Tangent Dirac structures and submanifolds
by Izu Vaisman
ABSTRACT. We write down the local equations that characterize the submanifolds
of a Dirac manifold
which have a normal bundle that is either a coisotropic or an isotropic submanifold of
endowed with the tangent Dirac structure. In the Poisson case, these formulas prove again a result of Xu: the submanifold
has a normal bundle which is a coisotropic submanifold of
with the tangent Poisson structure iff
is a Dirac submanifold. In the presymplectic case it is the isotropy of the normal bundle which characterizes the corresponding notion of a Dirac submanifold. On the way, we give a simple definition of the tangent Dirac structure, we make new remarks about it, and we establish characteristic, local formulas for various interesting classes of submanifolds of a Dirac manifold.
1 Introduction
The framework of the present paper is the
category. We will denote by
spaces of differential
-forms, by
spaces of
-vector fields, by
spaces of differentiable cross sections of vector bundles, and we will use the Einstein summation convention.
The Dirac structures were introduced in the study of constrained systems and unify Poisson and presymplectic geometry [1] . We will recall their definition later. A pair
that consists of an
-dimensional manifold
and a Dirac structure
on
is called a Dirac manifold. We started the study of submanifolds of a Dirac manifold in [11] , where we noticed the classes of properly normalized, totally Dirac and cosymplectic submanifolds. These classes extend the Poisson-Dirac with Dirac projection, Lie-Dirac (Dirac) and cosymplectic submanifolds of a Poisson manifold, respectively [13, 3] .
In [13] , Xu proved that the Dirac submanifolds of the Poisson manifold
are characterized by the nice property of having a normal bundle which is a coisotropic submanifold of the tangent manifold
endowed with the tangent Poisson structure. The latter is defined by the complete lift [14] of the bivector field
. All the terms of Xu's result, including the notion of a tangent Dirac structure [2] , are also defined for Dirac manifolds, and Xu's result indicates that one may expect interesting connections between the geometry of a submanifold
of a Dirac manifold
and the geometry of a normal bundle of
in the tangent manifold
. This is the motivation of the present paper.
We will discuss the geometric configuration of Xu's result in the general case of a Dirac manifold. The terms of the theorem are either new or not popular, and are based on either new or not popular geometric constructions.
Accordingly, it is an objective of the paper to explain these terms in detail.
Particularly, we recall the general construction of the vertical and complete lifts of tensor fields from a manifold
to the total space of the tangent bundle
, and the main properties of these operations [14] . We use these lifts in order to give a simple definition of the tangent Dirac structure and make some new remarks about it.
Then, we turn to submanifolds. We define various classes of submanifolds of a Dirac manifold and characterize them via local coordinates and bases.
Furthermore, we obtain the local conditions that characterize submanifolds
of
with a normal bundle
which is either a coisotropic or an isotropic submanifold of
. These formulas imply the result proven by Xu in the case of Poisson manifolds. Another consequence of the established formulas is that the analogs of Dirac submanifolds of a presymplectic manifold
are characterized by the existence of a normal bundle which is isotropic in
.
2 Complete and vertical lifts revisited
Let
be an
-dimensional, differentiable manifold and
be the total space of its tangent bundle. In the space of differentiable functions
one has the important linear subspace
of the fiberwise linear functions, the latter being functions of the form
|
(2.1)
|
In particular, if we denote by
local coordinates on
and by
the corresponding natural coordinates on the fibers of
(i.e., coordinates of tangent vectors with respect to the bases
), we have
. Hence, locally,
is functionally spanned by the set of functions
, where
is the natural projection and
. In what follows the function
will be denoted again by
.
Two other ingredients also are of great importance in the geometry of
.
