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

Viscosity subsoltions of Hamilton-Jacobi equations and Invariant sets of contact Hamilton systems

Xiang Shu    Jun Yan    Kai Zhao
Abstract

The objective of this paper is to present some results about viscosity subsolutions of the contact Hamiltonian-Jacobi equations on connected, closed manifold MM

H(x,xu,u)=0,xM.H(x,\partial_{x}u,u)=0,\quad x\in M.

Based on implicit variational principles introduced in [12, 14], we focus on the monotonicity of the solution semigroups on viscosity subsolutions and the positive invariance of the epigraph for viscosity subsolutions. Besides, we show a similar consequence for strict viscosity subsolutions on MM.

Keywords. Hamilton-Jacobi equations, viscosity subsolutions, weak KAM theory

Xiang Shu: School of Mathematical Sciences, Fudan University, Shanghai 200433, China; e-mail: [email protected]
      Jun Yan: School of Mathematical Sciences, Fudan University, Shanghai 200433, China; e-mail: [email protected]
      Kai Zhao: School of Mathematical Sciences, Fudan University, Shanghai 200433, China; e-mail: zhao_\_[email protected]
Mathematics Subject Classification (2020): 37J51; 35F21; 35D40

1 Introduction

Suppose MM is a closed (i.e.,compact, without boundary) connected and smooth manifold. Let H(x,p,u)H(x,p,u): TM×T^{*}M\times\mathbb{R}\rightarrow\mathbb{R} be a C3C^{3} function satisfying:

  • (H1)

    Positive Definiteness: 2Hp2(x,p,u)\frac{\partial^{2}H}{\partial p^{2}}(x,p,u) is positive definite on TM×T^{*}M\times\mathbb{R};

  • (H2)

    Superlinearity: For every (x,u)M×(x,u)\in M\times\mathbb{R}, lim|p|+H(x,p,u)|p|=+\lim_{|p|\to+\infty}\frac{H(x,p,u)}{|p|}=+\infty;

  • (H3)

    Lipschitz Continuity: There exists a constant λ>0\lambda>0 such that |Hu(x,p,u)|λ\Big{|}\frac{\partial H}{\partial u}(x,p,u)\Big{|}\leqslant\lambda for any (x,p,u)TM×(x,p,u)\in T^{*}M\times\mathbb{R};

We consider the viscosity solutions of the following first-order partial differential equation

H(x,xu,u)=0,xM.H(x,\partial_{x}u,u)=0,\quad x\in M. (HJs)

Let J1(M,)J^{1}(M,\mathbb{R}) denote the manifold of 1-jet of functions on MM. The standard contact form on J1(M,)J^{1}(M,\mathbb{R}) is the 1-form α=dupdx\alpha=du-pdx. Every C2C^{2} function H(x,u,p)H(x,u,p) determinates a unique vector field XHX_{H} defined by the conditions

XHα=Huα,α(XH)=H.\mathcal{L}_{X_{H}}\alpha=-\frac{\partial H}{\partial u}\alpha,\quad\alpha(X_{H})=-H.

where XH\mathcal{L}_{X_{H}} denotes the Lie derivative along the contact vector field XHX_{H}. In Darboux coordinates, the contact vector field XHX_{H} generated by HH is formulated by:

XH:{x˙=Hp(x,p,u),p˙=Hx(x,p,u)Hu(x,p,u)p,(x,p,u)TM×,u˙=Hp(x,p,u)pH(x,p,u).X_{H}:\left\{\begin{aligned} \dot{x}&=\frac{\partial H}{\partial p}(x,p,u),\\ \dot{p}&=-\frac{\partial H}{\partial x}(x,p,u)-\frac{\partial H}{\partial u}(x,p,u)\cdot p,\quad(x,p,u)\in T^{*}M\times\mathbb{R},\\ \dot{u}&=\frac{\partial H}{\partial p}(x,p,u)\cdot p-H(x,p,u).\end{aligned}\right. (1.1)

HH is called a contact Hamiltonian, and XHX_{H} is called a contact Hamiltonian vector field. Due to the fundamental theorems for ordinary differential equations , for each (x,p,u)TM×(x,p,u)\in T^{*}M\times\mathbb{R}, there exists a unique integral curve of (1.1) through the point, which is denoted by ΦHt(x,p,u)\Phi_{H}^{t}(x,p,u). In this paper, we need a extra condition on ΦHt\Phi_{H}^{t}.

  • (SH1)

    Completeness: Every maximal integral curve of the contact Hamiltonian flow ΦHt\Phi_{H}^{t} has all of \mathbb{R} as its domain of definition.

Remark 1.1.

In Appendix A, we can give a common sufficient condition (A.1) to ensure (SH1) holds.

Actually, it is common to find non-trivial examples satisfying all of our assumptions. For instance,

Example 1.2.

Consider the C3C^{3}-smooth Hamiltonian H:TM×H:T^{*}M\times\mathbb{R}\to\mathbb{R} defined by

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

where A(x)A(x) is positive definite and V(x,u):M×V(x,u):M\times\mathbb{R}\to\mathbb{R} satisfies |Vu(x,u)|λ|\frac{\partial V}{\partial u}(x,u)|\leqslant\lambda for any (x,u)M×(x,u)\in M\times\mathbb{R}.

In this paper, we assume that H(x,p,u)H(x,p,u) satisfies (H1)-(H3) and (SH1). Let L(x,x˙,u)L(x,\dot{x},u) be the Legendre transform of H(x,p,u)H(x,p,u),i.e.,

L(x,x˙,u)=suppTxM{px˙H(x,p,u)}.L(x,\dot{x},u)=\sup_{p\in T_{x}^{*}M}\{p\cdot\dot{x}-H(x,p,u)\}.

Let us recall two semigroups of operators introduced in [13]. Define a family of nonlinear operators {Tt}t0\{T^{-}_{t}\}_{t\geqslant 0} from C(M,)C(M,\mathbb{R}) to itself as follows. For each φC(M,)\varphi\in C(M,\mathbb{R}), we denote the unique continuous function on (x,t)M×[0,+)(x,t)\in M\times[0,+\infty) by (x,t)Ttφ(x)(x,t)\mapsto T^{-}_{t}\varphi(x) satisfying that

Ttφ(x)=infγ{φ(γ(0))+0tL(γ(τ),γ˙(τ),Tτφ(γ(τ)))𝑑τ},T^{-}_{t}\varphi(x)=\inf_{\gamma}\left\{\varphi(\gamma(0))+\int_{0}^{t}L(\gamma(\tau),\dot{\gamma}(\tau),T^{-}_{\tau}\varphi(\gamma(\tau)))d\tau\right\}, (1.2)

where the infimum is taken among the absolutely continuous curves γ:[0,t]M\gamma:[0,t]\to M with γ(t)=x\gamma(t)=x. It was also proved in [13] that {Tt}t0\{T^{-}_{t}\}_{t\geqslant 0} is a semigroup of operators and the function (x,t)Ttφ(x)(x,t)\mapsto T^{-}_{t}\varphi(x) is a viscosity solution of the evolutionary first-order partial differential equation

{tu+H(x,xu,u)=0,(x,t)M×(0,),u(x,0)=φ(x),xM,\begin{cases}\partial_{t}u+H(x,\partial_{x}u,u)=0,\quad(x,t)\in M\times(0,\infty),\\ u(x,0)=\varphi(x),\quad x\in M,\end{cases} (HJe)

We call {Tt}t0\{T^{-}_{t}\}_{t\geqslant 0} the backward solution semigroup.

