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

On a local systolic inequality for odd-symplectic forms

Gabriele Benedetti and Jungsoo Kang
Abstract

The aim of this paper is to formulate a local systolic inequality for odd-symplectic forms (also known as Hamiltonian structures) and to establish it in some basic cases.

Let Ω\Omega be an odd-symplectic form on an oriented closed manifold Σ\Sigma of odd dimension. We say that Ω\Omega is Zoll if the trajectories of the flow given by Ω\Omega are the orbits of a free S1S^{1}-action. After defining the volume of Ω\Omega and the action of its periodic orbits, we prove that the volume and the action satisfy a polynomial equation, provided Ω\Omega is Zoll. This builds the equality case of a conjectural systolic inequality for odd-symplectic forms close to a Zoll one. We prove the conjecture when the S1S^{1}-action yields a flat S1S^{1}-bundle or Ω\Omega is quasi-autonomous. In particular the conjecture is established in dimension three.

This new inequality recovers the contact systolic inequality as well as the inequality between the minimal action and the Calabi invariant for Hamiltonian isotopies C1C^{1}-close to the identity on a closed symplectic manifold. Applications to the study of periodic magnetic geodesics on closed orientable surfaces is given in the companion paper [BK19b].

1 Introduction

In this article we formulate a local systolic-diastolic inequality for odd-symplectic forms and establish it in some situations. Before discussing the precise set-up, we recall two inequalities which motivate us to study the local systolic-diastolic inequality for odd-symplectic forms and which arise as extreme cases of this inequality. Throughout the paper, we normalise the circle S1S^{1} to have length one, namely S1:=/S^{1}:={\mathbb{R}}/{\mathbb{Z}}.

1.1 Two known inequalities

Contact systolic-diastolic inequality

Let Σ\Sigma be a connected closed manifold of dimension 2n+12n+1. Let α\alpha be a contact form on Σ\Sigma, namely α\alpha is a one-form such that α(dα)n\alpha\wedge({\mathrm{d}}\alpha)^{n} is nowhere vanishing. Orienting Σ\Sigma so that α(dα)n\alpha\wedge({\mathrm{d}}\alpha)^{n} is positive, we define the total volume of Σ\Sigma with respect to α\alpha by

Volume(α):=Σα(dα)n>0.\mathrm{Volume}(\alpha):=\int_{\Sigma}\alpha\wedge({\mathrm{d}}\alpha)^{n}>0. (1.1)

We consider the Reeb vector field RαR_{\alpha} on Σ\Sigma characterised by the relations dα(Rα,)=0{\mathrm{d}}\alpha(R_{\alpha},\cdot)=0 and α(Rα)=1\alpha(R_{\alpha})=1 and denote by Φα\Phi_{\alpha} the generated flow.

A contact form α\alpha_{*} on Σ\Sigma is called Zoll if all orbits of Φα\Phi_{\alpha_{*}} are periodic with the same prime period T(α)T(\alpha_{*}). This is equivalent to saying that Φα\Phi_{\alpha_{*}} induces a free circle action on Σ\Sigma with period T(α)T(\alpha_{*}). In this case, the quotient is a closed symplectic manifold (M,ω)(M,\omega) and the quotient map 𝔭:ΣM\mathfrak{p}:\Sigma\to M is a non-trivial circle bundle. More precisely, we have 𝔭ω=dα\mathfrak{p}^{*}\omega={\mathrm{d}}\alpha_{*} so that 1T(α)ω-\tfrac{1}{T(\alpha_{*})}\omega represent the Euler class of the bundle. Vice versa, for every closed symplectic manifold (M,ω)(M,\omega) such that 1Tω\frac{1}{T}\omega represents an integer class for some T>0T>0, there exists a manifold Σ\Sigma and a Zoll contact form with period TT yielding (M,ω)(M,\omega) according to the construction above, see [BW58].

By [Bot80], if 𝔥\mathfrak{h} denotes the free homotopy class of prime periodic orbits of Φα\Phi_{\alpha_{*}}, then Φα\Phi_{\alpha} carries a periodic orbit in the class 𝔥\mathfrak{h}, if α\alpha is sufficiently close to α\alpha_{*} . We denote by Tmin(α,𝔥)T_{\min}(\alpha,\mathfrak{h}), respectively Tmax(α,𝔥)T_{\max}(\alpha,\mathfrak{h}) the minimal, respectively maximal, period of prime periodic orbits of Φα\Phi_{\alpha} in the class 𝔥\mathfrak{h}. If Σ\Sigma has dimension three, Tmin(α,𝔥)T_{\min}(\alpha,\mathfrak{h}) and Tmax(α,𝔥)T_{\max}(\alpha,\mathfrak{h}) satisfy the following contact systolic-diastolic inequality, which was originally conjectured by Álvarez-Paiva and Balacheff [ÁPB14].

Theorem 1.1.

[BK19a, Theorem 1.4] Let Σ\Sigma be a closed manifold of dimension three, and let |H1tor(Σ;)||H_{1}^{\mathrm{tor}}(\Sigma;{\mathbb{Z}})| be the order of the torsion part of H1(Σ,)H_{1}(\Sigma,{\mathbb{Z}}). If α\alpha_{*} is a Zoll contact form on Σ\Sigma, there exists a C2C^{2}-neighbourhood 𝒰\mathcal{U} of dα{\mathrm{d}}\alpha_{*} such that for every contact form α\alpha on Σ\Sigma with dα𝒰{\mathrm{d}}\alpha\in\mathcal{U}, there holds

Tmin(α,𝔥)21|H1tor(Σ;)|Volume(α)Tmax(α,𝔥)2,T_{\min}(\alpha,\mathfrak{h})^{2}\leq\frac{1}{|H_{1}^{\mathrm{tor}}(\Sigma;{\mathbb{Z}})|}\mathrm{Volume}(\alpha)\leq T_{\max}(\alpha,\mathfrak{h})^{2},

where any of the two equalities holds if and only if α\alpha is Zoll.

Remark 1.2.

When Σ=S3\Sigma=S^{3} the result above is due to Abbondandolo, Bramham, Hryniewicz and Salomão, see [ABHS18]. In that paper, forms α\alpha are also constructed, which are C0C^{0}-close to α\alpha_{*} and violate the systolic inequality. Furthermore, the same authors showed in [ABHS17] that, if (Σ,ξ)(\Sigma,\xi) is an arbitrary contact closed three-manifold, then, for every C>0C>0, there exists a contact form αC\alpha_{C} with kerαC=ξ\ker\alpha_{C}=\xi such that Tmin(αC)2>CVolume(αC)T_{\min}(\alpha_{C})^{2}>C\cdot\mathrm{Volume}(\alpha_{C}).

Symplectic systolic-diastolic inequality

Let (M,ω)(M,\mathrm{\omega}) be a connected closed symplectic manifold of dimension 2n2n, and let φ:MM\varphi:M\to M be a Hamiltonian diffeomorphism generated by a Hamiltonian H:M×[0,1]H:M\times[0,1]\to{\mathbb{R}}. We orient MM so that ωn\mathrm{\omega}^{n} is positive. One can define the Hamiltonian action 𝒜H:Λshort(M)\mathcal{A}_{H}:\Lambda_{\mathrm{short}}(M)\to{\mathbb{R}} on the space of “short” one-periodic curves q:S1Mq:S^{1}\to M by

𝒜H(γ):=D2q^ω+01H(q(t),t)dt,\mathcal{A}_{H}(\gamma):=\int_{D^{2}}{\widehat{q}}^{\hskip 2.0pt*}\mathrm{\omega}+\int_{0}^{1}H(q(t),t){\mathrm{d}}t, (1.2)

where q^:D2M\widehat{q}:D^{2}\to M is a “small” capping disc for qΛshort(M)q\in\Lambda_{\mathrm{short}}(M) (see Section 7 for a precise definition). The minimal and maximal Hamiltonian actions of fixed points of φ\varphi, whose associated curve is short, are given by

min𝒜H:=infγCrit𝒜H𝒜H(γ),max𝒜H:=supγCrit𝒜H𝒜H(γ).\min\mathcal{A}_{H}:=\inf_{\gamma\in\mathrm{Crit\,}\mathcal{A}_{H}}\mathcal{A}_{H}(\gamma),\qquad\max\mathcal{A}_{H}:=\sup_{\gamma\in\mathrm{Crit\,}\mathcal{A}_{H}}\mathcal{A}_{H}(\gamma).

The Calabi invariant of HH with respect to ω\mathrm{\omega} is defined by

CALω(H):=M×[0,1]Hωndt.\mathrm{CAL}_{\mathrm{\omega}}(H):=\int_{M\times[0,1]}H\,\omega^{n}\wedge{\mathrm{d}}t. (1.3)

One can easily see that the minimal/maximal actions and the Calabi invariant are related in the following way. It is one particular case of Proposition 7.16.

Proposition 1.3.

Assume that a Hamiltonian diffeomorphism φ:(M,ω)(M,ω)\varphi:(M,\mathrm{\omega})\to(M,\mathrm{\omega}) is generated by a quasi-autonomous Hamiltonian H:M×[0,1]H:M\times[0,1]\to{\mathbb{R}}. For instance, φ\varphi is the time-one map of a Hamiltonian isotopy C1C^{1}-close to the identity. Then, there holds

min𝒜HCALω(H)Mωnmax𝒜H,\min\mathcal{A}_{H}\leq\frac{\mathrm{CAL}_{\mathrm{\omega}}(H)}{\int_{M}\mathrm{\omega}^{n}}\leq\max\mathcal{A}_{H},

and any of the two equalities holds if and only if H(q,t)=h(t)H(q,t)=h(t) for all (q,t)M×[0,1](q,t)\in M\times[0,1] and some function h:[0,1]h:[0,1]\to{\mathbb{R}}.∎

Remark 1.4.

Proposition 1.3 is false for general Hamiltonian diffeomorphisms. Indeed, adapting [ABHS18, Proposition 2.28], one can construct a time-one map φ\varphi of a Hamiltonian isotopy C0C^{0}-close to the identity not satisfying the inequality above. A proper way to go beyond the quasi-autonomous case might be to take into account not only the fixed points but all the periodic points of φ\varphi as explored by Hutchings in [Hut16, Theorem 1.2].

Remark 1.5.

In [BK19a] we saw how the use of a global surface of section reduces Theorem 1.1 to a generalised version of Proposition 1.3 for surfaces with boundary, where ω\omega is symplectic in the interior and vanishes of order one at the boundary.

To give a unified model for the above two situations, we notice that the Hamiltonian diffeomorphism φ:MM\varphi:M\to M can be seen as the return map of a flow ΦX\Phi^{X} on the manifold M×S1M\times S^{1}. If H:M×S1H:M\times S^{1}\to{\mathbb{R}} is a Hamiltonian one-periodic in time generating φ\varphi as time one-map, then X(q,t):=t+XHt(q)X(q,t):=\partial_{t}+X_{H_{t}}(q), for all (q,t)M×S1(q,t)\in M\times S^{1}. Here, XHtX_{H_{t}} is the Hamiltonian vector field tangent to M×{t}M\times\{t\} associated with the function Ht:=H(,t)H_{t}:=H(\cdot,t). Thus, we view Proposition 1.3 as giving a systolic-diastolic inequality for those flows on the trivial S1S^{1}-bundle 𝔭:M×S1M\mathfrak{p}:M\times S^{1}\to M, which are obtained by lifting a Hamiltonian isotopy of MM. On the other hand, Theorem 1.1 yields a systolic-diastolic inequality for Reeb flows on the non-trivial S1S^{1}-bundles obtained from the Boothby-Wang construction. We will show that these two types of flows are particular incarnations of the flow induced by an odd-symplectic structure (also known as Hamiltonian structure, see [CM05]) on an oriented odd-dimensional closed manifold which is the total space of an arbitrary oriented circle bundle over a symplectic manifold. Moreover, we will formulate a conjectural systolic-diastolic inequality in this setting which recovers the contact (Theorem 1.1) and the symplectic one (Proposition 1.3) as two extreme cases.

1.2 Odd-symplectic systolic geometry

Let (Σ,𝔬Σ)(\Sigma,\mathfrak{o}_{\Sigma}) be a connected oriented closed manifold of dimension 2n+12n+1, and let CHdR2(Σ)C\in H^{2}_{\mathrm{dR}}(\Sigma) be a class in its de Rham cohomology. We write Ω1(Σ)\Omega^{1}(\Sigma) for the set of one-forms on Σ\Sigma and ΞC2(Σ)\Xi^{2}_{C}(\Sigma) for the set of closed two-forms on Σ\Sigma representing the class CC. We pick an auxiliary element Ω0ΞC2(Σ)\Omega_{0}\in\Xi^{2}_{C}(\Sigma) and get a surjective map

Ω1(Σ)ΞC2(Σ),αΩα:=Ω0+dα,\Omega^{1}(\Sigma)\to\Xi^{2}_{C}(\Sigma),\qquad\alpha\mapsto\Omega_{\alpha}:=\Omega_{0}+{\mathrm{d}}\alpha, (1.4)

whose kernel is the space of closed one-forms on Σ\Sigma. Using this map, we can associate to αΩ1(Σ)\alpha\in\Omega^{1}(\Sigma) a real number Vol(α)\mathrm{Vol}(\alpha), which generalises both the contact volume (1.1) and the Calabi invariant (1.3). Namely, we define a functional Vol:Ω1(Σ)\mathrm{Vol}:\Omega^{1}(\Sigma)\to{\mathbb{R}} by

Vol(0)=0,dαVolβ=ΣβΩαn,βΩ1(Σ).\mathrm{Vol}(0)=0,\qquad{\mathrm{d}}_{\alpha}\mathrm{Vol}\cdot\beta=\int_{\Sigma}\beta\wedge\Omega_{\alpha}^{n},\quad\forall\,\beta\in\Omega^{1}(\Sigma).

For instance, when n=1n=1, we have

Vol(α)=12Σαdα+ΣαΩ0.\mathrm{Vol}(\alpha)=\frac{1}{2}\int_{\Sigma}\alpha\wedge{\mathrm{d}}\alpha+\int_{\Sigma}\alpha\wedge\Omega_{0}.

We also remark that, when C=0C=0 and Ω0=0\Omega_{0}=0, the function Vol\mathrm{Vol} recovers the Chern-Simons action for principal S1S^{1}-bundles over Σ\Sigma (see [CS74] and Remark 2.1). For n=1n=1, Ω0=0\Omega_{0}=0 and dα=ιFμ{\mathrm{d}}\alpha=\iota_{F}\mu for some vector field FF on Σ\Sigma, whose flow preserves a global volume form μ\mu on Σ\Sigma, Vol(α)\mathrm{Vol}(\alpha) recovers the helicity of FF, see [Pat09].

If Cn=0C^{n}=0, then we can use the map (1.4) to push forward Vol\mathrm{Vol} to a volume functional

𝔙𝔬𝔩:ΞC2(Σ).{\mathfrak{Vol}}:\Xi^{2}_{C}(\Sigma)\to{\mathbb{R}}.

If Cn0C^{n}\neq 0, this procedure does not work, since there exists τHdR1(Σ)\tau\in H^{1}_{\mathrm{dR}}(\Sigma) such that τCn0\tau\cup C^{n}\neq 0. In this case, we say that αΩ1(Σ)\alpha\in\Omega^{1}(\Sigma) is normalised, if Vol(α)=0\mathrm{Vol}(\alpha)=0. For every ΩΞC2(Σ)\Omega\in\Xi^{2}_{C}(\Sigma), there exists αΩ1(Σ)\alpha\in\Omega^{1}(\Sigma) normalised such that Ω=Ωα\Omega=\Omega_{\alpha}. Therefore, in this situation we can just work with normalised forms and declare 𝔙𝔬𝔩:ΞC2(Σ){\mathfrak{Vol}}:\Xi^{2}_{C}(\Sigma)\to{\mathbb{R}} to be identically zero. In both cases, the volume functional is invariant under diffeomorphisms Ψ:ΣΣ\Psi:\Sigma\to\Sigma isotopic to the identity (see Proposition 2.8):

𝔙𝔬𝔩(ΨΩ)=𝔙𝔬𝔩(Ω).{\mathfrak{Vol}}(\Psi^{*}\Omega)={\mathfrak{Vol}}(\Omega).

Having identified what the volume should be, we want to introduce the set 𝒳(Ω)\mathcal{X}(\Omega) of closed characteristics of Ω\Omega generalising periodic Reeb and Hamiltonian orbits. To this purpose, we consider the possibly singular distribution kerΩΣ\ker\Omega\to\Sigma called the characteristic distribution and define

𝒳(Ω):={γ:S1Σ|γ˙kerΩ}/,\mathcal{X}(\Omega):=\Big{\{}\gamma:S^{1}\hookrightarrow\Sigma\ \Big{|}\ \dot{\gamma}\in\ker\Omega\Big{\}}\Big{/}\!\sim\,,

where γ1γ2\gamma_{1}\sim\gamma_{2} if and only if γ1\gamma_{1} and γ2\gamma_{2} coincide up to an orientation-preserving reparametrisation of S1S^{1}. The distribution kerΩ\ker\Omega is co-oriented by Ωn\Omega^{n} and we orient it using the given orientation 𝔬Σ\mathfrak{o}_{\Sigma} on Σ\Sigma.

We single out the forms in ΞC2(Σ)\Xi^{2}_{C}(\Sigma), whose characteristic distribution is one-dimensional. They are of special importance, as their closed characteristics are the periodic orbits of a flow.

Definition 1.6.

A two-form Ω\Omega on Σ\Sigma is said to be odd-symplectic, if it is closed and maximally non-degenerate, namely if its characteristic distribution is an (oriented) real line bundle over Σ\Sigma. We write 𝒮C(Σ)\mathcal{S}_{C}(\Sigma) for the subset of all odd-symplectic forms in ΞC2(Σ)\Xi^{2}_{C}(\Sigma).

In 𝒮C(Σ)\mathcal{S}_{C}(\Sigma) we hope to find elements whose characteristic distribution is as simple as possible. To this purpose we introduce the set of oriented S1S^{1}-bundles with total space Σ\Sigma:

𝔓(Σ):={𝔭:ΣM|𝔭is an oriented S1-bundle}.\mathfrak{P}(\Sigma):=\big{\{}\mathfrak{p}:\Sigma\to M\ \big{|}\ \mathfrak{p}\ \text{is an oriented $S^{1}$-bundle}\big{\}}.
Definition 1.7.

An odd-symplectic form Ω\Omega is said to be Zoll if the oriented leaves of its characteristic distribution are the fibres of some 𝔭Ω:ΣMΩ\mathfrak{p}_{\Omega}:\Sigma\to M_{\Omega} in 𝔓(Σ)\mathfrak{P}(\Sigma). We write 𝒵C(Σ){\mathcal{Z}}_{C}(\Sigma) for the set of Zoll forms in 𝒮C(Σ)\mathcal{S}_{C}(\Sigma).

Example 1.8.

Let (N1,σ1)(N_{1},\sigma_{1}) and (N2,σ2)(N_{2},\sigma_{2}) be two connected closed symplectic manifolds and suppose that [σ1]HdR2(N1)[\sigma_{1}]\in H^{2}_{\mathrm{dR}}(N_{1}) is an integral cohomology class. We build the product symplectic manifold (N1×N2,σ1σ2)(N_{1}\times N_{2},\sigma_{1}\oplus\sigma_{2}) and consider the oriented S1S^{1}-bundle 𝔭:ΣN1×N2\mathfrak{p}:\Sigma\to N_{1}\times N_{2} whose Euler class is [σ10]-[\sigma_{1}\oplus 0]. Then, Ω:=𝔭(σ1σ2)\Omega:=\mathfrak{p}^{*}(\sigma_{1}\oplus\sigma_{2}) is a Zoll form with cohomology class C:=𝔭[0σ2]C:=\mathfrak{p}^{*}[0\oplus\sigma_{2}].

Odd-symplectic Zoll forms Ω\Omega with vanishing cohomology class are exactly the differentials of Zoll contact forms α\alpha. We can use this observation to extend the classification of Zoll contact forms on three-manifolds contained in [BK19a, Proposition 1.2] to Zoll odd-symplectic forms.

Proposition 1.9.

Let Σ\Sigma be a connected oriented closed three-manifold. There is a Zoll odd-symplectic form on Σ\Sigma if and only if Σ\Sigma is the total space of an oriented S1S^{1}-bundle over a connected oriented closed surface MM. When the bundle is non-trivial, then every Zoll odd-symplectic form is the differential of a Zoll contact form. When the bundle is trivial, then the group H1(Σ;)H_{1}(\Sigma;{\mathbb{Z}}) is free with rank equal to 2genus(M)+12\,\mathrm{genus}(M)+1, and if Ω\Omega and Ω\Omega^{\prime} are Zoll odd-symplectic forms on Σ\Sigma, there exists a real number T>0T>0 and a diffeomorphism Ψ:ΣΣ\Psi:\Sigma\to\Sigma such that ΨΩ=TΩ\Psi^{*}\Omega^{\prime}=T\Omega. Moreover,

  • if Σ=S2×S1\Sigma=S^{2}\times S^{1}, the set 𝒵C(Σ){\mathcal{Z}}_{C}(\Sigma) has exactly two connected components, CHdR2(Σ)0\forall\,C\in H^{2}_{{\mathrm{dR}}}(\Sigma)\setminus 0;

  • if Σ=𝕋2×S1\Sigma={\mathbb{T}}^{2}\times S^{1}, the set 𝒵C(Σ){\mathcal{Z}}_{C}(\Sigma) is non-empty and connected, CHdR2(Σ)0\forall\,C\in H^{2}_{{\mathrm{dR}}}(\Sigma)\setminus 0;

  • if Σ=M×S1\Sigma=M\times S^{1} with χ(M)<0\chi(M)<0, there is a one-dimensional subspace LHdR2(Σ)L\subset H^{2}_{{\mathrm{dR}}}(\Sigma) such that 𝒵C(Σ){\mathcal{Z}}_{C}(\Sigma) is non-empty if and only if CL0C\in L\setminus 0; in this case the set 𝒵C(Σ){\mathcal{Z}}_{C}(\Sigma) is connected.

Remark 1.10.

The classification of Zoll odd-symplectic forms up to diffeomorphism on a three-manifold is equivalent to the classification of bundles in the set {𝔭Ω|Ω𝒵(Σ)}\{\mathfrak{p}_{\Omega}\ |\ \Omega\in{\mathcal{Z}}(\Sigma)\} up to isomorphism. Analogously, the connected components of 𝒵C(Σ){\mathcal{Z}}_{C}(\Sigma) on a three-manifold are in bijection with the connected components of {𝔭Ω|Ω𝒵C(Σ)}\{\mathfrak{p}_{\Omega}\ |\ \Omega\in{\mathcal{Z}}_{C}(\Sigma)\}. This is due to the fact that the map 𝒵C(Σ)𝔓(Σ){\mathcal{Z}}_{C}(\Sigma)\to\mathfrak{P}(\Sigma), given by Ω𝔭Ω\Omega\mapsto\mathfrak{p}_{\Omega}, has contractible fibres.

Let us assume that 𝒵C(Σ){\mathcal{Z}}_{C}(\Sigma) is not empty and take Ω𝒵C(Σ)\Omega_{*}\in{\mathcal{Z}}_{C}(\Sigma) with associated S1S^{1}-bundle

𝔭Ω:ΣM.\mathfrak{p}_{\Omega_{*}}:\Sigma\to M_{*}.

This implies that there exists a positive symplectic form ω\mathrm{\omega}_{*} on MM_{*} such that Ω=𝔭Ωω\Omega_{*}=\mathfrak{p}_{\Omega_{*}}^{*}\mathrm{\omega}_{*}. We set c:=[ω]HdR2(M)c_{*}:=[\mathrm{\omega}_{*}]\in H^{2}_{\mathrm{dR}}(M_{*}). We write 𝔥[S1,Σ]\mathfrak{h}\in[S^{1},\Sigma] for the free-homotopy class of the oriented 𝔭Ω\mathfrak{p}_{\Omega_{*}}-fibres and eHdR2(M)e_{*}\in H^{2}_{\mathrm{dR}}(M_{*}) for minus the real Euler class of 𝔭Ω\mathfrak{p}_{\Omega_{*}}. Let 𝔓0(Σ)\mathfrak{P}^{0}(\Sigma) be the connected component of 𝔭Ω\mathfrak{p}_{\Omega_{*}} inside 𝔓(Σ)\mathfrak{P}(\Sigma). As we did for the volume, the action will be computed with respect to some reference object, which we now define.

Definition 1.11.

A weakly Zoll pair is a couple (𝔭,c)(\mathfrak{p},c), where 𝔭:ΣM\mathfrak{p}:\Sigma\to M is an element in 𝔓(Σ)\mathfrak{P}(\Sigma) and cHdR2(M)c\in H^{2}_{\mathrm{dR}}(M) is a cohomology class. We write (Σ)\mathfrak{Z}(\Sigma) for the set of weakly Zoll pairs and C(Σ)\mathfrak{Z}_{C}(\Sigma) for the subset of those pairs (𝔭,c)(\mathfrak{p},c) such that C=𝔭cHdR2(Σ)C=\mathfrak{p}^{*}c\in H_{\mathrm{dR}}^{2}(\Sigma). We denote by C0(Σ)\mathfrak{Z}_{C}^{0}(\Sigma) the set of those (𝔭,c)C(Σ)(\mathfrak{p},c)\in\mathfrak{Z}_{C}(\Sigma) such that 𝔭𝔓0(Σ)\mathfrak{p}\in\mathfrak{P}^{0}(\Sigma).

Remark 1.12.

If Ω𝒮C(Σ)\Omega\in\mathcal{S}_{C}(\Sigma) is Zoll, then (𝔭Ω,[ωΩ])C(Σ)(\mathfrak{p}_{\Omega},[\mathrm{\omega}_{\Omega}])\in\mathfrak{Z}_{C}(\Sigma) is a weakly Zoll pair, where ωΩ\mathrm{\omega}_{\Omega} is the closed two-form on MΩM_{\Omega} such that Ω=𝔭ΩωΩ\Omega=\mathfrak{p}_{\Omega}^{*}\mathrm{\omega}_{\Omega}. Conversely, if (𝔭,c)C(Σ)(\mathfrak{p},c)\in\mathfrak{Z}_{C}(\Sigma) is a weakly Zoll pair, we can consider any closed two-form ω\mathrm{\omega} on MM such that c=[ω]c=[\mathrm{\omega}] and build the two-form 𝔭ωΞC2(Σ)\mathfrak{p}^{*}\mathrm{\omega}\in\Xi^{2}_{C}(\Sigma), which is Zoll exactly when ω\mathrm{\omega} is symplectic.

Let us now fix a reference weakly Zoll pair

(𝔭0,c0)C0(Σ),𝔭0:ΣM0.(\mathfrak{p}_{0},c_{0})\in\mathfrak{Z}_{C}^{0}(\Sigma),\qquad\mathfrak{p}_{0}:\Sigma\to M_{0}.

We write e0HdR2(M0)e_{0}\in H^{2}_{{\mathrm{dR}}}(M_{0}) for minus the real Euler class of 𝔭0\mathfrak{p}_{0} and we have the equivalence

Cn0e0=0.C^{n}\neq 0\qquad\Longleftrightarrow\qquad e_{0}=0.

Let Λ𝔥(Σ)\Lambda_{\mathfrak{h}}(\Sigma) be the space of one-periodic curves in the class 𝔥\mathfrak{h}, and let Λ~𝔥(Σ)\widetilde{\Lambda}_{\mathfrak{h}}(\Sigma) be the space of homotopies of paths {γr}r[0,1]\{\gamma_{r}\}_{r\in[0,1]} inside Λ𝔥(Σ)\Lambda_{\mathfrak{h}}(\Sigma) such that γ0\gamma_{0} is some oriented 𝔭0\mathfrak{p}_{0}-fibre. The admissible homotopies are allowed to move γ0\gamma_{0} inside the set of oriented 𝔭0\mathfrak{p}_{0}-fibres but have to fix the periodic curve γ1\gamma_{1}, so that the natural projection Λ~𝔥(Σ)Λ𝔥(Σ)\widetilde{\Lambda}_{\mathfrak{h}}(\Sigma)\to\Lambda_{\mathfrak{h}}(\Sigma), [γr]γ1[\gamma_{r}]\mapsto\gamma_{1} is a covering map.

We pick a closed two-form ω0Ξc02(M0)\mathrm{\omega}_{0}\in\Xi^{2}_{c_{0}}(M_{0}) and choose Ω0:=𝔭0ω0ΞC2(Σ)\Omega_{0}:=\mathfrak{p}_{0}^{*}\mathrm{\omega}_{0}\in\Xi^{2}_{C}(\Sigma) as our auxiliary element for the computation of the volume. One sees that 𝔙𝔬𝔩:ΞC2(Σ){\mathfrak{Vol}}:\Xi^{2}_{C}(\Sigma)\to{\mathbb{R}} depends only on (𝔭0,c0)(\mathfrak{p}_{0},c_{0}) and not on ω0\mathrm{\omega}_{0}. We associate to Ω=Ω0+dα\Omega=\Omega_{0}+{\mathrm{d}}\alpha the action functional

𝒜~Ω:Λ~𝔥(Σ),𝒜~Ω([γr]):=[0,1]×S1ΓΩ+S1γ0α,\widetilde{\mathcal{A}}_{\Omega}:\widetilde{\Lambda}_{\mathfrak{h}}(\Sigma)\to{\mathbb{R}},\qquad\widetilde{\mathcal{A}}_{\Omega}\big{(}[\gamma_{r}]\big{)}:=\int_{[0,1]\times S^{1}}\Gamma^{*}\Omega+\int_{S^{1}}\gamma_{0}^{*}\alpha,

where α\alpha is chosen to be normalised if Cn0C^{n}\neq 0, and Γ:[0,1]×S1\Gamma:[0,1]\times S^{1} is the cylinder traced by the path {γr}\{\gamma_{r}\}. Like the volume, the functional 𝒜~Ω\widetilde{\mathcal{A}}_{\Omega} depends only on (𝔭0,c0)(\mathfrak{p}_{0},c_{0}) and not on ω0\mathrm{\omega}_{0}. If [γr][\gamma_{r}] is a critical point of 𝒜~Ω\widetilde{\mathcal{A}}_{\Omega}, then γ1𝒳(Ω)\gamma_{1}\in\mathcal{X}(\Omega), provided γ1\gamma_{1} is embedded. Furthermore, the action is invariant under isotopies {Ψr:ΣΣ}\{\Psi_{r}:\Sigma\to\Sigma\} starting at the identity (Proposition 6.10):

𝒜~Ψ1Ω([Ψr1γr])=𝒜~Ω([γr]),[γr]Λ~𝔥(Σ).\widetilde{\mathcal{A}}_{\Psi_{1}^{*}\Omega}\big{(}[\Psi^{-1}_{r}\circ\gamma_{r}]\big{)}=\widetilde{\mathcal{A}}_{\Omega}\big{(}[\gamma_{r}]\big{)},\qquad\forall\,[\gamma_{r}]\in\widetilde{\Lambda}_{\mathfrak{h}}(\Sigma).

In general, 𝒜~Ω\widetilde{\mathcal{A}}_{\Omega} does not descend to a functional on Λ𝔥(Σ)\Lambda_{\mathfrak{h}}(\Sigma). More precisely, by Lemma 6.3 this happens if and only if c0|π2(M0)=ae0|π2(M0)c_{0}|_{\pi_{2}(M_{0})}=ae_{0}|_{\pi_{2}(M_{0})} for some aa\in{\mathbb{R}}. However, as we see now, we can define an action on the set of Zoll forms Ω\Omega with (𝔭Ω,[ωΩ])C0(Σ)(\mathfrak{p}_{\Omega},[\mathrm{\omega}_{\Omega}])\in\mathfrak{Z}_{C}^{0}(\Sigma). In this case, the volume of Ω\Omega can be expressed as a polynomial function of the action.

Definition 1.13.

The Zoll polynomial P:P:{\mathbb{R}}\to{\mathbb{R}} associated with (𝔭0,c0)(\mathfrak{p}_{0},c_{0}) is given by

P(0)=0,dPdA(A)=(Ae0+c0)n,[M0].P(0)=0,\qquad\frac{{\mathrm{d}}P}{{\mathrm{d}}A}(A)=\langle(Ae_{0}+c_{0})^{n},[M_{0}]\rangle.

For instance, when n=1n=1, the polynomial reads

P(A)=12e0,[M0]A2+c0,[M0]A.P(A)=\frac{1}{2}\langle e_{0},[M_{0}]\rangle A^{2}+\langle c_{0},[M_{0}]\rangle A.
Theorem 1.14.

There is a well-defined volume function

𝔙𝔬𝔩:C(Σ),𝔙𝔬𝔩(𝔭,c):=𝔙𝔬𝔩(𝔭ω){\mathfrak{Vol}}:\mathfrak{Z}_{C}(\Sigma)\to{\mathbb{R}},\qquad{\mathfrak{Vol}}(\mathfrak{p},c):={\mathfrak{Vol}}(\mathfrak{p}^{*}\mathrm{\omega})

and well-defined action functional

𝒜:C0(Σ),𝒜(𝔭,c):=𝒜~𝔭ω([δr]).\mathcal{A}:\mathfrak{Z}_{C}^{0}(\Sigma)\to{\mathbb{R}},\qquad\mathcal{A}(\mathfrak{p},c):=\widetilde{\mathcal{A}}_{\mathfrak{p}^{*}\mathrm{\omega}}\big{(}[\delta_{r}]\big{)}.

Here, ω\mathrm{\omega} is any closed two-form on MM in the class cc, {δr}\{\delta_{r}\} is any path of periodic curves such that δr\delta_{r} is an oriented 𝔭r\mathfrak{p}_{r}-fibre, where {𝔭r}\{\mathfrak{p}_{r}\} is any path of oriented S1S^{1}-bundles from 𝔭0\mathfrak{p}_{0} to 𝔭1=𝔭\mathfrak{p}_{1}=\mathfrak{p}. Moreover, there holds

P(𝒜(𝔭,c))=𝔙𝔬𝔩(𝔭,c),(𝔭,c)C0(Σ).P\big{(}\mathcal{A}(\mathfrak{p},c)\big{)}={\mathfrak{Vol}}(\mathfrak{p},c),\qquad\forall\,(\mathfrak{p},c)\in\mathfrak{Z}_{C}^{0}(\Sigma).

If A:=𝒜(𝔭Ω,c)A_{*}:=\mathcal{A}(\mathfrak{p}_{\Omega_{*}},c_{*}), then dPdA(A)>0\tfrac{{\mathrm{d}}P}{{\mathrm{d}}A}(A_{*})>0. In particular, the polynomial PP is non-zero.

From this result, we can generalise the equality cases in Theorem 1.1 and Proposition 1.3.

Corollary 1.15.

Let Ω𝒵C(Σ)\Omega\in{\mathcal{Z}}_{C}(\Sigma) be a Zoll odd-symplectic form such that 𝔭Ω𝔓0(Σ)\mathfrak{p}_{\Omega}\in\mathfrak{P}^{0}(\Sigma). If we set 𝒜(Ω):=𝒜(𝔭Ω,[ωΩ])\mathcal{A}(\Omega):=\mathcal{A}(\mathfrak{p}_{\Omega},[\mathrm{\omega}_{\Omega}]), then

P(𝒜(Ω))=𝔙𝔬𝔩(Ω).P(\mathcal{A}(\Omega))={\mathfrak{Vol}}(\Omega).

In what follows, we describe a conjectural systolic-diastolic inequality for odd-symplectic forms close to Ω\Omega_{*} and with class CHdR2(Σ)C\in H^{2}_{{\mathrm{dR}}}(\Sigma). To this end, we fix a finite open covering {Bi}\{B_{i}\} of MM_{*} by balls so that all their pairwise intersections are also contractible. Let Λ(𝔭Ω)\Lambda(\mathfrak{p}_{\Omega_{*}}) be the space of periodic curves γΛ𝔥(Σ)\gamma\in\Lambda_{\mathfrak{h}}(\Sigma) with the property that 𝔭Ω(γ)\mathfrak{p}_{\Omega_{*}}(\gamma) is contained in some BiB_{i} and there is a path {γrshort}Λ~𝔥(Σ)\{\gamma_{r}^{\mathrm{short}}\}\in\widetilde{\Lambda}_{\mathfrak{h}}(\Sigma) entirely contained in 𝔭Ω1(Bi)\mathfrak{p}_{\Omega_{*}}^{-1}(B_{i}) with γ1short=γ\gamma^{\mathrm{short}}_{1}=\gamma. If ΩΞC2(Σ)\Omega\in\Xi^{2}_{C}(\Sigma), we set

𝒜Ω:Λ(𝔭Ω),𝒜Ω(γ):=𝒜~Ω([δr#γrshort]),γ1short=γ.\mathcal{A}_{\Omega}:\Lambda(\mathfrak{p}_{\Omega_{*}})\to{\mathbb{R}},\qquad\mathcal{A}_{\Omega}(\gamma):=\widetilde{\mathcal{A}}_{\Omega}\big{(}[\delta_{r}\#\gamma^{\mathrm{short}}_{r}]\big{)},\qquad\gamma^{\mathrm{short}}_{1}=\gamma.

Here, the symbol #\# denotes the concatenation of paths, {δr}\{\delta_{r}\} is any path of periodic curves such that δ1=γ0short\delta_{1}=\gamma_{0}^{\mathrm{short}}, and for every r[0,1]r\in[0,1], δr\delta_{r} is an oriented 𝔭r\mathfrak{p}_{r}-fibre, where {𝔭r}\{\mathfrak{p}_{r}\} is a path of oriented S1S^{1}-bundles connecting 𝔭0\mathfrak{p}_{0} with 𝔭1=𝔭Ω\mathfrak{p}_{1}=\mathfrak{p}_{\Omega_{*}}. We define

𝒜min(Ω):=infγ𝒳(Ω)Λ(𝔭Ω)𝒜Ω(γ),𝒜max(Ω):=supγ𝒳(Ω)Λ(𝔭Ω)𝒜Ω(γ).\mathcal{A}_{\min}(\Omega):=\inf_{\gamma\in\mathcal{X}(\Omega)\cap\Lambda(\mathfrak{p}_{\Omega_{*}})}\mathcal{A}_{\Omega}(\gamma),\qquad\mathcal{A}_{\max}(\Omega):=\sup_{\gamma\in\mathcal{X}(\Omega)\cap\Lambda(\mathfrak{p}_{\Omega_{*}})}\mathcal{A}_{\Omega}(\gamma).

By [Gin87, Section III] or [ÁPB14, Section 3.2], if Ω𝒮C(Σ)\Omega\in\mathcal{S}_{C}(\Sigma) is C1C^{1}-close to Ω\Omega_{*}, the set 𝒳(Ω)Λ(𝔭Ω)\mathcal{X}(\Omega)\cap\Lambda(\mathfrak{p}_{\Omega_{*}}) is compact and non-empty. Furthermore, the numbers 𝒜min(Ω)\mathcal{A}_{\min}(\Omega) and 𝒜max(Ω)\mathcal{A}_{\max}(\Omega) are finite and vary C1C^{1}-continuously with Ω\Omega.

Conjecture 1 (Local systolic-diastolic inequality for odd-symplectic forms).

Let Ω\Omega_{*} be a Zoll odd-symplectic form with cohomology class CHdR2(Σ)C\in H^{2}_{\mathrm{dR}}(\Sigma). There is a Ck1C^{k-1}-neighbourhood 𝒰\mathcal{U} of Ω\Omega_{*} in 𝒮C(Σ)\mathcal{S}_{C}(\Sigma) with k2k\geq 2 such that

P(𝒜min(Ω))𝔙𝔬𝔩(Ω)P(𝒜max(Ω)),Ω𝒰.P(\mathcal{A}_{\min}(\Omega))\leq{\mathfrak{Vol}}(\Omega)\leq P(\mathcal{A}_{\max}(\Omega)),\qquad\forall\,\Omega\in\mathcal{U}.

The equality holds in any of the two inequalities above, if and only if Ω\Omega is Zoll.

Remark 1.16.

If the real Euler class of the bundle associated with Ω\Omega_{*} vanishes, then the inequality in the conjecture is equivalent to

𝒜min(Ω)0𝒜max(Ω),Ω𝒰,\mathcal{A}_{\min}(\Omega)\leq 0\leq\mathcal{A}_{\max}(\Omega),\qquad\forall\,\Omega\in\mathcal{U},

with any of the equalities holding if and only if Ω\Omega is Zoll.

Remark 1.17.

The volume, the action and the Zoll polynomial depend on the choice of reference pair (𝔭0,c0)C0(Σ)(\mathfrak{p}_{0},c_{0})\in\mathfrak{Z}_{C}^{0}(\Sigma). However, thanks to Theorem 1.14 we will observe in Remark 7.7 that the functional

(P𝒜Ω𝔙𝔬𝔩(Ω)):Λ(𝔭Ω)\big{(}P\circ\mathcal{A}_{\Omega}-{\mathfrak{Vol}}(\Omega)\big{)}:\Lambda(\mathfrak{p}_{\Omega_{*}})\to{\mathbb{R}}

is independent of such a choice. Hence, Conjecture 1 remains the same when we use another reference pair in C0(Σ)\mathfrak{Z}_{C}^{0}(\Sigma).

Remark 1.18.

As mentioned at the beginning of this subsection, Conjecture 1 recovers the contact and symplectic systolic-diastolic inequality. In the contact case, we have Ω=dα\Omega_{*}={\mathrm{d}}\alpha_{*} for a contact form α\alpha_{*} so that C=0C=0. Up to multiplication by a positive constant, we can assume that T(α)=1T(\alpha_{*})=1. We take 𝔬Σ=𝔬α\mathfrak{o}_{\Sigma}=\mathfrak{o}_{\alpha_{*}} and c0=0c_{0}=0. There holds

P(A)=1n+1e0n,[M0]An+1,P(A)=\frac{1}{n+1}\langle e_{0}^{n},[M_{0}]\rangle A^{n+1},

where e0n,[M0]=[ωn],[M]>0\langle e_{0}^{n},[M_{0}]\rangle=\langle[\mathrm{\omega}_{*}^{n}],[M_{*}]\rangle>0. When n=1n=1, we have e0,[M0]=|H1tor(Σ;)|\langle e_{0},[M_{0}]\rangle=|H_{1}^{\mathrm{tor}}(\Sigma;{\mathbb{Z}})|. If Ω=dα\Omega={\mathrm{d}}\alpha is C0C^{0}-close to Ω\Omega_{*}, we can take α\alpha to be a contact form. We have

𝔙𝔬𝔩(dα)=1nVolume(α),𝒜dα(γ)=S1γα=T(γ),γΛ𝔥(Σ).{\mathfrak{Vol}}({\mathrm{d}}\alpha)=\frac{1}{n}\mathrm{Volume}(\alpha),\qquad\mathcal{A}_{{\mathrm{d}}\alpha}(\gamma)=\int_{S^{1}}\gamma^{*}\alpha=T(\gamma),\quad\forall\,\gamma\in\Lambda_{\mathfrak{h}}(\Sigma).

Thus, there holds Tmin(α,𝔥)𝒜min(dα)𝒜max(dα)Tmax(α,𝔥)T_{\min}(\alpha,\mathfrak{h})\leq\mathcal{A}_{\min}({\mathrm{d}}\alpha)\leq\mathcal{A}_{\max}({\mathrm{d}}\alpha)\leq T_{\max}(\alpha,\mathfrak{h}), and Conjecture 1 implies

Tmin(α,𝔥)n1e0n,[M0]Volume(α)Tmax(α,𝔥)nT_{\min}(\alpha,\mathfrak{h})^{n}\leq\frac{1}{\langle e_{0}^{n},[M_{0}]\rangle}\mathrm{Volume}(\alpha)\leq T_{\max}(\alpha,\mathfrak{h})^{n}

with equality cases exactly when the contact form α\alpha is Zoll.

