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

Topological and control theoretic properties of Hamilton-Jacobi equations via Lax-Oleinik commutators

Piermarco Cannarsa, Wei Cheng and Jiahui Hong Dipartimento di Matematica, Università di Roma “Tor Vergata”, Via della Ricerca Scientifica 1, 00133 Roma, Italy [email protected] Department of Mathematics, Nanjing University, Nanjing 210093, China [email protected] School of Mathematical Sciences, Shanghai Jiao Tong University, Shanghai 200240, China [email protected]
Abstract.

In the context of weak KAM theory, we discuss the commutators {TtTt+}t0\{T^{-}_{t}\circ T^{+}_{t}\}_{t\geqslant 0} and {Tt+Tt}t0\{T^{+}_{t}\circ T^{-}_{t}\}_{t\geqslant 0} of Lax-Oleinik operators. We characterize the relation TtTt+=IdT^{-}_{t}\circ T^{+}_{t}=Id for both small time and arbitrary time tt. We show this relation characterizes controllability for evolutionary Hamilton-Jacobi equation. Based on our previous work on the cut locus of viscosity solution, we refine our analysis of the cut time function τ\tau in terms of commutators Tt+TtTt+TtT^{+}_{t}\circ T^{-}_{t}-T^{+}_{t}\circ T^{-}_{t} and clarify the structure of the super/sub-level set of the cut time function τ\tau.

Key words and phrases:
Hamilton-Jacobi equation, cut locus, weak KAM theory, controllability
2010 Mathematics Subject Classification:
35F21, 49L25, 37J50

1. Introduction

Suppose MM is a smooth connected and compact manifold without boundary with TMTM and TMT^{*}M the tangent and cotangent bundle of MM respectively. Let HH be a Hamiltonian on MM. The study of the forward Hamilton-Jacobi equation

{Dtu(t,x)+H(x,Dxu(t,x))=0,t[0,T],xM,u(0,x)=u0(x),\begin{cases}D_{t}u(t,x)+H(x,D_{x}u(t,x))=0,\qquad t\in[0,T],x\in M,\\ u(0,x)=u_{0}(x),\end{cases} (1.1)

and the backward Hamilton-Jacobi equation

{Dtu(t,x)H(x,Dxu(t,x))=0,t[0,T],xM,u(T,x)=u0(x),\begin{cases}-D_{t}u(t,x)-H(x,D_{x}u(t,x))=0,\qquad t\in[0,T],x\in M,\\ u(T,x)=u_{0}(x),\end{cases} (1.2)

plays an important rôle in many fields such as the calculus of variation and optimal control ([5, 16]), optimal transport ([26, 1]), Hamiltonian dynamical systems and PDE ([23, 19]). In principle, the solution u(t,x)u(t,x) has a representation formula by Lax-Oleinik evolution, under very general regularity assumptions on HH and u0u_{0}. More precisely, for any ϕ:M\phi:M\to\mathbb{R}, we define the abstract Lax-Oleinik operators as

Tt+ϕ(x)=supyM{ϕ(y)ct(x,y)},Ttϕ(x)=infyM{ϕ(y)+ct(y,x)},t>0,yM,\displaystyle\begin{split}T^{+}_{t}\phi(x)=&\,\sup_{y\in M}\{\phi(y)-c_{t}(x,y)\},\\ T^{-}_{t}\phi(x)=&\,\inf_{y\in M}\{\phi(y)+c_{t}(y,x)\},\end{split}\qquad t>0,y\in M,

where ct(x,y):[0,+)×M×Mc_{t}(x,y):[0,+\infty)\times M\times M is the (dynamical) cost function. If ct(x,y)=At(x,y)c_{t}(x,y)=A_{t}(x,y) is the action of the Lagrangian LL associated with HH, then the solution of (1.1) has the representation u(t,x)=Ttu0(x)u^{-}(t,x)=T^{-}_{t}u_{0}(x) and the solution of (1.2) has the representation u+(t,x)=TTt+u0(x)u^{+}(t,x)=T^{+}_{T-t}u_{0}(x) and the uniqueness issues of (1.1) and (1.2) are well established. However, we will only touch the part for the case ct(x,y)=At(x,y)c_{t}(x,y)=A_{t}(x,y), the fundamental solution of (1.1) and (1.2), in the current paper.

In this paper, we always suppose H:TMH:T^{*}M\to\mathbb{R} is a Tonelli Hamiltonian and L:TML:TM\to\mathbb{R} is the associated Tonelli Lagrangian. We take the cost function ct(x,y)=At(x,y)c_{t}(x,y)=A_{t}(x,y) with

At(x,y)=infγΓx,yt0tL(γ,γ˙)+c[H]ds,\displaystyle A_{t}(x,y)=\inf_{\gamma\in\Gamma^{t}_{x,y}}\int^{t}_{0}L(\gamma,\dot{\gamma})+c[H]\ ds,

where c[H]c[H] is the Mañé’s critical value and Γx,yt\Gamma^{t}_{x,y} is the set of absolutely continuous curves γ:[0,t]M\gamma:[0,t]\to M connecting xx to yy. For convenience, we always suppose c[H]=0c[H]=0.

To understand the relation between the solutions of (1.1) and (1.2), we will study the Lax-Oleinik commutators {TtTt+}t0\{T^{-}_{t}\circ T^{+}_{t}\}_{t\geqslant 0} and {Tt+Tt}t0\{T^{+}_{t}\circ T^{-}_{t}\}_{t\geqslant 0} instead of the semigroups {Tt±}t0\{T^{\pm}_{t}\}_{t\geqslant 0}. A key observation is to characterize a function ϕ:M\phi:M\to\mathbb{R} satisfying the relation

TtTt+ϕ(x)=ϕ(x),t>0,xM.T^{-}_{t}\circ T^{+}_{t}\phi(x)=\phi(x),\quad t>0,x\in M. (1.3)

Our first result is

Attainable set (Theorem 3.5) : Suppose ϕSCL(M)\phi\in\text{\rm SCL}\,(M), the space of semiconcave functions with linear modulus on MM, and t>0t>0. Then the following statements are equivalent.

  1. (1)

    TtTt+ϕ=ϕT^{-}_{t}\circ T^{+}_{t}\phi=\phi.

  2. (2)

    There exists a lower semicontinuous function ψ:M\psi:M\to\mathbb{R} such that ϕ=Ttψ\phi=T^{-}_{t}\psi.

  3. (3)

    For any xMx\in M and pDϕ(x)p\in D^{*}\phi(x), let γ(s)=πxΦHs(x,p)\gamma(s)=\pi_{x}\Phi_{H}^{s}(x,p), s[t,0]s\in[-t,0]111Dϕ(x)D^{*}\phi(x) is the set of reachable gradients of ϕ\phi at xx and ΦHs\Phi_{H}^{s} is the Hamiltonian flow of HH.. Then

    Tt+ϕ(γ(t))=ϕ(x)t0L(γ,γ˙)𝑑s.\displaystyle T^{+}_{t}\phi(\gamma(-t))=\phi(x)-\int^{0}_{-t}L(\gamma,\dot{\gamma})\ ds.

It is worth noting that Theorem 3.5 can be viewed as a controllability result for equation (1.1). More precisely, our result ensures that a function ϕ\phi can be reached by some initial data ψ\psi for TtT^{-}_{t} if and only if ϕ\phi satisfies condition (1.3). We also provide a detailed description of the relation between this controllability result and underlying dynamics in Section 3.4. This kind of problems have been already addressed in the literature. For instance, in [6], for autonomous equations of the form

Dtu+H(Dxu)=0,(t,x)(0,T)×n,D_{t}u+H(D_{x}u)=0,\quad(t,x)\in(0,T)\times\mathbb{R}^{n}, (1.4)

it shown that any solution uu is of class C1C^{1} if solving the equation backward from u(T,)u(T,\cdot) one finds the value u(0,)u(0,\cdot) at t=0t=0 (compare with Theorem 3.5). As for controllability, let us recall Proposition 9 in [3] describes a subset of C(M,)C(M,\mathbb{R}) that is attainable for the negative Lax-Oleinik seigroup associated with the autonomous equation (1.4) (see also Proposition 5 in [2]) for a local version of this result that applies to more general Hamiltonians). A similar attainability result for (1.4) is obtained in [18] for nonsmooth Hamiltonians. In the case t1t\ll 1, (1.3) is closely related to a theorem by Marie-Claude Arnaud ([4]) on the evolution of the 1-graph of a semiconcave function under the Hamiltonian flow, as well as Lasry-Lions type regularization in the context of weak KAM theory as first studied by Patrick Bernard ([7]) (see Proposition 3.2).

The following result gives a new characterization of weak KAM solutions of the stationary Hamilton-Jacobi

H(x,Du(x))=0,xM.H(x,Du(x))=0,\qquad x\in M. (1.5)

In other words, weak KAM solutions are exactly those functions that are reachable for TtT^{-}_{t} for all t>0t>0.

Reversibility (Theorem 3.6) : Suppose ϕSCL(M)\phi\in\text{\rm SCL}\,(M). Then TtTt+ϕ=ϕT^{-}_{t}\circ T^{+}_{t}\phi=\phi for all t0t\geqslant 0 if and only if ϕ\phi is a weak KAM solution of (1.5).

We remark that TtTt+ϕ=ϕT^{-}_{t}\circ T^{+}_{t}\phi=\phi implies that ϕ\phi is semiconcave automatically. If the equality TtTt+ϕ=ϕT^{-}_{t}\circ T^{+}_{t}\phi=\phi holds, then there exists ψ:M\psi:M\to\mathbb{R} which is semiconvex such that

ϕ=Ttψ,ψ=Tt+ϕ.\displaystyle\phi=T^{-}_{t}\psi,\quad\psi=T^{+}_{t}\phi.

Such a pair (ϕ,ψ)(\phi,\psi) is also called an admissible Kantorovich pair (for the cost function ct(x,y)=At(x,y)c_{t}(x,y)=A_{t}(x,y)) in the theory of optimal transport (see [10, 11, 26]).

The second part of this paper is on the structure of the cut locus of weak KAM solutions to (1.5). Let uu be a weak KAM solution of (1.5). For any xMx\in M, the cut time function of uu is defined by

τ(x):=sup{t0:γC1([0,t],M),γ(0)=x,u(γ(t))u(x)=At(x,γ(t))}\displaystyle\tau(x):=\sup\{t\geqslant 0:\exists\gamma\in C^{1}([0,t],M),\gamma(0)=x,u(\gamma(t))-u(x)=A_{t}(x,\gamma(t))\}

Moreover, τ(x)\tau(x) can be related to the commutators of the Lax-Oleinik semigroups as follows (see Section 3.3)

τ(x)=sup{t0:(TtTt+Tt+Tt)u(x)=0}.\displaystyle\tau(x)=\sup\{t\geqslant 0:(T^{-}_{t}\circ T^{+}_{t}-T^{+}_{t}\circ T^{-}_{t})u(x)=0\}.

Then, we define the cut locus of uu, Cut(u)\mbox{\rm Cut}\,(u), and the Aubry set of uu, (u)\mathcal{I}\,(u), as follows

Cut(u)={x:τ(x)=0},(u)={x:τ(x)=+}.\displaystyle\mbox{\rm Cut}\,(u)=\{x:\tau(x)=0\},\qquad\mathcal{I}\,(u)=\{x:\tau(x)=+\infty\}.

In [14, 15] we proved that the complement of (u)\mathcal{I}\,(u) is homotopically equivalent to Cut(u)\mbox{\rm Cut}\,(u).

Given a weak KAM solution uu of (1.5), we define

G(u):=\displaystyle G^{*}(u):= {(x,p):xM,pDu(x)TxM},\displaystyle\,\{(x,p):x\in M,p\in D^{*}u(x)\subset T^{*}_{x}M\},
G#(u):=\displaystyle G^{\#}(u):= {(x,p):xM,pD+u(x)Du(x)TxM}.\displaystyle\,\{(x,p):x\in M,p\in D^{+}u(x)\setminus D^{*}u(x)\subset T^{*}_{x}M\}.

The following result characterizes the super/sub-level sets of the cut time function τ\tau for t>0t>0.

Super/sub-level sets of τ\tau (Theorem 3.11)

  1. (1)

    For any t>0t>0, the set {xM:τ(x)t}\{x\in M:\tau(x)\geqslant t\} is bi-Lipschitz homeomorphic to G(u)G^{*}(u).

  2. (2)

    There exists t0>0t_{0}>0 such that for all 0<t<t00<t<t_{0}, the set {xM:τ(x)<t}\{x\in M:\tau(x)<t\} is bi-Lipschitz homeomorphic to G#(u)G^{\#}(u).

Finally, we remark that these results have essential applications to our recent work on the intrinsic construction of generalized characteristics and strict singular characteristics [12]. These results clarify the relations among certain features of Hamilton-Jacobi equations such as irreversibility, non-commutativity, and singularity.

The paper is organized as follows. In section 2, we give a brief introduction to weak KAM theory and Hamilton-Jacobi equations. In Section 3, we discuss equality TtTt+ϕ=ϕT^{-}_{t}\circ T^{+}_{t}\phi=\phi for both small time and arbitrary time tt, from both functional level and underlying dynamics. We also discuss the long time behavior of the operators TtTt+T^{-}_{t}\circ T^{+}_{t}. We finally analyze the commutator of Tt+T^{+}_{t} and TtT^{-}_{t} and its implications in the structure of the cut locus.

