This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Convexity properties of presymplectic moment maps

Yi Lin yilin@georgiasouthern.edu Department of Mathematical Sciences, Georgia Southern University, Statesboro, GA 30460 USA  and  Reyer Sjamaar sjamaar@math.cornell.edu Department of Mathematics, Cornell University, Ithaca, NY 14853-4201, USA
(Date: August 10, 2025)
Abstract.

The convexity and Morse-theoretic properties of moment maps in symplectic geometry typically fail for presymplectic manifolds. We find a condition on presymplectic moment maps that prevents these failures. Our result applies for instance to Prato’s quasifolds and to Hamiltonian actions on contact manifolds and cosymplectic manifolds.

1. Introduction

1.1. This paper deals with a topic in transverse geometry: in the context of a manifold XX with a (regular) foliation \mathscr{F} and a symplectic structure transverse to the foliation we develop analogues of a few basic results of symplectic geometry. While statements such as the Darboux theorem remain valid, one quickly discovers counterexamples to naive parallels of the convexity theorems of Hamiltonian compact Lie group actions proved by Atiyah [1], Guillemin and Sternberg [8], and Kirwan [15]. Some such counterexamples are catalogued in Section 4. Our main contribution is to state a condition under which these convexity theorems are true in the transversely symplectic setting. The condition, which we call cleanness of the group action, and special cases of which have been found earlier by other authors, such as He [10] and Ishida [14], is that there should exist an ideal of the Lie algebra of the group, called the null ideal, which at every point of the manifold spans the tangent space of the intersection of the group orbit with the leaf of \mathscr{F}.

We state and prove our convexity theorem in Section 2. It is formulated in terms of presymplectic structures instead of the equivalent language of transversely symplectic foliations. The adjective “presymplectic” has conflicting meanings in the current literature. We will use it for a closed 22-form of constant rank.

1.2. It was proved by Atiyah [1] and Guillemin and Sternberg [8] that the components of a symplectic moment map are Morse-Bott functions, an observation that lies at the heart of all subsequent developments in equivariant symplectic geometry. In the hope of opening the way to similar applications to the topology of presymplectic Lie group actions we show in Section 3 that under the cleanness assumption the components of a presymplectic moment map are Morse-Bott as well. In Section 4 we discuss some examples of the convexity theorem, including orbifolds, contact manifolds and Prato’s quasifolds [21].

1.3. To what extent does the moment polytope of the action of a compact Lie group GG on a transversely symplectic manifold XX depend only on the leaf space X/X/\mathscr{F}? If the foliation \mathscr{F} is fibrating, the leaf space is a symplectic manifold and the moment polytope of XX is the same as that of X/X/\mathscr{F}. In this case the moment polytope of XX is therefore completely determined by X/X/\mathscr{F}. But if the null foliation is not fibrating, the leaf space is often a messy topological space from which one cannot hope to recover the polytope. The structure of the leaf space can be enriched to that of an étale symplectic stack 𝒳\mathscr{X} (as in Lerman and Malkin [17], but without the Hausdorff condition), which is equipped with a Hamiltonian action of a stacky Lie group 𝒢\mathscr{G} (namely the “quotient” of GG by the in general non-closed normal subgroup generated by the null ideal). We will show in a future paper [13] that the moment polytope, reinterpreted as a stacky polytope, is an intrinsic invariant of the 𝒢\mathscr{G}-action on 𝒳\mathscr{X}.

1.4. The authors are grateful to Zuoqin Wang and Jiang-Hua Lu for helpful discussions. We draw the reader’s attention to Ratiu and Zung’s paper [22], which overlaps with ours and appeared on the arXiv at nearly the same time. We have omitted from this revised version some material developed more fully by them.

2. Presymplectic convexity

2.1.  A presymplectic manifold is a paracompact CC^{\infty}-manifold equipped with a closed 22-form of constant rank. A Hamiltonian action on a presymplectic manifold (X,ω)(X,\omega) consists of two pieces of data: a smooth action of a Lie group GG on XX and a smooth moment map Φ:X𝔤\Phi\colon X\to\mathfrak{g}^{*}. Here 𝔤=Lie(G)\mathfrak{g}=\operatorname{\mathrm{Lie}}(G) denotes the Lie algebra of GG and 𝔤\mathfrak{g}^{*} the dual vector space of 𝔤\mathfrak{g}. These data are subject to the following requirements: the GG-action should preserve the presymplectic form (i.e. gω=ωg^{*}\omega=\omega for all gGg\in G) and Φ\Phi should be an equivariant map satisfying dΦξ=ι(ξX)ωd\Phi^{\xi}=\iota(\xi_{X})\omega for all ξ𝔤\xi\in\mathfrak{g}. Here ξX\xi_{X} denotes the vector field on XX induced by ξ𝔤\xi\in\mathfrak{g}, and Φξ\Phi^{\xi} is the function on XX defined by Φξ(x)=Φ(x),ξ\Phi^{\xi}(x)=\langle\Phi(x),\xi\rangle for xXx\in X.

In the remainder of this section XX will denote a fixed manifold with a presymplectic form ω\omega and GG will denote a fixed compact connected Lie group acting on XX in a Hamiltonian fashion with moment map Φ\Phi. We will refer to XX as a presymplectic Hamiltonian GG-manifold. As far as we know, this notion was first introduced (under a different name) by Souriau [25, § 11]. The main goal of this section is to establish the following theorem. This result is very similar to the symplectic case, but the presence of the null foliation causes some interesting new phenomena.

2.2 Theorem (presymplectic convexity theorem).

Assume that the GG-action on XX is clean. Assume also that the manifold XX is connected and that the moment map Φ:X𝔤\Phi\colon X\to\mathfrak{g}^{*} is proper. Choose a maximal torus TT of GG and a closed Weyl chamber CC in 𝔱\mathfrak{t}^{*}, where 𝔱=Lie(T)\mathfrak{t}=\operatorname{\mathrm{Lie}}(T), and define Δ(X)=Φ(X)C\Delta(X)=\Phi(X)\cap C.

  1. (i)

    The fibres of Φ\Phi are connected and Φ:XΦ(X)\Phi\colon X\to\Phi(X) is an open map.

  2. (ii)

    Δ(X)\Delta(X) is a closed convex polyhedral set.

  3. (iii)

    Δ(X)\Delta(X) is rational if and only if the null subgroup N(X)N(X) of GG is closed.

This statement contains several undefined terms, which we proceed to explain in 2.22.2 below. In 2.6.52.10.4 we make further preliminary comments and state a sequence of auxiliary results. The proof of the theorem is in 2.11.2. Examples are presented in Section 4.

2.3.  A convex polyhedral set in a finite-dimensional real vector space is an intersection of a locally finite number of closed half-spaces. A convex polyhedron is an intersection of a finite number of closed half-spaces. A convex polytope is a bounded convex polyhedron. If the manifold XX is compact, the set Δ(X)\Delta(X) defined in Theorem 2.2 is a convex polytope.

  • \manfntsymbol127 2.4. Let EE be a finite-dimensional real vector space equipped with a Q-structure. We call a convex polyhedral subset of EE rational if it can be written as the locally finite intersection of half-spaces, each of which is given by an inequality of the form η,a\langle\eta,\cdot\rangle\geq a with rational normal vector ηE(Q)\eta\in E^{*}({\text{\bf Q}}) and aRa\in{\text{\bf R}}. This is nonstandard terminology. The more usual definition requires the scalars aa to be rational as well. Our notion of rationality is equivalent to the normal fan of the polyhedral set being rational.

    We call a convex polyhedral subset of 𝔱\mathfrak{t}^{*} rational if it is rational with respect to the Q-structure QZ𝔛(T){\text{\bf Q}}\otimes_{\text{\bf Z}}\mathfrak{X}^{*}(T). Here 𝔛(T)𝔱\mathfrak{X}^{*}(T)\subseteq\mathfrak{t}^{*} denotes the character lattice of the torus TT, i.e. the lattice dual to the exponential lattice 𝔛(T)=ker(exp:𝔱T)\mathfrak{X}_{*}(T)=\ker(\exp\colon\mathfrak{t}\to T).

2.5. The null ideal sheaf.  The subbundle ker(ω)\ker(\omega) of the tangent bundle TXTX is involutive (see e.g. [2, § 3]) and therefore, by Frobenius’ theorem, integrates to a (regular) foliation =X\mathscr{F}=\mathscr{F}_{X}, called the null foliation of ω\omega. We call ker(ω)\ker(\omega) the tangent bundle of the foliation and denote it usually by TT\mathscr{F}. The leaves of \mathscr{F} are not necessarily closed; indeed the case where they are not closed is the focus of our attention.

Let UU be an open subset of XX. Define 𝔫(U)\mathfrak{n}(U) to be the Lie subalgebra of 𝔤\mathfrak{g} consisting of all ξ𝔤\xi\in\mathfrak{g} with the property that the 11-form dΦξ=ι(ξX)ωd\Phi^{\xi}=\iota(\xi_{X})\omega vanishes on the GG-invariant open set GUG\cdot U. Equivalently, ξ\xi is in 𝔫(U)\mathfrak{n}(U) if and only if the moment map component Φξ\Phi^{\xi} is locally constant on GUG\cdot U, which is the case if and only if the induced vector field ξX\xi_{X} is tangent everywhere on GUG\cdot U to the foliation \mathscr{F}. Define N(U)N(U) to be the connected immersed (but not necessarily closed) Lie subgroup of GG whose Lie algebra is 𝔫(U)\mathfrak{n}(U). If the leaves of |GU\mathscr{F}|_{G\cdot U} are closed subsets of GUG\cdot U, then the subgroup N(U)N(U) is closed, but we do not assume this to be the case. For all gGg\in G and ξ𝔫(U)\xi\in\mathfrak{n}(U) we have

ι((Adg(ξ))X)ω=ι(g(ξX))ω=(g1)(ι(ξX)ω)=0\iota((\operatorname{\mathrm{Ad}}_{g}(\xi))_{X})\omega=\iota(g_{*}(\xi_{X}))\omega=(g^{-1})^{*}(\iota(\xi_{X})\omega)=0

on GUG\cdot U, and therefore the adjoint action of GG on 𝔤\mathfrak{g} preserves 𝔫(U)\mathfrak{n}(U). In particular the subalgebra 𝔫(U)\mathfrak{n}(U) of 𝔤\mathfrak{g} is an ideal and the subgroup N(U)N(U) of GG is normal.

The assignment 𝔫:U𝔫(U)\mathfrak{n}\colon U\mapsto\mathfrak{n}(U) is a presheaf on XX. Its associated sheaf 𝔫~\tilde{\mathfrak{n}} is a subsheaf of ideals of the constant sheaf 𝔤\mathfrak{g}. We have 𝔫~(U)=V𝔫(V)\tilde{\mathfrak{n}}(U)=\prod_{V}\mathfrak{n}(V), where the product is over all connected components VV of UU. The restriction morphisms of the presheaf 𝔫\mathfrak{n} are injective, which implies that for a decreasing sequence of open sets U1U2UnU_{1}\supseteq U_{2}\supseteq\cdots\supseteq U_{n}\supseteq\cdots the sequence of ideals 𝔫(U1)𝔫(U2)𝔫(Un)\mathfrak{n}(U_{1})\subseteq\mathfrak{n}(U_{2})\subseteq\dots\subseteq\mathfrak{n}(U_{n})\subseteq\dots is increasing and therefore eventually constant. Hence the sheaf 𝔫~\tilde{\mathfrak{n}} is constructible and its stalk 𝔫x=𝔫~x\mathfrak{n}_{x}=\tilde{\mathfrak{n}}_{x} at xx is equal to 𝔫(U)\mathfrak{n}(U) for all sufficiently small open neighbourhoods UU of xx. Similarly, the presheaf N:UN(U)N\colon U\mapsto N(U) sheafifies to a constructible subsheaf N~\tilde{N} of normal subgroups of the constant sheaf GG, whose stalk Nx=N~xN_{x}=\tilde{N}_{x} at xx is equal to N(U)N(U) for any suitably small open UU containing xx. We will call 𝔫~=𝔫~X\tilde{\mathfrak{n}}=\tilde{\mathfrak{n}}_{X} the null ideal sheaf and N~=N~X\tilde{N}=\tilde{N}_{X} the null subgroup sheaf of the presymplectic Hamiltonian action.

