8 March 2005
2000 Mathematics Subject Classification. 53C15. First author partially supported by NSF grant DMS 0305939, and second by NSF grant DMS 0204448.
On nearly semifree circle actions
Dusa McDuff
Susan Tolman
Department of Mathematics, Stony Brook University, Stony Brook, NY 11794-3651, USA E-mail address : dusa@math.sunysb.edu Department of Mathematics, University of Illinois at Urbana–Champaign, IL, USA E-mail address : tolman@math.uiuc.edu
-
Abstract.
Recall that an effective circle action is semifree if the stabilizer subgroup of each point is connected. We show that if
is a coadjoint orbit of a compact Lie group
then every element of
may be represented by a semifree
-action. A theorem of McDuff–Slimowitz then implies that
injects into
, which answers a question raised by Weinstein. We also show that a circle action on a manifold
which is semifree near a fixed point
cannot contract in a compact Lie subgroup
of the diffeomorphism group unless the action is reversed by an element of
that fixes the point
. Similarly, if a circle acts in a Hamiltonian fashion on a manifold
and the stabilizer of every point has at most two components, then the circle cannot contract in a compact Lie subgroup of the group of Hamiltonian symplectomorphism unless the circle is reversed by an element of
1 Introduction
This paper is an attempt to understand topological properties of Lie group actions. Its starting point was the following theorem of McDuff–Slimowitz [1] concerning circle subgroups of
, the group of symplectomorphisms of a symplectic manifold
. Recall that an effective circle action is semifree if the stabilizer subgroup of each point in
is either the circle itself or the trivial group. Also, we say that a circle subgroup
of a topological group
is essential in
if it represents a nonzero element in
and inessential in
otherwise.
Theorem 1.1.
Any semifree circle action on a closed symplectic manifold
is essential in
.
This is obvious if the action is not Hamiltonian since in this case the flux homomorphism
does not vanish on
. However, if the action is Hamiltonian with generating Hamiltonian
then the result is not so easy: the proof in [1] involved studying the Hofer length of the corresponding paths
in
.
The first result in this paper uses the theorem above to answer a question posed by Alan Weinstein in [4] . Let
be a semisimple Lie group with Lie algebra
. Let
be a coadjoint orbit, together with the Kostant–Kirillov symplectic form
. If the coadjoint action of
on
is effective, then
is naturally a subgroup of
, the group of Hamiltonian symplectomorphisms of
. This inclusion induces a natural map from the fundamental group of
to the fundamental group of
. Weinstein asks when this map is injective. We prove that this map is injective for all compact semisimple Lie groups. In [3] Vina established a special case of this result by quite different methods.
Theorem 1.2.
Let a compact semisimple Lie group
act effectively on a coadjoint orbit
. Then the inclusion
induces an injection from
to
In view of Theorem 1.1 , this is an immediate consequence of the following result, which we prove in Section 2 .
Proposition 1.3.
Let a compact semisimple Lie group
act effectively on a coadjoint orbit
. Then every nontrivial element in
may be represented by a circle that acts semifreely on
.
Theorem 1.1 immediately implies that if a compact Lie group
acts effectively on a closed symplectic manifold
, then any semifree circle subgroup
is essential in
. The other results in the paper generalize this claim. Observe that Theorem 1.1 does not immediately extend to the smooth (non-symplectic) category. For example, Claude LeBrun pointed out to us that the circle action on
induced by the diagonal action of
on
is semifree but gives a nullhomotopic loop since
. Nevertheless, the semifree condition does have consequences in the smooth category, even if the action is only semifree on a neighborhood of a component of the fixed point set; we shall say that such components are semifree. Further, given a circle subgroup
we say that
reverses
in
if
for all
. Finally, a component
of the fixed point set
of
is symmetric in
if there is an element
whose action on
fixes
pointwise and which reverses
.
Theorem 1.4.
Let
be a circle subgroup of a compact Lie group
which acts effectively on a connected manifold
. If there is a semifree component of the fixed point set
which is not symmetric in
, then
is essential in
.
Example 1.5.
