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

A geometric approach to Mather quotient problem

Wei Cheng Wei Cheng, School of Mathematics, Nanjing University, 22 Hankou Road, Nanjing, 210093, Jiangsu, China. [email protected]  and  Wenxue Wei Wenxue Wei, School of Mathematics, Nanjing University, 22 Hankou Road, Nanjing, 210093, Jiangsu, China. [email protected]
Abstract.

Let (M,g)(M,g) be a closed, connected and orientable Riemannian manifold with nonnegative Ricci curvature. Consider a Lagrangian L(x,v):TML(x,v):TM\to\mathbb{R} defined by L(x,v):=12gx(v,v)ω(v)+cL(x,v):=\frac{1}{2}g_{x}(v,v)-\omega(v)+c, where cc\in\mathbb{R} and ω\omega is a closed 1-form. From the perspective of differential geometry, we estimate the Laplacian of the weak KAM solution uu to the associated Hamilton-Jacobi equation H(x,du)=c[L]H(x,du)=c[L] in the barrier sense. This analysis enables us to prove that each weak KAM solution uu is constant if and only if ω\omega is a harmonic 1-form. Furthermore, we explore several applications to the Mather quotient and Mañé’s Lagrangian.

Key words and phrases:
Aubry-Mather theory, Mather quotient, Riccati equation, Harmonic one-form
2010 Mathematics Subject Classification:
35F21, 49L25, 37J50

1. Introduction

This paper focuses on a significant problem in Aubry-Mather theory originally posed by John Mather, concerning the Mather quotient. The Aubry-Mather theory employs variational methods to study Hamiltonian dynamical systems. Mather developed a theory to study the dynamics of the associated Euler-Lagrangian flow in frame of Tonelli theory in calculus of variations, by introduced certain invariant sets of the global Lagrangian dynamical systems such as Aubry set, Mather set, etc. (as detailed in [19, 20] and further elaborated in [18]).

1.1. Mather quotient

Concentrating on time-independent case, we suppose MM is a closed and connected smooth manifold with TMTM and TMT^{*}M its tangent and cotangent bundle respectively. A function L(x,v):TML(x,v):TM\to\mathbb{R} is called a Tonelli Lagrangian if LL is of class CrC^{r} (r3r\geqslant 3) and L(x,)L(x,\cdot) is strictly convex and uniformly superlinear on TxMT_{x}M for all xMx\in M. The Tonelli Hamiltonian H:TMH:T^{*}M\to\mathbb{R} associated to a Tonelli Lagrangian LL is defined by H(x,p)=supvTxM{p(v)L(x,v)}H(x,p)=\sup_{v\in T_{x}M}\{p(v)-L(x,v)\}, (x,p)TM(x,p)\in T^{*}M. In [20] Mather introduced the Peierls’ barrier function h:M×Mh:M\times M\to\mathbb{R}, h(x,y)=lim inft+{At(x,y)+c[L]t}h(x,y)=\liminf_{t\to+\infty}\{A_{t}(x,y)+c[L]t\}, x,yMx,y\in M. Here At(x,y)=infρ0tL(ρ,ρ˙)𝑑sA_{t}(x,y)=\inf_{\rho}\int^{t}_{0}L(\rho,\dot{\rho})\ ds where the infimum is taken over the family of absolutely continuous curve ρ:[0,t]M\rho:[0,t]\to M connecting x=ρ(0)x=\rho(0) to y=ρ(t)y=\rho(t), and c[L]c[L]\in\mathbb{R}, the Mañé’s critical value, is the unique constant such that hh is finite-valued (see [17]). The projected Aubry set is defined by 𝒜(L)={xM:h(x,x)=0}\mathcal{A}(L)=\{x\in M:h(x,x)=0\}. In [20], Mather also introduced a pseudo-metric δ\delta on 𝒜(L)\mathcal{A}(L) by

δ(x,y)=h(x,y)+h(y,x).\displaystyle\delta(x,y)=h(x,y)+h(y,x).

The relation xyδ(x,y)=0x\sim y\Leftrightarrow\delta(x,y)=0 gives an equivalence relation on 𝒜(L)\mathcal{A}(L). The associated quotient space (𝒜(L),,δ)(\mathcal{A}(L),\sim,\delta) is the so-called Mather quotient.

In [21], Mather showed that if MM has dimension 22 or if the Lagrangian is the kinetic energy associated to a Riemannian metric on MM with dimM3\dim M\leqslant 3, then the Mather quotient is totally disconnected, i.e. every connected component consists of a single point. Unfortunately, this does not hold in higher dimensions (see [4, 22]). The totally disconnectedness of Mather quotient is closely related to the upper semi-continuity of the Aubry set and has been studied in several earlier works such as [1, 16, 21, 26]. Those works consider this problem from either topological or variational points of view.

Certain Morse-Sard type results on this problem can be found in [16]. The authors proved

Theorem 1.1.

Let LL be a Tonelli Lagrangian on a closed smooth manifold MM. Then, it satisfies the Mather disconnectedness condition (i.e. for every pair u1,u2u_{1},u_{2} of weak KAM solutions, the image (u1u2)(𝒜(L))(u_{1}-u_{2})(\mathcal{A}(L))\subset\mathbb{R} is totally disconnected) in the following five cases:

  1. (i)

    The dimension of MM is 1 or 2 .

  2. (ii)

    The dimension of MM is 3, and 𝒜~(L)\tilde{\mathcal{A}}(L), the Aubry set111In the context of weak KAM theory, 𝒜~(L)=u{(x,v)TM:dxu=Lv(x,v)}TM\tilde{\mathcal{A}}(L)=\bigcap_{u}\{(x,v)\in TM:d_{x}u=L_{v}(x,v)\}\subset TM with uu taken over all C1C^{1} subsolution of (HJ), contains no fixed point of the associated Euler-Lagrange flow ΦtL\Phi_{t}^{L} (which is defined in Section 2).

  3. (iii)

    The dimension of MM is 3, and LL is of class C3,1C^{3,1}.

  4. (iv)

    The Lagrangian is of class Ck,1C^{k,1}, with k2dimM3k\geqslant 2\dim M-3, and every point of 𝒜~(L)\tilde{\mathcal{A}}(L) is fixed under the Euler-Lagrange flow ΦtL\Phi_{t}^{L} .

  5. (v)

    The Lagrangian is of class Ck,1C^{k,1}, with k8dimM8k\geqslant 8\dim M-8, and either each point of 𝒜~(L)\tilde{\mathcal{A}}(L), is fixed under the Euler-Lagrange flow ΦtL\Phi_{t}^{L} or its orbit in the 𝒜~(L)\tilde{\mathcal{A}}(L) is periodic with strictly positive period.

In [1], Bernard listed several so-called coincidence hypothesis and shown that the Mather disconnectedness condition implies the property that the Mather quotient is totally disconnected. He also obtained the upper semi-continuity of the Aubry set under these conditions.

In [26] the author proved

Theorem 1.2.

Let MM be a closed connected smooth manifold with dimension n1n\geqslant 1 and let LL be a Tonelli Lagrangian such that

ΛL:={(x,Lv(x,0)):xM}\displaystyle\Lambda_{L}:=\{(x,L_{v}(x,0)):x\in M\}

is a Lagrangian submanifold of TMT^{*}M and L(x,0)Cr(M)L(x,0)\in C^{r}(M), with r2n2r\geqslant 2n-2 and Lv(x,0)C2(M)L_{v}(x,0)\in C^{2}(M). Then, for every [ω][\omega] in the Liouville class of ΛL\Lambda_{L} and Lω(x,v):=L(x,v)ω(v)L_{\omega}(x,v):=L(x,v)-\omega(v), the Mather quotient (𝒜(Lω),,δ)(\mathcal{A}(L_{\omega}),\sim,\delta) is totally disconnected.

From late 1990’s, Albert Fathi developed celebrated weak KAM theory which serves as a conceptual bridge between the Aubry-Mather theory and the realm of partial differential equations (PDEs). Fathi proved there exists a unique constant c[L]c[L], exactly the Mañé’s critical value, such that the Hamilton-Jacobi equation

H(x,dxu)=c[L],xMH(x,d_{x}u)=c[L],\qquad x\in M (HJ)

admits a weak solution uu which is a common fixed point of the Lax-Oleinik semigroup Tt+c[L]tT^{-}_{t}+c[L]t for t0t\geqslant 0 (See more details in Section 2). Such weak solutions are called weak KAM solutions. Weak KAM theory enables the application of PDEs and tools from differential geometry.

If XX is a CkC^{k} vector field on a Riemannian manifold (M,g)(M,g) with k2k\geqslant 2, introduced by Ricardo Mañé in [18], the Mañé Lagrangian LX:TML_{X}:TM\to\mathbb{R} associated to XX is defined by

LX(x,v)=12gx(vX,vX),(x,v)TM.\displaystyle L_{X}(x,v)=\frac{1}{2}g_{x}(v-X,v-X),\quad\forall(x,v)\in TM.

In [16], the authors also obtained

Proposition 1.3.

Let LX:TML_{X}:TM\to\mathbb{R} be the Mañé Lagrangian associated to a CkC^{k} vector field XX on a closed connected Riemannian manifold (M,g)(M,g) with k2k\geqslant 2. Assume that LXL_{X} satisfies the Mather disconnectedness condition. Then we have the following:

  1. (i)

    The projected Aubry set 𝒜(LX)\mathcal{A}(L_{X}) is the set of chain-recurrent points of the flow of XX on MM.

  2. (ii)

    The constants are the only weak KAM solutions of (HJ) associated to LXL_{X} if and only if every point of MM is chain-recurrent under the flow of XX.

Theorem 1.4.

Let XX be a CkC^{k} vector field, with k2k\geqslant 2, on a closed connected Riemannian manifold (M,g)(M,g). Assume that one of the following conditions hold:

  1. (i)

    The dimension of MM is 1 or 2.

  2. (ii)

    The dimension of MM is 3, and the vector field XX never vanishes.

  3. (iii)

    The dimension of MM is 3, and XX is of class C3,1C^{3,1}.

Then the projected Aubry set 𝒜(LX)\mathcal{A}(L_{X}) of the Mañé Lagrangian LX:TML_{X}:TM\to\mathbb{R} associated to XX is the set of chain-recurrent points of the flow of XX on MM. Moreover, the constants are the only weak KAM solutions of (HJ) associated to LXL_{X} if and only if every point of MM is chain-recurrent under the flow of XX.

In a recent work ([10]), applying Bernard-Contreras’s theorem ([3]) the authors proved that there exists a residual subset 𝒢C(M)\mathcal{G}\subset C^{\infty}(M) such that, if L(x,v)=12ef(x)gx(v,v)L(x,v)=\frac{1}{2}e^{f(x)}g_{x}(v,v) with f𝒢f\in\mathcal{G}, then, for any [ω]H1(M,)[\omega]\in H^{1}(M,\mathbb{R}) and Lω(x,v):=L(x,v)ω(v)L_{\omega}(x,v):=L(x,v)-\omega(v), the Mather quotient (𝒜(Lω),,δ)(\mathcal{A}(L_{\omega}),\sim,\delta) has a finite number of elements.

1.2. Mather quotient, Ricci curvature and Harmonic 1-form

In this paper, we adopt a novel geometric perspective to examine Mather’s problem, with a particular focus on the Ricci curvature of the kinetic Riemannian metric. A central objective of our research is to provide an estimation of the Laplacian of the weak KAM solution in relation to the kinetic energy Lagrangian. This Laplacian estimation is intrinsically linked to the core aspects of Mather’s problem.

In his seminal work [19], John Mather observed the invariance of the Euler-Lagrange flow under transformations induced by adding exact 1-forms, and noted that the Aubry set is determined solely by the de Rham cohomology class. Furthermore, leveraging Hodge’s theorem, we understand that on a compact, oriented, smooth manifold, the Hodge cohomology 1(M,)\mathcal{H}^{1}(M,\mathbb{R}) is isomorphic to the de Rham cohomology H1(M,)H^{1}(M,\mathbb{R}). This isomorphism enables us to employ the Hodge cohomology to delve into the rigidity properties of the Aubry set, particularly in the context of the kinetic energy Lagrangian, under the condition that the manifold MM possesses nonnegative Ricci curvature.

The method used in this paper draw inspiration from the celebrating splitting theorem of Gromoll and Cheeger. In the realm of differential geometry, a particularly effective approach to estimating the Laplacian of solutions to the Hamilton-Jacobi equation involves the utilization of the Riccati equation. Through the viewpoint of differential geometry, we provide several estimates for the Laplacian of the weak KAM solution of (HJ). Furthermore, we establish the following results.

Theorem 1.5.

Suppose (M,g)(M,g) is a closed connected Riemannian manifold with nonnegative Ricci curvature. Then, for each weak KAM solution uu of (HJ) associated to the mechanical Lagrangian L(x,v)=12gx(v,v)+f(x)L(x,v)=\frac{1}{2}g_{x}(v,v)+f(x), we have

Δu(x)nk,xM\displaystyle\Delta u(x)\leqslant\sqrt{-nk},\qquad\forall x\in M

in the barrier sense, where n=dimM2n=\dim M\geqslant 2 and kk is some nonpositive number such that Δf(x)k\Delta f(x)\geqslant k for all xMx\in M.

Theorem 1.6.

Suppose L(x,v)=12gx(v,v)f(x)ω(v)L(x,v)=\frac{1}{2}g_{x}(v,v)-f(x)-\omega(v) is a Tonelli Lagrangian on a closed connected Riemannian manifold (M,g)(M,g) with ω\omega is a closed 1-form. Set

E={(x,v)TM:H(x,Lv(x,v))=c[L]},\displaystyle E=\{(x,v)\in TM:H(x,L_{v}(x,v))=c[L]\},

where H:TMH:T^{*}M\rightarrow\mathbb{R} is the Tonelli Hamiltonian associated to LL. If

(x,v)ERic(ρ˙(s))+Δf(ρ(s))0\displaystyle(x,v)\in E\qquad\Longrightarrow\qquad\operatorname{Ric}(\dot{\rho}(s))+\Delta f(\rho(s))\geqslant 0

where ρ(s)=πΦsL(x,v)\rho(s)=\pi\circ\Phi_{s}^{L}(x,v), s(,0]s\in(-\infty,0], then, for any weak KAM solution uu of (HJ), we have

Δu(x)divω#(x),xM\displaystyle\Delta u(x)\leqslant-\operatorname{div}\omega^{\#}(x),\quad\forall x\in M

in the barrier sense.

The theorems above implies some consequences on the Mather quotient and Mañé Lagrangian.

Theorem 1.7.

Given an orientable connected closed Riemannian manifold (M,g)(M,g) with nonnegative Ricci curvature. Let ω\omega be a closed 1-form on MM and let X:=ω#X:=\omega^{\#} be its corresponding vector field. Then, for every constant cc\in\mathbb{R}, each weak KAM solution uu of (HJ) associated to the Lagrangian L(x,v):=12gx(v,v)ω(v)+cL(x,v):=\frac{1}{2}g_{x}(v,v)-\omega(v)+c is constant if and only if ω\omega is a harmonic 1-form.

Corollary 1.8.

Given an orientable connected closed Riemannian manifold (M,g)(M,g) with nonnegative Ricci curvature. Let L(x,v)=12gx(v,v)L(x,v)=\frac{1}{2}g_{x}(v,v) be the kinetic energy associated to the Riemannian metric. Then, for each [ω]H1(M,)[\omega]\in H^{1}(M,\mathbb{R}) and Lω(x,v):=L(x,v)ω(v)L_{\omega}(x,v):=L(x,v)-\omega(v), the projected Aubry set 𝒜(Lω)=M\mathcal{A}(L_{\omega})=M, and the set-valued map H1(M,)[ω]𝒜(Lω){H}^{1}(M,\mathbb{R})\ni[\omega]\rightrightarrows\mathcal{A}(L_{\omega}) is constant. Moreover, the Mather quotient (𝒜(Lω),,δ)(\mathcal{A}(L_{\omega}),\sim,\delta) associated to the Lagrangian LωL_{\omega} is a singleton.