2.6. Clean actions.  Let xXx\in X and let UU be a GG-invariant open neighbourhood of xx. The N(U)N(U)-action maps each leaf of the foliation |U\mathscr{F}|_{U} to itself. Therefore the orbit N(U)xN(U)\cdot x is contained in Gx(x)G\cdot x\cap\mathscr{F}(x), where (x)\mathscr{F}(x) denotes the leaf of xx. Infinitesimally, the tangent space Tx(N(U)x)T_{x}(N(U)\cdot x) is contained in Tx(Gx)TxT_{x}(G\cdot x)\cap T_{x}\mathscr{F}. Taking UU to be small enough we have N(U)=NxN(U)=N_{x} and so Tx(Nxx)Tx(Gx)TxT_{x}(N_{x}\cdot x)\subseteq T_{x}(G\cdot x)\cap T_{x}\mathscr{F}. We call the GG-action on XX clean at xx if this inclusion is an equality, i.e.

(2.6.1) Tx(Nxx)=Tx(Gx)Tx.T_{x}(N_{x}\cdot x)=T_{x}(G\cdot x)\cap T_{x}\mathscr{F}.

Cleanness is a GG-invariant condition: if the action is clean at xx, then it is clean at gxgx for every gGg\in G. Cleanness is a local condition: the GG-action on XX is clean at xx if and only if the GG-action on UU is clean at xx for some GG-invariant open set UU containing xx. Cleanness is not necessarily an open condition. (See Example 2.6.5.)

We will state some criteria for the action to be clean in terms of the induced GG-action on the leaf space. Since the GG-action on XX preserves the form ω\omega, it sends leaves to leaves, and therefore descends to a continuous action on the leaf space X/X/\mathscr{F} (equipped with its quotient topology). Let GxG_{x} be the stabilizer of xXx\in X and 𝔤x\mathfrak{g}_{x} its Lie algebra. Let x¯=(x)\bar{x}=\mathscr{F}(x) denote the leaf of xx considered as a point in the leaf space. Let Gx¯G_{\bar{x}} be the stabilizer of x¯\bar{x} with respect to the induced GG-action on X/X/\mathscr{F}. We equip Gx¯G_{\bar{x}} with its induced Lie group structure (see e.g. [3, § III.4.5]), which makes it an immersed (but not necessarily embedded) subgroup of GG. Its Lie algebra 𝔤x¯\mathfrak{g}_{\bar{x}} consists of all ξ𝔤\xi\in\mathfrak{g} satisfying ξX(x)Tx\xi_{X}(x)\in T_{x}\mathscr{F}. By definition we have

(2.6.2) Gx¯x=Gx(x)G_{\bar{x}}\cdot x=G\cdot x\cap\mathscr{F}(x)

and

(2.6.3) 𝔫(U)=xGU𝔤x¯\mathfrak{n}(U)=\bigcap_{x\in G\cdot U}\mathfrak{g}_{\bar{x}}

for all open UXU\subseteq X. (Thus N(X)N(X) is the identity component of the subgroup of GG that acts trivially on the leaf space X/X/\mathscr{F}.) The foliation \mathscr{F}, being GG-invariant, induces a foliation of each orbit GxG\cdot x. This induced foliation is equal to the null foliation of the form ω\omega restricted to GxG\cdot x. The leaves of the induced foliation are the connected components of intersections of the form Gx(y)G\cdot x\cap\mathscr{F}(y). We see from \oldtagform@2.6.2 that the leaves of the induced foliation can also be described as the left translates of the connected components of Gx¯xG_{\bar{x}}\cdot x.

2.6.4 Lemma.

For every xXx\in X the following conditions are equivalent.

  1. (i)

    The GG-action is clean at xx;

  2. (ii)

    there exist vectors ξ1\xi_{1}, ξ2,\xi_{2},\dots, ξk𝔤\xi_{k}\in\mathfrak{g} and a GG-invariant open neighbourhood UU of xx with the property that ξ1,X\xi_{1,X}, ξ2,X,\xi_{2,X},\dots, ξk,X\xi_{k,X} are tangent to \mathscr{F} on UU and ξ1,X(x)\xi_{1,X}(x), ξ2,X(x),\xi_{2,X}(x),\dots, ξk,X(x)\xi_{k,X}(x) span Tx(Gx)TxT_{x}(G\cdot x)\cap T_{x}\mathscr{F};

  3. (iii)

    the leaves of the foliation of the GG-orbit GxG\cdot x induced by \mathscr{F} are NxN_{x}-orbits;

  4. (iv)

    Tx(Nxx)=Tx(Gx¯x)T_{x}(N_{x}\cdot x)=T_{x}(G_{\bar{x}}\cdot x);

  5. (v)

    the orbit NxxN_{x}\cdot x is an open subset of the orbit Gx¯xG_{\bar{x}}\cdot x;

  6. (vi)

    𝔤x¯=𝔤x+𝔫x\mathfrak{g}_{\bar{x}}=\mathfrak{g}_{x}+\mathfrak{n}_{x};

  7. (vii)

    the Gx¯G_{\bar{x}}-action is clean at xx.

Proof.

Condition \oldtagform@ii is a straightforward reformulation of the definition of cleanness. Next we show that \oldtagform@i \iff \oldtagform@iii. Let x\mathscr{F}^{x} be the induced foliation of the orbit GxG\cdot x. Let gGg\in G and y=gxy=gx. The leaf x(y)=gx(x)\mathscr{F}^{x}(y)=g\cdot\mathscr{F}^{x}(x) contains the orbit Nxy=Nxgx=gNxxN_{x}\cdot y=N_{x}\cdot gx=g\cdot N_{x}\cdot x. Since NxN_{x} is connected, the reverse inclusion x(y)Nxy\mathscr{F}^{x}(y)\subseteq N_{x}\cdot y holds if and only if NxxN_{x}\cdot x is open in x(x)\mathscr{F}^{x}(x), which is the case if and only if Tx(Nxx)=Tx(Gx)TxT_{x}(N_{x}\cdot x)=T_{x}(G\cdot x)\cap T_{x}\mathscr{F}. The equivalence \oldtagform@i \iff \oldtagform@iv is immediate from \oldtagform@2.6.1 and \oldtagform@2.6.2. Since the orbit NxxN_{x}\cdot x is always a subset of Gx¯xG_{\bar{x}}\cdot x, the equivalence \oldtagform@iv \iff \oldtagform@v is immediate. Condition \oldtagform@iv is equivalent to 𝔤x¯/𝔤x=(𝔫x+𝔤x)/𝔤x\mathfrak{g}_{\bar{x}}/\mathfrak{g}_{x}=(\mathfrak{n}_{x}+\mathfrak{g}_{x})/\mathfrak{g}_{x}, which is equivalent to \oldtagform@vi. Finally, \oldtagform@vii is a reformulation of \oldtagform@iv. ∎

2.6.5 Example.

Let X=Y×VX=Y\times V and ω=ωY0\omega=\omega_{Y}\oplus 0, where (Y,ωY)(Y,\omega_{Y}) is a symplectic Hamiltonian GG-manifold and VV is a real GG-module. Then the leaves of the null foliation \mathscr{F} are the fibres of the projection XYX\to Y. Let x=(y,v)Xx=(y,v)\in X. Then (x)={y}×V\mathscr{F}(x)=\{y\}\times V and therefore for ξ𝔤\xi\in\mathfrak{g} the tangent vector to the orbit ξX,xTx(Gx)\xi_{X,x}\in T_{x}(G\cdot x) is tangent to the leaf if and only if ξY,y=0\xi_{Y,y}=0. This shows that

(2.6.6) Tx(Gx)Tx={0}×Tv(Gyv)TyY×V.T_{x}(G\cdot x)\cap T_{x}\mathscr{F}=\{0\}\times T_{v}(G_{y}\cdot v)\subseteq T_{y}Y\times V.

The stabilizer of the leaf (x)\mathscr{F}(x) is Gx¯=GyG_{\bar{x}}=G_{y}. It follows from \oldtagform@2.6.3 that for all open UXU\subseteq X we have

𝔫(U)=(y,v)GU𝔤x¯=(y,v)GU𝔤y=𝔨,\mathfrak{n}(U)=\bigcap_{(y,v)\in G\cdot U}\mathfrak{g}_{\bar{x}}=\bigcap_{(y,v)\in G\cdot U}\mathfrak{g}_{y}=\mathfrak{k},

where 𝔨=ker(𝔤Γ(TY))\mathfrak{k}=\ker(\mathfrak{g}\to\Gamma(TY)) denotes the kernel of the infinitesimal action of 𝔤\mathfrak{g} on YY. Hence 𝔫~X\tilde{\mathfrak{n}}_{X} is the constant sheaf with stalk 𝔨\mathfrak{k} and Nx=KN_{x}=K, where K=exp(𝔨)K=\exp(\mathfrak{k}). Therefore

(2.6.7) Tx(Nxx)={0}×Tv(Kv)TyY×V.T_{x}(N_{x}\cdot x)=\{0\}\times T_{v}(K\cdot v)\subseteq T_{y}Y\times V.

Comparing \oldtagform@2.6.6 with \oldtagform@2.6.7 we see that the action is clean at x=(y,v)x=(y,v) if v=0v=0, but usually not at other points. For instance, if the action on YY is effective (K={1}K=\{1\}), the action is not clean at xx soon as dim(Gyv)1\dim(G_{y}\cdot v)\geq 1.

Two extreme cases of the cleanness condition merit attention. We say the GG-action is leafwise transitive at xx if (x)=Nxx\mathscr{F}(x)=N_{x}\cdot x. Leafwise transitivity at a point is a GG-invariant, local and open condition. We say that the action is leafwise nontangent at xx if Tx(Gx)Tx=0T_{x}(G\cdot x)\cap T_{x}\mathscr{F}=0. Leafwise nontangency forces the stalk NxN_{x} to be a subgroup of the stabilizer GxG_{x}. Leafwise nontangency at a point is a GG-invariant local condition, but not necessarily an open condition. Either condition implies cleanness.

We say that the GG-action on XX is clean (resp. leafwise transitive, resp. leafwise nontangent) if it is clean (resp. leafwise transitive, resp. leafwise nontangent) at all points of XX. Leafwise transitivity guarantees that the null foliation is Riemannian, a property that is frequently useful in applications.

2.7.  In the symplectic case (ker(ω)=0\ker(\omega)=0) the foliation \mathscr{F} is trivial, the sheaf 𝔫~\tilde{\mathfrak{n}} is constant with stalk equal to the kernel of the infinitesimal action 𝔤Γ(TX)\mathfrak{g}\to\Gamma(TX), and the cleanness condition is automatically fulfilled, so Theorem 2.2 reduces to the convexity theorems of Atiyah [1], Guillemin and Sternberg [8], and Kirwan [15]. A novel feature of our more general theorem is that the polyhedral set Δ(X)\Delta(X) may be irrational. It is however “rational” in the weak sense that its normal vectors are contained in the quasi-lattice 𝔛(T)/(𝔛(T)𝔫(X))\mathfrak{X}_{*}(T)/(\mathfrak{X}_{*}(T)\cap\mathfrak{n}(X)) in the quotient space 𝔱/𝔫(X)\mathfrak{t}/\mathfrak{n}(X), as we shall explain in 2.12.1.