The first is the Euler vector field
of infinitesimal homotheties of the fibers, which is characterized by
|
(2.2)
|
and has the local expression
The second is the tangent structure tensor field
, which is characterized by
|
(2.4)
|
where
, and has the local expression
|
(2.5)
|
If, as usual, we denote by
the subbundle tangent to the fibers, called the vertical bundle, we have
|
(2.6)
|
and
has a vanishing Nijenhuis tensor:
|
(2.7)
|
Firstly, using the ingredients introduced above and denoting
, we get
Proposition 2.1.
[
14]
There exists a unique homomorphism of real tensor algebras that sends a tensor
to a tensor
, where
and
, called the vertical lift, such that
|
(2.8)
|
where
,
, and
is any vector such that
. Moreover, the vertical lift of a differentiable tensor field is a differentiable tensor field and, for differential forms, the vertical lift commutes with the exterior differential.
-
Proof.
We notice that
there exists a unique
such that
. Then, we define
|
(2.9)
|
The assertions about tensor fields and differential forms follows from the fact that in the case of a differential form
(functions included) one has
, and in the case of a vector field
one has
.
Secondly, we define an operation on tensor fields known as the complete lift [14] . For any vector field
, the flow
lifts to a local
-parameter Lie group
on the manifold
, which is defined by a vector field
called the complete lift of
. The local expression of
is
|
(2.10)
|
and it is easy to check that [14]
|
(2.11)
|
|
(2.12)
|
|
(2.13)
|
Furthermore, one has
Proposition 2.2.
[
14]
There exists a unique homomorphism of real linear spaces which sends a tensor field
to a field
, called the complete lift of
, such that,
,
,
,
is given by ( 2.11 ), and
|
(2.14)
|
-
Proof.
Notice that the lift
of a function was chosen such that for any vector field seen as
the pull back
. The definition of
and condition ( 2.14 ) compel us to define the complete lift of a
-form
by
|
(2.15)
|
The corresponding local coordinate expression is
|
(2.16)
|
and for any vector field
the pull back of
is given by
, where
denotes the Lie derivative. Finally, condition ( 2.14 ) uniquely defines the complete lift of an arbitrary tensor field because this condition is compatible with associativity.
We also indicate the following properties of the complete lift:
a) [14] The complete lift of a
-form
is a
-form
and
. Indeed, a straightforward calculation shows that this condition holds for functions and
-forms. Then, the condition for an arbitrary form follows by expressing the latter locally as a real linear combination of exterior products of
-forms and using ( 2.14 ).
b) [14] The Lie derivative of a tensor field
has the following lifts:
|
(2.17)
|
It is enough to check ( 2.17 ) for functions, vector fields and
-forms, and this can be done with the already explained formulas ( 2.8 ) ( 2.13 ). Then, the general result follows from ( 2.14 ).
c) The complete lift of a
-vector field
is a
-vector field
and the Schouten-Nijenhuis bracket satisfies the condition
. This follows by expressing
locally as real linear combinations of exterior products of vector fields, using the expression of the Schouten-Nijenhuis bracket of such exterior products (e.g., [9] ) and ( 2.13 ), ( 2.14 ).
As an application of property c), if
is a Poisson bivector field on
, i.e.
, then
is a Poisson bivector field on the manifold
. The Poisson structure defined by
is called the tangent Poisson structure and it was used by many authors ([2, 12] , etc.).
d) The complete lifts
of all the vector fields
span a generalized foliation of
such that, if we identify
with the zero section of
, the leaves through points of
are the connected components of
and the leaf through
,
, is the connected component of
in
. Indeed, by ( 2.10 ), at a point
we have
and at a point
such that, for instance,
, we have
Finally, we refer the reader to [6] and the references therein for generalizations of the lift operations discussed above.
3 Tangent Dirac structures
Now, we will use the complete and vertical lifts in order to define the notion of a tangent Dirac structure, first introduced by Courant [2] , and make some new remarks about it.
The Dirac structures are defined as a class of subbundles of the vector bundle
. The bundle
has several interesting geometric objects. The first is the non degenerate metric of zero signature
|
(3.1)
|
where
are tangent vectors and
are tangent covectors at
.
