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

11institutetext: Xifeng Su22institutetext: Laboratory of Mathematics and Complex Systems (Ministry of Education), School of Mathematical Sciences, Beijing Normal University, Beijing 100875, People’s Republic of China
Tel: +86-15910594182
22email: [email protected], [email protected]
33institutetext: Jianlu Zhang 44institutetext: Hua Loo-Keng Key Laboratory of Mathematics & Mathematics Institute, Academy of Mathematics and systems science, Chinese Academy of Sciences, Beijing 100190, China
Tel: +86-182-1038-3625
44email: [email protected]

Essential forward weak KAM solution for the convex Hamilton-Jacobi equation

Xifeng Su    Jianlu Zhang
(Received: date / Accepted: date)
Abstract

For a convex, coercive continuous Hamiltonian on a compact closed Riemannian manifold MM, we construct a unique forward weak KAM solution of

H(x,dxu)=c(H)H(x,d_{x}u)=c(H)

by a vanishing discount approach, where c(H)c(H) is the Mañé critical value. We also discuss the dynamical significance of such a special solution.

Keywords:
discounted equation, weak KAM solution, Aubry Mather theory, Hamilton-Jacobi equation, viscosity solution
MSC:
35B40, 37J50, 49L25,37J51

1 Introduction

For a compact connected manifold MM without boundary, the Hamiltonian is usually mentioned as a continuous function defined on its cotangent bundle TMT^{*}M. In LPV , the authors firstly proposed the ergodic approximation technique, to consider the existence of so called viscosity solutions to the Hamilton-Jacobi equation

H(x,dxu)=c(H),xMH(x,d_{x}u)=c(H),\quad x\in M (HJ0)

for the Mañé critical value

c(H):=inf{c|ωC(M,) such that H(x,dxω)c,a.e.xM}.c(H):=\inf\{c\in\mathbb{R}|\exists\ \omega\in C(M,\mathbb{R})\text{ such that }H(x,d_{x}\omega)\leq c,\quad a.e.\ x\in M\}.

The Hamiltonian they concerned satisfies

  • (Coercivity) H(x,p)H(x,p) is coercive in pTxMp\in T_{x}^{*}M, uniformly w.r.t. xMx\in M.

  • (Convexity) H(x,p)H(x,p) is convex in pTxMp\in T_{x}^{*}M for all xMx\in M.

They perturbed (HJ0) by the following discounted equation

λu+H(x,dxu)=c(H),xM,λ>0\displaystyle\lambda u+H(x,d_{x}u)=c(H),\quad x\in M,\lambda>0 (1)

of which the Comparison Principle is allowed. Therefore, the viscosity solution uλu_{\lambda}^{-} of (1) is unique. In DFIZ , they established the convergence of uλu_{\lambda}^{-} as λ0+\lambda\rightarrow 0_{+}, to a specified viscosity solution u0u_{0}^{-} of (HJ0) which can be characterized by the combination of subsolutions of (HJ0), or Peierls barrier

h:M×Mh^{\infty}:M\times M\rightarrow\mathbb{R}

w.r.t. the projected Mather measures 𝔐\mathfrak{M} of (HJ0), see Appendix A for the relevant definitions of 𝔐,h\mathfrak{M},\ h^{\infty}, subsolutions etc.

In this paper, we consider a negative limit technique and try to find another specified solution of (HJ0). Precisely, we consider

λu+H(x,dxu)=c(H),xM,λ>0-\lambda u+H(x,d_{x}u)=c(H),\quad x\in M,\lambda>0 (HJλ)

of which a unique forward λ\lambda-weak KAM solution uλ+u_{\lambda}^{+} can be found (see Remark 2.3 for the definition). As λ0+\lambda\rightarrow 0_{+}, we get the following conclusion:

Theorem 1.1

Let H:TM,(x,p)H(x,p)H:T^{*}M\rightarrow\mathbb{R},(x,p)\mapsto H(x,p) be a continuous Hamiltonian coercive and convex in pp. For λ>0\lambda>0, the unique forward λ\lambda-weak KAM solution uλ+u_{\lambda}^{+} of (HJλ) uniformly converges as λ0+\lambda\rightarrow 0_{+}, to a unique forward 00-weak KAM solution111see Definition A.6 u0+u_{0}^{+} of (HJ0), which can be interpreted as

u0+(x)=inf+\displaystyle u_{0}^{+}(x)=\inf\mathcal{F}_{+} (2)

with

+:={w is a subsolution of (HJ0|Mw𝑑μ0,μ𝔐}\displaystyle\mathcal{F}_{+}:=\Big{\{}w\text{ is a subsolution of (\ref{eq:hj}) }\Big{|}\int_{M}wd\mu\geq 0,\ \forall\mu\in\mathfrak{M}\Big{\}} (3)

and

u0+(x)=infμ𝔐Mh(x,y)𝑑μ(y).\displaystyle u_{0}^{+}(x)=-\inf_{\mu\in\mathfrak{M}}\int_{M}h^{\infty}(x,y)d\mu(y). (4)
Remark 1.2

The novelty of this paper is that we adapt a symmetric Lagrangian skill to our C0C^{0}-setting. The lack of regularity invalidates a bunch of important properties of the Mather measures, Peierls barrier etc., so we have to find substitutes in the weak sense.

Besides, we mention that u0+(x)-u_{0}^{+}(x) is a viscosity solution of the symmetric equation

H(x,dxu(x))=c(H).H(x,-d_{x}u(x))=c(H).

Comparing to the backward 00-weak KAM solutions, the notion of viscosity solutions is more familiar to PDE specialists, although both are proved to be equivalent in F .

