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

A universal extension of helicity to topological flows

Oliver Edtmair and Sobhan Seyfaddini
(August 16, 2025)
Abstract

Helicity is a fundamental conserved quantity in physical systems governed by vector fields whose evolution is described by volume-preserving transformations on a three-manifold. Notable examples include inviscid, incompressible fluid flows, modeled by the three-dimensional Euler equations, and conducting plasmas, described by the magnetohydrodynamics (MHD) equations.

A key property of helicity is its invariance under volume-preserving diffeomorphisms. In an influential article from 1973, Arnold—having provided an ergodic interpretation of helicity as the “asymptotic Hopf invariant”—posed the question of whether this invariance persists under volume-preserving homeomorphisms. More generally, he asked whether helicity can be extended to topological volume-preserving flows. We answer both questions affirmatively for flows without rest points.

Our approach reformulates Arnold’s question in the framework of what we call C0C^{0} Hamiltonian structures. This perspective enables us to leverage recent developments in C0C^{0} symplectic geometry, particularly results concerning the algebraic structure of the group of area-preserving homeomorphisms.

1 Introduction

Helicity is a fundamental conserved quantity in physical systems where divergence-free vector fields evolve under volume-preserving transformations on a three-manifold. Notable examples include vorticity fields in inviscid, incompressible fluids, governed by the three-dimensional Euler equations, and magnetic fields in plasmas, described by the magnetohydrodynamics (MHD) equations. Although the concept of helicity was introduced by Woltjer in the study of magnetohydrodynamics [56], the term “helicity” was coined by Moffatt [35] in his work on hydrodynamics. Building on Moreau [41], Moffatt derived the conservation of helicity from the Helmholtz–Kelvin law of vorticity transport [26, 27, 52]. Owing to its physical significance, helicity remains an active topic of research in experimental physics; see, for example, [47]. For historical overviews, we refer to [36, 38, 37].

This article addresses questions raised by Arnold [1] concerning topological properties of helicity.

1.1 Helicity and Arnold’s questions

Let (Y3,μ)(Y^{3},\mu) be a closed smooth 33-manifold equipped with a volume form μ\mu, and let XX be a smooth vector field on YY that preserves μ\mu. That is, the flow φXt\varphi^{t}_{X} of XX satisfies (φXt)μ=μ(\varphi^{t}_{X})^{*}\mu=\mu for all tt. This condition is equivalent to the 22-form

ωιXμ\omega\coloneqq\iota_{X}\mu

being closed. The vector field XX is said to be exact if ω\omega is exact.

The helicity of an exact volume-preserving vector field XX is defined as

(X)Yαdα,\mathcal{H}(X)\coloneqq\int_{Y}\alpha\wedge d\alpha,

where α\alpha is any primitive 11-form of ω\omega and YY is oriented via the volume form μ\mu. This integral turns out to be independent of the choice of primitive 11-form α\alpha.

A simple but fundamental property of helicity is that it is preserved under volume- and orientation-preserving diffeomorphisms; that is,

(fX)=(X)for every fDiff+(Y,μ).\mathcal{H}(f_{*}X)=\mathcal{H}(X)\qquad\text{for every $f\in\operatorname{Diff}^{+}(Y,\mu)$.}

In fact, any other functional on the space of exact volume-preserving vector fields that is invariant under the action of Diff+(Y,μ)\operatorname{Diff}^{+}(Y,\mu) and satisfies certain natural regularity conditions must be a function of helicity [29, 16].

Two volume-preserving smooth vector fields X1X_{1} and X2X_{2} on (Y,μ)(Y,\mu) are said to be topologically conjugate if their flows are conjugate via a volume- and orientation-preserving homeomorphism, i.e. if there exists fHomeo+(Y,μ)f\in\operatorname{Homeo}^{+}(Y,\mu) such that

fφX1tf1=φX2tfor all t.f\circ\varphi_{X_{1}}^{t}\circ f^{-1}=\varphi_{X_{2}}^{t}\qquad\text{for all $t\in\mathbb{R}$.}

For examples of volume-preserving smooth vector fields that are topologically conjugate but not smoothly (or even C1C^{1}) conjugate, see [42, §10 & 11].

In his 1973 article [2], Arnold, having derived an ergodic interpretation of helicity, posed the following questions concerning the topological invariance of helicity.

Question 1.1.

Let X1X_{1} and X2X_{2} be two exact volume-preserving smooth vector fields which are topologically conjugate. Is it true that (X1)=(X2)\mathcal{H}(X_{1})=\mathcal{H}(X_{2})?

Question 1.2.

Does helicity admit an extension to topological volume-preserving flows? Here, a topological flow refers to a continuous flow and is not necessarily generated by a vector field.

In this paper we address the above questions for fixed-point-free volume-preserving flows, which is the setting considered by Ghys in [22, §5.4].

We provide a preliminary version of our results here, postponing the statement of our main result, Theorem 1.5, to Section 1.5, as it requires some preparatory material.

Theorem 1.3.

Two nowhere vanishing, exact, volume-preserving, smooth vector fields which are topologically conjugate have the same helicity.

Moreover, helicity admits an extension to fixed-point-free, exact, volume-preserving, topological flows whose flow lines have zero measure. This extension is invariant under conjugation by volume- and orientation-preserving homeomorphisms and compatible with the Calabi invariant in the sense of Eq. (1.1), and it is uniquely determined by these properties.

We briefly comment on the assumptions in the theorem. The notion of exactness for topological flows generalizes its smooth counterpart and is essential for defining helicity, even in the smooth setting; see Definitions 7.2 & 8.20. The condition that flow lines have zero measure is quite natural; it rules out certain pathological situations (see Example 8.12). The same assumption appears also in other works [20, 10]. The significance of the fixed-point-free condition is that it yields a C0C^{0} Hamiltonian structure, enabling us to build on recent advances in C0C^{0} symplectic topology and the understanding of the algebraic structure of groups of area-preserving homeomorphisms [13, 11, 14].

Compatibility with Calabi. We briefly explain this here, leaving further details to Sections 1.4 & 1.5. Let (Σ,ωΣ)(\Sigma,\omega_{\Sigma}) denote an open surface equipped with an area form ωΣ\omega_{\Sigma}. We denote its group of Hamiltonian diffeomorphisms by Ham(Σ)\operatorname{Ham}(\Sigma). It was proven recently [14, 34] that the Calabi homomorphism CalΣ:Ham(Σ)\mathrm{Cal}_{\Sigma}:\operatorname{Ham}(\Sigma)\rightarrow{\mathbb{R}} admits infinitely many extensions to the group of Hamiltonian homeomorphisms Ham¯(Σ)\overline{\operatorname{Ham}}(\Sigma). Pick one such extension

Cal¯Σ:Ham¯(Σ).\overline{\mathrm{Cal}}_{\Sigma}:\overline{\operatorname{Ham}}(\Sigma)\rightarrow{\mathbb{R}}.

The Calabi extensions for different surfaces Σ\Sigma can be picked such that they satisfy the naturality property (1.9) below; we prove this fact, which is of independent interest and also crucial for our arguments, in Theorem 1.4.

Now, fix a topological volume-preserving flow ψt\psi^{t} on (Y,μ)(Y,\mu) and suppose that we have a topological volume-preserving embedding

α:((0,1)×Σ),dtωΣ)(Y,μ)\alpha:((0,1)\times\Sigma),dt\wedge\omega_{\Sigma})\hookrightarrow(Y,\mu)

which intertwines the flow on (0,1)×Σ(0,1)\times\Sigma generated by the vector field t\partial_{t} and the flow ψt\psi^{t}. Consider a C0C^{0} Hamiltonian isotopy φtHam¯(Σ)\varphi^{t}\in\overline{\operatorname{Ham}}(\Sigma) and note that its suspension to (0,1)×Σ(0,1)\times\Sigma is volume preserving. We refer to the tuple 𝒫:=(Σ,ωΣ,α,φt)\mathcal{P}:=(\Sigma,\omega_{\Sigma},\alpha,\varphi^{t}) as a plug. Given a plug 𝒫\mathcal{P}, one can define a new volume-preserving flow ψt#𝒫\psi^{t}\#\mathcal{P} on YY by replacing the flow ψt\psi^{t} inside im(α)\operatorname{im}(\alpha) with the suspension of φt\varphi^{t}. The compatibility condition between our helicity extension ¯\overline{\mathcal{H}} and the Calabi extension Cal¯Σ\overline{\operatorname{Cal}}_{\Sigma} is given by

¯(ψt#𝒫)=¯(ψt)+Cal¯Σ(φ1).\overline{\mathcal{H}}(\psi^{t}\#\mathcal{P})=\overline{\mathcal{H}}(\psi^{t})+\overline{\mathrm{Cal}}_{\Sigma}(\varphi^{1}). (1.1)

1.2 Context and Historical Background

Our story begins in 1973, when Arnold [1] interpreted the helicity of a vector field as the average asymptotic linking number of flow lines, which motivated Questions 1.1 & 1.2; see Sections 2-4 of the English translation [2] and Problem 1973-23 in [3]. Two trajectories, beginning at two randomly chosen points in space, are followed for a long time and then closed into loops using a well-chosen system of short geodesic arcs.111The existence of the system of geodesic arcs is a subtle point which was proven rigorously in [54]. The linking number of these loops—averaged over time and all pairs of initial points—converges to the helicity as the time tends to infinity. Since linking numbers are preserved under homeomorphisms, at first glance this seems to imply topological invariance of helicity. However, as remarked by Ghys in [21], “… one should be cautious that a homeomorphism might entangle the small geodesic arcs that were used to close the trajectories.”

Since their formulation, Questions 1.1 and 1.2 have reappeared in various works [4, 20, 21, 51, 42], including Arnold and Khesin’s textbook [4, III, Problem 4.8], Ghys’ plenary ICM address [21, Section 1.4], and Tao’s blog [51]. The case of flows without rest points is emphasized again in [22, Sec. 5.4]. One reason for this ongoing interest is that real-world flows often lack smoothness. Therefore, a positive answer to these questions would highlight the significance of helicity as a meaningful invariant, even in low-regularity settings. As Tao notes [51], “This would be of interest in fluid equations, as it would suggest that helicity remains invariant even after the development of singularities in the flow.”

We should mention that Arnold’s influential article has prompted substantial further work in various directions; see, for example, [19, 9, 20, 10, 46, 28, 32, 42, 30]. Among these, Gambaudo & Ghys [20] and Müller & Spaeth [42] contain results towards Questions 1.1 & 1.2. Both cases in [20, 42] involve exact volume-preserving flows without fixed points, which are covered by Theorem 1.3. The Gambaudo–Ghys approach establishes the topological invariance of helicity for certain suspension flows by relating it to the Calabi invariant—a connection that is also central to our work.

The extension of helicity to low-regularity settings is the subject of ongoing research in fluid dynamics: In [23], Giri, Kwon and Novack extend helicity to a class of weak solutions of the Euler equations with regularity H12εH^{\frac{1}{2}-\varepsilon}. Using convex integration, they further show that helicity need not be conserved for such solutions. This does not contradict our result, since the vector fields they construct are too irregular to generate a flow. It would be interesting to compare their extension of helicity with ours on the overlap of their respective domains.

1.3 Hamiltonian structures

Smooth Hamiltonian Structures. Let YY be an oriented smooth 33-manifold. A Hamiltonian structure on YY is a closed and maximally nondegenerate 22-form ω\omega on YY. The prototypical example is that of the standard Hamiltonian structure on 3{\mathbb{R}}^{3}, which is given by the 22-form ωstd:=dxdy\omega_{\operatorname{std}}:=dx\wedge dy, where 3=×2{\mathbb{R}}^{3}={\mathbb{R}}\times{\mathbb{R}}^{2} is equipped with the coordinates (t,x,y)(t,x,y). Every smooth Hamiltonian structure is locally diffeomorphic to ωstd\omega_{\operatorname{std}}; see Lemma 4.3.

Maximal non-degeneracy of ω\omega means that kerω\ker\omega defines a line field on YY, called the characteristic line field, which integrates to a 11-dimensional foliation on YY, called the characteristic foliation. The Hamiltonian structure ω\omega naturally equips this foliation with a coorientation and a transverse measure locally diffeomorphic to the standard 22-dimensional Lebesgue measure. Conversely, every such foliation uniquely determines a Hamiltonian structure.

The relevance of Hamiltonian structures to our discussion is as follows: Equip YY with a volume form μ\mu. Then, a vector field XX is nowhere vanishing and volume-preserving if and only if ωιXμ\omega\coloneqq\iota_{X}\mu is a Hamiltonian structure. In this case, the flow φXt\varphi_{X}^{t} is tangent to the characteristic foliation of ω\omega.

We say a Hamiltonian structure ω\omega is exact if ω\omega is an exact 22-form. We define the helicity of an exact Hamiltonian structure on an oriented closed smooth 33-manifold YY to be

(ω)Yαdα,\mathcal{H}(\omega)\coloneqq\int_{Y}\alpha\wedge d\alpha,

where α\alpha is any choice of primitive 11-form of ω\omega. As already mentioned, this integral is independent of the choice of α\alpha. By definition, a nowhere vanishing volume-preserving vector field XX is exact if and only if the induced Hamiltonian structure ω\omega is exact. If this is the case, we have

(X)=(ω).\mathcal{H}(X)=\mathcal{H}(\omega).

This perspective allows us to reformulate Question 1.1 concerning the topological invariance of helicity in terms of Hamiltonian structures. Specifically, the question becomes whether two exact Hamiltonian structures related by pullback under an orientation-preserving homeomorphism necessarily have the same helicity. To make sense of pullbacks by homeomorphisms, we rely on the interpretation of Hamiltonian structures as cooriented measured foliations.


C0C^{0} Hamiltonian Structures. Let YY be an oriented topological 33-manifold. A C0C^{0} Hamiltonian structure on YY is a cooriented, 11-dimensional C0C^{0} foliation equipped with a transverse measure, locally modeled on 3{\mathbb{R}}^{3} with the characteristic foliation induced by ωstd\omega_{\operatorname{std}}. A more detailed definition, formalized in terms of C0C^{0} Hamiltonian atlases, is given in Section 5. We will see in Section 8.2 that C0C^{0} Hamiltonian structures serve the same role for topological volume-preserving flows (whose flow lines have zero measure) as smooth Hamiltonian structures do for smooth volume-preserving flows. This allows us to recast Question 1.2 as a question about extending helicity to exact222We explain how to interpret exactness for C0C^{0} Hamiltonian structures in Section 7.1. C0C^{0} Hamiltonian structures.

1.4 Plugs and the Calabi invariant in the C0C^{0} setting

In the smooth setting, Gambaudo–Ghys [20] discovered a natural connection between helicity and the Calabi invariant. We construct our extensions of helicity such that this connection continues to hold in the C0C^{0} setting.

Let (Σ,ωΣ)(\Sigma,\omega_{\Sigma}) be an open surface with an area form, and let Ham(Σ)\operatorname{Ham}(\Sigma) and Ham¯(Σ)\overline{\operatorname{Ham}}(\Sigma) denote, respectively, the groups of compactly supported Hamiltonian diffeomorphisms and homeomorphisms of (Σ,ωΣ)(\Sigma,\omega_{\Sigma}); see Section 2. We denote by CalΣ:Ham(Σ)\operatorname{Cal}_{\Sigma}:\operatorname{Ham}(\Sigma)\rightarrow{\mathbb{R}} the Calabi homomorphism [6], which is defined as follows: for φ=φH1Ham(Σ)\varphi=\varphi_{H}^{1}\in\operatorname{Ham}(\Sigma), generated by a compactly supported Hamiltonian HC([0,1]×Σ)H\in C^{\infty}([0,1]\times\Sigma), we have333Different conventions exist for defining the Calabi invariant; we include the factor 2 for convenience.

CalΣ(φ)=2[0,1]×ΣH𝑑tωΣ.\operatorname{Cal}_{\Sigma}(\varphi)=2\int_{[0,1]\times\Sigma}Hdt\wedge\omega_{\Sigma}.

The 22-form ωΣ\omega_{\Sigma} induces a natural Hamiltonian structure on (0,1)×Σ(0,1)\times\Sigma, which we denote by the same symbol ωΣ\omega_{\Sigma}. Given an oriented 33-manifold YY equipped with a C0C^{0} Hamiltonian structure Ω\Omega, we define a plug to be a tuple

𝒫(Σ,ωΣ,α,(φt)t[0,1])\mathcal{P}\coloneqq(\Sigma,\omega_{\Sigma},\alpha,(\varphi^{t})_{t\in[0,1]})

where α:((0,1)×Σ),ωΣ)(Y,Ω)\alpha:((0,1)\times\Sigma),\omega_{\Sigma})\hookrightarrow(Y,\Omega) is an embedding of C0C^{0} Hamiltonian structures and φtHam¯(Σ)\varphi^{t}\in\overline{\operatorname{Ham}}(\Sigma) is a C0C^{0} Hamiltonian isotopy. Given a plug 𝒫\mathcal{P}, we define an operation, called plug insertion, which creates a new C0C^{0} Hamiltonian structure Ω#𝒫\Omega\#\mathcal{P} on YY by replacing the C0C^{0} Hamiltonian structure Ω\Omega inside im(α)\operatorname{im}(\alpha) with the pushforward of the C0C^{0} Hamiltonian structure on (0,1)×Σ(0,1)\times\Sigma induced by ωΣ\omega_{\Sigma} via the embedding αΦ\alpha\circ\Phi, where Φ:(0,1)×Σ(0,1)×Σ\Phi:(0,1)\times\Sigma\rightarrow(0,1)\times\Sigma is defined by Φ(t,p)(t,φt(p))\Phi(t,p)\coloneqq(t,\varphi^{t}(p)).444Plug insertion does not affect exactness of Hamiltonian structures, in smooth and C0C^{0} settings. The effect of this operation on the characteristic foliations is as follows: Before plug insertion, the characteristic leaves of Ω\Omega inside im(α)\operatorname{im}(\alpha) are of the form α((0,1)×{p})\alpha((0,1)\times\{p\}) for pΣp\in\Sigma. After plug insertion, the characteristic leaves of Ω#𝒫\Omega\#\mathcal{P} are of the form

{α(t,φt(p))t(0,1)}\{\alpha(t,\varphi^{t}(p))\mid t\in(0,1)\}

for pΣp\in\Sigma.

In the smooth setting, the following important identity, which can be deduced from [20], relates the helicities of Ω\Omega and Ω#𝒫\Omega\#\mathcal{P}:

(Ω#𝒫)=(Ω)+Cal(φ1)\mathcal{H}(\Omega\#\mathcal{P})=\mathcal{H}(\Omega)+\operatorname{Cal}(\varphi^{1}) (1.2)

The Calabi homomorphism was recently extended to the group of Hamiltonian homeomorphisms [14, 34]. In light of this, it is natural to require that identity (1.2) continues to hold for an extension of helicity to C0C^{0} Hamiltonian structures. However, in the smooth setting, the Calabi homomorphism satisfies a certain naturality property—implicit in identity (1.2)—which was not known to hold in the C0C^{0} category. Below, we establish this naturality in the C0C^{0} setting via Theorem 1.4—a result of independent interest that plays a key role in our main theorem.

A universal extension of Calabi.

In this subsection, Σ,Σ1,Σ2\Sigma,\Sigma_{1},\Sigma_{2} denote non-empty, open, connected surfaces equipped with area forms. A smooth area- and orientation-preserving embedding ι:Σ1Σ2\iota:\Sigma_{1}\hookrightarrow\Sigma_{2} induces an injective homomorphism Ham(ι):Ham(Σ1)Ham(Σ2)\operatorname{Ham}(\iota):\operatorname{Ham}(\Sigma_{1})\rightarrow\operatorname{Ham}(\Sigma_{2}) via pushforward. The Calabi homomorphism satisfies the following naturality property:

CalΣ1=CalΣ2Ham(ι)\operatorname{Cal}_{\Sigma_{1}}=\operatorname{Cal}_{\Sigma_{2}}\circ\operatorname{Ham}(\iota) (1.3)

Note, moreover, that the embedding ι\iota induces a homomorphism Hamab(ι)\operatorname{Ham}^{\operatorname{ab}}(\iota) at the level of abelianizations

Hamab(ι):Hamab(Σ1)Hamab(Σ2).\operatorname{Ham}^{\operatorname{ab}}(\iota):\operatorname{Ham}^{\operatorname{ab}}(\Sigma_{1})\rightarrow\operatorname{Ham}^{\operatorname{ab}}(\Sigma_{2}). (1.4)

Here GabG/[G,G]G^{\operatorname{ab}}\coloneqq G/[G,G] denotes the abelianization of a group GG. It follows from Banyaga’s work [5] that the homomorphism (1.4) is an isomorphism and is independent of the embedding ι\iota. Hence the group Hamab(Σ)\operatorname{Ham}^{\operatorname{ab}}(\Sigma) does not depend on Σ\Sigma up to canonical isomorphism. In fact, the Calabi homomorphism induces a natural isomorphism

CalΣ:Hamab(Σ).\operatorname{Cal}_{\Sigma}:\operatorname{Ham}^{\operatorname{ab}}(\Sigma)\overset{\cong}{\longrightarrow}{\mathbb{R}}.

For our arguments, we need extensions of the Calabi homomorphism

Cal¯Σ:Ham¯(Σ)\overline{\operatorname{Cal}}_{\Sigma}:\overline{\operatorname{Ham}}(\Sigma)\rightarrow{\mathbb{R}}

that satisfy an analogue of the naturality property (1.3) in the C0C^{0} setting. Although [14, 34] construct infinitely many C0C^{0} extensions of Calabi, none are canonical,555The constructions in [14, 34] rely on the axiom of choice in an essential way; see [14, Remark 5.1]. and it is unclear whether they satisfy the desired properties.

We obtain suitable extensions of Calabi as a consequence of the theorem below. For every area- and orientation-preserving topological embedding of surfaces ι:Σ1Σ2\iota:\Sigma_{1}\hookrightarrow\Sigma_{2}, let Ham¯ab(ι):Ham¯ab(Σ1)Ham¯ab(Σ2)\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota):\overline{\operatorname{Ham}}^{\operatorname{ab}}(\Sigma_{1})\rightarrow\overline{\operatorname{Ham}}^{\operatorname{ab}}(\Sigma_{2}) denote the induced homomorphism at the level of abelianizations.

Theorem 1.4.

The homomorphism

Ham¯ab(ι):Ham¯ab(Σ1)Ham¯ab(Σ2)\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota):\overline{\operatorname{Ham}}^{\operatorname{ab}}(\Sigma_{1})\rightarrow\overline{\operatorname{Ham}}^{\operatorname{ab}}(\Sigma_{2})

is an isomorphism. Moreover, it does not depend on the choice of area- and orientation-preserving embedding ι:Σ1Σ2\iota:\Sigma_{1}\hookrightarrow\Sigma_{2}, i.e. if ι:Σ1Σ2\iota^{\prime}:\Sigma_{1}\hookrightarrow\Sigma_{2} is another area- and orientation-preserving embedding, then

Ham¯ab(ι)=Ham¯ab(ι).\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota^{\prime})=\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota).

We state and prove a generalization of Theorem 1.4 in Section 3; see Theorem 3.1.

Set Ham¯ab(𝔻)\mathcal{R}\coloneqq\overline{\operatorname{Ham}}^{\operatorname{ab}}({\mathbb{D}}), where 𝔻2{\mathbb{D}}\subset{\mathbb{R}}^{2} denotes the open unit disc. By Theorem 1.4, we may canonically identify

Ham¯ab(Σ)\overline{\operatorname{Ham}}^{\operatorname{ab}}(\Sigma)\cong\mathcal{R} (1.5)

for every non-empty, connected, open surface Σ\Sigma.

It was an open question—known as the simplicity conjecture—whether the group Ham¯(𝔻)\overline{\operatorname{Ham}}({\mathbb{D}}) is simple. This question was recently resolved in the negative in [13], which showed that Ham¯(𝔻)\overline{\operatorname{Ham}}({\mathbb{D}}) is not simple. Moreover, it is shown in [13, Cor. 1.3] that Ham¯(𝔻)\overline{\operatorname{Ham}}({\mathbb{D}}) is not perfect either. In other words, the group \mathcal{R} is non-trivial.

It can be further deduced from [14, 34] that the natural homomorphism

Hamab(Σ)Ham¯ab(Σ)\operatorname{Ham}^{\operatorname{ab}}(\Sigma)\rightarrow\overline{\operatorname{Ham}}^{\operatorname{ab}}(\Sigma) (1.6)

is injective and not surjective; see Section 3. The injectivity of the homomorphism (1.6), a key input from C0C^{0} symplectic topology, is essential for our proof of the topological invariance of helicity. Recalling that the Calabi homomorphism induces an identification Hamab(Σ)\operatorname{Ham}^{\operatorname{ab}}(\Sigma)\cong{\mathbb{R}}, the homomorphism (1.6) allows us to naturally view {\mathbb{R}} as a subgroup

.{\mathbb{R}}\subset\mathcal{R}. (1.7)

In view of the natural commutative diagram

Ham¯(Σ){\overline{\operatorname{Ham}}(\Sigma)}Ham¯ab(Σ){\overline{\operatorname{Ham}}^{\operatorname{ab}}(\Sigma)}{\mathcal{R}}Ham(Σ){\operatorname{Ham}(\Sigma)}Hamab(Σ){\operatorname{Ham}^{\operatorname{ab}}(\Sigma)}{{\mathbb{R}}}\scriptstyle{\cong}CalΣ\scriptstyle{\operatorname{Cal}_{\Sigma}}\scriptstyle{\cong}

we call the natural homomorphism

Cal¯Σ:Ham¯(Σ)Ham¯ab(Σ)\overline{\operatorname{Cal}}_{\Sigma}:\overline{\operatorname{Ham}}(\Sigma)\rightarrow\overline{\operatorname{Ham}}^{\operatorname{ab}}(\Sigma)\cong\mathcal{R} (1.8)

the universal (or \mathcal{R}-valued) extension of the Calabi homomorphism.666Note that this amounts to a change of notation from (1.1). From this point on-wards, Cal¯Σ\overline{\mathrm{Cal}}_{\Sigma} will always denote the universal extension of Calabi, unless otherwise stated.777Recall that the standing assumption of this subsection is that surfaces are connected. We will encounter non-connected surfaces Σ\Sigma later on, and in this case we define Cal¯Σ:Ham¯(Σ)\overline{\mathrm{Cal}}_{\Sigma}:\overline{\operatorname{Ham}}(\Sigma)\rightarrow\mathcal{R} to be the sum of all Cal¯S\overline{\mathrm{Cal}}_{S} where SS ranges over the components of Σ\Sigma. Its restriction to Ham(Σ)\operatorname{Ham}(\Sigma) agrees with the usual smooth Calabi homomorphism CalΣ\operatorname{Cal}_{\Sigma} via the natural inclusion {\mathbb{R}}\subset\mathcal{R}. Moreover, by Theorem 1.4 it satisfies the naturality property

Cal¯Σ1=Cal¯Σ2Ham¯(ι)\overline{\mathrm{{Cal}}}_{\Sigma_{1}}=\overline{\mathrm{{Cal}}}_{\Sigma_{2}}\circ\overline{\operatorname{Ham}}(\iota) (1.9)

for any area- and orientation-preserving topological embedding ι:Σ1Σ2\iota:\Sigma_{1}\hookrightarrow\Sigma_{2}.

The extension Cal¯Σ\overline{\operatorname{Cal}}_{\Sigma} is universal in the following sense: Consider an arbitrary extension of abelian groups A{\mathbb{R}}\subset A. Then every system of extensions

Cal¯ΣA:Ham¯(Σ)A\overline{\operatorname{Cal}}_{\Sigma}^{A}:\overline{\operatorname{Ham}}(\Sigma)\rightarrow A

of Calabi, one for every surface Σ\Sigma and subject to the naturality condition (1.9), arises as

Cal¯ΣA=pCal¯Σ\overline{\operatorname{Cal}}^{A}_{\Sigma}=p\circ\overline{\operatorname{Cal}}_{\Sigma}

for a unique group homomorphism p:Ap:\mathcal{R}\rightarrow A over {\mathbb{R}}. In particular, for every choice of projection pr:\operatorname{pr}:\mathcal{R}\rightarrow{\mathbb{R}}, we obtain a system of {\mathbb{R}}-valued Calabi extensions prCal¯Σ\operatorname{pr}\circ\overline{\operatorname{Cal}}_{\Sigma} subject to (1.9), and every such system uniquely arises this way. Note that a projection pr:\operatorname{pr}:\mathcal{R}\rightarrow{\mathbb{R}} is equivalent to a choice of a single Calabi extension Cal¯𝔻:Ham¯(𝔻)\overline{\operatorname{Cal}}_{\mathbb{D}}^{{\mathbb{R}}}:\overline{\operatorname{Ham}}({\mathbb{D}})\rightarrow{\mathbb{R}} for the open unit disc 𝔻{\mathbb{D}}. Thanks to [14], such extensions are abundant, but unfortunately not very canonical since they are found using the axiom of choice. In fact, as pointed out in [14, Remark 5.1], there are models of set theory in which the axiom of choice is false and every group homomorphism between Polish groups is automatically continuous. Since there does not exist a C0C^{0} continuous extension of the Calabi homomorphism to Ham¯(𝔻)\overline{\operatorname{Ham}}({\mathbb{D}}), in these models there is no extension as a group homomorphism at all.

For this reason, we prefer to work with the universal Calabi extension and construct a universal \mathcal{R}-valued extension of helicity to exact C0C^{0} Hamiltonian structures. It is always possible to obtain an {\mathbb{R}}-valued helicity extension from this if one wishes to, but this requires a non-canonical choice.

1.5 Main Result: A universal extension of helicity

We are now in position to state our main result, which defines helicity ¯(Ω)\overline{\mathcal{H}}(\Omega) for an arbitrary exact C0C^{0} Hamiltonian structure Ω\Omega on an oriented closed topological 33-manifold YY. The notion of exactness for C0C^{0} Hamiltonian structures is defined as the vanishing of a certain cohomology class Flux¯(Ω)H2(Y;)\overline{\mathrm{Flux}}(\Omega)\in H^{2}(Y;{\mathbb{R}}). In particular, if YY is a rational homology three-sphere, then every C0C^{0} Hamiltonian structure is exact. This is discussed in detail in Section 7.1.

Recall that Ω#𝒫\Omega\#\mathcal{P} denotes the plug insertion operation introduced in Section 1.4. Moreover, fΩf^{*}\Omega stands for the pullback of Ω\Omega under a homeomorphism ff.

Theorem 1.5.

There is a unique way of assigning a \mathcal{R}-valued helicity ¯(Ω)\overline{\mathcal{H}}(\Omega)\in\mathcal{R} to every exact C0C^{0} Hamiltonian structure Ω\Omega on an oriented closed topological 33-manifold YY such that the following conditions are satisfied:

  1. 1.

    Extension: If ω\omega and YY are smooth, then

    ¯(ω)=(ω),\overline{\mathcal{H}}(\omega)=\mathcal{H}(\omega)\in{\mathbb{R}},

    where we view {\mathbb{R}} as a subgroup of \mathcal{R} via the natural inclusion (1.7).

  2. 2.

    Invariance: We have

    ¯(fΩ)=¯(Ω),\overline{\mathcal{H}}(f^{*}\Omega)=\overline{\mathcal{H}}(\Omega),

    for any orientation-preserving homeomorphism ff.

  3. 3.

    Calabi Compatibility: For every plug 𝒫=(Σ,ωΣ,α,φt)\mathcal{P}=(\Sigma,\omega_{\Sigma},\alpha,\varphi^{t}), have

    ¯(Ω#𝒫)=¯(Ω)+Cal¯Σ(φ1).\overline{\mathcal{H}}(\Omega\#\mathcal{P})=\overline{\mathcal{H}}(\Omega)+\overline{\mathrm{Cal}}_{\Sigma}(\varphi^{1}). (1.10)

As already mentioned, every choice of projection pr:\operatorname{pr}:\mathcal{R}\rightarrow{\mathbb{R}} gives rise to a real-valued extension of helicity.

We point out that it follows from the Extension and Invariance properties in Theorem 1.5 that two smooth Hamiltonian structures which are conjugated by an orientation-preserving homeomorphism have the same helicity. The proof of this makes essential use of the injectivity of the natural map Hamab(𝔻)Ham¯ab(𝔻){\mathbb{R}}\cong\operatorname{Ham}^{\operatorname{ab}}({\mathbb{D}})\rightarrow\overline{\operatorname{Ham}}^{\operatorname{ab}}({\mathbb{D}})\cong\mathcal{R}.

Let us elaborate further on the universality of our helicity extension ¯\overline{\mathcal{H}} and the naturality of the properties (Extension, Invariance, and Calabi) that characterize it in Theorem 1.5. The properties Extension and Invariance can be viewed as minimal requirements that any reasonable extension of helicity to C0C^{0} Hamiltonian structures ought to satisfy. In light of identity (1.2), which describes the helicity change under plug insertion in the smooth setting, the Calabi property—though less obvious—is likewise a natural condition to impose.

To further motivate this property, observe that identity (1.2) has the following direct consequence: the helicity change (ω#𝒫)(ω)\mathcal{H}(\omega\#\mathcal{P})-\mathcal{H}(\omega) resulting from the insertion of a smooth plug 𝒫=(Σ,ωΣ,α,φt)\mathcal{P}=(\Sigma,\omega_{\Sigma},\alpha,\varphi^{t}) depends only on the time-one map φ1\varphi^{1}, and is independent of both the ambient Hamiltonian structure ω\omega and the embedding α\alpha of the plug. We refer to this fundamental property as plug homogeneity. Remarkably, imposing plug homogeneity—together with Extension and Invariance—already forces any helicity extension to be a function of our universal extension ¯\overline{\mathcal{H}}.

Proposition 1.6.

Let A{\mathbb{R}}\subset A be an extension of abelian groups. Let ¯A(Ω)A\overline{\mathcal{H}}^{A}(\Omega)\in A be an AA-valued extension of helicity to exact C0C^{0} Hamiltonian structures satisfying the Extension and Invariance properties in Theorem 1.5. Moreover, assume that ¯A\overline{\mathcal{H}}^{A} satisfies plug homogeneity. That is, for any C0C^{0} Hamiltonian structure Ω\Omega and for any Ω\Omega-plug 𝒫=(Σ,ωΣ,α,φt)\mathcal{P}=(\Sigma,\omega_{\Sigma},\alpha,\varphi^{t}), the helicity difference

¯A(Ω#𝒫)¯A(Ω)A\overline{\mathcal{H}}^{A}(\Omega\#\mathcal{P})-\overline{\mathcal{H}}^{A}(\Omega)\in A

only depends on φ1\varphi^{1} and is independent of Ω\Omega and α\alpha. Then there exists a unique homomorphism p:Ap:\mathcal{R}\rightarrow A over {\mathbb{R}} such that ¯A=p¯\overline{\mathcal{H}}^{A}=p\circ\overline{\mathcal{H}}.

This means that, once plug homogeneity is accepted as a fundamental property of helicity, one is essentially forced to arrive at our universal \mathcal{R}-valued helicity extension ¯\overline{\mathcal{H}}. Moreover, insisting on plug homogeneity makes the problem of defining an {\mathbb{R}}-valued extension of helicity equivalent to specifying a projection pr:\operatorname{pr}:\mathcal{R}\rightarrow{\mathbb{R}}, which, as explained in Subsection 1.4, cannot be done without making non-canonical choices. This indicates that \mathcal{R} is indeed the natural target group for a helicity extension to exact C0C^{0} Hamiltonian structures.

Remark 1.7.

We end our introduction with the following remarks.

  1. 1.

    The \mathcal{R}-valued helicity extension from Theorem 1.5 provides an obstruction to smoothability of C0C^{0} Hamiltonian structures: if ¯(Ω)\overline{\mathcal{H}}(\Omega) lies in \mathcal{R}\setminus\mathbb{R}, then Ω\Omega is not homeomorphic to any smooth Hamiltonian structure.

  2. 2.

    In a sequel to this work, we will investigate helicity from the point of view of characteristic classes of foliations and Haefliger structures.

  3. 3.

    While our results affirmatively answer Arnold’s Questions 1.1 and 1.2 for flows without fixed points, they remain open for flows with fixed points. In future work, we will address topological invariance of helicity in the presence of certain types of singularities, in particular the generic case of non-degenerate singularities. At present, it is unclear whether sufficiently complicated singular sets may destroy topological invariance.

Structure of the paper.

In Section 2, we review some preliminaries concerning area-preserving homeomorphisms, foliations, and transverse measures.

In Section 3, we prove a generalization of Theorem 1.4, which leads to the construction of our universal \mathcal{R}-valued extension of the Calabi homomorphism satisfying the naturality property stated in Section 1.4.

In Sections 4 and 5, we provide precise definitions of smooth and C0C^{0} Hamiltonian structures, along with the plug insertion operation.

Section 6 presents two central results—Theorems 6.1 and 6.11—which describe the structure of C0C^{0} Hamiltonian structures and their homeomorphisms. These theorems roughly state that such structures and maps can be smoothed up to the insertion of a carefully constructed plug.

In Section 7, we extend the definitions of flux and helicity to C0C^{0} Hamiltonian structures, thereby completing the proof of Theorem 1.5. We also provide a proof of Proposition 1.6.

Finally, in Section 8, we carefully explain the relationship between Hamiltonian structures and volume-preserving flows, both in the smooth and in the C0C^{0} setting. This is essential for deducing Theorem 1.3, which is formulated for topological volume-preserving flows, from our main result, Theorem 1.5, stated for C0C^{0} Hamiltonian structures.


Convention: Throughout this paper, all manifolds will be oriented and all (local) homeomorphisms between manifolds will be orientation preserving, unless specified otherwise.

Acknowledgments

We are grateful to Vikram Giri and Hyunju Kwon for helpful explanations about their work [23]. We thank Étienne Ghys, Boris Khesin and Chi Cheuk Tsang for their interest and comments. S.S. is indebted to Boris Khesin for introducing him to Arnold’s questions and for enlightening discussions.

O.E. is supported by Dr. Max Rössler, the Walter Haefner Foundation, and the ETH Zürich Foundation. S.S. is partially supported by ERC Starting Grant number 851701 and a start up grant from ETH-Zürich.

2 Preliminaries

We recall some preliminaries concerning area-preserving homeomorphisms and foliations, which will be needed in the forthcoming sections.

2.1 Area-preserving homeomorphisms

Let (Σ,ω)(\Sigma,\omega) be a smooth 22-manifold equipped with an area form. We assume that Σ\Sigma does not have boundary, but we allow it to be open.

Every compactly supported smooth Hamiltonian H:[0,1]×ΣH:[0,1]\times\Sigma\rightarrow{\mathbb{R}} induces a time-dependent Hamiltonian vector field XHX_{H} generating a compactly supported Hamiltonian isotopy (φHt)t[0,1](\varphi_{H}^{t})_{t\in[0,1]}. We adopt the sign convention that XHX_{H} is characterized by the identity ιXHtω=dHt\iota_{X_{H_{t}}}\omega=dH_{t}. The compactly supported Hamiltonian diffeomorphism group of (Σ,ω)(\Sigma,\omega) is denoted by Ham(Σ,ω)\operatorname{Ham}(\Sigma,\omega).

