A universal extension of helicity to topological flows
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 Hamiltonian structures. This perspective enables us to leverage recent developments in 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 be a closed smooth -manifold equipped with a volume form , and let be a smooth vector field on that preserves . That is, the flow of satisfies for all . This condition is equivalent to the -form
being closed. The vector field is said to be exact if is exact.
The helicity of an exact volume-preserving vector field is defined as
where is any primitive -form of and is oriented via the volume form . This integral turns out to be independent of the choice of primitive -form .
A simple but fundamental property of helicity is that it is preserved under volume- and orientation-preserving diffeomorphisms; that is,
In fact, any other functional on the space of exact volume-preserving vector fields that is invariant under the action of and satisfies certain natural regularity conditions must be a function of helicity [29, 16].
Two volume-preserving smooth vector fields and on are said to be topologically conjugate if their flows are conjugate via a volume- and orientation-preserving homeomorphism, i.e. if there exists such that
For examples of volume-preserving smooth vector fields that are topologically conjugate but not smoothly (or even ) 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 and be two exact volume-preserving smooth vector fields which are topologically conjugate. Is it true that ?
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 Hamiltonian structure, enabling us to build on recent advances in 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 denote an open surface equipped with an area form . We denote its group of Hamiltonian diffeomorphisms by . It was proven recently [14, 34] that the Calabi homomorphism admits infinitely many extensions to the group of Hamiltonian homeomorphisms . Pick one such extension
The Calabi extensions for different surfaces 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 on and suppose that we have a topological volume-preserving embedding
which intertwines the flow on generated by the vector field and the flow . Consider a Hamiltonian isotopy and note that its suspension to is volume preserving. We refer to the tuple as a plug. Given a plug , one can define a new volume-preserving flow on by replacing the flow inside with the suspension of . The compatibility condition between our helicity extension and the Calabi extension is given by
(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 . 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 be an oriented smooth -manifold. A Hamiltonian structure on is a closed and maximally nondegenerate -form on . The prototypical example is that of the standard Hamiltonian structure on , which is given by the -form , where is equipped with the coordinates . Every smooth Hamiltonian structure is locally diffeomorphic to ; see Lemma 4.3.
Maximal non-degeneracy of means that defines a line field on , called the characteristic line field, which integrates to a -dimensional foliation on , called the characteristic foliation. The Hamiltonian structure naturally equips this foliation with a coorientation and a transverse measure locally diffeomorphic to the standard -dimensional Lebesgue measure. Conversely, every such foliation uniquely determines a Hamiltonian structure.
The relevance of Hamiltonian structures to our discussion is as follows: Equip with a volume form . Then, a vector field is nowhere vanishing and volume-preserving if and only if is a Hamiltonian structure. In this case, the flow is tangent to the characteristic foliation of .
We say a Hamiltonian structure is exact if is an exact -form. We define the helicity of an exact Hamiltonian structure on an oriented closed smooth -manifold to be
where is any choice of primitive -form of . As already mentioned, this integral is independent of the choice of . By definition, a nowhere vanishing volume-preserving vector field is exact if and only if the induced Hamiltonian structure is exact. If this is the case, we have
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.
Hamiltonian Structures. Let be an oriented topological -manifold. A Hamiltonian structure on is a cooriented, -dimensional foliation equipped with a transverse measure, locally modeled on with the characteristic foliation induced by . A more detailed definition, formalized in terms of Hamiltonian atlases, is given in Section 5. We will see in Section 8.2 that 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 Hamiltonian structures in Section 7.1. Hamiltonian structures.
1.4 Plugs and the Calabi invariant in the 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 setting.
Let be an open surface with an area form, and let and denote, respectively, the groups of compactly supported Hamiltonian diffeomorphisms and homeomorphisms of ; see Section 2. We denote by the Calabi homomorphism [6], which is defined as follows: for , generated by a compactly supported Hamiltonian , we have333Different conventions exist for defining the Calabi invariant; we include the factor 2 for convenience.
The -form induces a natural Hamiltonian structure on , which we denote by the same symbol . Given an oriented -manifold equipped with a Hamiltonian structure , we define a plug to be a tuple
where is an embedding of Hamiltonian structures and is a Hamiltonian isotopy. Given a plug , we define an operation, called plug insertion, which creates a new Hamiltonian structure on by replacing the Hamiltonian structure inside with the pushforward of the Hamiltonian structure on induced by via the embedding , where is defined by .444Plug insertion does not affect exactness of Hamiltonian structures, in smooth and settings. The effect of this operation on the characteristic foliations is as follows: Before plug insertion, the characteristic leaves of inside are of the form for . After plug insertion, the characteristic leaves of are of the form
for .
In the smooth setting, the following important identity, which can be deduced from [20], relates the helicities of and :
(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 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 category. Below, we establish this naturality in the 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, denote non-empty, open, connected surfaces equipped with area forms. A smooth area- and orientation-preserving embedding induces an injective homomorphism via pushforward. The Calabi homomorphism satisfies the following naturality property:
(1.3) |
Note, moreover, that the embedding induces a homomorphism at the level of abelianizations
(1.4) |
Here denotes the abelianization of a group . It follows from Banyaga’s work [5] that the homomorphism (1.4) is an isomorphism and is independent of the embedding . Hence the group does not depend on up to canonical isomorphism. In fact, the Calabi homomorphism induces a natural isomorphism
For our arguments, we need extensions of the Calabi homomorphism
that satisfy an analogue of the naturality property (1.3) in the setting. Although [14, 34] construct infinitely many 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 , let denote the induced homomorphism at the level of abelianizations.
Theorem 1.4.
The homomorphism
is an isomorphism. Moreover, it does not depend on the choice of area- and orientation-preserving embedding , i.e. if is another area- and orientation-preserving embedding, then
Set , where denotes the open unit disc. By Theorem 1.4, we may canonically identify
(1.5) |
for every non-empty, connected, open surface .
It was an open question—known as the simplicity conjecture—whether the group is simple. This question was recently resolved in the negative in [13], which showed that is not simple. Moreover, it is shown in [13, Cor. 1.3] that is not perfect either. In other words, the group is non-trivial.
It can be further deduced from [14, 34] that the natural homomorphism
(1.6) |
is injective and not surjective; see Section 3. The injectivity of the homomorphism (1.6), a key input from symplectic topology, is essential for our proof of the topological invariance of helicity. Recalling that the Calabi homomorphism induces an identification , the homomorphism (1.6) allows us to naturally view as a subgroup
(1.7) |
In view of the natural commutative diagram
we call the natural homomorphism
(1.8) |
the universal (or -valued) extension of the Calabi homomorphism.666Note that this amounts to a change of notation from (1.1). From this point on-wards, 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 later on, and in this case we define to be the sum of all where ranges over the components of . Its restriction to agrees with the usual smooth Calabi homomorphism via the natural inclusion . Moreover, by Theorem 1.4 it satisfies the naturality property
(1.9) |
for any area- and orientation-preserving topological embedding .
The extension is universal in the following sense: Consider an arbitrary extension of abelian groups . Then every system of extensions
of Calabi, one for every surface and subject to the naturality condition (1.9), arises as
for a unique group homomorphism over . In particular, for every choice of projection , we obtain a system of -valued Calabi extensions subject to (1.9), and every such system uniquely arises this way. Note that a projection is equivalent to a choice of a single Calabi extension for the open unit disc . 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 continuous extension of the Calabi homomorphism to , 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 -valued extension of helicity to exact Hamiltonian structures. It is always possible to obtain an -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 for an arbitrary exact Hamiltonian structure on an oriented closed topological -manifold . The notion of exactness for Hamiltonian structures is defined as the vanishing of a certain cohomology class . In particular, if is a rational homology three-sphere, then every Hamiltonian structure is exact. This is discussed in detail in Section 7.1.
Recall that denotes the plug insertion operation introduced in Section 1.4. Moreover, stands for the pullback of under a homeomorphism .
Theorem 1.5.
There is a unique way of assigning a -valued helicity to every exact Hamiltonian structure on an oriented closed topological -manifold such that the following conditions are satisfied:
- 1.
-
2.
Invariance: We have
for any orientation-preserving homeomorphism .
-
3.
Calabi Compatibility: For every plug , have
(1.10)
As already mentioned, every choice of projection 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 .
Let us elaborate further on the universality of our helicity extension 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 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 resulting from the insertion of a smooth plug depends only on the time-one map , and is independent of both the ambient Hamiltonian structure and the embedding 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 .
Proposition 1.6.
Let be an extension of abelian groups. Let be an -valued extension of helicity to exact Hamiltonian structures satisfying the Extension and Invariance properties in Theorem 1.5. Moreover, assume that satisfies plug homogeneity. That is, for any Hamiltonian structure and for any -plug , the helicity difference
only depends on and is independent of and . Then there exists a unique homomorphism over such that .
This means that, once plug homogeneity is accepted as a fundamental property of helicity, one is essentially forced to arrive at our universal -valued helicity extension . Moreover, insisting on plug homogeneity makes the problem of defining an -valued extension of helicity equivalent to specifying a projection , which, as explained in Subsection 1.4, cannot be done without making non-canonical choices. This indicates that is indeed the natural target group for a helicity extension to exact Hamiltonian structures.
Remark 1.7.
We end our introduction with the following remarks.
-
1.
The -valued helicity extension from Theorem 1.5 provides an obstruction to smoothability of Hamiltonian structures: if lies in , then is not homeomorphic to any smooth Hamiltonian structure.
-
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.
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 -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 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 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 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 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 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 be a smooth -manifold equipped with an area form. We assume that does not have boundary, but we allow it to be open.
Every compactly supported smooth Hamiltonian induces a time-dependent Hamiltonian vector field generating a compactly supported Hamiltonian isotopy . We adopt the sign convention that is characterized by the identity . The compactly supported Hamiltonian diffeomorphism group of is denoted by .
Let denote the group of all compactly supported homeomorphisms of preserving the measure induced by . We topologize as the direct limit of all the subgroups of homeomorphisms supported inside some compact subset , equipped with the compact-open topology. The group of Hamiltonian homeomorphisms is defined to be the closure of inside . It is topologized as a subspace of .
Let denote the identity component of . Then agrees with the kernel of the mass flow homomorphism [18, 48]
where is a discrete subgroup whose precise definition will not be relevant for us.
Whenever there is no risk of confusion, we will abbreviate and .
It is known that and , equipped with the topology, are both simply connected, unless is a sphere. If is a sphere, the fundamental group is . In the case of , a proof of these facts is sketched in [44, Sec. 7.2] for closed . The arguments therein can be adapted to the case of open . As for , it can be shown that every loop in is homotopic a loop in using the following two facts: first, every homeomorphism in can be written as a limit of elements in ; and second, and 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 be a surface with area form.