1.1 Dynamic interpretation of u0+u_{0}^{+}

Now the vanishing discount approach supplies us with a pair of solutions of (HJ0):

{u0(x)=infμ𝔐Mh(y,x)𝑑μ(y),u0+(x)=infμ𝔐Mh(x,y)𝑑μ(y).\left\{\begin{aligned} &u_{0}^{-}(x)=\inf_{\mu\in\mathfrak{M}}\int_{M}h^{\infty}(y,x)d\mu(y),\\ &u_{0}^{+}(x)=-\inf_{\mu\in\mathfrak{M}}\int_{M}h^{\infty}(x,y)d\mu(y).\end{aligned}\right. (5)
Definition 1.3 (Conjugated Pair)

A backward 00-weak KAM solution uu^{-} of (HJ0) is conjugated to a forward 00-weak KAM solution u+u^{+}, if

  • u=u+u^{-}=u^{+} on the projected Mather set \mathcal{M} (see Definition A.4).

  • uu+u^{-}\geq u^{+} on MM.

In F , above definition is actually proposed to C2C^{2}-Tonelli Hamiltonian222H:(x,p)TMH:(x,p)\in T^{*}M\rightarrow\mathbb{R} is called Tonelli, if it’s positive definite and superlinear in pp. However, we have no difficulty to reserve this concept to the C0C^{0}-case by means of the evidence in DS . Moreover, For the following typical Hamiltonians, (u0,u0+)(u_{0}^{-},u_{0}^{+}) indeed forms a conjugated pair:

  1. 1.

    (uniquely ergodic) Suppose 𝔐\mathfrak{M} consists of a uniquely ergodic projected Mather measure (generic for C2C^{2}-Tonelli Hamiltonians, see Mn ), then

    dc(x,y):=h(x,y)+h(y,x)0d_{c}(x,y):=h^{\infty}(x,y)+h^{\infty}(y,x)\geq 0

    for all x,yMx,y\in M, and ‘==’ holds for x,yx,y\in\mathcal{M} (due to the definition of the Mather measure in Appendix A ). So (u0,u0+)(u_{0}^{-},u_{0}^{+}) is a conjugated pair.

  2. 2.

    (mechanical system) For a mechanical Hamiltonian

    H(x,p)=12p,p+V(x),H(x,p)=\frac{1}{2}\langle p,p\rangle+V(x),

    we can easily get c(H)=maxxMV(x)c(H)=\max_{x\in M}V(x), then the associated L(x,v)+c(H)0L(x,v)+c(H)\geq 0 on TMTM. Consequently, h:M×Mh^{\infty}:M\times M\rightarrow\mathbb{R} is nonnegative, so u0u0+u_{0}^{-}\geq u_{0}^{+} on MM. On the other side, due to the definition of 𝔐\mathfrak{M}, all the Mather measures are supported by equilibriums. So u0=u0+u_{0}^{-}=u_{0}^{+} on \mathcal{M}. In summary (u0,u0+)(u_{0}^{-},u_{0}^{+}) is a conjugated pair.

  3. 3.

    (constant subsolution) Such a case is also discussed in DFIZ . If H(x,0)c(H)H(x,0)\leq c(H) for all xMx\in M, i.e. constant is a subsolution of (HJ0), then due to the Young Inequality we get

    L(x,v)+c(H)L(x,v)+H(x,0)v,0=0L(x,v)+c(H)\geq L(x,v)+H(x,0)\geq\langle v,0\rangle=0

    for all (x,v)TM(x,v)\in TM, by a similar analysis like the case of mechanical systems (u0,u0+)(u_{0}^{-},u_{0}^{+}) proves to be a conjugated pair.

1.2 Organization of the article.

In Sec. 2, we prove some variational properties for nonsmooth Lagrangians. In Sec. 3, we prove the convergence of uλ+u_{\lambda}^{+} as λ0+\lambda\rightarrow 0_{+} and give a representative formula for the limit. For the consistency and readability of the article, some preliminary materials are moved to Appendix.

Acknowledgement. J. Zhang is supported by the National Natural Science Foundation of China (Grant No. 11901560). X. Su is supported by the National Natural Science Foundation of China (Grant No. 11971060, 11871242).

2 Nonsmooth symmetric Lagrangians

With the same adaption as in DFIZ , without loss of generality we can assume H(x,p)H(x,p) is superlinear in pp, i.e.

  • (Superlinearity) lim|p|+H(x,p)/|p|=+\lim_{|p|\rightarrow+\infty}H(x,p)/|p|=+\infty, for any xMx\in M.

In that case, by Fenchel’s formula (see 10), the Hamiltonian has an associated Lagrangian L:(x,v)TML:(x,v)\in TM\rightarrow\mathbb{R} which is superlinear and convex in the fibers of the tangent bundle. Consequently, we can propose a symmetrical Lagrangian L^(x,v):=L(x,v)\widehat{L}(x,v):=L(x,-v), of which the following fundamental facts hold:

Lemma 2.1
  • (i)

    The conjugated Hamiltonian H^:TM\widehat{H}:T^{*}M\rightarrow\mathbb{R} of L^(x,v)\widehat{L}(x,v) satisfies H^(x,p)=H(x,p)\widehat{H}(x,p)=H(x,-p) for all (x,p)TM(x,p)\in T^{*}M. Therefore, H^\widehat{H} is also continuous, superlinear and convex.

  • (ii)

    H^(x,dxω)cH(x,dx(ω))c\widehat{H}(x,d_{x}\omega)\leq c\iff H(x,d_{x}(-\omega))\leq c

  • (iii)

    c(H^)=c(H)c(\widehat{H})=c(H).

  • (iv)

    The projected Mather measure set 𝔐^\widehat{\mathfrak{M}} (associated with H^(x,p)\widehat{H}(x,p)) keeps the same with 𝔐\mathfrak{M}.