Similarly, one can define another semigroup of operators {Tt+}t0\{T^{+}_{t}\}_{t\geqslant 0}, called the forward solution semigroup, by

Tt+φ(x)=supγ{φ(γ(t))0tL(γ(τ),γ˙(τ),Ttτ+φ(γ(τ)))𝑑τ},T^{+}_{t}\varphi(x)=\sup_{\gamma}\left\{\varphi(\gamma(t))-\int_{0}^{t}L(\gamma(\tau),\dot{\gamma}(\tau),T^{+}_{t-\tau}\varphi(\gamma(\tau)))d\tau\right\},

where the supremum is taken among the absolutely continuous curves γ:[0,t]M\gamma:[0,t]\to M with γ(0)=x\gamma(0)=x.

It is well known that (HJs) admits a viscosity solution through Perron’s method following Ishii[9], and the fundamental idea of the proof is to find a special subset of viscosity subsolutions. Actually, viscosity subsolutions decide most of the properties of viscosity solutions, and we wish to clarify the properties of viscosity subsolutions. Recently, there are some dynamical results on the Aubry-Mather theory and the weak KAM theory for contact Hamiltonian systems [13, 12, 14, 15, 10],where variational principles [13, 5, 4] played essential roles.

In this paper, our aim is to state some necessary and sufficient conditions for the viscosity subsolutions. To be specific, we can show that the viscosity subsolution is completely decided by a positive invariant set under the contact Hamiltonian flow XHX_{H}, or equivalently by the monotonicity along the solution semigroups Tt,Tt+T^{-}_{t},T^{+}_{t} , namely

Theorem A.

Suppose that φ(x)C(M,)\varphi(x)\in C(M,\mathbb{R}), the following statements are equivalent:

  1. (1)

    φ(x)\varphi(x) is a viscosity subsoltion of equation (HJs), i.e. H(x,p,φ(x))0H(x,p,\varphi(x))\leqslant 0 for any xM,pD+φ(x)x\in M,p\in D^{+}\varphi(x).

  2. (2)

    Ttφ(x)φ(x)T^{-}_{t}\varphi(x)\geqslant\varphi(x) , for any t0,xMt\geqslant 0,x\in M;

  3. (3)

    Tt+φ(x)φ(x)T^{+}_{t}\varphi(x)\leqslant\varphi(x) for any t0,xMt\geqslant 0,x\in M;

  4. (4)

    ΦHt(Γφ)Γφ\Phi^{t}_{H}(\Gamma_{\varphi})\subset\Gamma_{\varphi} for any t0t\geqslant 0, where Γφ:={(x,p,u),uφ(x)}\Gamma_{\varphi}:=\{(x,p,u),u\geqslant\varphi(x)\}.

In [7], a function φ(x)C(M,)\varphi(x)\in C(M,\mathbb{R}) is called a viscosity subsolution of (HJs)\eqref{eq:intro_HJs} is strict at x0Mx_{0}\in M if there exists an open neighborhood Vx0V_{x_{0}} of x0x_{0}, and cx0<0c_{x_{0}}<0 such that u|Vx0u|{V}_{x_{0}} is a viscosity subsolution of H(x,dxu)=cx0H(x,d_{x}u)=c_{x_{0}} on Vx0V_{x_{0}}. We define that φ(x)\varphi(x) is a strict viscosity subsolution of (HJs)\eqref{eq:intro_HJs} on MM if φ(x)\varphi(x) is a viscosity subsolution of (HJs)\eqref{eq:intro_HJs} which is strict at each xMx\in M. In Theorem B, we give the necessary and sufficient conditions for the strict viscosity subsolution .

Theorem B.

Suppose that φ(x)C(M,)\varphi(x)\in C(M,\mathbb{R}), the following statements are equivalent:

  1. (1)

    φ(x)\varphi(x) is a strict viscosity subsoltion of equation (HJs) on MM.

  2. (2)

    Ttφ(x)>φ(x)T_{t}^{-}\varphi(x)>\varphi(x) for any t0,xMt\geqslant 0,x\in M and there exists c>0c>0 such that

    limt0+1t(Ttφ(x)φ(x))c,xM.\quad\lim_{t\to 0^{+}}\frac{1}{t}\Big{(}T_{t}^{-}\varphi(x)-\varphi(x)\Big{)}\geqslant c,\quad\forall x\in M. (1.3)
  3. (3)

    Tt+φ(x)<φ(x)T_{t}^{+}\varphi(x)<\varphi(x) for any t0,xMt\geqslant 0,x\in M and there exists c>0c>0 such that

    limt0+1t(Tt+φ(x)φ(x))c,xM.\lim_{t\to 0^{+}}\frac{1}{t}\Big{(}T_{t}^{+}\varphi(x)-\varphi(x)\Big{)}\leqslant-c,\quad\forall x\in M. (1.4)
Corollary C.

Ttφ(x)>φ(x)T_{t}^{-}\varphi(x)>\varphi(x) for any t0,xMt\geqslant 0,x\in M if and only if ΦHt(Γφ)Γ̊φ\Phi^{t}_{H}(\Gamma_{\varphi})\subset\mathring{\Gamma}_{\varphi} for any t>0t>0.

The rest of this paper is organized as follows. We give some preliminaries and prove Theorem A in Section 2. The proof of Theorem B and Corollary C are given in Section 3.

2 Proof of theorem A

In this section, we first recall the definition of the viscosity solution of equation(HJs) and implicit action functions and some properties of them. Then we give the proof of theorem A.

2.1 Preliminaries

Definition 2.1 (Viscosity solutions of equation (HJs)).

Let UU be an open subset UMU\subset M.

  • (1)

    A function u:Uu:U\to\mathbb{R} is called a viscosity subsolution of equation (HJs), if for every C1C^{1} function ϕ:U\phi:U\rightarrow\mathbb{R} and every point x0Ux_{0}\in U such that uϕu-\phi has a local maximum at x0x_{0}, we have H(x0,dxϕ(x0),u(x0))c;H(x_{0},d_{x}\phi(x_{0}),u(x_{0}))\leqslant c;

  • (2)

    A function u:Uu:U\to\mathbb{R} is called a viscosity supersolution of equation (HJs), if for every C1C^{1} function ψ:U\psi:U\rightarrow\mathbb{R} and every point y0Uy_{0}\in U such that uψu-\psi has a local minimum at y0y_{0}, we have H(y0,dxψ(y0),u(y0))c;H(y_{0},d_{x}\psi(y_{0}),u(y_{0}))\geqslant c;

  • (3)

    A function u:Uu:U\to\mathbb{R} is called a viscosity solution of equation (HJs), if it is both a viscosity subsolution and a viscosity supersolution.

