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

Variational construction of connecting orbits between Legendrian graphs

Liang Jin   Jun Yan  and  Kai Zhao
Abstract

Motivated by the problem of global stability of thermodynamical equilibria in non-equilibrium thermodynamics formulated in a recent paper [12], we introduce some mechanisms for constructing semi-infinite orbits of contact Hamiltonian systems connecting two Legendrian graphs from the viewpoint of Aubry-Mather theory and weak KAM theory.

Jin Liang: Department of Mathematics, Nanjing University of Science and Technology, Nanjing 210094, China; Fakultät für Mathematik, Ruhr-Universität Bochum, Universitätstraβ\betae 150, D-44801 Bochum, Germany
e-mail: [email protected]; [email protected]
      Jun Yan: School of Mathematical Sciences, Fudan University, Shanghai 200433, China; e-mail: [email protected]
      Kai Zhao: School of Mathematical Sciences, Fudan University, Shanghai 200433, China; e-mail: zhao_\_[email protected]

1 Introduction

Let MM be a closed, connected smooth Riemannian manifold and Σ=J1M=TM×\Sigma=J^{1}M=T^{\ast}M\times\mathbb{R} the manifold of 1-jets of functions on MM. We use either σ\sigma or the coordinates (q,p,u)(q,p,u) to denote points in Σ\Sigma, where (q,p)(q,p) is the usual coordinates on TMT^{\ast}M and uu\in\mathbb{R}. The kernel of the Gibbs 11-form α=dupdq\alpha=du-pdq defines the standard contact structure ξTΣ\xi\subset T\Sigma, which makes (Σ,ξ)(\Sigma,\xi) into a canonical contact manifold. In the following context,

  • πu:ΣTM;σ=(q,p,u)(q,p)\pi^{u}:\Sigma\rightarrow T^{\ast}M;\quad\sigma=(q,p,u)\mapsto(q,p) denotes the projection forgetting uu-component,

  • πq:ΣM;σ=(q,p,u)q\pi_{q}:\Sigma\rightarrow M;\quad\sigma=(q,p,u)\mapsto q denotes the projection onto qq-component.

1.1 Contact structure and classical thermodynamics

According to V.I.Arnold [1], “the first person who understood the significance of contact geometry for physics and thermodynamics” is J.W.Gibbs. As is well known, see [27, Part III, Chapter 1] and the references therein, Gibbs laid the foundation of classical thermodynamics by using only the functions of thermodynamic state, such as entropy and energy, as coordinates to describe the thermodynamic process, and then give the mathematical formulation of two principles of classical thermodynamics with the help of the 1-form α\alpha: any equilibrium process corresponds to an oriented path γ\gamma in Σ\Sigma and takes place in such a way that γ˙ker(α)=ξ\dot{\gamma}\in\ker(\alpha)=\xi. From Gibbs’ formulation, one naturally thinks of the sets of equilibrium states of a thermodynamic system as integral manifolds of the contact structure ξ\xi, especially

Definition 1.1 (Legendrian submanifolds).

An integral submanifold ΛΣ\Lambda\subset\Sigma of ξ\xi whose dimension is maximal, i.e., dim Λ=\Lambda= dim MM, is called Legendrian. Furthermore, Λ\Lambda is called a Legendrian graph if πq|Λ:ΛM\pi_{q}|_{\Lambda}:\Lambda\rightarrow M is a diffeomorphism.

It is necessary to

Remark 1.2.

[2, Lecture 2, Theorem 4] Legendrian graphs coincide with 1-graphs of functions: for every Legendrian graph Λ\Lambda, there is a function uC1(M,)u\in C^{1}(M,\mathbb{R}) such that

Λ=Λu:={(q,dqu(q),u(q)):qM},\Lambda=\Lambda_{u}:=\{\,(q,d_{q}u(q),u(q)):\,q\in M\,\},

and we shall use the notation Λu\Lambda_{u} if we want to emphasis the generating function.

Therefore, (Σ,ξ)(\Sigma,\xi) serves as the phase space in the geometrical description of classical (or equilibrium) thermodynamics. After the fundamental works of Gibbs, the classical (or equilibrium) thermodynamics, which is the theory of properties of matter in a state of thermodynamics equilibrium and mainly deals with the reversible evolution of equilibrium states, became the study of contact geometry of the phase space.

1.2 Contact Hamiltonian systems and non-equilibrium thermodynamics

Unfortunately, systems found in nature are rarely in thermodynamic equilibrium, mainly due to the exchange of matter and energy with the environment and the chemical reactions inside the system. Does this mean that Gibbs’ framework is nonsense to such systems? The discoveries of 20 century’s thermodynamics probably give an answer from negative direction.

For instance, physical experiments show that when a thermodynamic system in an equilibrium state undergoes a perturbation, it probably moves to a non-equilibrium state and then enter an interesting relaxation process that driving the system gradually returns to the original equilibrium. The past 30 years have witnessed a trend to interpret this relaxation processes via contact Hamiltonian flows, an analogy of Hamiltonian flow on contact manifolds [3, 12, 15, 22, 23]. One reason to choose such flows comes from the fact that they are transformations of Σ\Sigma preserving the contact structure ξ\xi and thus sets of equilibrium states (means that the flow transforms Legendrian submanifolds into Legendrian submanifolds) as well. As a consequence, the generating vector field XX satisfies the equation

Xα=fα,\mathcal{L}_{X}\alpha=f\alpha, (1.1)

where X\mathcal{L}_{X} denotes the Lie derivative along the vector field XX and ff a nowhere vanishing function on Σ\Sigma. It turns out that XX is uniquely determined by the function H=iXαH=i_{X}\alpha, called contact Hamiltonian, on Σ\Sigma, where iXi_{X} denotes the inner product of a vector field and a 1-form. Set q˙=iXdq,p˙=iXdp,u˙=iXdu\dot{q}=i_{X}dq,\,\dot{p}=i_{X}dp,\,\dot{u}=i_{X}du, the equation (1.1) reads in the coordinates (q,p,u)(q,p,u) as the contact Hamiltonian system

X:{q˙=Hp(q,p,u),p˙=Hq(q,p,u)Hu(q,p,u)p,u˙=Hp(q,p,u)pH(q,p,u).X:\begin{cases}\dot{q}=\frac{\partial H}{\partial p}(q,p,u),\\ \dot{p}=-\frac{\partial H}{\partial q}(q,p,u)-\frac{\partial H}{\partial u}(q,p,u)p,\\ \dot{u}=\frac{\partial H}{\partial p}(q,p,u)\cdot p-H(q,p,u).\end{cases} (1.2)

In the following context, we shall use the notation XHX_{H} instead of XX to emphasis the role of HH, and the corresponding phase flow is denoted by φHt\varphi^{t}_{H}. In particular, if the Hamiltonian is independent of uu, then (1.2) reduces to the classical Hamiltonian system.

Once and for all, we assume there is a Legendrian submanifold

  • ΛΣ\Lambda_{-}\subset\Sigma coincides with the pre-assigned set of original equilibrium states.

The ingredients of the geometric model of relaxation process are included in the following

Definition 1.3.

For HC(Σ,)H\in C^{\infty}(\Sigma,\mathbb{R}), the pair (H,Λ)(H,\Lambda_{-}) is called a non-equilibrium thermodynamic system (a system for short) if HH generates a complete phase flow φHt\varphi^{t}_{H} and H|Λ0H|_{\Lambda_{-}}\equiv 0. It follows that Λ\Lambda_{-} is an invariant manifold under φHt\varphi^{t}_{H}.

Remark 1.4.

In the language of non-equilibrium thermodynamics, the phase flow φHt\varphi^{t}_{H} represents the thermodynamic process, which can be chosen by determining the contact Hamiltonian HH from different physical considerations. However, the above definition indicates the choice of HH can not be arbitrary and should guarantee that the process preserves the set of original equilibrium states.

1.3 Prigogine’s question on the global stability of thermodynamic equilibrium

In his 1977 Nobel lecture [21, Page 269], Ilya.Prigogine attributed the local stability of thermodynamic equilibrium in the relaxation processes to the fact that uu (called thermodynamic potential) serves as a Lyapunov function near the equilibrium and then raise the global question:

  • Can we extrapolate this stability property further away from equilibrium?

This note is motivated by the latest work [12] of M.Entov and L.Polterovich, in which the authors reformulate Prigogine’s question into a question concerning contact Hamiltonian dynamics:

Question 1.5.

[12, Section 3, Question 3.1] Given a system (H,Λ)(H,\Lambda_{-}) and a subset Σ0Σ\Sigma_{0}\subset\Sigma of the phase space, does there exist an initial condition σ0Σ0\sigma_{0}\in\Sigma_{0} whose trajectory in the thermodynamical process generated by HH asymptotically converges to the equilibrium submanifold Λ\Lambda_{-}?

In [12], the authors also offered their answer to Question 1.5 for the case that Σ0=Λ0\Sigma_{0}=\Lambda_{0} is a Legendrian submanifold under the assumption proposed by physicists working in non-equilibrium thermodynamics:

  • (H1)

    Λ\Lambda_{-} is a local attractor, i.e., there is a neighborhood 𝒪\mathcal{O}_{-} of Λ\Lambda_{-} in Σ\Sigma such that for any σ𝒪\sigma\in\mathcal{O}_{-}, the orbit {φHtσ:t0}𝒪\{\varphi^{t}_{H}\sigma\,:\,t\geqslant 0\}\subset\mathcal{O}_{-} and the set ω(σ)\omega(\sigma) is a non-empty subset of Λ\Lambda_{-}.

In fact, they constructed semi-infinte orbits of φHt\varphi^{t}_{H} starting from Λ0\Lambda_{0} and converges to Λ\Lambda_{-} when tt goes to infinity. The methods used there is based on an existence mechanism, for finite time-length trajectories of (1.2) between Legendrian submanifolds, called interlinking established from Legendrian Contact Homology in hard contact geometry [11, 13].

As is indicated in the last section, Hamiltonian system could be seen as a special case of (1.2). Topics related to understanding the chaotic behavior of orbits in Hamiltonian system lie at the centre in this field. For instance, one of the main goal of the celebrated Aubry-Mather theory [17, 18] is to answer

[18] whether there exists an orbit which in the infinite past tends to one region of phase space and in the infinite future tends to another region of phase space” and “the possibility of finding an orbit which visits a prescribed sequence of regions of phase space in turn”.

Thus the construction of semi-infinite orbits asymptotic to certain invariant sets plays the role of building block for studying Hamiltonian dynamics from the above viewpoint. This fact also suggests the importance of Question 1.5 in the study of contact Hamiltonian dynamics.

1.4 Main results

In recent years, the study of contact Hamiltonian system from the viewpoint of Aubry-Mather theory [17, 18] and weak KAM theory [14] has fruitful consequences, including the proof of vanishing discount limit [9] raised in the homogenization problem of Hamilton-Jacobi equations [16], and attracts much interests. In their ground breaking paper [19], the authors focus on developing Aubry-Mather theory for certain contact Hamiltonian called conformally symplectic, i.e., H(q,p,u)=λu+h(q,p)H(q,p,u)=\lambda u+h(q,p) with λ>0\lambda>0 being a constant. They successfully apply their results to investigate the global dissipative dynamics of the system.

In the series of works [24]-[26], the authors found an implicitly defined variational principle for general contact Hamiltonian systems, and use it to built the variational theory in the spirit of [14] and [17, 18]. In this note, we concentrate on the application of this theory to problems with more dynamical ingredients. Precisely, the aim of this note is to provide our answers to Question 1.5 by methods developed in particular in [24]-[26]. With slight abuse of notation, we use ||q|\cdot|_{q} to denote the norm induced on the cotangent space TqMT^{\ast}_{q}M. In the following context, we shall restrict ourselves to consider the case when

  • Λ,Λ0\Lambda_{-},\Lambda_{0} are Legendrian graphs, i.e., there is u,u0C(M,)u_{-},u_{0}\in C^{\infty}(M,\mathbb{R}) such that

    Λ=Λu,Λ0=Λu0.\Lambda_{-}=\Lambda_{u_{-}},\quad\Lambda_{0}=\Lambda_{u_{0}}.

    It follows from Definition 1.3 that uu_{-} is a smooth solution to the equation

    H(q,dqu(q),u(q))=0,for anyqM.H(q,d_{q}u(q),u(q))=0,\quad\text{for any}\,\,q\in M. (HJs)

and a non-equilibrium thermodynamics system (H,Λ)(H,\Lambda_{-}) with the Hamiltonian satisfying

  • (H2)

    2Hp2(q,p,u)\frac{\partial^{2}H}{\partial p^{2}}(q,p,u) is positive definite for every (q,p,u)Σ(q,p,u)\in\Sigma and for every (q,u)M×(q,u)\in M\times\mathbb{R},

    lim|p|q+H(q,p,u)|p|q+.\lim_{|p|_{q}\rightarrow+\infty}\frac{H(q,p,u)}{|p|_{q}}\rightarrow+\infty.

Now we introduce the following

Definition 1.6.

A sub (resp. super)-deformation of uu_{-} is a function VC(M×[0,1],)V\in C^{\infty}(M\times[0,1],\mathbb{R}) such that

  • V(,1)=uV(\cdot,1)=u_{-},

  • H|ΛV(,s)<0(resp.>0)for alls[0,1)H|_{\Lambda_{V(\cdot,s)}}<0\,(\text{resp.}\,>0)\quad\text{for all}\,\,s\in[0,1).

The name comes from the fact that for s[0,1),V(,s)s\in[0,1),V(\cdot,s) is a strict subsolution (resp. supersolution) to the equation (HJs), see Definition 2.2 in Section 2.

For σΣ,ω(σ)\sigma\in\Sigma,\omega(\sigma) denotes the omega-limit set of σ\sigma under φHt\varphi^{t}_{H} (in general maybe empty!). The first result concerns the construction of semi-infinite orbits connecting Λ0\Lambda_{0} to Λ\Lambda_{-} by using Definition 1.6.

Theorem 1.7.

Let (H,Λ)(H,\Lambda_{-}) be a system with HC(Σ,)H\in C^{\infty}(\Sigma,\mathbb{R}) satisfies (H2). Assume one of the following conditions

  1. (a)(a)

    there is a sub-deformation V:M×[0,1]V:M\times[0,1]\rightarrow\mathbb{R} of uu_{-} such that V(,0)u0uV(\cdot,0)\leq u_{0}\leq u_{-},

  2. (b)(b)

    there is a super-deformation V:M×[0,1]V:M\times[0,1]\rightarrow\mathbb{R} of uu_{-} such that V(,0)u0uV(\cdot,0)\geq u_{0}\geq u_{-},

  3. (c)(c)

    there are a sub-deformation V¯:M×[0,1]\underline{V}:M\times[0,1]\rightarrow\mathbb{R} and a super-deformation V¯:M×[0,1]\overline{V}:M\times[0,1]\rightarrow\mathbb{R} of uu_{-} such that

    V¯(,0)min{u0,u},max{u0,u}V¯(,0).\underline{V}(\cdot,0)\leqslant\min\{u_{0},u_{-}\},\quad\max\{u_{0},u_{-}\}\leqslant\overline{V}(\cdot,0).

then it follows that

  1. (1)

    Λt0φHt(Λ0)¯\Lambda_{-}\subset\overline{\cup_{t\geqslant 0}\varphi^{t}_{H}(\Lambda_{0})},

  2. (2)

    and there is σ0Λ0\sigma_{0}\in\Lambda_{0} such that ω(σ0)Λ\omega(\sigma_{0})\subset\Lambda_{-}.

