1. Introduction
This paper is part of a series of papers devoted to the study of the geometry of singular spaces in terms of the theory of differential spaces, which were introduced by Sikorski [
1], see also [
2]. In this theory, geometric information about a space
S is encoded in a ring
of real valued functions, which are deemed to be smooth. In particular, we are concerned with the class of subcartesian spaces introduced by Aronszajn [
3]. A Hausdorff differential space
S is
subcartesian if every point
x of
S has a neighborhood
U that is diffeomorphic to a subset
V of a Euclidean (Cartesian) space
. The restriction of
to
U is isomorphic to the restriction of
to
V, see [
4].
Palais [
5] introduced the notion of a slice for an action of a not necessarily compact Lie group
G on a manifold
M. Since then, the structure of the space
of orbits of a proper action of
G on
M has been investigated by many mathematicians. In [
6] Duistermaat showed that
is a subcartesian differential space with differential structure
consisting of push forwards of smooth
G invariant functions on
M by the
G orbit mapping
. On a smooth manifold
M, there are two equivalent definitions of a vector field, namely as a derivation of
, or as a generator of a local one parameter local group of diffeomorphisms of
M. Choosing one, and proving the other is a matter of preference. On a subcartesian differential space
S, which is not in general a manifold, these notions differ. We use the term
vector field on
S for a generator of a local one parameter group of local diffeomorphisms and denote the class of all vector fields on
S by
. A key reason for the choice made in this paper is the special case of the orbit space of a proper action. The class of derivations of
is, in general, larger than the class
. For
we show that a derivation
Y of
is in
, i.e., is a vector field on
, if and only if there exists a
G invariant vector field
X on
M such that
Y is
related to
X, that is,
, where
is the
G orbit map.
In the literature, there has been extensive discussion about the notion of a differential form on a singular space, see Smith [
7], Marshall [
8], and Sjamaar [
9]. Here, in our search for an intrinsic notion of a differential form, we have been led to see them as multilinear maps on vector fields. In the case of a 1-form
on
with
a linear mapping
over the ring
of smooth functions on
, which is to say
for every
. With this definition we show that every differential 1-form on
pulls back under the orbit map
to a semi-basic
G invariant 1-form on
M. Furthermore, every
G invariant semi-basic differential 1-form on
M is the pull back by
of a differential 1-form on
.
We define a differential exterior algebra of differential forms on the orbit space, which satisfies a version of de Rham’s theorem. Our version is larger than Smith’s as it includes forms that are not Smith forms, see
Section 6. It also handles singular orbit spaces of a proper action of a Lie group on a smooth manifold. The Lie group need not be compact and the orbit space need not be smooth, both of which Koszul hypothesized in [
10].
We now give a section by section description of the contents of this paper.
Section 2 deals with basic properties of a proper action of a Lie group
G on a smooth manifold
M and the differential structure of the orbit space
. We introduce the reader to the theory of subcartesian differential spaces in the context of the orbit space
. The differential geometry of
M is described in terms of its smooth structure given by the ring
of smooth functions on
M. The differential geometry of the orbit space
, which may have singularities, is similarly described in terms of the ring
of smooth functions on
, which is isomorphic to the ring
of smooth,
G-invariant functions on
M. Since a proper action has an invariant Riemannian metric, several results are proved using properties of the geodesics of the metric. Additionally, certain objects are shown to be smooth submanifolds.
In
Section 3 we study vector fields on the orbit space
. In the case of the manifold
M, derivations of the ring
are vector fields on
M, and they generate a local one parameter group of local diffeomorphisms of
M. In the case of the ring
, not all derivations of
generate local one parameter groups of local diffeomorphisms of
. The derivations of
which generate local one parameter groups of local diffeomorphisms of
. This is the key idea of this paper. We establish that every vector field on the quotient
is covered by a
G-invariant vector field on
M. It is well known that the space
is stratified, see [
6,
11,
12]. We show that every vector field on
defines a vector field tangent to each stratum of
.
In
Section 4, we define differential 1-forms on the orbit space
as linear mappings on the space
of smooth vector fields on
. The most important consequence of this definition relates to pulling back 1-forms from
to
M. In particular, our notion of a differential 1-form is intrinsic.
In
Section 5 to prove a version of de Rham’s theorem we enlarge the algebra of differential 1-forms to
k-forms with an exterior derivative operator. The key techinical point is that everything is developed in terms of the Lie derivative of vector fields. Almost all of this section looks the same as that on manifolds.
In
Section 6 we give all the details of the simplest nontrivial example. This example reveals that differential forms in our sense are not the same as those of Smith [
7].
2. Basic Properties
This section gives some of the basic properties of smooth vector fields on the orbit space of a proper action of a Lie group on a smooth manifold.
Let
M be a connected smooth manifold with a proper action
of a Lie group
G on
M, and let
be the orbit map of the
G action
.
Let be the algebra of smooth G invariant functions on M and let be the algebra of functions on such that lies in . The map is a bijective algebra isomorphism, whose inverse is .
Proposition 1. The orbit space with the differential structure is a locally closed subcartesian differential space.