2.8.  Other antecedents of Theorem 2.2 can be found in the papers of Prato [21] and Ishida [14] and in He’s thesis [10]. These authors impose opposite versions of our cleanness condition: Prato and Ishida deal with leafwise transitive torus actions, while He studies certain leafwise nontangent torus actions. It was our attempt to unify their results that led to this paper. It was observed by He [10, Ch. 4] that in the absence of any cleanness hypothesis the convexity of the image may fail. We give further counterexamples in 4. However, cleanness is not necessary for convexity to hold. For instance, let ZZ be a GG-manifold and suppose we have an equivariant surjective submersion f:ZXf\colon Z\to X. Then fωf^{*}\omega is a presymplectic form on ZZ and fΦf^{*}\Phi is a moment map for the GG-action on ZZ. Clearly ZZ has the same moment map image as XX. But the action on ZZ is rarely clean, even if the action on XX is clean.

2.9. The leaf space as a Hamiltonian space.  The null subgroup N(X)N(X) acts trivially on the leaf space X/X/\mathscr{F}, so the induced GG-action descends to an action of the (in general non-Hausdorff) quotient group G/N(X)G/N(X) on the (in general non-Hausdorff) space X/X/\mathscr{F}. The moment map also descends because of the following basic proposition. Here we denote by Cbas(X)C^{\smash{\infty}}_{\smash{{\mathrm{bas}}}}(X) the set of basic smooth functions on XX, i.e. those that are constant along the leaves. By the affine span of a subset AA of a vector space EE we mean the smallest affine subspace of EE that contains AA, i.e. the intersection of all affine subspaces of EE that contain AA. We denote by FEF^{\circ}\subseteq E^{*} the annihilator of a linear subspace FF of EE.

2.9.1 Proposition.
  1. (i)

    The moment map Φ\Phi is constant along the leaves of \mathscr{F}.

  2. (ii)

    The moment map Φ\Phi induces a morphism of Poisson algebras Φ:𝔤Cbas(X)\Phi^{*}\colon\mathfrak{g}\to C^{\smash{\infty}}_{\smash{{\mathrm{bas}}}}(X).

  3. (iii)

    If XX is connected, the affine span of the image Φ(X)\Phi(X) is of the form λ+𝔫(X)\lambda+\mathfrak{n}(X)^{\circ} for some element λ𝔤\lambda\in\mathfrak{g}^{*} which is fixed under the coadjoint GG-action.

Proof.
\oldtagform@

i Let xXx\in X, vTx=ker(ωx)v\in T_{x}\mathscr{F}=\ker(\omega_{x}), and ξ𝔤\xi\in\mathfrak{g}. Then dΦξ(v)=ωx(ξX(x),v)=0d\Phi^{\xi}(v)=\omega_{x}(\xi_{X}(x),v)=0, so Φξ\Phi^{\xi} is constant on the leaf (x)\mathscr{F}(x) for all ξ\xi.

\oldtagform@

ii It follows from \oldtagform@i that Φ\Phi pulls back smooth functions on 𝔤\mathfrak{g}^{*} to basic functions on XX. That Cbas(X)C_{\smash{{\mathrm{bas}}}}^{\smash{\infty}}(X) is a Poisson algebra is discussed for example in [5, § 2.2]. That Φ\Phi^{*} preserves the Poisson bracket follows from the equivariance of Φ\Phi as in the non-degenerate case.

\oldtagform@

iii Let ξ𝔤\xi\in\mathfrak{g}. Regarding ξ\xi as a linear function on 𝔤\mathfrak{g}^{*} we have: ξ\xi is constant on Φ(X)\Phi(X) \iff Φξ\Phi^{\xi} is constant on XX \iff dΦξ=0d\Phi^{\xi}=0 \iff ξ𝔫(X)\xi\in\mathfrak{n}(X). Thus the affine span of Φ(X)\Phi(X) is equal to λ+𝔫(X)\lambda+\mathfrak{n}(X)^{\circ}, where λ=Φ(x)\lambda=\Phi(x) for some xXx\in X. Since the moment map is equivariant, this affine subspace contains the coadjoint orbit of λ\lambda. It follows from Lemma B.2, applied to the GG-module 𝔤\mathfrak{g}^{*} and the submodule 𝔫(X)\mathfrak{n}(X)^{\circ}, that λ+𝔫(X)=λ0+𝔫(X)\lambda+\mathfrak{n}(X)^{\circ}=\lambda_{0}+\mathfrak{n}(X)^{\circ}, where λ0\lambda_{0} is GG-fixed. ∎

So if XX is connected we can replace Φ\Phi with Φλ\Phi-\lambda to obtain a new equivariant moment map which maps XX into 𝔫(X)(𝔤/𝔫(X))\mathfrak{n}(X)^{\circ}\cong(\mathfrak{g}/\mathfrak{n}(X))^{*} and which descends to a continuous map

Φ:X/(𝔤/𝔫(X)).\Phi_{\mathscr{F}}\colon X/\mathscr{F}\longrightarrow(\mathfrak{g}/\mathfrak{n}(X))^{*}.

This suggests the point of view that Φ\Phi_{\mathscr{F}} is the “moment map” for a “Hamiltonian action” on the “symplectic leaf space” X/X/\mathscr{F} of the “Lie group” G/N(X)G/N(X), whose “tangent Lie algebra” is 𝔤/𝔫(X)\mathfrak{g}/\mathfrak{n}(X), and that Δ(X)=Φ(X/)C\Delta(X)=\Phi_{\mathscr{F}}(X/\mathscr{F})\cap C is the “moment polytope” for this action. In the forthcoming paper [13] we will justify this point of view in terms of the language of Lie groupoids or differentiable stacks: we will “integrate” the foliated manifold (X,)(X,\mathscr{F}) to a Lie groupoid XX_{\bullet} and the Lie algebra homomorphism 𝔫(X)𝔤\mathfrak{n}(X)\to\mathfrak{g} to a Lie 22-group GG_{\bullet} which acts on XX_{\bullet}, and show that the moment polytope is a Morita invariant of the GG_{\bullet}-action on XX_{\bullet}.

Here we point out just one manifestation of this Morita invariance. Suppose that the null subgroup N(X)N(X) admits a complement in the sense that there exists an immersed Lie subgroup KK of GG with the property that KN(X)=GKN(X)=G and 𝔨+𝔫(X)=𝔤\mathfrak{k}+\mathfrak{n}(X)=\mathfrak{g}. Let 𝔪(X)=𝔨𝔫(X)\mathfrak{m}(X)=\mathfrak{k}\cap\mathfrak{n}(X) be the null ideal of the KK-action on XX. Then the Lie 22-group N(X)GN(X)\to G is Morita equivalent to the Lie 22-group M(X)KM(X)\to K, where M(X)M(X) is the immersed Lie subgroup of KK generated by exp(𝔪(X))\exp(\mathfrak{m}(X)). The Lie algebras 𝔤/𝔫(X)\mathfrak{g}/\mathfrak{n}(X) and 𝔨/𝔪(X)\mathfrak{k}/\mathfrak{m}(X) are isomorphic, so it follows immediately from Proposition 2.9.1 that under the projection 𝔤𝔨\mathfrak{g}^{*}\to\mathfrak{k}^{*} the GG-moment map image of XX maps bijectively to the KK-moment map image. We can play GG and KK off against each other in two opposite ways. (1) If the GG-action is clean, then the KK-action may not be clean, but even so the convexity theorem guarantees the convexity of the KK-moment map image. (2) If GG is not compact, but KK is compact and acts cleanly, then the convexity theorem guarantees the convexity of the GG-moment map image.

2.10. A foliated slice theorem.  The first step towards the proof of Theorem 2.2 is to construct equivariant foliation charts. We would like to choose a GG-invariant transverse section to the foliation at a point xx, but a transverse section YY has the property TxYTx=0T_{x}Y\cap T_{x}\mathscr{F}=0, and therefore cannot be GG-invariant unless the action is leafwise nontangent at xx. Instead we do the next best thing, which is to choose a GG-invariant presymplectic submanifold YY that is transverse to \mathscr{F} at xx and has the weaker property that TxYTxTx(Gx)T_{x}Y\cap T_{x}\mathscr{F}\subseteq T_{x}(G\cdot x). The slice theorem for compact Lie group actions says that xx has an invariant open neighbourhood which is equivariantly diffeomorphic to a homogeneous vector bundle E=G×GxVE=G\times^{G_{x}}V, where VV is the GxG_{x}-module TxX/Tx(Gx)T_{x}X/T_{x}(G\cdot x). The following refinement of the slice theorem states that we can single out a direct summand V1V_{1} of VV such that the corresponding subbundle G×GxV1G\times^{G_{x}}V_{1} is a transversal YY of the desired type. We write points in homogeneous bundles such as EE as equivalence classes [g,v][g,v] of pairs (g,v)G×V(g,v)\in G\times V.

2.10.1 Theorem.

Let xXx\in X and let H=GxH=G_{x} be the stabilizer of xx. Define the HH-modules

V=TxXTx(Gx),V0=Tx(Gx)+TxTx(Gx),V1=V/V0=TxXTx(Gx)+Tx,V=\frac{T_{x}X}{T_{x}(G\cdot x)},\quad V_{0}=\frac{T_{x}(G\cdot x)+T_{x}\mathscr{F}}{T_{x}(G\cdot x)},\quad V_{1}=V/V_{0}=\frac{T_{x}X}{T_{x}(G\cdot x)+T_{x}\mathscr{F}},

and the corresponding GG-homogeneous vector bundles

E=G×HV,E0=G×HV0,E1=E/E0=G×HV1.E=G\times^{H}V,\qquad E_{0}=G\times^{H}V_{0},\qquad E_{1}=E/E_{0}=G\times^{H}V_{1}.

Choose an HH-invariant inner product on TxXT_{x}X and identify VV0V1V\cong V_{0}\oplus V_{1} as an orthogonal direct sum of HH-modules and EE0E1E\cong E_{0}\oplus E_{1} as an orthogonal direct sum of vector bundles. There exists a GG-equivariant open embedding χ:EX\chi\colon E\to X which sends [1,0][1,0] to xx and has the following properties:

  1. (i)

    Y=χ(E1)Y=\chi(E_{1}) is a presymplectic Hamiltonian GG-manifold with presymplectic form ωY=ω|Y\omega_{Y}=\omega|_{Y} and moment map ΦY=Φ|Y\Phi_{Y}=\Phi|_{Y};

  2. (ii)

    U=χ(E)U=\chi(E) is a GG-equivariant tubular neighbourhood of YY with tubular neighbourhood projection p:UYp\colon U\to Y corresponding to the orthogonal projection EE1E\to E_{1}; every fibre of pp is contained in a leaf of \mathscr{F} and p1(x)V0p^{-1}(x)\cong V_{0};

  3. (iii)

    pωY=ωUp^{*}\omega_{Y}=\omega_{U} and pΦY=ΦUp^{*}\Phi_{Y}=\Phi_{U}, where ωU=ω|U\omega_{U}=\omega|_{U} and ΦU=Φ|U\Phi_{U}=\Phi|_{U};

  4. (iv)

    YY is transverse to the foliation \mathscr{F};

  5. (v)

    TxYTxTx(Gx)T_{x}Y\cap T_{x}\mathscr{F}\subseteq T_{x}(G\cdot x).

Now assume that the GG-action is clean at xx. Then χ\chi can be chosen in such a way that in addition to \oldtagform@i\oldtagform@v the following conditions are satisfied:

  1. (vi)

    the GG-action on YY is leafwise transitive;

  2. (vii)

    the null ideal sheaves 𝔫~Y\tilde{\mathfrak{n}}_{Y} and 𝔫~U\tilde{\mathfrak{n}}_{U} are constant with stalk equal to 𝔫X,x\mathfrak{n}_{X,x}.

Proof.

