Normal Forms and Lie Groupoid Theory
Normal Forms and Lie Groupoid Theory
pr
=
V // U
One can also assume that there is some extra geometric structure be-
having well with respect to the submersion, and then ask if one can achieve
linearization of both the submersion and the extra geometric structure.
For example, if one assumes that 2 (M ) is a closed 2-form such that the
pullback of to each fiber is non-degenerate, then one can show that is a
locally trivial symplectic fibration (see, e.g., [12]). We will come back to this
later, for now we recall another basic linearization theorem:
Theorem 2 (Reeb). Let F be a foliation of M and let L0 be a compact
leaf of F with finite holonomy. Then there exists a saturated neighborhood
L0 U M , a hol(L0 )-invariant neighborhood 0 V x0 (L0 ), and a
diffeomorphism:
h
=
f0
L V // U M
hol(L0 )
Here Kx0 acts on the normal space x0 (Ox0 ) via the normal isotropy
representation. If the action is locally free then the orbits form a foliation, the
isotropy groups Kx are finite and hol(Ox ) is a quotient of Kx . Moreover, the
action of Kx on a slice descends to the linear holonomy action of hol(Ox ). The
slice theorem is then a special case of the Reeb stability theorem. However, in
general, the isotropy groups can have positive dimension and the two results
look apparently quite different.
Again, both in the case of foliations and in the case of group actions, we
could consider extra geometric structures (e.g., a metric or a symplectic form)
and ask for linearization taking into account this extra geometric structure.
One can find such linearization theorems in the literature (e.g., the local
normal form theorem for Hamiltonian actions [10]). Let us mention one such
recent result from Poisson geometry, due to Crainic and Marcut [4]:
Theorem 4 (Local normal form around symplectic leaves). Let (M, ) be a
Poisson manifold and let S M be a compact symplectic leaf. If the Poisson
homotopy bundle G y P S is a smooth compact manifold with vanishing
second de Rham cohomology group, then there is a neighborhood S U M ,
and a Poisson diffeomorphism:
: (U, |U ) (P G g, lin ).
We will not discuss here the various terms appearing in the statement
of this theorem, referring the reader to the original work [4]. However, it
should be clear that this result has the same flavor as the previous ones: some
compactness type assumption around a leaf/orbit leads to linearization or a
normal form of the geometric structure in a neighborhood of the leaf/orbit.
Although all these results have the same flavor, they do look quite dif-
ferent. Moreover, the proofs that one can find in the literature of these lin-
earization results are also very distinct. So it may come as a surprise that
they are actually just special cases of a very general linearization theorem.
In order to relate all these linearization theorems, and to understand the
significance of the assumptions one can find in their statements, one needs
a language where all these results fit into the same geometric setup. This
language exists and it is a generalization of the usual Lie theory from groups
to groupoids. We will recall it in the next lecture. After that, we will be in
shape to state the general linearization theorem and explain how the results
stated before are special instances of it.
Let us spell out this definition. We have a set of objects M , and a set
of arrows G. For each arrow g G we can associate its source s(g) and its
target t(g), resulting in two maps s, t : G M . We also write g : x y
for an arrow with source x and target y.
For any pair of composable arrows we have a product or composition
map:
m : G(2) G, (g, h) 7 gh.
In general, we will denote by G(n) the set of n strings of composable arrows:
G(n) := {(g1 , . . . , gn ) : s(gi ) = t(gi+1 )}.
The multiplication satisfies the associativity property:
(gh)k = g(hk), (g, h, k) G(3) .
For each object x M there is an identity arrow 1x and the identity
property holds:
1t(g) g = g = g1s(g) , g G.
It gives rise to an identity section u : M G, x 7 1x .
For each arrow g G there is an inverse arrow g 1 G, for which the
inverse property holds:
gg 1 = 1t(g) , g 1 g = 1s(g) , g G.
This gives rise to the inverse map : G G, g 7 g 1 .
Definition 2. A morphism of groupoids is a functor F : G H.
