: 10.1016/S0021-7824(02)01264-3.
  • Golse, F., Saint-Raymond, L. (2002). Velocity averaging in L1 for the transport equation. C. R. Acad. Sci. Paris, Ser. 1. 334(7):557–562. DOI: 10.1016/S1631-073X(02)02302-6.
  • Golse, F., Saint-Raymond, L. (2004). The Navier–Stokes limit of the Boltzmann equation for bounded collision kernels. Invent. Math. 155(1):81–161. DOI: 10.1007/s00222-003-0316-5.
  • Han-Kwan, D. (2010). L1 averaging lemma for transport equations with Lipschitz force fields. Kinet. Relat. Models. 3(4):669–683.
  • Cardaliaguet, P., Mészáros, A., Santambrogio, F. (2016). First order mean field games with density constraints: pressure equals price. SIAM J. Control Optim. 54(5):2672–2709. DOI: 10.1137/15M1029849.
  • Prosinski, A., Santambrogio, F. (2017). Global-in-time regularity via duality for congestion-penalized mean field games. Stochastics. 89(6-7):923–942. DOI: 10.1080/17442508.2017.1282958.
  • Santambrogio, F. (2018). Regularity via duality in calculus of variations and degenerate elliptic PDEs. J. Math. Anal. Appl. 457(2):1649–1674. DOI: 10.1016/j.jmaa.2017.01.030.
  • Bouchitté, G., Buttazzo, G., Seppecher, P. (1997). Energies with respect to a measure and applications to low-dimensional structures. Calc. Var. 5(1):37–54. DOI: 10.1007/s005260050058.
  • Ambrosio, L., Gigli, N., Savaré, G. (2008). Gradient flows in metric spaces and in the space of probability measures. 2nd ed. Birkhäuser Verlag, Basel: Lecture Notes in Mathematics ETH Zürich.
  • Ekeland, I., Témam, R. (1999). Convex Analysis and Variational Problems, Volume 28 of Classics in Applied Mathematics. Philadelphia, PA: Society for Industrial and Applied Mathematics (SIAM).
  • Dunford, N., Schwartz, J. T. (1988). Linear Operators, Part 1: General Theory. Wiley Classics Library. John Wiley.
  • Kosygina, E., Varadhan, S. R. S. (2008). Homogenization of Hamilton-Jacobi-Bellman equations with respect to time-space shifts in a stationary ergodic medium. Comm. Pure Appl. Math. 61(6):816–847. DOI: 10.1002/cpa.20220.
  • DiPerna, R. J., Lions, P.-L. (1989). Ordinary differential equations, transport theory and Sobolev spaces. Invent Math. 98(3):511–547. DOI: 10.1007/BF01393835.
  • Bouchut, F. (2001). Renormalized solutions to the Vlasov equation with coefficients of bounded variation. Arch. Ration. Mech. Anal. 157(1):75–90.

Appendix A.

Time regularity

In this appendix, we collect some facts about the regularity with respect to time of solutions u of (A.1) tuv·Dxu+H(x,v,Dvu)β,in D((0,T)×M×Rd).(A.1)

By this we mean that, for any non-negative test function 0ϕCc((0,T)×M×Rd), (A.2) 0TM×Rd(tϕ+v·Dxϕ)u+ϕ H(x,v,Dvu)dxdvdt0TM×Rdβϕdxdvdt.(A.2)

What we discuss is close to the standard theory of distributional solutions. However, in our case technical difficulties arise since, firstly, Equation(A.1) is an inequality and, secondly, we wish to work on the atypical domain Um0. We therefore found it useful to clarify several points. Our main goal is to give a precise sense to the specification of boundary data for this problem at time t = T, and to give a meaning to u0 (the ‘value of u at time t = 0’), which appears in the functional A˜ defining the variational problem.

Throughout this appendix we impose the following summability conditions on the pair (u,β)Lloc1(Um0)×Lloc1(Um0) and that H satisfies Equation(2.1).

Assumption 3.

The pair (u,β)Lloc1(Um0)×Lloc1(Um0) satisfies the following assumptions:

  • The positive part of β satisfies β+Lq([0,T]×M×Rd);

  • DvuLlocr(Um0);

Under Assumption 3, by a density argument the weak form Equation(A.2) extends additionally to test functions in Cc1((0,T)×M×Rd).

Lemma A.1.