Choose a foliation chart (O,ζ)(O,\zeta) at xx in the following manner: start with an HH-invariant open neighbourhood OO of xx and a chart ζ˘:OTxX\breve{\zeta}\colon O\to T_{x}X centred at xx which satisfies Txζ˘=idTxXT_{x}\breve{\zeta}=\operatorname{\mathrm{id}}_{T_{x}X} and which maps the leaf of each yOy\in O onto the affine subspace through ζ˘(y)\breve{\zeta}(y) parallel to TxT_{x}\mathscr{F}. In other words, ζ˘(O(y))=ζ˘(y)+Tx\breve{\zeta}(O\cap\mathscr{F}(y))=\breve{\zeta}(y)+T_{x}\mathscr{F} for all yy, and in particular ζ˘(O(x))=Tx\breve{\zeta}(O\cap\mathscr{F}(x))=T_{x}\mathscr{F}. Let dhdh be the normalized Haar measure on HH and put ζ(y)=Hhζ˘(h1y)𝑑h\zeta(y)=\int_{H}h\breve{\zeta}(h^{-1}y)\,dh for yOy\in O. Then ζ:OTxX\zeta\colon O\to T_{x}X is HH-equivariant and Txζ=idTxXT_{x}\zeta=\operatorname{\mathrm{id}}_{T_{x}X}. For all yOy\in O and zO(y)z\in O\cap\mathscr{F}(y) we have

ζ(z)ζ(y)=Hh(ζ˘(h1z)ζ˘(h1y))𝑑hTx,\zeta(z)-\zeta(y)=\int_{H}h\bigl{(}\breve{\zeta}(h^{-1}z)-\breve{\zeta}(h^{-1}y)\bigr{)}\,dh\in T_{x}\mathscr{F},

because ζ˘(h1z)ζ˘(h1y)Tx\breve{\zeta}(h^{-1}z)-\breve{\zeta}(h^{-1}y)\in T_{x}\mathscr{F} and HH preserves the subspace TxT_{x}\mathscr{F} of TxXT_{x}X. This shows that ζ(O(y))\zeta(O\cap\mathscr{F}(y)) is contained in the affine subspace ζ(y)+Tx\zeta(y)+T_{x}\mathscr{F}. After replacing OO with a smaller open set and after rescaling ζ\zeta if necessary we obtain an HH-equivariant chart ζ:OTxX\zeta\colon O\to T_{x}X centred at xx with the property that ζ(O(y))=ζ(y)+Tx\zeta(O\cap\mathscr{F}(y))=\zeta(y)+T_{x}\mathscr{F} for all yOy\in O. The normal bundle of the orbit GxG/HG\cdot x\cong G/H is the homogeneous vector bundle EE, whose fibre VV we identify with the HH-submodule Tx(Gx)T_{x}(G\cdot x)^{\perp} of TxXT_{x}X (the orthogonal complement of Tx(Gx)T_{x}(G\cdot x) with respect to the HH-invariant inner product on TxXT_{x}X). Likewise we identify

V0Tx(Gx)TxandV1Tx(Gx)(Tx).V_{0}\cong T_{x}(G\cdot x)^{\perp}\cap T_{x}\mathscr{F}\qquad\text{and}\qquad V_{1}\cong T_{x}(G\cdot x)^{\perp}\cap(T_{x}\mathscr{F})^{\perp}.

Then V=V0V1V=V_{0}\oplus V_{1} is an orthogonal direct sum and the map χV=ζ1|V\chi_{V}=\zeta^{-1}|_{V} is an HH-equivariant embedding of the fibre VV into OO which maps every affine subspace of VV parallel to V0V_{0} into a leaf of \mathscr{F}. The map χV\chi_{V} extends uniquely to a GG-equivariant open immersion χ:EX\chi\colon E\to X, the image of which is the open set U=χ(E)=GOU=\chi(E)=G\cdot O and which maps each fibre of the orthogonal projection EE1E\to E_{1} into a leaf of \mathscr{F}. Let us choose OO so small that χ\chi is an embedding. Then UU is a tubular neighbourhood of the orbit GxG\cdot x. We claim that χ:EX\chi\colon E\to X satisfies requirements \oldtagform@i\oldtagform@v. The sum E=E1E2E=E_{1}\oplus E_{2} is a vector bundle over E1E_{1} with fibre V0V_{0}, which proves \oldtagform@ii. The form ω\omega restricts to forms ωU\omega_{U} on U=χ(E)U=\chi(E) and ωY\omega_{Y} on Y=χ(E1)Y=\chi(E_{1}). The map Φ\Phi restricts to maps ΦU:U𝔤\Phi_{U}\colon U\to\mathfrak{g}^{*} and ΦY:Y𝔤\Phi_{Y}\colon Y\to\mathfrak{g}^{*}. Since ω\omega is \mathscr{F}-basic (i.e. ι(v)ω=ι(v)dω=0\iota(v)\omega=\iota(v)d\omega=0 for all vectors vv tangent to \mathscr{F}) and each fibre of pp is contained in a leaf of \mathscr{F}, we have pωY=ωUp^{*}\omega_{Y}=\omega_{U} and pΦY=ΦUp^{*}\Phi_{Y}=\Phi_{U}, which proves \oldtagform@iii. This also implies that ωY\omega_{Y} is of the same rank as ω\omega. In particular ωY\omega_{Y} is of constant rank, which proves \oldtagform@i, and YY is transverse to \mathscr{F}, which proves \oldtagform@iv. The restricted foliation Y=|Y\mathscr{F}_{Y}=\mathscr{F}|_{Y} is the null foliation of ωY\omega_{Y}. The tangent space to YY at xx is TxY=Tx(Gx)V1T_{x}Y=T_{x}(G\cdot x)\oplus V_{1}, so the tangent space to Y\mathscr{F}_{Y} at xx is the subspace of TxYT_{x}Y given by

TxY=ker(ωY,x)=TxYTx=(Tx(Gx)V1)Tx.T_{x}\mathscr{F}_{Y}=\ker(\omega_{Y,x})=T_{x}Y\cap T_{x}\mathscr{F}=\bigl{(}T_{x}(G\cdot x)\oplus V_{1}\bigr{)}\cap T_{x}\mathscr{F}.

Since V1V_{1} is orthogonal to both Tx(Gx)T_{x}(G\cdot x) and TxT_{x}\mathscr{F} this gives

TxY=Tx(Gx)Tx,T_{x}\mathscr{F}_{Y}=T_{x}(G\cdot x)\cap T_{x}\mathscr{F},

which implies \oldtagform@v. Now assume the action is clean at xx. Then we obtain

TxY=Tx(NX,xx).T_{x}\mathscr{F}_{Y}=T_{x}(N_{X,x}\cdot x).

Let us choose OO so small that NX,x=NX(U)N_{X,x}=N_{X}(U); then NX,xNX,yN_{X,x}\subseteq N_{X,y} for all yYy\in Y. The orbit NX,xyN_{X,x}\cdot y is an immersed submanifold of YY diffeomorphic to a homogeneous space of NX,xN_{X,x}. Because YY is a GG-homogeneous vector bundle over GxG\cdot x, we have a GG-equivariant projection YGxY\to G\cdot x. It follows that dim(NX,xy)dim(NX,xx)\dim(N_{X,x}\cdot y)\geq\dim(N_{X,x}\cdot x) for all yYy\in Y. On the other hand, since NX,xN_{X,x} is contained in NX,yN_{X,y}, NX,xyN_{X,x}\cdot y is contained in the leaf Y(y)\mathscr{F}_{Y}(y), whose dimension is independent of yy. So we see that dim(NX,xy)=dim(NX,xx)\dim(N_{X,x}\cdot y)=\dim(N_{X,x}\cdot x) and TyY=Ty(NX,xy)T_{y}\mathscr{F}_{Y}=T_{y}(N_{X,x}\cdot y) for all yYy\in Y. This proves \oldtagform@vi, but it actually proves something a bit stronger, namely Ty(Gy¯y)=Ty(NX,xy)T_{y}(G_{\bar{y}}\cdot y)=T_{y}(N_{X,x}\cdot y) for all yYy\in Y. This means 𝔤y¯/𝔤y=(𝔫X,x+𝔤y)/𝔤y\mathfrak{g}_{\bar{y}}/\mathfrak{g}_{y}=(\mathfrak{n}_{X,x}+\mathfrak{g}_{y})/\mathfrak{g}_{y}, i.e.

(2.10.2) 𝔤y¯=𝔤y+𝔫X,x\mathfrak{g}_{\bar{y}}=\mathfrak{g}_{y}+\mathfrak{n}_{X,x}

for all yYy\in Y. Let zUz\in U and put y=p(z)Yy=p(z)\in Y. To prove \oldtagform@vii it is enough to show that 𝔫X,z=𝔫Y,y=𝔫X,x\mathfrak{n}_{X,z}=\mathfrak{n}_{Y,y}=\mathfrak{n}_{X,x}. Let ξ𝔤\xi\in\mathfrak{g}. By definition we have ξ𝔫X,z\xi\in\mathfrak{n}_{X,z} if and only if ι(ξU)ωU=0\iota(\xi_{U^{\prime}})\omega_{U^{\prime}}=0 for some open neighbourhood UUU^{\prime}\subseteq U of zz. Since the vector fields ξU\xi_{U} and ξY\xi_{Y} are pp-related and pωY=ωUp^{*}\omega_{Y}=\omega_{U}, we have ι(ξU)ωU=0\iota(\xi_{U^{\prime}})\omega_{U^{\prime}}=0 if and only if ι(ξp(U))ωp(U)=0\iota(\xi_{p(U^{\prime})})\omega_{p(U^{\prime})}=0, i.e. ξ𝔫X,y\xi\in\mathfrak{n}_{X,y}. This proves 𝔫X,z=𝔫Y,y\mathfrak{n}_{X,z}=\mathfrak{n}_{Y,y}. Taking z=xz=x yields 𝔫X,x=𝔫Y,x\mathfrak{n}_{X,x}=\mathfrak{n}_{Y,x}. We finish by showing that 𝔫Y,y=𝔫Y,x\mathfrak{n}_{Y,y}=\mathfrak{n}_{Y,x}. We replace OO, if necessary, by a smaller open neighbourhood of xx such that YY has the property that 𝔫Y(Y)=𝔫Y,x\mathfrak{n}_{Y}(Y)=\mathfrak{n}_{Y,x}. Then we have 𝔫Y,y𝔫Y,x\mathfrak{n}_{Y,y}\supseteq\mathfrak{n}_{Y,x}. Supposing our assertion 𝔫Y,y=𝔫Y,x\mathfrak{n}_{Y,y}=\mathfrak{n}_{Y,x} to be false, we can find ξ𝔤\xi\in\mathfrak{g} such that ξ𝔫y𝔫x\xi\in\mathfrak{n}_{y}\setminus\mathfrak{n}_{x}. This means that the vector field ξY\xi_{Y} is tangent to Y\mathscr{F}_{Y} in an invariant neighbourhood U2U_{2} of yy in YY but not tangent to Y\mathscr{F}_{Y} in a invariant neighbourhood U1U_{1} of xx in YY. In other words,

ξY(w)Tw(Gw)TwY=Tw(Gw¯w)\xi_{Y}(w)\in T_{w}(G\cdot w)\cap T_{w}\mathscr{F}_{Y}=T_{w}(G_{\bar{w}}\cdot w)

for all wU2w\in U_{2}, but

ξY(v)Tv(Gv)TvY=Tv(Gv¯v)\xi_{Y}(v)\not\in T_{v}(G\cdot v)\cap T_{v}\mathscr{F}_{Y}=T_{v}(G_{\bar{v}}\cdot v)

for some vU1v\in U_{1}. Because of \oldtagform@2.10.2 this means that ξ𝔤w+𝔫X,x\xi\in\mathfrak{g}_{w}+\mathfrak{n}_{X,x} for all wU2w\in U_{2} but ξ𝔤v+𝔫X,x\xi\not\in\mathfrak{g}_{v}+\mathfrak{n}_{X,x} for some vU1v\in U_{1}. But, the group GG being compact, by choosing ww to be generic with respect to the GG-action we can arrange for 𝔤w\mathfrak{g}_{w} to be a subalgebra of 𝔤v\mathfrak{g}_{v}, which is a contradiction. Therefore 𝔫Y,y=𝔫Y,x\mathfrak{n}_{Y,y}=\mathfrak{n}_{Y,x}. ∎

2.10.3 Corollary.
  1. (i)

    If the action is clean at xXx\in X, then the sheaves 𝔫~\tilde{\mathfrak{n}} and N~\tilde{N} are constant on a neighbourhood of xx.