-
1.
Suppose is a finite open cover of , and for each , let be an open neighborhood of the identity. Then any Hamiltonian homeomorphism that lies in a sufficiently small neighborhood of the identity can be written as a composition , where each .
-
2.
Let be an arbitrary open cover of . Then, for every Hamiltonian homeomorphism , there exist finitely many open sets and Hamiltonian homeomorphisms such that .
Proposition 2.2.
Let be a surface equipped with an area form for each . Suppose is an area-preserving homeomorphism. Let be an open subset on which is smooth. Let be a distance function on that induces its topology, and let be a continuous, non-negative function satisfying . Then there exists an area-preserving diffeomorphism such that
2.2 Foliations and transverse measures
Let be an oriented manifold for . For , let denote the unit ball. We define a -dimensional foliation on to be a decomposition of into connected subsets , called leaves, such that can be covered by coordinate charts
called foliation charts, with the property that, for every leaf , the connected components of are of the form . The collection of the foliation charts defining forms a foliated atlas. Since the manifold is oriented, we may assume without loss of generality that the foliated atlas on is an oriented atlas of , meaning that we require the transition maps between the foliation charts to be orientation preserving. Note that this does not imply that is oriented.
We will consider foliations which are equipped with two additional structures: a transverse measure and a transverse orientation, also referred to as a coorientation. A transverse measure is specified by a foliated atlas of consisting of foliation charts and a finite Borel measure on for every chart such that the following is true: For , let be a chart and consider the transition map
Then, for every point , there exist an open neighborhood of in , an open neighborhood of in , and a topological embedding such that the restriction of to is of the form and moreover
Requiring that the topological embeddings are orientation preserving determines a transverse orientation, or coorientation, of the foliation .
Given a foliation , a transversal is a compact topological submanifold with boundary that is transverse to the foliation . This means that every point in is contained in a foliation chart such that is contained in a transverse slice of the form . A transverse measure induces a finite Borel measure on every transversal and maps between subsets of transversals obtained by sliding along leaves of are measure preserving. In fact, one can equivalently define a transverse measure on to be an assignment of a finite Borel measure to every transversal subject to the condition that sliding along leaves of preserves measure. A transverse orientation admits a similar description: it is an assignment of an orientation to every transversal such that sliding along leaves of preserves orientation.
Finally, note that since the ambient manifold is oriented, a coorientation of induces a natural orientation on the foliation . Vice-versa, the orientation on together with an orientation of determine a coorientation of .
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 -valued universal extension of the Calabi homomorphism which satisfies the naturality property stated in Section 1.4.
Let be the category whose objects are pairs , where is a non-empty, connected, smooth surface without boundary, and is an area form on . Morphisms in are smooth embeddings . We include both open and closed surfaces, without requiring finite area in the open case. Let and denote the full subcategories of open and closed surfaces, respectively. When there is no risk of confusion, we simply write for .
Recall that denotes the group of all compactly supported Hamiltonian diffeomorphisms of . Every area-preserving embedding induces a group homomorphism
given by pushforward of compactly supported Hamiltonian diffeomorphisms via . We can therefore regard as a functor from to the category of all groups. We define
to be the abelianization of . Since abelianization forms a functor from the category of groups to the category of abelian groups, is a functor from to the category of abelian groups.
It was proved by Banyaga [5] that is perfect for closed . In other words,
for every closed surface . For open surfaces (and more generally for exact connected symplectic manifolds of arbitrary dimension), Banyaga showed that . Moreover, an explicit isomorphism between and is induced by the Calabi homomorphism
whose definition we now recall. For a given , pick such that . Then,
This turns out to be independent of the choice of the Hamiltonian . Banyaga proved that is a surjective group homomorphism whose kernel is given by the commutator subgroup of . In other words, the Calabi homomorphism descends to an isomorphism .
The Calabi homomorphism is natural in the following sense: If is an area-preserving embedding between open surfaces , then
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 is an isomorphism for every open surface imply that is an isomorphism for all embeddings between open surfaces. Moreover, one can deduce that is independent of the choice of embedding .
Recall from Section 2 that denotes the group of compactly supported Hamiltonian homeomorphisms of . As before, this assignment defines a functor from the category 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 , which has the same objects as but allows all area-preserving topological embeddings as morphisms.
There is a natural inclusion , which corresponds to a natural transformation from the functor to the composition of the inclusion with the functor .
As mentioned in the introduction, for every area-preserving topological embedding of surfaces , we let denote the induced homomorphism at the level of abelianizations, where
As in the smooth setting, constitutes a functor from to the category of abelian groups.
We prove the following generalization of Theorem 1.4.
Theorem 3.1.
For , let be non-empty connected surfaces without boundary and equipped with area forms. Let be an area-preserving topological embedding. Then, the induced map
does not depend on the choice of embedding , i.e. if is any other embedding, then
Moreover,
-
i.
If and are both open or both closed, then is an isomorphism.
-
ii.
If is open and is closed, then is surjective and its kernel is given by the image of in .
Before presenting the proof of the above, we remark on some of its consequences.
For every surface , the abelian group 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 can be deduced from Theorem 3.1.
It is interesting to point out that non-perfectness of for one single open surface , for example , implies non-perfectness for all open surfaces via Theorem 3.1. Similarly, non-perfectness of for one single closed surface implies non-perfectness for all closed surfaces.
As already explained in the introduction, we define
Theorem 3.1 allows us to canonically identify
for every open surface .
Despite significant interest, the algebraic structure of remains rather mysterious. Most of what is known follows from the existence of a certain sequence of quasimorphisms
called link spectral invariants; see [11, 14]. Consider the space of real valued sequences and let denote the subspace of sequences converging to zero. Then the link spectral invariants induce a map
Since the sequence of defects of the quasimorphisms behaves like as goes to infinity, this map is a group homomorphism. Moreover, it fits into the following commutative diagram:
Here the homomorphism maps to the equivalence class of the constant sequence . 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
is injective. By Theorem 3.1 the same is then true for any other open surface as well. As discussed in the introduction, after identifying via the Calabi homomorphism, we can therefore regard
as a subgroup. It is proven in [14, Proposition 5.3] that the homomorphism is surjective. In particular, this implies that is a proper subgroup of .
Let be an open surface. The restriction of the natural map
to agrees with the usual Calabi homomorphism via the natural inclusion . We refer to this map as the universal -valued extension of Calabi. If is not connected, we define to be the sum of all where ranges over the components of . In view of Theorem 3.1, it is clear that we have the naturality property
(3.1) |
for any area-preserving topological embedding .
Every projection gives rise to a real-valued extension of the Calabi homomorphism for any open surface . Conversely, every real-valued Calabi extension arises this way. Clearly, the real-valued Calabi extensions arising from a single choice of projection satisfy the naturality condition (3.1).
Projections can be obtained as follows: The homomorphism descends to a surjective homomorphism . Viewing as a subgroup of via the inclusion , this is actually a homomorphism over . Composing with any projection yields a projection . There exist infinitely many projections , 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 -valued extension of the Calabi homomorphism to for one single open surface , for example , 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 ): We show that is independent of . Consider two area-preserving topological embeddings . Let . We need to show that in . By Proposition 2.1, we may pick a finite collection of relatively compact open discs and Hamiltonian homeomorphisms such that . Since is connected, we can also find Hamiltonian homeomorphisms such that . In we can then compute
as desired.
Step 2 (surjectivity): We show that is surjective. By Step 1, we may replace by a smooth area-preserving embedding. After identifying with its image under , we can regard as an open subset of . Given an arbitrary element , we need to show that the class has a representative in . By Proposition 2.1, there exist finitely many relatively compact open discs , each sufficiently small such that there exists such that , and Hamiltonian homeomorphisms such that . Set . Observe that in we have the identity
Therefore, is the desired representative of .
Step 3 (injectivity in the case of open surfaces): Assume that both and are open. We show that is injective. As in Step 2, we may assume that is smooth. We identify with its image under and view it as an open subset of . Our goal is to construct an inverse
of . By the universal property of abelianization, defining a group homomorphism is equivalent to defining a group homomorphism . Given , we define as follows. By Proposition 2.1, we may pick a fragmentation such that for some open disc which is sufficiently small such that there exists an area-preserving embedding . We set
Since is abelian, the order of the product of course does not matter. We need to check that this definition of is independent of the choice of fragmentation of . Postponing the proof of this claim, let us first argue that indeed induces a homomorphism which is an inverse of . Observe that is a group homomorphism because given fragmentations of , one can obtain a fragmentation of by concatenation. Consider . We may pick a fragmentation such that all the discs are contained in . We can then take the embeddings to be inclusions. With these choices, we see that . Since we already know that is surjective, this implies that is an isomorphism and that is its inverse.
Showing that the definition of is independent of choices reduces to verifying the following claim:
Let be open discs which are small enough such that they admit area-preserving embeddings into . Let be Hamiltonian homeomorphisms. Then
Fix discs and Hamiltonian homeomorphisms as in the claim and assume that . Set for . Our strategy is to construct, for , Hamiltonian homeomorphisms such that the following properties are satisfied:
-
1.
for all .
-
2.
for all .
-
3.
is smooth for all .
Once we have such homeomorphisms , the desired claim follows from the injectivity of . In order to see this, note that by Step 1 we may assume that the embeddings are smooth. By property 3, all are smooth as well. We can therefore compute
where we use that is independent of the embedding . By injectivity of , this means that is equal to the identity in . But since is the image of in , we see that is equal to the identity as well. It is then immediate from property 2 that is equal to the identity, as desired.
It remains to construct the homeomorphisms satisfying properties 1, 2, and 3. Fix and assume that the homeomorphisms have already been constructed for . We explain how to construct . For , we simply set . Note that all of these homeomorphisms are smooth. In order to simplify notation, we set
Note that is smooth and that
(3.2) |
Set
It follows from smoothness of and identity (3.2) that is smooth. As a consequence, any point at which is not smooth must be contained in . By Proposition 2.2, we can find which agrees with outside an arbitrarily small neighborhood of and is arbitrarily close to . Here we use that because is a disc, every compactly supported area-preserving diffeomorphism of is automatically contained in . We can then write , where is an area-preserving homeomorphism which is supported in an arbitrarily small neighborhood of and is arbitrarily close to the identity. Since is contained in , we can assume that . We claim that we can in addition assume that . Indeed, if this is not the case, simply pick close to the identity such that the mass flow homomorphism agrees on and . Then replace and by and , respectively. To summarize, we have constructed a factorization
(3.3) |
Moreover, we can take to be arbitrarily close to the identity. Now set
Observe that
Since is close to the identity, Proposition 2.1 allows us to fragment
(3.4) |
Using that is disjoint from by definition, we obtain
(3.5) |
We set
This concludes the construction of the Hamiltonian homeomorphisms and we need to check that they satisfy properties 1, 2, and 3. First, note that for all and for all 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 for , we need to check that
Substituting , we see that this is equivalent to showing thatv
It follows from identities (3.3) and (3.4) that
Thus it suffices to prove
In order to see this, simply observe that, for every , there exists such that . This concludes the proof that the homeomorphisms satisfy property 2 and thus the proof that the definition of the map is independent of choices.
