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

Global propagation of singularities for discounted Hamilton-Jacobi equations

Cui Chen and Jiahui Hong and Kai Zhao School of Mathematical Sciences, Jiangsu University, Zhenjiang 212013, China [email protected] Department of Mathematics, Nanjing University, Nanjing 210093, China [email protected] School of Mathematical Sciences, Fudan University, Shanghai 200433, China zhao_\_[email protected]
Abstract.

The main purpose of this paper is to study the global propagation of singularities of viscosity solution to discounted Hamilton-Jacobi equation

λv(x)+H(x,Dv(x))=0,xn.\lambda v(x)+H(x,Dv(x))=0,\quad x\in\mathbb{R}^{n}. (HJλ)

We reduce the problem for equation (HJλ) into that for a time-dependent evolutionary Hamilton-Jacobi equation. We proved that the singularities of the viscosity solution of (HJλ) propagate along locally Lipschitz singular characteristics which can extend to ++\infty. We also obtained the homotopy equivalence between the singular set and the complement of associated the Aubry set with respect to the viscosity solution of equation (HJλ).

Key words and phrases:
Hamilton-Jacobi equation, viscosity solutions, singularities

1. introduction

It is commonly accepted that, in optimal control, a crucial role is played by the Hamilton-Jacobi equation

{Dtu(t,x)+H(t,x,Dxu(t,x))=0(t,x)+×n,u(0,x)=u0(x)xn.\begin{cases}D_{t}u(t,x)+H(t,x,D_{x}u(t,x))=0\quad&(t,x)\in\mathbb{R}^{+}\times\mathbb{R}^{n},\\ u(0,x)=u_{0}(x)&x\in\mathbb{R}^{n}.\end{cases} (HJe)

It is well known that the singularities of such solutions propagate locally along generalized characteristics. The evidence of irreversibility for the Hamilton-Jacobi equation is the propagation of singularities. Once a singularity is created, it will propagate forward in time up to ++\infty. For a comprehensive survey of this topic, the readers can refer to [7].

The theory of local propagation of the singularities of the viscosity solutions of (HJe) has been established in [2] by introducing the notion of generalized characteristics (see also [14], [26]). Other progress on the local propagation includes the strict singular characteristics ([22], see also [24]). A recent remarkable result by Cannarsa and Cheng established the relation between generalized characteristics and strict singular characteristics on 2\mathbb{R}^{2} ([6]).

In the paper [4], Cannarsa and Cheng introduced an intrinsic method and obtained a global propagation result for time-independent Hamiltonian (see also [12]). By a procedure of sup-convolution with the kernel the fundamental solutions of associated autonomous Hamilton-Jacobi equations, they constructed a global singular arc 𝐲x(t):[0,t0]n\mathbf{y}_{x}(t):[0,t_{0}]\to\mathbb{R}^{n} from an initial singular point xx and t0t_{0} is independent of the initial point xx. The uniformness of such t0t_{0} holds because of uniform conditions (L1)-(L3) in [4] . In [5], they ask the following problem 𝐀𝟑\mathbf{A3}:

Can we drop the uniformness requirement of such t0 to obtain a global result ?\displaystyle\text{Can we drop the uniformness requirement of such }t_{0}\text{ to obtain a global result ?}

The first task of this paper is to drop the uniformness of t0t_{0} which can not be guaranteed by, e.g., the so called Fathi-Maderna conditions ([21]) which we will use for our purpose. In this paper, we showed that the answer to problem 𝐀𝟑\mathbf{A3} is affirmative for time-dependent case and discounted case.

There is a very natural connection between the discounted Hamilton-Jacobi equation (HJλ) and the evolutionary Hamilton-Jacobi equation (HJe) using a conformal Hamiltonian (see, for instance, [23]) or a contact Hamiltonian (see, for instance, [15]). More precisely, if vv is the unique viscosity solution of (HJλ), we define

u(t,x)=eλtv(x),xn,t>0.\displaystyle u(t,x)=e^{\lambda t}v(x),\qquad x\in\mathbb{R}^{n},t>0.

Then u(t,x)u(t,x) is a viscosity solution of (HJe) with a time-dependent Hamiltonian in the form eλtH(x,p/eλt)e^{\lambda t}H(x,p/e^{\lambda t}). Notice that vv and uu share the singularity. Thus, we can discuss the problem of propagation of singularities for equation (HJe) instead of equation (HJλ). We developed the intrinsic method in [4] adapt to our problem which has more technical difficulty comparing to the time-independent case ([9]).

Now we introduce the associated Lagrangian as

L(s,x,v)=suppn{pvH(s,x,p)},s>0,xn,vn.\displaystyle L(s,x,v)=\sup_{p\in\mathbb{R}^{n}}\{p\cdot v-H(s,x,p)\},\qquad s>0,x\in\mathbb{R}^{n},v\in\mathbb{R}^{n}.

To deal with evolutionary Hamilton-Jacobi equation (HJe), we suppose L=L(s,x,v):×n×nL=L(s,x,v):\mathbb{R}\times\mathbb{R}^{n}\times\mathbb{R}^{n}\to\mathbb{R} is of class C2C^{2} and satisfies the following assumptions:

  1. (L1)

    L(s,x,)L(s,x,\cdot) is strict convex on n\mathbb{R}^{n} for all s[0,]s\in[0,\infty], xnx\in\mathbb{R}^{n} .

  2. (L2)

    For any fixed T>0T>0, there exist cT>0c_{T}>0 and two superlinear and nondecreasing function θ¯T,θT:[0,+)[0,+)\overline{\theta}_{T},\theta_{T}:[0,+\infty)\to[0,+\infty), such that θ¯T(|v|)L(s,x,v)θT(|v|)cT,\overline{\theta}_{T}(|v|)\geqslant L(s,x,v)\geqslant\theta_{T}(|v|)-c_{T}, for all (s,x,v)[0,T]×n×n.(s,x,v)\in[0,T]\times\mathbb{R}^{n}\times\mathbb{R}^{n}.

  3. (L3)

    There exists C~1,C~2:[0,+)\widetilde{C}_{1},\widetilde{C}_{2}:\mathbb{R}\rightarrow[0,+\infty) such that |Lt(s,x,v)|C~1(T)+C~2(T)L(s,x,v)|L_{t}(s,x,v)|\leqslant\widetilde{C}_{1}(T)+\widetilde{C}_{2}(T)L(s,x,v) for all (s,x,v)[0,T]×n×n(s,x,v)\in[0,T]\times\mathbb{R}^{n}\times\mathbb{R}^{n}.

We say that a curve γ:[a,b]n\gamma:[a,b]\to\mathbb{R}^{n} is λ\lambda-calibrated curve for equation (HJλ) if

eλbu(γ(b))eλau(γ(a))=abeλtL(γ(t),γ˙(t))𝑑t,e^{\lambda b}u(\gamma(b))-e^{\lambda a}u(\gamma(a))=\int_{a}^{b}e^{\lambda t}L(\gamma(t),\dot{\gamma}(t))\ dt, (1.1)

and a curve γ:[a,b]n\gamma:[a,b]\to\mathbb{R}^{n} is calibrated curve for equation (HJe) if

u(b,γ(b))u(a,γ(a))=abL(t,γ(t),γ˙(t))𝑑t.u(b,\gamma(b))-u(a,\gamma(a))=\int_{a}^{b}L(t,\gamma(t),\dot{\gamma}(t))\ dt. (1.2)

A point xnx\in\mathbb{R}^{n} is a cut point of uu if no backward λ\lambda-calibrated curve of equation (HJλ) with Hamilton HH ending at xx can be extended beyond xx. A point (t,x)+×n(t,x)\in\mathbb{R}^{+}\times\mathbb{R}^{n} is a cut point of uu if no backward calibrated curve of equation (HJe) with Hamilton HH ending at (t,x)(t,x) can be extended beyond (t,x)(t,x). In both cases, we denote by Cut(u)\mbox{\rm{Cut}}\,(u) the set of cut points of uu. If uu is a viscosity solution of (HJλ) or (HJe), a singularity of uu is a point where uu is not differentiable. We denote by Sing(u)\mbox{\rm{Sing}}(u) the set of singularities of uu. It is well known that Sing(u)Cut(u)Sing(u)¯\mbox{\rm{Sing}}\,(u)\subset\mbox{\rm{Cut}}\,(u)\subset\overline{\mbox{\rm{Sing}}\,(u)}.

Our main result for the time-dependent case is: Let LL be a Lagrangian which satisfies (L1)-(L3) and let HH be the associated Hamiltonian. Suppose u0=u(0,):nu_{0}=u(0,\cdot):\mathbb{R}^{n}\rightarrow\mathbb{R} is a Lipschitz continuous function. Then for any fixed (t0,x)Cut(u)+×n(t_{0},x)\in\mbox{\rm{Cut}}(u)\subset\mathbb{R}^{+}\times\mathbb{R}^{n}, there exists a curve 𝐱:n\mathbf{x}:\mathbb{R}\rightarrow\mathbb{R}^{n} with 𝐱(t0)=x\mathbf{x}(t_{0})=x, such that (s,𝐱(s))Sing(u)(s,\mathbf{x}(s))\in\mbox{\rm{Sing}}(u) for all s[t0,+)s\in[t_{0},+\infty). Moreover, If condition (A) (see Section 3) holds, then for any T>0T>0, 𝐱(s)\mathbf{x}(s) is a Lipschitz curve on s[t0,T]s\in[t_{0},T].

Similarly, for the discounted equation (HJλ) we denote by LL the associated Lagrangian of HH. We suppose L=L(x,v):n×nL=L(x,v):\mathbb{R}^{n}\times\mathbb{R}^{n}\to\mathbb{R} is of C2C^{2} class and satisfying the following assumptions:

  1. (L1’)

    L(x,)L(x,\cdot) is strictly convex for all xnx\in\mathbb{R}^{n}.

  2. (L2’)

    There exist c1,c20c_{1},c_{2}\geqslant 0 and two superlinear functions θ1,θ2:[0,+)[0,+)\theta_{1},\theta_{2}:[0,+\infty)\to[0,+\infty) such that

    θ2(|v|)+c2L(x,v)θ1(|v|)c1,(x,v)n×n.\displaystyle\theta_{2}(|v|)+c_{2}\geqslant L(x,v)\geqslant\theta_{1}(|v|)-c_{1},\qquad\forall(x,v)\in\mathbb{R}^{n}\times\mathbb{R}^{n}.

Our main result for the discounted case is: Let LL be a Lagrangian which satisfies (L1’)-(L2’) and HH be the associated Hamiltonian and λ>0\lambda>0. Suppose v:nv:\mathbb{R}^{n}\to\mathbb{R} is a Lipschitz continuous semiconcave viscosity solution of (HJλ). Then

  1. (1)

    for any fixed xCut(v)x\in\mbox{\rm{Cut}}(v), there exists a locally Lipschitz curve 𝐱:[0,+)n\mathbf{x}:[0,+\infty)\rightarrow\mathbb{R}^{n} with 𝐱(0)=x\mathbf{x}(0)=x, such that 𝐱(s)Sing(v)\mathbf{x}(s)\in\mbox{\rm{Sing}}(v) for all s[0,+)s\in[0,+\infty),

  2. (2)

    the inclusions

    Sing(v)Cut(v)(n\(v))Sing(v)¯n\(v)\mbox{\rm{Sing}}(v)\subset\mbox{\rm{Cut}}(v)\subset\Big{(}\mathbb{R}^{n}\backslash\mathcal{I}(v)\Big{)}\cap\overline{\mbox{\rm{Sing}}(v)}\subset\mathbb{R}^{n}\backslash\mathcal{I}(v)

    are all homotopy equivalences and the spaces Sing(u)\mbox{\rm{Sing}}\,(u) and Cut(u)\mbox{\rm{Cut}}\,(u) are all locally contractible.

It worth noting that the construction of the homotopy equivalence here we used is very similar to what used in [8], [11] and [9]. The general notion of the cut locus of uu for contact type Hamilton-Jacobi equation was studied in [16] recently for smooth initial data.

This paper is organized as follows. In Sect. 2, we introduce Lax-Oleinik operator associated to (HJe) and give our global result on the propagation of singularities along local Lipschitz curves under an extra condition (A). In Sect.3, we discuss the global propagation of singularities for discounted Hamiltonian (HJλ) and give homotopy equivalence results as an application.This paper contains three appendices which include some background materials and useful conclusions. In Appendix A, we collect some relevant regularity results with respect to the fundamental solution of (HJe). In Appendix B and Appendix C, we give the proof of Lemma 2.2, Lemma 2.6 and Lemma 3.8.

Acknowledgements. Cui Chen is partly supported by National Natural Science Foundation of China (Grant No. 11801223, 11871267).

2. Global propagation of singularities for time-dependent Hamiltonian

In this section, we will discuss the connection between sup-convolution, singularities and generalized characteristics for the following time-dependent Hamilton-Jacobi equation:

{Dtu(t,x)+H(t,x,Dxu(t,x))=0(t,x)+×n,u(0,x)=u0(x)xn.\begin{cases}D_{t}u(t,x)+H(t,x,D_{x}u(t,x))=0\quad&(t,x)\in\mathbb{R}^{+}\times\mathbb{R}^{n},\\ u(0,x)=u_{0}(x)&x\in\mathbb{R}^{n}.\end{cases} (HJe)

Let L(s,x,v)L(s,x,v) be the associated Lagrangian of H(s,x,p)H(s,x,p). We assume that L(s,x,v):[0,+)×n×nL(s,x,v):[0,+\infty)\times\mathbb{R}^{n}\times\mathbb{R}^{n}\to\mathbb{R} is a C2C^{2}-smooth function which satisfies the following standard assumptions:

  1. (L1)

    Lvv(s,x,v)>0L_{vv}(s,x,v)>0 for any (s,x,v)[0,+)×n×n(s,x,v)\in[0,+\infty)\times\mathbb{R}^{n}\times\mathbb{R}^{n}.

  2. (L2)

    For any fixed T>0T>0, there exist cT>0c_{T}>0 and two superlinear and nondecreasing functions θ¯T,θT:[0,+)[0,+)\overline{\theta}_{T},\theta_{T}:[0,+\infty)\to[0,+\infty), such that

    θ¯T(|v|)L(s,x,v)θT(|v|)cT,(s,x,v)[0,T]×n×n.\displaystyle\overline{\theta}_{T}(|v|)\geqslant L(s,x,v)\geqslant\theta_{T}(|v|)-c_{T},\quad(s,x,v)\in[0,T]\times\mathbb{R}^{n}\times\mathbb{R}^{n}.
  3. (L3)

    There exist two locally bounded functions C~1,C~2:[0,+)[0,+)\widetilde{C}_{1},\widetilde{C}_{2}:[0,+\infty)\rightarrow[0,+\infty) such that for any T>0T>0,

    |Lt(s,x,v)|C~1(T)+C~2(T)L(s,x,v),(s,x,v)[0,T]×n×n.|L_{t}(s,x,v)|\leqslant\widetilde{C}_{1}(T)+\widetilde{C}_{2}(T)L(s,x,v),\quad(s,x,v)\in[0,T]\times\mathbb{R}^{n}\times\mathbb{R}^{n}.

