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

Action and periodic orbits on annulus

Yanxia Deng, Zhihong Xia School of Mathematics (Zhuhai), Sun Yat-sen University, Zhuhai, Guangdong, China Department of Mathematics, Northwestern University, Evanston, IL 60208 USA [email protected], [email protected]
(Date: version: June 10, 2021)
Abstract.

We consider the classical problem of area-preserving maps on annulus 𝔸=S1×[0,1]\mathbb{A}=S^{1}\times[0,1] . Let f\mathcal{M}_{f} be the set of all invariant probability measures of an area-preserving, orientation preserving diffeomorphism ff on 𝔸\mathbb{A}. Given any μ1\mu_{1} and μ2\mu_{2} in f\mathcal{M}_{f}, Franks [2][3], generalizing Poincaré’s last geometric theorem (Birkhoff [1]), showed that if their rotation numbers are different, then ff has infinitely many periodic orbits. In this paper, we show that if μ1\mu_{1} and μ2\mu_{2} have different actions, even if they have the same rotation number, then ff has infinitely many periodic orbits. In particular, if the action difference is larger than one, then ff has at least two fixed points. The same result is also true for area-preserving diffeomorphisms on unit disk, where no rotation number is available.

1. Introduction

Let 𝔸=S1×[0,1]\mathbb{A}=S^{1}\times[0,1], where S1=/S^{1}=\mathbb{R}/\mathbb{Z}, be an annulus with the standard area form ω=dydx\omega=dy\wedge dx, where xS1,y[0,1]x\in S^{1},y\in[0,1], and let ff be an area-preserving, orientation preserving diffeomorphism on 𝔸\mathbb{A} that preserves each boundary component. Let β\beta be a primitive of ω\omega, i.e. dβ=ωd\beta=\omega. Since ff is area-preserving and preserves each boundary component, we know fββf^{*}\beta-\beta is an exact 1-form. There is a function gg on 𝔸\mathbb{A} such that

dg=fββ.dg=f^{*}\beta-\beta.

The real valued function gg is called the action function for ff. The choice of gg depends on two factors, the first one is the integration constant, which can be fixed by assigning a zero value at a particular point. The second factor is the choice of β\beta. Two different choices of β\beta differ by a closed 1-form, we will show how the action depends on β\beta (Proposition 4). Let f\mathcal{M}_{f} be the set of all ff-invariant probability measures on 𝔸\mathbb{A}. For any μf\mu\in\mathcal{M}_{f}, the mean action, or simply the action, of μ\mu is defined to be

𝒜(μ)=𝔸g𝑑μ.\mathcal{A}(\mu)=\int_{\mathbb{A}}gd\mu.

When μ\mu is the area form, the action is called the Calabi invariant (cf. [4]).

Another important quantity that is associated with an invariant measure μf\mu\in\mathcal{M}_{f} is its rotation number. It measures the average rotation around the annulus. The precise definition of rotation numbers, and rotation vectors for general manifold, will be given in the next section. Given any two invariant measures μ1\mu_{1} and μ2\mu_{2} in f\mathcal{M}_{f}, Franks [2, 3], generalizing Poincaré’s last geometric theorem (Poincaré-Birkhoff Theorem [1]), showed that if the rotation numbers of μ1\mu_{1} and μ2\mu_{2} are different, then ff has infinitely many periodic orbits. More precisely, for any rational number p/qp/q between the rotation numbers of any two invariant measures, where pp and qq are relatively prime, there are at least two distinct periodic orbits of period qq with rotation number p/qp/q.

Poincaré’s last geometric theorem, more generally Franks’ theorem, shows that the difference in rotation numbers forces the existence of many other periodic orbits. In this paper, we will show that the same is true for differences in actions. But first, we need to point out that the difference in actions 𝒜(μ1)𝒜(μ2)\mathcal{A}(\mu_{1})-\mathcal{A}(\mu_{2}) for two invariant measures, in general, depends on the choice of β\beta, defined by dβ=ωd\beta=\omega. Interestingly, this dependence is closely related to their rotation numbers. This dependence will be made precise in the next section (Proposition 4). In particular, we will show that if two invariant measures μ1\mu_{1} and μ2\mu_{2} have exactly the same rotation number, then the action difference 𝒜(μ1)𝒜(μ2)\mathcal{A}(\mu_{1})-\mathcal{A}(\mu_{2}) is independent of any choice of the 1-form β\beta such that dβ=ωd\beta=\omega.

For the problem on the annulus, to precisely state our results, we will fix the special 1-form β=ydx\beta=ydx, under the standard coordinate. The choice of this 1-form is equivalent to collapsing the lower boundary component into a point, effectively killing the underlying topology.

Our main result is

Theorem 1.

Let ff be an area-preserving, orientation preserving diffeomoprhism on 𝔸\mathbb{A}, isotopic to identity. Let μ1,μ2f\mu_{1},\mu_{2}\in\mathcal{M}_{f} be any two ff-invariant probability measures. Suppose that |𝒜(μ1)𝒜(μ2)|0|\mathcal{A}(\mu_{1})-\mathcal{A}(\mu_{2})|\neq 0, then ff has infinitely many distinct periodic points. More precisely, for any positive integer qq such that

q>1|𝒜(μ1)𝒜(μ2)|,q>\frac{1}{|\mathcal{A}(\mu_{1})-\mathcal{A}(\mu_{2})|},

ff has at least two distinct periodic orbits with period qq, and qq is the least period if it is a prime number. In particular, if |𝒜(μ1)𝒜(μ2)|>1|\mathcal{A}(\mu_{1})-\mathcal{A}(\mu_{2})|>1, then ff has at least two fixed points.

The same result is also true for area-preserving, orientation preserving diffeomorphism of the unit disk 𝔻={(x,y)2,x2+y21}\mathbb{D}=\{(x,y)\in\mathbb{R}^{2},\;x^{2}+y^{2}\leq 1\}.

As an application to our theorem, we consider a rigid rotation on the annulus with an irrational rotation number. The action function in this case is a constant. One can perturb the map, preserving the area, in a neighborhood of an essential simple closed curve to change the mean rotation number, therefore creating new periodic orbits by Franks’ theorem. Our result shows that this can be done locally. Pick any point in the interior of the annulus. For any small neighborhood around the point, we can easily change the map to increase or decrease the mean action of the map, consequently we have a different action with respect to the area, i.e., the Calabi invariant is now different from the actions of untouched orbits. Therefore, there must be infinitely many new periodic points. The details of this example will be given at the last section of this paper.

2. Action and rotation vectors

In this section, we give a general introduction of the action and rotation vectors of symplectic diffeomorphisms.

2.1. Action function

Let M2nM^{2n} be a 2n2n-dimensional symplectic manifold with a non-degenerate closed 2-form ω\omega, and let

f:MMf:M\rightarrow M

be a diffeomorphism preserving the symplectic form ω\omega. We further assume that ff is exact symplectic, i.e., ff is isotopic to identity; there is a 1-form β\beta such that ω=dβ\omega=d\beta, and there is a function gg on MM such that

dg=fββ.dg=f^{*}\beta-\beta.