Theorem 1.9.

Given an orientable connected closed Riemannian manifold (M,g)(M,g). Let L(x,v)=12gx(vω,vω)L(x,v)=\frac{1}{2}g_{x}(v-\omega^{\sharp},v-\omega^{\sharp}) be the Mañé’s Lagrangian with ω\omega a closed 1-form and ω#\omega^{\#} its corresponding vector field. Set

E={(x,v)TM:H(x,Lv(x,v))=c[L]},f(x)=12gx(ω,ω),\displaystyle E=\{(x,v)\in TM:H(x,L_{v}(x,v))=c[L]\},\qquad f(x)=\frac{1}{2}g_{x}(\omega^{\sharp},\omega^{\sharp}),

with H(x,p)H(x,p) the associated Tonelli Hamiltonian. If the following condition holds

(x,v)ERic(ρ˙(s))+Δf(ρ(s))0\displaystyle(x,v)\in E\qquad\Longrightarrow\qquad\operatorname{Ric}(\dot{\rho}(s))+\Delta f(\rho(s))\geqslant 0

where ρ(s)=πΦsL(x,v)\rho(s)=\pi\circ\Phi_{s}^{L}(x,v), s(,0]s\in(-\infty,0], then every weak KAM solution uu of (HJ) is constant if and only if ω\omega is a harmonic 1-form, and the Mather quotient (𝒜(L),,δ)(\mathcal{A}(L),\sim,\delta) is a singleton if and only if ω\omega is a harmonic 1-form. In particular, if (M,g)(M,g) has nonnegative Ricci curvature and ω\omega is a harmonic 1-form, each weak KAM solution uu of (HJ) is constant.

The paper is organized as follows: In Section 2, we review certain basic facts from Aubry-Mather theory and Riemannian geometry, with a particular emphasis on the characteristics of conjugate points. Section 3 is dedicated to the Riccati equation. In Section 4 we prove the main results of this paper. The paper also includes two appendices: one provides the proofs for the points mentioned in Section 2, and the other discusses the index form in the context of the Lagrangian framework.

2. Preliminaries and Notions

2.1. Facts from Aubry-Mather theory and weak KAM theory

We now recall the basic facts from Aubry-Mather theory and weak KAM theory (see [14, 17, 15] and more details on semiconcavity in [6, 7]).

If LL is a Tonelli Lagrangian, we define the generating function

At(x,y)=infρΓx,yt0tL(ρ(s),ρ˙(s))𝑑s,t>0,x,yM\displaystyle A_{t}(x,y)=\inf_{\rho\in\Gamma_{x,y}^{t}}\int_{0}^{t}L(\rho(s),\dot{\rho}(s))\ ds,\quad t>0,x,y\in M

where Γx,yt={ρAC([0,t],M):ρ(0)=x,ρ(t)=y}\Gamma_{x,y}^{t}=\{\rho\in AC([0,t],M):\rho(0)=x,\rho(t)=y\}. A minimal curve for At(x,y)A_{t}(x,y) is an absolutely continuous curve ρΓx,yt\rho\in\Gamma^{t}_{x,y} such that

At(x,y)=0tL(ρ(s),ρ˙(s))𝑑s.\displaystyle A_{t}(x,y)=\int_{0}^{t}L(\rho(s),\dot{\rho}(s))ds.

By classical Tonelli theory, the infimum in the definition of At(x,y)A_{t}(x,y) can be achieved and any minimal curve ρ\rho is as smooth as LL. In local charts, ρ\rho satisfies Euler-Lagrange equation

ddsLv(ρ(s),ρ˙(s))=Lx(ρ(s),ρ˙(s)),s[a,b],\displaystyle\frac{d}{ds}L_{v}(\rho(s),\dot{\rho}(s))=L_{x}(\rho(s),\dot{\rho}(s)),\qquad s\in[a,b], (E-L)

We call a C1C^{1} curve ρ:[a,b]M\rho:[a,b]\to M an extremal for the Lagrangian LL if it satisfies (E-L). It is well known that (E-L) defines a complete Euler-Lagrange flow ΦtL:TMTM\Phi_{t}^{L}:TM\to TM.

We denote by dxAt(,y)d_{x}A_{t}(\cdot,y) (resp. dyAt(x,)d_{y}A_{t}(x,\cdot)) the differential of At(x,y)A_{t}(x,y) with respect to the first (resp. second) variable. Similarly, the gradient of At(x,y)A_{t}(x,y) with respect to the first (resp. second) variable will be denoted by xAt(,y)\nabla_{x}A_{t}(\cdot,y) (resp. yAt(x,)\nabla_{y}A_{t}(x,\cdot)).

Proposition 2.1.

If LL is a Tonelli Lagrangian on the connected closed Riemannian manifold (M,g)(M,g). Then, the following statements are true:

  1. (1)

    At(x,y)A_{t}(x,y) is differentiable at yy if and only if there is unique minimal curve ρ:[0,t]M\rho:[0,t]\to M for At(x,y)A_{t}(x,y). Moreover, if At(x,y)A_{t}(x,y) is differentiable at yy, we have

    dyAt(x,y)=Lv(ρ(t),ρ˙(t)).\displaystyle d_{y}A_{t}(x,y)=L_{v}(\rho(t),\dot{\rho}(t)).
  2. (2)

    At(x,y)A_{t}(x,y) is differentiable at xx if and only if there is unique minimal curve ρ:[0,t]M\rho:[0,t]\to M for At(x,y)A_{t}(x,y). Moreover, if At(,y)A_{t}(\cdot,y) is differentiable at xx, we have

    dxAt(x,y)=Lv(ρ(0),ρ˙(0)).\displaystyle d_{x}A_{t}(x,y)=-L_{v}(\rho(0),\dot{\rho}(0)).
  3. (3)

    For any τ>0\tau>0 there exists a compact subset KτTMK_{\tau}\subset TM satisfies the following property: if ρ:[0,t]M\rho:[0,t]\to M is a minimal curve for At(x,y)A_{t}(x,y) with t>τt>\tau, then

    (ρ(s),ρ˙(s))Kτ,s[0,t].\displaystyle(\rho(s),\dot{\rho}(s))\in K_{\tau},\qquad\forall s\in[0,t].

Let uC0(M)u\in C^{0}(M) and t>0t>0. We define respectively the negative and positive Lax-Oleinik operators: for any xMx\in M,

Ttu(x)=infyM{u(y)+At(y,x)},Tt+u(x)=supyM{u(y)At(x,y)}.\displaystyle T_{t}^{-}u(x)=\inf_{y\in M}\{u(y)+A_{t}(y,x)\},\qquad T_{t}^{+}u(x)=\sup_{y\in M}\{u(y)-A_{t}(x,y)\}.

As usual we define T0±=idT^{\pm}_{0}=id. From weak KAM theory, a function uu is a weak KAM solution of (HJ) if and only if Ttu(x)+c[L]t=u(x)T^{-}_{t}u(x)+c[L]t=u(x) for all t0t\geqslant 0. This implies that, if u:Mu:M\to\mathbb{R} is a weak KAM solution of (HJ), then for any continuous piecewise C1C^{1} curve ρ:[a,b]M\rho:[a,b]\to M, a<ba<b,

u(ρ(b))u(ρ(a))abL(ρ(s),ρ˙(s))𝑑s+c[L](ba).\displaystyle u(\rho(b))-u(\rho(a))\leq\int_{a}^{b}L(\rho(s),\dot{\rho}(s))ds+c[L](b-a).

A curve ρ:[a,b]M\rho:[a,b]\to M is (u,L,c[L])(u,L,c[L])-calibrated on [a,b][a,b], or uu-calibrated for short, if for every t,s[a,b]t,s\in[a,b] with tst\leqslant s,

u(ρ(s))u(ρ(t))=tsL(ρ(z),ρ˙(z))𝑑z+c[L](st).\displaystyle u(\rho(s))-u(\rho(t))=\int_{t}^{s}L(\rho(z),\dot{\rho}(z))dz+c[L](s-t).

If uu is a weak KAM solution of (HJ), then for any xMx\in M, there exists a (u,L,c[L])(u,L,c[L])-calibrated curve ρ:(,0]M\rho:(-\infty,0]\to M such that ρ(0)=x\rho(0)=x. One can refer to [7, 24] for more in the case when uu is not differentiable at xx.

For the associated Tonelli Hamiltonian H:TMH:T^{*}M\to\mathbb{R} of LL, in local charts, we have following Hamiltonian ODE

{x˙=Hp(x,p),p˙=Hx(x,p).\left\{\begin{aligned} \dot{x}&=H_{p}(x,p),\\ \dot{p}&=-H_{x}(x,p).\end{aligned}\right. (2.1)

We call ΦtH\Phi_{t}^{H} the Hamiltonian flow associated with ΦtL\Phi_{t}^{L}. The Legendre transform

(x,v)=(x,Lv(x,v))\displaystyle\mathcal{L}(x,v)=(x,L_{v}(x,v))

define a diffeomorphism from TMTM to TMT^{*}M, and it establishes a correspondence between the Euler-Lagrange flow ΦtL\Phi_{t}^{L} and its corresponding Hamilton flow by

ΦtH=ΦtL1.\displaystyle\Phi_{t}^{H}=\mathcal{L}\circ\Phi_{t}^{L}\circ\mathcal{L}^{-1}.

2.2. Facts from Riemannian Geometry

Let us recall some basic facts about Riemannian geometry. For more details in Riemannian geometry we refer to [11, 23, 25].

Let \nabla be the Riemannian connection on (M,g)(M,g). The curvature tensor of the Riemannian connection \nabla is defined by

R:Γ(TM)×Γ(TM)×Γ(TM)\displaystyle R:\Gamma(TM)\times\Gamma(TM)\times\Gamma(TM)\to Γ(TM),\displaystyle\,\Gamma(TM),
(X,Y,Z)\displaystyle(X,Y,Z)\mapsto R(X,Y)Z=XYZYXZ[X,Y]Z.\displaystyle\,R(X,Y)Z=\nabla_{X}\nabla_{Y}Z-\nabla_{Y}\nabla_{X}Z-\nabla_{[X,Y]}Z.

The Ricci curvature at vTpMv\in T_{p}M is defined as

Ric(v)=tr(wR(w,v)v).\displaystyle\operatorname{Ric}(v)=\operatorname{tr}(w\mapsto R(w,v)v).

Next, we introduce some differential operators on Riemannian manifold. Given a Riemannian manifold (M,g)(M,g) with its Riemannian connection \nabla.

  1. The gradient of a function fC1(M)f\in C^{1}(M) is given by

    :C1(M)\displaystyle\nabla:C^{1}(M) C(M),\displaystyle\,\to C(M),
    f\displaystyle\nabla f\mapsto (df)#.\displaystyle\,(df)^{\#}.

    An equivalent definition is that f(x)\nabla f(x) is the unique vector in TxMT_{x}M such that

    df(v)=g(f,v)\displaystyle df(v)=g(\nabla f,v)

    for any vTxMv\in T_{x}M.

  2. The divergence of a vector field XΓ(TM)X\in\Gamma(TM) is

    div:Γ(TM)C(M)\displaystyle\operatorname{div}:\Gamma(TM)\to C^{\infty}(M)
    divXtr(YYX).\displaystyle\operatorname{div}X\mapsto\operatorname{tr}(Y\mapsto\nabla_{Y}X).
  3. The Laplacian operator is defined as

    Δ:C2(M)\displaystyle\Delta:C^{2}(M)\to C(M),\displaystyle\,C(M),
    Δf\displaystyle\Delta f\mapsto divf.\displaystyle\,\operatorname{div}\nabla f.
  4. When seen as a (1,1)(1,1) type tensor, the Hessian of fC2(M)f\in C^{2}(M) is given by

    Hessf:Γ(TM)\displaystyle\operatorname{Hess}f:\Gamma(TM)\to Γ(TM),\displaystyle\,\Gamma(TM),
    Hessf(X)\displaystyle\operatorname{Hess}f(X)\mapsto Xf.\displaystyle\,\nabla_{X}\nabla f.
  5. When viewed as a (0,2)(0,2) type tensor, the Hessian of fC2(M)f\in C^{2}(M) is

    2f:Γ(TM)×Γ(TM)\displaystyle\nabla^{2}f:\Gamma(TM)\times\Gamma(TM)\to C(M),\displaystyle\,C^{\infty}(M),
    2f(X,Y)\displaystyle\nabla^{2}f(X,Y)\mapsto g(Xf,Y).\displaystyle\,g(\nabla_{X}\nabla f,Y).

In fact, we can consider Hessian operators for a function ff that are not of class C2C^{2}, even not continuously differentiable. For more details about this we refer to [27].

Given a compact oriented Riemannian manifold (M,g)(M,g). Let Ωk(M)\Omega^{k}(M) be the space of kk-form on MM and let Ω(M)=kΩk(M)\Omega(M)=\cup_{k}\Omega^{k}(M). Recall that the Riemannian metric gg induces an inner product on TpMT_{p}^{*}M. Extending this inner product from TpMT_{p}^{*}M to its k-th exterior wedge k(TpM)\bigwedge^{k}(T_{p}^{*}M) one can obtain an inner product ,p\langle\cdot,\cdot\rangle_{p} on k(TpM)\bigwedge^{k}(T_{p}^{*}M). The inner product on Ωk(M)\Omega^{k}(M) is then defined as

Ωk(M)×Ωk(M)\displaystyle\Omega^{k}(M)\times\Omega^{k}(M)\to ,\displaystyle\,\mathbb{R},
(ω,η)\displaystyle(\omega,\eta)\mapsto Mω,ηpσ,\displaystyle\,\int_{M}\langle\omega,\eta\rangle_{p}\sigma,

where σ\sigma is the volume form associated to gg. If we require (ω,η)=0(\omega,\eta)=0 for ωΩk(M)\omega\in\Omega^{k}(M), ηΩl(M)\eta\in\Omega^{l}(M) with klk\neq l, we get an inner product (,)(\cdot,\cdot) on Ω(M)\Omega(M).

Since the exterior differential operator d:Ωk(M)Ωk+1(M)d:\Omega^{k}(M)\to\Omega^{k+1}(M) is a linear operator on the inner space (Ω(M),(,))(\Omega(M),(\cdot,\cdot)), one has a linear adjoint operator

δ:Ωk+1(M)Ωk(M)\displaystyle\delta:\Omega^{k+1}(M)\to\Omega^{k}(M)

of dd such that (dω,η)=(ω,δη)(d\omega,\eta)=(\omega,\delta\eta). The Hodge Laplacian is then defined by

Δ=dδ+δd.\displaystyle\varDelta=d\delta+\delta d.

This is a second order linear differential operator. By definition, ωΩ(M)\omega\in\Omega(M) is a harmonic form if Δω=0\varDelta\omega=0.

Now we list some basic facts about Hodge cohomology (see, for instance, [23, 28]).

Proposition 2.2.

Let (M,g)(M,g) be a compact oriented Riemannian manifold. The following statements hold true:

  1. (a)

    Every harmonic form is closed.

  2. (b)

    A closed 1-form ω\omega is harmonic if and only if divω=0\operatorname{div}\omega^{\sharp}=0.

  3. (c)

    (Hodge Theorem): The Hodge Cohomology (M,)\mathcal{H}(M,\mathbb{R}) is isomorphic to the De Rham Cohomology H(M,)H(M,\mathbb{R}).

  4. (d)

    (Bochner Theorem): If (M,g)(M,g) has nonnegative Ricci curvature, then g(ω,ω)g(\omega^{\sharp},\omega^{\sharp}) is constant for every harmonic 1-form ω\omega.

2.3. Conjugate Points and Jacobi Fields

In order to estimate the Laplacian of the generating function At(x,y)A_{t}(x,y), we need to discuss the conjugate points and the Jacobi fields. This topic is well known in Riemannian geometry. However, for the sake of convenience we shall deal with these points in the frame of Lagrange geometry (see, also [8] in the Hamiltonian frame).

Definition 2.3.

