Abstract

An 𝑓-structure on a manifold 𝑀 is an endomorphism field 𝜑 satisfying 𝜑3+𝜑=0. We call an f-structure regular if the distribution 𝑇=ker𝜑 is involutive and regular, in the sense of Palais. We show that when a regular f-structure on a compact manifold M is an almost 𝒮-structure, it determines a torus fibration of M over a symplectic manifold. When rank 𝑇=1, this result reduces to the Boothby-Wang theorem. Unlike similar results for manifolds with 𝒮-structure or 𝒦-structure, we do not assume that the f-structure is normal. We also show that given an almost 𝒮-structure, we obtain an associated Jacobi structure, as well as a notion of symplectization.

1. Introduction

Let (𝑀,𝜂) be a cooriented contact manifold. The Boothby-Wang theorem [1] tells us that if the Reeb field 𝜉 corresponding to the contact form 𝜂 is regular (in the sense of Palais [2]), then 𝑀 is a prequantum circle bundle 𝜋𝑀𝑁 over a symplectic manifold (𝑁,𝜔), where 𝜋𝜔=𝑑𝜂 and 𝜂 may be identified with the connection 1-form. Conversely, let 𝑀 be a prequantum circle bundle over a symplectic manifold (𝑁,𝜔), and let 𝜂 be a connection 1-form. Given a choice of compatible almost complex structure 𝐽 for 𝜔, let 𝐺(𝑋,𝑌)=𝜔(𝐽𝑋,𝑌) be the associated Riemannian metric on 𝑁, and let 𝜋 denote the horizontal lift of vector fields defined by 𝜂. We can then define an endomorphism field 𝜑Γ(𝑀,End(𝑇𝑀)) by𝜑𝑋=𝜋𝐽𝜋𝑋,(1.1)

and a Riemannian metric 𝑔 by 𝑔=𝜋𝐺+𝜂𝜂. If we let 𝜉 be the vertical vector field satisfying 𝜂(𝜉)=1, then (𝜑,𝜉,𝜂,𝑔) defines a contact metric structure on 𝑀 [3]. In particular, we note that 𝜑 is an 𝑓-structure on 𝑀. By construction, we have 𝜑2=Id𝑇𝑀+𝜂𝜉, from which it follows that 𝜑3+𝜑=0.

In [4, 5], Blair et al. consider compact Riemannian manifolds equipped with a regular normal 𝑓-structure 𝜑 and show that such manifolds are the total space of a principal torus bundle over a complex manifold 𝑁, and that in addition, 𝑁 is a Kähler manifold if the fundamental 2-form of the 𝑓-structure is closed (i.e., if 𝑀 is a 𝒦-manifold). Saenz argued in [6] that if this 𝒦-structure is an 𝒮-structure, then the symplectic form of the Kähler manifold 𝑁 is integral.

While the results in [5, 6] provide us with a generalization of the Boothby-Wang theorem, the proofs in [5] (and by extension, the argument in [6]) rely in several places on the assumption that the 𝑓-structure 𝜑 is normal. Since this assumption is not required in the original Boothby-Wang theorem, it is natural to ask what can be said if this assumption is dropped for 𝑓-structures of higher corank. In this note, we use a theorem of Tanno [7] to show that if 𝑀 is a compact almost 𝒮-manifold, in the sense of [8], then 𝑀 is a principal torus bundle over a symplectic manifold whose symplectic form is integral. (More precisely, the symplectic form will be a real multiple of an integral symplectic form.) Not surprisingly, this tells us that requiring 𝜑 to be normal is the same as demanding that the base of our torus bundle be Kähler.

This “generalized Boothby-Wang theorem’’ is one of a number of similarities between manifolds with almost 𝒮-structure and contact manifolds. In the final section of this paper we demonstrate two more. First, there is a natural notion of symplectization: given an almost 𝒮-manifold 𝑀, there is an open, conic, symplectic submanifold of 𝑇𝑀 whose base is 𝑀. Second, a choice of one-form (expressed in terms of the almost 𝒮-structure) allows us to define a Jacobi bracket on the algebra of smooth functions on 𝑀, giving us in particular a notion of Hamiltonian vector field on manifolds with almost 𝒮-structure.

2. Preliminaries

2.1. Regular Involutive Distributions

Let 𝐹𝑇𝑀 be an involutive distribution of rank 𝑘. We briefly recall the notion of a regular distribution in the sense of Palais and refer the reader to [2] for the details. Roughly speaking, the involutive distribution 𝐹 is regular if each point 𝑝𝑀 has a coordinate neighbourhood (𝑈,𝑥1,𝑥𝑛) such that 𝜕𝜕𝑥1𝑝𝜕,,𝜕𝑥𝑘𝑝(2.1)

forms a basis for 𝐹𝑝𝑇𝑝𝑀, and such that the integral submanifold of 𝐹 through 𝑝 intersects 𝑈 in only one 𝑘-dimensional slice. When 𝐹 is regular, the leaf space =𝑀/𝐹 is a smooth Hausdorff manifold, and the quotient mapping 𝜋𝐹𝑀 is smooth and closed. When 𝑀 is compact and connected, the leaves of 𝐹 are compact and isomorphic and are the fibres of the smooth fibration 𝜋𝐹𝑀.