The real valued function gg is called the action function for ff. The choice of gg depends on two factors, the first one is the integration constant. This can be fixed by assigning gg a particular value at a special point, say x0Mx_{0}\in M. The second factor is the choice of β\beta. Suppose

ω=dβ=dβ~,\omega=d\beta=d\tilde{\beta},

then ββ~\beta-\tilde{\beta} is a closed form. Let

[ββ~]H1(M,).[\beta-\tilde{\beta}]\in H^{1}(M,\mathbb{R}).

be the cohomology class of the difference. It turns out that this cohomology class plays an essential role.

First, let’s suppose that the cohomology class of ββ~\beta-\tilde{\beta} is trivial, then there is a function S:MS:M\rightarrow\mathbb{R}, such that

β~β=dS,\tilde{\beta}-\beta=dS,

then for any function g~\tilde{g} such that

dg~=fβ~β~,d\tilde{g}=f^{*}\tilde{\beta}-\tilde{\beta},

we have

dg~dg=(fβ~fβ)+(β~β)=fdSdS=d(SfS)d\tilde{g}-d{g}=(f^{*}\tilde{\beta}-f^{*}\beta)+(\tilde{\beta}-\beta)=f^{*}dS-dS=d(S\circ f-S)

or

g~g=(SfS)+C\tilde{g}-{g}=(S\circ f-S)+C

for some constant CC. Conversely, for any function SS on MM, let

g~=g+(SfS)+C,\tilde{g}={g}+(S\circ f-S)+C,

then g~\tilde{g} is also an action function for ff, with trivial cohomology class for ββ~\beta-\tilde{\beta} for corresponding 1-forms.

Two real valued functions g~\tilde{g} and gg are said to be cohomologous if there is a real valued function SS on MM such that

g~=g+(SfS).\tilde{g}={g}+(S\circ f-S).

It is important to note that for any ff-invariant measure μ\mu on MM,

M(SfS)𝑑μ=MSd(fμ)Sdμ=0,\int_{M}(S\circ f-S)d\mu=\int_{M}Sd(f^{*}\mu)-Sd\mu=0,

in particular,

M(SfS)ωn=0.\int_{M}(S\circ f-S)\omega^{n}=0.

It follows that, if gg and g~\tilde{g} are cohomologous, then the action defined by these action functions are exactly same.

Next, we define the mean action of ff, the Calabi invariant. If MM has bounded volume, we assume, without loss of generality, rescale ω\omega so that Mωn=1\int_{M}\omega^{n}=1. Then the mean action is simply

𝒜(f)=Mgωn.\mathcal{A}(f)=\int_{M}g\omega^{n}.

We remark that the action, and therefore the mean action, defined above is relative. It denpends on the integration constant CC. If neccessary, we will make proper choices to fix CC.

The action also depends on the first cohomology of MM. This dependence is more interesting and will be explored later in the section. For now, we will fix a 1-form β\beta.

It is a very useful and convenient fact that the mean action is additive for composition of diffeomorphisms.

Proposition 2.

Let f1f_{1} and f2f_{2} be exact symplectic diffeomorphisms of MM. Let g1g_{1}, g2g_{2} and g12g_{12} be action functions for f1f_{1}, f2f_{2} and f2f1f_{2}\circ f_{1} respectively. Suppose there is a point x0Mx_{0}\in M such that

g12(x0)=g1(x0)+g2(f1(x0)).g_{12}(x_{0})=g_{1}(x_{0})+g_{2}(f_{1}(x_{0})).

Then

𝒜(f2f1)=𝒜(f1)+𝒜(f2).\mathcal{A}(f_{2}\circ f_{1})=\mathcal{A}(f_{1})+\mathcal{A}(f_{2}).
Proof.

The action function g12g_{12} is define by

dg12=(f2f1)ββ,dg_{12}=(f_{2}\circ f_{1})^{*}\beta-\beta,

hence,

dg12={f1(f2β)f2β}+{f2ββ}dg_{12}=\left\{f_{1}^{*}(f_{2}^{*}\beta)-f_{2}^{*}\beta\right\}+\{f_{2}^{*}\beta-\beta\}
=dg~1+dg2=d\tilde{g}_{1}+dg_{2}

for some function g~1\tilde{g}_{1}. In fact

dg~1=f1(f2β)f2β={f1(f2ββ)(f2ββ)}+{f1ββ}d\tilde{g}_{1}=f_{1}^{*}(f_{2}^{*}\beta)-f_{2}^{*}\beta=\{f_{1}^{*}(f_{2}^{*}\beta-\beta)-(f_{2}^{*}\beta-\beta)\}+\{f_{1}^{*}\beta-\beta\}
={f1dg2dg2}+dg1=d(g2f1g2)+dg1.=\{f_{1}^{*}dg_{2}-dg_{2}\}+dg_{1}=d(g_{2}\circ f_{1}-g_{2})+dg_{1}.

Obviously, such g~1\tilde{g}_{1} exists, we may choose

g~1=(g2f1g2)+g1,\tilde{g}_{1}=(g_{2}\circ f_{1}-g_{2})+g_{1},

and then we may choose

g12=g~1+g2=(g2f1g2)+g1+g2=g2f1+g1,g_{12}=\tilde{g}_{1}+g_{2}=(g_{2}\circ f_{1}-g_{2})+g_{1}+g_{2}=g_{2}\circ f_{1}+g_{1},

i.e., we may choose the action function for f2f1f_{2}\circ f_{1} by adding the action function for f1f_{1} and the f1f_{1}-shifted action function for f2f_{2}. Under this choice, we have

(1) g12(x0)=g1(x0)+g2(f1(x0)).\displaystyle{g}_{12}(x_{0})=g_{1}(x_{0})+g_{2}(f_{1}(x_{0})).

Now

𝒜(f2f1)\displaystyle\mathcal{A}(f_{2}\circ f_{1}) =\displaystyle= Mg12ωn\displaystyle\int_{M}g_{12}\omega^{n}
=\displaystyle= M(g1+g2f1)ωn\displaystyle\int_{M}(g_{1}+g_{2}\circ f_{1})\omega^{n}
=\displaystyle= Mg1ωn+M(g2f1)ωn\displaystyle\int_{M}g_{1}\omega^{n}+\int_{M}(g_{2}\circ f_{1})\omega^{n}
=\displaystyle= Mg1ωn+Mg2(f1ωn)\displaystyle\int_{M}g_{1}\omega^{n}+\int_{M}g_{2}(f_{1}^{*}\omega^{n})
=\displaystyle= Mg1ωn+Mg2ωn\displaystyle\int_{M}g_{1}\omega^{n}+\int_{M}g_{2}\omega^{n}
=\displaystyle= 𝒜(f1)+𝒜(f2).\displaystyle\mathcal{A}(f_{1})+\mathcal{A}(f_{2}).

We remark that the above equality is independent of the choice of 1-form β\beta and therefore the choices of the action functions, as long as condition (1) holds.

