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

Proof of the C2C^{2} Mañé’s conjecture on surfaces.

Gonzalo Contreras CIMAT
A.P. 402, 36.000
Guanajuato. GTO
México.
gonzalo@cimat.mx
Abstract.

We prove that C2C^{2} generic hyperbolic Mañé sets contain a periodic periodic orbit. In dimension 2, adding a result in [7] which states that C2C^{2} generic Mañé sets are hyperbolic we obtain Mañé’s Conjecture for surfaces in the C2C^{2} topology: Given a Tonelli Lagrangian LL on a compact surface MM there is a C2C^{2} open and dense set of functions f:Mf:M\to{\mathbb{R}} such that the Mañé set of the Lagrangian L+fL+f is a hyperbolic periodic orbit.

Key words and phrases:
Mañé’s Conjecture, Aubry-Mather Theory, Lagrangian Systems.
1991 Mathematics Subject Classification:
37J51, 37D05
Partially supported by CONACYT, Mexico, grant A1-S-10145.

1. Introduction.

Let MM be a closed riemannian manifold. A Tonelli Lagrangian is a C2C^{2} function L:TML:TM\to{\mathbb{R}} that is

  1. (i)

    Convex: a>0\exists a>0(x,v),(x,w)TM\forall(x,v),(x,w)\in TM,   wvv2L(x,v)wa|w|x2w\cdot\partial^{2}_{vv}L(x,v)\cdot w\geq a|w|_{x}^{2}.

The uniform convexity assumption and the compactness of MM imply that LL is

  1. (ii)

    Superlinear: A>0\forall A>0 B>0\exists B>0 such that (x,v)TM\forall(x,v)\in TM: L(x,v)>A|v|xBL(x,v)>A\,|v|_{x}-B.

Given kk\in{\mathbb{R}}, the Mañé action potential is defined as Φk:M×M{}\Phi_{k}:M\times M\to{\mathbb{R}}\cup\{-\infty\},

(1) Φk(x,y):=infγ𝒞(x,y)k+L(γ,γ˙),\Phi_{k}(x,y):=\inf_{\gamma\in{\mathcal{C}}(x,y)}\int k+L(\gamma,{\dot{\gamma}}),

where

(2) 𝒞(x,y):={γ:[0,T]Mabsolutely continuous |T>0,γ(0)=x,γ(T)=y}.{\mathcal{C}}(x,y):=\{\gamma:[0,T]\to M\;{\text{absolutely continuous }}|\;T>0,\;\gamma(0)=x,\;\gamma(T)=y\;\}.

The Mañé critical value is

(3) c(L):=sup{k|x,yM:Φk(x,y)=}.c(L):=\sup\{\,k\in{\mathbb{R}}\;|\;\exists\,x,y\in M:\;\Phi_{k}(x,y)=-\infty\;\}.

See [13] for several characterizations of c(L)c(L).

A curve γ:M\gamma:{\mathbb{R}}\to M is semi-static if

s<tstc(L)+L(γ,γ˙)=Φc(L)(γ(s),γ(t)).\forall s<t\qquad\int_{s}^{t}c(L)+L(\gamma,{\dot{\gamma}})=\Phi_{c(L)}(\gamma(s),\gamma(t)).

Also γ:M\gamma:{\mathbb{R}}\to M is static if

s<tstc(L)+L(γ,γ˙)=Φc(L)(γ(t),γ(s)).\forall s<t\qquad\int_{s}^{t}c(L)+L(\gamma,{\dot{\gamma}})=-\Phi_{c(L)}(\gamma(t),\gamma(s)).

The Mañé set of LL is

𝒩(L):={(γ(t),γ˙(t))TM|t,γ:M is semi-static },{\mathcal{N}}(L):=\{(\gamma(t),{\dot{\gamma}}(t))\in TM\;|\;t\in{\mathbb{R}},\;\gamma:{\mathbb{R}}\to M\text{ is semi-static }\},

and the Aubry set is

𝒜(L):={(γ(t),γ˙(t))TM|t,γ:M is static }.{\mathcal{A}}(L):=\{(\gamma(t),{\dot{\gamma}}(t))\in TM\;|\;t\in{\mathbb{R}},\;\gamma:{\mathbb{R}}\to M\text{ is static }\}.

The Euler-Lagrange equation

ddtvL=xL\tfrac{d}{dt}\,\partial_{v}L=\partial_{x}L

defines the Lagrangian flow φt\varphi_{t} on TMTM. The energy function EL:TME_{L}:TM\to{\mathbb{R}},

EL(x,v):=vL(x,v)vL(x,v),E_{L}(x,v):=\partial_{v}L(x,v)\cdot v-L(x,v),

is invariant under the Lagrangian flow. The Mañé set 𝒩(L){\mathcal{N}}(L) is invariant under the Lagrangian flow and it is contained in the energy level :=EL1{c(L)}{\mathcal{E}}:=E_{L}^{-1}\{c(L)\} (see Mañé [24, p. 146] or [13]).

Let inv(L){\mathcal{M}}_{\text{inv}}(L) be the set of Borel probabilities in TMTM which are invariant under the Lagrangian flow. Define the action functional AL:inv(L){+}A_{L}:{\mathcal{M}}_{\text{inv}}(L)\to{\mathbb{R}}\cup\{+\infty\} as

AL(μ):=L𝑑μ.A_{L}(\mu):=\int L\,d\mu.

The set of minimizing measures is

min(L):=argmininv(L)AL,{\mathcal{M}}_{\min}(L):=\arg\min_{{\mathcal{M}}_{\text{inv}}(L)}A_{L},

and the Mather set (L){\mathcal{M}}(L) is the union of the support of minimizing measures:

(L):=μmin(L)supp(μ).{\mathcal{M}}(L):=\textstyle\bigcup\limits_{\mu\in{\mathcal{M}}_{\min}(L)}\operatorname{supp}(\mu).

Mañé proves (cf. Mañé [24, Thm. IV] also [11, p. 165]) that an invariant measure is minimizing if and only if it is supported in the Aubry set. Therefore we get the set of inclusions

(4) 𝒜𝒩.{\mathcal{M}}\subseteq{\mathcal{A}}\subseteq{\mathcal{N}}\subseteq{\mathcal{E}}.
1.1 Definition.

We say that 𝒩(L){\mathcal{N}}(L) is hyperbolic if there are sub-bundles EsE^{s}, EuE^{u} of T|𝒩(L)T{\mathcal{E}}|_{{\mathcal{N}}(L)} and T0>0T_{0}>0 such that

  1. (i)

    T|𝒩(L)=EsddtφtEuT{\mathcal{E}}|_{{\mathcal{N}}(L)}=E^{s}\oplus\langle\frac{d}{dt}\varphi_{t}\rangle\oplus E^{u}.

  2. (ii)

    DφT0|Es<1\left\|D\varphi_{T_{0}}|_{E^{s}}\right\|<1, DφT0|Eu<1\left\|D\varphi_{-T_{0}}|_{E^{u}}\right\|<1.

  3. (iii)

    t\forall t\in{\mathbb{R}}(Dφt)(Es)=Es(D\varphi_{t})^{*}(E^{s})=E^{s}, (Dφt)(Eu)=Eu(D\varphi_{t})^{*}(E^{u})=E^{u}.

Hyperbolicity for autonomous lagrangian or hamiltonian flows is always understood as hyperbolicity for the flow restricted to the energy level.

Fix a Tonelli Lagrangian L{L}. Let

k(L):={ϕCk(M,)|𝒩(L+ϕ) is hyperbolic },{\mathcal{H}}^{k}({L}):=\{\,\phi\in C^{k}(M,{\mathbb{R}})\;|\;{\mathcal{N}}({L}+\phi)\text{ is hyperbolic }\},

endowed with the CkC^{k} topology. By [14, lemma 5.2, p. 661] the map ϕ𝒩(L+ϕ)\phi\mapsto{\mathcal{N}}({L}+\phi) is upper semi-continuous and ϕc(L+ϕ)\phi\mapsto c({L}+\phi) is continuous [14, lemma 5.1]. This, together with the persistence of hyperbolicity (cf. [17, 5.1.8] or proposition A.1 below) imply that k(L){\mathcal{H}}^{k}({L}) is an open set for any k2k\geq 2.

In [14] theorem C shows that generically =𝒜=𝒩{\mathcal{M}}={\mathcal{A}}={\mathcal{N}} is the support of a single minimizing measure. Mañé [23, theorem F] proves that this measure is a strong limit of invariant probabilities supported on periodic orbits.

Let

𝒫2(L):={ϕC2(M,)|𝒩(L+ϕ) contains a periodic orbit or a singularity},{\mathcal{P}}^{2}({L}):=\{\,\phi\in C^{2}(M,{\mathbb{R}})\;|\;{\mathcal{N}}({L}+\phi)\text{ contains a periodic orbit or a singularity}\},

and let 𝒫2(L)¯\overline{{\mathcal{P}}^{2}({L})} be its closure in C2(M,)C^{2}(M,{\mathbb{R}}). We will prove

Theorem A.

2(L)𝒫2(L)¯{\mathcal{H}}^{2}({L})\subset\overline{{\mathcal{P}}^{2}({L})}.

In [12] we proved that if Γ𝒩(L)\Gamma\subset{\mathcal{N}}({L}) is a periodic orbit, adding a potential ϕ00\phi_{0}\geq 0 which is locally of the form ϕ0(x)=εd(x,π(Γ))2\phi_{0}(x)=\varepsilon\,d(x,\pi(\Gamma))^{2} makes Γ\Gamma a hyperbolic periodic orbit (or hyperbolic singularity) for the Lagrangian flow of L+ϕ0{L}+\phi_{0} and also 𝒩(L+ϕ0)=Γ{\mathcal{N}}({L}+\phi_{0})=\Gamma. Moreover [12, p. 934], Γ\Gamma has the locking property meaning that there is a C2C^{2} neighborhood 𝒰{\mathcal{U}} of ϕ0\phi_{0} such that for ϕ𝒰\phi\in{\mathcal{U}}, 𝒩(L+ϕ)=Γϕ{\mathcal{N}}(L+\phi)=\Gamma_{\phi} the continuation Γϕ\Gamma_{\phi} of the periodic orbit Γ\Gamma in the energy level EL+ϕ1{c(L+ϕ)}E_{L+\phi}^{-1}\{c(L+\phi)\}. This follows from the semicontinuity of ϕ𝒩(L+ϕ)\phi\mapsto{\mathcal{N}}(L+\phi) and the expansivity of Γ\Gamma. Therefore defining

𝒫2(L):={ϕC2(M,)|𝒩(L+ϕ) is a hyperbolic periodic orbit or singularity}{\mathcal{H}}{\mathcal{P}}^{2}({L}):=\{\,\phi\in C^{2}(M,{\mathbb{R}})\;|\;{\mathcal{N}}({L}+\phi)\text{ is a hyperbolic periodic orbit or singularity}\}

we get

Corollary B.

The set 𝒫2(L){\mathcal{H}}{\mathcal{P}}^{2}({L}) contains an open and dense set in 2(L){\mathcal{H}}^{2}({L}).

With A. Figalli and L. Rifford in [7] we prove

Theorem C.

If dimM=2\dim M=2 then 2(L){\mathcal{H}}^{2}({L}) is open and dense.

Thus for surfaces in the C2C^{2} topology we obtain Mañé’s Conjecture [24, p. 143]:

Corollary D.

If dimM=2\dim M=2 then 𝒫2(L){\mathcal{H}}{\mathcal{P}}^{2}({L}) contains an open and dense set in C2(M,)C^{2}(M,{\mathbb{R}}).

Observe that from the inclusions in (4), for potentials ϕ𝒫2(L)\phi\in{\mathcal{H}}{\mathcal{P}}^{2}({L}) the lagrangian L+ϕL+\phi has a unique minimizing measure and it is supported on a hyperbolic periodic orbit or a hyperbolic singularity. The set 𝒫2(L){\mathcal{H}}{\mathcal{P}}^{2}({L}) is open in the C2C^{2} topology, so we can approximate the lagrangian L{L} with a CC^{\infty} potential ϕ\phi to obtain a periodic minimizing measure, but the approximation is only proved to be C2C^{2} small.

Since in theorem A the Aubry set is hyperbolic, by the shadowing lemma 𝒜(L){\mathcal{A}}({L}) is accumulated by periodic orbits. The idea of the proof is to choose a special periodic orbit Γ\Gamma nearby 𝒜(L){\mathcal{A}}({L}) with small action and small period and prove that adding a channel ϕ\phi centered at Γ\Gamma, defined in (96) produces that Γ𝒜(L+ϕ)\Gamma\subset{\mathcal{A}}({L}+\phi).

Theorem A is the same as the main theorem in the manuscript [10] which will remain unpublished. The proof below uses a simplification devised by Huang, Lian, Ma, Xu, Zhang [20], [21], see also Bochi [2]. The proof in [10], [9] is based in the fact that generic hyperbolic Mañé sets have zero topological entropy. The following proof is based on the periodic orbit which is used to prove zero entropy. The point is that the estimates of Bressaud and Quas [6] for the action and period of optimal periodic orbits nearby 𝒜(L){\mathcal{A}}({L}) are so good that the cutting process in proposition 4.3 stops before the estimates get spoiled.

This proof owes a lot to the people working on ergodic optimization. Ergodic optimization was born as a baby version of Aubry-Mather theory adapted to symbolic dynamics [8]. Now the subject has matured enough to give the main ideas of the proof of an important conjecture in Aubry-Mather theory.

Main differences of lagrangian systems with ergodic optimization besides that the dynamical system depends on the lagrangian, are that perturbations need to be C2C^{2} instead of Lipschitz and that the perturbations ϕ\phi are defined in the configuration space and not in the phase space. These problems are solved by comparing the actions with static orbits and using Fathi’s differentiability estimates for weak KAM solutions, and observing that quasi minimizing objects inherit part of Mather’s graph property.

In section 3 we obtain periodic specifications in 𝒜(L){\mathcal{A}}(L) with exponentially small jumps and sub-exponential period. In section 4 proposition 4.3 we obtain periodic orbits Γ\Gamma nearby 𝒜(L){\mathcal{A}}(L) with small action compared to their self-approximations, called class I by Yuan-Hunt [31]. In section 5 we prove in proposition 5.3 that adding a channel ϕ\phi centered in Γ\Gamma we obtain Γ𝒜(L+ϕ)\Gamma\subset{\mathcal{A}}(L+\phi), thus proving theorem A. In symbolic dynamics this was known to Yuan-Hunt [31] but here we use the method of Quas-Siefken [29]. In appendix A we prove the refinement of the shadowing lemmas that we need.

2. Preliminars.

Let inv(L){\mathcal{M}}_{\text{inv}}(L) be the set of Borel probabilities in TMTM invariant under the Lagrangian flow. Denote by min(L){\mathcal{M}}_{\min}(L) the set of minimizing measures for the Lagrangian LL, i.e.

(5) min(L):={μinv(L)|TML𝑑μ=c(L)}.{\mathcal{M}}_{\min}(L):=\Big{\{}\,\mu\in{\mathcal{M}}_{\text{inv}}(L)\;\Big{|}\;\int_{TM}L\,d\mu=-c(L)\;\Big{\}}.

Their name is justified (cf. Mañé [24, Theorem II]) by

(6) c(L)=minμinv(L)TML𝑑μ=minμ𝒞(TM)TML𝑑μ.-c(L)=\min_{\mu\in{\mathcal{M}}_{\text{inv}}(L)}\int_{TM}L\;d\mu=\min_{\mu\in{\mathcal{C}}(TM)}\int_{TM}L\;d\mu.

Fathi and Siconolfi [16, Theorem 1.6] prove the second equality in (6) where the set of closed measures is defined by

𝒞(TM):={μ Borel probability on TM|ϕC1(M,)TM𝑑ϕ𝑑μ=0}.{\mathcal{C}}(TM):=\Big{\{}\,\mu\text{ Borel probability on }TM\;\Big{|}\;\forall\phi\in C^{1}(M,{\mathbb{R}})\;\int_{TM}d\phi\;d\mu=0\,\Big{\}}.

Given a closed curve γ:[0,T]M\gamma:[0,T]\to M, using the closed measure f𝑑μγ:=1T0Tf(γ,γ˙)𝑑t\int f\,d\mu_{\gamma}:=\frac{1}{T}\int_{0}^{T}f(\gamma,{\dot{\gamma}})\,dt in (6) we get

(7) γ closed curve in MAL+c(L)(γ)0.\gamma\text{ closed curve in $M$}\quad\Longrightarrow\quad A_{L+c(L)}(\gamma)\geq 0.

Recall that a curve γ:M\gamma:{\mathbb{R}}\to M is static for a Tonelli Lagrangian LL if

(8) s<tstL(γ,γ˙)=Φc(L)(γ(t),γ(s));s<t\quad\Longrightarrow\quad\int_{s}^{t}L(\gamma,{\dot{\gamma}})=-\Phi_{c(L)}(\gamma(t),\gamma(s));

equivalently (cf. Mañé [24, pp. 142–143]), if γ\gamma is semi-static and

(9) s<tΦc(L)(γ(s),γ(t))+Φc(L)(γ(t),γ(s))=0.s<t\quad\Longrightarrow\quad\Phi_{c(L)}(\gamma(s),\gamma(t))+\Phi_{c(L)}(\gamma(t),\gamma(s))=0.

The Aubry set is defined as

𝒜(L):={(γ(t),γ˙(t))|t,γ is static},{\mathcal{A}}(L):=\{\,(\gamma(t),{\dot{\gamma}}(t))\;|\;t\in{\mathbb{R}},\;\gamma\text{ is static}\;\},

its elements are called static vectors. In this section we prove that with this definition 𝒜(L){\mathcal{A}}(L) is invariant.

2.1 Lemma (A priori bound).

For C>0C>0 there exists A0=A0(C)>0A_{0}=A_{0}(C)>0 such that if γ:[0,T]M\gamma:[0,T]\to M is a solution of the Euler-Lagrange equation with AL(γ)<CTA_{L}(\gamma)<C\,T, then

|γ˙(t)|<A0 for all t[0,T].\left|{\dot{\gamma}}(t)\right|<A_{0}\qquad\text{ for all }t\in[0,T].
Proof:.

The Euler-Lagrange flow preserves the energy function

(10) EL:=vvLL.E_{L}:=v\cdot\partial_{v}L-L.

We have that

s0ddsEL(x,sv)|s\displaystyle\hskip-19.91684pt\forall s\geq 0\qquad\tfrac{d\,}{ds}E_{L}(x,sv)\big{|}_{s} =svvvL(x,v)vsa|v|x2.\displaystyle=s\,v\cdot\partial_{vv}L(x,v)\cdot v\geq s\,a|v|_{x}^{2}.
EL(x,v)\displaystyle E_{L}(x,v) =EL(x,0)+01ddsEL(x,sv)𝑑s\displaystyle=E_{L}(x,0)+\int_{0}^{1}\tfrac{d\,}{ds}E_{L}(x,sv)\,ds
(11) minyMEL(y,0)+12a|v|x2.\displaystyle\geq\min_{y\in M}E_{L}(y,0)+\tfrac{1}{2}a|v|_{x}^{2}.

Let

g(r):=sup{wvvL(x,v)w:|v|xr,|w|x=1}.g(r):=\sup\big{\{}w\cdot\partial_{vv}L(x,v)\cdot w\,:\,|v|_{x}\leq r,\,|w|_{x}=1\big{\}}.

Then g(r)ag(r)\geq a and

(12) EL(x,v)maxzMEL(z,0)+12g(|v|x)|v|x2.E_{L}(x,v)\leq\max_{z\in M}E_{L}(z,0)+\tfrac{1}{2}\,g(|v|_{x})\,|v|_{x}^{2}.

By the superlinearity there is B>0B>0 such that L(x,v)>|v|xBL(x,v)>|v|_{x}-B for all (x,v)TM(x,v)\in TM. Since AL(γ)<CTA_{L}(\gamma)<C\,T, the mean value theorem implies that there is t0]0,T[t_{0}\in]0,T[ such that |γ˙(t0)|<B+C|{\dot{\gamma}}(t_{0})|<B+C. Then (12) gives an upper bound on the energy of γ\gamma and (11) bounds the speed of γ\gamma.

For x,yMx,\,y\in M and T>0T>0 define

𝒞T(x,y):={γ:[0,T]M|γ is absolutely continuous,γ(0)=x,γ(T)=y}.{\mathcal{C}}_{T}(x,y):=\{\,\gamma:[0,T]\to M\,|\,\gamma\text{ is absolutely continuous},\gamma(0)=x,\,\gamma(T)=y\,\}.
2.2 Corollary.

There exists A1>0A_{1}>0 such that if x,yMx,\,y\in M and γ𝒞T(x,y)\gamma\in{\mathcal{C}}_{T}(x,y) is a solution of the Euler-Lagrange equation with

AL+c(γ)Φc(x,y)+max{T,d(x,y)},A_{L+c}(\gamma)\leq\Phi_{c}(x,y)+\max\{T,d(x,y)\},

where c=c(L)c=c(L), then

  1. (a)

    T1A1d(x,y)T\,\geq\,\tfrac{1}{A}_{1}\;d(x,y).

  2. (b)

    |γ˙(t)|A1|{\dot{\gamma}}(t)|\,\leq\,A_{1}   for all t[0,T]t\in[0,T].

Proof:.

First suppose that d(x,y)Td(x,y)\leq T. Then item (a) holds with A1=1A_{1}=1. Let

(13) (r):=|c|+sup{L(x,v)|(x,v)TM,|v|r}.\ell(r):=|c|+\sup\{\,L(x,v)\,|\,(x,v)\in TM,\,|v|\leq r\,\}.

Since d(x,y)Td(x,y)\leq T, there exists a C1C^{1} curve η:[0,T]M\eta:[0,T]\to M joining xx to yy with |η˙|1|{\dot{\eta}}|\leq 1. We have that

AL+c(γ)Φc(x,y)+TAL+c(η)+T((1)+c)T+T.A_{L+c}(\gamma)\leq\Phi_{c}(x,y)+T\leq A_{L+c}(\eta)+T\leq\big{(}\ell(1)+c\big{)}\,T+T.

Then item (b) holds for A1=A0(|(1)+c+1|)A_{1}=A_{0}(|\ell(1)+c+1|) where A0A_{0} is from Lemma 2.1.

Now suppose that d(x,y)Td(x,y)\geq T. Let η:[0,d(x,y)]M\eta:[0,d(x,y)]\to M be a minimal geodesic with |η˙|1|{\dot{\eta}}|\equiv 1 joining xx to yy. Let D:=(1)+c+2>0D:=\ell(1)+c+2>0. From the superlinearity property there is B>1B>1 such that

L(x,v)+c>D|v|B,(x,v)TM.L(x,v)+c>D\,|v|-B,\qquad\forall(x,v)\in TM.

Then

(14) [(1)+c]d(x,y)\displaystyle[\ell(1)+c]\;d(x,y) AL+c(η)Φc(x,y)\displaystyle\geq A_{L+c}(\eta)\geq\Phi_{c}(x,y)
(15) AL+c(γ)d(x,y)\displaystyle\geq A_{L+c}(\gamma)-d(x,y)
0T(D|γ˙|B)𝑑td(x,y)\displaystyle\geq\int_{0}^{T}\bigl{(}D\;|{\dot{\gamma}}|-B\,\bigr{)}\,dt-d(x,y)
Dd(x,y)BTd(x,y).\displaystyle\geq D\;d(x,y)-B\,T-d(x,y).

Hence

TD(1)c1Bd(x,y)=1Bd(x,y).T\geq\tfrac{D-\ell(1)-c-1}{B}\;d(x,y)=\tfrac{1}{B}\;d(x,y).

This implies item (a). From (14) and (15), we get that

AL(γ)\displaystyle A_{L}(\gamma) [(1)+c+1]d(x,y)cT,\displaystyle\leq\bigl{[}\,\ell(1)+c+1\,\bigr{]}\;d(x,y)-c\,T,
{B[(1)+c+1]c}T.\displaystyle\leq\bigl{\{}\,B\,[\,\ell(1)+c+1\,]-c\,\bigr{\}}\;T.

Then Lemma 2.1 completes the proof.

We say that a curve γ:[0,T]M\gamma:[0,T]\to M is a Tonelli minimizer if it minimizes the action functional on 𝒞T(γ(0),γ(T)){\mathcal{C}}_{T}(\gamma(0),\gamma(T)), i.e. if it is a minimizer with fixed endpoints and fixed time interval.

2.3 Corollary.

There is A>0A>0 such that if x,yMx,\,y\in M and ηn𝒞Tn(x,y)\eta_{n}\in{\mathcal{C}}_{T_{n}}(x,y), n+n\in{\mathbb{N}}^{+} is a Tonelli minimizer with

AL+c(ηn)Φc(x,y)+1n,A_{L+c}(\eta_{n})\leq\Phi_{c}(x,y)+\tfrac{1}{n},

then there is N0>0N_{0}>0 such that n>N0\forall n>N_{0}, t[0,Tn]\forall t\in[0,T_{n}], |η˙n(t)|<A|{\dot{\eta}}_{n}(t)|<A.

Proof:.

If d(x,y)>0d(x,y)>0 then for nn large enough d(x,y)>1nd(x,y)>\tfrac{1}{n}. In this case Corollary 2.2 implies the result with the constant A1A_{1}. If d(x,y)=0d(x,y)=0 let ξn:[0,Tn]{x}\xi_{n}:[0,T_{n}]\to\{x\} be the constant curve. Since ηn\eta_{n} is a Tonelli minimizer, we have that

AL(ηn)AL(ξn)=0TnL(x,0)𝑑t|L(x,0)|Tn.A_{L}(\eta_{n})\leq A_{L}(\xi_{n})=\int_{0}^{T_{n}}L(x,0)\,dt\leq|L(x,0)|\,T_{n}.

Lemma 2.1 implies that |η˙n|A0(C)|{\dot{\eta}}_{n}|\leq A_{0}(C) with C=supxM|L(x,0)|C=\sup_{x\in M}|L(x,0)|. Now take A=max{A0(C),A1}A=\max\{A_{0}(C),A_{1}\}.

2.4 Lemma.

If (x,v)(x,v) is a static vector then γ:M\gamma:{\mathbb{R}}\to M, γ(t)=πφt(x,v)\gamma(t)=\pi\varphi_{t}(x,v) is a static curve, i.e. the Aubry set 𝒜(L){\mathcal{A}}(L) is invariant.

Proof:.

Let γ(t)=πφt(x,v)\gamma(t)=\pi\,\varphi_{t}(x,v) and suppose that γ|[a,b]\gamma|_{[a,b]} is static. We have to prove that all γ|\gamma|_{\mathbb{R}} is static. Let ηn𝒞Tn(γ(b),γ(a))\eta_{n}\in{\mathcal{C}}_{T_{n}}(\gamma(b),\gamma(a)) be a Tonelli minimizer with

AL+c(ηn)<Φc(γ(b),γ(a))+1n.A_{L+c}(\eta_{n})<\Phi_{c}(\gamma(b),\gamma(a))+\tfrac{1}{n}.