Proof. See Corollary 4.11 of Duistermaat [
6] and page 72 of [
4]. □
Let X be a smooth vector field on a manifold M. X gives rise to a map , called the derivation associated to X. If we want to emphasize this action of vector fields on M, we say that they form the space of derivations of . If we want to emphasize that X generates a local one parameter group of local diffeomorphisms of M, we say that X is a vector field on M and write for the set of vector fields on M. For each smooth manifold M we have . However, these notions need not coincide for a subcartesian differential space.
Let
be a differential space with
X a derivation of
. Let
be a maximal integral curve of
X, which starts at
x. Here
is an interval containing 0. If
t,
s, and
lie in
, and if
and
, then
The map
may fail to be a local diffeomorphism of the differential space
S, see example 3.2.7 in ([
4], p. 37). A
vector field on a subcartesian differential space
S is a derivation
X of
such that for every
there is an open neighborhood
U of
x and
such that for every
the map
is defined on
U and its restriction to
U is a diffeomorphism from
U onto an open subset of
S. In other words, the derivation
X is a vector field on
S if
is a local one parameter group of local diffeomorphisms of
S.
Example 1. Consider with the structure of a differential subspace of . Let be the inclusion mapping. The differential structure of consists of , which is the restriction of a smooth function f on to . Let be a vector field on . Then for every and every the function is smooth. Restricting to points in we obtain . We now show that we can obtain by operations on . Let and let be a sequence of points in , which converges to . Then Thus, we show that for every . In other words, the restriction of the vector field X to is a derivation of . Thus, . However, no two distinct points of can be joined by a smooth curve. Hence only the derivation of that is identically 0 on admits integral curves, i.e., .
Let
be the set of smooth
G invariant vector fields on
M, that is,
Since
, we have
. Additionally, we may consider the space
of derivations of
. Clearly, we have
. For
let
be the
local flow of
X, i.e.,
is a differentiable mapping such that
Here
D is a
domain, i.e.,
D is the largest (in the sense of containment) open subset of
such that for each
the set
is an open interval containing 0. Moreover,
for every
and if
,
, and
, then
. Since
,
Thus, for all , if .
Proposition 2. Let . Then X induces a derivation of defined by This leads to the module homomorphism Proof. Let
and
,
. Then
Since Y is a linear mapping of into itself, it follows that it is a derivation of .
We now show that the map
is a module homomorphism. For
X,
and
we have
Hence the map
is linear. For every
Therefore the map given by Equation (
5) is a module homomorphism. □
The importance of the module homomorphism (
5) stems from the following result.
Proposition 3. Since M is a smooth manifold, . So implies that .
Proof. Because the orbit space
is locally closed and subcartesian, every maximal integral curve of
X projects under the
G orbit map to a maximal integral curve of
Y. It follows that
Y is a smooth vector field on
, see proposition 3.2.6 on page 34 of [
4]. □
The following example shows that not every derivation on is a vector field on .
Example 2. Consider the action on generated by . The algebra of smooth invariant functions is generated by the polynomial . The orbit map of the action is . The derivation of is not a smooth vector field on , because its maximal integral curve starting at , given by , is defined on , which is not an open interval that contains 0.
Fix
. Then
is the isotropy group of the action
at
m. It is a compact subgroup of
G, see Duistermaat and Kolk [
13]. Let
H be a compact subgroup of
G. The set
is a submanifold of
M, which is not necessarily connected. Hence its connected component are submanifolds. Connected components of
are
H invariant submanifolds of
M, see Duistermaat and Kolk [
13]. The conjugacy class in
G of a closed subgroup
H is denoted by
and is called a
type. The set
is called an
orbit type . Moreover, the
G invariance of
implies that each connected component of
is
G invariant. The orbit type
is associated to the type
. Let
and
be two types. Define the partial order ≤ by the condition:
if some
G conjugate of
is a subgroup of
. Since the orbit space
is connected, there is a unique maximal orbit type
.
Proposition 4. The maximal orbit type is open and dense in M.
Proof. See page 118 of Duistermaat and Kolk [
13]. □
Proposition 5. The orbit type is a smooth invariant submanifold of every vector field .
Proof. Let
. Then
for some
. Let
be the local 1 parameter group of local diffeomorphisms of
M generated by
. Then
since
. Thus,
. Conversely, suppose that
. Then
. So
that is,
. Thus,
. Consequently,
, i.e.,
. □
Let Y be a smooth vector field on , which is related to the smooth G invariant vector field X on M, i.e., for . Then for . The set is well defined, because for every . So is the domain of a local generator of Y, i.e., is a differentiable mapping such that for .
Please note that the orbit type need not be connected and its connected components may be of different dimensions. In the following we concentrate our attention on the properties of the connected components of , which we denote by .
Proposition 6. For every compact subgroup H of G the image of each connected component of the orbit type under the orbit mapping is a smooth submanifold of the differential space .
Proposition 7. The connected component of the orbit type of is an invariant manifold of every smooth vector field Y on .
Proof. Let be the local flow of the vector field Y. For each point there is an open neighborhood of y in such that for every the map is a local diffeomorphism onto its image. Hence for every . So is an invariant manifold of the vector field Y. □
Theorem 1. Let H be a compact subgroup of the Lie group G. Let Y be a smooth vector field on . Then on every connected component of the orbit type there is a smooth G invariant vector field X, which is related to .
We need the next few results to prove this theorem, which is the main result of this section.