Let (u,β)Lloc1(Um0)×Lloc1(Um0) be a distributional solution to Equation(A.1) satisfying Assumption 3. Then

  1. for any ϕCc1((0,T)×M×Rd) the function(0,T)tϕ(t),u(t):=M×Rdϕ(t,x,v)u(t,x,v)dxdvis of locally bounded variation and therefore has a right continuous representative with a countable number of jump discontinuities.

  2. There exists a path (0,T)tu˜tCc1(M×Rd) which is right continuous with respect to the weak-star topology on Cc1(M×Rd) and such that u˜t=ut as elements of (Cc1) for almost every t(0,T).

Proof.

Since β+tu+v·DxuH(x,v,Dvu) is a positive distribution, it is given by some Radon measure ν on (0,T)×M×Rd. We have (A.3) tu=v·Dxu+H(x,v,Dvu)β+ν=:μ,(A.3) which we will use to deduce weak time regularity for u.

Consider a test function ϕCc1((0,T)×M×Rd). The function (A.4) fϕ:(0,T)Rtϕ,u(t):=M×Rdϕ(t,x,v)u(t,x,v)dxdv(A.4) has distributional derivative (A.5) fϕ=ddtϕ,u=tϕ+v·Dxϕ,u+ϕ,H(x,v,Dvu)+νβ.(A.5)

By Assumption 3, u, β and H(x,v,Dvu) are all locally integrable functions on Um0 and so in particular on (0,T)×M×Rd. Thus the distributional derivative fϕ defined in Equation(A.5) is a Radon measure on (0,T), and so the path Equation(A.4) is of locally bounded variation.

It follows that fϕ has a unique right-continuous version. That is, there exists a set Eϕ[0,T] of full measure and a right continuous function f˜ϕ such that f˜ϕ(t)=fϕ(t) for all tEϕ. The function f˜ϕ satisfies f˜ϕ(t)=f˜ϕ(s)+fϕ((s,t]) for all 0<s<t<T.

Now consider ψCc1(M×Rd) (independent of time). The path tu(t),ψ has time derivative ddtψ,u(t)=v·xψ,u+ψ,H(x,v,Dvu)+νβ.

For each compact subset KM×Rd, we define the following Radon measure on (0,T): for A(0,T) Borel, μK(A):=supv:(x,v)K|v|uL1(A×K)+H(·,·,Dvu)βL1(A×K)+ν(A×K).

For the right-continuous versions we have, for all 0<s<t<T and all ψCc1(M×Rd) with support contained in K, (A.6) |f˜ψ(t)f˜ψ(s)|ψC1μK((s,t]).(A.6)

Using these estimates, it is possible to construct a right continuous version of u: that is, a path tu˜tCc1(M×Rd) that is right continuous with respect to the weak-star topology.

Such a construction is classical, but because of the lack of a precise reference in our context, we sketch the main ideas here. Take a countable dense set ZCc1(M×Rd); there is a full measure set E(0,T) such that ψ,u(t)=f˜ψ(t) for all tE and all ψZ, and moreover u(t)Lloc1(M×Rd) (the latter is true for almost all t since u is Lloc1). Then u˜(t),ψ:=f˜ψ(t) defines a bounded linear functional on Z, for all tE. The estimate Equation(A.6) can be used to show that this is in fact true for all t(0,T). The resulting functional u˜(t) extends by density to a continuous linear functional on Cc1(M×Rd). Then the estimate Equation(A.6) can be used to prove that u˜(t),ψ is right continuous for all ψCc1(M×Rd), not just on Z. □

Next, we construct the extension of u˜ to the boundaries t=0,T.

Definition A.2

(Transport shift). Let tR. The operator Tt:Cc(M×Rd)Cc(M×Rd) is defined byTtϕ(x,v)=ϕ(xtv,v).

Remark A.3

(Group property). For any s,tR,TsTt=Ts+t.

Lemma A.4.

Let u be a solution to Equation(A.1) and let u˜ be its right continuous representative, obtained in Lemma A.1. Let ψCc1(M×Rd) be non-negative. Consider the function (0,T)tTtψ,u˜(t). Then

  1. As t tends to T,Ttψ,u˜(t) either tends to a finite limit or to positive infinity.

  2. As t tends to 0+,Ttψ,u˜(t) either tends to a finite limit or to negative infinity.

Proof.

Observe that (t+v·x)Ttψ=0.