This means that we have a map F : G H between the sets of arrows
and a map f : M N between the sets of objects, making the following
diagram commute:
F //
G H
s t s t
M // N
f
such that F (gh) = F (g)F (h) if g, h G are composable, and F (1x ) = 1f (x)
for all x M .
We are interested in groupoids and morphisms of groupoids in the
smooth category:
Definition 3. A Lie groupoid is a groupoid G M whose spaces of arrows
and objects are both manifolds, the structure maps s, t, u, m, i are all smooth
maps and such that s and t are submersions. A morphisms of Lie groupoids
is a morphism of groupoids for which the underlying map F : G H is
smooth.
Before we give some examples of Lie groupoids, let us list a few basic
properties.
the unit map u : M G is an embedding and the inverse : G G is
a diffeomorphism.
Normal Forms and Lie Groupoid Theory 5
0 // T S // T M // (S) // 0
describe the normal bundle (L) as the associated bundle L e h hol(L,x) x (L).
The underlying foliation of this linear model is still the linear foliation of
(L). However, the linear model now depends on the germ of the foliation
around L, i.e., on the non-linear holonomy. Unlike the case of the fundamen-
tal groupoid, the knowledge of the Bott connection is not enough to build
this linear model.
Example 8 (Group actions). Let K be a Lie group that acts on a manifold
M and let Kx be the isotropy group of some x M . For each k Kx , the
map k : M M , y 7 ky, fixes x and maps the orbit Ox to itself. Hence,
dx k induces a linear action of Kx on the normal space x (Ox ), called the
normal isotropy representation. Using this representation, it is not hard to
check that we have a vector bundle isomorphism:
// K Kx x (Ox )
(Ox )
'' uu
Ox
where the action Kx y K x (Ox ) is given by k(g, v) := (gk 1 , kv). More-
over, one has an action of K on (Ox ), which under this isomorphism corre-
sponds to the action:
K y K Kx x (Ox ), k[(k , v)] = [(kk , v)].
Now consider the action Lie groupoid K M M . One checks that
the local linear model around the orbit Ox is just the action Lie groupoid
K (Ox ) (Ox ), which under the isomorphism above corresponds to the
action groupoid:
K (K Kx x (Ox )) K Kx x (Ox ).
As we have already mentioned above, the Linearization Theorem states
that, under appropriate conditions, the groupoid is locally isomorphic around
a saturated submanifold to its local model. In order to make precise the
expression locally isomorphic we introduce the following definition:
Definition 6. Let G M be a Lie groupoid and S M a saturated sub-
manifold. A groupoid neighborhood of GS S is a pair of open sets U S
and Ue GS such that U e U is a subgroupoid of G M . A groupoid
e e = GU .
neighborhood U U is said to be full if U
Our first version of the linearization theorem reads as follows:
Theorem 7 (Weak linearization). Let G M be a Lie groupoid with a 2-
metric (2) . Then G is weakly linearizable around any saturated submanifold
S M : there are groupoid neighborhoods U e U of GS S in G M and
e
V V of GS S in the local model (GS ) (S), and an isomorphism
of Lie groupoids:
(U e U)
= (Ve V ),
which is the identity on GS .
Normal Forms and Lie Groupoid Theory 11
we recover the slice theorem for actions of compact groups. For a general
proper action, the results above only give weak linearization, which does not
allow to deduce the slice theorem. However, due to the particular structure of
the action groupoid, every orbit has a saturated neighborhood and one has a
uniform bound for the injectivity radius of the 2-metric. This gives invariant
linearization and leads to the Slice Theorem for any proper action.
Example 12. Let (M, ) be a Poisson manifold. The cotangent bundle T M
has a natural Lie algebroid structure with anchor : T M T M given by
contraction by and Lie bracket on sections (i.e., 1-forms):
[, ] := L() L() d(, ).
