This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Convergence/divergence phenomena in the vanishing discount limit of Hamilton-Jacobi equations

Andrea Davini, Panrui Ni, Jun Yan and Maxime Zavidovique Andrea Davini
Dipartimento di Matematica
Sapienza Università di Roma
P.le Aldo Moro 2, 00185 Roma
Italy
[email protected] Panrui Ni
Sorbonne Université, Université de Paris Cité, CNRS, Institut de Mathématiques de Jussieu-Paris Rive Gauche
Paris 75005
France
[email protected] Jun Yan
School of Mathematical Sciences
Fudan University
Shanghai 200433
China
[email protected] Maxime Zavidovique
Sorbonne Université, Université de Paris Cité, CNRS, Institut de Mathématiques de Jussieu-Paris Rive Gauche
Paris 75005
France
[email protected]
Abstract.

We study the asymptotic behavior of solutions of an equation of the form

G(x,Dxu,λu(x))=c0in MG\big{(}x,D_{x}u,\lambda u(x)\big{)}=c_{0}\qquad\hbox{in $M$} (*)

on a closed Riemannian manifold MM, where GC(TM×)G\in C(T^{*}M\times\mathbb{R}) is convex and superlinear in the gradient variable, is globally Lipschitz but not monotone in the last argument, and c0c_{0} is the critical constant associated with the Hamiltonian H:=G(,,0)H:=G(\cdot,\cdot,0). By assuming that uG(,,0)\partial_{u}G(\cdot,\cdot,0) satisfies a positivity condition of integral type on the Mather set of HH, we prove that any equi-bounded family of solutions of (*Convergence/divergence phenomena in the vanishing discount limit of Hamilton-Jacobi equations) uniformly converges to a distinguished critical solution u0u_{0} as λ0+\lambda\to 0^{+}. We furthermore show that any other possible family of solutions uniformly diverges to ++\infty or -\infty. We then look into the linear case G(x,p,u):=a(x)u+H(x,p)G(x,p,u):=a(x)u+H(x,p) and prove that the family (uλ)λ(0,λ0)(u_{\lambda})_{\lambda\in(0,\lambda_{0})} of maximal solutions to (*Convergence/divergence phenomena in the vanishing discount limit of Hamilton-Jacobi equations) is well defined and equi-bounded for λ0>0\lambda_{0}>0 small enough. When aa changes sign and enjoys a stronger localized positivity assumption, we show that equation (*Convergence/divergence phenomena in the vanishing discount limit of Hamilton-Jacobi equations) does admit other solutions too, and that they all uniformly diverge to -\infty as λ0+\lambda\to 0^{+}. This is the first time that converging and diverging families of solutions are shown to coexist in such a generality.

Key words and phrases:
weak KAM Theory, vanishing discount problems, Mather measures, viscosity solution theory
1991 Mathematics Subject Classification:

Introduction

In this paper we are concerned with the asymptotic behavior of solutions of an equation of the form

G(x,Dxu,λu(x))=c0in MG\big{(}x,D_{x}u,\lambda u(x)\big{)}=c_{0}\qquad\hbox{in $M$} (Eλ)

posed on a closed Riemannian manifold MM, where GC(TM×)G\in C(T^{*}M\times\mathbb{R}) is convex and superlinear in the gradient variable, is globally Lipschitz in the last argument, and c0c_{0} is the critical constant associated with the Hamiltonian H:=G(,,0)H:=G(\cdot,\cdot,0). We refer the reader to Section 1.3 for the definition of c0c_{0} and of the other related objects coming from weak KAM Theory that will be mentioned in this introduction. The monotonicity condition on GG in the last argument, that is standard for these kind of equations, is dropped here in favor of the following much weaker integral condition

  • (L5)

    TMLGu(x,v,0)𝑑μ~(x,v)<0for all μ~𝔐~,\displaystyle\int_{TM}\dfrac{\partial L_{G}}{\partial u}(x,v,0)\,d\tilde{\mu}(x,v)<0\qquad\hbox{for all $\tilde{\mu}\in\widetilde{\mathfrak{M}}$,}

where LGL_{G} is the convex conjugate function of GG, 𝔐~\widetilde{\mathfrak{M}} denotes the set of Mather measures for LG(,,0)L_{G}(\cdot,\cdot,0), and GG (and hence LGL_{G}) satisfies a C1C^{1}-type regularity condition near u=0u=0, see conditions (G4) and (L4) in Section 2. Under these assumptions, we prove that any equi-bounded family of solutions of (Eλ) uniformly converges to a distinguished critical solution u0u_{0} as λ0+\lambda\to 0^{+}. We furthermore show that any family consisting of other possible solutions uniformly diverges to ++\infty or -\infty.

We underline that the conditions presented above on GG are not sufficient to guarantee the existence and uniqueness of viscosity solutions to (Eλ). This is due to the fact that, without global strict monotonicity of GG with respect to uu, there is no comparison principle. This also makes unenforceable Perron’s method, which is the technique customarily employed to prove existence of solutions.

The general issue of existence and uniqueness of such solutions is subsequently addressed in the paper in the linear case G(x,p,u):=a(x)u+H(x,p)G(x,p,u):=a(x)u+H(x,p) under the minimal hypotheses on aC(M)a\in C(M) and HC(TM)H\in C(T^{*}M) that guarantee that the conditions on GG and LGL_{G} mentioned above are in force. We prove that the family (uλ)λ(0,λ0)(u_{\lambda})_{\lambda\in(0,\lambda_{0})} of maximal solutions to (Eλ) is well defined and equi-bounded for λ0>0\lambda_{0}>0 small enough. When aa changes sign and a0a\geqslant 0 in a neighborhood of the projected Aubry set 𝒜\mathcal{A}, we show that equation (Eλ) does admit other solutions too that uniformly diverge to -\infty as λ0+\lambda\to 0^{+}. If we additionally assume a>0a>0 on 𝒜\mathcal{A}, we furthermore show that any family made up of solutions to (Eλ) that differ from the maximal ones uniformly diverge to -\infty. Incidentally, this completely solves the vanishing discount problem for this model case under the sole assumption that a>0a>0 on 𝒜\mathcal{A} in view of the results established in [41], where aa was additionally assumed nonnegative on MM.

Condition (L5) was introduced in [7] and therein employed to solve the vanishing discount problem for an equation of the form (Eλ) under the same set of assumptions considered herein, plus the additional requirement that GG is globally non-decreasing in uu. Condition (L5) can be read as a strict monotonicity condition on GG with respect to uu, and this is transparent in the linear case G(x,p,u):=a(x)u+H(x,p)G(x,p,u):=a(x)u+H(x,p). What we find striking about the output of our study is the fact that (L5) is a very weak requirement: it implies that aa has to be strictly positive only on some portions of the projected Mather set \mathcal{M}, where the latter is the minimal closed set that contains the projection of the supports of all Mather measures. This set \mathcal{M} can be very small, such as a finite set of points, see Remark 1.10. Furthermore, it has been conjectured by Mañé [33] that for generic Hamiltonians HH both 𝒜\mathcal{A} and \mathcal{M} coincide with the support of a closed curve. Many results have been obtained in this direction, see for example [8] and the references therein, showing that condition (L5) generically leaves a lot of space for aa to take negative values.

The main results proved in this paper keep holding when the superlinearity condition on GG is relaxed in favor of a simple coercivity. We have decided not to pursue this generalization here since that would add additional technicalities with the drawback of hiding the ideas at the base of our work, see Remark 1.12 for further details.

History of the problem

The so-called ergodic approximation is a technique introduced in [32] to study the existence of solutions of the Hamilton-Jacobi equation111All solutions in the paper are meant in the viscosity sense. The definition will be provided later.

H(x,Dxu)=cin MH(x,D_{x}u)=c\qquad\hbox{in $M$} (1)

on the flat dd-dimensional torus M:=𝕋dd/dM:=\mathbb{T}^{d}\simeq\mathbb{R}^{d}/\mathbb{Z}^{d}, where the Hamiltonian HH is a continuous function on TMT^{*}M, coercive in the gradient variable, uniformly with respect to xMx\in M, and cc is a real number. Let λ>0\lambda>0 and uλu_{\lambda} be the unique solution of

λu(x)+H(x,Dxu)=0in M.\lambda u(x)+H(x,D_{x}u)=0\qquad\hbox{in $M$}.

According to [32], the functions λuλ-\lambda u_{\lambda} uniformly converge on MM, as λ0+\lambda\to 0^{+}, to a constant c0c_{0}. Furthermore, the solutions (uλ)λ>0(u_{\lambda})_{\lambda>0} are equi-Lipschitz, yielding, by the Arzelá-Ascoli Theorem, that the functions uλminxMuλu_{\lambda}-\min_{x\in M}u_{\lambda} uniformly converge, along subsequences as λ0+\lambda\to 0^{+}, to a solution of (1) with cc equaling c0c_{0}. The constant c0c_{0} is called critical value of HH and is characterized by the property of being the unique constant cc\in\mathbb{R} such that (1) admits solutions. At that time, it was not clear if different converging sequences yield the same limit. Some constraints on the possible limit solutions were subsequently found in [22, 29], but the breakthrough came with the work [12], where the authors proved that the unique solution uλu_{\lambda} of

λu(x)+H(x,Dxu)=c0in M\lambda u(x)+H(x,D_{x}u)=c_{0}\qquad\hbox{in $M$} (HJλ{\textrm{HJ}}_{\lambda})

converges to a distinguished solution of

H(x,Dxu)=c0in M,H(x,D_{x}u)=c_{0}\qquad\hbox{in $M$}, (HJ0)

as λ0+\lambda\rightarrow 0^{+} under the sole additional assumption that HH is convex in the gradient variable. The proof relies on techniques and tools issued from weak KAM Theory, in particular on the concept of Mather measure, and it works whenever MM is a closed Riemannian manifold. This kind of problem is also known as the vanishing discount problem. When the convexity condition on HH is dropped, the functions uλu_{\lambda} may not converge, as it was pointed out in [44] through a counterexample posed on the 1-dimensional torus.

As a nonlinear generalization (see [6, 5, 23] and [39]), one can study the uniform convergence of the unique solution of

Hλ(x,Dxu,u(x))=c0in M,H_{\lambda}\big{(}x,D_{x}u,u(x)\big{)}=c_{0}\qquad\hbox{in $M$},

as λ0+\lambda\rightarrow 0^{+}, where Hλ(x,p,u)H_{\lambda}(x,p,u) is strictly increasing in uu, and uniformly converges to H(x,p)H(x,p) on compact sets as λ0+\lambda\rightarrow 0^{+}. This kind of problem is called the vanishing contact structure problem. The vanishing discount problem falls in this framework as a particular case by choosing Hλ(x,p,u):=λu+H(x,p)H_{\lambda}(x,p,u):=\lambda u+H(x,p).

The asymptotic convergence result has been subsequently established in many different situations. For the second order case, one can refer to [26, 27, 34, 43]. For the discrete case, one can refer to [11, 38, 42] and also [2] in the context of twist maps. For the similar problem in the mean field game theory, one can refer to [4]. For the weakly coupled Hamilton-Jacobi systems, one can refer to [17, 14, 24, 25]. For the non-compact setting, one can refer to [18, 28].

A natural and challenging question is to weaken the hypothesis on the monotonicity of the Hamiltonian. A first degenerate case was studied in [41], where the author considered the convergence of the solution of

λa(x)u(x)+H(x,Dxu)=c0in M\lambda a(x)u(x)+H(x,D_{x}u)=c_{0}\qquad\hbox{in $M$}

as λ0+\lambda\rightarrow 0^{+}, where a(x)0a(x)\geqslant 0 on MM, and a(x)>0a(x)>0 on the projected Aubry set of HH. Inspired by the works [5, 41], the authors studied in [7] the vanishing discount problem for contact Hamilton-Jacobi equations of the form (Eλ), where the positivity hypothesis on aa assumed in [41] is weakened and generalized by introducing the non-degeneracy integral condition (L5). This work highlights once more that the concept of Mather measure plays a central role in the convergence result.

It is worth pointing out that all works mentioned above required a global non-decreasing hypothesis of the Hamiltonians in uu. If the discounted equation is not increasing in the unknown function uu, solutions may even not exist, and, if they exist, they may be not unique. In [15, 40], the authors discussed the uniform convergence of the minimal solution of (HJλ{\textrm{HJ}}_{\lambda}) as λ0\lambda\rightarrow 0^{-}. For the non-monotone vanishing discount problem, the second author provided the first example in [35] of nonconvergence. In this example, there exist a convergent family of solutions and a divergent family of solutions at the same time. This phenomenon is new comparing with all the previous works in this direction. In the present paper, we show that the example in [35] is in fact a very general phenomenon when the Hamiltonian is continuous, convex and superlinear in the fibres.

Let us conclude by mentioning that the type of problems we study are also closely linked to optimization problems in economics. The discount factor then models the effect of time through interest rates or inflation. Negative interest rates or deflation have been studied by economists (see for instance [31]). Our results then give possible asymptotics in the presence of coexisting inflation and deflation.

Presentation of our results

We present here our main results. Section 2 contains our analysis on the asymptotic behavior of possible solutions of a general contact Hamilton-Jacobi equation of the form (Eλ) when the discount factor λ\lambda goes to 0. The Hamiltonian G(x,p,u)G(x,p,u) is assumed convex and superlinear in pp, and globally Lipschitz in uu, see conditions (G1-3) in Section 2. It is also assumed to satisfy a C1C^{1}-type regularity condition in uu near u=0u=0, see condition (G4) in Section 2. The latter is for instance satisfied when the map uG(x,p,u)u\mapsto G(x,p,u) is C1C^{1} in a neighborhood of u=0u=0 in the following sense:

  • (G4)

    there exists ε>0\varepsilon>0 such that Gu(x,p,u)\frac{\partial G}{\partial u}(x,p,u) exists for all (x,p,u)TM×(ε,ε)(x,p,u)\in T^{*}M\times(-\varepsilon,\varepsilon) and is continuous in TM×(ε,ε)T^{*}M\times(-\varepsilon,\varepsilon).

Let us consider

G(x,Dxu,λu(x))=c0in M,G\big{(}x,D_{x}u,\lambda u(x)\big{)}=c_{0}\qquad\hbox{in $M$}, (Eλ)

and the limit equation

G(x,Dxu,0)=c0in M,G(x,D_{x}u,0)=c_{0}\qquad\hbox{in $M$}, (E0{\textrm{E}}_{0})

where c0c_{0} is the critical value associated with H:=G(,,0)H:=G(\cdot,\cdot,0). We will furthermore assume the non-degeneracy integral condition (L5), where 𝔐~\widetilde{\mathfrak{M}} denotes the set of Mather measures for L:=LG(,,0)L:=L_{G}(\cdot,\cdot,0).

The main results contained in Section 2 can be summarized as follows, see Theorems 2.2 and 2.5.

Theorem 1.

Under the previous assumptions, there exist a viscosity solution u0u_{0} of (E0{\textrm{E}}_{0}) and functions φ:(0,1)(,+]\varphi:(0,1)\to\mathbb{(}-\infty,+\infty], ψ:(0,1)[,+)\psi:(0,1)\to\mathbb{[}-\infty,+\infty) and θ:(0,1)\theta:(0,1)\to\mathbb{R} with

limλ0ψ(λ)=,limλ0φ(λ)=+,limλ0θ(λ)=0\lim_{\lambda\to 0}\psi(\lambda)=-\infty,\quad\quad\lim_{\lambda\to 0}\varphi(\lambda)=+\infty,\quad\quad\lim_{\lambda\to 0}\theta(\lambda)=0

such that, if vλv_{\lambda} is a solution of (Eλ) for some λ>0\lambda>0, then either one of the following alternatives occurs:

  • (i)

    vλψ(λ)v_{\lambda}\leqslant\psi(\lambda);

  • (ii)

    vλφ(λ)v_{\lambda}\geqslant\varphi(\lambda);

  • (iii)

    vλu0θ(λ)\|v_{\lambda}-u_{0}\|_{\infty}\leqslant\theta(\lambda).

We stress that all these three behaviors can happen at the same time for a properly chosen Hamiltonian GG. Indeed, consider G(x,p,u):=sin(u)+px2G(x,p,u):=\sin(u)+\|p\|_{x}^{2}. It is easlily seen that GG verifies all of the above conditions. The limit Hamiltonian is H(x,p):=G(x,p,0)=px2H(x,p):=G(x,p,0)=\|p\|_{x}^{2}, its critical constant is c0=0c_{0}=0 and the constant functions are the only solutions to H(x,Dxu)=0H(x,D_{x}u)=0 in MM. Moreover, for all λ>0\lambda>0, the three constant functions uλ=0u_{\lambda}=0, and uλ±=±πλu_{\lambda}^{\pm}=\pm\frac{\pi}{\lambda} are solutions to the discounted equation.

Theorem 1 shows in great generality that the only possible asymptotic behavior of families of solutions (vλ)λ(0,λ0)(v_{\lambda})_{\lambda\in(0,\lambda_{0})} is, up to subsequences, either to uniformly diverge to ±\pm\infty, or to uniformly converge to a specific solution u0u_{0} of (E0{\textrm{E}}_{0}). We also provide two characterizations of u0u_{0}, see Theorems 2.2 and 2.3.

The general problem of existence and uniqueness of such solutions is addressed in Section 3. Here we consider the linear case G(x,p,u):=a(x)u+H(x,p)G(x,p,u):=a(x)u+H(x,p) under the minimal hypotheses on aC(M)a\in C(M) and HC(TM)H\in C(T^{*}M) that guarantee that the conditions on GG and LGL_{G} presented above are in force. The discounted equation is then

λa(x)u(x)+H(x,Dxu)=c0in M,\lambda a(x)u(x)+H(x,D_{x}u)=c_{0}\qquad\hbox{in $M$}, (HJλ{\textrm{HJ}}_{\lambda})

with limit equation

H(x,Dxu)=c0in M.H(x,D_{x}u)=c_{0}\qquad\hbox{in $M$}. (HJ0)

We prove the following existence and convergence result:

Theorem 2.

Under the previous hypotheses, there is λ0>0\lambda_{0}>0 such that, for all λ(0,λ0)\lambda\in(0,\lambda_{0}), the equation (HJλ{\textrm{HJ}}_{\lambda}) admits a maximal viscosity solution uλC(M)u_{\lambda}\in C(M). Moreover, the family (uλ)λ(0,λ0)(u_{\lambda})_{\lambda\in(0,\lambda_{0})} is equi-bounded, hence it uniformly converges to u0u_{0} as λ0+\lambda\to 0^{+}.

The previous theorem excludes the possibility of families of solutions that uniformly diverge to ++\infty, but leaves open the possibility of families uniformly diverging to -\infty. By strengthening the non-degeneracy integral condition (L5) with a pointwise positivity condition on aa on the projected Aubry set 𝒜\mathcal{A}, we are able to improve the statement of Theorem 1 as follows.

Theorem 3.

Let us additionally assume that a0a\geqslant 0 in a neighborhood of 𝒜\mathcal{A} and that there exist x0Mx_{0}\in M such that a(x0)<0a(x_{0})<0. Then there exists a family of solutions (vλ)λ(0,λ^)(v_{\lambda})_{\lambda\in(0,\hat{\lambda})} to (HJλ{\textrm{HJ}}_{\lambda}) for some λ^(0,1)\hat{\lambda}\in(0,1) uniformly diverging to -\infty as λ0+\lambda\to 0^{+}.

When the previous hypothesis is reinforced, we obtain a stronger conclusion:

Theorem 4.

Let us additionally assume that a>0a>0 on 𝒜\mathcal{A}. Then any family (vλ)λ(0,λ)(v_{\lambda})_{\lambda\in(0,\lambda^{\prime})} of solutions to (HJλ{\textrm{HJ}}_{\lambda}) satisfying vλuλv_{\lambda}\neq u_{\lambda} for all λ(0,λ)\lambda\in(0,\lambda^{\prime}), with λ(0,1)\lambda^{\prime}\in(0,1), uniformly diverges to -\infty.

Namely, in this last case, (uλ)λ(u_{\lambda})_{\lambda} is the only converging family of solutions.

Let us conclude this presentation by stressing that, combining this analysis with the results of [41], we fully understand the asymptotic behavior of solutions to (HJλ{\textrm{HJ}}_{\lambda}) when a>0a>0 on 𝒜\mathcal{A}. The output is the following:

  • (a)

    if a0a\geqslant 0 on the whole MM, (this situation was considered in [41]) then, for all λ>0\lambda>0, there is a unique solution uλu_{\lambda} to (HJλ{\textrm{HJ}}_{\lambda}), and the family (uλ)λ(u_{\lambda})_{\lambda} converges as λ0+\lambda\to 0^{+}.

  • (b)

    if there exist a point x0Mx_{0}\in M such that a(x0)<0a(x_{0})<0, (this situation is discussed in the present paper), then we can find a λ0>0\lambda_{0}>0 small enough such that equation (HJλ{\textrm{HJ}}_{\lambda}) admits at least two solutions for every λ(0,λ0)\lambda\in(0,\lambda_{0}). The family (uλ)λ(0,λ0)(u_{\lambda})_{\lambda\in(0,\lambda_{0})} of maximal solutions uniformly converges to u0u_{0} as λ0+\lambda\to 0^{+}. Any family of other solutions (vλ)λ(v_{\lambda})_{\lambda} uniformly diverges to -\infty as λ0+\lambda\to 0^{+}.

1. Preliminaries

1.1. Notation

  • \circ

    Throughout this paper, we assume that MM is a closed, connected and smooth Riemannian manifold.

  • \circ

    We fix gg an auxiliary Riemannian metric on MM. Let d(x,y)d(x,y) be the distance between xx and yy in MM induced by gg. By compactness of MM our results are independent on the choice of gg.

  • \circ

    Let diam(M)(M) be the diameter of MM.

  • \circ

    We denote by TMTM and TMT^{*}M the tangent and cotangent bundle over MM respectively. We denote by (x,p)(x,p) and (x,v)(x,v) points of TMT^{*}M and TMTM respectively.

  • \circ

    Let π:TM,TMM\pi:TM,T^{*}M\to M denote both canonical projections, the context will make it clear which one is considered.

  • \circ

    We denote by x\|\cdot\|_{x} the norm on both TxMT_{x}M and TxMT_{x}^{*}M induced by gg.

  • \circ

    If NN is a smooth manifold, we will denote by C(N)C(N) the Polish space of continuous functions from NN to \mathbb{R} endowed with the metric of local uniform convergence on NN. We will denote by Cc(N)C_{c}(N) (resp. C1(N)C^{1}(N)) the set of compactly supported (resp., C1C^{1}) functions from NN to \mathbb{R}.

  • \circ

    We denote by 𝒫(TM)\mathscr{P}(TM) the space of Borel probability measures on TMTM endowed with the weak-* topology coming from the dual (Cc(TM))\big{(}C_{c}(TM)\big{)}^{\prime}.

  • \circ

    We denote C(TM)C_{\ell}(TM) the set of continuous functions g:TMg:TM\to\mathbb{R} with at most linear growth meaning that

    sup(x,v)TM|g(x,v)|1+vx<+.\sup\limits_{(x,v)\in TM}\frac{|g(x,v)|}{1+\|v\|_{x}}<+\infty.

    This last quantity defines a norm g\|g\|_{\ell} on the vector space C(TM)C_{\ell}(TM).

  • \circ

    \mathbb{N} denotes the set of positive integers.

1.2. Viscosity solutions.

We start by recalling the notion of viscosity solution.

Definition 1.1.

Let G:TM×G:T^{*}M\times\mathbb{R}\to\mathbb{R} be a continuous function, cc\in\mathbb{R}, and consider the equation

G(x,Dxu,u(x))=cin M.G\big{(}x,D_{x}u,u(x)\big{)}=c\qquad\hbox{in $M$}. (1.1)
  • (a)

    We say that uC(M)u\in C(M) is a viscosity subsolution of (1.1), denoted by

    G(x,Dxu,u(x))cin M,G\big{(}x,D_{x}u,u(x)\big{)}\leqslant c\qquad\hbox{in $M$},

    if, for all φC1(M)\varphi\in C^{1}(M) and x0Mx_{0}\in M such that uφu-\varphi has a local maximum at x0x_{0}, we have G(x0,Dx0φ,u(x0))cG\big{(}x_{0},D_{x_{0}}\varphi,u(x_{0})\big{)}\leqslant c. Such a function φ\varphi is termed supertangent to uu at x0x_{0}.

  • (b)

    We say that uC(M)u\in C(M) is a viscosity supersolution of (1.1), denoted by

    G(x,Dxu,u(x))cin M,G\big{(}x,D_{x}u,u(x)\big{)}\geqslant c\qquad\hbox{in $M$},

    if, for all φC1(M)\varphi\in C^{1}(M) and x0Mx_{0}\in M such that uφu-\varphi has a local minimum at x0x_{0}, then G(x0,Dx0φ,u(x0))cG\big{(}x_{0},D_{x_{0}}\varphi,u(x_{0})\big{)}\geqslant c. Such a function φ\varphi is termed subtangent to uu at x0x_{0}.

  • (c)

    We say that uC(M)u\in C(M) is a viscosity solution of (1.1) if it is both a viscosity sub and supersolution.

In this paper, solutions, subsolutions, supersolutions will be always meant in the viscosity sense and implicitly assumed continuous. We recall that, if uu is C1C^{1} on an open set UU, then it is a viscosity solution (resp. subsolution, supersolution) in UU if and only if it is a pointwise solution (resp. subsolution, supersolution) in UU.

The following stability result is well known, see for instance [3].

Proposition 1.2.

Let (Gn)n(G_{n})_{n} and (un)n(u_{n})_{n} be two sequences of functions in C(TM×)C(T^{*}M\times\mathbb{R}) and C(M)C(M), respectively, such that unu_{n} is a subsolution (resp. supersolution, solution) of (1.1) with G:=GnG:=G_{n}, for each nn\in\mathbb{N}. If unuu_{n}\to u in C(M)C(M) and GnGG_{n}\to G in C(TM×)C(T^{*}M\times\mathbb{R}) as n+n\to+\infty, then uu is a subsolution (resp. supersolution, solution) of (1.1).

1.3. Weak KAM solutions and Aubry-Mather theory.

We assume H:TMH:T^{*}M\to\mathbb{R} is a continuous Hamiltonian satisfying

  • (H1)

    (Convexity) H(x,p)H(x,p) is convex in pp for all xMx\in M.

  • (H2)

    (Superlinearity) limpx+H(x,p)/px=+\lim\limits_{\|p\|_{x}\to+\infty}H(x,p)/\|p\|_{x}=+\infty.

Let L:TML:TM\to\mathbb{R} be the convex conjugate function of HH, i.e.,

L(x,v):=suppTxM(p(v)H(x,p)),(x,v)TM.L(x,v):=\sup_{p\in T^{*}_{x}M}\big{(}p(v)-H(x,p)\big{)},\quad(x,v)\in TM.

It is well known that the Lagrangian LL is a continuous function on TMTM and it is convex and superlinear in vv. The Fenchel inequality is a direct consequence of this definition:

L(x,v)+H(x,p)p(v),for all (x,v,p)M×TxM×TxM.L(x,v)+H(x,p)\geqslant p(v),\quad\hbox{for all $(x,v,p)\in M\times T_{x}M\times T^{*}_{x}M$.} (1.2)

Moreover, it can be proven that HH is itself the convex conjugate of LL, i.e.,

H(x,p)=supvTxM(p(v)L(x,v))for all (x,p)TM.H(x,p)=\sup_{v\in T_{x}M}\big{(}p(v)-L(x,v)\big{)}\qquad\hbox{for all $(x,p)\in T^{*}M$}. (1.3)

Let c0c_{0}\in\mathbb{R} denote the critical constant defined as follows:

c0=min{c:H(x,Dxu)=cin Madmits subsolutions}.c_{0}=\min\{c\in\mathbb{R}:\ H(x,D_{x}u)=c\ \ \hbox{in $M$}\ \ \ \textrm{admits\ subsolutions}\}. (1.4)

We present here some facts that we will need about the critical equation, i.e.,

H(x,Dxu)=c0in M.H(x,D_{x}u)=c_{0}\qquad\hbox{in $M$}. (HJ0)

Solutions, subsolutions and supersolutions of (HJ0) will be termed critical in the sequel.

Due to the convex character of HH, the following holds, see for instance [3, 19].

Proposition 1.3.

Let uC(M)u\in C(M). The following properties hold:

  • (i)

    if uu is the pointwise supremum (respectively, infimum) of a family of subsolutions (resp., supersolutions) to (HJ0), then uu is a subsolution (resp., supersolution) of (HJ0);

  • (ii)

    if uu is the pointwise infimum of a family of equi-Lipschitz subsolutions to (HJ0), then uu is a Lipschitz subsolution of (HJ0);

  • (iii)

    if uu is a convex combination of a family of equi-Lipschitz subsolutions to (HJ0), then uu is a Lipschitz subsolution of (HJ0).

More precisely, items (ii) and (iii) above require the convexity of HH in the momentum, while item (i) is a general fact.

Since we are assuming HH to be superlinear (hence coercive, which is enough), we also have the following characterization of critical subsolutions, see for instance [3, 19].

Proposition 1.4.

