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

Global Hypersurfaces of Section for Geodesic Flows on Convex Hypersurfaces

Sunghae Cho and Dongho Lee
Abstract.

We construct a global hypersurface of section for the geodesic flow of a convex hypersurface in Euclidean space admits an isometric involution. This generalizes the Birkhoff annulus to higher dimensions.

1. Introduction

Global surfaces of section, an idea introduced by Poincaré in his work on celestial mechanics [Poi87] and also explored by Birkhoff [Bir66], feature prominently in the literature on the 3-dimensional dynamics. They allow us to reduce the dynamics of vector fields on 3-manifolds to the dynamics of surface diffeomorphisms. Ghys [Ghy09] called the existence of a global surface of section as a paradise for dynamicists, since it eliminates technical difficulties and allows to investigate the pure nature of dynamical systems.

Beyond their versatile utility, the existence of the global surfaces of section already gives us some information about the dynamics. Unless the given 3-manifold fibers over a circle, a global surface of section must have a boundary which is a set of periodic orbits. Hence the existence of periodic orbits is an essential obstruction, which cannot be overcome easily. Kuperberg [Kup94] introduced an example of a nonvanishing vector field on S3S^{3} without periodic orbits. Ginzburg [Gin99] has given examples of the Hamiltonian system without periodic orbits. The horocycle flow on unit cotangent bundle of higher genus surfaces provides another class of Hamiltonian flows without periodic orbits.

This obstruction can be removed in the case of Reeb dynamics in dimension three. The Weinstein conjecture [Wei79] asserts the existence of at least one periodic Reeb orbit on any compact contact manifold. Hofer [Hof93] proved the Weinstein conjecture on S3S^{3}, and Taubes [Tau07] proved the Weinstein conjecture in dimension three in a full generality using embedded contact homology. The work of Hofer was followed by the joint work with Wysocki and Zehnder [HWZ98] constructing disk-like global surface of section on a dynamically convex S3S^{3}. Recent developments along this way can be found in [HSa11], [HMSa15], and [SaH18]. [CDHR23], [CM22] and [CKMS22] contain results about the existence of a Birkhoff section for Reeb flows in dimension three, a generalization of a global surface of section that allows immersed boundary.

The notion of global surface of section can be generalized to higher dimensions as mentioned by Birkhoff [Bir66]. However, there are only few results known in higher dimensional dynamics even in the case of Reeb flows, due to the difficulty caused by the instability of the boundary of global hypersurface of section. Moreno and van Koert [MvK22] resolved this problem by imposing a symmetry which guarantees the existence of high dimensional invariant set, and provided a concrete example of a global hypersurface of section arising from the restricted three-body problem.

Our main result concerns the existence of global hypersurfaces of section of the geodesic flows on compact convex hypersurfaces in Euclidean space which satisfying certain conditions.

Theorem 1.1.

Let Mn+1M\subset\mathbb{R}^{n+1} be a regular closed hypersurface and NMN\subset M be a codimension 1 submanifold. Assume that there exists a tubular neighborhood ν(N)\nu(N) of NN and an isometric involution i:ν(N)ν(N)i:\nu(N)\to\nu(N) whose fixed point locus is NN. In addition, assume that MM has positive sectional curvature. Then the geodesic flow on STMST^{*}M admits a global hypersurface of section

P={(x,y)STM:xN,y,νx0},P=\left\{(x,y)\in ST^{*}M:x\in N,\langle y,\nu_{x}\rangle\geq 0\right\},

where ν\nu is a fixed normal vector field of NN with respect to MM. Moreover, the return map Ψ:P̊P̊\Psi:\mathring{P}\to\mathring{P} extends smoothly to the boundary of PP.

We note that the positivity of sectional curvature is equivalent to the convexity of the hypersurface. We first prove the existence part in 3.10, and 4.5 establishes that we can extend the return map smoothly to the boundary. Note that if the dimension of MM is 2, our theorem reduces to a special case of the theorem of Birkhoff [Bir66] which established the existence of a global surface of section of geodesic flow of a 2-sphere with positive curvature. In this sense, our result can be regarded as a generalization of the Birkhoff annulus to higher dimensions.

1.1. Acknowledgement

Sunghae Cho and Dongho Lee were supported by National Science Foundation of Korea Grant NRF2023005562 funded by the Korean Government.

2. Preliminaries

2.1. Global Hypersurfaces of Section and Open Book Decompositions

Let YY be a smooth closed manifold, i.e.  a compact manifold without boundary, and XX be a non-vanishing vector field on YY. A global hypersurface of section for XX is an embedded submanifold PYP\subset Y of codimension 1 with (possibly empty) boundary P=B\partial P=B such that

  • \cdot

    the vector field XX is transverse to the interior P̊\mathring{P} of PP,

  • \cdot

    the boundary BB is XX-invariant; in other words, XX is tangent to BB,

  • \cdot

    for any pYp\in Y, there exist t+,t>0t_{+},t_{-}>0 such that Flt+X(p),FltX(p)PFl^{X}_{t_{+}}(p),Fl^{X}_{-t_{-}}(p)\in P.

If PP is a global hypersurface of section, we can define the (first) return time τp\tau_{p} for each pP̊p\in\mathring{P} by τp=min{t>0:FltX(p)P}\tau_{p}=\min\{t>0:Fl^{X}_{t}(p)\in P\} and the (first) return map by Ψ(p)=FlτpX(p)\Psi(p)=Fl^{X}_{\tau_{p}}(p).

Example 2.1.

Let YY be STSnST^{*}S^{n}, the unit cotangent bundle of the nn-sphere with the standard metric, and XX be the unit geodesic vector field on YY. Consider YY as a subset of Tn+1n+1×n+1T^{*}\mathbb{R}^{n+1}\simeq\mathbb{R}^{n+1}\times\mathbb{R}^{n+1} and use the coordinates (x,y)=(x0,,xn,y0,,yn)(x,y)=(x_{0},\cdots,x_{n},y_{0},\cdots,y_{n}). Consider the submanifold

P={(x,y)Y:x0=0,y00},P=\left\{(x,y)\in Y:x_{0}=0,y_{0}\geq 0\right\},

which is the set of upward directions on the equator. Then one can see that PP is a global hypersurface of section, the boundary of PP is the unit cotangent bundle of the equator of SnS^{n}, and the first return map is the identity.

Open book decompositions are closely related to the global hypersurfaces of section. An open book decomposition on a closed manifold YY is a pair (B,π)(B,\pi) of a codimension 2 closed submanifold BB and a map π\pi from YBY\setminus B to S1S^{1}\subset\mathbb{C} which satisfies the following.

  • \cdot

    The normal bundle of BB is trivial. We call BB the binding. Let ξ:B×D2ν(B)\xi:B\times D^{2}\to\nu(B) be a fixed trivialization of the normal bundle, which is identified with a tubular beighborhood, of BB.

  • \cdot

    The map π\pi is a fiber bundle such that (πξ)(b;r,θ)=eiθ(\pi\circ\xi)(b;r,\theta)=e^{i\theta} on ν(B)B\nu(B)\setminus B, where (r,θ)(r,\theta) is a polar coordinate on D2D^{2}. We call the closure of each fiber π1(eiθ)¯=Pθ\overline{\pi^{-1}(e^{i\theta})}=P_{\theta} the page. Note that Pθ=B\partial P_{\theta}=B for any θ\theta.

Let XX be a vector field on YY, and (B,π)(B,\pi) be an open book on YY. If XX is transverse to each page PθP_{\theta} and tangent to BB, then we say XX is adapted to (B,π)(B,\pi).

Lemma 2.2.

Let YY be a closed manifold, BYB\subset Y be a codimension 2 closed submanifold, XX be a vector field on YY, and π\pi be a map from YBY\setminus B to S1S^{1}. Assume that

  1. (1)

    dpπ(X)>0d_{p}\pi(X)>0 for any pYp\in Y,

  2. (2)

    BB has a trivial tubular neighborhood, say B×D2B\times D^{2}, such that π(b;r,θ)=eiθ\pi(b;r,\theta)=e^{i\theta},

  3. (3)

    XX is tangent to BB.