Let Homeoc(Σ,ω)\operatorname{Homeo}_{c}(\Sigma,\omega) denote the group of all compactly supported homeomorphisms of Σ\Sigma preserving the measure induced by ω\omega. We topologize Homeoc(Σ,ω)\operatorname{Homeo}_{c}(\Sigma,\omega) as the direct limit of all the subgroups HomeoK(Σ,ω)\operatorname{Homeo}_{K}(\Sigma,\omega) of homeomorphisms supported inside some compact subset KΣK\subset\Sigma, equipped with the compact-open topology. The group of Hamiltonian homeomorphisms Ham¯(Σ,ω)\overline{\operatorname{Ham}}(\Sigma,\omega) is defined to be the closure of Ham(Σ,ω)\operatorname{Ham}(\Sigma,\omega) inside Homeoc(Σ,ω)\operatorname{Homeo}_{c}(\Sigma,\omega). It is topologized as a subspace of Homeoc(Σ,ω)\operatorname{Homeo}_{c}(\Sigma,\omega).

Let Homeo0(Σ,ω)\operatorname{Homeo}_{0}(\Sigma,\omega) denote the identity component of Homeoc(Σ,ω)\operatorname{Homeo}_{c}(\Sigma_{,}\omega). Then Ham¯(Σ,ω)\overline{\operatorname{Ham}}(\Sigma,\omega) agrees with the kernel of the mass flow homomorphism [18, 48]

θ:Homeo0(Σ,ω)H1(Σ;)/Γ,\theta:\operatorname{Homeo}_{0}(\Sigma,\omega)\rightarrow H^{1}(\Sigma;{\mathbb{R}})/\Gamma,

where ΓH1(Σ;)\Gamma\subset H^{1}(\Sigma;{\mathbb{R}}) is a discrete subgroup whose precise definition will not be relevant for us.

Whenever there is no risk of confusion, we will abbreviate Ham(Σ)Ham(Σ,ω)\operatorname{Ham}(\Sigma)\coloneqq\operatorname{Ham}(\Sigma,\omega) and Ham¯(Σ)Ham¯(Σ,ω)\overline{\operatorname{Ham}}(\Sigma)\coloneqq\overline{\operatorname{Ham}}(\Sigma,\omega).

It is known that Ham(Σ)\operatorname{Ham}(\Sigma) and Ham¯(Σ)\overline{\operatorname{Ham}}(\Sigma), equipped with the C0C^{0} topology, are both simply connected, unless Σ\Sigma is a sphere. If Σ\Sigma is a sphere, the fundamental group is /2\mathbb{Z}/2\mathbb{Z}. In the case of Ham(Σ)\operatorname{Ham}(\Sigma), a proof of these facts is sketched in [44, Sec. 7.2] for closed Σ\Sigma. The arguments therein can be adapted to the case of open Σ\Sigma. As for Ham¯(Σ)\overline{\operatorname{Ham}}(\Sigma), it can be shown that every loop in Ham¯(Σ)\overline{\operatorname{Ham}}(\Sigma) is homotopic a loop in Ham(Σ)\operatorname{Ham}(\Sigma) using the following two facts: first, every homeomorphism in Ham¯(Σ)\overline{\operatorname{Ham}}(\Sigma) can be written as a C0C^{0} limit of elements in Ham(Σ)\operatorname{Ham}(\Sigma); and second, Ham(Σ)\operatorname{Ham}(\Sigma) and Ham¯(Σ)\overline{\operatorname{Ham}}(\Sigma) are both locally path connected; see, for example, [49, Cor. 2] or [50, Lem. 3.2].

The following two propositions will be used in the upcoming sections. Both are variants of known fragmentation and approximation results for area-preserving maps and can be established using classical techniques—specifically, fragmentation techniques and approximation of homeomorphisms by diffeomorphisms in the area-preserving setting; see, e.g., [33, 17]. We therefore omit the proofs.

Proposition 2.1.

Let (Σ,ω)(\Sigma,\omega) be a surface with area form.

  1. 1.

    Suppose U1,,UnU_{1},\dots,U_{n} is a finite open cover of Σ\Sigma, and for each ii, let 𝒰iHam¯(Ui)\mathcal{U}_{i}\subset\overline{\operatorname{Ham}}(U_{i}) be an open neighborhood of the identity. Then any Hamiltonian homeomorphism φHam¯(Σ)\varphi\in\overline{\operatorname{Ham}}(\Sigma) that lies in a sufficiently small neighborhood of the identity can be written as a composition φ=φ1φn\varphi=\varphi_{1}\circ\cdots\circ\varphi_{n}, where each φi𝒰iHam¯(Ui)\varphi_{i}\in\mathcal{U}_{i}\subset\overline{\operatorname{Ham}}(U_{i}).

  2. 2.

    Let 𝒱\mathcal{V} be an arbitrary open cover of Σ\Sigma. Then, for every Hamiltonian homeomorphism φHam¯(Σ)\varphi\in\overline{\operatorname{Ham}}(\Sigma), there exist finitely many open sets U1,,Un𝒱U_{1},\dots,U_{n}\in\mathcal{V} and Hamiltonian homeomorphisms φiHam¯(Ui)\varphi_{i}\in\overline{\operatorname{Ham}}(U_{i}) such that φ=φ1φn\varphi=\varphi_{1}\circ\cdots\circ\varphi_{n}.

Proposition 2.2.

Let (Σi,ωi)(\Sigma_{i},\omega_{i}) be a surface equipped with an area form for each i{1,2}i\in\{1,2\}. Suppose φ:(Σ1,ω1)(Σ2,ω2)\varphi:(\Sigma_{1},\omega_{1})\to(\Sigma_{2},\omega_{2}) is an area-preserving homeomorphism. Let UΣ1U\subset\Sigma_{1} be an open subset on which φ\varphi is smooth. Let dd be a distance function on Σ2\Sigma_{2} that induces its topology, and let ρ:Σ10\rho:\Sigma_{1}\to\mathbb{R}_{\geq 0} be a continuous, non-negative function satisfying ρ1({0})U\rho^{-1}(\{0\})\subset U. Then there exists an area-preserving diffeomorphism ψ:(Σ1,ω1)(Σ2,ω2)\psi:(\Sigma_{1},\omega_{1})\to(\Sigma_{2},\omega_{2}) such that

d(φ(x),ψ(x))ρ(x)for all xΣ1.d(\varphi(x),\psi(x))\leq\rho(x)\quad\text{for all }x\in\Sigma_{1}.

2.2 Foliations and transverse measures

Let MnM^{n} be an oriented CrC^{r} manifold for r0{}r\in{\mathbb{Z}}_{\geq 0}\cup\{\infty\}. For k0k\geq 0, let BkkB^{k}\subset{\mathbb{R}}^{k} denote the unit ball. We define a dd-dimensional CrC^{r} foliation \mathcal{F} on MM to be a decomposition of MM into connected subsets LML\subset M, called leaves, such that MM can be covered by CrC^{r} coordinate charts

ϕ:UBd×Bndn,\phi:U\rightarrow B^{d}\times B^{n-d}\subset{\mathbb{R}}^{n},

called foliation charts, with the property that, for every leaf LL, the connected components of ϕ(LU)\phi(L\cap U) are of the form Bd×{}B^{d}\times\{*\}. The collection of the foliation charts {(U,ϕ)}\{(U,\phi)\} defining \mathcal{F} forms a foliated atlas. Since the manifold MM is oriented, we may assume without loss of generality that the foliated atlas on MM is an oriented atlas of MM, meaning that we require the transition maps between the foliation charts to be orientation preserving. Note that this does not imply that \mathcal{F} is oriented.

We will consider foliations \mathcal{F} which are equipped with two additional structures: a transverse measure Λ\Lambda and a transverse orientation, also referred to as a coorientation. A transverse measure Λ\Lambda is specified by a foliated atlas of MM consisting of foliation charts ϕ:UBd×Bnd\phi:U\rightarrow B^{d}\times B^{n-d} and a finite Borel measure λϕ\lambda_{\phi} on BndB^{n-d} for every chart ϕ\phi such that the following is true: For i{1,2}i\in\{1,2\}, let ϕi:UiBd×Bnd\phi_{i}:U_{i}\rightarrow B^{d}\times B^{n-d} be a chart and consider the transition map

ϕ21:ϕ1(U1U2)ϕ2(U1U2).\phi_{21}:\phi_{1}(U_{1}\cap U_{2})\rightarrow\phi_{2}(U_{1}\cap U_{2}).

Then, for every point p0=(x0,y0)ϕ1(U1U2)p_{0}=(x_{0},y_{0})\in\phi_{1}(U_{1}\cap U_{2}), there exist an open neighborhood VV of x0x_{0} in BdB^{d}, an open neighborhood WW of y0y_{0} in BndB^{n-d}, and a topological embedding ι:WBnd\iota:W\hookrightarrow B^{n-d} such that the restriction of ϕ21\phi_{21} to V×WV\times W is of the form ϕ21(x,y)=(,ι(y))\phi_{21}(x,y)=(*,\iota(y)) and moreover

ιλϕ1|W=λϕ2|ι(W).\iota_{*}\lambda_{\phi_{1}}|_{W}=\lambda_{\phi_{2}}|_{\iota(W)}.

Requiring that the topological embeddings ι\iota are orientation preserving determines a transverse orientation, or coorientation, of the foliation \mathcal{F}.

Given a foliation \mathcal{F}, a transversal is a compact topological submanifold with boundary TndMT^{n-d}\subset M that is transverse to the foliation \mathcal{F}. This means that every point in TT is contained in a foliation chart ϕ:UBd×Bnd\phi:U\rightarrow B^{d}\times B^{n-d} such that ϕ(TU)\phi(T\cap U) is contained in a transverse slice of the form {}×Bnd\{*\}\times B^{n-d}. A transverse measure Λ\Lambda induces a finite Borel measure on every transversal TT and maps between subsets of transversals obtained by sliding along leaves of \mathcal{F} are measure preserving. In fact, one can equivalently define a transverse measure Λ\Lambda on \mathcal{F} to be an assignment of a finite Borel measure to every transversal TT subject to the condition that sliding along leaves of \mathcal{F} preserves measure. A transverse orientation admits a similar description: it is an assignment of an orientation to every transversal TT such that sliding along leaves of \mathcal{F} preserves orientation.

Finally, note that since the ambient manifold MM is oriented, a coorientation of \mathcal{F} induces a natural orientation on the foliation \mathcal{F}. Vice-versa, the orientation on MM together with an orientation of \mathcal{F} determine a coorientation of \mathcal{F}.

3 Universal extension of the Calabi homomorphism

In this section, we prove Theorem 3.1, which generalizes Theorem 1.4 of the introduction. As a consequence, we deduce the existence of a \mathcal{R}-valued universal extension of the Calabi homomorphism which satisfies the naturality property stated in Section 1.4.

Let 𝒮\mathcal{S} be the category whose objects are pairs (Σ,ω)(\Sigma,\omega), where Σ\Sigma is a non-empty, connected, smooth surface without boundary, and ω\omega is an area form on Σ\Sigma. Morphisms in 𝒮\mathcal{S} are smooth embeddings ι:(Σ1,ω1)(Σ2,ω2)\iota:(\Sigma_{1},\omega_{1})\hookrightarrow(\Sigma_{2},\omega_{2}). We include both open and closed surfaces, without requiring finite area in the open case. Let 𝒮op\mathcal{S}_{\operatorname{op}} and 𝒮cl\mathcal{S}_{\operatorname{cl}} denote the full subcategories of open and closed surfaces, respectively. When there is no risk of confusion, we simply write Σ\Sigma for (Σ,ω)(\Sigma,\omega).

Recall that Ham(Σ)\operatorname{Ham}(\Sigma) denotes the group of all compactly supported Hamiltonian diffeomorphisms of Σ\Sigma. Every area-preserving embedding ι:Σ1Σ2\iota:\Sigma_{1}\hookrightarrow\Sigma_{2} induces a group homomorphism

Ham(ι):Ham(Σ1)Ham(Σ2)Ham(ι)(φ)ιφ\operatorname{Ham}(\iota):\operatorname{Ham}(\Sigma_{1})\rightarrow\operatorname{Ham}(\Sigma_{2})\qquad\operatorname{Ham}(\iota)(\varphi)\coloneqq\iota_{*}\varphi

given by pushforward of compactly supported Hamiltonian diffeomorphisms via ι\iota. We can therefore regard Ham\operatorname{Ham} as a functor from 𝒮\mathcal{S} to the category of all groups. We define

Hamab(Σ)Ham(Σ)/[Ham(Σ),Ham(Σ)]\operatorname{Ham}^{\operatorname{ab}}(\Sigma)\coloneqq\operatorname{Ham}(\Sigma)/[\operatorname{Ham}(\Sigma),\operatorname{Ham}(\Sigma)]

to be the abelianization of Ham(Σ)\operatorname{Ham}(\Sigma). Since abelianization forms a functor from the category of groups to the category of abelian groups, Hamab\operatorname{Ham}^{\operatorname{ab}} is a functor from 𝒮\mathcal{S} to the category of abelian groups.

It was proved by Banyaga [5] that Ham(Σ)\operatorname{Ham}(\Sigma) is perfect for closed Σ\Sigma. In other words,

Hamab(Σ)=1\operatorname{Ham}^{\operatorname{ab}}(\Sigma)=1

for every closed surface Σ𝒮cl\Sigma\in\mathcal{S}_{\operatorname{cl}}. For open surfaces Σ𝒮op\Sigma\in\mathcal{S}_{\operatorname{op}} (and more generally for exact connected symplectic manifolds of arbitrary dimension), Banyaga showed that Hamab(Σ)\operatorname{Ham}^{\operatorname{ab}}(\Sigma)\cong{\mathbb{R}}. Moreover, an explicit isomorphism between Hamab(Σ)\operatorname{Ham}^{\operatorname{ab}}(\Sigma) and {\mathbb{R}} is induced by the Calabi homomorphism

CalΣ:Ham(Σ),\operatorname{Cal}_{\Sigma}:\operatorname{Ham}(\Sigma)\rightarrow{\mathbb{R}},

whose definition we now recall. For a given φHam(Σ)\varphi\in\operatorname{Ham}(\Sigma), pick H:[0,1]×ΣH:[0,1]\times\Sigma\rightarrow{\mathbb{R}} such that φH1=φ\varphi_{H}^{1}=\varphi. Then,

CalΣ(φ)=2[0,1]×ΣH𝑑tω.\operatorname{Cal}_{\Sigma}(\varphi)=2\int_{[0,1]\times\Sigma}Hdt\wedge\omega.

This turns out to be independent of the choice of the Hamiltonian HH. Banyaga proved that CalΣ\operatorname{Cal}_{\Sigma} is a surjective group homomorphism whose kernel is given by the commutator subgroup of Ham(Σ)\operatorname{Ham}(\Sigma). In other words, the Calabi homomorphism descends to an isomorphism CalΣ:Hamab(Σ)\operatorname{Cal}_{\Sigma}:\operatorname{Ham}^{\operatorname{ab}}(\Sigma)\rightarrow{\mathbb{R}}.

The Calabi homomorphism is natural in the following sense: If ι:Σ1Σ2\iota:\Sigma_{1}\hookrightarrow\Sigma_{2} is an area-preserving embedding between open surfaces Σ1,Σ2𝒮op\Sigma_{1},\Sigma_{2}\in\mathcal{S}_{\operatorname{op}}, then

CalΣ2Ham(ι)=CalΣ1.\operatorname{Cal}_{\Sigma_{2}}\circ\operatorname{Ham}(\iota)=\operatorname{Cal}_{\Sigma_{1}}.

This follows from the observation that the integral of a compactly supported Hamiltonian does not change if we push it forward via an area-preserving embedding. Naturality of the Calabi homomorphism and the fact that CalΣ:Hamab(Σ)\operatorname{Cal}_{\Sigma}:\operatorname{Ham}^{\operatorname{ab}}(\Sigma)\rightarrow{\mathbb{R}} is an isomorphism for every open surface Σ𝒮op\Sigma\in\mathcal{S}_{\operatorname{op}} imply that Hamab(ι)\operatorname{Ham}^{\operatorname{ab}}(\iota) is an isomorphism for all embeddings ι\iota between open surfaces. Moreover, one can deduce that Hamab(ι)\operatorname{Ham}^{\operatorname{ab}}(\iota) is independent of the choice of embedding ι\iota.

Recall from Section 2 that Ham¯(Σ)\overline{\operatorname{Ham}}(\Sigma) denotes the group of compactly supported Hamiltonian homeomorphisms of Σ\Sigma. As before, this assignment defines a functor from the category 𝒮\mathcal{S} to the category of groups. Moreover, since Hamiltonian homeomorphisms can be pushed forward via area-preserving topological embeddings, we can extend this functor to a larger category 𝒮¯\overline{\mathcal{S}}, which has the same objects as 𝒮\mathcal{S} but allows all area-preserving topological embeddings as morphisms.

There is a natural inclusion Ham(Σ)Ham¯(Σ)\operatorname{Ham}(\Sigma)\subset\overline{\operatorname{Ham}}(\Sigma), which corresponds to a natural transformation from the functor Ham\operatorname{Ham} to the composition of the inclusion 𝒮𝒮¯\mathcal{S}\hookrightarrow\overline{\mathcal{S}} with the functor Ham¯\overline{\operatorname{Ham}}.

As mentioned in the introduction, for every area-preserving topological embedding of surfaces ι:Σ1Σ2\iota:\Sigma_{1}\hookrightarrow\Sigma_{2}, we let Ham¯ab(ι):Ham¯ab(Σ1)Ham¯ab(Σ2)\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota):\overline{\operatorname{Ham}}^{\operatorname{ab}}(\Sigma_{1})\rightarrow\overline{\operatorname{Ham}}^{\operatorname{ab}}(\Sigma_{2}) denote the induced homomorphism at the level of abelianizations, where

Ham¯ab(Σ)Ham¯(Σ)/[Ham¯(Σ),Ham¯(Σ)].\overline{\operatorname{Ham}}^{\operatorname{ab}}(\Sigma)\coloneqq\overline{\operatorname{Ham}}(\Sigma)/[\overline{\operatorname{Ham}}(\Sigma),\overline{\operatorname{Ham}}(\Sigma)].

As in the smooth setting, Ham¯ab\overline{\operatorname{Ham}}^{\operatorname{ab}} constitutes a functor from 𝒮¯\overline{\mathcal{S}} to the category of abelian groups.

We prove the following generalization of Theorem 1.4.

Theorem 3.1.

For i{1,2}i\in\{1,2\}, let (Σi,ωi)(\Sigma_{i},\omega_{i}) be non-empty connected surfaces without boundary and equipped with area forms. Let ι:Σ1Σ2\iota:\Sigma_{1}\hookrightarrow\Sigma_{2} be an area-preserving topological embedding. Then, the induced map

Ham¯ab(ι):Ham¯ab(Σ1)Ham¯ab(Σ2)\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota):\overline{\operatorname{Ham}}^{\operatorname{ab}}(\Sigma_{1})\rightarrow\overline{\operatorname{Ham}}^{\operatorname{ab}}(\Sigma_{2})

does not depend on the choice of embedding ι:Σ1Σ2\iota:\Sigma_{1}\hookrightarrow\Sigma_{2}, i.e. if ι:Σ1Σ2\iota^{\prime}:\Sigma_{1}\hookrightarrow\Sigma_{2} is any other embedding, then

Ham¯ab(ι)=Ham¯ab(ι).\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota^{\prime})=\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota).

Moreover,

  • i.

    If Σ1\Sigma_{1} and Σ2\Sigma_{2} are both open or both closed, then Ham¯ab(ι)\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota) is an isomorphism.

  • ii.

    If Σ1\Sigma_{1} is open and Σ2\Sigma_{2} is closed, then Ham¯ab(ι)\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota) is surjective and its kernel is given by the image of Hamab(Σ1)\operatorname{Ham}^{\operatorname{ab}}(\Sigma_{1}) in Ham¯ab(Σ1)\overline{\operatorname{Ham}}^{\operatorname{ab}}(\Sigma_{1}).

Before presenting the proof of the above, we remark on some of its consequences.

For every surface Σ𝒮\Sigma\in\mathcal{S}, the abelian group Ham¯ab(Σ)\overline{\operatorname{Ham}}^{\operatorname{ab}}(\Sigma) is non-trivial. This was proven in the case of the disc in [13] and for closed surfaces and open surfaces which are the interiors of compact surfaces with boundary in [12, 11]. For general open surfaces, possibly of infinite area, non-triviality of Ham¯ab(Σ)\overline{\operatorname{Ham}}^{\operatorname{ab}}(\Sigma) can be deduced from Theorem 3.1.

It is interesting to point out that non-perfectness of Ham¯(Σ)\overline{\operatorname{Ham}}(\Sigma) for one single open surface Σ𝒮op\Sigma\in\mathcal{S}_{\operatorname{op}}, for example Σ=𝔻\Sigma={\mathbb{D}}, implies non-perfectness for all open surfaces via Theorem 3.1. Similarly, non-perfectness of Ham¯(Σ)\overline{\operatorname{Ham}}(\Sigma) for one single closed surface Σ𝒮cl\Sigma\in\mathcal{S}_{\operatorname{cl}} implies non-perfectness for all closed surfaces.

As already explained in the introduction, we define

Ham¯ab(𝔻).\mathcal{R}\coloneqq\overline{\operatorname{Ham}}^{\operatorname{ab}}(\mathbb{D}).

Theorem 3.1 allows us to canonically identify

Ham¯ab(Σ)\mathcal{R}\cong\overline{\operatorname{Ham}}^{\operatorname{ab}}(\Sigma)

for every open surface Σ𝒮op\Sigma\in\mathcal{S}_{\operatorname{op}}.

Despite significant interest, the algebraic structure of \mathcal{R} remains rather mysterious. Most of what is known follows from the existence of a certain sequence of quasimorphisms

cd:Ham¯(𝔻)c_{d}:\overline{\operatorname{Ham}}({\mathbb{D}})\rightarrow{\mathbb{R}}

called link spectral invariants; see [11, 14]. Consider the space {\mathbb{R}}^{\mathbb{N}} of real valued sequences and let NN\subset{\mathbb{R}}^{\mathbb{N}} denote the subspace of sequences converging to zero. Then the link spectral invariants induce a map

c:Ham¯(𝔻)/Nφ[c1(φ),c2(φ),c3(φ),].c:\overline{\operatorname{Ham}}({\mathbb{D}})\rightarrow{\mathbb{R}}^{\mathbb{N}}/N\qquad\varphi\mapsto[c_{1}(\varphi),c_{2}(\varphi),c_{3}(\varphi),\dots].

Since the sequence of defects of the quasimorphisms cdc_{d} behaves like O(d1)O(d^{-1}) as dd goes to infinity, this map cc is a group homomorphism. Moreover, it fits into the following commutative diagram:

Ham¯(𝔻){\overline{\operatorname{Ham}}({\mathbb{D}})}Ham¯ab(𝔻){\overline{\operatorname{Ham}}^{\operatorname{ab}}({\mathbb{D}})}/N{{\mathbb{R}}^{\mathbb{N}}/N}Ham(𝔻){\operatorname{Ham}({\mathbb{D}})}Hamab(𝔻){\operatorname{Ham}^{\operatorname{ab}}({\mathbb{D}})}{{\mathbb{R}}}c\scriptstyle{c}Cal𝔻\scriptstyle{\operatorname{Cal}_{\mathbb{D}}}\scriptstyle{\cong}Δ\scriptstyle{\Delta}

Here the homomorphism Δ\Delta maps xx\in{\mathbb{R}} to the equivalence class of the constant sequence (x)n(x)_{n\in{\mathbb{N}}}. Commutativity of this diagram is equivalent to the important Weyl law satisfied by the link spectral invariants, which says that they asymptotically recover the Calabi invariant.

It follows from this commutative diagram that the natural map

Hamab(𝔻)Ham¯ab(𝔻)\operatorname{Ham}^{\operatorname{ab}}({\mathbb{D}})\rightarrow\overline{\operatorname{Ham}}^{\operatorname{ab}}({\mathbb{D}})

is injective. By Theorem 3.1 the same is then true for any other open surface Σ\Sigma as well. As discussed in the introduction, after identifying Hamab(𝔻)\operatorname{Ham}^{\operatorname{ab}}({\mathbb{D}})\cong{\mathbb{R}} via the Calabi homomorphism, we can therefore regard

{\mathbb{R}}\subset\mathcal{R}

as a subgroup. It is proven in [14, Proposition 5.3] that the homomorphism cc is surjective. In particular, this implies that {\mathbb{R}} is a proper subgroup of \mathcal{R}.

Let Σ𝒮op\Sigma\in\mathcal{S}_{\operatorname{op}} be an open surface. The restriction of the natural map

Cal¯Σ:Ham¯(Σ)Ham¯ab(Σ).\overline{\mathrm{Cal}}_{\Sigma}:\overline{\operatorname{Ham}}(\Sigma)\rightarrow\overline{\operatorname{Ham}}^{\operatorname{ab}}(\Sigma)\cong\mathcal{R}.

to Ham(Σ)\operatorname{Ham}(\Sigma) agrees with the usual Calabi homomorphism via the natural inclusion {\mathbb{R}}\subset\mathcal{R}. We refer to this map as the universal \mathcal{R}-valued extension of Calabi. If Σ\Sigma is not connected, we define Cal¯Σ:Ham¯(Σ)\overline{\mathrm{Cal}}_{\Sigma}:\overline{\operatorname{Ham}}(\Sigma)\rightarrow\mathcal{R} to be the sum of all Cal¯S\overline{\mathrm{Cal}}_{S} where SS ranges over the components of Σ\Sigma. In view of Theorem 3.1, it is clear that we have the naturality property

Cal¯Σ1=Cal¯Σ2Ham¯(ι)\overline{\mathrm{{Cal}}}_{\Sigma_{1}}=\overline{\mathrm{{Cal}}}_{\Sigma_{2}}\circ\overline{\operatorname{Ham}}(\iota) (3.1)

for any area-preserving topological embedding ι:Σ1Σ2\iota:\Sigma_{1}\hookrightarrow\Sigma_{2}.

Every projection p:p:\mathcal{R}\rightarrow{\mathbb{R}} gives rise to a real-valued extension pCal¯Σp\circ\overline{\operatorname{Cal}}_{\Sigma} of the Calabi homomorphism CalΣ\operatorname{Cal}_{\Sigma} for any open surface Σ\Sigma. Conversely, every real-valued Calabi extension arises this way. Clearly, the real-valued Calabi extensions arising from a single choice of projection p:p:\mathcal{R}\rightarrow{\mathbb{R}} satisfy the naturality condition (3.1).

Projections p:p:\mathcal{R}\rightarrow{\mathbb{R}} can be obtained as follows: The homomorphism cc descends to a surjective homomorphism c:/Nc:\mathcal{R}\rightarrow{\mathbb{R}}^{\mathbb{N}}/N. Viewing {\mathbb{R}} as a subgroup of /N{\mathbb{R}}^{\mathbb{N}}/N via the inclusion Δ\Delta, this is actually a homomorphism over {\mathbb{R}}. Composing cc with any projection /N{\mathbb{R}}^{\mathbb{N}}/N\rightarrow{\mathbb{R}} yields a projection p:p:\mathcal{R}\rightarrow{\mathbb{R}}. There exist infinitely many projections /N{\mathbb{R}}^{\mathbb{N}}/N\rightarrow{\mathbb{R}}, but the choice of any such projection requires the axiom of choice.

Finally, note that one interesting consequence of the discussion here is that having a {\mathbb{R}}-valued extension of the Calabi homomorphism to Ham¯(Σ)\overline{\operatorname{Ham}}(\Sigma) for one single open surface Σ\Sigma, for example Σ=𝔻\Sigma={\mathbb{D}}, implies the existence of such extensions for all open surfaces. The fact that Calabi admits such extensions was proven for surfaces obtained as the interior of a compact surface with non-empty boundary in [11] in the genus zero case and by Mak–Trifa in [34] in the case of arbitrary genus. Here we obtain extensions in full generality, even for surfaces of infinite area.

Proof of Theorem 3.1.

The proof proceeds in several steps.

Step 1 (independence of ι\iota): We show that Ham¯ab(ι)\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota) is independent of ι\iota. Consider two area-preserving topological embeddings ι,ι:Σ1Σ2\iota,\iota^{\prime}:\Sigma_{1}\hookrightarrow\Sigma_{2}. Let φHam¯(Σ1)\varphi\in\overline{\operatorname{Ham}}(\Sigma_{1}). We need to show that [ιφ]=[ιφ][\iota_{*}\varphi]=[\iota_{*}^{\prime}\varphi] in Ham¯ab(Σ2)\overline{\operatorname{Ham}}^{\operatorname{ab}}(\Sigma_{2}). By Proposition 2.1, we may pick a finite collection of relatively compact open discs D1,,DnΣ1D_{1},\dots,D_{n}\Subset\Sigma_{1} and Hamiltonian homeomorphisms φiHam¯(Di)\varphi_{i}\in\overline{\operatorname{Ham}}(D_{i}) such that φ=φ1φn\varphi=\varphi_{1}\circ\cdots\circ\varphi_{n}. Since Σ2\Sigma_{2} is connected, we can also find Hamiltonian homeomorphisms αiHam¯(Σ2)\alpha_{i}\in\overline{\operatorname{Ham}}(\Sigma_{2}) such that αiι|Di=ι|Di\alpha_{i}\circ\iota|_{D_{i}}=\iota^{\prime}|_{D_{i}}. In Ham¯ab(Σ2)\overline{\operatorname{Ham}}^{\operatorname{ab}}(\Sigma_{2}) we can then compute

[ιφ]=[ιφ1ιφn]=[α1(ιφ1)α11αn(ιφn)αn1]=[ιφ1ιφn]=[ιφ],[\iota_{*}^{\prime}\varphi]=[\iota_{*}^{\prime}\varphi_{1}\circ\cdots\circ\iota_{*}^{\prime}\varphi_{n}]=[\alpha_{1}(\iota_{*}\varphi_{1})\alpha_{1}^{-1}\circ\cdots\circ\alpha_{n}(\iota_{*}\varphi_{n})\alpha_{n}^{-1}]=[\iota_{*}\varphi_{1}\circ\cdots\circ\iota_{*}\varphi_{n}]=[\iota_{*}\varphi],

as desired.

Step 2 (surjectivity): We show that Ham¯ab(ι)\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota) is surjective. By Step 1, we may replace ι\iota by a smooth area-preserving embedding. After identifying Σ1\Sigma_{1} with its image under ι\iota, we can regard Σ1\Sigma_{1} as an open subset of Σ2\Sigma_{2}. Given an arbitrary element φHam¯(Σ2)\varphi\in\overline{\operatorname{Ham}}(\Sigma_{2}), we need to show that the class [φ]Ham¯ab(Σ2)[\varphi]\in\overline{\operatorname{Ham}}^{\operatorname{ab}}(\Sigma_{2}) has a representative in Ham¯(Σ1)\overline{\operatorname{Ham}}(\Sigma_{1}). By Proposition 2.1, there exist finitely many relatively compact open discs D1,,DnΣ2D_{1},\dots,D_{n}\Subset\Sigma_{2}, each sufficiently small such that there exists αiHam(Σ2)\alpha_{i}\in\operatorname{Ham}(\Sigma_{2}) such that αi(Di)Σ1\alpha_{i}(D_{i})\Subset\Sigma_{1}, and Hamiltonian homeomorphisms φiHam¯(Di)\varphi_{i}\in\overline{\operatorname{Ham}}(D_{i}) such that φ=φ1φn\varphi=\varphi_{1}\circ\cdots\varphi_{n}. Set ψiαiφiαi1Ham¯(Σ1)\psi_{i}\coloneqq\alpha_{i}\varphi_{i}\alpha_{i}^{-1}\in\overline{\operatorname{Ham}}(\Sigma_{1}). Observe that in Ham¯ab(Σ2)\overline{\operatorname{Ham}}^{\operatorname{ab}}(\Sigma_{2}) we have the identity

[φ]=[φ1φn]=[ψ1ψn].[\varphi]=[\varphi_{1}\circ\cdots\circ\varphi_{n}]=[\psi_{1}\circ\cdots\circ\psi_{n}].

Therefore, ψψ1ψnHam¯(Σ1)\psi\coloneqq\psi_{1}\circ\cdots\circ\psi_{n}\in\overline{\operatorname{Ham}}(\Sigma_{1}) is the desired representative of [φ][\varphi].

Step 3 (injectivity in the case of open surfaces): Assume that both Σ1\Sigma_{1} and Σ2\Sigma_{2} are open. We show that Ham¯ab(ι)\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota) is injective. As in Step 2, we may assume that ι\iota is smooth. We identify Σ1\Sigma_{1} with its image under ι\iota and view it as an open subset of Σ2\Sigma_{2}. Our goal is to construct an inverse

F:Ham¯ab(Σ2)Ham¯ab(Σ1)F:\overline{\operatorname{Ham}}^{\operatorname{ab}}(\Sigma_{2})\rightarrow\overline{\operatorname{Ham}}^{\operatorname{ab}}(\Sigma_{1})

of Ham¯ab(ι)\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota). By the universal property of abelianization, defining a group homomorphism F:Ham¯ab(Σ2)Ham¯ab(Σ1)F:\overline{\operatorname{Ham}}^{\operatorname{ab}}(\Sigma_{2})\rightarrow\overline{\operatorname{Ham}}^{\operatorname{ab}}(\Sigma_{1}) is equivalent to defining a group homomorphism F~:Ham¯(Σ2)Ham¯ab(Σ1)\tilde{F}:\overline{\operatorname{Ham}}(\Sigma_{2})\rightarrow\overline{\operatorname{Ham}}^{\operatorname{ab}}(\Sigma_{1}). Given φHam¯(Σ2)\varphi\in\overline{\operatorname{Ham}}(\Sigma_{2}), we define F~(φ)\tilde{F}(\varphi) as follows. By Proposition 2.1, we may pick a fragmentation φ=φnφ1\varphi=\varphi_{n}\circ\cdots\circ\varphi_{1} such that φiHam¯(Di)\varphi_{i}\in\overline{\operatorname{Ham}}(D_{i}) for some open disc DiΣ2D_{i}\Subset\Sigma_{2} which is sufficiently small such that there exists an area-preserving embedding ιi:DiΣ1\iota_{i}:D_{i}\hookrightarrow\Sigma_{1}. We set

F~(φ)i=1nHam¯ab(ιi)([φi])=i=1n[(ιi)φi]Ham¯ab(Σ1).\tilde{F}(\varphi)\coloneqq\prod\limits_{i=1}^{n}\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota_{i})([\varphi_{i}])=\prod\limits_{i=1}^{n}[(\iota_{i})_{*}\varphi_{i}]\in\overline{\operatorname{Ham}}^{\operatorname{ab}}(\Sigma_{1}).

Since Ham¯ab(Σ1)\overline{\operatorname{Ham}}^{\operatorname{ab}}(\Sigma_{1}) is abelian, the order of the product of course does not matter. We need to check that this definition of F~\tilde{F} is independent of the choice of fragmentation of φ\varphi. Postponing the proof of this claim, let us first argue that F~\tilde{F} indeed induces a homomorphism FF which is an inverse of Ham¯ab(ι)\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota). Observe that F~\tilde{F} is a group homomorphism because given fragmentations of φ,ψHam¯(Σ2)\varphi,\psi\in\overline{\operatorname{Ham}}(\Sigma_{2}), one can obtain a fragmentation of φψ\varphi\circ\psi by concatenation. Consider φHam¯(Σ1)\varphi\in\overline{\operatorname{Ham}}(\Sigma_{1}). We may pick a fragmentation φ=φnφ1\varphi=\varphi_{n}\circ\cdots\circ\varphi_{1} such that all the discs DiD_{i} are contained in Σ1\Sigma_{1}. We can then take the embeddings ιi:DiΣ1\iota_{i}:D_{i}\hookrightarrow\Sigma_{1} to be inclusions. With these choices, we see that FHam¯ab(ι)([φ])=[φ]F\circ\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota)([\varphi])=[\varphi]. Since we already know that Ham¯ab(ι)\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota) is surjective, this implies that Ham¯ab(ι)\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota) is an isomorphism and that FF is its inverse.

Showing that the definition of F~\tilde{F} is independent of choices reduces to verifying the following claim:

Let D1,,DnΣ2D_{1},\dots,D_{n}\Subset\Sigma_{2} be open discs which are small enough such that they admit area-preserving embeddings ιi:DiΣ1\iota_{i}:D_{i}\hookrightarrow\Sigma_{1} into Σ1\Sigma_{1}. Let φiHam¯(Di)\varphi_{i}\in\overline{\operatorname{Ham}}(D_{i}) be Hamiltonian homeomorphisms. Then

φnφ1=idiHam¯ab(ιi)([φi])=id.\varphi_{n}\circ\cdots\circ\varphi_{1}=\operatorname{id}\qquad\Longrightarrow\qquad\prod_{i}\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota_{i})([\varphi_{i}])=\operatorname{id}.

Fix discs DiD_{i} and Hamiltonian homeomorphisms φiHam¯(Di)\varphi_{i}\in\overline{\operatorname{Ham}}(D_{i}) as in the claim and assume that φnφ1=id\varphi_{n}\circ\cdots\circ\varphi_{1}=\operatorname{id}. Set φi(0)φi\varphi_{i}^{(0)}\coloneqq\varphi_{i} for 1in1\leq i\leq n. Our strategy is to construct, for 1i,jn1\leq i,j\leq n, Hamiltonian homeomorphisms φi(j)Ham¯(Di)\varphi_{i}^{(j)}\in\overline{\operatorname{Ham}}(D_{i}) such that the following properties are satisfied:

  1. 1.

    φn(j)φ1(j)=id\varphi_{n}^{(j)}\circ\cdots\circ\varphi_{1}^{(j)}=\operatorname{id} for all jj.

  2. 2.

    iHam¯ab(ιi)([φi(j)])=iHam¯ab(ιi)([φi(j1)])\prod_{i}\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota_{i})([\varphi_{i}^{(j)}])=\prod_{i}\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota_{i})([\varphi_{i}^{(j-1)}]) for all jj.

  3. 3.

    φi(j)Ham(Di)\varphi_{i}^{(j)}\in\operatorname{Ham}(D_{i}) is smooth for all iji\leq j.