By Corollary 2.3, for nn large enough, |η˙n|<A|{\dot{\eta}}_{n}|<A. We can assume that η˙n(0)w{\dot{\eta}}_{n}(0)\to w. Let ξ(s)=πφs(w)\xi(s)=\pi\,\varphi_{s}(w). If wγ˙(b)w\neq{\dot{\gamma}}(b) then for some ε>0\varepsilon>0 the curve γ|[bε,b]ξ|[0,ε]\gamma|_{[b-\varepsilon,b]}*\xi|_{[0,\varepsilon]} is not C1C^{1}, and hence it can not be a (Tonelli) minimizer of AL+cA_{L+c} in 𝒞2ε(γ(bε),ξ(ε)){\mathcal{C}}_{2\varepsilon}\big{(}\gamma(b-\varepsilon),\xi(\varepsilon)\big{)}. Thus

Φc(γ(bε),ξ(ε))<AL+c(γ|[bε,b])+AL+c(ξ|[0,ε]).\Phi_{c}(\gamma(b-\varepsilon),\xi(\varepsilon))<A_{L+c}(\gamma|_{[b-\varepsilon,b]})+A_{L+c}(\xi|_{[0,\varepsilon]}).
Φc(γ(a),γ(a))Φc(γ(a),γ(bε))+Φc(γ(bε),ξ(ε))+Φc(ξ(ε),γ(a))\displaystyle\Phi_{c}(\gamma(a),\gamma(a))\leq\Phi_{c}(\gamma(a),\gamma(b-\varepsilon))+\Phi_{c}(\gamma(b-\varepsilon),\xi(\varepsilon))+\Phi_{c}(\xi(\varepsilon),\gamma(a))
<AL+c(γ[a,bε])+AL+c(γ|[bε,b])+AL+c(ξ|[0,ε])+lim infnAL+c(ηn|[ε,Tn])\displaystyle\;<A_{L+c}(\gamma_{[a,b-\varepsilon]})+A_{L+c}(\gamma|_{[b-\varepsilon,b]})+A_{L+c}(\xi|_{[0,\varepsilon]})+\liminf_{n}A_{L+c}(\eta_{n}|_{[\varepsilon,T_{n}]})
AL+c(γ|[a,b])+limnAL+c(ηn|[0,ε]ηn|[ε,Tn])\displaystyle\;\leq A_{L+c}(\gamma|_{[a,b]})+\lim_{n}A_{L+c}\bigl{(}\,\eta_{n}|_{[0,\varepsilon]}*\eta_{n}|_{[\varepsilon,T_{n}]}\bigr{)}
=Φc(γ(b),γ(a))+Φc(γ(b),γ(a))=0.\displaystyle\;=-\Phi_{c}(\gamma(b),\gamma(a))+\Phi_{c}(\gamma(b),\gamma(a))=0.

Thus there is a closed curve, from γ(a)\gamma(a) to itself, with negative L+cL+c action, and also negative L+kL+k action for some k>c(L)k>c(L). Concatenating the curve with itself many times shows that Φk(γ(a),γ(a))=\Phi_{k}(\gamma(a),\gamma(a))=-\infty. By (3) this implies that kc(L)k\leq c(L), which is a contradiction. Thus w=γ˙(b)w={\dot{\gamma}}(b) and similarly limnη˙n(Tn)=γ˙(a)\lim_{n}{\dot{\eta}}_{n}(T_{n})={\dot{\gamma}}(a).

If lim supTn<+\limsup T_{n}<+\infty, we can assume that τ=limnTn>0\tau=\lim_{n}T_{n}>0 exists. In this case γ\gamma is a semi-static periodic orbit of period τ+ba\tau+b-a and then γ|\gamma|_{\mathbb{R}} is static.

Now suppose that limnTn=+\lim_{n}T_{n}=+\infty. If s>0s>0, we have that

AL+c\displaystyle A_{L+c} (γ|[as,b+s])+Φc(γ(b+s),γ(as))\displaystyle(\gamma|_{[a-s,b+s]})+\Phi_{c}(\gamma(b+s),\gamma(a-s))\leq
limn{AL+c(ηn|[Tns,Tn])+AL+c(γ|[a,b])+AL+c(ηn|[0,s])}+Φc(γ(b+s),γ(as))\displaystyle\leq\lim_{n}\big{\{}\,A_{L+c}(\eta_{n}|_{[T_{n}-s,T_{n}]})+A_{L+c}(\gamma|_{[a,b]})\begin{aligned} &+A_{L+c}(\eta_{n}|_{[0,s]})\,\big{\}}\\ &+\Phi_{c}(\gamma(b+s),\gamma(a-s))\end{aligned}
Φc(γ(b),γ(a))+limn{AL+c(ηn|[0,s])+AL+c(ηn|[s,Tns])+AL+c(ηn|[Tns,Tn])}\displaystyle\leq\begin{aligned} &-\Phi_{c}(\gamma(b),\gamma(a))\\ &+\lim_{n}\big{\{}\,A_{L+c}(\eta_{n}|_{[0,s]})+A_{L+c}(\eta_{n}|_{[s,T_{n}-s]})+A_{L+c}(\eta_{n}|_{[T_{n}-s,T_{n}]})\,\big{\}}\end{aligned}
Φc(γ(b),γ(a))+Φc(γ(b),γ(a))=0.\displaystyle\leq-\Phi_{c}(\gamma(b),\gamma(a))+\Phi_{c}(\gamma(b),\gamma(a))=0.

Thus γ[as,b+s]\gamma_{[a-s,b+s]} is static for all s>0s>0.

3. Optimal specifications.

Here lemma 3.1 and proposition 3.2 follow arguments by X. Bressaud and A. Quas [6].

Let A{0,1}M×MA\in\{0,1\}^{M\times M} be a M×MM\times M matrix of with entries in {0,1}\{0,1\}. The subshift of finite type ΣA\Sigma_{A} associated to AA is the set

ΣA={(xi)i{0,1}|iA(xi,xi+1)=1},\Sigma_{A}=\big{\{}\;(x_{i})_{i\in{\mathbb{Z}}}\in\{0,1\}^{\mathbb{Z}}\;\big{|}\quad\forall i\in{\mathbb{Z}}\quad A(x_{i},x_{i+1})=1\;\big{\}},

endowed with the metric

d(x,y)=2i,i=max{k|xi=yi|i|k}d(x,y)=2^{-i},\qquad i=\max\{\;k\in{\mathbb{N}}\;|\;x_{i}=y_{i}\;\;\forall|i|\leq k\;\}

and the shift transformation

σ:ΣAΣA,iσ(x)i=xi+1.\sigma:\Sigma_{A}\to\Sigma_{A},\qquad\forall i\in{\mathbb{Z}}\quad\sigma(x)_{i}=x_{i+1}.

3.1 Lemma.

Let ΣA\Sigma_{A} be a shift of finite type with MM symbols and topological entropy hh. Then ΣA\Sigma_{A} contains a periodic orbit of period at most 1+Me1h1+M\text{\rm\large e}^{1-h}.

Proof:.

Let k+1k+1 be the period of the shortest periodic orbit in ΣA\Sigma_{A}. We claim that a word of length kk in ΣA\Sigma_{A} is determined by the set of symbols that it contains. First note that since there are no periodic orbits of period kk or less, any allowed kk-word must contain kk distinct symbols. Now suppose that uu and vv are two distinct words of length kk in ΣA\Sigma_{A} containing the same symbols. Then, since the words are different, there is a consecutive pair of symbols, say aa and bb, in vv which occur in the opposite order (not necessarily consecutively) in uu. Then the infinite concatenation of the segment of uu starting at bb and ending at aa gives a word in ΣA\Sigma_{A} of period at most kk, which contradicts the choice of kk.

It follows that there are at most (Mk)\binom{M}{k} words of length kk. Using the basic properties of topological entropy [22, 4.1.8]

ehk(Mk)Mkk!(Mek)k.\displaystyle\text{\rm\large e}^{hk}\leq\binom{M}{k}\leq\frac{M^{k}}{k!}\leq\left(\frac{M\text{\rm\large e}}{k}\right)^{k}.

Taking kk-th roots, we see that kMe1hk\leq M\text{\rm\large e}^{1-h}.

From now on we assume that the Mañé set 𝒩(L){\mathcal{N}}(L) is hyperbolic. The definition of a specification or pseudo-orbit appears in A.12 in appendix A.

3.2 Proposition.

There are C,λ>0C,\lambda>0 such that for T>1T>1 large there is (Θ,𝔗)=({θi},{ti})(\Theta,\mathfrak{T})=(\{\theta_{i}\},\{t_{i}\}) a periodic T-specification in 𝒜(L){\mathcal{A}}(L), with P=PTP=P_{T} jumps (θi,ti)=(θi+P,ti+P)(\theta_{i},t_{i})=(\theta_{i+P},t_{i+P}), and period 4TPT\leq 4TP_{T} such that

(16) limT1TlogPT=0,\displaystyle\lim\nolimits_{T\to\infty}\tfrac{1}{T}\log P_{T}=0,
(17) imodPTd(φti(θi),φti(θi1))CeλT.\displaystyle\forall i\in{\mathbb{Z}}\mod P_{T}\qquad d\big{(}\varphi_{t_{i}}(\theta_{i}),\varphi_{t_{i}}(\theta_{i-1})\big{)}\leq C\,\text{\rm\large e}^{-\lambda T}.
Proof:.

Given a specification (Θ,𝔗)(\Theta,\mathfrak{T}) in 𝒜(L){\mathcal{A}}(L) write ξi:[ti,ti+1]𝒜(L)\xi_{i}:[t_{i},t_{i+1}]\to{\mathcal{A}}(L), ξi(s)=φs(θi)\xi_{i}(s)=\varphi_{s}(\theta_{i}); and ζi:[0,ti+1ti]𝒜(L)\zeta_{i}:[0,t_{i+1}-t_{i}]\to{\mathcal{A}}(L), ζi(s)=ξi(s+ti)\zeta_{i}(s)=\xi_{i}(s+t_{i}). We identify (Θ,𝔗)(\Theta,\mathfrak{T}), {ξi}\{\xi_{i}\}, {ζi}\{\zeta_{i}\} as the same specification. We will extend the definition of ξi\xi_{i}, ζi\zeta_{i} to larger intervals, with the same formula, as needed.

Let T>0T>0 and let δ>0\delta>0 be smaller than half of an expansivity constant A.8 for 𝒜(L){\mathcal{A}}(L) and smaller than β0\beta_{0} in proposition A.7 applied to 𝒜(L){\mathcal{A}}(L). Let G=GTG=G_{T} be a minimal (2T,δ)(2T,\delta)-spanning set for 𝒜(L){\mathcal{A}}(L), i.e.

(18) 𝒜(L)θGB(θ,2T,δ),{\mathcal{A}}(L)\subset\bigcup_{\theta\in G}B(\theta,2T,\delta),

where B(θ,2T,δ)B(\theta,2T,\delta) is the dynamic ball

B(θ,2T,δ)={ϑTM|d(φs(θ),φs(ϑ))δs[0,2T]},B(\theta,2T,\delta)=\{\,\vartheta\in TM\;|\;d(\varphi_{s}(\theta),\varphi_{s}(\vartheta))\leq\delta\;\;\;\forall s\in[0,2T]\;\},

and no proper subset of GG satisfies (18). Let ΣG\Sigma\subset G^{\mathbb{Z}} be the bi-infinite subshift of finite type with symbols in GG and matrix A{0,1}G×GA\in\{0,1\}^{G\times G} defined by

(19) A(θ,ϑ)=1φ2T(θ)B(ϑ,2T,δ).A(\theta,\vartheta)=1\qquad\Longleftrightarrow\qquad\varphi_{2T}(\theta)\in B(\vartheta,2T,\delta).

Given N+N\in{\mathbb{N}}^{+}, let SNS_{N} be a maximal (2NT,2δ)(2NT,2\delta)-separated set in 𝒜(L){\mathcal{A}}(L), i.e.

(20) θ,ϑSN,θϑϑB(θ,2NT,2δ),\theta,\,\vartheta\in S_{N},\quad\theta\neq\vartheta\quad\Longrightarrow\quad\vartheta\notin B(\theta,2NT,2\delta),

and SNS_{N} is a maximal subset of 𝒜(L){\mathcal{A}}(L) with property (20).

Given θSN\theta\in S_{N} let I(θ)I(\theta) be an itinerary in Σ\Sigma corresponding to θ\theta, i.e.

nφ2nT(θ)B(I(θ)n,2T,δ),I(θ)nGT.\forall n\in{\mathbb{Z}}\qquad\varphi_{2nT}(\theta)\in B(I(\theta)_{n},2T,\delta),\qquad I(\theta)_{n}\in G_{T}.

If θ,ϑSN\theta,\vartheta\in S_{N} are different points, then by (20) there are 0n<N0\leq n<N and s[0,2T]s\in[0,2T] such that d(φ2nT+s(θ),φ2nT+s(ϑ))>2δd(\varphi_{2nT+s}(\theta),\varphi_{2nT+s}(\vartheta))>2\delta. Thus I(θ)nI(ϑ)nI(\theta)_{n}\neq I(\vartheta)_{n}, i.e. I(θ)I(\theta), I(ϑ)I(\vartheta) belong to different NN-cylinders in Σ\Sigma. Therefore

CN:=#(N-cylinders in Σ)#SN.C_{N}:=\#(\text{$N$-cylinders in $\Sigma$})\geq\#S_{N}.

Since 2δ2\delta is smaller than an h-expansivity constant for 𝒜(L){\mathcal{A}}(L), see remark A.9, its topological entropy can be calculated using (n,2δ)(n,2\delta)-separated (or (n,δ)(n,\delta)-spanning) sets, htop(φ,𝒜(L))=h(φ,𝒜(L),2δ)h_{\rm top}(\varphi,{\mathcal{A}}(L))=h(\varphi,{\mathcal{A}}(L),2\delta) (c.f. Bowen [3] Thm. 2.4, p. 327), thus

h(Σ)lim supNlog#CNN2Tlim supNlog#SN2NT=2Thtop(𝒜(L))=:2Th.h(\Sigma)\geq\limsup_{N}\frac{\log\#C_{N}}{N}\geq 2T\limsup_{N}\frac{\log\#S_{N}}{2NT}=2T\,h_{top}({\mathcal{A}}(L))=:2Th.

There is KTK_{T} with sub-exponential growth in TT such that #GTKTe2Th\#G_{T}\leq K_{T}\,\text{\rm\large e}^{2Th}. Then Lemma 3.1 gives a periodic orbit Θ\Theta in Σ\Sigma with

(21) P:=period(Θ)1+KTe2The12Th1+KTe.P:={\rm period}(\Theta)\leq 1+K_{T}\,\text{\rm\large e}^{2Th}\;\text{\rm\large e}^{1-2Th}\leq 1+K_{T}\,\text{\rm\large e}.

By Proposition A.7, if θ,ϑ𝒜(L)\theta,\,\vartheta\in{\mathcal{A}}(L) and ϑB(θ,2T,δ)\vartheta\in B(\theta,2T,\delta) then there is

(22) |v|=|v(ϑ,θ)|<Dδ|v|=|v(\vartheta,\theta)|<D\,\delta

such that

(23) |s|Td(φs+v+T(ϑ),φs+T(θ))Dδeλ(T|s|).\forall\;|s|\leq T\quad d\big{(}\varphi_{s+v+T}(\vartheta),\varphi_{s+T}(\theta)\big{)}\leq D\,\delta\,\text{\rm\large e}^{-\lambda(T-|s|)}.

Given a sequence (θi)iΣ(\theta_{i})_{i\in{\mathbb{Z}}}\in\Sigma, define a specification (ζi|[0,2T+vi])i(\zeta_{i}|_{[0,2T+v_{i}]})_{i\in{\mathbb{Z}}} in 𝒜(L){\mathcal{A}}(L) by vi:=v(φ2T(θi),θi+1)v_{i}:=v(\varphi_{2T}(\theta_{i}),\theta_{i+1}) from (22), and ζi(s):=φs+T(θi)\zeta_{i}(s):=\varphi_{s+T}(\theta_{i}). From (19) we have that φ2T(θi)B(θi+1,2T,δ)\varphi_{2T}(\theta_{i})\in B(\theta_{i+1},2T,\delta). Then by (23), with ϑ=φ2T(θi)\vartheta=\varphi_{2T}(\theta_{i}), θ=θi+1\theta=\theta_{i+1} and s=0s=0

(24) d(ζi(2T+vi),ζi+1(0))=d(φ3T+vi(θi),φT(θi+1))DδeλT.d(\zeta_{i}(2T+v_{i}),\zeta_{i+1}(0))=d(\varphi_{3T+v_{i}}(\theta_{i}),\varphi_{T}(\theta_{i+1}))\leq D\,\delta\,\text{\rm\large e}^{-\lambda T}.

For the sequence ΘΣ\Theta\in\Sigma in (21) we have that (ξi)i(\xi_{i})_{i\in{\mathbb{Z}}} a periodic DδeλTD\delta\,\text{\rm\large e}^{-\lambda T}-possible specification with PP jumps, and period

(25) period({ξi})(2T+Dδ)(1+KTe)4T(1+KTe).{\rm period}(\{\xi_{i}\})\leq(2T+D\delta)\,(1+K_{T}\,\text{\rm\large e})\leq 4T(1+K_{T}\,\text{\rm\large e}).

4. Optimal periodic orbits.

A dominated function for LL is a function u:Mu:M\to{\mathbb{R}} such that for any γ:[0,T]M\gamma:[0,T]\to M absolutely continuous and 0s<tT0\leq s<t\leq T we have

(26) u(γ(t))u(γ(s))st[c(L)+L(γ,γ˙)].u(\gamma(t))-u(\gamma(s))\leq\int_{s}^{t}\big{[}c(L)+L(\gamma,{\dot{\gamma}})\big{]}.

We say that the curve γ\gamma calibrates uu if the equality holds in (26) for every 0s<tT0\leq s<t\leq T. Dominated functions always exist, for example, by the triangle inequality for Mañé’s potential Φc\Phi_{c}, the functions up(x):=Φc(p,x)u_{p}(x):=\Phi_{c}(p,x) are dominated for every pMp\in M. The definition of the Hamiltonian HH associated to LL implies that any C1C^{1} function u:Mu:M\to{\mathbb{R}} which satisfies

xM,H(x,dxu)c(L)\forall x\in M,\quad H(x,d_{x}u)\leq c(L)

is dominated.

4.1 Lemma.

If uu is a dominated function and γ\gamma is a static curve then γ\gamma calibrates uu.

Proof:.

Recall from (8), (9) that γ\gamma is static iff for all s<ts<t we have

(27) st[c(L)+L(γ,γ˙)]=ϕc(L)(γ(t),γ(s))=ϕc(L)(γ(s),γ(t)).\int_{s}^{t}\big{[}c(L)+L(\gamma,{\dot{\gamma}})\big{]}=-\phi_{c(L)}(\gamma(t),\gamma(s))=\phi_{c(L)}(\gamma(s),\gamma(t)).

If uu is dominated, γ\gamma is static and s<ts<t we have that

u(γ(t))u(γ(s))+ϕc(L)(γ(s),γ(t))=u(γ(s))ϕc(L)(γ(t),γ(s)).\displaystyle u(\gamma(t))\leq u(\gamma(s))+\phi_{c(L)}(\gamma(s),\gamma(t))=u(\gamma(s))-\phi_{c(L)}(\gamma(t),\gamma(s)).

Using again the domination of uu and then the previous inequality we get

u(γ(s))u(γ(t))+ϕc(L)(γ(t),γ(s))u(γ(s)).u(\gamma(s))\leq u(\gamma(t))+\phi_{c(L)}(\gamma(t),\gamma(s))\leq u(\gamma(s)).

Therefore, using (27),

u(γ(t))=u(γ(s))ϕc(L)(γ(t),γ(s))=u(γ(s))+st[c(L)+L(γ,γ˙)].u(\gamma(t))=u(\gamma(s))-\phi_{c(L)}(\gamma(t),\gamma(s))=u(\gamma(s))+\int_{s}^{t}\big{[}c(L)+L(\gamma,{\dot{\gamma}})\big{]}.

4.2 Lemma.

There are K>0K>0 and δ0>0\delta_{0}>0 such that if (z,z˙)𝒜(L)(z,{\dot{z}})\in{\mathcal{A}}(L) is a static vector, uu is a dominated function and d(z,y)<δ0d(z,y)<\delta_{0}, then in local coordinates

(28) |u(y)u(z)vL(z,z˙)(yz)|K|yz|2,\big{|}u(y)-u(z)-\partial_{v}L(z,{\dot{z}})(y-z)\big{|}\leq K\,|y-z|^{2},

where yz:=(expz)1(y)y-z:=(\exp_{z})^{-1}(y).

Proof:.

Let 𝔼TM{\mathbb{E}}\subset TM be a compact subset such that EL1{c(L)}int𝔼E^{-1}_{L}\{c(L)\}\subset\operatorname{int}{\mathbb{E}}. Cover MM by a finite set 𝒪{\mathcal{O}} of charts. Fix 0<ε<10<\varepsilon<1 such that if γ:[ε,ε]M\gamma:[-\varepsilon,\varepsilon]\to M has velocity (γ,γ˙)𝔼(\gamma,{\dot{\gamma}})\in{\mathbb{E}} then γ([ε,ε])\gamma([-\varepsilon,\varepsilon]) lies inside the domain of a chart in 𝒪{\mathcal{O}}. There are δ1>0\delta_{1}>0 smaller than the Lebesgue number of the covering 𝒪{\mathcal{O}} and A>0A>0 such that if (x,v)𝔼(x,v)\in{\mathbb{E}} and max{|h|,|k|}δ1\max\{|h|,\,|k|\}\leq\delta_{1} then in the charts

(29) |L(x+h,v+k)L(x,v)DL(x,v)(h,k)|A(|h|2+|k|2).\big{|}L(x+h,v+k)-L(x,v)-DL(x,v)(h,k)\big{|}\leq A(|h|^{2}+|k|^{2}).

Let u:Mu:M\to{\mathbb{R}} be dominated and (z,z˙)𝒜(L)(z,{\dot{z}})\in{\mathcal{A}}(L). Recall that 𝒜(L)EL1{c(L)}𝔼{\mathcal{A}}(L)\subset E_{L}^{-1}\{c(L)\}\subset{\mathbb{E}}. Write γ(t):=πφtL(z,z˙)\gamma(t):=\pi\varphi^{L}_{t}(z,{\dot{z}}). By Lemma 2.4 the complete curve γ:M\gamma:{\mathbb{R}}\to M is static. By Lemma 4.1, γ\gamma calibrates uu. Let δ0:=εδ1\delta_{0}:=\varepsilon\,\delta_{1}. Let yMy\in M with |yz|<δ0|y-z|<\delta_{0} in a local chart. Define β:]ε,0]M\beta:]-\varepsilon,0]\to M by

β(t):=γ(t)+(t+εε)(yz).\beta(t):=\gamma(t)+\left(\tfrac{t+\varepsilon}{\varepsilon}\right)(y-z).

Then β(ε)=γ(ε)\beta(-\varepsilon)=\gamma(-\varepsilon), β(0)=y\beta(0)=y, β˙=γ˙+1ε(yz){\dot{\beta}}={\dot{\gamma}}+\tfrac{1}{\varepsilon}(y-z). In particular |β˙γ˙|1ε|yz|δ1|{\dot{\beta}}-{\dot{\gamma}}|\leq\tfrac{1}{\varepsilon}|y-z|\leq\delta_{1} and we can apply (29).

ε0L(β,β˙)ε0L(γ,γ˙)+ε0{Lx(γ,γ˙)(βγ)+Lv(γ,γ˙)(β˙γ˙)}+Aε(1+1ε2)|yz|2.\displaystyle\int_{-\varepsilon}^{0}L(\beta,{\dot{\beta}})\leq\int_{-\varepsilon}^{0}L(\gamma,{\dot{\gamma}})+\int_{-\varepsilon}^{0}\Big{\{}L_{x}(\gamma,{\dot{\gamma}})(\beta-\gamma)+L_{v}(\gamma,{\dot{\gamma}})({\dot{\beta}}-{\dot{\gamma}})\Big{\}}+A\varepsilon\big{(}1+\tfrac{1}{\varepsilon^{2}}\big{)}\,|y-z|^{2}.

Using that γ\gamma is a solution of the Euler-Lagrange equation ddtLv=Lx\tfrac{d\,}{dt}L_{v}=L_{x} and integrating by parts, we get that

ε0L(β,β˙)\displaystyle\int_{-\varepsilon}^{0}L(\beta,{\dot{\beta}}) ε0L(γ,γ˙)𝑑t+Lv(γ,γ˙)(βγ)|ε0+2Aε|yz|2,\displaystyle\leq\int_{-\varepsilon}^{0}L(\gamma,{\dot{\gamma}})\,dt+L_{v}(\gamma,{\dot{\gamma}})(\beta-\gamma)\Big{|}_{-\varepsilon}^{0}+\tfrac{2A}{\varepsilon}\,|y-z|^{2},
(30) ε0L(γ,γ˙)𝑑t+Lv(z,z˙)(yz)+2Aε|yz|2.\displaystyle\leq\int_{-\varepsilon}^{0}L(\gamma,{\dot{\gamma}})\,dt+L_{v}(z,{\dot{z}})(y-z)+\tfrac{2A}{\varepsilon}\,|y-z|^{2}.

Since uu is dominated and calibrated by γ|[ε,0]\gamma|_{[-\varepsilon,0]} we obtain one of the inequalities in (28):

u(y)\displaystyle u(y) u(γ(ε))+ε0c(L)+L(β,β˙)\displaystyle\leq u(\gamma(-\varepsilon))+\int_{-\varepsilon}^{0}c(L)+L(\beta,{\dot{\beta}})
u(γ(ε))+ε0{L(γ,γ˙)+c(L)}𝑑t+Lv(z,z˙)(yz)+2Aε|yz|2\displaystyle\leq u(\gamma(-\varepsilon))+\int_{-\varepsilon}^{0}\Big{\{}L(\gamma,{\dot{\gamma}})+c(L)\Big{\}}dt+L_{v}(z,{\dot{z}})(y-z)+\tfrac{2A}{\varepsilon}\,|y-z|^{2}
u(z)+Lv(z,z˙)(yz)+2Aε|yz|2.\displaystyle\leq u(z)+L_{v}(z,{\dot{z}})(y-z)+\tfrac{2A}{\varepsilon}\,|y-z|^{2}.

Now define α:[0,ε]M\alpha:[0,\varepsilon]\to M by

α(t):=γ(t)+(εtε)(yz).\alpha(t):=\gamma(t)+\left(\tfrac{\varepsilon-t}{\varepsilon}\right)(y-z).

A similar argument to (30) gives

0εL(α,α˙)𝑑t0εL(γ,γ˙)𝑑tLv(z,z˙)(yz)+2Aε|yz|2.\int_{0}^{\varepsilon}L(\alpha,{\dot{\alpha}})\,dt\leq\int_{0}^{\varepsilon}L(\gamma,{\dot{\gamma}})\,dt-L_{v}(z,{\dot{z}})(y-z)+\tfrac{2A}{\varepsilon}\,|y-z|^{2}.

Since uu is dominated we have that

u(γ(ε))\displaystyle u(\gamma(\varepsilon)) u(y)+0ε{L(α,α˙)+c(L)}\displaystyle\leq u(y)+\int_{0}^{\varepsilon}\Big{\{}L(\alpha,{\dot{\alpha}})+c(L)\Big{\}}
u(y)+0ε{L(γ,γ˙)+c(L)}𝑑tLv(z,z˙)(yz)+2Aε|yz|2.\displaystyle\leq u(y)+\int_{0}^{\varepsilon}\Big{\{}L(\gamma,{\dot{\gamma}})+c(L)\Big{\}}\,dt-L_{v}(z,{\dot{z}})(y-z)+\tfrac{2A}{\varepsilon}\,|y-z|^{2}.

Since uu is calibrated by γ|[0,ε]\gamma|_{[0,\varepsilon]} we have that

u(γ(ε))0ε{L(γ,γ˙)+c(L)}=u(z).u(\gamma(\varepsilon))-\int_{0}^{\varepsilon}\Big{\{}L(\gamma,{\dot{\gamma}})+c(L)\Big{\}}=u(z).

Thus we get the remaining inequality

u(z)u(y)Lv(z,z˙)(yz)+2Aε|yz|2.u(z)\leq u(y)-L_{v}(z,{\dot{z}})(y-z)+\tfrac{2A}{\varepsilon}\,|y-z|^{2}.

The set 𝒩(L){\mathcal{N}}(L) is hyperbolic for the Euler-Lagrange flow restricted to the energy level EL1{c(L)}E_{L}^{-1}\{c(L)\}. There is a neighborhood UU of 𝒩(L){\mathcal{N}}(L) in EL1{c(L)}E_{L}^{-1}\{c(L)\} such that the set

(31) Λ=+φt(U¯)\Lambda=\textstyle\bigcap_{-\infty}^{+\infty}\varphi_{t}(\overline{U})

is hyperbolic, cf. [17, prop. 5.1.8]. We can assume that 𝒜(L){\mathcal{A}}(L) has no periodic orbits. The neighborhood UU can be taken so small that any periodic orbit Γ\Gamma in Λ\Lambda has period

(32) per(Γ)>10.\operatorname{per}(\Gamma)>10.

For BTMB\subset TM write

c(B,𝒜(L)):=supθBd(θ,𝒜(L)).c(B,{\mathcal{A}}(L)):=\sup\nolimits_{\theta\in B}d(\theta,{\mathcal{A}}(L)).
4.3 Proposition.