The following are equivalent facts:

  • (i)

    vv is a viscosity subsolution of (HJ0);

  • (ii)

    vv is Lipschitz continuous and an almost everywhere subsolution of (HJ0), i.e.,

    H(x,Dxv)c0for a.e. xM.H(x,D_{x}v)\leqslant c_{0}\qquad\hbox{for a.e. $x\in M$.}

Moreover, the set of viscosity subsolutions of (HJ0) is equi-Lipschitz, with κc0:=sup{px:H(x,p)c0}\kappa_{c_{0}}:=\sup\{\|p\|_{x}\,:\,H(x,p)\leq c_{0}\} as a common Lipschitz constant.

For every t>0t>0, we define the minimal action function ht:M×Mh_{t}:M\times M\to\mathbb{R} as

ht(x,y)=infγt0[L(γ(s),γ˙(s))+c0]𝑑s,h_{t}(x,y)=\inf_{\gamma}\int_{-t}^{0}\big{[}L\big{(}\gamma(s),\dot{\gamma}(s)\big{)}+c_{0}\big{]}ds,

where γ:[t,0]M\gamma:[-t,0]\to M is taken among all absolutely continuous curves222In the paper, even if not explicitly stated, all curves considered are at least absolutely continuous. satisfying γ(t)=x\gamma(-t)=x and γ(0)=y\gamma(0)=y. The Peierls barrier is the function h:M×Mh:M\times M\to\mathbb{R} defined by

h(x,y):=lim inft+ht(x,y).h(x,y):=\liminf_{t\to+\infty}h_{t}(x,y). (1.5)

It satisfies the following properties, see for instance [16]:

Proposition 1.5.
  • (i)

    The Peierls barrier hh is finite valued and Lipschitz continuous.

  • (ii)

    If vv is a critical subsolution, then

    v(x)v(y)h(y,x),v(x)v(y)ht(y,x) for every x,yM and t>0.\qquad v(x)-v(y)\leqslant h(y,x),\quad v(x)-v(y)\leqslant h_{t}(y,x)\qquad\text{ for every $x,y\in M$ and $t>0$}.
  • (iii)

    For every fixed yMy\in M, the function h(y,)h(y,\cdot) is a critical solution.

  • (iv)

    For every fixed yMy\in M, the function h(,y)-h(\cdot,y) is a critical subsolution.

The projected Aubry set 𝒜\mathcal{A} is the closed set defined by

𝒜:={yM:h(y,y)=0}.\mathcal{A}:=\{y\in M\,:\,h(y,y)=0\,\}.

The following holds, see [19, 21]:

Theorem 1.6.

There exists a critical subsolution vv which is both strict and of class C in M𝒜M\setminus\mathcal{A}, meaning that

H(x,Dxv)<c0 for every xM𝒜.H(x,D_{x}v)<c_{0}\quad\text{ for every $x\in M\setminus\mathcal{A}$.}

In particular, the projected Aubry set 𝒜\mathcal{A} is nonempty.

The last assertion directly follows from the definition of c0c_{0}, see (1.4).

Proposition 1.7.

If uu and vv are respectively a sub and supersolution such that uvu\leqslant v on 𝒜\mathcal{A}, then uvu\leqslant v on the whole of MM. In particular, 𝒜\mathcal{A} is a uniqueness set for (HJ0), meaning that if two solutions coincide on 𝒜\mathcal{A}, then they are equal.

We will say that a Borel probability measure μ~\tilde{\mu} on TMTM is closed if it satisfies the following conditions:

  • (a)

    TMvx𝑑μ~(x,v)<+\displaystyle\int_{TM}\|v\|_{x}\,d\tilde{\mu}(x,v)<+\infty;

  • (b)

    for all function fC1(M)f\in C^{1}(M), we have TMDxf(v)𝑑μ~(x,v)=0\displaystyle\int_{TM}D_{x}f(v)\,d\tilde{\mu}(x,v)=0.

We will denote by 𝒫0\mathscr{P}_{0} the set of such measures.

We will furthermore denote by 𝒫\mathscr{P}_{\ell} the family of probability measures μ~\tilde{\mu} that satisfy condition (a) above. The inclusions 𝒫0𝒫(C(TM))\mathscr{P}_{0}\subset\mathscr{P}_{\ell}\subset\big{(}C_{\ell}(TM)\big{)}^{\prime} hold. We will endow 𝒫\mathscr{P}_{\ell} with the weak-* topology coming from the dual (C(TM))\big{(}C_{\ell}(TM)\big{)}^{\prime}. We refer the reader to [9] for more details on these families of measures.

Theorem 1.8.

The following holds

minμ~𝒫0TML(x,v)𝑑μ~=c0.\min_{\tilde{\mu}\in\mathscr{P}_{0}}\int_{TM}L(x,v)\,d\tilde{\mu}=-c_{0}.

Measures realizing the above minimum are called Mather measures for LL. We denote by 𝔐~\widetilde{\mathfrak{M}} the set of all Mather measures. This set is compact. The Mather set and the projected Mather set are defined as follows:

~:=μ~𝔐~supp(μ~)¯,:=π(~).\widetilde{\mathcal{M}}:=\overline{\bigcup_{\tilde{\mu}\in\widetilde{\mathfrak{M}}}\textrm{supp}(\tilde{\mu})},\qquad\mathcal{M}:=\pi\big{(}\widetilde{\mathcal{M}}\big{)}.

These sets are also compact, see [19] for a proof in the regular case.333For the present nonregular case, a proof of this can be found in Appendix A in the ArXiv version of [12]. Furthermore, the following holds, see [41, Proposition 3.13] for a proof in the nonregular case.

Theorem 1.9.

The following inclusion holds: 𝒜\mathcal{M}\subseteq\mathcal{A}.

Remark 1.10.

In the example of a mechanical Hamiltonian, i.e., H(x,p)=px2/2+V(x)H(x,p)=\|p\|_{x}^{2}/2+V(x), it is well known that c0=maxMVc_{0}=\max_{M}V, 𝒜={yM:V(y)=maxMV}\mathcal{A}=\{y\in M\,:\,V(y)=\max_{M}V\} and the Mather measures are convex combinations of delta Diracs concentrated at points (y,0)(y,0) with y𝒜y\in\mathcal{A}, so that \mathcal{M} is also equal to {yM:V(y)=maxMV}\{y\in M\,:\,V(y)=\max_{M}V\}.

We conclude this paragraph by a technical lemma that will be of crucial use (see [9, Theorem 2-4.1.3.]).

Lemma 1.11.

Let aa\in\mathbb{R}. The set {μ~𝒫:TML(x,v)𝑑μ~(x,v)a}\{\tilde{\mu}\in\mathscr{P}_{\ell}\,:\,\ \int_{TM}L(x,v)d\tilde{\mu}(x,v)\leqslant a\} is compact in 𝒫\mathscr{P}_{\ell}.

Note that, as LL is bounded below, the quantity TML(x,v)𝑑μ~(x,v){+}\int_{TM}L(x,v)d\tilde{\mu}(x,v)\in\mathbb{R}\cup\{+\infty\} is well defined for any Borel probability measure. We also remark that in [9, Theorem 2-4.1.3.] the result is proved for a particular subclass of measures, however the proof makes no use of this fact and proves the above result.

1.4. Hamiltonians depending on the unknown function

We recall here known results that can be found in [36, 37] and the references therein. In this section, we consider a continuous Hamiltonian G:TM×G:T^{*}M\times\mathbb{R}\to\mathbb{R} which satisfies the following conditions

  • (G1)

    (Lipschitz in uu) uG(x,p,u)u\mapsto G(x,p,u) is KK-Lipschitz continuous for some K>0K>0, uniformly in (x,p)TM(x,p)\in T^{*}M;

  • (G2)

    (Convexity in pp) pG(x,p,u)p\mapsto G(x,p,u) is convex for each (x,u)M×(x,u)\in M\times\mathbb{R};

  • (G3)

    (Superlinearity in pp) pG(x,p,u)p\mapsto G(x,p,u) is superlinear for each (x,u)M×(x,u)\in M\times\mathbb{R}.

Remark 1.12.

The results of this paper keep holding even when the superlinearity condition (G3) is weakened in favor of a simple coercivity. For instance, Theorems 2.2 and 2.3 can be easily generalized to this setting. Indeed, since we are dealing there with a family of equi-bounded, and hence equi-Lipschitz, solutions, see Lemma 2.6, we could employ the usual trick of modifying GG outside a compact subset of TM×T^{*}M\times\mathbb{R} to make it superlinear. This cannot be done in other parts of the paper since we are dealing with families of solutions that are neither equi-bounded nor equi-Lipschitz in general. And even when they are, as in Section 3.1, this needs to be proved. In fact, this is the core of the analysis performed in Section 3.1, which takes advantage of the fact that the Lagrangian associated with the Hamiltonian via the Fenchel duality is finite-valued. This is no longer true in the purely coercive case, even though the difficulties arising could be handled by showing that all the minimizing curves that come into play in our analysis are indeed supported on the set where the Lagrangian is finite. Yet, we believe that treating this more general case would bring additional technicalities that would have the effect of hiding the ideas at the base of this work. We prefer to leave the coercive case to a possible future work.

Let LG:TM×L_{G}:TM\times\mathbb{R}\to\mathbb{R} be the convex conjugate function of GG, i.e.,

LG(x,v,u):=suppTxM(p(v)G(x,p,u)).L_{G}(x,v,u):=\sup_{p\in T^{*}_{x}M}\big{(}p(v)-G(x,p,u)\big{)}.

Then it can be proven that LGL_{G} verifies similar properties (see [5, Lemma 4.1] with easy adaptations):

  • (L1)

    uLG(x,v,u)u\mapsto L_{G}(x,v,u) is KK-Lipschitz continuous uniformly in (x,v)TM(x,v)\in TM;

  • (L2)

    vLG(x,v,u)v\mapsto L_{G}(x,v,u) is convex for each (x,u)M×(x,u)\in M\times\mathbb{R};

  • (L3)

    vLG(x,v,u)v\mapsto L_{G}(x,v,u) is superlinear for each (x,u)M×(x,u)\in M\times\mathbb{R};

Definition 1.13.

Let G:TM×G:T^{*}M\times\mathbb{R}\to\mathbb{R} be a Hamiltonian satisfying (G1-3) and let LG:TM×L_{G}:TM\times\mathbb{R}\to\mathbb{R} be the associated Lagrangian. Let cc\in\mathbb{R}. A function uC(M)u\in C(M) satisfying the following two properties is called a backward (resp. forward) weak KAM solution of

G(x,Dxu,u(x))=cin M.G\big{(}x,D_{x}u,u(x)\big{)}=c\quad\hbox{in $M$}. (1.6)
  • (1)

    For each absolutely continuous curve γ:[t,t]M\gamma:[t^{\prime},t]\rightarrow M, we have

    u(γ(t))u(γ(t))tt[LG(γ(s),γ˙(s),u(γ(s)))+c]𝑑s.u\big{(}\gamma(t)\big{)}-u\big{(}\gamma(t^{\prime})\big{)}\leqslant\int_{t^{\prime}}^{t}\Big{[}L_{G}\Big{(}\gamma(s),\dot{\gamma}(s),u\big{(}\gamma(s)\big{)}\Big{)}+c\Big{]}ds.

    The above condition reads as uu is dominated by LG+cL_{G}+c and will be denoted by uLG+cu\prec L_{G}+c.

  • (2)

    For each xMx\in M, there exists an absolutely continuous curve γ:(,0]M\gamma_{-}:(-\infty,0]\rightarrow M (resp. γ+:[0,+)M\gamma_{+}:[0,+\infty)\to M) with γ(0)=x\gamma_{-}(0)=x (resp. γ+(0)=x\gamma_{+}(0)=x) such that

    u(x)u(γ(t))=t0[LG(γ(s),γ˙(s),u(γ(s)))+c]𝑑s,t<0\displaystyle u(x)-u\big{(}\gamma_{-}(t)\big{)}=\int_{t}^{0}\Big{[}L_{G}\Big{(}\gamma_{-}(s),\dot{\gamma}_{-}(s),u\big{(}\gamma_{-}(s)\big{)}\Big{)}+c\Big{]}ds,\quad\forall t<0
    (resp.u(γ+(t))u(x)=0t[LG(γ+(s),γ˙+(s),u(γ+(s)))+c]𝑑s,t>0).\displaystyle\textrm{\Big{(}resp.}\ u(\gamma_{+}(t))-u(x)=\int_{0}^{t}\Big{[}L_{G}\Big{(}\gamma_{+}(s),\dot{\gamma}_{+}(s),u\big{(}\gamma_{+}(s)\big{)}\Big{)}+c\Big{]}ds,\quad\forall t>0\textrm{\Big{)}.}

    The curves satisfying the above equality are called (u,LG,c)(u,L_{G},c)-calibrated curves.

By [37, Appendix D] and [36, Appendix A] we have

Lemma 1.14.
  • (i)

    If uC(M)u\in C(M) is a backward weak KAM solution of (1.6), it is a viscosity solution of (1.6).

  • (ii)

    The function wC(M)w\in C(M) is a viscosity subsolution of (1.6) if and only if wLG+cw\prec L_{G}+c. The latter is also equivalent to ww being Lipschitz continuous and verifying G(x,Dxw,w(x))cG\big{(}x,D_{x}w,w(x)\big{)}\leqslant c for almost every xMx\in M.

We then give an approximation result in this setting. It is an easy consequence of [20, Theorem 8.5].

Theorem 1.15.

Assume G:TM×G:T^{*}M\times\mathbb{R}\to\mathbb{R} is a continuous Hamiltonian verifying (G2). Let w:Mw:M\to\mathbb{R} be a Lipschitz function verifying G(x,Dxw,w(x))cG\big{(}x,D_{x}w,w(x)\big{)}\leqslant c for almost every xMx\in M. Then, for every ε>0\varepsilon>0, there is a wεC(M)w_{\varepsilon}\in C^{\infty}(M) such that wwε<ε\|w-w_{\varepsilon}\|_{\infty}<\varepsilon and

G(x,Dxwε,w(x))c+ε,G(x,Dxwε,wε(x))c+εfor all xM.G\big{(}x,D_{x}w_{\varepsilon},{w}(x)\big{)}\leqslant c+\varepsilon,\quad G\big{(}x,D_{x}w_{\varepsilon},w_{\varepsilon}(x)\big{)}\leqslant c+\varepsilon\qquad\hbox{for all $x\in M$}.

Let aC(M)a\in C(M). Assume there are two points x1x_{1} and x2x_{2} such that a(x1)>0a(x_{1})>0 and a(x2)<0a(x_{2})<0. Let cc\in\mathbb{R} and consider the equation

a(x)u(x)+H(x,Dxu)=cin M.a(x)u(x)+H(x,D_{x}u)=c\qquad\hbox{in $M$}. (1.7)

This is a particular case of the previous setting for G(x,p,u)=a(x)u+H(x,p)G(x,p,u)=a(x)u+H(x,p). In this case, the associated Lagrangian is LG(x,v,u)=L(x,v)a(x)uL_{G}(x,v,u)=L(x,v)-a(x)u. The following implicit Lax-Oleinik semigroup (Tt)t0(T^{-}_{t})_{t\geqslant 0} is a well defined semigroup of operators Tt:C(M)C(M)T^{-}_{t}:C(M)\to C(M) that verify, for all φC(M)\varphi\in C(M),

Ttφ(x)=infγ(t)=x{φ(γ(0))+0t[L(γ(τ),γ˙(τ))a(γ(τ))Tτφ(γ(τ))+c]𝑑τ},\displaystyle T^{-}_{t}\varphi(x)=\inf_{\gamma(t)=x}\left\{\varphi\big{(}\gamma(0)\big{)}+\int_{0}^{t}\bigg{[}L\big{(}\gamma(\tau),\dot{\gamma}(\tau)\big{)}-a\big{(}\gamma(\tau)\big{)}T^{-}_{\tau}\varphi\big{(}\gamma(\tau)\big{)}+c\bigg{]}{d}\tau\right\}, (1.8)

where the infimum is taken among absolutely continuous curves γ:[0,t]M\gamma:[0,t]\rightarrow M with γ(t)=x\gamma(t)=x. A similar property defines and characterizes the forward semigroup (Tt+)t>0(T^{+}_{t})_{t>0}, where, for all φC(M)\varphi\in C(M),

Tt+φ(x)=supγ(0)=x{φ(γ(t))0t[L(γ(τ),γ˙(τ))a(γ(τ))Ttτ+φ(γ(τ))+c]𝑑τ}.\displaystyle T^{+}_{t}\varphi(x)=\sup_{\gamma(0)=x}\left\{\varphi\big{(}\gamma(t)\big{)}-\int_{0}^{t}\bigg{[}L\big{(}\gamma(\tau),\dot{\gamma}(\tau)\big{)}-a\big{(}\gamma(\tau)\big{)}T^{+}_{t-\tau}\varphi\big{(}\gamma(\tau)\big{)}+c\bigg{]}{d}\tau\right\}. (1.9)
Lemma 1.16.

[36, Appendix A] If u0u_{0} is a subsolution of (1.7), then

Tt+u0u0Ttu0.T^{+}_{t}u_{0}\leqslant u_{0}\leqslant T^{-}_{t}u_{0}.

If u0u_{0} is a strict subsolution444Meaning that cc can be replaced with cεc-\varepsilon for some ε>0\varepsilon>0. of (1.7), then

Tt+u0<u0<Ttu0.T^{+}_{t}u_{0}<u_{0}<T^{-}_{t}u_{0}.
Lemma 1.17.

[36, Proposition 2.9, Proposition 3.5, Lemma 5.4] If u0u_{0} is a subsolution of (1.7), then the limit

u:=limt+Ttu0(resp.v+:=limt+Tt+u0)u_{-}:=\lim_{t\to+\infty}T^{-}_{t}u_{0}\quad(resp.\ v_{+}:=\lim_{t\to+\infty}T^{+}_{t}u_{0})

exists, and is a viscosity solution (resp. forward weak KAM solution) of (1.7). In addition,

v:=limt+Ttv+v_{-}:=\lim_{t\to+\infty}T^{-}_{t}v_{+}

is also a viscosity solution of (1.7), and there is a point x0Mx_{0}\in M such that v(x0)=v+(x0)v_{-}(x_{0})=v_{+}(x_{0}).

If u0u_{0} is a strict subsolution of (1.7), then uu_{-} is the maximal solution of (1.7), and vv_{-} is the minimal solution of (1.7).

2. General convergence/divergence results

In this section, we consider a continuous Hamiltonian G:TM×G:T^{*}M\times\mathbb{R}\to\mathbb{R} which satisfies the following conditions:

  • (G1)

    (Lipschitz in uu) uG(x,p,u)u\mapsto G(x,p,u) is KK-Lipschitz continuous uniformly in (x,p)TM(x,p)\in T^{*}M for some K>0K>0;

  • (G2)

    (Convexity in pp) vG(x,p,u)v\mapsto G(x,p,u) is convex for each (x,u)M×(x,u)\in M\times\mathbb{R};

  • (G3)

    (Superlinearity in pp) pG(x,p,u)p\mapsto G(x,p,u) is superlinear for each (x,u)M×(x,u)\in M\times\mathbb{R};

  • (G4)

    (Modulus continuity near u=0u=0) The partial derivative Gu(x,p,0)\frac{\partial G}{\partial u}(x,p,0) exists. For every compact subset STMS\subset TM, we can find a modulus of continuity555A modulus of continuity is a nondecreasing function η:(0,+)(0,+)\eta:(0,+\infty)\to(0,+\infty) such that η(s)0\eta(s)\to 0 as s0s\to 0. ηS\eta_{S} such that

    |G(x,p,u)G(x,p,0)Gu(x,p,0)u||u|ηS(|u|),(x,p)S.\bigg{|}G(x,p,u)-G(x,p,0)-\frac{\partial G}{\partial u}(x,p,0)u\bigg{|}\leqslant|u|\eta_{S}(|u|),\quad\forall(x,p)\in S.

The dependence of GG in the uu-variable is nonlinear, in general, as in [7], but we do not make any global monotonicity assumption. As established in [5, Lemma 4.1] the associated Lagrangian function LG:TM×L_{G}:TM\times\mathbb{R}\to\mathbb{R} defined by

LG(x,v,u):=suppTxM(p(v)G(x,p,u)),L_{G}(x,v,u):=\sup_{p\in T^{*}_{x}M}\big{(}p(v)-G(x,p,u)\big{)},

has similar properties:

  • (L1)

    (Lipschitz in uu) uLG(x,v,u)u\mapsto L_{G}(x,v,u) is KK-Lipschitz continuous uniformly in (x,v)TM(x,v)\in TM;

  • (L2)

    (Convexity in vv) vLG(x,v,u)v\mapsto L_{G}(x,v,u) is convex for each (x,u)M×(x,u)\in M\times\mathbb{R};

  • (L3)

    (Superlinearity in vv) vLG(x,v,u)v\mapsto L_{G}(x,v,u) is superlinear for each (x,u)M×(x,u)\in M\times\mathbb{R};

  • (L4)

    (Modulus continuity near u=0u=0) The partial derivative LGu(x,v,0)\frac{\partial L_{G}}{\partial u}(x,v,0) exists. For every compact subset STMS\subset TM, we can find a modulus of continuity ηS\eta_{S} such that

    |LG(x,v,u)LG(x,v,0)LGu(x,v,0)u||u|ηS(|u|),(x,v)S.\bigg{|}L_{G}(x,v,u)-L_{G}(x,v,0)-\frac{\partial L_{G}}{\partial u}(x,v,0)u\bigg{|}\leqslant|u|\eta_{S}(|u|),\quad\forall(x,v)\in S.

The last condition is easier to state with the Lagrangian function as it involves Mather measures defined on TMTM:

  • (L5)

    (Non-degeneracy condition) For all Mather measures μ~\tilde{\mu} of (x,v)LG(x,v,0)(x,v)\mapsto L_{G}(x,v,0), we have

    TMLGu(x,v,0)𝑑μ~<0.\int_{TM}\frac{\partial L_{G}}{\partial u}(x,v,0)\,d\tilde{\mu}<0.
Remark 2.1.

(1) It is classical in convex analysis that for all (x,v)TM(x,v)\in TM there is (x,p)TM(x,p)\in T^{*}M verifying LG(x,v,0)+G(x,p,0)=p(v)L_{G}(x,v,0)+G(x,p,0)=p(v). It is proved in [5, Lemma 4.1] that if (x,p)TM(x,p)\in T^{*}M and (x,v)TM(x,v)\in TM verify the previous formula then LGu(x,v,0)=Gu(x,p,0)\frac{\partial L_{G}}{\partial u}(x,v,0)=-\frac{\partial G}{\partial u}(x,p,0). Therefore, an equivalent formulation of (L5) is

  • (G5)

    (Non-degeneracy condition) For all Mather measures μ~\tilde{\mu} of (x,v)LG(x,v,0)(x,v)\mapsto L_{G}(x,v,0),

    TMGu(x,p(x,v),0)𝑑μ~>0.\int_{TM}\frac{\partial G}{\partial u}(x,p_{(x,v)},0)\,d\tilde{\mu}>0.

    where p(x,v)p_{(x,v)} is chosen so to satisfy LG(x,v,0)+G(x,p(x,v),0)=p(x,v)(v)L_{G}(x,v,0)+G(x,p_{(x,v)},0)=p_{(x,v)}(v) .

(2) A result of Mañé ([33]) asserts that a generic Hamiltonian HH has a unique Mather measure.666It has been conjectured by Mañé that for generic Hamiltonians HH this unique Mather measure is supported on a closed curve. Hence our condition of integral type is quite loose for such a generic HH. We also refer the reader to Remark 1.10 for an explicit example.

Consider

G(x,Dxu,λu(x))=c0in M,G\big{(}x,D_{x}u,\lambda u(x)\big{)}=c_{0}\qquad\hbox{in $M$}, (Eλ)

and the limit equation

G(x,Dxu,0)=c0in M.G(x,D_{x}u,0)=c_{0}\qquad\hbox{in $M$}. (E0{\textrm{E}}_{0})

Here we denote by c0c_{0} the critical value of H(x,p):=G(x,p,0)H(x,p):=G(x,p,0). We denote by 𝔐~\widetilde{\mathfrak{M}} (resp. ~\widetilde{\mathcal{M}}) the set of all Mather measures (resp. the Mather set) corresponding to HH. In the sequel, we will always assume that λ\lambda belongs to the interval (0,1)(0,1).

The main theorems of this section are the following.

Theorem 2.2.

Let conditions (L1–5) be in force. Let us assume that there exist an equi-bounded family (uλ)λ(0,λ0)(u_{\lambda})_{\lambda\in(0,\lambda_{0})} of solutions to (Eλ), for some λ0(0,1)\lambda_{0}\in(0,1). Then the functions uλu_{\lambda} uniformly converge in MM, as λ0+\lambda\to 0^{+}, to a solution u0u_{0} of the critical equation (E0{\textrm{E}}_{0}). Moreover, u0u_{0} is the largest subsolution ww of (E0{\textrm{E}}_{0}) satisfying

TMw(x)LGu(x,v,0)𝑑μ~(x,v)0,μ~𝔐~.\int_{TM}w(x)\frac{\partial L_{G}}{\partial u}(x,v,0)\,d\tilde{\mu}(x,v)\geqslant 0,\quad\forall\tilde{\mu}\in\widetilde{\mathfrak{M}}. (S)

Under an additional pointwise condition on LG(,,0)L_{G}(\cdot,\cdot,0) on the Mather set, the limit critical solution u0u_{0} can be characterized as follows.

Theorem 2.3.

Let conditions (L1–5) be in force, and let us furthermore assume that LGu(x,v,0)0\frac{\partial L_{G}}{\partial u}(x,v,0)\leqslant 0 for all (x,v)~(x,v)\in\widetilde{\mathcal{M}}. Then the function u0=limλ0+uλu_{0}=\lim\limits_{\lambda\to 0^{+}}u_{\lambda} obtained in Theorem 2.2 above can be characterized as follows:

u0(x)=infμ~𝔐~TMh(y,x)LGu(y,v,0)𝑑μ~(y,v)TMLGu(y,v,0)𝑑μ~(y,v),xM,u_{0}(x)=\inf_{\tilde{\mu}\in\widetilde{\mathfrak{M}}}\frac{\int_{TM}h(y,x)\frac{\partial L_{G}}{\partial u}(y,v,0)d\tilde{\mu}(y,v)}{\int_{TM}\frac{\partial L_{G}}{\partial u}(y,v,0)d\tilde{\mu}(y,v)},\qquad{x\in M},

where h(y,x)h(y,x) is the Peierls barrier defined in (1.5).

Remark 2.4.

Let us stress that the previous theorems hold with same proofs even if (uλ)λ(0,λ0)(u_{\lambda})_{\lambda\in(0,\lambda_{0})} is replaced by any family (uλ)λΛ(u_{\lambda})_{\lambda\in\Lambda} of solutions to (Eλ), where Λ\Lambda is a subset of (0,1)(0,1) having 0 as accumulation point. In particular, this holds for Λ:=(λn)n\Lambda:=(\lambda_{n})_{n} with λn0+\lambda_{n}\to 0^{+} as n+n\to+\infty.

Concerning the asymptotic behavior of other possible families of solutions to (Eλ), we have the following trichotomy result.

Theorem 2.5.

Let conditions (L1–5) be in force, and let u0=limλ0+uλu_{0}=\lim\limits_{\lambda\to 0^{+}}u_{\lambda} be the critical solution obtained in Theorem 2.2. There exist φ:(0,1)(,+]\varphi:(0,1)\to\mathbb{(}-\infty,+\infty], ψ:(0,1)[,+)\psi:(0,1)\to\mathbb{[}-\infty,+\infty) and θ:(0,1)\theta:(0,1)\to\mathbb{R} with

limλ0ψ(λ)=,limλ0φ(λ)=+,limλ0θ(λ)=0\lim_{\lambda\to 0}\psi(\lambda)=-\infty,\quad\quad\lim_{\lambda\to 0}\varphi(\lambda)=+\infty,\quad\quad\lim_{\lambda\to 0}\theta(\lambda)=0

such that, if vλv_{\lambda} is a solution of (Eλ) for some λ>0\lambda>0, then either one of the following alternatives occurs:

  • (i)

    vλψ(λ)v_{\lambda}\leqslant\psi(\lambda);

  • (ii)

    vλφ(λ)v_{\lambda}\geqslant\varphi(\lambda);

  • (iii)

    vλu0θ(λ)\|v_{\lambda}-u_{0}\|_{\infty}\leqslant\theta(\lambda).

We start our analysis by remarking that the solutions (uλ)λ(0,λ0)(u_{\lambda})_{\lambda\in(0,\lambda_{0})} are equi-Lipschitz continuous. In the remainder of the section, we will denote by C>0C>0 the following constant

C:=supλ(0,λ0)uλ.C:=\sup_{\lambda\in(0,\lambda_{0})}\|u_{\lambda}\|_{\infty}.
Lemma 2.6.

The bounded family (uλ)λ(0,λ0)(u_{\lambda})_{\lambda\in(0,\lambda_{0})} is equi-Lipschitz continuous.

Proof.