  • (v)

    The Peierls barrier function associated with L^(x,v)\widehat{L}(x,v) satisfies h^(y,x)=h(x,y)\widehat{h}^{\infty}(y,x)=h^{\infty}(x,y) for any x,yMx,y\in M.

Proof

(i). Due to (10) and the definition of L^\widehat{L}, we have

H^(x,p)=maxvTxM{p,vL^(x,v)}=maxvTxM{p,vL(x,v)}=maxwTxM{p,wL(x,w)}=H(x,p),.\begin{split}\widehat{H}(x,p)&=\max_{v\in T_{x}M}\{\langle p,v\rangle-\widehat{L}(x,v)\}\\ &=\max_{v\in T_{x}M}\{\langle p,v\rangle-L(x,-v)\}\\ &=\max_{w\in T_{x}M}\{\langle-p,w\rangle-L(x,w)\}=H(x,-p),\end{split}.

as is desired.

(ii). If ω\omega is a subsolution of H^(x,dxω)c\widehat{H}(x,d_{x}\omega)\leq c, then for any absolutely continuous γ:[T,T]M\gamma:[-T,T]\rightarrow M connecting x,yMx,y\in M, we get

ω(y)ω(x)TT(L^(γ,γ˙)+c)𝑑t.\omega(y)-\omega(x)\leq\int_{-T}^{T}\big{(}\widehat{L}(\gamma,\dot{\gamma})+c\big{)}dt.

Furthermore,

ω(x)(ω(y))\displaystyle-\omega(x)-(-\omega(y)) =\displaystyle= ω(y)ω(x)\displaystyle\omega(y)-\omega(x)
\displaystyle\leq TT(L^(γ,γ˙)+c)𝑑t\displaystyle\int_{-T}^{T}\big{(}\widehat{L}(\gamma,\dot{\gamma})+c\big{)}dt
=\displaystyle= TT(L(γ,γ˙)+c)𝑑t\displaystyle\int_{-T}^{T}\big{(}L(\gamma,-\dot{\gamma})+c\big{)}dt
=\displaystyle= TT(L(γ^,γ^˙)+c)𝑑t\displaystyle\int_{-T}^{T}\big{(}L(\widehat{\gamma},-\dot{\widehat{\gamma}})+c\big{)}dt

with γ^(t):=γ(t)\widehat{\gamma}(t):=\gamma(-t) for all t[T,T]t\in[-T,T]. As γ\gamma is arbitrarily chosen, so we get ωL+c-\omega\prec L+c444see Definition A.5, then H(x,dxω)cH(x,-d_{x}\omega)\leq c for a.e. xMx\in M. Similarly, ωL+c\omega\prec L+c indicates ωL^+c-\omega\prec\widehat{L}+c.

(iii). As c(H^)=inf{c|ωC(M,) such that ωL^+c}c(\widehat{H})=\inf\{c\in\mathbb{R}|\exists\ \omega\in C(M,\mathbb{R})\text{ such that }\omega\prec\widehat{L}+c\}, then due to (ii), c(H^)=c(H)c(\widehat{H})=c(H).

(iv). Due to Proposition 2-4.3 of CI , for any measure μ~𝔐~\widetilde{\mu}\in\widetilde{\mathfrak{M}}, there exists a sequence of closed measures μ~nc(TM)\widetilde{\mu}_{n}\in\mathbb{P}_{c}(TM) (defined in Appendix A), such that μ~n\widetilde{\mu}_{n} weakly converges to μ~\widetilde{\mu} and

limn+TML𝑑μ~n=TML𝑑μ~.\lim_{n\rightarrow+\infty}\int_{TM}Ld\widetilde{\mu}_{n}=\int_{TM}Ld\widetilde{\mu}.

Moreover, for each μ~n\widetilde{\mu}_{n} there exists an absolutely continuous curve γn:t[Tn,Tn]M\gamma_{n}:t\in[-T_{n},T_{n}]\rightarrow M with Tn+T_{n}\rightarrow+\infty as n+n\rightarrow+\infty, such that

TMg𝑑μn=12TnTnTng(γn(t),γ˙n(t))𝑑t,gCc(TM,).\int_{TM}gd\mu_{n}=\frac{1}{2T_{n}}\int_{-T_{n}}^{T_{n}}g(\gamma_{n}(t),\dot{\gamma}_{n}(t))dt,\quad\forall\\ g\in C_{c}(TM,\mathbb{R}).

Therefore, for γ^n(t):=γn(t)\widehat{\gamma}_{n}(t):=\gamma_{n}(-t), we have

c(H)\displaystyle-c(H) =\displaystyle= limn+12TnTnTnL(γn(t),γ˙n(t))𝑑t\displaystyle\lim_{n\rightarrow+\infty}\frac{1}{2T_{n}}\int_{-T_{n}}^{T_{n}}L(\gamma_{n}(t),\dot{\gamma}_{n}(t))dt
=\displaystyle= limn+12TnTnTnL(γ^n(t),γ^˙n(t))𝑑t\displaystyle\lim_{n\rightarrow+\infty}\frac{1}{2T_{n}}\int_{-T_{n}}^{T_{n}}L(\widehat{\gamma}_{n}(-t),-\dot{\widehat{\gamma}}_{n}(-t))dt
=\displaystyle= limn+12TnTnTnL^(γ^n(t),γ^˙n(t))𝑑t\displaystyle\lim_{n\rightarrow+\infty}\frac{1}{2T_{n}}\int_{-T_{n}}^{T_{n}}\widehat{L}(\widehat{\gamma}_{n}(-t),\dot{\widehat{\gamma}}_{n}(-t))dt
=\displaystyle= limn+12TnTnTnL^(γ^n(s),γ^˙n(s))𝑑s\displaystyle\lim_{n\rightarrow+\infty}\frac{1}{2T_{n}}\int_{-T_{n}}^{T_{n}}\widehat{L}(\widehat{\gamma}_{n}(s),\dot{\widehat{\gamma}}_{n}(s))ds
=\displaystyle= c(H^).\displaystyle-c(\widehat{H}).