For any ε>0\varepsilon>0 there is a periodic orbit ΓΛEL1{c(L)}\Gamma\subset\Lambda\subset E_{L}^{-1}\{c(L)\}, such that

(33) c(Γ,𝒜(L))<εγ(Γ)andAL+c(L)(Γ)<ε2γ(Γ)2,c(\Gamma,{\mathcal{A}}(L))<\varepsilon\,\gamma(\Gamma)\quad\text{and}\quad A_{L+c(L)}(\Gamma)<\varepsilon^{2}\,\gamma(\Gamma)^{2},

where   γ(Γ):=min{dTM(Γ(s),Γ(t)):|st|modper(Γ)1}\gamma(\Gamma):=\min\{d_{TM}(\Gamma(s),\Gamma(t)):|s-t|_{\rm mod\operatorname{per}(\Gamma)}\geq 1\,\}.

Proof:.

Let T>0T>0 be very large which will be chosen at the end of the proof. Let {ξi}i=0P1\{\xi_{i}\}_{i=0}^{P-1}, ξi(t)=φtti(θi)\xi_{i}(t)=\varphi_{t-t_{i}}(\theta_{i}), t[ti,ti+1[t\in[t_{i},t_{i+1}[ be the periodic specification from proposition 3.2. Define (x0,x˙0):𝒜(L)(x_{0},{\dot{x}}_{0}):{\mathbb{R}}\to{\mathcal{A}}(L) by x0(t)=π(ξi(t))x_{0}(t)=\pi(\xi_{i}(t)) if t[ti,ti+1[t\in[t_{i},t_{i+1}[ and x0(s+tPt0)=x0(s)x_{0}(s+t_{P}-t_{0})=x_{0}(s).

We will use repeatedly the constants from appendix A applied to the hyperbolic set Λ\Lambda from (31). We will show that if TT is chosen sufficiently large then the objects at each step are specifications and periodic orbits inside111Because they are (segments of) periodic orbits Γ\Gamma with c(Γ,𝒜(L))c(\Gamma,{\mathcal{A}}(L)) small, and hence ΓU\Gamma\subset U. Observe that Λ\Lambda is not necessarily locally maximal, then a priori shadowing objects could be outside Λ\Lambda. Λ\Lambda.

By the shadowing theorem A.13, there is a periodic Euler-Lagrange solution (y0,y˙0)(y_{0},{\dot{y}}_{0}) with energy c(L)c(L) and a continuous reparametrization σ(t)\sigma(t), with |σ(t)t|ECeλT|\sigma(t)-t|\leq EC\,\text{\rm\large e}^{-\lambda T} such that

td([x0(t),x˙0(t)],[y0(σ(t)),y˙0(σ(t))])<ECeλT.\forall t\qquad d\big{(}[x_{0}(t),{\dot{x}}_{0}(t)],[y_{0}(\sigma(t)),{\dot{y}}_{0}(\sigma(t))]\big{)}<E\cdot C\,\text{\rm\large e}^{-\lambda T}.

Then Y0(s):=(y0(s),y˙0(s))Y_{0}(s):=(y_{0}(s),{\dot{y}}_{0}(s)) is a periodic orbit with a period near σ(tP)σ(t0)\sigma(t_{P})-\sigma(t_{0}). We want a sequence of times sks_{k} nearby σ(tk)\sigma(t_{k}) such that sPs0s_{P}-s_{0} is a period for Y0(s)Y_{0}(s). Using canonical coordinates from A.3 define wkw^{k}\in{\mathbb{R}} small by

Y0(σ(tk)),θk\displaystyle\langle Y_{0}(\sigma(t_{k})),\theta_{k}\rangle =Wγs(Y0(σ(tk)))Wγuu(θk)\displaystyle=W^{s}_{\gamma}(Y_{0}(\sigma(t_{k})))\cap W^{uu}_{\gamma}(\theta_{k})
=Wγss(φwk(Y0(σ(tk))))Wγuu(θk).\displaystyle=W^{ss}_{\gamma}\big{(}\varphi_{w_{k}}(Y_{0}(\sigma(t_{k})))\big{)}\cap W_{\gamma}^{uu}(\theta_{k})\neq\emptyset.

Now let sk:=wk+σ(tk)s_{k}:=w_{k}+\sigma(t_{k}). Observe that the time shift wkw_{k} is determined by the sequence θk\theta_{k} which is periodic. Then the sequence sks_{k} is periodic with the period sPs0s_{P}-s_{0} of Y0Y_{0} and by proposition 3.2,

(34) per(y0):=period(y0)5TPT.\operatorname{per}(y_{0}):={\rm period}(y_{0})\leq 5TP_{T}.

By proposition A.7 there is D>0D>0 such that for TT large enough there are vk0v^{0}_{k} such that

(35) |vk0|DECeλT,\displaystyle|v^{0}_{k}|\leq DE\cdot C\text{\rm\large e}^{-\lambda T},
(36) s[sk,sk+1]d(Y0(s),φssi+vk0(θi))DECeλTeλmin{ssk,sk+1s}.\displaystyle\forall s\in[s_{k},s_{k+1}]\quad d\big{(}Y^{0}(s),\varphi_{s-s_{i}+v^{0}_{k}}(\theta_{i})\big{)}\leq DE\cdot C\,\text{\rm\large e}^{-\lambda T}\,\text{\rm\large e}^{-\lambda\min\{s-s_{k},\,s_{k+1}-s\}}.

Let zk0(s):=πφssi+vk0(θi)z^{0}_{k}(s):=\pi\varphi_{s-s_{i}+v^{0}_{k}}(\theta_{i}), s[sk,sk+1]s\in[s_{k},s_{k+1}]. Since by 2.4 𝒜(L){\mathcal{A}}(L) is invariant, we also have that (zk0,z˙k0)𝒜(L)(z_{k}^{0},{\dot{z}}_{k}^{0})\in{\mathcal{A}}(L).

By adding a constant to LL we can assume that

(37) c(L)=0.c(L)=0.

On local charts we have that

L(y0,y˙0)L(zk0,z˙k0)\displaystyle L(y_{0},{\dot{y}}_{0})\leq L(z^{0}_{k},{\dot{z}}^{0}_{k}) +xL(zk0,z˙k0)(y0zk0)+vL(zk0,z˙k0)(y˙0z˙k0)\displaystyle+\partial_{x}L(z_{k}^{0},{\dot{z}}_{k}^{0})(y_{0}-z^{0}_{k})+\partial_{v}L(z^{0}_{k},{\dot{z}}^{0}_{k})({\dot{y}}_{0}-{\dot{z}}^{0}_{k})
+K1d([y0(s),y˙0(s)],[zk0(s),z˙k0(s)])2.\displaystyle+K_{1}\,d\big{(}[y_{0}(s),{\dot{y}}_{0}(s)],[z^{0}_{k}(s),{\dot{z}}^{0}_{k}(s)]\big{)}^{2}.

Using that zk0z^{0}_{k} is an Euler-Lagrange solution we obtain

sksk+1L(y0,y˙0)\displaystyle\int_{s_{k}}^{s_{k+1}}L(y_{0},{\dot{y}}_{0}) [sksk+1L(zk0,z˙k0)]+vL(zk0,z˙k0)(y0zk0)|sksk+1+\displaystyle\leq\left[\int_{s_{k}}^{s_{k+1}}L(z^{0}_{k},{\dot{z}}^{0}_{k})\right]+\partial_{v}L(z_{k}^{0},{\dot{z}}_{k}^{0})(y_{0}-z_{k}^{0})\Big{|}_{s_{k}}^{s_{k+1}}+
(38) +K1sksk+1d([y0(s),y˙0(s)],[zk0(s),z˙k0(s)])2.\displaystyle+K_{1}\int_{s_{k}}^{s_{k+1}}d\big{(}[y_{0}(s),{\dot{y}}_{0}(s)],[z^{0}_{k}(s),{\dot{z}}^{0}_{k}(s)]\big{)}^{2}.

Write Zk0:=(zk0,z˙k0)Z^{0}_{k}:=(z^{0}_{k},{\dot{z}}^{0}_{k}). Then

AL(y0)\displaystyle A_{L}(y_{0})\leq k=0PT1AL(zk0)+\displaystyle\sum_{k=0}^{P_{T}-1}A_{L}(z^{0}_{k})\;+
(39) +\displaystyle+ k=0PT1{vL(zk0,z˙k0)(y0zk0)|sk+1vL(zk+10,z˙k+10)(y0zk+10)|sk+1}\displaystyle\sum_{k=0}^{P_{T}-1}\Big{\{}\partial_{v}L(z^{0}_{k},{\dot{z}}^{0}_{k})(y_{0}-z^{0}_{k})\Big{|}_{s_{k+1}}-\partial_{v}L(z^{0}_{k+1},{\dot{z}}^{0}_{k+1})(y_{0}-z^{0}_{k+1})\Big{|}_{s_{k+1}}\Big{\}}
+\displaystyle+ K1k=0PT1sksk+1d(Y0,Zk0)2𝑑s.\displaystyle K_{1}\sum_{k=0}^{P_{T}-1}\int_{s_{k}}^{s_{k+1}}d(Y_{0},Z^{0}_{k})^{2}\,ds.

From (36), for K2:=K1DECK_{2}:=K_{1}DEC, the last term satisfies

(40) K1k=0PT1sksk+1d(Y0,Zk0)2𝑑sPTK2e2λT.K_{1}\sum_{k=0}^{P_{T}-1}\int_{s_{k}}^{s_{k+1}}d(Y_{0},Z^{0}_{k})^{2}\,ds\leq P_{T}\,K_{2}\,\text{\rm\large e}^{-2\lambda T}.

Let uu be a dominated function. By Lemma 4.2, if (z,z˙)𝒜(L)(z,{\dot{z}})\in{\mathcal{A}}(L) is a static vector then

(41) |u(y)u(z)vL(z,z˙)(yz)|K3|yz|2.\big{|}u(y)-u(z)-\partial_{v}L(z,{\dot{z}})(y-z)\big{|}\leq K_{3}\,|y-z|^{2}.

By Lemma 4.1, uu is necessarily calibrated by static curves. Then using (41),

k=0PT1AL(zk0)\displaystyle\sum_{k=0}^{P_{T}-1}A_{L}(z^{0}_{k}) =ku(zk0(sk+1))u(zk0(sk))\displaystyle=\sum_{k}u(z^{0}_{k}(s_{k+1}))-u(z^{0}_{k}(s_{k}))
=ku(zk0(sk+1))u(zk+10(sk+1))\displaystyle=\sum_{k}u(z^{0}_{k}(s_{k+1}))-u(z^{0}_{k+1}(s_{k+1}))
=k{u(zk0)u(y0)+u(y0)u(zk+10)}|sk+1\displaystyle=\sum_{k}\big{\{}u(z^{0}_{k})-u(y_{0})+u(y_{0})-u(z^{0}_{k+1})\big{\}}\Big{|}_{s_{k+1}}
k{vL(zk0,z˙k0)(zk0y0)+vL(zk+10,z˙k+10)(y0zk+10)}|sk+1+\displaystyle\leq\sum_{k}\Big{\{}\partial_{v}L(z^{0}_{k},{\dot{z}}^{0}_{k})(z^{0}_{k}-y_{0})+\partial_{v}L(z^{0}_{k+1},{\dot{z}}^{0}_{k+1})(y_{0}-z^{0}_{k+1})\Big{\}}\Big{|}_{s_{k+1}}+
(42) +K3{|zk0y0|2+|y0zk+10|2}|sk+1.\displaystyle\hskip 56.9055pt+K_{3}\,\Big{\{}|z^{0}_{k}-y_{0}|^{2}+|y_{0}-z^{0}_{k+1}|^{2}\Big{\}}\Big{|}_{s_{k+1}}.

From (36) the last term satisfies

(43) k=0PT1K3{|zk0y0|2+|y0zk+10|2}|sk+1PTK4e2λT.\sum_{k=0}^{P_{T}-1}K_{3}\,\Big{\{}|z^{0}_{k}-y_{0}|^{2}+|y_{0}-z^{0}_{k+1}|^{2}\Big{\}}\Big{|}_{s_{k+1}}\leq P_{T}\,K_{4}\,\text{\rm\large e}^{-2\lambda T}.

Replacing estimate (42) for kAL(zk0)\sum_{k}A_{L}(z_{k}^{0}) in inequality (4) we obtain

(44) AL(y0)k=0PT1{K1sksk+1d(Y0,Zk0)2𝑑s+K3(|zk0y0|sk2+|zk0y0|sk+12)}.\displaystyle A_{L}(y_{0})\leq\sum_{k=0}^{P_{T}-1}\Big{\{}K_{1}\int_{s_{k}}^{s_{k+1}}d(Y_{0},Z^{0}_{k})^{2}\,ds+K_{3}\,\big{(}|z^{0}_{k}-y_{0}|^{2}_{s_{k}}+|z^{0}_{k}-y_{0}|^{2}_{s_{k}+1}\big{)}\Big{\}}.
Using (40) and (43) we have that
(45) AL(y0)sum in (44)K5PTe2λT=:A1(T).\displaystyle A_{L}(y_{0})\leq\text{sum in \eqref{ALY0int}}\leq K_{5}\,P_{T}\,\text{\rm\large e}^{-2\lambda T}=:A_{1}(T).

From (36) we get

(46) c(Y0,𝒜(L))<DECeλT.c(Y_{0},{\mathcal{A}}(L))<DE\cdot C\,\text{\rm\large e}^{-\lambda T}.

We can choose in (45) K5>(DEC)2K_{5}>(DEC)^{2}, so that

(47) max{AL(y0)12,c(Y0,𝒜(L))}<A1(T)12.\max\{A_{L}(y_{0})^{\frac{1}{2}},\,c(Y_{0},{\mathcal{A}}(L))\}<A_{1}(T)^{\frac{1}{2}}.

Also from (35),

(48) |vk0|A1(T)12.|v^{0}_{k}|\leq A_{1}(T)^{\frac{1}{2}}.

If Γ=Y0\Gamma=Y_{0} satisfies (33) then the proof finishes.

If Γ=Y0\Gamma=Y_{0} does not satisfy (33) then there are r1,r2r_{1},r_{2}, |r1r2|mod(sPs0)1|r_{1}-r_{2}|_{{\rm mod}\,(s_{P}-s_{0})}\geq 1 such that

(49) εd(Y0(r1),Y0(r2))\displaystyle\varepsilon\,d(Y_{0}(r_{1}),Y_{0}(r_{2})) c(Y0,𝒜(L)),or\displaystyle\leq c(Y_{0},{\mathcal{A}}(L)),\qquad\text{or}
ε2d(Y0(r1),Y0(r2))2\displaystyle\varepsilon^{2}\,d(Y_{0}(r_{1}),Y_{0}(r_{2}))^{2} AL(y0),\displaystyle\leq A_{L}(y_{0}),

using (37). Shifting the initial point of Y0Y_{0}, we can assume that r1<r2r_{1}<r_{2} and r2r112per(y0)r_{2}-r_{1}\leq\tfrac{1}{2}\operatorname{per}(y_{0}). If for some i,ji,j we have that |rjsi|1|r_{j}-s_{i}|\leq 1 we replace rjr_{j} by sis_{i} and shift the other rkr_{k} accordingly. By Gronwall’s inequality the distance d(Y0(r1),Y0(r2))d(Y_{0}(r_{1}),Y_{0}(r_{2})) increases at most by a multiple, say B0>1B_{0}>1. This insures that the times {r1,r2,s1,sP1}\{r_{1},r_{2},s_{1},\dots s_{P-1}\} are all separated (mod(sPs0))({\rm mod}\,(s_{P}-s_{0})) at least by 1. With this modification we get

(50) r2r112per(y0)+2.r_{2}-r_{1}\leq\tfrac{1}{2}\operatorname{per}(y_{0})+2.

In the following iteration process we will compare distances of a periodic orbit Yi(s)Y_{i}(s) with a time shifted periodic orbit Yi1(s+vi)Y_{i-1}(s+v^{i}). We will ensure in (73) that all the time shifts used are smaller than 1. We will take all the time shifts into account using Gronwall’s inequality by adding a multiple B0>1B_{0}>1 to the distance estimates. Write

(51) D0:=B0DE>1.D_{0}:=B_{0}\cdot DE>1.

Let Y1=(y1,y˙1)Y_{1}=(y_{1},{\dot{y}}_{1}) be the closed orbit which shadows the periodic specification Y0|[r1,r2]Y_{0}|_{[r_{1},r_{2}]}. By (46), for TT large, Y0ΛY_{0}\subset\Lambda and hence by (32), per(y0)>10\operatorname{per}(y_{0})>10. Then for R=54R=\frac{5}{4}, using (34), we have that

(52) per(y1)12per(y0)+3R1per(y0)R1(5TPT).\operatorname{per}(y_{1})\leq\tfrac{1}{2}\;\operatorname{per}(y_{0})+3\leq R^{-1}\operatorname{per}(y_{0})\leq R^{-1}(5TP_{T}).

By theorem A.13, proposition A.7 and (51), there is v1v^{1} such that

(53) |v1|D0d(Y0(r1),Y0(r2)),\displaystyle|v^{1}|\leq D_{0}\cdot d(Y_{0}(r_{1}),Y_{0}(r_{2})),
(54) s[r1,r2]d(Y1(s),Y0(s+v1))D0eλmin{sr1,r2s}d(Y0(r1),Y0(r2)),\displaystyle\forall s\in[r_{1},r_{2}]\qquad d(Y_{1}(s),Y_{0}(s+v^{1}))\leq D_{0}\,\text{\rm\large e}^{-\lambda\min\{s-r_{1},r_{2}-s\}}\,d(Y_{0}(r_{1}),Y_{0}(r_{2})),

From (49) and (47),

(55) d(Y0(r1),Y0(r2))ε1A1(T)12.d(Y_{0}(r_{1}),Y_{0}(r_{2}))\leq\varepsilon^{-1}\,A_{1}(T)^{\frac{1}{2}}.

Using (54), (55), (49), (47) and D0ε1>1D_{0}\,\varepsilon^{-1}>1,

c(Y1,𝒜(L))\displaystyle c(Y_{1},{\mathcal{A}}(L)) D0d(Y0(r1),Y0(r2))+c(Y0,𝒜(L))\displaystyle\leq D_{0}\cdot d(Y_{0}(r_{1}),Y_{0}(r_{2}))+c(Y_{0},{\mathcal{A}}(L))
D0ε1c(Y0,𝒜(L))+A1(T)12\displaystyle\leq D_{0}\cdot\varepsilon^{-1}c(Y_{0},{\mathcal{A}}(L))+A_{1}(T)^{\frac{1}{2}}
2D0ε1A1(T)12.\displaystyle\leq 2D_{0}\,\varepsilon^{-1}\,A_{1}(T)^{\frac{1}{2}}.
(56) c(Y1,𝒜(L))\displaystyle c(Y_{1},{\mathcal{A}}(L)) B4A1(T)12using (65).\displaystyle\leq B_{4}\,A_{1}(T)^{\frac{1}{2}}\qquad\text{using \eqref{B3}.}
(57) |v1|+|vk0|\displaystyle|v^{1}|+|v^{0}_{k}| B4A1(T)12using (53), (48).\displaystyle\leq B_{4}\,A_{1}(T)^{\frac{1}{2}}\qquad\text{using \eqref{v1}, \eqref{v0k2}.}

In order to estimate the action of Y1Y_{1} we need to compare it with a specification in 𝒜(L){\mathcal{A}}(L). Write zk1(s)=zk0(s+v1)z^{1}_{k}(s)=z^{0}_{k}(s+v^{1}) and Zk1=(zk1,z˙k1)Z^{1}_{k}=(z^{1}_{k},{\dot{z}}^{1}_{k}). We cut the specification {Zk1|[sk,sk+1]}\{Z^{1}_{k}|_{[s_{k},s_{k+1}]}\} at r1r_{1} and r2r_{2} and remain with the periodic specification of period r2r112per(y0)+2r_{2}-r_{1}\leq\tfrac{1}{2}\operatorname{per}(y_{0})+2, and jumps in r1<si<si+1<<sj<r2r_{1}<s_{i}<s_{i+1}<\ldots<s_{j}<r_{2} where si1r1<sis_{i-1}\leq r_{1}<s_{i} and sj<r2sj+1s_{j}<r_{2}\leq s_{j+1}.

For s[sk,sk+1]s\in[s_{k},s_{k+1}] we have that

d(Y1(s),Zk1(s)))\displaystyle d(Y_{1}(s),Z_{k}^{1}(s))) d(Y1(s),Y0(s+v1))+d(Y0(s+v1),Zk1(s)).\displaystyle\leq d(Y_{1}(s),Y_{0}(s+v^{1}))+d(Y_{0}(s+v^{1}),Z_{k}^{1}(s)).
(58) d(Y1(),Zk1())2\displaystyle d(Y_{1}(\cdot),Z_{k}^{1}(\cdot))^{2} 2d(Y1(),Y0())2+2d(Y0(),Zk1())2,\displaystyle\leq 2\,d(Y_{1}(\cdot),Y_{0}(\cdot))^{2}+2\,d(Y_{0}(\cdot),Z_{k}^{1}(\cdot))^{2},
(59) 2(D0)2e2λmin{sr1,r2s}ε2A1(T)+using (54), (49), (47)\displaystyle\leq 2\,(D_{0})^{2}\,\text{\rm\large e}^{-2\lambda\min\{s-r_{1},r_{2}-s\}}\varepsilon^{-2}\,A_{1}(T)\,+\qquad\text{using~{}\eqref{dY0Y1}, \eqref{dy012}, \eqref{A1Tmax} }
+2(D0C)2e2λTe2λmin{ssk,sk+1s},using (36), (51).\displaystyle\quad+2\,(D_{0}C)^{2}\,\text{\rm\large e}^{-2\lambda T}\,\text{\rm\large e}^{-2\lambda\min\{s-s_{k},\,s_{k+1}-s\}},\qquad\text{using \eqref{dx0y0}, \eqref{D0}.}

where the omitted arguments in (58) are the same as in the previous inequality.

Repeating the estimates in (38), (4),  (42) for the intervals between r1,si,,sj,r2r_{1},s_{i},\ldots,s_{j},r_{2}, we get

(60) AL(y1)\displaystyle A_{L}(y_{1}) r1sk<r2{K1sksk+1d(Y1,Zk1)2𝑑s+K3(|y1zk1|sk2+|y1zk|sk+12)}.\displaystyle\leq\sum_{r_{1}\leq s_{k}<r_{2}}\Big{\{}K_{1}\int_{s_{k}}^{s_{k+1}}d(Y_{1},Z_{k}^{1})^{2}\,ds+K_{3}\big{(}|y_{1}-z_{k}^{1}|^{2}_{s_{k}}+|y_{1}-z_{k}|^{2}_{s_{k+1}}\big{)}\Big{\}}.

Using (58) we can separate the sums in (60) into two sums. The sums with terms 2d(Y0,Zk1)22\,d(Y_{0},Z_{k}^{1})^{2} or 2|y0zk1|22\,|y_{0}-z_{k}^{1}|^{2} are about half of the terms in (44), with a shift of v1v^{1}, plus a term for the new jump at (r1,r2)(r_{1},r_{2}). The total number of jumps is 12PT+1PT\leq\frac{1}{2}P_{T}+1\leq P_{T}, then the same estimate (45) gives:

(61) r1sk<r2sksk+12K1d(Y0,Zk)2𝑑s+2K3(|y0zk1|sk2+|y0zk1|sk+12)2A1(T).\displaystyle\sum_{r_{1}\leq s_{k}<r_{2}}\int_{s_{k}}^{s_{k+1}}2K_{1}\,d(Y_{0},Z_{k})^{2}\,ds+2K_{3}\big{(}|y_{0}-z_{k}^{1}|^{2}_{s_{k}}+|y_{0}-z^{1}_{k}|^{2}_{s_{k+1}}\big{)}\leq 2\,A_{1}(T).

The other sum uses terms with 2d(Y1(s),Y0(s+v1))22\,d(Y_{1}(s),Y_{0}(s+v^{1}))^{2} which are bounded in (59). Abbreviating the time shift v1v^{1}, this sum writes

r1sk<r2\displaystyle\sum_{r_{1}\leq s_{k}<r_{2}} sksk+12K1d(Y1,Y0)2𝑑s+2K3(|y1y0|sk2+|y1y0|sk+12)\displaystyle\int_{s_{k}}^{s_{k+1}}2K_{1}\,d(Y_{1},Y_{0})^{2}\,ds+2K_{3}\big{(}|y_{1}-y_{0}|^{2}_{s_{k}}+|y_{1}-y_{0}|^{2}_{s_{k+1}}\big{)}
r1r22K1d(Y1,Y0)2𝑑s+r1sk<r22K3(|y1y0|sk2+|y1y0|sk+12)\displaystyle\leq\int_{r_{1}}^{r_{2}}2K_{1}\,d(Y_{1},Y_{0})^{2}\,ds+\sum_{r_{1}\leq s_{k}<r_{2}}2K_{3}\big{(}|y_{1}-y_{0}|^{2}_{s_{k}}+|y_{1}-y_{0}|^{2}_{s_{k+1}}\big{)}
2K1(D0)2B1ε2A1(T)+2K3(D0)2 2B2ε2A1(T)\displaystyle\leq 2K_{1}\,(D_{0})^{2}B_{1}\,\varepsilon^{-2}\,A_{1}(T)+2K_{3}\,(D_{0})^{2}\,2B_{2}\,\varepsilon^{-2}A_{1}(T)
(62) B3A1(T),\displaystyle\leq B_{3}\,A_{1}(T),

using (54), (55), where

B1:=+e2λ|s|𝑑s=1λ,B2:=1+ne2λn,\displaystyle\textstyle B_{1}:=\int_{-\infty}^{+\infty}\text{\rm\large e}^{-2\lambda|s|}ds=\frac{1}{\lambda},\qquad B_{2}:=1+\sum_{n\in{\mathbb{N}}}\text{\rm\large e}^{-2\lambda n},
(63) B3=B3(ε):=2(K1+K3)(D0)2(B1+2B2)ε2.\displaystyle B_{3}=B_{3}(\varepsilon):=2\,(K_{1}+K_{3})\,(D_{0})^{2}(B_{1}+2B_{2})\,\varepsilon^{-2}.

Adding (61) and (62) we get

(64) AL(y1)sum in (60)B4A1(T),A_{L}(y_{1})\leq\text{sum in \eqref{aly1k3}}\leq B_{4}\,A_{1}(T),

where

(65) B4:=max{B3+4,2D0ε1}>4.B_{4}:=\max\{B_{3}+4,2D_{0}\,\varepsilon^{-1}\}>4.

If Y1=ΓY_{1}=\Gamma satisfies (33) the proof finishes. If not there are r3<r4r_{3}<r_{4}, r4r312per(y1)+2r_{4}-r_{3}\leq\tfrac{1}{2}\operatorname{per}(y_{1})+2, such that

d(Y1(r3),Y1(r4))\displaystyle d(Y_{1}(r_{3}),Y_{1}(r_{4})) ε1max{AL(y1)12,c(Y1,𝒜(L))}\displaystyle\leq\varepsilon^{-1}\max\{A_{L}(y_{1})^{\frac{1}{2}},c(Y_{1},{\mathcal{A}}(L))\}
(66) ε1B4A1(T)12using (64), (56).\displaystyle\leq\varepsilon^{-1}B_{4}\,A_{1}(T)^{\frac{1}{2}}\qquad\text{using \eqref{AY1}, \eqref{cY1A}.}

We shadow the specification Y1|[r3,r4]Y_{1}|_{[r_{3},r_{4}]} by a periodic orbit Y2=(y2,y˙2)Y_{2}=(y_{2},{\dot{y}}_{2}) with

(67) |v2|D0d(Y1(r3),Y1(r4)),\displaystyle|v^{2}|\leq D_{0}\cdot d(Y_{1}(r_{3}),Y_{1}(r_{4})),
s[r3,r4]d(Y2(s),Y1(s+v2))D0eλmin{sr3,r4s}d(Y1(r3),Y1(r4)),\displaystyle\forall s\in[r_{3},r_{4}]\qquad d(Y_{2}(s),Y_{1}(s+v^{2}))\leq D_{0}\,\text{\rm\large e}^{-\lambda\min\{s-r_{3},r_{4}-s\}}\,d(Y_{1}(r_{3}),Y_{1}(r_{4})),
per(y2)12per(y1)+3R2per(y0).\displaystyle\operatorname{per}(y_{2})\leq\tfrac{1}{2}\operatorname{per}(y_{1})+3\leq R^{-2}\operatorname{per}(y_{0}).