For any 0s<t<+0\leqslant s<t<+\infty and x,ynx,y\in\mathbb{R}^{n}, we define the fundamental solution of the Hamilton-Jacobi equation (HJe) as follows:

As,t(x,y)=infξΓx,ys,tstL(τ,ξ(τ),ξ˙(τ))𝑑τ.A_{s,t}(x,y)=\inf_{\xi\in\Gamma^{s,t}_{x,y}}\int_{s}^{t}L(\tau,\xi(\tau),\dot{\xi}(\tau))d\tau. (2.1)

where

Γx,ys,t={ξW1,1([0,t],n):γ(s)=x,γ(t)=y}\Gamma^{s,t}_{x,y}=\{\xi\in W^{1,1}([0,t],\mathbb{R}^{n}):\gamma(s)=x,\gamma(t)=y\}

We call ξΓx,ys,t\xi\in\Gamma^{s,t}_{x,y} a minimizer for As,t(x,y)A_{s,t}(x,y) if As,t(x,y)=stL(τ,ξ(τ),ξ˙(τ))𝑑τ.A_{s,t}(x,y)=\int_{s}^{t}L(\tau,\xi(\tau),\dot{\xi}(\tau))d\tau. The existence of minimizers in (2.1) is a well known result in Tonelli’s theory (see, for instance, [17]). Moreover, we have the following proposition:

Proposition 2.1.

Suppose LL satisfies (L1)-(L3). Then for any 0s<t<+0\leqslant s<t<+\infty and x,ynx,y\in\mathbb{R}^{n}, there exists ξΓx,ys,t\xi\in\Gamma^{s,t}_{x,y} such that ξ\xi is a minimizer for As,t(x,y)A_{s,t}(x,y) and the following properties hold:

  1. (1)

    ξ\xi is of class C2C^{2} and satisfies

    ddsLv(τ,ξ(τ),ξ˙(τ))=Lx(τ,ξ(τ),ξ˙(τ)),τ[s,t].\frac{d}{ds}L_{v}(\tau,\xi(\tau),\dot{\xi}(\tau))=L_{x}(\tau,\xi(\tau),\dot{\xi}(\tau)),\quad\forall\tau\in[s,t].
  2. (2)

    Let p(τ)=Lv(τ,ξ(τ),ξ˙(τ))p(\tau)=L_{v}(\tau,\xi(\tau),\dot{\xi}(\tau)) for τ[s,t]\tau\in[s,t]. Then (ξ,p)(\xi,p) satisfies

    {ξ˙(τ)=Hp(τ,ξ(τ),p(τ)),p˙(τ)=Hx(τ,ξ(τ),p(τ)),τ[s,t].\begin{cases}\dot{\xi}(\tau)=H_{p}(\tau,\xi(\tau),p(\tau)),\\ \dot{p}(\tau)=-H_{x}(\tau,\xi(\tau),p(\tau)),\end{cases}\qquad\forall\tau\in[s,t]. (2.2)

In Appendix A, we collect some relevant regularity results with respect to the fundamental solution As,t(x,y)A_{s,t}(x,y). The proofs of these regularity results are similar to those in [4] for autonomous case.

2.1. Semiconcave functions

Let Ωn\Omega\subset\mathbb{R}^{n} be a convex open set. We recall that a function u:Ωu:\Omega\to\mathbb{R} is said to be semiconcave (with linear modulus) if there exists a constant C>0C>0 such that

λu(x)+(1λ)u(y)u(λx+(1λ)y)C2λ(1λ)|xy|2,x,yΩ,λ[0,1].\lambda u(x)+(1-\lambda)u(y)-u(\lambda x+(1-\lambda)y)\leqslant\frac{C}{2}\lambda(1-\lambda)|x-y|^{2},\quad\forall x,y\in\Omega,\lambda\in[0,1].

For any continuous function u:nu:\mathbb{R}^{n}\to\mathbb{R} and xnx\in\mathbb{R}^{n}, we denote

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{\}},

which are called the subdifferential and superdifferential of uu at xx, respectively. Let now u:nu:\mathbb{R}^{n}\to\mathbb{R} be locally Lipschitz and xnx\in\mathbb{R}^{n}. We call pnp\in\mathbb{R}^{n} a reachable gradient of uu at xx if there exists a sequence {xk}\{x_{k}\} such that uu is differentiable at xkx_{k} for all kk\in\mathbb{N} and

limkxk=x,limkDu(xk)=p.\lim_{k\to\infty}x_{k}=x,\qquad\lim_{k\to\infty}Du(x_{k})=p.

The set of all reachable gradients of uu at xx is denoted by Du(x)D^{*}u(x).

2.2. Lax-Oleinik operator in time-dependent case and a priori estimate

Let f:nf:\mathbb{R}^{n}\to\mathbb{R} be a Lipschitz function. For any 0t1<t2<+0\leqslant t_{1}<t_{2}<+\infty and x1,x2nx_{1},x_{2}\in\mathbb{R}^{n}, we define the Lax-Oleinik operator

Tt1,t2f(x2):=infzn{f(z)+At1,t2(z,x2)},T^{-}_{t_{1},t_{2}}f(x_{2}):=\inf_{z\in\mathbb{R}^{n}}\{f(z)+A_{t_{1},t_{2}}(z,x_{2})\}, (2.3)
Tt1,t2+f(x1):=supyn{f(y)At1,t2(x1,y)},T^{+}_{t_{1},t_{2}}f(x_{1}):=\sup_{y\in\mathbb{R}^{n}}\{f(y)-A_{t_{1},t_{2}}(x_{1},y)\}, (2.4)

and denote

Z(f,t1,t2,x2)={zn:Tt1,t2f(x2)=f(z)+At1,t2(z,x2)},Y(f,t1,x1,t2)={yn:Tt1,t2+f(x1)=f(y)At1,t2(x1,y)}.\begin{split}Z(f,t_{1},t_{2},x_{2})=\{z\in\mathbb{R}^{n}:T^{-}_{t_{1},t_{2}}f(x_{2})=f(z)+A_{t_{1},t_{2}}(z,x_{2})\},\\ Y(f,t_{1},x_{1},t_{2})=\{y\in\mathbb{R}^{n}:T^{+}_{t_{1},t_{2}}f(x_{1})=f(y)-A_{t_{1},t_{2}}(x_{1},y)\}.\end{split} (2.5)

From Appendix B, we have the following a priori estimates:

Lemma 2.2.

(proved in Appendix B)  Suppose LL satisfies (L1)-(L3) and ff is a Lipschitz function on n\mathbb{R}^{n}. Then for any fixed T>0T>0, there exists a constant λ1(T,Lip[f])>0\lambda_{1}(T,\mbox{\rm{Lip}}[f])>0 such that for any 0t1<t2T0\leqslant t_{1}<t_{2}\leqslant T and x1,x2nx_{1},x_{2}\in\mathbb{R}^{n}

  1. (1)

    Z(f,t1,t2,x2)Z(f,t_{1},t_{2},x_{2})\neq\emptyset, and for any zt1,t2,x2Z(f,t1,t2,x2)z_{t_{1},t_{2},x_{2}}\in Z(f,t_{1},t_{2},x_{2}),

    |zt1,t2,x2x2|λ1(T,Lip[f])(t2t1).|z_{t_{1},t_{2},x_{2}}-x_{2}|\leqslant\lambda_{1}(T,\mbox{\rm{Lip}}[f])(t_{2}-t_{1}).
  2. (2)

    Y(f,t1,x1,t2)Y(f,t_{1},x_{1},t_{2})\neq\emptyset, and for any yt1,t2,x1Y(f,t1,x1,t2)y_{t_{1},t_{2},x_{1}}\in Y(f,t_{1},x_{1},t_{2}),

    |yt1,t2,x2x1|λ1(T,Lip[f])(t2t1).|y_{t_{1},t_{2},x_{2}}-x_{1}|\leqslant\lambda_{1}(T,\mbox{\rm{Lip}}[f])(t_{2}-t_{1}).

where λ1(T,K)=θT(K+1)+cT+θ¯T(0)\lambda_{1}(T,K)=\theta_{T}^{*}(K+1)+c_{T}+\overline{\theta}_{T}(0) for T>0T>0 and K0K\geqslant 0.

For 0t1<t2<+0\leqslant t_{1}<t_{2}<+\infty and x1,x2nx_{1},x_{2}\in\mathbb{R}^{n}, denote

Γ,x2t1,t2={ξW1,1([t1,t2],n):ξ(t2)=x2},\displaystyle\Gamma^{t_{1},t_{2}}_{\cdot,x_{2}}=\{\xi\in W^{1,1}([t_{1},t_{2}],\mathbb{R}^{n}):\xi(t_{2})=x_{2}\},
Γx1,t1,t2={ξW1,1([t1,t2],n):ξ(t1)=x1}.\displaystyle\Gamma^{t_{1},t_{2}}_{x_{1},\cdot}=\{\xi\in W^{1,1}([t_{1},t_{2}],\mathbb{R}^{n}):\xi(t_{1})=x_{1}\}.
Lemma 2.3.

Suppose LL satisfies (L1)-(L3), ff is a Lipschitz function on n\mathbb{R}^{n} and 0t1<t2<+0\leqslant t_{1}<t_{2}<+\infty, x1,x2nx_{1},x_{2}\in\mathbb{R}^{n}.

  1. (1)

    If ξΓ,x2t1,t2\xi\in\Gamma^{t_{1},t_{2}}_{\cdot,x_{2}} is a minimizer for Tt1,t2f(x2)T_{t_{1},t_{2}}^{-}f(x_{2}), then p(t1)=Lv(t1,ξ(t1),ξ˙(t1))Df(ξ(t1)).p(t_{1})=L_{v}(t_{1},\xi(t_{1}),\dot{\xi}(t_{1}))\in D^{-}f(\xi(t_{1})).

  2. (2)

    If ξΓx1,t1,t2\xi\in\Gamma^{t_{1},t_{2}}_{x_{1},\cdot} is a maximizer for Tt1,t2+f(x1)T_{t_{1},t_{2}}^{+}f(x_{1}), then p(t2)=Lv(t2,ξ(t2),ξ˙(t2))D+f(ξ(t2)).p(t_{2})=L_{v}(t_{2},\xi(t_{2}),\dot{\xi}(t_{2}))\in D^{+}f(\xi(t_{2})).

From now on, suppose u0u_{0} is a Lipschitz function on n\mathbb{R}^{n} and denote

u(t,x)=T0,tu0(x)=infzn{u0(z)+A0,t(z,x)},(t,x)[0,+)×n.u(t,x)=T^{-}_{0,t}u_{0}(x)=\inf_{z\in\mathbb{R}^{n}}\{u_{0}(z)+A_{0,t}(z,x)\},\quad(t,x)\in[0,+\infty)\times\mathbb{R}^{n}. (2.6)

Actually, we also have the following representation:

u(t,x)=infξΓ,x0,t{u0(ξ(0))+0tL(τ,ξ(τ),ξ˙(τ))𝑑τ}.u(t,x)=\inf_{\xi\in\Gamma^{0,t}_{\cdot,x}}\big{\{}u_{0}(\xi(0))+\int_{0}^{t}L(\tau,\xi(\tau),\dot{\xi}(\tau))d\tau\big{\}}. (2.7)
Proposition 2.4.

[13] The following properties hold.

  1. (1)

    u(t,x)u(t,x) is a viscosity solution of (HJe).

  2. (2)

    u(t,x)u(t,x) is locally linear semiconcave on (0,+)×n(0,+\infty)\times\mathbb{R}^{n}. More precisely, for any 0<T1<T20<T_{1}<T_{2} and R>0R>0, there exists C(T1,T2,R)>0C(T_{1},T_{2},R)>0 such that u(t,x)u(t,x) is linearly semiconcave on (T1,T2)×B(0,R)(T_{1},T_{2})\times B(0,R) with semiconcavity constant C(T1,T2,R)C(T_{1},T_{2},R). Moreover, C(T1,T2,R)C(T_{1},T_{2},R) is continuous with respect to RR.

  3. (3)

    For any (t,x)(0,)×n(t,x)\in(0,\infty)\times\mathbb{R}^{n} and any minimizer ξΓ,x0,t\xi\in\Gamma^{0,t}_{\cdot,x} of (2.7), uu is differentiable at (τ,ξ(τ))(\tau,\xi(\tau)) for all τ(0,t)\tau\in(0,t).

Moreover, we have the following result

Proposition 2.5.

[13, Thm 6.4.9] For any (t,x)(0,+)×n(t,x)\in(0,+\infty)\times\mathbb{R}^{n}, (q,p)Du(t,x)(q,p)\in D^{*}u(t,x) if and only if there exists a minimizer γΓ,x0,t\gamma\in\Gamma^{0,t}_{\cdot,x} of (2.7) such that and p=Lv(t,x,γ˙(t))p=L_{v}(t,x,\dot{\gamma}(t)) and q=H(t,x,p)q=-H(t,x,p).

Lemma 2.6.

(proved in Appendix B)  For any fixed T>0T>0, there exists F0(T)0F_{0}(T)\geqslant 0 such that uu is a Lipschitz function on (0,T]×n(0,T]\times\mathbb{R}^{n} and Lip[u]F0(T)\mbox{\rm Lip}\,[u]\leqslant F_{0}(T).

Due to Lemma 2.6 and Lemma 2.2, we obtain the following estimation:

Corollary 2.7.

For any T>0T>0, 0t1<t2T0\leqslant t_{1}<t_{2}\leqslant T and x1nx_{1}\in\mathbb{R}^{n}, we have

Y(u(t2,),t1,x1,t2),\displaystyle Y(u(t_{2},\cdot),t_{1},x_{1},t_{2})\neq\emptyset,

and for any yt1,t2,x1Y(u(t2,),t1,x1,t2)y_{t_{1},t_{2},x_{1}}\in Y(u(t_{2},\cdot),t_{1},x_{1},t_{2}),

|yt1,t2,x1x1|λ2(T)(t2t1).|y_{t_{1},t_{2},x_{1}}-x_{1}|\leqslant\lambda_{2}(T)(t_{2}-t_{1}). (2.8)

where λ2(T):=λ1(T,F0(T))\lambda_{2}(T):=\lambda_{1}(T,F_{0}(T)) for T>0T>0 with λ1\lambda_{1} defined in Lemma 2.2 and F0F_{0} defined in Lemma 2.6.

2.3. Propagation of singularities

Recall that xnx\in\mathbb{R}^{n} is a cut point of uu if no backward calibrated curve ending at xx can be extended beyond xx. We denote by Cut(u)\mbox{\rm{Cut}}\,(u) the set of cut points of uu. It is well known that Sing(u)Cut(u)Sing(u)¯\mbox{\rm{Sing}}\,(u)\subset\mbox{\rm{Cut}}\,(u)\subset\overline{\mbox{\rm{Sing}}\,(u)}. In the following proposition 2.8, we construct a singular arc starting from any cut point of u(t,x)u(t,x).

Proposition 2.8.