In the symplectic case, 𝔭Ω\mathfrak{p}_{\Omega_{*}} is trivial so that Σ=M×S1\Sigma=M_{*}\times S^{1} and we call tS1t\in S^{1} the global angular coordinate. We take Ω=Ω+d(Hdt)\Omega=\Omega_{*}+{\mathrm{d}}(H{\mathrm{d}}t) for some function H:M×S1H:M_{*}\times S^{1}\to{\mathbb{R}}. The characteristic distribution of Ω\Omega is generated by the vector field t+XHt\partial_{t}+X_{H_{t}}, where XHtX_{H_{t}} is the ω\omega_{*}-Hamiltonian vector field of Ht:=H(,t)H_{t}:=H(\cdot,t) tangent to M×{t}M_{*}\times\{t\}. Then, 𝒜Ω\mathcal{A}_{\Omega} is the Hamiltonian action defined in (1.2), c0n,[M0]=[ωn],[M]>0\langle c_{0}^{n},[M_{0}]\rangle=\langle[\mathrm{\omega}_{*}^{n}],[M_{*}]\rangle>0, and

P(A)=c0n,[M0]A,Vol(Hdt)=M×S1(Hdt)ωn=CALω(H).P(A)=\langle c_{0}^{n},[M_{0}]\rangle A,\qquad\mathrm{Vol}(H{\mathrm{d}}t)=\int_{M_{*}\times S^{1}}(H{\mathrm{d}}t)\wedge\omega_{*}^{n}=\mathrm{CAL}_{\mathrm{\omega}_{*}}(H).

Conjecture 1 generalises Proposition 1.3, since 𝒜min(Ω)=min𝒜H\mathcal{A}_{\min}(\Omega)=\min\mathcal{A}_{H} and 𝒜max(Ω)=max𝒜H\mathcal{A}_{\max}(\Omega)=\max\mathcal{A}_{H}, if we take the Hamiltonian HH to be normalised, namely CALω(H)=0\mathrm{CAL}_{\mathrm{\omega}_{*}}(H)=0.

Inspired by the Hamiltonian case, we will look first at a special class of ΩΞC2(Σ)\Omega\in\Xi^{2}_{C}(\Sigma) when studying Conjecture 1. Here, we assume without loss of generality that 𝔭0=𝔭Ω\mathfrak{p}_{0}=\mathfrak{p}_{\Omega_{*}}. We fix an S1S^{1}-connection form η\eta for the bundle 𝔭0\mathfrak{p}_{0}. This means that η\eta restricts to the angular form on each fibre and dη=𝔭0κ{\mathrm{d}}\eta=\mathfrak{p}_{0}^{*}\kappa for some κΞe02(M0)\kappa\in\Xi^{2}_{e_{0}}(M_{0}). We consequently choose the form ω0\mathrm{\omega}_{0} so that ω=ω0+Aκ\mathrm{\omega}_{*}=\mathrm{\omega}_{0}+A_{*}\kappa, where A=𝒜(Ω)A_{*}=\mathcal{A}(\Omega_{*}). We say that a form ΩΞC2(Σ)\Omega\in\Xi^{2}_{C}(\Sigma) is an H-form, if there exists a function H:ΣH:\Sigma\to{\mathbb{R}} such that

Ω=Ω0+d(Hη).\Omega=\Omega_{0}+{\mathrm{d}}(H\eta).

We call HH a defining Hamiltonian for Ω\Omega. If HH is C0C^{0}-close to AA_{*}, then Ω\Omega is odd-symplectic, and if HH is CkC^{k}-close to AA_{*}, then Ω\Omega is Ck1C^{k-1}-close to Ω\Omega_{*}. An H-form is called quasi-autonomous if there exist qmin,qmaxM0q_{\min},q_{\max}\in M_{0} and defining Hamiltonians HminH_{\min} and HmaxH_{\max} such that

minΣHmin=Hmin(z),z𝔭01(qmin),maxΣHmax=Hmax(z),z𝔭01(qmax).\min_{\Sigma}H_{\min}=H_{\min}(z),\quad\forall\,z\in\mathfrak{p}^{-1}_{0}(q_{\min}),\qquad\max_{\Sigma}H_{\max}=H_{\max}(z),\quad\forall\,z\in\mathfrak{p}^{-1}_{0}(q_{\max}).

We say that HH is quasi-autonomous, if the corresponding Ω\Omega is quasi-autonomous. We establish Conjecture 1 for quasi-autonomous forms close to a Zoll odd-symplectic form.

Proposition 1.19.

There exists a C2C^{2}-neighbourhood \mathcal{H} of the constant AA_{*} in the space of quasi-autonomous functions over Σ\Sigma such that if HH\in\mathcal{H} and Ω=Ω+d(Hη)\Omega=\Omega_{*}+{\mathrm{d}}(H\eta) then

P(𝒜min(Ω))𝔙𝔬𝔩(Ω)P(𝒜max(Ω)),P\big{(}\mathcal{A}_{\min}(\Omega)\big{)}\leq{\mathfrak{Vol}}(\Omega)\leq P\big{(}\mathcal{A}_{\max}(\Omega)\big{)},

and any of the two equalities holds if and only if Ω\Omega is Zoll. In the Zoll case, HH is invariant under the holonomy of η\eta. In particular, it is constant if e00e_{0}\neq 0.

Let 𝔭0:M0×S1M0\mathfrak{p}_{0}:M_{0}\times S^{1}\to M_{0} be the trivial bundle and take η=dt\eta={\mathrm{d}}t, where tS1t\in S^{1} is the angular coordinate. In this case, we show in Proposition 3.4 that if ΩΞC2(Σ)\Omega\in\Xi^{2}_{C}(\Sigma) is C2C^{2}-close to Ω\Omega_{*}, then Ω\Omega is an H-form with a C2C^{2}-small defining Hamiltonian HH, after applying a diffeomorphism of Σ\Sigma isotopic to the identity. Therefore, in this setting the conjecture follows from Proposition 1.3 (or Proposition 1.19) and the invariance of the volume and the action under diffeomorphisms. From this result and the topological Lemma 4.6, Conjecture 1 for bundles with e0=0e_{0}=0 can readily be proven.

Theorem 1.20.

The local systolic-diastolic inequality for odd-symplectic forms, whose associated bundle has vanishing real Euler class, holds true in the C2C^{2}-topology, .

If Ω\Omega_{*} is any Zoll odd-symplectic form on a closed three-manifold Σ\Sigma, then either the Euler class of its associated bundle vanishes or Ω=dα\Omega_{*}={\mathrm{d}}\alpha_{*} for some Zoll contact form α\alpha_{*}. Hence, Theorem 1.20 together with [BK19a, Theorem 1.4] establish Conjecture 1 in dimension three.

Corollary 1.21.

The local systolic-diastolic inequality for odd-symplectic forms holds true in the C2C^{2}-topology on closed three-manifolds.∎

Remark 1.22.

More generally, one could try to formulate a systolic-diastolic inequality in a neighbourhood of an odd-symplectic form Ω\Omega_{*}, whose closed characteristics are tangent to a (not necessarily free) S1S^{1}-action on Σ\Sigma, cf. [Tho76] and [BR94].

The example of magnetic geodesics

Our main motivation to study the systolic-diastolic inequality for odd-symplectic forms comes from twisted cotangent bundles, where the magnetic form is symplectic. Let (N,σ)(N,\sigma) be a connected closed symplectic manifold of dimension 2m2m. Let 𝔭TN:TNN\mathfrak{p}_{{\mathrm{T}}^{*}N}:{\mathrm{T}}^{*}N\to N be the cotangent bundle map and consider the twisted symplectic form

dλcan+𝔭TNσΩ2(TN),{\mathrm{d}}\lambda_{\mathrm{can}}+\mathfrak{p}_{{\mathrm{T}}^{*}N}^{*}\sigma\in\Omega^{2}({\mathrm{T}}^{*}N),

where λcan\lambda_{\mathrm{can}} is the canonical one-form on TN{\mathrm{T}}^{*}N. We fix a σ\sigma-compatible almost complex structure JJ on NN with associated metric gJg_{J}. The structure JJ turns 𝔭TN\mathfrak{p}_{{\mathrm{T}}^{*}N} into a complex vector bundle and we denote by 𝔭J:(TN)N\mathfrak{p}_{J}:\mathbb{P}_{\mathbb{C}}({\mathrm{T}}^{*}N)\to N its projectivisation. For all Riemannian metrics gg on NN in the same conformal class of gJg_{J}, let SgNS_{g}^{*}N be the unit co-sphere bundle of gg, and write 𝔭g:SgNN\mathfrak{p}_{g}:S^{*}_{g}N\to N and 𝔦g:SgNTN\mathfrak{i}_{g}:S_{g}^{*}N\hookrightarrow{\mathrm{T}}^{*}N for the natural projection and inclusion. The one-parameter group tetJt\mapsto e^{tJ} acts fibrewise on TN{\mathrm{T}}^{*}N and yields a free S1S^{1}-action on SgNS^{*}_{g}N, since JJ is gg-orthogonal. The quotient is naturally identified with (TN)\mathbb{P}_{\mathbb{C}}({\mathrm{T}}^{*}N), so that we have an oriented S1S^{1}-bundle 𝔭\mathfrak{p} making the following diagram commute

TN\textstyle{{\mathrm{T}}^{*}N\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔭TN\scriptstyle{\mathfrak{p}_{{\mathrm{T}}^{*}N}}SgN\textstyle{S^{*}_{g}N\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔭\scriptstyle{\mathfrak{p}}𝔭g\scriptstyle{\mathfrak{p}_{g}}𝔦g\scriptstyle{\ \,\mathfrak{i}_{g}}(TN).\textstyle{\mathbb{P}_{\mathbb{C}}({\mathrm{T}}^{*}N)\ignorespaces\ignorespaces\ignorespaces\ignorespaces.}𝔭J\scriptstyle{\mathfrak{p}_{J}}N\textstyle{N}

Therefore, Ωg,σ:=𝔦g(dλcan+𝔭TNσ)\Omega_{g,\sigma}:=\mathfrak{i}_{g}^{*}({\mathrm{d}}\lambda_{\mathrm{can}}+\mathfrak{p}_{{\mathrm{T}}^{*}N}^{*}\sigma) is an odd-symplectic two-form on SgNS^{*}_{g}N with cohomology class C:=𝔭g[σ]C:=\mathfrak{p}_{g}^{*}[\sigma]. The class CC vanishes if and only if NN is a surface different from the two-torus. In general, it is well-known that there exists a Zoll form 𝔭ωσ\mathfrak{p}^{*}\mathrm{\omega}_{\sigma} in the class CC. More specifically, ωσ\mathrm{\omega}_{\sigma} is the symplectic form on (TN)\mathbb{P}_{\mathbb{C}}({\mathrm{T}}^{*}N) defined as

ωσ:=aωFS+𝔭Jσ,\mathrm{\omega}_{\sigma}:=a\,\mathrm{\omega}_{\mathrm{FS}}+\mathfrak{p}_{J}^{*}\sigma,

for some a>0a>0 small enough (see [Voi07, Proposition 3.18] when (N,σ,J)(N,\sigma,J) is Kähler). Here, 𝔭ωFS=dη\mathfrak{p}^{*}\mathrm{\omega}_{\mathrm{FS}}={\mathrm{d}}\eta, where η\eta is a connection form for 𝔭\mathfrak{p}, which, for all xNx\in N, restricts on the fibre 𝔭g1(x)S2m1\mathfrak{p}_{g}^{-1}(x)\cong S^{2m-1} to the standard contact form on the sphere. The 𝔭\mathfrak{p}-fibres are almost tangent to the characteristic distribution of Ωg,σ\Omega_{g,\sigma}, if σ\sigma is very big. However, the form Ωg,σ\Omega_{g,\sigma} and 𝔭ωσ\mathfrak{p}^{*}\mathrm{\omega}_{\sigma} are remarkably not close to each other, if m>1m>1.

The relevance of this example stems from the fact that the characteristics of Ωg,σ\Omega_{g,\sigma} are the tangent lifts of the magnetic geodesics of the pair (g,σ)(g,\sigma). A curve c:Nc:{\mathbb{R}}\to N is called a magnetic geodesic if g(c˙,c˙)1g(\dot{c},\dot{c})\equiv 1 and there holds

c˙gc˙=fJc˙,\nabla^{g}_{\dot{c}}\dot{c}=-fJ\dot{c},

where g\nabla^{g} is the Levi-Civita connection of gg and f:N(0,)f:N\to(0,\infty) is the conformal factor given by fg=gJf\cdot g=g_{J}. We refer to the work of Kerman in [Ker99] for results on the existence of closed magnetic geodesics in this setting.

In the companion paper [BK19b], we use Corollary 1.21 and Proposition 1.9 to establish a systolic-diastolic inequality for magnetic geodesics when NN is a surface, namely m=1m=1.

1.3 Structure of the paper

  • In Section 2, we introduce the volume of a closed two-form on an odd-dimensional oriented closed manifold and explore its invariance properties.

  • We define odd-symplectic forms and H-forms in Section 3. Under certain conditions, we prove a stability result for H-forms in the set of odd-symplectic forms.

  • We devote Section 4 to review some basic facts about oriented S1S^{1}-bundles, which will be crucially used in the following discussion.

  • In Section 5, we define weakly Zoll pairs and Zoll odd-symplectic forms. We also prove Proposition 1.9, which classifies Zoll odd-symplectic forms on oriented three-manifolds.

  • In Section 6, we introduce the action of a closed two-form on an odd-dimensional manifold which is the total space of some oriented S1S^{1}-bundle. We prove Theorem 6.14, which is just a reformulation of Theorem 1.14 showing that the action and the volume of a weakly Zoll pair are related through the Zoll polynomial.

  • We formulate the local systolic-diastolic inequality for odd-symplectic forms in Section 7. We prove it for quasi-autonomous H-forms (Proposition 7.16 corresponding to Proposition 1.19) or when the real Euler class of the bundle vanishes (end of Subsection 7.3 corresponding to Theorem 1.20 and Proposition 1.3).


Acknowledgements. This work is part of a project in the Collaborative Research Center TRR 191 - Symplectic Structures in Geometry, Algebra and Dynamics funded by the DFG. It was initiated when the authors worked together at the University of Münster and partially carried out while J.K. was affiliated with the Ruhr-University Bochum. We thank Peter Albers, Kai Zehmisch, and the University of Münster for having provided an inspiring academic environment. We are also grateful to Michael Kapovich for the characterisation contained in Lemma 4.6. G.B. would like to express his gratitude to Hans-Bert Rademacher and the whole Differential Geometry group at the University of Leipzig. J.K. was supported by Samsung Science and Technology Foundation under Project Number SSTF-BA1801-01.

Notations

In the following discussion (Σ,𝔬Σ)(\Sigma,\mathfrak{o}_{\Sigma}) is a connected oriented closed manifold of dimension 2n+12n+1. We endow Σ\Sigma with an auxiliary Riemannian metric in order to measure distances and norms of various objects defined on Σ\Sigma. We denote by rr a real number in [0,1][0,1]. We will often consider paths r𝐨rr\mapsto\mathbf{o}_{r} with values in some target space 𝐎\mathbf{O}. We use the short-hand {𝐨r}𝐎\{\mathbf{o}_{r}\}\subset\mathbf{O} for such a path and [𝐨r][\mathbf{o}_{r}] for a homotopy class of paths with conveniently chosen boundary conditions. If {𝐨r}\{\mathbf{o}_{r}\} and {𝐨r}\{\mathbf{o}^{\prime}_{r}\} are two paths with 𝐨1=𝐨0\mathbf{o}_{1}=\mathbf{o}^{\prime}_{0}, we write {𝐨r}#{𝐨r}\{\mathbf{o}_{r}\}\#\{\mathbf{o}_{r}^{\prime}\} for the concatenated path. Finally, if a mathematical object is expressed by a certain symbol, we add a prime to it in order to denote another object of the same kind.

2 The volume of a closed two-form

2.1 Definition of the volume

Let Ξk(Σ)\Xi^{k}(\Sigma) be the set of closed kk-forms on Σ\Sigma and ΞC2(Σ)\Xi_{C}^{2}(\Sigma) the set of elements in Ξ2(Σ)\Xi^{2}(\Sigma) representing the de Rham cohomology class CHdR2(Σ)C\in H_{\mathrm{dR}}^{2}(\Sigma). The set ΞC2(Σ)\Xi^{2}_{C}(\Sigma) is affine with underlying vector space

Ω¯1(Σ):=Ω1(Σ)Ξ1(Σ).\bar{\Omega}^{1}(\Sigma):=\frac{\Omega^{1}(\Sigma)}{\Xi^{1}(\Sigma)}.

Indeed, for every ΩΞC2(Σ)\Omega\in\Xi^{2}_{C}(\Sigma), we have the surjective map

BΩ:Ω1(Σ)ΞC2(Σ),αΩ+dα.B_{\Omega}:\Omega^{1}(\Sigma)\longrightarrow\Xi^{2}_{C}(\Sigma),\qquad\alpha\mapsto\Omega+{\mathrm{d}}\alpha.

This map passes to the quotient under Ξ1(Σ)\Xi^{1}(\Sigma) and yields a bijection

B¯Ω:Ω¯1(Σ)ΞC2(Σ)\bar{B}_{\Omega}:\bar{\Omega}^{1}(\Sigma)\longrightarrow\Xi^{2}_{C}(\Sigma)

so that we have an identification

TΩΞC2(Σ)=Ω¯1(Σ).\mathrm{T}_{\Omega}\Xi^{2}_{C}(\Sigma)=\bar{\Omega}^{1}(\Sigma).

We define now an exact one-form volΩ1(Ω1(Σ))\mathrm{vol}\in\Omega^{1}(\Omega^{1}(\Sigma)). To this purpose, we fix a reference two-form Ω0ΞC2(Σ)\Omega_{0}\in\Xi^{2}_{C}(\Sigma), and for every αΩ1(Σ)\alpha\in\Omega^{1}(\Sigma), we use the short-hand

Ωα:=BΩ0(α)=Ω0+dα.\Omega_{\alpha}:=B_{\Omega_{0}}(\alpha)=\Omega_{0}+{\mathrm{d}}\alpha.

For every βΩ1(Σ)TαΩ1(Σ)\beta\in\Omega^{1}(\Sigma)\cong{\mathrm{T}}_{\alpha}\Omega^{1}(\Sigma), we set

volαβ:=ΣβΩαn.\mathrm{vol}_{\alpha}\cdot\beta:=\int_{\Sigma}\beta\wedge\Omega_{\alpha}^{n}.

In the lemma below, we show that the one-form vol\mathrm{vol} admits the primitive functional

Vol:Ω1(Σ),Vol(α):=01(ΣαΩrαn)dr.\mathrm{Vol}:\Omega^{1}(\Sigma)\to{\mathbb{R}}\,,\qquad\mathrm{Vol}(\alpha):=\int_{0}^{1}\left(\int_{\Sigma}\alpha\wedge\Omega_{r\alpha}^{n}\right){\mathrm{d}}r. (2.1)
Remark 2.1.

Exchanging the order of integration, we can rewrite

Vol(α)=ΣCS(α),CS(α):=01αΩrαndr.\mathrm{Vol}(\alpha)=\int_{\Sigma}\mathrm{CS}(\alpha),\qquad\mathrm{CS}(\alpha):=\int_{0}^{1}\alpha\wedge\Omega^{n}_{r\alpha}\,{\mathrm{d}}r.

When C=0C=0 and Ω0=0\Omega_{0}=0, the form (n+1)CS(α)(n+1)\mathrm{CS}(\alpha) reduces to the Chern-Simons form for principal S1S^{1}-bundles with invariant polynomial P(x1,,xn+1)=x1xn+1P(x_{1},\ldots,x_{n+1})=x_{1}\cdot\ldots\cdot x_{n+1} [CS74, equation (3.1)]. Furthermore, (n+1)Vol(α)(n+1)\mathrm{Vol}(\alpha) corresponds to the cohomology class in [CS74, Theorem 3.9, case 2l1=n2l-1=n] (where MM therein is our Σ\Sigma and ll is our n+1n+1).

Lemma 2.2.

The functional Vol\mathrm{Vol} associated with Ω0\Omega_{0} is uniquely characterised by the properties

Vol(0)=0,dVol=vol.\mathrm{Vol}(0)=0,\qquad{\mathrm{d}}\mathrm{Vol}=\mathrm{vol}.

The following formula holds (see also [CS74, equation (3.5)]):

Vol(α)=j=0n1j+1(nj)Σα(dα)jΩ0nj.\mathrm{Vol}(\alpha)\ =\ \sum_{j=0}^{n}\frac{1}{j+1}\binom{n}{j}\int_{\Sigma}\alpha\wedge({\mathrm{d}}\alpha)^{j}\wedge\Omega_{0}^{n-j}.
Proof.

Differentiating under the integral sign we write dαVol(β)=01V(r)dr{\mathrm{d}}_{\alpha}\mathrm{Vol}(\beta)=\int_{0}^{1}V(r){\mathrm{d}}r and compute

V(r)\displaystyle V(r) =Σ(βΩrαn+αΩrαn1nrdβ)\displaystyle=\int_{\Sigma}\Big{(}\beta\wedge\Omega_{r\alpha}^{n}+\alpha\wedge\Omega_{r\alpha}^{n-1}\wedge nr{\mathrm{d}}\beta\Big{)}
=Σ(βΩrαn+nrΩrαn1(βdαd(αβ)))\displaystyle=\int_{\Sigma}\Big{(}\beta\wedge\Omega_{r\alpha}^{n}+nr\Omega_{r\alpha}^{n-1}\wedge\big{(}\beta\wedge{\mathrm{d}}\alpha-{\mathrm{d}}(\alpha\wedge\beta)\big{)}\Big{)}
=Σ(βΩrαn+rβΩrαn1(ndα))\displaystyle=\int_{\Sigma}\Big{(}\beta\wedge\Omega_{r\alpha}^{n}+r\beta\wedge\Omega_{r\alpha}^{n-1}\wedge\big{(}n{\mathrm{d}}\alpha\big{)}\Big{)}
=ddr(ΣrβΩrαn),\displaystyle=\frac{{\mathrm{d}}}{{\mathrm{d}}r}\left(\int_{\Sigma}r\beta\wedge\Omega_{r\alpha}^{n}\right),

where in the third equality we used Stokes’ Theorem. We conclude that

dαVolβ=01ddr(ΣrβΩrαn)dr=Σ1βΩ1αnΣ0βΩ0αn=volαβ.{\mathrm{d}}_{\alpha}\mathrm{Vol}\cdot\beta=\int_{0}^{1}\frac{{\mathrm{d}}}{{\mathrm{d}}r}\left(\int_{\Sigma}r\beta\wedge\Omega_{r\alpha}^{n}\right){\mathrm{d}}r=\int_{\Sigma}1\cdot\beta\wedge\Omega_{1\cdot\alpha}^{n}-\int_{\Sigma}0\cdot\beta\wedge\Omega_{0\cdot\alpha}^{n}=\mathrm{vol}_{\alpha}\cdot\beta.

The formula for Vol\mathrm{Vol} follows by expanding the binomial Ωrαn=(Ω0+rdα)n\Omega_{r\alpha}^{n}=\big{(}\Omega_{0}+r{\mathrm{d}}\alpha\big{)}^{n} in (2.1) and integrating each term. ∎

Remark 2.3.

If dimΣ=3\dim\Sigma=3, namely n=1n=1, the formula for Vol\mathrm{Vol} turns into

Vol(α)=Σα(Ω0+12dα).\mathrm{Vol}(\alpha)=\int_{\Sigma}\alpha\wedge\big{(}\Omega_{0}+\tfrac{1}{2}{\mathrm{d}}\alpha\big{)}.

Let us specify the dependence of Vol\mathrm{Vol} on the reference form Ω0\Omega_{0}.

Lemma 2.4.

Let Vol:Ω1(Σ)\mathrm{Vol}^{\prime}:\Omega^{1}(\Sigma)\to{\mathbb{R}} be the volume functional associated with another reference two-form Ω0ΞC2(Σ)\Omega_{0}^{\prime}\in\Xi^{2}_{C}(\Sigma). If αΩ1(Σ)\alpha^{\prime}\in\Omega^{1}(\Sigma) is such that Ω0=Ω0+dα\Omega_{0}^{\prime}=\Omega_{0}+{\mathrm{d}}\alpha^{\prime}, then

Vol(α)=Vol(α)+Vol(αα),αΩ1(Σ).\mathrm{Vol}(\alpha)=\mathrm{Vol}(\alpha^{\prime})+\mathrm{Vol}^{\prime}(\alpha-\alpha^{\prime}),\qquad\forall\,\alpha\in\Omega^{1}(\Sigma).
Proof.

For every α,βΩ1(Σ)\alpha,\beta\in\Omega^{1}(\Sigma), we have

volαα1β=Σβ(Ω0+dαdα)n=Σβ(Ω0+dα)n=volαβ,\mathrm{vol}^{\prime}_{\alpha-\alpha_{1}}\cdot\beta=\int_{\Sigma}\beta\wedge(\Omega_{0}^{\prime}+{\mathrm{d}}\alpha-{\mathrm{d}}\alpha^{\prime})^{n}=\int_{\Sigma}\beta\wedge(\Omega_{0}+{\mathrm{d}}\alpha)^{n}=\mathrm{vol}_{\alpha}\cdot\beta,

where vol:=dVol\mathrm{vol}^{\prime}:={\mathrm{d}}\mathrm{Vol}^{\prime}. Therefore, from the fundamental theorem of calculus, we get

Vol(α)Vol(α)=01volr(αα)+α(αα)dr\displaystyle\mathrm{Vol}(\alpha)-\mathrm{Vol}(\alpha^{\prime})=\int_{0}^{1}\mathrm{vol}_{r(\alpha-\alpha^{\prime})+\alpha^{\prime}}\cdot(\alpha-\alpha^{\prime}){\mathrm{d}}r =01volr(αα)(αα)dr\displaystyle=\int_{0}^{1}\mathrm{vol}^{\prime}_{r(\alpha-\alpha^{\prime})}\cdot(\alpha-\alpha^{\prime}){\mathrm{d}}r
=Vol(αα).\displaystyle=\mathrm{Vol}^{\prime}(\alpha-\alpha^{\prime}).\qed

We now wish to use the map BΩ0B_{\Omega_{0}} to push the volume form and functional to the space of closed two-forms in the class CC. To this purpose, we observe that Vol\mathrm{Vol} behaves linearly under the addition of closed one-forms. More precisely, for all αΩ1(Σ)\alpha\in\Omega^{1}(\Sigma) we have

Vol(α+τ)Vol(α)=Vol(τ)=[τ]Cn,[Σ],τΞ1(Σ),\begin{split}\mathrm{Vol}(\alpha+\tau)-\mathrm{Vol}(\alpha)=\mathrm{Vol}(\tau)=\langle[\tau]\cup C^{n},[\Sigma]\rangle,\quad\forall\,\tau\in\Xi^{1}(\Sigma),\end{split} (2.2)

and similarly,

volατ=[τ]Cn,[Σ],τΞ1(Σ).\begin{split}\mathrm{vol}_{\alpha}\cdot\tau=\langle[\tau]\cup C^{n},[\Sigma]\rangle,\quad\forall\,\tau\in\Xi^{1}(\Sigma).\end{split} (2.3)

These formulae suggest to treat two separate cases.

Case 1: Cn=0C^{n}=0.

The volume functional Vol:Ω1(Σ)\mathrm{Vol}:\Omega^{1}(\Sigma)\to{\mathbb{R}} passes to the quotient by Ξ1(Σ)\Xi^{1}(\Sigma) according to (2.2):

𝔙𝔬𝔩:ΞC2(Σ),𝔙𝔬𝔩(Ω0)=0.{\mathfrak{Vol}}:\Xi^{2}_{C}(\Sigma)\to{\mathbb{R}},\qquad{\mathfrak{Vol}}(\Omega_{0})=0. (2.4)

Following Lemma 2.4, if 𝔙𝔬𝔩{\mathfrak{Vol}}^{\prime} denotes the functional obtained from Ω0\Omega_{0}^{\prime}, then there holds

𝔙𝔬𝔩(Ω)=𝔙𝔬𝔩(Ω0)+𝔙𝔬𝔩(Ω),ΩΞC2(Σ).{\mathfrak{Vol}}(\Omega)={\mathfrak{Vol}}(\Omega_{0}^{\prime})+{\mathfrak{Vol}}^{\prime}(\Omega),\qquad\forall\,\Omega\in\Xi^{2}_{C}(\Sigma).

This means that volΩ1(Ω1(Σ))\mathrm{vol}\in\Omega^{1}(\Omega^{1}(\Sigma)) descends to a one-form

𝔳𝔬𝔩Ω1(ΞC2(Σ)),d𝔙𝔬𝔩=𝔳𝔬𝔩,{\mathfrak{vol}}\in\Omega^{1}(\Xi^{2}_{C}(\Sigma)),\qquad{\mathrm{d}}{\mathfrak{Vol}}={\mathfrak{vol}},

which is independent of the reference two-form Ω0\Omega_{0}.

Case 2: Cn0C^{n}\neq 0.

The functional Vol\mathrm{Vol} does not descend to ΞC2(Σ)\Xi^{2}_{C}(\Sigma). Indeed, by Poincaré duality, there exists a form τΞ1(Σ)\tau\in\Xi^{1}(\Sigma) such that

[τ]Cn0.[\tau]\cup C^{n}\neq 0. (2.5)

However, we can use the volume function to define a distinguished class of one-forms.

Definition 2.5.

We say that αΩ1(Σ)\alpha\in\Omega^{1}(\Sigma) is a normalised one-form, if the implication

Cn0Vol(α)=0C^{n}\neq 0\quad\Longrightarrow\quad\mathrm{Vol}(\alpha)=0

holds true (in particular, when Cn=0C^{n}=0 as in Case 1, all one-forms are normalised).

The inclusion Vol1(0)Ω1(Σ)\mathrm{Vol}^{-1}(0)\subset\Omega^{1}(\Sigma) induces a bijection

Vol1(0)Ω¯1(Σ)B¯Ω0ΞC2(Σ),\frac{\mathrm{Vol}^{-1}(0)}{\sim}\longrightarrow\bar{\Omega}^{1}(\Sigma)\stackrel{{\scriptstyle\bar{B}_{\Omega_{0}}}}{{\cong}}\Xi^{2}_{C}(\Sigma), (2.6)

where αα′′\alpha^{\prime}\sim\alpha^{\prime\prime} if and only if α′′αΞ1(Σ)\alpha^{\prime\prime}-\alpha^{\prime}\in\Xi^{1}(\Sigma). Indeed, by (2.2) any class α¯:=α+Ξ1(Σ)Ω¯1(Σ)\bar{\alpha}:=\alpha+\Xi^{1}(\Sigma)\in\bar{\Omega}^{1}(\Sigma) has a normalised representative α=α+sτ\alpha^{\prime}=\alpha+s\tau, for some ss\in{\mathbb{R}}. We also observe that, if α′′\alpha^{\prime\prime} is another normalised representative of α¯\bar{\alpha}, then, again by (2.2), we have

[α′′α]Cn=0.[\alpha^{\prime\prime}-\alpha^{\prime}]\cup C^{n}=0. (2.7)

In view of (2.6), we define the volume form and functional on ΞC2(Σ)\Xi^{2}_{C}(\Sigma) trivially by setting

𝔙𝔬𝔩:ΞC2(Σ),𝔙𝔬𝔩:=0,𝔳𝔬𝔩Ω1(ΞC2(Σ)),𝔳𝔬𝔩:=0.{\mathfrak{Vol}}:\Xi^{2}_{C}(\Sigma)\to{\mathbb{R}},\quad{\mathfrak{Vol}}:=0,\qquad\qquad{\mathfrak{vol}}\in\Omega^{1}(\Xi^{2}_{C}(\Sigma)),\quad{\mathfrak{vol}}:=0.

If α\alpha is a normalised one-form representing α¯Ω¯1(Σ)\bar{\alpha}\in\bar{\Omega}^{1}(\Sigma), then the inclusion kervolαΩ1(Σ)\ker\mathrm{vol}_{\alpha}\subset\Omega^{1}(\Sigma) induces the surjection kervolαΩ¯1(Σ)Tα¯Ω¯1(Σ)\ker\mathrm{vol}_{\alpha}\to\bar{\Omega}^{1}(\Sigma)\cong{\mathrm{T}}_{\bar{\alpha}}\bar{\Omega}^{1}(\Sigma) due to the existence of τ\tau satisfying (2.5), and in turn the isomorphism

kervolαker(Ξ1(Σ)[]CnHdR2n+1(Σ))Tα¯Ω¯1(Σ)TΩαΞC2(Σ),\frac{\ker\mathrm{vol}_{\alpha}}{\ker\Big{(}\Xi^{1}(\Sigma)\xrightarrow{[\,\cdot\,]\cup C^{n}}H^{2n+1}_{\mathrm{dR}}(\Sigma)\Big{)}}\longrightarrow{\mathrm{T}}_{\bar{\alpha}}\bar{\Omega}^{1}(\Sigma)\cong{\mathrm{T}}_{\Omega_{\alpha}}\Xi^{2}_{C}(\Sigma),

thanks to (2.3) and (2.7).

Remark 2.6.

For contact forms and Hamiltonian systems the volume functional recovers the following well-known formulae.

  • (Contact forms) This is an instance of Case 1. Let Ω0=0\Omega_{0}=0 and αΩ1(Σ)\alpha\in\Omega^{1}(\Sigma) be a contact form. If we endow Σ\Sigma with the orientation 𝔬α\mathfrak{o}_{\alpha}, we recover the contact volume up to a constant factor:

    𝔙𝔬𝔩(dα)=Vol(α)=1n+1Σα(dα)n=1n+1Volume(α).{\mathfrak{Vol}}({\mathrm{d}}\alpha)=\mathrm{Vol}(\alpha)=\frac{1}{n+1}\int_{\Sigma}\alpha\wedge({\mathrm{d}}\alpha)^{n}=\frac{1}{n+1}\mathrm{Volume}(\alpha).
  • (Hamiltonian systems) This is an instance of Case 2. Let Σ=M×S1\Sigma=M\times S^{1} and Ω0=𝔭ω0\Omega_{0}=\mathfrak{p}^{*}\mathrm{\omega}_{0} for some symplectic form ω0\mathrm{\omega}_{0} on MM, where 𝔭:M×S1M\mathfrak{p}:M\times S^{1}\to M is the projection on the first factor. Let α=Hdϕ\alpha=H{\mathrm{d}}\phi, where H:M×S1H:M\times S^{1}\to{\mathbb{R}} and ϕ\phi is the coordinate on S1S^{1}. Then, the volume functional recovers the Calabi invariant of HH

    Vol(Hdϕ)=M×S1(Hdϕ)ω0n=CALω0(H)\mathrm{Vol}(H{\mathrm{d}}\phi)=\int_{M\times S^{1}}(H{\mathrm{d}}\phi)\wedge\mathrm{\omega}_{0}^{n}=\mathrm{CAL}_{\mathrm{\omega}_{0}}(H)

    and α\alpha is normalised if and only if the Calabi invariant of HH vanishes.

2.2 The volume is invariant under pull-back and isotopies

In this subsection, we prove two invariance results for the volume. Recall that Ω0\Omega_{0} is the fixed reference form in ΞC2(Σ)\Xi_{C}^{2}(\Sigma) for CHdR2(Σ)C\in H^{2}_{\mathrm{dR}}(\Sigma).

Proposition 2.7.

Let Σ\Sigma and Σ\Sigma^{\text{\tiny$\vee$}} be two closed oriented manifolds of dimension 2n+12n+1 and Π:ΣΣ\Pi:\Sigma^{\text{\tiny$\vee$}}\to\Sigma a map of degree degΠ\deg\Pi\in{\mathbb{Z}}. Let C:=ΠCHdR2(Σ)C^{\text{\tiny$\vee$}}:=\Pi^{*}C\in H^{2}_{\mathrm{dR}}(\Sigma^{\text{\tiny$\vee$}}) and Ω0:=ΠΩ0ΞC2(Σ)\Omega_{0}^{\text{\tiny$\vee$}}:=\Pi^{*}\Omega_{0}\in\Xi^{2}_{C^{\text{\tiny$\vee$}}}(\Sigma^{\text{\tiny$\vee$}}). If Vol,𝔙𝔬𝔩\mathrm{Vol},\,{\mathfrak{Vol}} are the volumes associated with Ω0\Omega_{0} and Vol,𝔙𝔬𝔩\mathrm{Vol}^{\text{\tiny$\vee$}},\,{\mathfrak{Vol}}^{\text{\tiny$\vee$}} the volumes associated with Ω0\Omega^{\text{\tiny$\vee$}}_{0}, there holds

VolΠ=(degΠ)Vol,𝔙𝔬𝔩Π=(degΠ)𝔙𝔬𝔩.\mathrm{Vol}^{\text{\tiny$\vee$}}\circ\Pi^{*}=(\deg\Pi)\cdot\mathrm{Vol},\qquad{\mathfrak{Vol}}^{\text{\tiny$\vee$}}\circ\Pi^{*}=(\deg\Pi)\cdot{\mathfrak{Vol}}.
Proof.

The statement follows immediately from the definition of the volume and the observation that for all top dimensional forms μ\mu on Σ\Sigma there holds

ΣΠμ=degΠΣμ.\int_{\Sigma^{\text{\tiny$\vee$}}}\Pi^{*}\mu=\deg\Pi\cdot\int_{\Sigma}\mu.\qed

For the second result, we need a little preparation. Let Ψ:ΣΣ\Psi:\Sigma\to\Sigma be a diffeomorphism isotopic to the identity idΣ\mathrm{id}_{\Sigma}. Since the pull-back Ψ\Psi^{*} of Ψ\Psi acts as the identity in cohomology, there is a normalised one-form θΩ1(Σ)\theta\in\Omega^{1}(\Sigma) (determined up to a normalised closed one-form) such that

dθ=ΨΩ0Ω0.{\mathrm{d}}\theta=\Psi^{*}\Omega_{0}-\Omega_{0}. (2.8)

We define

Ψ^θ:Ω1(Σ)Ω1(Σ),Ψ^θ(α):=θ+Ψα,\widehat{\Psi}^{*}_{\theta}:\Omega^{1}(\Sigma)\to\Omega^{1}(\Sigma),\qquad\widehat{\Psi}^{*}_{\theta}(\alpha):=\theta+\Psi^{*}\alpha,

so that the following diagram commutes:

Ω1(Σ)\textstyle{\Omega^{1}(\Sigma)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ψ^θ\scriptstyle{\widehat{\Psi}_{\theta}^{*}}BΩ0\scriptstyle{B_{\Omega_{0}}}Ω1(Σ).\textstyle{\Omega^{1}(\Sigma)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\,.}BΩ0\scriptstyle{B_{\Omega_{0}}}ΞC2(Σ)\textstyle{\Xi^{2}_{C}(\Sigma)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ψ\scriptstyle{\Psi^{*}}ΞC2(Σ)\textstyle{\Xi^{2}_{C}(\Sigma)} (2.9)

Let {Ψr}\{\Psi_{r}\} be an isotopy with Ψ0=idΣ\Psi_{0}=\mathrm{id}_{\Sigma} and Ψ1=Ψ\Psi_{1}=\Psi. This gives rise to a smooth family of one-forms {θr}\{\theta_{r}\} satisfying

θ0=0,dθr=ΨrΩ0Ω0,r[0,1].\theta_{0}=0,\qquad{\mathrm{d}}\theta_{r}=\Psi^{*}_{r}\Omega_{0}-\Omega_{0},\quad\forall\,r\in[0,1].

To construct such a family, we just take the time-dependent vector field XrX_{r} generating the isotopy {Ψr}\{\Psi_{r}\} and let {θr}\{\theta_{r}\} be the unique path such that

θ0=0,θ˙r=Ψr(ιXrΩ0).\theta_{0}=0,\qquad\dot{\theta}_{r}=\Psi_{r}^{*}(\iota_{X_{r}}\Omega_{0}).

To ease the notation, if αΩ1(Σ)\alpha\in\Omega^{1}(\Sigma), we use the short-hand

Ψ^r(α):=Ψ^θr(α)=θr+Ψrα.\widehat{\Psi}^{*}_{r}(\alpha):=\widehat{\Psi}_{{\theta_{r}}}^{*}(\alpha)=\theta_{r}+\Psi^{*}_{r}\alpha. (2.10)

Every θr\theta_{r} is normalised, i.e. Vol(θr)=0\mathrm{Vol}(\theta_{r})=0. Indeed, Vol(θ0)=Vol(0)=0\mathrm{Vol}(\theta_{0})=\mathrm{Vol}(0)=0 and

ddrVol(θr)=volθr(θ˙r)=ΣΨr(ιXrΩ0)(ΨrΩ0)n=Σ(ιXrΩ0)Ω0n=0.\frac{{\mathrm{d}}}{{\mathrm{d}}r}\mathrm{Vol}(\theta_{r})=\mathrm{vol}_{\theta_{r}}(\dot{\theta}_{r})=\int_{\Sigma}\Psi_{r}^{*}(\iota_{X_{r}}\Omega_{0})\wedge(\Psi^{*}_{r}\Omega_{0})^{n}=\int_{\Sigma}(\iota_{X_{r}}\Omega_{0})\wedge\Omega_{0}^{n}=0.

In particular, the form τ1:=θθ1\tau_{1}:=\theta-\theta_{1} is closed and normalised and as observed in (2.7), there holds

[τ1]Cn=0[\tau_{1}]\cup C^{n}=0

(note that this condition is automatically satisfied if Cn=0C^{n}=0).

Proposition 2.8.

If Ψ\Psi is a diffeomorphism isotopic to idΣ\mathrm{id}_{\Sigma} and θ\theta is any normalised one-form satisfying (2.8), there holds

VolΨ^θ=Vol.\mathrm{Vol}\circ\widehat{\Psi}^{*}_{\theta}=\mathrm{Vol}.

As a consequence, the set of normalised one-forms is Ψ^θ\widehat{\Psi}_{\theta}^{*}-invariant and we have

𝔙𝔬𝔩Ψ=𝔙𝔬𝔩.{\mathfrak{Vol}}\circ\Psi^{*}={\mathfrak{Vol}}.
Proof.

For any αΩ1(Σ)\alpha\in\Omega^{1}(\Sigma), we have

ddrΨ^r(α)=Ψr(ιXrΩ0)+Ψr(ιXrdα+d(α(Xr)))=Ψr(ιXrΩα+d(α(Xr))).\frac{{\mathrm{d}}}{{\mathrm{d}}r}\widehat{\Psi}^{*}_{r}(\alpha)=\Psi_{r}^{*}(\iota_{X_{r}}\Omega_{0})+\Psi_{r}^{*}\Big{(}\iota_{X_{r}}{\mathrm{d}}\alpha+{\mathrm{d}}\big{(}\alpha(X_{r})\big{)}\Big{)}=\Psi_{r}^{*}\Big{(}\iota_{X_{r}}\Omega_{\alpha}+{\mathrm{d}}\big{(}\alpha(X_{r})\big{)}\Big{)}.

Using this relation, we compute

ddrVol(Ψ^r(α))\displaystyle\frac{{\mathrm{d}}}{{\mathrm{d}}r}\mathrm{Vol}\big{(}\widehat{\Psi}^{*}_{r}(\alpha)\big{)} =volΨ^r(α)ddrΨ^r(α)\displaystyle=\mathrm{vol}_{\widehat{\Psi}^{*}_{r}(\alpha)}\cdot\frac{{\mathrm{d}}}{{\mathrm{d}}r}\widehat{\Psi}^{*}_{r}(\alpha)
=ΣΨr(ιXrΩα+d(α(Xr)))(ΨrΩα)n\displaystyle=\int_{\Sigma}\Psi_{r}^{*}\Big{(}\iota_{X_{r}}\Omega_{\alpha}+{\mathrm{d}}\big{(}\alpha(X_{r})\big{)}\Big{)}\wedge(\Psi_{r}^{*}\Omega_{\alpha})^{n}
=ΣιXrΩαΩαn+Σd(α(Xr))Ωαn\displaystyle=\int_{\Sigma}\iota_{X_{r}}\Omega_{\alpha}\wedge\Omega_{\alpha}^{n}+\int_{\Sigma}{\mathrm{d}}\big{(}\alpha(X_{r})\big{)}\wedge\Omega_{\alpha}^{n}
=0.\displaystyle=0.