That indicates Sμ~S^{*}\widetilde{\mu} is a Mather measure for L^(x,v)\widehat{L}(x,v), where S:TMTMS:TM\rightarrow TM is a diffeomorphism defined by S(x,v)=(x,v)S(x,v)=(x,-v). Namely, we have

TMg(x,v)𝑑Sμ~(x,v):=TMg(x,v)𝑑μ~(x,v),gCc(TM,).\int_{TM}g(x,v)dS^{*}\widetilde{\mu}(x,v):=\int_{TM}g(x,-v)d\widetilde{\mu}(x,v),\quad\forall g\in C_{c}(TM,\mathbb{R}).

Due to (A) and SS=idS\circ S=id,

Mf(x)𝑑μ(x)\displaystyle\int_{M}f(x)d\mu(x) =\displaystyle= TMfπ(x,v)𝑑μ~(x,v)\displaystyle\int_{TM}f\circ\pi(x,v)d\widetilde{\mu}(x,v)
=\displaystyle= TMfπ(x,v)𝑑Sμ~(x,v)\displaystyle\int_{TM}f\circ\pi(x,-v)dS^{*}\widetilde{\mu}(x,v)
=\displaystyle= Mf(x)𝑑πSμ~(x,v)\displaystyle\int_{M}f(x)d\pi^{*}S^{*}\widetilde{\mu}(x,v)

for all fC(M,)f\in C(M,\mathbb{R}), then πSμ~=μ𝔐\pi^{*}S^{*}\widetilde{\mu}=\mu\in\mathfrak{M}. So 𝔐^=𝔐\widehat{\mathfrak{M}}=\mathfrak{M}.

(v). Due to the definition of the Peierls barrier function, we calculate

h^(y,x)\displaystyle\widehat{h}^{\infty}(y,x) =\displaystyle= lim inft+(infξCac([0,t],M)ξ(0)=yξ(t)=x0tL^(ξ(s),ξ˙(s))𝑑s+c(H)t)\displaystyle\liminf_{t\rightarrow+\infty}\left(\inf_{\begin{subarray}{c}\xi\in C^{ac}([0,t],M)\\ \xi(0)=y\\ \xi(t)=x\end{subarray}}\int_{0}^{t}\widehat{L}(\xi(s),\dot{\xi}(s))ds+c(H)t\right)
=\displaystyle= lim inft+(infγCac([0,t],M)γ(t)=yγ(0)=x0tL(γ(ts),dγ(ts)ds)𝑑s+c(H)t)\displaystyle\liminf_{t\rightarrow+\infty}\left(\inf_{\begin{subarray}{c}\gamma\in C^{ac}([0,t],M)\\ \gamma(t)=y\\ \gamma(0)=x\end{subarray}}\int_{0}^{t}L\big{(}\gamma(t-s),-\frac{d\gamma(t-s)}{ds}\big{)}ds+c(H)t\right)
=\displaystyle= lim inft+(infγCac([0,t],M)γ(0)=xγ(t)=y0tL(γ(τ),γ˙(τ))𝑑τ+c(H)t)\displaystyle\liminf_{t\rightarrow+\infty}\left(\inf_{\begin{subarray}{c}\gamma\in C^{ac}([0,t],M)\\ \gamma(0)=x\\ \gamma(t)=y\end{subarray}}\int_{0}^{t}L(\gamma(\tau),\dot{\gamma}(\tau))d\tau+c(H)t\right)
=\displaystyle= h(x,y),\displaystyle h^{\infty}(x,y),

where ξ(s):=γ(ts)\xi(s):=\gamma(t-s) for s[0,t]s\in[0,t]. ∎

Now we can propose a function

uλ+(x):=infγCac([0,+),M)γ(0)=x0+eλt(L(γ(t),γ˙(t))+c(H))𝑑t,\displaystyle u_{\lambda}^{+}(x):=-\inf_{\begin{subarray}{c}\gamma\in C^{ac}([0,+\infty),M)\\ \gamma(0)=x\end{subarray}}\int_{0}^{+\infty}e^{-\lambda t}\big{(}L(\gamma(t),\dot{\gamma}(t))+c(H)\big{)}\ dt, (6)

of which the following properties hold:

Proposition 2.2
  1. (1)

    For λ(0,1]\lambda\in(0,1], uλ+u_{\lambda}^{+} is uniformly Lipschitz and equi-bounded on MM, with the Lipschitz (resp. equi-bound) constant depending only on LL.

  2. (2)

    For every xMx\in M, there exists a forward curve γλ,x+:[0,+)M\gamma_{\lambda,x}^{+}:[0,+\infty)\rightarrow M which achieves the minimum of (6).