Suppose LL is a Tonelli Lagrangian satisfying (L1)-(L3), HH is the associated Hamiltonian and u0:nu_{0}:\mathbb{R}^{n}\to\mathbb{R} is a Lipschitz function.Then for any fixed xnx\in\mathbb{R}^{n}and 0<t0<T0<t_{0}<T, there exist tx,T(0,T)t_{x,T}\in(0,T) which only depends on x,Tx,T such that for all t0t1Ttx,Tt_{0}\leqslant t_{1}\leqslant T-t_{x,T} and x1B(x,λ2(T)t1)x_{1}\in B(x,\lambda_{2}(T)t_{1}), there is a unique maximum point yt1,t,x1y_{t_{1},t,x_{1}} of u(t,)At1,t(x1,)u(t,\cdot)-A_{t_{1},t}(x_{1},\cdot) for t[t1,t1+tx,T]t\in[t_{1},t_{1}+t_{x,T}] and the curve

𝐲t1,t1+tx,T,x1(t):={x1ift=t1,yt1,t,x1ift(t1,t1+tx,T].\mathbf{y}_{t_{1},t_{1}+t_{x,T},x_{1}}(t):=\begin{cases}x_{1}\quad&\mbox{\rm{if}}\ \ t=t_{1},\\ y_{t_{1},t,x_{1}}\quad&\mbox{\rm{if}}\ \ t\in(t_{1},t_{1}+t_{x,T}].\end{cases} (2.9)

satisfies 𝐲t1,t1+tx,T,x1(t)B(x,λ2(T)T)\mathbf{y}_{t_{1},t_{1}+t_{x,T},x_{1}}(t)\in B(x,\lambda_{2}(T)T) for any t[t1,t1+tx,T]t\in[t_{1},t_{1}+t_{x,T}], where λ2\lambda_{2} is defined in Corollary 2.7. Moreover, if (t1,x1)Cut(u)(t_{1},x_{1})\in\mbox{\rm{Cut}}(u), then (t,𝐲(t))Sing(u)(t,\mathbf{y}(t))\in\mbox{\rm{Sing}}(u) for all t[t1,t1+tx,T]t\in[t_{1},t_{1}+t_{x,T}].

Proof.

For any fixed xn,T>0x\in\mathbb{R}^{n},T>0 , by proposition 2.4 (2), there exists C(x,T)>0C(x,T)>0 such that it is a semiconcavity constant for uu on [t0,t0+T]×B(x,λ2(T)T)[t_{0},t_{0}+T]\times B(x,\lambda_{2}(T)T). By (3) of Proposition A.2, there exists C2(x,T)>0C_{2}(x,T)>0 such that it is a uniformly convexity constant for At1,t(x1,)A_{t_{1},t}(x_{1},\cdot) on [t0,t0+T]×B(x,λ2(T)T)[t_{0},t_{0}+T]\times B(x,\lambda_{2}(T)T).

Therefore, u(t,)At1,t(x1,)u(t,\cdot)-A_{t_{1},t}(x_{1},\cdot) is strictly concave on B¯(x,λ2(T)t)\overline{B}(x,\lambda_{2}(T)t) for all t[t1,t1+tx,T]t\in[t_{1},t_{1}+t_{x,T}] provided that we further restrict tx,Tt_{x,T} in order to have

tx,T:=C2(x,T)2C(x,T).t_{x,T}:=\frac{C_{2}(x,T)}{2C(x,T)}.

We now proof that yt1,t,x1y_{t_{1},t,x_{1}} is a singular point of uu for every t(t1,t1+tx,T]t\in(t_{1},t_{1}+t_{x,T}]. Let ξt1,t,x1Γx1,yt1,t,x1t1,t\xi_{t_{1},t,x_{1}}\in\Gamma^{t_{1},t}_{x_{1},y_{t_{1},t,x_{1}}} be the unique minimizer for At1,t(x1,yt1,t,x1)A_{t_{1},t}(x_{1},y_{t_{1},t,x_{1}}) and let

pt1,t,x1(s):=Lv(s,ξt1,t,x1(s),ξ˙t1,t,x1(s)),s[t1,t1+tx,T],p_{t_{1},t,x_{1}}(s):=L_{v}(s,\xi_{t_{1},t,x_{1}}(s),\dot{\xi}_{t_{1},t,x_{1}}(s)),\quad s\in[t_{1},t_{1}+t_{x,T}],

be the associated dual arc. We claim that

pt1,t,x1(t)D+u(t,yt1,t,x1)\Du(t,yt1,t,x1)p_{t_{1},t,x_{1}}(t)\in D^{+}u(t,y_{t_{1},t,x_{1}})\backslash D^{*}u(t,y_{t_{1},t,x_{1}}) (2.10)

which in turn yields yt1,t,x1Sing(u)y_{t_{1},t,x_{1}}\in\mbox{\rm{Sing}}(u). Indeed, if pt1,t,x1(t)Du(t,yt1,t,x1)p_{t_{1},t,x_{1}}(t)\in D^{*}u(t,y_{t_{1},t,x_{1}}), then by Proposition 2.5, there would exist a C2C^{2} curve γt1,t,x1:(,t]n\gamma_{t_{1},t,x_{1}}:(-\infty,t]\rightarrow\mathbb{R}^{n} solving the minimum problem

minγW1,1([τ,t];n){τtL(s,γ(s),γ˙(s))𝑑s+u(γ(τ)):γ(t)=yt1,t,x1}\min_{\gamma\in W^{1,1}([\tau,t];\mathbb{R}^{n})}\left\{\int_{\tau}^{t}L(s,\gamma(s),\dot{\gamma}(s))\ ds+u(\gamma(\tau)):\gamma(t)=y_{t_{1},t,x_{1}}\right\} (2.11)

for all τt\tau\leqslant t. It is easily to checked that γt1,t,x1\gamma_{t_{1},t,x_{1}} and ξt1,t,x1\xi_{t_{1},t,x_{1}} coincide on [t1,t][t_{1},t] since both of them are extremal curves for LL and satisfy the same endpoint condition at γt1,t,x\gamma_{t_{1},t,x} i.e.

Lv(t,ξt1,t,x1(t),ξ˙t1,t,x1(t))=pt1,t,x1(t)=Lv(t,γt1,t,x(t),γ˙t1,t,x(t)).L_{v}(t,\xi_{t_{1},t,x_{1}}(t),\dot{\xi}_{t_{1},t,x_{1}}(t))=p_{t_{1},t,x_{1}}(t)=L_{v}(t,\gamma_{t_{1},t,x}(t),\dot{\gamma}_{t_{1},t,x}(t)).

This leads to a contradiction since (t1,x1)Cut(u)(t_{1},x_{1})\in\mbox{\rm{Cut}}(u) while uu should be smooth at (t1,γt1,t,x1(t1))(t_{1},\gamma_{t_{1},t,x_{1}}(t_{1})) and γt1,t,x(τ)\gamma_{t_{1},t,x}(\tau) is a backward calibrated curve for τ[t1,t]\tau\in[t_{1},t]. Thus, (2.10) holds true and (t,𝐲t1,t,x1(t))Sing(u)(t,\mathbf{y}_{t_{1},t,x_{1}}(t))\in\mbox{\rm{Sing}}(u) for all t(t1,t1+tx,T]t\in(t_{1},t_{1}+t_{x,T}]. ∎

By Proposition 2.8, for any t>0t>0 and xnx^{\prime}\in\mathbb{R}^{n} , there exist a t>tt^{\prime}>t which depends on xx^{\prime} such that argsupyn{u(s,y)At,s(x,y)}\arg\sup_{y\in\mathbb{R}^{n}}\{u(s,y)-A_{t,s}(x^{\prime},y)\} is singleton for any t<stt<s\leqslant t^{\prime}. We can denote that

Y(t,t,x):=Y(u(t,),t,t,x)=argsupyn{u(t,y)At,t(x,y)}=yt,t,x.Y(t,t^{\prime},x^{\prime}):=Y\big{(}u(t^{\prime},\cdot),t,t^{\prime},x^{\prime}\big{)}=\arg\sup_{y\in\mathbb{R}^{n}}\{u(t^{\prime},y)-A_{t,t^{\prime}}(x^{\prime},y)\}=y_{t,t^{\prime},x^{\prime}}.

By (2.8), it implies that

|Y(t,t,x)x|λ2(T)(tt)t0<ttT.|Y(t,t^{\prime},x^{\prime})-x^{\prime}|\leqslant\lambda_{2}(T)(t^{\prime}-t)\quad\forall\ t_{0}<t\leqslant t^{\prime}\leqslant T. (2.12)
Theorem 2.9.

Suppose LL is a Tonelli Lagrangian satisfying (L1)-(L3), HH is the associated Hamiltonian and u0:nu_{0}:\mathbb{R}^{n}\to\mathbb{R} is a Lipschitz function. Then for any fixed (t0,x)Cut(u)(t_{0},x)\in\mbox{\rm{Cut}}(u) and T>t0T>t_{0}, there exists a curve 𝐱:[t0,+)n\mathbf{x}:[t_{0},+\infty)\to\mathbb{R}^{n} with 𝐱(t0)=x\mathbf{x}(t_{0})=x, such that (s,𝐱(s))Sing(u)(s,\mathbf{x}(s))\in\mbox{\rm{Sing}}(u) for all s[t0,+)s\in[t_{0},+\infty).

Proof.

For any fixed (t0,x)Sing(u)(t_{0},x)\in\mbox{\rm{Sing}}(u) and T>0T>0, we denote that

{Ωn}n:=B(x,λ2(nT)nT),\{\Omega_{n}\}_{n\in\mathbb{N}}:=B(x,\lambda_{2}(nT)\ nT),

with Ω¯nΩn+1\overline{\Omega}_{n}\subset\Omega_{n+1} and Ω¯n\overline{\Omega}_{n} is compact for all nn, in addition, n=nΩ¯n\mathbb{R}^{n}=\bigcup_{n}\overline{\Omega}_{n}.

𝐒𝐭𝐞𝐩𝐈\mathbf{Step\ I} :Uniform Lipschitz estimation of connections of YY.

For any s>0s>0, there are a sequence of points {xj}1jn\{x_{j}\}_{1\leqslant j\leqslant n} and time {sj}1jn\{s_{j}\}_{1\leqslant j\leqslant n} with xj=Y(sj1,sj,xj1)x_{j}=Y(s_{j-1},s_{j},x_{j-1}) for any 1jn1\leqslant j\leqslant n and x0=xx_{0}=x, for t0s1s2snssTTt_{0}\leqslant s_{1}\leqslant s_{2}\leqslant\dots\leqslant s_{n}\leqslant s\leqslant\big{\lceil}\frac{s}{T}\big{\rceil}T.

By (2.12) , we have that

|Y(sn,s,xn)x|\displaystyle|Y(s_{n},s,x_{n})-x|\leqslant |Y(sn,s,xn)xn|+j=1n|Y(sj1,sj,xj1)|\displaystyle\,|Y(s_{n},s,x_{n})-x_{n}|+\sum_{j=1}^{n}|Y(s_{j-1},s_{j},x_{j-1})|
\displaystyle\leqslant λ2(sTT)(ssn+j=1n(sjsj1))\displaystyle\,\lambda_{2}\left(\big{\lceil}\frac{s}{T}\big{\rceil}T\right)\left(s-s_{n}+\sum^{n}_{j=1}(s_{j}-s_{j-1})\right)
=\displaystyle= λ2(sTT)(st0),\displaystyle\,\lambda_{2}\left(\big{\lceil}\frac{s}{T}\big{\rceil}T\right)(s-t_{0}),

which means that for any 1jn1\leqslant j\leqslant n,

Y(sj,s,xj)ΩsT,s[sj,sj+1],Y(s_{j},s,x_{j})\in\Omega_{\lceil\frac{s}{T}\rceil},\quad s\in[s_{j},s_{j+1}],

i.e. |Y(sj,s,xj)x|λ2(sTT)(st0)|Y(s_{j},s,x_{j})-x|\leqslant\lambda_{2}\left(\big{\lceil}\frac{s}{T}\big{\rceil}T\right)(s-t_{0}) for any 1jn1\leqslant j\leqslant n.

𝐒𝐭𝐞𝐩𝐈𝐈\mathbf{Step\ II}: Construction of curve 𝐱\mathbf{x}.

For x1:=xCut(u)Ω1x_{1}:=x\in\mbox{\rm{Cut}}(u)\cap\Omega_{1} without loss of generality, then there exists t1:=tx,T>0t_{1}:=t_{x,T}>0 such that Y(t0,,x)Y(t_{0},\cdot,x) is defined on [t0,t0+t1][t_{0},t_{0}+t_{1}] by Proposition 2.8. One can extend YY by induction.

For Y(t0,t0+t1,x)Ω1Y(t_{0},t_{0}+t_{1},x)\in\Omega_{1}, then we define Y(t0,s,x)=Y(t0+t1,s,Y(t0,t0+t1,x))Y(t_{0},s,x)=Y(t_{0}+t_{1},s,Y(t_{0},t_{0}+t_{1},x)) for all s[t0+t1,t0+2t1]s\in[t_{0}+t_{1},t_{0}+2t_{1}] ; inductively, if Y(t0,,x)Y(t_{0},\cdot,x) is defined on [t0,t0+kt1][t_{0},t_{0}+kt_{1}] such that Y(t0,t0+kt1,x)Ω1Y(t_{0},t_{0}+kt_{1},x)\in\Omega_{1} , then we define that Y(t0,s,x)=Y(t0+kt1,s,Y(t0,t0+kt1,x))Y(t_{0},s,x)=Y(t_{0}+kt_{1},s,Y(t_{0},t_{0}+kt_{1},x)) for all s[t0+kt1,t0+(k+1)t1]s\in[t_{0}+kt_{1},t_{0}+(k+1)t_{1}]. Now, let

k1=Tt0t1,k_{1}=\Big{\lfloor}\frac{T-t_{0}}{t_{1}}\Big{\rfloor}, (2.13)

then t0+k1t1Tt_{0}+k_{1}t_{1}\leqslant T which implies Y(t0,s,x)Ω1Y(t_{0},s,x)\in\Omega_{1} for any s[t0,t0+k1t1]s\in[t_{0},t_{0}+k_{1}t_{1}] by Step I .

In a similar way, x2:=Y(t0,t0+k1t1,x)Ω1Sing(u)Ω2Sing(u)x_{2}:=Y(t_{0},t_{0}+k_{1}t_{1},x)\in\Omega_{1}\cap\mbox{\rm{Sing}}(u)\subset\Omega_{2}\cap\mbox{\rm{Sing}}(u), we define

k2=2Tk1t1t0t2,k_{2}=\Big{\lfloor}\frac{2T-k_{1}t_{1}-t_{0}}{t_{2}}\Big{\rfloor},

where t2:=tx,2Tt1t_{2}:=t_{x,2T}\leqslant t_{1} is determined by applying Proposition 2.8 to Ω2\Omega_{2}.We also conclude that t0+k1t1+k2t22Tt_{0}+k_{1}t_{1}+k_{2}t_{2}\leqslant 2T which implies that Y(t0,s,x)Ω2Y(t_{0},s,x)\in\Omega_{2} for all s[t0,t0+k1t1+k2t2]s\in[t_{0},t_{0}+k_{1}t_{1}+k_{2}t_{2}] by Step I.