The second is
given by
|
(3.2)
|
which is a so-called para-Hermitian structure [4] . The third object is the non degenerate
-form
|
(3.3)
|
Definition 3.1.
A maximal
-isotropic subbundle
is called an almost Dirac structure on
.
The almost Dirac structure may be interpreted in terms of
alone.
Namely [1] ,
yields the pair
, where
is the generalized distribution defined as the natural projection of
on
and,
,
is the
-form induced by
of ( 3.3 ) (
, the value produced by ( 3.3 ) does not depend on the choice of
). Conversely, the pair
allows us to reconstruct
as follows
|
(3.4)
|
The next important thing for the bundle
is the Courant bracket, which is the operation defined on
by [1]
|
(3.5)
|
where
are vector fields and
are differential
-forms on
,
is the usual Lie bracket and
denotes the Lie derivative. The Courant bracket is skew-symmetric but satisfies a more complicated than the Jacobi identity.
Definition 3.2.
An almost Dirac structure
is called a Dirac structure on
if
is closed by Courant brackets.
In [1] , it was proven that the almost Dirac structure is Dirac iff the equivalent pair
satisfies the following conditions: i)
is a generalized foliation, ii) the form
is closed along the leaves of
. This means that the leaves of
are presymplectic manifolds (of a non constant rank!) and are called the presymplectic leaves of
. If the leaves are symplectic
is equivalent with a Poisson structure. Namely, if
is the corresponding Poisson bivector field, the Dirac structure is
|
(3.6)
|
If the leaves are the connected components of
,
is a presymplectic structure on
with the presymplectic form
such that
|
(3.7)
|
Another fundamental property of a Dirac structure is that the restriction of the Courant bracket ( 3.5 ) to
makes
into a Lie algebroid where the anchor is projection on
(e.g., [9] ).
In [2] , a Dirac structure of
was lifted to the manifold
. In what follows, we give a simpler definition of this lift. For this purpose we look at the locally free
-module sheaf
of rank
of the germs of cross sections of
, where
is the sheaf of germs of differentiable functions on
(e.g., [8] ). Then, we get
Proposition 3.1.
The
-module sheaf
spanned by the germs
,
,
is locally free of rank
and it is isomorphic with the sheaf of germs of cross sections of a Dirac structure
on
.
-
Proof.
Firstly, we show that the sheaf
is locally free of rank
. If
is a local basis for the sheaf
on
, an arbitrary germ
is of the form
whence
| |
| |
This shows that
is a local basis of
. (If we assume that
and use formulas ( 2.10 ), ( 2.16 ), linear independence follows from that of
.) Thus [8] ,
is isomorphic with the sheaf of germs of cross sections of the vector bundle with local trivialization bases
, which may be identified with a vector subbundle
of
.
Furthermore, if we denote by indices
and
, respectively, objects on the two manifolds, formulas ( 3.1 ), ( 3.3 ) and ( 2.12 ), ( 2.15 ) give
|
(3.8)
|
and similar formulas relate
to
. These formulas ensure the isotropy property for
.
Finally, from ( 3.8 ) for
, ( 2.11 ) ( 2.13 ) and property b) of the complete lift given at the end of Section 1, we get the following formulas for Courant brackets:
|
(3.9)
|
These formulas, and the fact that a Dirac structure is a Lie algebroid, ensure that
is closed by Courant brackets.
Definition 3.3.
The Dirac structure
of
provided by Proposition 3.1 is called the tangent Dirac structure of the Dirac structure
of
.
The comparison of the generating pairs of
with the bases produced by the computations of [2] or with the alternative definition of the reviewer of that paper [5] shows that the tangent Dirac structure of Definition 3.3 is the same as that of [2] .
Remark 3.1.
The tangent Dirac structure is invariant by the operator
of the tangent structure of the manifold
. Indeed, the action of
is defined by
, and the definitions of Section 1 yield
Example 3.1.