Acknowledgements. Piermarco Cannarsa was supported in part by the National Group for Mathematical Analysis, Probability and Applications (GNAMPA) of the Italian Istituto Nazionale di Alta Matematica “Francesco Severi” and by the Excellence Department Project awarded to the Department of Mathematics, University of Rome Tor Vergata, CUP E83C23000330006. Wei Cheng is partly supported by National Natural Science Foundation of China (Grant No. 12231010). The authors also appreciate Kai Zhao for helpful discussion.

2. Preliminaries

Table 1. Notation
MM compact and connected smooth manifold without boundary
C(M)C(M) the space of continuous functions on MM
SCL(M)\text{SCL}\,(M) the class of semiconcave functions with linear modulus on MM
C1,1(M)C^{1,1}(M) the space of C1C^{1} functions on MM with Lipschitz continuous differentials
c[H]c[H] Mañé’s critical value with respect to HH
D±ϕ(x)D^{\pm}\phi(x) the superdifferential and subdifferential of a function ϕ\phi
Dϕ(x)D^{*}\phi(x) the set of reachable gradients of a function ϕ\phi
ΦHt\Phi^{t}_{H} the Hamiltonian flow associated to the Hamiltonian HH
γ1γ2¯\overline{\gamma_{1}\gamma_{2}} a curve which is the juxtaposition of two curves γ1\gamma_{1} and γ2\gamma_{2}
(u)\mathcal{I}\,(u) the Aubry set of an individual weak KAM solution uu
Cut(u)\text{Cut}\,(u) the cut locus of an individual weak KAM solution uu
Sing(ϕ)\text{Sing}\,(\phi) the set of the points of non-differentiability of a function ϕ\phi
τ(x)\tau(x) the cut time function of a given weak KAM solution

For any smooth connected and closed manifold MM, we call H=H(x,p):TMH=H(x,p):T^{*}M\to\mathbb{R} a Tonelli Hamiltonian if HH is of class C2C^{2}, the function H(x,)H(x,\cdot) is strictly convex and uniformly superlinear. The associated Lagrangian is defined by Legendre transformation:

L(x,v)=suppTxM{p,vxH(x,p)},xM,vTxM.\displaystyle L(x,v)=\sup_{p\in T^{*}_{x}M}\{\langle p,v\rangle_{x}-H(x,p)\},\quad x\in M,v\in T_{x}M.

As a real-valued function on TMTM, LL is called a Tonelli Lagrangian. That is LL is also of class C2C^{2} and the function L(x,)L(x,\cdot) is strictly convex and uniformly superlinear. From classical calculus of variation, we say that a curve γ:[a,b]M\gamma:[a,b]\to M is an extremal if it satisfies the Euler-Lagrange equations locally. We denote by {ΦHt}t\{\Phi^{t}_{H}\}_{t\in\mathbb{R}} the flow associated to the Hamiltonian vector field of HH.

Given a Tonelli Hamiltonian HH with its associated Tonelli Lagrangian LL. Mañé’s critical value of HH is the unique real number c[H]c[H] such that the Hamilton-Jacobi equation

H(x,Du(x))=c[H],xMH(x,Du(x))=c[H],\quad x\in M (2.1)

admits a viscosity solution. We always suppose c[H]=0c[H]=0 in this paper for convenience. For ϕC(M)\phi\in C(M), we define the Lax-Oleinik operators as

Tt+ϕ(x)=supyM{ϕ(y)At(x,y)},Ttϕ(x)=infyM{ϕ(y)+At(y,x)},t>0,xM,\displaystyle\begin{split}T^{+}_{t}\phi(x)=&\,\sup_{y\in M}\{\phi(y)-A_{t}(x,y)\},\\ T^{-}_{t}\phi(x)=&\,\inf_{y\in M}\{\phi(y)+A_{t}(y,x)\},\end{split}\qquad t>0,x\in M,

A function ϕ:M\phi:M\to\mathbb{R} is called a negative (resp. positive) weak KAM solution if Ttϕ=ϕT^{-}_{t}\phi=\phi (resp. Tt+ϕ=ϕT^{+}_{t}\phi=\phi) for all t0t\geqslant 0. It is known that weak KAM solutions and viscosity solutions coincide. We usually omit the adjective “negative” if we do not mention the positive one in the context. A pair of functions (u,u+)(u^{-},u^{+}) on MM is called a weak KAM pair if uu^{-} and u+u^{+} are negative and positive weak KAM solutions, respectively, and they coincide on the projected Mather set (see more in [19, Section 5.1], [20]).

Any viscosity sub-solution of (2.1) is called a critical sub-solution. The following proposition characterizes critical sub-solutions.

Proposition 2.1 ([19]).

Given ϕC(M,)\phi\in C(M,\mathbb{R}), the following properties are equivalent:

  1. (a)

    ϕ\phi is a sub-solution of (2.1).

  2. (b)

    The inequality ϕ(y)ϕ(x)At(x,y)\phi(y)-\phi(x)\leqslant A_{t}(x,y) holds for each t>0t>0 and each (x,y)M×M(x,y)\in M\times M.

  3. (c)

    The function [0,)tTtϕ(x)[0,\infty)\ni t\mapsto T^{-}_{t}\phi(x) is non-decreasing for each xMx\in M.

  4. (d)

    The function [0,)tTt+ϕ(x)[0,\infty)\ni t\mapsto T^{+}_{t}\phi(x) is non-increasing for each xMx\in M.

Let Ωn\Omega\subset\mathbb{R}^{n} be a bounded and open set, a function ϕ:Ω\phi:\Omega\to\mathbb{R} is a semiconcave function (with linear modulus) if there is a constant C>0C>0 such that

λϕ(x)+(1λ)ϕ(y)ϕ(λx+(1λ)y)C2λ(1λ)|xy|2\lambda\phi(x)+(1-\lambda)\phi(y)-\phi(\lambda x+(1-\lambda)y)\leqslant\frac{C}{2}\lambda(1-\lambda)|x-y|^{2} (2.2)

for any x,yΩx,y\in\Omega and λ[0,1]\lambda\in[0,1]. Any constant CC that satisfies the above inequality is called a semiconcavity constant for uu in Ω\Omega.

Let ϕ:Ωn\phi:\Omega\subset\mathbb{R}^{n}\to\mathbb{R} be a continuous function. We recall that, for any xΩx\in\Omega, the closed convex sets

Dϕ(x)\displaystyle D^{-}\phi(x) ={pn:lim infyxϕ(y)ϕ(x)p,yx|yx|0},\displaystyle=\left\{p\in\mathbb{R}^{n}:\liminf_{y\to x}\frac{\phi(y)-\phi(x)-\langle p,y-x\rangle}{|y-x|}\geqslant 0\right\},
D+ϕ(x)\displaystyle D^{+}\phi(x) ={pn:lim supyxϕ(y)ϕ(x)p,yx|yx|0}.\displaystyle=\left\{p\in\mathbb{R}^{n}:\limsup_{y\to x}\frac{\phi(y)-\phi(x)-\langle p,y-x\rangle}{|y-x|}\leqslant 0\right\}.

are called the subdifferential and superdifferential of ϕ\phi at xx, respectively.

Let now ϕ:Ω\phi:\Omega\to\mathbb{R} be locally Lipschitz. We recall that a vector pnp\in\mathbb{R}^{n} is said to be a reachable (or limiting) gradient of ϕ\phi at xx if there exists a sequence {xn}Ω{x}\{x_{n}\}\subset\Omega\setminus\{x\}, converging to xx, such that ϕ\phi is differentiable at xkx_{k} for each kk\in\mathbb{N} and limkDϕ(xk)=p\lim_{k\to\infty}D\phi(x_{k})=p. The set of all reachable gradients of ϕ\phi at xx is denoted by Dϕ(x)D^{\ast}\phi(x).

Proposition 2.2 ([16]).

Let ϕ:Ω\phi:\Omega\to\mathbb{R} be a continuous function. If there exists a constant C>0C>0 such that, for any xΩx\in\Omega, there exists pnp\in\mathbb{R}^{n} such that

ϕ(y)ϕ(x)+p,yx+C2|yx|2,yΩ,\phi(y)\leqslant\phi(x)+\langle p,y-x\rangle+\frac{C}{2}|y-x|^{2},\quad\forall y\in\Omega, (2.3)

then ϕ\phi is semiconcave with constant CC and pD+ϕ(x)p\in D^{+}\phi(x). Conversely, if ϕ\phi is semiconcave in Ω\Omega with constant CC, then (2.3) holds for any x,yΩx,y\in\Omega such that [x,y]Ω[x,y]\subseteq\Omega and pD+ϕ(x)p\in D^{+}\phi(x).

Proposition 2.2 is useful to understand semiconcave functions and their superdifferential on manifolds. Suppose MM is endowed with a Riemannian metric gg, with dd the associated Riemannian distance. We call ϕ:M\phi:M\to\mathbb{R} semiconcave with constant CC if

tϕ(x)+(1t)ϕ(y)ϕ(γ(t))t(1t)Cd2(x,y),x,yM,γΓx,y,\displaystyle t\phi(x)+(1-t)\phi(y)-\phi(\gamma(t))\leqslant t(1-t)Cd^{2}(x,y),\forall x,y\in M,\gamma\in\Gamma_{x,y},

where Γx,y\Gamma_{x,y} is the set of all geodesics γ:[0,1]M\gamma:[0,1]\to M connecting xx to yy. For any xMx\in M, we say that pxD+ϕ(x)p_{x}\in D^{+}\phi(x) if pxp_{x} is a one form such that

ϕ(y)ϕ(x)+px(γ˙(0))+C2d2(x,y),γΓx,y.\displaystyle\phi(y)\leqslant\phi(x)+p_{x}(\dot{\gamma}(0))+\frac{C}{2}d^{2}(x,y),\qquad\forall\gamma\in\Gamma_{x,y}.

One can show that this definition of semiconcavity is independent of the choice of the Riemannian metric, even though semiconcavity constants may change. The reader can refer to [26, 25, 22] for more on semiconcave and convex functions on manifolds.

The following Moreau-Yosida-Lasry-Lions regularization type result in the context of weak KAM theory is due to Patrick Bernard.

Proposition 2.3 ([7]).

Let ϕSCL(M)\phi\in\text{\rm SCL}\,(M). Then there exists t0>0t_{0}>0 such that Tt0+ϕC1,1(M)T^{+}_{t_{0}}\phi\in C^{1,1}(M).

Lemma 2.4.

Suppose f,g:nf,g:\mathbb{R}^{n}\to\mathbb{R}, fgf\geqslant g. If x0nx_{0}\in\mathbb{R}^{n} is such that f(x0)=g(x0)f(x_{0})=g(x_{0}), then D+f(x0)D+g(x0)D^{+}f(x_{0})\subset D^{+}g(x_{0}), Dg(x0)Df(x0)D^{-}g(x_{0})\subset D^{-}f(x_{0}).

Proof.

If pD+f(x0)p\in D^{+}f(x_{0}), then for any θn\theta\in\mathbb{R}^{n},

lim supt0+g(x0+tθ)g(x0)tlim supt0+f(x0+tθ)f(x0)tp,θ,\displaystyle\limsup_{t\to 0^{+}}\frac{g(x_{0}+t\theta)-g(x_{0})}{t}\leqslant\limsup_{t\to 0^{+}}\frac{f(x_{0}+t\theta)-f(x_{0})}{t}\leqslant\langle p,\theta\rangle,

which implies that pD+g(x0)p\in D^{+}g(x_{0}). Thus, D+f(x0)D+g(x0)D^{+}f(x_{0})\subset D^{+}g(x_{0}), Dg(x0)Df(x0)D^{-}g(x_{0})\subset D^{-}f(x_{0}) by a similar reasoning. ∎

Proposition 2.5.

Suppose u:nu:\mathbb{R}^{n}\to\mathbb{R} is semiconcave and v:nv:\mathbb{R}^{n}\to\mathbb{R} is semiconvex with the same constant CC. Assume uvu\geqslant v, and A={xn:u(x)=v(x)}A=\{x\in\mathbb{R}^{n}:u(x)=v(x)\}. Then

  1. (1)

    Du(x)=Dv(x)Du(x)=Dv(x), xA\forall x\in A.

  2. (2)

    Du=DvDu=Dv are 4C4C-Lipschitz on AA.

Proof.

For any xAx\in A, Lemma 2.4 implies D+u(x)D+v(x)D^{+}u(x)\subset D^{+}v(x). Since uu is semiconcave with linear modulus, we have that D+u(x)D^{+}u(x)\neq\varnothing. Thus, Dv(x)Dv(x) exists and D+u(x)D+v(x)={Dv(x)}D^{+}u(x)\subset D^{+}v(x)=\{Dv(x)\}. So, Du(x)=Dv(x)Du(x)=Dv(x). This completes the proof of (1).