Once we have such homeomorphisms φi(j)\varphi_{i}^{(j)}, the desired claim follows from the injectivity of Hamab(ι):Hamab(Σ1)Hamab(Σ2)\operatorname{Ham}^{\operatorname{ab}}(\iota):\operatorname{Ham}^{\operatorname{ab}}(\Sigma_{1})\rightarrow\operatorname{Ham}^{\operatorname{ab}}(\Sigma_{2}). In order to see this, note that by Step 1 we may assume that the embeddings ιi:DiΣ1\iota_{i}:D_{i}\hookrightarrow\Sigma_{1} are smooth. By property 3, all φi(n)\varphi_{i}^{(n)} are smooth as well. We can therefore compute

Hamab(ι)(iHamab(ιi)([φi(n)]))=i[φi(n)]=id,\operatorname{Ham}^{\operatorname{ab}}(\iota)\left(\prod_{i}\operatorname{Ham}^{\operatorname{ab}}(\iota_{i})([\varphi_{i}^{(n)}])\right)=\prod_{i}[\varphi_{i}^{(n)}]=\operatorname{id},

where we use that Hamab(κ)\operatorname{Ham}^{\operatorname{ab}}(\kappa) is independent of the embedding κ\kappa. By injectivity of Hamab(ι)\operatorname{Ham}^{\operatorname{ab}}(\iota), this means that iHamab(ιi)([φi(n)])\prod_{i}\operatorname{Ham}^{\operatorname{ab}}(\iota_{i})([\varphi_{i}^{(n)}]) is equal to the identity in Hamab(Σ1)\operatorname{Ham}^{\operatorname{ab}}(\Sigma_{1}). But since iHam¯ab(ιi)([φi(n)])\prod_{i}\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota_{i})([\varphi_{i}^{(n)}]) is the image of iHamab(ιi)([φi(n)])\prod_{i}\operatorname{Ham}^{\operatorname{ab}}(\iota_{i})([\varphi_{i}^{(n)}]) in Ham¯ab(Σ1)\overline{\operatorname{Ham}}^{\operatorname{ab}}(\Sigma_{1}), we see that iHam¯ab(ιi)([φi(n)])\prod_{i}\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota_{i})([\varphi_{i}^{(n)}]) is equal to the identity as well. It is then immediate from property 2 that iHam¯ab(ιi)([φi])\prod_{i}\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota_{i})([\varphi_{i}]) is equal to the identity, as desired.

It remains to construct the homeomorphisms φi(j)\varphi_{i}^{(j)} satisfying properties 1, 2, and 3. Fix 1jn1\leq j\leq n and assume that the homeomorphisms φi(k)\varphi_{i}^{(k)} have already been constructed for k<jk<j. We explain how to construct φi(j)\varphi_{i}^{(j)}. For 1ij11\leq i\leq j-1, we simply set φi(j)φi(j1)\varphi_{i}^{(j)}\coloneqq\varphi_{i}^{(j-1)}. Note that all of these homeomorphisms are smooth. In order to simplify notation, we set

ψiφi(j1)for jin andψφj1(j1)φ1(j1).\psi_{i}\coloneqq\varphi_{i}^{(j-1)}\quad\text{for $j\leq i\leq n$ and}\quad\psi\coloneqq\varphi_{j-1}^{(j-1)}\circ\cdots\circ\varphi_{1}^{(j-1)}.

Note that ψ\psi is smooth and that

ψnψjψ=id.\psi_{n}\circ\cdots\circ\psi_{j}\circ\psi=\operatorname{id}. (3.2)

Set

Kisupp(ψiψj+1)for j+1in.K_{i}\coloneqq\operatorname{supp}(\psi_{i}\circ\cdots\circ\psi_{j+1})\qquad\text{for $j+1\leq i\leq n$}.

It follows from smoothness of ψ\psi and identity (3.2) that ψnψj+1ψj\psi_{n}\circ\cdots\circ\psi_{j+1}\circ\psi_{j} is smooth. As a consequence, any point at which ψj\psi_{j} is not smooth must be contained in supp(ψj)ψj1(Kn)\operatorname{supp}(\psi_{j})\cap\psi_{j}^{-1}(K_{n}). By Proposition 2.2, we can find ψjsmHam(Dj)\psi_{j}^{\operatorname{sm}}\in\operatorname{Ham}(D_{j}) which agrees with ψj\psi_{j} outside an arbitrarily small neighborhood of supp(ψj)ψj1(Kn)\operatorname{supp}(\psi_{j})\cap\psi_{j}^{-1}(K_{n}) and is arbitrarily C0C^{0} close to ψj\psi_{j}. Here we use that because DjD_{j} is a disc, every compactly supported area-preserving diffeomorphism of DjD_{j} is automatically contained in Ham(Dj)\operatorname{Ham}(D_{j}). We can then write ψj=αψjsm\psi_{j}=\alpha\circ\psi_{j}^{\operatorname{sm}}, where α\alpha is an area-preserving homeomorphism which is supported in an arbitrarily small neighborhood of supp(ψj)Kn\operatorname{supp}(\psi_{j})\cap K_{n} and is arbitrarily C0C^{0} close to the identity. Since supp(ψj)Kn\operatorname{supp}(\psi_{j})\cap K_{n} is contained in Dj(Dj+1Dn)D_{j}\cap(D_{j+1}\cup\cdots\cup D_{n}), we can assume that αHomeo0(Dj(Dj+1Dn),ω2)\alpha\in\operatorname{Homeo}_{0}(D_{j}\cap(D_{j+1}\cup\cdots\cup D_{n}),\omega_{2}). We claim that we can in addition assume that αHam¯(Dj(Dj+1Dn))\alpha\in\overline{\operatorname{Ham}}(D_{j}\cap(D_{j+1}\cup\cdots\cup D_{n})). Indeed, if this is not the case, simply pick βDiff0(Dj(Dj+1Dn),ω2)\beta\in\operatorname{Diff}_{0}(D_{j}\cap(D_{j+1}\cup\cdots\cup D_{n}),\omega_{2}) close to the identity such that the mass flow homomorphism agrees on α\alpha and β\beta. Then replace α\alpha and ψjsm\psi_{j}^{\operatorname{sm}} by αβ1\alpha\circ\beta^{-1} and βψjsm\beta\circ\psi_{j}^{\operatorname{sm}}, respectively. To summarize, we have constructed a factorization

ψj=αψjsmwith ψjsmHam(Dj) and αHam¯(Dj(Dj+1Dn)).\psi_{j}=\alpha\circ\psi_{j}^{\operatorname{sm}}\quad\text{with $\psi_{j}^{\operatorname{sm}}\in\operatorname{Ham}(D_{j})$ and $\alpha\in\overline{\operatorname{Ham}}(D_{j}\cap(D_{j+1}\cup\dots\cup D_{n}))$}. (3.3)

Moreover, we can take α\alpha to be arbitrarily close to the identity. Now set

UiDj(DiKi1)for j+1<in andUj+1DjDj+1.U_{i}\coloneqq D_{j}\cap(D_{i}\setminus K_{i-1})\quad\text{for $j+1<i\leq n$ and}\quad U_{j+1}\coloneqq D_{j}\cap D_{j+1}.

Observe that

Uj+1Un=Dj(Dj+1Dn).U_{j+1}\cup\cdots\cup U_{n}=D_{j}\cap(D_{j+1}\cup\cdots\cup D_{n}).

Since αHam¯(Dj(Dj+1Dn))\alpha\in\overline{\operatorname{Ham}}(D_{j}\cap(D_{j+1}\cup\dots\cup D_{n})) is close to the identity, Proposition 2.1 allows us to fragment

α=αnαj+1with αiHam¯(Ui).\alpha=\alpha_{n}\circ\cdots\circ\alpha_{j+1}\qquad\text{with $\alpha_{i}\in\overline{\operatorname{Ham}}(U_{i})$.} (3.4)

Using that UiU_{i} is disjoint from Ki1K_{i-1} by definition, we obtain

ψnψj+1α=(ψnαn)(ψj+1αj+1).\psi_{n}\circ\cdots\circ\psi_{j+1}\circ\alpha=(\psi_{n}\circ\alpha_{n})\circ\cdots\circ(\psi_{j+1}\circ\alpha_{j+1}). (3.5)

We set

φi(j)ψiαifor j+1in andφj(j)ψjsm.\varphi_{i}^{(j)}\coloneqq\psi_{i}\circ\alpha_{i}\quad\text{for $j+1\leq i\leq n$ and}\quad\varphi_{j}^{(j)}\coloneqq\psi_{j}^{\operatorname{sm}}.

This concludes the construction of the Hamiltonian homeomorphisms φi(j)\varphi_{i}^{(j)} and we need to check that they satisfy properties 1, 2, and 3. First, note that φi(j)Ham¯(Di)\varphi_{i}^{(j)}\in\overline{\operatorname{Ham}}(D_{i}) for all ii and φi(j)Ham(Di)\varphi_{i}^{(j)}\in\operatorname{Ham}(D_{i}) for all iji\leq j by construction. In particular, property 3 is satisfied. It is immediate from (3.2), (3.3), and (3.5) that property 1 holds. It remains to show property 2. Since φi(j)=φi(j1)\varphi_{i}^{(j)}=\varphi_{i}^{(j-1)} for i<ji<j, we need to check that

i=jnHam¯ab(ιi)([ψi])=Ham¯ab(ιj)([ψjsm])i=j+1nHam¯ab(ιi)([ψiαi]).\prod\limits_{i=j}^{n}\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota_{i})([\psi_{i}])=\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota_{j})([\psi_{j}^{\operatorname{sm}}])\cdot\prod\limits_{i=j+1}^{n}\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota_{i})([\psi_{i}\circ\alpha_{i}]).

Substituting Ham¯ab(ιi)([ψiαi])=Ham¯ab(ιi)([ψi])Ham¯ab(ιi)([αi])\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota_{i})([\psi_{i}\circ\alpha_{i}])=\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota_{i})([\psi_{i}])\cdot\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota_{i})([\alpha_{i}]), we see that this is equivalent to showing thatv

Ham¯ab(ιj)([ψj])=Ham¯ab(ιj)([ψjsm])i=j+1nHam¯ab(ιi)([αi]).\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota_{j})([\psi_{j}])=\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota_{j})([\psi_{j}^{\operatorname{sm}}])\cdot\prod\limits_{i=j+1}^{n}\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota_{i})([\alpha_{i}]).

It follows from identities (3.3) and (3.4) that

Ham¯ab(ιj)([ψj])=Ham¯ab(ιj)([ψjsm])i=j+1nHam¯ab(ιj)([αi]).\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota_{j})([\psi_{j}])=\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota_{j})([\psi_{j}^{\operatorname{sm}}])\cdot\prod\limits_{i=j+1}^{n}\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota_{j})([\alpha_{i}]).

Thus it suffices to prove

Ham¯ab(ιi)([αi])=Ham¯ab(ιj)([αi])for all j+1in.\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota_{i})([\alpha_{i}])=\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota_{j})([\alpha_{i}])\qquad\text{for all $j+1\leq i\leq n$.}

In order to see this, simply observe that, for every ii, there exists βHam¯(Σ1)\beta\in\overline{\operatorname{Ham}}(\Sigma_{1}) such that (ιi)(αi)=β(ιj)(αi)β1(\iota_{i})_{*}(\alpha_{i})=\beta\circ(\iota_{j})_{*}(\alpha_{i})\circ\beta^{-1}. This concludes the proof that the homeomorphisms φi(j)\varphi_{i}^{(j)} satisfy property 2 and thus the proof that the definition of the map F~\tilde{F} is independent of choices.

Step 4 (embeddings of open into closed surfaces): Now consider an embedding ι:Σ1Σ2\iota:\Sigma_{1}\hookrightarrow\Sigma_{2} of an open surface Σ1\Sigma_{1} into a closed surface Σ2\Sigma_{2}. Again, we assume that ι\iota is smooth and identify Σ1\Sigma_{1} with its image under ι\iota. Since the group of Hamiltonian diffeomorphisms Ham(Σ2)\operatorname{Ham}(\Sigma_{2}) of the closed surface Σ2\Sigma_{2} is perfect, the image of Hamab(Σ1)\operatorname{Ham}^{\operatorname{ab}}(\Sigma_{1}) in Ham¯ab(Σ1)\overline{\operatorname{Ham}}^{\operatorname{ab}}(\Sigma_{1}) is contained in the kernel of Ham¯ab(ι)\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota). Therefore, Ham¯ab(ι)\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota) induces a homomorphism

G:Ham¯ab(Σ1)/Hamab(Σ1)Ham¯ab(Σ2).G:\overline{\operatorname{Ham}}^{\operatorname{ab}}(\Sigma_{1})/\operatorname{Ham}^{\operatorname{ab}}(\Sigma_{1})\rightarrow\overline{\operatorname{Ham}}^{\operatorname{ab}}(\Sigma_{2}).

Our goal is to show that this is an isomorphism. Again, the strategy is to construct an inverse. Our construction is very similar to the one in Step 3. Given φHam¯(Σ2)\varphi\in\overline{\operatorname{Ham}}(\Sigma_{2}), we pick a fragmentation φ=φnφ1\varphi=\varphi_{n}\circ\cdots\circ\varphi_{1} with φiHam¯(Di)\varphi_{i}\in\overline{\operatorname{Ham}}(D_{i}) for open discs DiΣ2D_{i}\Subset\Sigma_{2} which are sufficiently small such that they admit embeddings ιi:DiΣ1\iota_{i}:D_{i}\hookrightarrow\Sigma_{1}. Given these choices, we set

F~(φ)[iHam¯ab(ιi)([φi])]Ham¯ab(Σ1)/Hamab(Σ1).\tilde{F}(\varphi)\coloneqq\left[\prod_{i}\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota_{i})([\varphi_{i}])\right]\in\overline{\operatorname{Ham}}^{\operatorname{ab}}(\Sigma_{1})/\operatorname{Ham}^{\operatorname{ab}}(\Sigma_{1}).

Again, the main difficulty is to show that this definition does not depend on the choice of fragmentation. Once we know this, we can argue as in Step 3 that F~\tilde{F} descends to a homomorphism

F:Ham¯ab(Σ2)Ham¯ab(Σ1)/Hamab(Σ1)F:\overline{\operatorname{Ham}}^{\operatorname{ab}}(\Sigma_{2})\rightarrow\overline{\operatorname{Ham}}^{\operatorname{ab}}(\Sigma_{1})/\operatorname{Ham}^{\operatorname{ab}}(\Sigma_{1})

which is an inverse to GG. In order to show that F~\tilde{F} is independent of the choice of fragmentation, we have to show that if DiΣ2D_{i}\Subset\Sigma_{2} are open discs admitting embeddings ιi:DiΣ1\iota_{i}:D_{i}\hookrightarrow\Sigma_{1}, which we can assume to be smooth, and if φiHam¯(Di)\varphi_{i}\in\overline{\operatorname{Ham}}(D_{i}) are Hamiltonian homeomorphisms such that φnφ1=id\varphi_{n}\circ\cdots\circ\varphi_{1}=\operatorname{id}, then [iHam¯ab(ιi)([φi])]=id\left[\prod_{i}\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota_{i})([\varphi_{i}])\right]=\operatorname{id} in the quotient Ham¯ab(Σ1)/Hamab(Σ1)\overline{\operatorname{Ham}}^{\operatorname{ab}}(\Sigma_{1})/\operatorname{Ham}^{\operatorname{ab}}(\Sigma_{1}). Going through the argument in Step 3 word by word, we see that there exist Hamiltonian diffeomorphisms φi(n)Ham(Di)\varphi_{i}^{(n)}\in\operatorname{Ham}(D_{i}) such that φn(n)φ1(n)=id\varphi_{n}^{(n)}\circ\cdots\circ\varphi_{1}^{(n)}=\operatorname{id} and

iHam¯ab(ιi)([φi(n)])=iHam¯ab(ιi)([φi])\prod_{i}\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota_{i})([\varphi_{i}^{(n)}])=\prod_{i}\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota_{i})([\varphi_{i}])

in Ham¯ab(Σ1)\overline{\operatorname{Ham}}^{\operatorname{ab}}(\Sigma_{1}). Since the φi(n)\varphi_{i}^{(n)} are Hamiltonian diffeomorphisms and the embeddings ιi\iota_{i} can assumed to be smooth, the left hand side of this identity is contained in Hamab(Σ1)\operatorname{Ham}^{\operatorname{ab}}(\Sigma_{1}). Thus the right hand side represents the identity in the quotient Ham¯ab(Σ1)/Hamab(Σ1)\overline{\operatorname{Ham}}^{\operatorname{ab}}(\Sigma_{1})/\operatorname{Ham}^{\operatorname{ab}}(\Sigma_{1}), which is exactly what we need to show for concluding that F~\tilde{F} is well-defined and independent of choices. This finishes the proof that the kernel of Ham¯ab(ι)\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota) is given by the image of Hamab(Σ1)\operatorname{Ham}^{\operatorname{ab}}(\Sigma_{1}) in the case of an open surface Σ1\Sigma_{1} embedded into a closed surface Σ2\Sigma_{2}.

Step 5 (two closed surfaces): Any area-preserving embedding ι\iota of a closed connected surface into another is necessarily an area-preserving homeomorphism, which clearly implies that Ham¯ab(ι)\overline{\operatorname{Ham}}^{\operatorname{ab}}(\iota) is an isomorphism. This concludes the proof of the theorem. ∎

4 Smooth Hamiltonian structures

In this section, we introduce smooth Hamiltonian structures and discuss basic notions such as flux, helicity, and plugs.

4.1 Basic definitions

Let WW be an oriented smooth 33-manifold without boundary and possibly open. A Hamiltonian structure on WW is a closed, maximally nondegenerate 22-form ω\omega on WW.

The line bundle kerωTW\operatorname{ker}\omega\subset TW is called the characteristic line bundle of ω\omega. The foliation generated by this line bundle is called the characteristic foliation of ω\omega. It is naturally cooriented by ω\omega. Together with the ambient orientation of WW, this coorientation induces a natural orientation of the characteristic foliation.

The 22-form ω\omega also induces a transverse measure on the characteristic foliation. Locally on every smooth transversal, the induced measure is diffeomorphic to the standard 22-dimensional Lebesgue measure. The (cooriented and measured) characteristic foliation uniquely determines the 22-form ω\omega. We can therefore equivalently think of a Hamiltonian structure as a 11-dimensional cooriented and measured smooth foliation with the property that the induced measure on transversals is locally diffeomorphic to the 22-dimensional Lebesgue measure.

A diffeomorphism of Hamiltonian structures (W0,ω0)(W_{0},\omega_{0}) and (W1,ω1)(W_{1},\omega_{1}) is a diffeomorphism f:W0W1f:W_{0}\rightarrow W_{1} which satisfies fω1=ω0f^{*}\omega_{1}=\omega_{0}.

Example 4.1.

Consider 3=×2{\mathbb{R}}^{3}={\mathbb{R}}\times{\mathbb{R}}^{2} equipped with coordinates (t,x,y)(t,x,y) and oriented with respect to these coordinates.

We will refer to the closed, maximally nondegenerate 22-form ωstd=dxdy\omega_{\operatorname{std}}=dx\wedge dy as the standard Hamiltonian structure on 3{\mathbb{R}}^{3}.

For a>0a>0, let B(a)2B(a)\subset{\mathbb{R}}^{2} denote the ball of area aa and consider the product (1,1)×B(a)×2(-1,1)\times B(a)\subset{\mathbb{R}}\times{\mathbb{R}}^{2}. The restriction of the standard Hamiltonian structure on 3{\mathbb{R}}^{3} induces a Hamiltonian structure on (1,1)×B(a)×2(-1,1)\times B(a)\subset{\mathbb{R}}\times{\mathbb{R}}^{2}. We will abbreviate it by the same symbol ωstd\omega_{\operatorname{std}} and refer to it as the standard Hamiltonian structure on (1,1)×B(a)(-1,1)\times B(a).

Example 4.2.

Consider a surface Σ\Sigma, without boundary, equipped with an area form ωΣ\omega_{\Sigma}. Let pr2:(0,1)×ΣΣ\operatorname{pr}_{2}:(0,1)\times\Sigma\rightarrow\Sigma be the projection onto the second factor. We equip (0,1)×Σ(0,1)\times\Sigma with the orientation induced by the volume form dtpr2ωΣdt\wedge\operatorname{pr}_{2}^{*}\omega_{\Sigma}, where tt denotes the coordinate on (0,1)(0,1). Then, pr2ωΣ\operatorname{pr}_{2}^{*}\omega_{\Sigma} is a Hamiltonian structure on (0,1)×Σ(0,1)\times\Sigma for which the oriented leaves of the characteristic foliation are given by the flow lines of the vector field t\partial_{t}. This Hamiltonian structure will be denoted by ((0,1)×Σ,ωΣ)((0,1)\times\Sigma,\omega_{\Sigma}).

Now, let (φt)t[0,1](\varphi^{t})_{t\in[0,1]} be a smooth isotopy in Ham(Σ,ωΣ)\operatorname{Ham}(\Sigma,\omega_{\Sigma}) such that φt=idΣ\varphi^{t}=\operatorname{id}_{\Sigma} for tt close to 0 and φt=φ1\varphi^{t}=\varphi^{1} for tt close to 11. Let H:[0,1]×ΣH:[0,1]\times\Sigma\rightarrow{\mathbb{R}} be the unique compactly supported Hamiltonian generating φt\varphi^{t}. Then, the 22-form

dHdt+ωΣdH\wedge dt+\omega_{\Sigma}

is another Hamiltonian structure on (0,1)×Σ(0,1)\times\Sigma. It agrees with ωΣ\omega_{\Sigma} outside a compact set and the leaves of its characteristic foliation are of the form {(t,φt(p))t(0,1)}\{(t,\varphi^{t}(p))\mid t\in(0,1)\} for pΣp\in\Sigma.

Hamiltonian structures satisfy a version of the Darboux neighborhood theorem which we state here as a lemma.

Lemma 4.3.

Let ω\omega be a Hamiltonian structure on WW. For every point pp in (W,ω)(W,\omega), there exists a neighborhood UU such that (U,ω)(U,\omega) is diffeomorphic to ((1,1)×B(a),ωstd)((-1,1)\times B(a),\omega_{\operatorname{std}}), for some a>0a>0.

Remark 4.4.

For every Hamiltonian structure ω\omega on WW, there exist a symplectic 44-manifold (M,ω)(M,\omega^{\prime}) and an embedding ι:(W,ω)(M,ω)\iota:(W,\omega)\hookrightarrow(M,\omega^{\prime}) such that ιω=ω\iota^{*}\omega^{\prime}=\omega. Moreover, by Gotay’s theorem [24], any two such embeddings ιi:(W,ω)(Mi,ωi),i{1,2}\iota_{i}:(W,\omega)\rightarrow(M_{i},\omega_{i}),\;i\in\{1,2\}, are neighborhood equivalent in the following sense: There exist open neighborhoods ιi(Wi)UiMi\iota_{i}(W_{i})\subset U_{i}\subset M_{i} and a symplectomorphism ϕ:(U1,ω1)(U2,ω2)\phi:(U_{1},\omega_{1})\rightarrow(U_{2},\omega_{2}) such that ϕι1=ι2\phi\circ\iota_{1}=\iota_{2}.

4.2 Flux and Helicity

Let (Y,ω)(Y,\omega) be a closed oriented 33-manifold YY equipped with a Hamiltonian structure ω\omega.

Definition 4.5.

We define the flux of the Hamiltonian structure ω\omega to be the cohomology class

Flux(ω)[ω]H2(Y;).\operatorname{Flux}(\omega)\coloneqq[\omega]\in H^{2}(Y;{\mathbb{R}}).

We say that a Hamiltonian structure ω\omega is exact if Flux(ω)=0\operatorname{Flux}(\omega)=0.

Recall that the helicity of an exact Hamiltonian structure ω\omega is defined to be the quantity

(ω)Yαω,\mathcal{H}(\omega)\coloneqq\int_{Y}\alpha\wedge\omega,

where α\alpha is any primitive of ω\omega. We check in the lemma below that helicity is well-defined.

Lemma 4.6.

Let ω\omega be an exact Hamiltonian structure on YY and let α\alpha and β\beta be two 11-forms such that dα=dβ=ωd\alpha=d\beta=\omega. Then,

Yαω=Yβω.\int_{Y}\alpha\wedge\omega=\int_{Y}\beta\wedge\omega.
Proof.

The 11-form αβ\alpha-\beta is closed and the 22-form ω\omega is exact. Hence, the 33-form (αβ)ω(\alpha-\beta)\wedge\omega is exact and consequently, by Stokes’ theorem, we have

Y(αβ)ω=0.\int_{Y}(\alpha-\beta)\wedge\omega=0.

4.3 Plugs

We introduce the concept of plugs, which can be inserted into a given Hamiltonian structure to generate a new one.

Definition 4.7.

Let ω\omega be a Hamiltonian structure on WW. A plug is a tuple 𝒫=(Σ,ωΣ,α,(φt)t[0,1])\mathcal{P}=(\Sigma,\omega_{\Sigma},\alpha,(\varphi^{t})_{t\in[0,1]}) consisting of

  1. 1.

    an open surface Σ\Sigma, not necessarily connected;

  2. 2.

    an area form ωΣ\omega_{\Sigma} on Σ\Sigma;

  3. 3.

    a smooth embedding of Hamiltonian structures α:((0,1)×Σ,ωΣ)(W,ω)\alpha:((0,1)\times\Sigma,\omega_{\Sigma})\hookrightarrow(W,\omega);

  4. 4.

    a smooth isotopy (φt)t[0,1](\varphi^{t})_{t\in[0,1]} in Ham(Σ)\operatorname{Ham}(\Sigma) such that φt=idΣ\varphi^{t}=\operatorname{id}_{\Sigma} for tt close to 0 and φt=φ1\varphi^{t}=\varphi^{1} for tt close to 11.

We define the Calabi invariant of the plug 𝒫=(Σ,ωΣ,α,(φt)t[0,1])\mathcal{P}=(\Sigma,\omega_{\Sigma},\alpha,(\varphi^{t})_{t\in[0,1]}) to be

Cal(𝒫):=CalΣ(φ1).\mathrm{Cal}(\mathcal{P}):=\mathrm{Cal}_{\Sigma}(\varphi^{1}). (4.1)

Given a plug 𝒫\mathcal{P}, we define a new Hamiltonian structure ω#𝒫\omega\#\mathcal{P} which coincides with ω\omega outside of im(α)\operatorname{im}(\alpha) and with α(dHdt+ωΣ)\alpha_{*}(dH\wedge dt+\omega_{\Sigma}) inside im(α)\operatorname{im}(\alpha), where HH is the unique compactly supported Hamiltonian generating φt\varphi^{t}. Note that ω#𝒫\omega\#\mathcal{P} is well-defined because dHdtdH\wedge dt vanishes near the boundary of im(α)\operatorname{im}(\alpha).

The Hamiltonian structure ω#𝒫\omega\#\mathcal{P} has the following equivalent definition which is better suited for the C0C^{0} setting which we consider below. Define Φ:(0,1)×Σ(0,1)×Σ\Phi:(0,1)\times\Sigma\rightarrow(0,1)\times\Sigma by Φ(t,p)(t,φt(p))\Phi(t,p)\coloneqq(t,\varphi^{t}(p)). Then ω#𝒫\omega\#{\mathcal{{P}}} is defined to agree with ω\omega outside of im(α)\operatorname{im}(\alpha) and with αΦωΣ\alpha_{*}\Phi_{*}\omega_{\Sigma} inside im(α)\operatorname{im}(\alpha). Note that before plug insertion, the characteristic leaves of ω\omega inside im(α)\operatorname{im}(\alpha) are of the form α((0,1)×{p})\alpha((0,1)\times\{p\}) for pΣp\in\Sigma. After plug insertion, the characteristic leaves of ω#𝒫\omega\#\mathcal{P} are of the form

{α(t,φt(p))t(0,1)}\{\alpha(t,\varphi^{t}(p))\mid t\in(0,1)\}

for pΣp\in\Sigma.

Although the Hamiltonian structure ω#𝒫\omega\#\mathcal{P} depends on the entire isotopy φt\varphi^{t}, its diffeomorphism type depends only on the time-1 map φ1\varphi^{1}.

Lemma 4.8.

Consider two plugs 𝒫1=(Σ,ωΣ,α,(φ1t)t[0,1])\mathcal{P}_{1}=(\Sigma,\omega_{\Sigma},\alpha,(\varphi_{1}^{t})_{t\in[0,1]}) and 𝒫2=(Σ,ωΣ,α,(φ2t)t[0,1])\mathcal{P}_{2}=(\Sigma,\omega_{\Sigma},\alpha,(\varphi_{2}^{t})_{t\in[0,1]}). If φ11=φ21\varphi_{1}^{1}=\varphi_{2}^{1}, then there exists a diffeomorphism of Hamiltonian structures ψ:(Y,ω#𝒫1)(Y,ω#𝒫2)\psi:(Y,\omega\#\mathcal{P}_{1})\rightarrow(Y,\omega\#\mathcal{P}_{2}) which is supported inside im(α)\mathrm{im}(\alpha) and is isotopic to the identity through diffeomorphisms supported inside im(α)\mathrm{im}(\alpha).

We omit the proof of the above lemma, as its generalization in the C0C^{0} setting is proven below in Lemma 5.7.

The next two lemmas describe the effect of plugs on flux and helicity. Lemma 4.10 may be viewed as a special case of [20, Théorème 3.1].

Lemma 4.9.

Let ω\omega be a Hamiltonian structure on a closed 33-manifold YY and let 𝒫\mathcal{P} be an ω\omega-plug. Then,

Flux(ω)=Flux(ω#𝒫).\operatorname{Flux}(\omega)=\operatorname{Flux}(\omega\#\mathcal{P}).
Proof.

Write 𝒫=(Σ,ωΣ,α,(φt)t[0,1])\mathcal{P}=(\Sigma,\omega_{\Sigma},\alpha,(\varphi^{t})_{t\in[0,1]}). Let HH be the unique compactly supported Hamiltonian generating φt\varphi^{t}. Then the 11-form α(Hdt)\alpha_{*}(Hdt) is compactly supported in im(α)\operatorname{im}(\alpha) and can hence be extended to a 11-form on YY by setting it equal to zero outside of im(α)\operatorname{im}(\alpha). By definition, the Hamiltonian structures ω\omega and ω#𝒫\omega\#\mathcal{P} differ by the exact 22-form dα(Hdt)d\alpha_{*}(Hdt). This clearly implies that Flux(ω)=Flux(ω#𝒫)\operatorname{Flux}(\omega)=\operatorname{Flux}(\omega\#\mathcal{P}). ∎

Lemma 4.10.

Let YY be a closed 33-manifold and let ω\omega be an exact Hamiltonian structure on YY. Let 𝒫\mathcal{P} be an ω\omega-plug. Then

(ω#𝒫)=(ω)+Cal(𝒫).\mathcal{H}(\omega\#\mathcal{P})=\mathcal{H}(\omega)+\mathrm{Cal}(\mathcal{P}).
Proof.

As in the proof of Lemma 4.9, note that ω\omega and ω#𝒫\omega\#\mathcal{P} differ by the exact 22-form dα(Hdt)d\alpha_{*}(Hdt). Let λ\lambda be an arbitrary primitive 11-form of ω\omega. We compute

(ω#𝒫)\displaystyle\mathcal{H}(\omega\#\mathcal{P}) =Y(λ+α(Hdt))(ω+dα(Hdt))\displaystyle=\int_{Y}(\lambda+\alpha_{*}(Hdt))\wedge(\omega+d\alpha_{*}(Hdt))
=(ω)+(0,1)×ΣH𝑑tωΣ+αλd(Hdt)+Hdtd(Hdt).\displaystyle=\mathcal{H}(\omega)+\int_{(0,1)\times\Sigma}Hdt\wedge\omega_{\Sigma}+\alpha^{*}\lambda\wedge d(Hdt)+Hdt\wedge d(Hdt). (4.2)

Note that Hdtd(Hdt)=HdtdHdtHdt\wedge d(Hdt)=Hdt\wedge dH\wedge dt vanishes identically. Moreover,

d(αλHdt)=ωΣHdtαλd(Hdt).d(\alpha^{*}\lambda\wedge Hdt)=\omega_{\Sigma}\wedge Hdt-\alpha^{*}\lambda\wedge d(Hdt).

By Stokes’ theorem, the integral (0,1)×Σd(αλHdt)\int_{(0,1)\times\Sigma}d(\alpha^{*}\lambda\wedge Hdt) vanishes, which implies that

(0,1)×Σαλd(Hdt)=(0,1)×ΣωΣHdt.\int_{(0,1)\times\Sigma}\alpha^{*}\lambda\wedge d(Hdt)=\int_{(0,1)\times\Sigma}\omega_{\Sigma}\wedge Hdt.

Combining this with (4.2), we see that

(ω#𝒫)=(ω)+2(0,1)×ΣH𝑑tωΣ=(ω)+Cal(φH1)=(ω)+Cal(𝒫).\mathcal{H}(\omega\#\mathcal{P})=\mathcal{H}(\omega)+2\int_{(0,1)\times\Sigma}Hdt\wedge\omega_{\Sigma}=\mathcal{H}(\omega)+\operatorname{Cal}(\varphi_{H}^{1})=\mathcal{H}(\omega)+\operatorname{Cal}(\mathcal{P}).

5 C0C^{0} Hamiltonian structures

In this section, we introduce C0C^{0} Hamiltonian structures and C0C^{0} plugs.

5.1 Basic definitions

Recall the definition of the standard Hamiltonian structure on 3{\mathbb{R}}^{3} from Example 4.1. A C0C^{0} Hamiltonian structure on a topological 33-manifold WW is a 11-dimensional cooriented C0C^{0} foliation, equipped with a transverse measure, which is locally modeled on 3{\mathbb{R}}^{3} with the characteristic foliation induced by ωstd\omega_{\operatorname{std}}. Explicitly, this notion can be formalized in terms of C0C^{0} Hamiltonian atlases.

Definition 5.1.

Consider 3=×2{\mathbb{R}}^{3}={\mathbb{R}}\times{\mathbb{R}}^{2} equipped with the smooth Hamiltonian structure ωstd\omega_{\operatorname{std}}. Let ψ:UV\psi:U\rightarrow V be a homeomorphism between two open subsets of 3{\mathbb{R}}^{3}. We say that ψ\psi preserves the standard Hamiltonian structure ωstd\omega_{\operatorname{std}} on 3{\mathbb{R}}^{3} if, for every point pUp\in U, there exist an open neighborhood of pp of the form I×BI\times B, where II\subset{\mathbb{R}} is an open interval and B2B\subset{\mathbb{R}}^{2} is a ball, and a continuous area- and orientation-preserving embedding a:B2a:B\hookrightarrow{\mathbb{R}}^{2} such that

ψ(t,z)=(,a(z))for all (t,z)I×B.\psi(t,z)=(*,a(z))\qquad\text{for all $(t,z)\in I\times B$.}

A C0C^{0} Hamiltonian atlas on a topological 33-manifold is an atlas whose transition maps are homeomorphisms preserving the standard Hamiltonian structure on 3{\mathbb{R}}^{3}. Two such atlases are considered equivalent if the transition maps between the two atlases preserve the standard Hamiltonian structure on 3{\mathbb{R}}^{3}. A C0C^{0} Hamiltonian structure Ω\Omega on WW is an equivalence class of C0C^{0} Hamiltonian atlases.

Definition 5.2.

A homeomorphism of C0C^{0} Hamiltonian structures ψ:(W0,Ω0)(W1,Ω1)\psi:(W_{0},\Omega_{0})\rightarrow(W_{1},\Omega_{1}) is a homeomorphism ψ:W0W1\psi:W_{0}\rightarrow W_{1} which, in local charts, preserves the standard Hamiltonian structure on 3{\mathbb{R}}^{3}. A topological embedding of C0C^{0} Hamiltonian structures α:(W0,Ω0)(W1,Ω1)\alpha:(W_{0},\Omega_{0})\hookrightarrow(W_{1},\Omega_{1}) is a topological embedding α:W0W1\alpha:W_{0}\hookrightarrow W_{1} which is a homeomorphism of the C0C^{0} Hamiltonian structures (W0,Ω0)(W_{0},\Omega_{0}) and (α(W0),Ω1)(\alpha(W_{0}),\Omega_{1}).

Let ψ:W0W1\psi:W_{0}\rightarrow W_{1} be a homeomorphism. Given a C0C^{0} Hamiltonian structure Ω1\Omega_{1} on W1W_{1}, we can pull back its defining atlas, via ψ\psi, to obtain a C0C^{0} Hamiltonian structure on W0W_{0} which we abbreviate by ψΩ1\psi^{*}\Omega_{1} and refer to as the pullback of Ω1\Omega_{1} via ψ\psi. We define the pushforward of a C0C^{0} Hamiltonian structure Ω0\Omega_{0} on W0W_{0} to be the pullback of Ω0\Omega_{0} via ψ1\psi^{-1}; we abbreviate this by ψΩ0\psi_{*}\Omega_{0}.

Remark 5.3.

A smooth Hamiltonian structure ω\omega on a smooth 33-manifold WW gives rise to a C0C^{0} Hamiltonian structure. Indeed, consider the atlas of WW consisting of all Darboux charts. The transition maps of this atlas are diffeomorphisms ψ:UV\psi:U\rightarrow V between open subsets of 3{\mathbb{R}}^{3} which satisfy ψωstd=ωstd\psi^{*}\omega_{\operatorname{std}}=\omega_{\operatorname{std}} as smooth 22-forms. Note that for a diffeomorphism ψ\psi, this condition is equivalent to preserving the standard Hamiltonian structure ωstd\omega_{\operatorname{std}} on 3{\mathbb{R}}^{3} in the sense of Definition 5.1. Hence our atlas consisting of smooth Darboux charts is a C0C^{0} Hamiltonian atlas.

Remark 5.4.

Let WW be a topological 33-manifold. In what follows, by a smooth Hamiltonian structure ω\omega on WW we mean a choice of smooth structure on WW and a maximally nondegenerate closed 22-form. Alternatively, a smooth Hamiltonian structure on a topological 33-manifold can be specified by an atlas on WW whose transition maps are diffeomorphisms preserving the standard Hamiltonian structure on 3{\mathbb{R}}^{3}.

5.2 C0C^{0} plugs

Analogous to the smooth plugs from Section 4.3, we introduce the concept of C0C^{0} plugs, which can be inserted into C0C^{0} Hamiltonian structures to generate new C0C^{0} Hamiltonian structures. We also define the Calabi invariant of C0C^{0} plugs.

Throughout the rest of this section, let WW denote a topological 33-manifold without boundary and possibly open.

Definition 5.5.

