Topical Review The following article is Open access

A review on coisotropic reduction in symplectic, cosymplectic, contact and co-contact Hamiltonian systems

and

Published 5 April 2024 © 2024 The Author(s). Published by IOP Publishing Ltd
, , Citation Manuel de León and Rubén Izquierdo-López 2024 J. Phys. A: Math. Theor. 57 163001 DOI 10.1088/1751-8121/ad37b2

1751-8121/57/16/163001

Abstract

In this paper we study coisotropic reduction in different types of dynamics according to the geometry of the corresponding phase space. The relevance of coisotropic reduction is motivated by the fact that these dynamics can always be interpreted as Lagrangian or Legendrian submanifolds. Furthermore, Lagrangian or Legendrian submanifolds can be reduced by a coisotropic one.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 4.0 license. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

The introduction of symplectic geometry in the study of Hamiltonian systems was a tremendous breakthrough, both in quantitative and qualitative aspects. For example, we have the results in the reduction of the original Hamiltonian system when in the presence of symmetries, or the so-called coisotropic reduction [1, 5, 16, 42]. Another relevant example, in the quantitative aspects, is the development of geometric integrators that respect geometric aspects and prove to be more efficient than the traditional ones (see for instance [44, 50]). It has also had a major influence on the study of completely integrable systems and Hamilton–Jacobi theory [1, 5, 12, 18]. In addition, the so-called geometric quantization relies on symplectic geometry [39, 54].

Regarding the reduction in the presence of symmetries, the most relevant result is the so-called Marsden–Weinstein symplectic reduction theorem [43] (a preliminary version can be found in Meyer [46]) using the momentum mapping, a natural extension of the classical linear and angular momentum. The reduced manifold is obtained using a regular value of the momentum mapping and the corresponding isotropy group, and the dynamics is projected to this reduced manifold, gaining for integration a smaller number of degrees of freedom. This theorem has been extended to many other contexts: cosymplectic, contact, and more general settings (see [2, 3, 15, 17, 29, 40, 45, 63] and the references therein). For a recent review on reduction by symmetries in cosymplectic geometry we refer to [23]. When the reduced space is not a manifold, we can reduce the algebra of observables [53] (see also [36]). This reduction recovers the Poisson algebra of the reduced space in Marsden–Weinstein reduction.

Related to the geometric reduction is Noether's theorem (in fact, this reduction is a generalisation of it), which states that a symmetry of the system produces a conserved quantity [6]. The introduction of geometric structures has revealed itself in a plethora of results relating symmetries and conserved quantities [16].

Furthermore, Lagrangian submanifolds play a crucial role, since it is easy to check that the image of a Hamiltonian vector field XH in a symplectic manifold $(M, \omega)$ can be interpreted as a Lagrangian submanifold of the symplectic manifold $(TM, \omega^c)$, where ωc is the complete or tangent lift of ω to the tangent bundle TM. This result has its equivalent in Lagrangian mechanics, and has led to the so-called Tulczyjew triples, which elegantly relate the different Lagrangian submanifolds that appear in Lagrangian and Hamiltonian descriptions of mechanics via the Legendre transformation [16, 58, 59]. This interpretation of the dynamics as a Lagrangian submanifold has been extended to other scenarios, including the Tulczyjew triple [21, 22, 2628, 34, 35, 63]. Lagrangian submanifolds are also relevant to develop the so-called Hamilton–Jacobi theory since they provide the geometric setting for solutions of the Hamilton–Jacobi problem (see [25] for a recent topical review on the subject). In this sense, we follow Weinstein's creed: 'Everything is a Lagrangian submanifold' [61].

Moreover, coisotropic submanifolds play a relevant role both in the theory of constraints and in the theory of quantization. For instance, coisotropic submanifolds are precisely the first class constraints considered by Dirac in [24] (see also [7]), where he developed the constraint algorithm for singular Lagrangians in the Hamiltonian setting. The constraint algorithm has been developed in geometrical terms in [32, 33]. In this approach, the phase space of a singular system is a presymplectic manifold and, in [31], Gotay showed that every presymplectic manifold $(P, \Omega)$ may be imbedded in a symplectic manifold $(M, \omega)$ as a closed coisotropic submanifold. More precisely, there exists an imbedding $j: P \rightarrow M$ such that

  • j(P) is closed in M;
  • $j^\ast \omega = \Omega$;
  • $TP^{\perp} \subseteq j_\ast (TP) $.

An alternative approach to the usual treatments of singular Lagrangians based on a Hamiltonian regularization scheme inspired on the coisotropic embedding of presymplectic systems was developed in [37].

The ideas to develop the coisotropic reduction procedure came from Weinstein [62] and were also partially inspired by Roels and Weisntein [49] and Marsden and Weinstein [43].

Coisotropic reduction works when we give a coisotropic submanifold N of a symplectic manifold $(M, \omega)$ and we consider (if it is well defined) the quotient manifold $N/(TN)^\perp $, where $(TN)^\perp$ is the symplectic complement of TN. Being involutive, this distribution along N defines a foliation. The corresponding leaf space inherits a reduced symplectic form from the symplectic structure given on M. If in addition we have a Lagrangian submanifold L with clean intersection with N, then $L \cap N$ projects into a Lagrangian submanifold of the quotient (see [1, 61]). Coisotropic reduction can be also combined with symplectic reduction to develop a reduction procedure for the Hamilton–Jacobi equation in presence of symmetries (see [18]).

Coisotropic reduction has been extended to the field of contact manifolds (with the interest of being in a dissipative context) [15, 57], but it has not been studied in sufficient detail in the case of cosymplectic manifolds nor in that of co-contact manifolds, the latter the natural settings to study time-dependent Hamiltonian contact systems [13, 19].

The objectives of this paper are twofold. On the one hand, to develop in detail the coisotropic reduction in the case of cosymplectic manifolds and those of co-contact, covering a gap in the literature. On the other hand, to present a survey that brings together in one place the different cases that appear in the study of Hamiltonian systems in classical mechanics.

The paper is structured as follows. Sections 2 and 3 are devoted to the main ingredients concerning symplectic Hamiltonian systems and the classical coisotropic reduction procedure. In order to go to the cosymplectic setting, we recall some general notions in Poisson structures (section 4) and then we consider the case of coisotropic reduction in the cosymplectic setting in section 5 (remember that this is the scenario to develop time dependent Hamiltonian systems). Contact manifolds require a more general notion than Poisson structures; indeed, they are examples of Jacobi structures, so that we give some fundamental notions in section 6. The coisotropic reduction scheme developed in contact manifolds is the subject of section 7, which is very different to the cosymplectic case since we are in presence of dissipative systems. We emphasize these differences in section 8, where we study the corresponding Lagrangian settings. To combine dissipative systems with Hamiltonians depending also on time, we consider cocontact manifolds in section 9, and develop there the corresponding coisotropic reduction procedure. Finally, we discuss a recent generalization of contact and cosymplectic systems called stable Hamiltonian systems in section 10.

2. Symplectic vector spaces

We refer to [1, 5, 16, 30, 42, 60] for the main definitions and results.

Definition 2.1 (Symplectic vector space). A symplectic vector space is a pair $(V, \omega)$ where V is a finite dimensional vector space and ω is a non-degenerate 2-form, called the symplectic form. Here, non-degeneracy means that the map

is an isomorphism.

For every non-degenerate 2-form on V there exist a basis $(x_i, y^i)$ whith i taking values from 1 to n such that, making use of the summation convention, $\omega = x^i \wedge y_i$, where $(x^i, y_i)$ is the dual basis. This implies that a symplectic vector space is necessarily of even dimension 2n.

Definition 2.2 (ω-orthogonal). Let $W \subseteq V$ be a subspace of V. We define its ω-orthogonal complement as

Note that $W^{\perp_\omega} = \operatorname{Ker} (i^* \flat_\omega)$ where $i :W \hookrightarrow V$ is the natural inclusion. Using the non-degeneracy of ω, this implies that $\operatorname{dim} W^{\perp_\omega} = \operatorname{dim} V - \operatorname{dim} W$, a result which will be useful throughout this paper.

The antisymmetry of ω gives rise to a wide variety of situations. In particular, we say that $W \subseteq V$ is:

  • i)  
    Isotropic if $W \subseteq W^{\perp_\omega}$ (if W is isotropic, necessarily $\operatorname{dim} W \unicode{x2A7D} n$);
  • ii)  
    Coisotropic if $W^{\perp_\omega} \subseteq W$ (if W is coisotropic, necessarily $\operatorname{dim} W \unicode{x2A7E} n$);
  • iii)  
    Lagrangian if W is isotropic and has an isotropic complement (if W is Lagrangian, necessarily $\operatorname{dim} W = n$);
  • iv)  
    Symplectic if $V = W \oplus W^{\perp_\omega}$.

A subspace W is Lagrangian if and only if $W = W^{\perp_\omega}$. This implies that Lagrangian subspaces are the isotropic subspaces of maximal dimension and the coisotropic subspaces of minimal dimension.

It can be easily checked that the symplectic complement has the following properties:

  • i)  
    $(W_1 \cap W_2)^{\perp_\omega} = W_1^{\perp_\omega} + W_2^{\perp_\omega};$
  • ii)  
    $(W_1 + W_2)^{\perp_\omega} = W_1^{\perp_\omega} \cap W_2^{\perp_\omega};$
  • iii)  
    $(W^{\perp_\omega})^{\perp_\omega} = W.$

3. Coisotropic reduction in symplectic geometry

Definition 3.1 (Symplectic manifold). A symplectic manifold is pair $(M , \omega)$ where M is a manifold and ω is a closed 2-form such that $(T_qM, \omega_q)$ is a symplectic vector space, for every $q \in M$. As in the linear case, for the existence of such a form, M needs to have even dimension 2n.

Every symplectic manifold is locally isomorphic, that is, there exists a set of canonical coordinates around each point:

Theorem 3.1 (Darboux theorem). Let $(M, \omega)$ be a symplectic manifold and $q \in M$. There exist a coordinate system $(q^i,p_i)$ around q such that $\omega = \mathrm{d}q^i \wedge \mathrm{d}p_ i$. These coordinates are called Darboux coordinates.

This non-degenerate form induces a bundle isomorphism between the tangent and cotangent bundles of M point-wise, namely

Definition 3.2 (Hamiltonian vector field). Given $H \in \mathcal{C}^\infty(M)$, we define the Hamiltonian vector field of H as

where $\sharp_\omega = \flat_\omega^{-1}.$ We say that a vector field X is Hamiltonian if $X = X_H$ for some function H and say that X is locally Hamiltonian if $X = X_H$ for some local function defined in a neighborhood of every point of the manifold.

Remark 3.1. Notice that a vector field is locally Hamiltonian if and only if $\flat_\omega(X)$ is closed, and Hamiltonian if and only if $\flat_\omega(X)$ is exact.

Locally, Hamiltonian vector fields have the expression

Then, the integral curves of the Hamiltonian vector field $X_H, (q^i(t), p_i(t))$, satisfy the local differential equations

which are the Hamilton's equations of motion.

The definitions of the different cases of subspaces given in the linear case can be extended point-wise to submanifolds $N \hookrightarrow M$. Consequently, we say that $N \hookrightarrow M$ is:

  • i)  
    Isotropic if $T_qN \subseteq T_qM$ is for every $q \in N$;
  • ii)  
    Coisotropic if $T_qN \subseteq T_qM $ is for every $q \in N$;
  • iii)  
    Lagrangian if N is isotropic and there is a isotropic subbundle (where we understand isotropic point-wise) $E \subseteq TM | N$ such that $TM = TN \oplus E$ (here $\oplus$ denotes the Whitney sum). This is exactly the point-wise definition of a Lagrangian subspace asking for the coisotropic complement to vary smoothly;
  • iv)  
    Symplectic if $T_qN \subseteq T_qM$ is for every $q \in N$.

These definitions extend naturally to distributions.

Just like in the linear case, a submanifold $N \hookrightarrow M$ is Lagrangian if and only if it is isotropic (or coisotropic) and has maximal (or minimal) dimension. This is a useful characterization that we will use several times in the rest of the paper.

Lemma 3.1. Let $i: L \rightarrow M$ be a submanifold of dimension n. Then, L is a Lagrangian submanifold of $(M, \omega)$ if and only if $i^*\omega = 0$.

Proof. It is trivial, since Lagrangian submanifolds are the isotropic submanifolds of maximal dimension, say n. □

3.1. Hamiltonian vector fields as Lagrangian submanifolds

Definition 3.3. Let $(M , \omega)$ be a symplectic manifold. Define the tangent symplectic structure on TM as $\omega_0 = -\mathrm{d}\lambda_0$ where $\lambda_0 = \flat_\omega^*\lambda_M$, and λM is the Liouville 1-form on the cotangent bundle.