Remark 1.8.

Conclusion (1) means that every point on the set of equilibria can be approximated by some finite time-length trajectories of the thermodynamic process initiating from Λ0\Lambda_{0}.

If the local stability of Λ\Lambda_{-} is assumed, then the conditions (a) and (b) can be replaced by some condition depending only on 0-jets. This is included in

Theorem 1.9.

Let (H,Λ)(H,\Lambda_{-}) be a system with HC(Σ,)H\in C^{\infty}(\Sigma,\mathbb{R}) satisfies (H1)-(H2). Assume one of the following conditions

  1. (a)(a^{\prime})

    there is a continuous function V:M×[0,1]V:M\times[0,1]\rightarrow\mathbb{R} such that V(,0)C(M,)V(\cdot,0)\in C^{\infty}(M,\mathbb{R}) with H|ΛV(,0)<0H|_{\Lambda_{V(\cdot,0)}}<0 and

    H(q,dqu(q),V(q,s))<0,qMandV(,0)u0u=V(,1),H(q,d_{q}u_{-}(q),V(q,s))<0,\quad\forall q\in M\quad\text{and}\quad V(\cdot,0)\leq u_{0}\leq u_{-}=V(\cdot,1),
  2. (b)(b^{\prime})

    there is a continuous function V:M×[0,1]V:M\times[0,1]\rightarrow\mathbb{R} such that V(,0)C(M,)V(\cdot,0)\in C^{\infty}(M,\mathbb{R}) with H|ΛV(,0)>0H|_{\Lambda_{V(\cdot,0)}}>0 and

    H(q,dqu(q),V(q,s))>0,qMandV(,0)u0u=V(,1),H(q,d_{q}u_{-}(q),V(q,s))>0,\quad\forall q\in M\quad\text{and}\quad V(\cdot,0)\geq u_{0}\geq u_{-}=V(\cdot,1),
  3. (c)(c^{\prime})

    there are continuous functions V¯,V¯:M×[0,1]\underline{V},\overline{V}:M\times[0,1]\rightarrow\mathbb{R} such that V¯(,0),V¯(,0)C(M,)\underline{V}(\cdot,0),\overline{V}(\cdot,0)\in C^{\infty}(M,\mathbb{R}) with H|ΛV¯(,0)<0,H|ΛV¯(,0)>0H|_{\Lambda_{\underline{V}(\cdot,0)}}<0,H|_{\Lambda_{\overline{V}(\cdot,0)}}>0 and

    H(q,dqu(q),V¯(q,s))<0,H(q,dqu(q),V¯(q,s))>0,\displaystyle H(q,d_{q}u_{-}(q),\underline{V}(q,s))<0,\quad\quad H(q,d_{q}u_{-}(q),\overline{V}(q,s))>0,
    V¯(q,0)min{u0(q),u(q)},max{u0(q),u(q)}V¯(q,0).\displaystyle\underline{V}(q,0)\leqslant\min\{u_{0}(q),u_{-}(q)\},\quad\,\max\{u_{0}(q),u_{-}(q)\}\leqslant\overline{V}(q,0).

is satisfied, then there is σ0Λ0\sigma_{0}\in\Lambda_{0} such that ω(σ0)Λ\omega(\sigma_{0})\subset\Lambda_{-}.

Remark 1.10.

The conditions (a)(c)(a)-(c), (a)(c)(a^{\prime})-(c^{\prime}) listed above are stated in a homotopy flavor. They give examples, in our informal opinion, of “weak” version of interlink property employed in [12] that are more easy to verify directly on the contact Hamiltonian.

1.5 Organization of the paper

The remaining of this paper is organized as follows. In Section 2, we briefly recall some necessary tools from [24]-[26] and give an extension of characteristic theory, which is crucial in our proof of Theorem 1.7. Section 3 is devoted to the construction of semi-infinite orbit asymptotically converges to Λ\Lambda_{-} when the solution semigroup associated to the evolutionary Hamilton-Jacobi equation converges. In Section 4, our homotopy criteria are verified to guarantee the convergence of the solution semigroup with initial data u0u_{0}. We also illustrate our results on some examples, including some generalizations of those from [12], in the last section.

2 Global characteristics theory via variational methods

To extend the characteristic theory to the global setting, we recall the variational methods developed for evolutionary Hamilton-Jacobi equation (including viscosity solutions theory) and the associated contact Hamiltonian system. A global version of characteristic theory is obtained by showing that viscosity solutions propagate along action minimizing orbits of (1.2). Notice that many objects discussed in this section is non-smooth in the classical viewpoint, thus necessary smoothness to guarantee the validity of definitions and theorems is presented in an accurate way.

The classical characteristics theory connects the local solvability of the Cauchy problem of the evolutionary Hamilton-Jacobi equation

{tU+H(q,qU,U)=0,(q,t)M×(0,+),U(,0)=v,qM,\begin{cases}\partial_{t}U+H(q,\partial_{q}U,U)=0,\quad\,(q,t)\in M\times(0,+\infty),\\ \hskip 54.00002ptU(\cdot,0)=v,\hskip 26.49997ptq\in M,\end{cases} (HJe)

to the study of contact Hamiltonian system (1.2) near Λv\Lambda_{v}, here vC2(M)v\in C^{2}(M) is a smooth initial data. More precisely, if one assume U:M×[0,+)U:M\times[0,+\infty)\rightarrow\mathbb{R} is a C2C^{2} solution to (HJe), then every trajectory φHtσ:=(q(t),p(t),u(t)),t0\varphi^{t}_{H}\sigma:=(q(t),p(t),u(t)),t\geq 0 of (1.2), called characteristic, starting from σΛv\sigma\in\Lambda_{v} satisfies the identities

p(t)=qU(q(t),t),u(t)=U(q(t),t).p(t)=\partial_{q}U(q(t),t),\quad u(t)=U(q(t),t). (2.1)

Equivalently, for every t0,{φHtσ:σΛv}=ΛU(,t)t\geq 0,\{\varphi^{t}_{H}\sigma:\sigma\in\Lambda_{v}\}=\Lambda_{U(\cdot,t)}. Following this spirit, one arrives at

Theorem 2.1.

[2, Lecture 2, Theorem 3] or [10, Chapter 3, Theorem 2] For any vC2(M,)v\in C^{2}(M,\mathbb{R}), there are δ>0\delta>0 and a solution UC2(M×[0,δ],)U\in C^{2}(M\times[0,\delta],\mathbb{R}) to (HJe), so that

  1. (1)

    for any t[0,δ],πqφHt:ΛvMt\in[0,\delta],\pi_{q}\circ\varphi^{t}_{H}:\Lambda_{v}\rightarrow M is a diffeomorphism,

  2. (2)

    any characteristic segment φHtσ=(q(t),p(t),u(t)),σΛv,t[0,δ]\varphi^{t}_{H}\sigma=(q(t),p(t),u(t)),\sigma\in\Lambda_{v},t\in[0,\delta] satisfies the identities (2.1).

Notice that in general, given a smooth initial data vv, Theorem 2.1 only allow us to construct smooth solution to (HJe) locally. The reason comes from the fact that, after the projection by πq\pi_{q}, the characteristics starting from Λv\Lambda_{v} may intersect at some large tt. Thus even for smooth initial data vv, there does not exist a global solution UC2(M×[0,),)U\in C^{2}(M\times[0,\infty),\mathbb{R}) to (HJe). To construct solutions to (HJe), it is necessary to extend the notion of ‘solutions’ to include non-smooth functions. The right one, namely viscosity solution, was firstly introduced by M.Crandall and P.L.Lions in [7], and is now widely accepted as the natural framework for the theory of Hamilton-Jacobi equations and certain second order PDEs.

Definition 2.2.

A continuous function u:Mu:M\rightarrow\mathbb{R} is called a viscosity subsolution (resp. supersolution) of (HJs) if for any qMq\in M and ϕC1(M,)\phi\in C^{1}(M,\mathbb{R}) such that uϕu-\phi attains a local maximum (resp. minimum) at qq,

H(q,dqϕ(q),ϕ(q))(resp.)  0;H(q,d_{q}\phi(q),\phi(q))\leq(\text{resp.}\geq)\,\,0; (2.2)

uu is called a viscosity solution if it is both a viscosity sub and supersolution of (HJs). Moreover, a viscosity subsolution is said to be strict if the inequality \leqslant (2.2) is replaced by << at any qMq\in M.

A continuous function U:M×[0,)U:M\times[0,\infty)\rightarrow\mathbb{R} is called a viscosity subsolution (resp. supersolution) of (HJe) if U(,0)(resp.)vU(\cdot,0)\leq(\text{resp.}\geq)\,\,v on MM and for any (q,t)M×(0,+)(q,t)\in M\times(0,+\infty) and Φ\Phi a C1C^{1} function defined on a neighborhood of (q,t)(q,t) such that UΦU-\Phi attains a local maximum (resp. minimum) at (q,t)(q,t),

tΦ(q,t)+H(q,qΦ(q,t),U(q,t))(resp.)  0;\partial_{t}\Phi(q,t)+H(q,\partial_{q}\Phi(q,t),U(q,t))\leq(\text{resp.}\geq)\,\,0;

UU is called a viscosity solution if it is both a viscosity sub and supersolution of (HJe).

From now on, solutions to (HJs) and (HJe) are always understood in the viscosity sense. Now we begin to give a brief summary of results in [24]-[26] concerning the variational part of the theory of viscosity solutions to (HJe) and (HJs).

2.1 A variational principle associated to (HJe)

Let TMTM denote the tangent bundle of MM. A point of TMTM will be denoted by (q,q˙)(q,\dot{q}), where qMq\in M and q˙TqM\dot{q}\in T_{q}M. Recall that pTqMp\in T^{\ast}_{q}M is a linear form on TqMT_{q}M, we use ,\langle\cdot,\cdot\rangle to denote the canonical pairing between tangent and cotangent bundle. For a contact Hamiltonian HC3(Σ,)H\in C^{3}(\Sigma,\mathbb{R}) satisfying (H2), we define the corresponding Lagrangian L:TM×L:TM\times\mathbb{R}\rightarrow\mathbb{R} by

L(q,q˙,u)=suppTqM{p,q˙H(q,p,u)},L(q,\dot{q},u)=\sup_{p\in T_{q}^{\ast}M}\{\langle p,\dot{q}\rangle-H(q,p,u)\},

i.e., LL is the convex dual of HH with respect to pp. The following action functions provide a formulation of the variational principle defined by the equation (HJe). Notice that the action function is implicitly defined since HH depends on the uu-variable. In [5][6], the authors show that Hoglotz’ variational principle also is a effective tool.

Proposition 2.3.

[26, Theorem 2.1, 2.2] Given any (q0,u0)M×(q_{0},u_{0})\in M\times\mathbb{R}, there exist two continuous functions hq0,u0(q,t)h_{q_{0},u_{0}}(q,t) and hq0,u0(q,t)h^{q_{0},u_{0}}(q,t) called the backward and forward action function respectively, defined on M×(0,+)M\times(0,+\infty) by

hq0,u0(q,t)=\displaystyle h_{q_{0},u_{0}}(q,t)= infγ(t)=qγ(0)=q0{u0+0tL(γ(τ),γ˙(τ),hq0,u0(γ(τ),τ))𝑑τ},\displaystyle\inf_{\begin{subarray}{c}\gamma(t)=q\\ \gamma(0)=q_{0}\end{subarray}}\Big{\{}u_{0}+\int_{0}^{t}L(\gamma(\tau),\dot{\gamma}(\tau),h_{q_{0},u_{0}}(\gamma(\tau),\tau))\ d\tau\Big{\}}, (2.3)
hq0,u0(q,t)=\displaystyle h^{q_{0},u_{0}}(q,t)= supγ(t)=q0γ(0)=q{u00tL(γ(τ),γ˙(τ),hq0,u0(γ(τ),tτ))𝑑τ},\displaystyle\sup_{\begin{subarray}{c}\gamma(t)=q_{0}\\ \gamma(0)=q\end{subarray}}\Big{\{}u_{0}-\int_{0}^{t}L(\gamma(\tau),\dot{\gamma}(\tau),h^{q_{0},u_{0}}(\gamma(\tau),t-\tau))\ d\tau\Big{\}}, (2.4)

where the infimum and supremum are taken among Lipschitz continuous curves γ:[0,t]M\gamma:[0,t]\rightarrow M and are achieved. Moreover, if γ1\gamma_{1} and γ2\gamma_{2} achieve the infimum in (2.3) and supremum in (2.4) respectively, then γ1,γ2C1([0,t],M)\gamma_{1},\gamma_{2}\in C^{1}([0,t],M). Set

q1(τ):=γ1(τ),u1(τ):=hq0,u0(γ1(τ),τ),p1(τ):=Lq˙(γ1(τ),γ˙1(τ),u1(τ)),\displaystyle q_{1}(\tau):=\gamma_{1}(\tau),\quad u_{1}(\tau):=h_{q_{0},u_{0}}(\gamma_{1}(\tau),\tau),\hskip 30.00005ptp_{1}(\tau):=\frac{\partial L}{\partial\dot{q}}(\gamma_{1}(\tau),\dot{\gamma}_{1}(\tau),u_{1}(\tau)),
q2(τ):=γ2(τ),u2(τ):=hq0,u0(γ2(τ),tτ)),p2(τ):=Lq˙(γ2(τ),γ˙2(τ),u2(τ)),\displaystyle q_{2}(\tau):=\gamma_{2}(\tau),\quad u_{2}(\tau):=h^{q_{0},u_{0}}(\gamma_{2}(\tau),t-\tau)),\quad p_{2}(\tau):=\frac{\partial L}{\partial\dot{q}}(\gamma_{2}(\tau),\dot{\gamma}_{2}(\tau),u_{2}(\tau)),

then (q1(τ),p1(τ),u1(τ))(q_{1}(\tau),p_{1}(\tau),u_{1}(\tau)) and (q2(τ),p2(τ),u2(τ))(q_{2}(\tau),p_{2}(\tau),u_{2}(\tau)) satisfy (1.2) with

q1(0)=q0,q1(t)=q,limτ0+u1(τ)=u0,\displaystyle q_{1}(0)=q_{0},\quad q_{1}(t)=q,\quad\lim_{\tau\to 0^{+}}u_{1}(\tau)=u_{0},
q2(0)=q,q2(t)=q0,limτtu2(τ)=u0.\displaystyle q_{2}(0)=q,\quad q_{2}(t)=q_{0},\quad\lim_{\tau\to t^{-}}u_{2}(\tau)=u_{0}.