Let Ω\Omega be a C0C^{0} Hamiltonian structure on WW. An Ω\Omega-plug (or simply a plug, if Ω\Omega is clear from the context) is a tuple 𝒫=(Σ,ωΣ,α,(φt)t[0,1])\mathcal{P}=(\Sigma,\omega_{\Sigma},\alpha,(\varphi^{t})_{t\in[0,1]}) consisting of

  1. 1.

    a smooth, open surface Σ\Sigma, not necessarily connected;

  2. 2.

    a smooth area form ωΣ\omega_{\Sigma} on Σ\Sigma;

  3. 3.

    a topological embedding of C0C^{0} Hamiltonian structures α:((0,1)×Σ,ωΣ)(W,Ω)\alpha:((0,1)\times\Sigma,\omega_{\Sigma})\hookrightarrow(W,\Omega);

  4. 4.

    a continuous isotopy (φt)t[0,1](\varphi^{t})_{t\in[0,1]} in Ham¯(Σ,ωΣ)\overline{\operatorname{Ham}}(\Sigma,\omega_{\Sigma}) such that φt=idΣ\varphi^{t}=\operatorname{id}_{\Sigma} for tt close to 0 and φt=φ1\varphi^{t}=\varphi^{1} for tt close to 11.

We define the Calabi invariant of 𝒫\mathcal{P} to be

Cal¯(𝒫)Cal¯Σ(φ1).\overline{\operatorname{Cal}}(\mathcal{P})\coloneqq\overline{\operatorname{Cal}}_{\Sigma}(\varphi^{1})\in\mathcal{R}.

Here, we recall that Cal¯Σ:Ham¯(Σ)\overline{\operatorname{Cal}}_{\Sigma}:\overline{\operatorname{Ham}}(\Sigma)\rightarrow\mathcal{R} denotes the universal \mathcal{R}-valued extension of the Calabi homomorphism; Section 3.

We now describe how an Ω\Omega-plug 𝒫\mathcal{P} can be inserted into Ω\Omega yielding a new C0C^{0} Hamiltonian structure Ω#𝒫\Omega\#{\mathcal{P}}. Define Φ:(0,1)×Σ(0,1)×Σ\Phi:(0,1)\times\Sigma\rightarrow(0,1)\times\Sigma by

Φ(t,p)(t,φt(p)).\Phi(t,p)\coloneqq(t,\varphi^{t}(p)).

Inside im(α)\operatorname{im}(\alpha), we define Ω#𝒫\Omega\#\mathcal{{P}} as the pushforward (αΦ)ωΣ(\alpha\circ\Phi)_{*}\omega_{\Sigma}, where ωΣ\omega_{\Sigma} is regarded as a C0C^{0} Hamiltonian structure on (0,1)×Σ(0,1)\times\Sigma. Note that (αΦ)ωΣ(\alpha\circ\Phi)_{*}\omega_{\Sigma} agrees with Ω\Omega near the boundary of im(α)\operatorname{im}(\alpha). We define Ω#𝒫\Omega\#\mathcal{P} to agree with Ω\Omega outside im(α)\operatorname{im}(\alpha). The effect of plug insertion on the characteristic foliation is as follows: Before plug insertion, the characteristic leaves of Ω\Omega inside im(α)\operatorname{im}(\alpha) are of the form α((0,1)×{p})\alpha((0,1)\times\{p\}) for pΣp\in\Sigma. After plug insertion, the characteristic leaves of ω#𝒫\omega\#\mathcal{P} are of the form

{α(t,φt(p))t(0,1)}\{\alpha(t,\varphi^{t}(p))\mid t\in(0,1)\}

for pΣp\in\Sigma.

The inverse of an Ω\Omega-plug 𝒫\mathcal{P} is defined to be the Ω#𝒫\Omega\#\mathcal{P}-plug

𝒫¯(Σ,ωΣ,αΦ,((φt)1)t[0,1]).\overline{\mathcal{P}}\coloneqq(\Sigma,\omega_{\Sigma},\alpha\circ\Phi,((\varphi^{t})^{-1})_{t\in[0,1]}). (5.1)

Note that if 𝒫\mathcal{P} is an Ω\Omega-plug, then Ω#𝒫#𝒫¯=Ω\Omega\#\mathcal{P}\#\overline{\mathcal{P}}=\Omega and 𝒫¯¯=𝒫\overline{\overline{\mathcal{P}}}=\mathcal{P}. A plug 𝒫\mathcal{P} is called trivial if φt=id\varphi^{t}=\operatorname{id} for all t[0,1]t\in[0,1]. If 𝒫\mathcal{P} is trivial, then Ω#𝒫=Ω\Omega\#\mathcal{P}=\Omega. If ψ:(W,Ω)(W,Ω)\psi:(W^{\prime},\Omega^{\prime})\rightarrow(W,\Omega) is a homeomorphism of C0C^{0} Hamiltonian structures, then the pull back ψ𝒫\psi^{*}\mathcal{P} of the Ω\Omega-plug 𝒫\mathcal{P} via ψ\psi is the Ω\Omega^{\prime}-plug defined by

ψ𝒫(Σ,ωΣ,ψ1α,(φt)t[0,1]).\psi^{*}\mathcal{P}\coloneqq(\Sigma,\omega_{\Sigma},\psi^{-1}\circ\alpha,(\varphi^{t})_{t\in[0,1]}). (5.2)

We observe that ψ\psi is a homeomorphism of C0C^{0} Hamiltonian structures

ψ:(W,Ω#ψ𝒫)(W,Ω#𝒫).\psi:(W^{\prime},\Omega^{\prime}\#\psi^{*}\mathcal{P})\rightarrow(W,\Omega\#\mathcal{P}). (5.3)
Remark 5.6.

Consider an Ω\Omega-plug 𝒫=(Σ,ωΣ,α,(φt)t[0,1])\mathcal{P}=(\Sigma,\omega_{\Sigma},\alpha,(\varphi^{t})_{t\in[0,1]}). It is always possible to slightly shrink the image of 𝒫\mathcal{P} without affecting Ω#𝒫\Omega\#\mathcal{P} or the Calabi invariant Cal¯(𝒫)\overline{\operatorname{Cal}}(\mathcal{P}) as follows: Pick a relatively compact open subset ΣΣ\Sigma^{\prime}\Subset\Sigma such that the entire isotopy φt\varphi^{t} is supported in Σ\Sigma^{\prime}. Let ε>0\varepsilon>0 such that φt\varphi^{t} is equal to the identity for all t(0,2ε]t\in(0,2\varepsilon] and equal to φ1\varphi^{1} for all t[12ε,1)t\in[1-2\varepsilon,1). Let τ:(0,1)(ε,1ε)\tau:(0,1)\rightarrow(\varepsilon,1-\varepsilon) be an orientation-preserving diffeomorphism which agrees with the identity on the interval [2ε,12ε][2\varepsilon,1-2\varepsilon]. Define an embedding of C0C^{0} Hamiltonian structures α:((0,1)×Σ,ωΣ)(W,Ω)\alpha^{\prime}:((0,1)\times\Sigma^{\prime},\omega_{\Sigma})\hookrightarrow(W,\Omega) by setting α(t,p)α(τ(t),p)\alpha^{\prime}(t,p)\coloneqq\alpha(\tau(t),p). Set 𝒫(Σ,ωΣ,α,(φt)t[0,1])\mathcal{P}^{\prime}\coloneqq(\Sigma^{\prime},\omega_{\Sigma},\alpha^{\prime},(\varphi^{t})_{t\in[0,1]}). The closure of the image of 𝒫\mathcal{P}^{\prime} is clearly contained in the image of 𝒫\mathcal{P}. Moreover, Ω#𝒫=Ω#𝒫\Omega\#\mathcal{P}^{\prime}=\Omega\#\mathcal{P}. Finally, by the naturality of the \mathcal{R}-valued extension of Calabi, we have Cal¯(𝒫)=Cal¯(𝒫)\overline{\operatorname{Cal}}(\mathcal{P^{\prime}})=\overline{\operatorname{Cal}}(\mathcal{P}).

Analogously to Lemma 4.8 in the smooth setting, although the C0C^{0} Hamiltonian structure ω#𝒫\omega\#\mathcal{P} depends on the entire isotopy φt\varphi^{t}, its homeomorphism type depends only on the time-1 map φ1\varphi^{1}.

Lemma 5.7.

Let Ω\Omega be a C0C^{0} Hamiltonian structure on WW. For i{0,1}i\in\{0,1\}, let 𝒫i=(Σ,ωΣ,α,(φit)t[0,1])\mathcal{P}_{i}=(\Sigma,\omega_{\Sigma},\alpha,(\varphi_{i}^{t})_{t\in[0,1]}) be two Ω\Omega-plugs. If φ01=φ11\varphi_{0}^{1}=\varphi_{1}^{1}, then there exists a homeomorphism of C0C^{0} Hamiltonian structures ψ:(W,Ω#𝒫0)(W,Ω#𝒫1)\psi:(W,\Omega\#\mathcal{P}_{0})\rightarrow(W,\Omega\#\mathcal{P}_{1}) which is supported inside im(α)\operatorname{im}(\alpha) and isotopic to the identity through homeomorphisms supported in im(α)\operatorname{im}(\alpha).

Proof.

Define Φi:(0,1)×Σ(0,1)×Σ\Phi_{i}:(0,1)\times\Sigma\rightarrow(0,1)\times\Sigma by Φi(t,p)(t,φit(p))\Phi_{i}(t,p)\coloneqq(t,\varphi_{i}^{t}(p)). Then

Φ1Φ01:((0,1)×Σ,(Φ0)ωΣ)((0,1)×Σ,(Φ1)ωΣ)\Phi_{1}\circ\Phi_{0}^{-1}:((0,1)\times\Sigma,(\Phi_{0})_{*}\omega_{\Sigma})\rightarrow((0,1)\times\Sigma,(\Phi_{1})_{*}\omega_{\Sigma})

is a homeomorphism of C0C^{0} Hamiltonian structures. Since φ01=φ11\varphi_{0}^{1}=\varphi_{1}^{1}, it is compactly supported. We can therefore define the desired homeomorphism ψ:(W,Ω#𝒫0)(W,Ω#𝒫1)\psi:(W,\Omega\#\mathcal{P}_{0})\rightarrow(W,\Omega\#\mathcal{P}_{1}) by setting ψαΦ1Φ01α1\psi\coloneqq\alpha\circ\Phi_{1}\circ\Phi_{0}^{-1}\circ\alpha^{-1} inside im(α)\operatorname{im}(\alpha) and extending it to be the identity on the complement of this set.

Now, the surface Σ\Sigma is open and therefore not S2S^{2}. Hence Ham¯(Σ)\overline{\operatorname{Ham}}(\Sigma) is simply connected, as explained in Section 2. It follows from simply connectedness of Ham¯(Σ)\overline{\operatorname{Ham}}(\Sigma) that the homeomorphism ψ\psi is isotopic to the identity via homeomorphisms supported in im(α)\operatorname{im}(\alpha). ∎

The next lemma states that sliding a plug along the characteristic foliation has no effect on the homeomorphism type of the C0C^{0} Hamiltonian structure.

Lemma 5.8.

Let Ω\Omega be a C0C^{0} Hamiltonian structure on WW. For i{0,1}i\in\{0,1\}, let 𝒫i=(Σ,ωΣ,αi,(φt)t[0,1])\mathcal{P}_{i}=(\Sigma,\omega_{\Sigma},\alpha_{i},(\varphi^{t})_{t\in[0,1]}) be two Ω\Omega-plugs. Assume that there exists a continuous family (αs)s[0,1](\alpha_{s})_{s\in[0,1]} of topological embeddings of C0C^{0} Hamiltonian structures αs:((0,1)×Σ,ωΣ)(W,Ω)\alpha_{s}:((0,1)\times\Sigma,\omega_{\Sigma})\hookrightarrow(W,\Omega) connecting α0\alpha_{0} to α1\alpha_{1} such that αs(t,p)\alpha_{s}(t,p) is contained in the same characteristic leaf of Ω\Omega as α0(t,p)\alpha_{0}(t,p) for all s[0,1]s\in[0,1] and (t,p)(0,1)×Σ(t,p)\in(0,1)\times\Sigma. Then, there exists a homeomorphism of C0C^{0} Hamiltonian structures ψ:(W,Ω#𝒫0)(W,Ω#𝒫1)\psi:(W,\Omega\#{\mathcal{P}_{0}})\rightarrow(W,\Omega\#{\mathcal{P}_{1}}) which is supported inside Us[0,1]im(αs)U\coloneqq\bigcup_{s\in[0,1]}\operatorname{im}(\alpha_{s}) and isotopic to the identity through homeomorphisms supported in UU.

Proof.

Define Φ:(0,1)×Σ(0,1)×Σ\Phi:(0,1)\times\Sigma\rightarrow(0,1)\times\Sigma by Φ(t,p)(t,φt(p))\Phi(t,p)\coloneqq(t,\varphi^{t}(p)). Let K(0,1)×ΣK\subset(0,1)\times\Sigma be a compact subset such that ΦωΣ\Phi_{*}\omega_{\Sigma} agrees with ωΣ\omega_{\Sigma} on the complement of KK. As a consequence of the assumption on the embeddings αs\alpha_{s}, we can find a family of homeomorphisms ψs:(W,Ω)(W,Ω)\psi_{s}:(W,\Omega)\rightarrow(W,\Omega) starting at the identity obtained by moving points along the characteristic leaves of Ω\Omega such that ψs\psi_{s} is supported inside UU and such that ψsα0(t,p)=αs(t,p)\psi_{s}\circ\alpha_{0}(t,p)=\alpha_{s}(t,p) for all (t,p)K(t,p)\in K. Then, ψψ1\psi\coloneqq\psi_{1} is a homeomorphism (W,Ω#𝒫0)(W,Ω#𝒫1)(W,\Omega\#{\mathcal{P}_{0}})\rightarrow(W,\Omega\#{\mathcal{P}_{1}}) supported inside UU. By construction, ψ\psi is isotopic to the identity through homeomorphisms supported in UU. ∎

6 Smoothings modulo plugs

In this section we prove Theorems 6.1 & 6.11 on smoothing C0C^{0} Hamiltonian structures and their homeomorphisms.

6.1 Smoothing C0C^{0} Hamiltonian structures

We prove in this section that every C0C^{0} Hamiltonian structure on a closed 33-manifold can be obtained from a smooth Hamiltonian structure via the insertion of a C0C^{0} plug. This result, stated below, plays a crucial role in the remainder of the paper.

Theorem 6.1.

Let Ω\Omega be a C0C^{0} Hamiltonian structure on a closed topological 33-manifold YY. Then, there exist a smooth Hamiltonian structure ω\omega on YY and a C0C^{0} ω\omega-plug 𝒫\mathcal{P} such that Ω=ω#𝒫\Omega=\omega\#\mathcal{P}.

Remark 6.2.

Recall that a smooth Hamiltonian structure ω\omega on YY includes the choice of a smooth structure on YY, see Remark 5.4. Thus Theorem 6.1 in particular contains the statement that the closed topological 33-manifold YY can be smoothed. It is of course well-known that every topological 33-manifold admits a smoothing. This follows from theorems of Moise [39, Theorem 1 & 3] (see also [40, §23, Theorem 1 & §35, Theorem 3]), saying that every topological 33-manifold can be triangulated and that every triangulated 33-manifold is piecewise linear, and the fact that every piecewise linear 33-manifold can be smoothed, see e.g. [53, Theorem 3.10.8]. As we will see, the existence of a C0C^{0} Hamiltonian structure Ω\Omega on YY actually simplifies the problem of smoothing the underlying manifold YY. Essentially, it reduces the problem to smoothing a topological 22-manifold.

Remark 6.3.

Theorem 6.1 is clearly false if we do not allow for the insertion of a C0C^{0} plug, i.e. there exist C0C^{0} Hamiltonian structures which cannot be globally upgraded to smooth Hamiltonian structures. Indeed, consider a closed surface (Σ,ωΣ)(\Sigma,\omega_{\Sigma}) and an area-preserving homeomorphism φ\varphi of Σ\Sigma which is not C0C^{0} conjugate to any area-preserving diffeomorphism. For example, we can take φ\varphi to have infinite topological entropy. The Hamiltonian structure ωΣ\omega_{\Sigma} on [0,1]×Σ[0,1]\times\Sigma descends to a C0C^{0} Hamiltonian structure on the mapping torus

Yφ[0,1]×Σ/(1,p)(0,φ(p)).Y_{\varphi}\coloneqq[0,1]\times\Sigma/\sim\qquad(1,p)\sim(0,\varphi(p)).

This C0C^{0} Hamiltonian structure is not homeomorphic to any smooth Hamiltonian structure.

Let YY and Ω\Omega be as in the statement of Theorem 6.1. Let φt\varphi^{t} be a fixed-point-free topological flow whose flow lines equipped with the orientation induced by φt\varphi^{t} agree with the oriented characteristic leaves of Ω\Omega. The existence of φt\varphi^{t} is guaranteed by Proposition 8.6.

Definition 6.4.

We say a subset A(Y,Ω)A\subset(Y,\Omega) is exhaustive if there exists a constant T>0T>0 such that, for any point pYp\in Y, there exist times T<t<0<t+<T-T<t_{-}<0<t_{+}<T such that φt±(p)A\varphi^{t_{\pm}}(p)\in A.

We remark that whether or not A(Y,Ω)A\subset(Y,\Omega) is exhaustive is independent of the choice of the flow φt\varphi^{t}.

Remark 6.5.

Our proof of Theorem 6.1 shows the following slightly stronger statement, which we record here for later reference: Let (Y,Ω)(Y,\Omega) be a closed topological 33-manifold equipped with a C0C^{0} Hamiltonian structure Ω\Omega. Let U(Y,Ω)U\subset(Y,\Omega) be an arbitrary exhaustive open set. Then we can find a smooth Hamiltonian structure ω\omega on YY and an ω\omega-plug 𝒫\mathcal{P} whose image is contained in UU such that Ω=ω#𝒫\Omega=\omega\#\mathcal{P}.

The following lemma guarantees the existence of an exhaustive flow box.

Lemma 6.6.

Let Ω\Omega be a C0C^{0} Hamiltonian structure on a closed topological 33-manifold YY, and let UYU\subset Y be an open exhaustive subset. Then there exist an open surface (Σ,ωΣ)(\Sigma,\omega_{\Sigma}) equipped with an area form, a compact subsurface with boundary KΣK\subset\Sigma, and a topological embedding of C0C^{0} Hamiltonian structures

α:((0,1)×Σ,ωΣ)(U,Ω)\alpha:((0,1)\times\Sigma,\omega_{\Sigma})\hookrightarrow(U,\Omega)

such that the image α((0,1)×K)\alpha((0,1)\times K) is exhaustive.

Proof.

Since UU is exhaustive, we can fix T>0T>0 as in Definition 6.4. Consider pYp\in Y. Then there exist T<t<0<t+<T-T<t_{-}<0<t_{+}<T such that φt±(p)U\varphi^{t_{\pm}}(p)\in U. For a>0a>0 sufficiently small, we can find embeddings of C0C^{0} Hamiltonian structures α±:((1,1)×B(2a),ωstd)(U,Ω)\alpha_{\pm}:((-1,1)\times B(2a),\omega_{\operatorname{std}})\hookrightarrow(U,\Omega) such that α±(0,0)=φt±(p)\alpha_{\pm}(0,0)=\varphi^{t_{\pm}}(p). Moreover, we can find an open neighborhood VpV_{p} of pp with the property that if we start at any point qVpq\in V_{p} and follow the flow φt\varphi^{t} forward/backward in time, we meet α±({0}×B(a))\alpha_{\pm}(\{0\}\times B(a)) within time at most TT.

Since YY is compact, we can cover it by finitely many neighborhoods VpV_{p}. This implies that there exists a finite collection of embeddings of C0C^{0} Hamiltonian structures αi:((1,1)×B(2ai),ωstd)(U,Ω)\alpha_{i}:((-1,1)\times B(2a_{i}),\omega_{\operatorname{std}})\hookrightarrow(U,\Omega) with ii ranging from 11 to nn such that the union iαi({0}×B(ai))\bigcup_{i}\alpha_{i}(\{0\}\times B(a_{i})) is an exhaustive set. We remark that the surfaces αi({0}×B(ai))\alpha_{i}(\{0\}\times B(a_{i})) are not necessarily pairwise disjoint.

Our next step is to construct, for each ii, finitely many closed discs of the form {sij}×Dij(1,1)×B(2ai)\{s_{i}^{j}\}\times D_{i}^{j}\subset(-1,1)\times B(2a_{i}) such that B(ai)jDijB(a_{i})\subset\bigcup_{j}D_{i}^{j} and the images αi({sij}×Dij)\alpha_{i}(\{s_{i}^{j}\}\times D_{i}^{j}), ranging over all i,ji,j, are pairwise disjoint.

The construction of these discs relies on Claim 6.7, stated below. Let a>0a>0 and consider a closed subset S((1,1)×B(2a),ωstd)S\subset((-1,1)\times B(2a),\omega_{\operatorname{std}}) which is a 22-dimensional topological submanifold, possibly with boundary, and which is transverse to the characteristic foliation on (1,1)×B(2a)(-1,1)\times B(2a) in the following sense: near every point of SS there exists a local C0C^{0} Hamiltonian chart in which both SS and the characteristic foliation are smooth and in which SS is transverse to the characteristic foliation in the usual sense.

Claim 6.7.

We can find finitely many pairwise disjoint closed discs in (1,1)×B(2a)(-1,1)\times B(2a) of the form {sj}×Dj\{s^{j}\}\times D^{j} which are all disjoint from SS and such that B(a)jDjB(a)\subset\bigcup_{j}D^{j}.

Proof of Claim 6.7.

Indeed, consider an arbitrary point pB(2a)p\in B(2a). It follows from the transversality assumption on SS that the characteristic leaf (1,1)×{p}(-1,1)\times\{p\} contains a point in the complement of SS. This means that we can find a small closed disc DB(2a)D\subset B(2a) containing pp and s(1,1)s\in(-1,1) such that {s}×D\{s\}\times D is disjoint from SS. Since B¯(a)\overline{B}(a) is compact, we can find finitely many discs {sj}×Dj\{s^{j}\}\times D^{j} in the complement of SS such that the DjD^{j} cover B(a)B(a). We can in addition make these discs pairwise disjoint by slightly perturbing the heights sjs^{j} so that the sjs^{j} are all distinct. This proves the claim. ∎

Returning to the construction of the discs {sij}×Dij\{s_{i}^{j}\}\times D_{i}^{j}, let 1in1\leq i\leq n and assume that the discs {skj}×Dkj\{s_{k}^{j}\}\times D_{k}^{j} have already been construction for all k<ik<i. Now simply set

Sαi1(k<ijαk({skj}×Dkj))(1,1)×B(2ai)S\coloneqq\alpha_{i}^{-1}(\bigcup_{k<i}\bigcup_{j}\alpha_{k}(\{s_{k}^{j}\}\times D_{k}^{j}))\subset(-1,1)\times B(2a_{i})

and apply the above claim to find the discs {sij}×Dij\{s_{i}^{j}\}\times D_{i}^{j}. This concludes our construction of the discs {sij}×Dij\{s_{i}^{j}\}\times D_{i}^{j}. We observe that since iαi({0}×B(ai))\bigcup_{i}\alpha_{i}(\{0\}\times B(a_{i})) is an exhaustive set and B(ai)B(a_{i}) is contained in the union jDij\bigcup_{j}D_{i}^{j}, the set i,jαi({sij}×Dij)\bigcup_{i,j}\alpha_{i}(\{s_{i}^{j}\}\times D_{i}^{j}) is exhaustive as well.

For each disc DijD_{i}^{j}, pick a slightly larger open disc BijB_{i}^{j}. Let ε>0\varepsilon>0 be small. Consider the embedding

α~:i,j(sij,sij+ε)×BijU\tilde{\alpha}:\bigsqcup_{i,j}(s_{i}^{j},s_{i}^{j}+\varepsilon)\times B_{i}^{j}\hookrightarrow U

which on each component is given by the restriction of the corresponding embedding αi\alpha_{i}. Here we have to choose ε>0\varepsilon>0 sufficiently small and take the enlarged discs BijB_{i}^{j} sufficiently close to the original discs DijD_{i}^{j} to ensure that this indeed defines an embedding and has image contained in UU.

We define KK to be the disjoint union of all the discs DijD_{i}^{j} and Σ\Sigma to be the disjoint union of all the discs BijB_{i}^{j}. Then the desired embedding α:(0,1)×ΣU\alpha:(0,1)\times\Sigma\hookrightarrow U is obtained from the above embedding α\alpha by reparametrizing the intervals (sij,sij+ε)(s_{i}^{j},s_{i}^{j}+\varepsilon) to (0,1)(0,1). ∎

We now present a proof of Theorem 6.1.

Proof of Theorem 6.1.

By Lemma 6.6, applied to the exhaustive set U=YU=Y, we may pick an open surface (Σ,ωΣ)(\Sigma,\omega_{\Sigma}), a compact subsurface with boundary KΣK\subset\Sigma, and an embedding of C0C^{0} Hamiltonian structures α:((0,1)×Σ,ωΣ)(Y,Ω)\alpha:((0,1)\times\Sigma,\omega_{\Sigma})\hookrightarrow(Y,\Omega) such that α((0,1)×K)\alpha((0,1)\times K) is exhaustive. In order to prove Remark 6.5, we note that if we are given an arbitrary exhaustive open subset U(Y,Ω)U\subset(Y,\Omega), we can choose α\alpha such that its image is contained in UU. Let 0<t0<t1<10<t_{0}<t_{1}<1 and define the compact interval I[t0,t1]I\coloneqq[t_{0},t_{1}]. Then the set Aα(I×K)A\coloneqq\alpha(I\times K) is also exhaustive. Up to possibly enlarging KK, we can assume that the interior of AA is still exhaustive.

Set WYAW\coloneqq Y\setminus A and equip it with the restriction of the C0C^{0} Hamiltonian structure Ω\Omega. Our goal is to construct a smoothing of the C0C^{0} Hamiltonian structure (W,Ω)(W,\Omega). The crucial observation that allows us to achieve this is that the characteristic foliation of (W,Ω)(W,\Omega) exhibits no complicated recurrent behavior. The closure (with respect to YY) of any characteristic leaf of (W,Ω)(W,\Omega) is an embedded compact interval which starts at α({t1}×K)\alpha(\{t_{1}\}\times K) and ends at α({t0}×K)\alpha(\{t_{0}\}\times K).

Let \mathcal{L} denote the leaf space of the characteristic foliation of (W,Ω)(W,\Omega) and let pr:W\operatorname{pr}:W\rightarrow\mathcal{L} denote the natural projection.

Claim 6.8.

The leaf space \mathcal{L} has the structure of a 22-dimensional, topological, possibly non-Hausdorff manifold equipped with a measure which in local charts is homeomorphic to the standard Lebesgue measure on 2{\mathbb{R}}^{2}. This means that \mathcal{L} admits an atlas whose transition maps are area-preserving homeomorphisms between open subsets of 2{\mathbb{R}}^{2}.

Remark 6.9.

It is not surprising to encounter non-Hausdorff manifolds in the context of leaf spaces, see for example the work of Haefliger and Reeb [25], which studies non-Hausdorff 11-manifolds in connection with foliations of 2{\mathbb{R}}^{2}.

Proof.

Consider a point pWp\in W. Since WW is an open subset of YY, for b>0b>0 sufficiently small, we can chose an embedding of C0C^{0} Hamiltonian structures β:((1,1)×B(b),ωstd)(W,Ω)\beta:((-1,1)\times B(b),\omega_{\operatorname{std}})\hookrightarrow(W,\Omega) such that β(0,0)=p\beta(0,0)=p. We show that after possibly shrinking bb, we can assume that the restriction of the projection pr:W\operatorname{pr}:W\rightarrow\mathcal{L} to β({0}×B(b))\beta(\{0\}\times B(b)) is injective. Indeed, assume by contradiction that this is not the case. Then there exist a sequence of points pkβ({0}×B(k1b))p_{k}\in\beta(\{0\}\times B(k^{-1}b)) and a sequence of times tk{0}t_{k}\in{\mathbb{R}}\setminus\{0\} such that φtk(pk)β({0}×B(k1b))\varphi^{t_{k}}(p_{k})\in\beta(\{0\}\times B(k^{-1}b)) and such φt(pk)\varphi^{t}(p_{k}) remains in WW as tt ranges from 0 to tkt_{k}. Here, φt\varphi^{t} is a choice of flow as in Definition 6.4. Since the complement of WW is exhaustive, the sequence tkt_{k} must be bounded. Clearly, the sequence must also be bounded away from zero. After passing to a subsequence, we can therefore assume that tkt_{k} converges to t{0}t_{*}\in{\mathbb{R}}\setminus\{0\}. This implies that pp is a periodic orbit of φt\varphi^{t} of period tt_{*}. Moreover, this periodic orbit must be contained in the closure of WW since it is the limit of flow line segments contained in WW. But this contradicts the fact that the complement of the closure of WW (i.e. the interior of AA) is exhaustive.

The above discussion yields a continuous injective map

γ:B(b){0}×B(b)(1,1)×B(b)𝛽Wpr\gamma:B(b)\cong\{0\}\times B(b)\subset(-1,1)\times B(b)\overset{\beta}{\rightarrow}W\overset{\operatorname{pr}}{\rightarrow}\mathcal{L}

mapping 0 to pr(p)\operatorname{pr}(p). We claim that this map is open. This amounts to showing that pr1(γ(V))W\operatorname{pr}^{-1}(\gamma(V))\subset W is open for every open subset VB(b)V\subset B(b). In order to see this, consider a point qpr1(γ(V))q\in\operatorname{pr}^{-1}(\gamma(V)). The point qq being in pr1(γ(V))\operatorname{pr}^{-1}(\gamma(V)) means that there exists a line segment contained in a characteristic leaf of (W,Ω)(W,\Omega) which connects qq to a point in β({0}×V)\beta(\{0\}\times V). Since WW and VV are open, every point qq^{\prime} in a neighborhood of qq can also be connected to β({0}×V)\beta(\{0\}\times V) via a line segment contained in a leaf of (W,Ω)(W,\Omega). Thus pr1(γ(V))\operatorname{pr}^{-1}(\gamma(V)) contains a neighborhood of qq, showing that pr1(γ(V))\operatorname{pr}^{-1}(\gamma(V)) is open.

We conclude that γ\gamma is a homeomorphism between B(b)B(b) and an open neighborhood of pr(p)\operatorname{pr}(p) in \mathcal{L}. Since pWp\in W was chosen arbitrarily, this implies that \mathcal{L} is a 22-dimensional, topological, possibly non-Hausdorff manifold. The transverse measure on the characteristic foliation of (W,Ω)(W,\Omega) clearly descends to a measure on \mathcal{L} locally homeomorphic to the standard 22-dimensional Lebesgue measure. ∎

Claim 6.10.

\mathcal{L} has a smoothing, i.e. it admits an atlas whose transition maps are area-preserving diffeomorphisms between open subsets of 2{\mathbb{R}}^{2}.

Proof.

It is well-known that every Hausdorff topological 22-manifold admits a smoothing. Indeed, by a theorem of Radó [45] (see also [40, §8, Theorem 3]) every Hausdorff topological 22-manifold admits a piecewise linear structure. Moreover, piecewise linear 22-manifolds, which are automatically Hausdorff, can be smoothed, see e.g. [53, Theorem 3.10.8]. Claim 6.10 cannot be directly deduced from these results since it involves non-Hausdorff manifolds and area-preserving homeomorphisms/diffeomorphisms. For this reason, we provide a proof of Claim 6.10. As we will now explain, Claim 6.10 essentially boils down to the fact that area-preserving homeomorphisms can be approximated by area-preserving diffeomorphisms, see Proposition 2.2.

Note that \mathcal{L} admits a finite open covering by charts φi:UiVi2\varphi_{i}:U_{i}\rightarrow V_{i}\subset{\mathbb{R}}^{2} with 1in1\leq i\leq n. In the following, it will be useful to be able to replace these charts by smaller charts φi|Ui:Uiφi(Ui)\varphi_{i}|_{U_{i}^{\prime}}:U_{i}^{\prime}\rightarrow\varphi_{i}(U_{i}^{\prime}) for relatively compact open subsets UiUiU_{i}^{\prime}\Subset U_{i} which still cover \mathcal{L}. We can choose our initial charts φi\varphi_{i} such that this is possible. Note that in this situation we can shrink the charts in such a way that it is possible to repeat the chart shrinking for the resulting cover by charts.

For all i,ji,j, we set Vjiφi(UiUj)V_{ji}\coloneqq\varphi_{i}(U_{i}\cap U_{j}) and let

φji:VjiVijφjiφjφi1\varphi_{ji}:V_{ji}\rightarrow V_{ij}\qquad\varphi_{ji}\coloneqq\varphi_{j}\circ\varphi_{i}^{-1}

denote the transition map, which is an area-preserving homeomorphism. Our goal is to modify the charts φi\varphi_{i} such that the transition maps φji\varphi_{ji} become smooth for all i,ji,j. Let 1j<in1\leq j<i\leq n. Suppose we have already modified the charts in such a way that φk\varphi_{\ell k} is smooth for all k,<ik,\ell<i and such that φki\varphi_{ki} is smooth for all k<jk<j. We explain how to modify the charts to make φji\varphi_{ji} smooth without destroying smoothness of the transition maps we have already made smooth. Set VVjik<jVkiV\coloneqq V_{ji}\cap\bigcup_{k<j}V_{ki}. It follows from our assumptions that the restriction of φji\varphi_{ji} to VV is smooth. After slightly shrinking all chart neighborhoods UkU_{k}, we can assume that φji\varphi_{ji} is smooth on an open neighborhood of the closure of VV inside VjiV_{ji}. Pick a non-negative continuous function ρ:Vji0\rho:V_{ji}\rightarrow{\mathbb{R}}_{\geq 0} such that VV is contained in ρ1({0})\rho^{-1}(\{0\}) and ρ1({0})\rho^{-1}(\{0\}) is contained in the smooth locus of φji\varphi_{ji}. Moreover, assume that ρ\rho decays to 0 towards the boundary of VjiV_{ji}. By Proposition 2.2, we can find an area-preserving diffeomorphism φji:VjiVij\varphi_{ji}^{\prime}:V_{ji}\rightarrow V_{ij} such that |φji(p)φji(p)|ρ(p)|\varphi_{ji}^{\prime}(p)-\varphi_{ji}(p)|\leq\rho(p) for all pVjip\in V_{ji}. Since ρ\rho decays to 0 towards the boundary of VjiV_{ji}, the homeomorphism φji1φji\varphi_{ji}^{-1}\circ\varphi_{ji}^{\prime} of VjiV_{ji} extends to an area-preserving homeomorphism χ\chi of ViV_{i} which agrees with the identity on the complement of VjiV_{ji}. Now define the chart φiχ1φi\varphi_{i}^{\prime}\coloneqq\chi^{-1}\circ\varphi_{i}. The transition map between φi\varphi_{i}^{\prime} and φj\varphi_{j} is simply given by φji\varphi_{ji}^{\prime} and is therefore smooth. Moreover, note that since φji\varphi_{ji} and φji\varphi_{ji}^{\prime} agree on VV, the homeomorphism χ\chi restricts to the identity on VkiV_{ki} for all k<jk<j. Thus the transition map φki\varphi_{ki} does not change and remains smooth. Now, simply replace φi\varphi_{i} by φi\varphi_{i}^{\prime}. This concludes our construction of a smoothing of \mathcal{L}

We will now use the smooth structure on \mathcal{L} to construct a smoothing of (W,Ω)(W,\Omega), i.e. we will construct a C0C^{0} Hamiltonian atlas for (W,Ω)(W,\Omega) whose transition maps are smooth.

Recall that we have a fixed-point-free flow φt\varphi^{t} on YY whose flow lines along with their natural orientations agree with the oriented leaves of the characteristic foliation of Ω\Omega. Let a>0a>0 be sufficiently small and pick finitely many topological embeddings ιi:B(a)W\iota_{i}:B(a)\hookrightarrow W such that the images of prιi\operatorname{pr}\circ\iota_{i} cover \mathcal{L} and such that each prιi\operatorname{pr}\circ\iota_{i} is an area-preserving embedding which is smooth with respect to the smooth structure on \mathcal{L} constructed in Claim 6.10. Given two embeddings ιi\iota_{i} and ιj\iota_{j}, we define

Vji(prιi)1(prιi(B(a))prιj(B(a)))B(a).V_{ji}\coloneqq(\operatorname{pr}\circ\iota_{i})^{-1}(\operatorname{pr}\circ\iota_{i}(B(a))\cap\operatorname{pr}\circ\iota_{j}(B(a)))\subset B(a).

Note that, for every point pVjip\in V_{ji}, there exists a unique transfer time tji(p)t_{ji}(p)\in{\mathbb{R}} such that φtji(p)(ιi(p))im(ιj)\varphi^{t_{ji}(p)}(\iota_{i}(p))\in\operatorname{im}(\iota_{j}) and φt(ιi(p))W\varphi^{t}(\iota_{i}(p))\in W for all tt ranging from 0 to tji(p)t_{ji}(p). The functions tji:Vjit_{ji}:V_{ji}\rightarrow{\mathbb{R}} are continuous. We can perturb the embeddings ιi\iota_{i}, without changing prιi\operatorname{pr}\circ\iota_{i}, such that all the tjit_{ji} become smooth. This can be done via an elementary smoothing process similar to the one described in the proof of Claim 6.10. In place of Proposition 2.2, the smoothing here relies on the simpler fact that real-valued continuous functions can be approximated by smooth ones. We omit the details of this.

Now, we say an embedding ι:B(b)W\iota:B(b)\hookrightarrow W to be smooth if prι\operatorname{pr}\circ\iota is a smooth area-preserving embedding and if, for every embedding ιi\iota_{i}, the locally defined transfer time between the image of ι\iota and the image of ιi\iota_{i} is smooth. Note that by construction, if the transfer time between the image of ι\iota and the image of ιi\iota_{i} is smooth near some point pB(b)p\in B(b), then the transfer time between the image of ι\iota and the image of any other ιj\iota_{j} is smooth near pp as well. In particular, this implies that we can find a smooth embedding ι\iota passing through any point in WW.

Finally, consider all C0C^{0} Hamiltonian embeddings of the form

ψ:((ε,ε)×B(b),ωstd)(W,Ω)ψ(t,p)=φt(ι(p))\psi:((-\varepsilon,\varepsilon)\times B(b),\omega_{\operatorname{std}})\hookrightarrow(W,\Omega)\qquad\psi(t,p)=\varphi^{t}(\iota(p))

for some ε>0\varepsilon>0 and some smooth embedding ι:B(b)W\iota:B(b)\hookrightarrow W. We can cover WW by images of embeddings of this form. The inverses of these embeddings yield the desired collection of C0C^{0} Hamiltonian charts with smooth transition maps.