Step 4 (embeddings of open into closed surfaces): Now consider an embedding of an open surface into a closed surface . Again, we assume that is smooth and identify with its image under . Since the group of Hamiltonian diffeomorphisms of the closed surface is perfect, the image of in is contained in the kernel of . Therefore, induces a homomorphism
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 , we pick a fragmentation with for open discs which are sufficiently small such that they admit embeddings . Given these choices, we set
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 descends to a homomorphism
which is an inverse to . In order to show that is independent of the choice of fragmentation, we have to show that if are open discs admitting embeddings , which we can assume to be smooth, and if are Hamiltonian homeomorphisms such that , then in the quotient . Going through the argument in Step 3 word by word, we see that there exist Hamiltonian diffeomorphisms such that and
in . Since the are Hamiltonian diffeomorphisms and the embeddings can assumed to be smooth, the left hand side of this identity is contained in . Thus the right hand side represents the identity in the quotient , which is exactly what we need to show for concluding that is well-defined and independent of choices. This finishes the proof that the kernel of is given by the image of in the case of an open surface embedded into a closed surface .
Step 5 (two closed surfaces): Any area-preserving embedding of a closed connected surface into another is necessarily an area-preserving homeomorphism, which clearly implies that 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 be an oriented smooth -manifold without boundary and possibly open. A Hamiltonian structure on is a closed, maximally nondegenerate -form on .
The line bundle is called the characteristic line bundle of . The foliation generated by this line bundle is called the characteristic foliation of . It is naturally cooriented by . Together with the ambient orientation of , this coorientation induces a natural orientation of the characteristic foliation.
The -form also induces a transverse measure on the characteristic foliation. Locally on every smooth transversal, the induced measure is diffeomorphic to the standard -dimensional Lebesgue measure. The (cooriented and measured) characteristic foliation uniquely determines the -form . We can therefore equivalently think of a Hamiltonian structure as a -dimensional cooriented and measured smooth foliation with the property that the induced measure on transversals is locally diffeomorphic to the -dimensional Lebesgue measure.
A diffeomorphism of Hamiltonian structures and is a diffeomorphism which satisfies .
Example 4.1.
Consider equipped with coordinates and oriented with respect to these coordinates.
We will refer to the closed, maximally nondegenerate -form as the standard Hamiltonian structure on .
For , let denote the ball of area and consider the product . The restriction of the standard Hamiltonian structure on induces a Hamiltonian structure on . We will abbreviate it by the same symbol and refer to it as the standard Hamiltonian structure on .
Example 4.2.
Consider a surface , without boundary, equipped with an area form . Let be the projection onto the second factor. We equip with the orientation induced by the volume form , where denotes the coordinate on . Then, is a Hamiltonian structure on for which the oriented leaves of the characteristic foliation are given by the flow lines of the vector field . This Hamiltonian structure will be denoted by .
Now, let be a smooth isotopy in such that for close to and for close to . Let be the unique compactly supported Hamiltonian generating . Then, the -form
is another Hamiltonian structure on . It agrees with outside a compact set and the leaves of its characteristic foliation are of the form for .
Hamiltonian structures satisfy a version of the Darboux neighborhood theorem which we state here as a lemma.
Lemma 4.3.
Let be a Hamiltonian structure on . For every point in , there exists a neighborhood such that is diffeomorphic to , for some .
Remark 4.4.
For every Hamiltonian structure on , there exist a symplectic -manifold and an embedding such that . Moreover, by Gotay’s theorem [24], any two such embeddings , are neighborhood equivalent in the following sense: There exist open neighborhoods and a symplectomorphism such that .
4.2 Flux and Helicity
Let be a closed oriented -manifold equipped with a Hamiltonian structure .
Definition 4.5.
We define the flux of the Hamiltonian structure to be the cohomology class
We say that a Hamiltonian structure is exact if .
Recall that the helicity of an exact Hamiltonian structure is defined to be the quantity
where is any primitive of . We check in the lemma below that helicity is well-defined.
Lemma 4.6.
Let be an exact Hamiltonian structure on and let and be two -forms such that . Then,
Proof.
The -form is closed and the -form is exact. Hence, the -form is exact and consequently, by Stokes’ theorem, we have
∎
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 be a Hamiltonian structure on . A plug is a tuple consisting of
-
1.
an open surface , not necessarily connected;
-
2.
an area form on ;
-
3.
a smooth embedding of Hamiltonian structures ;
-
4.
a smooth isotopy in such that for close to and for close to .
We define the Calabi invariant of the plug to be
(4.1) |
Given a plug , we define a new Hamiltonian structure which coincides with outside of and with inside , where is the unique compactly supported Hamiltonian generating . Note that is well-defined because vanishes near the boundary of .
The Hamiltonian structure has the following equivalent definition which is better suited for the setting which we consider below. Define by . Then is defined to agree with outside of and with inside . Note that before plug insertion, the characteristic leaves of inside are of the form for . After plug insertion, the characteristic leaves of are of the form
for .
Although the Hamiltonian structure depends on the entire isotopy , its diffeomorphism type depends only on the time-1 map .
Lemma 4.8.
Consider two plugs and . If , then there exists a diffeomorphism of Hamiltonian structures which is supported inside and is isotopic to the identity through diffeomorphisms supported inside .
We omit the proof of the above lemma, as its generalization in the 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 be a Hamiltonian structure on a closed -manifold and let be an -plug. Then,
Proof.
Write . Let be the unique compactly supported Hamiltonian generating . Then the -form is compactly supported in and can hence be extended to a -form on by setting it equal to zero outside of . By definition, the Hamiltonian structures and differ by the exact -form . This clearly implies that . ∎
Lemma 4.10.
Let be a closed -manifold and let be an exact Hamiltonian structure on . Let be an -plug. Then
5 Hamiltonian structures
In this section, we introduce Hamiltonian structures and plugs.
5.1 Basic definitions
Recall the definition of the standard Hamiltonian structure on from Example 4.1. A Hamiltonian structure on a topological -manifold is a -dimensional cooriented foliation, equipped with a transverse measure, which is locally modeled on with the characteristic foliation induced by . Explicitly, this notion can be formalized in terms of Hamiltonian atlases.
Definition 5.1.
Consider equipped with the smooth Hamiltonian structure . Let be a homeomorphism between two open subsets of . We say that preserves the standard Hamiltonian structure on if, for every point , there exist an open neighborhood of of the form , where is an open interval and is a ball, and a continuous area- and orientation-preserving embedding such that
A Hamiltonian atlas on a topological -manifold is an atlas whose transition maps are homeomorphisms preserving the standard Hamiltonian structure on . Two such atlases are considered equivalent if the transition maps between the two atlases preserve the standard Hamiltonian structure on . A Hamiltonian structure on is an equivalence class of Hamiltonian atlases.
Definition 5.2.
A homeomorphism of Hamiltonian structures is a homeomorphism which, in local charts, preserves the standard Hamiltonian structure on . A topological embedding of Hamiltonian structures is a topological embedding which is a homeomorphism of the Hamiltonian structures and .
Let be a homeomorphism. Given a Hamiltonian structure on , we can pull back its defining atlas, via , to obtain a Hamiltonian structure on which we abbreviate by and refer to as the pullback of via . We define the pushforward of a Hamiltonian structure on to be the pullback of via ; we abbreviate this by .
Remark 5.3.
A smooth Hamiltonian structure on a smooth -manifold gives rise to a Hamiltonian structure. Indeed, consider the atlas of consisting of all Darboux charts. The transition maps of this atlas are diffeomorphisms between open subsets of which satisfy as smooth -forms. Note that for a diffeomorphism , this condition is equivalent to preserving the standard Hamiltonian structure on in the sense of Definition 5.1. Hence our atlas consisting of smooth Darboux charts is a Hamiltonian atlas.
Remark 5.4.
Let be a topological -manifold. In what follows, by a smooth Hamiltonian structure on we mean a choice of smooth structure on and a maximally nondegenerate closed -form. Alternatively, a smooth Hamiltonian structure on a topological -manifold can be specified by an atlas on whose transition maps are diffeomorphisms preserving the standard Hamiltonian structure on .
5.2 plugs
Analogous to the smooth plugs from Section 4.3, we introduce the concept of plugs, which can be inserted into Hamiltonian structures to generate new Hamiltonian structures. We also define the Calabi invariant of plugs.
Throughout the rest of this section, let denote a topological -manifold without boundary and possibly open.
Definition 5.5.
Let be a Hamiltonian structure on . An -plug (or simply a plug, if is clear from the context) is a tuple consisting of
-
1.
a smooth, open surface , not necessarily connected;
-
2.
a smooth area form on ;
-
3.
a topological embedding of Hamiltonian structures ;
-
4.
a continuous isotopy in such that for close to and for close to .
We define the Calabi invariant of to be
Here, we recall that denotes the universal -valued extension of the Calabi homomorphism; Section 3.
We now describe how an -plug can be inserted into yielding a new Hamiltonian structure . Define by
Inside , we define as the pushforward , where is regarded as a Hamiltonian structure on . Note that agrees with near the boundary of . We define to agree with outside . The effect of plug insertion on the characteristic foliation is as follows: Before plug insertion, the characteristic leaves of inside are of the form for . After plug insertion, the characteristic leaves of are of the form
for .
The inverse of an -plug is defined to be the -plug
(5.1) |
Note that if is an -plug, then and . A plug is called trivial if for all . If is trivial, then . If is a homeomorphism of Hamiltonian structures, then the pull back of the -plug via is the -plug defined by
(5.2) |
We observe that is a homeomorphism of Hamiltonian structures
(5.3) |
Remark 5.6.
Consider an -plug . It is always possible to slightly shrink the image of without affecting or the Calabi invariant as follows: Pick a relatively compact open subset such that the entire isotopy is supported in . Let such that is equal to the identity for all and equal to for all . Let be an orientation-preserving diffeomorphism which agrees with the identity on the interval . Define an embedding of Hamiltonian structures by setting . Set . The closure of the image of is clearly contained in the image of . Moreover, . Finally, by the naturality of the -valued extension of Calabi, we have .
Analogously to Lemma 4.8 in the smooth setting, although the Hamiltonian structure depends on the entire isotopy , its homeomorphism type depends only on the time-1 map .