Therefore, by induction, for any ii\in\mathbb{N}, there exists ti:=tx,iTti1t_{i}:=t_{x,iT}\leqslant t_{i-1} is determined by applying Proposition 2.8 to Ωi\Omega_{i} with 0<titi10<t_{i}\leqslant t_{i-1} , let

ki=iTj=1i1kjtjt0ti,k_{i}=\Big{\lfloor}\frac{iT-\sum_{j=1}^{i-1}k_{j}t_{j}-t_{0}}{t_{i}}\Big{\rfloor}, (2.14)

We also conclude that t0+j=1ikjtjiTt_{0}+\sum_{j=1}^{i}k_{j}t_{j}\leqslant iT which implies that Y(t0,s,x)ΩiY(t_{0},s,x)\in\Omega_{i} for all s[t0,t0+j=1ikjtj]s\in[t_{0},t_{0}+\sum_{j=1}^{i}k_{j}t_{j}] by Step I.

Denote that xi:=Y(t0,t0+j=0i1kjtj,x)Ωi1Sing(u)ΩiSing(u).x_{i}:=Y(t_{0},t_{0}+\sum_{j=0}^{i-1}k_{j}t_{j},x)\in\Omega_{i-1}\cap\mbox{\rm{Sing}}(u)\subset\Omega_{i}\cap\mbox{\rm{Sing}}(u). This makes us to define an arc 𝐱:[0,t¯)n\mathbf{x}:[0,\overline{t})\rightarrow\mathbb{R}^{n} by

𝐱(s):=Y(t0,s,x)=Y(t0+j=0i1kjtj,s,xi)s[t0+j=0i1kjtj,t0+j=0ikjtj]\mathbf{x}(s):=Y(t_{0},s,x)=Y(t_{0}+\sum_{j=0}^{i-1}k_{j}t_{j},s,x_{i})\quad\forall s\in\Big{[}t_{0}+\sum_{j=0}^{i-1}k_{j}t_{j},t_{0}+\sum_{j=0}^{i}k_{j}t_{j}\Big{]}

where t¯:=t0+j=0kjtj\overline{t}:=t_{0}+\sum^{\infty}_{j=0}k_{j}t_{j}. It is clear that 𝐱\mathbf{x} is a generalized characteristic defined on [0,t¯)[0,\overline{t}) and 𝐱(s)Sing(u)\mathbf{x}(s)\in\mbox{\rm{Sing}}(u) for all s[0,t¯)s\in[0,\overline{t}), by Proposition 2.8.

𝐒𝐭𝐞𝐩𝐈𝐈𝐈\mathbf{Step\ III}: Estimation of time t¯\overline{t}.

To finish the proof, we only need to show that t¯=\overline{t}=\infty. Indeed, since T>t1t2t3T>t_{1}\geqslant t_{2}\geqslant t_{3}\geqslant\dots, we have that

t¯>\displaystyle\overline{t}> t0+j=1nkjtj=t0+j=1n1kjtj+nTj=1n1kjtjt0tntn\displaystyle\,t_{0}+\sum^{n}_{j=1}k_{j}t_{j}=t_{0}+\sum^{n-1}_{j=1}k_{j}t_{j}+\Big{\lfloor}\frac{nT-\sum_{j=1}^{n-1}k_{j}t_{j}-t_{0}}{t_{n}}\Big{\rfloor}t_{n}
\displaystyle\geqslant t0+j=1n1kjtj+(nTj=1n1kjtjt0tn1)tn\displaystyle\,t_{0}+\sum^{n-1}_{j=1}k_{j}t_{j}+\left(\frac{nT-\sum_{j=1}^{n-1}k_{j}t_{j}-t_{0}}{t_{n}}-1\right)t_{n}
=\displaystyle= nTtnnTt1asn.\displaystyle\,nT-t_{n}\geqslant nT-t_{1}\rightarrow\infty\quad\text{as}\ n\rightarrow\infty.

Therefore t¯=\overline{t}=\infty. ∎

Remark 2.10.

Example 5.6.7 of [13] showed that there exists a counterexample for global propagation of singularities without condition (L2).

2.4. Local Lipschitz of singular curve

  1. (A)

    For any given T+T\in\mathbb{R}^{+}, there exists K(T)+K(T)\in\mathbb{R}^{+} such that viscosity solution u(t,x)u(t,x) of (HJe) is differentiable on (t,x)[0,T]×n(t,x)\in[0,T]\times\mathbb{R}^{n} and

    |Dtu(t,x)Dsu(s,y)|K(T)(|ts|+|xy|),x,yn,s,t[0,T].|D_{t}u(t,x)-D_{s}u(s,y)|\leqslant K(T)\Big{(}|t-s|+|x-y|\Big{)},\quad\forall x,y\in\mathbb{R}^{n},s,t\in[0,T].
Theorem 2.11.

If condition (A) holds, then for any T>0T>0 and 𝐱(s)\mathbf{x}(s) (see Theorem 2.9) is a Lipschitz curve on s[0,T]s\in[0,T].

The proof of the theorem above is a direct consequence of following Lemma.

Lemma 2.12.

For any T>0T>0 and (t0,x)[0,T]×n(t_{0},x)\in[0,T]\times\mathbb{R}^{n}, let tx,Tt_{x,T} and 𝐲:[t0,t0+tx,T]n\mathbf{y}:[t_{0},t_{0}+t_{x,T}]\to\mathbb{R}^{n} be given by Proposition 2.8. If condition (A) holds, then 𝐲\mathbf{y} is Lipschitz on [t0,t0+tx,T][t_{0},t_{0}+t_{x,T}].

Proof.

Without loss of generality, we assume that t0=0t_{0}=0. Let ξt:=ξ0,t,xΓx,𝐲(t)0,t\xi_{t}:=\xi_{0,t,x}\in\Gamma_{x,\mathbf{y}(t)}^{0,t}, ξs:=ξ0,s,xΓx,𝐲(s)0,s\xi_{s}:=\xi_{0,s,x}\in\Gamma_{x,\mathbf{y}(s)}^{0,s} and η:=ξ0,t,xΓx,𝐲(s)0,t\eta:=\xi_{0,t,x}\in\Gamma_{x,\mathbf{y}(s)}^{0,t} be minimizers for A0,t(x,𝐲(t)),A0,s(x,𝐲(s))A_{0,t}(x,\mathbf{y}(t)),A_{0,s}(x,\mathbf{y}(s)) and A0,t(x,𝐲(s))A_{0,t}(x,\mathbf{y}(s)) respectively. Setting pt=Lv(t,ξt(t),ξ˙t(t)),ps=Lv(s,ξs(s),ξ˙s(s))p_{t}=L_{v}(t,\xi_{t}(t),\dot{\xi}_{t}(t)),p_{s}=L_{v}(s,\xi_{s}(s),\dot{\xi}_{s}(s)) and p=Lv(t,η(t),η˙(t))p=L_{v}(t,\eta(t),\dot{\eta}(t)), we have (qs,ps)D+u(s,𝐲(s))(q_{s},p_{s})\in D^{+}u(s,\mathbf{y}(s)) and (qt,pt)D+u(t,𝐲(t))(q_{t},p_{t})\in D^{+}u(t,\mathbf{y}(t)). Hence, by Proposition A.2, there exists C3(x,T)>0C_{3}(x,T)>0 such that

C2(x,T)ts|𝐲(t)𝐲(s)|2\displaystyle\,\frac{C_{2}(x,T)}{t-s}|\mathbf{y}(t)-\mathbf{y}(s)|^{2}
\displaystyle\leqslant ptp,𝐲(t)𝐲(s)=ptps,𝐲(t)𝐲(s)+psp,𝐲(t)𝐲(s)\displaystyle\,\langle p_{t}-p,\mathbf{y}(t)-\mathbf{y}(s)\rangle=\langle p_{t}-p_{s},\mathbf{y}(t)-\mathbf{y}(s)\rangle+\langle p_{s}-p,\mathbf{y}(t)-\mathbf{y}(s)\rangle
\displaystyle\leqslant C3(x,T)ts|𝐲(t)𝐲(s)||ts|+ptps,𝐲(t)𝐲(s)\displaystyle\,\frac{C_{3}(x,T)}{t-s}\cdot|\mathbf{y}(t)-\mathbf{y}(s)|\cdot|t-s|+\langle p_{t}-p_{s},\mathbf{y}(t)-\mathbf{y}(s)\rangle
+qtqs,tsqtqs,ts\displaystyle\,\quad\quad\quad\quad+\langle q_{t}-q_{s},t-s\rangle-\langle q_{t}-q_{s},t-s\rangle
\displaystyle\leqslant C3(x,T)ts|𝐲(t)𝐲(s)||ts|+C1(x,T)(|𝐲(t)𝐲(s)|2+|ts|2)\displaystyle\,\frac{C_{3}(x,T)}{t-s}\cdot|\mathbf{y}(t)-\mathbf{y}(s)|\cdot|t-s|+C_{1}(x,T)(|\mathbf{y}(t)-\mathbf{y}(s)|^{2}+|t-s|^{2})
qtqs,ts\displaystyle\,\quad\quad\quad\quad\quad\quad\quad\quad-\langle q_{t}-q_{s},t-s\rangle

where C2(x,T)>0C_{2}(x,T)>0 is a uniformly convexity constant in Proposition A.2 (3).

Actually, by condition (A), u(t,x)u(t,x) is differentiable with respect to tt, and

|qtqs|=|Dtu(t,x)Dsu(s,x)|K(T)(|ts|+|𝐲(t)𝐲(s)|).|q_{t}-q_{s}|=|D_{t}u(t,x)-D_{s}u(s,x)|\leqslant K(T)\big{(}|t-s|+|\mathbf{y}(t)-\mathbf{y}(s)|\big{)}.

Therefore,

C2(x,T)ts|𝐲(t)𝐲(s)|2\displaystyle\frac{C_{2}(x,T)}{t-s}|\mathbf{y}(t)-\mathbf{y}(s)|^{2}\leqslant C3(x,T)ts|𝐲(t)𝐲(s)||ts|+C1(x,T)(|𝐲(t)𝐲(s)|2+|ts|2)\displaystyle\,\frac{C_{3}(x,T)}{t-s}\cdot|\mathbf{y}(t)-\mathbf{y}(s)|\cdot|t-s|+C_{1}(x,T)(|\mathbf{y}(t)-\mathbf{y}(s)|^{2}+|t-s|^{2})
+K(T)|𝐲(t)𝐲(s)||ts|+K(T)|ts|2.\displaystyle\,\quad\quad\quad\quad+K(T)\cdot|\mathbf{y}(t)-\mathbf{y}(s)|\cdot|t-s|+K(T)|t-s|^{2}.

That is,

(C2(x,T)tsC1(x,T))|𝐲(t)𝐲(s)ts|2(C3(x,T)ts+K(T))|𝐲(t)𝐲(s)ts|C1+K(T).\displaystyle\Big{(}\frac{C_{2}(x,T)}{t-s}-C_{1}(x,T)\Big{)}\Big{|}\frac{\mathbf{y}(t)-\mathbf{y}(s)}{t-s}\Big{|}^{2}-\Big{(}\frac{C_{3}(x,T)}{t-s}+K(T)\Big{)}\Big{|}\frac{\mathbf{y}(t)-\mathbf{y}(s)}{t-s}\Big{|}\leqslant C_{1}+K(T).

Let tst-s be sufficiently small such that

tsC2(x,T)2C1(x,T),t-s\leqslant\frac{C_{2}(x,T)}{2C_{1}(x,T)},

then there exists a constant C4C_{4} which only depends on x,Tx,T such that

|𝐲(t)𝐲(s)ts|C4(x,T).\Big{|}\frac{\mathbf{y}(t)-\mathbf{y}(s)}{t-s}\Big{|}\leqslant C_{4}(x,T).

More precisely, we can take C4(x,T):=C3(x,T)C2(x,T)+K(T)+C1(x,T)+K(T)2C1(x,T)C_{4}(x,T):=\frac{C_{3}(x,T)}{C_{2}(x,T)}+\frac{K(T)+\sqrt{C_{1}(x,T)+K(T)}}{2C_{1}(x,T)}. ∎

3. Global propagation of singularities for discounted Hamiltonian

3.1. Global propagation of singularities for discounted Hamiltonian

For λ>0\lambda>0, we consider the Hamilton-Jacobi equation with discounted factor

λv(x)+H(x,Dv(x))=0,xn.\lambda v(x)+H(x,Dv(x))=0,\quad x\in\mathbb{R}^{n}. (HJλ)

where HH is a Tonelli Hamiltonian.

Lemma 3.1.

[15, Proposition 3.3] v(x)v(x) is a viscosity solution of (HJλ) if and only if u(t,x)=eλtv(x)u(t,x)=e^{\lambda t}v(x) is a viscosity solution of the Hamilton-Jacobi equation

{Dtu+H^(t,x,Dxu)=0,(t,x)(0,+)×nu(0,x)=v(x),xn,\begin{cases}D_{t}u+\widehat{H}(t,x,D_{x}u)=0,&(t,x)\in(0,+\infty)\times\mathbb{R}^{n}\\ u(0,x)=v(x),&x\in\mathbb{R}^{n},\end{cases} (3.1)

where H^(t,x,p)=eλtH(x,eλtp)\widehat{H}(t,x,p)=e^{\lambda t}H(x,e^{-\lambda t}p). Moreover, for any (t,x)(0+)×n(t,x)\in(0+\infty)\times\mathbb{R}^{n}, we have

xSing(v)(t,x)Sing(u),xCut(v)(t,x)Cut(u).x\in\mbox{\rm{Sing}}(v)\Leftrightarrow(t,x)\in\mbox{\rm{Sing}}(u),\quad x\in\mbox{\rm{Cut}}(v)\Leftrightarrow(t,x)\in\mbox{\rm{Cut}}(u). (3.2)
Remark 3.2.

Since γ:[a,b]n\gamma:[a,b]\to\mathbb{R}^{n} is a calibrated curve of equation (HJλ) with Hamiltonian HH if and only if γ:[a,b]n\gamma:[a,b]\to\mathbb{R}^{n} is a calibrated curve of equation (3.1) with Hamiltonian H^\widehat{H}, then by the definition of Cut(u)\mbox{\rm{Cut}}(u) in Definition 3.4, xCut(v)x\in\mbox{\rm{Cut}}(v) and (t,x)Cut(u)(t,x)\in\mbox{\rm{Cut}}(u) are equivalent.

Theorem 3.3.

Let HH be a Tonelli Hamiltonian and λ>0\lambda>0. Suppose v:nv:\mathbb{R}^{n}\to\mathbb{R} is the Lipschitz continuous viscosity solution of (HJλ). Then for any fixed xCut(v)x\in\mbox{\rm{Cut}}(v), there exists a locally Lipschitz curve 𝐱:[0,+)n\mathbf{x}:[0,+\infty)\rightarrow\mathbb{R}^{n} with 𝐱(0)=x\mathbf{x}(0)=x, such that 𝐱(τ)Sing(v)\mathbf{x}(\tau)\in\mbox{\rm{Sing}}(v) for all τ[0,+)\tau\in[0,+\infty).

Proof.