  2. (ii)

    Suppose that the GG-action on XX is clean and that XX is connected. Then the sheaves 𝔫~\tilde{\mathfrak{n}} and N~\tilde{N} are constant. It follows that Tx(N(X)x)=Tx(Gx)TxT_{x}(N(X)\cdot x)=T_{x}(G\cdot x)\cap T_{x}\mathscr{F} for all xXx\in X.

  3. (iii)

    If XX is connected and the GG-action on XX is leafwise transitive, then (x)=N(X)x\mathscr{F}(x)=N(X)\cdot x for all xx.

Proof.
\oldtagform@

i is a restatement of Theorem 2.10.1\oldtagform@vii. \oldtagform@ii follows from \oldtagform@i and the monotonicity property 𝔫(U1)𝔫(U2)\mathfrak{n}(U_{1})\supseteq\mathfrak{n}(U_{2}) for U1U2U_{1}\subseteq U_{2} of the presheaf 𝔫\mathfrak{n}. \oldtagform@iii follows immediately from \oldtagform@ii. ∎

If the GG-action is leafwise nontangent at xx, the transversal YY of Theorem 2.10.1 is a section of the foliation and therefore symplectic, which shows that XX is near xx a GG-equivariant bundle over a symplectic Hamiltonian GG-manifold. In particular, if xx is a fixed point the action is leafwise nontangent at xx and we obtain the following linearization or equivariant Darboux theorem. In the statement of this theorem we regard the tangent space TxXT_{x}X as a presymplectic manifold with constant presymplectic form ωx\omega_{x}. It follows from Corollary B.5 that at a fixed point xx the moment map has a well-defined Hessian Tx2Φ:TxX𝔤T^{2}_{x}\Phi\colon T_{x}X\to\mathfrak{g}^{*}.

2.10.4 Corollary (equivariant presymplectic Darboux theorem).

Let xXx\in X be a fixed point of GG. Then xx has a GG-invariant neighbourhood that is isomorphic as a presymplectic Hamiltonian GG-manifold to a GG-invariant neighbourhood of the origin in the tangent space TxXT_{x}X, equipped with the presymplectic structure ωx\omega_{x} and the moment map λ+Tx2Φ\lambda+T_{x}^{2}\Phi, where λ=Φ(x)(𝔤)G\lambda=\Phi(x)\in(\mathfrak{g}^{*})^{G} and the Hessian is given by Tx2Φξ(v)=12ωx(ξ(v),v)T_{x}^{2}\Phi^{\xi}(v)=\frac{1}{2}\omega_{x}(\xi(v),v).

Proof.

Since Gx=GG_{x}=G, we have E=V=TxXE=V=T_{x}X, E0=V0=TxE_{0}=V_{0}=T_{x}\mathscr{F}, and E1=V1=TxX/TxE_{1}=V_{1}=T_{x}X/T_{x}\mathscr{F}. Moreover, TxYTx=0T_{x}Y\cap T_{x}\mathscr{F}=0, so Y=χ(V1)Y=\chi(V_{1}) is symplectic. Therefore U=χ(V)=χ(V0×V1)=V0×YU=\chi(V)=\chi(V_{0}\times V_{1})=V_{0}\times Y. Applying the symplectic equivariant Darboux theorem (see [9, Theorem 22.2]) to the symplectic Hamiltonian GG-manifold YY and the fixed point xYx\in Y, we find that a GG-invariant neighbourhood of xx in XX is presymplectically and GG-equivariantly isomorphic to a neighbourhood of the origin in VV0V1V\cong V_{0}\oplus V_{1}. The action on VV being linear, the moment map on VV is quadratic with constant term λ\lambda and homogeneous quadratic part Tx2Φ=ΦVT_{x}^{2}\Phi=\Phi_{V}, where ΦV\Phi_{V} is as in Lemma B.1. ∎

2.11. Symplectization.  The second step towards the proof of Theorem 2.2 is symplectization. Let TT^{*}\mathscr{F} be the vector bundle dual to TT\mathscr{F} and let pr:TX\operatorname{\mathrm{pr}}\colon T^{*}\mathscr{F}\to X be the bundle projection. We choose a GG-invariant Riemannian metric on XX. (Recall our standing hypothesis that GG is compact.) Let TXTTX\to T\mathscr{F} be the orthogonal projection onto the subbundle TT\mathscr{F} of TXTX (with respect to the metric) and let j:TTXj\colon T^{*}\mathscr{F}\to T^{*}X be the dual embedding. Let ω0\omega_{0} be the standard symplectic form on the cotangent bundle TXT^{*}X and let Ω=prω+jω0\Omega=\operatorname{\mathrm{pr}}^{*}\omega+j^{*}\omega_{0}. The 22-form Ω\Omega on TT^{*}\mathscr{F} is symplectic in a neighbourhood of the zero section XX and the embedding XTX\to T^{*}\mathscr{F} is coisotropic. See [7] for these facts. The GG-action on TT^{*}\mathscr{F} is Hamiltonian with moment map Ψ:T𝔤\Psi\colon T^{*}\mathscr{F}\to\mathfrak{g}^{*} given by

Ψ=prΦ+jΦ0,\Psi=\operatorname{\mathrm{pr}}^{*}\Phi+j^{*}\Phi_{0},

where Φ0:TX𝔤\Phi_{0}\colon T^{*}X\to\mathfrak{g}^{*} is the moment map for the cotangent action given by the dual pairing Φ0ξ(y)=y,ξX(x)\Phi_{0}^{\xi}(y)=\langle y,\xi_{X}(x)\rangle for yTxXy\in T^{*}_{x}X. The germ at XX of the Hamiltonian GG-manifold TT^{*}\mathscr{F} is called the symplectization of XX. The next result says that in the leafwise transitive case every fibre of Φ\Phi is a fibre of Ψ\Psi and that the image of Φ\Phi is the intersection of the image of Ψ\Psi with an affine subspace.

2.11.1 Proposition.

Assume that XX is connected and that the GG-action on XX is leafwise transitive. Let λ(𝔤)G\lambda\in(\mathfrak{g}^{*})^{G} be as in Proposition 2.9.1\oldtagform@iii.

  1. (i)

    X=Ψ1(λ+𝔫(X))X=\Psi^{-1}(\lambda+\mathfrak{n}(X)^{\circ}).

  2. (ii)

    Ψ:T𝔤\Psi\colon T^{*}\mathscr{F}\to\mathfrak{g}^{*} intersects the affine subspace λ+𝔫(X)\lambda+\mathfrak{n}(X)^{\circ} cleanly.

  3. (iii)

    Φ(X)=Ψ(T)(λ+𝔫(X))\Phi(X)=\Psi(T^{*}\mathscr{F})\cap(\lambda+\mathfrak{n}(X)^{\circ}).

Proof.
\oldtagform@

i Let xXx\in X. Then Φ0(j(x))=0\Phi_{0}(j(x))=0, so

Ψ(x)=Φ(x)+Φ0(j(x))=Φ(x),\Psi(x)=\Phi(x)+\Phi_{0}(j(x))=\Phi(x),

and so Ψ(x)\Psi(x) is in λ+𝔫(X)\lambda+\mathfrak{n}(X)^{\circ} by Proposition 2.9.1\oldtagform@iii. This shows that XΨ1(λ+𝔫(X))X\subseteq\Psi^{-1}(\lambda+\mathfrak{n}(X)^{\circ}). Conversely, let zTz\in T^{*}\mathscr{F} and suppose that Ψ(z)λ+𝔫(X)\Psi(z)\in\lambda+\mathfrak{n}(X)^{\circ}. Put x=pr(z)Xx=\operatorname{\mathrm{pr}}(z)\in X and y=j(z)TxXy=j(z)\in T_{x}^{*}X. Then Ψ(z)=Φ(x)+Φ0(y)\Psi(z)=\Phi(x)+\Phi_{0}(y), so Φ0(y)=Ψ(z)Φ(x)𝔫(X)\Phi_{0}(y)=\Psi(z)-\Phi(x)\in\mathfrak{n}(X)^{\circ}. It follows that y,ξX(x)=0\langle y,\xi_{X}(x)\rangle=0 for every ξ𝔫(X)\xi\in\mathfrak{n}(X). In other words, yTxXy\in T_{x}^{*}X annihilates the tangent space to the N(X)N(X)-orbit N(X)xN(X)\cdot x. By Corollary 2.10.3\oldtagform@ii we have N(X)x=(x)N(X)\cdot x=\mathscr{F}(x) because the action is leafwise transitive. Therefore yy annihilates all of TxT_{x}\mathscr{F}. But yy is in the image of jj, which is a splitting of the natural surjection TXTT^{*}X\to T^{*}\mathscr{F}, and therefore yim(j)(Tx)=0y\in\operatorname{\mathrm{im}}(j)\cap(T_{x}\mathscr{F})^{\circ}=0. We conclude that z=xXz=x\in X.

\oldtagform@

ii We have just shown that Ψ1(λ+𝔫(X))\Psi^{-1}(\lambda+\mathfrak{n}(X)^{\circ}) is equal to XX and is therefore a submanifold of M=TM=T^{*}\mathscr{F}. It remains only to show that XX has the correct tangent bundle, namely TX=(TΨ)1(𝔫(X))TX=(T\Psi)^{-1}(\mathfrak{n}(X)^{\circ}). Let xXx\in X. Regarding xx as a point in the zero section of TXT^{*}X, we have Tx(TX)=TxXTxXT_{x}(T^{*}X)=T_{x}X\oplus T^{*}_{x}X and

TxM=TxXTx,TxΨ=prTxΦ+jTxΦ0,T_{x}M=T_{x}X\oplus T^{*}_{x}\mathscr{F},\qquad T_{x}\Psi=\operatorname{\mathrm{pr}}^{*}T_{x}\Phi+j^{*}T_{x}\Phi_{0},

where now pr\operatorname{\mathrm{pr}} stands for the projection TxMTxXT_{x}M\to T_{x}X and jj for the inclusion TxMTx(TX)T_{x}M\to T_{x}(T^{*}X). The derivative at xx of Φ0:TX𝔤\Phi_{0}\colon T^{*}X\to\mathfrak{g}^{*} is the linear map TxΦ0:TxXTxX𝔤T_{x}\Phi_{0}\colon T_{x}X\oplus T^{*}_{x}X\to\mathfrak{g}^{*} given by TxΦ0(u,v),ξ=v,ξX(x)\langle T_{x}\Phi_{0}(u,v),\xi\rangle=\langle v,\xi_{X}(x)\rangle for uTxXu\in T_{x}X, vTxXv\in T^{*}_{x}X and ξ𝔤\xi\in\mathfrak{g}. Now let wTxMw\in T_{x}M and suppose TxΨ(w)𝔫(X)T_{x}\Psi(w)\in\mathfrak{n}(X)^{\circ}. Put u=pr(w)TxXu=\operatorname{\mathrm{pr}}(w)\in T_{x}X and v=j(w)Tx(TX)v=j(w)\in T_{x}(T^{*}X). Then TxΦ0(v)=TxΨ(w)TxΦ(u)𝔫(X)T_{x}\Phi_{0}(v)=T_{x}\Psi(w)-T_{x}\Phi(u)\in\mathfrak{n}(X)^{\circ}, so v,ξX(x)=0\langle v,\xi_{X}(x)\rangle=0 for all ξ𝔫(X)\xi\in\mathfrak{n}(X). As in the proof of \oldtagform@i we deduce from this that v=0v=0, i.e. w=uTxXw=u\in T_{x}X. This proves (TxΨ)1(𝔫(X))TxX(T_{x}\Psi)^{-1}(\mathfrak{n}(X)^{\circ})\subseteq T_{x}X. The reverse inclusion follows from the fact that XX is contained in Ψ1(λ+𝔫(X))\Psi^{-1}(\lambda+\mathfrak{n}(X)^{\circ}).