A Poisson structure of
defined by the bivector field
with
is equivalent with the Dirac structure
given by ( 3.6 ).
From ( 2.12 ) and ( 2.15 ), it follows easily:
Hence, the tangent Dirac structure of
is the Poisson structure defined on
by the bivector field
, which is the usual definition of a tangent Poisson structure. Similarly, if
has a presymplectic structure defined by the closed
-form
, this structure may be seen as the Dirac structure
given by ( 3.7 ) and the tangent Dirac structure of
is the presymplectic structure defined on
by
.
Example 3.2.
The construction of the tangent Dirac structure extends to complex Dirac structures
. Such a structure
is a generalized complex structure of
if
[
7]
. By looking at a complex basis
of
, it follows easily that if
is a generalized complex structure the same is true for its tangent structure
. Therefore, the tangent manifold of a generalized complex manifold is a generalized complex manifold, again, in a canonical way. If
has a usual complex structure,
is the direct sum of the holomorphic tangent bundle of
and the anti-holomorphic cotangent bundle [
7]
, and
has the similar structure for the usual complex structure of
. On the other hand, if the generalized complex structure is defined by a symplectic form
on
,
is the complexification of the Dirac structure
of ( 3.7 ) [
7]
, and the generalized complex structure of
is defined by the symplectic form
.
Now, we will give some more results about the tangent Dirac structure.
Proposition 3.2.
If
is a presymplectic leaf of
on
with the presymplectic form
and if
, the presymplectic leaf of
through
is the tangent manifold
, and its presymplectic form is
, where the complete lift is from
to
.
-
Proof.
Obviously, the tangent space of the presymplectic leaf
of
at a point
,
, is spanned by the vectors
,
, where
is a vector field tangent to the presymplectic leaf
.
In particular, if
, we get the first part of the proposition. Furthermore, if
is the presymplectic form of the leaf
we have [1]
|
(3.10)
|
where
,
and the form
is defined by formula ( 3.3 ). Then, the definitions and properties of the complete and vertical lifts yield
|
(3.11)
|
and we are done.
We recall that a Poisson structure
is called homogeneous if there exists a vector field
such that
It is well known that the tangent Poisson structure
of any Poisson structure
of
is homogeneous with
, where
is the Euler vector field ( 2.3 ). The generalization of homogeneity to Dirac structures
is the condition
|
(3.13)
|
which reduces to ( 3.12 ) in the Poisson case [11] . Now, we get
Proposition 3.3.
For an arbitrary Dirac structure
of
, the tangent Dirac structure
is homogeneous with
.
-
Proof.
The Euler field
satisfies ( 2.2 ) and also
|
(3.14)
|
which follows by easy, local coordinates calculations. Thus, the result is proven if we show that
the pairs
are
-orthogonal to the pairs
. The examination of the corresponding scalar products shows that this is the case indeed.
Remark 3.2.
A Dirac structure
of
yields a pointwisely defined pull back
to
by the natural projection
, which is defined by
|
(3.15)
|
From ( 3.15 ), it follows that the equivalent, locally free sheaf of rank
is spanned by
where
and
is vertical on
.
If
is a local basis of
,
is a local basis of
, and we see that
is a differentiable Dirac structure on
, which is different from the tangent Dirac structure. In particular, if
where
is a Poisson bivector field,
is not a Poisson structure. We might say that
is the complete lift of
and
is the vertical lift. The presymplectic leaves of
are the restrictions
of the tangent bundle
to the presymplectic leaves
of
and the corresponding presymplectic form is the vertical lift of the presymplectic form of
.
4 Submanifolds of a Dirac manifold
We begin by defining various classes of submanifolds of a Dirac manifold.
More details and motivation on that may be found in [11] . For simplicity, all the submanifolds are assumed to be embedded submanifolds.
Definition 4.1.
