(Grundlehren Der Mathematischen Wissenschaften 347) Camille Laurent-Gengoux, Anne Pichereau, Pol Vanhaecke (Auth.) - Poisson Structures-Springer-Verlag Berlin Heidelberg (2013)
(Grundlehren Der Mathematischen Wissenschaften 347) Camille Laurent-Gengoux, Anne Pichereau, Pol Vanhaecke (Auth.) - Poisson Structures-Springer-Verlag Berlin Heidelberg (2013)
Series editors
Editor-in-Chief
A. Chenciner J. Coates S.R.S. Varadhan
For further volumes:
www.springer.com/series/138
Camille Laurent-Gengoux r Anne Pichereau r
Pol Vanhaecke
Poisson Structures
Camille Laurent-Gengoux Pol Vanhaecke
CNRS UMR 7122, Laboratoire CNRS UMR 7348, Lab. Mathématiques
de Mathématiques et Applications
Université de Lorraine Université de Poitiers
Metz, France Futuroscope Chasseneuil, France
Anne Pichereau
CNRS UMR 5208, Institut
Camille Jordan
Université Jean Monnet
Saint-Etienne, France
Poisson structures naturally appear in very different forms and contexts. Symplec-
tic manifolds, Lie algebras, singularity theory, r-matrices, for example, all lead to a
certain type of Poisson structures, sharing several features, despite the distances be-
tween the mathematics they originate from. This observation motivated us to bring
the different worlds in which Poisson structures live together in a single volume,
providing a multitude of entrances to the book, and hence to the subject. Thus, the
idea of the body of the book, Part II, was born.
It is well known, i.e., it is commonly agreed upon by the experts, that results and
techniques which are valid for one type of Poisson structure ought to apply, mutatis
mutandis, to other types of Poisson structures. When starting to write Part I of the
book, we soon realized that not everything could be derived from the general con-
cept of a Poisson algebra, so we faced the challenge of presenting the concepts and
the results in detail, for the algebraic, algebraic-geometric and geometric contexts,
without copy-pasting large parts of the text two or three times. But as the writing
moved on, both the algebraic and geometric contexts kept imposing themselves; fi-
nally, each found its proper place, appearing as being complementary and dual to
the other, rather than a consequence or rephrasing, one of the other. It added, unex-
pectedly, an extra dimension to the book.
It was pointed out by one of the referees that the main applications of Poisson
structures should be present in a book which has Poisson structures as its main
subject. This was quite another challenge, giving rise to the third and final part of
the book, Part III, undoubtedly an important addition.
Several years were necessary for this project, a big part was done at Poitiers,
when the three of us were appointed to the “Laboratoire de Mathématiques et Ap-
plications”. We were spending long hours together in what our colleagues called
The Aquarium (for an explanation of the name, look up “poisson” in a French dic-
tionary). When two of us moved away from Poitiers, we sometimes worked to-
gether in other places, which always offered us a pleasant and stimulating working
atmosphere. Thus, we are happy to acknowledge the hospitality of our colleagues
from the mathematics departments at the universities of Poitiers, Antwerp, Coimbra,
vii
viii Preface
Saint-Etienne and at the CRM in Barcelona and the Max-Planck Institute in Bonn.
Each one of us was partially supported by an ANR contract (TcChAm, DPSing and
GIMP, respectively), which made it possible to keep working on a blackboard, rather
than seeing each other on a computer screen.
We learned Poisson structures from our teachers, friends and collaborators: Mark
Adler, Paulo Antunes, Pantelis Damianou, Rui Fernandes, Benoit Fresse, Eva Mi-
randa, Joana Nunes da Costa, Marco Pedroni, Michael Penkava, Claude Roger,
Pierre van Moerbeke, Yvette Kosmann-Schwarzbach, Hervé Sabourin, Mathieu
Stiénon, Friedrich Wagemann, Alan Weinstein, Ping Xu and Marco Zambon. The
multitude of different points of view on the subject, which we learned from them,
have given a quite specific flavor to the book. Even if a book cannot replace what one
learns through lectures, discussions and collaborations, it is our hope that this book
may further transfer what we learned from them and that it invites other researchers
to explore the subject of Poisson structures, which turns out to be as diverse and rich
as the colorful fauna and flora which inhabit the bottom of our oceans.
ix
x Contents
2.3.1
Real Poisson Structures Associated to Holomorphic
Poisson Structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
2.3.2 Holomorphic Poisson Structures on Smooth Affine
Poisson Varieties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
2.4 Other Constructions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
2.4.1 Field Extension . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
2.4.2 Localization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
2.4.3 Germification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
2.5 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
2.6 Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
4 Poisson (Co)Homology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
4.1 Lie Algebra and Poisson Cohomology . . . . . . . . . . . . . . . . . . . . . . . . . 92
4.1.1 Lie Algebra Cohomology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
4.1.2 Poisson Cohomology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
4.2 Lie Algebra and Poisson Homology . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
4.2.1 Lie Algebra Homology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
4.2.2 Poisson Homology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
4.3 Operations in Homology and Cohomology . . . . . . . . . . . . . . . . . . . . . 101
4.4 The Modular Class of a Poisson Manifold . . . . . . . . . . . . . . . . . . . . . . 102
4.4.1 The Modular Vector Field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
4.4.2 The Modular Class . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
4.4.3 The Divergence of the Poisson Bracket . . . . . . . . . . . . . . . . . . 107
4.4.4 Unimodular Poisson Manifolds . . . . . . . . . . . . . . . . . . . . . . . . 109
4.5 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
4.6 Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
Contents xi
5 Reduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
5.1 Lie Groups and (Their) Lie Algebras . . . . . . . . . . . . . . . . . . . . . . . . . . 114
5.1.1 Lie Groups and Lie Algebras . . . . . . . . . . . . . . . . . . . . . . . . . . 114
5.1.2 Lie Group Actions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
5.1.3 Adjoint and Coadjoint Action . . . . . . . . . . . . . . . . . . . . . . . . . . 119
5.1.4 Invariant Bilinear Forms and Invariant Functions . . . . . . . . . . 121
5.2 Poisson Reduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
5.2.1 Algebraic Poisson Reduction . . . . . . . . . . . . . . . . . . . . . . . . . . 123
5.2.2 Poisson Reduction for Poisson Manifolds . . . . . . . . . . . . . . . . 127
5.3 Poisson–Dirac Reduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
5.3.1 Algebraic Poisson–Dirac Reduction . . . . . . . . . . . . . . . . . . . . . 134
5.3.2 Poisson–Dirac Reduction for Poisson Manifolds . . . . . . . . . . 137
5.3.3 The Transverse Poisson Structure . . . . . . . . . . . . . . . . . . . . . . . 143
5.4 Poisson Structures and Group Actions . . . . . . . . . . . . . . . . . . . . . . . . . 148
5.4.1 Poisson Actions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148
5.4.2 Poisson Actions and Quotient Spaces . . . . . . . . . . . . . . . . . . . 149
5.4.3 Fixed Point Sets as Poisson–Dirac Submanifolds . . . . . . . . . . 150
5.4.4 Reduction with Respect to a Momentum Map . . . . . . . . . . . . 153
5.5 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
5.6 Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
Part II Examples
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 439
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 449
Poisson structures appear in a large variety of different contexts, ranging from string
theory, classical/quantum mechanics and differential geometry to abstract algebra,
algebraic geometry and representation theory. In each one of these contexts, it turns
out that the Poisson structure is not a theoretical artifact, but a key element which,
unsolicited, comes along with the problem which is investigated and its delicate
properties are in basically all cases decisive for the solution to the problem.
∂H ∂H
q̇i = , ṗi = − , i = 1, . . . , r , (0.2)
∂ pi ∂ qi
For every solution t → (q(t), p(t)) to these differential equations, it follows from
(0.1) and (0.3) that
d
F(q(t), p(t)) = {F, H} (q(t), p(t)) , (0.4)
dt
for all smooth functions F on R2r .
xv
xvi Introduction
and such that the functions H1 , . . . , Hr (including the Hamiltonian H) depend on the
coordinates ρ1 , . . . , ρr only. According to (0.2), Hamilton’s equations take in these
coordinates the simple form
∂H ∂H
θ̇i = , ρ̇i = − =0, i = 1, . . . , r ,
∂ ρi ∂ θi
which entails that the ρi are constant and hence that the θi are affine functions of
time (since H does not depend on θ1 , . . . , θr ).
Thus, in a nutshell, the equations of motion of classical mechanics come with a
Poisson bracket, the Poisson bracket is decisive for their integrability and it permits
us to construct coordinates in which the problem (including the Poisson bracket)
takes a very simple form, from which the solutions and their characteristics can be
read off at once.
Introduction xvii
Abstraction and generalization. Before giving another example of the key rôle
played by Poisson brackets, we explain the algebraic and geometric abstraction of
the classical Poisson bracket (0.1). It is well known that Poisson’s theorem is ex-
plained by the Jacobi identity
satisfies the Jacobi identity if and only if the functions xi j satisfy the identity
d
∂ xi j ∂ x jk ∂ xki
∑ k ∂ x i ∂ x j ∂ x = 0 .
x + x + x
=1
This fact was observed by Lie, who also pointed out that under the assumption of
constancy of the rank (which he assumes implicitly), there exist local coordinates
q1 , . . . , qr , p1 , . . . , pr , z1 , . . . , zs , where d = 2r + s, such that (0.6) takes the form
r
∂F ∂G ∂G ∂F
{F, G} = ∑ − ,
i=1 ∂ qi ∂ pi ∂ qi ∂ pi
Moreover, this Lie algebra structure on C∞ (U) and the associative structure
on C∞ (U) (pointwise multiplication of functions) are compatible in the sense that
t : A × A → A
(0.9)
(a, b) → a t b
such that, for each fixed t ∈ F, the product t is associative, with 0 being the original
associative commutative product on A , i.e., (A , ·) = (A , 0 ). We refer to the family
of associative products t on A as a deformation of A . It is assumed that the family
depends nicely on t, in a way which we do not make precise here. We consider, for
every t ∈ F, the commutator of t , which is defined for all a, b ∈ A by
[a, b]t := a t b − b t a .
Introduction xix
Since t is associative, the commutator [· , ·]t satisfies for each t ∈ F the following
identities, valid for all a, b, c ∈ A :
Then the first equation in (0.10) implies that {· , ·} satisfies the Leibniz prop-
erty (0.8), while the second equation in (0.10) implies that {· , ·} satisfies the Jacobi
identity. Combined, it means that (A , ·, {· , ·}) is a Poisson algebra. Thus, a fam-
ily of deformations of a given commutative associative algebra leads naturally to a
Poisson bracket on this algebra! In the particular case of the associative algebras of
operators of quantum mechanics, in which t is a multiple of Planck’s constant h̄, the
operators become in the limit h̄ → 0 commuting classical variables and the above
procedure yields a Poisson bracket on the algebra of classical variables.
The main problem of deformation theory is to invert this procedure; in the context
of physics, this is called the quantization problem or, more precisely, the problem
of quantization by deformation. In its algebraic version, the first question is whether
given a Poisson algebra (A , ·, {· , ·}) there exists a family of associative products t
on A , such that (A , ·) = (A , 0 ) and such that {· , ·}, obtained from t by the above
procedure, is a Poisson bracket on A . The second question is to classify all such de-
formations, for a given (A , ·, {· , ·}). The physical interpretation of these questions
is that, starting from the algebra of classical variables, which inherits from phase
space a Poisson bracket, as explained above in the context of mechanical systems,
one wishes to (re-)construct the corresponding algebra of quantum mechanical op-
erators.
A mathematical simplification of this problem is to ask whether the above pro-
cedure can be inverted formally. Stated in a precise way, this means the following.
Let (A , ·, {· , ·}) be a Poisson algebra and consider the vector space A [[ν ]] of for-
mal power series in one variable ν , whose coefficients belong to A . Let μ denote
an associative product on A [[ν ]], i.e., an F[[ν ]]-bilinear map from A [[ν ]] to itself,
which is associative. Given a, b ∈ A , we can write
μ (a, b) = ∑ μk (a, b) ν k ,
k∈N
∂ kF ∂ kG νk
μ (F, G) := ∑ ∑ Ji1 , j1 . . . Jik , jk
∂ xi1 . . . ∂ xir ∂ x j1 . . . ∂ x jr k!
,
k∈N 1i1 , j1 ,...,ik , jk 2r
0 1r
where J is the skew-symmetric matrix of size 2r, given by J := . The
−1r 0
Moyal product, also called the Moyal–Weyl product, defines an associative product
on C∞ (R2r )[[ν ]], whose leading terms satisfy μ0 (F, G) = FG and
r
∂F ∂G ∂G ∂F
μ1 (F, G) − μ1 (G, F) = ∑ − ,
j=1 ∂qj ∂ pj ∂qj ∂ pj
where the latter right-hand side is the Poisson bracket (0.1). This shows that the
classical Poisson bracket admits a formal deformation.
It was shown by De Wilde and Lecomte in [57] that the Poisson algebra of an
arbitrary symplectic manifold (regular Poisson manifolds of maximal rank) admits
a formal deformation. A geometrical proof of their result, which also works in the
case of arbitrary regular Poisson manifolds, was given by Fedosov [73]. Finally, it
was Kontsevich [107] who proved, using ideas which come from string theory, that
the algebra of functions on any Poisson manifold admits a formal deformation. He
proved in fact a quite stronger result, which says that for a given Poisson manifold
(M, π ), the equivalence classes of formal deformations of the algebra of smooth
functions on M are classified by the equivalence classes of formal deformations of
the Poisson structure π .
Summarizing, when one deforms a commutative associative algebra, a Poisson
structure shows up and it plays a dominant rôle in the entire deformation process.
The examples. The main idea which led us to the writing of this book, and to its ac-
tual structure, is that Poisson structures come in big classes (families) where roughly
speaking each class has its own tools and is related to a very definite part of mathe-
matics, hence leading to specific questions. As a consequence, the main part of the
book (about half of it) is Part II, dedicated to six classes of examples. They make
their appearance in six different chapters, namely Chapters 6–11, as follows.
• Constant Poisson structures, such as Poisson’s original bracket (0.1), are accord-
ing to the Darboux theorem the model (normal form) for Poisson structures on real
or complex manifolds, in the neighborhood of any point where the rank is locally
constant. Regular Poisson manifolds and symplectic manifolds (such as cotangent
bundles and Kähler manifolds) are the main examples, which exhibit even in the
absence of singularities some of the main phenomena in Poisson geometry, which
distinguish it from Riemannian geometry and show its pertinence for classical and
quantum mechanics. Constant and regular Poisson structures are studied in Chap-
ter 6.
Introduction xxi
• Linear Poisson structures are in one-to-one correspondence with Lie algebras and
many features of Lie algebras are most naturally approached in terms of the canon-
ical Poisson bracket on the dual of a (finite-dimensional) Lie algebra, the so-called
Lie–Poisson structure. The coadjoint orbits, for example, of a Lie algebra are the
symplectic leaves of the Lie–Poisson structure, showing on the one hand that they
are even dimensional, and on the other hand that they carry a canonical symplectic
structure, the Kostant–Kirillov–Souriau symplectic structure. Linear Poisson struc-
tures appear also on an arbitrary Poisson manifold (M, π ) at any point x where the
Poisson structure vanishes: the tangent space Tx M inherits a linear Poisson struc-
ture, the linearization of π at x. It leads to the linearization problem, which inquires
if π and its linearization at x are isomorphic, at least on a neighborhood of x. Linear
Poisson structures and their relation to Lie algebras are the subject of Chapter 7.
• Many Poisson structures of interest are neither constant nor linear (or affine). In
fact, while the basic properties of the latter Poisson structures can be traced back to
(known) properties of bilinear forms and of Lie algebras, new phenomena appear
when considering quadratic Poisson structures (i.e., homogeneous Poisson struc-
tures of degree two) and higher degree Poisson structures, which often turn out to
be weight homogeneous. A prime example of this is the transverse Poisson struc-
ture to an adjoint orbit in a semi-simple Lie algebra. Moreover, this Poisson structure
arises in the case of the subregular orbit from a Nambu–Poisson structure, defined
by the invariant functions of the Lie algebra. Higher degree Poisson structures are
studied in Chapter 8.
• In dimension two, every bivector field is a Poisson structure, yet many questions
about Poisson structures on surfaces are non-trivial. For example, their local clas-
sification has up to now only been accomplished under quite strong regularity as-
sumptions on the singular locus of the Poisson structure. In dimension three, the
Jacobi identity can be stated as an integrability condition of a distribution (or of
a one-form), which eventually leads to the symplectic foliation. A surface in C3
which is defined by the zero locus of a polynomial ϕ inherits a Poisson structure
from the standard Nambu–Poisson structure on C3 , with ϕ as Casimir. In the case
of singular surfaces C2 /G, where G is a finite subgroup of SL2 (C), this Poisson
structure coincides with the Poisson structure obtained from the canonical symplec-
tic structure on C2 by reduction. Thus, rather than being trivial, Poisson structures in
dimensions two and three form a rich playground on which a variety of phenomena
can be observed. Poisson structures in dimensions two and three are discussed in
Chapter 9.
• In a Lie-theoretical context, a linear Poisson structure, different from the canoni-
cal Lie–Poisson structure, often pops up. This Poisson structure appears on the Lie
algebra g at hand, so that the dual g∗ of g comes equipped with a Lie algebra struc-
ture, or it appears on g∗ , so that g is equipped with a second Lie algebra structure.
The underlying operator, which relates either of these Lie structures with the orig-
inal Lie bracket on g is in the first case an element r ∈ g ⊗ g, called an r-matrix,
while it is in the second case a (vector space) endomorphism R of g, called an R-
matrix. These structures appeared first in the theory of integrable systems, but are
xxii Introduction
nowadays equally important in the theory of Lie–Poisson groups (see the next item)
and of quantum groups. In the context of Lie algebras which come from associative
algebras, r-matrices and R-matrices also lead to a quadratic Poisson structure on
the Lie algebra, which plays an important rôle in the theory of Lie–Poisson groups;
they also lead to a cubic Poisson structure, whose virtue seems at this point still very
mysterious. Poisson structures coming from r-matrices or R-matrices are the subject
of Chapter 10.
• A Poisson–Lie group is a Lie group G, equipped with a Poisson structure π for
which the group multiplication G × G → G is a Poisson map. The Poisson struc-
ture π leads to a Lie algebra structure on the dual of the Lie algebra g of G, making g
into a Lie bialgebra, a structure generalizing the structure which comes from an r-
matrix. Conversely, every finite-dimensional Lie bialgebra is obtained in this way
from a Poisson–Lie group, a result which extends Lie’s third theorem (every finite-
dimensional Lie algebra is the Lie algebra of a Lie group). Poisson–Lie groups have
many applications, for example in the description of the Schubert cells of a Lie
group. Poisson–Lie groups are discussed in Chapter 11.
jects and properties which are recalled in these appendices are used throughout the
book.
Chapters 1 and 2 contain the basic definitions and constructions. A good famil-
iarity with these chapters should suffice for reading a good part of any other chapter
of the book. Chapter 3 deals with multi-derivations and their geometrical analog,
multivector fields, and with Kähler forms, which are the algebraic analogs of dif-
ferential forms. Chapter 4 deals with Poisson cohomology and Chapter 5 is devoted
to reduction in the context of Poisson structures. At the end of each of these five
chapters, we give a list of exercises, so that the reader can test his understanding of
the theory.
Applications. The third part of the book contains the two major applications, which
we discussed above: integrable systems are discussed in Chapter 12, while de-
formation quantization is detailed in Chapter 13. They are however not the only
applications which are given in the book. In fact, we end each example chapter
(Chapters 6–11) with an application of the class of Poisson structures, studied in
that chapter.
What is absent from this book. The main topic about Poisson structures which is
absent from this book is what should be called “Poisson geometry”. By this we mean
the global geometry of Poisson structures, which involves the integration of Poisson
brackets, Poisson connections, stability of symplectic leaves and related topics. For
this, we refer to [51, 53, 74].
Equally absent from this book are the many generalizations of Poisson structures.
Quasi-Poisson structures and Poisson structures with background are not only math-
ematically speaking very interesting, see [112], they have in addition non-trivial ap-
plications in physics, see [119, 162, 176], and are themselves a particular case of
Dirac structures [48, 89]. The theory of quantum groups, initiated in the seminal
ICM talk [59] by Drinfel’d, has several connections with the topics which are de-
veloped in this book, but adding several chapters to the book would not have been
sufficient to give a fair account of this subject; we simply refer to the excellent books
[40, 103].
Equally absent from this book is the theory of Lie algebroids, introduced by
Pradines [171], which are both a particular case and a generalization of Poisson
manifolds. In particular, after the pioneering work of [47], the problem of the in-
tegration of Poisson manifolds, which claims that symplectic Lie groupoids are to
Poisson manifolds what Lie groups are to Lie algebras, has gained a long attention.
For a definite solution of this problem, given by Cranic and Fernandes, see [49, 50].
The notes and the references. When writing the history of the theory of Poisson
brackets, giving each of the main results in the theory and examples the “right”
credit is a challenge which is beyond the scope of this book. We claim no originality
in this book, but since none of the proofs which are given are copy-and-paste proofs
from existing proofs, we did not think it was necessary to cite the proof in the lit-
erature which is closest to the proof that we give. Also, many properties of Poisson
xxiv Introduction
structures, and many examples, have been through a long series of transformations
and generalizations, think for example of Weinstein’s splitting theorem or the Pois-
son and Poisson–Dirac reduction theorems; structuring the long list of important
intermediate results which led to the final results is certainly interesting from the
epistemological point of view, but this was not our main focus when writing the
book.
However, at the end of each chapter, we have put a section called “Notes”, in
which we indicate some references, including the main references which we know
of, for further reading on what has been treated in the chapter. These notes are also
used to situate, a posteriori, the treated subjects in the scientific literature and to trace
some of their historical developments. Finally, the notes are also used to indicate
some further connections to other fields of mathematics or physics. Since Poisson
structures are a very active field of research, it is clear that the list of connections
with other fields is not exhaustive.
Part I
Theoretical Background
Chapter 1
Poisson Structures: Basic Definitions
The compatibility between the two algebra structures is precisely what allows us
to associate to elements of A , derivations of A , derivations being the algebraic
analogs of vector fields.
Poisson algebras and their morphisms are defined in Section 1.1. In Section 1.2,
respectively Section 1.3, we introduce Poisson varieties, respectively Poisson man-
ifolds, and their morphisms. Sections 1.2 and 1.3 can be read independently and
in either order; the reader is invited to discover, by comparing them, up to which
point Poisson varieties and Poisson manifolds can be treated uniformly, and how
quickly the techniques and results diverge, past this point. In Section 1.4 we spe-
cialize the results of Sections 1.2 and 1.3 to the case of Poisson structures on a
finite-dimensional vector space; such a space can indeed be viewed as an affine
variety or as a manifold, depending on the choice of algebra of functions on it.
Throughout the present chapter, we fix a ground field F of characteristic zero,
which the reader may think of as being R or C, especially in the context of varieties.
In this section we introduce the general notions of a Poisson algebra and of a mor-
phism of Poisson algebras. The Hamiltonian operator, which associates to an ele-
ment of the Poisson algebra a derivation of the Poisson algebra, will be introduced
and its basic algebraic properties will be given.
We start with the algebraic notion of a Poisson algebra over a field F which, as we
have already said, is always assumed to be of characteristic zero.
Definition 1.1. A Poisson algebra is an F-vector space A equipped with two mul-
tiplications (F, G) → F · G and (F, G) → {F, G}, such that
(1) (A , ·) is a commutative associative algebra over F, with unit 1;
(2) (A , {· , ·}) is a Lie algebra over F;
(3) The two multiplications are compatible in the sense that
{F, {G, H}} + {G, {H, F}} + {H, {F, G}} = 0 (1.2)
is verified for all triples (F, G, H) of elements of A . In order to shorten the formu-
las, (1.2) is often written as
and similarly for other formulas which involve sums over three terms with cycli-
cally permuted variables. The algebraic terminology which expresses (1.1) is that
for every H ∈ A the linear map F → {F, H} is a derivation of A . Namely, a linear
map V : A → A is called a derivation of A (with values in A ) if
We now turn to the notion of a morphism between two Poisson algebras (defined
over the same field F). Since a Poisson algebra is an algebra in two different ways,
it is natural to demand that a morphism of Poisson algebras be a morphism with
respect to both algebra structures.
Definition 1.2. Let (Ai , ·i , {· , ·}i ), i = 1, 2, be two Poisson algebras over F. A linear
map φ : A1 → A2 which satisfies, for all F, G ∈ A1 ,
(1) φ (F ·1 G) = φ (F) ·2 φ (G);
(2) φ ({F, G}1 ) = {φ (F), φ (G)}2 ,
is called a morphism of Poisson algebras.
It is clear that if φ : A1 → A2 is a morphism of Poisson algebras and φ is bijective,
then φ −1 : A2 → A1 is also a morphism of Poisson algebras. We say then that φ is
an isomorphism of Poisson algebras.
6 1 Poisson Structures: Basic Definitions
As we will see in the case of Poisson manifolds and of Poisson varieties, one is
often led to morphisms of the associative algebras which underlie the two Poisson
algebras. In this case, we only have to consider condition (2) in the above definition,
which says that φ is a morphism of Lie algebras.
When one deals with subalgebras or ideals of Poisson algebras, it is important
to distinguish which multiplication (maybe both) is considered. Our convention is
that subalgebra and ideal refer to the associative multiplication, Lie subalgebra and
Lie ideal refer to the Lie bracket, and Poisson subalgebra and Poisson ideal refer to
both. This means that a vector subspace B ⊂ A is a Poisson subalgebra of A =
(A , ·, {· , ·}) if
B·B ⊂ B and {B, B} ⊂ B , (1.5)
while a vector subspace I ⊂ A is a Poisson ideal of A if
I ·A ⊂ I and {I , A } ⊂ I . (1.6)
In the first case, B becomes itself a Poisson algebra, when equipped with the restric-
tions of both multiplications to B and the inclusion map ı : B → A is a morphism of
Poisson algebras. In the second case, A /I inherits a Poisson bracket from A , such
that the canonical projection p : A → A /I is a morphism of Poisson algebras. We
will come back at length to the notions of Poisson subalgebra and Poisson ideal in
Chapter 2, when we have developed sufficient concepts and tools to illustrate these
notions geometrically.
In view of Definitions 1.1 and 1.2 we get, for a fixed field F, a category whose
objects are the Poisson algebras over F and whose morphisms are the morphisms of
Poisson algebras.
Definition 1.3. Let (A , ·, {· , ·}) be a Poisson algebra and let H ∈ A . The derivation
XH := {· , H} of A is called a Hamiltonian derivation and we call H a Hamiltonian,
associated to XH . We write
X : A → Ham(A , {· , ·})
(1.7)
H → XH := {· , H} .
An element H ∈ A whose Hamiltonian derivation is zero, XH = 0, is called a
Casimir and we denote
Cas (A , {· , ·}) := {H ∈ A | XH = 0}
for the F-vector space of Casimirs. When no confusion can arise, we write Ham(A )
for Ham(A , {· , ·}) and Cas(A ) for Cas(A , {· , ·}). It is clear that Cas(A ) is the
center of the Lie algebra (A , {· , ·}).
The defining properties of the Poisson bracket lead to the following proposition.
Proposition 1.4. Let (A , ·, {· , ·}) be a Poisson algebra.
(1) Cas(A ) is a subalgebra of (A , ·), which contains the image of F in A , under
the natural inclusion a → a · 1;
(2) If A has no zero divisors, then Cas(A ) is integrally closed in A ;
(3) Ham(A ) is not an A -module (in general); instead, XFG = FXG + GXF , for
every F, G ∈ A ;
(4) Ham(A ) is a Cas(A )-module;
(5) The map A → X1 (A ) which is defined by H → −XH is a morphism of Lie
algebras; as a consequence, Ham(A ) is a Lie subalgebra of X1 (A );
(6) The Lie algebra sequence
−X
0 −→ Cas(A ) −→ A −→ Ham(A ) −→ 0
1 As we said in the introduction to this chapter, Sections 1.2 and 1.3 can be read independently.
1.2 Poisson Varieties 9
Proof. The proof is based on the standard argument that two biderivations (or p-
derivations, in general) of some commutative associative algebra A are equal as
soon as they agree on a system of generators for A . Since we will use this argument
several times, we spell it out in detail. Both sides of (1.8) are bilinear in F and G,
so it suffices to show (1.8) when F and G are monomials in x1 , . . . , xd . If F or G is a
monomial of total degree 0, then the right-hand side in (1.8) is obviously zero, but
also the left-hand side is zero, because constant functions are Casimirs (item (1) in
Proposition 1.4). Also, the fact that (1.8) holds when F and G are both monomials of
degree 1, is clear. Suppose now that (1.8) holds when deg(F)+deg(G) n, for some
n 2; we show that it holds for all F and G such that deg(F) + deg(G) = n + 1. Let
F and G be non-constant monomials, such that deg(F) + deg(G) = n + 1. By skew-
symmetry, we may assume that deg(F) > 1. There exist monomials F1 , F2 ∈ A ,
with deg(F1 ) < deg(F) and deg(F2 ) < deg(F), such that F = F1 F2 . Since {· , ·} is a
biderivation and in view of the recursion hypothesis, we have that
Ψ ∗ (FG) = (FG) ◦ Ψ = (F ◦ Ψ ) (G ◦ Ψ ) = (Ψ ∗ F) (Ψ ∗ G) .
In this section, we show how a Poisson structure on an affine variety can be encoded
in a matrix. We first consider the case of a Poisson structure {· , ·} on the affine
space Fd (with its algebra of regular functions A := F[x1 , . . . , xd ]). The 2
d structure
functions xi j ∈ A , which are defined for 1 i, j d by xi j := xi , x j , satisfy
xi j = −x ji , (1.9)
d ∂ x jk
∂ xi j ∂ xki
∑ xk ∂ x + xi ∂ x + x j ∂ x = 0 , (1.10)
=1
for all 1 i, j, k d. Formula (1.10) is obtained by writing the Jacobi identity (1.2)
for the triple (xi , x j , xk ) in the form
xi j , xk + x jk , xi + xki , x j = 0 ,
so that the Jacobi identity is satisfied for all triples of elements of A = F[x1 , . . . , xd ]
if and only if (1.10) holds, i.e., if and only if the Jacobi identity holds for every triple
(xi , x j , xk ), with 1 i < j < k d.
Summarizing, we have the following proposition.
Proposition 1.8. If X = (xi j ) is a skew-symmetric d × d matrix, with elements in
A = F[x1 , . . . , xd ], then
d
∂F ∂G
{F, G} := ∑ xi j (1.12)
i, j=1 ∂ xi ∂ x j
defines a Poisson bracket on A (with Poisson matrix X) if and only if every triple
(xi , x j , xk ), with 1 i < j < k d, satisfies the Jacobi identity
xi , x j , xk + x j , xk , xi + {xk , xi } , x j = 0 ,
We
would like to define a skew-symmetric biderivation {· , ·} of A by setting
F, G := {F, G}0 , for all F, G ∈ F[x1 , . . . , xd ]. The problem is that, in general, this
is not well-defined. Namely, for every K ∈ I one has that K = 0, so that a necessary
condition for the bracket {· , ·} on A to be well-defined, is that for every K ∈ I and
1 i d,
d
∂K
{xi , K}0 = ∑ xi j ∈I . (1.14)
j=1 ∂xj
12 1 Poisson Structures: Basic Definitions
for every 1 i < j < k d. Again, notice that this condition is independent of the
chosen elements xi j ∈ F[x1 , . . . , xd ], which represent xi j .
• Suppose that {· , ·} is a Poisson bracket on A = F[x1 , . . . , xd ]/I . We can
choose elements
xi j ∈ F[x1 , . . . , xd ] such that x ji = −xi j for all i and j, and such
that xi j = xi , x j . Define a skew-symmetric biderivation {· , ·}0 on F[x1 , . . . , xd ]
by (1.13). Since {xi , ·} is a derivation of A , we have that
d ∂K d
∂K
0 = xi , K = ∑ xi , x j
∂xj
= ∑ xi j ∂ x j = {xi , K}0 ,
j=1 j=1
for every K ∈ I , so that condition (1.14) holds. This means that {· , ·} , defined
for all F, G ∈ A by F, G := {F, G}0 , is a biderivation of A , which coincides
with {· , ·} on a system of generators (x1 , . .. , xd ) of A . Thus, {· , ·} = {· , ·} and
we can compute the Poisson bracket F, G by using (1.13), as in the case of a
polynomial algebra.
We summarize these results in the following proposition.
Proposition 1.9. Let M be an affine variety in Fd , with algebra of regular functions
A = F[x1 , . . . , xd ]/I ,
d ∂F ∂G
F, G = ∑ xi , x j
∂ xi ∂ x j
, (1.18)
i, j=1
have the same rank. Since x0 can be written as a polynomial in the generators
x1 , . . . , xd of F (M), say x0 = F(x1 , . . . , xd ), we have in view of (1.18),
d ∂F
{xi , x0 } (m) = ∑ xi , x j (m)
∂xj
(m) ,
j=1
and so the zeroth column of X (m) is a linear combination of the other columns of
X (m), i.e., of the columns of X(m). Thus, X (m) and X(m) have the same rank.
Definition 1.11. For a Poisson variety (M, {· , ·}) and a point m ∈ M, the rank of the
Poisson matrix of {· , ·} with respect to an arbitrary system of generators of F (M),
evaluated at m, is called the rank of {· , ·} at m, denoted Rkm {· , ·}. The maximum
maxm∈M Rkm {· , ·} is called the rank of {· , ·}, denoted Rk {· , ·}. A point m ∈ M is
said to be a regular point of (M, {· , ·}), if Rkm {· , ·} = Rk {· , ·}, otherwise it is said
to be a singular point of (M, {· , ·}). The set of singular points of (M, {· , ·}) is called
the singular locus of (M, {· , ·}).
A more intrinsic definition of the rank of a Poisson structure can be given in terms of
the (Zariski) cotangent space to M at m. Recall that this vector space is intrinsically
defined as Im /Im2 , where Im denotes the ideal of F (M), consisting of all functions
which vanish at m; its dual space is the (Zariski) tangent space to M at m. We
denote the tangent space to M at m by Tm M, while the cotangent space is denoted
by Tm∗ M. Taking generators of F (M) which vanish at m, the ideal Im corresponds
to elements of F (M) “without constant term”, so that F (M)/Im F, in a natural
way (i.e., by evaluation at m). A point m of M where the dimension of Tm M attains
its minimal value is called a smooth point of M. The smooth points of M form a
Zariski open subset of M and the dimension of Tm M in a smooth point m of M is
called the dimension of M, denoted dim M.
By the biderivation property (1.1), we have that Im2 , Im ⊂ Im . Therefore, the
bracket {·, ·} induces for every m ∈ M a well-defined skew-symmetric bilinear map:
I m Im F (M)
πm : × → F. (1.19)
Im2 Im2 Im
i.e., the resulting matrix is nothing but the Poisson matrix of {· , ·}, with respect to
x1 , . . . , xd , evaluated at m. It follows that the rank of {· , ·} at m can also be defined
as the rank of the intrinsically defined bilinear form πm on Tm∗ M.
The main properties of the rank are given in the following proposition.
1.3 Poisson Manifolds 15
X :M→ gld
(1.20)
m → xi , x j (m) 1i, jd
it is open; since the topology which is considered here is the Zariski topology, these
open subsets are dense as soon as they are non-empty. This yields the proof of (3).
Finally, let m be an arbitrary point of M and denote the rank of {· , ·} at m by 2r.
Consider a point m of M(r) , which is a smooth point of M; such a point exists
because M(r) and the set of smooth points of M are both dense subsets of M. At
such a point m , the dimension of the (co)tangent space coincides with the dimension
of M, so that
Rkm {· , ·} Rkm {· , ·} dim Im /Im2 = dim M .
from differential geometry which we will use, referring the reader who needs more
details to Appendix B at the end of the book.
We consider both real manifolds and complex manifolds, since many constructions
in Poisson geometry apply in the same way if all manifolds are considered real or
if they are all considered complex; we will therefore make simple statements such
as “Let M and N be two manifolds and let Ψ : M → N be a map”, which the reader
may specialize to “Let M and N be two differentiable manifolds and let Ψ : M → N
be a smooth map” or to “Let M and N be two complex manifolds and let Ψ : M → N
be a holomorphic map”. Similarly, dim M stands for the real or complex dimension,
according to the context, F (M) stands for the appropriate algebra of functions, F
stands for R or C, and so on. In both cases, the coordinates, real or complex, will be
denoted by x = (x1 , . . . , xd ), where d = dim M.
For m ∈ M the tangent space to M at m is denoted by Tm M. Elements of Tm M
are by definition pointwise derivation at m, i.e., they are linear forms on the vector
space of all function germs at m, satisfying, for all functions F and G, defined on a
neighborhood of m in M,
in this formula, Fm stands for the germ of F at m. The dual space to Tm M is the
cotangent space, denoted Tm∗ M. The canonical pairing between Tm M and Tm∗ M is
denoted by · , ·. A map Ψ : M → N between manifolds leads for every m ∈ M
to a linear map TmΨ : Tm M → TΨ (m) N, called the tangent map of Ψ at m. Upon
identifying the tangent spaces to F with F, the tangent map leads to the differential
of a function; specifically, if F is a function on M, then we view the differential of
F at m, denoted dm F as a linear function on Tm M, that is, as an element of Tm∗ M.
We denote by X1 (M) the F (M)-module of vector fields on M and for V ∈ X1 (M)
and m ∈ M we denote by Vm the value of V at m (which is an element of Tm M). We
often use (and define) vector fields on M through their action on (local) functions:
when U is a non-empty open subset of M and V is a vector field on M (or on U)
then V [F] denotes the function on U, defined by
V [F] : U → F
(1.21)
m → Vm Fm .
Notice that we use square brackets to denote the action of a vector field on a func-
tion, as we did in the case of a derivation; this notation will be extended to bivector
fields in this section, and to multivector fields in Chapter 3.
We now come to the definition of a bivector field on a manifold. To do this, we
first introduce the notion of a pointwise biderivation.
Definition 1.13. Let M be a manifold and let m ∈ M. A bilinear map
Fm (M) Fm (M)
Bm : × →F
∼ ∼
is a pointwise biderivation of F (M) at m, if for all functions F, G and H, defined in
a neighborhood of m in M,
form:2
∂ ∂
P= ∑ P[xi , x j ] ∧
∂ xi ∂ x j
. (1.23)
1i< jd
In differential geometric terms, one often refers to bivector fields as being (skew-
symmetric) tensors (of type (2, 0)), which refers to the fact that they are F (M)-linear
objects. In view of the biderivation property for a bivector field, this may seem
absurd, but the point is that, as a tensor, a bivector field acts on a pair of differential
one-forms, rather than on a pair of functions. To explain this, notice that the value
of a bivector field P on two elements F and G of F (M), at m ∈ M, depends on the
differentials dm F and dm G only. Indeed, (1.23) yields
d
∂F ∂G
P[F, G](m) = ∑ P[xi , x j ](m)
∂ xi
(m)
∂xj
(m) , (1.24)
i, j=1
We will in the sequel usually make no distinction between the bivector field P and
the (2, 0)-tensor P̂, and denote both by the same letter P. Note also that the tensorial
interpretation, or equivalently (1.23), implies that we may specialize P to points
m ∈ M, giving a skew-symmetric, F-bilinear form
for all m ∈ M.
2 See (B.8) for the coordinate expression of a vector field, of which (1.23) is a direct generalization.
3 For a linear map φ : V → W between two vector spaces, the linear map ∧2 φ : ∧2V → ∧2W is
defined by (∧2 φ )(v ∧ w) := φ (v) ∧ φ (w), for all v, w ∈ V . See Appendix A for details and general-
izations.
1.3 Poisson Manifolds 19
Remark 1.14. Vector fields on a manifold M can be defined as sections of its tan-
gent bundle, T M, the vector bundle over M, whose fiber over m ∈ M is the vector
space Tm M. Similarly, bivector fields on M can be defined as (smooth or holomor-
phic) sections of ∧2 T M → M, where the latter vector bundle has as fiber over m ∈ M
the vector space ∧2 Tm M.
We are now ready to define the notion of a Poisson structure on a (real or complex)
manifold M. The idea is that a bivector field π on M leads for every open subset U
of M to a skew-symmetric product on F (U), which is demanded to make F (U)
into a Poisson algebra.
Definition 1.15. Let π be a bivector field on a manifold M. We say that π is a Pois-
son structure on M if for every open subset U of M, the restriction of π to U makes
F (U) into a Poisson algebra. We then call (M, π ) a Poisson manifold.
The Poisson structure π of a Poisson manifold (M, π ) will often be denoted by {· , ·},
in particular the Poisson bracket π [F, G] of two functions F, G, defined on an open
subset U of M, will usually be denoted by {F, G}. The notation {F, G} for π [F, G]
will be referred to as the bracket notation.
Let (U, x) be a coordinate neighborhood of a d-dimensional Poisson manifold
(M, π ). On U, the bivector field π can, according to (1.23), be written as
∂ ∂
π= ∑ xi , x j
∂ x i
∧
∂ xj
, (1.26)
1i< jd
so that the matrix X := xi , x j 1i, jd encodes the restriction of π to U. This
matrix X, whose elements belong to F (U), is called the Poisson matrix of π with
respect to the coordinates x1 , . . . , xd . It is the differential geometric analog of the
Poisson matrix, introduced in Section 1.2.2 in the context of affine Poisson varieties.
Evaluating the Poisson matrix X at m ∈ M, we get the matrix of the skew-symmetric
bilinear map πm : Tm∗ M × Tm∗ M → F, defined by the Poisson structure at m.
Proposition 1.16. Let π = {· , ·} be a bivector field on a manifold M of dimension d.
Then π is a Poisson structure on M if and only if one of the following equivalent
conditions holds.
(i) For every open subset U of M and for all functions F, G, H ∈ F (U), the
Jacobi identity holds:
(ii) For some collection of coordinate charts (U, x) of M which cover M and for
all functions F, G, H ∈ F (U), the Jacobi identity (1.27) holds;
20 1 Poisson Structures: Basic Definitions
(iii) For some collection of coordinate charts (U, x) of M which cover M and for
all i, j, k with 1 i < j < k d the following equality holds:
d
∂ xi j ∂ x jk ∂ xki
∑ xk
∂ x
+ xi
∂ x
+ x j
∂ x
=0, (1.28)
=1
where xi j := xi , x j , for 1 i, j d;
(iv) For some collection of coordinate charts (U, x) of M which cover M and for
all i, j, k with 1 i < j < k d, the Jacobi identity holds:
xi , x j , xk + x j , xk , xi + {xk , xi } , x j = 0 . (1.29)
Proof. Clearly, (i) holds if and only if π makes every F (U) into a Poisson algebra,
hence π is a Poisson structure if and only if (i) holds. (ii) is a consequence of (i), by
specialization. The equivalence of (ii) and (iii) follows from the fact that
{{F, G} , H} + (F, G, H)
d
∂ xi j ∂ F ∂ G ∂ H
= ∑ xk
∂ x ∂ xi ∂ x j ∂ xk
+ (F, G, H) (1.30)
i, j,k,=1
d
d
∂ xi j ∂F ∂G ∂H
= ∑ ∑ xk + (i, j, k) ,
i, j,k=1 =1 ∂ x ∂ xi ∂ x j ∂ xk
where we used (1.26) to compute the Poisson brackets. Notice that this computation
shows that the value of (1.30) at a point m ∈ U depends only on F, G and H in a
neighborhood of m, in other words on the germs Fm , Gm and Hm . Therefore, if the
Jacobi identity holds on a coordinate neighborhood (U, x), then it will hold on every
open subset V ⊂ U, which leads to the implication (ii) ⇒ (i). The equivalence of
(iii) and (iv) is clear because (1.28) is just (1.29) written out explicitly, using (1.26).
Remark 1.17. In the case of a real manifold, every germ is the germ of a globally
defined function, hence it suffices to verify the Jacobi identity for all triples of func-
tions on M. For complex manifolds, this is not sufficient, see Remark B.4.
We now turn to the notion of a Poisson map between two Poisson manifolds.
Definition 1.18. A map Ψ : M → N between Poisson manifolds (M, π ) and (N, π )
is called a Poisson map if for all open subsets U ⊂ M and V ⊂ N, with Ψ (U) ⊂ V , the
induced map Ψ ∗ : F (V ) → F (U), which is defined for all F ∈ F (V ) by Ψ ∗ (F) :=
F ◦ Ψ , is a morphism of Poisson algebras, i.e., for all functions F, G ∈ F (V ),
∧2 (T Ψ )π = π , (1.32)
so that ∧2 (TmΨ )πm = πΨ (m) if and only if {F ◦ Ψ , G ◦ Ψ } (m) = {F, G} (Ψ (m)) for
all functions F and G, defined in a neighborhood of Ψ (m). This proves that Ψ is a
Poisson map if and only if ∧2 (T Ψ )π = π .
It follows that the first order differential equation, which describes the integral
curves of XH in a coordinate chart (U, x), is given by
22 1 Poisson Structures: Basic Definitions
dxi d ∂H
dt
= {xi , H} = ∑ xi , x j
∂xj
, i = 1, . . . , d . (1.34)
j=1
Definition 1.20. For a Poisson manifold (M, π ) and a point m ∈ M the integer Rk πm ,
which is also the rank of every Poisson matrix of π at m, is called the rank of π at m,
denoted Rkm π . One says that (M, π ) is a regular Poisson manifold when its rank is
constant (independent of m ∈ M), and that the rank is locally constant at m, when it
is constant on some neighborhood of m in M. The rank of π is said to be maximal at
m ∈ M, when it coincides with the dimension of M. The maximum maxm∈M Rkm π
is called the rank of π , denoted Rk π . A point m ∈ M is said to be a regular point
of (M, π ), if Rkm π = Rk π , otherwise it is said to be a singular point of (M, {· , ·}).
The set of singular points of (M, π ) is called the singular locus of (M, π ).
The main properties of the rank are given in the following proposition.
Proposition 1.21. Let (M, π ) be a Poisson manifold and let m ∈ M.
(1) Rkm π is even and is equal to dim Hamm (M);
(2) For every s ∈ N, the subset M(s) of M, defined by
is open (and dense in the complex case); in particular, the subset of points
m ∈ M such that Rkm π = Rk π is open.
Proof. The rank of π at m is the rank of the bilinear map πm , so it is the dimension
of the image of the linear map Tm∗ M → Tm M, defined by ξ → πm (ξ , ·). In turn,
the latter equals the dimension of Hamm (M), since every ξ ∈ Tm∗ M can be written
as dm F, for some function F, defined in a neighborhood of m. Also, since πm is
skew-symmetric, its rank is even. This shows (1). The set Rs of d × d matrices of
rank greater than or equal to 2s is an open subset of gld . Let m be a point in M(s) and
let (x1 , . . . , xd ) be local coordinates on a neighborhood U of m. The restriction to U
of M(s) is open since it is the inverse image of Rs by the continuous map X : U → gld
defined by m → xi , x j (m) 1i, jd . This yields the proof of (2).
In the real case, the rank of a Poisson structure does not necessarily attain its maxi-
mum on a dense open subset. This is shown in the following example.
Example 1.22. Let ϕ be a smooth function on R2 which is positive on the interior
of the unit disk and which is zero elsewhere. Then the rank of the Poisson structure
on R2 , which is defined by the bivector field ϕ ∂ /∂ x ∧ ∂ /∂ y, is of rank 2 only on
the inside of the disk.
1.3 Poisson Manifolds 23
To finish this section, we show that the flow of a (locally) Hamiltonian vector field
leaves the Poisson structure invariant.
Proposition 1.23. Let (M, π ) be a Poisson manifold. The Lie derivative of π with
respect to every (locally) Hamiltonian vector field is zero. As a consequence, the
flow of each (locally) Hamiltonian vector field preserves the Poisson structure.
for all F, G ∈ F (U) (see Section 3.3.4 for generalizations). For P := π = {· , ·} and
V := XH , with H ∈ F (U), this yields
A vector field V is called a Poisson vector field if taking the Lie derivative with
respect to V kills the Poisson structure, i.e., if LV π = 0. Proposition 1.23 implies
that all Hamiltonian vector fields are Poisson vector fields. This property will be
reformulated in cohomological terms in Section 4.1.
This section deals with Weinstein’s splitting theorem, which states that, in the neigh-
borhood of a point where the rank of the Poisson structure is 2r, the Poisson mani-
fold is a product of a symplectic manifold of dimension 2r, and a Poisson manifold
which has rank zero at the origin. It implies the Darboux theorem, which states that
all symplectic manifolds of the same dimension are locally isomorphic, “there are
no local invariants in symplectic geometry”, as well as a generalization to regular
Poisson manifolds.
Theorem 1.25 (Weinstein’s splitting theorem). Let (M, π ) be a (real or com-
plex) Poisson manifold, let m ∈ M be an arbitrary point and denote the rank
24 1 Poisson Structures: Basic Definitions
where the functions φk are (smooth or holomorphic) functions, which depend
on z = (z1 , . . . , zs ) only, and which vanish when z = 0. Such local coordinates
q1 , . . . , qr , p1 , . . . , pr , z1 , . . . , zs are called splitting coordinates, centered at m.
Proof. We use induction on r. For the case r = 0, it is clear that for every Poisson
manifold (M, π ) and for every point m such that the rank of π at m is zero, an
arbitrary system of local coordinates (z1 , . . . , zd ), centered at m, works (d := dim M).
Let r ∈ N∗ and assume that Theorem 1.25 holds true for every Poisson manifold, at
every point where the rank is 2(r − 1). Let (M, π ) be a Poisson manifold and let
m ∈ M be a point for which Rkm π = 2r. We will show that the theorem holds for
(M, π ) at m.
Since Rkm π > 0, there exists a function p on a neighborhood of m, whose Hamil-
tonian vector field X p does not vanish at m ∈ M; we may suppose that p(m) = 0.
Since X p (m) = 0, there exists by the straightening theorem (Theorem B.7) a sys-
tem of coordinates (q, y2 , . . . , yd ) on a neighborhood U of m, centered at m, such
that X p = ∂ /∂ q. Writing {· , ·} = π , it follows that
∂q
{q, p} = X p [q] = =1, [Xq , X p ] = X{p,q} = −X1 = 0 and X p [yi ] = 0 ,
∂q
∂ d
∂
X q = ξ1 + ∑ ξi ,
∂ q i=2 ∂ yi
d
∂p
−1 = {p, q} (m) = Xq [p](m) = ∑ ξi (m) (m) ,
i=2 ∂ yi
so that the vector field Xq = ∑di=2 ξi ∂ /∂ yi is independent of q and does not van-
ish at m. Applying the straightening theorem once more, we may introduce a sys-
tem of coordinates (q, p , y3 , . . . , yd ) on a neighborhood of m, centered at m, where
p , y3 , . . . , yd depend on y2 , . . . , yd only, with
∂ d
∂
= − ∑ ξi = −Xq .
∂p
i=2 ∂ yi
1.3 Poisson Manifolds 25
∂p
= −Xq [p] = {q, p} = 1 ,
∂ p
{q, p} = 1 ,
{q, yi } = −Xq [yi ] = ∂ yi /∂ p = 0 ,
{p, yi } = −X p [yi ] = −∂ yi /∂ q = 0 ,
∂ ∂ ∂ ∂
π= ∧ + ∑ {yk , y } ∧ . (1.37)
∂ q ∂ p 3k<d ∂ yk ∂ y
To show that {yk , y } is independent of p and q, for all values of k, , we just have
to check that {{yk , y } , p} = {{yk , y } , q} = 0, an easy consequence of the Jacobi
identity for π . The Jacobi identity also yields that the second term in (1.37) defines
a Poisson structure π on a neighborhood V of the origin o of Fd−2 . The Poisson
matrix of π with respect to q, p, y3 , . . . , yd has a diagonal block form, where the
lower block is the Poisson matrix of π with respect to y3 , . . . , yd . Thus, π has
rank 2(r − 1) at o, and by the induction hypothesis there exist local coordinates
q2 , . . . , qr , p2 , . . . , pr , z1 , . . . , zs centered at o, such that π takes on V the following
form:
r
∂ ∂ ∂ ∂
π = ∑ ∧ + ∑ φk (z) ∧ .
i=2 ∂ qi ∂ pi 1k<s ∂ zk ∂ z
In terms of the system of coordinates (q1 , q2 , . . . , qr , p1 , p2 , . . . , pr , z1 , . . . , zs ), which
is centered at m, π takes the required form (1.36), where we have set q1 := q and
p1 := p.
For a given point m in a Poisson manifold M, splitting coordinates are not unique.
We will see however in Section 5.3.3 that the Poisson structure, which is defined
in a neighborhood of z = 0 in Fs by the second term in (1.36), is unique, up to
isomorphism.
In terms of the splitting coordinates q1 , . . . , qr , p1 , . . . , pr , z1 , . . . , zs , the Poisson
structure (1.36) has the block form
⎛ ⎞
0 1r 0
⎝ −1r 0 0 ⎠
0 0 Φ
26 1 Poisson Structures: Basic Definitions
where Φ ∈ Mats (F (M)) is given by Φi j := zi , z j and 1r denotes the identity
matrix of size r. It follows that Φ = 0 in a neighborhood of m, when the rank of
π is locally constant (= 2r) at m. This leads to the following strengthening of the
splitting theorem for points at which the rank is locally constant.
Theorem 1.26 (Darboux theorem). Let (M, π ) be a (real or complex) Poisson
manifold of dimension d, and suppose that m is a point where the rank of π is
locally constant and equal to 2r. There exists a coordinate neighborhood U of m
with coordinates (q1 , . . . , qr , p1 , . . . , pr , z1 , . . . , zs ) such that, on U,
r
∂ ∂
π=∑ ∧ . (1.38)
i=1 ∂ qi ∂ pi
where each submanifold inherits a Poisson structure from the ambient Poisson struc-
ture; since this induced Poisson structure is of maximal rank, it is symplectic,4 so
one usually refers to this decomposition as the symplectic foliation of the Poisson
manifold. We combine two approaches, which are both due to Weinstein [199], to
obtain the symplectic foliation; both have their advantage, a fact that we exploit to
get at the decomposition in a natural way and to prove its properties with minimal
effort. For another approach, using the theory of generalized distributions, we refer
to [125]. Throughout this section (M, π ) is an arbitrary Poisson manifold.
The idea which underlies the decomposition is based on the following two ob-
servations. First, recall from Proposition 1.23 that the flow of every Hamiltonian
vector field preserves the Poisson structure, hence its rank at each point. Second,
the rank of the Poisson structure at a point is precisely the dimension of the space
of Hamiltonian vector fields, at that point. Thus, the flow of all Hamiltonian vector
fields, starting from a given point m ∈ M where the rank of the Poisson structure
is 2r, should trace out locally a submanifold of dimension 2r, on which the Poisson
structure has constant rank 2r. This leads, for m ∈ M, to the following definition of
a subset of M,
Sm (M) := m ∈ M | ∃ a piecewise Hamiltonian path in M from m to m .
manifold.
28 1 Poisson Structures: Basic Definitions
when the manifold topology on N is the induced topology). However, locally, with
respect to the manifold topology on N, every point in N has a neighborhood, which
sits in M in the same way an open subset of Fd sits in Fd (d = dim N d = dim M).
To construct this immersed submanifold of M, through m, we choose on a neigh-
borhood U of m, splitting coordinates (at m) which we denote by x = (q1 , . . . , qr ,
p1 , . . . , pr , z1 , . . . , zs ) and we define
Sm (U, x) := m ∈ U | z1 (m ) = · · · = zs (m ) = 0 .
where the functions φk (z) vanish for z = 0, it follows that at points m ∈ Sm (U, x),
the Poisson structure takes the form ∑ri=1 ∂∂qi ∧ ∂∂pi , which defines a Poisson structure
on Sm (U, x), which we denote by πS . The map ı : (Sm (U, x), πS ) → (M, π ) is a
Poisson map, since for all functions F, G ∈ F (U), and for every m ∈ Sm (U, x), we
have that
r
∂F ∂G ∂F ∂G
{F, G} (ı(m )) = {F, G} (m ) = ∑
(m ) (m ) − (m ) (m )
i=1 ∂ qi ∂ pi ∂ pi ∂ qi
= {F ◦ ı, G ◦ ı}S (m ) .
In the language of Section 2.2, (Sm (U, x), πS ) is a Poisson submanifold of (M, π ).
Notice that πS has constant rank 2r = dim Sm (U, x). In the following proposition
we show that, for each m ∈ M, the subsets Sm (U) and Sm (U, x) coincide, for small
enough U.
Proposition 1.29. Let m be a point in a Poisson manifold (M, π ) and let x =
(q1 , . . . , qr , p1 , . . . , pr , z1 , . . . , zs ) be splitting coordinates at m, on a coordinate neigh-
borhood U, with x(U) convex. Then Sm (U) = Sm (U, x).
Proof. For F ∈ F (U) and m ∈ Sm (U, x) we have, in view of (1.40), that
s
∂F
XF [zi ](m ) = {zi , F} (m ) = ∑ φi (m ) ∂ z (m ) = 0 ,
=1
since the functions φi vanish at all points m of Sm (U, x). It follows that all Hamil-
tonian vector fields of U are tangent to Sm (U, x) at points m ∈ Sm (U, x). Thus,
a (piecewise) Hamiltonian path in U which starts at m must stay in Sm (U, x). It
follows that Sm (U) ⊂ Sm (U, x). In order to show the other inclusion, suppose that
m ∈ Sm (U, x). We show that there exists a Hamiltonian path from m to m in U.
Consider the (linear) function H on U, given by
1.3 Poisson Manifolds 29
r
H := ∑ (qi (m )pi − pi (m )qi ) ,
i=1
Theorem 1.30. Every Poisson manifold (M, π ) is the disjoint union of immersed
submanifolds, whose tangent spaces are spanned by the Hamiltonian vector fields
of (M, π ). The Poisson structure, restricted to each of these submanifolds yields
a Poisson structure of maximal rank (symplectic structure). This decomposition is
called the symplectic foliation of M and the immersed submanifolds are called the
symplectic leaves of M. For m ∈ M the symplectic leaf which contains m is given by
Sm (M) = m ∈ M | ∃ a piecewise Hamiltonian path in M from m to m .
The first part of the theorem can also be restated by saying that the (singular)
distribution, defined by the Hamiltonian vector fields, is integrable (admits an inte-
30 1 Poisson Structures: Basic Definitions
gral manifold through each point), see [125, Appendix 3]. The theorem leads to the
following description of the symplectic leaves of a Poisson manifold.
Proposition 1.31. Let (M, π ) be a Poisson manifold of dimension d and rank 2r.
(1) A function F on M is a Casimir function if and only if F is constant on each
symplectic leaf;
(2) The (non-empty open subsets of the) symplectic leaves are the smallest em-
bedded manifolds of M which are Poisson submanifolds (i.e., such that the
inclusion map is a Poisson map);
(3) If Y is a Poisson vector field on M which is tangent to every symplectic leaf
of M, then Y is Hamiltonian in the neighborhood of every point m ∈ M where
the rank of π is 2r.
Proof. The fact that the manifold Sm (M) admits locally (in the sense of the topol-
ogy on Sm (M) which we have constructed) the description as a submanifold
Sm (U, x), leads at once to the proof of (1) and (2). We proceed to the proof of (3).
If the rank of π at m is 2r, so that m is a regular point of π , then there exists local
coordinates q1 , . . . , qr , p1 , . . . , pr , z1 , . . . , zd−2r in a contractible neighborhood U of m
with respect to which the Poisson structure π is given by:
r
∂ ∂
π=∑ ∧ . (1.41)
i=1 ∂ q i ∂ pi
The vector fields ∂∂q , ∂∂p , . . . , ∂∂qr , ∂∂pr span the symplectic leaves of π on U. There-
1 1
fore, every vector field Y , which is tangent to the symplectic leaves of π , is of the
form
r
∂ r
∂
Y = ∑ Fi + ∑ Gi (1.42)
i=1 ∂ p i i=1 ∂ qi
for some smooth or holomorphic functions F1 , . . . , Fr , G1 , . . . , Gr , defined on U. Sup-
pose now that Y is a Poisson vector field, so that LY π = 0. Then one computes
easily from (1.35), (1.41) and (1.42) that
∂ Fi ∂ Fj ∂ Gi ∂Gj ∂ Fi ∂ Gi
= , = and =− ,
∂qj ∂ qi ∂ pj ∂ pi ∂ pj ∂qj
Locally, the regular leaves (i.e., the leaves which pass through a point where the
rank is maximal, hence locally constant) are given as the level sets of the (local)
Casimir functions. This is shown in the following proposition.
Proposition 1.32. Let (M, π ) be a Poisson manifold of dimension d, let U be a non-
empty open subset of M and let F1 , . . . , Fs ∈ F (U), satisfying:
(1) The rank of π is constant on U and is equal to d − s;
(2) The functions F1 , . . . , Fs are Casimirs of the restriction of π to U;
(3) For every point m of U, the differentials dm F1 , . . . , dm Fs are independent.
Then the symplectic foliation of the restriction of π to U coincides with the foliation
which is defined on U by the map F := (F1 , . . . , Fs ) : M → Fs .
In good cases most or all of the symplectic leaves are level sets of the Casimir
functions, but this is not true in general.
Example 1.33. Consider the Poisson structure ∂ /∂ x ∧ ∂ /∂ y on R3 , where x, y, z are
the natural coordinates on R3 . Since translations over a constant vector are obviously
Poisson maps, it descends to a Poisson structure on every torus R3 /Λ , where Λ is a
lattice in R3 . Unless Λ is very special, all symplectic leaves are dense on the torus,
hence they cannot be the level sets of a (Casimir) function.
{F, G}1 (m) = {F ◦ ıS , G ◦ ıS }1,S (m) = {F ◦ ıS , G ◦ ıS }2,S (m) = {F, G}2 (m) .
This applies to every point m of M, so that {F, G}1 (m) = {F, G}2 (m), for all m ∈ M,
which was to be shown.
32 1 Poisson Structures: Basic Definitions
In this case, the skew-symmetric matrix (xi j )1i, jd is the Poisson matrix of π with
respect to the coordinates x1 , . . . , xd .
At every point m ∈ V , the tangent space TmV is in a canonical way isomorphic to V ,
and similarly the cotangent space Tm∗V is canonically isomorphic to V ∗ . Therefore,
one usually thinks of the Poisson bivector at m, which is a skew-symmetric bilinear
map πm : Tm∗V × Tm∗V → F, as a skew-symmetric bilinear map V ∗ × V ∗ → F, as an
element of V ∧V , or as a linear map V ∗ → V . The rank of the latter linear map is the
rank of the Poisson structure at m.
It is clear that the results in this section are also valid when the vector space V is
replaced by a non-empty open subset of V .
34 1 Poisson Structures: Basic Definitions
1.5 Exercises
for every m ∈ M1 .
5. Given a skew-symmetric biderivation {· , ·} on a commutative associative alge-
bra A , define its Jacobiator J as the skew-symmetric tri-linear map A 3 → A ,
given by
J (F, G, H) := {F, {G, H}} + {G, {H, F}} + {H, {F, G}} .
Show that J is a derivation in each of its arguments (as in (1.4); in the language of
Section 3.1, it is a triderivation) and derive from it an alternative proof of Proposi-
tion 1.8.
6. Consider on F2 the bivector field π , defined by {x, y} := x2 , and let Y denote the
vector field on F2 , defined by Y [x] := 0 and Y [y] := x. Show that Y is a Poisson
vector field, but that there exists no neighborhood of o = (0, 0) on which Y is a
Hamiltonian vector field (see item (3) of Proposition 1.31).
7. The following is an open problem: given a Poisson algebra (A , ·, {· , ·}) of fi-
nite type, do there exist generators x1 , . . . , xd of A and representatives xi j = −x ji ∈
F[x1 , . . . , xd ] of xi , x j such that
d
∂ xi j ∂ x jk ∂ xki
∑ xk ∂ x + xi ∂ x + x j ∂ x = 0 ,
=1
1.6 Notes 35
for every 1 i, j, k d; in other words, such that the matrix X := (xi j )1i, jd is a
Poisson matrix?
1.6 Notes
In the mathematics and physics literature, Poisson structures are most often consid-
ered in the case of smooth manifolds. In this context, Poisson structures are usually
defined as sections of the exterior square of the tangent bundle of the manifold (see
Vaisman [194]), as Lie algebra structures on the algebra of functions on the man-
ifold (see Cannas da Silva–Weinstein [34] or Dufour–Zung [63]) or as sheafs of
Poisson algebras on the manifold (see Przybylski [172]). Abstract Poisson algebras
are considered in Bhaskara–Viswanath [23]; see also Huebschmann [96]. For more
information on the general theory of real manifolds, we refer to Warner [198] or to
Spivak [187]; for complex manifolds, see Wells [203]. For a gentle introduction to
algebraic geometry, in particular the link between commutative associative algebras
and affine varieties, see Perrin [165] or Shafarevich [182].
The main result in this chapter is Weinstein’s splitting theorem (Theorem 1.25),
from which we derived the symplectic foliation. This general theorem admits further
generalizations: for example, an equivariant version is given in Dufour–Zung [63],
while a generalization to arbitrary Lie algebroids is given in Fernandes [75]. See
Laurent–Miranda–Vanhaecke [120] for the Carathéodory–Jacobi–Lie theorem for
Poisson manifolds, which was stated without proof in Theorem 1.28 above; for reg-
ular Poisson manifolds, this theorem goes back to Lie and Engel [128].
Chapter 2
Poisson Structures: Basic Constructions
In this chapter we give a few basic, general constructions which allow one to build
new Poisson structures from given ones. These constructions are fundamental and
will be used throughout the book. More advanced constructions, which all fall under
the general concept of reduction, will be discussed in detail in Chapter 5, while
several constructions which are specific to a particular class of examples will be
given in Part II.
As in the previous chapter, all Poisson algebras will be defined over an arbitrary
field F of characteristic zero, our Poisson varieties will be affine varieties over F
and our Poisson manifolds will be either real smooth manifolds (F = R) or complex
manifolds (F = C). We will adhere to our habit of reformulating all our algebraic
constructions in geometrical terms, or, conversely, present the geometrical construc-
tion in general algebraic terms. We stress that although the geometrical and algebraic
constructions are based on the same idea, their concrete implementation is usually
quite different.
Section 2.1 deals with the tensor product of Poisson algebras, which geomet-
rically corresponds to the construction of a Poisson structure on the product of
two Poisson manifolds. We investigate in Section 2.2 the notion of a Poisson ideal,
whose geometrical counterpart is that of a Poisson submanifold. A Poisson struc-
ture cannot be restricted to any submanifold; a necessary and sufficient conditions
for this is that all Hamiltonian vector fields be tangent to the submanifold. In Sec-
tion 2.3, we show on the one hand how real and complex Poisson structures are
related, and on the other hand how complex algebraic and holomorphic Poisson
structures are related. In Section 2.4, we assemble a few constructions which are of
a different nature, namely changing the base field, localization and germification.
Unless otherwise stated, F stands throughout this chapter for an arbitrary field of
characteristic zero.
Let (A1 , ·1 ) and (A2 , ·2 ) be two commutative associative algebras over F, with unit.
Their tensor product A1 ⊗ A2 is itself a commutative associative algebra with unit,
with respect to the F-bilinear product, which is defined for F1 , G1 ∈ A1 and for
F2 , G2 ∈ A2 by:
In this formula, and in the formulas which follow, we use the obvious abbreviations
F1 G1 for F1 ·1 G1 and similarly for F2 G2 . The algebras A1 and A2 are naturally
identified with subalgebras of A1 ⊗ A2 via the inclusion maps ji : Ai → A1 ⊗ A2
(i = 1, 2), given by j1 (F) = F ⊗ 1 for every F ∈ A1 , and similarly j2 (F) = 1 ⊗ F
for every F ∈ A2 . Also, to a derivation V of A1 we can associate in a natural way a
derivation V of A1 ⊗ A2 (with values in A1 ⊗ A2 ), such that V [j1 (F)] = j1 (V [F]),
for every F ∈ A1 . It is defined by
V : A1 ⊗ A2 → A1 ⊗ A2
(2.2)
F1 ⊗ F2 → V [F1 ] ⊗ F2 .
Notice that V [j2 (F)] = 0, for every F ∈ A2 . To see that V is indeed a derivation of
A1 ⊗ A2 , use the fact that V is a derivation of A1 , combined with (2.1). Similarly, to
a derivation W of A2 , we can associate in a natural way a derivation W of A1 ⊗ A2 ,
such that W[j2 (F)] = j2 (W[F]), for every F ∈ A2 .
By extending this construction to the case of skew-symmetric biderivations, we
can construct a skew-symmetric biderivation of A1 ⊗ A2 (with values in A1 ⊗ A2 ),
starting from a skew-symmetric biderivation of A1 (or of A2 ). We use this in the
following proposition to construct a Poisson structure on A1 ⊗ A2 , starting from a
Poisson structure on A1 and a Poisson structure on A2 .
2.1 Tensor Products and Products 39
Proposition 2.1. Let (Ai , ·i , {· , ·}i ) be two Poisson algebras over F, where i = 1, 2.
The tensor product A1 ⊗ A2 admits a unique Poisson bracket {· , ·}, making the
canonical inclusions ji : Ai → A1 ⊗ A2 into Poisson morphisms with Poisson-
commuting images, {j1 (A1 ), j2 (A2 )} = 0. For Fi , Gi ∈ Ai , this Poisson bracket is
given by
For every Hamiltonian derivation XH of A1 , with H ∈ A1 , X H is a Hamiltonian
derivation of A1 ⊗ A2 , more precisely,
X H = XH⊗1 ,
π [F1 ⊗ 1, G1 ⊗ 1] = π1 [F1 , G1 ] ⊗ 1 ,
π [F1 ⊗ 1, 1 ⊗ G2 ] = 0 ,
π [1 ⊗ F2 , 1 ⊗ G2 ] = 1 ⊗ π2 [F2 , G2 ] .
By the biderivation property, these formulas can be summarized in the single for-
mula
for every Fi , Gi ∈ Ai , (i = 1, 2). It follows that π satisfies the Jacobi identity, be-
cause (2.5), combined with the fact that π is a biderivation, implies that
π [π [F1 ⊗ F2 , G1 ⊗ G2 ], H1 ⊗ H2 ]+ (F, G, H)
= π1 [π1 [F1 , G1 ], H1 ] ⊗ (F2 G2 H2 ) + (F1 G1 H1 ) ⊗ π2 [π2 [F2 , G2 ], H2 ]+ (F, G, H) .
so that X H = XH⊗1 , as was to be shown.
Given two affine varieties M1 and M2 , their product M1 × M2 is an affine variety,
with algebra of functions F (M1 × M2 ) F (M1 ) ⊗ F (M2 ). Under the latter iso-
morphism, the natural projection maps pi : M1 × M2 → Mi , with i = 1, 2, are dual
to the inclusion maps ji : F (Mi ) → F (M1 ) ⊗ F (M2 ); specifically, the following
triangle of algebra homomorphisms is commutative:
p∗i
F (Mi ) F (M1 × M2 )
ji
F (M1 ) ⊗ F (M2 )
Translated in the language of varieties, Proposition 2.1 yields the following result.
2.1 Tensor Products and Products 41
Fig. 2.1 On the product of two affine varieties M1 and M2 , functions which are constant on the
fibers of the canonical projection map p1 : M1 × M2 → M1 are of the form p∗1 F = F ◦ p1 , for some
F ∈ F (M1 ). Every vector field on M1 × M2 which annihilates all these functions is tangent to the
fibers of p1 .
Proposition 2.2. Let (Mi , πi ) be two affine Poisson varieties, where i = 1, 2. The
product variety M1 × M2 has a unique Poisson structure π = {· , ·} such that the
projection maps pi : M1 × M2 → Mi (i = 1, 2) are Poisson maps and such that
for all functions F1 ∈ F (M1 ) and F2 ∈ F (M2 ). Each Hamiltonian vector field on M1
(or M2 ) extends to a Hamiltonian vector field on M1 × M2 , tangent to the fibers of
p2 (or p1 ).
The Poisson variety (M1 × M2 , π ) is called the product Poisson variety of (M1 , π1 )
and (M2 , π2 ), while π is called its product Poisson structure.
To finish this section, we describe the Poisson matrix of the product Poisson struc-
ture on M1 × M2 , when M1 and M2 are affine Poisson varieties. We show that, in
42 2 Poisson Structures: Basic Constructions
which yields the first diagonal block of the Poisson matrix of {· , ·}. Similarly, the
fact that p2is a Poisson map yields the second diagonal block. The other entries
p∗1 xi , p∗2 y j of the Poisson matrix are zero, as follows from (2.6).
By a slight abuse of notation, one usually simply writes xi for ∗ ∗
p1 xi and y j for p2 y j .
X 0
Then the Poisson matrix (2.7) takes the simple form and the product Pois-
0Y
son structure π = {· , ·} can be explicitly written in the following form
∂ ∂ ∂ ∂
π= ∑ xi j ∧ + ∑ yi j
∂ xi ∂ x j 1i< ∂ y i
∧
∂ yj
.
1i< jd1 jd2
Fig. 2.2 A vector field on M1 can be naturally lifted to a vector field on M1 × M2 , tangent to the
fibers of p2 : M1 × M2 → M2 .
This means that for every F ∈ F (U1 ) and G ∈ F (U2 ), where U2 is an open subset
of M2 ,
V[F ◦ p1 ] = V [F] ◦ p1 and V[G ◦ p2 ] = 0 , (2.8)
since for all m = (m1 , m2 ) ∈ U1 ×U2 ,
and similarly for the second equality in (2.8). These equalities can be used to
write V in terms of local coordinates, showing in particular that V is a vector field
(i.e., it is smooth/holomorphic). Similarly, every vector field W on an open subset
on M1 ×U2 . See Fig. 2.2.
U2 of M2 leads to a vector field W
By extending this construction to the case of bivector fields, we can construct
a bivector field on open subsets of M1 × M2 , starting from a bivector field defined
on open subsets of M1 (or M2 ). We will use this in the following proposition to
construct a Poisson structure on M1 × M2 , starting from a Poisson structure on M1
and a Poisson structure on M2 .
44 2 Poisson Structures: Basic Constructions
Proposition 2.5. Let (M1 , π1 ) and (M2 , π2 ) be two Poisson manifolds. The product
manifold M1 ×M2 has a unique Poisson structure {· , ·} = π , such that the projection
maps pi : M1 × M2 → Mi (i = 1, 2) are Poisson maps and such that, for all open
subsets Ui ⊂ Mi and for every Fi ∈ F (Ui ) with i = 1, 2, the functions F1 ◦ p1 and
F2 ◦ p2 are in involution (on U1 ×U2 ),
{F1 ◦ p1 , F2 ◦ p2 } = 0 . (2.9)
It follows that the bivector field π := π1 + π2 , where π2 is constructed in the same
way as π1 , but using Tm2 ım1 instead of Tm1 ım2 , has the following properties:
π [F1 ◦ p1 , G1 ◦ p1 ] = π1 [F1 , G1 ] ◦ p1 ,
π [F1 ◦ p1 , G2 ◦ p2 ] = 0 , (2.10)
π [F2 ◦ p2 , G2 ◦ p2 ] = π2 [F2 , G2 ] ◦ p2 .
Moreover, the first and third equations in (2.10) express that p1 and p2 are Poisson
maps, while the second one says that π satisfies (2.9).
If a Poisson structure on M1 × M2 satisfies the requirements of Proposition 2.5,
then it satisfies the equations (2.10), from which one can obtain explicit formu-
las for it, in local coordinates, which provides uniqueness of such a Poisson struc-
ture. In order to show how these explicit formulas are obtained from (2.10), let
(U, x) and (V, y) be coordinate charts of M1 and M2 , where x = (x1 , . . . , xd1 ) and
y = (y1 , . . . , yd2 ). By a slight abuse of notation, we write the corresponding coordi-
nates on U ×V as x1 , . . . , xd1 , y1 , . . . , yd2 . According to (1.26) and (2.10), π is given
in terms of these coordinates by
∂ ∂ ∂ ∂
π= ∑ π [xi , x j ] ∧ + ∑ π [yi , y j ]
∂ xi ∂ x j 1i< ∂ y i
∧
∂ yj
1i< jd1 jd2
∂ ∂ ∂ ∂
= ∑ π1 [xi , x j ] ◦ p1 ∧ + ∑ π2 [yi , y j ] ◦ p2
∂ xi ∂ x j 1i< jd2
∧
∂ yi ∂ y j
,
1i< jd1
so that
π(m1 ,m2 ) (F, G) = (π1 )m1 (F ◦ ım2 , G ◦ ım2 ) + (π2 )m2 (F ◦ ım1 , G ◦ ım1 ) ,
which yields, in the bracket notation, the following formula for the product bracket
46 2 Poisson Structures: Basic Constructions
{F, G} (m1 , m2 ) = {F ◦ ım2 , G ◦ ım2 }1 (m1 ) + F ◦ ım1 , G ◦ ım1 2 (m2 ) , (2.12)
The simplest sub-objects of Poisson manifolds are Poisson submanifolds, which are
submanifolds, equipped with a Poisson structure, with respect to which the inclu-
sion map is a Poisson map. For a submanifold, several conditions will be given,
which are equivalent to the fact that it is a Poisson submanifold, including the well
known condition that all Hamiltonian vector fields (of locally defined functions) be
tangent to the submanifold. The algebraic condition, which corresponds to it, is that
the ideal of functions which vanish on the submanifold, is a Poisson ideal. In gen-
eral algebraic terms, for an ideal I of a Poisson algebra (A , ·, {· , ·}), one has the
following fact: A /I has a Poisson bracket which makes the canonical projection
into a morphism of Poisson algebras if and only if I is a Poisson ideal; if such a
Poisson bracket on A /I exists, then it is unique.
We give in Section 2.2.1 the algebraic construction, and we reformulate the result
in geometrical terms for subvarieties of affine Poisson varieties. In Section 2.2.2, we
give a purely geometrical construction for (immersed or embedded) submanifolds of
a Poisson manifold. For the latter, we also relate the Poisson submanifold condition
to the intersection of the submanifold with the symplectic leaves.