First, let
act on
by the defining representation on the first two copies of
, and let
be the circle subgroup given by
. This action has a semifree fixed point, namely
. Moreover, this circle subgroup is inessential in
. Therefore, by the theorem above, there exists
which reverses the circle action and fixes
. In fact, we can take any
which lies in the normalizer
but not in
itself. Note that
for any such
.
In contrast, consider the natural action of
on
, and let
be the circle subgroup given by
. This action is semifree and essential, but is not reversed by any
. To see this, note that the circle has order
in
, whereas every circle that can be reversed has order
or
.
One can weaken the semifree hypothesis in the above theorem, at the cost of adding a global isotropy assumption and working once more in the symplectic category. We say that a circle action has at most twofold isotropy if every point which is not either fixed or free has stabilizer
Recall, also, that a symplectic action of
on
is Hamiltonian if it is given by an equivariant moment map
.
Theorem 1.6.
Let
be a circle subgroup of a compact Lie group
which acts effectively on a connected symplectic manifold
. If
has at most twofold isotropy and if there is no
which reverses
, then
is essential in
.
Example 1.7.
This theorem does not extend to circle actions which have at most threefold isotropy. For example, the action of
on
given by
is inessential in
. However, since
and
are not diffeomorphic, this action has no reversor.
We also need the symplectic hypothesis. To see this, consider the obvious action of
on
. The subgroup
acts with at most twofold isotropy but has no reversor.
Remark 1.8.
If
is a simple group, we do not need to assume that
is symplectic in Theorem 1.6 ; we only need to assume that there exists a point
which is fixed by a maximal torus containing
but is not fixed by all of
. Note that, in contrast, in the example above, the only points fixed by
are fixed by all of
.
Remark 1.9.
If
is any circle subgroup of
– or indeed a subgroup of any simple group of type
,
, or
– then there exists a
which reverses
. In this case, Theorem 1.6 is trivial and the force of Theorem 1.4 is that we can choose
so that it also fixes
.
Remark 1.10.
In the proof of the above theorems, we pick a maximal torus
which contains
. The reversor
that we construct lies in the normalizer
and has the property that
lies in
. However, as we saw in Example 1.5 ,
may not be equal to the identity.
Theorems 1.4 and 1.6 have the following easy corollaries:
Corollary 1.11.
Consider a Hamiltonian circle action
on a closed symplectic manifold
with moment map
, normalized so that
. If
is a semifree fixed component, then
is essential in every compact subgroup
that contains it, unless there is a symplectomorphism
of
that fixes
and reverses
. In this case, all the following hold:
-
(1)
for all
.
-
(2)
There is a one-to-one correspondence between the positive weights at
and the negative weights at
, and vice versa.
-
(3)
induces an isomorphism on the image of the restriction map in equivariant cohomology
.
-
(4)
. In particular,
-
(a)
.
-
(b)
The sum of the weights at
is zero.
Corollary 1.12.
Consider a Hamiltonian circle action
on a closed symplectic manifold
with moment map
, normalized so that
. If the action has at most twofold isotropy, then
is essential in every compact subgroup
that contains it, unless there is a symplectomorphism
of
that reverses
. In this case, all the following hold:
-
(1)
for all
.
-
(2)
There is a one-to-one correspondence between the positive weights at
and the negative weights at
, and vice versa.
-
(3)
induces an isomorphism on the image of the restriction map in equivariant cohomology
.
It is unknown whether the existence of such
is necessary for
to be inessential in
. We make partial progress towards answering this question in [2] .
All the results in this paper are proved by a case by case study of the structure of semisimple Lie algebras.
2 Coadjoint orbits
In this section, we prove Proposition 1.3 . We begin with a brief review of a few facts about Lie groups.
Each simply connected compact semisimple Lie group is a product of simple factors, and its center is the product of the centers of its simple factors. Moreover, since its Lie algebra splits into a corresponding sum, the coadjoint orbits also are products of coadjoint orbits of simple groups.
Therefore, we may assume that
is simple.
Let
be a compact simple Lie group. Let
denote the universal cover of
, and
denote the quotient of
by its center. Let
denote the Lie algebra of
, and let
denote the Lie algebra of a maximal torus
.