We recall that for any xMx\in M and continuous function uu, the closed convex sets

Du(x)={pTxM:liminfyxu(y)u(x)p,yx|yx|0},\displaystyle D^{-}u(x)=\Big{\{}p\in T^{*}_{x}M:\lim\inf_{y\to x}\frac{u(y)-u(x)-\langle p,y-x\rangle}{|y-x|}\geqslant 0\Big{\}},
D+u(x)={pTxM:limsupyxu(y)u(x)p,yx|yx|0}.\displaystyle D^{+}u(x)=\Big{\{}p\in T^{*}_{x}M:\lim\sup_{y\to x}\frac{u(y)-u(x)-\langle p,y-x\rangle}{|y-x|}\leqslant 0\Big{\}}.

are called the subdifferential and superdifferential of uu at xx , respectively, see [2, 3, 6, 8] for more details on this notion and its relationship with viscosity solutions. For instance, it is proved in [6, Prop 3.1.7] that pD+u(x)p\in D^{+}u(x) if and only if p=Dϕ(x)p=D\phi(x) for some C1C^{1} function ϕ\phi such that uϕu-\phi attains a local maximum at xx. Thus, we can present the other notion of the viscosity subsolution(supersolution) as

H(x,p,u)0(0),pD+u(x)(Du(x)).H(x,p,u)\leqslant 0(\ \geqslant 0\ ),\quad\forall p\in D^{+}u(x)\ (\ D^{-}u(x)\ ).

Let us recall two implicit action functions introduced in [12][14] which can give the representation formulae for TtT_{t}^{-} and Tt+T_{t}^{+} by [14, Proposition 4.1]:

Ttφ(x)=infyMhy,φ(y)(x,t),Tt+φ(x)=supyMhy,φ(y)(x,t),(x,t)M×(0,+),T^{-}_{t}\varphi(x)=\inf_{y\in M}h_{y,\varphi(y)}(x,t),\quad T^{+}_{t}\varphi(x)=\sup_{y\in M}h^{y,\varphi(y)}(x,t),\quad(x,t)\in M\times(0,+\infty), (2.1)

where the continuous functions

hx0,u0(x,t):M××M×(0,+),hx0,u0(x,t):M××M×(0,+)\displaystyle h_{x_{0},u_{0}}(x,t):M\times\mathbb{R}\times M\times(0,+\infty)\to\mathbb{R},\qquad h^{x_{0},u_{0}}(x,t):M\times\mathbb{R}\times M\times(0,+\infty)\to\mathbb{R}
(x0,u0,x,t)hx0,u0(x,t)(x0,u0,x,t)hx0,u0(x,t)\displaystyle(x_{0},u_{0},x,t)\mapsto h_{x_{0},u_{0}}(x,t)\qquad\qquad\qquad\qquad(x_{0},u_{0},x,t)\mapsto h^{x_{0},u_{0}}(x,t)

were introduced in Proposition 2.2 and 2.3, called forward and backward implicit action functions respectively. There are various properties of implicit action functions in [12, 13, 14].

Proposition 2.2.

[12] Properties of backward implicit action function hx0,u0(x,t)h_{x_{0},u_{0}}(x,t):

  1. (1)

    (Minimality) Given x0,xMx_{0},x\in M and u0u_{0}\in\mathbb{R} and t>0t>0,let Sx0,u0x,tS^{x,t}_{x_{0},u_{0}} be the set of the solutions (x(s),p(s),u(s))(x(s),p(s),u(s)) of (1.1) on [0,t][0,t] with x(0)=x0,x(t)=x,u(0)=u0x(0)=x_{0},x(t)=x,u(0)=u_{0}.Then

    hx0,u0(x,t):=inf{u(t):(x(s),p(s),u(s))Sx0,u0x,t},h_{x_{0},u_{0}}(x,t):=\inf\{u(t):(x(s),p(s),u(s))\in S^{x,t}_{x_{0},u_{0}}\}, (2.2)

    for any (x,t)M×(0,+)(x,t)\in M\times(0,+\infty).

  2. (2)

    (Monotonicity) Given x0M,u1<u2x_{0}\in M,u_{1}<u_{2}\in\mathbb{R}, we have hx0,u1(x,t)<hx0,u2(x,t)h_{x_{0},u_{1}}(x,t)<h_{x_{0},u_{2}}(x,t) for any t>0,xMt>0,x\in M.

  3. (3)

    (Lipschitz continuity)The function (x0,u0,x,t)hx0,u0(x,t)(x_{0},u_{0},x,t)\mapsto h_{x_{0},u_{0}}(x,t) is locally Lipschitz continuous on M××M×(0,+)M\times\mathbb{R}\times M\times(0,+\infty).

  4. (4)

    (Implicit variational) For any given x0Mx_{0}\in M and u0u_{0}\in\mathbb{R},

    hx0,u0(x,t)=u0+infγ(t)=xγ(0)=x00tL(γ(τ),γ˙(τ),hx0,u0(γ(τ),τ))𝑑τ,h_{x_{0},u_{0}}(x,t)=u_{0}+\inf_{\begin{subarray}{c}\gamma(t)=x\\ \gamma(0)=x_{0}\end{subarray}}\int_{0}^{t}L(\gamma(\tau),\dot{\gamma}(\tau),h_{x_{0},u_{0}}(\gamma(\tau),\tau))\ d\tau, (2.3)

    where the infimum is taken among the Lipschitz continuous curves γ:[0,t]M\gamma:[0,t]\rightarrow M and can be achieved.

Proposition 2.3.

[14] Properties of forward implicit action function hx0,u0(x,t)h^{x_{0},u_{0}}(x,t):

  1. (1)

    (Maximality) Given x0,xMx_{0},x\in M and u0u_{0}\in\mathbb{R} and t>0t>0,let Sx,tx0,u0S_{x,t}^{x_{0},u_{0}} be the set of the solutions (x(s),p(s),u(s))(x(s),p(s),u(s)) of (1.1) on [0,t][0,t] with x(0)=x,x(t)=x0,u(t)=u0x(0)=x,x(t)=x_{0},u(t)=u_{0}. Then

    hx0,u0(x,t):=sup{u(0):(x(s),p(s),u(s))Sx,tx0,u0},h^{x_{0},u_{0}}(x,t):=\sup\{u(0):(x(s),p(s),u(s))\in S_{x,t}^{x_{0},u_{0}}\}, (2.4)

    for any (x,t)M×(0,+)(x,t)\in M\times(0,+\infty).

  2. (2)

    (Monotonicity) Given x0M,u1<u2x_{0}\in M,u_{1}<u_{2}\in\mathbb{R}, we have hx0,u1(x,t)<hx0,u2(x,t)h^{x_{0},u_{1}}(x,t)<h^{x_{0},u_{2}}(x,t) for any t>0,xMt>0,x\in M.

  3. (3)

    (Lipschitz continuity) The function (x0,u0,x,t)hx0,u0(x,t)(x_{0},u_{0},x,t)\mapsto h^{x_{0},u_{0}}(x,t) is locally Lipschitz continuous on M××M×(0,+)M\times\mathbb{R}\times M\times(0,+\infty).