We conclude from (2.2) and the relation [τ1]Cn=0[\tau_{1}]\cup C^{n}=0 that

Vol(Ψ^θ(α))=Vol(Ψ^1(α)+τ1)=Vol(Ψ^1(α))+[τ1]Cn,[Σ]=Vol(Ψ^0(α))=Vol(α).\mathrm{Vol}\big{(}\widehat{\Psi}^{*}_{\theta}(\alpha)\big{)}=\mathrm{Vol}\big{(}\widehat{\Psi}^{*}_{1}(\alpha)+\tau_{1}\big{)}=\mathrm{Vol}\big{(}\widehat{\Psi}^{*}_{1}(\alpha)\big{)}+\langle[\tau_{1}]\cup C^{n},[\Sigma]\rangle=\mathrm{Vol}\big{(}\widehat{\Psi}^{*}_{0}(\alpha)\big{)}=\mathrm{Vol}(\alpha).

This proves the first identity. The second identity follows from the first one and the commutation relation BΩ0Ψ^θ=ΨBΩ0B_{\Omega_{0}}\circ\widehat{\Psi}_{\theta}^{*}=\Psi^{*}\circ B_{\Omega_{0}} in (2.9). ∎

3 Odd-symplectic forms

3.1 A couple of definitions

For a closed two-form ΩΞ2(Σ)\Omega\in\Xi^{2}(\Sigma) on Σ\Sigma, we consider the (possibly singular) distribution kerΩΣ\ker\Omega\to\Sigma defined by

kerΩ:={(z,u)TΣ|Ωz(u,v)=0,vTzΣ}.\ker\Omega:=\big{\{}(z,u)\in{\mathrm{T}}\Sigma\ \big{|}\ \Omega_{z}(u,v)=0,\;\forall\,v\in{\mathrm{T}}_{z}\Sigma\big{\}}.

The distribution kerΩ\ker\Omega is naturally co-oriented by Ωn\Omega^{n} and we orient it combining the orientation on Σ\Sigma with such a co-orientation.

Definition 3.1.

We say that ΩΞ2(Σ)\Omega\in\Xi^{2}(\Sigma) is odd-symplectic if kerΩΣ\ker\Omega\to\Sigma is a one-dimensional distribution and denote the set of odd-symplectic forms by 𝒮(Σ)Ξ2(Σ)\mathcal{S}(\Sigma)\subset\Xi^{2}(\Sigma). If CHdR2(Σ)C\in H^{2}_{\mathrm{dR}}(\Sigma), we write 𝒮C(Σ):=𝒮(Σ)ΞC2(Σ)\mathcal{S}_{C}(\Sigma):=\mathcal{S}(\Sigma)\cap\Xi^{2}_{C}(\Sigma).

Remark 3.2.

Any of the following conditions gives an equivalent definition of ΩΞ2(Σ)\Omega\in\Xi^{2}(\Sigma) being odd-symplectic:

  • Ωn\Omega^{n} is nowhere vanishing.

  • There exists a one-form σ\sigma on Σ\Sigma such that σΩn\sigma\wedge\Omega^{n} is a volume form.

  • There exists a nowhere vanishing one-form σ\sigma such that Ω|kerσ\Omega|_{\ker\sigma} is a non-degenerate bilinear form on kerσ\ker\sigma.

Definition 3.3.

Given a reference form Ω0ΞC2(Σ)\Omega_{0}\in\Xi^{2}_{C}(\Sigma) and a pair of one-forms α0,σ0Ω1(Σ)\alpha_{0},\sigma_{0}\in\Omega^{1}(\Sigma), we can build an affine map

C(Σ)\displaystyle C^{\infty}(\Sigma) ΞC2(Σ),\displaystyle\longrightarrow\Xi^{2}_{C}(\Sigma),
H\displaystyle H Ωα0+Hσ0=Ω0+d(α0+Hσ0).\displaystyle\longmapsto\Omega_{\alpha_{0}+H\sigma_{0}}=\Omega_{0}+{\mathrm{d}}(\alpha_{0}+H\sigma_{0}).

We refer to the image of this map, as the set of H-forms (with respect to Ω0\Omega_{0}, α0\alpha_{0}, and σ0\sigma_{0}).

3.2 A stability result

As we will see in Section 7, given Ω0\Omega_{0}, α0\alpha_{0} and σ0\sigma_{0} as in Definition (3.3), it is extremely helpful to see, if an odd-symplectic form Ω1𝒮C(Σ)\Omega_{1}\in\mathcal{S}_{C}(\Sigma) admits a diffeomorphism Ψ:ΣΣ\Psi:\Sigma\to\Sigma isotopic to idΣ\mathrm{id}_{\Sigma} such that ΨΩ1\Psi^{*}\Omega_{1} is an H-form. We provide a criterion for such a diffeomorphism to exist.

Proposition 3.4.

Let {Ωr=Ω0+dαr}\{\Omega_{r}=\Omega_{0}+{\mathrm{d}}\alpha_{r}\} be a path in 𝒮C(Σ)\mathcal{S}_{C}(\Sigma). Let {σr}\{\sigma_{r}\} be a path in Ω1(Σ)\Omega^{1}(\Sigma) such that σrΩrn\sigma_{r}\wedge\Omega_{r}^{n} is a volume form on Σ\Sigma, which exists according to Remark 3.2. Let us denote by Ir:(kerσr)kerσrI_{r}:(\ker\sigma_{r})^{*}\to\ker\sigma_{r} the inverse of the map vιvΩαr|kerσrv\mapsto\iota_{v}\Omega_{\alpha_{r}}|_{\ker\sigma_{r}}. If

(σ˙rιIr(α˙r)dσr)|kerσr=0,\big{(}\dot{\sigma}_{r}-\iota_{I_{r}(\dot{\alpha}_{r})}{\mathrm{d}}\sigma_{r}\big{)}\big{|}_{\ker\sigma_{r}}=0, (3.1)

then there exist an isotopy {Ψr:ΣΣ}\{\Psi_{r}:\Sigma\to\Sigma\} with Ψ0=idΣ\Psi_{0}=\mathrm{id}_{\Sigma} generated by a time-dependent vector field XrkerσrX_{r}\in\ker\sigma_{r} and a path of functions {Hr:Σ}\{H_{r}:\Sigma\to{\mathbb{R}}\} such that, for every r[0,1]r\in[0,1],

Ψ^rαr=α0+Hrσ0+dKr,Kr:=0rαr(Xr)Ψrdr.\widehat{\Psi}_{r}^{*}\alpha_{r}=\alpha_{0}+H_{r}\sigma_{0}+{\mathrm{d}}K_{r},\qquad K_{r}:=\int_{0}^{r}\alpha_{r^{\prime}}(X_{r^{\prime}})\circ\Psi_{r^{\prime}}{\mathrm{d}}r^{\prime}.

Here Ψ^r\widehat{\Psi}_{r}^{*} is the map in (2.10). In particular, we have

Ψ1Ω1=Ω0+d(α0+H1σ0),Vol(α1)=Vol(α0+H1σ0).\Psi^{*}_{1}\Omega_{1}=\Omega_{0}+{\mathrm{d}}(\alpha_{0}+H_{1}\sigma_{0}),\qquad\mathrm{Vol}(\alpha_{1})=\mathrm{Vol}(\alpha_{0}+H_{1}\sigma_{0}).
Proof.

The argument is a refinement of the Gray stability theorem, see e.g. [Gei08, Theorem 2.2.2]. We introduce a vector field VrV_{r} on Σ\Sigma uniquely determined by the relations

ιVrΩαr=0,σr(Vr)=1.\iota_{V_{r}}\Omega_{\alpha_{r}}=0,\qquad\sigma_{r}(V_{r})=1.

We define a path of vector fields rXrr\mapsto X_{r} on Σ\Sigma satisfying the relations

Xrkerσr,(ιXrΩαr+α˙r)|kerσr=0,X_{r}\in\ker\sigma_{r},\qquad\big{(}\iota_{X_{r}}\Omega_{\alpha_{r}}+\dot{\alpha}_{r}\big{)}\big{|}_{\ker\sigma_{r}}=0, (3.2)

which exists and is unique by (i). Applying (ii), we also have

(ιXrdσr+σ˙r)|kerσr=0.\big{(}\iota_{X_{r}}{\mathrm{d}}\sigma_{r}+\dot{\sigma}_{r}\big{)}\big{|}_{\ker\sigma_{r}}=0. (3.3)

Let rΨrr\mapsto\Psi_{r} be the isotopy on Σ\Sigma obtained by integrating XrX_{r} and setting Ψ0=idΣ\Psi_{0}=\mathrm{id}_{\Sigma}. We construct an auxiliary path of functions rJrr\mapsto J_{r} through

J0=0,J˙r=(dσr(Xr,Vr)+σ˙r(Vr))Ψr.J_{0}=0,\qquad\dot{J}_{r}=\big{(}{\mathrm{d}}\sigma_{r}(X_{r},V_{r})+\dot{\sigma}_{r}(V_{r})\big{)}\circ\Psi_{r}.

Combining the last equation with (3.3), we arrive at

ιXrdσr+σ˙r=(J˙rΨr1)σr.\iota_{X_{r}}{\mathrm{d}}\sigma_{r}+\dot{\sigma}_{r}=(\dot{J}_{r}\circ\Psi_{r}^{-1})\sigma_{r}.

With this piece of information, we can compute

ddr(Ψrσr)=Ψr(ιXrdσr+σ˙r)=Ψr(J˙rΨr1σr)=J˙r(Ψrσr).\frac{{\mathrm{d}}}{{\mathrm{d}}r}\big{(}\Psi_{r}^{*}\sigma_{r}\big{)}=\Psi_{r}^{*}\big{(}\iota_{X_{r}}{\mathrm{d}}\sigma_{r}+\dot{\sigma}_{r}\big{)}=\Psi_{r}^{*}\big{(}\dot{J}_{r}\circ\Psi_{r}^{-1}\sigma_{r}\big{)}=\dot{J}_{r}\big{(}\Psi_{r}^{*}\sigma_{r}\big{)}.

Since reJrσ0r\mapsto e^{J_{r}}\sigma_{0} also satisfies the same ordinary differential equation and coincides with Ψrσr\Psi_{r}^{*}\sigma_{r} at r=0r=0, we see that

Ψrσr=eJrσ0.\Psi_{r}^{*}\sigma_{r}=e^{J_{r}}\sigma_{0}. (3.4)

Next we define the path of functions {Hr:Σ}\{H_{r}:\Sigma\to{\mathbb{R}}\} by

H0=0,H˙r=(α˙r(Vr)Ψr)eJr.H_{0}=0,\qquad\dot{H}_{r}=\big{(}\dot{\alpha}_{r}(V_{r})\circ\Psi_{r}\big{)}e^{J_{r}}. (3.5)

Conditions (3.2) and (3.5) together can be rephrased as

ιXrΩαr+α˙r=((H˙reJr)Ψr1)σr.\iota_{X_{r}}\Omega_{\alpha_{r}}+\dot{\alpha}_{r}=\big{(}(\dot{H}_{r}e^{-J_{r}})\circ\Psi_{r}^{-1}\big{)}\sigma_{r}. (3.6)

We are ready to prove the formula for Ψ^rαr\widehat{\Psi}^{*}_{r}\alpha_{r} in the statement of the proposition. The identity clearly holds for r=0r=0 and we just need to show that the rr-derivatives of both sides coincide for every rr:

ddr(Ψ^rαr)=Ψr(d(αr(Xr))ιXrdαr+α˙r)+θ˙r\displaystyle\frac{{\mathrm{d}}}{{\mathrm{d}}r}\big{(}\widehat{\Psi}_{r}^{*}\alpha_{r}\big{)}=\Psi_{r}^{*}\Big{(}{\mathrm{d}}\big{(}\alpha_{r}(X_{r})\big{)}\iota_{X_{r}}{\mathrm{d}}\alpha_{r}+\dot{\alpha}_{r}\Big{)}+\dot{\theta}_{r} =d(αr(Xr)Ψr)+Ψr(ιXrΩαr+α˙r)\displaystyle={\mathrm{d}}\big{(}\alpha_{r}(X_{r})\circ\Psi_{r}\big{)}+\Psi_{r}^{*}\big{(}\iota_{X_{r}}\Omega_{\alpha_{r}}+\dot{\alpha}_{r}\big{)}
=dK˙r+H˙reJrΨrσr\displaystyle={\mathrm{d}}\dot{K}_{r}+\dot{H}_{r}e^{-J_{r}}\Psi_{r}^{*}\sigma_{r}
=dK˙r+H˙rσ0,\displaystyle={\mathrm{d}}\dot{K}_{r}+\dot{H}_{r}\sigma_{0},

where we used that θ˙r=ΨrιXrΩ0\dot{\theta}_{r}=\Psi_{r}^{*}\iota_{X_{r}}\Omega_{0} and equations (3.4) and (3.6). Finally, Proposition 2.8 yields the identity Vol(α1)=Vol(α0+H1σ0)\mathrm{Vol}(\alpha_{1})=\mathrm{Vol}(\alpha_{0}+H_{1}\sigma_{0}). ∎

Remark 3.5.

The form σrΩrn\sigma_{r}\wedge\Omega_{r}^{n} is a volume form and condition (3.1) is satisfied in the following two extreme cases:

  1. (a)

    The form Ω0\Omega_{0} vanishes and αr=Trσr\alpha_{r}=T_{r}\sigma_{r} is a contact form for every rr, where {Tr}\{T_{r}\} is some path of real numbers. In this case, KrK_{r} and θr\theta_{r} vanish so that we have the usual Gray stability theorem

    Ψrαr=(1+1T0Hr)α0.\Psi^{*}_{r}\alpha_{r}=(1+\tfrac{1}{T_{0}}H_{r})\alpha_{0}.
  2. (b)

    The form σ0\sigma_{0} is closed, there holds σr=σ0\sigma_{r}=\sigma_{0} for all rr, and Ωr\Omega_{r} non-degenerate on kerσ0\ker\sigma_{0}.

Remark 3.6.

It would be interesting to find a condition not involving rr-derivatives implying (3.1). Such a condition can indeed be found in cases (a) or (b) in the remark above, where the odd-symplectic forms Ωr\Omega_{r} are actually stable according to [CM05, Section 2.1].

4 Oriented S1S^{1}-bundles

4.1 S1S^{1}-bundles and free S1S^{1}-actions

Let 𝔓(Σ)\mathfrak{P}(\Sigma) be the space of all oriented S1S^{1}-bundles having Σ\Sigma as total space, up to equivalence. Here we say that two bundles are equivalent if they have the same oriented fibres. If 𝔭:ΣM\mathfrak{p}:\Sigma\to M is an element of 𝔓(Σ)\mathfrak{P}(\Sigma), then MM is a closed manifold of dimension 2n2n. We write Ξ2(M)\Xi^{2}(M) for the space of closed two-forms on MM and Ξc2(M)\Xi^{2}_{c}(M) for the set of those ωΞ2(M)\mathrm{\omega}\in\Xi^{2}(M) with [ω]=c[\mathrm{\omega}]=c. We orient MM combining the orientation on Σ\Sigma with the orientation of the 𝔭\mathfrak{p}-fibres. We denote by 𝔭-\mathfrak{p} the bundle obtained from 𝔭\mathfrak{p} by reversing the orientation.

Definition 4.1.

If 𝔭\mathfrak{p} belongs to 𝔓(Σ)\mathfrak{P}(\Sigma), then 𝔭1(pt)Σ\mathfrak{p}^{-1}({\mathrm{pt}})\subset\Sigma denotes an oriented 𝔭\mathfrak{p}-fibre and [𝔭1(pt)]H1(Σ;)[\mathfrak{p}^{-1}({\mathrm{pt}})]_{\mathbb{Z}}\in H_{1}(\Sigma;{\mathbb{Z}}) its integral homology class.

Definition 4.2.

A bundle map between oriented S1S^{1}-bundles 𝔭:ΣM\mathfrak{p}:\Sigma\to M and 𝔭:ΣM\mathfrak{p}^{\text{\tiny$\vee$}}:\Sigma^{\text{\tiny$\vee$}}\to M^{\text{\tiny$\vee$}} is a map Π:ΣΣ\Pi:\Sigma^{\text{\tiny$\vee$}}\to\Sigma diffeomorphically sending every oriented fibre of 𝔭\mathfrak{p}^{\text{\tiny$\vee$}} to an oriented fibre of 𝔭\mathfrak{p}. In this case, we write 𝔭=Π𝔭\mathfrak{p}^{\text{\tiny$\vee$}}=\Pi^{*}\mathfrak{p}. A bundle map yields a quotient map π:MM\pi:M^{\text{\tiny$\vee$}}\to M between the base manifolds so that 𝔭Π=π𝔭\mathfrak{p}\circ\Pi=\pi\circ\mathfrak{p}^{\text{\tiny$\vee$}}. If Π\Pi is also a diffeomorphism, we say that Π\Pi is a bundle isomorphism. When Σ=Σ\Sigma^{\text{\tiny$\vee$}}=\Sigma, we write 𝔭=𝔭\mathfrak{p}^{\text{\tiny$\vee$}}=\mathfrak{p}^{\prime}, where 𝔭:ΣM\mathfrak{p}^{\prime}:\Sigma\to M^{\prime}, and denote by Ψ:ΣΣ\Psi:\Sigma\to\Sigma a bundle isomorphism between 𝔭\mathfrak{p} and 𝔭\mathfrak{p}^{\prime} with quotient map ψ:MM\psi:M^{\prime}\to M. We summarise the properties of Π\Pi and Ψ\Psi in two commutative diagrams.

Σ\textstyle{\Sigma^{\text{\tiny$\vee$}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔭\scriptstyle{\mathfrak{p}^{\text{\tiny$\vee$}}}Π\scriptstyle{\Pi}Σ,\textstyle{\Sigma\ignorespaces\ignorespaces\ignorespaces\ignorespaces\,,}𝔭\scriptstyle{\mathfrak{p}}M\textstyle{M^{\text{\tiny$\vee$}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π\scriptstyle{\pi}M\textstyle{M}    Σ\textstyle{\Sigma\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔭\scriptstyle{\mathfrak{p}^{\prime}}Ψ\scriptstyle{\Psi}\scriptstyle{\sim}Σ.\textstyle{\Sigma\ignorespaces\ignorespaces\ignorespaces\ignorespaces\,.}𝔭\scriptstyle{\mathfrak{p}}M\textstyle{M^{\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ\scriptstyle{\psi}\scriptstyle{\sim}M\textstyle{M} (4.1)

Let 𝔘(Σ)\mathfrak{U}(\Sigma) be the space of free S1S^{1}-actions on Σ\Sigma. For 𝔲𝔘(Σ)\mathfrak{u}\in\mathfrak{U}(\Sigma), we denote by V=V𝔲V=V_{\mathfrak{u}} the fundamental vector field on Σ\Sigma generated by 𝔲\mathfrak{u}. We have a natural map

𝔘(Σ)𝔓(Σ),\mathfrak{U}(\Sigma)\to\mathfrak{P}(\Sigma), (4.2)

which associates to a free S1S^{1}-action the quotient map onto its orbit space. It is a classical fact that this map is surjective. Indeed, let us fix 𝔭𝔓(Σ)\mathfrak{p}\in\mathfrak{P}(\Sigma). Using a partition of unity, we find a vector field XX on Σ\Sigma, positively tangent to the 𝔭\mathfrak{p}-fibres at every point. If T(z)>0T(z)>0 is the period of the periodic orbit of XX starting at zΣz\in\Sigma, then the rescaled vector field V:=TXV:=TX yields the desired S1S^{1}-action. The set 𝔘(Σ)\mathfrak{U}(\Sigma) carries a natural C1C^{1}-topology as a closed subset of the space Maps(S1×Σ,Σ)\mathrm{Maps}(S^{1}\times\Sigma,\Sigma) and we endow 𝔓(Σ)\mathfrak{P}(\Sigma) with the quotient topology brought by the map (4.2). We write 𝔘(𝔭)\mathfrak{U}(\mathfrak{p}) for the fibre above 𝔭𝔓(Σ)\mathfrak{p}\in\mathfrak{P}(\Sigma). This is a convex space in the following sense. If V0V_{0} and V1=T1V0V_{1}=T_{1}V_{0} are the fundamental vector fields of 𝔲0,𝔲1𝔘(𝔭)\mathfrak{u}_{0},\mathfrak{u}_{1}\in\mathfrak{U}(\mathfrak{p}), then

Vr:=V1r+(1r)T1,r[0,1]V_{r}:=\frac{V_{1}}{r+(1-r)T_{1}},\qquad\forall r\in[0,1]

is the fundamental vector field of some 𝔲r𝔘(𝔭)\mathfrak{u}_{r}\in\mathfrak{U}(\mathfrak{p}).

Finally, if Λ(Σ)\Lambda(\Sigma) denotes the space one-periodic curves in Σ\Sigma, we have a map

ȷ𝔭:ΣΛ(Σ)\jmath_{\mathfrak{p}}:\Sigma\to\Lambda(\Sigma) (4.3)

associating to a point the 𝔲\mathfrak{u}-orbit through it for some 𝔲𝔘(𝔭)\mathfrak{u}\in\mathfrak{U}(\mathfrak{p}). Up to an orientation-preserving change of parametrisation of the elements of Λ(Σ)\Lambda(\Sigma), this map depends only on 𝔭\mathfrak{p}.

Definition 4.3.

If 𝔭:ΣM\mathfrak{p}:\Sigma\to M belongs to 𝔓(Σ)\mathfrak{P}(\Sigma) and 𝔲\mathfrak{u} belongs to 𝔘(𝔭)\mathfrak{U}(\mathfrak{p}), we write

eH2(M;)e_{\mathbb{Z}}\in H^{2}(M;{\mathbb{Z}})

for minus the Euler class of 𝔲\mathfrak{u}, as defined, for example in [Che77]. This class is independent of 𝔲𝔘(𝔭)\mathfrak{u}\in\mathfrak{U}(\mathfrak{p}), and therefore, we refer to it as minus the Euler class of 𝔭\mathfrak{p}.

Remark 4.4.

The bundle 𝔭\mathfrak{p} is trivial (namely admits a global section) if and only if e=0e_{\mathbb{Z}}=0. If 𝔭:ΣM\mathfrak{p}^{\text{\tiny$\vee$}}:\Sigma^{\text{\tiny$\vee$}}\to M^{\text{\tiny$\vee$}} is another bundle with minus Euler class eH2(M;)e^{\text{\tiny$\vee$}}_{\mathbb{Z}}\in H^{2}(M^{\text{\tiny$\vee$}};{\mathbb{Z}}), then

𝔭=Π𝔭e=πe,\mathfrak{p}^{\text{\tiny$\vee$}}=\Pi^{*}\mathfrak{p}\quad\Longrightarrow\quad e^{\text{\tiny$\vee$}}_{\mathbb{Z}}=\pi^{*}e_{\mathbb{Z}},

where Π\Pi is a bundle map as in (4.1).

The inclusion map {\mathbb{Z}}\hookrightarrow{\mathbb{R}} induces a map on the level of homology and cohomology and we write ee and [𝔭1(pt)][\mathfrak{p}^{-1}({\mathrm{pt}})] for the images of ee_{\mathbb{Z}} and [𝔭1(pt)][\mathfrak{p}^{-1}({\mathrm{pt}})]_{\mathbb{Z}}, respectively:

H2(M;)\displaystyle H^{2}(M;{\mathbb{Z}}) H2(M;)HdR2(M),\displaystyle\longrightarrow H^{2}(M;{\mathbb{R}})\cong H^{2}_{\mathrm{dR}}(M),\qquad H1(Σ;)\displaystyle H_{1}(\Sigma;{\mathbb{Z}}) H1(Σ;).\displaystyle\longrightarrow H_{1}(\Sigma;{\mathbb{R}}).
e\displaystyle e_{\mathbb{Z}} e\displaystyle\longmapsto e [𝔭1(pt)]\displaystyle[\mathfrak{p}^{-1}({\mathrm{pt}})]_{\mathbb{Z}} [𝔭1(pt)]\displaystyle\longmapsto[\mathfrak{p}^{-1}({\mathrm{pt}})]

From the piece of the Gysin exact sequence HdR0(M)eHdR2(M)𝔭HdR2(Σ)H^{0}_{\operatorname{dR}}(M)\xrightarrow{\cup e}H_{\operatorname{dR}}^{2}(M)\xrightarrow{\mathfrak{p}^{*}}H^{2}_{\operatorname{dR}}(\Sigma), we know that

e=ker(𝔭:HdR2(M)HdR2(Σ)).{\mathbb{R}}\cdot e=\ker\Big{(}\mathfrak{p}^{*}:H^{2}_{\operatorname{dR}}(M)\to H^{2}_{\operatorname{dR}}(\Sigma)\Big{)}. (4.4)

Moreover, by the universal coefficient theorem, we get

e is torsione=0,[𝔭1(pt)] is torsion[𝔭1(pt)]=0.e_{\mathbb{Z}}\text{ is torsion}\ \ \iff\ \ e=0,\qquad\qquad[\mathfrak{p}^{-1}({\mathrm{pt}})]_{\mathbb{Z}}\text{ is torsion}\ \ \iff\ \ [\mathfrak{p}^{-1}({\mathrm{pt}})]=0.

The Chern-Weil theory yields a convenient description of the real cohomology class ee. Let 𝒦(𝔲)\mathcal{K}(\mathfrak{u}) denote the space of connection one-forms associated with 𝔲𝔘(𝔭)\mathfrak{u}\in\mathfrak{U}(\mathfrak{p}):

𝒦(𝔲):={ηΩ1(Σ)|η(V)=1,dη=𝔭κη for some κηΩ2(M)}.\mathcal{K}(\mathfrak{u}):=\big{\{}\eta\in\Omega^{1}(\Sigma)\ \big{|}\ \eta(V)=1,\;{\mathrm{d}}\eta=\mathfrak{p}^{*}\kappa_{\eta}\text{ for some }\kappa_{\eta}\in\Omega^{2}(M)\big{\}}.

This space is non-empty, and for every η𝒦(𝔲)\eta\in\mathcal{K}(\mathfrak{u}), the form κη\kappa_{\eta} is closed and we have

e=[κη]HdR2(M).e=[\kappa_{\eta}]\in H^{2}_{{\mathrm{dR}}}(M).

Conversely, for any κΞ2(M)\kappa\in\Xi^{2}(M) with e=[κ]e=[\kappa], there is ηκ𝒦(𝔲)\eta_{\kappa}\in\mathcal{K}(\mathfrak{u}) with κηκ=κ\kappa_{\eta_{\kappa}}=\kappa. We denote by

𝒦(𝔭):=𝔲𝔘(𝔭)𝒦(𝔲)\mathcal{K}(\mathfrak{p}):=\bigcup_{\mathfrak{u}\in\mathfrak{U}(\mathfrak{p})}\mathcal{K}(\mathfrak{u})

the set of connection one-forms for 𝔭\mathfrak{p}.

4.2 Flat S1S^{1}-bundles and the local structure of 𝔓(Σ)\mathfrak{P}(\Sigma)

This subsection is devoted to the proof of three lemmas. In the first two, we study flat bundles, i.e. with e=0e=0. In the last one, we prove a theorem of Weinstein [Wei74] showing that the space 𝔓(Σ)\mathfrak{P}(\Sigma) is locally trivial.

Lemma 4.5.

Let 𝔭:ΣM\mathfrak{p}:\Sigma\to M be an oriented S1S^{1}-bundle. We have an equivalence

e=0[𝔭1(pt)]0.e=0\qquad\Longleftrightarrow\qquad[\mathfrak{p}^{-1}({\mathrm{pt}})]\neq 0. (4.5)

If cHdR2(M)c\in H^{2}_{\mathrm{dR}}(M) and we define C:=𝔭cHdR2(Σ)C:=\mathfrak{p}^{*}c\in H^{2}_{\mathrm{dR}}(\Sigma), there holds

PD(Cn)=cn,[M][𝔭1(pt)]H1(Σ;).\mathrm{PD}(C^{n})=\langle c^{n},[M]\rangle\cdot[\mathfrak{p}^{-1}({\mathrm{pt}})]\,\in\,H_{1}(\Sigma;{\mathbb{R}}). (4.6)

In particular, if cn0c^{n}\neq 0, then

(i)\displaystyle(i) ker(HdR1(Σ),[𝔭1(pt)])=ker(HdR1(Σ)()CnHdR2n+1(Σ)),\displaystyle\ \ \ker\Big{(}H^{1}_{\operatorname{dR}}(\Sigma)\xrightarrow{\langle\,\cdot\,,[\mathfrak{p}^{-1}({\mathrm{pt}})]\rangle}{\mathbb{R}}\Big{)}=\ker\Big{(}H^{1}_{\operatorname{dR}}(\Sigma)\xrightarrow{(\,\cdot\,)\cup C^{n}}H^{2n+1}_{\mathrm{dR}}(\Sigma)\Big{)}, (4.7)
(ii)\displaystyle(ii) Cn0e=0.\displaystyle\ \ C^{n}\neq 0\quad\Longleftrightarrow\quad e=0.
Proof.

The proof of (4.5) relies on the Gysin sequence and of its Poincaré dual

HdR2n2(M)\textstyle{H^{2n-2}_{\operatorname{dR}}(M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}e\scriptstyle{\cup e}PD\scriptstyle{\mathrm{PD}}HdR2n(M)\textstyle{H_{\operatorname{dR}}^{2n}(M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}PD\scriptstyle{\mathrm{PD}}𝔭\scriptstyle{\mathfrak{p}^{*}}PD\scriptstyle{\mathrm{PD}}HdR2n(Σ)\textstyle{H^{2n}_{\operatorname{dR}}(\Sigma)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔭\scriptstyle{\mathfrak{p}_{*}}PD\scriptstyle{\mathrm{PD}}H2n1(M),\textstyle{H^{2n-1}(M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces,}PD\scriptstyle{\mathrm{PD}}H2(M;)\textstyle{H_{2}(M;{\mathbb{R}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}e\scriptstyle{\cap e}H0(M;)\textstyle{H_{0}(M;{\mathbb{R}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}PD(𝔭)\scriptstyle{\mathrm{PD}(\mathfrak{p}^{*})}H1(Σ;)\textstyle{H_{1}(\Sigma;{\mathbb{R}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H1(M;)\textstyle{H_{1}(M;{\mathbb{R}})}

where PD\mathrm{PD} denotes Poincaré duality and 𝔭\mathfrak{p}_{*} stands for the integration along fibres. We claim that PD(𝔭)\mathrm{PD}(\mathfrak{p}^{*}) sends the class of a point [pt][{\mathrm{pt}}] to [𝔭1(pt)][\mathfrak{p}^{-1}({\mathrm{pt}})]. To this purpose, let μΞ2n(M)\mu\in\Xi^{2n}(M) be such that PD([μ])=[pt]\mathrm{PD}([\mu])=[{\mathrm{pt}}], so that Mμ=1\int_{M}\mu=1. Going around the central square in the diagram above, the claim follows, if we can show that PD(𝔭[μ])=[𝔭1(pt)]\mathrm{PD}(\mathfrak{p}^{*}[\mu])=[\mathfrak{p}^{-1}({\mathrm{pt}})]. This last equality is true since for all τΞ1(Σ)\tau\in\Xi^{1}(\Sigma) we have

[τ],[𝔭1(pt)]=𝔭[τ]=𝔭[τ]Mμ=M𝔭[τ]μ=()Στ𝔭μ=[τ]𝔭[μ],[Σ],\langle[\tau],[\mathfrak{p}^{-1}({\mathrm{pt}})]\rangle=\mathfrak{p}_{*}[\tau]=\mathfrak{p}_{*}[\tau]\cdot\int_{M}\mu=\int_{M}\mathfrak{p}_{*}[\tau]\cdot\mu\stackrel{{\scriptstyle(\star)}}{{=}}\int_{\Sigma}\tau\wedge\mathfrak{p}^{*}\mu=\langle[\tau]\cup\mathfrak{p}^{*}[\mu],[\Sigma]\rangle,

where in ()(\star) we have used Fubini’s Theorem. Therefore, [𝔭1(pt)]0[\mathfrak{p}^{-1}({\mathrm{pt}})]\neq 0 if and only if the map e\cap e is equal to zero. This happens if and only if e,H2(M;)=0\langle e,H_{2}(M;{\mathbb{R}})\rangle=0 as one sees identifying H0(M;)H_{0}(M;{\mathbb{R}}) with {\mathbb{R}}. Finally, from the universal coefficient theorem, e,H2(M;)=0\langle e,H_{2}(M;{\mathbb{R}})\rangle=0 if and only if e=0e=0.

Let us show (4.6). If ωΞc2(M)\mathrm{\omega}\in\Xi^{2}_{c}(M) and τΞ1(Σ)\tau\in\Xi^{1}(\Sigma) is arbitrary, we have

cn,[M][τ],[𝔭1(pt)]=M𝔭[τ]ωn=Στ𝔭ωn=[τ]Cn,[Σ].\langle c^{n},[M]\rangle\cdot\langle[\tau],[\mathfrak{p}^{-1}({\mathrm{pt}})]\rangle=\int_{M}\mathfrak{p}_{*}[\tau]\cdot\mathrm{\omega}^{n}=\int_{\Sigma}\tau\wedge\mathfrak{p}^{*}\mathrm{\omega}^{n}=\langle[\tau]\cup C^{n},[\Sigma]\rangle.

Let us assume that cn,[M]0\langle c^{n},[M]\rangle\neq 0. Looking again at the last equation, we readily see that (i) in (4.7) holds. The equivalence in (ii) stems from a combination of (4.6) and (4.5)

Cn0PD(Cn)0[𝔭1(pt)]0e=0.C^{n}\neq 0\quad\iff\quad\mathrm{PD}(C^{n})\neq 0\quad\iff\quad[\mathfrak{p}^{-1}({\mathrm{pt}})]\neq 0\quad\iff\quad e=0.\qed

We now show that a flat bundle can always be pulled back to a trivial one.

Lemma 4.6.

Let 𝔭:ΣM\mathfrak{p}:\Sigma\to M be an oriented S1S^{1}-bundle. The class eHdR2(M)e\in H^{2}_{\mathrm{dR}}(M) vanishes if and only if there exists a finite cyclic cover π:MM\pi:M^{\text{\tiny$\vee$}}\to M such that the pull-back bundle 𝔭:ΣM\mathfrak{p}^{\text{\tiny$\vee$}}:\Sigma^{\text{\tiny$\vee$}}\to M^{\text{\tiny$\vee$}} of 𝔭\mathfrak{p} by π\pi is trivial. In this case, the order of ee_{\mathbb{Z}} in H2(M;)H^{2}(M;{\mathbb{Z}}) equals the minimum degree of a cover with the properties above.

Proof.

We preliminarily observe that if π:MM\pi:M^{\text{\tiny$\vee$}}\to M is a finite cover of degree kk and the bundle 𝔭:ΣM\mathfrak{p}^{\text{\tiny$\vee$}}:\Sigma^{\text{\tiny$\vee$}}\to M^{\text{\tiny$\vee$}} is the pull-back of 𝔭\mathfrak{p} through π\pi with minus the Euler class ee_{\mathbb{Z}}^{\text{\tiny$\vee$}}, then by Remark 4.4 there holds

πe=ke,\pi_{*}e_{\mathbb{Z}}^{\text{\tiny$\vee$}}=k\cdot e_{\mathbb{Z}},

where π:H2(M;)H2(M;)\pi_{*}:H^{2}(M^{\text{\tiny$\vee$}};{\mathbb{Z}})\to H^{2}(M;{\mathbb{Z}}) is the transfer map (see [Hat02, Chapter 3.G]). Therefore, if we suppose that 𝔭\mathfrak{p}^{\text{\tiny$\vee$}} is trivial, we deduce that e=0e_{\mathbb{Z}}^{\text{\tiny$\vee$}}=0, and hence, ke=0k\cdot e_{\mathbb{Z}}=0. In particular, the order of ee_{\mathbb{Z}} divides kk.

Conversely, let kk be a positive integer such that ke=0k\cdot e_{\mathbb{Z}}=0 and take 𝔲𝔘(𝔭)\mathfrak{u}\in\mathfrak{U}(\mathfrak{p}). Let kS1\tfrac{{\mathbb{Z}}}{k{\mathbb{Z}}}\to S^{1} be the canonical homomorphism jj/kj\to j/k, and we denote by 𝔲k\mathfrak{u}_{k} the free k\tfrac{{\mathbb{Z}}}{k{\mathbb{Z}}}-action on Σ\Sigma obtained combining this homomorphism with 𝔲\mathfrak{u}. Let Σk=Σ/𝔲k\Sigma_{k}=\Sigma/\mathfrak{u}_{k} be the quotient by this action and denote by Πk:ΣΣk\Pi_{k}:\Sigma\to\Sigma_{k} the associated quotient map. The natural map 𝔭k:ΣkM\mathfrak{p}_{k}:\Sigma_{k}\to M is an oriented S1S^{1}-bundle. Let (e)kH2(M;)(e_{\mathbb{Z}})_{k}\in H^{2}(M;{\mathbb{Z}}) denote minus the Euler class of 𝔭k\mathfrak{p}_{k}. Since S1/kS1S^{1}/\tfrac{{\mathbb{Z}}}{k{\mathbb{Z}}}\cong S^{1} through the map ϕkϕ\phi\mapsto k\cdot\phi, we see that (e)k=ke=0(e_{\mathbb{Z}})_{k}=k\cdot e_{\mathbb{Z}}=0. Therefore, we have a section 𝔰k:MΣk\mathfrak{s}_{k}:M\to\Sigma_{k} for 𝔭k\mathfrak{p}_{k} and we take a connected component MΣM^{\text{\tiny$\vee$}}\subset\Sigma of Πk1(𝔰k(M))\Pi_{k}^{-1}(\mathfrak{s}_{k}(M)). If 𝔦:MΣ\mathfrak{i}:M^{\text{\tiny$\vee$}}\to\Sigma is the inclusion map, we define π:MM\pi:M^{\text{\tiny$\vee$}}\to M as π:=𝔭𝔦\pi:=\mathfrak{p}\circ\mathfrak{i}. The map π\pi is a covering map, whose deck transformation group is given by the sub-action of 𝔲k\mathfrak{u}_{k} that leaves MM^{\text{\tiny$\vee$}} invariant. The deck transformation group is isomorphic to j\tfrac{{\mathbb{Z}}}{j{\mathbb{Z}}}, where jj divides kk, since it is a subgroup of the cyclic group j\tfrac{{\mathbb{Z}}}{j{\mathbb{Z}}}. Let Π:ΣΣ\Pi:\Sigma^{\text{\tiny$\vee$}}\to\Sigma be a covering map lifting π\pi. We sum up the construction above in the commutative diagram

Σ\textstyle{\Sigma^{\text{\tiny$\vee$}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔭\scriptstyle{\mathfrak{p}^{\text{\tiny$\vee$}}}Π\scriptstyle{\Pi}Σ\textstyle{\Sigma\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔭\scriptstyle{\mathfrak{p}}Πk\scriptstyle{\Pi_{k}}Σk.\textstyle{\Sigma_{k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces.}𝔭k\scriptstyle{\mathfrak{p}_{k}}M\textstyle{M^{\text{\tiny$\vee$}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔰\scriptstyle{\mathfrak{s}^{\text{\tiny$\vee$}}}𝔦\scriptstyle{\mathfrak{i}}π\scriptstyle{\pi}M\textstyle{M\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔰k\scriptstyle{\mathfrak{s}_{k}}

We construct a section 𝔰:MΣ\mathfrak{s}^{\text{\tiny$\vee$}}:M^{\text{\tiny$\vee$}}\to\Sigma^{\text{\tiny$\vee$}} tautologically, as follows. Let us consider an arbitrary qMq^{\text{\tiny$\vee$}}\in M^{\text{\tiny$\vee$}}. By definition of pull-back bundle, the 𝔭\mathfrak{p}^{\text{\tiny$\vee$}}-fibre over qq^{\text{\tiny$\vee$}} is identified through Π\Pi with the 𝔭\mathfrak{p}-fibre over π(q)\pi(q^{\text{\tiny$\vee$}}). Since 𝔦(q)Σ\mathfrak{i}(q^{\text{\tiny$\vee$}})\in\Sigma belongs to the 𝔭\mathfrak{p}-fibre over π(q)\pi(q^{\text{\tiny$\vee$}}), we can define 𝔰(q)\mathfrak{s}^{\text{\tiny$\vee$}}(q^{\text{\tiny$\vee$}}) through the equation

Π(𝔰(q))=𝔦(q).\Pi(\mathfrak{s}^{\text{\tiny$\vee$}}(q^{\text{\tiny$\vee$}}))=\mathfrak{i}(q^{\text{\tiny$\vee$}}).

This proves the existence of a cover as in the statement of the lemma and shows that the degree kk^{\text{\tiny$\vee$}} is less than or equal to the order of ee_{\mathbb{Z}}. ∎

If 𝔭𝔓(Σ)\mathfrak{p}\in\mathfrak{P}(\Sigma) and {Ψr:ΣΣ}\{\Psi_{r}:\Sigma\to\Sigma\} is an isotopy with Ψ0=idΣ\Psi_{0}=\mathrm{id}_{\Sigma}, then {𝔭r:=Ψr𝔭}\{\mathfrak{p}_{r}:=\Psi_{r}^{*}\mathfrak{p}\} yields a path in 𝔓(Σ)\mathfrak{P}(\Sigma). The next lemma shows that all paths in 𝔓(Σ)\mathfrak{P}(\Sigma) arise in this way.

Lemma 4.7.

The following two statements hold:

  1. (i)

    Every oriented S1S^{1}-bundle 𝔭0𝔓(Σ)\mathfrak{p}_{0}\in\mathfrak{P}(\Sigma) has a C1C^{1}-neighbourhood 𝒲\mathcal{W} and a continuous map (𝔭𝒲)(Ψ𝔭Diff(Σ))(\mathfrak{p}\in\mathcal{W})\mapsto(\Psi_{\mathfrak{p}}\in\mathrm{Diff}(\Sigma)) such that

    Ψ𝔭0=idΣ,Ψ𝔭𝔭=𝔭0,𝔭𝒲.\bullet\ \ \Psi_{\mathfrak{p}_{0}}=\mathrm{id}_{\Sigma},\qquad\qquad\bullet\ \ \Psi_{\mathfrak{p}}^{*}\mathfrak{p}=\mathfrak{p}_{0},\quad\forall\,\mathfrak{p}\in\mathcal{W}.
  2. (ii)

    If {𝔭r:ΣMr}\{\mathfrak{p}_{r}:\Sigma\to M_{r}\} is a path in 𝔓(Σ)\mathfrak{P}(\Sigma), there exists an isotopy {Ψr}\{\Psi_{r}\} of Σ\Sigma such that

    Ψ0=idΣ,Ψr𝔭r=𝔭0,r[0,1].\qquad\bullet\ \ \Psi_{0}=\mathrm{id}_{\Sigma},\ \qquad\qquad\bullet\ \ \Psi_{r}^{*}\mathfrak{p}_{r}=\mathfrak{p}_{0},\quad\forall\,r\in[0,1].
Proof.