A submanifold
is properly normalizable if there exists a normal bundle
of
such that
|
(4.1)
|
where the projections are defined by the decomposition
.
If ( 4.1 ) holds, the pair
is a properly normalized submanifold of
.
It follows easily that condition ( 4.1 ) is equivalent with
|
(4.2)
|
and
is a differentiable Dirac structure of
equal to the pull back of
by the embedding of
in
[11] . Therefore, a properly normalizable submanifold has a well defined induced Dirac structure. If
comes from a Poisson structure a properly normalizable submanifold is a Poisson-Dirac submanifold which admits a Dirac projection in the sense of [3] . If
comes from a presymplectic form
the submanifold
is properly normalizable iff there exists a normal bundle
of
which is
-orthogonal to
.
In [11] we have defined an interesting invariant of a properly normalized submanifold called the second fundamental form of
in
. This invariant associates with every pair
a
-form
with the value on
given by
|
(4.3)
|
where
extend
from
to
and
. The result is independent of the choice of the extensions because the right hand side of ( 4.3 ) is
-linear in all arguments.
In order to define another class of submanifolds we notice the existence of the field of subspaces along
|
(4.4)
|
(
denotes the annihilator space). This field may not be differentiable and the subspaces may have various dimensions and may intersect
. For these reasons we will say that
is the pseudo-normal field of
with respect to
.
Definition 4.2.
The submanifold
is a cosymplectic submanifold if the pseudo-normal field
is a differentiable normal bundle
of
in
.
will be called the natural normal bundle of the cosymplectic submanifold
.
In [11] , it was proven that, if
is a cosymplectic submanifold,
is a properly normalized submanifold, that
is cosymplectic iff
|
(4.5)
|
that the induced Dirac structure of a cosymplectic submanifold is Poisson and, along
, one has
|
(4.6)
|
where
is the bivector field of the induced Poisson structure and
is a
-form the graph of which is
. If
comes from a Poisson structure the notion of a cosymplectic submanifold is the known one [13] . If
comes from a presymplectic form
the submanifold
is cosymplectic iff the pull back of
to
is non degenerate.
Because of ( 4.5 ) we give the following definition:
Definition 4.3.
The function
defined by
is called the cosymplecticity default of the submanifold
of
.
It turns out that the second fundamental form of a cosymplectic manifold vanishes [11] . This property is the source of
Definition 4.4.
A submanifold
of a Dirac manifold
is a totally Dirac submanifold if it is properly normalizable by a certain normal bundle
and the second fundamental form of
is zero.
One can see [11] that, if
comes from a Poisson structure, a totally Dirac submanifold is just a Dirac submanifold in the sense of Xu [13] (in [3] these were called Lie-Dirac submanifolds). Indeed, if
is the Poisson bivector field, the second fundamental form of
becomes
(then
because of the proper normalization property), and, if
, we are in the case where
is soldered to
[10] . Similarly, if
where
is a presymplectic form on
, for the properly normalized submanifold
we get
and the vanishing of this form, together with
means that
is soldered to
in the sense of [10] .
Finally, in accordance with the Poisson case, we also define
Definition 4.5.
A submanifold
of a Dirac manifold
is coisotropic if the conditions
and
imply
.
Dually,
is an isotropic submanifold of
if the conditions
and
imply
.
In the presymplectic case the coisotropy and isotropy properties are the classical ones (i.e.,
and
, respectively, where
is the presymplectic form). In the Poisson case,
is coisotropic iff,
,
is a coisotropic subspace of
, respectively,
is an isotropic submanifold of
,
being the symplectic leaf through
. Obviously,
is a coisotropic submanifold of
iff its pseudo-normal field satisfies the condition
.
Now, we shall explain how to represent a Dirac structure
of
in the neighborhood of a point
of a submanifold
of
by means of local bases.