Then

c(Y2,𝒜(L))\displaystyle c(Y_{2},{\mathcal{A}}(L)) D0d(Y1(r3),Y1(r4))+c(Y1,𝒜(L))\displaystyle\leq D_{0}\cdot d(Y_{1}(r_{3}),Y_{1}(r_{4}))+c(Y_{1},{\mathcal{A}}(L))
2D0ε1B4A1(T)12using (66), (56)\displaystyle\leq 2D_{0}\,\varepsilon^{-1}B_{4}\,A_{1}(T)^{\frac{1}{2}}\qquad\text{using \eqref{cr3r4}, \eqref{cY1A}}
(B4)2A1(T)12using (65).\displaystyle\leq(B_{4})^{2}\,A_{1}(T)^{\frac{1}{2}}\qquad\text{using \eqref{B3}.}
|v2|+|v1|+|vk0|\displaystyle|v^{2}|+|v^{1}|+|v^{0}_{k}| (B4)2A1(T)12similarly, using (67), (57).\displaystyle\leq(B_{4})^{2}\,A_{1}(T)^{\frac{1}{2}}\qquad\text{similarly, using \eqref{v2}, \eqref{v1v0k}.}

We need to compare Y2Y_{2} with a specification in 𝒜(L){\mathcal{A}}(L). Write zk2(s)=zk1(s+v2)z_{k}^{2}(s)=z_{k}^{1}(s+v^{2}) and Zk2=(zk2,z˙k2)Z_{k}^{2}=(z_{k}^{2},{\dot{z}}_{k}^{2}). Then

(68) AL(y2)\displaystyle A_{L}(y_{2}) r3k<r4sksk+1K1d(Y2,Zk2)2𝑑s+K3(|y2zk2|sk2+|y2zk2|sk+12).\displaystyle\leq\sum_{r_{3}\leq k<r_{4}}\int_{s_{k}}^{s_{k+1}}K_{1}\,d(Y_{2},Z_{k}^{2})^{2}\,ds+K_{3}\,\big{(}|y_{2}-z_{k}^{2}|_{s_{k}}^{2}+|y_{2}-z_{k}^{2}|_{s_{k+1}}^{2}\big{)}.
(69) d(Y2,Zk2)22d(Y2,Y1)2+2d(Y1,Zk2)2.d(Y_{2},Z^{2}_{k})^{2}\leq 2\,d(Y_{2},Y_{1})^{2}+2\,d(Y_{1},Z^{2}_{k})^{2}.
(70) r3r42K1d(Y2,\displaystyle\int_{r_{3}}^{r_{4}}2K_{1}\,d(Y_{2}, Y1)2ds+r3sk<r42K3(|y2y1|sk2+|y2y1|sk+12)\displaystyle Y_{1})^{2}\,ds+\sum_{r_{3}\leq s_{k}<r_{4}}2K_{3}\big{(}|y_{2}-y_{1}|^{2}_{s_{k}}+|y_{2}-y_{1}|^{2}_{s_{k+1}}\big{)}
{2K1B1(D0)2+2K3(D0)2 2B2}d(Y1(r3),Y1(r4))2\displaystyle\leq\big{\{}2K_{1}\,B_{1}\,(D_{0})^{2}+2K_{3}\,(D_{0})^{2}\,2B_{2}\big{\}}\;d(Y_{1}(r_{3}),Y_{1}(r_{4}))^{2}
2(K1+K3)(D0)2(B1+2B2)ε2(B4)2A1(T)using (66),\displaystyle\leq 2(K_{1}+K_{3})(D_{0})^{2}(B_{1}+2B_{2})\,\varepsilon^{-2}(B_{4})^{2}A_{1}(T)\qquad\text{using~{}\eqref{cr3r4}},
(71) B3(B4)2A1(T)using (63).\displaystyle\leq B_{3}\,(B_{4})^{2}\,A_{1}(T)\qquad\text{using~{}\eqref{B2}.}

From (69) and (68) we have that

AL(y2)\displaystyle AL(y_{2}) sum in (70)+2sum in (60)\displaystyle\leq\text{sum in~{}\eqref{sum30}}+2\,\text{sum in~{}\eqref{aly1k3}}
(B4)3A1(T)using (71), (64), (65),\displaystyle\leq(B_{4})^{3}\,A_{1}(T)\qquad\text{using \eqref{B2B3}, \eqref{AY1}, \eqref{B3},}
(B4)4A1(T).\displaystyle\leq(B_{4})^{4}A_{1}(T).
per(y2)\displaystyle\operatorname{per}(y_{2}) R2per(y0)R2(5TPT).\displaystyle\leq R^{-2}\operatorname{per}(y_{0})\leq R^{-2}(5TP_{T}).

At the nn-th iteration we have

AL(yn)\displaystyle A_{L}(y_{n}) (B4)2nA1(T),\displaystyle\leq(B_{4})^{2n}\,A_{1}(T),
(72) per(yn)\displaystyle\operatorname{per}(y_{n}) Rnper(y0)Rn(5TPT).\displaystyle\leq R^{-n}\operatorname{per}(y_{0})\leq R^{-n}(5TP_{T}).
c(Yn,𝒜(L))(B4)nA1(T)12.\displaystyle c(Y_{n},{\mathcal{A}}(L))\leq(B_{4})^{n}A_{1}(T)^{\frac{1}{2}}.
|vk0|+i=1n|vi|(B4)nA1(T)12.\displaystyle|v^{0}_{k}|+\operatorname*{{\textstyle\sum}}_{i=1}^{n}|v^{i}|\leq(B_{4})^{n}A_{1}(T)^{\frac{1}{2}}.

Let α2>0\alpha_{2}>0 be such that {θTM:d(θ,𝒜(L))<α2}U\{\,\theta\in T^{*}M:d(\theta,{\mathcal{A}}(L))<\alpha_{2}\,\}\subset U, where UU is from (31). This process can be repeated as long as c(Yn,𝒜(L))<α2c(Y_{n},{\mathcal{A}}(L))<\alpha_{2} holds and (33) is not satisfied. The resulting periodic YnY_{n} is in Λ\Lambda and hence by (32) it has period larger than 10. Thus the process stops at an iterate NN where the period in (72) is larger than 1. This is

N\displaystyle N\leq logRper(y0)logR(5TPT),\displaystyle\log_{R}\operatorname{per}(y_{0})\leq\log_{R}(5TP_{T}),
AL(yN)(B4)2NA1(T)\displaystyle A_{L}(y_{N})\leq(B_{4})^{2N}A_{1}(T) (5TPT)2logRB4K5PTe2λTusing (45),\displaystyle\leq(5TP_{T})^{2\log_{R}B_{4}}\cdot K_{5}\,P_{T}\,\text{\rm\large e}^{-2\lambda T}\qquad\text{using \eqref{A1T},}
c(YN,𝒜(L))\displaystyle c(Y_{N},{\mathcal{A}}(L)) (5TPT)logRB4K5PTeλT,\displaystyle\leq(5TP_{T})^{\log_{R}B_{4}}\sqrt{K_{5}P_{T}}\;\text{\rm\large e}^{-\lambda T},
(73) |vk0|+i=1n|vi|\displaystyle|v^{0}_{k}|+\operatorname*{{\textstyle\sum}}_{i=1}^{n}|v^{i}| (5TPT)logRB4K5PTeλT.\displaystyle\leq(5TP_{T})^{\log_{R}B_{4}}\sqrt{K_{5}P_{T}}\;\text{\rm\large e}^{-\lambda T}.

Since by (16), PTP_{T} has sub-exponential growth in TT, we have that c(YN,𝒜(L))c(Y_{N},{\mathcal{A}}(L)) and AL(yN)A_{L}(y_{N}) can be made arbitrarily small by choosing TT sufficiently large. Then the process stops not because c(YN,𝒜(L))c(Y_{N},{\mathcal{A}}(L)) is large, but because (33) holds.

5. The perturbed minimizers.

The following Crossing Lemma is extracted for Mather [25] with the observation that the estimates can be taken uniformly on a C2C^{2} neighbourhood of LL.

5.1 Lemma (Mather [25, p. 186]).

If K>0K>0, then there exist ε\varepsilon, δ\delta, η\eta, ζ>0\zeta>0 and

(74) C>1,C>1,

such that if ϕC2<ζ\left\|\phi\right\|_{C^{2}}<\zeta, and α,γ:[t0ε,t0+ε]M\alpha,\gamma:[t_{0}-\varepsilon,t_{0}+\varepsilon]\to M are solutions of the Euler-Lagrange equation for L+ϕL+\phi with dα(t0)\left\|d\alpha(t_{0})\right\|, dγ(t0)K\left\|d\gamma(t_{0})\right\|\leq K, d(α(t0),γ(t0))δd\big{(}\alpha(t_{0}),\gamma(t_{0})\big{)}\leq\delta, and

d(dα(t0),dγ(t0))Cd(α(t0),γ(t0)),d\big{(}d\alpha(t_{0}),d\gamma(t_{0})\big{)}\geq C\;d\big{(}\alpha(t_{0}),\gamma(t_{0})\big{)},

then there exist C1C^{1} curves a,c:[t0ε,t0+ε]Ma,c:[t_{0}-\varepsilon,t_{0}+\varepsilon]\to M such that a(t0ε)=α(t0ε)a(t_{0}-\varepsilon)=\alpha(t_{0}-\varepsilon), a(t0+ε)=γ(t0+ε)a(t_{0}+\varepsilon)=\gamma(t_{0}+\varepsilon), c(t0ε)=γ(t0ε)c(t_{0}-\varepsilon)=\gamma(t_{0}-\varepsilon), c(t0+ε)=α(t0+ε)c(t_{0}+\varepsilon)=\alpha(t_{0}+\varepsilon), and

(75) AL+ϕ(α)+AL+ϕ(γ)AL+ϕ(a)AL+ϕ(c)>ηd(dα(t0),dγ(t0))2.A_{L+\phi}(\alpha)+A_{L+\phi}(\gamma)-A_{L+\phi}(a)-A_{L+\phi}(c)>\eta\;d\big{(}d\alpha(t_{0}),d\gamma(t_{0})\big{)}^{2}.
5.2 Lemma.

Given a Tonelli lagrangian L0L_{0} and a compact subset ΔTM\Delta\subset TM, there are ε>0\varepsilon>0, K>0K>0 and δ1>0\delta_{1}>0 such that for any Tonelli lagrangian LL with (LL0)|Bε(Δ)C2<ε\left\|(L-L_{0})|_{B_{\varepsilon}(\Delta)}\right\|_{C^{2}}<\varepsilon, Bε(Δ):={θTM:d(θ,Δ)<ε}B_{\varepsilon}(\Delta):=\{\theta\in TM:d(\theta,\Delta)<\varepsilon\}, and any T>0T>0:

  1. (a)

    If xC1([0,T],M)x\in C^{1}([0,T],M) is a solution of the Euler-Lagrange equation for LL with (x,x˙)Δ(x,{\dot{x}})\in\Delta and zC1([0,T],M)z\in C^{1}([0,T],M) satisfies

    d([z(t),z˙(t)],[x(t),x˙(t)])4ρδ1t[0,T],d\big{(}[z(t),{\dot{z}}(t)],[x(t),{\dot{x}}(t)]\big{)}\leq 4\rho\leq\delta_{1}\qquad\forall t\in[0,T],

    then

    (76) |0TL(z,z˙)dt0TL(x,x˙)dtvL(x,x˙)(zx)|0T|K(1+T)ρ2,\left|\int_{0}^{T}L(z,{\dot{z}})\,dt-\int_{0}^{T}L(x,{\dot{x}})\,dt-\partial_{v}L(x,{\dot{x}})\cdot(z-x)\Big{|}_{0}^{T}\right|\leq K\,(1+T)\,\rho^{2},

    where zx:=(expx)1(z)z-x:=(\exp_{x})^{-1}(z).

  2. (b)

    If xC1([0,T],M)x\in C^{1}([0,T],M) is a solution of the Euler-Lagrange equation for LL with (x,x˙)Δ(x,{\dot{x}})\in\Delta and the curves w1,w2,zC1([0,T],M)w_{1},\,w_{2},\,z\in C^{1}([0,T],M) satisfy w1(0)=x(0)w_{1}(0)=x(0), w1(T)=z(T)w_{1}(T)=z(T), w2(0)=z(0)w_{2}(0)=z(0), w2(T)=x(T)w_{2}(T)=x(T), and for all ξ{z,w1,w2}\xi\in\{z,\,w_{1},\,w_{2}\} we have

    d([ξ(t),ξ˙(t)],[x(t),x˙(t)])4ρδ1t[0,T],d\big{(}[\xi(t),{\dot{\xi}}(t)],[x(t),{\dot{x}}(t)]\big{)}\leq 4\rho\leq\delta_{1}\qquad\forall t\in[0,T],

    then

    |AL(x)+AL(z)AL(w1)AL(w2)|3Kρ2(1+T).\left|A_{L}(x)+A_{L}(z)-A_{L}(w_{1})-A_{L}(w_{2})\right|\leq 3K\rho^{2}(1+T).

Proof:

  1. (a)

    We use a coordinate system on a tubular neighbourhood of x([0,T])x([0,T]) with a bound in the C2C^{2} norm independent of TT and of x˙(0){\dot{x}}(0). In case xx has self-intersections or short returns the coordinate system is an immersion.

    We have that

    L(z,z˙)L(x,x˙)=xL(x,x˙)(zx)+vL(x,x˙)(z˙x˙)+O(ρ2),\displaystyle L(z,{\dot{z}})-L(x,{\dot{x}})=\partial_{x}L(x,{\dot{x}})(z-x)+\partial_{v}L(x,{\dot{x}})({\dot{z}}-{\dot{x}})+O(\rho^{2}),

    here O(ρ2)Kρ2O(\rho^{2})\leq K\,\rho^{2} where KK depends on the second derivatives of LL on a small neighbourhood of the compact Δ\Delta and hence it can be taken uniform on a C2C^{2} neighbourhood of LL. Since xx satisfies the Euler-Lagrange equation for LL,

    L(z,z˙)L(x,x˙)=ddt[vL(x,x˙)(zx)]+O(ρ2).\displaystyle L(z,{\dot{z}})-L(x,{\dot{x}})=\tfrac{d}{dt}[\partial_{v}L(x,{\dot{x}})(z-x)]+O(\rho^{2}).

    This implies (76).

  2. (b)

    By item (a)

    AL(w1)AL(x)\displaystyle A_{L}(w_{1})-A_{L}(x) vL(x,x˙)(w1x)|0T+Kρ2(1+T)\displaystyle\leq\partial_{v}L(x,{\dot{x}})(w_{1}-x)\Big{|}_{0}^{T}+K\rho^{2}(1+T)
    vL(x(T),x˙(T))(z(T)x(T))+Kρ2(1+T).\displaystyle\leq\partial_{v}L(x(T),{\dot{x}}(T))(z(T)-x(T))+K\rho^{2}(1+T).
    AL(w2)AL(x)\displaystyle A_{L}(w_{2})-A_{L}(x) vL(x(0),x˙(0))(z(0)x(0))+Kρ2(1+T).\displaystyle\leq-\partial_{v}L(x(0),{\dot{x}}(0))(z(0)-x(0))+K\rho^{2}(1+T).
    AL(x)AL(z)\displaystyle A_{L}(x)-A_{L}(z) vL(x,x˙)(zx)|0T+Kρ2(1+T)\displaystyle\leq-\partial_{v}L(x,{\dot{x}})(z-x)\Big{|}_{0}^{T}+K\rho^{2}(1+T)
    vL(x(T),x˙(T))(z(T)x(T))\displaystyle\leq-\partial_{v}L(x(T),{\dot{x}}(T))(z(T)-x(T))
    +vL(x(0),x˙(0))(z(0)x(0))+Kρ2(1+T).\displaystyle\hskip 10.0pt+\partial_{v}L(x(0),{\dot{x}}(0))(z(0)-x(0))+K\rho^{2}(1+T).

    Adding these inequalities we get

    AL(w1)+AL(w2)AL(x)AL(z)3Kρ2(1+T).A_{L}(w_{1})+A_{L}(w_{2})-A_{L}(x)-A_{L}(z)\leq 3K\rho^{2}(1+T).

    The remaining inequality is obtained similarly. ∎

The following proposition has its origin in Yuan and Hunt [31], the present proof uses some arguments by Quas and Siefken [29]. Proposition 5.3 together with proposition 4.3 imply theorem A.

5.3 Proposition.

Suppose that for every δ>0\delta>0 there is a periodic orbit ΓΛEL1{c(L)}\Gamma\subset\Lambda\subset E_{L}^{-1}\{c(L)\} such that

(77) c(Γ,𝒜(L))<δγ(Γ)andAL+c(L)(Γ)<δ2γ(Γ)2,c(\Gamma,{\mathcal{A}}(L))<\delta\,\gamma(\Gamma)\qquad\text{and}\qquad A_{L+c(L)}(\Gamma)<\delta^{2}\,\gamma(\Gamma)^{2},

where   γ(Γ):=min{dTM(Γ(s),Γ(t)):|st|mod(perΓ)1}\gamma(\Gamma):=\min\{d_{TM}(\Gamma(s),\Gamma(t)):|s-t|_{\rm mod(\operatorname{per}{\Gamma})}\geq 1\,\}.

Then for any ε>0\varepsilon>0 there is ϕC2(M,)\phi\in C^{2}(M,{\mathbb{R}}) with ϕC2<ε\left\|\phi\right\|_{C^{2}}<\varepsilon such that Γ𝒜(L+ϕ)\Gamma\subset{\mathcal{A}}(L+\phi), where Γ\Gamma is one of the periodic orbits in (77).

Idea of the Proof:

We choose δ=δ(ε)\delta=\delta(\varepsilon) sufficiently small and use the periodic orbit Γ\Gamma given by the hypothesis. We perturb the Lagrangian by a potential ϕ\phi which is a non-negative channel centered at π(Γ)\pi(\Gamma) defined in (96). The curve Γ\Gamma is a periodic orbit for the flows of LL and of L+ϕL+\phi. We show that Γ\Gamma is contained in the Aubry set 𝒜(L+ϕ){\mathcal{A}}(L+\phi) by proving that any semi-static curve x:],0]Mx:]-\infty,0]\to M for L+ϕL+\phi has

α-limit of (x,x˙)=Γ;\alpha\text{-limit of }(x,{\dot{x}})=\Gamma;

because by Mañé [24, Theorem V.(c)], α\alpha-limits of semi-static orbits are static. This is done by calculating the action of each segment of the semi-static which is spent outside of a small neighbourhood of Γ\Gamma, and proving that it has a uniform positive lower bound. Since the total action of a semi-static is finite, the quantity of those segments is finite. Thus the semi-static eventually stays forever in a small neighbourhood of Γ\Gamma. The expansivity of Λ(L+ϕ)𝒩(L+ϕ)\Lambda(L+\phi)\supset{\mathcal{N}}(L+\phi) implies that the α\alpha-limit of the semi-static is Γ\Gamma.

5.4 Lemma.

If 𝒜(L){\mathcal{A}}(L) has no periodic orbits and Γn\Gamma_{n} is a sequence of periodic orbits with

c(Γn,𝒜(L))<δndiam(𝒜(L)),\displaystyle c(\Gamma_{n},{\mathcal{A}}(L))<\delta_{n}\cdot\operatorname{diam}({\mathcal{A}}(L)),
γn:=min{d(Γn(s),Γn(t)):|st|mod(perΓn)1}.\displaystyle\gamma_{n}:=\min\{d(\Gamma_{n}(s),\Gamma_{n}(t)):|s-t|_{\rm{mod(per\,}\Gamma_{n})}\geq 1\,\}.

Then  limnδn=0\lim_{n}\delta_{n}=0\Longrightarrowlimnγn=0\lim_{n}\gamma_{n}=0.

Proof:.

Let TnT_{n} be the period of Γn\Gamma_{n}. First we prove that limnTn=\lim_{n}T_{n}=\infty. If not, we can extract a subsequence where θ:=limnΓn(0)𝒜(L)\theta:=\lim_{n}\Gamma_{n}(0)\in{\mathcal{A}}(L) and S:=limnTnS:=\lim_{n}T_{n} exist. Then θ\theta is a periodic point in 𝒜(L){\mathcal{A}}(L) which contradicts the hypothesis.

Consider the points Γn(4m)\Gamma_{n}(4m), 0mMn:=[14Tn]0\leq m\leq M_{n}:=[\tfrac{1}{4}T_{n}], mm\in{\mathbb{N}}. Since limnTn=\lim_{n}T_{n}=\infty, the quantity MnM_{n} of these points tends to infinity. Therefore

γnminm1m2d(Γn(m1),Γn(m2))𝑛0.\gamma_{n}\leq\min_{m_{1}\neq m_{2}}d(\Gamma_{n}(m_{1}),\Gamma_{n}(m_{2}))\overset{n}{\longrightarrow}0.


Proof of Proposition 5.3:

By adding a constant to LL we can assume that

(78) c(L)=0.c(L)=0.

Fix K1>0K_{1}>0 such that

(79) [ELc(L)+1][|v|K1].[E_{L}\leq c(L)+1]\subset[|v|\leq K_{1}].

Bernard [1] after Fathi and Siconolfi [16] proves that there is a C1+LipC^{1+\operatorname{Lip}} critical subsolution uu of the Hamilton-Jacobi equation for LL, H(x,dxu)c(L)H(x,d_{x}u)\leq c(L). Thus

(80) Ldu0.L-du\geq 0.

By Gronwall’s inequality and the continuity of Mañé’s critical value c(L)c(L) (see [14, Lemma 5.1]) there is α>0\alpha>0 and γ0\gamma_{0} such that if ϕC21\left\|\phi\right\|_{C^{2}}\leq 1, 0<γ<γ00<\gamma<\gamma_{0} and Γ\Gamma is a periodic orbit for L+ϕL+\phi with energy smaller than c(L+ϕ)+1c(L+\phi)+1 then

(81) d(φsL+ϕ(ϑ),Γ)γ4 and d(φtL+ϕ(ϑ),Γ)γ3|ts|>α.d(\varphi^{L+\phi}_{s}(\vartheta),\Gamma)\leq\frac{\gamma}{4}\quad\text{ and }\quad d(\varphi^{L+\phi}_{t}(\vartheta),\Gamma)\geq\frac{\gamma}{3}\quad\Longrightarrow\quad|t-s|>\alpha.

The graph property states that the projection π:𝒜(L)M\pi:{\mathcal{A}}(L)\to M has a Lipschitz inverse (see Mañé [24]). The Lipschitz constant is the same as CC in Mather’s Crossing Lemma 5.1. The Aubry set has energy c(L)c(L) and c(L+ϕ)c(L+\phi) is continuous on ϕ\phi. Then one can choose

(82) ε1<ζ\varepsilon_{1}<\zeta

and KK, C>1C>1 in Lemma 5.1 such that if ϕC2<ε1\left\|\phi\right\|_{C^{2}}<\varepsilon_{1} then 𝒜(L+ϕ){\mathcal{A}}(L+\phi) is a graph with Lipschitz constant CC.

By the upper semicontinuity of the Mañé set [14, lemma 5.2] we can choose a neighbourhood UU of 𝒩(L){\mathcal{N}}(L) and 0<ε2<ε10<\varepsilon_{2}<\varepsilon_{1} such that if ϕC2<ε2\left\|\phi\right\|_{C^{2}}<\varepsilon_{2} then the set

Λ(ϕ):=tφtL+ϕ(U¯)\Lambda(\phi):=\bigcap_{t\in{\mathbb{R}}}\varphi_{-t}^{L+\phi}(\overline{U})

is hyperbolic and contains 𝒩(L+ϕ){\mathcal{N}}(L+\phi). Take 0<ε3<ε20<\varepsilon_{3}<\varepsilon_{2} such that Λ(ϕ)\Lambda(\phi) has uniform constants of hyperbolicity (A.7), expansivity (A.8, A.9) and canonical coordinates (A.3) for all ϕC2<ε3<ε2\left\|\phi\right\|_{C^{2}}<\varepsilon_{3}<\varepsilon_{2}.

Write

(83) γδ:=γ(Γ).{\gamma_{\delta}}:=\gamma(\Gamma).

We can assume that 𝒜(L){\mathcal{A}}(L) has no periodic points. By lemma 5.4, γδ\gamma_{\delta} is small when δ\delta is small. Given 0<ε<ε30<\varepsilon<\varepsilon_{3}, choose 0<δε0<\delta\ll\varepsilon and a periodic orbit Γ\Gamma satisfying (77) with δ\delta and γδ\gamma_{\delta} so small that for all ϕC2<ε3\left\|\phi\right\|_{C^{2}}<\varepsilon_{3},

(84) γδ\displaystyle{\gamma_{\delta}} <ϵ0where ϵ0 is a flow expansivity constant for Λ(ϕ) as in A.8 and A.9.\displaystyle<\epsilon_{0}\hskip 6.0pt\text{where $\epsilon_{0}$ is a flow expansivity constant for $\Lambda(\phi)$ as in \ref{dfe} and \ref{rue}. }
(85) 2γδ\displaystyle 2{\gamma_{\delta}} <δ1with δ1:=δ[K1] from lemma 5.1, where K1 is from (79).\displaystyle<\delta_{1}\hskip 5.0pt\text{with }\delta_{1}:=\delta[K_{1}]\text{ from lemma~{}\ref{CL}, where $K_{1}$ is from~{}\eqref{dK1}. }
(86) γδ\displaystyle{\gamma_{\delta}} <β0where β0 is from proposition A.7 for Λ(ϕ).\displaystyle<\beta_{0}\hskip 4.0pt\text{where $\beta_{0}$ is from proposition~{}\ref{B71} for $\Lambda(\phi)$.}
(87) γδ\displaystyle{\gamma_{\delta}} <η0where η0 is from the canonical coordinates in A.4 for Λ(ϕ),\displaystyle<\eta_{0}\hskip 9.0pt\text{where $\eta_{0}$ is from the canonical coordinates in \ref{B4} for $\Lambda(\phi)$,}

and such that writing

(88) γ¯δ:=γδ3C(B+1)<12γδ,{\overline{\gamma}_{\delta}}:=\frac{{\gamma_{\delta}}}{3C(B+1)}<\tfrac{1}{2}\,{\gamma_{\delta}},

we have that

(89) γ¯δ<γ0 where γ0 is from (81),\displaystyle{\overline{\gamma}_{\delta}}<\gamma_{0}\hskip 56.9055pt\text{ where $\gamma_{0}$ is from~{}\eqref{tima},}

and there is ρ\rho,

(90) δγδ<ρ<14γ¯δ1\delta\,{\gamma_{\delta}}<\rho<\tfrac{1}{4}{\overline{\gamma}_{\delta}}\ll 1

such that

(91) 14ερ2>δ2(γδ)2,\displaystyle\tfrac{1}{4}\,\varepsilon\,\rho^{2}>\delta^{2}\,({\gamma_{\delta}})^{2},
(92) Cρ>1η1δγδ,\displaystyle C\rho>\tfrac{1}{\sqrt{\eta_{1}}}\,{\delta\,\gamma_{\delta}},
(93) (132ε(γ¯δ)2δ2(γδ)2)α6KD2C2(B+1)2ρ23δ2(γδ)2>0,\displaystyle\big{(}\tfrac{1}{32}\,\varepsilon\,({\overline{\gamma}_{\delta}})^{2}-\delta^{2}({\gamma_{\delta}})^{2}\big{)}\alpha-6KD^{2}C^{2}(B+1)^{2}\rho^{2}-3\,\delta^{2}({\gamma_{\delta}})^{2}>0,

where BB is from Lemma A.4, C=C[K1]C=C[K_{1}] and η1=η[K1]\eta_{1}=\eta[K_{1}] are from Lemma 5.1 with

(94) C>1,C>1,

DD is from Proposition A.7 and KK is from Lemma 5.2 applied to the compact Δ=[ELc(L)+5]\Delta=[E_{L}\leq c(L)+5]. Inequality (93) implies

(95) 132ε(γ¯δ)2>δ2(γδ)2.\tfrac{1}{32}\,\varepsilon\,({\overline{\gamma}_{\delta}})^{2}>\delta^{2}({\gamma_{\delta}})^{2}.