It follows that ddtTtψ,u˜(t)=Ttψ,H(x,v,Dvu)+νβ. Then the negative part of the time derivative satisfies [ddtTtψ,u˜(t)]Ttψ,CH+β+L1(0,T).

Thus Ttψ,u˜(t) can be written as the difference of monotone functions, where the decreasing part is absolutely continuous on (0,T) and can be extended to finite limits at the endpoints. By monotonicity, the increasing part either has a finite limit at t = T, or tends to positive infinity; similarly, at t = 0 it either has a finite limit or tends to negative infinity. □

Definition A.5

(Weak traces). For any ψT,ψ0Cc1(M×Rd), let(A.7) ψT,uT:=limtTTtTψT,u˜(t),ψ0,u0:=limt0+Ttψ0,u˜(t).(A.7)

These define linear maps from Cc1(M×Rd) to R{} in the case of u0, and R{+} in the case of uT.

We now suppose that, in addition to the weak Hamilton-Jacobi inequality in the interior Equation(A.2), u satisfies the following: for all non-negative test functions ϕCc1((0,T]×M×Rd) (A.8) 0TM×Rdu[tϕ+divx(vϕ)]+ϕH(x,v,Dvu)dxdvdt0TM×Rdβϕdxdvdt+M×RdβTϕTdxdv.(A.8)

In our setting, we will have that βTLloc1(M×Rd) is a given function whose positive part satisfies (βT)+Ls(M×Rd). In this case, we show below that the time trace uT, enjoys some more properties.

Lemma A.6.

If u satisfies Equation(A.8) with βTM(M×Rd), then uT as defined in Equation(A.7) is a bounded linear functional on Cc1(M×Rd) and uTβT in the sense of distributions: that is, for all ψTCc1(M×Rd) non-negative,ψT,uTψT,βT.

In particular, we have that ψ,uT=+ does not occur for any ψCc1(M×Rd).

Remark A.7.

Since then βTuT is a positive distribution, if βTLloc1(M×Rd) then we in fact have that uT is represented by a signed Radon measure with absolutely continuous positive part.

Proof of Lemma A.6.

In what follows, we will use the right continuous representative of u constructed in Lemma A.1. By the abuse of notation, we write simply u for u˜. Fix ψTCc1(M×Rd) non-negative. For each ε>0 small consider a smooth, non-negative test function ηε:[0,T]R, chosen such that η(t)=0 for all 0tTε and the derivative ηε satisfies 0ηεε1,ηε(t)={0t[0,Tε]ε1t[Tε+ε2,Tε2].

Note that as a consequence of the fundamental theorem of calculus, one has limε0ηε(T)=1.

We define the following non-negative test function ϕεCc1((0,T]×M×Rd): ϕε(t,x,v)=ηε(t)ψT(x+(Tt)v,v).

Substituting this choice of ϕ into Equation(A.8), we obtain (A.9) TεTηε(t)M×Rdu(t)ψT(x+(Tt)v,v)dxdvdtηε(T)M×RdβTψTdxdv+α1(ε).(A.9) where α1(ε)=TεTM×RdβϕεdxdvdtTεTM×RdϕεH(x,v,Dvu)dxdvdt.

Note that limε0α1(ε)=0, since β and H(x,v,Dvu) are locally integrable and ϕε are bounded in L, uniformly in ε.

We exclude the possibility that ψT,uT=+. Indeed, if this occurs, then for any M > 0, there exists ε such that for any t[Tε,T], M<M×Rdu(t)ψT(x+(Tt)v,v)dxdv.

Then by bounding the left hand side of inequality Equation(A.9) from below we obtain that for any ε<ε, M(12ε)ηε(T)M×RdβTψTdxdv+α1(ε).

Taking the limit ε0 gives MM×RdβTψTdxdv.

Since this holds for any M > 0, we derive a contradiction. Thus—using also Lemma A.4uT is in fact a linear map from Cc1(M×Rd) to R. We note also that the map tM×Rdu(t)ψT(x+(Tt)v,v)dxdv extends to a function that is bounded and continuous (from the left) at t = T.

Next, we show that uTβT as functionals on Cc1(M×Rd). We have (A.10) 1εTεTM×RduψT(x+(Tt)v,v)dxdvdtηε(T)M×RdβTψTdxdv+α2(ε),(A.10) where α2(ε):=α1(ε)+TεT(ε1ηε)M×RduψT(x+(Tt)v,v)dxdvdt.