  4. (4)

    (Implicit variational)For any given x0Mx_{0}\in M and u0u_{0}\in\mathbb{R},

    hx0,u0(x,t)=u0infγ(0)=xγ(t)=x00tL(γ(τ),γ˙(τ),hx0,u0(γ(τ),tτ))𝑑τ,h^{x_{0},u_{0}}(x,t)=u_{0}-\inf_{\begin{subarray}{c}\gamma(0)=x\\ \gamma(t)=x_{0}\end{subarray}}\int_{0}^{t}L(\gamma(\tau),\dot{\gamma}(\tau),h^{x_{0},u_{0}}(\gamma(\tau),t-\tau))\ d\tau,

    where the infimum is taken among the Lipschitz continuous curves γ:[0,t]M\gamma:[0,t]\rightarrow M and can be achieved.

The relation between hx0,u0(x,t)h_{x_{0},u_{0}}(x,t) and hx0,u0(x,t)h^{x_{0},u_{0}}(x,t) was shown as follows:

Proposition 2.4.

[14, Prop 3.5] (Equivalence) hx0,u0(x,t)=uhx,u(x0,t)=u0.h_{x_{0},u_{0}}(x,t)=u\Leftrightarrow h^{x,u}(x_{0},t)=u_{0}.

2.2 Proof of Theorem A

Before turning to the proof of the main Theorem A, we will need one more preliminary.

Lemma 2.5.

Given φC(M,)\varphi\in C(M,\mathbb{R}), we have

ΦHt(Γφ)ΓTtφ,t0.\Phi^{t}_{H}(\Gamma_{\varphi})\subset\Gamma_{T_{t}^{-}\varphi},\quad t\geqslant 0.
Proof.

For any (x1,p1,u1)ΦHt(Γφ)(x_{1},p_{1},u_{1})\in\Phi^{t}_{H}(\Gamma_{\varphi}), we assume that:

(x1,p1,u1)=ΦHt(x0,p0,u0),u0φ(x0)(x_{1},p_{1},u_{1})=\Phi^{t}_{H}(x_{0},p_{0},u_{0}),\quad u_{0}\geqslant\varphi(x_{0})

Thus, by (2.1) and Proposition 2.2, it have

Ttφ(x1)=\displaystyle T_{t}^{-}\varphi(x_{1})= infyMhy,φ(y)(x1,t)hx0,φ(x0)(x1,t)hx0,u0(x1,t)\displaystyle\,\inf_{y\in M}h_{y,\varphi(y)}(x_{1},t)\leqslant h_{x_{0},\varphi(x_{0})}(x_{1},t)\leqslant h_{x_{0},u_{0}}(x_{1},t)
=\displaystyle= inf{u(t):(x(s),u(s),p(s))Sx0,u0x1,t}u1.\displaystyle\,\inf\{u(t):(x(s),u(s),p(s))\in S^{x_{1},t}_{x_{0},u_{0}}\}\leqslant u_{1}.

It follows that ΦHt(Γφ)ΓTtφ\Phi^{t}_{H}(\Gamma_{\varphi})\subset\Gamma_{T_{t}^{-}\varphi} for any t0t\geqslant 0.

We get back to the proof of Theorem A and assume that φ(x)\varphi(x) is Lipschitz continuous at first.

Proof of Theorem A: (1)(2)(1)\Rightarrow(2): We claim that: If φ(x)\varphi(x) is a viscosity subsoltion of (HJs),then for any piecewise C1C^{1} curve γ:[a,b]M\gamma:[a,b]\rightarrow M,

φ(γ(b))φ(γ(a))abL(γ(t),γ˙(t),φ(γ(t)))𝑑t,\varphi(\gamma(b))-\varphi(\gamma(a))\leqslant\int_{a}^{b}L(\gamma(t),\dot{\gamma}(t),\varphi(\gamma(t)))dt,

If φ(x)\varphi(x) is differentiable at x0x_{0}, then H(x0,xφ(x0),φ(x0))0H(x_{0},\partial_{x}\varphi(x_{0}),\varphi(x_{0}))\leqslant 0. Following the method mentioned in [7, Prop 4.2.3], we can always choose a sequence of piecewise C1C^{1} curves γn:[a,b]M\gamma_{n}:[a,b]\to M, such that φ\varphi is differentiable on γn(t)\gamma_{n}(t) for almost every t[a,b]t\in[a,b], γn(a)=γ(a)\gamma_{n}(a)=\gamma(a),γn(b)=γ(b)\gamma_{n}(b)=\gamma(b) and γn\gamma_{n} converges in the C1C^{1} topology to γ\gamma, then we obtain

φ(γ(b))φ(γ(a))=lim infnabdφ(γn(t))dt𝑑t=lim infnabxφ(γn(t)),γ˙n(t)𝑑t\displaystyle\,\varphi(\gamma(b))-\varphi(\gamma(a))=\liminf_{n\to\infty}\int_{a}^{b}\frac{d\varphi(\gamma_{n}(t))}{dt}dt=\liminf_{n\to\infty}\int_{a}^{b}\langle\partial_{x}\varphi(\gamma_{n}(t)),\dot{\gamma}_{n}(t)\rangle dt
\displaystyle\leqslant lim infnabL(γn(t),γ˙n(t),φ(γn(t)))+H(γn(t),γ˙n(t),φ(γn(t)))dt\displaystyle\,\liminf_{n\to\infty}\int_{a}^{b}L(\gamma_{n}(t),\dot{\gamma}_{n}(t),\varphi(\gamma_{n}(t)))+H(\gamma_{n}(t),\dot{\gamma}_{n}(t),\varphi(\gamma_{n}(t)))dt
\displaystyle\leqslant lim infnabL(γn(t),γ˙n(t),φ(γn(t)))𝑑t\displaystyle\,\liminf_{n\to\infty}\int_{a}^{b}L(\gamma_{n}(t),\dot{\gamma}_{n}(t),\varphi(\gamma_{n}(t)))dt
=\displaystyle= abL(γ(t),γ˙(t),φ(γ(t)))𝑑t,\displaystyle\,\int_{a}^{b}L(\gamma(t),\dot{\gamma}(t),\varphi(\gamma(t)))dt,

which completes our claim.

Then we want to show TtφφT_{t}^{-}\varphi\geqslant\varphi for any t0t\geqslant 0. By contradiction, we assume that there exists x0Mx_{0}\in M and t>0t>0 such that Ttφ(x0)<φ(x0)T^{-}_{t}\varphi(x_{0})<\varphi(x_{0}).

Let γ:[0,t]M\gamma:[0,t]\to M be a minimizer of (1.2) with γ(t)=x0\gamma(t)=x_{0}, i.e.

Ttφ(x0)=φ(γ(0))+0tL(γ(s),γ˙(s),Tsφ(γ(s)))𝑑s.T_{t}^{-}\varphi(x_{0})=\varphi(\gamma(0))+\int_{0}^{t}L(\gamma(s),\dot{\gamma}(s),T_{s}^{-}\varphi(\gamma(s)))\ ds.

Let G(s)=φ(γ(s))Tsφ(γ(s))G(s)=\varphi(\gamma(s))-T_{s}^{-}\varphi(\gamma(s)), since G(0)=0G(0)=0 and G(t)>0G(t)>0, then there exists 0t0<t0\leqslant t_{0}<t such that G(t0)=0G(t_{0})=0 and G(s)>0,s(t0,t]G(s)>0,\forall s\in(t_{0},t] .