Suppose L:TML:TM\to\mathbb{R} is a Tonelli Lagrangian on a closed connected Riemannian manifold (M,g)(M,g) and ρ:[a,b]M\rho:[a,b]\to M is an extremal. A variation of extremal curves along ρ\rho is a map ρ(t,s)C2([a,b]×(ε,ε))\rho(t,s)\in C^{2}([a,b]\times(-\varepsilon,\varepsilon)) satisfying

  1. (1)

    ρ(t,0)=ρ(t)\rho(t,0)=\rho(t) for all t[a,b]t\in[a,b].

  2. (2)

    ρ(,s):[a,b]M\rho(\cdot,s):[a,b]\to M is an extremal curve for each s(ε,ε)s\in(-\varepsilon,\varepsilon).

Let J:[a,b]TMJ:[a,b]\to TM be a vector field along ρ\rho. We say that JJ is a Jacobi field if one can find a variation ρ(t,s)C2([a,b]×(ε,ε))\rho(t,s)\in C^{2}([a,b]\times(-\varepsilon,\varepsilon)) of extremal curves along ρ\rho such that

J(t)=s|s=0ρ(t,s).\displaystyle J(t)=\left.\frac{\partial}{\partial s}\right|_{s=0}\rho(t,s).
Proposition 2.4.

Let L:TML:TM\to\mathbb{R} be a Tonelli Lagrangian on a closed connected Riemannian manifold (M,g)(M,g). If ρ:[a,b]M\rho:[a,b]\to M is an extremal and J:[a,b]TMJ:[a,b]\to TM is a vector field along ρ\rho, then JJ is a Jacobi field along ρ\rho if and only if JJ solves the second order linear Jacobi equation

ddt(Lvx(ρ(t),ρ˙(t))J(t)+Lvv(ρ(t),ρ˙(t))J˙(t))=Lxx(ρ(t),ρ˙(t))J(t)+Lxv(ρ(t),ρ˙(t))J˙(t).\frac{d}{dt}(L_{vx}(\rho(t),\dot{\rho}(t))J(t)+L_{vv}(\rho(t),\dot{\rho}(t))\dot{J}(t))=L_{xx}(\rho(t),\dot{\rho}(t))J(t)+L_{xv}(\rho(t),\dot{\rho}(t))\dot{J}(t). (2.2)

in local chart.

Proposition 2.5.

Suppose J:[a,b]TMJ:[a,b]\to TM is a Jacobi field along ρ:[a,b]M\rho:[a,b]\to M such that J(0)=0J(0)=0. Then

J(s)=z|z=0πΦsaL(ρ(a),ρ˙(a)+zρ˙(a)J).\displaystyle J(s)=\left.\partial_{z}\right|_{z=0}\pi\circ\Phi_{s-a}^{L}(\rho(a),\dot{\rho}(a)+z\nabla_{\dot{\rho}(a)}J).
Definition 2.6.

If L:TML:TM\to\mathbb{R} is a Tonelli Lagrangian and ρ:[a,b]M\rho:[a,b]\to M is an extremal. The point (ρ(b),ρ˙(b))(\rho(b),\dot{\rho}(b)) is said to be conjugate to (ρ(a),ρ˙(a))(\rho(a),\dot{\rho}(a)) if there exists a nonzero Jacobi field JJ along ρ\rho such that

ρ(a)=ρ(b)=0.\displaystyle\rho(a)=\rho(b)=0.

In general, a Tonelli Lagrangian L:TML:TM\to\mathbb{R} is not necessarily symmetrical. So one can define the reverse of LL by L˘(x,v)=L(x,v)\breve{L}(x,v)=L(x,-v). Simultaneously, one gets the reverse Hamiltonian H˘(x,p):=H(x,p)\breve{H}(x,p):=H(x,-p) which is exactly the Hamiltonian associated to L˘\breve{L}. The next proposition clarifies the relation of conjugacy with respect to LL and L˘\breve{L} respectively.

Proposition 2.7.

Let L:TML:TM\to\mathbb{R} be a Tonelli Lagrangian and let L˘(x,v):=L(x,v)\breve{L}(x,v):=L(x,-v) for all (x,v)TM(x,v)\in TM. If ρ:[a,b]M\rho:[a,b]\to M is an extremal, then, (ρ(b),ρ˙(b))(\rho(b),\dot{\rho}(b)) is conjugate to (ρ(a),ρ˙(a))(\rho(a),\dot{\rho}(a)) with respect to LL if and only if (ρ(a),ρ˙(a))(\rho(a),-\dot{\rho}(a)) is conjugate to (ρ(b),ρ˙(b))(\rho(b),-\dot{\rho}(b)) with respect to L˘\breve{L}.

Proposition 2.8.

Given a mechanical Lagrangian L(x,v)=12gx(v,v)f(x)L(x,v)=\frac{1}{2}g_{x}(v,v)-f(x) on a connected closed manifold (M,g)(M,g). Let ω\omega be a closed 1-form on MM and let X:=ω#X:=\omega^{\#} be its corresponding vector field. Then, any minimizer ρ:[0,t]M\rho:[0,t]\to M of At(ρ(0),ρ(t))A_{t}(\rho(0),\rho(t)) associated to the Lagrangian

Lω(x,v):=L(x,v)ω(v)=L(x,v)gx(X,v)\displaystyle L_{\omega}(x,v):=L(x,v)-\omega(v)=L(x,v)-g_{x}(X,v)

solves

ρ˙ρ˙=f,\nabla_{\dot{\rho}}\dot{\rho}=-\nabla f, (2.3)

and each Jacobi field JJ along ρ\rho satisfies

ρ˙ρ˙J+R(J,ρ˙)ρ˙+Hessf(J)=0.\nabla_{\dot{\rho}}\nabla_{\dot{\rho}}J+R(J,\dot{\rho})\dot{\rho}+\operatorname{Hess}f(J)=0. (2.4)

Equation (2.3) is equivalent to the Euler-Lagrange equation (E-L). Indeed, we have L(x,v)=12gij(x)vivjf(x)ωlvlL(x,v)=\frac{1}{2}g_{ij}(x)v^{i}v^{j}-f(x)-\omega_{l}v^{l} in local chart. Here and after we use the Einstein summation convention. Then, we have

Lvk=gki(x)viωk,\displaystyle\,L_{v_{k}}=g_{ki}(x)v^{i}-\omega_{k},
Lxk=12gijxk(x)vivjfxk(x)ωlxk(x)vl.\displaystyle\,L_{x_{k}}=\frac{1}{2}\frac{\partial g_{ij}}{\partial x_{k}}(x)v^{i}v^{j}-\frac{\partial f}{\partial x_{k}}(x)-\frac{\partial\omega_{l}}{\partial x_{k}}(x)v^{l}.

The Euler-Lagrange equation (E-L) tells us

ddtLvk(ρ(t),ρ˙(t))=Lxk(ρ(t),ρ˙(t))\displaystyle\frac{d}{dt}L_{v_{k}}(\rho(t),\dot{\rho}(t))=L_{x_{k}}(\rho(t),\dot{\rho}(t))

which yields to

12gijxk(ρ)ρ˙iρ˙jfxk(ρ)ωlxk(ρ)ρ˙l=\displaystyle\frac{1}{2}\frac{\partial g_{ij}}{\partial x_{k}}(\rho)\dot{\rho}^{i}\dot{\rho}^{j}-\frac{\partial f}{\partial x_{k}}(\rho)-\frac{\partial\omega_{l}}{\partial x_{k}}(\rho)\dot{\rho}^{l}= gkixj(ρ)ρ˙iρ˙j+gki(ρ)ρ¨iωkxj(ρ)ρ˙j\displaystyle\,\frac{\partial g_{ki}}{\partial x_{j}}(\rho)\dot{\rho}^{i}\dot{\rho}^{j}+g_{ki}(\rho)\ddot{\rho}^{i}-\frac{\partial\omega_{k}}{\partial x_{j}}(\rho)\dot{\rho}^{j}
=\displaystyle= 12gkixj(ρ)ρ˙iρ˙j+12gkjxi(ρ)ρ˙iρ˙j+gki(ρ)ρ¨iωkxj(ρ)ρ˙j.\displaystyle\,\frac{1}{2}\frac{\partial g_{ki}}{\partial x_{j}}(\rho)\dot{\rho}^{i}\dot{\rho}^{j}+\frac{1}{2}\frac{\partial g_{kj}}{\partial x_{i}}(\rho)\dot{\rho}^{i}\dot{\rho}^{j}+g_{ki}(\rho)\ddot{\rho}^{i}-\frac{\partial\omega_{k}}{\partial x_{j}}(\rho)\dot{\rho}^{j}.

Since ω\omega is closed, we have

ωlxk(ρ)ρ˙l=ωjxk(ρ)ρ˙j=ωkxj(ρ)ρ˙j.\displaystyle\frac{\partial\omega_{l}}{\partial x_{k}}(\rho)\dot{\rho}^{l}=\frac{\partial\omega_{j}}{\partial x_{k}}(\rho)\dot{\rho}^{j}=\frac{\partial\omega_{k}}{\partial x_{j}}(\rho)\dot{\rho}^{j}.

This means that

12gijxk(ρ)ρ˙iρ˙jfxk(ρ)=12gkixj(ρ)ρ˙iρ˙j+12gkjxi(ρ)ρ˙iρ˙j+gki(ρ)ρ¨i.\displaystyle\frac{1}{2}\frac{\partial g_{ij}}{\partial x_{k}}(\rho)\dot{\rho}^{i}\dot{\rho}^{j}-\frac{\partial f}{\partial x_{k}}(\rho)=\frac{1}{2}\frac{\partial g_{ki}}{\partial x_{j}}(\rho)\dot{\rho}^{i}\dot{\rho}^{j}+\frac{1}{2}\frac{\partial g_{kj}}{\partial x_{i}}(\rho)\dot{\rho}^{i}\dot{\rho}^{j}+g_{ki}(\rho)\ddot{\rho}^{i}.

Hence, we obtain

ρ¨k=gkm12{gmixjρ˙iρ˙j+gmjxiρ˙iρ˙jgijxmρ˙iρ˙j}gkmfxm=Γijkρ˙iρ˙jgkmfxm,\begin{split}\ddot{\rho}^{k}=&\,-g^{km}\frac{1}{2}\left\{\frac{\partial g_{mi}}{\partial x_{j}}\dot{\rho}^{i}\dot{\rho}^{j}+\frac{\partial g_{mj}}{\partial x_{i}}\dot{\rho}^{i}\dot{\rho}^{j}-\frac{\partial g_{ij}}{\partial x_{m}}\dot{\rho}^{i}\dot{\rho}^{j}\right\}-g^{km}\frac{\partial f}{\partial x_{m}}\\ =&\,-\Gamma_{ij}^{k}\dot{\rho}^{i}\dot{\rho}^{j}-g^{km}\frac{\partial f}{\partial x_{m}},\end{split} (2.5)

where

Γijk=gkm12{gmjxi+gmixjgijxm}\displaystyle\Gamma_{ij}^{k}=g^{km}\frac{1}{2}\left\{\frac{\partial g_{mj}}{\partial x_{i}}+\frac{\partial g_{mi}}{\partial x_{j}}-\frac{\partial g_{ij}}{\partial x_{m}}\right\}

are Christoffel symbols. Notice that equation (2.3) is also equivalent to equation (2.5) in local chart. Therefore, equation (2.3) is equivalent to the Euler-Lagrange equation (E-L).

Lemma 2.9.

Suppose L:TML:TM\to\mathbb{R} is a Tonelli Lagrangian on a closed connected Riemannian manifold (M,g)(M,g). Let ρ:[a,b]M\rho:[a,b]\to M be an extremal. Then (ρ(b),ρ˙(b))(\rho(b),\dot{\rho}(b)) is not conjugate to (ρ(a),ρ˙(a))(\rho(a),\dot{\rho}(a)), s>ts>t if and only if dρ˙(a)(πΦbaL)d_{\dot{\rho}(a)}(\pi\circ\Phi_{b-a}^{L}) is non-degenerate.

Now we come to the connection between the conjugate points and the differentiability of At(x,)A_{t}(x,\cdot).

Proposition 2.10.

Suppose ρ:[0,t]M\rho:[0,t]\to M is a minimal curve for At(x,y)A_{t}(x,y) and (ρ(t),ρ(t))(\rho(t),\rho(t)) is not conjugate to (ρ(0),ρ˙(0))(\rho(0),\dot{\rho}(0)). Then At(x,)A_{t}(x,\cdot) is of class CrC^{r} in a neighborhood UU of ρ(t)\rho(t) provided At(x,)A_{t}(x,\cdot) is differentiable at ρ(t)\rho(t). Moreover,

πΦtL:(πΦtL)1(U)U\displaystyle\pi\circ\Phi_{t}^{L}:(\pi\circ\Phi_{t}^{L})^{-1}(U)\to U

is a Cr1C^{r-1} diffeomorphism and the curve ρz(s):=πΦsL(x,vz)\rho_{z}(s):=\pi\circ\Phi_{s}^{L}(x,v_{z}) is the unique minimal curve for At(x,z)A_{t}(x,z) where TxMvz=(πΦtL)1(z)T_{x}M\ni v_{z}=(\pi\circ\Phi_{t}^{L})^{-1}(z).

Now we introduce the definition of the cut points which play an important role in calculus of variations.

Definition 2.11.

Suppose ρ:[0,t]M\rho:[0,t]\to M is a minimal curve for At(x,y)A_{t}(x,y). Then, ρ(t)\rho(t) is a cut point of ρ(0)\rho(0) if the curve ρ¯(s)=πΦsL(x,ρ˙(0))\bar{\rho}(s)=\pi\circ\Phi_{s}^{L}(x,\dot{\rho}(0)), s[0,τ]s\in[0,\tau], is not a minimal curve for Aτ(x,ρ¯(τ))A_{\tau}(x,\bar{\rho}(\tau)) for any τ>t\tau>t.

Lemma 2.12.

If ρ:[0,t]M\rho:[0,t]\to M is a minimal curve for At(x,y)A_{t}(x,y) and ρ(t)\rho(t) is a cut point of ρ(0)\rho(0), then either

  1. (i)

    (ρ(t),ρ˙(t))(\rho(t),\dot{\rho}(t)) is conjugate to (ρ(0),ρ˙(0))(\rho(0),\dot{\rho}(0)), or

  2. (ii)

    there exists another minimizer ρ~:[0,t]M\tilde{\rho}:[0,t]\rightarrow M of At(x,y)A_{t}(x,y).

Theorem 2.13.

Let L:TML:TM\to\mathbb{R} be a Tonelli Lagrangian on a closed connected Riemannian manifold (M,g)(M,g) and let uu be a weak KAM solution of (HJ). Suppose that xMx\in M and ρ:(,0]M\rho:(-\infty,0]\to M is a (u,L,c[L])(u,L,c[L])-calibrated curve ending at ρ(0)=x\rho(0)=x.

  1. (1)

    (ρ(0),ρ˙(0))(\rho(0),\dot{\rho}(0)) is not conjugate to (ρ(s),ρ˙(s))(\rho(-s),\dot{\rho}(-s)) for any s>0s>0.

  2. (2)

    For every τ>0\tau>0, Aτ(ρ(τ),)A_{\tau}(\rho(-\tau),\cdot) is of C2C^{2} in a neighborhood of ρ(0)=x\rho(0)=x and ρ(0)\rho(0) is not a cut point of ρ(τ)\rho(-\tau).

Proof.