Recall that λM is defined as follows:

where $X_{\alpha_z} \in T_{\alpha_x}(T^*M)$, $\alpha_x \in T^*_xM$, and $\pi_M : T^*M \longrightarrow M$ is the canonical projection. The Liouville 1-form can be also defined as the unique 1-form λM on $T^*M$ such that, for every 1-form $\alpha:M \rightarrow T^*M$,

In coordinates $(q^i, p_i, \dot{q}^i, \dot{p}_i)$ the tangent symplectic structure is

Proposition 3.1. Let $X: M \rightarrow TM$ be a vector field. Then

Proof. This is a straight-forward verification. Let $v \in T_qM$, then we have

 □

Proposition 3.2. Let $X: M \rightarrow TM$ be a vector field. Then X is locally Hamiltonian if and only if X(M) is a Lagrangian submanifold of $(TM, \omega_0)$.

Proof. We only check that X(M) is isotropic using lemma 3.1, since $\operatorname{dim} X(M) = \operatorname{dim} M = \frac{1}{2} \operatorname{dim} TM.$ In fact:

which gives the characterization. □

We can also check this last proposition easily in coordinates. Indeed, let

We have

and thus, X defines a Lagrangian submanifold if and only if

Taking

and

these conditions become

This implies $G^i = \displaystyle \frac{\partial H}{\partial x^i},$ for some local function H. It is clear that locally, we have $X = X_H$.

3.2. Coisotropic reduction

Now, given a coisotropic submanifold $N \hookrightarrow M$, we define the distribution $(TN)^{\perp_\omega}$ on N as the subbundle of $TM |_N$ consisting of all ω-orthogonal spaces $(T_qN)^{\perp_\omega}$. Note that this distribution is regular and its rank is $\operatorname{dim} M - \operatorname{dim} N$.

Proposition 3.3. Let $(M, \omega)$ be a symplectic manifold and $N \hookrightarrow M$ be a coisotropic submanifold. The distribution $q \mapsto (T_qN)^{\perp_\omega}$ is involutive.

Proof. Let $X,Y$ be vector fields along N with values in $TN^{\perp_\omega}$ and Z be any other vector field tangent to N. Since ω is closed we have

since $X,Y$ belong to the orthogonal complement of TN. We conclude that $\omega([X,Y],Z) = 0$ for every field Z tangent to N, that is, $[X,Y] \in (TN)^{\perp_\omega}$. □

Since the distribution is involutive and regular, the Frobenius' Theorem guarantees the existence of a maximal regular foliation $\mathcal{F}$ of N, that is, a decomposition of N into maximal submanifolds tangent to the distribution. In what follows, we suppose that $N / \mathcal{F}$ (the space of all leaves) admits a manifold structure so that the projection

is a submersion. The main result is the Weinstein reduction theorem [62]:

Theorem 3.2 (Coisotropic reduction in the symplectic setting). Let $(M , \omega)$ be a symplectic manifold and $N \hookrightarrow M$ be a coisotropic submanifold. If $N / \mathcal{F}$ (the spaces of all leaves under the distribution $q \mapsto (T_qN)^{\perp_\omega}$) admits a manifold structure such that $N \xrightarrow{\pi} N / \mathcal{F}$ is a submersion, there exist an unique 2-form ωN on $N / \mathcal{F}$ that defines a symplectic manifold structure such that, if $N \xrightarrow{i}M$ is the natural inclusion, then $i^* \omega = \pi ^* \omega_N$. The following diagram summarizes the situation:

Standard image High-resolution image

Proof. Uniqueness is guaranteed from the imposed relation since it forces us to define

where $[u]: = T\pi(q) \cdot u$. We only need to check that this is a well-defined closed form and that it is non-degenerate.

We begin showing that our definition does not depend on the representative of the vector $[u]$. For this, it is sufficient to observe that $(\omega_N)_{[q]}([u],[v]) = 0$ whenever u is a vector in the distribution.

Furthermore,

for every vector field X in N with values in $(TN^{\perp_\omega})$, and this implies the independence of the point (for every two points in the same leaf of the foliation can be joined by a finite union of flows of such fields).

It is clearly non-degenerate and it is closed, since $\mathrm{d}\pi^* \omega_N = i_*\mathrm{d}\omega = 0$ and π is a submersion. □

3.3. Projection of Lagrangian submanifolds

Definition 3.4 (Clean intersection). We say that two submanifolds $L, N \hookrightarrow M$ have clean intersection if $L \cap N \hookrightarrow M$ is again a submanifold and $T_q(L \cap N) = T_q L \cap T_qN$, for every $q \in L \cap N$.

Proposition 3.4. Let $L \hookrightarrow M$ be a Lagrangian submanifold and $N \hookrightarrow M$ a coisotropic submanifold. If they have clean intersection and $L_N: = \pi(L \cap N)$ is a submanifold of $N/\mathcal{F}$, LN is Lagrangian.

Proof. It is sufficient to see that is isotropic and that it has maximal dimension in $N / \mathcal{F}$. It is isotropic since $[u] \in T_q(L _ N)$ implies $\omega_N([u],[v]) = \omega(u,v) = 0$, for every $[v] \in T_q(L _ N)$. Now, since $\operatorname{Ker} \mathrm{d}_q\pi = (T_qN)^{\perp_\omega}$, the kernel-range formula yields

Equation (1)

Furthermore,

Equation (2)

because L is Lagrangian and N coisotropic. Substituting (2) in (1) we obtain

which is exactly $\frac{1}{2} \operatorname{dim} N/ \mathcal{F}$, as a direct calculation shows. □

3.4. An example

As an example of coisotropic reduction, let us take

and

Since it has codimension 1, it defines a coisotropic submanifold if we endow M with the natural symplectic form $\omega : = \mathrm{d} q^i \wedge \mathrm{d} p_i.$ It is easy to check that $(TN)^{\perp_\omega}$ is generated by

where $(q^i, p_i)$ are canonical coordinates in $\mathbb{R}^{2(n +1)}$. Therefore, the leaves of the corresponding foliation are precisely the orbits of the previous vector field in $\mathbb{S}^{2n +1}$. Making the identification

two points $x, y \in \mathbb{S}^{2n +1}$ are in the same orbit if and only if there exist some $\alpha \in \mathbb{C}$ ($|\alpha| = 1$) such that

This implies that

is the complex projective space of complex dimension n, $\mathbb{P}^n \mathbb{C}$. Therefore, we conclude that through coisotropic reduction we can define a natural symplectic structure on $\mathbb{P}^n \mathbb{C},$ for every n.

The above example is taken from exercise 5.3B in [1].

4. Poisson structures

A symplectic structure $(M,\omega)$ induces a Lie algebra structure on the ring of functions $\mathcal{C}^\infty(M)$.

Definition 4.1 (Poisson bracket). Let $(M,\omega)$ be a symplectic manifold and $f, g \in \mathcal{C}^\infty(M)$. We define the Poisson bracket of $f,g$ as the function

It is easily checked that in Darboux coordinates the Poisson bracket is

Proposition 4.1. The Poisson bracket satifies the following properties:

  • i)  
    It is bilinear with respect to $\mathbb{R}$;
  • ii)  
    $\{f,g\cdot h\} = g \cdot \{f,h\} +\{f,g\} \cdot h $ (the Leibniz rule);
  • iii)  
    $\{f,\{g,h\}\} + \{h,\{f,g\}\} + \{g,\{h,f\}\} = 0$ (the Jacobi identity).

Taking into consideration the previous definition, we can generalize the notion of symplectic manifolds as follows:

Definition 4.2 (Poisson manifold). A Poisson manifold is a pair $(P, \{\cdot, \cdot \})$ where P is a manifold and $\{\cdot , \cdot \}$ is an antisymmetric bracket in the ring of functions $\mathcal{C}^\infty(P)$ satisfying the Leibniz rule and the Jacobi identity.

Definition 4.3 (Hamiltonian vector field, Characteristic distribution). Given $H \in \mathcal{C}^\infty(M),$ the Leibniz rule implies that $\{H,\cdot\}$ defines a derivation on $C^\infty(M)$ an thus is associated to a unique vector field XH , which will be called the Hamiltonian vector field of H. The collection of all Hamiltonian vector fields generates the characteristic distribution, namely

The definition of Hamiltonian vector field implies that $\{f,g\}$ only depends on the values of $\mathrm{d}f, \mathrm{d}g$ and thus we can define a bivector field

where $\mathrm{d}f = \alpha, \mathrm{d}g = \beta.$ We have $\{f,g\} = \Lambda(\mathrm{d}f,\mathrm{d}g).$ Λ also satisfies the partial differential equation $[\Lambda, \Lambda] = 0$, where $[\cdot, \cdot]$ is the Schouten–Nijenhuis bracket [60]. This last property is actually equivalent to the Jacobi identity, that is, given a bivector field Λ, $\{f,g\}: = \Lambda(\mathrm{d}f,\mathrm{d}g)$ defines a Poisson structure if and only if $[\Lambda, \Lambda] = 0.$

Definition 4.4. Let $(P, \{\cdot, \cdot \})$ be a Poisson manifold. Then we define

Notice that $\operatorname{Im} \sharp_\Lambda = \mathcal{S}$, the characteristic distribution.

Remark 4.1. In the case of symplectic manifolds $\sharp_ \omega = \sharp_\Lambda$, and the characteristic distribution is the whole tangent bundle; however, in the general setting $\sharp_\Lambda$ need not be a bundle isomorphism. Actually, if $\sharp_\Lambda$ is a bundle isomorphism, it arises form a symplectic structure defined as $\omega(v,w): = \Lambda(\sharp_\Lambda^{-1}(v), \sharp_\Lambda^{-1}(w))$ [16].

This type of distributions is in general not of constant rank, so we cannot directly apply the Frobenius' theorem. But there is an extension of the result, due to Stefan [55] and Sussmann [56] (independently) that works for generalised distributions, locally generated by vector fields that leave the distribution invariant. This is the situation for characteristic distributions in the case of Poisson manifolds (see [42]).

So, the characteristic distribution is involutive and each leaf S of the foliation admits a symplectic structure defining for $f,g \in \mathcal{C}^\infty(S)$ and $q \in S$,

for arbitrary extensions $\widetilde f,\widetilde g \in\mathcal{C}^\infty(P)$ of $f,g$ respectively. It can be easily checked that this definition does not depend on the chosen functions and that it defines a non-degenerate Poisson structure and thus, S is a symplectic manifold [60]. The symplectic form is given by

where $f, g \in C^\infty(S).$

Definition 4.5 (Λ-orthogonal). Let $\Delta_q \subseteq T_qP$ be a subspace on a Poisson manifold $(P, \{\cdot, \cdot \})$. We define the Λ-orthogonal complement $\Delta_q^{\perp_\Lambda} = \sharp_\Lambda(\Delta_q^0)$ where $\Delta_q^0$ is the annihilator of $\Delta_q$, that is, $\Delta_q^0: = \{\alpha \in T^*_qP \,\, | \, \, \alpha = 0 \,\, \text{in} \,\, \Delta_q\}.$

Just as in the symplectic scenario, we say that a subspace $\Delta_q\subseteq T_qP$ is

  • i)  
    Isotropic if $\Delta_q \subseteq \Delta_q^{\perp_\Lambda}$ for every $q \in P$;
  • ii)  
    Coisotropic if $\Delta_q^{\perp_\Lambda} \subseteq \Delta_q$ for every $q \in P$;
  • iii)  
    Lagrangian if $\Delta_q = \Delta_q^{\perp_\Lambda} \cap \mathcal{S}_q$ for every $q \in P$. Notice that this is equivalent to $\Delta_q \cap \mathcal{S}_q$ being Lagrangian in each symplectic vector space $\mathcal{S}_q.$

The Λ-orthogonal complement satisfies the following properties:

  • $i)$  
    $(W_1 \cap W_2)^{\perp_\Lambda} = W_1^{\perp_\Lambda} + W_2^{\perp_\Lambda};$
  • $ii)$  
    $(W_1 + W_2)^{\perp_\Lambda} \subseteq W_1^{\perp_\Lambda} \cap W_2^{\perp_\Lambda}.$

Remark 4.2. For symplectic manifolds, the above definitions coincide with the ones previously given.

5. Coisotropic reduction in cosymplectic geometry

Cosymplectic structures are relevant precisely because they are the natural arena to develop time-dependent Lagrangian and Hamiltonian mechanics [16].

Definition 5.1 (Cosymplectic manifold). A cosymplectic manifold is a triple $(M,\Omega, \theta)$ where M is a $(2n + 1)$-manifold, θ is a closed 1-form and Ω is a closed 2-form such that $\theta \wedge \Omega^n\neq 0$.

Similar to the symplectic setting, there exist canonical coordinates, which will be called Darboux coordinates $(q^i,p_i,t)$ such that $\Omega = \mathrm{d}q^i \wedge \mathrm{d}p_i$ and $\theta = \mathrm{d}t$. The existence of such coordinate charts is proven in [30].

There are two natural distributions defined on M:

  • i)  
    The horizontal distribution $\mathcal{H} : =\operatorname{Ker} \theta$;
  • ii)  
    The vertical distribution $\mathcal{V}: = \operatorname{Ker} \Omega$.

These distributions induce the following types of tangent vectors in each tangent space. A vector $v \in T_qM$ will be called:

  • i)  
    Horizontal if $v \in \mathcal{H}_q$;
  • ii)  
    Vertical if $v \in \mathcal{V}_q$.