Lemma 5.7.
Let be a Hamiltonian structure on . For , let be two -plugs. If , then there exists a homeomorphism of Hamiltonian structures which is supported inside and isotopic to the identity through homeomorphisms supported in .
Proof.
Define by . Then
is a homeomorphism of Hamiltonian structures. Since , it is compactly supported. We can therefore define the desired homeomorphism by setting inside and extending it to be the identity on the complement of this set.
Now, the surface is open and therefore not . Hence is simply connected, as explained in Section 2. It follows from simply connectedness of that the homeomorphism is isotopic to the identity via homeomorphisms supported in . ∎
The next lemma states that sliding a plug along the characteristic foliation has no effect on the homeomorphism type of the Hamiltonian structure.
Lemma 5.8.
Let be a Hamiltonian structure on . For , let be two -plugs. Assume that there exists a continuous family of topological embeddings of Hamiltonian structures connecting to such that is contained in the same characteristic leaf of as for all and . Then, there exists a homeomorphism of Hamiltonian structures which is supported inside and isotopic to the identity through homeomorphisms supported in .
Proof.
Define by . Let be a compact subset such that agrees with on the complement of . As a consequence of the assumption on the embeddings , we can find a family of homeomorphisms starting at the identity obtained by moving points along the characteristic leaves of such that is supported inside and such that for all . Then, is a homeomorphism supported inside . By construction, is isotopic to the identity through homeomorphisms supported in . ∎
6 Smoothings modulo plugs
In this section we prove Theorems 6.1 & 6.11 on smoothing Hamiltonian structures and their homeomorphisms.
6.1 Smoothing Hamiltonian structures
We prove in this section that every Hamiltonian structure on a closed -manifold can be obtained from a smooth Hamiltonian structure via the insertion of a plug. This result, stated below, plays a crucial role in the remainder of the paper.
Theorem 6.1.
Let be a Hamiltonian structure on a closed topological -manifold . Then, there exist a smooth Hamiltonian structure on and a -plug such that .
Remark 6.2.
Recall that a smooth Hamiltonian structure on includes the choice of a smooth structure on , see Remark 5.4. Thus Theorem 6.1 in particular contains the statement that the closed topological -manifold can be smoothed. It is of course well-known that every topological -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 -manifold can be triangulated and that every triangulated -manifold is piecewise linear, and the fact that every piecewise linear -manifold can be smoothed, see e.g. [53, Theorem 3.10.8]. As we will see, the existence of a Hamiltonian structure on actually simplifies the problem of smoothing the underlying manifold . Essentially, it reduces the problem to smoothing a topological -manifold.
Remark 6.3.
Theorem 6.1 is clearly false if we do not allow for the insertion of a plug, i.e. there exist Hamiltonian structures which cannot be globally upgraded to smooth Hamiltonian structures. Indeed, consider a closed surface and an area-preserving homeomorphism of which is not conjugate to any area-preserving diffeomorphism. For example, we can take to have infinite topological entropy. The Hamiltonian structure on descends to a Hamiltonian structure on the mapping torus
This Hamiltonian structure is not homeomorphic to any smooth Hamiltonian structure.
Let and be as in the statement of Theorem 6.1. Let be a fixed-point-free topological flow whose flow lines equipped with the orientation induced by agree with the oriented characteristic leaves of . The existence of is guaranteed by Proposition 8.6.
Definition 6.4.
We say a subset is exhaustive if there exists a constant such that, for any point , there exist times such that .
We remark that whether or not is exhaustive is independent of the choice of the flow .
Remark 6.5.
Our proof of Theorem 6.1 shows the following slightly stronger statement, which we record here for later reference: Let be a closed topological -manifold equipped with a Hamiltonian structure . Let be an arbitrary exhaustive open set. Then we can find a smooth Hamiltonian structure on and an -plug whose image is contained in such that .
The following lemma guarantees the existence of an exhaustive flow box.
Lemma 6.6.
Let be a Hamiltonian structure on a closed topological -manifold , and let be an open exhaustive subset. Then there exist an open surface equipped with an area form, a compact subsurface with boundary , and a topological embedding of Hamiltonian structures
such that the image is exhaustive.
Proof.
Since is exhaustive, we can fix as in Definition 6.4. Consider . Then there exist such that . For sufficiently small, we can find embeddings of Hamiltonian structures such that . Moreover, we can find an open neighborhood of with the property that if we start at any point and follow the flow forward/backward in time, we meet within time at most .
Since is compact, we can cover it by finitely many neighborhoods . This implies that there exists a finite collection of embeddings of Hamiltonian structures with ranging from to such that the union is an exhaustive set. We remark that the surfaces are not necessarily pairwise disjoint.
Our next step is to construct, for each , finitely many closed discs of the form such that and the images , ranging over all , are pairwise disjoint.
The construction of these discs relies on Claim 6.7, stated below. Let and consider a closed subset which is a -dimensional topological submanifold, possibly with boundary, and which is transverse to the characteristic foliation on in the following sense: near every point of there exists a local Hamiltonian chart in which both and the characteristic foliation are smooth and in which is transverse to the characteristic foliation in the usual sense.
Claim 6.7.
We can find finitely many pairwise disjoint closed discs in of the form which are all disjoint from and such that .
Proof of Claim 6.7.
Indeed, consider an arbitrary point . It follows from the transversality assumption on that the characteristic leaf contains a point in the complement of . This means that we can find a small closed disc containing and such that is disjoint from . Since is compact, we can find finitely many discs in the complement of such that the cover . We can in addition make these discs pairwise disjoint by slightly perturbing the heights so that the are all distinct. This proves the claim. ∎
Returning to the construction of the discs , let and assume that the discs have already been construction for all . Now simply set
and apply the above claim to find the discs . This concludes our construction of the discs . We observe that since is an exhaustive set and is contained in the union , the set is exhaustive as well.
For each disc , pick a slightly larger open disc . Let be small. Consider the embedding
which on each component is given by the restriction of the corresponding embedding . Here we have to choose sufficiently small and take the enlarged discs sufficiently close to the original discs to ensure that this indeed defines an embedding and has image contained in .
We define to be the disjoint union of all the discs and to be the disjoint union of all the discs . Then the desired embedding is obtained from the above embedding by reparametrizing the intervals to . ∎
We now present a proof of Theorem 6.1.
Proof of Theorem 6.1.
By Lemma 6.6, applied to the exhaustive set , we may pick an open surface , a compact subsurface with boundary , and an embedding of Hamiltonian structures such that is exhaustive. In order to prove Remark 6.5, we note that if we are given an arbitrary exhaustive open subset , we can choose such that its image is contained in . Let and define the compact interval . Then the set is also exhaustive. Up to possibly enlarging , we can assume that the interior of is still exhaustive.
Set and equip it with the restriction of the Hamiltonian structure . Our goal is to construct a smoothing of the Hamiltonian structure . The crucial observation that allows us to achieve this is that the characteristic foliation of exhibits no complicated recurrent behavior. The closure (with respect to ) of any characteristic leaf of is an embedded compact interval which starts at and ends at .
Let denote the leaf space of the characteristic foliation of and let denote the natural projection.
Claim 6.8.
The leaf space has the structure of a -dimensional, topological, possibly non-Hausdorff manifold equipped with a measure which in local charts is homeomorphic to the standard Lebesgue measure on . This means that admits an atlas whose transition maps are area-preserving homeomorphisms between open subsets of .
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 -manifolds in connection with foliations of .
Proof.
Consider a point . Since is an open subset of , for sufficiently small, we can chose an embedding of Hamiltonian structures such that . We show that after possibly shrinking , we can assume that the restriction of the projection to is injective. Indeed, assume by contradiction that this is not the case. Then there exist a sequence of points and a sequence of times such that and such remains in as ranges from to . Here, is a choice of flow as in Definition 6.4. Since the complement of is exhaustive, the sequence must be bounded. Clearly, the sequence must also be bounded away from zero. After passing to a subsequence, we can therefore assume that converges to . This implies that is a periodic orbit of of period . Moreover, this periodic orbit must be contained in the closure of since it is the limit of flow line segments contained in . But this contradicts the fact that the complement of the closure of (i.e. the interior of ) is exhaustive.
The above discussion yields a continuous injective map
mapping to . We claim that this map is open. This amounts to showing that is open for every open subset . In order to see this, consider a point . The point being in means that there exists a line segment contained in a characteristic leaf of which connects to a point in . Since and are open, every point in a neighborhood of can also be connected to via a line segment contained in a leaf of . Thus contains a neighborhood of , showing that is open.
We conclude that is a homeomorphism between and an open neighborhood of in . Since was chosen arbitrarily, this implies that is a -dimensional, topological, possibly non-Hausdorff manifold. The transverse measure on the characteristic foliation of clearly descends to a measure on locally homeomorphic to the standard -dimensional Lebesgue measure. ∎
Claim 6.10.
has a smoothing, i.e. it admits an atlas whose transition maps are area-preserving diffeomorphisms between open subsets of .
Proof.
It is well-known that every Hausdorff topological -manifold admits a smoothing. Indeed, by a theorem of Radó [45] (see also [40, §8, Theorem 3]) every Hausdorff topological -manifold admits a piecewise linear structure. Moreover, piecewise linear -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 admits a finite open covering by charts with . In the following, it will be useful to be able to replace these charts by smaller charts for relatively compact open subsets which still cover . We can choose our initial charts 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 , we set and let
denote the transition map, which is an area-preserving homeomorphism. Our goal is to modify the charts such that the transition maps become smooth for all . Let . Suppose we have already modified the charts in such a way that is smooth for all and such that is smooth for all . We explain how to modify the charts to make smooth without destroying smoothness of the transition maps we have already made smooth. Set . It follows from our assumptions that the restriction of to is smooth. After slightly shrinking all chart neighborhoods , we can assume that is smooth on an open neighborhood of the closure of inside . Pick a non-negative continuous function such that is contained in and is contained in the smooth locus of . Moreover, assume that decays to towards the boundary of . By Proposition 2.2, we can find an area-preserving diffeomorphism such that for all . Since decays to towards the boundary of , the homeomorphism of extends to an area-preserving homeomorphism of which agrees with the identity on the complement of . Now define the chart . The transition map between and is simply given by and is therefore smooth. Moreover, note that since and agree on , the homeomorphism restricts to the identity on for all . Thus the transition map does not change and remains smooth. Now, simply replace by . This concludes our construction of a smoothing of ∎
We will now use the smooth structure on to construct a smoothing of , i.e. we will construct a Hamiltonian atlas for whose transition maps are smooth.