We now turn to the proof of (2). For any x1,x2Ax_{1},x_{2}\in A, θn\theta\in\mathbb{R}^{n}, we have

Du(x1)Du(x2),θDv(x1)Du(x2),θ\displaystyle\,\langle Du(x_{1})-Du(x_{2}),\theta\rangle\leqslant\langle Dv(x_{1})-Du(x_{2}),\theta\rangle
\displaystyle\leqslant v(x1+θ)v(x1)+C2|θ|2u(x2+θ)+u(x2)+C2|θ|2\displaystyle\,v(x_{1}+\theta)-v(x_{1})+\frac{C}{2}|\theta|^{2}-u(x_{2}+\theta)+u(x_{2})+\frac{C}{2}|\theta|^{2}
\displaystyle\leqslant [12v(x1)12v(x1+2θ)+v(x1+θ)]+[12v(x1)12v(2x2x1)+v(x2)]\displaystyle\,\big{[}-\frac{1}{2}v(x_{1})-\frac{1}{2}v(x_{1}+2\theta)+v(x_{1}+\theta)\big{]}+\big{[}-\frac{1}{2}v(x_{1})-\frac{1}{2}v(2x_{2}-x_{1})+v(x_{2})\big{]}
+[12u(x1+2θ)+12u(2x2x1)u(x2+θ)]+C|θ|2\displaystyle\,+\big{[}\frac{1}{2}u(x_{1}+2\theta)+\frac{1}{2}u(2x_{2}-x_{1})-u(x_{2}+\theta)\big{]}+C|\theta|^{2}
\displaystyle\leqslant C2|θ|2+C2|x2x1|2+C2|x2x1θ|2+C|θ|2\displaystyle\,\frac{C}{2}|\theta|^{2}+\frac{C}{2}|x_{2}-x_{1}|^{2}+\frac{C}{2}|x_{2}-x_{1}-\theta|^{2}+C|\theta|^{2}
\displaystyle\leqslant C|x2x1|2+C|x2x1||θ|+2C|θ|2.\displaystyle\,C|x_{2}-x_{1}|^{2}+C|x_{2}-x_{1}|\cdot|\theta|+2C|\theta|^{2}.

It follows that

|Du(x1)Du(x2)|=1|x2x1|max|θ|=|x2x1|Du(x1)Du(x2),θ\displaystyle\,|Du(x_{1})-Du(x_{2})|=\frac{1}{|x_{2}-x_{1}|}\max_{|\theta|=|x_{2}-x_{1}|}\langle Du(x_{1})-Du(x_{2}),\theta\rangle
\displaystyle\leqslant 1|x2x1|(C|x2x1|2+C|x2x1||x2x1|+2C|x2x1|2)\displaystyle\,\frac{1}{|x_{2}-x_{1}|}(C|x_{2}-x_{1}|^{2}+C|x_{2}-x_{1}|\cdot|x_{2}-x_{1}|+2C|x_{2}-x_{1}|^{2})
=\displaystyle=  4C|x2x1|.\displaystyle\,4C|x_{2}-x_{1}|.

Therefore, Du=DvDu=Dv is 4C4C-Lipschitz on AA. ∎

3. Lax-Oleinik commutators, controllability and singularity

In this section we consider the Hamilton-Jacobi equation

H(x,Du(x))=0,xM.H(x,Du(x))=0,\qquad x\in M. (HJs)

3.1. Commutators of Lax-Oleinik semigroup

As in Proposition 2.1, the Lax-Oleinik semigroups {Tt±ϕ}\{T^{\pm}_{t}\phi\} have monotonicity properties in tt only when ϕ\phi is a continuous critical sub-solution of (HJs). In this section, we will consider the operators TtTt+T^{-}_{t}\circ T^{+}_{t} and Tt+TtT^{+}_{t}\circ T^{-}_{t} instead. We begin by analyzing the property

TtTt+ϕ(x)=ϕ(x),t>0,xM,\displaystyle T^{-}_{t}\circ T^{+}_{t}\phi(x)=\phi(x),\quad t>0,x\in M,

for ϕSCL(M)\phi\in\text{\rm SCL}\,(M). The study of the operators TtTt+T^{-}_{t}\circ T^{+}_{t} and Tt+TtT^{+}_{t}\circ T^{-}_{t} has already been introduced in the context of weak KAM theory in [8, 9]. However, for application to controllability and singularity issues, we need to complete the results by Bernard with some new observations.

Lemma 3.1.
  1. (1)

    If ϕ:M\phi:M\to\mathbb{R} is upper semicontinuous, then TtTt+ϕϕT^{-}_{t}\circ T^{+}_{t}\phi\geqslant\phi for all t>0t>0, and TtTt+ϕT^{-}_{t}\circ T^{+}_{t}\phi is non-decreasing with respect to tt. Moreover, for any t>0t>0 and xMx\in M,

    TtTt+ϕ(x)=ϕ(x)\displaystyle T^{-}_{t}\circ T^{+}_{t}\phi(x)=\phi(x)

    if and only if there exists γC1([0,t],M)\gamma\in C^{1}([0,t],M) with γ(t)=x\gamma(t)=x such that

    Tt+ϕ(γ(0))=ϕ(x)0tL(γ,γ˙)𝑑s.T^{+}_{t}\phi(\gamma(0))=\phi(x)-\int^{t}_{0}L(\gamma,\dot{\gamma})\ ds. (3.1)
  2. (2)

    If ϕ:M\phi:M\to\mathbb{R} is lower semicontinuous, then Tt+TtϕϕT^{+}_{t}\circ T^{-}_{t}\phi\leqslant\phi for all t>0t>0, and Tt+TtϕT^{+}_{t}\circ T^{-}_{t}\phi is non-increasing with respect to tt. Moreover, for any t>0t>0 and xMx\in M,

    Tt+Ttϕ(x)=ϕ(x)\displaystyle T^{+}_{t}\circ T^{-}_{t}\phi(x)=\phi(x)

    if and only if there exists γC1([0,t],M)\gamma\in C^{1}([0,t],M) with γ(0)=x\gamma(0)=x such that

    Ttϕ(γ(t))=ϕ(x)+0tL(γ,γ˙)𝑑s.T^{-}_{t}\phi(\gamma(t))=\phi(x)+\int^{t}_{0}L(\gamma,\dot{\gamma})\ ds. (3.2)
  3. (3)

    If ϕ:M\phi:M\to\mathbb{R} is upper semicontinuous and t>0t>0, then

    Tt+TtTt+ϕ=Tt+ϕ.\displaystyle T^{+}_{t}\circ T^{-}_{t}\circ T^{+}_{t}\phi=T^{+}_{t}\phi.

    If ϕ:M\phi:M\to\mathbb{R} is lower semicontinuous and t>0t>0, then

    TtTt+Ttϕ=Ttϕ.\displaystyle T^{-}_{t}\circ T^{+}_{t}\circ T^{-}_{t}\phi=T^{-}_{t}\phi.
Proof.

(1) Let ϕ:M\phi:M\to\mathbb{R} be upper semicontinuous. For any xMx\in M and t0t\geqslant 0 we have that

TtTt+ϕ(x)=\displaystyle T^{-}_{t}\circ T^{+}_{t}\phi(x)= infyM{Tt+ϕ(y)+At(y,x)}\displaystyle\,\inf_{y\in M}\{T^{+}_{t}\phi(y)+A_{t}(y,x)\}
=\displaystyle= infyM{supzM{ϕ(z)At(y,z)}+At(y,x)}\displaystyle\,\inf_{y\in M}\{\sup_{z\in M}\{\phi(z)-A_{t}(y,z)\}+A_{t}(y,x)\}
\displaystyle\geqslant ϕ(x)\displaystyle\,\phi(x)

by taking z=xz=x. Thus, if t2t1>0t_{2}\geqslant t_{1}>0, then

Tt2Tt2+ϕ=Tt1Tt2t1Tt2t1+Tt1+ϕTt1Tt1+ϕ.\displaystyle T^{-}_{t_{2}}\circ T^{+}_{t_{2}}\phi=T^{-}_{t_{1}}\circ T^{-}_{t_{2}-t_{1}}\circ T^{+}_{t_{2}-t_{1}}\circ T^{+}_{t_{1}}\phi\geqslant T^{-}_{t_{1}}\circ T^{+}_{t_{1}}\phi.

Now, let t>0t>0 and let xMx\in M be such that TtTt+ϕ(x)=ϕ(x)T^{-}_{t}\circ T^{+}_{t}\phi(x)=\phi(x). Then there exists a C1C^{1} curve γ:[0,t]M\gamma:[0,t]\to M, γ(t)=x\gamma(t)=x such that

ϕ(x)=TtTt+ϕ(x)=Tt+ϕ(γ(0))+0tL(γ,γ˙)𝑑s.\displaystyle\phi(x)=T^{-}_{t}\circ T^{+}_{t}\phi(x)=T^{+}_{t}\phi(\gamma(0))+\int^{t}_{0}L(\gamma,\dot{\gamma})\ ds.

Thus, (3.1) holds. Conversely, if there exists a C1C^{1} curve γ:[0,t]M\gamma:[0,t]\to M, with γ(t)=x\gamma(t)=x, which satisfies (3.1), then

ϕ(x)=Tt+ϕ(γ(0))+0tL(γ,γ˙)𝑑sTtTt+ϕ(x).\displaystyle\phi(x)=T^{+}_{t}\phi(\gamma(0))+\int^{t}_{0}L(\gamma,\dot{\gamma})\ ds\geqslant T^{-}_{t}\circ T^{+}_{t}\phi(x).

The inequality above is an equality since TtTt+ϕϕT^{-}_{t}\circ T^{+}_{t}\phi\geqslant\phi. This completes the proof of (1). The proof of (2) is similar and will be omitted.

(3) Suppose ϕ:M\phi:M\to\mathbb{R} is upper semicontinuous and t>0t>0. By (1), we have that

Tt+TtTt+ϕ=Tt+(TtTt+)ϕTt+ϕ.\displaystyle T^{+}_{t}\circ T^{-}_{t}\circ T^{+}_{t}\phi=T^{+}_{t}\circ(T^{-}_{t}\circ T^{+}_{t})\phi\geqslant T^{+}_{t}\phi.

On the other hand, property (2) implies that

Tt+TtTt+ϕ=(Tt+Tt)Tt+ϕTt+ϕ.\displaystyle T^{+}_{t}\circ T^{-}_{t}\circ T^{+}_{t}\phi=(T^{+}_{t}\circ T^{-}_{t})\circ T^{+}_{t}\phi\leqslant T_{t}^{+}\phi.

Thus, there holds Tt+TtTt+ϕ=Tt+ϕT^{+}_{t}\circ T^{-}_{t}\circ T^{+}_{t}\phi=T^{+}_{t}\phi. Similarly, if ϕ:M\phi:M\to\mathbb{R} is lower semicontinuous and t>0t>0, then TtTt+Ttϕ=TtϕT^{-}_{t}\circ T^{+}_{t}\circ T^{-}_{t}\phi=T^{-}_{t}\phi. ∎

Before proceeding further, we recall a result from [4] which describes the evolution of the 1-graph

graph(D+ϕ)={(x,p):xM,pD+ϕ(x)TM}\displaystyle\text{\rm graph}\,(D^{+}\phi)=\{(x,p):x\in M,p\in D^{+}\phi(x)\subset T^{*}M\}

under the Hamiltonian flow {ΦHt}\{\Phi_{H}^{t}\} for short time (see (3) of the following proposition).

Proposition 3.2.

Let ϕSCL(M)\phi\in\text{\rm SCL}\,(M) and let t0>0t_{0}>0 be such that Tt0+ϕC1,1(M)T^{+}_{t_{0}}\phi\in C^{1,1}(M). Then the following holds true.

  1. (1)

    Tt+ϕC1,1(M)T^{+}_{t}\phi\in C^{1,1}(M) for all t(0,t0]t\in(0,t_{0}] and TtTt+ϕ=ϕT^{-}_{t}\circ T^{+}_{t}\phi=\phi for all t[0,t0]t\in[0,t_{0}].

  2. (2)

    Tt+ϕ=Tt0tTt0+ϕT^{+}_{t}\phi=T^{-}_{t_{0}-t}\circ T^{+}_{t_{0}}\phi for all t[0,t0]t\in[0,t_{0}].

  3. (3)

    (Arnaud) We have that

    graph(DTt+ϕ)=ΦHt(graph(D+ϕ)),t(0,t0].\text{\rm graph}\,(DT^{+}_{t}\phi)=\Phi_{H}^{-t}(\text{\rm graph}\,(D^{+}\phi)),\qquad\forall t\in(0,t_{0}]. (3.3)
  4. (4)

    Let u(t,x)=Tt+ϕ(x)u(t,x)=T^{+}_{t}\phi(x) for (t,x)[0,t0]×M(t,x)\in[0,t_{0}]\times M. Then uu is of class Cloc1,1C^{1,1}_{\rm loc} on (0,t0)×M(0,t_{0})\times M and it is a viscosity solution of the Hamilton-Jacobi equation

    {DtuH(x,Dxu)=0,(t,x)(0,t0)×M;u(0,x)=ϕ(x),xM.\begin{cases}D_{t}u-H(x,D_{x}u)=0,\qquad(t,x)\in(0,t_{0})\times M;\\ u(0,x)=\phi(x),\qquad x\in M.\end{cases} (HJ+e{}_{e}^{+})
Proof.

Suppose ϕSCL(M)\phi\in\text{\rm SCL}\,(M), t0>0t_{0}>0 and Tt0+ϕC1,1(M)T^{+}_{t_{0}}\phi\in C^{1,1}(M). The fact that Tt+ϕC1,1(M)T^{+}_{t}\phi\in C^{1,1}(M) for all t(0,t0]t\in(0,t_{0}] is known ([7]). Fix xMx\in M and t(0,t0]t\in(0,t_{0}]. There exists yMy\in M such that

TtTt+ϕ(x)=Tt+ϕ(y)+At(y,x).\displaystyle T^{-}_{t}\circ T^{+}_{t}\phi(x)=T^{+}_{t}\phi(y)+A_{t}(y,x).