In Darboux coordinates, these distributions are locally generated as follows:

Just as before, we can define a bundle isomorphism between the tangent and cotangent bundles:

Its inverse is denoted by $\sharp_{\theta, \Omega}$.

The vector field defined as $\mathcal{R} : = \sharp_{\theta, \Omega}(\theta)$ is called the Reeb vector field. The Reeb vector field is locally given by

Let H be a differentiable function on M. We define the following vector fields

  • i)  
    The gradient vector field $\operatorname{grad} H : = \sharp_{\theta, \Omega} (\mathrm{d}H)$;
  • ii)  
    The Hamiltonian vector field $X_H : = \operatorname{grad} H - \mathcal{R}(H) \mathcal{R}$;
  • iii)  
    The evolution vector field $\mathcal{E}_H : = X_H + \mathcal{R}.$

These vector fields have the local expressions:

Notice that an integral curve $(q^i(\lambda), p_i(\lambda), t(\lambda)$ of the evolution vector field satisfies the following differential equations

which immediately give time-dependent Hamilton's equations

since we have $t = \lambda+const.$

Notice that the horizontal distribution $\mathcal{H}$ is the distribution generated by all Hamiltonian vector fields. Just as in the symplectic case, we can define a Poisson bracket:

Definition 5.2 (Poisson bracket). Let $\{\cdot, \cdot \}$ be the bracket in the ring $\mathcal{C}^{\infty}(M)$ given by

We can easily check that this is indeed a Poisson structure by observing that in coordinates it is given by

Thus, the coordinate expression of the Poisson tensor is

So, the Hamiltonian vector fields coincide with the ones provided by this induced Poisson structure, following the notions and results given in section 4. In particular, given $\Delta_q \subseteq T_qM$, we have

Note that $\operatorname{Ker} \sharp_\Lambda = \langle \theta \rangle $ and that $\operatorname{Im} \sharp_\Lambda = \mathcal{H},$ that is, $\mathcal{H}$ is the characteristic distribution of the Poisson structure induced by $(\theta, \Omega)$. This implies the following result:

Proposition 5.1.  $i: L \rightarrow M$ is a Lagrangian submanifold if and only if

for every $q \in L.$

Proof. It follows from the definition of Lagrangian submanifold (section 4) and the fact that $\mathcal{H}$ is the characteristic distribution on M. □

It is also easy to see that

observing that we have $\Omega(X_f,X_g) = \Omega(\operatorname{grad} f, \operatorname{grad} g)$.

5.1. Gradient, Hamiltonian and evolution vector fields as Lagrangian submanifolds

Definition 5.3. Given a cosymplectic manifold $(M, \Omega, \theta)$, we define the symplectic structure on TM as $\Omega_0 : = - \mathrm{d}\lambda_0$, where $\lambda_0 = \flat_{\theta,\Omega}^*\lambda_M$, λM being the Liouville 1-form in the cotangent bundle $T^*M$.

There is another expression of Ω0, namely

as one can verify [11]. Here, $\alpha^v, \alpha^c$ denote the complete and vertical lifts of a form α on M to its tangent bundle TM [16]. This implies that in the induced coordinates in TM, $(q^i,p_i, t, \dot{q}^i, \dot{p}_i, \dot{t})$,

Proposition 5.2. Let $(M, \Omega, \theta)$ be a cosymplectic manifold and $X: M \rightarrow TM$ a vector field. Then X(M) is a Lagrangian submanifold of $(TM, \Omega_0)$ if and only if X is locally a gradient vector field.

Proof. It is easily checked that $X^*\lambda_0 = \flat_{\theta,\Omega}(X)$ (just like in proposition 3.1) and then, X(M) is Lagrangian if and only if

that is, X is locally a gradient vector field. □

We can also check this in coordinates. Indeed, let

be a vector field on M. $X: M \hookrightarrow TM$ defines a Lagrangian submanifold if and only if $X^*\Omega_0 = 0.$ An easy calculation gives

Therefore, X defines a Lagrangian submanifold of $(TM, \Omega_0)$ if and only if

Equation (3)

Equation (4)

Equation (5)

Equation (6)

Equation (7)

The equations above can be summarized by taking

since they translate to

We conclude that $G^i = \displaystyle \frac{\partial H}{\partial x^i}$, for some local function H, that is, locally, $X = \operatorname{grad} H$.

In general, the Hamiltonian and evolution vector field do not define Lagrangian submanifolds in $(TM,\Omega_0)$. However, modifying the form we can achieve this. First, let us study the Hamiltonian vector field XH . We have

The form defined as

is a symplectic form and has the local expression

Also,

We have proved that XH defines a Lagrangian submanifold of $(TM, \Omega_H).$ Furthermore, since

it follows that the evolution vector field $\mathcal{E}_H$ also defines a Lagrangian submanifold of $(TM , \Omega_H)$.

This also gives a way of interpreting both vector fields as Lagrangian submanifolds of a the cosymplectic submanifold $(TM \times \mathbb{R}, \Omega_H, \mathrm{d}s),$ taking the coordinate in $\mathbb{R}$ to be constant.

5.2. Coisotropic reduction

We can interpret the orthogonal complement defined by the Poisson structure using the cosymplectic structure. We note that $\Omega|_\mathcal{H}$ defined as Ω restricted to the distribution $\mathcal{H}$ induces a symplectic vector space in each $\mathcal{H}_q$ and thus we have a symplectic vector bunde $\mathcal{H} \rightarrow M$. If $\Delta_q \subseteq \mathcal{H}_q$, we have the $\Omega|_\mathcal{H}$-orthogonal complement

Proposition 5.3. Let $\Delta_q \subseteq T_qM$. Then $\Delta_q^{\perp_\Lambda} = (\Delta_q\cap \mathcal{H})^{\perp_{\Omega|_{\mathcal{H}}}}.$

Proof. Let $v \in \Delta_q^{\perp_\Lambda}$, that is, $v = \sharp_\Lambda(\alpha)$ with $\alpha \in \Delta_q^0$. This implies that v is horizontal. We only need to check that $\Omega(v,w) = 0$ for every $w \in \Delta_q \cap \mathcal{H}_q$. Indeed, since $\theta(w) = 0$,

Now we compare dimensions. We distinguish two cases, if $\theta \in \Delta_q^0$, we have

which is exactly $\operatorname{dim}(\Delta_q \cap \mathcal{H}_q)^{\perp_{\Omega|_\mathcal{H}}}$, for $\Delta_q \subseteq \mathcal{H}_q$ and $(\mathcal{H}_q, \Omega |_\mathcal{H})$ is symplectic. Now, if $\theta \not \in \Delta_q^0$, then

and, since $\Delta_q \not \subseteq \mathcal{H}_q$, we have $\operatorname{dim} (\Delta_q \cap \mathcal{H}_q) = \operatorname{dim} \Delta_q - 1$ which implies that $\operatorname{dim} (\Delta_q \cap \mathcal{H}_q)^{\perp_{\Omega|_\mathcal{H}}} = 2n + 1 - \operatorname{dim} \Delta. $ □

This last proposition clarifies the situation. The Λ-orthogonal of a subspace Δ is just the symplectic orthogonal of the intersection with the symplectic leaf. This means that coisotropic reduction in cosymplectic geometry will be performed in each leaf of the characteristic distribution $\mathcal{H}$. Also, because the Λ-orthogonal complement is just the symplectic complement of the intersection with $\mathcal{H}$, we have the following properties:

  • i)  
    $(\Delta_1 \cap \Delta_2)^{\perp_\Lambda} = \Delta_1^{\perp_\Lambda} + \Delta_2^{\perp_\Lambda}.$
  • ii)  
    $(\Delta_1 + \Delta_2)^{\perp_\Lambda} = \Delta_1^{\perp_\Lambda} \cap \Delta_2^{\perp_\Lambda}.$
  • iii)  
    $(\Delta^{\perp_\Lambda})^{\perp_\Lambda} = \Delta \cap \mathcal{H}.$

It will also be important to distinguish submanifolds $N \hookrightarrow M$ acording to the position relative to the distributions $\mathcal{H},\mathcal{V}$.

Definition 5.4 (Horizontal, non-horizontal and vertical submanifolds). Let $i: N \hookrightarrow M$ be a submanifold. N will be called a:

  • i)  
    Horizontal submanifold if $T_qN \subseteq \mathcal{H}_q$ for every $q \in N$;
  • ii)  
    Non-horizontal submanifold if $T_qN \not \subseteq \mathcal{H}_q$ for every $q \in N$;
  • iii)  
    Vertical submanifold if the Reeb vector field is tangent to N, that is, $\mathcal{R}(q) \in T_qN$ for every $q \in N$.

Remark 5.1. Note that if $N\hookrightarrow M$ is a vertical submanifold, then N is non-horizontal.

Lagrangian submanifolds are characterized as follows:

Lemma 5.1. Let $L\hookrightarrow M$ be a Lagrangian submanifold and $q \in L$. Then

  • i)  
    If $T_qL \subseteq \mathcal{H}_q$, $\operatorname{dim} T_qL^{\perp_\Lambda} = \operatorname{dim} M - \operatorname{dim} L -1$.
  • ii)  
    If $T_qL \not \subseteq \mathcal{H}_q$, $\operatorname{dim} T_qL^{\perp_\Lambda} = \operatorname{dim} M - \operatorname{dim} L$.

and, in either case, $\operatorname{dim} T_qL^{\perp_\Lambda} = n$, where $\operatorname{dim} M = 2n +1$.

Proof. 

  • i)  
    Since $\theta \in T_qL^0$ we have
  • ii)  
    It follows from the previous calculation using that $\theta \not \in T_qL^0$ because

The proof of the equality $\operatorname{dim} T_qL^{\perp_\Lambda} = n$ is straightforward using that $T_qL \cap \mathcal{H}_q$ is a Lagrangian subspace of $(\mathcal{H}_q, \Omega|_\mathcal{H})$. □

Lemma 5.1 guarantees that either $\operatorname{dim} L = n$, in which case L is horizontal, or $\operatorname{dim} L = n+1$, in which case L is non-horizontal. We have the following useful characterization of Lagrangian submanifolds:

Lemma 5.2. Let $L \hookrightarrow M$ be a submanifold. We have

  • i)  
    If $\operatorname{dim} L = n$, then L is Lagrangian if and only if $i^* \theta = 0, i^* \Omega = 0$.
  • ii)  
    If L is non-horizontal and $\operatorname{dim} L = n+1$, then L is Lagrangian if and only if $i^*\Omega = 0$.

Proof. Both assertions are proved by a comparison of dimensions. □

Proposition 5.4. Let $ i:N \hookrightarrow M $ be a coisotropic submanifold. Then the distribution $(TN)^{\perp_\Lambda}$ is involutive.

Proof. We start proving that $\mathcal{H}$ is an involutive distribution. Let $X,Y$ be vector fields tangent to $\mathcal{H}$. Since θ is closed we have

that is, $[X,Y]$ is tangent to $\mathcal{H}.$

Denote $\Omega_0: = i^* \Omega$. Let $X,Y$ be vector fields in N tangent to $(TN)^{\perp_\Lambda}$. Using proposition 5.3, $[X,Y] \in (TN)^{\perp_\Lambda}$ if and only if $[X,Y] \in (TN \cap \mathcal{H})^{\perp_{\Omega |_\mathcal{H}}}$. In order to see this, we take an arbitrary vector field Z on N tangent to $\mathcal{H}$ and check that $\Omega_0([X,Y],Z) = 0$. Because Ω is closed, we have

where we have used that $X,Y,Z,[Y,Z], [X,Z]$ are horizontal (since $\mathcal{H}$ is involutive) and that $X,Y \in (TN \cap \mathcal{H})^{\perp_{\Omega |_\mathcal{H}}}$. □

5.3. Vertical coisotropic reduction

We shall now study coisotropic reduction of a vertical submanifold $N \hookrightarrow M$. Let $q \in N$. We have $\operatorname{dim} (T_qN)^0 = \operatorname{dim} M - \operatorname{dim} N$. Since N is vertical, $\theta \not \in (T_qN)^0$ and we have

In particular, $(TN)^{\perp_\Lambda}$ is a regular distribution.

Theorem 5.1 (Vertical coisotropic reduction in the cosymplectic setting). Let $(M, \Omega, \theta)$ be a cosymplectic manifold and $i:N \hookrightarrow M$ be an coisotropic vertical submanifold. Denote by $\mathcal{F}$ the maximal foliation of the involutive regular distribution $(TN)^{\perp_\Lambda}$. If the space of all leaves $N/ \mathcal{F}$ admits a manifold structure such that the projection $\pi : N \rightarrow N/ \mathcal{F}$ is a submersion, then there exist unique θN , ωN such that

and they define a cosymplectic structre on $N / \mathcal{F}.$ The following diagram summarizes the situation:

Standard image High-resolution image

Proof. Uniqueness is clear from the imposed relation. Denote $\Omega_0: = i^* \Omega$, $\theta_0 : = i^*\theta$. We only need to verify that the following forms are closed, well defined and define a cosymplectic structure:

where $[u]: = T\pi(q) \cdot u \in T_{[q]}N/ \mathcal{F}$. If they were well defined, it is clear that they are smooth and closed since $\pi^* \mathrm{d} \theta_N = \mathrm{d} \theta_0 = 0$, $\pi^*\Omega_N = \mathrm{d} \Omega_0 = 0$ and π is a submersion.