As a direct consequence of Proposition 2.3, we obtain

Corollary 2.4.

Given q0,qM,u0q_{0},q\in M,u_{0}\in\mathbb{R} and t>0t>0, set (q(τ),p(τ),u(τ))=φHτ(q(0),p(0),u(0))(q(\tau),p(\tau),u(\tau))=\varphi^{\tau}_{H}(q(0),p(0),u(0)) and

Sq0,u0q,t={(q(τ),p(τ),u(τ)),τ[0,t]:q(0)=q0,q(t)=q,u(0)=u0},S^{q,t}_{q_{0},u_{0}}=\big{\{}(q(\tau),p(\tau),u(\tau)),\tau\in[0,t]\,:\,q(0)=q_{0},q(t)=q,u(0)=u_{0}\big{\}},
Sq,tq0,u0={(q(τ),p(τ),u(τ)),τ[0,t]:q(0)=q,q(t)=q0,u(t)=u0},S^{q_{0},u_{0}}_{q,t}=\big{\{}(q(\tau),p(\tau),u(\tau)),\tau\in[0,t]\,:\,q(0)=q,q(t)=q_{0},u(t)=u_{0}\big{\}},

then for any (q,t)M×(0,+)(q,t)\in M\times(0,+\infty),

hq0,u0(q,t)=inf{u(t):(q(τ),p(τ),u(τ))Sq0,u0q,t},h_{q_{0},u_{0}}(q,t)=\inf\,\{u(t):(q(\tau),p(\tau),u(\tau))\in S^{q,t}_{q_{0},u_{0}}\}, (2.5)
hq0,u0(q,t)=sup{u(0):(q(τ),p(τ),u(τ))Sq,tq0,u0}.h^{q_{0},u_{0}}(q,t)=\sup\{u(0):(q(\tau),p(\tau),u(\tau))\in S^{q_{0},u_{0}}_{q,t}\}. (2.6)

We collect some fundamental properties of the action functions here, which are frequently used in the later context. For details and proofs of these properties, we refer to the paper [25].

Proposition 2.5.

[25] The backward and forward action functions satisfy

  1. (1)

    (u0u_{0}-monotonicity) Given q0M,u1<u2q_{0}\in M,u_{1}<u_{2}\in\mathbb{R},  for all (q,t)M×(0,)(q,t)\in M\times(0,\infty),

    hq0,u1(q,t)<hq0,u2(q,t),hq0,u1(q,t)<hq0,u2(q,t).h_{q_{0},u_{1}}(q,t)<h_{q_{0},u_{2}}(q,t),\quad h^{q_{0},u_{1}}(q,t)<h^{q_{0},u_{2}}(q,t).
  2. (2)

    (Markov property) Given (x0,u0)M×(x_{0},u_{0})\in M\times\mathbb{R},  for all t,τ>0t,\tau>0 and qMq\in M,

    hq0,u0(q,t+τ)=infq1Mhq1,hq0,u0(q1,t)(q,τ),hq0,u0(q,t+τ)=supq1Mhq1,hq0,u0(q1,t)(q,τ).\begin{split}&h_{q_{0},u_{0}}(q,t+\tau)=\inf_{q_{1}\in M}h_{q_{1},h_{q_{0},u_{0}}(q_{1},t)}(q,\tau),\\ &h^{q_{0},u_{0}}(q,t+\tau)=\sup_{q_{1}\in M}h^{q_{1},h^{q_{0},u_{0}}(q_{1},t)}(q,\tau).\end{split} (2.7)

    Moreover, the infimum is attained at q1q_{1} if and only if there exists a C1C^{1} minimizer γ\gamma of hq0,u0(q,t+τ)h_{q_{0},u_{0}}(q,t+\tau) with γ(t)=q1\gamma(t)=q_{1}, the supremum is attained at q1q_{1} if and only if there exists a C1C^{1} minimizer γ\gamma of hq0,u0(q,t+τ)h^{q_{0},u_{0}}(q,t+\tau) with γ(t)=q1\gamma(t)=q_{1}.

  3. (3)

    (Lipschitz continuity) The functions

    (q0,u0,q,t)hq0,u0(q,t),(q0,u0,q,t)hq0,u0(q,t)(q_{0},u_{0},q,t)\mapsto h_{q_{0},u_{0}}(q,t),\quad(q_{0},u_{0},q,t)\mapsto h^{q_{0},u_{0}}(q,t)

    are locally Lipschitz continuous on the domain M××M×(0,+)M\times\mathbb{R}\times M\times(0,+\infty).

It turns out that any solution to (HJe) can be expressed by action functions. The representation involves some families of nonlinear operators which we now introduce.

Definition 2.6 (Solution semigroups).

For each vC(M,)v\in C(M,\mathbb{R}) and (q,t)M×(0,+)(q,t)\in M\times(0,+\infty), define

Ttv(q):=infq0Mhq0,v(q0)(q,t),Tt+v(q):=supq0Mhq0,v(q0)(q,t).\begin{split}T^{-}_{t}v(q):=\inf_{q_{0}\in M}h_{q_{0},v(q_{0})}(q,t),\\ T^{+}_{t}v(q):=\sup_{q_{0}\in M}h^{q_{0},v(q_{0})}(q,t).\end{split} (2.8)

In addition, we set T0±v(q)=v(q)T^{\pm}_{0}v(q)=v(q), then for t0,Tt±:vTt±vt\geq 0,\,\,T^{\pm}_{t}:v\mapsto T^{\pm}_{t}v maps C(M,)C(M,\mathbb{R}) to itself.

The above definition allow us to deduce some properties of solutions semigroups from Proposition 2.5 as corollaries. In particular, we have

Proposition 2.7.

[25, Proposition 4.3] Two families of operator {Tt±}t0\{T^{\pm}_{t}\}_{t\geqslant 0} defined above satisfy

  1. (1)

    (monotonicity) For initial data v,vC(M,)v,v^{\prime}\in C(M,\mathbb{R}) with v<vv<v^{\prime} (resp. vvv\leqslant v^{\prime}) on MM, then for all qMq\in M,

    Tt±v(q)<Tt±v(q),resp.(Tt±v(q)Tt±v(q)).T^{\pm}_{t}v(q)<T^{\pm}_{t}v^{\prime}(q),\quad\text{resp.}\,\,(T^{\pm}_{t}v(q)\leqslant T^{\pm}_{t}v^{\prime}(q)). (2.9)
  2. (2)

    (Semigroup property) For any t,τ0t,\tau\geqslant 0,

    Tt+τ±=Tt±Tτ±,T^{\pm}_{t+\tau}=T^{\pm}_{t}\circ T^{\pm}_{\tau}, (2.10)

    so that the families of operators {Tt±}t0\{T^{\pm}_{t}\}_{t\geq 0} form two semigroups acting on C(M,)C(M,\mathbb{R}).

  3. (3)

    (Continuity 1) For any (q,t)M×(0,+)(q,t)\in M\times(0,+\infty), the functions

    (q,t)Tt±v(q),(q,t)\mapsto T^{\pm}_{t}v(q),

    are locally Lipschitz continuous and limt0+Tt±v(q)=v(q)\lim_{t\rightarrow 0^{+}}T^{\pm}_{t}v(q)=v(q) for all qMq\in M.

  4. (4)

    (Continuity 2) For any t0t\geqslant 0, the maps

    vTt±vv\mapsto T^{\pm}_{t}v

    are continuous with respect to \|\cdot\|_{\infty} defined on C(M,)C(M,\mathbb{R}).

It turns out that the notion of subsolution (resp. strict subsolution) is equivalent to the tt-monotonicity (-strict monotonicity) of the solution semigroups. The following proposition can be easily seen from the form of equation (HJe).

Proposition 2.8.

Let vC(M,)v\in C(M,\mathbb{R}) be a subsolution (resp. strict subsolution) to (HJs), then

  1. (1)

    for any qMq\in M and t0t\geqslant 0 (resp. t>0t>0), v(q)v(q)\leqslant  (resp. <<)  Ttv(q)T^{-}_{t}v(q),

  2. (2)

    for any qMq\in M and t0t\geqslant 0 (resp. t>0t>0), v(q)v(q)\geqslant  (resp. >>)  Tt+v(q)T^{+}_{t}v(q).

2.2 Solution semigroups and their characteristics

For a general initial data vC(M,)v\in C(M,\mathbb{R}), we define U:M×[0,)U:M\times[0,\infty)\rightarrow\mathbb{R} by

U(q,t):=Ttv(q)=infqMhq,v(q)(q,t).U(q,t):=T^{-}_{t}v(q)=\inf_{q^{\prime}\in M}h_{q^{\prime},v(q^{\prime})}(q,t). (2.11)

By Proposition 2.5 (3), fixing (q,t)M×(0,+)(q,t)\in M\times(0,+\infty), the map

qhq,v(q)(q,t)q^{\prime}\mapsto h_{q^{\prime},v(q^{\prime})}(q,t)

is continuous. Then there is a q0Mq_{0}\in M such that U(q,t)=hq0,v(q0)(q,t)U(q,t)=h_{q_{0},v(q_{0})}(q,t). Due to the properties of backward action function, we have

Lemma 2.9.

For any minimizer γ:[0,t]M\gamma:[0,t]\rightarrow M of hq0,v(q0)(q,t)h_{q_{0},v(q_{0})}(q,t),

U(γ(τ),τ)=hq0,v(q0)(γ(τ),τ).U(\gamma(\tau),\tau)=h_{q_{0},v(q_{0})}(\gamma(\tau),\tau).
Proof.

By (2.11), we only need to show that for any τ[0,t]\tau\in[0,t],

U(γ(τ),τ)hq0,v(q0)(γ(τ),τ).U(\gamma(\tau),\tau)\geq h_{q_{0},v(q_{0})}(\gamma(\tau),\tau).

We argue by contradiction. Assume there is τ0(0,t)\tau_{0}\in(0,t) such that

u¯:=U(γ(τ0),τ0)<u¯:=hq0,v(q0)(γ(τ0),τ0),\underline{u}:=U(\gamma(\tau_{0}),\tau_{0})<\bar{u}:=h_{q_{0},v(q_{0})}(\gamma(\tau_{0}),\tau_{0}),

then to complete the proof, it is necessary to see that

U(q,t)=Ttτ0U(,τ0)(q)hγ(τ0),u¯(q,tτ0)<hγ(τ0),u¯(q,tτ0)=hq0,v(q0)(q,t)=U(q,t).U(q,t)=T^{-}_{t-\tau_{0}}U(\cdot,\tau_{0})(q)\leqslant h_{\gamma(\tau_{0}),\underline{u}}(q,t-\tau_{0})<h_{\gamma(\tau_{0}),\bar{u}}(q,t-\tau_{0})=h_{q_{0},v(q_{0})}(q,t)=U(q,t).

Here, the first equality follows from property (2.10) and the second equality is a consequence of Proposition 2.5 (2) and the fact that γ\gamma is a minimizer of hq0,v(q0)(q,t)h_{q_{0},v(q_{0})}(q,t); the second inequality is deduced from Proposition 2.5 (1). ∎

The following well-known theorem gives the name of the operator families defined in Definition 2.6.

Proposition 2.10.

[25, Proposition 4.4] UU is the unique solution to (HJe). In a similar fashion,

U(q,t)=Tt+v(q)U(q,t)=-T^{+}_{t}v(q)

is the unique solution to

{tU+H˘(q,qU,U)=0,(q,t)M×(0,+),U(,0)=v,qM,\begin{cases}\partial_{t}U+\breve{H}(q,\partial_{q}U,U)=0,\quad\,(q,t)\in M\times(0,+\infty),\\ \hskip 54.00002ptU(\cdot,0)=v,\hskip 26.49997ptq\in M,\end{cases} (2.12)

where H˘(q,p,u)=H(q,p,u)\breve{H}(q,p,u)=H(q,-p,-u). Due to these facts, we call {Tt}t0\{T^{-}_{t}\}_{t\geqslant 0} the backward solution semigroup and {Tt+}t0\{T^{+}_{t}\}_{t\geqslant 0} the forward solution semigroup to (HJe).

We need the fact that the qq-projection of the characteristic ensured by Theorem 2.1 is a minimizer in sense of Proposition 2.3. Notice that by Theorem 2.1, the map πqφHδ:ΛvM\pi_{q}\circ\varphi^{\delta}_{H}:\Lambda_{v}\rightarrow M is a diffeomorphism, we use (πqφHδ)1:MΛv(\pi_{q}\circ\varphi^{\delta}_{H})^{-1}:M\rightarrow\Lambda_{v} to denote its inverse.

Lemma 2.11.

For any q1Mq_{1}\in M and σ0=(q0,dqv(q0),v(q0))=(πqφHδ)1(q1)\sigma_{0}=(q_{0},d_{q}v(q_{0}),v(q_{0}))=(\pi_{q}\circ\varphi^{\delta}_{H})^{-1}(q_{1}), set

φHτσ0=(q(τ),p(τ),u(τ))forτ[0,δ],\varphi^{\tau}_{H}\sigma_{0}=(q(\tau),p(\tau),u(\tau))\quad\text{for}\,\,\tau\in[0,\delta],

then for all τ[0,δ],u(τ)=hq0,v(q0)(q(τ),τ)\tau\in[0,\delta],u(\tau)=h_{q_{0},v(q_{0})}(q(\tau),\tau) and

hq0,v(q0)(q1,δ)=v(q0)+0δL(q(τ),q˙(τ),hq0,v(q0)(q(τ),τ))𝑑τ.h_{q_{0},v(q_{0})}(q_{1},\delta)=v(q_{0})+\int_{0}^{\delta}L(q(\tau),\dot{q}(\tau),h_{q_{0},v(q_{0})}(q(\tau),\tau))\ d\tau.
Proof.

By Theorem 2.1, UC2(M×[0,δ],)U\in C^{2}(M\times[0,\delta],\mathbb{R}) and (q(τ),p(τ),u(τ))(q(\tau),p(\tau),u(\tau)) satisfy

p(τ)=qU(q(τ),τ),u(τ)=U(q(τ),τ),τ[0,δ].p(\tau)=\partial_{q}U(q(\tau),\tau),\quad u(\tau)=U(q(\tau),\tau),\quad\tau\in[0,\delta]. (2.13)

and the boundary conditions read as

q(0)=q0,q(δ)=q1,u(0)=v(q0).q(0)=q_{0},\quad q(\delta)=q_{1},\quad u(0)=v(q_{0}). (2.14)

It follows from Proposition 2.10 that

hq0,v(q0)(q(τ),τ)Tτv(q(τ))=U(q(τ),τ)=u(τ).h_{q_{0},v(q_{0})}(q(\tau),\tau)\geq T^{-}_{\tau}v(q(\tau))=U(q(\tau),\tau)=u(\tau).