\oldtagform@

iii follows immediately from \oldtagform@i. ∎

The next result, which is essentially due to Guillemin and Sternberg, is a partial converse to Proposition 2.11.1 as well as a useful source of examples.

2.11.2 Proposition.

Let (M,ωM)(M,\omega_{M}) be a symplectic Hamiltonian GG-manifold with moment map ΦM:M𝔤\Phi_{M}\colon M\to\mathfrak{g}^{*}. Let λ𝔤\lambda\in\mathfrak{g}^{*} and let 𝔞\mathfrak{a} be an ideal of 𝔤\mathfrak{g} satisfying 𝔨𝔞𝔤λ\mathfrak{k}\subseteq\mathfrak{a}\subseteq\mathfrak{g}_{\lambda}, where 𝔨\mathfrak{k} is the kernel of the infinitesimal action 𝔤Γ(TM)\mathfrak{g}\to\Gamma(TM). Assume that ΦM\Phi_{M} intersects the affine subspace λ+𝔞\lambda+\mathfrak{a}^{\circ} cleanly. Then X=ΦM1(λ+𝔞)X=\Phi_{M}^{-1}(\lambda+\mathfrak{a}^{\circ}) is a coisotropic submanifold of MM preserved by the action of GG. Therefore XX is a presymplectic Hamiltonian GG-manifold with presymplectic form ω=ωM|X\omega=\omega_{M}|_{X} and moment map Φ=ΦM|X\Phi=\Phi_{M}|_{X}. The action on XX is leafwise transitive and the null ideal 𝔫(X)\mathfrak{n}(X) of XX is equal to 𝔞\mathfrak{a}.

Proof.

It follows from Lemma B.3 that λ+𝔞\lambda+\mathfrak{a}^{\circ} is preserved by the coadjoint action. Therefore XX is preserved by GG and, by [9, Theorem 26.4], XX is coisotropic and the leaves of the null foliation of the presymplectic form ω\omega are the orbits of the AA-action on XX, where AA is the connected immersed normal subgroup corresponding to the ideal 𝔞\mathfrak{a}. Hence the action is leafwise transitive. By Proposition 2.9.1\oldtagform@iii the affine span of Φ(X)\Phi(X) is the affine subspace λ+𝔫(X)\lambda+\mathfrak{n}(X)^{\circ} and the affine span of Φ(M)\Phi(M) is λ+𝔨\lambda+\mathfrak{k}^{\circ}. Therefore

λ+𝔫(X)=(λ+𝔞)(λ+𝔨)=λ+𝔞.\lambda+\mathfrak{n}(X)^{\circ}=(\lambda+\mathfrak{a}^{\circ})\cap(\lambda+\mathfrak{k}^{\circ})=\lambda+\mathfrak{a}^{\circ}.

We conclude that 𝔫(X)=𝔞\mathfrak{n}(X)=\mathfrak{a}. ∎

2.12. Proof of the convexity theorem.  First we prove the following local version of the presymplectic convexity theorem. Recall that TT denotes a maximal torus of GG, 𝔱\mathfrak{t}^{*} the dual of its Lie algebra 𝔱\mathfrak{t}, and CC a closed chamber in 𝔱\mathfrak{t}^{*}. Recall also that every coadjoint orbit intersects the chamber CC in exactly one point and that the inclusion C𝔤C\to\mathfrak{g}^{*} induces a homeomorphism C𝔤/Ad(G)C\to\mathfrak{g}^{*}/\operatorname{\mathrm{Ad}}^{*}(G), the quotient of 𝔤\mathfrak{g}^{*} by the coadjoint action. (See e.g. [3, § IX.5.2].) We identify 𝔤/Ad(G)\mathfrak{g}^{*}/\operatorname{\mathrm{Ad}}^{*}(G) with CC via this homeomorphism and denote by ϕ:XC\phi\colon X\to C the composition of the moment map Φ:X𝔤\Phi\colon X\to\mathfrak{g}^{*} with the quotient map 𝔤C\mathfrak{g}^{*}\to C. Then the intersection Δ(X)=Φ(X)C\Delta(X)=\Phi(X)\cap C is nothing but the image ϕ(X)\phi(X).

2.12.1 Theorem (local presymplectic convexity theorem).

Assume that the GG-action is clean at xXx\in X. Then there exist a rational convex polyhedral cone Δx\Delta_{x} in 𝔱\mathfrak{t}^{*} with apex ϕ(x)\phi(x) and a basis of GG-invariant open neighbourhoods UU of xx in XX with the following properties:

  1. (i)

    the fibres of the map ϕ|U\phi|_{U} are connected;

  2. (ii)

    ϕ:UΔx(ϕ(x)+𝔫x)\phi\colon U\to\Delta_{x}\cap\bigl{(}\phi(x)+\mathfrak{n}_{x}^{\circ}\bigr{)} is an open map.

Proof.

Let us replace xx by a suitable GG-translate in order that Φ(x)=ϕ(x)\Phi(x)=\phi(x). Choose a transversal YY at xx and a GG-invariant tubular neighbourhood UU of YY as in Theorem 2.10.1. Since GG is connected, we may assume YY and UU to be connected. Then Φ(U)ϕ(x)+𝔫x\Phi(U)\subseteq\phi(x)+\mathfrak{n}_{x}^{\circ} by Proposition 2.9.1\oldtagform@iii, and therefore ϕ(U)ϕ(x)+𝔫x\phi(U)\subseteq\phi(x)+\mathfrak{n}_{x}^{\circ}. Let ϕU=ϕ|U\phi_{U}=\phi|_{U} and ϕY=ϕ|Y\phi_{Y}=\phi|_{Y}. Let MM be the symplectization of YY as defined in 2.10.4. Let Ψ:M𝔤\Psi\colon M\to\mathfrak{g}^{*} the moment map for the GG-action on MM and ψ:MC\psi\colon M\to C the composition of Ψ\Psi with the orbit map 𝔤C\mathfrak{g}^{*}\to C. It follows from Theorem 2.10.1 and Proposition 2.11.1 that for all νϕ(U)\nu\in\phi(U)

(2.12.2) ϕU1(ν)=p1(ϕY1(ν)),ϕY1(ν)=ψ1(ν),\phi_{U}^{-1}(\nu)=p^{-1}(\phi_{Y}^{-1}(\nu)),\qquad\phi_{Y}^{-1}(\nu)=\psi^{-1}(\nu),

and that

(2.12.3) ϕU=ϕYp,ϕY=ψ|Y,Y=ψ1(ϕ(x)+𝔫x).\phi_{U}=\phi_{Y}\circ p,\qquad\phi_{Y}=\psi|_{Y},\qquad Y=\psi^{-1}\bigl{(}\phi(x)+\mathfrak{n}_{x}^{\circ}\bigr{)}.

The local convexity theorem in the symplectic case (see e.g. [24, Theorem 6.5]) states that the fibres of ψ\psi are connected and that ψ:MΔx\psi\colon M\to\Delta_{x} is an open mapping to a rational convex polyhedral cone Δx\Delta_{x} in 𝔱\mathfrak{t}^{*} with apex ϕ(x)=ψ(x)\phi(x)=\psi(x). Taking this into account, we see from \oldtagform@2.12.2 that the fibres of ϕU\phi_{U} contract onto fibres of ψ\psi and are therefore connected as well, and we see from \oldtagform@2.12.3 that ϕU\phi_{U} is an open map to Δx(ϕ(x)+𝔫x)\Delta_{x}\cap\bigl{(}\phi(x)+\mathfrak{n}_{x}^{\circ}\bigr{)}. ∎

Proof of Theorem 2.2.

Theorem 2.12.1 asserts that the family of convex cones Δx(ϕ(x)+𝔫x)\Delta_{x}\cap\bigl{(}\phi(x)+\mathfrak{n}_{x}^{\circ}\bigr{)}, where xx ranges over XX, is a system of local convexity data for the quotient map ϕ\phi in the sense of Hilgert, Neeb and Planck [12, Definition 3.3]. Parts \oldtagform@i and \oldtagform@ii of Theorem 2.2 now follow from the local-to-global principle due to these authors, [12, Theorem 3.10]. It remains to prove part \oldtagform@iii. The local-to-global principle also tells us that for each xXx\in X the cone Δx(ϕ(x)+𝔫x)\Delta_{x}\cap\bigl{(}\phi(x)+\mathfrak{n}_{x}^{\circ}\bigr{)} is equal to the intersection of all supporting half-spaces of the convex set Δ(X)\Delta(X) at the point ϕ(x)\phi(x). A closed convex set is equal to the intersection of all its supporting half-spaces, and therefore

Δ(X)=xXΔx(ϕ(x)+𝔫x).\Delta(X)=\bigcap_{x\in X}\Delta_{x}\cap\bigl{(}\phi(x)+\mathfrak{n}_{x}^{\circ}\bigr{)}.

Since the action is clean, by Corollary 2.10.3\oldtagform@ii we have 𝔫x=𝔫(X)\mathfrak{n}_{x}=\mathfrak{n}(X) for all xx. By Proposition 2.9.1\oldtagform@iii there is an Ad(G)\operatorname{\mathrm{Ad}}^{*}(G)-fixed λ𝔤\lambda\in\mathfrak{g}^{*} such that Φ(X)λ+𝔫(X)=ϕ(x)+𝔫(X)\Phi(X)\subseteq\lambda+\mathfrak{n}(X)^{\circ}=\phi(x)+\mathfrak{n}(X)^{\circ} for all xx. After shifting the moment map we may assume λ=0\lambda=0. Then

(2.12.4) Δ(X)=xXΔx𝔫(X),\Delta(X)=\bigcap_{x\in X}\Delta_{x}\cap\mathfrak{n}(X)^{\circ},

the intersection of the locally finite family of rational cones Δx\Delta_{x} with the linear subspace 𝔫(X)\mathfrak{n}(X)^{\circ}. Now suppose that the null subgroup N(X)N(X) is closed. Then by Corollary A.2 the subspace 𝔫(X)𝔱\mathfrak{n}(X)\cap\mathfrak{t} of 𝔱\mathfrak{t} is rational and therefore the subspace 𝔫(X)𝔱\mathfrak{n}(X)^{\circ}\cap\mathfrak{t}^{*} of 𝔱\mathfrak{t}^{*} is rational. Because of this and \oldtagform@2.12.4 the polyhedral set Δ(X)\Delta(X) is rational. Conversely, suppose that Δ(X)\Delta(X) is rational. Let 𝔷\mathfrak{z} be the centre of 𝔤\mathfrak{g} and pr:𝔤𝔷\operatorname{\mathrm{pr}}\colon\mathfrak{g}^{*}\to\mathfrak{z}^{*} the projection dual to the inclusion of 𝔷\mathfrak{z} into 𝔤[𝔤,𝔤]𝔷\mathfrak{g}\cong[\mathfrak{g},\mathfrak{g}]\oplus\mathfrak{z}. Then the image pr(Δ(X))\operatorname{\mathrm{pr}}(\Delta(X)) is a rational polyhedral subset of 𝔷\mathfrak{z}. The chamber CC of 𝔤\mathfrak{g}^{*} is of the form C=C×𝔷C=C^{\prime}\times\mathfrak{z}^{*}, where CC^{\prime} is a chamber of [𝔤,𝔤][\mathfrak{g},\mathfrak{g}]^{*}, so we see that pr(Δ(X))\operatorname{\mathrm{pr}}(\Delta(X)) is equal to the image of XX under the moment map ΦZ\Phi_{Z} for the action of the central subtorus Z=expG(𝔷)Z=\exp_{G}(\mathfrak{z}) of GG. By Proposition 2.9.1\oldtagform@iii the affine span of ΦZ(X)=pr(Δ(X))\Phi_{Z}(X)=\operatorname{\mathrm{pr}}(\Delta(X)) is equal to (𝔫(X)𝔷)(\mathfrak{n}(X)\cap\mathfrak{z})^{\circ}. It follows that 𝔫(X)𝔷\mathfrak{n}(X)\cap\mathfrak{z} is a rational subspace of 𝔷\mathfrak{z}. By Corollary A.2 we conclude that N(X)N(X) is a closed subgroup of GG. ∎

2.13. Irrationality.  The following result is a consequence of the proof of Theorem 2.2\oldtagform@iii.

2.13.1 Corollary.

Under the hypotheses of Theorem 2.2, the polyhedral set Δ(X)\Delta(X) is rational if and only if the moment map image for the action of the central subtorus Z=exp(𝔷)Z=\exp(\mathfrak{z}) is a rational polyhedral subset of 𝔷\mathfrak{z}^{*}. In particular Δ(X)\Delta(X) is always rational if GG is semisimple.