Let ϕ:M[0,1]\phi:M\to[0,1] be a CC^{\infty} function such that ϕC2<10ε\left\|\phi\right\|_{C^{2}}<10\,\varepsilon and

(96) 0ϕ(x)={0if xπ(Γ),14ερ2if d(x,πΓ)ρ,132ε(γ¯δ)2if d(x,πΓ)14γ¯δ.0\leq\phi(x)=\begin{cases}0&\text{if }x\in\pi(\Gamma),\\ \geq\tfrac{1}{4}\,\varepsilon\,\rho^{2}&\text{if }d(x,\pi\Gamma)\geq\rho,\\ \tfrac{1}{32}\,\varepsilon\,({\overline{\gamma}_{\delta}})^{2}&\text{if }d(x,\pi\Gamma)\geq\tfrac{1}{4}{\overline{\gamma}_{\delta}}.\end{cases}

Using uu from (80) write

(97) 𝕃:=L+ϕ+c(L+ϕ)du.{\mathbb{L}}:=L+\phi+c(L+\phi)-du.

The Euler-Lagrange flow of 𝕃{\mathbb{L}} and the sets 𝒜(𝕃){\mathcal{A}}({\mathbb{L}}), 𝒩(𝕃){\mathcal{N}}({\mathbb{L}}) are the same as those of L+ϕL+\phi. In particular the hyperbolicity constants (84)–(87) and Lipschitz graphs contants (94) remain valid for 𝕃{\mathbb{L}}.

Claim 5.4.1:

If δ\delta is small enough then

  1. (1)

    We have that

    infd(s,t)mod T1d(πΓ(s),πΓ(t))>34γ¯δ.\inf\limits_{d(s,t)_{\text{mod }T}\geq 1}d\big{(}\pi\Gamma(s),\pi\Gamma(t)\big{)}>\tfrac{3}{4}\,{\overline{\gamma}_{\delta}}.

    In particular the neighbourhood B(πΓ,38γ¯δ)B(\pi\Gamma,\tfrac{3}{8}{\overline{\gamma}_{\delta}}) of πΓ\pi\Gamma of radius 38γ¯δ\frac{3}{8}{\overline{\gamma}_{\delta}} has no self intersections, i.e. it is homeomorphic to S1×]0,1[dimM1S^{1}\times]0,1[^{\dim M-1}.

  2. (2)

    If x:],0]Mx:]-\infty,0]\to M is a semi-static orbit for 𝕃{\mathbb{L}} then for all t1t\leq-1

    (98) either d([x(t),x˙(t)],Γ)δγ(Γ)η1 or d([x(t),x˙(t)],Γ)Cd(x(t),πΓ),\displaystyle\text{either }\quad d\big{(}[x(t),{\dot{x}}(t)],\Gamma\big{)}\leq\tfrac{\delta\,\gamma(\Gamma)}{\sqrt{\eta_{1}}}\quad\text{ or }\quad d\big{(}[x(t),{\dot{x}}(t)],\Gamma\big{)}\leq C\,d\big{(}x(t),\pi\Gamma\big{)},
    (99) or d(x(t),πΓ)δ1,\displaystyle\text{or }\quad d(x(t),\pi\Gamma)\geq\delta_{1},

    where η1=η(K1)\eta_{1}=\eta(K_{1}), C=C(K1)C=C(K_{1}) and δ1=δ1(K1)\delta_{1}=\delta_{1}(K_{1}) are from Lemma 5.1 for K=K1K=K_{1} from (79).

Proof:

Let T=per(Γ)T=\operatorname{per}(\Gamma) be the period of Γ\Gamma.

(1). Given s,t[0,T]s,t\in[0,T], by (77) there are θs,θt𝒜(L)\theta_{s},\theta_{t}\in{\mathcal{A}}(L) such that

d(πΓ(s),πθs)\displaystyle d(\pi\Gamma(s),\pi\theta_{s}) d(Γ(s),θs)<δγ(Γ),\displaystyle\leq d(\Gamma(s),\theta_{s})<\delta\,\gamma(\Gamma),
d(πΓ(t),πθt)\displaystyle d(\pi\Gamma(t),\pi\theta_{t}) d(Γ(t),θt)<δγ(Γ).\displaystyle\leq d(\Gamma(t),\theta_{t})<\delta\,\gamma(\Gamma).

If d(s,t)mod T1d(s,t)_{\text{\rm mod }T}\geq 1 then

d(θs,θt)\displaystyle d(\theta_{s},\theta_{t}) d(Γ(s),Γ(t))d(Γ(s),θs)d(Γ(t),θt)\displaystyle\geq d(\Gamma(s),\Gamma(t))-d(\Gamma(s),\theta_{s})-d(\Gamma(t),\theta_{t})
>γ(Γ)2δγ(Γ).\displaystyle>\gamma(\Gamma)-2\delta\,\gamma(\Gamma).

Since θs,θt𝒜(L)\theta_{s},\theta_{t}\in{\mathcal{A}}(L), by the graph property 5.1 for 𝒜(L){\mathcal{A}}(L) and (88), (74) we have that

d(πθs,πθt)1Cd(θs,θt)γ(Γ)(12δ)C>γ¯δ2δγδ.d(\pi\theta_{s},\pi\theta_{t})\geq\tfrac{1}{C}\,d(\theta_{s},\theta_{t})\geq\frac{\gamma(\Gamma)(1-2\delta)}{C}>{\overline{\gamma}_{\delta}}-2\delta\,{\gamma_{\delta}}.

Then

d(πΓ(s),πΓ(t))\displaystyle d(\pi\Gamma(s),\pi\Gamma(t)) d(πθs,πθt)d(πΓ(s),πθs)d(πΓ(t),πθt)\displaystyle\geq d(\pi\theta_{s},\pi\theta_{t})-d(\pi\Gamma(s),\pi\theta_{s})-d(\pi\Gamma(t),\pi\theta_{t})
>γ¯δ4δγδ>34γ¯δ.\displaystyle>{\overline{\gamma}_{\delta}}-4\delta\,{\gamma_{\delta}}>\tfrac{3}{4}\,{\overline{\gamma}_{\delta}}.

(2). Suppose by contradiction that there exists t1t\leq-1 such that

(100) d(x(t),πΓ)<δ1 and \displaystyle d(x(t),\pi\Gamma)<\delta_{1}\qquad\text{ and }
(101) d([x(t),x˙(t)],Γ)2>δ2γ(Γ)2η1 and d([x(t),x˙(t)],Γ)>Cd(x(t),πΓ).\displaystyle d\big{(}[x(t),{\dot{x}}(t)],\Gamma\big{)}^{2}>\frac{\delta^{2}\,\gamma(\Gamma)^{2}}{\eta_{1}}\quad\text{ and }\quad d\big{(}[x(t),{\dot{x}}(t)],\Gamma\big{)}>C\,d\big{(}x(t),\pi\Gamma\big{)}.

First we check that we can apply the Crossing Lemma 5.1 to 𝕃{\mathbb{L}}. Given γ:[0,S]M\gamma:[0,S]\to M we have that

γc(L+ϕ)du=Sc(L+ϕ)u(γ(S))+u(γ(0))\oint_{\gamma}c(L+\phi)-du=S\,c(L+\phi)-u(\gamma(S))+u(\gamma(0))

depends only on the time interval SS and the endpoints of γ\gamma. Thus instead of 𝕃{\mathbb{L}} in (97), it is enough to apply Lemma 5.1 to L+ϕL+\phi, for whom it holds if ϕC2<ε1<ζ\left\|\phi\right\|_{C^{2}}<\varepsilon_{1}<\zeta by (82).

Now we check the speed hypothesis in Lemma 5.1. Observe that

E𝕃=v𝕃v𝕃=EL+ϕc(L+ϕ)=ELϕc(L+ϕ),E_{\mathbb{L}}=v\,{\mathbb{L}}_{v}-{\mathbb{L}}=E_{L+\phi}-c(L+\phi)=E_{L}-\phi-c(L+\phi),

and that by (6)

c(𝕃)=c(L+ϕ+c(L+ϕ))=0.c({\mathbb{L}})=c\big{(}L+\phi+c(L+\phi)\big{)}=0.

Therefore

𝒩(𝕃)[E𝕃=c(𝕃)][EL=ϕ+c(L+ϕ)].{\mathcal{N}}({\mathbb{L}})\subset[E_{\mathbb{L}}=c({\mathbb{L}})]\subset[E_{L}=\phi+c(L+\phi)].

If ϕ\phi is small enough

ϕ+c(L+ϕ)<c(L)+1,\phi+c(L+\phi)<c(L)+1,

and then x˙(t)𝒩(𝕃)[ELc(L)+1]{\dot{x}}(t)\in{\mathcal{N}}({\mathbb{L}})\subset[E_{L}\leq c(L)+1]. By hypothesis in 5.3, Γ[EL=c(L)]\Gamma\subset[E_{L}=c(L)]. Therefore by (79),

tx˙(t),Γ(t)[ELc(L)+1][|v|K1].\forall t\qquad{\dot{x}}(t),\;\Gamma(t)\in[E_{L}\leq c(L)+1]\subset[|v|\leq K_{1}].

Finally we check the distance hypothesis in Lemma 5.1. Let t0t_{0} be such that d(x(t),πΓ)=d(x(t),π(Γ(t0)))d(x(t),\pi\Gamma)=d(x(t),\pi(\Gamma(t_{0}))). By (100) and the definition of δ1\delta_{1} in (85) we can apply Lemma 5.1 for 𝕃{\mathbb{L}} and K=K1K=K_{1} from (79), to xx and πΓ\pi\Gamma at x(t)x(t) and π(Γ(t0))\pi(\Gamma(t_{0})). Also note that by (101) we have that, as required in Lemma 5.1,

d([x(t),x˙(t)],Γ(t0))d([x(t),x˙(t)],Γ)>Cd(x(t),πΓ)=Cd(x(t),πΓ(t0)).d\big{(}[x(t),{\dot{x}}(t)],\Gamma(t_{0})\big{)}\geq d\big{(}[x(t),{\dot{x}}(t)],\Gamma\big{)}>C\,d\big{(}x(t),\pi\Gamma\big{)}=C\,d\big{(}x(t),\pi\Gamma(t_{0})\big{)}.

Using 0<ε10<\varepsilon\leq 1 from Lemma 5.1 we obtain C1C^{1} curves w1,w2:[ε,ε]Mw_{1},\,w_{2}:[-\varepsilon,\varepsilon]\to M with w1(ε)=x(tε)w_{1}(-\varepsilon)=x(t-\varepsilon), w1(ε)=πΓ(t0+ε)w_{1}(\varepsilon)=\pi\Gamma(t_{0}+\varepsilon), w2(ε)=πΓ(t0ε)w_{2}(-\varepsilon)=\pi\Gamma(t_{0}-\varepsilon), w2(ε)=x(t+ε)w_{2}(\varepsilon)=x(t+\varepsilon) such that

A𝕃(w1)+A𝕃(w2)<A𝕃(πΓ|[t0ε,t0+ε])+A𝕃(x|[tε,t+ε])η1d([x(t),x˙(t)],Γ(t0))2.A_{\mathbb{L}}(w_{1})+A_{\mathbb{L}}(w_{2})<A_{\mathbb{L}}(\pi\Gamma|_{[t_{0}-\varepsilon,t_{0}+\varepsilon]})+A_{\mathbb{L}}(x|_{[t-\varepsilon,t+\varepsilon]})-\eta_{1}\,d([x(t),{\dot{x}}(t)],\Gamma(t_{0}))^{2}.

Since ϕ0\phi\geq 0 and (78) we have that

(102) c(L+ϕ)c(L)=0.c(L+\phi)\leq c(L)=0.

Using (77), ϕ|πΓ0\phi|_{\pi\Gamma}\equiv 0 and that πΓ\pi\Gamma is a closed curve we have that

A𝕃(πΓ)=AL+c(L+ϕ)(πΓ)AL+c(L)(πΓ)<δ2γ(Γ)2.A_{\mathbb{L}}(\pi\Gamma)=A_{L+c(L+\phi)}(\pi\Gamma)\leq A_{L+c(L)}(\pi\Gamma)<\delta^{2}\,\gamma(\Gamma)^{2}.

We compute the action of the curve w1πΓ|[t0+ε,t0+Tε]w2w_{1}*\pi\Gamma|_{[t_{0}+\varepsilon,t_{0}+T-\varepsilon]}*w_{2} which joins x(tε)x(t-\varepsilon) to x(t+ε)x(t+\varepsilon).

A𝕃(\displaystyle A_{\mathbb{L}}( w1)+A𝕃(πΓ|[t0+ε,t0+Tε])+A𝕃(w2)<\displaystyle w_{1})+A_{\mathbb{L}}(\pi\Gamma|_{[t_{0}+\varepsilon,t_{0}+T-\varepsilon]})+A_{\mathbb{L}}(w_{2})<
<A𝕃(x|[tε,t+ε])+A𝕃(πΓ|[t0ε,t0+ε])+𝔸𝕃(πΓ|[t0+ε,t0+Tε])η1d([x(t),x˙(t)],Γ(t0))2\displaystyle<A_{\mathbb{L}}(x|_{[t-\varepsilon,t+\varepsilon]})+A_{\mathbb{L}}(\pi\Gamma|_{[t_{0}-\varepsilon,t_{0}+\varepsilon]})+{\mathbb{A}}_{\mathbb{L}}(\pi\Gamma|_{[t_{0}+\varepsilon,t_{0}+T-\varepsilon]})-\eta_{1}\,d([x(t),{\dot{x}}(t)],\Gamma(t_{0}))^{2}
<A𝕃(x|[tε,t+ε])+δ2γ(Γ)2η1d([x(t),x˙(t)],Γ)2\displaystyle<A_{\mathbb{L}}(x|_{[t-\varepsilon,t+\varepsilon]})+\delta^{2}\,\gamma(\Gamma)^{2}-\eta_{1}\,d([x(t),{\dot{x}}(t)],\Gamma)^{2}
<A𝕃(x|[tε,t+ε]),using (101).\displaystyle<A_{\mathbb{L}}(x|_{[t-\varepsilon,t+\varepsilon]}),\qquad\text{using~{}\eqref{posso3}.}

This contradicts the assumption that xx is semi-static for 𝕃{\mathbb{L}}.

\triangle

Since we can assume that 𝒜(L){\mathcal{A}}(L) has no periodic orbits, if δ\delta is small enough

(103) T:=per(Γ)>1.T:=\operatorname{per}(\Gamma)>1.

Observe that Γ\Gamma is also a periodic orbit for L+ϕL+\phi. Let μΓ\mu_{\Gamma} be the invariant probability supported on Γ\Gamma. Using (6), (78), (77) we have that

c(L+ϕ)\displaystyle c(L+\phi) (L+ϕ)𝑑μΓ=L𝑑μΓ\displaystyle\geq-\int(L+\phi)\,d\mu_{\Gamma}=-\int L\;d\mu_{\Gamma}
(104) 1Tδ2γ(Γ)2.\displaystyle\geq-\tfrac{1}{T}\,\delta^{2}\,\gamma(\Gamma)^{2}.

We will prove that any semi-static curve x:],0]Mx:]-\infty,0]\to M for L+ϕL+\phi has α\alpha-limit{(x,x˙)}\{(x,{\dot{x}})\} =Γ=\Gamma. Since α\alpha-limits of semi-static orbits are static (Mañé [24, Theorem V.(c)]), this implies that Γ𝒜(L+ϕ)\Gamma\subset{\mathcal{A}}(L+\phi). Thus finishing the proof of Proposition 5.3.

Since by (84), the number γ¯δ{\overline{\gamma}_{\delta}} is smaller than the flow expansivity constant of 𝒩(L+ϕ){\mathcal{N}}(L+\phi), it is enough to prove that the tangent (x,x˙)(x,{\dot{x}}) of any semi-static curve x:],0]Mx:]-\infty,0]\to M spends only a bounded time outside the 38γ¯δ\tfrac{3}{8}{\overline{\gamma}_{\delta}}-neighbourhood of Γ\Gamma.

Let x:],0]Mx:]-\infty,0]\to M be a semi-static curve for L+ϕL+\phi. Let θ:=(x(0),x˙(0))\theta:=(x(0),{\dot{x}}(0)) and let ψt=φtL+ϕ\psi_{t}=\varphi_{t}^{L+\phi} be the lagrangian flow of L+ϕL+\phi. By (85) and (88) we have that

(105) d(x(t),πΓ)δ1d(x(t),πΓ)>14γ¯δ.d(x(t),\pi\Gamma)\geq\delta_{1}\quad\Longrightarrow\quad d(x(t),\pi\Gamma)>\tfrac{1}{4}{\overline{\gamma}_{\delta}}.

By (98)-(99) and (92) we have that

(106) d(ψt(θ),Γ)>Cρ&d(x(t),πΓ)<δ1d(x(t),πΓ)1Cd(ψt(θ),Γ).d(\psi_{t}(\theta),\Gamma)>C\rho\quad\&\quad d(x(t),\pi\Gamma)<\delta_{1}\quad\Longrightarrow\quad d(x(t),\pi\Gamma)\geq\tfrac{1}{C}\,d(\psi_{t}(\theta),\Gamma).

By (90) and (88) we have that 14γδ>Cρ\tfrac{1}{4}\gamma_{\delta}>C\rho. And then from (105) and (106) we get

(107) d(ψt(θ),Γ)14γδ(>14Cγ¯δ)d(x(t),πΓ)>14γ¯δ.d(\psi_{t}(\theta),\Gamma)\geq\tfrac{1}{4}{\gamma_{\delta}}\quad\left(>\tfrac{1}{4}C\,{\overline{\gamma}_{\delta}}\right)\quad\Longrightarrow\quad d(x(t),\pi\Gamma)>\tfrac{1}{4}\,{\overline{\gamma}_{\delta}}.

Also, from (105), (106) and (90) we have that

(108) d(ψt(θ),Γ)>Cρd(x(t),πΓ)>ρ.d(\psi_{t}(\theta),\Gamma)>C\rho\quad\Longrightarrow\quad d(x(t),\pi\Gamma)>\rho.

Then by (108), (96), (104), (103) and (91), we have that

(109) d(ψt(θ),Γ)>Cρϕ(x(t))+c(L+ϕ)14ερ2δ2γδ2=:a0>0.d(\psi_{t}(\theta),\Gamma)>C\rho\quad\Longrightarrow\quad\phi(x(t))+c(L+\phi)\geq\tfrac{1}{4}\,\varepsilon\rho^{2}-\delta^{2}\,\gamma_{\delta}^{2}=:a_{0}>0.

For ξΛ(ϕ)\xi\in\Lambda(\phi) consider the local invariant manifolds

Wηs(ξ)\displaystyle W^{s}_{\eta}(\xi) :={ζE𝕃1{c(𝕃)}:t0d(ψt(ζ),ψt(ξ))η},\displaystyle:=\{\,\zeta\in E_{\mathbb{L}}^{-1}\{c({\mathbb{L}})\}\;:\;\forall t\geq 0\quad d(\psi_{t}(\zeta),\psi_{t}(\xi))\leq\eta\,\},
Wηss(ξ)\displaystyle W^{ss}_{\eta}(\xi) :={ζWηs(ξ):limt+d(ψt(ζ),ψt(ξ))=0},\displaystyle:=\{\,\zeta\in W^{s}_{\eta}(\xi)\;:\;\lim_{t\to+\infty}d(\psi_{t}(\zeta),\psi_{t}(\xi))=0\,\},
Wηu(ξ)\displaystyle W^{u}_{\eta}(\xi) :={ζE𝕃1{c(𝕃)}:t0d(ψt(ζ),ψt(ξ))η},\displaystyle:=\{\,\zeta\in E_{{\mathbb{L}}}^{-1}\{c({\mathbb{L}})\}\;:\;\forall t\leq 0\quad d(\psi_{t}(\zeta),\psi_{t}(\xi))\leq\eta\,\},
Wηuu(ξ)\displaystyle W^{uu}_{\eta}(\xi) :={ζWηu(ξ):limtd(ψt(ζ),ψt(ξ))=0}.\displaystyle:=\{\,\zeta\in W^{u}_{\eta}(\xi)\;:\;\lim_{t\to-\infty}d(\psi_{t}(\zeta),\psi_{t}(\xi))=0\,\}.

Also consider the canonical coordinates as in A.4 on Λ(ϕ)\Lambda(\phi), i.e. there are η0,η>0\eta_{0},\,\eta>0 such that if ξ,ζΛ(ϕ)\xi,\,\zeta\in\Lambda(\phi) and d(ξ,ζ)<η0d(\xi,\zeta)<\eta_{0} then there is v=v(ξ,ζ)v=v(\xi,\zeta)\in{\mathbb{R}}, |v|η|v|\leq\eta such that

(110) ξ,ζ:=Wηss(ψv(ξ))Wηuu(ζ).\displaystyle\langle\xi,\zeta\rangle:=W^{ss}_{\eta}(\psi_{v}(\xi))\cap W^{uu}_{\eta}(\zeta)\neq\emptyset.

We use the canonical coordinates to parametrize the approaches of ψt(θ)\psi_{t}(\theta) to Γ\Gamma in the following way. By (87), γδ<η0{\gamma_{\delta}}<\eta_{0}. The local weak stable manifold of Γ\Gamma

Wηs(Γ):=ξΓWηs(ξ)=ξΓWηss(ξ)W^{s}_{\eta}(\Gamma):=\textstyle\bigcup_{\xi\in\Gamma}W^{s}_{\eta}(\xi)=\bigcup_{\xi\in\Gamma}W^{ss}_{\eta}(\xi)

forms a cylinder homeomorphic to Γ()×]0,1[dimM1\Gamma({\mathbb{R}})\times]0,1[^{\dim M-1}. When d(ψt(θ),Γ())<γδd(\psi_{t}(\theta),\Gamma({\mathbb{R}}))<{\gamma_{\delta}} the strong local unstable manifold Wηuu(ψt(θ))W^{uu}_{\eta}(\psi_{t}(\theta)) intersects this cylinder transversely and defines a unique time parameter v(t)v(t) (mod TT) such that

(111) Wηss(Γ(v(t)))Wηuu(ψt(θ))0.W^{ss}_{\eta}(\Gamma(v(t)))\cap W^{uu}_{\eta}(\psi_{t}(\theta))\neq 0.

Since the family of strong invariant manifolds is invariant under each iterate ψt\psi_{t} we have that if d(ψt(θ),Γ())<γδd(\psi_{t}(\theta),\Gamma({\mathbb{R}}))<{\gamma_{\delta}} for all t[a,b]t\in[a,b] then

s[0,ba]v(a+s)=v(a)+s.\forall s\in[0,b-a]\qquad v(a+s)=v(a)+s.

Let BB be from Lemma A.4. Write θ=(x(0),x˙(0))\theta=(x(0),{\dot{x}}(0)) and define Sk(θ)S_{k}(\theta), Tk(θ)T_{k}(\theta) recursively by

(112) S0(θ)\displaystyle S_{0}(\theta) :=0,\displaystyle:=0,
Tk(θ)\displaystyle T_{k}(\theta) :=sup{t<Sk1(θ)|d(ψt(θ),Γ(v(t)))C(B+1)ρ},\displaystyle:=\sup\,\big{\{}\,t<S_{k-1}(\theta)\;\big{|}\;d\big{(}\psi_{t}(\theta),\Gamma(v(t))\big{)}\leq C(B+1)\rho\,\big{\}},
Ck(θ)\displaystyle C_{k}(\theta) :=sup{t<Tk(θ)|d(ψt(θ),Γ())=13γδ},\displaystyle:=\sup\,\big{\{}\,t<T_{k}(\theta)\;\big{|}\;d(\psi_{t}(\theta),\Gamma({\mathbb{R}}))=\tfrac{1}{3}{\gamma_{\delta}}\,\big{\}},
Sk(θ)\displaystyle S_{k}(\theta) :=inf{t>Ck(θ)|d(ψt(θ),Γ(v(t)))C(B+1)ρ}.\displaystyle:=\inf\big{\{}\,t>C_{k}(\theta)\;\big{|}\;d\big{(}\psi_{t}(\theta),\Gamma(v(t))\big{)}\leq C(B+1)\rho\,\big{\}}.
Refer to caption
Figure 1. This figure illustrates the distance of the orbit of θ\theta to the periodic orbit Γ\Gamma and the choice of SkS_{k}, TkT_{k} and CkC_{k}.
Claim 5.4.2:
  1. (1)

    If Sk1(θ)>S_{k-1}(\theta)>-\infty then Tk(θ)>T_{k}(\theta)>-\infty.

  2. (2)

    If Tk(θ)>T_{k}(\theta)>-\infty then Tk+1(θ)Ck(θ)T_{k+1}(\theta)\leq C_{k}(\theta).

  3. (3)

    If Ck1(θ)>C_{k-1}(\theta)>-\infty then d[ψTk(θ)(θ),Γ(v(Tk(θ)))]=C(B+1)ρd\big{[}\psi_{T_{k}(\theta)}(\theta),\Gamma(v(T_{k}(\theta)))\big{]}=C(B+1)\rho.

  4. (4)

    If Ck(θ)>C_{k}(\theta)>-\infty then Ck(θ)<Sk(θ)Tk(θ)C_{k}(\theta)<S_{k}(\theta)\leq T_{k}(\theta).

  5. (5)

    If the sequence {Tk}\{T_{k}\} is finite, then α-limit(x,x˙)=Γ\alpha\text{-limit}(x,{\dot{x}})=\Gamma.

  6. (6)

    If t[Sk(θ),Tk(θ)]t\in[S_{k}(\theta),T_{k}(\theta)] then d(ψt(θ),Γ())13γδd(\psi_{t}(\theta),\Gamma({\mathbb{R}}))\leq\tfrac{1}{3}{\gamma_{\delta}}.

Proof:

(1). Suppose by contradiction that Sk1(θ)>S_{k-1}(\theta)>-\infty but Tk(θ)=T_{k}(\theta)=-\infty. Let ΦkL\Phi^{L}_{k} be the action potential (1) for LL. Since Φc(L)L\Phi^{L}_{c(L)} is Lipschitz, it is bounded on M×MM\times M.

tSk1(θ)L(x,x˙)=tSk1(θ){c(L)+L(x,x˙)}\displaystyle\int_{-t}^{S_{k-1}(\theta)}L(x,{\dot{x}})=\int_{-t}^{S_{k-1}(\theta)}\big{\{}c(L)+L(x,{\dot{x}})\big{\}} Φc(L)L(x(t),x(Sk1(θ)))\displaystyle\geq\Phi^{L}_{c(L)}\big{(}x(-t),x(S_{k-1}(\theta))\big{)}
(113) infy,zMΦc(L)L(y,z)=:b0>.\displaystyle\geq\inf_{y,z\in M}\Phi^{L}_{c(L)}(y,z)=:b_{0}>-\infty.

Recall that η\eta is from the canonical coordinates A.3 for Λ(ϕ)\Lambda(\phi) as in (110) and satisfies (87). Since Tk(θ)=T_{k}(\theta)=-\infty we have that for all t<Sk1(θ)t<S_{k-1}(\theta) either

(114) d(ψt(θ),Γ())>η>γδ or d(\psi_{t}(\theta),\Gamma({\mathbb{R}}))>\eta>{\gamma_{\delta}}\qquad\text{ or }
(115) d(ψt(θ),Γ())η but d(ψt(θ),Γ(v(t)))>C(B+1)ρ.d(\psi_{t}(\theta),\Gamma({\mathbb{R}}))\leq\eta\quad\text{ but }\quad d(\psi_{t}(\theta),\Gamma(v(t)))>C(B+1)\rho.

In the case (115) let s(t)s(t) be such that d(ψt(θ),Γ(s(t)))=d(ψt(θ),Γ())ηd(\psi_{t}(\theta),\Gamma(s(t)))=d(\psi_{t}(\theta),\Gamma({\mathbb{R}}))\leq\eta. We have that

Γ(s(t)),ψt(θ)\displaystyle\langle\Gamma(s(t)),\psi_{t}(\theta)\rangle =Wηs(Γ(s(t)))Wηuu(ψt(θ))\displaystyle=W^{s}_{\eta}(\Gamma(s(t)))\cap W^{uu}_{\eta}(\psi_{t}(\theta))
=Wηs(Γ(v(t)))Wηuu(ψt(θ))=Γ(v(t)),ψt(θ)\displaystyle=W^{s}_{\eta}(\Gamma(v(t)))\cap W^{uu}_{\eta}(\psi_{t}(\theta))=\langle\Gamma(v(t)),\psi_{t}(\theta)\rangle
(116) =Wηss(Γ(v(t)))Wηuu(ψt(θ)).\displaystyle=W^{ss}_{\eta}(\Gamma(v(t)))\cap W^{uu}_{\eta}(\psi_{t}(\theta)).