Then (B,π)(B,\pi) is an open book decomposition of YY, which XX is adapted to.

Proof.

By (1), π\pi is a submersion. We can remove an open tubular neighborhood ν(B)\nu(B) of BB on which (2) holds, then π|Yν(B)\pi|_{Y\setminus\nu(B)} is still a submersion. In particular, (2) guarantees that π|ν(B)\pi|_{\partial\nu(B)} is a submersion. Since π|Yν(B)\pi|_{Y\setminus\nu(B)} is proper, we can apply the Ehresmann fibration theorem and conclude that π\pi defines a fiber bundle. With (2), we can see that (B,π)(B,\pi) is an open book decomposition on YY.

Since dπ(X)0d\pi(X)\neq 0, XX cannot be tangent to the level sets of π\pi. This means that XX is transverse to each page. With (3), we can see that XX is adapted to (B,π)(B,\pi). ∎

If a vector field XX is adapted to (B,π)(B,\pi), then each page Pθ=π1(eiθ)P_{\theta}=\pi^{-1}(e^{i\theta}) can be regarded as a candidate for the global hypersurface of section. There might exist an orbit of XX which does not return to the page in a finite time. Such an orbit should be asymptotic to the boundary as tt becomes large. A discussion about a case of dimension 3 can be found in [HWZ98]. In 3.5, we will show that the case of unbounded return time does not appear in our setting.

2.2. Symplectic Manifolds and Contact Manifolds

Let WW be a manifold without boundary and ω\omega be a 2-form on WW. If ω\omega is a non-degenerate closed form, we say ω\omega is a symplectic form and (W,ω)(W,\omega) is a symplectic manifold. A symplectomorphism is a diffeomorphism between symplectic manifolds which preserves the symplectic form. Let WWW^{\prime}\subset W be a submanifold such that ω|W\omega|_{W^{\prime}} is also a symplectic form on WW^{\prime}. We call WW^{\prime} a symplectic submanifold. A diffeomorphism between manifolds induces a symplectomorphism between cotangent bundles via pullback. Also, if NN is a submanifold of MM, then TNT^{*}N is a symplectic submanifold of TMT^{*}M.

Let (W,ω)(W,\omega) be a symplectic manifold with a boundary, and assume that ω\omega has a primitive λ\lambda, i.e.  ω=dλ\omega=d\lambda. A vector field XX such that iXω=λi_{X}\omega=\lambda is called Liouville vector field. If a Liouville vector field XX exists and defines an outward vector field on the boundary, we call (W,λ)(W,\lambda) a Liouville domain.

Given a smooth function H:WH:W\to\mathbb{R}, we can associate a vector field XHX_{H} to HH via

(iXH)ω()=ω(XH,)=dH().\left(i_{X_{H}}\right)\omega(-)=\omega(X_{H},-)=dH(-).

This is well-defined by non-degeneracy of ω\omega. We say XHX_{H} is a Hamiltonian vector field, and in this sense, we call HH a Hamiltonian function or simply Hamiltonian. If a diffeomorphism can be written as a time 1-flow of a Hamiltonian vector field, we say it is a Hamiltonian diffeomorphism. Note that we can also use time-dependent Hamiltonian H:W×H:W\times\mathbb{R}\to\mathbb{R} and get time-dependent Hamiltonian diffeomorphism. By definition, a Hamiltonian vector field XHX_{H} vanishes at a point pp if and only if dpH=0d_{p}H=0. A Hamiltonian diffeomorphism is a symplectomorphism, as (FlXH)ω=XHω=diXHω=0\left(Fl^{X_{H}}\right)^{*}\omega=\mathcal{L}_{X_{H}}\omega=di_{X_{H}}\omega=0 by the Cartan magic formula. An important observation is that the Hamiltonian vector field of an autonomous Hamiltonian is tangent to the regular level set of the generating Hamiltonian. In other words, the value of HH is preserved under the flow of XHX_{H}. More detailed explanations with examples about symplectic geometry and Hamiltonian mechanics can be found in [Arn89], [CdS01], [Ber01], [HZ11] or [MS17].

Let H,F:WH,F:W\to\mathbb{R} be Hamiltonians. The Poisson bracket is defined by {H,F}:=ω(XH,XF)\{H,F\}:=\omega(X_{H},X_{F}). It’s clear that the Poisson bracket is alternating. If (W,ω)=(Tn,ωstd)(W,\omega)=(T^{*}\mathbb{R}^{n},\omega_{\mathrm{std}}), we have the following formula, which can be found for example in the chapter 1 of [MS17]

{H,F}=jHyjFxjHxjFxj.\{H,F\}=\sum_{j}\frac{\partial H}{\partial y_{j}}\frac{\partial F}{\partial x_{j}}-\frac{\partial H}{\partial x_{j}}\frac{\partial F}{\partial x_{j}}.

Let YY be a (2n+1)(2n+1)-dimensional manifold, and ξ\xi be a 2n2n-dimensional distribution on YY. Then we can locally write ξ\xi as kerα\ker\alpha for some 1-form α\alpha. Assume that ξ\xi is coorientable, i.e.  TW/ξTW/\xi is orientable. Then we can find a globally defined 1-form α\alpha. If α(dα)n\alpha\wedge(d\alpha)^{n} is a volume form, we say α\alpha is a contact form, ξ=kerα\xi=\ker\alpha is a contact structure, and (W,ξ)(W,\xi) is a contact manifold. A contact form α\alpha defines a unique vector field RR such that α(R)=1\alpha(R)=1, iRdα=0i_{R}d\alpha=0. We call RR the Reeb vector field. Note that the Reeb vector field is always non-vanishing. A standard example of a contact manifold is the boundary of a Liouville domain (W,λ)(W,\lambda) with contact form λ|W\lambda|_{\partial W}, and a regular level set of Hamiltonian function H:(W,dλ)H:(W,d\lambda)\to\mathbb{R} with the same contact form. In the second case, the Hamiltonian vector field XHX_{H} restricted to the regular level set of HH is the Reeb vector field. We can consider an open book decomposition (B,π)(B,\pi) on a contact manifold (Y,kerα)(Y,\ker\alpha) which the Reeb vector field adapted to. The close relationship between contact structures on a manifold and an open book decomposition is explored in [Gir02].

2.3. Geodesic Flow as a Hamiltonian Flow

Let (M,g)(M,g) be a complete Riemannian manifold. For each xMx\in M and vTpMv\in T_{p}M, we have a unique geodesic γx,v\gamma_{x,v} with the initial condition γx,v(0)=x\gamma_{x,v}(0)=x, γ˙x,v(0)=v\dot{\gamma}_{x,v}(0)=v. The geodesic flow is a 1-parameter family of diffeomorphisms on TMTM defined by Φt(x,v)=(γx,v(t),γ˙x,v(t))\Phi_{t}(x,v)=(\gamma_{x,v}(t),\dot{\gamma}_{x,v}(t)). By differentiating Φt\Phi_{t} by tt, we get the geodesic vector field on TMTM which generates the geodesic flow.

The metric gg on TMTM induces the dual metric gg^{*} on TMT^{*}M by natural pairing, and we can also define the (co-)geodesic flow and (co-)geodesic vector field on TMT^{*}M.

Proposition 2.3.

The geodesic vector field on TMT^{*}M is a Hamiltonian vector field with Hamiltonian

H(x,y):=12(yg21).H(x,y):=\frac{1}{2}\left(||y||_{g^{*}}^{2}-1\right).

where TMT^{*}M is equipped with a canonical symplectic form ω=dxidyi\omega=\sum dx_{i}\wedge dy_{i}

Proof.

See the proof of Theorem 2.3.1 in [FvK18] or Theorem 1.5.2 in [Gei08]. ∎

The regular level set H1(0)H^{-1}(0) is a unit cotangent bundle STMST^{*}M, which naturally is a contact manifold, whose Reeb vector field is Hamiltonian vector field XHX_{H} restricted to STMST^{*}M.