So we see that the irrationality of presymplectic moment polytopes is essentially an abelian phenomenon.

As explained in 2.6.5, Δ(X)\Delta(X) is best regarded intrinsically as a subset of 𝔱0\mathfrak{t}_{0}^{*}, where 𝔱0\mathfrak{t}_{0} is the quotient 𝔱/(𝔱𝔫(X))\mathfrak{t}/(\mathfrak{t}\cap\mathfrak{n}(X)). The vector spaces 𝔱0\mathfrak{t}_{0} and 𝔱0\mathfrak{t}_{0}^{*} have no natural Q-structure (except when the null subgroup N(X)N(X) is closed), but as a substitute we have the quasi-lattice Λ=im(𝔛(T)𝔱0)\Lambda=\operatorname{\mathrm{im}}\bigl{(}\mathfrak{X}_{*}(T)\to\mathfrak{t}_{0}\bigr{)}. Here 𝔛(T)=ker(exp:𝔱T)\mathfrak{X}_{*}(T)=\ker(\exp\colon\mathfrak{t}\to T) is the exponential lattice of the maximal torus, and by a quasi-lattice in a vector space VV we mean a finitely generated additive subgroup that spans VV over R. The rank of the quasi-lattice Λ\Lambda is dim(𝔱0)\geq\dim(\mathfrak{t}_{0}), where equality holds if and only if 𝔱𝔫(X)\mathfrak{t}\cap\mathfrak{n}(X) is rational, i.e. N(X)N(X) is closed. If we regard the polyhedral set Δ(X)\Delta(X) as a subset of 𝔱0\mathfrak{t}_{0}^{*}, the normal vectors to its facets are in 𝔱0=𝔱0\mathfrak{t}_{\smash{0}}^{\smash{*}*}=\mathfrak{t}_{0}. It follows from \oldtagform@2.12.4 that these normal vectors are contained in the quasi-lattice Λ\Lambda.

2.13.2 Corollary.

Under the hypotheses of Theorem 2.2, the polyhedral subset Δ(X)𝔱0\Delta(X)\subseteq\mathfrak{t}_{0}^{*} is the intersection of a locally finite collection of halfspaces of the form η,a\langle\eta,\cdot\rangle\geq a, where aRa\in{\text{\bf R}} and η\eta is in the quasi-lattice Λ=im(𝔛(T)𝔱0)\Lambda=\operatorname{\mathrm{im}}\bigl{(}\mathfrak{X}_{*}(T)\to\mathfrak{t}_{0}\bigr{)}.

3. Morse functions and the abelian case

3.1. This section is a discussion of Morse-theoretic properties of presymplectic moment maps and of the presymplectic convexity theorem in the abelian case. The main result is Theorem 3.4.6, which asserts that the components of a presymplectic moment map are Morse-Bott functions under the assumption that the action is clean.

We keep the notational conventions of Section 2: XX is a manifold with presymplectic form ω\omega and GG is a compact connected Lie group acting on MM in a Hamiltonian fashion with moment map Φ\Phi. We denote the null foliation of ω\omega by \mathscr{F}, the null ideal of an open subset UU by 𝔫(U)\mathfrak{n}(U) and the null subgroup by N(U)N(U). We denote the leaf of xXx\in X by (x)\mathscr{F}(x). When we think of the leaf as a point in the leaf space X/X/\mathscr{F} we denote it by x¯\bar{x}.

3.2.  Let EE be a finite-dimensional real vector space and σ\sigma a presymplectic form on EE. We call an inner product on EE compatible with σ\sigma if on the linear subspace F=ker(σ)F=\ker(\sigma)^{\perp} orthogonal to ker(σ)\ker(\sigma) the inner product is compatible with the symplectic form σ|F\sigma|_{F} in the usual sense, namely that the endomorphism JJ of FF determined by u,v=σ(Ju,v)\langle u,v\rangle=\sigma(Ju,v) for all uu, vFv\in F defines an orthogonal complex structure. On the subspace FF the symplectic form and the compatible inner product combine to give a Hermitian inner product. Compatible inner products always exist and, if EE is a presymplectic HH-module for some compact Lie group HH, can be chosen to be HH-invariant. A choice of such an inner product makes the subspace FF a unitary HH-module.

3.3.  We say that a subset ZZ of XX is a submanifold at xXx\in X if xx has an open neighbourhood UU with the property that ZUZ\cap U is a submanifold of UU. Let f:XRf\colon X\to{\text{\bf R}} be smooth and let Γf={xXTxf=0}\Gamma_{f}=\{\,x\in X\mid T_{x}f=0\,\} be its critical set. We say a critical point xΓfx\in\Gamma_{f} is nondegenerate in the sense of Bott if Γf\Gamma_{f} is a submanifold at xx and the kernel of the Hessian Tx2f:TxXRT_{x}^{2}f\colon T_{x}X\to{\text{\bf R}} is equal to the tangent space TxΓfT_{x}\Gamma_{f}. The index of ff at xx is the dimension of a maximal negative definite subspace for the quadratic form Tx2fT_{x}^{2}f. We say ff is a Morse-Bott function if all its critical points are nondegenerate in the sense of Bott.

3.4.  For ξ𝔤\xi\in\mathfrak{g} we denote the critical set ΓΦξ\Gamma_{\Phi^{\xi}} of the function Φξ\Phi^{\xi} by X[ξ]X^{[\xi]}. For a subalgebra 𝔥\mathfrak{h} of 𝔤\mathfrak{g} we define the critical set of 𝔥\mathfrak{h} to be X[𝔥]=ξ𝔥X[ξ]X^{[\mathfrak{h}]}=\bigcap_{\xi\in\mathfrak{h}}X^{[\xi]}, the common critical set of the functions Φξ\Phi^{\xi} with ξ𝔥\xi\in\mathfrak{h}. In the symplectic case the critical set of 𝔥\mathfrak{h} is the fixed point manifold of the subgroup generated by 𝔥\mathfrak{h}. The next lemma says that for this to remain true in the presymplectic case we have to replace fixed points by fixed leaves.

3.4.1 Lemma.
  1. (i)

    Let 𝔥\mathfrak{h} be a Lie subalgebra of 𝔤\mathfrak{g} and HH the connected immersed subgroup of GG with Lie algebra 𝔥\mathfrak{h}. Then

    X[𝔥]={xX𝔥𝔤x¯}={xXHx¯=x¯}.X^{[\mathfrak{h}]}=\{\,x\in X\mid\mathfrak{h}\subseteq\mathfrak{g}_{\bar{x}}\,\}=\{\,x\in X\mid H\cdot\bar{x}=\bar{x}\,\}.
  2. (ii)

    X[ξ]={xXξ𝔤x¯}={xXξX(x)Tx}X^{[\xi]}=\{\,x\in X\mid\xi\in\mathfrak{g}_{\bar{x}}\,\}=\{\,x\in X\mid\xi_{X}(x)\in T_{x}\mathscr{F}\,\} for all ξ𝔤\xi\in\mathfrak{g}.

Proof.
\oldtagform@

i Let xXx\in X. Since HH is connected, we have Hx¯=x¯H\cdot\bar{x}=\bar{x} \iff HGx¯H\subseteq G_{\bar{x}} \iff 𝔥𝔤x¯\mathfrak{h}\subseteq\mathfrak{g}_{\bar{x}}, which proves the second equality. Moreover, Hx¯=x¯H\cdot\bar{x}=\bar{x} \iff Tx(Hx)TxT_{x}(H\cdot x)\subseteq T_{x}\mathscr{F}. Applying Lemma B.4\oldtagform@i to the subgroup HH we see that Tx(Hx)TxT_{x}(H\cdot x)\subseteq T_{x}\mathscr{F} is equivalent to TxΦξ=0T_{x}\Phi^{\xi}=0 for all ξ𝔥\xi\in\mathfrak{h}, which proves the first equality.

\oldtagform@

ii follows from \oldtagform@i applied to the Lie subalgebra spanned by ξ\xi. ∎

The critical set X[ξ]X^{[\xi]} is unaffected if we perturb ξ\xi in the direction of the null ideal. Specifically, if UU is an open subset and ζ\zeta is in the null ideal 𝔫(U)\mathfrak{n}(U), then dΦζ=0d\Phi^{\zeta}=0 on UU. Therefore

(3.4.2) X[ξ]U=X[η]UandTx2Φξ=Tx2ΦηX^{[\xi]}\cap U=X^{[\eta]}\cap U\quad\text{and}\quad T_{x}^{2}\Phi^{\xi}=T_{x}^{2}\Phi^{\eta}

for all xUx\in U and for all ξ\xi, η𝔤\eta\in\mathfrak{g} satisfying ξη𝔫(U)\xi-\eta\in\mathfrak{n}(U). Similarly,

(3.4.3) X[𝔥]U=X[𝔥+𝔫(U)]UX^{[\mathfrak{h}]}\cap U=X^{[\mathfrak{h}+\mathfrak{n}(U)]}\cap U

for all subalgebras 𝔥\mathfrak{h}. The critical set of 𝔥\mathfrak{h} is preserved by the subgroup HN(X)HN(X) generated by the Lie subalgebra 𝔥+𝔫(X)\mathfrak{h}+\mathfrak{n}(X).

The next assertion implies that if the GG-action is clean the critical set is a submanifold and its normal bundle is symplectic. This can easily be false without the cleanness assumption. (See 4 for counterexamples.)

3.4.4 Proposition.

Let 𝔥\mathfrak{h} be a Lie subalgebra of 𝔤\mathfrak{g} and let xX[𝔥]x\in X^{[\mathfrak{h}]}. Assume that the GG-action on XX is clean at xx. Then X[𝔥]X^{[\mathfrak{h}]} is a submanifold at xx. The subspace of TxXT_{x}X orthogonal to TxX[𝔥]T_{x}X^{[\mathfrak{h}]} with respect to a GxG_{x}-invariant compatible inner product on TxXT_{x}X is symplectic.

Proof.