For each x,yMx,y\in M, we denote by d:=d(x,y)d:=d(x,y) the distance between them. Take a geodesic ζ:[0,d]M\zeta:[0,d]\rightarrow M satisfying ζ(0)=x\zeta(0)=x and ζ(d)=y\zeta(d)=y with constant speed ζ˙ζ=1\|\dot{\zeta}\|_{\zeta}=1. Denote Lλ(x,v):=LG(x,v,λuλ)L_{\lambda}(x,v):=L_{G}(x,v,\lambda u_{\lambda}). Since uλLλ+c0u_{\lambda}\prec L_{\lambda}+c_{0}, we have

uλ(y)uλ(x)\displaystyle u_{\lambda}(y)-u_{\lambda}(x) 0d[LG(ζ(s),ζ˙(s),λuλ(ζ(s)))+c0]𝑑s\displaystyle\leqslant\int_{0}^{d}\bigg{[}L_{G}\Big{(}\zeta(s),\dot{\zeta}(s),\lambda u_{\lambda}\big{(}\zeta(s)\big{)}\Big{)}+c_{0}\bigg{]}ds
(maxxM,vx1|LG(x,v,0)|+λ0KC+c0)d(x,y).\displaystyle\leqslant\bigg{(}\max_{x\in M,\|v\|_{x}\leqslant 1}|L_{G}(x,v,0)|+\lambda_{0}KC+c_{0}\bigg{)}d(x,y).

The assertion follows by exchanging the role of xx and yy. ∎

From the fact that uλu_{\lambda} is Lipschitz for every fixed λ(0,λ0)\lambda\in(0,\lambda_{0}) we deduce the following fact.

Lemma 2.7.

Let λ(0,λ0)\lambda\in(0,\lambda_{0}). Then

TMLG(x,v,λuλ(x))𝑑μ~(x,v)c0for all μ~𝒫0.\int_{TM}L_{G}\big{(}x,v,\lambda u_{\lambda}(x)\big{)}d\tilde{\mu}(x,v)\geqslant-c_{0}\qquad\hbox{for all $\tilde{\mu}\in\mathscr{P}_{0}$.}
Proof.

Pick μ~𝒫0\tilde{\mu}\in\mathscr{P}_{0} and choose ε>0\varepsilon>0. By applying Theorem 1.15 to the Hamiltonian (x,p)G(x,p,λuλ(x))(x,p)\mapsto G\big{(}x,p,\lambda u_{\lambda}(x)\big{)} and by choosing w:=uλw:=u_{\lambda}, we infer that there exists wεC(M)w_{\varepsilon}\in C^{\infty}(M) such that G(x,Dxwε,λuλ(x))c0+εG\big{(}x,D_{x}w_{\varepsilon},\lambda u_{\lambda}(x)\big{)}\leqslant c_{0}+\varepsilon for all xMx\in M. By definition of LGL_{G} we have that

G(x,Dxwε,λuλ(x))+LG(x,v,λuλ(x))Dxwε(v)for all (x,v)TM.G\big{(}x,D_{x}w_{\varepsilon},\lambda u_{\lambda}(x)\big{)}+L_{G}\big{(}x,v,\lambda u_{\lambda}(x)\big{)}\geqslant D_{x}w_{\varepsilon}(v)\qquad\hbox{for all $(x,v)\in TM$}.

By integrating this inequality with respect to μ~\tilde{\mu} we get

0=TMDxwε(v)𝑑μ~(x,v)TM(LG(x,v,λuλ(x))+G(x,Dxwε,λuλ(x)))𝑑μ~(x,v)TM(LG(x,v,λuλ(x))+c0+ε)𝑑μ~(x,v)=TMLG(x,v,λuλ(x))𝑑μ~(x,v)+c0+ε.0=\int_{TM}D_{x}w_{\varepsilon}(v)\,d\tilde{\mu}(x,v)\\ \leqslant\int_{TM}\Big{(}L_{G}\big{(}x,v,\lambda u_{\lambda}(x)\big{)}+G\big{(}x,D_{x}w_{\varepsilon},\lambda u_{\lambda}(x)\big{)}\Big{)}d\tilde{\mu}(x,v)\\ \leqslant\int_{TM}\Big{(}L_{G}\big{(}x,v,\lambda u_{\lambda}(x)\big{)}+c_{0}+\varepsilon\Big{)}d\tilde{\mu}(x,v)=\int_{TM}L_{G}\big{(}x,v,\lambda u_{\lambda}(x)\big{)}d\tilde{\mu}(x,v)+c_{0}+\varepsilon.

The result follows letting ε0+\varepsilon\to 0^{+}. ∎

By the Arzelá-Ascoli Theorem and Lemma 2.6, any sequence (uλn)n(u_{\lambda_{n}})_{n} with λn0+\lambda_{n}\to 0^{+} admits a subsequence which uniformly converges to a continuous function uu^{*}. By the stability of viscosity solution, see Proposition 1.2, uu^{*} is a solution of (E0{\textrm{E}}_{0}). In the following, we are going to show the uniqueness of the possible limit uu^{*}, thus establishing the convergence result. Define 𝒮\mathcal{S} the set of all subsolutions ww of (E0{\textrm{E}}_{0}) satisfying condition (S). We define

u0(x):=supw𝒮w(x).u_{0}(x):=\sup_{w\in\mathcal{S}}w(x). (2.1)

A priori, 𝒮\mathcal{S} may be empty, and, even if 𝒮\mathcal{S} is not empty, u0u_{0} might be ++\infty. Both these circumstances will be excluded under the hypotheses of Theorem 2.2.

Lemma 2.8.

Any accumulation point uu^{*} of the family (uλ)λ(0,λ0)(u_{\lambda})_{\lambda\in(0,\lambda_{0})} as λ0+\lambda\to 0^{+} satisfies

TMLGu(x,v,0)u(x)𝑑μ(x,v)0,μ~𝔐~.\int_{TM}\frac{\partial L_{G}}{\partial u}(x,v,0)u^{*}(x)d\mu(x,v)\geqslant 0,\quad\forall\tilde{\mu}\in\widetilde{\mathfrak{M}}.

In particular, 𝒮\mathcal{S}\not=\varnothing and uu0u^{*}\leqslant u_{0}.

Proof.

Recall that CC is the uniform bound of (uλ)λ(0,λ0)(u_{\lambda})_{\lambda\in(0,\lambda_{0})}. Let μ~𝔐~\tilde{\mu}\in\widetilde{\mathfrak{M}}. For λ(0,λ0)\lambda\in(0,\lambda_{0}) we have

c0\displaystyle-c_{0} TMLG(x,v,λuλ(x))𝑑μ~\displaystyle\leqslant\int_{TM}L_{G}\big{(}x,v,\lambda u_{\lambda}(x)\big{)}d\tilde{\mu}
TM[LG(x,v,0)+λLGu(x,v,0)uλ(x)+λCη~(λC)]𝑑μ~\displaystyle\leqslant\int_{TM}\bigg{[}L_{G}(x,v,0)+\lambda\frac{\partial L_{G}}{\partial u}(x,v,0)u_{\lambda}(x)+\lambda C\eta_{\widetilde{\mathcal{M}}}(\lambda C)\bigg{]}d\tilde{\mu}
=c0+TM[λLGu(x,v,0)uλ(x)+λCη~(λC)]𝑑μ~,\displaystyle=-c_{0}+\int_{TM}\bigg{[}\lambda\frac{\partial L_{G}}{\partial u}(x,v,0)u_{\lambda}(x)+\lambda C\eta_{\widetilde{\mathcal{M}}}(\lambda C)\bigg{]}d\tilde{\mu},

which implies

TMLGu(x,v,0)uλ(x)𝑑μ~Cη~(λC).\int_{TM}\frac{\partial L_{G}}{\partial u}(x,v,0)u_{\lambda}(x)d\tilde{\mu}\geqslant-C\eta_{\widetilde{\mathcal{M}}}(\lambda C).

The conclusion follows by sending λ0+\lambda\to 0^{+}. ∎

In what follows, we will use the notation Lλ(x,v):=LG(x,v,λuλ)L_{\lambda}(x,v):=L_{G}(x,v,\lambda u_{\lambda}).

Lemma 2.9.

For xMx\in M and λ(0,λ0)\lambda\in(0,\lambda_{0}), let γλx:(,0]M\gamma^{x}_{\lambda}:(-\infty,0]\to M be a (uλ,Lλ,c0)(u_{\lambda},L_{\lambda},c_{0})-calibrated curve with γλx(0)=x\gamma^{x}_{\lambda}(0)=x. Then there exists κ^>0\hat{\kappa}>0, independent of λ(0,λ0)\lambda\in(0,\lambda_{0}) and of xMx\in M, such that γλx\gamma^{x}_{\lambda} is κ^\hat{\kappa}-Lipschitz continuous for every λ(0,λ0)\lambda\in(0,\lambda_{0}) and xMx\in M.

Proof.

By Lemma 2.6, there is κ>0\kappa>0 independent of λ\lambda such that uλu_{\lambda} is κ\kappa-Lipschitz continuous. By superlinearity of LGL_{G}, for each T>0T>0, there is CTC_{T}\in\mathbb{R} such that

LG(x,v,0)Tvx+CT.L_{G}(x,v,0)\geqslant T\|v\|_{x}+C_{T}. (2.2)

Thus, we have for 0t>s0\geqslant t>s

κd(γλx(t),γλx(s))\displaystyle\ \kappa d\big{(}\gamma^{x}_{\lambda}(t),\gamma^{x}_{\lambda}(s)\big{)} uλ(γλx(t))uλ(γλx(s))=st[LG(γλx(τ),γ˙λx(τ),λuλ(γλx(τ)))+c0]𝑑τ\displaystyle\geqslant u_{\lambda}\big{(}\gamma^{x}_{\lambda}(t)\big{)}-u_{\lambda}\big{(}\gamma^{x}_{\lambda}(s)\big{)}=\int_{s}^{t}\bigg{[}L_{G}\Big{(}\gamma^{x}_{\lambda}(\tau),\dot{\gamma}^{x}_{\lambda}(\tau),\lambda u_{\lambda}\big{(}\gamma^{x}_{\lambda}(\tau)\big{)}\Big{)}+c_{0}\bigg{]}d\tau
st((κ+1)γ˙λx(τ)γλx(τ)+Cκ+1)𝑑τ+(c0λ0KC)(ts)\displaystyle\geqslant\int_{s}^{t}\big{(}(\kappa+1)\|\dot{\gamma}^{x}_{\lambda}(\tau)\|_{\gamma^{x}_{\lambda}(\tau)}+C_{\kappa+1}\big{)}d\tau+(c_{0}-\lambda_{0}KC)(t-s)
(κ+1)d(γλx(t),γλx(s))+(Cκ+1+c0λ0KC)(ts),\displaystyle\geqslant(\kappa+1)d\big{(}\gamma^{x}_{\lambda}(t),\gamma^{x}_{\lambda}(s)\big{)}+(C_{\kappa+1}+c_{0}-\lambda_{0}KC)(t-s),

which implies

d(γλx(t),γλx(s))(λ0KCc0Cκ+1)(ts).d\big{(}\gamma^{x}_{\lambda}(t),\gamma^{x}_{\lambda}(s)\big{)}\leqslant(\lambda_{0}KC-c_{0}-C_{\kappa+1})(t-s).

The proof is now complete. ∎

From now on, we denote by γλx\gamma^{x}_{\lambda} the calibrated curve considered in Lemma 2.9. By the compactness of 𝔐~\widetilde{\mathfrak{M}}, there are two constants ε1>0\varepsilon_{1}>0 and ε2>0\varepsilon_{2}>0 with

ε2<infμ~𝔐~TMLGu(x,v,0)𝑑μ~supμ~𝔐~TMLGu(x,v,0)𝑑μ~<ε1,-\varepsilon_{2}<\inf_{\tilde{\mu}\in\widetilde{\mathfrak{M}}}\int_{TM}\frac{\partial L_{G}}{\partial u}(x,v,0)d\tilde{\mu}\leqslant\sup_{\tilde{\mu}\in\widetilde{\mathfrak{M}}}\int_{TM}\frac{\partial L_{G}}{\partial u}(x,v,0)d\tilde{\mu}<-\varepsilon_{1}, (2.3)

We derive from this the following asymptotic informations on the calibrated curves γλx\gamma^{x}_{\lambda}, cf. [7, Corollary 7.4].

Lemma 2.10.

There exist λ¯(0,λ0)\bar{\lambda}\in(0,\lambda_{0}) and T0>0T_{0}>0 such that, for all λ(0,λ¯)\lambda\in(0,\bar{\lambda}) and for all xMx\in M, we have

ε2<1baabLGu(γλx(s),γ˙λx(s),0)𝑑s<ε1,-\varepsilon_{2}<\frac{1}{b-a}\int_{a}^{b}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds<-\varepsilon_{1}, (2.4)

for all a<b0a<b\leqslant 0 with baT0b-a\geqslant T_{0}.

Proof.

Let us prove the left-hand inequality in (2.4). We argue by contradiction. Assume there is a sequence λn0\lambda_{n}\to 0 and a sequence bnan+b_{n}-a_{n}\to+\infty such that

1bnananbnLGu(γλnx(s),γ˙λnx(s),0)𝑑sε2.\frac{1}{b_{n}-a_{n}}\int_{a_{n}}^{b_{n}}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda_{n}}(s),\dot{\gamma}^{x}_{\lambda_{n}}(s),0)ds\leqslant-\varepsilon_{2}. (2.5)

Define for all nn\in\mathbb{N} a probability measure μ~n\tilde{\mu}_{n} on TMTM by

TMg(x,v)𝑑μ~n:=1bnananbng(γλnx(s),γ˙λnx(s))𝑑s,gCc(TM).\int_{TM}g(x,v)d\tilde{\mu}_{n}:=\frac{1}{b_{n}-a_{n}}\int_{a_{n}}^{b_{n}}g\big{(}\gamma^{x}_{\lambda_{n}}(s),\dot{\gamma}^{x}_{\lambda_{n}}(s)\big{)}ds,\quad\forall g\in C_{c}(TM).

Here Cc(TM)C_{c}(TM) is the set of all continuous functions defined on TMTM with compact supports. By Lemma 2.9, all the measures μ~n\tilde{\mu}_{n} have support in the compact set

{(x,v)TM:vxλ0KCc0Cκ+1}.\{(x,v)\in TM:\ \|v\|_{x}\leqslant\lambda_{0}KC-c_{0}-C_{\kappa+1}\}.

Up to extracting a subsequence if necessary, we can assume that μ~n\tilde{\mu}_{n} converges to μ~\tilde{\mu} in the weak-* topology on Cc(TM)C_{c}(TM). Note that μ~\tilde{\mu} is also compactly supported. For fC1(M)f\in C^{1}(M), we have

TMDxf(v)𝑑μ~n(x,v)\displaystyle\int_{TM}D_{x}f(v)d\tilde{\mu}_{n}(x,v) =1bnananbnDγλnx(s)f(γ˙λnx(s))𝑑s\displaystyle=\frac{1}{b_{n}-a_{n}}\int_{a_{n}}^{b_{n}}D_{\gamma^{x}_{\lambda_{n}}(s)}f\big{(}\dot{\gamma}^{x}_{\lambda_{n}}(s)\big{)}ds
=1bnan(f(γλnx(bn))f(γλnx(an)))0,\displaystyle=\frac{1}{b_{n}-a_{n}}\Big{(}f\big{(}\gamma^{x}_{\lambda_{n}}(b_{n})\big{)}-f\big{(}\gamma^{x}_{\lambda_{n}}(a_{n})\big{)}\Big{)}\to 0,

as n+n\to+\infty, which implies that μ~\tilde{\mu} is closed. Since γλnx\gamma^{x}_{\lambda_{n}} is a calibrated curve, we have

TM[LG(x,v,λnuλn(x))+c0]𝑑μ~n(x,v)\displaystyle\int_{TM}\bigg{[}L_{G}\big{(}x,v,\lambda_{n}u_{\lambda_{n}}(x)\big{)}+c_{0}\bigg{]}d\tilde{\mu}_{n}(x,v)
=1bnananbn[LG(γλnx(s),γ˙λnx(s),λnuλn(γλnx(s)))+c0]𝑑s\displaystyle=\frac{1}{b_{n}-a_{n}}\int_{a_{n}}^{b_{n}}\bigg{[}L_{G}\Big{(}\gamma^{x}_{\lambda_{n}}(s),\dot{\gamma}^{x}_{\lambda_{n}}(s),\lambda_{n}u_{\lambda_{n}}\big{(}\gamma^{x}_{\lambda_{n}}(s)\big{)}\Big{)}+c_{0}\bigg{]}ds
=1bnan(uλn(γλnx(bn))uλn(γλnx(an))).\displaystyle=\frac{1}{b_{n}-a_{n}}\Big{(}u_{\lambda_{n}}\big{(}\gamma^{x}_{\lambda_{n}}(b_{n})\big{)}-u_{\lambda_{n}}\big{(}\gamma^{x}_{\lambda_{n}}(a_{n})\big{)}\Big{)}.

Since uλu_{\lambda} is bounded, letting n+n\to+\infty, we find that

TMLG(x,v,0)𝑑μ~=c0.\int_{TM}L_{G}(x,v,0)d\tilde{\mu}=-c_{0}.

Therefore, the limit μ~\tilde{\mu} is a Mather measure of HH. By (2.5), we get

TMLGu(x,v,0)𝑑μ~(x,v)ε2,\int_{TM}\frac{\partial L_{G}}{\partial u}(x,v,0)d\tilde{\mu}(x,v)\leqslant-\varepsilon_{2},

which contradicts (2.3). The right-hand side inequality in (2.4) can be proved similarly. ∎

In the following, we will denote by λ¯>0\bar{\lambda}>0 and T0>0T_{0}>0 the constants given by Lemma 2.10.

Lemma 2.11.

Let λ(0,λ¯)\lambda\in(0,\bar{\lambda}) and xMx\in M. The following holds:

  • (i)

    for any t(,T0]t\in(-\infty,-T_{0}], we have

    ε2tt0LGu(γλx(s),γ˙λx(s),0)𝑑sε1t.\varepsilon_{2}t\leqslant\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds\leqslant\varepsilon_{1}t.

    As a consequence, eλ0LGu(γλx(s),γ˙λx(s),0)𝑑s=0.\displaystyle e^{\lambda\int_{-\infty}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)\,ds}=0.

  • (ii)

    For any TT0T\geqslant T_{0}, we have

    eλε2T0eλε2Tλε2T0eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s𝑑t1λε1+eλKT01λK.\frac{e^{-\lambda\varepsilon_{2}T_{0}}-e^{-\lambda\varepsilon_{2}T}}{\lambda\varepsilon_{2}}\leqslant\int_{-T}^{0}e^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}dt\leqslant\frac{1}{\lambda\varepsilon_{1}}+\frac{e^{\lambda KT_{0}}-1}{\lambda K}.

    In particular,

    eλε2T0λε20eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s𝑑t1λε1+eλKT01λK.\frac{e^{-\lambda\varepsilon_{2}T_{0}}}{\lambda\varepsilon_{2}}\leqslant\int_{-\infty}^{0}e^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}dt\leqslant\frac{1}{\lambda\varepsilon_{1}}+\frac{e^{\lambda KT_{0}}-1}{\lambda K}. (2.6)
Proof.

Item (i) is a direct consequence of Lemma 2.10. It remains to prove Item (ii). By Item (i) and (L1) we have

T0eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s𝑑t=TT0eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s𝑑t+T00eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s𝑑t\displaystyle\int_{-T}^{0}e^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}dt=\int_{-T}^{-T_{0}}\!\!e^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}dt+\int_{-T_{0}}^{0}\!\!e^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}dt
TT0eλε1t𝑑t+T00eλKt𝑑t=eλε1T0eλε1Tλε1+eλKT01λK1λε1+eλKT01λK.\displaystyle\leqslant\int_{-T}^{-T_{0}}e^{\lambda\varepsilon_{1}t}dt+\int_{-T_{0}}^{0}e^{-\lambda Kt}dt=\frac{e^{-\lambda\varepsilon_{1}T_{0}}-e^{-\lambda\varepsilon_{1}T}}{\lambda\varepsilon_{1}}+\frac{e^{\lambda KT_{0}}-1}{\lambda K}\leqslant\frac{1}{\lambda\varepsilon_{1}}+\frac{e^{\lambda KT_{0}}-1}{\lambda K}.

For the other side, we have

T0eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s𝑑t\displaystyle\int_{-T}^{0}e^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}dt TT0eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s𝑑t\displaystyle\geqslant\int_{-T}^{-T_{0}}e^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}dt
TT0eλε2t𝑑t=eλε2T0eλε2Tλε2.\displaystyle\geqslant\int_{-T}^{-T_{0}}e^{\lambda\varepsilon_{2}t}dt=\frac{e^{-\lambda\varepsilon_{2}T_{0}}-e^{-\lambda\varepsilon_{2}T}}{\lambda\varepsilon_{2}}.

Let T+T\to+\infty, we then get (2.6). ∎

We proceed by associating to each calibrated curve a probability measure on TMTM. These probability measures will play a key role in the proof of the convergence result stated in Theorem 2.2. They can be regarded as a generalization to the case at issue of the analogous measures first introduced in [11, formula (3.5)]. They already appeared in this exact form in [5, 41, 7].

Definition 2.12.

We define probability measures μ~λx\tilde{\mu}^{x}_{\lambda} on TMTM by

TMf(y,v)𝑑μ~λx(y,v)=0f(γλx(t),γ˙λx(t))eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s𝑑t0eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s𝑑t,fCc(TM).\int_{TM}f(y,v)d\tilde{\mu}^{x}_{\lambda}(y,v)=\frac{\int_{-\infty}^{0}f\big{(}\gamma^{x}_{\lambda}(t),\dot{\gamma}^{x}_{\lambda}(t)\big{)}e^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}dt}{\int_{-\infty}^{0}e^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}dt},\quad\forall f\in C_{c}(TM).

By (2.6), the measure μ~λx\tilde{\mu}^{x}_{\lambda} is well-defined for λ(0,λ¯)\lambda\in(0,\bar{\lambda}).

The following holds, cf. [11, Proposition 3.6], [5, Proposition 4.5], [41, Proposition 5.8], [7, Proposition 7.5].

Lemma 2.13.

The family (μ~λx)λ(0,λ¯)(\tilde{\mu}^{x}_{\lambda})_{\lambda\in(0,\bar{\lambda})} has support contained in a common compact subset of TMTM, in particular it is relatively compact in 𝒫(TM)\mathscr{P}(TM). Furthermore, if μ~λnxμ~\tilde{\mu}^{x}_{\lambda_{n}}\stackrel{{\scriptstyle\ast}}{{\rightharpoonup}}\tilde{\mu} in 𝒫(TM)\mathscr{P}(TM) for λn0+\lambda_{n}\to 0^{+}, then μ~\tilde{\mu} is a Mather measure.

Proof.

The first part is a direct consequence of Lemma 2.9. It remains to prove that the limit μ~\tilde{\mu} is a Mather measure. We first prove that μ~\tilde{\mu} is closed. For fC1(M)f\in C^{1}(M), we have, by integrating by parts,

TMDxf(v)𝑑μ~λx(y,v)\displaystyle\int_{TM}D_{x}f(v)d\tilde{\mu}^{x}_{\lambda}(y,v) =0ddt(f(γλx(t)))eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s𝑑t0eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s𝑑t\displaystyle=\frac{\int_{-\infty}^{0}\frac{d}{dt}\Big{(}f\big{(}\gamma^{x}_{\lambda}(t)\big{)}\Big{)}e^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}dt}{\int_{-\infty}^{0}e^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}dt}
=f(x)0f(γλx(t))ddt(eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s)𝑑t0eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s𝑑t.\displaystyle=\frac{f(x)-\int_{-\infty}^{0}f\big{(}\gamma^{x}_{\lambda}(t)\big{)}\frac{d}{dt}\Big{(}e^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}\Big{)}dt}{{\int_{-\infty}^{0}e^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}dt}}.

By (L1) and eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s>0e^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}>0, we have

|0f(γλx(t))ddt(eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s)𝑑t|\displaystyle\bigg{|}\int_{-\infty}^{0}f\big{(}\gamma^{x}_{\lambda}(t)\big{)}\frac{d}{dt}\Big{(}e^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}\Big{)}dt\bigg{|}
=|0λf(γλx(t))LGu(γλx(t),γ˙λx(t),0)eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s𝑑t|\displaystyle=\bigg{|}\int_{-\infty}^{0}\lambda f\big{(}\gamma^{x}_{\lambda}(t)\big{)}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(t),\dot{\gamma}^{x}_{\lambda}(t),0)e^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}dt\bigg{|}
λfK0eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s𝑑tK(1ε1+eλKT01K)f,\displaystyle\leqslant\lambda\|f\|_{\infty}K\int_{-\infty}^{0}e^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}dt\leqslant K\bigg{(}\frac{1}{\varepsilon_{1}}+\frac{e^{\lambda KT_{0}}-1}{K}\bigg{)}\|f\|_{\infty},

where, for the last inequality, we have used (2.6). By (2.6) again, we have

TMDxf(v)𝑑μ~λx(y,v)\displaystyle\int_{TM}D_{x}f(v)d\tilde{\mu}^{x}_{\lambda}(y,v) λε2eλε2T0K(1ε1+eλKT0K)f0,\displaystyle\leqslant\frac{\lambda\varepsilon_{2}}{e^{-\lambda\varepsilon_{2}T_{0}}}K\bigg{(}\frac{1}{\varepsilon_{1}}+\frac{e^{\lambda KT_{0}}}{K}\bigg{)}\|f\|_{\infty}\to 0,

as λ0+\lambda\to 0^{+}.

We then prove that μ~\tilde{\mu} is minimizing. Since tuλ(γλx(t))t\mapsto u_{\lambda}\big{(}\gamma^{x}_{\lambda}(t)\big{)} is Lipschitz continuous, and γλx\gamma^{x}_{\lambda} is a (uλ,Lλ,c0)(u_{\lambda},L_{\lambda},c_{0})-calibrated curve, for a.e. t<0t<0 we have

ddtuλ(γλx(t))=LG(γλx(t),γ˙λx(t),λuλ(γλx(t)))+c0.\frac{d}{dt}u_{\lambda}\big{(}\gamma^{x}_{\lambda}(t)\big{)}=L_{G}\Big{(}\gamma^{x}_{\lambda}(t),\dot{\gamma}^{x}_{\lambda}(t),\lambda u_{\lambda}\big{(}\gamma^{x}_{\lambda}(t)\big{)}\Big{)}+c_{0}.

Then

TM(LG(x,v,0)+c0)𝑑μ~λx(x,v)\displaystyle\int_{TM}\big{(}L_{G}(x,v,0)+c_{0}\big{)}d\tilde{\mu}^{x}_{\lambda}(x,v)
=0(LG(γλx(t),γ˙λx(t),0)+c0)eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s𝑑t0eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s𝑑t\displaystyle=\frac{\int_{-\infty}^{0}\big{(}L_{G}(\gamma^{x}_{\lambda}(t),\dot{\gamma}^{x}_{\lambda}(t),0)+c_{0}\big{)}e^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}dt}{\int_{-\infty}^{0}e^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}dt}
=0(LG(γλx(t),γ˙λx(t),λuλ(γλx(t)))+c0Δλ(t))eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s𝑑t0eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s𝑑t\displaystyle=\frac{\int_{-\infty}^{0}\Big{(}L_{G}\Big{(}\gamma^{x}_{\lambda}(t),\dot{\gamma}^{x}_{\lambda}(t),\lambda u_{\lambda}\big{(}\gamma^{x}_{\lambda}(t)\big{)}\Big{)}+c_{0}-\Delta_{\lambda}(t)\Big{)}e^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}dt}{\int_{-\infty}^{0}e^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}dt}
=0(ddt(uλ(γλx(t)))Δλ(t))eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s𝑑t0eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s𝑑t,\displaystyle=\frac{\int_{-\infty}^{0}\Big{(}\frac{d}{dt}\Big{(}u_{\lambda}\big{(}\gamma^{x}_{\lambda}(t)\big{)}\Big{)}-\Delta_{\lambda}(t)\Big{)}e^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}dt}{\int_{-\infty}^{0}e^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}dt},

where

Δλ(t)=LG(γλx(t),γ˙λx(t),λuλ(γλx(t)))LG(γλx(t),γ˙λx(t),0).\Delta_{\lambda}(t)=L_{G}\Big{(}\gamma^{x}_{\lambda}(t),\dot{\gamma}^{x}_{\lambda}(t),\lambda u_{\lambda}\big{(}\gamma^{x}_{\lambda}(t)\big{)}\Big{)}-L_{G}\big{(}\gamma^{x}_{\lambda}(t),\dot{\gamma}^{x}_{\lambda}(t),0\big{)}.

Similarly to the first part of the proof, we have

0ddtuλ(γλx(t))eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s𝑑t0eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s𝑑tλε2eλε2T0K(1ε1+eλKT0K)C0,\displaystyle\frac{\int_{-\infty}^{0}\frac{d}{dt}u_{\lambda}\big{(}\gamma^{x}_{\lambda}(t)\big{)}e^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}dt}{\int_{-\infty}^{0}e^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}dt}\leqslant\frac{\lambda\varepsilon_{2}}{e^{-\lambda\varepsilon_{2}T_{0}}}K\bigg{(}\frac{1}{\varepsilon_{1}}+\frac{e^{\lambda KT_{0}}}{K}\bigg{)}C\to 0,

as λ0+\lambda\to 0^{+}, where CC is a uniform bound on the uλ\|u_{\lambda}\|_{\infty}.