{· , ·}, i.e., such that p is a Poisson morphism. The fact that p is surjective implies
that, if such a bracket exists, then it is unique.
Suppose that such a Poisson bracket {· , ·}0 exists. The morphism property
F, G 0
= {F, G} (2.13)
for all F, G ∈ A .
We now turn to the case of affine Poisson varieties, which will bring us to the notion
of a Poisson subvariety. We first recall that if N and M are affine varieties which
belong to the same affine space Fd , then N is said to be an affine subvariety of M if
N ⊂ M. Then N is the zero locus of a prime ideal IN in F (M) and F (N) is in a
natural way isomorphic to F (M)/IN , with which it will in the sequel be identified.
We denote by κ : F (M) → F (N) the canonical projection, which we think of as a
restriction map.
48 2 Poisson Structures: Basic Constructions
Translating the equivalent algebraic conditions, given at the beginning of this sec-
tion, to the case of varieties, we have that for a derivation V of F (M), the following
conditions are equivalent:
(i) There exists a derivation V0 of F (N) such that κ (V [F]) = V0 [κ (F)], for every
F ∈ F (M), which can be expressed as the commutativity of the following
diagram:
V
F (M) F (M)
κ κ
F (N) F (N)
V0
(ii) V maps the prime ideal IN of all functions which vanish on N to itself: for
every F ∈ IN , V [F] ∈ IN .
As we will see in the next section, these two equivalent conditions on V correspond
in the context of manifolds to the fact that V , viewed as a vector field on M, is
tangent to N, and V0 is the restriction of V to N. In the case of varieties, we therefore
also think of V0 as being the restriction of V to N and, by a slight abuse of language,
we say that V is tangent to N (at points of N), if V satisfies V [IN ] ⊂ IN .
Definition 2.9. Let (M, {· , ·}) be a Poisson variety and let N ⊂ M be a subvariety.
We say that N is a Poisson subvariety of M if there exists a Poisson structure on N
which turns the inclusion map ı : N → M into a Poisson map.
so that (XF )ı(n0 ) ∈ Tn0 ı Tn0 N0 = Tn N, for all n0 ∈ N0 , which means that XF is
tangent to N. This shows that all (locally defined) Hamiltonian vector fields are
tangent to N.
(ii) =⇒ (iii). Let n ∈ N and consider the symplectic leaf Sn (M) which con-
tains n (see Section 1.3.4). Since ı is an immersion, there exists an open subset V of
N0 which contains n0 , where n = ı(n0 ), and an open subset U of M, such that ı(V ) is
a closed, embedded submanifold of U. We also consider the symplectic leaf Sn (U),
which is an open subset of Sn (M) (recall that it is, in general, strictly smaller than
Sn (M) ∩ U). We claim that Sn (U) ⊂ ı(V ). To show this, it suffices to show that
every Hamiltonian path in U which starts in ı(V ), stays in ı(V ). Consider a Hamil-
tonian path γ : I → U, where I is an open neighborhood of 0 in F (which is R or C),
with γ (0) ∈ ı(V ). It is the integral curve of a Hamiltonian vector field XF , with
F ∈ F (U ), which is tangent to ı(V ), at all points of ı(V ), where U is an open
subset of U, containing γ (I). Suppose that γ (I) is not entirely contained in ı(V ).
Consider
I := {t ∈ I | γ (t) ∈ ı(V )} ,
which is an open subset of I, since γ is an integral curve of a vector field which is
tangent to ı(V ). Let I0 denote the connected component of I which contains 0. Let t
be an arbitrary element of the (topological) boundary of I0 and let n := γ (t ). Then
2.2 Poisson Ideals and Poisson Submanifolds 51
πN : N0 → ∧2 T N0
(2.15)
n0 → (πN )n0
where we have used the definition (2.14) of πN to obtain the last line. Since n0 →
π [F, G](ı(n0 )) is clearly a smooth/holomorphic function on V ⊂ N0 , it follows that
πN is a (smooth/holomorphic) bivector field on N0 . The same computation, with
three functions F0 , G0 and H0 , defined on a neighborhood of n0 in N0 , yields
∂ ∂ d d
∂ ∂
π= ∑ αi j ∧ + ∑ ∑ αi j ∧
∂ xi ∂ x j i=1 j=s+1 ∂ xi ∂ x j
,
1i< js
This section is devoted to three constructions which are intimately related to the real
or holomorphic nature of Poisson structures.
where the latter isomorphism consists of writing a C-valued function as the sum
of its real and its imaginary parts. It is clear that every coordinate chart (U, z)
of M, with z = (z1 , . . . , zd ) leads to a coordinate chart (UR , Z) of MR , with Z =
∞
√where xk , yk ∈ C (UR ) are the real smooth functions obtained
(x1 , . . . , xd , y1 , . . . , yd ),
by writing zk = xk + −1 yk under the decomposition (2.16).
Let V be a holomorphic vector field on M. According to (B.8), in a coordinate
chart (U, z), we can write
2.3 Real and Holomorphic Poisson Structures 53
d
∂ 1 d ∂ √ ∂
V = ∑ V [zk ] ∂ zk
= ∑ V [zk ]
2 k=1 ∂ xk
− −1
∂ yk
. (2.17)
k=1
Decomposing V [zk ] into its real and imaginary parts, leads to the definition of
two derivations of C∞ (UR ), the real and imaginary parts of V , denoted by ℜ(V )
and ℑ(V ), by putting:
1 d ∂ ∂
ℜ(V ) := ∑ ℜ(V [zk ]) ∂ xk + ℑ(V [zk ]) ∂ yk
2 k=1
, (2.18)
1 d ∂ ∂
ℑ(V ) := ∑ ℑ(V [zk ]) ∂ xk − ℜ(V [zk ]) ∂ yk
2 k=1
. (2.19)
∂G ∂H ∂G ∂H
= , =− ,
∂ xk ∂ yk ∂ yk ∂ xk
1 1
ℜ(V )[F] = V [F], ℜ(V )[F] = V [F] . (2.20)
2 2
These equations characterize ℜ(V ) since (2.18) can be derived from it. Similarly,
the imaginary part ℑ(V ) is characterized by the fact that
√ √
−1 −1
ℑ(V )[F] = − V [F], ℑ(V )[F] = V [F] ,
2 2
for all F ∈ H (U). These characterizations imply that ℜ(V ) and ℑ(V ) are inde-
pendent of the choice of the coordinates and therefore, define (smooth) vector fields
on MR .
An analogous construction works in the case of bivector fields. Indeed, let P be a
bivector field of M. According to (1.23), we can write P in a coordinate chart (U, z)
as
∂ ∂
P= ∑ P[zk , z ] ∧
∂ zk ∂ z
1k<d
1 ∂ √ ∂ ∂ √ ∂
= ∑ P[zk , z ]
4 1k<d ∂ xk
− −1
∂ yk
∧
∂ x
− −1
∂ y
.
As above, decomposing P[zk , z ] into its real and imaginary parts, leads to the defi-
nition of two skew-symmetric biderivations ℜ(P) and ℑ(P) of C∞ (UR ), by putting,
54 2 Poisson Structures: Basic Constructions
1 ∂ ∂ ∂ ∂
ℜ(P) := ∑
4 1k<d
ℜ(P[zk , z ]) ∧ − ∧
∂ xk ∂ x ∂ yk ∂ y
1 ∂ ∂ ∂ ∂
+ ∑ ℑ(P[zk , z ]) ∂ xk ∧ ∂ y + ∂ yk ∧ ∂ x
4 1k<d
,
and similarly for ℑ(P). By definition, ℜ(P) and ℑ(P) are skew-symmetric bideriva-
tions of C∞ (UR ) and they uniquely extend to skew-symmetric biderivations of
C∞ (UR , C), by C-linearity. Then, ℜ(P) is, as in (2.20), characterized by the fol-
lowing properties: for all F, G ∈ H (U),
1
ℜ(P)[F, G] = P[F, G] ,
2
ℜ(P)[F, G] = 0 , (2.21)
1
ℜ(P)[F, G] = P[F, G] .
2
There is an analogous characterization for the imaginary part ℑ(P). Because of
these characterizations, we can conclude, as in the case of vector fields, that ℜ(P)
and ℑ(P) are well-defined bivector fields of MR . We show in the next proposition
that, if the bivector field P has the additional property of being a Poisson structure,
then ℜ(P) is also Poisson structure. One shows similarly that ℑ(P) is a Poisson
structure.
Proposition 2.13. If (M, π ) is a complex Poisson manifold of dimension d, then the
bivector field ℜ(π ) is a Poisson structure on MR . In local coordinates, ℜ(π ) is given
by:
1 ∂ ∂ ∂ ∂
ℜ(π ) = ∑ ℜ(πk ) ∂ xk ∧ ∂ x − ∂ yk ∧ ∂ y
4 1k<d
1 ∂ ∂ ∂ ∂
+ ∑
4 1k<d
ℑ( π k ) ∧ + ∧
∂ xk ∂ y ∂ yk ∂ x
√
where zk = xk + −1yk and πk := π (zk , z ), for 1 k, d.
1
ℜ(π )[ℜ(π )[F, G], H] = {{F, G} , H} ,
4
ℜ(π )[ℜ(π )[F, G], H] = 0 ,
ℜ(π )[ℜ(π )[F, G], H] = 0 ,
1
ℜ(π )[ℜ(π )[F, G], H] = {{F, G} , H} ,
4
2.3 Real and Holomorphic Poisson Structures 55
where we have written {· , ·} for π . Since π satisfies the Jacobi identity, these equa-
tions imply that ℜ(π ) satisfies the Jacobi identity, when applied to arbitrary triples
of holomorphic or anti-holomorphic functions on U. In particular, for every coordi-
nate chart (UR , Z), the Jacobi identity will be satisfied by ℜ(π ) for the coordinate
functions xk and y , because they can be written as C-linear combinations of the
holomorphic and anti-holomorphic functions zk and zk . Then, according to Proposi-
tion 1.16, ℜ(π ) is a bivector field of MR , satisfying the Jacobi identity, i.e., it is a
Poisson structure on MR .
equals d − s is a (Zariski) open subset of N, called the smooth part of N, and de-
noted N sm . The points of N sm are called smooth points of N. By the implicit function
theorem, N admits in the neighborhood of every smooth point the structure of a com-
plex manifold, which is the unique structure for which the functions x1 |N , . . . , xd |N
are holomorphic. In particular, N sm is a complex manifold. For every point n in N sm
we can choose s functions among x1 |N , . . . , xd |N , which constitute local coordinates
of N in a neighborhood of n in N sm and the other functions xi |N are then expressible
as holomorphic functions in terms of these local coordinates.
Suppose now that N comes equipped with a Poisson structure, i.e., (N, {· , ·}) is
a (complex) affine Poisson variety, N ⊂ Cd . Since {· , ·} associates to two regular
functions1 on N a regular function on N, we may choose, forevery i, j, with 1
i, j d, a representative xi j of the Poisson bracket xi |N , x j |N , i.e., an element of
C[x1 , . . . , xd ] such that xi j |N = xi |N , x j |N . Since all xi j are polynomials, the skew-
symmetric biderivation
∂ ∂
∑ xi j ∂ xi ∧ ∂ x j
1i< jd
on Cd .
56 2 Poisson Structures: Basic Constructions
∂ F̃ ∂ G̃
{F, G}H (n ) := ∑ xi j (n )
∂ xi
(n )
∂xj
(n ).
1i< jd
It defines the unique bivector field on V such that xi |N , x j |N = xi |N , x j |N on V ,
H
for all 1 i, j d, hence it is a well-defined holomorphic
bivector field on N sm .
Since {· , ·} satisfies the Jacobi identity for all triples xi |N , x j |N , xk|N , with 1 i <
j < k d, the same holds true for {· , ·}H , which means that the Jacobi identity
holds for {· , ·}H , since in a neighborhood of every point, a subset of x1 , . . . , xd
yields a system of local coordinates. This shows that {· , ·} leads to a holomorphic
Poisson structure {· , ·}H on N sm . In particular, if N is a smooth Poisson variety, so
that N sm = N, the above construction yields a holomorphic Poisson structure on N,
starting from the Poisson bracket on the algebra of regular functions on N.
We finish this chapter with a few isolated constructions, which are of a different
nature.
a(b ⊗ F) := (ab) ⊗ F ,
(2.22)
(a ⊗ F)(b ⊗ G) := (ab) ⊗ (FG) .
Combining (2.22) and (2.23) shows that every F-Poisson algebra (A , ·, {· , ·}), leads
in a natural way to an F̂-Poisson algebra (F̂ ⊗F A , ·, {· , ·}). For example, it allows
2.4 Other Constructions 57
2.4.2 Localization
(F, H) ∼ (F , H ) ⇔ FH = F H . (2.24)
showing that VS is indeed determined by its values on A . In order to show that every
derivation V of A extends to a derivation of A /S one takes (2.25) as a definition
of VS . One needs to check that (2.25) is well-defined: to do this, suppose that F1 H2 =
F2 H1 , with Fi ∈ A and Hi ∈ S. First, notice that the equality F1 H2 = F2 H1 implies
that V [F1 H2 − F2 H1 ] = 0, which we write as
This shows that VS is well-defined. It is easy to verify that (2.25) and the fact that V
is a derivation, imply that, for all F, F ∈ A and H, H ∈ S, one has that
F F F F F F
VS = VS + VS , (2.27)
HH H H H H
{{H, K} , F} + (H, K, F)
= ,
(HK)2
In the case of an affine variety M there are two important cases of localization. The
first one consists in taking for S the set Sm of all functions F ∈ F (M) which do not
vanish at a given point m. Then the localization F (M)/Sm is the algebra of rational
functions F/G, with F, G ∈ F (M) and G(m) = 0. By the above, every Poisson
structure π on M will induce a Poisson structure πSm on this algebra of functions. Of
course, the algebra of all rational functions on M also inherits a Poisson structure
from M, leading to a Poisson algebra which contains all (F (M)/Sm , πSm ) as Poisson
subalgebras.
The second special case of localization consists in taking a non-constant func-
tion G ∈ A and considering for S the set SG := {Gn | n ∈ N}. Then F (M)/SG is
the algebra of rational functions F/Gn , with F ∈ F (M) and n ∈ N. Picking an in-
determinate ξ we have an algebra isomorphism
F (M)[ξ ]
ρ: → F (M)/SG
Gξ − 1
n n
∑ Fi ξ i → ∑ Fi /Gi .
i=0 i=0
2.4.3 Germification
where F and G are arbitrary representatives of Fm and Gm ; also x1 , . . . , xd are local
coordinates, in a neighborhood of m and xi j := xi , x j . Evaluating {Fm , Gm }m at m
yields a skew-symmetric bilinear map Tm∗ M × Tm∗ M → F, which is precisely the
pointwise biderivation at m, associated to the Poisson structure, which we denoted
earlier by πm . Although we have used the same notation πm for the germ of the
Poisson structure at m and for the Poisson bivector at m, one should not forget that
the germ at m contains more information than the Poisson bivector at m.
2.5 Exercises
1. Determine the algebra of Casimirs of the Poisson tensor product of two Poisson
algebras.
2. Let A and B be two Poisson algebras and let ρ : A → B ⊗ A be a Poisson
morphism, where the right-hand side is equipped with the product bracket. Prove
that
A B := {F ∈ A | ρ (F) = 1 ⊗ F}
is a Poisson subalgebra of A .
3. Prove Proposition 2.14 in the case in which A may have zero divisors (in
this case, the definition (2.24) of the equivalence relation ∼ is replaced by (a, s) ∼
(a , s ) ⇔ ∃u ∈ S such that u(as − a s) = 0).
4. Reprove Proposition 2.14 by first showing that a biderivation of A /S vanishes,
as soon as it vanishes on A .
5. Let (M1 , π1 ) and (M2 , π2 ) be two Poisson manifolds. Let Ψ : M1 → M2 be a
Poisson map. Show that if Ψ (M1 ) is a submanifold of M2 , then it is a Poisson sub-
manifold.
6. Let (M, π ) be a complex Poisson manifold. Show that every linear combination
of the Poisson structures ℜ(π ) and ℑ(π ), defined in Section 2.3.1, is a Poisson
structure on MR .
7. Let π be a Poisson structure on an open subset U of Rd , where d 2, and let F
be a smooth function on U. Show that if π has rank at most 2 at every point of U,
then F π is a Poisson structure on U.
2.6 Notes 61
8. We assume in this exercise that the reader is familiar with the notion of sym-
plectic manifold (see Section 6.3). Let (M, ω ) be a symplectic manifold, and denote
by π the canonical Poisson structure, associated to ω . Prove that the open subsets
of M are the only Poisson submanifolds of (M, π ).
9. The purpose of this exercise is to establish the existence of non-trivial Poisson
structures on every (smooth real) manifold of dimension at least 2. Throughout the
exercise we consider Rd with its algebra of smooth functions.
a. Show that, if d 2, then there exists a Poisson structure on Rd which has
constant rank 2 in a neighborhood of 0 and which has rank zero outside some
compact subset of Rd ;
b. Show that, if d 4, then there exists a Poisson structure on Rd which has
constant rank 4 in a neighborhood of 0 and which has rank at most 2 outside
some compact subset of Rd . Using Exercise 6, conclude that there exists such
a Poisson structure, which vanishes outside some compact subset of Rd ;
c. Generalizing parts a. and b., show that for every s ∈ N, such that 2s d, there
exists a Poisson structure on Rd which has constant rank 2s in a neighborhood
of 0 and which vanishes outside some compact subset of Rd ;
d. Using a coordinate chart, conclude that every manifold of dimension d 2,
admits a non-trivial Poisson structure; it can be chosen such that its rank, at a
given point, equals a given even integer 2s d;
e. The following is an open problem: show that for every Poisson structure π on
Rd there exists a Poisson structure on Rd which coincides with π in a neigh-
borhood of the origin of 0 and which vanishes outside some compact subset
of Rd .
2.6 Notes
The constructions which were given in this chapter are so basic that it is difficult
to give the original references. Most of the geometrical constructions are at least
implicit in Lichnerowicz’s seminal paper [126], in which Poisson manifolds (of
constant rank) were introduced. The algebraic formulation of these constructions
follows easily. For the link between real and holomorphic Poisson structures, ex-
plained in Section 2.3.1, see Laurent–Stiénon–Xu [121].
Chapter 3
Multi-Derivations and Kähler Forms
The Jacobi identity, which is a key element in the definition of a Poisson bracket on
an algebra A (or a Poisson structure on a manifold M), can be naturally formulated
in terms of the Schouten bracket on the space X• (A ) of all skew-symmetric multi-
derivations of A (or on the space X• (M) of all multivector fields on M). In a sense,
the Schouten bracket is dual to the differential on the space of Kähler forms Ω • (A )
(or on Ω • (M), the space of differential forms on M). See Table 3.1 for an overview
of the algebraic structures which will be explained in detail in this chapter, with
special emphasis on their connections with Poisson structures.
Table 3.1 A summary of the operations which will be considered in this chapter. In the table,
P ∈ X p , where p > 0.
X• (A ), X• (M) Ω • (A ), Ω • (M)
Algebraic multi-derivation Kähler form
Geometric multivector field differential form
Wedge product ∧ : X• × X• → X• ∧ : Ω• ×Ω• → Ω•
Differential – d : Ω • → Ω •+1
Schouten bracket [· , ·]S : X• × X• → X•−1 –
Lie derivative LP : X• → X•+p−1 LP : Ω • → Ω •−p+1
Internal product – ıP : Ω • → Ω •−p
Since we are using mostly the algebraic properties of these operations, our con-
structions will be mainly algebraic, but they will always be reformulated in geo-
metrical terms. Multi-derivations are introduced in Section 3.1 and Kähler forms in
Section 3.2. The Schouten bracket and the closely related notion of (generalized)
Lie derivative will be introduced in Section 3.3.
Unless otherwise stated, F stands throughout this chapter for an arbitrary field of
characteristic zero.
3.1.1 Multi-Derivations
(FP)[F1 , . . . , Fp ] := F P[F1 , . . . , Fp ] ,
There are two natural composition laws on the graded A -module X• (A ) of skew-
symmetric multi-derivations of a commutative associative F-algebra A . The first
one, introduced here, is the wedge product. The second one is a graded Lie bracket,
the Schouten bracket, which will be introduced in Section 3.3.
We first recall the notion of a shuffle. For p, q ∈ N, a (p, q)-shuffle is a permuta-
tion σ of the set {1, . . . , p + q}, such that σ (1) < · · · < σ (p) and σ (p + 1) < · · · <
σ (p + q). The set of all (p, q)-shuffles is denoted by S p,q . For a shuffle σ ∈ S p,q ,
we denote the signature of σ (as a permutation) by sgn(σ ). It is also convenient to
define S p,−1 := 0/ and S−1,q := 0, / for p, q ∈ N.
For P ∈ X (A ) and Q ∈ X (A ), their wedge product P ∧ Q ∈ X p+q (A ) is the
p q
(P ∧ Q)[F1 , . . . , Fp+q ] :=
(3.3)
∑ sgn(σ ) P[Fσ (1) , . . . , Fσ (p) ] Q[Fσ (p+1) , . . . , Fσ (p+q) ] ,
σ ∈S p,q
P ∧ Q = (−1) pq Q ∧ P . (3.4)
ıF : X• (A ) → X•−1 (A ) ,
ıF (P ∧ Q) = ıF P ∧ Q + (−1) p P ∧ ıF Q . (3.6)
This follows immediately from (3.3), upon using that for every shuffle σ ∈ S p,q , one
has that either σ (1) = 1 or σ (p + 1) = 1. Also, each derivation V ∈ X1 (A ) defines
a graded F-linear map of degree 0,
LV : X• (A ) → X• (A ) ,
which is called the Lie derivative with respect to V . For P ∈ X p (A ), its Lie deriva-
tive LV P ∈ X p (A ) is defined as follows:
p
LV P[F1 , . . . , Fp ] := V [P[F1 , . . . , Fp ]] − ∑ P[F1 , . . . , V [Fi ], . . . , Fp ] , (3.7)
i=1
LV (P ∧ Q) = LV P ∧ Q + P ∧ LV Q , (3.8)
We have already pointed out in Section B.2 that, in the context of manifolds, the
analog of a derivation, respectively skew-symmetric biderivation, is a vector field,
respectively a bivector field. Similarly, multivector fields are the geometrical analogs
of skew-symmetric multi-derivations. Since the basic definitions and properties of
multivector fields are a direct generalization of what we have seen for bivector fields
in Section B.2, the explanations and justifications which are given in this section are
kept to a minimum.
Let M be a manifold (a real manifold when F = R, a complex manifold when
F = C) and let p ∈ N∗ and m ∈ M. Generalizing Definition 1.13, we call a skew-
symmetric p-linear map
Fm (M) p
Ψm : →F
∼
a skew-symmetric pointwise p-derivation at m, if for all functions F, G, H, . . . , K,
defined in a neighborhood of m in M,
Ψm (Fm Gm , Hm , . . . , Km )
= F(m) Ψm (Gm , Hm , . . . , Km ) + G(m) Ψm (Fm , Hm , . . . , Km ) .
where 1 i1 < i2 < · · · < i p dim M; in this formula, the wedge product is the
wedge product of (linear) maps, as in the previous section (see for example Eq. 3.3).
A map, which assigns to every point m ∈ M an element Pm of ∧ p Tm M is called a
p-vector field on M if for every open subset U ⊂ M and for all functions F1 , . . . , Fp ∈
F (U), one has that P[F1 , . . . , Fp ] ∈ F (U), where P[F1 , . . . , Fp ] is the function on U,
defined for all m ∈ U by P[F1 , . . . , Fp ](m) := Pm ((F1 )m , . . . , (Fp )m ), which we also
write as Pm [F1 , . . . , Fp ]. Thus, on every coordinate chart (U, x), the p-vector field P
becomes a skew-symmetric p-derivation of the algebra of functions F (U) and P
admits the local coordinate expression
∂ ∂
P= ∑ P[xi1 , . . . , xi p ]
∂ x i1
∧···∧
∂ xi p
, (3.9)
1i1 <···<i p d
to define a multivector field for all local functions, defined on open subsets which
cover the manifold M.
We denote the F (M)-module of p-vector fields on M by X p (M) and we de-
fine the graded F (M)-module X• (M) := ⊕ p∈N X p (M), which is the geometrical
analog of X• (A ). The operations which we have introduced in Section 3.1.2 for
skew-symmetric p-derivations, namely the wedge product ∧, the contraction ıF by
a function F and the Lie derivative LV , are introduced in the same way for p-vector
fields, upon replacing the functions on which they act by local functions. Clearly,
all properties of these operations can be repeated word-for-word, as they are purely
algebraic. However, in a geometrical approach, the formula (3.7) for the Lie deriva-
tive of a p-vector field should be viewed as a property of the Lie derivative, not as its
definition. To give the geometrical definition of the Lie derivative, let us first recall
that if V and W are vector fields on a manifold M, then their commutator [V , W ] is
the vector field on M, whose value at m ∈ M is given by
d
[V , W ]m := (LV W )m := TΦ−t (m) Φt WΦ−t (m) , (3.10)
dt |t=0
where Φt is the local flow of V . Notice that the tangent map TΦ−t (m) Φt of the map
Φt at Φ−t (m) sends tangent vectors at Φ−t (m) (here, the vector WΦ−t (m) ) to tangent
vectors at m, so taking the derivative in (3.10) amounts to differentiating a curve
in Tm M. More generally, for P a p-vector field on a manifold M and V a vector
field on M, the Lie derivative of P with respect to V is defined as the p-vector field,
whose value at m ∈ M is given by
d
(LV P)m := ∧ p TΦ−t (m) Φt PΦ−t (m) . (3.11)
dt |t=0
In this section, we introduce the objects which are, in a sense, dual to skew-
symmetric multi-derivations of a (commutative associative) algebra A : the Kähler
forms of A . In the case of a manifold M, the dual objects to multivector fields are
differential forms. The modules of Kähler forms and of differential forms will be
described in this section, together with their algebraic structure.
3.2 Kähler Forms and Differential Forms 69
d
A Ω 1 (A )
V Vˆ
A
In order to prove the universal property, notice that the commutativity of the diagram
amounts to saying that Vˆ (dF) = V [F], for all F ∈ A . It is then clear that Vˆ should
be defined as the (unique) A -linear map, such that
X1 (A ) HomA (Ω 1 (A ), A ) , (3.13)
The elements of Ω p (A ), with p > 0, are called Kähler p-forms, or simply Kähler
forms. As a vector space, respectively as an A -module, Ω p (A ) is generated by
elements of the form GdF1 ∧· · ·∧dFp , respectively of the form dF1 ∧· · ·∧dFp , where
G, F1 , . . . , Fp ∈ A . Being, by definition, an exterior algebra, it is an associative,
graded commutative A -algebra with product ∧. The F-linear map d : A → Ω 1 (A )
extends by functoriality of ∧ to an F-linear map ∧• d : ∧• A → Ω • (A ). Explicitly,
it is given by
∧• d(F1 ∧ · · · ∧ Fp ) := dF1 ∧ · · · ∧ dFp , (3.14)
where F1 , . . . , Fp ∈ A . It is clear that ∧• d is a homomorphism between the algebras
(∧• A , ∧) and (Ω • (A ), ∧). For every skew-symmetric p-derivation P of A , there
exists a unique A -linear map P̂ : Ω p (A ) → A , such that the following triangle is
commutative:
∧pd
∧pA ∧ p Ω 1 (A ) = Ω p (A )
P P̂
A
X p (A ) HomA (Ω p (A ), A ) . (3.16)
for all G, F1 , . . . , Fq ∈ A . With this notation, the map P → P̂ yields a natural iso-
morphism
3.2 Kähler Forms and Differential Forms 71
X• (A ) → HomA (Ω p (A ), A ) .
p∈N
Ψ : X p (A ) ⊗A Ω p (A ) → A
(3.18)
P⊗ω → ω , P := P̂(ω )
Ψ1 : X p (A ) → HomA (Ω p (A ), A ) ,
Ψ2 : Ω p (A ) → HomA (X p (A ), A ) .
We now consider the algebra and coalgebra structures on Ω • (A ), which are closely
related to the corresponding structures on ∧• A , where A is as before a commu-
tative associative F-algebra. We assume that the reader is familiar with the facts,
recalled in Appendix A (in particular, in Sections A.3 and A.4), that (∧• A , ∧) is an
associative, graded commutative F-algebra, and that (∧• A , Δ ) is a coassociative F-
coalgebra. Since Ω • (A ) is an exterior algebra (over A ), we have, as we mentioned
in the previous section, a wedge product on Ω • (A ), also denoted by ∧, which
makes (Ω • (A ), ∧) into an associative, graded commutative A -algebra. Also, it
leads to an A -linear coassociative coproduct
Δ : Ω • (A ) → Ω • (A ) ⊗A Ω • (A ) ,
which is defined in the same way as in the case of ∧• A (see (A.17)). Namely, Δ p,q
is given, for p, q ∈ N, and F1 , . . . , Fp+q ∈ A by
(1Ω • (A ) ⊗ Δ ) ◦ Δ = (Δ ⊗ 1Ω • (A ) ) ◦ Δ . (3.20)
ν : A ⊗A Ω • (A ) → Ω • (A )
(3.22)
F ⊗ω → Fω .
We used ν in (3.21) only for the case of ω ∈ Ω 0 (A ) = A , but we will need its
general form later on; in particular we will need the following properties of ν :
P̂ ◦ ν = ν ◦ (1A ⊗ P̂) ,
(3.23)
Δ ◦ ν = (ν ⊗ 1Ω • (A ) ) ◦ (1A ⊗ Δ ) ,
where P ∈ X• (A ). The proof of these properties follows at once from the definition
(3.22) of ν .
d : Ω • (A ) → Ω •+1 (A ) ,
by putting
d(G dF1 ∧ · · · ∧ dFp ) := dG ∧ dF1 ∧ · · · ∧ dFp , (3.24)
for all G, F1 , . . . , Fp ∈ A , where p ∈ N. It is called the (algebraic) de Rham differ-
ential. It is a graded derivation, of degree 1, of (Ω • (A ), ∧), as
d d d
0 Ω 0 (A ) Ω 1 (A ) Ω 2 (A ) ···
is a complex, called the (algebraic) de Rham complex. A Kähler p-form ω for which
dω = 0 is called closed, while it is called exact if it is contained in the image of d. By
the above, exact Kähler forms are closed, which leads, for p ∈ N, to the cohomology
3.2 Kähler Forms and Differential Forms 73
space2
p Ker(d : Ω p (A ) → Ω p+1 (A ))
HdR (A ) := ,
Im(d : Ω p−1 (A ) → Ω p (A ))
where it is understood that HdR0 (A ) := Ker(d : A → Ω 1 (A )) = F. The F-vector
p
space HdR (A ) is called the p-th de Rham cohomology space and the graded vector
space p
•
HdR (A ) := HdR (A )
p∈N
• (A ) × H • (A ) → H • (A )
∧ : HdR dR dR
([ω ], [η ]) → [ω ∧ η ]
where [ω ] denotes the de Rham cohomology class of a closed Kähler form ω . It fol-
• (A ), ∧) is, just like (Ω • (A ), ∧), an associative, graded commutative
lows that (HdR
F-algebra.
To close this section, we wish to point out that Ω • leads to a covariant functor
from the category of commutative associative F-algebras to the category of graded
F-vector spaces (associative, graded commutative F-algebras, if you wish). Namely,
let φ : A → B be an algebra homomorphism, where A and B are commutative
associative algebras. For every p ∈ N, the induced linear map Ω p (φ ) : Ω p (A ) →
Ω p (B) is defined as follows:
Ω p (φ ) : Ω p (A ) → Ω p (B)
G dF1 ∧ · · · ∧ dFp → φ (G) dφ (F1 ) ∧ · · · ∧ dφ (Fp ) .
From this definition, it is easily verified that Ω • is indeed a covariant functor. Since
Ω • (φ ) and d commute, Ω • (φ ) induces a graded linear map of degree 0:
• • •
HdR (φ ) : HdR (A ) → HdR (B) .
• is also a (covariant) functor.
It follows that HdR
2 Notice that it is only a vector space and not an A -module, as the differential d is only F-linear.
74 3 Multi-Derivations and Kähler Forms
There are several equivalent ways of defining the notion of a differential form on a
manifold. The most economical definition of a differential p-form on a real manifold
is that it is an F (M)-linear map from X p (M) to F (M). In other words, it is a skew-
symmetric, F (M)-p-linear map from X1 (M) to F (M). This definition does not
work in the case of a complex manifold since, as we have already said, the space of
holomorphic functions on M may consist of constant functions only. To remedy this,
one considers local functions and local vector fields. Namely, given a manifold M, a
map which assigns to every m ∈ M an element ωm ∈ ∧ p Tm∗ M, is called a differential
p-form if for every open subset U ⊂ M and for all vector fields V1 , . . . , V p ∈ X1 (U),
one has that ω (V1 , . . . , V p ) ∈ F (U), where ω (V1 , . . . , V p ) is the function on U,
defined by
ω (V1 , . . . , V p )(m) := ωm ((V1 )m , . . . , (V p )m ) ,
for all m ∈ U. Thus, on every coordinate chart (U, x), the differential p-form ω
becomes a skew-symmetric F (U)-p-linear map from F (U) to itself. It follows
that ω admits the local coordinate expression
∂ ∂
ω= ∑ ω ∂ xi , . . . , ∂ xi dxi1 ∧ · · · ∧ dxi p , (3.26)
1i <···<i p d
1 1 p
where d denotes the dimension of M. Just like a p-vector field, a differential p-form
is a tensorial object: for m ∈ M, the value of ω (V1 , . . . , V p )(m) depends only on the
values at m of the vector fields V1 , . . . , V p , which is a useful fact for constructing
differential p-forms.
The F (M)-module of differential p-forms on M is denoted by Ω p (M) and we
define the graded F (M)-module
Ω • (M) := Ω p (M) ,
p∈N
for ω ∈ Ω p (M) and η ∈ Ω • (M), just like in (3.25). In a coordinate chart (U, x), it
leads to the classical formula for the differential of a differential p-form, namely,
3.3 The Schouten Bracket 75
and d is called the de Rham differential. It has formally the same properties as the
algebraic de Rham differential, in particular it leads to a complex, known as the de
Rham complex, whose cohomology
Ker(d : Ω p (M) → Ω p+1 (M))
• p
HdR (M) := HdR (M) := ,
p∈N p∈N
Im(d : Ω p−1 (M) → Ω p (M))
is called the de Rham cohomology of M. In the manifold case, the properties of the
de Rham differential are usually derived from the following explicit formula: dω is
given, for ω ∈ Ω p (M), by
p
dω (V0 , V1 , . . . , V p ) = ∑ (−1)i Vi ω V0 , V1 , . . . , Vi , . . . , V p (3.28)
i=0
+ ∑ j , . . . , V p ,
(−1)i+ j ω [Vi , V j ] , V0 , . . . , Vi , . . . , V
0i< jp
In this section we introduce a few extra natural operations on the graded algebras
of skew-symmetric multi-derivations and of Kähler forms on a commutative asso-
ciative algebra A . The first one is an action of X• (A ) on Ω • (A ), which gen-
eralizes the natural pairing between multi-derivations and Kähler forms, which we
encountered in Remark 3.3. The second operation is the Schouten bracket which ex-
tends the classical commutator of derivations to the case of skew-symmetric multi-
derivations. It will lead at the end of the section to the so-called Gerstenhaber al-
gebra structure on X• (A ) and to two generalizations of the Lie derivative, acting
in one case on the algebra of multi-derivations, in the other case on the algebra of
Kähler forms, where the derivative is in both cases taken with respect to a multi-
derivation. In the geometrical setting, this yields for every manifold M on the one
hand an action of the graded algebra of multivector fields X• (M) on the graded al-
gebra of differential forms Ω • (M) and on the other hand a graded Lie bracket, the
Schouten bracket, on X• (M).
76 3 Multi-Derivations and Kähler Forms
ıP : Ω • (A ) → Ω •−p (A ) ,
ıP (dF1 ∧ · · · ∧ dFk ) := ∑ sgn(σ )P[Fσ (1) , . . . , Fσ (p) ] dFσ (p+1) ∧ · · · ∧ dFσ (k)
σ ∈S p,k−p
(3.29)
when k p, and otherwise ıP ω := 0; for F ∈ X0 (A ) = A , (3.29) has to be under-
stood as ıF ω := F ω . The reader may wish to write the definition of ıP differently,
using
P[Fσ (1) , . . . , Fσ (p) ] = P̂ dFσ (1) ∧ · · · ∧ dFσ (p) .
With this notation, it is obvious that the restriction of ıP to Ω p (A ) is just P̂, which
was defined in (3.15); the restriction to the spaces Ω q (A ), with q = p, is however
different, since P̂ acts trivially on them (see (3.17)). It also follows that ıP can be
written, for every P ∈ X• (A ), in terms of the coproduct Δ , defined in (3.19), as
ıP = ν ◦ (P̂ ⊗ 1Ω • (A ) ) ◦ Δ . (3.30)
Q̂ ◦ ıP = Q̂ ◦ ν ◦ (P̂ ⊗ 1Ω • (A ) ) ◦ Δ
= ν ◦ (1A ⊗ Q̂) ◦ (P̂ ⊗ 1Ω • (A ) ) ◦ Δ
= ν ◦ (P̂ ⊗ Q̂) ◦ Δ
= P∧Q ,
which yields (3.31), when applied to ω ∈ Ω p+q (A ). Equation (3.31) explains the
notation ıP : as in the case of ıF (F ∈ A ), defined in (3.5), ıP provides ω with the
p-derivation P as its first p arguments. Yet, when considering ıF G for F, G ∈ A
one should decide from the context whether G is interpreted as a 0-derivation or as
a 0-form, since in the first case ıF G = 0 (as defined in (3.5)), while in the second
case ıF G = FG (as defined in (3.29)).
In the following proposition, we establish the main properties of ı.
3.3 The Schouten Bracket 77
Proof. We give three different proofs of (1). The first one is a quick proof, but it
is not valid for general algebras, because we assume that HomA (X p (A ), A )
Ω p (A ), for all p ∈ N (see Remark 3.3 at the end of Section 3.2.2). Applying (3.31)
three times, we find for every ω ∈ Ω k (A ) and for homogeneous elements P, Q and
R of X• (A ), whose degrees sum up to k,
ıQ∧P ω , R = ω , (Q ∧ P) ∧ R = ω , Q ∧ (P ∧ R)
(3.33)
= ıQ ω , P ∧ R = ıP (ıQ ω ), R .
Under the above assumption, the fact that (3.33) is valid for all R, implies that
ıQ∧P ω = ıP (ıQ ω ), for all ω ∈ Ω • (A ), which leads to (1). Our second proof is
based on the fact that Δ is coassociative. To simplify the notation, we will write 1
for 1Ω • (A ) ; using (3.30), the properties (3.23) of ν , the coassociativity (3.20) of Δ
and (3.21) (in that order), we find
ıP ◦ ıQ = ν ◦ (P̂ ⊗ 1) ◦ Δ ◦ ν ◦ (Q̂ ⊗ 1) ◦ Δ
= ν ◦ (P̂ ⊗ 1) ◦ (ν ⊗ 1) ◦ (1A ⊗ Δ ) ◦ (Q̂ ⊗ 1) ◦ Δ
= ν ◦ (ν ⊗ 1) ◦ (1A ⊗ P̂ ⊗ 1) ◦ (Q̂ ⊗ 1 ⊗ 1) ◦ (1 ⊗ Δ ) ◦ Δ
= ν ◦ (ν ⊗ 1) ◦ (Q̂ ⊗ P̂ ⊗ 1) ◦ (Δ ⊗ 1) ◦ Δ
= ν ◦ ((ν ◦ (Q̂ ⊗ P̂) ◦ Δ ) ⊗ 1) ◦ Δ
= ν◦ Q ∧P⊗1 ◦Δ
= ıQ∧P .
This terminates the second proof. For the third proof, we show by direct computation
that
(ıP ◦ ıQ ) (dF1 ∧ · · · ∧ dFk ) = ıQ∧P (dF1 ∧ · · · ∧ dFk ) (3.34)
for all F1 , . . . , Fk . Since ıP is A -linear for all P ∈ X p (A ), this will provide a third
proof of (1). Since ıP ω = 0 for ω ∈ Ω q (A ) when q < p, where P ∈ X p (A ), it
suffices to prove (3.34) when k p + q. On the one hand,
ıP ◦ ıQ (dF1 ∧ · · · ∧ dFk )
⎛ ⎞
where Sq,p,k−p−q is the set of all permutations σ of the set {1, . . . , k} which are
increasing on the intervals 1, . . . , q and q + 1, . . . , q + p and q + p + 1, . . . , k. On the
other hand, (3.29) and (3.3) lead to
which yields the third proof. The point of this computation is that an element of
Sq,p,k−p−q can be uniquely obtained from a shuffle σ ∈ Sq,k−q , followed by a shuffle
of the set σ (q + 1), . . . , σ (k), or from a shuffle ρ ∈ Sq+p,k−p−q , followed by a shuffle
of the set ρ (1), . . . , ρ (q + p).
The proof of (2) is immediate from (1), since the graded commutator is given by
[ıP , ıQ ] = ıP ◦ıQ −(−1) pq ıQ ◦ıP = ıQ∧P −(−1) pq ıP∧Q , and since P∧Q = (−1) pq Q∧ P
(graded commutativity, see (3.4)).
We now prove (3). Since, for every Q ∈ X• (A ), the map ıQ is A -linear and
the wedge product is also A -bilinear, we can suppose that ω is of the form ω =
dF1 ∧ · · · ∧ dFk , with k ∈ N∗ , k p − 1 and F1 , . . . , Fk ∈ A . In this case,
since every (p, k − p + 1)-shuffle of {0, . . . , k} has 0 as the first element in its first
part or in its second part. The first sum is exactly ı(ıF P) (dF1 ∧ · · · ∧ dFk ) = ı(ıF P) ω ,
while the second sum is given by (−1) p dF ∧ ıP (dF1 ∧ · · · ∧ dFk ) = (−1) p dF ∧ ıP ω .
This yields a proof of Eq. (3.32).
P̃ : ∧• A → ∧•−p+1 (A )
∧• d ∧• d
d◦ıP
Ω • (A ) Ω •−p+1 (A )
ıP d
•−p
Ω (A )
[· , ·]S : X p (A ) × Xq (A ) → X p+q−1 (A ) ,
for p, q ∈ N, which therefore only respects the grading up to a shift over 1, suggest-
ing to shift the degree of all multi-derivations by 1. Namely, we define X p−1 (A ) :=
X p (A ), for p ∈ N, and we call p := p − 1 the shifted degree of an element P of
X p (A ) = X p (A ). Thus, elements of A have shifted degree −1, derivations of A
have shifted degree 0, and so on. Notice that p + q − 1 = p + q, so that the Schouten
bracket is a family of maps
[· , ·]S : X p (A ) × Xq (A ) → X p+q (A ) ,
A priori, we have only that [P, Q]S ∈ Hom(A ⊗(p+q+1) , A ), but the reader will eas-
ily check that indeed [P, Q]S ∈ X p+q+1 (A ) = X p+q (A ), by showing that [P, Q]S is
skew-symmetric and verifies the derivation property (3.2). In the case of a manifold,
the Schouten bracket of multivector fields is defined as in (3.36), but replacing the
functions Fi by local functions.
The Schouten bracket can be seen as a generalization of many classical elemen-
tary operations on functions, derivations and multi-derivations. First, let Q := F ∈ A
and P ∈ X p (A ), then [P, F]S = ıF P (see (3.5)). Second, let P := V ∈ X1 (A ) and Q ∈
Xq (A ), then [V , Q]S = LV Q, the Lie derivative of Q with respect to V (see (3.7)).
Third, the Schouten bracket of two skew-symmetric biderivations P, Q ∈ X2 (A ) is
given by
so that
ı[P,Q]S (dF ∧ ω )
= ı(ıF [P,Q]S ) ω − (−1) p+q dF ∧ ı[P,Q]S ω
= ı(ıF [P,Q]S ) ω − (−1) p+q dF ∧ ([[ıP , d] , ıQ ] ω )
= (−1)q ı[ıF P,Q]S ω + ı[P,ıF Q]S ω − (−1) p+q dF ∧ ([[ıP , d] , ıQ ] ω ) , (3.40)
to obtain the last line, which is similar to (3.6), and is a direct consequence of the
definition of [P, Q]S . Equation (3.40) has to be compared to [[ıP , d] , ıQ ] (dF ∧ ω ), i.e.,
to
(ıP dıQ − (−1) p dıQ∧P − (−1) pq ıQ dıP )(dF ∧ ω ) . (3.41)
Using (3.32) twice and (3.25) once, we find
If we substitute these results in (3.41), then we find nine terms which can, by using
the induction hypothesis twice, be written as follows:
(ıP dı(ıF Q) − (−1) p dı(ıF Q∧P) − (−1) pq ı(ıF Q) dıP ) ω = ı[P,ıF Q]S ω
(−1)q (ı(ıF P) dıQ + (−1) p dı(Q∧ıF P) − (−1) pq ıQ dı(ıF P) ) ω = (−1)q ı[ıF P,Q]S ω
dF ∧ −(−1) p+q ıP dıQ + (−1)q dıQ∧P − (−1) pq ıQ dıP ω =
−(−1) p+q dF ∧ ([[ıP , d] , ıQ ] ω ) .
Cartan’s formula leads to a simple proof of the main properties of the Schouten
bracket, as given in the following proposition.
Proposition 3.7. Let A be a commutative associative algebra. The Schouten bracket
[· , ·]S : X• (A ) × X• (A ) → X• (A ) ,
defines a graded Lie algebra structure on X• (A ), meaning that, for all homoge-
neous elements P, Q and R in X• (A ), of respective shifted degree p, q and r,
(1) [P, Q]S ∈ X p+q (A );
(2) [P, Q]S = −(−1) pq [Q, P]S , (graded skew-symmetry);
(3) (−1) pr [P, [Q, R]S ]S + (P, Q, R) = 0 (graded Jacobi identity).
Moreover, the associative product ∧ and the Lie bracket [· , ·]S are compatible in the
sense that [·, P]S is a graded derivation of (X• (A ), ∧) of degree p:
with P, Q and R as above. In view of the graded skew-symmetry of [· , ·]S , one also
has, for such elements,
3.3 The Schouten Bracket 83
Proof. (1) and (2) follow at once from the explicit formula for the Schouten bracket
(Eq. (3.36)). Notice that the graded skew-symmetry also follows from Cartan’s for-
mula (3.38). Indeed, the graded Jacobi identity for the graded commutator [· , ·],
which proves that [Q, P]S = −(−1) pq [P, Q]S , since P → ıP is injective. We next prove
the graded Jacobi identity for [· , ·]S . Let P ∈ X p (A ) and Q ∈ Xq (A ). We associate
to these multi-derivations, the coderivations P̃ (of degree p) and Q̃ (of degree q) of
the coalgebra (∧• A , Δ ), as in (3.35). We claim that
[P̃, Q̃] = [P, Q]S . (3.44)
(P̃ ◦ Q̃)(F1 , . . . , Fk )
[P̃, Q̃](F1 , . . . , Fk ) = [P, Q]S [F1 , . . . , Fk ] = [P, Q]S (F1 , . . . , Fk ) .
and the graded Jacobi identity for the Schouten bracket follows from the graded
Jacobi identity for the commutator of coderivations.
The Leibniz property of the Schouten bracket follows easily from Cartan’s for-
mula. Indeed, for P, Q and R as in (3.43), Cartan’s formula and (1) of Proposition 3.4
imply that
which yields (3.43), and hence also (3.42). We have used the formula3
[ψ , φ1 ◦ φ2 ] = [ψ , φ1 ] ◦ φ2 + (−1)d1 d φ1 ◦ [ψ , φ2 ] , (3.45)
For any Lie algebra (g, [· , ·]) there is a graded Lie algebra structure on ∧• g, which is
the natural extension of the Lie bracket [· , ·] on g. It is very similar to the Schouten
3 The formula is a graded version of the well-known matrix identity [x, yz] = [x, y]z + y[x, z]; also,
it is half of the Jacobi identity (A.13).
3.3 The Schouten Bracket 85
[[x1 ∧ · · · ∧ x p , y1 ∧ · · · ∧ yq ]] (3.46)
p q
:= (−1) pq ∑ ∑ (−1)i+ j [xi , y j ] ∧ x1 ∧ · · · ∧ xi ∧ · · · ∧ x p ∧ y1 ∧ · · · ∧ y j ∧ · · · ∧ yq .
i=1 j=1
Then (∧• g, ∧, [[· , ·]]) is a Gerstenhaber algebra. The graded Lie bracket [[· , ·]] on ∧• g
is called the algebraic Schouten bracket.
Proof. The proof of each one of the properties of (∧• g, ∧, [[· , ·]]) is either immediate
from the definitions or follows from a direct computation. Using the graded Leibniz
identity
[[Y ∧ Z, X]] = [[Y, X]] ∧ Z + (−1) pqY ∧ [[Z, X]] , (3.47)
valid for all X ∈ ∧ p g, Y ∈ ∧q g and Z ∈ ∧• g, the graded Jacobi identity can be
proven easily by recursion, since for three elements of shifted degree zero, it is just
the Jacobi identity in g, while if at least one element is a monomial of higher shifted
degree, say X, it can be written as X = X ∧ x, with x ∈ g and X ∈ ∧• g, of lower
degree. For a quick proof of (3.47), one clearly only needs to check that the signs
in the right-hand side of (3.47) are correct, which follows at once from the graded
skew-symmetry of [[· , ·]].
LP : X• (A ) → X•+p (A )
(3.48)
Q → [P, Q]S .
86 3 Multi-Derivations and Kähler Forms
The properties of the Schouten bracket, given in Proposition 3.7, lead at once to the
following properties for this generalized Lie derivative.
Proposition 3.10. Let A be a commutative associative algebra and let P, Q and R
be homogeneous elements of X• (A ) of shifted degrees p, q and r. Then
(1) LP Q ∈ X p+q (A ) ;
(2) LP (Q ∧ R) = Q ∧ LP R + (−1) pr (LP Q) ∧ R ;
(3) LP [Q, R]S = [LP Q, R]S + (−1) pq [Q, LP R]S .
In the geometrical context, the generalized Lie derivative becomes a Lie deriva-
tive with respect to an arbitrary p-vector field P; it associates to a q-vector field a
(q + p − 1)-vector field, i.e., it raises the degree of a multivector field by p = p − 1.
Since the calculus of differential forms has been more popularized than the cal-
culus of multivector fields, the notion of the Lie derivative of a differential form
(rather than of a multivector field) with respect to a vector field is better known.
We recall this definition and generalize it to the Lie derivative of a differential form
with respect to a multivector field. We start from the geometrical definition, which
is similar to the geometrical definition (3.11). Namely, let V be a vector field on a
manifold M and let ω be a differential p-form on M. Then LV ω is the differential
p-form on M, whose value at m ∈ M is given by
d
(LV ω )m = Φ ∗ ωΦt (m) ,
dt |t=0 t
where Φt is the local flow of V and Φt∗ the pull-back of ω by Φt , so that Φt∗ (ωΦt (m) )
is a differential p-form on Tm M (depending on t). The classical4 Cartan formula says
that
LV = ıV ◦ d + d ◦ ıV , (3.49)
a formula which we can take as a definition of the Lie derivative in the alge-
braic setting. Writing (3.49) as LV = [ıV , d], where we recall that [· , ·] denotes
the graded commutator (see (A.12)), leads to the following natural generalization
of the Lie derivative to an action of a multivector field (or multi-derivation) on
a differential form (Kähler form), namely we define the Lie derivative with re-
spect to a skew-symmetric multi-derivation P ∈ X p (A ) as the graded F-linear map
Ω • (A ) → Ω •−p (A ) of degree −p, given by
LP := [ıP , d] ,
where we recall that ıP was defined in (3.29). Explicitly, this means that for ω ∈
Ω • (A ),
LP ω := ıP (dω ) − (−1) p d(ıP ω ) . (3.50)
Comparing the Definitions (3.48) and (3.50) of the generalized Lie derivative,
one notices that there is some ambiguity in the definition of the generalized Lie
4 The classical Cartan formula should not be confused with Cartan’s formula (3.38), which relates
the Schouten bracket to the de Rham differential.
3.4 Exercises 87
which proves (3). Finally, using ıP∧Q = ıQ ◦ ıP (see Proposition 3.4) and (3.45),
3.4 Exercises
1. Prove the explicit formula (3.28) for a Kähler k-form ω on an arbitrary commu-
tative associative algebra A ; the arguments Vi are in this context arbitrary deriva-
tions of A .
2. Give an alternative proof of the graded Jacobi identity for the Schouten bracket
by using Cartan’s formula (3.38).
3. Let A be an arbitrary commutative associative algebra. Prove the following
formulas, for F ∈ A and ω ∈ Ω • (A ):
LFP ω = F LP ω − (−1) p dF ∧ ıP ω ,
LP (F ω ) = F LP ω + ı(ıF P) ω .
88 3 Multi-Derivations and Kähler Forms
F[x, y, z]
A := .
yx, yz, y2
∂ ∂
a. Show that P := y ∧ induces a non-zero skew-symmetric biderivation P
∂y ∂z
of A ;
b. Show that P does not belong to the exterior algebra ∧2 X1 (A ).
3.5 Notes
For the algebraic analog, which is the calculus of Kähler forms, we refer to
Hartshorne [93].
Multivector fields are rarely explicitly treated in books on differential geome-
try; even if those fields themselves are particular cases of tensor fields, everywhere
treated in detail, their algebraic structure does not derive from the general tensor
formalism. The calculus of multivector fields is presented in detail in Vaisman [194]
and in Bhaskara–Viswanath [23]. Cannas da Silva–Weinstein [34] concentrate on
the structure of the space of multivector fields as a Gerstenhaber algebra, which
they present as a particular case of the Gerstenhaber algebra of a Lie algebroid.
The Schouten bracket, also called the Schouten–Nijenhuis bracket, was intro-
duced by Schouten in [177, 178]; a generalization of the Schouten bracket, which
also generalizes the Fröhlicher–Nijenhuis bracket on vector valued differential
forms, is given in Vinogradov [197].
Chapter 4
Poisson (Co)Homology
Table 4.1 A summary of the notations which we use for Lie and Poisson homology and cohomol-
ogy. (A , ·, π ) is an arbitrary Poisson algebra over F, and g is a Lie algebra with a representation
space V .
In Sections 4.1 and 4.2, we construct the various complexes which lead to the ho-
mologies and cohomologies which we will consider. In Section 4.3 we describe a
few natural operations in Poisson cohomology and homology and we show in Sec-
tion 4.4 that the Poisson cohomology and homology spaces of a Poisson manifold
are, under certain conditions, isomorphic to each other.
Throughout this chapter, F is an arbitrary field of characteristic zero.
Before constructing the Poisson cohomology complex, and deriving the cohomol-
ogy spaces from it, we recall Lie algebra cohomology, which is formally very sim-
ilar to Poisson cohomology, and more widely known (see [156]). The basic notions
which we recall will also be useful when we compare Poisson and Lie algebra co-
homology in the case of the canonical Poisson structure, defined on the dual of a
Lie algebra (see Chapter 7).
Let (g, [· , ·]) be a Lie algebra over F. We recall that a representation of g is a Lie
algebra homomorphism ρ : g → End(V ), where V is an F-vector space, called the
representation space and the Lie algebra structure on End(V ) is given by the com-
mutator: φ ◦ ψ − ψ ◦ φ , for φ , ψ ∈ End(V ). The vector space V , together with the
representation ρ , is called a g-module. We write ρx for ρ (x), where x ∈ g. We fix a
representation (ρ ,V ) of g and we introduce, for k ∈ N, the F-vector space
where the symbol over an argument means that we omit that argument. Ele-
ments of Ker δLk are called Lie k-cocycles, while elements in Im δLk−1 are called Lie
k-coboundaries. The Jacobi identity for [· , ·] and the fact that ρ is a Lie algebra ho-
momorphism permits one to show that δLk ◦ δLk−1 = 0, for every k ∈ N∗ . In words:
every coboundary is a cocycle. The converse is not true in general and the cohomol-
ogy spaces precisely measure the obstruction. Namely, for k ∈ N∗ , we define
and HL0 (g;V ) := Ker δL0 . Putting these spaces together, we obtain a graded F-vector
space
HL• (g;V ) := HLk (g;V ) ,
k∈N
For example, for k = 0, the coboundary operator is given by δL0 (v)(x) = ρx v, for
v ∈ V C0 (g;V ) and x ∈ g. It follows that HL0 (g;V ) = Ker δL0 is given by
HL0 (g;V ) = Ker ρx ,
x∈g
In this case, the Lie algebra cohomology of g is called the trivial Lie algebra coho-
mology of g and is simply denoted by HL• (g). It follows from (4.2) that Im δL0 = {0}
and that Ker δL1 = {c ∈ g∗ | ∀x, y ∈ g, c([x, y]) = 0} , so that
∗
HL1 (g) g/ [g, g] .
For a semi-simple Lie algebra g, for example, it is well known that [g, g] = g (see
[190, Ch. 20]). Therefore, HL1 (g) = 0 when g is semi-simple, which is a special case
of Whitehead’s lemma (Lemma 4.1).
Example 4.3. Let ρ := ad, the adjoint representation of g on itself, which is given
by ρx = adx := [x, · ], for x ∈ g. Then
Ck (g; g) = Hom(∧k g, g)
for x, y ∈ g and c ∈ Hom(g, g). Since δL0 (x) = − adx , for x ∈ g C0 (g; g), the space
of 1-coboundaries Im δL0 consists of all derivations of g of the form adx , where x ∈ g;
these derivations are called inner derivations. It follows that HCE1 (g) is the quotient
of the space of all derivations of g, by the space of all inner derivations of g. The
elements of this quotient are called the exterior derivations of g.
The spaces HCE 2 (g) and H 3 (g) play an important rôle in the theory of formal
CE
deformations of Lie algebras. This fact is quite similar (by forgetting the multi-
derivation aspect) to what happens in the case of Poisson algebras. Therefore, we
will not make it explicit in the Lie algebra case, but in the Poisson algebra case, see
Section 13.2.
We now define Poisson cohomology for a Poisson algebra (A , ·, {· , ·}); the Poisson
bracket on A will also be denoted by π , so that π (F, G) = {F, G}, for F, G ∈ A .
For k ∈ N, the space of k-cochains of the Poisson cohomology complex is by
definition Xk (A ), the F-vector space of skew-symmetric k-derivations of A (see
Section 3.1.1). The Poisson coboundary operator is the Poisson analog of the
Chevalley–Eilenberg coboundary operator (4.3), where Hom(∧k g, g) has been re-
placed by Xk (A ), and where the Lie bracket [· , ·] is the Poisson bracket {· , ·}.
Namely, the graded F-linear map (of degree 1) δπ : X• (A ) → X•+1 (A ) is defined,
for Q ∈ Xq (A ), where q ∈ N, by
q
δπq (Q)[F0 , . . . , Fq ] := ∑ (−1)i i , . . . , Fq ]
Fi , Q[F0 , . . . , F (4.4)
i=0
+ ∑ (−1)i+ j Q i , . . . , Fj , . . . , Fq ,
Fi , Fj , F0 , . . . , F
0i< jq
The elements of Ker δπq are called Poisson q-cocycles, while the elements of Im δπq−1
are called Poisson q-coboundaries. Elements of the q-th Poisson cohomology space
are Poisson q-cocycles modulo Poisson q-coboundaries,
Remark 4.4. Let (A , ·, {· , ·}) be a Poisson algebra over F. Forgetting the asso-
ciative product, (A , {· , ·}) is simply a Lie algebra and it makes sense to con-
sider C• (A , A ), with the Chevalley–Eilenberg coboundary operator (4.3). This
yields precisely (4.4), so that the Poisson coboundary operator of (A , ·, {· , ·}) is
the Chevalley–Eilenberg coboundary operator, restricted to the multi-derivations
of (A , ·).
For small q, the Poisson coboundary operator δπq has a natural interpretation, which
yields a natural interpretation for the Poisson cohomology space Hπq (A ). One easily
reads off from (4.4) that for F ∈ X0 (A ) = A and for V ∈ X1 (A ),
Hπ0 (A ) = Cas(A ) ,
Poisson 1-cocycles are Poisson derivations and Poisson 1-coboundaries are Hamil-
tonian derivations. Denoting the space of all Poisson derivations of A by P(A ), it
follows that
P(A )
Hπ1 (A ) = .
Ham(A )
Poisson 2-cocycles Q ∈ X2 (A ) are skew-symmetric biderivations which satisfy
[π , Q]S = 0, i.e., they are the ones which are compatible with {· , ·}, see Section 3.3.2;
Poisson 2-coboundaries are biderivations, obtained as a Lie derivative of the Poisson
structure. It follows that
skew-symmetric biderivations compatible with {· , ·}
Hπ2 (A ) = .
Lie derivatives of {· , ·}
As we will see in Section 13.2, Hπ2 (A ) and Hπ3 (A ) show up naturally in deforma-
tion theory.
There is a natural morphism from de Rham cohomology to Poisson cohomology,
which we explain here in the algebraic context; this morphism is defined in the same
way in the case of a smooth manifold. We first construct an A -linear map
π : Ω 1 (A ) → X1 (A )
(4.8)
G dF → GXF .
It is clear that this map is well-defined; for example, we have for all F, G ∈ A that
where we used (3) of Proposition 1.4 in the last step. The map π extends naturally
to an A -linear map
∧• π : Ω • (A ) → ∧• X1 (A ) → X• (A ) ,
d
Ω q (A ) Ω q+1 (A )
∧q π ∧q+1 π
q
δπ
Xq (A ) Xq+1 (A )
q
It leads for every q ∈ N to an F-linear map HdR (A ) → Hπq (A ).
where we also used that each Hamiltonian vector field XFi is a Poisson 1-cocycle,
[XFi , π ]S = 0. This leads to the commutativity of the above diagram. It follows that
∧q π sends q-cocycles to q-cocycles and sends q-coboundaries to q-coboundaries.
Thus, it induces a map in cohomology.
Remark 4.6. Contrary to what we will see for Poisson homology in Section 4.2,
Poisson cohomology is not a functor: a homomorphism ϕ : A → B between two
Poisson algebras does not lead in general to a homomorphism between the spaces
of all multi-derivations of A and B, not even between their corresponding Poisson
cohomology spaces. In the case of manifolds, for example, it is well known that
vector fields cannot be transferred, in either direction, by a smooth map, which is
not a diffeomorphism.
properties, as we will see. As in the previous section, we start with the Lie algebra
case.
The basic setup of Lie algebra homology is the same as in the case of Lie algebra
cohomology: (g, [· , ·]) is a Lie algebra and (ρ ,V ) is a representation of g. For k ∈ N,
we define
Ck (g;V ) := V ⊗ ∧k g ,
whose elements are called V -valued k-chains of g. Note that C0 (g,V ) V , since
∧0 g = F. The Lie boundary operator ∂kL : Ck (g;V ) → Ck−1 (g;V ) is the linear map
given, for all v ∈ V and x1 , . . . , xk ∈ g, by:
∂kL v ⊗ (x1 ∧ · · · ∧ xk )
k
= ∑ (−1)i ρxi v ⊗ x1 ∧ · · · ∧ xi ∧ · · · ∧ xk
i=1
+ ∑ (−1)i+ j v ⊗ [xi , x j ] ∧ x1 ∧ · · · ∧ xi ∧ · · · ∧ xj ∧ · · · ∧ xk
1i< jk
∂1L (v ⊗ x) = −ρx v .
Therefore, the subspace Im ∂1L of V is denoted by gV and H0L (g;V ) = V /gV , since
∂0L = 0. When V = g and ρ = ad, as in Example 4.3, the boundary operator takes
the following form
4.2 Lie Algebra and Poisson Homology 99
∂kL x0 ⊗ (x1 ∧ · · · ∧ xk )
k
= ∑ (−1)i [xi , x0 ] ⊗ x1 ∧ · · · ∧ xi ∧ · · · ∧ xk (4.9)
i=1
+ ∑ (−1)i+ j x0 ⊗ [xi , x j ] ∧ x1 ∧ · · · ∧ xi ∧ · · · ∧ xj ∧ · · · ∧ xk .
1i< jk
In this case, gV = [g, g], so that H0CE (g) := H0L (g; g) = g/[g, g]. This space also
appears when one computes H1L (g; F), where ρ : g → F is the trivial representation,
as in Example 4.2. Indeed, in this case,
+ ∑ i ∧ · · · ∧ dF
(−1)i+ j F0 d Fi , Fj ∧ dF1 ∧ · · · ∧ dF j ∧ · · · ∧ dFk ,
1i< jk
∂ π = Lπ = [ıπ , d] = ıπ ◦ d − d ◦ ıπ .
[∂ π , d] = ∂ π ◦ d + d ◦ ∂ π = 0 , (4.10)
π π π
[∂ , ıπ ] = ∂ ◦ ıπ − ıπ ◦ ∂ = 0 , (4.11)
Proof. For the biderivation π = {· , ·}, the internal product ıπ is, according to (3.29),
explicitly given by
where F0 , . . . , Fk ∈ A , so that
and
+ ∑ i . . . dF
(−1)i+ j+1 F0 d Fi , Fj ∧ dF1 ∧ · · · ∧ dF j ∧ · · · ∧ dFk .
1i< jk
Taking the difference of these two formulas leads at once to ∂ π = [ıπ , d] = Lπ . The
properties of the Lie derivative (Proposition 3.11) imply that ∂ π ◦ d + d ◦ ∂ π = 0,
and that
[∂ π , ıπ ] = [Lπ , ıπ ] = ı[π ,π ]S = 0 ,
where we used in the last equality that π is a Poisson structure. This yields (4.10)
and (4.11). Similarly, since Lπ has degree −1,
2 Lπ ◦ Lπ = [Lπ , Lπ ] = L[π ,π ]S = 0 ,
and we call elements of Ker ∂kπ Poisson k-cycles and elements of Im ∂k+1
π Poisson
k-boundaries. Putting the Poisson homology spaces together leads to the graded
vector space H•π (A ) := ⊕k∈N Hkπ (A ). In the case of a Poisson manifold (M, π ),
4.3 Operations in Homology and Cohomology 101
A
H0π (A ) = .
{A , A }
First, it follows from these formulas that if Q and R are Poisson cocycles, then Q ∧ R
and [Q, R]S are Poisson cocycles. This means that both operations can be restricted to
cocycles. Second, it follows from the same formulas that either product of a Poisson
cocycle with a Poisson coboundary is a Poisson coboundary. This means that, for
Poisson cocycles Q and R, the cohomology class of Q ∧ R and the cohomology
class of [Q, R]S are independent of the representatives Q and R of their cohomology
classes. Thus, we have two well-defined products in Poisson cohomology, which we
also denote by ∧ and [· , ·]S . Obviously, their algebraic properties are the same as on
X• (A ) and we obtain:
Proposition 4.9. For every Poisson algebra (A , ·, π ), (Hπ• (A ), ∧, [· , ·]S ) is a Ger-
stenhaber algebra.
We now consider the action of Poisson cohomology on Poisson homology, which
comes from the internal product and from the Lie derivative. Remembering that
the Poisson boundary operator ∂ π is the Lie derivative Lπ , we take from Proposi-
tion 3.11 the formulas which compute the graded commutator of a Lie derivative
with an internal product, or with another Lie derivative, namely,
It follows from these formulas, as above, that ı and L induce indeed actions of the
Poisson cohomology of A on the Poisson homology of A .
For a symplectic1 manifold (M, ω ), the answer to the above question is yes. In-
deed, since the symplectic structure ω is preserved by all Hamiltonian flows, the
same is true for the Liouville form ω d on M, where dim M = 2d. Thus, every sym-
plectic manifold is a unimodular Poisson manifold. Many properties of symplectic
manifolds generalize to the case of unimodular Poisson manifolds. For example, for
unimodular Poisson manifolds the Poisson cohomology and the Poisson homology
spaces are isomorphic to each other (Theorem 4.18).
We first list a few basic facts about volume forms on a real orientable manifold
M. Suppose that λ is an arbitrary volume form on M. Recall that this means that
λ is a differential d-form on M, where d := dim M, vanishing at no point of M,
and that such a form exists on a manifold if and only if the manifold is orientable;
moreover, two such forms differ only by a factor, which is a nowhere vanishing
smooth function on M. We have that dλ = 0 and that dF ∧ λ = 0, for every F ∈
F (M). Combined with (3.32), this yields
V [F]λ = dF ∧ ıV λ , (4.13)
for every vector field V ∈ X1 (M). Also, if π is a Poisson structure on M, then (4.12)
specializes to
ıXF λ = dF ∧ ıπ λ . (4.14)
The Lie derivative of λ with respect to some vector field is again a top form, hence it
is a multiple of λ . For F ∈ F (M), let us denote by DF ∈ F (M) the smooth function
on M defined by LXF λ = DF λ . We show that the map F → DF is a derivation.
Indeed, for all F, G ∈ F (M), we have that
Definition 4.10. Let (M, π ) be a real Poisson manifold, equipped with a volume
form λ . The vector field Φλ , defined by
for all F ∈ F (M), is called the modular vector field of π with respect to λ .
When the chosen volume form is clear from the context, we simply write Φ for Φλ .
Notice that the modular vector field Φ can also be defined as the unique vector field
on M, satisfying
Lπ λ = ı Φ λ . (4.17)
This is seen by writing Φ [F]λ in two different ways. On the one hand, (4.13) tells
us that Φ [F]λ = dF ∧ ıΦ λ , for all F ∈ F (M). On the other hand, (4.16), (4.14) and
the formula Lπ = ıπ ◦ d − d ◦ ıπ imply that
for all F ∈ F (M). It follows that dF ∧ıΦ λ = dF ∧Lπ λ for every F ∈ F (M), which
leads to (4.17).
In view of Definition 4.10, the question which we posed at the beginning of this
section now becomes:
Question: “Let M be a Poisson manifold. Does there exist a volume form λ on M,
with respect to which the modular vector field Φ is zero?”
By giving a cohomological interpretation of the modular vector field, we will be
able to answer the question in the next section.
The modular vector field which we constructed in the previous section depends on
the choice of volume form on M. We will now show that when the volume form is
changed, the modular vector field will be modified by a Hamiltonian vector field.
We also show that the modular vector field is a Poisson vector field, so that it defines
a Poisson cohomology class which, by the above, is independent of the choice of
volume form.
Proposition 4.11. Let (M, {· , ·}) be a real Poisson manifold, which is orientable.
For λ a volume form on M, the modular vector field Φ has the following properties:
(1) Φ is a Poisson vector field, i.e., it is a Poisson 1-cocycle;
(2) The Poisson cohomology class of Φ is independent of the chosen volume
form λ ;
(3) If Φ is a Hamiltonian vector field, i.e., if the Poisson cohomology class of Φ
is zero, then there exists a volume form λ , with respect to which the modular
vector field is zero (and therefore λ is preserved by all the Hamiltonian flows).
4.4 The Modular Class of a Poisson Manifold 105
Proof. In order to prove (1) we need to verify that Φ [{F, G}] = {Φ [F], G} +
{F, Φ [G]}, for all F, G ∈ F (M). The basic properties of the Poisson bracket and
the Lie derivative yield:
Φ [{F, G}] λ = LX{F,G} λ = L[XG ,XF ] λ = LXG , LXF λ
= LXG (Φ [F]λ ) − LXF (Φ [G]λ )
= ({Φ [F], G} + {F, Φ [G]}) λ ,
This proves (1). Let us consider now another volume form μ on M and let us denote
by Φλ and Φμ the modular class of {· , ·} with respect to λ , respectively μ . There
exists a nowhere vanishing function ρ ∈ F (M) such that μ = ρ λ . Using that λ and
μ are both top forms and (4.13), we compute, for an arbitrary F ∈ F (M),
to conclude that
1
Φμ [F] = Φλ [F] − Xρ [F] ,
ρ
for all F ∈ F (M), and hence that
1
Φμ = Φλ − Xρ = Φλ − Xln |ρ | .
ρ
This shows that the vector fields Φλ and Φμ differ by a Hamiltonian vector field,
i.e., a Poisson 1-coboundary. The cohomology class of the modular form is therefore
independent of the chosen volume form, which proves (2). The previous calculation
shows that if this cohomology class is trivial, i.e., Φ = XF for some F ∈ F (M),
then μ := exp(F)λ is a volume form whose modular vector field is zero.
r
s
λ := (dqi ∧ dpi ) dz j ,
i=1 j=1
and we denote
s
λk := (dqi ∧ dpi ) dz j ,
i=k j=1
from which it follows that LXF λ = 0, i.e., the modular vector field Φ is zero on U.
Then (2) of Proposition 4.11 implies that all modular vector fields are Hamiltonian
on U, hence they are tangent on U to all symplectic leaves. The second property is
obvious, since the modular class depends on {· , ·} only.
Example 4.15. If m is a point where the rank of the Poisson structure is not locally
constant, then it may happen that there does not exist a neighborhood of m on which
the modular vector field is Hamiltonian. In fact, the modular vector field may be
non-vanishing at a point where the rank of the Poisson structure vanishes. Consider
for example the Poisson structure {· , ·} = x ∂∂x ∧ ∂∂y on R2 , equipped with the volume
∂
form dx ∧ dy. The rank at the origin is zero, while Φ = ∂ y . In particular, the modular
class of this Poisson structure is not zero.
We now show that the modular vector field is (up to a sign) the divergence of the
Poisson structure (with respect to λ ). The divergence operator (with respect to λ ) is
the graded F-linear map (of degree −1),
X• (M)
Ω d−• (M)
Div d (4.18)
where d := dim M. The diagram involves the star operator , which is the family of
F (M)-linear isomorphisms : Xk (M) → Ω d−k (M), defined for Q ∈ Xk (M) by
Q := ıQ λ . (4.19)
108 4 Poisson (Co)Homology
For example, taking the standard volume form on Rd , i.e., λ = dx1 ∧ · · · ∧ dxd
the star of a vector field V = ∑di=1 Fi ∂ /∂ xi , respectively of a bivector field P =
∑i< j Pi j ∂ /∂ xi ∧ ∂ /∂ x j is given by
d
V = ∑ (−1)i−1 Fi dx1 ∧ · · · ∧ dx
i ∧ · · · ∧ dxd ,
i=1
P = ∑ i ∧ · · · ∧ dx
(−1)i+ j−1 Pi j dx1 ∧ · · · ∧ dx j ∧ · · · ∧ dxd ,
1i< jd
In general, the divergence of a bivector field, written in terms of vector fields, can
be computed using the following proposition.
Proposition 4.16. Let λ be a volume form on an orientable manifold M and let V
and W be two vector fields on M. The divergence of V ∧ W with respect to λ is
given by
Div(V ∧ W ) = Div(W )V − Div(V )W − [V , W ] . (4.23)
Proof. According to the definition of the (generalized) Lie derivative (3.50) and its
properties (4) and (2) in Proposition 3.11, we have that
Proposition 4.17. Let (M, π ) be a real Poisson manifold, which is assumed ori-
entable. For every volume form λ on M,
Φ = − Div(π ) . (4.24)
Since is an isomorphism, this shows that the modular vector field is minus the
divergence of π . Since Div ◦ Div = −1 ◦ d ◦ d ◦ = 0, the modular vector field is
divergence-free.
According to (4.25), the modular vector field can also be defined by the formula
dπ = − Φ ;
forgetting the stars, the modular vector field is the differential of the Poisson bracket.
The string (4.25) also contains the identity Lπ λ = Φ , which amounts to
∂ π (λ ) = Φ . (4.26)
We now prove that, for unimodular Poisson manifolds, the Poisson homology and
cohomology spaces are isomorphic. Let (M, π ) be a d-dimensional Poisson mani-
fold. We first show how the Poisson boundary and coboundary operators are related
via the modular vector field Φ . To do this, we specialize Cartan’s formula (3.38)
[[ıP , d] , ıQ ] = ı[P,Q]S ,
which we also write, using ıQ ◦ ıP = ıP∧Q , valid for P, Q ∈ X• (M) (see Proposi-
tion 3.4), as follows,
110 4 Poisson (Co)Homology
The latter formula says that the modular vector field Φ measures the non-commuta-
tivity of the following diagram
X• (M)
Ω d−• (M)
Φ ∧
δπ ∂π (4.29)
The reader is invited to compare the diagrams (4.29) and (4.18), keeping in mind
that both depend on the volume form λ , but only the second one depends on the
Poisson structure π . Equation (4.28) leads at once to the following duality theorem.
Proposition 4.18. For every unimodular Poisson manifold (M, {· , ·}) of dimension
d, the Poisson cohomology spaces and Poisson homology spaces are isomorphic:
π
Hπk (M) Hd−k (M), 0kd. (4.30)
4.5 Exercises
1. Show that ∧ does not induce a product in Poisson homology (Hint: compute
∂ π (dF ∧ dG) for F, G ∈ A ).
2. Show that the Poisson cohomology spaces are modules over the algebra of
Casimirs.
3. Find a formula for the Poisson cohomology of the tensor product of two Poisson
algebras.
4. Prove property (2) of Proposition 4.11 by using the star operator (4.19).
5. Suppose that g is a Lie algebra, equipped with a symmetric bilinear form · | ·
which is ad-invariant, i.e., [x, y] | z = x | [y, z] for all x, y, z ∈ g. Consider the ele-
4.6 Notes 111
ment C ∈ Hom(∧3 g, F), defined by C(x, y, z) := [x, y] | z. Show that C is a 3-cocycle
in the trivial Lie algebra cohomology.
6. Consider the Poisson structure π on F2 given by
∂ ∂
π := x ∧ .
∂x ∂y
a. Show that the modular class of π is not trivial;
b. Compute the Poisson homology and the Poisson cohomology of π ;
c. Conclude from part b. that the conclusion of Proposition 4.18 does not hold for
π (F2 )
the Poisson manifold (F2 , π ), i.e., that the vector spaces Hπk (F2 ) and H2−k
are not isomorphic, for k = 0, 1, 2.
∂ ∂ ∂ ∂ ∂ ∂
π := x2 ∧ + y2 ∧ + z2 ∧ .