Lemma 1. The G orbit through is a submanifold of M.
Proof. Let
be a slice to the
G action
at
m. By Bochner’s lemma, see ([
14], p. 306), there is a local diffeomorphism
, which sends
to
and intertwines the
action
on
with the
H action
. Since
is
H invariant, it follows that
is a local diffeomorphism which sends
to
m. Let
L be a complement of
in
, where
is the Lie algbebra of
H. The map
which sends
to
m is a local diffeomorphism that sends an open neighborhood of
in
onto an open neighborhood of
m in
. Thus,
is a smooth submanifold of
M near
m. For every
, since
is a diffeomorphism of
M, the map
is a local diffeomorphism of
in
onto an open neighborhood of
in
. Thus,
is a submanifold of
M. □
Lemma 2. For each connected component of the orbit type the mapis a smooth surjective submersion, whose typical fiber is an orbit of the G action Φ restricted to . Proof. The orbit map
is a surjective smooth map of the smooth manifold
M onto the differential space
. Hence its restriction
to the connected component
of the orbit type
and the codomain to
is a smooth map of the smooth manifold
onto the differential space
. By Proposition 6, the differential space
is a smooth manifold. Hence
is a smooth map of the smooth manifold
onto the smooth manifold
. At
, the fiber
is the
G orbit in
through
m, which is a smooth submanifold of
. We have
, using the restriction of the
G invariant Riemannian metric on
M to
, see Palais [
5]. Because the vector space
is isomorphic to
, with the same dimension, it follows that the map
is surjective. Consequently, the map
is a submersion. □
Next we construct a connection on the fibration
and then review some geometric facts about geodesics. Because the
G action
(
1) on the smooth manifold
M is proper, it has a
G invariant Riemannian metric. Let
be the restriction of this metric to the smooth submanifold
. For each
this yields the
G invariant decomposition
where
and
, using the metric
on
. The distributions
and
are smooth. Moreover, for every
the map
where
, is an isomorphism of vector spaces. Thus, Equations (
9) and (
10) define an Ehresmann connection
on the fibration
. Because
and
imply
and
, the distributions
and
are
G invariant. Thus, the connection
is
G invariant.
Let
and
be the tangent and cotangent bundle projection maps, respectively. The metric
on
defines a vector bundle isomorphism
where
. The inverse of
is
. The metric
determines the Hamiltonian function
which gives rise to the Hamiltonian system
, where
is the canonical symplectic form on
. The Hamiltonian vector field
on
is defined by
. For
let
be the local flow of the vector field
, which is defined for
t in an open interval
in
containing 0. For
the curve
given by
is a
geodesic on
, starting at
, for the metric
. There is an open tubular neighborhood
U of the zero section of the cotangent bundle
such that the local flow
is defined for all
. For each
, the
exponential map
is a diffeomorphism onto
, where
is a suitable open neighborhood of
, see Brickell and Clark [
15].
Next we reduce the
G symmetry of the Hamiltonian system
. Because the metric
on
is
G invariant, the smooth Hamiltonian
E (
11) on
is
G invariant, it induces a metric
on
such that
, where
is orthogonal projection. The smooth Hamiltonian
E (
11) on
is
G invariant, and hence induces a smooth Hamiltonian function
Since the
G orbit map
is smooth, symplectic reduction of the Hamiltonian system
leads to the
G reduced Hamiltonian system
. The reduced system has a Hamiltonian vector field
defined by
. Its local flow
is
related to the local flow
of
, i.e.,
. The curve
given by
is a geodesic on
starting at
for the reduced metric
. Here
is the cotangent bundle projection map. Please note that
, where
. There is an open neighborhood
of the zero section of
such that the local flow
of the reduced vector field
is defined for all
. The reduced exponential map
is a diffeomorphism onto
, where
.
Proposition 8. The fibration is locally trivial.
Proof. For some
, the open ball
is a subset of
. Then
is a submanifold of
containing
m. Look at the geodesic
on
given by
starting at
. One has
. To see this, observe that there is a
such that
, since
is a diffeomorphism. So
Thus, is a diffeomorphism of the open ball of radius contained in onto a submanifold .
For every
let
be the geodesic in
joining
to
, i.e.,
. Because
is a diffeomorphism, the vector
is uniquely determined by
. Let
where
is the horizontal lift of the geodesic
using the connection
. Here
with
,
, and
. The map
is
parallel translation of the fiber
along the geodesic
joining
to
in
using the connection
.
Consider the mappings
and the projection mapping
Then
is a local trivialization of the fibration defined by the mapping
because for every
and every
We now show that
is a diffeomorphism. Define the smooth maps
and
The following calculation shows that
.
Additionally,
that is,
. Thus,
is a diffeomorphism.
The preceding argument can be repeated at each point of . Hence the fibration is locally trivial. □
Corollary 1. The locally trivial fibration defined by (8) has a local trivialization where V is an open G invariant subset of with k, and . Proof. Suppose that
for some
,
and some
. Then
So
. In other words, the
G orbit of
in
is
. Hence
for every
. Thus, the map
(
13) is given by
Clearly,
is a diffeomorphism. It intertwines the
G action
with the
G action
and satisfies
, where
. Hence
is a local trivialization. □
Lemma 3. Every smooth vector field on is related to a smooth G invariant vector field on .