  3. (3)

    (Domination) uλ+u_{\lambda}^{+} is λ\lambda-dominated by LL and is denoted by uλ+λL+c(H)u_{\lambda}^{+}\prec_{-\lambda}L+c(H), i.e., for any (x,y)M×M(x,y)\in M\times M and ba0b-a\geq 0, we have

    eλbuλ+(y)eλauλ+(x)infγCac([a,b],M)γ(a)=x,γ(b)=yabeλs(L(γ,γ˙)+c(H))𝑑se^{-\lambda b}u_{\lambda}^{+}(y)-e^{-\lambda a}u_{\lambda}^{+}(x)\leq\inf_{\begin{subarray}{c}\gamma\in C^{ac}([a,b],M)\\ \gamma(a)=x,\gamma(b)=y\end{subarray}}\int_{a}^{b}e^{-\lambda s}\Big{(}L(\gamma,\dot{\gamma})+c(H)\Big{)}ds
  4. (4)

    (Calibration) For any t0t\geq 0,

    uλ+(x)=eλtuλ+(γλ,x+(t))+t0eλs(L(γλ,x+(s),γ˙λ,x+(s))+c(H))𝑑s.u_{\lambda}^{+}(x)=e^{-\lambda t}u_{\lambda}^{+}(\gamma_{\lambda,x}^{+}(t))+\int_{t}^{0}e^{-\lambda s}\Big{(}L(\gamma_{\lambda,x}^{+}(s),\dot{\gamma}_{\lambda,x}^{+}(s))+c(H)\Big{)}ds.

    Such a curve γλ,x+\gamma_{\lambda,x}^{+} is called a forward calibrated curve of uλ+u_{\lambda}^{+}.

  5. (5)

    uλ+(x)-u_{\lambda}^{+}(x) is the viscosity solution of the following symmetrical H-J equation

    λu+H^(x,xu)=c(H),λ>0.\displaystyle\lambda u+\widehat{H}(x,\partial_{x}u)=c(H),\quad\lambda>0. (7)
Remark 2.3 (forward λ\lambda-weak KAM solution)

A continuous function w:Mw:M\rightarrow\mathbb{R} is called a forward λ\lambda-weak KAM solution of (HJλ) if it satisfies item (3) and (4) of Proposition 2.2. Due to Property (5) of Proposition 2.2, such a forward λ\lambda-weak KAM solution is unique.

Proof of Proposition 2.2: (5) By a simple transformation, we can see

uλ+(x)=infξCac((,0],M)ξ(0)=x0eλt(L^(ξ(t),ξ˙(t))+c(H^))𝑑t.\displaystyle-u_{\lambda}^{+}(x)=\inf_{\begin{subarray}{c}\xi\in C^{ac}((-\infty,0],M)\\ \xi(0)=x\end{subarray}}\int^{0}_{-\infty}e^{-\lambda t}\big{(}\widehat{L}(\xi(t),\dot{\xi}(t))+c(\widehat{H})\big{)}\ dt.

Due to Appendix 2 of DFIZ , uλ+-u_{\lambda}^{+} is a viscosity solution of (7), which is unique due to the Comparison Principle.

(1) For any viscosity solution u0(x)u_{0}(x) of (HJ0), we get

u¯0(x):=u0(x)u00u0(x)+u0:=u¯0(x),xM.\underline{u}_{0}(x):=u_{0}(x)-\|u_{0}\|\leq 0\leq u_{0}(x)+\|u_{0}\|:=\bar{u}_{0}(x),\quad\forall x\in M.

Consequently, u¯0\underline{u}_{0} (resp. u¯0\bar{u}_{0}) is a subsolution (resp. supersolution) of (7). Due to the Comparison Principle again, for any λ>0\lambda>0 and any viscosity solution ωλ\omega_{\lambda} of (7) satisfies u¯0ωλu¯0\underline{u}_{0}\leq\omega_{\lambda}\leq\bar{u}_{0}. So we get the equi-boundedness of {uλ+}λ(0,1]\{u_{\lambda}^{+}\}_{\lambda\in(0,1]}.

Let γ:[0,d(x,y)]M\gamma:[0,d(x,y)]\rightarrow M be the geodesic joining yy to xx parameterized by the arc-length, where d:M×Md:M\times M\rightarrow\mathbb{R} is the Euclidean distance. For every absolute continuous curve ξ:[0,+)M\xi:[0,+\infty)\rightarrow M with ξ(0)=x\xi(0)=x, we define a curve

η(t)={γ(t),t[0,d(x,y)],ξ(td(x,y)),t[d(x,y),+).\eta(t)=\left\{\begin{aligned} &\gamma(t),\qquad\qquad\qquad t\in[0,d(x,y)],\\ &\xi(t-d(x,y)),\qquad t\in[d(x,y),+\infty).\end{aligned}\right. (8)

Then we have

uλ+(y)0+eλt(L(η(t),η˙(t))+c(H))𝑑t0d(x,y)eλt(L(γ(t),γ˙(t))+c(H))𝑑t+d(x,y)+eλt(L(ξ(td(x,y))),ξ˙(td(x,y)))+c(H))dt0d(x,y)eλt(L(γ(t),γ˙(t))+c(H))𝑑t+eλd(x,y)0+eλt(L(ξ(t)),ξ˙(t))+c(H))dt.\begin{split}-u_{\lambda}^{+}(y)&\leq\int_{0}^{+\infty}e^{-\lambda t}\big{(}L(\eta(t),\dot{\eta}(t))+c(H)\big{)}\ dt\\ &\leq\int_{0}^{d(x,y)}e^{-\lambda t}\big{(}L(\gamma(t),\dot{\gamma}(t))+c(H)\big{)}\ dt\\ &\quad+\int_{d(x,y)}^{+\infty}e^{-\lambda t}\big{(}L(\xi(t-d(x,y))),\dot{\xi}(t-d(x,y)))+c(H)\big{)}\ dt\\ &\leq\int_{0}^{d(x,y)}e^{-\lambda t}\big{(}L(\gamma(t),\dot{\gamma}(t))+c(H)\big{)}\ dt\\ &\quad+e^{\lambda d(x,y)}\int_{0}^{+\infty}e^{-\lambda t}\big{(}L(\xi(t)),{\color[rgb]{1,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{1,0,0}\dot{\xi}(t)})+c(H)\big{)}\ dt.\end{split}

By minimizing with respect to all ξCac([0,+))\xi\in C^{ac}\big{(}[0,+\infty)\big{)} with ξ(0)=x\xi(0)=x, we obtain

uλ+(y)eλd(x,y)uλ+(x)+0d(x,y)eλt(L(γ(t),γ˙(t))+c(H))𝑑t.-u_{\lambda}^{+}(y)\leq-e^{\lambda d(x,y)}u_{\lambda}^{+}(x)+\int_{0}^{d(x,y)}e^{-\lambda t}\big{(}L(\gamma(t),\dot{\gamma}(t))+c(H)\big{)}\ dt.