In general, this Lie algebroid fails to be integrable. However, under the as-
sumptions of the local normal form theorem stated in Lecture 1, this groupoid
is integrable, and then its source 1-connected integration is an s-proper Lie
groupoid whose orbits are the symplectic leaves. This groupoid can then be
linearized around a symplectic leaf, but this linearization does not yet yield
the local canonical form for the Poisson structure.
It turns out that the source 1-connected integration is a symplectic Lie
groupoid, i.e., there is a symplectic structure on its space of arrows which is
compatible with multiplication. One can apply a Moser type trick to further
bring the symplectic structure on the local normal form to a canonical form,
which then yields the canonical form of the Poisson structure. The details of
this approach, which differ from the original proof of the canonical form due
to Crainic and Marcut, can be found in [4].
where, for each n N, the face maps i : G(n) G(n1) , i = 0, . . . , n and the
degeneracy maps i : G(n) G(n+1) , i = 1, . . . , n, are defined as follows: for
a n-string of composable arrows the i-th face map associates the (n1)-string
of composable arrows obtained by omitting the i-object:
g1 gn
! g1 gi gi+1 gn
xx xx xx yy xx
i =
while the i-th degeneracy map inserts into a n-string of composable arrows
an identity at the i-th entry:
1xi
g1 gn
! g1 gn
xx xx xx xx
i =
For a Lie groupoid there is, additionally, a natural action of Sn+1 on G(n) : for
a string of n-composable arrows (g1 , . . . , gn ) choose (n+1) arrows (h0 , . . . , hn ),
all with the same source, so that:
h0 hn
h1 h2 hn1
xxyy g1 xx~~ g2
(( xx gn
Then the Sn+1 -action on the arrows (h0 , . . . , hn ) by permutation gives a well
defined Sn+1 -action on G(n) . Notice that this action permutes the face maps
i , since there are maps i : Sn+1 Sn such that:
i = i () (i) .
Definition 7. A n-metric (n N) on a groupoid G M is a Riemannian
metric (n) on G(n) which is Sn+1 -invariant and for which all the face maps
i : G(n) G(n1) are Riemmanian submersions.
Actually, it is enough in this definition to ask that one of the face maps
is a Riemannian submersion: the assumption that the action of Sn+1 is by
isometries implies that if one face map is a Riemannian submersion then all
the face maps are Riemannian submersions.
For any n 1, the metrics induced on G(n1) by the different face maps
i : G(n) G(n1) coincide, giving a well defined metric (n1) , which is a
(n 1)-metric. Obviously, one can repeat this process, so that a n-metric on
G(n) determines a k-metric on G(k) for all 0 k n.
Example 13 (0-metrics). When n = 0 we adopt the convention that (0) is a
metric on M = G(0) which makes the orbit space a Riemannian foliation and
is invariant under the action of G on the normal space to the orbits, i.e., such
that each arrow Tg acts by isometries on (O). Such 0-metrics were studied
by Pflaum, Posthuma and Tang in [15], in the case of proper groupoids. One
can think of such metrics as determining a metric on the (possibly singular)
14 Rui Loja Fernandes
orbit space M/G. Indeed, it is proved in [15] that the distance on the orbit
space determined by (0) enjoys some nice properties.
Example 14 (1-metrics). A 1-metric is just a metric (1) on the space of arrows
G = G(1) for which the source and target maps are Riemannian submersions
and inversion is an isometry. These metrics were studied by Gallego et al.
in [9]. A 1-metric induces a 0-metric (0) on M = G(0) , for which the orbit
foliation is Riemannian.
Example 15 (2-metrics). A 2-metric is a metric (2) in the space of com-
posable arrows G(2) , which is invariant under the S3 -action generated by the
involution (g1 , g2 ) 7 (g21 , g11 ) and the 3-cycle (g1 , g2 ) 7 ((g1 g2 )1 , g1 ), and
for which multiplication is a Riemannian submersion. Hence, a 2-metric on
G(2) induces a 1-metric on G(1) . These metrics were introduced in [6].
