Gelfand–Cetlin abelianizations of symplectic quotients
Abstract.
We show that generic symplectic quotients of a Hamiltonian -space by the action of a compact connected Lie group are also symplectic quotients of the same manifold by a compact torus. The torus action in question arises from certain integrable systems on , the dual of the Lie algebra of . Examples of such integrable systems include the Gelfand–Cetlin systems of Guillemin–Sternberg in the case of unitary and special orthogonal groups, and certain integrable systems constructed for all compact connected Lie groups by Hoffman–Lane. Our abelianization result holds for smooth quotients, and more generally for quotients which are stratified symplectic spaces in the sense of Sjamaar–Lerman.
Key words and phrases:
symplectic quotient, Gelfand–Cetlin system, stratified symplectic space1991 Mathematics Subject Classification:
53D20 (primary); 17B80 (secondary)1. Introduction
Let be a compact connected Lie group with Lie algebra . Suppose that is a Hamiltonian -space, i.e. a symplectic manifold equipped with a symplectic action of and equivariant moment map The symplectic or Marsden–Weinstein [9] quotient of by at level is the topological space
where is the -stabilizer of If acts freely on then is a smooth symplectic manifold. In the absence of this freeness assumption, is a stratified symplectic space in the sense of Sjamaar-Lerman [10].
The purpose of this paper is to show that certain integrable systems on allow us to express generic symplectic quotients of a Hamiltonian -space as symplectic quotients of the same manifold by the action of a compact torus. Such integrable systems include the Gelfand–Cetlin systems constructed by Guillemin–Sternberg [4, 5] for unitary and special orthogonal groups, as well Hoffman–Lane’s more recent generalizations of Gelfand–Cetlin systems [8] to arbitrary Lie type.
An example of our main result arises in classical mechanics [3]. Suppose that we are given a Hamiltonian -space with an invariant Hamiltonian function . The -action gives rise to two Poisson-commuting conserved quantities: the total angular momentum, and the angular momentum in some fixed direction in the Lie algebra of corresponding to a choice of maximal torus. These quantities give the components of a moment map for a densely defined 2-torus action on coming from the Gelfand–Cetlin system of Guillemin–Sternberg for the case of .111The square of the total angular momentum is a smooth function, but the orbits of its Hamiltonian flow do not have constant period, and so it does not generate a circle action. Taking the square root gives a function whose Hamiltonian flow generates a circle action, but which is only continuous, not differentiable, at zero. It therefore does not define a Hamiltonian flow at zero. Our main result shows that for a non-zero value of the angular momentum, the symplectic quotient coincides with an appropriate symplectic quotient of under the densely defined torus action. See [3] for more examples of these techniques.
1.1. Main result
We introduce the notion of a Gelfand–Cetlin datum in Definition 1. This amounts to being a continuous map on that restricts to a Poisson moment map for a Hamiltonian action of a compact torus on an open dense subset , along with some extra conditions that capture salient properties of the classical Gelfand–Cetlin systems. One of these conditions is that the open, symplectic submanifold be a Hamiltonian -space with moment map for any Hamiltonian -space with moment map . The results of Guillemin–Sternberg [4, 5] imply that Gelfand–Cetlin data exist for all unitary and special orthogonal groups, while more recent results of Hoffman–Lane [8] imply that such data exist in all Lie types.
The following is the main result of our paper.
Theorem.
Let be a compact connected Lie group, and a Hamiltonian -space with moment map . Suppose that is a Gelfand–Cetlin datum, and consider a point .
-
(i)
The torus acts freely on if and only if acts freely on . In this case, there is a canonical symplectomorphism .
-
(ii)
There is a canonical isomorphism of stratified symplectic spaces.
Part (ii) is strictly more general than (i). Part (i) is included for the sake of exposition and accessibility.
One may regard this theorem as an approach to abelianizing the generic symplectic quotients of a Hamiltonian -space , i.e. to presenting such quotients as symplectic quotients by a compact torus. An alternative approach to abelianization is pursued in the work of Guillemin–Jeffrey–Sjamaar [2].