Statement (1) is a direct consequence of Proposition B.4. To prove (2), it is sufficient to prove xx is a point of differentiability of Aτ(ρ(τ),)A_{\tau}(\rho(-\tau),\cdot), for any τ>0\tau>0, by Proposition 2.10 and Lemma 2.12. Otherwise, there exists another minimizer α:[0,τ]M\alpha:[0,\tau]\to M of Aτ(ρ(τ),x)A_{\tau}(\rho(-\tau),x). Then, the speed curve of

α1(t)={ρ(t),t[τ1,τ]α(t+τ),t[τ,0]\displaystyle\alpha_{1}(t)=\begin{cases}\rho(t),&t\in[-\tau-1,-\tau]\\ \alpha(t+\tau),&t\in[-\tau,0]\end{cases}

satisfies (E-L) which contradicts to the Cauchy-Lipschitz Theorem. ∎

3. Riccati Equation

Now we turn to the associated Riccati equation which will help us to estimate the Laplacian of the fundamental solution At(x,y)A_{t}(x,y) and the Laplacian of weak KAM solution in the barrier sense. Our approach is inspired by the Gromoll and Cheeger’s splitting theorem on a non-compact manifold with nonnegative Ricci curvature. The properties of rays ensure the non-existence of conjugate point (or cut point) in positive direction. Fortunately, in the compact case, Fathi’s weak KAM theorem constructs the backward calibrated curves which play the same role as the rays. In principle, the key point in the following discussion is to avoid the trouble of the existence of conjugate points.

In this section, we suppose that dimM=n2\dim M=n\geqslant 2 and the Lagrangian LL has the form

L(x,v)=12gx(v,v)f(x)ω(v)=12gx(v,v)f(x)gx(X,v)\displaystyle L(x,v)=\frac{1}{2}g_{x}(v,v)-f(x)-\omega(v)=\frac{1}{2}g_{x}(v,v)-f(x)-g_{x}(X,v)

where ω\omega is a closed 1-form on MM and X:=ω#X:=\omega^{\#} be its corresponding vector field.

Theorem 3.1.

If ρ:[0,t]M\rho:[0,t]\to M is a minimal curve for At(ρ(0),ρ(t))A_{t}(\rho(0),\rho(t)) and (ρ(t),ρ˙(t))(\rho(t),\dot{\rho}(t)) is not conjugate to (ρ(0),ρ˙(0))(\rho(0),\dot{\rho}(0)), then we have

Θ˙(s)+1nΘ2(s)+Ric(ρ˙(s))+Δf(ρ(s))0,s(0,t],\dot{\Theta}(s)+\frac{1}{n}\Theta^{2}(s)+\operatorname{Ric}(\dot{\rho}(s))+\Delta f(\rho(s))\leqslant 0,\quad s\in(0,t], (3.1)

where Θ(s)=ΔyAs(ρ(0),ρ(s))+divX(ρ(s)).\Theta(s)=\Delta_{y}A_{s}(\rho(0),\rho(s))+\operatorname{div}X(\rho(s)).

Proof.

Let (e1,e2,,en)(e_{1},e_{2},\cdots,e_{n}) be an orthonormal basis of Tρ(0)MT_{\rho(0)}M and let us parallel transport along ρ\rho to define a new family (e1(s),e2(s),,en(s))(e_{1}(s),e_{2}(s),\ldots,e_{n}(s)) in Tρ(s)MT_{\rho(s)}M. Denote πΦsL(x,ρ˙(0)+zei)\pi\circ\Phi_{s}^{L}(x,\dot{\rho}(0)+ze_{i}) by ρi(s,z),i=1,2,,n\rho_{i}(s,z),i=1,2,\cdots,n. By Proposition 2.5,

Ji(s)=z|z=0ρi(s,z),i=1,2,,n,\displaystyle J_{i}(s)=\left.\partial_{z}\right|_{z=0}\rho_{i}(s,z),\qquad i=1,2,\ldots,n,

are Jacobi fields with Ji(0)=0,ρ˙Ji(0)=eiJ_{i}(0)=0,\nabla_{\dot{\rho}}J_{i}(0)=e_{i}. Let Ji(s)=j=1naij(s)ej(s),i=1,2,,nJ_{i}(s)=\sum_{j=1}^{n}a_{ij}(s)e_{j}(s),i=1,2,\cdots,n, then

j=1na¨ij(s)ej(s)+j=1naij(s)R(ej(s),ρ˙(s))ρ˙(s)+j=1naij(s)Hessf(ej(s))=0\displaystyle\sum_{j=1}^{n}\ddot{a}_{ij}(s)e_{j}(s)+\sum_{j=1}^{n}a_{ij}(s)R(e_{j}(s),\dot{\rho}(s))\dot{\rho}(s)+\sum_{j=1}^{n}a_{ij}(s)\operatorname{Hess}f(e_{j}(s))=0

by (2.4). This implies that

a¨ik(s)+j=1naij(s)g(R(ej(s),ρ˙(s))ρ˙(s),ek(s))+j=1naij(s)g(Hessf(ej(s)),ek(s))=0\displaystyle\ddot{a}_{ik}(s)+\sum_{j=1}^{n}a_{ij}(s)g(R(e_{j}(s),\dot{\rho}(s))\dot{\rho}(s),e_{k}(s))+\sum_{j=1}^{n}a_{ij}(s)g(\operatorname{Hess}f(e_{j}(s)),e_{k}(s))=0

for i,k=1,2,,ni,k=1,2,\ldots,n. Therefore, we have the following matrix Riccati equation

A¨(s)+A(s)R(s)+A(s)2f(s)=0,\displaystyle\ddot{A}(s)+A(s)R(s)+A(s)\nabla^{2}f(s)=0,

where

A(s)=\displaystyle A(s)= (aij(s))1i,jn,\displaystyle\,(a_{ij}(s))_{1\leqslant i,j\leqslant n},
R(s)=\displaystyle R(s)= (g(R(ei(s),ρ˙(s))ρ˙(s),ej(s))1i,jn,\displaystyle\,(g(R(e_{i}(s),\dot{\rho}(s))\dot{\rho}(s),e_{j}(s))_{1\leqslant i,j\leqslant n},
2f(s)=\displaystyle\nabla^{2}f(s)= g(Hessf(ei(s)),ej(s))1i,jn.\displaystyle\,g(\operatorname{Hess}f(e_{i}(s)),e_{j}(s))_{1\leqslant i,j\leqslant n}.

We now claim that

ρi˙Ji(s)=aij(s)HessAsx(ej(s)),\displaystyle\nabla_{\dot{\rho_{i}}}J_{i}(s)=a_{ij}(s)\operatorname{Hess}A_{s}^{x}(e_{j}(s)),

where Asx():=As(ρ(0),)A_{s}^{x}(\cdot):=A_{s}(\rho(0),\cdot).

Indeed, by Proposition 2.10 we have

ρ˙Ji(s)=\displaystyle\nabla_{\dot{\rho}}J_{i}(s)= sρizρi(s,0)=zρisρi(s,0)=zρisπΦsL(x,ρ˙(0)+zei)(s,0)\displaystyle\,\nabla_{\partial_{s}\rho_{i}}\partial_{z}\rho_{i}(s,0)=\nabla_{\partial_{z}\rho_{i}}\partial_{s}\rho_{i}(s,0)=\nabla_{\partial_{z}\rho_{i}}\frac{\partial}{\partial s}\pi\circ\Phi_{s}^{L}(x,\dot{\rho}(0)+ze_{i})(s,0)
=\displaystyle= zρiΦsL(x,ρ˙(0)+zei)(s,0)=zρi(ρi(s,z),ρ˙i(s,z))(s,0),\displaystyle\,\nabla_{\partial_{z}\rho_{i}}\Phi_{s}^{L}(x,\dot{\rho}(0)+ze_{i})(s,0)=\nabla_{\partial_{z}\rho_{i}}(\rho_{i}(s,z),\dot{\rho}_{i}(s,z))(s,0),

where for any r[0,s]r\in[0,s], ρi(r,z)=πΦrL(ρ(0),ρ˙(0)+zei)\rho_{i}(r,z)=\pi\circ\Phi_{r}^{L}(\rho(0),\dot{\rho}(0)+ze_{i}) is the unique minimal curve for As(ρ(0),πΦsL(x,ρ˙(0)+zei))A_{s}(\rho(0),\pi\circ\Phi_{s}^{L}(x,\dot{\rho}(0)+ze_{i})) for zz small enough.

By Proposition 2.1 we have Lv(ρi(s,z),ρ˙i(s,z))=dyAs(ρ(0),ρi(s,z))L_{v}(\rho_{i}(s,z),\dot{\rho}_{i}(s,z))=d_{y}A_{s}(\rho(0),\rho_{i}(s,z)), and this implies that

(ρi(s,z),ρ˙i(s,z))=1(ρi(s,z),dyAs(ρ(0),ρi(s,z)))=yAs(ρ(0),ρi(s,z))+X(ρi(s,z)),\displaystyle(\rho_{i}(s,z),\dot{\rho}_{i}(s,z))=\mathcal{L}^{-1}(\rho_{i}(s,z),d_{y}A_{s}(\rho(0),\rho_{i}(s,z)))=\nabla_{y}A_{s}(\rho(0),\rho_{i}(s,z))+X(\rho_{i}(s,z)),

where H(x,p)=12gx(p+ω,p+ω)+f(x)H(x,p)=\frac{1}{2}g_{x}^{*}(p+\omega,p+\omega)+f(x) is the Tonelli Hamiltonian associated to LL. Therefore,

ρ˙Ji(s)=\displaystyle\nabla_{\dot{\rho}}J_{i}(s)= zρi(ρi(s,z),ρ˙i(s,z))(s,0)=Ji(s)(yAsx+X)\displaystyle\,\nabla_{\partial_{z}\rho_{i}}(\rho_{i}(s,z),\dot{\rho}_{i}(s,z))(s,0)=\nabla_{J_{i}(s)}(\nabla_{y}A_{s}^{x}+X)
=\displaystyle= j=1naij(s)ej(s)(yAsx+X)=j=1naij(s)(HessAsx(ej(s))+ej(s)X).\displaystyle\,\nabla_{\sum_{j=1}^{n}a_{ij}(s)e_{j}(s)}(\nabla_{y}A_{s}^{x}+X)=\sum_{j=1}^{n}a_{ij}(s)(\operatorname{Hess}A_{s}^{x}(e_{j}(s))+\nabla_{e_{j}(s)}X).

Notice that

g(ρ˙Ji(s),ek(s))=\displaystyle g(\nabla_{\dot{\rho}}J_{i}(s),e_{k}(s))= j=1naij(s){g(HessAsx(ej(s),ek(s))+g(ej(s)X,ek(s))}\displaystyle\,\sum_{j=1}^{n}a_{ij}(s)\{g(\operatorname{Hess}A_{s}^{x}(e_{j}(s),e_{k}(s))+g(\nabla_{e_{j}(s)}X,e_{k}(s))\}
=\displaystyle= j=1na˙ij(s)g(ej(s),ek(s)).\displaystyle\sum_{j=1}^{n}\dot{a}_{ij}(s)g(e_{j}(s),e_{k}(s)).

Thus, we obtain

A(s)(2Asx+B(s))=A˙(s),\displaystyle A(s)(\nabla^{2}A_{s}^{x}+B(s))=\dot{A}(s),

where 2Asx=(g(HessAsx(ei(s)),ej(s))1i,jn\nabla^{2}A_{s}^{x}=(g(\operatorname{Hess}A_{s}^{x}(e_{i}(s)),e_{j}(s))_{1\leqslant i,j\leqslant n} and B(s)=(g(ei(s)X,ej(s)))1i,jnB(s)=(g(\nabla_{e_{i}(s)}X,e_{j}(s)))_{1\leqslant i,j\leqslant n}.

Since (ρ(s),ρ˙(s))(\rho(s),\dot{\rho}(s)) is not conjugate to (ρ(0),ρ˙(0))(\rho(0),\dot{\rho}(0)), Ji(s)=dρ˙(0)(πΦsL)(ei)J_{i}(s)=d_{\dot{\rho}(0)}(\pi\circ\Phi_{s}^{L})(e_{i}) are linear independent and A(s)A(s) is invertible for each s(0,t]s\in(0,t]. Let Λ(s)=A1(s)A˙(s)=2Asx+B(s)\Lambda(s)=A^{-1}(s)\dot{A}(s)=\nabla^{2}A_{s}^{x}+B(s). We get that

Λ˙(s)=\displaystyle\dot{\Lambda}(s)= A1(s)A˙(s)A1(s)A˙(s)+A1(s)A¨(s)\displaystyle\,-A^{-1}(s)\dot{A}(s)A^{-1}(s)\dot{A}(s)+A^{-1}(s)\ddot{A}(s)
=\displaystyle= Λ2(s)+A1(s)(A(s)R(s)A(s)2f(s))\displaystyle\,-\Lambda^{2}(s)+A^{-1}(s)(-A(s)R(s)-A(s)\nabla^{2}f(s))
=\displaystyle= Λ2(s)R(s)2f(s)\displaystyle\,-\Lambda^{2}(s)-R(s)-\nabla^{2}f(s)

and trΛ(s)=ΔyAs(ρ(0),ρ(s))+divX(ρ(s))\operatorname{tr}\Lambda(s)=\Delta_{y}A_{s}(\rho(0),\rho(s))+\operatorname{div}X(\rho(s)). We rewrite the equality above as

Λ˙(s)+Λ2(s)+R(s)+2f(s)=0.\dot{\Lambda}(s)+\Lambda^{2}(s)+R(s)+\nabla^{2}f(s)=0. (3.2)

By taking trace of (3.2), we arrive at

ddt(trΛ(s))+tr(Λ2(s))+Ric(ρ˙(s))+Δf(ρ(s))=0.\displaystyle\frac{d}{dt}(\operatorname{tr}\Lambda(s))+\operatorname{tr}(\Lambda^{2}(s))+\operatorname{Ric}(\dot{\rho}(s))+\Delta f(\rho(s))=0.

Set trΛ(s)=Θ(s)\operatorname{tr}\Lambda(s)=\Theta(s). Then (3.1) follows by recalling the inequality tr(Λ2(s))1ntr2(Λ(s))\operatorname{tr}(\Lambda^{2}(s))\geqslant\frac{1}{n}\operatorname{tr}^{2}(\Lambda(s)). ∎

Next, we give some basic comparison estimates of the Riccati equation that will be needed later.

Lemma 3.2.

Consider a C1C^{1} function α:(0,t)\alpha:(0,t)\to\mathbb{R} such that

α˙(s)+1nα2(s)+k0,lims0+s2α(s)=0.\displaystyle\dot{\alpha}(s)+\frac{1}{n}\alpha^{2}(s)+k\leqslant 0,\qquad\lim_{s\to 0^{+}}s^{2}\alpha(s)=0.

Then,

α(s){nkcot(k/ns)ifk>0,s<min{t,π/k/n},n/sifk=0,nkcoth(k/ns)ifk<0.\displaystyle\alpha(s)\leqslant\begin{cases}\sqrt{nk}\cot(\sqrt{k/n}s)&\text{if}\ k>0,s<\min\{t,\pi/\sqrt{k/n}\},\\ n/s&\text{if}\ k=0,\\ \sqrt{-nk}\operatorname{coth}(\sqrt{-k/n}s)&\text{if}\ k<0.\end{cases}

for any s(0,t)s\in(0,t).

Proof.

Set

Sn,k(s):={n/ksin(k/ns)ifk>0,sifk=0,n/ksinh(k/ns)ifk<0.\displaystyle S_{n,k}(s):=\begin{cases}\sqrt{n/k}\sin(\sqrt{k/n}s)&\text{if}\ k>0,\\ s&\text{if}\ k=0,\\ \sqrt{-n/k}\operatorname{sinh}(\sqrt{-k/n}s)&\text{if}\ k<0.\end{cases}

then, β(t):=nS˙n,k(s)/Sn,k(s)\beta(t):=n\dot{S}_{n,k}(s)/S_{n,k}(s) solves the Riccati equation

α˙(s)+1nα2(s)+k=0,s(0,t).\displaystyle\dot{\alpha}(s)+\frac{1}{n}\alpha^{2}(s)+k=0,\quad s\in(0,t).