Combining (2.14) and Corollary 2.4 gives

hq0,v(q0)(q(τ),τ)=inf{u(τ):(q(t),p(t),u(t))Sq0,v(q0)q(τ),τ}u(τ)h_{q_{0},v(q_{0})}(q(\tau),\tau)=\inf\,\{u(\tau):(q(t),p(t),u(t))\in S^{q(\tau),\tau}_{q_{0},v(q_{0})}\}\leq u(\tau)

and therefore for τ[0,δ]\tau\in[0,\delta],

u(τ)=hq0,v(q0)(q(τ),τ).u(\tau)=h_{q_{0},v(q_{0})}(q(\tau),\tau).

Now we can compute as

hq0,v(q0)(q1,δ)U(q1,δ)=U(q(δ),δ)\displaystyle h_{q_{0},v(q_{0})}(q_{1},\delta)\geq U(q_{1},\delta)=U(q(\delta),\delta)
=\displaystyle=\, U(q(0),0)+0δtU(q(τ),τ)+qU(q(τ),τ),q˙(τ)dτ\displaystyle U(q(0),0)+\int^{\delta}_{0}\partial_{t}U(q(\tau),\tau)+\langle\partial_{q}U(q(\tau),\tau),\dot{q}(\tau)\rangle\,d\tau
=\displaystyle=\, v(q0)+0δtU(q(τ),τ)+p(τ),pH(q(τ),p(τ),u(τ))dτ\displaystyle v(q_{0})+\int^{\delta}_{0}\partial_{t}U(q(\tau),\tau)+\langle p(\tau),\partial_{p}H(q(\tau),p(\tau),u(\tau))\rangle\,d\tau
=\displaystyle=\, v(q0)+0δtU(q(τ),τ)+H(q(τ),p(τ),u(τ))+L(q(τ),q˙(τ),u(τ))dτ\displaystyle v(q_{0})+\int^{\delta}_{0}\partial_{t}U(q(\tau),\tau)+H(q(\tau),p(\tau),u(\tau))+L(q(\tau),\dot{q}(\tau),u(\tau))\,d\tau
=\displaystyle=\, v(q0)+0δL(q(τ),q˙(τ),u(τ))𝑑τ\displaystyle v(q_{0})+\int^{\delta}_{0}L(q(\tau),\dot{q}(\tau),u(\tau))\,d\tau
=\displaystyle=\, v(q0)+0δL(q(τ),q˙(τ),hq0,v(q0)(q(τ),τ))𝑑τ.\displaystyle v(q_{0})+\int^{\delta}_{0}L(q(\tau),\dot{q}(\tau),h_{q_{0},v(q_{0})}(q(\tau),\tau))\,d\tau.

Here, the third equality uses (2.13) and the equations (1.2) for characteristics, the fourth equality follows from the knowledge of Legendre-Fenchel inequality in convex analysis, i.e.,

q˙=pH(q,p,u)p,q˙=H(q,p,u)+L(q,q˙,u),\dot{q}=\partial_{p}H(q,p,u)\Leftrightarrow\langle p,\dot{q}\rangle=H(q,p,u)+L(q,\dot{q},u),

the fifth equality is due to (2.13) and the fact that U(q,t)U(q,t) is a solution to (HJe), precisely

tU(q(τ),τ)+H(q(τ),p(τ),u(τ))\displaystyle\,\partial_{t}U(q(\tau),\tau)+H(q(\tau),p(\tau),u(\tau))
=\displaystyle= tU(q(τ),τ)+H(q(τ),qU(q(τ),τ),U(q(τ),τ))=0.\displaystyle\,\partial_{t}U(q(\tau),\tau)+H(q(\tau),\partial_{q}U(q(\tau),\tau),U(q(\tau),\tau))=0.

Combining the above inequality and (2.3), we complete the proof. ∎

The above lemma justifies the fact that the characteristics initiating from Λv\Lambda_{v} from which the local smooth solution is constructed by Theorem 2.1 are actually action minimizers of TtvT^{-}_{t}v. This is true not only for local solutions, in fact we have

Theorem 2.12.

Assume vC2(M,)v\in C^{2}(M,\mathbb{R}). For any (q,t)M×(0,+)(q,t)\in M\times(0,+\infty), there is σ0Λv\sigma_{0}\in\Lambda_{v} such that the characteristic segment φHτσ0=(q(τ),p(τ),u(τ)),τ[0,t]\varphi^{\tau}_{H}\sigma_{0}=(q(\tau),p(\tau),u(\tau)),\tau\in[0,t] satisfies

U(q(τ),τ)=u(τ).U(q(\tau),\tau)=u(\tau).
Proof.

For tδt\leq\delta, Theorem 2.12 reduces to Theorem 2.1 and there is nothing to prove. For t>δt>\delta, we use Definition 2.6 and (2.10) to write

U(q,t)=Ttv(q)=TtδTδv(q)=infqMhq,Tδv(q)(q,tδ).U(q,t)=T^{-}_{t}v(q)=T^{-}_{t-\delta}\circ T^{-}_{\delta}v(q)=\inf_{q^{\prime}\in M}h_{q^{\prime},T^{-}_{\delta}v(q^{\prime})}(q,t-\delta).

Proposition 2.5 (3) and 2.7 (3) imply hq,Tδv(q)(q,t)h_{q^{\prime},T^{-}_{\delta}v(q^{\prime})}(q,t) is Lipschitz continuous in qq^{\prime}. Since MM is compact, the above infimum is attained at q=q1Mq^{\prime}=q_{1}\in M. Set u1=Tδv(q1)u_{1}=T^{-}_{\delta}v(q_{1}), then according to Proposition 2.3, there is a minimizer γ1:[0,tδ]M\gamma_{1}:[0,t-\delta]\rightarrow M with γ1(0)=q1\gamma_{1}(0)=q_{1} and

U(q,t)=hq1,u1(q,tδ)=u1+0tδL(γ1(τ),γ1˙(τ),hq1,u1(γ1(τ),τ))𝑑τ.U(q,t)=h_{q_{1},u_{1}}(q,t-\delta)=u_{1}+\int_{0}^{t-\delta}L(\gamma_{1}(\tau),\dot{\gamma_{1}}(\tau),h_{q_{1},u_{1}}(\gamma_{1}(\tau),\tau))\ d\tau. (2.15)

For τ[δ,t]\tau\in[\delta,t], we set q1(τ)=γ1(τδ)q_{1}(\tau)=\gamma_{1}(\tau-\delta) and

u1(τ):=hq1,u1(γ1(τδ),τδ),p1(τ):=Lq˙(γ1(τδ),γ˙1(τδ),u1(τ)),u_{1}(\tau):=h_{q_{1},u_{1}}(\gamma_{1}(\tau-\delta),\tau-\delta),\quad p_{1}(\tau):=\frac{\partial L}{\partial\dot{q}}(\gamma_{1}(\tau-\delta),\dot{\gamma}_{1}(\tau-\delta),u_{1}(\tau)),

then Proposition 2.3 also implies that (q1(τ),p1(τ),u1(τ))(q_{1}(\tau),p_{1}(\tau),u_{1}(\tau)) satisfies (1.2) and

q1(δ)=q1,q1(t)=q,limτtu1(τ)=u1.\displaystyle q_{1}(\delta)=q_{1},\quad q_{1}(t)=q,\quad\lim_{\tau\rightarrow t^{-}}u_{1}(\tau)=u_{1}.

By Lemma 2.11, there is a unique σ0=(q0,dqv(q0),v(q0))Λv\sigma_{0}=(q_{0},d_{q}v(q_{0}),v(q_{0}))\in\Lambda_{v} such that

πqφHδ(σ0)=q1,\pi_{q}\circ\varphi^{\delta}_{H}(\sigma_{0})=q_{1},

and if φHτσ0=(q0(τ),p0(τ),u0(τ)),τ[0,δ]\varphi^{\tau}_{H}\sigma_{0}=(q_{0}(\tau),p_{0}(\tau),u_{0}(\tau)),\tau\in[0,\delta], then

hq0,v(q0)(q1,δ)=v(q0)+0δL(q0(τ),q˙0(τ),hq0,v(q0)(q0(τ),τ))𝑑τh_{q_{0},v(q_{0})}(q_{1},\delta)=v(q_{0})+\int_{0}^{\delta}L(q_{0}(\tau),\dot{q}_{0}(\tau),h_{q_{0},v(q_{0})}(q_{0}(\tau),\tau))\ d\tau (2.16)

and

hq0,v(q0)(q1,δ)=u0(δ)=U(q1,δ)=Tδv(q1)=u1.h_{q_{0},v(q_{0})}(q_{1},\delta)=u_{0}(\delta)=U(q_{1},\delta)=T^{-}_{\delta}v(q_{1})=u_{1}.

Claim:  for τ[δ,t]\tau\in[\delta,t],

hq0,v(q0)(q1(τ),τ)=hq1,u1(q1(τ),τδ).h_{q_{0},v(q_{0})}(q_{1}(\tau),\tau)=h_{q_{1},u_{1}}(q_{1}(\tau),\tau-\delta). (2.17)

Proof of the claim:  It follows from Proposition 2.5 (2) that

hq0,v(q0)(q1(τ),τ)=infqMhq,hq0,v(q0)(q,δ)(q1(τ),τδ)hq1,u1(q1(τ),τδ).h_{q_{0},v(q_{0})}(q_{1}(\tau),\tau)=\inf_{q^{\prime}\in M}h_{q^{\prime},h_{q_{0},v(q_{0})}(q^{\prime},\delta)}(q_{1}(\tau),\tau-\delta)\leq h_{q_{1},u_{1}}(q_{1}(\tau),\tau-\delta).

Assume for some τ0(δ,t)\tau_{0}\in(\delta,t),

u2:=hq0,v(q0)(q1(τ0),τ0)<hq1,u1(q1(τ0),τ0δ):=u¯2,u_{2}:=h_{q_{0},v(q_{0})}(q_{1}(\tau_{0}),\tau_{0})<h_{q_{1},u_{1}}(q_{1}(\tau_{0}),\tau_{0}-\delta):=\bar{u}_{2},

then by Proposition 2.5 (1) and the fact that γ1\gamma_{1} is a minimizer of hq1,u1(q,tδ)h_{q_{1},u_{1}}(q,t-\delta),

hq0,v(q0)(q,t)hq1(τ0),u2(q,tτ0)<hq1(τ0),u¯2(q,tτ0)=hq1,u1(q,tδ)=U(q,t).h_{q_{0},v(q_{0})}(q,t)\leq h_{q_{1}(\tau_{0}),u_{2}}(q,t-\tau_{0})<h_{q_{1}(\tau_{0}),\bar{u}_{2}}(q,t-\tau_{0})=h_{q_{1},u_{1}}(q,t-\delta)=U(q,t).

This contradicts to Proposition 2.10 since

U(q,t)=Ttv(q)=infqMhq,v(q)(q,t)hq0,v(q0)(q,t).U(q,t)=T^{-}_{t}v(q)=\inf_{q^{\prime}\in M}h_{q^{\prime},v(q^{\prime})}(q,t)\leq h_{q_{0},v(q_{0})}(q,t).

Define q:[0,t]Mq:[0,t]\rightarrow M by

q(τ):={q0(τ),τ[0,δ];q1(τ),τ[δ,t],q(\tau):=\begin{cases}q_{0}(\tau),\quad\tau\in[0,\delta];\\ q_{1}(\tau),\quad\tau\in[\delta,t],\end{cases}

then combining (2.15)-(2.17), we obtain

hq0,v(q0)(q,t)=hq1,u1(q,tδ)=u1+0tδL(γ1(τ),γ1˙(τ),hq1,u1(γ1(τ),τ))𝑑τ\displaystyle\,h_{q_{0},v(q_{0})}(q,t)=h_{q_{1},u_{1}}(q,t-\delta)=u_{1}+\int_{0}^{t-\delta}L(\gamma_{1}(\tau),\dot{\gamma_{1}}(\tau),h_{q_{1},u_{1}}(\gamma_{1}(\tau),\tau))\ d\tau
=\displaystyle= u1+δtL(q1(τ),q1˙(τ),hq1,u1(q1(τ),τδ))𝑑τ\displaystyle\,\,u_{1}+\int_{\delta}^{t}L(q_{1}(\tau),\dot{q_{1}}(\tau),h_{q_{1},u_{1}}(q_{1}(\tau),\tau-\delta))\ d\tau
=\displaystyle= hq0,v(q0)(q1,δ)+δtL(q1(τ),q1˙(τ),hq0,v(q0)(q1(τ),τ))𝑑τ\displaystyle\,\,h_{q_{0},v(q_{0})}(q_{1},\delta)+\int_{\delta}^{t}L(q_{1}(\tau),\dot{q_{1}}(\tau),h_{q_{0},v(q_{0})}(q_{1}(\tau),\tau))\ d\tau
=\displaystyle= v(q0)+0δL(q0(τ),q˙0(τ),hq0,v(q0)(q0(τ),τ))𝑑τ+δtL(q1(τ),q1˙(τ),hq0,v(q0)(q1(τ),τ))𝑑τ\displaystyle\,\,v(q_{0})+\int_{0}^{\delta}L(q_{0}(\tau),\dot{q}_{0}(\tau),h_{q_{0},v(q_{0})}(q_{0}(\tau),\tau))\ d\tau+\int_{\delta}^{t}L(q_{1}(\tau),\dot{q_{1}}(\tau),h_{q_{0},v(q_{0})}(q_{1}(\tau),\tau))\ d\tau
=\displaystyle= v(q0)+0tL(q(τ),q˙(τ),hq0,v(q0)(q(τ),τ))𝑑τ,\displaystyle\,\,v(q_{0})+\int_{0}^{t}L(q(\tau),\dot{q}(\tau),h_{q_{0},v(q_{0})}(q(\tau),\tau))\ d\tau,

This shows that q:[0,t]Mq:[0,t]\rightarrow M is a minimizer of hq0,v(q0)(q,t)h_{q_{0},v(q_{0})}(q,t), thus by Proposition 2.3,

u(τ)=hq0,v(q0)(γ(τ),τ),p(τ)=Lq˙(q(τ),q˙(τ),u(τ))u(\tau)=h_{q_{0},v(q_{0})}(\gamma(\tau),\tau),\quad p(\tau)=\frac{\partial L}{\partial\dot{q}}(q(\tau),\dot{q}(\tau),u(\tau))

is a C1C^{1} characteristic starting from σ0\sigma_{0}. Invoking Lemma 2.9, we find

u(τ)={u0(τ)=U(q0(τ),τ)=U(q(τ),τ),τ[0,δ];hq1,u1(q1(τ),τδ)=U(q1(τ),τ)=U(q(τ),τ),τ[δ,t].u(\tau)=\begin{cases}u_{0}(\tau)=U(q_{0}(\tau),\tau)=U(q(\tau),\tau),\quad\tau\in[0,\delta];\\ h_{q_{1},u_{1}}(q_{1}(\tau),\tau-\delta)=U(q_{1}(\tau),\tau)=U(q(\tau),\tau),\quad\tau\in[\delta,t].\end{cases}

Remark 2.13.