1.2. Organization
Section 2 briefly establishes some of our conventions concerning Lie theory and Hamiltonian geometry. Section 3 subsequently motivates and contextualizes the notion of a Gelfand–Cetlin datum. Our main result is then proved in Section 4 for smooth quotients. A generalization to stratified symplectic spaces is formulated and proved in Section 5.
Acknowledgements
The authors would like to thank Megumi Harada and Jeremy Lane for exceedingly useful conversations. P.C. acknowledges support from a Utah State University startup grant, while J.W. acknowledges support from Simons Collaboration Grant # 579801.
2. Background and conventions
This section establishes some of our notation and conventions regarding Lie theory and Hamiltonian geometry.
2.1. Tori
The Lie algebra of the unitary group is the real vector space of purely imaginary numbers. We will identify this vector space with in the obvious way. It follows that is the Lie algebra of for all non-negative integers , and that
is the exponential map for . In certain contexts, we will implicitly use the dot product to regard as the dual of the Lie algebra of .
2.2. General compact connected Lie groups
Let be a compact connected Lie group with Lie algebra and rank . One has the adjoint representations and , as well as the coadjoint representations and . The -representations induce -actions on and , and thereby give rise to stabilizer subgroups
of for all and . On the other hand, the -representations allow us to define centralizers
for all and . It follows that (resp. ) is the Lie algebra of (resp. ). Let us also note that and for all and . The regular loci
are open, dense, -invariant subsets of and , respectively. An element (resp. ) then belongs to (resp. ) if and only if (resp. ) is a Cartan subalgebra of . This is equivalent to (resp. ) being a maximal torus of .
A few remarks on maximal tori and and Cartan subalgebras are warranted. Let be a Cartan subalgebra, and write for the maximal torus with Lie algebra . The exponential map then restricts to a surjective homomorphism of abelian groups. The kernel of the latter is a free -submodule of with rank equal to . It follows that the same is true of
2.3. Hamiltonian geometry
Let be a Poisson manifold, i.e. is a Poisson bivector field on the manifold . Note that may be regarded as a skew-symmetric bilnear map from two copies of to the trivial rank- vector bundle over . Contracting with cotangent vectors in the first argument then determines a vector bundle morphism . One calls non-degenerate if is an isomorphism. In this case, for a unique symplectic form , where is the vector bundle morphism obtained by contracting with tangent vectors in the first argument. This process gives rise to a bijective correspondence between symplectic structures on and non-degenerate Poisson structures on . We will thereby make no distinction between symplectic manifolds and non-degenerate Poisson manifolds.
If is a Poisson manifold, then can be recovered from the Poisson bracket that it induces. This bracket associates to smooth functions the smooth function
At the same time, one defines the Hamiltonian vector field of a smooth function by . It follows that
for all smooth functions .
Now let be a compact connected Lie group with Lie algebra and exponential map . If acts smoothly on a manifold , then each determines a generating vector field by
for all . A Poisson manifold with a smooth -action will be called a Poisson Hamiltonian -space if is -invariant and comes equipped with a moment map. This last term refers to -equivariant smooth map satisfying for all , where is the result of pairing with pointwise. We will reserve the term Hamiltonian -space for a Poisson Hamiltonian -space whose underlying Poisson structure is symplectic.
It will be advantageous to recall the Poisson Hamiltonian -space structure on . The Poisson bracket on is given by
for all smooth functions and points , where denote the differentials of at , respectively. One finds that is a Poisson Hamiltonian -space with respect to the coadjoint action, and with the identity serving as the Poisson moment map.
3. Gelfand–Cetlin data
In this section, we define Gelfand–Cetlin data and introduce their main properties. This begins with the definition itself in 3.1. The existence of Gelfand–Cetlin data is addressed in 3.2, while concrete techniques for constructing such data are discussed in 3.3 and 3.4. In 3.5, we describe concrete Gelfand–Cetlin data for unitary groups.