Notice that the first 3 stages of the nerve
1
// s //
G(2) m //
// G // M
2 t
completely determine the remaining G(n) , for n 3. Hence, one should expect
that n-metrics, for n 3, are determined by their 2-metrics. In fact, one has
the following properties, whose proof can be found in [7, 8]:
there is at most one 3-metric inducing a given 2-metric and every 3-
metric has a unique extension to an n-metric for every n 3.
there are examples of groupoids which admit an n-metric, but do not
admit a n + 1-metric, for n = 0, 1, 2.
uniqueness fails in low degrees: one can have, e.g., two different 2-metrics
on G(2) inducing the same 1-metric on G(1) .
The geometric realization of the nerve of a groupoid G M is usu-
ally denoted by BG, can be seen as the classifying space of principal G-
bundles (see [11]). Two Morita equivalent groupoids G1 M1 and G2 M2
(see [5] or [13]), give rise to homotopy equivalent spaces BG1 and BG2 . An
alternative point of view, is to think of BG as a geometric stack with atlas
G M , and two atlas represent the same stack if they are Morita equivalent
groupoids. The following result shows that one may think of a n-metric as a
metric in BG:
Proposition 10 ( [7, 8]). If G M and H N are Morita equivalent
groupoids, then G admits a n-metric if and only if H admits a n-metric.
In fact, it is shown in [7, 8] that it is possible to transport a n-metric
via a Morita equivalence. This constructions depends on some choices, but
the transversal component of the n-metric is preserved. We refer to those
references for a proof and more details.
Since a n-metric determines a k-metric, for all 0 k n, a necessary
condition for a groupoid G M to admit a n-metric is that the orbit foliation
can be made Riemannian. This places already some strong restrictions on the
Normal Forms and Lie Groupoid Theory 15
class of groupoids admitting a n-metric. So one may wonder when can one
find such metrics. One important result in this direction is that any proper
groupoid admits such a metric:
Theorem 11. A proper Lie groupoid G M admits n-metrics for all n 0.
Proof. The proof uses the following trick, called in [6] the gauge trick. For
each n, consider the manifold G[n] Gn of n-tuples of arrows with the same
source, and the map
(n) : G[n+1] G(n) , (h0 , h1 . . . , hn ) 7 (h0 h11 , h1 h1 1
2 , . . . , hn1 hn ).
The fibers of (n) coincide with the orbits of the right-multiplication action,
f f h0
[n+1]
G xG o
o oo
(h0 , . . . , hn ) k = (h0 k, . . . , hn k). rrrr k
xxrr
hn
and this action is free and proper, hence defining a principal G-bundle. The
strategy is to define a metric on G[n+1] in a such way that (n) becomes a
Riemannian submersion, and that the resulting metric on G(n) is a n-metric.
The group Sn+1 acts on the manifold G[n+1] by permuting its coordi-
nates, and this action covers the action Sn+1 y G(n) , so the map (n) is
Sn+1 equivariant. On the other hand, there are (n + 1) left groupoid actions
G y G[n+1] , each consisting in left multiplication on a given coordinate.
oo ff h0 ff h0 ff h0
k
o o o o
k o o ... o o
rrr rrr rr
r
xxrrr xxrrr k xxrrr
hn hn oo hn
These left actions commute with the above right action and cover (n + 1)-
principal actions G y G(n) , with projection the face maps i : G(n)
G(n1) .
Now for a proper groupoid one can use averaging to construct a metric
on G which is invariant under the left action of G on itself by left translations.
The product metric on G[n+1] is invariant both under the (n+1) left G-actions
above and the Sn+1 -action. In general, it will not be invariant under the right
G-action G y G[n+1] G(n) , but we can average it to obtain a new metric
which is invariant under all actions. It follows that the resulting metric on
G(n) is an n-metric.
Remark 12. The maps (n) : G[n+1] G(n) , n = 0, 1, . . . that appear in
this proof form the simplicial model for the universal principal G-bundle
: EG BG. This bundle plays a key role in many different constructions
associated with the groupoid (see [7]).