We have now constructed a smoothing of (W,Ω)(W,\Omega); however, it does not in general extend to a global smoothing of (Y,Ω)(Y,\Omega). The problem is the following: Let Sα((0,t0)×Σ)S_{-}\subset\alpha((0,t_{0})\times\Sigma) and S+α((t1,1)×Σ)S_{+}\subset\alpha((t_{1},1)\times\Sigma) be two surfaces which are transverse to the characteristic foliation and smoothly embedded with respect to the smoothing of (W,Ω)(W,\Omega) constructed above. Moreover, suppose that each leaf segment of the form α((0,1)×{p})\alpha((0,1)\times\{p\}) for pΣp\in\Sigma intersects each of the surfaces S±S_{\pm} exactly once. Traversing α((0,1)×Σ)\alpha((0,1)\times\Sigma) along the characteristic foliation then yields a homeomorphism SS+S_{-}\rightarrow S_{+}. This homeomorphism is not going to be smooth in general and this is exactly the obstruction to extending the smoothing of (W,Ω)(W,\Omega) to all of YY. We may, however, modify Ω\Omega by inserting an Ω\Omega-plug of the form 𝒬=(Σ,ωΣ,α,(ψt)t[0,1])\mathcal{Q}=(\Sigma,\omega_{\Sigma},\alpha,(\psi^{t})_{t\in[0,1]}) such that the resulting map SS+S_{-}\rightarrow S_{+} becomes smooth. We may then extend our smoothing of (W,Ω)(W,\Omega) to a global smoothing of the C0C^{0} Hamiltonian structure Ω#𝒬\Omega\#\mathcal{Q} on YY.

Now, simply set ω=Ω#𝒬\omega=\Omega\#\mathcal{Q} and 𝒫𝒬¯\mathcal{P}\coloneqq\overline{\mathcal{Q}}, where 𝒬¯\overline{\mathcal{Q}} denotes the inverse plug of 𝒬\mathcal{Q}. Then, ω#𝒫=Ω#𝒬#𝒫=Ω\omega\#\mathcal{P}=\Omega\#\mathcal{Q}\#\mathcal{P}=\Omega, as desired.

To prove the strengthened statement in Remark 6.5, the image of 𝒫\mathcal{P} is contained in the given exhaustive open set UU because the embedding α\alpha was chosen to have image contained in UU. ∎

6.2 Smoothing C0C^{0} Hamiltonian homeomorphisms

We prove in this section that homeomorphisms of C0C^{0} Hamiltonian structures on closed 33-manifolds may be smoothed, up to insertion of plugs.

Theorem 6.11.

For i{1,2}i\in\{1,2\}, let YiY_{i} be a closed 33-manifold and let ωi\omega_{i} be a smooth Hamiltonian structure on YiY_{i}. Let 𝒫i\mathcal{P}_{i} be a C0C^{0} ωi\omega_{i}-plug and assume that

ψ:(Y1,ω1#𝒫1)(Y2,ω2#𝒫2)\psi:(Y_{1},\omega_{1}\#\mathcal{P}_{1})\rightarrow(Y_{2},\omega_{2}\#\mathcal{P}_{2})

is a homeomorphism of C0C^{0} Hamiltonian structures. Then, there exist smooth ωi\omega_{i}-plugs 𝒬i\mathcal{Q}_{i} for i{1,2}i\in\{1,2\} and a diffeomorphism of Hamiltonian structures

φ:(Y1,ω1#𝒬1)(Y2,ω2#𝒬2)\varphi:(Y_{1},\omega_{1}\#\mathcal{Q}_{1})\rightarrow(Y_{2},\omega_{2}\#\mathcal{Q}_{2})

with the following properties:

  1. 1.

    Cal(𝒬1)Cal¯(𝒫1)=Cal(𝒬2)Cal¯(𝒫2)\operatorname{Cal}(\mathcal{Q}_{1})-\overline{\operatorname{Cal}}(\mathcal{P}_{1})=\operatorname{Cal}(\mathcal{Q}_{2})-\overline{\operatorname{Cal}}(\mathcal{P}_{2})

  2. 2.

    φ\varphi is homotopic to ψ\psi through homeomorphisms.

Remark 6.12.

As an immediate corollary of Theorem 6.11, one obtains that two exact smooth Hamiltonian structures ω1\omega_{1} and ω2\omega_{2} on YY which are homeomorphic have the same helicity. Indeed, we can apply Theorem 6.11 with empty plugs 𝒫i\mathcal{P}_{i}. Then the resulting smooth ωi\omega_{i}-plugs 𝒬i\mathcal{Q}_{i} have the same (smooth) Calabi invariant, i.e. Cal(𝒬1)=Cal(𝒬2)\operatorname{Cal}(\mathcal{Q}_{1})=\operatorname{Cal}(\mathcal{Q}_{2}). Since the smooth Hamiltonian structures ω1#𝒬1\omega_{1}\#\mathcal{Q}_{1} and ω2#𝒬2\omega_{2}\#\mathcal{Q}_{2} are diffeomorphic, they have the same helicity. Therefore, we can conclude that (ω1)=(ω2)\mathcal{H}(\omega_{1})=\mathcal{H}(\omega_{2}) by Lemma 4.10.

The proof of Theorem 6.11, which takes up the rest of this section, requires some preparatory lemmas.

Let (Σ,ω)(\Sigma,\omega) be a surface equipped with an area-form ω\omega. Let CΣC_{\Sigma} denote the cylinder (0,3)×Σ(0,3)\times\Sigma; we equip it with the smooth Hamiltonian structure induced by ω\omega. We also introduce the notation

CΣ(0,1)×Σ,CΣ0(1,2)×Σ,CΣ+(2,3)×Σ.C^{-}_{\Sigma}\coloneqq(0,1)\times\Sigma,\qquad C^{0}_{\Sigma}\coloneqq(1,2)\times\Sigma,\qquad C^{+}_{\Sigma}\coloneqq(2,3)\times\Sigma.

Figure 6.1 illustrates CΣC_{\Sigma} and may aid the reader during the proof of Lemma 6.13. In that proof, it is helpful to visualize CΣC^{-}_{\Sigma} and CΣ+C^{+}_{\Sigma} as being thin relative to CΣ0C^{0}_{\Sigma}. This can be made precise via a reparametrization of the interval (0,3)(0,3), but we omit the details to keep the notation light and the argument clear.

Refer to caption
CΣC_{\Sigma}^{-}
CΣ+C_{\Sigma}^{+}
CΣ0C_{\Sigma}^{0}
Refer to caption
Figure 6.1: A depiction of CΣC_{\Sigma}. The dotted vertical arrow indicates the direction of the characteristic foliation in CΣC_{\Sigma}. It is helpful to imagine CΣ,CΣ+C^{-}_{\Sigma},C^{+}_{\Sigma} as thin relative to CΣ0C^{0}_{\Sigma}.
Lemma 6.13.

Let YY be a closed topological 33-manifold equipped with a C0C^{0} Hamiltonian structure Ω\Omega. Let U,VYU,V\subset Y be two open sets such that UV=YU\cup V=Y. Assume that UU is exhaustive in the sense of Definition 6.4.

Then there exist open surfaces with area forms (Σ1,ω1),,(Σn,ωn)(\Sigma_{1},\omega_{1}),\ldots,(\Sigma_{n},\omega_{n}), relatively compact open subsets SiΣiS_{i}\Subset\Sigma_{i}, and topological embeddings of C0C^{0} Hamiltonian structures

ιi:(CΣi,ωi)(V,Ω)\iota_{i}:(C_{\Sigma_{i}},\omega_{i})\hookrightarrow(V,\Omega)

satisfying the following properties:

  1. 1.

    YUiιi(CSi0)Y\setminus U\subset\bigcup_{i}\iota_{i}(C_{S_{i}}^{0})

  2. 2.

    ιi(CΣi±)¯UV\overline{\iota_{i}(C_{\Sigma_{i}}^{\pm})}\subset U\cap V for all ii.

  3. 3.

    ιj(CΣj±)ιi(CΣi)=\iota_{j}(C_{\Sigma_{j}}^{\pm})\cap\iota_{i}(C_{\Sigma_{i}})=\emptyset for all j<ij<i.

Moreover, if the restriction of Ω\Omega to VV is smooth, we can take all ιi\iota_{i} to be smooth as well.

Proof.

Let pYUp\in Y\setminus U be an arbitrary point. Let LL denote the characteristic leaf passing through pp. By the exhaustiveness assumption on UU, there exists a topologically embedded closed interval ILVI\subset L\cap V such that pp is contained in the interior of II and the endpoints of II are contained in UVU\cap V. Using sufficiently small local transverse sections of the characteristic foliation at the endpoints of II, one can construct an open surface with area form (Σ,ω)(\Sigma,\omega) and a topological embedding ι:(CΣ,ω)(V,Ω)\iota:(C_{\Sigma},\omega)\hookrightarrow(V,\Omega) containing pp in its image such that ι(CΣ±)¯UV\overline{\iota(C_{\Sigma}^{\pm})}\subset U\cap V. If the restriction of Ω\Omega to VV is smooth, we can construct ι\iota to be smooth as well.

Since the set YUY\setminus U is compact, we can cover it with the images of finitely many such cylinder embeddings. We may therefore pick n0n\geq 0 and tuples (Σi,Si,ιi)(\Sigma_{i},S_{i},\iota_{i}) for 1in1\leq i\leq n such that

YUiιi(CSi0)andιi(CΣi±)¯UV.Y\setminus U\subset\bigcup_{i}\iota_{i}(C_{S_{i}}^{0})\qquad\text{and}\qquad\overline{\iota_{i}(C_{\Sigma_{i}}^{\pm})}\subset U\cap V.

It remains to explain how to achieve property 3 in the statement of the lemma. First, note that there is enough flexibility in the construction of (Σi,Si,ιi)(\Sigma_{i},S_{i},\iota_{i}) to make sure that the intersection between ιi(CΣi)¯\overline{\iota_{i}(C_{\Sigma_{i}}^{*})} and ιj(CΣj)¯\overline{\iota_{j}(C_{\Sigma_{j}}^{\bullet})} for 1i,jn1\leq i,j\leq n and ,{+,}*,\bullet\in\{+,-\} is non-empty if and only if i=ji=j and =*=\bullet.

Moreover, we can arrange (Σi,Si,ιi)(\Sigma_{i},S_{i},\iota_{i}) to have the property that, for all j<ij<i and {+,}*\in\{+,-\} such that the intersection ιj(CΣj)ιi(CΣi)\iota_{j}(C_{\Sigma_{j}}^{*})\cap\iota_{i}(C_{\Sigma_{i}}) is non-empty, there exist an open subset EΣiE\subset\Sigma_{i} and an open interval J(1,2)J\Subset(1,2) such that

ιi1(ιj(CΣj))=J×E.\iota_{i}^{-1}(\iota_{j}(C_{\Sigma_{j}}^{*}))=J\times E.

In order to resolve, and remove, the intersection between ιj(CΣj)\iota_{j}(C_{\Sigma_{j}}^{*}) and ιi(CΣi)\iota_{i}(C_{\Sigma_{i}}), we replace (Σi,Si,ιi)(\Sigma_{i},S_{i},\iota_{i}) by three tuples (Σik,Sik,ιik)(\Sigma_{i}^{k},S_{i}^{k},\iota_{i}^{k}) with 1k31\leq k\leq 3 as described in detail below. At the end of this process, the cylinder ιi(CΣi)\iota_{i}(C_{\Sigma_{i}}) will be replaced by three cylinders ιi1(CΣi1),ιi2(CΣi2),ιi3(CΣi3)\iota_{i}^{1}(C_{\Sigma_{i}^{1}}),\iota_{i}^{2}(C_{\Sigma_{i}^{2}}),\iota_{i}^{3}(C_{\Sigma_{i}^{3}}), which will be constructed in such a way that they do not intersect ιj(CΣj)\iota_{j}(C_{\Sigma_{j}}^{*}). This, of course, requires removing some portions of ιi(CΣi)\iota_{i}(C_{\Sigma_{i}}), and also ιi(CSi0)\iota_{i}(C_{S_{i}}^{0}). The removed portions will all be contained in UU and hence property 1 will continue to hold. We will also ensure that property 2 continues to hold. However, this is less complicated.

Refer to caption
CΣC_{\Sigma_{-}}
CΣ+C_{\Sigma_{+}}
CΣ0C_{\Sigma_{0}}
Refer to caption
CΣjC_{\Sigma_{j}}
CΣiC_{\Sigma_{i}}
Refer to caption
CΣjC_{\Sigma_{j}}
Refer to caption
CΣi1C_{\Sigma_{i}^{1}}
CΣi2C_{\Sigma_{i}^{2}}
CΣi3C_{\Sigma_{i}^{3}}
Figure 6.2: Removing intersections: On the left, CΣj+C_{\Sigma_{j}}^{+} intersects CΣiC_{\Sigma_{i}}, with i>ji>j. On the right, CΣiC_{\Sigma_{i}} is replaced with with CΣi1,CΣi2,CΣi3C_{\Sigma_{i}^{1}},C_{\Sigma_{i}^{2}},C_{\Sigma_{i}^{3}}.

We now proceed with a detailed description of the above process, which is also depicted in Figure 6.2. Pick open sets

Si1Σi1ΣiandSi2=Si3Σi2=Σi3ΣiS_{i}^{1}\Subset\Sigma_{i}^{1}\Subset\Sigma_{i}\quad\text{and}\quad S_{i}^{2}=S_{i}^{3}\Subset\Sigma_{i}^{2}=\Sigma_{i}^{3}\Subset\Sigma_{i}

such that

SiSi1Si2,EΣi1=andιi(J×Σi2)¯UV.S_{i}\subset S_{i}^{1}\cup S_{i}^{2},\quad E\cap\Sigma_{i}^{1}=\emptyset\quad\text{and}\quad\overline{\iota_{i}(J\times\Sigma_{i}^{2})}\subset U\cap V.

Moreover, we can arrange Σi2=Σi3\Sigma_{i}^{2}=\Sigma_{i}^{3} to be contained in an arbitrarily small neighborhood of the closure of EE inside Σi\Sigma_{i}. Note that (1,2)J(1,2)\setminus J has two components (1,t](1,t_{-}] and [t+,2)[t_{+},2). We pick orientation-preserving embeddings

τ2:[0,3](1,t)andτ3:[0,3](t+,2)\tau_{2}:[0,3]\rightarrow(1,t_{-})\quad\text{and}\quad\tau_{3}:[0,3]\rightarrow(t_{+},2)

such that

τ2([0,1])\displaystyle\tau_{2}([0,1]) (1,1+ε)\displaystyle\subset(1,1+\varepsilon) τ2([2,3])\displaystyle\tau_{2}([2,3]) (tε,t)\displaystyle\subset(t_{-}-\varepsilon,t_{-})
τ3([0,1])\displaystyle\tau_{3}([0,1]) (t+,t++ε)\displaystyle\subset(t_{+},t_{+}+\varepsilon) τ3([2,3])\displaystyle\tau_{3}([2,3]) (2ε,2)\displaystyle\subset(2-\varepsilon,2)

for some small ε>0\varepsilon>0. We define

ιi1:CΣi1\displaystyle\iota_{i}^{1}:C_{\Sigma_{i}^{1}} V\displaystyle\rightarrow V ιi1(t,p)\displaystyle\iota_{i}^{1}(t,p) ιi(t,p)\displaystyle\coloneqq\iota_{i}(t,p)
ιi2:CΣi2\displaystyle\iota_{i}^{2}:C_{\Sigma_{i}^{2}} V\displaystyle\rightarrow V ιi2(t,p)\displaystyle\iota_{i}^{2}(t,p) ι(τ2(t),p)\displaystyle\coloneqq\iota(\tau_{2}(t),p)
ιi3:CΣi3\displaystyle\iota_{i}^{3}:C_{\Sigma_{i}^{3}} V\displaystyle\rightarrow V ιi3(t,p)\displaystyle\iota_{i}^{3}(t,p) ιi(τ3(t),p).\displaystyle\coloneqq\iota_{i}(\tau_{3}(t),p).

If ε\varepsilon is chosen sufficiently small and Σi2=Σi3\Sigma_{i}^{2}=\Sigma_{i}^{3} is contained in a sufficiently small neighborhood of the closure of EE in Σi\Sigma_{i}, then the six sets ιik(CΣik±)¯\overline{\iota_{i}^{k}(C_{\Sigma_{i}^{k}}^{\pm})} for 1k31\leq k\leq 3 are contained in UVU\cap V and are disjoint from ι(CΣ±)¯\overline{\iota_{\ell}(C_{\Sigma_{\ell}}^{\pm})} for any i\ell\neq i. They are also pairwise disjoint. Moreover, ιi(CSi0)U\iota_{i}(C_{S_{i}}^{0})\setminus U is contained in kιik(CSik0)\bigcup_{k}\iota_{i}^{k}(C_{S_{i}^{k}}^{0}), which implies that

YU1k3ιik(CSii0)iι(CS0).Y\setminus U\subset\bigcup_{1\leq k\leq 3}\iota_{i}^{k}(C_{S_{i}^{i}}^{0})\cup\bigcup_{\ell\neq i}\iota_{\ell}(C_{S_{\ell}}^{0}).

The three sets ιik(CΣik)\iota_{i}^{k}(C_{\Sigma_{i}^{k}}) for 1k31\leq k\leq 3 are disjoint from ιj(CΣj)\iota_{j}(C_{\Sigma_{j}}^{*}). In addition, ιik(CΣik±)\iota_{i}^{k}(C_{\Sigma_{i}^{k}}^{\pm}) is disjoint from ιi(CΣi)\iota_{i}^{\ell}(C_{\Sigma_{i}^{\ell}}) for 1k<31\leq k<\ell\leq 3. We discard the tuple (Si,Σi,ιi)(S_{i},\Sigma_{i},\iota_{i}) and instead insert the three tuples (Si1,Σi1,ιi1),(Si2,Σi2,ιi2),(Si3,Σi3,ιi3)(S_{i}^{1},\Sigma_{i}^{1},\iota_{i}^{1}),(S_{i}^{2},\Sigma_{i}^{2},\iota_{i}^{2}),(S_{i}^{3},\Sigma_{i}^{3},\iota_{i}^{3}) between (Σi1,Si1,ιi1)(\Sigma_{i-1},S_{i-1},\iota_{i-1}) and (Σi+1,Si+1,ιi+1)(\Sigma_{i+1},S_{i+1},\iota_{i+1}). Then we adjust the indexing of our new list of tuples. As observed above, properties 1 and 2 in Lemma 6.13 are still satisfied. Applying the above construction finitely many times, we can get rid of all non-empty intersections of the form ιj(CΣj±)ιi(CΣi)\iota_{j}(C_{\Sigma_{j}}^{\pm})\cap\iota_{i}(C_{\Sigma_{i}}) for j<ij<i and therefore also achieve property 3. ∎

Definition 6.14.

Let U,VU,V be open subset of 3{\mathbb{R}}^{3}. Suppose that ψ:(U,ωstd)(V,ωstd)\psi:(U,\omega_{\operatorname{std}})\rightarrow(V,\omega_{\operatorname{std}}) is a homeomorphism of Hamiltonian structures. We say that ψ\psi is transversely smooth if, for every point pUp\in U, there exist an open neighborhood of pp of the form I×BI\times B, where II\subset{\mathbb{R}} is an open interval and B2B\subset{\mathbb{R}}^{2} is a ball, and a smooth area-preserving embedding a:B2a:B\hookrightarrow{\mathbb{R}}^{2} such that

ψ(t,z)=(,a(z))for all (t,z)I×B.\psi(t,z)=(*,a(z))\qquad\text{for all $(t,z)\in I\times B$.}

Let ψ:(Y,ω0)(Y,ω1)\psi:(Y,\omega_{0})\rightarrow(Y,\omega_{1}) be a homeomorphism of Hamiltonian structures, where ω0,ω1\omega_{0},\omega_{1} are smooth. We say that ψ\psi is transversely smooth if it is transversely smooth in smooth local charts.

Lemma 6.15.

For i{1,2}i\in\{1,2\}, let (Yi,ωi)(Y_{i},\omega_{i}) be a closed 33-manifold with a smooth Hamiltonian structure. Suppose that φ0:(Y1,ω1)(Y2,ω2)\varphi_{0}:(Y_{1},\omega_{1})\to(Y_{2},\omega_{2}) is a transversely smooth homeomorphism of Hamiltonian structures. Then, there exists a diffeomorphism of Hamiltonian structures φ1:(Y1,ω1)(Y2,ω2)\varphi_{1}:(Y_{1},\omega_{1})\rightarrow(Y_{2},\omega_{2}) which is isotopic to φ0\varphi_{0} through homeomorphisms of Hamiltonian structures φt:(Y1,ω1)(Y2,ω2)\varphi_{t}:(Y_{1},\omega_{1})\rightarrow(Y_{2},\omega_{2}). Moreover, the homeomorphisms φt\varphi_{t} all induce the same homeomorphism between the leaf spaces of (Y1,ω1)(Y_{1},\omega_{1}) and (Y2,ω2)(Y_{2},\omega_{2}).

Proof.

Using smooth local charts on (Y1,ω1)(Y_{1},\omega_{1}) and (Y2,ω2)(Y_{2},\omega_{2}), one can reduce the proof of the lemma to the following claim.

Claim 6.16.

Consider open intervals IJI\Subset J\subset{\mathbb{R}} and open balls BC2B\Subset C\subset{\mathbb{R}}^{2}. Let UJ×CU\subset J\times C be an open subset. Let pr2:3=×22\operatorname{pr}_{2}:{\mathbb{R}}^{3}={\mathbb{R}}\times{\mathbb{R}}^{2}\rightarrow{\mathbb{R}}^{2} denote the projection onto the second factor. Consider an embedding ψ0:J×C3\psi_{0}:J\times C\hookrightarrow{\mathbb{R}}^{3} with the property that pr2ψ0=pr2|J×C\operatorname{pr}_{2}\circ\psi_{0}=\operatorname{pr}_{2}|_{J\times C}. Assume that the restriction of ψ0\psi_{0} to UU is a smooth embedding.

Then there exists a family of embeddings ψt:J×C3\psi_{t}:J\times C\hookrightarrow{\mathbb{R}}^{3}, for t[0,1]t\in[0,1], satisfying pr2ψt=pr2\operatorname{pr}_{2}\circ\psi_{t}=\operatorname{pr}_{2} and agreeing with ψ0\psi_{0} outside some compact subset of J×CJ\times C, such that the restriction of ψ1\psi_{1} to U(I×B)U\cup(I\times B) is a smooth embedding.

In order to see that this claim implies the lemma, note that since φ0\varphi_{0} is transversely smooth, in suitable smooth local Darboux coordinates it precisely takes the form of an embedding ψ0\psi_{0} as in the claim. Therefore, the assertion of the claim allows us to go through a finite covering of (Y1,ω1)(Y_{1},\omega_{1}) by Darboux charts and step by step make φ0\varphi_{0} smooth in each of the charts.

We briefly outline the proof of the claim. The embedding ψ0:J×C3\psi_{0}:J\times C\hookrightarrow{\mathbb{R}}^{3} is of the form

ψ0(s,z)=(f0(s,z),z)\psi_{0}(s,z)=(f_{0}(s,z),z)

where (s,z)(s,z) denotes a point in ×2{\mathbb{R}}\times{\mathbb{R}}^{2}. For each fixed zCz\in C, the function f0(,z)f_{0}(\cdot,z) is strictly increasing. The restriction of f0f_{0} to UU is smooth and the partial derivative sf0\partial_{s}f_{0} is strictly positive in this region. We can find a function f1:J×Cf_{1}:J\times C\rightarrow{\mathbb{R}} agreeing with f0f_{0} outside some compact subset of J×CJ\times C such that f1(,z)f_{1}(\cdot,z) is strictly increasing for every zz and such that the restriction of f1f_{1} to U(I×B)U\cup(I\times B) is smooth and has strictly positive partial derivative sf1\partial_{s}f_{1} in this region. Moreover, we can find a path of functions ft(s,z)f_{t}(s,z) connecting f0f_{0} to f1f_{1} such that each ftf_{t} agrees with f0f_{0} outside a compact subset of J×CJ\times C and such that ft(,z)f_{t}(\cdot,z) is strictly increasing for all zz. We can then define

ψt(s,z)=(ft(s,z),z),\psi_{t}(s,z)=(f_{t}(s,z),z),

which satisfies the requirements of the claim. ∎

Remark 6.17.

Let (Y,ω)(Y,\omega) be a 33-manifold equipped with a smooth Hamiltonian structure and let 𝒫=(Σ,ωΣ,α,(φt)t[0,1])\mathcal{P}=(\Sigma,\omega_{\Sigma},\alpha,(\varphi^{t})_{t\in[0,1]}) be a C0C^{0} ω\omega-plug. Using Proposition 2.2, it is possible to find an area-preserving homeomorphism γ\gamma of (Σ,ωΣ)(\Sigma,\omega_{\Sigma}), not necessarily compactly supported, such that α(id×γ):((0,1)×Σ,ωΣ)(Y,ω)\alpha\circ(\operatorname{id}\times\gamma):((0,1)\times\Sigma,\omega_{\Sigma})\hookrightarrow(Y,\omega) is a transversely smooth embedding of Hamiltonian structures. By an argument similar to the proof of Lemma 6.15, it is possible to modify the embedding α(id×γ)\alpha\circ(\operatorname{id}\times\gamma) by slightly sliding it along the leaves of the characteristic foliation of (Y,ω)(Y,\omega) to obtain a smooth embedding of Hamiltonian structures α:((0,1)×Σ,ωΣ)(Y,ω)\alpha^{\prime}:((0,1)\times\Sigma,\omega_{\Sigma})\hookrightarrow(Y,\omega). Define the ω\omega-plug

𝒫(Σ,ωΣ,α,(γ1φtγ)t[0,1]).\mathcal{P^{\prime}}\coloneqq(\Sigma,\omega_{\Sigma},\alpha^{\prime},(\gamma^{-1}\circ\varphi^{t}\circ\gamma)_{t\in[0,1]}).

One can check, using Lemma 5.8, that ω#𝒫\omega\#\mathcal{P}^{\prime} is homeomorphic to ω#𝒫\omega\#\mathcal{P} via a homeomorphism isotopic to the identity. Moreover, it follows from the naturality of the \mathcal{R}-valued extension of the Calabi homomorphism that Cal¯(𝒫)=Cal¯(𝒫)\overline{\operatorname{Cal}}(\mathcal{P}^{\prime})=\overline{\operatorname{Cal}}(\mathcal{P}). The upshot of this remark is that we can replace any C0C^{0} ω\omega-plug 𝒫\mathcal{P} by an ω\omega-plug whose underlying embedding α:((0,1)×Σ,ωΣ)(Y,ω)\alpha:((0,1)\times\Sigma,\omega_{\Sigma})\hookrightarrow(Y,\omega) is smooth.

Proof of Theorem 6.11.

First, let us consider the special case where 𝒫2\mathcal{P}_{2} is a trivial plug. In this simplified situation, we will construct a smooth plug 𝒬1\mathcal{Q}_{1} satisfying Cal(𝒬1)=Cal¯(𝒫1)\operatorname{Cal}(\mathcal{Q}_{1})=\overline{\operatorname{Cal}}(\mathcal{P}_{1}) and a diffeomorphism of Hamiltonian structures

φ:(Y1,ω1#𝒬1)(Y2,ω2)\varphi:(Y_{1},\omega_{1}\#\mathcal{Q}_{1})\rightarrow(Y_{2},\omega_{2})

which is isotopic to ψ\psi through homeomorphisms. Note that this in particular implies that Cal¯(𝒫1)\overline{\operatorname{Cal}}(\mathcal{P}_{1})\in{\mathbb{R}}\subset\mathcal{R}.

Let UY1U\subset Y_{1} denote the image of the plug 𝒫1\mathcal{P}_{1}. After possibly adding trivial components to the plug 𝒫1\mathcal{P}_{1}, we can assume that UU is exhaustive in the sense of Definition 6.4. This is possible by Lemma 6.6. Note that the addition of trivial components to 𝒫1\mathcal{P}_{1} does not change the Calabi invariant of the plug. Pick an open neighborhood VV of Y1UY_{1}\setminus U such that ω1#𝒫1\omega_{1}\#\mathcal{P}_{1} agrees with ω1\omega_{1} on VV and is thus smooth in this region. We apply Lemma 6.13. This yields smooth open surfaces (Σ1,σ1),,(Σn,σn)(\Sigma_{1},\sigma_{1}),\ldots,(\Sigma_{n},\sigma_{n}), relatively compact open subsets SiΣiS_{i}\Subset\Sigma_{i}, and smooth embeddings of Hamiltonian structures ιi:(CΣi,σi)(V,ω1#𝒫1)\iota_{i}:(C_{\Sigma_{i}},\sigma_{i})\hookrightarrow(V,\omega_{1}\#\mathcal{P}_{1}) satisfying all the properties listed in Lemma 6.13.

In what follows, we will construct C0C^{0} ω1#𝒫1\omega_{1}\#\mathcal{P}_{1}-plugs 𝒵1,𝒵1+,,𝒵n,𝒵n+\mathcal{Z}_{1}^{-},\mathcal{Z}_{1}^{+},\ldots,\mathcal{Z}_{n}^{-},\mathcal{Z}_{n}^{+} which are of the form

𝒵i=(Σi,σi,ιi|CΣi,φit)\displaystyle\mathcal{Z}_{i}^{-}=(\Sigma_{i},\sigma_{i},\iota_{i}|_{C_{\Sigma_{i}}^{-}},\;\varphi_{i}^{t}\,)
𝒵i+=(Σi,σi,ιi|CΣi+,(φit)1)\displaystyle\mathcal{Z}_{i}^{+}=(\Sigma_{i},\sigma_{i},\iota_{i}|_{C_{\Sigma_{i}}^{+}},(\varphi_{i}^{t})^{-1})

where φit\varphi_{i}^{t} is an isotopy in Ham¯(Σi)\overline{\operatorname{Ham}}(\Sigma_{i}) which will be chosen below. By the properties listed in Lemma 6.13, the 2n2n plugs 𝒵i±\mathcal{Z}_{i}^{\pm} are pairwise disjoint and contained in UVU\cap V. Let Ωi\Omega_{i} denote the C0C^{0} Hamiltonian structure obtained by inserting the plugs 𝒵j±\mathcal{Z}_{j}^{\pm} for 1ji1\leq j\leq i into ω1#𝒫1\omega_{1}\#\mathcal{P}_{1}. We will show that for appropriately chosen φit\varphi_{i}^{t} there exists a homeomorphism of C0C^{0} Hamiltonian structures

ψi:(Y1,Ωi)(Y2,ω2)\psi_{i}:(Y_{1},\Omega_{i})\rightarrow(Y_{2},\omega_{2})

whose restriction to jiιj(CSj0)U¯\bigcup_{j\leq i}\iota_{j}(C_{S_{j}}^{0})\setminus\overline{U} is transversely smooth. Note that the assertion on transverse smoothness of ψi\psi_{i} makes sense because the restriction of Ωi\Omega_{i} to the complement of U¯\overline{U} is a smooth Hamiltonian structure.

We will now describe an inductive procedure for constructing φit\varphi_{i}^{t} and ψi\psi_{i} as above. Let 1in1\leq i\leq n and suppose that φk\varphi_{k} and ψk\psi_{k} have already been constructed for all k<ik<i. We explain how to construct φi\varphi_{i} and ψi\psi_{i}. Consider the composition

ψi1ιi:(CΣi,σi)(Y2,ω2),\psi_{i-1}\circ\iota_{i}:(C_{\Sigma_{i}},\sigma_{i})\hookrightarrow(Y_{2},\omega_{2}),

which is a topological embedding of C0C^{0} Hamiltonian structures. Note that the leaf space of (im(ψi1ιi),ω2)(\operatorname{im}(\psi_{i-1}\circ\iota_{i}),\omega_{2}) has the structure of a smooth 22-dimensional manifold with area form. We abbreviate it by (,λ)(\mathcal{L},\lambda). The embedding ψi1ιi\psi_{i-1}\circ\iota_{i} induces an area-preserving homeomorphism γ:(Σi,σi)(,λ)\gamma:(\Sigma_{i},\sigma_{i})\rightarrow(\mathcal{L},\lambda). Let pr2:CΣi=(0,3)×ΣiΣi\operatorname{pr}_{2}:C_{\Sigma_{i}}=(0,3)\times\Sigma_{i}\rightarrow\Sigma_{i} denote the projection onto the second factor. Note that the restriction of γ\gamma to the set

Epr2(ιi1(j<iιj(CSj0))U¯)E\coloneqq\operatorname{pr}_{2}(\iota_{i}^{-1}(\bigcup_{j<i}\iota_{j}(C_{S_{j}}^{0}))\setminus\overline{U})

is smooth by the transverse smoothness assumption on ψk\psi_{k} for k<ik<i.

Using Proposition 2.2, one can find an area-preserving homeomorphism γ~:(Σi,σi)(,λ)\tilde{\gamma}:(\Sigma_{i},\sigma_{i})\rightarrow(\mathcal{L},\lambda) which agrees with γ\gamma outside some compact set and whose restriction to SiES_{i}\cup E is smooth. Indeed, pick a continuous function ρ:Si0\rho:S_{i}\rightarrow{\mathbb{R}}_{\geq 0} which decays to zero towards the boundary of SiS_{i}, whose zero set ρ1({0})\rho^{-1}(\{0\}) is contained in EE, and which has the property that there exists an open neighborhood FF of SiE\partial S_{i}\cap E in EE such that FSiρ1({0})F\cap S_{i}\subset\rho^{-1}(\{0\}). Now apply Proposition 2.2 to γ|Si\gamma|_{S_{i}} with this function ρ\rho. This yields an area-preserving diffeomorphism γ~:Siγ(Si)\tilde{\gamma}:S_{i}\rightarrow\gamma(S_{i}). Since ρ\rho decays to zero near the boundary of SiS_{i}, we can extend γ~\tilde{\gamma} to an area-preserving homeomorphism γ~:Σi\tilde{\gamma}:\Sigma_{i}\rightarrow\mathcal{L} by setting it equal to γ\gamma outside of SiS_{i}. Since FSiρ1({0})F\cap S_{i}\subset\rho^{-1}(\{0\}), we see that γ~\tilde{\gamma} agrees with γ\gamma on FF and is therefore smooth in this region. We conclude that the restriction of γ~\tilde{\gamma} to SiES_{i}\cup E is smooth.

Moreover, we can pick γ~\tilde{\gamma} such that γ~1γ\tilde{\gamma}^{-1}\circ\gamma is in Ham¯(Σi)\overline{\operatorname{Ham}}(\Sigma_{i}). Indeed, note that since γ~\tilde{\gamma} is a C0C^{0} approximation of γ\gamma, the homeomorphism γ~1γ\tilde{\gamma}^{-1}\circ\gamma is contained in the identity component Homeo0(Σi,σi)\operatorname{Homeo}_{0}(\Sigma_{i},\sigma_{i}). If γ~1γ\tilde{\gamma}^{-1}\circ\gamma is not in Ham¯(Σi)\overline{\operatorname{Ham}}(\Sigma_{i}), simply pick a compactly supported area-preserving diffeomorphism βDiff0(,λ)\beta\in\operatorname{Diff}_{0}(\mathcal{L},\lambda) such that the mass flow homomorphism on Homeo0(,λ)\operatorname{Homeo}_{0}(\mathcal{L},\lambda) takes the same value on β\beta and γγ~1\gamma\circ\tilde{\gamma}^{-1}. Then replace γ~\tilde{\gamma} by βγ~\beta\circ\tilde{\gamma}.

Since γ~1γHam¯(Σi)\tilde{\gamma}^{-1}\circ\gamma\in\overline{\operatorname{Ham}}(\Sigma_{i}), we can find an isotopy φitHam¯(Σi)\varphi_{i}^{t}\in\overline{\operatorname{Ham}}(\Sigma_{i}) whose time-1 map satisfies

φi1=γ~1γ.\varphi_{i}^{1}=\tilde{\gamma}^{-1}\circ\gamma.

The map ψi\psi_{i} is constructed by modifying ψi1\psi_{i-1} inside ιi(CΣi)\iota_{i}(C_{\Sigma_{i}}), hence it agrees with ψi1\psi_{i-1} outside ιi(CΣi)\iota_{i}(C_{\Sigma_{i}}). We describe the construction of ψi\psi_{i} on each of ιi(CΣi),ιi(CΣi0),ιi(CΣi+)\iota_{i}(C_{\Sigma_{i}}^{-}),\iota_{i}(C_{\Sigma_{i}}^{0}),\iota_{i}(C_{\Sigma_{i}}^{+}). Inside ιi(CΣi0)\iota_{i}(C_{\Sigma_{i}}^{0}), we define ψi\psi_{i} such that

ψiιi(t,z)=ψi1ιi(t,γ1γ~(z))for (t,z)CΣi0.\psi_{i}\circ\iota_{i}(t,z)=\psi_{i-1}\circ\iota_{i}(t,\gamma^{-1}\circ\tilde{\gamma}(z))\qquad\text{for $(t,z)\in C_{\Sigma_{i}}^{0}$.}

Inside ιi(CΣi)\iota_{i}(C_{\Sigma_{i}}^{-}), we define ψi\psi_{i} such that

ψiιi(t,z)=ψi1ιi(t,(φit)1(z))for (t,z)CΣi.\psi_{i}\circ\iota_{i}(t,z)=\psi_{i-1}\circ\iota_{i}(t,(\varphi_{i}^{t})^{-1}(z))\qquad\text{for $(t,z)\in C_{\Sigma_{i}}^{-}$.}

Inside ιi(CΣi+)\iota_{i}(C_{\Sigma_{i}}^{+}), we define ψi\psi_{i} such that

ψiιi(t,z)=ψi1ιi(t,φit2(z))for (t,z)CΣi+.\psi_{i}\circ\iota_{i}(t,z)=\psi_{i-1}\circ\iota_{i}(t,\varphi_{i}^{t-2}(z))\qquad\text{for $(t,z)\in C_{\Sigma_{i}}^{+}$.}

One can verify that the above yield a well-defined map ψi\psi_{i} on ιi(CΣi)\iota_{i}(C_{\Sigma_{i}}) which coincides with ψi1\psi_{i-1} near the boundary of ιi(CΣi)\iota_{i}(C_{\Sigma_{i}}). The plugs 𝒵i±\mathcal{Z}_{i}^{\pm} were defined precisely such that ψi\psi_{i} is a homeomorphism of C0C^{0} Hamiltonian structures ψi:(Y1,Ωi)(Y2,ω2)\psi_{i}:(Y_{1},\Omega_{i})\rightarrow(Y_{2},\omega_{2}). Finally, note that the map Σi\Sigma_{i}\rightarrow\mathcal{L} induced by ψiιi|CΣi0\psi_{i}\circ\iota_{i}|_{C_{\Sigma_{i}}^{0}} is precisely given by γ~\tilde{\gamma}. Since the restriction of γ~\tilde{\gamma} to SiES_{i}\cup E is smooth, we see that the restriction of ψi\psi_{i} to jiιj(CΣj0)U¯\bigcup_{j\leq i}\iota_{j}(C_{\Sigma_{j}}^{0})\setminus\overline{U} is transversely smooth. Since ψi1\psi_{i-1} is isotopic to ψ\psi by assumption, so is ψi\psi_{i}. This concludes our construction of ψi\psi_{i} and the isotopy φit\varphi_{i}^{t}.