∂y ∂z ∂z ∂x ∂x ∂y
4.6 Notes
In Chapter 2, we have given a few basic constructions which allow one to build
new Poisson structures from given ones. In the present chapter, we explain a few
more advanced constructions, which all fall under the general concept of reduc-
tion. Roughly speaking, reduction means that the object under study (here a Poisson
structure), is replaced by an object of the same type, but on a manifold of smaller
dimension, in classical terms “with fewer degrees of freedom”. These constructions
are based on geometrical considerations, yet their algebraic formulation highlights
some key elements of the reduction mechanism. We therefore present all construc-
tions in both algebraic and geometrical terms.
We present two constructions for reducing Poisson structures. Geometrically
speaking, the first one, called Poisson reduction, deals with Poisson structures on
quotients of Poisson manifolds, or, more generally on quotients of (coisotropic) sub-
manifolds of Poisson manifolds. Poisson reduction is presented in Section 5.2. The
second construction, called Poisson–Dirac reduction, concerns Poisson structures
on submanifolds, which are not necessarily Poisson submanifolds; a typical exam-
ple is a submanifold which is transverse to a symplectic leaf, leading to the notion
of transverse Poisson structure. Poisson–Dirac reduction is discussed in Section 5.3.
As an application of these constructions, we consider in Section 5.4 Poisson va-
rieties and Poisson manifolds on which a group acts; we show how a Poisson struc-
ture is inherited on the quotient space, the subspace of fixed points of the action, and
on the reduced space of the so-called momentum map. For the convenience of the
reader, and in order to fix both our notations and conventions, the basic facts on Lie
groups, and their relation with Lie algebras are recalled in the first section of this
chapter.
As in the previous chapters, all Poisson algebras are defined over an arbitrary
field F of characteristic zero, all Poisson varieties are affine varieties over F and the
Poisson manifolds are real or complex manifolds.
We recall in this section the natural correspondence between Lie groups and Lie al-
gebras, we give the basic definitions and properties of group actions and we explain
construction of invariant (multi-)vector fields on a Lie group.
A Lie group is a set G endowed with two structures, a group structure and a manifold
structure, which are required to be compatible, in the sense that the product map (the
group structure)
μ : G × G → G : (g, h) → gh
is assumed to be smooth (or holomorphic). Then the inverse map
inv : G → G : g → g−1
Lg : G → G : h → gh ,
Rg : G → G : h → hg ,
Cg : G → G : h → ghg−1 .
The tangent maps of μ and inv can be expressed in terms of these diffeomorphisms:
since μ (g, h) = Lg (h) = Rh (g) for all g, h ∈ G, the tangent map of μ at (g, h) is given
by
T(g,h) μ (v, w) = Tg Rh (v) + Th Lg (w) (5.1)
for all (v, w) ∈ T(g,h) (G × G) Tg G × Th G. Differentiating the identities Lg ◦ Lg−1 =
1G at e and μ (g, inv(g)) = e at g ∈ G, leads, upon using (5.1), to the following
expression for the tangent map of inv at g ∈ G,
Th Lg (←
−
x h ) = Th Lg (Te Lh (x)) = Te (Lg ◦ Lh )(x) = Te Lgh (x) = ←
−
x gh ,
(5) Let G be a connected and simply connected Lie group with Lie algebra g, and
let h be a Lie subalgebra of g. There exists a unique connected Lie subgroup H
of G with Lie algebra h.
In a different vein, a Lie group and its Lie algebra are related as follows: there exists
a diffeomorphism exp between a neighborhood of the origin o in g and a neighbor-
hood of the unit e in G, having the fundamental property that for every x ∈ g, the
integral curve of ← −
x which starts from o is given (for small |t|) by t → exp(tx). In
fact, one possible construction of the diffeomorphism exp is by integrating, for x in
a neighborhood of o in g, the left-invariant vector fields ← −
x . We refer to exp as the
exponential map.
We end this section with a few comments on algebraic groups, which for us will
always be complex (defined over C). By definition, a (complex) algebraic group is
an algebraic variety with a group structure, such that the product map and the inverse
map are morphisms (of algebraic varieties). Much of what we discussed in this
section carries over to algebraic groups, for example there is canonically associated
to every algebraic group G a Lie algebra, whose underlying vector space g is the
(Zariski) tangent space to G at the unit e of G. However, there is in general no
algebraic analog of the exponential map and parts (3) and (4) of Lie’s theorem do
not hold in general. When dealing with quotients of algebraic varieties by algebraic
group actions, we will often assume that the algebraic group is reductive, which
means that all its finite-dimensional representations are completely reducible; as we
will explain at the end of Section 5.1.2, such quotients are rather well-behaved.
for all m ∈ M. The virtue of the latter group action is that it is a representation
of G (on F (M)); a representation of a Lie group G on a vector space V is a group
action χ : G × V → V with the property that for every g ∈ G the map χg : V → V ,
defined for all x ∈ g by χg (x) := χ (g, x), is a linear map (an endomorphism of V ).
The action of an algebraic group on an algebraic variety is defined in an analogous
way, demanding now that the map χ is a morphism of algebraic varieties.
The main group actions which we will consider in this chapter are the group
actions of G on itself, given by
5.1 Lie Groups and (Their) Lie Algebras 117
L : G × G → G : (g, h) → gh ,
R : G × G → G : (g, h) → hg−1 ,
C : G × G → G : (g, h) → ghg−1 .
These actions are called left translation, right translation and conjugation (in that
order). Notice that taking the inverse of g in the definition of R is essential to make
it into an action (otherwise it would be a so-called right action); we stress that it
amounts to R(g, h) = Rg−1 (h) for all g, h ∈ G.
For a fixed group action χ of G on a manifold M, the fundamental vector field
associated to x ∈ Te G, denoted x, is the vector field on M, whose value at m ∈ M is
given by
d
xm [F] := F(χ (exp(−tx), m)) , (5.5)
dt |t=0
for all functions F, defined on a neighborhood of m in M. Writing g = exp(tx),
one sees that Eq. (5.5) is the infinitesimal version of (5.4). Applied to the case of
right translation (recall that R(g, h) = Rg−1 h = hg−1 ), this yields for all functions F,
defined in a neighborhood of g in G,
d d
xg [F] = F(R(exp(−tx), g)) = F(g exp(tx))
dt |t=0 dt |t=0
d
= ((F ◦ Lg ) exp(tx)) = x[F ◦ Lg ]
dt |t=0
= Te Lg (x)[F] = ← −
x g [F] .
We conclude that xg = ← −
x g , for all g ∈ G, i.e., x = ←
−
x . In words: the left-invariant
vector field on G, whose value at e is x, is the fundamental vector field of right
translation, corresponding to x. Similarly, the fundamental vector field of left trans-
lation on G, corresponding to x ∈ g, is given by −− →x . Also, since Cg = Rg−1 ◦ Lg ,
the fundamental vector field of conjugation on G, corresponding to x ∈ g, is given
by ←−
x −→ −x . The identity
inv ◦ Rg = Lg−1 ◦ inv
implies that a vector field V on G is right-invariant if and only if the vector field
on G, defined1 by inv∗ V , is left-invariant. It follows, using Te inv = −1Te G , that
inv∗ ←
−
x = −− →
x , for every x ∈ g.
Given a group action of a Lie group G on a manifold M, the map g → X1 (M),
which associates, as above, to x ∈ g the fundamental vector field x on M, admits
as a natural generalization a map ∧• g → X• (M), which associates to a multivector
X of g a multivector field X on M. For X ∈ ∧ p g of the form X = x1 ∧ · · · ∧ x p , the
p-vector field X is defined as
1Recall that for a diffeomorphism Φ : M → N between manifolds and for V a vector field on M,
we denote by Φ∗ V the vector field on N, which is defined by (Φ∗ V )Φ (m) := Tm Φ (Vm ), for all
m ∈ M, and similarly for multivector fields on M.
118 5 Reduction
X := x1 ∧ · · · ∧ x p
and is called the fundamental p-vector field, corresponding to X. In the case of right
←− →
−
and left translations on a Lie group G, we write X , respectively (−1) p X for the
fundamental multivector field, associated to X ∈ ∧ g. Equivalently, for X ∈ ∧ p g, the
p
←
− →
− ←
− →
−
vector fields X and X can be defined by X g := ∧ p (Te Lg )X and X g := ∧ p (Te Rg )X,
←−
for all g ∈ G. As above, it follows from this definition that X is a left-invariant
→
−
multivector field and that X is a right-invariant multivector field.
←−
We show in the following proposition that the map, which sends X ∈ ∧• g to X ,
is a homomorphism of Gerstenhaber algebras. 2
Proposition 5.2. Let G be a Lie group with Lie algebra g. For all X,Y ∈ ∧• g, we
have
←−−− ← − ← − ←−−− ← −← −
X ∧Y = X ∧ Y , [[X,Y ]] = X , Y S , (5.6)
←
−
so the map X → X preserves the wedge product and the Schouten bracket. Up to a
→
−
sign, the same is true for the map X → X : for all X,Y ∈ ∧• g, we have
−−−→ − → → − −−−→ →
− −→
X ∧Y = X ∧ Y , [[X,Y ]] = − X , Y S . (5.7)
Moreover, ←
−−→
X,Y S=0, (5.8)
for all X,Y ∈ ∧• g.
equations in (5.7). We now prove (5.8). In view of the graded Leibniz rule for the
Schouten bracket, it suffices to prove (5.8) when X and Y are elements of g, which
we write as x and y. The vector fields ← −
x and − →
y are the fundamental vector fields
of the actions of right, respectively left, translation of G on itself. These actions
commute, hence the flows of ← −
x and − →y also commute. The commutator [← −
x ,−
→
y ] is
therefore zero.
For algebraic groups there is, in general, no exponential map, but for linear ac-
tions on vector spaces (representations), such as the actions which we will con-
sider in the next section, taking the differential of the action, viewed as a map
χ : G → End(V ), yields a linear vector field on V , which is the analog (up to a
sign) of the fundamental vector field, defined in (5.5).
2 Recall that (∧• g, ∧, [[· , ·]]) is a Gerstenhaber algebra, just like (X• (M), ∧, [· , ·]S ), where M is an
arbitrary manifold (see Proposition 3.9).
5.1 Lie Groups and (Their) Lie Algebras 119
To finish this section, we recall some rather general conditions under which the
quotient space M/G of a group action G × M → M is of the same type as M, i.e., is
an affine variety or a manifold.
(1) Suppose that M is an affine variety on which a reductive algebraic group G acts.
Then the algebra of G-invariant functions F (M)G is finitely generated, hence is the
algebra of regular functions on an affine variety, which can be identified with the
orbit space M/G. Since, in particular, every finite group G is reductive, the quotient
of an affine variety by a finite group is an affine variety.
(2) Suppose that M is a manifold and that G is a Lie group, which acts properly
and locally freely on M. Then the orbit space M/G has a unique manifold structure
for which the canonical map p : M → M/G is smooth. Interesting particular cases
include free actions of a finite group and locally free actions of a compact group on
a manifold. We recall that an action χ : G × M → M is called a proper action, if
the map
χ × p2 : G × M → M × M
(5.9)
(g, m) → (gm, m)
is a proper map, i.e., the inverse image of every compact subset of M × M is a
compact subset of G × M. Also, χ is called a locally free action if, for every m ∈ M,
there exists a neighborhood U of the identity in G such that the restriction to U of
the map g → gm is injective.
We next consider the adjoint action of a Lie group G on its Lie algebra g, and more
generally on the space ∧• g; we will also consider its infinitesimal version, which is
a Lie algebra action of g on itself, and more generally on ∧• g. Dually, we will also
consider the corresponding group action of G and the corresponding Lie algebra
action of g on its dual vector space g∗ and on ∧• g∗ . The formulas are most naturally
written in terms on the natural pairing · , · between ∧ p g∗ and ∧ p g, given for all
ξ1 , . . . , ξ p ∈ g∗ and for all x1 , . . . , x p ∈ g by
ξ1 [x1 ] · · · ξ1 [x p ]
. .. .
ξ1 ∧ · · · ∧ ξ p , x1 , . . . , x p := .. . (5.10)
ξ p [x1 ] · · · ξ p [x p ]
The resulting map (g, ξ ) → Ad∗g ξ is a group action of G on g∗ , called the coad-
joint action of G on g∗ , denoted Ad∗ . As in the case of the adjoint action, Ad ∗ is a
representation of G, the coadjoint representation of G on g.
These adjoint and coadjoint actions extend to group actions of G on ∧• g and
on ∧• g∗ , also denoted by Ad and Ad∗ , upon setting
ad := Te Ad : g → End(g) ,
called the adjoint representation of g. Writing adx for the image of x ∈ g by ad,
it is a fundamental fact that the endomorphism adx is given by adx (y) = [x, y],
for all y ∈ g. It follows that ad is indeed a Lie algebra representation: ad[x,y] =
adx ◦ ady − ady ◦ adx , for all x, y ∈ g. More generally, taking the tangent map of
Ad : G → End(∧• g) at e we obtain the adjoint representation of g on ∧• g, still
denoted by ad. Equation (5.11) implies that
p
adx (x1 ∧ · · · ∧ x p ) = ∑ x1 ∧ · · · ∧ [x, xi ] ∧ · · · ∧ x p ,
i=1
for all x1 , . . . , x p ∈ g hence that adx X = [[x, X]], for all X ∈ ∧• g. In view of the graded
Jacobi identity for the algebraic Schouten bracket [[· , ·]], we have for all x, y ∈ g and
for all X ∈ ∧• g that
ad[x,y] X = [[[x, y] , X]] = [[x, [[y, X]]]] − [[y, [[x, X]]]] = adx (ady X) − ady (adx X) ,
5.1 Lie Groups and (Their) Lie Algebras 121
so that ad[x,y] = adx ◦ ady − ady ◦ adx , which shows that ad is indeed a Lie algebra rep-
resentation of g (on ∧• g). Similarly, we obtain the coadjoint representation by taking
the tangent map of Ad∗ : G → End(∧• g∗ ) at e, which yields ad∗ : g → End(∧• g∗ ).
As above, ad∗ is for all x ∈ g a derivation of ∧• g∗ and ad∗x is related to adx by
∧ p (Te Lg )X = ∧ p (Te Rg )X ,
Let G be a Lie group and let g be its Lie algebra. A bilinear form · | · on g is called
Ad-invariant, if for all g ∈ G and for all x, y ∈ g,
Adg x | Adg y = x | y .
If · | · is Ad-invariant, then
for all x, y, z ∈ g; a bilinear form on g satisfying the latter condition is called ad-
invariant. When G is connected, every bilinear form which is ad-invariant is Ad-
invariant, so the two notions of invariance need not be distinguished.
We often use a non-degenerate symmetric bilinear form to identify g with its dual
vector space g∗ . Non-degeneracy of a bilinear form · | · means that the linear map
χ : g → g∗ , which sends x ∈ g to the linear form x | · : y → x | y, is an isomorphism.
Its inverse will be denoted by ψ . Explicitly, χ is given by
χ (x), y = x | y = χ (y), x ,
for all x, y ∈ g. If · | · is moreover Ad-invariant, then for every g ∈ G and for all
x, y ∈ g it follows from the definitions that
∗
Adg (χ (x)), y = χ (x), Adg−1 y = x | Adg−1 y
= Adg x | y = χ (Adg x), y ,
g
χ
g∗
ad∗dξ F ξ = 0 , (5.15)
where we used in the last step that dtd |t=0 F(Ad∗exptz ξ ) = 0, itself a direct conse-
quence of the Ad∗ -invariance of F. If F : g∗ → F is an Ad∗ -invariant function and
if the bilinear form · | · is Ad-invariant, then the commutativity of (5.14) leads for
every g ∈ G to
F ◦ χ ◦ Adg = F ◦ Ad∗g ◦χ = F ◦ χ ,
so that F ◦ χ is Ad-invariant. Similarly, · | · is Ad-invariant and F ◦ χ is Ad-
invariant, then F is Ad∗ -invariant. Thus, χ establishes a one-to-one correspondence
5.2 Poisson Reduction 123
N ⊂ M
p (5.17)
We first describe Poisson reduction in purely algebraic terms. Dualizing the dia-
gram (5.17), we consider a Poisson algebra (A , ·, {· , ·}), an ideal I of (A , ·), the
canonical map κ : A → A /I and a subalgebra B of A /I , as in the following
diagram:
124 5 Reduction
A
κ (5.18)
B ⊂ A /I
N (I ) = {F ∈ A | {I , F} ⊂ I } . (5.20)
Proof. Suppose first that conditions (1) and (2) are satisfied. We wish to define a
Poisson bracket on B by using Eq. (5.21). Let F, G ∈ B and let F̃, G̃ be representa-
tives in A of F and G; both F̃ and G̃ are determined uniquely up to an element of I .
According to condition (1), F̃ and G̃ belong to N (I ), so their bracket F̃, G̃ is,
modulo I , independent of the chosen representatives F̃ andG̃ for F and G. Con-
dition (2) implies that F̃, G̃ ∈ κ −1 (B), so that κ F̃, G̃ ∈ B. It follows that
(5.21) defines a map {· , ·}B : B × B → B. Since κ is an algebra homomorphism,
5.2 Poisson Reduction 125
where F̃, G̃ and H̃ are arbitrary representatives of F, G and H. The Jacobi identity for
{· , ·}B therefore follows from the Jacobi identity for the bracket {· , ·}. We conclude
that (A , I , B) is Poisson reducible, in particular the reduced Poisson bracket on
B is given by (5.21).
Suppose now that (A , I , B) is Poisson reducible, so that B admits a Pois-
son bracket {· , ·}B which satisfies (5.19). Since F̃ and G̃ are arbitrary
elements of
κ −1 (B), the latter formula implies that κ κ −1 (B), κ −1 (B) ⊂ B, which yields
−1
F̃ ∈ κ (B) and
condition (2). For arbitrary elements G̃ ∈ I we have from (5.19),
since κ (G̃) = 0, that κ F̃, G̃ = 0, so that κ −1 (B), I ⊂ I , which yields
condition (1).
Notice that since κ −1 (B) is a subalgebra of (A , ·) and N (I ) is a Poisson
subalgebra of (A , ·, {· , ·}), the two conditions (1) and (2) in Proposition 5.5 can be
summarized in the single condition:
κ −1 (B) is a Poisson subalgebra of N (I ) .
Moreover, viewing the canonical map κ as a map κ : κ −1 (B) → B, the for-
mula (5.21) for the reduced Poisson bracket can be simply stated as κ being a
morphism of Poisson algebras, where κ −1 (B), which is a Poisson subalgebra
of N (I ), and hence of A , is equipped with the Poisson bracket which it inherits
from A .
We next consider Poisson reduction in the setting of affine varieties. We do this
by replacing each of the algebras in (5.18) by an affine variety, keeping in mind that
under the basic correspondence between algebras of finite type and affine varieties,
surjective algebra homomorphisms are replaced by injective maps between varieties
(with the direction of the arrow being reversed) while injective algebra homomor-
phisms are replaced by dominant maps between varieties, i.e., maps whose image
is (Zariski) dense in the target variety. We are thus led to consider the following
diagram,
N ⊂ M
p (5.22)
P
where M, N and P are affine varieties and p is a dominant map.
Definition 5.6. Let N be a subvariety of an affine Poisson variety (M, {· , ·}) and let
p : N → P be a dominant map, where P is also an affine variety. We say that the
triple (M, N, P) is Poisson reducible if there exists a Poisson structure {· , ·}P on P,
such that, for every n ∈ N,
126 5 Reduction
for all F, G ∈ F (P) and for all extensions F̃, G̃ ∈ F (M) of F ◦ p and G ◦ p. The
Poisson structure {· , ·}P is called a reduced Poisson structure.
It is clear from the definitions that a triple of affine varieties (M, N, P) is Poisson
reducible if and only if the triple of algebras (A , I , B) := (F (M), IN , F (P)) is
Poisson reducible, where IN denotes the ideal of N. The geometric interpretation
of the normalizer N (IN ) of IN , defined in (5.20), is that
Proof. If (i) and (ii) are equivalent, then condition (1) in Proposition 5.7 is clearly
satisfied. For condition (2), let F, G ∈ F (P) and let F̃, G̃ ∈ F (M) be extensions of
F ◦ p and G ◦ p. According to (i) ⇒ (ii), XF̃ [IN ] ⊂ IN and XG̃ [IN ] ⊂ IN , where,
as before, we denote the ideal of all functions on M which vanish on N by IN . Using
the Jacobi identity for {· , ·} in the form of Proposition 1.4 (5), it follows that
X{F̃,G̃} [IN ] ⊂ XF̃ G̃, IN + XG̃ F̃, IN ⊂ XF̃ [IN ] + XG̃ [IN ] ⊂ IN ,
which implies, in view of (ii) ⇒ (i), that the restriction of F̃, G̃ to N is of the
form H ◦ p, for some H ∈ F (P).
The sufficient condition for Poisson reducibility, given in Corollary 5.8, is not a
necessary one. For example, let P and Q be affine Poisson varieties and consider the
triple (P × Q, P × Q, P), where P × Q is equipped with the product Poisson structure.
If Q is of positive dimension (i.e., is not a point), then conditions (i) and (ii) in the
corollary are not equivalent. However, the triple is reducible, since condition (1) in
Proposition 5.7 is trivially satisfied, while condition (2) is a consequence of the fact
that the projection map p2 : P × Q → Q is a Poisson map (see Proposition 2.2).
We now turn to the case of Poisson manifolds. The basic setup and the basic defi-
nition are very similar to the case of affine Poisson varieties (Section 5.2.1). First,
recall that, by definition, a map p : N → P between manifolds is a submersion if the
tangent map Tn p : Tn N → Tp(n) P is surjective for every n ∈ N; in particular, every
submersion is an open map (it maps open subsets of N to open subsets of P).
Definition 5.9. Let (M, {· , ·}) be a Poisson manifold, N an (immersed or embedded)
submanifold of M, P a manifold and p : N → P a surjective submersion, as in the
following diagram:
N ⊂ M
p (5.24)
P
We say that the triple (M, N, P) is Poisson reducible if there exists a Poisson struc-
ture πP = {· , ·}P on P, such that for all open subsets V ⊂ N and U ⊂ M, with
V ⊂ U ∩ N, and for all functions F, G ∈ F (p(V )), one has
at all points n of V , where F̃, G̃ ∈ F (U) are arbitrary extensions of F ◦ p|V and
G ◦ p|V . The Poisson structure πP on P is called the reduced Poisson structure.
Remark 5.10. Recall from Section 2.2.2 that an immersed submanifold comes with
its own topology and differential structure, so the open subset V of N need not be
of the form U ∩ N for some open subset U of M. However, when N is an embedded
submanifold of M, it carries the induced topology from M and the condition in the
above definition simplifies to: for every open subset U ⊂ M and for all functions
F, G ∈ F (p(U ∩ N)), one has (5.25), for all n ∈ U ∩ N, where F̃, G̃ ∈ F (U) are
arbitrary extensions of F ◦ p|U∩N and G ◦ p|U∩N .
Proof. We first point out that for an open subset V ⊂ N and a function F ∈ F (V )
the following conditions are equivalent:
(i) F is of the form H ◦ p for some function H ∈ F (p(V ));
(ii) F is constant on all the fibers of p.
The only thing which is not obvious in this equivalence is the fact that the func-
tion H, constructed in (ii) ⇒ (i), is smooth/holomorphic; it is a consequence of the
implicit function theorem, applied to the surjective submersion p.
Suppose that (M, N, P) is Poisson reducible. Let V ⊂ N and U ⊂ M be open sub-
sets, with V ⊂ U ∩ N, and suppose that F̃ ∈ F (U) is constant on the fibers of p,
when restricted to V . Then F̃|V = F ◦ p|V for some function F ∈ F (p(V )). For an
arbitrary n ∈ V and an arbitrary function G̃, defined on a neighborhood U of n
in M, and vanishing on some neighborhood of n in N, we have in view of (5.25) that
F̃, G̃ (n) = {F, G}P (p(n)) = 0, since G = 0, in a neighborhood of p(n). There-
fore, XF̃ is tangent to N at n, leading to condition (1). Condition (2) follows also
5.2 Poisson Reduction 129
from (5.25), because the left-hand side of (5.25) depends only on p(n), i.e., on the
fiber which contains n, rather than on n itself.
Suppose now that conditions (1) and (2) hold. For every w ∈ P, we define a skew-
symmetric biderivation (πP )w at w by: for all germs Fw , Gw at w,
where n is an arbitrary point of N for which p(n) = w and F̃n , G̃n are germs at n
of arbitrary functions F̃, G̃, defined on a neighborhood of n in M, and satisfying
F̃|V = F ◦ p|V and G̃|V = G ◦ p|V for some neighborhood V of n in N. The right-hand
side in (5.28) is independent of the choice of n (with p(n) = w) because, according
to (2), the value of F̃, G̃ at points n of V depends on p(n) only. Similarly, it
is independent of the chosen F̃ and G̃ because, according to (1), if G̃ is such that
G̃|V = 0 and F̃ is such that F̃|V = F ◦ p|V , then F̃, G̃ (n) = 0. Thus πP is well-
defined at every point w ∈ P and it inherits from π the biderivation property at
each point. We need to show that the bivector field πP , defined on P by w → (πP )w
is smooth/holomorphic. Let w ∈ P, let n ∈ p−1 (w) and let V be a neighborhood
of n in N, having the property that every function in F (V ) extends to a function
in F (U), where U is a neighborhood of V in M. Such neighborhoods V and U exist
because N is an immersed submanifold of M. For every pair of functions F, G ∈
F (p(V )), we then have, using the equivalence (i)⇔(ii), given at the beginning of
the proof, that
{F, G}P ◦ p = F̃, G̃ | = H̃ ◦ p ,
V
for some function H̃ ∈ F (p(V )), in particular {F, G}P ∈ F (p(V )). This shows that
πP is a smooth/holomorphic bivector field in a neighborhood of w; since w was
arbitrary, πP is a smooth/holomorphic bivector field on P. In the bracket notation, it
is given by (5.26), for arbitrary points p(n) in P and for arbitrary functions F and G,
defined in a neighborhood of p(n) in P. Notice that (5.26) implies that
for arbitrary functions F, G, H, defined in a neighborhood of p(n), and F̃, G̃, H̃ ex-
tensions as before. It follows from (5.29) that the Jacobi identity for {· , ·} implies
the Jacobi identity for {· , ·}P . Also, (5.27) is an immediate consequence of (5.26).
tion F̃, defined on a neighborhood of V in M, and such that F̃|N∩V = 0, the Hamilto-
nian vector field XF̃ is tangent to N, at every point of V .
The notion of a coisotropic submanifold generalizes the notion of a Poisson sub-
manifold, since the Hamiltonian vector fields, associated to (local) functions which
vanish on a Poisson submanifold, are not only tangent to the Poisson submanifold,
but they actually vanish on it.
The condition that a submanifold N of a Poisson manifold is coisotropic can be
stated pointwise. To do this, we introduce, for every m ∈ M, the linear map πm :
Tm∗ M → Tm M, which corresponds3 to the bivector πm ∈ ∧2 Tm M, and for n ∈ N, the
subspace Tn⊥ N of Tn∗ M, which consists of the annihilator of Tn N in Tn∗ M,
By the implicit function theorem, every element of Tn⊥ N can be realized as the
differential at n of some function F̃, defined on a neighborhood of n in M, and whose
restriction to N vanishes. It leads to the following two pointwise characterizations
of a coisotropic submanifold.
Lemma 5.13. Let N be a submanifold of a Poisson manifold (M, π ). The following
conditions on N are equivalent.
(i) N is a coisotropic submanifold;
(ii) For every n ∈ N, πn (Tn⊥ N, Tn⊥ N) = {0};
(iii) For every n ∈ N, πn (Tn⊥ N) ⊂ Tn N.
Similarly, the stronger condition that all Hamiltonian vector fields XF̃ are tangent to
the fibers pn := p−1 (p(n)) of the surjective map p : N → P can be stated as follows:
for every n ∈ N,
πn (Tn⊥ N) ⊂ Tn pn .
In the following proposition we show that for a triple (M, N, P) as above, if all the
tangent spaces Tn pn are spanned by Hamiltonian vector fields which come from
functions which vanish on N, then (M, N, P) is Poisson reducible.
3Our sign convention is that πm (dm H)[F] = πm [Hm , Fm ] = {H, F} (m), for all functions F and H,
defined on a neighborhood of m.
5.2 Poisson Reduction 131
Proof. Notice first that in view of (iii) ⇒ (i) in Lemma 5.13, N is a coisotropic
submanifold of M. To show that (M, N, P) is Poisson reducible, we verify both con-
ditions (1) and (2) in Proposition 5.11.
(1) The skew-symmetry of πn , combined with (5.30), yields that πn (Tn⊥ pn ) ⊂ Tn N,
for every n ∈ N. It implies that the Hamiltonian vector field of every function, de-
fined on a neighborhood of n in M, whose restriction to N is constant on the fibers
of p, is tangent to N, which is condition (1).
(2) Let U ⊂ M and V ⊂ N be non-empty open subsets, with V ⊂ U ∩ N. Let F̃, G̃ ∈
F (U) be two functions whose restrictions to V are constant on the fibers of p. We
show that the restriction of F̃, G̃ to V is constant on the fibers of p. Let n ∈ V
and v ∈ Tn pn be arbitrary. Then v can be extended, on a neighborhood of n in M,
to a Hamiltonian vector field, which is tangent to the fibers of p. Indeed, by (5.30),
there exists a function H̃ on a neighborhood U ⊂ U of n in M, whose restriction
to V := V ∩ U is zero, with (XH̃ )n = −πn (dn H̃) = v. Since H̃ is zero on V , its
Hamiltonian vector field XH̃ is indeed, again by (5.30), tangent to the fibers of p at
every point of V . It follows that the function XH̃ [F̃] is zero on V , because F̃|V is
constant on the fibers of p. By (5.30), XH̃ [F̃], G̃ vanishes on N (recall that G̃ is
also constant on the fibers of p). Similarly, XH̃ [G̃], F̃ (n) = 0. Using the Jacobi
identity, it follows that
v F̃, G̃ = XH̃ F̃, G̃ (n) = XH̃ [F̃], G̃ (n) + F̃, XH̃ [G̃] (n) = 0 .
where Ã, D and F are square matrices of size s, t and u respectively. Each of the
blocks Ã, B, C, . . . is a matrix-valued function on U, whose restriction to U ∩ N
is denoted by Ã|N , B|N , C|N , . . . We summarize the conditions related to Poisson
reducibility, expressed in terms of the Poisson matrix X, in Table 5.1.
The proof of each one of the lines of the table is a direct consequence of the
definitions. To start with, N ∩U is coisotropic in U ⊂ M if and only if every Hamil-
tonian vector field XG , with G a function on U which vanishes on N ∩U, is tangent
to N at every point of N ∩U. Since N ∩U is the zero locus of the local coordinates
z1 , . . . , zs , this happens if and only if the Hamiltonian vector fields Xzi are tangent
to N ∩U at every point of N ∩U for all i = 1, . . . , u, i.e. if and only if Xzi [z j ](n ) = 0
for all i, j = 1, . . . , s and n ∈ N ∩U. This is equivalent to saying that F|N = 0.
Let us explain the second line in the table. Condition (1) in Proposition 5.11
demands that the Hamiltonian vector field XG̃ is tangent to N for every function
G̃ whose restriction to N is of the form G ◦ p for some function G defined in a
5.2 Poisson Reduction 133
Table 5.1 A summary of the conditions related to Poisson reducibility, expressed in terms of the
Poisson matrix X.
C|N = F|N = 0
(M, N, P) is Poisson reducible
Ã|N depends only on x̃1 , . . . , x̃s
C|N = F|N = 0
Condition (5.30) in Proposition 5.14
Rk(E(n )) = t for every n ∈ N ∩U
Let (A , ·, {· , ·}) be a Poisson algebra and let I be an ideal of (A , ·). Consider the
composition of algebra homomorphisms
κ A
N (I ) A (5.31)
I
N (I ) := {F ∈ A | {F, I } ⊂ I } ,
which is a Poisson subalgebra of A . The kernel of the composite map in (5.31) is the
Poisson ideal N (I ) ∩ I of N (I ). It follows that N (I )/(N (I ) ∩ I ) is itself
a Poisson algebra. Notice that it is isomorphic to a subalgebra of A /I , although
the latter is not a Poisson algebra, in general. It leads to the following definition.
Definition 5.17. Let (A , ·, {· , ·}) be a Poisson algebra. An ideal I of (A , ·) is said
to be a Poisson–Dirac ideal if one of the following equivalent conditions holds:
(i) The natural inclusion of algebras N (I )/(N (I ) ∩ I ) → A /I is an iso-
morphism;
(ii) The composition of algebra homomorphisms N (I ) → A → A /I is sur-
jective;
(iii) I + N (I ) = A .
The three conditions in the definition are clearly equivalent. Condition (i) implies
that A /I is a Poisson algebra, isomorphic to N (I )/(N (I ) ∩ I ), as stated in
the following proposition.
Proposition 5.18. Let I be a Poisson–Dirac ideal of a Poisson algebra (A , ·, {· , ·}).
Then A /I has a unique Poisson bracket which makes the surjective morphism
N (I ) → A → A /I into a morphism of Poisson algebras. This bracket is said
to be obtained by Poisson–Dirac reduction.
When I is a Poisson–Dirac ideal, the different Poisson algebras which appear in
Poisson–Dirac reduction can be summarized in the following commutative diagram,
5.3 Poisson–Dirac Reduction 135
N (I ) A
κ
(5.32)
N (I )
A /I
N (I ) ∩ I
where all maps are morphisms of Poisson algebras, except (in general) for the
canonical projection κ : A → A /I .
We now consider Poisson–Dirac reduction for affine Poisson varieties. Suppose
that N ⊂ M is a subvariety of an affine Poisson variety (M, {· , ·}). We denote the
ideal of N (in F (M)) by IN . Recall from (5.23) that N (IN ) consists of all func-
tions on M whose Hamiltonian vector field is tangent to N. Translating (ii) of Defi-
nition 5.17 into geometrical terms, leads to the following definition.
Definition 5.19. Let (M, {· , ·}) be an affine Poisson variety. A subvariety N ⊂ M is
said to be a Poisson–Dirac subvariety if every function on N is the restriction of a
function on M whose Hamiltonian vector field is tangent to N.
Thus, a subvariety N is, by definition, a Poisson–Dirac subvariety if and only if its
ideal IN is a Poisson–Dirac ideal. In view of Definition 5.17 this leads to other, al-
gebraic, characterizations of Poisson–Dirac subvarieties and Proposition 5.18 shows
that a Poisson–Dirac subvariety inherits a Poisson structure from its ambient Pois-
son variety. That is the content of the following proposition.
Proposition 5.20. Let N be a Poisson–Dirac subvariety of an affine Poisson variety
(M, {· , ·}). There exists a unique Poisson structure {· , ·}N on N such that, for every
F, G ∈ F (N), their Poisson bracket is given by
{F, G}N = F̃, G̃ N (5.33)
QN M
DN N
where
Ai j = Fi , Fj , Bik = {Fi , Gk } , Dk = {Gk , G } ,
for 1 i, j s and 1 k, t. Since N is a Poisson–Dirac subvariety, the functions
F1 , . . . , Fs , restricted to N, generate F (N). Also, the Poisson matrix of (N, {· , ·}N ),
5.3 Poisson–Dirac Reduction 137
W V U
F|
V
F F̃
F
Every function F, defined on an open subset of N, admits a local extension at every
point of its domain of definition.
Definition 5.23. Let (M, {· , ·}) be a Poisson manifold and let N be an (immersed or
embedded) submanifold of M. We say that N is a Poisson–Dirac submanifold of M,
if for every n ∈ N, every function F, defined in a neighborhood of n in N, admits a
local extension (F̃,U,V ) at n, such that XF̃ is tangent to N at every point of V .
Note that, in particular, a Poisson submanifold N of a Poisson manifold (M, {· , ·}) is
a Poisson–Dirac submanifold, since all Hamiltonian vector fields are tangent to N.
We show in the following proposition that every Poisson–Dirac submanifold in-
herits a Poisson structure from its ambient Poisson manifold.
Proposition 5.24. Let (M, {· , ·}) be a Poisson manifold and let N be a Poisson–
Dirac submanifold of M. There exists on N a unique Poisson structure {· , ·}N such
that, for all open subsets V ⊂ N and U ⊂ M with V ⊂ U, for every function F̃, G̃ ∈
138 5 Reduction
F (U) whose Hamiltonian vector fields are tangent to N at every point of V , and for
every n ∈ V , we have
{F, G}N (n) = F̃, G̃ (n) , (5.34)
where F, G ∈ F (V ) are the restrictions of F̃, G̃ to V . In particular, locally, every
Hamiltonian vector field on N is the restriction to N of a Hamiltonian vector field
on M.
Proof. For every n ∈ N, we define a skew-symmetric pointwise biderivation (πN )n
by its values on germs Fn , Gn at n,
for every n ∈ V . To check that the bivector field is smooth/holomorphic, notice that
the above extensions (F̃,U,V ) and (G̃,U,V ) are local extensions of F and G at
every n ∈ V , with Hamiltonian vector fields tangent to N at every point of V , so
that (5.36) is not only valid for n, but for every n ∈ V . Since the right-hand side of
(5.36) is a smooth or holomorphic function of n, and since F and G are arbitrary
functions, we can conclude that the bivector field πN is also smooth or holomorphic
(on V , hence on N).
Let, as above, F̃, G̃, H̃ be three functions, defined on an open subset U of M,
assume that their Hamiltonian vector fields are tangent to N at every point of an open
subset V of N, and let F, G, H denote their restrictions to V . The Hamiltonian vector
field of F̃, G̃ is tangent to N, at every point of V , since X{F̃,G̃} = − [XF̃ , XG̃ ]
(item (5) in Proposition 1.4). As a consequence, F̃, G̃ ,U,V is a local extension
of {F, G}N at n, whose Hamiltonian vector field is tangent to N at every point of V ,
hence
5.3 Poisson–Dirac Reduction 139
where xi j (n ) := xi , x j (n ), for 1 i < j d, and where the first sum is the re-
duced Poisson structure at n. Equation (5.38) implies that, in well-chosen local co-
ordinates, the Poisson matrix of π at points of N takes a block-diagonal form. We
show in the following proposition that this fact characterizes Poisson–Dirac sub-
manifolds.
Proposition 5.25. Let (M, π ) be a Poisson manifold of dimension d and let N be a
submanifold of M. Then N is a Poisson–Dirac submanifold of M if and only if there
exists for each n ∈ N a coordinate chart (U, x) of M, adapted at N, and centered
at n with the following property: if the Poisson matrix X := ( xi , x j )1i, jd of π in
terms of these coordinates is written in block form
A B
X= ,
−B D
In particular, the tangent vector (XF̃ )n in Tn M belongs to Tn N if and only if it is
equal to (XF )n . Since, by Theorem 1.30, the tangent space of a symplectic leaf at
a given point n is the space of all Hamiltonian vectors (XF )n at n , an element in
Tn Sn (M) belongs to Tn N if and only if it belongs to Tn Sn (N). In short, we have,
for all n ∈ Sn (N),
Tn Sn (N) = Tn Sn (M) ∩ Tn N . (5.40)
In general geometrical terms, (5.39) and (5.40) say that we have three submanifolds
M0 , M1 and M2 of M, such that
for all m0 ∈ M0 . According to the implicit function theorem, there exist in this case
neighborhoods U0 , U1 and U2 of m0 in M0 , M1 and M2 respectively such that U0 =
U1 ∩U2 . In particular, there is a neighborhood of m0 in M1 ∩ M2 , which is contained
in M0 . Applied to (5.39) and (5.40), this means that Sn (N) is an open subset in C .
Let us show that Sn (N) is also a closed subset of C . Let n be a point in the closure
of Sn (N) in C . There is, for the above reasons, an open subset W of C contained in
Sn (N). Since W has a non-empty intersection with Sn (N), the symplectic leaves
Sn (N) and Sn (N) coincide, so that in particular n ∈ Sn (N), i.e., Sn (N) is a closed
subset of C . In conclusion, Sn (N) is, for every n ∈ N, an open and closed subset
of C , the connected component of Sn (M) ∩ N, which contains n.
We now explain how to compute the Poisson matrix of a Poisson structure obtained
by Poisson–Dirac reduction out of the Poisson matrix of the Poisson structure on the
ambient space. According to (5.34), if x̃1 , . . . , x̃s are functions whose Hamiltonian
vector fields are tangent to N, at every point in a neighborhood V of n in N, and
such that their restrictions x1 , . . . , xs to V are local coordinates for N, then the matrix
x̃i , x̃ j | is the Poisson matrix of {· , ·}N with respect to the coordinates
V 1i, js
x1 , . . . , xs . We show in the following proposition how to construct such functions and
we derive from it an explicit formula for the latter Poisson structure.
Proposition 5.27. Let (M, π ) be a Poisson manifold and let N be an s-dimensional
submanifold of M. Suppose that (U, x) is a coordinate chart of M, adapted to N
and centered at some point n ∈ N. LetV denote an open subset of N, containing n,
having the property that x1 |V , . . . , xs |V is a system of local coordinates on V . Let X
denote the Poisson matrix xi , x j 1i, jd of π with respect to these coordinates.
142 5 Reduction
where A and D are square matrices, of size s, respectively d − s, while B has size
s × (d − s). If the matrix D(m) is invertible for every m ∈ U, then
(1) V is a Poisson–Dirac submanifold of (M, π );
(2) The Poisson matrix XN of the reduced Poisson structure on V is given, at
n ∈ V , by
XN (n ) = A(n ) + B(n )D(n )−1 B(n ) .
Let (M, π ) be a Poisson manifold and let m ∈ M be a point where the rank of π is 2r.
Suppose that q1 , . . . , qr , p1 , . . . , pr , z1 , . . . , zs are splitting coordinates, defined on a
neighborhood U of m in M, and centered at m. According to Weinstein’s splitting
theorem (Theorem 1.25), this means that, on U,
r
∂ ∂ ∂ ∂
π=∑ ∧ + ∑ φk (z) ∧ ,
i=1 ∂ q i ∂ p i 1k<s ∂ zk ∂ z
where the functions φk are (smooth or holomorphic) functions, which depend on
z = (z1 , . . . , zs ) only, and which vanish when z = 0. In view of Definition 5.23, the
embedded submanifold N of M, defined by
∂ ∂
πN = ∑ φk (z) ∧
∂ zk ∂ z
,
1k<s
with the understanding that (z1 , . . . , zs ) stands for the restriction of (z1 , . . . , zs ) to N.
The submanifold N of M which appears in the latter example is transverse at m
to the symplectic leaf Sm through m, since Tm N ⊕ Tm Sm = Tm M. Recall that two
submanifolds N and O of a manifold M are said to be transverse at a point m ∈ M if
m ∈ N ∩ O and Tm N + Tm O = Tm M; when N and O have complementary dimension
in M, this is equivalent to Tm N ⊕ Tm Om = Tm M. The following theorem states that,
in a neighborhood of m in M, every submanifold of M of dimension dim N, which
is transverse to Sm at m, is in a neighborhood of m a Poisson–Dirac submanifold
of M and is isomorphic, as a Poisson manifold, to a neighborhood of m in N.
Theorem 5.28. Let (M, π ) be a Poisson manifold and let m ∈ M. The rank of π at m
is denoted by 2r. Suppose that q1 , . . . , qr , p1 , . . . , pr , z1 , . . . , zs are splitting coordi-
nates, defined on a neighborhood U of m in M, and centered at m. We denote by N0
the Poisson–Dirac submanifold of M, defined by
where π = ∑1k<s φk (z) ∂∂z ∧ ∂∂z is a Poisson structure on U, whose structure
k
functions φk depend only on the variables z = (z1 , . . . , zs ) and vanish for z = 0.
These coordinates are adapted to the symplectic leaf Sm , since Sm is defined on U
by z = 0. Consider an arbitrary s-dimensional submanifold N1 of M, transverse at m
to the symplectic leaf Sm of π , and of dimension s = dim N0 . Since N1 is of codi-
mension 2r in M, N1 is in a neighborhood of m in M given as the intersection of 2r
hypersurfaces
H1 (q, p, z) = · · · = H2r (q, p, z) = 0 . (5.42)
Since N1 is transverse to Sm at m, these hypersurfaces are transverse to Sm at m
and their tangent spaces have Tm N1 as their common intersection. By the implicit
function theorem, applied to (5.42), there exist functions F1 , . . . , Fr , G1 , . . . , Gr , such
that N1 is, in a neighborhood of m, given by
qi = Fi (z1 , . . . , zs ) , pi = Gi (z1 , . . . , zs ) ,
for i = 1, . . . , r. Letting
qi , pj (m) = δi, j , qi , qj (m) = pi , pj (m) = 0 ,
where {· , ·} := π and δi, j stands for the Kronecker delta. As a consequence, the
matrix ⎛ ⎞
r r
qi , qj (m ) qi , pj (m )
⎜ i, j=1 i, j=1 ⎟
⎝ r r ⎠
pi , q j (m )
pi , p j (m )
i, j=1 i, j=1
The proof of (2) uses some of the objects which were constructed in the first
part of the proof, namely the open neighborhood U of m in M and the functions
F1 , G1 , . . . , Fr , Gr which are defined on U . Consider the product manifold U × F,
which we equip with its product Poisson structure (the Poisson structure on F is
trivial because F is one-dimensional). The splitting coordinates which we have con-
structed on U , plus the natural coordinate on F which we denote by τ , yield coor-
dinates on U × F. We define on U × F the following 2r functions
Since
for all 1 i, j r and for all τ ∈ F, this linear system has a unique solution, which is
defined on a neighborhood U ×W of m × [0, 1] in U × F, and which is smooth. No-
tice that since all coefficients in (5.44) are independent of the coordinates pi and qi ,
146 5 Reduction
∂ Qi
V [Qi ] (m , τ ) = (m , τ ) + XH [Qi ](m , τ ) = Fi (m ) − Fi (m ) = 0 ,
∂τ
and similarly V [Pi ] (m , τ ) = 0, for all (m , τ ) ∈ N and all i = 1, . . . , r. The time 1
flow of V maps U × {0} to a neighborhood of (m, 1) in U × {1}. Since it pre-
serves N , it maps (N0 ∩U ) × {0} to a neighborhood of (m, 1) in N1 × {1}, as was
to be shown.
It follows from Theorem 5.28 that all s-dimensional submanifolds which are trans-
verse to the symplectic leaf which contains m, are isomorphic as Poisson manifolds
in a neighborhood of m, when they are equipped with their reduced Poisson struc-
ture.
We show in the following proposition that all s-dimensional submanifolds which
are transverse to a given symplectic leaf S of M are isomorphic as Poisson mani-
folds in a neighborhood of their intersection point with S , when they are equipped
with their reduced Poisson structure.
Proposition 5.29. Let (M, π ) be a Poisson manifold and let S be a symplectic leaf
of (M, π ). We denote the dimension of S by 2r and we let s := dim M − 2r. Assume
that we are given:
(1) Two points m, m which belong to S ;
(2) Two s-dimensional submanifolds N and N of M, which are transverse to S
at m and at m respectively.
Then there exist neighborhoods V and V of m and m in N and N respectively, such
that:
(1) V and V are Poisson–Dirac submanifolds of (M, π );
(2) Equipped with their reduced Poisson structure, V and V are isomorphic as
Poisson manifolds.
Proof. Let m and m be two points of M which belong to the same symplectic leaf
S of (M, π ). By definition, such points can be joined by a piecewise Hamiltonian
5.3 Poisson–Dirac Reduction 147
path. It is therefore sufficient to prove the proposition for points m and m which
can be joined by a Hamiltonian path. Assume therefore that U is an open subset
of M, that H is a function on U and that the flow Φt of XH is defined for t in a
neighborhood of [0, 1] and for all points of U, with in particular Φ1 (m) = m . Since
Φt is the flow of a Hamiltonian vector field, Φ1 is a Poisson diffeomorphism be-
tween a neighborhood U of m and a neighborhood Φ1 (U) of Φ1 (m) = m . Since
Φ1 preserves the Poisson structure, it sends the symplectic leaf S through m to the
symplectic leaf S through Φ1 (m); moreover, since Φ1 is a diffeomorphism, it sends
every submanifold, transverse to S at m, to a submanifold transverse to S at m ,
and Φ1 realizes a Poisson diffeomorphism between them, when both are equipped
with the Poisson structure which they inherit from (M, π ) as Poisson–Dirac subman-
ifolds. Since we know from item (2) of Theorem 5.28 that the transverse Poisson
structure is independent of the chosen transversal through a given point, this shows
in addition that the transverse Poisson structure is independent of the chosen point
on a given symplectic leaf of (M, π ).
Theorem 5.28 and Proposition 5.29 motivate and justify the following definition.
We have seen two particular situations in which one constructs a (natural) rep-
resentative of the transverse Poisson structure to a symplectic leaf S in a Pois-
son manifold (M, π ). The first one is Weinstein’s splitting theorem (Theorem 1.25),
which allows us to write π in terms of splitting coordinates on a neighborhood of a
point m ∈ S as
r
∂ ∂ ∂ ∂
π=∑ ∧ + ∑ φk (z) ∧ , (5.45)
i=1 ∂ qi ∂ pi 1k<s ∂ zk ∂ z
where the functions φk are (smooth or holomorphic) functions, which depend on
z = (z1 , . . . , zs ) only, and which vanish when z = 0; in this case, the second sum in
(5.45) is a Poisson structure, which is a representative for the transverse Poisson
structure to the symplectic leaf S , while the first sum in (5.45) yields the canonical
Poisson structure on the symplectic leaf S . This shows that the decomposition of
Poisson structures, which appears in Weinstein’s splitting theorem, is unique. The
second situation is when one selects a submanifold N of M, which is transverse to
the symplectic leaf S ; in this case, Theorem 5.28 says that N is in a neighborhood V
of the intersection point {m} := S ∩ N a Poisson–Dirac submanifold of (M, π ) and
the reduced Poisson structure on V is a representative for the transverse Poisson
structure to the symplectic leaf S .
148 5 Reduction
Poisson manifolds often come with symmetry groups which lead, in favorable cases,
to new Poisson manifolds. If a Lie group G acts on a Poisson manifold, preserving
the Poisson structure, then there are three natural constructions which lead to new
Poisson manifolds:
• By Poisson reduction, the quotient M/G, assumed to be a manifold, inherits a
Poisson structure;
• By Poisson–Dirac reduction, the set of fixed points of the action, assumed to be
a manifold, inherits a Poisson structure;
• By Poisson reduction, the action yields a Poisson structure on certain quotients
of the fibers of the momentum map, assuming that the latter map exists.
Similarly, such new Poisson structures arise in the case of an affine Poisson variety,
when an algebraic group is acting. As indicated in the three items above, this section
is an application of the previous sections, in the particular case of a Poisson manifold
(or affine Poisson variety) upon which a group is acting.
so that a Poisson action is in this case precisely an action preserving the Poisson
structure. In general, however, a Poisson action does not preserve the Poisson struc-
ture.
5.4 Poisson Structures and Group Actions 149
Proof. In the case of an affine Poisson variety, M/G is an affine variety. According
to Proposition 5.7, in which (1) is automatically satisfied, (M, M, M/G) is Poisson
reducible as soon as the G-invariant functions on M form a Lie subalgebra of F (M),
in formulas
F (M)G , F (M)G ⊂ F (M)G . (5.46)
where we used in the last step that p2 is a Poisson map (see Proposition 2.2). This
shows, in view of the above characterization of G-invariant functions, that {F, G} ∈
F (M)G . Hence, F (M)G is indeed a Lie subalgebra of (M, {· , ·}), which achieves
the proof for the case of affine Poisson varieties.
In the manifold case, M/G is a manifold since the action is proper and locally
free. Functions which are defined on open subsets of M are constant on the fibers
of p : M → M/G if and only if they are G-invariant. Therefore, (M, M, M/G) is,
in view of Proposition 5.11, Poisson reducible if and only if the Poisson bracket
of every pair of G-invariant functions, defined on an arbitrary open subset of M,
is G-invariant. This is checked precisely as in (5.47), taking for F and G arbitrary
functions, which are defined on an open subset of M.
quotient N/G does. This is the content of the following generalization of Proposi-
tion 5.33, which is proved in exactly the same way as a direct application of Propo-
sition 5.7 in the case of an affine Poisson variety, and of Proposition 5.11 in the case
of a Poisson manifold.
Proposition 5.34. Let χ : G × M → M be a group action and let N be a G-invariant
subset of M; denote by p : N → N/G the quotient map. Suppose that either one of
the following is satisfied:
(1) (M, {· , ·}) and (G, {· , ·}G ) are affine Poisson varieties, with G being a reduc-
tive algebraic group, and N is an affine subvariety of M;
(2) (M, {· , ·}) and (G, {· , ·}G ) are Poisson manifolds, with G being a Lie group,
N is a submanifold of M and the action of G on N is proper and locally free.
Suppose that, in addition, the following conditions hold:
(1) χ is a Poisson action;
(2) For every function F, whose restriction to N is G-invariant, the Hamiltonian
vector field XF is tangent to N; it suffices that this property holds for F ∈
F (M) in the affine variety case, but it is required to hold for F defined on
arbitrary open subsets in the manifold case.
Then N/G carries a unique Poisson structure {· , ·}N/G such that for every open
subset V of N/G and for every pair of functions F, G ∈ F (V ), and for all n ∈
p−1 (V ),
F̃, G̃ (n) = {F, G}N/G (p(n)) ,
In this section we show that, under some general assumptions, if a group action on
a Poisson variety or manifold leaves the Poisson structure invariant, then the fixed
point set of the action is a Poisson–Dirac subvariety or submanifold. We do this in
the case of a finite or linear algebraic group, acting on an affine variety; see the end
of this section for the case of Lie groups acting on manifolds.
Let G be a finite group or linear algebraic group, acting on an affine variety M.
We write both χg (m) and gm for the action of g ∈ G on m ∈ M, while the induced
action on F ∈ F (M) is denoted by ψg (F), as in (5.4). We denote the algebra of
invariant functions by F (M)G ,
M G := {m ∈ M | gm = m for all g ∈ G} .
Δ := {(m, m) | m ∈ M} .
Proof. Let F ∈ F (M)G be a G-invariant function. For all G ∈ F (M) and all g ∈ G,
we have
where we used in the last step that every map ψg is a morphism of Poisson algebras.
Since I (M G ) is generated by all functions of the form G − ψg (G), with g ∈ G and
G ∈ F (M), the inclusion F (M)G , I (M G ) ⊂ I (M G ) follows, i.e., F (M)G is
contained in the normalizer of I (M G ). In view of (iii) in Definition 5.17, I (M G )
will be a Poisson–Dirac ideal of F (M) if F (M)G + I (M G ) = F (M).
We consider two particular cases of the above proposition, which both lead to the
result that the fixed point set of the group action is a Poisson–Dirac subvariety. We
first consider the simpler case in which G is a finite group.
Proposition 5.36. Let G be a finite group, acting on an affine Poisson variety
(M, {· , ·}). If the action preserves the Poisson structure, then the fixed point set
M G is a (not necessarily irreducible) Poisson–Dirac subvariety of (M, {· , ·}). For
F, G ∈ F (M G ), their reduced bracket is given, at n ∈ M G , by
152 5 Reduction
1 1
{F, G}M G (n) =
|G|2 g ∑ ψg1 (F̃), ψg2 (G̃) (n) = ∑ ψg (F̃), G̃ (n)
|G| g∈G
1 ,g2 ∈G
Proof. By averaging the action over the group, one constructs a linear projection
ρ : F (M) → F (M)G ; explicitly, ρ (F) is given, for F ∈ F (M), by
1
ρ (F) := ∑ ψg (F) ,
|G| g∈G
(5.50)
1
F − ρ (F) = ∑ (F − ψg (F)) ,
|G| g∈G
1
{F, G}M G (n) = ρ (F̃), ρ (G̃) (n) =
|G|2 g ∑ ψg1 (F̃), ψg2 (G̃) (n) ,
1 ,g2 ∈G
for all n ∈ M G . Since G̃ − ρ (G̃) ∈ I (M G ), the latter line can also be written as
1
{F, G}M G (n) = ρ (F̃), G̃ (n) = ∑ ψg (F̃), G̃ (n) ,
|G| g∈G
for all n ∈ M G .
As a general principle in invariant theory, results which hold for the action of a
finite group, acting on an affine variety, can be extended to the action of a reductive
group, acting on an affine variety, by replacing the above averaging operator with
the Reynolds operator (see [190]). It leads to the following result.
Proposition 5.37. Let G be a reductive group, acting on an affine Poisson variety
(M, π ). If the action preserves the Poisson structure, then the fixed point set M G is
a (not necessarily irreducible) Poisson–Dirac subvariety of (M, π ).
For a differential geometric analog of the proposition, see [54, Th. 2.1].
5.4 Poisson Structures and Group Actions 153
Let G be a Lie group, with Lie algebra g. Suppose that G is acting on a Poisson
manifold (M, π ). It is natural to demand that the fundamental vector fields x, with
x ∈ g, are Hamiltonian. It leads to the following definition.
Definition 5.38. Let G be a Lie group acting on a Poisson manifold (M, π ). The
action is said to be a Hamiltonian action if there exists a Lie algebra homomorphism
μ̃ : g → F (M)
(5.51)
x→ μ̃x
g X1 (M)
−X (5.52)
μ̃
F (M)
where the horizontal arrow is the map x → x, and where we recall that −X is a
morphism of Lie algebras (see (5) of Proposition 1.4 and (1.7)). We call μ̃ a co-
momentum map of the action. The dual map,
μ : M → g∗
(5.53)
m → μ (m) ,
where μ (m) ∈ g∗ is the linear map x → μ̃x (m), is said to be a momentum map of the
action.
We show in the following proposition that, under some assumptions, the orbit space
μ −1 (0)/G inherits a Poisson structure from M.
Proposition 5.39. Let G be a connected Lie group and let (M, π ) be a Poisson man-
ifold. We assume that G acts locally freely and properly on M and that the action is
Hamiltonian. Let μ denote a momentum map of the action. If 0 is a regular value of
μ , then
(1) μ −1 (0) is an embedded submanifold of M, invariant under the action of G;
(2) μ −1 (0)/G is a manifold;
(3) The triple (M, μ −1 (0), μ −1 (0)/G) is Poisson reducible.
The different spaces involved in the theorem are represented in the following dia-
gram:
154 5 Reduction
μ −1 (0) M
p (5.54)
μ −1 (0)/G
N= m ∈ M | μ̃y (m) = 0 ,
y∈g
and this vector space is invariant for all vector fields x, since,
for every n ∈ N, where we recall that pn is the fiber of p which passes through n.
Since 0 ∈ g∗ is a regular value for the momentum map μ , the space Tn⊥ N is, for
every n ∈ N, given by
Tn⊥ N = {dn μ̃x | x ∈ g} .
Since πn (dn μ̃x ) = − Xμ̃x n = xn , this means that
which proves (5.55), and hence that (M, μ −1 (0), μ −1 (0)/G) is Poisson reducible.
Remark 5.41. Also, one can generalize Proposition 5.39 and obtain a Poisson struc-
ture on μ −1 (ξ )/Gξ for every regular value ξ ∈ g∗ of μ , where Gξ is the sub-
group of G, defined in terms of the adjoint action Ad : G × g → g, by Gξ :=
g ∈ G | ξ ◦ Adg = ξ . See [144] for this and for other generalizations.
5.5 Exercises
1. Let G be a Lie group, with Lie algebra g. For r ∈ ∧2 g, show that the following
are equivalent:
(i) →
−r is a Poisson structure on G;
(ii) ← −
r is a Poisson structure on G;
(iii) [[r, r]] = 0.
Show that, when these equivalent conditions are satisfied, the Poisson structures −
→
r
←
−
and r are compatible. For r of the form r = x ∧ y, where x, y ∈ g, show that these
conditions are satisfied if and only if x and y span a Lie subalgebra of g.
2. Let M be a manifold on which a Lie group G acts. For x ∈ g, the Lie algebra
of G, denote by x the fundamental vector field, associated to x. Let r = ∑ki=1 xi ∧ yi ∈
∧2 g, satisfying [[r, r]] = 0 and consider the bivector field r := ∑ki=1 xi ∧ yi on M.
a. Show that r is a Poisson structure on M;
b. Prove that the action of (G, −
→
r −← −
r ) on (M, r) is a Poisson action;
c. Suppose that M is a vector space and the action of G on M is a representation.
Show that r is a quadratic Poisson structure (see Section 8.2).
3. Let M1 and M2 be Poisson manifolds and endow M1 ×M2 with the product Pois-
son bracket. Show that for every m1 ∈ M1 (respectively m2 ∈ M2 ), the submanifold
{m1 } × M2 (respectively M1 × {m2 }) is a Poisson–Dirac submanifold of M1 × M2 .
4. Let M be a Poisson manifold, and let N be a Poisson–Dirac submanifold of M.
Suppose that V is a Poisson vector field on M, which is tangent to N at every point
of N. Show that the restriction of V to N is a Poisson vector field.
5. Let P and M be Poisson manifolds and let φ : P → M be a submersive Poisson
map. Let N be a submanifold of M.
a. Show that if N is a coisotropic submanifold of M, then φ −1 (N) is a coisotropic
submanifold of P;
b. Show, by giving a counterexample, that if N is a Poisson–Dirac submanifold
of M, then φ −1 (N) need not be a Poisson–Dirac submanifold of P.
156 5 Reduction
6. We assume in this exercise that the reader is familiar with the notion of sym-
plectic manifold (see Section 6.3). Let (M, ω ) be a symplectic manifold, and denote
by π the canonical Poisson structure, associated to ω . Let N be a submanifold of M.
Show that the following are equivalent:
(i) N is a Poisson–Dirac submanifold of (M, π );
(ii) The restriction of ω to N is non-degenerate.
Conclude that for symplectic manifolds the notions of symplectic submanifold and
Poisson–Dirac submanifold coincide.
7. We assume in this exercise that the reader is familiar with the Lie–Poisson struc-
ture on the dual of a Lie algebra (see Chapter 7). Let g be a finite-dimensional
Lie algebra and let g = gξ ⊕ m be a vector space decomposition of g, where
gξ := {x ∈ g | ad∗x ξ = 0} denotes the centralizer of an element ξ ∈ g∗ . We consider
the affine subspace N := ξ + m⊥ of g∗ , which is isomorphic to g∗ξ , via the affine
map χ , which is defined, for all η ∈ N, by χ (η ) := (η − ξ )|g .
ξ
F ν G := FG + ν {F, G}
for all F, G ∈ F (M) (see Exercise 8 of Chapter 3, where this product is proven to
be associative). Suppose that
defines the structure of a left module over the F[ν ]-algebra (F (M)[ν ]/ν 2 , ν ) on
F (N)[ν ]/ν 2 , satisfying Φ (F, H) = F|N H (mod ν ) for every F ∈ F (M) and for
every H ∈ F (N). Let σ : F (M) × F (N) → F (N) be the bilinear map, given for
all F ∈ F (M) and H ∈ F (N) by
Φ (F, H) − F|N H
σ (F, H) = .
ν
ν =0
5.6 Notes
Lie groups and Lie algebras are standard tools in geometry and in physics; in turn,
the study of their structure and their representations is strongly inspired by ideas
which come from geometry and physics. A quick introduction to Lie groups is given
in Warner [198]; see Bump [29] for a more extensive treatment. For the theory of
Lie algebras, we refer to Humphreys [99]. A great introduction to both Lie groups
and Lie algebras and their representation theory is given in Fulton–Harris [80].
The starting point of reduction theory comes from the fact that integrals of motion
permit to eliminate degrees of freedom in Hamiltonian systems. One may cite here
Noether’s theorem, which is the predecessor of the notion of momentum map and
the theory of Dirac constraints, which eventually led to the notion of Poisson–Dirac
reduction. The modern, geometrical development of these ideas first appears in the
seminal papers by Śniatycki–Tulczyjew [184], Meyer [148] and Marsden–Weinstein
[146]. See Kosmann–Schwarzbach [114], Weinstein [199] and the introduction of
Ortega–Ratiu [160] for more details on the history of the subject; in the latter in-
troduction one also finds a detailed list of topics in the theory of moment maps and
reduction.
In our presentation of reduction, we have separated the Poisson reduction of
coisotropic submanifolds from the theory of reduction of Poisson–Dirac subman-
ifolds, which are both particular cases of the so-called Marsden–Ratiu reduction
procedure [144]. The notion of momentum map, presented here, has also been gen-
eralized in many different ways, see for example Alekseev–Malkin–Meinrenken [9]
for Lie group valued momentum mappings and Lu [133] for Poisson–Lie group val-
ued momentum mappings. These generalizations and their associated reduction can
be nicely understood with the help of Lie groupoids, see Xu [208].
We have not discussed the geometry of the image of a momentum map (which is,
under some assumptions, a convex polytope). See Guillemin [91] and the references
therein.
Part II
Examples
Chapter 6
Constant Poisson Structures, Regular
and Symplectic Manifolds
has the property that the Poisson bracket {F, G} of two linear functions F and G is a
constant function on R2r . This property leads to the following definition, which will
cover a first class of basic examples of Poisson structures.
Definition 6.1. A Poisson structure π on a finite-dimensional vector space V is
called a constant Poisson structure, if for each pair of linear functions F and G
on V , their Poisson bracket π [F, G], also denoted by {F, G}, is a constant function
on V .
Let X = (xi j ) ∈ Matd (F) be an arbitrary skew-symmetric matrix, where d ∈ N∗ . We
show that X defines a Poisson structure on Fd , i.e., a Poisson bracket on F (Fd ),
where we recall that, for a finite-dimensional F-vector space V , the notation F (V )
stands for the algebra of polynomial functions on V , in general, but may also be
chosen as the algebra of smooth functions on V , when F = R, or as the algebra
of holomorphic functions on V , when F = C. In order to construct this Poisson
structure, consider the skew-symmetric biderivation of F (Fd ), defined for all F, G ∈
F (Fd ), by
d
∂F ∂G
{F, G} := ∑ xi j , (6.3)
i, j=1 ∂ xi ∂ x j
where 2r is the rank of X and hence of the Poisson bracket π (at every point). It
means that, in terms of new linear coordinates on V , defined in terms of x1 , . . . , xd
by
(q1 , . . . , qr , p1 , . . . , pr , z1 , . . . , zd−2r ) := (x1 , . . . , xd )A
one obtains the following brackets
qi , p j = δi, j , qi , q j = pi , p j = {qi , zk } = {pi , zk } = {zk , z } = 0 ,
∂H ∂H
q̇i = , ṗi = − , żk = 0 ,
∂ pi ∂ qi
where q̇i is a commonly used shorthand for XH [qi ] = {qi , H}, and similarly for the
other variables.
In order to obtain a more intrinsic formulation of the notion of a constant Pois-
son structure on a finite-dimensional vector space V , we consider the vector space
∧2V = V ∧V , whose elements are called bivectors of V . A bivector b of V is often
viewed as a linear map, or as a bilinear form, on V ∗ . This is done as follows: writing
b = ∑si=1 vi ∧ wi , where vi , wi ∈ V , we define1 the linear map b : V ∗ → V , by
1 In terms of internal products, as recalled in Appendix A, b (ξ ) = ıξ b, which shows that b , as
given by (6.5), is well-defined.
164 6 Constant Poisson Structures, Regular and Symplectic Manifolds
s
b (ξ ) = ∑ (ξ , vi wi − ξ , wi vi ) , (6.5)
i=1
The square brackets in this formula have the following meaning: for v ∈ V , v[·] is the
unique derivation of F (V ) such that v[ξ ] = ξ , v for all linear functions ξ on V .
Geometrically, we think of v[·] as the directional derivative, defined by a vector v.
Equation (6.9) is a coordinate-free analog of (6.4).
The fact that every bivector on V defines a Poisson structure on V , motivates the
following definition.
6.2 Regular Poisson Manifolds 165
By the above, we have the following abstract version and amplification of Proposi-
tion 6.2.
for all F, G ∈ F (V ).
Comparing the definitions, it is clear that every constant Poisson structure on a con-
nected manifold is regular: taking around an arbitrary
point m ∈ M a coordinate
chart (U, x) such that all Poisson brackets xi , x j are constant on U, we see that
the rank of π is constant on U, since it is given by the rank ofthe Poisson matrix
of π with respect to x, whose entries are the Poisson brackets xi , x j ; since M is
connected, it follows that the rank of π takes the same value at every point of M,
i.e., π is a regular Poisson structure. It is a fundamental fact that the converse is also
true, leading to the following proposition.
Proposition 6.8. Let (M, π ) be a Poisson manifold, which is assumed to be con-
nected. Then (M, π ) is a regular Poisson manifold if and only if π is a constant
Poisson structure.
166 6 Constant Poisson Structures, Regular and Symplectic Manifolds
Proof. Let (M, π ) be a regular Poisson manifold and let m ∈ M. We need to show
thatthere exists a coordinate chart (U, x), centered at m, such that all Poisson brack-
ets xi , x j are constant (functions) on U. Since, by assumption, the rank of π is
constant at m, this is an immediate consequence of the Darboux theorem, which
we proved in Section 1.3.3 (Theorem 1.26). According to this theorem, if the rank
of a Poisson structure is constant (say, equal to 2r) in the neighborhood of a point
m ∈ M, then there exists in a neighborhood of m a coordinate chart (U, x), with
x = (q1 , . . . , qr , p1 , . . . , pr , z1 , . . . , zd−2r ), centered at m, in which the Poisson struc-
ture takes the canonical form
qi , q j = pi , p j = {qi , zk } = {pi , zk } = {zk , z } = 0 , qi , p j = δi, j ,
The study of regular Poisson manifolds is in general much simpler than the study of
arbitrary Poisson manifolds. Being locally the product of a symplectic manifold and
Rs or Cs with the zero Poisson structure, their (local) study can often be reduced
to the case of symplectic manifolds. But sometimes, it is the regularity itself which
can be exploited, as we show in the following example.
Example 6.9. According to Proposition 1.21, the rank of a Poisson manifold (M, π )
at a point m is equal to the dimension of the vector space of Hamiltonian vector
fields at that point, Rkm M = dim Hamm (M). Therefore, if the rank is constant,
the Hamiltonian vector fields define a distribution of rank Rk M. It is differen-
tiable, because it is defined by smooth vector fields. According to Proposition 1.4,
[XF , XG ] = −X{F,G} , so the distribution is integrable in the sense of Frobenius, so
through every point there passes a unique integral manifold of the distribution, by
Frobenius’ theorem. By definition, all Hamiltonian vector fields of (M, π ) are tan-
gent to this integral manifold, at each of its points. By Proposition 2.10, this integral
manifold carries a (unique) Poisson structure of maximal rank. In this way we have
recovered, by simple means, the symplectic foliation2 of a Poisson manifold, in the
regular case.
The most important class of regular Poisson manifolds consists of the symplectic
manifolds (real or complex). We first consider the particular case of a symplectic
vector space.
2 In this case it is indeed a foliation, and not a singular foliation.
6.3 Symplectic Manifolds 167
where we used in the last line that (XF )m = −π (dm F) (see (6.8)). The latter defi-
nition of (XF )m is, in view of the definition of π , equivalent to saying that (XF )m
is the unique element of TmV V such that
ω ((XF )m , ·) = dm F . (6.11)
ω (XF , ·) = dF .
v ∈ V.
168 6 Constant Poisson Structures, Regular and Symplectic Manifolds
vector space (V, ω ), where ω ∈ Hom(V ∧ V, F). We leave this as an exercise to the
reader; a globalization of this statement will be worked out in the next section.
Proof. Let us first assume that ω is closed. Notice that one important consequence
of this assumption is that ω is conserved by the flow of every Hamiltonian vector
field. To see this, use Cartan’s formula (3.50) for the Lie derivative LV ,
LV = ıV d + dıV (6.16)
which is a special case of (2) in Proposition 3.11, simplifies for Hamiltonian vector
fields to
which is precisely the Jacobi identity, as follows from (6.15). Thus, we have shown
that if ω is closed, so that (M, π ) is a symplectic manifold, then π is a Poisson
structure on M.
Assume now that ω is an almost symplectic structure, and that the bivector
field π , defined by (6.13), satisfies the Jacobi identity. Using the explicit for-
mula (3.28) for computing the de Rham differential, we obtain, for a triple of Hamil-
tonian vector fields as above,
as follows from (6.14), (6.15) and item (5) in Proposition 1.4. Since ωm is non-
degenerate for every m ∈ M, we have that Tm M = Hamm (π ) for every m ∈ M; the
above computation therefore shows that dω vanishes on every triple of vector fields,
i.e., that ω is closed, hence that (M, ω ) is a symplectic manifold.
Proposition 6.11. Suppose that (M, π ) is a regular Poisson manifold, whose rank is
equal to the dimension of M. There exists on M a (unique) symplectic structure ω ,
whose canonical Poisson structure is π .
In particular, for local functions F and G on M, one has π (FdG) = −FXG . Simi-
larly, one uses the inverse of πm to associate to each vector field on M a differential
one-form on M; in the compact notation, the differential one-form associated to
−1
V ∈ X1 (M) is π (V ). We use πm to define a non-degenerate differential two-
−1
form ω on M: for vector fields V1 and V2 , let ω (V1 , V2 ) := − π (V1 ), V2 .
Then
−1
ω (XF , XG ) := − π (XF ), XG = dF, XG = {F, G} ,
are isomorphic,
k π
HdR (M) Hπk (M) Hd−k (M) .
Proof. Since every symplectic vector space is even-dimensional, every tangent
space of a symplectic manifold is even-dimensional, hence also the symplectic man-
ifold itself. In order to prove (2), we assume that M is a real manifold (complex
manifolds are always orientable). Since ω is non-degenerate (at every point) the
differential d-form Λ := ω d/2 is nowhere vanishing, hence it is a volume form and
M is orientable. The volume form Λ is called the Liouville volume form of (M, ω ).
It is preserved by the flow of every Hamiltonian vector field XF , since
d/2
LXF Λ = LXF ω d/2 = ∑ ω i−1 ∧ LXF ω ∧ ω d/2−i = 0 ,
i=1
where we have used in the last step that LXF ω = 0 (see the proof of Proposi-
tion 6.10). Thus, (M, π ) is unimodular; see Section 4.4.4, where it is shown that
for a unimodular Poisson manifold, the Poisson cohomology spaces Hπk (M) and the
Poisson homology spaces Hd−k π (M) (with 0 k d) are isomorphic vector spaces,
that at every point m, the inverse of πm is just ωm ; for a local section V of T M
−1
(i.e., vector field), the inverse of π is given by π (V ) = ω (V , ·). The natural
•
extension of π to Ω (M) yields, for every k with 0 k d, a map
d
Ω k (M) Ω k+1 (M)
∧k π ∧k+1 π
−δπk
Xk (M) Xk+1 (M)
where we have used in the last step that XG1 , . . . , XGk are Poisson 1-cocycles. The
commutativity of the diagram implies that the Poisson complex (X• (M), [· , ·]S ) and
the de Rham complex (Ω • (M), d) are isomorphic. In particular, they have the same
cohomology.
The resulting inequality RkΨ (m) π2 Rkm π1 , is in sharp contrast with the inequality
RkΨ (m) π2 Rkm π1 , which is valid for a Poisson map Ψ (see Exercise 4 of Chap-
ter 1). Thus, a symplectic map Ψ : M1 → M2 between two symplectic manifolds
(M1 , ω1 ) and (M2 , ω2 ) is not necessarily a Poisson map, as a map Ψ : (M1 , π1 ) →
(M2 , π2 ), where πi is the canonical Poisson structure, associated to ωi , for i = 1, 2. It
follows that the map, which associates to a symplectic manifold the corresponding
Poisson manifold, is not functorial.
172 6 Constant Poisson Structures, Regular and Symplectic Manifolds
where we have used in the last step that w is tangent to N. On the other hand, using
(5.27) it follows that
6.3 Symplectic Manifolds 173
Comparing (6.19) and (6.20) we see that (ı∗ ω )n = (p∗ ωP )n for every n ∈ N, which
we needed to show.
Remark 6.15. The fact that the reduced Poisson structure on P is of maximal rank
(symplectic) can also be read off from its Poisson matrix. Recall from Remark 5.16
that in well-chosen coordinates on an open subset U of M, centered at a point n ∈ N,
the Poisson matrix of (M, π ) is given by
⎛ ⎞
à B C
X = ⎝ −B D E ⎠
−C −E F
and that the condition (6.18) implies that C|N = F|N = 0 (see the last line of Ta-
ble 5.1). Recall also that the Poisson matrix of the reduced Poisson structure πP is
given by the matrix A, defined by Ã|N = A ◦ p. Notice that, in the present case, E is
a square matrix, because the dimension t of the fibers of p : N → P is equal to the
codimension of N in M, a consequence of condition (6.18), combined with the fact
that π is of maximal rank. For every n ∈ N ∩U,
det X(n ) = (det E(n ))2 det Ã(n ) = (det E(n ))2 det A(p(n )) ,
so that det A(p(n )) = 0. This gives an alternative proof of the fact that A, and
hence πP , is of maximal rank at every point of P. Notice that the argument also
reproves (in the symplectic case) that Rk(E(n )) = t for every n ∈ N ∩U (see again
the last line of Table 5.1).
Combining the above proposition with Proposition 5.39 leads to the following result.
Corollary 6.16. Let G be a connected Lie group and let (M, ω ) be a symplectic
manifold. We assume that G acts locally freely and properly on M and that the
action is Hamiltonian. If 0 is a regular value of its momentum map μ , then
(1) μ −1 (0) is an embedded submanifold of M, invariant under the action of G;
(2) μ −1 (0)/G is a manifold;
(3) The triple (M, μ −1 (0), μ −1 (0)/G) is Poisson reducible.
The reduced Poisson structure on μ −1 (0)/G is regular and of maximal rank, i.e.,
it is the canonical Poisson structure, associated to a symplectic structure ω0 on
μ −1 (0)/G. Also, ı∗ ω = p∗ ω0 , where p : μ −1 (0) → μ −1 (0)/G is the quotient map.