In particular, a vector field 𝑋 on 𝑀 is regular if each 𝑝𝑀 has a neighbourhood 𝑈 through which the integral curve of 𝑋 through 𝑝 passes only once. If 𝑀 is compact, the integral curves of a regular vector field are thus diffeomorphic to circles. Applying this fact to the Reeb vector field of a contact manifold gives part of the proof of the Boothby-Wang theorem.

2.2. 𝑓-Structures

An 𝑓-structure on 𝑀 is an endomorphism field 𝜑Γ(𝑀,End𝑇𝑀) such that 𝜑3+𝜑=0.(2.2) Such structures were introduced by Yano in [9]; many of the facts regarding 𝑓-structures are collected in the book [10]. By a result of Stong [11], every 𝑓-structure is of constant rank. If rank𝜑=dim𝑀, then 𝜑 is an almost complex structure on 𝑀, while if rank𝜑=dim𝑀1, then 𝜑 determines an almost contact structure on 𝑀.

It is easy to check that the operators 𝑙=𝜑2 and 𝑚=𝜑2+Id𝑇𝑀 are complementary projection operators; letting 𝐸=𝑙(𝑇𝑀)=im𝜑 and 𝑇=𝑚(𝑇𝑀)=ker𝜑, we obtain the splitting𝑇𝑀=𝐸𝑇=im𝜑ker𝜑(2.3) of the tangent bundle. Since (𝜑|𝐸)2=Id𝐸, 𝜑 is necessarily of even rank. When the corank of 𝜑 is equal to one, the distribution 𝑇 is automatically trivial and involutive. However, if rank𝑇>1, this need not be the case, and one often makes additional simplifying assumptions about 𝑇. An 𝑓-structure such that 𝑇 is trivial is called an 𝑓-structure with parallelizable kernel (or 𝑓pk-structure for short) in [8]. We will assume that an 𝑓·pk-structure includes a choice of a trivializing frame {𝜉𝑖} and corresponding coframe {𝜂𝑖} for 𝑇, with𝜂𝑖𝜉𝑗=𝛿𝑖𝑗𝜉,𝜑𝑖=𝜂𝑗𝜑=0,𝜑2=Id+𝜂𝑖𝜉𝑖.(2.4)

(This is known as an 𝑓-structure with complemented frames in [4]; such a choice of frame and coframe always exists.) Given an 𝑓·pk-structure, it is always possible [10] to find a Riemannian metric 𝑔 that is compatible with (𝜑,𝜉𝑖,𝜂𝑗) in the sense that, for all 𝑋,𝑌Γ(𝑀,𝑇𝑀), we have𝑔(𝑋,𝑌)=𝑔(𝜑𝑋,𝜑𝑌)+𝑘𝑖=1𝜂𝑖(𝑋)𝜂𝑖(𝑌).(2.5) Following [8], we will call the 4-tuple (𝜑,𝜉𝑖,𝜂𝑗,𝑔) a metric 𝑓·pk structure. Given a metric 𝑓·pk-structure (𝜑,𝜉𝑖,𝜂𝑗,𝑔), we can define the fundamental 2-form Φ𝑔𝒜2(𝑀) by Φ𝑔(𝑋,𝑌)=𝑔(𝜑𝑋,𝑌).(2.6)

Remark 2.1. Our definition of Φ𝑔 is chosen to agree with our preferred sign conventions in symplectic geometry; however, many authors place 𝜑 in the second slot, so our convention here uses the opposite sign of that found for example in [5, 8].

We will call an 𝑓-structure 𝜑  regular if the distribution 𝑇=ker𝜑 is regular in the sense of Palais [2]. An 𝑓pk-structure is regular if the vector fields 𝜉𝑖 are regular and independent. An 𝑓·pk-structure is called normal [4] if the tensor 𝑁 defined by []+𝑁=𝜑,𝜑𝑘𝑖=1𝑑𝜂𝑖𝜉𝑖(2.7) vanishes identically. Here [𝜑,𝜑] denotes the Nijenhuis torsion of 𝜑, which is given by[]𝜑,𝜑(𝑋,𝑌)=𝜑2[]+[][][].𝑋,𝑌𝜑𝑋,𝜑𝑌𝜑𝜑𝑋,𝑌𝜑𝑋,𝜑𝑌(2.8)

When 𝜑 is normal, the +𝑖-eigenbundle of 𝜑 (extended by linearity to 𝑇𝑀) defines a CR structure 𝐸1,0𝑇𝑀. Regular normal 𝑓-structures are studied in [5], where it is proved that a compact manifold with regular normal 𝑓-structure is a principal torus bundle over a complex manifold 𝑁. If the fundamental 2-form Φ𝑔 of a normal 𝑓-structure is closed, then the 𝑓-structure is called a 𝒦-structure, and 𝑀 a 𝒦-manifold. For a compact regular 𝒦-manifold 𝑀, the base 𝑁 of the torus fibration is a Kähler manifold. A special case of a 𝒦-manifold is an 𝒮-manifold. On an 𝒮 manifold, there exist constants 𝛼1,,𝛼𝑘 such that 𝑑𝜂𝑖=𝛼𝑖Φ𝑔 for 𝑖=1,,𝑘. Two commonly considered cases are the case 𝛼𝑖=0 for all 𝑖, and the case 𝛼𝑖=1 for all 𝑖. In the language of CR geometry, the former case is analogous to a “Levi-flat’’ CR manifold,

while the latter defines an analogue of a strongly pseudoconvex CR manifold (typically, strongly pseudoconvex CR manifolds are assumed to be of “hypersurface type,’’ meaning that the complementary distribution 𝑇 has rank one; see [12]).