For notational convenience, we write a point (x1,,xn)(x_{1},\ldots,x_{n}) in n\mathbb{R}^{n} by x\vec{x} so that x=(x0,x)x=(x_{0},\vec{x}) is a point in n+1\mathbb{R}^{n+1}. For a function f:n+1f:\mathbb{R}^{n+1}\to\mathbb{R}, we denote the partial derivative xif\partial_{x_{i}}f by fif_{i}, and the gradient vector field (f0,,fn)(f_{0},\ldots,f_{n}) by f\nabla f. We will also write the Hessian matrix of ff which is computed in n+1\mathbb{R}^{n+1} by Hess(f)=(xixjf)ij=(fij)\text{Hess}(f)=(\partial_{x_{i}}\partial_{x_{j}}f)_{ij}=(f_{ij}). We use coordinate (y0,,yn)(y_{0},\cdots,y_{n}) for the cotangent fiber Txn+1T^{*}_{x}\mathbb{R}^{n+1}, and also for the cotangent fiber of a hypersurface contained in n+1\mathbb{R}^{n+1}.

Let f:n+1f:\mathbb{R}^{n+1}\to\mathbb{R} be a smooth function, and 0 be a regular value. The level set M=f1(0)M=f^{-1}(0) is an nn-dimensional Riemannian manifold, whose metric is inherited from n+1\mathbb{R}^{n+1}. We can embed TMT^{*}M into Tn+1T^{*}\mathbb{R}^{n+1} by

TM={(x,y)Tn+1:f(x)=0,yf=0}.T^{*}M=\left\{(x,y)\in T^{*}\mathbb{R}^{n+1}:f(x)=0,\,\,y\cdot\nabla f=0\right\}.

Here, \cdot is a standard inner product on n+1\mathbb{R}^{n+1}, and we identified TMT^{*}M to TMTM by the metric on MM. Let H~=12(y21)\tilde{H}=\frac{1}{2}\left(||y||^{2}-1\right) on Tn+1T^{*}\mathbb{R}^{n+1}, and H=H~|TWH=\tilde{H}|_{T^{*}W}. Define f~,g:Tn+1\tilde{f},g:T^{*}\mathbb{R}^{n+1}\to\mathbb{R} by

f~(x,y)=f(x),g(x,y)=yf\tilde{f}(x,y)=f(x),\quad g(x,y)=y\cdot\nabla f

so that TMT^{*}M is an intersection f~1(0)g1(0)\tilde{f}^{-1}(0)\cap g^{-1}(0).

Proposition 2.4.

Let WW be a symplectic manifold, and H~\tilde{H} be a Hamiltonian on WW. Consider smooth functions f,g:Wf,g:W\to\mathbb{R} with c1,c2c_{1},c_{2} as the regular values of f,gf,g such that V=f1(c1)g1(c2)WV=f^{-1}(c_{1})\cap g^{-1}(c_{2})\subset W is a symplectic submanifold of codimension 2. Let H=H~|VH=\tilde{H}|_{V}. Then the Hamiltonian vector field XHX_{H} is given by

XH=XH~{g,H~}{g,f}Xf{f,H~}{f,g}Xg.X_{H}=X_{\tilde{H}}-\frac{\{g,\tilde{H}\}}{\{g,f\}}X_{f}-\frac{\{f,\tilde{H}\}}{\{f,g\}}X_{g}.
Proof.

We can write XHX_{H} as XH=XH~+aXf+bXgX_{H}=X_{\tilde{H}}+aX_{f}+bX_{g} for some functions a,ba,b. Since XHX_{H} is defined on the level set of ff and gg, we must have XH(f)=0=XH(g)X_{H}(f)=0=X_{H}(g). We also have Xf(f)=0=Xg(g)X_{f}(f)=0=X_{g}(g) from the definition. It follows that

0\displaystyle 0 =XH~(f)+bXg(f)={f,H~}+b{f,g},\displaystyle=X_{\tilde{H}}(f)+bX_{g}(f)=\{f,\tilde{H}\}+b\{f,g\},
0\displaystyle 0 =XH~(g)+aXf(g)={g,H~}+a{g,f}.\displaystyle=X_{\tilde{H}}(g)+aX_{f}(g)=\{g,\tilde{H}\}+a\{g,f\}.

Putting a,ba,b into the first equation yields the result. ∎

Remark 2.5.

The formula of 2.4 can be directly generalized to the case of any 2k2k-functions, say V=f11(c1)f2k1(c2k)V=f_{1}^{-1}(c_{1})\cap\cdots\cap f_{2k}^{-1}(c_{2k}). The coefficients of XfiX_{f_{i}}’s in that case will be a solution of a linear equation consisting of Poisson brackets.

From a straightforward computation, we have

XH~=jyjxj,Xf~=jfxjyj.X_{\tilde{H}}=\sum_{j}y_{j}\frac{\partial}{\partial x_{j}},\quad X_{\tilde{f}}=-\sum_{j}\frac{\partial f}{\partial x_{j}}\frac{\partial}{\partial y_{j}}.

Using the formula for the Poisson bracket, we have

{f~,g}\displaystyle\{\tilde{f},g\} =fxjgyj=(fxj)2=f2,\displaystyle=\sum\frac{\partial f}{\partial x_{j}}\frac{\partial g}{\partial y_{j}}=\sum\left(\frac{\partial f}{\partial x_{j}}\right)^{2}=||\nabla f||^{2},
{f~,H~}\displaystyle\{\tilde{f},\tilde{H}\} =fxjH~yj=fxjyj=yf=0,\displaystyle=\sum\frac{\partial f}{\partial x_{j}}\frac{\partial\tilde{H}}{\partial y_{j}}=\sum\frac{\partial f}{\partial x_{j}}y_{j}=y\cdot\nabla f=0,
{g,H~}\displaystyle\{g,\tilde{H}\} =gxjH~yj=jxj(iyifxi)yj=i,j2fxixjyiyj=Hess(f)(y,y).\displaystyle=\sum\frac{\partial g}{\partial x_{j}}\frac{\partial\tilde{H}}{\partial y_{j}}=\sum_{j}\frac{\partial}{\partial x_{j}}\left(\sum_{i}y_{i}\frac{\partial f}{\partial x_{i}}\right)y_{j}=\sum_{i,j}\frac{\partial^{2}f}{\partial x_{i}\partial x_{j}}y_{i}y_{j}=\mathrm{Hess}(f)(y,y).

To sum up, we have the following.

Theorem 2.6.

Let f:n+1f:\mathbb{R}^{n+1}\to\mathbb{R} be a smooth function, and cc be a regular value of ff. Let M=f1(c)M=f^{-1}(c). Then, the geodesic vector field on TMT^{*}M is given by

XH=jyjxjHess(f)x(y,y)f(x)2jfj(x)yj.X_{H}=\sum_{j}y_{j}\frac{\partial}{\partial x_{j}}-\frac{\mathrm{Hess}(f)_{x}(y,y)}{\|\nabla f(x)\|^{2}}\sum_{j}f_{j}(x)\frac{\partial}{\partial y_{j}}.

Alternatively, this formula can be derived by orthogonal projection.

2.4. Sectional Curvature of Hypersurfaces

An important quantity associated to the Riemannian manifold (M,g)(M,g) is the sectional curvature KM(,)K_{M}(-,-), which is defined for a point xMx\in M and two vectors v,wv,w in TpMT_{p}M. This can be regarded as the Gauss curvature of a 2-dimensional subspace ‘spanned’ by v,wv,w. For the precise definition and properties of sectional curvature, we refer to [Mil63], [KN63], [Spi79], or [dC92]. Let NMN\subset M be an oriented hypersurface, and ν\nu be the unit normal vector field on NN.

Proposition 2.7.

[[dC92], Theorem 6.2.5] Let S(v)=vνS(v)=\nabla_{v}\nu. Then the following formula holds.

KN(v,w)=KM(v,w)+S(v),vS(w),wS(v),w2v,vw,wv,w2K_{N}(v,w)=K_{M}(v,w)+\frac{\langle S(v),v\rangle\langle S(w),w\rangle-\langle S(v),w\rangle^{2}}{\langle v,v\rangle\langle w,w\rangle-\langle v,w\rangle^{2}}

The formula becomes simpler in the case of a hypersurface in the Euclidean space, whose sectional curvature vanishes.