Pick an arbitrary connection η𝒦(𝔭0)\eta\in\mathcal{K}(\mathfrak{p}_{0}) and an arbitrary Riemannian metric on MM with injectivity radius ρinj\rho_{\mathrm{inj}} and distance function dist:M×M[0,){\mathrm{dist}}:M\times M\to[0,\infty). We define the open subset of M×ΣM\times\Sigma

W:={(q,z)M×Σ|dist(q,𝔭0(z))<ρinj/2}W:=\Big{\{}(q,z)\in M\times\Sigma\ \Big{|}\ {\mathrm{dist}}(q,\mathfrak{p}_{0}(z))<\rho_{\mathrm{inj}}/2\Big{\}}

and let 𝔮M:WM\mathfrak{q}_{M}:W\to M and 𝔷Σ:WΣ\mathfrak{z}_{\Sigma}:W\to\Sigma be the projections on the two factors. We define 𝒲\mathcal{W} as the space of all oriented S1S^{1}-bundles 𝔭\mathfrak{p} such that the following two properties hold:

  • For every z1,z2z_{1},z_{2} in Σ\Sigma belonging to the same 𝔭\mathfrak{p}-fibre, we have dist(𝔭0(z1),𝔭0(z2))<ρinj/2{\mathrm{dist}}(\mathfrak{p}_{0}(z_{1}),\mathfrak{p}_{0}(z_{2}))<\rho_{\mathrm{inj}}/2;

  • For every zΣz\in\Sigma, the disc DzΣD_{z}\subset\Sigma is transverse to the orbits of 𝔭\mathfrak{p}. Here, DzD_{z} is the union of all the η\eta-horizontal lifts through zz of the geodesic rays in MM emanating from 𝔭0(z)\mathfrak{p}_{0}(z) and with length ρinj\rho_{\mathrm{inj}}.

For every 𝔭𝒲\mathfrak{p}\in\mathcal{W}, we construct a map of fibre bundles over MM given by

S:WTM,S(q,z):=(q,v(q,z)),v(q,z):=S1expq1𝔭0γz(t)dt,S:W\to{\mathrm{T}}M,\qquad S(q,z):=\big{(}q,v(q,z)\big{)},\quad v(q,z):=\int_{S^{1}}\exp_{q}^{-1}\circ\,\mathfrak{p}_{0}\circ\gamma_{z}(t){\mathrm{d}}t,

where γz:S1Σ\gamma_{z}:S^{1}\to\Sigma is the oriented 𝔭\mathfrak{p}-fibre passing through zz. By definition, SS is constant along the 𝔭\mathfrak{p}-fibres. Let 0MTM0_{M}\subset{\mathrm{T}}M be the zero section and set G:=S1(0M)G:=S^{-1}(0_{M}). We claim that

  • SS is transverse to 0M0_{M},

  • 𝔭M:=𝔮M|G:GM\mathfrak{p}_{M}:=\mathfrak{q}_{M}|_{G}:G\to M is an S1S^{1}-bundle map,

  • ΨΣ:=𝔷Σ|G:GΣ\Psi_{\Sigma}:=\mathfrak{z}_{\Sigma}|_{G}:G\to\Sigma is a diffeomorphism.

This is clear if 𝔭=𝔭0\mathfrak{p}=\mathfrak{p}_{0}, since in this case S(q,z)=expq1(𝔭0(z))S(q,z)=\exp_{q}^{-1}(\mathfrak{p}_{0}(z)) and G={(q,z)|𝔭0(z)=q}G=\{(q,z)\ |\ \mathfrak{p}_{0}(z)=q\}. Therefore, up to shrinking 𝒲\mathcal{W}, it also holds for 𝔭𝒲\mathfrak{p}\in\mathcal{W}, as SS depends continuously on 𝔭\mathfrak{p} and being transverse, or a submersion or a diffeomorphism is a C1C^{1}-open condition [Hir94, Chapter 2]. Thus, we get an S1S^{1}-bundle 𝔭:=𝔭MΨΣ1:ΣM\mathfrak{p}^{\prime}:=\mathfrak{p}_{M}\circ\Psi_{\Sigma}^{-1}:\Sigma\to M, which is equivalent to 𝔭\mathfrak{p}. We now define a map Φ𝔭:ΣΣ\Phi_{\mathfrak{p}}:\Sigma\to\Sigma such that Φ𝔭𝔭0=𝔭\Phi_{\mathfrak{p}}^{*}\mathfrak{p}_{0}=\mathfrak{p}. For every zΣz\in\Sigma, Φ𝔭(z)\Phi_{\mathfrak{p}}(z) is the parallel transport with respect to η\eta along the radial geodesic on MM connecting 𝔭0(z)\mathfrak{p}_{0}(z) with 𝔭(z)\mathfrak{p}^{\prime}(z). The dependence of Φ𝔭\Phi_{\mathfrak{p}} from 𝔭\mathfrak{p} is continuous and clearly Φ𝔭0=idΣ\Phi_{\mathfrak{p}_{0}}=\mathrm{id}_{\Sigma}. As being a diffeomorphism is a C1C^{1}-open condition, we have Φ𝔭Diff(Σ)\Phi_{\mathfrak{p}}\in\mathrm{Diff}(\Sigma) and (i) is proved by setting Ψ𝔭=Φ𝔭1\Psi_{\mathfrak{p}}=\Phi_{\mathfrak{p}}^{-1}.

For part (ii), we break the given path into short paths and apply the first part. ∎

5 Weakly Zoll pairs and Zoll odd-symplectic forms

5.1 Definitions and first properties

Definition 5.1.

A couple (𝔭,c)(\mathfrak{p},c) is called a weakly Zoll pair if 𝔭:ΣM\mathfrak{p}:\Sigma\to M is an element of 𝔓(Σ)\mathfrak{P}(\Sigma) and cHdR2(M)c\in H^{2}_{\mathrm{dR}}(M). We say that ΩΞ2(Σ)\Omega\in\Xi^{2}(\Sigma) is associated with (𝔭,c)(\mathfrak{p},c), if Ω=𝔭ω\Omega=\mathfrak{p}^{*}\mathrm{\omega} for some ωΞc2(M)\mathrm{\omega}\in\Xi^{2}_{c}(M). We denote by (Σ)\mathfrak{Z}(\Sigma) the set of weakly Zoll pairs and by C(Σ)\mathfrak{Z}_{C}(\Sigma) the subset of those (𝔭,c)(Σ)(\mathfrak{p},c)\in\mathfrak{Z}(\Sigma) with 𝔭c=CHdR2(Σ)\mathfrak{p}^{*}c=C\in H^{2}_{\mathrm{dR}}(\Sigma).

Let (𝔭0,c0)C(Σ)(\mathfrak{p}_{0},c_{0})\in\mathfrak{Z}_{C}(\Sigma) with 𝔭0:ΣM0\mathfrak{p}_{0}:\Sigma\to M_{0} and take any ω0Ξc02(M0)\mathrm{\omega}_{0}\in\Xi^{2}_{c_{0}}(M_{0}). We consider the volume functionals Vol:Ω1(Σ)\mathrm{Vol}:\Omega^{1}(\Sigma)\to{\mathbb{R}} and 𝔙𝔬𝔩:ΞC2(Σ){\mathfrak{Vol}}:\Xi^{2}_{C}(\Sigma)\to{\mathbb{R}} defined in Section 2 with respect to Ω0:=𝔭0ω0ΞC2(Σ)\Omega_{0}:=\mathfrak{p}_{0}^{*}\mathrm{\omega}_{0}\in\Xi^{2}_{C}(\Sigma). We use 𝔙𝔬𝔩{\mathfrak{Vol}} to get a volume on C(Σ)\mathfrak{Z}_{C}(\Sigma) denoted with the same name:

𝔙𝔬𝔩:C(Σ),𝔙𝔬𝔩(𝔭,c):=𝔙𝔬𝔩(𝔭ω),ωΞc2(M).{\mathfrak{Vol}}:\mathfrak{Z}_{C}(\Sigma)\to{\mathbb{R}},\qquad{\mathfrak{Vol}}(\mathfrak{p},c):={\mathfrak{Vol}}(\mathfrak{p}^{*}\mathrm{\omega}),\quad\mathrm{\omega}\in\Xi_{c}^{2}(M).
Lemma 5.2.

The following three statements hold.

  1. (i)

    The volume functional 𝔙𝔬𝔩:C(Σ){\mathfrak{Vol}}:\mathfrak{Z}_{C}(\Sigma)\to{\mathbb{R}} is well-defined.

  2. (ii)

    For all ζΩ1(M0)\zeta\in\Omega^{1}(M_{0}), 𝔭0ζ\mathfrak{p}_{0}^{*}\zeta is 𝔭0ω0\mathfrak{p}_{0}^{*}\mathrm{\omega}_{0}-normalised, namely

    Vol(𝔭0ζ)=0.\mathrm{Vol}(\mathfrak{p}_{0}^{*}\zeta)=0.
  3. (iii)

    The functions 𝔙𝔬𝔩:ΞC2(Σ){\mathfrak{Vol}}:\Xi^{2}_{C}(\Sigma)\to{\mathbb{R}}, 𝔙𝔬𝔩:C(Σ){\mathfrak{Vol}}:\mathfrak{Z}_{C}(\Sigma)\to{\mathbb{R}} depend only on the pair (𝔭0,c0)(\mathfrak{p}_{0},c_{0}) and not on the chosen ω0Ξc02(M0)\mathrm{\omega}_{0}\in\Xi^{2}_{c_{0}}(M_{0}).

Proof.

All three items will stem out a preliminary result. Let (𝔭,c)C(Σ)(\mathfrak{p},c)\in\mathfrak{Z}_{C}(\Sigma) with ωΞc2(M)\mathrm{\omega}\in\Xi^{2}_{c}(M) and pick αωΩ1(Σ)\alpha_{\mathrm{\omega}}\in\Omega^{1}(\Sigma) such that 𝔭0ω0+dαω=𝔭ω\mathfrak{p}_{0}^{*}\mathrm{\omega}_{0}+{\mathrm{d}}\alpha_{\mathrm{\omega}}=\mathfrak{p}^{*}\mathrm{\omega}. We claim that

Vol(αω+𝔭ζ)=Vol(αω),ζΩ1(M).\mathrm{Vol}(\alpha_{\mathrm{\omega}}+\mathfrak{p}^{*}\zeta)=\mathrm{Vol}(\alpha_{\mathrm{\omega}}),\qquad\forall\,\zeta\in\Omega^{1}(M). (5.1)

Indeed, if we set αr:=αω+r𝔭ζ\alpha_{r}:=\alpha_{\mathrm{\omega}}+r\mathfrak{p}^{*}\zeta so that 𝔭0ω0+dαr=𝔭(ω+rdζ)\mathfrak{p}_{0}^{*}\mathrm{\omega}_{0}+{\mathrm{d}}\alpha_{r}=\mathfrak{p}^{*}(\mathrm{\omega}+r{\mathrm{d}}\zeta) and α˙r=𝔭ζ\dot{\alpha}_{r}=\mathfrak{p}^{*}\zeta, then

Vol(αω+𝔭ζ)Vol(αω)=01volαrα˙rdr=01(Σ𝔭ζ𝔭(ω+rdζ)n)dr=0.\mathrm{Vol}(\alpha_{\mathrm{\omega}}+\mathfrak{p}^{*}\zeta)-\mathrm{Vol}(\alpha_{\mathrm{\omega}})=\int_{0}^{1}\mathrm{vol}_{\alpha_{r}}\cdot\dot{\alpha}_{r}{\mathrm{d}}r=\int_{0}^{1}\Big{(}\int_{\Sigma}\mathfrak{p}^{*}\zeta\wedge\mathfrak{p}^{*}(\mathrm{\omega}+r{\mathrm{d}}\zeta)^{n}\Big{)}{\mathrm{d}}r=0.

If ω\mathrm{\omega}^{\prime} is another form in Ξc2(M)\Xi^{2}_{c}(M), then there is ζΩ1(M)\zeta\in\Omega^{1}(M) with ωω=dζ\mathrm{\omega}^{\prime}-\mathrm{\omega}={\mathrm{d}}\zeta and (5.1) implies

𝔙𝔬𝔩(𝔭ω)=𝔙𝔬𝔩(𝔭ω),{\mathfrak{Vol}}(\mathfrak{p}^{*}\mathrm{\omega}^{\prime})={\mathfrak{Vol}}(\mathfrak{p}^{*}\mathrm{\omega}),

which establishes item (i). Choosing (𝔭,c)=(𝔭0,c0)(\mathfrak{p},c)=(\mathfrak{p}_{0},c_{0}) and α=0\alpha=0 in (5.1), we get item (ii). For item (iii), we take another form ω0Ξc02(M0)\mathrm{\omega}_{0}^{\prime}\in\Xi^{2}_{c_{0}}(M_{0}) and write ω0ω0=dζ0\mathrm{\omega}_{0}^{\prime}-\mathrm{\omega}_{0}={\mathrm{d}}\zeta_{0} for some ζ0Ω1(M0)\zeta_{0}\in\Omega^{1}(M_{0}). Applying Lemma 2.4 with Ω0=𝔭0ω0\Omega_{0}=\mathfrak{p}_{0}^{*}\mathrm{\omega}_{0} and Ω0=𝔭0ω0\Omega_{0}^{\prime}=\mathfrak{p}^{*}_{0}\mathrm{\omega}_{0}^{\prime} together with item (ii), we deduce

Vol(α)=Vol(α𝔭0ζ0),αΩ1(Σ),\mathrm{Vol}(\alpha)=\mathrm{Vol}^{\prime}(\alpha-\mathfrak{p}_{0}^{*}\zeta_{0}),\qquad\forall\,\alpha\in\Omega^{1}(\Sigma),

where Vol\mathrm{Vol}^{\prime} is the volume functional associated with Ω0\Omega_{0}^{\prime}. This implies 𝔙𝔬𝔩(Ω)=𝔙𝔬𝔩(Ω){\mathfrak{Vol}}(\Omega)={\mathfrak{Vol}}^{\prime}(\Omega) for all ΩΞC2(Σ)\Omega\in\Xi^{2}_{C}(\Sigma). ∎

We have a canonical projection

𝔓:(Σ)𝔓(Σ),𝔓(𝔭,c)=𝔭.\mathfrak{P}:\mathfrak{Z}(\Sigma)\to\mathfrak{P}(\Sigma),\qquad\mathfrak{P}(\mathfrak{p},c)=\mathfrak{p}.

Let us fix a class CHdR2(Σ)C\in H^{2}_{\mathrm{dR}}(\Sigma) and a connected component 𝔓0(Σ)\mathfrak{P}^{0}(\Sigma) of 𝔓(Σ)\mathfrak{P}(\Sigma). We define

C0(Σ):=𝔓1(𝔓0(Σ))C(Σ).\mathfrak{Z}_{C}^{0}(\Sigma):=\mathfrak{P}^{-1}(\mathfrak{P}^{0}(\Sigma))\cap\mathfrak{Z}_{C}(\Sigma).

We consider the restriction of 𝔓\mathfrak{P} on this set

𝔓C0:C0(Σ)𝔓0(Σ).\mathfrak{P}^{0}_{C}:\mathfrak{Z}_{C}^{0}(\Sigma)\to\mathfrak{P}^{0}(\Sigma).

By (4.4), for every (𝔭,c)C0(Σ)(\mathfrak{p},c)\in\mathfrak{Z}_{C}^{0}(\Sigma), we have a surjective map

(𝔓C0)1(𝔭),A(𝔭,Ae+c),{\mathbb{R}}\to(\mathfrak{P}^{0}_{C})^{-1}(\mathfrak{p}),\qquad A\mapsto(\mathfrak{p},Ae+c),

where ee is minus the real Euler class of 𝔭\mathfrak{p}. Finally, we define the evaluation map

ev:C0(Σ),ev(𝔭,c):=cn,[M].{\mathrm{ev}}:\mathfrak{Z}^{0}_{C}(\Sigma)\to{\mathbb{R}},\qquad{\mathrm{ev}}(\mathfrak{p},c):=\langle c^{n},[M]\rangle.
Definition 5.3.

We say that C0(Σ)\mathfrak{Z}^{0}_{C}(\Sigma) is non-degenerate, if the map ev:C0(Σ){\mathrm{ev}}:\mathfrak{Z}^{0}_{C}(\Sigma)\to{\mathbb{R}} is non-zero, namely if there exists (𝔭,c)C0(Σ)(\mathfrak{p},c)\in\mathfrak{Z}^{0}_{C}(\Sigma) with ev(𝔭,c)0{\mathrm{ev}}(\mathfrak{p},c)\neq 0.

Lemma 5.4.

Let 𝔭0:ΣM0\mathfrak{p}_{0}:\Sigma\to M_{0} be an element in 𝔓0(Σ)\mathfrak{P}^{0}(\Sigma), and let e0HdR2(M0)e_{0}\in H^{2}_{\mathrm{dR}}(M_{0}) be minus the real Euler class of 𝔭0\mathfrak{p}_{0}. For every 𝔭:ΣM\mathfrak{p}:\Sigma\to M in 𝔓0(Σ)\mathfrak{P}^{0}(\Sigma), there exists a diffeomorphism Ψ:ΣΣ\Psi:\Sigma\to\Sigma isotopic to idΣ\mathrm{id}_{\Sigma} such that Ψ𝔭=𝔭0\Psi^{*}\mathfrak{p}=\mathfrak{p}_{0}. If ψ:M0M\psi:M_{0}\to M is the quotient map of Ψ\Psi (see (4.1)), then, for every (𝔭,c)(𝔓C0)1(𝔭)(\mathfrak{p},c)\in(\mathfrak{P}_{C}^{0})^{-1}(\mathfrak{p}) and AA\in{\mathbb{R}}, there holds

(i)\displaystyle(i) ψ(Ae+c)\displaystyle\psi^{*}(Ae+c) =Ae0+ψc,\displaystyle=Ae_{0}+\psi^{*}c,
(ii)\displaystyle(ii) ev(𝔭,Ae+c)\displaystyle{\mathrm{ev}}(\mathfrak{p},Ae+c) =ev(𝔭0,Ae0+ψc),\displaystyle={\mathrm{ev}}(\mathfrak{p}_{0},Ae_{0}+\psi^{*}c),
(iii)\displaystyle(iii) 𝔙𝔬𝔩(𝔭,Ae+c)\displaystyle{\mathfrak{Vol}}(\mathfrak{p},Ae+c) =𝔙𝔬𝔩(𝔭0,Ae0+ψc).\displaystyle={\mathfrak{Vol}}(\mathfrak{p}_{0},Ae_{0}+\psi^{*}c).
Proof.

Since 𝔓0(Σ)\mathfrak{P}^{0}(\Sigma) is connected, the existence of a diffeomorphism Ψ\Psi as in the statement follows from Lemma 4.7.(ii). By Re, we see that ψe=e0\psi^{*}e=e_{0}, which immediately implies item (i). For item (ii), we observe that [M]=ψ[M0][M]=\psi_{*}[M_{0}] since ψ\psi is an orientation-preserving diffeomorphism, and compute

ev(𝔭,Ae+c)=(Ae+c)n,ψ[M0]=ψ(Ae+c)n,[M0]=ev(𝔭0,Ae0+ψc).{\mathrm{ev}}(\mathfrak{p},Ae+c)=\langle(Ae+c)^{n},\psi_{*}[M_{0}]\rangle=\langle\psi^{*}(Ae+c)^{n},[M_{0}]\rangle={\mathrm{ev}}(\mathfrak{p}_{0},Ae_{0}+\psi^{*}c).

For the last relation, we take any ωAΞ2(M)\mathrm{\omega}_{A}\in\Xi^{2}(M) such that [ωA]=Ae+c[\mathrm{\omega}_{A}]=Ae+c. Then ΩA:=𝔭ωA\Omega_{A}:=\mathfrak{p}^{*}\mathrm{\omega}_{A} is associated with (𝔭,Ae+c)(\mathfrak{p},Ae+c). The form ΨΩAΞC2(Σ)\Psi^{*}\Omega_{A}\in\Xi^{2}_{C}(\Sigma) is associated with (𝔭0,Ae0+ψc)(\mathfrak{p}_{0},Ae_{0}+\psi^{*}c). Indeed, ΨΩA=Ψ𝔭ωA=𝔭0ψωA\Psi^{*}\Omega_{A}=\Psi^{*}\mathfrak{p}^{*}\mathrm{\omega}_{A}=\mathfrak{p}_{0}^{*}\psi^{*}\mathrm{\omega}_{A} and [ψωA]=ψ(Ae+c)=Ae0+ψc[\psi^{*}\mathrm{\omega}_{A}]=\psi^{*}(Ae+c)=Ae_{0}+\psi^{*}c. Thus, from Proposition 2.8, we derive

𝔙𝔬𝔩(𝔭,Ae+c)=𝔙𝔬𝔩(ΩA)=𝔙𝔬𝔩(ΨΩA)=𝔙𝔬𝔩(𝔭0,Ae0+ψc).{\mathfrak{Vol}}(\mathfrak{p},Ae+c)={\mathfrak{Vol}}(\Omega_{A})={\mathfrak{Vol}}(\Psi^{*}\Omega_{A})={\mathfrak{Vol}}(\mathfrak{p}_{0},Ae_{0}+\psi^{*}c).\qed
Corollary 5.5.

If C0(Σ)\mathfrak{Z}_{C}^{0}(\Sigma) is non-empty, the four statements below hold.

  1. (i)

    The real Euler class of some bundle in 𝔓0(Σ)\mathfrak{P}^{0}(\Sigma) vanishes, if and only if the real Euler class of every element of 𝔓0(Σ)\mathfrak{P}^{0}(\Sigma) vanishes.

  2. (ii)

    The set C0(Σ)\mathfrak{Z}^{0}_{C}(\Sigma) is non-degenerate, if and only if, for every 𝔭𝔓0(Σ)\mathfrak{p}\in\mathfrak{P}^{0}(\Sigma) there exists a pair (𝔭,c)C0(Σ)(\mathfrak{p},c)\in\mathfrak{Z}^{0}_{C}(\Sigma) such that ev(𝔭,c)0{\mathrm{ev}}(\mathfrak{p},c)\neq 0.

  3. (iii)

    If the real Euler class of the bundles in 𝔓0(Σ)\mathfrak{P}^{0}(\Sigma) vanishes, then 𝔓C0\mathfrak{P}^{0}_{C} is a diffeomorphism and ev:C0(Σ){\mathrm{ev}}:\mathfrak{Z}^{0}_{C}(\Sigma)\to{\mathbb{R}} is a constant map.

  4. (iv)

    If the real Euler class of the bundles in 𝔓0(Σ)\mathfrak{P}^{0}(\Sigma) does not vanish, then 𝔓C0\mathfrak{P}^{0}_{C} has the structure of an affine {\mathbb{R}}-bundle. The {\mathbb{R}}-action on some (𝔭,c)C0(Σ)(\mathfrak{p},c)\in\mathfrak{Z}^{0}_{C}(\Sigma) is given by

    A(𝔭,c)=(𝔭,Ae+c),A.A\cdot(\mathfrak{p},c)=(\mathfrak{p},Ae+c),\qquad\forall\,A\in{\mathbb{R}}.
Proof.

Items (i) and (ii) are direct consequences of items (i) and (ii) in Lemma 5.4.

To prove item (iii), let us assume that bundles in 𝔓0(Σ)\mathfrak{P}^{0}(\Sigma) have vanishing real Euler class. By (4.4), the 𝔓C0\mathfrak{P}^{0}_{C}-fibres are sets containing only one element. This shows that 𝔓C0\mathfrak{P}^{0}_{C} is a diffeomorphism. This together with item (ii) in Lemma 5.4 yields that ev{\mathrm{ev}} is constant.

We suppose that the real Euler class does not vanish and prove item (iv). By (4.4), the 𝔓C0\mathfrak{P}^{0}_{C}-fibres are in bijection with {\mathbb{R}} through the maps A(𝔭,Ae+c)A\mapsto(\mathfrak{p},Ae+c). We construct an explicit local trivialisation of the bundle structure. Lemma 4.7.(i) yields a neighbourhood 𝒲\mathcal{W} of 𝔭0𝔓0(Σ)\mathfrak{p}_{0}\in\mathfrak{P}^{0}(\Sigma) and a map 𝔭Ψ𝔭\mathfrak{p}\mapsto\Psi_{\mathfrak{p}} from 𝒲\mathcal{W} to a neighbourhood of idΣ\mathrm{id}_{\Sigma} inside Diff(Σ)\mathrm{Diff}(\Sigma) such that Ψ𝔭0=idΣ\Psi_{\mathfrak{p}_{0}}=\mathrm{id}_{\Sigma} and Ψ𝔭𝔭=𝔭0\Psi_{\mathfrak{p}}^{*}\mathfrak{p}=\mathfrak{p}_{0}. We define Φ𝔭:=Ψ𝔭1\Phi_{\mathfrak{p}}:=\Psi_{\mathfrak{p}}^{-1} and let ϕ𝔭:MM0\phi_{\mathfrak{p}}:M\to M_{0} be the quotient map. If (𝔭0,c0)(𝔓C0)1(𝔭0)(\mathfrak{p}_{0},c_{0})\in(\mathfrak{P}_{C}^{0})^{-1}(\mathfrak{p}_{0}), then the map

𝒲×(𝔓C0)1(𝒲),(𝔭,A)=(𝔭,Ae+ϕ𝔭c0)\mathcal{W}\times{\mathbb{R}}\to(\mathfrak{P}_{C}^{0})^{-1}(\mathcal{W}),\qquad(\mathfrak{p},A)=(\mathfrak{p},Ae+\phi_{\mathfrak{p}}^{*}c_{0})

provides a local trivialisation around 𝔭0\mathfrak{p}_{0}. ∎

Definition 5.6.

A two-form Ω\Omega on Σ\Sigma is called Zoll, if it is odd-symplectic and there exists 𝔭Ω:ΣMΩ\mathfrak{p}_{\Omega}:\Sigma\to M_{\Omega} in 𝔓(Σ)\mathfrak{P}(\Sigma) such that the oriented 𝔭Ω\mathfrak{p}_{\Omega}-fibres are positively tangent to kerΩ\ker\Omega. In this case, we say that 𝔭Ω\mathfrak{p}_{\Omega}, which is determined up to equivalence, is the bundle associated with Ω\Omega. We write 𝒵(Σ){\mathcal{Z}}(\Sigma) for the set of Zoll (odd-symplectic) forms on Σ\Sigma and 𝒵C(Σ){\mathcal{Z}}_{C}(\Sigma) for the subset of those forms with class CHdR2(Σ)C\in H^{2}_{\mathrm{dR}}(\Sigma).

If Ω\Omega is Zoll, then the form Ω\Omega descends to the base manifold MΩM_{\Omega}, since dΩ=0{\mathrm{d}}\Omega=0. Namely, there exists ωΩΩ2(MΩ)\mathrm{\omega}_{\Omega}\in\Omega^{2}(M_{\Omega}) such that

Ω=𝔭ΩωΩ.\Omega=\mathfrak{p}_{\Omega}^{*}\mathrm{\omega}_{\Omega}.

The form ωΩ\mathrm{\omega}_{\Omega} is closed as well, since 𝔭Ω\mathfrak{p}^{*}_{\Omega} is injective. Furthermore, ωΩ\mathrm{\omega}_{\Omega} is symplectic since Ω\Omega is odd-symplectic and the orientations induced by ωΩ\mathrm{\omega}_{\Omega} and by 𝔭Ω\mathfrak{p}_{\Omega} on MΩM_{\Omega} coincide. We have a natural inclusion

𝒵(Σ)(Σ),Ω(𝔭Ω,[ωΩ]).{\mathcal{Z}}(\Sigma)\to\mathfrak{Z}(\Sigma),\qquad\Omega\mapsto(\mathfrak{p}_{\Omega},[\mathrm{\omega}_{\Omega}]).

As ωΩ\mathrm{\omega}_{\Omega} is positive symplectic, we deduce that

ev(𝔭Ω,[ωΩ])>0.{\mathrm{ev}}(\mathfrak{p}_{\Omega},[\mathrm{\omega}_{\Omega}])>0.
Remark 5.7.

The inequality above implies that for Ω𝒵(Σ)\Omega\in\mathcal{Z}(\Sigma) the component C0(Σ)\mathfrak{Z}^{0}_{C}(\Sigma) which (±𝔭Ω,[ωΩ])(\pm\mathfrak{p}_{\Omega},[\mathrm{\omega}_{\Omega}]) is belonging to, is non-degenerate.

Remark 5.8.

By [BW58], Zoll forms with vanishing cohomology class are just the exterior differentials of Zoll contact forms. Indeed, if Ω=𝔭ΩωΩ\Omega=\mathfrak{p}_{\Omega}^{*}\mathrm{\omega}_{\Omega} is exact, then the cohomology class of ωΩ\mathrm{\omega}_{\Omega} is non-zero, as ωΩ\mathrm{\omega}_{\Omega} is symplectic, and lies in the kernel of the map 𝔭Ω\mathfrak{p}^{*}_{\Omega}. In view of (4.4), this means that there exists T{0}T\in{\mathbb{R}}\setminus\{0\} such that [1TωΩ]=e[\tfrac{1}{T}\omega_{\Omega}]=e. Then, there exists a connection η𝒦(𝔭Ω)\eta\in\mathcal{K}(\mathfrak{p}_{\Omega}) with κη=1TωΩ\kappa_{\eta}=\tfrac{1}{T}\omega_{\Omega} and α:=Tη\alpha:=T\eta is a Zoll contact form with dα=Ω{\mathrm{d}}\alpha=\Omega. The orientations 𝔬Σ\mathfrak{o}_{\Sigma} and 𝔬α\mathfrak{o}_{\alpha} coincide exactly when T>0T>0.

Before moving further, it is worthwhile to briefly discuss stability properties of Zoll forms. Let {Ωr}\{\Omega_{r}\} be a path in 𝒵C(Σ){\mathcal{Z}}_{C}(\Sigma) with the corresponding path of associated bundles {𝔭r}\{\mathfrak{p}_{r}\}. If the real Euler class of the bundles vanishes, we aim at finding an isotopy {Ψr}\{\Psi_{r}\} of Σ\Sigma, such that

ΨrΩr=Ω0.\Psi_{r}^{*}\Omega_{r}=\Omega_{0}.

If the real Euler class of the bundles does not vanish, we aim at finding an isotopy {Ψr}\{\Psi_{r}\} of Σ\Sigma, a path of real numbers {Ar}\{A_{r}\} with A0=0A_{0}=0, and η0𝒦(𝔭0)\eta_{0}\in\mathcal{K}(\mathfrak{p}_{0}) such that

ΨrΩr=Ω0+Ardη0.\Psi_{r}^{*}\Omega_{r}=\Omega_{0}+A_{r}{\mathrm{d}}\eta_{0}.

In this last case, it seems unlikely that all paths admit such an expression. That this happens if C=0C=0 is a result of Weinstein [Wei74]. The stability for e=0e=0 is reminiscent of [Gin87].

Proposition 5.9.

Let {Ωr}\{\Omega_{r}\} be a path in 𝒵C(Σ){\mathcal{Z}}_{C}(\Sigma) with the path of associated bundles {𝔭r}\{\mathfrak{p}_{r}\}.

  • If C=0C=0, there is η0𝒦(𝔭0)\eta_{0}\in\mathcal{K}(\mathfrak{p}_{0}), an isotopy {Ψr}\{\Psi_{r}\} of Σ\Sigma and real numbers {Tr}\{T_{r}\} such that

    ΨrΩr=Trdη0=Ω0+(TrT0)dη0.\Psi_{r}^{*}\Omega_{r}=T_{r}{\mathrm{d}}\eta_{0}=\Omega_{0}+(T_{r}-T_{0}){\mathrm{d}}\eta_{0}.
  • If e=0e=0, there is an isotopy {Ψr}\{\Psi_{r}\} of Σ\Sigma such that ΨrΩr=Ω0\Psi_{r}^{*}\Omega_{r}=\Omega_{0}.

Proof.

By Lemma 4.7, we can suppose in both cases that {Ωr=𝔭0ωr}\{\Omega_{r}=\mathfrak{p}_{0}^{*}\mathrm{\omega}_{r}\}, where {ωr}\{\mathrm{\omega}_{r}\} is a path of symplectic forms on MM. If C=0C=0, [KN96] yield {ηr}𝒦(𝔭0)\{\eta_{r}\}\subset\mathcal{K}(\mathfrak{p}_{0}) and non-zero real numbers {Tr}\{T_{r}\} such that Ωr=d(Trηr)\Omega_{r}={\mathrm{d}}(T_{r}\eta_{r}) and ηr=η0+𝔭ζr\eta_{r}=\eta_{0}+\mathfrak{p}^{*}\zeta_{r}, where {ζr}\{\zeta_{r}\} are one-forms on MM. Setting αr:=Trηr\alpha_{r}:=T_{r}\eta_{r} and applying the Gray stability theorem from Remark 3.5.(a), we see that there exists an isotopy {Ψr}\{\Psi_{r}\} and a path {Hr:Σ}\{H_{r}:\Sigma\to{\mathbb{R}}\} with H0=0H_{0}=0 such that

Ψrαr=(1+1T0Hr)α0.\Psi_{r}^{*}\alpha_{r}=(1+\tfrac{1}{T_{0}}H_{r})\alpha_{0}.

Since α˙r=T˙rηr+Tr𝔭ζ˙r\dot{\alpha}_{r}=\dot{T}_{r}\eta_{r}+T_{r}\mathfrak{p}^{*}\dot{\zeta}_{r}, equation (3.5) implies H˙r=T˙r\dot{H}_{r}=\dot{T}_{r} and hence Hr=TrT0H_{r}=T_{r}-T_{0}, which readily implies the desired formula.

If e=0e=0, then ωr=ω0+dζr\mathrm{\omega}_{r}=\mathrm{\omega}_{0}+{\mathrm{d}}\zeta_{r} due to (4.4). By Remark 3.5.(b) and Proposition 3.4 with α0=0\alpha_{0}=0, it follows that ΨrΩr=Ω0+d(Hrη0)\Psi_{r}^{*}\Omega_{r}=\Omega_{0}+{\mathrm{d}}(H_{r}\eta_{0}), for some paths {Ψr}\{\Psi_{r}\} and {Hr}\{H_{r}\} as above. Again by equation (3.5), we conclude that Hr=0H_{r}=0 for all rr. ∎

We are now ready to classify Zoll odd-symplectic forms on a three-dimensional manifold, as promised in the introduction.

5.2 Classification of Zoll odd-symplectic forms in dimension three

Proof of Proposition 1.9

Let Σ\Sigma have dimension three, let bΣb_{\Sigma} be the rank of the free part of H1(Σ;)H_{1}(\Sigma;{\mathbb{Z}}), and let |H1tor(Σ;)||H_{1}^{\mathrm{tor}}(\Sigma;{\mathbb{Z}})| denote the cardinality of its torsion subgroup. Let Ω𝒵C(Σ)\Omega\in{\mathcal{Z}}_{C}(\Sigma) be a Zoll form with cohomology class CC and let 𝔭=𝔭Ω\mathfrak{p}=\mathfrak{p}_{\Omega} its associated bundle. If C=0C=0, we know from Remark 5.8 that Ω\Omega is the differential of a Zoll contact form. In particular, Σ\Sigma is the total space of a non-trivial oriented S1S^{1}-bundle, bΣb_{\Sigma} is even and we already treated this case in [BK19a, Proposition 1.2].

Suppose that C0C\neq 0. In this case, e=0e=0 by equivalence (ii) in (4.7) . Since MM is a surface this implies that e=0e_{\mathbb{Z}}=0, and hence 𝔭\mathfrak{p} is trivial. Therefore, ΣM×S1\Sigma\cong M\times S^{1} and we see that

bΣ=1+2genus(M),|H1tor(Σ;)|=1.b_{\Sigma}=1+2\,\mathrm{genus}(M),\qquad|H_{1}^{\mathrm{tor}}(\Sigma;{\mathbb{Z}})|=1. (5.2)

In particular, bΣb_{\Sigma} is odd. Let Ω𝒵C(Σ)\Omega^{\prime}\in{\mathcal{Z}}_{C^{\prime}}(\Sigma) be another Zoll form with class CC^{\prime}. Since bΣb_{\Sigma} is odd, then C0C^{\prime}\neq 0 and 𝔭:ΣM\mathfrak{p}^{\prime}:\Sigma\to M^{\prime} is the trivial bundle. We write Ω=𝔭ω\Omega=\mathfrak{p}^{*}\mathrm{\omega} and Ω=𝔭ω\Omega^{\prime}=\mathfrak{p}^{\prime*}\mathrm{\omega}^{\prime}, where ω\mathrm{\omega} and ω\mathrm{\omega}^{\prime} are symplectic forms on MM and MM^{\prime} respectively. From (5.2), MM and MM^{\prime} have the same genus, so that there is a diffeomorphism ψ:MM\psi:M\to M^{\prime}. Since ω\omega and ω\omega^{\prime} are symplectic on MM and MM^{\prime} and HdR2(M;)HdR2(M;)H^{2}_{\mathrm{dR}}(M;{\mathbb{R}})\cong{\mathbb{R}}\cong H^{2}_{\mathrm{dR}}(M^{\prime};{\mathbb{R}}), by Moser’s trick, we can assume that

ψω=Tω,for some T>0.\psi^{*}\omega^{\prime}=T\omega,\qquad\text{for some }T>0.

As both 𝔭\mathfrak{p} and 𝔭\mathfrak{p}^{\prime} are trivial bundles, it is immediate to find a diffeomorphism

Ψ:ΣΣ\Psi:\Sigma\to\Sigma (5.3)

lifting ψ\psi and such that Ψ𝔭=𝔭\Psi^{*}\mathfrak{p}^{\prime}=\mathfrak{p}. Therefore, ff preserves the orientation of Σ\Sigma and ΨΩ=TΩ\Psi^{*}\Omega^{\prime}=T\Omega.

We now want to describe the connected components of the space of Zoll forms on Σ\Sigma with fixed cohomology class. In order to do so, we first determine the connected components of the space of oriented S1S^{1}-bundles 𝔓(Σ)\mathfrak{P}(\Sigma) with the help of classical results in low-dimensional topology. As a preliminary observation, we point out that if 𝔭𝔓(Σ)\mathfrak{p}\in\mathfrak{P}(\Sigma), then the class [𝔭1(pt)]H1(Σ;)[\mathfrak{p}^{-1}({\mathrm{pt}})]_{\mathbb{Z}}\in H_{1}(\Sigma;{\mathbb{Z}}) is primitive, since its intersection number with a global section of 𝔭\mathfrak{p} is equal to 11. We distinguish three cases: M=S2M=S^{2}, M=𝕋2M={\mathbb{T}}^{2}, and genus(M)2\mathrm{genus}(M)\geq 2.

Case 1: ΣS2×S1\Sigma\cong S^{2}\times S^{1}. We regard S2S^{2} as the unit sphere in 3{\mathbb{R}}^{3}. We recall that the mapping class group of orientation-preserving diffeomorphisms of S2×S1S^{2}\times S^{1} is given by

MCG(S2×S1)2[Ψ1]2[Ψ2].\mathrm{MCG}(S^{2}\times S^{1})\cong\tfrac{{\mathbb{Z}}}{2{\mathbb{Z}}}\,[\Psi_{1}]\oplus\tfrac{{\mathbb{Z}}}{2{\mathbb{Z}}}\,[\Psi_{2}].

Here, the generators Ψ1,Ψ2:S2×S1S2×S1\Psi_{1},\Psi_{2}:S^{2}\times S^{1}\to S^{2}\times S^{1} are given by

Ψ1(q,ϕ):=(q,ϕ),Ψ2(q,ϕ):=(ρϕ(q),ϕ),(q,ϕ)S2×S1,\Psi_{1}(q,\phi):=(-q,-\phi),\qquad\Psi_{2}(q,\phi):=(\rho_{\phi}(q),\phi),\qquad\forall\,(q,\phi)\in S^{2}\times S^{1},

where ρϕ:S2S2\rho_{\phi}:S^{2}\to S^{2} is the rotation of angle 2πϕ2\pi\phi around the north pole. We consider the standard projection 𝔭+:S1×S2S2\mathfrak{p}_{+}:S^{1}\times S^{2}\to S^{2} along S1S^{1} and we define 𝔭:=Ψ1𝔭+\mathfrak{p}_{-}:=\Psi_{1}^{*}\mathfrak{p}_{+}.

We claim that 𝔓(S2×S1)\mathfrak{P}(S^{2}\times S^{1}) has four connected components containing respectively 𝔭+\mathfrak{p}_{+}, Ψ2𝔭+\Psi_{2}^{*}\mathfrak{p}_{+}, 𝔭\mathfrak{p}_{-} and Ψ2𝔭\Psi_{2}^{*}\mathfrak{p}_{-}. To this purpose, we observe that [𝔭+1(pt)]=[(Ψ2𝔭+)1(pt)][\mathfrak{p}_{+}^{-1}({\mathrm{pt}})]_{\mathbb{Z}}=[(\Psi_{2}^{*}\mathfrak{p}_{+})^{-1}({\mathrm{pt}})]_{\mathbb{Z}} and [𝔭1(pt)]=[(Ψ2𝔭)1(pt)][\mathfrak{p}_{-}^{-1}({\mathrm{pt}})]_{\mathbb{Z}}=[(\Psi_{2}^{*}\mathfrak{p}_{-})^{-1}({\mathrm{pt}})]_{\mathbb{Z}} are distinct and yield the two primitive homology classes in H1(S2×S1;)H_{1}(S^{2}\times S^{1};{\mathbb{Z}})\cong{\mathbb{Z}}. Therefore, we just need to show that 𝔭+\mathfrak{p}_{+}, Ψ2𝔭+\Psi_{2}^{*}\mathfrak{p}_{+} are not in the same connected component. If, by contradiction, there were a path in 𝔓(S2×S1)\mathfrak{P}(S^{2}\times S^{1}) from 𝔭+\mathfrak{p}_{+} to Ψ2𝔭+\Psi_{2}^{*}\mathfrak{p}_{+}, then by Lemma 4.7.(ii) there would exist a diffeomorphism Ψ2:S2×S1S2×S1\Psi_{2}^{\prime}:S^{2}\times S^{1}\to S^{2}\times S^{1} isotopic to Ψ2\Psi_{2} such that 𝔭+=(Ψ2)𝔭+\mathfrak{p}_{+}=(\Psi_{2}^{\prime})^{*}\mathfrak{p}_{+}. This forces Ψ2\Psi_{2}^{\prime} to be of the form

Ψ2(q,ϕ)=(ψ(q),ϕ(x,ϕ))\Psi_{2}^{\prime}(q,\phi)=(\psi(q),\phi^{\prime}(x,\phi))

where ϕϕ(q,ϕ)\phi\mapsto\phi^{\prime}(q,\phi) is an orientation-preserving diffeomorphism of S1S^{1}, for all qS2q\in S^{2}, and ψ\psi is an orientation-preserving diffeomorphism of S2S^{2}. However, every orientation-preserving diffeomorphisms of S2S^{2} is isotopic to the identity and the set Diff+(S1)\mathrm{Diff}_{+}(S^{1}) of orientation-preserving diffeomorphisms of S1S^{1} is homotopy equivalent to S1S^{1}, so that π2(Diff+(S1))=0\pi_{2}(\mathrm{Diff}_{+}(S^{1}))=0. Thus, the map Ψ2\Psi_{2}^{\prime} would be isotopic to the identity, which is impossible, as it is isotopic to Ψ2\Psi_{2}.

Let 𝔭\mathfrak{p} be an arbitrary element in 𝔓(S2×S1)\mathfrak{P}(S^{2}\times S^{1}). We have either [𝔭1(pt)]=[𝔭+1(pt)][\mathfrak{p}^{-1}(\mathrm{pt})]_{\mathbb{Z}}=[\mathfrak{p}^{-1}_{+}(\mathrm{pt})]_{\mathbb{Z}} or [𝔭1(pt)]=[𝔭1(pt)][\mathfrak{p}^{-1}(\mathrm{pt})]_{\mathbb{Z}}=[\mathfrak{p}^{-1}_{-}(\mathrm{pt})]_{\mathbb{Z}}. We have seen in (5.3) that there exists an isomorphism of oriented S1S^{1}-bundles Ψ:S2×S1S2×S1\Psi:S^{2}\times S^{1}\to S^{2}\times S^{1} preserving the orientation and such that Ψ𝔭+=𝔭\Psi^{*}\mathfrak{p}_{+}=\mathfrak{p}. Therefore, if [𝔭1(pt)]=[𝔭+1(pt)][\mathfrak{p}^{-1}(\mathrm{pt})]_{\mathbb{Z}}=[\mathfrak{p}^{-1}_{+}(\mathrm{pt})]_{\mathbb{Z}}, then Ψ\Psi is either isotopic to idS2×S1\mathrm{id}_{S^{2}\times S^{1}} or to Ψ2\Psi_{2} and 𝔭\mathfrak{p} is either homotopic to 𝔭+\mathfrak{p}_{+} or to Ψ2𝔭+\Psi_{2}^{*}\mathfrak{p}_{+}; if [𝔭1(pt)]=[𝔭1(pt)][\mathfrak{p}^{-1}(\mathrm{pt})]_{\mathbb{Z}}=[\mathfrak{p}^{-1}_{-}(\mathrm{pt})]_{\mathbb{Z}}, then Ψ\Psi is either isotopic to Ψ1\Psi_{1} or to Ψ1Ψ2\Psi_{1}\circ\Psi_{2} and 𝔭\mathfrak{p} is either homotopic to 𝔭\mathfrak{p}_{-} or to Ψ2𝔭\Psi_{2}^{*}\mathfrak{p}_{-}.