By using the variable transformation s=τ+1[1,+)s=\tau+1\in[1,+\infty), we only need to find a locally Lipschitz curve 𝐱:[1,+)n\mathbf{x}:[1,+\infty)\rightarrow\mathbb{R}^{n} with 𝐱(1)=x\mathbf{x}(1)=x, such that 𝐱(s)Sing(v)\mathbf{x}(s)\in\mbox{\rm{Sing}}(v) for all s[1,+)s\in[1,+\infty). Due to (3.2), xCut(v)x\in\mbox{\rm{Cut}}(v) implies (1,x)Cut(u)(1,x)\in\mbox{\rm{Cut}}(u). It is easy to check that L^(t,x,v)=eλtL(x,v)\widehat{L}(t,x,v)=e^{\lambda t}L(x,v) satisfies (L1)-(L3). Therefore, by Theorem 2.9, there exists a curve 𝐱:[1,+)n\mathbf{x}:[1,+\infty)\rightarrow\mathbb{R}^{n} with 𝐱(1)=x\mathbf{x}(1)=x, such that (s,𝐱(s))Sing(u)(s,\mathbf{x}(s))\in\mbox{\rm{Sing}}(u) for all s(1,+)s\in(1,+\infty), that is, 𝐱(τ)Sing(v)\mathbf{x}(\tau)\in\mbox{\rm{Sing}}(v) for all τ(0,+)\tau\in(0,+\infty).

It remains to show that the curve 𝐱(τ):[0,+)n\mathbf{x}(\tau):[0,+\infty)\rightarrow\mathbb{R}^{n} is locally Lipschitz. Notice that u(t,x)=eλtv(x)u(t,x)=e^{\lambda t}v(x) is differentiable with respect to tt and

Dtu(t,x)=λeλtv(x)=λu(t,x),(t,x)(0,+)×n.\displaystyle D_{t}u(t,x)=\lambda e^{\lambda t}v(x)=\lambda u(t,x),\qquad(t,x)\in(0,+\infty)\times\mathbb{R}^{n}.

For any (t,x)(0,+)×n(t,x)\in(0,+\infty)\times\mathbb{R}^{n} and (s,y)(0,+)×n(s,y)\in(0,+\infty)\times\mathbb{R}^{n}, by Lemma 2.6, we have

|Dtu(t,x)Dsu(s,y)|=|λu(t,x)λu(s,y)|F0(T)(|ts|+|xy|),\displaystyle|D_{t}u(t,x)-D_{s}u(s,y)|=|\lambda u(t,x)-\lambda u(s,y)|\leqslant F_{0}(T)(|t-s|+|x-y|),

which implies condition (A) holds. Hence 𝐱:[0,+)n\mathbf{x}:[0,+\infty)\rightarrow\mathbb{R}^{n} is locally Lipschitz by Lemma 2.12. ∎

3.2. Homotopy equivalence

Now, suppose u:nu:\mathbb{R}^{n}\to\mathbb{R} is the Lipschitz viscosity solution of

λu(x)+H(x,du(x))=0,xn,\lambda u(x)+H(x,du(x))=0,\qquad x\in\mathbb{R}^{n}, (HJλ)

where HH is a Tonelli Hamiltonian and λ>0\lambda>0.

Definition 3.4.

(Aubry set): We define (u)\mathcal{I}(u)111If MM is compact, (u)\mathcal{I}(u) is not empty and can be characterized by conjugate pairs for contact Hamiltonian systems with increasing condition in [25]. For noncompact case, the question is still open if (u)\mathcal{I}(u)\neq\emptyset . , the Aubry set of uu, as

(u)={xn: there exists a calibrated curve γ:(,+)n with γ(0)=x}\displaystyle\mathcal{I}(u)=\{x\in\mathbb{R}^{n}:\text{ there exists a calibrated curve }\gamma:(-\infty,+\infty)\to\mathbb{R}^{n}\text{ with }\gamma(0)=x\}

In general we have the following inclusions:

Sing(u)Cut(u)n(u),Sing(u)Cut(u)Sing(u)¯.\mbox{\rm{Sing}}\,(u)\subset\mbox{\rm{Cut}}\,(u)\subset\mathbb{R}^{n}\setminus\mathcal{I}(u),\quad\mbox{\rm{Sing}}\,(u)\subset\mbox{\rm{Cut}}\,(u)\subset\overline{\mbox{\rm{Sing}}\,(u)}.
Theorem 3.5.

The inclusions

Sing(u)Cut(u)(n(u))Sing(u)¯n(u)\mbox{\rm{Sing}}\,(u)\subset\mbox{\rm{Cut}}\,(u)\subset\Big{(}\mathbb{R}^{n}\setminus\mathcal{I}(u)\Big{)}\cap\overline{\mbox{\rm{Sing}}\,(u)}\subset\mathbb{R}^{n}\setminus\mathcal{I}(u)

are all homotopy equivalences.

This theorem obviously implies the following corollary (see, for instance, [18])

Corollary 3.6.

For every connected component CC of n(u)\mathbb{R}^{n}\setminus\mathcal{I}(u), these three intersections Sing(u)C\mbox{\rm{Sing}}\,(u)\cap C, Cut(u)C\mbox{\rm{Cut}}\,(u)\cap C and Sing(u)¯C\overline{\mbox{\rm{Sing}}\,(u)}\cap C are path connected.

Theorem 3.7.

[8, Thm. 1.3] The spaces Sing(u)\mbox{\rm{Sing}}\,(u) and Cut(u)\mbox{\rm{Cut}}\,(u) are locally contractible.

The proof of Theorem 3.5 and 3.7 needs the following Lemma.

Lemma 3.8.

There exists a continuous homotopy F:n×[0,+)nF:\mathbb{R}^{n}\times[0,+\infty)\to\mathbb{R}^{n} with the following properties:

  1. (a)

    for all xnx\in\mathbb{R}^{n}, we have F(x,0)=xF(x,0)=x;

  2. (b)

    if F(x,s)Sing(u)F(x,s)\notin\mbox{\rm{Sing}}(u) for some s>0s>0 and xnx\in\mathbb{R}^{n}, then the curve σF(x,σ)\sigma\mapsto F(x,\sigma) is calibrated on [0,s][0,s];

  3. (c)

    if there exists a calibrated curve γ:[0,s]n\gamma:[0,s]\to\mathbb{R}^{n} with γ(0)=x\gamma(0)=x, then σF(x,σ)=γ(σ)\sigma\mapsto F(x,\sigma)=\gamma(\sigma), for every σ[0,s]\sigma\in[0,s].

The proof of Lemma 3.8 is in Appendix C. These properties imply:

Lemma 3.9.
  1. (1)

    F(Cut(u)×(0,+))Sing(u)F(\mbox{\rm Cut}\,(u)\times(0,+\infty))\subset\mbox{\rm{Sing}}(u);

  2. (2)

    if F(x,s)Sing(u)F(x,s)\notin\mbox{\rm{Sing}}\,(u) for all s[0,+)s\in[0,+\infty), then x(u)x\in\mathcal{I}(u) and sF(x,s)s\mapsto F(x,s), s[0,+)s\in[0,+\infty) is a forward calibrated curve with F(x,0)=xF(x,0)=x;

  3. (3)

    if x(u)x\notin\mathcal{I}(u), then F(x,s)(u)F(x,s)\notin\mathcal{I}(u) for every s[0,+)s\in[0,+\infty).

Now, for xnx\in\mathbb{R}^{n}, we define τ(x)\tau(x) to be the supremum of the t0t\geqslant 0 such that there exists a calibrated curve γ:[0,t]n\gamma:[0,t]\to\mathbb{R}^{n} with γ(0)=x\gamma(0)=x.

Lemma 3.10.
  1. (i)

    τ(x)=0\tau(x)=0 if and only if xCut(u)x\in\mbox{\rm Cut}\,(u);

  2. (ii)

    τ(x)=+\tau(x)=+\infty if and only if x(u)x\in\mathcal{I}(u);

  3. (iii)

    the function τ\tau is upper semi-continuous.

Proof.

(i) and (ii) follows directly from the definition of Cut(u)\mbox{\rm Cut}\,(u) and (u)\mathcal{I}(u). It remains to prove (iii). Indeed, we only need to prove that for any τ>0\tau^{\prime}>0 the set {xn:τ(x)τ}\{x\in\mathbb{R}^{n}:\tau(x)\geqslant\tau^{\prime}\} is closed. Take any sequence xix_{i} such that τ(xi)τ\tau(x_{i})\geqslant\tau^{\prime} and xix0x_{i}\to x_{0}, and let γi:[0,τ]n\gamma_{i}:[0,\tau^{\prime}]\to\mathbb{R}^{n}, γi(0)=xi\gamma_{i}(0)=x_{i} be the associated calibrated curves. By taking a subsequence, we can assume that

limiDu(xi)=p0Du(x0).\displaystyle\lim_{i\to\infty}Du(x_{i})=p_{0}\in D^{*}u(x_{0}).

Notice that γi:[0,τ]n\gamma_{i}:[0,\tau^{\prime}]\to\mathbb{R}^{n} is the solution of (2.2) with initial condition γi(0)=xi\gamma_{i}(0)=x_{i}, pi(0)=Du(xi)p_{i}(0)=Du(x_{i}). Let γ0:[0,τ]n\gamma_{0}:[0,\tau^{\prime}]\to\mathbb{R}^{n} be the solution of (2.2) with initial condition γ0(0)=x0\gamma_{0}(0)=x_{0}, p0(0)=p0p_{0}(0)=p_{0}. It follows that γi\gamma_{i} converges to γ0\gamma_{0} in C2C^{2} topology. Thus, we have

eλτu(γ0(τ))\displaystyle e^{\lambda\tau^{\prime}}u(\gamma_{0}(\tau^{\prime})) =limieλτu(γi(τ))=limiu(γi(0))+0τeλtL(γi(t),γ˙i(t))𝑑t\displaystyle=\lim_{i\to\infty}e^{\lambda\tau^{\prime}}u(\gamma_{i}(\tau^{\prime}))=\lim_{i\to\infty}u(\gamma_{i}(0))+\int_{0}^{\tau^{\prime}}e^{\lambda t}L(\gamma_{i}(t),\dot{\gamma}_{i}(t))\ dt
=u(γ0(0))+0τeλtL(γ0(t),γ˙0(t))𝑑t.\displaystyle=u(\gamma_{0}(0))+\int_{0}^{\tau^{\prime}}e^{\lambda t}L(\gamma_{0}(t),\dot{\gamma}_{0}(t))\ dt.

This implies γ0:[0,τ]n\gamma_{0}:[0,\tau^{\prime}]\to\mathbb{R}^{n} is a calibrated curve and τ(x0)τ\tau(x_{0})\geqslant\tau^{\prime}. Therefore, the set {xn:τ(x)τ}\{x\in\mathbb{R}^{n}:\tau(x)\geqslant\tau^{\prime}\} is closed and the function τ\tau is upper semi-continuous. ∎

Proof of Theorem 3.5.

By Lemma 3.10, the function τ\tau is upper semi-continuous and finite on n(u)\mathbb{R}^{n}\setminus\mathcal{I}(u). Thus, by Proposition 7.20 in [3], we can find a continuous function α:n(u)(0,+)\alpha:\mathbb{R}^{n}\setminus\mathcal{I}(u)\to(0,+\infty) with α>τ\alpha>\tau on n(u)\mathbb{R}^{n}\setminus\mathcal{I}(u). We now define G:(n(u))×[0,1]n(u)G:(\mathbb{R}^{n}\setminus\mathcal{I}(u))\times[0,1]\to\mathbb{R}^{n}\setminus\mathcal{I}(u) by

G(x,s)=F(x,sα(x)).G(x,s)=F(x,s\alpha(x)).

Due to Lemma 3.8, Lemma 3.9 and the continuity of α\alpha, the map G(x,s)G(x,s) is a homotopy of n(u)\mathbb{R}^{n}\setminus\mathcal{I}(u) into itself, such that G(n(u),1)Sing(u)G(\mathbb{R}^{n}\setminus\mathcal{I}(u),1)\subset\mbox{\rm Sing}\,(u) and G(Cut(u),(0,1])Sing(u)G(\mbox{\rm Cut}\,(u),(0,1])\subset\mbox{\rm Sing}\,(u). Therefore, the time one map of GG gives a homotopy inverse for each one of the inclusions

Sing(u)Cut(u)Sing(u)¯(n(u))n(u).\displaystyle\mbox{\rm Sing}\,(u)\subset\mbox{\rm Cut}\,(u)\subset\overline{\mbox{\rm Sing}\,(u)}\cap(\mathbb{R}^{n}\setminus\mathcal{I}(u))\subset\mathbb{R}^{n}\setminus\mathcal{I}(u).

3.3. genuine propagation of singularities

To study genuine propagation of singularities, we have to check that the singular arc 𝐱\mathbf{x} in Theorem 3.3 is not a fixed point. As we show below, the following condition about strong critical point (see, for instance [14],[4]) can be useful for this purpose.

Definition 3.11.

We say that xnx\in\mathbb{R}^{n} is a strong critical point of a viscosity solution vv of (HJλ) if

0λv(x)+Hp(x,D+v(x)).0\in\lambda v(x)+H_{p}(x,D^{+}v(x)).
Corollary 3.12.

Let 𝐱:[0,+)n\mathbf{x}:[0,+\infty)\to\mathbb{R}^{n} be the singular curve in Theorem 3.3. If xx is not a strong critical point of vv, then there exists t>0t>0 such that 𝐱(s)x0\mathbf{x}(s)\neq x_{0} for all s(0,t]s\in(0,t].

3.4. Existence of global Lipschitz viscosity solution of (HJλ)

We assume L=L(x,v):n×nL=L(x,v):\mathbb{R}^{n}\times\mathbb{R}^{n}\to\mathbb{R} is a function of class C2C^{2} satisfying:

  1. (L1’)

    L(x,)L(x,\cdot) is strictly convex for all xnx\in\mathbb{R}^{n}.

  2. (L2’)

    There exist c1,c20c_{1},c_{2}\geqslant 0 and two superlinear functions θ1,θ2:[0,+)[0,+)\theta_{1},\theta_{2}:[0,+\infty)\to[0,+\infty) such that

    θ2(|v|)+c2L(x,v)θ1(|v|)c1,(x,v)n×n.\displaystyle\theta_{2}(|v|)+c_{2}\geqslant L(x,v)\geqslant\theta_{1}(|v|)-c_{1},\qquad\forall(x,v)\in\mathbb{R}^{n}\times\mathbb{R}^{n}.

The associated Hamiltonian H:n×n×H:\mathbb{R}^{n}\times\mathbb{R}^{n}\times\mathbb{R}\to\mathbb{R} is defined by

H(x,p)=supvn{p,vL(x,v)},(x,p)n×n.\displaystyle H(x,p)=\sup_{v\in\mathbb{R}^{n}}\{\langle p,v\rangle-L(x,v)\},\qquad(x,p)\in\mathbb{R}^{n}\times\mathbb{R}^{n}.
Theorem 3.13.

Suppose L:n×nL:\mathbb{R}^{n}\times\mathbb{R}^{n}\to\mathbb{R} satisfies (L1’)-(L2’) and λ>0\lambda>0. Then there exists u0:nu_{0}:\mathbb{R}^{n}\to\mathbb{R} such that u0u_{0} is the unique bounded and Lipschitz viscosity solution of (HJλ) on n\mathbb{R}^{n}.

Remark 3.14.