Finally, to bound the error term we use (L1) to find

|Δλ(t)|=|LG(γλx(t),γ˙λx(t),λuλ(γλx(t)))L(γλx(t),γ˙λx(t),0)|λKC.\displaystyle|\Delta_{\lambda}(t)|=\Big{|}L_{G}\Big{(}\gamma^{x}_{\lambda}(t),\dot{\gamma}^{x}_{\lambda}(t),\lambda u_{\lambda}\big{(}\gamma^{x}_{\lambda}(t)\big{)}\Big{)}-L\big{(}\gamma^{x}_{\lambda}(t),\dot{\gamma}^{x}_{\lambda}(t),0\big{)}\Big{|}\leqslant\lambda KC.

Therefore, as λ0+\lambda\to 0^{+} along the sequence (λn)(\lambda_{n}), we conclude that μ~\tilde{\mu} is a Mather measure.∎

The following lemma will be crucial for the proof of the convergence result, cf. [11, Lemma 3.7], [5, Lemma 4.7], [7, Lemma 7.7].

Lemma 2.14.

Let ww be any subsolution of (E0{\textrm{E}}_{0}). For every xMx\in M and λ(0,λ¯)\lambda\in(0,\bar{\lambda}), we have

uλ(x)w(x)TMw(y)LGu(y,v,0)𝑑μ~λx(y,v)TMLGu(y,v,0)𝑑μ~λx(y,v)+Rλ(x),u_{\lambda}(x)\geqslant w(x)-\frac{\int_{TM}w(y)\frac{\partial L_{G}}{\partial u}(y,v,0)d\tilde{\mu}^{x}_{\lambda}(y,v)}{\int_{TM}\frac{\partial L_{G}}{\partial u}(y,v,0)d\tilde{\mu}^{x}_{\lambda}(y,v)}+R_{\lambda}(x),

where limλ0+Rλ(x)=0\lim\limits_{\lambda\to 0^{+}}R_{\lambda}(x)=0.

Proof.

For ε>0\varepsilon>0, using Theorem 1.15 we take wεC(M)w_{\varepsilon}\in C^{\infty}(M) such that wεwε\|w_{\varepsilon}-w\|_{\infty}\leqslant\varepsilon and

G(x,Dxwε,0)c0+ε,xM.G(x,D_{x}w_{\varepsilon},0)\leqslant c_{0}+\varepsilon,\quad\forall x\in M.

Using the Fenchel inequality we have for all (x,v)TM(x,v)\in TM,

LG(x,v,0)\displaystyle L_{G}(x,v,0) LG(x,v,0)+G(x,Dxwε,0)c0ε\displaystyle\geqslant L_{G}(x,v,0)+G(x,D_{x}w_{\varepsilon},0)-c_{0}-\varepsilon
Dxwε(v)c0ε.\displaystyle\geqslant D_{x}w_{\varepsilon}(v)-c_{0}-\varepsilon.

Since tuλ(γλx(t))t\mapsto u_{\lambda}\big{(}\gamma^{x}_{\lambda}(t)\big{)} is Lipschitz continuous, and γλx\gamma^{x}_{\lambda} is a (uλ,Lλ,c0)(u_{\lambda},L_{\lambda},c_{0})-calibrated curve, for a.e. t<0t<0 we have

ddtuλ(γλx(t))\displaystyle\frac{d}{dt}u_{\lambda}\big{(}\gamma^{x}_{\lambda}(t)\big{)} =LG(γλx(t),γ˙λx(t),λuλ(γλx(t)))+c0\displaystyle=L_{G}\Big{(}\gamma^{x}_{\lambda}(t),\dot{\gamma}^{x}_{\lambda}(t),\lambda u_{\lambda}\big{(}\gamma^{x}_{\lambda}(t)\big{)}\Big{)}+c_{0} (2.7)
LG(γλx(t),γ˙λx(t),λuλ(γλx(t)))LG(γλx(t),γ˙λx(t),0)+Dγλx(t)wε(γ˙λx(t))ε\displaystyle\geqslant L_{G}\Big{(}\gamma^{x}_{\lambda}(t),\dot{\gamma}^{x}_{\lambda}(t),\lambda u_{\lambda}\big{(}\gamma^{x}_{\lambda}(t)\big{)}\Big{)}-L_{G}\big{(}\gamma^{x}_{\lambda}(t),\dot{\gamma}^{x}_{\lambda}(t),0\big{)}+D_{\gamma^{x}_{\lambda}(t)}w_{\varepsilon}(\dot{\gamma}^{x}_{\lambda}(t))-\varepsilon
=ddtwε(γλx(t))+λLGu(γλx(s),γ˙λx(s),0)uλ(γλx(t))ε+Ωλ,x(t),\displaystyle=\frac{d}{dt}w_{\varepsilon}\big{(}\gamma^{x}_{\lambda}(t)\big{)}+\lambda\frac{\partial L_{G}}{\partial u}\big{(}\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0\big{)}u_{\lambda}\big{(}\gamma^{x}_{\lambda}(t)\big{)}-\varepsilon+\Omega_{\lambda,x}(t),

where

Ωλ,x(t):=\displaystyle\Omega_{\lambda,x}(t):= LG(γλx(t),γ˙λx(t),λuλ(γλx(t)))\displaystyle L_{G}\Big{(}\gamma^{x}_{\lambda}(t),\dot{\gamma}^{x}_{\lambda}(t),\lambda u_{\lambda}\big{(}\gamma^{x}_{\lambda}(t)\big{)}\Big{)}
LG(γλx(t),γ˙λx(t),0)λLGu(γλx(s),γ˙λx(s),0)uλ(γλx(t)).\displaystyle-L_{G}\big{(}\gamma^{x}_{\lambda}(t),\dot{\gamma}^{x}_{\lambda}(t),0\big{)}-\lambda\frac{\partial L_{G}}{\partial u}\big{(}\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0\big{)}u_{\lambda}\big{(}\gamma^{x}_{\lambda}(t)\big{)}.

Let us estimate the error term Ωλ,x(t)\Omega_{\lambda,x}(t). Let us set S:={(x,v)TM:vxκ^}S:=\{(x,v)\in TM\,:\,\|v\|_{x}\leqslant\hat{\kappa}\,\}, where κ^>0\hat{\kappa}>0 is the Lipschitz constant of the curves {γλx:λ(0,λ0)}\{\gamma^{x}_{\lambda}\,:\,\lambda\in(0,\lambda_{0})\,\}, according to Lemma 2.9. By (L4) we have

|Ωλ,x(t)|λCηS(λC)for all t0 and λ(0,λ0),|\Omega_{\lambda,x}(t)|\leqslant\lambda C\eta_{S}(\lambda C)\qquad\hbox{for all $t\leqslant 0$ and $\lambda\in(0,\lambda_{0})$,} (2.8)

where CC is a uniform bound on the uλ\|u_{\lambda}\|_{\infty}. By multiplying both sides of (2.7) by eλt0LGu(γλx(s),γ˙λx(s),0)𝑑se^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds} and by rearranging terms, we obtain, for a.e. t<0t<0,

ddt(uλ(γλx(t))eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s)(ddtwε(γλx(t))ε+Ωλ,x(t))eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s.\displaystyle\frac{d}{dt}\bigg{(}u_{\lambda}\big{(}\gamma^{x}_{\lambda}(t)\big{)}e^{\lambda\int_{t}^{0}\!\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}\bigg{)}\geqslant\bigg{(}\frac{d}{dt}w_{\varepsilon}\big{(}\gamma^{x}_{\lambda}(t)\big{)}-\varepsilon+\Omega_{\lambda,x}(t)\bigg{)}e^{\lambda\int_{t}^{0}\!\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}.

Integrating the above inequality over the interval (T,0](-T,0] where TT0T\geqslant T_{0} as stated in Lemma 2.11, and using an integration by parts, we have

uλ(x)uλ(γλx(T))eλT0LGu(γλx(s),γ˙λx(s),0)𝑑s\displaystyle u_{\lambda}(x)-u_{\lambda}\big{(}\gamma^{x}_{\lambda}(-T)\big{)}e^{\lambda\int_{-T}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}
wε(x)wε(γλx(T))eλT0LGu(γλx(s),γ˙λx(s),0)𝑑s\displaystyle\quad\geqslant w_{\varepsilon}(x)-w_{\varepsilon}\big{(}\gamma^{x}_{\lambda}(-T)\big{)}e^{\lambda\int_{-T}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}
T0wε(γλx(t))ddt(eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s)𝑑t+T0(Ωλ,x(t)ε)eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s.\displaystyle\quad-\int_{-T}^{0}w_{\varepsilon}\big{(}\gamma^{x}_{\lambda}(t)\big{)}\frac{d}{dt}\bigg{(}e^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}\bigg{)}dt+\int_{-T}^{0}(\Omega_{\lambda,x}(t)-\varepsilon)e^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}.

Letting ε0+\varepsilon\to 0^{+} it follows that,

uλ(x)\displaystyle u_{\lambda}(x) w(x)(w+C)eλT0LGu(γλx(s),γ˙λx(s),0)𝑑s\displaystyle\geqslant w(x)-(\|w\|_{\infty}+C)e^{\lambda\int_{-T}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}
T0w(γλx(t))ddt(eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s)𝑑t+T0Ωλ,x(t)eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s.\displaystyle-\int_{-T}^{0}w\big{(}\gamma^{x}_{\lambda}(t)\big{)}\frac{d}{dt}\bigg{(}e^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}\bigg{)}dt+\int_{-T}^{0}\Omega_{\lambda,x}(t)e^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}.

From Lemma 2.11 we infer that the maps

teλt0LGu(γλx(s),γ˙λx(s),0)𝑑sandtddt(eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s)t\mapsto e^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}\qquad\hbox{and}\qquad t\mapsto\frac{d}{dt}\bigg{(}e^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}\bigg{)}

are in L1((,0])L^{1}\big{(}(-\infty,0]\big{)} and converge to 0 as tt\to-\infty. By taking also into account (2.7), we can send T+T\to+\infty in the above inequality, to get, by the Dominated Convergence Theorem,

uλ(x)\displaystyle u_{\lambda}(x) w(x)\displaystyle\geqslant w(x)
0w(γλx(t))ddt(eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s)𝑑t+0Ωλ,x(t)eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s\displaystyle-\int_{-\infty}^{0}w\big{(}\gamma^{x}_{\lambda}(t)\big{)}\frac{d}{dt}\bigg{(}e^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}\bigg{)}dt+\int_{-\infty}^{0}\Omega_{\lambda,x}(t)e^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}
=:w(x)Iλ+Rλ(x).\displaystyle=:w(x)-I_{\lambda}+R_{\lambda}(x).

By Definition 2.12, we have

Iλ\displaystyle I_{\lambda} =λ0w(γλx(t))LGu(γλx(s),γ˙λx(s),0)eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s𝑑t\displaystyle=-\lambda\int_{-\infty}^{0}w\big{(}\gamma^{x}_{\lambda}(t)\big{)}\frac{\partial L_{G}}{\partial u}\big{(}\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0\big{)}e^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}dt
=λ(0eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s𝑑t)TMw(y)LGu(y,v,0)𝑑μ~λx(y,v).\displaystyle=-\lambda\Big{(}\int_{-\infty}^{0}e^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}dt\Big{)}\int_{TM}w(y)\frac{\partial L_{G}}{\partial u}(y,v,0)\,d\tilde{\mu}^{x}_{\lambda}(y,v).

According to Lemma 2.11-(i), for λ(0,λ0)\lambda\in(0,\lambda_{0}), we derive that

λ0eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s𝑑t\displaystyle\lambda\int_{-\infty}^{0}e^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}dt =λ0eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s𝑑t0ddt(eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s)𝑑t\displaystyle=\frac{\lambda\int_{-\infty}^{0}e^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}dt}{\int_{-\infty}^{0}\frac{d}{dt}\bigg{(}e^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}\bigg{)}dt}
=0eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s𝑑t0LGu(γλx(t),γ˙λx(t),0)eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s𝑑t\displaystyle=-\frac{\int_{-\infty}^{0}e^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}dt}{\int_{-\infty}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}({t}),\dot{\gamma}^{x}_{\lambda}({t}),0)e^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}dt}
=1TMLGu(y,v,0)𝑑μ~λx(y,v).\displaystyle=-\frac{1}{\int_{TM}\frac{\partial L_{G}}{\partial u}(y,v,0)d\tilde{\mu}^{x}_{\lambda}(y,v)}.

By (2.7) we also have

|Rλ(x)|λCηS(λC)0eλt0LGu(γλx(s),γ˙λx(s),0)𝑑s(1ε1+eλKT01K)CηS(λC).|R_{\lambda}(x)|\leqslant\lambda C\eta_{S}(\lambda C)\int_{-\infty}^{0}e^{\lambda\int_{t}^{0}\frac{\partial L_{G}}{\partial u}(\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s),0)ds}\leqslant\bigg{(}\frac{1}{\varepsilon_{1}}+\frac{e^{\lambda KT_{0}}-1}{K}\bigg{)}C\eta_{S}(\lambda C).

The assertion follows by sending λ0+\lambda\to 0^{+}. ∎

We are now in position to prove the first two main theorems of this section.

Proof of Theorem 2.2.

Let uu^{*} be an accumulation point of the (uλ)λ(0,λ0)(u_{\lambda})_{\lambda\in(0,\lambda_{0})} as λ0+\lambda\to 0^{+}. By Lemma 2.8, we know that u𝒮u^{*}\in\mathcal{S}, so uu0u^{*}\leqslant u_{0} in MM. Let us prove that uu0u^{*}\geqslant u_{0}. Fix xMx\in M and pick w𝒮w\in\mathcal{S}. From Lemmas 2.13 and 2.14 we infer that

u(x)w(x)TMw(y)LGu(y,v,0)𝑑μ~TMLGu(y,v,0)𝑑μ~,u^{*}(x)\geqslant w(x)-\frac{\int_{TM}w(y)\frac{\partial L_{G}}{\partial u}(y,v,0)d\tilde{\mu}}{\int_{TM}\frac{\partial L_{G}}{\partial u}(y,v,0)d\tilde{\mu}},

for some Mather measure μ~\tilde{\mu}. Since ww satisfies the constraint (S), we get from this that u(x)w(x)u^{*}(x)\geqslant w(x), hence u(x)supw𝒮w(x)=:u0(x)u^{*}(x)\geqslant\sup\limits_{w\in\mathcal{S}}w(x)=:u_{0}(x). By the arbitrariness of the choice of xMx\in M, we get that u0u_{0} is finite-valued and that uu0u^{*}\geqslant u_{0} in MM. We conclude that u0u_{0} is the unique accumulation point of the family of functions (uλ)λ(0,λ0)(u_{\lambda})_{\lambda\in(0,\lambda_{0})} as λ0+\lambda\to 0^{+}. The proof is complete. ∎

Proof of Theorem 2.3.

We denote

u^0(x):=infμ~𝔐~TMh(y,x)LGu(y,v,0)𝑑μ~(y,v)TMLGu(y,v,0)𝑑μ~(y,v),xM.\hat{u}_{0}(x):=\inf_{\tilde{\mu}\in\widetilde{\mathfrak{M}}}\frac{\int_{TM}h(y,x)\frac{\partial L_{G}}{\partial u}(y,v,0)d\tilde{\mu}(y,v)}{\int_{TM}\frac{\partial L_{G}}{\partial u}(y,v,0)d\tilde{\mu}(y,v)},\qquad x\in M.

Note that u^0\hat{u}_{0} is finite-valued, as u^0minM×Mh>\hat{u}_{0}\geq\min\limits_{M\times M}h>-\infty. We start by remarking that u^0\hat{u}_{0} is a subsolution of (E0{\textrm{E}}_{0}). Indeed, for every fixed μ~~\tilde{\mu}\in\widetilde{\mathcal{M}}, the function hμ~:M,xTMh(y,x)𝑑μ~(y)h_{\tilde{\mu}}:M\to\mathbb{R},\ x\mapsto\int_{TM}h(y,x)\,d{\tilde{\mu}}(y) is a convex combination of the family of critical solutions (hy)yM(h_{y})_{y\in M}, where hy(x)=h(y,x)h_{y}(x)=h(y,x). By the convexity of HH in the momentum and the equi-Lipschitz character of the critical subsolutions, see Propositions 1.3 and 1.4, it follows that each hμ~h_{\tilde{\mu}} is a critical subsolution. By Proposition 1.3 again, we infer that a finite valued infimum of critical subsolutions is itself a critical subsolution. Therefore u^0\hat{u}_{0} is a critical subsolution.

Let us now prove that u0u^0u_{0}\leqslant\hat{u}_{0}. By Proposition 1.4 we know that u0(x)u0(y)+h(y,x)u_{0}(x)\leqslant u_{0}(y)+h(y,x) for all x,yMx,y\in M. Let us integrate this inequality with respect to a Mather measure μ~𝔐~\tilde{\mu}\in\widetilde{\mathfrak{M}}. By assumption and by definition of Mather set, we have LGu(x,v,0)0\frac{\partial L_{G}}{\partial u}(x,v,0)\leqslant 0 for all (x,v)supp(μ~)(x,v)\in\textrm{supp}(\tilde{\mu}). We infer

u0(x)TMLGu(y,v,0)𝑑μ~(y,v)\displaystyle u_{0}(x)\int_{TM}\frac{\partial L_{G}}{\partial u}(y,v,0)\,d\tilde{\mu}(y,v)
TMu0(y)LGu(y,v,0)𝑑μ~(y,v)+TMh(y,x)LGu(y,v,0)𝑑μ~(y,v)\displaystyle\geqslant\int_{TM}u_{0}(y)\frac{\partial L_{G}}{\partial u}(y,v,0)\,d\tilde{\mu}(y,v)+\int_{TM}h(y,x)\frac{\partial L_{G}}{\partial u}(y,v,0)\,d\tilde{\mu}(y,v)
TMh(y,x)LGu(y,v,0)𝑑μ~(y,v),\displaystyle\geqslant\int_{TM}h(y,x)\frac{\partial L_{G}}{\partial u}(y,v,0)\,d\tilde{\mu}(y,v),

where, for the last inequality, we used that u0u_{0} satisfies (S). From this we get

u0(x)TMh(y,x)LGu(y,v,0)𝑑μ~(y,v)TMLGu(y,v,0)𝑑μ~(y,v)for all xM.u_{0}(x)\leqslant\dfrac{\int_{TM}h(y,x)\frac{\partial L_{G}}{\partial u}(y,v,0)\,d\tilde{\mu}(y,v)}{\int_{TM}\frac{\partial L_{G}}{\partial u}(y,v,0)\,d\tilde{\mu}(y,v)}\qquad\hbox{for all $x\in M$.}

By taking the inf with respect to μ~~\tilde{\mu}\in{\widetilde{\mathcal{M}}} of the right-hand side term in the above inequality, we get u0u^0u_{0}\leqslant\hat{u}_{0}.

Last, we prove that u0u^0u_{0}\geqslant\hat{u}_{0}. For every fixed z𝒜z\in\mathcal{A}, set

Uz(x):=h(x,z)+u^0(z),xM.U_{z}(x):=-h(x,z)+\hat{u}_{0}(z),\qquad x\in M.

Then UzU_{z} is a subsolution of (E0{\textrm{E}}_{0}), by Proposition 1.5. Furthermore, for every xMx\in M and μ~𝔐~\tilde{\mu}\in\widetilde{\mathfrak{M}}, we have

TMUz(x)LGu(x,v,0)𝑑μ~(x,v)TMLGu(x,v,0)𝑑μ~(x,v)=TMh(x,z)LGu(x,v,0)𝑑μ~(x,v)TMLGu(x,v,0)𝑑μ~(x,v)+u^0(z)0,\frac{\int_{TM}U_{z}(x)\,\frac{\partial L_{G}}{\partial u}(x,v,0)d\tilde{\mu}(x,v)}{\int_{TM}\frac{\partial L_{G}}{\partial u}(x,v,0)d\tilde{\mu}(x,v)}=-\frac{\int_{TM}h(x,z)\,\frac{\partial L_{G}}{\partial u}(x,v,0)d\tilde{\mu}(x,v)}{\int_{TM}\frac{\partial L_{G}}{\partial u}(x,v,0)d\tilde{\mu}(x,v)}+\hat{u}_{0}(z)\leqslant 0,

which implies that Uz𝒮U_{z}\in\mathcal{S}. Then u0(x)Uz(x)u_{0}(x)\geqslant U_{z}(x) for all xMx\in M. In particular, by taking x=z𝒜x=z\in\mathcal{A}, we get

u0(z)Uz(z)=h(z,z)+u^0(z)=u^0(z).u_{0}(z)\geqslant U_{z}(z)=-h(z,z)+\hat{u}_{0}(z)=\hat{u}_{0}(z).

We have thus shown that u0u^0u_{0}\geqslant\hat{u}_{0} on 𝒜\mathcal{A}, hence u0u^0u_{0}\geqslant\hat{u}_{0} on MM according according to Proposition 1.7. ∎

We proceed by proving the third main theorem of this section, namely the trichotomy result stated in Theorem 2.5. We start by establishing a sort of Harnack-type inequality for subsolutions of (Eλ).

Lemma 2.15.

There exists a constant A+>0A_{+}>0 such that, if λ(0,1)\lambda\in(0,1) and wλ:Mw_{\lambda}:M\to\mathbb{R} is a subsolution to (Eλ), then

minxMwλ(x)maxxMwλ(x)minxMwλ(x)+A+(1+λminxM|wλ(x)|).\min_{x\in M}w_{\lambda}(x)\leqslant\max_{x\in M}w_{\lambda}(x)\leqslant\min_{x\in M}w_{\lambda}(x)+A_{+}\Big{(}1+\lambda\min_{x\in M}|w_{\lambda}(x)|\Big{)}. (2.9)
Proof.

The left hand side inequality is obvious. Let us prove the right hand side inequality. Let xλMx_{\lambda}\in M such that Mλ=wλ(xλ)=maxwλM_{\lambda}=w_{\lambda}(x_{\lambda})=\max w_{\lambda}. Let yλMy_{\lambda}\in M such that mλ=wλ(yλ)=minwλm_{\lambda}=w_{\lambda}(y_{\lambda})=\min w_{\lambda}. Denote dλ:=d(yλ,xλ)d_{\lambda}:=d(y_{\lambda},x_{\lambda}). Let ζ:[0,dλ]M\zeta:[0,d_{\lambda}]\rightarrow M be a geodesic satisfying ζ(0)=yλ\zeta(0)=y_{\lambda} and ζ(dλ)=xλ\zeta(d_{\lambda})=x_{\lambda} with constant speed 1. By wλLλ+c0w_{\lambda}\prec L_{\lambda}+c_{0} and (L1), we get

wλ(ζ(s))\displaystyle w_{\lambda}\big{(}\zeta(s)\big{)} wλ(yλ)+0s[LG(ζ(τ),ζ˙(τ),λwλ(ζ(τ)))+c0]𝑑τ\displaystyle\leqslant w_{\lambda}(y_{\lambda})+\int_{0}^{s}\Big{[}{L_{G}}\Big{(}\zeta(\tau),\dot{\zeta}(\tau),\lambda w_{\lambda}\big{(}\zeta(\tau)\big{)}\Big{)}+c_{0}\Big{]}d\tau
mλ+0s[LG(ζ(τ),ζ˙(τ),λmλ)+c0+λK(wλ(ζ(τ))mλ)]𝑑τ\displaystyle\leqslant m_{\lambda}+\int_{0}^{s}\Big{[}{L_{G}}\big{(}\zeta(\tau),\dot{\zeta}(\tau),\lambda m_{\lambda}\big{)}+c_{0}+\lambda K\big{(}w_{\lambda}\big{(}\zeta(\tau)\big{)}-m_{\lambda}\big{)}\Big{]}d\tau
mλ+0s[LG(ζ(τ),ζ˙(τ),0)+c0+λK|mλ|]𝑑τ+λK0s[wλ(ζ(τ))mλ]𝑑τ\displaystyle\leqslant m_{\lambda}+\int_{0}^{s}\Big{[}{L_{G}}\big{(}\zeta(\tau),\dot{\zeta}(\tau),0\big{)}+c_{0}+\lambda K|m_{\lambda}|\Big{]}d\tau+\lambda K\int_{0}^{s}\Big{[}w_{\lambda}\big{(}\zeta(\tau)\big{)}-m_{\lambda}\Big{]}d\tau
mλ+(CLG+λK|mλ|)dλ+λK0s[wλ(ζ(τ))mλ]𝑑τ\displaystyle\leqslant m_{\lambda}+(C_{L_{G}}+\lambda K|m_{\lambda}|)d_{\lambda}+\lambda K\int_{0}^{s}\Big{[}w_{\lambda}\big{(}\zeta(\tau)\big{)}-m_{\lambda}\Big{]}d\tau

where CLG:=maxxM,vx1|LG(x,v,0)+c0|C_{L_{G}}:=\max\limits_{x\in M,\|v\|_{x}\leqslant 1}|{L_{G}}(x,v,0)+c_{0}|. By the Gronwall inequality we infer

wλ(ζ(s))mλ(CLG+λK|mλ|)dλeλKs(CLG+λK|mλ|)dλeλKdλ,s(0,dλ].w_{\lambda}\big{(}\zeta(s)\big{)}-m_{\lambda}\leqslant(C_{L_{G}}+\lambda K|m_{\lambda}|)d_{\lambda}e^{\lambda Ks}\leqslant(C_{L_{G}}+\lambda K|m_{\lambda}|)d_{\lambda}e^{\lambda Kd_{\lambda}},\quad\forall s\in(0,d_{\lambda}].

Taking s=dλs=d_{\lambda}, and recalling that λ(0,1)\lambda\in(0,1),we have

Mλ=wλ(xλ)mλ+(CLG+λK|mλ|)dλeλKdλmλ+(CLG+λK|mλ|)D¯eKD¯,M_{\lambda}=w_{\lambda}(x_{\lambda})\leqslant m_{\lambda}+(C_{L_{G}}+\lambda K|m_{\lambda}|)d_{\lambda}e^{\lambda Kd_{\lambda}}\leqslant m_{\lambda}+(C_{L_{G}}+\lambda K|m_{\lambda}|)\overline{D}e^{K\overline{D}},

where D¯:=\overline{D}:=diam(M)(M). The result follows taking A+=max(CLG,K)D¯eKD¯A_{+}=\max(C_{L_{G}},K)\overline{D}e^{K\overline{D}}. ∎

As a consequence, we derive the following key proposition. It will be also used in Section 3 to show the existence of diverging families of solutions.

Proposition 2.16.

Let Λ\Lambda be a subset of (0,1)(0,1) having 0 as accumulation point. Let (wλ)λΛ(w_{\lambda})_{\lambda\in\Lambda} be a family of subsolutions of (Eλ).

  • (i)

    If, for each λΛ\lambda\in\Lambda, there is a point xλMx_{\lambda}\in M such that wλ(xλ)w_{\lambda}(x_{\lambda})\to-\infty as λ0+\lambda\to 0^{+}, λΛ\lambda\in\Lambda, then wλw_{\lambda} uniformly converges to -\infty as λ0+\lambda\to 0^{+}, λΛ\lambda\in\Lambda.

  • (ii)

    If, for each λΛ\lambda\in\Lambda, there is a point xλMx_{\lambda}\in M such that wλ(xλ)+w_{\lambda}(x_{\lambda})\to+\infty as λ0+\lambda\to 0^{+}, λΛ\lambda\in\Lambda, then wλw_{\lambda} uniformly converges to ++\infty as λ0+\lambda\to 0^{+}, λΛ\lambda\in\Lambda.

Proof.

As in the previous proof, we denote Mλ=maxwλM_{\lambda}=\max w_{\lambda} and mλ=minwλm_{\lambda}=\min w_{\lambda}.

Let us prove assertion (i). The hypothesis wλ(xλ)w_{\lambda}(x_{\lambda})\to-\infty is equivalent to mλm_{\lambda}\to-\infty as λ0+\lambda\to 0^{+}, λΛ\lambda\in\Lambda. Then restricting to λΛ(0,1)\lambda\in\Lambda\cap(0,1), (2.9) implies that mλMλm_{\lambda}\sim M_{\lambda} hence MλM_{\lambda}\to-\infty as λ0+\lambda\to 0^{+}, λΛ\lambda\in\Lambda.

Let us prove item (ii). The hypothesis wλ(xλ)+w_{\lambda}(x_{\lambda})\to+\infty is equivalent to Mλ+M_{\lambda}\to+\infty as λ0+\lambda\to 0^{+}, λΛ\lambda\in\Lambda. Then restricting to λΛ(0,1)\lambda\in\Lambda\cap(0,1), (2.9) implies that mλMλm_{\lambda}\sim M_{\lambda} hence mλ+m_{\lambda}\to+\infty as λ0+\lambda\to 0^{+}, λΛ\lambda\in\Lambda. ∎

We are now in position to prove Theorem 2.5. We recall that, in what follows, u0u_{0} still denotes the critical solution provided by Theorem 2.2

Proof of Theorem 2.5.

Let us denote by SλS_{\lambda} the set of solutions to (Eλ). Define

Sλ:={vλis a solution of (Eλ) withvλ<u01}S_{\lambda}^{-}:=\{v_{\lambda}\ \textrm{is\ a\ solution\ of\ (\ref{E})\ with}\ v_{\lambda}<u_{0}-1\}

and

ψ(λ):=sup{vλ(x):vλSλ,xM}.\psi(\lambda):=\sup\{v_{\lambda}(x):\ v_{\lambda}\in S_{\lambda}^{-},\ x\in M\}.

If Sλ=S_{\lambda}^{-}=\varnothing, we set ψ(λ)=\psi(\lambda)=-\infty.