Proof. To see this let
, the Lie algebra of
G. Let
, where
is left translation on
G by
g. By construction
. So
is a vector field on
, which is smooth. For every
and every
one has
So
is a
G invariant vector field on
. Moreover,
So the vector field on and the vector field on U are related. □
Lemma 4. Every smooth vector field on U is related to a smooth G invariant vector field X on V.
Proof. Pull the vector field
on
back by the trivialization
(
13) to a vector field
X on
V. Since
intertwines the
G action
on
V with the
G action
on
, the vector field
X is
G invariant. For
and
one has
Thus, the G invariant vector field X on V is related to the vector field on U. □
Proof of Theorem 1. We just have to piece the local bits together. Cover the orbit type
by
, where
is a local trivialization of the bundle
. Let
Y be a smooth vector field on
. Since
and
is an open mapping,
is an open subset of
. Because
covers
, it follows that
is an open covering of
. Applying Lemma 3 to the smooth vector field
and then using Lemma 4, we obtain a
G invariant vector field
on
, which is
related to the vector field
on
. Since
Y is a smooth vector field on
, on
, where
i,
, one has
. So on
one has
. Thus, the
G invariant vector fields
piece together to give a smooth
G invariant vector field
X on
. Since
is
related to the vector field
, the vector field
X on
is
related to the vector field
Y on
. □
3. Vector Fields on
We start with a local argument in a neighborhood of a point
with compact isotropy group
H. By Bochner’s lemma there is a local diffeomorphism
, which sends
to
and intertwines the linear
H action
with the
H action
. Because the
G action
on the smooth manifold
M is proper, it has a
G invariant Riemannian metric
. Using the restriction of
to
, we define
. Then there is an
H invariant open ball
centered at
with
B contained in the domain of the local diffeomorphism
such that
is a slice to the
G action on
M at
m.
We now construct a model for the
H orbit space
of the restriction to
B of the linear action
on
H on
. Let
be a basis of the vector space
E. Hence
E is isomorphic to
. Let
be coordinates on
with respect to the basis
. Let
be a set of generators for the algebra of
H invariant polynomials on
. Let
be coordinates on
. The orbit map of the
H action on
is
By Schwarz’ theorem
is
H invariant, i.e.,
, if and only if there is a function
such that
for every
. Smooth functions on the orbit space
are restrictions to
of smooth functions on
. For every
the pull back
by the orbit map
is given by
The
H orbit map
is a smooth map of differential spaces. We are interested in
, the space of
H orbits on the open ball
B in
. Restricting
to the domain
and the codomain
gives
which is a surjective smooth map of differential spaces.
Lemma 5. Let be the orbit mapping of a linear action of a compact Lie group H on an open ball B in . For every smooth vector field Y on there is a smooth H invariant vector field X on B, which is related to Y, i.e., for every .
Proof. Let Y be a smooth vector field on . Since is a differential subspace of , in coordinates on we may write , where is the restriction to of a smooth function on .
We begin the proof by showing that the orbit space
is connected. Observe that the open ball
is centered at the origin and the action of
H on
B is the restriction of the linear action of
H on
, see Equation (
16). The linearity of the action of
H on
B implies that it commutes with scalar multiplication. Moreover, the origin
is
H invariant so that it is an orbit of
H. Hence
. Let
. For each
and every
we have
. Therefore the line segment
joins
H orbits through the points
and
. Thus, the
H orbit through
and the
H orbit through
belong to the same connected component of
. This implies that
is connected.
The connectedness of
ensures that there is a unique principal type
whose corresponding orbit type
is open and dense in
B, see Duistermaat and Kolk ([
13], p.118). Moreover, the orbit space
is a connected smooth manifold, and
is a locally trivial fibration, whose fiber over
is the
H orbit
. Hence for every
there exists an open neighborhood
V of
in
such that
is trivial. In other words, there is a diffeomorphism
such that
, where
is projection on the second factor. This implies that there is a smooth
H invariant vector field
on
, which is
related to the restriction
of
Y to
.
Repeating the above argument at each point leads to a covering of by H invariant open subsets of on which there exists an H invariant vector field , which is related to the restriction of the vector field Y to . Using an H invariant partition of unity on , we obtain a vector field on , which is related to , i.e., .
The module
of
H invariant smooth vector fields on
B is finitely generated by polynomial vector fields, see [
16], and we denote a generating set by
. Hence, every
H invariant smooth vector field
on
B is of the form
for some
. Similarly, every
K invariant smooth vector field on
can be written as
, where
. Since
is open and dense in
B, a generic
need not extend to a smooth function on
B. Therefore a generic vector field on
need not extend to a smooth vector field on
B. On the other hand, the
H invariant vector field
on
is obtained above from a smooth bounded vector field
Y on
Therefore
for each
where every
, and
is the restriction of
to
Since
is open and dense in
B, we may define
provided that
exists and is unique. Since the vector fields
are smooth on
B,
Moreover, since
is open and dense in
it is open and dense in
, the closure of
B. In Equation (
19), each function
is the restriction to
of a smooth function
on
Moreover, the choice of polynomial basis
ensures that the right-hand side of Equation (
19) extends to the closure
of
B. Hence all the the limits
in Equation (
20) exist, and
is defined for all
.