G(s)=φ(γ(s))Tsφ(γ(s))=φ(γ(s))φ(γ(t0))t0sL(γ(τ),γ˙(τ),Tτφ(γ(τ)))𝑑τ\displaystyle\,G(s)=\varphi(\gamma(s))-T_{s}^{-}\varphi(\gamma(s))=\varphi(\gamma(s))-\varphi(\gamma(t_{0}))-\int_{t_{0}}^{s}L(\gamma(\tau),\dot{\gamma}(\tau),T_{\tau}^{-}\varphi(\gamma(\tau)))\ d\tau
\displaystyle\leqslant t0sL(γ(τ),γ˙(τ),φ(γ(τ)))L(γ(τ),γ˙(τ),Tτφ(γ(τ)))dτλt0sG(τ)𝑑τ,\displaystyle\,\int_{t_{0}}^{s}L(\gamma(\tau),\dot{\gamma}(\tau),\varphi(\gamma(\tau)))-L(\gamma(\tau),\dot{\gamma}(\tau),T_{\tau}^{-}\varphi(\gamma(\tau)))\ d\tau\leqslant\lambda\int_{t_{0}}^{s}G(\tau)\ d\tau,

where λ\lambda is the Lipschitz constant of LL with respect to uu. From Gronwall inequality ,it follows that G(s)0G(s)\leqslant 0 for any s[t0,t]s\in[t_{0},t]. It derives a contradiction, which follows that TtφφT_{t}^{-}\varphi\geqslant\varphi for any t0t\geqslant 0.

(2)(1)(2)\Rightarrow(1): Since TtφφT_{t}^{-}\varphi\geqslant\varphi for any t0t\geqslant 0, by (2.1), we have

Ttφ(y)=infxMhx,φ(x)(y,t)φ(y).T_{t}^{-}\varphi(y)=\inf_{x\in M}h_{x,\varphi(x)}(y,t)\geqslant\varphi(y).

For any fixed xMx\in M and any C1C^{1} curve ξ:[0,t]M\xi:[0,t]\to M with ξ(0)=x\xi(0)=x and ξ(t)=y\xi(t)=y, due to (2.3), we have

φ(y)hx,φ(x)(y,t)=φ(x)+infγ(t)=yγ(0)=x0tL(γ(τ),γ˙(τ),hx,φ(x)(γ(τ),τ))𝑑τ,φ(x)+0tL(ξ(τ),ξ˙(τ),hx,φ(x)(ξ(τ),τ))𝑑τ.\begin{split}\varphi(y)\leqslant&\,h_{x,\varphi(x)}(y,t)=\varphi(x)+\inf_{\begin{subarray}{c}\gamma(t)=y\\ \gamma(0)=x\end{subarray}}\int_{0}^{t}L(\gamma(\tau),\dot{\gamma}(\tau),h_{x,\varphi(x)}(\gamma(\tau),\tau))\ d\tau,\\ \leqslant&\,\varphi(x)+\int_{0}^{t}L(\xi(\tau),\dot{\xi}(\tau),h_{x,\varphi(x)}(\xi(\tau),\tau))\ d\tau.\end{split} (2.5)

For any differentiable point xMx\in M of φ\varphi,

limt0+1t[φ(ξ(t))φ(ξ(0))]\displaystyle\lim_{t\rightarrow 0^{+}}\frac{1}{t}[\varphi(\xi(t))-\varphi(\xi(0))]\leqslant limt0+1t0tL(ξ(s),ξ˙(s),hx,φ(x)(ξ(s),s))𝑑s,\displaystyle\,\lim_{t\rightarrow 0^{+}}\frac{1}{t}\int_{0}^{t}L(\xi(s),\dot{\xi}(s),h_{x,\varphi(x)}(\xi(s),s))\ ds,

which leads to

xφ(x),ξ˙(0)L(ξ(0),ξ˙(0),hx,φ(x)(ξ(0),0))=L(x,ξ˙(0),φ(x)).\langle\partial_{x}\varphi(x),\dot{\xi}(0)\rangle\leqslant L(\xi(0),\dot{\xi}(0),h_{x,\varphi(x)}(\xi(0),0))=L(x,\dot{\xi}(0),\varphi(x)).

By taking ξ˙(0)=pH(x,φ(x),xφ(x))\dot{\xi}(0)=\partial_{p}H(x,\varphi(x),\partial_{x}\varphi(x)), we get

L(x,ξ˙(0),φ(x))+H(x,xφ(x),φ(x))=xφ(x),ξ˙(0),L(x,\dot{\xi}(0),\varphi(x))+H(x,\partial_{x}\varphi(x),\varphi(x))=\langle\partial_{x}\varphi(x),\dot{\xi}(0)\rangle,

which implies H(x,xφ(x),φ(x))0.H(x,\partial_{x}\varphi(x),\varphi(x))\leqslant 0.

Since φ\varphi is Lipschitz on MM, φ\varphi is differentiable almost everywhere, then φ(x):M\varphi(x):M\rightarrow\mathbb{R} is an almost everywhere subsolution. As a result, it has to be a viscosity subsolution of (HJs), where the equivalence between almost everywhere subsolutions and viscosity subsolutions was proved in bunch of references [2, 3, 11].

(2)(4)(2)\Rightarrow(4): From Lemma 2.5, we have ΦHt(Γφ)ΓTtφ\Phi^{t}_{H}(\Gamma_{\varphi})\subset\Gamma_{T_{t}^{-}\varphi} for any t0t\geqslant 0. Due to (2), it get ΓTtφΓφ\Gamma_{T_{t}^{-}\varphi}\subset\Gamma_{\varphi} for any t0t\geqslant 0. Hence, ΦHt(Γφ)Γφ\Phi^{t}_{H}(\Gamma_{\varphi})\subset\Gamma_{\varphi} for any t0t\geqslant 0.

(4)(2)(4)\Rightarrow(2): For any given xMx\in M and t0t\geqslant 0, by Lemma 2.5, there exists pTxMp\in T_{x}M such that

(x,p,u)ΦHt(Γφ), and u=Ttφ(x).(x,p,u)\in\Phi^{t}_{H}(\Gamma_{\varphi}),\text{ and }u=T^{-}_{t}\varphi(x).

Since (x,p,u)ΦHt(Γφ)Γφ(x,p,u)\in\Phi^{t}_{H}(\Gamma_{\varphi})\subset\Gamma_{\varphi}, it follows that Ttφ(x)=uφ(x)T^{-}_{t}\varphi(x)=u\geqslant\varphi(x) for any t0,xMt\geqslant 0,x\in M.

(2)(3)(2)\Rightarrow(3): Given t>0t>0,due to (2.1), it gets that

Ttφ(y)=infxMhx,φ(x)(y,t)φ(y)yM,T^{-}_{t}\varphi(y)=\inf_{x\in M}h_{x,\varphi(x)}(y,t)\geqslant\varphi(y)\quad\forall y\in M,

which implies that hx,φ(x)(y,t)φ(y)h_{x,\varphi(x)}(y,t)\geqslant\varphi(y) for any x,yMx,y\in M.