The theorem shows that even for large tt, the solution of (HJe) can also be traced by characteristics starting from the 1-graph of the initial data. The main difference from the case when tt is small, treated in Theorem 2.1, is that the map πqφHt:ΛvM\pi_{q}\circ\varphi^{t}_{H}:\Lambda_{v}\rightarrow M is only a surjection rather than a diffeomorphism.

3 Connecting Legendrian graph and equilibria

Let uC(M,)u_{-}\in C^{\infty}(M,\mathbb{R}) be a classical solution to (HJs), then (H,Λ)(H,\Lambda_{-}) is a non-equilibrium thermodynamic system in the sense of Definition 1.3. This section is devoted to establishing abstract mechanisms for the existence of connecting orbits of an arbitrary Legendrian graph to the set of equilibria Λ\Lambda_{-}. These mechanisms are based on the large time behavior of solution semigroups.

3.1 Large time behavior of solution semigroups

According to their definitions, {Tt±}t0\{T^{\pm}_{t}\}_{t\geqslant 0} act on C(M,)C(M,\mathbb{R}), the space of continuous functions on MM. We shall focus on the fixed points of such actions and introduce

Definition 3.1.

A continuous function u(u_{-}\,\,( resp. u+)u_{+}) is called a fixed point of {Tt}t0(\{T^{-}_{t}\}_{t\geqslant 0}\,\,( resp. {Tt+}t0)\{T^{+}_{t}\}_{t\geqslant 0}) if

Ttu=u,(resp.Tt+u+=u+.)for anyt0.T^{-}_{t}u_{-}=u_{-},\quad(\text{resp.}\,\,T^{+}_{t}u_{+}=u_{+}.)\quad\text{for any}\,\,t\geq 0.

We use 𝒮(\mathcal{S}_{-}\,\,(resp. 𝒮+)\mathcal{S}_{+}) to denote the set of fixed points of {Tt}t0(\{T^{-}_{t}\}_{t\geqslant 0}\,\,(resp. {Tt+}t0)\{T^{+}_{t}\}_{t\geqslant 0}).

Remark 3.2.

𝒮\mathcal{S}_{-} is the analogy of weak KAM solutions [14] with uu-independent Hamiltonian.

As an easy consequence of Proposition 2.10, we have

Proposition 3.3.

u𝒮u_{-}\in\mathcal{S}_{-} if and only if uu_{-} is a solution to (HJs). Similarly, u+𝒮+u_{+}\in\mathcal{S}_{+} if and only if u+-u_{+} is a solution to the equation

H˘(q,dqu(q),u(q))=0,for any qM;\breve{H}(q,d_{q}u(q),u(q))=0,\quad\text{for any }q\in M; (3.1)

If for some initial data vC(M,)v\in C(M,\mathbb{R}), the uniform limit u:=limtTtvu_{-}:=\lim_{t\rightarrow\infty}T^{-}_{t}v exists, then for any s0s\geqslant 0, we deduce from Proposition 2.7 (3) and (5) of {Tt}t0\{T^{-}_{t}\}_{t\geqslant 0} that

Tsu=Ts(limtTtv)=limtTsTtv=limtTs+tv=u,T^{-}_{s}u_{-}=T^{-}_{s}(\lim_{t\rightarrow\infty}T^{-}_{t}v)=\lim_{t\rightarrow\infty}T^{-}_{s}\circ T^{-}_{t}v=\lim_{t\rightarrow\infty}T^{-}_{s+t}v=u_{-},

so that u𝒮u_{-}\in\mathcal{S}_{-}. Similar conclusion holds with - replaced by ++ in the above discussion. From PDE aspects, the existence of uniform limits limtTt±v\lim_{t\rightarrow\infty}T^{\pm}_{t}v is usually studied under the subject of large time behaviors.

3.2 Construction of connecting orbits I

In this part, we wish to extract a mechanism for producing semi-infinite connecting orbits between the Legendrian graph Λ0\Lambda_{0} and the states of equilibrium Λ\Lambda_{-} when uu_{-} is the uniform limit of the solution semigroup initiating from the smooth data u0u_{0}. In fact, we could obtain the following

Theorem 3.4.

Assume u0,uC(M,)u_{0},u_{-}\in C^{\infty}(M,\mathbb{R}) and (H,Λ)(H,\Lambda_{-}) is a system. If the equality

limt+Ttu0(q)=u(q)\lim_{t\rightarrow+\infty}T^{-}_{t}u_{0}(q)=u_{-}(q) (3.2)

holds uniformly for all qMq\in M, then

  1. (1)

    Λt0φHt(Λ0)¯\Lambda_{-}\subset\overline{\cup_{t\geqslant 0}\varphi^{t}_{H}(\Lambda_{0})},

  2. (2)

    there is σ0Λ0\sigma_{0}\in\Lambda_{0} such that ω(σ0)\omega(\sigma_{0}) is a nonempty subset of Λ\Lambda_{-}.

Proof.

(1)   For any σ=(q¯,dqu(q¯),u(q¯))Λ\sigma=(\bar{q},d_{q}u_{-}(\bar{q}),u_{-}(\bar{q}))\in\Lambda_{-}, by Theorem 2.12, we choose, for n1,σn=(qn,pn,un)Λ0n\geq 1,\sigma_{n}=(q_{n},p_{n},u_{n})\in\Lambda_{0} such that the corresponding characteristic segments

φHtσn=(qn(t),pn(t),un(t)),t[0,n]\varphi^{t}_{H}\sigma_{n}=(q_{n}(t),p_{n}(t),u_{n}(t)),\quad t\in[0,n]

satisfies the identity

un(t)=Ttu0(qn(t))=hqn(tτ),un(tτ)(qn(t),τ),limnqn(n)=q¯,u_{n}(t)=T^{-}_{t}u_{0}(q_{n}(t))=h_{q_{n}(t-\tau),u_{n}(t-\tau)}(q_{n}(t),\tau),\quad\lim_{n\rightarrow\infty}q_{n}(n)=\bar{q}, (3.3)

for all τ[0,t]\tau\in[0,t], where we use (2.17). Combining the above equations with the assumption (3.2),

limnun(n)=limnTnu0(qn(n))=u(q¯).\lim_{n\rightarrow\infty}u_{n}(n)=\lim_{n\rightarrow\infty}T^{-}_{n}u_{0}(q_{n}(n))=u_{-}(\bar{q}). (3.4)

Due to the uniform Lipschitz property of {Tnu0}n1\{T_{n}u_{0}\}_{n\geq 1}, the characteristic segments {φHtσn}n1\{\varphi^{t}_{H}\sigma_{n}\}_{n\geqslant 1} are uniformly bounded in Σ\Sigma, thus the sequence {pn(n)}n1\{p_{n}(n)\}_{n\geqslant 1} are relatively compact.

Claim:  limnpn(n)=dqu(q¯)\lim_{n\rightarrow\infty}p_{n}(n)=d_{q}u_{-}(\bar{q}).

Proof of the claim: We argue by contradiction to assume that for a subsequence {nj}\{n_{j}\}\subset\mathbb{N} with limjnj=+\lim_{j\rightarrow\infty}n_{j}=+\infty such that limjpnj(nj):=p¯dqu(q¯)\lim_{j\rightarrow\infty}p_{n_{j}}(n_{j}):=\bar{p}\neq d_{q}u_{-}(\bar{q}). Set σ¯=(q¯,p¯,u(q¯))\bar{\sigma}=(\bar{q},\bar{p},u_{-}(\bar{q})) and

φH1σ¯=(q1,p1,u1),φH1σ=(q1,p1,u1).\varphi^{-1}_{H}{\bar{\sigma}}=(q_{-1},p_{-1},u_{-1}),\quad\varphi^{1}_{H}\sigma=(q_{1},p_{1},u_{1}).

Due to the invariance of Λ\Lambda_{-} under φHt,u1=u(q1),p1=dqu(q1)\varphi^{t}_{H},u_{1}=u_{-}(q_{1}),p_{1}=d_{q}u_{-}(q_{1}). Since φHtσnj\varphi^{t}_{H}\sigma_{n_{j}} are characteristics, we could apply the theorem of continuous dependence of solutions on initial data to obtain limjqnj(nj1)=q1\lim_{j\rightarrow\infty}q_{n_{j}}(n_{j}-1)=q_{-1} and arguing as (3.4) to obtain u1=limjunj(nj1)=u(q1)u_{-1}=\lim_{j\rightarrow\infty}u_{n_{j}}(n_{j}-1)=u_{-}(q_{-1}). Combining the above equality and (3.3), we deduce that

u(q¯)=limjunj(nj)=limjhqnj(nj1),unj(nj1)(qnj(nj),1)=hq1,u(q1)(q¯,1),u_{-}(\bar{q})=\lim_{j\rightarrow\infty}u_{n_{j}}(n_{j})=\lim_{j\rightarrow\infty}h_{q_{n_{j}}(n_{j}-1),u_{n_{j}}(n_{j}-1)}(q_{n_{j}}(n_{j}),1)=h_{q_{-1},u_{-}(q_{-1})}(\bar{q},1),

where the last equality follows from Proposition 2.5. Now we compute as

u(q1)=T2u(q1)hq1,u(q1)(q1,2)<hq¯,hq1,u(q1)(q¯,1)(q1,1)=hq¯,u(q¯)(q1,1)=u(q1),u_{-}(q_{1})=T^{-}_{2}u_{-}(q_{1})\leq h_{q_{-1},u_{-}(q_{-1})}(q_{1},2)<h_{\bar{q},h_{q_{-1},u_{-}(q_{-1})}(\bar{q},1)}(q_{1},1)=h_{\bar{q},u_{-}(\bar{q})}(q_{1},1)=u_{-}(q_{1}),

where the first equality uses the fact that u𝒮u_{-}\in\mathcal{S}_{-} and the second (strict) inequality uses the assumption p¯dqu(q¯)\bar{p}\neq d_{q}u_{-}(\bar{q}) and the Markov property, i.e., Proposition 2.5 (2), in fact the concatenate curve constructed from the minimizers of hq1,u(q1)(q¯,1)h_{q_{-1},u_{-}(q_{-1})}(\bar{q},1) and hq¯,hq1,u(q1)(q¯,1)(q1,1)h_{\bar{q},h_{q_{-1},u_{-}(q_{-1})}(\bar{q},1)}(q_{1},1) has a corner at q¯\bar{q}, thus can not be the minimizer (must be C1C^{1}) of hq1,u(q1)(q1,2)h_{q_{-1},u_{-}(q_{-1})}(q_{1},2). The situation is depicted below and it leads to a contradiction.

[Uncaptioned image]

(2)  As in the proof of (1), we choose characteristic segments

σn(τ):=φHτσn=(qn(τ),pn(τ),un(τ)),τ[0,n]\sigma_{n}(\tau):=\varphi^{\tau}_{H}\sigma_{n}=(q_{n}(\tau),p_{n}(\tau),u_{n}(\tau)),\quad\tau\in[0,n]

with σn=(qn,pn,un)Λ0\sigma_{n}=(q_{n},p_{n},u_{n})\in\Lambda_{0} and, up to a subsequence,

limnσn=σ0=(q0,dqu0(q0),u0(q0))Λ0,un(τ)=Tτu0(qn(τ)),\lim_{n\rightarrow\infty}\sigma_{n}=\sigma_{0}=(q_{0},d_{q}u_{0}(q_{0}),u_{0}(q_{0}))\in\Lambda_{0},\quad u_{n}(\tau)=T^{-}_{\tau}u_{0}(q_{n}(\tau)), (3.5)

here we add no assumption on the convergence of qq-component. By continuous dependence of solutions of (1.2) on the initial data, σn(τ)\sigma_{n}(\tau) converges on compact intervals to a semi-infinite characteristic

σ:[0,+)Σwithσ(τ)=(q(τ),p(τ),u(τ))andσ(0)=σ0.\sigma:[0,+\infty)\rightarrow\Sigma\quad\text{with}\quad\sigma(\tau)=(q(\tau),p(\tau),u(\tau))\quad\text{and}\quad\sigma(0)=\sigma_{0}.

It follows from (3.5) and the continuity of the function Tτu0,u(τ)=Tτu0(q(τ))T^{-}_{\tau}u_{0},u(\tau)=T^{-}_{\tau}u_{0}(q(\tau)) for all τ[0,+)\tau\in[0,+\infty). Due to the uniform Lipschitz property of {Tnu0}n1\{T_{n}u_{0}\}_{n\geq 1}, the characteristics σn\sigma_{n} are uniformly bounded. This fact shows that σ\sigma is bounded and ω(σ0)\omega(\sigma_{0}) is nonempty. Now we prove that ω(σ0)Λ\omega(\sigma_{0})\subset\Lambda_{-}.

For any σ¯=(q¯,p¯,u¯)ω(σ0)\bar{\sigma}=(\bar{q},\bar{p},\bar{u})\in\omega(\sigma_{0}),

Claim 1:  u¯=u(q¯)\bar{u}=u_{-}(\bar{q}). By definition, here is non-negative sequence {tj}j1\{t_{j}\}_{j\geq 1} with limjtj=+\lim_{j\rightarrow\infty}t_{j}=+\infty and (q¯,p¯,u¯)=limj(q(tj),p(tj),u(tj))=limj(q(tj),p(tj),Ttju0(q(tj)))(\bar{q},\bar{p},\bar{u})=\lim_{j\rightarrow\infty}(q(t_{j}),p(t_{j}),u(t_{j}))=\lim_{j\rightarrow\infty}(q(t_{j}),p(t_{j}),T^{-}_{t_{j}}u_{0}(q(t_{j}))). Thus for any ϵ>0\epsilon>0, there is N1N_{1}\in\mathbb{N} such that for j>N1j>N_{1},

|u¯u(tj)|<ϵ,|u(q(tj))u(q¯)|<ϵ.|\bar{u}-u(t_{j})|<\epsilon,\quad|u_{-}(q(t_{j}))-u_{-}(\bar{q})|<\epsilon.

On the other hand, (3.2) implies that there is N2N_{2}\in\mathbb{N} such that for j>N2j>N_{2},

|Ttju0(q)u(q)|<ϵ,for allqM.|T^{-}_{t_{j}}u_{0}(q)-u_{-}(q)|<\epsilon,\quad\text{for all}\quad q\in M.