Recall that we have a fixed-point-free flow on whose flow lines along with their natural orientations agree with the oriented leaves of the characteristic foliation of . Let be sufficiently small and pick finitely many topological embeddings such that the images of cover and such that each is an area-preserving embedding which is smooth with respect to the smooth structure on constructed in Claim 6.10. Given two embeddings and , we define
Note that, for every point , there exists a unique transfer time such that and for all ranging from to . The functions are continuous. We can perturb the embeddings , without changing , such that all the 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 to be smooth if is a smooth area-preserving embedding and if, for every embedding , the locally defined transfer time between the image of and the image of is smooth. Note that by construction, if the transfer time between the image of and the image of is smooth near some point , then the transfer time between the image of and the image of any other is smooth near as well. In particular, this implies that we can find a smooth embedding passing through any point in .
Finally, consider all Hamiltonian embeddings of the form
for some and some smooth embedding . We can cover by images of embeddings of this form. The inverses of these embeddings yield the desired collection of Hamiltonian charts with smooth transition maps.
We have now constructed a smoothing of ; however, it does not in general extend to a global smoothing of . The problem is the following: Let and be two surfaces which are transverse to the characteristic foliation and smoothly embedded with respect to the smoothing of constructed above. Moreover, suppose that each leaf segment of the form for intersects each of the surfaces exactly once. Traversing along the characteristic foliation then yields a homeomorphism . This homeomorphism is not going to be smooth in general and this is exactly the obstruction to extending the smoothing of to all of . We may, however, modify by inserting an -plug of the form such that the resulting map becomes smooth. We may then extend our smoothing of to a global smoothing of the Hamiltonian structure on .
Now, simply set and , where denotes the inverse plug of . Then, , as desired.
To prove the strengthened statement in Remark 6.5, the image of is contained in the given exhaustive open set because the embedding was chosen to have image contained in . ∎
6.2 Smoothing Hamiltonian homeomorphisms
We prove in this section that homeomorphisms of Hamiltonian structures on closed -manifolds may be smoothed, up to insertion of plugs.
Theorem 6.11.
For , let be a closed -manifold and let be a smooth Hamiltonian structure on . Let be a -plug and assume that
is a homeomorphism of Hamiltonian structures. Then, there exist smooth -plugs for and a diffeomorphism of Hamiltonian structures
with the following properties:
-
1.
-
2.
is homotopic to through homeomorphisms.
Remark 6.12.
As an immediate corollary of Theorem 6.11, one obtains that two exact smooth Hamiltonian structures and on which are homeomorphic have the same helicity. Indeed, we can apply Theorem 6.11 with empty plugs . Then the resulting smooth -plugs have the same (smooth) Calabi invariant, i.e. . Since the smooth Hamiltonian structures and are diffeomorphic, they have the same helicity. Therefore, we can conclude that by Lemma 4.10.
The proof of Theorem 6.11, which takes up the rest of this section, requires some preparatory lemmas.
Let be a surface equipped with an area-form . Let denote the cylinder ; we equip it with the smooth Hamiltonian structure induced by . We also introduce the notation
Figure 6.1 illustrates and may aid the reader during the proof of Lemma 6.13. In that proof, it is helpful to visualize and as being thin relative to . This can be made precise via a reparametrization of the interval , but we omit the details to keep the notation light and the argument clear.
Lemma 6.13.
Let be a closed topological -manifold equipped with a Hamiltonian structure . Let be two open sets such that . Assume that is exhaustive in the sense of Definition 6.4.
Then there exist open surfaces with area forms , relatively compact open subsets , and topological embeddings of Hamiltonian structures
satisfying the following properties:
-
1.
-
2.
for all .
-
3.
for all .
Moreover, if the restriction of to is smooth, we can take all to be smooth as well.
Proof.
Let be an arbitrary point. Let denote the characteristic leaf passing through . By the exhaustiveness assumption on , there exists a topologically embedded closed interval such that is contained in the interior of and the endpoints of are contained in . Using sufficiently small local transverse sections of the characteristic foliation at the endpoints of , one can construct an open surface with area form and a topological embedding containing in its image such that . If the restriction of to is smooth, we can construct to be smooth as well.
Since the set is compact, we can cover it with the images of finitely many such cylinder embeddings. We may therefore pick and tuples for such that
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 to make sure that the intersection between and for and is non-empty if and only if and .
Moreover, we can arrange to have the property that, for all and such that the intersection is non-empty, there exist an open subset and an open interval such that
In order to resolve, and remove, the intersection between and , we replace by three tuples with as described in detail below. At the end of this process, the cylinder will be replaced by three cylinders , which will be constructed in such a way that they do not intersect . This, of course, requires removing some portions of , and also . The removed portions will all be contained in and hence property 1 will continue to hold. We will also ensure that property 2 continues to hold. However, this is less complicated.
We now proceed with a detailed description of the above process, which is also depicted in Figure 6.2. Pick open sets
such that
Moreover, we can arrange to be contained in an arbitrarily small neighborhood of the closure of inside . Note that has two components and . We pick orientation-preserving embeddings
such that
for some small . We define
If is chosen sufficiently small and is contained in a sufficiently small neighborhood of the closure of in , then the six sets for are contained in and are disjoint from for any . They are also pairwise disjoint. Moreover, is contained in , which implies that
The three sets for are disjoint from . In addition, is disjoint from for . We discard the tuple and instead insert the three tuples between and . 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 for and therefore also achieve property 3. ∎
Definition 6.14.
Let be open subset of . Suppose that is a homeomorphism of Hamiltonian structures. We say that is transversely smooth if, for every point , there exist an open neighborhood of of the form , where is an open interval and is a ball, and a smooth area-preserving embedding such that
Let be a homeomorphism of Hamiltonian structures, where are smooth. We say that is transversely smooth if it is transversely smooth in smooth local charts.
Lemma 6.15.
For , let be a closed -manifold with a smooth Hamiltonian structure. Suppose that is a transversely smooth homeomorphism of Hamiltonian structures. Then, there exists a diffeomorphism of Hamiltonian structures which is isotopic to through homeomorphisms of Hamiltonian structures . Moreover, the homeomorphisms all induce the same homeomorphism between the leaf spaces of and .
Proof.
Using smooth local charts on and , one can reduce the proof of the lemma to the following claim.
Claim 6.16.
Consider open intervals and open balls . Let be an open subset. Let denote the projection onto the second factor. Consider an embedding with the property that . Assume that the restriction of to is a smooth embedding.
Then there exists a family of embeddings , for , satisfying and agreeing with outside some compact subset of , such that the restriction of to is a smooth embedding.
In order to see that this claim implies the lemma, note that since is transversely smooth, in suitable smooth local Darboux coordinates it precisely takes the form of an embedding as in the claim. Therefore, the assertion of the claim allows us to go through a finite covering of by Darboux charts and step by step make smooth in each of the charts.
We briefly outline the proof of the claim. The embedding is of the form
where denotes a point in . For each fixed , the function is strictly increasing. The restriction of to is smooth and the partial derivative is strictly positive in this region. We can find a function agreeing with outside some compact subset of such that is strictly increasing for every and such that the restriction of to is smooth and has strictly positive partial derivative in this region. Moreover, we can find a path of functions connecting to such that each agrees with outside a compact subset of and such that is strictly increasing for all . We can then define
which satisfies the requirements of the claim. ∎
Remark 6.17.
Let be a -manifold equipped with a smooth Hamiltonian structure and let be a -plug. Using Proposition 2.2, it is possible to find an area-preserving homeomorphism of , not necessarily compactly supported, such that 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 by slightly sliding it along the leaves of the characteristic foliation of to obtain a smooth embedding of Hamiltonian structures . Define the -plug
One can check, using Lemma 5.8, that is homeomorphic to via a homeomorphism isotopic to the identity. Moreover, it follows from the naturality of the -valued extension of the Calabi homomorphism that . The upshot of this remark is that we can replace any -plug by an -plug whose underlying embedding is smooth.
Proof of Theorem 6.11.
First, let us consider the special case where is a trivial plug. In this simplified situation, we will construct a smooth plug satisfying and a diffeomorphism of Hamiltonian structures
which is isotopic to through homeomorphisms. Note that this in particular implies that .
Let denote the image of the plug . After possibly adding trivial components to the plug , we can assume that is exhaustive in the sense of Definition 6.4. This is possible by Lemma 6.6. Note that the addition of trivial components to does not change the Calabi invariant of the plug. Pick an open neighborhood of such that agrees with on and is thus smooth in this region. We apply Lemma 6.13. This yields smooth open surfaces , relatively compact open subsets , and smooth embeddings of Hamiltonian structures satisfying all the properties listed in Lemma 6.13.
In what follows, we will construct -plugs which are of the form
where is an isotopy in which will be chosen below. By the properties listed in Lemma 6.13, the plugs are pairwise disjoint and contained in . Let denote the Hamiltonian structure obtained by inserting the plugs for into . We will show that for appropriately chosen there exists a homeomorphism of Hamiltonian structures
whose restriction to is transversely smooth. Note that the assertion on transverse smoothness of makes sense because the restriction of to the complement of is a smooth Hamiltonian structure.
We will now describe an inductive procedure for constructing and as above. Let and suppose that and have already been constructed for all . We explain how to construct and . Consider the composition
which is a topological embedding of Hamiltonian structures. Note that the leaf space of has the structure of a smooth -dimensional manifold with area form. We abbreviate it by . The embedding induces an area-preserving homeomorphism . Let denote the projection onto the second factor. Note that the restriction of to the set
is smooth by the transverse smoothness assumption on for .
Using Proposition 2.2, one can find an area-preserving homeomorphism which agrees with outside some compact set and whose restriction to is smooth. Indeed, pick a continuous function which decays to zero towards the boundary of , whose zero set is contained in , and which has the property that there exists an open neighborhood of in such that . Now apply Proposition 2.2 to with this function . This yields an area-preserving diffeomorphism . Since decays to zero near the boundary of , we can extend to an area-preserving homeomorphism by setting it equal to outside of . Since , we see that agrees with on and is therefore smooth in this region. We conclude that the restriction of to is smooth.
Moreover, we can pick such that is in . Indeed, note that since is a approximation of , the homeomorphism is contained in the identity component . If is not in , simply pick a compactly supported area-preserving diffeomorphism such that the mass flow homomorphism on takes the same value on and . Then replace by .
Since , we can find an isotopy whose time-1 map satisfies
The map is constructed by modifying inside , hence it agrees with outside . We describe the construction of on each of . Inside , we define such that
Inside , we define such that
Inside , we define such that
One can verify that the above yield a well-defined map on which coincides with near the boundary of . The plugs were defined precisely such that is a homeomorphism of Hamiltonian structures . Finally, note that the map induced by is precisely given by . Since the restriction of to is smooth, we see that the restriction of to is transversely smooth. Since is isotopic to by assumption, so is . This concludes our construction of and the isotopy .