174 6 Constant Poisson Structures, Regular and Symplectic Manifolds
Let (V, ω ) be a symplectic C-vector space of dimension 2d. We denote by SP(V ) the
group of all linear symplectomorphisms of V , i.e., the group of all endomorphisms
of V which preserve ω . Recall from Section 6.3.1 that the symplectic two-form ω
leads to a (constant) Poisson structure of rank 2d on V , which we called the canon-
ical Poisson of (V, ω ). Let G be a finite subgroup of SP(V ). The purpose of the
present section is to study the singular affine variety V /G, both as an affine variety
and as a Poisson variety. The case where the dimension of V is equal to 2 will be
studied in more detail in Section 9.2.4.
Let us specialize to the present setting what has been said at the very end of Sec-
tion 5.1.2 about the quotient of an affine variety by a finite group. Since G is a finite
subgroup of SP(V ), it acts on V , and the quotient space V /G is an affine variety,
whose algebra of regular functions F (V /G) is naturally identified with the algebra
F (V )G of all G-invariant polynomials on V . The canonical projection p : V → V /G
corresponds, dually, to the inclusion of algebras F (V /G) F (V )G → F (V ).
Moreover, since every linear symplectomorphism of (V, ω ) preserves the canoni-
cal Poisson structure of (V, ω ), the finite subgroup G ⊂ SP(V ) acts by Poisson iso-
morphisms. As a consequence, according to the first item of Proposition 5.33, V /G
carries a unique Poisson structure with respect to which p : V → V /G is a Poisson
map. This justifies the following definition.
Definition 6.17. Let (V, ω ) be a symplectic C-vector space and let G be a finite
subgroup of SP(V ). The unique Poisson structure {· , ·}V /G on V /G with respect
to which the natural projection map p : V → V /G is a Poisson map, is called the
quotient Poisson structure associated to G. The pair (V /G, {· , ·}V /G ) is called the
quotient Poisson variety associated to G.
Since the quotient space V /G is, in general, singular, it makes no sense to ask
whether it is a symplectic manifold. However, the second item of the next proposi-
tion implies that V /G is as symplectic as it can be, i.e., {· , ·}V /G is of maximal rank
at every non-singular point of V /G.
Proposition 6.18. Let (V, ω ) be a symplectic C-vector space of dimension 2d and
let G = {e} be a finite subgroup of SP(V ), the group of all symplectomorphisms
of V . Denote by p the natural projection map p : V → V /G and by U the (non-
empty) Zariski open subset of all x ∈ V for which the stabilizer Gx is trivial.
(1) A point in V /G is a non-singular point if and only if it belongs to p(U);
(2) The rank of {· , ·}V /G is equal to 2d at every non-singular point of V /G.
As a consequence, the restriction of {· , ·}V /G to p(U) is a holomorphic Poisson
structure, which is the canonical Poisson structure associated to a symplectic holo-
morphic two-form on p(U).
Proof. It is a classical result (valid for the action of an arbitrary finite group on a
vector space) that, for every x ∈ V , the point p(x) is a non-singular point of V /G
6.3 Symplectic Manifolds 175
if and only if its stabilizer Gx is a complex reflection group, i.e., a group generated
by elements for which the space of fixed points is a hyperplane (see [42]). But
the set of fixed points of a linear symplectomorphism cannot be a hyperplane. As
a consequence, in the present case, the point p(x) is non-singular if and only if
Gx = {e}. This proves (1).
The open subset U ⊂ V is a G-invariant subset, the action of G on U is free
since the stabilizer of every point of U is by definition trivial and, according to Sec-
tion 2.3.2, U is equipped with a holomorphic Poisson structure. This implies on the
one hand that the projection map p : U → U/G = p(U) is a local biholomorphism
(when both sets are considered as holomorphic manifolds), and on the other hand,
since p is moreover a Poisson map between Poisson manifolds, that p preserves the
rank of the Poisson structure at each point, so that the rank of {· , ·}V /G is 2d at every
point of p(U). Therefore, {· , ·}V /G restricts to a regular holomorphic Poisson struc-
ture of maximal rank on p(U). According to Proposition 6.11, it is the canonical
Poisson structure, associated to a symplectic structure on the holomorphic manifold
p(U). This completes the proof of (2).
Remark 6.19. For every λ ∈ C∗ , the map x → λ −1 x leads to a regular map of V /G
to itself. Since x → λ −1 x multiplies the canonical Poisson structure on V by λ 2 , this
induced automorphism also transforms the quotient Poisson structure {· , ·}V /G into
λ 2 {· , ·}V /G . As a consequence, for every non-zero μ ∈ C∗ , the Poisson varieties
(V /G, {· , ·}V /G ) and (V /G, μ {· , ·}V /G ) are isomorphic as Poisson varieties.
The case where the dimension of V is 2 will be studied in more detail in Sec-
tion 9.2.4. We give here an immediate corollary of Proposition 6.18, which will
be useful then.
Corollary 6.20. Let V be a symplectic C-vector space of dimension 2 and let G =
{e} be a finite subgroup of SP(V ). Then the image of the origin o of V through the
canonical projection p : V → V /G is the only point of V /G which is singular. It is
also the only point where the rank of the quotient Poisson structure is zero.
Proof. For a vector space V of dimension 2, the group SP(V ) coincides with the
group SL(V ) of all linear maps of determinant +1. Every element of a finite sub-
group G of SL(V ) is of finite order, hence is diagonalizable. Clearly, a diagonal-
izable element of SL(V ), different from the unit, admits o ∈ V as its unique fixed
point, hence U = V \ {o}. Clearly, the rank of {· , ·}V /G at p(o) is zero. According
to Proposition 6.18, the rank of {· , ·}V /G is 2 at every other point of V /G.
It is shown in Section 1.3.4 that every Poisson manifold admits a (singular) foliation
whose leaves carry a symplectic structure. Thus, the number of examples of sym-
plectic manifolds is abundant. There are however two large classes of symplectic
176 6 Constant Poisson Structures, Regular and Symplectic Manifolds
manifolds which need special attention: cotangent bundles and Kähler manifolds.
The first class, considered in this section, is of fundamental importance in classical
mechanics, see e.g. [1, 15] and [125].
The idea that cotangent bundles are symplectic manifolds comes from classical
mechanics: local coordinates on M and the corresponding momenta, which are lin-
ear coordinates on the fibers of ρ : T ∗ M → M yield canonical coordinates on phase
space. To make this precise, let us first show how to build natural local coordi-
nates on T ∗ M from local coordinates on M. Suppose that x1 , . . . , xd are coordinates
on U ⊂ M; it yields half of a system of coordinates on T ∗U ⊂ T ∗ M by letting
qi := xi ◦ ρ . For the other half, let pi denote the function on T ∗U, which is defined
by
∂
pi (ξm ) := ξm , , (6.21)
∂ xi m
where ξm ∈ T ∗U is in the fiber over m, i.e., ρ (ξm ) = m. The functions pi and xi are
dual in the following sense: for m ∈ U and 1 i, j d,
∂
pi (dm x j ) = dm x j , = δi, j .
∂ xi m
for every ξm ∈ Tm∗ M. The reader will easily verify that the canonical Poisson bracket
of ω satisfies the following formulas: if F1 , F2 ∈ F (M) and V1 , V2 ∈ X1 (M), then
F̃1 , F̃2 = 0,
V˜1 , V˜2 = [V1 , V2 ] and
V˜1 , F̃1 = V1 [F1 ] .
In fact, these formulas can be used to define ω (see [3, Ch. 2]).
We now turn to a second class of examples, which is important from the point of
view of complex geometry. Let M be a complex manifold and let us denote the
underlying real manifold by MR . Suppose that ds2 is a Hermitian metric on M. This
means that, in terms of a system of local (holomorphic) coordinates z1 , . . . , zd ,
ds2 = ∑ hi j dzi ⊗ dz j ,
1i, jd
where (hi j ) is a positive definite Hermitian matrix. The imaginary part (up to a sign)
of such a metric is a non-degenerate differential 2-form on MR , given by
√
−1
ω :=
2 1i,∑ hi j dzi ∧ dz j .
jd
One says that the metric ds2 is a Kähler metric if ω is closed, i.e., if ω is a symplectic
structure. Then (MR , ω ) is called a Kähler manifold and ω is called the associated
differential 2-form of the metric. Notice that the metric can be recovered from its
associated differential 2-form.
Two facts underlie the importance of Kähler manifolds. The first one is that the
complex projective space PN admits a Kähler metric. The second fact is that every
178 6 Constant Poisson Structures, Regular and Symplectic Manifolds
submanifold of a Kähler manifold is itself a Kähler manifold. This means that every
non-singular complex projective variety has a Kähler structure.
6.4 Notes
Together with symplectic manifolds, considered in the previous chapter, Lie alge-
bras provide the first examples of Poisson manifolds. Namely, the dual g∗ of a finite-
dimensional Lie algebra g admits a natural Poisson structure, called its Lie–Poisson
structure, which provides new insights and technical tools in the study of Lie al-
gebras. For example, the coadjoint orbits in g∗ are precisely the symplectic leaves
of the Lie–Poisson structure, showing in particular that coadjoint orbits are even-
dimensional.
Lie–Poisson structures are linear Poisson structures, in the sense that they are
Poisson structures on a vector space V , for which the Poisson bracket of every pair
of linear functions on V (elements of V ∗ ) is a linear function on V . Also, every linear
Poisson structure (on a finite-dimensional vector space) is a Lie–Poisson structure
and we have a functorial equivalence between linear Poisson structures and Lie
algebra structures.
In many cases, one considers the Poisson structure on the Lie algebra g itself,
rather than on g∗ . This is usually done by identifying g with g∗ , using a non-
degenerate Ad-invariant symmetric bilinear form on g, which exists for example
for every semisimple Lie algebra. The coadjoint orbits are then identified with the
adjoint orbits and the Hamiltonian vector fields on g take a natural form, a so-called
Lax form.
The Lie–Poisson structure on g∗ is introduced in Section 7.1, while the induced
Poisson structure on g is discussed in Section 7.2. The main properties of the Lie–
Poisson structure are given in Section 7.3. A variant of Lie–Poisson structures,
namely affine (= linear + constant) Poisson structures and their Lie theoretical
interpretation is discussed in Section 7.4. We finish this chapter with a short in-
troduction to the linearization of Poisson structures (in the neighborhood of a point
where the rank is zero).
Unless otherwise stated, F denotes an arbitrary field of characteristic zero.
Suppose that g is a finite-dimensional Lie algebra over F, with Lie bracket [· , ·]. We
denote the dual vector space to g by g∗ . We associate to each element1 e of g, a
linear function e∗ : g∗ → F, defined by
e∗ : g ∗ → F
ξ → ξ , e := ξ (e) .
for all e, f ∈ g. On the one hand, this implies that Definition (7.1) is independent
of the chosen basis (e1 , . . . , ed ). On the other hand, it yields, for all i, j, k with
1 i, j, k d,
xi , x j , xk = [[ei , e j ], ek ]∗ ,
so that the Jacobi identity for [· , ·] implies that xi , x j , xk + (i, j, k) = 0 for all
1 i < j < k d. According to Proposition 1.8, this shows that {· , ·} is a Poisson
structure on F[x1 , . . . , xd ]. Similarly, according to (iii) in Proposition 1.36, Eq. (7.1)
also defines a Poisson structure on the algebra of smooth functions on g∗ , when
F = R, or on the algebra of holomorphic functions on g∗ , when F = C; in either
case, it is the unique Poisson structure which satisfies (7.2). Thus, F (g∗ ) inherits a
Poisson structure from the Lie bracket on g, irrespective of whether we take F (g∗ )
to be the algebra of polynomial, smooth, or holomorphic functions on g∗ .
An intrinsic formula for {F, G} can be written down in terms of the differentials
of F and G. Since the differential of F ∈ F (g∗ ) at ξ is a linear map dξ F : Tξ g∗ → F,
and since Tξ g∗ is naturally isomorphic with g∗ , we can think of dξ F as being an
element of (g∗ )∗ , i.e., as an element of g, since g and its bidual are canonically
isomorphic (recall that g is finite-dimensional; the canonical isomorphism g → (g∗ )∗
1 We use in this section letters e and f to denote elements of a Lie algebra g, to reserve the letters
x, y, z for elements of the bidual (g∗ )∗ . After identification of g with its bidual, the letters x, y, z
become elements of g, just like everywhere else in the book.
7.1 The Lie–Poisson Structure on g∗ 181
Definition 7.1. Let (g, [· , ·]) be a finite-dimensional Lie algebra and let F (g∗ ) de-
note the algebra of polynomial, smooth or holomorphic functions on g∗ . The Poisson
structure on g∗ , defined for F, G ∈ F (g∗ ), at ξ ∈ g∗ by
{F, G} (ξ ) := ξ , [dξ F, dξ G] (7.5)
d
xi , x j = [ei , e j ]∗ = ∑ ckij xk .
k=1
In this notation, Eq. (7.1) for the Poisson structure takes the form
d
∂F ∂G
{F, G} := ∑ ckij xk
∂ xi ∂ x j
.
i, j,k=1
For monomials of degree 1, the Poisson bracket {· , ·} on Sg is just the usual Lie
bracket, {z, z } := [z, z ]. For monomials of higher degree, one uses the biderivation
property, which gives
p q
z1 . . . z p , z1 . . . zq = ∑ ∑ z1 . . . zk . . . z p z1 . . . z . . . zq zk , z .
k=1 =1
From this formula, one easily gives a direct proof of the Jacobi identity for the
bracket {· , ·} on Sg.
ξ , ψ (η ) = ψ (ξ ) | ψ (η ) = η , ψ (ξ ) , (7.8)
dχ (x) (F ◦ ψ ) = dx F ◦ ψ ,
where we have used (7.8) in the last step. Equation (7.9) is usually written in the
following form,
{F, G}g (x) = x | [∇x F, ∇x G] , (7.10)
where ∇x F, the gradient of F at x (with respect to · | ·) is defined, for F ∈ F (g)
and x ∈ g by
∇x F | y = dx F, y , (7.11)
for every y ∈ g, which is equivalent to saying that ∇x F = ψ (dx F). Since F is a
function on a vector space, (7.11) can also be written in the following form,
d
∇x F | y = F(x + ty) , (7.12)
dt |t=0
A⊥ := {x ∈ g | ∀y ∈ A, x | y = 0} . (7.13)
Proposition 7.5. Let (g, [· , ·]) be a finite-dimensional Lie algebra, equipped with a
non-degenerate, symmetric bilinear form · | ·, and suppose that h ⊂ g is a Lie ideal
of g. Then its orthogonal, h⊥ , is a Poisson submanifold of g. For F, G ∈ F (h⊥ ),
their Poisson bracket {F, G}h⊥ is given, at x ∈ h⊥ , by
{F, G}h⊥ (x) = F̃, G̃ g (x) , (7.14)
We now show that the Hamiltonian vector fields XH on (g, {· , ·}) can be writ-
ten down in a simple form, if · | · is ad-invariant (i.e., when g is a quadratic Lie
algebra). Recall from Section 5.1.4 that this means that x | [y, z] = [x, y] | z for
all x, y, z ∈ g. Then we have that (7.10) can be written, for H = G, as
Comparing this to
we find that ẋ := (XH )x , the Hamiltonian vector field XH at x, is given by the Lax
equation
ẋ = [∇x H, x] . (7.15)
Lax equations are very important in the theory of integrable systems, as one easily
shows that their flows are isospectral, hence admit many constants of motion (see
Section 12.2.5 below).
7.3 Properties of the Lie–Poisson Structure 185
In this section we first study the symplectic foliation of a Lie–Poisson structure and
show how it is related to the coadjoint action. We then explain how the Poisson
cohomology of a Lie–Poisson structure is related to Lie algebra cohomology. In
Section 7.3.3, we relate the modular class of a Lie–Poisson structure to the modular
form of a Lie algebra.
We have seen in Section 1.3.3 that a Poisson manifold naturally decomposes into
symplectic manifolds. We now show that in the case of a Lie–Poisson structure
186 7 Linear Poisson Structures and Lie Algebras
Proof. We first show that at each point ξ ∈ g∗ the vector space of all fundamental
vector fields of the coadjoint action of G on g∗ at ξ coincides with Hamξ (g∗ , {· , ·}),
the vector space of all Hamiltonian vector fields of the Lie–Poisson structure on g∗ ,
at ξ . To show it, rewrite (7.5) as
(XG )ξ [F] = ξ , − addξ G dξ F = ad∗dξ G ξ , dξ F
(XG )ξ = ad∗dξ G ξ .
as we needed to show. Since the coadjoint orbits of G are connected, this shows that
the leaves of the symplectic foliation of the Lie–Poisson structure on g∗ are precisely
the coadjoint orbits. It follows that the Ad∗ -invariant functions are precisely the
functions which are constant on every leaf of the symplectic foliation. According to
Proposition 1.31, they are the Casimir functions of the Lie–Poisson structure on g∗ .
Fig. 7.1 For r 0 the sphere x2 + y2 + z2 = r2 is a symplectic leaf of the Lie–Poisson structure
on (so3 (R))∗ .
ture, the symplectic foliation can be rather complicated. This is illustrated in the
following examples of (real) three-dimensional Lie algebras.
Example 7.9. We start with the Lie–Poisson structure of (so3 (R))∗ , which is defined
by the following brackets
Example 7.10. We next consider the Lie–Poisson structure of (sl2 (R))∗ , where the
standard sl2 (R) brackets
Fig. 7.2 Every point of the plane x = 0 is a symplectic leaf for the Lie–Poisson structure on the
dual of the Heisenberg algebra. The planes x = c, with c = 0, are the other symplectic leaves.
Clearly, C := 4xy + z2 is a Casimir function and the rank of the Poisson structure
is two at every point, except at the origin; also, the differential of C is non-zero,
except at the origin. As in the previous example, Proposition 1.32 implies that the
symplectic leaves are the connected components of the fibers of C, namely,
(1) The one-sheeted hyperboloids 4xy + z2 = c, for c > 0;
(2) The two components of the two-sheeted hyperboloids 4xy + z2 = c, for c < 0;
(3) The two components of the cone z2 = −4xy minus the origin;
(4) The origin.
Notice that the previous example and the present example are geometrically
very different, yet the underlying Lie algebras have the same complexification,
namely sl2 (C).
Example 7.11. Our third example is the Heisenberg algebra, with Lie brackets
It is immediate that the Casimir functions are precisely the functions which are
independent of y and of z. The Poisson structure has now rank zero on the plane
x = 0, so that each point of this plane is a symplectic leaf; together, these points
form the zero locus of the Casimir function x. The other symplectic leaves are two-
dimensional, they are the planes x = c, with c ∈ R∗ . See Fig. 7.2.
Example 7.12. In the fourth example we show that, for a linear Poisson structure
on R3 , non-trivial smooth Casimir functions may even not exist locally (at some
points). Take on R3 , with respect to the system of coordinates (x, y, z), the following
Poisson matrix, ⎛ ⎞
0 0 x
⎜ ⎟
⎜ 0 0 y⎟ (7.16)
⎝ ⎠.
−x −y 0
7.3 Properties of the Lie–Poisson Structure 189
Fig. 7.3 For the open book foliation, which is the symplectic foliation of the Lie–Poisson structure,
given by the Poisson matrix (7.16) the leaves are the points of the Z-axis and the half-planes,
obtained by removing the Z-axis from the planes passing through it.
Example 7.13. We consider, in the last example, for a fixed α ∈ C, the basic Lie
brackets
[x, y] = 0 , [y, z] = y , [z, x] = α x ,
which lead to the following Poisson matrix for the corresponding Lie–Poisson struc-
ture on C3 , ⎛ ⎞
0 0 −α x
⎜ ⎟
⎜ 0 0 y ⎟.
⎝ ⎠
α x −y 0
∂F ∂F ∂F
= αx −y =0,
∂z ∂x ∂y
We show in the following proposition that, for every q ∈ N, the cohomology space
HLq (g, F (g∗ )) is canonically isomorphic to the q-th Poisson cohomology space
of (g∗ , π ).
Proposition 7.14. Let g be a finite-dimensional Lie algebra. The dual g∗ of g is
equipped with its algebra F (g∗ ) of polynomial, holomorphic or smooth functions.
The Lie–Poisson bracket on F (g∗ ) is denoted by π . For every q ∈ N there is a
canonical isomorphism
HLq (g, F (g∗ )) Hπq (g∗ ) .
Proof. We show that the F (g∗ )-valued Lie algebra cohomology complex of g is
isomorphic to the Poisson cohomology complex of (g∗ , π ). Let Q be a Poisson q-
cochain of (F (g∗ ), π ), i.e., Q ∈ Xq (g∗ ). We associate to Q the element Ψq (Q) ∈
Cq (g; F (g∗ )) = Hom(∧q g, F (g∗ )), defined by
Ψq (Q) : ∧q g → F (g∗ )
(7.18)
x1 ∧ · · · ∧ xq → Q[x1∗ , . . . , xq∗ ] .
7.3 Properties of the Lie–Poisson Structure 191
It is easy to see that for every q ∈ N the resulting linear map Ψq : Xq (g∗ ) →
Hom(∧q g, F (g∗ )) is anisomorphism. Moreover, a direct comparison of (4.4)
and (7.17), using xi∗ , x∗j = [xi , x j ]∗ , shows that the following diagram is commu-
tative. q
δπ
Xq (g∗ ) Xq+1 (g∗ )
Ψq Ψq+1
q
δL
Hom(∧q g, F (g∗ )) Hom(∧q+1 g, F (g∗ ))
where the first map is an isomorphism, but the second map is in general neither
injective nor surjective. In the case of the above representation of g on F (g∗ ), it
leads for every q ∈ N to the linear map
since the invariant elements of F (g∗ ) are the functions F ∈ F (g∗ ) such that
{x∗ , F} = 0 for every x ∈ g, i.e., the Casimir functions of (g∗ , π ). Recall from Propo-
sition 7.7 that these functions are the Ad∗ -invariant functions. For a semi-simple Lie
algebra, a non-trivial property on its Lie algebra cohomology (with coefficients in a
finite-dimensional representation) implies that ıq is an isomorphism, for every q ∈ N,
as we show in the following proposition.
Proposition 7.15. Let g be a semi-simple Lie algebra. We denote by F (g∗ ) the al-
gebra of polynomial functions on g∗ and by Cas(g∗ ) ⊂ F (g∗ ) the subalgebra of
polynomial Casimir functions. For every q ∈ N, the map ıq , given by (7.20), is an
isomorphism:
ıq : HLq (g) ⊗ Cas(g∗ ) Hπq (g∗ ) .
Proof. We denote by Fk (g∗ ) the vector space of homogeneous polynomials of de-
gree k on g∗ , where k ∈ N. It is a subrepresentation of F (g∗ ), and
F (g∗ ) = Fk (g∗ ) ,
k∈N
which implies that the F (g∗ )-valued Lie algebra cohomology of g decomposes as
a direct sum, q
HLq (g, F (g∗ )) = HL (g, Fk (g∗ )) ,
k∈N
192 7 Linear Poisson Structures and Lie Algebras
for every q ∈ N. Moreover, for every q, k ∈ N, the linear map ıq restricts to yield a
map
ıq,k : HLq (g) ⊗ Cask (g∗ ) → HLq (g, Fk (g∗ )) .
Now, according to [95, Th. 11], for every finite-dimensional representation V of a
semi-simple Lie algebra g, the natural linear map HLq (g) ⊗V g → HLq (g,V ) described
in (7.19) is bijective for every q ∈ N. In particular, for every k ∈ N, ıq,k is an iso-
morphism, which, in turn, implies that ıq is an isomorphism, as was to be shown.
so that μg vanishes on [g, g]. Every Lie algebra which satisfies [g, g] = g is therefore
unimodular; in particular, if g is semi-simple, then g is unimodular. If g is nilpotent,
then g is unimodular, since adx is nilpotent, hence traceless, for every x ∈ g. For
a different reason, quadratic Lie algebras are unimodular as well. However, solv-
able Lie algebras are not unimodular in general, for instance, the two-dimensional
Lie algebra defined, on a basis (e, f ) of g by [e, f ] := f , is not unimodular, since
Trace(ade ) = 1.
In the following proposition, we relate the modular form μg of g with the modular
vector field Φ of the Poisson manifold (g∗ , π ), where π stands for the canonical Lie–
Poisson structure on g∗ . The volume form which we choose on g∗ is any translation
invariant volume form (see Remark 4.13). Φ is a constant vector field on g∗ , hence
corresponds in a canonical way to an element of g∗ .
Proposition 7.16. Let g be a finite-dimensional Lie algebra. Interpreted as an ele-
ment of g∗ , the modular vector field of (g∗ , π ) with respect to any translation invari-
ant volume form is, up to a sign the modular form of g.
Proof. Let (e1 , . . . , ed ) be a basis of g and let (ξ1 , . . . , ξd ) be the dual basis of g∗ .
We use the functions x1 = e∗1 , . . . , xd = e∗d as linear coordinates on g∗ . According
7.4 Affine Poisson Structures 193
∂ ∂ ∂
Div(π ) = ∑ xi , x j = ∑ [ei , e j ]∗ , ξ j .
1i, jd ∂ x j ∂ xi 1i, jd ∂ xi
In view of (3.37), the first term in this expression is a linear function on V , while the
second term is a constant function on V . Thus, [π , π ]S = 0 if and only if [π1 , π1 ]S = 0
(i.e., π1 is a linear Poisson structure) and [π0 , π1 ]S = 0 (i.e., π0 and π1 are compati-
ble). It follows that the map, which sends an affine Poisson structure π to its constant
and linear parts (π0 , π1 ), yields the announced one-to-one correspondence.
194 7 Linear Poisson Structures and Lie Algebras
see Example 4.2, in particular formula (4.2) for the Lie algebra coboundary opera-
tor δL , in the case of a trivial representation. It follows that [π0 , π1 ]S = 0 if and only
if c is a Lie 2-cocycle, δL2 (c) = 0.
In view of the above proposition, affine Poisson structures are also called modified
canonical Poisson structures or modified Lie–Poisson structures.
If g is semisimple, then HL2 (g) is trivial (see Lemma 4.1), so that, if c is a Lie 2-
cocycle, then c is a Lie 2-coboundary, c = δL1 (c ), written out, c(xi , x j ) = c ([x j , xi ])
and we see that the affine Poisson structure, corresponding to ([· , ·], c), is nothing but
the Lie–Poisson structure associated to [· , ·], with xi∗ replaced by xi∗ +c (xi ), i.e., both
Poisson structures are the same up to an affine change of variables (a translation).
Affine Poisson structures are, in general, also closely related to Lie–Poisson
structures in a different way. Namely, given a Lie algebra (g, [· , ·]) and a Lie 2-
cocycle c, as in Proposition 7.19, we consider on g × F the Lie bracket
[[(x1 , a1 ), (x2 , a2 )]c , (x3 , a3 )]c = [([x1 , x2 ] , c(x1 , x2 )), (x3 , a3 )]c
= ([[x1 , x2 ] , x3 ] , c([x1 , x2 ] , x3 )) ,
so that the Jacobi identity for [· , ·] and the cocycle condition for c imply the Jacobi
identity for [· , ·]c , i.e., (g × F, [· , ·]c ) is a Lie algebra. In the following proposition we
show that the Lie–Poisson structure {· , ·}c on (g × F)∗ , restricted to a hyperplane,
is isomorphic to the affine Poisson structure {· , ·} on g, defined by the 2-cocycle c.
Proposition 7.20. Let (g, [· , ·]) be a finite-dimensional Lie algebra and let c be a 2-
cocycle in the trivial Lie algebra cohomology of g. Denote the Lie–Poisson structure
on (g × F)∗ , associated to (g × F, [· , ·]c ), by {· , ·}c , and denote by {· , ·} the affine
Poisson structure on g∗ , with constant term c.
(1) The linear function 1∗ : (g × F)∗ → F, defined by φ → φ (0, 1), is a Casimir
function of {· , ·}c ;
(2) The hyperplane N := {φ ∈ (g × F)∗ | φ (0, 1) = 1} is a Poisson submanifold
of (g × F)∗ ;
(3) The isomorphism Ψ : N → g∗ , defined for φ ∈ N by φ → φ (·, 0), is a Poisson
isomorphism between (N, {· , ·}c ) and (g∗ , {· , ·}).
Proof. We first show that 1∗ is a Casimir function of {· , ·}c . Notice that 1∗ = (0, 1),
as an element of g × F. For (ξ , a) ∈ g∗ × F, it follows that d(ξ ,a) 1∗ = 1∗ = (0, 1),
which belongs to the center of [· , ·]c (see (7.22)). It follows that for every function
F on (g × F)∗ ,
{F, 1∗ }c (ξ , a) = (ξ , a), [d(ξ ,a) F, (0, 1)]c = 0 .
This shows (1). Since the hyperplane N, as defined in (2), is a level set of the
Casimir function 1∗ , it is a Poisson submanifold of (g × F)∗ , which yields (2).
It is clear that Ψ is an isomorphism between the affine space N and the vec-
tor space g∗ . In order to prove that Ψ is a Poisson morphism, it is sufficient to
prove that {x∗ ◦ Ψ , y∗ ◦ Ψ }c = {x∗ , y∗ } ◦ Ψ , for every x, y ∈ g, where, as above,
stars stand for the corresponding linear functions on g∗ . By a slight abuse of no-
tation, we will also consider x∗ as a linear function on (g × F)∗ , where we put
x∗ (φ ) := φ (x, 0), for all φ ∈ (g × F)∗ . With this abuse of notation, x∗ ◦ Ψ = x∗ ,
and {x∗ , y∗ } ◦ Ψ = [x, y]∗ + c(x, y). For {x∗ ◦ Ψ , y∗ ◦ Ψ }c , which is now written
as {x∗ , y∗ }c , let φ ∈ N and compute
{x∗ , y∗ }c (φ ) = φ , dφ x∗ , dφ y∗ c = φ , [(x, 0), (y, 0)]c
= φ , ([x, y], c(x, y)) = [x, y]∗ (φ ) + c(x, y) ,
As we have shown earlier in this section, every affine Poisson structure which is
constructed by using a Lie 2-cocycle which is a coboundary, is isomorphic to the
196 7 Linear Poisson Structures and Lie Algebras
Example 7.21. Consider on R3 the linear and affine Poisson structures π1 and π ,
whose Poisson matrices are given by
⎛ ⎞ ⎛ ⎞
0 0 −x 0 1 −x
⎜ ⎟ ⎜ ⎟
⎜0 0 y ⎟ and ⎜ −1 0 y ⎟.
⎝ ⎠ ⎝ ⎠
x −y 0 x −y 0
The underlying Lie algebra of both Poisson structures is defined by the brackets
[x, y] = 0, and [y, z] = y and [z, x] = x. The corresponding Lie 2-cocycle is given by
c(x, y) = 1 and c(x, z) = c(y, z) = 0. Notice that π is a regular Poisson structure,
of rank two, while the rank of π1 vanishes at the origin. Clearly, xy is a Casimir
function of π1 , while xy + z is a Casimir function of π . The symplectic leaves of π1
are the points on the Z-axis, the two half-planes obtained by removing the Z-axis
from the plane x = 0, the two half-planes obtained by removing the Z-axis from
the plane y = 0, and the connected components of the hyperbolic cylinders xy = c,
with c ∈ R∗ . For the affine Poisson structure π , the symplectic leaves are the parallel
hyperboloids xy + z = c, with c ∈ R.
Proposition 7.22. Let (M, π ) be a real or complex Poisson manifold, and suppose
that m ∈ M is a point at which π vanishes. There exists a unique linear Poisson
structure {· , ·}1 = π1 on Tm M, such that, for every neighborhood U of m in M and
for all F, G ∈ F (U),
{dm F, dm G}1 = dm {F, G} , (7.23)
where the elements dm F, dm G and dm {F, G} of Tm∗ M are viewed as (linear) functions
on Tm M.
7.5 The Linearization of Poisson Structures 197
and similarly for the partial derivatives of G at m. The Jacobi identity for the Poisson
structure π implies the Jacobi identity for π1 since it follows at once from (7.23) that
on TΨ (m) M ,
{L1 , L2 }1 ◦ dmΨ = {L1 ◦ dmΨ , L2 ◦ dmΨ }1 . (7.25)
Given such functions L1 and L2 , there exist functions F and G, defined on a neigh-
borhood of Ψ (m) in M , such that dΨ (m) F = L1 and dΨ (m) G = L2 . Since Ψ is a
Poisson map, {F, G} ◦ Ψ = {F ◦ Ψ , G ◦ Ψ } , which we differentiate at m, giving
Let us prove the claim. First, if do F = 0, then {do x, do y}1 = do F = 0, so that the
linearized Poisson structure is trivial; in view of Example 7.24, π is in this case not
linearizable at o. Assume now that do F = 0 and consider the vector field
∂F ∂ ∂F ∂
V := − ,
∂x ∂y ∂y ∂x
which does not vanish at o. By the straightening theorem (Theorem B.7), there exists
a function G, defined in a neighborhood of o, such that V [G] = 1. Since
∂F ∂G ∂F ∂G
− = V [G] = 1 ,
∂x ∂y ∂y ∂x
we have on the one hand that do F and do G are linearly independent, and on the other
hand that {F, G} = F. Thus, (F, G) is a system of coordinates on a neighborhood
of o and π is linear in terms of these coordinates. According to Example 7.25, π is
linearizable at o.
Remark 7.28. For i = 1, 2, let (Mi , πi ) be a Poisson manifold, whose Poisson struc-
ture vanishes at mi ∈ Mi . Let Ψ : M1 → M2 be an isomorphism of Poisson manifolds
with Ψ (m1 ) = m2 . Then the tangent map TmΨ : Tm1 M1 → Tm2 M2 is an isomorphism
of Poisson manifolds, where both tangent spaces are equipped with the linearized
Poisson structures of π1 and π2 .
In particular, if two finite-dimensional vector spaces V1 and V2 , equipped with a
linear Poisson structure, are isomorphic as Poisson manifolds via a map Ψ which
sends the origin o of V1 to the origin of V2 , then the linear map doΨ : V1 ToV1 →
ToV2 V2 is a Poisson isomorphism. In other words, if there exists (in the neighbor-
hood of the origin) a Poisson isomorphism between two linear Poisson structures,
then a linear Poisson isomorphism between them also exists.
The condition that a Poisson structure π which vanishes at m ∈ M is linearizable
at m can be expressed as the existence of a neighborhood U of m in M and of a
diffeomorphism Ψ between U and a neighborhood of the origin o in Tm M, with
Ψ (m) = o, such that
{ ξ ◦ Ψ , η ◦ Ψ } = {ξ , η }1 ◦ Ψ ,
Example 7.30. It is clear from the definition that every Poisson structure which is
linearizable at a point is at that point linearizable up to order , for all ∈ N∗ .
However, the converse is not true, in general. To show this, we define a (smooth)
−2
Poisson structure π on R2 by {x, y} := f (x), where f (x) := e−x for x = 0 and
f (0) := 0. The fact that f vanishes at 0, together with all its derivatives at 0, implies
(+1)
that {x, y} ∈ Io (R2 ), for all ∈ N, where o denotes the origin of R2 . It follows
on the one hand that the linearized Poisson structure at o is trivial, and on the other
hand that π is linearizable up to order , for all ∈ N∗ . However, the Poisson struc-
ture π is not linearizable at o, since there is no neighborhood of o on which π is
zero.
d
xi , x j − ∑ ckij xk ∈ Im (U) ,
(2)
(7.26)
k=1
d
x̃i , x̃ j 1
= ∑ ckij x̃k , (7.27)
k=1
d
yi , y j = ∑ ckij yk , (7.28)
k=1
d
yi , y j − ∑ ckij yk ∈ Im
(+1)
(V ) . (7.29)
k=1
The inclusion (7.26) holds for all 1 i, j d if and only if π vanishes at m. In this
case, the linearized Poisson structure π1 at m is given by (7.27), where x̃1 , . . . , x̃d are
the linear coordinates on Tm M, defined by x̃i := dm xi , for i = 1, . . . , d. Still assuming
that π vanishes at m, there exists a coordinate chart (V, y) for M, centered at m ∈ M,
such that (7.28) (respectively (7.29)) holds for all 1 i, j d, if and only if π is
linearizable at m (respectively π is linearizable at m up to order ).
According to Proposition 7.3, there is a one-to-one correspondence between lin-
ear Poisson structures on Tm M and Lie algebra brackets on Tm∗ M. It follows that there
corresponds to the linearized Poisson structure at m (where m is a point at which π
vanishes) a Lie algebra structure [· , ·] on Tm∗ M. It satisfies, by construction,
Proof. We do not prove Conn’s results (1) and (2), as the proof is long and rather
technical (see [45, 46]); instead, we prove the analogous (but easier to prove) re-
sult (3). We do this by recursion on . Every Poisson structure which vanishes at m
is linearizable up to order 1, so that the statement holds true for = 1. Assume
now that π is linearizable up to order . According to Definition 7.29, there exists a
neighborhood U of m in M and a diffeomorphism Ψ between U and a neighborhood
of the origin o in Tm M, with Ψ (m) = o, such that
(+1)
{x ◦ Ψ , y ◦ Ψ } − [x, y] ◦ Ψ ∈ Im (U) , (7.30)
for all x, y ∈ Tm∗ M; we write here, and in the rest of the proof, the elements of
Tm∗ M with latin letters x, y, z, because we view Tm∗ M = g as a Lie algebra. For
x ∈ g, we write Fx as a shorthand for the element x ◦ Ψ of F (U) and we denote
(+1) (+2)
I := Im (U) and I := Im (U). Notice that I /I is finite-dimensional,
because I /I is isomorphic to the space of homogeneous polynomials of degree
+ 1 in dim M variables. With these notations, (7.30) can be written as
Fx , Fy − F[x,y] ∈ I . (7.31)
χ : g × (I /I ) → I /I
(7.32)
(x, p(F)) →
x · p(F) := p({Fx , F}) ,
c : g∧g → I /I
(7.34)
(x, y) → p( Fx , Fy − F[x,y] ) .
Since HL2 (g; I /I ) is trivial, there exists a linear map b : g → I /I , such that c
is the coboundary of b,
which is zero in view of (7.35). This proves (7.36). Since x → Fx is a linear map,
there exists a map Ψ+1 from U to a neighborhood of o in Tm M, such that Fx =
x ◦ Ψ+1 , for every x ∈ g. Since dmΨ = dmΨ+1 , there exists a neighborhood U ⊂ U
7.6 Notes 203
Remark 7.32. It is clear from the above proof that the open subsets on which the
diffeomorphisms Ψ are defined form a sequence which decreases with , and that
the intersection of all these open subsets may be reduced to a point. Notice that,
even if this intersection is not reduced to a point, the Poisson structure may not be
linearizable at m (see Example 7.30).
7.6 Notes
The fact that Lie–Poisson structures are in one-to-one correspondence with Lie al-
gebra structures is implicit in Lie’s works, in which one also finds the relation to the
coadjoint representation of the adjoint group of the Lie algebra. In the first half of
the twentieth century, the theory of Lie algebras and Lie groups became a subject of
its own, with focus on applications to theoretical physics and differential geometry.
The cited observations of Lie were rediscovered and reinforced in the works of Kir-
illov [105, 106], Kostant [117] and Souriau [185], who clearly exhibit the symplectic
structure on the coadjoint orbits which is inherited from the Lie–Poisson structure;
this structure is nowadays known as the Kostant–Kirillov–Souriau symplectic struc-
ture. For general information about Lie algebras and their representation theory, we
refer to Jacobson [102] or to the first chapter of Bourbaki [27].
Exactly as Lie algebra structures are in one-to-one correspondence with linear
Poisson structures, Lie algebroids are in one-to-one correspondence with fiberwise
linear Poisson structures [64], i.e., Poisson structures on a vector bundle which have
the property that the bracket of any two fiberwise linear functions is a fiberwise
linear function, while the bracket of a fiberwise linear function with a basic function
is a basic function. See [53] for an introduction to Lie algebroids.
The linearization problem was first posed and studied by Weinstein [199]. The
main results, which give conditions under which a Poisson structure is linearizable,
were obtained by Conn [45, 46]; a geometrical proof of these results is given in [52].
See Dufour–Zung [63] for a detailed account on the linearization problem.
Chapter 8
Higher Degree Poisson Structures
Constant and linear Poisson structures, studied in the previous two chapters, are
particular cases of a bigger class of Poisson structures, namely the (weight) ho-
mogeneous Poisson structures on a finite-dimensional vector space V : constant and
linear Poisson structures are the homogeneous Poisson structures of degree zero and
one, respectively. Like in the constant or linear case, (weight) homogeneous Poisson
structures are easier to study than general Poisson structures. For example, homo-
geneity of a Poisson structure permits one to split the determination of its Poisson
cohomology into its homogeneous parts. Also, weight homogeneous Poisson struc-
tures turn out to be useful for studying general Poisson structures. We will see an
example of this in Section 9.1.3 of the next chapter, when classifying Poisson struc-
tures on F2 , with a simple singularity.
Among distinguished classes of weight homogeneous Poisson structures, one
finds – besides constant and linear Poisson structures – quadratic Poisson struc-
tures (i.e., homogeneous Poisson structures of degree two), rank 2 Poisson structures
which arise from Nambu–Poisson structures and the transverse Poisson structures
to adjoint orbits in a semi-simple Lie algebra.
Section 8.1 is devoted to the study of general properties of (weight) homoge-
neous Poisson structures, while we study in Section 8.2 a few special properties and
characterizations of quadratic Poisson structures: their multiplicativity, their mod-
ular class and the decomposition of a quadratic Poisson structure, using a related
unimodular Poisson structure. In Section 8.3 we define the notion of a Nambu–
Poisson structure and we explain the construction which leads to a large class of
(weight homogeneous) rank 2 Poisson structures. Finally, Section 8.4 deals with the
transverse Poisson structure to a nilpotent orbit in a semi-simple Lie algebra.
Unless otherwise stated, F denotes an arbitrary field of characteristic zero.
In this section, we define and study polynomial Poisson structures on a vector space,
with special emphasis on homogeneous and weight homogeneous Poisson struc-
tures.
Throughout the section, we consider Poisson structures and, more generally, mul-
tivector fields on a finite-dimensional vector space V , which is equipped with its al-
gebra of functions, denoted by F (V ). If F is an arbitrary field, this algebra is the al-
gebra of polynomial functions on V ; however, if F = R (respectively F = C), F (V )
may stand either for the algebra of smooth (respectively holomorphic) or polyno-
mial functions on V , depending on the context. It is clear that, in either case, F (V )
contains all polynomial functions on V . For p ∈ N, the corresponding space of p-
vector fields on V (skew-symmetric p-derivations of F (V )) is denoted by X p (V ).
The origin of V will be denoted by o.
Example 8.2. Constant Poisson structures (see Chapter 6) are polynomial Poisson
structures of degree zero, while linear and affine Poisson structures (see Chapter 7)
are polynomial Poisson structures of degree one.
It is clear that when F (V ) is the algebra of polynomial functions on V , then all
Poisson structures on V are polynomial Poisson structures on V .
E [F] = LE F = r F . (8.5)
(LE P − (r − p) P) [F1 , . . . , Fp ]
p
= E [P[F1 , . . . , Fp ]] − ∑ P[F1 , . . . , E [F ], . . . , Fp ] − (r − p) P[F1 , . . . , Fp ]
=1
p
= E [P[F1 , . . . , Fp ]] − ∑ deg(F ) − (r − p) P[F1 , . . . , Fp ] .
=1
Example 8.7. Consider the bivector field π := V1 ∧ V2 , where V1 and V2 are two
homogeneous vector fields on a finite-dimensional vector space V . Then π is homo-
geneous of weight ϖ (V1 ) + ϖ (V2 ), because
Since
[π , π ]S = [V1 ∧ V2 , V1 ∧ V2 ]S = −2 V1 ∧ V2 ∧ [V1 , V2 ] ,
we have that π is a (homogeneous) Poisson structure on V if and only if the vec-
tor fields V1 , V2 , [V1 , V2 ] are linearly dependent at all points of V . This condition is
automatically satisfied when V1 is the Euler vector field E and V2 = V is an ar-
bitrary homogeneous vector field, since then [E , V ] = LE V = ϖ (V ) V , which is
a multiple of V . The Poisson structure E ∧ V is homogeneous of weight ϖ (V ),
since ϖ (E ) = 0.
To finish this section, we give a proposition whose main content is that an isomor-
phism between two homogeneous Poisson structures on a vector space can always
be realized by a linear isomorphism.
Proposition 8.8. Let π1 and π2 be two homogeneous Poisson structures on a finite-
dimensional vector space V . Then the following are equivalent:
(i) There exists a Poisson isomorphism L : (V, π1 ) → (V, π2 ) which is a linear
map;
(ii) There exists a Poisson isomorphism Ψ : (V, π1 ) → (V, π2 ) with Ψ (o) = o;
(iii) There exists a Poisson isomorphism Ψ : (U1 , π1 ) → (U2 , π2 ) with Ψ (o) = o,
where U1 ,U2 are neighborhoods of the origin o in V .
Proof. The implications (i) =⇒ (ii) =⇒ (iii) are clear. Assume now that (iii) holds,
and let Ψ : (U1 , π1 ) → (U2 , π2 ) be a Poisson isomorphism with Ψ (o) = o. We are
going to prove that the differential L = doΨ of Ψ at the origin is a linear Poisson
isomorphism between (V, π1 ) and (V, π2 ). This can be done by comparing the Taylor
expansions at o of both sides of the following identity, which is valid for all functions
F, G ∈ F (V2 )
{F, G}2 ◦ Ψ = {F ◦ Ψ , G ◦ Ψ }1
where {· , ·}1 = π1 and {· , ·}2 = π2 are the Poisson brackets corresponding to π1
and π2 . Since π1 and π2 are homogeneous of degree k, the Taylor expansions at
the origin o of both sides of this equality vanish for degrees 0, . . . , k − 1, while the
components of degree k are equal to {do F, do G}2 ◦ L and {do F ◦ L, do G ◦ L}1 respec-
tively. Since every linear function on V is the differential at the origin of a function
in F (V ), we obtain that for all linear functions F , G ∈ F (V ),
F , G 2 ◦ L = F ◦ L, G ◦ L 1
which, in turn, implies that L is a Poisson map. Moreover, since L = doΨ is invert-
ible, L is in fact a Poisson isomorphism.
8.1 Polynomial and (Weight) Homogeneous Poisson Structures 211
We will use the above proposition in Section 9.2.3, when we classify the set of
all quadratic Poisson structures on C3 .
for all λ ∈ F. In this case, the integer r, which is unique if F = 0, is called the weight
of F and is denoted by ϖ (F). It is easy to see that, if F, G ∈ F (Fd ) are weight
homogeneous with respect to ϖ , then their product FG is also weight homogeneous
with respect to ϖ , of weight ϖ (FG) = ϖ (F) + ϖ (G).
Remark 8.9. It is clear that a function F on Fd is homogeneous if and only if it is
weight homogeneous with respect to ϖ = (1, . . . , 1), and in this case, the weight
of F is equal to the degree of F.
Taking F := xi in (8.8), we see that each coordinate function xi , with 1 i d is
weight homogeneous of weight ϖi , so that we often refer to ϖi as being the weight
of xi .
Similarly to the case of homogeneous functions (see Proposition 8.3), we show
in the following proposition that a smooth or holomorphic function which is weight
homogeneous, is a polynomial function.
Proposition 8.10. If F is a smooth function on Rd (respectively a holomorphic func-
tion on Cd ) and F is weight homogeneous with respect to some weight vector, then F
is a polynomial function.
Proof. Let ϖ = (ϖ1 , . . . , ϖd ) be a weight vector for Fd (F = R or F = C) and assume
that F ∈ F (Fd ) satisfies (8.8), for all λ ∈ F. Let us consider s ∈ N∗ , such that, for all
1 j1 , . . . , js d, we have that ϖ j1 + · · · + ϖ js > r. By differentiating both sides of
equation (8.3) with respect to xi1 , . . . , xis , for all 1 i1 , . . . , is d, we obtain that all
functions ∂ xi ∂...F∂ xi have negative weight, hence are zero on Fd . This implies that F
s
1 s
is necessarily a polynomial function on Fd .
212 8 Higher Degree Poisson Structures
∂ ∂
Eϖ := ϖ1 x1 + · · · + ϖ d xd . (8.9)
∂ x1 ∂ xd
Notice that the weighted Euler vector field Eϖ is weight homogeneous with re-
spect to ϖ . It is clear from (8.8) that a function F on Fd is weight homogeneous of
weight r if and only if
Eϖ [F] = LEϖ F = r F . (8.10)
This formula, which generalizes the classical Euler formula (8.5), is called the
weighted Euler formula.
Proposition 8.12. Let ϖ = (ϖ1 , . . . , ϖd ) be a weight vector for Fd and let P be a
non-zero p-vector field on Fd , where p ∈ N. The following conditions on P are
equivalent.
(i) P is weight homogeneous with respect to ϖ ;
(ii) There exists an integer r ∈ Z, such that for all 1 i1 < · · · < i p d, the
function P[xi1 , . . . , xi p ] is zero, or is weight homogeneous of weight r + ϖi1 +
· · · + ϖi p ;
(iii) There exists an integer r ∈ Z, such that for all weight homogeneous elements
F1 , . . . , Fp of F (Fd ), one has that P[F1 , . . . , Fp ] is zero, or is weight homoge-
neous of weight r + ϖ (F1 ) + · · · + ϖ (Fp );
(iv) There exists an integer r ∈ Z, such that LEϖ P = rP.
The (possibly negative) integer r, which appears in (ii)–(iv) is the same and is
called 2 the weight of P, denoted ϖ (P).
Proof. Since the proof is very similar to the proof of Proposition 8.4, we will not
repeat it here; we only focus on the implication (i)⇒(ii), which requires some extra
explanation. Suppose that V is a vector field on Fd and that for some 1 i < j d
the functions V [xi ] and V [x j ] are different from zero. We claim that if V is weight
2 In order that the formulas which involve the weight are valid also for the zero p-vector field, we
adopt, as for homogeneous multivector fields, the convention that both the degree and weight of
the zero p-vector field are −∞.
8.2 Quadratic Poisson Structures 213
ϖi (ϖ j − 1) + ϖ (V [xi ]) = ϖ j (ϖi − 1) + ϖ (V [x j ]) ,
Since LEϖ Eϖ = [Eϖ , Eϖ ] = 0, the weighted Euler vector field is weight homoge-
neous of weight zero.
Finally, we specialize the definition of weight homogeneity to the case of Poisson
structures.
Definition 8.13. A Poisson structure π on Fd is said to be weight homogeneous if for
all F, G ∈ F (Fd ) which are weight homogeneous with respect to ϖ , their Poisson
bracket {F, G} is weight homogeneous with respect to ϖ .
Writing a Poisson structure π on Fd in the form
∂ ∂
π= ∑ xi j ∧
∂ xi ∂ x j
,
1i< jd
LEϖ [π ] = ϖ (π ) π .
∂ ∂
π= ∑ xi j ∧
∂ xi ∂ x j
, (8.12)
1i< jd
where all xi j := xi , x j , with 1 i < j d, are homogeneous polynomial functions
of degree 2 on V . Notice that a Poisson structure π on V is quadratic if and only if the
Poisson bracket of every pair of linear functions on V is a homogeneous polynomial
function of degree 2 on V .
To start with, we give two important classes of quadratic Poisson structures.
∂ ∂
π := ∑ ai j xi x j ∧
∂ xi ∂ x j
. (8.13)
1i< jd
which is zero since A is skew-symmetric. This proves the Jacobi identity for every
triple of coordinate functions, which shows, in view of Proposition 1.36, that the
bivector field π is a Poisson structure on Fd . A quadratic Poisson structure of the
form (8.13) is called a diagonal Poisson structure.
Notice that the rank of this Poisson structure is equal to the rank of the matrix A.
More precisely, for every m ∈ Fd , the rank of π at m is equal to the rank of the
matrix, obtained by removing from the matrix A the k-th line and the k-th column
for all k for which xk (m) = 0.
[E ∧ V , E ∧ V ]S = −2 E ∧ V ∧ [E , V ] = 0 ,
Proof. The equivalence between (i) and (ii) is a particular case of the equivalence
between (i) and (iv) in Proposition 8.4 with p = 2. Let π = {· , ·} be a Poisson
structure on V . The multiplication by a ∈ F∗ is a Poisson map if and only if
for all linear functions F, G ∈ V ∗ and all v ∈ V , which is equivalent to saying that
{F, G} is a homogeneous polynomial function of degree 2, for all F, G ∈ V ∗ . By
Proposition 8.4, the latter means that π is quadratic. This shows that (i) and (iii) are
equivalent.
Proof. Let p be a polynomial which satisfies (8.14), for all a, b ∈ F. The degree r of
p is at most 2, because (a, b) → p(ab) is a polynomial function (in two variables)
of total degree 2r, while the total degree of a2 p(b) + b2 p(a) is at most r + 2. But
(8.14) implies that p(0) = p(1) = p(−1) = 0, so that p is the zero polynomial.
We now establish the link between multiplicative and quadratic Poisson structures.
216 8 Higher Degree Poisson Structures
for all a ∈ F and v ∈ V . This shows that {F, G} is a homogeneous polynomial of de-
gree 2. Since this holds for all linear functions F and G on V , the Poisson structure π
is quadratic.
Proposition 8.18 takes in the smooth or holomorphic context the following form.
Proof. The proof of the proposition goes along the same lines as the proof of Propo-
sition 8.18; the only difference is that one needs to reprove Lemma 8.17 in that
context, namely, to prove that there exists no non-zero smooth or holomorphic func-
tion p satisfying (8.14). Deriving condition (8.14) three times with respect to a,
yields that if p satisfies (8.14) for all a, b ∈ F, it also satisfies bp (ab) = p (a), for
all a, b ∈ F. Plugging b = 0 in it, we obtain that p (a) = 0 for all a ∈ F, and there-
fore that p is a polynomial. According to Lemma 8.17, p is the zero polynomial.
Remark 8.20. It is clear that the only differentiability condition which we use on
the function p in Proposition 8.19 is that p is of class C3 in a neighborhood of o
in V . Therefore, Proposition 8.19 can be easily generalized to multiplicative bivector
8.2 Quadratic Poisson Structures 217
∂ ∂
π= ∑ xi j ∧
∂ xi ∂ x j
,
1i< jd
and
{x1 ◦ μ , x2 ◦ μ } = {y1 z1 , y1 z2 + z1 y2 }
= z21 {y1 , y2 } + y21 {z1 , z2 }
= z21 y21 ln(|y1 |) + y21 z21 ln(|z1 |) .
Example 8.22. Let V := Matd (F) the vector space of square matrices of size d. For
1 i, j d, we consider on V the linear function, defined for all x = (xi j )di, j=1 ∈ V
by ξi j (x) := xi j . These functions provide a set of linear coordinates on V . For i and j
as above, let εi, j := 1 if i > j and εi, j := −1 if i < j and εi, j := 0 if i = j. Then the
brackets, defined for 1 i, j, k, d, by
218 8 Higher Degree Poisson Structures
ξi j , ξk := (εi,k + ε j, )ξi ξk j
Example 8.25. Let V be a d-dimensional vector space and let V be a linear vector
field on V . As we have seen in Example 8.15, π := E ∧ V is a quadratic Poisson
structure on V . In order to compute its divergence, notice that Div(E ) = d, which
follows from (4.21), and that [E , V ] = 0, because the weight of a linear vector field
is 0. Substituted in equation (4.23) for the divergence of a wedge product, one finds
terms involves only the modular vector field of π , while the other term is unimodu-
lar. This decomposition is very useful, as we will see in Section 9.2.3.
We first recall two more facts from Section 4.4.3, namely that, given a volume
form λ on V , the star operator is the family of isomorphisms
: Xq (V ) → Ω d−q (V ) ,
Div = −1 ◦ d ◦ .
[E ∧ Φπ , π ]S = (LE π ) ∧ Φπ − E ∧ (LΦπ π ) = 0 ,
where we have used in the last step that both E and Φπ are Poisson vector fields
(with respect to π ). The modular vector field Φπ of the Poisson structure π :=
E ∧ Φπ is given by
where we have used (8.18), together with the fact that Div(Φπ ) = 0, which holds
because Φπ is itself the divergence of a bivector field. Since the quadratic Poisson
structures π and π = E ∧ Φπ are compatible, and since their modular vector fields
are proportional (see (8.19)), there exists a linear combination of π and E ∧ Φπ
which is a unimodular Poisson structure: R := π − d1 E ∧ Φπ is a Poisson structure
on V which is quadratic, compatible with π and unimodular.
Let us denote by α the differential (d − 2)-form on V , defined by α := R = ıR λ .
It is a closed form since R is unimodular. It remains to be checked that LΦπ R = 0
and that LΦπ α = 0. First, we have that LΦπ π = 0 (since Φπ is a Poisson vector
field), and we also have that LΦπ (E ∧ Φπ ) = 0 (since the vector fields E and Φπ
commute). We therefore have LΦπ R = 0. It follows from this, upon using (2) of
220 8 Higher Degree Poisson Structures
where we used in the last step that [Φπ , R]S = LΦπ R = 0 and that LΦπ λ =
Div(Φπ ) λ = 0 (see (4.20)). This finishes the proof of the announced properties
of α and of R = −1 α .
The quadratic Poisson structures whose modular vector field satisfies a genericity
assumption, stated below, are diagonal, when expressed in a system of well-chosen
coordinates. We devote the end of this section to explain and to prove this. First, we
introduce the required notions. Let V be a linear vector field on a d-dimensional
vector space V . Since V sends linear functions on V (elements of V ∗ ) to linear
functions on V , we can restrict V to V ∗ , which yields an endomorphism V (1) of V ∗ ,
V (1) : V ∗ → V ∗
(8.20)
F → V [F] .
∂ ∂
π= ∑ ai j xi x j ∧
∂ xi ∂ x j
, (8.22)
1i< jd
where (ai j ) ∈ Matd (F) is a skew-symmetric matrix. The system of linear coordinates
(x1 , . . . , xd ) is then said to be log-canonical for π .
8.2 Quadratic Poisson Structures 221
The above-mentioned converse, which yields a sufficient condition for the diago-
nalizability of a quadratic Poisson structure is given in the following proposition.
Proof. Since the modular vector field Φπ is linear, it restricts to a linear endomor-
(2)
phism Φπ of the space A2 of all homogeneous polynomial functions of degree 2
on V ,
(2)
Φπ : A2 → A2
F → Φπ [F] .
Suppose that Φπ is diagonalizable and let x1 , . . . , xd be independent eigenvectors
of Φπ , with eigenvalues a1 , . . . , ad . We use x1 , . . . , xd as linear coordinates on V .
The d+1 2 polynomial functions {xi x j | 1 i j d} form a basis of A2 , and
(2)
each one of them is an eigenvector of Φπ , because Φπ is a derivation: xi x j is an
(2)
eigenvector of Φπ , corresponding to the eigenvalue ai + a j , for all 1 i j d.
If we assume that the d+1
2 sums ai + a j | 1 i j d are all different (which
(2)
is assumption (2)), then all eigenvalues of Φπ are different, hence the eigenspace
(2)
decomposition of A2 with respect to Φπ is given by
A2 = Cxi x j . (8.23)
1i jd
We show that this implies that π is diagonalizable. To do this, we use that the mod-
ular vector field Φπ is a Poisson vector field: writing {· , ·} for π and specializing
∂F ∂ G ∂ ϕ1 ∂ ϕd−2
···
∂ x1 ∂ x1 ∂ x1 ∂ x1
∂ (F, G, ϕ1 , . . . , ϕd−2 ) .. .. ..
{F, G} := χ =χ . . . . (8.24)
∂ (x1 , . . . , xd )
∂F ∂ G ∂ ϕ1 ∂ ϕd−2
···
∂ xd ∂ xd ∂ xd ∂ xd
It leads to a map F (V ) × F (V ) → F (V ), which is a skew-symmetric biderivation
of F (V ) and which, up to a non-zero constant, does not depend on the chosen linear
coordinates x1 , . . . , xd , since (8.24) can also be written as
dF ∧ dG ∧ dϕ1 ∧ · · · ∧ dϕd−2
{F, G} = χ . (8.25)
dx1 ∧ dx2 ∧ · · · ∧ dxd
Moreover, it defines a Poisson structure on V , a fact which we will prove in this sec-
tion. When the functions ϕ1 , . . . , ϕd−2 are polynomial or weight homogeneous, then
so is the corresponding Poisson structure. In order to prove that {· , ·} defines a Pois-
son structure on V , we briefly introduce the notion of a Nambu–Poisson structure,
which generalizes the notion of a Poisson structure.
Definition 8.30. For q 2, a q-vector field {· , . . . , ·} on a finite-dimensional vector
space V is called a Nambu–Poisson structure of order q on V if it satisfies the so-
called fundamental identity:
Remark 8.31. The definition is easily generalized to arbitrary (real or complex) man-
ifolds by demanding that the q-vector field satisfies the fundamental identity for lo-
cal functions, which yields the notion of a Nambu–Poisson manifold. Replacing the
algebra of functions F (V ) by an arbitrary commutative associative algebra, leads
to the abstract notion of a Nambu–Poisson algebra, which generalizes the notion of
a Poisson algebra, as given in Definition 1.1.
for all F, G1 , G2 ∈ F (V ), which is precisely the Jacobi identity for {· , ·}. Therefore,
a Nambu–Poisson structure of order 2 is the same as a Poisson structure. In this
sense, Nambu–Poisson structures are a generalization of Poisson structures.
Given a Nambu–Poisson structure {· , . . . , ·} of order q 2 on V , one can associate
to q − 1 elements F1 , . . . , Fq−1 of F (V ), a vector field XF1 ,...,Fq−1 on V by defining,
for G ∈ F (V ):
XF1 ,...,Fq−1 [G] := {G, F1 , . . . , Fq−1 } .
It is called the Hamiltonian vector field associated to F1 , . . . , Fq−1 . The fundamental
identity (8.26) says that all Hamiltonian vector fields XF1 ,...,Fq−1 are derivations of
(F (V ), {· , . . . , ·}).
We show in the following proposition how Nambu–Poisson structures of or-
der q − 1 can be constructed from Nambu–Poisson structures of order q.
Proposition 8.33. Let {· , . . . , ·} be a Nambu–Poisson structure of order q 3 on a
finite-dimensional vector space V . Given H ∈ F (V ), define a (q − 1)-vector field
on V by
{· , . . . , ·}H := {· , . . . , ·, H} .
Then {· , . . . , ·}H is a Nambu–Poisson structure of order q − 1 on V .
Proof. The fundamental identity (8.26) for {· , . . . , ·}H takes the form
which is precisely (8.27), since the skew-symmetry of {· , . . . , ·} implies that the last
term in (8.28) is equal to zero.
We show in the following proposition that for d-vector fields on a d-dimensional
vector space V , the fundamental identity is automatically satisfied.
Proposition 8.34. Let V be a d-dimensional vector space, with d 2. Every d-
vector field on V is a Nambu–Poisson structure of order d on V .
Proof. Let {· , . . . , ·} be a d-vector field on V . We need to prove that {· , . . . , ·} sat-
isfies the fundamental identity (8.26). To do this, we consider the multilinear map
N : F (V )2d−1 → F (V ), which is defined by
224 8 Higher Degree Poisson Structures
N (F1 , . . . , Fd−1 , G1 , . . . , Gd )
:= {F1 , . . . , Fd−1 , {G1 , . . . , Gd }}
d
− ∑ {G1 , . . . , Gi−1 , {F1 , . . . , Fd−1 , Gi } , Gi+1 , . . . , Gd } ,
i=1
F (V )d → F (V )
(8.29)
(G1 , . . . , Gd ) → N (F1 , . . . , Fd−1 , G1 , . . . , Gd )
∂ ∂
{· , . . . , ·} = χ ∧···∧ , (8.31)
∂ x1 ∂ xd
so that (8.30), which we need to prove, takes the form
d
∂
{F1 , . . . , Fd−1 , χ } = χ ∑ ∂ xi {F1 , . . . , Fd−1 , xi } . (8.32)
i=1
∂ F1 ∂ Fd−1 ∂χ
···
∂ x1 ∂ x1 ∂ x1
{F1 , . . . , Fd−1 , χ } = χ .
.. .. ..
. .
∂ F1 ∂ Fd−1 ∂χ
···
∂ xd ∂ xd ∂ xd
dF1 ∧ · · · ∧ dFd−1 ∧ dχ
=χ
dx1 ∧ · · · ∧ dxd
d χ dF1 ∧ · · · ∧ dFd−1
= (−1)d−1 χ
dx1 ∧ · · · ∧ dxd
d i ∧ · · · ∧ dxd
d Ci dx1 ∧ · · · ∧ dx
= (−1) d−1
χ ∑ dx1 ∧ · · · ∧ dxd
i=1
8.3 Nambu–Poisson Structures 225
d
∂ Ci
= χ ∑ (−1)d−i ,
i=1 ∂ xi
where
∂ ∂ ∂
Ci := χ (dF1 ∧ · · · ∧ dFd−1 ) ,..., ,...,
∂ x1 ∂ xi ∂ xd
∂ ∂ ∂ ∂
= χ (dF1 ∧ · · · ∧ dFd−1 ∧ dxi ) ,..., ,..., ,
∂ x1 ∂ xi ∂ xd ∂ xi
∂ ∂
= (−1) χ (dF1 ∧ · · · ∧ dFd−1 ∧ dxi )
d−i
,...,
∂ x1 ∂ xd
= (−1)d−i {F1 , . . . , Fd−1 , xi } .
∂ (F, G, ϕ1 , . . . , ϕd−2 )
{F, G} := χ , (8.33)
∂ (x1 , . . . , xd )
∂ (F, G, ϕ1 , . . . , ϕd−2 )
{F, G} := {F, G, ϕ1 , . . . , ϕd−2 } = χ . (8.34)
∂ (x1 , . . . , xd )
It follows at once from Eq. (8.33) that, if the functions χ , ϕ1 , . . . , ϕd−2 are polyno-
mial functions on V , then the Poisson structure π is a polynomial Poisson struc-
ture on V . If, moreover, for some fixed weights ϖ1 , . . . , ϖd of the linear coordinates
x1 , . . . , xd on V , the functions χ , ϕ1 , . . . , ϕd−2 are weight homogeneous, then it also
follows from this formula that the Poisson structure π is a weight homogeneous
Poisson structure on V , of weight
d−2 d
ϖ (π ) = ϖ ( χ ) + ∑ ϖ (ϕi ) − ∑ ϖ i .
i=1 i=1
∂ ∂
{· , . . . , ·} = χ ∧···∧ ,
∂ x1 ∂ xd
and the formula for the Poisson structure is given by
∂ (F, G, ϕ1 , . . . , ϕd−2 )
{F, G} := χ ,
∂ (x1 , . . . , xd )
dF1 ∧ · · · ∧ dFd
{F1 , . . . , Fd } := χ ,
λ
for all (local) functions F1 , . . . , Fd on M. In particular, if we fix d − 1 functions
χ , ϕ1 , . . . , ϕd−2 on M and a volume form λ , then a Poisson structure π on M is
defined by
dF ∧ dG ∧ dϕ1 ∧ · · · ∧ dϕd−2
{F, G} := χ ,
λ
for all (local) functions F, G on M. The functions ϕ1 , . . . , ϕd−2 are Casimir functions
of π and the rank of π is two at all points of M, except at the zeros of χ and at
the points where the differentials of the latter Casimir functions are dependent. In
this case, a simple proof of the fact that π is a Poisson structure can be given by
extending, at points where the rank is two, the Casimir functions ϕ1 , . . . , ϕd−2 to a
system of local coordinates.
8.4 Transverse Poisson Structures to Adjoint Orbits 227
In this section, we recall a few facts which we will use about complex semi-simple
Lie algebra (see [99, 190] for details and proofs). Let (g, [· , ·]) be a semi-simple Lie
algebra over F = C; we denote the algebra of polynomial functions or of holomor-
phic functions on g by F (g). The Killing form of g is the symmetric bilinear form
on g, defined for all x, y ∈ g by
where we recall that adx : g → g is the linear map, defined by adx z := [x, z], for
all z ∈ g. For a subspace n of g, the orthogonal complement of n in g with respect to
the Killing form is denoted by n⊥ . For x ∈ g, we denote g(x) := {y ∈ g | [x, y] = 0},
the centralizer of x in g. An element x ∈ g is said to be nilpotent if the endomorphism
adx of g is nilpotent, i.e., if there exists a k ∈ N, such that adkx = 0. According to the
Jacobson–Morozov theorem, if e ∈ g is nilpotent, then there exist h, f ∈ g such that
g= gi ,
i∈Z
where gi is the eigenspace of adh which corresponds to the eigenvalue i. For exam-
ple, e ∈ g2 and f ∈ g−2 . For x ∈ gi and y ∈ g j , we have an inclusion adx ◦ ady (gk ) ⊂
gi+ j+k , for every k, so that
gi | g j = 0 if i + j = 0 . (8.36)
where each Vn j is a simple s-module, with adh -eigenvalues (also called adh -weights)
n j , n j − 2, n j − 4, . . . , −n j , so that dimVn j = n j + 1. It follows that
s
∑ n j = dim g − s . (8.38)
j=1
Moreover, s = dim g(e), since the centralizer g(e) of e is generated by the vectors of
highest adh -weight of each Vn j .
We show in this section that every S-triple (h, e, f ) of g leads to a system of linear
coordinates on g, centered at e, such that the Lie–Poisson structure on g is weight
homogeneous with respect to some weight vector ϖ , which has a natural Lie alge-
braic interpretation: up to a shift of 2, it consists of the weights of g as an s-module.
For every t ∈ C∗ , choose a complex number λt such that exp(−λt ) = t. Such a
choice cannot be made in such a way that λt depends in a continuous way on t, but
that will be irrelevant for what follows. We denote the adjoint group of g by G and
we consider the map
λ : C∗ → G
t → exp(λt h) .
For a fixed t, the map Adλ (t) : g → g is well-defined (independent of the choice made
for λt ) and it acts on each of the eigenspaces gi of adh as a homothecy with ratio t −i ,
∇z F | u = dz F, u ,
If x ∈ gi and y ∈ g j , with i + j = −2, then (8.36) implies that e | [x, y] = 0, so that
ϖ ( Fx , Fy g ) − ϖ (Fx ) − ϖ (Fy ) = ϖ (F[x,y] ) − ϖ (Fx ) − ϖ (Fy )
= i + j + 2 − (i + 2) − ( j + 2) = −2 .
This result extends to the case i + j = −2, since then ϖ (F[x,y] ) = i + j + 2 = 0, which
is the weight of the constant function e | [x, y], and then ϖ ({Fx , Fy }g ) − ϖ (Fx ) −
ϖ (Fy ) = −(i + 2) − ( j + 2) = −2. This shows, according to (8.11), that the Lie–
Poisson structure on g is weight homogeneous of degree −2 with respect to the
weight vector (i1 + 2, . . . , id + 2). Notice that this weight vector contains, in general,
both positive and negative integers.
Proof. First, let us point out that an adh -invariant complementary subspace n to g(e)
exists, because the centralizer g(e) is generated by the highest weight vectors of
each Vn j in the decomposition (8.37) of g. Also, it is clear that
Te g = Te O ⊕ Te N ,
Since =1 ν
∑2r + ∑si=1 ni
is the sum of all adh -weights, it is zero, which yields in
combination with (8.38)
2r s
∑ (ν + 1) = 2r − ∑ ni = 2r − (dim g − s) = 0 .
=1 i=1
This shows that every term of det(D), restricted to N, is of weight zero, hence det(D)
is constant; since det(D(e)) = 0, this constant is different from zero and the trans-
verse Poisson structure is polynomial.
To finish, we show that the Poisson structure {· , ·}N is weight homogeneous of
weight −2 with respect to ϖ = (n1 + 2, . . . , ns + 2). According to (8.42) we need to
show that for all 1 i, j s the functions Ai j and (BD−1 B )i j are weight homoge-
neous of degree ϖ (qi ) + ϖ (q j ) − 2 = ni + n j + 2. For Ai j this is clear, since A is part
of the Poisson matrix of the Lie–Poisson structure on g, which we have seen to be
weight homogeneous of weight −2. Similarly, we have that ϖ (Bim ) = ni + νm + 2.
Since
ϖ (Bim D−1 −1
m B j ) = ni + n j + νm + ν + 4 + ϖ (Dm ) ,
232 8 Higher Degree Poisson Structures
ϖ (D−1
m ) = −νm − ν − 2 . (8.43)
Dαβ
A typical term of the element D−1 m of the inverse matrix D
−1 is of the form
det(D) ,
where Dαβ = Dα1 β1 . . . Dα2r−1 β2r−1 , with
{α1 , . . . , α2r−1 } = {1, . . . , 2r} \ {} and {β1 , . . . , β2r−1 } = {1, . . . , 2r} \ {m} .
Since det(D) has weight zero, we need to show that Dαβ has weight −νm − ν − 2,
which is done in precisely the same way as the above proof that det(D) has weight
zero. This gives us (8.43).
8.5 Notes
For higher order Poisson structures, starting with quadratic Poisson structures, there
is no general theory and there is no immediate interpretation, as there is in the
case of constant Poisson structures (in terms of bivectors) and in the case of lin-
ear Poisson structures (in terms of Lie brackets). For a few subclasses of quadratic
Poisson structures there is, however, a quite general theory and an interpretation.
Namely, multiplicative quadratic Poisson structures are closely related to Poisson–
Lie groups: the datum of a skew-symmetric r-matrix on an associative algebra leads
to multiplicative quadratic Poisson structure, which makes the set of invertible el-
ements into a Poisson–Lie group; see Section 11.1.6 for details. It also leads to a
cubic Poisson structure, whose virtue seems at this point still very mysterious. Pois-
son structures coming from r-matrices or R-matrices are the subject of Chapter 10;
see Section 10.3.
See Dufour–Zung [63] for an extensive account on Nambu–Poisson structures.
Transverse Poisson structures to general adjoint orbits in a semi-simple Lie alge-
bra, with special emphasis to transverse Poisson structures to subregular orbits, are
studied in Damianou–Sabourin–Vanhaecke [56].
Chapter 9
Poisson Structures in Dimensions Two and Three
Low-dimensional Poisson manifolds are often used as toy models, to obtain a better
understanding of the theory of Poisson manifolds, as well as to illustrate their unex-
pected complexity. In the two-dimensional case, for example, describing all Poisson
structures on the affine space F2 is deceivingly simple, because the Jacobi identity is
satisfied for all bivector fields; however, their local classification is non-trivial, and
has up to now only been accomplished under quite strong regularity assumptions
on the singular locus of the Poisson structure, which can be identified with the zero
locus of a local function on the manifold. As a result, the study of Poisson structures
in two dimensions already takes us to the non-trivial singularity theory of functions
of two variables!
Dimension three is the smallest dimension in which the Jacobi identity for a
bivector field is not always satisfied. In this dimension, the Jacobi identity can be
stated as an integrability condition of a distribution, which eventually leads to the
symplectic foliation, or as the integrability condition of a differential one-form, dual
to the Poisson structure with respect to a volume form (assuming that the manifold
is orientable). Despite the extra complexity which comes from the extra dimension,
Poisson structures in dimension three have one key property in common with Pois-
son structures in dimension two: their rank is equal to two (except when the Poisson
structure is trivial). Many results about three-dimensional Poisson manifolds are es-
sentially true because the Poisson structure is of rank two. A key example of such a
result is the characterization of a regular Poisson structure of rank two in terms of
its symplectic foliation.
Poisson structures in dimensions two and three are presented in different sections
(Sections 9.1 and 9.2). In both cases, we pay special attention to the global/local and
geometrical/algebraic aspects of these Poisson structures and we prove at the end of
each section a (partial) classification result.
Unless otherwise stated, F is an arbitrary field of characteristic zero.
∂ ∂ ∂
P= ∑ Pi jk ∧ ∧
∂ xi ∂ x j ∂ xk
,
1i< j<kd
∂
Fig. 9.1 The singular locus of the Poisson structure π = χ ∂x ∧ ∂∂y is the zero locus of χ , i.e., the
curve defined by χ = 0.
values zero and two at the points of M, and the singular locus of (M, π ) coincides
with the set of points where π vanishes (assuming that π is non-trivial).
Finally, consider F2 , with standard coordinates (x, y), equipped with its algebra
of functions F (F2 ); recall that F (F2 ) denotes the algebra of polynomial functions
on F2 if F is an arbitrary field, but can also be the algebra of smooth or holomorphic
functions on F2 if F = R or F = C. Every skew-symmetric biderivation on F (F2 )
or bivector field on F2 (i.e., every Poisson structure on F2 ) is of the form
∂ ∂
π=χ ∧ , where χ ∈ F (F2 ) , (9.1)
∂x ∂y
and by the above, every such π is a Poisson structure on F2 . Hence, there is a one-
to-one correspondence between the following three objects: the algebra F (F2 ) of
all functions on F2 , the vector space X2 (F2 ) of all bivector fields on F2 and the set
of all Poisson structures on F2 . The singular locus of π coincides with the zero locus
of χ , which is the curve {χ = 0} ⊂ F2 . See Fig. 9.1.
Let M be a (real or complex) manifold of dimension two and let F (M) denote its
algebra of smooth or holomorphic functions. Let (U, (x, y)) be a coordinate chart
of M. According to (1.23), every bivector field on U can be written as
∂ ∂
π=χ ∧ , where χ = π [x, y] ∈ F (U) . (9.2)
∂x ∂y
We have seen in Section 9.1.1 that every bivector field on U is a Poisson structure.
The rank of π is two at all points m ∈ U where χ (m) = 0 and is zero at all other
points of U.
236 9 Poisson Structures in Dimensions Two and Three
Let us show that, although the Jacobi identity is trivially satisfied in dimension
two, and although all Poisson structures in dimension two are locally given by the
simple formula (9.2), two Poisson structures in dimension two may, both from the
algebraic and geometric points of view, be very different in the neighborhood of
their singularities.