We claim that limλ0ψ(λ)=\lim\limits_{\lambda\to 0}\psi(\lambda)=-\infty. Otherwise, there is A>0A>0, sequences λn0+\lambda_{n}\to 0^{+}, vλnSλnv_{\lambda_{n}}\in S_{\lambda_{n}} and ynMy_{n}\in M such that

vλn(yn)>A,n.v_{\lambda_{n}}(y_{n})>-A,\quad\forall n\in\mathbb{N}. (2.10)

We claim that vλnv_{\lambda_{n}} is uniformly bounded from below. If not, there is znMz_{n}\in M such that

vλn(zn)as n+.v_{\lambda_{n}}(z_{n})\to-\infty\qquad\hbox{as $n\to+\infty$}.

By Proposition 2.16, vλnv_{\lambda_{n}} uniformly converges to -\infty, which contradicts (2.10). Then, according to Theorem 2.2 and Remark 2.4, the sequence vλnv_{\lambda_{n}} uniformly converges. to the only possible limit u0u_{0}, which contradicts the fact that vλnSλnv_{\lambda_{n}}\in S_{\lambda_{n}} for all nn\in\mathbb{N}.

In a similar manner, define

Sλ+:={vλis a solution of (Eλ) withvλ>u0+1}S_{\lambda}^{+}:=\{v_{\lambda}\ \textrm{is\ a\ solution\ of\ (\ref{E})\ with}\ v_{\lambda}>u_{0}+1\}

and

φ(λ):=inf{vλ(x):vλSλ+,xM}.\varphi(\lambda):=\inf\{v_{\lambda}(x):\ v_{\lambda}\in S_{\lambda}^{+},\ x\in M\}.

If Sλ+=S_{\lambda}^{+}=\varnothing, we set φ(λ)=+\varphi(\lambda)=+\infty. The same proof yields that limλ0φ(λ)=+\lim\limits_{\lambda\to 0}\varphi(\lambda)=+\infty.

Define

θ(λ):=sup{|vλ(x)u0(x)|,vλSλ(SλSλ+),xM}.\theta(\lambda):=\sup\{|v_{\lambda}(x)-u_{0}(x)|,\ v_{\lambda}\in S_{\lambda}\setminus(S_{\lambda}^{-}\cup S_{\lambda}^{+}),\ x\in M\}.

We claim that limλ0θ(λ)=0\lim\limits_{\lambda\to 0}\theta(\lambda)=0. Otherwise, there is ε>0\varepsilon>0, a sequence of discount factors λn0+\lambda_{n}\to 0^{+} and of solutions vλnSλn(SλnSλn+)v_{\lambda_{n}}\in S_{\lambda_{n}}\setminus(S^{-}_{\lambda_{n}}\cup S^{+}_{\lambda_{n}}) and ynMy_{n}\in M such that

|vλn(yn)u0(yn)|>ε,n.|v_{\lambda_{n}}(y_{n})-u_{0}(y_{n})|>\varepsilon,\quad\forall n. (2.11)

Since vλnSλn(SλnSλn+)v_{\lambda_{n}}\in S_{\lambda_{n}}\setminus(S^{-}_{\lambda_{n}}\cup S^{+}_{\lambda_{n}}), there is also a point xnx_{n} such that

|vλn(xn)u0(xn)|1.|v_{\lambda_{n}}(x_{n})-u_{0}(x_{n})|\leqslant 1.

Similarly to the first step, we infer from Proposition 2.16 that vλnv_{\lambda_{n}} is uniformly bounded. Then, according to Theorem 2.2 and Remark 2.4, the sequence vλnv_{\lambda_{n}} uniformly converges to the only possible limit u0u_{0}. But again this yields a contradiction as vλnu0>ε\|v_{\lambda_{n}}-u_{0}\|_{\infty}>\varepsilon for all nn\in\mathbb{N}. This concludes the proof. ∎

3. The linear case

In this section we continue our analysis on the vanishing discount problem in the case when GG depends linearly on uu. Hence we will consider a Hamilton-Jacobi equation with discount factor λ>0\lambda>0 of the form

λa(x)u(x)+H(x,Dxu)=c0in M,\lambda a(x)u(x)+H(x,D_{x}u)=c_{0}\qquad\hbox{in $M$}, (HJλ{\textrm{HJ}}_{\lambda})

where we assume that H:TMH:T^{*}M\to\mathbb{R} is a continuous function satisfying (H1)-(H2) (convexity and superlinearity) and the coefficient aa is a continuous function on MM satisfying the following condition:

TMa(x)𝑑μ~>0for all μ~𝔐~,\int_{TM}a(x)d\tilde{\mu}>0\qquad\hbox{for all $\tilde{\mu}\in\widetilde{\mathfrak{M}}$}, (a1a1)

where 𝔐~\widetilde{\mathfrak{M}} denotes the compact and convex subset of 𝒫(TM)\mathscr{P}(TM) made up by Mather measures associated with HH. Without any loss of generality, we shall restrict to the case λ(0,1)\lambda\in(0,1). Equations of the form (HJλ{\textrm{HJ}}_{\lambda}) can be regarded as a model example for the more general Hamilton-Jacobi equations of contact type considered in Section 2, cf. equation (Eλ). The full Hamiltonian is then given by G(x,p,u)=a(x)u+H(x,p)G(x,p,u)=a(x)u+H(x,p). The Hamiltonian GG then fulfills all the hypotheses of the previous section. In particular, it is Lipschitz with respect to uu with Lipschitz constant a\|a\|_{\infty}.

We will be concerned with the issue of existence of solutions to equation (HJλ{\textrm{HJ}}_{\lambda}), at least for small values of λ(0,1)\lambda\in(0,1), and their asymptotic behavior as λ0+\lambda\to 0^{+}. In Section 3.1 we show the existence of the maximal solution to (HJλ{\textrm{HJ}}_{\lambda}) for values of λ(0,1)\lambda\in(0,1) small enough. These solutions are shown to be equi-bounded and equi-Lispchitz, therefore, in view of the results established in Section 2, they converge to a solution of the limit critical equation

H(x,Du)=c0in M.H(x,Du)=c_{0}\quad\hbox{in $M$.} (HJ0)

In Section 3.2, under further natural conditions on the coefficient aa, we investigate the existence of possible other families of solutions of (HJλ{\textrm{HJ}}_{\lambda}) and we describe their asymptotic behavior as λ0+\lambda\to 0^{+}.

3.1. A converging family of solutions

We recall that L:TML:TM\to\mathbb{R} is the Lagrangian associated with HH via the Legendre transform, namely

L(x,v):=suppTxM(p(v)H(x,p))for all (x,v)TM.L(x,v):=\sup_{p\in T^{*}_{x}M}\big{(}p(v)-H(x,p)\big{)}\qquad\hbox{for all $(x,v)\in TM$}.

The constant c0c_{0}\in\mathbb{R} is the critical value of HH. We define the following value function, for xMx\in M,

Vλ(x):=infγlim supt+t0eλs0a(γ(τ))𝑑τ(L(γ(s),γ˙(s))+c0)𝑑s,V_{\lambda}(x):=\inf_{\gamma}\limsup_{t\to+\infty}\int_{-t}^{0}e^{-\lambda\int_{s}^{0}a(\gamma(\tau))d\tau}\Big{(}L\big{(}\gamma(s),\dot{\gamma}(s)\big{)}+c_{0}\Big{)}ds,

where the infimum is taken among absolutely continuous curves γ:(,0]M\gamma:(-\infty,0]\to M with γ(0)=x\gamma(0)=x. This definition is inspired by the formula given in [41, Theorem 4.8] to represent the unique solution of (HJλ{\textrm{HJ}}_{\lambda}) in the case a0a\geqslant 0 in MM and a>0a>0 on 𝒜\mathcal{A}.

The following holds.

Theorem 3.1.

There exists λ0(0,1)\lambda_{0}\in(0,1) such that for every λ(0,λ0)\lambda\in(0,\lambda_{0}) the following holds:

  • (i)

    the value function Vλ:MV_{\lambda}:M\to\mathbb{R} is finite-valued;

  • (ii)

    the functions {Vλ:λ(0,λ0)}\{V_{\lambda}\,:\,\lambda\in(0,\lambda_{0})\,\} are equi-bounded and equi-Lipschitz;

  • (iii)

    the value function VλV_{\lambda} is the maximal subsolution of (HJλ{\textrm{HJ}}_{\lambda}). In particular, it is the maximal solution of (HJλ{\textrm{HJ}}_{\lambda});

  • (iv)

    for every xMx\in M, there exists a curve γλx:(,0]M\gamma^{x}_{\lambda}:(-\infty,0]\to M such that

    Vλ(x)=0eλs0a(γλx(τ))𝑑τ(L(γλx(s),γ˙λx(s))+c0)𝑑s.V_{\lambda}(x)=\int_{-\infty}^{0}e^{-\lambda\int_{s}^{0}a(\gamma^{x}_{\lambda}(\tau))d\tau}\Big{(}L\big{(}\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s)\big{)}+c_{0}\Big{)}ds.

    Furthermore, the curve γλx\gamma^{x}_{\lambda} is κ^\hat{\kappa}-Lipschitz, for some κ^>0\hat{\kappa}>0 independent of λ(0,λ0)\lambda\in(0,\lambda_{0}) and xMx\in M.

Remark 3.2.

For every λ>0\lambda>0, define

cλ:=infuC(M)supxM{H(x,Du)+λa(x)u}.c_{\lambda}:=\inf_{u\in C^{\infty}(M)}\sup_{x\in M}\big{\{}H(x,Du)+\lambda a(x)u\big{\}}.

According to [30, 36], if (HJλ{\textrm{HJ}}_{\lambda}) has solutions, then c0cλc_{0}\geqslant c_{\lambda}. Theorem 3.1 implies that c0cλc_{0}\geqslant c_{\lambda} for every λ>0\lambda>0 small enough if (a1a1) holds. We also point out that the value provided by the above inf-sup with λ=0\lambda=0 agrees with the critical constant c0c_{0}, as it is well known, see for instance [21, 10].

According to Theorem 2.2, the functions VλV_{\lambda} uniformly converge to a critical solution u0u_{0} as λ0+\lambda\to 0^{+}, where u0u_{0} is the maximal subsolution ww of (HJ0) satisfying

TMw(x)a(x)𝑑μ~(x,v)0,μ~𝔐~.\int_{TM}w(x)a(x)d\tilde{\mu}(x,v)\leqslant 0,\quad\forall\tilde{\mu}\in\widetilde{\mathfrak{M}}.

We can furthermore strengthen the conclusion of Theorem 2.5 on the asymptotic behavior, as λ0+\lambda\to 0^{+}, of all possible families of solutions to equation (HJλ{\textrm{HJ}}_{\lambda}). Theorem 3.1 in fact rules out the possibility that there exist families of solutions that diverge to ++\infty. The precise statement is the following.

Theorem 3.3.

There exist ψ:(0,1)[,+)\psi:(0,1)\to\mathbb{[}-\infty,+\infty) and θ:(0,1)\theta:(0,1)\to\mathbb{R} with

limλ0ψ(λ)=,limλ0θ(λ)=0\lim_{\lambda\to 0}\psi(\lambda)=-\infty,\quad\lim_{\lambda\to 0}\theta(\lambda)=0

such that, if vλv_{\lambda} is a solution of (HJλ{\textrm{HJ}}_{\lambda}), then either vλψ(λ)v_{\lambda}\leqslant\psi(\lambda) or vλu0θ(λ)\|v_{\lambda}-u_{0}\|_{\infty}\leqslant\theta(\lambda) for all λ(0,1)\lambda\in(0,1).

The remainder of this subsection is devoted to prove the statement of Theorem 3.1. This will be obtained via a series of intermediate results.

A key step in order to establish that the value function is a (Lispchitz) viscosity subsolution to (HJ0) is to prove that it satisfies the Dynamic Programming Principle. Please note that the next proposition is valid even when VλV_{\lambda} is not finite-valued.

Proposition 3.4.

(Dynamic Programming Principle). Let λ(0,1)\lambda\in(0,1). For each absolutely continuous curve γ:[b1,b2]M\gamma:[{b_{1}},b_{2}]\to M, we have

eλb20a(γ(τ))𝑑τVλ(γ(b2))eλb10a(γ(τ))𝑑τVλ(γ(b1))+b1b2eλs0a(γ(τ))𝑑τ(L(γ(s),γ˙(s))+c0)𝑑s.\displaystyle e^{-\lambda\int_{b_{2}}^{0}a(\gamma(\tau))d\tau}V_{\lambda}\big{(}\gamma(b_{2})\big{)}\leqslant e^{-\lambda\int_{b_{1}}^{0}a(\gamma(\tau))d\tau}V_{\lambda}\big{(}\gamma({b_{1}})\big{)}+\int_{{b_{1}}}^{b_{2}}e^{-\lambda\int_{s}^{0}a(\gamma(\tau))d\tau}\Big{(}L\big{(}\gamma(s),\dot{\gamma}(s)\big{)}+c_{0}\Big{)}ds.
Proof.

We start by claiming that we can reduce to the case b2=0{b_{2}}=0, without any loss of generality. Indeed, by multiplying the above inequality for eλb20a(γ(τ))𝑑τe^{\lambda\int_{b_{2}}^{0}a(\gamma(\tau))d\tau}, we get

Vλ(γ(b2))eλb1b2a(γ(τ))𝑑τVλ(γ(b1))+b1b2eλsb2a(γ(τ))𝑑τ(L(γ(s),γ˙(s))+c0)𝑑s.\displaystyle V_{\lambda}\big{(}\gamma({b_{2}})\big{)}\leqslant e^{-\lambda\int_{b_{1}}^{b_{2}}a(\gamma(\tau))d\tau}V_{\lambda}\big{(}\gamma({b_{1}})\big{)}+\int_{{b_{1}}}^{b_{2}}e^{-\lambda\int_{s}^{b_{2}}a(\gamma(\tau))d\tau}\Big{(}L\big{(}\gamma(s),\dot{\gamma}(s)\big{)}+c_{0}\Big{)}ds.

The claim follows by making the change of variables τ:=τb2\tau^{\prime}:=\tau-{b_{2}}, s:=sb2s^{\prime}:=s-{b_{2}} and by replacing γ\gamma with γb2():=γ(+b2)\gamma_{-{b_{2}}}(\cdot):=\gamma(\cdot+{b_{2}}).

Let us then prove the assertion with b2=0{b_{2}}=0. For each ξ:(,0]M\xi:(-\infty,0]\to M with ξ(0)=γ(b1)\xi(0)=\gamma({b_{1}}), we define

γ¯(t):={γ(t)if t[b1,0]ξ(tb1)if t(,b1)\bar{\gamma}(t):=\begin{cases}\gamma(t)&\hbox{if $t\in[{b_{1}},0]$}\\ \xi(t-{b_{1}})&\hbox{if $t\in(-\infty,{b_{1}})$}\end{cases}

Then, by definition of the value function, we have

Vλ(γ(0))\displaystyle V_{\lambda}\big{(}\gamma(0)\big{)} lim supt+t0eλs0a(γ¯(τ))𝑑τ(L(γ¯(s),γ¯˙(s))+c0)𝑑s\displaystyle\leqslant\limsup_{t\to+\infty}\int_{-t}^{0}e^{-\lambda\int_{s}^{0}{a}(\bar{\gamma}(\tau))d\tau}\Big{(}L\big{(}\bar{\gamma}(s),\dot{\bar{\gamma}}(s)\big{)}+c_{0}\Big{)}ds
=b10eλs0a(γ(τ))𝑑τ(L(γ(s),γ˙(s))+c0)𝑑s\displaystyle=\int_{{b_{1}}}^{0}e^{-\lambda\int_{s}^{0}a(\gamma(\tau))d\tau}\Big{(}L\big{(}\gamma(s),\dot{\gamma}(s)\big{)}+c_{0}\Big{)}ds
+eλb10a(γ(τ))𝑑τlim supt+t0eλs0a(ξ(τ))𝑑τ(L(ξ(s),ξ˙(s))+c0)𝑑s.\displaystyle\quad+e^{-\lambda\int_{b_{1}}^{0}a(\gamma(\tau))d\tau}\limsup_{t\to+\infty}\int_{-t}^{0}e^{-\lambda\int_{s}^{0}a(\xi(\tau))d\tau}\Big{(}L\big{(}\xi(s),\dot{\xi}(s)\big{)}+c_{0}\Big{)}ds.

Taking the infimum among all ξ\xi, we get the assertion. ∎

We now proceed to show VλV_{\lambda} is a bounded function on MM, at least for λ(0,1)\lambda\in(0,1) small enough. We start by proving the following upper bound.

Lemma 3.5.

There is a constant C^v+>0\widehat{C}_{v}^{+}>0 such that Vλ(x)C^v+/λV_{\lambda}(x)\leqslant\widehat{C}_{v}^{+}/\lambda for all xMx\in M and λ(0,1)\lambda\in(0,1).

Proof.

By condition (a1a1), there is a point x0Mx_{0}\in M such that a(x0)>0a(x_{0})>0. We denote d:=d(x0,x)d:=d(x_{0},x), D¯:=diam(M)\overline{D}:=\textrm{diam}(M). We take a geodesic ζ:[d,0]M\zeta:[-d,0]\to M with ζ(d)=x0\zeta(-d)=x_{0} and ζ(0)=x\zeta(0)=x. Define

γ¯(t)={ζ(t),dt0,x0,t<d\bar{\gamma}(t)=\begin{cases}\zeta(t),&-d\leqslant t\leqslant 0,\\ x_{0},&t<-d\\ \end{cases}

and set CL:=maxxM,vx1|L(x,v)+c0|C_{L}:=\max\limits_{x\in M,\|v\|_{x}\leqslant 1}|L(x,v)+c_{0}|. Then

Vλ(x)\displaystyle V_{\lambda}(x) lim supt+t0eλs0a(γ¯(τ))𝑑τ(L(γ¯(s),γ¯˙(s))+c0)𝑑s\displaystyle\leqslant\limsup_{t\to+\infty}\int_{-t}^{0}e^{-\lambda\int_{s}^{0}a(\bar{\gamma}(\tau))d\tau}\Big{(}L\big{(}\bar{\gamma}(s),\dot{\bar{\gamma}}(s)\big{)}+c_{0}\Big{)}ds
=d0eλs0a(ζ(τ))𝑑τ(L(ζ(s),ζ˙(s))+c0)𝑑s\displaystyle=\int_{-d}^{0}e^{-\lambda\int_{s}^{0}a(\zeta(\tau))d\tau}\Big{(}L\big{(}\zeta(s),\dot{\zeta}(s)\big{)}+c_{0}\Big{)}ds
+eλd0a(ζ(τ))𝑑τlim supt+tdeλsda(x0)𝑑τ(L(x0,0)+c0)𝑑s\displaystyle\quad+e^{-\lambda\int_{-d}^{0}a(\zeta(\tau))d\tau}\limsup_{t\to+\infty}\int_{-t}^{-d}e^{-\lambda\int_{s}^{-d}a(x_{0})d\tau}\big{(}L(x_{0},0)+c_{0}\big{)}ds
CL(d0eλas𝑑s+eλaddeλa(x0)(d+s)𝑑s)\displaystyle\leqslant C_{L}\left(\int_{-d}^{0}e^{-\lambda\|a\|_{\infty}s}ds+e^{\lambda\|a\|_{\infty}d}\int_{-\infty}^{-d}e^{\lambda a(x_{0})(d+s)}ds\right)
=CL(eλad1λa+eλadλa(x0))2CLeaD¯λa(x0).\displaystyle=C_{L}\left(\frac{e^{\lambda\|a\|_{\infty}d}-1}{\lambda\|a\|_{\infty}}+\frac{e^{\lambda\|a\|_{\infty}d}}{\lambda a(x_{0})}\right)\leqslant\frac{2C_{L}\,e^{\|a\|_{\infty}\overline{D}}}{\lambda a(x_{0})}.\qed

Now we know that the infimum in Vλ(x)V_{\lambda}(x) is taken among curves in

Γλ:={γ:(,0]M:lim supt+t0eλs0a(γ(τ))𝑑τ(L(γ(s),γ˙(s))+c0)𝑑sC^v+λ}.\displaystyle\quad\Gamma_{\lambda}:=\bigg{\{}\gamma:(-\infty,0]\to M:\ \limsup_{t\to+\infty}\int_{-t}^{0}e^{-\lambda\int_{s}^{0}a(\gamma(\tau))d\tau}\Big{(}L\big{(}\gamma(s),\dot{\gamma}(s)\big{)}+c_{0}\Big{)}ds\leqslant\frac{\widehat{C}_{v}^{+}}{\lambda}\bigg{\}}.

For γ:(,0]M\gamma:(-\infty,0]\to M and t>0t>0, we define

α(λ,γ,t):=t0eλs0a(γ(τ))𝑑τ𝑑s\alpha(\lambda,\gamma,t):=\int_{-t}^{0}e^{-\lambda\int_{s}^{0}a(\gamma(\tau))d\tau}ds

and

β(λ,γ,t):=eλt0a(γ(s))𝑑s.\beta(\lambda,\gamma,t):=e^{-\lambda\int_{-t}^{0}a(\gamma(s))ds}.

Since ddtα(λ,γ,t)=β(λ,γ,t)>0\frac{d}{dt}\alpha(\lambda,\gamma,t)=\beta(\lambda,\gamma,t)>0, tα(λ,γ,t)t\mapsto\alpha(\lambda,\gamma,t) is increasing. Then limt+α(λ,γ,t)\lim\limits_{t\to+\infty}\alpha(\lambda,\gamma,t) exists for each λ\lambda and γ\gamma, and may equal ++\infty. We will also need the following auxiliary remark.

Lemma 3.6.

For every λ>0\lambda>0 and for all curves γ\gamma, we have

λa+1α(λ,γ,t)β(λ,γ,t)α(λ,γ,t)λa+1α(λ,γ,t),t>0.-\lambda\|a\|_{\infty}+\frac{1}{\alpha(\lambda,\gamma,t)}\leqslant\frac{\beta(\lambda,\gamma,t)}{\alpha(\lambda,\gamma,t)}\leqslant\lambda\|a\|_{\infty}+\frac{1}{\alpha(\lambda,\gamma,t)},\quad\forall t>0.
Proof.

A direct calculation shows that

ddtβ(λ,γ,t)ddtα(λ,γ,t)=λa(γ(t)).\frac{\frac{d}{dt}\beta(\lambda,\gamma,t)}{\frac{d}{dt}\alpha(\lambda,\gamma,t)}=-\lambda a\big{(}\gamma(-t)\big{)}.

By the Cauchy mean value theorem, there is ξ(0,t)\xi\in(0,t) such that

β(λ,γ,t)β(λ,γ,0)α(λ,γ,t)α(λ,γ,0)=β(λ,γ,t)1α(λ,γ,t)=ddtβ(λ,γ,ξ)ddtα(λ,γ,ξ)=λa(γ(ξ)),\frac{\beta(\lambda,\gamma,t)-\beta(\lambda,\gamma,0)}{\alpha(\lambda,\gamma,t)-\alpha(\lambda,\gamma,0)}=\frac{\beta(\lambda,\gamma,t)-1}{\alpha(\lambda,\gamma,t)}=\frac{\frac{d}{dt}\beta(\lambda,\gamma,\xi)}{\frac{d}{dt}\alpha(\lambda,\gamma,\xi)}=-\lambda a\big{(}\gamma(-\xi)\big{)},

which implies the conclusion. ∎

We distill in the next lemma an argument that we will repeatedly use in the sequel.

Lemma 3.7.

Assume there are sequences λn0+\lambda_{n}\to 0^{+}, γn:(,0]M\gamma_{n}:(-\infty,0]\to M, tn(0,+)t_{n}\in(0,+\infty) and θn0+\theta_{n}\to 0^{+} such that

α(λn,γn,tn):=tn0eλns0a(γn(τ))𝑑τ𝑑s+,as n+,\displaystyle\qquad\alpha(\lambda_{n},\gamma_{n},t_{n}):=\int_{-t_{n}}^{0}e^{-\lambda_{n}\int_{s}^{0}a(\gamma_{n}(\tau))d\tau}ds\to+\infty,\qquad\hbox{as $n\to+\infty$,} (3.1)
1α(λn,γn,tn)tn0eλns0a(γn(τ))𝑑τ(L(γn(s),γ˙n(s))+c0)𝑑sθn,n.\displaystyle\qquad\frac{1}{\alpha(\lambda_{n},\gamma_{n},t_{n})}\int_{-t_{n}}^{0}e^{-\lambda_{n}\int_{s}^{0}a(\gamma_{n}(\tau))d\tau}\Big{(}L\big{(}\gamma_{n}(s),\dot{\gamma}_{n}(s)\big{)}+c_{0}\Big{)}ds\leqslant\theta_{n},\quad\forall n\in\mathbb{N}. (3.2)

Define a probability measure μ~n𝒫(TM)\tilde{\mu}_{n}\in\mathscr{P}(TM) by

TMf𝑑μ~n:=tn0eλns0a(γn(τ))𝑑τf(γn(s),γ˙n(s))𝑑sα(λn,γn,tn),fC(TM).\int_{TM}f\,d\tilde{\mu}_{n}:=\frac{\int_{-t_{n}}^{0}e^{-\lambda_{n}\int_{s}^{0}a(\gamma_{n}(\tau))d\tau}f\big{(}\gamma_{n}(s),\dot{\gamma}_{n}(s)\big{)}ds}{\alpha(\lambda_{n},\gamma_{n},t_{n})},\qquad\forall f\in C_{\ell}(TM). (\star)

Then the set (μ~n)n(\tilde{\mu}_{n})_{n} is relatively compact in 𝒫\mathscr{P}_{\ell} (for the weak-* topology coming from C(TM)C_{\ell}(TM)) and any of its accumulation points is a Mather measure associated with LL.

Proof.

Note that, due to (3.1), we have tn+t_{n}\to+\infty as n+n\to+\infty. Since LL is uniformly superlinear in the fibers, there exists a constant C10C_{1}\geqslant 0 such that

maxnθnTM(L(x,v)+c0)𝑑μ~n(x,v)TM(vxC1)𝑑μ~nfor all n,\max_{n\in\mathbb{N}}\,\theta_{n}\geqslant\int_{TM}\big{(}L(x,v)+c_{0}\big{)}\,d\tilde{\mu}_{n}(x,v)\geqslant\int_{TM}(\|v\|_{x}-C_{1})d\tilde{\mu}_{n}\qquad\hbox{for all $n\in\mathbb{N}$,}

which readily implies that the sequence (μ~n)n(\tilde{\mu}_{n})_{n} is a well defined sequence in 𝒫\mathscr{P}_{\ell}. The asserted precompactness of (μ~n)n(\tilde{\mu}_{n})_{n} in 𝒫(TM)\mathscr{P}(TM) follows from Lemma 1.11.

Let μ~\tilde{\mu} be a limit point of a subsequence of (μ~n)n(\tilde{\mu}_{n})_{n}, that we will not relabel to ease notations. We are going to show that μ~\tilde{\mu} is a Mather measure.

The measure μ~\tilde{\mu} is closed: we pick ϕC1(M)\phi\in C^{1}(M). An integration by parts shows that

TMDxϕ(v)𝑑μ~n(x,v)\displaystyle\int_{TM}D_{x}\phi(v)\,d\tilde{\mu}_{n}(x,v)
=tn0dϕdt(γn(s))eλns0a(γn(τ))𝑑τ𝑑sα(λn,γn,tn)\displaystyle=\frac{\int_{-t_{n}}^{0}\frac{d\phi}{dt}\big{(}\gamma_{n}(s)\big{)}\,e^{-\lambda_{n}\int_{s}^{0}a(\gamma_{n}(\tau))d\tau}\,ds}{\alpha(\lambda_{n},\gamma_{n},t_{n})} (3.3)
=ϕ(γn(0))eλntn0a(γn(τ))𝑑τϕ(γn(tn))α(λn,γn,tn)λnTMa(x)ϕ(x)𝑑μ~n(x,v).\displaystyle=\frac{\phi\big{(}\gamma_{n}(0)\big{)}-e^{-\lambda_{n}\int_{-t_{n}}^{0}a(\gamma_{n}(\tau))d\tau}\phi\big{(}\gamma_{n}(-t_{n})\big{)}}{\alpha(\lambda_{n},\gamma_{n},t_{n})}-\lambda_{n}\int_{TM}a(x)\phi(x)\,d\tilde{\mu}_{n}(x,v).

We infer

|TMDxϕ(v)𝑑μ~n|\displaystyle\bigg{|}\int_{TM}D_{x}\phi(v)d\tilde{\mu}_{n}\bigg{|} ϕα(λn,γn,tn)+eλntn0a(γn(τ))𝑑τα(λn,γn,tn)ϕ+λnaϕ\displaystyle\leqslant\frac{\|\phi\|_{\infty}}{\alpha(\lambda_{n},\gamma_{n},t_{n})}+\frac{e^{-\lambda_{n}\int_{-t_{n}}^{0}a(\gamma_{n}(\tau))d\tau}}{\alpha(\lambda_{n},\gamma_{n},t_{n})}\|\phi\|_{\infty}+\lambda_{n}\|a\|_{\infty}\|\phi\|_{\infty}
=(1α(λn,γn,tn)+β(λn,γn,tn)α(λn,γn,tn)+λna)ϕ.\displaystyle=\bigg{(}\frac{1}{\alpha(\lambda_{n},\gamma_{n},t_{n})}+\frac{\beta(\lambda_{n},\gamma_{n},t_{n})}{\alpha(\lambda_{n},\gamma_{n},t_{n})}+\lambda_{n}\|a\|_{\infty}\bigg{)}\|\phi\|_{\infty}.