Let us first check that these definitions do not depend on the chosen representatives of the vectors. It suffices to observe that for vectors in the distribution, say $v \in (T_qN)^{\perp_\Lambda}$, we have $i_v \Omega_0 = 0$ and $i_v \theta_0 = 0$. This easily follows from proposition 5.3 using that the horizontal proyection of every vector $u \in T_qN$ is tangent to N (here we use the condition $\mathcal{R}(q) \in T_qN$). To see the independence of the point in the leaf chosen, it is enough to observe that

for every vector field on N tangent to the distribution $(T_qN)^{\perp_\Lambda}$ (since every two points in the same leaf of the foliation can be joined by a finite union of flows of such fields). Indeed, we have

Now we check that they define a cosymplectic structure. Assuming $k = \operatorname{dim} N$ and $2n +1 = \operatorname{dim} M$, from the remark above we have

and hence, $(N/\mathcal{F},\Omega_N,\theta_N)$ is a cosymplectic manifold if and only if

which is equivalent to $\theta_0 \wedge \Omega_0^{k-n-1} \neq 0$, because π is a submersion. For every point $q \in N$, TqN can be decomposed in

where $T_qN_\mathcal{H} = (T_qN) \cap \mathcal{H}_q$. It is easy to see that $(T_qN)_\mathcal{H}$ is a coisotropic subspace of $(\mathcal{H}_q, \Omega|_\mathcal{H})$. This implies (using symplectic reduction) that there are $\operatorname{dim}(T_qN)_\mathcal{H} - \operatorname{dim} (T_qN)_\mathcal{H}^{\perp_{\widetilde \Omega}} = k - 1 - (2n +1 -k) = 2k - 2n - 2$ horizontal vectors, say $u_1, \dots, u_{2k-2n-2}$ such that $\Omega_0^{k-n-1}(u_1, \dots, u_{2k-2n-2})\neq 0$. Taking the last vector to be $\mathcal{R}(q)$, it is clear that

 □

5.3.1. Projection of Lagrangian submanifolds.

Now we will prove that Lagrangian submanifolds $L \hookrightarrow M$ project to Lagrangian submanifolds in $N/\mathcal{F}.$

Proposition 5.5 (Projection of horizontal Lagrangian submanifolds is Lagrangian). Under the hypotheses of theorem 5.1, let $L \hookrightarrow M$ be an horizontal Lagrangian submanifold such that L and N have clean intersection. If $ L_N: = \pi(L \cap N)$ is a submanifold of $N/ \mathcal{F}$, then LN is Lagrangian.

Proof. Let $\mathcal{H}_N$ be the horizontal distribution in $N/\mathcal{F}$. It is clear that LN is horizontal, because $T_{[q]}L_N = T\pi(q)(T_qL \cap T_qN)$. Using proposition 5.3 we have

We will check that

and prove that $\operatorname{dim} L_N = \operatorname{dim} N- n - 1$, which with lemma 5.2 together with the calculation of the dimension of $N / \mathcal{F}$ done in theorem 5.1, yields the result.

Let $[v], [w] \in T_{[q]}L_N$. Then

since L is Lagrangian. Because $[w]$ is arbitrary, this last calculation implies that $[v] \in T_{[q]}L_N^{\perp_{\Omega_N|_{\mathcal{H}_N}}} \subseteq T_{[q]}L_N^{\perp_{\Lambda_N}}$. Now,

Equation (8)

Furthermore, since L is Lagrangian and horizontal, $(T_qL \cap (T_qN)^{\perp_\Lambda})^{\perp_\Lambda} = T_qL \cap \mathcal{H}_q + (T_qN^{\perp_\Lambda})^{\perp_\Lambda} = T_qL + (T_qN^{\perp_\Lambda})^{\perp_\Lambda}$ and thus (using that $T_qL \cap (T_qN^{\perp_\Lambda})^{\perp_\Lambda}$ is necessarily horizontal),

Since $\operatorname{dim}(T_qL\cap \mathcal{H}_q + (T_qN^{\perp_\Lambda})^{\perp_\Lambda}) = \operatorname{dim}(T_qL\cap \mathcal{H}_q + T_qN) - 1$ (which comes from the fact that $\operatorname{dim} (T_qN^{\perp_\Lambda})^{\perp_\Lambda} = \operatorname{dim} T_qN - 1$) we have

Equation (9)

Substituting (9) in (8) and using $\operatorname{dim} L = n$, we conclude

 □

Proposition 5.6 (Projection of non-horizontal Lagrangian submanifold is Lagrangian). Under the hypotheses of theorem 5.1, let $L \hookrightarrow M$ be a non-horizontal Lagragian submanifold. If L and N have clean intersection and $L_N: = \pi(L \cap N) \hookrightarrow N/\mathcal{F}$ is a submanifold, then LN is Lagrangian.

Proof. The proof follows the same lines as that of proposition 5.5. That LN is isotropic follows easily from proposition 5.3. However, in order to calculate $\operatorname{dim} L_N$, we need to distinguish whether $L \cap N$ is horizontal or not.

  • i)  
    If $L\cap N$ is horizontal, we need to check that $\operatorname{dim} L_N = \operatorname{dim} N - n - 1$, since LN is horizontal. Because $(T_qL \cap T_qN^{\perp_\Lambda})^{\perp_\Lambda} = T_qL \cap \mathcal{H}_q + (T_qN^{\perp_\Lambda})^{\perp_\Lambda}$, we have
    It is easy to check that $\operatorname{dim}(T_qL\cap \mathcal{H}_q + (T_qN^{\perp_\Lambda})^{\perp_\Lambda}) = \operatorname{dim}(T_qL\cap \mathcal{H}_q + T_qN) - 1$ and thus,
    We conclude that
    where we have used that $T_qL \cap \mathcal{H}_q \cap T_qN = T_qL \cap T_qN$, since $N \cap L$ is horizontal, and $\operatorname{dim} T_qL \cap \mathcal{H} _q = \operatorname{dim} L - 1$, because $T_qL \not \subseteq \mathcal{H}_q$.
  • ii)  
    If $L\cap N$ is not horizontal, we need to check that $\operatorname{dim} L_N = \operatorname{dim} N - n$. This follows from the same calculation done in i), using that

 □

5.4. Horizontal coisotropic reduction

We will restrict the study to horizontal coisotropic submanifolds $N\hookrightarrow M$, that is, manifolds satisfying $T_qN \subseteq \mathcal{H}_q$ for every $q\in N$. Note that in this case the distribution $(TN)^{\perp_\Lambda}$ is also regular, since

Theorem 5.2 (Horizontal coisotropic reduction in the cosymplectic setting). Let $(M, \Omega, \theta)$ be a cosymplectic manifold and $i: N \hookrightarrow M$ be an horizontal coisotropic submanifold. Denote by $\mathcal{F}$ the space of leaves determined by the regular and involutive distribution $(TN)^{\perp_\Lambda}.$ If $N/ \mathcal{F}$ admits a manifold structure such that $\pi : N \rightarrow N/ \mathcal{F}$ is a submersion, then there exists a unique 2- form $\Omega_N$ in $N/\mathcal{F}$ such that

and $(N/\mathcal{F}, \Omega_N)$ is a symplectic manifold.

Proof. Since N is horizontal and the horizontal distribution is integrable, N will be contained in an unique symplectic leaf and thus, we are performing symplectic reduction. The proof is just repeating what has been done in theorem 3.2. □

We can generalize this process to arbitrary submanifolds. Let $N \hookrightarrow M$ be a coisotropic submanifold. Since in general we cannot guarantee the well-definedness of the 2-form in the quotient, we will reduce the intersection of N with each one of the symplectic leaves. It is clear that $TN\cap \mathcal{H}$ is an involutive distribution, since TN and $\mathcal{H}$ are. If this distribution was regular, for every $q \in N$ there would exist an unique maximal leaf of the distribution, say Sq . Notice that $S_q \hookrightarrow M$ is an horizontal submanifold. We can perform coisotropic reduction on each of these submanifolds.

5.4.1. Projection of Lagrangian submanifolds

Proposition 5.7. Let $L \hookrightarrow M$ be a Lagrangian submanifold. If L and N have clean intersection and $L_N: = \pi(L\cap N)$ is a submanifold, then LN is Lagrangian.

Proof. Let $q \in N \cap L$, we have to prove that

is a Lagrangian subspace of $T_{[q]} N/\mathcal{F} = T_{[q]}N / \left( T_{q} N\right)^{\perp_{\Lambda}}.$ Since N is horizontal, $T_q N \subseteq \mathcal{H}_q$. Now, from proposition 5.3, we know that

that is, Tq N is a coisotropic subspace of $\mathcal{H}_q$, with its natural symplectic structure. A similar argument shows that Tq L is a Lagrangian subspace of $\mathcal{H}_q.$ Now, from symplectic reduction (linear symplectic reduction) we conclude that

is a Lagrangian submanifold of $T_qN/\left( T_{q} N\right)^{\perp_{\Lambda}} = T_{[q]} N/ \mathcal{F}$, with the symplectic structure induced by $\Omega_q|_{\mathcal{H}}$, which coincides sith the symplectic structure induced by $\Omega_q$, as a quick check shows. □

6. Jacobi structures

Contact and cocontact manifolds are not Poisson manifolds. However, there is still a Lie bracket defined in the algebra of functions, as we will see. This bracket induces what is called a Jacobi manifold. In this section we define and study such structures (see [38, 42] for more details).

Definition 6.1 (Jacobi Manifold). A Jacobi structure on a manifold M is a Lie bracket defined in the algebra of functions $(\mathcal{C}^\infty(M), \{\cdot, \cdot \})$ that satisfies the weak Leibniz rule, that is,

Every Jacobi bracket can be uniquely expressed as

where Λ is a bivector field (called the Jacobi tensor) and E is a vector field. Λ and E satisfy the equalities

where $[\cdot, \cdot]$ is the Schouten–Nijenhuis bracket. Conversely, given a bivector field Λ and a vector field E,

defines a Jacobi bracket if and only if both equalities above hold.

Remark 6.1. It is clear that Poisson manifolds are Jacobi manifolds, taking E = 0.

The Jacobi tensor allows us to define the morphism

Define the Λ-orthogonal of distributions Δ as

We can define the Hamiltonian vector field defined by a function H as

Just like in the Poisson case, we say that a distribution Δ is:

  • i)  
    Isotropic if $\Delta \subseteq \Delta^{\perp_\Lambda}$;
  • ii)  
    Coisotropic if $\Delta^{\perp_\Lambda} \subseteq \Delta$;
  • iii)  
    Legendrian if $\Delta^{\perp_\Lambda} = \Delta$.

These definitions extend naturally to submanifolds.

Remark 6.2. As in the case of Poisson manifolds, a Jacobi structure on a manifold M defines a characteristic distribution ${\cal S}$ as follows: ${\cal S}_x$ is the vector subspace of Tx M generated by the values of all Hamiltonian vector fields at x and the vector field E evaluated at x. This is again an involutive distribution in the sense of Stefan and Sussmann [55, 56], and the leaves of the corresponding foliation are contact manifolds if the leaf has odd dimension, and locally conformal symplectic manifolds, if the leaf has even dimension. The definition of the Jacobi bracket on the leaves follows the same path that in the case of Poisson manifolds.

7. Coisotropic reduction in contact geometry

Contact manifolds are the natural setting for Hamiltonian systems with dissipation, instead of symplectic Hamiltonian systems where the antisymmetry of the symplectic form provides conservative properties [8, 15]. In the Lagrangian picture, contact Lagrangian systems correspond to the so-called Lagrangians depending on the action, and instead of Hamilton's principle, one has to use the so-called Herglotz principle to obtain the dynamics [20].

Definition 7.1 (Contact manifold). A contact manifold is a couple $(M, \eta)$ where M is a $(2n+1)$-dimensional manifold, η is a 1-form and $\eta \wedge ( \mathrm{d}\eta)^n \neq 0$.

In this case we also have Darboux coordinates $(q^i,p_i,z)$ in M [30] such that

We have also have a bundle isomorphism defined as in the cosymplectic case

its inverse $\sharp_\eta = \flat_ \eta^{-1}$, and a couple of natural distributions:

  • i)  
    The horizontal distribution $\mathcal{H}: = \operatorname{Ker} \eta$;
  • ii)  
    The vertical distribution $\mathcal{V}: = \operatorname{Ker} d\eta$.

We can find different types of tangent vectors at a point $q \in M$. Indeed, a tangent vector $v \in T_qM$ will be called

  • i)  
    Horizontal if $v \in \mathcal{H}_q$;
  • ii)  
    Vertical if $v \in \mathcal{V}_q$.