3.1. Definition and relation to integrable systems
Let be a compact connected Lie group with Lie algebra and rank . Consider the quantities
and introduce the following tori of small, intermediate, and big ranks:
The respective Lie algebras of these tori are
Definition 1.
A Gelfand–Cetlin datum is a pair , consisting of a continuous map and open dense subset that satisfy the following conditions:
-
(i)
are -invariant on and smooth on ;
-
(ii)
is a -basis of the lattice for all ;
-
(iii)
;
-
(iv)
is a smooth submersion and moment map for a Poisson Hamiltonian -space structure on ;
-
(v)
is a principal -bundle;
-
(vi)
if is a Hamiltonian -space with moment map , then
is a moment map for a Hamiltonian -space structure on .
In this case, we adopt the notation
We also refer to the elements of as the strongly regular elements of .
Remark 2.
It is instructive to consider this definition in relation to the theory of completely integrable systems. One is thereby led to the following result.
Proposition 3.
Let be a Gelfand–Cetlin datum. If is a coadjoint orbit and , then
is a completely integrable system, principal -bundle, and moment map for a Hamiltonian action of on .
Proof.
Note that the Hamiltonian vector field of any smooth function is tangent to . It follows that is stable under the action of on . Definition 1(iv) now implies that is a Hamiltonian -space with moment map . We conclude that is a moment map for the Hamiltonian action of on .
Since is constant-valued on , Definition 1(v) tells us that is a principal -bundle. It therefore remains only to prove that . This follows immediately from the fact that . ∎
3.2. Existence of Gelfand–Cetlin data
It is natural to wonder about the generality in which Gelfand–Cetlin data exist. The earliest constructions are due to Guillemin–Sternberg [4, 5], and apply to all unitary groups and special orthogonal groups . The underlying techniques are based on Thimm’s method, as described in [5]. Further details are outlined in Sections 3.3–3.5 of this paper. The case of symplectic groups is considerably more subtle and addressed in Harada’s paper [6].
Some recent work of Hoffman–Lane [8] implies the existence of Gelfand–Cetlin data for an arbitrary compact connected Lie group ; the reader is referred to [8, Section 6.2] for the relevant details. This Hoffman–Lane paper is part of a broader program aimed at generalizing the results of Harada–Kaveh [7].
3.3. Construction of Gelfand–Cetlin data: integrality
We now discuss the construction of functions satisfying Conditions (i) and (ii) in Definition 1. Let be a compact connected Lie group with Lie algebra and rank . Choose a -invariant inner product on , a Cartan subalgebra , and a closed, fundamental Weyl chamber . The chamber is known to be a fundamental domain for the adjoint action of on . Our inner product thereby identifies with a fundamental domain for the -action on . We may therefore define a continuous surjection by the property that for all , where is the coadjoint orbit of . One sometimes calls the sweeping map on with respect to ; its fibers are exactly the coadjoint orbits of , and is the interior of . One also finds that the commutative diagram
is Cartesian. The left vertical map
is a easily seen to be a smooth, surjective submersion.
Now choose a -basis of the -submodule . Note that the pairing between and allows one to regard as functions on . With this in mind, each determines a function
The previous paragraph implies that are smooth on , while being -invariant and continuous as functions on . Given any , the differentials may be described as follows.
Proposition 4.
If , then is a -basis of the lattice .
Proof.
Choose for which . Since each function is -invariant, one has
for all . We also observe that restricts to a -module isomorphism . It therefore suffices to take and prove that for all .
Assume that , and fix . Our invariant inner product allows us to regard as a subspace of . We then note that , and that is contained in the kernel of . Let us also note that is the annihilator of in , and as such is contained in the kernel of . These last two sentences reduce us to proving that for all . On the other hand, we have . It therefore suffices to prove that for all . But this is an immediate consequence of the following two observations: is an open subset of , and for all . ∎
3.4. Construction of Gelfand–Cetlin data: Thimm’s method
Retain the objects and notation discussed in Section 3.3. Let be a descending filtration of by connected closed subgroups with respective Lie algebras . Let us also choose a Cartan subalgebra and closed, fundamental Weyl chamber for each . Our -invariant inner product on gives rise to a -module isomorphism , by means of which and correspond to subsets and of . As in Section 3.3, one may define a continuous surjection by the condition that for all .