Thus for j>max{N1,N2}j>\max\{N_{1},N_{2}\},

|u¯u(q¯)|\displaystyle|\bar{u}-u_{-}(\bar{q})| |u¯u(tj)|+|u(tj)u(q(tj))|+|u(q(tj))u(q¯)|\displaystyle\,\leq|\bar{u}-u(t_{j})|+|u(t_{j})-u_{-}(q(t_{j}))|+|u_{-}(q(t_{j}))-u_{-}(\bar{q})|
|u¯u(tj)|+|Ttju0(q(tj))u(q(tj))|+|u(q(tj))u(q¯)|\displaystyle\,\leq|\bar{u}-u(t_{j})|+|T^{-}_{t_{j}}u_{0}(q(t_{j}))-u_{-}(q(t_{j}))|+|u_{-}(q(t_{j}))-u_{-}(\bar{q})|
<3ϵ,\displaystyle\,<3\epsilon,

this shows the first claim.

Claim 2:  p¯=dqu(q¯)\bar{p}=d_{q}u_{-}(\bar{q}). Notice that if σ¯ω(σ0)\bar{\sigma}\in\omega(\sigma_{0}), then for τ\tau\in\mathbb{R},

σ¯(τ)=φHτx¯=(q¯(τ),p¯(τ),u¯(τ))ω(σ0).\bar{\sigma}(\tau)=\varphi^{\tau}_{H}\bar{x}=(\bar{q}(\tau),\bar{p}(\tau),\bar{u}(\tau))\in\omega(\sigma_{0}).

By the first claim, we have u¯(τ)=u(q¯(τ))\bar{u}(\tau)=u_{-}(\bar{q}(\tau)) for every τ\tau\in\mathbb{R}. For τ\tau\in\mathbb{R}, we set

(q~(τ),p~(τ),u~(τ))=φHτ(q¯,dqu(q¯),u¯).(\tilde{q}(\tau),\tilde{p}(\tau),\tilde{u}(\tau))=\varphi^{\tau}_{H}(\bar{q},d_{q}u_{-}(\bar{q}),\bar{u}).

It follows that q¯(0)=q~(0)=q¯,u¯(0)=u~(0)=u(q¯)\bar{q}(0)=\tilde{q}(0)=\bar{q},\bar{u}(0)=\tilde{u}(0)=u_{-}(\bar{q}). Since uu_{-} is a C2C^{2} solution to (HJs),

(q~(τ),p~(τ),u~(τ))Λ.(\tilde{q}(\tau),\tilde{p}(\tau),\tilde{u}(\tau))\in\Lambda_{-}.

We argue as in (1) to assume that p¯dqu(q¯)\bar{p}\neq d_{q}u_{-}(\bar{q}). Since u𝒮u_{-}\in\mathcal{S}_{-}, then

u¯=u(q¯)=T1u(q¯)hq¯(1),u(q¯(1))(q¯,1)u¯,\bar{u}=u_{-}(\bar{q})=T^{-}_{1}u_{-}(\bar{q})\leq h_{\bar{q}(-1),u_{-}(\bar{q}(-1))}(\bar{q},1)\leq\bar{u},

where the last inequality follows from Corollary 2.4, and

u(q~(1))\displaystyle u_{-}(\tilde{q}(1)) =T2u(q~(1))hq¯(1),u(q¯(1))(q~(1),2)\displaystyle\,=T^{-}_{2}u_{-}(\tilde{q}(1))\leq h_{\bar{q}(-1),u_{-}(\bar{q}(-1))}(\tilde{q}(1),2)
<hq¯,hq¯(1),u(q¯(1))(q¯,1)(q~(1),1)=hq¯,u¯(q~(1),1)\displaystyle\,<h_{\bar{q},h_{\bar{q}(-1),u_{-}(\bar{q}(-1))}(\bar{q},1)}(\tilde{q}(1),1)=h_{\bar{q},\bar{u}}(\tilde{q}(1),1)
=hq~(0),u(q~(0))(q~(1),1)u~(1)=u(q~(1)),\displaystyle\,=h_{\tilde{q}(0),u_{-}(\tilde{q}(0))}(\tilde{q}(1),1)\leq\tilde{u}(1)=u_{-}(\tilde{q}(1)),

where the second inequality uses Markov property, i.e. Proposition 2.5 (ii), of action function and the last inequality is again a consequence of Corollary 2.4. This leads to the contradiction we desire. ∎

Remark 3.5.

The above mechanism has the advantages that it requires no more information about the local dynamics of Λ\Lambda_{-}. However, the condition (3.2) asserts the convergence of the orbit {Ttu0}\{T^{-}_{t}u_{0}\} to a fixed solution uu_{-}, which may hold only for a small part of initial data if there is no additional assumption on HH.

3.3 Construction of connecting orbits II

In this second part, we try to weaken the condition (3.2) under the local stability assumption on the set of equilibria. As is mentioned in the introduction, such assumption is widely adopted by physicists working in non-equilibrium thermodynamics.

Theorem 3.6.

Assume u0,uC(M,),(H,Λ)u_{0},u_{-}\in C^{\infty}(M,\mathbb{R}),(H,\Lambda_{-}) is a system satisfying (H1) and the limit

u(q)=lim inftTtu0(q)u_{\ast}(q)=\liminf_{t\rightarrow\infty}\,\,\,T^{-}_{t}u_{0}(q) (3.6)

exists uniformly in qMq\in M with

  1. (1)

    uuu_{\ast}\geq u_{-} on MM,

  2. (2)

    the set {qM:u(q)=u(q)}\{q\in M\,:\,u_{\ast}(q)=u_{-}(q)\} is nonempty,

then there exists σ0Λ0\sigma_{0}\in\Lambda_{0} such that ω(σ0)\omega(\sigma_{0}) is a nonempty subset of Λ\Lambda_{-}.

Proof.

Since Λ\Lambda_{-} is a local attractor, it remains to show the existence of σ0Λ0\sigma_{0}\in\Lambda_{0} such that for some T>0T>0,

φHTσ0𝒪.\varphi^{T}_{H}\sigma_{0}\in\mathcal{O}_{-}.

For each n1n\geq 1, we choose qnM,{tn}nq_{n}\in M,\{t_{n}\}_{n\in\mathbb{N}} such that limnTtnu0(q)=lim inftTtu0(q)\lim_{n\rightarrow\infty}T^{-}_{t_{n}}u_{0}(q)=\liminf_{t\rightarrow\infty}T^{-}_{t}u_{0}(q) and

Ttnu0(qn)u(qn)=minqM{Ttnu0(q)u(q)}.T^{-}_{t_{n}}u_{0}(q_{n})-u_{-}(q_{n})=\min_{q\in M}\{T^{-}_{t_{n}}u_{0}(q)-u_{-}(q)\}. (3.7)

Since Ttnu0uT^{-}_{t_{n}}u_{0}-u_{-} is a semi-concave function on MM, it is differentiable at the minimal point qnq_{n} and

dqTtnu0(qn)=dqu(qn).d_{q}T^{-}_{t_{n}}u_{0}(q_{n})=d_{q}u_{-}(q_{n}). (3.8)

We set σn:=(qn,dqTtnu0(qn),Ttnu0(qn))\sigma_{n}:=(q_{n},d_{q}T^{-}_{t_{n}}u_{0}(q_{n}),T^{-}_{t_{n}}u_{0}(q_{n})), then

Claim: Any limit point of the sequence {σn}n1\{\sigma_{n}\}_{n\geq 1} belongs to Λ\Lambda_{-}.

Up to a subsequence, we may assume limnqn=q¯\lim_{n\rightarrow\infty}q_{n}=\bar{q} and limnTtnu0(qn)\lim_{n\rightarrow\infty}T^{-}_{t_{n}}u_{0}(q_{n}) exists. Applying (3.8) and the fact that uC(M,)u_{-}\in C^{\infty}(M,\mathbb{R}),

limndqTtnu0(qn)=dqu(q¯).\lim_{n\rightarrow\infty}d_{q}T^{-}_{t_{n}}u_{0}(q_{n})=d_{q}u_{-}(\bar{q}).

By the definition of u,qnu_{\ast},q_{n} as well as the equi-Lipschitz continuity of {Ttnu0}n1\{T^{-}_{t_{n}}u_{0}\}_{n\geq 1}, for any qMq\in M,

u(q)u(q)\displaystyle\,\,\,u_{\ast}(q)-u_{-}(q)
=\displaystyle= limn[Ttnu0(q)u(q)]limn[Ttnu0(qn)u(qn)]\displaystyle\,\lim_{n\rightarrow\infty}[T^{-}_{t_{n}}u_{0}(q)-u_{-}(q)]\geq\lim_{n\rightarrow\infty}[T^{-}_{t_{n}}u_{0}(q_{n})-u_{-}(q_{n})]
=\displaystyle= limn[Ttnu0(q¯)u(q¯)]=u(q¯)u(q¯).\displaystyle\,\lim_{n\rightarrow\infty}[T^{-}_{t_{n}}u_{0}(\bar{q})-u_{-}(\bar{q})]=u_{\ast}(\bar{q})-u_{-}(\bar{q}).

By the fact that uuu_{\ast}\geq u_{-} and the set {qM:u(q)=u(q)}\{q\in M:u_{\ast}(q)=u_{-}(q)\} is nonempty, we have

u(q¯)=u(q¯),limnTtnu0(qn)=u(q¯).u_{\ast}(\bar{q})=u_{-}(\bar{q}),\quad\lim_{n\rightarrow\infty}T^{-}_{t_{n}}u_{0}(q_{n})=u_{-}(\bar{q}).

This verifies the claim. Thus for all nn sufficiently large, σn𝒪\sigma_{n}\in\mathcal{O}_{-}. To conclude this case, we fix such an σn\sigma_{n} and use Theorem 2.12 to find σn(0):=φHtnσnΛ0\sigma_{n}(0):=\varphi^{-t_{n}}_{H}\sigma_{n}\in\Lambda_{0} such that the corresponding characteristic segments φHτσn(0),τ[0,tn]\varphi^{\tau}_{H}\sigma_{n}(0),\tau\in[0,t_{n}] connects Λ0\Lambda_{0} with σn\sigma_{n}. ∎

4 Homotopic criteria for the convergence of solution semigroup

Due to the abstract mechanisms established in the last section, the problem of constructing semi-infinite orbits connecting an arbitrary Legendrian graph Λ0\Lambda_{0} to the equilibria submanifold Λ\Lambda_{-} is reduced to the study of large time behavior of the generating data u0u_{0}. The main goal of this section is to provide some criteria to ensure the required behavior of u0u_{0}. We divide this section in two parts, according to whether the local stability is assumed for the set of equilibria.

4.1 Homotopy method I

Due to the definition of sub and super-deformation defined in the introduction, one could use the tt-monotonicity of solution semigroups, indicated in Proposition 2.8, to show the following

Theorem 4.1.

Assume u0C(M,)u_{0}\in C(M,\mathbb{R}) and (H,Λ)(H,\Lambda_{-}) is a system. If there exists

  1. (a)

    a sub-deformation V:M×[0,1]V:M\times[0,1]\rightarrow\mathbb{R} of uu_{-} such that

    V(q,0)u0(q)u(q)for any qM,V(q,0)\leq u_{0}(q)\leq u_{-}(q)\quad\text{for any }q\in M,
  2. (b)

    a super-deformation V:M×[0,1]V:M\times[0,1]\rightarrow\mathbb{R} of uu_{-} such that

    V(q,0)u0(q)u(q)for any qM,V(q,0)\geq u_{0}(q)\geq u_{-}(q)\quad\text{for any }q\in M,
  3. (c)

    a sub-deformation V¯:M×[0,1]\underline{V}:M\times[0,1]\rightarrow\mathbb{R} and a super-deformation V¯:M×[0,1]\overline{V}:M\times[0,1]\rightarrow\mathbb{R} of uu_{-} such that

    V¯(,0)min{u0,u},max{u0,u}V¯(,0).\underline{V}(\cdot,0)\leqslant\min\{u_{0},u_{-}\},\quad\max\{u_{0},u_{-}\}\leqslant\overline{V}(\cdot,0).

then   limt+Ttu0=u\lim_{t\rightarrow+\infty}T^{-}_{t}u_{0}=u_{-}.

Proof.

We shall show the conclusion holds under the assumption (a) and the proof under (b) is completely similar. Due to Proposition 2.7 (1), we obtain that

TtV(q,0)Ttu0(q)Ttu(q)=u(q),for any qM.T^{-}_{t}V(q,0)\leq T^{-}_{t}u_{0}(q)\leq T^{-}_{t}u_{-}(q)=u_{-}(q),\quad\text{for any }\,\,q\in M. (4.1)

Thus to prove the conclusion, it is enough to prove

limt+TtV(,0)=u.\lim_{t\rightarrow+\infty}T^{-}_{t}V(\cdot,0)=u_{-}. (4.2)

Since V(,0)V(\cdot,0) is a subsolution to (HJs), Proposition 2.8 guarantees that for 0tτ0\leq t\leq\tau,

TtV(,0)TτV(,0).T^{-}_{t}V(\cdot,0)\leq T^{-}_{\tau}V(\cdot,0).

Combining (4.1) and the equi-Lipschitz property of {TtV(,0)}t1\{T^{-}_{t}V(\cdot,0)\}_{t\geq 1},

  • the uniform limit v=limt+TtV(,0)v_{-}=\lim_{t\rightarrow+\infty}T^{-}_{t}V(\cdot,0) exists and v𝒮v_{-}\in\mathcal{S}_{-} with V(,0)vuV(\cdot,0)\leq v_{-}\leq u_{-} on MM.

Assume vuv_{-}\neq u_{-}, then there are s0[0,1),q0Ms_{0}\in[0,1),q_{0}\in M such that

V(q,s0)v(q)for anyqM,andV(q0,s0)=v(q0)<u(q0).V(q,s_{0})\leq v_{-}(q)\quad\text{for any}\,\,q\in M,\quad\text{and}\quad V(q_{0},s_{0})=v_{-}(q_{0})<u_{-}(q_{0}).

Then by Definition 1.6, as a subsolution to (HJs), V(,s0)V(\cdot,s_{0}) is strict and applying Proposition 2.8 again,

Ttv(q0)TtV(q0,s0)>V(q0,s0)=v(q0)T^{-}_{t}v_{-}(q_{0})\geq T^{-}_{t}V(q_{0},s_{0})>V(q_{0},s_{0})=v_{-}(q_{0})

for all t>0t>0, which contradicts the fact that v𝒮v_{-}\in\mathcal{S}_{-}.

(c): Applying (a),(b) above to the assumption

V¯(,0)u0¯:=min{u0,u}uu0¯:=max{u0,u}V¯(,0),\underline{V}(\cdot,0)\leqslant\underline{u_{0}}:=\min\{u_{0},u_{-}\}\leqslant u_{-}\leqslant\overline{u_{0}}:=\max\{u_{0},u_{-}\}\leqslant\overline{V}(\cdot,0),

we arrive at

limt+Ttu0¯=u,limt+Ttu0¯=u.\lim_{t\rightarrow+\infty}T^{-}_{t}\underline{u_{0}}=u_{-},\quad\lim_{t\rightarrow+\infty}T^{-}_{t}\overline{u_{0}}=u_{-}.