This time, however, we cannot define a canonical Poisson structure since the bivector field

is not a Poisson tensor. In fact,

where $\mathcal{R}$ is the Reeb vector field defined as $\mathcal{R} : = \sharp_{\eta}(\eta)$ (locally $\mathcal{R} = \frac{\partial}{\partial z}$). This is easily seen performing a direct calculation in Darboux coordinates using the local expresion

This defines a Jacobi structure in M taking Λ as above and $E = - \mathcal{R}$ (see section 6). The Jacobi bracket is locally expressed as

The morphism induced by the Jacobi tensor Λ satisfies

Remark 7.1. Although notation between cosymplectic and contact manifolds is similar, they are different in nature. In cosymplectic geometry, we had a closed 1-form θ, and a closed 2-form Ω satisfying the non-degeneracy condition. In contact geometry we have a 1-form η and a (closed) 2-form $d \eta$ which also satisfies the non-degeneracy condition. In each tangent space, these structures will be isomorphic. Indeed, we can always assume that $p_i = 0$ at certain $q \in M$, which would give $\eta = \mathrm{d}z$ in said point. However, they are far from locally isomorphic. In cosymplectic manifolds, the horizontal distribution is involutive, but in contact manifolds is not (this is the key to obtain dissipative dynamics). In section 8 we will enphasize on the differences between these geoemtries by studying the variational principle in contact mechanics.

7.1. Hamiltonian and evolution vector fields as Lagrangian and Legendrian submanifolds

Definition 7.2 (Hamiltonian vector field). Let $H \in \mathcal{C}^\infty(M)$. Define the Hamiltonian vector field of H as

Locally, it has the local expression

The dynamics corresponding to a Hamiltonian vector field XH are determined by

For instance, if we take as Hamiltonian

the equations determined by XH are

which are precisely the equations for the damped harmonic oscillator.

We can define a symplectic structure in TM taking $\Omega_0 : = \flat_\eta^* \Omega_M$, where $\Omega_M$ is the canonical symplectic structure in $T^*M$. In local coordinates $(q^i, p_i, z, \dot{q}^i, \dot{p}_i, \dot{z})$, it has the expression

Definition 7.3 (Gradient vector field). Given a Hamiltonian H on M, define the gradient vector field of H as

Locally, it is given by

We have the following relation between both vector fields

Just like in the previous sections, a vector field $X: M \rightarrow TM$ is locally a gradient vector field if and only if it defines a Lagrangian submanifold in $(TM , \Omega_0)$. The proof is straight-forward, checking that

We can also interpret the Hamiltonian vector field XH as a Lagrangian submanifold of TM, but we need to modify slightly the symplectic form. It is easy to verify that

therefore, taking

we have

It is clear that $\Omega_H$ is a symplectic form. We have proved:

Proposition 7.1. The Hamiltonian vector field $X_H: M \rightarrow TM$ defines a Lagrangian submanifold of the symplectic manifold $(TM, \Omega_H)$.

Now we study evolution vector fields, an important vector field in the application of contact geometry to thermodynamics.

Definition 7.4. Given a Hamiltonian H, we define the evolution vector field as

Locally, the evolution vector field is written

Let us see how we can modify the symplectic form Ω0 in such a way that $\mathcal{E}_H$ defines a Lagrangian submanifold. We have

and thus, $\mathcal{E}_H$ defines a Lagrangian submanifold of $(TM , \widetilde \Omega_H),$ where

We can also interpret Hamiltonian and evolution vector fields as Legendrian submanifolds of a certain contact structure defined on $TM \times \mathbb{R}.$

Definition 7.5. Let $(M, \eta)$ be a contact manifold. Define the contact form on $TM \times \mathbb{R}$ as

where ηc , ηv are the complete and vertical lifts [16]. It is easily checked that $\hat{\eta}$ defines a contact structure [15].

In local coordinates it has the expression:

We have the following [15]:

Proposition 7.2. Let $X_H: M \rightarrow TM$ be a Hamiltonian vector field. Then, the submanifold defined by the immersion

is a Legendrian submanifold of $(TM\times \mathbb{R}, \hat{\eta})$.

Proof. Using the properties of complete and vertical lifts we have

Using lemma 7.2 it will be sufficient to see that $\mathcal{L}_{X_H} \eta = - \mathcal{R}(H).$ This is a straight-forward verification since

 □

7.2. An overview of the use of contact geometry in thermodynamics

A thermodynamical system is characterized by the following variables

corresponding to energy, temperature, entropy, and the generalized pressures and volumes, respectively. We will denote $q^i = V^i$, following the notation used so far. The energy of the system is a function

The first law of thermodynamics can be written as

where $\delta \mathcal{Q}$ and $\delta \mathcal{W}$ are one forms representing the heat and work, respectively. We will assume that these forms have local expressions

for some functions $P_i, T$ that physically represent the conjugate variables to qi (pressure, if qi represented volume), and S (temperature, if S was the entropy), respectively. Then, the first law of thermodynamics reads

If we were studying an isolated system, energy would be conserved, that is

or, dividing by T,

Identifying $p_i = P_i/T$, after a change of variables if necessary, we obtain

Defining

the canonical contact form in $T^\ast Q \times \mathbb{R},$ we conclude that isolated processes take place in Legendrian submanifolds of $T^\ast Q \times \mathbb{R},$ motivating their study.

It turns out that the integral curves of the evolution vector field $\mathcal{E}_H$ can be interpreted as an isolated system

Proposition 7.3. The integral curves of $\mathcal{E}_H$ satisfy

Furthermore, locally

Proof. Indeed, by definition we know that

Therefore,

Now, the last part is a consequence of the equality

 □

For more details on the subject, we refer to [4, 47, 48, 51, 52], (see also two modern approaches [9, 10]).

7.3. Coisotropic reduction

Coisotropic reduction in contact manifolds has been developed in [15] (see also [41, 57]).

The following definition will result useful. Given a subspace $\Delta_q \subseteq T_qM$, we define the $\mathrm{d}\eta$-orthogonal complement as

Proposition 7.4. Let $\Delta_q \subseteq T_qM$ be a subspace. Then

Furthermore, if $\mathcal{R}(q) \in \Delta_q$ or $\Delta_q \subseteq \mathcal{H}_q$, the equality holds.

Proof. Let $v \in \Delta_q^{\perp_{d\eta}}\cap \mathcal{H}_q$ and take $\alpha : = i_v \mathrm{d} \eta$. It is clear that $\alpha \in \Delta_q^0$. We will prove that $\sharp_\Lambda(-\alpha) = v$. Indeed, for $\beta \in T^*_qM$, since $b_\eta(v) = \alpha$ (as a direct calculation shows), we have

that is, $v = \sharp_\Lambda(-\alpha)$.

Now, if $\mathcal{R}(q) \in \Delta_q$, we compare dimensions. Since $\operatorname{Ker} \sharp_\Lambda = \langle \eta\rangle$ and $\eta \not \in \Delta_q^0$, we have

Furthermore, $\Delta^{\perp_{\mathrm{d}\eta}}\cap \mathcal{H}_q$ has the same dimension, since

This latter is just the symplectic complement in $(\mathcal{H}_q, \mathrm{d}\eta |_{\mathcal{H}})$ and hence,

Now, if $\Delta_q \subseteq \mathcal{H}_q$, $\eta \in \Delta_q^0$ and, thus,

Since $\Delta_q^{\perp_{\mathrm{d}\eta}} \cap \mathcal{H}_q$ is just the symplectic complement of $\Delta_q$ we have

 □

This proposition allows us to characterize Legendrian submanifolds:

Lemma 7.1. If $L\hookrightarrow M$ is a Legendrian submanifold, then L is horizontal and $\operatorname{dim} L = n$ (where $\operatorname{dim} M = 2n+1$). Furthermore, if L is horizontal and isotropic (or coisotropic) with $\operatorname{dim} L = n$, L is Legendrian.

Proof. Since $\sharp_\Lambda$ takes values in $\mathcal{H}$, it is clear that every Legendrian submanifold is horizontal. Since L is horizontal,

From the previous equation and using that $T_qL^{\perp_\Lambda} = T_qL$, we deduce that $\operatorname{dim} L = \operatorname{dim} M - \operatorname{dim} L - 1$. This implies $\operatorname{dim} L = n$. The last property is easily seen via a direct comparison of dimensions. □

We will also need characterization of isotropic submanifolds in contact geometry:

Lemma 7.2. A submanifold $N \hookrightarrow M$ is isotropic if and only if $i^* \eta = 0$.

Proof. Necessity is clear, since $\sharp_\Lambda$ takes values in $\mathcal{H}$. Now suppose that N is horizontal. We have $i^*\eta = 0$ and thus, $i^* \mathrm{d}\eta = 0$. This implies that $T_qN \subseteq T_qN^{\perp_{\mathrm{d}\eta}} \cap \mathcal{H}_q \subseteq T_qN^{\perp_\Lambda},$ using proposition 7.4. □

Proposition 7.5. Let $i:N \hookrightarrow M$ be a coisotropic submanifold such that $\mathcal{R}(q) \in T_qN$ for every $q\in N$ or $T_qN \subseteq \mathcal{H}_q$ for every $q \in N$. Define $\eta_0: = i^*\eta$. Then

Proof. Let $q\in N$. Proposition 7.4 implies that

But, since Tq N is coisotropic, then it is just $\operatorname{Ker} \mathrm{d}\eta_0 \cap \operatorname{Ker} \eta_0.$ □

We then have the following result:

Proposition 7.6. Let $i: N \hookrightarrow M$ be a coisotropic subamnifold such that $\mathcal{R}(q) \in T_qN$ for every $q \in N$ or $T_qN \subseteq \mathcal{H}_q$ for every $q \in N$. Then, the distribution $TN^{\perp_\Lambda}$ defined by $q \mapsto T_qN^{\perp_\Lambda}$ is involutive.

Proof. Denote $\eta_0: = i^*\eta$ and let $X, Y$ be vector fields along N taking values in $TN^{\perp_\Lambda}$. Proposition 7.5 implies that

It suffices to check that

Indeed, taking Z an arbitrary vector field in N, we have

where we have used that $X,Y \in \operatorname{Ker} \mathrm{d} \eta_0.$ In a similar way we obtain

that is, $[X,Y] \in \operatorname{Ker} \mathrm{d}\eta_0 \cap \operatorname{Ker} \eta_0 = TN^{\perp_\Lambda}.$ □

7.4. Vertical coisotropic reduction

We will restrict the study to vertical submanifolds, that is, submanifolds satisfying $\mathcal{R}(q) \in T_qN$, for every $q \in N$. Notice that if N is a coisotropic vertical submanifold, the distribution $(TN)^{\perp_\Lambda}$ is regular of rank

Theorem 7.1 (Vertical coisotropic reduction in the contact setting). Let $(M, \eta)$ be a contact manifold and $i :N \hookrightarrow M$ be a coisotropic submanifold such that $\mathcal{R}(q) \in T_qN$ for every $q \in N$. If the space of all leaves $N /\mathcal{F}$ admits a manifold structure such that the projection $\pi: N \rightarrow N/\mathcal{F}$ is a submersion, then there exists a unique 1-form ηN in $N/\mathcal{F}$ such that $\pi^*\eta_N = i_*\eta $ and $(N/\mathcal{F},\eta_N)$ is a contact manifold.

Proof. Denote $\eta_0: = i^*\eta$. Uniqueness is clear from the imposed relation since it forces us to define

It only remains to check well-definedness and that it defines a contact manifold. That this definition does not depend on the chosen representative vector is clear since a vector tangent to the distribution is necessarily in the kernel of η. Furthermore, if X is a vector field tangent to the distribution $TN^{\perp_\Lambda}$, proposition 7.5 implies

since $X \in \operatorname{Ker} \mathrm{d}\eta_0 \cap \operatorname{Ker} \eta_0$.

To check that it is a contact manifold, we calculate the dimension of $N /\mathcal{F}$. We have that $\operatorname{rank} \, (T_qN)^{\perp_\Lambda} = \operatorname{dim} M - \operatorname{dim} N$. We conclude, taking $k: = \operatorname{dim} N$, that

and therefore, $(N/\mathcal{F}, \eta_N)$ is a contact manifold if and only if

Since π is a submersion, this is equivalent to $\eta_0 \wedge ( \mathrm{d}\eta_0)^{k-n-1}\neq0$. This is straightforward using proposition 7.5. □

7.4.1. Projection of Legendrian submanifolds.

Now we check that the image of a Legendrian submanifold $L \hookrightarrow M$ under the projection $\pi: N \rightarrow N/\mathcal{F}$ is again a Legendrian submanifold.

Proposition 7.7. Let $L \hookrightarrow M$ be a Legendrian submanifold such that L and N have clean intersection. If $ L_N: = \pi(L \cap N)$ is a submanifold of $N/ \mathcal{F}$, then LN is Legendrian.

Proof. It suffices to check that LN is horizontal, isotropic and $\operatorname{dim} L_N = \operatorname{dim} N - n - 1$ using lemma 7.2. Since L is horizontal, LN is horizontal and thus, LN is isotropic.