Let u=hx,φ(x)(y,t)u=h_{x,\varphi(x)}(y,t), and by Proposition 2.4, if follows that hy,u(x,t)=φ(x)h^{y,u}(x,t)=\varphi(x). From (2) of Proposition 2.3, we have hy,φ(y)(x,t)φ(x)h^{y,\varphi(y)}(x,t)\leqslant\varphi(x) for any x,yMx,y\in M. Thus, by (2.1), it follows that Tt+φ(x)=supyMhy,φ(y)(x,t)φ(x)T^{+}_{t}\varphi(x)=\sup_{y\in M}h^{y,\varphi(y)}(x,t)\leqslant\varphi(x) for any xMx\in M.

(3)(2)(3)\Rightarrow(2): The proof is similar to the process above, and we omit it here. ∎

Remark 2.6.

Recall that any viscosity subsolution of (HJs) has to be Lipschitz continuous by [1, Lemma 2.2], and it is easy to verify that φ(x)\varphi(x) is Lipschitz if it satisfies (2). As the statement above, it is sufficient to assume that φ(x)\varphi(x) is continuous when we adapt it into Theorem A.

3 Proof of Theorem B and Corollary C

Proof of Theorem B: (1)(2)(1)\Rightarrow(2): By Remark 2.6, φ(x)C(M)\varphi(x)\in C(M) is a Lipschitz strict viscosity subsoltion of equation (HJs) and then for any x0Mx_{0}\in M, there exists an open neighborhood Vx0V_{x_{0}} of x0x_{0}, and cx0<0c_{x_{0}}<0 such that u|Vx0u|{V}_{x_{0}} is a viscosity subsolution of H(x,dxu)=cx0H(x,d_{x}u)=c_{x_{0}} on Vx0V_{x_{0}}.

Due to MM is compact, every covering of MM contains a finite subcollection covering, then there exists c>0c>0 such that φ(x)C(M)\varphi(x)\in C(M) is a Lipschitz strict viscosity subsoltion of (HJs) with H+cH+c.

Similar with Theorem A, we can choose a sequence of piecewise C1C^{1} curves γn:[a,b]M\gamma_{n}:[a,b]\to M, such that φ\varphi is differentiable on γn(t)\gamma_{n}(t) on [a,b][a,b], and γn\gamma_{n} converges in the C1C^{1} topology to γ\gamma, then we obtain

φ(γ(b))φ(γ(a))=limn+φ(γn(b))φ(γn(a))=limn+abdxφ(γn(t)),γ˙n(t)𝑑tlimn+abL(γn(t),γ˙n(t),φ(γn(t)))+H(γn(t),γ˙n(t),φ(γn(t)))dtlimn+abL(γn(t),γ˙n(t),φ(γn(t)))cdt=abL(γ(t),γ˙(t),φ(γ(t)))cdt.\begin{split}&\,\varphi(\gamma(b))-\varphi(\gamma(a))=\lim_{n\to+\infty}\varphi(\gamma_{n}(b))-\varphi(\gamma_{n}(a))=\lim_{n\to+\infty}\int_{a}^{b}\langle d_{x}\varphi(\gamma_{n}(t)),\dot{\gamma}_{n}(t)\rangle dt\\ \leqslant&\,\lim_{n\to+\infty}\int_{a}^{b}L(\gamma_{n}(t),\dot{\gamma}_{n}(t),\varphi(\gamma_{n}(t)))+H(\gamma_{n}(t),\dot{\gamma}_{n}(t),\varphi(\gamma_{n}(t)))dt\\ \leqslant&\,\lim_{n\to+\infty}\int_{a}^{b}L(\gamma_{n}(t),\dot{\gamma}_{n}(t),\varphi(\gamma_{n}(t)))-c\ dt\\ =&\,\int_{a}^{b}L(\gamma(t),\dot{\gamma}(t),\varphi(\gamma(t)))-c\ dt.\end{split} (3.1)

By Theorem A, φ(x)\varphi(x) is a viscosity subsoltion of equation (HJs) and TtφφT_{t}^{-}\varphi\geqslant\varphi for any t0t\geqslant 0. We claim that

Ttφ(x)φ(x)+c1eλtλxM,t[0,+).T_{t}^{-}\varphi(x)\geqslant\varphi(x)+c\cdot\frac{1-e^{-\lambda t}}{\lambda}\quad\forall x\in M,t\in[0,+\infty). (3.2)

By contradiction, we assume that there exists xMx\in M and t>0t>0 such that Ttφ(x)<φ(x)+c1eλtλT^{-}_{t}\varphi(x)<\varphi(x)+c\cdot\frac{1-e^{-\lambda t}}{\lambda}.Due to (1.2), there exists γ:[0,t]M\gamma:[0,t]\to M be a minimizer of (1.2) with γ(t)=x\gamma(t)=x, i.e.

Ttφ(x)=φ(γ(0))+0tL(γ(τ),γ˙(τ),Tτφ(γ(τ)))𝑑τ.T_{t}^{-}\varphi(x)=\varphi(\gamma(0))+\int_{0}^{t}L(\gamma(\tau),\dot{\gamma}(\tau),T_{\tau}^{-}\varphi(\gamma(\tau)))\ d\tau. (3.3)

Let G(s)=φ(γ(s))+c1eλsλTsφ(γ(s))G(s)=\varphi(\gamma(s))+c\cdot\frac{1-e^{-\lambda s}}{\lambda}-T_{s}^{-}\varphi(\gamma(s)), since G(0)=0G(0)=0 and G(t)>0G(t)>0, then there exists 0t0<t0\leqslant t_{0}<t such that G(t0)=0G(t_{0})=0 and G(s)>0G(s)>0 for any s(t0,t]s\in(t_{0},t]. According to (3.1) and (3.3), it shows that

G(s)=\displaystyle G(s)= G(s)G(t0)\displaystyle\,G(s)-G(t_{0})
=\displaystyle= φ(γ(s))+c1eλsλTsφ(γ(s))φ(γ(t0))c1eλt0λ+Tt0φ(γ(t0))\displaystyle\,\varphi(\gamma(s))+c\cdot\frac{1-e^{-\lambda s}}{\lambda}-T_{s}^{-}\varphi(\gamma(s))-\varphi(\gamma(t_{0}))-c\cdot\frac{1-e^{-\lambda t_{0}}}{\lambda}+T_{t_{0}}^{-}\varphi(\gamma(t_{0}))
=\displaystyle= φ(γ(s))φ(γ(t0))t0sL(γ(τ),γ˙(τ),Tτφ(γ(τ)))𝑑τ+ceλt0eλsλ\displaystyle\,\varphi(\gamma(s))-\varphi(\gamma(t_{0}))-\int_{t_{0}}^{s}L(\gamma(\tau),\dot{\gamma}(\tau),T_{\tau}^{-}\varphi(\gamma(\tau)))\ d\tau+c\cdot\frac{e^{-\lambda t_{0}}-e^{-\lambda s}}{\lambda}
\displaystyle\leqslant t0s[L(γ(τ),γ˙(τ),φ(γ(τ)))L(γ(τ),γ˙(τ),Tτφ(γ(τ)))]𝑑τc(st0)+ceλt0eλsλ\displaystyle\,\int_{t_{0}}^{s}\Big{[}L(\gamma(\tau),\dot{\gamma}(\tau),\varphi(\gamma(\tau)))-L(\gamma(\tau),\dot{\gamma}(\tau),T_{\tau}^{-}\varphi(\gamma(\tau)))\Big{]}\ d\tau-c(s-t_{0})+c\cdot\frac{e^{-\lambda t_{0}}-e^{-\lambda s}}{\lambda}
\displaystyle\leqslant λt0s|Tτφ(γ(τ))φ(γ(τ))|𝑑τc(st0)+ceλt0eλsλ\displaystyle\,\lambda\int_{t_{0}}^{s}\Big{|}T_{\tau}^{-}\varphi(\gamma(\tau))-\varphi(\gamma(\tau))\Big{|}\ d\tau-c(s-t_{0})+c\cdot\frac{e^{-\lambda t_{0}}-e^{-\lambda s}}{\lambda}
<\displaystyle< λt0sc1eλτλ𝑑τc(st0)+ceλt0eλsλ=0.\displaystyle\,\lambda\int_{t_{0}}^{s}c\cdot\frac{1-e^{-\lambda\tau}}{\lambda}d\tau-c(s-t_{0})+c\cdot\frac{e^{-\lambda t_{0}}-e^{-\lambda s}}{\lambda}=0.