Due to u0¯u0u0¯\underline{u_{0}}\leqslant u_{0}\leqslant\overline{u_{0}}, one follows that limt+Ttu0=u\lim_{t\rightarrow+\infty}T^{-}_{t}u_{0}=u_{-}. ∎

4.2 Homotopy method II

If the priori stability of the manifold of equilibria is assumed, then we could replace the existence of sub   (resp. super)-deformation in Theorem 4.1 by a more flexible condition. So we propose a second

Theorem 4.2.

Assume u0C(M,)u_{0}\in C(M,\mathbb{R}) and (H,Λ)(H,\Lambda_{-}) is a system. Assume one of the following conditions

  1. (a)(a^{\prime})

    there is a continuous function V:M×[0,1]V:M\times[0,1]\rightarrow\mathbb{R} with V(,0)V(\cdot,0) is a C2C^{2} subsolution to (HJs) and

    H(q,dqu(q),V(q,s))<0,qMandV(,0)u0u=V(,1),H(q,d_{q}u_{-}(q),V(q,s))<0,\quad\forall q\in M\quad\text{and}\quad V(\cdot,0)\leq u_{0}\leq u_{-}=V(\cdot,1),
  2. (b)(b^{\prime})

    there is a continuous function V:M×[0,1]V:M\times[0,1]\rightarrow\mathbb{R} with V(,0)V(\cdot,0) is a C2C^{2} supersolution to (HJs) and

    H(q,dqu(q),V(q,s))>0,qMandV(,0)u0u=V(,1),H(q,d_{q}u_{-}(q),V(q,s))>0,\quad\forall q\in M\quad\text{and}\quad V(\cdot,0)\geq u_{0}\geq u_{-}=V(\cdot,1),
  3. (c)(c^{\prime})

    there are continuous functions V¯,V¯:M×[0,1]\underline{V},\overline{V}:M\times[0,1]\rightarrow\mathbb{R} such that V¯(,0),V¯(,0)\underline{V}(\cdot,0),\overline{V}(\cdot,0) are a C2C^{2} subsolution and a C2C^{2} supersolution to (HJs) respectively and for all qMq\in M,

    H(q,dqu(q),V¯(q,s))<0,H(q,dqu(q),V¯(q,s))>0,\displaystyle H(q,d_{q}u_{-}(q),\underline{V}(q,s))<0,\quad\quad H(q,d_{q}u_{-}(q),\overline{V}(q,s))>0,
    V¯(q,0)min{u0(q),u(q)},max{u0(q),u(q)}V¯(q,0).\displaystyle\underline{V}(q,0)\leqslant\min\{u_{0}(q),u_{-}(q)\},\quad\max\{u_{0}(q),u_{-}(q)\}\leqslant\overline{V}(q,0).

is satisfied, then the uniform limit u(q)=lim inftTtu0(q)u_{\ast}(q)=\liminf_{t\rightarrow\infty}T^{-}_{t}u_{0}(q) exists and the set

{qM:u(q)=u(q)}is nonempty.\{q\in M\,:\,u_{\ast}(q)=u_{-}(q)\}\quad\text{is nonempty.}
Proof.

(a)(a^{\prime}): Since V(,0)u0uV(\cdot,0)\leq u_{0}\leq u_{-} on MM, due to Proposition 2.7 (1), we have

TtV(q,0)Ttu0(q)Ttu(q)=u(q),for any qM.T^{-}_{t}V(q,0)\leq T^{-}_{t}u_{0}(q)\leq T^{-}_{t}u_{-}(q)=u_{-}(q),\quad\text{for any }\,\,q\in M. (4.3)

Using the assumption that V(,0)V(\cdot,0) is a subsolution to (HJs), Proposition 2.8 implies that for 0tτ0\leq t\leq\tau,

TtV(,0)TτV(,0).T^{-}_{t}V(\cdot,0)\leq T^{-}_{\tau}V(\cdot,0). (4.4)

Combining (4.3) and the equi-Lipschitz property of {TtV(,0)}t1\{T^{-}_{t}V(\cdot,0)\}_{t\geq 1}, the uniform limit

v:=limt+TtV(,0)v_{\ast}:=\lim_{t\rightarrow+\infty}T^{-}_{t}V(\cdot,0)

exists with v𝒮v_{\ast}\in\mathcal{S}_{-} and

V(,0)vuonM.V(\cdot,0)\leq v_{\ast}\leq u_{-}\quad\text{on}\,\,M.

We now show that v=uv_{\ast}=u_{-} by contradiction. Assume there is q0Mq_{0}\in M such that

u(q0)v(q0)=maxqM[u(q0)v(q0)]>0.u_{-}(q_{0})-v_{\ast}(q_{0})=\max_{q\in M}[u_{-}(q_{0})-v_{\ast}(q_{0})]>0. (4.5)

By the continuity of the function VV, there are s0[0,1),δ>0s_{0}\in[0,1),\delta>0 such that V(q0,s0)=v(q0)V(q_{0},s_{0})=v_{\ast}(q_{0}) and

supq˙Tq0M[dqu(q0),q˙L(q0,q˙,v(q0))]\displaystyle\,\sup_{\dot{q}\in T_{q_{0}}M}[\langle d_{q}u_{-}(q_{0}),\dot{q}\rangle-L(q_{0},\dot{q},v_{\ast}(q_{0}))]
=\displaystyle= H(q0,dqu(q0),v(q0))=H(q0,dqu(q0),V(q0,s0))=δ<0.\displaystyle\,H(q_{0},d_{q}u_{-}(q_{0}),v_{\ast}(q_{0}))=H(q_{0},d_{q}u_{-}(q_{0}),V(q_{0},s_{0}))=-\delta<0.

By Proposition 2.3 and Definition 2.6, there is a γC1([1,0],M)\gamma\in C^{1}([-1,0],M) with γ(0)=q0\gamma(0)=q_{0} such that

Ttv(q0)=hγ(t),v(γ(t))(q0,t)=v(γ(t))+t0L(γ(τ),γ˙(τ),hγ(t),v(γ(t))(γ(τ),τ))𝑑τ.T^{-}_{t}v_{\ast}(q_{0})=h_{\gamma(-t),v_{\ast}(\gamma(-t))}(q_{0},t)=v_{\ast}(\gamma(-t))+\int_{-t}^{0}L(\gamma(\tau),\dot{\gamma}(\tau),h_{\gamma(-t),v_{\ast}(\gamma(-t))}(\gamma(\tau),\tau))\ d\tau.

holds for all t[0,1]t\in[0,1]. Due to the fact that v𝒮v_{\ast}\in\mathcal{S}_{-},  for t>0t>0 sufficiently small,

u(q0)v(q0)=u(q0)Ttv(q0)\displaystyle\,u_{-}(q_{0})-v_{\ast}(q_{0})=u_{-}(q_{0})-T^{-}_{t}v_{\ast}(q_{0})
=\displaystyle= [u(γ(t))v(γ(t))]+[u(γ(0))u(γ(t))]t0L(γ(τ),γ˙(τ),hγ(t),v(γ(t))(γ(τ),τ)dτ\displaystyle\,[u_{-}(\gamma(-t))-v_{\ast}(\gamma(-t))]+[u_{-}(\gamma(0))-u_{-}(\gamma(-t))]-\int_{-t}^{0}L(\gamma(\tau),\dot{\gamma}(\tau),h_{\gamma(t),v_{\ast}(\gamma(t))}(\gamma(\tau),\tau)\ d\tau
=\displaystyle= [u(γ(t))v(γ(t))]+t[dqu(q0),γ˙(0)L(q0,γ˙(0),v(q0))]+o(t)\displaystyle\,[u_{-}(\gamma(-t))-v_{\ast}(\gamma(-t))]+t\,[\langle d_{q}u_{-}(q_{0}),\dot{\gamma}(0)\rangle-L(q_{0},\dot{\gamma}(0),v_{\ast}(q_{0}))]+o(t)
\displaystyle\leq [u(γ(t))v(γ(t))]δt+o(t)\displaystyle\,[u_{-}(\gamma(-t))-v_{\ast}(\gamma(-t))]-\delta t+o(t)
<\displaystyle< u(γ(t))v(γ(t))u(q0)v(q0),\displaystyle\,u_{-}(\gamma(-t))-v_{\ast}(\gamma(-t))\leq u_{-}(q_{0})-v_{\ast}(q_{0}),

where the last inequality uses (4.5). This leads to the contradiction.

To complete the proof, we use (4.3) to obtain

u=vulim suptTtu0u,u_{-}=v_{\ast}\leqslant u_{\ast}\leqslant\limsup_{t\rightarrow\infty}T^{-}_{t}u_{0}\leqslant u_{-},

and it follows that u(q)=v(q)=u(q)u_{\ast}(q)=v_{\ast}(q)=u_{-}(q) for any qMq\in M.

(b)(b^{\prime}):  By reversing the signs in the inequalities (4.3)-(4.4) and arguing with the same reasoning as in the beginning of the above proof, the uniform limit v:=limt+TtV(,0)v_{\ast}:=\lim_{t\rightarrow+\infty}T^{-}_{t}V(\cdot,0) exists with v𝒮v_{\ast}\in\mathcal{S}_{-} and

V(,0)vuonM.V(\cdot,0)\geq v_{\ast}\geq u_{-}\quad\text{on}\,\,M.

It follows from the fact that V(0,)u0V(0,\cdot)\geqslant u_{0} and Proposition 2.7 (1) that

vlim suptTtu0lim inftTtu0=uuv_{\ast}\geqslant\limsup_{t\rightarrow\infty}T^{-}_{t}u_{0}\geqslant\liminf_{t\rightarrow\infty}T^{-}_{t}u_{0}=u_{\ast}\geqslant u_{-}

To complete the proof, it is enough show that the set {qM:v(q)=u(q)}\{q\in M\,:\,v_{\ast}(q)=u_{-}(q)\} is nonempty, we argue by contradiction. Assume v>uv_{\ast}>u_{-} on MM and there is q0Mq_{0}\in M with

v(q0)u(q0)=minqM[v(q)u(q)]>0.v_{\ast}(q_{0})-u_{-}(q_{0})=\min_{q\in M}\,\,[v_{\ast}(q)-u_{-}(q)]>0. (4.6)

By the continuity of the function VV, there is s0[0,1)s_{0}\in[0,1) such that V(q0,s0)=v(q0)V(q_{0},s_{0})=v_{\ast}(q_{0}) and

supq˙Tq0M[dqu(q0),q˙L(q0,q˙,v(q0))]=H(q0,dqu(q0),v(q0))=H(q0,dqu(q0),V(q0,s0))>0.\sup_{\dot{q}\in T_{q_{0}}M}[\langle d_{q}u_{-}(q_{0}),\dot{q}\rangle-L(q_{0},\dot{q},v_{\ast}(q_{0}))]=H(q_{0},d_{q}u_{-}(q_{0}),v_{\ast}(q_{0}))=H(q_{0},d_{q}u_{-}(q_{0}),V(q_{0},s_{0}))>0.

Thus there is q˙0Tq0M\dot{q}_{0}\in T_{q_{0}}M such that

δ:=dqu(q0),q˙0L(q0,q˙0,v(q0))>0.\delta:=\langle d_{q}u_{-}(q_{0}),\dot{q}_{0}\rangle-L(q_{0},\dot{q}_{0},v_{\ast}(q_{0}))>0.

Setting (q(t),p(t),u(t))=φHt(q0,p0,v(q0))(q(t),p(t),u(t))=\varphi^{t}_{H}(q_{0},p_{0},v_{\ast}(q_{0})), where p0=Lq˙(q0,q˙0,v(q0))p_{0}=\frac{\partial L}{\partial\dot{q}}(q_{0},\dot{q}_{0},v_{\ast}(q_{0})), then for t>0t>0 sufficiently small, we have

u(t)u(q(t))=[u(0)+tu˙(0)+o(t)][u(q(0))+tdqu(q(0)),q˙(0)+o(t)]\displaystyle\,u(t)-u_{-}(q(t))=[u(0)+t\dot{u}(0)+o(t)]-[u_{-}(q(0))+t\langle d_{q}u_{-}(q(0)),\dot{q}(0)\rangle+o(t)]
=\displaystyle= [v(q0)u(q0)][dqu(q0),q˙0L(q0,q˙0,v(q0))]t+o(t)\displaystyle\,[v_{\ast}(q_{0})-u_{-}(q_{0})]-[\langle d_{q}u_{-}(q_{0}),\dot{q}_{0}\rangle-L(q_{0},\dot{q}_{0},v_{\ast}(q_{0}))]t+o(t)
=\displaystyle= [v(q0)u(q0)]δt+o(t),\displaystyle\,[v_{\ast}(q_{0})-u_{-}(q_{0})]-\delta t+o(t),

where for the second equation, the charactersitic system (1.2) as well as the definition of p0p_{0} are used. Since v𝒮v_{\ast}\in\mathcal{S}_{-}, it follows from Definition 2.6 and Corollary 2.4 that

v(q(t))u(q(t))=Ttv(q(t))u(q(t))\displaystyle\,v_{\ast}(q(t))-u_{-}(q(t))=T^{-}_{t}v_{\ast}(q(t))-u_{-}(q(t))
\displaystyle\leq hq0,v(q0)(q(t),t)u(q(t))u(t)u(q(t))\displaystyle\,h_{q_{0},v_{\ast}(q_{0})}(q(t),t)-u_{-}(q(t))\leq u(t)-u_{-}(q(t))
=\displaystyle= [v(q0)u(q0)]δt+o(t)\displaystyle\,[v_{\ast}(q_{0})-u_{-}(q_{0})]-\delta t+o(t)
<\displaystyle< v(q0)u(q0),\displaystyle\,v_{\ast}(q_{0})-u_{-}(q_{0}),

which contradicts (4.6).

(c)(c^{\prime}): The proof is completely similar to that of Theorem 4.1 (c). ∎

5 Sample examples and applications of the main theorems

This section includes direct consequences of the connecting mechanisms as well as interpretations of them on certain classes of contact systems. As concluding remarks, we shall discuss the relationship of our results with those obtained in the paper [12].

5.1 Monotone Hamiltonian

Let us begin with the class of contact Hamiltonian HH that are strictly increasing in uu, i.e.,

  • (M)

    Hu(q,p,u)>0\frac{\partial H}{\partial u}(q,p,u)>0 for every (q,p,u)Σ(q,p,u)\in\Sigma.