Using a tubular neighborhood of
with fibers tangent to a chosen normal bundle
, we get local coordinates
;
around
such that
are coordinates along
and
are coordinates along the tubular fibers. Then the local equations of
are
, and
|
(4.7)
|
|
(4.8)
|
On the coordinate neighborhood obtained above (shrunken if necessary), we may consider a basis of
that consists of
independent pairs
where
|
(4.9)
|
In these, and in all the formulas that follow, our convention is that any index of coordinates
takes the same values as the index
and any index of coordinates
takes the same values as the index
.
Of course, these bases must satisfy the conditions implied by the definition of a Dirac structure namely, isotropy:
|
(4.10)
|
and integrability:
|
(4.11)
|
We also add that the tangent distribution of the presymplectic foliation of
is
|
(4.12)
|
and the presymplectic form
is determined by
|
(4.13)
|
In what follows we write down the characteristic conditions for the various classes of submanifolds. Definition 4.1 shows that
is a properly normalizable submanifold of
iff it has a normal bundle
such that there exist local bases ( 4.9 ) of
which satisfy the conditions
|
(4.14)
|
We call them adapted bases. Furthermore,
is cosymplectic in
iff there exist
and bases that satisfy ( 4.14 ) and the supplementary conditions
|
(4.15)
|
This is an immediate consequence of formula ( 4.6 ).
Back to a properly normalized submanifold
, with ( 4.3 ) we can compute the components of the second fundamental form with respect to adapted bases, and the result is
|
(4.16)
|
Therefore, following Definition 4.4 , the characterization of a totally Dirac submanifold consists of conditions ( 4.14 ) and the annulation of the components ( 4.16 ).
Remark 4.1.
The skew-symmetry of the second fundamental form
is an immediate consequence of the first condition ( 4.10 ). The vanishing of
for a cosymplectic submanifold follows from the first condition ( 4.11 ). Indeed, if
is a cosymplectic submanifold, the
-component of
behaves like a Poisson structure (see ( 4.6 )), and the fact that
|
(4.17)
|
belongs to
along
implies the annulation of the
-form component of ( 4.17 ) when calculated on a vector field of the form
and evaluated at
. This exactly yields
.
Finally, Definition 4.5 shows that a submanifold
is coisotropic if the coefficients of the formulas ( 4.9 ) are such that
one has
|
(4.18)
|
The condition for an isotropic submanifold is obtained by reversing the sense of the implication in ( 4.18 ).
Now, we have all the ingredients required to discuss the result of Xu quoted in Introduction in the framework of Dirac manifolds. We will deduce the conditions for a submanifold
of a Dirac manifold
to have a normal bundle
which is either a coisotropic or an isotropic submanifold of
, and obtain some geometric conclusions of these conditions.
We consider a point
, a normal bundle
of
in
, and the local coordinates and bases of formulas ( 4.9 ) around
. Then, if we denote by
the corresponding natural coordinates on the fibers of
, the submanifold
has the local equations
, and
|
(4.19)
|
|
(4.20)
|
With the bases ( 4.9 ) the tangent Dirac structure
is locally spanned by
, where
|
(4.21)
|
|
(4.22)
|
By writing down a linear combination of these pairs with coefficients
we get a local cross section
of
which has the property
iff
|
(4.23)
|
The same cross section
satisfies the condition
iff
|
(4.24)
|
Therefore, we have proven
Proposition 4.1.
A submanifold
of a Dirac manifold
has a normal bundle
which is coisotropic in
iff, around the points of
,
has bases ( 4.9 ) such that the equations ( 4.24 ) are a consequence of the equations ( 4.23 ). Similarly,
is isotropic in
iff the equations ( 4.23 ) are a consequence of the equations ( 4.24 ).
As consequences of Proposition 4.1 we get
Proposition 4.2.
[
13]
A submanifold
of a Poisson manifold
with the Poisson bivector field
is totally Dirac iff it has a normal bundle
which is a coisotropic submanifold of
.
-
Proof.