Inspired by (3.10) in [29], we have

ddt(Sn,k2(αβ))=\displaystyle\frac{d}{dt}(S_{n,k}^{2}(\alpha-\beta))=  2Sn,kS˙n,k(αβ)+Sn,k2(α˙β˙)\displaystyle\,2S_{n,k}\dot{S}_{n,k}(\alpha-\beta)+S_{n,k}^{2}(\dot{\alpha}-\dot{\beta})
\displaystyle\leqslant  2Sn,kS˙n,k(αβ)+Sn,k2(1nα2k+1nβ2+k)\displaystyle\,2S_{n,k}\dot{S}_{n,k}(\alpha-\beta)+S_{n,k}^{2}(-\frac{1}{n}\alpha^{2}-k+\frac{1}{n}\beta^{2}+k)
=\displaystyle=  2Sn,kS˙n,k(αβ)+1nSn,k2(β2α2)\displaystyle\,2S_{n,k}\dot{S}_{n,k}(\alpha-\beta)+\frac{1}{n}S_{n,k}^{2}(\beta^{2}-\alpha^{2})
=\displaystyle= 2nSn,k2β(αβ)+1nSn,k2(β2α2)\displaystyle\,\frac{2}{n}S_{n,k}^{2}\beta(\alpha-\beta)+\frac{1}{n}S_{n,k}^{2}(\beta^{2}-\alpha^{2})
=\displaystyle= 1nSn,k2(βα)20.\displaystyle\,-\frac{1}{n}S_{n,k}^{2}(\beta-\alpha)^{2}\leqslant 0.

Together with the condition that lims0+Sn,k2α=lims0+Sn,k2β=0\lim_{s\to 0^{+}}S_{n,k}^{2}\alpha=\lim_{s\to 0^{+}}S_{n,k}^{2}\beta=0, we have α(s)β(s)\alpha(s)\leqslant\beta(s) for s(0,t)s\in(0,t). ∎

4. Main Results

In this section, we show certain rigidity results for the weak KAM solutions and Aubry sets under certain curvature hypothesis. Moreover, we give some applications to the Mather quotient and Mañé’s Lagrangian.

Now, we recall the notion of Laplacian of a continuous function in the barrier sense.

Definition 4.1.

Let f:Mf:M\to\mathbb{R} be a continuous function on a Riemannian manifold (M,g)(M,g).

  1. (1)

    A C2C^{2} function f^:M\hat{f}:M\to\mathbb{R} is said to be a support function from above of ff at pMp\in M if f^(p)=f(p)\hat{f}(p)=f(p) and f^(x)f(x)\hat{f}(x)\geqslant f(x) in some neighborhood of pp.

  2. (2)

    We say Δf(p)B\Delta f(p)\leqslant B\in\mathbb{R} in the barrier sense if for every ϵ>0\epsilon>0, one can find a C2C^{2} support function fϵf_{\epsilon} from above of ff at pp such that Δfϵ(p)B+ϵ\Delta f_{\epsilon}(p)\leqslant B+\epsilon.

  3. (3)

    A continuous function f:Mf:M\to\mathbb{R} is said to be superharmonic if Δf(p)0\Delta f(p)\leqslant 0 in the barrier sense for each pMp\in M.

  4. (4)

    Similarly, we say that a continuous function f:Mf:M\to\mathbb{R} is subharmonic if f-f is superharmonic.

The following maximal principle was proved by Calabi in [5]. A fundamental proof can be aslo found in [13].

Theorem 4.2.

If f:Mf:M\to\mathbb{R} is a superharmonic function, then ff is constant in a neighborhood of every local minimum. In particular, ff is constant if ff has a global minimum.

Lemma 4.3.

Let L:TML:TM\to\mathbb{R} be a Tonelli Lagrangian on a closed connected Riemannian manifold (M,g)(M,g) and let uu be a weak KAM solution of (HJ). For any point xx and any (u,L,c[L])(u,L,c[L])-calibrated curve ρ:(,0]M\rho:(-\infty,0]\to M ending at xx, the function

u(ρ(t))+At(ρ(t),)+c[L]t\displaystyle u(\rho(-t))+A_{t}(\rho(-t),\cdot)+c[L]t

defined on MM is a support function from above of uu at xx for any t>0t>0.

Proof.

Since ρ\rho is a (u,L,c[L])(u,L,c[L])-calibrated curve, we have

u(x)=Ttu(x)+c[L]t=u(ρ(t))+At(ρ(t),x)+c[L]t\displaystyle u(x)=T_{t}^{-}u(x)+c[L]t=u(\rho(-t))+A_{t}(\rho(-t),x)+c[L]t

for any t>0t>0. In addition,

u(y)=Ttu(y)+c[L]tu(ρ(t))+At(ρ(t),y)+c[L]t,yM.\displaystyle u(y)=T_{t}^{-}u(y)+c[L]t\leqslant u(\rho(-t))+A_{t}(\rho(-t),y)+c[L]t,\qquad\forall y\in M.

By Proposition B.4, (x.ρ˙(0))(x.\dot{\rho}(0)) is not conjugate to (ρ(t),ρ˙(t))(\rho(-t),\dot{\rho}(-t)) and Theorem 2.13 ensures that the function u^()=u(ρ(t))+At(ρ(t),)+c[L]t\hat{u}(\cdot)=u(\rho(-t))+A_{t}(\rho(-t),\cdot)+c[L]t is of C2C^{2} in a neighborhood of xx. Thus, u^\hat{u} is a support function from above of uu at xx. ∎

Lemma 4.4.

Assume that L:TML:TM\to\mathbb{R} is a Tonelli Lagrangian on a closed connected Riemannian manifold (M,g).(M,g). If ρ:[0,t]M\rho:[0,t]\to M is a minimal curve for At(ρ(0),ρ(t))A_{t}(\rho(0),\rho(t)) satisfying |ρ˙|<K|\dot{\rho}|<K for some constant K>0K>0, then there exist τ>0\tau>0 and C1,C2>0C_{1},C_{2}>0 such that

|ΔyAs(ρ(0),ρ(s))|C1s+C2\displaystyle|\Delta_{y}A_{s}(\rho(0),\rho(s))|\leqslant\frac{C_{1}}{s}+C_{2}

for all s(0,τ)s\in(0,\tau), with constants τ\tau, C1C_{1} and C2C_{2} depend only on KK.

Proof.

Choose a coordinate chart (U,ϕ)(U,\phi) of ρ(0)\rho(0), then one can find τ1>0\tau^{1}>0 so that ρ(s)U\rho(s)\in U for all s(0,τ1)s\in(0,\tau^{1}). This allows us to reduce to the case when M=nM=\mathbb{R}^{n}.

Note that Lv(ρ,ρ˙)L_{v}(\rho,\dot{\rho}) is bounded. Applying Lemma 2.4 in [2] one can find τ^=τ^(K)(0,τ1)\hat{\tau}=\hat{\tau}(K)\in(0,\tau^{1}) and C^(K)>0\hat{C}(K)>0 such that

|dy2As(ρ(0),ρ(s))|C^Ks,s(0,τ^)\displaystyle|d^{2}_{y}A_{s}(\rho(0),\rho(s))|\leqslant\frac{\hat{C}_{K}}{s},\qquad\forall s\in(0,\hat{\tau})

Recall that in local chart the Laplacian operator has the representation

Δu=1detgxi(gijdetguxj).\displaystyle\Delta u=\frac{1}{\sqrt{\det g}}\frac{\partial}{\partial x^{i}}\big{(}g^{ij}\sqrt{\det g}\frac{\partial u}{\partial x^{j}}\big{)}.

Together with the boundedness of dyAs(ρ(0),ρ(s))=Lv(ρ(s),ρ˙(s))d_{y}A_{s}(\rho(0),\rho(s))=L_{v}(\rho(s),\dot{\rho}(s)), there are τ(0,τ^)\tau\in(0,\hat{\tau}) and C1C^K,C2>0C_{1}\geqslant\hat{C}_{K},C_{2}>0 such that

|ΔyAs(ρ(0),ρ(s))|C1s+C2,s(0,τ),\displaystyle|\Delta_{y}A_{s}(\rho(0),\rho(s))|\leqslant\frac{C_{1}}{s}+C_{2},\qquad\forall s\in(0,\tau),

with constants τ,C1,C2\tau,C_{1},C_{2} depending only on KK. ∎

Proof of Theorem 1.5.

For all xMx\in M and each (u,L,c[L])(u,L,c[L])-calibrated curve ρ:(,0]M\rho:(-\infty,0]\to M ending at x,x, we claim that ΔyAt(ρ(t),x)nkcoth(k/nt)\Delta_{y}A_{t}(\rho(-t),x)\leqslant\sqrt{-nk}\operatorname{coth}(\sqrt{-k/n}t) for every t>0t>0.

Notice that

ρ~(s)=ρ(st),s[0,t]\displaystyle\tilde{\rho}(s)=\rho(s-t),\qquad s\in[0,t]

is a minimal curve for At(ρ~(0),ρ~(t))A_{t}(\tilde{\rho}(0),\tilde{\rho}(t)). By Theorem 2.13, (ρ~(t),ρ~˙(t))(\tilde{\rho}(t),\dot{\tilde{\rho}}(t)) is not conjugate to (ρ~(0),ρ~˙(0))(\tilde{\rho}(0),\dot{\tilde{\rho}}(0)). Applying Theorem 3.1 we obtain

Θ˙(s)+1nΘ2(s)+Ric(ρ~˙(s))+Δf(ρ~(s))0,s(0,t],\displaystyle\dot{\Theta}(s)+\frac{1}{n}\Theta^{2}(s)+\operatorname{Ric}(\dot{\tilde{\rho}}(s))+\Delta f(\tilde{\rho}(s))\leqslant 0,\qquad s\in(0,t],

where Θ(s)=ΔyAs(ρ~(0),ρ~(s))\Theta(s)=\Delta_{y}A_{s}(\tilde{\rho}(0),\tilde{\rho}(s)). Since MM has nonnegative Ricci curvature and Δfk\Delta f\geqslant k, we arrive at

Θ˙(s)+1nΘ2(s)+kΘ˙(s)+1nΘ2(s)+Δf(ρ~(s))0,s(0,t].\displaystyle\dot{\Theta}(s)+\frac{1}{n}\Theta^{2}(s)+k\leqslant\dot{\Theta}(s)+\frac{1}{n}\Theta^{2}(s)+\Delta f(\tilde{\rho}(s))\leqslant 0,\qquad s\in(0,t].

Using that lims0+s2Θ=0\lim_{s\to 0^{+}}s^{2}\Theta=0 (see Lemma 4.4) together with Lemma 3.2 we discover

Θ(t)nkcoth(k/nt),t(0,+).\displaystyle\Theta(t)\leqslant\sqrt{-nk}\operatorname{coth}(\sqrt{-k/n}t),\qquad t\in(0,+\infty).

From the claim we deduce that

Δu(x)nk\displaystyle\Delta u(x)\leqslant\sqrt{-nk}

in the barrier sense. ∎

Proof of Theorem 1.6 .

For each point xMx\in M, there exists a (u,L,c[L])(u,L,c[L])-calibrated curve ρ:(,0]M\rho:(-\infty,0]\to M such that

ρ(0)=x,H(x,Lv(ρ(0),ρ˙(0))=c[L],ρ(s)=πΦsL(ρ(0),ρ˙(0)).\displaystyle\rho(0)=x,\ H(x,L_{v}(\rho(0),\dot{\rho}(0))=c[L],\ \rho(s)=\pi\circ\Phi_{s}^{L}(\rho(0),\dot{\rho}(0)).

Then, for each t>0t>0, the curve

ρ~(s)=ρ(st),s[0,t],\displaystyle\tilde{\rho}(s)=\rho(s-t),\qquad s\in[0,t],

is a minimal curve for At(ρ~(0),ρ~(t))A_{t}(\tilde{\rho}(0),\tilde{\rho}(t)) and (ρ~(t),ρ~˙(t))(\tilde{\rho}(t),\dot{\tilde{\rho}}(t)) is not conjugate to (ρ~(0),ρ~˙(0))(\tilde{\rho}(0),\dot{\tilde{\rho}}(0)). By Theorem 3.1 we have that

Θ˙(s)+1nΘ2(s)+Ric(ρ~˙(s))+Δf(ρ~(s))0,s(0,t],\displaystyle\dot{\Theta}(s)+\frac{1}{n}\Theta^{2}(s)+\operatorname{Ric}(\dot{\tilde{\rho}}(s))+\Delta f(\tilde{\rho}(s))\leqslant 0,\qquad s\in(0,t],

where Θ(s)=ΔyAs(ρ~(0),ρ~(s))+divω#(ρ~(s))\Theta(s)=\Delta_{y}A_{s}(\tilde{\rho}(0),\tilde{\rho}(s))+\operatorname{div}\omega^{\#}(\tilde{\rho}(s)).

Invoking Lemma 3.2 and the fact lims0+s2Θ=0\lim_{s\rightarrow 0^{+}}s^{2}\Theta=0 (see Lemma 4.4) we obtain

Θ(t)=ΔyAt(ρ~(0),ρ~(t))+divω#(x)=ΔyAt(ρ(t),x)+divω#(x)n/t,t>0.\displaystyle\Theta(t)=\Delta_{y}A_{t}(\tilde{\rho}(0),\tilde{\rho}(t))+\operatorname{div}\omega^{\#}(x)=\Delta_{y}A_{t}(\rho(-t),x)+\operatorname{div}\omega^{\#}(x)\leqslant n/t,\qquad\forall t>0.

Hence, Δu(x)divω#(x)\Delta u(x)\leqslant-\operatorname{div}\omega^{\#}(x) in the barrier sense. ∎

Proof of Theorem 1.7.

We first prove the sufficiency. If ω\omega is a harmonic 1-form, then divω#0\operatorname{div}\omega^{\#}\equiv 0 by Proposition 2.2 (2). Since (M,g)(M,g) has nonnegative Ricci curvature, by Theorem 1.6 each weak KAM solution uu of (HJ) is superharmonic. Hence, Theorem 4.2 implies that uu must be constant.

Now we turn to prove the necessity. If each weak KAM solution uu of (HJ) is constant, Theorem 1.6 shows that 0divω#(x)0\leqslant-\operatorname{div}\omega^{\#}(x) for all xMx\in M. By Stoke’s Theorem we can obtain that divω#0\operatorname{div}\omega^{\#}\equiv 0. Hence, ω\omega is a harmonic 1-form by Proposition 2.2 (2). ∎

Remark 4.5.

If (M,g)(M,g) is the flat tours 𝕋n\mathbb{T}^{n} with dimension n2n\geqslant 2, then the de Rham Cohomology H1(𝕋n,)=nH^{1}(\mathbb{T}^{n},\mathbb{R})=\mathbb{R}^{n}. The authors in [9] proved that each weak KAM solution associated to the Lagrangian

L(x,v)=12v,v2ω(v),[ω]H1(𝕋n,)=n,\displaystyle L(x,v)=\frac{1}{2}\langle v,v\rangle^{2}-\omega(v),\quad[\omega]\in H^{1}(\mathbb{T}^{n},\mathbb{R})=\mathbb{R}^{n},

must be constant (see section 5.5 in [9]).

Proof of Corollary 1.8.

By Proposition 2.2 (3), the inclusion

i:1(M,)\displaystyle i:\mathcal{H}^{1}(M,\mathbb{R}) H1(M,)\displaystyle\to H^{1}(M,\mathbb{R})
ω\displaystyle\omega [ω]\displaystyle\mapsto[\omega]

is an isomorphism. For each [ω]H1(M,)[\omega]\in H^{1}(M,\mathbb{R}), one can choose a representative element ω~\tilde{\omega} which is a harmonic 1-form. Thus, we can reduce to the case when ω1(M,)\omega\in\mathcal{H}^{1}(M,\mathbb{R}).