To summarize, at this point we have a homeomorphism of C0C^{0} Hamiltonian structures

ψn:(Y1,Ωn)(Y2,ω2)\psi_{n}:(Y_{1},\Omega_{n})\rightarrow(Y_{2},\omega_{2})

which, by property 1 in Lemma 6.13, is transversely smooth on Y1U¯Y_{1}\setminus\overline{U}. Moreover, Ωn\Omega_{n} is obtained from ω1#𝒫1\omega_{1}\#\mathcal{P}_{1} via the insertion of the plugs 𝒵1±,,𝒵n±\mathcal{Z}_{1}^{\pm},\ldots,\mathcal{Z}_{n}^{\pm}.

Since the plugs 𝒵1±,,𝒵n±\mathcal{Z}_{1}^{\pm},\ldots,\mathcal{Z}_{n}^{\pm} are pairwise disjoint and all contained in UVU\cap V, there exist a single ω1\omega_{1}-plug 𝒬1\mathcal{Q}_{1}, contained in UU, and a homeomorphism χ:(Y1,ω1#𝒬1)(Y1,Ωn)\chi:(Y_{1},\omega_{1}\#\mathcal{Q}_{1})\rightarrow(Y_{1},\Omega_{n}) which is isotopic to the identity and compactly supported inside UU. The Calabi invariant of 𝒬1\mathcal{Q}_{1} is given by

Cal¯(𝒬1)=Cal¯(𝒫1)+i(Cal¯(𝒵i)+Cal¯(𝒵i+))=Cal¯(𝒫1),\overline{\operatorname{Cal}}(\mathcal{Q}_{1})=\overline{\operatorname{Cal}}(\mathcal{P}_{1})+\sum_{i}(\overline{\operatorname{Cal}}(\mathcal{Z}_{i}^{-})+\overline{\operatorname{Cal}}(\mathcal{Z}_{i}^{+}))=\overline{\operatorname{Cal}}(\mathcal{P}_{1}),

where we use naturality of the tautological extension of the Calabi homomorphism and the identity

Cal¯(𝒵i)+Cal¯(𝒵i+)=Cal¯(φi)+Cal¯(φi1)=0.\overline{\operatorname{Cal}}(\mathcal{Z}_{i}^{-})+\overline{\operatorname{Cal}}(\mathcal{Z}_{i}^{+})=\overline{\operatorname{Cal}}(\varphi_{i})+\overline{\operatorname{Cal}}(\varphi_{i}^{-1})=0.

Like ψn\psi_{n}, the homeomorphism

ψnχ:(Y1,ω1#𝒬1)(Y2,ω2)\psi_{n}\circ\chi:(Y_{1},\omega_{1}\#\mathcal{Q}_{1})\rightarrow(Y_{2},\omega_{2})

is transversely smooth on the complement of U¯\overline{U}. Write 𝒬1:=(Σ,σ,α,(θt)t[0,1])\mathcal{Q}_{1}:=(\Sigma,\sigma,\alpha,(\theta^{t})_{t\in[0,1]}) for the data of the plug 𝒬1\mathcal{Q}_{1}. Note that we can assume that the embedding α\alpha is smooth with respect to the smooth Hamiltonian structure ω1\omega_{1}, see Remark 6.17.

Claim 6.18.

The map θ1\theta^{1} is smooth.

Proof.

Pick δ>0\delta>0 sufficiently small such that θt\theta^{t} is equal to id\operatorname{id} for t[0,2δ)t\in[0,2\delta) and equal to θ1\theta^{1} for t(12δ,1]t\in(1-2\delta,1]. Note that ω1#𝒬1\omega_{1}\#\mathcal{Q}_{1} agrees with ω1\omega_{1} on α((0,2δ)×Σ)\alpha((0,2\delta)\times\Sigma) and on α((12δ)×Σ)\alpha((1-2\delta)\times\Sigma) and that ψnχ\psi_{n}\circ\chi is transversely smooth on these sets.

Set Sα({δ}×Σ)S_{-}\coloneqq\alpha(\{\delta\}\times\Sigma) and S+α({1δ}×Σ)S_{+}\coloneqq\alpha(\{1-\delta\}\times\Sigma). Traversing im(α)\operatorname{im}(\alpha) along the characteristic foliation of ω1#𝒬1\omega_{1}\#\mathcal{Q}_{1} induces a homeomorphism

f:SS+.f:S_{-}\rightarrow S_{+}.

Similarly, traversing im(ψnχα)\operatorname{im}(\psi_{n}\circ\chi\circ\alpha) along the the characteristic foliation of ω2\omega_{2} induces a homeomorphism

f:ψnχ(S))ψnχ(S+).f^{\prime}:\psi_{n}\circ\chi(S_{-}))\rightarrow\psi_{n}\circ\chi(S_{+}).

These homeomorphisms fit into the following commutative diagram:

ψnχ(S){\psi_{n}\circ\chi(S_{-})}ψnχ(S+){\psi_{n}\circ\chi(S_{+})}S{S_{-}}S+{S_{+}}{δ}×Σ{\{\delta\}\times\Sigma}{1δ}×Σ{\{1-\delta\}\times\Sigma}Σ{\Sigma}Σ{\Sigma}f\scriptstyle{f^{\prime}}f\scriptstyle{f}ψnχ\scriptstyle{\psi_{n}\circ\chi}ψnχ\scriptstyle{\psi_{n}\circ\chi}α\scriptstyle{\alpha}α\scriptstyle{\alpha}\scriptstyle{\cong}θ1\scriptstyle{\theta^{1}}\scriptstyle{\cong}

Since ψnχ\psi_{n}\circ\chi is transversely smooth on a neighborhood of SS_{-} and S+S_{+} and ω2\omega_{2} is a smooth Hamiltonian structure, we see that θ1\theta^{1} is smooth. ∎

As a consequence of Claim 6.18, we may modify the plug 𝒬1\mathcal{Q}_{1} by replacing θt\theta^{t} with a smooth Hamiltonian isotopy with the same time-1 map is θ1\theta^{1}. By Lemma 5.7, modifying the plug in this way yields a smooth Hamiltonian structure homeomorphic to ω1#𝒬1\omega_{1}\#\mathcal{Q}_{1} via a homeomorphism supported in UU and isotopic to the identity. Moreover, the two plugs have the same Calabi invariant. We will refer to this new smooth plug by the same symbol 𝒬1\mathcal{Q}_{1}. We then have a homeomorphism of Hamiltonian structures

φ:(Y1,ω1#𝒬1)(Y2,ω2)\varphi:(Y_{1},\omega_{1}\#\mathcal{Q}_{1})\rightarrow(Y_{2},\omega_{2})

which is isotopic to ψ\psi and transversely smooth outside of U¯\overline{U}. Since both ω1#𝒬1\omega_{1}\#\mathcal{Q}_{1} and ω2\omega_{2} are smooth and no characteristic leaf of ω1#𝒬1\omega_{1}\#\mathcal{Q}_{1} is trapped in UU, we can conclude that φ\varphi is transversely smooth everywhere. By Lemma 6.15, we can replace φ\varphi by a diffeomorphism of Hamiltonian structures, still isotopic to ψ\psi. This concludes the proof of Theorem 6.11 in the special case that 𝒫2\mathcal{P}_{2} is a trivial plug.

It remains to treat the general case. Let 𝒫¯2\overline{\mathcal{P}}_{2} be the inverse plug of 𝒫2\mathcal{P}_{2} and let ψ𝒫¯2\psi^{*}\overline{\mathcal{P}}_{2} be its pull back via ψ\psi; see Equations (4.1) and (5.2). Then, as noted in (5.3), ψ\psi is a homeomorphism of C0C^{0} Hamiltonian structures

ψ:(Y1,ω1#𝒫1#ψ𝒫¯2)(Y2,ω2#𝒫2#𝒫¯2)=(Y2,ω2).\psi:(Y_{1},\omega_{1}\#\mathcal{P}_{1}\#\psi^{*}\overline{\mathcal{P}}_{2})\rightarrow(Y_{2},\omega_{2}\#\mathcal{P}_{2}\#\overline{\mathcal{P}}_{2})=(Y_{2},\omega_{2}).

We also have

Cal¯(ψ𝒫¯2)=Cal¯(𝒫¯2)=Cal¯(𝒫2).\overline{\operatorname{Cal}}(\psi^{*}\overline{\mathcal{P}}_{2})=\overline{\operatorname{Cal}}(\overline{\mathcal{P}}_{2})=-\overline{\operatorname{Cal}}(\mathcal{P}_{2}).

As we explain below, we can slide the plug ψ𝒫¯2\psi^{*}\overline{\mathcal{P}}_{2} along the characteristic foliation of ω1#𝒫1\omega_{1}\#\mathcal{P}_{1} so that it becomes disjoint from 𝒫1\mathcal{P}_{1}, as described in Lemma 5.8.

Let 𝒮\mathcal{S} denote the resulting plug disjoint from 𝒫1\mathcal{P}_{1}. It is both an ω1\omega_{1}- and an ω1#𝒫1\omega_{1}\#\mathcal{P}_{1}-plug and has Calabi invariant Cal¯(𝒮)=Cal¯(𝒫2)\overline{\operatorname{Cal}}(\mathcal{S})=-\overline{\operatorname{Cal}}(\mathcal{P}_{2}). Let 𝒫1𝒮\mathcal{P}_{1}\sqcup\mathcal{S} be the ω1\omega_{1}-plug given by the disjoint union of 𝒫1\mathcal{P}_{1} and 𝒮\mathcal{S}. By Lemma 5.8, (Y1,ω1#𝒫1#ψ𝒫¯2)(Y_{1},\omega_{1}\#\mathcal{P}_{1}\#\psi^{*}\overline{\mathcal{P}}_{2}) is homeomorphic to (Y1,ω1#(𝒫1𝒮))(Y_{1},\omega_{1}\#(\mathcal{P}_{1}\sqcup\mathcal{S})) via a homeomorphism isotopic to the identity. Thus (Y1,ω1#(𝒫1𝒮))(Y_{1},\omega_{1}\#(\mathcal{P}_{1}\sqcup\mathcal{S})) is homeomorphic to (Y2,ω2)(Y_{2},\omega_{2}) via a homeomorphism isotopic to ψ\psi. By the special case of Theorem 6.11 already treated above, we can find a smooth ω1\omega_{1}-plug 𝒬1\mathcal{Q}_{1} such that there exists a diffeomorphism of smooth Hamiltonian structures φ:(Y1,ω1#𝒬1)(Y2,ω2)\varphi:(Y_{1},\omega_{1}\#\mathcal{Q}_{1})\rightarrow(Y_{2},\omega_{2}), still isotopic to ψ\psi, and such that

Cal(𝒬1)=Cal¯(𝒫1𝒮)=Cal¯(𝒫1)Cal¯(𝒫2).\operatorname{Cal}(\mathcal{Q}_{1})=\overline{\operatorname{Cal}}(\mathcal{P}_{1}\sqcup\mathcal{S})=\overline{\operatorname{Cal}}(\mathcal{P}_{1})-\overline{\operatorname{Cal}}(\mathcal{P}_{2}).

Now simply define 𝒬2\mathcal{Q}_{2} to be the empty plug. Then the tuple (𝒬1,𝒬2,φ)(\mathcal{Q}_{1},\mathcal{Q}_{2},\varphi) clearly satisfies all assertions of Theorem 6.11.

It thus remains to explain why we can always reduce to a situation where the plug ψ𝒫¯2\psi^{*}\overline{\mathcal{P}}_{2} can be made disjoint from 𝒫1\mathcal{P}_{1} by sliding it along the characteristic foliation of ω1#𝒫1\omega_{1}\#\mathcal{P}_{1}.

For i{1,2}i\in\{1,2\}, write 𝒫i=(Σi,ωΣi,αi,(φit)t[0,1])\mathcal{P}_{i}=(\Sigma_{i},\omega_{\Sigma_{i}},\alpha_{i},(\varphi_{i}^{t})_{t\in[0,1]}). Note that by Remark 6.17, we can assume that αi\alpha_{i} are smooth embeddings of Hamiltonian structures. After possibly shrinking the images of 𝒫i\mathcal{P}_{i} as in Remark 5.6, we can assume that there exist surfaces (Σi,ωΣi)(\Sigma_{i}^{\prime},\omega_{\Sigma_{i}^{\prime}}) containing (Σi,ωΣi)(\Sigma_{i},\omega_{\Sigma_{i}}) as relatively compact open subsets such that αi\alpha_{i} extends to a smooth embedding of Hamiltonian structures

αi:((1,2)×Σi,ωΣi)(Yi,ωi).\alpha_{i}^{\prime}:((-1,2)\times\Sigma_{i}^{\prime},\omega_{\Sigma_{i}^{\prime}})\hookrightarrow(Y_{i},\omega_{i}).

Consider a compact surface with smooth boundary SΣ1S\subset\Sigma_{1}^{\prime} containing Σ¯1\overline{\Sigma}_{1} in its interior. Let t0(1,2)t_{0}\in(-1,2) and let U(1,2)×Σ1U\subset(-1,2)\times\Sigma_{1}^{\prime} be an arbitrary open neighborhood of {t0}×S\{t_{0}\}\times S. Then there exists a compactly supported homeomorphism of C0C^{0} Hamiltonian structures

Ψ:((1,2)×Σ1,ωΣ1)((1,2)×Σ1,ωΣ1)\Psi:((-1,2)\times\Sigma_{1}^{\prime},\omega_{\Sigma_{1}^{\prime}})\rightarrow((-1,2)\times\Sigma_{1}^{\prime},\omega_{\Sigma_{1}^{\prime}})

such that [0,1]×Σ¯1Ψ(U)[0,1]\times\overline{\Sigma}_{1}\subset\Psi(U). Indeed, any compactly supported homeomorphism of ((1,2)×Σ1,ωΣ1)((-1,2)\times\Sigma_{1}^{\prime},\omega_{\Sigma_{1}^{\prime}}) takes the form Ψ(t,p)=(fp(t),p)\Psi(t,p)=(f_{p}(t),p) for a family of compactly supported homeomorphisms fpf_{p} of (1,2)(-1,2) which agree with the identity for pp outside some compact subset of Σ1\Sigma_{1}^{\prime}. One can arrange fpf_{p} such that, for every pSp\in S, the homeomorphism fpf_{p} maps an arbitrarily small neighborhood of t0t_{0} to the interval (1/2,3/2)(-1/2,3/2). The image of UU under the resulting homeomorphism Ψ\Psi will then contain [0,1]×Σ¯1[0,1]\times\overline{\Sigma}_{1}, as desired. Let us also point out that the group of compactly supported homeomorphisms of ((1,2)×Σ1,ωΣ1)((-1,2)\times\Sigma_{1}^{\prime},\omega_{\Sigma_{1}^{\prime}}) is connected. In particular, Ψ\Psi is isotopic to the identity through compactly supported homeomorphisms of C0C^{0} Hamiltonian structures.

We can extend the isotopy (φ1t)t[0,1](\varphi_{1}^{t})_{t\in[0,1]} in Ham¯(Σ1)\overline{\operatorname{Ham}}(\Sigma_{1}) to an isotopy (φ1t)t(1,2)(\varphi_{1}^{t})_{t\in(-1,2)} in Ham¯(Σ1)\overline{\operatorname{Ham}}(\Sigma_{1}^{\prime}) which is equal to the identity outside of Σ1\Sigma_{1}, and agrees with the identity for t<0t<0 and with φ11\varphi_{1}^{1} for t>1t>1. We can then define the homeomorphism

Φ1:(1,2)×Σ1(1,2)×Σ1Φ1(t,p)(t,φ1t(p)),\Phi_{1}:(-1,2)\times\Sigma_{1}^{\prime}\rightarrow(-1,2)\times\Sigma_{1}^{\prime}\qquad\Phi_{1}(t,p)\coloneqq(t,\varphi_{1}^{t}(p)),

see Subsection 5.2. Recall that by definition, α1\alpha_{1}^{\prime} is an embedding of C0C^{0} Hamiltonian structures

αi:((1,2)×Σ1,(Φ1)ωΣ1)(Y1,ω1#𝒫1).\alpha_{i}^{\prime}:((-1,2)\times\Sigma_{1}^{\prime},(\Phi_{1})_{*}\omega_{\Sigma_{1}^{\prime}})\hookrightarrow(Y_{1},\omega_{1}\#\mathcal{P}_{1}).

The homeomorphism Φ1\Phi_{1} leaves {t0}×S\{t_{0}\}\times S invariant. The above discussion thus implies that for every open neighborhood UU of {t0}×S\{t_{0}\}\times S, there exists a compactly supported homeomorphism Ψ\Psi of ((1,2)×Σ1,(Φ1)ωΣ1)((-1,2)\times\Sigma_{1}^{\prime},(\Phi_{1})_{*}\omega_{\Sigma_{1}^{\prime}}) isotopic to the identity through such homeomorphisms such that Ψ(U)\Psi(U) contains [0,1]×Σ¯1[0,1]\times\overline{\Sigma}_{1}. From this observation, we obtain the following statement: Suppose that the closure of the image of ψ𝒫¯2\psi^{*}\overline{\mathcal{P}}_{2} is disjoint from α1({t0}×S)\alpha_{1}^{\prime}(\{t_{0}\}\times S). Then one can slide the plug ψ𝒫¯2\psi^{*}\overline{\mathcal{P}}_{2} along the characteristic foliation of ω1#𝒫1\omega_{1}\#\mathcal{P}_{1} to make it disjoint from 𝒫1\mathcal{P}_{1}.

We therefore need to explain how to reduce to the case where the closure of the image of ψ𝒫¯2\psi^{*}\overline{\mathcal{P}}_{2} is disjoint from α1({t0}×S)\alpha_{1}^{\prime}(\{t_{0}\}\times S). We set Tψ(α1({t0}×S))T\coloneqq\psi(\alpha_{1}^{\prime}(\{t_{0}\}\times S)). This defines a compact surface in (Y2,ω2#𝒫2)(Y_{2},\omega_{2}\#\mathcal{P}_{2}) which is transverse to the characteristic foliation, in the sense described in the proof of Claim 6.7. In what follows, we identify ((1,2)×Σ2,ωΣ2)((-1,2)\times\Sigma_{2}^{\prime},\omega_{\Sigma_{2}^{\prime}}) with its image under the embedding α2\alpha_{2}^{\prime} and work in these coordinates. Since TT is transverse to the characteristic foliation, we can find finitely many open discs B1,,BnΣ2B_{1},\dots,B_{n}\Subset\Sigma_{2}^{\prime} covering Σ¯2\overline{\Sigma}_{2} and numbers 1<s1<<sn<0-1<s_{1}<\dots<s_{n}<0 such that the closures of the discs {si}×Bi\{s_{i}\}\times B_{i} are disjoint from TT (cf. proof of Claim 6.7). Now, choose embeddings of C0C^{0} Hamiltonian structures

βi:((0,1)×Bi,ωBi)((1,0)×Σ2,ωΣ2)(Y,ω2)\beta_{i}:((0,1)\times B_{i},\omega_{B_{i}})\hookrightarrow((-1,0)\times\Sigma_{2}^{\prime},\omega_{\Sigma_{2}^{\prime}})\subset(Y,\omega_{2})

such that the closures of their images are pairwise disjoint and also disjoint from TT.

Pick a smooth isotopy (φ~2t)t[0,1](\tilde{\varphi}_{2}^{t})_{t\in[0,1]} in Ham(Σ2)\operatorname{Ham}(\Sigma_{2}) which C0C^{0} approximates the continuous isotopy (φ2t)t[0,1](\varphi_{2}^{t})_{t\in[0,1]}; the existence of (φ~2t)(\tilde{\varphi}_{2}^{t}) is well-known and it can be deduced from the following two facts: first, every homeomorphism in Ham¯\overline{\operatorname{Ham}} can be written as a C0C^{0} limit of elements in Ham\operatorname{Ham}; and second, Ham\operatorname{Ham} and Ham¯\overline{\operatorname{Ham}} are both locally path connected; see, for example, [49, Cor. 2] or [50, Lem. 3.2]. We can take γ\gamma to be arbitrarily C0C^{0} close to the identity by choosing better approximations φ~2t\tilde{\varphi}_{2}^{t}. By Proposition 2.1, we can therefore find a fragmentation γ=γnγ1\gamma=\gamma_{n}\circ\cdots\circ\gamma_{1} into Hamiltonian homeomorphisms γiHam¯(Bi)\gamma_{i}\in\overline{\operatorname{Ham}}(B_{i}) which are C0C^{0} close to the identity. For each ii, pick a C0C^{0} small isotopy (γit)t[0,1](\gamma_{i}^{t})_{t\in[0,1]} connecting the identity to γi\gamma_{i}. We define ω2\omega_{2}-plugs

𝒬2(Σ2,ωΣ2,α2,(φ~2t)t[0,1])and𝒵i(Bi,ωBi,βi,(γit)t[0,1]).\mathcal{Q}_{2}\coloneqq(\Sigma_{2},\omega_{\Sigma_{2}},\alpha_{2},(\tilde{\varphi}_{2}^{t})_{t\in[0,1]})\qquad\text{and}\qquad\mathcal{Z}_{i}\coloneqq(B_{i},\omega_{B_{i}},\beta_{i},(\gamma_{i}^{t})_{t\in[0,1]}).

Note that these plugs are disjoint. Moreover, 𝒬2\mathcal{Q}_{2} is a smooth plug because both α2\alpha_{2} and φ~2t\tilde{\varphi}_{2}^{t} are smooth. We also define the ω2\omega_{2}-plug 𝒵\mathcal{Z} to be the disjoint union of the plugs 𝒵i\mathcal{Z}_{i}. Using Lemma 5.7 and the identity φ21=φ~21γn1γ11\varphi_{2}^{1}=\tilde{\varphi}_{2}^{1}\circ\gamma_{n}^{1}\circ\cdots\circ\gamma_{1}^{1}, we see that there exists a homeomorphism of C0C^{0} Hamiltonian structures

χ:(Y2,ω2#𝒫2)(Y2,ω2#(𝒬2𝒵))\chi:(Y_{2},\omega_{2}\#\mathcal{P}_{2})\rightarrow(Y_{2},\omega_{2}\#(\mathcal{Q}_{2}\sqcup\mathcal{Z}))

isotopic to the identity. Moreover, since φ~2t\tilde{\varphi}_{2}^{t} is C0C^{0} close to φ2t\varphi_{2}^{t} and all γit\gamma_{i}^{t} are C0C^{0} close to the identity, it is clear from the proof of Lemma 5.7 that the homeomorphsim χ\chi can be taken C0C^{0} close to the identity. In particular, we can assume that χ(T)\chi(T) is disjoint from the closure of the image of 𝒵\mathcal{Z}. Now consider the composite

χψ:(Y1,ω1#𝒫1)(Y2,(ω2#𝒬2)#𝒵)\chi\circ\psi:(Y_{1},\omega_{1}\#\mathcal{P}_{1})\rightarrow(Y_{2},(\omega_{2}\#\mathcal{Q}_{2})\#\mathcal{Z})

and note that (χψ)𝒵¯(\chi\circ\psi)^{*}\overline{\mathcal{Z}} is disjoint from α1({t0}×S)\alpha_{1}^{\prime}(\{t_{0}\}\times S). As discussed above, we can slide (χψ)𝒵¯(\chi\circ\psi)^{*}\overline{\mathcal{Z}} away from 𝒫1\mathcal{P}_{1} along the characteristic foliation of ω1#𝒫1\omega_{1}\#\mathcal{P}_{1}. Hence, there exist a smooth ω1\omega_{1}-plug 𝒬1\mathcal{Q}_{1} of Calabi invariant

Cal(𝒬1)=Cal¯(𝒫1)Cal¯(𝒵)\operatorname{Cal}(\mathcal{Q}_{1})=\overline{\operatorname{Cal}}(\mathcal{P}_{1})-\overline{\operatorname{Cal}}(\mathcal{Z}) (6.1)

and a diffeomorphism of Hamiltonian structures φ:(Y1,ω1#𝒬1)(Y2,ω2#𝒬2)\varphi:(Y_{1},\omega_{1}\#\mathcal{Q}_{1})\rightarrow(Y_{2},\omega_{2}\#\mathcal{Q}_{2}) isotopic to χψ\chi\circ\psi and hence also to ψ\psi. Now observe that by naturality of our \mathcal{R}-valued extension of the Calabi homomorphism, we have

Cal¯(𝒫2)=Cal(𝒬2)+Cal¯(𝒵).\overline{\operatorname{Cal}}(\mathcal{P}_{2})=\operatorname{Cal}(\mathcal{Q}_{2})+\overline{\operatorname{Cal}}(\mathcal{Z}).

Together with identity (6.1), this readily yields the desired identity

Cal(𝒬1)Cal¯(𝒫1)=Cal(𝒬2)Cal¯(𝒫2).\operatorname{Cal}(\mathcal{Q}_{1})-\overline{\operatorname{Cal}}(\mathcal{P}_{1})=\operatorname{Cal}(\mathcal{Q}_{2})-\overline{\operatorname{Cal}}(\mathcal{P}_{2}).

This completes the proof of Theorem 6.11 in the general case. ∎

7 Flux and helicity of C0C^{0} Hamiltonian structures

In this section, we use the results of Section 6 to extend the definitions of flux and helicity to C0C^{0} Hamiltonian structures. We prove our main result, Theorem 1.5, on the existence and uniqueness of a universal \mathcal{R}-valued helicity extension. Moreover, we prove Proposition 1.6.

7.1 Flux

We generalize the notion of flux, which we introduced for smooth Hamiltonian structures in Section 4.5, to C0C^{0} Hamiltonian structures.

We begin with the following lemma, which is a straightforward consequence of Theorem 6.11.

Lemma 7.1.

For i{1,2}i\in\{1,2\}, let YiY_{i} be a closed 33-manifold and let ωi\omega_{i} be a smooth Hamiltonian structure on YiY_{i}. Let 𝒫i\mathcal{P}_{i} be a C0C^{0} ωi\omega_{i}-plug and assume that

ψ:(Y1,ω1#𝒫1)(Y2,ω2#𝒫2)\psi:(Y_{1},\omega_{1}\#\mathcal{P}_{1})\rightarrow(Y_{2},\omega_{2}\#\mathcal{P}_{2})

is a homeomorphism of C0C^{0} Hamiltonian structures. Then,

Flux(ω1)=ψFlux(ω2).\operatorname{Flux}(\omega_{1})=\psi^{*}\operatorname{Flux}(\omega_{2}).
Proof.

By Theorem 6.11, we can find a smooth ωi\omega_{i}-plug 𝒬i\mathcal{Q}_{i} for i{1,2}i\in\{1,2\} and a diffeomorphism of Hamiltonian structures

φ:(Y1,ω1#𝒬1)(Y2,ω2#𝒬2)\varphi:(Y_{1},\omega_{1}\#\mathcal{Q}_{1})\rightarrow(Y_{2},\omega_{2}\#\mathcal{Q}_{2})

which is isotopic to ψ\psi. Clearly, we have

Flux(ω1#𝒬1)=φFlux(ω2#𝒬2).\operatorname{Flux}(\omega_{1}\#\mathcal{Q}_{1})=\varphi^{*}\operatorname{Flux}(\omega_{2}\#\mathcal{Q}_{2}).

Recall from Lemma 4.9 that Flux(ωi#𝒬i)=Flux(ωi)\operatorname{Flux}(\omega_{i}\#\mathcal{Q}_{i})=\operatorname{Flux}(\omega_{i}). The assertion of the lemma now follows from the observation that the actions of φ\varphi and ψ\psi on cohomology agree because these two maps are isotopic. ∎

Definition 7.2.

Let Ω\Omega be a C0C^{0} Hamiltonian structure on a closed topological 33-manifold YY. By Theorem 6.1, we can find a smooth Hamiltonian structure ω\omega on YY and a C0C^{0} ω\omega-plug 𝒫\mathcal{P} such that Ω=ω#𝒫\Omega=\mathcal{\omega}\#\mathcal{P}. We define the flux of Ω\Omega to be the cohomology class

Flux¯(Ω)Flux(ω)H2(Y;),\overline{\operatorname{Flux}}(\Omega)\coloneqq\operatorname{Flux}(\omega)\in H^{2}(Y;{\mathbb{R}}),

which is independent of choices by Lemma 7.1. We say that a C0C^{0} Hamiltonian structure Ω\Omega is exact if Flux¯(Ω)=0\overline{\operatorname{Flux}}(\Omega)=0.

Note that if ω\omega is a smooth Hamiltonian structure, then Flux¯(ω)=Flux(ω)\overline{\operatorname{Flux}}(\omega)=\operatorname{Flux}(\omega) because we can simply take 𝒫\mathcal{P} to be the trivial plug in Definition 7.2. Moreover, it is immediate from Lemma 7.1 that if ψ:(Y1,Ω1)(Y2,Ω2)\psi:(Y_{1},\Omega_{1})\rightarrow(Y_{2},\Omega_{2}) is a homeomorphism of C0C^{0} Hamiltonian structures, then Flux¯(Ω1)=ψFlux¯(Ω2)\overline{\operatorname{Flux}}(\Omega_{1})=\psi^{*}\overline{\operatorname{Flux}}(\Omega_{2}).

Lemma 7.3.

Let Ω\Omega be a C0C^{0} Hamiltonian structure on a closed topological manifold YY and let 𝒫\mathcal{P} be an Ω\Omega-plug. Then

Flux¯(Ω)=Flux¯(Ω#𝒫).\overline{\operatorname{Flux}}(\Omega)=\overline{\operatorname{Flux}}(\Omega\#\mathcal{P}).
Proof.

After possibly slightly shrinking the image of the plug 𝒫\mathcal{P}, see Remark 5.6, we can assume that the complement of the image of 𝒫\mathcal{P} contains an exhaustive open set UU. By Theorem 6.1 and Remark 6.5, we can find a smooth Hamiltonian structure ω\omega on YY and a C0C^{0} ω\omega-plug 𝒬\mathcal{Q} whose image is contained in UU such that Ω=ω#𝒬\Omega=\omega\#\mathcal{Q}. Let 𝒵\mathcal{Z} be the disjoint union of the plugs 𝒬\mathcal{Q} and 𝒫\mathcal{P}. Now 𝒵\mathcal{Z} is an ω\omega-plug and ω#𝒵=Ω#𝒫\omega\#\mathcal{Z}=\Omega\#\mathcal{P}. Using the definition of Flux¯\overline{\operatorname{Flux}}, we can therefore compute

Flux¯(Ω#𝒫)=Flux¯(ω#𝒵)=Flux(ω)=Flux¯(ω#𝒬)=Flux¯(Ω).\overline{\operatorname{Flux}}(\Omega\#\mathcal{P})=\overline{\operatorname{Flux}}(\omega\#\mathcal{Z})=\operatorname{Flux}(\omega)=\overline{\operatorname{Flux}}(\omega\#\mathcal{Q})=\overline{\operatorname{Flux}}(\Omega).

7.2 Helicity

The goal of this section is to prove Theorem 1.5.

The uniqueness of the extension ¯\overline{\mathcal{H}} in Theorem 1.5 is a consequence of Theorem 6.1. Indeed, consider an arbitrary exact C0C^{0} Hamiltonian structure Ω\Omega on a closed topological 33-manifold YY. By Theorem 6.1, there exist a smooth Hamiltonian structure ω\omega on YY and an ω\omega-plug 𝒫\mathcal{P} such that Ω=ω#𝒫\Omega=\omega\#\mathcal{P}. By Lemma 7.3, the smooth Hamiltonian structure ω\omega is exact. It follows from properties 1, 2, and 3 in Theorem 1.5 that we must have

¯(Ω)=¯(ω#𝒫)=¯(ω)+Cal¯(𝒫)=(ω)+Cal¯(𝒫).\overline{\mathcal{H}}(\Omega)=\overline{\mathcal{H}}(\omega\#\mathcal{P})=\overline{\mathcal{H}}(\omega)+\overline{\operatorname{Cal}}(\mathcal{P})=\mathcal{H}(\omega)+\overline{\operatorname{Cal}}(\mathcal{P}). (7.1)

This proves uniqueness.

We turn to the proof of existence. We begin with the following lemma, which is a corollary of Theorem 6.11.

Lemma 7.4.

For i{1,2}i\in\{1,2\}, let YiY_{i} be a closed 33-manifold and let ωi\omega_{i} be an exact smooth Hamiltonian structure on YiY_{i}. Let 𝒫i\mathcal{P}_{i} be a C0C^{0} ωi\omega_{i}-plug and assume that

ψ:(Y1,ω1#𝒫1)(Y2,ω2#𝒫2)\psi:(Y_{1},\omega_{1}\#\mathcal{P}_{1})\rightarrow(Y_{2},\omega_{2}\#\mathcal{P}_{2})

is a homeomorphism of C0C^{0} Hamiltonian structures. Then

(ω1)+Cal¯(𝒫1)=(ω2)+Cal¯(𝒫2).\mathcal{H}(\omega_{1})+\overline{\operatorname{Cal}}(\mathcal{P}_{1})=\mathcal{H}(\omega_{2})+\overline{\operatorname{Cal}}(\mathcal{P}_{2}).
Proof.

By Theorem 6.11, we can find a smooth ωi\omega_{i}-plug 𝒬i\mathcal{Q}_{i} for i{1,2}i\in\{1,2\} and a diffeomorphism of Hamiltonian structures

φ:(Y1,ω1#𝒬1)(Y1,ω1#𝒬1)\varphi:(Y_{1},\omega_{1}\#\mathcal{Q}_{1})\rightarrow(Y_{1},\omega_{1}\#\mathcal{Q}_{1})

such that

Cal(𝒬1)Cal¯(𝒫1)=Cal(𝒬2)Cal¯(𝒫2)\operatorname{Cal}(\mathcal{Q}_{1})-\overline{\operatorname{Cal}}(\mathcal{P}_{1})=\operatorname{Cal}(\mathcal{Q}_{2})-\overline{\operatorname{Cal}}(\mathcal{P}_{2})

Since φ\varphi is a diffeomorphism, we clearly have (ω1#𝒬1)=(ω2#𝒬2)\mathcal{H}(\omega_{1}\#\mathcal{Q}_{1})=\mathcal{H}(\omega_{2}\#\mathcal{Q}_{2}). It follows from Lemma 4.10 that (ωi#𝒬i)=(ωi)+Cal(𝒬i)\mathcal{H}(\omega_{i}\#\mathcal{Q}_{i})=\mathcal{H}(\omega_{i})+\operatorname{Cal}(\mathcal{Q}_{i}). The assertion of the lemma is an immediate consequence of these facts. ∎

Let Ω\Omega be an arbitrary exact C0C^{0} Hamiltonian structure on a closed topological 33-manifold YY. By Theorem 6.1, we may write Ω=ω#𝒫\Omega=\omega\#\mathcal{P} for a smooth Hamiltonian structure ω\omega and a C0C^{0} ω\omega-plug 𝒫\mathcal{P}. By Lemma 7.3, the smooth Hamiltonian structure ω\omega is exact. The idea is to use identity (7.1) as a definition, i.e. to set

¯(Ω)(ω)+Cal¯(𝒫).\overline{\mathcal{H}}(\Omega)\coloneqq\mathcal{H}(\omega)+\overline{\operatorname{Cal}}(\mathcal{P}). (7.2)

By Lemma 7.4, this definition is independent of choices. It remains to check that our extension ¯\overline{\mathcal{H}} satisfies properties 1, 2, and 3 in Theorem 1.5. For property 1, observe that if Ω=ω\Omega=\omega is already smooth, we can simply take 𝒫\mathcal{P} to be the trivial plug in (7.2). Property 2 is immediate from Lemma 7.4. Finally, suppose that Ω\Omega is an exact C0C^{0} Hamiltonian structure and that 𝒫\mathcal{P} is an Ω\Omega-plug. After possibly slightly shrinking the image of 𝒫\mathcal{P}, we can assume that the complement of 𝒫\mathcal{P} contains an exhaustive open set UU. By Theorem 6.1 and Remark 6.5, we can write Ω=ω#𝒬\Omega=\omega\#\mathcal{Q} for a smooth Hamiltonian structure ω\omega on YY and a C0C^{0} ω\omega-plug 𝒬\mathcal{Q} whose image is contained in UU. By Lemma 7.3, the Hamiltonian structure ω\omega is exact. The disjoint union 𝒵\mathcal{Z} of 𝒫\mathcal{P} and 𝒬\mathcal{Q} is an ω\omega-plug and we have Ω#𝒫=ω#𝒵\Omega\#\mathcal{P}=\omega\#\mathcal{Z}. Moreover, we have Cal¯(𝒵)=Cal¯(𝒬)+Cal¯(𝒫)\overline{\operatorname{Cal}}(\mathcal{Z})=\overline{\operatorname{Cal}}(\mathcal{Q})+\overline{\operatorname{Cal}}(\mathcal{P}). We can therefore compute

¯(Ω#𝒫)=ω+Cal¯(𝒵)=ω+Cal¯(𝒬)+Cal¯(𝒫)=¯(Ω)+Cal¯(𝒫),\overline{\mathcal{H}}(\Omega\#\mathcal{P})=\mathcal{\omega}+\overline{\operatorname{Cal}}(\mathcal{Z})=\mathcal{\omega}+\overline{\operatorname{Cal}}(\mathcal{Q})+\overline{\operatorname{Cal}}(\mathcal{P})=\overline{\mathcal{H}}(\Omega)+\overline{\operatorname{Cal}}(\mathcal{P}),

verifying property 3. This concludes the proof of Theorem 1.5.∎

7.3 Universality

The goal of this subsection is to prove Proposition 1.6. Let A{\mathbb{R}}\subset A be an arbitrary extension of abelian groups and let ¯A\overline{\mathcal{H}}^{A} be an AA-valued extension of helicity to C0C^{0} Hamiltonian structures. Assume that ¯A\overline{\mathcal{H}}^{A} satisfies the Extension and Invariance properties in Theorem 1.5. Moreover, assume that ¯A\overline{\mathcal{H}}^{A} satisfies plug homogeneity. Our task is to show that there exists a unique group homomorphism p:Ap:\mathcal{R}\rightarrow A over {\mathbb{R}} such that ¯A=p¯\overline{\mathcal{H}}^{A}=p\circ\overline{\mathcal{H}}.