To summarize, at this point we have a homeomorphism of Hamiltonian structures
which, by property 1 in Lemma 6.13, is transversely smooth on . Moreover, is obtained from via the insertion of the plugs .
Since the plugs are pairwise disjoint and all contained in , there exist a single -plug , contained in , and a homeomorphism which is isotopic to the identity and compactly supported inside . The Calabi invariant of is given by
where we use naturality of the tautological extension of the Calabi homomorphism and the identity
Like , the homeomorphism
is transversely smooth on the complement of . Write for the data of the plug . Note that we can assume that the embedding is smooth with respect to the smooth Hamiltonian structure , see Remark 6.17.
Claim 6.18.
The map is smooth.
Proof.
Pick sufficiently small such that is equal to for and equal to for . Note that agrees with on and on and that is transversely smooth on these sets.
Set and . Traversing along the characteristic foliation of induces a homeomorphism
Similarly, traversing along the the characteristic foliation of induces a homeomorphism
These homeomorphisms fit into the following commutative diagram:
Since is transversely smooth on a neighborhood of and and is a smooth Hamiltonian structure, we see that is smooth. ∎
As a consequence of Claim 6.18, we may modify the plug by replacing with a smooth Hamiltonian isotopy with the same time-1 map is . By Lemma 5.7, modifying the plug in this way yields a smooth Hamiltonian structure homeomorphic to via a homeomorphism supported in 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 . We then have a homeomorphism of Hamiltonian structures
which is isotopic to and transversely smooth outside of . Since both and are smooth and no characteristic leaf of is trapped in , we can conclude that is transversely smooth everywhere. By Lemma 6.15, we can replace by a diffeomorphism of Hamiltonian structures, still isotopic to . This concludes the proof of Theorem 6.11 in the special case that is a trivial plug.
It remains to treat the general case. Let be the inverse plug of and let be its pull back via ; see Equations (4.1) and (5.2). Then, as noted in (5.3), is a homeomorphism of Hamiltonian structures
We also have
As we explain below, we can slide the plug along the characteristic foliation of so that it becomes disjoint from , as described in Lemma 5.8.
Let denote the resulting plug disjoint from . It is both an - and an -plug and has Calabi invariant . Let be the -plug given by the disjoint union of and . By Lemma 5.8, is homeomorphic to via a homeomorphism isotopic to the identity. Thus is homeomorphic to via a homeomorphism isotopic to . By the special case of Theorem 6.11 already treated above, we can find a smooth -plug such that there exists a diffeomorphism of smooth Hamiltonian structures , still isotopic to , and such that
Now simply define to be the empty plug. Then the tuple clearly satisfies all assertions of Theorem 6.11.
It thus remains to explain why we can always reduce to a situation where the plug can be made disjoint from by sliding it along the characteristic foliation of .
For , write . Note that by Remark 6.17, we can assume that are smooth embeddings of Hamiltonian structures. After possibly shrinking the images of as in Remark 5.6, we can assume that there exist surfaces containing as relatively compact open subsets such that extends to a smooth embedding of Hamiltonian structures
Consider a compact surface with smooth boundary containing in its interior. Let and let be an arbitrary open neighborhood of . Then there exists a compactly supported homeomorphism of Hamiltonian structures
such that . Indeed, any compactly supported homeomorphism of takes the form for a family of compactly supported homeomorphisms of which agree with the identity for outside some compact subset of . One can arrange such that, for every , the homeomorphism maps an arbitrarily small neighborhood of to the interval . The image of under the resulting homeomorphism will then contain , as desired. Let us also point out that the group of compactly supported homeomorphisms of is connected. In particular, is isotopic to the identity through compactly supported homeomorphisms of Hamiltonian structures.
We can extend the isotopy in to an isotopy in which is equal to the identity outside of , and agrees with the identity for and with for . We can then define the homeomorphism
see Subsection 5.2. Recall that by definition, is an embedding of Hamiltonian structures
The homeomorphism leaves invariant. The above discussion thus implies that for every open neighborhood of , there exists a compactly supported homeomorphism of isotopic to the identity through such homeomorphisms such that contains . From this observation, we obtain the following statement: Suppose that the closure of the image of is disjoint from . Then one can slide the plug along the characteristic foliation of to make it disjoint from .
We therefore need to explain how to reduce to the case where the closure of the image of is disjoint from . We set . This defines a compact surface in which is transverse to the characteristic foliation, in the sense described in the proof of Claim 6.7. In what follows, we identify with its image under the embedding and work in these coordinates. Since is transverse to the characteristic foliation, we can find finitely many open discs covering and numbers such that the closures of the discs are disjoint from (cf. proof of Claim 6.7). Now, choose embeddings of Hamiltonian structures
such that the closures of their images are pairwise disjoint and also disjoint from .
Pick a smooth isotopy in which approximates the continuous isotopy ; the existence of is well-known and it can be deduced from the following two facts: first, every homeomorphism in can be written as a limit of elements in ; and second, and are both locally path connected; see, for example, [49, Cor. 2] or [50, Lem. 3.2]. We can take to be arbitrarily close to the identity by choosing better approximations . By Proposition 2.1, we can therefore find a fragmentation into Hamiltonian homeomorphisms which are close to the identity. For each , pick a small isotopy connecting the identity to . We define -plugs
Note that these plugs are disjoint. Moreover, is a smooth plug because both and are smooth. We also define the -plug to be the disjoint union of the plugs . Using Lemma 5.7 and the identity , we see that there exists a homeomorphism of Hamiltonian structures
isotopic to the identity. Moreover, since is close to and all are close to the identity, it is clear from the proof of Lemma 5.7 that the homeomorphsim can be taken close to the identity. In particular, we can assume that is disjoint from the closure of the image of . Now consider the composite
and note that is disjoint from . As discussed above, we can slide away from along the characteristic foliation of . Hence, there exist a smooth -plug of Calabi invariant
(6.1) |
and a diffeomorphism of Hamiltonian structures isotopic to and hence also to . Now observe that by naturality of our -valued extension of the Calabi homomorphism, we have
Together with identity (6.1), this readily yields the desired identity
This completes the proof of Theorem 6.11 in the general case. ∎
7 Flux and helicity of Hamiltonian structures
In this section, we use the results of Section 6 to extend the definitions of flux and helicity to Hamiltonian structures. We prove our main result, Theorem 1.5, on the existence and uniqueness of a universal -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 Hamiltonian structures.
We begin with the following lemma, which is a straightforward consequence of Theorem 6.11.
Lemma 7.1.
For , let be a closed -manifold and let be a smooth Hamiltonian structure on . Let be a -plug and assume that
is a homeomorphism of Hamiltonian structures. Then,
Proof.
By Theorem 6.11, we can find a smooth -plug for and a diffeomorphism of Hamiltonian structures
which is isotopic to . Clearly, we have
Recall from Lemma 4.9 that . The assertion of the lemma now follows from the observation that the actions of and on cohomology agree because these two maps are isotopic. ∎
Definition 7.2.
Note that if is a smooth Hamiltonian structure, then because we can simply take to be the trivial plug in Definition 7.2. Moreover, it is immediate from Lemma 7.1 that if is a homeomorphism of Hamiltonian structures, then .
Lemma 7.3.
Let be a Hamiltonian structure on a closed topological manifold and let be an -plug. Then
Proof.
After possibly slightly shrinking the image of the plug , see Remark 5.6, we can assume that the complement of the image of contains an exhaustive open set . By Theorem 6.1 and Remark 6.5, we can find a smooth Hamiltonian structure on and a -plug whose image is contained in such that . Let be the disjoint union of the plugs and . Now is an -plug and . Using the definition of , we can therefore compute
∎
7.2 Helicity
The goal of this section is to prove Theorem 1.5.
The uniqueness of the extension in Theorem 1.5 is a consequence of Theorem 6.1. Indeed, consider an arbitrary exact Hamiltonian structure on a closed topological -manifold . By Theorem 6.1, there exist a smooth Hamiltonian structure on and an -plug such that . By Lemma 7.3, the smooth Hamiltonian structure is exact. It follows from properties 1, 2, and 3 in Theorem 1.5 that we must have
(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 , let be a closed -manifold and let be an exact smooth Hamiltonian structure on . Let be a -plug and assume that
is a homeomorphism of Hamiltonian structures. Then
Proof.
Let be an arbitrary exact Hamiltonian structure on a closed topological -manifold . By Theorem 6.1, we may write for a smooth Hamiltonian structure and a -plug . By Lemma 7.3, the smooth Hamiltonian structure is exact. The idea is to use identity (7.1) as a definition, i.e. to set
(7.2) |
By Lemma 7.4, this definition is independent of choices. It remains to check that our extension satisfies properties 1, 2, and 3 in Theorem 1.5. For property 1, observe that if is already smooth, we can simply take to be the trivial plug in (7.2). Property 2 is immediate from Lemma 7.4. Finally, suppose that is an exact Hamiltonian structure and that is an -plug. After possibly slightly shrinking the image of , we can assume that the complement of contains an exhaustive open set . By Theorem 6.1 and Remark 6.5, we can write for a smooth Hamiltonian structure on and a -plug whose image is contained in . By Lemma 7.3, the Hamiltonian structure is exact. The disjoint union of and is an -plug and we have . Moreover, we have . We can therefore compute
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 be an arbitrary extension of abelian groups and let be an -valued extension of helicity to Hamiltonian structures. Assume that satisfies the Extension and Invariance properties in Theorem 1.5. Moreover, assume that satisfies plug homogeneity. Our task is to show that there exists a unique group homomorphism over such that .
Let be a non-empty connected open surface with area form. Let us begin by defining a map
as follows. Let be a Hamiltonian homeomorphism. Pick an arbitrary isotopy in connecting the identity to . Moreover, pick an arbitrary Hamiltonian structure such that there exists an embedding of Hamiltonian structures . Define the -plug and set
By plug homogeneity, this is well-defined and independent of choices.
We claim that is a group homomorphism. Indeed, consider two Hamiltonian homeomorphisms . Pick Hamiltonian isotopies and starting at the identity and ending at and , respectively. We define the concatenation by
Now pick a Hamiltonian structure and an embedding of Hamiltonian structures . Define the plug . By construction, it is possible to split the plug into two disjoint plugs and inserting the isotopies and , respectively, such that . We can then compute
Here we use plug homogeneity and the definition of . This shows that is a homomorphism.
Next, we claim that the restriction of to the group of Hamiltonian diffeomorphisms agrees with the Calabi homomorphism via the inclusion . Suppose that . Then we can choose the Hamiltonian isotopy, the Hamiltonian structure, and the plug embedding involved in the definition of to be smooth. Let and be the resulting smooth Hamiltonian structure and plug, respectively. We compute
Here the second equality follows from the assumption that satisfies the Extension property in Theorem 1.5. The third equality follow from Lemma 4.10.