A refinement of the notion of 𝒮-structure was introduced in [8]: a metric 𝑓pk-structure (𝜑,𝜉𝑖,𝜂𝑗,𝑔) which is not necessarily normal is called an almost 𝒮-structure if 𝑑𝜂𝑖=Φ𝑔 for each 𝑖=1,,𝑘. An 𝑓-structure 𝜑 is called CR-integrable in [8] if the +𝑖-eigenbundle 𝐸1,0𝑇𝑀 of 𝜑 is involutive (and hence, defines a CR structure). It is shown in [8] that an 𝑓pk-structure is CR-integrable if and only if the tensor 𝑁 given by (2.7) satisfies 𝑁(𝑋,𝑌)=0 for all 𝑋,𝑌Γ(𝑀,𝐸), where 𝐸=im𝜑, whereas for a normal 𝑓pk-structure, 𝑁 must vanish for all 𝑋,𝑌Γ(𝑀,𝑇𝑀). In [13] it is proved that a CR-integrable almost 𝒮-manifold admits a canonical connection analogous to the Tanaka-Webster connection of a strongly pseudoconvex CR manifold. For the relationship between this connection and the 𝜕𝑏 operator of the corresponding tangential Cauchy-Riemann complex, as well as an application of this relationship to defining an analogue of geometric quantization for almost 𝒮-manifolds, see [14].

In this paper, we will define an almost 𝒦-structure to be a metric 𝑓pk-structure for which 𝑑Φ𝑔=0, and we will define an almost 𝒮-structure more generally to be an almost 𝒦-structure such that 𝑑𝜂𝑖=𝛼𝑖Φ𝑔 for constants 𝛼𝑖, for 𝑖=1,,𝑘.

3. Properties of Almost 𝒦 and Almost 𝒮-Structures

Let (𝜑,𝜉𝑖,𝜂𝑖) be an 𝑓pk-structure on a compact, connected manifold 𝑀. Let 𝑔 be a Riemannian metric satisfying the compatibility condition (2.5), and let Φ𝑔 denote the corresponding fundamental 2-form. Let 𝐸=im𝜑, and 𝑇=ker𝜑 denote the distribution spanned by the 𝜉𝑖. It is easy to check that the distributions 𝐸 and 𝑇 are orthogonal with respect to 𝑔, and that the restriction of Φ𝑔 to 𝐸𝐸 is nondegenerate, from which we have the following lemma.

Lemma 3.1. 𝑋Γ(𝑀,𝑇) if and only if 𝜄(𝑋)Φ𝑔=0.

Proposition 3.2. Let (𝜑,𝜉𝑖,𝜂𝑖,𝑔) be a metric 𝑓·pk-structure. Then 𝑇=ker𝜑 is involutive whenever 𝑑Φ𝑔=0.

Proof. Let 𝑋,𝑌Γ(𝑀,𝑇), and let 𝑍Γ(𝑀,𝑇𝑀). Then, using Lemma 3.1 above, we have 𝑑Φ𝑔(𝑋,𝑌,𝑍)=𝑋Φ𝑔(𝑌,𝑍)+𝑌Φ𝑔(𝑍,𝑋)+𝑍Φ𝑔(𝑋,𝑌)Φ𝑔([]𝑋,𝑌,𝑍)Φ𝑔([]𝑌,𝑍,𝑋)Φ𝑔([]𝑍,𝑋,𝑌)=Φ𝑔([]𝑋,𝑌,𝑍).(3.1) Therefore, if 𝑑Φ𝑔=0, then 𝜄([𝑋,𝑌])Φ𝑔=0, and thus [𝑋,𝑌]Γ(𝑀,𝑇), which proves the proposition.

Let us now suppose that (𝜑,𝜉𝑖,𝜂𝑖,𝑔) is an almost 𝒮-structure, so that the 1-forms 𝜂𝑖 satisfy 𝑑𝜂𝑖=𝛼𝑖Φ𝑔 for constants 𝛼𝑖, some of which may be zero. The following results were proved in [8] in the case that 𝛼𝑖=1 for all 𝑖; we easily see that the results remain true in our more general setting.

Proposition 3.3. If (𝜑,𝜉𝑖,𝜂𝑗,𝑔) is an almost 𝒮-structure, then (𝜉𝑖)𝜉𝑗=[𝜉𝑖,𝜉𝑗]=0 for all 𝑖,𝑗=1,,𝑘.

Proof. Since the fundamental 2-form Φ𝑔 of an almost 𝒮-structure is closed, the distribution 𝑇 is involutive. Thus we may write [𝜉𝑖,𝜉𝑗𝑐]=𝑎𝑖𝑗𝜉𝑎. But for any 𝑎,𝑖,𝑗{1,,𝑘}, we have 𝑐𝑎𝑖𝑗=𝜂𝑎𝜉𝑖,𝜉𝑗=𝜉𝑖𝜂𝑎𝜉𝑗𝜉𝑗𝜂𝑎𝜉𝑖𝑑𝜂𝑎𝜉𝑖,𝜉𝑗=𝛼𝑎Φ𝑔𝜉𝑖,𝜉𝑗=0.(3.2)

Proposition 3.4. If (𝜑,𝜉𝑖,𝜂𝑗,𝑔) is an almost 𝒮-structure, then (𝜉𝑖)𝜂𝑗=0 for all 𝑖,𝑗=𝑖,,𝑘.