We need to show that this definition of
X on
B depends only on
. Since each
is continuous on
B and its first partial derivatives are bounded on
, it follows that
are uniformly continuous on
In particular, if
is a smooth curve, such that
and
, then
Thus, the values of on B are uniquely determined by . Repeating this argument for all the first-order partial derivatives of , we deduce that the first-order partial derivatives of on B are uniquely determined by and its first partial derivatives. Continuing this process for every partial derivative of every order shows that the restriction of to B is uniquely determined by .
The above argument applies to each of the functions for in Equation (19) and ensures that the H invariant vector field (19), thought of as the smooth section , extends to a smooth H invariant map , which is related to the section . It remains to show that X is a vector field on B.
By construction
is an
H invariant vector field on an open dense subset
of
B, which is
related to the vector field
. The closure of
in
B is the union of orbit types
, where
. Suppose that
, where
for all
. Then
and
because
is the restriction to
of a smooth, and hence continuous, vector field on
. By Proposition 7, for every orbit type
, the manifold
is an invariant manifold of the vector field
Y. So
is a vector field on
. Hence for every
,
Therefore X is a smooth vector field on B, which is related to the vector field Y on . □
The aim of the rest of this section is to prove.
Theorem 2. Let be a proper action of a Lie group G on a connected smooth manifold M with orbit map . Every smooth vector field on the locally closed subcartesian differential space is π related to a smooth G invariant vector field on M.
First we prove.
Lemma 6. Let be a slice to the G action Φ at and suppose that is a smooth H invariant vector field on some H invariant open neighborhood of m in . Here H is the isotropy group at m. Then the vector field extends to a smooth G invariant vector field X on M.
Proof. Let
be an
H invariant open subset of
containing
m. Because
is a slice,
is a
G invariant open subset of
M, which contains the
G orbit
. On
define the vector field
. We check that
X is well defined. Suppose that
, where
g,
and
,
. Since
is a slice, it follows that
. Hence
where the last equality above follows because the vector field
is
H invariant. So the vector field
X on
is well defined and by definition is
G invariant.
Next we show that
X is smooth. Let
L be a complement to the Lie algebra
of the Lie group
H in the Lie algebra
of the Lie group
G. For every
and
consider the map
, which sends the identity element
of
G to
. It is a local diffeomorphism, since its tangent
is the identity map. Thus, there are open subsets
,
, and
of
,
, and
, respectively, such that
. Hence every
may be written uniquely as
for some
and some
. For every
we have
Consider the local diffeomorphism
Then
. So if
, then
. Let
W be a neighborhood of
such that
restricted to
W yields a diffeomorphism
. For
let
be the integral curve of the vector field
starting at
s. Since
is a
G invariant extension of
to a vector field on
(whose smoothness we want to prove) of a smooth
H invariant vector field
on
, it follows that
is a smooth vector field on
. Therefore
for all
. Consider a curve
in
starting at
defined by
. Using Equation (
22) we obtain
for all
. Since the family of curves
depends smoothly on
and
U is an open subset of
containing
, it follows that
is a smooth vector field on
U. For any
there exists a
such that the open set
contains
. Since
X is
G invariant, smoothness of
X on
U ensures that
X is smooth on
. Hence
X is a smooth vector field on
.
The above argument can be repeated at each point . This leads to a covering of M by open G invariant subsets , where is a slice at for the action of G on M and I is an index set. If Y is a vector field on , then for each there exists a G invariant vector field on that is related to the restriction of Y to . Using a G invariant partition of unity on M subordinate to the covering , we can glue the pieces together to obtain a smooth G invariant vector field X on M, which is related to the vector field Y on . □
Proof of Theorem 2. Applying Lemma 6 to the push forward by the local diffeomorphism , given by the Bochner lemma, of the vector field on B constructed in Lemma 5, proves Theorem 2. □
Proposition 9. If Y is a derivation of , which is π related to a derivation X of , then Y is a smooth vector field on .
Proof. Since M is a smooth manifold, X is a smooth G invariant vector field on M, which is related to derivation Y of . Thus, the image under of a maximal integral curve of X on M, is a maximal integral curve of Y on . Hence Y is a smooth vector field on the locally closed subcartesian differential space . □
4. Differential -Forms on the Orbit Space
In this section we define the notion of a differential 1-form on the orbit space of a proper group action on a smooth manifold M with orbit map . We show that the differential 1-forms on together with the exterior derivative generate a differential exterior algebra.
Theorem 2 and Proposition 9 show that
Y is a vector field on
if and only if there is a
G invariant vector field
X on
M, which is
related to
Y, i.e., every integral curve of
Y is the image under the map
of an integral curve of
X. Let
be the set of differential 1-forms on
, i.e., the set of linear mappings
which are linear over the ring
, i.e.,
for every
and every
.
In order to prove some basic properties of differential 1-forms on
, we need to prove some properties of the
G orbit map
(
2).
The map
where
and
Y is the vector field on
constructed in Proposition 2, is the
tangent of the map
at
. To show that
is well defined we argue as follows. Suppose that
, where
. Then
since
is a linear map.
Lemma 7. For each where is the Lie algebra of G. Proof. By definition
. Thus,
The curve is an integral curve of starting at m. So for every . Thus, , i.e., . Consequently, .