To illustrate this, let (M, π ) be a two-dimensional Poisson manifold and let m be
an arbitrary point of M. We write π in a neighborhood of m in M as
∂ ∂
π=χ ∧ ,
∂x ∂y
where x, y are local coordinates, vanishing at m, and where χ := π [x, y]. We distin-
guish the following three cases:
(1) If χ (m) = 0, which means that the rank of π at m is two, then there exists,
according to Weinstein’s splitting theorem (Theorem 1.25), a coordinate neighbor-
hood U of m in M, with coordinates x, y, centered at m, such that, on U,
∂ ∂
π= ∧ .
∂x ∂y
∂ ∂ ∂ ∂
π1 := (x2 − y2 ) ∧ and π2 := x2 ∧ .
∂x ∂y ∂x ∂y
Since the singular locus of π1 consists of the pair of lines y = ±x, while the singular
locus of π2 consists of the (double) line x2 = 0, it is clear that these Poisson struc-
tures are (even locally, in the neighborhood of a point of their singular locus) not
isomorphic. This simple example illustrates the fact that there are at least as many
non-isomorphic Poisson structures on the plane, as there are non-isomorphic alge-
braic curves in the plane. In fact there are more. In order to see this, consider the
following two Poisson structures:
∂ ∂ ∂ ∂
π1 := (x2 − y2 ) ∧ , and π3 := 2(x2 − y2 ) ∧ .
∂x ∂y ∂x ∂y
Since they are proportional, they have the same zero locus, but they are not isomor-
phic as Poisson structures, a fact which can easily be checked by using Proposi-
tion 8.8.
9.1 Poisson Structures in Dimension Two 237
As we will see in the next section, there is a partial local classification of Poisson
structures on C2 (or R2 ), under the strong hypothesis that the singular locus of the
Poisson structure admits a simple singularity (at the origin).
In this section, we consider the classification problem for Poisson structures in di-
mension two. Since the Jacobi identity is trivially satisfied in dimension two, the
classification amounts to the simpler problem of classifying bivector fields in di-
mension two, yet only partial results are known. The classification theorem which
we will present below has the following restrictions:
(1) The Poisson structures and the coordinate transformations which are consid-
ered are formal;
(2) The Poisson structures which are considered, are assumed to have a singular
locus which admits a simple singularity at the origin;
(3) The ground field is C.
As is made explicit in the notes at the end of the chapter, some of these restrictions
can be weakened.
We first give an overview of the results and the ideas which will be presented in
this section; full proofs will be given in Subsections A, B, C and D below. We begin
by giving the definition of a formal Poisson structure on C2 .
Definition 9.1. A formal Poisson structure on C2 is a Poisson bracket on the algebra
of all formal power series in two variables C[[x, y]].
As we show in Subsection 9.1.3.1, every formal Poisson structure on C2 is of the
form α ∂∂x ∧ ∂∂y , with α ∈ C[[x, y]], and every such biderivation is a formal Poisson
structure on C2 .
By definition, a formal coordinate transformation of C2 is an algebra isomor-
phism Φ : C[[x, y]] → C[[x, y]] such that the formal power series Φ (x) and Φ (y) have
no constant term. Two formal power series α , β ∈ C[[x, y]] are said to be equivalent
if there exists a formal coordinate transformation Φ of C2 , such that Φ (α ) = β . The
following formal version of the inverse function theorem holds: if Φ is an algebra
homomorphism such that the formal power series Φ (x) and Φ (y) have no constant
term and whose Jacobian
⎛ ⎞
∂ Φ (x) ∂ Φ (x)
⎜ ∂x ∂y ⎟
Jac Φ := det ⎜⎝ ∂ Φ (y) ∂ Φ (y) ⎠
⎟
∂x ∂y
is invertible, i.e., its constant term is different from zero, then Φ is a formal coordi-
nate transformation.
238 9 Poisson Structures in Dimensions Two and Three
∂ ∂
Φ∗ π = (Φ∗ π )[x, y] ∧ ,
∂x ∂y
it follows that, if we write π in the form π = α ∂∂x ∧ ∂∂y , then Φ∗ π takes the form
∂ ∂
Φ∗ π = Φ −1 (α Jac(Φ )) ∧ . (9.3)
∂x ∂y
∂ ∂
χ (1 + ψ ) ∧ ,
∂x ∂y
∂ ∂
χ (1 + ψ ) ∧ ,
∂x ∂y
where χ is one of the weight homogeneous polynomials which appear in Arnold’s
classification theorem of formal power series with a simple singularity (Theo-
rem 9.7) and ψ ∈ C[x, y] is an arbitrary weight homogeneous polynomial of weight
ϖ (χ ) − ϖ1 − ϖ2 . For each χ in Arnold’s list, one easily determines all possibili-
ties for ψ , which leads to the following classification theorem of formal Poisson
structures on C2 , which admit a simple singularity at the origin.
Theorem 9.4. Let π = α ∂∂x ∧ ∂∂y be a formal Poisson structure on C2 , where
α ∈ C[[x, y]] is a formal power series, which has a simple singularity at the ori-
gin. Then, up to a non-zero multiplicative constant, π is formally isomorphic to a
Poisson structure of the form η ∂∂x ∧ ∂∂y , where η ∈ C[x, y] is one the polynomials
which appear in the third column of the following table.
Type k η
A2k k1 x2 + y2k+1
A2k−1 k1 x2 + y2k 1 + λ yk−1
D2k k2 x2 y + y2k−1 1 + λ x + μ yk−1
2
D2k+1 k2 x y + y2k (1 + λ x)
E6 x 3 + y4
3
E7 x + xy3 1 + λ y2
E8 x 3 + y5
Remark 9.5. The Poisson structures which appear in the classification theorem are
not claimed to be all non-isomorphic. Special care has to be taken with respect to the
phrase “up to a non-zero multiplicative constant”, which appears in its statement. In
general, a formal Poisson structure π on C2 is formally isomorphic to every Poisson
structure of the form λ π , with λ ∈ C∗ , but there are exceptions. We already gave
in Section 9.1.2 an example of two proportional quadratic Poisson structures which
are non-isomorphic. In order to make a more general statement, let χ be a weight
homogeneous polynomial of C[x, y], with respect to the weight vector (ϖ1 , ϖ2 ) and
240 9 Poisson Structures in Dimensions Two and Three
∂ ∂
(Φa )∗ π = aϖ1 +ϖ2 −ϖ (χ ) χ ∧ .
∂x ∂y
In this subsection, we recall some basic notions about formal power series (in two
variables) and we introduce some concepts which are needed to state the theorem of
Arnold about the classification of formal power series with a simple singularity.
A formal power series α ∈ C[[x, y]] can be written in a unique way as α =
∑i, j∈N αi j xi y j , so the datum of a family (αi j )i, j∈N in C is equivalent to the datum
of a formal power series in C[[x, y]]. This elementary fact is important when deal-
ing with the question whether certain operations yield well-defined power series: a
formal power series is well-defined if for every i, j ∈ N the coefficient ai j is well-
defined. For example, the product of two power series is well-defined, because each
term of the product is given by a finite sum. Similarly, if α ∈ C[[x, y]] is a formal
power series without constant term, then
αi
exp(α ) := ∑
i∈N i!
is a well-defined power series. A formal power series α ∈ C[[x, y]] can be evaluated
at (x, y) = (0, 0), which yields the constant term of the series, denoted α (0, 0), but
can in general not be evaluated at other values of x, y. With the standard addition and
product of power series, it is clear that C[[x, y]] forms an associative algebra, whose
invertible elements are the series α for which α (0, 0) = 0.
We define the order of a formal power series α ∈ C[[x, y]] as the minimum of
the total degrees of each of its terms, denoted ord(α ). When we are in a weight
homogeneous context, we will also consider the order of a power series α with
respect to a weight vector ϖ = (ϖ1 , ϖ2 ), where ϖ1 and ϖ2 are the weights of the
9.1 Poisson Structures in Dimension Two 241
variables x and y (see Section 8.1.3). Then ordϖ (α ), the order of α with respect to
ϖ , is by definition the minimum of the weights of each of the terms of α .
As we said previously, a formal coordinate transformation of C2 is an algebra iso-
morphism Φ : C[[x, y]] → C[[x, y]] so that the constant terms of Φ (x) and of Φ (y) are
zero. Also, recall that two formal power series α , β ∈ C[[x, y]] are said to be equiva-
lent if there exists a formal coordinate transformation Φ of C2 such that Φ (α ) = β .
Notice that if α and β are two equivalent formal series of C[[x, y]], they have the
same constant term.
Definition 9.6. A formal power series α ∈ C[[x, y]] satisfying α (0, 0) = 0 is said to
have a simple singularity at the origin, if a small enough neighborhood of α (for the
Whitney topology, see [17]) intersects only a finite number of orbits associated to
the equivalence of formal power series.
We are now able to state Arnold’s classification in the case of formal power series in
C[[x, y]], with a simple singularity. The names of the singularity types are borrowed
from the classification theory of simple Lie algebras; see [147] for an explanation.
Theorem 9.7 (Arnold’s theorem). Let α ∈ C[[x, y]] be a formal power series with-
out constant term, which has a simple singularity at (0, 0). Then α is equivalent to
one of the weight homogeneous polynomials χ which appear in the third column of
the following table.
Type k χ
Ak k1 x2 + yk+1
Dk k4 x2 y + yk−1
E6 x 3 + y4
E7 x3 + xy3
E8 x 3 + y5
where α , β ∈ C[[x, y]] are arbitrary. We show in the following proposition that there
are no other skew-symmetric k-derivations of C[[x, y]] (k 1).
Proposition 9.8. The spaces of skew-symmetric multi-derivations of C[[x, y]] are
given by
∂ ∂
X1 C[[x, y]] = α +β | α , β ∈ C[[x, y]] ,
∂x ∂y
∂ ∂
X2 C[[x, y]] = α ∧ | α ∈ C[[x, y]] ,
∂x ∂y
Xk C[[x, y]] = {0}, for all k 3 .
Proof. Let us prove this proposition for the case of the derivations of C[[x, y]],
the case of the skew-symmetric biderivations and skew-symmetric k-derivations
of C[[x, y]] (for k 3) being analogous. To do that, we first point out that ev-
ery derivation W of C[[x, y]] satisfies the following property: if γ ∈ C[[x, y]], then
ord(W [γ ]) is either −∞ (i.e. W [γ ] = 0) or ord(W [γ ]) ord(γ ) − 1. This can be
done easily by recursion on ord(γ ), first in the particular case where γ ∈ C[[x]]
or γ ∈ C[[y]], and secondly for an arbitrary γ ∈ C[[x, y]], using the fact that every
γ ∈ C[[x, y]], with ord(γ ) 1, can be written in at least one of the following two
forms:
• γ = xγ1 + γ2 , with γ1 ∈ C[[x, y]], ord(γ1 ) = ord(γ ) − 1 and γ2 ∈ C[[y]] ;
• γ = yγ1 + γ2 , with γ1 ∈ C[[x, y]], ord(γ1 ) = ord(γ ) − 1 and γ2 ∈ C[[x]] .
Now, let V ∈ X1 (C[[x, y]]) and consider the formal power series α := V [x] and
β := V [y]. We prove that V = α ∂∂x + β ∂∂y , by showing that W := V − α ∂∂x − β ∂∂y
is zero. First, W is a derivation of C[[x, y]] and W [x] = W [y] = 0, hence W vanishes
on every polynomial in C[x, y]. Let now γ ∈ C[[x, y]] be an arbitrary formal power
series. Suppose that W [γ ] = 0 and let s := ord(W [γ ]). We decompose the formal
power series γ as follows: γ = γs + γ̂s , where γs ∈ C[x, y] is a polynomial of total
degree at most s + 1 and γ̂s ∈ C[[x, y]] is a formal power series of order at least s + 2.
Since W vanishes on all polynomials, W [γ ] = W [γs ] + W [γ̂s ] = W [γ̂s ], so that
which is impossible. Therefore, W [γ ] = 0 for all formal power series γ and the
equality V = α ∂∂x + β ∂∂y follows.
is zero, which implies that every such biderivation defines a formal Poisson struc-
ture on C2 . Thus, there is a one-to-one correspondence between the following three
objects: the algebra C[[x, y]] of all formal power series in two variables, the vector
space X2 (C[[x, y]]) of all skew-symmetric biderivations of C[[x, y]] and the set of
all formal Poisson structures on C2 . Notice that the series α in (9.5) is given by
α = π [x, y], so that a formal Poisson structure π on C2 is completely determined by
its value on the variables x and y. Given the above one-to-one correspondence, it is
natural to transport the property that a function has a simple singularity, to the case
of a formal Poisson structure on C2 .
Definition 9.9. Let π = α ∂∂x ∧ ∂∂y be a formal Poisson structure of C2 , satisfying
α (0, 0) = 0. If α admits a simple singularity at the origin, then π is said to admit a
simple singularity at the origin.
To finish this subsection, we introduce a certain class of formal coordinate trans-
formations of C2 , which are constructed from derivations of C[[x, y]]. Let V be a
derivation of C[[x, y]] and assume that for every β ∈ C[[x, y]] the series V [β ] is ei-
ther zero or of order larger than the order of β , with respect to a weight vector ϖ .
For every k 1 and for every β ∈ C[[x, y]], the formal power series V k [β ] is either
zero or of order at least k + ordϖ (β ). Therefore, for every i, j ∈ N, the coefficient
of xi y j in the infinite sum
1
exp (V ) (β ) := ∑ k! V k [β ]
k∈N
Clearly, the formal power series exp (V ) (x) and exp (V ) (y) vanish at the ori-
gin. We show that exp (V ) is an algebra homomorphism; i.e., that exp(V )(β γ ) =
exp(V )(β ) exp(V )(γ ), for all β , γ ∈ C[[x, y]]. Since V is a derivation of C[[x, y]],
we have for all β , γ ∈ C[[x, y]],
i
i
V (β γ ) = ∑
i
V i−k (β ) V k (γ ) ,
k=0 k
so that
i
1 i 1 i
exp(V )(β γ ) = ∑ V (β γ ) = ∑ ∑ V i−k (β ) V k (γ )
i∈N i! i∈N k=0 i! k
i
1
= ∑ ∑ (i − k)! k! V i−k (β ) V k (γ )
i∈N k=0
1
= ∑ j! k!
V j (β ) V k (γ ) = exp(V )(β ) exp(V )(γ ) .
j,k∈N
244 9 Poisson Structures in Dimensions Two and Three
The aim of this subsection is to prove Proposition 9.3. Throughout the subsection,
ϖ = (ϖ1 , ϖ2 ) is a fixed weight vector; weight homogeneity of a polynomial and the
order of a formal power series are to be understood as being with respect to ϖ . As
before, Eϖ denotes the weighted Euler derivation and we denote |ϖ | := ϖ1 + ϖ2 .
We start with three technical lemmas.
Proof. According to (9.3), (Φt )∗ π = γ ∂∂x ∧ ∂∂y , where γ ∈ C[[x, y]] is given by
Φt−1 (χ ) = λt χ ⇐⇒ μt χ = Φt (χ ) ,
9.1 Poisson Structures in Dimension Two 245
1
where μt := is a formal power series in C[[x, y]] (depending on t), since
Φt (λt )
Φt (λt )(0, 0) = λt (0, 0) = 0 for all t. Our problem is therefore reduced to constructing
a one-parameter family (μt )t∈C of formal power series in C[[x, y]], such that Φt (χ ) =
μt χ , such that each coefficient of the power series μt depends polynomially on t, and
such that μ0 = 1 and μt (0, 0) = 0 for all t ∈ C. Under the latter hypothesis on μt , we
have the following equivalences:1
·
1
μt χ = Φt (χ ) ⇐⇒ Φt (χ ) = 0
μt
μ̇t 1
⇐⇒ − 2 Φt (χ ) + Φt (α Eϖ [χ ]) = 0
μt μt
−1 μ̇t
⇐⇒ α Eϖ [χ ] = Φt χ
μt
μ̇t
⇐⇒ ϖ (χ ) α = Φt−1 ,
μt
where we have used in the second equivalence that (Φt )t∈C is the formal flow of
α Eϖ , and the weighted Euler formula (8.10) in the last equivalence. Our problem is
thereby reduced to solving the linear differential equation
is the unique solution. It is for each t a well-defined formal power series (since the
t
formal power series Φτ (α )dτ does not have a constant term) whose coefficients
0
depend polynomially on t, and it satisfies μt (0, 0) = 1, for all t ∈ C. This proves the
announced existence of the family (μt )t∈C , and hence of the family (λt )t∈C , which
proves that χ divides (Φt )∗ π , for all t.
∂ ∂
Lemma 9.11. Let π = α ∂x ∧ ∂y be a formal Poisson structure on C2 and let us
write α as
α = χs0 +|ϖ | + · · · + χsr +|ϖ | + α̂ ,
where s0 < s1 < · · · < sr ∈ Z and where χs0 +|ϖ | , . . . , χsr +|ϖ | are the first r + 1
non-zero weight homogeneous terms of the formal power series α , of weight
s0 + |ϖ |, . . . , sr + |ϖ | and α̂ is a formal power series in C[[x, y]] of order at least
with β̂ ∈ C[[x, y]] and of order at least sr + |ϖ | + 1. Stated briefly, Φ eliminates the
term χsr +|ϖ | without modifying the terms of lower weight.
Proof. Let π = α ∂∂x ∧ ∂∂y and suppose that there exists a weight homogeneous
derivation V , satisfying (9.8), where χs0 +|ϖ | and χsr +|ϖ | are the parts of α of weight
s0 + |ϖ |, respectively of weight sr + |ϖ |. Then V1 and V2 are weight homogeneous
polynomials of weight ϖ (V1 ) = sr − s0 + ϖ1 and ϖ (V2 ) = sr − s0 + ϖ2 . It implies
that Φ := exp(V ) is a well-defined formal coordinate transformation of C2 ; explic-
itly, Φ (x) and Φ (y) are of the form
Φ (x) = x + V1 + Vˆ1 ,
(9.10)
Φ (y) = y + V2 + Vˆ2 ,
where Vˆi is a formal power series in C[[x, y]] with ordϖ Vˆi > sr − s0 + ϖi for i =
1, 2, so that
∂ V1 ∂ V2
ϖ =ϖ = sr − s0 ,
∂x ∂y
∂ Vˆ1 ∂ Vˆ2
ordϖ > sr − s0 , ordϖ > sr − s0 ,
∂x ∂y
and
1 + ∂ V1 + ∂ V1 ∂ V1 ∂ Vˆ1
ˆ
+
∂x ∂x ∂y ∂y
Jac Φ = .
∂ V2 ∂ Vˆ2 ∂ V2 ∂ V2
ˆ
1+
∂x + ∂x ∂y
+
∂y
It follows that the terms of α Jac Φ , whose weights are at most sr +|ϖ |, are precisely
the following:
∂ V1 ∂ V2
χs0 +|ϖ | , . . . , χsr−1 +|ϖ | , χsr +|ϖ | + χs0 +|ϖ | + χ .
∂x ∂ y s0 +|ϖ |
9.1 Poisson Structures in Dimension Two 247
Let β be the formal power series, defined by Φ∗ π = β ∂∂x ∧ ∂∂y . We denote its term
of weight i by ψi . According to (9.3), Φ (β ) = α Jac Φ , with Φ of the form (9.10).
The terms of Φ (β ), and hence of β , of weight i, smaller than sr + |ϖ |, coincide with
the corresponding terms of α Jac Φ . Specifically, all are zero except for
Recall that V satisfies (9.8). Applied to x and y, this condition means that
∂ V1 ∂ V2
V χs0 +|ϖ | − + χs0 +|ϖ | = χsr +|ϖ | ,
∂x ∂y
which leads, according to equation (9.12), to ψsr +|ϖ | = 0. Combined with (9.11),
this shows that β is indeed of the desired form (9.9).
The following lemma is an analog of Euler’s weighted formula for bivector fields
on C2 .
Lemma 9.12. Let ψ ∈ C[x, y] be a polynomial and let π = χ ∂∂x ∧ ∂∂y be a Poisson
structure on C2 . If ψ and π are weight homogeneous with respect to some weight
vector ϖ = (ϖ1 , ϖ2 ), then
Lψ Eϖ π = (ϖ (χ ) − |ϖ | − ϖ (ψ )) ψ π . (9.13)
Proof (of Proposition 9.3). Let π = χβ ∂∂x ∧ ∂∂y be a formal Poisson structure on C2 ,
where χ ∈ C[x, y] and β ∈ C[[x, y]] are as stated. Since β verifies β (0, 0) = 1, we
can write β = 1 + ψs + βs , where ψs ∈ C[x, y] is a weight homogeneous polynomial
of weight s and βs ∈ C[[x, y]] is a formal power series of order at least s + 1. Let
us first show that s can be assumed to be at least equal to ϖ (χ ) − |ϖ |. Assume that
s < ϖ (χ ) − |ϖ | and consider the Lie derivative
∂ ∂ ∂ ∂
Lψs Eϖ χ ∧ = (ϖ (χ ) − |ϖ | − s) χψs ∧ ,
∂x ∂y ∂x ∂y
We are finally ready to give the proof of the classification theorem of formal Poisson
structures on C2 , which admit a simple singularity.
Proof (of Theorem 9.4). Let π = α ∂∂x ∧ ∂∂y be a formal Poisson structure on C2 ,
where α ∈ C[[x, y]] is a formal power series, which has a simple singularity at the
origin. According to Arnold’s theorem (Theorem 9.7) there exists a formal coordi-
nate transformation Φ , such that Φ −1 (α ) = χ , where χ is a weight homogeneous
polynomial, with respect to some weight vector ϖ = (ϖ1 , ϖ2 ). According to (9.3),
Φ∗ π = χβ ∂∂x ∧ ∂∂y , where β ∈ C[[x, y]] is a formal power series, with β (0, 0) = 0.
Using Proposition 9.3, we can conclude that, up to a constant, π is isomorphic to
χ (1 + ψ ) ∂∂x ∧ ∂∂y , where χ is one of the weight homogeneous polynomials which
appear in Theorem 9.7 and where ψ is a weight homogeneous polynomial of weight
ϖ (χ ) − |ϖ |. It suffices now to consider χ in the list of weight homogeneous poly-
nomials of Theorem 9.7 and construct the most general weight homogeneous poly-
nomial ψ ∈ C[x, y] of weight ϖ (χ ) − |ϖ |, which amounts to finding all monomials
xi y j , with i, j ∈ N satisfying
ϖ1 i + ϖ2 j = ϖ ( χ ) − ϖ1 − ϖ2 . (9.14)
Table 9.1 The different types of formal simple singularities which appear in Theorem 9.7, with
the corresponding weight vector ϖ = (ϖ1 , ϖ2 ), the integers ϖ1 and ϖ2 being the weights of the
variables x and y. For each type, equation (9.14) and the monomials corresponding to the solutions
of this equation are given.
In this section, we study Poisson structures on the affine space F3 , which is equipped
with its algebra of functions, denoted by F (F3 ). If F is an arbitrary field, this al-
gebra is the algebra of polynomial functions on F3 ; however, if F = R (respectively
F = C), then F (F3 ) may stand either for the algebra of smooth (respectively holo-
morphic) or polynomial functions on F3 , depending on the context. The standard
coordinates on F3 are denoted by (x, y, z). Every bivector field π on F3 decom-
poses as
∂ ∂ ∂ ∂ ∂ ∂
π = P1 ∧ + P2 ∧ + P3 ∧ , (9.15)
∂y ∂z ∂z ∂x ∂x ∂y
where P1 := {y, z}, P2 := {z, x} and P3 := {x, y}. It follows from item (iii) in Propo-
sition 1.16 that this bivector field satisfies the Jacobi identity, and is therefore a
Poisson structure on F3 , if and only if
which is equivalent to
∂ P3 ∂ P2 ∂ P1 ∂ P3 ∂ P2 ∂ P1
P1 − + P2 − + P3 − =0. (9.16)
∂y ∂z ∂z ∂x ∂x ∂y
→
− −
→
P | curl P = 0 . (9.17)
By (9.17), the assumption that π is a Poisson structure gives the vanishing of the
first term on the right-hand side of this equation, while the vanishing of the second
term follows from the identity A | B × A = 0, valid for all A, B ∈ F3 . This shows
→
−
that condition (9.17) is satisfied for χ P , so that the bivector field χπ is a Poisson
structure on F3 .
Another classical identity which involves the curl operator is
curl grad ϕ = 0 ,
valid for every function ϕ ∈ F (F3 ). It implies, in view of (9.17), that when the
→
− →
−
vector field P , associated to a bivector field π , is the gradient of a function, P =
grad ϕ , then π is a Poisson structure. Substituted in (9.15), this means that
∂ϕ ∂ ∂ ∂ϕ ∂ ∂ ∂ϕ ∂ ∂
πϕ := ∧ + ∧ + ∧ (9.19)
∂x ∂y ∂z ∂y ∂z ∂x ∂z ∂x ∂y
dF ∧ dG ∧ dϕ
{F, G}ϕ = .
dx ∧ dy ∧ dz
In particular, ϕ is a Casimir function for πϕ . Proposition 9.13 implies the following
statement.
Proposition 9.14. Let ϕ ∈ F (F3 ) be an arbitrary function on F3 and consider the
bivector field πϕ on F3 , defined by (9.19). For every function χ ∈ F (F3 ), the bivec-
tor field χπϕ is a Poisson structure on F3 and it admits ϕ as a Casimir function.
252 9 Poisson Structures in Dimensions Two and Three
Proof. The bivector field π is a Poisson structure on F3 if and only if the Schouten
bracket [π , π ]S is zero. Since
1
[π , π ]S [F, G, H] = {{F, G} , H} + {{G, H} , F} + {{H, F} , G} ,
2
for all functions F, G, H ∈ F (F3 ), we have that, if ϕ is a Casimir function of π , then
[π , π ]S [F, G, ϕ ] = 0, for all F, G ∈ F (F3 ). Let us show that this implies that, if ϕ is a
Casimir function of π and dϕ is non-zero on a dense subset of F3 , then [π , π ]S = 0,
so that π is a Poisson structure on F3 . Since [π , π ]S is a trivector field on F3 , it can
be written as
∂ ∂ ∂
[π , π ]S = ψ ∧ ∧ , (9.21)
∂x ∂y ∂z
where ψ ∈ F (F3 ). Plugging F := x and G := y in [π , π ]S [F, G, ϕ ] = 0 we obtain,
in view of (9.21), ψ ∂∂ϕz = 0. Similarly, one obtains ψ ∂∂ϕx = ψ ∂∂ϕy = 0, by plugging
in the two other pairs of variables. Since, by assumption, dϕ is non-zero on a dense
subset of F3 , it follows that ψ = 0 on F3 , hence that the trivector field [π , π ]S is zero
on F3 . This proves (1).
We next prove (2). We assume that π is a polynomial Poisson structure on F3 and
we define polynomial functions P1 , P2 , P3 ∈ F (F3 ) by
∂ ∂ ∂ ∂ ∂ ∂
π = P1 ∧ + P2 ∧ + P3 ∧ .
∂y ∂z ∂z ∂x ∂x ∂y
Suppose that ϕ is a polynomial Casimir function of π . We apply the equation
{F, ϕ } = 0, which is valid for all F ∈ F (F3 ), successively to F := x, F := y and
F := z. It gives the following relations between the functions P1 , P2 and P3 :
∂ϕ ∂ϕ ∂ϕ ∂ϕ ∂ϕ ∂ϕ
P2 = P3 , P3 = P1 , P1 = P2 . (9.22)
∂z ∂y ∂x ∂z ∂y ∂x
9.2 Poisson Structures in Dimension Three 253
∂ϕ ∂ϕ ∂ϕ
Let us assume that the partial derivatives ∂ x , ∂ y , ∂ z are relatively prime, and let Q
denote the greatest common divisor of ∂ x and ∂∂ϕy . Then it follows from the third
∂ϕ
∂ϕ ∂ϕ
QP1 = χ and QP2 = χ . (9.23)
∂x ∂y
In view of the first (or the second) equality in (9.22), this implies that
∂ϕ
QP3 = χ . (9.24)
∂z
∂ϕ ∂ϕ ∂ϕ
P1 = χ , P2 = χ and P3 = χ .
∂x ∂y ∂z
This proves (2).
∂ϕ ∂ϕ ∂ϕ
ϖ (ϕ )ϕ = ϖ 1 x + ϖ2 y + ϖ3 z , (9.25)
∂x ∂y ∂z
where ϖ1 , ϖ2 and ϖ3 are the weights of the variables x, y and z. Suppose that Q ∈
F (F3 ) is a non-constant irreducible polynomial function which divides the partial
derivatives ∂∂ϕx , ∂∂ϕy and ∂∂ϕz . It follows from (9.25) that Q divides ϕ . Since Q is non-
∂Q ∂Q ∂Q ∂Q
constant, at least one of the partial derivatives ∂x , ∂y and ∂z , say ∂x , is different
∂ϕ
from zero. Since the irreducible polynomial Q divides ϕ and its square Q2
∂x ,
also divides ϕ , so that ϕ is not square-free. Therefore, if ϕ is non-constant, weight
homogeneous and square-free, then ∂∂ϕx , ∂∂ϕy and ∂∂ϕz are relatively prime.
Note that it appeared already in Example 7.12. The associatedvector field is given
→
− →
− →
−
by P : (y, −x, 0), whose curl is (0, 0, −2), so that P | curl P = 0, which means,
254 9 Poisson Structures in Dimensions Two and Three
∂ϕ ∂ϕ ∂ϕ
= 0 and x +y =0.
∂z ∂x ∂y
The first equation means that ϕ does not depend on the variable z, while the second
one implies, in view of the Euler formula (8.5), that ϕ is homogeneous of degree 0
in the two remaining variables. According to Proposition 8.3, every homogeneous
function of degree zero is constant, so every Casimir function of π is constant. It
follows from Proposition 9.14 that π is not of the form χπϕ , with χ , ϕ ∈ F (F3 ).
Proof. Suppose that π is a bivector field of rank two on M. Let F and G be arbitrary
functions, defined on a non-empty open subset U of M. Let m be an arbitrary point
of U. We show that
Suppose first that the tangent vectors (XF )m and (XG )m are proportional, say
(XG )m = a (XF )m for some a ∈ F. Then the right-hand side of (9.27) vanishes
at m, while
so the left-hand side of (9.27) also vanishes at m. Thus, equation (9.27) is satisfied.
Assume now that the tangent vectors (XF )m and (XG )m are not proportional.
Since the rank of π at m is two, these tangent vectors span πm (Tm∗ M), so that for
every function H, defined on a neighborhood of m ∈ M, there exist a, b ∈ F such
that
(XH )m = a (XF )m + b (XG )m . (9.28)
Applying (9.28) to F yields
and we obtain that {F, H} (m) = b {F, G} (m). Similarly, applying (9.28) to G yields
{G, H} (m) = −a {F, G} (m). Substituted in (9.28), we obtain
Since H is arbitrary, this proves (9.27) also when (XF )m and (XG )m are not pro-
portional, thereby proving (1).
We now prove (2). If π is a Poisson structure of rank 2 on M, then {πm (Tm∗ M) |
m ∈ M(2) }, the distribution on M(2) , defined by all Hamiltonian vector fields, is
involutive because [XF , XG ] = −X{F,G} for all functions F and G. Conversely, let
π be a bivector field of rank two on M and assume that the distribution {πm (Tm∗ M) |
m ∈ M(2) } is involutive. Let m be a point of M. If πm is zero, then the Schouten
bracket [π , π ]S , evaluated at m, vanishes as well. If πm is different from zero, so that
m ∈ M(2) , then there exist functions F, G, defined in a neighborhood U of m, such
XF
that {F, G} = 0 on U. According to (1), π = V ∧ XG on U, where V := {F,G} .A
direct computation, with the help of the graded Leibniz identity (3.42), gives
1
[π , π ]S = V ∧ [V , XG ] ∧ XG .
2
Since XG and V are tangent to the distribution {πm (Tm∗ M) | m ∈ M(2) }, assumed to
be involutive, the bracket [V , XG ] is also tangent to the latter distribution. Hence,
([π , π ]S )m ∈ ∧3 πm (Tm∗ M), for every m ∈ M(2) . But the latter space is zero, since
256 9 Poisson Structures in Dimensions Two and Three
Proof. A proof can be given by picking local coordinates and using Proposi-
tion 9.13. We give a more geometrical proof. Consider the open subsets M(2) and
of M consisting of all points where the rank of π , respectively of ϕπ , is two. It
M(2)
⊂M
is clear that M(2) (2) and that for every point m ∈ M(2) , the vector spaces
the former is also involutive, which in view of the same proposition, implies that ϕπ
is a Poisson structure on M.
Proof. Suppose that π1 and π2 are Poisson structures which have the same sym-
plectic leaves. For m ∈ M, denote by Sm their common symplectic leaf through m.
If both Poisson structures are regular of rank 2, then on the one hand, Tm Sm is a
two-dimensional vector space, so that ∧2 Tm Sm is one-dimensional, while on the
other hand both (π1 )m and (π2 )m are non-zero elements of ∧2 Tm Sm , so there exists
a unique χ (m) ∈ F∗ , such that π1 (m) = χ (m)π2 (m). Thus, under the above assump-
tions, there exists a nowhere vanishing function χ on M, such that π1 = χπ2 ; since χ
is given locally as the quotient of two non-vanishing functions, χ belongs to F (M).
Q := ıQ λ .
9.2 Poisson Structures in Dimension Three 257
α ∧ dα = 0 . (9.29)
In this case, for every m ∈ M at which the rank of π is two, the kernel of αm is
the tangent space of the symplectic leaf of π through m;
(2) The bivector field π is a unimodular Poisson structure if and only if there
exists a volume form λ on M, such that the differential one-form α := ıπ λ is
closed.
Proof. Recall from Cartan’s formula (3.38) that if P and Q are two multivector fields
on M, then
[[ıP , d] , ıQ ] = ı[P,Q]S .
Specializing to P := π and Q := π , and applied to the volume form λ , it yields
Now, for every bivector field P and for every differential 2-form β , the fact that λ is
a top form implies that2
(ıP β )λ = β ∧ ıP λ . (9.32)
In particular, (ıπ dα )λ = dα ∧ ıπ λ = dα ∧ α , so that (9.31) gives eventually:
2 dα ∧ α = ı[π ,π ]S λ λ .
and π ∧ XF = 0 because the rank of π is 2 (see (9.26)), so that (XF )m ∈ Ker αm for
every m ∈ M and for every function F, defined in a neighborhood of m in M. This
completes the proof of (1).
We now prove (2). Recall from (4.24) that the modular vector field Φ of π with
respect to a given volume form λ , can be expressed in terms of the divergence of π
as Φ = − Div(π ). According to the definition of Div (see (4.18)) and the definition
α := ıπ λ , we have that
1
πΦ ,ϕ := πϕ − E ∧ Φ , (9.34)
3
where E is the Euler vector field, and πϕ is the Poisson structure defined in (9.19).
Proof. To start with, let us show that, given a linear vector field Φ on F3 and a cubic
function ϕ on F3 , which satisfy (9.33), the bivector field πΦ ,ϕ defined by (9.34) is
a Poisson structure. First, both E ∧ Φ and πϕ are Poisson structures according to
Example 8.15 and Proposition 9.14 respectively. As a consequence,
2 2 2
π Φ , ϕ , π Φ , ϕ S = − π ϕ , E ∧ Φ S = − LE π ϕ ∧ Φ + L Φ π ϕ ∧ E .
3 3 3
It suffices to prove that both of the remaining terms vanish. For the first term, by
Proposition 8.4, the Lie derivative with respect to the Euler vector field of every
quadratic Poisson structure is zero, so that LE πϕ = 0. The vanishing of the sec-
ond term follows at once by applying the following identity, valid for every vector
field V on F3 ,
LV πϕ = Div(V ) πϕ + πV [ϕ ] (9.35)
to V := Φ , upon using (9.33). We prove the identity (9.35) by showing that
First, by item (2) in Proposition 3.11 and by (4.20), the left-hand side of (9.36) is
given by
We are left with the task of showing that the assignment, defined by (9.34) is
injective. Since the divergence of πϕ vanishes and according to Example 8.25,
1 1
Div(πΦ ,ϕ ) = Div − E ∧ Φ = Φ − Div(Φ ) E = Φ ,
3 3
so that we can recover Φ as the modular vector field of πΦ ,ϕ . Since the cubic poly-
nomial ϕ can also be uniquely recovered from πϕ , both Φ and ϕ can be uniquely
recovered from πΦ ,ϕ , which proves injectivity, hence bijectivity of the assignment,
defined by (9.34). This completes the proof.
Proposition 9.21 leads to an explicit description of the family of all quadratic
Poisson structures on F3 , as given in the following theorem.
Theorem 9.22. Let π be a quadratic Poisson structure on C3 . There exist linear
coordinates on C3 in which π takes the form
1
π = πϕ − E ∧ Φ ,
3
where (Φ , ϕ ) corresponds to one of the types I–VII in Table 9.2, where πϕ is the
Poisson structure
∂ϕ ∂ ∂ ∂ϕ ∂ ∂ ∂ϕ ∂ ∂
πϕ := ∧ + ∧ + ∧ ,
∂x ∂y ∂z ∂y ∂z ∂x ∂z ∂x ∂y
Proof. Recall from Section 8.2 that linear vector fields on a finite-dimensional vec-
tor space V are in one-to-one correspondence with endomorphisms of the dual vec-
tor space V ∗ : more precisely, to a linear vector field Φ on V corresponds, as in
(8.20), the endomorphism V (1) of V ∗ defined by
Φ (1) : V ∗ → V ∗
(9.38)
F → Φ [F] .
Table 9.2 Up to linear isomorphism, every quadratic Poisson structure on C3 is of the form πΦ ,ϕ ,
where Φ and ϕ are given in the second and third columns.
Type Φ ϕ Parameters
a1 , a2 , a3 ∈ C∗
all ai different
I a1 x ∂∂x + a2 y ∂∂y + a3 z ∂∂z bxyz
a1 + a2 + a3 = 0;
b∈C
a ∈ C∗
II ax ∂∂x + ay ∂∂y − 2az ∂∂z z G(x, y)
G quadratic function
then A := (ai j ) ∈ Matd (C) is the matrix of Φ (1) in terms of the basis (x1 , . . . , xd )
for V ∗ . Using the classical formula (4.21) for the divergence of a vector field, we
find
d
∂ d
Div(Φ ) = ∑ (ai j xi ) = ∑ aii ,
i, j=1 ∂ x j i=1
so that the divergence of Φ is a constant function, whose value is the trace of the
endomorphism Φ (1) . In particular, a linear vector field Φ on C3 is divergence-free
if and only if the corresponding endomorphism Φ (1) of C3 is traceless. There are
seven types of such endomorphisms:
• If all eigenvalues of Φ (1) are different, then there are two possibilities, according
to whether one or zero of these eigenvalues are zero, leading to types I and III
in Table 9.3;
262 9 Poisson Structures in Dimensions Two and Three
Table 9.3 There are, up to isomorphism, seven types of traceless endomorphisms of C3 . For each
type, its matrix takes in terms of a well-chosen basis the form indicated in the second column of the
table. The range of the parameters which appear in the matrices is indicated in the third column.
a1 , a2 , a3 ∈ C∗
I diag(a1 , a2 , a3 ) all ai different
a1 + a2 + a3 = 0
II diag(a, a, −2a) a ∈ C∗
⎛ ⎞
a1 0
⎜ ⎟
IV ⎝0 a 0 ⎠ a ∈ C∗
0 0 −2a
⎛ ⎞
010
⎜ ⎟
V ⎝0 0 1⎠ −
000
⎛ ⎞
001
⎜ ⎟
VI ⎝0 0 0⎠ −
000
VII 0 −
• If Φ (1) has precisely two different eigenvalues, then they are both different from
zero, leading to types II or IV, according to whether Φ (1) is diagonalizable or
not;
• If all eigenvalues of Φ (1) are equal, then they are all zero, so that Φ (1) is a
nilpotent endomorphism. This leads to three types, the types V–VII.
The conditions on the parameters in Table 9.3 are chosen such that every traceless
endomorphism of C3 appears exactly once in the table.
It is clear that the matrices which are given in the second column of Table 9.3 cor-
respond to the linear vector fields on C3 , listed in the second column of Table 9.2.
Combining Proposition 9.21 and Lemma 9.23 leads to the announced classifica-
tion of the set of quadratic Poisson structures on C3 .
9.2 Poisson Structures in Dimension Three 263
∂ϕ ∂ ∂ ∂ϕ ∂ ∂ ∂ϕ ∂ ∂
πϕ := ∧ + ∧ + ∧ , (9.39)
∂x ∂y ∂z ∂y ∂z ∂x ∂z ∂x ∂y
Table 9.4 Every finite subgroup of SL2 (C) is conjugated to precisely one
√
of the subgroups,
√
given
2π −1 2π −1
in terms of generators in the second column of the table, where ε := e 8 and η := e 5 .
⎛ √ ⎞ ⎛ ⎞
−1π
⎜e k 0 ⎟ ⎜ 0 1 ⎟
Dk k4 ⎝ √
− −1π
⎠ and ⎝ ⎠
0 e k −1 0
⎞ ⎛
⎜ ε 7 ε7
⎟
E6 generators of D4 and √12 ⎝ ⎠
ε5 ε
⎞ ⎛
⎜ε 0 ⎟
E7 generators of E6 and ⎝ ⎠
0 ε7
⎛ ⎞ ⎛ ⎞
⎜ −η ⎜η +η
3 0 4 1
⎟ 1 ⎟
E8 ⎝ ⎠ and η 2 −η 3 ⎝ ⎠
0 −η 2 1 −η − η 4
Table 9.5 For each of the finite subgroups G of SL2 (C), the quotient surface C2 /G is the zero
locus of a weight homogeneous polynomial, given in the second column of the table. The third
column indicates the weight vector with respect to which each of these polynomials is weight
homogeneous.
Type of G ϕ Weights
Ak k1 x2 + y2 − zk+1 (k + 1, k + 1, 2)
Dk k 4 x2 + y2 z + zk−1 (k − 1, k − 2, 2)
E6 x 2 + y 3 + z4 (6, 4, 3)
E7 x2 + y3 + yz3 (9, 6, 4)
E8 x 2 + y 3 + z5 (15, 10, 6)
Note that this means that, for each group G which appears in the table, the algebra
of G-invariant polynomials is an algebra with three weight homogeneous generators
P1 , P2 , P3 , and that every polynomial R ∈ C[X,Y, Z], satisfying R(P1 , P2 , P3 ) = 0, is a
multiple of ϕ . The isomorphism C2 /G Σϕ is induced by the map C2 → C3 , given
by (P1 , P2 , P3 ), as expressed in the following commutative diagram:
(P1 ,P2 ,P3 )
C2 C3
p ı (9.40)
C2 /G
Σϕ
Recall from Section 5.1.2 that F (C2 )G F (C2 /G) (since G is a finite group). On
the other hand, since Σϕ is the affine surface of C3 , defined by the polynomial ϕ ,
its algebra of functions F (Σϕ ) is given by C[x, y, z]/ ϕ . It follows that we have an
algebra isomorphism
F (C2 )G C[x, y, z]/ ϕ . (9.41)
We show in the following proposition that the isomorphism C2 /G Σϕ is actu-
ally an isomorphism of Poisson varieties.
Proposition 9.24. Let G be a finite subgroup of SL2 (C) and let {· , ·}C2 /G denote
the Poisson structure on C2 /G, induced by the action of G on C2 . The Poisson sur-
266 9 Poisson Structures in Dimensions Two and Three
face (C2 /G, {· , ·}C2 /G ) is isomorphic, as a Poisson variety, to the Poisson surface
(Σϕ , {· , ·}ϕ ), where ϕ is the polynomial, associated to G, as indicated in Table 9.5.
{F, G}ϕ
χ= ,
{F, G}ϕ
so that χ is actually a rational function on Σϕ . Since its divisor of zeros and its di-
visor of poles are both discrete (they contain at most the point p(o)), χ is constant.
This implies that (Σϕ , {· , ·}ϕ ) and (C2 /G, μ {· , ·}C2 /G ) are isomorphic Poisson va-
rieties for some μ ∈ C∗ . In view of Remark 6.19, this constant μ can be chosen
equal to 1.
9.3 Notes 267
9.3 Notes
The classification in dimension two given in this chapter follows Monnier [150, 152],
a variant of Arnold’s proof [16]. In these references, the proof is more general than
the proof given here, since the case when R is taken as the ground field is also
considered, and with some extra work it is shown that the classification is actually
non-formal, i.e., the formal coordinate transformations can be realized by smooth
(or analytic) local coordinate transformations, when the given Poisson structure is
smooth (or analytic). The question of having a global rather than a local classi-
fication in the context of varieties or manifolds remains a real challenge; for an
important contribution in this direction, see [173], in which a classification of topo-
logically stable Poisson structures on real surfaces is given.
The classification of quadratic Poisson structures in dimension three, detailed in
Section 9.2.3 above, has been obtained independently by Dufour–Haraki [60] and
Liu–Xu [129]. For the relation between Poisson structures and differential forms, in
the presence of a volume form, see [88].
Chapter 10
R-Brackets and r-Brackets
The theory of integrable systems gave birth to interesting constructions, which yield
new Poisson structures on a given (finite-dimensional) Lie algebra g, upon using
some extra structure on the Lie algebra, roughly speaking a matrix. There are two
distinct formalisms for doing this, depending on whether this matrix is viewed as
a linear map R : g → g, in which case R is called an R-matrix, or as an element
r ∈ g ⊗ g, in which case r is called an r-matrix. In their basic form, the new Poisson
structures are Lie–Poisson structures, so it will be assumed that the reader is familiar
with the basic theory of Lie–Poisson structures, as developed in Chapter 7.
For a given linear map R : g → g, which satisfies a condition which will be ex-
plained in detail, a new Lie bracket on g is defined by letting [x, y]R := 12 ([Rx, y] +
[x, Ry]), for all x, y ∈ g. The new Lie bracket on g leads to a Poisson structure on g∗ ,
which is referred to as an R-bracket on g∗ . Upon identifying g with g∗ , which is usu-
ally done by using a non-degenerate symmetric bilinear form, the R-bracket on g∗
becomes a Lie–Poisson structure on g.
A tensor r ∈ g ⊗ g can be viewed as a zero-chain on g, with values in g ⊗ g, so
its coboundary δL0 (r) : g → g ⊗ g admits as transpose a linear map γr : g∗ ⊗ g∗ → g∗ .
When r satisfies a condition, which will also be explained in detail, γr defines a Lie
algebra structure on g∗ , hence leads directly to a Lie–Poisson bracket on g, i.e., no
identification of g with g∗ is needed to define this Lie–Poisson structure on g. This
Lie–Poisson bracket is called an r-bracket.
Section 10.1 is devoted to the construction of the R-bracket on g and on g∗ , with
special emphasis on the main example, which comes from Lie algebra splittings. In
Section 10.2, we discuss the case of r-brackets and we show how R-brackets and
r-brackets are related. The construction of higher degree brackets, from an R-matrix
or r-matrix, is given in Section 10.3. For the connection with Poisson–Lie groups,
we refer to Chapter 11.
In this chapter, F is an arbitrary field of characteristic zero, g is a finite-
dimensional Lie algebra over F and F (g) denotes the algebra of polynomial func-
tions on g, but it may also be taken as the algebra of smooth, respectively holomor-
phic, functions on g, when F = R, respectively when F = C.
We first give the general definition of an R-matrix and discuss its relation with the
Yang–Baxter equation.
Definition 10.1. Let (g, [· , ·]) be a Lie algebra and let R : g → g be a linear map. If
the skew-symmetric bilinear map [· , ·]R : g × g → g, defined for x, y ∈ g by
1
[x, y]R := ([Rx, y] + [x, Ry]) , (10.1)
2
defines a Lie bracket on g, then R is called an R-matrix for g. Then g, equipped with
the two Lie algebra structures [· , ·] and [· , ·]R , is called a double Lie algebra.
The standard terminology R-matrix and the notation Rx instead of R(x) are used
because one often thinks of R (and also x) as being a matrix.
We show in the following proposition that the conditions which express that a
linear map R : g → g is an R-matrix for g, can be rewritten in a symmetric form,
using a bilinear map on g, which depends quadratically on R.
Proposition 10.2. Let (g, [· , ·]) be a Lie algebra. A linear map R : g → g is an R-
matrix for g if and only if
where we have used the Jacobi identity for [· , ·] in the last step. This shows that the
Jacobi identity for [· , ·]R is equivalent to (10.2).
It follows from the proposition that, if there exists a constant c ∈ F such that
BR (x, y) = −c [x, y] for all x, y ∈ g, then R is an R-matrix for g. Given c ∈ F, the
equation
∀x, y ∈ g : BR (x, y) = −c [x, y] , (10.4)
viewed as an equation in R, is called the modified Yang–Baxter equation. When g is
a complex (respectively real) Lie algebra and c = 0, then a rescaling of R in (10.4)
yields c = 1 (respectively c = ±1). When c = 0, one speaks of the Yang–Baxter
equation. Sometimes one adds the adjective classical, to distinguish these equations
from the quantum Yang–Baxter equation. We warn the reader that the terminol-
ogy Yang–Baxter equation, with or without adjectives, is used for many equations
which are closely related to, but not equivalent to, the above Yang–Baxter equation
(see [108]).
Given a Lie algebra (g, [· , ·]), a linear map T : g → g is called an intertwining
operator of g if T [x, y] = [T x, y] = [x, Ty], for all x, y ∈ g, where we have written T x
as a shorthand for T (x). An intertwining operator of g can be used to construct a
new R-matrix for g, starting from a given R-matrix for g, as given in the following
proposition.
Proposition 10.3. Let (g, [· , ·]) be a Lie algebra and let R be an R-matrix for g. If
T : g → g is an intertwining operator of g, then the linear map R ◦ T is an R-matrix
for g.
for all x, y, z ∈ g, where we have used that BR◦T (x, y) = BR (T x, Ty), which is valid
because T is an intertwining operator of g. If T is invertible, then (10.5) is equiva-
lent to
T [BR (T x, Ty) , z] + (x, y, z) = 0 , (10.6)
which holds for all x, y, z ∈ g, because T is an intertwining operator of g and in view
of (10.2). If T is not invertible, pick an arbitrary α ∈ F and replace in (10.6) T by
T + α 1g . The identity
The main, and historically first, example of an R-matrix comes from a Lie algebra
splitting. If (g, [· , ·]) is a Lie algebra and g is written as a vector space direct sum
g = g+ ⊕ g− of two Lie subalgebras g+ and g− of g, then g+ ⊕ g− is called a Lie
algebra splitting. It leads to two projections P+ : g → g+ and P− : g → g− . For
x ∈ g, we abbreviate P+ (x) to x+ and P− (x) to x− . A Lie algebra splitting leads to
an R-matrix for g, as described by the following proposition.
Proposition 10.4. Let (g, [· , ·]) be a Lie algebra and let g = g+ ⊕g− be a Lie algebra
splitting, with corresponding projections P± : g → g± . The involutive endomorphism
R := P+ − P− of g is a solution of the modified Yang–Baxter equation (10.4), with
c = 1, namely
[Rx, Ry] − R ([Rx, y] + [x, Ry]) = − [x, y] , (10.8)
for all x, y ∈ g. In particular, R is an R-matrix for g.
while
where we have used in the last step that g+ and g− are Lie subalgebras of g. This
proves that R satisfies (10.8).
10.1 Linear R-Brackets 273
Notice that, when R = P+ − P− comes from a Lie algebra splitting, then an alterna-
tive formula for the R-bracket [· , ·]R is given, for x, y ∈ g, by
This elementary fact has important consequences, as we will see in the next sec-
tion. Notice that (10.9) can be used to give an elementary direct proof of the Jacobi
identity for [· , ·]R , without using Propositions 10.2 and 10.4.
Remark 10.5. More generally, suppose that g = g+ ⊕ g0 ⊕ g− is a vector space
decomposition of a Lie algebra g, where g+ and g− are Lie subalgebras, g0 is
abelian and is contained in the normalizer of g+ and of g− , i.e., [g0 , g± ] ⊂ g± . Let
R ∈ End(g) be defined for x ∈ g by Rx := x+ − x− , where x+ , x0 and x− stand for
the projections of x on g+ , g0 and g− respectively. Then one finds, as in the proof of
Proposition 10.4, that
for all x, y ∈ g. It follows that, as in the case of a Lie algebra splitting, R is a solution
of the modified Yang–Baxter equation (10.4), with c = 1.
If we assume that the Lie algebra (g, [· , ·]) is finite-dimensional, every R-matrix for g
leads to a Lie–Poisson structure on g∗ , called an R-bracket; the trivial case R = 1g
is usually excluded, because it corresponds to the canonical Lie–Poisson structure
on g∗ . According to (7.5), it is given for all F, G ∈ F (g∗ ) at ξ ∈ g∗ by
{F, G}R (ξ ) := ξ , dξ F, dξ G R
1
= ξ , R dξ F , dξ G + dξ F, R dξ G ,
2
where we recall that dξ F is viewed in this formula as an element of g, under the
canonical isomorphism between g and its bidual. Also, · , · stands for the canonical
pairing between g∗ and g. When R = P+ − P− comes from a Lie algebra splitting
g = g+ ⊕ g− , this formula can, according to (10.9), also be written as
{F, G}R (ξ ) = ξ , dξ F + , dξ G + − ξ , dξ F − , dξ G − . (10.10)
1
{F, G}R,g (x) = x | [R (∇x F) , ∇x G] + [∇x F, R (∇x G)] ,
2
for F, G ∈ F (g) and x ∈ g. When R comes from a Lie algebra splitting g = g+ ⊕ g− ,
this bracket becomes
{F, G}R,g (x) = x | (∇x F)+ , (∇x G)+ − x | (∇x F)− , (∇x G)− , (10.11)
Proof. It is clear from formula (10.9) for the R-bracket which comes from a Lie
algebra splitting that [g+ , g]R ⊂ g+ and that [g− , g]R ⊂ g− . Therefore, g+ and g−
are both Lie ideals of (g, [· , ·]R ) and we may conclude, according to Proposition 7.5,
that g⊥ ⊥
+ and g− are both Poisson submanifolds of (g, {· , ·}R,g ). Equation (10.12)
follows at once upon combining Eqs. (7.14) and (10.11).
We finish this section by giving two simple, but important, examples of R-brackets.
Example 10.7. Consider the Lie algebra splitting of g := gld (F), given by
g = Δd ⊕ Δd< = g+ ⊕ g− ,
where g+ := Δd (respectively g− := Δd< ) denotes the Lie algebra of upper trian-
gular matrices (respectively of strictly lower triangular matrices). Letting x | y :=
< ⊥
Trace(xy), we have that g⊥ − = Δd = Δd , the vector space of all lower triangular
matrices in g. If F is a linear function on g, then ∇x F = ∇F is independent of x ∈ g
and it follows from (7.12) that it is given by ∇F | y = F(y), for all y ∈ g. Denoting
by xi j the linear function which picks the element (i, j) of a matrix, it follows that
∇xi j = E ji , the matrix whose only non-zero entry is a 1 at position ( j, i). According
to (10.12), the Poisson bracket on Δd is therefore given by
xi j , xk − (x) = x | [E ji , Ek ] ,
where j i and k.
10.2 Linear r-Brackets 275
Example 10.8. Another natural Lie algebra splitting of gld (F) is given by gld (F) =
Δd ⊕ g− , where g− stands for the Lie subalgebra of skew-symmetric matrices. Us-
ing again x | y := Trace(xy), we have that g⊥
− is the vector space of all symmetric
d × d matrices, which inherits a Poisson structure from the Lie algebra splitting.
In the previous section, we have seen that a linear map R : g → g leads, under certain
conditions, to a new Lie–Poisson structure on g∗ , and hence on g, when g∗ and
g have been identified. In the present section, we describe a similar construction,
starting from an element r ∈ g ⊗ g, which gives, under certain conditions, directly a
Lie–Poisson structure on g.
The Jacobi identity for [· , ·] implies that the linear map ad : g × T • g → T • g, defined
by (10.13), is a representation of g on T • g. Combined with the natural projection
map T • g → ∧• g, we obtain the adjoint representation of g on ∧• g which, as we have
seen in Section 5.1.3, can be written in terms of the algebraic Schouten bracket:
adx Y = [[x,Y ]] for all x ∈ g and all Y ∈ ∧• g. Elements Y of T • g or of ∧• g for which
276 10 R-Brackets and r-Brackets
adx Y = 0, for all x ∈ g, are called ad-invariant; for example, elements of degree one
are ad-invariant if and only if they belong to the center of g.
It is clear from (10.13) that adx can be restricted to g ⊗ g, which yields a rep-
resentation of g on g ⊗ g. It allows us to consider, for p ∈ N, the vector space
C p (g) := Hom(∧ p g, g ⊗ g), of skew-symmetric p-linear maps on g, with values in
the representation space g ⊗ g. According to Section 4.1.1, C• (g) carries a cobound-
ary operator δL : C• (g) → C•+1 (g). Applied to r ∈ g ⊗ g C0 (g), we obtain a linear
map γr := δL0 (r) : g → g ⊗ g, which is given by
γr (x) = adx r ,
for all x ∈ g and ξ , η ∈ g∗ . We will be interested in the case in which the bilinear
map γr defines a Lie algebra structure on g∗ .
Definition 10.9. Let (g, [· , ·]) be a finite-dimensional Lie algebra. A bivector r ∈
g ⊗ g is called an r-matrix for g if the transpose γr of γr = δL0 (r) : g → g ⊗ g is a
Lie bracket on g∗ . In this case, r leads to a Lie bracket on g∗ , denoted by [· , ·]r , and
defined by1
[ξ , η ]r , x := ξ ∧ η , adx r (10.15)
for all ξ , η ∈ g∗ and x ∈ g. This bracket is called the Lie bracket, associated with r,
and (g, [· , ·], [· , ·]r ) is called a coboundary Lie bialgebra.
As the terminology suggests, coboundary Lie bialgebras are a particular case of
Lie bialgebras, but we will only consider this more general class when we discuss
Poisson–Lie groups, see Chapter 11.
We will give necessary and sufficient conditions on r ∈ g⊗g so that (g, [· , ·], [· , ·]r )
is a coboundary Lie bialgebra. Our proof will be based on the following proposition.
so that φ̃ ◦ φ̃ = 0 if and only if φ satisfies the Jacobi identity. Step (∗) requires a
proof. Let X ∈ ∧2V , which we suppose to be of the form X = v ∧ w, with v, w ∈ V
(in general, an element of ∧2V is a linear combination of such monomials). In view
of the graded derivation property (10.16) we have, for all ξ , η , ζ ∈ V ∗ , that
φ̃ (X), ξ ∧ η ∧ ζ
= φ (v) ∧ w − φ (w) ∧ v, ξ ∧ η ∧ ζ
= φ (v), ξ ∧ η w, ζ − φ (w), ξ ∧ η v, ζ + (ξ , η , ζ )
= v, φ (ξ ∧ η ) w, ζ − w, φ (ξ ∧ η ) v, ζ + (ξ , η , ζ )
= v ∧ w, φ (ξ ∧ η ) ∧ ζ + (ξ , η , ζ )
= X, φ (ξ ∧ η ) ∧ ζ + (ξ , η , ζ ) .
where we have used that (adx r)± = adx r± ; to verify the latter, write r± as a sum of
elements of the form y ⊗ z ± z ⊗ y. It follows from (10.18) that γr is skew-symmetric
if and only if r+ is ad-invariant.
Suppose now that r+ is ad-invariant. Then γr (x) = adx r = adx r− = γr− (x) for
all x ∈ g, so that γr depends only on a := r− . Thus we are given a skew-symmetric
element a ∈ g ∧ g and we wish to prove that the Jacobi identity for γa is equivalent
to the ad-invariance of [[a, a]]. In view of Proposition 10.10, it suffices to show that
the ad-invariance of [[a, a]] is equivalent to γ̃a ◦ γ̃a = 0 (on ∧• g; equivalently, on g).
Combining (10.13) and (3.46), we find that
for all x ∈ g. Since γ̃a and −[[a, ·]] are both graded derivations of degree 1 of ∧• g, it
follows that γ̃a (X) = −[[a, X]], for all X ∈ ∧• g. Therefore, γ̃a ◦ γ̃a = 0 if and only if
[[a, [[a, x]]]] = 0, for all x ∈ g. In view of the graded Jacobi identity for [[· , ·]],
so that [[a, [[a, x]]]] = 0, for all x ∈ g, is equivalent to [[x, [[a, a]]]] = 0, for all x ∈ g,
which is precisely the ad-invariance of [[a, a]].
The simplest way to satisfy both conditions in Proposition 10.11 is to demand that
r is skew-symmetric and satisfies [[r, r]] = 0.
defines a Lie algebra structure on g∗ . Then g inherits a linear Poisson structure {· , ·}r
from [· , ·]r , since g (g∗ )∗ is then the dual of a Lie algebra. According to (7.4), the
Poisson bracket of two functions F, G ∈ F (g) is given by
{F, G}r (x) = [dx F, dx G]r , x = dx F ∧ dx G, adx r− , (10.20)
adx r− = ∑ ai j ei ∧ e j .
1i< jd
Combined with (10.20), this means that the coefficients of the matrix can be com-
puted by
ai j = ξi ∧ ξ j , adx r− = ξi , ξ j r (x) ,
as was to be shown.
Ξ ⊗, Ξ r
(x) := ∑ ξi , ξ j r
(x) ei ∧ e j .
1i< jd
In this formula, (e1 , . . . , ed ) is a basis of g and (ξ1 , . . . , ξd ) is its dual basis. Equa-
tion (10.21) is known as the first Russian formula.
The first Russian formula is very useful for explicit computations, in particular
when g is realized as a Lie subalgebra of glN (F) MatN (F). Then this formula
can be thought of as an equality of matrices of size N 2 . Indeed, there is a standard
algebra isomorphism between MatN (F) ⊗ MatN (F) and MatN 2 (F); to describe it,
we use the classical elementary matrices Ei j , whose only non-zero entry is a 1 at
position (i, j), but we use the convention that numbering starts from zero, in order
to simplify the formula. Then the N 2 elements Ei j with 0 i, j < N form a basis
of MatN (F) and the N 4 elements Ei j ⊗ Ek , with 0 i, j, k, < N are a basis of
MatN (F) ⊗ MatN (F). The isomorphism MatN (F) ⊗ MatN (F) → MatN 2 (F) is given
by mapping Ei j ⊗ Ek to EiN+k, jN+ . It is an isomorphism of associative algebras,
since
(Ei j ⊗ Ek )(Ei j ⊗ Ek ) = δ j,i δ,k Ei j ⊗ Ek ,
while
EiN+k, jN+ Ei N+k , j N+ = δ j,i δ,k EiN+k, j N+ .
2Since g is finite-dimensional, g is isomorphic to a subalgebra of glN (F), for some N, which is the
Lie algebra of the associative algebra Matn (F) of square matrices of size N.
280 10 R-Brackets and r-Brackets
χ ⊗ 1g : g ⊗ g → g∗ ⊗ g End(g) ,
Rx | y = x | R∗ y ,
for all x, y ∈ g, which is, in view of the definition (10.1) of [· , ·]R , equivalent to
showing that, for all x, y, z ∈ g,
1
[Rx, y] + [x, Ry] | z = [χ (x), χ (y)]r , z . (10.24)
2
With the above formulas for R, the left-hand side in (10.24) can be written as
10.2 Linear r-Brackets 281
We leave it as an exercise for the reader to check that, under the assumptions of
Proposition 10.13, the operator BR , defined in (10.3), is expressible in terms of the
bracket [[r− , r− ]], defined in (3.46), as
z | BR (x, y) = χ (x) ∧ χ (y) ∧ χ (z), [[r− , r− ]] ,
for all x, y, z ∈ g. It implies that [[r− , r− ]] = 0 if and only if BR (x, y) = 0 for all x, y ∈ g,
i.e., if and only if R is a solution of the Yang–Baxter equation.
Example 10.14. Let (g, [· , ·]) be a finite-dimensional Lie algebra, equipped with a
non-degenerate symmetric bilinear form · | ·. Suppose that g = g+ ⊕ g− is a Lie
algebra splitting, where each of the factors is isotropic: g+ | g+ = g− | g− = {0} .
Let R be the linear endomorphism of g, given by R := P+ − P− , where P+ and P− are
the projections onto g+ and g− respectively. According to Proposition 10.4, R is an
R-matrix. For x ∈ g, we denote x+ := P+ (x) and x− := P− (x). Then the Lie bracket
associated with R is given, for all x, y ∈ g, by
1 d
r= ∑ e i ∧ εi .
2 i=1
(10.26)
Example 10.15. The above example can be easily generalized to the case of a de-
composition g = g+ ⊕ g0 ⊕ g− , as in Remark 10.5. It is assumed, as in the cited
remark, that g+ and g− are Lie subalgebras of g and that g0 is abelian and is con-
tained in the normalizer of g+ and of g− . Moreover, it is assumed that g comes
equipped with a non-degenerate symmetric bilinear form · | · with respect to which
both g+ and g− are isotropic and orthogonal to g0 . Recall that the endomorphism R
of g, defined for x ∈ g by Rx := x+ − x− , where x+ , x0 and x− stand for the projec-
tions of x on g+ , g0 and g− respectively, is a solution of the modified Yang–Baxter
equation (10.4), with c = 1. In view of the above assumptions,
1 d
r = ∑ e i ∧ εi .
2 i=1
as a Lie subalgebra. Consider on sld (C) the R-matrix√associated to the above de-
composition, given in Example 10.15 and let R := 2 −1R, which is an R-matrix
for sld (C), because R is a multiple of R. Notice that R leaves sud invariant, since
√ √
x+x = 0 implies that x+ +x− = 0, which in turns yields −1Rx+ −1(Rx) = 0.
It follows that R can be restricted to sud and is an R-matrix for sud . We con-
sider the symmetric bilinear form on sud , given by (x, y) → x | y := ℜ(Trace(xy)).
It is non-degenerate, hence induces an isomorphism χ between sud and its dual,
and R is skew-symmetric with respect to · | ·. Define, for all 1 i < j d,
the skew-symmetric matrices Ai j := √12 (Ei j − E ji ), the symmetric matrices Si j :=
√ √
√−1 (Ei j
+ E ji ) and the diagonal matrices Hi := √−1 (Eii − Ei+1,i+1 ). These matri-
2 2
ces form a basis of sud , with the set of all Si j and Ak forming an orthogonal set,
10.3 R-Brackets and r-Brackets of Higher Degree 283
with Si j | Si j = Ai j | Ai j = −1. The restriction of R to sud vanishes on diagonal
matrices, and satisfies
so that it is given by
R = 2 ∑ χ (Si j ) ⊗ Ai j − 2 ∑ χ (Ai j ) ⊗ Si j .
1i< jd 1i< jd
r = ∑ Si j ∧ Ai j .
1i< jd
In this section we show that, under some assumptions, an R-matrix leads not only
to a linear Poisson structure (see Section 10.1), but also to a quadratic and a cubic
Poisson structure. For each of these Poisson structures, the finite-dimensional Lie
algebra g, which underlies the construction, is assumed to be the Lie algebra of an
associative algebra: we suppose that we have an associative product on g and the Lie
bracket on g is defined, for all x, y ∈ g, by [x, y] := xy − yx. It is moreover assumed
that g comes equipped with a non-degenerate symmetric bilinear map · | ·, which
satisfies, for all x, y, z ∈ g,
xy | z = x | yz . (10.27)
Notice that this property implies that · | · is ad-invariant.
Example 10.17. The key example which the reader should keep in mind is that of
g := gld (F), which is the Lie algebra of Matd (F), the associative algebra of d × d
matrices with coefficients in F. The symmetric bilinear form · | ·, which is defined
by x | y := Trace(xy), for all x, y ∈ g, satisfies (10.27).
{F, G}Q
R (x) :=
1
2 [x, ∇x F] | R(x ∇x G + ∇x G x)
(10.28)
− 12 [x, ∇x G] | R(x ∇x F + ∇x F x) ,
284 10 R-Brackets and r-Brackets
for all F, G ∈ F (g). For the linear functions Fi (i = 1, 2, 3), their gradient ∇x Fi ∈ g is
independent of x; we therefore simply denote ∇x Fi by fi . When computing (10.30),
we will need ∇x {F1 , F2 }+ and ∇x {F1 , F2 }− . Since {F1 , F2 }+ and {F1 , F2 }− depend
quadratically on x, these gradients are easily computed from (10.31), giving
∇x {F1 , F2 }+ = R+ (x f1 ) f2 + f1 R+ ( f2 x) − f2 R+ ( f1 x) − R+ (x f2 ) f1 ,
(10.32)
∇x {F1 , F2 }− = R− ( f1 x) f2 − R− ( f2 x) f1 + f1 R− (x f2 ) − f2 R− (x f1 ) .
Using in two subsequent steps that R− is skew-symmetric and then that R− satisfies
the modified Yang–Baxter equation, we compute
{F1 , F2 }+ , F3 +
+ (1, 2, 3)
= [R+ ( f1 x), R+ ( f2 x)] | x f3 − [R+ (x f1 ), R+ (x f2 )] | f3 x + (1, 2, 3) ,
(10.33)
and the following two other terms:
where the Lie bracket [· , ·]R− is defined as in (10.1), using R− instead of R: for
x, y ∈ g,
1
[x, y]R− := ([R− x, y] + [x, R− y]) .
2
Clearly, (10.33) and (10.34) sum up to zero (which yields (10.30)), if
1
R+ : (g, [· , ·]R− ) → (g, [· , ·])
2
is a homomorphism of Lie algebras. Written out, this means that it suffices to prove
that
[R+ x, R+ y] − R+ [R− x, y] − R+ [x, R− y] = 0 , (10.35)
for all x, y ∈ g. We define a linear map B∗R : g ⊗ g → g by requiring that for all
x, y, z ∈ g,
BR (x, y) | z = x | B∗R (y, z) . (10.36)
Explicitly, B∗R is given, for x, y ∈ g, by
Notice that
so that
1 1
BR− (x, y) + B∗R (y, x) − B∗R (x, y)
2 2 (10.37)
= [R+ x, R+ y] − R+ [R− x, y] − R+ [x, R− y] ,
which is the left-hand side of (10.35). Since R− is a solution of the modified Yang–
Baxter equation (with constant c), we have that BR− (x, y) = −c [x, y], for all x, y ∈ g.
But R is also a solution of the modified Yang–Baxter equation (with the same con-
stant c) which, combined with (10.36) yields that B∗R (x, y) = −c [x, y], for all x, y ∈ g.
Substituted in (10.37), this yields (10.35). Thus, (10.33) and (10.34) sum up to zero,
which proves (10.30), hence {· , ·}Q R is a (quadratic) Poisson structure.
286 10 R-Brackets and r-Brackets
1
dx F, (XH )x = {F, H}Q
R (x) = [x, ∇x F] | R(x ∇x H + ∇x H x)
2
= [x, ∇x F] | R(x ∇x H) = ∇x F | [R(x ∇x H), x]
= dx F, [R(x ∇x H), x] ,
for all F, G ∈ F (g). In the particular case where R comes from an r-matrix r ∈ g ⊗ g,
as explained in Section 10.2.3, this formula takes a particularly simple form. In
order to establish it, we decompose r− = ∑α sα ⊗ tα and use, as in the proof of
Proposition 10.13, the convention that a sum over α is implicit in all formulas which
contain α . Recall that with this notation,
for all x, y ∈ g. For F, G ∈ F (g), the bracket (10.38) can be rewritten, using (10.39),
as
∑
Q Q
Ξ ⊗, Ξ r
(x) := ξi , ξ j R
(x) ei ∧ e j ,
1i< jd
which is known as the second Russian formula. As in the case of the first Russian
formula, this formula can be thought of in terms of matrices in MatN 2 (F), upon using
the standard algebra isomorphism between MatN (F) ⊗ MatN (F) and MatN 2 (F).
For the cubic R-bracket, the context is the same as in the case of the quadratic
R-bracket, but the assumptions are weaker.
{F, G}CR (x) := [x, ∇x F] | R(x ∇x G x) − [x, ∇x G] | R(x ∇x F x) , (10.41)
ẋ = [R(x∇x H x), x] .
Proof. The proposition can be proved in the same way as in the case of the quadratic
R-bracket (Proposition 10.18). Actually, the proof of the Jacobi identity for {· , ·}CR
requires less computation, as one does not need to split up R into its symmetric
and skew-symmetric parts and as the formula for the cubic bracket (10.41) has only
half as many terms as the formula for the quadratic bracket (10.28). For linear func-
tions F1 , F2 , F3 ∈ F (g), whose gradients (at any point) are denoted by f1 , f2 , f3 , first
compute from (10.41) that at x ∈ g one has
To finish this section, we show how the linear, quadratic and cubic R-brackets are
related. The Lie algebra g is, as above, the Lie algebra of an associative algebra,
with unit e. On g, we can therefore consider the vector fields V0 , V1 and V2 , given at
all x ∈ g by
288 10 R-Brackets and r-Brackets
for every x ∈ g. In particular, V1 is the Euler vector field, which was defined in
Section 8.1.2. For i = 0, 1, 2, we write Li for the Lie derivative with respect to Vi .
Proposition 10.20. Let (g, [· , ·]) be the Lie algebra of an associative algebra, with
unit e, which is equipped with a non-degenerate symmetric bilinear map · | ·, which
satisfies xy | z = x | yz, for all x, y, z ∈ g. Let R be a solution of the modified Yang–
Baxter equation, with constant c ∈ F, and suppose that its skew-symmetric part R−
is also a solution of the modified Yang–Baxter equation, with the same constant c.
Then the linear, quadratic and cubic Poisson structures on g, associated with R,
namely {· , ·}R,g , {· , ·}Q C
R and {· , ·}R , are related by Lie derivatives, as indicated in
the following commutative diagram:
−L2 −L2 L2
{· , ·}R,g {· , ·}Q
R {· , ·}CR 0
−L1 L1 (10.42)
1L 1L
L0 2 0 2 0
0 {· , ·}R,g {· , ·}Q
R {· , ·}CR
Moreover, for a fixed such R, the three Poisson structures {· , ·}R,g , {· , ·}Q
R and
C
{· , ·}R are compatible.
Proof. Since V1 is the Euler vector field, the Lie derivatives of the linear, quadratic
and cubic structures with respect to V1 give a multiple of these Poisson structures, up
to a constant, which is respectively −1, 0 and 1 (see Proposition 8.4). This yields the
two vertical arrows in the diagram. For each of the Lie derivatives L0 and L2 , the
computation is very similar, so we will only do it for one of them, namely we show
R . Let F1 , F2 be linear functions on g and let f 1 := ∇F1
that −L2 {· , ·}R,g = {· , ·}Q
and f2 := ∇F2 , as before. According to the formula (3.7) for the Lie derivative of a
biderivation (p = 2), we need to show that
{F1 , F2 }Q
R = −V2 {F1 , F2 }R,g + {V2 [F1 ], F2 }R,g + {F1 , V2 [F2 ]}R,g . (10.43)
To do this, first notice that for a linear function F on g, one has ∇x F | y
= F(y),
independently of x ∈ g (so we write f for ∇x F) and V2 [F](x) = F(x2 ) = f | x2 . It
follows that, for all x, y ∈ g,
d
∇x (V2 [F]) | y = F((x + ty)2 ) = F(xy) + F(yx) = x f + f x | y .
dt |t=0
10.4 Notes 289
∇x (V2 [F]) = x f + f x .
One computes similarly that ∇x (V1 [F]) = f and that ∇x (V0 [F]) = 0, but these for-
mulas are only needed for the verification of the other cases. Since {F1 , F2 }R,g is a
linear function on g, the right-hand side of (10.43), evaluated at x, is given by
− x2 | [ f1 , f2 ]R + x | [∇x (V2 [F1 ]), f2 ]R + x | [ f1 , ∇x (V2 [F2 ])]R
= − x2 | [ f1 , f2 ]R + x | [x f1 + f1 x, f2 ]R + x | [ f1 , x f2 + f2 x]R
1 1
= [x, f1 ] | R(x f2 + f2 x) − [x, f2 ] | R(x f1 + f1 x) ,
2 2
where we used in the last step that πRQ and πR are compatible, and that [V2 , πR ]S =
−πRQ (which is a Poisson structure!).
10.4 Notes
R-matrices and r-matrices have their origins in the theory of integrable systems,
but also play a fundamental rôle in the theory of quantum groups. They first ap-
peared in the contexts of the Adler–Kostant–Symes theorem, exhibiting the Lie al-
gebraic structures which underlie the integrability of both the Toda lattices and the
Korteweg–de Vries equation, see Adler [2].
The term “quantum group” is a generic name which is used for several types
of non-commutative algebras, usually Hopf algebras, i.e., algebras (rather than
groups!) endowed with an algebra structure on the dual. Quantum groups are in
3 In general, the notion “is compatible with” is not transitive.
290 10 R-Brackets and r-Brackets
A Poisson structure on a Lie group G makes G into a Poisson manifold. For reasons
which will be given later in this chapter, one demands the following compatibility
relation between the Poisson structure on G and the group structure of G.
Definition 11.1. A Poisson structure π on a (real or complex) Lie group G is said
to be multiplicative if the product map μ : G × G → G is a Poisson map, where
G × G is endowed with the product Poisson structure. The pair (G, π ) is then called
a Poisson–Lie group.
A map Φ : G → H between two Poisson–Lie groups (G, π ) and (H, π ) is called
a Poisson–Lie group homomorphism if it is both a Poisson map and a group homo-
morphism.
Proposition 11.2. Let G be a Lie group. For a Poisson structure π on G, the follow-
ing three conditions are equivalent:
(i) π is multiplicative;
(ii) For all g, h ∈ G:
for all g, h ∈ G.
Proof. We first prove that (i) and (ii) are equivalent, which amounts to proving that
μ : G × G → G is a Poisson map if and only if (11.1) holds for all g, h ∈ G. Let us
denote the product Poisson structure on G × G by Π ; according to (2.11), the value
of Π at (g, h) ∈ G × G is given by the bivector
where ıh and ıg are the inclusions of G into G × G given by ıh : g → (g, h) and
ıg : h → (g, h) respectively. According to (1.32), μ is a Poisson map if and only if
which is precisely the right-hand side of (11.1). This proves that (i) and (ii) are
equivalent.