Proof. We have (𝜉)𝜂𝑗=𝑑(𝜂𝑗(𝜉𝑖))+𝜄(𝜉𝑖)𝑑𝜂𝑗=𝛼𝑗(𝜄(𝜉𝑖)Φ𝑔)=0.

We remark that several other results from [8] hold in this more general setting, but they are not needed here. To conclude this section, we state a theorem due to Tanno [7].

Theorem 3.5. For a regular and proper vector field 𝑋 on a manifold 𝑀, the following are equivalent. (i)The period function 𝜆𝑋 of 𝑋 is constant.(ii)There exists a 1-form 𝜂 such that 𝜂(𝑋)=1 and (𝑋)𝜂=0.(iii)There exists a Riemannian metric 𝑔 such that 𝑔(𝑋,𝑋)=1 and (𝑋)𝑔=0.

In the above theorem, the period function 𝜆𝑋𝑀 is defined by𝜆𝑋(𝑝)=inf{𝑡>0exp(𝑡𝑋)𝑝=𝑝}.(3.3)

If 𝑀 is noncompact, the value 𝜆𝑋(𝑝)= is possible. Part (iii) of the above tells us that 𝑋 is a unit Killing field for the metric 𝑔. Using this result, Tanno was able to give a simple proof (which is reproduced in [3]) of the Boothby-Wang theorem [1].

4. The Structure of Regular Almost 𝒮-Manifolds

As noted above, from [5], a compact manifold with regular normal 𝑓-structure is a principal torus bundle over a complex manifold 𝑁, and 𝑁 is Kähler if 𝑀 is a 𝒦-manifold. If 𝑀 is an 𝒮-manifold with Φ𝑔=𝑑𝜂𝑖 for each 𝑖, then by [6], the symplectic form on 𝑁 is integral. We now dispense with the requirement that the 𝑓-structure on 𝑀 be normal, and state a similar result for almost 𝒮-manifolds.

Theorem 4.1. Let 𝑀 be a compact manifold of dimension 2𝑛+𝑘 equipped with a regular almost 𝒮-structure ̃𝜉(𝜑,𝑖,̃𝜂𝑖,̃𝑔) of rank  2𝑛. Then there exists an almost 𝒮-structure (𝜑,𝜉𝑖,𝜂𝑖,𝑔) on 𝑀 for which the vector fields 𝜉1,,𝜉𝑘 are the infinitesimal generators of a free and effective 𝕋𝑘-action on 𝑀. Moreover, the quotient 𝑁=𝑀/𝕋𝑘 is a smooth symplectic manifold of dimension 2𝑛, and if the 𝛼𝑖 such that 𝑑̃𝜂𝑖=𝛼𝑖Φ̃𝑔 are not all zero, then the symplectic form on 𝑁 is a real multiple of an integral symplectic form.