Using coordinates as in ( 4.9 ) we may write
|
(4.25)
|
Accordingly, the bases ( 4.9 ) may be taken under the form
Then, the equations ( 4.23 ) become
, and the equations ( 4.24 ) become
for
. Obviously, the first system implies the second iff
for
. These exactly are the conditions for
to be a (totally) Dirac submanifold [13] .
Similarly, we have
Proposition 4.3.
A submanifold
of a presymplectic manifold
with the closed
-form
is totally Dirac iff it has a normal bundle
which is an isotropic submanifold of
.
-
Proof.
With the same notation, we have
and the bases
Then, the system ( 4.23 ) becomes
for
, and the system ( 4.24 ) becomes
. The latter conditions imply the former iff
and
for
.
These are the conditions that characterize a totally Dirac submanifold of a presymplectic manifold.
Proposition 4.4.
If
is a cosymplectic submanifold of the Dirac manifold
, the cosymplecticity default of the natural normal bundle
seen as a submanifold of
satisfies the inequalities
.
-
Proof.
The local cross sections of
must satisfy both ( 4.23 ) and ( 4.24 ), which, modulo ( 4.14 ) and ( 4.15 ), include the conditions
and do not restrict the coefficients
. Therefore,
has a basis which consists of the pairs
and of linear combinations of
.
Remark 4.2.
The questions discussed above may also be considered for the vertical lift
defined in Remark 3.2 , instead of the tangent structure
. The local bases of
are
.
The conditions ( 4.23 ) are to be replaced by
|
(4.26)
|
The conditions ( 4.24 ) are to be replaced by
|
(4.27)
|
From these formulas, we see that
is never coisotropic in
.
References
-
T. J. Courant, Dirac Manifolds, Transactions Amer. Math. Soc., 319 (1990), 631-661.
-
T. J. Courant, Tangent Dirac structures, J. Phys. A, 23 (1990), no. 22, 5153 5168.
-
M. Crainic and R. L. Fernandez, Integrability of Poisson brackets, J. Differential Geom., 66 (2004), 71-137.
-
V. Cruceanu, P. Fortuny and P. M. Gadea, A survey on paracomplex geometry, Rocky Mountain J. of Math., 26 (1996), 1-33.
-
M. de Leon, Math. Reviews 1085863 (92d:58064).
-
J. Grabowski and P. Urbański, Tangent and cotangent lifts and graded Lie algebras associated with Lie algebroids, Ann. Global Analysis Geom., 15 (1997), 447-486.
-
M. Gualtieri, Generalized complex geometry, Oxford University DPhil thesis, 107 pages, arXiv:math.DG/0401221.
-
Tennison, B. R., Sheaf theory, London Math. Soc. Lecture Notes Series 20, Cambridge Univ. Press, Cambridge, 1975.
-
I. Vaisman, Lectures on the geometry of Poisson manifolds, Progress in Math., vol. 118, Birkhäuser, Basel, 1994.
-
I. Vaisman, Dirac submanifolds of Jacobi manifolds. In: The Breadth of Symplectic and Poisson Geometry, Festschrift in Honor of Alan Weinstein (J. E. Marsden and T. Ratiu, eds.), Progress in Math., vol. 232, p. 603-622, Birkhäuser, Boston, 2005.
-
I. Vaisman, Foliation coupling Dirac structures, arXiv:math.SG/0412318.
-
A. Weinstein, Lagrangian mechanics and grupoids, Mechanics day (Waterloo ON, 1992), Fields Inst. Commun., 7, AMS, Providence, RI, 1996, 207-231.
-
P. Xu, Dirac submanifolds and Poisson involutions, Ann. Sci. Ec. Norm. Sup., 36 (2003), 403-430.
-
K. Yano and S. Ishihara, Tangent and Cotangent Bundles, M. Dekker, Inc., New York, 1973.
Department of Mathematics
|
University of Haifa, Israel
|
E-mail: vaisman@math.haifa.ac.il
|