Still, there are many examples of groupoids, which are not necessarily
proper, but admit a n-metric. Some are given in the next set of examples.
16 Rui Loja Fernandes
exp(1)
Ve // G
exp(0)
V // M
It follows that the exponential maps of (1) and (0) give the desired weak
linearization, i.e., a groupoid isomorphism
exp
(Ve V )
= (exp(Ve ) exp(V )).
Normal Forms and Lie Groupoid Theory 17
Acknowledgments
I would specially like to thank Matias del Hoyo, for much of the work reported
in these notes rests upon our ongoing collaboration [68]. He and my two PhD
students, Daan Michiels and Joel Villatoro, made several very usual com-
ments on a preliminary version of these notes, that helped improving them.
I would also like to thank Marius Crainic, Ioan Marcut, David Martinez-
Torres and Ivan Struchiner, for many fruitful discussions which helped shape
my view on the linearization problem. Finally, thanks go also to the organiz-
ers of the XXXIII WGMP, for the invitation to deliver these lectures and for
a wonderful stay in Bialowieza.
References
[1] M. Crainic and R.L. Fernandes; Lectures on integrability of Lie brackets; Geom.
Topol. Monogr. 17 (2011), 1-107.
[2] M. Crainic and R. L. Fernandes; Integrability of Lie brackets; Ann. of Math. (2)
157 (2003), 575-620.
[3] M. Crainic and I. Struchiner; On the Linearization Theorem for proper Lie
groupoids; Ann. Scient. Ec. Norm. Sup. 4e serie, 46 (2013), 723-746.
[4] M. Crainic and I. Marcut; A Normal Form Theorem around Symplectic Leaves;
J. Differential Geom. 92 (2012), no. 3, 417-461.
[5] M. del Hoyo; Lie groupoids and their orbispaces; Portugal. Math. 70 (2013),
161-209.
[6] M. del Hoyo and R.L. Fernandes; Riemannian metrics on Lie groupoids;
preprint arXiv:1404.5989.
[7] M. del Hoyo and R.L. Fernandes; Simplicial metrics on Lie groupoids; Work
in progress.
[8] M. del Hoyo and R.L. Fernandes; A stacky version of Ehresmanns Theorem;
Work in progress.
[9] E. Gallego, L. Gualandri, G. Hector and A. Reventos; Groupodes Rieman-
niens; Publ. Mat. 33 (1989), no. 3, 417422.
[10] V. Guillemin and S. Sternberg; Symplectic techniques in physics, Cambridge
Univ. Press, Cambridge, 1984.
[11] A. Haefliger, Homotopy and integrability, in ManifoldsAmsterdam 1970 (Proc.
Nuffic Summer School), Lecture Notes in Mathematics, Vol. 197 197, Berlin,
New York: Springer-Verlag, 133163.
[12] D. McDuff and D. Salamon; Introduction to symplectic topology, 2nd edition,
Oxford Mathematical Monographs, Oxford University Press, New York, 1998.
[13] I. Moerdijk and J. Mrcun; Introduction to foliations and Lie groupoids; Cam-
bridge Stud. Adv. Math. 91, Cambridge University Press, Cambridge, 2003.
[14] P. Molino; Riemannian foliations, Progress in mathematics vol. 73, Birkhauser,
Basel, 1988.
[15] M. Pflaum, H. Poshtuma and X. Tang; Geometry of orbit spaces of proper Lie
groupoids; to appear in J. Reine Angew. Math.. Preprint arXiv:1101.0180.
18 Rui Loja Fernandes
[16] A. Weinstein; Linearization problems for Lie algebroids and Lie groupoids,
Lett. Math. Phys. 52 (2000), 93-102.
[17] A. Weinstein; Linearization of regular proper groupoids, J. Inst. Math. Jussieu.
1 (2002), 493-511.
[18] N.T. Zung; Proper groupoids and momentum maps: linearization, affinity and
convexity, Ann. Scient. Ec. Norm. Sup. 4e serie, 39 (2006), 841-869.