Comparing dimensions, we have

Equation (10)

Now, since $(T_qL \cap T_qN^{\perp_\Lambda})^{\perp_\Lambda} = T_qL + (T_qN^{\perp_\Lambda})^{\perp_\Lambda}$ and $(T_qL \cap T_qN)^{\perp_\Lambda}$ is horizontal, we have

Equation (11)

Using $\operatorname{dim}(T_qL \cap T_qN^{\perp_\Lambda}) = \operatorname{dim}(T_qL + T_qN) - 1$ and substituting (11)) in (10), we obtain

 □

7.5. Horizontal coisotropic reduction

We will restrict the study to horizontal coisotropic submanifolds $N \rightarrow M$, that is, manifolds satisfying $T_qN \subseteq \mathcal{H}_q$ for every $q\in N$.

Remark 7.2. Notice that in this case reduction is trivial, since the only coisotropic horizontal submanifolds of a contact manifold are those that are Legendrian. This would imply

making the resulting manifold trivial.

Given an arbitrary coisotropic submanifold $N \hookrightarrow M$, we cannot guarantee the well-definedness of the 2-form in the quotient $N/\mathcal{F}$ (actually, in the contact setting, we cannot even guarantee the integrability of $TN^{\perp_\Lambda}$) so this time (referring to horizontal reduction in cosymplectic geometry) we cannot obtain a foliation of N in symplectic leaves, since $TN \cap \mathcal{H}|N$ is not integrable in the general setting.

Remark 7.3. The triviality of this case makes the projection of Lagrangian submanifolds trivial.

8. An interlude: different geometries provide different dynamics

In this section we will show how different geometric structures on the same phase space can provide different dynamics for the same Lagrangian or Hamiltonian function. This fact explains the different equations of motion between the cosymplectic case (time-dependent Lagrangians) and the contact case (action-dependent Lagrangians). At the end of the section we will also see that Hamilton's principle can be generalised to the so-called Herglotz principle, which gives a new way to obtain these different dynamics.

8.1. The Lagrangian picture

In order to distinguish the different dynamics discussed in this paper and their physical nature, it is useful to recall their Lagrangian formulation.

Assume that $L : TQ \longrightarrow \mathbb{R}$ is a Lagrangian function, where Q is a n-dimensional configuration manifold. Then, $L = L(q^i, \dot{q}^i)$, where $(q^i)$ are coordinates in Q and $(q^i, \dot{q}^i)$ are the induced bundle coordinates in TQ (positions and velocities). We will assume that L is regular, that is, the Hessian matrix

is regular. Using the canonical endomorphism S on TQ locally defined by

One can construct a 1-form λL defined by

and the 2-form

Then, ωL is symplectic if and only if L is regular.

In that case, we have the corresponding vector bundle isomorphism

and the Hamiltonian vector field

defined by

where $E_L = \Delta(L) -L$ is the energy, and $\Delta = \dot{q}^i \, \frac{\partial}{\partial \dot{q}^i}$ is the Liouville vector field on TQ.

The vector field ξL (the Euler–Lagrange vector field) is a second order differential equation, that is, its integral curves are just the tangent lifts of its projections to Q. These projections are called the solutions of ξL , and satisfy the usual Euler–Lagrange equations

Equation (12)

Next, we recall here the geometric formalism for time-dependent Lagrangian systems. In this case, we also have a regular Lagrangian $L : TQ \times \mathbb{R} \longrightarrow \mathbb{R}$, and we consider the cosymplectic structure given by the pair $(\Omega_L, \mathrm{d}z)$ (in this case z represents time), where

It is esay to check that L is regular if and only if

In that case, we have a cosymplectic structure and, defining $(W^{ij})$ to be the inverse matrix of

The Reeb vector field of the cosymplectic manifold $(TQ, \Omega_L, \mathrm{d}z)$ is locally given by

Recall that we had the vector bundle isomorphism

which gives the vector fields defined in section 5. In particular, we have the evolution vector field

where EL is the energy of the system,

Now, if $(q^i(t), \dot{q}^i(t), z(t))$ is an integral curve of ${\mathcal E}_L$, then its projection to Q satisfies the usual Euler–Lagrange equations

Equation (13)

since $z = t + constant$.

Finally, let $L : TQ \times \mathbb{R} \longrightarrow \mathbb{R}$ be a Lagrangian function, $L = L(q^i, \dot{q}^i, z)$, where z is a global coordinate on $\mathbb{R}$, representing action.

We will assume that L is regular, that is, the Hessian matrix

is regular. So, we construct a 1-form λL defined by

where now S and $S^*$ are the natural extension of S.

Now, the 1-form

is a contact form on $TQ \times \mathbb{R}$ if and only if L is regular, and then

The corresponding Reeb vector field is

The energy of the system is defined as in the precedent cases by

where $\Delta = \dot{q}^i \, \frac{\partial}{\partial \dot{q}^i}$ is the Liouville vector field on TQ extended in the usual way to $TQ \times \mathbb{R}$. The vector bundle isomorphism of contact manifolds defined in section 7 is given by

We shall denote its inverse by ${\sharp_{\eta_L}} = ({\flat_{\eta_L})}^{-1}$.

Denote by ξL the unique vector field defined by the equation

Equation (14)

Note that this is precisely the Hamiltonian vector field of EL defined in section 7. A direct computation from equation (14) shows that, if $(q^i(t), \dot{q}^i(t), z(t))$ is an integral curve of $\bar{\xi}_L$, we obtain

with an additional equation $\dot{z} = L$. These equations correspond to the generalized Euler–Lagrange equations considered by G Herglotz in 1930.

Equation (15)

Notice that Herglotz equations depend on the action, so this type of Lagrangians are called in physics dependent on the action.

8.2. The Herglotz principle

In order to give additional differences between the usual Hamilton principle and Herglotz principle, it is interesting to recall briefly the last one and how it is a natural generalization of the former one.

Let $L:TQ \times \mathbb{R} \to \mathbb{R}$ be a Lagrangian function.

Fix $q_1,q_2 \in Q$ and an interval $[a,b] \subset \mathbb{R}$. We denote by $\Omega(q_1,q_2, [a,b]) \subseteq({\cal C}^\infty([a,b]\to Q))$ the space of smooth curves ξ such that $\xi(a) = q_1$ and $\xi(b) = q_2$. This space has the structure of an infinite dimensional smooth manifold whose tangent space at ξ is given by the set of vector fields over ξ that vanish at the endpoints, that is,

Equation (16)

We will consider the following maps. Fix an initial action $c \in \mathbb{R}$. Let

Equation (17)

be the operator that assigns to each curve ξ the curve $\mathcal{Z}(\xi)$ that solves the following ODE:

Equation (18)

Now we define the action functional as the map which assigns to each curve the solution to the previous ODE evaluated at the endpoint:

Equation (19)

that is, $\mathcal{A} = ev_b \circ \mathcal{Z}$, where $ev_b: \zeta \mapsto \zeta(b)$ is the evaluation map at b. We have

Theorem 8.1. Let $L: TQ \times \mathbb{R} \to \mathbb{R}$ be a Lagrangian function and let $\xi\in \Omega(q_1,q_2, [a,b])$ be a curve in Q. Then, $(\xi,\dot\xi, \mathcal{Z}(\xi))$ satisfies the Herglotz's equations if and only if ξ is a critical point of $\mathcal{A}$.

Remark 8.1. This theorem generalizes Hamilton's Variational Principle. In the case that the Lagrangian is independent of the $\mathbb{R}$ coordinate (i.e. $L(q^i,\dot{q}^i,z) = \hat{L}(q^i, \dot{q}^i)$) the contact Lagrange equations reduce to the usual Euler–Lagrange equations. In this situation, we can integrate the ODE of (19) and we get

Equation (20)

that is, the usual Euler–Lagrange action up to a constant.

8.3. The Legendre transformation and the Hamiltonian picture

Given a Lagrangian function $L : TQ \times \mathbb{R} \longrightarrow \mathbb{R}$ we can define the Legendre transformation

given by

where

A direct computation shows that

where λQ is the canonical Liouville form on $T^*Q$, here extended to the product manifold $ T^*Q \times \mathbb{R}$. If the Lagrangian does not depend on time, then the Legendre transformation reduces to a bundle morphism

and then the pull-back of the canonical symplectic form ωQ on $T^*Q$ is just ωL . In addition, the energy EL corresponds via the Legendre transformation to the Hamiltonian energy H such that

Since the corresponding geometric structures (symplectic, cosymplectic and contact) on both sides are preserved by the Legendre transformation, one concludes that corresponding dynamics are $\mathbb{F}L$-related (see [14] for the details).

9. Coisotropic reduction in cocontact geometry

Cocontact manifolds have been introduced in [13] just to provide a setting for dissipative systems which also depend on time. In geometric terms, we are combining cosymplectic and contact structures.

Definition 9.1 (Cocontact manifold). A cocontact manifold is a triple $(M,\theta, \eta)$, where M is a $(2n + 2)$-dimensional manifold, θ is a closed 1-form, η is a 1-form and, $\theta \wedge \eta \wedge (\mathrm{d}\eta)^n \neq 0$ is a volume form.

The bundle isomorphism in this case is defined as

and its inverse is denoted by $\sharp_{\theta,\eta} = \flat_{\theta,\eta}^{-1}.$

In cocontact geometry there exists as well a set of canonical coordinates $(q^i,p_i,z,t)$, which will be called Darboux coordinates, such that

We can define as well the Reeb vector fields as

which can be expressed locally as

We also have vertical and horizontal distributions:

  • i)  
    The z-horizontal distribution, $\mathcal{H}_z : = \operatorname{Ker} \eta$;
  • ii)  
    The t-horizontal distribution, $\mathcal{H}_t : = \operatorname{Ker} \theta$;
  • iii)  
    The tz-horizontal distribution $\mathcal{H}_{tz} : = \mathcal{H}_t \cap \mathcal{H}_z$;
  • iv)  
    The t-vertical distribution, $\mathcal{V}_t : = \langle \mathcal{R}_t \rangle$;
  • v)  
    The z-vertical distribution, $\mathcal{V}_z : = \langle \mathcal{R}_z \rangle.$

9.1. Hamiltonian vector fields as Lagrangian and Legendrian submanifolds

Just like in previous sections, define the gradient vector field of certain Hamiltonian $H\in \mathcal{C}^\infty(M)$ as

Locally, the gradient vector field is expressed:

We can define a symplectic structure in TM taking

where $\Omega_M$ is the canonical symplectic form on the cotangent bundle. In the induced coordinates $(q^i, p_i, z, \dot{q}^i, \dot{p}_i, \dot{z}),$ Ω0 takes the form

It is easy to verify that a vector field $X: M \rightarrow TM$ is a locally gradient vector field if and only if it defines a Lagrangian submanifold in $(TM , \Omega_0).$

Definition 9.2 (Hamiltonian vector field). Given a Hamiltonian H on M, define its Hamiltonian vector field as

The Hamiltonian vector field has the local expression

Integral curves of this vector field satisfy

which are equivalent to the time-dependent version of the contact equations:

In general, XH does not define a Lagrangian submanifold of $(TM, \Omega_0)$; but, just like in the cosymplectic and contact scenario, we can achieve this by modifying the symplectic form. Indeed, since

defining

we have that XH defines a Lagrangian submanifold of $(TM, \Omega_H).$

Now we interpret the Hamiltonian vector field XH as a Legendrian submanifold of $TM \times \mathbb{R} \times \mathbb{R}$ with the cocontact structure given by the forms

where (s, e) are the parameters in $\mathbb{R} \times \mathbb{R}$. In local coordinates $(q^i, p_i, z, t, \dot{q}^i, \dot{p}_i, \dot{z}, \dot{t}, s,e)$, these forms have the expression:

It is easy to see that these forms define a cocontact structure. Now, given a vector field $X: M \rightarrow TM$ and two functions $f, g$ on M, define

Applying the properties of complete and vertical lifts, namely $X^*\alpha^c = \mathcal{L}_X(\alpha)$, we have

and

Proposition 9.1. Let H be a Hamiltonian on M. Then $X_H \times \mathcal{R}_z(H) \times 0$ defines a Legendrian submanifold of $(TM \times \mathbb{R} \times \mathbb{R}, \widetilde \theta, \widetilde \eta).$

Proof. Using the observation above and lemma 9.1, it is sufficient to observe that

 □

9.2. Coisotropic reduction

A cocontact manifold is also a Jacobi manifold defining

and thus, we have the Λ -orthogonal and the corresponding definitions of isotropic, coisotropic or Legendrian submanifolds and distributions.

Notice that $\mathcal{H}_t$ is an integrable distribution and that each leaf of its foliation inherits a contact structure. Indeed, $\mathcal{H}_t$ is the characteristic distribution defined by the Jacobi structure, $\mathcal{S}$.

Now we give a symplectic interpretation of the Λ-orthogonal. Notice that the restriction of $\mathrm{d}\eta$ to $\mathcal{H}_{tz}$ defines a symplectic structure on the distribution. Denote by $\perp_{\mathrm{d}\eta|}$ its symplectic orthogonal. The Λ-orthogonal is just the symplectic orthogonal of the intersection with $\mathcal{H}_{tz}$.