Suppose that is an area- and orientation-preserving embedding of open surfaces. Then we have
(7.3) |
where is the homomorphism between Hamiltonian homeomorphism groups obtained by pushforward via ; see Section 3. In order to see this, consider a Hamiltonian structure and an embedding . Given , we can form a plug by inserting the homeomorphism via the embedding . We can form a second plug by inserting the homeomorphism via the embedding . Note that the Hamiltonian structures obtained by inserting and agree, i.e. . Identity (7.3) is then immediate from the definition of .
Combining the above observations, we see that the homomorphisms descend to a well-defined group homomorphism
over which is independent of . We need to check that .
First, observe that since is a homomorphism over , and both and satisfy the Extension property in Theorem 1.5, we have
for every smooth Hamiltonian structure .
It is immediate from the definition of that satisfies the following version of the Calabi property in Theorem 1.5: for every Hamiltonian structure and every -plug , we have
Now consider an arbitrary Hamiltonian structure . By Theorem 6.1, we can write for some smooth Hamiltonian structure and a plug . We compute
Here the second equality uses the Calabi property for , the third equality uses the Extension property and the fact that is a homomorphism over , and the fourth equality uses the Calabi property of .
This concludes the proof of the existence part of Proposition 1.6. For the uniqueness part, simply observe that every element of is attained as the universal -valued helicity of some Hamiltonian structure .∎
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 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 -manifold . Suppose is a volume form on that is compatible with its orientation. Let denote a Hamiltonian structure, i.e., a closed maximally nondegenerate -form. Finally, let be a smooth, fixed-point-free flow on , generated by a nowhere vanishing vector field .
Consider equipped with coordinates . Let be the standard volume form and equip with the orientation induced by this volume form. Let be the flow generated by the vector field . Recall that the standard Hamiltonian structure is given by the -form .
Lemma 8.1.
The following statements are equivalent:
-
1.
We have .
-
2.
The triple is locally diffeomorphic to the triple .
Proof.
Note that and, moreover, this condition is preserved by diffeomorphisms and can be checked locally. Hence statement 2 implies statement 1.
Conversely, suppose that . Given an arbitrary point , we may pick sufficiently small and an embedding such that and the pull back agrees with the the standard area form on . For sufficiently small, extend to an embedding via the flow . It can be checked that the pull back of via is given by . ∎
Definition 8.2.
We call a compatible triple if the equivalent conditions in Lemma 8.1 are satisfied.
Note that given , there exists a unique vector field satisfying and we can extend to a compatible triple by taking to be the flow generated by . Below we formulate criteria for extension of or to compatible triples.
Proposition 8.3.
The pair extends to a compatible triple if and only if it is locally diffeomorphic to . The analogous statements hold for the pairs and .
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 clearly extend to a compatible triple. This shows that being locally diffeomorphic to a pair contained in the triple 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 uniquely determine the third.
Proof.
This is a consequence of the following fact: Let be a -dimensional vector space and suppose , , and are non-zero and satisfy the identity . Then, any two components of the triple uniquely determine the third. ∎
Proposition 8.5.
-
1.
The pair extends to a compatible triple if and only if the flow preserves .
-
2.
The pair extends to a compatible triple if and only if the flow is positively tangent to the characteristic foliation of .
Proof.
If is a compatible triple, then preserves and the flow is positively tangent to the characteristic foliation of . This shows the “only if” direction in both statements. Conversely, if preserves , then is a maximally nondegenerate closed -form and we can extend to a compatible triple by defining the Hamiltonian structure to consist of the -form .
Finally, if the flow is positively tangent to the characteristic foliation of , then there exists a unique volume form satisfying . ∎
8.2 The case
The goal of this section is to explain the relationship between Hamiltonian structures and volume-preserving flows in the setting.
Topological flows and foliations
Let be a closed oriented topological -manifold, for now not equipped with a measure. Consider a topological flow on , i.e. a continuous map such that is a homeomorphism of for each and, moreover,
It will be convenient to use the notation . In the following, we always assume that is fixed-point-free, i.e. that there does not exist a point such that for all .
Proposition 8.6.
The flow lines of a fixed-point-free topological flow on form an oriented -dimensional foliation . Conversely, for every oriented foliation on , there exists a fixed-point-free topological flow such that
Proof.
We begin by proving that the flow lines of a fixed-point-free flow on form an oriented -dimensional foliation . This actually makes important use of the fact that has dimension three; see Remark 8.8 below. Let be an arbitrary point. It was shown by Whitney [55] (see also [8, Lemma 1 and Corollary 1]) that admits a local cross section through which is a -dimensional topological disc. This means that there exist and a topologically embedded -dimensional disc containing such that
is an embedding. By construction, this embedding maps straight lines into flow lines of . Its inverse is therefore a topological foliation chart. Since was arbitrary, we can cover by such foliation charts.
Next, we show that every oriented foliation admits a fixed-point-free flow such that . Our strategy is to construct a special metric on with the property that the leaves of the characteristic foliation of are rectifiable curves with respect to the metric . We then use the metric to obtain, for each point , a unique continuous curve , where denotes the characteristic leaf containing , with the following properties:
-
•
parametrized by arc length,
-
•
,
-
•
is an orientation-preserving homeomorphism if is an open leaf and an orientation-preserving covering if is a closed leaf.
In this situation, we can define the flow by simply setting .
We now describe the construction of the metric . For , we consider foliation charts of the form
where denotes the disc of radius . The oriented leaves of the foliation within this chart are given by
for . For some , we can pick a finite collection of charts
such that the sets form an open covering of . Moreover, fix a smooth map
with the property that there exist an open set and a point such that the restriction of to is a diffeomorphism and . For , we define
Note that is continuous and that its restriction to is an embedding. For , we define the pseudo-metric
where denotes the standard norm on . Suppose that are two distinct points. There exists such that . For this , we must have which shows that
defines a metric on .
Next, we show that every leaf of the characteristic foliation of is rectifiable with respect to the metric . We recall the necessary preliminaries. A continuous curve
is called rectifiable if its length
(8.1) |
is finite, where the supremum is taken over all partitions of the interval . We say is rectifiable if the restriction of to any closed interval is rectifiable. If is rectifiable, then any reparametrization of it is rectifiable and has the same length.
Now, consider a leaf of the foliation and pick a continuous map which is an orientation-preserving homeomorphism if is open and an orientation-preserving covering if is closed. The claim below shows that is a rectifiable curve.
Claim 8.7.
For each compact interval , we have . Moreover, we have
(8.2) |
Proof.
For each , we can write
where is a countable collection of pairwise disjoint finite open intervals . Every compact subset of intersects only finitely many of the intervals . Given a point , let be a parametrization of the line segment . Since is a smooth map, we have , where stands for length measured with respect to the standard metric on . From this we can conclude that , where denotes length with respect to the pseudo-metric , defined analogously to (8.1) . Observe that if is an interval disjoint from , then . Using the fact that intersects only finitely many of the intervals , we conclude that .
Since the restriction of to is an embedding, there exists a positive number such that for all . If is an interval such that , we then have . Since the sets cover , every is contained in some interval 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 is a point and is the leaf containing , there exists a unique curve which is an orientation-preserving homeomorphism, or an orientation-preserving covering if is closed, such that is parametrized by arc length and . As already mentioned, we can then define . It remains to check that is continuous. A priori, we must also show that is a homeomorphism for each , however, this follows immediately from continuity as is the inverse of .
Fix a point and an arbitrary . Our goal is to show that if is close to , then is close to . First note that if is sufficiently close to and denotes the leaf containing , we can find a continuous map which is an orientation-preserving covering of such that and such that is close to . It then suffices to show that is close to being parametrized by arc length, i.e. that for every compact interval , the length is close to . We provide an outline of an argument proving this: For , let be an embedded continuous line segment with image contained in for . If and are close, then and are close. Note that for every segment of traversing there is a corresponding -close-by segment of traversing . This implies that is close to for all and for every interval . Thus and are close as well. This concludes the proof of continuity of . ∎
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 define a foliation. However, this dimensional restriction is not required for the second part of the proposition: in higher-dimensional manifolds, every oriented 1-dimensional foliation still admits a fixed-point-free flow tracing out its leaves.
We now describe an example of a fixed-point-free topological flow on a topological -manifold which does not admit a local cross section homeomorphic to a -dimensional ball everywhere. Start with a -dimensional homology sphere which is not homeomorphic to the -sphere . Consider the suspension , where are collapsed to points . While this space is a homology manifold, it is not a topological manifold. The two points do not have neighborhoods homeomorphic to the -dimensional ball. However, by the double suspension theorem [7], the double suspension is homeomorphic to . This also implies that the product is a topological -manifold. Let be the fixed-point-free flow on which rotates the factor and restricts to the identity on the factor. Now observe that cannot have a local cross section homeomorphic to at any point contained in . If it did, one would obtain a neighborhood of in homeomorphic to .
The Oxtoby–Ulam theorem
Let be a compact metric space and let be a Borel measure on of finite mass, i.e. a measure defined on the Borel -algebra of satisfying . The support of is the set of all points such that every open neighborhood of has strictly positive measure. The measure is said to have full support if . 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 is non-atomic if for every point .
Remark 8.9.
An atom of a general measure space is often defined to be a measurable set such that and such that every measurable set satisfies or . A measure is then called non-atomic if it does not have any atoms. A simple argument (see [31, 2.IV]) shows that if is an inner and outer regular finite Borel measure, then 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 for all points . 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 be a compact connected topological manifold, possibly with boundary. For , let be a finite, non-atomic Borel measure on of full support. Moreover, assume that . Then, there exists a homeomorphism such that if and only if . If this is the case, the homeomorphism can be chosen such that .
Volume-preserving topological flows and measured foliations
As above, let be a fixed-point-free topological flow on a closed oriented topological -manifold . Let be the foliation induced by ; see Proposition 8.6.
Lemma 8.11.
Suppose that is a finite Borel measure on preserved by . Then and induce a transverse measure on characterized by the condition that, for every transversal , every Borel subset , and every sufficiently small, we have
(8.3) |
Proof.
Pick such that the restriction of to is an embedding. Since is preserved by , the expression on the right hand side of (8.3) is independent of . Therefore, (8.3) yields a well-defined finite Borel measure on . We need to show that maps between subsets of transversals obtained by sliding along leaves of are measure preserving. This can be reduced to showing that if is a transversal, is a Borel subset, is a continuous function with small norm, and is small, then
This identity follows from the assumption that preserves . ∎
Let us say that a transverse measure on some foliation has full support if the induced measure on any transversal has full support. Similarly, we say that is non-atomic if the induced measure on any transversal is non-atomic.