Let (Σ,ωΣ)(\Sigma,\omega_{\Sigma}) be a non-empty connected open surface with area form. Let us begin by defining a map

p~Σ:Ham¯(Σ)A\tilde{p}_{\Sigma}:\overline{\operatorname{Ham}}(\Sigma)\rightarrow A

as follows. Let φHam¯(Σ)\varphi\in\overline{\operatorname{Ham}}(\Sigma) be a Hamiltonian homeomorphism. Pick an arbitrary isotopy φt\varphi^{t} in Ham¯(Σ)\overline{\operatorname{Ham}}(\Sigma) connecting the identity to φ\varphi. Moreover, pick an arbitrary C0C^{0} Hamiltonian structure (Y,Ω)(Y,\Omega) such that there exists an embedding of C0C^{0} Hamiltonian structures α:((0,1)×Σ,ωΣ)(Y,Ω)\alpha:((0,1)\times\Sigma,\omega_{\Sigma})\hookrightarrow(Y,\Omega). Define the Ω\Omega-plug 𝒫(Σ,ωΣ,α,(φt)t[0,1])\mathcal{P}\coloneqq(\Sigma,\omega_{\Sigma},\alpha,(\varphi^{t})_{t\in[0,1]}) and set

p~Σ(φ)¯A(Ω#𝒫)¯A(Ω)A.\tilde{p}_{\Sigma}(\varphi)\coloneqq\overline{\mathcal{H}}^{A}(\Omega\#\mathcal{P})-\overline{\mathcal{H}}^{A}(\Omega)\in A.

By plug homogeneity, this is well-defined and independent of choices.

We claim that p~Σ\tilde{p}_{\Sigma} is a group homomorphism. Indeed, consider two Hamiltonian homeomorphisms φ,ψHam¯(Σ)\varphi,\psi\in\overline{\operatorname{Ham}}(\Sigma). Pick Hamiltonian isotopies (φt)t[0,1](\varphi^{t})_{t\in[0,1]} and (ψt)t[0,1](\psi^{t})_{t\in[0,1]} starting at the identity and ending at φ\varphi and ψ\psi, respectively. We define the concatenation (χt)t[0,1](\chi^{t})_{t\in[0,1]} by

χt{φ2tt[0,1/2]ψ2t1φ1t[1/2,1].\chi^{t}\coloneqq\begin{cases}\varphi^{2t}&t\in[0,1/2]\\ \psi^{2t-1}\circ\varphi^{1}&t\in[1/2,1].\end{cases}

Now pick a C0C^{0} Hamiltonian structure (Y,Ω)(Y,\Omega) and an embedding of C0C^{0} Hamiltonian structures α:((0,1)×Σ,ωΣ)(Y,Ω)\alpha:((0,1)\times\Sigma,\omega_{\Sigma})\hookrightarrow(Y,\Omega). Define the plug 𝒵(Σ,ωΣ,α,(χt)t[0,1])\mathcal{Z}\coloneqq(\Sigma,\omega_{\Sigma},\alpha,(\chi^{t})_{t\in[0,1]}). By construction, it is possible to split the plug 𝒵\mathcal{Z} into two disjoint plugs 𝒫\mathcal{P} and 𝒬\mathcal{Q} inserting the isotopies φt\varphi^{t} and ψt\psi^{t}, respectively, such that Ω#𝒵=Ω#(𝒫𝒬)\Omega\#\mathcal{Z}=\Omega\#(\mathcal{P}\sqcup\mathcal{Q}). We can then compute

p~Σ(ψφ)\displaystyle\tilde{p}_{\Sigma}(\psi\circ\varphi) =\displaystyle= ¯A(Ω#𝒵)¯A(Ω)\displaystyle\overline{\mathcal{H}}^{A}(\Omega\#\mathcal{Z})-\overline{\mathcal{H}}^{A}(\Omega)
=\displaystyle= ¯A((Ω#𝒫)#𝒬)¯A(Ω)\displaystyle\overline{\mathcal{H}}^{A}((\Omega\#\mathcal{P})\#\mathcal{Q})-\overline{\mathcal{H}}^{A}(\Omega)
=\displaystyle= ¯A((Ω#𝒫)#𝒬)¯A(Ω#𝒫)+¯A(Ω#𝒫)¯A(Ω)\displaystyle\overline{\mathcal{H}}^{A}((\Omega\#\mathcal{P})\#\mathcal{Q})-\overline{\mathcal{H}}^{A}(\Omega\#\mathcal{P})+\overline{\mathcal{H}}^{A}(\Omega\#\mathcal{P})-\overline{\mathcal{H}}^{A}(\Omega)
=\displaystyle= p~Σ(ψ)+p~Σ(φ).\displaystyle\tilde{p}_{\Sigma}(\psi)+\tilde{p}_{\Sigma}(\varphi).

Here we use plug homogeneity and the definition of p~Σ\tilde{p}_{\Sigma}. This shows that p~Σ\tilde{p}_{\Sigma} is a homomorphism.

Next, we claim that the restriction of p~Σ\tilde{p}_{\Sigma} to the group of Hamiltonian diffeomorphisms Ham(Σ)\operatorname{Ham}(\Sigma) agrees with the Calabi homomorphism CalΣ\operatorname{Cal}_{\Sigma} via the inclusion A{\mathbb{R}}\subset A. Suppose that φHam(Σ)\varphi\in\operatorname{Ham}(\Sigma). Then we can choose the Hamiltonian isotopy, the Hamiltonian structure, and the plug embedding involved in the definition of p~Σ\tilde{p}_{\Sigma} to be smooth. Let (Y,ω)(Y,\omega) and 𝒫\mathcal{P} be the resulting smooth Hamiltonian structure and plug, respectively. We compute

p~Σ(φ)\displaystyle\tilde{p}_{\Sigma}(\varphi) =\displaystyle= ¯A(ω#𝒫)¯A(ω)\displaystyle\overline{\mathcal{H}}^{A}(\omega\#\mathcal{P})-\overline{\mathcal{H}}^{A}(\omega)
=\displaystyle= (ω#𝒫)(ω)\displaystyle\mathcal{H}(\omega\#\mathcal{P})-\mathcal{H}(\omega)
=\displaystyle= CalΣ(φ).\displaystyle\operatorname{Cal}_{\Sigma}(\varphi).

Here the second equality follows from the assumption that ¯A\overline{\mathcal{H}}^{A} satisfies the Extension property in Theorem 1.5. The third equality follow from Lemma 4.10.

Suppose that ι:Σ1Σ2\iota:\Sigma_{1}\hookrightarrow\Sigma_{2} is an area- and orientation-preserving embedding of open surfaces. Then we have

p~Σ2Ham¯(ι)=p~Σ1,\tilde{p}_{\Sigma_{2}}\circ\overline{\operatorname{Ham}}(\iota)=\tilde{p}_{\Sigma_{1}}, (7.3)

where Ham¯(ι)\overline{\operatorname{Ham}}(\iota) is the homomorphism between Hamiltonian homeomorphism groups obtained by pushforward via ι\iota; see Section 3. In order to see this, consider a C0C^{0} Hamiltonian structure (Y,Ω)(Y,\Omega) and an embedding α:((0,1)×Σ2,ωΣ2)(Y,Ω)\alpha:((0,1)\times\Sigma_{2},\omega_{\Sigma_{2}})\hookrightarrow(Y,\Omega). Given φHam¯(Σ1)\varphi\in\overline{\operatorname{Ham}}(\Sigma_{1}), we can form a plug 𝒫2\mathcal{P}_{2} by inserting the homeomorphism Ham¯(ι)(φ)\overline{\operatorname{Ham}}(\iota)(\varphi) via the embedding α\alpha. We can form a second plug 𝒫1\mathcal{P}_{1} by inserting the homeomorphism φ\varphi via the embedding α(id(0,1)×ι)\alpha\circ(\operatorname{id}_{(0,1)}\times\iota). Note that the C0C^{0} Hamiltonian structures obtained by inserting 𝒫1\mathcal{P}_{1} and 𝒫2\mathcal{P}_{2} agree, i.e. Ω#𝒫1=Ω#𝒫2\Omega\#\mathcal{P}_{1}=\Omega\#\mathcal{P}_{2}. Identity (7.3) is then immediate from the definition of p~Σj\tilde{p}_{\Sigma_{j}}.

Combining the above observations, we see that the homomorphisms p~Σ\tilde{p}_{\Sigma} descend to a well-defined group homomorphism

p:Ham¯ab(Σ)Ap:\mathcal{R}\cong\overline{\operatorname{Ham}}^{\operatorname{ab}}(\Sigma)\rightarrow A

over {\mathbb{R}} which is independent of Σ\Sigma. We need to check that ¯A=p¯\overline{\mathcal{H}}^{A}=p\circ\overline{\mathcal{H}}.

First, observe that since pp is a homomorphism over {\mathbb{R}}, and both ¯A\overline{\mathcal{H}}^{A} and ¯\overline{\mathcal{H}} satisfy the Extension property in Theorem 1.5, we have

¯A(ω)=p¯(ω)=(ω)\overline{\mathcal{H}}^{A}(\omega)=p\circ\overline{\mathcal{H}}(\omega)=\mathcal{H}(\omega)

for every smooth Hamiltonian structure ω\omega.

It is immediate from the definition of pp that ¯A\overline{\mathcal{H}}^{A} satisfies the following version of the Calabi property in Theorem 1.5: for every C0C^{0} Hamiltonian structure Ω\Omega and every Ω\Omega-plug 𝒫\mathcal{P}, we have

¯A(Ω#𝒫)=¯A(Ω)+pCal¯(𝒫).\overline{\mathcal{H}}^{A}(\Omega\#\mathcal{P})=\overline{\mathcal{H}}^{A}(\Omega)+p\circ\overline{\operatorname{Cal}}(\mathcal{P}).

Now consider an arbitrary C0C^{0} Hamiltonian structure (Y,Ω)(Y,\Omega). By Theorem 6.1, we can write Ω=ω#𝒫\Omega=\omega\#\mathcal{P} for some smooth Hamiltonian structure ω\omega and a C0C^{0} plug 𝒫\mathcal{P}. We compute

¯A(Ω)\displaystyle\overline{\mathcal{H}}^{A}(\Omega) =\displaystyle= ¯A(ω#𝒫)\displaystyle\overline{\mathcal{H}}^{A}(\omega\#\mathcal{P})
=\displaystyle= ¯A(ω)+pCal¯(𝒫)\displaystyle\overline{\mathcal{H}}^{A}(\omega)+p\circ\overline{\operatorname{Cal}}(\mathcal{P})
=\displaystyle= p((ω)+Cal¯(𝒫))\displaystyle p(\mathcal{H}(\omega)+\overline{\operatorname{Cal}}(\mathcal{P}))
=\displaystyle= p¯(ω#𝒫)\displaystyle p\circ\overline{\mathcal{H}}(\omega\#\mathcal{P})
=\displaystyle= p¯(Ω).\displaystyle p\circ\overline{\mathcal{H}}(\Omega).

Here the second equality uses the Calabi property for ¯A\overline{\mathcal{H}}^{A}, the third equality uses the Extension property and the fact that pp is a homomorphism over {\mathbb{R}}, and the fourth equality uses the Calabi property of ¯\overline{\mathcal{H}}.

This concludes the proof of the existence part of Proposition 1.6. For the uniqueness part, simply observe that every element of \mathcal{R} is attained as the universal \mathcal{R}-valued helicity ¯(Ω)\overline{\mathcal{H}}(\Omega) of some C0C^{0} Hamiltonian structure Ω\Omega.∎

8 The relationship between Hamiltonian structures and volume-preserving flows

As mentioned earlier, there exists a close relationship between Hamiltonian structures and volume-preserving flows. The goal of this section is to clarify this relation, particularly in the C0C^{0} setting. This is needed for deducing Theorem 1.3, which is stated for volume-preserving flows, from our main result Theorem 1.5, which is stated for Hamiltonian structures. Theorem 1.3 is proven in Section 8.4.

8.1 The smooth case

Consider a closed oriented smooth 33-manifold YY. Suppose μ\mu is a volume form on YY that is compatible with its orientation. Let ω\omega denote a Hamiltonian structure, i.e., a closed maximally nondegenerate 22-form. Finally, let φ\varphi be a smooth, fixed-point-free flow on YY, generated by a nowhere vanishing vector field XX.

Consider 3=×2{\mathbb{R}}^{3}={\mathbb{R}}\times{\mathbb{R}}^{2} equipped with coordinates (t,x,y)(t,x,y). Let μstd=dtdxdy\mu_{\operatorname{std}}=dt\wedge dx\wedge dy be the standard volume form and equip 3{\mathbb{R}}^{3} with the orientation induced by this volume form. Let φstd\varphi_{\operatorname{std}} be the flow generated by the vector field Xstd=tX_{\operatorname{std}}=\partial_{t}. Recall that the standard Hamiltonian structure ωstd\omega_{\operatorname{std}} is given by the 22-form ωstd=dxdy\omega_{\operatorname{std}}=dx\wedge dy.

Lemma 8.1.

The following statements are equivalent:

  1. 1.

    We have ιXμ=ω\iota_{X}\mu=\omega.

  2. 2.

    The triple (φ,μ,ω)(\varphi,\mu,\omega) is locally diffeomorphic to the triple (φstd,μstd,ωstd)(\varphi_{\operatorname{std}},\mu_{\operatorname{std}},\omega_{\operatorname{std}}).

Proof.

Note that ιXstdμstd=ωstd\iota_{X_{\operatorname{std}}}\mu_{\operatorname{std}}=\omega_{\operatorname{std}} and, moreover, this condition is preserved by diffeomorphisms and can be checked locally. Hence statement 2 implies statement 1.

Conversely, suppose that ιXμ=ω\iota_{X}\mu=\omega. Given an arbitrary point pYp\in Y, we may pick a>0a>0 sufficiently small and an embedding ι:B(a)Y\iota:B(a)\hookrightarrow Y such that ι(0)=p\iota(0)=p and the pull back ιω\iota^{*}\omega agrees with the the standard area form on B(a)B(a). For ε>0\varepsilon>0 sufficiently small, extend ι\iota to an embedding ι:(ε,ε)×B(a)Y\iota:(-\varepsilon,\varepsilon)\times B(a)\hookrightarrow Y via the flow φ\varphi. It can be checked that the pull back of (φ,μ,ω)(\varphi,\mu,\omega) via ι\iota is given by (φstd,μstd,ωstd)(\varphi_{\operatorname{std}},\mu_{\operatorname{std}},\omega_{\operatorname{std}}). ∎

Definition 8.2.

We call (φ,μ,ω)(\varphi,\mu,\omega) a compatible triple if the equivalent conditions in Lemma 8.1 are satisfied.

Note that given (μ,ω)(\mu,\omega), there exists a unique vector field XX satisfying ιXμ=ω\iota_{X}\mu=\omega and we can extend (μ,ω)(\mu,\omega) to a compatible triple by taking φ\varphi to be the flow generated by XX. Below we formulate criteria for extension of (φ,μ)(\varphi,\mu) or (φ,ω)(\varphi,\omega) to compatible triples.

Proposition 8.3.

The pair (φ,μ)(\varphi,\mu) extends to a compatible triple if and only if it is locally diffeomorphic to (φstd,μstd)(\varphi_{\operatorname{std}},\mu_{\operatorname{std}}). The analogous statements hold for the pairs (φ,ω)(\varphi,\omega) and (μ,ω)(\mu,\omega).

Proof.

The “only if” part of the statement is clear in view of the characterization of compatibility given in Lemma 8.1. For the converse direction, note that any two components of (φstd,μstd,ωstd)(\varphi_{\operatorname{std}},\mu_{\operatorname{std}},\omega_{\operatorname{std}}) clearly extend to a compatible triple. This shows that being locally diffeomorphic to a pair contained in the triple (φstd,μstd,ωstd)(\varphi_{\operatorname{std}},\mu_{\operatorname{std}},\omega_{\operatorname{std}}) allows to locally extend to a compatible triple. But in view of the uniqueness of extensions to a compatible triple proved in Lemma 8.4, stated and proven below, local extendibility implies global extendibility. ∎

Lemma 8.4.

Any two components of a compatible triple (φ,μ,ω)(\varphi,\mu,\omega) uniquely determine the third.

Proof.

This is a consequence of the following fact: Let VV be a 33-dimensional vector space and suppose vVv\in V, σΛ3V\sigma\in\Lambda^{3}V^{*}, and τΛ2V\tau\in\Lambda^{2}V^{*} are non-zero and satisfy the identity ιvσ=τ\iota_{v}\sigma=\tau. Then, any two components of the triple (v,σ,τ)(v,\sigma,\tau) uniquely determine the third. ∎

Proposition 8.5.
  1. 1.

    The pair (φ,μ)(\varphi,\mu) extends to a compatible triple if and only if the flow φ\varphi preserves μ\mu.

  2. 2.

    The pair (φ,ω)(\varphi,\omega) extends to a compatible triple if and only if the flow φ\varphi is positively tangent to the characteristic foliation of ω\omega.

Proof.

If (φ,μ,ω)(\varphi,\mu,\omega) is a compatible triple, then φ\varphi preserves μ\mu and the flow φ\varphi is positively tangent to the characteristic foliation of ω\omega. This shows the “only if” direction in both statements. Conversely, if φ\varphi preserves μ\mu, then ιXμ\iota_{X}\mu is a maximally nondegenerate closed 22-form and we can extend (φ,μ)(\varphi,\mu) to a compatible triple by defining the Hamiltonian structure ω\omega to consist of the 22-form ωιXμ\omega\coloneqq\iota_{X}\mu.

Finally, if the flow φ\varphi is positively tangent to the characteristic foliation of ω\omega, then there exists a unique volume form μ\mu satisfying ιXμ=ω\iota_{X}\mu=\omega. ∎

8.2 The C0C^{0} case

The goal of this section is to explain the relationship between Hamiltonian structures and volume-preserving flows in the C0C^{0} setting.

Topological flows and C0C^{0} foliations

Let YY be a closed oriented topological 33-manifold, for now not equipped with a measure. Consider a topological flow φ\varphi on YY, i.e. a continuous map φ:×YY\varphi:{\mathbb{R}}\times Y\rightarrow Y such that φ(t,)\varphi(t,\cdot) is a homeomorphism of YY for each tt\in{\mathbb{R}} and, moreover,

φ(t,φ(s,p))=φ(t+s,p)for all t,s and pY.\varphi(t,\varphi(s,p))=\varphi(t+s,p)\qquad\text{for all $t,s\in{\mathbb{R}}$ and $p\in Y$.}

It will be convenient to use the notation φt(p)φ(t,p)\varphi^{t}(p)\coloneqq\varphi(t,p). In the following, we always assume that φ\varphi is fixed-point-free, i.e. that there does not exist a point pYp\in Y such that φt(p)=p\varphi^{t}(p)=p for all tt\in{\mathbb{R}}.

Proposition 8.6.

The flow lines of a fixed-point-free topological flow φ\varphi on YY form an oriented 11-dimensional C0C^{0} foliation φ\mathcal{F}_{\varphi}. Conversely, for every oriented C0C^{0} foliation \mathcal{F} on YY, there exists a fixed-point-free topological flow φ\varphi such that =φ.\mathcal{F}=\mathcal{F}_{\varphi}.

Proof.

We begin by proving that the flow lines of a fixed-point-free flow φ\varphi on YY form an oriented 11-dimensional C0C^{0} foliation φ\mathcal{F}_{\varphi}. This actually makes important use of the fact that YY has dimension three; see Remark 8.8 below. Let pYp\in Y be an arbitrary point. It was shown by Whitney [55] (see also [8, Lemma 1 and Corollary 1]) that φ\varphi admits a local cross section through pp which is a 22-dimensional topological disc. This means that there exist ε>0\varepsilon>0 and a topologically embedded 22-dimensional disc BYB\subset Y containing pp such that

(ε,ε)×BY(t,q)φt(q)(-\varepsilon,\varepsilon)\times B\rightarrow Y\qquad(t,q)\mapsto\varphi^{t}(q)

is an embedding. By construction, this embedding maps straight lines (ε,ε)×{}(-\varepsilon,\varepsilon)\times\{*\} into flow lines of φ\varphi. Its inverse is therefore a topological foliation chart. Since pYp\in Y was arbitrary, we can cover YY by such foliation charts.

Next, we show that every oriented C0C^{0} foliation \mathcal{F} admits a fixed-point-free flow φ\varphi such that =φ\mathcal{F}=\mathcal{F}_{\varphi}. Our strategy is to construct a special metric dd on YY with the property that the leaves of the characteristic foliation of Ω\Omega are rectifiable curves with respect to the metric dd. We then use the metric dd to obtain, for each point pYp\in Y, a unique continuous curve γp:LY\gamma_{p}:{\mathbb{R}}\rightarrow L\subset Y, where LL denotes the characteristic leaf containing pp, with the following properties:

  • γp\gamma_{p} parametrized by arc length,

  • γp(0)=p\gamma_{p}(0)=p,

  • γp:L\gamma_{p}:{\mathbb{R}}\rightarrow L is an orientation-preserving homeomorphism if LL is an open leaf and an orientation-preserving covering if LL is a closed leaf.

In this situation, we can define the flow φ\varphi by simply setting φt(p)γp(t)\varphi^{t}(p)\coloneqq\gamma_{p}(t).

We now describe the construction of the metric dd. For δ>0\delta>0, we consider foliation charts of the form

Cδ:=(0,1)×Dδ,C_{\delta}:=(0,1)\times D_{\delta},

where DδD_{\delta} denotes the disc of radius δ\delta. The oriented leaves of the foliation within this chart are given by

(0,1)×{p},(0,1)\times\{p\},

for pDδp\in D_{\delta}. For some δ>0\delta>0, we can pick a finite collection of charts

ιi:C2δY1in\iota_{i}:C_{2\delta}\hookrightarrow Y\qquad 1\leq i\leq n

such that the sets ιi(Cδ)\iota_{i}(C_{\delta}) form an open covering of YY. Moreover, fix a smooth map

f:C2δS34f:C_{2\delta}\rightarrow S^{3}\subset{\mathbb{R}}^{4}

with the property that there exist an open set CδUC2δC_{\delta}\Subset U\Subset C_{2\delta} and a point S3*\in S^{3} such that the restriction of ff to UU is a diffeomorphism f|U:US3{}f|_{U}:U\rightarrow S^{3}\setminus\{*\} and f(C2δU)=f(C_{2\delta}\setminus U)=*. For 1in1\leq i\leq n, we define

fi:YS3fi(p){f(ιi1(p))if p in im(ιi)otherwise.f_{i}:Y\rightarrow S^{3}\qquad f_{i}(p)\coloneqq\begin{cases}f(\iota_{i}^{-1}(p))&\text{if $p$ in $\operatorname{im}(\iota_{i})$}\\ *&\text{otherwise}.\end{cases}

Note that fif_{i} is continuous and that its restriction to ιi(Cδ)\iota_{i}(C_{\delta}) is an embedding. For 1in1\leq i\leq n, we define the pseudo-metric

di:Y×Y0di(p,q)|fi(p)fi(q)|,d_{i}:Y\times Y\rightarrow{\mathbb{R}}_{\geq 0}\qquad d_{i}(p,q)\coloneqq|f_{i}(p)-f_{i}(q)|,

where |||\cdot| denotes the standard norm on 4{\mathbb{R}}^{4}. Suppose that pqYp\neq q\in Y are two distinct points. There exists ii such that pιi(Cδ)p\in\iota_{i}(C_{\delta}). For this ii, we must have fi(p)fi(q)f_{i}(p)\neq f_{i}(q) which shows that

di=1ndid\coloneqq\sum\limits_{i=1}^{n}d_{i}

defines a metric on YY.

Next, we show that every leaf LL of the characteristic foliation of Ω\Omega is rectifiable with respect to the metric dd. We recall the necessary preliminaries. A continuous curve

γ:[a,b]Y\gamma:[a,b]\to Y

is called rectifiable if its length

(γ):=sup{i=1nd(γ(ti1),γ(ti))}\ell(\gamma):=\sup\left\{\sum_{i=1}^{n}d(\gamma(t_{i-1}),\gamma(t_{i}))\right\} (8.1)

is finite, where the supremum is taken over all partitions a=t0<t1<<tn=ba=t_{0}<t_{1}<\cdots<t_{n}=b of the interval [a,b][a,b]. We say γ:Y\gamma:{\mathbb{R}}\rightarrow Y is rectifiable if the restriction of γ\gamma to any closed interval is rectifiable. If γ\gamma is rectifiable, then any reparametrization of it is rectifiable and has the same length.

Now, consider a leaf LL of the foliation \mathcal{F} and pick a continuous map γ:L\gamma:{\mathbb{R}}\rightarrow L which is an orientation-preserving homeomorphism if LL is open and an orientation-preserving covering if LL is closed. The claim below shows that γ\gamma is a rectifiable curve.

Claim 8.7.

For each compact interval [a,b][a,b]\subset{\mathbb{R}}, we have (γ|[a,b])<\ell(\gamma|_{[a,b]})<\infty. Moreover, we have

limb+(γ|[a,b])=+andlima(γ|[a,b])=+.\lim_{b\rightarrow+\infty}\ell(\gamma|_{[a,b]})=+\infty\qquad\text{and}\qquad\lim_{a\rightarrow-\infty}\ell(\gamma|_{[a,b]})=+\infty. (8.2)
Proof.

For each ii, we can write

Iiγ1(im(ιi))=jIijI_{i}\coloneqq\gamma^{-1}(\operatorname{im}(\iota_{i}))=\bigcup_{j}I_{i}^{j}

where (Iij)j(I_{i}^{j})_{j} is a countable collection of pairwise disjoint finite open intervals IijI_{i}^{j}\subset{\mathbb{R}}. Every compact subset of {\mathbb{R}} intersects only finitely many of the intervals IijI_{i}^{j}. Given a point zD2δz\in D_{2\delta}, let ηz\eta_{z} be a parametrization of the line segment (0,1)×{z}C2δ(0,1)\times\{z\}\subset C_{2\delta}. Since ff is a smooth map, we have std(fηz)<+\ell_{\operatorname{std}}(f\circ\eta_{z})<+\infty, where std\ell_{\operatorname{std}} stands for length measured with respect to the standard metric on 4{\mathbb{R}}^{4}. From this we can conclude that i(γ|Iij)<+\ell_{i}(\gamma|_{I_{i}^{j}})<+\infty, where i\ell_{i} denotes length with respect to the pseudo-metric did_{i}, defined analogously to (8.1) . Observe that if JJ is an interval disjoint from IiI_{i}, then i(γ|J)=0\ell_{i}(\gamma|_{J})=0. Using the fact that [a,b][a,b] intersects only finitely many of the intervals IijI_{i}^{j}, we conclude that (γ|[a,b])=ii(γ|[a,b])<+\ell(\gamma|_{[a,b]})=\sum_{i}\ell_{i}(\gamma|_{[a,b]})<+\infty.

Since the restriction of ff to CδC_{\delta} is an embedding, there exists a positive number c>0c>0 such that (fηz)>c\ell(f\circ\eta_{z})>c for all zDδz\in D_{\delta}. If IijI_{i}^{j} is an interval such that γ(Iij)ιi(Cδ)\gamma(I_{i}^{j})\subset\iota_{i}(C_{\delta}), we then have i(γ|Iij)>c\ell_{i}(\gamma|_{I_{i}^{j}})>c. Since the sets ιi(Cδ)\iota_{i}(C_{\delta}) cover YY, every tt\in{\mathbb{R}} is contained in some interval IijI_{i}^{j} with this property. The limits (8.2) are an immediate consequence. This concludes the proof of the claim. ∎

It follows from Claim 8.7 that if pYp\in Y is a point and LL is the leaf containing pp, there exists a unique curve γp:LY\gamma_{p}:{\mathbb{R}}\rightarrow L\subset Y which is an orientation-preserving homeomorphism, or an orientation-preserving covering if LL is closed, such that γp\gamma_{p} is parametrized by arc length and γp(0)=p\gamma_{p}(0)=p. As already mentioned, we can then define φ(t,p)γp(t)\varphi(t,p)\coloneqq\gamma_{p}(t). It remains to check that φ:×MM\varphi:{\mathbb{R}}\times M\rightarrow M is continuous. A priori, we must also show that φ(t,)\varphi(t,\cdot) is a homeomorphism for each tt\in{\mathbb{R}}, however, this follows immediately from continuity as φ(t,)\varphi(-t,\cdot) is the inverse of φ(t,)\varphi(t,\cdot).

Fix a point pYp\in Y and an arbitrary T>0T>0. Our goal is to show that if qYq\in Y is close to pp, then γq|[T,T]\gamma_{q}|_{[-T,T]} is C0C^{0} close to γp|[T,T]\gamma_{p}|_{[-T,T]}. First note that if qq is sufficiently close to pp and LqL_{q} denotes the leaf containing qq, we can find a continuous map β:LqY\beta:{\mathbb{R}}\rightarrow L_{q}\subset Y which is an orientation-preserving covering of LqL_{q} such that β(0)=q\beta(0)=q and such that β|[T,T]\beta|_{[-T,T]} is C0C^{0} close to γp|[T,T]\gamma_{p}|_{[-T,T]}. It then suffices to show that β\beta is close to being parametrized by arc length, i.e. that for every compact interval [a,b][T,T][a,b]\subset[-T,T], the length (β|[a,b])\ell(\beta|_{[a,b]}) is close to |ba||b-a|. We provide an outline of an argument proving this: For i{1,2}i\in\{1,2\}, let ηi:[0,1]C2δ\eta_{i}:[0,1]\rightarrow C_{2\delta} be an embedded continuous line segment with image contained in (0,1)×{zi}(0,1)\times\{z_{i}\} for ziD2δz_{i}\in D_{2\delta}. If η1\eta_{1} and η2\eta_{2} are C0C^{0} close, then (fη1)\ell(f\circ\eta_{1}) and (fη2)\ell(f\circ\eta_{2}) are close. Note that for every segment of γp\gamma_{p} traversing ιi(C2δ)\iota_{i}(C_{2_{\delta}}) there is a corresponding C0C^{0}-close-by segment of β\beta traversing ιi(C2δ)\iota_{i}(C_{2\delta}). This implies that i(γp|[a,b])\ell_{i}(\gamma_{p}|_{[a,b]}) is close to i(β|[a,b])\ell_{i}(\beta|_{[a,b]}) for all ii and for every interval [a,b][T,T][a,b]\subset[-T,T]. Thus |ba|=(γp|[a,b])|b-a|=\ell(\gamma_{p}|_{[a,b]}) and (β|[a,b])\ell(\beta|_{[a,b]}) are close as well. This concludes the proof of continuity of φ\varphi. ∎

Remark 8.8.

The assumption that Y has dimension 3 plays a crucial role in our proof of the first part of Proposition 8.6, namely that the flow lines of φ\varphi define a C0C^{0} foliation. However, this dimensional restriction is not required for the second part of the proposition: in higher-dimensional manifolds, every oriented 1-dimensional C0C^{0} foliation still admits a fixed-point-free flow tracing out its leaves.

We now describe an example of a fixed-point-free topological flow ψ\psi on a topological 55-manifold ZZ which does not admit a local cross section homeomorphic to a 44-dimensional ball B4B^{4} everywhere. Start with a 33-dimensional homology sphere MM which is not homeomorphic to the 33-sphere S3S^{3}. Consider the suspension SM=[1,1]×M/SM=[-1,1]\times M/\sim, where {±1}×M\{\pm 1\}\times M are collapsed to points p±p_{\pm}. While this space is a homology manifold, it is not a topological manifold. The two points p±p_{\pm} do not have neighborhoods homeomorphic to the 44-dimensional ball. However, by the double suspension theorem [7], the double suspension S2MS^{2}M is homeomorphic to S5S^{5}. This also implies that the product ZS1×SMZ\coloneqq S^{1}\times SM is a topological 55-manifold. Let ψ\psi be the fixed-point-free flow on ZZ which rotates the S1S^{1} factor and restricts to the identity on the SMSM factor. Now observe that ψ\psi cannot have a local cross section homeomorphic to B4B^{4} at any point contained in S1×{p±}S^{1}\times\{p_{\pm}\}. If it did, one would obtain a neighborhood of p±p_{\pm} in SMSM homeomorphic to B4B^{4}.

The Oxtoby–Ulam theorem

Let XX be a compact metric space and let μ\mu be a Borel measure on XX of finite mass, i.e. a measure μ\mu defined on the Borel σ\sigma-algebra of XX satisfying μ(X)<+\mu(X)<+\infty. The support supp(μ)\operatorname{supp}(\mu) of μ\mu is the set of all points xXx\in X such that every open neighborhood of xx has strictly positive measure. The measure μ\mu is said to have full support if supp(μ)=X\operatorname{supp}(\mu)=X. This is equivalent to requiring that the measure of any non-empty open subset is non-negative. Following [15, Def. 2.15], we say that μ\mu is non-atomic if μ({x})=0\mu(\{x\})=0 for every point xXx\in X.

Remark 8.9.

An atom of a general measure space (X,Σ,μ)(X,\Sigma,\mu) is often defined to be a measurable set AΣA\in\Sigma such that μ(A)>0\mu(A)>0 and such that every measurable set BAB\subset A satisfies μ(B)=0\mu(B)=0 or μ(B)=μ(A)\mu(B)=\mu(A). A measure is then called non-atomic if it does not have any atoms. A simple argument (see [31, 2.IV]) shows that if μ\mu is an inner and outer regular finite Borel measure, then μ\mu is non-atomic in this sense if and only if it is non-atomic in the sense defined above, i.e. if and only if μ({x})=0\mu(\{x\})=0 for all points xx. Note that any finite Borel measure on a compact metric space is inner and outer regular [15, Prop. 2.3]. This means that in all situations of interest to us in this paper, the two notions of being non-atomic agree.

The following theorem was proved in [43] by Oxtoby and Ulam, who also give credit to von Neumann for an independent and unpublished proof. Our formulation of the theorem follows [18].

Theorem 8.10.

Let MnM^{n} be a compact connected topological manifold, possibly with boundary. For i{1,2}i\in\{1,2\}, let μi\mu_{i} be a finite, non-atomic Borel measure on MM of full support. Moreover, assume that μi(M)=0\mu_{i}(\partial M)=0. Then, there exists a homeomorphism f:MMf:M\rightarrow M such that fμ1=μ2f_{*}\mu_{1}=\mu_{2} if and only if μ1(M)=μ2(M)\mu_{1}(M)=\mu_{2}(M). If this is the case, the homeomorphism ff can be chosen such that f|M=idMf|_{\partial M}=\operatorname{id}_{\partial M}.

Volume-preserving topological flows and measured C0C^{0} foliations

As above, let φ\varphi be a fixed-point-free topological flow on a closed oriented topological 33-manifold YY. Let φ\mathcal{F}_{\varphi} be the C0C^{0} foliation induced by φ\varphi; see Proposition 8.6.

Lemma 8.11.

Suppose that μ\mu is a finite Borel measure on YY preserved by φ\varphi. Then φ\varphi and μ\mu induce a transverse measure Λφ,μ\Lambda_{\varphi,\mu} on φ\mathcal{F}_{\varphi} characterized by the condition that, for every transversal ΣY\Sigma\subset Y, every Borel subset AΣA\subset\Sigma, and every ε>0\varepsilon>0 sufficiently small, we have

Λφ,μ(A)=ε1μ({φt(p)pA,0<t<ε}).\Lambda_{\varphi,\mu}(A)=\varepsilon^{-1}\mu(\{\varphi^{t}(p)\mid p\in A,0<t<\varepsilon\}). (8.3)
Proof.

Pick ε0>0\varepsilon_{0}>0 such that the restriction of φ\varphi to (ε0,ε0)×Σ(-\varepsilon_{0},\varepsilon_{0})\times\Sigma is an embedding. Since μ\mu is preserved by φ\varphi, the expression on the right hand side of (8.3) is independent of ε(0,ε0)\varepsilon\in(0,\varepsilon_{0}). Therefore, (8.3) yields a well-defined finite Borel measure on Σ\Sigma. We need to show that maps between subsets of transversals obtained by sliding along leaves of φ\mathcal{F}_{\varphi} are measure preserving. This can be reduced to showing that if ΣY\Sigma\subset Y is a transversal, AΣA\subset\Sigma is a Borel subset, τ:A\tau:A\rightarrow{\mathbb{R}} is a continuous function with small LL^{\infty} norm, and ε>0\varepsilon>0 is small, then

μ({φt(p)pA,0<t<ε})=μ({φt(p)pA,τ(p)<t<τ(p)+ε}).\mu(\{\varphi^{t}(p)\mid p\in A,0<t<\varepsilon\})=\mu(\{\varphi^{t}(p)\mid p\in A,\tau(p)<t<\tau(p)+\varepsilon\}).

This identity follows from the assumption that φ\varphi preserves μ\mu. ∎

Let us say that a transverse measure Λ\Lambda on some foliation \mathcal{F} has full support if the induced measure on any transversal has full support. Similarly, we say that Λ\Lambda is non-atomic if the induced measure on any transversal is non-atomic.

Now suppose that φ\varphi is a fixed-point-free topological flow preserving a finite Borel measure μ\mu. Let (φ,Λφ,μ)(\mathcal{F}_{\varphi},\Lambda_{\varphi,\mu}) be the induced measured foliation; see Lemma 8.11. It is straightforward to see that μ\mu has full support if and only if the induced transverse measure Λφ,μ\Lambda_{\varphi,\mu} has full support. Since μ\mu has finite mass and φ\varphi does not have fixed points, the measure μ\mu is automatically non-atomic. However, it is possible that the transverse measure Λφ,μ\Lambda_{\varphi,\mu} has atoms.

Example 8.12.

Consider Y=S3Y=S^{3} with the Hopf flow φ\varphi, i.e. the flow tracing out the fibers of the Hopf fibration at unit speed. Let μ3\mu_{3} be the standard 33-dimensional Lebesgue measure on S3S^{3}, scaled to have total volume 11. The Hopf flow φ\varphi preserves μ3\mu_{3}. Moreover, pick a Hopf fiber FF and let μ1\mu_{1} be the 11-dimensional Lebesgue measure supported on this fiber FF, again scaled to have total volume 11. Then μ12(μ1+μ3)\mu\coloneqq\frac{1}{2}(\mu_{1}+\mu_{3}) is a finite Borel measure on YY invariant under φ\varphi. The induced transverse measure Λφ,μ\Lambda_{\varphi,\mu} on the Hopf fibration φ\mathcal{F}_{\varphi} is not non-atomic: any intersection point of a transversal with the special fiber FF yields an atom.

In this example, the flow φ\varphi is smooth, but the measure μ\mu is not. We can turn the situation around using the Oxtoby-Ulam theorem: Since both μ\mu and μ3\mu_{3} have full support, are non-atomic, and have total volume 11, there exists a homeomorphism ff of YY such that fμ=μ3f_{*}\mu=\mu_{3}. Let fφf_{*}\varphi be the pushforward of φ\varphi under ff. Then fφf_{*}\varphi is a fixed-point-free topological flow on YY which preserves the standard Lebesgue measure μ3\mu_{3}. However, fφf_{*}\varphi has a periodic orbit f(F)f(F) of positive measure, causing the induced transverse measure Λfφ,μ3\Lambda_{f_{*}\varphi,\mu_{3}} on fφ\mathcal{F}_{f_{*}\varphi} to be not non-atomic.