Therefore, we have

uλ+(x)uλ+(y)(1eλd(x,y))uλ+(x)+0d(x,y)eλt(L(γ(t),γ˙(t))+c(H))𝑑t1eλd(x,y)λ(|λuλ+(x)|+C1)(C+C1)d(x,y),\begin{split}u_{\lambda}^{+}(x)-u_{\lambda}^{+}(y)&\leq\big{(}1-e^{\lambda d(x,y)}\big{)}u_{\lambda}^{+}(x)+\int_{0}^{d(x,y)}e^{-\lambda t}\big{(}L(\gamma(t),\dot{\gamma}(t))+c(H)\big{)}\ dt\\ &\leq\frac{1-e^{\lambda d(x,y)}}{\lambda}\left(|\lambda u_{\lambda}^{+}(x)|+C_{1}\right)\leq(C+C_{1})d(x,y),\end{split}

where C:=max{|maxxML(x,0)+c(H)|,|min(x,v)TML(x,v)+c(H)}C:=\max\{|\max_{x\in M}L(x,0)+c(H)|,|\min_{(x,v)\in TM}L(x,v)+c(H)\} and C1:=max{L(x,v):xM,v1}C_{1}:=\max\{L(x,v)~{}:~{}x\in M,\|v\|\leq 1\}. By exchanging the role of xx and yy, we get the other inequality, which shows that uλ+u_{\lambda}^{+} is uniformly Lipschitz and the Lipschitz constant is independent of λ\lambda.

(2), (3) & (4) By a similar analysis as Propositions 6.2–6.3 in DFIZ , all these three items can be easily proved. ∎

3 Discounted limit of forward λ\lambda-weak KAM solutions

Recall that u^λ:=uλ+\widehat{u}_{\lambda}^{-}:=-u_{\lambda}^{+} is the unique viscosity solution of (7),

u^λ(x)=infγ(0)=x0eλt(L^(γ,γ˙)+c(H))𝑑t.\widehat{u}_{\lambda}^{-}(x)=\inf_{\gamma(0)=x}\int_{-\infty}^{0}e^{\lambda t}\big{(}\widehat{L}(\gamma,\dot{\gamma})+c(H)\big{)}dt.

So the following conclusion holds instantly:

Lemma 3.1

DFIZ As λ0+\lambda\rightarrow 0_{+}, u^λ\widehat{u}_{\lambda}^{-} converges to a unique function u^0\widehat{u}_{0}^{-} which is a viscosity solution of the following

H^(x,xu)=c(H)\displaystyle\widehat{H}(x,\partial_{x}u)=c(H) (9)

with the following two different interpretations:

  • u^0\widehat{u}_{0}^{-} is the maximal subsolution w:Mw:M\rightarrow\mathbb{R} of (9) such that for any projected Mather measure μ^𝔐^\widehat{\mu}\in\widehat{\mathfrak{M}}, Mw𝑑μ^0\int_{M}w\cdot d{\widehat{\mu}}\leq 0.

  • u^0\widehat{u}_{0}^{-} is the infimum of functions h^μ\widehat{h}_{\mu}^{\infty} defined by

    h^μ(x):=Mh^(y,x)𝑑μ^(y),μ^𝔐^.\widehat{h}_{\mu}^{\infty}(x):=\int_{M}\widehat{h}^{\infty}(y,x)d\widehat{\mu}(y),\quad\widehat{\mu}\in\widehat{\mathfrak{M}}.

Due to Lemma 2.1 and Lemma 3.1, we get

u0+:=limλ0+uλ+=limλ0+u^λ=u^0u_{0}^{+}:=\lim_{\lambda\rightarrow 0_{+}}u_{\lambda}^{+}=-\lim_{\lambda\rightarrow 0_{+}}\widehat{u}_{\lambda}^{-}=-\widehat{u}_{0}^{-}

which is uniquely established and interpreted as the following:

Lemma 3.2

u0+u_{0}^{+} is a forward 00-weak KAM solution of (HJ0).

Proof

As u^0\widehat{u}_{0}^{-} is the viscosity solution of (9), then u^0L^+c(H)\widehat{u}_{0}^{-}\prec\widehat{L}+c(H) due to Proposition 5.3 of DFIZ . On the other side, due to

U(x,t):=infγCac([0,t],M)γ(0)=x{u^0(γ(t))+t0L^(γ(τ),γ˙(τ))+c(H)dτ},t0,U(x,t):=\inf_{\begin{subarray}{c}\gamma\in C^{ac}([0,t],M)\\ \gamma(0)=x\end{subarray}}\{\widehat{u}_{0}^{-}(\gamma(-t))+\int^{0}_{-t}\widehat{L}(\gamma(\tau),\dot{\gamma}(\tau))+c(H)\mbox{d}\tau\},\quad\forall t\geq 0,

is the unique viscosity solution of the Cauchy problem

{tu+H^(x,dxu)=c(H)u(x,0)=u^0(x),t0,\left\{\begin{aligned} &\partial_{t}u+\widehat{H}(x,d_{x}u)=c(H)\\ &u(x,0)=\widehat{u}_{0}^{-}(x),\quad t\geq 0,\end{aligned}\right.