Let be the ranks of , respectively. Let us also choose a -basis of the lattice for each . As in Section 3.3, we define the functions
for . The same section implies that are -invariant and continuous on , as well as smooth on . We also have the following equivalent version of Proposition 4.
Proposition 5.
If and , then is a -basis of the lattice .
Let denote the transpose of the inclusion for each . Consider the functions on defined by
for and . It will be convenient to enumerate these functions as
(3.1) |
where .
The discussion preceding Proposition 5 implies that is smooth on the open subset
Let us also define
These last few sentences give context for the following consequence of [4, Theorem 3.4].
Proposition 6.
Let and be as defined above.
-
(i)
The restriction is a moment map for a Poisson Hamiltonian -space structure on .
-
(ii)
If is a Hamiltonian -space with moment map , then
is a moment map for a Hamiltonian -space structure on .
This result has the following immediate connection to Definition 1: the pair is a Gelfand–Cetlin datum if and only if , is dense in , and is a principal -bundle. Guillemin–Sternberg [4, 5] explicitly show these conditions to be achievable for and . Our next section outlines the details of the Guillemin–Sternberg construction for .
3.5. Example of Gelfand–Cetlin datum: the Gelfand-Cetlin system on
Fix a positive integer . Consider the Lie group of unitary matrices, and its Lie algebra of skew-Hermitian matrices. Let us also consider the real -module of Hermitian matrices. In what follows, we will freely identify with by means of the non-degenerate, -invariant bilinear form
(3.2) |
Given an integer , define the subgroup
The descending chain then induces such a chain on the level of Lie algebras. We have
for all , where the second equation implicitly uses (3.2). The transpose of then sends to the submatrix in the bottom right-hand corner of .
Now note that
is a Cartan subalgebra for each . Let be the result of setting and for , noting that is a -basis of . At the same time, consider the fundamental Weyl chamber
for each . Under the pairing (3.2), corresponds to the cone
We then have sweeping maps and compositions for and , as in Section 3.4. The functions
from Section 3.4 are therefore given by the following condition: if and , then are the eigenvalues of the submatrix in the bottom right-hand corner of .
Observe that the number of maps is
Our enumeration (3.1) therefore takes the form
Let us also consider the open dense subset
of . By the paragraph following the proof of Proposition 6, is a Gelfand–Cetlin datum if and only if is a principal bundle for . This later condition is verified in [4, Section 5].
4. The abelianization theorem
This section is devoted to the proof of our abelianization theorem for smooth quotients. Some preliminary results are established in 4.1 and 4.2, while the main proof appears in 4.3.
4.1. The universal maximal torus
Adopt the notation and conventions in Section 3.1, and let be a Gelfand–Cetlin datum. Given any , Definition 1(ii) tells us that is a basis of . This basis determines a vector space isomorphism
The torus is then a universal maximal torus in the following sense.
Proposition 7.
If , then integrates to a Lie group isomorphism .
Proof.
It suffices to prove that restricts to a -module isomorphism from the kernel of to . This is an immediate consequence of Proposition 4. ∎
Proposition 8.
Let be a Hamiltonian -space with moment map . Suppose that . We then have for all and , where the left and right-hand sides denote the actions of on and on , respectively.
Proof.
Let be the generating vector field on determined by via the action of on . Write for the generating vector field on determined by through the action of on . It suffices to prove that for all and . Setting and letting denote the standard basis vector, this is equivalent to establishing that for all and . On the other hand, (resp. ) is the Hamiltonian vector field on (resp. ) associated to (resp. the component of ). This further reduces us to proving that . But it is clear that
for all and . ∎
4.2. Some supplementary results
We now prove two supplementary facts needed to establish the main result of this paper. We continue with the notation and conventions of Sections 3.1 and 4.1.
Proposition 9.
Let be a Hamiltonian -space with moment map . Suppose that .