Proposition 9.2. Given a distribution Δ on a cocontact manifold $(M, \eta, \theta)$,

Proof. We check one inclusion and compare dimensions:

Let $\alpha \in \Delta^0_q$ and $u \in \Delta^0_q \cap (\mathcal{H}_{tz})_q.$ We will see that $d\eta_q(u,\sharp_\Lambda(\alpha)) = 0.$ Indeed,

that is, $\Delta^{\perp_\Lambda} \subseteq (\Delta \cap (\mathcal{H}_{tz})_q)^{\perp_{\mathrm{d}\eta|}}.$

Now we compare both dimensions. Let $k: = \operatorname{dim} \Delta, r_q : = \operatorname{dim}(\Delta_q^0 \cap\langle \theta_q, \eta_q\rangle).$ Since

we have

which implies that

It only remains to observe that

and that

 □

Now we can give a characterization of Legendrian submanifolds:

Lemma 9.1.  $i:L \rightarrow M$ is a Legendrian submanifold ($(TL)^{\perp_\Lambda} = TL$) if and only if $\operatorname{dim} L = n$ and

Proof. Necessity is clear, since L is necessarily horizontal. Sufficiency follows from $i^* \mathrm{d}\eta = 0$ which, together with $\operatorname{dim} L = n$, implies that Tq L is a Lagrangian submanifold of $(\mathcal{H}_{tz})_q$, for every $q\in L$. Using proposition 9.2, we have the equivalence. □

This allows us to express the Λ-orthogonal complement of a coisotropic distribution in a more convenient way:

Corollary 9.1. Let Δ be a coisotropic distribution on M ($\Delta^{\perp_\Lambda}\subseteq \Delta$), then

where η0 and θ0 are the restrictions of η and θ to Δ, respectively.

Corollary 9.2. Let $i: N \hookrightarrow M$ be a coisotropic submanifold of $(M,\theta, \eta).$ Then the distribution $TN^{\perp_\Lambda}$ is involutive.

Proof. Denote $\eta_0 : = i^*\eta$, $\theta_0 : = i^*\theta$ and let $X,Y$ be vector fields on N tangent to the distribution $TN^{\perp_\Lambda}$ and Z be an arbitrary tz-horizontal vector field on N. Using corollary 9.2, we only need to check that $[X,Y] \in \operatorname{Ker} \mathrm{d}\eta_0|_{\Delta \cap \mathcal{H}_{tz}} \cap \operatorname{Ker} \eta_0 \cap \operatorname{Ker} \theta_0.$ Indeed, expanding the expressions $0 = \operatorname{d}^2 \eta_0(X,Y,Z)$, $0 = \mathrm{d}\eta_0(X,Y)$, $0 = \mathrm{d}\theta_0(X,Y)$; we obtain

 □

Now, given a coisotropic submanifold $N \hookrightarrow M$, since $TN^{\perp_\Lambda}$ is involutive, it provides a maximal foliation, $\mathcal{F}$. We assume that $N/\mathcal{F}$ inherits a manifold structure such that the canonical projection $\pi: N \rightarrow N/\mathcal{F}$ is a submersion.

Just like in the previous cases, for the well-definedness and non-degeneracy of the forms in the quotient, we need to restrict the coisotropic submanifolds we are studying. Consequently, we will say that a submanifold $N \hookrightarrow M$ is:

  • i)  
    t-vertical (resp. z-vertical) if $\mathcal{V}_t\subseteq TN$ (resp. $\mathcal{V}_z\subseteq TN$).
  • ii)  
    tz-vertical, if it is both t-vertical and z-vertical.
  • iii)  
    t-horizontal (resp. z-horizontal) if $\mathcal{H}_t \subseteq TN$ (resp. $\mathcal{H}_z \subseteq TN$).
  • iv)  
    tz-horizontal if it is both t-horizontal and z-horizontal, that is, if $TN \subseteq \mathcal{H}_{tz}$.

9.3.  $tz-$vertical reduction

Let $i: N \hookrightarrow M$ be a tz-vertical submanifold. It is easy to check that under these conditions

and that $(TN)^{\perp_\Lambda}$ is a regular distribution of rank

Theorem 9.1 (tz-vertical coisotropic reduction). Let $i: N \hookrightarrow M$ be a tz-vertical submanifold of a cocontact manifold $(M, \theta, \eta).$ Denote by $\mathcal{F}$ the maximal foliation induced by the integrable distribution $TN^{\perp_\Lambda}$ on N. If $N/\mathcal{F}$ admits a manifold structure such that the canonical projection $\pi: N \rightarrow N/\mathcal{F}$ defines a submersion, then there exists unique forms $\theta_N, \eta_N$ on $N/\mathcal{F}$ such that $(N/\mathcal{F}, \theta_N, \eta_N)$ defines a cocontact structure and

Before proving the theorem, let us calculate the dimension of the quotient. Let $k + 2: = \operatorname{dim} N$. We have

and, therefore,

Proof. Uniqueness is clear, since π is a submersion. We only need to check the well-definedness taking

Independence of the vector is clear, using proposition 9.2. For independence on the point, let X be a vector field on N tangent to the distribution. It is easy to check that

and thus, well-definedness follows. For non-degeneracy, it is enough to proof that

This follows easily from $TN^{\perp_\Lambda} = \operatorname{Ker} \mathrm{d}\eta_0 \cap \operatorname{Ker} \eta_0 \cap \operatorname{Ker} \theta_0.$ □

9.3.1. Projection of Legendrian submanifolds

Proposition 9.3. Let $L \hookrightarrow M$ be a Legendrian submanifold and $i: L \hookrightarrow M$ be a tz-vertical coisotropic submanifold. If L and N have clean intersection and $L_N = \pi(L \cap N)$ is a submanifold in $N/\mathcal{F}$, LN is Legendrian.

Proof. Using lemma 9.1, LN is clearly isotropic. Now we need to check that $\operatorname{dim} L_N = k - n$, given that $\operatorname{dim} N/\mathcal{F} = 2(k-n) + 2.$ We have

Equation (21)

Furthermore, since $(TL \cap (TN)^{\perp_\Lambda}) = TL + (TN^{\perp_\Lambda})^{\perp_\Lambda}$ and $TL \cap (TN)^{\perp_\Lambda}$ is tz-horizontal,

Equation (22)

Now, using the Grassman formula:

Equation (23)

Equation (24)

Equation (25)

From (22) and (25) we obtain

Equation (26)

Substituting in (21) yields $\operatorname{dim} \pi(L \cap N) = k - n.$ □

9.4.  $t-$vertical, $z-$horizontal reduction

Suppose $i: N \hookrightarrow M$ is a t-vertical and z-horizontal coisotropic submanifold. This time we have

since η0 = 0 implies $d\eta _0 = 0$. We conclude that $(TN)^{\perp_\Lambda} = TN \cap \mathcal{H}_{tz}$, which implies that

This means that reduction is trivial, leaving the trivial cosymplectic submanifold of dimension 1:

Theorem 9.2. Let $i: N \hookrightarrow M$ be a t-vertical, z-horizontal coisotropic submanifold of a coconatct manifold $(M , \theta, \eta)$. Denote by $\mathcal{F}$ the maximal foliation defined by the distribution $(TN)^{\perp_\Lambda}$. If $N/\mathcal{F}$ has a manifold structure such that the canonical projection $\pi: N \rightarrow N/\mathcal{F}$ defines a submersion, then $N/\mathcal{F}$ is one-dimensional and there exists and unique volume form θN on $N/\mathcal{F}$ such that

Remark 9.1. Given the triviality of reduction in the t-vertical and z-horizontal case, projection of Legendrian submanifolds in M will always result in 0-dimensional Lagrangian submanifolds in $N/\mathcal{F}$.

9.5.  $z-$vertical, $t-$horizontal reduction

Let $i: N \rightarrow M$ be a z-vertical and t-horizontal coisotropic submanifold. It is easy to check that this time we have the equality

Since $\mathcal{H}_t$ is integrable, we have that coisotropic reduction of N is actually happening in one of the leaves of the foliation that inhertits a contact structure from the cocontact structure. We conclude, from theorem 7.1:

Theorem 9.3. Let $i: N \rightarrow M$ be a z-vertical and t-horizontal coisotropic submanifold of a cocontact manifold $(M, \theta, \eta)$. Denote by $\mathcal{F}$ the maximal foliation on N defined by the distribution $TN^{\perp_\Lambda}$. If $N/\mathcal{F}$ has a manifold structure such that the canonical projection $\pi: N \rightarrow N/\mathcal{F}$ defines a submersion, then there exists an unique form ηN such that $(N/\mathcal{F}, \eta_N)$ is a contact manifold and

9.5.1. Projection of Legendrian submanifolds

Proposition 9.4. Let $L \hookrightarrow M$ be a Legendrian submanifold. If L and N have clean intersection and $L_N = \pi(L \cap N)$ is a submanifold in $N/\mathcal{F}$, LN is Legendrian in $(N/\mathcal{F}, \theta_N)$.

Proof. It is clearly horizontal and, therefore, using lemma 7.2, it is isotropic. Now, supposing $k = \operatorname{dim} N$, we only need to check that

since $\operatorname{dim} N/\mathcal{F} = 2(k - n) + 1.$ This is straight-forward, following the same steps given in proposition 9.3. □

9.6.  $tz-$horizontal reduction

Let $i: N \hookrightarrow M$ be a tz-horizontal coisotropic submanifold. Since η0 = 0, $\mathrm{d}\eta_0 = 0$ and θ0 = 0, we have

wich implies that

leaving a trivial symplectic manifold, having as many points as path components of N. This means that if $N/\mathcal{F}$ admits a manifold structure, it will be a symplectic manifold.

Remark 9.2. Proceeding as in the previous cases we immediately obtain that the projection of Legendrian submanifolds is trivial.

10. Coisotropic reduction in stable Hamiltonian structures

There were several attempts to combine cosymplectic and contact structures. The first one is due to Albert [3], using a combination of a 1-form and a 2-form; however, the setting is not useful for us since the lack of integrability. The second attempt in this direction is due to Acakpo [2], which is studied in this section.

Definition 10.1 (Stable Hamiltonian structure). A stable Hamiltonian structure (SHS) is a triple $(M, \omega, \lambda)$ where M is a $2n + 1$ dimensional manifold, ω is a closed 2-form and λ is a 1-form such that

There exists, just like in the previous cases, a natural isomorphism

and its inverse $\sharp_{\lambda, \omega} : = \flat_{\lambda, \omega} ^{-1}.$ Let us perform some calculations in coordinates. Since $\mathrm{d}\omega$ is closed and of constant range 2n, around any point, there exists coordinates $(q^i, p_i, z)$ such that

(see [30]). In this coordinate chart λ will have an expression of the form

Since

we conclude that c ≠ 0. Let $\varphi_t(q^i, p_i, z)$ be the (local) flow of the vector field $\displaystyle{\frac{1}{c} \frac{\partial }{\partial z}}$. Fix some value z0, and define the map

It is clear that this defines a local diffeomorphism. Take the new set of coordinates to be

We have

Therefore, in the new coordinate chart,

We conclude:

Proposition 10.1. Around every point of M there exists a coordinate chart $(q^i, p_i, z)$ such that

We call these coordinates Darboux coordinates.

In Darboux coordinates, the condition $\operatorname{Ker} \text{d}\omega \subseteq \text{d}\lambda$ translates to

Also, the musical isomorphisms take the expression:

Imitating the definitions in the contact case, we can define a bivector field on M as

and the morphism

with the induced Λ-orthogonal complement for distributions

In coordinates $(q^i, p_i, z)$ the bivector field Λ takes the local form:

We also have the distributions

  • i)  
    $\mathcal{H}_q: = \operatorname{Ker} \lambda_q,$
  • ii)  
    $ \mathcal{V}_q: = \operatorname{Ker} \omega_q,$

and the Reeb vector field $\mathcal{R}_q : = \sharp_{\lambda, \omega}(\lambda_q).$ Locally,

Equation (27)

Equation (28)

Equation (29)

A natural question to ask is wether the bivector field Λ arises from a Jacobi bracket. We have

Taking an arbitrary vector field

$(\Lambda, E)$ defines a Jacobi structure if and only if

It is easily checked that the first equality holds when

Equation (30)

Equation (31)

Equation (32)

Equation (33)

Equation (34)

for certain local unique funtion f. It is easy to check that this relations translate intrinsically to

Now, let us compute $[E, \Lambda]$ for $E = f \displaystyle{\frac{\partial }{\partial z}}.$

Therefore, $[E, \Lambda] = 0$ if and only if

Equation (35)

This is easily seen to be equivalent to

We have concluded the following:

Proposition 10.2. The bivector field Λ arises from a Jacobi structure if and only if there exists some $f \in \mathcal{C}^\infty(M)$ such that

And, in that case, the Jacobi structure is defined by the pair $(\Lambda, f \mathcal{R})$.