Proof. By assumption, the vector fields ̃𝜉1̃𝜉,,𝑘 are regular, independent, and proper, and by Proposition 3.2, the distribution ̃𝜉𝑇=span{1̃𝜉,,𝑘} is involutive. Thus, by the results of Palais, 𝑁=𝑀/𝑇 is a smooth manifold, and 𝜋𝑀𝑁 is a smooth fibration whose fibres are the leaves of the distribution 𝑇. Since 𝑀 is compact, the fibres are compact and isomorphic [2]. For each 𝑖=1,,𝑘, we have ̃𝜂𝑖(̃𝜉𝑖)=1 and ̃𝜉(𝑖)̃𝜂𝑖=0. Thus, by Theorem 3.5, the period functions 𝜆𝑖̃𝜉=𝜆𝑖 are constant. We rescale by setting 𝜉𝑖=𝜆𝑖̃𝜉𝑖 and 𝜂𝑖=(1/𝜆𝑖)̃𝜂𝑖. We still have 𝜂𝑖(𝜉𝑗)=𝛿𝑖𝑗, and note that the associated metric 𝑔 for which (𝜑,𝜉𝑖,𝜂𝑖,𝑔) is an almost 𝒮-structure differs from ̃𝑔 only along 𝑇, so that Φ𝑔=Φ̃𝑔. Each 𝜉𝑖 now has period 1, and since the vector fields 𝜉𝑖 all commute, they are the generators of a free and effective 𝕋𝑘-action on 𝑀. The argument for local triviality is the same as in [5], so we do not repeat it here. Thus, we have that 𝑀 is a principal 𝕋𝑘-bundle over 𝑁=𝑀/𝑇. The infinitesimal action of 𝑘 is given by 𝑡𝑋=1,,𝑡𝑘𝑋𝑀=𝑡𝑖𝜉𝑖,(4.1) from which we see that 𝜼=(𝜂1,,𝜂𝑘) is a connection 1-form on 𝑀: we have 𝜄(𝑋𝑀)𝜂=𝑋 and (𝑋𝑀)𝜂=0 for all 𝑋𝑘.
Now, we note that the fundamental 2-form Φ𝑔 is horizontal and invariant, since 𝜄(𝑋)Φ𝑔=(𝑋)Φ𝑔=0 for all 𝑋Γ(𝑀,𝑇), and thus there exists a 2-form Ω on 𝑁 such that 𝜋Ω=Φ𝑔. Since 𝜋𝑑Ω=𝑑Φ𝑔=0, Ω is closed, and since 𝜋Ω𝑛=Φ𝑛𝑔0, Ω is nondegenerate, and hence symplectic.
Finally, let us suppose that one of the 𝛼𝑖 is nonzero; without loss of generality, let us say 𝛼10. By the same argument as above, the vector fields 𝜉2,,𝜉𝑘 generate a free 𝕋𝑘1-action on 𝑀, giving us a fibration 𝑝𝑀𝑃. Now, since (𝜉𝑖)𝜉1=(𝜉𝑖)𝜂1=0 for 𝑖=2,,𝑘, the vector field 𝜉1 and 1-form 𝜂1 are invariant under the 𝕋𝑘1-action. We can thus define a 1-form 𝜂 on 𝑃 by 𝜂(𝑋)=𝜂1(̃𝑝𝑋), where ̃𝑝𝑋 denotes the horizontal lift of 𝑋 with respect to the connection 1-form defined by 𝜂2,,𝜂𝑘, and a vector field 𝜉 on 𝑃 by 𝜉=𝑝𝜉1. Note that 𝑑𝜂(𝑋,𝑌)=𝑑𝜂1(̃𝑝𝑋,̃𝑝𝑌). We then have 𝜂(𝜉)=1, and (𝜉)𝜂=𝜄(𝜉1)𝑑𝜂1=0, so that Theorem 3.5 applies to the pair (𝜂,𝜉). It follows that 𝜉 generates a free action of 𝑆1=/ on 𝑃, giving us the 𝕋1-bundle structure 𝑞𝑃𝑁. Since 𝜋=𝑞𝑝, it follows that𝑑𝜂(𝑋,𝑌)=𝑑𝜂1𝛼(̃𝑝𝑋,̃𝑝𝑌)=1𝜆1𝜋Ω𝛼(̃𝑝𝑋,̃𝑝𝑌)=1𝜆1𝑞Ω(𝑋,𝑌).(4.2) Thus, 𝑃 is a Boothby-Wang fibration over (𝑁,(𝛼1/𝜆1)Ω), from which it follows that the symplectic form (𝛼1/𝜆)Ω must be integral (see [15]), and hence Ω is a real multiple of an integral symplectic form.

Remark 4.2. Note that since the last part of the argument is valid for any pair of nonzero constants 𝛼𝑖,𝛼𝑗, from which it follows that for each 𝑖,𝑗 for which 𝛼𝑖 and 𝛼𝑗 are nonzero, we must have 𝛼𝑖/𝜆𝑖𝜆𝑗/𝛼𝑗.

Conversely, we have the following theorem.

Theorem 4.3. Suppose that 𝑀 is a principal 𝕋𝑘-bundle over a symplectic manifold (𝑁,𝜔), equipped with connection 1-form 𝜼=(𝜂1,,𝜂𝑘) such that there exist constants 𝛼1,,𝛼𝑘 for which 𝑑𝜂𝑖=𝛼𝑖𝜋𝜔. Then 𝑀 admits an almost 𝒮-structure.

Proof. The proof is essentially the same as the proof given in [4] when 𝑁 is Kähler, if we omit the proof of normality. Given a choice of compatible almost complex structure 𝐽 and associated metric 𝐺, we can define an 𝑓-structure 𝜑 by 𝜑𝑋=𝜋𝐽𝜋𝑋, where 𝜋 denotes the horizontal lift with respect to 𝜼. If we let 𝜉1,,𝜉𝑘 denote vertical vectors such that 𝜂𝑖(𝜉𝑗)=𝛿𝑖𝑗, and define the metric 𝑔 by 𝑔(𝑋,𝑌)=𝜋𝜂𝐺(𝑋,𝑌)+𝑖(𝑋)𝜂𝑖(𝑌),(4.3) then it is straightforward to check that the data (𝜑,𝜉𝑖,𝜂𝑗,𝑔) defines an almost 𝒮-structure on 𝑀. (Note that Φ𝑔=𝜋𝜔, so that 𝑑𝜂𝑖=𝛼𝑖Φ𝑔.)

Remark 4.4. We can also use the results of Tanno [7] to show that the vector fields 𝜉1,,𝜉𝑘 of an almost 𝒮-structure are Killing. Let 𝜋 denote the horizontal lift defined by 𝜼. Then we can define a Riemannian metric 𝐺 on 𝑁 by 𝐺(𝑋,𝑌)=𝑔(𝜋𝑋,𝜋𝑌) for any 𝑋,𝑌Γ(𝑁,𝑇𝑁), where 𝑔 is the metric of the almost 𝒮-structure on 𝑀. It follows that 𝑔=𝜋𝜂𝐺+𝑖𝜂𝑖, whence 𝑔(𝜉𝑖,𝜉𝑖)=1 and (𝜉𝑖)𝑔=0 for 𝑖=1,,𝑘. Moreover, the endomorphism field 𝐽Γ(𝑁,End(𝑇𝑁)) defined by 𝐽𝑋=𝜋𝜑𝜋𝑋 is easily seen to be an almost complex structure on 𝑁 that is compatible with 𝐺, and the symplectic form Ω then satisfies Ω(𝑋,𝑌)=𝐺(𝑋,𝐽𝑌).