For the second term here we have |TεT(ε1ηε)M×RduψT(x+(Tt)v,v)dxdvdt||TεTε+ε2|ε1ηε||u(t),ψT(x+(Tt)v,v)|dt|+|Tε2T|ε1ηε||u(t),ψT(x+(Tt)v,v)|dt|2εu(t),ψT(x+(Tt)v,v)L[Tε,T], which converges to zero as ε0 since the trajectory u(t),ψT(x+(Tt)v,v) is bounded near t = T. Thus limε0α2(ε)=0.

Taking the limit ε0 in inequality Equation(A.10), we conclude that ψT,uTψT,βT.

Since βTuT is a positive linear functional on Cc1(M×Rd), it is bounded, and therefore uT is also a bounded linear functional on Cc1(M×Rd).

Corollary A.8.

If u satisfies Equation(A.8) with βTM(M×Rd) then, in the notation of Equationequation (A.3), the measure ν extends to a finite Radon measure on (0,T]×M×Rd given by ν(A)=ν(A1(0,T)).

Proof.

We show that, for any non-negative test function ϕCc1((0,T]×M×Rd), (A.11) ν,ϕ1(0,T)<+.(A.11)

It suffices to prove Equation(A.11) for test functions of the form ϕ(t,x,v)=θ(t)Ttψ(x,v), where θCc1(0,T] and ψCc1(M×Rd).

Then ddtϕ,u(t)=θ(t)Ttψ,u+ϕ,H(x,v,Dvu)+νβθ(t)Ttψ,u+ϕ,νCHβ+.

It follows that (once again using the right continuous version of u), for t < T, ν,ϕ1(0,t]ϕ,u(t)0tM×Rdθ(s)Tsψ,u+ϕ(CH+β+)dxdvds.

Taking the limit tT, by definition of uT, ν,ϕ1(0,T)ϕ,uT0TM×Rdθ(s)Tsψ,u+ϕ(CH+β+)dxdvds<+.

We now discuss the trace of u at t = 0: u0 as defined in Definition A.7. ψ,u0 is defined for all ψCc1(M×Rd). Our aim is to give a meaning to the quantity m0,u0, which appears in the definition of the functional A˜. In the case where m0Cc1(M×Rd) this is straightforward, noting that we allow the possible value . We now consider the more general case where m0C(M×Rd).

Lemma A.9.

Assume that, for all ϕCc1({m0>0}),ϕ,u0. Then u0 is represented by a Radon measure on {m0>0}. Furthermore, the positive part (u0)+ has the property that{m0>0}m0d(u0)+(x,v)<+.

Proof.

Let ϕCc1(M×Rd) be non-negative. Since ddtTtϕ,u˜(t)Ttϕ,CH+β+, we have ϕ,u00TTtϕ,CH+β+dt+TTϕ,βT=:Sϕ.

The right hand side is linear in ϕ and satisfies |0TTtϕ,CH+β+dt+TTϕ,βT|ϕL(CH+β+L1(K[0,T])+βTL1(KT)); here KT denotes the set KT:={(x+vT,v):(x,v)K},

where K is the support of ϕ, and K[0,T] is the set K[0,T]:={(x+vt,v):(x,v)K,t[0,T]}.

Thus S defines a bounded linear functional on Cc(M×Rd). In particular it is a distribution; moreover, it is represented by a signed Radon measure.

Observe next that Su0 is a positive linear functional on Cc1({m0>0}), and thus bounded and a distribution. By positivity it is given by a Radon measure ν0 on {m0>0}. We deduce that u0=Sν0.

That is, u0 is a signed Radon measure.

Moreover, from the definition of the Hahn-Jordan decomposition we have the following estimate for the positive part: ϕ,(u0)+0TTtϕ,CH+β+dt+TTϕ,(βT)+.

Let ϕnCc(M×Rd) be an increasing sequence of functions such that ϕn converges to m0 as n. Since supnϕn,(u0)+0TTtm0,CH+β+dt+TTm0,(βT)+CHTm0L1+T1/qβ+Lqm0Lq+(βT)+Lsm0Ls, we conclude that m0,(u0)+ is finite. □

Based on the previous lemma, we make the following definition.

Definition A.10.

We define m0,u0 as follows:

  1. If there exists ϕCc1({m0>0}) such that ϕ,u0=, then we define m0,u0=+.

  2. Otherwise, let ϕnCc({m0>0}) be an increasing sequence of functions such that ϕn converges to m0 as n, and define m0,u0=limnϕn,u0.