In Theorem 3.3, we suppose u:nu:\mathbb{R}^{n}\to\mathbb{R} is a globally Lipschitz continuous viscosity solution of equation (HJλ). Actually, the viscosity solutions of equation (HJλ) are not always globally Lipschitz continuous. There is a counterexample as follows:

u(x)+12|Du(x)|2=0,xn,u(x)+\frac{1}{2}|Du(x)|^{2}=0,\quad x\in\mathbb{R}^{n}, (3.3)

Obviously, u1(x)=12x2u_{1}(x)=-\frac{1}{2}x^{2} and u2(x)0u_{2}(x)\equiv 0 are both viscosity solutions of (3.3). But u1(x)=12x2u_{1}(x)=-\frac{1}{2}x^{2} is not globally Lipschitz on n\mathbb{R}^{n} and u2(x)0u_{2}(x)\equiv 0 is the unique globally Lipschitz viscosity solution of 3.3.

For any function u:nu:\mathbb{R}^{n}\to\mathbb{R} and t>0t>0, we define the Lax-Oleinik operator (See [10])

Ttu(x)=infξΓ,x0,t{eλtu(ξ(0))+0teλ(st)L(ξ,ξ˙)𝑑s},xn.T_{t}^{-}u(x)=\inf_{\xi\in\Gamma_{\cdot,x}^{0,t}}\Big{\{}e^{-\lambda t}u(\xi(0))+\int_{0}^{t}e^{\lambda(s-t)}L(\xi,\dot{\xi})\ ds\Big{\}},\qquad x\in\mathbb{R}^{n}. (3.4)

Recall some properties of TtT_{t}^{-} as follows:

Lemma 3.15.
  1. (1)

    For any u:nu:\mathbb{R}^{n}\to\mathbb{R} and t1,t2>0t_{1},t_{2}>0, we have

    Tt1Tt2u=Tt1+t2u.\displaystyle T_{t_{1}}^{-}T_{t_{2}}^{-}u=T_{t_{1}+t_{2}}^{-}u.
  2. (2)

    Set ui:n,i=1,2u_{i}:\mathbb{R}^{n}\to\mathbb{R},\ i=1,2. If u1u2u_{1}\leqslant u_{2}, then there holds

    Ttu1Ttu2,t>0\displaystyle T_{t}^{-}u_{1}\leqslant T_{t}^{-}u_{2},\qquad\forall t>0
  3. (3)

    Suppose ui:n,i=1,2u_{i}:\mathbb{R}^{n}\to\mathbb{R},\ i=1,2 are bounded on n\mathbb{R}^{n}. Then we have

    Ttu1Ttu2eλtu1u2,t>0.\displaystyle\|T_{t}^{-}u_{1}-T_{t}^{-}u_{2}\|_{\infty}\leqslant e^{-\lambda t}\|u_{1}-u_{2}\|_{\infty},\qquad\forall t>0.
  4. (4)

    For any t>0t>0, uTtuu\leqslant T_{t}^{-}u if and only if for any absolutely continuous curve γ:[a,b]n\gamma:[a,b]\to\mathbb{R}^{n}, there holds

    eλbu(γ(b))eλau(γ(a))+abeλtL(γ(t),γ˙(t))𝑑t.e^{\lambda b}u(\gamma(b))\leqslant e^{\lambda a}u(\gamma(a))+\int_{a}^{b}e^{\lambda t}L(\gamma(t),\dot{\gamma}(t))\ dt. (3.5)
Lemma 3.16.

There exists positive constants K1=c1/λK_{1}=c_{1}/\lambda, K2=(θ2(0)+c2)/λK_{2}=(\theta_{2}(0)+c_{2})/\lambda such that

K1Tt(K1)(x)K2,t>0,xn.-K_{1}\leqslant T_{t}^{-}(-K_{1})(x)\leqslant K_{2},\qquad\forall t>0,x\in\mathbb{R}^{n}.
Proof.

On the one hand, for any t>0t>0, xnx\in\mathbb{R}^{n}, consider the curve ξ(s)x\xi(s)\equiv x in (3.4), then we obtain that

Tt(K1)(x)\displaystyle T_{t}^{-}(-K_{1})(x)\leqslant eλtK1+0teλ(st)L(x,0)𝑑s\displaystyle\,-e^{-\lambda t}K_{1}+\int_{0}^{t}e^{\lambda(s-t)}L(x,0)\ ds
\displaystyle\leqslant 0te(st)λ(θ2(0)+c2)𝑑s\displaystyle\,\int_{0}^{t}e^{(s-t)\lambda}(\theta_{2}(0)+c_{2})\ ds
\displaystyle\leqslant θ2(0)+c2λ=K2.\displaystyle\,\frac{\theta_{2}(0)+c_{2}}{\lambda}=K_{2}.

On the other hand, let ηΓ,x0,t\eta\in\Gamma_{\cdot,x}^{0,t} be a minimizer for (3.4). It follows that

Tt(K1)(x)=\displaystyle T_{t}^{-}(-K_{1})(x)= eλtK1+0teλ(st)L(η,η˙)𝑑s\displaystyle\,-e^{-\lambda t}K_{1}+\int_{0}^{t}e^{\lambda(s-t)}L(\eta,\dot{\eta})\ ds
\displaystyle\geqslant eλtK1+0teλ(st)(θ1(|η˙(s)|)c1)𝑑s\displaystyle\,-e^{-\lambda t}K_{1}+\int_{0}^{t}e^{\lambda(s-t)}\big{(}\theta_{1}(|\dot{\eta}(s)|)-c_{1}\big{)}\ ds
\displaystyle\geqslant eλtK1c10teλ(st)𝑑s\displaystyle\,-e^{-\lambda t}K_{1}-c_{1}\int_{0}^{t}e^{\lambda(s-t)}\ ds
=\displaystyle= eλtK1(1eλt)K1=K1.\displaystyle\,-e^{-\lambda t}K_{1}-(1-e^{-\lambda t})K_{1}=-K_{1}.

This completes our proof. ∎

Similar to [1, lemma 2.2], we show that bounded subsolutions is Lipschitz on n\mathbb{R}^{n} as follows:

Lemma 3.17.

Suppose u:nu:\mathbb{R}^{n}\to\mathbb{R} is bounded on n\mathbb{R}^{n} and uTtuu\leqslant T_{t}^{-}u for any t>0t>0. Then uu is Lipschitz on n\mathbb{R}^{n}.

Proof.

For x,ynx,y\in\mathbb{R}^{n}, xyx\neq y, consider the curve ξ(s)=x+s(yx)|yx|,s[0,|yx|]\xi(s)=x+s\frac{(y-x)}{|y-x|},\ s\in[0,|y-x|]. Then uTtuu\leqslant T_{t}^{-}u implies that

u(y)eλ|yx|u(x)+0|yx|eλ(s|yx|)L(x+s(yx)|yx|,(yx)|yx|)𝑑s,u(y)\leqslant e^{-\lambda|y-x|}u(x)+\int_{0}^{|y-x|}e^{\lambda(s-|y-x|)}L\Big{(}x+s\frac{(y-x)}{|y-x|},\frac{(y-x)}{|y-x|}\Big{)}\ ds, (3.6)

It follows that

u(y)u(x)\displaystyle\,u(y)-u(x)
\displaystyle\leqslant 1eλ|yx|λ(λu(x))+0|yx|eλ(s|yx|)L(x+s(yx)|yx|,(yx)|yx|)𝑑s\displaystyle\,\frac{1-e^{-\lambda|y-x|}}{\lambda}\cdot(-\lambda u(x))+\int_{0}^{|y-x|}e^{\lambda(s-|y-x|)}L\Big{(}x+s\frac{(y-x)}{|y-x|},\frac{(y-x)}{|y-x|}\Big{)}\ ds
=\displaystyle= 0|yx|eλ(s|yx|)[L(x+s(yx)|yx|,(yx)|yx|)λu(x)]𝑑s\displaystyle\,\int_{0}^{|y-x|}e^{\lambda(s-|y-x|)}\Bigg{[}L\Big{(}x+s\frac{(y-x)}{|y-x|},\frac{(y-x)}{|y-x|}\Big{)}-\lambda u(x)\Bigg{]}\ ds
\displaystyle\leqslant 0|yx|eλ(s|yx|)(θ2(1)+c2+λu)𝑑s\displaystyle\,\int_{0}^{|y-x|}e^{\lambda(s-|y-x|)}\Big{(}\theta_{2}(1)+c_{2}+\lambda\|u\|_{\infty}\Big{)}\ ds
\displaystyle\leqslant (θ2(1)+c2+λu)|yx|.\displaystyle\,(\theta_{2}(1)+c_{2}+\lambda\|u\|_{\infty})|y-x|.

Similarly, there holds u(x)u(y)(θ2(1)+c2+λu)|yx|u(x)-u(y)\leqslant(\theta_{2}(1)+c_{2}+\lambda\|u\|_{\infty})|y-x|. Therefore,

|u(y)u(x)|(θ2(1)+c2+λu)|yx|,x,yn.\displaystyle|u(y)-u(x)|\leqslant(\theta_{2}(1)+c_{2}+\lambda\|u\|_{\infty})|y-x|,\qquad\forall x,y\in\mathbb{R}^{n}.

Following Fathi ([20, 19]), uC0(n,)u\in C^{0}(\mathbb{R}^{n},\mathbb{R}) is called a weak-KAM solution of (HJλ) if

Ttu=u,t>0.\displaystyle T_{t}^{-}u=u,\qquad\forall t>0.
Lemma 3.18.

Suppose u:nu:\mathbb{R}^{n}\to\mathbb{R} is bounded on n\mathbb{R}^{n}. Then uu is a weak-KAM solution of (HJλ) if and only if uu is a viscosity solution of (HJλ).

Proof of Theorem 3.13.

Due to Lemma 3.16, we have that

K1Tt(K1)(x)K2,t>0,xn,\displaystyle-K_{1}\leqslant T_{t}^{-}(-K_{1})(x)\leqslant K_{2},\qquad\forall t>0,x\in\mathbb{R}^{n},

and Tt(K1)T_{t}^{-}(-K_{1}) is non-decreasing for t>0t>0 by (1) and (4) of Lemma 3.15. It follows that T1(K1)(K1)K1+K2\|T_{1}^{-}(-K_{1})-(-K_{1})\|_{\infty}\leqslant K_{1}+K_{2}. Using Lemma 3.15 (3), we obtain that for any kk\in\mathbb{N},

Tk+1(K1)Tk(K1)\displaystyle\|T_{k+1}^{-}(-K_{1})-T_{k}^{-}(-K_{1})\|_{\infty}\leqslant ekλT1(K1)(K1)\displaystyle\,e^{-k\lambda}\|T_{1}^{-}(-K_{1})-(-K_{1})\|_{\infty}
\displaystyle\leqslant ekλ(K1+K2),\displaystyle\,e^{-k\lambda}(K_{1}+K_{2}),

which implies

k=1+Tk+1(K1)Tk(K1)k=1+ekλ(K1+K2)=eλ(K1+K2)1eλ<+.\displaystyle\sum_{k=1}^{+\infty}\|T_{k+1}^{-}(-K_{1})-T_{k}^{-}(-K_{1})\|_{\infty}\leqslant\sum_{k=1}^{+\infty}e^{-k\lambda}(K_{1}+K_{2})=\frac{e^{-\lambda}(K_{1}+K_{2})}{1-e^{-\lambda}}<+\infty.

Therefore, there exists a unique u0:nu_{0}:\mathbb{R}^{n}\to\mathbb{R} such that

limt+Tt(K1)u0=0\lim_{t\to+\infty}\|T_{t}^{-}(-K_{1})-u_{0}\|_{\infty}=0 (3.7)

with

K1u0(x)K2,xn.-K_{1}\leqslant u_{0}(x)\leqslant K_{2},\qquad\forall x\in\mathbb{R}^{n}. (3.8)

For any t>0t^{\prime}>0, by (3.7) and Lemma 3.15 (3) we have

Ttu0=Ttlimt+Tt(K1)=limt+TtTt(K1)=u0.T_{t^{\prime}}^{-}u_{0}=T_{t^{\prime}}^{-}\lim_{t\to+\infty}T_{t}^{-}(-K_{1})=\lim_{t\to+\infty}T_{t^{\prime}}^{-}T_{t}^{-}(-K_{1})=u_{0}. (3.9)

(3.8), (3.9) and Lemma 3.17 implies u0u_{0} is Lipschitz on n\mathbb{R}^{n}. Now we know that u0u_{0} is a weak-KAM solution of (HJλ) which is bounded and Lipschitz on n\mathbb{R}^{n}. By Lemma 3.18, u0u_{0} is also a viscosity solution of (HJλ). The uniqueness of u0u_{0} is a direct consequence of Lemma 3.15 (3). This completes the proof of Theorem 3.13. ∎

Appendix A Regularity properties of fundamental solutions

Here we collect some relevant regularity results with respect to the fundamental solution of (HJe). The proofs of these regularity results are similar to those in [4] in autonomous case.

Proposition A.1.

Suppose L satisfies condition (L1)-(L3). Then for any T>0T>0, 0s<tT0\leqslant s<t\leqslant T, x,ynx,y\in\mathbb{R}^{n}, and any minimizer ξΓx,ys,t\xi\in\Gamma^{s,t}_{x,y} for As,t(x,y)A_{s,t}(x,y), we have

supτ[s,t]|ξ˙(τ)|κ(T,|xy|ts).\displaystyle\sup_{\tau\in[s,t]}|\dot{\xi}(\tau)|\leqslant\kappa(T,\frac{|x-y|}{t-s}).

where κ:(0,+)×(0,+)(0,+)\kappa:(0,+\infty)\times(0,+\infty)\rightarrow(0,+\infty) is nondecreasing.

Now, for (s,x)[0,+)×n(s,x)\in[0,+\infty)\times\mathbb{R}^{n}, λ>0\lambda>0 and τ>0\tau>0, let

Sλ(s,x,τ)={(t,y)×n:s<t<s+τ,|yx|<λ(ts)}.\displaystyle S_{\lambda}(s,x,\tau)=\{(t,y)\in\mathbb{R}\times\mathbb{R}^{n}:s<t<s+\tau,|y-x|<\lambda(t-s)\}.
Proposition A.2.