Being semiconcave and attaining its minimum at yy, Tt+ϕ()+At(,x)T_{t}^{+}\phi(\cdot)+A_{t}(\cdot,x) must be differentiable at yy. Since At(,x)A_{t}(\cdot,x) is itself differentiable by Proposition A.1, Tt+ϕT_{t}^{+}\phi must be differentiable at yy as well. By Fermat’s rule we have

Lv(ξ(0),ξ˙(0))=DyAt(y,x)=DTt+ϕ(y),\displaystyle L_{v}(\xi(0),\dot{\xi}(0))=-D_{y}A_{t}(y,x)=DT^{+}_{t}\phi(y),

where ξΓy,xt\xi\in\Gamma^{t}_{y,x} is the unique minimal curve for At(y,x)A_{t}(y,x). On the other hand, there exists a unique zMz\in M such that Tt+ϕ(y)=ϕ(z)At(y,z)T^{+}_{t}\phi(y)=\phi(z)-A_{t}(y,z) since Tt+ϕC1,1(M)T^{+}_{t}\phi\in C^{1,1}(M). Then,

DTt+ϕ(y)=Lv(η(0),η˙(0))\displaystyle DT^{+}_{t}\phi(y)=L_{v}(\eta(0),\dot{\eta}(0))

where ηΓy,zt\eta\in\Gamma^{t}_{y,z} is the unique minimal curve for At(y,z)A_{t}(y,z). Thus, ξ(0)=η(0)=y\xi(0)=\eta(0)=y and Lv(η(0),η˙(0))=Lv(ξ(0),ξ˙(0))L_{v}(\eta(0),\dot{\eta}(0))=L_{v}(\xi(0),\dot{\xi}(0)). Since ξ\xi and η\eta satisfy the Euler-Lagrange equation with the same initial conditions, we conclude that ξη\xi\equiv\eta, and so that z=xz=x. Therefore,

TtTt+ϕ(x)=ϕ(z)At(y,z)+At(y,x)=ϕ(x).\displaystyle T^{-}_{t}\circ T^{+}_{t}\phi(x)=\phi(z)-A_{t}(y,z)+A_{t}(y,x)=\phi(x).

and (1) follows.

By (1) we conclude that

Tt0tTt0+ϕ=Tt0tTt0t+Tt+ϕ=Tt+ϕ,t(0,t0].T^{-}_{t_{0}-t}\circ T^{+}_{t_{0}}\phi=T^{-}_{t_{0}-t}\circ T^{+}_{t_{0}-t}\circ T^{+}_{t}\phi=T^{+}_{t}\phi,\qquad t\in(0,t_{0}]. (3.4)

The proof is completed noting that Statement (3) is known (see [4]) and (4) is obvious. ∎

For any ϕSCL(M)\phi\in\text{\rm SCL}\,(M), define two functions τ1,τ2:SCL(M)[0,]\tau_{1},\tau_{2}:\text{\rm SCL}\,(M)\to[0,\infty]

τ1(ϕ):=sup{t>0:Tt+ϕC1,1(M)},τ2(ϕ):=sup{t0:TtTt+ϕ=ϕ}.\tau_{1}(\phi):=\,\sup\{t>0:T^{+}_{t}\phi\in C^{1,1}(M)\},\qquad\tau_{2}(\phi):=\,\sup\{t\geqslant 0:T^{-}_{t}\circ T^{+}_{t}\phi=\phi\}. (3.5)

Let πx:TMM\pi_{x}:T^{*}M\to M be the canonical projection onto MM.

Corollary 3.3.

Suppose ϕSCL(M)\phi\in\text{\rm SCL}\,(M).

  1. (1)

    τ2(ϕ)τ1(ϕ)>0\tau_{2}(\phi)\geqslant\tau_{1}(\phi)>0.

  2. (2)

    We have

    {t>0:Tt+ϕC1,1(M)}=\displaystyle\{t>0:T^{+}_{t}\phi\in C^{1,1}(M)\}= (0,τ1(ϕ)),\displaystyle\,(0,\tau_{1}(\phi)),
    {t0:TtTt+ϕ=ϕ}=\displaystyle\{t\geqslant 0:T^{-}_{t}\circ T^{+}_{t}\phi=\phi\}= {[0,τ2(ϕ)],τ2(ϕ)<;[0,),τ2(ϕ)=.\displaystyle\,\begin{cases}[0,\tau_{2}(\phi)],&\tau_{2}(\phi)<\infty;\\ [0,\infty),&\tau_{2}(\phi)=\infty.\end{cases}
Proof.

(1) is a consequence of Proposition 3.2, and (2) is easily follows from the definition of τ1(ϕ)\tau_{1}(\phi) and τ2(ϕ)\tau_{2}(\phi). ∎

Corollary 3.4.

Suppose ϕSCL(M)\phi\in\text{\rm SCL}\,(M), τ1(ϕ)<\tau_{1}(\phi)<\infty. Then Tτ1(ϕ)+ϕT^{+}_{\tau_{1}(\phi)}\phi is semiconvex and the following properties hold.

  1. (1)

    Tt+ϕ=Tτ1(ϕ)tTτ1(ϕ)+ϕT^{+}_{t}\phi=T^{-}_{\tau_{1}(\phi)-t}\circ T^{+}_{\tau_{1}(\phi)}\phi for all t[0,τ1(ϕ)]t\in[0,\tau_{1}(\phi)], and Tt+ϕC1,1(M)T^{+}_{t}\phi\in C^{1,1}(M) for all t(0,τ1(ϕ))t\in(0,\tau_{1}(\phi)).

  2. (2)

    graph(DTt+ϕ)=ΦHt(graph(D+ϕ))=ΦHτ1(ϕ)t(graph(DTτ1(ϕ)+ϕ))\text{\rm graph}\,(DT^{+}_{t}\phi)=\Phi_{H}^{-t}(\text{\rm graph}\,(D^{+}\phi))=\Phi_{H}^{\tau_{1}(\phi)-t}(\text{\rm graph}\,(D^{-}T^{+}_{\tau_{1}(\phi)}\phi)) for all t(0,τ1(ϕ))t\in(0,\tau_{1}(\phi)).

  3. (3)

    Let u(t,x)=Tt+ϕ(x)u(t,x)=T^{+}_{t}\phi(x) for (t,x)[0,τ1(ϕ)]×M(t,x)\in[0,\tau_{1}(\phi)]\times M. Then uu is of class Cloc1,1C^{1,1}_{\rm loc} on (0,τ1(ϕ))×M(0,\tau_{1}(\phi))\times M and it is a viscosity solution of the Hamilton-Jacobi equation

    {DtuH(x,Dxu)=0,(t,x)(0,τ1(ϕ))×M;u(0,x)=ϕ(x),xM.\displaystyle\begin{cases}D_{t}u-H(x,D_{x}u)=0,\qquad(t,x)\in(0,\tau_{1}(\phi))\times M;\\ u(0,x)=\phi(x),\qquad x\in M.\end{cases}
Proof.

The proof follows from Proposition 3.2 and the definition of τ1(ϕ)\tau_{1}(\phi) and τ2(ϕ)\tau_{2}(\phi). ∎

The following proposition characterizes the condition TtTt+ϕ=ϕT^{-}_{t}\circ T^{+}_{t}\phi=\phi for general t>0t>0.

Theorem 3.5.

Suppose ϕSCL(M)\phi\in\text{\rm SCL}\,(M) and t0>0t_{0}>0. Then the following statements are equivalent.

  1. (1)

    Tt0Tt0+ϕ=ϕT^{-}_{t_{0}}\circ T^{+}_{t_{0}}\phi=\phi.

  2. (2)

    There exists a lower semicontinuous function ψ:M\psi:M\to\mathbb{R} such that ϕ=Tt0ψ\phi=T^{-}_{t_{0}}\psi.

  3. (3)

    For any xMx\in M and pDϕ(x)p\in D^{*}\phi(x), we have that

    Tt0+ϕ(γ(t0))=ϕ(x)t00L(γ,γ˙)𝑑s.\displaystyle T^{+}_{t_{0}}\phi(\gamma(-t_{0}))=\phi(x)-\int^{0}_{-t_{0}}L(\gamma,\dot{\gamma})\ ds.

    where γ(s)=πxΦHs(x,p)\gamma(s)=\pi_{x}\Phi_{H}^{s}(x,p), s[t0,0]s\in[-t_{0},0].

Proof.

Taking ψ=Tt0+ϕ\psi=T^{+}_{t_{0}}\phi, it follows that (1) implies (2). Now, suppose there exists a lower semicontinuous function ψ:M\psi:M\to\mathbb{R} such that ϕ=Tt0ψ\phi=T^{-}_{t_{0}}\psi. By Lemma 3.1 (3) we obtain

Tt0Tt0+ϕ=Tt0Tt0+Tt0ψ=Tt0ψ=ϕ.\displaystyle T^{-}_{t_{0}}\circ T^{+}_{t_{0}}\phi=T^{-}_{t_{0}}\circ T^{+}_{t_{0}}\circ T^{-}_{t_{0}}\psi=T^{-}_{t_{0}}\psi=\phi.

The proof of the equivalence of (1) and (2) is thus complete.

Since (3)(1)(3)\Rightarrow(1) is obvious in view of Lemma 3.1, it is sufficient to prove that (1)(3)(1)\Rightarrow(3). We suppose that xMx\in M, pDϕ(x)p\in D^{*}\phi(x) and let γ(s)=πxΦHs(x,p)\gamma(s)=\pi_{x}\Phi_{H}^{s}(x,p), s[t0,0]s\in[-t_{0},0]. Since Tt0Tt0+ϕ=ϕT^{-}_{t_{0}}\circ T^{+}_{t_{0}}\phi=\phi, we have that pDTt0Tt0+ϕ(x)p\in D^{*}T^{-}_{t_{0}}\circ T^{+}_{t_{0}}\phi(x). Thus, γ\gamma is an extremal such that

ϕ(x)=Tt0Tt0+ϕ(x)=Tt0+ϕ(γ(t0))+t00L(γ,γ˙)𝑑s\displaystyle\phi(x)=T^{-}_{t_{0}}\circ T^{+}_{t_{0}}\phi(x)=T^{+}_{t_{0}}\phi(\gamma(-t_{0}))+\int^{0}_{-t_{0}}L(\gamma,\dot{\gamma})\ ds

thanks to [16, Theorem 6.4.9]. This leads to our conclusion. ∎

3.2. Long time behavior of the operators TtTt+T^{-}_{t}\circ T^{+}_{t}.

Recall that τ1\tau_{1} and τ2\tau_{2} are defined in (3.5).

Theorem 3.6.

Suppose ϕSCL(M)\phi\in\text{\rm SCL}\,(M).

  1. (1)

    τ2(ϕ)=+\tau_{2}(\phi)=+\infty if and only if ϕ\phi is a weak KAM solution of (HJs).

  2. (2)

    τ1(ϕ)=+\tau_{1}(\phi)=+\infty if and only if ϕC1,1(M)\phi\in C^{1,1}(M) and satisfies (HJs).

Proof.

Suppose ϕ\phi is a weak KAM solution of (HJs). Then Ttϕ=ϕT^{-}_{t}\phi=\phi for all t0t\geqslant 0. By Proposition 3.5 we have that TtTt+ϕ=ϕT^{-}_{t}\circ T^{+}_{t}\phi=\phi for all t0t\geqslant 0. It follows that τ2(ϕ)=+\tau_{2}(\phi)=+\infty.

Conversely, suppose τ2(ϕ)=+\tau_{2}(\phi)=+\infty. For any xMx\in M, pDϕ(x)p\in D^{*}\phi(x) let γ(s)=πxΦHs(x,p)\gamma(s)=\pi_{x}\Phi_{H}^{s}(x,p) with s(,0]s\in(-\infty,0]. By Theorem 3.5 we conclude that

Tt+ϕ(γ(t))=ϕ(x)t0L(γ,γ˙)𝑑s,t0.\displaystyle T^{+}_{t}\phi(\gamma(-t))=\phi(x)-\int^{0}_{-t}L(\gamma,\dot{\gamma})\ ds,\quad\forall t\geqslant 0.

Thus, At(γ(t),x)=t0L(γ,γ˙)𝑑sA_{t}(\gamma(-t),x)=\int^{0}_{-t}L(\gamma,\dot{\gamma})\ ds for all t0t\geqslant 0. By Lemma A.2 we have that H(x,p)=0H(x,p)=0. This implies that ϕ\phi satisfies (HJs) almost everywhere. Since HH is convex with respect to the pp-variable, then ϕ\phi is a viscosity solution of (HJs) (see [16, Proposition 5.3.1]). This completes the proof of (1).

Now, we turn to the proof of (2). If ϕC1,1(M)\phi\in C^{1,1}(M) satisfies (HJs), then Tt+ϕ=ϕC1,1(M)T^{+}_{t}\phi=\phi\in C^{1,1}(M) for all t>0t>0, which implies τ1(ϕ)=+\tau_{1}(\phi)=+\infty. If τ1(ϕ)=+\tau_{1}(\phi)=+\infty, then for any xMx\in M, pD+ϕ(x)p\in D^{+}\phi(x) let γ(s)=πxΦHs(x,p)\gamma(s)=\pi_{x}\Phi_{H}^{s}(x,p), s(,0]s\in(-\infty,0]. Theorem 3.5 leads to

Tt+ϕ(γ(t))=ϕ(x)t0L(γ,γ˙)𝑑s,t0.\displaystyle T^{+}_{t}\phi(\gamma(-t))=\phi(x)-\int^{0}_{-t}L(\gamma,\dot{\gamma})\ ds,\quad\forall t\geqslant 0.