This is well-defined (allowing for the possible value +) by Lemma A.9.

Lemma A.11.

Suppose that the assumptions of Lemma A.6 hold and suppose in addition that(A.12) 0TM×Rdu[tϕ+divx(vϕ)]+ϕH(x,v,Dvu)dxdvdt0TM×Rdβϕdxdvdt+M×RdβTϕTdxdvM×Rdβ0ϕ0dxdv,(A.12)

holds for all ϕCc1(Um0), where β0M({m0>0}) is also given. Then for the trace u0 of the right continuous version of u we haveβ0u0,in D({m0>0}),and in particular u0,ψ for any ψCc1({m0>0}).

If in addition we suppose that β0 is such that β0,m0 is meaningful and finite, then u0,m0 is finite andβ0,m0u0,m0.

Proof.

The proof of this result follows the same lines as the proof of Lemma A.6, so we point out only the main differences. Let u stand for the right continuous representative constructed in Lemma A.1. Fix ψ0Cc1({m0>0}) non-negative. For each ε>0 small consider a smooth, non-negative test function ηε:[0,T]R, chosen such that η(t)=0 for all εtT and the derivative ηε satisfies ε1ηε0,ηε(t)={ε1t[ε2,εε2],0t[ε,T].

Note that as a consequence of the fundamental theorem of calculus, one has limε0ηε(0)=1.

We define the following non-negative test function ϕεCc1(Um0): ϕε(t,x,v)=ηε(t)ψT(xtv,v).

Substituting this choice of ϕ into Equation(A.12), we obtain 0εηε(t)M×Rdu(t)ψT(xtv,v)dxdvdtηε(0)β0,ψ0+α1(ε). where α1(ε)=0εM×Rdβϕεdxdvdt0εM×RdϕεH(x,v,Dvu)dxdvdt.

As before, we note that limε0α1(ε)=0. We exclude the possibility that u0,ψ0=. For this, we rewrite the previous inequality as ηε(0)β0,ψ0α1(ε)0εηε(t)M×Rdu(t)ψT(xtv,v)dxdvdt, and use the same arguments as when proving uT,ψT+ in the proof of Lemma A.6.

Therefore, u0 defines a linear map on Cc1({m0>0}). Having this, we can show the inequality β0,ψ0u0,ψ0 in the same way as corresponding inequality in Lemma A.6.

Now, using the Definition A.10, u0,m0 is meaningful, having also the possibility that it is . However, if the additional assumption that β0,m0 is finite takes place, taking a an increasing sequence of test functions, we find that β0,m0u0,m0, so clearly, the latter term cannot be .

Through similar arguments it is possible to justify the existence of weak time traces for competitors m in Problem 3.3, thereby giving meaning to the initial value problem {tm+v·Dxm+divvw=0,in D((0,T)×M×Rd)m|t=0=m0.

Recall that in Remark 3.4 we established that, in the cases of interest to us, there exists a function VLr(mdxdvdt) such that w=Vm, and so we may assume that m is a distributional solution of the following equation: tm+divx(vm)+divv(Vm)=0, with |V|mL1([0,T]×M×Rd).

This setting is much more standard since here the time derivatives ddtϕ,mt will be in L1[0,T] for any ϕCc1(M×Rd) rather than measures, that is, we expect absolutely continuous rather than right continuous trajectories. Moreover we can work on the whole space [0,T]×M×Rd rather than only the reachable set Um0.

Deducing that m has a narrowly continuous representative is essentially an application of [Citation56, Lemma 8.1.2]. However, since we do not necessarily have 0TM×Rd|v|mdxdvdt<+, due to the unbounded drift v, we cannot immediately apply this lemma. Below we briefly sketch the adaptation to our case.

Lemma A.12.

(See [Citation56, Lemma 8.1.2]). Let 0m(L1Lq)([0,T]×M×Rd) be a non-negative function satisfyingtm+divx(vm)+divv(Vm)=0

in the sense of distributions on (0,T)×M×Rd, where V is given, such that |V|mL1([0,T]×M×Rd).

Then there exists a continuous curve m˜:[0,T](Cc1(M×Rd)) such that m˜t=mt for almost all t[0,T]. Thus m˜0 is well-defined as an element of (Cc1(M×Rd)) (or in fact, by positivity, a Radon measure).

Furthermore, if m˜0 is a probability measure, then m˜ extends uniquely to a narrowly continuous curve in the space of probability measures, i.e. m˜C([0,T];P(M×Rd)).