Now suppose that is a fixed-point-free topological flow preserving a finite Borel measure . Let be the induced measured foliation; see Lemma 8.11. It is straightforward to see that has full support if and only if the induced transverse measure has full support. Since has finite mass and does not have fixed points, the measure is automatically non-atomic. However, it is possible that the transverse measure has atoms.
Example 8.12.
Consider with the Hopf flow , i.e. the flow tracing out the fibers of the Hopf fibration at unit speed. Let be the standard -dimensional Lebesgue measure on , scaled to have total volume . The Hopf flow preserves . Moreover, pick a Hopf fiber and let be the -dimensional Lebesgue measure supported on this fiber , again scaled to have total volume . Then is a finite Borel measure on invariant under . The induced transverse measure on the Hopf fibration is not non-atomic: any intersection point of a transversal with the special fiber yields an atom.
In this example, the flow is smooth, but the measure is not. We can turn the situation around using the Oxtoby-Ulam theorem: Since both and have full support, are non-atomic, and have total volume , there exists a homeomorphism of such that . Let be the pushforward of under . Then is a fixed-point-free topological flow on which preserves the standard Lebesgue measure . However, has a periodic orbit of positive measure, causing the induced transverse measure on to be not non-atomic.
Note that we can produce even more pathological examples. Let be a dense sequence of Hopf fibers. For each , let denote the -dimensional Lebesgue measure on of total volume . Define the -invariant measure . This measure has full support, is non-atomic, and has total volume . Again by Oxtoby-Ulam, the measure is homeomorphic to the standard -dimensional Lebesgue measure , i.e. there exists a homeomorphism of Y such that . Then is a fixed-point-free topological flow on which preserves and has the property that the complement of some countable sequence of periodic orbits has measure zero.
Lemma 8.13.
Suppose that is a finite Borel measure preserved by a fixed-point-free topological flow . Then the following statements are equivalent:
-
1.
is non-atomic.
-
2.
All flow lines of have vanishing measure.
-
3.
All periodic orbits of have vanishing measure.
Proof.
If is a transversal and is a point, then if and only if the flow line of through has positive measure. This shows the equivalence of the first two statements.
We argue that any flow line of of positive measure is necessarily periodic. Indeed, if is a non-periodic flow line through a point , then is the disjoint union of the countable collection of flow line segments for . All of these flow line segments have the same volume. Since is finite, the total volume of must be finite as well. But this implies that all flow line segments and therefore itself have vanishing volume. ∎
Compatible triples
Consider a closed oriented topological -manifold . Fix a finite Borel measure and a Hamiltonian structure on . Let denote a fixed-point-free topological flow on .
The Hamiltonian structure determines a -dimensional foliation on equipped with a transverse measure . Moreover, determines a coorientation of , which in combination with the orientation on determines an orientation .
Lemma 8.14.
The following statements are equivalent:
-
1.
The flow preserves and coincides with , as measured, cooriented (hence, oriented) foliations.
-
2.
The triple is locally homeomorphic to the triple
Proof.
Note that the triple satisfies all conditions in statement 1. Since all of these conditions can be checked locally, statement 2 implies statement 1. Conversely, suppose that satisfies statement 1. Every point in has a neighborhood which can be parametrized via an orientation-preserving topological embedding of Hamiltonian structures
Note that maps line segments into flow lines of respecting orientation. For sufficiently small, define the embedding
This embedding continues to pull back to and in addition pulls back to . We conclude that the transverse measure agrees with . Thus agrees with on all sets of the form where is an interval and is a Borel set. This determines and we conclude that is a local homeomorphism between and . ∎
Definition 8.15.
We call a compatible triple if the equivalent conditions in Lemma 8.14 are satisfied.
We now formulate criteria guaranteeing that any of the pairs , or extend to compatible triples.
Proposition 8.16.
The pair extends to a compatible triple if and only if it is locally homeomorphic to . The analogous statements hold for the pairs and .
Proof.
Lemma 8.17.
Any two components of a compatible triple uniquely determine the third.
Proof.
It suffices to check the above uniqueness locally. This means we need to verify that if is a compatible triple on (an open subset of) two of whose components coincide with the corresponding components of , then . In view of the characterization of compatibility given by statement 2, it suffices to check that if is a homeomorphism of preserving two members of the triple , then it preserves the third. This is elementary and we omit the details. ∎
Proposition 8.18.
-
1.
The pair extends to a compatible triple if and only if the flow of preserves and the flow lines of have zero measure.
-
2.
The pair extends to a compatible triple if and only if the flow of preserves and the induced transverse measure on is of full support and non-atomic.
-
3.
The pair extends to a compatible triple if and only if the oriented flow lines of agree with the oriented characteristic leaves of .
Proof.
The first and the second items are equivalent by Lemma 8.13. We will prove the second item.
If the pair extends to a compatible triple, then is locally homeomorphic to (the measured foliation of) , which clearly implies that is of full support and non-atomic. Conversely, suppose that preserves and is of full support and non-atomic. Given an arbitrary point in , pick a local cross section through that point which is homeomorphic to a disc; as mentioned earlier, the existence of the cross section was proven by Whitney [55]. Since the measure on induced by is of full support and non-atomic, it follows from the Oxtoby-Ulam theorem that there exists an area-preserving parametrization . In addition, we can pick such that it is orientation preserving with respect to the orientation of induced by the flow and the orientation of . Extend to an embedding via the flow . Clearly, pulls back to . Moreover, it pulls back to . Hence equipped with the orientation and coorientation induced by and the ambient orientation of is a Hamiltonian structure extending to a compatible triple. We have proven the second statement.
We turn now to the third statement. If extends to a compatible triple, then clearly the oriented flow lines of agree with the oriented characteristic leaves of by the characterization of compatibility given in statement 1 in Lemma 8.14. Conversely, suppose that the oriented flow lines of agree with the oriented characteristic leaves of . As we saw in the proof of Lemma 8.14, we can parametrize a neighborhood of any point in via an embedding of Hamiltonian structures that also pulls back the flow to . By Proposition 8.16, this means that extends to a compatible triple. ∎
8.3 Mass flow and flux
In this section, we briefly review the mass flow homomorphism and show that for a volume-preserving flow which is a component of a compatible triple , the flux admits a description in terms of mass flow.
Consider a closed and oriented topological manifold of dimension . Suppose that is a finite non-atomic Borel measure on of full support. Recall that there is a mass flow homomorphism, introduced in [18, 48],
where denotes the universal cover of the identity component of the group of volume-preserving homeomorphisms of and the identification is given by Poincaré duality.
We briefly recall the definition of . Let denote the circle. Then we have an isomorphism and therefore . Given an element represented by a volume-preserving isotopy starting at the identity, defining therefore amounts to defining a group homomorphism
Let be a homotopy class represented by a map and consider the homotopy of maps , which at coincides with the constant map zero. There exists a unique lift to a homotopy of maps , which at coincides with the constant map zero. Then,
(8.4) |
which turns out to be independent of the choice of representative of and . This defines , which we may view as taking values in , via Poincaré duality.
Definition 8.20.
Let be a flow. We say is exact if the element of determined by has vanishing mass flow.
We now restrict our attention to the case where is a closed, oriented 3-manifold, which we will refer to from this point onward as .
Proposition 8.21.
Suppose that is a compatible triple. Then, for every , we have
(8.5) |
Proof.
We begin by proving the identity (8.5) for smooth compatible triples . Let be a smooth map. Throughout this proof, we will view as mapping into which we identify with . Viewing the right hand side of identity (8.5) as a homomorphism , we compute
where 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 preserves , which implies that is independent of . The fifth equality holds because being a -form necessarily vanishes and hence we have
The sixth equality uses that is a compatible triple. In the final equality, is the de Rham cohomology class represented by the -form . This class agrees with the image of the cohomology class represented by the circle map under the natural map . Using the identifications
this shows identity (8.5) in the smooth case.
In the general case, where 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.
a smooth compatible triple on a smooth, closed and oriented -manifold ,
-
2.
an open surface with area form consisting of finitely many disc components, and a smooth embedding satisfying for all ,
-
3.
an isotopy in ,
such that the compatible triple obtained from by inserting888Here, by inserting the isotopy into we mean modifying the compatible triple 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 via the embedding is homeomorphic to .
Given an arbitrary homotopy class in , we may represent it by a smooth map with the property that . This makes use of the fact consists of finitely many disc components. For sufficiently small, we then have
This implies that
for all sufficiently small. Because mass flow is a homomorphism, we conclude that this identity holds for all . Since is a smooth compatible triple, we have
by the smooth case treated above. It follows from Definition 7.2, which defines , that
Combining the above identities, we conclude that
Since is homeomorphic to , this implies that
∎
8.4 Proof of Theorem 1.3
Let be a closed -manifold equipped with a volume form . We will assume throughout this section that is equipped with the orientation induced by .
We denote by the set consisting of all volume-preserving topological flows which are fixed-point-free and whose flow lines have vanishing measure. Let be the subset consisting of exact flows (see Definition 8.20); in other words, consists of flows satisfying the assumptions of Theorem 1.3. Similarly, let denote the set of all Hamiltonian structures on , and let denote the subset of exact Hamiltonian structures (see Definition 7.2).
Take any . Then, by Proposition 8.18 and Lemma 8.14, there exists a unique Hamiltonian structure such that the triple is a compatible triple in the sense of Definition 8.15. This yields a mapping from to , which preserves mass flow/flux by Proposition 8.21. Hence it induces a mapping from to
The above yields a -valued extension of helicity, which we continue to denote by : for , we simply define
where the right-hand-side in the above is the universal -valued extension of helicity given by Theorem 1.5. The three properties listed in Theorem 1.5 translate as follows for this -valued extension of helicity .
- 1.
-
2.
Conjugation invariance: we have
for any orientation- and volume-preserving homeomorphism .
-
3.
Calabi Compatibility: For every plug , we have
where 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, is uniquely determined by the above properties. Indeed, by Claim 8.22 the flow is topologically conjugate to a flow of the form where 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 -valued extension of helicity to , satisfying the above three properties, must coincide with .
Now, to obtain a real-valued extension of helicity, as stated in Theorem 1.3, simply pick to be a choice of projection. Then, the composition is an -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 . 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 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 -manifolds. V. The triangulation theorem and Hauptvermutung. Ann. of Math. (2), 56:96–114, 1952.
- [40] E. E. Moise. Geometric topology in dimensions and , 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 -fragmentation property for hamiltonian diffeomorphisms and homeomorphisms on surfaces. arXiv:2403.15767, 2024.
- [50] S. Seyfaddini. -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 -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