Thus At(γ(t),x)=t0L(γ,γ˙)𝑑sA_{t}(\gamma(-t),x)=\int^{0}_{-t}L(\gamma,\dot{\gamma})\ ds for all t0t\geqslant 0. By Lemma A.2 we have that H(x,p)=0H(x,p)=0 for all pD+ϕ(x)p\in D^{+}\phi(x). Since H(x,)H(x,\cdot) is strictly convex, we deduce that D+ϕ(x)D^{+}\phi(x) is a singleton, that is, p=Dϕ(x)p=D\phi(x). Therefore, ϕ\phi is a C1C^{1} solution of (HJs). It is well known that any C1C^{1} solution ϕ\phi of (HJs) is of class C1,1C^{1,1}. This leads to our conclusion. ∎

Still assuming c[H]=0c[H]=0 we recall that the Peierls barrier in classical weak KAM theory is defined by

h(x,y)=lim inft+At(x,y),x,yM.h(x,y)=\liminf_{t\to+\infty}A_{t}(x,y),\qquad x,y\in M. (3.6)

For ϕC(M)\phi\in C(M), let

Sϕ=limt+Ttϕ,S+ϕ=limt+Tt+ϕ.\displaystyle S^{-}\phi=\lim_{t\to+\infty}T^{-}_{t}\phi,\quad S^{+}\phi=\lim_{t\to+\infty}T^{+}_{t}\phi.

The liminf in (3.6) is indeed a limit and the limits in the definition of S±S^{\pm} exist by a well known result of Fathi (see, [19, 21]).

Proposition 3.7.

Let ϕC(M)\phi\in C(M).

  1. (1)

    We have the following relations:

    Sϕ(x)=infyM{ϕ(y)+h(y,x)},S+ϕ(x)=supyM{ϕ(y)h(x,y)},xM.\displaystyle S^{-}\phi(x)=\inf_{y\in M}\{\phi(y)+h(y,x)\},\qquad S^{+}\phi(x)=\sup_{y\in M}\{\phi(y)-h(x,y)\},\qquad x\in M.
  2. (2)

    We also have

    SS+ϕ=limt+TtTt+ϕ,S+Sϕ=limt+Tt+Ttϕ.\displaystyle S^{-}S^{+}\phi=\lim_{t\to+\infty}T^{-}_{t}T^{+}_{t}\phi,\qquad S^{+}S^{-}\phi=\lim_{t\to+\infty}T^{+}_{t}T^{-}_{t}\phi.
  3. (3)

    SS+ϕϕS^{-}S^{+}\phi\geqslant\phi. For any xMx\in M, SS+ϕ(x)=ϕ(x)S^{-}S^{+}\phi(x)=\phi(x) if and only if there exists yMy\in M such that

    S+ϕ(y)=ϕ(x)h(y,x).\displaystyle S^{+}\phi(y)=\phi(x)-h(y,x).
  4. (4)

    S+SϕϕS^{+}S^{-}\phi\leqslant\phi. For any xMx\in M, S+Sϕ(x)=ϕ(x)S^{+}S^{-}\phi(x)=\phi(x) if and only if there exists yMy\in M such that

    Sϕ(y)=ϕ(x)+h(x,y).\displaystyle S^{-}\phi(y)=\phi(x)+h(x,y).
Proof.

We know from Fleming’s lemma [19, Theorem 4.4.3] that At(,)A_{t}(\cdot,\cdot) is equi-Lipschitz for t1t\geqslant 1, and the convergence in (3.6) is uniform. Thus, for any xMx\in M, we have

Sϕ(x)\displaystyle S^{-}\phi(x) =limt+Ttϕ(x)=limt+infyM{ϕ(y)+At(y,x)}\displaystyle=\lim_{t\to+\infty}T^{-}_{t}\phi(x)=\lim_{t\to+\infty}\inf_{y\in M}\{\phi(y)+A_{t}(y,x)\}
=infyMlimt+{ϕ(y)+At(y,x)}=infyM{ϕ(y)+h(y,x)}.\displaystyle=\inf_{y\in M}\lim_{t\to+\infty}\{\phi(y)+A_{t}(y,x)\}=\inf_{y\in M}\{\phi(y)+h(y,x)\}.

The proof of the second equality in (1) is similar and this completes the proof of (1).

Notice that by the non-expansive property of Lax-Oleinik semigroups ([19]), for any t>0t>0 there holds

TtS+ϕTtTt+ϕ\displaystyle\|T^{-}_{t}S^{+}\phi-T^{-}_{t}T^{+}_{t}\phi\|_{\infty} =TtS+ϕTt(Tt+ϕ)\displaystyle=\|T^{-}_{t}S^{+}\phi-T^{-}_{t}(T^{+}_{t}\phi)\|_{\infty}
S+ϕ(Tt+ϕ)0,ast+.\displaystyle\leqslant\|S^{+}\phi-(T^{+}_{t}\phi)\|_{\infty}\to 0,\quad\text{as}\ t\to+\infty.

Therefore, we have SS+ϕ=limt+TtS+ϕ=limt+TtTt+ϕS^{-}S^{+}\phi=\lim_{t\to+\infty}T^{-}_{t}S^{+}\phi=\lim_{t\to+\infty}T^{-}_{t}T^{+}_{t}\phi. The proof of the second equality of (2) is similar.

The proof of (3) and (4) is similar to that of Lemma 3.1 and will be omitted. ∎

3.3. Cut locus

Recall Mather’s barrier function B:MB:M\to\mathbb{R} introduced in [24],

B(x):=u(x)u+(x),xM,\displaystyle B(x):=u^{-}(x)-u^{+}(x),\qquad x\in M,

where (u,u+)(u^{-},u^{+}) is a weak KAM pair. We introduce an analogue of the barrier function BB. For any weak KAM solution uu of (HJs), let B:[0,+)×MB:[0,+\infty)\times M\to\mathbb{R},

B(t,x):=u(x)Tt+u(x),(t,x)[0,+)×M.\displaystyle B(t,x):=u(x)-T^{+}_{t}u(x),\qquad(t,x)\in[0,+\infty)\times M.

Due to Theorem 3.6 we have

B(t,x)=(TtTt+Tt+Tt)u(x).\displaystyle B(t,x)=(T^{-}_{t}\circ T^{+}_{t}-T^{+}_{t}\circ T^{-}_{t})u^{-}(x).

We denote by Cut(u)\mbox{\rm Cut}\,(u) the cut locus of uu. That is, xCut(u)x\in\mbox{\rm Cut}\,(u) if any calibrated curve ending at xx can not be extended further as a calibrated curve. More precisely, let τ:M\tau:M\to\mathbb{R} be the cut time function of uu, for any xMx\in M,

τ(x):=\displaystyle\tau(x):= sup{t0:γC1([0,t],M),γ(0)=x,u(γ(t))u(x)=At(x,γ(t))}.\displaystyle\,\sup\{t\geqslant 0:\exists\gamma\in C^{1}([0,t],M),\gamma(0)=x,u(\gamma(t))-u(x)=A_{t}(x,\gamma(t))\}.

Then Cut(u)={xM:τ(x)=0}\mbox{\rm Cut}\,(u)=\{x\in M:\tau(x)=0\}. Recall that

(u)={xM:there exists a u-calibrated curve γ:[0,+)M such that γ(0)=x}\displaystyle\mathcal{I}(u)=\{x\in M:\text{there exists a }u\text{-calibrated curve }\gamma:[0,+\infty)\to M\text{ such that }\gamma(0)=x\}

The following properties of the cut locus and cut time function are essentially known.

Proposition 3.8 ([14, 15]).

Suppose uu is a weak KAM solution of (HJs).

  1. (1)

    Given t>0t>0 and xMx\in M, then Tt+u(x)=u(x)T^{+}_{t}u(x)=u(x) if and only if there exists a uu-calibrated curve γ:[0,t]M\gamma:[0,t]\to M such that γ(0)=x\gamma(0)=x.

  2. (2)

    τ(x)=sup{t0:B(t,x)=0}\tau(x)=\sup\{t\geqslant 0:B(t,x)=0\} for all xMx\in M.

  3. (3)

    τ\tau is upper-semicontinuous and Cut(u)\mbox{\rm Cut}\,(u) is a GδG_{\delta}-set.

  4. (4)

    Cut(u)={x:τ(x)=0}\mbox{\rm Cut}\,(u)=\{x:\tau(x)=0\} and (u)={x:τ(x)=+}\mathcal{I}(u)=\{x:\tau(x)=+\infty\}.

Given a weak KAM pair (u,u+)(u^{-},u^{+}) of (HJs), we define

u(t,x)=Tt+u(x),(t,x)(,0]×M.\displaystyle u(t,x)=T^{+}_{-t}u^{-}(x),\qquad(t,x)\in(-\infty,0]\times M.

It is clear that u+(x)u(t,x)u(x)u^{+}(x)\leqslant u(t,x)\leqslant u^{-}(x) for all (t,x)(,0]×M(t,x)\in(-\infty,0]\times M. Define

G(u):=\displaystyle G^{*}(u^{-}):= {(x,p):xM,pDu(x)TxM},\displaystyle\,\{(x,p):x\in M,p\in D^{*}u^{-}(x)\subset T^{*}_{x}M\},
G#(u):=\displaystyle G^{\#}(u^{-}):= {(x,p):xM,pD+u(x)Du(x)TxM},\displaystyle\,\{(x,p):x\in M,p\in D^{+}u^{-}(x)\setminus D^{*}u^{-}(x)\subset T^{*}_{x}M\},

then graph(D+u)=G(u)G#(u)\text{graph}\,(D^{+}u^{-})=G^{*}(u^{-})\cup G^{\#}(u^{-}). Let

(u)=\displaystyle\mathscr{F}^{*}(u^{-})= {γ:(,0]M:γ(s)=πxΦHs(x,p),(x,p)G(u)}\displaystyle\,\{\gamma:(-\infty,0]\to M:\gamma(s)=\pi_{x}\Phi_{H}^{s}(x,p),(x,p)\in G^{*}(u^{-})\}
𝒜(u)=\displaystyle\mathcal{A}^{*}(u^{-})= {(t,x)(,0]×M:γ(u),γ(t)=x}\displaystyle\,\{(t,x)\in(-\infty,0]\times M:\exists\gamma\in\mathscr{F}^{*}(u^{-}),\gamma(t)=x\}
𝒜#(u)=\displaystyle\mathcal{A}^{\#}(u^{-})= ((,0]×M)𝒜(u).\displaystyle\,((-\infty,0]\times M)\setminus\mathcal{A}^{*}(u^{-}).
Proposition 3.9.
  1. (1)

    For any (t,x)𝒜(u)({0}×M)(t,x)\in\mathcal{A}^{*}(u^{-})\setminus(\{0\}\times M) there exists a unique γ(u)\gamma\in\mathscr{F}^{*}(u^{-}) with γ(t)=x\gamma(t)=x. Let p=Lv(γ,γ˙)p=L_{v}(\gamma,\dot{\gamma}), then we have

    u(t,x)=u(x)=u(γ(0))t0L(γ,γ˙)𝑑s,\displaystyle u(t,x)=u^{-}(x)=u^{-}(\gamma(0))-\int^{0}_{t}L(\gamma,\dot{\gamma})\ ds,
    Dxu(t,x)=Du(x)=p(t),\displaystyle D_{x}u(t,x)=Du^{-}(x)=p(t),
    Dtu(t,x)=H(x,Dxu(t,x))=0.\displaystyle D_{t}u(t,x)=-H(x,D_{x}u(t,x))=0.
  2. (2)

    For any (t,x)𝒜#(u)(t,x)\in\mathcal{A}^{\#}(u^{-}), let γ:[t,0]M\gamma:[t,0]\to M, γ(t)=x\gamma(t)=x be a maximizer for u(t,x)u(t,x), that is,

    u(t,x)=u(γ(0))t0L(γ,γ˙)𝑑s,\displaystyle u(t,x)=u^{-}(\gamma(0))-\int_{t}^{0}L(\gamma,\dot{\gamma})\ ds,

    and let p=Lv(γ,γ˙)p=L_{v}(\gamma,\dot{\gamma}). Then (γ(0),p(0))G#(u)(\gamma(0),p(0))\in G^{\#}(u^{-}), γ(0)Sing(u)\gamma(0)\in\mbox{\rm Sing}\,(u^{-}), H(x,p(t))<0H(x,p(t))<0, and

    u+(x)<u(t,x)<u(x).\displaystyle u^{+}(x)<u(t,x)<u^{-}(x).
Proof.

The uniqueness of γ\gamma follows from the fact that calibrated curves cannot cross in the interior of the time interval. Now we have

u(x)=u(γ(0))t0L(γ,γ˙)𝑑su(t,x)u(x),\displaystyle u^{-}(x)=u^{-}(\gamma(0))-\int_{t}^{0}L(\gamma,\dot{\gamma})\ ds\leqslant u(t,x)\leqslant u^{-}(x),

which implies u(t,x)=u(x)u(t,x)=u^{-}(x). Notice that uu and uu^{-} are differentiable on the interior of the calibrated curve. Thus, Dxu(t,x)=Du(x)=p(t)D_{x}u(t,x)=Du^{-}(x)=p(t). Finally, (γ(0),p(0))G(u)(\gamma(0),p(0))\in G^{*}(u^{-}) implies

Dtu(t,x)=H(x,Dxu(t,x))=H(γ(t),p(t))=H(γ(0),p(0))=0.\displaystyle D_{t}u(t,x)=-H(x,D_{x}u(t,x))=-H(\gamma(t),p(t))=-H(\gamma(0),p(0))=0.