Corollary 2.8.

[[dC92], Theorem 6.2.5] Let f:n+1f:\mathbb{R}^{n+1}\to\mathbb{R} be a function, and M=f1(c)M=f^{-1}(c) be a regular hypersurface. Let v,wv,w be orthonormal vector fields on TMTM. Then we have the following.

KM(v,w)=S(v),vS(w),wS(v),w2.K_{M}(v,w)=\langle S(v),v\rangle\langle S(w),w\rangle-\langle S(v),w\rangle^{2}.

3. A Global Hypersurface of Section of the Geodesic Flow on a Convex Hypersurface

3.1. Setting

We consider the case f:n+1f:\mathbb{R}^{n+1}\to\mathbb{R} is a smooth function and 0 is a regular value. Let MM be the regular level set f1(0)f^{-1}(0). We assume the following conditions for ff.

  1. (A1)

    For any point (x0,x)(x_{0},\vec{x}) in MM such that |x0||x_{0}| is small enough, f(x0,x)=f(x0,x)f(x_{0},\vec{x})=f(-x_{0},\vec{x}).

  2. (A2)

    The Hessian of ff is positive definite. Explicitly, for any (x,y)TM(x,y)\in T^{*}M, y0y\neq 0, Hess(f)x(y,y)>0\mathrm{Hess}(f)_{x}(y,y)>0.

Here is a classical theorem which explains the role of (A1). For the proof, see [KN69] Chapter VII.8.

Lemma 3.1.

Let (M,g)(M,g) be a Riemannian manifold and NN be a closed submanifold. Assume that there exist a tubular neighborhood ν(N)\nu(N) of NN and an isometric involution i:ν(N)ν(N)i:\nu(N)\to\nu(N), i.e., ii is an isometry, iIdi\neq\mathrm{Id} and i2=Idi^{2}=\mathrm{Id}. In addition, assume that NN is the fixed point locus of ii. Then NN is a totally geodesic submanifold. In other words, the geodesic vector field of TMT^{*}M restricted to TNT^{*}N is tangent to TNT^{*}N.

Let NN be the codimension 11 submanifold N=M{(x,y):x0=0}.N=M\cap\{(x,y):x_{0}=0\}. Then it is clear that NN is a fixed point set of a locally defined isometric involution iM(x0,x)=(x0,x)i_{M}(x_{0},\vec{x})=(-x_{0},\vec{x}), and it follows that NN is a totally geodesic submanifold from 3.1. Let Y=STMY=ST^{*}M and B=STNB=ST^{*}N. Then BB can be written as

B={(x,y)STM:x0=0,y0=0}.B=\{(x,y)\in ST^{*}M:x_{0}=0,y_{0}=0\}.

We note the triviality of the normal bundle of BB in YY.

Lemma 3.2.

Let MM be a Riemannian manifold diffeomorphic to an nn-sphere, and NMN\subset M be a codimension 11 closed submanifold. Then STNST^{*}N has a trivial normal bundle in STMST^{*}M.

Proof.

We’ll first show that the normal bundle ν(N)\nu(N) of NN is trivial, which is equivalent to the orientability of NN. Suppose not, then for any section ss of ν(N)\nu(N) which meets zero section transversely, there exists a loop γ\gamma such that s|γs|_{\gamma} has odd number of zeros. It follows that the intersection form [N][γ][N]\cdot[\gamma] in 2\mathbb{Z}_{2}-coordinate is not 0, which is a contradiction because Hn1(Sn;2)=H1(Sn;2)=0H_{n-1}(S^{n};\mathbb{Z}_{2})=H_{1}(S^{n};\mathbb{Z}_{2})=0.

Now, let νM(N)N×(ϵ,ϵ)M\nu_{M}(N)\simeq N\times(-\epsilon,\epsilon)\subset M. For xNx\in N, we have TxM=TxNT_{x}M=\mathbb{R}\oplus T_{x}N, implying

STxM={(t,v)TxN:t2+v2=1},ST_{x}M=\left\{(t,v)\in\mathbb{R}\oplus T_{x}N:t^{2}+\lVert v\rVert^{2}=1\right\},

where \lVert\cdot\rVert is the metric inherited from MM. Thus, the normal fiber at p=(x,v)STNp=(x,v)\in ST^{*}N is

νSTM(STN)p=νSTxM(STxN)vνM(N)xνSn1(Sn)vνM(N)x,\nu_{ST^{*}M}(ST^{*}N)_{p}=\nu_{ST_{x}^{*}M}(ST_{x}^{*}N)_{v}\oplus\nu_{M}(N)_{x}\simeq\nu_{S^{n-1}}(S^{n})_{v}\oplus\nu_{M}(N)_{x},

where Sn1S^{n-1} is embedded in SnS^{n} along the equator. ∎

Condition (A2) implies that ff is a convex function, and MM bounds a compact convex domain. It follows that MM is diffeomorphic to SnS^{n}, where the diffeomorphism can be explicitly written as xx/xx\mapsto x/||x||, assuming that 0 is in the interior of the bounding region of MM. Since MM is convex and 0 is in the interior of the domain MM bounds, this is well-defined.

3.2. Existence of a Global Hypersurface of Section

With the setting of the previous section, we define a map π:YBS1\pi:Y\setminus B\to S^{1}\subset\mathbb{C} by

π(x,y)=x0+iy0|x0+iy0|.\pi(x,y)=\frac{x_{0}+iy_{0}}{|x_{0}+iy_{0}|}.

The angular form is defined by

Θ=idlogπ=y0dx0x0dy0x02+y02=θx02+y02.\Theta=i\cdot d\log\pi=\frac{y_{0}dx_{0}-x_{0}dy_{0}}{x_{0}^{2}+y_{0}^{2}}=\frac{\theta}{x_{0}^{2}+y_{0}^{2}}.

Put XHX_{H} which was computed in 2.6 into θ\theta, we have the following.

θ(XH)=y02+x02Hess(f)x(y,y)f(x)2f0(x)x0.\theta(X_{H})=y_{0}^{2}+x_{0}^{2}\frac{\mathrm{Hess}(f)_{x}(y,y)}{\|\nabla f(x)\|^{2}}\frac{f_{0}(x)}{x_{0}}.

We define a function A=A(x,y)A=A(x,y) by

A(x,y)=Hess(f)x(y,y)f(x)2f0(x)x0A(x,y)=\frac{\mathrm{Hess}(f)_{x}(y,y)}{\|\nabla f(x)\|^{2}}\frac{f_{0}(x)}{x_{0}}

so that θ(XH)=A(x,y)x02+y02\theta(X_{H})=A(x,y)x_{0}^{2}+y_{0}^{2}.

Proposition 3.3.

Under the assumptions (A1) and (A2), there exists ε>0\varepsilon>0, which only depends on the function ff, such that A(x,y)>εA(x,y)>\varepsilon for any (x,y)YB(x,y)\in Y\setminus B.

Proof.

Since 0 is a regular value of ff and YY is compact, there exists some C>0C>0 such that 0<f2C0<||\nabla f||^{2}\leq C on YY. Also, by the compactness of YY, condition (A2) implies that there exists some δ>0\delta>0 such that Hess(f)x(y,y)>δ\mathrm{Hess}(f)_{x}(y,y)>\delta for any (x,y)Y(x,y)\in Y.

By condition (A1), we have f0|x0=0=0f_{0}|_{x_{0}=0}=0. Thus, we can perform a Taylor expansion with respect to x0x_{0}:

f(x0,x)\displaystyle f(x_{0},\vec{x}) =f(0,x)+12f00(0,x)x02+O(x03),\displaystyle=f(0,\vec{x})+\frac{1}{2}f_{00}(0,\vec{x})x_{0}^{2}+O(x_{0}^{3}),
f0(x0,x)x0\displaystyle\frac{f_{0}(x_{0},\vec{x})}{x_{0}} =f00(0,x)+O(x0).\displaystyle=f_{00}(0,\vec{x})+O(x_{0}).