Now, we claim that h(x,y)0h(x,y)\equiv 0 for any (x,y)M×M(x,y)\in M\times M. Since

u(y)u(x)h(x,y),(x,y)M×M\displaystyle u(y)-u(x)\leqslant h(x,y),\qquad\forall(x,y)\in M\times M

for any weak KAM solution uu of (HJ). We observe that h(x,y)0h(x,y)\geqslant 0 since uu must be constant by Theorem 1.7. Now taking a point z𝒜(Lω)z\in\mathcal{A}(L_{\omega}) we have h(x,y)h(x,z)+h(z,y)h(x,y)\leqslant h(x,z)+h(z,y) for all x,yMx,y\in M. Notice that hz():=h(z,)h_{z}(\cdot):=h(z,\cdot) is a weak KAM solution of (HJ) and h(z,z)=0h(z,z)=0. It yields that hz()=h(z,)0h_{z}(\cdot)=h(z,\cdot)\equiv 0 by Theorem 1.7. In addition, because hz():=h(,z)h^{z}(\cdot):=h(\cdot,z) is a weak KAM solution of (HJ) associated to the Lagrangian

L˘(x,v):=L(x,v)=12gx(v,v)+ω(v)\breve{L}(x,v):=L(x,-v)=\frac{1}{2}g_{x}(v,v)+\omega(v)

and ω-\omega is also a harmonic 1-form, we know from Theorem 1.7 that hz()0.h^{z}(\cdot)\equiv 0. Therefore,

0h(x,y)h(x,z)+h(z,y)=0,(x,y)M×M.\displaystyle 0\leqslant h(x,y)\leqslant h(x,z)+h(z,y)=0,\quad\forall(x,y)\in M\times M.

The fact that h(x,y)=0h(x,y)=0 for every (x,y)M×M(x,y)\in M\times M implies that h(x,x)0h(x,x)\equiv 0 for all xMx\in M and δ(x,y)=h(x,y)+h(y,x)0\delta(x,y)=h(x,y)+h(y,x)\equiv 0 for any (x,y)M×M.(x,y)\in M\times M. This completes the proof. ∎

Now we recall the following result obtained in [16, Corollary 4.1.13].

Proposition 4.6.

Let L:TML:TM\to\mathbb{R} be a Tonelli Lagrangian on a closed connected Riemannian manifold (M,g).(M,g). Then, the Mather quotient (𝒜(L),,δ)(\mathcal{A}(L),\sim,\delta) is a singleton if and only if any two weak KAM solutions of (HJ) differ by a constant.

Proof of Theorem 1.9.

Indeed, the first part of the proof proof has the same reasoning as that of Theorem 1.7.

If ω\omega is a harmonic 1-form, then divω#(x)=0\operatorname{div}\omega^{\#}(x)=0 for all xMx\in M. Invoking Theorem 1.6, each weak KAM solution uu of (HJ) is superharmonic. Therefore, uu must be constant.

Conversely, if each weak KAM solution uu of (HJ) is constant, Theorem 1.6 implies that 0divω#(x)0\leqslant-\operatorname{div}\omega^{\#}(x) for all xMx\in M. Appying Stoke’s Theorem one can derive that divω#0\operatorname{div}\omega^{\#}\equiv 0. Thus ω\omega is a harmonic 1-form.

Notice that the constant 0 is a weak KAM solution of (HJ). Together with Proposition 4.6, we find that the Mather quotient (𝒜(L),,δ)(\mathcal{A}(L),\sim,\delta) is a singleton if and only if ω\omega is a harmonic 1-form.

Finally, we turn to the rest of the proof. Since (M,g)(M,g) has nonnegative Ricci curvature and ω\omega is a harmonic 1-form, by Proposition 2.2 (d) we obtain that f(x)=12gx(ω,ω)f(x)=\frac{1}{2}g_{x}(\omega^{\sharp},\omega^{\sharp}) is constant. This implies that Ric(v)+Δf(x)0\operatorname{Ric}(v)+\Delta f(x)\geqslant 0 for all (x,v)TM(x,v)\in TM. Hence, uu must be constant by the first part of the proof. ∎

Appendix A proofs of statements on conjugate points and Jacobi equation

Proof of Proposition 2.4.

If JJ is a Jacobi field along ρ\rho, then one can find a variation ρ(t,s)C2([a,b]×(ε,ε))\rho(t,s)\in C^{2}([a,b]\times(-\varepsilon,\varepsilon)) of extremal curves along ρ\rho such that

J(t)=s|s=0ρ(t,s).\displaystyle J(t)=\left.\frac{\partial}{\partial s}\right|_{s=0}\rho(t,s).

Thus we have

ddtLv(ρ(t,s),ρ˙(t,s))=Lx(ρ(t,s),ρ˙(t,s))\displaystyle\frac{d}{dt}L_{v}(\rho(t,s),\dot{\rho}(t,s))=L_{x}(\rho(t,s),\dot{\rho}(t,s))

in local chart. Taking partial derivative with respect to ss at s=0s=0, we get

ddt(Lvx(ρ(t),ρ˙(t))J(t)+Lvv(ρ(t),ρ˙(t))J˙(t))=Lxx(ρ(t),ρ˙(t))J(t)+Lxv(ρ(t),ρ˙(t))J˙(t).\displaystyle\frac{d}{dt}(L_{vx}(\rho(t),\dot{\rho}(t))J(t)+L_{vv}(\rho(t),\dot{\rho}(t))\dot{J}(t))=L_{xx}(\rho(t),\dot{\rho}(t))J(t)+L_{xv}(\rho(t),\dot{\rho}(t))\dot{J}(t).

Now we turn to prove the sufficiency. Suppose JJ solves 2.2 in local chart. We want to define a variation ρ(t,s)\rho(t,s) such that

J(t)=s|s=0ρ(t,s).\displaystyle J(t)=\left.\frac{\partial}{\partial s}\right|_{s=0}\rho(t,s).

For this, choose a smooth curve α:(ε,ε)M\alpha:(-\varepsilon,\varepsilon)\to M and a smooth vector field VV along α\alpha satisfying

α(0)=ρ(a),α˙(0)=J(a),\displaystyle\alpha(0)=\rho(a),\ \dot{\alpha}(0)=J(a),
V(0)=ρ˙(a),α˙V(0)=ρ˙J(a).\displaystyle V(0)=\dot{\rho}(a),\ \nabla_{\dot{\alpha}}V(0)=\nabla_{\dot{\rho}}J(a).

Now we define

ρ(t,s):=πΦtaL(α(s),V(s))C2([a,b]×(ε,ε)).\displaystyle\rho(t,s):=\pi\circ\Phi_{t-a}^{L}(\alpha(s),V(s))\in C^{2}([a,b]\times(-\varepsilon,\varepsilon)).

Then we have

t|t=aρ(t,s)=V(s),s|s=0ρ(a,s)=α˙(0).\displaystyle\left.\frac{\partial}{\partial t}\right|_{t=a}\rho(t,s)=V(s),\ \left.\frac{\partial}{\partial s}\right|_{s=0}\rho(a,s)=\dot{\alpha}(0).

Set

W(t):=s|s=0ρ(t,s).\displaystyle W(t):=\left.\frac{\partial}{\partial s}\right|_{s=0}\rho(t,s).

Note that

W(a)=s|s=0ρ(a,s)=α˙(0)=J(a)\displaystyle W(a)=\left.\frac{\partial}{\partial s}\right|_{s=0}\rho(a,s)=\dot{\alpha}(0)=J(a)

and J,WJ,W solves 2.2 in local coordinate systems. If we can show that ρ˙W(a)=ρ˙J(a)\nabla_{\dot{\rho}}W(a)=\nabla_{\dot{\rho}}J(a), it then follows from Cauchy-Lipschitz theorem that J(t)=W(t)J(t)=W(t). Indeed, we have

ρ˙W(a)=tρsρ(a,0)=sρtρ(a,0)=α˙V(0)=ρ˙J(a).\displaystyle\nabla_{\dot{\rho}}W(a)=\nabla_{\partial_{t}\rho}\partial_{s}\rho(a,0)=\nabla_{\partial_{s}\rho}\partial_{t}\rho(a,0)=\nabla_{\dot{\alpha}}V(0)=\nabla_{\dot{\rho}}J(a).

It follows that WJW\equiv J as claimed. ∎

Proof of Proposition 2.5.

Consider the C2C^{2} variation

ρ(s,z)=πΦsaL(ρ(a),ρ˙(a)+zρ˙(a)J)C2([a,b]×(ε,ε)).\displaystyle\rho(s,z)=\pi\circ\Phi_{s-a}^{L}(\rho(a),\dot{\rho}(a)+z\nabla_{\dot{\rho}(a)}J)\in C^{2}([a,b]\times(-\varepsilon,\varepsilon)).

We have

z|z=0ρ(a,z)=z|z=0ρ(a)=0,sρzρ(a,0)=zρsρ(a,0)=z|z=0ρ˙(a)+zρ˙(a)J=ρ˙(a)J.\begin{gathered}\left.\partial_{z}\right|_{z=0}\rho(a,z)=\left.\partial_{z}\right|_{z=0}\rho(a)=0,\\ \nabla_{\partial_{s}\rho}\partial_{z}\rho(a,0)=\nabla_{\partial_{z}\rho}\partial_{s}\rho(a,0)=\left.\partial_{z}\right|_{z=0}\dot{\rho}(a)+z\nabla_{\dot{\rho}(a)}J=\nabla_{\dot{\rho}(a)}J.\end{gathered}

Notice that z|z=0ρ(s,z)\left.\partial_{z}\right|_{z=0}\rho(s,z) solves the Jacobi equation (2.2) in local chart and

z|z=0ρ(a,z)=J(a)=0,\displaystyle\,\left.\partial_{z}\right|_{z=0}\rho(a,z)=J(a)=0,
sρzρ(a,0)=ρ˙(a)J.\displaystyle\,\nabla_{\partial_{s}\rho}\partial_{z}\rho(a,0)=\nabla_{\dot{\rho}(a)}J.

We obtain that

J(s)=z|z=0ρ(s,z),s[a,b]\displaystyle J(s)=\left.\partial_{z}\right|_{z=0}\rho(s,z),\quad s\in[a,b]

by Cauchy-Lipschitz theorem. ∎

Proof of Proposition 2.7.

We only need to prove the necessity since its sufficiency can be proved similarly.

Suppose that (ρ(b),ρ˙(b))(\rho(b),\dot{\rho}(b)) is conjugate to (ρ(b),ρ˙(b))(\rho(b),\dot{\rho}(b)) with respect to LL, then there exists a nonzero Jacobi field JJ along ρ\rho such that ρ(a)=ρ(b)=0\rho(a)=\rho(b)=0. Set ρ˘(t)=ρ(a+bt)\breve{\rho}(t)=\rho(a+b-t), t[a,b]t\in[a,b]. Then (ρ˘(t),ρ˘˙(t))(\breve{\rho}(t),\dot{\breve{\rho}}(t)) is a trajectory of the Euler-Lagrange flow associated to L˘\breve{L}.

Since J:[a,b]TMJ:[a,b]\to TM is a Jacobi field along ρ(t)\rho(t) with respect to LL, one has that J˘(t):=J(a+bt)\breve{J}(t):=J(a+b-t) is a Jacobi field along ρ˘(t)\breve{\rho}(t) with respect to L˘\breve{L}. Indeed, by a direct computation we have

ddt(L˘vx(ρ˘(t),ρ˘˙(t))J˘(t)+L˘vv(ρ˘(t),ρ˘˙(t))J˘˙(t))=L˘xx(ρ˘(t),ρ˘˙(t))J˘(t)+L˘xv(ρ˘(t),ρ˘˙(t))J˘˙(t)\displaystyle\frac{d}{dt}(\breve{L}_{vx}(\breve{\rho}(t),\dot{\breve{\rho}}(t))\breve{J}(t)+\breve{L}_{vv}(\breve{\rho}(t),\dot{\breve{\rho}}(t))\dot{\breve{J}}(t))=\breve{L}_{xx}(\breve{\rho}(t),\dot{\breve{\rho}}(t))\breve{J}(t)+\breve{L}_{xv}(\breve{\rho}(t),\dot{\breve{\rho}}(t))\dot{\breve{J}}(t)

in local chart. This implies that there is a nonzero Jacobi field J˘\breve{J} along ρ˘\breve{\rho} with respect to L˘\breve{L} satisfying ρ˘(a)=ρ˘(b)=0\breve{\rho}(a)=\breve{\rho}(b)=0. It follows that (ρ˘(a),ρ˘˙(a))(\breve{\rho}(a),\dot{\breve{\rho}}(a)) is conjugate to (ρ˘(b),ρ˘˙(b))(\breve{\rho}(b),\dot{\breve{\rho}}(b)) with respect to L˘\breve{L}. In other words, (ρ(a),ρ˙(a))(\rho(a),-\dot{\rho}(a)) is conjugate to (ρ(b),ρ˙(b))(\rho(b),-\dot{\rho}(b)) with respect to L˘\breve{L}. ∎

Proof of Proposition 2.8.

For any variation ρ(s,z)C([0,t]×(ε,ε)M)\rho(s,z)\in C^{\infty}([0,t]\times(-\varepsilon,\varepsilon)\to M) of ρ(s)\rho(s) such that

ρ(s,0)=ρ(s),ρ(0,z)ρ(0),ρ(t,z)ρ(t),\displaystyle\rho(s,0)=\rho(s),\ \rho(0,z)\equiv\rho(0),\ \rho(t,z)\equiv\rho(t),

we have

ddz|z=00tL(ρ,sρ)+ω(sρ)ds=ddz|z=00tL(ρ,sρ)𝑑s,\displaystyle\left.\frac{d}{dz}\right|_{z=0}\int_{0}^{t}L(\rho,\partial_{s}\rho)+\omega(\partial_{s}\rho)ds=\left.\frac{d}{dz}\right|_{z=0}\int_{0}^{t}L(\rho,\partial_{s}\rho)ds,

since ω\omega is closed. Therefore,

0=\displaystyle 0= ddz|z=00tL(ρ,sρ)+ω(sρ)ds=ddz|z=00tL(ρ,sρ)𝑑s\displaystyle\,\left.\frac{d}{dz}\right|_{z=0}\int_{0}^{t}L(\rho,\partial_{s}\rho)+\omega(\partial_{s}\rho)ds=\left.\frac{d}{dz}\right|_{z=0}\int_{0}^{t}L(\rho,\partial_{s}\rho)ds
=\displaystyle= ddz|z=00t12g(sρ(s,z),sρ(s,z))f(ρ(s,z))ds\displaystyle\,\left.\frac{d}{dz}\right|_{z=0}\int_{0}^{t}\frac{1}{2}g(\partial_{s}\rho(s,z),\partial_{s}\rho(s,z))-f(\rho(s,z))ds
=\displaystyle= 0tg(zρsρ|z=0,ρ˙)g(f(ρ),zρ|z=0)ds\displaystyle\,\int_{0}^{t}g(\left.\nabla_{\partial_{z}\rho}\partial_{s}\rho\right|_{z=0},\dot{\rho})-g(\nabla f(\rho),\left.\partial_{z}\rho\right|_{z=0})ds
=\displaystyle= 0tg(sρzρ|z=0,ρ˙)g(f(ρ),zρ|z=0)ds\displaystyle\,\int_{0}^{t}g(\left.\nabla_{\partial_{s}\rho}\partial_{z}\rho\right|_{z=0},\dot{\rho})-g(\nabla f(\rho),\left.\partial_{z}\rho\right|_{z=0})ds
=\displaystyle= 0tddsg(zρ|z=0,ρ˙)g(zρ|z=0,ρ˙ρ˙)g(f(ρ),zρ|z=0)ds\displaystyle\,\int_{0}^{t}\frac{d}{ds}g(\left.\partial_{z}\rho\right|_{z=0},\dot{\rho})-g(\left.\partial_{z}\rho\right|_{z=0},\nabla_{\dot{\rho}}\dot{\rho})-g(\nabla f(\rho),\left.\partial_{z}\rho\right|_{z=0})ds
=\displaystyle= 0tg(f(ρ)ρ˙ρ˙,zρ|z=0)𝑑s.\displaystyle\,\int_{0}^{t}g(-\nabla f(\rho)-\nabla_{\dot{\rho}}\dot{\rho},\left.\partial_{z}\rho\right|_{z=0})ds.

which yields that

ρ˙ρ˙=f.\displaystyle\nabla_{\dot{\rho}}\dot{\rho}=-\nabla f.