Remark 10.1. Notice that we recover the cosymplectic scenario when f = 0 and the contact scenario when f = 1 (because the definition of Λ in contact geometry is the opposite of the definition we gave in SHS).

Let us return to the study of coisotropic reduction. It is easy to see that ω induces a symplectic form in $\mathcal{H}$, $\omega|_\mathcal{H}$. This induces the symplectic orthogonal for $\Delta_q \subset \mathcal{H}_q$:

A distribution Δ in M will be called:

  • i)  
    Isotropic if $\Delta \subseteq \Delta^{\perp_\Lambda};$
  • ii)  
    Coisotropic if $\Delta^{\perp_\Lambda} \subseteq \Delta;$
  • iii)  
    Lagrangian if $\Delta^{\perp_\Lambda} = \Delta^{\perp_\Lambda} \cap \mathcal{H}.$

We have the following equality:

Proposition 10.3. Let Δ be a distribution on M. Then

Proof. The proof follows the same lines as that of proposition 5.3 □

Just like in previous sections, we say that a Lagrangian submanifold $L \hookrightarrow M$ is horizontal if $T_qL \subseteq \mathcal{H}_q\, \forall q \in L$ and say that it is non-horizontal if $T_qL \not \subseteq \mathcal{H}_q \, \forall q \in L.$ We have the following characterization:

Lemma 10.1. Let $L \hookrightarrow M$ be an isotropic (or coisotropic) submanifold. We have

  • i)  
    If L is horizontal and $\operatorname{dim} L = n$, then L is Lagrangian.
  • ii)  
    If L is non-horizontal and $\operatorname{dim} L = n+1$, then L is Lagrangian.

Proof. The proof is similar to the proof of lemma 5.2, since we only need to check the condition at each tangent space. □

Now, given a coisotropic submanifold $N \hookrightarrow M$ (that is, $(T_qN)^{\perp_\Lambda} \subseteq T_qN$), the distribution $(TN)^{\perp_\Lambda}$ is not necessarily integrable and we shall assume it in what follows:

10.1. Gradient and Hamiltonian vector fields as Lagrangian submanifolds

We can define a symplectic structure on TM taking

where $\Omega_M$ is the canonical symplectic form on $T^*M$. In Darboux coordinates it has the expression:

Definition 10.2 (Gradient vector field). Given a Hamiltonian $H \in \mathcal{C}^ \infty(M),$ define the gradient vector field of H as

In Darboux coordinates, the gradient vector field is written

It is easily checked that $X: M \rightarrow TM$ is locally a gradient vector field if and only if X(M) is a Lagrangian submanifold of $(TM, \Omega_0)$. Indeed, we have the equality

When Λ comes from a Jacobi bracket on M, that is, when

for some function f on M, we have the Hamiltonian vector field of the Jacobi structure $(\Lambda, f \mathcal{R})$:

In Darboux coordinates it has the expression:

Let us interpret the Hamiltonian vector field as a Lagrangian submanifold of TM, with some symplectic form. First, observe that

Therefore, defining the symplectic form

we have that XH defines a Lagrangian submanifold of $(TM, \Omega_H).$

10.2. Vertical coisotropic reduction

Theorem 10.1 (Vertical coisotropic reduction in stable Hamiltonian structures). Let $i: N \hookrightarrow M$ be a vertical coisotropic submanifold such that $(TN)^{\perp_\Lambda}$ defines an integrable distribution. Let $\mathcal{F}$ be the set of leaves and suppose that $N/\mathcal{F}$ admits a manifold structure such that the canonical projection $\pi: N \rightarrow N/\mathcal{F}$ defines a submersion. If $i^* \mathrm{d}\lambda = 0$ in $TN \cap \mathcal{H}$, then $N/\mathcal{F}$ admits an unique stable Hamiltonian system structure $(\omega_N, \lambda_N)$ such that $\pi^*\omega_N = i^*\omega$ and $\pi^*\lambda_N = i^*\lambda$. The following diagram summarizes the situation:

Standard image High-resolution image

Proof. The proof is similar to the the proof of theorem 5.1. Asking $i^*d\lambda = 0$ is necessary to guarantee the well-definedness of λN in the quotient using

where $\lambda_0 = i^*\lambda$. It would only remain to check that

Indeed, since $\operatorname{Ker} \omega_N = \langle \mathcal{\mathcal{R}_N} \rangle$, and $\mathcal{R}_N = \pi_* \mathcal{R}$, it follows from

 □

10.2.1. Projection of Lagrangian submanifolds.

We have the result:

Proposition 10.4 (Projection of Lagrangian submanifolds). Let $i: L \hookrightarrow M$ be a Lagrangian submanifold. If L and N have clean intersection and $\pi(L \cap N)$ is a submanifold in $N /\mathcal{F}$, then it is Lagrangian.

Proof. The proof is similar to the the proof of propositions 5.5 and 5.6, since the proof reduces to the study of each tangent space. □

10.3. Horizontal coisotropic reduction

Theorem 10.2 (Horizontal coisotropic reduction in stable Hamiltonian structures). Let $i: N \rightarrow M $ be a coisotropic horizontal submanifold such that $(TN)^{\perp_\Lambda}$ defines an integrable distribution. Let $\mathcal{F}$ be the set of leaves of the foliation and suppose that $N/\mathcal{F}$ admits a manifold structure such that the canonical projection $\pi: N \rightarrow N/\mathcal{F}$ defines a submersion. Then $N/\mathcal{F}$ admits and unique symplectic structure ωN such that $\pi^*\omega_N = i^*\omega$. The following diagram summarizes the situation:

Standard image High-resolution image

Proof. The proof is similar to the the proof of theorem 3.2. □

10.3.1. Projection of Lagrangian submanifolds

Proposition 10.5 (Projection of Lagrangian submanifolds). Let $L \hookrightarrow M$ be a Lagrangian submanifold. If L and N have clean intersection and $\pi(L \cap N)$ is a submanifold in $N/\mathcal{L}$, then it is Lagrangian.

Proof. The proof is similar to the the proof of proposition 3.4 since we only need to check it in every tangent space. □

11. Conclusions

In this paper we have reviewed the concept of coisotropic and Lagrangian (and Legendrian) submanifolds in different geometric settings. We have shown the connection of these geometric constructions with the different dynamics that usually appear in mechanics. All these different geometries (symplectic, cosymplectic, contact, cocontact) can be classified within Jacobi geometry (in some case, Poisson, a particular case of Jacobi structure). This approach allows in a simple way to see these particular situations from a more general point of view. Sometimes it is important to stand at a certain altitude in order to realise that these particular situations respond to the same geometrical pattern.

In each case we have introduced the notions of coisotropic and Lagrangian submanifold and studied in detail the corresponding coisotropic reduction theorems. The interpretations of the different types of vector fields in the different types of geometry as Lagrangian or Legendrian submanifolds are summarized in table 1. Also, the results on coisotropic reduction are summarized in table 2.

Table 1. Interpretation of vector fields as Lagrangian or Legendrian submanifolds.

GeometryVector fieldType of submanifoldAmbient manifold
SymplecticHamiltonianLagrangian $(TM, \omega_0)$, symplectic
$(M, \omega)$ $X_H = \sharp_\omega(\mathrm{d}H)$   $\omega_0 = \flat_\omega^* \omega_M$
CosymplecticGradientLagrangian $(TM, \Omega_0)$, symplectic
$(M, \Omega, \theta)$ $\operatorname{grad} H = \sharp_{\theta, \Omega}(\mathrm{d}H)$   $\Omega_0 = \flat_{\theta, \Omega}^* \Omega_M$
 HamiltonianLagrangian $(TM, \Omega_H)$, symplectic
  $X_H = \operatorname{grad} H - \mathcal{R}(H) \mathcal{R}$   $\Omega_H = \Omega_0 + (\operatorname{d}(\mathcal{R}(H)) \wedge \theta)^v$
  Lagrangian $(TM \times \mathbb{R},\Omega_H,\mathrm{d}s)$, cosymplectic
 EvolutionLagrangian $(TM, \Omega_H)$, symplectic
  $\mathcal{E}_H = \operatorname{grad} H + \mathcal{R}$   
ContactGradientLagrangian $(TM, \Omega_0)$, symplectic
$(M, \eta)$ $\operatorname{grad} H = \sharp_{\eta}(\mathrm{d}H)$   $\Omega_0 = \flat_\eta ^*\Omega_M$
 HamiltonianLagrangian $(TM, \Omega_H)$, symplectic
  $X_H = \operatorname{grad} H - (\mathcal{R}(H) + H)\mathcal{R}$   $\Omega_H = \Omega_0 - \operatorname{d}(\mathcal{R}(H) \eta + H \eta) ^v$
  Legendrian $(TM \times \mathbb{R}, \hat{\eta})$, contact
    $\hat{\eta} = \eta^c + t \eta^v$
 EvolutionLagrangian $(TM\times \mathbb{R}, \widetilde \Omega_H)$, symplectic
  $\mathcal{E}_H = X_H + H \mathcal{R}$   $\widetilde \Omega_H = \Omega_0 - \operatorname{d}(\mathcal{R}(H) \eta)^v$
CocontactGradientLagrangian $(TM, \Omega_0)$, symplectic
$(M, \theta, \eta)$ $\operatorname{grad} H = \sharp_{\theta, \eta}(\mathrm{d}H)$   $\Omega_0 = \flat_{\eta, \theta}^* \Omega_M$
 HamiltonianLagrangian $(TM, \Omega_H)$, symplectic
  $X_H = \operatorname{grad}(H) - (\mathcal{R}_z(H) + H)\mathcal{R}_z$   $\Omega_H = \Omega_0 - \operatorname{d}(\mathcal{R}_z(H) \eta + H\eta)^v$
  $\quad \quad +(1 - \mathcal{R}_t(H))\mathcal{R}_t$   $\quad \quad + \mathrm{d}(\theta - \mathcal{R}_t(H)\theta)^v$
  Legendrian $(TM \times \mathbb{R} \times \mathbb{R}, \widetilde \theta, \widetilde \eta)$, cocontact
    $\widetilde \eta = \eta^c + s \eta ^v + \theta^c + e \theta ^v$
    $\widetilde \theta = \theta^c$
SHSGradientLagrangian $(TM, \Omega_0)$, symplectic
$(M , \omega, \lambda)$ $\operatorname{grad} H = \sharp_{\lambda, \omega}({\mathrm{d}H})$   $\Omega_0 = \flat_{\lambda, \omega} ^*\Omega_M$
If $d\lambda = f \omega,$ HamiltonianLagrangian $(TM, \Omega_H)$, symplectic
$\sharp_{\lambda, \Omega}(df) \in \mathcal{V} $ $X_H = -\operatorname{grad} H + (\mathcal{R}(H) + f H) \mathcal{R}$   $\Omega_H = \Omega_0 + \operatorname{d}(\mathcal{R}(H)\lambda + fH\lambda)^v$

Table 2. Summary of results on coisotropic reduction.

 Coisotropic submanifold Projection of Lagrangian
Geometry $N \hookrightarrow M$ Resulting manifoldand Legendrian submanifolds
SymplecticArbitrary $(N/\mathcal{F}, \omega_N)$, symplectic L Lagrangian $\implies$ LN Lagrangian
CosymplecticVertical $(N/\mathcal{F}, \theta_N, \Omega_N)$, cosymplectic L Lagrangian $\implies$ LN Lagrangian
 Horizontal $(N, \Omega_N)$, symplectic L Lagrangian $\implies$ LN Lagrangian
 ArbitraryFoliation consisting of symplectic 
  manifolds of $N/\mathcal{F}$  
ContactVertical $(N/\mathcal{F}, \eta_N)$, contact L Legendrian $\implies$ LN Legendrian
 Horizontal $\operatorname{dim} N/\mathcal{F} = 0$  
Cocontact tz-vertical $(N/\mathcal{F}, \theta_N, \eta_N)$, cocontact L Legendrian $\implies$ LN Legendrian
  t-vertical, z-horizontal $\operatorname{dim} N/\mathcal{F} = 1$, $\theta_N \neq 0$  
  z-vertical, t-horizontal $(N/\mathcal{F}, \eta_N),$ contact L Legendrian $\implies$ LN Legendrian
  tz-horizontal $\operatorname{dim} N/\mathcal{F} = 0$  
SHSVertical $(N/\mathcal{F}, \omega_N, \lambda_N)$, stable Hamiltonian L Lagrangian $\implies$ LN Lagrangian
 Horizontal $(N/\mathcal{F}, \omega_N)$, symplectic L Lagrangian $\implies$ LN Lagrangian

Acknowledgments

We acknowledge the financial support of Grant PID2019-106715GBC21, the Severo Ochoa Programme for Centres of Excellence in R&D (CEX2019-000904-S), and JAE Intro Programme 2022 (Becas de Introducción a la Investigación para estudiantes universitarios). Finally, we also thank the referees for their corrections and suggestions.

Data availability statement

All data that support the findings of this study are included within the article (and any supplementary files).

Please wait… references are loading.
10.1088/1751-8121/ad37b2