These Hamiltonian could be seen as a generalization of discounted Hamiltonian and the corresponding systems model the motions of particles in mechanical systems with friction. For a contact Hamiltonian system defined by monotone Hamiltonian, we could obtain the following conclusion which relates to the fact that Λ\Lambda_{-} is part of the maximal attractor for φHt\varphi^{t}_{H} on Σ\Sigma.

Corollary 5.1.

Assume HH satisfies (H2) and (M) and there is a CC^{\infty} function u:Mu_{-}:M\rightarrow\mathbb{R} such that (H,Λ)(H,\Lambda_{-}) constitutes a system. Then for every CC^{\infty} function u0u_{0}, there is σ0Λ0\sigma_{0}\in\Lambda_{0} such that ω(σ0)Λ\omega(\sigma_{0})\subset\Lambda_{-}.

Proof.

Due to compactness of MM, there is c>0c>0 such that for all qMq\in M,

u(q)cu0(q)u(q)+c.u_{-}(q)-c\leq u_{0}(q)\leq u_{-}(q)+c.

Notice that V(q,t)(V_{-}(q,t)( resp.V+(q,t))V_{+}(q,t)) :M×[0,1]:M\times[0,1]\rightarrow\mathbb{R} defined by

V(q,t)=u(q)(1t)c,(resp.V+(q,t)=u(q)+(1t)c)V_{-}(q,t)=u_{-}(q)-(1-t)c,\quad(\text{resp.}\,\,V_{+}(q,t)=u_{-}(q)+(1-t)c)

are a sub (resp. super)-deformation of uu_{-} respectively with V±(q,0)=u(q)±cV_{\pm}(q,0)=u_{-}(q)\pm c. Then by the monotonicity of solution semigroup, we have for t0t\geq 0,

TtV(q,0)Ttu0(q)TtV+(q,0)T^{-}_{t}V_{-}(q,0)\leq T^{-}_{t}u_{0}(q)\leq T^{-}_{t}V_{+}(q,0)

By Theorem 4.1, we have

limtTtV±(q,0)=u(q)and solimtTtu0(q)=u(q).\lim_{t\rightarrow\infty}T^{-}_{t}V_{\pm}(q,0)=u_{-}(q)\quad\text{and so}\quad\lim_{t\rightarrow\infty}T^{-}_{t}u_{0}(q)=u_{-}(q).

Now we apply Theorem 3.4 to complete the proof. ∎

5.2 Contact Möbius model

In this part, we apply our connecting mechanism to give an analysis of an interesting model raised in [12].

Example 5.2.

[12, Example 2.12] Let M=𝕊1M=\mathbb{S}^{1} and Σ=J1M\Sigma=J^{1}M the corresponding phase space with the canonical contact structure defined in the introduction. The Hamiltonian F:ΣF_{\mathcal{M}}:\Sigma\rightarrow\mathbb{R} defined by

F(q,p,u)=p2+u21F_{\mathcal{M}}(q,p,u)=p^{2}+u^{2}-1

induced an integrable contact Hamiltonian flow on Σ\Sigma. Precisely, since the Hamiltonian is independent of qq, the contact Hamiltonian vector field (1.2) can be projected to the (p,u)(p,u)-plane as

{p˙=2pu,u˙=p2u2+1.\begin{cases}\dot{p}=-2pu,\\ \dot{u}=p^{2}-u^{2}+1.\end{cases}

If one introduce the complex coordinate z=u+1pz=u+\sqrt{-1}\,p on the (p,u)(p,u)-plane with the real cylinder l={p=0}l=\{p=0\}, the flow defined by (5.1) is described as the one-parameter subgroup of the Möbius transformations PSL(2,)PSL(2,\mathbb{R}) admitting an unstable fixed point at w=1w=-1 and a stable point at w=1w=1. In the complex coordinates, the solutions reads as

w(t)=w(0)cosht+sinhtw(0)sinht+cosht,w(0)=u(0)+1p(0).w(t)=\frac{w(0)\cosh t+\sinh t}{w(0)\sinh t+\cosh t},\quad w(0)=u(0)+\sqrt{-1}\,p(0). (5.1)

From the above formula, one could see that the phase flow of (5.1) is incomplete. For ϵ>0\epsilon>0, we choose a cut-off function a:[0,+)[1,+)a:[0,+\infty)\rightarrow[-1,+\infty) with a(s)>0a^{\prime}(s)>0 for all ss and

limsa(s)=a>1,a(s)=s1,fors[0,1+ϵ],\lim_{s\rightarrow\infty}a(s)=a_{\infty}>1,\quad a(s)=s-1,\quad\text{for}\,\,s\in[0,1+\epsilon],

to construct a new Hamiltonian H(q,p,u)=a(p2+u2)H_{\mathcal{M}}(q,p,u)=a(p^{2}+u^{2}) to make the flow complete with the dynamics in the disk {p2+u2<1+ϵ}\{p^{2}+u^{2}<1+\epsilon\} unchanged. The authors focus on the following fact since it contains some ingredients for the mechanism of their constructions [12, Theorem 2.9].

  • ()(\clubsuit)

    Along the real cylinder l={p=0}l=\{p=0\}, for cc\in\mathbb{R}, if we define Legendrian graphs

    Λ={(q,0,1):q𝕊1},Λc:={(q,0,c):q𝕊1},\Lambda_{-}=\{(q,0,1)\,:\,q\in\mathbb{S}^{1}\},\quad\Lambda_{c}:=\{(q,0,c)\,:\,q\in\mathbb{S}^{1}\},

    then Λ\Lambda_{-} is a local attractor for φHt\varphi^{t}_{H} and Λc\Lambda_{c} admits trajectories of the contact Hamiltonian flow starting on KcK_{c} and converge asymptotically to Λ\Lambda_{-} for c>1c>1 but not for c<1c<-1.

The author have shown that the Legendrian submanifolds (Λc,Λ)(\Lambda_{c},\Lambda_{-}) is interlinked for c>1c>1 but not for c<1c<-1. This fact explains why the connecting mechanism [12, Theorem 2.9] works only for c>1c>1.

To give an interpretation of the fact ()(\clubsuit) from our viewpoint, we shall focus on the dynamics in a neighborhood of the unit disk on (p,u)(p,u)-plane. In this region, H=F=u2+p21H_{\mathcal{M}}=F_{\mathcal{M}}=u^{2}+p^{2}-1 satisfying (H2). According to description of ()(\clubsuit), we divide the analysis into two cases: for an initial data u0C(𝕊1,)u_{0}\in C^{\infty}(\mathbb{S}^{1},\mathbb{R}) with

  1. 1.

    u0>1u_{0}>-1 (not necessarily constant) and Λ0{p2+u2<1+ϵ}\Lambda_{0}\subset\{p^{2}+u^{2}<1+\epsilon\}, then

    1<U¯:=minqM{u0(q),u(q)}u0U¯:=maxq{u0(q),u(q)}1+ϵ.-1<\underline{U}:=\min_{q\in M}\{u_{0}(q),u_{-}(q)\}\leqslant u_{0}\leqslant\overline{U}:=\max_{q}\{u_{0}(q),u_{-}(q)\}\leqslant\sqrt{1+\epsilon}. (5.2)

    Then one easily constructs

    • a sub-deformation V¯:M×[0,1],V¯(q,λ)=(1λ)U¯+λ,\underline{V}:M\times[0,1]\rightarrow\mathbb{R},\quad\underline{V}(q,\lambda)=(1-\lambda)\underline{U}+\lambda,

    • a super-deformation V¯:M×[0,1],V¯(q,λ)=(1λ)U¯+λ.\overline{V}:M\times[0,1]\rightarrow\mathbb{R},\quad\overline{V}(q,\lambda)=(1-\lambda)\overline{U}+\lambda.

    Now we apply Theorem 1.7 (c) to get

    1. (1)

      Λt0φHt(Λ0)¯\Lambda_{-}\subset\overline{\cup_{t\geqslant 0}\varphi^{t}_{H}(\Lambda_{0})},

    2. (2)

      and there is σ0Λ0\sigma_{0}\in\Lambda_{0} such that ω(σ0)Λ\omega(\sigma_{0})\subset\Lambda_{-}.

  2. 2.

    u0c<1u_{0}\equiv c<-1 and Λ0{p2+u2<1+ϵ}\Lambda_{0}\subset\{p^{2}+u^{2}<1+\epsilon\}, then u0<uu_{0}<u_{-} and any CC^{\infty} function V:M×[0,1]V:M\times[0,1]\rightarrow\mathbb{R} with V(,0)u0V(\cdot,0)\leqslant u_{0} satisfies

    H|ΛV(,0)>0,H(,0,V(,0))>0.H_{\mathcal{M}}|_{\Lambda_{V(\cdot,0)}}>0,\quad H_{\mathcal{M}}(\cdot,0,V(\cdot,0))>0.

    From these facts, it is easily deduced that the deformations listed in the conditions (a)(c),(a)(c)(a)-(c),(a^{\prime})-(c^{\prime}) of Theorem 1.7-1.9 do not exist, thus showing the necessity of these conditions.

By studying the phase portrait of the system defined by HH_{\mathcal{M}}, we found that if the initial data u0<1u_{0}<-1 but is not constant on 𝕊1\mathbb{S}^{1}, there is an semi-infinite orbit initiating from Λ0\Lambda_{0} and converge asymptotically to Λ\Lambda_{-}. This phenomenon is detected neither by the mechanisms formulated in [12], since in this case (Λ0,Λ)(\Lambda_{0},\Lambda_{-}) is not interlinked, nor by our results. So it is natural to ask

Question 5.3.

Is there an abstract mechanism, for some suitable setting including Example 5.2 as a special case, for the existence of such orbits?

Acknowledgments

All of the author are supported in part by the National Natural Science Foundation of China (Grant No. 12171096). L. Jin is also supported in part by the NSFC (Grant No. 11901293, 11971232). J. Yan is also supported in part by the NSFC (Grant No. 11790272). The first author would like to thank Dr. S.Suhr for his kind invitation and RUB (Ruhr-Universität Bochum) for its hospitality, where part of this work is done.

References

  • [1] V.I.Arnold: Contact geometry: the geometrical method of Gibbs’s thermodynamics. Proceedings of the Gibbs Symposium (New Haven, CT, 1989), 163-179, Amer. Math. Soc., Providence, RI, (1990).
  • [2] V.I.Arnold: Lectures on partial differential equations. Translated from the second Russian edition by Roger Cooke. Universitext. Springer-Verlag, Berlin; Publishing House PHASIS, Moscow, x+157 pp, (2004).
  • [3] Bravetti, A., Lopez-Monsalvo, C. S., Nettel, F.; Contact symmetries and Hamiltonian thermodynamics, Ann. of Physics 361 (2015), 377-400.
  • [4] P.Cannarsa, C.Sinestrari: Semiconcave functions, Hamilton-Jacobi equations, and optimal control. Progress in Nonlinear Differential Equations and their Applications, 58. Birkhäuser Boston, Inc., Boston, MA, xiv+304 pp, (2004).
  • [5] P. Cannarsa, W. Cheng, K. Wang and J. Yan, Herglotz’ generalized variational principle and contact type Hamilton-Jacobi equations, Trends in Control Theory and Partial Differential Equations, 39–67. Springer INdAM Ser., 32, Springer, Cham, 2019.
  • [6] P. Cannarsa, W. Cheng, L. Jin, K. Wang and J. Yan, Herglotz’ variational principle and Lax-Oleinik evolution, J. Math. Pures Appl. 141 (2020), 99–136.
  • [7] M.Crandall,P.L.Lions: Viscosity solutions of Hamilton-Jacobi equations. Trans. Amer. Math. Soc. 277 ,(1983) 1–42.
  • [8] M-O.Czarnecki,L.Rifford: Approximation and regularization of Lipschitz functions: convergence of the gradients. Trans. Amer. Math. Soc. 358 (2006), 4467-4520.
  • [9] A.Davini, A.Fathi, R.Iturriaga, M. Zavidovique: Convergence of the solutions of the discounted Hamilton-Jacobi equation: convergence of the discounted solutions. Invent. Math. 206 (2016),29-55.
  • [10] L.C.Evans: Partial differential equations. Second edition. Graduate Studies in Mathematics, 19. American Mathematical Society, Providence, RI, xxii+749 pp, (2010).
  • [11] M.Entov, L. Polterovich:Leonid Legendrian persistence modules and dynamics. J. Fixed Point Theory Appl. 24 (2022), no. 2, Paper No. 30, 54 pp. 53D42 (37J55)
  • [12] M.Entov, L.Polterovich: Contact topology and non-equilibrium thermodynamics. arXiv:2101.03770.
  • [13] M.Entov, L.Polterovich: Lagrangian tetragons and instabilities in Hamiltonian dynamics, Nonlinearity 30 (2017), 13-34.
  • [14] A.Fathi: Weak KAM Theorem in Lagrangian Dynamics. preliminary version 10, Lyon, unpublished (2008).
  • [15] A. Ghosh, B. Chandrasekhar. Contact geometry and thermodynamics of black holes in AdS spacetimes, Physical Review D 100, 126020 (2019).
  • [16] P. L. Lions, G. Papanicolaou, S. Varadhan. Homogenization of Hamilton-Jacobi equations, Unpublished Work. (1987).
  • [17] J.N.Mather: Action minimizing invariant measures for positive definite Lagrangian systems. Math. Z. 207, 169-207 (1991)
  • [18] J.N.Mather: Variational construction of connecting orbits. Ann. Inst. Fourier (Grenoble) 43 (1993), no. 5, 1349-1386.
  • [19] M. Stefano, A. Sorrentino:Aubry-Mather theory for conformally symplectic systems. Communications in Mathematical Physics 354 (2017), 775-808.
  • [20] P.R.Ni, L.Wang: A nonlinear semigroup approach to Hamilton-Jacobi equations-revisited, arxiv: 2202.11315v2.
  • [21] I.Prigogine: Time, structure, and fluctuations, Science 201: 4358 (1978), 777-785.
  • [22] S. G. Rajeev; A Hamilton-Jacobi formalism for thermodynamics. Annals of Physics 323 (2008), 2265-2285.
  • [23] S. Goto:Nonequilibrium thermodynamic process with hysteresis and metastable states-A contact Hamiltonian with unstable and stable segments of a Legendre submanifold. J. Math. Physics 63,053302 (2022) .
  • [24] K.Wang, L.Wang and J.Yan: Implicit variational principle for contact Hamiltonian systems. Nonlinearity 30 (2017), 492–515.
  • [25] K.Wang, L.Wang and J.Yan: Variational principle for contact Hamiltonian systems and its applications. J. Math. Pures Appl. 123 (2019), 167–200.
  • [26] K.Wang, L.Wang and J.Yan: Aubry-Mather theory for contact Hamiltonian systems. Commun. Math. Phys. 366 (2019), 981–1023.
  • [27] V.Zorich: Mathematical analysis of problems in the natural sciences. Translated from the 2008 Russian original by Gerald Gould. Springer, Heidelberg, xii+135 pp, (2008).