Suppose ρ(s,z)C2([a,b]×(ε,ε),M)\rho(s,z)\in C^{2}([a,b]\times(-\varepsilon,\varepsilon),M) is a variation such that ρ(,z)\rho(\cdot,z), z(ε,ε)z\in(-\varepsilon,\varepsilon) satisfy (2.3). Then we have

sρsρzρ=\displaystyle\nabla_{\partial_{s}\rho}\nabla_{\partial_{s}\rho}\partial_{z}\rho= sρzρsρ\displaystyle\,\nabla_{\partial_{s}\rho}\nabla_{\partial_{z}\rho}\partial_{s}\rho
=\displaystyle= zρsρsρ+R(sρ,zρ)sρ\displaystyle\,\nabla_{\partial_{z}\rho}\nabla_{\partial_{s}\rho}\partial_{s}\rho+R(\partial_{s}\rho,\partial_{z}\rho)\partial_{s}\rho
=\displaystyle= zρf+R(sρ,zρ)sρ\displaystyle\,-\nabla_{\partial_{z}\rho}\nabla f+R(\partial_{s}\rho,\partial_{z}\rho)\partial_{s}\rho
=\displaystyle= Hessf(zρ)+R(sρ,zρ)sρ.\displaystyle\,-\operatorname{Hess}f(\partial_{z}\rho)+R(\partial_{s}\rho,\partial_{z}\rho)\partial_{s}\rho.

Taking z=0z=0 we obtain that

ρ˙ρ˙J+R(J,ρ˙)ρ˙+Hessf(J)=0,\displaystyle\nabla_{\dot{\rho}}\nabla_{\dot{\rho}}J+R(J,\dot{\rho})\dot{\rho}+\operatorname{Hess}f(J)=0,

where J=zρ|z=0J=\left.\partial_{z}\rho\right|_{z=0}. ∎

Proof of Lemma 2.9.

Suppose J:[a,b]TMJ:[a,b]\to TM is a Jacobi field along ρ\rho with J(a)=0J(a)=0. Set (x,v)=(ρ(a),ρ˙(a))(x,v)=(\rho(a),\dot{\rho}(a)), w=ρ˙(a)Jw=\nabla_{\dot{\rho}(a)}J. Then we have

z|z=0πΦτaL(x,v+zw)=dv(πΦτaL)(w)=J(τ),τ[a,b].\displaystyle\left.\partial_{z}\right|_{z=0}\pi\circ\Phi_{\tau-a}^{L}(x,v+zw)=d_{v}(\pi\circ\Phi_{\tau-a}^{L})(w)=J(\tau),\qquad\tau\in[a,b].

Observe that JJ is nonzero if and only if w0w\neq 0. Therefore, (ρ(b),ρ˙(b))(\rho(b),\dot{\rho}(b)) is conjugate to (ρ(a),ρ˙(a))(\rho(a),\dot{\rho}(a)) if and only if

J(b)=dv(πΦbaL)(w)=0.\displaystyle J(b)=d_{v}(\pi\circ\Phi_{b-a}^{L})(w)=0.

i.e., if and only if dv(πΦbaL)d_{v}(\pi\circ\Phi_{b-a}^{L}) is degenerate. Hence, (ρ(b),ρ˙(b))(\rho(b),\dot{\rho}(b)) is not conjugate to (ρ(a),ρ˙(a))(\rho(a),\dot{\rho}(a)) if and only if dv(πΦbaL)d_{v}(\pi\circ\Phi_{b-a}^{L}) is non-degenerate. ∎

Proof of Proposition 2.10.

Since (ρ(t),ρ˙(t))(\rho(t),\dot{\rho}(t)) is not conjugate to (ρ(0),ρ˙(0))(\rho(0),\dot{\rho}(0)), we can find a neighborhood WW of ρ˙(0)\dot{\rho}(0) such that πΦtL|W\left.\pi\circ\Phi_{t}^{L}\right|_{W} is a C1C^{1} diffeomorphism.

Now we claim that there exists a neighborhood UU of ρ(t)\rho(t) satisfying the following property: if zUz\in U and ρz\rho_{z} is a minimizer of At(x,z)A_{t}(x,z), then we have ρ˙z(0)W\dot{\rho}_{z}(0)\in W. If no such neighborhood exists, we can find a sequence {zi}\{z_{i}\} converging to ρ(t)\rho(t) with the following property: for each ziz_{i} we can find a minimal curve ρzi\rho_{z_{i}} for At(x,zi)A_{t}(x,z_{i}) with ρ˙zi(0)W\dot{\rho}_{z_{i}}(0)\notin W. By (3) of Proposition 2.1 we can just suppose ρ˙zi(0)v1\dot{\rho}_{z_{i}}(0)\to v_{1} as i+i\to+\infty. By continuous dependence,

0tL(ΦsL(x,ρ˙zi(0))ds=At(x,zi)\displaystyle\int_{0}^{t}L(\Phi_{s}^{L}(x,\dot{\rho}_{z_{i}}(0))ds=A_{t}(x,z_{i})

converges to

0tL(ΦsL(x,v1))𝑑s=At(x,ρ(t))\displaystyle\int_{0}^{t}L(\Phi_{s}^{L}(x,v_{1}))ds=A_{t}(x,\rho(t))

as i+i\to+\infty.

This implies that ΦsL(x,v1):[0,t]M\Phi_{s}^{L}(x,v_{1}):[0,t]\to M is a minimal curve for At(x,ρ(t))A_{t}(x,\rho(t)). Hence, v1=ρ˙(0)v_{1}=\dot{\rho}(0) since At(x,)A_{t}(x,\cdot) is differentiable at ρ(t)\rho(t). But this contradicts the assumption that ρ˙zi(0)W\dot{\rho}_{z_{i}}(0)\notin W. From this claim we have that for each zUz\in U,

At(x,z)=0tL(ΦsL(x,vz))𝑑s,\displaystyle A_{t}(x,z)=\int_{0}^{t}L(\Phi_{s}^{L}(x,v_{z}))ds,

where vzv_{z} is uniquely determined by the condition vzWv_{z}\in W, πΦtL(x,vz)=z\pi\circ\Phi_{t}^{L}(x,v_{z})=z. In addition, we have dyAt(x,z)=Lv(ΦtL(x,vz))d_{y}A_{t}(x,z)=L_{v}(\Phi_{t}^{L}(x,v_{z})) by Proposition 2.1. Therefore. At(x,)A_{t}(x,\cdot) is of CrC^{r} in UU since UzvzU\ni z\mapsto v_{z} is a Cr1C^{r-1} diffeomorphism. ∎

Proof of Lemma 2.12.

Let ρ~(s)=πΦsL(ρ(0),ρ˙(0))\tilde{\rho}(s)=\pi\circ\Phi_{s}^{L}(\rho(0),\dot{\rho}(0)), s[0,+)s\in[0,+\infty). Take a minimal curve ρτ:[0,t+τ]M\rho_{\tau}:[0,t+\tau]\to M for At+τ(x,ρ~(t+τ))A_{t+\tau}(x,\tilde{\rho}(t+\tau)) for each fixed τ>0\tau>0. By (3) of Proposition 2.1 we assume that ρ˙τi(0)\dot{\rho}_{\tau_{i}}(0) converges to vv as i+i\to+\infty where {τi}\{\tau_{i}\} is a positive sequence converging to 0 .

By continuous dependence,

0t+τiL(ΦsL(ρτi(0),ρ˙τi(0)))𝑑s=At+τi(x,ρ~(t+τ))\displaystyle\int_{0}^{t+\tau_{i}}L(\Phi_{s}^{L}(\rho_{\tau_{i}}(0),\dot{\rho}_{\tau_{i}}(0)))ds=A_{t+\tau_{i}}(x,\tilde{\rho}(t+\tau))

converges to

0tL(ΦsL(x,v)ds=At(x,y)\displaystyle\int_{0}^{t}L(\Phi_{s}^{L}(x,v)ds=A_{t}(x,y)

as i+i\to+\infty.

Then, we can find another minimal curve ρ~(s)=πΦsL(x,v)\tilde{\rho}(s)=\pi\circ\Phi_{s}^{L}(x,v), s[0,t]s\in[0,t], forAt(x,y)A_{t}(x,y) if vρ˙(0)v\neq\dot{\rho}(0). If ρ˙(0)=v\dot{\rho}(0)=v, we must show that (ρ(t),ρ˙(t))(\rho(t),\dot{\rho}(t)) is conjugate to (ρ(0),ρ˙(0))(\rho(0),\dot{\rho}(0)). Otherwise, we suppose that (ρ(t),ρ˙(t))(\rho(t),\dot{\rho}(t)) is not conjugate to (ρ(0),ρ˙(0))(\rho(0),\dot{\rho}(0)). Then, dρ˙(0)(πΦtL)d_{\dot{\rho}(0)}(\pi\circ\Phi_{t}^{L}) is non-degenerate. Hence, there exists an open neighborhood (tϵ,t+ϵ)×W(t-\epsilon,t+\epsilon)\times W of (t,ρ˙(0))(t,\dot{\rho}(0)) such that dw(πΦrL)d_{w}(\pi\circ\Phi_{r}^{L}) is non-degenerate for every (r,w)(tϵ,t+ϵ)×W(r,w)\in(t-\epsilon,t+\epsilon)\times W. By constant rank theorem we can know that πΦrL:WM\pi\circ\Phi_{r}^{L}:W\to M is a Cr1C^{r-1} diffeomorphism for all r(tϵ,t+ϵ)r\in(t-\epsilon,t+\epsilon).

Together with

πΦt+τiL(x,ρ˙τi(0))=ρ~(t+τi)=πΦt+τiL(x,ρ˙(0)),\displaystyle\pi\circ\Phi_{t+\tau_{i}}^{L}(x,\dot{\rho}_{\tau_{i}}(0))=\tilde{\rho}(t+\tau_{i})=\pi\circ\Phi_{t+\tau_{i}}^{L}(x,\dot{\rho}(0)),

we obtain that ρ˙(0)=ρ˙τi(0)\dot{\rho}(0)=\dot{\rho}_{\tau_{i}}(0) for τi(0,ϵ)\tau_{i}\in(0,\epsilon). This means that ρ~|[0,t+τi]\tilde{\rho}_{\left.\right|_{[0,t+\tau_{i}]}} is a minimal curve for At+τi(x,ρ~(t+τi))A_{t+\tau_{i}}(x,\tilde{\rho}(t+\tau_{i})) which leads to a contradiction. ∎

Appendix B Index Form

It is well known that any two points in the interior of a minimal curve ρ:[0,t]M\rho:[0,t]\to M for At(x,y)A_{t}(x,y), with ρ(0)=x\rho(0)=x and ρ(t)=y\rho(t)=y, are not conjugate to each other. This is also true for a point γ(t)\gamma(t) with t(0,t)t\in(0,t) and a endpoint x=γ(0)x=\gamma(0) or y=γ(t)y=\gamma(t) even if xx and yy is conjugate to each other (see, for instance, [8, Corollary 4.2]). We still give a detailed proof of the form of statement we need in Lagrangian scheme. For the following notion of index form, see [12].

Definition B.1.

Suppose L:TML:TM\rightarrow\mathbb{R} is a Tonelli Lagrangian defined on an open subset MM of n\mathbb{R}^{n}. Let ρ:[a,b]M\rho:[a,b]\rightarrow M be a C1C^{1} curve satisfying (E-L). For any two continuous piecewise C2C^{2} vector fields η,θ\eta,\theta along ρ\rho, define the index form I(η,θ)I(\eta,\theta) by

I(η,θ)=abη˙Lvvθ˙+η˙Lvxθ+ηLxvθ˙+ηLxxθdt.\displaystyle I(\eta,\theta)=\int_{a}^{b}\dot{\eta}^{\top}L_{vv}\dot{\theta}+\dot{\eta}^{\top}L_{vx}\theta+\eta^{\top}L_{xv}\dot{\theta}+\eta^{\top}L_{xx}\theta dt.
Lemma B.2.

Assume that L:TML:TM\rightarrow\mathbb{R} is a Tonelli Lagrangian defined on an open subset MM of n\mathbb{R}^{n}. Let ρ:[a,b]M\rho:[a,b]\rightarrow M be a C1C^{1} curve satisfying (E-L) and let η,θ\eta,\theta are continuous piecewise C2C^{2} vector fields along ρ\rho. For any partition a=t0<t1<<tk+1=ba=t_{0}<t_{1}<\cdots<t_{k+1}=b such that the vector fields η,θ\eta,\theta along ρ\rho are of C2C^{2} on [ti,ti+1],i=1,2,k[t_{i},t_{i+1}],i=1,2,\cdots k, we have

I(η,θ)=i=1k+1(η˙Lvv+ηLxv)θ|ti+ti+1\displaystyle I(\eta,\theta)=\sum_{i=1}^{k+1}\left.(\dot{\eta}^{\top}L_{vv}+\eta^{\top}L_{xv})\theta\right|_{t_{i}^{+}}^{t_{i+1}^{-}}

where η|[ti,ti+1]\eta|_{[t_{i},t_{i+1}]} are Jacobi fields along ρ\rho.

Proof.

Since η|[ti,ti+1]\eta|_{[t_{i},t_{i+1}]} are Jacobi fields along ρ\rho, we have

ddt(Lvxη+Lvvη˙)=Lxxη+Lxvη˙,t[ti,ti+1].\displaystyle\frac{d}{dt}(L_{vx}\eta+L_{vv}\dot{\eta})=L_{xx}\eta+L_{xv}\dot{\eta},\quad t\in[t_{i},t_{i+1}].

Therefore

I(η,θ)=\displaystyle I(\eta,\theta)= abη˙Lvvθ˙+η˙Lvxθ+ηLxvθ˙+ηLxxθdt\displaystyle\int_{a}^{b}\dot{\eta}^{\top}L_{vv}\dot{\theta}+\dot{\eta}^{\top}L_{vx}\theta+\eta^{\top}L_{xv}\dot{\theta}+\eta^{\top}L_{xx}\theta dt
=\displaystyle= ab(η˙Lvv+ηLxv)θ˙+(η˙Lvx+ηLxx)θdt\displaystyle\int_{a}^{b}(\dot{\eta}^{\top}L_{vv}+\eta^{\top}L_{xv})\dot{\theta}+(\dot{\eta}^{\top}L_{vx}+\eta^{\top}L_{xx})\theta dt
=\displaystyle= abddt(η˙Lvv+ηLxv)θ+(η˙Lvx+ηLxx)θdt\displaystyle\int_{a}^{b}-\frac{d}{dt}(\dot{\eta}^{\top}L_{vv}+\eta^{\top}L_{xv})\theta+(\dot{\eta}^{\top}L_{vx}+\eta^{\top}L_{xx})\theta dt
+i=1k(η˙Lvv+ηLxv)θ|ti+ti+1\displaystyle+\sum_{i=1}^{k}\left.(\dot{\eta}^{\top}L_{vv}+\eta^{\top}L_{xv})\theta\right|_{t_{i}^{+}}^{t_{i+1}^{-}}
=\displaystyle= i=1k(η˙Lvv+ηLxv)θ|ti+ti+1.\displaystyle\sum_{i=1}^{k}\left.(\dot{\eta}^{\top}L_{vv}+\eta^{\top}L_{xv})\theta\right|_{t_{i}^{+}}^{t_{i+1}^{-}}.

This completes the proof. ∎

Proposition B.3.

Let L:TML:TM\rightarrow\mathbb{R} be a Tonelli Lagrangian on an open subset MM of n\mathbb{R}^{n} and let ρ:[a,b]M\rho:[a,b]\rightarrow M be a C1C^{1} curve satisfying (E-L). If

α:[a,b]×(ε,ε)\displaystyle\alpha:[a,b]\times(-\varepsilon,\varepsilon)\to M\displaystyle\,M
(t,s)\displaystyle(t,s)\mapsto α(t,s).\displaystyle\,\alpha(t,s).

is a continuous piecewise C3C^{3} variation of ρ\rho such that

  1. (i)

    α(t,0)=ρ(t),t[a,b]\alpha(t,0)=\rho(t),t\in[a,b],

  2. (ii)

    there exists a partition a=t0<t1<<tk+1=ba=t_{0}<t_{1}<\cdots<t_{k+1}=b such that α\alpha is of C3C^{3} on [ti,ti+1]×(ε,ε),i=1,2,,k[t_{i},t_{i+1}]\times(-\varepsilon,\varepsilon),i=1,2,\cdots,k.