where λ\lambda is the Lipschitz constant of LL with respect to uu. It derives a contradiction, which follows that

Ttφ(x)φ(x)+c1eλtλ>φ(x)xM,t(0,+).T_{t}^{-}\varphi(x)\geqslant\varphi(x)+c\cdot\frac{1-e^{-\lambda t}}{\lambda}>\varphi(x)\quad\forall x\in M,t\in(0,+\infty).

Moreover, we have

limt0+1t(Ttφ(x)φ(x))limt0+c1eλtλt=c>0,xM.\lim_{t\to 0^{+}}\frac{1}{t}\Big{(}T_{t}^{-}\varphi(x)-\varphi(x)\Big{)}\geqslant\lim_{t\to 0^{+}}c\cdot\frac{1-e^{-\lambda t}}{\lambda t}=c>0,\quad\forall x\in M.

(2)(1)(2)\Rightarrow(1): Due to (2.1),

limt0+1t(hy,φ(y)(x,t)φ(x))limt0+1t(Ttφ(x)φ(x))c>0,xM.\lim_{t\to 0^{+}}\frac{1}{t}\Big{(}h_{y,\varphi(y)}(x,t)-\varphi(x)\Big{)}\geqslant\lim_{t\to 0^{+}}\frac{1}{t}\Big{(}T_{t}^{-}\varphi(x)-\varphi(x)\Big{)}\geqslant c>0,\quad\forall x\in M. (3.4)

For any fixed xMx\in M and any C1C^{1} curve ξ:[0,t]M\xi:[0,t]\to M with ξ(0)=y\xi(0)=y and ξ(t)=x\xi(t)=x, due to (2.3), we have

hy,φ(y)(x,t)=\displaystyle h_{y,\varphi(y)}(x,t)= φ(y)+infγ(t)=xγ(0)=y0tL(γ(τ),γ˙(τ),hy,φ(y)(γ(τ),τ))𝑑τ,\displaystyle\,\varphi(y)+\inf_{\begin{subarray}{c}\gamma(t)=x\\ \gamma(0)=y\end{subarray}}\int_{0}^{t}L(\gamma(\tau),\dot{\gamma}(\tau),h_{y,\varphi(y)}(\gamma(\tau),\tau))\ d\tau,
\displaystyle\leqslant φ(y)+0tL(ξ(τ),ξ˙(τ),hx,φ(x)(ξ(τ),τ))𝑑τ.\displaystyle\,\varphi(y)+\int_{0}^{t}L(\xi(\tau),\dot{\xi}(\tau),h_{x,\varphi(x)}(\xi(\tau),\tau))\ d\tau.

For any differentiable point xMx\in M of φ\varphi,putting it into (3.4), we obtian that

c+limt0+1t[φ(ξ(t))φ(ξ(0))]\displaystyle c+\lim_{t\rightarrow 0^{+}}\frac{1}{t}[\varphi(\xi(t))-\varphi(\xi(0))]\leqslant limt0+1t0tL(ξ(s),ξ˙(s),hx,φ(x)(ξ(s),s))𝑑s,\displaystyle\,\lim_{t\rightarrow 0^{+}}\frac{1}{t}\int_{0}^{t}L(\xi(s),\dot{\xi}(s),h_{x,\varphi(x)}(\xi(s),s))\ ds,

which leads to

c+xφ(x),ξ˙(0)L(ξ(0),ξ˙(0),hx,φ(x)(ξ(0),0))=L(x,ξ˙(0),φ(x)).c+\langle\partial_{x}\varphi(x),\dot{\xi}(0)\rangle\leqslant L(\xi(0),\dot{\xi}(0),h_{x,\varphi(x)}(\xi(0),0))=L(x,\dot{\xi}(0),\varphi(x)).

By taking ξ˙(0)=pH(x,φ(x),xφ(x))\dot{\xi}(0)=\partial_{p}H(x,\varphi(x),\partial_{x}\varphi(x)), we get

L(x,ξ˙(0),φ(x))+H(x,xφ(x),φ(x))=xφ(x),ξ˙(0),L(x,\dot{\xi}(0),\varphi(x))+H(x,\partial_{x}\varphi(x),\varphi(x))=\langle\partial_{x}\varphi(x),\dot{\xi}(0)\rangle,

which implies H(x,xφ(x),φ(x))c.H(x,\partial_{x}\varphi(x),\varphi(x))\leqslant-c.

(2)(3)(2)\Leftrightarrow(3): It is similar with (2)(3)(2)\Leftrightarrow(3) of Theorem A. ∎

Proof of Corollary C: It is a strict version of Theorem A. On one hand, due to Lemma 2.5, ΦHt(Γφ)ΓTtφ\Phi_{H}^{t}(\Gamma_{\varphi})\subset\Gamma_{T_{t}^{-}\varphi}. Thus, by Ttφ>φT_{t}^{-}\varphi>\varphi, we have ΓTtφΓ̊φ\Gamma_{T_{t}^{-}\varphi}\subset\mathring{\Gamma}_{\varphi}. On the other hand, for any given xMx\in M and t>0t>0, by (2.1) and (2.2), there exists pTxMp\in T_{x}M such that (x,p,u)ΦHt(Γφ)(x,p,u)\in\Phi^{t}_{H}(\Gamma_{\varphi}) and u=Ttφ(x)u=T^{-}_{t}\varphi(x). Since (x,p,u)ΦHt(Γφ)Γ̊φ(x,p,u)\in\Phi^{t}_{H}(\Gamma_{\varphi})\subset\mathring{\Gamma}_{\varphi}, it follows that Ttφ(x)=u>φ(x).T^{-}_{t}\varphi(x)=u>\varphi(x). Hence, Ttφ(x)>φ(x)T_{t}^{-}\varphi(x)>\varphi(x) for any t>0t>0 and xMx\in M. ∎

Appendix A Completeness of the flow

Lemma A.1.