This proves the proposition. ∎

2.2. Action on invariant measures

Let

f:MMf:M\rightarrow M

be an exact symplectic diffeomorphism and let

g:Mg:M\rightarrow\mathbb{R}

be a fixed action function for ff. Let f\mathcal{M}_{f} be the set of all ff-invariant probability measures on MM. For any μf\mu\in\mathcal{M}_{f}, the mean action, or simply the action, of μ\mu is defined to be

𝒜(μ)=Mg𝑑μ.\mathcal{A}(\mu)=\int_{M}gd\mu.

In particular, for any periodic orbit γ={p0,p1,,pk=p0}\gamma=\{p_{0},p_{1},\ldots,p_{k}=p_{0}\}, where fi(p0)=pif^{i}(p_{0})=p_{i}, for i=1,2,,ki=1,2,\ldots,k, the corresponding invariant measure is

μγ=1k(δp0+δp1++δpk1),\mu_{\gamma}=\frac{1}{k}(\delta_{p_{0}}+\delta_{p_{1}}+\ldots+\delta_{p_{k-1}}),

and the mean action on γ\gamma is

𝒜(γ)=𝒜(μγ)=1k(g(p0)+g(p1)++g(pk1)).\mathcal{A}(\gamma)=\mathcal{A}(\mu_{\gamma})=\frac{1}{k}(g(p_{0})+g(p_{1})+\ldots+g(p_{k-1})).

The action we defined so far is for exact symplectic diffeomorphisms. However, it is a well-known fact that if MM is compact, then there is no 1-form β\beta such that dβ=ωd\beta=\omega, hence no symplectic diffeomorphism on a compact manifold is exact. A class of symplectic diffeomorphisms that shares many properties with exact symplectic diffeomorphisms is the Hamiltonian diffeomorphisms. These are the time-1 maps of periodic Hamiltonian flows. On compact surfaces, one can blow up a contractible fixed point, whose existence is guarantied by Arnold’s conjecture, and then consider the exact area-preserving map on a compact surface with boundary.

On a compact symplectic manifold, we can also take another approach. Since action on periodic orbits, and in extension, invariant measures, is relative, we are more interested in the differences between two invariant measures.

Let p1p_{1} and p2p_{2} be two fixed points for a symplectic diffeomorphism f:MMf:M\rightarrow M. We assume that ff is isotopic to identity, but not necessarily exact. We say that p1p_{1} and p2p_{2} are homologous, if the loops γ~1\tilde{\gamma}_{1} and γ~2\tilde{\gamma}_{2}, starting with p1p_{1} and p2p_{2} respectively, following the isotopy of ff back to p1p_{1} and p2p_{2} respectively, are homologous, i.e.,

[γ~1]=[γ~2]H1(M,)[\tilde{\gamma}_{1}]=[\tilde{\gamma}_{2}]\in H_{1}(M,\mathbb{R})

Let γ\gamma be a curve on MM connecting p1p_{1} and p2p_{2}, then it is easy to see that γ\gamma and f(γ)f(\gamma) are homologous, there is a disk DD on MM such that

D=f(γ)γ.\partial D=f(\gamma)-\gamma.

Finally, we define the difference of the actions between p1p_{1} and p2p_{2} to be

𝒜(p2)𝒜(p1)=Dω.\mathcal{A}(p_{2})-\mathcal{A}(p_{1})=\int_{D}\omega.

We remark that if ff is exact symplectic or Hamiltonian, then the above definition is independent of the choice of γ\gamma. However, if ff isotopic to identity, but not Hamiltonian, then it does depend on the homotopy class of the curve γ\gamma.

The action difference can be extended to periodic points easily. If γ1\gamma_{1} and γ2\gamma_{2} are periodic orbits with common period kk. Assume that γ1\gamma_{1} and γ2\gamma_{2} are homologous. Let γ\gamma be a curve from one point in the orbit of γ1\gamma_{1} to a point in the orbit of γ2\gamma_{2} and let DD be a disk DD on MM such that

D=fk(γ)γ,\partial D=f^{k}(\gamma)-\gamma,

then

𝒜(γ2)𝒜(γ1)=1kDω.\mathcal{A}(\gamma_{2})-\mathcal{A}(\gamma_{1})=\frac{1}{k}\int_{D}\omega.

Again, this can be extended to two invariant measures μ1\mu_{1} and μ2\mu_{2}. For this, we need to define the homology class of invariant measures. This turns out to be exactly the rotation vectors, to be defined next. Two invariant measures are said to be homologous if they have the same rotation vectors.

2.3. Rotation vectors

Let f:MMf:M\rightarrow M be a symplectic diffeomorphism, isotopic to identity. Let γ={p0,p1,,pk=p0}\gamma=\{p_{0},p_{1},\ldots,p_{k}=p_{0}\}, be a periodic orbit, where fi(p0)=pif^{i}(p_{0})=p_{i}, for i=1,2,,ki=1,2,\ldots,k. Let γ~(t)\tilde{\gamma}(t) be a closed curve on MM obtained by isotopy, with γ~(i)=pi\tilde{\gamma}(i)=p_{i}, for i=1,2,,ki=1,2,\ldots,k. We define the rotation vector of γ\gamma to be the homology class of γ~\tilde{\gamma} divided by its period,

ρ(γ)=1k[γ~]H1(M,).\rho(\gamma)=\frac{1}{k}[\tilde{\gamma}]\in H_{1}(M,\mathbb{R}).

To generalize the concept of rotation vector, for any closed 1-form α\alpha on MM, we define the bi-linear form <,><\cdot,\,\cdot>^{*} by

<γ,α>=1kγ~α.<\gamma,\,\alpha>^{*}=\frac{1}{k}\oint_{\tilde{\gamma}}\alpha.

It is easy to see that above pairing depends only on the homology class of γ~\tilde{\gamma}, in H1(M,)H_{1}(M,\mathbb{R}), and the cohomology class of α\alpha, in H1(M,)H^{1}(M,\mathbb{R}). This paring equivalently defines, for any periodic orbit γ\gamma, the rotation vector ρ(γ)H1(M,)\rho(\gamma)\in H_{1}(M,\mathbb{R}), by the following equation:

<ρ(γ),[α]>=<γ,α>=1kγ~α<\rho(\gamma),[\alpha]>=<\gamma,\alpha>^{*}=\frac{1}{k}\oint_{\tilde{\gamma}}\alpha

where the first pairing is the canonical pairing between homology and cohomology of the manifold MM.

This definition of rotation vector can be naturally extended to ff-invariant measures in f\mathcal{M}_{f}. Fix a closed 1-form α\alpha, for any point xMx\in M, let γ~(t,x),t\tilde{\gamma}(t,x),t\in\mathbb{R} be the curve in MM by connecting orbit of xx by the isotopy in such a way that γ~(i,x)=fi(x)\tilde{\gamma}(i,x)=f^{i}(x). Define, if exists,

ρ(x,α)=limT1Tγ~(t,x):t[0,T]α.\rho(x,\alpha)=\lim_{T\rightarrow\infty}\frac{1}{T}\int_{\tilde{\gamma}(t,x):t\in[0,T]}\alpha.