Proof.

Since m,|V|mL1([0,T]×M×Rd), for any compact set KM×Rd we have 0TK(|v|+|V|)mdxdvdt<+.

It follows that, as in the proof of [Citation56, Lemma 8.1.2], we may select a dense subset {ϕn}nN of Cc1(M×Rd) and take a version m˜t of mt such that tϕn,m˜t is continuous with respect to t for all n and m˜t=mt for almost all t[0,T], and a define a unique weak-* continuous extension of m˜ to (Cc1(M×Rd)). Thus m˜0 is well-defined as the unique element of (Cc1(M×Rd)) satisfying ϕ,m˜0:=limt0ϕ,m˜tfor allϕCc1(M×Rd).

Moreover, since m˜ is non-negative, in fact m˜t is a Radon measure on M×Rd for all t[0,T].

Furthermore, it follows from the continuity of m˜ in the weak-* sense of (Cc1(M×Rd)), and the fact that m˜ is locally finite, that, for any (N.B. now time-dependent) ϕCc1([0,T]×M×Rd), the path tϕ(t,·),m˜t is continuous. Thus we may also use the final argument from [Citation56, Lemma 8.1.2] (similar to our argument for the time traces at the boundary in Lemma A.6) to prove the following equality (c.f. [Citation56, EquationEquation (8.1.4)]): for any ϕCc1([0,T]×M×Rd) and any 0t1t2T, (A.13) ϕ(t2,·),m˜t2ϕ(t1,·),m˜t1=t1t2tϕ+v·Dxϕ+V·Dvϕ,m˜tdt.(A.13)

Next we wish to show that, if m˜0 is a probability measure, then m˜t is a probability measure for all t[0,T]. If this is the case, then we may apply [Citation56, Remark 5.1.6]—if m˜tnm˜t in the sense of distributions as n tends to infinity, then this convergence also holds in the narrow sense—to deduce that m˜t is a narrowly continuous path in the space of probability measures, as desired.

To do this we use the argument of Lemma A.4 to avoid the need for vm to be integrable. First, fix a sequence ζRCc(M×Rd) of smooth, compactly supported functions, approximating the constant function 1 in a monotone limit as R tends to infinity—that is, let ζR satisfy the assumptions given in EquationEquation (5.2). We note in particular that DζRLC/R for some constant C > 0 independent of R. Then, for each t, consider the test function TttζR (recall T from Definition A.2), which satisfies (t+v·Dx)TttζR=0,TttζR|t=t=ζR.

Using Equation(A.13) with t1=0,t2=t and ϕ=TttζR, we find that ζR,m˜t=TtζR,m˜0+0tDv(TttζR),Vm˜tdt.

We observe that, for all t[0,t], |Dv(TttζR)|=|[(tt)Dx+Dv]ζR|(1+t)DζRLC(1+t)R.

Thus (since m˜t=mt for almost all t) |0tDv(TttζR),Vm˜tdt|C(1+t)R0TM×Rd|V|mtdxdvdt.

The right hand side tends to zero as R tends to infinity, since |V|mL1([0,T]×M×Rd). Moreover, since m˜0 is a probability measure and TtζR increases monotonically to 1 pointwise as R tends to infinity, it follows that limR0TtζR,m˜0=1.

Finally, since m˜t is a Radon measure and limRζR,m˜t is a monotone limit, m˜t(M×Rd)=limRζR,m˜t=1.

That is, m˜t is a probability measure for all t[0,T]. This completes the proof. □

Appendix B.

Truncations and maxima

Given a distributional solution to the Hamilton-Jacobi inequality (B.1) tuv·Dxu+H(x,v,Dvu)β,in D((0,T)×M×Rd)uTβT,in D(M×Rd)(B.1) in the sense of Definition 3.9, we show that the truncations of u from below, that is, the functions max{u,l} for some l < 0, satisfy a similar inequality. In a similar vein, we also show that given u1 and u2 both satisfying Equation(B.1), their maximum satisfies the same inequality Equation(B.1).

Lemma B.1.

Let uLloc1((0,T)×M×Rd) satisfy Equation(B.1) in the sense of distributions. Assume that βLloc1((0,T)×M×Rd) and DvuLlocr((0,T)×M×Rd). Then ul:=(ul)+ satisfies tulv·Dxul+H(x,v,Dvul)1{u>l}β1{u>l},in D((0,T)×M×Rd),[(ul)+]T(βTl)+in D(M×Rd).