In order to show that (ii) and (iii) are equivalent, we fix g, h ∈ G and we apply
the isomorphism ∧2 (Tgh R(gh)−1 ) to both sides of (11.1), giving
and we show that the latter equation is, term by term, equation (11.3). For the first
two terms, this follows from the definition of Ψ and from the identity
R(gh)−1 ◦ Rh = Rg−1 , namely
For the third and final term, recall that the map Cg is conjugation by g, so that
Cg = Rg−1 ◦ Lg , leading to the identity R(gh)−1 ◦ Lg = Cg ◦ Rh−1 , which in turn yields
where we have used in the last step that ∧(TeCg )(x) = Adg x for all x ∈ g and g ∈ G.
This proves that condition (11.1) is equivalent to condition (11.3).
In item (iii) of Proposition 11.2, we used the term cocycle, which is borrowed from
the terminology used when introducing the concept of cohomology for Lie groups.
We will not develop Lie group cohomology in this book, we rather restrict ourselves
to giving the general definition of a cocycle of a group (strictly speaking a 1-cocycle)
and proving the one result on group cocycles which we will use.
294 11 Poisson–Lie Groups
for all g, h ∈ G.
TgΨ ◦ Te Lg = ρg ◦ TeΨ ,
so that the vanishing of TeΨ implies the vanishing of TgΨ ◦ Te Lg , and therefore
of TgΨ , for all g ∈ G. Since G is assumed to be connected, this implies that Ψ is
a constant function. Since Ψ (e) = 0, as follows from (11.6), Ψ vanishes at each
point, which shows (1). If W is a G-invariant subspace, then we have an induced
representation G × V /W → V /W and Ψ induces a cocycle of G in V /W . In view
of (1), the latter cocycle is zero as soon as its tangent map at e is zero, i.e., as soon
as TeΨ (x) ∈ W for all x ∈ g, which proves (2).
See [181] for more information on group cohomology.
which yields, upon putting g := exp(tx) and taking the derivative with respect to t
at t = 0,
de {F, G} , x = de F ∧ de G, TeΨ (x) , (11.8)
where we have used that Ψ (e) = 0. Comparing (11.7) with (11.8) leads to the desired
equality
(π1 )x = TeΨ (x) , (11.9)
proving (2). To say that inv is an anti-Poisson map means that inv : (G, π ) →
(G, −π ) is a Poisson map: for every g ∈ G, ∧2 (Tg inv) πg = −πg−1 . To prove that
the latter holds, let g ∈ G, substitute g−1 for h in (11.1), and use πe = 0, to find that
Combined with Tg inv = −Te Lg−1 ◦ Tg Rg−1 (see (5.2)), this leads to
Proposition 11.8. Let (G, π ) be a Poisson–Lie group and let (H, π ) be a Poisson–
Lie subgroup of G. Suppose that H is a closed subgroup of G, so that the left coset
space H \G is a smooth manifold. There exists a unique Poisson structure on H \G,
such that the canonical projection p : G → H\G is a Poisson map. Moreover, the
right action of G on H\G is a (right) Poisson action.
Proof. Consider the left action of H on G which is the restriction of the product
map on G. Since H is a Poisson submanifold of G, the product H × G is a Poisson
submanifold of G × G, hence the left action of H on G is a Poisson action. Since
the quotient space of this action is precisely H\G, according to Proposition 5.33,
the quotient space H\G inherits a unique Poisson structure from (G, π ) such that
the canonical projection p : G → H\G is a Poisson map. Since the product map
μ : G × G → G is a Poisson map, the induced map (H\G) × G → (H\G) is also a
Poisson map, hence the (right) action of G on H\G is a Poisson action.
for all v, w ∈ V and for all linear functions F, G ∈ V ∗ . We have on the one hand that
and on the other hand, since F and F ◦ Rw differ by a constant (namely F(w)), that
{F, G} (v) = dv F ∧ dv G, πv
= dv (F ◦ Rw ) ∧ dv (G ◦ Rw ), πv (11.13)
= dv+w F ∧ dv+w G, ∧2 (Tv Rw ) πv .
298 11 Poisson–Lie Groups
for all v, w ∈ V . Since the latter condition is precisely condition (ii) in Proposi-
tion 11.2, this shows that π is linear if and only if π is multiplicative.
Since the dual of a finite-dimensional Lie algebra admits a canonical linear Pois-
son structure (the Lie–Poisson structure), it follows from the above proposition that
the dual of a finite-dimensional Lie algebra is in a canonical way a Poisson–Lie
group.
Proof. It follows immediately from Proposition 5.2 that the Schouten bracket of
π := ←−
a −→ −
a with itself is given by
[π , π ]S = [←−
a −→ −a ,←−
a −→ −a ]S
= [ a , a ]S − [ a , −
←− ←
− ←− →a ]S − [→
−
a ,←
−
a ]S + [→
−
a ,−
→
a ]S
←− ←
− →
− →
−
= [ a , a ]S + [ a , a ]S
←−−− −−−→
= [[a, a]] − [[a, a]] .
In view of item (3) in Proposition 5.3, we can conclude that π , as defined above, is
a Poisson structure on G if and only if [[a, a]] is Ad-invariant. It remains to be shown
that π is moreover multiplicative. The map Ψ : G → ∧2 g associated with π as in
item (iii) in Proposition 11.2 (i.e. defined, for all g ∈ G, by Ψ (g) := ∧2 (Tg Rg−1 ) πg )
is given by
Ψ (g) = Adg a − a , (11.14)
11.1 Multiplicative Poisson Structures and Poisson–Lie Groups 299
which, in view of the equivalence of items (i) and (iii) in Proposition 11.2, implies
that π is multiplicative. Therefore, (G, π ) is a Poisson–Lie group if and only if
[[a, a]] is Ad-invariant. If G is connected, then this is, according to Proposition 5.3,
equivalent to [[a, a]] being ad-invariant, i.e., to a being an r-matrix.
The previous proposition leads to the following definition.
Definition 11.11. A Poisson–Lie group (G, π ) is said to be a coboundary Poisson–
Lie group if π is of the form π = ←
−
a −→
−
a for some element a ∈ ∧2 g.
For a given coboundary Poisson–Lie group (G, π ) there may exist two different
elements a1 , a2 ∈ ∧2 g such that π = ←a−1 − −
→
a1 and π = ←
a−2 − −
→
a2 . However, in such a
case, we have
a−−−−→ ←−−−−
1 − a2 = a 1 − a2
(1) Vinv is a (non-empty) Zariski open subset of V and the restriction of μ to Vinv
defines the structure of a Lie group on Vinv , whose Lie algebra is isomorphic
to V , with Lie bracket given by [x, y] = μ (x, y) − μ (y, x) for all x, y ∈ V ;
(2) If π is a multiplicative Poisson structure on (V, μ ), in the sense of Sec-
tion 8.2.1, then the restriction π of π to Vinv is a multiplicative Poisson struc-
ture on the Lie group Vinv , making (Vinv , π ) into a Poisson–Lie group;
(3) Let a ∈ ∧2V be a skew-symmetric r-matrix for V . The quadratic bivector
field π on V , whose value at x ∈ V is given by
πx := 2 [x ⊗ x, a] (11.16)
Proof. Since μ is assumed to be associative, the invertible elements are precisely the
elements v ∈ V for which the linear map Lv : V → V , defined by Lv (w) := μ (v, w),
is an isomorphism, i.e.,
Vinv = {v ∈ V | det Lv = 0} .
←
− F(v + t μ (v, x)) − F(v)
x [F](v) = xe [F ◦ Lv ] = lim = F(μ (v, x)),
t→0 t
so that ←
−
x [F] = F ◦ Rx . It follows that
11.1 Multiplicative Poisson Structures and Poisson–Lie Groups 301
[←
−
x ,←
−
y ] [F] = ←
−
x [←
−
y [F]] − ←
−
y [←
−
x [F]]
= F ◦ (Ry ◦ Rx − Rx ◦ Ry )
= F ◦ Rμ (x,y)−μ (y,x)
←−−−−−−−−−−−−
= (μ (x, y) − μ (y, x))[F],
←−−− −−−→
z1 ∧ z2 x = xz1 ∧ xz2 , z1 ∧ z2 x = z1 x ∧ z2 x ,
2We know from Section 10.3, in particular from the second Russian formula (10.40), that π is a
Poisson structure, but this fact is not used in the proof.
302 11 Poisson–Lie Groups
is an r-matrix on the Lie algebra Matd (F). With the help of this r-matrix, a Pois-
son structure π on Matd (F) can be constructed as in Proposition 11.13. Let us ex-
press the coefficients of the Poisson matrix of π with respect to the linear coordi-
nates ξi j . For all i, j, k, = 1, . . . , d, and all x ∈ Matd (F), we find using the explicit
formula (11.19) for r and (11.18) for the corresponding Poisson structure,
ξi j , ξk r (x) = 2 ξi j ∧ ξkl , [x ⊗ x, r]
= ∑ ξi j ∧ ξk , [x ⊗ x, Eab ∧ Eba ]
a<b
= ∑ ξi j ∧ ξk , xEab ∧ xEba − Eab x ∧ Eba x
a<b
= ∑ (δ j,b δ,a − δ j,a δ,b − δi,a δk,b + δi,b δk,a ) ξi (x)ξk j (x)
a<b
= (ε j, + εi,k ) ξi (x)ξk j (x) ,
for all x ∈ g, and for all ξ , η ∈ g∗ . We will use the natural bilinear form · | · d on d,
defined for all x, y ∈ g and for all ξ , η ∈ g∗ by
11.2 Lie Bialgebras 303
[(x, ξ ), (y, η )]d := ([x, y]g +ad∗ξ y−ad∗η x, [ξ , η ]g∗ +ad∗x η −ad∗y ξ ) , (11.23)
If any of these three equivalent conditions is satisfied, the Lie bracket on d, which
satisfies the conditions demanded in (i), is unique and is given by (11.23); moreover,
it makes (d, · | · d ) into a quadratic Lie algebra.
The proof of this proposition will make use of the following lemma.
Lemma 11.16. Let g be a finite-dimensional vector space, equipped with Lie brack-
ets [· , ·]g and [· , ·]g∗ on g, respectively on g∗ . The skew-symmetric bilinear map
[· , ·]d : d × d → d, defined in (11.23), has the following properties:
(1) For all d1 , d2 , d3 ∈ d:
(2) [· , ·]d satisfies the Jacobi identity if and only if for all x, y ∈ g → d, and for all
ξ , η ∈ g∗ → d:
Since the latter expression is invariant with respect to a cyclic permutation of the
indices 1, 2, 3, it follows that
which proves (1). Let ϕ denote the 4-linear map from d to itself, defined for all
d1 , . . . , d4 ∈ d by
ϕ (d1 , d2 , d3 , d4 ) := Jd (d1 , d2 , d3 ) | d4 d ,
Since Jd = 0 if and only if [· , ·]d satisfies the Jacobi identity, and since · | · d is
non-degenerate, ϕ = 0 if and only if [· , ·]d satisfies the Jacobi identity.
In order to prove (2), we therefore only need to show that (11.26) holds for all
x, y ∈ g and all ξ , η ∈ g∗ , if and only if ϕ = 0. Since Jd (d1 , d2 , d3 ) = 0 when
d1 , d2 , d3 all belong to g or all belong to g∗ , we have that ϕ (d1 , d2 , d3 , d4 ) = 0 when
d1 , d2 , d3 all belong to g or all belong to g∗ , whatever the value of d4 ∈ d. The
skew-symmetry of Jd , combined with (11.25), implies that ϕ is skew-symmetric.
Therefore, ϕ (d1 , d2 , d3 , d4 ) = 0 as soon as at least three of the di all belong to g,
or all belong to g∗ . It follows, since ϕ is 4-linear, that ϕ = 0 if and only if
ϕ (d1 , d2 , d3 , d4 ) = 0, whenever two of the di belong to g and the two other belong
to g∗ ; but that is, in view of (11.25), precisely the condition that (11.26) holds for
all x, y ∈ g and all ξ , η ∈ g∗ . This establishes the proof of (2).
We are now ready to give the proof of Proposition 11.15.
Proof. We first prove that (i) and (ii) are equivalent. Suppose first that (ii) is sat-
isfied. Then the Lie bracket, defined by the explicit formula (11.23) has in view
of (1) in Lemma 11.16 the properties announced in (i). Suppose next that there ex-
ists a Lie bracket [· , ·] on d, with the properties announced in item (i). We show that
[· , ·] = [· , ·]d , i.e., [· , ·] is given by the explicit formula (11.23). In view of (i), [· , ·]
and [· , ·]d agree when both arguments are taken either from g or from g∗ , so we only
need to show that
[x, η ] = (− ad∗η x, ad∗x η ) , (11.28)
for all x ∈ g and all η ∈ g∗ . For (z, γ ) ∈ g × g∗ , we have in view of (11.25), combined
with the properties of [· , ·], assumed in (i), that
Since · | · d is non-degenerate, this shows (11.28), hence that if there exists a Lie
bracket on d, satisfying (i), then it is given by (11.23). In particular, it shows that (i)
and (ii) are equivalent.
We now prove the equivalence of items (ii) and (iii). In view of item (2) of
Lemma 11.16, we need to show that (11.26) holds for all x, y ∈ g, and for all
ξ , η ∈ g∗ , if and only if (11.24) holds for all x, y ∈ g. We claim that, to do this,
it suffices to show that the following formulas hold, for all x, y ∈ g, and for all
ξ , η ∈ g∗ :
ξ ∧ η , [[δ (x), y]] = ad∗y ξ , ad∗η x − ad∗y η , ad∗ξ x . (11.29)
Indeed, using (11.21 ), (11.29) and the definitions of [· , ·]d and · | · d , we find that
ξ ∧ η , δ ([x, y]g ) − [[δ (x), y]] − [[x, δ (y)]]
= [ξ , η ]g∗ , [x, y]g − ad∗y ξ , ad∗η x + ad∗y η , ad∗ξ x
+ ad∗x ξ , ad∗η y − ad∗x η , ad∗ξ y
= [x, y]d | [ξ , η ]d d+ [y, ξ ]d | [x, η ]d d+ [ξ , x]d | [y, η ]d d ,
for all x, y ∈ g, and for all ξ , η ∈ g∗ . In order to show (11.29), use (5.13) and the
derivation property of the coadjoint action of g on ∧2 g∗ :
ξ ∧ η , [[δ (x), y]] = − ξ ∧ η , ady δ (x) = ad∗y (ξ ∧ η ), δ (x)
= ad∗y ξ ∧ η + ξ ∧ ad∗y η , δ (x)
= ad∗y ξ , η g∗ + ξ , ad∗y η g∗ , x
= ad∗y ξ , ad∗η x − ad∗y η , ad∗ξ x .
This proves (11.29) and hence also the equivalence of (ii) and (iii).
for some λ ∈ F∗ . In this case δ (x) = λ x ∧ y and δ (y) = 0, so that we arrive at the
same conclusion.
Remark 11.20. For every λ , μ ∈ F and every Lie bialgebra (g, [· , ·]g , [· , ·]g∗ ), the
triple (g, λ [· , ·]g , μ [· , ·]g∗ ) is a Lie bialgebra. It is clear that it is isomorphic to the
original one if λ = μ ∈ F∗ , under the isomorphism λ 1g : g → g. However, in general,
it is not the case when λ = μ , even if both λ and μ are different from 0. For instance,
for two different values of the parameter λ , the Lie bialgebra structures described
by (11.30) are not isomorphic.
Proof. For h a subspace of g, let ı : h → g denote the inclusion map, with transpose
ı : g∗ → h∗ . If a Lie sub-bialgebra structure on h exists, making the inclusion map
into a homomorphism of Lie bialgebras, then the Lie bracket on h is the restriction
of the Lie bracket on g (and condition (1) is satisfied); also, ı : g∗ → h∗ is a Lie
algebra homomorphism, so that its kernel h⊥ is a Lie ideal of g∗ (and condition (2)
is satisfied).
Conversely, if the two conditions are satisfied, the Lie brackets on g and on g∗
induce Lie brackets on h (by restriction) and on h∗ (by taking the quotient with
respect to h⊥ ), hence equip h and h∗ with two Lie brackets [· , ·]h and [· , ·]h∗ . Since
the transpose δ of [· , ·]h∗ is the restriction of δ to h, item (iii) in Proposition 11.15
implies that these brackets define a Lie bialgebra structure on h. By construction, ı
is a homomorphism of Lie bialgebras.
Proposition 11.23. Let (g, [· , ·]g , [· , ·]g∗ ) be a Lie bialgebra, where g is a finite-
dimensional vector space. The triple (g∗ , [· , ·]g∗ , [· , ·]g ) is a Lie bialgebra, called the
dual of the Lie bialgebra (g, [· , ·]g , [· , ·]g∗ ). The natural isomorphism S : g × g∗ →
g∗ × g, given for all (x, ξ ) ∈ g × g∗ by S(x, ξ ) := (ξ , x) is a Lie algebra isomorphism
of the doubles of these two Lie bialgebras.
Proof. In item (i) in Proposition 11.15, the Lie algebras (g, [· , ·]g ) and (g∗ , [· , ·]g∗ )
play a symmetric rôle, hence the first part of the proposition is clear. The isomor-
phism S : g × g∗ → g∗ × g which sends (x, ξ ) to (ξ , x) is then clearly a Lie alge-
bra isomorphism between the doubles of the Lie bialgebras (g, [· , ·]g , [· , ·]g∗ ) and
(g∗ , [· , ·]g∗ , [· , ·]g ), which is the second part of the proposition. Notice that S arises
as the composition of the isomorphism g × g∗ (g × g∗ )∗ , induced by · | · d , with
the natural isomorphism (g × g∗ )∗ g∗ × g.
It is clear that the dual of the dual of a (finite-dimensional) Lie bialgebra is just the
Lie bialgebra itself. It is also obvious from Definition 11.17 that the transpose of a
homomorphism of Lie bialgebras is a Lie bialgebra homomorphism between their
dual Lie bialgebras. We leave it as an exercise to the reader to show that, if h is a
Lie sub-bialgebra of g, then there exists a (unique) Lie bialgebra structure on the
quotient space g∗ /h⊥ such that the projection map g∗ → g∗ /h⊥ is a Lie bialgebra
homomorphism.
308 11 Poisson–Lie Groups
Proposition 11.24. Let (g, [· , ·]g ) be a Lie algebra, and let r ∈ g ⊗ g be an r-matrix
of g. Then the coboundary Lie bialgebra (g, [· , ·]g , [· , ·]r ) is a Lie bialgebra.
Proof. Consider the coboundary Lie bialgebra (g, [· , ·]g , [· , ·]r ), where r is an r-
matrix of g. Since both [· , ·]g and [· , ·]r are Lie brackets, we only need to verify
that condition (iii) in Proposition 11.15 holds, where we recall from (11.21) that
δ : g → g ∧ g is the linear map, defined by
ξ ∧ η , δ (x) = [ξ , η ]r , x ,
so that δ (x) = adx r− , where r− denotes the skew-symmetric part of r. Using the
fact that ad is a Lie algebra representation of g (on g ∧ g),
δ ([x, y]g ) = ad[x,y]g r− = adx ady r− − ady adx r− = adx δ (y) − ady δ (x)
for all x, y ∈ g. This proves that condition (iii) in Proposition 11.15 holds, hence that
(g, [· , ·]g , [· , ·]r ) is a Lie bialgebra.
Not every Lie bialgebra is a coboundary Lie bialgebra. When the Lie algebra g is
abelian, recall from Example 11.18 that (g∗ , [· , ·]g = 0, [· , ·]g∗ ) is a Lie bialgebra for
every Lie bracket [· , ·]g∗ on g∗ . Such a Lie bialgebra cannot be a coboundary Lie
bialgebra, unless [· , ·]g∗ is trivial.
We show in the following proposition that the double of a Lie bialgebra is a
coboundary Lie bialgebra. Thus we can embed every Lie bialgebra in a coboundary
Lie bialgebra, a fact which will be useful when we prove, in the next section, that
every Lie bialgebra is the Lie bialgebra of a Poisson–Lie group.
Proposition 11.25. Let (g, [· , ·]g , [· , ·]g∗ ) be a finite-dimensional Lie bialgebra. On
its double (d, [· , ·]d ), there exists a skew-symmetric r-matrix a ∈ ∧2 d, with corre-
sponding Lie bracket [· , ·]a on d∗ , such that (g, [· , ·]g , [· , ·]g∗ ) is a Lie sub-bialgebra
of (d, [· , ·]d , [· , ·]a ). In terms of a pair of dual bases (e1 , . . . , ed ) and (ε1 , . . . , εd ) for
g and for g∗ respectively, a is given by a = 12 ∑di=1 ei ∧ εi .
11.2 Lie Bialgebras 309
for all (ξ , y), (η , z) ∈ g∗ × g and for all x ∈ g. This is done by explicitly computing
the left-hand side of the latter equation, namely
1 d
= ∑ (ξ , y) ∧ (η , z), [(x, 0), (ei , 0)]d ∧ (0, εi ) + (ei , 0) ∧ [(x, 0), (0, εi )]d
2 i=1
1 d
= ∑ (ξ , y) ∧ (η , z), ([x, ei ]g , 0) ∧ (0, εi ) + (ei , 0) ∧ (− ad∗εi x, ad∗x εi )
2 i=1
1 d
= ∑ ξ , [x, ei ]g εi , z − η , [x, ei ]g εi , y
2 i=1
1 d
+ ∑
2 i=1
ξ , ei ad∗x εi , z − η , ad∗εi x − η , ei ad∗x εi , y − ξ , ad∗εi x
1 d
=− ∑
2 i=1
ξ , ei η , ad∗εi x − η , ei ξ , ad∗εi x
= [ξ , η ]g∗ , x .
Definition 11.27. Let (E, [· , ·]) be a finite-dimensional quadratic Lie algebra, whose
bilinear form is denoted by · | · . Let V and W be vector subspaces of E. The triple3
((E, [· , ·], · | · ),V,W ) is said to be a Manin triple if
(1) E = V ⊕W ;
(2) V and W are Lie subalgebras of E;
(3) V and W are isotropic with respect to · | · .
Proof. Let (g, [· , ·]g , [· , ·]g∗ ) be a finite-dimensional Lie bialgebra, with double
d := g × g∗ . Identify, as before, g and g∗ as Lie subalgebras of d, via the natural
identifications g g × {0} and g∗ {0} × g∗ , so that d = g ⊕ g∗ . Clearly, both g
and g∗ are isotropic with respect to · | · d , so that ((d, [· , ·]d , · | · d ), g, g∗ ) is a Manin
triple. For the converse, let ((E, [· , ·]E , · | · E ),V,W ) be a Manin triple, with E finite-
dimensional. Then · | · E induces an isomorphism between V ∗ and W , so we obtain
a Lie algebra structure on V and on V ∗ . In view of (i) in Proposition 11.15 and be-
cause (E, · | · E ) is quadratic, (V, [· , ·]V , χ∗ ([· , ·]W )) is a Lie bialgebra. It is easily
verified that both constructions are inverse to each other.
3 The triple is usually simply written as (E,V,W ).
11.2 Lie Bialgebras 311
Notice that, inverting the rôles of V and W in a Manin triple, corresponds at the
bialgebra level to taking the dual.
Example 11.29. We construct a natural Manin triple which corresponds to Exam-
ple 10.15. Thus, we assume that (g, [· , ·]) is a finite-dimensional Lie algebra and
that g = g+ ⊕ g0 ⊕ g− is a Lie algebra decomposition, where g+ and g− are Lie
subalgebras of g and that g0 is abelian and is contained in the normalizer of g+
and in the normalizer of g− . Moreover, it is assumed that g comes equipped with
a non-degenerate symmetric bilinear form · | · with respect to which both g+ and
g− are isotropic and orthogonal to g0 . Recall that the endomorphism R of g, de-
fined for x ∈ g by Rx := x+ − x− , where x+ , x0 and x− stand for the projections
of x on g+ , g0 and g− respectively, is a skew-symmetric solution to the modified
Yang–Baxter equation (10.4), with c = 1, hence leads to a skew-symmetric r-matrix
for g, making (g, [· , ·], [· , ·]r ) into a Lie bialgebra. Let E := g × g, equipped with
the product Lie bracket, denoted by [· , ·] , and with the ad-invariant symmetric non-
degenerate bilinear form · | · , which is defined, for all x1 , x2 , y2 , y2 ∈ g, by
W := {(x + z, y − z) | x ∈ g+ , y ∈ g− and z ∈ g0 } .
ψ (x + z, y − z), ψ (x + z , y − z ) R
= x − y + 2z, x − y + 2z R
1
= ( x + y, x − y + 2z + x − y + 2z, x + y )
2
312 11 Poisson–Lie Groups
= x, x − y, y + x, z + z, x + y, z + z, y
= x + z, x + z − y − z, y − z
= ψ ( x + z, x + z , y − z, y − z )
= ψ ( (x + z, y − z), (x + z , y − z ) ) .
Notice that, considering the particular case of g0 = {0}, the constructed Manin triple
corresponds to the case of a Lie algebra splitting (see Example 10.14).
Example 11.30. We also construct the Manin triple which corresponds to Exam-
ple 10.16. To do this, we consider the complex Lie algebra sld (C) as a quadratic
Lie algebra over R, equipped
with the bilinear form, defined for x, y ∈ sld (C) by
x | y ℑ := ℑ Trace(xy) . We consider two Lie subalgebras: t+ , the space of upper
triangular matrices admitting only real coefficients on the diagonal, and sud . We
have that sld (C) = sud ⊕ t+ . For all x, y ∈ sud ,
We have seen in Proposition 11.25 that the double of a Lie bialgebra is also a Lie
bialgebra. We construct in the following proposition the Manin triple of this double
Lie bialgebra in terms of the Manin triple of the original bialgebra.
Proposition 11.31. Let ((E, [· , ·], · | · ),V,W ) be a Manin triple, where E is finite-
dimensional. Let E × E be equipped with the product Lie bracket, which we denote
by [· , ·] , and with the non-degenerate ad-invariant symmetric bilinear form · | · ,
given for (x1 , y1 ), (x2 , y2 ) ∈ E × E by
Let
V := {(u, u) ∈ E × E | u ∈ E} ,
(11.33)
W := W ×V .
11.2 Lie Bialgebras 313
Then ((E × E, [· , ·] , · | · ),V ,W ) is a Manin triple and the Lie bialgebra associ-
ated to it is the double of the Lie bialgebra associated to (E,V,W ).
Proof. Let ((E, [· , ·], · | · ),V,W ) be a Manin triple and let (g, [· , ·]g , [· , ·]g∗ ) be its
associated Lie bialgebra. According to Proposition 11.28, we may identify (E, [· , ·])
with (d = g × g∗ , [· , ·]d ); under this identification, (V, [· , ·]|V ) and (W, [· , ·]|W ) are
identified with (g, [· , ·]g ) and (g∗ , [· , ·]g∗ ) respectively; the inner product · | · on
E corresponds to the natural inner product (11.22) on d. According to Proposi-
tion 11.25, d is a coboundary Lie bialgebra, where the Lie bracket on d∗ is the
r-bracket a which is associated to the Lie algebra splitting d = g∗ ⊕ g, i.e., the
Lie algebra splitting E = W ⊕ V . According to Example 11.29, in the particular
case of g0 = {0}, the Manin triple of such a coboundary Lie algebra is given by
((E × E, [· , ·] , · | · ),V ,W ), where [· , ·] is the product Lie bracket on E × E and
where · | · and V , W are given by (11.32) and (11.33).
Proof. Let [· , ·]g∗ be the Lie bracket on g∗ which corresponds to the linear Pois-
son structure π = {· , ·} on g. For F, G ∈ F (g), their Poisson bracket at x ∈ g is,
according to (7.4) and (11.21), given by
{F, G} (x) = [dx F, dx G]g∗ , x = dx F ∧ dx G, δ (x) ,
ψ : (g, [· , ·]g ) → (h, [· , ·]h ) is a homomorphism of Lie algebras and ψ : (g, {· , ·}g ) →
(h, {· , ·}h ) is a Poisson map. Indeed, the latter condition is equivalent to the fact that
ψ : (h∗ , [· , ·]h∗ ) → (g∗ , [· , ·]g∗ ) is a Lie algebra homomorphism. It follows that Lie
sub-bialgebras can be characterized as follows:
Proposition 11.33. Let (g, [· , ·]g , [· , ·]g∗ ) be a Lie bialgebra, and denote by π the
linear Poisson structure on g corresponding to [· , ·]g∗ . A subspace h ⊂ g is a Lie
sub-bialgebra if and only if the following two conditions are satisfied:
(1) h is a Lie subalgebra of the Lie algebra (g, [· , ·]g );
(2) h is a Poisson submanifold of the Poisson manifold (g, π ).
In this section, we relate the objects which appeared in the two previous sections,
namely we establish a natural correspondence between Poisson–Lie groups and
finite-dimensional Lie bialgebras. We first show that the tangent space to the unit
of a Poisson–Lie group inherits the structure of a Lie bialgebra (Section 11.3.1).
We then show that every finite-dimensional Lie bialgebra is the Lie bialgebra
of a Poisson–Lie group. This is first done in Section 11.3.2 for coboundary Lie
bialgebras and then in general in Section 11.3.3, using the fact that every finite-
dimensional Lie bialgebra can be embedded in a coboundary Lie bialgebra (its dou-
ble, see Proposition 11.25).
Proposition 11.34. Let (G, π ) be a Poisson–Lie group, with Lie algebra (g, [· , ·]g ),
and with linearized bracket [· , ·]g∗ .
(1) The triple (g, [· , ·]g , [· , ·]g∗ ) is a Lie bialgebra;
(2) Let Φ be a Poisson–Lie group homomorphism from (G, π ) to a Poisson–Lie
group (H, π ), with corresponding Lie bialgebra (h, [· , ·]h , [· , ·]h∗ ). The tan-
gent map Te Φ of Φ at the unit e of G is a Lie bialgebra homomorphism from
(g, [· , ·]g , [· , ·]g∗ ) to (h, [· , ·]h , [· , ·]h∗ ).
11.3 Poisson–Lie Groups and Lie Bialgebras 315
Proof. Let (G, π ) be a Poisson–Lie group with Lie algebra (g, [· , ·]g ) and let Ψ
denote the cocycle G → ∧2 g, defined by (11.2). We show that the cocycle con-
dition (11.3) implies condition (ii) in Proposition 11.32, which guarantees that the
triple (g, [· , ·]g , [· , ·]g∗ ) is a Lie bialgebra, where [· , ·]g∗ denotes the linearized bracket
of (G, π ). Consider the identity
which is valid for all g, h ∈ G, as follows from the cocycle condition (11.3), upon
using that Ψ (e) = 0. Differentiating (11.35) leads to an identity for the linearized
Poisson bracket π1 , since
(π1 )x = TeΨ (x) (11.36)
for x ∈ g (see (2) in Proposition 11.5). In order to obtain this identity, replace h
by exp(ty) in (11.35) and take the derivative at t = 0. In view of (11.36), this leads to
for all y ∈ g and g ∈ G. We replace now in (11.37) g by exp(tx) and we take the
derivative at t = 0; we claim that we obtain
which in view of Proposition 11.32, shows that the triple (g, [· , ·]g ), [· , ·]g∗ ) is a Lie
bialgebra.
Let us show in detail how (11.38) is obtained from (11.37). Since π1 is a linear
Poisson structure, the map g → ∧2 g, defined by x → (π1 )x is a linear map. It follows
that, for every function z : F → g,
d
(π1 )z(t) = (π1 ) d | z(t) .
dt t=0 dt t=0
Applied to the left-hand side of (11.37), with g = exp(tx), this yields the left-hand
side of (11.38), in view of the definition of ad. Also,
d
Ad (π1 )y = adx (π1 )y ,
dt t=0 exp(tx)
which yields the first term in the right-hand side of (11.37). Finally, since Ψ (e) = 0,
d d
ad Ψ (exp(tx)) = ad Ψ (exp(tx))
dt t=0 dt t=0
Adexp(tx) y y
which is the second term in the right-hand side of (11.37). This completes the proof
of (11.38), and hence of item (1).
316 11 Poisson–Lie Groups
The second item of the previous proposition implies that the Lie bialgebra of a
Poisson–Lie subgroup of (G, π ) is a Lie sub-bialgebra of the bialgebra of (G, π ).
This is stated, together with its converse, in the following proposition.
Proposition 11.35. Let (G, π ) be a Poisson–Lie group, whose Lie bialgebra is de-
noted by (g, [· , ·]g , [· , ·]g∗ ). For a connected Lie subgroup H of G, with Lie algebra h,
the following conditions are equivalent:
(i) H is a Poisson–Lie subgroup of (G, π );
(ii) h is a Lie sub-bialgebra of (g, [· , ·]g , [· , ·]g∗ ).
Assuming that these equivalent conditions hold, denote the Lie brackets of the Lie
bialgebra h by [· , ·]h and [· , ·]h∗ , and denote the Poisson structure on H, which
makes H into a Poisson submanifold of (G, π ), by π . Then (h, [· , ·]h , [· , ·]h∗ ) is the
Lie bialgebra of (H, π ).
Proof. The implication (i) ⇒ (ii) follows at once from Proposition 11.34, since
the inclusion map of a Poisson–Lie subgroup is a Poisson–Lie group homomor-
phism. Conversely, assume that h is a Lie sub-bialgebra of the Lie bialgebra
(g, [· , ·]g , [· , ·]g∗ ) of (G, π ). Then the transpose δ : g → ∧2 g of [· , ·]g∗ maps h to
∧2 h. According to item (2) in Proposition 11.5 and the proof of Proposition 11.32,
δ (x) = TeΨ (x) for all x ∈ g, where Ψ is defined as usual by (11.2). It follows that
TeΨ (x) ∈ ∧2 h for all x ∈ h. According to Proposition 11.7, this shows that the con-
nected Lie subgroup H of G is a Poisson–Lie subgroup. Thus, condition (ii) implies
condition (i).
We show in this section that the Lie bialgebra of a coboundary Poisson–Lie group
is a coboundary Lie bialgebra, which justifies why the same adjective “cobound-
ary” is used for both structures. We also show the following converse: every
finite-dimensional coboundary Lie bialgebra is the Lie bialgebra of a coboundary
Poisson–Lie group.
We first show that there corresponds to every coboundary Poisson–Lie group a
coboundary Lie bialgebra. Let (G, π ) be a coboundary Poisson–Lie group. By defi-
nition, there exists a ∈ ∧2 g such that π = ← −
a −−→a , where (g, [· , ·]) is the Lie algebra
of G. According to Proposition 11.10, [[a, a]] is Ad-invariant, hence is ad-invariant.
Proposition 10.11 implies that a is a skew-symmetric r-matrix, i.e., the associated
bracket [· , ·]a is a Lie bracket on g∗ , making (g, [· , ·], [· , ·]a ) into a coboundary Lie
11.3 Poisson–Lie Groups and Lie Bialgebras 317
bialgebra (see Definition 10.9). The next proposition shows that this coboundary Lie
bialgebra is the Lie bialgebra of the Poisson–Lie group (G, π ).
Proposition 11.36. Let (G, π ) be a coboundary Poisson–Lie group, with π = ← −
a −→−
a
for some r-matrix a ∈ ∧2 g, where (g, [· , ·]g ) denotes the Lie algebra of G. The Lie
bialgebra of (G, π ) is the coboundary Lie bialgebra (g, [· , ·]g , [· , ·]a ).
Corollary 11.37. Let (g, [· , ·]g , [· , ·]a ) be a finite-dimensional coboundary Lie bial-
gebra. If G is a connected Lie group, with Lie algebra (g, [· , ·]g ), then G ad-
mits the structure of a coboundary Poisson–Lie group, whose Lie bialgebra is
(g, [· , ·]g , [· , ·]a ). In particular, every finite-dimensional coboundary Lie bialgebra
is the Lie bialgebra of a coboundary Poisson–Lie group.
Proof. We only need to prove the first part of the corollary, since the last state-
ment follows from it at once upon using the fact that every Lie algebra is the Lie
algebra of a connected Lie group (see Theorem 5.1). Let (g, [· , ·]g , [· , ·]a ) be a finite-
dimensional coboundary Lie bialgebra, with a ∈ ∧2 g so that [[a, a]] is ad-invariant.
Let G be a connected Lie group whose Lie algebra is g. Since G is assumed to be
connected, [[a, a]] is Ad-invariant. Setting π := ←
−
a −−
→
a , Propositions 11.10 and 11.36
imply that (G, π ) is a Poisson–Lie group, whose Lie bialgebra is (g, [· , ·]g , [· , ·]a ), as
was to be shown.
Corollary 11.38. Let (g, [· , ·]g , [· , ·]g∗ ) be a finite-dimensional Lie bialgebra and
let D be a connected Lie group whose Lie algebra is the double (d, [· , ·]d ) of
(g, [· , ·]g , [· , ·]g∗ ). Then D admits a unique Poisson structure πD , making (D, πD ) into
a coboundary Poisson–Lie group, whose Lie bialgebra is (d, [· , ·]d , [· , ·]a ). In terms
of a pair of dual bases (e1 , . . . , ed ) and (ε1 , . . . , εd ) for g and for g∗ respectively, the
skew-symmetric r-matrix a of d is given by a = 12 ∑di=1 ei ∧ εi .
Proof. According to Proposition 11.25, the double (d, [· , ·]d ) is a coboundary Lie
bialgebra; as before, we denote the corresponding Lie structure on d∗ by [· , ·]a . If D
is a connected Lie group with Lie algebra (d, [· , ·]d ), then D admits, according to
Corollary 11.37, the structure of a coboundary Poisson–Lie group, whose Lie bial-
gebra is (d, [· , ·]d , [· , ·]a ). As we will see in Theorem 11.39, the Poisson structure
on D which integrates the given Lie bialgebra structure is unique.
318 11 Poisson–Lie Groups
We show in this section that every finite-dimensional Lie bialgebra can be inte-
grated. We mean by this that there exists for every finite-dimensional Lie bialgebra
(g, [· , ·]g , [· , ·]g∗ ) a (not necessarily unique) Poisson–Lie group (G, π ) whose Lie
bialgebra is isomorphic to (g, [· , ·]g , [· , ·]g∗ ). For the particular case of a coboundary
Lie bialgebras, this was proved in Corollary 11.37.
Theorem 11.39. Let (g, [· , ·]g , [· , ·]g∗ ) be a finite-dimensional Lie bialgebra and
let G be a connected Lie group with Lie algebra (g, [· , ·]g ).
(1) There exists at most one Poisson structure π on G such that (G, π ) is a
Poisson–Lie group, whose Lie bialgebra is (g, [· , ·]g , [· , ·]g∗ );
(2) If G is simply connected, then there exists on G a unique Poisson struc-
ture π , such that (G, π ) is a Poisson–Lie group, whose Lie bialgebra is
(g, [· , ·]g , [· , ·]g∗ );
(3) In particular, every finite-dimensional Lie bialgebra is the Lie bialgebra of a
Poisson–Lie group.
Proof. Let π (1) and π (2) be two multiplicative Poisson structures on the connected
Lie group G, leading to the same Lie bialgebra, i.e., their linearized Poisson struc-
tures at e coincide. In view of (2) in Proposition 11.5, TeΨ (1) (x) = TeΨ (2) (x), for all
(i)
x ∈ g, where Ψ (i) : G → ∧2 g is defined by Ψ (i) (g) := ∧2 (Tg Rg−1 )πg , for i = 1, 2. It
follows that Ψ (1) − Ψ (2) , which is a cocycle (since both Ψ (1) and Ψ (2) are cocycles),
has zero tangent map at e. By Proposition 11.4, this implies that Ψ (1) − Ψ (2) = 0,
and hence that π (1) = π (2) . This shows item (1).
Let (g, [· , ·]g , [· , ·]g∗ ) be a finite-dimensional Lie bialgebra. According to Propo-
sition 11.25, (g, [· , ·]g , [· , ·]g∗ ) is a Lie sub-bialgebra of a coboundary Lie bialgebra,
which we denote by (d, [· , ·]d , [· , ·]a ), because the underlying Lie algebra is the dou-
ble of (g, [· , ·]g , [· , ·]g∗ ). In view of Proposition 11.37, (d, [· , ·]d , [· , ·]a ) can be inte-
grated into a Poisson–Lie group (D, πD ), which is the only fact about d which we
will use in this proof. We denote by Ĝ the unique connected Lie subgroup of D,
whose Lie algebra is g. According to Proposition 11.35, the subgroup Ĝ inherits a
multiplicative Poisson structure π̂ from πD , whose corresponding Lie bialgebra is
(g, [· , ·]g , [· , ·]g∗ ). This shows (3).
We show now that (Ĝ, π̂ ) can be replaced by the connected and simply con-
nected Lie group G whose Lie algebra is g, which is the content of (2). In view
of Lie’s theorem (see item (4) of Theorem 5.1), there exists a Lie group homomor-
phism Φ : G → Ĝ, whose induced Lie algebra homomorphism is the identity map.
We claim that there is a unique Poisson structure π on G which turns Φ : G → Ĝ into
a Poisson map, and we show that π is multiplicative. Then (G, π ) is a Poisson–Lie
group and Φ is a homomorphism of Poisson–Lie groups, so that the induced
Lie bialgebra homomorphism, which is the identity map, is a homomorphism of Lie
bialgebras, which means that the Lie bialgebra of G is (g, [· , ·]g , [· , ·]g∗ ).
Let us prove the claim. Differentiating the identity Φ = RΦ (g) Φ Rg−1 at g ∈ G
shows that the tangent maps of Φ at g ∈ G and at e are related by Tg Φ = Tê RΦ (g) ◦
11.3 Poisson–Lie Groups and Lie Bialgebras 319
Te Φ ◦ Tg Rg−1 (in this formula, e and ê are the units of the Lie groups G and Ĝ
respectively). Since the three linear maps which appear in the right-hand side of
this formula are invertible, so is Tg Φ , hence Φ is a local diffeomorphism. We can
therefore define π for all g ∈ G by setting
Φ ×Φ Φ
Ĝ × Ĝ μ̂
Ĝ
Combined with the fact that μ is a Poisson map, this shows that μ̂ is Poisson map.
As a consequence, π is multiplicative, as remained to be shown.
Having established the fact that every Lie bialgebra can be integrated into a Poisson–
Lie group, we now show that every Lie bialgebra homomorphism can be integrated
into a morphism of Poisson–Lie groups.
Theorem 11.40. Let (g, [· , ·]g , [· , ·]g∗ ) and (h, [· , ·]h , [· , ·]h∗ ) be two finite-dimensi-
onal Lie bialgebras and let (G, π ) and (H, π ) be Poisson–Lie groups which in-
tegrate them, where G is chosen connected and simply connected. For every Lie
bialgebra homomorphism φ : g → h, there exists a unique Poisson–Lie group homo-
morphism Φ : G → H whose induced Lie bialgebra homomorphism is φ , i.e., such
that Te Φ = φ , where e denotes the unit element of G.
for all g, h ∈ G. We first show that Ξ = 0. In view of item (1) of Proposition 11.4,
it suffices to show that Te Ξ (x) = 0 for all x ∈ g. To prove the latter, recall from
320 11 Poisson–Lie Groups
item (2) in Proposition 11.5 that TeΨ (x) = (π1 )x , for all x ∈ g, where π1 denotes the
linearized Poisson structure of π at e, and similarly for Ψ and π1 . It allows us to
compute
where the last equality, which states that φ is a Poisson map, is tantamount to the
fact that the transpose map φ : (h∗ , [· , ·]h∗ ) → (g∗ , [· , ·]g∗ ) is a Lie algebra homo-
morphism (see Section 11.2.6). It follows that Ξ = 0, leading to the equality
∧2 φ (Ψ (g)) = Ψ (Φ (g)) ,
valid for all g ∈ G, which can be written, using Te Φ = φ and using the defini-
tion (11.2) of the cocycle Ψ , as
Φ ◦ Rg−1 = RΦ (g−1 ) ◦ Φ ,
which, substituted in (11.40), leads to ∧2 (Tg Φ )πg = πΦ (g) , since in a Lie group,
right translation by an element of the group is a diffeomorphism. In view of Propo-
sition 1.19, the latter equality implies that Φ is a Poisson map, as was to be shown.
The notion of duality for Lie bialgebra leads to a notion of duality for connected,
simply connected Poisson–Lie groups.
Definition 11.41. Two Poisson–Lie groups (G, π ) and (G∗ , π ∗ ) are said to be dual
to each other if their Lie bialgebras are dual to each other.
Let (G, π ) be a Poisson–Lie group and let (g, [· , ·]g , [· , ·]g∗ ) denote its Lie bialge-
bra, whose dual bialgebra is (g∗ , [· , ·]g∗ , [· , ·]g ). Every Poisson–Lie group (G∗ , π ∗ )
which integrates the latter Lie bialgebra is dual to (G, π ). A canonical dual is picked
by demanding that G∗ be connected and simply connected. Indeed, there is up to
isomorphism a unique connected and simply connected Lie group G∗ whose Lie
algebra is (g∗ , [· , ·]g∗ ) (item (3) of Theorem 5.1) and on G∗ there exists a unique
Poisson structure π ∗ such that (G∗ , π ∗ ) integrates (g, [· , ·]g , [· , ·]g∗ ). This Poisson–
Lie group is called the dual of (G, π ). For (G, π ) a connected and simply connected
Poisson–Lie group, taking twice the dual gives back the Poisson–Lie group (G, π ).
Example 11.42. Let G be a connected, simply connected Lie group, with Lie algebra
(g, [· , ·]g ). Consider g∗ as a Lie group, where the group operation is addition. In
view of Corollary 11.9, g∗ becomes a Poisson–Lie group, when it is equipped with
the linear Poisson structure π on g∗ corresponding to the Lie bracket [· , ·]g . Its Lie
11.4 Dressing Actions and Symplectic Leaves 321
bialgebra is (g∗ , 0, [· , ·]g ), with dual Lie bialgebra (g, [· , ·]g , 0). It follows that the
(connected and simply connected) dual to the Poisson–Lie group (g∗ , π ) is the Lie
group G, equipped with the trivial Poisson structure.
We show in this section that the symplectic leaves of a Poisson–Lie group can be
described by using dressing actions, a notion which we first define in the general
context of Lie groups.
Let G be a connected Lie group and let G1 and G2 be two Lie subgroups of G.
The triple (G, G1 , G2 ) is said to be complete if the map G1 × G2 → G given by
(g1 , g2 ) → g1 g2 is a diffeomorphism. Since (g1 g2 )−1 = g−1 −1
2 g1 , for all g1 , g2 ∈ G,
the triple (G, G1 , G2 ) is complete if and only if the triple (G, G2 , G1 ) is complete. It
follows that, when (G, G1 , G2 ) is complete, there exists for every (g1 , g2 ) ∈ G1 ×G2
a unique pair (g1 , g2 ) ∈ G1 × G2 , such that g1 g2 = g2 g1 . It is customary to denote
the elements g1 and g2 by gg12 and gg21 respectively.
The associativity of the group structure on G implies that the assignments
−1
(g, h) → hg and (g, h) → gh are group actions. Indeed, the relation (g1 h1 )g2 =
g1 (h1 g2 ) implies, for g1 , h1 ∈ G1 and g2 ∈ G2 , that
h
(g 1 )
gg21 h1 (g1 h1 )g2 = g1 gh21 hg12 = (gh21 )g1 g1 2 hg12 ,
valid for all g1 , h1 ∈ G1 and g2 ∈ G2 . This shows that the assignment (g1 , g2 ) → gg21
defines a group action of G1 on (the manifold) G2 . Similarly, one derives from the
relation g1 (g2 h2 ) = (g1 g2 )h2 , for g1 ∈ G1 and g2 , h2 ∈ G2 , the relation
g−1
which says that the assignment (g2 , g1 ) → g12 defines a group action of G2 on (the
manifold) G1 . These two group actions are respectively called the dressing action
of G1 on G2 and the dressing action of G2 on G1 .
In the case of Poisson–Lie groups, the notion of dressing action leads to the
following definition.
Definition 11.44. Let (G, π ) be a connected Poisson–Lie group with Lie bialgebra
(g, [· , ·]g , [· , ·]g∗ ). A Lie group D is called a dressing group of (G, π ) if the following
three conditions are satisfied.
(1) The Lie algebra of D is the double (d, [· , ·]d ) of (g, [· , ·]g , [· , ·]g∗ );
(2) G is the connected Lie subgroup of D corresponding to the Lie subalgebra
g ⊂ d;
(3) (D, G, G∗ ) is complete, where G∗ stands for the connected Lie subgroup of D
corresponding to the Lie subalgebra g∗ ⊂ d.
The Lie subgroup G∗ of D is then called the dual of (G, π ) with respect to the
dressing group D. The Poisson structure on G∗ as a Poisson submanifold of the
Poisson–Lie group (D, πD ) is denoted by π ∗ .
Example 11.45. A Lie bialgebra structure on g := sud was given in Example 10.16,
We showed in Example 11.30 that its double d = g ⊕ g∗ is isomorphic to the Lie
algebra sld (C) of traceless complex matrices, identification under which the Lie
algebras g and g∗ become the Lie algebras sud and t+ , where t+ is the Lie alge-
bra of T+ , i.e., the vector space of upper triangular matrices of trace zero with real
coefficients on the diagonal. Consider the Poisson bracket π on SUd which makes
(SUd , π ) into a Poisson–Lie group, integrating the latter bialgebra structure on sud .
Since (SLd (C), SUd , T+ ) is complete, as we have seen in Example 11.43, condi-
tions (1)–(3) in the above definition are verified, so that SLd (C) is a dressing group
of (SUd , π ).
(g, h) → gh, is a Poisson diffeomorphism. Since the same is true for the diffeomor-
phism G∗ × G → D, given by (h, g) → hg, the map G × G∗ → G∗ × G, defined by
(g, h) → (hg , gh ) is a Poisson diffeomorphism. This proves (1). As a corollary, the
map G×G∗ → G∗ , defined by (g, h) → hg , is a Poisson map, meaning that the dress-
ing action of G on G∗ is a Poisson map, proving (2). Since the inverse map h → h−1
is a Poisson map from (G∗ , −π ∗ ) to (G∗ , π ∗ ) (see Proposition 11.5, item (3)), the
−1
assignment (g, h) → gh is also a Poisson map, provided that G × G∗ is endowed
with the product Poisson structure (π , −π ∗ ), which means precisely that the dress-
ing action of G∗ on G is Poisson, provided that we replace the Poisson structure
π ∗ of G∗ by its opposite. This proves (3).
We now prove (4). For all g ∈ G and t ∈ F we have, by definition of the dressing
action
−1
g exp(−t α ) = h(t)gexp(−t α ) = h(t)gexp(t α ) , (11.44)
where h(t) = exp (−t α )g . Note that h(0) = e, and that h(t) is a path in G∗ . Multi-
plying (11.44) by g−1 and taking the derivative with respect to t at t = 0, we obtain
d exp(t α )−1 −1 d
− Adg α = g g + h(t)
dt t=0 dt t=0
d
= Tg Rg−1 α g + h(t) , (11.45)
dt t=0
where the last equality follows from the definition of the fundamental vector field
of a Lie group action. Since h(t) ∗
∗
is apath in G , (11.45) is the decomposition of
− Adg α in g ⊕ g , so that −Pg Adg α = Tg Rg−1 α g , which amounts to (11.43).
We can now describe the symplectic leaves of a Poisson–Lie group which admits
a dressing group, in terms of the dressing action.
Theorem 11.47. Let (G, π ) be a connected Poisson–Lie group. Assume that D is a
dressing group of (G, π ), and let G∗ ⊂ D be the dual of G with respect to D. Then
the symplectic leaves of (G, π ) are the orbits of the dressing action of G∗ on G.
Proof. Recall from Corollary 11.38 that D admits a Poisson structure πD , making
(D, πD ) into a coboundary Poisson–Lie group. This Poisson structure is associated
to the skew-symmetric r-matrix a := 12 ∑di=1 ei ∧ εi ∈ ∧2 d, where (e1 , . . . , ed ) and
(ε1 , . . . , εd ) are dual bases of g and g∗ . In view of (11.14), one has for every h ∈ D,
1 d 1 d
∧2 (Tg Rg−1 ) πg = ∑
2 i=1
Adg ei ∧ Adg εi − ∑ ei ∧ εi .
2 i=1
The left-hand side of this equation belongs to ∧2 g, hence is fixed by the linear map
∧2 Pg (with Pg the projection of d = g ⊕ g∗ onto g). It follows that
324 11 Poisson–Lie Groups
1 d
∧2 (Tg Rg−1 ) πg = ∑ Adg ei ∧ Pg(Adg εi ) ,
2 i=1
(11.46)
which, in turn, gives the following expression of the Poisson bivector π , evaluated
at the point g,
1 d −
πg = ∑ ( ← ei )g ∧ εi g , (11.47)
2 i=1
where we used item (3) of Proposition 11.46 as well as the identity Te Rg (Adg x) =
Te Lg (x) = ←
−x g for every x ∈ g.
We use the explicit expression (11.47) of πg to show that the symplectic leaf of π
through g coincides with the orbit through g of the dressing action of G∗ on G.
Since G∗ is connected and since (ε1 , . . . , εd ) is a basis of g∗ , it amounts to showing
that the image of the linear map πg : Tg∗ G → Tg G, defined by πg , is the vector sub-
space of Tg G generated by ε1 g , . . . , εd g . To show this, it suffices according to (11.47)
to show that the matrix (ai j )i, j=1,...,d expressing the elements εi g , i = 1, . . . , d, in
terms of the basis of tangent vectors (← e−1 )g , . . . , (←
e−
d )g is skew-symmetric, since
these elements are then the columns of the Poisson matrix of π at g, expressed
in terms of the latter basis. Notice that, in view of (11.46), this matrix is also the
matrix which expresses the elements Pg (Adg εi ), i = 1, . . . , g, in terms of the vec-
tors Adg e1 , . . . , Adg ed ,
d
Pg (Adg εi ) = ∑ ai j Adg e j ,
j=1
so that
ai j = Ad∗g ε j , Pg (Adg εi ) = Pg∗ (Adg ε j ), Pg (Adg εi ) ,
where Pg∗ is the projection of d = g ⊕ g∗ onto g∗ ; we also used the fact that
ad∗x ε j = Pg∗ ([x, (y, ε j )]d ) for all x, y ∈ g, which follows from the definition (11.23)
of [· , ·]d . Hence, to show that the matrix (ai j )i, j=1,...,d is skew-symmetric, we need
to show that
Pg∗ (Adg εi ), Pg (Adg ε j ) + Pg∗ (Adg ε j ), Pg (Adg εi ) = 0 (11.48)
in view of the Ad-invariance of · | · d . This proves (11.48), hence yields that the
matrix (ai j )i, j=1,...,d is skew-symmetric, which was to be shown.
11.5 Notes 325
11.5 Notes
Poisson structures can be seen as the semi-classical limit of quantum structures, i.e.,
as being what remains of quantization when h̄2 (but not h̄) is considered to be
small enough to be ignored (see Chapter 13). This is especially true for Poisson–Lie
groups which first appeared as semi-classical limits of quantum groups, in the pio-
neering work [58] of Drinfel’d. A second motivation for the introduction of Poisson–
Lie groups was the study of integrable systems associated to infinite-dimensional
Lie algebras, such as the Korteweg–de Vries equation, where Poisson–Lie groups
provide the hidden symmetry, see Semenov–Tian–Shansky [180].
Poisson–Lie groups have proved to be interesting on their own. First, they natu-
rally appear when studying the singularities of several highly singular spaces asso-
ciated to semi-simple Lie groups, like Bruhat and Schubert cells, see Lu–Weinstein
[132, 136]. Also, a Poisson–Lie group is the first instance of a more general ob-
ject, namely a Poisson–Lie groupoid, see Mackenzie–Xu [140], which appears natu-
rally when studying the dynamical Yang–Baxter equation. Exactly like Poisson–Lie
groups and Lie bialgebras are in one-to-one correspondence, Poisson–Lie groupoids
are in one-to-one correspondence with Lie bialgebroids, see Mackenzie–Xu [140].
One of the first systematic studies of Poisson–Lie groups is the PhD thesis of
Lu [132], which remains an excellent introduction to the subject, especially to those
who want to learn more about the relation with Lie groupoids and enlarged defi-
nitions of moment maps. It also deals with a slightly more general object: affine
Poisson structures on Lie groups. Another useful reference is the review article by
Kosmann–Schwarzbach [110], which insists in particular on the link with Lie bial-
gebras.
Part III
Applications
Chapter 12
Liouville Integrable Systems
In this chapter we present the main application of Poisson structures: the theory of
integrable Hamiltonian systems.
Poisson’s theorem, which states that the Poisson bracket of two constants of mo-
tion (of a Hamiltonian system) is a constant of motion, is from the modern point of
view a direct consequence of the formalism, but was in Poisson’s time considered
as a fundamental step towards the explicit integration of Hamilton’s equations, as
for doing this by the known methods, a large number of constants of motion was
required.
The fundamental rôle of Poisson structures in integrability was emphasized by
Liouville. As he showed, having half of the dimension of phase space of independent
constants of motion is sufficient for integrating the equations of motion, under the
assumption that these constants of motion are in involution (Poisson commute). Pre-
cisely, integration means here integration by quadratures. A geometrical counterpart
to this result states that, under some topological assumptions, the motion of such a
Hamiltonian system evolves on tori (the so-called Liouville tori), whose dimension
is half the dimension of phase space, and the motion on them is quasi-periodic. A
more elaborate version of this theorem is the action-angle theorem, which yields a
canonical model for Liouville integrable systems, in the neighborhood of a Liouville
torus.
Liouville’s theorems and the action-angle theorems will be presented here in
the context of general Poisson manifolds. These manifolds generalize the phase
spaces, considered in the classical results, namely the cotangent bundle of the con-
figuration space, equipped with its canonical symplectic structure. We give the ba-
sic definitions and properties of functions in involution and of the map, associ-
ated to them, in Section 12.1. Several constructions of functions in involution are
given in Section 12.2: Poisson’s theorem, the Hamiltonian form of Noether’s theo-
rem, bi-Hamiltonian vector fields, Thimm’s method, Lax equations and the Adler–
Kostant–Symes theorem. The action-angle theorem for Poisson manifolds is stated
and proved in Section 12.3.
In this section we give the basic definitions and the (geometrical) properties of
functions in involution. Throughout the section, all manifolds considered are real
(F = R) or complex (F = C). For such a manifold M, we denote its algebra of
smooth or holomorphic functions by F (M).
Definition 12.1. Let (M, π ) be a Poisson manifold and let F1 and F2 be two elements
of F (M). We say that F1 and F2 are in involution if their Poisson bracket is zero,
{F1 , F2 } = 0. More generally, let F = (F1 , . . . , Fs ) be an s-tuple of elements of F (M).
We say that F is involutive if for all 1 i, j s, the functions Fi and Fj are in
involution.
Classically, a function F ∈ F (M) which is constant on all the integral curves of a
Hamiltonian vector field XH , where H is a given function on a Poisson manifold
(M, π ), is called a constant of motion (for H). Since XH [F] = {F, H}, we have that
F is a constant of motion for H if and only if F and H are in involution. Therefore,
the notion of involutive s-tuples generalizes the notion of constant of motion.
An s-tuple F = (F1 , . . . , Fs ) of elements of F (M) defines a map M → Fs , which
will be denoted by the same letter F. We say that F is independent if the open
subset UF of M, defined by
UF := {m ∈ M | dm F1 ∧ · · · ∧ dm Fs = 0} , (12.1)
where 2r stands for the rank of π . Notice that, if F is independent, then the open
subset UF ∩ M(r) of M is non-empty.
When dealing with involutive functions, the reader should have the geometrical
picture in mind, described by the following proposition (see also Fig. 12.1).
Proposition 12.2. Let (M, π ) be a Poisson manifold and assume that F = (F1 , . . . , Fs )
is an involutive and independent s-tuple of elements of F (M). The Hamiltonian vec-
tor fields XFi , 1 i s, commute and they are tangent to the smooth fibers of the
map F : M → Fs .
Proof. The Hamiltonian vector fields, which are associated to functions in involu-
tion, commute, because [XF , XG ] = −X{F,G} for all F, G ∈ F (M). Assume that
F = (F1 , . . . , Fs ) is involutive and let 1 i s; since XFi [Fj ] = Fj , Fi = 0 for all
1 j s, the vector field XFi is tangent to the smooth level sets of F1 , . . . , Fs , that
is, to the smooth fibers of F.
The number of independent functions on a manifold M is bounded by the dimension
of M. In the case of a Poisson manifold, the upper bound for the number of indepen-
dent functions in involution is much lower, as given by the following proposition.
12.1 Functions in Involution 331
Fig. 12.1 The Hamiltonian vector fields of the components of F = (F1 , . . . , Fs ) are tangent to the
smooth fibers F−1 (c) of F and they commute.
Proposition 12.3. Let (M, π ) be a Poisson manifold of dimension d and rank 2r and
suppose that F = (F1 , . . . , Fs ) is an independent s-tuple of elements of F (M).
(1) If F1 , . . . , Fs are Casimirs, then s d − 2r;
(2) If F is involutive, then s d − r;
(3) If F is involutive, with s = d − r, then
dim span (XF1 )m , . . . , (XFs )m = r
Proof. For m ∈ M, let πm denote the linear map Tm∗ M → Tm M, which corresponds
to the Poisson structure. Explicitly, for H ∈ F (M),
Recall that the rank of πm is Rkm π . For every H ∈ Cas(M) the cotangent vector
dm H belongs to Ker πm , whose dimension is d − Rkm π . Let F = (F1 , . . . , Fs ) be
independent and let m be an element of the non-empty (open) subset UF ∩ M(r)
of M. Suppose first that each element of F is a Casimir. Since dm F1 , . . . , dm Fs are
independent elements of Ker πm , we have that
and (1) follows. Next, suppose that F is involutive and consider Fm , the fiber of F,
passing through m. Since m ∈ UF ∩ M(r) , the restriction of Fm to a neighborhood U
of m is a submanifold of dimension d − s of U, passing
through m. This dimension
is an upper bound for the dimension dm of span (XF1 )m , . . . , (XFs )m , because
these s vectors are tangent to that fiber at m. Moreover, dm s − dim Ker πm =
s + 2r − d, because the differential one-forms dF1 , . . . , dFs are independent at m.
Combining the two inequalities for dm , we get
332 12 Liouville Integrable Systems
Poisson’s theorem yields new constants of motion (of a given Hamiltonian H), start-
ing from two given constants of motion.
Theorem 12.4 (Poisson’s theorem). Let (M, π ) be a Poisson manifold and let F, G
and H be elements of F (M). If F and G are constants of motion for H, then their
Poisson bracket {F, G} is also a constant of motion for H.
Proof. The proof is an easy consequence of the Jacobi identity, valid for all
F, G, H ∈ F (M):
{F, {G, H}} + {G, {H, F}} + {H, {F, G}} = 0 . (12.3)
Namely, if F and G are constants of motion for H, then the first two terms in (12.3)
vanish. Then the remaining term also vanishes, that is, {F, G} is a constant of motion
for H.
In practice, Poisson’s theorem often does not yield a function which is independent
of the given constants of motion; as we will see, it happens in many important cases
that the Poisson bracket of two constants of motion is actually equal to zero.
12.2 Constructions of Functions in Involution 333
Since m ∈ M is arbitrary, this shows that the function μ̃x is a constant of motion
for H.
It follows that, since Fλ is a Casimir for the Poisson pencil, the sequence of functions
(Fi )i∈Z satisfies (12.5) for all i ∈ Z, hence is a bi-Hamiltonian hierarchy. By the
above, (Fi )i∈Z is involutive.
{F ◦ Ψ , G ◦ Ψ }1 = {F, G}2 ◦ Ψ = 0 .
12.2 Constructions of Functions in Involution 335
The second one is that Lie algebra homomorphisms between finite-dimensional Lie
algebras induce Poisson maps: let φ : g1 → g2 be a Lie algebra homomorphism,
where (g1 , [· , ·]1 ) and (g2 , [· , ·]2 ) are finite-dimensional Lie algebras and consider
the transpose map φ : g∗2 → g∗1 , which assigns to ξ2 ∈ g∗2 the linear form φ (ξ2 ) :
g1 → F, defined for all x ∈ g1 by
φ (ξ2 ), x := ξ2 , φ (x) ,
where we recall that · , · stands for the canonical pairing between a vector space
and its dual. Then φ is a Poisson map, when both g∗1 and g∗2 are equipped with their
Lie–Poisson structure (denoted {· , ·}i , for i = 1, 2). To check this, it is sufficient to
check that if x, y ∈ g1 (g∗1 )∗ , then
x◦φ ,y◦φ = {x, y}1 ◦ φ . (12.6)
2
Recall that under the canonical isomorphism g1 (g∗1 )∗ , the Lie–Poisson bracket
[x, y]1 corresponds to {x, y}1 , and similarly for g2 . Also, under the isomorphism
(g∗2 )∗ g2 , the linear map x ◦ φ : g∗2 → F corresponds to φ (x), for all x ∈ g1 .
Therefore, (12.6) is equivalent to saying that φ is a Lie algebra homomorphism.
Combining both observations, it is clear that Lie algebra inclusions lead to func-
tions in involution (on the dual of the larger Lie algebra). It has the following non-
trivial consequence: if g0 is a Lie subalgebra of g, then the Ad∗ -invariant functions
on g∗0 provide involutive functions on g∗ ; notice that while the Hamiltonian vector
fields of these functions are trivial on g∗0 , they are non-trivial, in general, on g∗ .
Applied to a sequence of subalgebras, a family of functions in involution can be
constructed on the dual of a Lie algebra (equipped with its Lie–Poisson structure).
One usually refers to this technique as Thimm’s method. Explicitly, let g be a finite-
dimensional Lie algebra and suppose that
g1 ⊂ g2 ⊂ · · · ⊂ gk ⊂ gk+1 = g
[g1 , g2 ] ⊂ g2 . (12.7)
Then for every function F1 on g∗1 and for every Ad∗ -invariant function F2 on g∗2 , the
functions F1 ◦ ı1 and F2 ◦ ı2 on g∗ are in involution.
336 12 Liouville Integrable Systems
Let M ⊆ gld (F) be an affine subspace of the Lie algebra of (d × d)-matrices with
entries in F and let V be a vector field on M. Suppose that there exists a matrix-
valued function φ : M → gld (F) such that
for all x ∈ M. Then V is said to be written in Lax form and (12.8) is called a Lax
equation. As we have seen in Section 7.2, if g is a quadratic Lie algebra, identified
with its dual using its bilinear form, and equipped with its Lie–Poisson structure,
then every Hamiltonian vector field on g is of this form. We claim that all coefficients
of the characteristic polynomial det(λ 1d − x), which we view as functions (of x)
on M, are constants of motion of V . To see this, let x(t) be an integral curve of V ,
defined for t in a neighborhood of 0, and denote by y(t) the value of φ at x(t) ∈ M.
Take an arbitrary i > 0 and use Trace(AB) = Trace(BA) to compute
d dx
Trace xi (t) = i Trace xi−1 (t) (t)
dt dt
i
= i Trace x (t)y(t) − xi−1 (t)y(t)x(t) = 0 .
Thus, the functions Hi : x → Trace xi are constants of motion of V . The same is true
for each coefficient of the characteristic polynomial of x, viewed as a function on M,
because each such coefficient is a polynomial in the functions Hi . In particular, if
V = XH is a Hamiltonian vector field, with Hamiltonian H, then H is in involution
with each Hi , and with each coefficient of the characteristic polynomial of x.
12.2 Constructions of Functions in Involution 337
1
[x, y]R := ([Rx, y] + [x, Ry]) ,
2
defines a Lie bracket on g. If g is quadratic, i.e., g is equipped with an ad-invariant
non-degenerate symmetric bilinear form · | · , then the original Lie bracket [· , ·] and
the R-bracket [· , ·]R lead to two Lie–Poisson structures {· , ·} and {· , ·}R on g. In
order to describe these, recall that to a function F ∈ F (g) and an element x ∈ g, an
element ∇x F of g is associated by setting ∇x F | · = dx F, that is,
∇x F | y = dx F, y ,
for all y ∈ g. Using this notation, the two Lie–Poisson structures on g take the form
for F, G ∈ F (g) at x ∈ g (see Section 7.2). The involutivity theorem yields functions
in involution, coming from the R-bracket; moreover, their Hamiltonian vector fields
can be written in a Lax form.
Theorem 12.7 (Involutivity theorem). Let (g, [· , ·]) be a finite-dimensional quadra-
tic Lie algebra, whose bilinear form is denoted by · | · . Let R be an R-matrix
for g and denote the Lie–Poisson bracket on g, which corresponds to the R-bracket,
by {· , ·}R . If F and H are Ad-invariant functions on g, then
(1) {F, H}R = 0;
(2) The Hamiltonian vector field of H with respect to {· , ·}R is given, at x ∈ g, by
1
(XH )x = − [x, R(∇x H)] . (12.10)
2
Proof. Recall that the Ad-invariance of H means that [∇x H, x] = 0 for all x ∈ g (see
Section 5.1.4). For functions F and H on g, their R-bracket is given, at x ∈ g, by
338 12 Liouville Integrable Systems
1 1
{F, H}R (x) = x | [R (∇x F) , ∇x H] + x | [∇x F, R (∇x H)]
2 2
1 1
= R (∇x F) | [∇x H, x] + R (∇x H) | [x, ∇x F] ,
2 2
which is zero when F and H are both Ad-invariant; if only H is assumed Ad-
invariant, only the first term vanishes, leading to
1 1
{F, H}R (x) = ∇x F | [R (∇x H) , x] = dx F, [R (∇x H) , x] ,
2 2
which proves (12.10).
We will now refine the involutivity theorem to obtain the (historically older) Adler–
Kostant–Symes theorem, in which the R-matrix comes from a Lie algebra splitting.
Recall from Section 10.1.2 that if (g, [· , ·]) is a finite-dimensional Lie algebra and g
is written as a vector space direct sum g = g+ ⊕ g− of two Lie subalgebras g+ and
g− of g, then g+ ⊕ g− is called a Lie algebra splitting, and that such a splitting
leads to a natural R-matrix. Denoting the projections of x ∈ g on g+ and g− by x+ ,
respectively x− , the endomorphism R : g → g is defined by
Rx := x+ − x− ,
{F, G}R (x) = x | (∇x F)+ , (∇x G)+ − x | (∇x F)− , (∇x G)− .
[ ε , g+ ] ⊂ g ⊥
+, [ ε , g− ] ⊂ g ⊥
−. (12.11)
1
XHε (y) = − [y, R(∇y H)] = ±[y, (∇y H)∓ ] ,
2
where y ∈ ε + g⊥ −;
(4) Let G be a Lie group, whose Lie algebra is g, and let G+ and G− denote the
Lie subgroups of G, corresponding to g+ , respectively g− . For x0 ∈ g and for
|t| small, let g+ (t) and g− (t) denote the smooth curves in G+ respectively G−
which solve the factorization problem1
d
∇ε +x F | y = F(ε + x + ty) = 0 ,
dt |t=0
for all y ∈ g⊥
− . It follows from ad-invariance of · | · and from the assumptions (12.11)
on ε that
Both of the above terms are equal to zero: the first one because ∇ε +x F ∈ g− and
the second one because x ∈ g⊥ − , combined with the fact that g− is a Lie subalge-
bra (of g). This proves that all Hamiltonian vector fields XH are tangent to ε + g⊥ −,
hence that ε +g⊥ − is a Poisson submanifold of (g, {· , ·}R ), which is the content of (1).
Since (ε + g⊥ − , {· , ·}ε ) is a Poisson submanifold of (g, {· , ·}R ), the involutivity of
two Ad-invariant functions F and H on g (Theorem 12.7) implies the involutivity
of Fε and Hε on ε + g⊥ − , which proves (2); similarly, the first equality in (3) fol-
lows from (1) and from the involutivity theorem (Theorem 12.7), while the second
equality is a direct consequence of the obvious formulas
We now turn to (4), the integral curves of XH . We first show that Adg+ (t) x0 =
Adg− (t) x0 (the second equality in (12.13)). Since Ad is a group homomorphism, the
factorization (12.12) implies that
In this section, we formulate and prove two basic theorems in the theory of inte-
grable systems: the Liouville theorem and the action-angle theorem. Classically,
these theorems are considered for Liouville integrable systems on symplectic man-
ifolds; we consider these theorems here in the more natural context of Poisson
manifolds. We follow [3] in Section 12.3.1 and [120] in Sections 12.3.2 to 12.3.4.
Throughout the section, all manifolds are real and F (M) stands for the algebra of
smooth functions on the (real) manifold M.
We first give the definition of a Liouville integrable system on a real Poisson mani-
fold.
Definition 12.9. Let (M, π ) be a real Poisson manifold of rank 2r and of dimen-
sion d. Let F = (F1 , . . . , Fs ) be an s-tuple of smooth functions on M and suppose
that
(1) F is independent;
(2) F is in involution;
(3) r + s = d.
Then (M, π , F) is called a Liouville integrable system of dimension d and rank 2r.
Viewed as a map, F : M → Rs is called the momentum map of (M, π , F).
Suppose that (M, π , F) is a Liouville integrable system, where (M, π ) is a Poisson
manifold of rank 2r. We recall that M(r) denotes the open subset of M where the rank
of π is equal to 2r, and UF denotes the dense open subset of M, which consists of
all points of M where the differentials of the elements of F are linearly independent.
We consider on the non-empty open subset UF ∩ M(r) of M two (regular) foliations:
(1) The Hamiltonian vector fields XF1 , . . . , XFs define, according to Proposi-
tions 12.2 and 12.3, on the non-empty open subset UF ∩ M(r) of M an inte-
grable distribution D of rank r. The integral manifolds of D are the leaves
of a foliation, which we denote by F ; the leaf of F , passing through
m ∈ UF ∩ M(r) , is denoted by Fm , and is called the invariant manifold of F,
through m;
(2) Since, by definition, the restriction F of F to UF ∩ M(r) is a submersion, its
fibers define a foliation on UF ∩ M(r) , which is denoted by F . Thus, the leaves
of F are the connected components of the fibers of the map
F = (F1 , . . . , Fs ) : UF ∩ M(r) → Rs .
We show in the following proposition that both foliations F and F coincide, lead-
ing to a description of the invariant manifolds of F.
342 12 Liouville Integrable Systems
Proposition 12.10. Let (M, π , F) be a Liouville integrable system and let F and F
denote the foliations, defined above.
(1) The foliations F and F coincide;
(2) For m ∈ UF ∩ M(r) , the invariant manifold Fm of F is the connected compo-
nent of the fiber of F which contains m. In particular, Fm is an (embedded)
submanifold of M.
Proof. We prove item (1); item (2) is an immediate consequence of it. Since all
leaves of F and of F are r-dimensional (and are, by definition, connected), it suf-
fices to show for every m ∈ UF ∩ M(r) the following inclusion of tangent spaces:
Tm F ⊂ Tm F . According to Proposition 12.2, the fact that F1 , . . . , Fs are pairwise
in involution implies that each of the Hamiltonian vector fields XF1 , . . . , XFs , at m,
belong to Tm F . Since Tm F is spanned by these vector fields, the required inclusion
follows.
We now come to the Liouville theorem for Liouville integrable systems on Poisson
manifolds.
Proof. We suppose that the flow Φ (i) of each of the integrable vector fields XFi is
complete on Fm , i.e., is defined for all t ∈ R. We may order the functions Fi such
that the r vector fields XF1 , . . . , XFr are independent at m. These vector fields are
then independent at every point of Fm . Indeed, since the vector fields XFi pairwise
commute,
r
LXF j (XF1 ∧ · · · ∧ XFr ) = ∑ XF1 ∧ · · · ∧ XFj , XFi ∧ · · · ∧ XFr = 0 ,
i=1
for j = 1, . . . , s. It means that XF1 ∧ · · · ∧ XFr is conserved by the flow of each one of
the vector fields XF1 , . . . , XFs . In particular, since this r-vector field is non-vanishing
at m, it is non-vanishing on the entire invariant manifold Fm . As a consequence, Fm
is a leaf of the distribution, defined by the first r vector fields XF1 , . . . , XFr , in a
neighborhood of Fm .
The completeness and commutativity of the vector fields XF1 , . . . , XFr on Fm
imply that we can define an action Rr × Fm → Fm by
12.3 The Liouville Theorem and the Action-Angle Theorem 343
(r)
(t1 , . . . ,tr ), m → Φt1 ◦ Φt2 ◦ · · · ◦ Φtr m .
(1) (2)
Since Fm is the integral manifold through m of the distribution defined by the first
r integrable vector fields, the action is transitive on Fm and Fm becomes a homo-
geneous space. The action is also locally free, because the vector fields XFi are
independent at every point of Fm . Therefore the stabilizer is a discrete subgroup
Hm of Rr and Fm is diffeomorphic to Rr /Hm . If Fm is compact, then Hm must be a
lattice, so Rr /Hm is a torus, smoothly embedded into M. Otherwise Hm is a discrete
subgroup whose rank q is at most r − 1 and Rr /Hm is isomorphic to Rr−q × Tq . By
construction, the vector fields XFi are mapped to translation-invariant vector fields
in both cases.
F|U
pB
Bd−r
Moreover, U and φ can be chosen such that the last d − 2r components of φ are
Casimirs of π , restricted to U.
pB
F
Bsm
Since Fm is compact, it is covered by finitely many of the sets Vm , say Vm1 , . . . ,Vm .
Thus, if every pair of the diffeomorphisms φm1 , . . . , φm agrees on the intersection
of their domain of definition (whenever non-empty), we can define a global dif-
feomorphism on a neighborhood U of Fm , whose image is the intersection of the
concentric balls Bsm1 , . . . , Bsm . In order to ensure that these diffeomorphisms agree,
we need to chose them in a more specific way. This is done by choosing an arbitrary
Riemannian metric on M. Using the exponential map, defined by the metric, we
can identify a neighborhood of the zero section in the normal bundle of Fm , with
a neighborhood of Fm in M; in particular, for every m ∈ Fm there exist neighbor-
hoods Um of m in M and Vm of m in Fm , with smooth maps ψm : Um → Vm , which
have the important virtue that they agree on the intersection of their domains. Upon
shrinking the open subsets Um , if necessary, the maps φm := ψm × (F1 , . . . , Fs ) are a
choice of diffeomorphisms, defined on a neighborhood U of Fm , with the required
properties. This leads to the diffeomorphism φ : U → Fm × Bs , which we view as a
diffeomorphism φ : U → Tr × Bs , using the fact that Fm is diffeomorphic with Tr ,
as stated in the Liouville theorem.