Then we have

2s2|s=0abL(α,αt)𝑑t=I(αs|s=0,αs|s=0)+i=1kLv2αs2|ti+ti+1.\displaystyle\left.\frac{\partial^{2}}{\partial s^{2}}\right|_{s=0}\int_{a}^{b}L(\alpha,\frac{\partial\alpha}{\partial t})\ dt=I\bigg{(}\left.\frac{\partial\alpha}{\partial s}\right|_{s=0},\left.\frac{\partial\alpha}{\partial s}\right|_{s=0}\bigg{)}+\sum_{i=1}^{k}\left.L_{v}\frac{\partial^{2}\alpha}{\partial s^{2}}\right|_{t_{i}^{+}}^{t_{i+1}^{-}}.
Proof.

Notice that

sabL(α,αt)𝑑t=abLxαs+Lv2αstdt.\displaystyle\frac{\partial}{\partial s}\int_{a}^{b}L(\alpha,\frac{\partial\alpha}{\partial t})dt=\int_{a}^{b}L_{x}\frac{\partial\alpha}{\partial s}+L_{v}\frac{\partial^{2}\alpha}{\partial s\partial t}\,dt.

Then,

2s2abL(α,αt)𝑑t=\displaystyle\frac{\partial^{2}}{\partial s^{2}}\int_{a}^{b}L(\alpha,\frac{\partial\alpha}{\partial t})dt= abLxxαsαs+Lxv2αstαs+Lvxαs2αst+Lvv2αst2αst\displaystyle\int_{a}^{b}L_{xx}\frac{\partial\alpha}{\partial s}\cdot\frac{\partial\alpha}{\partial s}+L_{xv}\frac{\partial^{2}\alpha}{\partial s\partial t}\cdot\frac{\partial\alpha}{\partial s}+L_{vx}\frac{\partial\alpha}{\partial s}\cdot\frac{\partial^{2}\alpha}{\partial s\partial t}+L_{vv}\frac{\partial^{2}\alpha}{\partial s\partial t}\cdot\frac{\partial^{2}\alpha}{\partial s\partial t}
+Lx2αs2+Lv3αs2tdt\displaystyle+L_{x}\frac{\partial^{2}\alpha}{\partial s^{2}}+L_{v}\frac{\partial^{3}\alpha}{\partial s^{2}\partial t}dt
=\displaystyle= abLxxαsαs+Lxv2αstαs+Lvxαs2αst+Lvv2αst2αst\displaystyle\int_{a}^{b}L_{xx}\frac{\partial\alpha}{\partial s}\cdot\frac{\partial\alpha}{\partial s}+L_{xv}\frac{\partial^{2}\alpha}{\partial s\partial t}\cdot\frac{\partial\alpha}{\partial s}+L_{vx}\frac{\partial\alpha}{\partial s}\cdot\frac{\partial^{2}\alpha}{\partial s\partial t}+L_{vv}\frac{\partial^{2}\alpha}{\partial s\partial t}\cdot\frac{\partial^{2}\alpha}{\partial s\partial t}
+Lx2αs2(ddtLv)2αs2dt+i=1kLv2αs2|ti+ti+1.\displaystyle+L_{x}\frac{\partial^{2}\alpha}{\partial s^{2}}-\Big{(}\frac{d}{dt}L_{v}\Big{)}\frac{\partial^{2}\alpha}{\partial s^{2}}dt+\sum_{i=1}^{k}\left.L_{v}\frac{\partial^{2}\alpha}{\partial s^{2}}\right|_{t_{i}^{+}}^{t_{i+1}^{-}}.

Taking s=0s=0, we obtain

2s2|s=0abL(α,αt)𝑑t=I(αs|s=0,αs|s=0)+i=1kLv2αs2|ti+ti+1.\displaystyle\left.\frac{\partial^{2}}{\partial s^{2}}\right|_{s=0}\int_{a}^{b}L(\alpha,\frac{\partial\alpha}{\partial t})dt=I\bigg{(}\left.\frac{\partial\alpha}{\partial s}\right|_{s=0},\left.\frac{\partial\alpha}{\partial s}\right|_{s=0}\bigg{)}+\sum_{i=1}^{k}\left.L_{v}\frac{\partial^{2}\alpha}{\partial s^{2}}\right|_{t_{i}^{+}}^{t_{i+1}^{-}}.

Proposition B.4.

Suppose L:TML:TM\rightarrow\mathbb{R} is a Tonelli Lagrangian defined on a closed connected Riemannian manifold (M,g)(M,g). Let ρ:[a,b]M\rho:[a,b]\rightarrow M be a C1C^{1} curve such that

Aba(ρ(z),ρ(b))=abL(ρ(s),ρ˙(s))𝑑s.\displaystyle A_{b-a}(\rho(z),\rho(b))=\int_{a}^{b}L(\rho(s),\dot{\rho}(s))ds.

Then, (ρ(b),ρ˙(b))(\rho(b),\dot{\rho}(b)) is not conjugate to (ρ(z),ρ˙(z))(\rho(z),\dot{\rho}(z)) for each z(a,b)z\in(a,b).

Proof.

Since there exists a smooth map σ:[a,b]×UM\sigma:[a,b]\times U\rightarrow M with UU an open neighborhood of 0n0\in\mathbb{R}^{n}, such that yσ(t,y)y\mapsto\sigma(t,y), mapping 0 to ρ(t)\rho(t), is a diffeomorphism from UU to an open neighborhood of ρ(t)M\rho(t)\in M, the computations can be reduced to the case that MM is an open subset of n\mathbb{R}^{n}.

Now, suppose that (ρ(b),ρ˙(b))(\rho(b),\dot{\rho}(b)) is conjugate to (ρ(z),ρ˙(z))(\rho(z),\dot{\rho}(z)) for some z(a,b)z\in(a,b). Then there exists a nonzero Jacobi field J:[z,b]TMJ:[z,b]\rightarrow TM along ρ\rho with J(z)=J(b)=0J(z)=J(b)=0. Consider the continuous piecewise C2C^{2} vector field

J^(t)={0, if t[a,z),J(t), if t[z,b],\displaystyle\hat{J}(t)=\left\{\begin{array}[]{cc}0,&\text{ if }t\in[a,z),\\ J(t),&\text{ if }t\in[z,b],\end{array}\right.

and a smooth vector field EE along ρ\rho such that E(a)=E(b)=0E(a)=E(b)=0 and E(z)=J˙(z)0E(z)=\dot{J}(z)\neq 0. Set V(t)=J^(t)+λE(t),t[a,b]V(t)=\hat{J}(t)+\lambda E(t),\ t\in[a,b] and

α(t,s)=ρ(t)+sV(t),(t,s)[a,b]×(ε,ε).\displaystyle\alpha(t,s)=\rho(t)+sV(t),\quad(t,s)\in[a,b]\times(-\varepsilon,\varepsilon).

Then we have α(t,0)=ρ(t),α(a,s)=ρ(a),α(b,s)=ρ(b)\alpha(t,0)=\rho(t),\alpha(a,s)=\rho(a),\alpha(b,s)=\rho(b) and αs=V\frac{\partial\alpha}{\partial s}=V. For the action

Aα(s):=abL(α(t,s),αt(t,s))𝑑t,\displaystyle A_{\alpha}(s):=\int_{a}^{b}L(\alpha(t,s),\frac{\partial\alpha}{\partial t}(t,s))dt,

we obtain that

ddsAα(0)\displaystyle\frac{d}{ds}A_{\alpha}(0) =abLxαs+Lv2αstdt\displaystyle=\int_{a}^{b}L_{x}\frac{\partial\alpha}{\partial s}+L_{v}\frac{\partial^{2}\alpha}{\partial s\partial t}dt
=abLxαs(ddtLv)αsdt+Lvαs|az+Lvαs|z+b=0.\displaystyle=\int_{a}^{b}L_{x}\frac{\partial\alpha}{\partial s}-\big{(}\frac{d}{dt}L_{v}\big{)}\frac{\partial\alpha}{\partial s}dt+\left.L_{v}\frac{\partial\alpha}{\partial s}\right|_{a}^{z^{-}}+\left.L_{v}\frac{\partial\alpha}{\partial s}\right|_{z^{+}}^{b}=0.

Moreover, by Proposition B.3 and Lemma B.2 one has

d2ds2Aα(0)\displaystyle\frac{d^{2}}{ds^{2}}A_{\alpha}(0) =I(αs|s=0,αs|s=0)+Lv2αs2|az+Lv2αs2|z+b=I(V,V)\displaystyle=I\big{(}\left.\frac{\partial\alpha}{\partial s}\right|_{s=0},\left.\frac{\partial\alpha}{\partial s}\right|_{s=0}\big{)}+\left.L_{v}\frac{\partial^{2}\alpha}{\partial s^{2}}\right|_{a}^{z^{-}}+\left.L_{v}\frac{\partial^{2}\alpha}{\partial s^{2}}\right|_{z^{+}}^{b}=I(V,V)
=I(J^+λE,J^+λE)\displaystyle=I(\hat{J}+\lambda E,\hat{J}+\lambda E)
=I(J^,J^)+2λI(J^,E)+λ2I(E,E)\displaystyle=I(\hat{J},\hat{J})+2\lambda I(\hat{J},E)+\lambda^{2}I(E,E)
=2λI(J^,E)+λ2I(E,E)\displaystyle=2\lambda I(\hat{J},E)+\lambda^{2}I(E,E)
=2λ(J˙(z)LvvJ˙(z))+λ2I(E,E).\displaystyle=-2\lambda(\dot{J}^{\top}(z)L_{vv}\dot{J}(z))+\lambda^{2}I(E,E).

Therefore, ρ¨(0)<0\ddot{\rho}(0)<0 if 0<λ<2J˙(z)LvvJ˙(z)/|I(E,E)|0<\lambda<2\dot{J}^{\top}(z)L_{vv}\dot{J}(z)/|I(E,E)|. This implies that ρ[a,b]\rho_{\mid[a,b]} is not minimizing. This leads to a contradiction. ∎

With similar argument one also has the following statement.

Corollary B.5.

Suppose L:TML:TM\rightarrow\mathbb{R} is a Tonelli Lagrangian defined on a closed connected Riemannian manifold (M,g)(M,g). Let ρ:[a,b]M\rho:[a,b]\to M be a C1C^{1} curve such that

Aba(ρ(a),ρ(z))=abL(ρ(s),ρ˙(s))𝑑s.\displaystyle A_{b-a}(\rho(a),\rho(z))=\int_{a}^{b}L(\rho(s),\dot{\rho}(s))\ ds.

Then, (ρ(z),ρ˙(z))(\rho(z),\dot{\rho}(z)) is not conjugate to (ρ(a),ρ˙(a))(\rho(a),\dot{\rho}(a)) for each z(a,b)z\in(a,b).

References

  • [1] Patrick Bernard. On the Conley decomposition of Mather sets. Rev. Mat. Iberoam., 26(1):115–132, 2010.
  • [2] Patrick Bernard. The Lax-Oleinik semi-group: a Hamiltonian point of view. Proc. Roy. Soc. Edinburgh Sect. A, 142(6):1131–1177, 2012.
  • [3] Patrick Bernard and Gonzalo Contreras. A generic property of families of Lagrangian systems. Ann. of Math. (2), 167(3):1099–1108, 2008.
  • [4] D. Burago, S. Ivanov, and B. Kleiner. On the structure of the stable norm of periodic metrics. Math. Res. Lett., 4(6):791–808, 1997.
  • [5] E. Calabi. An extension of E. Hopf’s maximum principle with an application to Riemannian geometry. Duke Math. J., 25:45–56, 1958.
  • [6] Piermarco Cannarsa and Wei Cheng. Generalized characteristics and Lax-Oleinik operators: global theory. Calc. Var. Partial Differential Equations, 56(5):Art. 125, 31, 2017.
  • [7] Piermarco Cannarsa and Carlo Sinestrari. Semiconcave functions, Hamilton-Jacobi equations, and optimal control, volume 58 of Progress in Nonlinear Differential Equations and their Applications. Birkhäuser Boston, Inc., Boston, MA, 2004.
  • [8] Gonzalo Contreras and Renato Iturriaga. Convex Hamiltonians without conjugate points. Ergodic Theory Dynam. Systems, 19(4):901–952, 1999.
  • [9] Gonzalo Contreras and Renato Iturriaga. Global minimizers of autonomous Lagrangians. 22o Colóquio Brasileiro de Matemática. Instituto de Matemática Pura e Aplicada (IMPA), Rio de Janeiro, 1999.
  • [10] Gonzalo Contreras and José Antônio G. Miranda. On finite quotient Aubry set for generic geodesic flows. Math. Phys. Anal. Geom., 23(2):Paper No. 14, 11, 2020.
  • [11] Manfredo Perdigão do Carmo. Riemannian geometry. Mathematics: Theory & Applications. Birkhäuser Boston, Inc., Boston, MA, 1992.
  • [12] J. J. Duistermaat. On the Morse index in variational calculus. Advances in Math., 21(2):173–195, 1976.
  • [13] Jost Eschenburg and Ernst Heintze. An elementary proof of the Cheeger-Gromoll splitting theorem. Ann. Global Anal. Geom., 2(2):141–151, 1984.
  • [14] Albert Fathi. Weak KAM theorem in Lagrangian dynamics. Cambridge University Press, Cambridge (to appear).
  • [15] Albert Fathi and Alessio Figalli. Optimal transportation on non-compact manifolds. Israel J. Math., 175:1–59, 2010.
  • [16] Albert Fathi, Alessio Figalli, and Ludovic Rifford. On the Hausdorff dimension of the Mather quotient. Comm. Pure Appl. Math., 62(4):445–500, 2009.
  • [17] Albert Fathi and Antonio Siconolfi. Existence of C1C^{1} critical subsolutions of the Hamilton-Jacobi equation. Invent. Math., 155(2):363–388, 2004.
  • [18] Ricardo Mañé. On the minimizing measures of Lagrangian dynamical systems. Nonlinearity, 5(3):623–638, 1992.
  • [19] John N. Mather. Action minimizing invariant measures for positive definite Lagrangian systems. Math. Z., 207(2):169–207, 1991.
  • [20] John N. Mather. Variational construction of connecting orbits. Ann. Inst. Fourier (Grenoble), 43(5):1349–1386, 1993.
  • [21] John N. Mather. Total disconnectedness of the quotient Aubry set in low dimensions. volume 56, pages 1178–1183. 2003. Dedicated to the memory of Jürgen K. Moser.
  • [22] John N. Mather. Examples of Aubry sets. Ergodic Theory Dynam. Systems, 24(5):1667–1723, 2004.
  • [23] Peter Petersen. Riemannian geometry, volume 171 of Graduate Texts in Mathematics. Springer, Cham, third edition, 2016.
  • [24] Ludovic Rifford. On viscosity solutions of certain Hamilton-Jacobi equations: regularity results and generalized Sard’s theorems. Comm. Partial Differential Equations, 33(1-3):517–559, 2008.
  • [25] Takashi Sakai. Riemannian geometry, volume 149 of Translations of Mathematical Monographs. American Mathematical Society, Providence, RI, 1996.
  • [26] Alfonso Sorrentino. On the total disconnectedness of the quotient Aubry set. Ergodic Theory Dynam. Systems, 28(1):267–290, 2008.
  • [27] Cédric Villani. Optimal transport: old and new, volume 338 of Grundlehren der Mathematischen Wissenschaften. Springer-Verlag, Berlin, 2009.
  • [28] Frank W. Warner. Foundations of differentiable manifolds and Lie groups, volume 94 of Graduate Texts in Mathematics. Springer-Verlag, New York-Berlin, 1983.
  • [29] Guofang Wei and Will Wylie. Comparison geometry for the Bakry-Emery Ricci tensor. J. Differential Geom., 83(2):377–405, 2009.