To prove the reverse inclusion, we argue as follows. Since
, it follows that
, where
. Let
be the maximal orbit type of the proper
G action on
M. The maximal orbit type
is a dense open subset of
M, whose boundary
contains
, since the orbit types of the
G action stratify
M. Suppose that
is a nonzero vector in
. There is a vector field
X on
M with
with an integral curve
starting at
such that
for some
. We may suppose that
. Since
is a smooth submanifold of the differential space
, the curve
is a smooth integral curve of the vector field
such that
. Hence on
the curve
on the smooth manifold
is smooth. Thus,
since
. Thus, the curve
is constant, since
is a smooth manifold. Because the curve
is continuous on
, we obtain
. Hence
for all
. But
. So
. Hence
where the equality follows from Equation (
24). This verifies Equation (
23). □
A differential 1-form on M is semi-basic with respect to the G action if and only if for every , the Lie algebra of G.
Proposition 10. For every , the differential 1-form on M is G invariant and semi-basic.
Proof. By definition of
the map
is linear, since the map
is linear. Moreover, for any
Thus, . For every one has , because . So is a semi-basic 1-form on M. □
Proposition 11. Let ϑ be a G invariant semi-basic differential 1-form on M. Then there is a 1-form θ on such that .
Proof. Given
, there is an
, which is
related to
Y, i.e.,
for every
. It is clear that the definition of
needs to be
It remains to show that
is well defined. Since the 1-form
and the vector field
X are
G invariant, we obtain
for every
. Thus, the function
is smooth and
G invariant. We now show that the mapping
, where
is given in Equation (
26), is well defined. Suppose that
such that
is
related to
Y. Then
for every
. So
, by Proposition 10. Thus,
since the 1-form
on
M is semi-basic. This shows that the map
is well defined. From Equation (
26) it follows that
is a linear mapping and that
for every
. Hence
is a differential 1-form on
, i.e.,
. Every
is
related to a
. Thus,
, that is,
. □
5. De Rham’s Theorem
In this section we construct an exterior algebra of differential forms on the orbit space with an exterior derivative and show that de Rham’s theorem holds for the sheaf of differential exterior algebras.
Let
. A differential
ℓ-form
on
is an element of
, the set of alternating
ℓ multilinear real valued mappings on
, namely
which is linear over
, that is,
for every
and every
. A differential 0-form on
is a smooth function on
. Let
be the real vector space of differential
ℓ-forms on
. For each
let
.
Proposition 12. Let with . Then the ℓ-form , the set of semi-basic G invariant ℓ-forms on M. Herefor every and every for . Proof. The proof is analogous to the proof of Proposition 10 for 1-forms on and is omitted. □
Proposition 13. Let , where . Then there is an ℓ-form such that .
Proof. The proof is analogous to the proof of Proposition 11 for G invariant semi-basic 1-forms on M and is omitted. □
We now define the exterior algebra of differential forms on . Let and . The exterior product is the form on corresponding to the G invariant semi-basic -form on M. Then is an exterior algebra of differential forms on .
The exterior derivative operator on is defined in terms of the Lie bracket of vector fields on . If Y, , then there are , , each of which is related to Y and , respectively. Their Lie bracket . Then there is a vector field on , which is related to . Define . The following lemma shows that this Lie bracket is well defined.
Lemma 8. For every and every Y, Proof. We compute.
from which Equation (
27) follows, because the orbit map
is surjective. □
Corollary 2. is a Lie bracket on .
Proof. The corollary follows from a computation using Equation (
27). We give another argument. Bilinearity of the Lie bracket is straightforward to verify. We need only show that the Jacobi identity holds. We compute.
which is the Jacobi identity on
. □
Let
be an
ℓ-form on
. Inductively define the exterior derivative
of
as the
-form given by
Here
is the
form on
defined by
for
. To complete the definition of exterior derivative, we define
on 0-forms. This we do as follows. Let
. Define the 1-form
by
, for every
, every
, and every
.
Lemma 9. Let . Then Proof. Suppose that
is an
ℓ-form on
. Pulling back the forms on both sides of Equation (
28) by the orbit map
gives
By induction, assume that Equation (
29) holds for all forms of degree strictly less than
ℓ. Then
Now
, where
, since
Thus, Equation (
30) reads
□
Lemma 10. If and , then Proof. On
M we have
which implies that Equation (
32) holds, since the orbit map
is surjective. □
Lemma 11. for every .
Proof. Suppose that
. Then
is an
ℓ-form on
M. Because
M is a smooth manifold, one has
. By Lemma 8
. So
Since the G orbit map is surjective, implies .
We now treat the case when
. Let
and let
,
. Then
□
We prove an equivariant version of the Poincaré lemma in .
Lemma 12. Let G be a Lie group, which acts linearly on by . Let H be a compact subgroup of G. Let β be an H invariant closed ℓ-form with on an open H invariant ball B centered at the origin of , whose closure is compact. Suppose that β is semi-basic with respect to the G action Φ, i.e., for every , the Lie algebra of G. Here . Then there is an H invariant -form α on B, which is semi-basic with respect to the G action Φ, such that .