It is easy to see that ρ(x,α)=ρ(x,α)\rho(x,\alpha)=\rho(x,\alpha^{\prime}), if [α]=[α]H1(M,)[\alpha]=[\alpha^{\prime}]\in H^{1}(M,\mathbb{R}). This is because that if α\alpha is exact, α=dS\alpha=dS for some function S:MS:M\rightarrow\mathbb{R}, then

ρ(x,dS)=limT1Tγ(t,x):t[0,T]𝑑S=limT1T(S(γ(T))S(x))=0.\rho(x,dS)=\lim_{T\rightarrow\infty}\frac{1}{T}\int_{\gamma(t,x):t\in[0,T]}dS=\lim_{T\rightarrow\infty}\frac{1}{T}(S(\gamma(T))-S(x))=0.

For any invariant measure μf\mu\in\mathcal{M}_{f}, by Birkhoff Ergodic Theorem, for any fixed α\alpha, for μa.e.xM\mu-a.e.\,x\in M, the limit exists, ρ(x,α)\rho(x,\alpha) is well-defined. Therefore, it is also well-defined for a finite set of basis vectors [α][\alpha] in H1(M,)H^{1}(M,\mathbb{R}). Hence, for μa.e.xM\mu-a.e.\,x\in M, ρ(x,α)\rho(x,\alpha) is well-defined for all closed 1-forms α\alpha. Moreover, by Birkhoff Ergodic Theorem,

ρ(x,α)𝑑μ=(γ(t,x):t[0,1]α)𝑑μ\int\rho(x,\alpha)d\mu=\int\left(\int_{\gamma(t,x):t\in[0,1]}\alpha\right)d\mu

The above equation is linear in both α\alpha and μ\mu, it depends only on the cohomology class of α\alpha, therefore, it define a pairing between μ\mu and cohomology elements in H1(M,)H^{1}(M,\mathbb{R}). This pairing defines the rotation vector ρ(μ)H1(M,)\rho(\mu)\in H_{1}(M,\mathbb{R}).

Definition 3.

For any μMf\mu\in M_{f} and any closed 1-form α\alpha, [α]H1(M,)[\alpha]\in H^{1}(M,\mathbb{R}), the rotation vector of μ\mu, ρ(μ)H1(M,)\rho(\mu)\in H_{1}(M,\mathbb{R}) is defined by the following equation

<ρ(μ),[α]>=(γ(t,x):t[0,1]α)𝑑μ<\rho(\mu),[\alpha]>=\int\left(\int_{\gamma(t,x):t\in[0,1]}\alpha\right)d\mu\in\mathbb{R}

where the left hand side is the canonical pairing between homology and cohomology of the manifold MM.

As an example, if ff is a Hamiltonian diffeomorphism, then ωnf\omega^{n}\in\mathcal{M}_{f} and if MM is compact without boundary, then

ρ(ωn)=0.\rho(\omega^{n})=0.

This is not true in general for non-Hamiltonian symplectic diffeomorphisms. An easy example is T(a,b):𝕋2𝕋2T_{(a,b)}:\mathbb{T}^{2}\rightarrow\mathbb{T}^{2} given by

T(a,b)(x,y)=(x+a,y+b) mod 2,T_{(a,b)}(x,\,y)=(x+a,\,y+b)\mbox{ mod }\mathbb{Z}^{2},

for (a,b)2(a,b)\notin\mathbb{Z}^{2}.

This is also not true in general for Hamiltonian diffeomorphisms on manifold with boundaries, for example the annulus, 𝔸\mathbb{A}.

We now return to actions on invariant measures. As we have hinted before, the action we defined depends on the choice of 1-form β\beta. It turns out that this dependence, interestingly, is closely related to the rotation vector we just defined. Let α\alpha be a closed 1-form on MM, let

β~=β+α,\tilde{\beta}=\beta+\alpha,

then dβ~=dβ=ωd\tilde{\beta}=d\beta=\omega. Let g~\tilde{g} and gg be action functions defined by β~\tilde{\beta} and β\beta respectively. To fix the integration constants, pick a point x0Mx_{0}\in M, let

g(x)=x0x(f(β)β){g}(x)=\int_{x_{0}}^{x}(f^{*}({\beta})-\beta)

and likewise

g~(x)=x0x(f(β~)β)~=g(x)+x0x(fαα)\tilde{g}(x)=\int_{x_{0}}^{x}(f^{*}(\tilde{\beta})-\tilde{\beta)}=g(x)+\int_{x_{0}}^{x}(f^{*}\alpha-\alpha)

All the integrals are path independent. For any invariant measure μ\mu, its action under g~\tilde{g},

𝒜~(μ)\displaystyle\widetilde{\mathcal{A}}(\mu) =\displaystyle= Mg~𝑑μ=Mg𝑑μ+M(x0x(fαα))𝑑μ\displaystyle\int_{M}\tilde{g}d\mu=\int_{M}gd\mu+\int_{M}\left(\int_{x_{0}}^{x}(f^{*}\alpha-\alpha)\right)d\mu
=\displaystyle= 𝒜(μ)+M(γ(t,x):t[0,1]α)𝑑μM(γ(t,x0):t[0,1]α)𝑑μ\displaystyle\mathcal{A}(\mu)+\int_{M}\left(\int_{\gamma(t,x):t\in[0,1]}\alpha\right)d\mu-\int_{M}\left(\int_{\gamma(t,x_{0}):t\in[0,1]}\alpha\right)d\mu
=\displaystyle= 𝒜(μ)+<ρ(μ),α>Cx0,α,\displaystyle\mathcal{A}(\mu)+<\rho(\mu),\alpha>-C_{x_{0},\alpha},

where Cx0,αC_{x_{0},\alpha} is a constant depending on x0x_{0} and α\alpha, but not on μ\mu. Here we used the Stokes’ theorem and Birkhoff Ergodic Theorem.

The constant Cx0,αC_{x_{0},\alpha} is zero, if our refence point x0x_{0} is a contractible fixed point.

Since the action is typically used in the relative sense, for comparison between two different invariant measures, we have the following proposition.

Proposition 4.

Let μ1,μ2f\mu_{1},\mu_{2}\in\mathcal{M}_{f} be two invariant probability measures and let 𝒜~\widetilde{\mathcal{A}} and 𝒜\mathcal{A} be two actions defined by 1-forms β~\tilde{\beta} and β\beta respectively. Let α=β~β\alpha=\tilde{\beta}-\beta, then

𝒜~(μ1)𝒜~(μ2)=𝒜(μ1)𝒜(μ2)+<ρ(μ1)ρ(μ2),[α]>\widetilde{\mathcal{A}}(\mu_{1})-\widetilde{\mathcal{A}}(\mu_{2})=\mathcal{A}(\mu_{1})-\mathcal{A}(\mu_{2})+<\rho(\mu_{1})-\rho(\mu_{2}),\,[\alpha]>

In particular, if two invariant measures μ1\mu_{1} and μ2\mu_{2} have exactly the same rotation vector, then the action difference 𝒜(μ1)𝒜(μ2)\mathcal{A}(\mu_{1})-\mathcal{A}(\mu_{2}) is independent of any choice of 1-form β\beta and the integration constant.