Remark 4.5. If 𝑀 is only an almost 𝒦-manifold, it is not clear that we can expect any analogous result to hold, since the proof in [5] for a 𝒦-manifold does not work without normality, and Tanno’s theorem cannot be applied if (𝜉𝑖)𝜂𝑗0 for all 𝑖,𝑗, and this need not hold if 𝑑𝜂𝑗 is not a multiple of Φ𝑔.

Remark 4.6. If 𝑀 is noncompact, then as noted below the statement of Tanno’s theorem, the period 𝜆𝑖 of one of the 𝜉𝑖 could be infinite, in which case 𝜉𝑖 generates an -action on 𝑀 instead of an 𝑆1-action.

5. Symplectization and Jacobi Structures

We conclude this paper with a discussion of the relationship between almost 𝒮-structures and related geometries intended to reinforce the view that almost 𝒮-structures deserve to be viewed as higher corank analogues of contact structures. (However, see also [16] for the notion of 𝑘-contact structures, which, from the point of view of Heisenberg calculus, are also deserving of the title of higher corank contact structure. From this perspective, almost 𝒮-structures are perhaps more analogous to contact metric structures, or even strongly pseudoconvex CR structures, although they are not CR-integrable in general.)

Recall that a stable complex structure on a manifold 𝑀 is a complex structure defined on the fibres of 𝑇𝑀𝑘 for some 𝑘. Given an 𝑓pk-structure (𝜑,𝜉𝑖,𝜂𝑗) on 𝑀, we obtain a stable complex structure 𝐽Γ(𝑀,End(𝑇𝑀𝑘)) by setting 𝐽𝑋=𝜑𝑋 for 𝑋Γ(𝑀,𝐸), and defining 𝐽𝜉𝑖=𝜏𝑖 and 𝐽𝜏𝑖=𝜉𝑖, where 𝜏1,,𝜏𝑘 is a basis for 𝑘. As explained in [17], a stable complex structure determines a Spin𝑐-structure on 𝑀.

Alternatively, (and with some abuse of notation), we can think of the above complex structure on each fibre 𝑇𝑥𝑀×𝑘 as coming from an almost complex structure on 𝑀×𝑘 obtained from the 𝑓-structure 𝜑. With this point of view, we note that it is possible to define a “symplectization’’ analogous to the symplectization of a cooriented contact manifold, provided that our 𝑓pk-structure is an almost 𝒮-structure, with at least one of the 𝛼𝑗 (such that 𝑑𝜂𝑗=𝛼𝑗Φ𝑔) nonzero. As above, we let 𝑇𝑀=𝐸𝑇 denote the splitting of the tangent bundle determined by the 𝑓-structure, and let 𝐸0𝑇=span{𝜂𝑖}𝑀×𝑘 denote the annihilator of 𝐸. It is then possible to find an open connected symplectic submanifold 𝐸0+ of 𝑇𝑀 whose tangent bundle is 𝑇𝑥𝑀×𝑘. For concreteness, let us use the identification 𝐸0𝑀×𝑘, and with respect to coordinates (𝑥,𝑡1,,𝑡𝑘), let 𝛼=𝑘𝑖=1𝑡𝑖𝜂𝑖,(5.1)and define 𝜔=𝑑𝛼. (We are abusing notation here slightly; technically we should write 𝜋𝜂𝑖 in place of 𝜂𝑖, where 𝜋𝑀×𝑘𝑀 is the projection onto the first factor.) Using the fact that 𝑑𝜂𝑖=𝛼𝑖Φ𝑔 for each 𝑖, we have 𝜂𝜔=𝑗𝑑𝑡𝑗+𝑡𝑗𝛼𝑗Φ𝑔.(5.2)Define 𝜏𝐶(𝐸0) to be the function given in coordinates by 𝛼𝜏=𝑗𝑡𝑗. Note that since 𝜂𝑖𝜂𝑖=𝑑𝑡𝑖𝑑𝑡𝑖=0, we have 𝑘𝑖=1𝜂𝑗𝑑𝑡𝑗𝑘=𝑘!𝜂1𝑑𝑡1𝜂𝑘𝑑𝑡𝑘.(5.3)We also note that Φ𝑚𝑔=0 for 𝑚>𝑛. Thus, using the binomial theorem, we find that the top-degree form 𝜔𝑛+𝑘 has only one nonzero term; namely, 𝜔𝑛+𝑘=(𝑛+𝑘)!𝜂𝑛!1𝑑𝑡1𝜂𝑘𝑑𝑡𝑘𝜏Φ𝑔𝑛.(5.4)

Thus, 𝜔𝑛+𝑘 is a volume form on the open subset 𝐸0+ of 𝐸0 defined by 𝜏>0, and hence 𝜔 is a symplectic form on 𝐸0+.

Next, we will show that for certain choices of section 𝜂Γ(𝑀,𝐸0) we obtain a Jacobi structure on 𝑀 defined in a manner analogous to the Jacobi structure associated to a choice of contact form on a contact manifold. We recall that a Jacobi structure on 𝑀 is given by a Lie bracket {,} on 𝐶(𝑀) such that for any 𝑓,𝑔𝐶(𝑀) the support of {𝑓,𝑔} is contained in the intersection of the supports of 𝑓 and 𝑔. Jacobi structures were introduced independently by Kirillov [18] and Lichnerowicz [19]; a good introduction can be found in [20].