whereas u^0(x)\widehat{u}_{0}^{-}(x) is also a viscosity solution of the Cauchy problem. So it follows that U(x,t)=u^0(x)U(x,t)=\widehat{u}_{0}^{-}(x) for all xM,t>0x\in M,t>0. Hence, by the same analysis as in the proof of Proposition 6.2 of DFIZ , for any xMx\in M, there exists a curve γx:(,0]M\gamma_{x}^{-}:(-\infty,0]\rightarrow M absolutely continuous and ending with xx, such that

u^0(γx(t))u^0(γx(s))=stL^(γx(τ),γ˙x(τ))+c(H)dτ\widehat{u}_{0}^{-}(\gamma_{x}^{-}(t))-\widehat{u}_{0}^{-}(\gamma_{x}^{-}(s))=\int_{s}^{t}\widehat{L}(\gamma_{x}^{-}(\tau),\dot{\gamma}_{x}^{-}(\tau))+c(H)\mbox{d}\tau

for all st0s\leq t\leq 0. After all, u^0\widehat{u}_{0}^{-} has to be a backward 00-weak KAM solution of (9). Consequently, u0+=u^0u_{0}^{+}=-\widehat{u}_{0}^{-} has to be a forward 00-weak KAM solution of (HJ0).∎

Proof of Theorem 1.1: It’s a direct corollary from Lemma 2.1, 3.1 and 3.2.∎

Appendix A Aubry-Mather theory of nonsmooth convex Hamiltonians

As is known, the continuous, superlinear, convex H(x,p)H(x,p) has a dual Lagrangian

L(x,v):=maxpTx𝕋n{p,vH(x,p)},(x,v)TM\displaystyle L(x,v):=\max_{p\in T_{x}^{*}\mathbb{T}^{n}}\{\langle p,v\rangle-H(x,p)\},\quad(x,v)\in TM (10)

which is also continuous, superlinear and convex in vv. Consequently, for any x,yMx,y\in M and t>0t>0, the action function

ht(x,y):=infγCac([0,t],M)γ(0)=x,γ(t)=y0t(L(γ,γ˙)+c(H))𝑑s\displaystyle h^{t}(x,y):=\inf_{\begin{subarray}{c}\gamma\in C^{ac}([0,t],M)\\ \gamma(0)=x,\gamma(t)=y\end{subarray}}\int_{0}^{t}\big{(}L(\gamma,\dot{\gamma})+c(H)\big{)}ds (11)

always attains its infimum at an absolutely continuous minimizing curve γmin:[0,t]M\gamma_{\min}:[0,t]\rightarrow M due to the Tonelli Theorem. In Ma2 , the Peierls barrier function

h(x,y):=lim inft+ht(x,y)\displaystyle h^{\infty}(x,y):=\liminf_{t\rightarrow+\infty}h^{t}(x,y) (12)

is proved to be well-defined and continuous on M×MM\times M.

Definition A.1

Ma2 The projected Aubry set is defined by

𝒜:={xM:h(x,x)=0}\mathcal{A}:=\{x\in M:h^{\infty}(x,x)=0\}

Consider TMTM (resp. MM) as a measurable space and (TM)\mathbb{P}(TM) (resp. (M)\mathbb{P}(M)) by the set of all Borel probability measures on it. A measure on TMTM is denoted by μ~\widetilde{\mu}, and we remove the tilde if we project it to MM. We say that a sequence {μ~n}n\{\widetilde{\mu}_{n}\}_{n} of probability measures weakly converges to a probability measure μ~\widetilde{\mu} if

limn+TMf(x,v)𝑑μ~n(x,v)=TMf(x,v)𝑑μ~(x,v)\lim_{n\rightarrow+\infty}\int_{TM}f(x,v)d\widetilde{\mu}_{n}(x,v)=\int_{TM}f(x,v)d\widetilde{\mu}(x,v)

for any fCc(TM,)f\in C_{c}(TM,\mathbb{R}). Accordingly, the deduced probability measure μn\mu_{n} weakly converges to μ\mu, i.e.

limn+Mf(x)𝑑μn(x)\displaystyle\lim_{n\rightarrow+\infty}\int_{M}f(x)d\mu_{n}(x) :=\displaystyle:= limn+TMf(x)𝑑πμ~n(x)\displaystyle\lim_{n\rightarrow+\infty}\int_{TM}f(x)d\pi^{*}\widetilde{\mu}_{n}(x)
=\displaystyle= limn+TMfπ(x,v)𝑑μ~n(x,v)\displaystyle\lim_{n\rightarrow+\infty}\int_{TM}f\circ\pi(x,v)d\widetilde{\mu}_{n}(x,v)
=\displaystyle= TMfπ(x,v)𝑑μ~(x,v)\displaystyle\int_{TM}f\circ\pi(x,v)d\widetilde{\mu}(x,v) (14)
=\displaystyle= Mf(x)dπμ~(x)=:Mf(x)dμ(x)\displaystyle\int_{M}f(x)d\pi^{*}\widetilde{\mu}(x)=:\int_{M}f(x)d\mu(x) (15)

for any fC(M,)f\in C(M,\mathbb{R}).

Definition A.2

A probability measure μ~\widetilde{\mu} on TMTM is closed if it satisfies:

  • TM|v|𝑑μ~(x,v)<+\int_{TM}|v|d\widetilde{\mu}(x,v)<+\infty;

  • TMϕ(x),v𝑑μ~(x,v)=0\int_{TM}\langle\nabla\phi(x),v\rangle d\widetilde{\mu}(x,v)=0 for every ϕC1(M,)\phi\in C^{1}(M,\mathbb{R}).