A similar result holds for the truncation ul=(ul)++l. Moreover it suffices to consider the case l = 0.

Lemma B.2.

Let u1, u2 satisfy Assumption 3 and Equation(B.1). Then u=max{u1,u2} also satisfies Equation(B.1).

The result here is in the spirit of renormalization [Citation60]. Bouchut [Citation61, Theorem 1.1] proved a chain rule for the kinetic transport operator, i.e. the identity (B.2) (t+v·Dx)h(u)=h(u)(t+v·Dx)u(B.2) that applies when h is a Lipschitz function and tu+v·DxuLloc1. However, since in our case tu+v·Dxu may only be a measure, we are not able to use this result directly, or indeed prove a chain rule with equality as in EquationEquation (B.2). Nevertheless, the ideas of the proofs in [Citation24, Citation61] can be used to obtain the inequality that is sufficient for our case.

The argument proceeds in several steps.

B.1. Extension

We define the following time shift and extension of u on the time interval (2η,T+2η) for η>0: u˜(t,x,v)={u(t,x,v),t(0,T)0,t[T,T+2η]. Then, defining β˜(t,x,v)={β(t,x,v),t(0,T)H(x,v,0),t[T,T+2η], u˜ satisfies the following inequality in the sense of distributions on (0,T+2η)×M×Rd (see the similar construction in [Citation21, Section 6.3]): tu˜v·Dxu˜+H(x,v,Dvu˜)β˜+βT t1[T,T+2η],in D((0,T+2η)×M×Rd).

B.2. Fenchel’s inequality

Since L is the Fenchel conjugate of H, for any continuously differentiable vector field aCb1 we have tu˜v·Dxu˜a·Dvu˜L(x,v,a)+β˜+βT t1[T,T+2η],in D((0,T+2η)×M×Rd).

B.3. Regularisation

Fix non-negative, symmetric, unit mass mollifiers χ,ψCc(Rd) and θCc(R). Assume that θ is supported on the set [1,1]. Then define, for ε,δ,η>0, χε(v):=εdχ(vε),ψδ(x):=δdψ(xδ),θη(t):=η1θ(tη).

Then define the full mollifier φ by φ(t,x,v)=θη(t)ψδ(x)χε(v).

Then the regularization uη,ε,δ:=u˜φ satisfies the following inequality, in a pointwise sense on (η,T+η]×M×Rd: (t+v·Dx+a·Dv)uη,ε,δLη,ε,δ+βη,ε,δ+φ(tT,·)x,vβT+Eη,ε,δ, where βη,ε,δ:=β˜φ and Lη,ε,δ:=L(x,v,a)φ, while Eη,ε,δ denotes the commutator Eη,ε,δ:=θηt[χεvψδx(v·Dx+a·Dv)u(v·Dx+a·Dv)χεvψδxu].

With the choice ε=δ2 this error converges to zero in Lloc1 by Lemma 5.3 and [Citation60].

B.4. The maximum function

We fix a smooth approximation of the functions x+=max{x,0} and max{x1,x2}. First, for each α>0 we fix a smooth function γα(x) approximating x+ in such a way that 0γα(x)x+ and 0γα(x)1 for all α>0 and limα0γα(x)=x+ for all x and limα0γα(x)=1 for x > 0. Note in particular that then γα(x)=0 for all α>0 and x0, so that γα converges pointwise to the function 1{x>0} as α tends to zero.

We similarly define an approximation hα of the maximum function by hα(x1,x2):=x2+γα(x1x2).

Observe that hα satisfies 0xihα1 for i = 1, 2, and x1hα(x1,x2)+x2hα(x1,x2)=1.

B.5. Inequality for the truncations and maximum

Since γα is non-negative, (B.3) (t+v·Dx+a·Dv)uη,ε,δ γα(uη,ε,δ)[βη,ε,δ+Lη,ε,δ+φ(tT,·)x,vβT+Eη,ε,δ]γα(uη,ε,δ).(B.3)

Thus, since γαC1 and uη,ε,δ is smooth in all variables, applying the usual chain rule (t+v·Dx+a·Dv)γα(uη,ε,δ)[βη,ε,δ+Lη,ε,δ+φ(tT,·)x,vβT+Eη,ε,δ]γα(uη,ε,δ). Similarly, given two subsolutions u1 and u2, (t+v·Dx+a·Dv)hα(uη,ε,δ1,uη,ε,δ2)[βη,ε,δ+Lη,ε,δ+φ(tT,·)x,vβT+Eη,ε,δ1]1hα(uη,ε,δ1,uη,ε,δ2)+[βη,ε,δ+Lη,ε,δ+φ(tT,·)x,vβT+Eη,ε,δ2]2hα(uη,ε,δ1,uη,ε,δ2).