Now, we turn to prove (2). First, by calculus of variations we know that p(0)D+u(γ(0))p(0)\in D^{+}u^{-}(\gamma(0)). The definition of 𝒜#(u)\mathcal{A}^{\#}(u^{-}) implies that p(0)Du(γ(0))p(0)\notin D^{*}u^{-}(\gamma(0)). It follows that (γ(0),p(0))G#(u)(\gamma(0),p(0))\in G^{\#}(u^{-}). Therefore, we have that γ(0)Sing(u)\gamma(0)\in\mbox{\rm Sing}\,(u^{-}), and H(x,p(t))=H(γ(0),p(0))<0H(x,p(t))=H(\gamma(0),p(0))<0. It is well known that u+(x)u(t,x)u(x)u^{+}(x)\leqslant u(t,x)\leqslant u^{-}(x). If u(t,x)=u(x)u(t,x)=u^{-}(x), then we can choose γ:[t,0]M\gamma:[t,0]\to M, γ(t)=x\gamma(t)=x to be a calibrated curve for u(t,x)u(t,x). Thus,

u(x)=u(t,x)=u(γ(0))t0L(γ,γ˙)𝑑s.\displaystyle u^{-}(x)=u(t,x)=u^{-}(\gamma(0))-\int_{t}^{0}L(\gamma,\dot{\gamma})\ ds.

So, γ\gamma is in fact a calibrated curve for uu^{-}, which implies Lv(γ(0),γ˙(0))Du(γ(0))L_{v}(\gamma(0),\dot{\gamma}(0))\in D^{*}u^{-}(\gamma(0)). As a result, (t,x)𝒜(u)(t,x)\in\mathcal{A}^{*}(u^{-}). This leads to a contradiction. So we have u(t,x)<u(x)u(t,x)<u^{-}(x). If u+(x)=u(t,x)u^{+}(x)=u(t,x), we know that (0,Du+(x))Du(t,x)=coDu(t,x)(0,D^{-}u^{+}(x))\in D^{-}u(t,x)=\mbox{\rm co}\,D^{*}u(t,x). Since H(x,p(t))<0H(x,p(t))<0 for all calibrated curves of u(t,x)u(t,x), it is easy to see that

q=H(x,p)>0,(q,p)Du(t,x).\displaystyle q=-H(x,p)>0,\qquad\forall(q,p)\in D^{*}u(t,x).

This leads to a contradiction with (0,Du+(x))coDu(t,x)(0,D^{-}u^{+}(x))\in\mbox{\rm co}\,D^{*}u(t,x). Therefore, u+(x)<u(t,x)u^{+}(x)<u(t,x). ∎

Proposition 3.10.
  1. (1)

    For any t0t\geqslant 0 and xMx\in M,

    (t,x)𝒜(u)Tt+u(x)=u(x)τ(x)t.\displaystyle(-t,x)\in\mathcal{A}^{*}(u^{-})\quad\Longleftrightarrow\quad T^{+}_{t}u^{-}(x)=u^{-}(x)\quad\Longleftrightarrow\quad\tau(x)\geqslant t.

    That is

    𝒜(u)=\displaystyle\mathcal{A}^{*}(u^{-})= {(t,x)(,0]×M:τ(x)t}\displaystyle\,\{(t,x)\in(-\infty,0]\times M:\tau(x)\geqslant-t\}
    =\displaystyle= {(t,x)(,0]×M:u(t,x)=u(x)}.\displaystyle\,\{(t,x)\in(-\infty,0]\times M:u(t,x)=u^{-}(x)\}.

    Moreover, Du=DuDu^{-}=Du is locally Lipschitz on 𝒜(u)({0}×M)\mathcal{A}^{*}(u^{-})\setminus(\{0\}\times M).

  2. (2)

    For any t0t\geqslant 0 and xMx\in M,

    (t,x)𝒜#(u)Tt+u(x)<u(x)τ(x)<t.\displaystyle(-t,x)\in\mathcal{A}^{\#}(u^{-})\quad\Longleftrightarrow\quad T^{+}_{t}u^{-}(x)<u^{-}(x)\quad\Longleftrightarrow\quad\tau(x)<t.

    That is

    𝒜#(u)=\displaystyle\mathcal{A}^{\#}(u^{-})= {(t,x)(,0]×M:τ(x)<t}\displaystyle\,\{(t,x)\in(-\infty,0]\times M:\tau(x)<-t\}
    =\displaystyle= {(t,x)(,0]×M:u(t,x)<u(x)}.\displaystyle\,\{(t,x)\in(-\infty,0]\times M:u(t,x)<u^{-}(x)\}.
Proof.

It is sufficient to only prove (1) since (2) is an immediate consequence of (1). The first equivalence follows from Proposition 3.9 and the second one from Proposition 3.8 (2). These equivalences imply that

𝒜(u)=\displaystyle\mathcal{A}^{*}(u^{-})= {(t,x)(,0]×M:τ(x)t}\displaystyle\,\{(t,x)\in(-\infty,0]\times M:\tau(x)\geqslant-t\}
=\displaystyle= {(t,x)(,0]×M:u(t,x)=u(x)}.\displaystyle\,\{(t,x)\in(-\infty,0]\times M:u(t,x)=u^{-}(x)\}.

By Proposition 2.5, we know that Du=DuDu^{-}=Du is locally Lipschitz on 𝒜(u)({0}×M)\mathcal{A}^{*}(u^{-})\setminus(\{0\}\times M). ∎

Theorem 3.11.
  1. (1)

    For any t>0t>0, there is a bi-Lipschitz homeomorphism between {xM:τ(x)t}\{x\in M:\tau(x)\geqslant t\} and G(u)G^{*}(u^{-}).

  2. (2)

    For any t(0,τ1(u)]t\in(0,\tau_{1}(u^{-})], there is a bi-Lipschitz homeomorphism between {xM:τ(x)<t}\{x\in M:\tau(x)<t\} and G#(u)G^{\#}(u^{-}).

Proof.

For any t>0t>0, Proposition 3.10 (1) implies there is a C1C^{1} diffeomorphism between {(x,Du(x)):τ(x)t}\{(x,Du^{-}(x)):\tau(x)\geqslant t\} and ΦHt(G(u))\Phi_{H}^{-t}(G^{*}(u^{-})), and the projection

πx:{(x,Du(x)):τ(x)t}{x:τ(x)t}\displaystyle\pi_{x}:\{(x,Du^{-}(x)):\tau(x)\geqslant t\}\to\{x:\tau(x)\geqslant t\}

is a bi-Lipschitz homeomorphism. Therefore, {xM:τ(x)t}\{x\in M:\tau(x)\geqslant t\} is bi-Lipschitz homeomorphic to G(u)G^{*}(u^{-}).

Now we turn to the proof of (2). For any t(0,τ1(u)]t\in(0,\tau_{1}(u^{-})], Proposition 3.10 (2) and Proposition 3.2 imply that ΦHt\Phi_{H}^{-t} determines a C1C^{1} diffeomorphism between {(x,Du(x)):τ(x)<t}\{(x,Du^{-}(x)):\tau(x)<t\} and ΦHt(G#(u))\Phi_{H}^{-t}(G^{\#}(u^{-})), and the projection

πx:{(x,Du(x)):τ(x)<t}{x:τ(x)<t}\displaystyle\pi_{x}:\{(x,Du^{-}(x)):\tau(x)<t\}\to\{x:\tau(x)<t\}

is a bi-Lipschitz homeomorphism. Therefore, {xM:τ(x)<t}\{x\in M:\tau(x)<t\} is bi-Lipschitz homeomorphic to G#(u)G^{\#}(u^{-}). ∎

Corollary 3.12.

Suppose ψ:M\psi:M\to\mathbb{R} is lower semicontinuous, t0>0t_{0}>0. Let

uψ(t,x)=Tt0+tψ(x),(t,x)[t0,0]×M.\displaystyle u_{\psi}(t,x)=T^{-}_{t_{0}+t}\psi(x),\qquad(t,x)\in[-t_{0},0]\times M.

Then the following statements are equivalent

  1. (1)

    u=Tt0ψu^{-}=T^{-}_{t_{0}}\psi.

  2. (2)

    uψ:[t0,0]×Mu_{\psi}:[-t_{0},0]\times M\to\mathbb{R} is a viscosity solution to the Hamilton-Jacobi equation

    {Dtu+H(x,Dxu)=0,(t,x)[t0,0)×M,u(0,x)=u(x),xM.\displaystyle\begin{cases}D_{t}u+H(x,D_{x}u)=0,&(t,x)\in[-t_{0},0)\times M,\\ u(0,x)=u^{-}(x),&x\in M.\end{cases}
  3. (3)

    ψ(x)=u(x)\psi(x)=u^{-}(x) for all xx satisfying τ(x)t0\tau(x)\geqslant t_{0}, and Tt0+u(x)ψ(x)T^{+}_{t_{0}}u^{-}(x)\leqslant\psi(x) otherwise.

Proof.

The conclusion follows directly from Proposition 3.6, Proposition 3.10, and Corollary 3.15 below. ∎

3.4. More on controllability and underlying dynamics

In fact, the equivalence between (1) and (2) in Theorem 3.5 implies that TtTt+ϕ=ϕT^{-}_{t}\circ T^{+}_{t}\phi=\phi is a necessary and sufficient condition for ϕ\phi to be reachable in time tt from some initial data ψ\psi assigned to the evolutionary Hamilton-Jacobi equation

{Dtu(t,x)+H(x,Dxu(t,x))=0,u(0,x)=ψ(x).\displaystyle\begin{cases}D_{t}u(t,x)+H(x,D_{x}u(t,x))=0,\\ u(0,x)=\psi(x).\end{cases}

To understand more on this controllability problem and the underlying dynamics in the context of weak KAM theory, we introduce the following sets, directly related to Theorem 3.5. Recall τ2(ϕ)>0\tau_{2}(\phi)>0 and suppose 0<tτ2(ϕ)0<t\leqslant\tau_{2}(\phi). We define

(ϕ,t):=\displaystyle\mathscr{F}(\phi,t):= {γC1([t,0],M):TtTt+ϕ(γ(0))=Tt+ϕ(γ(t))+t0L(γ,γ˙)𝑑s},\displaystyle\,\bigg{\{}\gamma\in C^{1}([-t,0],M):T^{-}_{t}\circ T^{+}_{t}\phi(\gamma(0))=T^{+}_{t}\phi(\gamma(-t))+\int^{0}_{-t}L(\gamma,\dot{\gamma})\ ds\bigg{\}},
𝒜(ϕ,t):=\displaystyle\mathcal{A}(\phi,t):= {(s,γ(s))[t,0]×M:γ(ϕ,t)},\displaystyle\,\{(s,\gamma(s))\in[-t,0]\times M:\gamma\in\mathscr{F}(\phi,t)\},
(ϕ,t):=\displaystyle\mathscr{F}^{*}(\phi,t):= {γC1([t,0],M):γ(s)=πxΦHs(x,p),s[t,0],xM,pDϕ(x)},\displaystyle\,\bigg{\{}\gamma\in C^{1}([-t,0],M):\gamma(s)=\pi_{x}\Phi^{s}_{H}(x,p),s\in[-t,0],x\in M,p\in D^{*}\phi(x)\bigg{\}},
𝒜(ϕ,t):=\displaystyle\mathcal{A}^{*}(\phi,t):= {(s,γ(s))[t,0]×M:γ(ϕ,t)}.\displaystyle\,\{(s,\gamma(s))\in[-t,0]\times M:\gamma\in\mathscr{F}^{*}(\phi,t)\}.

Given t0(0,τ2(ϕ)]t_{0}\in(0,\tau_{2}(\phi)] and ϕSCL(M)\phi\in\text{\rm SCL}\,(M), we set

u˘(t,x)=Tt+ϕ(x),u(t,x)=Tt0+tTt0+ϕ(x),(t,x)[t0,0]×M.\breve{u}(t,x)=T^{+}_{-t}\phi(x),\qquad u(t,x)=T^{-}_{t_{0}+t}\circ T^{+}_{t_{0}}\phi(x),\qquad(t,x)\in[-t_{0},0]\times M. (3.7)
Proposition 3.13.

Suppose ϕSCL(M)\phi\in\text{\rm SCL}\,(M), t0(0,τ2(ϕ)]t_{0}\in(0,\tau_{2}(\phi)] and u,u˘u,\breve{u} are defined in (3.7).