Suppose that HH satisfies (H1)-(H3) and there exists a continuous function A(h):+A(h):\mathbb{R}\to\mathbb{R}^{+} such that

|pHp(x,p,u)|A(H(x,p,u))(1+|u|),(x,p,u)TM×.\Big{|}p\cdot\frac{\partial H}{\partial p}(x,p,u)\Big{|}\leqslant A\big{(}H(x,p,u)\big{)}(1+|u|),\quad\forall(x,p,u)\in T^{*}M\times\mathbb{R}. (A.1)

Then condition (SH1) holds, i.e. the contact vector field ΦHt\Phi_{H}^{t} generates a complete flow on TM×T^{*}M\times\mathbb{R}.

Proof of Lemma A.1: Denote that H(t):=H(ΦHt(x,p,u))=H(x(t),p(t),u(t))H(t):=H(\Phi_{H}^{t}(x,p,u))=H(x(t),p(t),u(t)) for any tt\in\mathbb{R}. By (1.1), one can compute that

ddtH(ΦHt(x,p,u))=[Hxx˙+Hpp˙+Huu˙](x(t),p(t),u(t))\displaystyle\frac{d}{dt}H(\Phi_{H}^{t}(x,p,u))=\Big{[}\frac{\partial H}{\partial x}\cdot\dot{x}+\frac{\partial H}{\partial p}\cdot\dot{p}+\frac{\partial H}{\partial u}\cdot\dot{u}\Big{]}(x(t),p(t),u(t))
=\displaystyle= [HxHpHp(Hx+Hup)+Hu(Hpp˙H)](x(t),p(t),u(t))\displaystyle\,\Big{[}\frac{\partial H}{\partial x}\cdot\frac{\partial H}{\partial p}-\frac{\partial H}{\partial p}\cdot(\frac{\partial H}{\partial x}+\frac{\partial H}{\partial u}\cdot p)+\frac{\partial H}{\partial u}\cdot(\frac{\partial H}{\partial p}\cdot\dot{p}-H)\Big{]}(x(t),p(t),u(t))
=\displaystyle= Hu(x(t),p(t),u(t))H(x(t),p(t),u(t)).\displaystyle\,-\frac{\partial H}{\partial u}(x(t),p(t),u(t))\cdot H(x(t),p(t),u(t)).

which together with (H3), implies that |H(x(t),p(t),u(t))|eλ|t||H(x(0),p(0),u(0))||H(x(t),p(t),u(t))|\leqslant e^{\lambda|t|}|H(x(0),p(0),u(0))| .
From (SH1), |pHp|A(H)(1+|u|)|p\cdot\frac{\partial H}{\partial p}|\leqslant A(H)(1+|u|), then by (1.1) it obtain

|u˙|=|pHpH|A(H)(1+|u|)+|H|A(h0)|u|+A(h0)+eλ|t||H(0)|.\displaystyle|\dot{u}|=\Big{|}p\cdot\frac{\partial H}{\partial p}-H\Big{|}\leqslant A(H)(1+|u|)+|H|\leqslant A(h_{0})|u|+A(h_{0})+e^{\lambda|t|}|H(0)|.

where A(h0)=sup|h|eλ|t|H(0)A(h)A(h_{0})=\sup_{|h|\leqslant e^{\lambda|t|}H(0)}A(h). It shows that |u(s)||u(s)| is bounded on [t,t][-t,t] for any t0t\geqslant 0. If there does not exist a flow ΦHt(x,p,u)\Phi_{H}^{t}(x,p,u) for t(,+)t\in(-\infty,+\infty), it means that the solution (x(t),p(t),u(t))(x(t),p(t),u(t)) can not be extended further for some finite t0t_{0}. It implies that |p(t0)||p(t_{0})| would blow up to infinite for some finite t0t_{0}, which leads to the boundless of |H(t0)||H(t_{0})| because of (H2). It contradicts that |H(t)||H(t)| is bounded by finite tt. Therefore, the contact vector field ΦHt\Phi_{H}^{t} generates a complete flow.


Acknowledgements

Jun Yan is supported by NSFC Grant No. 11631006, 11790273.All the authors are grateful to Prof. Lin Wang for helpful suggestions.

References

  • [1] Y. Achdou, G. Barles, H. Ishii, and G. L. Litvinov. Hamilton-Jacobi equations: approximations, numerical analysis and applications, volume 2074 of Lecture Notes in Mathematics. Springer, Heidelberg; Fondazione C.I.M.E., Florence, 2013. Lecture Notes from the CIME Summer School held in Cetraro, August 29–September 3, 2011, Edited by Paola Loreti and Nicoletta Anna Tchou, Fondazione CIME/CIME Foundation Subseries.
  • [2] M. Bardi and I. Capuzzo-Dolcetta. Optimal Control and Viscosity Solutions of Hamilton-Jacobi-Bellman Equations. Systems and Control: Foundations & Applications. Birkhäuser Boston Inc., 1997.
  • [3] G. Barles. Solutions de viscosité des équations de Hamilton-Jacobi, volume 17 of Mathématiques & Applications. Springer-Verlag, Paris, 1994.
  • [4] P. Cannarsa, W. Cheng, L. Jin, K. Wang, and J. Yan. Herglotz’ variational principle and Lax-Oleinik evolution. J. Math. Pures Appl. (9), 141:99–136, 2020.
  • [5] P. Cannarsa, W. Cheng, K. Wang, and J. Yan. Herglotz’ generalized variational principle and contact type Hamilton-Jacobi equations. In Trends in control theory and partial differential equations, volume 32 of Springer INdAM Ser., pages 39–67. Springer, Cham, 2019.
  • [6] P. Cannarsa and C. Sinestrari. Semiconcave functions, Hamilton-Jacobi equations, and optimal control, volume 58 of Progress in Nonlinear Differential Equations and their Applications. Birkhäuser Boston, Inc., Boston, MA, 2004.
  • [7] A. Fathi. Weak KAM theorem in Lagrangian dynamics. Cambridge University Press, Cambridge.
  • [8] 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(6):1193–1236, 2012.
  • [9] H. Ishii. Perron’s method for hamilton-jacobi equations. Duke Mathematical Journal, 55(2):369–384, 1987.
  • [10] P. Ni, L. Wang, and J. Yan. A representation formula of the viscosity solution of the contact-type Hamilton-Jacobi equation and its applications. 01 2021.
  • [11] A. Siconolfi. Hamilton-Jacobi equations and weak KAM theory, volume 1–3. Springer, New York, 2012.
  • [12] K. Wang, L. Wang, and J. Yan. Implicit variational principle for contact Hamiltonian systems. Nonlinearity, 30(2):492–515, 2017.
  • [13] K. Wang, L. Wang, and J. Yan. Aubry–Mather Theory for Contact Hamiltonian Systems. Comm. Math. Phys., 366(3):981–1023, 2019.
  • [14] K. Wang, L. Wang, and J. Yan. Variational principle for contact Hamiltonian systems and its applications. J. Math. Pures Appl. (9), 123:167–200, 2019.
  • [15] K. Wang, L. Wang, and J. Yan. Weak kam solutions of hamilton-jacobi equations with decreasing dependence on unknown functions. Journal of Differential Equations, 286:411–432, 6 2021.