Again, we assume 𝑀 is equipped with an almost 𝒮-structure with the constants 𝛼𝑗 such that 𝑑𝜂𝑗=𝛼𝑗Φ𝑔 not all zero. Our first goal is to define a notion of a Hamiltonian vector field 𝑋𝑓 associated to each function 𝑓𝐶(𝑀). To begin with, let 𝑏𝜉=𝑗𝜉𝑗 be an arbitrary section of 𝑇=ker𝜑, and let 𝑐𝜂=𝑗𝜂𝑗 be an arbitrary section of 𝐸0𝑇. We will narrow down the possibilities for 𝜉 and 𝜂 as we consider the properties we wish the vector fields 𝑋𝑓 to satisfy. The idea is to generalize the approach used to define Hamiltonian vector fields on a contact manifold (𝑀,𝜂). Recall that on manifold equipped with a contact form 𝜂, where we define Φ=𝑑𝜂, the Reeb vector field 𝜉 is defined by 𝜄(𝜉)𝜂=1 and 𝜄(𝜉)Φ=0. A contact Hamiltonian vector field 𝑋𝑓 satisfies the equations 𝜄(𝑋𝑓)𝜂=𝑓 and 𝜄(𝑋𝑓)Φ=𝑑𝑓(𝜉𝑓)𝜂. Lichnerowicz showed in [21] that these are the necessary and sufficient conditions for each 𝑋𝑓 to be an infinitesimal symmetry of the contact structure: it follows that for each 𝑓𝐶(𝑀), (𝑋𝑓)𝜂=(𝜉𝑓)𝜂.

We wish to impose similar conditions on 𝜉, 𝜂 and (the yet to be defined) 𝑋𝑓 in the case of almost 𝒮-manifolds. We already know that 𝜄(𝜉)Φ𝑔=0, by Lemma 3.1, so we begin by adding the requirement that 𝑏𝜂(𝜉)=𝑗𝑐𝑗=1. Next, we give our definition of a Hamiltonian vector field.

Definition 5.1. Let 𝜂 and 𝜉 be as above. For any 𝑓𝐶(𝑀), we define the Hamiltonian vector field associated to 𝑓 by the equations 𝜄𝑋𝑓𝜂𝑗=𝛼𝑗𝑓,for𝑗=1,,𝑘,(5.5)𝜄𝑋𝑓Φ𝑔=𝑑𝑓(𝜉𝑓)𝜂.(5.6)

Remark 5.2. Note that the above equations uniquely define 𝑋𝑓, by the nondegeneracy of the restriction of Φ to 𝐸=im𝜑. The constants 𝛼𝑗 are the same ones such that 𝑑𝜂𝑗=𝛼𝑗Φ𝑔. One can check that if we began with 𝑎𝑗 in place of the 𝛼𝑗, we would be forced to take 𝑎𝑗=𝛼𝑗 for consistency reasons. (In particular, this will be necessary if the bracket we define below is to be a Lie bracket.) Moreover, this gives us the identity 𝑋𝑓𝜂𝑗=𝛼𝑗(𝜉𝑓)𝜂(5.7) for each 𝑗=1,,𝑘; we would otherwise have an unwanted term of the form (𝑎𝑗𝛼𝑗)𝑑𝑓. Note that on the right-hand side of the above equation we have 𝜂 and not 𝜂𝑗; this is unavoidable with our definition of 𝑋𝑓.

We can fix the coefficients of 𝜉 by requiring that 𝜉 be the Hamiltonian vector field associated to the constant function 1, as is standard for Jacobi structures (see [20]). It is easy to see that (5.5) then immediately forces us to take 𝛼𝜉=𝑗𝜉𝑗; that is, the coefficients 𝑏𝑗 are equal the constants 𝛼𝑗. Thus, 𝜉 is essentially determined by the almost 𝒮-structure, although 𝜂 is constrained only by the condition 𝜂(𝜉)=1, so the Jacobi structure we define below cannot be considered entirely canonical (as one might expect). From the requirement that 𝜂(𝜉)=1, it follows that for each 𝑓𝐶(𝑀), we have 𝑋𝑓𝑐𝜂=𝑗𝑋𝑓𝜂𝑗=𝑐𝑗𝛼𝑗(𝜉𝑓)𝜂=(𝜉𝑓)𝜂,(5.8)again in analogy with the contact case. Note that the normalization 𝜂(𝜉)=1 also implies that 𝑑𝜂=Φ𝑔. We are now ready to define our bracket on 𝐶(𝑀).