Since Hess(f)x(y,y)>δ\mathrm{Hess}(f)_{x}(y,y)>\delta, we have f00(x)=Hess(f)x((y0,0),(y0,0))>δf_{00}(x)=\mathrm{Hess}(f)_{x}((y_{0},0),(y_{0},0))>\delta for any xx. We can choose a small η>0\eta>0 such that for |x0|<η|x_{0}|<\eta, f0/x0>δ/2f_{0}/x_{0}>\delta/2. Consequently, we have A(x,y)>δ2/2CA(x,y)>\delta^{2}/2C for |x0|<η|x_{0}|<\eta.

For |x0|η|x_{0}|\geq\eta, we only need to bound f0/x0f_{0}/x_{0}. However, f0(0,x)=0f_{0}(0,\vec{x})=0 and f00>δf_{00}>\delta implies that f0(x0,x)>0f_{0}(x_{0},\vec{x})>0 if x0>0x_{0}>0 and f0(x0,x)<0f_{0}(x_{0},\vec{x})<0 if x0<0x_{0}<0. By the compactness of Y{|x0|η}Y\cap\{|x_{0}|\geq\eta\}, there exists some δ1>0\delta_{1}>0 such that f0/x0>δ1f_{0}/x_{0}>\delta_{1}. This implies A(x,y)>δδ1/CA(x,y)>\delta\delta_{1}/C for |x0|η|x_{0}|\geq\eta. Taking ε=min(δ2/2C,δδ1/C)\varepsilon=\min(\delta^{2}/2C,\delta\delta_{1}/C), we obtain the desired lower bound. ∎

Theorem 3.4.

Under the assumptions (A1) and (A2), π:YBS1\pi:Y\setminus B\to S^{1} defines an open book decomposition, which the geodesic vector field is adapted to.

Proof.

We will apply 2.2 to this situation. From 3.3, it is evident that there exists ε>0\varepsilon>0 such that

Θ(XH)=A(x,y)x02+y02x02+y02>ε\Theta(X_{H})=\frac{A(x,y)x_{0}^{2}+y_{0}^{2}}{x_{0}^{2}+y_{0}^{2}}>\varepsilon

for any (x,y)YB(x,y)\in Y\setminus B. Since Θ=idlogπ\Theta=i\cdot d\log\pi, the first condition on π\pi and XHX_{H} in 2.2 is satisfied.

Now consider the trivial tubular neighborhood of BB, the existence of which is guaranteed by 3.2, denoted as

ν(B)\displaystyle\nu(B) B×D2\displaystyle\simeq B\times D^{2}
(x,y)\displaystyle(x,y) (x,y;x0,y0)\displaystyle\mapsto(\vec{x},\vec{y};x_{0},y_{0})

where x0,y0x_{0},y_{0} are small enough. Note that π(b,r,θ)=eiθ\pi(b,r,\theta)=e^{i\theta}.

Lastly, since NN is a totally geodesic submanifold by (A1), the geodesic vector field XHX_{H} is tangent to STN=BST^{*}N=B. Therefore, we can apply 2.2 and obtain the desired result. ∎

Theorem 3.5.

Under the assumptions (A1) and (A2), the geodesic flow on Y=STMY=ST^{*}M admits a global hypersurface of section, which is given by

P={(x,y)Y:x0=0,y00}.P=\{(x,y)\in Y:x_{0}=0,y_{0}\geq 0\}.
Proof.

The hypersurface PP corresponds to a page π1(i)\pi^{-1}(i) of the open book constructed in 3.4. To conclude, we need to demonstrate that the return time is bounded. It suffices to show that there exists t>0t\in\mathbb{R}_{>0} such that π(FltXH(x,y))=π(x,y)\pi(Fl^{X_{H}}_{t}(x,y))=\pi(x,y). From 3.3, we observe the existence of ε>0\varepsilon>0 such that Θ(XH)>ε\Theta(X_{H})>\varepsilon. Thus, we have

02π/εidlogπ(XH)>02π/εε𝑑t>2π.\int_{0}^{2\pi/\varepsilon}i\cdot d\log\pi(X_{H})>\int_{0}^{2\pi/\varepsilon}\varepsilon dt>2\pi.

By the intermediate value theorem, we can conclude that there exists a positive number τ<2π/ε\tau<2\pi/\varepsilon such that 0τidlogπ(XH)=2π\int_{0}^{\tau}i\cdot d\log\pi(X_{H})=2\pi, which means that π(FlτXH(x,y))=π(x,y).\pi(Fl^{X_{H}}_{\tau}(x,y))=\pi(x,y). It means that τ\tau is a bounded positive finite return time for (x,y)(x,y). The boundedness of the negative return time can be demonstrated similarly. ∎

Corollary 3.6.

Under the assumptions (A1) and (A2), the geodesic flow on MM admits an S1S^{1}-family of global hypersurfaces of section.

Proof.

We have an open book decomposition with bounded return time, so we can choose any page of the open book, let’s say π1(eiθ)=Pθ\pi^{-1}(e^{i\theta})=P_{\theta}, as a global hypersurface of section. ∎

Note that if we take n=2n=2, then MM is 2-sphere so P=Pπ/2P=P_{\pi/2} in 3.5 is the Birkhoff annulus. In this sense, our construction can be regarded as a generalization of the Birkhoff annulus.

3.3. Relation with Sectional Curvature

We can apply 2.8 on our hypersurface MM. The unit normal vector ν\nu is ff\frac{\nabla f}{||\nabla f||}, and we have

S(xi)=xiff=j(fijfkfkfkifjf3)xj.S\left(\frac{\partial}{\partial x_{i}}\right)=\frac{\partial}{\partial x_{i}}\frac{\nabla f}{||\nabla f||}=\sum_{j}\left(\frac{f_{ij}}{||\nabla f||}-\frac{\sum_{k}f_{k}f_{ki}f_{j}}{||\nabla f||^{3}}\right)\frac{\partial}{\partial x_{j}}.

It follows that

S(v,w)=i,jvifijwifi,j,kvifikfkfjwjf3=Hess(f)(v,w)fS(v,w)=\frac{\sum_{i,j}v_{i}f_{ij}w_{i}}{||\nabla f||}-\frac{\sum_{i,j,k}v_{i}f_{ik}f_{k}f_{j}w_{j}}{||\nabla f||^{3}}=\frac{\mathrm{Hess}(f)(v,w)}{||\nabla f||}

for v,wTMv,w\in TM. The second term vanishes because wTMw\in TM implies that wf=jwjfj=0w\cdot\nabla f=\sum_{j}w_{j}f_{j}=0.

Put this into 2.8, and we get the following.

Proposition 3.7.

Let v,wv,w be orthogonal unit tangent vectors of a point xMx\in M. Then,

KM(v,w)=Hess(f)x(v,v)Hess(f)x(w,w)Hess(f)x(v,w)2f(x)2.K_{M}(v,w)=\frac{\mathrm{Hess}(f)_{x}(v,v)\mathrm{Hess}(f)_{x}(w,w)-\mathrm{Hess}(f)_{x}(v,w)^{2}}{||\nabla f(x)||^{2}}.

Hence, the sign of KMK_{M} depends on the sign of the (2×22\times 2)-minors of Hess(f)\mathrm{Hess}(f). Since the formula holds for any vv and ww, we have the following.

Corollary 3.8.

The sectional curvature KM(σ)xK_{M}(\sigma)_{x} is positive for any point xMx\in M and plane σTxM\sigma\subset T_{x}M if and only if Hess(f)x\mathrm{Hess}(f)_{x} is either positive definite or negative definite.

Proof.

Since Hess(f)x\mathrm{Hess}(f)_{x} is symmetric bilinear form, there exists a basis 𝔅\mathfrak{B} of TxMT_{x}M that diagonalizes Hess(f)x\mathrm{Hess}(f)_{x}, and it continuously varies with xx. Let

[Hess(f)x]𝔅=diag(λ1,,λn).[\mathrm{Hess}(f)_{x}]_{\mathfrak{B}}=\mathrm{diag}(\lambda_{1},\cdots,\lambda_{n}).