Thus (t+v·Dx+a·Dv)hα(uη,ε,δ1,uη,ε,δ2)βη,ε,δ+Lη,ε,δ+φ(tT,·)x,vβT+Eη,ε,δ11hα(uη,ε,δ1,uη,ε,δ2)+Eη,ε,δ22hα(uη,ε,δ1,uη,ε,δ2).

B.6. Limits

We now take the limit as the smoothing parameters tend to zero in the previous inequalities. This procedure yields the proofs of the Lemmas B.1 and B.2. We continue to choose ε=δ2 to ensure convergence of the commutator. We detail the procedure in the case of the truncation Equation(B.3); the case of the maximum function is similar.

Proof of Lemma B.1.

We first test the inequality Equation(B.3) with an arbitrary non-negative smooth function ζCc((0,T]×M×Rd). We fix an extension of ζ to a function ζCc((0,T+1]×M×Rd) and consider integrating over [0,T+η]×M×Rd. For all η>0 small enough that the support of ζ is contained in (η,T+1]×M×Rd, 0T+ηM×Rdγα(uη,ε,δ)(tζ+v·Dxζ+divv(aζ))dxdvdt0T+ηM×Rd[βη,ε,δ+Lη,ε,δ+Eη,ε,δ]γα(uη,ε,δ)ζdxdvdt+TηT+ηM×Rdγα(uη,ε,δ(t,x,v))θη(Tt)χεvψδxβTζ(t,x,v)dxdvdt.

We have used that uη,ε,δ(T+η,x,v)=0. Since 0γα1, we may estimate the boundary term from above to obtain 0T+ηM×Rdγα(uη,ε,δ)(tζ+vDxζ+divv(aζ))dxdvdt0T+ηM×Rd[βη,ε,δ+Lη,ε,δ+Eη,ε,δ]γα(uη,ε,δ)ζdxdvdt+M×Rd[βT(x,v)]+φζ(T,x,v)dxdv.

Since u˜,β˜,L(x,v,a)Lloc1, uη,ε,δ,βη,ε,δ,Lη,ε,δ converge respectively to these strongly in Lloc1 by standard results on convolutions. We have already noted that Eη,ε,δ converges to zero in Lloc1. We therefore also obtain pointwise convergence along a subsequence. Similarly, φζ converges to ζ pointwise since ζ is smooth. From this we obtain convergence of all terms, by continuity of γα and γα and applying dominated convergence. Hence we obtain the following inequality: 0TM×Rdγα(u)(tζ+vDxζ+divv(aζ))dxdvdt0TM×Rd[β+L(x,v,a)]γα(u)ζdxdvdt+M×Rd[βT(x,v)]+ζ(T,x,v)dxdv.

The convergences as α0 all follow by dominated convergence: for example, since uLloc1((0,T]×M×Rd) and ζ has support contained in (0,T]×M×Rd, we have γα(u)(tζ+v·Dxζ+divv(aζ))|u||(t+v·Dx)ζ|L1.

A similar argument is used for the term involving βT.

For the remaining term, use that |γα|1 (the bound being uniform in α>0), and both β and H(x,v,Dvu) are in Lloc1((0,T]×M×Rd) by assumption. Then 0TM×Rdu+(tζ+vDxζ+divv(aζ))dxdvdt0TM×Rd[β+L(x,v,a)]1{u>0}ζdxdvdt+M×Rd[βT]+ζT dxdv.

Finally, taking a sequence of vector fields a converging in Lr((0,T)×M×Rd) to DpH(x,v,Dvu+), we conclude that 0TM×Rdu+(tζ+vDxζ)dxdvdt0TM×Rd[βH(x,v,Dvu+)]1{u>0}ζdxdvdt+M×Rd[βT]+ζT dxdv, that is, the following holds in the sense of distributions: tu+v·Dxu++H(x,v,Dvu+)1{u>0}β1{u>0},in D((0,T)×M×Rd)[u+]T(βT)+in D(M×Rd).