2.4. Actions on the annulus

We now restrict ourselves to two specific spaces, disk 𝔻\mathbb{D} and annulus 𝔸\mathbb{A}. First let ff be an orientation-preserving, area-preserving diffeomorphism of a unit disk 𝔻\mathbb{D} in 2\mathbb{R}^{2}. Clearly, ff is exact symplectic. The first homology of 𝔻\mathbb{D} is trivial, so there is no ambiguity in the the action function. The action, particularly the difference in actions of two invariant measures, has a very clear geometric meaning. Take for example the simple case of two fixed points. Let p1p_{1} and p2p_{2} be two points in 𝔻\mathbb{D} fixed by ff. Pick any curve γ\gamma connecting p1p_{1} to p2p_{2}. There is a unique signed disk U𝔻U\subset\mathbb{D} such that

U=f(γ)γ.\partial U=f(\gamma)-\gamma.

Then simply, we have

𝒜(p2)𝒜(p1)=Uω𝔻,\mathcal{A}(p_{2})-\mathcal{A}(p_{1})=\int_{U}\omega_{\mathbb{D}},

where ω𝔻\omega_{\mathbb{D}} is the nomalized standard area form on 𝔻\mathbb{D}.

If p1p_{1} and p2p_{2} are periodic points of period kk, then we consider fkf^{k}, then the difference in action with respect to ff is that of fkf^{k} divided by kk. As for two invariant ergodic measures μ1\mu_{1} and μ2\mu_{2}, we can take generic points for these measures, approximate them by periodic points, then take limits.

The situation for the annulus is a little more complicated. Let ff be an area-preserving, orientation preserving diffeomoprhism on 𝔸=S1×[0,1]\mathbb{A}=S^{1}\times[0,1], isotopic to identity. It is easy to see that ff is exact symplectic. Suppose p1,p2𝔸p_{1},p_{2}\in\mathbb{A} are two fixed points of ff, and let γ\gamma be a curve on 𝔸\mathbb{A} connecting p1p_{1} and p2p_{2}. If p1p_{1} and p2p_{2} have exactly the same rotation number, then their orbits are homologous, this will be the same case as for the disks, their action difference in action is

𝒜(p2)𝒜(p1)=Dω,\mathcal{A}(p_{2})-\mathcal{A}(p_{1})=\int_{D}\omega,

where DD is a disk on 𝔸\mathbb{A} such that D=f(γ)γ\partial D=f(\gamma)-\gamma. This action difference is independent of the choice of β\beta in the definition of action function.

If p1p_{1} and p2p_{2} are not homologous (i.e. when they have different rotation numbers), now f(γ)f(\gamma) and γ\gamma does not bound an area, there is no natural way to define the region DD as in the previous case. By Proposition 4, the choice of 1-form β\beta makes difference in the action. To remove the ambiguities, we can collapse the boundary A0A_{0} to a point, i.e. consider the quotient space 𝔸/\mathbb{A}/\sim, where the equivalence class is defined by (x1,0)(x2,0)(x_{1},0)\sim(x_{2},0) for all xiS1x_{i}\in S^{1}. The resulting space is the closed unit disk 𝔻\mathbb{D}, and we still use ff to denote the corresponding map on 𝔻\mathbb{D} and ω𝔻\omega_{\mathbb{D}} the corresponding area form. For any fixed points p1,p2p_{1},p_{2} of f:𝔻𝔻f:\mathbb{D}\to\mathbb{D}, the curves f(γ)f(\gamma) and γ\gamma always bound a region UU, i.e. U=f(γ)γ\partial U=f(\gamma)-\gamma, thus the above defined action difference on 𝔸\mathbb{A} is the same as Uω𝔻\int_{U}\omega_{\mathbb{D}}, independent of whether p1p_{1} and p2p_{2} have the same rotation number.

Equivalently, if we add the boundary component A0=S1×{0}A_{0}=S^{1}\times\{0\} to γ\gamma and f(γ)f(\gamma) in 𝔸\mathbb{A}, then together they always bound a region, i.e. there is a signed region UU such that U=f(γ)γ+kA0\partial U=f(\gamma)-\gamma+kA_{0}, for some integer kk depending on the rotation numbers of p1p_{1} and p2p_{2}. We can define

𝒜(p2)𝒜(p1)=Uω.\mathcal{A}(p_{2})-\mathcal{A}(p_{1})=\int_{U}\omega.

The above geometric construction can be achieved by making proper choice of the primitive of ω\omega on the annulus. Let (x,y)(x,y) be the natural coordinate system on 𝔸=S1×[0,1]\mathbb{A}=S^{1}\times[0,1], the standard area form is ω=dydx\omega=dy\wedge dx. We choose and fix β=ydx\beta=ydx. This 1-form is exactly the pullback of the 1-form on the disk 𝔻\mathbb{D} with respect to collapsing of the bounday component A0A_{0}.

By fixing the 1-form, β=ydx\beta=ydx, we have fixed the action function gg up to a constant. For any diffeomorphism f:𝔸𝔸f:\mathbb{A}\to\mathbb{A} and invariant measures μ1\mu_{1} and μ2\mu_{2} in f\mathcal{M}_{f}, the action difference of any two invariant measures

𝒜(μ2)𝒜(μ1)=𝔸g𝑑μ2𝔸g𝑑μ1\mathcal{A}(\mu_{2})-\mathcal{A}(\mu_{1})=\int_{\mathbb{A}}gd\mu_{2}-\int_{\mathbb{A}}gd\mu_{1}

is well-defined.

3. Proof of the main theorem

We first prove some special cases of Theorem 1. Notice that ff restricted to each of the boundary 𝔸0=S1×{0}\mathbb{A}_{0}=S^{1}\times\{0\} and 𝔸1=S1×{1}\mathbb{A}_{1}=S^{1}\times\{1\} is an orientation preserving circle diffeomorphism. Let ρ0\rho_{0} and ρ1\rho_{1} be the rotation numbers on the boundaries respectively. If ρi=pq\rho_{i}=\frac{p}{q} (i=0,1)(i=0,1) is rational, then there is a periodic point with period qq, in particular, it supports an atomic invariant measure; otherwise, there is an invariant measure supported on the boundary.

Lemma 5.

Let ff be an area-preserving, orientation preserving diffeomoprhism on 𝔸\mathbb{A}, isotopic to identity. Let μ0,μ1f\mu_{0},\mu_{1}\in\mathcal{M}_{f} be invariant measures with supports in 𝔸0\mathbb{A}_{0} and 𝔸1\mathbb{A}_{1} respectively, and |𝒜(μ0)𝒜(μ1)|0|\mathcal{A}(\mu_{0})-\mathcal{A}(\mu_{1})|\neq 0. Then there exists an interval with length |𝒜(μ0)𝒜(μ1)||\mathcal{A}(\mu_{0})-\mathcal{A}(\mu_{1})| such that for any rational number pq\frac{p}{q} in the interval, there are at least two distinct periodic orbits of period qq with rotation number pq\frac{p}{q}.