Since KM>0K_{M}>0, we must have λiλj>0\lambda_{i}\lambda_{j}>0 for any i,ji,j. It means that λi\lambda_{i} all have the same sign, which implies that Hess(f)x\mathrm{Hess}(f)_{x} is either positive definite or negative definite. Since the process depends on xx continuously, the sign of λi\lambda_{i} cannot change since λi0\lambda_{i}\neq 0. The converse follows from the formula in 3.7. ∎

The following corollary can also be found in various literature, for example [Sac60].

Corollary 3.9.

Let f:n+1f:\mathbb{R}^{n+1}\to\mathbb{R} and M=f1(0)n+1M=f^{-1}(0)\subset\mathbb{R}^{n+1} be a regular hypersurface. Assume that MM has a positive sectional curvature. Then MM is diffeomorphic to the nn-sphere.

Proof.

By 3.8, Hess(f)\mathrm{Hess}(f) is either positive definite or negative definite. If Hess(f)\mathrm{Hess}(f) is positive definite, f1(,0]f^{-1}(-\infty,0] is convex and we can get the result. If Hess(f)\mathrm{Hess}(f) is negative definite, then for f¯=f\bar{f}=-f, M=f¯1(0)M=\bar{f}^{-1}(0), and Hess(f¯)\mathrm{Hess}(\bar{f}) is positive definite, and we can use the same argument. ∎

Now we can formulate 3.5 in another form.

Theorem 3.10.

Let M=f1(0)n+1M=f^{-1}(0)\subset\mathbb{R}^{n+1} be a hypersurface satisfying (A1), and MM has a positive sectional curvature. Then there exists an open book decomposition (π,B)(\pi,B) of STMST^{*}M to which the geodesic vector field XHX_{H} is adapted, as given in 3.4. Moreover, there exists a global hypersurface of section PP, as given in 3.5.

Proof.

Since KM>0K_{M}>0, 3.8 implies that ff is either positive definite or negative definite. If ff is positive definite, we’re done. If ff is negative definite, define f~=f\tilde{f}=-f. Then f~\tilde{f} is positive definite, f~1(0)=f1(0)\tilde{f}^{-1}(0)=f^{-1}(0), and the condition (A1) still holds. Moreover, the formula in 2.6 for ff and f~\tilde{f} are the same. Thus we get the result. ∎

3.4. Topology of the Global Hypersurface of Section

Using the same notations from the previous sections, we will explore the topology of NN and PP.

Lemma 3.11.

Under the assumptions (A1) and (A2), NN is diffeomorphic to Sn1S^{n-1}.

Proof.

We can express NN as

N={(0,x)n+1:f(0,x)=0}{(0,x)n+1:xn}n.N=\left\{(0,\vec{x})\in\mathbb{R}^{n+1}:f(0,\vec{x})=0\right\}\subset\left\{(0,\vec{x})\in\mathbb{R}^{n+1}:\vec{x}\in\mathbb{R}^{n}\right\}\simeq\mathbb{R}^{n}.

The convexity of ff is preserved if we restrict ff to the subspace {(0,x):xn}\left\{(0,\vec{x}):\vec{x}\in\mathbb{R}^{n}\right\}, so we obtain the result. ∎

Proposition 3.12.

The global hypersurface of section PP constructed in 3.5 is diffeomorphic to T1Sn1T_{\leq 1}^{*}S^{n-1}, which is the subset of TSn1T^{*}S^{n-1} consisting of covectors of length 1\leq 1. The boundary B=PB=\partial P is homeomorphic to STSn1ST^{*}S^{n-1}.

Proof.

Let pp be a diffeomorphism from the upper hemisphere HnSnH^{n}\subset S^{n} to the closed disk DnD^{n}. Define a map ϕ:PT1N\phi:P\to T^{*}_{\leq 1}N by ϕ(x,y)=(x,p(y))\phi(x,y)=(x,p(y)), then it’s clear that ϕ\phi is a diffeomorphism. With 3.11, we can conclude the result. ∎

4. Return Map

4.1. General Properties

In this subsection, we investigate the general properties of the global hypersurfaces of sections for Reeb vector fields. Let α\alpha be a contact form on a manifold YY and RR be its Reeb vector field. Assume that there exists an open book decomposition (B,π)(B,\pi) of YY to which RR is adapted. Let P=Pθ=π1(eiθ)P=P_{\theta}=\pi^{-1}(e^{i\theta}) be a page, and assume that the first return map Ψ:P̊P̊\Psi:\mathring{P}\to\mathring{P} is well-defined, i.e.  the return time τ\tau is bounded for each point pPp\in P.

Lemma 4.1.

The interior of PP is a symplectic manifold with symplectic form dαd\alpha.

Proof.

Let dimY=2n+1\dim Y=2n+1. Since YY is contact, αdαn\alpha\wedge d\alpha^{n} never vanishes. Since P̊\mathring{P} is transverse to RR, we can take a local frame (X1,,X2n)(X_{1},\cdots,X_{2n}) of P̊\mathring{P} such that (R,X1,,X2n)(R,X_{1},\cdots,X_{2n}) is a local frame of YY, and

αdαn(R,X1,,X2n)=α(R)dαn(X1,,X2n)0.\alpha\wedge d\alpha^{n}(R,X_{1},\cdots,X_{2n})=\alpha(R)d\alpha^{n}(X_{1},\cdots,X_{2n})\neq 0.

It means that dαd\alpha is a non-degenerate closed 2-form on P̊\mathring{P}, so it’s a symplectic form. ∎

Proposition 4.2.

For the return map Ψ:P̊P̊\Psi:\mathring{P}\to\mathring{P}, we have

Ψαα=dτ.\Psi^{*}\alpha-\alpha=d\tau.

In particular, Ψ\Psi is a symplectomorphism.

Proof.

This is a generalization of the well-known fact in dimension 3, which can be found, for example, in [ABHSa17] or [FvK18]. We have FlτpR(p)=Ψ(p)Fl^{R}_{\tau_{p}}(p)=\Psi(p). Differentiating both sides and plugging in a vector field XX, we have

dFlτpR(p)X+(dτp(X))R=dpΨ(X).dFl^{R}_{\tau_{p}}(p)X+(d\tau_{p}(X))R=d_{p}\Psi(X).

Since FlRFl^{R} preserves α\alpha, we have that

Ψα(X)=α(dΨ(X))=(FlR)α(X)+α(R)dτ(X)=α(X)+dτ(X).\Psi^{*}\alpha(X)=\alpha(d\Psi(X))=(Fl^{R})^{*}\alpha(X)+\alpha(R)d\tau(X)=\alpha(X)+d\tau(X).

The second statement follows if we take exterior derivative on both sides. ∎

Remark 4.3.

If we equip dαd\alpha to the whole PP, which is a manifold with boundary, then (P,dα)(P,d\alpha) is not a Liouville domain. More precisely, dαd\alpha degenerates at the boundary of PP. The Reeb vector field RR is tangent to P\partial P and iRdα=0i_{R}d\alpha=0, indicating that dαd\alpha degenerates on the boundary. But even though dαd\alpha degenerates at the boundary, (P,ker(α|P))(\partial P,\ker(\alpha|_{\partial P})) is a contact manifold.

4.2. Extension to the Boundary

In the preceding sections, we constructed a global hypersurface of section PP for the geodesic flow on a convex hypersurface M=f1(0)M=f^{-1}(0) in n+1\mathbb{R}^{n+1} and we obtained a return map Ψ:P̊P̊\Psi:\mathring{P}\to\mathring{P}. As demonstrated in 4.1, P̊\mathring{P} is a symplectic manifold, with its symplectic form being the restriction of the standard symplectic form on TMT^{*}M. Furthermore, as indicated in 4.2, Ψ\Psi is a symplectomorphism.