Case 2: Σ𝕋2×S1=𝕋3\Sigma\cong{\mathbb{T}}^{2}\times S^{1}={\mathbb{T}}^{3}. To every orientation-preserving diffeomorphism Ψ\Psi on 𝕋3{\mathbb{T}}^{3}, we can associate the induced map in homology

H1(Ψ):H1(𝕋3;)H1(𝕋3;).H_{1}(\Psi):H_{1}({\mathbb{T}}^{3};{\mathbb{Z}})\to H_{1}({\mathbb{T}}^{3};{\mathbb{Z}}).

As H1(𝕋3;)3H_{1}({\mathbb{T}}^{3};{\mathbb{Z}})\cong{\mathbb{Z}}^{3}, we can identify H1(Ψ)H_{1}(\Psi) with an element of SL(3;)\mathrm{SL}(3;{\mathbb{Z}}). Conversely, every ASL(3;)A\in\mathrm{SL}(3;{\mathbb{Z}}) gives an orientation-preserving diffeomorphism ΨA:𝕋3𝕋3\Psi_{A}:{\mathbb{T}}^{3}\to{\mathbb{T}}^{3} with H1(ΨA)=AH_{1}(\Psi_{A})=A. Moreover, the mapping class group is computed explicitly through the isomorphism

MCG(𝕋3)SL(3;),ΨH1(Ψ).\mathrm{MCG}({\mathbb{T}}^{3})\to\mathrm{SL}(3;{\mathbb{Z}}),\qquad\Psi\mapsto H_{1}(\Psi). (5.4)

Let 𝔭0:T3𝕋2×S1𝕋2\mathfrak{p}_{0}:{\mathrm{T}}^{3}\cong{\mathbb{T}}^{2}\times S^{1}\to{\mathbb{T}}^{2} be the standard projection along S1S^{1}. For every primitive class 𝔥H1(𝕋3;){\mathfrak{h}}\in H_{1}({\mathbb{T}}^{3};{\mathbb{Z}}), there exists A𝔥SL(3;)A_{\mathfrak{h}}\in\mathrm{SL}(3;{\mathbb{Z}}) with A𝔥𝔥=[𝔭1(pt)]A_{\mathfrak{h}}\cdot\mathfrak{h}=[\mathfrak{p}^{-1}({\mathrm{pt}})]_{\mathbb{Z}}, so that the fibres of the oriented S1S^{1}-bundle

𝔭𝔥:=ΨA𝔥𝔭0\mathfrak{p}_{\mathfrak{h}}:=\Psi_{A_{\mathfrak{h}}}^{*}\mathfrak{p}_{0}

lie in the homology class 𝔥{\mathfrak{h}}. The map A𝔥A_{\mathfrak{h}} is not unique. However, if A𝔥A_{\mathfrak{h}}^{\prime} is another such map, then ΨA𝔥𝔭0\Psi_{A_{\mathfrak{h}}^{\prime}}^{*}\mathfrak{p}_{0} and ΨA𝔥𝔭0\Psi_{A_{\mathfrak{h}}}^{*}\mathfrak{p}_{0} are equivalent bundles, and up to equivalence, 𝔭𝔥\mathfrak{p}_{\mathfrak{h}} depends only on 𝔥\mathfrak{h}.

On the other hand, if 𝔭𝔓(𝕋3)\mathfrak{p}\in\mathfrak{P}({\mathbb{T}}^{3}), we claim that 𝔭[𝔭1(pt)]\mathfrak{p}_{[\mathfrak{p}^{-1}({\mathrm{pt}})]_{\mathbb{Z}}} and 𝔭\mathfrak{p} are isotopic. Indeed, let Ψ:𝕋3𝕋3\Psi:{\mathbb{T}}^{3}\to{\mathbb{T}}^{3} be an orientation-preserving diffeomorphism with Ψ𝔭0=𝔭\Psi^{*}\mathfrak{p}_{0}=\mathfrak{p}, as in (5.3). The equality H1(Ψ)([𝔭1(pt)])=[𝔭01(pt)]H_{1}(\Psi)([\mathfrak{p}^{-1}({\mathrm{pt}})]_{\mathbb{Z}})=[\mathfrak{p}^{-1}_{0}({\mathrm{pt}})]_{\mathbb{Z}} implies that H1(Ψ)=A[𝔭1(pt)]H_{1}(\Psi)=A_{[\mathfrak{p}^{-1}({\mathrm{pt}})]_{\mathbb{Z}}} so that there holds 𝔭[𝔭1(pt)]=ΨH1(Ψ)𝔭0\mathfrak{p}_{[\mathfrak{p}^{-1}({\mathrm{pt}})]_{\mathbb{Z}}}=\Psi_{H_{1}(\Psi)}^{*}\mathfrak{p}_{0}. Moreover, since Ψ\Psi and ΨH1(Ψ)\Psi_{H_{1}(\Psi)} are isotopic due to (5.4), ΨH1(Ψ)𝔭0\Psi_{H_{1}(\Psi)}^{*}\mathfrak{p}_{0} is homotopic to 𝔭\mathfrak{p} and the claim is proven.

Case 3: ΣM×S1\Sigma\cong M\times S^{1} with genus(M)2\mathrm{genus}(M)\geq 2. Let 𝔭+:M×S1M\mathfrak{p}_{+}:M\times S^{1}\to M be the oriented S1S^{1}-bundles given by the standard projection and set 𝔭:=𝔭+\mathfrak{p}_{-}:=-\mathfrak{p}_{+}. The bundles 𝔭+\mathfrak{p}_{+} and 𝔭\mathfrak{p}_{-} are not homotopic since the homotopy classes of their fibres are different. If 𝔭\mathfrak{p} is any element of 𝔓(M×S1)\mathfrak{P}(M\times S^{1}), by [Wal67, Satz (5.5)], there exists a diffeomorphism Ψ:M×S1M×S1\Psi:M\times S^{1}\to M\times S^{1} isotopic to the identity such that either Ψ𝔭=𝔭+\Psi^{*}\mathfrak{p}=\mathfrak{p}_{+} or Ψ𝔭=𝔭\Psi^{*}\mathfrak{p}=\mathfrak{p}_{-}. Thus, 𝔭\mathfrak{p} is either isotopic to 𝔭\mathfrak{p}_{-} or to 𝔭+\mathfrak{p}_{+}.

This finishes the description of the connected components of 𝔓(Σ)\mathfrak{P}(\Sigma) in the three cases. To determine the connected components of the space of Zoll forms with fixed cohomology class, we consider the map assigning to each Zoll odd-symplectic form its associated bundle

𝔓𝒵:𝒵(Σ)𝔓(Σ),Ω𝔭Ω\mathfrak{P}_{{\mathcal{Z}}}:{\mathcal{Z}}(\Sigma)\to\mathfrak{P}(\Sigma),\qquad\Omega\mapsto\mathfrak{p}_{\Omega}

If 𝔭=𝔓𝒵(Ω)\mathfrak{p}=\mathfrak{P}_{{\mathcal{Z}}}(\Omega), then PD([Ω])\mathrm{PD}([\Omega]) is a positive multiple of [𝔭1(pt)][\mathfrak{p}^{-1}({\mathrm{pt}})] by (4.6). Moreover, the map 𝔓𝒵\mathfrak{P}_{{\mathcal{Z}}} is surjective, since the quotient manifold of any bundle is an orientable surface, and therefore, possesses a positive symplectic form. In particular, if 𝔭𝔓(Σ)\mathfrak{p}\in\mathfrak{P}(\Sigma) and CHdR2(Σ)C\in H^{2}_{\mathrm{dR}}(\Sigma) are such that PD(C)\mathrm{PD}(C) is a positive multiple of [𝔭1(pt)][\mathfrak{p}^{-1}({\mathrm{pt}})], then there exists Ω𝒵C(Σ)𝒵(Σ)\Omega\in{\mathcal{Z}}_{C}(\Sigma)\subset{\mathcal{Z}}(\Sigma) such that 𝔓𝒵(Ω)=𝔭\mathfrak{P}_{{\mathcal{Z}}}(\Omega)=\mathfrak{p}. This shows that

  • if M=S2M=S^{2} or 𝕋2{\mathbb{T}}^{2}, then 𝒵C(Σ){\mathcal{Z}}_{C}(\Sigma) is not empty, for all C0C\neq 0;

  • if MM has higher genus, then 𝒵C(Σ){\mathcal{Z}}_{C}(\Sigma) is not empty if and only if CC is a non-zero element in 𝔭+(HdR2(M))\mathfrak{p}_{+}^{*}\big{(}H^{2}_{\mathrm{dR}}(M)\big{)}.

We fix now some CHdR2(Σ)C\in H^{2}_{\mathrm{dR}}(\Sigma) for which 𝒵C(Σ){\mathcal{Z}}_{C}(\Sigma) is non-empty and consider 𝔓𝒵,C:=𝔓𝒵|𝒵C(Σ)\mathfrak{P}_{{\mathcal{Z}},C}:=\mathfrak{P}_{{\mathcal{Z}}}|_{{\mathcal{Z}}_{C}(\Sigma)}. As its non-empty fibres are convex, the connected components of 𝒵C(Σ){\mathcal{Z}}_{C}(\Sigma) correspond through 𝔓𝒵,C\mathfrak{P}_{{\mathcal{Z}},C} to the connected components of 𝔓𝒵,C(𝒵C(Σ))\mathfrak{P}_{{\mathcal{Z}},C}({\mathcal{Z}}_{C}(\Sigma)). The latter set is the union of those connected components of 𝔓(Σ)\mathfrak{P}(\Sigma), whose elements 𝔭\mathfrak{p} satisfy PD(C)=a[𝔭1(pt)]\mathrm{PD}(C)=a[\mathfrak{p}^{-1}({\mathrm{pt}})] for some a>0a>0. By the discussion above, this shows that:

  • If MM is the two-sphere, then the set 𝔓𝒵,C(𝒵C(Σ))\mathfrak{P}_{{\mathcal{Z}},C}({\mathcal{Z}}_{C}(\Sigma)) has two connected components since [𝔭+1(pt)]=[(Ψ2𝔭+)1(pt)][\mathfrak{p}_{+}^{-1}({\mathrm{pt}})]=[(\Psi_{2}^{*}\mathfrak{p}_{+})^{-1}({\mathrm{pt}})] and [𝔭1(pt)]=[(Ψ2𝔭)1(pt)][\mathfrak{p}_{-}^{-1}({\mathrm{pt}})]=[(\Psi_{2}^{*}\mathfrak{p}_{-})^{-1}({\mathrm{pt}})];

  • If MM has positive genus, then the set 𝔓𝒵,C\mathfrak{P}_{{\mathcal{Z}},C} is connected.

The proof of the proposition is completed.∎

6 Action of closed two-forms

6.1 The action form

Definition 6.1.

If Ω\Omega is a closed two-form on Σ\Sigma, an embedded one-periodic curve γ:S1Σ\gamma:S^{1}\to\Sigma with γ˙kerΩ\dot{\gamma}\in\ker\Omega is said to be a closed characteristics of Ω\Omega. We write 𝒳(Ω)\mathcal{X}(\Omega) for the set of closed characteristics of Ω\Omega, up to orientation-preserving reparametrisations of S1S^{1}.

Fix a free-homotopy class 𝔥[S1,Σ]{\mathfrak{h}}\in[S^{1},\Sigma] and let Λ𝔥(Σ)\Lambda_{\mathfrak{h}}(\Sigma) be the space of one-periodic curves in Σ\Sigma with class 𝔥\mathfrak{h}. In what follows, given a pair (C,𝔥)(C,\mathfrak{h}), we shall study a variational principle on Λ𝔥(Σ)\Lambda_{\mathfrak{h}}(\Sigma), which detects, for all ΩΞC2(Σ)\Omega\in\Xi^{2}_{C}(\Sigma), the elements of 𝒳(Ω)\mathcal{X}(\Omega) belonging to Λ𝔥(Σ)\Lambda_{\mathfrak{h}}(\Sigma). To this purpose, we define the action form 𝔞=𝔞(Ω)Ω1(Λ𝔥(Σ))\mathfrak{a}=\mathfrak{a}(\Omega)\in\Omega^{1}(\Lambda_{\mathfrak{h}}(\Sigma)) by

𝔞γ(ξ):=S1Ω(ξ(t),γ˙(t))dt,γΛ𝔥(Σ),ξTγΛ𝔥(Σ).\mathfrak{a}_{\gamma}(\xi):=\int_{S^{1}}\Omega(\xi(t),\dot{\gamma}(t))\,{\mathrm{d}}t,\qquad\forall\,\gamma\in\Lambda_{\mathfrak{h}}(\Sigma),\ \forall\,\xi\in\mathrm{T}_{\gamma}\Lambda_{\mathfrak{h}}(\Sigma).

Any C1C^{1}-path {γr}Λ𝔥(Σ)\{\gamma_{r}\}\subset\Lambda_{\mathfrak{h}}(\Sigma) can also be interpreted as a cylinder Γ:[0,1]×S1Σ\Gamma:[0,1]\times S^{1}\to\Sigma with Γ(r,t)=γr(t)\Gamma(r,t)=\gamma_{r}(t), so that

01{γr}𝔞(Ω)=[0,1]×S1ΓΩ.\int_{0}^{1}\{\gamma_{r}\}^{*}\mathfrak{a}(\Omega)=\int_{[0,1]\times S^{1}}\Gamma^{*}\Omega. (6.1)

Furthermore, we have the following classical observation.

Lemma 6.2.

The action form 𝔞(Ω)\mathfrak{a}(\Omega) is closed. An embedded periodic curve γΛ𝔥(Σ)\gamma\in\Lambda_{\mathfrak{h}}(\Sigma) is a closed characteristic of Ω\Omega if and only if 𝔞γ(Ω)=0\mathfrak{a}_{\gamma}(\Omega)=0.∎

If Ω\Omega and Ω\Omega^{\prime} are forms in ΞC2(Σ)\Xi^{2}_{C}(\Sigma) with ΩΩ=dα\Omega^{\prime}-\Omega={\mathrm{d}}\alpha, then there holds

𝔞(Ω)𝔞(Ω)=dα,α(γ):=S1γα.\mathfrak{a}(\Omega^{\prime})-\mathfrak{a}(\Omega)={\mathrm{d}}\mathcal{B}_{\alpha},\qquad\mathcal{B}_{\alpha}(\gamma):=\int_{S^{1}}\gamma^{*}\alpha.

Thus, it makes sense to define 𝔞(C):=[𝔞(Ω)]HdR1(Λ𝔥(Σ))\mathfrak{a}(C):=[\mathfrak{a}(\Omega)]\in H^{1}_{\operatorname{dR}}(\Lambda_{\mathfrak{h}}(\Sigma)). We have a homomorphism

HdR2(Σ)HdR1(Λ𝔥(Σ)),C𝔞(C).H^{2}_{{\mathrm{dR}}}(\Sigma)\to H^{1}_{\operatorname{dR}}(\Lambda_{\mathfrak{h}}(\Sigma)),\qquad C\mapsto\mathfrak{a}(C). (6.2)

If 𝔞(Ω)\mathfrak{a}(\Omega) admits a primitive functional, then the zeros of 𝔞(Ω)\mathfrak{a}(\Omega) are critical points of the primitive. In the next lemma, we give a criterion ensuring that 𝔞(Ω)\mathfrak{a}(\Omega) is exact on Λ𝔥(Σ)\Lambda_{\mathfrak{h}}(\Sigma), in the case that there exists an oriented S1S^{1}-bundle with fibres in 𝔥{\mathfrak{h}}. Below, we regard classes in HdR2(M)H^{2}_{{\mathrm{dR}}}(M) as real homomorphisms on π2(M)\pi_{2}(M) through the canonical map π2(M)H2(M;)\pi_{2}(M)\to H_{2}(M;{\mathbb{Z}}).

Lemma 6.3.

Let CHdR2(Σ)C\in H^{2}_{\mathrm{dR}}(\Sigma) and (𝔭,c)C(Σ)(\mathfrak{p},c)\in\mathfrak{Z}_{C}(\Sigma) such that 𝔭:ΣM\mathfrak{p}:\Sigma\to M has minus the real Euler class eHdR2(M)e\in H^{2}_{\mathrm{dR}}(M) and the oriented 𝔭\mathfrak{p}-fibres have class 𝔥[S1,Σ]{\mathfrak{h}}\in[S^{1},\Sigma]. There holds

𝔞(C)=0inHdR1(Λ𝔥(Σ))kere|π2(M)kerc|π2(M).\mathfrak{a}(C)=0\;\ \mathrm{in}\ H^{1}_{\mathrm{dR}}(\Lambda_{\mathfrak{h}}(\Sigma))\qquad\Longleftrightarrow\qquad\ker e|_{\pi_{2}(M)}\subseteq\ker c|_{\pi_{2}(M)}.
Proof.

Let ω\mathrm{\omega} be any element in Ξc2(M)\Xi^{2}_{c}(M) and define Ω:=𝔭ω\Omega:=\mathfrak{p}^{*}\mathrm{\omega}. The cohomology class 𝔞(C)\mathfrak{a}(C) is trivial if and only if its integral over any one-periodic curve Γ:S1Λ𝔥(Σ)\Gamma:S^{1}\to\Lambda_{\mathfrak{h}}(\Sigma) vanishes. Choosing any oriented fibre γ:S1Σ\gamma:S^{1}\to\Sigma such that [γ]=𝔥[\gamma]=\mathfrak{h}, we may assume that Γ(0)=Γ(1)=γ\Gamma(0)=\Gamma(1)=\gamma, up to homotopy, since 𝔞\mathfrak{a} is closed and therefore the integral of 𝔞\mathfrak{a} depends only on the homotopy class of Γ\Gamma. In view of (6.1), 𝔞(C)HdR1(Λ𝔥(Σ))\mathfrak{a}(C)\in H^{1}_{{\mathrm{dR}}}(\Lambda_{\mathfrak{h}}(\Sigma)) is trivial if and only if, for any such Γ\Gamma,

[0,1]×S1ΓΩ=0.\int_{[0,1]\times S^{1}}\Gamma^{*}\Omega=0.

As 𝔭γ\mathfrak{p}\circ\gamma is a constant curve, we think of Γ¯:=𝔭Γ\bar{\Gamma}:=\mathfrak{p}\circ\Gamma as a map from S2S^{2} into MM with homotopy class [Γ¯]π2(M)[\bar{\Gamma}]\in\pi_{2}(M). We have

[0,1]×S1ΓΩ=S2Γ¯ω=c,[Γ¯].\int_{[0,1]\times S^{1}}\Gamma^{*}\Omega=\int_{S^{2}}\bar{\Gamma}^{*}\mathrm{\omega}=\langle c,[\bar{\Gamma}]\rangle.

Hence, the lemma follows if we show that the map Γ[Γ¯]\Gamma\mapsto[\bar{\Gamma}] is onto kere|π2(M)\ker e|_{\pi_{2}(M)}. To see that the image [Γ¯][\bar{\Gamma}] is indeed in kere|π2(M)\ker e|_{\pi_{2}(M)}, we compute

e,[Γ¯]=S2Γ¯κη=[0,1]×S1Γ(dη)=S1γηS1γη=0,\langle e,[\bar{\Gamma}]\rangle=\int_{S^{2}}\bar{\Gamma}^{*}\kappa_{\eta}=\int_{[0,1]\times S^{1}}\Gamma^{*}({\mathrm{d}}\eta)=\int_{S^{1}}\gamma^{*}\eta-\int_{S^{1}}\gamma^{*}\eta=0,

where η𝒦(𝔭)\eta\in\mathcal{K}(\mathfrak{p}) is any connection for 𝔭\mathfrak{p}.

Then, we show that for any υ:S2M\upsilon:S^{2}\to M with e,[υ]=0\langle e,[\upsilon]\rangle=0, there is Γ:[0,1]×S1Σ\Gamma:[0,1]\times S^{1}\to\Sigma with Γ(0,)=Γ(1,)=γ\Gamma(0,\cdot)=\Gamma(1,\cdot)=\gamma such that Γ¯=𝔭Γ=υ\bar{\Gamma}=\mathfrak{p}\circ\Gamma=\upsilon. By the naturality of the Euler class, the restriction of 𝔭\mathfrak{p} over υ\upsilon admits a global section Υ:S2Σ\Upsilon:S^{2}\to\Sigma. It satisfies 𝔭Υ=υ\mathfrak{p}\circ\Upsilon=\upsilon. Using the quotient map [0,1]×S1S2[0,1]\times S^{1}\to S^{2}, we lift Υ\Upsilon to a map Υ:[0,1]×S1Σ\Upsilon^{\prime}:[0,1]\times S^{1}\to\Sigma. Up to modifying υ\upsilon within its homotopy class, we can assume that Υ(0,)=Υ(1,)=γ(0)\Upsilon^{\prime}(0,\cdot)=\Upsilon^{\prime}(1,\cdot)=\gamma(0). Finally, we define

Γ:[0,1]×S1Σ,Γ(s,t)=𝔲(t,Υ(s,t)),\Gamma:[0,1]\times S^{1}\to\Sigma,\qquad\Gamma(s,t)=\mathfrak{u}(t,\Upsilon^{\prime}(s,t)),

where 𝔲𝔘(𝔭)\mathfrak{u}\in\mathfrak{U}(\mathfrak{p}). It follows that Γ(0,)=Γ(1,)=γ\Gamma(0,\cdot)=\Gamma(1,\cdot)=\gamma and Γ¯=Υ¯=Υ¯=υ\bar{\Gamma}=\bar{\Upsilon}^{\prime}=\bar{\Upsilon}=\upsilon, as required. ∎

Remark 6.4.

The condition 𝔞(C)=0\mathfrak{a}(C)=0 in HdR1(Λ𝔥(Σ))H^{1}_{\mathrm{dR}}(\Lambda_{\mathfrak{h}}(\Sigma)) depends only on CC and 𝔥\mathfrak{h}. Therefore, the condition kere|π2(M)kerc|π2(M)\ker e|_{\pi_{2}(M)}\subseteq\ker c|_{\pi_{2}(M)} is also independent of the chosen pair (𝔭,c)C(Σ)(\mathfrak{p},c)\in\mathfrak{Z}_{C}(\Sigma) with the property that the 𝔭\mathfrak{p}-fibres are in the class 𝔥\mathfrak{h}. This last statement can also be seen directly combining Remark 4.4, equation (4.4), and item (i) in Lemma 5.4.

6.2 The action functional on a covering space

In view of Lemma 6.3 we cannot expect the existence of a primitive functional of the action form 𝔞\mathfrak{a} in general. One standard way to resolve this problem is to find a primitive functional on a suitable covering space of Λ𝔥(Σ)\Lambda_{\mathfrak{h}}(\Sigma). We define a natural covering space in (6.5) below. However, it will have the small disadvantage that in some cases the primitive functional depends on the choice of a one-form α\alpha such that Ω=Ω0+dα\Omega=\Omega_{0}+{\mathrm{d}}\alpha, where Ω0\Omega_{0} is a reference two-form in the same cohomology class of Ω\Omega. We overcome this nuisance through the notion of non-degeneracy, which we introduced in Definition 5.3.

As before, let CC be a class in HdR2(Σ)H^{2}_{\mathrm{dR}}(\Sigma) and 𝔓0(Σ)\mathfrak{P}^{0}(\Sigma) be a connected component of 𝔓(Σ)\mathfrak{P}(\Sigma). We write 𝔥[S1,Σ]\mathfrak{h}\in[S^{1},\Sigma] for the class of the oriented fibres of any bundle in 𝔓0(Σ)\mathfrak{P}^{0}(\Sigma). For the rest of this section, we work under the assumption that

C0(Σ)=𝔓1(𝔓0(Σ))C(Σ)is non-empty and non-degenerate,\mathfrak{Z}_{C}^{0}(\Sigma)=\mathfrak{P}^{-1}(\mathfrak{P}_{0}(\Sigma))\cap\mathfrak{Z}_{C}(\Sigma)\ \ \text{is non-empty and non-degenerate}, (6.3)

where 𝔓:(Σ)𝔓(Σ)\mathfrak{P}:\mathfrak{Z}(\Sigma)\to\mathfrak{P}(\Sigma) is the standard projection. We fix a reference pair (𝔭0,c0)C0(Σ)(\mathfrak{p}_{0},c_{0})\in\mathfrak{Z}_{C}^{0}(\Sigma) with 𝔭0:ΣM0\mathfrak{p}_{0}:\Sigma\to M_{0} and denote by e0HdR2(M0)e_{0}\in H^{2}_{{\mathrm{dR}}}(M_{0}) minus the real Euler class of 𝔭0\mathfrak{p}_{0}. From (4.3), we have an embedding

ȷ𝔭0:ΣΛ𝔥(Σ),\jmath_{\mathfrak{p}_{0}}:\Sigma\to\Lambda_{\mathfrak{h}}(\Sigma), (6.4)

where ȷ𝔭0(z)\jmath_{\mathfrak{p}_{0}}(z) is a parametrisation of the oriented 𝔭0\mathfrak{p}_{0}-fibre passing through zΣz\in\Sigma. By definition, there holds ȷ𝔭0(z)𝒳(Ω0)\jmath_{\mathfrak{p}_{0}}(z)\in\mathcal{X}(\Omega_{0}). We consider the following covering space of Λ𝔥(Σ)\Lambda_{\mathfrak{h}}(\Sigma):

Λ~𝔥(Σ):={{γr}|γ0ȷ𝔭0(Σ),γrΛ𝔥(Σ),r[0,1]}/,\widetilde{\Lambda}_{\mathfrak{h}}(\Sigma):=\Big{\{}\{\gamma_{r}\}\ \Big{|}\ \gamma_{0}\in\jmath_{\mathfrak{p}_{0}}(\Sigma),\ \gamma_{r}\in\Lambda_{\mathfrak{h}}(\Sigma),\ \forall\,r\in[0,1]\Big{\}}\,\Big{/}\!\sim\,, (6.5)

where {γr0}{γr1}\{\gamma_{r}^{0}\}\sim\{\gamma_{r}^{1}\}, if there is a homotopy {γrs}s[0,1]\{\gamma_{r}^{s}\}_{s\in[0,1]} such that

γ0sȷ𝔭0(Σ),γ10=γ1s=γ11,s[0,1].\gamma_{0}^{s}\in\jmath_{\mathfrak{p}_{0}}(\Sigma),\quad\gamma_{1}^{0}=\gamma_{1}^{s}=\gamma_{1}^{1},\quad\forall\,s\in[0,1]. (6.6)

We denote the elements of Λ~𝔥(Σ)\widetilde{\Lambda}_{\mathfrak{h}}(\Sigma) as [γr][\gamma_{r}] so that the covering map is given by

Λ~𝔥(Σ)Λ𝔥(Σ),[γr]γ1.\widetilde{\Lambda}_{\mathfrak{h}}(\Sigma)\to{\Lambda}_{\mathfrak{h}}(\Sigma),\qquad[\gamma_{r}]\mapsto\gamma_{1}.

We further choose some ω0Ξc02(M0)\omega_{0}\in\Xi^{2}_{c_{0}}(M_{0}) and set as reference form

Ω0:=𝔭0ω0ΞC2(Σ).\Omega_{0}:=\mathfrak{p}_{0}^{*}\mathrm{\omega}_{0}\in\Xi^{2}_{C}(\Sigma).

Let Vol:Ω1(Σ)\mathrm{Vol}:\Omega^{1}(\Sigma)\to{\mathbb{R}} and 𝔙𝔬𝔩:ΞC2(Σ){\mathfrak{Vol}}:\Xi^{2}_{C}(\Sigma)\to{\mathbb{R}} be the volume functionals associated with Ω0\Omega_{0}. As observed in Lemma 5.2, 𝔙𝔬𝔩{\mathfrak{Vol}} depends only on (𝔭0,ω0)(\mathfrak{p}_{0},\mathrm{\omega}_{0}) but not on ω0\mathrm{\omega}_{0}.

Let αΩ1(Σ)\alpha\in\Omega^{1}(\Sigma) and recall the notation Ωα=Ω0+dαΞC2(Σ)\Omega_{\alpha}=\Omega_{0}+{\mathrm{d}}\alpha\in\Xi^{2}_{C}(\Sigma). We define the action

𝒜~α:Λ~𝔥(Σ),𝒜~α([γr]):=S1γ0α+[0,1]×S1ΓΩα,\widetilde{\mathcal{A}}_{\alpha}:\widetilde{\Lambda}_{\mathfrak{h}}(\Sigma)\to{\mathbb{R}}\,,\qquad\widetilde{\mathcal{A}}_{\alpha}\big{(}[\gamma_{r}]\big{)}:=\int_{S^{1}}\gamma_{0}^{*}\alpha+\int_{[0,1]\times S^{1}}\Gamma^{*}\Omega_{\alpha}, (6.7)

where, as before, Γ:[0,1]×S1Σ\Gamma:[0,1]\times S^{1}\to\Sigma is the cylinder associated with {γr}\{\gamma_{r}\}.

The action is well-defined as one sees by integrating 0=dΩα0={\mathrm{d}}\Omega_{\alpha} over a homotopy satisfying (6.6) and then applying Stokes’ Theorem. Decomposing Ωα=Ω0+dα\Omega_{\alpha}=\Omega_{0}+{\mathrm{d}}\alpha in the second integrand above and using Stokes’ Theorem, we can rewrite 𝒜~α\widetilde{\mathcal{A}}_{\alpha} as

𝒜~α([γr])=S1γ1α+D2Γ¯ω0,\widetilde{\mathcal{A}}_{\alpha}\big{(}[\gamma_{r}]\big{)}=\int_{S^{1}}\gamma_{1}^{*}\alpha+\int_{D^{2}}\bar{\Gamma}^{*}\omega_{0}, (6.8)

where Γ¯=𝔭0Γ:D2M0\bar{\Gamma}=\mathfrak{p}_{0}\circ\Gamma:D^{2}\to M_{0}. A straightforward computation shows that

d[γr]𝒜~α([ξr])=𝔞γ1(ξ1),[γr]Λ~𝔥(Σ),[ξr]T[γr]Λ~𝔥(Σ),{\mathrm{d}}_{[\gamma_{r}]}\widetilde{\mathcal{A}}_{\alpha}\big{(}[\xi_{r}]\big{)}=\mathfrak{a}_{\gamma_{1}}\big{(}\xi_{1}\big{)},\qquad\forall\,[\gamma_{r}]\in\widetilde{\Lambda}_{\mathfrak{h}}(\Sigma),\ \forall\,[\xi_{r}]\in{\mathrm{T}}_{[\gamma_{r}]}\widetilde{\Lambda}_{\mathfrak{h}}(\Sigma),

where 𝔞=𝔞(Ωα)\mathfrak{a}=\mathfrak{a}(\Omega_{\alpha}). Hence, Lemma 6.2 can be rephrased as follows.

Corollary 6.5.

Let [γr]Λ~𝔥(Σ)[\gamma_{r}]\in\widetilde{\Lambda}_{\mathfrak{h}}(\Sigma) with γ1Λ𝔥(Σ)\gamma_{1}\in\Lambda_{\mathfrak{h}}(\Sigma) embedded. Then [γr][\gamma_{r}] is a critical point of 𝒜~α\widetilde{\mathcal{A}}_{\alpha} if and only if γ1𝒳(Ωα)\gamma_{1}\in\mathcal{X}(\Omega_{\alpha}).∎

If α\alpha and α\alpha^{\prime} are one-forms such that Ωα=Ω=Ωα\Omega_{\alpha}=\Omega=\Omega_{\alpha^{\prime}}, then τ:=αα\tau:=\alpha^{\prime}-\alpha is closed and we have

𝒜~α=𝒜~α+[τ],[𝔭1(pt)].\widetilde{\mathcal{A}}_{\alpha^{\prime}}=\widetilde{\mathcal{A}}_{\alpha}+\langle[\tau],[\mathfrak{p}^{-1}({\mathrm{pt}})]\rangle.\vskip 8.0pt (6.9)

Case 1: e00e_{0}\neq 0.

By Lemma 4.5, we have [𝔭1(pt)]=0[\mathfrak{p}^{-1}({\mathrm{pt}})]=0. The action functional depends only on the two-form ΩΞC2(Σ)\Omega\in\Xi^{2}_{C}(\Sigma) due to (6.9). Therefore, for any αΩ1(Σ)\alpha\in\Omega^{1}(\Sigma) with Ω=Ωα\Omega=\Omega_{\alpha}, we can set

𝒜~Ω:=𝒜~α.\widetilde{\mathcal{A}}_{\Omega}:=\widetilde{\mathcal{A}}_{\alpha}.

In this situation, the non-degeneracy (6.3) of C0(Σ)\mathfrak{Z}_{C}^{0}(\Sigma) is not needed to associate the action functional with elements in ΞC2(Σ)\Xi^{2}_{C}(\Sigma), as opposed to the next case.

Case 2: e0=0e_{0}=0.

Here, the action functionals of α\alpha and α\alpha^{\prime} might be different. Nevertheless, as C0(Σ)\mathfrak{Z}_{C}^{0}(\Sigma) is non-degenerate, we have c0n,[M0]=ev(𝔭0,c0)0\langle c_{0}^{n},[M_{0}]\rangle={\mathrm{ev}}(\mathfrak{p}_{0},c_{0})\neq 0. Thus, by (2.2) and (4.6), we can write

[τ],[𝔭1(pt)]=[τ]Cn,[Σ]c0n,[M0]=Vol(α)Vol(α)c0n,[M0],\langle[\tau],[\mathfrak{p}^{-1}({\mathrm{pt}})]\rangle=\frac{\langle[\tau]\cup C^{n},[\Sigma]\rangle}{\langle c^{n}_{0},[M_{0}]\rangle}=\frac{\mathrm{Vol}(\alpha^{\prime})-\mathrm{Vol}(\alpha)}{\langle c_{0}^{n},[M_{0}]\rangle},

so that if α\alpha and α\alpha^{\prime} have the same volume, they also have the same action by (6.9). We set

𝒜~Ω:=𝒜~α\widetilde{\mathcal{A}}_{\Omega}:=\widetilde{\mathcal{A}}_{\alpha} (6.10)

for a normalised αΩ1(Σ)\alpha\in\Omega^{1}(\Sigma), i.e. Vol(α)=0\mathrm{Vol}(\alpha)=0, with Ω=Ωα\Omega=\Omega_{\alpha}.

Remark 6.6.

This remark is parallel to Remark 2.6. For contact forms and Hamiltonian systems, the action functional recovers the following well-known formulae.

  • (Contact forms) Let Ω0=0\Omega_{0}=0 and αΩ1(Σ)\alpha\in\Omega^{1}(\Sigma) be a (possibly contact) one-form. As C=0C=0, we have that e00e_{0}\neq 0 by Lemma 4.5.(ii). Thus, this is a special instance of Case 1. In fact, 𝔞(C)=0\mathfrak{a}(C)=0 due to (6.2), and the function 𝒜α:Λ𝔥(Σ)\mathcal{A}_{\alpha}:\Lambda_{\mathfrak{h}}(\Sigma)\to{\mathbb{R}}, given by

    𝒜α(γ)=S1γα,\mathcal{A}_{\alpha}(\gamma)=\int_{S^{1}}\gamma^{*}\alpha,

    is the unique primitive of 𝔞\mathfrak{a} such that 𝒜~dα([γr])=𝒜α(γ1)\widetilde{\mathcal{A}}_{{\mathrm{d}}\alpha}([\gamma_{r}])=\mathcal{A}_{\alpha}(\gamma_{1}), for every [γr]Λ~𝔥(Σ)[\gamma_{r}]\in\widetilde{\Lambda}_{\mathfrak{h}}(\Sigma).

  • (Hamiltonian systems) Let 𝔭0\mathfrak{p}_{0} be trivial, namely Σ=M0×S1\Sigma=M_{0}\times S^{1}, so that this is a special instance of Case 2. Assume Ω0=𝔭0ω0\Omega_{0}=\mathfrak{p}_{0}^{*}\mathrm{\omega}_{0}, where ω0\mathrm{\omega}_{0} is a symplectic form on M0M_{0}, and α=Hdt\alpha=H{\mathrm{d}}t, where tt is the angular coordinate on S1S^{1}. If γ1(t)=(q1(t),t)\gamma_{1}(t)=(q_{1}(t),t), then we get the classical Hamiltonian action functional

    𝒜~Hdt([γr]):=S1H(q1(t),t)dt+D2Γ¯ω0\widetilde{\mathcal{A}}_{H{\mathrm{d}}t}([\gamma_{r}]):=\int_{S^{1}}H(q_{1}(t),t)\,{\mathrm{d}}t+\int_{D^{2}}\bar{\Gamma}^{*}\mathrm{\omega}_{0}

    on the space of contractible curves with capping disc. Furthermore, the condition kere0|π2(M0)kerc0|π2(M0)\ker e_{0}|_{\pi_{2}(M_{0})}\subseteq\ker c_{0}|_{\pi_{2}(M_{0})} in Lemma 6.3 means that ω0\mathrm{\omega}_{0} is symplectically aspherical.

Next, we study the relation between the actions with respect to two different reference weakly Zoll pairs.

Proposition 6.7.

Let (𝔭0,c0)(\mathfrak{p}_{0}^{\prime},c_{0}^{\prime}) be another element in C0(Σ)\mathfrak{Z}_{C}^{0}(\Sigma). We write Λ~𝔥(Σ)\widetilde{\Lambda}_{\mathfrak{h}}^{\prime}(\Sigma) for the associated covering space of Λ𝔥(Σ)\Lambda_{\mathfrak{h}}(\Sigma) and Λ~𝔥(𝔭0,𝔭0)Λ~𝔥(Σ)\widetilde{\Lambda}_{\mathfrak{h}}(\mathfrak{p}_{0},\mathfrak{p}_{0}^{\prime})\subset\widetilde{\Lambda}_{\mathfrak{h}}(\Sigma) for the set of elements [δr][\delta_{r}] such that δ1ȷ𝔭0(Σ)\delta_{1}\in\jmath_{\mathfrak{p}_{0}^{\prime}}(\Sigma). We pick ω0Ξc02(M0)\mathrm{\omega}_{0}^{\prime}\in\Xi^{2}_{c_{0}^{\prime}}(M_{0}^{\prime}) and set Ω0:=(𝔭0)ω0\Omega_{0}^{\prime}:=(\mathfrak{p}_{0}^{\prime})^{*}\mathrm{\omega}_{0}^{\prime}. We choose αΩ1(Σ)\alpha^{\prime}\in\Omega^{1}(\Sigma) such that

Ω0=Ω0+dα,αis Ω0-normalised.\Omega_{0}^{\prime}=\Omega_{0}+{\mathrm{d}}\alpha^{\prime},\qquad\alpha^{\prime}\ \text{is $\Omega_{0}$-normalised}.

For any ΩΞC2(Σ)\Omega\in\Xi^{2}_{C}(\Sigma), we take αΩ1(Σ)\alpha\in\Omega^{1}(\Sigma) such that

Ω=Ω0+dα,αis Ω0-normalised.\Omega=\Omega_{0}+{\mathrm{d}}\alpha,\qquad\ \ \alpha\ \text{is $\Omega_{0}$-normalised}.

Then, there holds Ω=Ω0+d(αα)\Omega=\Omega_{0}^{\prime}+{\mathrm{d}}(\alpha-\alpha^{\prime}) and the one-form αα\alpha-\alpha^{\prime} is Ω0\Omega_{0}^{\prime}-normalised. Moreover, if we denote by 𝒜~Ω\widetilde{\mathcal{A}}^{\prime}_{\Omega} the action of Ω\Omega with respect to Ω0\Omega_{0}^{\prime}, then

𝒜~Ω([{δr}#{γr}])=𝒜~Ω0([δr])+𝒜~Ω([γr]),\widetilde{\mathcal{A}}_{\Omega}\big{(}[\{\delta_{r}\}\#\{\gamma_{r}\}]\big{)}=\widetilde{\mathcal{A}}_{\Omega_{0}^{\prime}}\big{(}[\delta_{r}]\big{)}+\widetilde{\mathcal{A}}_{\Omega}^{\prime}\big{(}[\gamma_{r}]\big{)}, (6.11)

for every [δr]Λ~𝔥(𝔭0,𝔭0)[\delta_{r}]\in\widetilde{\Lambda}_{\mathfrak{h}}(\mathfrak{p}_{0},\mathfrak{p}_{0}^{\prime}) and [γr]Λ~𝔥(Σ)[\gamma_{r}]\in\widetilde{\Lambda}_{\mathfrak{h}}^{\prime}(\Sigma). Here, the concatenation is made by choosing any representative {γr}\{\gamma_{r}\} of [γr][\gamma_{r}] with γ0=δ1\gamma_{0}=\delta_{1}.

Proof.

The one-form αα\alpha-\alpha^{\prime} is Ω0\Omega^{\prime}_{0}-normalised thanks to Lemma 2.4. Let Γ,Δ:[0,1]×S1Σ\Gamma,\Delta:[0,1]\times S^{1}\to\Sigma be the cylinders traced by the paths {γr}\{\gamma_{r}\} and {δr}\{\delta_{r}\}, respectively. Let us show equation (6.11) using (6.7) and (6.8):

𝒜~Ω0([δr])+𝒜~Ω([γr])\displaystyle\widetilde{\mathcal{A}}_{\Omega_{0}^{\prime}}\big{(}[\delta_{r}]\big{)}+\widetilde{\mathcal{A}}_{\Omega}^{\prime}\big{(}[\gamma_{r}]\big{)} =S1δ1α+[0,1]×S1ΔΩ0+S1γ0(αα)+[0,1]×S1ΓΩ\displaystyle=\int_{S^{1}}\delta_{1}^{*}\alpha^{\prime}+\int_{[0,1]\times S^{1}}\Delta^{*}\Omega_{0}+\int_{S^{1}}\gamma_{0}^{*}(\alpha-\alpha^{\prime})+\int_{[0,1]\times S^{1}}\Gamma^{*}\Omega
=[0,1]×S1ΔΩ0+S1(γ1)α+[0,1]×S1ΓΩ0+[0,1]×S1Γ(dα)\displaystyle=\int_{[0,1]\times S^{1}}\Delta^{*}\Omega_{0}+\int_{S^{1}}(\gamma_{1})^{*}\alpha+\int_{[0,1]\times S^{1}}\Gamma^{*}\Omega_{0}+\int_{[0,1]\times S^{1}}\Gamma^{*}({\mathrm{d}}\alpha)
=[0,1]×S1(Δ#Γ)Ω0+S1γ0α+S1γ1αS1γ0α\displaystyle=\int_{[0,1]\times S^{1}}(\Delta\#\Gamma)^{*}\Omega_{0}+\int_{S^{1}}\gamma_{0}^{*}\alpha+\int_{S^{1}}\gamma_{1}^{*}\alpha-\int_{S^{1}}\gamma_{0}^{*}\alpha
=𝒜~Ω([{δr}#{γr}]).\displaystyle=\widetilde{\mathcal{A}}_{\Omega}\big{(}[\{\delta_{r}\}\#\{\gamma_{r}\}]\big{)}.\qed
Corollary 6.8.