Suppose L satisfies condition (L1)-(L3).Then for any fixed T>0T>0, R>0R>0 and λ>0\lambda>0, there exists t0(s,x,T,R,λ)>0t_{0}(s,x,T,R,\lambda)>0 such that for any (s,x)[0,T]×B¯(0,R)(s,x)\in[0,T]\times\bar{B}(0,R)

  1. (1)

    The function (t,y)As,t(x,y)(t,y)\mapsto A_{s,t}(x,y) is semiconcave on the cone Sλ(s,x,t0(T,R,λ))S_{\lambda}(s,x,t_{0}(T,R,\lambda)) and there exists C0(T,R,λ)>0C_{0}(T,R,\lambda)>0 such that for all (t,y)Sλ(s,x,t0(T,R,λ))(t,y)\in S_{\lambda}(s,x,t_{0}(T,R,\lambda)), h[0,12(ts))h\in[0,\frac{1}{2}(t-s)) and zB(0,λ(ts))z\in B(0,\lambda(t-s)) we have that

    As,t+h(x,y+z)+As,th(x,yz)2As,t(x,y)C0(T,R,λ)ts(h2+|z|2).\displaystyle A_{s,t+h}(x,y+z)+A_{s,t-h}(x,y-z)-2A_{s,t}(x,y)\leqslant\frac{C_{0}(T,R,\lambda)}{t-s}(h^{2}+|z|^{2}).
  2. (2)

    The function (t,y)As,t(x,y)(t,y)\mapsto A_{s,t}(x,y) is semiconvex on the cone Sλ(s,x,t0(T,R,λ))S_{\lambda}(s,x,t_{0}(T,R,\lambda)) and there exists C1(T,R,λ)>0C_{1}(T,R,\lambda)>0 such that for all (t,y)Sλ(s,x,t0(T,R,λ))(t,y)\in S_{\lambda}(s,x,t_{0}(T,R,\lambda)), h[0,12(ts))h\in[0,\frac{1}{2}(t-s)) and zB(0,λ(ts))z\in B(0,\lambda(t-s)) we have that

    As,t+h(x,y+z)+As,th(x,yz)2As,t(x,y)C1(T,R,λ)ts(h2+|z|2).\displaystyle A_{s,t+h}(x,y+z)+A_{s,t-h}(x,y-z)-2A_{s,t}(x,y)\geqslant-\frac{C_{1}(T,R,\lambda)}{t-s}(h^{2}+|z|^{2}).
  3. (3)

    For all t(s,s+τ]t\in(s,s+\tau], the function As,t(x,)A_{s,t}(x,\cdot) is uniformly convex on B(x,λ(ts))B(x,\lambda(t-s)), and there exists C2(T,R,λ)>0C_{2}(T,R,\lambda)>0 such that for all yB(x,λ(ts))y\in B(x,\lambda(t-s)) and zB(0,λ(ts))z\in B(0,\lambda(t-s)) we have that

    As,t(x,y+z)+As,t(x,yz)2As,t(x,y)C2(T,R,λ)ts|z|2.\displaystyle A_{s,t}(x,y+z)+A_{s,t}(x,y-z)-2A_{s,t}(x,y)\geqslant\frac{C_{2}(T,R,\lambda)}{t-s}|z|^{2}.

    Moreover, C(T,R,λ)C(T,R,\lambda) is continuous with respect to RR.

Remark A.3.

In this paper, for any fixed T>0T>0, we choose λ:=λ2(T)\lambda:=\lambda_{2}(T) and R:=λ2(T)TR:=\lambda_{2}(T)T, where λ2(T)\lambda_{2}(T) is defined in Lemma 2.6. Assume that xSλ(0,0,s)x\in S_{\lambda}(0,0,s), then by Lemma 2.6 and the definition of Sλ(s,x,τ)S_{\lambda}(s,x,\tau), we can let t0(s,x,T,R)=Tst_{0}(s,x,T,R)=T-s and Ci(T,R,λ)C_{i}(T,R,\lambda) only depends on TT for i=1,2,3i=1,2,3.

Moreover, if we consider the domain [0,T]×B¯(x0,R)[0,T]\times\bar{B}(x_{0},R) for any x0nx_{0}\in\mathbb{R}^{n}, then Ci(T,R,λ)C_{i}(T,R,\lambda) only depends on initial point x0x_{0} and TT for i=1,2,3i=1,2,3.

Proposition A.4.

Suppose L satisfies condition (L1)-(L3). Then for any fixed T>0T>0, R>0R>0, λ>0\lambda>0 and (s,x)[0,T]×B¯(0,R)(s,x)\in[0,T]\times\bar{B}(0,R), the function (t,y)As,t(x,y)(t,y)\mapsto A_{s,t}(x,y) is of class Cloc1,1C^{1,1}_{loc} on the cone Sλ(s,x,t0(T,R,λ))S_{\lambda}(s,x,t_{0}(T,R,\lambda)), where t0(T,R,λ)t_{0}(T,R,\lambda) is that in Proposition A.2. Moreover, we have

DyAs,t(x,y)=Lv(t,ξ(t),ξ˙(t)),\displaystyle D_{y}A_{s,t}(x,y)=L_{v}(t,\xi(t),\dot{\xi}(t)),
DxAs,t(x,y)=Lv(s,ξ(s),ξ˙(s)),\displaystyle D_{x}A_{s,t}(x,y)=-L_{v}(s,\xi(s),\dot{\xi}(s)),
DtAs,t(x,y)=Es,t,x,y,\displaystyle D_{t}A_{s,t}(x,y)=-E_{s,t,x,y},

where ξΓx,ys,t\xi\in\Gamma_{x,y}^{s,t} is the unique minimizer for As,t(x,y)A_{s,t}(x,y) and

Es,t,x,y:=H(t,ξ(t),p(t))\displaystyle E_{s,t,x,y}:=H(t,\xi(t),p(t))

is the energy of the Hamilton trajectory (ξ,p)(\xi,p) with

p(τ):=Lv(τ,ξ(τ),ξ˙(τ)),τ[s,t].p(\tau):=L_{v}(\tau,\xi(\tau),\dot{\xi}(\tau)),\quad\tau\in[s,t].

Appendix B Proof of Lemma 2.2 and Lemma 2.6

The convex conjugate of a superlinear function θT\theta_{T} is defined as

θT(s)=supr0{rsθT(r)},s0.\theta_{T}^{*}(s)=\sup_{r\geqslant 0}\{rs-\theta_{T}(r)\},\quad s\geqslant 0.

In view of the superlinear growth of θT\theta_{T}, it is clear that θT\theta_{T}^{*} is well defined and satisfies

θT(r)+θT(s)rs,r,s0,\theta_{T}(r)+\theta_{T}^{*}(s)\geqslant rs,\quad r,s\geqslant 0,

which in turn can be used to show that θT(s)/s+\theta_{T}^{*}(s)/s\rightarrow+\infty as s+s\rightarrow+\infty.

Proof of Lemma 2.2.

For item (1), let k=Lip[f]+1k=\mbox{\rm{Lip}}[f]+1. Then for any xnx\in\mathbb{R}^{n}, 0t1<t2T0\leqslant t_{1}<t_{2}\leqslant T and znz\in\mathbb{R}^{n}, we have

At1,t2(z,x)=\displaystyle A_{t_{1},t_{2}}(z,x)= infξΓz,xt1,t2t1t2L(s,ξ,ξ˙)𝑑s\displaystyle\,\inf_{\xi\in\Gamma_{z,x}^{t_{1},t_{2}}}\int_{t_{1}}^{t_{2}}L(s,\xi,\dot{\xi})\ ds
\displaystyle\geqslant infξΓz,xt1,t2t1t2θT(|ξ˙|)𝑑scT(t2t1)\displaystyle\,\inf_{\xi\in\Gamma_{z,x}^{t_{1},t_{2}}}\int_{t_{1}}^{t_{2}}\theta_{T}(|\dot{\xi}|)\ ds-c_{T}(t_{2}-t_{1})
\displaystyle\geqslant infξΓz,xt1,t2kt1t2|ξ˙|𝑑s(θT(k)+cT)(t2t1)\displaystyle\,\inf_{\xi\in\Gamma_{z,x}^{t_{1},t_{2}}}k\int_{t_{1}}^{t_{2}}|\dot{\xi}|\ ds-(\theta_{T}^{*}(k)+c_{T})(t_{2}-t_{1})
\displaystyle\geqslant k|zx|(θT(k)+cT)(t2t1).\displaystyle\,k|z-x|-(\theta_{T}^{*}(k)+c_{T})(t_{2}-t_{1}).

Therefore,

f(x)+At1,t2(x,x)f(z)At1,t2(z,x)\displaystyle f(x)+A_{t_{1},t_{2}}(x,x)-f(z)-A_{t_{1},t_{2}}(z,x)
\displaystyle\leqslant Lip[f]|zx|k|zx|+(θT(k)+cT)(t2t1)+t1t2L(s,x,0)𝑑s\displaystyle\,\mbox{\rm Lip}[f]\cdot|z-x|-k|z-x|+(\theta_{T}^{*}(k)+c_{T})\ (t_{2}-t_{1})+\int_{t_{1}}^{t_{2}}L(s,x,0)\ ds
\displaystyle\leqslant |zx|+[θT(k)+cT+θ¯T(0)](t2t1).\displaystyle\,-|z-x|+[\theta_{T}^{*}(k)+c_{T}+\overline{\theta}_{T}(0)]\ (t_{2}-t_{1})\ .

Now, taking λ1=θT(k)+cT+θ¯T(0)\lambda_{1}=\theta_{T}^{*}(k)+c_{T}+\overline{\theta}_{T}(0), it follows that

Λt1,t2x:={z:f(z)+At1,t2(z,x)f(x)+At1,t2(x,x)}B¯(x,λ1(t2t1)).\Lambda_{t_{1},t_{2}}^{x}:=\{z:f(z)+A_{t_{1},t_{2}}(z,x)\leqslant f(x)+A_{t_{1},t_{2}}(x,x)\}\subset\overline{B}\big{(}x,\lambda_{1}(t_{2}-t_{1})\big{)}. (B.1)

Therefore Λt1,t2x\Lambda_{t_{1},t_{2}}^{x} is compact and the infimum in (2.3) is attained, i.e., Z(f,t1,t2,x)Z(f,t_{1},t_{2},x)\neq\emptyset. Moreover, due to (B.1), for any zt1,t2,xZ(f,t1,t2,x)z_{t_{1},t_{2},x}\in Z(f,t_{1},t_{2},x), we have

|zt1,t2,xx|λ1(t2t1).|z_{t_{1},t_{2},x}-x|\leqslant\lambda_{1}(t_{2}-t_{1}).

For item (2), A similar result holds for the sup-convolution defined in (2.4). ∎

Proof of Lemma 2.6.

Set (t,x)(0,T]×n(t,x)\in(0,T]\times\mathbb{R}^{n} is a differentiable point of uu. Due to Proposition 2.1 and Proposition 2.5, the solution of (2.2) with terminal condition

{ξ(t)=xp(t)=u(t,x)\begin{cases}\xi(t)=x\\ p(t)=\nabla u(t,x)\end{cases}

is the unique minimizer for u(t,x)u(t,x). Lemma 2.2 (1) implies |ξ(0)x|λ1(T,Lip[u0])t|\xi(0)-x|\leqslant\lambda_{1}(T,\mbox{\rm Lip}[u_{0}])t. Now, denote that

E(s):=H(s,ξ(s),p(s)),s[0,t].\displaystyle E(s):=H(s,\xi(s),p(s)),\qquad s\in[0,t].

By Lemma 2.3, we know that p(0)Du0(ξ(0))p(0)\in D^{-}u_{0}(\xi(0)). This implies |p(0)|Lip[u0]|p(0)|\leqslant\mbox{\rm Lip}\,[u_{0}] and

E(0)=H(0,ξ(0),p(0))θT(|p(0)|)+cTθT(Lip(u0))+cT.\displaystyle E(0)=H(0,\xi(0),p(0))\leqslant\theta^{*}_{T}(|p(0)|)+c_{T}\leqslant\theta^{*}_{T}(\mbox{\rm Lip}\,(u_{0}))+c_{T}.

Notice that

ddsE(s)=ddsH(s,ξ(s),p(s))=Ht+Hxξ˙(s)+Hpp˙(s)\displaystyle\frac{d}{ds}E(s)=\frac{d}{ds}H(s,\xi(s),p(s))=H_{t}+H_{x}\cdot\dot{\xi}(s)+H_{p}\cdot\dot{p}(s)
=\displaystyle= Ht+HxHp+Hp(Hx)=Ht(s,ξ(s),p(s))=Lt(s,ξ(s),ξ˙(s)).\displaystyle H_{t}+H_{x}\cdot H_{p}+H_{p}\cdot(-H_{x})=H_{t}(s,\xi(s),p(s))=-L_{t}(s,\xi(s),\dot{\xi}(s)).

Thus, we have

E(t)\displaystyle E(t) =E(0)+0tddsE(s)𝑑s\displaystyle=E(0)+\int_{0}^{t}\frac{d}{ds}E(s)\ ds
=E(0)+0tLt(s,ξ(s),ξ˙(s))𝑑s\displaystyle=E(0)+\int_{0}^{t}L_{t}(s,\xi(s),\dot{\xi}(s))\ ds
E(0)+0t(C~1(T)+C~2(T)L(s,ξ(s),ξ˙(s)))𝑑s\displaystyle\leqslant E(0)+\int_{0}^{t}\Big{(}\widetilde{C}_{1}(T)+\widetilde{C}_{2}(T)L(s,\xi(s),\dot{\xi}(s))\Big{)}\ ds
E(0)+tC~1(T)+C~2(T)0tL(s,ξ(0)+s(xξ(0)),xξ(0)t)𝑑s\displaystyle\leqslant E(0)+t\ \widetilde{C}_{1}(T)+\widetilde{C}_{2}(T)\int_{0}^{t}L(s,\xi(0)+s(x-\xi(0)),\frac{x-\xi(0)}{t})\ ds
E(0)+C~1(T)T+C~2(T)0tθ¯T(|xξ(0)t|)𝑑s\displaystyle\leqslant E(0)+\widetilde{C}_{1}(T)T+\widetilde{C}_{2}(T)\int_{0}^{t}\overline{\theta}_{T}(|\frac{x-\xi(0)}{t}|)\ ds
θT(Lip(u0))+cT+C~1(T)T+C~2(T)Tθ¯T(λ1(T,Lip[u0])t).\displaystyle\leqslant\theta^{*}_{T}(\mbox{\rm Lip}\,(u_{0}))+c_{T}+\widetilde{C}_{1}(T)T+\widetilde{C}_{2}(T)T\cdot\overline{\theta}_{T}(\lambda_{1}(T,\mbox{\rm Lip}[u_{0}])t).

Since

|u(t,x)|θ¯T(|u(t,x)|)+θ¯T(1)H(t,x,u(t,x))+θ¯T(1)=E(t)+θ¯T(1),\displaystyle|\nabla u(t,x)|\leqslant\overline{\theta}_{T}^{*}(|\nabla u(t,x)|)+\overline{\theta}_{T}(1)\leqslant H(t,x,\nabla u(t,x))+\overline{\theta}_{T}(1)=E(t)+\overline{\theta}_{T}(1),

it follows that

|u(t,x)|\displaystyle|\nabla u(t,x)|\leqslant θT(Lip(u0))+cT+C~1(T)T+C~2(T)Tθ¯T(λ1(T,Lip[u0])t)+θ¯T(1)\displaystyle\theta^{*}_{T}(\mbox{\rm Lip}\,(u_{0}))+c_{T}+\widetilde{C}_{1}(T)T+\widetilde{C}_{2}(T)T\cdot\overline{\theta}_{T}(\lambda_{1}(T,\mbox{\rm Lip}[u_{0}])t)+\overline{\theta}_{T}(1)
:=\displaystyle:= F1(T).\displaystyle F_{1}(T).

By proposition 2.4 (1), we obtain

|ut(t,x)|\displaystyle|u_{t}(t,x)| =|H(t,x,u(t,x))|θT(|u(t,x)|)+c0+|θ¯T(|u(t,x)|)|\displaystyle=|-H(t,x,\nabla u(t,x))|\leqslant\theta_{T}^{*}(|\nabla u(t,x)|)+c_{0}+|\bar{\theta}_{T}^{*}(|\nabla u(t,x)|)|
θT(F1(T))+c0+|θ¯T(F1(T))|:=F2(T).\displaystyle\leqslant\theta_{T}^{*}(F_{1}(T))+c_{0}+|\bar{\theta}_{T}^{*}(F_{1}(T))|:=F_{2}(T).