-
(i)
If and satisfy , then .
-
(ii)
The saturation of under the action of on is .
Proof.
To verify (i), let and be such that . Let us also observe that is -equivariant when restricted to a map . These last two sentences imply that . Since acts freely on , we must have .
We now verify (ii). To this end, note that is a -invariant subset of that contains . This implies that the saturation of is contained in . For the opposite inclusion, suppose that . Definition 1(v) tells us that for some . By the equivariance property of mentioned in the previous paragraph, we must have . This completes the proof of (ii). ∎
Fix . In light of the previous proposition, we may define the map
The following result is immediate consequence of the previous proposition.
Corollary 10.
If , then is a homeomorphism.
4.3. Proof of the abelianization theorem
Theorem 11.
Let be a Hamiltonian -space with moment map . Suppose that .
-
(i)
The stabilizer acts freely on if and only if acts freely on .
-
(ii)
In the case of (i), there is a canonical symplectomorphism .
Proof.
We begin by verifying (i). In light of Proposition 7, the multiplication map
is a Lie group isomorphism. We also note that the action of on and multiplication action of on itself define an action on . By Proposition 8, the homeomorphism is equivariant in the following sense:
for all and . It follows that acts freely on if and only if acts freely on .
We now prove (ii). By Corollary 10, the inclusion descends to a diffeomorphism
We also note that the -action on induces a residual action of the subtorus on . Proposition 8 then tells us that
for all and . The map therefore descends to a diffeomorphism
It therefore suffices to prove that pulls the symplectic form on back to the symplectic form on .
We have a commutative diagram
where and are the canonical quotient maps and is the inclusion. We also have inclusion maps and . Another consideration is that (resp. ) is the unique -form on (resp. ) for which (resp. ), where is the symplectic form on . It therefore suffices to prove that . On the other hand, our commutative diagram implies that
This completes the proof. ∎
5. Generalization to stratified symplectic spaces
We now provide a generalization of Theorem 11 in the realm of stratified symplectic spaces [10]. In 5.1, we recall the immediately pertinent parts of Sjamaar and Lerman’s more general theory of stratified symplectic spaces. The generalization of Theorem 11 to stratified symplectic spaces appears in 5.2.
5.1. Stratified symplectic spaces
Let be a topological space on which a compact torus acts continuously. Given a closed subgroup , let
be the locus of points with -stabilizer equal to . Denote by the set of all closed subgroups for which .
Now let be a compact connected Lie group with Lie algebra . Suppose that is a Hamiltonian -space with moment map . As discussed in the introduction to this paper, is a stratified symplectic space [10] for all . This means that is naturally partitioned into symplectic manifolds satisfying certain compatibility conditions. While we refer the reader to [10, Definition 1.12] for a precise definition and description of stratified symplectic spaces, the following exposition will be sufficient for our purposes.
Fix a point , and recall that is a maximal torus. Adopt the more parsimonious notation
and note that is the disjoint union
The arguments in the proof of [10, Theorem 2.1] imply that each subset is a locally closed, -invariant submanifold of . These arguments also imply that the topological quotient carries a unique manifold structure for which the canonical map is a surjective submersion. One further consequence of [10, Theorem 2.1] is the existence of a symplectic form on such that is the pullback of along the inclusion . It follows that is a disjoint union
(5.1) |
of symplectic manifolds, called the symplectic strata of .
Remark 12.
The quotients need not be manifolds in the traditional sense of the term; each may have connected components of different dimensions. To obtain a stratification into genuine symplectic manifolds, one must refine (5.1) and declare the symplectic strata to be the connected components of the quotients . The distinction between (5.1) and this refined stratification will not materially affect any argument in this paper.
Definition 13.
Let and be compact connected Lie groups with respective Lie algebras and . Suppose that (resp. ) is a Hamiltonian -space (resp. Hamiltonian -space) with moment map (resp. ). Take and . A pair of maps and will be called an isomorphism of stratified symplectic spaces if the following conditions are satisfied:
-
(i)
is a homeomorphism;
-
(ii)
is a bijection;
-
(iii)
restricts to a symplectomorphism for each .