  1. (1)

    For any (t,x)𝒜(ϕ,t0)({t0,0}×M)(t,x)\in\mathcal{A}(\phi,t_{0})\setminus(\{-t_{0},0\}\times M), there exists a unique γ(ϕ,t0)\gamma\in\mathscr{F}(\phi,t_{0}) such that γ(t)=x\gamma(t)=x. Moreover, if p=Lv(γ,γ˙)p=L_{v}(\gamma,\dot{\gamma}) is the dual arc of γ\gamma, then

    u(t,x)=\displaystyle u(t,x)= u˘(t,x)=ϕ(γ(0))t0L(γ,γ˙)𝑑s,\displaystyle\,\breve{u}(t,x)=\phi(\gamma(0))-\int^{0}_{t}L(\gamma,\dot{\gamma})\ ds,
    Dxu(t,x)=\displaystyle D_{x}u(t,x)= Dxu˘(t,x)=p(t).\displaystyle\,D_{x}\breve{u}(t,x)=p(t).
  2. (2)

    u˘u\breve{u}\leqslant u, and

    𝒜(ϕ,t0)={(t,x)[t0,0]×M:u(t,x)=u˘(t,x)}.\displaystyle\mathcal{A}(\phi,t_{0})=\{(t,x)\in[-t_{0},0]\times M:u(t,x)=\breve{u}(t,x)\}.

    Du=Du˘Liploc(𝒜(ϕ,t0)({t0,0}×M))Du=D\breve{u}\in\text{\rm Lip}_{loc}\,(\mathcal{A}(\phi,t_{0})\setminus(\{-t_{0},0\}\times M)). In particular, 𝒜(ϕ,t0)=[t0,0]×M\mathcal{A}(\phi,t_{0})=[-t_{0},0]\times M if t0(0,τ1(ϕ)]t_{0}\in(0,\tau_{1}(\phi)].

  3. (3)

    (ϕ,t0)(ϕ,t0)\mathscr{F}^{*}(\phi,t_{0})\subset\mathscr{F}(\phi,t_{0}) and 𝒜(ϕ,t0)𝒜(ϕ,t0)\mathcal{A}^{*}(\phi,t_{0})\subset\mathcal{A}(\phi,t_{0}).

Proof.

(1) The uniqueness of γ\gamma follows from the fact that minimizers cannot cross. Notice that γ\gamma is a maximizer for Tt+ϕ(x)T^{+}_{-t}\phi(x) and a minimizer for Tt0+tTt0+ϕ(x)T^{-}_{t_{0}+t}T^{+}_{t_{0}}\phi(x). Thus, we have

u˘(t,x)\displaystyle\breve{u}(t,x) =Tt+ϕ(x)=ϕ(γ(0))t0L(γ,γ˙)𝑑s\displaystyle=T^{+}_{-t}\phi(x)=\phi(\gamma(0))-\int_{t}^{0}L(\gamma,\dot{\gamma})\ ds
=Tt0+ϕ(γ(t0))+t0tL(γ,γ˙)𝑑s=Tt0+tTt0+ϕ(x)=u(t,x).\displaystyle=T^{+}_{t_{0}}\phi(\gamma(-t_{0}))+\int_{-t_{0}}^{t}L(\gamma,\dot{\gamma})\ ds=T^{-}_{t_{0}+t}T^{+}_{t_{0}}\phi(x)=u(t,x).

Since value functions uu and u˘\breve{u} are differentiable in the interior of the minimizer γ\gamma, we have that

p(t)=Dxu(t,x)=Dxu˘(t,x).\displaystyle p(t)=D_{x}u(t,x)=D_{x}\breve{u}(t,x).

(2) u˘u\breve{u}\leqslant u is a direct consequence of Lemma 3.1 (1). In (1) we have already proved that

𝒜(ϕ,t0){(t,x)[t0,0]×M:u(t,x)=u˘(t,x)}.\displaystyle\mathcal{A}(\phi,t_{0})\subset\{(t,x)\in[-t_{0},0]\times M:u(t,x)=\breve{u}(t,x)\}.

On the other hand, for any (t,x)[t0,0]×M(t,x)\in[-t_{0},0]\times M such that u˘(t,x)=u(t,x)\breve{u}(t,x)=u(t,x), there exists γ1:[t,0]M\gamma_{1}:[t,0]\to M, γ1(t)=x\gamma_{1}(t)=x such that

u˘(t,x)=ϕ(γ1(0))t0L(γ1,γ˙1)𝑑s,\displaystyle\breve{u}(t,x)=\phi(\gamma_{1}(0))-\int_{t}^{0}L(\gamma_{1},\dot{\gamma}_{1})\ ds,

and there exists γ2:[t0,t]M\gamma_{2}:[-t_{0},t]\to M, γ2(t)=x\gamma_{2}(t)=x such that

u(t,x)=Tt0+ϕ(γ2(t0))+t0tL(γ2,γ˙2)𝑑s\displaystyle u(t,x)=T^{+}_{t_{0}}\phi(\gamma_{2}(-t_{0}))+\int_{-t_{0}}^{t}L(\gamma_{2},\dot{\gamma}_{2})\ ds

Let γ=γ2γ1¯:[t0,0]M\gamma=\overline{\gamma_{2}\gamma_{1}}:[-t_{0},0]\to M, then γ(t)=x\gamma(t)=x and

Tt0+ϕ(γ(t0))=ϕ(γ(0))t00L(γ,γ˙)𝑑s,\displaystyle T^{+}_{t_{0}}\phi(\gamma(-t_{0}))=\phi(\gamma(0))-\int_{-t_{0}}^{0}L(\gamma,\dot{\gamma})\ ds,

that is, γ(ϕ,t0)\gamma\in\mathscr{F}(\phi,t_{0}). This implies

{(t,x)[t0,0]×M:u(t,x)=u˘(t,x)}𝒜(ϕ,t0).\displaystyle\{(t,x)\in[-t_{0},0]\times M:u(t,x)=\breve{u}(t,x)\}\subset\mathcal{A}(\phi,t_{0}).

Notice that u˘\breve{u} is locally semiconvex and uu is locally semiconcave on (t0,0)×M(-t_{0},0)\times M. Proposition 2.5 leads to

Du=Du˘Liploc(𝒜(ϕ,t0)({t0,0}×M))\displaystyle Du=D\breve{u}\in\text{\rm Lip}_{loc}\,(\mathcal{A}(\phi,t_{0})\setminus(\{-t_{0},0\}\times M))

If t0(0,τ1(ϕ)]t_{0}\in(0,\tau_{1}(\phi)], it follows from Proposition 3.2 (2) that 𝒜(ϕ,t0)=[t0,0]×M\mathcal{A}(\phi,t_{0})=[-t_{0},0]\times M.

(3) For any γ(ϕ,t0)\gamma\in\mathscr{F}^{*}(\phi,t_{0}), let p=Lv(γ,γ˙)p=L_{v}(\gamma,\dot{\gamma}). Then, p(0)Dϕ(γ(0))=DTt0Tt0+ϕ(γ(0))p(0)\in D^{*}\phi(\gamma(0))=D^{*}T^{-}_{t_{0}}T^{+}_{t_{0}}\phi(\gamma(0)). Thus, γ\gamma is a minimizer for Tt0Tt0+ϕ(γ(0))T^{-}_{t_{0}}T^{+}_{t_{0}}\phi(\gamma(0)), that is,

ϕ(γ(0))=Tt0Tt0+ϕ(γ(0))=Tt0+ϕ(γ(t0))+t00L(γ,γ˙)𝑑s,\displaystyle\phi(\gamma(0))=T^{-}_{t_{0}}T^{+}_{t_{0}}\phi(\gamma(0))=T^{+}_{t_{0}}\phi(\gamma(-t_{0}))+\int_{-t_{0}}^{0}L(\gamma,\dot{\gamma})\ ds,

which implies that γ(ϕ,t0)\gamma\in\mathscr{F}(\phi,t_{0}). Therefore, (ϕ,t0)(ϕ,t0)\mathscr{F}^{*}(\phi,t_{0})\subset\mathscr{F}(\phi,t_{0}) and 𝒜(ϕ,t0)𝒜(ϕ,t0)\mathcal{A}^{*}(\phi,t_{0})\subset\mathcal{A}(\phi,t_{0}). ∎

Recall that, for any ϕSCL(M)\phi\in\text{\rm SCL}\,(M) and t0>0t_{0}>0, Tt0Tt0+ϕ=ϕT^{-}_{t_{0}}\circ T^{+}_{t_{0}}\phi=\phi if and only if ϕ=Tt0ψ\phi=T^{-}_{t_{0}}\psi for some lower semicontinuous function ψ\psi. We define

(ϕ,ψ,t0)=\displaystyle\mathscr{F}(\phi,\psi,t_{0})= {γ:[t0,0]M:ϕ(γ(0))=ψ(γ(t0))+t00L(γ,γ˙)𝑑s},\displaystyle\,\left\{\gamma:[-t_{0},0]\to M:\phi(\gamma(0))=\psi(\gamma(-t_{0}))+\int^{0}_{-t_{0}}L(\gamma,\dot{\gamma})\ ds\right\},
𝒜(ϕ,ψ,t0)=\displaystyle\mathcal{A}(\phi,\psi,t_{0})= {(t,γ(t)):γ(ϕ,ψ,t0),t[t0,0]}.\displaystyle\,\{(t,\gamma(t)):\gamma\in\mathscr{F}(\phi,\psi,t_{0}),t\in[-t_{0},0]\}.

and

uψ(t,x):=Tt0+tψ(x),(t,x)[t0,0]×M.u_{\psi}(t,x):=T^{-}_{t_{0}+t}\psi(x),\qquad(t,x)\in[-t_{0},0]\times M. (3.8)
Proposition 3.14.

Suppose ϕSCL(M)\phi\in\text{\rm SCL}\,(M), t0>0t_{0}>0 and ϕ=Tt0ψ\phi=T^{-}_{t_{0}}\psi. Let uu and u˘\breve{u} be defined in (3.7), and let uψu_{\psi} be defined in (3.8).

  1. (1)

    We have the following relations:

    (ϕ,t0)(ϕ,ψ,t0)(ϕ,t0),𝒜(ϕ,t0)𝒜(ϕ,ψ,t0)𝒜(ϕ,t0).\displaystyle\mathscr{F}^{*}(\phi,t_{0})\subset\mathscr{F}(\phi,\psi,t_{0})\subset\mathscr{F}(\phi,t_{0}),\qquad\mathcal{A}^{*}(\phi,t_{0})\subset\mathcal{A}(\phi,\psi,t_{0})\subset\mathcal{A}(\phi,t_{0}).
  2. (2)

    uψuu˘u_{\psi}\geqslant u\geqslant\breve{u} and

    𝒜(ϕ,ψ,t0)={(t,x)[t0,0]×M:uψ(t,x)=u(t,x)=u˘(t,x)}.\displaystyle\mathcal{A}(\phi,\psi,t_{0})=\{(t,x)\in[-t_{0},0]\times M:u_{\psi}(t,x)=u(t,x)=\breve{u}(t,x)\}.

    Moreover, Duψ=Du=Du˘Liploc(𝒜(ϕ,ψ,t0)({t0,0}×M))Du_{\psi}=Du=D\breve{u}\in\text{\rm Lip}_{loc}\,(\mathcal{A}(\phi,\psi,t_{0})\setminus(\{-t_{0},0\}\times M)).

Proof.

(1) The proof of (ϕ,t0)(ϕ,ψ,t0)\mathscr{F}^{*}(\phi,t_{0})\subset\mathscr{F}(\phi,\psi,t_{0}) is similar to the one of Proposition 3.13 (3). For any γ(ϕ,ψ,t0)\gamma\in\mathscr{F}(\phi,\psi,t_{0}), we have that

ψ(γ(t0))=ϕ(γ(0))t00L(γ,γ˙)𝑑sTt0+ϕ(γ(t0)),\displaystyle\psi(\gamma(-t_{0}))=\phi(\gamma(0))-\int_{-t_{0}}^{0}L(\gamma,\dot{\gamma})\ ds\leqslant T^{+}_{t_{0}}\phi(\gamma(-t_{0})),

while Lemma 3.1 (2) leads to Tt0+ϕ=Tt0+Tt0ψψT^{+}_{t_{0}}\phi=T^{+}_{t_{0}}T^{-}_{t_{0}}\psi\leqslant\psi. It follows that

Tt0+ϕ(γ(t0))=ψ(γ(t0))=ϕ(γ(0))t00L(γ,γ˙)𝑑s,\displaystyle T^{+}_{t_{0}}\phi(\gamma(-t_{0}))=\psi(\gamma(-t_{0}))=\phi(\gamma(0))-\int_{-t_{0}}^{0}L(\gamma,\dot{\gamma})\ ds,

which implies γ(ϕ,t0)\gamma\in\mathscr{F}(\phi,t_{0}). Therefore, (ϕ,ψ,t0)(ϕ,t0)\mathscr{F}(\phi,\psi,t_{0})\subset\mathscr{F}(\phi,t_{0}). Similarly, we have that

𝒜(ϕ,t0)𝒜(ϕ,ψ,t0)𝒜(ϕ,t0).\displaystyle\mathcal{A}^{*}(\phi,t_{0})\subset\mathcal{A}(\phi,\psi,t_{0})\subset\mathcal{A}(\phi,t_{0}).

(2) Since Tt0+ϕψT^{+}_{t_{0}}\phi\leqslant\psi, we have that

uψ(t,x)=Tt0+tψ(x)Tt0+tTt0+ϕ(x)=u(t,x),(t,x)[t0,0]×M.\displaystyle u_{\psi}(t,x)=T^{-}_{t_{0}+t}\psi(x)\geqslant T^{-}_{t_{0}+t}T^{+}_{t_{0}}\phi(x)=u(t,x),\ \forall(t,x)\in[-t_{0},0]\times M.