Let’s denote by c(TM)\mathbb{P}_{c}(TM) the set of all closed measures on TMTM, then the following conclusion is proved in DFIZ :

Theorem A.3

minμ~c(TM)TML(x,v)𝑑μ~(x,v)=c(H)\min_{\widetilde{\mu}\in\mathbb{P}_{c}(TM)}\int_{TM}L(x,v)d\widetilde{\mu}(x,v)=-c(H). Moreover, the minimizer is called a Mather measure and we denote by 𝔐~\widetilde{\mathfrak{M}} the set of them. Similarly, we can project 𝔐~\widetilde{\mathfrak{M}} to 𝔐(M)\mathfrak{M}\subset\mathbb{P}(M) w.r.t. π:TMM\pi:TM\rightarrow M, which contains all the projected Mather measures.

Definition A.4

Ma The Mather set is defined by

~:=μ~𝔐~supp(μ~)¯TM\widetilde{\mathcal{M}}:=\overline{\bigcup_{\tilde{\mu}\in\widetilde{\mathfrak{M}}}supp(\tilde{\mu})}\subset TM

and the projected Mather set :=π~\mathcal{M}:=\pi\widetilde{\mathcal{M}} is accordingly defined.

Definition A.5 (subsolution)

A function u:Mu:M\rightarrow\mathbb{R} is called a viscosity subsolution, or subsolution for short of

H(x,dxu)=c,xMH(x,d_{x}u)=c,\quad x\in M

(denoted by uL+cu\prec L+c), if u(y)u(x)ht(x,y)+(cc(H))tu(y)-u(x)\leq h^{t}(x,y)+(c-c(H))t for all (x,y)M×M(x,y)\in M\times M and t0t\geq 0.

Definition A.6

A function u:Mu:M\rightarrow\mathbb{R} is called a backward (resp. forward) 00-weak KAM solution of (HJ0) if it satisfies:

  • uL+c(H)u\prec L+c(H), i.e. for any two points (x,y)M×M(x,y)\in M\times M and any absolutely continuous curve γ:[a,b]M\gamma:[a,b]\rightarrow M connecting them, we have

    u(y)u(x)ab(L(γ,γ˙)+c(H))𝑑t.u(y)-u(x)\leq\int_{a}^{b}\big{(}L(\gamma,\dot{\gamma})+c(H)\big{)}dt.
  • for any xMx\in M there exists a curve γx:(,0]M\gamma_{x}^{-}:(-\infty,0]\rightarrow M (resp. γx+:[0,+)M\gamma_{x}^{+}:[0,+\infty)\rightarrow M) ending with (resp. starting from) xx, such that for any s<t0s<t\leq 0 (resp. 0s<t0\leq s<t),

    u(γx(t))u(γx(s))=st(L(γx,γ˙x)+c(H))𝑑tu(\gamma_{x}^{-}(t))-u(\gamma_{x}^{-}(s))=\int_{s}^{t}\big{(}L(\gamma_{x}^{-},\dot{\gamma}_{x}^{-})+c(H)\big{)}dt
    (resp.u(γx+(t))u(γx+(s))=st(L(γx+,γ˙x+)+c(H))dt).\bigg{(}resp.\quad u(\gamma_{x}^{+}(t))-u(\gamma_{x}^{+}(s))=\int_{s}^{t}\big{(}L(\gamma_{x}^{+},\dot{\gamma}_{x}^{+})+c(H)\big{)}dt\bigg{)}.

References

  • (1) Crandall, M. G., Evans, L. C., & Lions, P.-L. Some properties of viscosity solutions of Hamilton-Jacobi equations. Trans. Amer. Math. Soc. 282 (1984), no. 2 487-502.
  • (2) Contreras G. & Iturriaga R., Global minimizers of autonomous Lagrangians, 22 Colóquio Brasileiro de Matemática. In: 22nd Brazilian Mathematics Colloquium, Instituto de Matemática Pura e Aplicada (IMPA), Rio de Janeiro (1999).
  • (3) Gonzalo Contreras & Gabriel P. Paternain, Connecting orbits between static classes for generic Lagrangian systems, Topology 41 (2002), no. 4, 645-666.
  • (4) Contreras G, Iturriaga R, Paternain G.P. & Paternain M. Lagrangian graphs, minimizing measures and Mañé’s critical values. Geometric and Functional Analysis, 8(1998), no. 5, 788-809.
  • (5) Davini A, Fathi A, Itturiaga R. & Zavidovique M., Convergence of the solutions of the discounted Hamilton-Jacobi equation, Invent. Math. 206 (2016), no. 1, 29-55.
  • (6) Davini A. & Siconolfi A., A generalized dynamical approach to the large time behavior of solutions of Hamilton-Jacobi equations, SIAM J. Math. Anal., Vol. 38, No.2, pp478-502, 2006.
  • (7) Fathi A., weak KAM theorems in Lagrangian dynamics, version 10 unpublished, 2008.
  • (8) Lions P.-L., Papanicolaou G. & Varadhan S. Homogenization of Hamilton-Jacobi equation, unpublished preprint, (1987)
  • (9) Mañé R., Generic properties and problems of minimizing measures of Lagrangian systems. Nonlinearity 9 (1996), no. 2, 273-310.
  • (10) Mather J., Action minimizing invariant measures for positive definite Lagrangian systems. Math. Z. 207 (1991), no. 2, 169-207.
  • (11) Mather J., Variational construction of connecting orbits. Ann. Inst. Fourier (Grenoble) 43 (1993), no. 5, 1349-1386.