Proof.

Let 𝔸~=×[0,1]\tilde{\mathbb{A}}=\mathbb{R}\times[0,1] be the standard universal cover of 𝔸\mathbb{A} and f~:𝔸~𝔸~\tilde{f}:\tilde{\mathbb{A}}\to\tilde{\mathbb{A}} a lift of ff. Let p0,p1p_{0},p_{1} be any two points on the boundaries 𝔸~0,𝔸~1\tilde{\mathbb{A}}_{0},\tilde{\mathbb{A}}_{1}, respectively, and let γ\gamma be a simple curve connecting p0p_{0} and p1p_{1}. Let UU be the region bounded by γ\gamma, f~(γ)\tilde{f}(\gamma), hp0h_{p_{0}} and hp1h_{p_{1}}, where hpih_{p_{i}} is the segment from pip_{i} to f~(pi)\tilde{f}(p_{i}) on the boundaries 𝔸~i\tilde{\mathbb{A}}_{i}, (i=0,1)(i=0,1) respectively. Therefore we have

U=γf~(γ)+hp1hp0.\partial U=\gamma-\tilde{f}(\gamma)+h_{p_{1}}-h_{p_{0}}.

Using the Stoke’s theorem and the definition of the action function we have

Uω=g(p0)g(p1)+hp1βhp0β.\int_{U}\omega=g(p_{0})-g(p_{1})+\int_{h_{p_{1}}}\beta-\int_{h_{p_{0}}}\beta.

Since ff is area-preserving, Uω\int_{U}\omega is independent of the choice of p0,p1p_{0},p_{1} and the curve γ\gamma, in particular, Uω=ρ(ω)\int_{U}\omega=\rho(\omega), the mean rotation number of f~\tilde{f} for the invariant measure induced by ω=dydx\omega=dy\wedge dx. Since β=ydx\beta=ydx, we have

ρ(ω)=Uω=𝒜(μ0)𝒜(μ1)+ρ1.\rho(\omega)=\int_{U}\omega=\mathcal{A}(\mu_{0})-\mathcal{A}(\mu_{1})+\rho_{1}.

Therefore, the rotation set of f~:𝔸~𝔸~\tilde{f}:\tilde{\mathbb{A}}\to\tilde{\mathbb{A}} contains a closed interval with length |𝒜(μ0)𝒜(μ1)||\mathcal{A}(\mu_{0})-\mathcal{A}(\mu_{1})|. By Franks’ theorem [3], for any rational number pq\frac{p}{q} in the interval, there are at least two distinct periodic orbits of period qq with rotation number pq\frac{p}{q}. In particular, for any positive integer q>1|𝒜(μ0)𝒜(μ1)|q>\frac{1}{|\mathcal{A}(\mu_{0})-\mathcal{A}(\mu_{1})|}, there is a rational number with denominator qq contained in the interval, thus there must be at least two distinct periodic orbits with period qq. Moreover, if qq is prime, then it is the least period. ∎

Now, let’s consider the case with one of the invariant measure supported on the boundary and the other is atomic at a fixed point.

Lemma 6.

Let ff be an area-preserving, orientation preserving diffeomoprhism on 𝔸\mathbb{A}, isotopic to identity. Let μ0f\mu_{0}\in\mathcal{M}_{f} be an invariant measure with support in 𝔸0\mathbb{A}_{0}, and pp is a fixed point of ff such that |𝒜(μ0)𝒜(p)|0|\mathcal{A}(\mu_{0})-\mathcal{A}(p)|\neq 0. Then for any positive integer qq such that

q>1|𝒜(μ0)𝒜(p)|,q>\frac{1}{|\mathcal{A}(\mu_{0})-\mathcal{A}(p)|},

ff has at least two distinct periodic orbits with period qq, and qq is the least period if it is a prime number.

Proof.

If pp is on 𝔸1\mathbb{A}_{1}, then it reduces to Lemma 5, we assume p𝔸1p\notin\mathbb{A}_{1}.

Let p0p_{0} be a point in the support of μ0\mu_{0}, and let γ\gamma be a simple curve connecting p0p_{0} and pp, which does not touch 𝔸1\mathbb{A}_{1}. Let UU be the region bounded by γ\gamma, f(γ)f(\gamma), hp0h_{p_{0}}, where hp0h_{p_{0}} is the segment from p0p_{0} to f(p0)f(p_{0}) on the boundary 𝔸0\mathbb{A}_{0}. Using the Stoke’s theorem and the definition of the action function we have

Uω=g(p0)g(p)hp0β=g(p0)g(p).\int_{U}\omega=g(p_{0})-g(p)-\int_{h_{p_{0}}}\beta=g(p_{0})-g(p).

Since ff is area-preserving, we have

Uω=𝒜(μ0)𝒜(p).\int_{U}\omega=\mathcal{A}(\mu_{0})-\mathcal{A}(p).

Collapse the boundary 𝔸1\mathbb{A}_{1} to a point and denote it by the point A1A_{1} in the resulting disk. Since on this disk,

A10t1ht(γ),A_{1}\notin\bigcup_{0\leq t\leq 1}h_{t}(\gamma),

where ht,t[0,1]h_{t},t\in[0,1] denotes the isotopy of ff, we can blow up pp relative to p1p_{1}. Remove the point pp and add a boundary circle CC. Denote the resulting annulus by 𝔸\mathbb{A}^{\prime} and the corresponding map by f:𝔸𝔸f^{\prime}:\mathbb{A}^{\prime}\to\mathbb{A}^{\prime}. Note that A1A_{1} is a contractible fixed point for ff^{\prime}, hence its rotation number is zero.

In this process the transformations are area-preserving. If γ\gamma^{\prime} is a curve connecting the boundaries of 𝔸\mathbb{A}^{\prime}, the area between γ\gamma^{\prime}, f(γ)f^{\prime}(\gamma^{\prime}) and the two boundaries of 𝔸\mathbb{A}^{\prime} will be equal to Uω=𝒜(μ0)𝒜(p)\int_{U}\omega=\mathcal{A}(\mu_{0})-\mathcal{A}(p) of the original map ff. Thus the mean rotation number of ff^{\prime} is 𝒜(μ0)𝒜(p)\mathcal{A}(\mu_{0})-\mathcal{A}(p). Since the rotation number of the fixed point A1A_{1} is zero, we get that the rotation set of ff^{\prime} contains an interval of length |𝒜(μ0)𝒜(μ1)||\mathcal{A}(\mu_{0})-\mathcal{A}(\mu_{1})|. Therefore the conclusion follows similar to Lemma 5.

Remark 7.

The above result also holds if we change the assumption of μ0\mu_{0} to μ1f\mu_{1}\in\mathcal{M}_{f} whose support is in 𝔸1\mathbb{A}_{1}. In this case we have

Uω=𝒜(p)𝒜(μ1)+ρ(μ1).\int_{U}\omega=\mathcal{A}(p)-\mathcal{A}(\mu_{1})+\rho(\mu_{1}).