The action 𝒜~Ω\widetilde{\mathcal{A}}_{\Omega} depends only on the fixed reference pair (𝔭0,c0)C0(Σ)(\mathfrak{p}_{0},c_{0})\in\mathfrak{Z}_{C}^{0}(\Sigma), and not on the specific choice of ω0Ξc02(M0)\mathrm{\omega}_{0}\in\Xi^{2}_{c_{0}}(M_{0}).

Proof.

Let ω0\mathrm{\omega}_{0}^{\prime} be another element in Ξc02(M0)\Xi^{2}_{c_{0}}(M_{0}). By (6.11) with (𝔭0,c0)=(𝔭0,c0)(\mathfrak{p}_{0}^{\prime},c_{0}^{\prime})=(\mathfrak{p}_{0},c_{0}) and [δr]=[γ0][\delta_{r}]=[\gamma_{0}], it suffices to show that 𝒜~𝔭0ω0([γ0])=0\widetilde{\mathcal{A}}_{\mathfrak{p}_{0}^{*}\mathrm{\omega}_{0}^{\prime}}([\gamma_{0}])=0. To this purpose, let ζΩ1(M0)\zeta\in\Omega^{1}(M_{0}) be such that ω0ω0=dζ\mathrm{\omega}_{0}^{\prime}-\mathrm{\omega}_{0}={\mathrm{d}}\zeta. By Lemma 5.2.(ii), 𝔭0ζ\mathfrak{p}_{0}^{*}\zeta is 𝔭0ω0\mathfrak{p}_{0}^{*}\mathrm{\omega}_{0}-normalised and we can use it to compute 𝒜~𝔭0ω0\widetilde{\mathcal{A}}_{\mathfrak{p}_{0}^{*}\mathrm{\omega}_{0}^{\prime}}:

𝒜~𝔭0ω0([γ0])=S1γ0(𝔭0ζ)=0.\widetilde{\mathcal{A}}_{\mathfrak{p}_{0}^{*}\mathrm{\omega}_{0}^{\prime}}([\gamma_{0}])=\int_{S^{1}}\gamma_{0}^{*}(\mathfrak{p}^{*}_{0}\zeta)=0.\qed

6.3 The action is invariant under pull-back and isotopies

In this subsection, we prove two invariance results for the action. For the first one, we consider an additional connected oriented closed manifold Σ\Sigma^{\text{\tiny$\vee$}} of dimension 2n+12n+1. We suppose that there are a bundle 𝔭0:ΣM0\mathfrak{p}_{0}^{\text{\tiny$\vee$}}:\Sigma^{\text{\tiny$\vee$}}\to M_{0}^{\text{\tiny$\vee$}} in 𝔓0(Σ)\mathfrak{P}^{0}(\Sigma^{\text{\tiny$\vee$}}) and a bundle map Π:ΣΣ\Pi:\Sigma^{\text{\tiny$\vee$}}\to\Sigma with 𝔭0=Π𝔭0\mathfrak{p}_{0}^{\text{\tiny$\vee$}}=\Pi^{*}\mathfrak{p}_{0}. Let 𝔓0(Σ)\mathfrak{P}^{0}(\Sigma^{\text{\tiny$\vee$}}) be the connected component of 𝔓(Σ)\mathfrak{P}(\Sigma^{\text{\tiny$\vee$}}) containing 𝔭0\mathfrak{p}_{0}^{\text{\tiny$\vee$}}. Let π0:M0M0\pi_{0}:M_{0}^{\text{\tiny$\vee$}}\to M_{0} be the map fitting into the first commutative diagram in (4.1). If we set

c0:=π0c0HdR2(M0),C:=ΠCHdR2(Σ),c_{0}^{\text{\tiny$\vee$}}:=\pi_{0}^{*}c_{0}\in H^{2}_{\mathrm{dR}}(M_{0}^{\text{\tiny$\vee$}}),\qquad C^{\text{\tiny$\vee$}}:=\Pi^{*}C\in H^{2}_{\mathrm{dR}}(\Sigma^{\text{\tiny$\vee$}}),

then (𝔭0,c0)C0(Σ):=(𝔓)1(𝔓0(Σ))C(Σ)(\mathfrak{p}_{0}^{\text{\tiny$\vee$}},c_{0}^{\text{\tiny$\vee$}})\in\mathfrak{Z}^{0}_{C^{\text{\tiny$\vee$}}}(\Sigma^{\text{\tiny$\vee$}}):=(\mathfrak{P}^{\text{\tiny$\vee$}})^{-1}(\mathfrak{P}^{0}(\Sigma^{\text{\tiny$\vee$}}))\cap\mathfrak{Z}_{C^{\text{\tiny$\vee$}}}(\Sigma^{\text{\tiny$\vee$}}). In particular, C0(Σ)\mathfrak{Z}_{C^{\text{\tiny$\vee$}}}^{0}(\Sigma^{\text{\tiny$\vee$}}) is non-empty. We write 𝔓C0:C0(Σ)𝔓0(Σ)\mathfrak{P}_{C^{\text{\tiny$\vee$}}}^{0}:\mathfrak{Z}_{C^{\text{\tiny$\vee$}}}^{0}(\Sigma^{\text{\tiny$\vee$}})\to\mathfrak{P}^{0}(\Sigma^{\text{\tiny$\vee$}}) for the projection and ev:C0(Σ){\mathrm{ev}}^{\text{\tiny$\vee$}}:\mathfrak{Z}_{C^{\text{\tiny$\vee$}}}^{0}(\Sigma^{\text{\tiny$\vee$}})\to{\mathbb{R}} for the evaluation. Furthermore, we abbreviate

ω0:=π0ω0Ξc02(M0),Ω0:=(𝔭0)ω0=ΠΩ0ΞC2(Σ),\mathrm{\omega}_{0}^{\text{\tiny$\vee$}}:=\pi_{0}^{*}\mathrm{\omega}_{0}\in\Xi^{2}_{c_{0}^{\text{\tiny$\vee$}}}(M_{0}^{\text{\tiny$\vee$}}),\qquad\Omega_{0}^{\text{\tiny$\vee$}}:=(\mathfrak{p}_{0}^{\text{\tiny$\vee$}})^{*}\mathrm{\omega}_{0}^{\text{\tiny$\vee$}}=\Pi^{*}\Omega_{0}\in\Xi^{2}_{C^{\text{\tiny$\vee$}}}(\Sigma^{\text{\tiny$\vee$}}),

so that Ω0\Omega_{0}^{\text{\tiny$\vee$}} is associated with the weakly Zoll pair (𝔭0,c0)(\mathfrak{p}_{0}^{\text{\tiny$\vee$}},c_{0}^{\text{\tiny$\vee$}}). We note that

π0e0=e0,π0(Ae0+c0)=Ae0+c0,A,\bullet\quad\pi_{0}^{*}e_{0}=e_{0}^{\text{\tiny$\vee$}},\qquad\qquad\bullet\quad\pi_{0}^{*}(Ae_{0}+c_{0})=Ae^{\text{\tiny$\vee$}}_{0}+c_{0}^{\text{\tiny$\vee$}},\quad\forall\,A\in{\mathbb{R}}, (6.12)

where e0e_{0}^{\text{\tiny$\vee$}} is minus the real Euler class of 𝔭0\mathfrak{p}_{0}^{\text{\tiny$\vee$}}. If 𝔥\mathfrak{h}^{\text{\tiny$\vee$}} denotes the free-homotopy class of oriented fibres of elements in 𝔓0(Σ)\mathfrak{P}^{0}(\Sigma^{\text{\tiny$\vee$}}), we write 𝒜~α:Λ~𝔥(Σ)\widetilde{\mathcal{A}}^{\text{\tiny$\vee$}}_{\alpha^{\text{\tiny$\vee$}}}:\widetilde{\Lambda}_{\mathfrak{h}^{\text{\tiny$\vee$}}}(\Sigma^{\text{\tiny$\vee$}})\to{\mathbb{R}} for the action of αΩ1(Σ)\alpha^{\text{\tiny$\vee$}}\in\Omega^{1}(\Sigma^{\text{\tiny$\vee$}}) with respect to Ω0\Omega_{0}^{\text{\tiny$\vee$}}. The bundle map Π\Pi yields a map between the spaces

Π:Λ𝔥(Σ)Λ𝔥(Σ),Πγ:=Πγ\Pi_{*}:\Lambda_{\mathfrak{h}^{\text{\tiny$\vee$}}}(\Sigma^{\text{\tiny$\vee$}})\to\Lambda_{\mathfrak{h}}(\Sigma),\qquad\Pi_{*}\gamma^{\text{\tiny$\vee$}}:=\Pi\circ\gamma^{\text{\tiny$\vee$}}

and between their covering spaces

Π~:Λ~𝔥(Σ)Λ~𝔥(Σ),Π~[γr]:=[Πγr].\widetilde{\Pi}_{*}:\widetilde{\Lambda}_{\mathfrak{h}^{\text{\tiny$\vee$}}}(\Sigma^{\text{\tiny$\vee$}})\to\widetilde{\Lambda}_{\mathfrak{h}}(\Sigma),\qquad\widetilde{\Pi}_{*}[\gamma_{r}^{\text{\tiny$\vee$}}]:=[\Pi\circ\gamma_{r}^{\text{\tiny$\vee$}}].
Proposition 6.9.

The set C0(Σ)\mathfrak{Z}_{C^{\text{\tiny$\vee$}}}^{0}(\Sigma^{\text{\tiny$\vee$}}) is non-degenerate if and only if degπ00\deg\pi_{0}\neq 0. In this case, 𝒜~Ω:Λ~𝔥(Σ)\widetilde{\mathcal{A}}^{\text{\tiny$\vee$}}_{\Omega^{\text{\tiny$\vee$}}}:\widetilde{\Lambda}_{\mathfrak{h}^{\text{\tiny$\vee$}}}(\Sigma^{\text{\tiny$\vee$}})\to{\mathbb{R}} is well-defined for all ΩΞC2(Σ)\Omega^{\text{\tiny$\vee$}}\in\Xi^{2}_{C^{\text{\tiny$\vee$}}}(\Sigma^{\text{\tiny$\vee$}}) and there holds

𝒜~ΠΩ=𝒜~ΩΠ~,ΩΞC2(Σ).\widetilde{\mathcal{A}}^{\text{\tiny$\vee$}}_{\Pi^{*}\Omega}=\widetilde{\mathcal{A}}_{\Omega}\circ\widetilde{\Pi}_{*},\qquad\forall\,\Omega\in\Xi^{2}_{C}(\Sigma).
Proof.

The map (𝔓C0)1(𝔭0)(𝔓C0)1(𝔭0)(\mathfrak{P}_{C}^{0})^{-1}(\mathfrak{p}_{0})\to(\mathfrak{P}_{C^{\text{\tiny$\vee$}}}^{0})^{-1}(\mathfrak{p}_{0}^{\text{\tiny$\vee$}}) given by (𝔭0,c)(𝔭0,π0c)(\mathfrak{p}_{0},c)\mapsto(\mathfrak{p}_{0}^{\text{\tiny$\vee$}},\pi_{0}^{*}c) is well-defined. By (4.4) and (6.12), it is also surjective. Since C0(Σ)\mathfrak{Z}_{C}^{0}(\Sigma) is non-degenerate and there holds

ev(𝔭0,π0c)=degπ0ev(𝔭0,c),(𝔭0,c)(𝔓C0)1(𝔭0),{\mathrm{ev}}^{\text{\tiny$\vee$}}(\mathfrak{p}_{0}^{\text{\tiny$\vee$}},\pi_{0}^{*}c)=\deg\pi_{0}\cdot{\mathrm{ev}}(\mathfrak{p}_{0},c),\qquad\forall\,(\mathfrak{p}_{0},c)\in(\mathfrak{P}_{C}^{0})^{-1}(\mathfrak{p}_{0}), (6.13)

we see that C0(Σ)\mathfrak{Z}_{C^{\text{\tiny$\vee$}}}^{0}(\Sigma^{\text{\tiny$\vee$}}) is non-degenerate exactly when degπ0\deg\pi_{0} is non-zero.

Let Ω=Ω0+dαΞC2(Σ)\Omega=\Omega_{0}+{\mathrm{d}}\alpha\in\Xi^{2}_{C}(\Sigma), where αΩ1(Σ)\alpha\in\Omega^{1}(\Sigma) is Ω0\Omega_{0}-normalised. Then, ΠΩ=Ω0+d(Πα)\Pi^{*}\Omega=\Omega_{0}^{\text{\tiny$\vee$}}+{\mathrm{d}}(\Pi^{*}\alpha) and Πα\Pi^{*}\alpha is Ω0\Omega_{0}^{\text{\tiny$\vee$}}-normalised by Proposition 2.7. For all [γr]Λ~𝔥(Σ)[\gamma_{r}^{\text{\tiny$\vee$}}]\in\widetilde{\Lambda}_{\mathfrak{h}^{\text{\tiny$\vee$}}}(\Sigma), we have

𝒜~ΠΩ([γr])=S1(γ0)Πα+[0,1]×S1(Γ)ΠΩ\displaystyle\widetilde{\mathcal{A}}^{\text{\tiny$\vee$}}_{\Pi^{*}\Omega}\big{(}[\gamma^{\text{\tiny$\vee$}}_{r}]\big{)}=\int_{S^{1}}(\gamma_{0}^{\text{\tiny$\vee$}})^{*}\Pi^{*}\alpha+\int_{[0,1]\times S^{1}}(\Gamma^{\text{\tiny$\vee$}})^{*}\Pi^{*}\Omega =S1(Πγ0)α+[0,1]×S1(ΠΓ)Ω\displaystyle=\int_{S^{1}}(\Pi\circ\gamma_{0}^{\text{\tiny$\vee$}})^{*}\alpha+\int_{[0,1]\times S^{1}}(\Pi\circ\Gamma^{\text{\tiny$\vee$}})^{*}\Omega
=𝒜~Ω(Π~[γr]).\displaystyle=\widetilde{\mathcal{A}}_{\Omega}\big{(}\widetilde{\Pi}_{*}[\gamma^{\text{\tiny$\vee$}}_{r}]\big{)}.\qed

For the second invariance result, we consider a diffeomorphism Ψ:ΣΣ\Psi:\Sigma\to\Sigma isotopic to idΣ\mathrm{id}_{\Sigma}. We define

Ψ:Λ𝔥(Σ)Λ𝔥(Σ),Ψγ:=(Ψ1)γ=Ψ1γ.\Psi^{*}:\Lambda_{\mathfrak{h}}(\Sigma)\to\Lambda_{\mathfrak{h}}(\Sigma),\qquad\Psi^{*}\gamma:=(\Psi^{-1})_{*}\gamma=\Psi^{-1}\circ\gamma.

Let {Ψr}\{\Psi_{r}\} be an isotopy of diffeomorphisms of Σ\Sigma with Ψ0=idΣ\Psi_{0}=\mathrm{id}_{\Sigma} and Ψ1=Ψ\Psi_{1}=\Psi. We denote by [Ψr][\Psi_{r}] the homotopy class with fixed end-points of {Ψr}\{\Psi_{r}\} in the space of isotopies. Given such a homotopy class, we define

[Ψr]:Λ~𝔥(Σ)Λ~𝔥(Σ),[Ψr][γr]:=[Ψr1γr].[\Psi_{r}]^{*}:\widetilde{\Lambda}_{\mathfrak{h}}(\Sigma)\to\widetilde{\Lambda}_{\mathfrak{h}}(\Sigma),\qquad[\Psi_{r}]^{*}[\gamma_{r}]:=\big{[}\Psi_{r}^{-1}\circ\gamma_{r}\big{]}.
Proposition 6.10.

For every ΩΞC2(Σ)\Omega\in\Xi^{2}_{C}(\Sigma) and every homotopy class [Ψr][\Psi_{r}] as above, there holds

𝒜~Ψ1Ω[Ψr]=𝒜~Ω.\widetilde{\mathcal{A}}_{\Psi^{*}_{1}\Omega}\circ[\Psi_{r}]^{*}=\widetilde{\mathcal{A}}_{\Omega}.
Proof.

We observe preliminarily that if XrX_{r} and YrY_{r} are the time-dependent vector fields generating Ψr\Psi_{r} and Ψr1\Psi_{r}^{-1}, we have the relation

Xr=dΨr(Yr)Ψr1.-X_{r}={\mathrm{d}}\Psi_{r}(Y_{r})\circ\Psi_{r}^{-1}.

Let α\alpha be a normalised one-form such that Ω=Ω0+dα\Omega=\Omega_{0}+{\mathrm{d}}\alpha and let {θr}\{\theta_{r}\} be the path of normalised one-forms introduced in Section 2.2 such that

ΨrΩ=Ω0+d(θr+Ψrα),θ˙r=Ψr(ιXrΩ0),θ0=0.\Psi_{r}^{*}\Omega=\Omega_{0}+{\mathrm{d}}(\theta_{r}+\Psi_{r}^{*}\alpha),\qquad\dot{\theta}_{r}=\Psi_{r}^{*}(\iota_{X_{r}}\Omega_{0}),\quad\theta_{0}=0.

For s[0,1]s\in[0,1], we define the truncation {Ψrs:=Ψrs}\{\Psi_{r}^{s}:=\Psi_{rs}\} and we write 𝔞s:=𝔞(ΨsΩ)\mathfrak{a}^{s}:=\mathfrak{a}(\Psi_{s}^{*}\Omega). For every [γr]Λ~𝔥(Σ)[\gamma_{r}]\in\widetilde{\Lambda}_{\mathfrak{h}}(\Sigma), we compute

dds(𝒜~ΨsΩ([Ψrs][γr]))=(dds𝒜~ΨsΩ)([Ψrs][γr])+d[Ψrsγr]𝒜~ΨsΩdds[Ψrsγr]=S1(Ψsγ1)(dds(θs+Ψsα))dt+𝔞Ψsγ1sddsΨsγ1=S1γ1(ιXsΩ+d(α(Xs)))dt+S1(ΨsΩ)(sΨsγ1,tΨsγ1)dt=S1γ1(ιXsΩ)dt+S1Ω(dΨss(Ψs1γ1),dΨst(Ψs1γ1))dt=S1Ω(Xsγ1,γ˙1)dt+S1Ω(Xsγ1,γ˙1)dt=0,\begin{split}\frac{{\mathrm{d}}}{{\mathrm{d}}s}\Big{(}\widetilde{\mathcal{A}}_{\Psi_{s}^{*}\Omega}\big{(}[\Psi_{r}^{s}]^{*}[\gamma_{r}]\big{)}\Big{)}&=\Big{(}\frac{{\mathrm{d}}}{{\mathrm{d}}s}\widetilde{\mathcal{A}}_{\Psi_{s}^{*}\Omega}\Big{)}\big{(}[\Psi_{r}^{s}]^{*}[\gamma_{r}]\big{)}+{\mathrm{d}}_{[\Psi_{rs}^{*}\gamma_{r}]}\widetilde{\mathcal{A}}_{\Psi_{s}^{*}\Omega}\cdot\frac{{\mathrm{d}}}{{\mathrm{d}}s}[\Psi_{rs}^{*}\gamma_{r}]\\ &=\int_{S^{1}}(\Psi_{s}^{*}\gamma_{1})^{*}\Big{(}\tfrac{{\mathrm{d}}}{{\mathrm{d}}s}(\theta_{s}+\Psi_{s}^{*}\alpha)\Big{)}\,{\mathrm{d}}t+\mathfrak{a}^{s}_{\Psi_{s}^{*}\gamma_{1}}\cdot\frac{{\mathrm{d}}}{{\mathrm{d}}s}\Psi_{s}^{*}\gamma_{1}\\ &=\int_{S^{1}}\gamma_{1}^{*}\Big{(}\iota_{X_{s}}\Omega+{\mathrm{d}}\big{(}\alpha(X_{s})\big{)}\Big{)}\,{\mathrm{d}}t+\int_{S^{1}}(\Psi_{s}^{*}\Omega)\Big{(}\tfrac{\partial}{\partial s}\Psi_{s}^{*}\gamma_{1},\tfrac{\partial}{\partial t}\Psi_{s}^{*}\gamma_{1}\Big{)}{\mathrm{d}}t\\ &=\int_{S^{1}}\gamma_{1}^{*}(\iota_{X_{s}}\Omega)\,{\mathrm{d}}t+\int_{S^{1}}\Omega\Big{(}{\mathrm{d}}\Psi_{s}\cdot\tfrac{\partial}{\partial s}\big{(}\Psi_{s}^{-1}\circ\gamma_{1}\big{)},{\mathrm{d}}\Psi_{s}\cdot\tfrac{\partial}{\partial t}\big{(}\Psi_{s}^{-1}\circ\gamma_{1}\big{)}\Big{)}{\mathrm{d}}t\\ &=\int_{S^{1}}\Omega(X_{s}\circ\gamma_{1},\dot{\gamma}_{1})\,{\mathrm{d}}t+\int_{S^{1}}\Omega(-X_{s}\circ\gamma_{1},\dot{\gamma}_{1})\,{\mathrm{d}}t\\ &=0,\end{split}

where in the second equality we used (6.8) to compute dds𝒜~ΨsΩ\tfrac{{\mathrm{d}}}{{\mathrm{d}}s}\widetilde{\mathcal{A}}_{\Psi_{s}^{*}\Omega}. Since 𝒜~Ψ0Ω[Ψ0]=𝒜~Ω\widetilde{\mathcal{A}}_{\Psi_{0}^{*}\Omega}\circ[\Psi_{0}]^{*}=\widetilde{\mathcal{A}}_{\Omega}, the proof is completed. ∎

6.4 The action of weakly Zoll pairs

Based on the action functional we have studied, we define the action functional on the space of weakly Zoll pairs:

𝒜:C0(Σ),𝒜(𝔭,c):=𝒜~𝔭ω([ȷ𝔭r(z0)])\mathcal{A}:\mathfrak{Z}_{C}^{0}(\Sigma)\to{\mathbb{R}},\qquad\mathcal{A}(\mathfrak{p},c):=\widetilde{\mathcal{A}}_{\mathfrak{p}^{*}\mathrm{\omega}}\big{(}[\jmath_{\mathfrak{p}_{r}}(z_{0})]\big{)} (6.14)

where z0z_{0} is a point in Σ\Sigma, ωΞc2(M)\mathrm{\omega}\in\Xi^{2}_{c}(M), and {𝔭r}\{\mathfrak{p}_{r}\} is a path in 𝔓0(Σ)\mathfrak{P}^{0}(\Sigma) starting at the reference bundle 𝔭0\mathfrak{p}_{0} and ending at 𝔭1=𝔭\mathfrak{p}_{1}=\mathfrak{p}. It will turn out that this action is well-defined without the need to pass to a covering space, even when the condition in Lemma 6.3 is not met. This fact is a striking consequence of the non-degeneracy (6.3) of C0(Σ)\mathfrak{Z}_{C}^{0}(\Sigma). A key role will be played by the following polynomial.

Definition 6.11.

Let Q:Q:{\mathbb{R}}\to{\mathbb{R}} be the auxiliary polynomial

Q(A):=ev(𝔭0,Ae0+c0)=(Ae0+c0)n,[M0],A.Q(A):={\mathrm{ev}}(\mathfrak{p}_{0},Ae_{0}+c_{0})=\langle(Ae_{0}+c_{0})^{n},[M_{0}]\rangle,\qquad\forall\,A\in{\mathbb{R}}.

The Zoll polynomial P:P:{\mathbb{R}}\to{\mathbb{R}} of the pair (𝔭0,c0)(\mathfrak{p}_{0},c_{0}) is given through

P(0)=0,dPdA=Q.P(0)=0,\qquad\frac{{\mathrm{d}}P}{{\mathrm{d}}A}=Q.
Remark 6.12.

The Zoll polynomial is non-constant by Corollary 5.5.(ii) since we have assumed that C0(Σ)\mathfrak{Z}_{C}^{0}(\Sigma) is non-degenerate. Furthermore, we have the explicit formula

P(A)=0AQ(A)dA=j=0n1j+1(nj)e0jc0nj,[M0]Aj+1,P(A)=\int_{0}^{A}Q(A^{\prime})\,{\mathrm{d}}A^{\prime}=\sum_{j=0}^{n}\frac{1}{j+1}\binom{n}{j}\langle e^{j}_{0}\cup c^{n-j}_{0},[M_{0}]\rangle A^{j+1},

which reduces to

P(A)=e0,[M0]A22+c0,[M0]A,forn=1.P(A)=\langle e_{0},[M_{0}]\rangle\frac{A^{2}}{2}+\langle c_{0},[M_{0}]\rangle A,\quad\text{for}\ \ n=1. (6.15)
Remark 6.13.

This remark is parallel to Remark 2.6 and 6.6. The Zoll polynomial has a simple form when any of c0c_{0} and e0e_{0} vanishes.

  • Let us assume that C=0C=0 and take c0=0c_{0}=0, which is relevant to the study of Zoll contact forms. We have

    P(A)=e0n,[M0]An+1n+1.P(A)=\langle e_{0}^{n},[M_{0}]\rangle\frac{A^{n+1}}{n+1}.
  • Let us assume that e0=0e_{0}=0, which is relevant to the study of Hamiltonian systems on MM (in which case 𝔭0\mathfrak{p}_{0} is trivial and ω0\mathrm{\omega}_{0} is a symplectic form). We have

    P(A)=c0n,[M0]A.P(A)=\langle c_{0}^{n},[M_{0}]\rangle A.
Theorem 6.14.

The functional 𝒜:C0(Σ)\mathcal{A}:\mathfrak{Z}_{C}^{0}(\Sigma)\to{\mathbb{R}} does not depend on any choice involved. Moreover, for any (𝔭,c)C0(Σ)(\mathfrak{p},c)\in\mathfrak{Z}_{C}^{0}(\Sigma), there holds

𝒜(𝔭,c)=0,ife0=0 or (𝔭,c)=(𝔭0,c0),{\mathcal{A}}(\mathfrak{p},c)=0,\quad\text{if}\ \ e_{0}=0\ \text{ or }\ (\mathfrak{p},c)=(\mathfrak{p}_{0},c_{0}), (6.16)

and

𝔙𝔬𝔩(𝔭,c)=P(𝒜(𝔭,c)).{\mathfrak{Vol}}(\mathfrak{p},c)=P\big{(}\mathcal{A}(\mathfrak{p},c)\big{)}. (6.17)

If Ψ:ΣΣ\Psi:\Sigma\to\Sigma is isotopic to idΣ\mathrm{id}_{\Sigma} satisfying Ψ𝔭=𝔭0\Psi^{*}\mathfrak{p}=\mathfrak{p}_{0} (which exists by Lemma 4.7) and ψ:M0M\psi:M_{0}\to M is its quotient map, then

ψc=𝒜(𝔭,c)e0+c0.\psi^{*}c={\mathcal{A}}(\mathfrak{p},c)e_{0}+c_{0}. (6.18)
Proof.

To ease the notation, we write (𝔭1,c1)(\mathfrak{p}_{1},c_{1}) instead of (𝔭,c)(\mathfrak{p},c) to denote an arbitrary element of C0(Σ)\mathfrak{Z}_{C}^{0}(\Sigma) with 𝔭1:ΣM1\mathfrak{p}_{1}:\Sigma\to M_{1} and c1HdR2(M1)c_{1}\in H^{2}_{\mathrm{dR}}(M_{1}). Moreover, {𝔭r}\{\mathfrak{p}_{r}\} will indicate any path connecting the reference bundle 𝔭0\mathfrak{p}_{0} with 𝔭1\mathfrak{p}_{1}. We divide the proof in five steps.

Step 1. For any path {𝔭r}\{\mathfrak{p}_{r}\}, the action value 𝒜~𝔭1ω1([ȷ𝔭r(z0)])\widetilde{\mathcal{A}}_{\mathfrak{p}_{1}^{*}\mathrm{\omega}_{1}}\big{(}[\jmath_{\mathfrak{p}_{r}}(z_{0})]\big{)} does not depend on the choice of z0Σz_{0}\in\Sigma and ω1Ξc12(M1)\mathrm{\omega}_{1}\in\Xi^{2}_{c_{1}}(M_{1}).

Let z1z_{1} be another point in Σ\Sigma, and let {zs}\{z_{s}\} be a path between z0z_{0} and z1z_{1} with s[0,1]s\in[0,1]. We claim that 𝒜~𝔭1ω1([ȷ𝔭r(zs)])\widetilde{\mathcal{A}}_{\mathfrak{p}_{1}^{*}\mathrm{\omega}_{1}}\big{(}[\jmath_{\mathfrak{p}_{r}}(z_{s})]\big{)} does not depend on ss. Indeed, since ȷ𝔭1(zs)𝒳(𝔭1ω1)\jmath_{\mathfrak{p}_{1}}(z_{s})\in\mathcal{X}({\mathfrak{p}_{1}^{*}\mathrm{\omega}_{1}}) for all s[0,1]s\in[0,1], we see that d[ȷ𝔭r(zs)]𝒜~𝔭1ω1=0{\mathrm{d}}_{[\jmath_{\mathfrak{p}_{r}}(z_{s})]}\widetilde{\mathcal{A}}_{\mathfrak{p}_{1}^{*}\mathrm{\omega}_{1}}=0, by Corollary 6.5, which in turn implies

dds𝒜~𝔭1ω1([ȷ𝔭r(zs)])=d[ȷ𝔭r(zs)]𝒜~𝔭1ω1dds[ȷ𝔭r(zs)]=0.\frac{{\mathrm{d}}}{{\mathrm{d}}s}\widetilde{\mathcal{A}}_{\mathfrak{p}_{1}^{*}\mathrm{\omega}_{1}}\big{(}[\jmath_{\mathfrak{p}_{r}}(z_{s})]\big{)}={\mathrm{d}}_{[\jmath_{\mathfrak{p}_{r}}(z_{s})]}\widetilde{\mathcal{A}}_{\mathfrak{p}_{1}^{*}\mathrm{\omega}_{1}}\cdot\frac{{\mathrm{d}}}{{\mathrm{d}}s}[\jmath_{\mathfrak{p}_{r}}(z_{s})]=0.

This shows the independence of the action value from z0z_{0}.

Next, we take another ω1Ξc12(M1)\mathrm{\omega}_{1}^{\prime}\in\Xi_{c_{1}}^{2}(M_{1}) such that ω1ω1=dζ\mathrm{\omega}_{1}^{\prime}-\mathrm{\omega}_{1}={\mathrm{d}}\zeta for some one-form ζ\zeta on M1M_{1}. By Lemma 5.2.(ii), 𝔭1ζ\mathfrak{p}_{1}^{*}\zeta is normalised with respect to 𝔭1ω1\mathfrak{p}_{1}^{*}\mathrm{\omega}_{1}. Applying equation (6.11) with Ω=𝔭1ω1\Omega=\mathfrak{p}_{1}^{*}\mathrm{\omega}_{1}^{\prime}, Ω0=𝔭1ω1\Omega_{0}^{\prime}=\mathfrak{p}_{1}^{*}\mathrm{\omega}_{1}, {δr}={ȷ𝔭r(z0)}\{\delta_{r}\}=\{\jmath_{\mathfrak{p}_{r}}(z_{0})\} and {γr}={γ0}\{\gamma_{r}\}=\{\gamma_{0}\} a constant path, we find

𝒜~𝔭1ω1([ȷ𝔭r(z0)])𝒜~𝔭1ω1([ȷ𝔭r(z0)])=𝒜~𝔭1ω1([γ0])=S1γ0(𝔭1ζ)=0.\widetilde{\mathcal{A}}_{\mathfrak{p}_{1}^{*}\mathrm{\omega}_{1}^{\prime}}\big{(}[\jmath_{\mathfrak{p}_{r}}(z_{0})]\big{)}-\widetilde{\mathcal{A}}_{\mathfrak{p}_{1}^{*}\mathrm{\omega}_{1}}\big{(}[\jmath_{\mathfrak{p}_{r}}(z_{0})]\big{)}=\widetilde{\mathcal{A}}_{\mathfrak{p}_{1}^{*}\mathrm{\omega}_{1}^{\prime}}^{\prime}\big{(}[\gamma_{0}]\big{)}=\int_{S^{1}}\gamma_{0}^{*}(\mathfrak{p}_{1}^{*}\zeta)=0.

Hence, the action depends on the cohomology class c1c_{1}, not on the representative.

Step 2. Let {𝔭r}\{\mathfrak{p}_{r}\} be a path and {Ψr}\{\Psi_{r}\} an isotopy of diffeomorphisms given by Lemma 4.7.(ii) such that Ψr𝔭r=𝔭0\Psi_{r}^{*}\mathfrak{p}_{r}=\mathfrak{p}_{0}. If ψ1:M0M1\psi_{1}:M_{0}\to M_{1} is the quotient map of Ψ1\Psi_{1}, then

(i)ψ1c1=𝒜~𝔭1ω1([ȷ𝔭r(z0)])e0+c0,(ii)𝒜~𝔭1ω1([ȷ𝔭r(z0)])=0,ife0=0.(i)\quad\psi_{1}^{*}c_{1}=\widetilde{\mathcal{A}}_{\mathfrak{p}_{1}^{*}\mathrm{\omega}_{1}}\big{(}[\jmath_{\mathfrak{p}_{r}}(z_{0})]\big{)}e_{0}+c_{0},\qquad(ii)\ \ \widetilde{\mathcal{A}}_{\mathfrak{p}_{1}^{*}\mathrm{\omega}_{1}}\big{(}[\jmath_{\mathfrak{p}_{r}}(z_{0})]\big{)}=0,\quad\text{if}\ \ e_{0}=0.

Setting γ1:=ȷ𝔭0Ψ11(z0)\gamma_{1}:=\jmath_{\mathfrak{p}_{0}}\circ\Psi_{1}^{-1}(z_{0}) and using Proposition 6.10, we compute

𝒜~𝔭1ω1([ȷ𝔭r(z0)])=𝒜~𝔭1ω1([Ψr1][ȷ𝔭0Ψr1(z0)])=𝒜~Ψ1𝔭1ω1([ȷ𝔭0Ψr1(z0)])=𝒜~Ψ1𝔭1ω1([γ1])\begin{split}\widetilde{\mathcal{A}}_{\mathfrak{p}_{1}^{*}\mathrm{\omega}_{1}}\big{(}[\jmath_{\mathfrak{p}_{r}}(z_{0})]\big{)}=\widetilde{\mathcal{A}}_{\mathfrak{p}_{1}^{*}\mathrm{\omega}_{1}}\big{(}[\Psi_{r}^{-1}]^{*}[\jmath_{\mathfrak{p}_{0}}\circ\Psi_{r}^{-1}(z_{0})]\big{)}&=\widetilde{\mathcal{A}}_{\Psi_{1}^{*}\mathfrak{p}_{1}^{*}\mathrm{\omega}_{1}}\big{(}[\jmath_{\mathfrak{p}_{0}}\circ\Psi_{r}^{-1}(z_{0})]\big{)}\\ &=\widetilde{\mathcal{A}}_{\Psi_{1}^{*}\mathfrak{p}_{1}^{*}\mathrm{\omega}_{1}}\big{(}[\gamma_{1}]\big{)}\end{split} (6.19)

where the last equality follows from [ȷ𝔭0Ψr1(z0)]=[γ1][\jmath_{\mathfrak{p}_{0}}\circ\Psi_{r}^{-1}(z_{0})]=[\gamma_{1}] (see (6.5)). We observe that Ψ1𝔭1ω1=𝔭0ψ1ω1\Psi_{1}^{*}\mathfrak{p}_{1}^{*}\mathrm{\omega}_{1}=\mathfrak{p}_{0}^{*}\psi_{1}^{*}\mathrm{\omega}_{1} and thus (𝔭0,ψ1c1)C0(Σ)(\mathfrak{p}_{0},\psi_{1}^{*}c_{1})\in\mathfrak{Z}_{C}^{0}(\Sigma). By items (iii) and (iv) in Corollary 5.5, there exists A1A_{1}\in{\mathbb{R}} such that

ψ1c1=A1e0+c0.\psi_{1}^{*}c_{1}=A_{1}e_{0}+c_{0}. (6.20)

If e00e_{0}\neq 0, A1A_{1} is uniquely defined by this property. If e0=0e_{0}=0, we simply impose A1:=0A_{1}:=0. Equation (6.20) implies that there are η𝒦(𝔭0)\eta\in\mathcal{K}(\mathfrak{p}_{0}) and ζΩ1(M0)\zeta\in\Omega^{1}(M_{0}) such that

Ψ1𝔭1ω1=𝔭0ω0+d(A1η+𝔭0ζ),𝔭0κ=dη\Psi_{1}^{*}\mathfrak{p}_{1}^{*}\mathrm{\omega}_{1}=\mathfrak{p}_{0}^{*}\mathrm{\omega}_{0}+{\mathrm{d}}(A_{1}\eta+\mathfrak{p}^{*}_{0}\zeta),\qquad\mathfrak{p}_{0}^{*}\kappa={\mathrm{d}}\eta

for some κΞe02(M0)\kappa\in\Xi^{2}_{e_{0}}(M_{0}). Since A1=0A_{1}=0, if e0=0e_{0}=0, the one-form A1η+𝔭0ζA_{1}\eta+\mathfrak{p}_{0}^{*}\zeta is 𝔭0ω0\mathfrak{p}_{0}^{*}\mathrm{\omega}_{0}-normalised by Lemma 5.2.(ii) and we conclude with (6.19) that

𝒜~𝔭1ω1([ȷ𝔭r(z0)])=𝒜~Ψ1𝔭1ω1([γ1])=S1γ1(A1η+𝔭0ζ)=A1.\widetilde{\mathcal{A}}_{\mathfrak{p}_{1}^{*}\mathrm{\omega}_{1}}\big{(}[\jmath_{\mathfrak{p}_{r}}(z_{0})]\big{)}=\widetilde{\mathcal{A}}_{\Psi_{1}^{*}\mathfrak{p}_{1}^{*}\mathrm{\omega}_{1}}\big{(}[\gamma_{1}]\big{)}=\int_{S^{1}}\gamma_{1}^{*}(A_{1}\eta+\mathfrak{p}^{*}_{0}\zeta)=A_{1}.

This proves both items in Step 2.

Step 3. For any path {𝔭r}\{\mathfrak{p}_{r}\}, there holds 𝔙𝔬𝔩(𝔭1,c1)=P(𝒜~𝔭1ω1([ȷ𝔭r(z0)])){\mathfrak{Vol}}(\mathfrak{p}_{1},c_{1})=P\big{(}\widetilde{\mathcal{A}}_{\mathfrak{p}_{1}^{*}\mathrm{\omega}_{1}}\big{(}[\jmath_{\mathfrak{p}_{r}}(z_{0})]\big{)}\big{)}.

From item (iii) in Lemma 5.4 with A=0A=0, we have

𝔙𝔬𝔩(𝔭1,c1)=𝔙𝔬𝔩(𝔭0,ψ1c1)=Vol(A1η+𝔭0ζ),{\mathfrak{Vol}}(\mathfrak{p}_{1},c_{1})={\mathfrak{Vol}}(\mathfrak{p}_{0},\psi_{1}^{*}c_{1})=\mathrm{Vol}(A_{1}\eta+\mathfrak{p}_{0}^{*}\zeta),

where A1=𝒜~𝔭1ω1([ȷ𝔭r(z0)])A_{1}=\widetilde{\mathcal{A}}_{\mathfrak{p}_{1}^{*}\mathrm{\omega}_{1}}\big{(}[\jmath_{\mathfrak{p}_{r}}(z_{0})]\big{)}, η\eta and ζ\zeta are from the previous step. Using the definition of the volume (2.1) and Fubini’s Theorem, we compute

Vol(A1η+𝔭0ζ)\displaystyle\mathrm{Vol}(A_{1}\eta+\mathfrak{p}_{0}^{*}\zeta) =01(Σ(A1η+𝔭0ζ)𝔭0(rA1κ+ω0)n)dr\displaystyle=\int_{0}^{1}\Big{(}\int_{\Sigma}(A_{1}\eta+\mathfrak{p}_{0}^{*}\zeta)\wedge\mathfrak{p}_{0}^{*}\big{(}rA_{1}\kappa+\mathrm{\omega}_{0}\big{)}^{n}\Big{)}{\mathrm{d}}r
=01(M0((𝔭0)(A1η+𝔭0ζ))(rA1κ+ω0)n)dr\displaystyle=\int_{0}^{1}\Big{(}\int_{M_{0}}\big{(}(\mathfrak{p}_{0})_{*}(A_{1}\eta+\mathfrak{p}_{0}^{*}\zeta)\big{)}\wedge\big{(}rA_{1}\kappa+\mathrm{\omega}_{0}\big{)}^{n}\Big{)}{\mathrm{d}}r
=01A1(rA1e0+c0)n,[M0]dr\displaystyle=\int_{0}^{1}A_{1}\big{\langle}(rA_{1}e_{0}+c_{0})^{n},[M_{0}]\big{\rangle}{\mathrm{d}}r
=01A1Q(rA1)dr\displaystyle=\int_{0}^{1}A_{1}Q(rA_{1}){\mathrm{d}}r
=P(A1)P(0)\displaystyle=P(A_{1})-P(0)
=P(A1).\displaystyle=P(A_{1}).

Step 4. If (𝔭1,c1)=(𝔭0,c0)(\mathfrak{p}_{1},c_{1})=(\mathfrak{p}_{0},c_{0}), then, for any path {𝔭r}\{\mathfrak{p}_{r}\}, we have 𝒜~𝔭1ω1([ȷ𝔭r(z0)])=0\widetilde{\mathcal{A}}_{\mathfrak{p}_{1}^{*}\mathrm{\omega}_{1}}\big{(}[\jmath_{\mathfrak{p}_{r}}(z_{0})]\big{)}=0.

Let ψ1\psi_{1} and A1=𝒜~𝔭1ω1([ȷ𝔭r(z0)])A_{1}=\widetilde{\mathcal{A}}_{\mathfrak{p}_{1}^{*}\mathrm{\omega}_{1}}\big{(}[\jmath_{\mathfrak{p}_{r}}(z_{0})]\big{)} as in Step 2. Since ψ1e0=e0\psi_{1}^{*}e_{0}=e_{0}, applying ψ1\psi_{1} iteratively to item (i) in Step 2, we obtain

(ψ1m)c0=mA1e0+c0,m.(\psi^{m}_{1})^{*}c_{0}=mA_{1}e_{0}+c_{0},\qquad\forall\,m\in{\mathbb{Z}}.

From item (ii) in Lemma 5.4 with A=0A=0, we deduce

Q(0)=ev(𝔭0,c0)=ev(𝔭0,(ψ1m)c0)=ev(𝔭0,mA1e0+c0)=Q(mA1),m.\displaystyle Q(0)={\mathrm{ev}}(\mathfrak{p}_{0},c_{0})={\mathrm{ev}}(\mathfrak{p}_{0},(\psi^{m}_{1})^{*}c_{0})={\mathrm{ev}}(\mathfrak{p}_{0},mA_{1}e_{0}+c_{0})=Q(mA_{1}),\qquad\forall\,m\in{\mathbb{Z}}.

We assume by contradiction that A10A_{1}\neq 0. Since QQ is a polynomial, the above identity implies that Q(A)Q0Q(A)\equiv Q_{0}, where Q0Q_{0}\in{\mathbb{R}}, so that P(A)=Q0AP(A)=Q_{0}A for all AA\in{\mathbb{R}}. Since C0(Σ)\mathfrak{Z}_{C}^{0}(\Sigma) is non-degenerate, the coefficient Q0Q_{0} is non-zero. This contradicts Step 3:

0=𝔙𝔬𝔩(𝔭0,c0)=P(A1)=Q0A1.0={\mathfrak{Vol}}(\mathfrak{p}_{0},c_{0})=P(A_{1})=Q_{0}A_{1}.

Step 5. End of the proof.