Let
,
, and
be the lattices consisting of vectors
whose exponential is the identity in
,
, and
, respectively. There is a one-to-one correspondence between
and circle subgroup of
,
and circle subgroups of
, and
and circle subgroups of
, given by sending
to
. Note that
. Because
is simply connected,
Let
denote the dual to
, and let
denote the set of roots of
, i.e. the nonzero weights of the adjoint action
on
, where
is the complexification of
. The lattice
is dual to the lattice in
generated by the roots, i.e.
precisely when
for all
. If we use the Killing form
to identify
and
, then
is generated by the set
Further the set of weights at any fixed point
for the action of
on
is a nonempty subset of the set of roots. Therefore the result will follow if we find a representative
for each nontrivial class in
such that
for every
.
We will check this on a case by case basis; in each case we will use the Killing form to identify
and
. Let
be the standard metric on
with the standard basis
, and define
(I) For the group
, where
,
and the roots are
for
. Hence
As representatives for the quotient
, we take
for
.
(II) For the group
, where
,
and the roots are
and
for
. Hence
As representatives of the quotient
, we take
and
.
(III) For the group
, where
,
and the roots are
and
for
. Hence
As representatives of the quotient
, we take
and
.
(IV) For the group
, where
,
and the roots are
for
. Hence
The quotient
is isomorphic to
if
is even, and to
if
is odd. Either way, as representatives of
, we take
,
,
and
.
(V, a) For the group
,
and the roots are
,
, and
for
and
distinct, where
. Hence,
As representatives of the quotient
, we take
,
, and
. (V, b) For the group
,
, and the roots are
, and
for
and
distinct. Hence
As representatives for the quotient
, we take
and
.
Every group of type
,
, and
is simply connected, so no further argument is necessary.
3 Lie Group Actions
This section contains proofs of Theorems 1.4 and 1.6 . We begin by stating a lemma about root systems, that is proved at the end. We shall always assume that the positive Weyl chamber is closed.
Lemma 3.1.
Let
be a simply connected compact simple Lie group. Let
be the Lie algebra of a maximal torus
. Let
be the integral lattice, let
denote the set of roots, and let
denote the Weyl group. Use the Killing form to identify
and
.
Fix
. Choose a positive Weyl chamber which contains
. Let
denote the highest root. Then the following claims hold:
(a) If
, then there exist orthogonal roots
so that
and so that
for all
.
(b) Let
be a set of roots which contains every root
such that
or
is also a root. Assume also that
is closed under addition, that is, it contains every root which can be written as the sum of roots in
. Then
contains all roots.
(c) If
and
is not an element of the Weyl group, then for every nonzero weight
there exists
so that
.
(d) If
is not an element of the Weyl group, then
is the only root which lies in the positive Weyl chamber.
Using this result, we can find find elements which reverse certain circle subgroups of simply connected compact simple Lie groups. Note that because
is simply connected, every circle subgroup of
is inessential in
.
Lemma 3.2.
Let
be a circle subgroup of a simply connected compact simple Lie group
.
(i) Let
be a nontrivial representation of
. If
acts semifreely on
then there exists
that reverses
.
(ii) Let
be a proper subgroup containing
. If the adjoint action of
on
is semifree, then there exists
that reverses
.
(iii) Let
be a proper subgroup containing a maximal torus which contains
. If the natural action of
on
has at most twofold isotropy, then there exists
that reverses
.
The assumption in (ii) above is a special case of (i) since the representation
is restricted; however, the conclusion is stronger since it asserts that the reversor lies in
. Statement (ii) and (iii) are also related: the former makes a strong assumption about the action induced by
on the tangent space to
at the fixed point
, the latter makes a weaker assumption about the action at all the fixed points on
.
We will now use the claims in Lemma 3.1 to prove Lemma 3.2 . Let
be a maximal torus which contains
. Let
denote the integral lattice. Let
be the vector corresponding to
. Choose a positive Weyl chamber which contains
. Let
denote the highest root.
Recall that the Weyl group
is the quotient
, where
is the the normalizer of
in
. Every root
gives rise to an element
whose action on
is given by
.