Here UU is the region such that U=γf(γ)+hp1\partial U=\gamma-f(\gamma)+h_{p_{1}}, where γ\gamma is a simple curve from pp to a point p1p_{1} in the support of μ1\mu_{1}. In this case, we collapse the boundary 𝔸0\mathbb{A}_{0} and blow up pp. In the resulting new annulus 𝔸\mathbb{A}^{\prime}, its mean rotation number is Uω=𝒜(p)𝒜(μ1)+ρ(μ1)\int_{U}\omega=\mathcal{A}(p)-\mathcal{A}(\mu_{1})+\rho(\mu_{1}), and the boundary 𝔸1\mathbb{A}_{1}^{\prime} has the same rotation number ρ(μ1)\rho(\mu_{1}) as 𝔸1\mathbb{A}_{1} in the old annulus. Thus the rotation set contains a closed interval with length |𝒜(μ0)𝒜(μ1)||\mathcal{A}(\mu_{0})-\mathcal{A}(\mu_{1})|.

Next, let’s consider the case where both measures are supported on periodic orbits.

Lemma 8.

Let ff be an area-preserving, orientation preserving diffeomoprhism on 𝔸\mathbb{A}, isotopic to identity. Suppose that γ1\gamma_{1} and γ2\gamma_{2} are periodic orbits with common period kk. If |𝒜(γ1)𝒜(γ2)|0|\mathcal{A}(\gamma_{1})-\mathcal{A}(\gamma_{2})|\neq 0, then for any positive integer qq such that

q>1|𝒜(γ1)𝒜(γ2)|,q>\frac{1}{|\mathcal{A}(\gamma_{1})-\mathcal{A}(\gamma_{2})|},

ff has at least two distinct periodic orbits with period qq, and qq is the least period if it is a prime number.

Proof.

Consider the map fkf^{k}, thus fkf^{k} has two fixed points p1,p2p_{1},p_{2}. We have

𝒜k(p1)𝒜k(p2)=k(𝒜(γ1)𝒜(γ2)),\mathcal{A}_{k}(p_{1})-\mathcal{A}_{k}(p_{2})=k(\mathcal{A}(\gamma_{1})-\mathcal{A}(\gamma_{2})),

where 𝒜k\mathcal{A}_{k} represents the action of fkf^{k}. Let F=fkF=f^{k}, it suffices to show that the rotation set of F:𝔸𝔸F:\mathbb{A}\to\mathbb{A} contains a closed interval with length k|𝒜(γ1)𝒜(γ2)|k|\mathcal{A}(\gamma_{1})-\mathcal{A}(\gamma_{2})|.

We assume p1,p2p_{1},p_{2} are not on the boundaries, otherwise it will reduce to the above lemmas. Collapse 𝔸0\mathbb{A}_{0} to a point so that we get a closed disk, next blow up p2p_{2} and add a circle CC to get an annulus 𝔸\mathbb{A}^{\prime}. Use the standard coordinates (x,y)S1×[0,1](x,y)\in S^{1}\times[0,1] for 𝔸\mathbb{A}^{\prime} such that the circle C=S1×{0}C=S^{1}\times\{0\}. Also let β=ydx\beta=ydx be the primitive for 𝔸\mathbb{A}^{\prime}, then the action difference between p1p_{1} and the boundary CC for F:𝔸𝔸F^{\prime}:\mathbb{A}^{\prime}\to\mathbb{A}^{\prime} is equal to the original action difference of p1,p2p_{1},p_{2} for F:𝔸𝔸F:\mathbb{A}\to\mathbb{A}. Apply the proof of Lemma 6 to F:𝔸𝔸F^{\prime}:\mathbb{A}^{\prime}\to\mathbb{A}^{\prime} completes the proof of Lemma 8.

Finally, let’s prove the general case where μ1,μ2\mu_{1},\mu_{2} are any two invariant measures in f\mathcal{M}_{f} with non-zero action difference.

Proof of Theorem 1.

By Ergodic decomposition theorem (cf. [5]), it suffices to prove the theorem by assuming μ1\mu_{1}, μ2\mu_{2} are ergodic invariant measures. We first assume both of the measures are non-atomic. For each fixed positive integer qq such that q>1|𝒜(μ1)𝒜(μ2)|q>\frac{1}{|\mathcal{A}(\mu_{1})-\mathcal{A}(\mu_{2})|}, choose ϵ=ϵ(q)>0\epsilon=\epsilon(q)>0 so that

q>1|𝒜(μ1)𝒜(μ2)|ϵ.q>\frac{1}{|\mathcal{A}(\mu_{1})-\mathcal{A}(\mu_{2})|-\epsilon}.

Let pi𝔸p_{i}\in\mathbb{A} be regular points of μi\mu_{i}, i=1,2i=1,2 respectively, then

limn1nj=0n1g(fj(pi))=𝔸g𝑑μi=𝒜(μi),i=1,2.\lim_{n\to\infty}\frac{1}{n}\sum_{j=0}^{n-1}g(f^{j}(p_{i}))=\int_{\mathbb{A}}gd\mu_{i}=\mathcal{A}(\mu_{i}),\quad i=1,2.

Notice that ff is area-preserving, thus the regular points pip_{i} are recurrent. Since both measures are non-atomic, there are small neighborhoods UiU_{i} of pip_{i} and sufficiently large kk\in\mathbb{N} such that fk(pi)Uif^{k}(p_{i})\in U_{i} and fj(Ui)Ui=f^{j}(U_{i})\cap U_{i}=\emptyset for all j=1,,k1j=1,\cdots,k-1. We will assume kk is large enough so that

k>1|𝒜(μ1)𝒜(μ2)|+10qk>\frac{1}{|\mathcal{A}(\mu_{1})-\mathcal{A}(\mu_{2})|}+10q

and

|1nj=0n1g(fj(pi))𝒜(μi)|<ε5,nk.|\frac{1}{n}\sum_{j=0}^{n-1}g(f^{j}(p_{i}))-\mathcal{A}(\mu_{i})|<\frac{\varepsilon}{5},\quad\forall n\geq k.

Let ViUiV_{i}\subset U_{i} be a disk containing pip_{i} and fk(pi)f^{k}(p_{i}) and we choose an isotopy ht:𝔸𝔸h_{t}:\mathbb{A}\to\mathbb{A} such that

  • (1)

    h0=id:𝔸𝔸h_{0}=\text{id}:\mathbb{A}\to\mathbb{A},

  • (2)

    ht(z)=zh_{t}(z)=z for all zz in 𝔸(V1V2)\mathbb{A}\setminus(V_{1}\cup V_{2}),

  • (3)

    h1(fk(pi))=pih_{1}(f^{k}(p_{i}))=p_{i}, i=1,2i=1,2.