It remains to be shown that φ can be chosen such that its last d − 2r = s − r com-
ponents are Casimirs. To do this, we consider on U the integrable distribution D
defined by all Hamiltonian vector fields on U; it has rank 2r and its leaves are the
symplectic leaves of (U, π ). It is clear that D ⊂ D , where we recall that D is the
distribution defined by the Hamiltonian vector fields XF1 , . . . , XFs . Consider the
submersive map
pB ◦ φ : U → Tr × Bs → Bs ,
whose fibers are by assumption the leaves of F , that is, the integral manifolds of D
(restricted to U), so that the kernel of the tangent map T (pB ◦ φ ) is precisely D. The
image of D by T (pB ◦ φ ) is therefore a (smooth) distribution D of rank r on Bs ,
which is integrable, since D is integrable. The foliation defined by the integral
manifolds of D is, in the neighborhood of the point pB (φ (m)), defined by s − r =
d − 2r independent functions z1 , . . . , zd−2r . Pulling them back to M, we get functions
z1 , . . . , zd−2r on a neighborhood V ⊂ U of Fm which are Casimir functions, because
they are constant on the leaves of D , which are the symplectic leaves of (U, π ).
12.3 The Liouville Theorem and the Action-Angle Theorem 345
In this section, we study Liouville integrable systems of the form (Tr × Bs , π , F),
where F = pB : Tr × Bs → Bs is the projection on the second factor and π is a
regular Poisson structure of rank 2r on Tr × Bs . Given such a Liouville integrable
system, there is no reason why the Hamiltonian vector fields XFi , which are tangent
to the fibers of F (which are tori), would be constant on these fibers, and even if they
were constant on these fibers, they might vary from one fiber to another in the sense
that their flow may not come from an action of the torus Tr on Tr × Bs . We show
in the following proposition that vector fields which satisfy the latter property can
be constructed by taking well-chosen linear combinations of the Hamiltonian vector
fields XFi , with as coefficients F-basic functions, i.e., functions of the form λ ◦ F,
where λ ∈ F (Bs ); equivalently, smooth functions on Tr × Bs which are constant on
the fibers of F.
Proposition 12.13. Let (Tr × Bs , π , F) be a Liouville integrable system, where π
has constant rank 2r and F = (F1 , . . . , Fs ) is the projection on the second compo-
nent. Suppose that among the s commuting vector fields XF1 , . . . , XFs , the first r
are independent at every point of Tr × Bs . There exists a ball Bs0 ⊂ Bs , concentric
with Bs , and there exist F-basic functions λij ∈ F (Bs0 ), such that the vector fields
V1 , . . . , Vr , defined by
r
Vi := ∑ λi j X F j ,
j=1
Φ : Rr × (Tr × Bs ) → Tr × Bs
(1) (r)
((t1 , . . . ,tr ), m) → Φt1 ◦ · · · ◦ Φtr (m) .
(i)
Since the vector fields XFi are pairwise commuting, the flows Φti pairwise com-
mute and Φ is an action of Rr on Tr × Bs . Since the vector fields XF1 , . . . , XFr are
independent at all points of Tr × Bs , the fibers of F, which are r-dimensional tori,
are the orbits of the action. For c ∈ Bs , let Λc denote the lattice of Rr , which is the
isotropy group of every point in F−1 (c); it is the period lattice of the action Φ , re-
stricted to F−1 (c). In order to simplify the notation, we assume that Bs is centered
at o.
Let mo be an arbitrary point of F−1 (o) and choose a basis (λ1 (o), . . . , λr (o)) of
the lattice Λo . For a fixed i, with 1 i r, for m in a neighborhood of mo in Tr × Bs
346 12 Liouville Integrable Systems
Since the action is locally free, the Jacobian condition is satisfied and we get by
solving for Li around λi (o) a smooth Rr -valued function λi (m), defined for m in
a neighborhood Wi of mo . Doing this for i = 1, . . . , r and setting W := ∩ri=1Wi , we
have that W is a neighborhood of mo , and on W we have functions λ1 (m), . . . , λr (m),
with the property that Φ (λi (m), m) = m for all m ∈ W and for all 1 i r. Thus,
λ1 (m), . . . , λr (m) belong to the lattice ΛF(m) for all m ∈ W and they form a basis
when m = mo ; by continuity, they form a basis of ΛF(m) for all m ∈ W .
The functions λi can be extended to a neighborhood of the torus F−1 (o). In
fact, the functions λi are F-basic, hence extend uniquely to F-basic functions on
F−1 (F(W )). We will use in the sequel the same notation λi for these extensions and
we write F−1 (F(W )) simply as W . Using these functions, we define the following
smooth map:
Φ̃ : Rr ×W → W
r
(12.15)
((t1 , . . . ,tr ), m) → Φ ∑ ti λi (m), m .
i=1
Since the functions λi are F-basic, the fact that Φ is an action implies that Φ̃ is an
action. This action has the extra feature that the stabilizer of every point in W is Zr .
Thus, Φ̃ induces an action of Tr on W , which we still denote by Φ̃ . By shrinking W ,
if necessary, we may assume that W is of the form F−1 (Bs0 ), where Bs0 is an open
ball, concentric with Bs , and contained in it. Thus we have a torus action
Φ̃ : Tr ×W → W
r
((t1 , . . . ,tr ), m) → Φ ∑ ti λi (m), m .
i=1
We formulate and prove in this section the action-angle theorem. The following
lemma plays a key rôle in the proof.
12.3 The Liouville Theorem and the Action-Angle Theorem 347
Lemma 12.14. Let M be a manifold, equipped with a vector field V and a bivector
field P. Suppose that V is complete and that each one of its integral curves has
period 1. If LV2 P = 0, then LV P = 0.
Proof. Let Q be the bivector field on M, defined by Q := LV P. We pick an arbitrary
point m and we show that Qm = 0. Denoting the flow of V by Φt , we have for all t
that
d
(Φt )∗ PΦ−t (m) = (Φt )∗ (LV P)Φ−t (m) = (Φt )∗ QΦ−t (m) = Qm , (12.16)
dt
where we used in the last step that the bivector field Q satisfies LV Q = 0. By inte-
grating (12.16),
(Φt )∗ PΦ−t (m) = Pm + tQm .
Evaluated at t = 1 this yields Qm = 0, since Φ1 = Φ−1 = 1M , as V has period 1.
FXG , F XG , π S S
= [FXG , XF ∧ XG ]S
= F [XG , XF ∧ XG ]S + XG [F, XF ∧ XG ]S = 0 ,
as was to be shown.
We proceed to construct the action coordinates: we show that the Poisson vector
fields Vi are actually Hamiltonian vector fields (upon shrinking Br , if necessary),
where the Hamiltonians can be taken as F-basic functions; these Hamiltonians will
yield the action coordinates. According to Theorem 1.28, we can construct on a
neighborhood U of m in Tr × Bs functions g1 , . . . , gr such that
( f1 , . . . , fr , g1 , . . . , gr , z1 , . . . , zs−r )
r
∂ λi j r r
∂
Xλ j = ∑ X fk , and Vi = ∑ λi j X f j = − ∑ λi j ∂ g j .
i
k=1 ∂ f k j=1 j=1
Expressing that Vi is a Poisson vector field, [Vi , π ]S = 0, then takes the explicit form
r
∂ ∂ ∂ r
∂ λi j ∂ ∂
− ∑ λi j , ∧
∂ g j ∂ f k ∂ gk
= ∑ ∂ f k ∂ g j ∧ ∂ gk = 0 ,
j,k=1 S j,k=1
Since the vector fields X f1 , . . . , X fr are linearly independent at all points of U , all
coefficients of (12.18) vanish and we get, for every i, j, k ∈ {1, . . . , r}:
∂ λi j ∂ λik
= . (12.19)
∂ fk ∂ fj
satisfy
∂ pi
λi j = , (12.21)
∂ fj
for all 1 i, j r. As above, we view these functions as functions on U . They are
Hamiltonians of the vector fields Vi , because
r
∂ pi r
X pi = ∑ ∂ fj Xfj = ∑ λij X f j = Vi . (12.22)
j=1 j=1
This shows that each one of the vector fields Vi is a Hamiltonian vector field on U .
Since the functions pi and z j are F-basic, each admits a unique extension to an
F-basic function on F(F−1 (U )), a neighborhood of Fm which we call U in the
sequel. On U , we still have Vi = X pi .
We proceed to the construction of the angle coordinates. In view of Theo-
rem 1.28, there exist on a neighborhood U ⊂ U of m in Tr × Bs , smooth functions
θ1 , . . . , θr such that
r
∂ ∂
π=∑ ∧ . (12.23)
j=1 ∂ θ j ∂ pj
∂ Fi
ṗ j = p j , Fi = − =0, j = 1, . . . , r ,
∂θj
∂ Fi
θ̇ j = θ j , Fi = = cj , j = 1, . . . , r ,
∂ pj
żk = 0 , k = 1, . . . , s − 2r ,
where c j depends on the p and z coordinates only. Integrating the first and third
equations, we find that all p j and zk are constant (which is not surprising since the
Fj are constants of motion and the zk are Casimirs), so that the c j are constant,
which yields upon using the second equation that the angle coordinates evolve lin-
early, θ j (t) = c j t, for j = 1, . . . , r. Thus, action-angle coordinates not only exhibit
an integrable system locally as a Hamiltonian system with linear flow(s) on a family
of tori, they are also the coordinates in which the equations of motion take their
simplest possible form.
The action-angle theorem admits a natural generalization to the case of non-
commutative integrable systems (see [120]). The classical examples of Liouville
integrable systems can be found in [204]. For a modern treatment of Liouville inte-
grability, including recently discovered examples, we refer to [3, 164].
12.4 Notes
The term “integrable system” appears in a large variety of mathematical and physi-
cal contexts, referring often more to phenomena than to a precise general definition.
The notion of a Liouville integrable system, detailed in this chapter, is a notable
exception. Liouville’s observation that “Liouville integrable systems” can be solved
by quadratures has been a main impulse to the development of Poisson geometry,
for a long time restricted to symplectic geometry (the geometry of the phase space
of a mechanical system). For an excellent account of what was classically known
about integrable systems, the reader should consult Whittaker [204]. The books by
Abraham–Marsden [1], Arnold [15] and Libermann–Marle [125] are all devoted
12.4 Notes 351
In this section, we introduce the notion of a formal deformation and of a k-th order
deformation of a commutative associative algebra and we show how these defor-
mations are related to Hochschild cohomology. We pick an indeterminate ν and we
denote by Fν the ring F[[ν ]] of formal power series in ν with coefficients in F.
Let A be a commutative associative algebra over F with unit. The product on A will
be denoted in its functional form by μ , or by a simple juxtaposition of the factors,
so we write FG for μ (F, G) for all F, G ∈ A . Associated to A , we consider the Fν -
algebra A [[ν ]] of formal power series in ν , with coefficients in A ; we also denote
this algebra by A ν . Thus, an element of A ν is a formal power series ∑i∈N Fi ν i ,
where all Fi ∈ A . The product which A ν inherits from A by “extension of scalars”,
as in Section 2.4.1, is given by
∑ Fi ν i · ∑ G jν j := ∑ Fi G j ν i+ j , (13.1)
i∈N j∈N i, j∈N
Condition (1) in the above definition implies that the map Φ is invertible, so the
notion of equivalence of formal deformations defines an equivalence relation on the
set of all formal deformations of a product μ . Also, if μ is a formal deformation
of μ and Φ : A ν → A ν is an Fν -linear map such that Φ (F) = F (mod ν ), for
every F ∈ A , then the map μ : A ν × A ν → A ν , defined for F, G ∈ A ν by
μ (F, G) := Φ μ (Φ −1 (F), Φ −1 (G)) ,
We show this in the case of k = 1, the case of arbitrary k being only notationally
more complicated. Suppose that Φ : A ν → A ν is an Fν -linear map. For F ∈ A ,
we can write Φ (F) in a unique way as
where every Φi (F) belongs to A . It follows that we can associate to every ele-
ment Φ of HomFν (A ν , A ν ) a sequence of linear maps (Φi )i∈N from A to A , that
is, an element of HomF (A , A )[[ν ]]. Conversely, given such a sequence of F-linear
maps (Φi )i∈N , the map Φ , defined by (13.2), extends in a unique way to an Fν -linear
map. By a slight abuse of notation, we keep the notation Φi for the Fν -linear exten-
sions of the maps Φi to A ν , which has the effect that (13.2) remains valid when F
belongs to A ν .
Formal deformations are often constructed term by term, which is done rigor-
ously by considering k-th order deformations. For k ∈ N, we consider the ring
Fνk := Fν /ν k+1 . Elements of this ring can be represented by polynomials in ν
of degree at most k and the product in this ring is the ordinary multiplication of
polynomials, followed by stripping off the terms whose degree is higher than k.
Starting from a commutative associative algebra A , we consider similarly
μ(k) = μ0 + μ1 ν + · · · + μk ν k , (13.3)
356 13 Deformation Quantization
where each μi is an F-bilinear map from A to itself, is called a k-th order deforma-
tion of μ if μ0 = μ and μ(k) is associative.
Under the natural quotient maps A ν → Akν and Ak+1 ν → A ν , formal deformations
k
(respectively (k + 1)-th order deformations) are mapped to k-th order deformations.
By a slight abuse of language, we say that an Fν -bilinear map μ : A ν × A ν → A ν
such that μ = μ (mod ν ) defines a k-th order deformation of μ if the induced
Fνk -bilinear map μ(k) : Akν × Akν → Akν is a k-th order deformation of μ . Clearly,
this is equivalent to saying that
for all F, G, H ∈ A .
The notion of equivalence of k-th order deformations of μ is obtained by replac-
ing Fν by Fνk and A ν by Akν in Definition 13.2, which deals with the notion of
equivalence of formal deformations. Again, an Fν -linear map A ν → A ν which in-
duces an equivalence of k-th order deformations will be said to define an equivalence
of k-th order deformations.
Let (A , μ ) be a commutative associative algebra and suppose that μ(1) = μ0 +
μ1 ν defines a first order deformation of μ = μ0 . For a, b, c ∈ A , the coefficient
in ν in the associativity equation μ(1) (μ(1) (a, b), c) = μ(1) (a, μ(1) (b, c)) (mod ν 2 )
is given by
μ1 (ab, c) + μ1 (a, b)c = aμ1 (b, c) + μ1 (a, bc) .
It follows that
μ1− (ab, c) + μ1− (a, b)c − aμ1− (b, c) − μ1− (a, bc) = 0 ,
1
μ1− (a, b) := (μ1 (a, b) − μ1 (b, a)) , (13.4)
2
for all a, b ∈ A . Performing a cyclic permutation of a, b, c in this equation leads to
the following three equations.
μ1− (ab, c) + μ1− (a, b)c − μ1− (b, c)a − μ1− (a, bc) = 0 ,
μ1− (bc, a) + μ1− (b, c)a − μ1− (c, a)b − μ1− (b, ac) = 0 ,
μ1− (ac, b) + μ1− (c, a)b − μ1− (a, b)c − μ1− (c, ab) = 0 ,
for all a, b and c in A , i.e., μ1− satisfies the Jacobi identity. In conclusion, if μ(2) =
μ0 + μ1 ν + μ2 ν 2 is a second order deformation of an associative algebra (A , μ ),
then (A , μ , μ1− ) is a Poisson algebra!
which is a graded Lie bracket, just like the Schouten bracket, after a shift of degree
(see Section 3.3.2). Thus, we denote k := k −1, and we define HCk−1 (V ) := HCk (V )
for k ∈ N so that HCk (V ) = HCk (V ). Elements of the latter space have degree k and
shifted degree k. With respect to the shifted degree, the Gerstenhaber bracket is a
graded Lie bracket, i.e., for all k, ∈ N∗ ,
and [· , ·]G is graded skew-symmetric and satisfies the graded Jacobi identity (see
Table A.2 in Appendix A).
Explicitly, the bracket [· , ·]G is defined, for φ ∈ HCk (V ) and ψ ∈ HC (V ), by the
following formula,
358 13 Deformation Quantization
is a coboundary operator, i.e., δμk+1 ◦ δμk = 0 for every k ∈ N. This operator is called
the Hochschild coboundary operator and the associated complex
δμk−1 δμk δμk+1
··· HCk−1 (A ) HCk (A ) HCk+1 (A ) ···
for k ∈ N∗ , while HH0μ (A ) := Ker δμ0 . The elements of Ker δμk are called Hochschild
k-cocycles and the elements of Im δμk−1 are called Hochschild k-coboundaries.
An explicit formula for the Hochschild coboundary operator δμ is obtained by
substituting μ for ψ in (13.5); writing FG for μ (F, G), it takes, for φ ∈ HCk (A ),
and F0 , . . . , Fk ∈ A , the simple form
Since
δμ0 (F)(G) = GF − FG ,
13.1 Deformations of Associative Products 359
for F, G ∈ A , one has that HH0μ (A ) = Ker δμ0 is the center of the associative alge-
bra (A , μ ). Since moreover
∑ [μi , μ j ]G = 0 ,
i+ j=k+1
i, j0
which we can write in terms of the Hochschild coboundary operator (13.6), associ-
ated to μ = μ0 , as
1
δμ2 (μk+1 ) = −[μk+1 , μ0 ]S =
2 ∑ [ μi , μ j ] G . (13.9)
i+ j=k+1
i, j1
1 In other words, [· , ·]G denotes the Gerstenhaber bracket on the Fν -module HC• (A ν ).
360 13 Deformation Quantization
This means that μ(k) can be extended to a deformation of μ of order k +1, if and only
if the right-hand side of (13.9) is a 3-coboundary for the Hochschild coboundary
operator δμ2 . Notice that it is a 3-cocycle, since
⎛ ⎞
⎜ ⎟
δμ3 ⎜
⎝ ∑ [μi , μ j ]G ⎟
⎠= ∑ μ , [μi , μ j ]G G
i+ j=k+1 i+ j=k+1
i, j1 i, j1
=− ∑ μi , [μ j , μ ]G G
+ [μ j , [μ , μi ]G ]G
i+ j=k+1
i, j1
=2 ∑ μi , δμ2 (μ j ) G
i+ j=k+1
i, j1
= ∑ μi , [μ j , μ ]G G
i+ j+=k+1
i, j,1
= 0,
where we have used the graded Jacobi identity for [· , ·]G to obtain the second and the
last lines. It follows that the obstruction to extending a deformation of μ of some
order to the next order lies in the third Hochschild cohomology space HH3μ (A ).
This statement is the content of the following proposition.
Proposition 13.5. Let (A , μ ) be a commutative associative algebra and suppose
that μ(k) = ∑ki=0 μi ν i defines a k-th order deformation of μ . Then μ(k) can be ex-
tended to a (k + 1)-th order deformation of μ if and only if the Hochschild 3-cocycle
∑ki=1 [μi , μk+1−i ]G is a Hochschild 3-coboundary.
It is clear from (13.9) that the term μk+1 ν k+1 which makes a k-th order deformation
of μ into a (k + 1)-th order deformation of μ is not unique (if it exists): one can
always add a Hochschild 2-cocycle to μk+1 and this is the only freedom in the choice
of μk+1 . If this 2-cocycle is a coboundary, then the two extended (k + 1)-th order
deformations are essentially the same, in the sense that they are equivalent (see
Definition 13.2). More generally, we have the following proposition.
Proposition 13.6. Let (A , μ ) be a commutative associative algebra and suppose
that μ(k) = ∑ki=0 μi ν i defines a k-th order deformation of μ , where k ∈ N∗ . Let k,
where ∈ N∗ . Let φ : A → A be an F-linear map, which we extend to an Fν -linear
map A ν → A ν , still denoted by φ , and let
μ() = μ0 + μ1 ν + μ2 ν 2 + · · · + μ−1 ν −1 + (μ + δμ1 (φ )) ν . (13.10)
extends to a k-th order deformation μ of μ , which is equivalent to μ .
Then, μ() (k) (k)
Φ : Aν → Aν
F → F − φ (F) ν
whose inverse is given by Φ −1 (F) = ∑i∈N φ i (F) ν i . The Fν -bilinear map μ(k)
, de-
= μ + δ 1 (φ ) ν , as
where we have used (13.8) in the last step. It follows that μ() () μ
in (13.10).
It follows from the proposition that in order to construct all possible deforma-
tions of μ , up to equivalence, one only has to consider, at every step k, as many
possibilities as there are elements in HH2μ (A ).
Let μ = ∑∞ ν
i=0 μi ν be an element of HC (A ), where μ0 = μ . As above, μ is
i 2
1
δμ2 (x) − [x, x]G = 0 , (13.11)
2
where, by a slight abuse of notation, δμ2 stands for the Fν -linear extension of δμ2
to HC2 (A ν ) HC2 (A )[[ν ]]. For future use, we stress that we consider only so-
lutions x of this equation which are formal series without constant term. Precisely,
x ∈ ν HC2 (A ν ). In terms of the terminology of Section 13.3, equation (13.11) is
called the Maurer–Cartan equation associated to the differential graded Lie algebra
( HC• (A ), [· , ·]G , δμ ).
362 13 Deformation Quantization
where for r, s ∈ Nd the coefficients φr,s are smooth functions on U, which van-
ish except for a finite number of (r, s). When all coefficients φr,s are constant,
we speak of a constant coefficient bi-differential operator. For each multi-index
r = (r1 , . . . , rd ) ∈ Nd , we used in (13.12) the notation
∂ rF ∂ r1 +···+rd F
:= r .
∂ xr ∂ x1r1 · · · ∂ xdd
cochains of C∞ (M), i.e., the Hochschild cochains which are given by multi-differen-
tial operators. The space of all differential Hochschild cochains of C∞ (M) is denoted
by HC•diff (M) := k∈N HCkdiff (M). Notice that the Gerstenhaber bracket can be re-
stricted to such cochains because the Gerstenhaber bracket of two multi-differential
operators is a multi-differential operator.
νi
eνπ /2 (F ⊗ G) := F ⊗ G + ∑∗ 2i i! π i (F ⊗ G) . (13.14)
i∈N
Explicitly, when Φ is of the form (13.12), with all φr,s constant, then Φ̂xy is given by
∂ r+s
Φ̂xy = ∑ φr,s
∂ x r ∂ ys
.
r,s∈Nd
F G = eν π̂xy /2 F(x)G(y)|y=x .
(F G) H = eν π̂xz /2 (F G)(x)H(z)|z=x
= eν π̂xz /2+ν π̂yz /2 eν π̂xy /2 F(x)G(y)H(z)|z=y=x
= eν (π̂xz +π̂yz +π̂xy )/2 F(x)G(y)H(z)|z=y=x
= F (G H) .
The computation demands some extra justification. We have used in the third equal-
ity that if φ1 and φ2 are commuting endomorphisms, then eφ1 +φ2 = eφ1 eφ2 ; this prop-
erty follows at once from the definition given in (13.14). We used in the second
equality that if V = Vx is a derivation of F (V ), then
V F(x, y)|y=x = Vx F(x, y)|y=x + Vy F(x, y)|y=x = (Vx + Vy )F(x, y)|y=x ,
which follows from the chain rule. This proves that is associative, hence is a formal
deformation of μ ; since for F, G ∈ F (V ) the coefficient of ν in F G is 12 π [F, G], it
follows that is a star product for (V, π ).
When the rank 2r of the constant Poisson structure π is equal to the dimension of V ,
there exists a system of linear coordinates (q1 , . . . , qr , p1 , . . . , pr ) on V , in which the
Poisson bracket of functions F, G ∈ F (V ) takes the canonical form
r
∂F ∂G ∂F ∂G
{F, G} = ∑ − ;
i=1 ∂ qi ∂ pi ∂ pi ∂ qi
see Proposition 6.5 for the existence of such coordinates. For such a choice of Pois-
son structure and such coordinates, the explicit formula (13.14) is referred to as the
Moyal–Weyl product, or simply the Moyal product.
13.1 Deformations of Associative Products 365
In this section we explain how the Moyal–Weyl product can be used to prove that
every symplectic manifold (M, ω ) admits a star product; more precisely, the alge-
bra of smooth functions on (M, ω ), equipped with its canonical Poisson structure,
admits a star product. The construction presented here is due to Fedosov; we refer
to his original paper [73] for proofs and details.
We first need to elaborate on the star product in the case of a symplectic vector
space (the Moyal–Weyl product, explained in Section 13.1.6 above). Let (V, ω ) be
a finite-dimensional symplectic real vector space and let π = {· , ·} denote the (con-
stant) Poisson structure on V which corresponds to ω (see Section 6.3.1). We denote
by Fˆ (V ) the space of all formal power series in the variables that constitute a basis
for V ∗ and we denote by Fˆ j (V ) the subspace of Fˆ (V ), consisting of those formal
power series whose valuation is at least j ∈ N. We also consider, for k ∈ N,
E k (V ) := ∑ ν i Fˆ j (V ) .
i, j∈N
2i+ jk
It is easy to see from the explicit formula (13.15) that the Moyal–Weyl product
on C∞ (V )ν extends to Fˆ (V )ν := Fˆ (V )[[ν ]]; moreover, for all k, ∈ N, we have
that
E k (V ) E (V ) ⊂ E k+ (V ) , (13.16)
k
which follows immediately from the fact that Fˆ (V ), Fˆ (V ) ⊂ Fˆ k+−2 (V ),
for all k, ∈ N∗ . To V we associate the associative algebra A(V ), defined by
A(V ) := E 0 (V ) ⊗ ∧•V ∗ ,
(F ⊗ ξ ) (G ⊗ η ) := (F G) ⊗ (ξ ∧ η ) ,
We say that a section of the canonical projection ∪m∈M A(Tm M) → M mapping each
element to its base point is smooth if its projection on the vector bundle
A(Tm M)
k (T M)
→M
m∈M
A m
is smooth, for every k ∈ N. For k ∈ N, the space of all smooth sections which take
values in Ak (Tm M), for every m ∈ M, is denoted by A k (M); we use the shorthand
A (M) for A 0 (M). The product of two sections a ∈ A k (M) and b ∈ A (M) is
the section of A k+ (M) whose value at m ∈ M is a(m) m b(m). Combined, the
linear maps P0 : A(Tm M) → R defined above lead to a linear map A (M) → F (M)ν ,
which we also denote by P0 .
Theorem 13.9. Let (M, ω ) be a symplectic manifold and let (A (M), ) be the al-
gebra constructed as above. There exists a derivation D : A (M) → A (M) with the
following property: for every F ∈ F (M)ν , there exists a unique aF ∈ A (M) such
that D(aF ) = 0 and P0 (aF ) = F.
ıZ dω (V , W ) = ω (∇Z V , W ) + ω (V , ∇Z W ) ,
[a ⊗ ξ , b ⊗ η ]m := (a b − b a) ⊗ ξ ∧ η ,
erty, one proves the existence and uniqueness of aF , as stated in Theorem 13.9.
It is clear that D is a derivation, because both the adjoint action and the covariant
derivatives are derivations.
Corollary 13.10. Let (M, ω ) be a symplectic manifold, let (A (M), ) be the al-
gebra constructed above and let D be a derivation of A (M) as in Theorem 13.9.
Denote by the bilinear map on F (M)ν defined for all F, G ∈ F (M)ν by
13.2 Deformations of Poisson Structures 367
F G := P0 (aF aG ) , (13.18)
where aF is the unique element in the kernel of D, for which P0 (aF ) = F, and simi-
larly for aG . Then is a star product for (M, π ), where π is the canonical Poisson
structure on M, associated to ω .
Proof. We first prove that the product, defined by (13.18), is associative. Let
B(M) ⊂ A (M) be the kernel of D; it is a subalgebra of A (M) since D is a
derivation. By Theorem 13.9, the assignment F → aF is a bijective linear map
from F (M)ν to B(M) with inverse map the restriction of P0 to B(M). It fol-
lows that the product on F (M)ν is obtained from the associative product
on B(M), via a bijection, hence is associative. We prove that is a star product
for (M, π ). Let F and G be smooth functions on M (in particular, they are inde-
pendent of ν ). A closer look at the construction of aF and aG in [73] implies that
aF = (F + dm F + Hessm F) ⊗ 1 + bF and aG = (G + dm G + Hessm G) ⊗ 1 + bG for
some bF , bG ∈ A 3 (M), where dm F and dm G are viewed as linear functions on
Tm M, and Hessm F and Hessm G are viewed as quadratic functions on Tm M. An
explicit computation then yields P0 (aF aG ) = FG + ν2 {F, G} (mod ν 2 ), which
completes the proof.
algebra is valid only for algebras over a field, but it is clear that one can define
Poisson algebras over an arbitrary commutative ring (with unit) R, by replacing in
Definition 1.1 “F-vector space A ” by “R-module A ”. It is also clear that the Jacobi
identity for π is satisfied if and only if [π , π ]S = 0, where the Schouten bracket
[· , ·]S is still given by (3.36), but for Fν -multi-linear maps, rather than F-multi-linear
maps; this is the main fact which we will use about this more general context.
Definition 13.12. Let (A , ·, π ) be a Poisson algebra and let π and π be two formal
deformations of π . Then π and π are said to be equivalent deformations if there
exists an Fν -linear map Φ : A ν → A ν , satisfying for all F, G ∈ A the following
conditions:
(1) Φ (F) = F (mod ν );
(2) Φ (FG) = Φ (F)Φ (G);
(3) Φ (π [F, G]) = π [Φ (F), Φ (G)].
Condition (1) implies that Φ is invertible; then (2) means that Φ is an automorphism
of the associative algebra (A , ·), hence of (A ν , ·) and (3) means that Φ is a Lie
algebra isomorphism (A ν , π ) → (A ν , π ).
It is clear that, if π is a formal deformation of π and Φ : A ν → A ν is an
ν
F -linear map satisfying items (1) and (2) of Definition 13.12, then the map π :
A ν × A ν → A ν , defined for F, G ∈ A ν by
π [F, G] := Φ π [Φ −1 (F), Φ −1 (G)] , (13.19)
π(k) = π0 + π1 ν + · · · + πk ν k , (13.20)
where we have used the Poisson coboundary operator (4.5), associated to the Pois-
son bracket π = π0 . As in the associative case (see Section 13.1.3), one checks by
using the graded Jacobi identity for [· , ·]S , that the right-hand side in (13.21) is a
Poisson 3-cocycle, so that the obstruction for extending a deformation of the Pois-
son bracket π , of some order to the next order lies in the third Poisson cohomology
space Hπ3 (A ). This shows the following proposition.
Proposition 13.14. Let (A , ·, π ) be a Poisson algebra and suppose that π(k) =
∑ki=0 πi ν i is a k-th order deformation of π . Then π(k) can be extended to a (k + 1)-th
order deformation of π if and only if the Poisson 3-cocycle ∑ki=1 [πi , πk+1−i ]S is a
Poisson 3-coboundary.
It is clear from (13.21) that the term πk+1 ν k+1 which makes a k-th order deforma-
tion of π into a (k + 1)-th order deformation of π is not unique (if it exists): one
can always add an arbitrary Poisson 2-cocycle (for π ) to πk+1 . If this cocycle is a
coboundary, then the two extended (k + 1)-th order deformations are essentially the
same, in the sense that they are equivalent (see Definition 13.12). A stronger state-
ment is given in the following proposition, which is the analog of Proposition 13.6.
Proposition 13.15. Let (A , ·, π ) be a Poisson algebra and suppose that π(k) =
∑ki=0 πi ν i is a k-th order deformation of π , where k ∈ N∗ and let k, where ∈ N∗ .
Let ϕ ∈ X1 (A ), which we extend to an Fν -linear map A ν → A ν , still denoted by ϕ ,
and let
π() = π0 + π1 ν + π2 ν 2 + · · · + π−1 ν −1 + (π + δπ1 (ϕ )) ν . (13.22)
extends to a k-th order deformation π of π , which is equivalent to π .
Then, π() (k) (k)
Φ : Aν → Aν
(−1)k k k (13.23)
F → e−ν ϕ (F) = ∑
ν ϕ (F) .
k∈N k!
It is clear that Φ is Fν -linear and that Φ (F) = F (mod ν ), for all F ∈ A . We show
that Φ (FG) = Φ (F)Φ (G) for all F, G ∈ A . To do this, we use that for all k ∈ N,
370 13 Deformation Quantization
k
ϕ (FG) = ∑
k
ϕ r (F) ϕ s (G) , (13.24)
r+s=k
s
r,s0
(−1)k k k
Φ (FG) = e−ν ϕ (FG) = ∑
ν ϕ (FG)
k∈N k!
(−1)k k k r
= ∑ ∑ k! s
ν ϕ (F) ϕ s (G)
k∈N r+s=k
r,s0
(−1)r+s (r+s) r
= ∑∑ s!r!
ν ϕ (F) ϕ s (G)
r∈N s∈N
(−1)r r r (−1)s s s
= ∑ ν ϕ (F) ∑ ν ϕ (G)
r∈N r! s∈N s!
= Φ (F) Φ (G) .
, which is defined for all F, G ∈ A ν by
As in (13.19), we have that π(k)
π(k) [F, G] := Φ π(k) Φ −1 (F), Φ −1 (G) ,
is a k-th order deformation of π which is equivalent to π(k) . Using the formula for
the inverse of Φ , namely
ν k k
Φ −1 (F) = eν ϕ (F) = ∑
ϕ (F) ,
k∈N k!
we can make the following computation in the algebra Aν (i.e., all equalities are
modulo ν +1 ):
(F, G) = Φ π
π() () F + ϕ [F]ν , G + ϕ [G]ν
= Φ π() [F, G] + (π [ϕ [F], G] + π [F, ϕ [G]])ν
= π() [F, G] + (π [ϕ [F], G] + π [F, ϕ [G]] − ϕ [π [F, G]])ν
= π() [F, G] + δπ1 (ϕ )[F, G] ν .
It follows from the proposition that in order to construct all possible deformations
of π , up to equivalence, one only has to consider, at every step k, as many possibili-
ties as there are elements in Hπ2 (A ). Notice that the associative analog of the above
proposition, which was given in Proposition 13.6, is easier to prove; the reason for
13.3 Differential Graded Lie Algebras 371
this is that in the Poisson case the map Φ which defines the equivalence needs to be
an automorphism of the associative algebra (Akν , ·).
Differential graded Lie algebras are graded Lie algebras which are equipped with
a derivation of degree one whose square is zero. A prime example, which we have
already encountered several times in this book, is the algebra of skew-symmetric
multi-derivations of a Poisson algebra (or multivector fields on a Poisson manifold),
where the graded Lie bracket is the Schouten bracket and the differential is (up to
a sign) the Poisson coboundary operator. Since in the study of a differential graded
Lie algebra its symmetric algebra plays an important rôle we first give the basic
definitions and properties of the symmetric algebra of a graded vector space.
since it only depends on the -tuple (i1 , . . . , i ) and not on the choice of x1 , . . . , x .
With X as above, we often write sgn(σ ; X) as a shorthand for sgn(σ ; i1 , . . . , i ). For
indices i1 , . . . , i such that Vi1 Vi2 . . .Vi = 0 we define sgn(σ ; i1 , . . . , i ) := 1 for all
σ ∈ S .
In what follows, the graded vector space of which we consider the symmetric
algebra will always be a graded Lie algebra, but in which we shift the grading by
one. Thus, if g = ∈N g is a graded Lie algebra, we consider on g two gradings:
for x ∈ g we define |x| := and d 0 (x) := − 1 = |x| − 1. Then g is a graded Lie
algebra with respect to the grading | · |, but when the grading d 0 is used, it is only
a graded vector space; when we use the latter grading, we will write g instead of g.
The symmetric algebra which plays an important rôle in what follows is the graded
commutative algebra S• g, where the grading on S• g is given, for a homogeneous
monomial x1 . . . x , by
d 0 (x1 . . . x ) := ∑ |xi | − = ∑ d 0 (xi ) . (13.28)
i=1 i=1
Of course, S• g inherits also a grading from the tensor algebra, which assigns to a
monomial x1 . . . x the degree ; we will not use this grading, except that we denote,
for ∈ N, by S g the span of all monomials of the form x1 . . . x , where x1 , . . . , x ∈ g.
In this section, we introduce the notion of a differential graded Lie algebra and of a
morphism of differential graded Lie algebras. The basic examples, given below, are
the Poisson complex, associated to a Poisson algebra, and the Hochschild complex,
associated to an associative algebra.
Definition 13.16. A differential graded Lie algebra is a triple (g, [· , ·], D), where
(g, [· , ·]) is a graded Lie algebra and D : g → g is a differential of g, i.e., D is a
graded linear map of degree one from g to g, D satisfies D ◦ D = 0 and D is a graded
derivation of degree one of (g, [· , ·]).
Expressed in terms of formulas, the fact that (g = k∈Z gk , [· , ·]) is a graded Lie
algebra means that [· , ·] satisfies, for all homogeneous elements x, y and z of g and
for all i, j ∈ Z, the following properties:
(1) [gi , g j ] ⊂ gi+ j ;
13.3 Differential Graded Lie Algebras 373
Dz (y) := [z, y]
Example 13.21. Every differential graded Lie algebra (g, [· , ·], D) leads to a coho-
mology space, which is itself a pointed differential graded Lie algebra. To see this,
consider Dk : gk → gk+1 , the restriction of the differential D to gk . Since D ◦ D = 0,
we have a complex
Dk−1 Dk Dk+1
··· gk−1 gk gk+1 ···
The cohomology of this complex is denoted by HD• (g) = k
k∈Z HD (g), where, for
all k ∈ Z,
HDk (g) := Ker Dk / Im Dk−1 .
D is a graded derivation of degree one of (g, [· , ·]), hence [· , ·] induces a Lie bracket
on HD• (g), also denoted by [· , ·] and the triple (HD• (g), [· , ·], 0) is a pointed differential
graded Lie algebra.
We now show that the differential and the bracket of a differential graded Lie algebra
(g, [· , ·], D) can be encoded in a single differential Λ on S• g. Let g be a graded vector
space, let [· , ·] : g × g → g be a graded skew-symmetric bilinear map of degree zero
and let D : g → g be a graded linear map of degree one. Two linear maps from S• g
to itself are defined by setting for every homogeneous monomial X = x1 . . . x ∈ S g,
ΛD (X) := ∑ sgn(σi ; X)D(xi )X̂i ,
i=1 (13.29)
∑
0 (x
Λ[· ,·] (X) := sgn(σi j ; X)(−1)d i) [xi , x j ] X̂i j ,
1i< j
where X̂i and X̂i j are obtained by removing from X the element xi , respectively the
elements xi and x j . For example, X̂i = x1 . . . xi−1 xi+1 . . . x . The permutation σi ∈ S
is the permutation which sends 1 to i and subtracts one from the integers 2, . . . , i and
fixes the other integers, i.e., it is the cycle (i, i − 1, . . . , 1); similarly, σi j is the per-
mutation which sends 1 to i, sends 2 to j, removes two from the integers 3, . . . , i + 1,
removes one from the integers i + 2, . . . , j and fixes the other integers. We leave it as
an exercise to the reader to check that the maps ΛD and Λ[· ,·] are well-defined. We
define Λ to be the sum of the two linear maps in (13.29),
where σi jk and X̂i jk are defined similarly to σi j and X̂i j , defined earlier in this section,
and
αi jk (X) := (−1)|xi ||xk | [[xi , x j ] , xk ] + (i, j, k) .
It follows that Λ[·2,·] = 0 if and only if [· , ·] satisfies the graded Jacobi identity. Finally,
where
βi j (X) := D([xi , x j ]) − [D(xi ), x j ] − (−1)|xi | [xi , D(x j )] ;
as a consequence, ΛD ◦ Λ[· ,·] + Λ[· ,·] ◦ ΛD = 0 if and only if D is a derivation of [· , ·].
The last three equivalences, combined, prove the equivalence of (i) and (ii).
We finish this section by giving the definition of a morphism of differential
graded Lie algebras.
Definition 13.23. Let (g, [· , ·]g , Dg ) and (h, [· , ·]h , Dh ) be two differential graded Lie
algebras. A graded linear map φ : g → h of degree zero is said to be a morphism of
differential graded Lie algebras, if φ is a homomorphism of graded Lie algebras
and φ ◦ Dg = Dh ◦ φ .
Let us denote by Λg and Λh the total differentials of (g, [· , ·]g , Dg ), respectively
of (h, [· , ·]h , Dh ). It follows easily from the definitions that a graded linear map of
degree zero φ : g → h is a morphism of differential graded Lie algebras if and only
if S• φ ◦ Λg = Λh ◦ S• φ , where S• φ : S• g → S• h is defined by
376 13 Deformation Quantization
for all x1 , . . . , x ∈ g.
this form, we will often write xk (or (x)k in the case of ambiguity) for the coefficient
of ν k in x, a notation which comes in particularly handy when picking a particular
coefficient of the series of a complex expression.
We keep the same notations [· , ·] and D for the Fν -linear extension of [· , ·] and
of D to gν . Similarly, when Λ denotes the total differential on S• g, then Λ will still
denote its extension to (S• g)ν := (S• g)[[ν ]]. Consider the subspace ν gν1 := ν g1 [[ν ]]
of gν , which consists of all formal power series in ν without constant term and with
coefficients in g1 , the degree one component of g. On ν gν1 we consider the equation
1
D(x) + [x, x] = 0 , (13.32)
2
which is called the Maurer–Cartan equation associated to (g, [· , ·], D) (or to g). We
denote by MC(g) the set of all elements x of ν gν1 which are solutions of the Maurer–
Cartan equation associated to g. We will give in Proposition 13.26 below two char-
acterizations of the solutions of the Maurer–Cartan equation. To do this, we define,
for x ∈ ν gν1 and for k ∈ N∗ ,
1
Obsk (x) := D(xk ) +
2 ∑ [xi , x j ] . (13.33)
i+ j=k
i, j1
Then (13.32) can be written out as the infinite list of equations Obsk (x) = 0,
where k ∈ N∗ . If x satisfies Obsk (x) = 0 for k = 1, . . . , we say that x is a solu-
tion of the Maurer–Cartan equation up to order and we denote by MC (g) the set
of all such elements. Clearly, x ∈ MC (g) if and only if x satisfies (13.32) modulo
13.3 Differential Graded Lie Algebras 377
ν +1 and x ∈ MC(g) if and only if x ∈ MC (g) for all ∈ N∗ . Moreover, we have
the following proposition.
Proposition 13.24. Let (g, [· , ·], D) be a differential graded Lie algebra. Suppose
that x ∈ MC (g) for some ∈ N∗ . Then Obs+1 (x) is a cocycle.
Proof. Since D is a graded derivation, satisfying D2 = 0, and since the Lie bracket
is graded skew-symmetric, we have that
1
D (Obs+1 (x)) =
2 ∑ ([D(x j ), xk ] − [x j , D(xk )]) = ∑ [D(x j ), xk ] .
j+k=+1 j+k=+1
j,k1 j,k1
x2 xk
ex := 1 + x + + · · · + + · · · ∈ (S• g)ν . (13.34)
2 k!
Notice that this infinite sum is a well-defined formal power series in ν with co-
efficients in S• g since at most the first k + 1 terms of the infinite sum in (13.34)
contribute to the coefficient of ν k in ex (recall that x is a formal power series in
ν without constant term). The total differential Λ , applied to ex , can be explicitly
computed, as is shown in the following lemma.
Lemma 13.25. Let (g, [· , ·], D) be a differential graded Lie algebra with total differ-
ential Λ : S• g → S• g. For x ∈ ν gν1 , one has
1
Λ (ex ) = D(x) + [x, x] ex . (13.35)
2
Proof. One computes directly that
Proposition 13.26. Let (g, [· , ·], D) be a differential graded Lie algebra with total
differential Λ : S• g → S• g. For x ∈ ν gν1 , the following statements are equivalent.
(i) x satisfies the Maurer–Cartan equation associated to g;
(ii) Λ (ex ) = 0;
(iii) Obsk (x) = 0 for every k ∈ N∗ .
In the case of a pointed differential graded Lie algebra (g, [· , ·], z), these conditions
are also equivalent to
(iv) [z + x, z + x] = 0.
We have introduced in Sections 13.1.1 and 13.2.1 the notion of equivalence of de-
formations in the case of associative products and of Poisson brackets. We now
introduce the corresponding notion for formal power series (without constant term)
with coefficients in the degree one component of an arbitrary pointed differential
graded Lie algebra.
Definition 13.27. Let (g, [· , ·], z) be a pointed differential graded Lie algebra. Two
formal power series without constant term x, y ∈ ν gν are said to be gauge equivalent
denoted by x ∼ y, if there exists ξ ∈ ν gν0 such that
z + y = eadξ (z + x) .
The definition requires some explanation. As the notation suggests,2 eadξ stands for
the endomorphism of gν , defined for x ∈ gν by
∞ adkξ x 1
e adξ
(x) := ∑ k!
= x + [ξ , x] + [ξ , [ξ , x]] + · · · .
2
(13.36)
k=0
The latter infinite sum is a well-defined formal power series in ν with coefficients
in g since at most the first k + 1 terms of the sum contribute to the coefficient of ν k .
Notice that it follows easily from (13.36) that
d adt ξ
e (x) = ξ , eadt ξ (x) , (13.37)
dt
for ξ ∈ ν gν0 and x ∈ gν . When we work in a fixed pointed differential graded Lie
algebra (g, [· , ·], z), we will make extensive use of the notation
∞
1 k
ξ x := eadξ (z + x) − z = x + ∑ adξ (z + x) ,
k=1 k!
2The endomorphism eadξ of gν , defined here, should not be confused with the exponential eξ ,
which is the element of (S• g)ν , defined in (13.34).
13.3 Differential Graded Lie Algebras 379
where x and ξ are as above. In terms of this notation, x, y ∈ ν gν are gauge equivalent
if and only if there exists ξ ∈ ν gν0 such that y = ξ x.
In order to show that the relation ∼ is an equivalence relation, we use the
Campbell–Hausdorff formula
where CH(u, v) is an infinite sum, each of whose terms consists only of repeated
brackets of u and v. The first few terms of CH(u, v) are given by
1 1
CH(u, v) = u + v + [u, v] + ([u, [u, v]] + [v, [v, u]]) + · · · . (13.39)
2 12
In particular, CH(u, −u) = 0 for all u. This formula is usually given for u and v in
a finite-dimensional Lie algebra, with exp being the exponential map between the
Lie algebra and its Lie group (see for example [29]), but the Campbell–Hausdorff
formula is essentially a formal identity, in particular it is equally valid when for
example u and v are formal power series (without constant term) with coefficients
in an arbitrary Lie algebra. In the present case, we apply it for u and v of the form
adξ , where ξ is a formal power series without constant term and the Lie bracket is
the graded commutator (of graded linear maps, such as adξ ). For ξ , η ∈ ν gν0 and
x ∈ ν gν , the Campbell–Hausdorff formula and the fact that ad is a representation
imply that
eadη eadξ (x) = eCH(adη ,adξ ) (x) = eadCH(η ,ξ ) (x) ,
so that
η (ξ x) = CH(η , ξ ) x . (13.40)
The fact that ∼ is an equivalence relation follows at once from it.
To finish this section, we show that the equivalence relation ∼ is compatible with
the Maurer–Cartan equation, i.e., that if x ∼ y, then x is a solution of the Maurer–
Cartan equation if and only if y is a solution of the Maurer–Cartan equation. To do
this, let x, y ∈ ν gν1 be two equivalent solutions of the Maurer–Cartan equation and let
ξ ∈ ν gν0 be such that y = ξ x. Since x is a solution of the Maurer–Cartan equation
if and only if [z + x, z + x] = 0, we need to show that
eadξ (z + x), eadξ (z + x) = 0 .
For s,t ∈ gν , the fact that adξ is a derivation of the Lie bracket implies that
∞
1 j ∞ j
1 j i
e adξ
[s,t] = ∑ adξ [s,t] = ∑ ∑ adξ s, adξj−i t
j=0 j! j=0 i=0 j! i
∞ j ∞
1 i 1 1 k 1
= ∑∑ adξ s, adξj−i t = ∑ adξ s, adξ t
j=0 i=0 i! ( j − i)! k,=0 k! !
380 13 Deformation Quantization
= eadξ s, eadξ t .
We may conclude that eadξ (z + x), eadξ (z + x) = eadξ [z + x, z + x] = 0, since x is a
solution of the Maurer–Cartan equation.
The upshot is that gauge equivalence defines an equivalence relation on the set of
solutions of the Maurer–Cartan equation associated to g. For x ∈ MC(g), we denote
by cl(x) its equivalence class modulo gauge equivalence.
where each one of the functions xk (t) is a polynomial function, with values in V . We
will often consider such paths satisfying x0 (t) = 0; they will be referred to as poly-
nomial paths in ν V ν . Polynomial paths in V ν admit well-defined derivatives with
respect to t, which are themselves also polynomial paths in V ν . Also, polynomial
paths in a (graded) algebra, such as (S•V )ν , form an algebra and, for every polyno-
mial path γ (t) in ν V ν , its exponential eγ (t) (defined as in (13.34)) is a polynomial
path in (S•V )ν .
Definition 13.28. Let (g, [· , ·], z) be a pointed differential graded Lie algebra, and
let Λ denote the total differential on (S• g)ν , as in (13.30). We say that two solutions
x, y ∈ ν gν1 of the Maurer–Cartan equation are path equivalent if there exists two
polynomial paths γ (t), η (t) in ν gν1 and ν gν0 respectively, such that γ (0) = x, γ (1) = y
and
d γ (t)
e = Λ η (t)eγ (t) . (13.41)
dt
We will show in Proposition 13.31 below that two solutions of the Maurer–Cartan
equation are gauge equivalent if and only if they are path equivalent. We will do this
by using the following two lemmas, the first one of which can be viewed as giving
yet another (equivalent!) notion of equivalence of solutions of the Maurer–Cartan
equation.
13.3 Differential Graded Lie Algebras 381
Lemma 13.29. Let (g, [· , ·], z) be a pointed differential graded Lie algebra and
suppose that x, y ∈ ν gν1 are two solutions of the Maurer–Cartan equation asso-
ciated to g. Let γ (t) and η (t) be polynomial paths in ν gν1 and ν gν0 respectively,
with γ (0) = x and γ (1) = y. Then condition (13.41) is equivalent to the following
condition:
d
γ (t) = [z + γ (t), η (t)] . (13.42)
dt
Proof. If condition
(13.41)
holds, then the component in gν = S1 gν ⊂ (S• g)ν of
d γ (t) γ
dt e − Λ η (t)e (t) vanishes. Since the components of γ (t) ∈ ν gν1 commute, this
component is given by
d d
γ (t) − ΛDz (η (t)) − Λ[· ,·] (η (t)γ (t)) = γ (t) − [z + γ (t), η (t)] ,
dt dt
condition (13.41) implies condition (13.42).
Conversely, suppose that γ (t) and η (t) satisfy condition (13.42), where γ (0) = x
is a solution of the Maurer–Cartan equation and γ (1) = y. We first show that γ (t)
is a solution of the Maurer–Cartan equation for all t, which we do by showing that
[z + γ (t), z + γ (t)] = 0 for all t. This is in turn done by showing that the polynomial
path F(t) := [z + γ (t), z + γ (t)] is the unique solution of a linear differential equa-
tion, with initial condition F(0) = 0. Taking the derivative of F and using (13.42)
and the graded Jacobi identity we find
d d d
F(t) = γ (t), z + γ (t) + z + γ (t), γ (t)
dt dt dt
= [[z + γ (t), η (t)] , z + γ (t)] + [z + γ (t), [z + γ (t), η (t)]]
= [[z + γ (t), z + γ (t)] , η (t)]
= [F(t), η (t)] ,
which yields the required linear differential equation. We write F(t) = ∑k1 Fk (t)ν k
and η (t) = ∑k1 ηk (t)ν k , where all coefficients Fk (t) and ηk (t) depend polynomi-
ally on t. Then the differential equation can be written out as an infinite list of
equations
k−1
d
Fk (t) = ∑ [Fk− (t), η (t)] and Fk (0) = 0 ,
dt =1
to find that
382 13 Deformation Quantization
Λ η (t)eγ (t) = [z + γ (t), η (t)] eγ (t) + Λ eγ (t) η (t)
= [z + γ (t), η (t)] eγ (t) ,
where we have used in the last step that γ (t) ∈ MC(g) for all t, combined with
item (ii) of Proposition 13.26. It follows that
d γ (t) d
e − Λ η (t)eγ (t) = γ (t) − [z + γ (t), η (t)] eγ (t) ,
dt dt
The second lemma which we will use to prove Proposition 13.31 below can
be seen as a generalization of the classical formula, giving the derivative of e f (t) ,
when f (t) takes values in a non-commutative algebra.
Lemma 13.30. Let (g, [· , ·]) be a graded Lie algebra and let ξ (t) be a polynomial
path in ν gν0 . Then
dead−ξ (t)
eadξ (t) ◦ = − ad 1 dξ (t) . (13.43)
dt ∑ (k+1)! adkξ (t) ( dt )
k∈N
Proof. Given a polynomial path ξ (t) in ν gν0 , we consider for fixed t the polynomial
path in End(g1 )ν , defined for s ∈ F by
dead−sξ (t)
Gs := eadsξ (t) ◦ + ad sk+1 dξ (t) . (13.44)
dt ∑ (k+1)!
adkξ (t) dt
k∈N
For the derivative of Bs , it follows at once from the second equation in (13.46) that
d
Bs (x) = eadsξ (t) ξ̇ (t), x ,
ds
so that, in view of the graded Jacobi identity,
sk+1
[ξ (t), Bs (x)] = ∑ ξ (t) ξ̇ (t), x + Bs ([ξ (t), x])
adk+1
k∈N (k + 1)!
d
= Bs (x) − ξ̇ (t), x + Bs ([ξ (t), x]) ,
ds
giving the following formula for the derivative of Bs :
d
Bs (x) = [ξ (t), Bs (x)] − Bs ([ξ (t), x]) + ξ̇ (t), x . (13.48)
ds
Equations (13.47) and (13.48), combined, yield the proof of (13.45), and hence
of (13.43).
Using the two lemmas, we show that gauge equivalence and path equivalence are
the same (for solutions of the Maurer–Cartan equation).
Proposition 13.31. Let (g, [· , ·], z) be a pointed differential graded Lie algebra. Two
solutions x, y ∈ ν gν1 of the Maurer–Cartan equation associated to g are gauge equiv-
alent if and only if they are path equivalent.
Proof. Let us first assume that x and y are gauge equivalent. Then there exists ξ ∈
ν gν0 such that y = ξ x. For t ∈ F, define
γ (t) := (t ξ ) x = eadt ξ (z + x) − z ,
and let η (t) := −ξ , which is a constant polynomial path. Then γ (0) = x, γ (1) = y
and it follows from (13.37) that γ (t) and η (t) satisfy condition (13.42). In view of
Lemma 13.29, x and y are path equivalent.
Conversely, let us assume that x and y are path equivalent, and let γ (t), η (t) be
polynomial paths in ν gν1 and ν gν0 respectively, such that γ (0) = x, γ (1) = y and such
that the equivalent conditions (13.41) and (13.42) are satisfied. In order to establish
that x, y are gauge equivalent, we will construct a polynomial path ξ (t) in ν gν0 such
that γ (t) = ξ (t) x, i.e.,
for all t ∈ F. It implies that y = γ (1) = ξ (1) x, so that x and y are gauge equivalent.
In order to satisfy (13.49) for t = 0, we require ξ (0) = 0. With this additional as-
sumption, the condition which we wish to impose on ξ (t) is dtd e− adξ (t) (z + γ (t)) = 0,
or equivalently:
d − adξ (t)
eadξ (t) e (z + γ (t)) = 0 .
dt
We can compute the left-hand side of the previous equation with the help of (13.43)
and (13.42) as follows:
1 dξ (t)
∑ (k + 1)! adkξ (t) dt
= −η (t) (13.50)
k∈N
and ξ (0) = 0. To see that such a polynomial path ξ (t) exists, substitute ξ (t) =
∑∈N∗ ξ (t)ν and η (t) = ∑∈N∗ η (t)ν in (13.50) and observe that the equation
decomposes into an infinite number of equations
We pointed out in Example 13.17 that for every Poisson algebra (A , ·, π ), the triple
(X• (A ), [· , ·]S , π ) is a pointed differential graded Lie algebra. Similarly, accord-
ing to Example 13.19, for every commutative associative algebra (A , μ ), the triple
( HC• (A ), [· , ·]G , μ ) is a pointed differential graded Lie algebra. Up to a sign, the
differentials Dπ and Dμ coincide with the Poisson and Hochschild coboundary op-
erators, namely
for all F, G ∈ A . Writing x for μ + x, the left-hand side of the above equation is
given by
eξ x e−ξ (F), e−ξ (G)
1 1 1 r s
= x (F, G) + ∑∗ ∑ (−1)s+t
r! s! t!
ξ x (ξ (F), ξ t (G))
k∈N r,s,t∈N
r+s+t=k
386 13 Deformation Quantization
1
= x (F, G) + ∑∗ k! adkξ (x )(F, G) = eadξ (x )(F, G)
k∈N
1
∑ ∑ sgn(σ ; X)
! ∏
Φi j (xσ (i1 +···+i j−1 +1) . . . xσ (i1 +···+i j ) ) , (13.52)
i1 ,...,i ∈N∗ σ ∈Si1 ,...,i j=1
i1 +···+i =k
Remark 13.35.
(1) Φ is the (unique) extension of Φ as a coderivation of S• g with values in S• h,
but this fact will not be used in what follows.
(2) It is clear that, when Φ is identically zero on Sk g for all k 2, i.e., when Φ is
simply a graded linear map of degree 0 from g to h, then Φ = S• Φ (see (13.31)).
(x1 . . . xk ) is zero for all > k.
(3) It is also clear that the component in S h of Φ
(4) The order of the terms in the product in (13.52) is important: the meaning of
the product is that the terms are ordered as follows: Φi1 Φi2 . . . Φi .
Definition 13.36. Let (g, [· , ·]g , Dg ) and (h, [· , ·]h , Dh ) be two differential graded Lie
algebras, whose total differentials are denoted by Λg and by Λh respectively. A
graded linear map Φ : S• g → h of degree 0 is called an L∞ -morphism between g
and h if
◦ Λg = Λh ◦ Φ
Φ . (13.53)
and
(xy) = Dh (Φ2 (xy)) + Dh (Φ1 (x))Φ1 (y)
Λh ◦ Φ
0
+ (−1)d (x) [Φ1 (x), Φ1 (y)]h + Φ1 (x)Dh (Φ1 (y)) .
Equating the right-hand sides of these two equations yields, upon using that Φ1 ◦
Dg = Dh ◦ Φ1 , the following equation, involving both the maps Φ1 and Φ2 :
3 Strictly speaking, Φ1 is a map from S1 g → S1 h, but we rather consider it as a map from g to h,
i.e., we consider g and h with their original grading, as differential graded Lie algebras; with this
convention Φ1 induces maps in cohomology with the following labeling: Φ1k : HDk g (g) → HDk h (h).
13.4 L∞ -Morphisms of Differential Graded Lie Algebras 389
The infinite sum makes sense because, for every k ∈ N∗ , at most the first k terms
contribute to the coefficient of ν k in Ω̃Φ (x). Notice that if x ∈ ν gν1 , then all elements
xk are formal power series whose coefficients are elements of degree 0 in S• g, since
xk is of degree 0 when x ∈ g1 (or x ∈ gν1 ). It follows that Φ̃k (xk ) ∈ ν hν1 for all k ∈ N∗ ,
hence Ω̃Φ (x) ∈ ν hν1 .
Sections 13.4.1 to 13.4.3 below are devoted to the proof of the following theorem.
Theorem 13.39. Let (g, [· , ·]g , zg ) and (h, [· , ·]h , zh ) be two pointed differential gra-
ded Lie algebras. If Φ : S• g → h is an L∞ -morphism, then
ΩΦ : MC(g)/ ∼ → MC(h)/ ∼
cl(x) → cl Ω̃Φ (x)
13.4.1 ΩΦ is Well-Defined
We start with a proposition which will be useful for showing that ΩΦ is well-defined.
Proposition 13.40. Let (g, [· , ·]g , Dg ) and (h, [· , ·]h , Dh ) be two differential graded
Lie algebras. Let Φ : S• g → h be a graded linear map of degree 0 and let Ω̃Φ be the
associated map, defined as in (13.56).
(1) If x ∈ ν gν1 , then
(ex ) = eΩ̃Φ (x) .
Φ (13.57)
(2) If x ∈ ν gν1 and ω ∈ ν gν , then
(ω e ) = ∑ 1
Φ x
Φk (ω x ) eΩ̃Φ (x) .
k−1
(13.58)
k∈N∗ (k − 1)!
390 13 Deformation Quantization
1 1
eΩ̃Φ (x) = 1 + ∑∗ ! ∑ Φi1 (xi1 ) . . . Φi (xi ) . (13.59)
∈N i1 ,...,i ∈N∗ 1 . . i !
i ! .
For every k ∈ N∗ and for every permutation σ ∈ Sk , one has sgn(σ ; xk ) = 1, because
x ∈ ν gν1 . Hence, by definition of Φ̂ , we have:
1
(ex ) =
Φ ∑ k! Φ (xk )
k∈N
1 1
= 1+ ∑∗ k! ∑∗ ∑ ! σ ∈S∑
Φi1 (xi1 ) . . . Φi (xi )
k∈N ∈N i1 +···+i =k i ,...,i
1
i1 ,...,i 1
1 1
= 1+ ∑∗ ! ∑ ∗ (i + · · · + i )! ∑ Φi1 (xi1 ) . . . Φi (xi )
∈N i1 ,...,i ∈N
1 σ ∈S i ,...,i
1
which is the same as the right-hand side of (13.59), since #Si1 ,...,i = (i1i+···+i
1 !...i !
)!
. This
proves (1).
In order to prove (2), we introduce a formal parameter t and we compute the
coefficient in t of both sides of the equality
The proof that this equality holds is the same as the proof of (13.57) since everything
is computed modulo t 2 , so that all products of elements of odd degree in h, which
are precisely the ones which do not commute, are discarded. For the same reason,
the coefficient of t in Φ (ex+t ω ) is given by Φ (ω ex ), which is the left-hand side
of (13.58), and the coefficient of t in e Φ Ω̃ (x+t ω ) is given by the coefficient of t of
eΩ̃Φ (x) eZ1 (t) , where Z1 (t) is the linear term in t of Ω̃Φ (x + t ω ), which is given by
k−1
1 1
Z1 (t) = ∑∗ k! ∑ Φk (x (t ω )xk−−1) = ∑∗ (k − 1)! Φk (ω xk−1 )t ,
k∈N =0 k∈N
so that the coefficient of t in eΩ̃Φ (x+t ω ) is given by the right-hand side of (13.58),
which establishes (2).
Proposition 13.41. Let (g, [· , ·]g , zg ) and (h, [· , ·]h , zh ) be two pointed differential
graded Lie algebras. Let Φ : S• g → h be an L∞ -morphism and let Ω̃Φ : ν gν → ν hν
be the associated map, defined as in (13.56).
(1) If x ∈ ν gν1 is a solution of the Maurer–Cartan equation associated to g, then
Ω̃Φ (x) is a solution of the Maurer–Cartan equation associated to h;
(2) If x, y ∈ ν gν1 are gauge equivalent solutions of the Maurer–Cartan equation
associated to g, then Ω̃Φ (x) and Ω̃Φ (y) are gauge equivalent;
13.4 L∞ -Morphisms of Differential Graded Lie Algebras 391
(3) If x ∈ ν gν1 and ξ ∈ ν k gν0 for some k ∈ N, then there exists s ∈ ν k hν0 , such that
Ω̃Φ (ξ x) = s Ω̃Φ (x). Moreover, s = Φ1 (ξ ) (mod ν k+1 ).
d γ (t)
e = Λg η (t)eγ (t) .
dt
Now Ω̃Φ (γ (t)) is a polynomial path in ν hν1 and χ (t) is a polynomial path in ν hν0 .
Equation (13.62) says that Ω̃Φ (x) = Ω̃Φ (γ (0)) and Ω̃Φ (y) = Ω̃Φ (γ (1)) are path
equivalent, hence gauge equivalent (by Proposition 13.31), which is the content
of (2).
According to Remark 13.32, if x ∈ ν gν1 and ξ ∈ ν k gν0 , then the polynomial
path η (t) which is used in establishing the path equivalence between x and ξ x
can be chosen in ν k gν0 , namely we can choose η (t) := −ξ . Then the polynomial
path χ (t), given by (13.63), belongs to ν k hν0 so that, again by Remark 13.32, the
gauge equivalence between Ω̃Φ (ξ x) and Ω̃Φ (x) can be realized by an element
s ∈ ν k hν0 . Since χ (t) = Φ1 (η (t)) (mod ν k+1 ), we have s = Φ1 (ξ ) (mod ν k+1 ),
which proves (3).
392 13 Deformation Quantization
13.4.2 Surjectivity of ΩΦ
Let ∈ N∗ and suppose that x ∈ MC (g). Then Dg (x) + 12 [x, x]g = 0 (mod ν +1 ),
hence the right-hand side of equation (13.67) is zero, modulo ν +1 . In view
of (13.65), the right-hand side of (13.66) is also zero, modulo ν +1 , so that
1
Dh (Ω̃Φ (x)) + Ω̃Φ (x), Ω̃Φ (x) h
=0 (mod ν +1 ) .
2
This shows that Ω̃Φ (x) ∈ MC (h), which proves the first part of the proposition.
For x ∈ MC (g), the first part of the proof and (13.66) imply that
Λh Φ (ex ) = Λh (eΩ̃Φ (x) ) = Obs+1 (Ω̃Φ (x)) . (13.68)
+1 +1
Moreover, letting ω := Dg (x) + 12 [x, x]g ∈ ν +1 gν2 we obtain, using (13.67) and
Proposition 13.40,
1
Λh Φ x (ω e ) = ∑
(e ) = Φ x
Φk (ω x ) eΩ̃Φ (x) .
k−1
(13.69)
k∈N ∗ (k − 1)!
Therefore,
13.4 L∞ -Morphisms of Differential Graded Lie Algebras 393
(ex )
Λh Φ = (Φ1 (ω ))+1 = Φ1 (ω+1 ) = Φ1 (Obs+1 (x)) ,
+1
in particular Obsk (x) − Obsk (x ) is a coboundary. This fact follows immediately
from the definition (13.33) of the operator Obsk .
Suppose that y ∈ MC(h) (in particular, all non-zero coefficients yk of y belong to
h1 ). We show that there exist formal power series x ∈ ν gν1 and ξ ∈ ν hν0 , such that
y = ξ Ω̃Φ (x). These series will be constructed term by term. Thus, we assume that
we have constructed the first k terms x1 , . . . , xk of x and the first k terms ξ1 , . . . , ξk
of ξ , such that
(ak ) xk := x1 ν + · · · + xk ν k ∈ MCk (g);
(bk ) ξ k := ξ1 ν + · · · + ξk ν k and xk satisfy the following equation:
We show how to construct an extra term xk+1 for x and ξk+1 for ξ , such that (ak+1 )
and (bk+1 ) are satisfied. Notice that (a0 ) and (b0 ) are trivially satisfied, since by
construction x0 = 0 and ξ 0 = 0, while the constant term of y is zero.
First we construct xk+1 ∈ g1 such that xk+1 := xk + xk+1 ν k+1 ∈ MCk+1 (g). In
view of Proposition 13.42, (bk ) and (13.70), there exists u ∈ h1 such that
Φ1 (Obsk+1 (xk )) = Obsk+1 (Ω̃Φ (xk )) = Obsk+1 (−ξ k ) y + D(u) = D(u) ,
where we used in the last step that y, and hence (−ξ k ) y, belongs to MC(h). Since
xk ∈ MCk (g), Proposition 13.24 implies that Obsk+1 (xk ) is a cocycle, so it defines
a cohomology class in HD2 zg (g); by the above computation, the image under Φ1 of
this cohomology class is trivial. Now Φ12 : HD2 zg (g) → HD2 z (h) is injective, since Φ
h
is an L∞ -quasi-isomorphism, hence the cohomology class of Obsk+1 (xk ) is trivial,
i.e., there exists xk+1
∈ g1 satisfying Obsk+1 (xk ) = xk+1 , zg g . Letting xk+1 := xk +
xk+1 ν k+1 , we have that
As xk+1 ∈ MCk+1 (g), Proposition 13.42 implies that Ω̃Φ (xk+1 ) ∈ MCk+1 (h), hence
ξ k Ω̃Φ (xk+1 ) ∈ MCk+1 (h), so that Obsk+1 ξ k Ω̃Φ (xk+1 ) = 0. Moreover, since
y ∈ MC(h), we also have Obsk+1 (y) = 0, so that (13.71) gives Dzh (yk+1 −Xk+1 ) = 0.
Since Φ11 : HD1 zg (g) → HD1 z (h) is surjective, there exists a cocycle xk+1 ∈ g1 and an
h
element ξk+1 ∈ h0 , such that
yk+1 − Xk+1 = Φ1 (xk+1 ) − Dzh (ξk+1 ) = Φ1 (xk+1 ) + ξk+1 , zh h
,
Remark 13.43. Notice that the fact that Φ1• is an isomorphism is not fully used in
the proof. In fact, we only used that Φ12 : HD2 zg (g) → HD2 z (h) is injective and that
h
Φ11 : HD1 zg (g) → HD1 z (h) is surjective.
h
13.4 L∞ -Morphisms of Differential Graded Lie Algebras 395
13.4.3 Injectivity of ΩΦ
In this section we prove that ΩΦ is injective, which is also part of the statement of
Theorem 13.39. To do this, we use the following lemma. Recall that, for a formal
power series x (in ν ), we denote by xk its coefficient of ν k .
Lemma 13.44. Let (g, [· , ·], z) be a pointed differential graded Lie algebra. Suppose
that x ∈ ν gν1 and that ξ ∈ ν k gν0 .
(1) (ξ x − x)k = −Dz (ξk );
(2) If s ∈ g0 is a cocycle, then (ξ + sν k ) x = ξ x (mod ν k+1 );
(3) If ξ x = x (mod ν k+1 ), then ξk is a cocycle;
(4) If ξk is a coboundary and x is solution of the Maurer–Cartan equation, then
there exists ξ ∈ ν k+1 gν0 , such that ξ x = ξ x.
Proof. Items (1)–(3) follow from the fact that, since ξ ∈ ν k gν0 , the term in ν in
ξ x is x when 1 < k and is xk + [ξk , z] = xk − Dz (ξk ) when = k.
For item (4), assume that ξk is a coboundary, ξk = Dz (η ) for some η ∈ g−1 , and
that x ∈ MC(g), so that [z + x, z + x] = 0. For w ∈ ν gν−1 , the graded Jacobi identity
implies that ad[w,z+x] (z + x) = − 12 [[z + x, z + x] , w] = 0, so that
1
[w, z + x] x = x + ∑∗ k! adk[w,z+x] (z + x) = x .
k∈N
We first show that the existence of these sequences satisfying the properties (Ak )
and (Bk ) proves that x and y are gauge equivalent; properties (Ck ) and (Dk ) will
turn out to be useful for constructing the sequences satisfying the former properties.
Consider the formal power series Ξ := limk→∞ Ξ (k) , where the polynomials Ξ (k) are
recursively defined by Ξ (k) := CH(ξ (k) ν k , Ξk−1 ) for k ∈ N∗ and Ξ0 := 0. Notice that
the limit is well-defined, because Ξ (k) = Ξ (k−1) (mod ν k ) for k ∈ N∗ , and that it
belongs to ν gν0 . In view of (A1 ),. . . ,(Ak ), (Bk ), and (13.40),
so that
(k)
Ω̃Φ (x(k) ) − Ω̃Φ (y) = Φ1 xk+1 − yk+1 . (13.74)
k+1
The left-hand side of (13.74) is a coboundary, in view of (Dk ) and item (1) of
Lemma 13.44. Since Φ11 : HD1 zg (g) → HD1 z (h) is injective, (x(k) − y)k+1 is also
h
(k)
a coboundary, i.e., there exists ξ (k+1) ∈ g0 such that Dzg (ξ (k+1) ) = xk+1 − yk+1 .
(k)
Now ((ξ (k+1) ν k+1 ) x(k) )k+1 = xk+1 + ξ (k+1) , zg , so that (ξ (k+1) ν k+1 ) x(k) = y
(mod ν k+2 ). Defining x(k+1) := (ξ (k+1) ν k+1 ) x(k) , it follows that x(k+1) and ξ (k+1)
satisfy (Ak+1 ) and (Bk+1 ).
(k)
The constructed element ξ (k+1) does not satisfy (Ck+1 ), namely uk+1 + Φ1 (ξ (k+1) )
need not be a coboundary, but it is a cocycle. In fact, in view of (Ak+1 ) and (Bk+1 ),
Ω̃Φ (ξ (k+1) ν k+1 x(k) ) = Φ1 (ξ (k+1) ν k+1 ) Ω̃Φ (x(k) ) (mod ν k+2 ) ,
13.5 Kontsevich’s Formality Theorem and Its Consequences 397
so that, still modulo ν k+2, the left-hand side of (13.75) is CH(u(k) , Φ1 (ξ (k+1) ν k+1 ))
Ω̃Φ (x(k) ). According to (13.39),
(k)
(CH(u(k) , Φ1 (ξ (k+1) ν k+1 )))k+1 = uk+1 + Φ1 (ξ (k+1) ) ,
(k)
so that we can apply item in (3) of Lemma 13.44 to (13.75), to conclude that uk+1 +
Φ1 (ξ (k+1) ) is a cocycle.
We now show how ξ (k+1) can be modified, so that (Ck+1 ) is satisfied, i.e., so that
(k)
uk+1 + Φ1 (ξ (k+1) ) is a coboundary. We will do this without much altering x(k+1) .
(k)
Since uk+1 + Φ1 (ξ (k+1) ) is a cocycle it defines a cohomology class in HD0 z (h);
h
since Φ10 : HD0 zg (g) → HD0 z (h) is surjective, this cohomology class is the image of
h
a cocycle κ ∈ g0 . Upon replacing ξ (k+1) by ξ (k+1) − κ , the new x(k+1) , defined by
(Ak+1 ), still satisfies (Bk+1 ), in view of item (2) of Lemma 13.44, i.e., removing a
cocycle from ξ (k+1) does not alter x(k+1) modulo ν k+1 . However, the cohomology
(k)
class of uk+1 + Φ1 (ξ (k+1) ) is now trivial, i.e., the latter element is a coboundary.
The last step of the construction is the construction of uk+1 , such that (Dk+1 ) is
satisfied. Item (3) in Proposition 13.41 yields the existence of s ∈ ν k+1 hν0 such that
Ω̃Φ (ξ (k+1) ν k+1 x(k) ) = s Ω̃Φ (x(k) ); by the same proposition, s = Φ1 (ξ (k+1) )
(mod ν k+2 ). Using (Ak+1 ) and (Dk ), it follows that
Ω̃Φ (x(k+1) ) = s Ω̃Φ (x(k) ) = s u(k) Ω̃Φ (y) = CH(s, u(k) ) Ω̃Φ (y) .
(k)
In view of (13.39), (CH(s, u(k) ))k+1 = Φ1 (ξ (k+1) ) + uk+1 , which is a coboundary,
according to (Ck+1 ). Item (4) in Lemma 13.44 implies that there exists u(k+1) in
ν k+2 hν0 such that
which is (Dk+1 ).
Remark 13.45. The fact that Φ1• is an isomorphism is not fully used in the proof of
the injectivity of ΩΦ . Indeed, we only used that Φ11 : HD1 zg (g) → HD1 z (h) is injec-
h
tive and that Φ10 : HD0 zg (g) → HD0 z (h) is surjective (see Remark 13.43 for a similar
h
remark concerning the proof of the surjectivity of ΩΦ ).
stead of giving a full proof of Kontsevich’s theorem, which is beyond the scope of
this book, we present some of the main ingredients, and we concentrate on a few of
its consequences, which have deep implications to both Poisson geometry and Lie
theory. The most popular consequence can be stated in a few words: the algebra of
smooth functions on a real Poisson manifold admits a formal deformation quantiza-
tion. This result will be proved in Theorem 13.51, as a consequence of Kontsevich’s
formality theorem (Theorem 13.46).
Let M be a real manifold and denote by μ the usual product on C∞ (M). Recall that
there are two pointed differential graded Lie algebras, which are naturally associated
to M:
• (X• (M), [· , ·]S , 0) is the graded vector space of multivector fields on M, with
shifted degree, equipped with the Schouten bracket and with the zero differen-
tial (see Example 13.18);
• ( HC•diff (M), [· , ·]G , μ ) is the graded vector space of differential Hochschild
cochains of C∞ (M), with shifted degree, equipped with the Gerstenhaber bracket
and with the Hochschild differential associated to μ (see Example 13.20).
First, we show how these differential graded Lie algebras are related. Let P be a
p-vector field on M, where p 1. A p-linear map φ (P) : (C∞ (M)) p → C∞ (M) is
naturally associated to P by setting, for all F1 , . . . , Fp ∈ C∞ (M),
1
φ (P)(F1 , . . . , Fp ) := P[F1 , . . . , Fp ] . (13.76)
p!
It leads for every p to a graded linear map of degree 0,
φ : X p (M) → HCdiff
p
(M) , (13.77)
where we used in the last equality that P is a derivation in each of its arguments.
It follows that the map φ takes values in the Hochschild cocycles of (C∞ (M), μ ).
In particular, φ induces a linear map H • φ between the cohomologies of the pointed
differential graded Lie algebras (X• (M), [· , ·]S , 0) and ( HC•diff (M), [· , ·]G , μ ). Ac-
cording to the Hochschild–Kostant–Rosenberg theorem (see [131]), H • φ is an iso-
13.5 Kontsevich’s Formality Theorem and Its Consequences 399
morphism of differential graded Lie algebras. The map φ itself is however not a
morphism of (differential) graded Lie algebras, since [P, Q]S = [P, Q]G for general
multivector fields P, Q as one easily checks from the definitions.