We apply Lemma A.4 with x:=Γ(s(t))x:=\Gamma(s(t)) and y:=ψt(θ)y:=\psi_{t}(\theta). Using (155) we have that

(117) d(y,ψv(x))d(y,x)+d(x,ψv(x))(1+B)d(y,x).d(y,\psi_{v}(x))\leq d(y,x)+d(x,\psi_{v}(x))\leq(1+B)\,d(y,x).

Observe that (116) implies that ψv(x)=Γ(v(t))\psi_{v}(x)=\Gamma(v(t)). Replacing xx and yy in (117) and using (115) we have that

d(ψt(θ),Γ())\displaystyle d(\psi_{t}(\theta),\Gamma({\mathbb{R}})) =d(ψt(θ),Γ(s(t)))11+Bd(ψt(θ),Γ(v(t)))\displaystyle=d(\psi_{t}(\theta),\Gamma(s(t)))\geq\tfrac{1}{1+B}\;d(\psi_{t}(\theta),\Gamma(v(t)))
(118) >Cρ.\displaystyle>C\rho.

Observe that by (90) and (88), in case (114) inequality (118) also holds. Therefore

(119) t<Sk1(θ)d(ψt(θ),Γ())>Cρ.\forall t<S_{k-1}(\theta)\qquad d(\psi_{t}(\theta),\Gamma({\mathbb{R}}))>C\rho.

Since xx is semi-static for L+ϕL+\phi we have for all t<Sk1(θ)-t<S_{k-1}(\theta) that

>supy,zMΦc(L+ϕ)L+ϕ(y,z)\displaystyle\infty>\sup_{y,z\in M}\Phi^{L+\phi}_{c(L+\phi)}(y,z) Φc(L+ϕ)L+ϕ(x(t),x(Sk1(θ)))\displaystyle\geq\Phi_{c(L+\phi)}^{L+\phi}\big{(}x(-t),x(S_{k-1}(\theta))\big{)}
=tSk1(θ)[L(x,x˙)+ϕ(x)+c(L+ϕ)]\displaystyle=\int_{-t}^{S_{k-1}(\theta)}\big{[}L(x,{\dot{x}})+\phi(x)+c(L+\phi)\big{]}
=tSk1(θ)L(x,x˙)+tSk1(θ)[ϕ(x)+c(L+ϕ)]\displaystyle=\int_{-t}^{S_{k-1}(\theta)}L(x,{\dot{x}})+\int_{-t}^{S_{k-1}(\theta)}\big{[}\phi(x)+c(L+\phi)\big{]}
(120) b0+a0(t+Sk1(θ)) by (113) and (119), (109).\displaystyle\geq b_{0}+a_{0}\big{(}t+S_{k-1}(\theta)\big{)}\qquad\text{ by \eqref{intb0} and \eqref{tsk}, \eqref{phirho}.}

By (109) we have that a0>0a_{0}>0. Letting t+t\to+\infty, inequality (120) gives a contradiction.

(2). Let

(121) f(t):=d(ψt(θ),Γ()) and g(t):=d(ψt(θ),Γ(v(t))),f(t):=d(\psi_{t}(\theta),\Gamma({\mathbb{R}}))\qquad\text{ and }\qquad g(t):=d(\psi_{t}(\theta),\Gamma(v(t))),

when gg is defined (in particular by (87) when f(t)<γδf(t)<{\gamma_{\delta}}). Then f(t)g(t)f(t)\leq g(t).

Suppose first that Ck(θ)=C_{k}(\theta)=-\infty. Then f(t)13γδf(t)\neq\tfrac{1}{3}{\gamma_{\delta}} for all t<Tk(θ)t<T_{k}(\theta). By hypothesis Tk(θ)>T_{k}(\theta)>-\infty, then f(Tk(θ))g(Tk(θ))C(B+1)ρf(T_{k}(\theta))\leq g(T_{k}(\theta))\leq C(B+1)\rho. By (90), C(B+1)ρ<13γδC(B+1)\rho<\tfrac{1}{3}\gamma_{\delta}, and hence f(t)<13γδf(t)<\tfrac{1}{3}{\gamma_{\delta}} for all t<Tk(θ)t<T_{k}(\theta). By (86) and Proposition A.7 with LL\to\infty we have that limtg(t)=0\lim_{t\to-\infty}g(t)=0. Then Sk(θ)=S_{k}(\theta)=-\infty and also Tk+1(θ)=T_{k+1}(\theta)=-\infty.

Now suppose that Ck(θ)>C_{k}(\theta)>-\infty. By the definition of Sk(θ)S_{k}(\theta) for all t]Ck(θ),Sk(θ)[t\in]C_{k}(\theta),S_{k}(\theta)[ we have that g(t)>C(B+1)ρg(t)>C(B+1)\rho. This implies that Tk+1(θ)Ck(θ)T_{k+1}(\theta)\leq C_{k}(\theta).

(3). Let f,gf,\,g be as in (121). By the hypothesis Ck1(θ)>C_{k-1}(\theta)>-\infty and by the definition of Ck1(θ)C_{k-1}(\theta), Ck1(θ)Tk1(θ)C_{k-1}(\theta)\leq T_{k-1}(\theta). Then f(Ck1(θ))=13γδf(C_{k-1}(\theta))=\tfrac{1}{3}{\gamma_{\delta}}. By (90), C(B+1)ρ<13γδC(B+1)\rho<\tfrac{1}{3}{\gamma_{\delta}} and then

(122) C(B+1)ρ<13γδ=f(Ck1(θ))g(Ck1(θ)).C(B+1)\rho<\tfrac{1}{3}{\gamma_{\delta}}=f(C_{k-1}(\theta))\leq g(C_{k-1}(\theta)).

By the definition of Sk1(θ)S_{k-1}(\theta) we have that Ck1(θ)Sk1(θ)C_{k-1}(\theta)\leq S_{k-1}(\theta). But by (122), g(Ck1(θ))13γδg(C_{k-1}(\theta))\geq\tfrac{1}{3}{\gamma_{\delta}}, and by the definition of Sk1(θ)S_{k-1}(\theta), if Sk1(θ)<+S_{k-1}(\theta)<+\infty then g(Sk1(θ))C(B+1)ρ<13γδg(S_{k-1}(\theta))\leq C(B+1)\rho<\tfrac{1}{3}{\gamma_{\delta}}. Therefore Ck1(θ)Sk1(θ)C_{k-1}(\theta)\neq S_{k-1}(\theta) and then

(123) Ck1(θ)<Sk1(θ)+.C_{k-1}(\theta)<S_{k-1}(\theta)\leq+\infty.

By (122) and the definition of Sk1(θ)S_{k-1}(\theta) we have that

t]Ck1(θ),Sk1(θ)[g(t)>C(B+1)ρ.\forall t\in]C_{k-1}(\theta),S_{k-1}(\theta)[\qquad g(t)>C(B+1)\rho.

This implies that Tk(θ)<Ck1(θ)T_{k}(\theta)<C_{k-1}(\theta), with strict inequality by (122). By (123) and item (1) we have that Ck1(θ)>C_{k-1}(\theta)>-\infty implies that Tk(θ)>T_{k}(\theta)>-\infty. Therefore

(124) <Tk(θ)<Ck1(θ)<Sk1(θ).-\infty<T_{k}(\theta)<C_{k-1}(\theta)<S_{k-1}(\theta).

The definition of Tk(θ)T_{k}(\theta) and the continuity of g(t)g(t) on its domain imply that

(125) g(Tk(θ))C(B+1)ρ.g(T_{k}(\theta))\leq C(B+1)\rho.

The domain of definition and continuity of gg contains f1(]0,γδ[)g1(]0,γδ[)f^{-1}(]0,{\gamma_{\delta}}[)\supset g^{-1}(]0,{\gamma_{\delta}}[). By the intermediate value theorem for gg on connected components of [gγδ][g\leq{\gamma_{\delta}}] and (124), (125), (122), the image g([Tk(θ),Ck1(θ)])g([T_{k}(\theta),C_{k-1}(\theta)]), and hence also g(],Sk1(θ)[)g(]-\infty,S_{k-1}(\theta)[), contain the closed interval [C(B+1)ρ,13γδ]\big{[}C(B+1)\rho,\tfrac{1}{3}{\gamma_{\delta}}\big{]}. Therefore, by the definition of Tk(θ)T_{k}(\theta), we have that g(Tk(θ))=C(B+1)ρg(T_{k}(\theta))=C(B+1)\rho.

(4). Let ff, gg be from (121). If Ck(θ)>C_{k}(\theta)>-\infty then by the definition of Ck(θ)C_{k}(\theta),

(126) Ck(θ)Tk(θ).C_{k}(\theta)\leq T_{k}(\theta).

Therefore Tk(θ)>T_{k}(\theta)>-\infty. Then the definition of Tk(θ)T_{k}(\theta) implies that

(127) g(Tk(θ))C(B+1)ρ.g(T_{k}(\theta))\leq C(B+1)\rho.

Since f(t)f(t) is continuous,

(128) f(Ck(θ))=13γδ.f(C_{k}(\theta))=\tfrac{1}{3}{\gamma_{\delta}}.

By (127),  (90) and (128) we have that

(129) g(Tk(θ))C(B+1)ρ<14γδ<13γδ=f(Ck(θ))g(Ck(θ)).g(T_{k}(\theta))\leq C(B+1)\rho<\tfrac{1}{4}\gamma_{\delta}<\tfrac{1}{3}{\gamma_{\delta}}=f(C_{k}(\theta))\leq g(C_{k}(\theta)).

This implies that Ck(θ)Tk(θ)C_{k}(\theta)\neq T_{k}(\theta). This together with (126) imply that

(130) Ck(θ)<Tk(θ).C_{k}(\theta)<T_{k}(\theta).

By (127) and (130) the value Sk(θ)S_{k}(\theta) is an infimum of a set which contains Tk(θ)T_{k}(\theta), therefore

(131) Sk(θ)Tk(θ).S_{k}(\theta)\leq T_{k}(\theta).

This proves the second inequality in item (4).

The first of the following inequalities follows from the definition of Sk(θ)S_{k}(\theta). The second inequality is (131). The third inequality follows from the definition of Tk(θ)T_{k}(\theta).

(132) Ck(θ)Sk(θ)Tk(θ)Sk1(θ).C_{k}(\theta)\leq S_{k}(\theta)\leq T_{k}(\theta)\leq S_{k-1}(\theta).

We get that

<Ck(θ)Sk(θ)Sk1(θ)S0(θ):=0<+.-\infty<C_{k}(\theta)\leq S_{k}(\theta)\leq S_{k-1}(\theta)\leq\cdots\leq S_{0}(\theta):=0<+\infty.

From the definition of Sk(θ)S_{k}(\theta) and Sk(θ)<+S_{k}(\theta)<+\infty, and then (128), we have that

g(Sk(θ))C(B+1)ρ<13γδ=f(Ck(θ))g(Ck(θ)).g(S_{k}(\theta))\leq C(B+1)\rho<\tfrac{1}{3}{\gamma_{\delta}}=f(C_{k}(\theta))\leq g(C_{k}(\theta)).

In particular Ck(θ)Sk(θ)C_{k}(\theta)\neq S_{k}(\theta). Thus from (132), Ck(θ)<Sk(θ)C_{k}(\theta)<S_{k}(\theta).

(5). If the sequence {Tk}\{T_{k}\} is finite, there is \ell\in{\mathbb{N}} such that T>T_{\ell}>-\infty and T+1=T_{\ell+1}=-\infty. Let f,gf,\,g be from (121). By item (2) we have that <T(θ)C1(θ)-\infty<T_{\ell}(\theta)\leq C_{\ell-1}(\theta). Then we can apply item (3) and use (90) to obtain

(133) f(T(θ))g(T(θ))=C(B+1)ρ<13γδ.f(T_{\ell}(\theta))\leq g(T_{\ell}(\theta))=C(B+1)\rho<\tfrac{1}{3}{\gamma_{\delta}}.

Since T+1(θ)=T_{\ell+1}(\theta)=-\infty, by item (1), S(θ)=S_{\ell}(\theta)=-\infty and by item (4), C(θ)=C_{\ell}(\theta)=-\infty. Since C(θ)=C_{\ell}(\theta)=-\infty we have that f(t)13γδf(t)\neq\tfrac{1}{3}{\gamma_{\delta}} for all t<T(θ)t<T_{\ell}(\theta). But by (133), f(T(θ))<13γδf(T_{\ell}(\theta))<\tfrac{1}{3}{\gamma_{\delta}}. Since f(t)f(t) is continuous, using (86) we get that

f(t)<13γδ<β0 for all t<T(θ).f(t)<\tfrac{1}{3}{\gamma_{\delta}}<\beta_{0}\qquad\text{ for all }t<T_{\ell}(\theta).

This implies that there is a continuous function s:],T(θ)]s:]-\infty,T_{\ell}(\theta)]\to{\mathbb{R}} such that

tTk(θ)d(ψt(θ),Γ(s(t)))β0.\forall t\leq T_{k}(\theta)\qquad d\big{(}\psi_{t}(\theta),\Gamma(s(t))\big{)}\leq\beta_{0}.

By Proposition A.7 there is vv\in{\mathbb{R}} and λ>0\lambda>0 such that

tT(θ)d(ψt(θ),Γ(t+v))Dβ0eλ(T(θ)t).\forall t\leq T_{\ell}(\theta)\qquad d(\psi_{t}(\theta),\Gamma(t+v))\leq D\,\beta_{0}\,\text{\rm\large e}^{-\lambda(T_{\ell}(\theta)-t)}.

This implies that limt+d(ψt(θ),Γ)=0\lim\limits_{t\to+\infty}d(\psi_{-t}(\theta),\Gamma)=0 and that α-limit(θ)=Γ()\alpha\text{-limit}(\theta)=\Gamma({\mathbb{R}}).

(6). By item (2), Ck1(θ)Tk(θ)>C_{k-1}(\theta)\geq T_{k}(\theta)>-\infty. By item (3) we have that f(Tk(θ))g(Tk(θ))=C(B+1)ρ<13γδf(T_{k}(\theta))\leq g(T_{k}(\theta))=C(B+1)\rho<\tfrac{1}{3}{\gamma_{\delta}}. By the definition of Ck(θ)C_{k}(\theta) we have that t]Ck(θ),Tk(θ)]\forall t\in]C_{k}(\theta),T_{k}(\theta)] f(t)13γδf(t)\neq\tfrac{1}{3}{\gamma_{\delta}}. Then by the continuity of f(t)f(t), t]Ck(θ),Tk(θ)]\forall t\in]C_{k}(\theta),T_{k}(\theta)] f(t)<13γδf(t)<\tfrac{1}{3}{\gamma_{\delta}}. Now it is enough to see that by item (4), [Sk(θ),Tk(θ)]]Ck(θ),Tk(θ)][S_{k}(\theta),T_{k}(\theta)]\subset]C_{k}(\theta),T_{k}(\theta)].

\triangle

Let

Bk(θ):=sup{t<Ck(θ)|d(ψt(θ),Γ())14γδ}.B_{k}(\theta):=\sup\big{\{}\;t<C_{k}(\theta)\;\big{|}\;d(\psi_{t}(\theta),\Gamma({\mathbb{R}}))\leq\tfrac{1}{4}{\gamma_{\delta}}\;\big{\}}.
Claim 5.4.3:
[Bk(θ),Ck(θ)][Tk+1(θ),Sk(θ)].[B_{k}(\theta),C_{k}(\theta)]\subset[T_{k+1}(\theta),S_{k}(\theta)].

Proof:

Let f,gf,\,g be as in (121). By the definition of Sk(θ)S_{k}(\theta) we have that Sk(θ)Ck(θ)S_{k}(\theta)\geq C_{k}(\theta). By the definition of Bk(θ)B_{k}(\theta) and (90), we have that

(134) g|]Bk,Ck[f|]Bk,Ck[>14γδ>C(B+1)ρ.g|_{]B_{k},C_{k}[}\geq f|_{]B_{k},C_{k}[}>\tfrac{1}{4}{\gamma_{\delta}}>C(B+1)\rho.

By the definition of Sk(θ)S_{k}(\theta) we have that

(135) g|]Ck,Sk[>C(B+1)ρ.g|_{]C_{k},S_{k}[}>C(B+1)\rho.

By the definition of Ck(θ)C_{k}(\theta) and the continuity of f(t)f(t) we have that

(136) g(Ck(θ))f(Ck(θ))=13γδ>C(B+1)ρ.g(C_{k}(\theta))\geq f(C_{k}(\theta))=\tfrac{1}{3}{\gamma_{\delta}}>C(B+1)\rho.

Joining (134), (135) and (136) we get that

g|]Bk,Sk[>C(B+1)ρ.g|_{]B_{k},S_{k}[}>C(B+1)\rho.

By the definition of Tk+1(θ)T_{k+1}(\theta) this implies that Tk+1(θ)Bk(θ)T_{k+1}(\theta)\leq B_{k}(\theta).

\triangle

If t[Bk(θ),Ck(θ)]t\in[B_{k}(\theta),C_{k}(\theta)], by the definition of Bk(θ)B_{k}(\theta) we have that

d(ψt(θ),Γ)14γδ.d(\psi_{t}(\theta),\Gamma)\geq\tfrac{1}{4}{\gamma_{\delta}}.

Then by (107),

(137) t[Bk(θ),Ck(θ)]d(x(t),πΓ)>14γ¯δ.t\in[B_{k}(\theta),C_{k}(\theta)]\quad\Longrightarrow\quad d(x(t),\pi\Gamma)>\tfrac{1}{4}{\overline{\gamma}_{\delta}}.

By the definition of Tk+1(θ)T_{k+1}(\theta) we have that

(138) t]Tk+1(θ),Sk(θ)[eitherd(ψt(θ),Γ(v(t)))>C(B+1)ρ\forall t\in]T_{k+1}(\theta),S_{k}(\theta)[\qquad\text{either}\qquad d\big{(}\psi_{t}(\theta),\Gamma(v(t))\big{)}>C(B+1)\rho

or d(ψt(θ),Γ())>η>Cρd(\psi_{t}(\theta),\Gamma({\mathbb{R}}))>\eta>C\rho (when v(t)v(t) does not exist). Here η>Cρ\eta>C\rho follows from (90), (88), (87). The arguments in (117)-(118) apply in the case (138) to obtain

(139) t]Tk+1(θ),Sk(θ)[d(ψt(θ),Γ)>Cρ.t\in]T_{k+1}(\theta),S_{k}(\theta)[\quad\Longrightarrow\quad d(\psi_{t}(\theta),\Gamma)>C\rho.

The continuity of ff and the definition of BkB_{k} and CkC_{k} give

(140) f(Bk)14γδ,f(Ck)=13γδ.f(B_{k})\leq\tfrac{1}{4}\,{\gamma_{\delta}},\qquad f(C_{k})=\tfrac{1}{3}{\gamma_{\delta}}.

From Claim 5.4.3, {(137), (96)}, (104), {(89), (140), (81)}, {(95), (103)} and {(139), (109)}, we have that

Tk+1(θ)Sk(θ)(ϕ+c(L+ϕ))\displaystyle\int_{T_{k+1}(\theta)}^{S_{k}(\theta)}\Big{(}\phi+c(L+\phi)\Big{)} Bk(θ)Ck(θ)(132ε(γ¯δ)21Tδ2γδ2)+[Tk+1,Sk][Bk,Ck](ϕ+c(L+ϕ))\displaystyle\geq\int_{B_{k}(\theta)}^{C_{k}(\theta)}\Big{(}\tfrac{1}{32}\,\varepsilon\,({\overline{\gamma}_{\delta}})^{2}-\tfrac{1}{T}\,\delta^{2}\,{\gamma_{\delta}}^{2}\Big{)}+\int_{[T_{k+1},S_{k}]\setminus[B_{k},C_{k}]}\Big{(}\phi+c(L+\phi)\Big{)}
(141) (132ε(γ¯δ)21Tδ2γδ2)α+0.\displaystyle\geq\big{(}\tfrac{1}{32}\,\varepsilon\,({\overline{\gamma}_{\delta}})^{2}-\tfrac{1}{T}\,\delta^{2}\,{\gamma_{\delta}}^{2}\big{)}\,\alpha+0.

Recall from (97) that

𝕃:=L+ϕ+c(L+ϕ)du,{\mathbb{L}}:=L+\phi+c(L+\phi)-du,

where uu is from (80). Observe that the lagrangian flow for 𝕃{\mathbb{L}} is the same as the lagrangian flow ψt\psi_{t} for L+ϕL+\phi. Also 𝒩(𝕃)=𝒩(L+ϕ){\mathcal{N}}({\mathbb{L}})={\mathcal{N}}(L+\phi) and 𝒜(𝕃)=𝒜(L+ϕ){\mathcal{A}}({\mathbb{L}})={\mathcal{A}}(L+\phi). Using (80) and (141),

Tk+1(θ)Sk(θ)𝕃(ψt(θ))𝑑t\displaystyle\int_{T_{k+1}(\theta)}^{S_{k}(\theta)}{\mathbb{L}}(\psi_{t}(\theta))\,dt =Tk+1(θ)Sk(θ)(Ldu)+Tk+1(θ)Sk(θ)(ϕ+c(L+ϕ))\displaystyle=\int_{T_{k+1}(\theta)}^{S_{k}(\theta)}(L-du)+\int_{T_{k+1}(\theta)}^{S_{k}(\theta)}\Big{(}\phi+c(L+\phi)\Big{)}
(142) 0+(132ε(γ¯δ)21Tδ2γδ2)α.\displaystyle\geq 0+\left(\tfrac{1}{32}\,\varepsilon\,({\overline{\gamma}_{\delta}})^{2}-\tfrac{1}{T}\,\delta^{2}\,{\gamma_{\delta}}^{2}\right)\alpha.

Case 1: Suppose that Tk(θ)Sk(θ)>T+2T_{k}(\theta)-S_{k}(\theta)>T+2.

Let mkm_{k}\in{\mathbb{N}} be such that

Sk(θ)+mkTTk(θ)1<Sk(θ)+(mk+1)T.S_{k}(\theta)+m_{k}T\leq T_{k}(\theta)-1<S_{k}(\theta)+(m_{k}+1)T.

Then mk1m_{k}\geq 1. Let Rk(θ):=Sk(θ)+mkTR_{k}(\theta):=S_{k}(\theta)+m_{k}T. Then 1Tk(θ)Rk(θ)<T+11\leq T_{k}(\theta)-R_{k}(\theta)<T+1. By Claim 5.4.2.(6), Γ\Gamma is γδ3\tfrac{{\gamma_{\delta}}}{3}-shadowed by ψ[Sk,Tk](θ)\psi_{[S_{k},T_{k}]}(\theta). Therefore by inequality (180) in Proposition A.7 there is vv\in{\mathbb{R}} such that t[Sk,Tk]\forall t\in[S_{k},T_{k}]

(143) d(ψt(θ),Γ(t+v))Deλmin{tSk,Tkt}[d(ψSk(θ),Γ(Sk+v))+d(ψTk(θ),Γ(Tk+v))].d(\psi_{t}(\theta),\Gamma(t+v))\leq D\,\text{\rm\large e}^{-\lambda\min\{t-S_{k},T_{k}-t\}}[d(\psi_{S_{k}}(\theta),\Gamma(S_{k}+v))+d(\psi_{T_{k}}(\theta),\Gamma(T_{k}+v))].

Also the choice of vv in Proposition A.7 is the same as in (111) so that

(144) t+v=v(t)t[Sk(θ),Tk(θ)].t+v=v(t)\qquad\forall t\in[S_{k}(\theta),T_{k}(\theta)].
Refer to caption
Figure 2. The auxiliary segments w1w_{1} an w2w_{2}.

By the definition of SkS_{k} and TkT_{k} in (112) and the continuity of g(t)g(t) on its domain we have that

(145) g(Sk)C(B+1)ρ,g(Tk)C(B+1)ρ.g(S_{k})\leq C(B+1)\rho,\qquad g(T_{k})\leq C(B+1)\rho.

By (143), (144) and (145) we have for s[0,1]s\in[0,1] that

d(ψs+Rk(θ)\displaystyle d\big{(}\psi_{s+R_{k}}(\theta) ,Γ(v(s+Rk)))\displaystyle,\Gamma(v(s+R_{k}))\big{)}\leq
Deλmin{s+RkSk,TksRk}[d(ψSk(θ),Γ(v(Sk)))+d(ψTk(θ),Γ(v(Tk)))]\displaystyle\leq D\text{\rm\large e}^{-\lambda\min\{s+R_{k}-S_{k},T_{k}-s-R_{k}\}}\big{[}d(\psi_{S_{k}}(\theta),\Gamma(v(S_{k})))+d(\psi_{T_{k}}(\theta),\Gamma(v(T_{k})))\big{]}
De0[g(Sk)+g(Tk)]2DC(B+1)ρ.\displaystyle\leq D\,\text{\rm\large e}^{0}\;[g(S_{k})+g(T_{k})]\leq 2DC(B+1)\rho.
d(Γ(v(s+\displaystyle d(\Gamma(v(s+ Sk)),ψs+Sk(θ))2DC(B+1)ρ.\displaystyle S_{k})),\psi_{s+S_{k}}(\theta))\leq 2DC(B+1)\rho.

From (144) we have that

v(s+Rk)=s+Rk+v=s+Sk+v+mkT=v(s+Sk)+mkT.v(s+R_{k})=s+R_{k}+v=s+S_{k}+v+m_{k}T=v(s+S_{k})+m_{k}T.

So that Γ(v(s+Rk))=Γ(v(s+Sk))\Gamma(v(s+R_{k}))=\Gamma(v(s+S_{k})). Adding the inequalities above we get

(146) s[0,1]d(ψs+Rk(θ),ψs+Sk(θ))4DC(B+1)ρ.\forall s\in[0,1]\qquad d(\psi_{s+R_{k}}(\theta),\psi_{s+S_{k}}(\theta))\leq 4DC(B+1)\rho.

In local coordinates about π(Γ)\pi(\Gamma) define

w1(s+Rk)\displaystyle w_{1}(s+R_{k}) =(1\displaystyle=(1- s)\displaystyle s)\, x(s+Rk)\displaystyle x(s+R_{k})\; +\displaystyle+ s\displaystyle s\, x(s+Sk),s[0,1];\displaystyle x(s+S_{k}),\qquad s\in[0,1];
w2(s+Sk)\displaystyle w_{2}(s+S_{k}) =\displaystyle= s\displaystyle s\, x(s+Rk)\displaystyle x(s+R_{k}) +\displaystyle+ (1\displaystyle\;(1- s)\displaystyle s)\, x(s+Sk),s[0,1].\displaystyle x(s+S_{k}),\qquad s\in[0,1].

By Lemma 5.2(b) and (146) we have that

AL+ϕ(x|[Sk,1+Sk])+AL+ϕ(x|[Rk,1+Rk])AL+ϕ(w1)+AL+ϕ(w2)6KD2C2(B+1)2ρ2.A_{L+\phi}(x|_{[S_{k},1+S_{k}]})+A_{L+\phi}(x|_{[R_{k},1+R_{k}]})\geq A_{L+\phi}(w_{1})+A_{L+\phi}(w_{2})-6KD^{2}C^{2}(B+1)^{2}\rho^{2}.

Since the pairs of segments {x|[Sk,1+Sk],x|[Rk,1+Rk]}\{\,x|_{[S_{k},1+S_{k}]},\,x|_{[R_{k},1+R_{k}]}\,\} and {w1,w2}\{\,w_{1},\,w_{2}\,\} have the same collections of endpoints

Sk1+Sk𝑑u(x˙)+Rk1+Rk𝑑u(x˙)=w1𝑑u+w2𝑑u.\int_{S_{k}}^{1+S_{k}}du({\dot{x}})+\int_{R_{k}}^{1+R_{k}}du({\dot{x}})=\oint_{w_{1}}du+\oint_{w_{2}}du.