To show that the functional 𝒜:C0(Σ)\mathcal{A}:\mathfrak{Z}^{0}_{C}(\Sigma)\to{\mathbb{R}} is well-defined, it remains to see that the action value 𝒜~𝔭1ω1([ȷ𝔭r(z0)])\widetilde{\mathcal{A}}_{\mathfrak{p}_{1}^{*}\mathrm{\omega}_{1}}\big{(}[\jmath_{\mathfrak{p}_{r}}(z_{0})]\big{)} does not depend on the choice of path {𝔭r}\{\mathfrak{p}_{r}\}. Indeed, if {𝔭r}\{\mathfrak{p}^{\prime}_{r}\} is another path with 𝔭0=𝔭0\mathfrak{p}^{\prime}_{0}=\mathfrak{p}_{0} and 𝔭1=𝔭1\mathfrak{p}_{1}^{\prime}=\mathfrak{p}_{1}, the concatenation {𝔭r}#{𝔭1r}\{\mathfrak{p}^{\prime}_{r}\}\#\{\mathfrak{p}_{1-r}\} forms a loop based at 𝔭0\mathfrak{p}_{0}. Applying Step 4 and the definition of the action (6.8), we conclude

𝒜~𝔭1ω1([ȷ𝔭r(z0)])𝒜~𝔭1ω1([ȷ𝔭r(z0)])=𝒜~𝔭0ω0([ȷ𝔭r#𝔭1r(z0)])=0.\widetilde{\mathcal{A}}_{\mathfrak{p}_{1}^{*}\mathrm{\omega}_{1}}\big{(}[\jmath_{\mathfrak{p}^{\prime}_{r}}(z_{0})]\big{)}-\widetilde{\mathcal{A}}_{\mathfrak{p}_{1}^{*}\mathrm{\omega}_{1}}\big{(}[\jmath_{\mathfrak{p}_{r}}(z_{0})]\big{)}=\widetilde{\mathcal{A}}_{\mathfrak{p}_{0}^{*}\mathrm{\omega}_{0}}\big{(}[\jmath_{\mathfrak{p}^{\prime}_{r}\#\mathfrak{p}_{1-r}}(z_{0})]\big{)}=0.

Finally, the identities claimed in the statement of the theorem follow directly from what we have proven. Equation (6.16) is a consequence of item (ii) in Step 2 and of Step 4. Equation (6.17) follows from Step 3. Item (i) in Step 2 implies (6.18). ∎

Remark 6.15.

If e0=0e_{0}=0, we could also have considered the space of weakly Zoll one-forms ¯C0(Σ)\bar{\mathfrak{Z}}_{C}^{0}(\Sigma), whose elements are pairs (𝔭,α)(\mathfrak{p},\alpha) with (𝔭,[ωα])C0(Σ)(\mathfrak{p},[\mathrm{\omega}_{\alpha}])\in\mathfrak{Z}_{C}^{0}(\Sigma), where 𝔭ωα=Ω0+dα\mathfrak{p}^{*}\mathrm{\omega}_{\alpha}=\Omega_{0}+{\mathrm{d}}\alpha, and α\alpha not necessarily normalised. We have an action functional

𝒜¯:¯C0(Σ),𝒜¯(𝔭,α):=𝒜~α([ȷ𝔭r(z0)]).\bar{\mathcal{A}}:\bar{\mathfrak{Z}}_{C}^{0}(\Sigma)\to{\mathbb{R}},\qquad\bar{\mathcal{A}}(\mathfrak{p},\alpha):=\widetilde{\mathcal{A}}_{\alpha}([\jmath_{\mathfrak{p}_{r}}(z_{0})]).

Since C0(Σ)\mathfrak{Z}_{C}^{0}(\Sigma) is non-degenerate, from Remark 6.13 and Theorem 6.14, we deduce

Vol(α)=c0n,[M0]𝒜¯(𝔭,α),(𝔭,α)¯C0(Σ).\mathrm{Vol}(\alpha)=\langle c_{0}^{n},[M_{0}]\rangle\cdot\bar{\mathcal{A}}(\mathfrak{p},\alpha),\qquad\forall\,(\mathfrak{p},\alpha)\in\bar{\mathfrak{Z}}_{C}^{0}(\Sigma).
Remark 6.16.

Let Ω𝒵C(Σ)\Omega\in{\mathcal{Z}}_{C}(\Sigma) be a Zoll odd-symplectic form such that the associated bundle 𝔭Ω\mathfrak{p}_{\Omega} belongs to 𝔓0(Σ)\mathfrak{P}^{0}(\Sigma). If we define

𝒜(Ω):=𝒜(𝔭Ω,[ωΩ]),Ω=𝔭ΩωΩ,\mathcal{A}(\Omega):=\mathcal{A}\big{(}\mathfrak{p}_{\Omega},[\mathrm{\omega}_{\Omega}]\big{)},\qquad\Omega=\mathfrak{p}_{\Omega}^{*}\mathrm{\omega}_{\Omega},

then Theorem 6.14 translates into

𝔙𝔬𝔩(Ω)=P(𝒜(Ω)),Ω𝒵C(Σ).{\mathfrak{Vol}}(\Omega)=P\big{(}\mathcal{A}(\Omega)\big{)},\qquad\forall\,\Omega\in{\mathcal{Z}}_{C}(\Sigma).

As we did for the volume and the action, we study how the Zoll polynomial behaves under pull-back and change of pair (𝔭0,c0)(\mathfrak{p}_{0},c_{0}) within C0(Σ)\mathfrak{Z}_{C}^{0}(\Sigma).

Proposition 6.17.

Let 𝔭0:ΣM0\mathfrak{p}_{0}^{\text{\tiny$\vee$}}:\Sigma^{\text{\tiny$\vee$}}\to M_{0}^{\text{\tiny$\vee$}} be an oriented S1S^{1}-bundle with maps Π:ΣΣ\Pi:\Sigma^{\text{\tiny$\vee$}}\to\Sigma and π0:M0M0\pi_{0}:M_{0}^{\text{\tiny$\vee$}}\to M_{0} as described in (4.1). If PP^{\text{\tiny$\vee$}} is the Zoll polynomial of the weakly Zoll pair (Π𝔭0,π0c0)=(𝔭0,c0)(\Pi^{*}\mathfrak{p}_{0},\pi_{0}^{*}c_{0})=(\mathfrak{p}_{0}^{\text{\tiny$\vee$}},c_{0}^{\text{\tiny$\vee$}}), there holds

P=(degπ0)P.P^{\text{\tiny$\vee$}}=(\deg\pi_{0})\cdot P.
Proof.

From equations (6.12) and (6.13), we see that Q=(degπ0)QQ^{\text{\tiny$\vee$}}=(\deg\pi_{0})\cdot Q, where QQ^{\text{\tiny$\vee$}} is the derivative of PP^{\text{\tiny$\vee$}}. Since P(0)=0=P(0)P(0)=0=P^{\text{\tiny$\vee$}}(0), the statement follows. ∎

Proposition 6.18.

Let (𝔭0,c0)(\mathfrak{p}_{0}^{\prime},c_{0}^{\prime}) be a pair in C0(Σ)\mathfrak{Z}_{C}^{0}(\Sigma) with 𝔭0:ΣM0\mathfrak{p}_{0}^{\prime}:\Sigma\to M_{0}^{\prime}. There holds

(c0)n,[M0]=ev(𝔭0,c0)=dPdA(𝒜(𝔭0,c0)).\langle(c_{0}^{\prime})^{n},[M_{0}^{\prime}]\rangle={\mathrm{ev}}(\mathfrak{p}_{0}^{\prime},c_{0}^{\prime})=\frac{{\mathrm{d}}P}{{\mathrm{d}}A}\big{(}\mathcal{A}(\mathfrak{p}_{0}^{\prime},c_{0}^{\prime})\big{)}.

If PP^{\prime} is the Zoll polynomial associated with (𝔭0,c0)(\mathfrak{p}_{0}^{\prime},c_{0}^{\prime}), then

P(A)=P(A+𝒜(𝔭0,c0))P(𝒜(𝔭0,c0)),A.P^{\prime}(A)=P\big{(}A+\mathcal{A}(\mathfrak{p}_{0}^{\prime},c_{0}^{\prime})\big{)}-P\big{(}\mathcal{A}(\mathfrak{p}_{0}^{\prime},c_{0}^{\prime})\big{)},\qquad\forall\,A\in{\mathbb{R}}.
Proof.

If QQ^{\prime} denotes the derivative of PP^{\prime}, by Lemma 5.4.(ii) and (6.18), we have

Q(A)=Q(A+𝒜(𝔭0,c0)).Q^{\prime}(A)=Q\big{(}A+\mathcal{A}(\mathfrak{p}_{0}^{\prime},c_{0}^{\prime})\big{)}.

Setting A=0A=0, we get the first part of the statement. For the second part, we integrate the above identity and use the normalization P(0)=0P^{\prime}(0)=0:

P(A)=0AQ(A)dA=0AQ(A+𝒜(𝔭0,c0))dA=P(A+𝒜(𝔭0,c0))P(𝒜(𝔭0,c0)).P^{\prime}(A)=\int_{0}^{A}Q^{\prime}(A^{\prime}){\mathrm{d}}A^{\prime}=\int_{0}^{A}Q\big{(}A^{\prime}+\mathcal{A}(\mathfrak{p}_{0}^{\prime},c_{0}^{\prime})\big{)}{\mathrm{d}}A^{\prime}=P\big{(}A+\mathcal{A}(\mathfrak{p}_{0}^{\prime},c_{0}^{\prime})\big{)}-P\big{(}\mathcal{A}(\mathfrak{p}_{0}^{\prime},c_{0}^{\prime})\big{)}.\qed

Combining the result above with the transformation rules for the action and the volume under change of reference weakly Zoll pair, we conclude the following invariance property.

Corollary 6.19.

Let PP^{\prime}, 𝒜~\widetilde{\mathcal{A}}, 𝔙𝔬𝔩{\mathfrak{Vol}}^{\prime} be the Zoll polynomial, the action and the volume associated with another reference pair (𝔭0,c0)C0(Σ)(\mathfrak{p}_{0}^{\prime},c_{0}^{\prime})\in\mathfrak{Z}_{C}^{0}(\Sigma). For every ΩΞC2(Σ)\Omega\in\Xi^{2}_{C}(\Sigma), there holds

P𝒜~Ω𝔙𝔬𝔩(Ω)=P𝒜~Ω𝔙𝔬𝔩(Ω)on Λ~𝔥(Σ).P^{\prime}\circ\widetilde{\mathcal{A}}_{\Omega}^{\prime}-{\mathfrak{Vol}}^{\prime}(\Omega)=P\circ\widetilde{\mathcal{A}}_{\Omega}-{\mathfrak{Vol}}(\Omega)\qquad\text{on }\ \widetilde{\Lambda}_{\mathfrak{h}}(\Sigma).
Proof.

We preliminarily observe that

𝒜~Ω=𝒜(𝔭0,c0)+𝒜~Ω\widetilde{\mathcal{A}}_{\Omega}=\mathcal{A}(\mathfrak{p}_{0}^{\prime},c_{0}^{\prime})+\widetilde{\mathcal{A}}_{\Omega}^{\prime}

thanks to Proposition 6.7. We also recall that Lemma 2.4 implies that

𝔙𝔬𝔩(Ω)=𝔙𝔬𝔩(𝔭0,c0)+𝔙𝔬𝔩(Ω).{\mathfrak{Vol}}(\Omega)={\mathfrak{Vol}}(\mathfrak{p}_{0}^{\prime},c_{0}^{\prime})+{\mathfrak{Vol}}^{\prime}(\Omega).

Using these identities together with Proposition 6.18 and Theorem 6.14, we readily compute

P(𝒜~Ω)𝔙𝔬𝔩(Ω)\displaystyle P(\widetilde{\mathcal{A}}_{\Omega})-{\mathfrak{Vol}}(\Omega) =P(𝒜~Ω)𝔙𝔬𝔩(𝔭0,c0)𝔙𝔬𝔩(Ω)\displaystyle=P(\widetilde{\mathcal{A}}_{\Omega})-{\mathfrak{Vol}}(\mathfrak{p}_{0}^{\prime},c_{0}^{\prime})-{\mathfrak{Vol}}^{\prime}(\Omega)
=P(𝒜~Ω+𝒜(𝔭0,c0))𝔙𝔬𝔩(𝔭0,c0)𝔙𝔬𝔩(Ω)\displaystyle=P\big{(}\widetilde{\mathcal{A}}_{\Omega}^{\prime}+\mathcal{A}(\mathfrak{p}_{0}^{\prime},c_{0}^{\prime})\big{)}-{\mathfrak{Vol}}(\mathfrak{p}_{0}^{\prime},c_{0}^{\prime})-{\mathfrak{Vol}}^{\prime}(\Omega)
=P(𝒜~Ω)+P(𝒜(𝔭0,c0))𝔙𝔬𝔩(𝔭0,c0)𝔙𝔬𝔩(Ω)\displaystyle=P^{\prime}(\widetilde{\mathcal{A}}_{\Omega}^{\prime})+P\big{(}\mathcal{A}(\mathfrak{p}_{0}^{\prime},c_{0}^{\prime})\big{)}-{\mathfrak{Vol}}(\mathfrak{p}_{0}^{\prime},c_{0}^{\prime})-{\mathfrak{Vol}}^{\prime}(\Omega)
=P(𝒜~Ω)𝔙𝔬𝔩(Ω).\displaystyle=P^{\prime}(\widetilde{\mathcal{A}}_{\Omega}^{\prime})-{\mathfrak{Vol}}^{\prime}(\Omega).\qed

7 A conjectural local systolic-diastolic inequality

The following is our setting to study a systolic-diastolic inequality for odd-symplectic forms. We assume that there exists a Zoll odd-symplectic form Ω𝒵C(Σ)\Omega_{*}\in{\mathcal{Z}}_{C}(\Sigma), for some CHdR2(Σ)C\in H^{2}_{\mathrm{dR}}(\Sigma). We denote the bundle associated with Ω\Omega_{*} by

𝔭1:=𝔭Ω:ΣM1:=MΩ\mathfrak{p}_{1}:=\mathfrak{p}_{\Omega_{*}}:\Sigma\to M_{1}:=M_{\Omega_{*}}

and let ωΞ2(M1)\mathrm{\omega}_{*}\in\Xi^{2}(M_{1}) be the symplectic form such that Ω=𝔭1ω\Omega_{*}=\mathfrak{p}_{1}^{*}\mathrm{\omega}_{*}. The space 𝔓0(Σ)\mathfrak{P}^{0}(\Sigma) is the connected component of 𝔭1\mathfrak{p}_{1} in 𝔓(Σ)\mathfrak{P}(\Sigma) and C0(Σ)\mathfrak{Z}_{C}^{0}(\Sigma) is the corresponding space of weakly Zoll pairs, which is non-empty and non-degenerate thanks to Remark 5.7. Namely, condition (6.3) is automatically satisfied and we can use all the results contained in Section 6.

We fix a reference weakly Zoll pair (𝔭0,c0)C0(Σ)(\mathfrak{p}_{0},c_{0})\in\mathfrak{Z}_{C}^{0}(\Sigma), where

𝔭0:ΣM0\mathfrak{p}_{0}:\Sigma\to M_{0}

and let ω0\mathrm{\omega}_{0} be some form in Ξc02(M0)\Xi^{2}_{c_{0}}(M_{0}). Let 𝔥[S1,Σ]\mathfrak{h}\in[S^{1},\Sigma] denote the free-homotopy class of the 𝔭0\mathfrak{p}_{0}-fibres. We define the volume and the action functionals

𝔙𝔬𝔩:ΞC2(Σ),𝒜:C0(Σ),𝒜~Ω:Λ~𝔥(Σ){\mathfrak{Vol}}:\Xi^{2}_{C}(\Sigma)\to{\mathbb{R}},\qquad\mathcal{A}:\mathfrak{Z}_{C}^{0}(\Sigma)\to{\mathbb{R}},\qquad\widetilde{\mathcal{A}}_{\Omega}:\widetilde{\Lambda}_{\mathfrak{h}}(\Sigma)\to{\mathbb{R}}

with respect to (𝔭0,c0)(\mathfrak{p}_{0},c_{0}). Due to Proposition 6.18, we have

dPdA(A)=[ω]n,[M1]>0,A:=𝒜(𝔭1,ω).\frac{{\mathrm{d}}P}{{\mathrm{d}}A}(A_{*})=\langle[\mathrm{\omega}_{*}]^{n},[M_{1}]\rangle>0,\qquad A_{*}:=\mathcal{A}(\mathfrak{p}_{1},\mathrm{\omega}_{*}). (7.1)
Definition 7.1.

We say that a finite cover {Bi}\{B_{i}\} of M1M_{1} is admissible, if all the sets BiB_{i} and their intersections BiBjB_{i}\cap B_{j} are homeomorphic to the open Euclidean ball, if non-empty. We write {Σi:=𝔭11(Bi)}\{\Sigma_{i}:=\mathfrak{p}_{1}^{-1}(B_{i})\} for the pull-back cover of Σ\Sigma, so that ΣiΣj\Sigma_{i}\cap\Sigma_{j} retracts to a 𝔭1\mathfrak{p}_{1}-fibre, if non-empty. We readily see that admissible finite covers exist and we fix one of them throughout the present section.

We also consider another oriented S1S^{1}-bundle 𝔭0:ΣM0\mathfrak{p}_{0}^{\text{\tiny$\vee$}}:\Sigma^{\text{\tiny$\vee$}}\to M_{0}^{\text{\tiny$\vee$}} with bundle map Π:ΣΣ\Pi:\Sigma^{\text{\tiny$\vee$}}\to\Sigma such that 𝔭0=Π𝔭0\mathfrak{p}_{0}^{\text{\tiny$\vee$}}=\Pi^{*}\mathfrak{p}_{0}. We further assume that Π\Pi is a finite covering map. This is equivalent to asking that the corresponding quotient map π0:M0M0\pi_{0}:M_{0}^{\text{\tiny$\vee$}}\to M_{0}, which fits into the first commutative diagram in (4.1), is a finite covering map. Thus, we endow M0M_{0}^{\text{\tiny$\vee$}} and Σ\Sigma^{\text{\tiny$\vee$}} with the orientations pulled back by π0\pi_{0} and Π\Pi, respectively.

The pull-back form Ω:=ΠΩ\Omega_{*}^{\text{\tiny$\vee$}}:=\Pi^{*}\Omega_{*} is Zoll odd-symplectic on Σ\Sigma^{\text{\tiny$\vee$}} and defines the oriented S1S^{1}-bundle 𝔭1:=Π𝔭1:ΣM1\mathfrak{p}_{1}^{\text{\tiny$\vee$}}:=\Pi^{*}\mathfrak{p}_{1}:\Sigma^{\text{\tiny$\vee$}}\to M_{1}^{\text{\tiny$\vee$}}. Since 𝔭0\mathfrak{p}_{0} and 𝔭1\mathfrak{p}_{1} are connected by a path of oriented S1S^{1}-bundles, the quotient map π1:M1M1\pi_{1}:M_{1}^{\text{\tiny$\vee$}}\to M_{1} is also an orientation-preserving finite covering map, which fits into the first commutative diagram in (4.1). Moreover, π1\pi_{1} evenly covers all the sets in {Bi}\{B_{i}\}, as they are contractible. Therefore, the family {Bj}\{B_{j}^{\text{\tiny$\vee$}}\}, whose elements are the connected components of the pre-images π11(Bi)\pi_{1}^{-1}(B_{i}), yields a finite admissible cover of M1M_{1}^{\text{\tiny$\vee$}}.

7.1 A local primitive for the action form

We have seen that the action form does not admit a global primitive on Λ𝔥(Σ)\Lambda_{\mathfrak{h}}(\Sigma) in general. However, we can always find a local primitive on the space of periodic curves that are close to ȷ𝔭1(Σ)\jmath_{\mathfrak{p}_{1}}(\Sigma). Let Λ(𝔭1)\Lambda(\mathfrak{p}_{1}) be the subset of Λ𝔥(Σ)\Lambda_{\mathfrak{h}}(\Sigma), whose elements γ\gamma are contained in some Σi\Sigma_{i} and are homotopic within Σi\Sigma_{i} to some 𝔭1\mathfrak{p}_{1}-fibre. This means that γ\gamma admits a short homotopy {γrshort}\{\gamma^{\mathrm{short}}_{r}\}. Namely, such a homotopy has the properties

γrshortΣi,r[0,1],γ0short=ȷ𝔭1(zi),γ1short=γ,\gamma^{\mathrm{short}}_{r}\subset\Sigma_{i},\ \ \forall\,r\in[0,1],\qquad\gamma^{\mathrm{short}}_{0}=\jmath_{\mathfrak{p}_{1}}(z_{i}),\ \ \gamma^{\mathrm{short}}_{1}=\gamma,

for some ziΣiz_{i}\in\Sigma_{i}. We choose any path {𝔭r}\{\mathfrak{p}_{r}\} from 𝔭0\mathfrak{p}_{0} and 𝔭1\mathfrak{p}_{1} in 𝔓0(Σ)\mathfrak{P}^{0}(\Sigma), and consider the map

Λ(𝔭1)Λ~𝔥(Σ),γ[{ȷ𝔭r(zi)}#{γrshort}]\Lambda(\mathfrak{p}_{1})\to\widetilde{\Lambda}_{\mathfrak{h}}(\Sigma),\qquad\gamma\mapsto\big{[}\{\jmath_{\mathfrak{p}_{r}}(z_{i})\}\#\{\gamma_{r}^{\mathrm{short}}\}\big{]}

Since the intersection of two elements of the finite admissible cover is contractible, this map does not depend on the particular short homotopy nor on the set Σi\Sigma_{i}. Composing this map with the functional 𝒜~Ω\widetilde{\mathcal{A}}_{\Omega} on Λ~𝔥(Σ)\widetilde{\Lambda}_{\mathfrak{h}}(\Sigma), we define the action functional on Λ(𝔭1)\Lambda(\mathfrak{p}_{1}):

𝒜Ω:Λ(𝔭1),𝒜Ω(γ):=𝒜~Ω([{ȷ𝔭r(zi)}#{γrshort}]).\mathcal{A}_{\Omega}:\Lambda(\mathfrak{p}_{1})\to{\mathbb{R}}\,,\qquad\mathcal{A}_{\Omega}(\gamma):=\widetilde{\mathcal{A}}_{\Omega}\Big{(}\big{[}\{\jmath_{\mathfrak{p}_{r}}(z_{i})\}\#\{\gamma_{r}^{\mathrm{short}}\}\big{]}\Big{)}. (7.2)

We note that d𝒜Ω=𝔞(Ω){\mathrm{d}}\mathcal{A}_{\Omega}=\mathfrak{a}(\Omega) on Λ(𝔭1)\Lambda(\mathfrak{p}_{1}).

Lemma 7.2.

The action value 𝒜Ω(γ)\mathcal{A}_{\Omega}(\gamma) does not depend on the choice of path {𝔭r}\{\mathfrak{p}_{r}\}.

Proof.

If {𝔭r}\{\mathfrak{p}_{r}^{\prime}\} is another path in 𝔓0(Σ)\mathfrak{P}^{0}(\Sigma) with 𝔭0=𝔭0\mathfrak{p}^{\prime}_{0}=\mathfrak{p}_{0} and 𝔭1=𝔭1\mathfrak{p}^{\prime}_{1}=\mathfrak{p}_{1}, then, by (6.8) and (6.16),

𝒜~Ω([{ȷ𝔭r(zi)}#{γrshort}])𝒜~Ω([{ȷ𝔭r(zi)}#{γrshort}])=𝒜(𝔭0,c0)=0.\widetilde{\mathcal{A}}_{\Omega}\Big{(}\big{[}\{\jmath_{\mathfrak{p}_{r}}(z_{i})\}\#\{\gamma_{r}^{\mathrm{short}}\}\big{]}\Big{)}-\widetilde{\mathcal{A}}_{\Omega}\Big{(}\big{[}\{\jmath_{\mathfrak{p}_{r}^{\prime}}(z_{i})\}\#\{\gamma_{r}^{\mathrm{short}}\}\big{]}\Big{)}={\mathcal{A}}(\mathfrak{p}_{0},c_{0})=0.\qed

We write 𝒳(Ω;𝔭1):=𝒳(Ω)Λ(𝔭1)\mathcal{X}(\Omega;\mathfrak{p}_{1}):=\mathcal{X}(\Omega)\cap\Lambda(\mathfrak{p}_{1}) for the set of closed characteristics of Ω\Omega in Λ(𝔭1)\Lambda(\mathfrak{p}_{1}). The systole over the set Λ(𝔭1)\Lambda(\mathfrak{p}_{1}) is the functional

𝒜min:ΞC2(Σ){,+},𝒜min(Ω):=infγ𝒳(Ω;𝔭1)𝒜Ω(γ),\mathcal{A}_{\min}:\Xi^{2}_{C}(\Sigma)\to{\mathbb{R}}\cup\{-\infty,+\infty\},\qquad\mathcal{A}_{\min}(\Omega):=\inf_{\gamma\in\mathcal{X}(\Omega;\mathfrak{p}_{1})}\mathcal{A}_{\Omega}(\gamma), (7.3)

while the diastole over the set Λ(𝔭1)\Lambda(\mathfrak{p}_{1}) is the functional

𝒜max:ΞC2(Σ){,+},𝒜max(Ω):=supγ𝒳(Ω;𝔭1)𝒜Ω(γ).\mathcal{A}_{\max}:\Xi^{2}_{C}(\Sigma)\to{\mathbb{R}}\cup\{-\infty,+\infty\}\,,\qquad\mathcal{A}_{\max}(\Omega):=\sup_{\gamma\in\mathcal{X}(\Omega;\mathfrak{p}_{1})}\mathcal{A}_{\Omega}(\gamma). (7.4)
Proposition 7.3.

Let Π:ΣΣ\Pi:\Sigma^{\text{\tiny$\vee$}}\to\Sigma be a bundle map with 𝔭0=Π𝔭0\mathfrak{p}_{0}^{\text{\tiny$\vee$}}=\Pi^{*}\mathfrak{p}_{0}, which is also a finite covering map. Let us consider the oriented S1S^{1}-bundle 𝔭1=Π𝔭1:ΣM1\mathfrak{p}_{1}^{\text{\tiny$\vee$}}=\Pi^{*}\mathfrak{p}_{1}:\Sigma^{\text{\tiny$\vee$}}\to M_{1}^{\text{\tiny$\vee$}}. Then, Π:(γΛ(𝔭1))(ΠγΛ(𝔭1))\Pi_{*}:(\gamma^{\text{\tiny$\vee$}}\in\Lambda(\mathfrak{p}_{1}^{\text{\tiny$\vee$}}))\mapsto(\Pi\circ\gamma^{\text{\tiny$\vee$}}\in\Lambda(\mathfrak{p}_{1})) is a (surjective) finite covering map and there holds

𝒜ΠΩ=𝒜ΩΠ,ΩΞC2(Σ),\mathcal{A}^{\text{\tiny$\vee$}}_{\Pi^{*}\Omega}=\mathcal{A}_{\Omega}\circ\Pi_{*},\qquad\forall\,\Omega\in\Xi^{2}_{C}(\Sigma),

where 𝒜ΠΩ\mathcal{A}^{\text{\tiny$\vee$}}_{\Pi^{*}\Omega} is the action functional on Λ(𝔭1)\Lambda(\mathfrak{p}_{1}^{\text{\tiny$\vee$}}) with reference pair (𝔭0,c0)(\mathfrak{p}_{0}^{\text{\tiny$\vee$}},c_{0}^{\text{\tiny$\vee$}}). As a consequence, the restriction map Π:𝒳(ΠΩ;𝔭1)𝒳(Ω;𝔭1)\Pi_{*}:\mathcal{X}(\Pi^{*}\Omega;\mathfrak{p}_{1}^{\text{\tiny$\vee$}})\rightarrow\mathcal{X}(\Omega;\mathfrak{p}_{1}) is also a finite cover and we have

𝒜min(ΠΩ)=𝒜min(Ω),𝒜max(ΠΩ)=𝒜max(Ω).\mathcal{A}^{\text{\tiny$\vee$}}_{\min}(\Pi^{*}\Omega)=\mathcal{A}_{\min}(\Omega),\qquad\mathcal{A}_{\max}^{\text{\tiny$\vee$}}(\Pi^{*}\Omega)=\mathcal{A}_{\max}(\Omega).
Proof.

The statement follows from Proposition 6.9, since Π\Pi^{*} lifts a path {𝔭r}\{\mathfrak{p}_{r}\} in 𝔓0(Σ)\mathfrak{P}^{0}(\Sigma) to a path {𝔭r}\{\mathfrak{p}_{r}^{\text{\tiny$\vee$}}\} in 𝔓0(Σ)\mathfrak{P}^{0}(\Sigma^{\text{\tiny$\vee$}}) and Π\Pi maps short homotopies for 𝔭1\mathfrak{p}_{1}^{\text{\tiny$\vee$}} to short homotopies for 𝔭1\mathfrak{p}_{1}. ∎

We recall some results from [Gin87], ensuring that the systole and the diastole of Ω\Omega are finite, if Ω\Omega is an odd-symplectic form sufficiently close to Ω\Omega_{*} in the same cohomology class. As usual, we measure distances and norms by fixing Riemannian metrics on the relevant spaces.

Proposition 7.4.

There exists a C1C^{1}-neighbourhood 𝒱\mathcal{V} of Ω\Omega_{*} in 𝒮C(Σ)\mathcal{S}_{C}(\Sigma) such that, for all Ω𝒱\Omega\in\mathcal{V}, the following properties hold:

  1. (i)

    There is a constant C>0C>0 depending on 𝒱\mathcal{V} but not on Ω\Omega such that

    supt(dist(qγ(0),qγ(t)))CΩΩC0,qγ:=𝔭1γ,γ𝒳(Ω;𝔭1).\sup_{t\in{\mathbb{R}}}\Big{(}{\mathrm{dist}}\big{(}q_{\gamma}(0),q_{\gamma}(t)\big{)}\Big{)}\leq C\|\Omega-\Omega_{*}\|_{C^{0}},\qquad q_{\gamma}:=\mathfrak{p}_{1}\circ\gamma,\qquad\forall\,\gamma\in\mathcal{X}(\Omega;\mathfrak{p}_{1}).
  2. (ii)

    The set 𝒳(Ω;𝔭1)\mathcal{X}(\Omega;\mathfrak{p}_{1}) is compact and non-empty and the restrictions 𝒜min|𝒱:𝒱\mathcal{A}_{\min}|_{\mathcal{V}}:\mathcal{V}\to{\mathbb{R}}, 𝒜max|𝒱:𝒱\mathcal{A}_{\max}|_{\mathcal{V}}:\mathcal{V}\to{\mathbb{R}} are Lipschitz-continuous real functions.

  3. (iii)

    If Ω𝒱\Omega\in\mathcal{V} is a Zoll form, then all its closed characteristics (up to orientation) lie in Λ(𝔭1)\Lambda(\mathfrak{p}_{1}), i.e. 𝒳(Ω)=𝒳(Ω;𝔭1)\mathcal{X}(\Omega)=\mathcal{X}(\Omega;\mathfrak{p}_{1}), and, recalling the notation 𝒜(Ω)=𝒜(𝔭Ω,[ωΩ])\mathcal{A}(\Omega)=\mathcal{A}(\mathfrak{p}_{\Omega},[\mathrm{\omega}_{\Omega}]),

    𝒜min(Ω)=𝒜(Ω)=𝒜max(Ω).\mathcal{A}_{\min}(\Omega)=\mathcal{A}(\Omega)=\mathcal{A}_{\max}(\Omega).
Proof.

Let η𝒦(𝔭1)\eta\in\mathcal{K}(\mathfrak{p}_{1}), and let VV be the vertical vector field with η(V)=1\eta(V)=1. If Ω\Omega is odd-symplectic, the equation

ιXΩ(ηΩn)=Ωn\iota_{X_{\Omega}}(\eta\wedge\Omega_{*}^{n})=\Omega^{n}

determines a nowhere vanishing section XΩkerΩX_{\Omega}\in\ker\Omega. The periodic orbits of the flow ΦXΩ\Phi^{X_{\Omega}} yield elements in 𝒳(Ω)\mathcal{X}(\Omega). Since V=XΩV=X_{\Omega_{*}}, we have

XΩVC1C1ΩΩC1\|X_{\Omega}-V\|_{C^{1}}\leq C_{1}\|\Omega-\Omega_{*}\|_{C^{1}}

for some constant C1>0C_{1}>0. Using this estimate, one can argue as in [BK19a, Lemma 3.3, Proposition 3.4] to show that there exists a C1C^{1}-neighbourhood 𝒱𝒮C(Σ)\mathcal{V}\subset\mathcal{S}_{C}(\Sigma) of Ω\Omega_{*} such that, for all Ω𝒱\Omega\in\mathcal{V} item (i) holds and T(γ)2T(\gamma)\leq 2, where T(γ)T(\gamma) is the period of γ𝒳(Ω;𝔭1)\gamma\in\mathcal{X}(\Omega;\mathfrak{p}_{1}).

One can follow [Gin87, Section III] or [ÁPB14, Section 3.2] to show that there exists a Lipschitz-continuous map 𝒴:𝒱×ΣΛ(𝔭1)\mathcal{Y}:\mathcal{V}\times\Sigma\to\Lambda(\mathfrak{p}_{1}) with the following properties. If one defines, for every Ω𝒱\Omega\in\mathcal{V}, the function S(Ω):ΣS(\Omega):\Sigma\to{\mathbb{R}} by S(Ω)(z):=𝒜Ω(𝒴(Ω,z))S(\Omega)(z):=\mathcal{A}_{\Omega}\big{(}\mathcal{Y}(\Omega,z)\big{)} for all zΣz\in\Sigma, then one gets a Lipschitz function S:𝒱C1(Σ)S:\mathcal{V}\to C^{1}(\Sigma) such that

CritS(Ω)𝒳(Ω;𝔭1),z𝒴(Ω,z)\mathrm{Crit\,}S(\Omega)\longrightarrow\mathcal{X}(\Omega;\mathfrak{p}_{1}),\qquad z\mapsto\mathcal{Y}(\Omega,z)

is a one-to-one correspondence. In particular,

minS(Ω)=𝒜min(Ω),maxS(Ω)=𝒜max(Ω).\min S(\Omega)=\mathcal{A}_{\min}(\Omega),\qquad\max S(\Omega)=\mathcal{A}_{\max}(\Omega).

Since the functions max,min:C1(Σ)\max,\min:C^{1}(\Sigma)\to{\mathbb{R}} are Lipschitz-continuous, this shows (ii).

To prove (iii), let ρLeb>0\rho_{\mathrm{Leb}}>0 be a Lebesgue number for the open cover {Bi}\{B_{i}\} of M1M_{1} and let CC be the constant given in (i). Up to shrinking 𝒱\mathcal{V}, we can assume that

CΩΩC1ρLeb,Ω𝒱.C\|\Omega-\Omega_{*}\|_{C^{1}}\leq\rho_{\mathrm{Leb}},\qquad\forall\,\Omega\in\mathcal{V}.

Let now Ω𝒱\Omega\in\mathcal{V} be a Zoll form with associated bundle 𝔭Ω\mathfrak{p}_{\Omega}, and consider the set

Σ(Ω;𝔭1):={zΣ|{tΦtXΩ(z)}Λ(𝔭1)},\Sigma(\Omega;\mathfrak{p}_{1}):=\big{\{}z\in\Sigma\ \big{|}\ \{t\mapsto\Phi_{t}^{X_{\Omega}}(z)\}\in\Lambda(\mathfrak{p}_{1})\big{\}},

which is non-empty since we proved that 𝒳(Ω;𝔭1)\mathcal{X}(\Omega;\mathfrak{p}_{1}) is non-empty. This set is open, as the sets BiB_{i} are open. Furthermore, from item (i) and the inequality CΩΩC1ρLebC\|\Omega-\Omega_{*}\|_{C^{1}}\leq\rho_{\mathrm{Leb}}, it is also closed. Since Σ\Sigma is connected, we deduce Σ(Ω;𝔭1)=Σ\Sigma(\Omega;\mathfrak{p}_{1})=\Sigma as desired.

To prove the equality between the actions in (iii), we observe that, since 𝔭Ω\mathfrak{p}_{\Omega} is C1C^{1}-close to 𝔭1\mathfrak{p}_{1}, there exist a path {𝔮r}\{\mathfrak{q}_{r}\} with 𝔮0=𝔭1\mathfrak{q}_{0}=\mathfrak{p}_{1}, 𝔮1=𝔭Ω\mathfrak{q}_{1}=\mathfrak{p}_{\Omega} and some ziΣiz_{i}\in\Sigma_{i} such that all the loops ȷ𝔮r(zi)\jmath_{\mathfrak{q}_{r}}(z_{i}) are contained in Σi\Sigma_{i}, and, hence, build a short path. Therefore,

𝒜Ω(ȷ𝔭Ω(zi))=𝒜~Ω([{ȷ𝔭r(zi)}#{ȷ𝔮r(zi)}])=𝒜~Ω([{ȷ(𝔭#𝔮)r(zi)}])=𝒜(Ω),\mathcal{A}_{\Omega}(\jmath_{\mathfrak{p}_{\Omega}}(z_{i}))=\widetilde{\mathcal{A}}_{\Omega}\Big{(}\big{[}\{\jmath_{\mathfrak{p}_{r}}(z_{i})\}\#\{\jmath_{\mathfrak{q}_{r}}(z_{i})\}\big{]}\Big{)}=\widetilde{\mathcal{A}}_{\Omega}\Big{(}\big{[}\{\jmath_{(\mathfrak{p}\#\mathfrak{q})_{r}}(z_{i})\}\big{]}\Big{)}=\mathcal{A}(\Omega),

where {(𝔭#𝔮)r}={𝔭r}#{𝔮r}\{(\mathfrak{p}\#\mathfrak{q})_{r}\}=\{\mathfrak{p}_{r}\}\#\{\mathfrak{q}_{r}\}. As far as the existence of {𝔮r}\{\mathfrak{q}_{r}\} is concerned, the first statement in Lemma 4.7 yields a diffeomorphism Ψ:ΣΣ\Psi:\Sigma\to\Sigma, which is C1C^{1}-close to idΣ\mathrm{id}_{\Sigma} and satisfies Ψ𝔭Ω=𝔭1\Psi^{*}\mathfrak{p}_{\Omega}=\mathfrak{p}_{1}. As a consequence, there exists an isotopy {Ψr}\{\Psi_{r}\} of diffeomorphisms C1C^{1}-close to idΣ\mathrm{id}_{\Sigma} with Ψ1=Ψ\Psi_{1}=\Psi. The path 𝔮r:=(Ψr1)𝔭0\mathfrak{q}_{r}:=(\Psi_{r}^{-1})^{*}\mathfrak{p}_{0} has the desired properties. ∎

It seems appropriate at this point to recall Conjecture 1 formulated in the Introduction.

Conjecture 2.

If Ω𝒵C(Σ)\Omega_{*}\in{\mathcal{Z}}_{C}(\Sigma) is a Zoll odd-symplectic form, there is a Ck1C^{k-1}-neighbourhood 𝒰\mathcal{U} of Ω\Omega_{*} in 𝒮C(Σ)\mathcal{S}_{C}(\Sigma), for some k2k\geq 2, such that

P(𝒜min(Ω))𝔙𝔬𝔩(Ω)P(𝒜max(Ω)),Ω𝒰.P(\mathcal{A}_{\min}(\Omega))\leq{\mathfrak{Vol}}(\Omega)\leq P(\mathcal{A}_{\max}(\Omega)),\qquad\forall\,\Omega\in\mathcal{U}.

The equality holds in any of the two inequalities above if and only if Ω\Omega is Zoll.

The order of the three terms in the inequality stems from the fact that PP is monotone increasing in a neighbourhood of AA_{*} by (7.1). Theorem 6.14 and Proposition 7.4.(iii) imply that if Ω\Omega is Zoll, then both equalities hold.

Corollary 7.5.

If 𝒱𝒮C(Σ)\mathcal{V}\subset\mathcal{S}_{C}(\Sigma) is the C1C^{1}-neighbourhood of a Zoll odd-symplectic form Ω\Omega_{*} given by Theorem 7.4, then there holds

P(𝒜min(Ω))=𝔙𝔬𝔩(Ω)=P(𝒜max(Ω)),Ω𝒱𝒵C(Σ).P(\mathcal{A}_{\min}(\Omega))={\mathfrak{Vol}}(\Omega)=P(\mathcal{A}_{\max}(\Omega)),\qquad\forall\,\Omega\in\mathcal{V}\cap\mathcal{Z}_{C}(\Sigma).
Remark 7.6.

If the real Euler class vanishes, then the inequality in Conjecture 2 becomes

𝒜min(Ω)0𝒜max(Ω),Ω𝒰.\mathcal{A}_{\min}(\Omega)\leq 0\leq\mathcal{A}_{\max}(\Omega),\qquad\forall\,\Omega\in\mathcal{U}.

Indeed, in this case 𝔙𝔬𝔩0{\mathfrak{Vol}}\equiv 0, P(A)=[ω]n,M1]AP(A)=\langle[\mathrm{\omega}_{*}]^{n},M_{1}]\rangle A and [ω]n,M1]>0\langle[\mathrm{\omega}_{*}]^{n},M_{1}]\rangle>0 by (7.1).

Remark 7.7.

The conjecture is independent of the chosen reference pair (𝔭0,c0)C0(Σ)(\mathfrak{p}_{0},c_{0})\in\mathfrak{Z}_{C}^{0}(\Sigma). Indeed, if (𝔭0,c0)C0(Σ)(\mathfrak{p}_{0}^{\prime},c_{0}^{\prime})\in\mathfrak{Z}_{C}^{0}(\Sigma) is another pair with corresponding objects 𝔙𝔬𝔩{\mathfrak{Vol}}^{\prime}, PP^{\prime} and 𝒜Ω\mathcal{A}_{\Omega}^{\prime}, then

P𝒜Ω𝔙𝔬𝔩(Ω)=P𝒜Ω𝔙𝔬𝔩(Ω)P^{\prime}\circ\mathcal{A}_{\Omega}^{\prime}-{\mathfrak{Vol}}^{\prime}(\Omega)=P\circ\mathcal{A}_{\Omega}-{\mathfrak{Vol}}(\Omega)

by Corollary 6.19.

In the next subsection, we will see that PP can be written as the integral over M1M_{1} of a function 𝒫:M1×\mathcal{P}:M_{1}\times{\mathbb{R}}\to{\mathbb{R}} so that the monotonicity of PP corresponds to the monotonicity of 𝒫\mathcal{P} in the second variable for an interval \mathcal{I} containing AA_{*}. This will enable us to establish the conjecture in some simple cases. Before doing that, we end this subsection by observing that Conjecture 1 is also consistent with the pull-back operation.

Proposition 7.8.

Let Π:ΣΣ\Pi:\Sigma^{\text{\tiny$\vee$}}\to\Sigma be a bundle map with 𝔭0=Π𝔭0\mathfrak{p}_{0}^{\text{\tiny$\vee$}}=\Pi^{*}\mathfrak{p}_{0}, which is also a finite covering map. If Conjecture 1 holds for Ω=ΠΩ𝒮C(Σ)\Omega_{*}^{\text{\tiny$\vee$}}=\Pi^{*}\Omega_{*}\in\mathcal{S}_{C^{\text{\tiny$\vee$}}}(\Sigma^{\text{\tiny$\vee$}}), then it holds for Ω𝒮C(Σ)\Omega_{*}\in\mathcal{S}_{C}(\Sigma).

Proof.