Proof. Let
X be a linear vector field on
all of whose eigenvalues are negative real numbers. By averaging over the compact group
H, we may assume that
X is
H invariant. Let
be the flow of
X, which maps
B into itself. Moreover,
. On
B one has
The
-form
on
B is
H invariant, since
commutes with the
H action on
B, and
is an
H invariant
-form on
B, because the vector field
X and the
ℓ-form
are both
H invariant. Thus,
on
B. Moreover,
is
G semi-basic, since
The last equality above follows because the ℓ-form is G semi-basic. □
Since is a locally contractible space, we have
Proposition 14 (Poincaré Lemma). Let be a closed ℓ-form on with . For each there is a contractible open neighborhood of and an -form ϕ on such that on .
Proof. Since the
G action
on
M is proper, it has a slice
at
m, where
. Using Bochner’s lemma there is an open neighborhood
of
m in
, which is the image of an
invariant open ball
, centered at the origin
whose closure is compact, under a diffeomorphism
. The diffeomorphism
intertwines the linear
H action
with the
action
on
. Let
be the semi-basic
G invariant form on
such that
. Since
is closed by hypothesis, it follows that the semi-basic
ℓ-form
on
is closed. Let
. Then
is a semi-basic
H invariant closed
ℓ-form on
. Under the map
the
ℓ-form
pulls back to a
G semi-basic
H invariant
ℓ-form
on
. By Lemma 12 there is a
G semi-basic
H invariant
-form
on
B such that
. Hence
is a semi-basic
H invariant
-form on
. The
-form
on
extends to a
G invariant
-form
on
defined by
for every
and every
. Arguing as in the proof of Lemma 6, it follows that
is a smooth
G invariant
-form on
. The form
is semi-basic. Moreover,
on
, since for every
one has
Let
. Since
is contractible and the
G orbit map
is continuous and open, it follows that the open neighborhood
of
is contractible. Since the
ℓ-form
is semi-basic, there is an
ℓ-form
on
such that
on
. On
we have
Because the orbit map is surjective, it follows that on , which proves the proposition. □
Lemma 13. Let and suppose that is a connected open neighborhood of such that , then is constant on .
Proof. It follows from our hypotheses that is a smooth G invariant function on the open connected component of containing m. Moreover, on we have . Since M is a smooth manifold, it follows that f is constant on . Hence is constant on the connected open set because is a continuous open map. □
To prove de Rham’s theorem, we will need some sheaf theory, which can be found in appendix C of Lukina, Takens, and Broer [
17]. Let
be an open covering of
. Because
is locally contractible, the open covering
has a good refinement
, that is, every
with
is locally contractible and
is either contractible or empty. In addition, because
is paracompact, every open covering has a locally finite subcovering. Since the
G action on
M is proper, the orbit space
has a
partition of unity subbordinate to the covering
.
Define the differential exterior algebra valued sheaf
over
by
whose sections are differential forms on
. The sheaf
induces the subsheaves
whose sections are differential
ℓ-forms on
. Please note that
is a smooth vector bundle, as is
Let
be the sheaf of locally constant
-valued functions on
. The two exact sequence of sheaves
are exact.
We say that the sheaf is fine if for every open subset of , every smooth function on and every smooth section of the sheaf , then .
Theorem 3. The sheaves Λ and of sections of the vector bundles Λ and are fine.
Proof. We treat the case of the sheaf . The proof for the sheaf is similar and is omitted. The definition of fineness holds by definition of differential form. □
Corollary 3. Λ and are fine sheaves of sections over , which is paracompact. Let be an open covering of . Then , the sheaf of cohomology group of with values in the sheaf Λ, vanishes for all . Similarly, for all .
We are now in position to formulate de Rham’s theorem. Let be the sheaf of differential ℓ-forms on and let be the sheaf homomorphism induced by exterior differentiation. For each let , whose elements are closed ℓ-forms on . By Lemma 13 . Define the de Rham cohomology group when and . Here is the set of sections of the sheaf .
Theorem 4 (de Rham’s Theorem). The sheaf cohomology of with coefficients in does not depend on the good covering of . Thus, for every the de Rham cohomology group is isomorphic to the sheaf cohomology group of the good covering with values in the sheaf of locally constant real valued functions.
Proof. We give a sketch, leaving out the homological algebra, which is standard. For more details, see [
17] or [
18]. Let
be a good covering of
. The Poincaré lemma holds on any finite intersection of contractible open sets in
, so the following sequence of sheaves is exact
where
is the inclusion mapping. This exact sequence gives rise to the long exact sequence of cohomology groups
where
,
, and
are homomorphisms on cohomology induced by the inclusion, exterior differentiation and coboundary homomorphisms, respectively. Since
is a fine sheaf, its cohomology vanishes for
and the above sequence falls apart into the exact sequence
and for every
the exact sequence
Now
. Applying the sequence (
34) consecutively gives
Exactness of the sequence (
35) gives
Here ≃ means is isomorphic to. □
Corollary 4. For the zeroth cohomology we have Our version of de Rham’s theorem is not the same as Smith’s version, since the
invariant semi-basic 1-form
in
Section 6 is not a Smith 1-form, see also Smith ([
7], p. 133). However, his cohomology and ours agree. Our results extend those of Koszul [
10], who hypothesized that
was a smooth manifold and that the group
G was compact.