Therefore, since c(L+ϕ)c(L+\phi) is constant,

(147) A𝕃(x|[Sk,1+Sk])+A𝕃(x|[Rk,1+Rk])A𝕃(w1)+A𝕃(w2)6KD2C2(B+1)2ρ2.A_{\mathbb{L}}(x|_{[S_{k},1+S_{k}]})+A_{\mathbb{L}}(x|_{[R_{k},1+R_{k}]})\geq A_{\mathbb{L}}(w_{1})+A_{\mathbb{L}}(w_{2})-6KD^{2}C^{2}(B+1)^{2}\rho^{2}.

The integral of dxud_{x}u on closed curves is zero. Therefore

(148) c(𝕃)=c(L+ϕ+c(L+ϕ))=0.c({\mathbb{L}})=c(L+\phi+c(L+\phi))=0.

Since w1x|[1+Sk,Rk]w_{1}*x|_{[1+S_{k},R_{k}]} is a closed curve and c(𝕃)=0c({\mathbb{L}})=0, using (7),

(149) A𝕃(w1)+A𝕃(x|[1+Sk,Rk])0.A_{\mathbb{L}}(w_{1})+A_{\mathbb{L}}(x|_{[1+S_{k},R_{k}]})\geq 0.

Using (80) and (104),

(150) 𝕃=(Ldu)+ϕ+c(L+ϕ)0+01Tδ2(γδ)2.{\mathbb{L}}=(L-du)+\phi+c(L+\phi)\geq 0+0-\tfrac{1}{T}\,\delta^{2}({\gamma_{\delta}})^{2}.

Since Tk(θ)Rk(θ)T+2T_{k}(\theta)-R_{k}(\theta)\leq T+2, using (103), on the curve w2x|[1+Rk,Tk]w_{2}*x|_{[1+R_{k},T_{k}]} we have that

(151) A𝕃(w2)+A𝕃(x|[1+Rk,Tk])T+2Tδ2(γδ)23δ2(γδ)2.A_{\mathbb{L}}(w_{2})+A_{\mathbb{L}}(x|_{[1+R_{k},T_{k}]})\geq-\tfrac{T+2}{T}\,\delta^{2}({\gamma_{\delta}})^{2}\geq-3\,\delta^{2}({\gamma_{\delta}})^{2}.

From (147), (149) and (151) we get that

A𝕃(x|[Sk,Tk])\displaystyle A_{\mathbb{L}}(x|_{[S_{k},T_{k}]}) A𝕃(w1)\displaystyle\geq A_{\mathbb{L}}(w_{1}) +A𝕃(w2)6KD2C2(B+1)2ρ2\displaystyle+A_{\mathbb{L}}(w_{2})-6KD^{2}C^{2}(B+1)^{2}\rho^{2}
+A𝕃(x|[1+Sk,Rk])+A𝕃(x|[1+Rk,Tk])\displaystyle+A_{\mathbb{L}}(x|_{[1+S_{k},R_{k}]})+A_{\mathbb{L}}(x|_{[1+R_{k},T_{k}]})
6KD\displaystyle\geq-6KD C22(B+1)2ρ23δ2(γδ)2.{}^{2}C^{2}(B+1)^{2}\rho^{2}-3\,\delta^{2}({\gamma_{\delta}})^{2}.

Case 2: If TkSkT+2T_{k}-S_{k}\leq T+2, from (150) we also have

A𝕃(x|[Sk,Tk])\displaystyle\hskip-56.9055ptA_{\mathbb{L}}(x|_{[S_{k},T_{k}]}) T+2Tδ2(γδ)23δ2(γδ)2\displaystyle\geq-\tfrac{T+2}{T}\,\delta^{2}({\gamma_{\delta}})^{2}\geq-3\,\delta^{2}({\gamma_{\delta}})^{2}
6KD2C2(B+1)2ρ23δ2(γδ)2.\displaystyle\geq-6KD^{2}C^{2}(B+1)^{2}\rho^{2}-3\,\delta^{2}({\gamma_{\delta}})^{2}.

Adding inequality (142) and using (93) we obtain a positive lower bound for the action independent of kk:

A𝕃(x|[Tk+1,Tk])(132ε(γ¯δ)2δ2(γδ)2)α12KD2C2(B+1)2ρ23δ2(γδ)2>0.A_{\mathbb{L}}(x|_{[T_{k+1},T_{k}]})\geq\left(\tfrac{1}{32}\,\varepsilon\,({\overline{\gamma}_{\delta}})^{2}-\delta^{2}\,({\gamma_{\delta}})^{2}\right)\alpha-12KD^{2}C^{2}(B+1)^{2}\rho^{2}-3\,\delta^{2}({\gamma_{\delta}})^{2}>0.

Since xx is semi-static for L+ϕL+\phi, and then also for 𝕃{\mathbb{L}}, and by (148) c(𝕃)=0c({\mathbb{L}})=0, the total action is finite:

A𝕃(x|],0])maxy,zMΦc(𝕃)𝕃(y,z)<+.A_{\mathbb{L}}(x|_{]-\infty,0]})\leq\max_{y,z\in M}\Phi^{\mathbb{L}}_{c({\mathbb{L}})}(y,z)<+\infty.

Therefore there must be at most finitely many TkT_{k}’s.

By item (5) in claim 5.4.2, we have that α\alpha-limit(x,x˙)=Γ(x,{\dot{x}})=\Gamma. Since α\alpha-limits of semi-static orbits are static (Mañé [24, theorem V.(c)]), we obtain that Γ𝒜(L+ϕ)\Gamma\subset{\mathcal{A}}(L+\phi). This finishes the proof of proposition 5.3.

Appendix A Shadowing

Let ψ\psi be the flow of a C1C^{1} vector field on a compact manifold MM. A compact ψ\psi-invariant subset ΛM\Lambda\subset M is hyperbolic for ψ\psi if the tangent bundle restricted to Λ\Lambda is decomposed as the Whitney sum TΛM=EsEEuT_{\Lambda}M=E^{s}\oplus E\oplus E^{u}, where EE is the 1-dimensional vector bundle tangent to the flow and there are constants C,λ>0C,\lambda>0 such that

  1. (a)

    Dψt(Es)=EsD\psi_{t}(E^{s})=E^{s}, Dψt(Eu)=EuD\psi_{t}(E^{u})=E^{u} for all tt\in{\mathbb{R}}.

  2. (b)

    |Dψt(v)|Ceλt|v||D\psi_{t}(v)|\leq C\,\text{\rm\large e}^{-\lambda t}|v| for all vEsv\in E^{s}, t0t\geq 0.

  3. (c)

    |Dψt(u)|Ceλt|u||D\psi_{-t}(u)|\leq C\,\text{\rm\large e}^{-\lambda t}|u| for all uEuu\in E^{u}, t0t\geq 0.

It follows from the definition that the hyperbolic splittig EsEEuE^{s}\oplus E\oplus E^{u} over Λ\Lambda is continuous.

From now on we shall assume that Λ\Lambda does not contain fixed points for ψ\psi. For xΛx\in\Lambda define the following stable and unstable sets:

Wss(x):\displaystyle W^{ss}(x): ={yM|d(ψt(x),ψt(y))0 as t+},\displaystyle=\{\,y\in M\;|\;d(\psi_{t}(x),\psi_{t}(y))\to 0\text{ as }t\to+\infty\,\},
Wεss(x):\displaystyle W^{ss}_{\varepsilon}(x): ={yWss(x)|d(ψt(x),ψt(y))εt0},\displaystyle=\{\,y\in W^{ss}(x)\;|\;d(\psi_{t}(x),\psi_{t}(y))\leq\varepsilon\;\;\forall t\geq 0\,\},
Wuu(x):\displaystyle W^{uu}(x): ={yM|d(ψt(x),ψt(y))0 as t+},\displaystyle=\{\,y\in M\;|\;d(\psi_{-t}(x),\psi_{-t}(y))\to 0\text{ as }t\to+\infty\,\},
(152) Wεuu(x):\displaystyle W^{uu}_{\varepsilon}(x): ={yWuu(x)|d(ψt(x),ψt(y))εt0},\displaystyle=\{\,y\in W^{uu}(x)\;|\;d(\psi_{-t}(x),\psi_{-t}(y))\leq\varepsilon\;\;\forall t\geq 0\,\},
Wεs(x):\displaystyle W^{s}_{\varepsilon}(x): ={yM|d(ψt(x),ψt(y))εt0},\displaystyle=\{\,y\in M\;|\;d(\psi_{t}(x),\psi_{t}(y))\leq\varepsilon\;\;\forall t\geq 0\,\},
Wεu(x):\displaystyle W^{u}_{\varepsilon}(x): ={yM|d(ψt(x),ψt(y))εt0}.\displaystyle=\{\,y\in M\;|\;d(\psi_{-t}(x),\psi_{-t}(y))\leq\varepsilon\;\;\forall t\geq 0\,\}.

Conditions {(a),(b),(c)} are equivalent to {(a),(d)}, where

  1. (d)

    There exists T>0T>0 such that DψT|Es<12\left\|D\psi_{T}|_{E^{s}}\right\|<\tfrac{1}{2}  and  DψT|Eu<12\left\|D\psi_{-T}|_{E^{u}}\right\|<\tfrac{1}{2}.

Let 𝔛k(M){\mathfrak{X}}^{k}(M) be the Banach manifold of the CkC^{k} vector fields on MM, k1k\geq 1. Let X=tψtX=\partial_{t}\psi_{t} be the vector field of ψt\psi_{t}. For Y𝔛k(M)Y\in{\mathfrak{X}}^{k}(M) denote by ψtY\psi^{Y}_{t} the flow of YY.

A.1 Proposition.

There are open sets X𝒰𝔛1(M)X\in{\mathcal{U}}\subset{\mathfrak{X}}^{1}(M) and ΛUM\Lambda\subset U\subset M such that for every Y𝒰Y\in{\mathcal{U}} the set ΛY:=tψtY(U¯)\Lambda_{Y}:=\bigcap_{t\in{\mathbb{R}}}\psi^{Y}_{t}(\overline{U}) is hyperbolic for the flow ψtY\psi^{Y}_{t} of YY, with uniform constants CC, λ\lambda, TT on (b), (c) and (d).

Proposition A.1 can be proven by a characterization of hyperbolicity using cones (cf. Hasselblatt-Katok [18, Proposition 17.4.4]) and obtaining uniform contraction (expansion) for a fixed iterate in ΛY\Lambda_{Y}. See Fisher-Hasselblatt [17] prop. 5.1.8 p. 256].

A.2.

Proposition [19, 5.6, p. 63], [4, 1.3], [17, 6.6.1].

There are constants C,λ>0C,\,\lambda>0 such that, for small ε\varepsilon,

  1. (a)

    d(ψt(x),ψt(y))Ceλtd(x,y)d\big{(}\psi_{t}(x),\psi_{t}(y)\big{)}\leq C\,\text{\rm\large e}^{-\lambda t}\,d(x,y) when xΛx\in\Lambda, yWεss(x)y\in W_{\varepsilon}^{ss}(x), t0t\geq 0.

  2. (b)

    d(ψt(x),ψt(y))Ceλtd(x,y)d\big{(}\psi_{-t}(x),\psi_{-t}(y)\big{)}\leq C\,\text{\rm\large e}^{-\lambda t}\,d(x,y) when xΛx\in\Lambda, yWεuu(x)y\in W_{\varepsilon}^{uu}(x), t0t\geq 0.

A.3.

Canonical Coordinates [28, 3.1], [19, 4.1], [30, 7.4], [4, 1.4], [5, 1.2], [17, 6.2.2]:

There are α,γ>0\alpha,\,\gamma>0 for which the following is true: If x,yΛx,y\in\Lambda and d(x,y)αd(x,y)\leq\alpha then there is a unique v=v(x,y)v=v(x,y)\in{\mathbb{R}} with |v|γ|v|\leq\gamma such that

(153) x,y:=Wγss(ψv(x))Wγuu(y).\langle x,y\rangle:=W_{\gamma}^{ss}(\psi_{v}(x))\cap W^{uu}_{\gamma}(y)\neq\emptyset.

This set consists of a single point, which we denote x,yM\langle x,y\rangle\in M. The maps vv and ,\langle\;,\;\rangle are continuous on the set {(x,y)|d(x,y)α}Λ×Λ\{\,(x,y)\;|\;d(x,y)\leq\alpha\,\}\subset\Lambda\times\Lambda.

A.4 Lemma.

There are η0>0\eta_{0}>0, B>1B>1, and open sets ΛU\Lambda\subset U, X𝒰𝔛k(M)X\in{\mathcal{U}}\subset{\mathfrak{X}}^{k}(M) such that
if d(x,y)η0d(x,y)\leq\eta_{0}, Y𝒰Y\in{\mathcal{U}}, x,yΛUYx,\,y\in\Lambda^{Y}_{U} and η=Bd(x,y)\eta=B\,d(x,y) then

(154) x,yWηss(ψvY(x))Wηuu(y)with |v(x,y)|η\displaystyle\langle x,y\rangle\in W^{ss}_{\eta}(\psi^{Y}_{v}(x))\cap W^{uu}_{\eta}(y)\qquad\text{with }\quad|v(x,y)|\leq\eta\qquad
(155) and d(x,ψvY(x))η.\displaystyle\text{and }\qquad d(x,\psi^{Y}_{v}(x))\leq\eta.
Proof:.

We have that x,x=x\langle x,x\rangle=x and v(x,x)=0v(x,x)=0. By uniform continuity, given δ>0\delta>0, for d(x,y)d(x,y) small enough

(156) d(x,y,x)δ,d(x,y,y)δ,d(\langle x,y\rangle,x)\leq\delta,\qquad d(\langle x,y\rangle,y)\leq\delta,

and v=v(x,y)v=v(x,y) is so small that

(157) d(ψv(x),x)δ.d(\psi_{v}(x),x)\leq\delta.

The continuity of the hyperbolic splitting implies that the angles (Es,Eu)\measuredangle(E^{s},E^{u}), (Y,Es)\measuredangle(Y,E^{s}) and (EsY,Eu)\measuredangle(E^{s}\oplus{\mathbb{R}}Y,E^{u}) are bounded away from zero, uniformly on ΛVY:=tψtY(V¯)\Lambda^{Y}_{V}:=\bigcap_{t\in{\mathbb{R}}}\psi^{Y}_{-t}(\overline{V}), for some VUV\supset U and all YY in an open set 𝒰0𝔛1(M){\mathcal{U}}_{0}\subset{\mathfrak{X}}^{1}(M) with X𝒰0X\in{\mathcal{U}}_{0}. There is β1>0\beta_{1}>0 such that if x,yΛUYx,\,y\in\Lambda^{Y}_{U} and d(x,y)<β1d(x,y)<\beta_{1} then

x,y=Wγs(x)Wγuu(y)V.\langle x,y\rangle=W^{s}_{\gamma}(x)\cap W^{uu}_{\gamma}(y)\in V.

The strong local invariant manifolds WγssW^{ss}_{\gamma}, WγuuW^{uu}_{\gamma} are tangent to EsE^{s}, EuE^{u} at ΛVY\Lambda^{Y}_{V} and for a fixed γ\gamma as C1C^{1} submanifolds they vary continuously on the base point xMx\in M and on the vector field in the C1C^{1} topology (cf. [15, Thm. 4.3],[19, Thm. 4.1]). There is a family of small cones EXu(x)Cu(x)TxME^{u}_{X}(x)\subset C^{u}(x)\subset T_{x}M, EXs(x)Cs(x)TxME^{s}_{X}(x)\subset C^{s}(x)\subset T_{x}M defined on a neighborhood WW of Λ\Lambda invariant under Dψ1YD\psi^{Y}_{-1} and Dψ1YD\psi^{Y}_{1} respectively, for YY in a C1C^{1} neighborhood 𝒲{\mathcal{W}} of XX. The exponential of these cones contain Wγuu(x)W^{uu}_{\gamma}(x) and Wγss(x)W^{ss}_{\gamma}(x) for xΛWYx\in\Lambda^{Y}_{W} and Y𝒲Y\in{\mathcal{W}}. The angles between these cones are uniformly bounded away from zero, so for example if zuWuu(x)z^{u}\in W^{uu}(x), zsWss(x)z^{s}\in W^{ss}(x) and d(zu,x)d(z^{u},x), d(zs,x)d(z^{s},x) are small, then d(zu,x)+d(zs,x)<A0d(zu,zs)d(z^{u},x)+d(z^{s},x)<A_{0}\,d(z^{u},z^{s}) for some A0>0A_{0}>0. We can construct similiar cones separating EuE^{u} from EsXE^{s}\oplus{\mathbb{R}}X.

Shrinking UU and 𝒰{\mathcal{U}} if necessary there are 0<β2<β10<\beta_{2}<\beta_{1} and A1,A2,A3>0A_{1},\,A_{2},\,A_{3}>0 such that if Y𝒰Y\in{\mathcal{U}}, x,yΛUYx,\,y\in\Lambda^{Y}_{U} and d(x,y)<β2d(x,y)<\beta_{2}, taking w:=x,yWγs(x)Wγuu(y)w:=\langle x,y\rangle\in W^{s}_{\gamma}(x)\cap W^{uu}_{\gamma}(y) and vv such that wWγss(ψvY(x))w\in W^{ss}_{\gamma}(\psi^{Y}_{v}(x)), i.e. ψvY(x)ψ[1,1]Y(x)Wγss(w)\psi^{Y}_{v}(x)\in\psi^{Y}_{[-1,1]}(x)\cap W^{ss}_{\gamma}(w), then

(158) d(x,w)+d(w,y)\displaystyle d(x,w)+d(w,y) A1d(x,y),\displaystyle\leq A_{1}\,d(x,y),
(159) d(x,ψvY(x))+d(ψvY(x),w)\displaystyle d(x,\psi^{Y}_{v}(x))+d(\psi^{Y}_{v}(x),w) A2d(x,w)A2A1d(x,y),\displaystyle\leq A_{2}\,d(x,w)\leq A_{2}A_{1}\,d(x,y),
|v|A3d(x,ψvY(x))\displaystyle|v|\leq A_{3}\,d(x,\psi^{Y}_{v}(x)) A3A2A1d(x,y).\displaystyle\leq A_{3}A_{2}A_{1}\,d(x,y).

We can assume that 𝒰0{\mathcal{U}}_{0} and UU are so small that the constants CC, λ\lambda, ε\varepsilon in Proposition A.2 can be taken uniform for all Y𝒰0Y\in{\mathcal{U}}_{0} and in ΛUY\Lambda^{Y}_{U}. By Proposition A.2, since w:=x,yWγss(ψv(x))w:=\langle x,y\rangle\in W^{ss}_{\gamma}(\psi_{v}(x)), we have that

t0d(ψtY(x,y),ψtY(ψvY(x)))\displaystyle\forall t\geq 0\qquad d\big{(}\psi^{Y}_{t}(\langle x,y\rangle),\psi^{Y}_{t}(\psi^{Y}_{v}(x))\big{)} Ceλtd(w,ψvY(x))\displaystyle\leq C\,\text{\rm\large e}^{-\lambda t}\,d(w,\psi_{v}^{Y}(x))
A2A1Ceλtd(x,y) using (159).\displaystyle\leq A_{2}A_{1}C\,\text{\rm\large e}^{-\lambda t}\,d(x,y)\qquad\text{ using~{}\eqref{a2xpw}.}

Take B1:=(1+A2)A1CB_{1}:=(1+A_{2})A_{1}C. Then if d(x,y)<β2d(x,y)<\beta_{2} and η=B1d(x,y)\eta=B_{1}\,d(x,y) we obtain that x,yWηss(ψvY(x))\langle x,y\rangle\in W^{ss}_{\eta}(\psi^{Y}_{v}(x)).

Since w=x,yWγuu(y)w=\langle x,y\rangle\in W^{uu}_{\gamma}(y) we have that

t0d(ψtY(x,y),ψtY(y))\displaystyle\forall t\geq 0\qquad d(\psi^{Y}_{-t}(\langle x,y\rangle),\psi^{Y}_{-t}(y)) Ceλtd(w,y)\displaystyle\leq C\,\text{\rm\large e}^{-\lambda t}\,d(w,y)
A1Ceλtd(x,y)using (158).\displaystyle\leq A_{1}C\,\text{\rm\large e}^{-\lambda t}d(x,y)\qquad\text{using }\eqref{a1xwy}.

Thus if η=B1d(x,y)\eta=B_{1}\,d(x,y) then x,yWηuu(y)\langle x,y\rangle\in W^{uu}_{\eta}(y).

By (156) and (157) there is 0<η0<β20<\eta_{0}<\beta_{2} such that if d(x,y)η0d(x,y)\leq\eta_{0} then d(w,x)d(w,x), d(w,y)d(w,y) and d(ψv(x),x)d(\psi_{v}(x),x) are small enough to satisfy the above inequalities. Now let

B:=max{2,B1,A3A2A1,A2A1}.B:=\max\{2,\,B_{1},\,A_{3}A_{2}A_{1},\,A_{2}A_{1}\}.

A.5 Proposition.

There are open sets X𝒰𝔛1(M)X\in{\mathcal{U}}\subset{\mathfrak{X}}^{1}(M) and ΛUM\Lambda\subset U\subset M and η0,γ>0\eta_{0},\gamma>0, B>1B>1 such that

η>0β=β(η)=1Bmin{η,η0}Y𝒰\forall\eta>0\qquad\exists\beta=\beta(\eta)=\tfrac{1}{B}\,\min\{\eta,\eta_{0}\}\qquad\forall Y\in{\mathcal{U}}

if ψt=ψtY\psi_{t}=\psi^{Y}_{t} is the flow of YY, x,yΩUY:=tψt(U¯)x,y\in\Omega^{Y}_{U}:=\bigcap_{t\in{\mathbb{R}}}\psi_{t}(\overline{U}) and s:s:{\mathbb{R}}\to{\mathbb{R}} continuous with s(0)=0s(0)=0 satisfy

(160) d(ψt+s(t)(y),ψt(x))β for |t|L,d(\psi_{t+s(t)}(y),\psi_{t}(x))\leq\beta\quad\text{ for }|t|\leq L,

then

(161) |s(t)|3η for all |t|L,|v(x,y)|η and |s(t)|\leq 3\eta\quad\text{ for all }|t|\leq L,\qquad|v(x,y)|\leq\eta\quad\text{ and }
|s|L,\displaystyle\forall|s|\leq L,\qquad d(ψs(y),ψs+v(x))Ceλ(L|s|)[d(ψL(w),ψL(y))+d(ψL(w),ψL+v(x))],\displaystyle d(\psi_{s}(y),\psi_{s+v}(x))\leq C\,\text{\rm\large e}^{-\lambda(L-|s|)}\,\big{[}d(\psi_{L}(w),\psi_{L}(y))+d(\psi_{-L}(w),\psi_{-L+v}(x))\big{]},
(162) where w:=x,y=Wγss(ψv(x))Wγuu(y).\displaystyle\text{where }\qquad w:=\langle x,y\rangle=W^{ss}_{\gamma}(\psi_{v}(x))\cap W^{uu}_{\gamma}(y).

also

(163) |s|L,d(ψs(y),ψsψv(x))Cγeλ(L|s|).\forall|s|\leq L,\qquad d(\psi_{s}(y),\psi_{s}\psi_{v}(x))\leq C\,\gamma\,e^{-\lambda(L-|s|)}.

In particular

d(y,ψv(x))CγeλL.d(y,\psi_{v}(x))\leq C\,\gamma\,e^{-\lambda L}.
Proof:.

Let γ=Bη0\gamma=B\eta_{0} with {η0,B}\{\eta_{0},B\} from A.4. We may assume that η\eta is so small that

(164) η<γ8,\displaystyle\eta<\tfrac{\gamma}{8},
(165) sup{d(ψu(x),x):xM,|u|4η}γ8.\displaystyle\sup\{\,d(\psi_{u}(x),x)\;:\;x\in M,\,|u|\leq 4\eta\,\}\leq\tfrac{\gamma}{8}.

Let

(166) β=β(η):=1Bmin{η,η0},\beta=\beta(\eta):=\tfrac{1}{B}\,\min\{\eta,\eta_{0}\},

where B>1B>1 and η0\eta_{0} are from lemma A.4. Consider xx, yy and s(t)s(t) as in the hypothesis. Since s(0)=0s(0)=0 we have that d(x,y)βd(x,y)\leq\beta. Using lemma A.4 we can define

(167) w:=x,y=Wηss(ψv(x))Wηuu(y),w:=\langle x,y\rangle=W^{ss}_{\eta}(\psi_{v}(x))\cap W^{uu}_{\eta}(y)\neq\emptyset,

we also have

(168) |v|=|v(x,y)|η.|v|=|v(x,y)|\leq\eta.

Define the sets

A\displaystyle A :={t[0,L]:|s(t)|3η\displaystyle:=\{\,t\in[0,L]\;:\;|s(t)|\geq 3\eta\; or d(ψt(y),ψt(w))12γ},\displaystyle\text{ or }\;d(\psi_{t}(y),\psi_{t}(w))\geq\tfrac{1}{2}{\gamma}\,\},
B\displaystyle B :={t[0,L]:|s(t)|3η\displaystyle:=\{\,t\in[0,L]\;:\;|s(-t)|\geq 3\eta\; or d(ψt+v(x),ψt(w))12γ}.\displaystyle\text{ or }\;d(\psi_{-t+v}(x),\psi_{-t}(w))\geq\tfrac{1}{2}\gamma\,\}.

Suppose that AA\neq\emptyset. Let t1:=infAt_{1}:=\inf A. Then d(ψt(y),ψt(w))12γd(\psi_{t}(y),\psi_{t}(w))\leq\tfrac{1}{2}\,\gamma, t[0,t1]\forall t\in[0,t_{1}]. Since wWηuu(y)w\in W^{uu}_{\eta}(y) and by (164), η<18γ\eta<\tfrac{1}{8}\gamma; from (152) we have that d(ψt(y),ψt(w))18γd(\psi_{t}(y),\psi_{t}(w))\leq\tfrac{1}{8}\gamma, t0\forall t\leq 0. Therefore

(169) d(ψt1r(y),ψt1r(w))12γ,r0.d(\psi_{t_{1}-r}(y),\psi_{t_{1}-r}(w))\leq\tfrac{1}{2}\,\gamma,\qquad\forall r\geq 0.

Since ss is continuous, s(0)=0s(0)=0 and t1At_{1}\in\partial A, we have that |s(t1)|3η|s(t_{1})|\leq 3\eta. Using (165) twice with u=|s(t1)|u=|s(t_{1})|, (169) and the triangle inequality we obtain

d(ψt1+s(t1)r(y),ψt1+s(t1)r(w))34γ,r0.d(\psi_{t_{1}+s(t_{1})-r}(y),\psi_{t_{1}+s(t_{1})-r}(w))\leq\tfrac{3}{4}\gamma,\qquad\forall r\geq 0.

Hence ψt1+s(t1)(w)Wγuu(ψt1+s(t1)(y))\psi_{t_{1}+s(t_{1})}(w)\in W^{uu}_{\gamma}(\psi_{t_{1}+s(t_{1})}(y)). From (167), wWηss(ψv(x))w\in W^{ss}_{\eta}(\psi_{v}(x)), and then

(170) d(ψr(w),ψr+v(x))η<γ8,r0.d(\psi_{r}(w),\psi_{r+v}(x))\leq\eta<\tfrac{\gamma}{8},\qquad\forall r\geq 0.

Since |s(t1)|3η|s(t_{1})|\leq 3\eta, using (165) twice with u=s(t1)u=s(t_{1}), and (170) with r=t1+p0r=t_{1}+p\geq 0, and the triangle inequality, we get

d(ψt1+s(t1)+p(w),ψt1+s(t1)+v+p(x))3γ8,p0.d(\psi_{t_{1}+s(t_{1})+p}(w),\psi_{t_{1}+s(t_{1})+v+p}(x))\ \leq\tfrac{3\gamma}{8},\qquad\forall p\geq 0.

Hence ψt1+s(t1)(w)Wγss(ψs(t1)+v(ψt1(x)))\psi_{t_{1}+s(t_{1})}(w)\in W^{ss}_{\gamma}(\psi_{s(t_{1})+v}(\psi_{t_{1}}(x))). We have shown that

(171) ψt1+s(t1)(w)Wγss(ψs(t1)+v(ψt1(x)))Wγuu(ψt1+s(t1)(y)).\psi_{t_{1}+s(t_{1})}(w)\in W^{ss}_{\gamma}(\psi_{s(t_{1})+v}(\psi_{t_{1}}(x)))\cap W^{uu}_{\gamma}(\psi_{t_{1}+s(t_{1})}(y)).