The following result, due to Kontsevich, states that, even if φ is not a mor-
phism of differential graded Lie algebras, it extends to an L∞ -morphism Φ between
(X• (M), [· , ·]S , 0) and ( HC•diff (M), [· , ·]G , μ ); since the induced map in cohomology
H • Φ1 is H • φ , which is according to the Hochschild–Kostant–Rosenberg theorem
an isomorphism, Φ is an L∞ -quasi-isomorphism (see Definition 13.37).
Theorem 13.46 (Kontsevich’s formality theorem). Let M be a real manifold and
denote by μ the usual product on C∞ (M). There exists an L∞ -quasi-isomorphism Φ
from (X• (M), [· , ·]S , 0) to ( HC•diff (M), [· , ·]G , μ ), such that the map Φ1 : X• (M) →
HC•diff (M) is the natural inclusion map φ , defined in (13.77).
For a proof of this theorem, which is quite involved and is beyond the scope of
this book, we refer to [38, 39, 107, 189] and [14]. We will give in the following
section some more details about the construction of the L∞ -quasi-isomorphism in
the case of M = Rd . But first, we say a few words about the origin of the name
of the theorem. Recall from Example 13.21 that the cohomology of a differen-
tial graded Lie algebra is itself also a (pointed) differential graded Lie algebra. In
the case of (X• (M), [· , ·]S , 0), the differential is trivial hence the latter differential
graded Lie algebra and its cohomology are isomorphic differential graded Lie al-
gebras. According to the Hochschild–Kostant–Rosenberg theorem, cited above, this
cohomology and the cohomology of ( HC•diff (M), [· , ·]G , μ ) are isomorphic differen-
tial graded Lie algebras. Thus, Kontsevich’s theorem says that there exists an L∞ -
quasi-isomorphism between ( HC•diff (M), [· , ·]G , μ ) and its cohomology. A differen-
tial graded Lie algebra which admits an L∞ -quasi-isomorphism to its cohomology
is called formal, a terminology which is borrowed from topology. Thus, Kontse-
vich’s theorem says that ( HC•diff (M), [· , ·]G , μ ) is formal, which explains the name
“formality theorem”.
Without going into technical details, we present in this section Kontsevich’s ex-
plicit construction of the L∞ -quasi-isomorphism Φ between (X• (M), [· , ·]S , 0) and
( HC•diff (M), [· , ·]G , μ ) in the particular case where M = Rd . In order to do this, we
first need to explain how multi-differential operators are naturally coded by graphs,
how weights are associated to these graphs and how they all sum up to yield a for-
mula for the L∞ -quasi-isomorphism Φ .
400 13 Deformation Quantization
We describe in this subsection the graphs which are used in Kontsevich’s construc-
tion. Recall that an oriented graph is a pair (V, A), where A is a subset of V × V .
The elements of V are called vertices and the elements of A arrows. For an arrow
a = (x, y) ∈ A, we call y its head, denoted h(a) and x its tail, denoted t(a). We will
assume that Γ contains no loops, which means that if (x, y) ∈ A, then x = y. For given
integers k, ∈ N, the vertex set of the graphs which are considered is the set Vk, ,
defined by
Vk, := {1, . . . , k} ∪ {1, . . . , } .
For reasons which will be clear later, it is demanded that these integers satisfy 2k +
− 2 0, that is the cases (k, ) = (0, 0) and (k, ) = (1, 0) are excluded. The k
vertices 1, . . . , k of Vk, are called vertices of the first type, while the vertices 1, . . . ,
of Vk, are called vertices of the second type. Another piece of data added to the
graphs which are considered here is an ordering of each of the sets of arrows with a
common tail, i.e., of each of the sets
Star(s) := {a ∈ A | t(a) = s} ,
for s ∈ {1, . . . , k}. Denoting the number of elements in Star(s) by ps , such an order-
ing amounts to a numbering σs of the elements of Star(s) by the integers 1, . . . , ps ;
using this numbering we write Star(s) as Star(s) = {σs1 , . . . , σsps } .
Definition 13.47. Let k, ∈ N, such that 2k + − 2 0. A triplet Γ = (Vk, , A, σ ) is
said to be a Kontsevich graph if
(1) The pair (Vk, , A) is an oriented graph without loops, called the underlying
graph of Γ ;
(2) For every arrow a ∈ A, its tail t(a) is a vertex of the first type;
(3) For every s ∈ {1, . . . , k}, σs is an ordering of Star(s).
The set of Kontsevich graphs Γ = (Vk, , A, σ ) is denoted by Gk, . The underlying
graph of some element G3,4 is depicted in Fig. 13.1.
Fig. 13.1 The elements in Gk, are triplets (V, A, σ ), where the set of vertices consists of k vertices
of the first type, denoted 1, . . . , k and vertices of the second type, denoted 1, . . . , . All arrows start
from vertices of the first type and are ordered by σ . Depicted here is the graph (V, A) that underlies
some element of G3,4 . For reasons that will become clear later, the graph is drawn in the upper
half-plane, with the vertices of the second type on its boundary.
Fig. 13.2 For p and q, points in the Poincaré half-plane H , θ (p, q) denotes the angle θ between
the vertical geodesic through p and the geodesic through p and q.
(dx)2 + (dy)2
(ds)2 = ,
y2
whose geodesics are the vertical half-lines, starting from the horizontal axis (the
line y = 0) and half-circles, with center on the horizontal axis. For p, q ∈ H :=
H ∪ {y = 0}, with p = q, we denote by θ (p, q) the angle (at p) between the verti-
cal geodesic through p and the geodesic through p and q (see Fig. 13.2). The angle
function θ leads, for each arrow a of Γ , to an analytic function θa on a configu-
ration space of points, whose construction we briefly recall (see the appendix by
A. Bruguières in [39] for details). With k, as above, consider Confk, , the space
of (k + )-tuples of distinct points, the first k points belonging to H and the re-
maining points belonging to the boundary of H . We write a point of Confk, as a
(k + )-tuple (z1 , . . . , zk , z1 , . . . , z ), since we think of such a point as corresponding
to the vertices of a graph of Gk, , drawn in H . The group G of affine transforma-
tions of C of the form z → λ z + μ , with λ ∈ R∗+ and μ ∈ R, acts in a natural way
on Confk, . This action is free and the quotient manifold
402 13 Deformation Quantization
Fig. 13.3 The graph Γ belongs to G1, . It has one vertex of the first type and vertices of the
second type. Its weight is 1/(!)2 .
Ck, := Confk, /G
k ' ( dθa
1
ϖΓ := ∏ .
s=1 (# Star(s))! a∈A
2π
Ck,
We do not discuss here the technical issues, related to the fact that Ck, has corners
and concerning the choice of ordering in the above wedge product. We simply point
out that, for dimensional reasons, ϖΓ = 0 whenever the number of arrows in Γ is
different from the dimension 2k + − 2 of Ck, .
Example 13.48. Up to the ordering of its arrows, there is a single graph Γ in G1,
which has arrows. It is depicted in Fig. 13.3. The Kontsevich weight of Γ is given
by
' (
'
1 1 1
ϖΓ = dθ(1,t) = dθ1 ∧ · · · ∧ dθ = .
! (2π ) t=1
! (2π ) (!)2
C1, 0θ1 <···<θ 2π
(13.78)
13.5 Kontsevich’s Formality Theorem and Its Consequences 403
Fig. 13.4 To each vertex of a graph one associates a multivector field on Rd (a 0-vector field, i.e.,
a function, when the vertex is of the second type). Depicted is the graph of Fig. 13.1, with the extra
data.
Similarly, we associate a function DΓγ, s (Qs ) to each vertex s of the first type, by
setting
⎛ ⎞
⎜ ∂ ⎟
DΓγ, s (Qs ) := ⎜
⎝ ∏ ⎟ Qs x 1 , . . . , x ps
∂ xγ (a) ⎠ γ ( σs ) γ ( σs ) ,
a∈A
h(a)=s
where we recall that σs1 , . . . , σsps = Star(s) is the set of arrows whose tail is s,
ordered by σs . When the above products are taken over the empty set, i.e., if t,
respectively s, is not the head of an arrow in A, then the derivatives reduce to a
zero-th order derivative, i.e., no derivatives are taken. Taking the product of the thus
404 13 Deformation Quantization
constructed k + functions and summing over all maps γ : A → {1, . . . , d}, leads to
the function
k
BΓ (Q1 , . . . , Qk )(F1 , . . . , F ) := ∑ ∏ DΓγ, s (Qs ) ∏ DΓγ,t (Ft ) .
γ :A→{1,...,d} s=1 t=1
∂ F1 ∂ F2 ∂ F
BΓ (Q)(F1 , . . . , F ) = ∑ Q[xγ1 , . . . , xγ ]
∂ xγ1 ∂ xγ2
···
∂ xγ
1γ1 ,...,γ d
= Q[F1 , . . . , F ],
Example 13.50. Similarly, we write down BΓ , where Γ is the graph of Fig. 13.4.
The arrows of the graph are labeled 1, . . . , 7, starting with the arrows whose tail is
Q1 ∈ X2 (Rd ), then those whose tail is Q2 ∈ X3 (Rd ), finally the ones whose tail is
Q3 ∈ X2 (Rd ). It leads to the following formula:
∂ Q̃1 ∂ F1 ∂ 2 F2 ∂ 2 F3 ∂ F4
BΓ (Q1 , Q2 , Q3 )(F1 , . . . , F4 ) = ∑ Q̃2 Q̃3 ,
1γ1 ,...,γ7 d ∂ xγ6 ∂ xγ1 ∂ xγ2 ∂ xγ3 ∂ xγ4 ∂ xγ7 ∂ xγ5
where
Q̃1 := Q1 [xγ1 , xγ2 ], Q̃2 := Q2 [xγ3 , xγ4 , xγ5 ], Q̃3 := Q3 [xγ6 , xγ7 ] .
We have introduced in the previous subsections all the ingredients which are
used in Kontsevich’s explicit construction of the L∞ -quasi-isomorphism Φ between
•
(X• (Rd ), [· , ·]S , 0) and ( HCdiff (Rd ), [· , ·]G , μ ). As above, we denote these pointed
differential graded Lie algebras by g and h respectively, and we denote the natu-
ral coordinates on Rd by x1 , . . . , xd . Recall that Φ : S• g → h is completely deter-
mined by the family of maps (Φk )k∈N , where for each k ∈ N, Φk is the restric-
tion of Φ to Sk g. They are defined as follows: for all p1 , . . . , pk ∈ N and for all
Q1 Q2 . . . Qk ∈ g p1 −1 · · · g pk −1 , the -differential operator Φk (Q1 Q2 . . . Qk ) ∈ h−1 ,
where := ∑ki=1 pi − 2k + 2, is defined by:
In this definition, the sum is taken over all Kontsevich graphs Γ = (Vk, , A) ∈ Gk,
and for each such graph Γ , ϖΓ and BΓ are the weight and the -differential operator
associated to Γ , as defined in Subsections 13.5.2.2 and 13.5.2.3.
A few comments about this definition are in order:
(1) As we have seen, ϖΓ = 0 when Γ is a graph in Gk, whose number of arrows
is different from 2k + − 2. This explain why the sum in (13.79) is only taken
over Gk, , where := ∑ki=1 pi − 2k + 2. Recall also that BΓ is zero when Γ has
a vertex i for which the number of arrows which start from i is not equal to pi .
(2) With respect to the grading which we introduced on S• g, the monomial
Q1 Q2 . . . Qk in (13.79) is of degree ∑ki=1 pi − 2k, while BΓ (Q1 Q2 . . . Qk ) is of
weight − 2, with as above. It follows that Φ is a graded map (of degree 0).
(3) Each Φk is well-defined: the order of Q1 , . . . , Qk in the right-hand side of
(13.79) is irrelevant, since both ϖΓ and BΓ are independent of the ordering
of the vertices of the first type.
(4) The above formula for Φk is particularly simple for k = 1. In fact, there are !
graphs Γ = (V1, , A, σ ) in G1, with arrows, differing only by their ordering
(as given by σ ). As we have seen in Example 13.48, the weight of these graphs
is 1/(!)2 ; according to Example 13.49, BΓ (Q) = Q. It follows that
406 13 Deformation Quantization
1
Φ1 (Q) = ∑ ϖΓ BΓ (Q) =
!
Q.
Γ ∈G1,
Theorem 13.51. Let M be a real manifold and denote by μ the usual product
on C∞ (M).
(1) Every Poisson structure π on M admits a star product, i.e., there exists a star
product μ = μ + ∑i∈N∗ μi ν i on C∞ (M), with μ1 = π2 ;
(2) There is a one-to-one correspondence between the set of equivalence classes
of formal deformations of the trivial Poisson structure on M and the set of
equivalence classes of star products on C∞ (M);
(3) The correspondence in (2) is natural in the following sense: if the equivalence
class of a formal deformation π = ∑i∈N∗ πi ν i of the zero Poisson structure
on M and the equivalence class of a star product μ = μ + ∑i∈N∗ μi ν i on M
correspond under (2), then π1 is the skew-symmetric part of 2μ1 .
Theorems 13.46 and 13.39, combined, show that there is a one-to-one corre-
spondence between the set of gauge equivalence classes of solutions of the Maurer–
Cartan equation associated to the differential graded Lie algebra (X• (M), [· , ·]S , 0)
and the set of gauge equivalence classes of solutions of the Maurer–Cartan equa-
tion associated to ( HC•diff (M), [· , ·]G , μ ). According to Proposition 13.33, the first
set is the set of equivalence classes of deformations of the trivial Poisson structure,
while the second set, after a translation by μ , is the set of equivalence classes of star
products on C∞ (M). This yields the one-to-one correspondence in (2).
Let π = ∑k∈N∗ πk ν k be a formal deformation of the trivial Poisson structure on
M and consider μ := μ + Ω̃Φ (π ), which is a star product on C∞ (M). Recall that
the equivalence class of π and the equivalence class of μ are related by the one-to-
one correspondence in (2). We have, as in (13.80), that μ = μ + π21 ν (mod ν 2 ).
Let μ = μ + ∑k∈N∗ μk ν k be an arbitrary star product on C∞ (M) which is equivalent
to μ , which means that μ − μ and μ − μ are gauge equivalent. Then there exists
ξ = ∑k∈N∗ ξk ν k in ν HC0diff (M)ν such that μ = eadξ (μ ), so that
π
μ = μ +
1
+ [ξ1 , μ ]G ν (mod ν 2 ) .
2
The bilinear map [ξ1 , μ ]G ∈ HC1diff (M) is symmetric, because
for all F, G ∈ C∞ (M). This shows that π1 is the skew-symmetric part of 2μ1 , which
is the content of (3).
If (M, π ) is a real Poisson manifold, one can also consider the formal deformations
of π rather than of the trivial Poisson structure on M. We show in the following
theorem how these deformations are related to star products on C∞ (M)M.
Theorem 13.52. Let (M, π ) be a real Poisson manifold and let μ denote the usual
product on C∞ (M). There is a one-to-one correspondence between the set of equiva-
lence classes of formal deformations of the Poisson structure π and the set of equiv-
alence classes of star products μ = μ + ∑k∈N∗ μk ν k on C∞ (M), for which π is the
skew-symmetric part of 2μ1 .
Proof. Let us denote by D0 the set of equivalence classes of formal deformations
of the trivial Poisson structure on M and by Dπ the set of equivalence classes of
formal deformations of the Poisson structure π on M. Let us also denote by Sμ the
set of equivalence classes of star products on C∞ (M) and finally by Sμ ,π the set of
equivalence classes of star products μ = μ + ∑k∈N∗ μk ν k on C∞ (M), such that the
skew-symmetric part of 2μ1 is π . Of course, one has Sμ ,π ⊆ Sμ and according to
item (2) of Theorem 13.51, there exists a bijection ψ : D0 → Sμ .
An element π = π + ∑k∈N∗ πk ν k ∈ X1 (M)ν is a formal deformation of π if
and only if [π , π ]S = 0, which is also equivalent to [π ν , π ν ]S = 0, meaning that
π ν = πν + ∑k∈N∗ πk ν k+1 is a formal deformation of the trivial Poisson structure.
Moreover, saying that two formal deformations π and π of π are equivalent means
408 13 Deformation Quantization
that π − π and π − π are two gauge equivalent solutions of the Maurer–Cartan
equation associated to the pointed differential graded Lie algebra (X• (M), [· , ·]S , π ),
i.e., that there exists ξ ∈ ν X0 (M)ν satisfying
Since this equation is equivalent to π ν = ν eadξ (π ) = eadξ (π ν ), two formal de-
formations π and π of π are equivalent if and only if π ν and π ν are equivalent
as formal deformations of the trivial Poisson structure. This fact permits us to iden-
tify ν Dπ with a subset of D0 . Now, according to item (3) of Theorem 13.51, we
have ψ (ν Dπ ) ⊆ Sμ ,π . Moreover, if μ = μ + ∑k∈N∗ μk ν k is an element of Sμ ,π ,
one has 2μ1− = π and, once more according to item (3) of Theorem 13.51, there
exists a formal deformation π = ∑k∈N∗ πk ν k of the trivial Poisson structure, whose
equivalence class is sent by ψ to μ and such that π = 2μ1− = π1 . This implies in
particular that ν1 π is a formal deformation of the Poisson structure π and that μ
lies in ψ (ν Dπ ). This shows that ψ (ν Dπ ) = Sμ ,π , so that ψ induces a bijection
between ν Dπ and Sμ ,π , hence between Dπ and Sμ ,π .
νk
μ (F, G) = μ (F, G) + ∑ Φk (π k )(F, G)
k1 k!
νk
= μ (F, G) + ∑ ∑ ϖΓ BΓ (π , . . . , π )(F, G) .
k1 k! Γ ∈Gk,2
In the above sum, only those graphs for which every vertex of the first type is the
tail of exactly two arrows contribute, since BΓ (π , . . . , π ) is equal to zero for all other
graphs.
13.6 Notes
The fact that every Poisson manifold admits a deformation quantization has been
shown by Kontsevich [107]. An alternative proof of Kontsevich’s result was given
by Tamarkin [189].
In a purely algebraic context, the problem of (formally) deforming commuta-
tive associative algebras was considered by Gerstenhaber in [82], who establishes
the link with Hochschild cohomology. Deformations of Poisson structures were in-
troduced in [126]. Deformations of Poisson manifolds, in which both the associa-
tive product and the Poisson bracket are deformed, are considered in Ginzburg–
Kaledin [85].
Appendix A
Multilinear Algebra
In this appendix, we recall some basic definitions and properties related to multi-
linear algebra, including (graded) algebra and coalgebra structures, derivations and
coderivations. We need them in two particular cases, namely for vector spaces over
a field F, and for A -modules, where A is a commutative associative algebra over a
field F. Notice that the latter modules can also be thought of as vector spaces over F,
a fact which we often use, since many operations which we consider are F-linear,
rather than A -linear. In order to cover both cases, we consider in this appendix
the structures which we need on modules over an arbitrary commutative ring R with
unit; the reader may find it useful to keep in mind the example of the C∞ (M)-module
of vector fields or differential forms over a manifold M.
Throughout the appendix, F denotes a field of characteristic zero and R denotes
an arbitrary commutative ring with unit, denoted by 1.
For R-modules V and W , the set of linear1 maps V → W is denoted by HomR (V,W ),
or by Hom(V,W ) when it is clear that V and W are considered as R-modules.
Hom(V,W ) is itself an R-module in a natural way. When V is a free R-module (for
example when R = F, so that V is an F-vector space) and B is a basis of V , every
map B → W extends to a unique element of Hom(V,W ) by linearity; when both V
and W are finite-dimensional F-vector spaces, and bases for V and W have been
fixed, we often think of elements of Hom(V,W ) as matrices (with dimW rows and
dimV columns).
The dual of an R-module V is the R-module V ∗ := Hom(V, R). For a finite-
dimensional F-vector space V , the dual V ∗ is an F-vector space, isomorphic to V ;
the isomorphism is however not canonical, since it depends on the choice of a basis
1 We rarely use the word R-linear, to avoid confusion with the terminology for multilinear maps,
i.e., k-linear maps, with k ∈ N∗ .
· , · : V ∗ ×V → R
(A.1)
(ξ , v) → ξ , v := ξ (v) ,
We will often deal with bilinear maps, sometimes with more general multilinear
maps between R-modules, where multilinear means (R-)linear in each of its argu-
ments, keeping the other arguments fixed. The language of tensor products, which
we introduce now, is very useful for this. Let V1 , V2 ,V and W be R-modules and
let p : V1 ×V2 → V be a bilinear map. If we compose p with a linear map V → W ,
then we obtain a bilinear map V1 × V2 → W ; one may wonder if, given V1 and V2 ,
there exists an R-module V , such that, for every R-module W , every bilinear map
V1 × V2 → W can be obtained in this way from a linear map on V → W . In fact,
there is a unique (up to isomorphism) such R-module, called the tensor product of
V1 and V2 , denoted by V1 ⊗R V2 , or V1 ⊗V2 , and it comes with a natural bilinear map
p : V1 × V2 → V1 ⊗ V2 . Formally, the stated property means that every bilinear map
φ : V1 × V2 → W factors uniquely via p, meaning that there exists a unique linear
map φ , such that φ ◦ p = φ . This property is displayed in the following diagram:
p
V1 ×V2 V1 ⊗V2
φ φ
In practice we do not make a distinction between the bilinear map φ and the linear
map φ , so we simply write φ ∈ Hom(V1 ⊗ V2 ,W ). A natural construction of the
tensor product V1 ⊗V2 is as the quotient of the free R-module which is generated by
all formal expressions v1 ⊗ v2 , with v1 ∈ V1 and v2 ∈ V2 , divided by the equivalence
relation defined by
(v1 + v1 ) ⊗ v2 = v1 ⊗ v2 + v1 ⊗ v2 ,
v1 ⊗ (v2 + v2 ) = v1 ⊗ v2 + v1 ⊗ v2 ,
a(v1 ⊗ v2 ) = (av1 ) ⊗ v2 = v1 ⊗ (av2 ),
where v1 , v1 ∈ V1 and v2 , v2 ∈ V2 and a ∈ R. One says that v1 ⊗ v2 is the ten-
sor product of v1 and v2 . The maps p and φ are in this notation simply given by
A.1 Tensor Algebra 413
so that, in the sequel, we will not make a notational distinction between the elements
in (A.2). The extension to several R-modules is clear. For a given R-module V , we
obtain a natural sequence of R-modules T kV := V ⊗k , where k = 0, 1, 2, . . . , which is
defined, as the notation suggests, by V ⊗k := V ⊗ · · · ⊗V (k factors), when k 1 and
V ⊗0 := R. Equipped with the product (X,Y ) → X ⊗Y , where X ∈ V ⊗k and Y ∈ V ⊗ ,
the R-module
∞
∞
T •V := T kV = V ⊗k
k=0 k=0
becomes a graded R-algebra, called the tensor algebra of V . See Section A.3 below
for the basic definitions on graded algebras.
Fixing one R-module Z, it is useful to think of taking the tensor product with Z
as a functor, which means on the one hand that a linear map φ ∈ Hom(V,W ) yields,
in a natural way, a linear map
1V ⊗ 1Z = 1V ⊗Z , (φ ◦ ψ ) ⊗ 1Z = (φ ⊗ 1Z ) ◦ (ψ ⊗ 1Z ) ,
where the latter product is the multiplication in A . When this multiplication has
extra properties, they yield similar properties for the product of maps: for example,
if A is commutative, then
φ1 ⊗ φ2 = (φ2 ⊗ φ1 ) ◦ S ,
(φ1 ⊗ φ2 ) ⊗ φ3 = φ1 ⊗ (φ2 ⊗ φ3 ) ,
where ∧kV := T kV /Nk , for k ∈ N. Notice that ∧kV = {0} as soon as k is bigger
than the (minimal) number of generators of V . We denote the quotient maps by
p : T •V → ∧•V and pk : T kV → ∧kV . Since N is a two-sided ideal of T •V , we have
that the associative product ⊗ on T •V induces an associative product on ∧•V , which
is denoted by ∧. Thus p(v1 ⊗ · · · ⊗ vk ) = p(v1 ) ∧ · · · ∧ p(vk ), which we also write as
v1 ∧· · ·∧vk , because p1 (i.e., the restriction of p to V ) is injective. One easily verifies
that, if X ∈ ∧iV and Y ∈ ∧ jV , then X ∧Y ∈ ∧i+ jV and
X ∧Y = (−1)i jY ∧ X . (A.6)
In the language of the next section, this property is called graded commutativity.
The associative, graded commutative R-algebra (∧•V, ∧) is called the exterior alge-
bra of V and elements of ∧•V are called multivectors. One similarly constructs the
symmetric algebra (S•V, ·) which is the associative commutative graded R-algebra
obtained as T •V /N , where N is the two-sided ideal of T •V , generated by all
v ⊗ w − w ⊗ v, where v, w ∈ V .
It is clear that every skew-symmetric k-linear map φ ∈ Hom(V ⊗k ,W ) vanishes
on Nk . Therefore, for R-modules V and W , we have that every skew-symmetric
k-linear map φ ∈ Hom(V ⊗k ,W ) corresponds in a canonical way to a linear map
φ : ∧kV → W , as in the following commutative diagram:
pk
V ⊗k ∧k V
φ φ
ρk− : ∧k V → T kV
1
k! σ∑
v1 ∧ · · · ∧ vk → sgn(σ )vσ (1) ⊗ · · · ⊗ vσ (k) . (A.7)
∈S k
It is easy to see that this map is well-defined and injective, and that its image consists
precisely of all skew-symmetric k-tensors. One similarly identifies the symmetric k-
tensors with SkV by using the symmetrization map
ρk+ : SkV → T kV ,
whose definition is formally the same as the above definition (A.7) of the skew-
symmetrization map, except that one leaves out the factor sgn(σ ).
There are two types of internal products related to the exterior algebra. Let V
and W be arbitrary R-modules. For X ∈ ∧ jV , the internal product ıX yields, for
every i ∈ N, a linear map
which is given by
ıX φ (Z) := φ (X ∧ Z)
where φ : ∧iV → W and Z ∈ ∧i− jV , assuming i j; otherwise, i.e., when i < j, then
ıX φ := 0. It is easily verified that
ıX∧Y = ıY ◦ ıX ,
ıφ : ∧ jV → ∧ j−iV ,
when i j, and ıφ is the zero map otherwise. In this formula, Si,k denotes the set
of all (i, k)-shuffles, i.e., all permutations σ ∈ Si+k for which σ (1) < · · · < σ (i) and
σ (i + 1) < · · · < σ (i + k); sgn(σ ) is the signature of σ as a permutation.
A.2 Exterior and Symmetric Algebra 417
ıφ ∧ψ = ıψ ◦ ıφ ,
ıφ ∧ψ (v1 , . . . , vk )
= ∑ sgn(σ )(φ ∧ ψ )(vσ (1) , . . . , vσ (i+ j) ) vσ (i+ j+1) ∧ · · · ∧ vσ (k)
σ ∈Si+ j,k−i− j
= (ıψ ◦ ıφ )(v1 , . . . , vk ) ,
where we have used the notation Si, j,k−i− j for the set of all permutations of {1, . . . , k}
that are increasing on the intervals 1, . . . , i and i + 1, . . . , i + j and i + j + 1, . . . , k.
We return to the general case for which A is not necessarily R, but an arbitrary
commutative associative R-algebra. We point out that, if φ ∈ Hom(∧iV, A ), ψ ∈
Hom(∧ jV, A ) and χ ∈ Hom(∧kV, A ), where i, j, k 0, then
(φ ∧ ψ ) ∧ χ = φ ∧ (ψ ∧ χ ) ,
part (i.e., on the last vs that appear out of the arguments of φ and ψ ). Moreover, for
φ ∈ Hom(∧iV, A ) and ψ ∈ Hom(∧ jV, A ), we have
φ ∧ ψ = (−1)i j ψ ∧ φ .
In the language of the next section, the product ∧ makes ⊕k∈N Hom(∧kV, A ) into a
graded R-algebra which is associative and graded commutative. Notice that, in the
notation which we use, if φ1 , . . . , φk ∈ Hom(V, A ) and v1 , . . . , vk ∈ V , then
φ1 ∧ · · · ∧ φk , v1 ∧ · · · ∧ vk = det φi , v j 1i, jk . (A.9)
In this section we recall the basic definitions of algebras and graded algebras. These
definitions will be dualized in the next section, to obtain the notions of a coalgebra
and of a graded coalgebra.
Let V be an R-module. An algebra structure on V is a bilinear map μ : V ×V → V ,
called a product. We also view μ as an element of Hom(V ⊗V,V ), and we say that
(V, μ ) is an R-algebra. Usually, μ is assumed to have additional properties; the typ-
ical extra properties that μ may be supposed to have are summarized in Table A.1.
Table A.1 A product μ on an R-module, which makes it into an R-algebra, is usually assumed to
have one or two additional properties, taken from the list which appears in this table. We write the
properties in their usual form (with u, v, w ∈ V ) and in their functional form; the latter is useful for
obtaining the “co”-version (see Section A.4). S is the twist map u ⊗ v → v ⊗ u and S is the cycle
map u ⊗ v ⊗ w → v ⊗ w ⊗ u.
may have are commutativity and associativity; some authors call (V, μ ) in this case
simply an algebra (or R-algebra), but we will not use this convention here since
our modules will usually have two algebra structures, one of which is a Lie algebra
structure, and the other one is associative and commutative.
Example A.1. A simple example of a Lie algebra structure is given by the vector
space Hom(W,W ), where W is an arbitrary vector space, equipped with the commu-
tator
[φ1 , φ2 ] := φ1 ◦ φ2 − φ2 ◦ φ1 ,
where φ1 , φ2 ∈ Hom(W,W ). The Jacobi identity for [· , ·] is a direct consequence of
the associativity of the composition of (linear) maps.
φ⊗φ φ (A.10)
μ
W ⊗W W
In the case of Lie algebras (V, [· , ·]) and (W, [· , ·] ), such a linear map φ is called
a Lie algebra homomorphism. The homomorphism property then takes the form
φ ([v1 , v2 ]) = [φ (v1 ), φ (v2 )] for all v1 , v2 ∈ V .
We now turn to the graded version of these definitions. For this, it is assumed
that V is a graded R-module,
V= Vi , (A.11)
i∈Z
where each of the subspaces Vi is invariant under the action of R. The notation V•
(or V • , when the subspaces are indexed by superscripts) is also used for V . In many
cases, one has V = ⊕i∈NVi , or even V = ⊕ki=0Vi , the other Vi being undefined, but
one easily arrives at the form (A.11) by defining Vi := {0}, for those values of i
where Vi was undefined. An element of Vi is called a homogeneous element of V of
degree i. A linear map φ : V → W between graded R-modules is said to be graded
of degree r if φ (Vi ) ⊂ Wi+r for every i ∈ Z. When V and W are written as V• and W• ,
the suggestive notation φ : V• → W•+r is also used. We denote the R-module of all
graded linear maps from V to W of degree r by Homr (V,W ). A graded product on V
is a product μ on V such that
Then V , equipped with μ , becomes a graded R-algebra and we have induced maps
μi, j : Vi ⊗V j → Vi+ j for all i, j. A graded linear map of degree zero φ : V → W which
420 A Multilinear Algebra
Table A.2 The graded analog of Table A.1 is displayed. The only difference between the graded
and ungraded notions lies in the signs; in fact, as there are no signs in the case of graded associa-
tivity, the notion of associativity and graded associativity coincide. For the graded Jacobi identity,
it is understood that when one sums over the three cyclic permutations of (i, j, k), the exponent
takes the consecutive values 0, 1 and 2. The elements u, v and w are assumed to be homogeneous
of respective degrees i, j and k.
Example A.2. To give a simple example of a graded Lie algebra, we define for
graded linear maps φi ∈ Homri (V,V ), where i = 1, 2, their graded commutator
[φ1 , φ2 ] as the graded linear map of degree r1 + r2 , given by
The graded R-module ⊕r∈Z Homr (V,V ), equipped with this bracket, is a graded
Lie algebra. For graded linear maps φ1 , φ2 , φ3 of degree r1 , r2 , r3 , the graded Jacobi
identity takes the following form
(−1)r1 r3 [φ1 , [φ2 , φ3 ]] + (−1)r2 r1 [φ2 , [φ3 , φ1 ]] + (−1)r3 r2 [φ3 , [φ1 , φ2 ]] = 0 , (A.13)
In the language of Section A.5, this means that [φ1 , ·], which is a graded linear map
of degree r1 , is a graded derivation of [· , ·].
A.3 Algebras and Graded Algebras 421
We have in this appendix already met the following three examples of graded
algebras, associated to an R-module V .
(T •V, ⊗) is a graded R-algebra, which is associative;
(∧•V, ∧) is a graded R-algebra, which is associative and graded commutative;
(S•V, ·) is a graded R-algebra, which is associative and commutative.
The grading on T •V comes from the natural decomposition T •V = ⊕i∈NV ⊗i ;
for ∧•V and S•V , the induced grading is used. The commutativity of the graded
algebra (S•V, ·) should not be confused with the graded commutativity of the
graded algebra (∧•V, ∧): in the former the arguments commute, but in the latter
they only commute up to a sign, see (A.6). We think of T • as a functor: given
a linear map φ ∈ Hom(V,W ), we obtain a homomorphism of graded algebras
T • φ : T •V → T •W, whose restriction to T kV is the linear map T kV → T kW , de-
fined by T k φ := φ ⊗ φ ⊗ · · · ⊗ φ , i.e.,
T • (1V ) = 1T •V , T • (ψ ◦ φ ) = T • (ψ ) ◦ T • (φ ) .
We now consider the “co”-versions of the concepts which were introduced in the
previous section. This is done in the customary way, namely we obtain the “co”-
versions by dualizing the definitions of the previous section, written in their func-
tional forms (see Tables A.1 and A.2), which is done by reversing all arrows and
switching the order of their composition. Let us first dualize the definition of an al-
gebra: a coalgebra structure on an R-module V is an element Δ of Hom(V,V ⊗V ),
called a coproduct. The most important additional properties which Δ may have, are
summarized in Table A.3.
Table A.3 Dualizing the usual properties of the product of an algebra, written in functional form,
as in Table A.1, we obtain the usual properties which the coproduct Δ , defining a coalgebra struc-
ture on an R-module, can have. As before, S is the twist map and S is the cycle map.
φ⊗φ φ
W ⊗W W
Δ
We now consider the graded versions of the above “co”-concepts. Let V be a graded
R-module, V = ⊕i∈ZVi . A graded coproduct is a coproduct Δ on V such that for
every k ∈ Z,
Δ (Vk ) ⊂ Vi ⊗V j ,
i+ j=k
is finitely supported, i.e., Δ (Vk ) has a non-trivial intersection with only a finite
number of Vi ⊗ V j . Notice that Δ (Vk ) is automatically finitely supported when
Vi = {0} for all i < 0. A graded R-module V , equipped with a graded coprod-
uct Δ , is called a graded R-coalgebra. For fixed i, j ∈ Z, we will need the linear map
Δi, j : Vi+ j → Vi ⊗V j , which is obtained by composing Δ , restricted to Vi+ j , with the
A.4 Coalgebras and Graded Coalgebras 423
Table A.4 This last table deals with the case of graded coalgebras. We recall that S is the twist
map and S is the cycle map (see Table A.2).
We have seen in the previous section that the graded R-modules T •V, ∧•V
and S•V have a natural (graded) algebra structure. We now show that they also have
a graded coalgebra structure; the latter structure is important at a few places in this
book. The coalgebra structure Δ on T •V is called de-concatenation and is defined,
for k ∈ N and for v1 , . . . , vk ∈ V , by
k
Δ (v1 ⊗ · · · ⊗ vk ) := ∑ (v1 ⊗ · · · ⊗ vi ) ⊗ (vi+1 ⊗ · · · ⊗ vk ) , (A.14)
i=0
Δi, j : V ⊗(i+ j) → V ⊗i ⊗V ⊗ j
(A.15)
v1 ⊗ · · · ⊗ vi+ j → (v1 ⊗ · · · ⊗ vi ) ⊗ (vi+1 ⊗ · · · ⊗ vi+ j ) .
We now turn to the natural coalgebra structure of the exterior algebra ∧•V . Let us
denote by δ the diagonal map V → V × V : v → (v, v). By functoriality of ∧• , it
induces a linear map
∧• δ : ∧•V → ∧• (V ×V ) ,
which we view as a linear map
δ δ ×1V
V ×V V ×V ×V
1V ×δ
∧• δ ∧• (δ ×1V )
∧• (V ×V ) ∧• (V ×V ×V )
∧• (1V ×δ )
Δ Δ ⊗1∧•V
The graded algebra structure on ∧•V ⊗ ∧•V ⊗ ∧•V is defined as in (A.16). An ex-
plicit formula for Δ is given by
Δi, j (v1 ∧ · · · ∧ vi+ j ) = ∑ sgn(σ )(vσ (1) ∧ · · · ∧ vσ (i) ) ⊗ (vσ (i+1) ∧ · · · ∧ vσ (i+ j) ) ,
σ ∈Si, j
(A.17)
when i, j ∈ N and Δi, j = 0 when i < 0 or j < 0. The above formula is similar to the
de-concatenation formula (A.14) which we have introduced in the case of the tensor
algebra.
for all v ∈ Vp and w ∈ Vq , where vw is a shorthand for μ (v, w). In functional notation
this means that for every p, q ∈ Z,
φ̃ : ∧•V → ∧•+rV
k
v1 ∧ · · · ∧ vk → ∑ (−1)i−1 φ (vi ) ∧ v1 ∧ · · · ∧ vi ∧ · · · ∧ vk .
i=1
426 A Multilinear Algebra
φ̃ (X ∧Y ) = φ̃ (X) ∧Y + (−1) pr X ∧ φ̃ (Y ),
which is an easy consequence of the fact that φ will either be applied to a factor
which appears in X, leaving Y untouched, or vice versa.
We now formulate the notion of a derivation for the case of a graded coalgebra.
As before, (V, Δ ) is a graded coalgebra, where V = ⊕i∈ZVi is a graded R-module.
Dualizing (A.18), a linear map φ : V → V of degree −r is called a graded coderiva-
tion of degree r if for every p, q ∈ Z,
as maps from Vp+r+q to Vp ⊗Vq . We denote the R-module of all graded coderivations
of degree r of V by CoDerr (V ). Again, it follows by direct computation that the
graded commutator of two graded coderivations of degrees r1 and r2 is a graded
coderivation of degree r1 + r2 . This implies that ⊕r∈Z CoDerr (V ) is also a graded
Lie algebra, with the graded commutator as Lie bracket. Like ⊕r∈Z Derr (V ), it is
a Lie subalgebra of the graded Lie algebra ⊕r∈Z Homr (V,V ), equipped with the
graded commutator. The following example is the “co”-version of Example A.3.
Example A.4. Let V be an R-module and consider the graded coalgebra ∧•V .
Equipped with the graded commutator, the R-module ⊕r∈Z CoDerr (∧•V ) is a graded
Lie algebra. Particular elements of CoDerr (∧•V ) can be constructed from linear
maps ∧r+1V → V : every linear map
φ : ∧r+1V → V
φ̃ : ∧ •V → ∧•−rV
v1 ∧ · · · ∧ vk → ∑ sgn(τ ) φ (vτ (1) , . . . , vτ (r+1) ) ∧ vτ (r+2) ∧ · · · ∧ vτ (k) .
τ ∈Sr+1,k−1−r
All coderivations of ∧•V are obtained in this way: if Φ : ∧•V → ∧•−rV is a coderiva-
tion of degree r of (∧•V, Δ ), then Φ = φ̃ , where φ : ∧r+1V → V is the restriction of
Φ to ∧r+1V . Indeed, since Φ and φ̃ agree on ∧kV , for k r + 1 (they are both zero
when k r, for degree reasons), they also agree, in view of (A.19), on ∧r+2V (take
p = q = 1 in (A.19)), and similarly for the higher exterior powers of V .
Appendix B
Real and Complex Differential Geometry
In this appendix we recall the basic notions of differential geometry: the definition
of a real manifold, of a complex manifold and of a vector field on such a manifold.
We also recall briefly the main properties of vector fields on manifolds: the existence
of integral curves of a vector field, the flow of a vector field, the bracket of vector
fields and the straightening theorem, which says that a vector field takes, in well-
chosen coordinates, a simple form. Our definition of vector fields on a manifold is
based on the concept of a pointwise derivation. This approach easily generalizes to
the introduction of the concept of a bivector field on a manifold, a crucial element
in the (geometrical!) definition of the notion of a Poisson structure on a (real or
complex) manifold (see Section 1.3).
y ◦ x−1 x(U∩V )
: x(U ∩V ) → y(U ∩V )
is a smooth map, see Fig. B.1. This homeomorphism is called a transition map. A
collection of compatible coordinate charts of M, whose domains cover M, is called
an atlas of M. For connected differentiable manifolds, the integer d is independent
of the coordinate chart; it is called the dimension of M, a terminology which we also
use in the non-connected case, when d is independent of the coordinate chart; it is
Fig. B.1 A manifold comes equipped with an atlas, a collection of coordinate charts (U, x), where
the transition maps y ◦ x−1 between coordinate charts (U, x) and (V, y) (with U ∩ V = 0)
/ are de-
manded to be smooth.
denoted by dim M. When the integer d is even, we may interpret the homeomor-
phisms x as taking values in Cd/2 ; in this case, if all transition maps are complex
analytic (holomorphic), then M is called a complex manifold and d/2 is called the
(complex) dimension of M. It is a trivial, but important, fact that every non-empty
open subset of a real or complex manifold is itself, in a natural way, a real or com-
plex manifold. In particular, every open subset of a (real or complex) vector space
is a (real or complex) manifold.
The main virtue of manifolds is that we can do calculus on them, hence also
analytic geometry. Roughly speaking, the coordinate charts allow us to identify ob-
jects on the manifold, locally, with standard objects on open subsets of Fd (F = R
or F = C) and the transition maps allow us to compare these standard objects. The
first object one thinks of is that of a smooth function: a function F : M → F on a
real (respectively complex) manifold M is called a smooth function (respectively a
holomorphic function) if for every coordinate chart (U, x) of M the function
F̃ = F ◦ x−1 : x(U) → F
Fig. B.2 Smooth functions on a manifold M are functions on M which are smooth in terms of local
coordinates.
One similarly defines the notion of a smooth map between real manifolds and a
holomorphic map between complex manifolds. It leads to two categories: the cat-
egory of real manifolds, whose objects are real manifolds with smooth maps as
morphisms, and the category of complex manifolds, whose objects are complex
manifolds and whose morphisms are holomorphic maps. Since complex manifolds
are in a natural way also real manifolds, and since holomorphic maps are smooth,
there is a natural forgetful functor from the latter category to the former.
Let M be a manifold and let m ∈ M be an arbitrary point. We define the tangent space
Tm M of M at m. To do this, we consider the set Fm (M) of all pairs (F,U), where U is
an open subset of M which contains m, and F is an element of F (U). Two elements
(F,U) and (G,V ) of Fm (M) are defined to be equivalent, denoted (F,U) ∼ (G,V ), if
there exists a pair (H,W ) ∈ Fm (M), such that W ⊂ U ∩V and H = F|W = G|W . For
(F,U) ∈ Fm (M), we denote its equivalence class by Fm and we call it the germ of F
(or of (F,U)) at m. It is clear that the quotient set Fm (M)/ ∼ of all function germs
at m inherits from F (M) the structure of an associative F-algebra. For example,
in the complex case, the algebra Fm (M)/ ∼ is isomorphic to the algebra of power
series in d variables, whose radius of convergence is positive. Notice that a function
germ Fm has a well-defined value at m, which is simply F(m).
Definition B.1. Let M be a manifold and let m ∈ M. A pointwise derivation δm of
F (M) at m is a linear function
430 B Real and Complex Differential Geometry
Fm (M)
δm : →F,
∼
satisfying, for all functions F and G, defined on a neighborhood of m in M,
dm F : Tm M → F
(B.2)
δm → δm Fm .
see Fig. B.3. This is well-defined, because the germ (G ◦ Ψ )m is independent of the
function G which represents the germ GΨ (m) . When F is a function, defined on a
neighborhood of m, the tangent map at m is a linear map Tm F : Tm M → TF(m) F,
and we can recover the differential dm F, upon composing Tm F with the canonical
isomorphism TF(m) F F, which will be explained below.
The tangent map obeys the usual rules of calculus: for example, if M, N and P
are manifolds, Ψ : M → N and Ξ : N → P are maps and m ∈ M, then
Fig. B.3 A map Ψ between two manifolds M and N leads for every point m ∈ M to a tangent map
TmΨ , which is a linear map between the tangent space Tm M and TΨ (m) N.
d
Fm = F(m) + ∑ ((xi )m − xi (m))Fm .
(i)
(B.4)
i=1
This shows that Tm Fd is spanned by the pointwise derivations (ei )m , so that the map,
defined by v → vm , is an isomorphism between Fd and Tm Fd .
If (U, x) is a coordinate chart of a manifold M, centered at m, then each of
the d vectors of the natural basis (e1 , . . . , ed ) of Fd leads to a pointwise derivation
of F (M) at m, defined by
∂
:= To x−1 (ei )o .
∂ xi m
∂
Fm = ∂i F̃(x(m)) ,
∂ xi m
for all δm ∈ Tm M. Thus, in our setup, we view tangent vectors as objects which act
on equivalence classes of functions, rather than viewing functions as objects which
define linear forms on equivalence classes of curves, although both points of view
are equivalent (see [198, Ch. 1]).
V [F](m) := Vm Fm ∈ F . (B.6)
We also write Vm [F] for Vm Fm , so that Vm [F] = V [F](m). We say that V is a smooth
vector field (respectively holomorphic vector field) on M if for every open subset
U ⊂ M and for every function F ∈ F (U), the function V [F], defined by (B.6),
belongs to F (U) (i.e., it is a smooth, respectively holomorphic function on U).
When the type of manifold which is considered is irrelevant or is clear from the
context, we simply say vector field for smooth or holomorphic vector field. Notice
that we use square brackets to denote the action of a vector field on a function.
With respect to pointwise multiplication, the vector fields on M form an F (M)-
module, which is denoted by X1 (M). Viewed as a vector space, X1 (M) is a Lie
algebra, where the Lie bracket is the commutator of vector fields, defined as follows.
Let V and W be vector fields on M, and let m ∈ M. For every function F, defined
in a neighborhood of m, letting
It is clear that vector fields can be restricted to open subsets; we usually do not
make a notational distinction between a vector field on M and its restriction to some
open subset of M. It is also clear from (B.6) that V defines, for every open subset U
of M, a derivation of F (U), i.e., we have
Remark B.4. It is shown in standard books on differential geometry that for a real
manifold M, the above natural correspondence between (smooth) vector fields on M
and derivations of F (M) is bijective. For complex manifolds however, this is not
true in general: think of a compact complex torus Cd /Z2d , which has non-trivial
holomorphic vector fields, but whose algebra of holomorphic functions consists of
constant functions only, so that all its derivations are trivial. The same phenomenon
occurs for skew-symmetric biderivations and bivector fields (e.g., Poisson struc-
tures), introduced in Chapter 1.
Remark B.5. The set of all tangent vectors at m, for m ranging through M, has a
natural vector bundle structure over M, denoted T M → M. The fiber over m is the
vector space Tm M and the vector fields on M can be defined as the (smooth, holo-
morphic) sections of T M → M. In abstract geometrical constructions, it is the latter
point of view on vector fields which is often the most appropriate.
∂
[F](m) := ∂i F̃(x(m)) , (B.7)
∂ xi
∂
pointwise derivations ∂ xi form a basis of Tm M, for every m ∈ U. Then V is
m
a (smooth) vector field on U if and only if the coefficients V (i) in this expression
belong to F (U). It is clear that these coefficients V (i) are given by V (i) = V [xi ].
By a slight abuse of language, we often refer to the expression
d
∂
V = ∑ V [xi ] (B.8)
i=1 ∂ xi
as a coordinate expression of V in the coordinate chart (U, x). It is very useful for
explicit computations, as it allows us to compute with vector fields on a manifold,
locally, in the same way as on Fd .
B.4 The Flow of a Vector Field 435
Fig. B.4 For a given vector field on a manifold, there passes through every point of the manifold a
unique integral curve.
φ1 | = φ2 | .
(I1 ∩I2 )0 (I1 ∩I2 )0
The map φ or the pair (I, φ ) is called an integral curve of V , passing through m, see
Fig. B.4. By a slight abuse of notation, the left-hand side in (B.9) is often denoted
by ddtφ (t ); using this notation (B.9) takes the more familiar form
dφ
(t ) = Vφ (t ) .
dt
The integral curves of a vector field depend smoothly on the initial data; this is stated
in a precise way in the following theorem (see [187, Ch. 5] for a proof).
Theorem B.6. Let V be a vector field on a manifold M and let m ∈ M. There exists
a neighborhood U of (0, m) in F × M and there exists a map Φ : U → M such that,
for every (0, m ) ∈ U, the restriction
Φ |I : Im → M , (B.10)
m
for every m ∈ U. When it is clear from the context, we often simply say distribution
for smooth (or holomorphic) distribution. A vector field V , defined on an open
subset U of M, is said to be adapted to D on U if Vm ∈ D(m) for every m ∈ U. A
distribution D on M is said to be involutive if for every open subset U of M, and for
every pair of vector fields on U, which are adapted to D on U, their Lie bracket is
also adapted to D on U. Involutivity of a distribution is a very strong condition, as
is plain from Frobenius’ theorem.
Theorem B.8 (Frobenius’ theorem). Let M be a d-dimensional manifold and sup-
pose that D is a k-dimensional distribution on M. If D is involutive, then every point
m ∈ M admits a coordinate chart (U, x), with m ∈ U, such that
∂ ∂
D(m ) = span ,..., , for every m ∈ U .
∂ x1 m ∂ xk m
B.5 The Frobenius Theorem 437
For a short and elementary proof, which is immediately adapted to the holomorphic
case, we refer to [41] or [138]. It is clear that Frobenius’ theorem can be seen as a
generalization of the straightening theorem (Theorem B.7).
References
40. V. Chari and A. Pressley. A guide to quantum groups. Cambridge University Press, Cam-
bridge, 1995. Corrected reprint of the 1994 original.
41. S. Chern and J. Wolfson. A simple proof of Frobenius theorem. In Manifolds and Lie groups
(Notre Dame, IN, 1980), volume 14 of Progr. Math., pages 67–69. Birkhäuser, Boston,
MA, 1981.
42. C. Chevalley. Invariants of finite groups generated by reflections. Amer. J. Math., 77:778–
782, 1955.
43. N. Chriss and V. Ginzburg. Representation theory and complex geometry. Birkhäuser Boston
Inc., Boston, MA, 1997.
44. D. Collingwood and W. McGovern. Nilpotent orbits in semisimple Lie algebras. Van Nos-
trand Reinhold Mathematics Series. Van Nostrand Reinhold Co., New York, 1993.
45. J. Conn. Normal forms for analytic Poisson structures. Ann. of Math. (2), 119(3):577–601,
1984.
46. J. Conn. Normal forms for smooth Poisson structures. Ann. of Math. (2), 121(3):565–593,
1985.
47. A. Coste, P. Dazord, and A. Weinstein. Groupoı̈des symplectiques. In Publications du
Département de Mathématiques. Nouvelle Série. A, Vol. 2, volume 87 of Publ. Dép. Math.
Nouvelle Sér. A, pages i–ii, 1–62. Univ. Claude-Bernard, Lyon, 1987.
48. Th. Courant. Dirac manifolds. Trans. Amer. Math. Soc., 319(2):631–661, 1990.
49. M. Crainic and R. Fernandes. Integrability of Lie brackets. Ann. of Math. (2), 157(2):575–
620, 2003.
50. M. Crainic and R. Fernandes. Integrability of Poisson brackets. J. Differential Geom.,
66(1):71–137, 2004.
51. M. Crainic and R. Fernandes. Stability of symplectic leaves. Invent. Math., 180(3):481–533,
2010.
52. M. Crainic and R. Fernandes. A geometric approach to Conn’s linearization theorem. Ann.
of Math. (2), 173(2):1121–1139, 2011.
53. M. Crainic and R. Fernandes. Lectures on integrability of Lie brackets. In Lectures on
Poisson geometry, volume 17 of Geom. Topol. Monogr., pages 1–107. Geom. Topol. Publ.,
Coventry, 2011.
54. P. Damianou and R. Fernandes. From the Toda lattice to the Volterra lattice and back. Rep.
Math. Phys., 50(3):361–378, 2002.
55. P. Damianou and R. Fernandes. Integrable hierarchies and the modular class. Ann. Inst.
Fourier (Grenoble), 58(1):107–137, 2008.
56. P. Damianou, H. Sabourin, and P. Vanhaecke. Transverse Poisson structures to adjoint orbits
in semisimple Lie algebras. Pacific J. Math., 232(1):111–138, 2007.
57. M. De Wilde and P. Lecomte. Existence of star-products and of formal deformations
of the Poisson Lie algebra of arbitrary symplectic manifolds. Lett. Math. Phys., 7(6):487–
496, 1983.
58. V. Drinfel’d. Hamiltonian structures on Lie groups, Lie bialgebras and the geometric meaning
of classical Yang-Baxter equations. Dokl. Akad. Nauk SSSR, 268(2):285–287, 1983.
59. V. Drinfel’d. Quantum groups. In Proceedings of the International Congress of Mathemati-
cians, Vol. 1, 2 (Berkeley, CA, 1986), pages 798–820, Providence, RI, 1987. Amer. Math.
Soc.
60. J.-P. Dufour and A. Haraki. Rotationnels et structures de Poisson quadratiques. C. R. Acad.
Sci. Paris Sér. I Math., 312(1):137–140, 1991.
61. J.-P. Dufour and M. Zhitomirskii. Classification of nonresonant Poisson structures. J. London
Math. Soc. (2), 60(3):935–950, 1999.
62. J.-P. Dufour and M. Zhitomirskii. Singularities and bifurcations of 3-dimensional Poisson
structures. Israel J. Math., 121:199–220, 2001.
63. J.-P. Dufour and N. T. Zung. Poisson structures and their normal forms, volume 242 of
Progress in Mathematics. Birkhäuser Verlag, Basel, 2005.
64. J.-P. Dufour and N.T. Zung. Linearization of Nambu structures. Compositio Math.,
117(1):77–98, 1999.
442 References
65. J. Duistermaat. On global action-angle coordinates. Comm. Pure Appl. Math., 33(6):687–
706, 1980.
66. J. Duistermaat and J. Kolk. Lie groups. Universitext. Springer-Verlag, Berlin, 2000.
67. Ch. Ehresmann. Les connexions infinitésimales dans un espace fibré différentiable. In
Colloque de topologie (espaces fibrés), Bruxelles, 1950, pages 29–55. Georges Thone,
Liège, 1951.
68. P. Etingof and V. Ginzburg. Noncommutative del Pezzo surfaces and Calabi-Yau algebras.
J. Eur. Math. Soc. (JEMS), 12(6):1371–1416, 2010.
69. P. Etingof and A. Varchenko. Geometry and classification of solutions of the classical dy-
namical Yang-Baxter equation. Comm. Math. Phys., 192(1):77–120, 1998.
70. P. Etingof and A. Varchenko. Solutions of the quantum dynamical Yang-Baxter equation and
dynamical quantum groups. Comm. Math. Phys., 196(3):591–640, 1998.
71. S. Evens, J.-H. Lu, and A. Weinstein. Transverse measures, the modular class and a coho-
mology pairing for Lie algebroids. Quart. J. Math. Oxford Ser. (2), 50(200):417–436, 1999.
72. L. Faddeev and L. Takhtajan. Hamiltonian methods in the theory of solitons. Classics in
Mathematics. Springer, Berlin, English edition, 2007. Translated from the 1986 Russian orig-
inal by A. Reyman.
73. B. Fedosov. A simple geometrical construction of deformation quantization. J. Differential
Geom., 40(2):213–238, 1994.
74. R. Fernandes. Connections in Poisson geometry. I. Holonomy and invariants. J. Differential
Geom., 54(2):303–365, 2000.
75. R. Fernandes. Lie algebroids, holonomy and characteristic classes. Adv. Math., 170(1):119–
179, 2002.
76. R. Fernandes and Ph. Monnier. Linearization of Poisson brackets. Lett. Math. Phys., 69:89–
114, 2004.
77. V. V. Fock and A. A. Rosly. Poisson structure on moduli of flat connections on Riemann
surfaces and the r-matrix. In Moscow Seminar in Mathematical Physics, volume 191 of Amer.
Math. Soc. Transl. Ser. 2, pages 67–86. Amer. Math. Soc., Providence, RI, 1999.
78. A. Fomenko. Integrability and nonintegrability in geometry and mechanics, volume 31 of
Mathematics and its Applications (Soviet Series). Kluwer Academic Publishers Group, Dor-
drecht, 1988. Translated from the Russian by M. V. Tsaplina.
79. J.-P. Françoise, G. L. Naber, and T. S. Tsun, editors. Encyclopedia of mathematical physics.
Vol. 1, 2, 3, 4, 5. Academic Press/Elsevier Science, Oxford, 2006.
80. W. Fulton and J. Harris. Representation theory, volume 129 of Graduate Texts in Mathemat-
ics. Springer-Verlag, New York, 1991. A first course, Readings in Mathematics.
81. I. Gel’fand and I. Dorfman. Hamiltonian operators and the classical Yang-Baxter equation.
Funktsional. Anal. i Prilozhen., 16(4):1–9, 96, 1982.
82. M. Gerstenhaber. On the deformation of rings and algebras. Ann. of Math. (2), 79:59–103,
1964.
83. V. Ginzburg. Momentum mappings and Poisson cohomology. Internat. J. Math., 7(3):329–
358, 1996.
84. V. Ginzburg. Coisotropic intersections. Duke Math. J., 140(1):111–163, 2007.
85. V. Ginzburg and D. Kaledin. Poisson deformations of symplectic quotient singularities. Adv.
Math., 186(1):1–57, 2004.
86. V. Ginzburg and A. Weinstein. Lie-Poisson structure on some Poisson Lie groups. J. Amer.
Math. Soc., 5(2):445–453, 1992.
87. J. Grabowski, G. Marmo, and P. W. Michor. Construction of completely integrable systems
by Poisson mappings. Modern Phys. Lett. A, 14(30):2109–2118, 1999.
88. J. Grabowski, G. Marmo, and A. M. Perelomov. Poisson structures: towards a classification.
Modern Phys. Lett. A, 8(18):1719–1733, 1993.
89. M. Gualtieri. Generalized complex geometry. PhD. Thesis, Oxford, 2004.
90. L. Guieu and C. Roger. L’algèbre et le groupe de Virasoro. Les Publications CRM, Montreal,
QC, 2007. Aspects géométriques et algébriques, généralisations. [Geometric and algebraic
aspects, generalizations], With an appendix by V. Sergiescu.
References 443
91. V. Guillemin. Moment maps and combinatorial invariants of Hamiltonian T n -spaces, vol-
ume 122 of Progress in Mathematics. Birkhäuser Boston Inc., Boston, MA, 1994.
92. V. Guillemin and S. Sternberg. Symplectic techniques in physics. Cambridge University
Press, Cambridge, second edition, 1990.
93. R. Hartshorne. Algebraic geometry. Springer-Verlag, New York, 1977. Graduate Texts in
Mathematics, No. 52.
94. N. Hitchin. Instantons, Poisson structures and generalized Kähler geometry. Comm. Math.
Phys., 265(1):131–164, 2006.
95. G. Hochschild and J.-P. Serre. Cohomology of Lie algebras. Ann. of Math. (2), 57:591–603,
1953.
96. J. Huebschmann. Poisson cohomology and quantization. J. Reine Angew. Math., 408:57–113,
1990.
97. J. Huebschmann. Poisson structures on certain moduli spaces for bundles on a surface. Ann.
Inst. Fourier (Grenoble), 45(1):65–91, 1995.
98. J. Huebschmann. Symplectic and Poisson structures of certain moduli spaces. I. Duke Math.
J., 80(3):737–756, 1995.
99. J. Humphreys. Introduction to Lie algebras and representation theory, volume 9 of Graduate
Texts in Mathematics. Springer-Verlag, New York, 1978. Second printing, revised.
100. C. Jacobi. Sur le mouvement d’un point et sur un cas particulier du problème des trois corps.
Compt. Rend., 3:59–61, 1836.
101. C. Jacobi. Note von der geodätischen linie auf einem ellipsoid und den verschiedenen an-
wendungen einer merkwürdigen analytischen substitution. J. Reine Angew. Math., 19:309–
313, 1839.
102. N. Jacobson. Lie algebras. Dover Publications Inc., New York, 1979. Republication of the
1962 original.
103. Ch. Kassel. Quantum groups, volume 155 of Graduate Texts in Mathematics. Springer-
Verlag, New York, 1995.
104. S. Khoroshkin, A. Radul, and V. Rubtsov. A family of Poisson structures on Hermitian sym-
metric spaces. Comm. Math. Phys., 152(2):299–315, 1993.
105. A. A. Kirillov. Unitary representations of nilpotent Lie groups. Uspehi Mat. Nauk,
17(4(106)):57–110, 1962.
106. A. A. Kirillov. Local Lie algebras. Uspehi Mat. Nauk, 31(4(190)):57–76, 1976.
107. M. Kontsevich. Deformation quantization of Poisson manifolds. Lett. Math. Phys.,
66(3):157–216, 2003 (q-alg/9709040, 1997).
108. Y. Kosmann-Schwarzbach. Quantum and classical Yang-Baxter equations. Modern Phys.
Lett. A, 5(13):981–990, 1990.
109. Y. Kosmann-Schwarzbach. From Poisson algebras to Gerstenhaber algebras. Ann. Inst.
Fourier (Grenoble), 46(5):1243–1274, 1996.
110. Y. Kosmann-Schwarzbach. Lie bialgebras, Poisson Lie groups and dressing transformations.
In Integrability of nonlinear systems (Pondicherry, 1996), volume 495 of Lecture Notes in
Phys., pages 104–170. Springer, Berlin, 1997.
111. Y. Kosmann-Schwarzbach. Odd and even Poisson brackets. In Quantum theory and symme-
tries (Goslar, 1999), pages 565–571. World Sci. Publ., River Edge, NJ, 2000.
112. Y. Kosmann-Schwarzbach. Quasi, twisted, and all that. . .in Poisson geometry and Lie al-
gebroid theory. In The breadth of symplectic and Poisson geometry, volume 232 of Progr.
Math., pages 363–389. Birkhäuser Boston, Boston, MA, 2005.
113. Y. Kosmann-Schwarzbach. Poisson manifolds, Lie algebroids, modular classes: a survey.
SIGMA Symmetry Integrability Geom. Methods Appl., 4:Paper 005, 30, 2008.
114. Y. Kosmann-Schwarzbach. The Noether theorems. Sources and Studies in the History of
Mathematics and Physical Sciences. Springer, New York, 2011. Invariance and conserva-
tion laws in the twentieth century, Translated, revised and augmented from the 2006 French
edition by Bertram E. Schwarzbach.
115. Y. Kosmann-Schwarzbach and F. Magri. Poisson-Nijenhuis structures. Ann. Inst. H. Poincaré
Phys. Théor., 53(1):35–81, 1990.
444 References
142. S. Majid. Foundations of quantum group theory. Cambridge University Press, Cambridge,
1995.
143. Ch.-M. Marle. The Schouten-Nijenhuis bracket and interior products. J. Geom. Phys., 23(3–
4):350–359, 1997.
144. J. Marsden and T. Ratiu. Reduction of Poisson manifolds. Lett. Math. Phys., 11(2):161–169,
1986.
145. J. Marsden and T. Ratiu. Introduction to mechanics and symmetry, volume 17 of Texts in
Applied Mathematics. Springer-Verlag, New York, second edition, 1999. A basic exposition
of classical mechanical systems.
146. J. Marsden and A. Weinstein. Reduction of symplectic manifolds with symmetry. Rep. Math-
ematical Phys., 5(1):121–130, 1974.
147. J. McKay. Graphs, singularities, and finite groups. In The Santa Cruz Conference on Finite
Groups (Univ. California, Santa Cruz, CA, 1979), volume 37 of Proc. Sympos. Pure Math.,
pages 183–186. Amer. Math. Soc., Providence, RI, 1980.
148. Kenneth R. Meyer. Symmetries and integrals in mechanics. In Dynamical systems (Proc.
Sympos., Univ. Bahia, Salvador, 1971), pages 259–272. Academic Press, New York, 1973.
149. E. Miranda and N.T. Zung. A note on equivariant normal forms of Poisson structures. Math.
Res. Lett., 13(5-6):1001–1012, 2006.
150. Ph. Monnier. Une cohomologie associée à une fonction. Applications aux cohomologies de
Poisson et de Nambu-poisson. PhD. Thesis, Université de Montpellier, 2001.
151. Ph. Monnier. Formal Poisson cohomology of quadratic Poisson structures. Lett. Math. Phys.,
59(3):253–267, 2002.
152. Ph. Monnier. Poisson cohomology in dimension two. Israel J. Math., 129:189–207, 2002.
153. Ph. Monnier and N. T. Zung. Levi decomposition for smooth Poisson structures. J. Differen-
tial Geom., 68(2):347–395, 2004.
154. D. Mumford, J. Fogarty, and F. Kirwan. Geometric invariant theory, volume 34 of Ergebnisse
der Mathematik und ihrer Grenzgebiete (2) [Results in Mathematics and Related Areas (2)].
Springer-Verlag, Berlin, third edition, 1994.
155. J. Nestruev. Smooth manifolds and observables, volume 220 of Graduate Texts in Mathe-
matics. Springer-Verlag, New York, 2003. Joint work of A. M. Astashov, A. B. Bocharov,
S. V. Duzhin, A. B. Sossinsky, A. M. Vinogradov and M. M. Vinogradov, Translated from
the 2000 Russian edition by Sossinsky, I. S. Krasil’schik and Duzhin.
156. A. Nijenhuis and R. Richardson. Deformations of Lie algebra structures. J. Math. Mech.,
17:89–105, 1967.
157. A. Odesskiı̆ and V. Rubtsov. Polynomial Poisson algebras with a regular structure of sym-
plectic leaves. Teoret. Mat. Fiz., 133(1):1321–1337, 2002.
158. P. Olver. Applications of Lie groups to differential equations, volume 107 of Graduate Texts
in Mathematics. Springer-Verlag, New York, second edition, 1993.
159. J.-P. Ortega and T. Ratiu. The optimal momentum map. In Geometry, mechanics, and dy-
namics, pages 329–362. Springer, New York, 2002.
160. J.-P. Ortega and T. Ratiu. Momentum maps and Hamiltonian reduction, volume 222 of
Progress in Mathematics. Birkhäuser Boston Inc., Boston, MA, 2004.
161. V. Ovsienko and K. Rozhe. Deformations of Poisson brackets and extensions of Lie algebras
of contact vector fields. Russian Math. Surveys, 47(6):135–191, 1992.
162. J.-S. Park. Topological open p-branes. In Symplectic geometry and mirror symmetry (Seoul,
2000), pages 311–384. World Sci. Publ., River Edge, NJ, 2001.
163. S. Pelap. Poisson (co)homology of polynomial Poisson algebras in dimension four:
Sklyanin’s case. J. Algebra, 322(4):1151–1169, 2009.
164. A. M. Perelomov. Integrable systems of classical mechanics and Lie algebras. Vol. I.
Birkhäuser Verlag, Basel, 1990. Translated from the Russian by A. G. Reyman.
165. D. Perrin. Algebraic geometry. Universitext. Springer-Verlag London Ltd., London, 2008.
An introduction, Translated from the 1995 French original by C. Maclean.
166. A. Pichereau. Poisson (co)homology and isolated singularities. J. Algebra, 299(2):747–777,
2006.
446 References
167. A. Pichereau. Formal deformations of Poisson structures in low dimensions. Pacific J. Math.,
239(1):105–133, 2009.
168. A. Pichereau and G. Van de Weyer. Double Poisson cohomology of path algebras of quivers.
J. Algebra, 319(5):2166–2208, 2008.
169. S. Poisson. Mémoire sur la variation des constantes arbitraires dans les questions de
mécanique. J. Ecole Polytec., 8:266–344, 1809.
170. A. Polishchuk. Algebraic geometry of Poisson brackets. J. Math. Sci. (New York),
84(5):1413–1444, 1997. Algebraic geometry, 7.
171. J. Pradines. Théorie de Lie pour les groupoı̈des différentiables. Calcul différentiel dans
la catégorie des groupoı̈des infinitésimaux. C. R. Acad. Sci. Paris Sér. A-B, 264:A245–
A248, 1967.
172. B. Przybylski. Complex Poisson manifolds. In Differential geometry and its applications
(Opava, 1992), volume 1 of Math. Publ., pages 227–241. Silesian Univ. Opava, Opava, 1993.
173. O. Radko. A classification of topologically stable Poisson structures on a compact oriented
surface. J. Symplectic Geom., 1(3):523–542, 2002.
174. C. Roger and P. Vanhaecke. Poisson cohomology of the affine plane. J. Algebra, 251(1):448–
460, 2002.
175. V. Roubtsov and T. Skrypnyk. Compatible Poisson brackets, quadratic Poisson algebras
and classical r-matrices. In Differential equations: geometry, symmetries and integrability,
volume 5 of Abel Symp., pages 311–333. Springer, Berlin, 2009.
176. P. Saksida. Lattices of Neumann oscillators and Maxwell-Bloch equations. Nonlinearity,
19(3):747–768, 2006.
177. J. A. Schouten. Ueber Differentialkomitanten zweier kontravarianter Grössen. Nederl. Akad.
Wetensch., Proc., 43:449–452, 1940.
178. J. A. Schouten. On the differential operators of first order in tensor calculus. Rapport ZA
1953-012. Math. Centrum Amsterdam, 1953.
179. M. Semenov-Tian-Shansky. What a classical r-matrix is. Functional Anal. Appl., 17(4):259–
272, 1983.
180. M. Semenov-Tian-Shansky. Dressing transformations and Poisson group actions. Publ. Res.
Inst. Math. Sci., 21(6):1237–1260, 1985.
181. J.-P. Serre. Local fields, volume 67 of Graduate Texts in Mathematics. Springer-Verlag, New
York, 1979. Translated from the French by M. J. Greenberg.
182. I. Shafarevich. Basic algebraic geometry. 1. Springer-Verlag, Berlin, second edition, 1994.
Varieties in projective space, Translated from the 1988 Russian edition and with notes by
M. Reid.
183. P. Slodowy. Simple singularities and simple algebraic groups, volume 815 of Lecture Notes
in Mathematics. Springer, Berlin, 1980.
184. J. Śniatycki and W. M. Tulczyjew. Generating forms of Lagrangian submanifolds. Indiana
Univ. Math. J., 22:267–275, 1972/73.
185. J.-M. Souriau. Quantification géométrique. Comm. Math. Phys., 1:374–398, 1966.
186. M. Spivak. Calculus on manifolds. A modern approach to classical theorems of advanced
calculus. W. A. Benjamin, Inc., New York-Amsterdam, 1965.
187. M. Spivak. A comprehensive introduction to differential geometry. Vol. I. Publish or Perish
Inc., Wilmington, DE, second edition, 1979.
188. T. A. Springer. Invariant theory. Lecture Notes in Mathematics, Vol. 585. Springer-Verlag,
Berlin, 1977.
189. D. E. Tamarkin. Operadic proof of M. Kontsevich’s formality theorem. ProQuest LLC, Ann
Arbor, MI, 1999. Thesis (Ph.D.)–The Pennsylvania State University.
190. P. Tauvel and R. Yu. Lie algebras and algebraic groups. Springer Monographs in Mathemat-
ics. Springer-Verlag, Berlin, 2005.
191. A. Thimm. Integrable geodesic flows on homogeneous spaces. Ergodic Theory Dynamical
Systems, 1(4):495–517 (1982), 1981.
References 447
gauge holomorphic
equivalence, 380 distribution, 436
equivalent, 378 function, 428
generalized Lie derivative, 85 map, 429
germ, 429 vector field, 433
germification, 59 homogeneous
Gerstenhaber element, 419
algebra, 84 function, 207
bracket, 357 monomial, 372
graded Poisson structure, 207
algebra, 419, 420 homology
coalgebra, 422 Lie algebra, 98
coderivation, 426 Poisson, 99
commutator, 420 homomorphism
coproduct, 422 algebra, 419
derivation, 425 coalgebra, 422
Lie algebra, 82, 420 Lie algebra, 419
Lie bracket, 420 Lie bialgebra, 305
linear map, 419 Poisson–Lie group, 292
module, 419
product, 419 ideal, 6
gradient, 183 Lie, 6
graph Poisson, 6
Kontsevich, 400 Poisson–Dirac, 134
underlying, 400 identity
weight, 400 fundamental, 222
group Jacobi, 4
algebraic, 116 immersed submanifold, 49
cocycle, 294 immersion, 49
cohomology, 294 independent functions, 330
dressing, 322 inner derivation, 94, 359
Lie, 114 integrable distribution, 29
Poisson–Lie, 292 integrable system
representation, 116 Liouville, 341
integral curve, 435
Hadamard lemma, 431 internal product, 76, 416
Hamiltonian, 6 intertwining, 271
action, 153 invariant
derivation, 6 bilinear form, 121
function, 21 density, 105
local, 21 element, 93
path, 27 function, 121
piecewise, 27 left, 115
vector, 22 manifold, 341
field, 6, 21, 223 right, 115
head, 400 submanifold, 149
Hermitian metric, 177 subset, 149
hierarchy subvariety, 149
bi-Hamiltonian, 334 involution, 34
Hochschild functions in, 26, 330
coboundary, 358 involutive
cocycle, 358 distribution, 436
cohomology, 358 functions, 330
complex, 358
Index 453
singular symmetric
distribution, 29 algebra, 371, 415
locus, 14, 22 tensor, 416
point, 14, 22 symmetrization, 416
singularity symplectic
Du Val, 265 connection, 366
Klein, 265 foliation, 26, 27, 29
simple, 241, 243 leaf, 29
skew-symmetric manifold, 168
biderivation, 5 almost, 168
tensor, 416 exact, 177
skew-symmetrization, 416 map, 171
smooth structure, 167, 168
distribution, 436 almost, 168
function, 428 vector space, 167
map, 429 symplectic connection, 366
point, 14, 55
vector field, 433 tail, 400
space tangent, 48, 49
cotangent, 14, 16, 430 bundle, 19
representation, 92 map, 16, 430
tangent, 14, 16, 49, 430 space, 14, 16, 49, 430
splitting tensor
coordinates, 24 algebra, 413
Lie algebra, 272, 338 product, 39, 412
theorem, 23 skew-symmetric, 416
star product, 362 symmetric, 416
equivalent, 362 theorem
straightening theorem, 436 action-angle, 347
structure Adler–Kostant–Symes, 337
constants, 182 Arnold, 241
function, 10 Conn’s linearization, 201
Nambu–Poisson, 222 Darboux, 26
Poisson, 8, 19 Frobenius, 436
symplectic, 167, 168 Jacobson–Morozov, 227
subalgebra, 6 Kontsevich’s formality, 399
Lie, 6 Lie, 115
Poisson, 6 Liouville, 342
subgroup Noether, 333
Lie, 114 Poisson, 34, 332
Poisson–Lie, 296 splitting, 23
submanifold, 49 straightening, 436
coisotropic, 129 Weinstein, 23
embedded, 27, 49 Thimm’s method, 335
immersed, 27, 49 total differential, 375
Poisson, 49 transition map, 427
Poisson–Dirac, 137 translation
submersion, 127 left, 117
subspace right, 117
isotropic, 281 transverse
subvariety coordinates, 347
affine, 47 Poisson structure, 147
Poisson, 10, 48 submanifolds, 143
Poisson–Dirac, 135 triple
Index 457
F, 3 Pm [F, G], 17
{· , ·}, 4 T M, 19
{F, G}, 4 Rkm π , 22
, 4 Rk π , 22
X1 (A ), 5 LV , 23
[V , W ], 5 o, 25
P[F, G], 5 1r , 26
V [F], 5 Sm (M), 27
X2 (A ), 5 F (V ), 32
XH , 6 X1 (M), 32
Ham(A , {· , ·}), 6 X2 (M), 32
Ham(A ), 6 J , 34
Cas(A , {· , ·}), 7 A1 ⊗ A2 , 39
Cas(A ), 7 N sm , 55
F (M), 8 [· , ·], 56
Ψ ∗, 9 [F, G], 56
dF, 10 P[F1 , . . . , Fp ], 64
Matd (A ), 10 X p (A ), 64
Rkm {· , ·}, 14 X• (A ), 64
Rk {· , ·}, 14 S p,q , 65
Tm M, 14 sgn, 65
Tm∗ M, 14 P ∧ Q, 65
πm , 14 ıF , 66
M(s) , 15 LV , 66
dim, 16 P[F1 , . . . , Fp ], 67
F (M), 16 Ω 1 (A ), 69
Tm M, 16 dF, 69
Fm , 16 Ω p (A ), 70
Tm∗ M, 16 Ω • (A ), 70
· , ·, 16 ω ∧ η , 71
ξ , v, 16 HdR p
(A ), 73
TmΨ , 16 •
HdR (A ), 73
dm F, 16 [ω ], 73
X1 (M), 16 Ω p (M), 74
Vm , 16 Ω • (M), 74
V [F], 16 d, 74
p
Vm [F], 17 HdR (M), 75
Sk , 415 δ , 423
sgn(σ ), 416 , 424
ıX , 416 Derr (V ), 425
ıφ , 416 CoDerr (V ), 426
Si, j , 416 F (M), 428
φ ∧ ψ , 417 Fm , 429
Si, j,k , 417 Tm M, 430
S , 418
Tm∗ M, 430
V • , 419
· , ·, 430
V• , 419
V• → W•+r , 419 ξ , v, 430
Homr (V,W ), 419 dm F, 430
μi, j , 419 TmΨ , 430
[· , ·], 420 V [F], 433
T • φ , 421 Vm [F], 433
Δ , 422 X1 (M), 433
Δi, j , 422 Ψ∗ V , 433