Definition 5.3. Let 𝑀 be a manifold with almost 𝒮-structure, with constants 𝛼𝑗 not all zero. Let 𝛼𝜉=𝑗𝜉𝑗, and let 𝜂 be a section of 𝐸0 such that 𝜂(𝜉)=1. We then define a bracket on 𝐶(𝑀) by 𝑋{𝑓,𝑔}=𝜄𝑓,𝑋𝑔𝜂.(5.9)

The bracket is clearly antisymmetric, and one checks (using the identity 𝜄([𝑋,𝑌])=[(𝑋),𝜄(𝑌)]) that{𝑓,𝑔}=𝑋𝑓𝑔𝑋𝑔𝑓+Φ𝑔𝑋𝑓,𝑋𝑔=𝑋𝑓𝑔(𝜉𝑓)𝑔.(5.10)

Note that since the definition of the Hamiltonian vector fields depended on the choice of 𝜂, the bracket depends on 𝜂, even though 𝜂 no longer appears explicitly in either of the above expressions for the bracket. From the latter equality we see that the support of {𝑓,𝑔} is contained in the support of 𝑔, and by antisymmetry it must be contained in the support of 𝑓 as well. Thus, the bracket given by (5.9) is a Jacobi bracket provided we can verify the Jacobi identity. Since the Jacobi identity is valid for the Lie bracket on vector fields, it suffices to prove the following proposition.

Proposition 5.4. Let {𝑓,𝑔} be the bracket on 𝐶(𝑀) given by (5.9). Then the vector field 𝑋{𝑓,𝑔} corresponding to the function {𝑓,𝑔} is given by 𝑋{𝑓,𝑔}=[𝑋𝑓,𝑋𝑔].

Lemma 5.5. For each 𝑖=1,,𝑘, we have [𝜉𝑖,𝑋𝑓]=𝑋𝜉𝑖𝑓.

Proof. From Propositions 3.3 and 3.4, we know that [𝜉𝑖,𝜉𝑗]=0 and (𝜉𝑖)𝜂𝑗=0 for any 𝑖,𝑗{1,,𝑘}; from the latter, it follows easily that (𝜉𝑖)Φ𝑔=0 as well. The result then follows from the uniqueness of Hamiltonian vector fields, since 𝜄𝜉𝑖,𝑋𝑓𝜂𝑗=𝜉𝑖𝑋,𝜄𝑓𝜂𝑗=𝛼𝑗𝜉𝑖𝜄𝜉𝑓,𝑖,𝑋𝑓Φ𝑔𝜉=𝑖𝜉(𝑑𝑓(𝜉𝑓)𝜂)=𝑑𝑖𝜉𝑓𝜉𝑖𝑓𝜂.(5.11)

Lemma 5.6. For each 𝑖=1,,𝑘, we have 𝜉𝑖{𝑓,𝑔}={𝜉𝑖𝑓,𝑔}+{𝑓,𝜉𝑖𝑔}.

Proof. We have, using Lemma 5.5 and the fact that [𝜉𝑖,𝜉]=0 in the second line, 𝜉𝑖{𝑓,𝑔}=𝜉𝑖𝑋𝑓𝑔𝜉𝑖((𝜉𝑓)𝑔)=𝑋𝑓𝜉𝑖𝜉𝑔(𝜉𝑓)𝑖𝑔+𝑋𝜉𝑖𝑓𝜉𝑔𝜉𝑖𝑔=𝑓𝑓,𝜉𝑖+𝜉𝑔𝑖.𝑓,𝑔(5.12)

Proof of Proposition 5.4. We need to show that 𝜄([𝑋𝑓,𝑋𝑔])𝜂𝑗=𝛼𝑗{𝑓,𝑔} for each 𝑗=1,,𝑘, and that 𝜄([𝑋𝑓,𝑋𝑔])Φ=𝑑{𝑓,𝑔}(𝜉{𝑓,𝑔})𝜂. First, since 𝜄(𝑋𝑔𝑐)𝜂=𝑗𝛼𝑗𝑔=𝑔, we have 𝜄𝑋𝑓,𝑋𝑔𝜂𝑗𝑋=𝑓𝜂𝑗𝑋𝑔𝑋𝜄𝑔𝑋𝑓𝜂𝑗=𝛼𝑗𝑋𝑓𝑋𝑔𝜄𝑔𝛼𝑗𝜉𝑓𝜂=𝛼𝑗{𝑓,𝑔}.(5.13) From Lemma 5.6, we have 𝜉{𝑓,𝑔}={𝑓,𝜉𝑔}{𝑔,𝜉𝑓}=𝑋𝑓(𝜉𝑔)𝑋𝑔(𝜉𝑓), and, thus, 𝜄𝑋𝑓,𝑋𝑔Φ𝑔𝑋=𝑓𝑋(𝑑𝑔(𝜉𝑔)𝜂)𝜄𝑔𝑑(𝜉𝑓)𝜂+(𝜉𝑓)Φ𝑔𝑋=𝑑𝑓𝑔𝑋𝑓(𝜉𝑔)(𝜉𝑔)(𝜉𝑓)𝜂+𝑋𝑔𝑋(𝜉𝑓)𝜂𝑔𝑑(𝜉𝑓)(𝜉𝑓)(𝑑𝑔(𝜉𝑔)𝜂)=𝑑𝑓𝑋𝑔(𝜉𝑓)𝑔𝑓(𝜉𝑔)𝑋𝑔𝜂(𝜉𝑓)=𝑑{𝑓,𝑔}𝜉{𝑓,𝑔}𝜂.(5.14)

Acknowledgments

The research for this paper was made possible by a postdoctoral fellowship from Natural Sciences and Engineering Research Council of Canada (NSERC), and by the University of California, Berkeley, the host institution for the fellowship. The author would like to thank Alan Weinstein for several useful discussions and suggestions which helped to improve the paper.