Now we investigate the boundary behavior of the return map in terms of the defining function ff. Given that (x0,y0)(\partial_{x_{0}},\partial_{y_{0}}) forms a normal symplectic frame of a contact submanifold B=STNB=ST^{*}N, it follows from Lemma 8.1 of [MvK22] the linearized Hamiltonian flow decomposes into two blocks; one corresponding to TBTB and the other to x0,y0\langle\partial_{x_{0}},\partial_{y_{0}}\rangle. Let ZN=(x0(t)y0(t))Z_{N}=\begin{pmatrix}x_{0}(t)\\ y_{0}(t)\end{pmatrix} be the normal part of a flowline linearized near BB. Then we have

Z˙N=JNSNZN\dot{Z}_{N}=J_{N}S_{N}Z_{N}

where JNJ_{N} is the normal part of the standard complex structure on TY=TTMTY=TT^{*}M and SNS_{N} is a matrix called the normal Hessian.

Proposition 4.4.

[MvK22] If the normal Hessian is positive definite, the return map extends to the boundary smoothly, and such an extension is unique.

Theorem 4.5.

Under the assumptions (A1) and (A2), or the assumption of 3.10, the return map Ψ:P̊P̊\Psi:\mathring{P}\to\mathring{P} can be smoothly extended to the boundary.

Proof.

Instead of working with a normal Hessian, we will use linearized flow. According to Proposition 8.2 of [MvK22], we must have

Θ(XH)=ZNtSNZNZNtZN+O(1)\Theta(X_{H})=\frac{Z_{N}^{t}S_{N}Z_{N}}{Z_{N}^{t}Z_{N}}+O(1)

near (x0,y0)=(0,0)(x_{0},y_{0})=(0,0). From the previous section, we also have

Θ(XH)=A(x,y)x02+y02x02+y02=ZNtdiag(Hess(f)x(y,y)f(x)2f00(x),1)ZNZNtZN+O(x0).\Theta(X_{H})=\frac{A(x,y)x_{0}^{2}+y_{0}^{2}}{x_{0}^{2}+y_{0}^{2}}=\frac{Z_{N}^{t}\mathrm{diag}\left(\frac{\mathrm{Hess}(f)_{x}(y,y)}{||\nabla f(x)||^{2}}f_{00}(x),1\right)Z_{N}}{Z_{N}^{t}Z_{N}}+O(x_{0}).

It follows that

SN=diag(Hess(f)x(y,y)f(x)2f00(x),1).S_{N}=\mathrm{diag}\left(\frac{\mathrm{Hess}(f)_{x}(y,y)}{||\nabla f(x)||^{2}}f_{00}(x),1\right).

From the condition (A2), we observe that Hess(f)(y,y)\mathrm{Hess}(f)(y,y) and f00f_{00} are always positive, and it follows that SNS_{N} is positive definite. From 4.4, we get the result. If we impose the positive sectional curvature condition, it follows that Hess(f)(y,y)\mathrm{Hess}(f)(y,y) and f00f_{00} have the same sign, which leads to the same result. ∎

Example 4.6.

3.4 can be applied to ellipsoids, since they are convex and satisfying the symmetry condition (A1). In particular, consider an ellipsoid given by

E={xn+1:x02a02+x12++xn21=0},E=\left\{x\in\mathbb{R}^{n+1}:\frac{x_{0}^{2}}{a_{0}^{2}}+x_{1}^{2}+\cdots+x_{n}^{2}-1=0\right\},

where a0>0a_{0}\in\mathbb{R}_{>0}. The global hypersurface of section constructed in 3.4 is the upper-hemisphere bundle on the equator,

P={((0,x),(y0,y))Tn+1:x=1,x,y=0,y02+y2=1,y00}.P=\left\{((0,\vec{x}),(y_{0},\vec{y}))\in T^{*}\mathbb{R}^{n+1}\,:\,||\vec{x}||=1,\,\langle\vec{x},\vec{y}\rangle=0,\,y_{0}^{2}+||\vec{y}||^{2}=1,\,y_{0}\geq 0\right\}.

The return map Ψ\Psi on PP can be computed in terms of elliptic integrals

Ψ((0,x),(y0,y))=((0,xcosG(y)+yysinG(y)),(y0,ycosG(y)yxsinG(y))),\Psi\left((0,\vec{x}),(y_{0},\vec{y})\right)=\left(\left(0,\vec{x}\cos{G(\|\vec{y}\|)}+\frac{\vec{y}}{\|\vec{y}\|}\sin{G(\|\vec{y}\|)}\right),\left(y_{0},\vec{y}\cos{G(\|\vec{y}\|)}-\|\vec{y}\|\vec{x}\sin{G(\|\vec{y}\|)}\frac{}{}\right)\right),

where GG is given by

G(t)=t(1a02)a0F(2π|(1a02)(1t2)a02)+ta0Π(1t2;2π|(1a02)(1t2)a02).G(t)=-\frac{t(1-a_{0}^{2})}{a_{0}}F\left(2\pi\,\Big{|}\,\frac{-(1-a_{0}^{2})(1-t^{2})}{a_{0}^{2}}\right)+\frac{t}{a_{0}}\Pi\left(1-t^{2};2\pi\,\Big{|}\,\frac{-(1-a_{0}^{2})(1-t^{2})}{a_{0}^{2}}\right).

where FF is the elliptic integral of the first kind and Π\Pi is the elliptic integral of the third kind.

This can be regarded as a specific case of a hypersurface of revolution. Let f:n+1f:\mathbb{R}^{n+1}\to\mathbb{R} satisfy (A1) globally, (A2) and the third condition; if x=x||\vec{x}||=||\vec{x}\,^{\prime}||, then f(x0,x)=f(x0,x)f(x_{0},\vec{x})=f(x_{0},\vec{x}\,^{\prime}). For convenience, we assume that f(0,x)=0f(0,\vec{x})=0 if and only if x=1||\vec{x}||=1. The hypersurface of revolution is given by a regular level set M=f1(0)M=f^{-1}(0). In this case, we can parametrize the set M{(x0,x1,0,,0)}M\cap\left\{(x_{0},x_{1},0,\cdots,0)\right\} by (a(ϕ),cosϕ,0,,0)(a(\phi),\cos\phi,0,\cdots,0), where aa is a function on ϕ\phi. Note that in the case of ellipsoid, a(ϕ)=a0sinϕa(\phi)=a_{0}\sin\phi. We can compute the return map with a help of the Clairaut integral [Arn89], which has a same form as in the case of ellipsoid with

G(t):=t02π(1t2)sin2σ+{a(arcsin(1t2sinσ))}21(1t2)sin2σ𝑑σ.G(t):=t\int_{0}^{2\pi}\frac{\sqrt{(1-t^{2})\sin^{2}{\sigma}+\{a^{\prime}(\arcsin{(\sqrt{1-t^{2}}\sin{\sigma})})\}^{2}}}{1-(1-t^{2})\sin^{2}{\sigma}}d\sigma.

It’s straightforward to see that the return map Ψ\Psi is a Hamiltonian diffeomorphism generated by H(x,y)=(G)(y)H(x,y)=\left(\int G\right)(||\vec{y}||).

Sending a(ϕ)a(\phi) to 0, the dynamics converges to a billiard on the unit disk D={xn:x1}D=\left\{\vec{x}\in\mathbb{R}^{n}:||\vec{x}||\leq 1\right\}, which is defined on a set

Y0={(x,y)Tn:x1,y=1,y points inward if x=1}.Y_{0}=\left\{(\vec{x},\vec{y})\in T^{*}\mathbb{R}^{n}:||\vec{x}||\leq 1,\,||\vec{y}||=1,\,\vec{y}\text{ points inward if }||\vec{x}||=1\right\}.

The global hypersurface of section converges to P0={(x,y)Y0:x=1}P_{0}=\left\{(\vec{x},\vec{y})\in Y_{0}:||\vec{x}||=1\right\}. We write y=yT+yN\vec{y}=\vec{y}_{T}+\vec{y}_{N} for (x,y)P0(\vec{x},\vec{y})\in P_{0}, where yT\vec{y}_{T} is a component tangent to D\partial D, and and yN\vec{y}_{N} is a component normal to D\partial D. Then the return map converges to the second iterate of the high-dimensional billiard map

Ψ(x,y)=(xcosG0(yT)+yTyTsinG0(yT),yTcosG0(yT)yTxsinG0(yT)),\Psi\left(\vec{x},\vec{y}\right)=\left(\vec{x}\cos{G_{0}(\|\vec{y}_{T}\|)}+\frac{\vec{y}_{T}}{\|\vec{y}_{T}\|}\sin{G_{0}(\|\vec{y}_{T}\|)},\,\vec{y_{T}}\cos{G_{0}(\|\vec{y}_{T}\|)}-\|\vec{y}_{T}\|\vec{x}\sin{G_{0}(\|\vec{y_{T}}\|)}\right),

where G0(t)=4arccostG_{0}(t)=4\arccos t, which agrees with the function GG with a(ϕ)=0a(\phi)=0.