Remark 14.
Assume that this definition is satisfied. Equip and with the refined stratifications discussed in Remark 12. By (ii) and (iii), the association defines a bijection from the set of symplectic strata to the set of symplectic strata in . Property (i) implies that this bijection is an isomorphism of partially ordered sets, i.e. any symplectic strata satisfying must also satisfy . We also know that restricts to a symplectomorphism for all symplectic strata , as follows from (iii). In other words, an isomorphism in the sense of Definition 13 gives rise to an isomorphism between the refined symplectic stratifications on and .
5.2. A more general abelianization theorem
Let us continue with the notation and conventions set in Section 4, as well as those in Section 5.1 concerning stratified symplectic spaces.
In preparation for our next proposition, we encourage the reader to recall Proposition 7 and Corollary 10.
Proposition 15.
Let be a Hamiltonian -space with moment map . Suppose that .
-
(i)
The association defines a bijection .
-
(ii)
If is a closed subgroup, then restricts to a diffeomorphism
Proof.
As in the proof of Theorem 11(i), we have
for all and . It follows that defines a bijection
and that restricts to a homeomorphism
for all closed subgroups . On the other hand, we clearly have a bijection
We also note that and for . These last three sentences combine to imply the desired results. ∎
The following is our generalization of Theorem 11 to stratified symplectic spaces.
Theorem 16.
If , then there is a canonical isomorphism of stratified symplectic spaces.
Proof.
By Corollary 10 and Proposition 15, the inclusion descends to a homeomorphism
whose restriction to is a diffeomorphism
for all . We also note that the -action on induces a residual action of the subtorus on . Proposition 8 then tells us that
for all and . The map therefore descends to a homeomorphism
whose restriction to is a diffeomorphism
for all .
Now consider the bijection
from Proposition 15(i). We claim that and define an isomorphism of stratified symplectic spaces, in the sense of Definition 13. In light of the previous paragraph, it suffices to prove the following for all : pulls the symplectic form on back to the symplectic form on .
Proposition 15(ii) implies that . This leads to the commutative diagram
where and are the canonical quotient maps and is the inclusion. We also have inclusion maps and . Another consideration is that (resp. ) is the unique -form on (resp. ) for which (resp. ). It therefore suffices to prove that . On the other hand, our commutative diagram implies that
This completes the proof. ∎
References
- [1] Duistermaat, J. J. On global action-angle coordinates. Comm. Pure Appl. Math. 33, 6 (1980), 687–706.
- [2] Guillemin, V., Jeffrey, L., and Sjamaar, R. Symplectic implosion. Transform. Groups 7, 2 (2002), 155–184.
- [3] Guillemin, V., and Sternberg, S. The moment map and collective motion. Ann. Physics 127, 1 (1980), 220–253.
- [4] Guillemin, V., and Sternberg, S. The Gel’fand-Cetlin system and quantization of the complex flag manifolds. J. Funct. Anal. 52, 1 (1983), 106–128.
- [5] Guillemin, V., and Sternberg, S. On collective complete integrability according to the method of Thimm. Ergodic Theory Dynam. Systems 3, 2 (1983), 219–230.
- [6] Harada, M. The symplectic geometry of the Gelfand-Cetlin-Molev basis for representations of . J. Symplectic Geom. 4, 1 (2006), 1–41.
- [7] Harada, M., and Kaveh, K. Integrable systems, toric degenerations and Okounkov bodies. Invent. Math. 202, 3 (2015), 927–985.
- [8] Hoffman, B., and Lane, J. Stratified gradient Hamiltonian vector fields and collective integrable systems. arXiv:2008.13656 (2022), 44pp.
- [9] Marsden, J. E., and Weinstein, A. Reduction of symplectic manifolds with symmetry. Rep. Mathematical Phys. 5, 1 (1974), 121–130.
- [10] Sjamaar, R., and Lerman, E. Stratified symplectic spaces and reduction. Ann. of Math. (2) 134, 2 (1991), 375–422.