6. An Example
In this section we give an example, which illustrates Theorem 2 and the construction of differential 1-forms on the orbit space of a proper group action on a smooth manifold.
Consider the
action on
generated by
The algebra of
invariant polynomials on
is generated by the polynomials
,
, and
, which are subject to the relation
Let
be the Hilbert map of the
action associated to the polynomial generators
,
, and
. The map
(
38) is the orbit map of the
action on
. The relation
defines the orbit space
as a closed semialgebraic subset
of
with coordinates
. Geometrically
is a cone in
with vertex
.
Because
is a compact Lie group, which acts linearly on
, Schwarz’ theorem [
19] implies that the space
of
invariant smooth functions on
is equal to
, where
if and only if there is an
such that
.
Lemma 14. Let satisfy for every . Then there are , such that for every .
Proof. Since
, it follows that
. Suppose that there is an integer
such that
for
and
. Then by Taylor’s theorem with integral remainder we have
for every
, where
and
for
. By hypothesis
So
If
k is odd Equation (
40) implies
for
. Consequently,
which proves the lemma when
k is odd. When
k is even, Equation (
40) reads
for
, which implies
for
. In particular,
, which contradicts our hypothesis.
Now suppose that
f is flat at
, i.e.,
for every
. Then
and
divide
f, i.e.,
and
are smooth functions on
. To see this note that
and
are smooth for all
. Since
f is flat at
, so are
and
. Clearly
. From
it follows that
. Similarly,
. □
Proposition 15. The module of invariant smooth vector fields on is generated by Proof. A smooth vector field
X on
may be written as
, where
f and
g are smooth.
if and only if
that is,
and
for every
. Using lemma 38 write
and
, where
,
,
, and
. Hence for every
we have
where
,
,
, and
. □
Lemma 15. The vector fieldson , where for , are σ related to the invariant vector fields (41) for . Proof. The calculation
gives the vector fields
on
. Since
for
, the vector fields
on
given by (
43) leave invariant the ideal
I of
generated by
. Hence for each
the vector field
define the vector field
on
, which is given in Equation (
42). The vector fields
are
related to the
invariant vector fields
(
41) for
, because
. □
Since the tangent to the Hilbert mapping
(
38) is defined and is surjective, the tangent bundle
of the semialgebraic variety
(
37) is the semialgebraic subset of
with coordinates
defined by Equation (
37) and
By Theorem 2 every smooth vector field on
is
related to a smooth
invariant vector field on
. Because the
module
of smooth
invariant vector fields on
is generated by the vector fields
for
given by Equation (
41), it follows that the
related vector fields
for
given by Equation (
42) generate the
module
of smooth vector fields on
.
Lemma 16. The differential 1-formsgenerate the module of invariant 1-forms on . Proof. We use the differential forms
instead of those given in (
44), because we then obtain
for
. Suppose that the 1-form
on
, where
for
, is invariant under the
action generated by
. Then for every
So
if and only if for
one has
for every
. By Lemma 14 if
for some
, then there are
for
such that
for every
. Consequently, for some
,
,
,
for every
. Here
,
,
, and
. This proves the lemma. □
For
define the 1-forms
on
by
see the proof of Proposition 10. The 1-forms
generate the
module of 1-forms on
, since the
invariant 1-forms
for
generate the
module
of
invariant 1-forms on
, see proposition 23. Every
invariant 1-form on
is semi-basic, since the Lie algebra of
is
.
Let be the 1-form on Σ defined by its valuesHere for . The 1-form is not the restriction of a 1-form on to Σ. Proof. Equation (
47) follows immediately from the definition of
given in Equation (
46).
We give three proofs of the assertion about .
- 1.
Consider the 1 form
on
. Then
The following argument shows that the 1 -form
is not smooth, because its coefficients are not smooth functions on
. First we need some geometric information about the
orbit space
defined by
with
and
. The only subgroups of
are the identity
and
. The isotropy group
at
is
if
and
if
. The corresponding orbit types are
and
, whose image under the orbit map
is
, the vectex of the cone
, and
, which is a smooth manifold. Thus,
is a smooth 1-form, whose pull back under
is the smooth 1-form
on
. The 1-form
does not extend to a smooth 1-form
because the functions
and
are not smooth at
, the vertex of the cone
. To see this let
. The closed line segment
, where
, lies in
and joins
to
. Now
. So
. Hence the function
is not continuous at
. A similiar argument shows that the function
is not continuous at
.
- 2.
The following argument shows that the 1-form
on
defined in Equation (
48) is not the restriction to
of any smooth 1-form on
. Suppose it is. Then
, for some
, where
I is the ideal of
generated by
. Using (
48) we obtain
which implies
Similarly,
Set
and
. Then
So Equation (
49a) holds. Multiplying (
49b) by
and (
49c) by
and adding gives
But
does not divide
, which does not lie in
I. Thus, Equation (
50) does not hold for any choice of
. Hence our hypothesis is false, i.e., the 1-form
on
is not the restriction to
of a 1-form on
.
- 3.
Our third proof is more analytic. The 1-form
(
48) on the orbit space
is not the restriction to
of a 1 form
on
, where
. Suppose that
, then
which does not vanish at
. However, the 2-form
vanishes at
, since the 1-forms
(
44) for
vanish at
. This is a contradiction, since
. □