By sending n+n\to+\infty and by using Lemma 3.6, hypothesis (3.1) and again Lemma 1.11, we derive

TMDxϕ(v)𝑑μ~(x,v)=limn+TMDxϕ(v)𝑑μ~n(x,v)=0.\int_{TM}D_{x}\phi(v)\,d\tilde{\mu}(x,v)=\lim_{n\to+\infty}\int_{TM}D_{x}\phi(v)\,d\tilde{\mu}_{n}(x,v)=0.

The measure μ~\tilde{\mu} is minimizing: by (3.2) we get

0=limn+θnlim infn+TM(L(x,v)+c0)𝑑μ~n(x,v)TM(L(x,v)+c0)𝑑μ~(x,v),0=\lim_{n\to+\infty}\theta_{n}\geqslant\liminf_{n\to+\infty}\int_{TM}\big{(}L(x,v)+c_{0}\big{)}\,d\tilde{\mu}_{n}(x,v)\geqslant\int_{TM}\big{(}L(x,v)+c_{0}\big{)}\,d\tilde{\mu}(x,v),

where the last inequality follows from the fact that LL is continuous and bounded from below and the measures μ~n\tilde{\mu}_{n} are weakly-* converging to μ~\tilde{\mu} in 𝒫\mathscr{P}_{\ell} (hence, narrowly converging), see for instance [1, Section 5.1.1].

In view of Theorem 1.8, we conclude that μ~\tilde{\mu} is a Mather measure. ∎

Let C>0C>0 be a fixed constant and define, for any fixed integer p0p\geqslant 0,

Γλ(p):={γ:(,0]M:lim supt+t0eλs0a(γ(τ))𝑑τ(L(γ(s),γ˙(s))+c0)𝑑sCλp}.\displaystyle\quad\Gamma_{\lambda}(p):=\bigg{\{}\gamma:(-\infty,0]\to M:\ \limsup_{t\to+\infty}\int_{-t}^{0}e^{-\lambda\int_{s}^{0}a(\gamma(\tau))d\tau}\big{(}L\big{(}\gamma(s),\dot{\gamma}(s)\big{)}+c_{0}\big{)}ds\leqslant\frac{C}{\lambda^{p}}\bigg{\}}.

The information provided by the next lemma will be crucial for our upcoming analysis.

Lemma 3.8.

Let p0p\geqslant 0 be a fixed integer and assume that Γλ(p)\Gamma_{\lambda}(p)\not=\varnothing for every λ(0,1)\lambda\in(0,1) small enough. Then there exists λ(p)(0,1)\lambda(p)\in(0,1) such that

A(p):=sup{λp+1α(λ,γ,t):λ(0,λ(p)),γΓλ(p),t>0}<+.\displaystyle A(p):=\sup\Big{\{}\lambda^{p+1}\alpha(\lambda,\gamma,t)\,:\,\lambda\in\big{(}0,\lambda(p)\big{)},\ \gamma\in\Gamma_{\lambda}(p),\ t>0\Big{\}}<+\infty.
Remark 3.9.

Note that Γλ(p+1)Γλ(p)\Gamma_{\lambda}(p+1)\supseteq\Gamma_{\lambda}(p) for every integer p0p\geqslant 0 and λ(0,1)\lambda\in(0,1). In particular, when p1p\geqslant 1 and C:=C^v+C:=\widehat{C}_{v}^{+}, we have Γλ(p)\Gamma_{\lambda}(p)\not=\varnothing for every λ(0,1)\lambda\in(0,1) in view of Lemma 3.5.

Proof.

Let p0p\geqslant 0 be a fixed integer. We argue by contradiction. Assume there exist sequences λn0+\lambda_{n}\to 0^{+}, γnΓλn(p)\gamma_{n}\in\Gamma_{\lambda_{n}}(p) and tn+t_{n}\to+\infty such that

λnp+1α(λn,γn,tn)+as n+,\lambda^{p+1}_{n}\alpha(\lambda_{n},\gamma_{n},t_{n})\to+\infty\qquad\hbox{as $n\to+\infty$,}

with

tn0eλns0a(γn(τ))𝑑τ(L(γn(s),γ˙n(s))+c0)𝑑sCλnp+1.\int_{-t_{n}}^{0}e^{-\lambda_{n}\int_{s}^{0}a(\gamma_{n}(\tau))d\tau}\Big{(}L\big{(}\gamma_{n}(s),\dot{\gamma}_{n}(s)\big{)}+c_{0}\Big{)}ds\leqslant\frac{C}{\lambda_{n}^{p}}+1.

In particular, conditions (3.1) and (3.2) in Lemma 3.7 are satisfied with

θn:=C/λnp+1α(λn,γn,tn).\theta_{n}:=\dfrac{C/\lambda_{n}^{p}+1}{\alpha(\lambda_{n},\gamma_{n},t_{n})}.

Let μ~n\tilde{\mu}_{n} be the probability measure defined in (\star3.7). According to Lemma 3.7, up to extraction of a subsequence (not relabeled), the measures μ~n\tilde{\mu}_{n} weakly converge to a Mather measure μ~\tilde{\mu}.

Now we choose a subsolution φ1\varphi\leqslant-1 of (HJ0). For every ε>0\varepsilon>0, there exists, in view of Theorem 1.15, a function φεC(M)\varphi_{\varepsilon}\in C^{\infty}(M) satisfying

φεφεandH(x,Dxφε)c0+εfor all xM.\|\varphi_{\varepsilon}-\varphi\|_{\infty}\leqslant\varepsilon\qquad\hbox{and}\qquad H\big{(}x,D_{x}\varphi_{\varepsilon}\big{)}\leqslant c_{0}+\varepsilon\quad\hbox{for all $x\in M$}. (3.4)

For ε>0\varepsilon>0 small, we have φε<0\varphi_{\varepsilon}<0. We get

θn\displaystyle\quad\ \theta_{n} TM(L(x,v)+c0)𝑑μ~n(x,v)\displaystyle\geqslant\int_{TM}\big{(}L(x,v)+c_{0}\big{)}\,d\tilde{\mu}_{n}(x,v)
TM(Dxφε(v)H(x,Dxφε)+c0)𝑑μ~n(x,v)TM(Dxφε(v)ε)𝑑μ~n(x,v)\displaystyle\geqslant\int_{TM}\Big{(}D_{x}\varphi_{\varepsilon}(v)-H\big{(}x,D_{x}\varphi_{\varepsilon}\big{)}+c_{0}\Big{)}\,d\tilde{\mu}_{n}(x,v)\geqslant\int_{TM}\big{(}D_{x}\varphi_{\varepsilon}(v)-\varepsilon\big{)}\,d\tilde{\mu}_{n}(x,v)
=φε(γn(0))eλntn0a(γn(τ))𝑑τφε(γn(tn))α(λn,γn,tn)λnTMa(x)φε(x)𝑑μ~n(x,v)ε\displaystyle=\frac{\varphi_{\varepsilon}\big{(}\gamma_{n}(0)\big{)}-e^{-\lambda_{n}\int_{-t_{n}}^{0}a(\gamma_{n}(\tau))d\tau}\varphi_{\varepsilon}\big{(}\gamma_{n}(-t_{n})\big{)}}{\alpha(\lambda_{n},\gamma_{n},t_{n})}-\lambda_{n}\int_{TM}a(x)\varphi_{\varepsilon}(x)\,d\tilde{\mu}_{n}(x,v)-\varepsilon
>φε(γn(0))α(λn,γn,tn)λnTMa(x)φε(x)𝑑μ~n(x,v)ε.\displaystyle>\frac{\varphi_{\varepsilon}\big{(}\gamma_{n}(0)\big{)}}{\alpha(\lambda_{n},\gamma_{n},t_{n})}-\lambda_{n}\int_{TM}a(x)\varphi_{\varepsilon}(x)\,d\tilde{\mu}_{n}(x,v)-\varepsilon.

The equality appearing above is obtained via an integration by parts (cf. proof of Lemma 3.7 equation (3.3), when we prove that μ~\tilde{\mu} is closed), while for the last inequality we have used that fact that φε<0\varphi_{\varepsilon}<0. Now we send ε0+\varepsilon\to 0^{+} and divide the above inequality by λn\lambda_{n}. We get

Cλnp+1α(λn,γn,tn)+1λnα(λn,γn,tn)φ(γn(0))λnα(λn,γn,tn)TMa(x)φ(x)𝑑μ~n.\frac{C}{\lambda_{n}^{p+1}\alpha(\lambda_{n},\gamma_{n},t_{n})}+\frac{1}{\lambda_{n}\alpha(\lambda_{n},\gamma_{n},t_{n})}\geqslant\frac{\varphi\big{(}\gamma_{n}(0)\big{)}}{\lambda_{n}\alpha(\lambda_{n},\gamma_{n},t_{n})}-\int_{TM}a(x)\varphi(x)d\tilde{\mu}_{n}.

Since λnα(λn,γn,tn)λnp+1α(λn,γn,tn)+\lambda_{n}\alpha(\lambda_{n},\gamma_{n},t_{n})\geqslant{\lambda}^{p+1}_{n}\alpha(\lambda_{n},\gamma_{n},t_{n})\to+\infty as n+n\to+\infty, we get

TMa(x)φ(x)𝑑μ~(x,v)0.\int_{TM}a(x)\varphi(x)\,d\tilde{\mu}(x,v)\geqslant 0.

Since for all m0m\geqslant 0 the function φm1\varphi-m\leqslant-1 is also a negative subsolution of (HJ0), by replacing φ\varphi with φm\varphi-m in the inequality above we obtain

φaTMa(x)φ(x)𝑑μ~(x,v)mTMa(x)𝑑μ~(x,v),for all m0,\|\varphi\|_{\infty}\|a\|_{\infty}\geqslant\int_{TM}a(x)\varphi(x)\,d\tilde{\mu}(x,v)\geqslant m\int_{TM}a(x)\,d\tilde{\mu}(x,v),\qquad\hbox{for all $m\geqslant 0$,}

which implies that TMa(x)𝑑μ~0\int_{TM}a(x)d\tilde{\mu}\leqslant 0. This contradicts assumption (a1a1). ∎

We now proceed to show that the value function VλV_{\lambda} is bounded from below, for every fixed λ(0,λ(1))\lambda\in\big{(}0,\lambda(1)\big{)}, where λ(1)>0\lambda(1)>0 is the value obtained according to Lemma 3.7 with p=1p=1 and C=C^v+C=\widehat{C}_{v}^{+}, where C^v+\widehat{C}_{v}^{+} is the constant provided by Lemma 3.5.

Lemma 3.10.

There exists a constant C^v>0\widehat{C}^{-}_{v}>0 independent of λ\lambda such that

Vλ(x)C^vλ2for all xM and λ(0,λ(1)).V_{\lambda}(x)\geqslant-\widehat{C}^{-}_{v}\lambda^{-2}\qquad\hbox{for all $x\in M$ and $\lambda\in\big{(}0,\lambda(1)\big{)}$.}
Proof.

Let us fix γΓλ\gamma\in\Gamma_{\lambda}. For every t>0t>0 we have

t0eλs0a(γ(τ))𝑑τ(L(γ(s),γ˙(s))+c0)𝑑sCLt0eλs0a(γ(τ))𝑑τ𝑑s,\int_{-t}^{0}e^{-\lambda\int_{s}^{0}a(\gamma(\tau))d\tau}\Big{(}L\big{(}\gamma(s),\dot{\gamma}(s)\big{)}+c_{0}\Big{)}ds\geqslant-C^{-}_{L}\int_{-t}^{0}e^{-\lambda\int_{s}^{0}a(\gamma(\tau))d\tau}ds,

where CL:=min{minTM(L(x,v)+c0),0}0C^{-}_{L}:=-\min\big{\{}\min_{TM}\big{(}L(x,v)+c_{0}\big{)},0\big{\}}\geqslant 0. We are now going to apply Lemma 3.8 with p=1p=1 and by choosing C:=C^v+C:=\widehat{C}_{v}^{+} in the definition of Γλ(1)\Gamma_{\lambda}(1), so that Γλ(1)=Γλ\Gamma_{\lambda}(1)=\Gamma_{\lambda}: from the above inequality we infer

t0eλs0a(γ(τ))dτ(L(γ(s),γ˙(s))+c0)dsCLA(1)λ2for all t>0.\int_{-t}^{0}e^{-\lambda\int_{s}^{0}a(\gamma(\tau))d\tau}\Big{(}L\big{(}\gamma(s),\dot{\gamma}(s)\big{)}+c_{0}\Big{)}ds\geqslant-\frac{C^{-}_{L}A(1)}{\lambda^{2}}\qquad\hbox{for all $t>0$}.

The assertion readily follows from the definition of VλV_{\lambda}. ∎

From the information gathered so far, we know that the value function VλV_{\lambda} is finite-valued on MM for every λ(0,λ(1))\lambda\in\big{(}0,\lambda(1)\big{)}. We now proceed to get bounds for VλV_{\lambda} from above and from below on MM independent of λ\lambda.

Lemma 3.11.

There is a constant C¯+v>0\overline{C}^{+}_{v}>0 such that

Vλ(x)C¯+vfor all x𝒜 and λ(0,λ(1)).V_{\lambda}(x)\leqslant\overline{C}^{+}_{v}\qquad\hbox{for all $x\in\mathcal{A}$ and $\lambda\in\big{(}0,\lambda(1)\big{)}$.}
Proof.

According to [13, Theorem 4.14 and Proposition 4.4], see also [16, Theorem 3.3], for every x𝒜x\in\mathcal{A} there exists a curve η:𝒜\eta:\mathbb{R}\to\mathcal{A} with η(0)=x\eta(0)=x such that, for every subsolution φ\varphi of (HJ0),

L(η(s),η˙(s))+c0=dds(φη)(s)for a.e. s.L\big{(}\eta(s),\dot{\eta}(s)\big{)}+c_{0}=\frac{d}{ds}(\varphi\mbox{\scriptsize$\circ$}\eta)(s)\quad\hbox{for a.e.\ $s\in\mathbb{R}$.} (3.5)

Let us denote by 𝒦\mathcal{K} the family of curves η:𝒜\eta:\mathbb{R}\to\mathcal{A} satisfying (3.5). We remark for further use that these curves are equi-Lipschitz, see for instance [13, Lemma 4.9].

Pick a subsolution φ0\varphi\leqslant 0 of (HJ0) and fix x𝒜x\in\mathcal{A}. From the definition of VλV_{\lambda} we get

Vλ(x)\displaystyle V_{\lambda}(x) \displaystyle\leqslant lim supt+t0eλs0a(η(τ))dτ(L(η(s),η˙(s))+c0)ds\displaystyle\limsup_{t\to+\infty}\int_{-t}^{0}e^{-\lambda\int_{s}^{0}a(\eta(\tau))d\tau}\Big{(}L\big{(}\eta(s),\dot{\eta}(s)\big{)}+c_{0}\Big{)}ds
=\displaystyle= lim supt+t0eλs0a(η(τ))dτdds(φη)(s)ds\displaystyle\limsup_{t\to+\infty}\int_{-t}^{0}e^{-\lambda\int_{s}^{0}a(\eta(\tau))\,d\tau}\,\frac{d}{ds}(\varphi\mbox{\scriptsize$\circ$}\eta)(s)\,ds
=\displaystyle= lim supt+{φ(η(0))eλt0a(η(τ))dτφ(η(t))λα(λ,η,t)TMa(x)φ(x)dμ~ηt},\displaystyle\limsup_{t\to+\infty}\bigg{\{}\varphi\big{(}\eta(0)\big{)}-e^{-\lambda\int_{-t}^{0}a(\eta(\tau))d\tau}\varphi\big{(}\eta(-t)\big{)}-\lambda\alpha(\lambda,\eta,t)\int_{TM}a(x)\varphi(x)d\tilde{\mu}^{\eta}_{t}\bigg{\}},

where μ~ηt𝒫(TM)\tilde{\mu}^{\eta}_{t}\in\mathscr{P}(TM) is defined by

TMf(x,v)dμ~ηt(x,v):=t0eλs0a(η(τ))dτf(η(s),η˙(s))dst0eλs0a(η(τ))dτds,fCc(TM).\int_{TM}f(x,v)\,d\tilde{\mu}^{\eta}_{t}(x,v):=\frac{\int_{-t}^{0}e^{-\lambda\int_{s}^{0}a(\eta(\tau))d\tau}f\big{(}\eta(s),\dot{\eta}(s)\big{)}ds}{\int_{-t}^{0}e^{-\lambda\int_{s}^{0}a(\eta(\tau))d\tau}ds},\qquad\forall f\in C_{c}(TM).

The second equality in (3.1) is derived via an integration by parts as for (3.3). By assumption (a1a1) and compactness of the family of Mather measures 𝔐~\widetilde{\mathfrak{M}}, there exists ε>0\varepsilon>0 such that

TMa(x)dμ~(x,v)>εfor all μ~𝔐~.\int_{TM}a(x)\,d\tilde{\mu}(x,v)>\varepsilon\qquad\hbox{for all $\tilde{\mu}\in\widetilde{\mathfrak{M}}$}. (3.7)

We show that there is T0>0T_{0}>0 such that, for all curves η𝒦\eta\in\mathcal{K}, we have

t0a(η(s))ds>εtfor all tT0.\int_{-t}^{0}a\big{(}\eta(s)\big{)}\,ds>\varepsilon t\qquad\hbox{for all $t\geqslant T_{0}$.}

We argue by contradiction. Assume there exist sequences ηn𝒦\eta_{n}\in\mathcal{K} and tn0t_{n}\to 0 such that

tn0a(ηn(s))dsεtn.\int_{-t_{n}}^{0}a\big{(}\eta_{n}(s)\big{)}\,ds\leqslant\varepsilon t_{n}. (3.8)

Define μ~n𝒫(TM)\tilde{\mu}_{n}\in\mathscr{P}(TM) by

TMf(x,v)dμ~n:=1tntn0f(ηn(s),η˙n(s))ds,fCc(TM).\int_{TM}f(x,v)d\tilde{\mu}_{n}:=\frac{1}{t_{n}}\int_{-t_{n}}^{0}f\big{(}\eta_{n}(s),\dot{\eta}_{n}(s)\big{)}\,ds,\quad\forall f\in C_{c}(TM).

Due to the fact that the curves (ηn)n(\eta_{n})_{n} are equi-Lipschitz, the measures (μ~n)n(\tilde{\mu}_{n})_{n} have equi-compact support. In particular, up to extracting a subsequence (not relabeled), they weakly converge to a probability measure μ~𝒫(TM)\tilde{\mu}\in\mathscr{P}(TM). Furthermore

limnTMf(x,v)dμ~n(x,v)=TMf(x,v)dμ~(x,v)fC(TM).\lim_{n}\int_{TM}f(x,v)\,d\tilde{\mu}_{n}(x,v)=\int_{TM}f(x,v)\,d\tilde{\mu}(x,v)\qquad\forall f\in C(TM).

We claim that μ~\tilde{\mu} is closed. Indeed, for every ϕC1(M)\phi\in C^{1}(M) we have

TMDxϕ(v)dμ~(x,v)\displaystyle\int_{TM}D_{x}\phi(v)\,d\tilde{\mu}(x,v) =\displaystyle= limnTMDxϕ(v)dμ~n(x,v)\displaystyle\lim_{n}\int_{TM}D_{x}\phi(v)\,d\tilde{\mu}_{n}(x,v)
=\displaystyle= limn1tntn0dds(ϕηn)(s)dslimn2ϕtn=0.\displaystyle\lim_{n}\frac{1}{t_{n}}\int_{-t_{n}}^{0}\frac{d}{ds}(\phi\mbox{\scriptsize$\circ$}\eta_{n})(s)\,ds\leqslant\lim_{n}\frac{2\|\phi\|_{\infty}}{t_{n}}=0.

We proceed to show that μ~\tilde{\mu} is minimizing, namely, a Mather measure. Pick a subsolution φ\varphi to (HJ0). By exploiting (3.5), we get

TM(L(x,v)+c0)dμ~\displaystyle\int_{TM}\big{(}L(x,v)+c_{0}\big{)}d\tilde{\mu} =limnTM(L(x,v)+c0)dμ~n=limn1tntn0(L(ηn(s),η˙n(s))+c0)ds\displaystyle=\lim_{n}\int_{TM}\big{(}L(x,v)+c_{0}\big{)}d\tilde{\mu}_{n}=\lim_{n}\frac{1}{t_{n}}\int_{-t_{n}}^{0}\!\!\Big{(}L\big{(}\eta_{n}(s),\dot{\eta}_{n}(s)\big{)}+c_{0}\Big{)}ds
=limn1tntn0dds(φηn)(s)ds=limnφ(ηn(0))φ(ηn(tn))tn=0.\displaystyle=\lim_{n}\frac{1}{t_{n}}\int_{-t_{n}}^{0}\frac{d}{ds}(\varphi\mbox{\scriptsize$\circ$}\eta_{n})(s)\,ds=\lim_{n}\frac{\varphi\big{(}\eta_{n}(0)\big{)}-\varphi\big{(}\eta_{n}(-t_{n})\big{)}}{t_{n}}=0.

By (3.8), we also have TMa(x)dμ~(x,v)ε,\int_{TM}a(x)\,d\tilde{\mu}(x,v)\leqslant\varepsilon, which leads to a contradiction with (3.7). Then for t>T0t>T_{0} we have

eλt0a(η(τ))dτeλεte^{-\lambda\int_{-t}^{0}a(\eta(\tau))d\tau}\leqslant e^{-\lambda\varepsilon t}

and

0<λα(λ,η,t)\displaystyle 0<\lambda\alpha(\lambda,\eta,t) =\displaystyle= λt0eλs0a(η(τ))dτds\displaystyle\lambda\int_{-t}^{0}e^{-\lambda\int_{s}^{0}a(\eta(\tau))d\tau}ds
=\displaystyle= λT00eλs0a(η(τ))dτds+λtT0eλs0a(η(τ))dτds\displaystyle\lambda\int_{-T_{0}}^{0}e^{-\lambda\int_{s}^{0}a(\eta(\tau))d\tau}ds+\lambda\int_{-t}^{-T_{0}}e^{-\lambda\int_{s}^{0}a(\eta(\tau))d\tau}ds
\displaystyle\leqslant λT00eλasds+λtT0eλεsds\displaystyle\lambda\int_{-T_{0}}^{0}e^{-\lambda\|a\|_{\infty}s}\,ds+\lambda\int_{-t}^{-T_{0}}e^{\lambda\varepsilon s}ds
\displaystyle\leqslant eλaT01a+eλεT0εeaT0a+1ε=:Cη.\displaystyle\frac{e^{\lambda{\|a\|_{\infty}}T_{0}}-1}{\|a\|_{\infty}}+\frac{e^{-\lambda\varepsilon T_{0}}}{\varepsilon}\leqslant\frac{e^{{\|a\|_{\infty}}T_{0}}}{\|a\|_{\infty}}+\frac{1}{\varepsilon}=:C_{\eta}.

By using these inequalities in (3.1) and recalling that φ0\varphi\leqslant 0, we finally get

Vλ(x)Cηaφ,V_{\lambda}(x)\leqslant C_{\eta}\|a\|_{\infty}\|\varphi\|_{\infty},

which gives the upper bound of VλV_{\lambda} on 𝒜\mathcal{A} independent of λ\lambda. ∎

By exploiting the fact that VλV_{\lambda} satisfies the Dynamic Programming Principle, we show that the partial upper bound obtained in Lemma 3.11 actually entails a uniform upper bound on the whole MM.

Proposition 3.12.

There is Cv+>0C_{v}^{+}>0 independent of λ\lambda such that

Vλ(x)Cv+for all xM and λ(0,λ(1)).V_{\lambda}(x)\leqslant C_{v}^{+}\qquad\hbox{for all $x\in M$ and $\lambda\in\big{(}0,\lambda(1)\big{)}$.}
Proof.

Fix xMx\in M and pick a point y𝒜y\in\mathcal{A}. Set d:=d(x,y)d:=d(x,y), D¯:=\overline{D}:=diam(M)(M) and CL:=maxxM,vx1|L(x,v)+c0|C_{L}:=\max\limits_{x\in M,\|v\|_{x}\leqslant 1}|L(x,v)+c_{0}| . Take a geodesic ζ:[d,0]M\zeta:[-d,0]\to M with ζ(d)=y\zeta(-d)=y, ζ(0)=x\zeta(0)=x and ζ˙ζ=1\|\dot{\zeta}\|_{\zeta}=1. By Proposition 3.4 we have

Vλ(x)\displaystyle V_{\lambda}(x) eλd0a(ζ(τ))dτVλ(y)+d0eλs0a(ζ(τ))dτ(L(ζ(s),ζ˙(s))+c0)ds\displaystyle\leqslant e^{-\lambda\int_{-d}^{0}a(\zeta(\tau))d\tau}V_{\lambda}(y)+\int_{-d}^{0}e^{-\lambda\int_{s}^{0}a(\zeta(\tau))d\tau}\Big{(}L\big{(}\zeta(s),\dot{\zeta}(s)\big{)}+c_{0}\Big{)}ds
eλadC¯v++CLdeλadeaD¯(C¯v++CLD¯),\displaystyle\leqslant e^{\lambda\|a\|_{\infty}d}\overline{C}_{v}^{+}+C_{L}de^{\lambda\|a\|_{\infty}d}\leqslant e^{\|a\|_{\infty}\overline{D}}(\overline{C}_{v}^{+}+C_{L}\overline{D}),

which gives the sought uniform upper bound of VλV_{\lambda}. ∎

Now that we know that VλV_{\lambda} is uniformly bounded from above, we can prove that VλV_{\lambda} is also uniformly bounded from below.

Proposition 3.13.

There exist λ¯(0,λ(1))\overline{\lambda}\in\big{(}0,\lambda(1)\big{)} and a constant Cv>0C^{-}_{v}>0 independent of λ\lambda such that

Vλ(x)Cvfor all xM and λ(0,λ¯).V_{\lambda}(x)\geqslant-C_{v}^{-}\qquad\hbox{for all $x\in M$ and $\lambda\in(0,\overline{\lambda})$.}

In particular, |Vλ(x)|Cv:=max{Cv,Cv+}|V_{\lambda}(x)|\leqslant C_{v}:=\max\{C_{v}^{-},C_{v}^{+}\} for all xMx\in M and λ(0,λ¯)\lambda\in(0,\overline{\lambda}).

Proof.

By Proposition 3.12, we know that, for λ<λ(1)\lambda<\lambda(1), the infimum in VλV_{\lambda} is taken among the set

Γ:={γ:(,0]M:lim supt+t0eλs0a(γ(τ))dτ(L(γ(s),γ˙(s))+c0)dsCv+}.\displaystyle\Gamma:=\bigg{\{}\gamma:(-\infty,0]\to M:\ \limsup_{t\to+\infty}\int_{-t}^{0}e^{-\lambda\int_{s}^{0}a(\gamma(\tau))d\tau}\Big{(}L\big{(}\gamma(s),\dot{\gamma}(s)\big{)}+c_{0}\Big{)}ds\leqslant{C}_{v}^{+}\bigg{\}}.

By Lemma 3.8, there exists λ(0)(0,1)\lambda(0)\in(0,1) (depending on C+vC^{+}_{v}) such that

A(0):=sup{λα(λ,γ,t):λ(0,λ(0)),γΓ,t>0}<+.A(0):=\sup\bigg{\{}\lambda\alpha(\lambda,\gamma,t):\ \lambda\in\big{(}0,\lambda(0)\big{)},\ \gamma\in\Gamma,\ t>0\,\bigg{\}}<+\infty.