Proof of Lemma 3.2 (i). Let
be a nontrivial representation of
. Assume that
acts semifreely on
.
Suppose first that
. By claim (a), there exist orthogonal roots
so that
. Since the roots are orthogonal, for each
the associated element of the Weyl group
takes
to
and leaves
fixed for all
. Hence, their product
takes
to
, and so reverses
.
So assume instead that
. If
is in the Weyl group, then statement (i) is trivial. So we assume that it is not. Let
act on
via restriction, and pick any nonzero weight
in the weight decomposition.
By claim (c), we can find some
such that
Since
also appears in the weight decomposition, this contradicts the assumption that the action of
on
is semifree.
Proof of Lemma 3.2 (ii). Let
be a proper subgroup which contains
. Assume that the adjoint action of
on
is semifree. Let
be the set of roots
so that the associated weight space
lies in
.
Clearly, if
, then
.
Suppose first that
. By claim (a), there exist orthogonal roots
so that
and so that
for every
. Since
,
lies in
for all
. Hence, the associated element of the Weyl group
lies in
for all
. Thus
must lie in
.
So assume instead that
. We see immediately that
and
lie in
. If
,
and
are all roots, then
. Hence, since
is closed under Lie bracket, if
and
are in
then
also, that is,
is closed under addition. Additionally, if
and
are roots such that
, then either
or
. If the former holds, then
and
lie in
. Since
is closed under addition, so do
and
.
The other case is identical. Thus, claim (b) implies that every root lies in
. This contradicts the claim that
is a proper subgroup.
Proof of Lemma 3.2 (iii). Let
be a proper subgroup which contains the maximal torus
, and assume that the natural action of
on
has at most twofold isotropy.
If
, then part (iii) follows by the argument used to prove part (i). So assume that
. We may also assume that
does not lie in the Weyl group, because otherwise the claim is trivial. Since
is proper, there exists at least one root
so that the associated weight space
is not contained in
. Then there is
so that the root
lies in the positive Weyl chamber. Hence by (d)
, and so
Choose
which descends to
. Then
is a fixed point for
, and
is one of the weights for
at this fixed point. This contradicts the fact that the action has at most twofold isotropy.
We are now ready to deduce Theorems 1.4 and 1.6 . In both cases, we will do this by proving the contrapositive, that is, we will assume that
is an inessential circle subgroup and use this to construct a reversor.
Let
denote the universal cover of
. Then
is the direct product of a compact simply connected semisimple Lie group and a vector space. Since
is inessential, it lifts to a circle subgroup of
. Since this lift must lie in the compact part of
, we may assume without loss of generality that
is a compact simply connected semisimple Lie group.
In fact, it is enough to prove these claims for the universal cover of
, as long as we no longer insist on an effective action but instead allow a finite number of elements of the group to act trivially on
. Thus we may assume that
is the product of compact simple and simply connected groups
. Let
be the projection of
to
. Without loss of generality, we may assume that
for all
.
Proof of Theorem 1.4 . Let
as above. Choose
and let
be the stabilizer of
. Then
. There exists a representation
of
, called the isotropy representation, so that a neighborhood of the
-orbit through
is equivariantly diffeomorphic to a neighborhood of the zero section of
. Fix some simple factor
, and let
.
Assume first that
is a proper subgroup. Note that
is invariant under the action of
. Thus, since
acts semifreely on
via the adjoint action,
acts semifreely on
. Thus, by Lemma 3.2 (ii) there exists an element
that reverses
.
So assume on the contrary that
. Let
be the projection of
onto the product of all the simple factors except
. Since
and
, we must have
. Hence
acts on
. For any integer
, let
denote the subspace of
on which
acts with weight
. Since
commutes with
,
is a representation of
. Since only a finite number of elements of
act trivially on
,
must act nontrivially on
, and hence also on
. Therefore, there is some
so that the representation of
on
is nontrivial. Because
is simple,
must act with both positive and negative weights on
. But the weights for the action of
on
are the weights for the action of
shifted by
. Hence, because
is a semifree fixed point component,
and the action of
on
is itself semifree.