Let us suppose that the conjecture is true for a neighbourhood 𝒰𝒱\mathcal{U}^{\text{\tiny$\vee$}}\subset\mathcal{V}^{\text{\tiny$\vee$}} around Ω\Omega_{*}^{\text{\tiny$\vee$}}, where 𝒱\mathcal{V}^{\text{\tiny$\vee$}} is given by Proposition 7.4. Thus, for all Ω𝒰\Omega^{\text{\tiny$\vee$}}\in\mathcal{U}^{\text{\tiny$\vee$}}, there holds

P(𝒜min(Ω))𝔙𝔬𝔩(Ω)P(𝒜max(Ω))P^{\text{\tiny$\vee$}}(\mathcal{A}_{\min}(\Omega^{\text{\tiny$\vee$}}))\leq{\mathfrak{Vol}}(\Omega^{\text{\tiny$\vee$}})\leq P^{\text{\tiny$\vee$}}(\mathcal{A}_{\max}(\Omega^{\text{\tiny$\vee$}})) (7.5)

and any of the two equalities holds if and only if Ω\Omega^{\text{\tiny$\vee$}} is Zoll. We claim that the local systolic inequality holds in the neighbourhood 𝒰:=(Π)1(𝒰)\mathcal{U}:=(\Pi^{*})^{-1}(\mathcal{U}^{\text{\tiny$\vee$}}) of Ω\Omega_{*}. Indeed, let us take an arbitrary Ω\Omega in 𝒰\mathcal{U}. This means that ΠΩ\Pi^{*}\Omega belongs to 𝒰\mathcal{U}^{\text{\tiny$\vee$}}. Applying Proposition 7.3, Proposition 6.17 and Proposition 2.7, it follows that

P(𝒜min(Ω))=P(𝒜min(ΠΩ))=1degπ0P(𝒜min(ΠΩ))1degΠ𝔙𝔬𝔩(ΠΩ)=𝔙𝔬𝔩(Ω),P(\mathcal{A}_{\min}(\Omega))=P(\mathcal{A}_{\min}(\Pi^{*}\Omega))=\frac{1}{\deg\pi_{0}}P^{\text{\tiny$\vee$}}(\mathcal{A}_{\min}(\Pi^{*}\Omega))\leq\frac{1}{\deg\Pi}{\mathfrak{Vol}}(\Pi^{*}\Omega)={\mathfrak{Vol}}(\Omega),

where we used degΠ=degπ0>0\deg\Pi=\deg\pi_{0}>0 and (7.5). An analogous inequality holds for 𝒜max(Ω)\mathcal{A}_{\max}(\Omega).

Let us assume now, for instance, that P(𝒜min(Ω))=𝔙𝔬𝔩(Ω)P(\mathcal{A}_{\min}(\Omega))={\mathfrak{Vol}}(\Omega). From the computation above, this implies P(𝒜min(ΠΩ))=𝔙𝔬𝔩(ΠΩ)P^{\text{\tiny$\vee$}}(\mathcal{A}_{\min}(\Pi^{*}\Omega))={\mathfrak{Vol}}(\Pi^{*}\Omega). Since the conjecture is true for 𝒰\mathcal{U}^{\text{\tiny$\vee$}}, we see that ΠΩ\Pi^{*}\Omega is Zoll and we call 𝔭\mathfrak{p}^{\text{\tiny$\vee$}} the associated bundle. All the fibres of 𝔭\mathfrak{p}^{\text{\tiny$\vee$}} lie in some Σj\Sigma^{\text{\tiny$\vee$}}_{j}, by Proposition 7.4. Since Π\Pi is injective on such sets, it follows that there exists a bundle 𝔭:ΣM\mathfrak{p}:\Sigma\to M such that 𝔭=Π𝔭\mathfrak{p}^{\text{\tiny$\vee$}}=\Pi^{*}\mathfrak{p}, which implies that Ω\Omega is a Zoll form with bundle 𝔭\mathfrak{p}. ∎

7.2 A second look at H-forms

For the rest of this section, we assume that 𝔭0=𝔭1=𝔭Ω\mathfrak{p}_{0}=\mathfrak{p}_{1}=\mathfrak{p}_{\Omega_{*}}, and take {𝔭r}\{\mathfrak{p}_{r}\} as the constant path. This does not cause any loss in generality thanks to Remark 7.7. We fix a free S1S^{1}-action 𝔲𝔘(𝔭0)\mathfrak{u}\in\mathfrak{U}(\mathfrak{p}_{0}) with generating vector field VV. Let us take η𝒦(𝔲)\eta\in\mathcal{K}(\mathfrak{u}) with curvature κΞ2(M0)\kappa\in\Xi^{2}(M_{0}). Namely, we have [κ]=e0[\kappa]=e_{0} and 𝔭0κ=dη\mathfrak{p}_{0}^{*}\kappa={\mathrm{d}}\eta. By Theorem 6.14, there is ω0Ξc02(M0)\mathrm{\omega}_{0}\in\Xi^{2}_{c_{0}}(M_{0}) such that

ω=Aκ+ω0.\mathrm{\omega}_{*}=A_{*}\kappa+\omega_{0}.

Moreover, we can exploit our freedom in choosing η\eta and c0c_{0} to get

ω=ω0,if e0=0,ω=Aκ,if C=0.\bullet\ \ \mathrm{\omega}_{*}=\mathrm{\omega}_{0},\quad\text{if $e_{0}=0$},\qquad\qquad\bullet\ \ \mathrm{\omega}_{*}=A_{*}\kappa,\quad\text{if $C=0$}.

Indeed, if e0=0e_{0}=0, we just take η\eta to be a flat connection, i.e. κ=0\kappa=0. If C=0C=0, we take c0=0c_{0}=0 so that A0A_{*}\neq 0 (as [ωn]0[\mathrm{\omega}_{*}^{n}]\neq 0) and pick η𝒦(𝔲)\eta\in\mathcal{K}(\mathfrak{u}) with the property that κ=1Aω\kappa=\tfrac{1}{A_{*}}\mathrm{\omega}_{*}, which is equivalent to ω0=0\mathrm{\omega}_{0}=0. We define the function

𝒬:M0×,𝒬(,A)ωn=(Aκ+ω0)n,A.\mathcal{Q}:M_{0}\times{\mathbb{R}}\to{\mathbb{R}},\qquad\mathcal{Q}(\,\cdot\,,A)\mathrm{\omega}^{n}_{*}=(A\kappa+\mathrm{\omega}_{0})^{n},\qquad\forall\,A\in{\mathbb{R}}.

The function 𝒫:M0×\mathcal{P}:M_{0}\times{\mathbb{R}}\to{\mathbb{R}} is determined by the properties

𝒫(q,0)=0,qM0,𝒫A(q,A)=𝒬(q,A),(q,A)M0×.\bullet\ \ \mathcal{P}(q,0)=0,\quad\forall\,q\in M_{0},\qquad\qquad\bullet\ \ \frac{\partial\mathcal{P}}{\partial A}(q,A)=\mathcal{Q}(q,A),\quad\forall\,(q,A)\in M_{0}\times{\mathbb{R}}.

There holds

Q(A)=M0𝒬(,A)ωn,P(A)=M0𝒫(,A)ωn.Q(A)=\int_{M_{0}}\mathcal{Q}(\,\cdot\,,A)\,\mathrm{\omega}^{n}_{*},\qquad P(A)=\int_{M_{0}}\mathcal{P}(\,\cdot\,,A)\mathrm{\omega}^{n}_{*}.

From the definition, 𝒬(,A)1\mathcal{Q}(\,\cdot\,,A_{*})\equiv 1 and we define \mathcal{I}\subset{\mathbb{R}} to be the maximal interval satisfying

A,minqM0𝒬(q,A)>0,A.\bullet\ \ A_{*}\in\mathcal{I},\qquad\qquad\bullet\ \ \min_{q\in M_{0}}\mathcal{Q}(q,A)>0,\quad\forall\,A\in\mathcal{I}.

Therefore, 𝒫\mathcal{P} is strictly increasing in the second coordinate on M0×M_{0}\times\mathcal{I}. This fact will be crucially used in Proposition 7.16.

Remark 7.9.

If C=0C=0, then 𝒬(,A)=(A/A)n\mathcal{Q}(\cdot,A)=(A/A_{*})^{n}, 𝒫(,A)=An+1(A/A)n+1\mathcal{P}(\cdot,A)=\tfrac{A_{*}}{n+1}(A/A_{*})^{n+1}, which implies =(0,+)\mathcal{I}=(0,+\infty) or =(,0)\mathcal{I}=(-\infty,0) depending on the sign of AA_{*}. If e0=0e_{0}=0, then 𝒬1\mathcal{Q}\equiv 1 and 𝒫(,A)A\mathcal{P}(\,\cdot\,,A)\equiv A, which implies =\mathcal{I}={\mathbb{R}}.

We now make use of the notion of H-form from Section 3.1. The H-forms we use below are associated with Ω0\Omega_{0}, α0=0\alpha_{0}=0 and σ0=η\sigma_{0}=\eta, the given S1S^{1}-connection. Namely, we deal with forms of the type

ΩHη=Ω0+d(Hη)ΞC2(Σ),H:Σ.\Omega_{H\eta}=\Omega_{0}+{\mathrm{d}}(H\eta)\in\Xi^{2}_{C}(\Sigma),\qquad H:\Sigma\to{\mathbb{R}}.
Remark 7.10.

Let us describe the fibres of the map HΩHηH\mapsto\Omega_{H\eta}. To this end, we compute

d(Hη)=dhHη+H𝔭0κ.{\mathrm{d}}(H\eta)={\mathrm{d}}^{h}H\wedge\eta+H\mathfrak{p}_{0}^{*}\kappa. (7.6)

Here dhH{\mathrm{d}}^{h}H is the horizontal part of dH{\mathrm{d}}H:

dhH:=dHdH(V)η.{\mathrm{d}}^{h}H:={\mathrm{d}}H-{\mathrm{d}}H(V)\eta.

Let us suppose now that d(Hη)=0{\mathrm{d}}(H\eta)=0. From (7.6), we deduce

dhH=0,H𝔭0κ=0.{\mathrm{d}}^{h}H=0,\qquad H\mathfrak{p}_{0}^{*}\kappa=0.

The first relation tells us that HH is invariant along curves tangent to kerη\ker\eta.

Thus, if e00e_{0}\neq 0, then HH is constant, as the holonomy in this case is the whole S1S^{1}. Evaluating the second relation at a point qM0q\in M_{0}, where κq0\kappa_{q}\neq 0, we conclude that H=0H=0.

On the other hand, if e0=0e_{0}=0, then we can take the quotient Πk:ΣΣk\Pi_{k}:\Sigma\to\Sigma_{k} by the holonomy of η\eta, which is a finite subgroup of S1S^{1} (see Lemma 4.6). The bundle ΣkM0\Sigma_{k}\to M_{0} is trivial and admits an angular function ϕk:ΣkS1\phi_{k}:\Sigma_{k}\to S^{1}. We define ϕ:ΣS1\phi:\Sigma\to S^{1} as ϕ:=ϕkΠk\phi:=\phi_{k}\circ\Pi_{k}. We conclude that for each HH with dhH=0{\mathrm{d}}^{h}H=0, there exists H¯:S1\bar{H}:S^{1}\to{\mathbb{R}} such that H=H¯ϕH=\bar{H}\circ\phi.

In our setting, the volume of an H-form can be computed in terms of the integration operator associated with the free S1S^{1}-action 𝔲\mathfrak{u}:

𝔲:C0(Σ)C0(M0),𝔲(K)(q):=(𝔭0)(Kη)(q)=S1K(ΦtV(z))dt,qM0,\mathfrak{u}_{*}:C^{0}(\Sigma)\to C^{0}(M_{0}),\qquad\mathfrak{u}_{*}(K)(q):=(\mathfrak{p}_{0})_{*}(K\eta)(q)=\int_{S^{1}}K\big{(}\Phi^{V}_{t}(z)\big{)}\,{\mathrm{d}}t,\quad\forall\,q\in M_{0},

where zz is any point in 𝔭01(q)\mathfrak{p}_{0}^{-1}(q) and (𝔭0)(\mathfrak{p}_{0})_{*} denotes the integration along the 𝔭0\mathfrak{p}_{0}-fibres.

Lemma 7.11.

For a function H:ΣH:\Sigma\to{\mathbb{R}}, there holds

Vol(Hη)=M0𝔲(𝒫(𝔭0,H))ωn.\mathrm{Vol}(H\eta)=\int_{M_{0}}\mathfrak{u}_{*}\big{(}\mathcal{P}(\mathfrak{p}_{0},H)\big{)}\,\omega_{*}^{n}.
Proof.

Recalling the definition of Vol\mathrm{Vol} from (2.1) and using Fubini’s Theorem, we compute

Vol(Hη)=01drΣHηΩrHηn\displaystyle\mathrm{Vol}(H\eta)=\int_{0}^{1}{\mathrm{d}}r\int_{\Sigma}H\eta\wedge\Omega_{rH\eta}^{n} =01drΣHη(𝔭0ω0+rH𝔭0κ)n\displaystyle=\int_{0}^{1}{\mathrm{d}}r\int_{\Sigma}H\eta\wedge\Big{(}\mathfrak{p}_{0}^{*}\mathrm{\omega}_{0}+rH\mathfrak{p}_{0}^{*}\kappa\Big{)}^{n}
=01drΣHη𝒬(𝔭0,rH)(𝔭0ω)n\displaystyle=\int_{0}^{1}{\mathrm{d}}r\int_{\Sigma}H\eta\wedge\mathcal{Q}(\mathfrak{p}_{0},rH)\big{(}\mathfrak{p}_{0}^{*}\mathrm{\omega}_{*}\big{)}^{n}
=Σ(01H𝒬(𝔭0,rH)dr)η(𝔭0ω)n\displaystyle=\int_{\Sigma}\Big{(}\int_{0}^{1}H\mathcal{Q}(\mathfrak{p}_{0},rH)\,{\mathrm{d}}r\Big{)}\eta\wedge\big{(}\mathfrak{p}_{0}^{*}\mathrm{\omega}_{*}\big{)}^{n}
=Σ𝒫(𝔭0,H)η(𝔭0ω)n\displaystyle=\int_{\Sigma}\mathcal{P}(\mathfrak{p}_{0},H)\eta\wedge\big{(}\mathfrak{p}_{0}^{*}\mathrm{\omega}_{*}\big{)}^{n}
=M0(𝔭0)(𝒫(𝔭0,H)η)ωn\displaystyle=\int_{M_{0}}(\mathfrak{p}_{0})_{*}\big{(}\mathcal{P}(\mathfrak{p}_{0},H)\eta\big{)}\wedge\mathrm{\omega}^{n}_{*}
=M0𝔲(𝒫(𝔭0,H))ωn,\displaystyle=\int_{M_{0}}\mathfrak{u}_{*}\big{(}\mathcal{P}(\mathfrak{p}_{0},H)\big{)}\,\mathrm{\omega}^{n}_{*},

where in the second equality, we have used that HηdHη=0H\eta\wedge{\mathrm{d}}H\wedge\eta=0. ∎

Remark 7.12.

When e0=0e_{0}=0, defining the Calabi invariant as the volume of HηH\eta

CALω(H):=Vol(Hη)=M0𝔲(H)ωn,\mathrm{CAL}_{\mathrm{\omega}_{*}}(H):=\mathrm{Vol}(H\eta)=\int_{M_{0}}\mathfrak{u}_{*}(H)\,\mathrm{\omega}^{n}_{*},

recovers the classical definition of the Calabi invariant for the trivial bundle M0×S1M0M_{0}\times S^{1}\to M_{0}. We say that HH is normalised with respect to ω\mathrm{\omega}_{*} if its Calabi invariant CALω(H)\mathrm{CAL}_{\mathrm{\omega}_{*}}(H) vanishes, which is equivalent to requiring that the one-form HηH\eta is normalised according to Definition 2.5. When H=hϕH=h\circ\phi, as in Remark 7.10, then HH is normalised if and only if the integral of hh is zero.

Lemma 7.13.

If a function H:ΣH:\Sigma\to{\mathbb{R}} takes values in \mathcal{I}, the H-form ΩHη\Omega_{H\eta} is odd-symplectic.

Proof.

To prove the statement, we show that ΩHη\Omega_{H\eta} is non-degenerate on the hyperplane distribution kerη\ker\eta. We compute preliminarily

Ω0+d(Hη)=Ω0+dhHη+H𝔭0κ=(𝔭0ω0+H𝔭0κ)+dhHη.\Omega_{0}+{\mathrm{d}}(H\eta)=\Omega_{0}+{\mathrm{d}}^{h}H\wedge\eta+H\mathfrak{p}_{0}^{*}\kappa=(\mathfrak{p}_{0}^{*}\mathrm{\omega}_{0}+H\mathfrak{p}_{0}^{*}\kappa)+{\mathrm{d}}^{h}H\wedge\eta.

Let zΣz\in\Sigma be arbitrary, and set q:=𝔭0(z)q:=\mathfrak{p}_{0}(z). We have an inverse z:TqM0(kerη)z\mathcal{L}_{z}:{\mathrm{T}}_{q}M_{0}\to(\ker\eta)_{z} for the projection d𝔭0:(kerη)zTqM0{\mathrm{d}}\mathfrak{p}_{0}:(\ker\eta)_{z}\to{\mathrm{T}}_{q}M_{0}, which we can use to pull back the restriction of forms on kerη\ker\eta to the tangent space of M0M_{0}:

z(Ω0+d(Hη))zn=(z(Ω0+d(Hη))z)n\displaystyle\mathcal{L}_{z}^{*}\Big{(}\Omega_{0}+{\mathrm{d}}(H\eta)\Big{)}_{\!z}^{n}=\Big{(}\mathcal{L}_{z}^{*}\big{(}\Omega_{0}+{\mathrm{d}}(H\eta)\big{)}_{z}\Big{)}^{n} =(z(𝔭0ω0+H𝔭0κ)z)n\displaystyle=\Big{(}\mathcal{L}_{z}^{*}\big{(}\mathfrak{p}_{0}^{*}\mathrm{\omega}_{0}+H\mathfrak{p}_{0}^{*}\kappa\big{)}_{z}\Big{)}^{n}
=(ω0+H(z)κ)qn\displaystyle=\Big{(}\mathrm{\omega}_{0}+H(z)\kappa\Big{)}_{\!q}^{n}
=𝒬(q,H(z))(ωn)q,\displaystyle=\mathcal{Q}(q,H(z))(\mathrm{\omega}_{*}^{n})_{q},

where we used z(dhHη)=0\mathcal{L}_{z}^{*}({\mathrm{d}}^{h}H\wedge\eta)=0. The last form is non-degenerate since 𝒬(q,H(z))>0\mathcal{Q}(q,H(z))>0. ∎

Combining this lemma with the stability property proved in Proposition 3.4, we arrive at the following corollary.

Corollary 7.14.

Suppose that 𝔭0:ΣM0\mathfrak{p}_{0}:\Sigma\to M_{0} is an oriented S1S^{1}-bundle with e0=0e_{0}=0 and choose η\eta to be a flat connection for 𝔭0\mathfrak{p}_{0}. Let Ω0=𝔭0ω0\Omega_{0}=\mathfrak{p}_{0}^{*}\mathrm{\omega}_{0} be given, where ω0\mathrm{\omega}_{0} is some symplectic form on M0M_{0}. For every sufficiently small C2C^{2}-neighbourhood 𝒰\mathcal{U} of Ω0\Omega_{0} in 𝒮C(Σ)\mathcal{S}_{C}(\Sigma), there exist a C2C^{2}-neighbourhood 𝒟\mathcal{D} of idΣ\mathrm{id}_{\Sigma} in Diff(Σ)\mathrm{Diff}(\Sigma) and a C2C^{2}-neighbourhood \mathcal{H} of the zero function in the space of normalised functions on Σ\Sigma with the following property:

Ω𝒰,Ψ𝒟,H,ΨΩ=Ω0+d(Hη).\forall\,\Omega\in\mathcal{U},\ \exists\,\Psi\in\mathcal{D},\ \exists\,H\in\mathcal{H},\qquad\Psi^{*}\Omega=\Omega_{0}+{\mathrm{d}}(H\eta).
Proof.

Let Ω𝒮C(Σ)\Omega\in\mathcal{S}_{C}(\Sigma) be C2C^{2}-close to Ω0\Omega_{0}. By standard elliptic arguments (see for instance [Nic07, Chapter 10]), we have Ω=Ω0+dα\Omega=\Omega_{0}+{\mathrm{d}}\alpha for a C2C^{2}-small normalised one-form α\alpha. We apply Proposition 3.4 with α0=0\alpha_{0}=0, σrη\sigma_{r}\equiv\eta and αr=rα\alpha_{r}=r\alpha so that Ωr=Ω0+rdα=Ω0+r(ΩΩ0)\Omega_{r}=\Omega_{0}+r{\mathrm{d}}\alpha=\Omega_{0}+r(\Omega-\Omega_{0}). Indeed, the hypothesis in Remark 3.5.(b) are met, as Ω\Omega is C0C^{0}-close to Ω0\Omega_{0} and σrη\sigma_{r}\equiv\eta is closed. Thus, we have the existence of a diffeomorphism Ψ:ΣΣ\Psi:\Sigma\to\Sigma isotopic to idΣ\mathrm{id}_{\Sigma} and of a normalised function H:ΣH:\Sigma\to{\mathbb{R}} such that ΨΩ=Ω0+d(Hη)\Psi^{*}\Omega=\Omega_{0}+{\mathrm{d}}(H\eta). More explicitly, from (3.2) and (3.5), we see that we have paths {Ψr}\{\Psi_{r}\} and {Hr}\{H_{r}\} with Ψ0=idΣ\Psi_{0}=\mathrm{id}_{\Sigma} and H0=0H_{0}=0 satisfying

Xrkerη,(ιXr(Ω0+r(ΩΩ0))+α)|kerη=0,H˙r=α(V)Ψr,X_{r}\in\ker\eta,\qquad\big{(}\iota_{X_{r}}\big{(}\Omega_{0}+r(\Omega-\Omega_{0})\big{)}+\alpha\big{)}\big{|}_{\ker\eta}=0,\qquad\dot{H}_{r}=\alpha(V)\circ\Psi_{r},

where XrX_{r} is the time-dependent vector field generating Ψr\Psi_{r} and VV is the vertical vector field of η\eta. From these relations, we see that Ψr\Psi_{r} is C2C^{2}-close to idΣ\mathrm{id}_{\Sigma} and HrH_{r} is C2C^{2}-small. ∎

7.3 The local systolic-diastolic inequality for quasi-autonomous H-forms

We introduce H-forms of special type for which we can prove the systolic-diastolic inequality.

Definition 7.15.

We say that an H-form ΩHη\Omega_{H\eta} is quasi-autonomous, if there are pairs qmin,qmaxM0q_{\min},q_{\max}\in M_{0} and Hmin,Hmax:ΣH_{\min},H_{\max}:\Sigma\to{\mathbb{R}} such that

(i)\displaystyle(i) ΩHminη=ΩHη=ΩHmaxη\displaystyle\quad\Omega_{H_{\min}\eta}=\Omega_{H\eta}=\Omega_{H_{\max}\eta}
(ii)\displaystyle(ii) minΣHmin=Hmin(z),z𝔭01(qmin),maxΣHmax=Hmax(z),z𝔭01(qmax).\displaystyle\quad\min_{\Sigma}H_{\min}=H_{\min}(z),\ \ \forall\,z\in\mathfrak{p}^{-1}_{0}(q_{\min}),\qquad\max_{\Sigma}H_{\max}=H_{\max}(z),\ \ \forall\,z\in\mathfrak{p}_{0}^{-1}(q_{\max}).

We say that H:ΣH:\Sigma\to{\mathbb{R}} is quasi-autonomous, if the form ΩHη\Omega_{H\eta} is quasi-autonomous. Thanks to Remark 7.10, we have Hmax=H=HmaxH_{\max}=H=H_{\max} if e00e_{0}\neq 0. We recover the standard definition of quasi-autonomous Hamiltonians [HZ11, p. 186] if 𝔭0\mathfrak{p}_{0} is trivial.

Proposition 7.16.

Let Ω𝒵C(Σ)\Omega_{*}\in{\mathcal{Z}}_{C}(\Sigma) be a Zoll odd-symplectic form with bundle 𝔭Ω=𝔭0\mathfrak{p}_{\Omega_{*}}=\mathfrak{p}_{0}. When e0=0e_{0}=0, the reference connection one-form η\eta for 𝔭0\mathfrak{p}_{0} is assumed to be flat . If H:ΣH:\Sigma\to{\mathbb{R}} is quasi-autonomous with values in \mathcal{I} and satisfies 𝒜min(ΩHη)\mathcal{A}_{\min}(\Omega_{H\eta})\in\mathcal{I}, 𝒜max(ΩHη)\mathcal{A}_{\max}(\Omega_{H\eta})\in\mathcal{I}, then

P(𝒜min(ΩHη))𝔙𝔬𝔩(ΩHη)P(𝒜max(ΩHη)).P(\mathcal{A}_{\min}(\Omega_{H\eta}))\leq{\mathfrak{Vol}}(\Omega_{H\eta})\leq P(\mathcal{A}_{\max}(\Omega_{H\eta})).

Moreover, let us consider the following three conditions:

  1. (a)

    HH is constant when e00e_{0}\neq 0, or factors through the map ϕ\phi of Remark 7.10 when e0=0e_{0}=0.

  2. (b)

    The form ΩHη\Omega_{H\eta} is Zoll.

  3. (c)

    Any of the two inequalities above is an equality.

We have (a)(b)(\mathrm{a})\Rightarrow(\mathrm{b}) and (a)(c)(\mathrm{a})\Leftrightarrow(\mathrm{c}). The implication (b)(c)(\mathrm{b})\Rightarrow(\mathrm{c}) holds if ΩHη\Omega_{H\eta} belongs to the neighbourhood 𝒱\mathcal{V} given in Proposition 7.4 or if

𝒜max(ΩHη)𝒜min(ΩHη)<inf{a>0|a=𝔞(C),[γr],for some[γr]π1(Λ𝔥(Σ))}.\mathcal{A}_{\max}(\Omega_{H\eta})-\mathcal{A}_{\min}(\Omega_{H\eta})\,<\,\inf\Big{\{}a>0\ \Big{|}\ a=\langle\mathfrak{a}(C),[\gamma_{r}]\rangle,\ \ \text{for some}\ \ [\gamma_{r}]\in\pi_{1}(\Lambda_{\mathfrak{h}}(\Sigma))\Big{\}}.
Proof.

From (7.6), we see that the oriented 𝔭0\mathfrak{p}_{0}-fibre over qminq_{\min} is a closed characteristic for ΩHη\Omega_{H\eta} with action minHmin\min H_{\min}. Therefore,

𝒜min(ΩHη)minHminHmin.\mathcal{A}_{\min}(\Omega_{H\eta})\leq\min H_{\min}\leq H_{\min}. (7.7)

Since 𝒫\mathcal{P} is increasing in the second coordinate AA when AA\in\mathcal{I}, we have by Lemma 7.11

P(𝒜min(ΩHη))P(minHmin)=M0𝔲(𝒫(𝔭0,minHmin))ωnM0𝔲(𝒫(𝔭0,Hmin))ωn=𝔙𝔬𝔩(ΩHη).\begin{split}P(\mathcal{A}_{\min}(\Omega_{H\eta}))\leq P(\min H_{\min})&=\int_{M_{0}}\mathfrak{u}_{*}\big{(}\mathcal{P}(\mathfrak{p}_{0},\min H_{\min})\big{)}\,\mathrm{\omega}_{*}^{n}\\ &\leq\int_{M_{0}}\mathfrak{u}_{*}\big{(}\mathcal{P}(\mathfrak{p}_{0},H_{\min})\big{)}\,\mathrm{\omega}_{*}^{n}\\[2.15277pt] &={\mathfrak{Vol}}(\Omega_{H\eta}).\end{split} (7.8)

In the case of e0=0e_{0}=0, this is true since =\mathcal{I}={\mathbb{R}} (see Remark 7.9) and we can assume that HminH_{\min} is normalised up to adding a constant. The inequality for 𝒜max\mathcal{A}_{\max} is obtained similarly.

Next, we show the implications between (a), (b), and (c). Assuming (a), we have

ΩHη={𝔭0(ω0+Hκ),ife00𝔭0ω,ife0=0,\Omega_{H\eta}=\begin{cases}\mathfrak{p}^{*}_{0}(\mathrm{\omega}_{0}+H\kappa),&\text{if}\ e_{0}\neq 0\\[2.15277pt] \mathfrak{p}^{*}_{0}\mathrm{\omega},&\text{if}\ e_{0}=0,\end{cases}

which is Zoll since the value of HH is in \mathcal{I}. This shows (a)(b)(\mathrm{a})\Rightarrow(\mathrm{b}). Moreover, if e00e_{0}\neq 0, then H=Hmin=HmaxH=H_{\min}=H_{\max} is constant. Therefore, all the inequalities in (7.7) and (7.8) are actually equalities, and (c) follows. If e0=0e_{0}=0, then H=H¯ϕH=\bar{H}\circ\phi and the normalised functions HminH_{\min} and HmaxH_{\max} are both equal to zero. By the same reason as before, (a) implies (c).

We now show that (c)(\mathrm{c}) implies (a)(\mathrm{a}). Assume for instance, that P(𝒜min(ΩHη))=𝔙𝔬𝔩(ΩHη)P(\mathcal{A}_{\min}(\Omega_{H\eta}))={\mathfrak{Vol}}(\Omega_{H\eta}). From inequality (7.8) and the fact that A𝒫(,A)A\mapsto\mathcal{P}(\,\cdot\,,A) is strictly increasing for AA\in\mathcal{I}, we have that Hmin(z)=minHminH_{\min}(z)=\min H_{\min}, for all zΣz\in\Sigma. Therefore, HminH_{\min} is constant. If e00e_{0}\neq 0, then H=HminH=H_{\min} and the conclusion follows. If e0=0e_{0}=0, then H=Hmin+(HHmin)H=H_{\min}+(H-H_{\min}), and therefore, HH factors through the map ϕ\phi, as HminH_{\min} is constant and HHminH-H_{\min} factors through ϕ\phi by Remark 7.10, since ΩHη=ΩHminη\Omega_{H\eta}=\Omega_{H_{\min}\eta}.

Finally, let us assume (b), namely that ΩHη\Omega_{H\eta} is Zoll. If ΩHη𝒱\Omega_{H\eta}\in\mathcal{V}, we know from Proposition 7.4, that 𝒳(ΩHη)Λ(𝔭0)\mathcal{X}(\Omega_{H\eta})\subset\Lambda(\mathfrak{p}_{0}). Therefore, 𝒜min(ΩHη)=𝒜max(ΩHη)\mathcal{A}_{\min}(\Omega_{H\eta})=\mathcal{A}_{\max}(\Omega_{H\eta}) due to Proposition 7.4.(iii) and (c) follows. On the other hand, let

ϵ:=𝒜max(ΩHη)𝒜min(ΩHη)0\epsilon:=\mathcal{A}_{\max}(\Omega_{H\eta})-\mathcal{A}_{\min}(\Omega_{H\eta})\geq 0

and suppose that ϵ\epsilon is strictly less than the positive generator of 𝔞(C),π1(Λ𝔥(Σ))\langle\mathfrak{a}(C),\pi_{1}(\Lambda_{\mathfrak{h}}(\Sigma))\rangle. Let 𝔭\mathfrak{p} be the bundle associated with the Zoll form ΩHη\Omega_{H\eta}, and let z0,z1Σz_{0},z_{1}\in\Sigma be two points such that ȷ𝔭(z0),ȷ𝔭(z1)𝒳(ΩHη;𝔭0)\jmath_{\mathfrak{p}}(z_{0}),\jmath_{\mathfrak{p}}(z_{1})\in\mathcal{X}(\Omega_{H\eta};\mathfrak{p}_{0}) and

𝒜ΩHη(ȷ𝔭(z0))=𝒜min(ΩHη),𝒜ΩHη(ȷ𝔭(z1))=𝒜max(ΩHη).\mathcal{A}_{\Omega_{H\eta}}(\jmath_{\mathfrak{p}}(z_{0}))=\mathcal{A}_{\min}(\Omega_{H\eta}),\qquad\mathcal{A}_{\Omega_{H\eta}}(\jmath_{\mathfrak{p}}(z_{1}))=\mathcal{A}_{\max}(\Omega_{H\eta}).

If szss\mapsto z_{s} is any path connecting z0z_{0} to z1z_{1}, we obtain an element [γr]π1(Λ𝔥(Σ))[\gamma_{r}]\in\pi_{1}(\Lambda_{\mathfrak{h}}(\Sigma)) by concatenating a short homotopy for ȷ𝔭(z0)\jmath_{\mathfrak{p}}(z_{0}), the path sȷ𝔭(zs)s\mapsto\jmath_{\mathfrak{p}}(z_{s}), and a reverse short homotopy for ȷ𝔭(z1)\jmath_{\mathfrak{p}}(z_{1}). This loop has the property that

ϵ=𝔞(C),[γr],\epsilon=\langle\mathfrak{a}(C),[\gamma_{r}]\rangle,

which implies that ϵ=0\epsilon=0 and condition (c) follows. ∎

Remark 7.17.

By Proposition 7.4, a quasi-autonomous function that is C2C^{2}-close to the constant A=𝒜(Ω)A_{*}=\mathcal{A}(\Omega_{*}) satisfies the hypotheses of Proposition 7.16. Moreover for such a function, conditions (a), (b), and (c) above are equivalent since the corresponding odd-symplectic form belongs to 𝒱\mathcal{V}.

Combining the result we have just proven with the theory of generating functions on closed symplectic manifolds, we can give a proof of Theorem 1.20, namely of the local systolic-diastolic inequality in the case of e0=0e_{0}=0.

Proof of Theorem 1.20

Let Ω\Omega_{*} be a Zoll-odd symplectic form such that the associated S1S^{1}-bundle 𝔭Ω\mathfrak{p}_{\Omega_{*}} has vanishing real Euler class. Due to Remark 7.7 we may assume that 𝔭Ω=𝔭0:ΣM0\mathfrak{p}_{\Omega_{*}}=\mathfrak{p}_{0}:\Sigma\to M_{0} and e0=0e_{0}=0. Let ω\mathrm{\omega}_{*} be the symplectic form on M0M_{0} such that 𝔭0ω=Ω\mathfrak{p}_{0}^{*}\mathrm{\omega}_{*}=\Omega_{*}. By Lemma 4.6, there exists a trivial oriented S1S^{1}-bundle 𝔭0:ΣM0\mathfrak{p}_{0}^{\text{\tiny$\vee$}}:\Sigma^{\text{\tiny$\vee$}}\to M_{0}^{\text{\tiny$\vee$}} and a bundle map Π:ΣΣ\Pi:\Sigma^{\text{\tiny$\vee$}}\to\Sigma such that 𝔭0=Π𝔭0\mathfrak{p}_{0}^{\text{\tiny$\vee$}}=\Pi^{*}\mathfrak{p}_{0}. Therefore, thanks to Proposition 7.8, it is enough to prove Theorem 1.20 for trivial bundles. Let 𝔭0:M0×S1M0\mathfrak{p}_{0}:M_{0}\times S^{1}\to M_{0} with angular function ϕ\phi. Assume that Ω\Omega is C2C^{2}-close to Ω\Omega_{*}. Due to Corollary 7.14, there is a diffeomorphism Ψ:ΣΣ\Psi:\Sigma\to\Sigma such that Ψ=Ψ1\Psi=\Psi_{1}, where {Ψr}\{\Psi_{r}\} is a C1C^{1}-small isotopy with Ψ0=idM0×S1\Psi_{0}=\mathrm{id}_{M_{0}\times S^{1}}, and

ΨΩ=ΩHdϕ\Psi^{*}\Omega=\Omega_{H{\mathrm{d}}\phi}

for a C2C^{2}-small normalised Hamiltonian H:M0×S1H:M_{0}\times S^{1}\to{\mathbb{R}}. Thus, 𝔙𝔬𝔩(Ω)=0=𝔙𝔬𝔩(ΩHdϕ){\mathfrak{Vol}}(\Omega)=0={\mathfrak{Vol}}(\Omega_{H{\mathrm{d}}\phi}), namely CALω(H)=0\mathrm{CAL}_{\mathrm{\omega}_{*}}(H)=0. Moreover, Ω\Omega and ΩHdϕ\Omega_{H{\mathrm{d}}\phi} have the same systole and diastole due to Proposition 6.10, as [Ψr][\Psi_{r}]^{*} preserves the set of short homotopies. The closed characteristics in 𝒳(ΩHdϕ;𝔭0)\mathcal{X}(\Omega_{H{\mathrm{d}}\phi};\mathfrak{p}_{0}) are curves γΛ(𝔭0)\gamma\in\Lambda(\mathfrak{p}_{0}) of the type γ(t)=(q(t),t)\gamma(t)=(q(t),t), for some loop q:S1Mq:S^{1}\to M with small capping disc q^:D2M\widehat{q}:D^{2}\to M. Thanks to (6.8), the action of these curves recovers the classical Hamiltonian action

𝒜Hdϕ(γ)=D2q^ω+S1H(q(t),t)dt\mathcal{A}_{H{\mathrm{d}}\phi}(\gamma)=\int_{D^{2}}\widehat{q}^{\hskip 2.0pt*}\mathrm{\omega}_{*}+\int_{S^{1}}H(q(t),t){\mathrm{d}}t

Let {φt}t[0,1]\{\varphi_{t}\}_{t\in[0,1]} be the Hamiltonian isotopy on (M0,ω)(M_{0},\mathrm{\omega}_{*}) up to time one generated by HH. The maps φt\varphi_{t} are C1C^{1}-close to the identity, as HH is C2C^{2}-small. From the theory of generating functions on arbitrary symplectic manifolds (see [LM95, Proposition 5.11] and [MS98, Proposition 9.31]), {φt}t[0,1]\{\varphi_{t}\}_{t\in[0,1]} is homotopic with fixed endpoints to a C1C^{1}-small Hamiltonian isotopy {φt}t[0,1]\{\varphi^{\prime}_{t}\}_{t\in[0,1]} generated by a quasi-autonomous normalised Hamiltonian H:M0×S1H^{\prime}:M_{0}\times S^{1}\to{\mathbb{R}}, namely 𝔙𝔬𝔩(ΩHdϕ)=0=CALω(H){\mathfrak{Vol}}(\Omega_{H^{\prime}{\mathrm{d}}\phi})=0=\mathrm{CAL}_{\mathrm{\omega}_{*}}(H^{\prime}). Adapting an argument in [Sch00], we have an action-preserving bijection

𝒳(ΩHdϕ;𝔭0)Fixφ1=Fixφ1𝒳(ΩHdϕ;𝔭0).\mathcal{X}(\Omega_{H{\mathrm{d}}\phi};\mathfrak{p}_{0})\cong\mathrm{Fix\,}\varphi_{1}=\mathrm{Fix\,}\varphi_{1}^{\prime}\cong\mathcal{X}(\Omega_{H^{\prime}{\mathrm{d}}\phi};\mathfrak{p}_{0}).

Therefore, ΩHdϕ\Omega_{H{\mathrm{d}}\phi} and ΩHdϕ\Omega_{H^{\prime}{\mathrm{d}}\phi} have the same systole and diastole. The local systolic-diastolic inequality holds for ΩHdϕ\Omega_{H^{\prime}{\mathrm{d}}\phi} due to Proposition 7.16, and hence also for ΩHdϕ\Omega_{H{\mathrm{d}}\phi} and Ω\Omega.∎

References

  • [ABHS17] A. Abbondandolo, B. Bramham, U.L. Hryniewicz, and P.A.S. Salomão, Contact forms with large systolic ratio in dimension three, to appear in Ann. Scuola Norm. Sup. Pisa Cl. Sci., arXiv:1709.01621, 2017.
  • [ABHS18]   , Sharp systolic inequalities for Reeb flows on the three-sphere, Invent. Math. 211 (2018), no. 2, 687–778.
  • [ÁPB14] J. C. Álvarez Paiva and F. Balacheff, Contact geometry and isosystolic inequalities, Geom. Funct. Anal. 24 (2014), no. 2, 648–669.
  • [BK19a] G. Benedetti and J. Kang, A local contact systolic inequality in dimension three, to appear in JEMS, arXiv:1902.01249, 2019.
  • [BK19b]   , On a systolic inequality for closed magnetic geodesics on surfaces, arXiv:1902.01262, 2019.
  • [Bot80] M. Bottkol, Bifurcation of periodic orbits on manifolds and Hamiltonian systems, J. Differential Equations 37 (1980), no. 1, 12–22.
  • [BR94] A. Banyaga and P. Rukimbira, Weak stability of almost regular contact foliations, J. Geom. 50 (1994), no. 1-2, 16–27.
  • [BW58] W. M. Boothby and H. C. Wang, On contact manifolds, Ann. of Math. (2) 68 (1958), 721–734.
  • [Che77] S. S. Chern, Circle bundles, 114–131. Lecture Notes in Math., Vol. 597.
  • [CM05] K. Cieliebak and K. Mohnke, Compactness for punctured holomorphic curves, J. Symplectic Geom. 3 (2005), no. 4, 589–654, Conference on Symplectic Topology.
  • [CS74] S. S. Chern and J. Simons, Characteristic forms and geometric invariants, Ann. of Math. (2) 99 (1974), 48–69.
  • [Gei08] H. Geiges, An introduction to contact topology, Cambridge Studies in Advanced Mathematics, vol. 109, Cambridge University Press, Cambridge, 2008.
  • [Gin87] V. L. Ginzburg, New generalizations of Poincaré’s geometric theorem, Funktsional. Anal. i Prilozhen. 21 (1987), no. 2, 16–22, 96.
  • [Hat02] A. Hatcher, Algebraic topology, Cambridge University Press, Cambridge, 2002.
  • [Hir94] M. W. Hirsch, Differential topology, Graduate Texts in Mathematics, vol. 33, Springer-Verlag, New York, 1994, Corrected reprint of the 1976 original.
  • [Hut16] M. Hutchings, Mean action and the Calabi invariant, J. Mod. Dyn. 10 (2016), 511–539.
  • [HZ11] H. Hofer and E. Zehnder, Symplectic invariants and Hamiltonian dynamics, Modern Birkhäuser Classics, Birkhäuser Verlag, Basel, 2011, Reprint of the 1994 edition.
  • [Ker99] E. Kerman, Periodic orbits of Hamiltonian flows near symplectic critical submanifolds, Internat. Math. Res. Notices (1999), no. 17, 953–969.
  • [KN96] S. Kobayashi and K. Nomizu, Foundations of differential geometry. Vol. I, Wiley Classics Library, John Wiley & Sons, Inc., New York, 1996, Reprint of the 1963 original, A Wiley-Interscience Publication.
  • [LM95] F. Lalonde and D. McDuff, Hofer’s LL^{\infty}-geometry: energy and stability of Hamiltonian flows. II, Invent. Math. 122 (1995), no. 1, 1–33, 35–69.
  • [MS98] D. McDuff and D. Salamon, Introduction to symplectic topology, second ed., Oxford Mathematical Monographs, The Clarendon Press, Oxford University Press, New York, 1998.
  • [Nic07] L. I. Nicolaescu, Lectures on the geometry of manifolds, second ed., World Scientific Publishing Co. Pte. Ltd., Hackensack, NJ, 2007.
  • [Pat09] G. P. Paternain, Helicity and the Mañé critical value, Algebr. Geom. Topol. 9 (2009), no. 3, 1413–1422.
  • [Sch00] M. Schwarz, On the action spectrum for closed symplectically aspherical manifolds, Pacific J. Math. 193 (2000), no. 2, 419–461.
  • [Tho76] C. B. Thomas, Almost regular contact manifolds, J. Differential Geometry 11 (1976), no. 4, 521–533.
  • [Voi07] C. Voisin, Hodge theory and complex algebraic geometry. I, english ed., Cambridge Studies in Advanced Mathematics, vol. 76, Cambridge University Press, Cambridge, 2007, Translated from the French by Leila Schneps.
  • [Wal67] F. Waldhausen, Eine Klasse von 33-dimensionalen Mannigfaltigkeiten. I, II, Invent. Math. 3 (1967), 308–333; ibid. 4 (1967), 87–117.
  • [Wei74] A. Weinstein, Application des opérateurs intégraux de Fourier aux spectres des variétés riemanniennes, C. R. Acad. Sci. Paris Sér. A 279 (1974), 229–230.

Ruprecht-Karls-Universität Heidelberg, Mathematisches Institut, Im Neuenheimer Feld 205, 69120 Heidelberg, Germany

Seoul National University, Department of Mathematical Sciences, Research Institute in Mathematics, Gwanak-Gu, Seoul 08826, South Korea

E-mail address: jungsoo.kang@snu.ac.kr