Let us set λ¯:=min{λ(0),λ(1)}\overline{\lambda}:=\min\{\lambda(0),\lambda(1)\}.777We recall that λ(1)(0,1)\lambda(1)\in(0,1) is the value obtained according to Lemma 3.7 with p=1p=1 and C=C^v+C=\widehat{C}_{v}^{+}, where C^v+\widehat{C}_{v}^{+} is the constant provided by Lemma 3.5. Let φ1\varphi\leqslant-1 be a subsolution of (HJ0). For each ε>0\varepsilon>0, we can take φεC(M)\varphi_{\varepsilon}\in C^{\infty}(M) satisfying φεφε\|\varphi_{\varepsilon}-\varphi\|_{\infty}\leqslant\varepsilon and H(x,Dxφε)c0+εH\big{(}x,D_{x}\varphi_{\varepsilon}\big{)}\leqslant c_{0}+\varepsilon for all xMx\in M, given by Theorem 1.15. For ε>0\varepsilon>0 small, we have φε<0\varphi_{\varepsilon}<0. For each γΓ\gamma\in\Gamma, we have

t0eλs0a(γ(τ))dτ(L(γ(s),γ˙(s))+c0)ds\displaystyle\int_{-t}^{0}e^{-\lambda\int_{s}^{0}a(\gamma(\tau))d\tau}\Big{(}L\big{(}\gamma(s),\dot{\gamma}(s)\big{)}+c_{0}\Big{)}ds
t0eλs0a(γ(τ))dτ(Dγ(s)φε(γ˙(s))H(γ(s),Dγ(s)φε)+c0)ds\displaystyle\geqslant\int_{-t}^{0}e^{-\lambda\int_{s}^{0}a(\gamma(\tau))d\tau}\Big{(}D_{\gamma(s)}\varphi_{\varepsilon}\big{(}\dot{\gamma}(s)\big{)}-H\big{(}\gamma(s),D_{\gamma(s)}\varphi_{\varepsilon}\big{)}+c_{0}\Big{)}ds
t0eλs0a(γ(τ))dτ(dφεdt(γ(s))ε)ds\displaystyle\geqslant\int_{-t}^{0}e^{-\lambda\int_{s}^{0}a(\gamma(\tau))d\tau}\bigg{(}\frac{d\varphi_{\varepsilon}}{dt}\big{(}\gamma(s)\big{)}-\varepsilon\bigg{)}ds
=φε(γ(0))eλt0a(γ(τ))dτφε(γ(t))α(λ,γ,t)TM(λa(x)φε(x)+ε)dμ~γt\displaystyle=\varphi_{\varepsilon}\big{(}\gamma(0)\big{)}-e^{-\lambda\int_{-t}^{0}a(\gamma(\tau))d\tau}\varphi_{\varepsilon}\big{(}\gamma(-t)\big{)}-\alpha(\lambda,\gamma,t)\int_{TM}\big{(}\lambda a(x)\varphi_{\varepsilon}(x)+\varepsilon\big{)}d\tilde{\mu}^{\gamma}_{t}
φε(γ(0))α(λ,γ,t)TM(λa(x)φε(x)+ε)dμ~γt,\displaystyle\geqslant\varphi_{\varepsilon}\big{(}\gamma(0)\big{)}-\alpha(\lambda,\gamma,t)\int_{TM}\big{(}\lambda a(x)\varphi_{\varepsilon}(x)+\varepsilon\big{)}d\tilde{\mu}^{\gamma}_{t},

where

TMf(x,v)dμ~γt(x,v):=t0eλs0a(γ(τ))dτf(γ(s),γ˙(s))dsα(λ,γ,t),fCc(TM).\int_{TM}f(x,v)d\tilde{\mu}^{\gamma}_{t}(x,v):=\frac{\int_{-t}^{0}e^{-\lambda\int_{s}^{0}a(\gamma(\tau))d\tau}f\big{(}\gamma(s),\dot{\gamma}(s)\big{)}ds}{\alpha(\lambda,\gamma,t)},\qquad\forall f\in C_{c}(TM).

We recall that α(λ,γ,t):=t0eλs0a(γ(τ))dτds\alpha(\lambda,\gamma,t):=\int_{-t}^{0}e^{-\lambda\int_{s}^{0}a(\gamma(\tau))d\tau}ds. Now let ε0+\varepsilon\to 0^{+} to get

t0eλs0a(γ(τ))dτ(L(γ(s),γ˙(s))+c0)dsφ(γ(0))λα(λ,γ,t)TMa(x)φ(x)dμ~γt.\int_{-t}^{0}e^{-\lambda\int_{s}^{0}a(\gamma(\tau))d\tau}\Big{(}L\big{(}\gamma(s),\dot{\gamma}(s)\big{)}+c_{0}\Big{)}ds\geqslant\varphi\big{(}\gamma(0)\big{)}-\lambda\alpha(\lambda,\gamma,t)\int_{TM}a(x)\varphi(x)d\tilde{\mu}^{\gamma}_{t}.

Sending t+t\to+\infty we get

lim supt+t0eλs0a(γ(τ))dτ(L(γ(s),γ˙(s))+c0)ds(1+A(0)a)φ.\limsup_{t\to+\infty}\int_{-t}^{0}e^{-\lambda\int_{s}^{0}a(\gamma(\tau))d\tau}\Big{(}L\big{(}\gamma(s),\dot{\gamma}(s)\big{)}+c_{0}\Big{)}ds\geqslant-\big{(}1+A(0)\|a\|_{\infty}\big{)}\|\varphi\|_{\infty}.

The bound from below readily follows from this by definition of VλV_{\lambda}. The last assertion is a consequence of Proposition 3.12 and of what we have just shown above. ∎

We now proceed to show that the value function is Lipschitz continuous. This is indeed a consequence of this more general result.

Proposition 3.14.

Let w:Mw:M\to\mathbb{R} be a bounded function. Assume that ww satisfies the Dynamic Programming Principle, i.e., for each absolutely continuous curve γ:[t,0]M\gamma:[-t,0]\to M, we have

w(γ(0))eλt0a(γ(τ))dτw(γ(t))t0eλs0a(γ(τ))dτ(L(γ(s),γ˙(s))+c0)ds.\displaystyle\ w\big{(}\gamma(0)\big{)}-e^{-\lambda\int_{-t}^{0}a(\gamma(\tau))d\tau}w\big{(}\gamma(-t)\big{)}\leqslant\int_{-t}^{0}e^{-\lambda\int_{s}^{0}a(\gamma(\tau))d\tau}\Big{(}L\big{(}\gamma(s),\dot{\gamma}(s)\big{)}+c_{0}\Big{)}ds. (DPP)

Then ww is Lipschitz continuous, with a Lipschitz constant that only depends on L,diam(M)L,\,\mbox{\rm diam}(M) and w\|w\|_{\infty}.

Proof.

Pick x,yMx,y\in M. Let ζ:[d,0]M\zeta:[-d,0]\to M be a geodesic with ζ(d)=y\zeta(-d)=y and ζ(0)=x\zeta(0)=x, where d:=d(x,y)d:=d(x,y). By (DPP) we have

w(x)w(y)\displaystyle w(x)-w(y) w(y)(1eλd0a(ζ(τ))dτ)+d0eλs0a(ζ(τ))dτ(L(ζ(s),ζ˙(s))+c0)ds.\displaystyle\leqslant-w(y)\bigg{(}1-e^{-\lambda\int_{-d}^{0}a(\zeta(\tau))d\tau}\bigg{)}+\int_{-d}^{0}e^{-\lambda\int_{s}^{0}a(\zeta(\tau))d\tau}\big{(}L\big{(}\zeta(s),\dot{\zeta}(s)\big{)}+c_{0}\Big{)}ds.

There is τ0(d,0)\tau_{0}\in(-d,0) such that d0a(ζ(τ))dτ=a(ζ(τ0))d\int_{-d}^{0}a\big{(}\zeta(\tau)\big{)}d\tau=a\big{(}\zeta(\tau_{0})\big{)}d. If a(ζ(τ0))0a\big{(}\zeta(\tau_{0})\big{)}\geqslant 0, we have

01eλa(ζ(τ0))deλa(ζ(τ0))d1eλad1.0\leqslant 1-e^{-\lambda a(\zeta(\tau_{0}))d}\leqslant e^{\lambda a(\zeta(\tau_{0}))d}-1\leqslant e^{\lambda\|a\|_{\infty}d}-1.

If a(ζ(τ0))<0a\big{(}\zeta(\tau_{0})\big{)}<0, we have

01eλa(ζ(τ0))d1eλad.0\geqslant 1-e^{-\lambda a(\zeta(\tau_{0}))d}\geqslant 1-e^{\lambda\|a\|_{\infty}d}.

Then

w(x)w(y)λweλad1λ+CLd0eλasds(aw+CL)eλad1λa.\displaystyle w(x)-w(y)\leqslant\lambda\|w\|_{\infty}\frac{e^{\lambda\|a\|_{\infty}d}-1}{\lambda}+C_{L}\int_{-d}^{0}e^{-\lambda\|a\|_{\infty}s}ds\leqslant(\|a\|_{\infty}\|w\|_{\infty}+C_{L})\frac{e^{\lambda\|a\|_{\infty}d}-1}{\lambda\|a\|_{\infty}}.

Since λ(0,1)\lambda\in(0,1) and dD¯d\leqslant\overline{D}, there is CD¯>0C_{\overline{D}}>0 such that eλad1CD¯λd~e^{\lambda\|a\|_{\infty}d}-1\leqslant C_{\overline{D}}\lambda\tilde{d}. Exchanging the role of xx and yy, we get the conclusion. ∎

As a consequence of the previous proposition and Proposition 3.13, we derive the following information.

Corollary 3.15.

There is κ>0\kappa>0 independent of λ\lambda such that the functions {Vλ:λ(0,λ¯)}\{V_{\lambda}\,:\,\lambda\in(0,\bar{\lambda})\} are κ\kappa-Lipschitz continuous.

Next, we show that the value function is a viscosity subsolution of the equation (HJλ{\textrm{HJ}}_{\lambda}). Indeed, the following result holds.

Proposition 3.16.

Let wC(M)w\in C(M). Then ww is a subsolution of (HJλ{\textrm{HJ}}_{\lambda}) if and only if (DPP) holds for every absolutely curve γ:[t,0]M\gamma:[-t,0]\to M and every t>0t>0.

Proof.

Let us first assume that ww is a viscosity subsolution of (HJ0). By Proposition 1.4, we derive that ww is Lipschitz continuous. Using Theorem 1.15, we take a sequence wnC1(M)w_{n}\in C^{1}(M) such that wnw1/n\|w_{n}-w\|_{\infty}\leqslant 1/n and

λa(x)wn(x)+H(x,Dxwn)c0+1nfor all xM.\lambda a(x)w_{n}(x)+H\big{(}x,D_{x}w_{n}\big{)}\leqslant c_{0}+\frac{1}{n}\qquad\hbox{for all $x\in M$}.

For γ:[t,0]M\gamma:[-t,0]\to M, we have, performing the now usual integration by parts and the Fenchel inequality (1.2),

wn(γ(0))eλt0a(γ(τ))dτwn(γ(t))=t0dds(eλs0a(γ(τ))dτwn(γ(s)))ds\displaystyle w_{n}\big{(}\gamma(0)\big{)}-e^{-\lambda\int_{-t}^{0}a(\gamma(\tau))d\tau}w_{n}\big{(}\gamma(-t)\big{)}=\int_{-t}^{0}\frac{d}{ds}\bigg{(}e^{-\lambda\int_{s}^{0}a(\gamma(\tau))d\tau}w_{n}\big{(}\gamma(s)\big{)}\bigg{)}ds
=t0eλs0a(γ(τ))dτ(λa(γ(s))wn(γ(s))+Dγ(s)wn(γ˙(s)))ds\displaystyle=\int_{-t}^{0}e^{-\lambda\int_{s}^{0}a(\gamma(\tau))d\tau}\bigg{(}\lambda a\big{(}\gamma(s)\big{)}w_{n}\big{(}\gamma(s)\big{)}+D_{\gamma(s)}w_{n}\big{(}\dot{\gamma}(s)\big{)}\bigg{)}ds
t0eλs0a(γ(τ))dτ(λa(γ(s))wn(γ(s))+H(γ(s),Dγ(s)wn)+L(γ(s),γ˙(s)))ds\displaystyle\leqslant\int_{-t}^{0}e^{-\lambda\int_{s}^{0}a(\gamma(\tau))d\tau}\bigg{(}\lambda a\big{(}\gamma(s)\big{)}w_{n}\big{(}\gamma(s)\big{)}+H\big{(}\gamma(s),D_{\gamma(s)}w_{n}\big{)}+L\big{(}\gamma(s),\dot{\gamma}(s)\big{)}\bigg{)}ds
t0eλs0a(γ(τ))dτ(c0+1n+L(γ(s),γ˙(s)))ds\displaystyle\leqslant\int_{-t}^{0}e^{-\lambda\int_{s}^{0}a(\gamma(\tau))d\tau}\bigg{(}c_{0}+\frac{1}{n}+L\big{(}\gamma(s),\dot{\gamma}(s)\big{)}\bigg{)}ds
1nt0eλasds+t0eλs0a(γ(τ))dτ(L(γ(s),γ˙(s))+c0)ds\displaystyle\leqslant\frac{1}{n}\int_{-t}^{0}e^{-\lambda\|a\|_{\infty}s}ds+\int_{-t}^{0}e^{-\lambda\int_{s}^{0}a(\gamma(\tau))d\tau}\Big{(}L\big{(}\gamma(s),\dot{\gamma}(s)\big{)}+c_{0}\Big{)}ds
=eλat1naλ+t0eλs0a(γ(τ))dτ(L(γ(s),γ˙(s))+c0)ds.\displaystyle=\frac{e^{\lambda\|a\|_{\infty}t}-1}{n\|a\|_{\infty}\lambda}+\int_{-t}^{0}e^{-\lambda\int_{s}^{0}a(\gamma(\tau))d\tau}\Big{(}L\big{(}\gamma(s),\dot{\gamma}(s)\big{)}+c_{0}\Big{)}ds.

The assertion follows by sending n+n\to+\infty.

Conversely, let us assume that ww satisfies (DPP). According to Proposition 3.14, ww is Lipschitz continuous. We only need to check if ww is a subsolution where ww is differentiable, thanks to Proposition 1.4. Let ww be differentiable at xx. We take a C1C^{1} curve γ:[t,0]M\gamma:[-t,0]\to M with γ(0)=x\gamma(0)=x and γ˙(0)=v\dot{\gamma}(0)=v. Then

w(γ(0))eλt0a(γ(τ))dτw(γ(t))t1tt0eλs0a(γ(τ))dτ(L(γ(s),γ˙(s))+c0)ds.\frac{w\big{(}\gamma(0)\big{)}-e^{-\lambda\int_{-t}^{0}a(\gamma(\tau))d\tau}w\big{(}\gamma(-t)\big{)}}{t}\leqslant\frac{1}{t}\int_{-t}^{0}e^{-\lambda\int_{s}^{0}a(\gamma(\tau))d\tau}\Big{(}L\big{(}\gamma(s),\dot{\gamma}(s)\big{)}+c_{0}\Big{)}ds.

Letting t0+t\to 0^{+}, we get

λa(x)w(x)+Dxw(v)L(x,v)c0.\lambda a(x)w(x)+D_{x}w(v)-L(x,v)\leqslant c_{0}.

Taking the supremum with respect to vv, we get, thanks to (1.3),

λa(x)w(x)+H(x,Dxw)c0,\lambda a(x)w(x)+H(x,D_{x}w)\leqslant c_{0},

which implies that ww is a subsolution. ∎

We proceed to show that the value function is the maximal viscosity subsolution of (HJλ{\textrm{HJ}}_{\lambda}), and hence a solution by maximality. We need an auxiliary lemma first.

Lemma 3.17.

Let λ>0\lambda>0 and let γ:(,0]M\gamma:(-\infty,0]\to M be an absolutely continuous curve. Let us assume that

supt>0α(λ,γ,t)<+.\sup_{t>0}\alpha(\lambda,\gamma,t)<+\infty. (3.9)

Then eλt0a(γ(τ))dτ0e^{-\lambda\int_{-t}^{0}a(\gamma(\tau))d\tau}\to 0  as t+t\to+\infty.

Proof.

We argue by contradiction. Assume there exist an increasing sequence tn+t_{n}\to+\infty and a δ(0,1)\delta\in(0,1) small enough such that

eλtn0a(γ(τ))dτδfor all n.e^{-\lambda\int_{-t_{n}}^{0}a(\gamma(\tau))d\tau}\geqslant\delta\qquad\qquad\hbox{for all $n\in\mathbb{N}$}.

Then, for all t(tn,tn+ln(δ1)λa)t\in\big{(}t_{n},t_{n}+\frac{\ln(\delta^{-1})}{\lambda\|a\|_{\infty}}\big{)}, we have

λttna(γ(τ))dτλa|ttn|lnδ,-\lambda\int_{-t}^{-t_{n}}a\big{(}\gamma(\tau)\big{)}d\tau\geqslant-\lambda\|a\|_{\infty}|t-t_{n}|\geqslant\ln\delta,

hence

eλt0a(γ(τ))dτ=eλttna(γ(τ))dτeλtn0a(γ(τ))dτδ2,for all n.e^{-\lambda\int_{-t}^{0}a(\gamma(\tau))d\tau}=e^{-\lambda\int_{-t}^{-t_{n}}a(\gamma(\tau))d\tau}e^{-\lambda\int_{-t_{n}}^{0}a(\gamma(\tau))d\tau}\geqslant\delta^{2},\qquad\hbox{for all $n\in\mathbb{N}$.}

Let us pick r>0r>0 with rln(δ1)λar\leqslant\frac{\ln(\delta^{-1})}{\lambda\|a\|_{\infty}}. Up to extracting a subsequence, we can assume that r|tn+1tn|r\leqslant|t_{n+1}-t_{n}| for all nn\in\mathbb{N}. Let us set t0:=0t_{0}:=0. We have

tn0eλs0a(γ(τ))dτdsi=0n1tirtieλs0a(γ(τ))dτdsrδ2n+\int_{-t_{n}}^{0}e^{-\lambda\int_{s}^{0}a(\gamma(\tau))d\tau}ds\geqslant\sum_{i=0}^{n-1}\int_{-t_{i}-r}^{-t_{i}}e^{-\lambda\int_{s}^{0}a(\gamma(\tau))d\tau}ds\geqslant r\delta^{2}n\to+\infty

as n+n\to+\infty, which contradicts (3.9). ∎

Let us prove the result announced above. We recall that the λ¯\bar{\lambda} appearing in the next two statements is the real number in (0,1)(0,1) provided by Proposition 3.13.

Proposition 3.18.

For every fixed λ(0,λ¯)\lambda\in(0,\bar{\lambda}), the value function VλV_{\lambda} is the maximal subsolution of (HJλ{\textrm{HJ}}_{\lambda}). In particular, it is a viscosity solution of (HJλ{\textrm{HJ}}_{\lambda}).

Proof.

Let ww be a subsolution of (HJλ{\textrm{HJ}}_{\lambda}). Let us fix xMx\in M. By definition, for all ε>0\varepsilon>0, there is γΓ\gamma\in\Gamma such that

Vλ(x)+ε\displaystyle V_{\lambda}(x)+\varepsilon >lim supT+T0eλs0a(γ(τ))dτ(L(γ(s),γ˙(s))+c0)ds\displaystyle>\limsup_{T\to+\infty}\int_{-T}^{0}e^{-\lambda\int_{s}^{0}a(\gamma(\tau))d\tau}\Big{(}L\big{(}\gamma(s),\dot{\gamma}(s)\big{)}+c_{0}\Big{)}ds
=t0eλs0a(γ(τ))dτ(L(γ(s),γ˙(s))+c0)ds\displaystyle=\int_{-t}^{0}e^{-\lambda\int_{s}^{0}a(\gamma(\tau))d\tau}\Big{(}L\big{(}\gamma(s),\dot{\gamma}(s)\big{)}+c_{0}\Big{)}ds
+eλt0a(γ(τ))dτlim supT+T0eλs0a(ξ(τ))dτ(L(ξ(s),ξ˙(s))+c0)ds,\displaystyle\quad+e^{-\lambda\int_{-t}^{0}a(\gamma(\tau))d\tau}\limsup_{T\to+\infty}\int_{-T}^{0}e^{-\lambda\int_{s}^{0}a(\xi(\tau))d\tau}\Big{(}L\big{(}\xi(s),\dot{\xi}(s)\big{)}+c_{0}\Big{)}ds,

where ξ(s):=γ(st)\xi(s):=\gamma(s-t) for s0s\leqslant 0. By Proposition 3.16, we have

t0eλs0a(γ(τ))dτ(L(γ(s),γ˙(s))+c0)dsw(x)eλt0a(γ(τ))dτw(γ(t)).\int_{-t}^{0}e^{-\lambda\int_{s}^{0}a(\gamma(\tau))d\tau}\Big{(}L\big{(}\gamma(s),\dot{\gamma}(s)\big{)}+c_{0}\Big{)}ds\geqslant w(x)-e^{-\lambda\int_{-t}^{0}a(\gamma(\tau))d\tau}w\big{(}\gamma(-t)\big{)}.

By the definition of VλV_{\lambda}, we have

eλt0a(γ(τ))dτlim supT+T0eλs0a(ξ(τ))dτ(L(ξ(s),ξ˙(s))+c0)dseλt0a(γ(τ))dτVλ(γ(t)).\displaystyle e^{-\lambda\int_{-t}^{0}a(\gamma(\tau))d\tau}\limsup_{T\to+\infty}\int_{-T}^{0}e^{-\lambda\int_{s}^{0}a(\xi(\tau))d\tau}\Big{(}L\big{(}\xi(s),\dot{\xi}(s)\big{)}+c_{0}\Big{)}ds\geqslant e^{-\lambda\int_{-t}^{0}a(\gamma(\tau))d\tau}V_{\lambda}\big{(}\gamma(-t)\big{)}.

Combining all the above inequalities, we conclude

Vλ(x)+ε>w(x)+eλt0a(γ(τ))dτ(Vλ(γ(t))w(γ(t))).V_{\lambda}(x)+\varepsilon>w(x)+e^{-\lambda\int_{-t}^{0}a(\gamma(\tau))d\tau}\Big{(}V_{\lambda}\big{(}\gamma(-t)\big{)}-w\big{(}\gamma(-t)\big{)}\Big{)}.

We know that supt>0α(λ,γ,t)<+\sup\limits_{t>0}\alpha(\lambda,\gamma,t)<+\infty, cf. proof of Proposition 3.13. By Lemma 3.17, we derive that eλt0a(γ(τ))dτ0e^{-\lambda\int_{-t}^{0}a(\gamma(\tau))d\tau}\to 0 as t+t\to+\infty. By finally sending ε0+\varepsilon\to 0^{+} we conclude that Vλ(x)w(x)V_{\lambda}(x)\geqslant w(x). This, together with Propositions 3.4 and 3.16, proves the asserted maximality of VλV_{\lambda}.

Now we show that VλV_{\lambda} is a viscosity solution of (HJλ{\textrm{HJ}}_{\lambda}). Since VλV_{\lambda} is the maximal subsolution, we only need to show that it is a supersolution. The argument is standard and depends on the bump construction. We give it below for the reader’s convenience. Assume, by contradiction, that the supersolution test fails at some point zMz\in M. This means that there is a strict subtangent ϕC1(M)\phi\in C^{1}(M) of VλV_{\lambda} at zz888Meaning that VλϕV_{\lambda}-\phi has a strict local minimum at zz. such that

λa(z)Vλ(z)+H(z,Dzϕ)<c0.\lambda a(z)V_{\lambda}(z)+H\big{(}z,D_{z}\phi\big{)}<c_{0}.

Up to adding a constant to ϕ\phi, we can assume that Vλ(z)=ϕ(z)V_{\lambda}(z)=\phi(z). Let Br(z)B_{r}(z) be the open ball centered at zz with the radius rr. Let us choose r>0r>0 and ε>0\varepsilon>0 small enough so that

ϕ+ε<Vλon Br(z)andλa(x)(ϕ(x)+ε)+H(x,Dxϕ)<c0xBr(z).\displaystyle\phi+\varepsilon<V_{\lambda}\ \hbox{on $\partial B_{r}(z)$}\quad\hbox{and}\quad\lambda a(x)(\phi(x)+\varepsilon)+H\big{(}x,D_{x}\phi\big{)}<c_{0}\ \ \forall x\in B_{r}(z). (3.10)

Set

V~λ(x):={max{Vλ(x),ϕ(x)+ε}if xBr(z).Vλ(x)if xMBr(z)\widetilde{V}_{\lambda}(x):=\begin{cases}\max\{V_{\lambda}(x),\phi(x)+\varepsilon\}&\hbox{if $x\in B_{r}(z)$.}\\ V_{\lambda}(x)&\hbox{if $x\in M\setminus B_{r}(z)$}\end{cases}

Due to (3.10), the function V~λ\widetilde{V}_{\lambda} is a subsolution of (HJλ{\textrm{HJ}}_{\lambda}) in Br(z)B_{r}(z) as the maximum of two subsolutions in that open ball, and it agrees with VλV_{\lambda} in an open neighborhood of MBr(z)M\setminus B_{r}(z). This readily implies that V~λ\widetilde{V}_{\lambda} is a subsolution of (HJλ{\textrm{HJ}}_{\lambda}) on the whole MM. Yet, we have V~λ(z)=φ(z)+ε>Vλ(z)\widetilde{V}_{\lambda}(z)=\varphi(z)+\varepsilon>V_{\lambda}(z), contradicting the fact that VλV_{\lambda} is the maximal subsolution of (HJλ{\textrm{HJ}}_{\lambda}). This shows that VλV_{\lambda} is indeed a solution to (HJλ{\textrm{HJ}}_{\lambda}) in MM. ∎

From now on, we denote by uλ(x)u_{\lambda}(x) the value function Vλ(x)V_{\lambda}(x), since it is the maximal solution.

Proposition 3.19.

There exists λ0(0,λ¯)\lambda_{0}\in(0,\bar{\lambda}) such that, for every fixed xMx\in M and λ(0,λ0)\lambda\in(0,\lambda_{0}), we can find a curve γxλ:(,0]M\gamma^{x}_{\lambda}:(-\infty,0]\to M with γxλ(0)=x\gamma^{x}_{\lambda}(0)=x such that

uλ(x)=0eλs0a(γxλ(τ))dτ(L(γxλ(s),γ˙xλ(s))+c0)ds.u_{\lambda}(x)=\int_{-\infty}^{0}e^{-\lambda\int_{s}^{0}a(\gamma^{x}_{\lambda}(\tau))d\tau}\Big{(}L\big{(}\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s)\big{)}+c_{0}\Big{)}ds.

Furthermore, the curve γxλ\gamma^{x}_{\lambda} is κ^\hat{\kappa}-Lipschitz continuous, for some constant κ^>0\hat{\kappa}>0 independent of λ(0,λ0)\lambda\in(0,\lambda_{0}) and xMx\in M.

Proof.

Let us fix λ(0,λ¯)\lambda\in(0,\bar{\lambda}). According to Proposition 3.18, uλu_{\lambda} is a viscosity, hence a weak KAM, solution of

λa(x)uλ(x)+H(x,Dxu)=c0in M.\lambda a(x)u_{\lambda}(x)+H(x,D_{x}u)=c_{0}\qquad\hbox{in $M$}.

By Lemma 2.9, we know that, for every fixed xMx\in M, there is a Lipschitz curve γxλ:(,0]M\gamma^{x}_{\lambda}:(-\infty,0]\to M with γxλ(0)=x\gamma^{x}_{\lambda}(0)=x such that

uλ(γxλ(b2))uλ(γxλ(b1))=b1b2[L(γxλ(s),γ˙xλ(s))+c0λa(γxλ(s))uλ(γxλ(s))]ds\displaystyle u_{\lambda}\big{(}\gamma^{x}_{\lambda}(b_{2})\big{)}-u_{\lambda}\big{(}\gamma^{x}_{\lambda}(b_{1})\big{)}=\int_{b_{1}}^{b_{2}}\bigg{[}L\big{(}\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s)\big{)}+c_{0}-\lambda a\big{(}\gamma^{x}_{\lambda}(s)\big{)}u_{\lambda}\big{(}\gamma^{x}_{\lambda}(s)\big{)}\bigg{]}ds (3.11)

for all b1<b20b_{1}<b_{2}\leqslant 0. The above equality with b1=tb_{1}=-t and b2=0b_{2}=0 can be restated as

u(t)=u(0)+0th(s)u(s)ds+0t(s)dsfor all t>0,u(t)=u(0)+\int_{0}^{t}h(s)u(s)ds+\int_{0}^{t}\ell(s)ds\qquad\hbox{for all $t>0$}, (3.12)

where

u(s):=uλ(γxλ(s)),h(s):=λa(γxλ(s)),(s):=L(γxλ(s),γ˙xλ(s))+c0.u(s):=-u_{\lambda}\big{(}\gamma^{x}_{\lambda}(-s)\big{)},\quad h(s):=\lambda a\big{(}\gamma^{x}_{\lambda}(-s)\big{)},\quad\ell(s):=L\big{(}\gamma^{x}_{\lambda}(-s),\dot{\gamma}^{x}_{\lambda}(-s)\big{)}+c_{0}.

Note that the function uu and hh are in L([0,t])L^{\infty}\big{(}[0,t]\big{)}. Since \ell is uniformly bounded from below, this implies, in view of (3.12), that \ell is in L1([0,t])L^{1}\big{(}[0,t]\big{)}. The operator 𝒯\mathcal{T} defined by

𝒯f(t)=u(0)+0th(s)f(s)ds+0t(s)ds\mathcal{T}f(t)=u(0)+\int_{0}^{t}h(s)f(s)ds+\int_{0}^{t}\ell(s)ds

is a contraction when λat<1\lambda\|a\|_{\infty}t<1, in particular there is a unique fixed point of 𝒯\mathcal{T}. This gives the existence of the local unique solution of (3.12), which is defined, for example, on [0,t0][0,t_{0}] with t0=12λat_{0}=\frac{1}{2\lambda\|a\|_{\infty}}. Since the time of local existence is independent of the initial data, the maximal solution is defined for all t>0t>0, and is given by

u(t)=u(0)e0th(s)ds+0t(s)etsh(τ)dτds,u(t)=u(0)e^{\int_{0}^{t}h(s)ds}+\int_{0}^{t}\ell(s)e^{\int^{t}_{s}h(\tau)d\tau}ds,

as an explicit computation shows. This gives directly

uλ(x)=eλt0a(γxλ(τ))dτuλ(γxλ(T))+t0eλs0a(γxλ(τ))dτ(L(γxλ(s),γ˙xλ(s))+c0)ds\displaystyle u_{\lambda}(x)=e^{-\lambda\int_{-t}^{0}a(\gamma^{x}_{\lambda}(\tau))d\tau}u_{\lambda}\big{(}\gamma^{x}_{\lambda}(-T)\big{)}+\!\!\int_{-t}^{0}e^{-\lambda\int_{s}^{0}a(\gamma^{x}_{\lambda}(\tau))d\tau}\Big{(}L\big{(}\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s)\big{)}+c_{0}\Big{)}ds (3.13)

for all t>0t>0. Due to the fact that the functions (uλ)λ(0,λ¯)(u_{\lambda})_{\lambda\in(0,\overline{\lambda})} are equi-bounded and equi-Lipschitz, this implies that the curve γλx\gamma_{\lambda}^{x} is κ^\hat{\kappa}-Lipschitz, with a Lipschitz constant κ^\hat{\kappa} that is independent of λ(0,λ¯)\lambda\in(0,\overline{\lambda}) and xMx\in M, see Lemma 2.9.