Therefore by Lemma 3.2 (i) there exists
that reverses
.
Since
reverses
for each
,
reverses
, as required.
Moreover, since
(in general they are not equal),
lies in
, and hence fixes
.
Proof of Theorem 1.6 . Fix some simple factor
. Let
be the Weyl group of
. Let
be a maximal torus of
containing
. Let
be the moment map for the
-action. Pick any
so that the one parameter subgroup generated by
is dense in
. Let
be any point which maps to the minimum value of
, the component of
in the direction
. By construction,
is a fixed point for
. Assume first that
, that is, the function
is nonnegative on
. Since the moment polytope
is invariant under the Weyl group
, this implies that
is also nonnegative on
for all
. Because
is simple and
is a generic point of
, for any nonzero
there exists an element
such that
. Applying this to
, we see that
must be the single point
, which is impossible, because the action is effective.
Therefore,
.
Now let us reconsider the action of
on
. Let
be the stabilizer of
in
, and let
. Since
acts with at most twofold isotropy on
,
acts with at most twofold isotropy on
. Since
is not zero, the stabilizer of
in
. is a proper subgroup of
. Since
is equivariant, this implies that
is a proper subgroup of
. By Lemma 3.2 (iii), this implies that there exists
which reverses
. Then
reverses
.
Proof of Lemma 3.1 . We now prove claims (a)-(d) on a case by case basis, using the classification of compact simple Lie groups. We will use the notation of § 2 . Note, however, that here
since
is simply connected.
(I) Recall that for the group
, where
,
, the roots are
for
, and the integral lattice is
. The positive Weyl chamber is1
The highest root is
.
If
, then
for all
. Since
and
for all
, there are an equal number of
's and
's, and the rest are
's. Hence,
is the sum of orthogonal roots of the form
.
Since
, this proves claim (a).
Since
, the roots
and
lie in
for all
. If neither
nor
is equal equal to
, then
is also in
. This proves claim (b).
We now prove (c). The weight lattice is
.
By permuting the coordinates of
, we may assume
. Since
, there exists
such that
; since this difference lies in
, it must be at least
. Since
and
,
. Therefore, either
or
. In the former case,
In the latter case, let
be obtained from
by the permutation which reverses the coordinates, so that
. Then
The only facts we have used are that
, that the Weyl group contains the permutation group
, and that
for any
.
Finally,
is the only root in the positive Weyl chamber.
(II) Recall that for the group
, where
,
, the roots are
and
for
, and the integral lattice is
The positive Weyl chamber is
. The highest root is
.
If
, then either
and
for all
, or
for all
. Either way, since
, we can write
as the sum of orthogonal roots
such that
.
Since
, the roots
and
lie in
for
or
. Since
, the roots
and
lie in
. Every root can be written as a sum of these roots.
Since
lies in the Weyl group, we are done.
(III) Recall that for the group
, where
,
, the roots are
and
for
, and the integral lattice is
. The positive Weyl chamber is
. The highest root is
.
If
, then
for all
. Since
, we can write
as half the sum of orthogonal roots of the form
. Note that
.
Since
, the roots
lie in
for
.
Every root can be written as a sum of these roots.
Since
lies in the Weyl group, we are done.
(IV) Recall that for the group
, where
,
, the roots are
for
, and the integral lattice is
The positive Weyl chamber is
. The highest root is
.
If
, then either
and
for all
, or
for all
. Either way, since
, we can write
as the sum of orthogonal roots
such that
. Since
, the roots
and
lie in
for
or
. Every root can be written as a sum of these roots.
Now assume that
. Consider a nonzero weight
By applying the Weyl group, we may assume
lies in the positive Weyl chamber. Since
also lies in the positive Weyl chamber,
for all
. Moreover, since
, and
,
. Therefore,
(Here, we have used that
.) Since
is nonzero, either
, or
. In either case,
. (In the first case, we use the fact that
and
implies that
.) Therefore,
, This proves claim (c).
Finally,
is the only root in the positive Weyl chamber.
(V, a) Recall that for the group
,
and the roots are
,
, and
for
and
distinct, where
.