Therefore,

|Du(t,x)|=(|u(t,x)|2+|ut(t,x)|2)12(F12(T)+F22(T))12:=F0(T).\displaystyle|Du(t,x)|=(|\nabla u(t,x)|^{2}+|u_{t}(t,x)|^{2})^{\frac{1}{2}}\leqslant\big{(}F_{1}^{2}(T)+F_{2}^{2}(T)\big{)}^{\frac{1}{2}}:=F_{0}(T).

Combing this with Proposition 2.4 (2), we conclude that uu is a Lipschitz function on (0,T]×n(0,T]\times\mathbb{R}^{n} and Lip[u]F0(T)\mbox{\rm Lip}\,[u]\leqslant F_{0}(T). ∎

Appendix C Proof of Lemma 3.8

We define F:n×[0,t]nF:\mathbb{R}^{n}\times[0,t]\to\mathbb{R}^{n} as:

F(x,s)=𝐱x(s),s[0,+),F(x,s)=\mathbf{x}_{x}(s),\quad s\in[0,+\infty),

where 𝐱x:[0,+)n\mathbf{x}_{x}:[0,+\infty)\to\mathbb{R}^{n} is that in Theorem 2.9.

Similar to [8, Lemma 2.1], F(x,s)F(x,s) has the properties (a), (b) and (c) stated in Lemma 3.8. Due to Theorem 3.3, 𝐱x(s)\mathbf{x}_{x}(s) is a locally Lipschitz curve. Thus, to prove that F(x,s)=𝐱x(s)F(x,s)=\mathbf{x}_{x}(s) is continuous, it remains to show 𝐱x(s)\mathbf{x}_{x}(s) is continuous with respect to xx. We prove it in Lemma C.2.

Lemma C.1.

For any fixed xnx\in\mathbb{R}^{n} and t>0t>0, let tx,T>0t_{x,T}>0 be defined in Lemma 2.8. Then for any 0<s<t<T0<s<t<T with tstx,Tt-s\leqslant t_{x,T}, the map zys,t,zz\mapsto y_{s,t,z} is Lipschitz on B(x,λ2(T)tx,T)B(x,\lambda_{2}(T)t_{x,T}) with Lipschitz constant K1(x,T)K_{1}(x,T) which only depends on x,Tx,T.

Proof.

For any x1,x2x_{1},x_{2} with |x1x2|<λ2(T)tx,T|x_{1}-x_{2}|<\lambda_{2}(T)t_{x,T} and |x1x2|<r|x_{1}-x_{2}|<r, we denote by ys,t,x1y_{s,t,x_{1}} and ys,t,x2y_{s,t,x_{2}} the unique maximizers of Ts,t+u(t,x1)T^{+}_{s,t}u(t,x_{1}) and Ts,t+u(t,x2)T^{+}_{s,t}u(t,x_{2}) respectively. Notice that As,t(x1,)A_{s,t}(x_{1},\cdot) is uniformly convex in the ball B(x1,2λ2(T)tx,T)B(x_{1},2\lambda_{2}(T)t_{x,T}) with convexity constant C2(x,T)/tC_{2}(x,T)/t for t(0,tT)t\in(0,t_{T}).Then we have

C2(x,T)t|ys,t,x1ys,t,x2|2As,t(x1,ys,t,x1)As,t(x1,ys,t,x2)DyAs,t(x1,ys,t,x2)(ys,t,x1ys,t,x2).\frac{C_{2}(x,T)}{t}|y_{s,t,x_{1}}-y_{s,t,x_{2}}|^{2}\leqslant A_{s,t}(x_{1},y_{s,t,x_{1}})-A_{s,t}(x_{1},y_{s,t,x_{2}})-D_{y}A_{s,t}(x_{1},y_{s,t,x_{2}})(y_{s,t,x_{1}}-y_{s,t,x_{2}}).

On the other hand, since C(T)C(T) is the semiconcave constant of u(t,)u(t,\cdot) for t[0,T]t\in[0,T], we have

As,t(x1,ys,t,x1)As,t(x1,ys,t,x2)u(t,ys,t,x1)u(t,ys,t,x2)\displaystyle\,A_{s,t}(x_{1},y_{s,t,x_{1}})-A_{s,t}(x_{1},y_{s,t,x_{2}})\leqslant u(t,y_{s,t,x_{1}})-u(t,y_{s,t,x_{2}})
\displaystyle\leqslant DyAs,t(x2,ys,t,x2)(ys,t,x1ys,t,x2)+C(x,T)|ys,t,x1ys,t,x2|2.\displaystyle\,D_{y}A_{s,t}(x_{2},y_{s,t,x_{2}})(y_{s,t,x_{1}}-y_{s,t,x_{2}})+C(x,T)|y_{s,t,x_{1}}-y_{s,t,x_{2}}|^{2}.

Notice that DyAt(x2,ys,t,x2)+u(t,ys,t,x2)D_{y}A_{t}(x_{2},y_{s,t,x_{2}})\in\nabla^{+}u(t,y_{s,t,x_{2}}). Thus,

C2(x,T)ts|ys,t,x1ys,t,x2|2\displaystyle\,\frac{C_{2}(x,T)}{t-s}|y_{s,t,x_{1}}-y_{s,t,x_{2}}|^{2}
\displaystyle\leqslant (DyAs,t(x2,ys,t,x2)DyAs,t(x1,ys,t,x2))(ys,t,x1ys,t,x2)+C(x,T)|ys,t,x1ys,t,x2|2\displaystyle\,\Big{(}D_{y}A_{s,t}(x_{2},y_{s,t,x_{2}})-D_{y}A_{s,t}(x_{1},y_{s,t,x_{2}})\Big{)}(y_{s,t,x_{1}}-y_{s,t,x_{2}})+C(x,T)|y_{s,t,x_{1}}-y_{s,t,x_{2}}|^{2}
\displaystyle\leqslant C0(x,T)ts|x1x2||ys,t,x1ys,t,x2|+C(x,T)|ys,t,x1ys,t,x2|2.\displaystyle\,\frac{C_{0}(x,T)}{t-s}|x_{1}-x_{2}|\cdot|y_{s,t,x_{1}}-y_{s,t,x_{2}}|+C(x,T)|y_{s,t,x_{1}}-y_{s,t,x_{2}}|^{2}.

Therefore, due to tx,T:=C2(x,T)2C(x,T)t_{x,T}:=\frac{C_{2}(x,T)}{2C(x,T)} , for any 0<tstx,T0<t-s\leqslant t_{x,T}, we have

|ys,t,x1ys,t,x2|C0(x,T)C2(x,T)C(x,T)(ts)|x1x2|2C0(x,T)C2(x,T)|x1x2|:=K1(x,T)|x1x2|.|y_{s,t,x_{1}}-y_{s,t,x_{2}}|\leqslant\frac{C_{0}(x,T)}{C_{2}(x,T)-C(x,T)\cdot(t-s)}|x_{1}-x_{2}|\leqslant\frac{2C_{0}(x,T)}{C_{2}(x,T)}|x_{1}-x_{2}|:=K_{1}(x,T)|x_{1}-x_{2}|.

Lemma C.2.

For any t>0t>0, 𝐱x(t)\mathbf{x}_{x}(t) is continuous with respect to xx.

Proof.

Let Ωn(x):={B(x,λ2(nT)nT)}\Omega_{n}(x):=\{B(x,\lambda_{2}(nT)nT)\} and assume tx1,Ttx2,Tt_{x_{1},T}\geqslant t_{x_{2},T}. By Theorem 2.9, we construct an new curve 𝐱~x1(t)\widetilde{\mathbf{x}}_{x_{1}}(t) defined by Ωn(x1)\Omega_{n}(x_{1}) and t~x1,T:=tx2,T\widetilde{t}_{x_{1},T}:=t_{x_{2},T}. Then, for (n1)Tt<nT(n-1)T\leqslant t<nT with nn\in\mathbb{N}, by Lemma C.1, we have

|𝐱~x1(t)𝐱x2(t)|K0(nT)|x1x2|,|x1x2|<K1(T)tx2,T.|\widetilde{\mathbf{x}}_{x_{1}}(t)-\mathbf{x}_{x_{2}}(t)|\leqslant K_{0}(nT)\cdot|x_{1}-x_{2}|,\quad\forall|x_{1}-x_{2}|<K_{1}(T)t_{x_{2},T}. (C.1)

where K0(T)=(K1(T))k1K_{0}(T)=(K_{1}(T))^{k_{1}} and k1k_{1} depends on t~x1,T=tx2,T\widetilde{t}_{x_{1},T}=t_{x_{2},T}.

On the other hand, due to Theorem 2.8, Proposition A.2 and Proposition 2.4(2),

tx,T=C2(x,T)2C(x,T)t_{x,T}=\frac{C_{2}(x,T)}{2C(x,T)}

is continuous with respect to xx. Therefore, for 0t<T0\leqslant t<T, we have

|𝐱~x1(t)𝐱x1(t)|K~0(T)|tx1,Ttx2,T|,|\widetilde{\mathbf{x}}_{x_{1}}(t)-\mathbf{x}_{x_{1}}(t)|\leqslant\widetilde{K}_{0}(T)\cdot|t_{x_{1},T}-t_{x_{2},T}|,

where K~0(T):=(K1(T))k1\widetilde{K}_{0}(T):=(K_{1}(T))^{k_{1}} and k1k_{1} depends on tx1,Tt_{x_{1},T}. Combining this with (C.1), one obtain that for any t[0,T)t\in[0,T) ,

|𝐱x2(t)𝐱x1(t)|\displaystyle|\mathbf{x}_{x_{2}}(t)-\mathbf{x}_{x_{1}}(t)|\leqslant |𝐱~x1(t)𝐱x2(t)|+|𝐱~x1(t)𝐱x1(t)|\displaystyle\,|\widetilde{\mathbf{x}}_{x_{1}}(t)-\mathbf{x}_{x_{2}}(t)|+|\widetilde{\mathbf{x}}_{x_{1}}(t)-\mathbf{x}_{x_{1}}(t)|
\displaystyle\leqslant K0(T)|x1x2|+K~0(T)|tx1,Ttx2,T|.\displaystyle\,K_{0}(T)\cdot|x_{1}-x_{2}|+\widetilde{K}_{0}(T)\cdot|t_{x_{1},T}-t_{x_{2},T}|.

Similarly, for t[(n1)T,nT)t\in[(n-1)T,nT), there exists constant K0(n,T)K_{0}(n,T) and K~0(n,T)\widetilde{K}_{0}(n,T) such that

|𝐱x2(t)𝐱x1(t)|K0(n,T)|x1x2|+K~0(n,T)|tx1,nTtx2,nT|.\displaystyle|\mathbf{x}_{x_{2}}(t)-\mathbf{x}_{x_{1}}(t)|\leqslant K_{0}(n,T)\cdot|x_{1}-x_{2}|+\widetilde{K}_{0}(n,T)\cdot|t_{x_{1},nT}-t_{x_{2},nT}|.

Since tx,nTt_{x,nT} is continuous with respect to xx for each nn\in\mathbb{N}, we conclude that for any t>0t>0, 𝐱x(t)\mathbf{x}_{x}(t) is continuous with respect to xx. ∎

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] P. Albano and P. Cannarsa. Propagation of singularities for solutions of nonlinear first order partial differential equations. Arch. Ration. Mech. Anal., 162(1):1–23, 2002.
  • [3] A. Brown and C. Pearcy. An Introduction to Analysis, volume 154. Graduate Texts in Mathematics, Heidelberg, 1995.
  • [4] P. Cannarsa and W. Cheng. Generalized characteristics and Lax-Oleinik operators: global theory. Calc. Var. Partial Differential Equations, 56(5):Art. 125, 31, 2017.
  • [5] P. Cannarsa and W. Cheng. On and beyond propagation of singularities of viscosity solutions. preprint, arXiv:1805.11583, 2018.
  • [6] P. Cannarsa and W. Cheng. Local singular characteristics on 2\mathbb{R}^{2}. Bollettino della Unione matematica italiana, pages 1–22, 2021.
  • [7] P. Cannarsa and W. Cheng. Singularities of solutions of Hamilton-Jacobi equations. Milan Journal of Mathematics, 2021.
  • [8] P. Cannarsa, W. Cheng, and A. 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.
  • [9] P. Cannarsa, W. Cheng, and A. Fathi. Singularities of solutions of time dependent Hamilton-Jacobi equations. Applications to Riemannian geometry. preprint, arXiv:1912.04863, 2019.
  • [10] 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.
  • [11] P. Cannarsa, W. Cheng, M. Mazzola, and K. Wang. Global generalized characteristics for the Dirichlet problem for Hamilton-Jacobi equations at a supercritical energy level. SIAM J. Math. Anal., 51(5):4213–4244, 2019.
  • [12] P. Cannarsa, M. Mazzola, and C. Sinestrari. Global propagation of singularities for time dependent Hamilton-Jacobi equations. Discrete Contin. Dyn. Syst., 35(9):4225–4239, 2015.
  • [13] 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.
  • [14] P. Cannarsa and Y. Yu. Singular dynamics for semiconcave functions. J. Eur. Math. Soc. (JEMS), 11(5):999–1024, 2009.
  • [15] C. Chen, W. Cheng, and Q. Zhang. Lasry-Lions approximations for discounted Hamilton-Jacobi equations. J. Differential Equations, 265(2):719–732, 2018.
  • [16] W. Cheng and J. Hong. Local strict singular characteristics: Cauchy problem with smooth initial data,. preprint, arXiv:2103.06217, 2021.
  • [17] F. H. Clarke and R. B. Vinter. Regularity properties of solutions to the basic problem in the calculus of variations. Trans. Amer. Math. Soc., 289(1):73–98, 1985.
  • [18] J. Dugundji. Topology. Allyn and Bacon, Inc., Boston, Mass., 1966.
  • [19] A. Fathi. Solutions KAM faibles conjuguées et barrières de Peierls. C. R. Acad. Sci. Paris Sér. I Math., 325(6):649–652, 1997.
  • [20] A. Fathi. Théorème KAM faible et théorie de Mather sur les systèmes lagrangiens. C. R. Acad. Sci. Paris Sér. I Math., 324(9):1043–1046, 1997.
  • [21] A. Fathi and E. Maderna. Weak KAM theorem on non compact manifolds. NoDEA Nonlinear Differential Equations Appl., 14(1-2):1–27, 2007.
  • [22] K. Khanin and A. Sobolevski. On dynamics of Lagrangian trajectories for Hamilton-Jacobi equations. Arch. Ration. Mech. Anal., 219(2):861–885, 2016.
  • [23] S. Marò and A. Sorrentino. Aubry-Mather theory for conformally symplectic systems. Comm. Math. Phys., 354(2):775–808, 2017.
  • [24] T. Strömberg. Propagation of singularities along broken characteristics. Nonlinear Anal., 85:93–109, 2013.
  • [25] K. Wang, L. Wang, and J. Yan. Aubry–Mather Theory for Contact Hamiltonian Systems. Comm. Math. Phys., 366(3):981–1023, 2019.
  • [26] Y. Yu. A simple proof of the propagation of singularities for solutions of Hamilton-Jacobi equations. Ann. Sc. Norm. Super. Pisa Cl. Sci. (5), 5(4):439–444, 2006.