References

  • [ABHSa17] Alberto Abbondandolo, Barney Bramham, Umberto L. Hryniewicz, and Pedro A. S. Salomão, A systolic inequality for geodesic flows on the two-sphere, Math. Ann. 367 (2017), no. 1-2, 701–753. MR 3606452
  • [Arn89] V. I. Arnold̀, Mathematical methods of classical mechanics, Graduate Texts in Mathematics, vol. 60, Springer-Verlag, New York, [1989?], Translated from the 1974 Russian original by K. Vogtmann and A. Weinstein, Corrected reprint of the second (1989) edition. MR 1345386
  • [Ber01] Rolf Berndt, An introduction to symplectic geometry, Graduate Studies in Mathematics, vol. 26, American Mathematical Society, Providence, RI, 2001, Translated from the 1998 German original by Michael Klucznik. MR 1793955
  • [Bir66] George D. Birkhoff, Dynamical systems, American Mathematical Society Colloquium Publications, vol. Vol. IX, American Mathematical Society, Providence, RI, 1966, With an addendum by Jurgen Moser. MR 209095
  • [CDHR23] Vincent Colin, Pierre Dehornoy, Umberto Hryniewicz, and Ana Rechtman, Generic properties of 33-dimensional Reeb flows: Birkhoff sections and entropy, 2023.
  • [CdS01] Ana Cannas da Silva, Lectures on symplectic geometry, Lecture Notes in Mathematics, vol. 1764, Springer-Verlag, Berlin, 2001. MR 1853077
  • [CKMS22] Gonzalo Contreras, Gerhard Knieper, Marco Mazzucchelli, and Benjamin H. Schulz, Surfaces of section for geodesic flows of closed surfaces, 2022.
  • [CM22] Gonzalo Contreras and Marco Mazzucchelli, Existence of Birkhoff sections for Kupka-Smale Reeb flows of closed contact 3-manifolds, Geom. Funct. Anal. 32 (2022), no. 5, 951–979. MR 4498837
  • [dC92] Manfredo Perdigão do Carmo, Riemannian geometry, portuguese ed., Mathematics: Theory & Applications, Birkhäuser Boston, Inc., Boston, MA, 1992. MR 1138207
  • [FvK18] Urs Frauenfelder and Otto van Koert, The restricted three-body problem and holomorphic curves, Pathways in Mathematics, Birkhäuser/Springer, Cham, 2018. MR 3837531
  • [Gei08] Hansjörg Geiges, An introduction to contact topology, Cambridge Studies in Advanced Mathematics, vol. 109, Cambridge University Press, Cambridge, 2008. MR 2397738
  • [Ghy09] Étienne Ghys, Right-handed vector fields & the Lorenz attractor, Jpn. J. Math. 4 (2009), no. 1, 47–61. MR 2491282
  • [Gin99] Viktor L. Ginzburg, Hamiltonian dynamical systems without periodic orbits, Northern California Symplectic Geometry Seminar, Amer. Math. Soc. Transl. Ser. 2, vol. 196, Amer. Math. Soc., Providence, RI, 1999, pp. 35–48. MR 1736212
  • [Gir02] Emmanuel Giroux, Géométrie de contact: de la dimension trois vers les dimensions supérieures, Proceedings of the International Congress of Mathematicians, Vol. II (Beijing, 2002), Higher Ed. Press, Beijing, 2002, pp. 405–414. MR 1957051
  • [HMSa15] Umberto Hryniewicz, Al Momin, and Pedro A. S. Salomão, A Poincaré-Birkhoff theorem for tight Reeb flows on S3S^{3}, Invent. Math. 199 (2015), no. 2, 333–422. MR 3302117
  • [Hof93] H. Hofer, Pseudoholomorphic curves in symplectizations with applications to the Weinstein conjecture in dimension three, Invent. Math. 114 (1993), no. 3, 515–563. MR 1244912
  • [HSa11] Umberto Hryniewicz and Pedro A. S. Salomão, On the existence of disk-like global sections for Reeb flows on the tight 3-sphere, Duke Math. J. 160 (2011), no. 3, 415–465. MR 2852366
  • [HWZ98] H. Hofer, K. Wysocki, and E. Zehnder, The dynamics on three-dimensional strictly convex energy surfaces, Ann. of Math. (2) 148 (1998), no. 1, 197–289. MR 1652928
  • [HZ11] Helmut Hofer and Eduard Zehnder, Symplectic invariants and Hamiltonian dynamics, Modern Birkhäuser Classics, Birkhäuser Verlag, Basel, 2011, Reprint of the 1994 edition. MR 2797558
  • [KN63] Shoshichi Kobayashi and Katsumi Nomizu, Foundations of differential geometry. Vol I, Interscience Publishers (a division of John Wiley & Sons, Inc.), New York-London, 1963. MR 152974
  • [KN69] by same author, Foundations of differential geometry. Vol. II, Interscience Tracts in Pure and Applied Mathematics, No. 15, vol. Vol. II, Interscience Publishers John Wiley & Sons, Inc., New York-London-Sydney, 1969. MR 238225
  • [Kup94] Krystyna Kuperberg, A smooth counterexample to the Seifert conjecture, Ann. of Math. (2) 140 (1994), no. 3, 723–732. MR 1307902
  • [Mil63] J. Milnor, Morse theory, Annals of Mathematics Studies, vol. No. 51, Princeton University Press, Princeton, NJ, 1963, Based on lecture notes by M. Spivak and R. Wells. MR 163331
  • [MS17] Dusa McDuff and Dietmar Salamon, Introduction to symplectic topology, third ed., Oxford Graduate Texts in Mathematics, Oxford University Press, Oxford, 2017. MR 3674984
  • [MvK22] Agustin Moreno and Otto van Koert, Global hypersurfaces of section in the spatial restricted three-body problem, Nonlinearity 35 (2022), no. 6, 2920–2970. MR 4443924
  • [Poi87] H. Poincaré, Les méthodes nouvelles de la mécanique céleste. Tome I, Les Grands Classiques Gauthier-Villars. [Gauthier-Villars Great Classics], Librairie Scientifique et Technique Albert Blanchard, Paris, 1987, Solutions périodiques. Non-existence des intégrales uniformes. Solutions asymptotiques. [Periodic solutions. Nonexistence of uniform integrals. Asymptotic solutions], Reprint of the 1892 original, With a foreword by J. Kovalevsky, Bibliothèque Scientifique Albert Blanchard. [Albert Blanchard Scientific Library]. MR 926906
  • [Sac60] Richard Sacksteder, On hypersurfaces with no negative sectional curvatures, Amer. J. Math. 82 (1960), 609–630. MR 116292
  • [SaH18] Pedro A. S. Salomão and Umberto L. Hryniewicz, Global surfaces of section for Reeb flows in dimension three and beyond, Proceedings of the International Congress of Mathematicians—Rio de Janeiro 2018. Vol. II. Invited lectures, World Sci. Publ., Hackensack, NJ, 2018, pp. 941–967. MR 3966795
  • [Spi79] Michael Spivak, A comprehensive introduction to differential geometry. Vol. II, second ed., Publish or Perish, Inc., Wilmington, DE, 1979. MR 532831
  • [Tau07] Clifford Henry Taubes, The Seiberg-Witten equations and the Weinstein conjecture, Geom. Topol. 11 (2007), 2117–2202. MR 2350473
  • [Wei79] Alan Weinstein, On the hypotheses of Rabinowitz’ periodic orbit theorems, J. Differential Equations 33 (1979), no. 3, 353–358. MR 543704