Since |s(t1)+v||s(t1)|+|v|4η<γ|s(t_{1})+v|\leq|s(t_{1})|+|v|\leq 4\eta<\gamma and by (160),

(172) d(ψt1+s(t1)(y),ψt1(x))β,d(\psi_{t_{1}+s(t_{1})}(y),\psi_{t_{1}}(x))\leq\beta,

equation (171) implies that

v(ψt1(x),ψt1+s(t1)(y))=s(t1)+v(x,y),\displaystyle v(\psi_{t_{1}}(x),\psi_{t_{1}+s(t_{1})}(y))=s(t_{1})+v(x,y),
ψt1+s(t1)(w)=ψt1(x),ψt1+s(t1)(y).\displaystyle\psi_{t_{1}+s(t_{1})}(w)=\langle\psi_{t_{1}}(x),\psi_{t_{1}+s(t_{1})}(y)\rangle.

By Lemma A.4, (172) and (166),

(173) |s(t1)+v|η and \displaystyle|s(t_{1})+v|\leq\eta\qquad\text{ and }
ψt1+s(t1)(w)Wηuu(ψt1+s(t1)(y)),in particular\displaystyle\psi_{t_{1}+s(t_{1})}(w)\in W^{uu}_{\eta}(\psi_{t_{1}+s(t_{1})}(y)),\;\text{in particular}
(174) d(ψt1+s(t1)(w),ψt1+s(t1)(y))η.\displaystyle d(\psi_{t_{1}+s(t_{1})}(w),\psi_{t_{1}+s(t_{1})}(y))\leq\eta.

Since |s(t1)|3η|s(t_{1})|\leq 3\eta, from (165), (174) and (164), we get that

d(ψt1(w),ψt1(y))η+2(γ8)3γ8.d(\psi_{t_{1}}(w),\psi_{t_{1}}(y))\leq\eta+2\left(\tfrac{\gamma}{8}\right)\leq\tfrac{3\gamma}{8}.

From (173) and (168) we have that

|s(t1)||s(t1)+v|+|v|2η.|s(t_{1})|\leq|s(t_{1})+v|+|v|\leq 2\eta.

These statements contradict t1At_{1}\in A. Hence A=A=\emptyset.

Similarly one shows that B=B=\emptyset. Since A=A=\emptyset, inequality (175) holds for all t[0,L]t\in[0,L]. From (167), wWηuu(y)w\in W^{uu}_{\eta}(y) and by (164), η<γ8\eta<\tfrac{\gamma}{8}; thus inequality (175) also holds for t0t\leq 0.

(175) tLd(ψt(y),ψt(w))<12γ.\forall t\leq L\qquad d(\psi_{t}(y),\psi_{t}(w))<\tfrac{1}{2}{\gamma}.

Therefore

(176) ψL(w)W12γuu(ψL(y)).\psi_{L}(w)\in W^{uu}_{\frac{1}{2}\gamma}(\psi_{L}(y)).

From Proposition A.2 we get

|s|Ld(ψs(w),ψs(y))Ceλ(L|s|)d(ψL(w),ψL(y)).\forall|s|\leq L\qquad d(\psi_{s}(w),\psi_{s}(y))\leq C\,\text{\rm\large e}^{-\lambda(L-|s|)}\,d(\psi_{L}(w),\psi_{L}(y)).

Similarly, B=B=\emptyset imples that

(177) ψL(w)W12γss(ψL+v(x)) and \psi_{-L}(w)\in W^{ss}_{\frac{1}{2}\gamma}(\psi_{-L+v}(x))\qquad\text{ and }
|s|Ld(ψs(w),ψs+v(x))Ceλ(L|s|)d(ψL(w),ψL+v(x)).\forall|s|\leq L\qquad d(\psi_{s}(w),\psi_{s+v}(x))\leq C\,\text{\rm\large e}^{-\lambda(L-|s|)}\,d(\psi_{-L}(w),\psi_{-L+v}(x)).

Adding these inequalities we obtain

|s|L\displaystyle\forall|s|\leq L\qquad d(ψs(y),ψs+v(x))Ceλ(L|s|)[d(ψL(w),ψL(y))+d(ψL(w),ψL+v(x))],\displaystyle d(\psi_{s}(y),\psi_{s+v}(x))\leq C\,\text{\rm\large e}^{-\lambda(L-|s|)}\,\big{[}d(\psi_{L}(w),\psi_{L}(y))+d(\psi_{-L}(w),\psi_{-L+v}(x))\big{]},
(178) where w:=x,y=Wγss(ψv(x))Wγuu(y).\displaystyle\text{where }\qquad w:=\langle x,y\rangle=W^{ss}_{\gamma}(\psi_{v}(x))\cap W^{uu}_{\gamma}(y).

This proves inequality (162).

From (168), |v(x,y)|η|v(x,y)|\leq\eta. The fact AB=A\cup B=\emptyset also gives |s(t)|3η|s(t)|\leq 3\eta for t[L,L]t\in[-L,L]. This proves (161). From (176), (177) and (178) we get inequality (163).

A.6 Proposition.

Let γ\gamma, η0\eta_{0} and β=β(η)\beta=\beta(\eta) be from Proposition A.5. Given η<min{η0,12γ}\eta<\min\{\eta_{0},\tfrac{1}{2}\gamma\}

  1. (a)

    If x,yΛx,\,y\in\Lambda and s:[0,+[s:[0,+\infty[\to{\mathbb{R}} continuous with s(0)=0s(0)=0 satisfy

    d(ψt+s(t)(y),ψt(x))βt0,d(\psi_{t+s(t)}(y),\psi_{t}(x))\leq\beta\qquad\forall t\geq 0,

    then |s(t)|3η|s(t)|\leq 3\eta for all t0t\geq 0 and there is |v(x,y)|η|v(x,y)|\leq\eta such that yWγss(ψv(x))y\in W^{ss}_{\gamma}(\psi_{v}(x)).

  2. (b)

    Similarly, if x,yΛx,\,y\in\Lambda, s:],0]s:]-\!\infty,0]\to{\mathbb{R}} is continuous with s(0)=0s(0)=0 and

    d(ψt+s(t)(y),ψt(x))βt0,d(\psi_{t+s(t)}(y),\psi_{t}(x))\leq\beta\qquad\forall t\leq 0,

    then |s(t)|3η|s(t)|\leq 3\eta for all t0t\leq 0 and there is |v(x,y)|η|v(x,y)|\leq\eta such that yWγuu(ψv(x))y\in W^{uu}_{\gamma}(\psi_{v}(x)).

Proof:.

We only prove item (a). The same proof as in Proposition A.5 shows that taking

w:=x,y=Wηss(ψv(x))Wηuu(y),w:=\langle x,y\rangle=W^{ss}_{\eta}(\psi_{v}(x))\cap W^{uu}_{\eta}(y)\neq\emptyset,

we have that |v|=|v(x,y)|η|v|=|v(x,y)|\leq\eta and

=A:={t[0,+[:|s(t)|3η or d(ψt(y),ψt(w))12γ}.\emptyset=A:=\{\,t\in[0,+\infty[\;:\;|s(t)|\geq 3\eta\;\text{ or }\;d(\psi_{t}(y),\psi_{t}(w))\geq\tfrac{1}{2}{\gamma}\,\}.

Therefore |s(t)|3η|s(t)|\leq 3\eta for all t0t\geq 0 and wW12γss(y)Wηss(ψv(x))w\in W^{ss}_{\frac{1}{2}\gamma}(y)\cap W^{ss}_{\eta}(\psi_{v}(x)). Since 12γ+η<γ\tfrac{1}{2}\gamma+\eta<\gamma we get that yWγss(ψv(x))y\in W^{ss}_{\gamma}(\psi_{v}(x)).

A.7 Proposition.

There are D>0D>0, β0>0\beta_{0}>0 and open sets X𝒰𝔛1(M)X\in{\mathcal{U}}\subset{\mathfrak{X}}^{1}(M), ΛUM\Lambda\subset U\subset M, such that

β]0,β0]Y𝒰,\forall\beta\in]0,\beta_{0}]\qquad\forall Y\in{\mathcal{U}},

if Y𝒰Y\in{\mathcal{U}}, ψt=ψtY\psi_{t}=\psi^{Y}_{t} is the flow of YY, x,yΛUY:=tψt(U¯)x,\,y\in\Lambda^{Y}_{U}:=\bigcap_{t\in{\mathbb{R}}}\psi_{t}(\overline{U}) and s:s:{\mathbb{R}}\to{\mathbb{R}} continuous with s(0)=0s(0)=0 satisfy

(179) d(ψt+s(t)(y),ψt(x))β for |t|L,d(\psi_{t+s(t)}(y),\psi_{t}(x))\leq\beta\quad\text{ for }|t|\leq L,

then |s(t)|Dβ|s(t)|\leq D\beta for all |t|L|t|\leq L and there is |v|=|v(x,y)|Dβ|v|=|v(x,y)|\leq D\beta such that

|s|L,d(ψs(y),ψs+v(x))Dβeλ(L|s|).\displaystyle\forall|s|\leq L,\qquad d(\psi_{s}(y),\psi_{s+v}(x))\leq D\,\beta\,\text{\rm\large e}^{-\lambda(L-|s|)}.

Moreover for all |s|L|s|\leq L,

(180) d(ψs(y),ψs+v(x))Deλ(L|s|)[d(ψL(y),ψL+v(x))+d(ψL(y),ψL+v(x))],d(\psi_{s}(y),\psi_{s+v}(x))\leq D\,\text{\rm\large e}^{-\lambda(L-|s|)}\,\big{[}d(\psi_{L}(y),\psi_{L+v}(x))+d(\psi_{-L}(y),\psi_{-L+v}(x))\big{]},

and vv is determined by

x,y=Wγss(ψv(x))Wγuu(y).\langle x,y\rangle=W^{ss}_{\gamma}(\psi_{v}(x))\cap W^{uu}_{\gamma}(y)\neq\emptyset.
Proof:.

Let CC, 𝒰{\mathcal{U}}, UU η0>0\eta_{0}>0 and BB be from Proposition A.5. The continuity of the hyperbolic splitting implies that the angle (Es,Eu)\measuredangle(E^{s},E^{u}) is bounded away from zero. As in the argument after (157), there are invariant families of cones separating EsE^{s} from EuE^{u} whose image under the exponential map contain the local invariant manifolds WγssW^{ss}_{\gamma}, WγuuW^{uu}_{\gamma}. And hence as in (158) there are A,β1>0A,\,\beta_{1}>0 such that if x,yΛUYx,\,y\in\Lambda^{Y}_{U}, d(x,y)<β1d(x,y)<\beta_{1} and

w=x,y=Wγss(ψv(x))Wγuu(y),w=\langle x,y\rangle=W^{ss}_{\gamma}(\psi_{v}(x))\cap W^{uu}_{\gamma}(y),

then

(181) d(w,ψv(x))+d(w,y)Ad(ψv(x),y).d(w,\psi_{v}(x))+d(w,y)\leq A\,d(\psi_{v}(x),y).

Suppose that 0<β<min{1Bη0,β1}0<\beta<\min\{\tfrac{1}{B}\eta_{0},\,\beta_{1}\} and xx, yy, s(t)s(t), ψtY\psi^{Y}_{t}, LL satisfy (179). Apply Proposition A.5 with η:=Bβ\eta:=B\beta.

Then |s(L)|3η|s(L)|\leq 3\eta, and

d(ψL(y),ψL(x))\displaystyle d(\psi_{L}(y),\psi_{L}(x)) d(ψL+s(L)(y),ψL(x))+|s(L)|Ysup\displaystyle\leq d(\psi_{L+s(L)}(y),\psi_{L}(x))+|s(L)|\cdot\|Y\|_{\sup}
β+3ηYsup<α,\displaystyle\leq\beta+3\eta\left\|Y\right\|_{\sup}<\alpha,

if β\beta is small enough. So that ψL(x),ψL(y)\langle\psi_{L}(x),\psi_{L}(y)\rangle is well defined. Similarly |s(L)|3η|s(-L)|\leq 3\eta and d(ψL(y),ψL(x))<αd(\psi_{-L}(y),\psi_{-L}(x))<\alpha. Since the time tt map ψt\psi_{t} preserves the family of strong invariant manifolds, in equation (162) we have that

ψL(w)\displaystyle\psi_{L}(w) =ψL(x),ψL(y)=Wγss(ψL+v(x))Wγuu(ψL(y)),\displaystyle=\langle\psi_{L}(x),\psi_{L}(y)\rangle=W^{ss}_{\gamma}(\psi_{L+v}(x))\cap W^{uu}_{\gamma}(\psi_{L}(y)),
ψL(w)\displaystyle\psi_{-L}(w) =ψL(x),ψL(y)=Wγss(ψL+v(x))Wγuu(ψL(y)).\displaystyle=\langle\psi_{-L}(x),\psi_{-L}(y)\rangle=W^{ss}_{\gamma}(\psi_{-L+v}(x))\cap W^{uu}_{\gamma}(\psi_{-L}(y)).

Therefore, using (181),

d(ψL(w),ψL(y))+d\displaystyle d(\psi_{L}(w),\psi_{L}(y))+d (ψL(w),ψL+v(x))\displaystyle(\psi_{-L}(w),\psi_{-L+v}(x))
(182) A[d(ψL+v(x),ψL(y))+d(ψL+v(x),ψL(y))],\displaystyle\leq A\big{[}d(\psi_{L+v}(x),\psi_{L}(y))+d(\psi_{-L+v}(x),\psi_{-L}(y))\big{]},
d(ψL+v(x),ψL(y))\displaystyle d(\psi_{L+v}(x),\psi_{L}(y)) d(ψL+v(x),ψL(x))+d(ψL(x),ψL+s(L)(y))+d(ψL+s(L)(y),ψL(y))\displaystyle\leq d(\psi_{L+v}(x),\psi_{L}(x))+d(\psi_{L}(x),\psi_{L+s(L)}(y))+d(\psi_{L+s(L)}(y),\psi_{L}(y))
|v|Ysup+β+|s(L)|Ysup\displaystyle\leq|v|\left\|Y\right\|_{\sup}+\beta+|s(L)|\,\left\|Y\right\|_{\sup}
B1β,\displaystyle\leq B_{1}\beta,

for some B1=B1(𝒰)>0B_{1}=B_{1}({\mathcal{U}})>0, because by Proposition A.5, |v|η|v|\leq\eta, |s(t)|3η|s(t)|\leq 3\eta and η=Bβ\eta=B\beta, so that

|v|Bβ,|s(t)|3Bβ.|v|\leq B\beta,\qquad|s(t)|\leq 3B\beta.

A similar estimate holds for d(ψL+v(x),ψL(y))d(\psi_{-L+v}(x),\psi_{-L}(y)) and hence from (182),

d(ψL(w),ψL(y))+d(ψL(w),ψL+v(x))2AB1β.d(\psi_{L}(w),\psi_{L}(y))+d(\psi_{-L}(w),\psi_{-L+v}(x))\leq 2AB_{1}\,\beta.

Replacing this in (162) we have that

|s|L,d(ψs(y),ψs+v(x))D1βeλ(L|s|),\forall|s|\leq L,\qquad d(\psi_{s}(y),\psi_{s+v}(x))\leq D_{1}\,\beta\,\text{\rm\large e}^{-\lambda(L-|s|)},

where D1=2AB1CD_{1}=2AB_{1}C.

By (182) and (162) we also have that

d(ψs(y),ψs+v(x))ACeλ(L|s|)[d(ψL(y),ψL+v(x))+d(ψL(y),ψL+v(x))].d(\psi_{s}(y),\psi_{s+v}(x))\leq AC\,\text{\rm\large e}^{-\lambda(L-|s|)}\,\big{[}d(\psi_{L}(y),\psi_{L+v}(x))+d(\psi_{-L}(y),\psi_{-L+v}(x))\big{]}.

Now take D:=max{D1,B, 3B,AC}D:=\max\{D_{1},\,B,\,3B,\,AC\,\}.

A.8 Definition.

We say that ψ|Λ\psi|_{\Lambda} is flow expansive if for every η>0\eta>0 there is α¯=α¯(η)>0\overline{\alpha}=\overline{\alpha}(\eta)>0 such that if xΛx\in\Lambda, yMy\in M and there is s:s:{\mathbb{R}}\to{\mathbb{R}} continuous with s(0)=0s(0)=0 and d(ψs(t)(y),ψt(x))α¯d(\psi_{s(t)}(y),\psi_{t}(x))\leq\overline{\alpha} for all tt\in{\mathbb{R}}, then y=ψv(x)y=\psi_{v}(x) for some |v|η|v|\leq\eta.

A.9 Remark.

Observe that Proposition A.5 implies uniform expansivity in a neighbourhood of (X,Λ)(X,\Lambda), namely there are neighbourhoods X𝒰𝔛1(M)X\in{\mathcal{U}}\subset{\mathfrak{X}}^{1}(M) and ΛUM\Lambda\subset U\subset M such that for every η>0\eta>0 there is α=α(η,𝒰,U)>0\alpha=\alpha(\eta,{\mathcal{U}},U)>0 such that if xΛUY:=tψtY(U¯)x\in\Lambda^{Y}_{U}:=\cap_{t\in{\mathbb{R}}}\psi^{Y}_{t}(\overline{U}), yMy\in M, s:(,0)(,0)s:({\mathbb{R}},0)\to({\mathbb{R}},0) continuous and t\forall t\in{\mathbb{R}}, d(ψs(t)Y(y),ψtY(x))<αd(\psi^{Y}_{s(t)}\big{(}y),\psi^{Y}_{t}(x)\big{)}<\alpha; then y=ψvY(x)y=\psi^{Y}_{v}(x) for some |v|<η|v|<\eta. See also Fisher-Hasselblatt [17, cor. 5.3.5].

This also implies uniform h-expansivity of their time-one maps as in Definition A.11.

A.10 Definition.

Let f:XXf:X\to X be a homeomorphism. For ε>0\varepsilon>0 and xXx\in X define

Γε(x,f):={yX|nd(fn(y),fn(x))ε}.\Gamma_{\varepsilon}(x,f):=\{\,y\in X\;|\;\forall n\in{\mathbb{Z}}\quad d(f^{n}(y),f^{n}(x))\leq\varepsilon\,\}.

We say that ff is entropy expansive or h-expansive if there is ε>0\varepsilon>0 such that

xXhtop(Γε(x,f),f)=0.\forall x\in X\qquad h_{\text{top}}(\Gamma_{\varepsilon}(x,f),f)=0.

Such an ε\varepsilon is called an h-expansive constant for ff.

A.11 Definition.

Let 𝒰{\mathcal{U}} be a topological subspace of C0(X,X)𝒰C^{0}(X,X)\supset{\mathcal{U}} and YXY\subseteq X compact. We say that 𝒰{\mathcal{U}} is uniformly h-expansive on YY if there is ε>0\varepsilon>0 such that

f𝒰yYhtop(Γε(y,f),f)=0.\forall f\in{\mathcal{U}}\quad\forall y\in Y\qquad h_{\text{top}}(\Gamma_{\varepsilon}(y,f),f)=0.

In our applications 𝒰{\mathcal{U}} will be a C1C^{1} neighbourhood of a diffeomorphism endowed with the C0C^{0} topology. An h-expansive homeomorphism corresponds to 𝒰={f}{\mathcal{U}}=\{f\}.

A.12 Definition.

Let L>0L>0, we say that (T,Γ)(T,\Gamma) is an LL-specification if

  1. (a)

    Γ={xi}iΛ\Gamma=\{x_{i}\}_{i\in{\mathbb{Z}}}\subset\Lambda.

  2. (b)

    T={ti}iT=\{t_{i}\}_{i\in{\mathbb{Z}}}\subset{\mathbb{R}} and ti+1tiLt_{i+1}-t_{i}\geq Li\forall i\in{\mathbb{Z}}.

We say that the specification (T,Γ)(T,\Gamma) is δ\delta-possible if

id(ψti(xi),ψti(xi1))δ.\forall i\in{\mathbb{Z}}\qquad d(\psi_{t_{i}}(x_{i}),\psi_{t_{i}}(x_{i-1}))\leq\delta.
A.13 Theorem.

Given >0\ell>0 there are δ0=δ0()>0\delta_{0}=\delta_{0}(\ell)>0 and E=E()>0E=E(\ell)>0 such that if 0<δ<δ00<\delta<\delta_{0} and (T,Γ)=({ti},{xi})i(T,\Gamma)=(\{t_{i}\},\{x_{i}\})_{i\in{\mathbb{Z}}} is a δ\delta-possible \ell-specification on Λ\Lambda then there exist yMy\in M and σ:\sigma:{\mathbb{R}}\to{\mathbb{R}} continuous, piecewise linear, strictly increasing with σ(t0)=t0\sigma(t_{0})=t_{0} and |σ(t)t|<Eδ|\sigma(t)-t|<E\,\delta such that

(183) it]ti,ti+1[d(ψσ(t)(y),ψt(xi))<Eδ.\forall i\in{\mathbb{Z}}\quad\forall t\in]t_{i},t_{i+1}[\qquad d\big{(}\psi_{\sigma(t)}(y),\psi_{t}(x_{i})\big{)}<E\,\delta.

Moreover, if the specification is periodic then yy is a periodic point for ψ\psi.

Theorem A.13 does not need that the hyperbolic set Λ\Lambda is locally maximal, but if not the point yy is not in Λ\Lambda. In Fisher-Hasselbaltt [17] the shadowing theorem A.13 is proved without a local maximality assumption and with the Lipschitz estimate (183) and σ(t)\sigma(t) a homeomorphism such that σ(t)t\sigma(t)-t has Lipschitz constant EδE\delta. But then proposition A.7 above proves the bound |σ(t)t|<Eδ|\sigma(t)-t|<E\delta and moreover, that σ(t)t\sigma(t)-t can be taken constant on each interval ]ti,ti+1[]t_{i},t_{i+1}[.

Theorem A.13 is proved in Bowen [4] (2.2) p. 6 with σ(t)t\sigma(t)-t constant on each ]ti,ti+1[]t_{i},t_{i+1}[ and without the estimate EδE\delta. A proof of theorem A.13 for flows without the local maximality hypothesis and with the explicit estimate EδE\delta appears in Palmer [26] theorem 9.3, p. 188. In [26], [27] the theorem requires an upper bound on the lengths of the intervals in TT. This is because there the theorem is proven also for perturbations of the flow. Indeed by Proposition A.7 longer intervals in TT improve the estimate.

References

  • [1] Patrick Bernard, Existence of C1,1C^{1,1} critical sub-solutions of the Hamilton-Jacobi equation on compact manifolds, Ann. Sci. École Norm. Sup. (4) 40 (2007), no. 3, 445–452, See also arXiv:math/0512018.
  • [2] Jairo Bochi, Genericity of periodic maximization: proof of Contreras’ theorem following Huang, Lian, Ma, Xu, and Zhang, manuscript, 2019.
  • [3] Rufus Bowen, Entropy-expansive maps, Trans. Amer. Math. Soc. 164 (1972), 323–331.
  • [4] by same author, Periodic orbits for hyperbolic flows, Amer. J. Math. 94 (1972), 1–30.
  • [5] by same author, Symbolic dynamics for hyperbolic flows, Amer. J. Math. 95 (1973), 429–460.
  • [6] Xavier Bressaud and Anthony Quas, Rate of approximation of minimizing measures, Nonlinearity 20 (2007), no. 4, 845–853.
  • [7] G. Contreras, A. Figalli, and L. Rifford, Generic hyperbolicity of Aubry sets on surfaces, Invent. Math. 200 (2015), no. 1, 201–261.
  • [8] G. Contreras, A. O. Lopes, and Ph. Thieullen, Lyapunov minimizing measures for expanding maps of the circle, Ergodic Theory Dynam. Systems 21 (2001), no. 5, 1379–1409.
  • [9] Gonzalo Contreras, Ground states are generically a periodic orbit, Invent. Math. 205 (2016), no. 2, 383–412.
  • [10] by same author, Generic Mañé Sets, Preprint, arXiv:1307.0559, arxiv.org, 2021.
  • [11] Gonzalo Contreras, Jorge Delgado, and Renato Iturriaga, Lagrangian flows: the dynamics of globally minimizing orbits. II, Bol. Soc. Brasil. Mat. (N.S.) 28 (1997), no. 2, 155–196, special issue in honor of Ricardo Mañé.
  • [12] Gonzalo Contreras and Renato Iturriaga, Convex Hamiltonians without conjugate points, Ergodic Theory Dynam. Systems 19 (1999), no. 4, 901–952.
  • [13] by same author, Global Minimizers of Autonomous Lagrangians, 22o22^{\text{o}} Coloquio Bras. Mat., IMPA, Rio de Janeiro, 1999.
  • [14] Gonzalo Contreras and Gabriel P. Paternain, Connecting orbits between static classes for generic Lagrangian systems, Topology 41 (2002), no. 4, 645–666.
  • [15] Sylvain Crovisier and Rafael Potrie, Introduction to partially hyperbolic dyanmics, lecture notes for a minicourse at ICTP School on Dynamical Systems 2015.
  • [16] Albert Fathi and Antonio Siconolfi, Existence of C1C^{1} critical subsolutions of the Hamilton-Jacobi equation, Invent. Math. 155 (2004), no. 2, 363–388.
  • [17] Todd Fisher and Boris Hasselblatt, Hiperbolic Flows, Zurich Lectures in Advanced Mathematics, European Mathematical Society, Berlin, 2019.
  • [18] Boris Hasselblatt and Anatole Katok, Introduction to the modern theory of dynamical systems, Cambridge University Press, Cambridge, 1995, With a supplementary chapter by Katok and Leonardo Mendoza.
  • [19] Morris W. Hirsch, Charles C. Pugh, and Michael Shub, Invariant manifolds, Springer-Verlag, Berlin, 1977, Lecture Notes in Mathematics, Vol. 583.
  • [20] Wen Huang, Zeng Lian, Xiao Ma, Leiye Xu, and Yiwei Zhang, Ergodic optimization theory for a class of typical maps, preprint arXiv:1904.01915, 2019.
  • [21] by same author, Ergodic optimization theory for Axiom A flows, preprint arXiv:1904.10608, 2019.
  • [22] Douglas Lind and Brian Marcus, An introduction to symbolic dynamics and coding, Cambridge Mathematical Library, Cambridge University Press, Cambridge, 2021, Second edition.
  • [23] Ricardo Mañé, Generic properties and problems of minimizing measures of Lagrangian systems, Nonlinearity 9 (1996), no. 2, 273–310.
  • [24] by same author, Lagrangian flows: the dynamics of globally minimizing orbits, International Conference on Dynamical Systems (Montevideo, 1995), Longman, Harlow, 1996, Reprinted in Bol. Soc. Brasil. Mat. (N.S.) 28 (1997), no. 2, 141–153., pp. 120–131.
  • [25] John N. Mather, Action minimizing invariant measures for positive definite Lagrangian systems, Math. Z. 207 (1991), no. 2, 169–207.
  • [26] Kenneth James Palmer, Shadowing in dynamical systems, Mathematics and its Applications, vol. 501, Kluwer Academic Publishers, Dordrecht, 2000, Theory and applications. MR 1885537
  • [27] by same author, Shadowing lemma for flows, Scholarpedia 4 (2009), no. 4, 7918.
  • [28] Charles Pugh and Michael Shub, The Ω\Omega-stability theorem for flows, Invent. Math. 11 (1970), 150–158.
  • [29] Anthony Quas and Jason Siefken, Ergodic optimization of super-continuous functions on shift spaces, Ergodic Theory and Dynamical Systems 32 (2012), no. 6, 2071–2082.
  • [30] Steve Smale, Differentiable dynamical systems. I: Diffeomorphisms; II: Flows; III: More on flows; IV: Other Lie groups, Bull. Am. Math. Soc. 73 (1967), 747–792, 795–804, 804–808; 808–817; Appendix to I: Anosov diffeomorphisms by John Mather, 792–795.
  • [31] Guo-Cheng Yuan and Brian R. Hunt, Optimal orbits of hyperbolic systems, Nonlinearity 12 (1999), no. 4, 1207–1224.