We want to show that there exists λ0(0,λ¯]\lambda_{0}\in(0,\bar{\lambda}] such that eλt0a(γxλ(τ))dτ0e^{-\lambda\int_{-t}^{0}a(\gamma^{x}_{\lambda}(\tau))d\tau}\to 0 as t+t\to+\infty whenever λ(0,λ0)\lambda\in(0,\lambda_{0}). According to Lemma 3.17, it suffices to show that there exists a λ0(0,λ¯]\lambda_{0}\in(0,\bar{\lambda}] such that

sup{λα(λ,γ,t):λ(0,λ0),γ𝒞λx,xM,t>0}<+,\sup\bigg{\{}\lambda\alpha(\lambda,\gamma,t):\ \lambda\in(0,\lambda_{0}),\gamma\in\mathscr{C}_{\lambda}^{x},\,x\in M,\,\ t>0\bigg{\}}<+\infty, (3.14)

where we have denoted by 𝒞xλ\mathscr{C}^{x}_{\lambda} the family of absolutely continuous curves γ:(,0]M\gamma:(-\infty,0]\to M with γ(0)=x\gamma(0)=x that satisfy (3.11). Notice in fact that (3.14) implies in particular that, for every fixed λ(0,λ0)\lambda\in(0,\lambda_{0}), condition (3.9) in Lemma 3.17 is met. We argue by contradiction. Let us assume the claim false. Then there exist sequences λn0+\lambda_{n}\to 0^{+}, tn(0,+)t_{n}\in(0,+\infty), xnMx_{n}\in M and γn𝒞λnxn\gamma_{n}\in\mathscr{C}_{\lambda_{n}}^{x_{n}} such that λnα(λn,γn,tn)+\lambda_{n}\alpha(\lambda_{n},\gamma_{n},t_{n})\to+\infty as n+n\to+\infty. Notice that the latter implies that tn+t_{n}\to+\infty and αn:=α(λn,γn,tn)+\alpha_{n}:=\alpha(\lambda_{n},\gamma_{n},t_{n})\to+\infty as n+n\to+\infty. Let μ~n\tilde{\mu}_{n} be the probability measure defined in (\star3.7). Then (3.13) can be restated as

TM(L(x,v)+c0)dμ~n(x,v)=1αn(uλn(γn(0))eλntn0a(γn(τ))dτuλn(γn(tn)))\displaystyle\int_{TM}\big{(}L(x,v)+c_{0}\big{)}\,d\tilde{\mu}_{n}(x,v)=\frac{1}{\alpha_{n}}\bigg{(}u_{\lambda_{n}}\big{(}\gamma_{n}(0)\big{)}-e^{-\lambda_{n}\int_{-t_{n}}^{0}a(\gamma_{n}(\tau))d\tau}u_{\lambda_{n}}\big{(}\gamma_{n}(-t_{n})\big{)}\bigg{)} (3.15)

for all nn\in\mathbb{N}. With the aid of Lemma 3.6, it is easy to check that the right-hand side of (3.15) goes to 0 as n+n\to+\infty. We can therefore apply Lemma 3.7 to infer that, up to extracting a subsequence, μ~n\tilde{\mu}_{n} weakly converges to a Mather measure μ~\tilde{\mu}.

Now we choose a subsolution φ1\varphi\leqslant-1 of (HJ0). For ε>0\varepsilon>0, we take φεC(M)\varphi_{\varepsilon}\in C^{\infty}(M) satisfying φεφε\|\varphi_{\varepsilon}-\varphi\|_{\infty}\leqslant\varepsilon and H(x,Dxφε)c0+εH\big{(}x,D_{x}\varphi_{\varepsilon}\big{)}\leqslant c_{0}+\varepsilon for all xMx\in M, given by Theorem 1.15. For ε\varepsilon small enough, we have φε<0\varphi_{\varepsilon}<0. By (3.15) we get

1αn(uλn(γn(0))eλntn0a(γn(τ))dτuλn(γn(tn)))=TM(L(x,v)+c0)dμ~n\displaystyle\frac{1}{\alpha_{n}}\bigg{(}u_{\lambda_{n}}\big{(}\gamma_{n}(0)\big{)}-e^{-\lambda_{n}\int_{-t_{n}}^{0}a(\gamma_{n}(\tau))d\tau}u_{\lambda_{n}}\big{(}\gamma_{n}(-t_{n})\big{)}\bigg{)}=\int_{TM}\big{(}L(x,v)+c_{0}\big{)}d\tilde{\mu}_{n}
TM(Dxφε(v)H(x,Dxφε)+c0)dμ~nTM(Dxφε(v)ε)dμ~n\displaystyle\geqslant\int_{TM}\Big{(}D_{x}\varphi_{\varepsilon}(v)-H\big{(}x,D_{x}\varphi_{\varepsilon}\big{)}+c_{0}\Big{)}\,d\tilde{\mu}_{n}\geqslant\int_{TM}\big{(}D_{x}\varphi_{\varepsilon}(v)-\varepsilon\big{)}\,d\tilde{\mu}_{n}
=φε(γn(0))eλntn0a(γn(τ))dτφε(γn(tn))α(λn,γn,tn)λnTMa(x)φε(x)dμ~nε\displaystyle=\frac{\varphi_{\varepsilon}\big{(}\gamma_{n}(0)\big{)}-e^{-\lambda_{n}\int_{-t_{n}}^{0}a(\gamma_{n}(\tau))d\tau}\varphi_{\varepsilon}\big{(}\gamma_{n}(-t_{n})\big{)}}{\alpha(\lambda_{n},\gamma_{n},t_{n})}-\lambda_{n}\int_{TM}a(x)\varphi_{\varepsilon}(x)\,d\tilde{\mu}_{n}-\varepsilon
φε(γn(0))αnλnTMa(x)φε(x)dμ~nε,\displaystyle\geqslant\frac{\varphi_{\varepsilon}\big{(}\gamma_{n}(0)\big{)}}{\alpha_{n}}-\lambda_{n}\int_{TM}a(x)\varphi_{\varepsilon}(x)\,d\tilde{\mu}_{n}-\varepsilon,

where for the first inequality we have used Fenchel’s inequality (1.2) and for the subsequent equality an integration by parts. Let ε0+\varepsilon\to 0^{+} to get

1αn(uλn(γn(0))eλntn0a(γn(τ))dτuλn(γn(tn)))φ(γn(0))αnλnTMa(x)φ(x)dμ~n.\displaystyle\frac{1}{\alpha_{n}}\bigg{(}u_{\lambda_{n}}\big{(}\gamma_{n}(0)\big{)}-e^{-\lambda_{n}\int_{-t_{n}}^{0}a(\gamma_{n}(\tau))d\tau}u_{\lambda_{n}}\big{(}\gamma_{n}(-t_{n})\big{)}\bigg{)}\geqslant\frac{\varphi\big{(}\gamma_{n}(0)\big{)}}{\alpha_{n}}-\lambda_{n}\int_{TM}a(x)\varphi(x)\,d\tilde{\mu}_{n}.

Now we divide by λn\lambda_{n}, we use the fact that λnαn+\lambda_{n}\alpha_{n}\to+\infty and Lemma 3.6 to get

TMa(x)φ(x)dμ~(x,v)auλ.\int_{TM}a(x)\varphi(x)\,d\tilde{\mu}(x,v)\geqslant-\|a\|_{\infty}\|u_{\lambda}\|_{\infty}.

Since for, all m0m\geqslant 0, φm0\varphi-m\leqslant 0 is also a subsolution of (HJ0), we get, by applying the previous inequality to φm\varphi-m, that

φTMa(x)φ(x)dμ~(x,v)mTMa(x)dμ~(x,v)auλ,m0,\|\varphi\|_{\infty}\geqslant\int_{TM}a(x)\varphi(x)\,d\tilde{\mu}(x,v)\geqslant m\int_{TM}a(x)\,d\tilde{\mu}(x,v)-\|a\|_{\infty}\|u_{\lambda}\|_{\infty},\quad\forall m\geqslant 0,

which implies that TMa(x)dμ~0\int_{TM}a(x)\,d\tilde{\mu}\leqslant 0. This leads to a contradiction with (a1a1) being μ~\tilde{\mu} a Mather measure, as we recalled above.

Let us fix λ(0,λ0)\lambda\in(0,\lambda_{0}) and xMx\in M. By sending t+t\to+\infty in (3.13) and by exploiting the information just gathered, we derive that

uλ(x)=limt+t0eλs0a(γxλ(τ))dτ(L(γxλ(s),γ˙xλ(s))+c0)ds.\displaystyle u_{\lambda}(x)=\lim_{t\to+\infty}\int_{-t}^{0}e^{-\lambda\int_{s}^{0}a(\gamma^{x}_{\lambda}(\tau))d\tau}\Big{(}L\big{(}\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s)\big{)}+c_{0}\Big{)}ds.

We conclude that

uλ(x)=limt+t0eλs0a(γxλ(τ))dτ(\displaystyle u_{\lambda}(x)=\lim_{t\to+\infty}\int_{-t}^{0}e^{-\lambda\int_{s}^{0}a(\gamma^{x}_{\lambda}(\tau))d\tau}\Big{(} L(γxλ(s),γ˙xλ(s))+c0)ds\displaystyle L\big{(}\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s)\big{)}+c_{0}\Big{)}ds (3.16)
=0eλs0a(γxλ(τ))dτ(L(γxλ(s),γ˙xλ(s))+c0)ds.\displaystyle=\int_{-\infty}^{0}e^{-\lambda\int_{s}^{0}a(\gamma^{x}_{\lambda}(\tau))d\tau}\Big{(}L\big{(}\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s)\big{)}+c_{0}\Big{)}ds.

Indeed (3.14) means that the positive function seλs0a(γxλ(τ))dτs\mapsto e^{-\lambda\int_{s}^{0}a(\gamma^{x}_{\lambda}(\tau))d\tau} is in L1((,0])L^{1}\big{(}(-\infty,0]\big{)}. Furthermore, the fact that the curve γxλ\gamma^{x}_{\lambda} is Lipschitz implies that the function sL(γxλ(s),γ˙xλ(s))s\mapsto L\big{(}\gamma^{x}_{\lambda}(s),\dot{\gamma}^{x}_{\lambda}(s)\big{)} is in L((,0])L^{\infty}\big{(}(-\infty,0]\big{)}. Equality (3.16) follows from this by making use of the Dominated Convergence Theorem. ∎

3.2. Diverging families of solutions

We have seen in Theorem 3.1 that, under the assumption (a1a1), the maximal solution uλu_{\lambda} of (HJλ{\textrm{HJ}}_{\lambda}) uniformly converges as λ0+\lambda\to 0^{+} to a critical solution u0u_{0}. In this section, we will show that if we furthermore assume

  • (aa2) 

    there is a point x0Mx_{0}\in M such that a(x0)<0a(x_{0})<0;

  • (aa3) 

    a(x)0a(x)\geqslant 0 in an open neighborhood UU of the projected Aubry set 𝒜\mathcal{A},

then there exists a family of solutions to (HJλ{\textrm{HJ}}_{\lambda}) that uniformly diverges to -\infty. Furthermore, when condition (a33) is reinforced in favor of the following one

  • (aa3)

    a(x)>0a(x)>0 on 𝒜\mathcal{A},

then all solutions to (HJλ{\textrm{HJ}}_{\lambda}) different from the maximal ones uniformly diverge to -\infty as λ0+\lambda\to 0^{+}. We summarize all this in the following statement. We refer the reader to Section 1.3 for the definition of the projected Aubry set. We underline that, in view of Theorem 1.9, condition (a33) is stronger than the integral condition (a1a1). Throughout this section, we will denote by uλu_{\lambda} the maximal solution of (HJλ{\textrm{HJ}}_{\lambda}).

Theorem 3.20.

Let us assume (aa2).

  • (i)

    If (aa3) holds, there is a family of solutions (vλ)λ(0,λ^)(v_{\lambda})_{\lambda\in(0,\hat{\lambda})} of (HJλ{\textrm{HJ}}_{\lambda}) for some λ^(0,1)\hat{\lambda}\in(0,1) uniformly diverging to -\infty as λ0+\lambda\to 0^{+}.

  • (ii)

    If (aa3\,{}^{\prime}) holds, then any family of solutions (vλ)λ(0,λ)(v_{\lambda})_{\lambda\in(0,\lambda^{\prime})} of (HJλ{\textrm{HJ}}_{\lambda}) with λ(0,1)\lambda^{\prime}\in(0,1) satisfying vλuλv_{\lambda}\not=u_{\lambda} for all λ(0,λ)\lambda\in(0,\lambda^{\prime}) uniformly diverges to -\infty as λ0+\lambda\to 0^{+}.

Proof.

By Theorem 1.6, there is a subsolution φ\varphi of (HJ0) which is CC^{\infty} and strict in M𝒜M\setminus\mathcal{A}. Therefore, for every open neighborhood UU of 𝒜\mathcal{A} there is a δ=δ(U)>0\delta=\delta(U)>0 depending on UU such that

H(x,Dxφ)c0δ<c0for all xMU.H\big{(}x,D_{x}\varphi\big{)}\leqslant c_{0}-\delta<c_{0}\quad\hbox{for all $x\in M\setminus U$.}

Let us prove (i). Let UU be the open neighborhood of 𝒜\mathcal{A} given by condition (aa3). Define

φλ:=φ1λ.\varphi_{\lambda}:=\varphi-\frac{1}{\sqrt{\lambda}}.

For λ^>0\hat{\lambda}>0 small enough, we have φλ<0\varphi_{\lambda}<0 for all λ(0,λ^)\lambda\in(0,\hat{\lambda}). Up to choosing a smaller λ^>0\hat{\lambda}>0 if necessary, we have

λa(x)φλ+H(x,Dxφλ)λa(x)φλ(x)+c0c0a.e. in U,\lambda a(x)\varphi_{\lambda}+H(x,D_{x}\varphi_{\lambda})\leqslant\lambda a(x)\varphi_{\lambda}(x)+c_{0}\leqslant c_{0}\quad\hbox{a.e. in $U$,}

and

λa(x)φλ+H(x,Dxφλ)λaφ+aλ+c0δ<c0in M\U,\lambda a(x)\varphi_{\lambda}+H(x,D_{x}\varphi_{\lambda})\leqslant\lambda\|a\|_{\infty}\|\varphi\|_{\infty}+\|a\|_{\infty}\sqrt{\lambda}+c_{0}-\delta<c_{0}\quad\hbox{in $M\backslash U$,}

for all λ(0,λ^)\lambda\in(0,\hat{\lambda}). Therefore, φλ\varphi_{\lambda} is a subsolution of (HJλ{\textrm{HJ}}_{\lambda}). Now we denote by Tλ,tT^{\lambda,-}_{t} (resp. Tλ,+tT^{\lambda,+}_{t}) the Lax-Oleinik semigroup defined in (1.8) (resp. the forward semigroup defined in (1.9)) associated to L(x,v)λa(x)uL(x,v)-\lambda a(x)u. Define

v+λ:=limt+Tλ,+tφλ,vλ:=limt+Tλ,tv+λ,v^{+}_{\lambda}:=\lim_{t\to+\infty}T^{\lambda,+}_{t}\varphi_{\lambda},\quad v_{\lambda}:=\lim_{t\to+\infty}T^{\lambda,-}_{t}v^{+}_{\lambda},

By Lemma 1.16, v+λφλv^{+}_{\lambda}\leqslant\varphi_{\lambda}, in particular v+λv^{+}_{\lambda} uniformly diverges to -\infty as λ0+\lambda\to 0^{+}. By Lemma 1.17, vλv_{\lambda} is a solution of (HJλ{\textrm{HJ}}_{\lambda}), and there is a point xλMx_{\lambda}\in M such that vλ(xλ)=v+λ(xλ)v_{\lambda}(x_{\lambda})=v^{+}_{\lambda}(x_{\lambda}) for each λ(0,λ^)\lambda\in(0,\hat{\lambda}). We derive that vλ(xλ)v_{\lambda}(x_{\lambda})\to-\infty as λ0+\lambda\to 0^{+}. By Proposition 2.16, we conclude that vλv_{\lambda} uniformly diverges to -\infty as λ0+\lambda\to 0^{+}.

Let us prove (ii). From assumption (aa3) we infer that there is an open neighborhood UU of 𝒜\mathcal{A} and θ>0\theta>0 such that a(x)θa(x)\geqslant\theta for all xUx\in U. One can easily check that there exist λ(0,1)\lambda^{\prime}\in(0,1) small enough and ε>0\varepsilon>0 such that

λa(x)φλ+H(x,Dxφλ)λa(x)φλ(x)+c0λaφλθ+c0c0εa.e. in U\displaystyle\lambda a(x)\varphi_{\lambda}+H(x,D_{x}\varphi_{\lambda})\leqslant\lambda a(x)\varphi_{\lambda}(x)+c_{0}\leqslant\lambda\|a\|_{\infty}\|\varphi\|_{\infty}-\sqrt{\lambda}\,\theta+c_{0}\leqslant c_{0}-\varepsilon\quad\hbox{a.e. in $U$}

and

λa(x)φλ+H(x,Dxφλ)λaφ+aλ+c0δc0εin M\U,\displaystyle\lambda a(x)\varphi_{\lambda}+H(x,D_{x}\varphi_{\lambda})\leqslant\lambda\|a\|_{\infty}\|\varphi\|_{\infty}+\|a\|_{\infty}\sqrt{\lambda}+c_{0}-\delta\leqslant c_{0}-\varepsilon\quad\hbox{in $M\backslash U$,}

for all λ(0,λ)\lambda\in(0,\lambda^{\prime}). Therefore, φλ\varphi_{\lambda} is a strict subsolution of (HJλ{\textrm{HJ}}_{\lambda}).
Claim: Let λ(0,λ)\lambda\in(0,\lambda^{\prime}). There is no solution vλuλv_{\lambda}\not=u_{\lambda} of (HJλ{\textrm{HJ}}_{\lambda}) satisfying φλvλuλ\varphi_{\lambda}\leqslant v_{\lambda}\leqslant u_{\lambda} in MM.

We argue by contradiction. Assume there is such a solution vλv_{\lambda}. We have

Tλ,tφλTλ,tvλTλ,tuλ=uλt>0.T^{\lambda,-}_{t}\varphi_{\lambda}\leqslant T^{\lambda,-}_{t}v_{\lambda}\leqslant T^{\lambda,-}_{t}u_{\lambda}=u_{\lambda}\qquad\forall t>0.

Since φλ\varphi_{\lambda} is a strict subsolution of (HJλ{\textrm{HJ}}_{\lambda}), by Lemma 1.17,

limt+Tλ,tφλ=uλ.\lim_{t\to+\infty}T^{\lambda,-}_{t}\varphi_{\lambda}=u_{\lambda}.

We get

limt+Tλ,tvλ=uλ,\lim_{t\to+\infty}T^{\lambda,-}_{t}v_{\lambda}=u_{\lambda},

which contradicts the fact that vλv_{\lambda} is a fixed point of Tλ,tT^{\lambda,-}_{t}.

Therefore, for each solution vλv_{\lambda} of (HJλ{\textrm{HJ}}_{\lambda}), there is a point xλMx_{\lambda}\in M such that

vλ(xλ)φλ(xλ)as λ0+.v_{\lambda}(x_{\lambda})\leqslant\varphi_{\lambda}(x_{\lambda})\to-\infty\quad\hbox{as $\lambda\to 0^{+}$}.

By Proposition 2.16, vλv_{\lambda} uniformly converges to -\infty as λ0+\lambda\to 0^{+}. ∎

Remark 3.21.

We note that the proof above also shows that the solutions vλv_{\lambda} diverge to -\infty with speed (at least) of order 1/λ-1/\sqrt{\lambda}.

Acknowledgements

Andrea Davini is a member of the INdAM Research Group GNAMPA. He is supported for this research by Sapienza Università di Roma - Research Funds 2021. Panrui Ni and Maxime Zavidovique are supported by ANR CoSyDy (ANR-CE40-0014). Jun Yan is supported by National Natural Science Foundation of China (Grant Nos. 12171096, 12231010).

Declarations

Conflict of interest statement: The author states that there is no conflict of interest.

Data availability statement: Data sharing not applicable to this article as no datasets were generated or analysed during the current study.

References

  • [1] L. Ambrosio, N. Gigli, and G. Savaré, Gradient flows in metric spaces and in the space of probability measures, Lectures in Mathematics ETH Zürich, Birkhäuser Verlag, Basel, 2005.
  • [2] M.-C. Arnaud and M. Zavidovique, Weak K.A.M. solutions and minimizing orbits of twist maps, Trans. Amer. Math. Soc., 376 (2023), pp. 8129–8171.
  • [3] G. Barles, Solutions de viscosité des équations de Hamilton-Jacobi, vol. 17 of Mathématiques & Applications (Berlin) [Mathematics & Applications], Springer-Verlag, Paris, 1994.
  • [4] P. Cardaliaguet and A. Porretta, Long time behavior of the master equation in mean field game theory, Analysis & PDE, 12 (2019), pp. 1397–1453.
  • [5] Q. Chen, Convergence of solutions of Hamilton–Jacobi equations depending nonlinearly on the unknown function, Advances in Calculus of Variations, (2021).
  • [6] Q. Chen, W. Cheng, H. Ishii, and K. Zhao, Vanishing contact structure problem and convergence of the viscosity solutions, Comm. Partial Differential Equations, 44 (2019), pp. 801–836.
  • [7] Q. Chen, A. Fathi, M. Zavidovique, and J. Zhang, Convergence of the solutions of the nonlinear discounted Hamilton–Jacobi equation: The central role of Mather measures, Journal de Mathématiques Pures et Appliquées, 181 (2024), pp. 22–57.
  • [8] G. Contreras, A. Figalli, and L. Rifford, Generic hyperbolicity of Aubry sets on surfaces, Invent. Math., 200 (2015), pp. 201–261.
  • [9] G. Contreras and R. Iturriaga, Global minimizers of autonomous Lagrangians, 22o Colóquio Brasileiro de Matemática. [22nd Brazilian Mathematics Colloquium], Instituto de Matemática Pura e Aplicada (IMPA), Rio de Janeiro, 1999.
  • [10] G. Contreras, R. Iturriaga, G. P. Paternain, and M. Paternain, Lagrangian graphs, minimizing measures and Mañé’s critical values, Geom. Funct. Anal., 8 (1998), pp. 788–809.
  • [11] A. Davini, A. Fathi, R. Iturriaga, and M. Zavidovique, Convergence of the solutions of the discounted equation: the discrete case, Math. Z., 284 (2016), pp. 1021–1034.
  • [12]  , Convergence of the solutions of the discounted Hamilton-Jacobi equation, Invent. Math., 206 (2016), pp. 29–55.
  • [13] A. Davini and A. Siconolfi, A generalized dynamical approach to the large time behavior of solutions of Hamilton-Jacobi equations, SIAM J. Math. Anal., 38 (2006), pp. 478–502.
  • [14] A. Davini, A. Siconolfi, and M. Zavidovique, Random Lax-Oleinik semigroups for Hamilton-Jacobi systems, J. Math. Pures Appl. (9), 120 (2018), pp. 294–333.
  • [15] A. Davini and L. Wang, On the vanishing discount problem from the negative direction, Discrete Contin. Dyn. Syst., 41 (2021), pp. 2377–2389.
  • [16] A. Davini and M. Zavidovique, Weak KAM theory for nonregular commuting Hamiltonians, Discrete Contin. Dyn. Syst. Ser. B, 18 (2013), pp. 57–94.
  • [17]  , Convergence of the solutions of discounted Hamilton-Jacobi systems, Advances in Calculus of Variations, 14 (2019), pp. 1–15.
  • [18] I. C. Dolcetta and A. Davini, On the vanishing discount approximation for compactly supported perturbations of periodic Hamiltonians: the 1d case, Comm. Partial Differential Equations, 48 (2023), p. 576–622.
  • [19] A. Fathi, Weak KAM from a PDE point of view: viscosity solutions of the Hamilton-Jacobi equation and Aubry set, Proc. Roy. Soc. Edinburgh Sect. A, 142 (2012), pp. 1193–1236.
  • [20] A. Fathi and E. Maderna, Weak KAM theorem on non compact manifolds, NoDEA Nonlinear Differential Equations Appl., 14 (2007), pp. 1–27.
  • [21] A. Fathi and A. Siconolfi, PDE aspects of Aubry-Mather theory for quasiconvex Hamiltonians, Calc. Var. Partial Differential Equations, 22 (2005), pp. 185–228.
  • [22] D. A. Gomes, Generalized Mather problem and selection principles for viscosity solutions and Mather measures, Adv. Calc. Var., 1 (2008), pp. 291–307.
  • [23] D. A. Gomes, H. Mitake, and H. V. Tran, The selection problem for discounted Hamilton-Jacobi equations: some non-convex cases, J. Math. Soc. Japan, 70 (2018), pp. 345–364.
  • [24] H. Ishii, The vanishing discount problem for monotone systems of Hamilton-Jacobi equations. Part 1: linear coupling, Mathematics in Engineering, 3 (2021), pp. 1–21.
  • [25] H. Ishii and L. Jin, The vanishing discount problem for monotone systems of Hamilton-Jacobi equations. Part 2: nonlinear coupling, Calc. Var., 59 (2020), pp. 1–28.
  • [26] H. Ishii, H. Mitake, and H. Tran, The vanishing discount problem and viscosity Mather measures. Part 1: The problem on a torus, J. Math. Pures Appl., 108 (2017), p. 125–149.
  • [27]  , The vanishing discount problem and viscosity Mather measures. part 2: Boundary value problems, J. Math. Pures Appl., 108 (2017), pp. 261–305.
  • [28] H. Ishii and A. Siconolfi, The vanishing discount problem for Hamilton-Jacobi equations in the Euclidean space, Comm. Partial Differential Equations, 45 (2020), pp. 525–560.
  • [29] R. Iturriaga and H. Sánchez-Morgado, Limit of the infinite horizon discounted Hamilton-Jacobi equation, Discrete Contin. Dyn. Syst. Ser. B, 15 (2011), pp. 623–635.
  • [30] L. Jin, J. Yan, and K. Zhao, Nonlinear semigroup approach to the Hamilton-Jacobi equation—a toy model, Minimax Theory Appl., 8 (2023), pp. 61–84.
  • [31] N. R. Kocherlakota, On the ‘discount’ factor in growth economies, Journal of Monetary Economics, 25 (1990), pp. 43–47.
  • [32] P.-L. Lions, G. Papanicolaou, and S. Varadhan, Homogenization of Hamilton-Jacobi equation. unpublished preprint, 1987.
  • [33] R. Mañé, Generic properties and problems of minimizing measures of Lagrangian systems, Nonlinearity, 9 (1996), pp. 273–310.
  • [34] H. Mitake and H. Tran, Selection problems for a discount degenerate viscous Hamilton-Jacobi equation, Adv. Math., 306 (2017), pp. 684–703.
  • [35] P. Ni, Multiple asymptotic behaviors of solutions in the generalized vanishing discount problem, Proc. Amer. Math. Soc., 151 (2023), pp. 5239–5250.
  • [36] P. Ni and L. Wang, A nonlinear semigroup approach to Hamilton-Jacobi equations—revisited, J. Differential Equations, 403 (2024), pp. 272–307.
  • [37] P. Ni, L. Wang, and J. Yan, A representation formula of the viscosity solution of the contact Hamilton-Jacobi equation and its applications, Chinese Ann. Math. Ser. B, (to appear).
  • [38] P. Ni and M. Zavidovique, Nonlinear and degenerate discounted approximation in discrete weak KAM theory, 2024.
  • [39] Y. Wang, J. Yan, and J. Zhang, Convergence of viscosity solutions of generalized contact Hamilton-Jacobi equations, Arch. Rational Mech. Anal., 241 (2021), pp. 1–18.
  • [40]  , On the negative limit of viscosity solutions for discounted Hamilton-Jacobi equations, J. Dyn. Differ. Equ., (2022).
  • [41] M. Zavidovique, Convergence of solutions for some degenerate discounted Hamilton–Jacobi equations, Analysis & PDE, 15 (2022), pp. 1287 – 1311.
  • [42]  , Discrete and Continuous Weak KAM Theory: an introduction through examples and its applications to twist maps, 2023.
  • [43] J. Zhang, Limit of solutions for semilinear Hamilton-Jacobi equations with degenerate viscosity, Advances in Calculus of Variations, (2023).
  • [44] B. Ziliotto, Convergence of the solutions of the discounted Hamilton-Jacobi equation: A counterexample, J. Math. Pures Appl., 128 (2019), pp. 330–338.