For any (t,x)𝒜(ϕ,ψ,t0)(t,x)\in\mathcal{A}(\phi,\psi,t_{0}), there exists γ(ϕ,ψ,t0)(ϕ,t0)\gamma\in\mathscr{F}(\phi,\psi,t_{0})\subset\mathscr{F}(\phi,t_{0}) such that γ(t)=x\gamma(t)=x. Thus, we have that

uψ(t,x)=ψ(γ(t0))+t0tL(γ,γ˙)𝑑s=ϕ(γ(0))t0L(γ,γ˙)𝑑s=u˘(t,x).\displaystyle u_{\psi}(t,x)=\psi(\gamma(-t_{0}))+\int_{-t_{0}}^{t}L(\gamma,\dot{\gamma})\ ds=\phi(\gamma(0))-\int_{t}^{0}L(\gamma,\dot{\gamma})\ ds=\breve{u}(t,x).

This implies that

𝒜(ϕ,ψ,t0){(t,x)[t0,0]×M:uψ(t,x)=u(t,x)=u˘(t,x)}.\displaystyle\mathcal{A}(\phi,\psi,t_{0})\subset\{(t,x)\in[-t_{0},0]\times M:u_{\psi}(t,x)=u(t,x)=\breve{u}(t,x)\}.

The remaining part of the proof is similar to Proposition 3.13 (2). ∎

Corollary 3.15.

Under the same assumptions of Proposition 3.14, the following statements are equivalent.

  1. (1)

    ϕ=Tt0ψ\phi=T^{-}_{t_{0}}\psi.

  2. (2)

    uψ:[t0,0]×Mu_{\psi}:[-t_{0},0]\times M\to\mathbb{R} is a viscosity solution of the Hamilton-Jacobi equation

    {Dtu+H(x,Dxu)=0,(t,x)[t0,0)×M,u(0,x)=ϕ(x),xM.\displaystyle\begin{cases}D_{t}u+H(x,D_{x}u)=0,&(t,x)\in[-t_{0},0)\times M,\\ u(0,x)=\phi(x),&x\in M.\end{cases}
  3. (3)

    ψ(x)=Tt0+ϕ(x)\psi(x)=T^{+}_{t_{0}}\phi(x) for all xπx(𝒜(ϕ,t0)({t0}×M))x\in\pi_{x}(\mathcal{A}^{*}(\phi,t_{0})\cap(\{-t_{0}\}\times M)), and Tt0+ϕ(x)ψ(x)T^{+}_{t_{0}}\phi(x)\leqslant\psi(x) otherwise.

Proof.

(1) \Leftrightarrow (2) is trivial. (1) \Rightarrow (3) follows directly from Proposition 3.14. We only need to prove that (3) \Rightarrow (1). ψTt0+ϕ\psi\geqslant T^{+}_{t_{0}}\phi implies Tt0ψTt0Tt0+ϕ=ϕT^{-}_{t_{0}}\psi\geqslant T^{-}_{t_{0}}T^{+}_{t_{0}}\phi=\phi. On the other hand, for any xMx\in M, choose pDϕ(x)=DTt0Tt0+ϕ(x)p\in D^{*}\phi(x)=D^{*}T^{-}_{t_{0}}T^{+}_{t_{0}}\phi(x) and let γ(s)=πxΦHs(x,p)\gamma(s)=\pi_{x}\Phi_{H}^{s}(x,p), s[t0,0]s\in[-t_{0},0]. It follows that

ϕ(x)=Tt0Tt0+ϕ(x)=Tt0+ϕ(γ(t0))+t00L(γ,γ˙)𝑑s.\displaystyle\phi(x)=T^{-}_{t_{0}}T^{+}_{t_{0}}\phi(x)=T^{+}_{t_{0}}\phi(\gamma(-t_{0}))+\int_{-t_{0}}^{0}L(\gamma,\dot{\gamma})\ ds.

Notice that γ(t0)πx(𝒜(ϕ,t0)({t0}×M))\gamma(-t_{0})\in\pi_{x}(\mathcal{A}^{*}(\phi,t_{0})\cap(\{-t_{0}\}\times M)), which implies Tt0+ϕ(γ(t0))=ψ(γ(t0))T^{+}_{t_{0}}\phi(\gamma(-t_{0}))=\psi(\gamma(-t_{0})). Thus, we have

ϕ(x)=ψ(γ(t0))+t00L(γ,γ˙)𝑑sTt0ψ(x).\displaystyle\phi(x)=\psi(\gamma(-t_{0}))+\int_{-t_{0}}^{0}L(\gamma,\dot{\gamma})\ ds\geqslant T^{-}_{t_{0}}\psi(x).

In conclusion, there holds Tt0ψ(x)=ϕ(x)T^{-}_{t_{0}}\psi(x)=\phi(x), xM\forall x\in M. ∎

Appendix A some facts from weak KAM theory

Proposition A.1.

Suppose LL is a Tonelli Lagrangian. Then for any λ>0\lambda>0

  1. (1)

    there exists a constant Cλ>0C_{\lambda}>0 such that for any xnx\in\mathbb{R}^{n}, t(0,2/3)t\in(0,2/3) the function yAt(x,y)y\mapsto A_{t}(x,y) defined on B(x,λt)B(x,\lambda t) is semiconcave with constant Cλt\frac{C_{\lambda}}{t} uniformly with respect to xx;

  2. (2)

    there exist Cλ>0C^{\prime}_{\lambda}>0 and tλ>0t_{\lambda}>0 such that the function yAt(x,y)y\mapsto A_{t}(x,y) is convex with constant Cλt\frac{C^{\prime}_{\lambda}}{t} on B(x,λt)B(x,\lambda t) with 0<ttλ0<t\leqslant t_{\lambda}. The constants CλC^{\prime}_{\lambda} and tλt_{\lambda} are independent of xx;

  3. (3)

    there exists tλ>0t^{\prime}_{\lambda}>0 such that the function yAt(x,y)y\mapsto A_{t}(x,y) is of class C2C^{2} on B(x,λt)B(x,\lambda t) with 0<ttλ0<t\leqslant t^{\prime}_{\lambda}. Moreover,

    DyAt(x,y)=\displaystyle D_{y}A_{t}(x,y)= Lv(ξ(t),ξ˙(t)),\displaystyle L_{v}(\xi(t),\dot{\xi}(t)),
    DxAt(x,y)=\displaystyle D_{x}A_{t}(x,y)= Lv(ξ(0),ξ˙(0)),\displaystyle-L_{v}(\xi(0),\dot{\xi}(0)),
    DtAt(x,y)=\displaystyle D_{t}A_{t}(x,y)= Et,x,y,\displaystyle-E_{t,x,y},

    where ξΓx,yt\xi\in\Gamma^{t}_{x,y} is the unique minimizer for At(x,y)A_{t}(x,y) and

    Et,x,y:=E(ξ(s),ξ˙(s))s[0,t].\displaystyle E_{t,x,y}:=E(\xi(s),\dot{\xi}(s))\qquad\forall\,s\in[0,t].

    We remark that E(x,v):=Lv(x,v)vL(x,v)E(x,v):=L_{v}(x,v)\cdot v-L(x,v) is the energy function in the Lagrangian formalism, and

    E(ξ(s),ξ˙(s))=H(ξ(s),p(s)),s[0,t],\displaystyle E(\xi(s),\dot{\xi}(s))=H(\xi(s),p(s)),\quad s\in[0,t],

    for the dual arc p(s)=Lv(ξ(s),ξ˙(s))p(s)=L_{v}(\xi(s),\dot{\xi}(s));

For the proof of the aforementioned Proposition, the readers can refer to [12] and [13].

Lemma A.2 ([17]).

For any ε>0\varepsilon>0 there exists r2(ε)>0r_{2}(\varepsilon)>0 such that if t>r2(ε)t>r_{2}(\varepsilon) and γΓx,yt\gamma\in\Gamma^{t}_{x,y} is a minimal curve for At(x,y)A_{t}(x,y), and p=Lv(γ,γ˙)p=L_{v}(\gamma,\dot{\gamma}) is the dual arc, then (recall that H(γ(s),p(s))=constH(\gamma(s),p(s))=const)

|H(γ(s),p(s))c[H]|<ε,s[0,t].\displaystyle|H(\gamma(s),p(s))-c[H]|<\varepsilon,\quad\forall s\in[0,t].

References

  • [1] Luigi Ambrosio, Elia Brué, and Daniele Semola. Lectures on optimal transport, volume 130 of Unitext. Springer, 2021.
  • [2] Fabio Ancona, Piermarco Cannarsa, and Khai T. Nguyen. Compactness estimates for Hamilton-Jacobi equations depending on space. Bull. Inst. Math. Acad. Sin. (N.S.), 11(1):63–113, 2016.
  • [3] Fabio Ancona, Piermarco Cannarsa, and Khai T. Nguyen. Quantitative compactness estimates for Hamilton-Jacobi equations. Arch. Ration. Mech. Anal., 219(2):793–828, 2016.
  • [4] M.-C. Arnaud. Pseudographs and the Lax-Oleinik semi-group: a geometric and dynamical interpretation. Nonlinearity, 24(1):71–78, 2011.
  • [5] Martino Bardi and Italo Capuzzo-Dolcetta. Optimal control and viscosity solutions of Hamilton-Jacobi-Bellman equations. Systems & Control: Foundations & Applications. Birkhäuser Boston, Inc., Boston, MA, 1997. With appendices by Maurizio Falcone and Pierpaolo Soravia.
  • [6] E. N. Barron, P. Cannarsa, R. Jensen, and C. Sinestrari. Regularity of Hamilton-Jacobi equations when forward is backward. Indiana Univ. Math. J., 48(2):385–409, 1999.
  • [7] Patrick Bernard. Existence of C1,1C^{1,1} critical sub-solutions of the Hamilton-Jacobi equation on compact manifolds. Ann. Sci. École Norm. Sup. (4), 40(3):445–452, 2007.
  • [8] Patrick Bernard. Lasry-Lions regularization and a lemma of Ilmanen. Rend. Semin. Mat. Univ. Padova, 124:221–229, 2010.
  • [9] Patrick Bernard. The Lax-Oleinik semi-group: a Hamiltonian point of view. Proc. Roy. Soc. Edinburgh Sect. A, 142(6):1131–1177, 2012.
  • [10] Patrick Bernard and Boris Buffoni. Optimal mass transportation and Mather theory. J. Eur. Math. Soc. (JEMS), 9(1):85–121, 2007.
  • [11] Patrick Bernard and Boris Buffoni. Weak KAM pairs and Monge-Kantorovich duality. In Asymptotic analysis and singularities—elliptic and parabolic PDEs and related problems, volume 47 of Adv. Stud. Pure Math., pages 397–420. Math. Soc. Japan, Tokyo, 2007.
  • [12] Piermarco Cannarsa and Wei Cheng. Generalized characteristics and Lax-Oleinik operators: global theory. Calc. Var. Partial Differential Equations, 56(5):Art. 125, 31, 2017.
  • [13] Piermarco Cannarsa and Wei Cheng. Singularities of Solutions of Hamilton–Jacobi Equations. Milan J. Math., 89(1):187–215, 2021.
  • [14] Piermarco Cannarsa, Wei Cheng, and Albert Fathi. On the topology of the set of singularities of a solution to the Hamilton-Jacobi equation. C. R. Math. Acad. Sci. Paris, 355(2):176–180, 2017.
  • [15] Piermarco Cannarsa, Wei Cheng, and Albert Fathi. Singularities of solutions of time dependent Hamilton-Jacobi equations. Applications to Riemannian geometry. Publ. Math. Inst. Hautes Études Sci., 133(1):327–366, 2021.
  • [16] Piermarco Cannarsa and Carlo 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.
  • [17] M. J. Dias Carneiro. On minimizing measures of the action of autonomous Lagrangians. Nonlinearity, 8(6):1077–1085, 1995.
  • [18] Carlos Esteve-Yagüe and Enrique Zuazua. Reachable set for Hamilton-Jacobi equations with non-smooth Hamiltonian and scalar conservation laws. Nonlinear Anal., 227:Paper No. 113167, 18, 2023.
  • [19] Albert Fathi. Weak KAM theorem in Lagrangian dynamics. Cambridge University Press, Cambridge (to appear).
  • [20] Albert Fathi. Solutions KAM faibles conjuguées et barrières de Peierls. C. R. Acad. Sci. Paris Sér. I Math., 325(6):649–652, 1997.
  • [21] Albert Fathi. Sur la convergence du semi-groupe de Lax-Oleinik. C. R. Acad. Sci. Paris Sér. I Math., 327(3):267–270, 1998.
  • [22] Albert Fathi and Alessio Figalli. Optimal transportation on non-compact manifolds. Israel J. Math., 175:1–59, 2010.
  • [23] Pierre-Louis Lions. Generalized solutions of Hamilton-Jacobi equations, volume 69 of Research Notes in Mathematics. Pitman (Advanced Publishing Program), Boston, Mass.-London, 1982.
  • [24] John N. Mather. Variational construction of connecting orbits. Ann. Inst. Fourier (Grenoble), 43(5):1349–1386, 1993.
  • [25] Constantin Udrişte and Constantin Udrişte. Convex functions and optimization methods on Riemannian manifolds, volume 297 of Mathematics and its Applications. Kluwer Academic Publishers Group, Dordrecht, 1994.
  • [26] Cédric Villani. Optimal transport: old and new, volume 338 of Grundlehren der Mathematischen Wissenschaften. Springer-Verlag, Berlin, 2009.