Let f1=h1ff_{1}=h_{1}\circ f and notice that f1k(pi)=pif_{1}^{k}(p_{i})=p_{i}, and f1j(pi)=fj(pi)f_{1}^{j}(p_{i})=f^{j}(p_{i}) for 1jk11\leq j\leq k-1. Let g1g_{1} be the action function for f1f_{1} and 𝒜1\mathcal{A}_{1} the corresponding action acts on invariant measures. If we choose UiU_{i} sufficiently small we can make

|𝒜1(μpi)𝒜(μi)|<ε2,i=1,2,|\mathcal{A}_{1}(\mu_{p_{i}})-\mathcal{A}(\mu_{i})|<\frac{\varepsilon}{2},\quad i=1,2,

where μpi\mu_{p_{i}} is the measure supported on {pi,f(pi),,fk1(pi)}\{p_{i},f(p_{i}),\cdots,f^{k-1}(p_{i})\} and each point has measure 1k\frac{1}{k}. Notice that each μpi\mu_{p_{i}} is an invariant measure for f1f_{1} since pip_{i} are kk-periodic for f1f_{1}. Now, consider the map f1:𝔸𝔸f_{1}:\mathbb{A}\to\mathbb{A}, which is isotopic to the identity and preserves orientation. Moreover, f1f_{1} has two periodic points p1,p2p_{1},p_{2} with period kk and

|𝒜1(μp1)𝒜1(μp2)||𝒜(μ1)𝒜(μ2)|ε.|\mathcal{A}_{1}(\mu_{p_{1}})-\mathcal{A}_{1}(\mu_{p_{2}})|\geq|\mathcal{A}(\mu_{1})-\mathcal{A}(\mu_{2})|-\varepsilon.

Then by Lemma 8, f1f_{1} has at least two periodic points with period q~\tilde{q} whenever

q~>1|𝒜(μ1)𝒜(μ2)|ε.\tilde{q}>\frac{1}{|\mathcal{A}(\mu_{1})-\mathcal{A}(\mu_{2})|-\varepsilon}.

Notice that k>1|𝒜(μ1)𝒜(μ2)|+10qk>\frac{1}{|\mathcal{A}(\mu_{1})-\mathcal{A}(\mu_{2})|}+10q. By the construction of the perturbation h1h_{1} whose support is in U1U2U_{1}\cup U_{2}, we know that any new periodic points that may be created by the perturbation must have periods no less than kk. Thus any periodic points of f1f_{1} with much smaller periods than kk must be periodic points of the original map ff in the beginning, hence ff has at least two periodic points with period qq and proving the theorem.

If one of the measure is non-atomic and the other is atomic, we only need to do perturbation for the non-atomic measure and then the proof is similar. If both measures are atomic, this reduces to the case of Lemma 8. ∎

4. Application and examples

In this section, as application of our main theorem, we give some examples.

Example 1 (Local perturbation and periodic points).

Let f:𝔸𝔸f:\mathbb{A}\to\mathbb{A} be a rigid irrational rotation, i.e.

f(x,y)=(x+a,y),a,f(x,y)=(x+a,y),\quad a\in\mathbb{R}\setminus\mathbb{Q},

then ff has no periodic points. In this case, fββ=0f^{*}\beta-\beta=0, thus the action function is a constant. Without loss of generality, we assume it’s the zero function, then the action of any μf\mu\in\mathcal{M}_{f} is zero.

Consider a local perturbation, i.e. an area-preserving, orientation preserving diffeomoprhism h:𝔸𝔸h:\mathbb{A}\to\mathbb{A} isotopic to identity, such that hh supports on a small open set U𝔸U\subset\mathbb{A}. Choose hh properly (cf. Example 2), we can make the mean action (i.e. the Calabi invariant) of hh strictly positive, thus by Proposition 2, the mean action of the perturbed map f:=fhf^{\prime}:=f\circ h satisfies

𝒜(f)=𝒜(f)+𝒜(h)=𝒜(h)0.\mathcal{A}(f^{\prime})=\mathcal{A}(f)+\mathcal{A}(h)=\mathcal{A}(h)\neq 0.

Notice that outside UU, hh is the identity map, thus the action of the boundary of ff^{\prime} is still zero. By Theorem 1, for any positive integer qq such that q>1|𝒜(h)|q>\frac{1}{|\mathcal{A}(h)|}, ff^{\prime} has at least two distinct periodic orbits with period qq.

In general, if the local perturbation changes the action, one can create periodic points, which could be very useful in many situations.

Example 2 (Mean action of a monotone twist map on the disk).

Let 𝔻\mathbb{D} be the unit disk with standard area form ω𝔻=1πrdrdθ\omega_{\mathbb{D}}=\frac{1}{\pi}rdrd\theta and let β𝔻=12πr2dθ\beta_{\mathbb{D}}=\frac{1}{2\pi}r^{2}d\theta. Let ϕ:[0,1]\phi:[0,1]\to\mathbb{R} be a smooth function such that

ϕ(r)0,ϕ(0)>0,ϕ(1)=0,\phi^{\prime}(r)\leq 0,\quad\phi(0)>0,\quad\phi(1)=0,

and ϕ(r)=0\phi(r)=0 for all rr near 11. Consider the map

h:𝔻𝔻h:\mathbb{D}\to\mathbb{D}
(r,θ)(r,θ+ϕ(r)).(r,\theta)\mapsto(r,\theta+\phi(r)).

Thus

hβ𝔻β𝔻=12πr2ϕ(r)dr=dg,h^{*}\beta_{\mathbb{D}}-\beta_{\mathbb{D}}=\frac{1}{2\pi}r^{2}\phi^{\prime}(r)dr=dg,

let

g(r,θ)=12π1rs2ϕ(s)𝑑s.g(r,\theta)=\frac{1}{2\pi}\int_{1}^{r}s^{2}\phi^{\prime}(s)ds.

The mean action of hh is equal to

𝒜(h)=gω𝔻=1π01r1rs2ϕ(s)𝑑s𝑑r>0.\mathcal{A}(h)=\int g\omega_{\mathbb{D}}=\frac{1}{\pi}\int_{0}^{1}r\int_{1}^{r}s^{2}\phi^{\prime}(s)dsdr>0.

We can use such a map on U=𝔻ϵU=\mathbb{D}_{\epsilon}, the disk of small radius ϵ>0\epsilon>0, to construct the desired local perturbation in Example 1.

Acknowledgement

The first author is supported by Sun Yat-sen University start-up grant No. 74120-18841290.

References

  • [1] Birkhoff, G. D. Proof of Poincaré’s geometric theorem. Trans. Amer. Math. Soc. 14, 1 (1913), 14–22.
  • [2] Franks, J. Generalizations of the Poincaré-Birkhoff theorem. Ann. of Math. (2) 128, 1 (1988), 139–151.
  • [3] Franks, J. Geodesics on S2S^{2} and periodic points of annulus homeomorphisms. Invent. Math. 108, 2 (1992), 403–418.
  • [4] Hutchings, M. Mean action and the Calabi invariant. J. Mod. Dyn. 10 (2016), 511–539.
  • [5] Mañé, R. Ergodic theory and differentiable dynamics, vol. 8 of Ergebnisse der Mathematik und ihrer Grenzgebiete (3) [Results in Mathematics and Related Areas (3)]. Springer-Verlag, Berlin, 1987. Translated from the Portuguese by Silvio Levy.