Note that we can produce even more pathological examples. Let (Fn)n1(F_{n})_{n\geq 1} be a dense sequence of Hopf fibers. For each nn, let μFn\mu_{F_{n}} denote the 11-dimensional Lebesgue measure on FnF_{n} of total volume 11. Define the φ\varphi-invariant measure μn12nμFn\mu\coloneqq\sum_{n\geq 1}2^{-n}\mu_{F_{n}}. This measure has full support, is non-atomic, and has total volume 11. Again by Oxtoby-Ulam, the measure μ\mu is homeomorphic to the standard 33-dimensional Lebesgue measure μ3\mu_{3}, i.e. there exists a homeomorphism gg of Y such that gμ=μ3g_{*}\mu=\mu_{3}. Then gφg_{*}\varphi is a fixed-point-free topological flow on YY which preserves μ3\mu_{3} and has the property that the complement of some countable sequence of periodic orbits has measure zero.

Lemma 8.13.

Suppose that μ\mu is a finite Borel measure preserved by a fixed-point-free topological flow φ\varphi. Then the following statements are equivalent:

  1. 1.

    Λφ,μ\Lambda_{\varphi,\mu} is non-atomic.

  2. 2.

    All flow lines of φ\varphi have vanishing measure.

  3. 3.

    All periodic orbits of φ\varphi have vanishing measure.

Proof.

If ΣY\Sigma\subset Y is a transversal and xΣx\in\Sigma is a point, then Λφ,μ({x})>0\Lambda_{\varphi,\mu}(\{x\})>0 if and only if the flow line of φ\varphi through xx has positive measure. This shows the equivalence of the first two statements.

We argue that any flow line LL of φ\varphi of positive measure is necessarily periodic. Indeed, if LL is a non-periodic flow line through a point xx, then LL is the disjoint union of the countable collection of flow line segments Liφ([i,i+1)×{x})L_{i}\coloneqq\varphi([i,i+1)\times\{x\}) for ii\in{\mathbb{Z}}. All of these flow line segments have the same volume. Since μ\mu is finite, the total volume of LL must be finite as well. But this implies that all flow line segments LiL_{i} and therefore LL itself have vanishing volume. ∎

Compatible triples

Consider a closed oriented topological 33-manifold YY. Fix a finite Borel measure μ\mu and a C0C^{0} Hamiltonian structure Ω\Omega on YY. Let φ\varphi denote a fixed-point-free topological flow on YY.

The C0C^{0} Hamiltonian structure Ω\Omega determines a 11-dimensional foliation Ω\mathcal{F}_{\Omega} on YY equipped with a transverse measure ΛΩ\Lambda_{\Omega}. Moreover, Ω\Omega determines a coorientation of Ω\mathcal{F}_{\Omega}, which in combination with the orientation on YY determines an orientation \mathcal{F}.

Lemma 8.14.

The following statements are equivalent:

  1. 1.

    The flow φ\varphi preserves μ\mu and (φ,Λφ,μ)(\mathcal{F}_{\varphi},\Lambda_{\varphi,\mu}) coincides with (Ω,ΛΩ)(\mathcal{F}_{\Omega},\Lambda_{\Omega}), as measured, cooriented (hence, oriented) foliations.

  2. 2.

    The triple (φ,μ,Ω)(\varphi,\mu,\Omega) is locally homeomorphic to the triple (φstd,μstd,ωstd).(\varphi_{\operatorname{std}},\mu_{\operatorname{std}},\omega_{\operatorname{std}}).

Proof.

Note that the triple (φstd,μstd,ωstd)(\varphi_{\operatorname{std}},\mu_{\operatorname{std}},\omega_{\operatorname{std}}) satisfies all conditions in statement 1. Since all of these conditions can be checked locally, statement 2 implies statement 1. Conversely, suppose that (φ,μ,Ω)(\varphi,\mu,\Omega) satisfies statement 1. Every point in YY has a neighborhood which can be parametrized via an orientation-preserving topological embedding of C0C^{0} Hamiltonian structures

ι:((1,1)×B(a),ωstd)(Y,Ω).\iota:((-1,1)\times B(a),\omega_{\operatorname{std}})\hookrightarrow(Y,\Omega).

Note that ι\iota maps line segments (1,1)×{}(-1,1)\times\{*\} into flow lines of φ\varphi respecting orientation. For ε>0\varepsilon>0 sufficiently small, define the embedding

α:(ε,ε)×B(a)Yα(t,p)φt(ι(0,p)).\alpha:(-\varepsilon,\varepsilon)\times B(a)\rightarrow Y\qquad\alpha(t,p)\coloneqq\varphi^{t}(\iota(0,p)).

This embedding continues to pull back Ω\Omega to ωstd\omega_{\operatorname{std}} and in addition pulls back φ\varphi to φstd\varphi_{\operatorname{std}}. We conclude that the transverse measure αΛφ,μ=Λφstd,αμ\alpha^{*}\Lambda_{\varphi,\mu}=\Lambda_{\varphi_{\operatorname{std}},\alpha^{*}\mu} agrees with αω=ωstd\alpha^{*}\omega=\omega_{\operatorname{std}}. Thus αμ\alpha^{*}\mu agrees with μstd\mu_{\operatorname{std}} on all sets of the form I×AI\times A where I(ε,ε)I\subset(-\varepsilon,\varepsilon) is an interval and AB(a)A\subset B(a) is a Borel set. This determines αμ=μstd\alpha^{*}\mu=\mu_{\operatorname{std}} and we conclude that α\alpha is a local homeomorphism between (φstd,μstd,ωstd)(\varphi_{\operatorname{std}},\mu_{\operatorname{std}},\omega_{\operatorname{std}}) and (φ,μ,ω)(\varphi,\mu,\omega). ∎

Definition 8.15.

We call (φ,μ,Ω)(\varphi,\mu,\Omega) a compatible triple if the equivalent conditions in Lemma 8.14 are satisfied.

We now formulate criteria guaranteeing that any of the pairs (φ,μ)(\varphi,\mu), (φ,Ω)(\varphi,\Omega) or (μ,Ω)(\mu,\Omega) extend to compatible triples.

Proposition 8.16.

The pair (φ,μ)(\varphi,\mu) extends to a compatible triple if and only if it is locally homeomorphic to (φstd,μstd)(\varphi_{\operatorname{std}},\mu_{\operatorname{std}}). The analogous statements hold for the pairs (φ,Ω)(\varphi,\Omega) and (μ,Ω)(\mu,\Omega).

Proof.

This can be proven through an argument analogous to the smooth case presented in Proposition 8.3; we omit the details here. The proof of Proposition 8.3 makes use of Lemma 8.4, whose C0C^{0} counterpart is Lemma 8.17, stated and proven below. ∎

Lemma 8.17.

Any two components of a compatible triple (φ,μ,Ω)(\varphi,\mu,\Omega) uniquely determine the third.

Proof.

It suffices to check the above uniqueness locally. This means we need to verify that if (φ,μ,Ω)(\varphi,\mu,\Omega) is a compatible triple on (an open subset of) 3{\mathbb{R}}^{3} two of whose components coincide with the corresponding components of (φstd,μstd,ωstd)(\varphi_{\operatorname{std}},\mu_{\operatorname{std}},\omega_{\operatorname{std}}), then (φ,μ,Ω)=(φstd,μstd,ωstd)(\varphi,\mu,\Omega)=(\varphi_{\operatorname{std}},\mu_{\operatorname{std}},\omega_{\operatorname{std}}). In view of the characterization of compatibility given by statement 2, it suffices to check that if ψ\psi is a homeomorphism of 3{\mathbb{R}}^{3} preserving two members of the triple (φstd,μstd,ωstd)(\varphi_{\operatorname{std}},\mu_{\operatorname{std}},\omega_{\operatorname{std}}), then it preserves the third. This is elementary and we omit the details. ∎

Proposition 8.18.
  1. 1.

    The pair (φ,μ)(\varphi,\mu) extends to a compatible triple if and only if the flow of φ\varphi preserves μ\mu and the flow lines of φ\varphi have zero measure.

  2. 2.

    The pair (φ,μ)(\varphi,\mu) extends to a compatible triple if and only if the flow of φ\varphi preserves μ\mu and the induced transverse measure Λφ,μ\Lambda_{\varphi,\mu} on φ\mathcal{F}_{\varphi} is of full support and non-atomic.

  3. 3.

    The pair (φ,Ω)(\varphi,\Omega) extends to a compatible triple if and only if the oriented flow lines of φ\varphi agree with the oriented characteristic leaves of Ω\Omega.

Proof.

The first and the second items are equivalent by Lemma 8.13. We will prove the second item.

If the pair (φ,μ)(\varphi,\mu) extends to a compatible triple, then (φ,Λφ,μ)(\mathcal{F}_{\varphi},\Lambda_{\varphi,\mu}) is locally homeomorphic to (the measured foliation of) ωstd\omega_{\operatorname{std}}, which clearly implies that Λφ,μ\Lambda_{\varphi,\mu} is of full support and non-atomic. Conversely, suppose that φ\varphi preserves μ\mu and Λφ,μ\Lambda_{\varphi,\mu} is of full support and non-atomic. Given an arbitrary point in YY, pick a local cross section BYB\subset Y through that point which is homeomorphic to a disc; as mentioned earlier, the existence of the cross section BB was proven by Whitney [55]. Since the measure on BB induced by Λφ,μ\Lambda_{\varphi,\mu} is of full support and non-atomic, it follows from the Oxtoby-Ulam theorem that there exists an area-preserving parametrization ι:(B(a),ωstd)(B,Λφ,μ)\iota:(B(a),\omega_{\operatorname{std}})\rightarrow(B,\Lambda_{\varphi,\mu}). In addition, we can pick ι\iota such that it is orientation preserving with respect to the orientation of BB induced by the flow φ\varphi and the orientation of YY. Extend ι\iota to an embedding α:(ε,ε)×B(a)Y\alpha:(-\varepsilon,\varepsilon)\times B(a)\hookrightarrow Y via the flow φ\varphi. Clearly, α\alpha pulls back φ\varphi to φstd\varphi_{\operatorname{std}}. Moreover, it pulls back (φ,Λφ,μ)(\mathcal{F}_{\varphi},\Lambda_{\varphi,\mu}) to ωstd\omega_{\operatorname{std}}. Hence Ω(φ,Λφ,μ)\Omega\coloneqq(\mathcal{F}_{\varphi},\Lambda_{\varphi,\mu}) equipped with the orientation and coorientation induced by φ\varphi and the ambient orientation of YY is a C0C^{0} Hamiltonian structure extending (φ,μ)(\varphi,\mu) to a compatible triple. We have proven the second statement.

We turn now to the third statement. If (φ,Ω)(\varphi,\Omega) extends to a compatible triple, then clearly the oriented flow lines of φ\varphi agree with the oriented characteristic leaves of Ω\Omega by the characterization of compatibility given in statement 1 in Lemma 8.14. Conversely, suppose that the oriented flow lines of φ\varphi agree with the oriented characteristic leaves of Ω\Omega. As we saw in the proof of Lemma 8.14, we can parametrize a neighborhood of any point in YY via an embedding of C0C^{0} Hamiltonian structures α:((ε,ε)×B(a),ωstd)(Y,Ω)\alpha:((-\varepsilon,\varepsilon)\times B(a),\omega_{\operatorname{std}})\hookrightarrow(Y,\Omega) that also pulls back the flow φ\varphi to φstd\varphi_{\operatorname{std}}. By Proposition 8.16, this means that (φ,Ω)(\varphi,\Omega) extends to a compatible triple. ∎

Remark 8.19.

We proved in Proposition 8.6 that every oriented C0C^{0} foliation admits a fixed-point free flow tracing out its leaves. Hence, as a consequence of the second item in Proposition 8.18, every C0C^{0} Hamiltonian structure Ω\Omega is a component of a compatible tripe (φ,μ,Ω)(\varphi,\mu,\Omega).

8.3 Mass flow and flux

In this section, we briefly review the mass flow homomorphism and show that for a volume-preserving flow φ\varphi which is a component of a compatible triple (φ,μ,Ω)(\varphi,\mu,\Omega), the flux Flux¯(Ω)\overline{\operatorname{Flux}}(\Omega) admits a description in terms of mass flow.

Consider a closed and oriented topological manifold MM of dimension nn. Suppose that μ\mu is a finite non-atomic Borel measure on MM of full support. Recall that there is a mass flow homomorphism, introduced in [18, 48],

θ~:Homeo~0(M,μ)H1(M;)Hn1(M;),\widetilde{\theta}:\widetilde{\operatorname{Homeo}}_{0}(M,\mu)\rightarrow H_{1}(M;{\mathbb{R}})\cong H^{n-1}(M;{\mathbb{R}}),

where Homeo~0(M,μ)\widetilde{\operatorname{Homeo}}_{0}(M,\mu) denotes the universal cover of the identity component of the group of volume-preserving homeomorphisms of (M,μ)(M,\mu) and the identification H1(M;)Hn1(M;)H_{1}(M;{\mathbb{R}})\cong H^{n-1}(M;{\mathbb{R}}) is given by Poincaré duality.

We briefly recall the definition of θ~\widetilde{\theta}. Let 𝕋/{\mathbb{T}}\coloneqq{\mathbb{R}}/{\mathbb{Z}} denote the circle. Then we have an isomorphism H1(M;)[M,𝕋]H^{1}(M;{\mathbb{Z}})\cong[M,{\mathbb{T}}] and therefore H1(M;)Hom([M,𝕋],)H_{1}(M;{\mathbb{R}})\cong\operatorname{Hom}([M,{\mathbb{T}}],{\mathbb{R}}). Given an element φ~Homeo~0(M,μ)\widetilde{\varphi}\in\widetilde{\operatorname{Homeo}}_{0}(M,\mu) represented by a volume-preserving isotopy (φt)t[0,1](\varphi_{t})_{t\in[0,1]} starting at the identity, defining θ~(φ~)\widetilde{\theta}(\widetilde{\varphi}) therefore amounts to defining a group homomorphism

θ~(φ~):[M,𝕋].\widetilde{\theta}(\widetilde{\varphi}):[M,{\mathbb{T}}]\rightarrow{\mathbb{R}}.

Let [f][M,𝕋][f]\in[M,{\mathbb{T}}] be a homotopy class represented by a map f:M𝕋f:M\rightarrow{\mathbb{T}} and consider the homotopy of maps fφtf:Y𝕋f\circ\varphi^{t}-f:Y\rightarrow{\mathbb{T}}, which at t=0t=0 coincides with the constant map zero. There exists a unique lift to a homotopy of maps fφtf¯:Y\overline{f\circ\varphi^{t}-f}:Y\rightarrow{\mathbb{R}}, which at t=0t=0 coincides with the constant map zero. Then,

θ~(φ~)([f])Yfφ1f¯𝑑μ,\widetilde{\theta}(\widetilde{\varphi})([f])\coloneqq\int_{Y}\overline{f\circ\varphi_{1}-f}d\mu, (8.4)

which turns out to be independent of the choice of representative of φ~\tilde{\varphi} and [f][f]. This defines θ~:Homeo~0(M,μ)H1(M;)\widetilde{\theta}:\widetilde{\operatorname{Homeo}}_{0}(M,\mu)\rightarrow H_{1}(M;{\mathbb{R}}), which we may view as taking values in Hn1(M;)H^{n-1}(M;{\mathbb{R}}), via Poincaré duality.

Definition 8.20.

Let φttHomeo0(M,μ){\varphi^{t}}_{t\in{\mathbb{R}}}\subset\operatorname{Homeo}_{0}(M,\mu) be a flow. We say φt\varphi^{t} is exact if the element of Homeo~0(M,μ)\widetilde{\operatorname{Homeo}}_{0}(M,\mu) determined by φtt[0,1]{\varphi^{t}}_{t\in[0,1]} has vanishing mass flow.

We now restrict our attention to the case where MM is a closed, oriented 3-manifold, which we will refer to from this point onward as YY.

Proposition 8.21.

Suppose that (φ,μ,Ω)(\varphi,\mu,\Omega) is a compatible triple. Then, for every T{0}T\in{\mathbb{R}}\setminus\{0\}, we have

1Tθ~([(φt)t[0,T]])=Flux¯(Ω).\frac{1}{T}\widetilde{\theta}([(\varphi^{t})_{t\in[0,T]}])=\overline{\operatorname{Flux}}(\Omega). (8.5)
Proof.

We begin by proving the identity (8.5) for smooth compatible triples (φ,μ,ω)(\varphi,\mu,\omega). Let f:Y𝕋f:Y\rightarrow{\mathbb{T}} be a smooth map. Throughout this proof, we will view θ~\widetilde{\theta} as mapping into H1(Y;)H_{1}(Y;{\mathbb{R}}) which we identify with Hom([Y,],)\operatorname{Hom}([Y,{\mathbb{Z}}],{\mathbb{R}}). Viewing the right hand side of identity (8.5) as a homomorphism [Y,𝕋][Y,{\mathbb{T}}]\rightarrow{\mathbb{R}}, we compute

T1θ~([(φt)t[0,T]])([f])\displaystyle T^{-1}\widetilde{\theta}([(\varphi^{t})_{t\in[0,T]}])([f]) =T1YfφTf¯μ\displaystyle=T^{-1}\int_{Y}\overline{f\circ\varphi^{T}-f}\,\mu
=T1Y(0T(ιXdf)φt𝑑t)μ\displaystyle=T^{-1}\int_{Y}\left(\int_{0}^{T}(\iota_{X}df)\circ\varphi^{t}\enspace dt\right)\mu
=T10T(Y(ιXdf)φtμ)𝑑t\displaystyle=T^{-1}\int_{0}^{T}\left(\int_{Y}(\iota_{X}df)\circ\varphi^{t}\enspace\mu\right)dt
=Y(ιXdf)μ\displaystyle=\int_{Y}(\iota_{X}df)\mu
=Y𝑑fιXμ\displaystyle=\int_{Y}df\wedge\iota_{X}\mu
=Y𝑑fω\displaystyle=\int_{Y}df\wedge\omega
=[df]Flux(ω),[Y],\displaystyle=\langle\,[df]\cup\operatorname{Flux}(\omega),[Y]\,\rangle,

where ,\langle\cdot,\cdot\rangle stands for the pairing of cohomology and homology. Here, the first identity uses the definition of the mass flow homomorphism (8.4). The fourth equality uses that φt\varphi^{t} preserves μ\mu, which implies that Y(ιXdf)φtμ\int_{Y}(\iota_{X}df)\circ\varphi^{t}\,\mu is independent of tt. The fifth equality holds because dfμdf\wedge\mu being a 44-form necessarily vanishes and hence we have

0=ιX(dfμ)=(ιXdf)μdfιXμ.0=\iota_{X}(df\wedge\mu)=(\iota_{X}df)\mu-df\wedge\iota_{X}\mu.

The sixth equality uses that (φ,μ,ω)(\varphi,\mu,\omega) is a compatible triple. In the final equality, [df]H1(Y;)[df]\in H^{1}(Y;{\mathbb{R}}) is the de Rham cohomology class represented by the 11-form dfdf. This class agrees with the image of the cohomology class represented by the circle map ff under the natural map H1(Y;)H1(Y;)H^{1}(Y;{\mathbb{Z}})\rightarrow H^{1}(Y;{\mathbb{R}}). Using the identifications

Hom([Y,],)Hom(H1(Y;),)H1(Y;R)H2(Y;),\operatorname{Hom}([Y,{\mathbb{Z}}],{\mathbb{R}})\cong\operatorname{Hom}(H^{1}(Y;{\mathbb{Z}}),{\mathbb{R}})\cong H_{1}(Y;R)\cong H^{2}(Y;{\mathbb{R}}),

this shows identity (8.5) in the smooth case.

In the general case, where (φ,μ,Ω)(\varphi,\mu,\Omega) is not assumed to be smooth, we make use of the following claim. Its proof closely parallels that of Theorem 6.1 and is therefore omitted.

Claim 8.22.

There exist

  1. 1.

    a smooth compatible triple (φ0,μ0,ω0)(\varphi_{0},\mu_{0},\omega_{0}) on a smooth, closed and oriented 33-manifold Y0Y_{0},

  2. 2.

    an open surface with area form (Σ,ωΣ)(\Sigma,\omega_{\Sigma}) consisting of finitely many disc components, and a smooth embedding ι:((0,ε)×Σ,ωΣ)(Y0,ω0)\iota:((0,\varepsilon)\times\Sigma,\omega_{\Sigma})\hookrightarrow(Y_{0},\omega_{0}) satisfying φ0s(ι(t,p))=ι(t+s,p)\varphi_{0}^{s}(\iota(t,p))=\iota(t+s,p) for all t<s<εt-t<s<\varepsilon-t,

  3. 3.

    an isotopy (ψt)t[0,ε](\psi_{t})_{t\in[0,\varepsilon]} in Ham¯(Σ)\overline{\operatorname{Ham}}(\Sigma),

such that the compatible triple (φ1,μ0,Ω1)(\varphi_{1},\mu_{0},\Omega_{1}) obtained from (φ0,μ0,ω0)(\varphi_{0},\mu_{0},\omega_{0}) by inserting888Here, by inserting the isotopy (ψt)t(\psi_{t})_{t} into (φ0,μ0,ω0)(\varphi_{0},\mu_{0},\omega_{0}) we mean modifying the compatible triple (φ0,μ0,ω0)(\varphi_{0},\mu_{0},\omega_{0}) in a manner analogous to the plug insertion operation, introduced in Section 5.2. We alluded to this in relation to Equation (1.1). the isotopy (ψt)t(\psi_{t})_{t} via the embedding ι\iota is homeomorphic to (φ,μ,Ω)(\varphi,\mu,\Omega).

Given an arbitrary homotopy class in [Y0,𝕋][Y_{0},{\mathbb{T}}], we may represent it by a smooth map f:Y0𝕋f:Y_{0}\rightarrow{\mathbb{T}} with the property that fι(t,p)=tf\circ\iota(t,p)=t. This makes use of the fact Σ\Sigma consists of finitely many disc components. For T>0T>0 sufficiently small, we then have

fφ0Tf¯=fφ1Tf¯.\overline{f\circ\varphi_{0}^{T}-f}=\overline{f\circ\varphi_{1}^{T}-f}.

This implies that

θ~([(φ0t)t[0,T]])=θ~([(φ1t)t[0,T]])\widetilde{\theta}([(\varphi_{0}^{t})_{t\in[0,T]}])=\widetilde{\theta}([(\varphi_{1}^{t})_{t\in[0,T]}])

for all T>0T>0 sufficiently small. Because mass flow is a homomorphism, we conclude that this identity holds for all T>0T>0. Since (φ0,μ0,ω0)(\varphi_{0},\mu_{0},\omega_{0}) is a smooth compatible triple, we have

T1θ~([(φ0t)t[0,T]])=Flux(ω0)T^{-1}\widetilde{\theta}([(\varphi_{0}^{t})_{t\in[0,T]}])=\operatorname{Flux}(\omega_{0})

by the smooth case treated above. It follows from Definition 7.2, which defines Flux¯\overline{\operatorname{Flux}}, that

Flux¯(Ω1)=Flux(ω0).\overline{\operatorname{Flux}}(\Omega_{1})=\operatorname{Flux}(\omega_{0}).

Combining the above identities, we conclude that

T1θ~([(φ1t)t[0,T]])=Flux¯(Ω1).T^{-1}\widetilde{\theta}([(\varphi_{1}^{t})_{t\in[0,T]}])=\overline{\operatorname{Flux}}(\Omega_{1}).

Since (φ,μ,Ω)(\varphi,\mu,\Omega) is homeomorphic to (φ1,μ1,Ω1)(\varphi_{1},\mu_{1},\Omega_{1}), this implies that

T1θ~([(φt)t[0,T]])=Flux¯(Ω).T^{-1}\widetilde{\theta}([(\varphi^{t})_{t\in[0,T]}])=\overline{\operatorname{Flux}}(\Omega).

8.4 Proof of Theorem 1.3

We will now deduce Theorem 1.3 from Theorem 1.5 and the content of Section 8.2.

Let (Y,μ)(Y,\mu) be a closed 33-manifold equipped with a volume form μ\mu. We will assume throughout this section that YY is equipped with the orientation induced by μ\mu.

We denote by 𝒮\mathcal{S} the set consisting of all volume-preserving topological flows φ\varphi which are fixed-point-free and whose flow lines have vanishing measure. Let 𝒮0𝒮\mathcal{S}_{0}\subset\mathcal{S} be the subset consisting of exact flows (see Definition 8.20); in other words, 𝒮0\mathcal{S}_{0} consists of flows satisfying the assumptions of Theorem 1.3. Similarly, let 𝒯\mathcal{T} denote the set of all C0C^{0} Hamiltonian structures on YY, and let 𝒯0𝒯\mathcal{T}_{0}\subset\mathcal{T} denote the subset of exact C0C^{0} Hamiltonian structures (see Definition 7.2).

Take any φ𝒮\varphi\in\mathcal{S}. Then, by Proposition 8.18 and Lemma 8.14, there exists a unique C0C^{0} Hamiltonian structure Ωφ\Omega_{\varphi} such that the triple (φ,μ,Ωφ)(\varphi,\mu,\Omega_{\varphi}) is a compatible triple in the sense of Definition 8.15. This yields a mapping from 𝒮\mathcal{S} to 𝒯\mathcal{T}, which preserves mass flow/flux by Proposition 8.21. Hence it induces a mapping from 𝒮0\mathcal{S}_{0} to 𝒯0\mathcal{T}_{0}

The above yields a \mathcal{R}-valued extension of helicity, which we continue to denote by ¯\overline{\mathcal{H}}: for φ𝒮0\varphi\in\mathcal{S}_{0}, we simply define

¯(φ):=¯(Ωφ),\overline{\mathcal{H}}(\varphi):=\overline{\mathcal{H}}(\Omega_{\varphi}),

where the right-hand-side in the above is the universal \mathcal{R}-valued extension of helicity given by Theorem 1.5. The three properties listed in Theorem 1.5 translate as follows for this \mathcal{R}-valued extension of helicity ¯:𝒮0\overline{\mathcal{H}}:\mathcal{S}_{0}\rightarrow\mathcal{R}.

  1. 1.

    Extension: If φ\varphi is smooth, then

    ¯(φ)=(φ),\overline{\mathcal{H}}(\varphi)=\mathcal{H}(\varphi)\in{\mathbb{R}},

    where we view {\mathbb{R}} as a subgroup of \mathcal{R} via the natural inclusion (1.7).

  2. 2.

    Conjugation invariance: we have

    ¯(fφf1)=¯(φ),\overline{\mathcal{H}}(f\varphi f^{-1})=\overline{\mathcal{H}}(\varphi),

    for any orientation- and volume-preserving homeomorphism ff.

  3. 3.

    Calabi Compatibility: For every plug 𝒫=(Σ,ωΣ,α,ψt)\mathcal{P}=(\Sigma,\omega_{\Sigma},\alpha,\psi^{t}), we have

    ¯(φ#𝒫)=¯(φ)+Cal¯Σ(ψ1),\overline{\mathcal{H}}(\varphi\#\mathcal{P})=\overline{\mathcal{H}}(\varphi)+\overline{\mathrm{Cal}}_{\Sigma}(\psi^{1}),

    where φ#𝒫\varphi\#\mathcal{P} refers to the analogue of the plug-insertion operation for flows, which we alluded to in relation to Equation (1.1).

Moreover, as in Theorem 1.5, ¯:𝒮0\overline{\mathcal{H}}:\mathcal{S}_{0}\rightarrow\mathcal{R} is uniquely determined by the above properties. Indeed, by Claim 8.22 the flow φ\varphi is topologically conjugate to a flow of the form ψ#𝒫\psi\#\mathcal{P} where ψ\psi is a smooth exact volume-preserving flow. Applying Theorem 6.1 and arguing as we did in the proof of Theorem 1.5, one can deduce that any other \mathcal{R}-valued extension of helicity to 𝒮0\mathcal{S}_{0}, satisfying the above three properties, must coincide with ¯\overline{\mathcal{H}}.

Now, to obtain a real-valued extension of helicity, as stated in Theorem 1.3, simply pick pr:\operatorname{pr}:\mathcal{R}\rightarrow{\mathbb{R}} to be a choice of projection. Then, the composition pr¯\operatorname{pr}\circ\overline{\mathcal{H}} is an {\mathbb{R}}-valued extension of helicity which satisfies the three properties stated in Theorem 1.3. The Calabi compatibility property holds with respect to the real-valued extension of Calabi prCal¯Σ:Ham¯(Σ)\operatorname{pr}\circ\overline{\mathrm{Cal}}_{\Sigma}:\overline{\operatorname{Ham}}(\Sigma)\rightarrow{\mathbb{R}}. The uniqueness part of the statement may be proven using Claim 8.22, as in the previous paragraph.

This completes the proof of Theorem 1.3.∎

References

  • [1] V. I. Arnold. The asymptotic Hopf invariant and its applications. Proc. Summer School in Diff. Equations at Dilizhan, 1973, Erevan (in Russian), 375, 1974.
  • [2] V. I. Arnold. The asymptotic Hopf invariant and its applications. volume 5, pages 327–345. 1986. Selected translations.
  • [3] V. I. Arnold. Arnold’s problems. Springer-Verlag, Berlin; PHASIS, Moscow, revised edition, 2004. With a preface by V. Philippov, A. Yakivchik and M. Peters.
  • [4] V. I. Arnold and B. Khesin. Topological methods in hydrodynamics, volume 125 of Applied Mathematical Sciences. Springer, Cham, second edition, [2021] ©2021.
  • [5] A. Banyaga. Sur la structure du groupe des difféomorphismes qui préservent une forme symplectique. Comment. Math. Helv., 53(2):174–227, 1978.
  • [6] E. Calabi. On the group of automorphisms of a symplectic manifold. In Problems in analysis (Sympos. in honor of Salomon Bochner, Princeton Univ., Princeton, N.J., 1969), pages 1–26. Princeton Univ. Press, Princeton, NJ, 1970.
  • [7] J. W. Cannon. Shrinking cell-like decompositions of manifolds. Codimension three. Ann. of Math. (2), 110(1):83–112, 1979.
  • [8] W. C. Chewning Jr. and R. S. Owen. Local sections of flows on manifolds. Proc. Amer. Math. Soc., 49:71–77, 1975.
  • [9] G. Contreras. Average linking numbers of closed orbits of hyperbolic flows. J. London Math. Soc. (2), 51(3):614–624, 1995.
  • [10] G. Contreras and R. Iturriaga. Average linking numbers. Ergodic Theory Dynam. Systems, 19(6):1425–1435, 1999.
  • [11] D. Cristofaro-Gardiner, V. Humilière, C. Y. Mak, S. Seyfaddini, and I. Smith. Quantitative Heegaard Floer cohomology and the Calabi invariant. Forum Math. Pi, 10:Paper No. e27, 59, 2022.
  • [12] D. Cristofaro-Gardiner, V. Humilière, and S. Seyfaddini. PFH spectral invariants on the two-sphere and the large scale geometry of Hofer’s metric. J. Eur. Math. Soc. (JEMS), 26(12):4537–4584, 2024.
  • [13] D. Cristofaro-Gardiner, V. Humilière, and S. Seyfaddini. Proof of the simplicity conjecture. Ann. of Math. (2), 199(1):181–257, 2024.
  • [14] D. Cristofaro-Gardiner, V. Humilière, C. Y. Mak, S. Seyfaddini, and I. Smith. Subleading asymptotics of link spectral invariants and homeomorphism groups of surfaces. arXiv:2206.10749.
  • [15] M. Denker, C. Grillenberger, and K. Sigmund. Ergodic theory on compact spaces, volume Vol. 527 of Lecture Notes in Mathematics. Springer-Verlag, Berlin-New York, 1976.
  • [16] A. Enciso, D. Peralta-Salas, and F. Torres de Lizaur. Helicity is the only integral invariant of volume-preserving transformations. Proc. Natl. Acad. Sci. USA, 113(8):2035–2040, 2016.
  • [17] M. Entov, L. Polterovich, and P. Py. On continuity of quasimorphisms for symplectic maps. In Perspectives in analysis, geometry, and topology, volume 296 of Progr. Math., pages 169–197. Birkhäuser/Springer, New York, 2012. With an appendix by Michael Khanevsky.
  • [18] A. Fathi. Structure of the group of homeomorphisms preserving a good measure on a compact manifold. Ann. Sci. École Norm. Sup. (4), 13(1):45–93, 1980.
  • [19] M. H. Freedman and Z. He. Divergence-free fields: energy and asymptotic crossing number. Ann. of Math. (2), 134(1):189–229, 1991.
  • [20] J. Gambaudo and É. Ghys. Enlacements asymptotiques. Topology, 36(6):1355–1379, 1997.
  • [21] É. Ghys. Knots and dynamics. In International Congress of Mathematicians. Vol. I, pages 247–277. Eur. Math. Soc., Zürich, 2007.
  • [22] É. Ghys. Le groupe des homéomorphismes de la sphère de dimension 2 qui respectent l’aire et l’orientation n’est pas un groupe simple [dáprès D. Cristofaro-Gardiner, V. Humilière et S. Seyfaddini]. Astérisque, (446):Exp. No. 1204, 251–284, 2023.
  • [23] V. Giri, H. Kwon, and M. Novack. Non-conservation of generalized helicity in the Euler equations. In preparation.
  • [24] M. J. Gotay. On coisotropic imbeddings of presymplectic manifolds. Proc. Amer. Math. Soc., 84(1):111–114, 1982.
  • [25] A. Haefliger and G. Reeb. Variétés (non séparées) à une dimension et structures feuilletées du plan. Enseign. Math. (2), 3:107–125, 1957.
  • [26] H. Helmholtz. On integrals of the hydrodynamical equations which express vortex motions. transl. P. G. Tait in Phil. Mag. 33:4 (1867), page 485–512, 1858.
  • [27] H. Helmholtz. Über Integrale der hydrodynamischen Gleichungen, welche den Wirbelbewegungen entsprechen. J. Reine Angew. Math., 55:25–55, 1858.
  • [28] B. Khesin. Geometry of higher helicities. Mosc. Math. J., 3(3):989–1011, 1200, 2003. Dedicated to Vladimir Igorevich Arnold on the occasion of his 65th birthday.
  • [29] B. Khesin, D. Peralta-Salas, and C. Yang. The helicity uniqueness conjecture in 3D hydrodynamics. Trans. Amer. Math. Soc., 375(2):909–924, 2022.
  • [30] B. Khesin and N. C. Saldanha. Relative helicity and tiling twist. Transactions of the American Mathematical Society, 2025.
  • [31] J. D. Knowles. On the existence of non-atomic measures. Mathematika, 14:62–67, 1967.
  • [32] D. Kotschick and T. Vogel. Linking numbers of measured foliations. Ergodic Theory Dynam. Systems, 23(2):541–558, 2003.
  • [33] F. Le Roux. Simplicity of Homeo(𝔻2,𝔻2,Area){\rm Homeo}(\mathbb{D}^{2},\partial\mathbb{D}^{2},{\rm Area}) and fragmentation of symplectic diffeomorphisms. J. Symplectic Geom., 8(1):73–93, 2010.
  • [34] C. Y. Mak and I. Trifa. Hameomorphism groups of positive genus surfaces. arXiv:2306.06377.
  • [35] H. K. Moffatt. The degree of knottedness of tangled vortex lines. J. Fluid Mech., 35:117–129, 1969.
  • [36] H. K. Moffatt. Some developments in the theory of turbulence. Journal of Fluid Mechanics, 106:27–47, 1981.
  • [37] H. K. Moffatt. Helicity and singular structures in fluid dynamics. Proceedings of the National Academy of Sciences, 111(10):3663–3670, 2014.
  • [38] H. K. Moffatt and A. Tsinober. Helicity in laminar and turbulent flow. In Annual review of fluid mechanics, Vol. 24, pages 281–312. Annual Reviews, Palo Alto, CA, 1992.
  • [39] E. E. Moise. Affine structures in 33-manifolds. V. The triangulation theorem and Hauptvermutung. Ann. of Math. (2), 56:96–114, 1952.
  • [40] E. E. Moise. Geometric topology in dimensions 22 and 33, volume Vol. 47 of Graduate Texts in Mathematics. Springer-Verlag, New York-Heidelberg, 1977.
  • [41] J. Moreau. Constantes d’un îlot tourbillonnaire en fluide parfait barotrope. C. R. Acad. Sci. Paris, 252:2810–2812, 1961.
  • [42] S. Müller and P. Spaeth. Helicity of vector fields preserving a regular contact form and topologically conjugate smooth dynamical systems. Ergodic Theory Dynam. Systems, 33(5):1550–1583, 2013.
  • [43] J. C. Oxtoby and S. M. Ulam. Measure-preserving homeomorphisms and metrical transitivity. Ann. of Math. (2), 42:874–920, 1941.
  • [44] L. Polterovich. The geometry of the group of symplectic diffeomorphisms. Lectures in Mathematics ETH Zürich. Birkhäuser Verlag, Basel, 2001.
  • [45] T. Radó. Über den Begriff der Riemannschen Fläche. Acta Universitatis Szegediensis, 2:101–121, 1925.
  • [46] T. Rivière. High-dimensional helicities and rigidity of linked foliations. Asian J. Math., 6(3):505–533, 2002.
  • [47] M. W. Scheeler, W. M. van Rees, H. Kedia, D. Kleckner, and W. T. M. Irvine. Complete measurement of helicity and its dynamics in vortex tubes. Science, 357(6350):487–491, 2017.
  • [48] S. Schwartzman. Asymptotic cycles. Ann. of Math. (2), 66:270–284, 1957.
  • [49] B. Serraille. The sharp C0{C}^{0}-fragmentation property for hamiltonian diffeomorphisms and homeomorphisms on surfaces. arXiv:2403.15767, 2024.
  • [50] S. Seyfaddini. C0C^{0}-limits of Hamiltonian paths and the Oh-Schwarz spectral invariants. Int. Math. Res. Not. IMRN, (21):4920–4960, 2013.
  • [51] T. Tao. 2006 ICM: Étienne ghys, “knots and dynamics”. https://terrytao.wordpress.com/2007/08/03/2006-icm-etienne-ghys-knots-and-dynamics/#more-168, 2007.
  • [52] W. Thomson. On vortex motion. Transactions of the Royal Society of Edinburgh, 25(1):217–260, 1868.
  • [53] W. P. Thurston. Three-dimensional geometry and topology. Vol. 1, volume 35 of Princeton Mathematical Series. Princeton University Press, Princeton, NJ, 1997.
  • [54] T. Vogel. On the asymptotic linking number. Proc. Amer. Math. Soc., 131(7):2289–2297, 2003.
  • [55] H. Whitney. Cross-sections of curves in 33-space. Duke Math. J., 4(1):222–226, 1938.
  • [56] L. Woltjer. A theorem on force-free magnetic fields. Proc. Nat. Acad. Sci. U.S.A., 44:489–491, 1958.

Oliver Edtmair
ETH-ITS, ETH Zürich, Scheuchzerstrasse 70, 8006 Zürich, Switzerland.
e-mail: oliver.edtmair@eth-its.ethz.ch

Sobhan Seyfaddini
D-MATH, ETH Zürich, Rämistrasse 101, 8092 Zürich, Switzerland.
e-mail: sobhan.seyfaddini@math.ethz.ch