Therefore
The positive Weyl chamber is
(Note that these conditions imply
.) The highest root is
.
Write
, where
. Assume that
. Combining the inequalities
,
,
, and
, we see that
. Since also
,
, and
, we have
. Moreover, in both cases, if the final inequality in the sentence is an equality, so are all the preceding ones. Since
,
. Since
,
or
. In the former case,
, so
, so
. Hence,
In the latter case,
is odd, so
implies that
. In this case,
. This proves claim (a).
Since
the roots
and
lie in
for all
. Moreover,
, so the roots
and
lie in
for all
.
Since, for example,
, it follows easily that
contains all roots.
Let
be a nonzero weight. Write
as before. By applying the Weyl group, we may assume that
is in the positive Weyl chamber. Since
, it is enough to show that
.
This fact now follows from the argument from
, since
, since the Weyl group contains the permutation group
, and since
must satisfy
.
Finally,
is the only root in the positive Weyl chamber.
(V, b) Recall that for the group
,
, the roots are
and
for
and
distinct, and the integral lattice is
The positive Weyl chamber is
(Note that this automatically implies that
.) The highest root is
.
Assume that
. Combining the inequalities
,
, and
for
and
, we see that
.
Since also
and
for
and
,
.
Moreover, in both cases, if the last inequality in the sentence is an equality, all the inequalities are equalities. Since
, the only possibilities are
or
In the former case, we must have
. In the latter case, the only possibilities are
, or
. The proves claim (a).
Since
, the roots
and
lie in
for all
Since
, the roots
and
also lie in
for all
. All roots can be written as a sum of these roots. This proves claim (b).
Since
,
for every
. Hence, the argument for claim (c) follows from the argument for
.
Finally,
is the only root in the positive Weyl chamber.
(V, c) For the group
,
} and the roots are
, and
for
and
distinct. Hence the integral lattice is
The positive Weyl chamber is
(Note that these conditions imply that
.) The highest root is
.
Assume that
. Combining the inequalities
we see that
. Since
and
,
.
Moreover, in both cases, if the last inequality in the sentence is an equality, all the inequalities are equalities. Since
, the only possibilities are
or
. In the former case,
. In the latter case,
. Claim (a) follows.
We now notice that
for all
and
. Therefore, the corresponding roots
,
,
, and
all lie in
.
Since every root can be written as a sum of these roots, claim (b) follows.
Since
,
for every
. Hence, the argument for claim (c) carries over from the argument for the group
.
Finally,
is the only root in the positive Weyl chamber.
(VI) For the group
,
. The roots are
,
for
, and
. Hence the integral lattice is
.
The positive Weyl chamber is
(Note that automatically
.) The highest root is
.
The argument for claim (a) carries over word for word from the argument for
.
Notice that if
or
| |
| |
| |
Hence, the corresponding roots all lie in
. Since every root can be written as the sum of these roots, this proves claim (b).
Since
lies in the Weyl group, we are done.
(VII) For the group
,
}. The roots are
and
for
and
distinct. The positive Weyl chamber is
. (Note that automatically
.) The integral lattice is
. The highest root is
.
The argument for claim (a) follows the argument for
word for word.
Since
, the roots
,
,
and
all lie in
. Since every root can be written as a sum of these roots, claim (b) follows.
Since
lies in the Weyl group, we are done.
References
-
D. McDuff and J. Slimowitz, Hofer–Zehnder capacity and length minimizing Hamiltonian paths, SG/0101085, Geom. Topol. 5 (2001), 799–830.
-
D. McDuff and S. Tolman, Topological properties of Hamiltonian circle actions, SG/0404338.
-
A. Vina, On the Homotopy of symplectomorphism groups of Homogeneous spaces, SG/0305407.
-
A. Weinstein, Cohomology of symplectomorphism groups and critical values of Hamiltonians, Math Z. 201 (1989), 75–82.
Department of Mathematics, Stony Brook University, Stony Brook, NY 11794-3651, USA E-mail address : dusa@math.sunysb.edu Department of Mathematics, University of Illinois at Urbana–Champaign, IL, USA E-mail address : tolman@math.uiuc.edu