By Lemmas 2.6.4\oldtagform@vi and 3.4.1\oldtagform@i the cleanness assumption implies that 𝔥𝔤x¯=𝔤x+𝔫x\mathfrak{h}\subseteq\mathfrak{g}_{\bar{x}}=\mathfrak{g}_{x}+\mathfrak{n}_{x}. That is to say, 𝔥+𝔫x=𝔣+𝔫x\mathfrak{h}+\mathfrak{n}_{x}=\mathfrak{f}+\mathfrak{n}_{x}, where 𝔣\mathfrak{f} is the subalgebra (𝔥+𝔫x)𝔤x(\mathfrak{h}+\mathfrak{n}_{x})\cap\mathfrak{g}_{x} of 𝔤x\mathfrak{g}_{x}. Choose an open neighbourhood UU of xx with 𝔫(U)=𝔫x\mathfrak{n}(U)=\mathfrak{n}_{x}. Then

X[𝔥]U=X[𝔥+𝔫x]U=X[𝔣+𝔫x]U=X[𝔣]UX^{[\mathfrak{h}]}\cap U=X^{[\mathfrak{h}+\mathfrak{n}_{x}]}\cap U=X^{[\mathfrak{f}+\mathfrak{n}_{x}]}\cap U=X^{[\mathfrak{f}]}\cap U

by \oldtagform@3.4.3, so we may just as well replace 𝔥\mathfrak{h} with 𝔣\mathfrak{f}. The advantage of the subalgebra 𝔣\mathfrak{f} is that it fixes xx and therefore acts linearly in a GxG_{x}-equivariant Darboux chart centred at xx. More precisely, the equivariant Darboux Theorem, Corollary 2.10.4, allows us to replace XX with the presymplectic GxG_{x}-module E=TxXE=T_{x}X and the functions Φη\Phi^{\eta} for η𝔣\eta\in\mathfrak{f} with the quadratic forms Tx2ΦηT_{x}^{2}\Phi^{\eta}. Choosing a GxG_{x}-invariant compatible inner product on EE, we have that E=E0E1E=E_{0}\oplus E_{1} is an orthogonal direct sum of a GxG_{x}-module E0E_{0} and a unitary GxG_{x}-module E1E_{1}. The critical set is then X[𝔣]=E0E1𝔣X^{[\mathfrak{f}]}=E_{0}\oplus E_{1}^{\mathfrak{f}}, where E1𝔣E_{1}^{\mathfrak{f}} denotes the 𝔣\mathfrak{f}-fixed subspace of E1E_{1}. This shows that X[𝔣]X^{[\mathfrak{f}]} is a submanifold. The orthogonal complement of X[𝔣]X^{[\mathfrak{f}]} is a unitary submodule of E1E_{1}, and in particular it is symplectic. ∎

Taking 𝔥=Rξ\mathfrak{h}={\text{\bf R}}\xi gives the following result.

3.4.5 Theorem.

Let ξ𝔤\xi\in\mathfrak{g} and let xX[ξ]x\in X^{[\xi]} be a critical point of Φξ\Phi^{\xi}. Assume that the GG-action on XX is clean at xx. Then xx is nondegenerate in the sense of Bott. Choose a GxG_{x}-invariant compatible inner product on TxXT_{x}X. Then the positive and negative subspaces of Tx2ΦξT_{x}^{2}\Phi^{\xi} are symplectic subspaces of TxXT_{x}X. In particular the index of Φξ\Phi^{\xi} at xx is even.

Proof.

It follows from Proposition 3.4.4 that X[ξ]X^{[\xi]} is a submanifold at xx. We argue nondegeneracy by writing ξ=η+ζ\xi=\eta+\zeta with η𝔤x\eta\in\mathfrak{g}_{x} and ζ𝔫x\zeta\in\mathfrak{n}_{x}. Then X[ξ]=X[η]X^{[\xi]}=X^{[\eta]} near xx and Tx2Φξ=Tx2ΦηT_{x}^{2}\Phi^{\xi}=T_{x}^{2}\Phi^{\eta} because of \oldtagform@3.4.2, and the vector field ηX\eta_{X} is linear in an equivariant Darboux chart at xx. With the same notation as in the proof of the proposition, the subspace of TxXT_{x}X orthogonal to X[η]X^{[\eta]} is the sum of the nonzero weight spaces of the unitary module E1E_{1}. In other words, (TxX[η])=E1+E1(T_{x}X^{[\eta]})^{\perp}=E_{1}^{+}\oplus E_{1}^{-}, where E1+E_{1}^{+}, resp. E1E_{1}^{-}, is spanned by all positive, resp. negative, weight vectors, i.e. vectors eE1e\in E_{1} satisfying ξ(e)=1αe\xi(e)=\sqrt{-1}\alpha e for some α>0\alpha>0, resp. α<0\alpha<0. On E1+E_{1}^{+} the Hessian of Φη\Phi^{\eta} is positive definite, on E1E_{1}^{-} it is negative definite. ∎

Assuming cleanness at all points of XX we obtain the next statement, which is the main result of this section.

3.4.6 Theorem.

Assume that the GG-action on XX is clean. Then for every ξ𝔤\xi\in\mathfrak{g} the component Φξ\Phi^{\xi} of the moment map is a Morse-Bott function. The positive and negative normal bundles of the critical set X[ξ]X^{[\xi]}, taken with respect to a GG-invariant compatible Riemannian metric on XX, are symplectic subbundles of TXTX orthogonal to the subbundle TT\mathscr{F}.

In the symplectic case this result goes back to Atiyah [1] and Guillemin and Sternberg [8]. For leafwise transitive Hamiltonian circle actions on K-contact manifolds the result was proved by Rukimbira [23]. His result was extended to leafwise transitive presymplectic Hamiltonian torus actions by Ishida [14].

3.5.  If GG is a torus and XX is compact symplectic, then the vertices of the moment polytope Φ(X)\Phi(X) are images of GG-fixed points. In the presymplectic case there may not be any fixed points. Instead one needs to consider the critical points of the moment map, or equivalently the GG-fixed leaves. Moreover, we can weaken the assumption that GG is abelian to the assumption that the quotient G/N(X)G/N(X) is abelian, or equivalently that the null ideal 𝔫(X)\mathfrak{n}(X) contains the derived subalgebra of 𝔤\mathfrak{g}. By Proposition 2.9.1, then the moment map image is contained in 𝔷(𝔤)\mathfrak{z}(\mathfrak{g})^{*}, the dual of the centre 𝔷(𝔤)\mathfrak{z}(\mathfrak{g}) of 𝔤\mathfrak{g}, and we have Δ(X)=Φ(X)\Delta(X)=\Phi(X).

3.5.1 Theorem (abelian convexity).

Assume that the null ideal 𝔫(X)\mathfrak{n}(X) contains the derived subalgebra [𝔤,𝔤][\mathfrak{g},\mathfrak{g}], that XX is compact, and that the GG-action on XX is clean. Then Φ(X[𝔤])\Phi\bigl{(}X^{[\mathfrak{g}]}\bigr{)} is a finite subset of 𝔤\mathfrak{g}^{*} and Φ(X)\Phi(X) is the convex hull of Φ(X[𝔤])\Phi\bigl{(}X^{[\mathfrak{g}]}\bigr{)}. For every vertex λ\lambda of Φ(X)\Phi(X) the fibre Φ1(λ)\Phi^{-1}(\lambda) is a connected component of X[𝔤]X^{[\mathfrak{g}]}.

Proof.

It follows from Theorem 2.2 that Φ(X)\Phi(X) is a convex polytope and therefore equal to the convex hull of its vertices. It follows from Proposition 3.4.4 that X[𝔤]X^{[\mathfrak{g}]} is a closed submanifold of XX, and therefore has a finite number of connected components. The moment map is constant on each component, so Φ(X[𝔤])\Phi(X^{[\mathfrak{g}]}) is finite. Let λ\lambda be a vertex of Φ(X)\Phi(X) and let xΦ1(λ)x\in\Phi^{-1}(\lambda). Then there exists an open subset Ξ\Xi of 𝔱\mathfrak{t} with the property that for every ξΞ\xi\in\Xi the function Φξ\Phi^{\xi} attains its global minimum at xx. Hence TxΦξ=0T_{x}\Phi^{\xi}=0 and Tx2ΦξT_{x}^{2}\Phi^{\xi} is positive semidefinite for all ξΞ\xi\in\Xi. Because Ξ\Xi spans 𝔱\mathfrak{t}, this implies that TxΦ=0T_{x}\Phi=0, i.e. xX[𝔤]x\in X^{[\mathfrak{g}]}. By Theorem 3.4.6, for all ξΞ\xi\in\Xi the Hessian of Φξ\Phi^{\xi} at xx is positive definite in the direction normal to X[𝔤]X^{[\mathfrak{g}]}. Computing in an equivariant Darboux chart UU at xx we see that the portion UΦ1(λ)U\cap\Phi^{-1}(\lambda) of the fibre Φ1(λ)\Phi^{-1}(\lambda) is contained in X[𝔤]X^{[\mathfrak{g}]}. Therefore the entire fibre is contained in X[𝔤]X^{[\mathfrak{g}]}. Since the fibre is connected (Theorem 2.2), it is equal to a component of X[𝔤]X^{[\mathfrak{g}]}. ∎

4. Examples

4.1. In this section GG denotes a compact connected Lie group, 𝔱\mathfrak{t} a Cartan subalgebra of 𝔤\mathfrak{g} and CC a chamber in 𝔱\mathfrak{t}^{*}.

4.2. Failure of convexity.  Z. He [10, Ch. 4] gave the first example of a presymplectic Hamiltonian torus action with a nonconvex moment map image. Here we show that such examples are ubiquitous. In particular presymplectic Hamiltonian actions are typically not clean. Our starting point is the following elementary fact, which is implicit in [9, Theorem 26.4]. (Cf. also Proposition 2.11.2.)

4.2.1 Lemma.

Let (M,πM)(M,\pi_{M}) and (N,πN)(N,\pi_{N}) be Poisson manifolds and ϕ:MN\phi\colon M\to N a Poisson morphism. Let YY be a coisotropic submanifold of NN that intersects ϕ\phi cleanly. Then X=ϕ1(Y)X=\phi^{-1}(Y) is a coisotropic submanifold of MM.

Proof.

That ϕ\phi is a Poisson morphism means by definition that the square

TM\textstyle{TM\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Tϕ\scriptstyle{T\phi}ϕTN\textstyle{\phi^{*}TN}TM\textstyle{T^{*}M\ignorespaces\ignorespaces\ignorespaces\ignorespaces}πM\scriptstyle{\pi_{M}^{\sharp}}ϕTN\textstyle{\phi^{*}T^{*}N\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Tϕ\scriptstyle{T^{*}\phi}ϕπN\scriptstyle{\phi^{*}\pi_{N}^{\sharp}}

commutes. That YY is coisotropic means that πN(TY)\pi_{N}^{\sharp}(T^{\circ}Y) is contained in TYTY, where TYT^{\circ}Y denotes the annihilator of TYTY in TM|YT^{*}M|_{Y}. That YY intersects ϕ\phi cleanly means that XX is a submanifold with tangent bundle TX=(Tϕ)1(ϕTY)TX=(T\phi)^{-1}(\phi^{*}TY). Therefore the annihilator of TXTX is TX=Tϕ(ϕTY)T^{\circ}X=T^{*}\phi(\phi^{*}T^{\circ}Y). From πN(TY)TY\pi_{N}^{\sharp}(T^{\circ}Y)\subseteq TY we infer

TϕπMTϕ(ϕTY)ϕTY,T\phi\circ\pi_{M}^{\sharp}\circ T^{*}\phi(\phi^{*}T^{\circ}Y)\subseteq\phi^{*}TY,

and hence Tϕ(πM(TX))ϕTYT\phi(\pi_{M}^{\sharp}(T^{\circ}X))\subseteq\phi^{*}TY. We conclude that πM(TX)TX\pi_{M}^{\sharp}(T^{\circ}X)\subseteq TX. ∎

Let us apply this result to a symplectic Hamiltonian GG-manifold MM with symplectic form ωM\omega_{M}, taking NN to be the linear Poisson manifold 𝔤\mathfrak{g}^{*} and ϕ\phi to be the moment map. Any GG-invariant submanifold YY of 𝔤\mathfrak{g}^{*} is coisotropic. Therefore, if YY intersects ϕ\phi cleanly, its preimage X=ϕ1(Y)X=\phi^{-1}(Y) is a coisotropic submanifold of MM. It follows that the closed 22-form ω=ωM|X\omega=\omega_{M}|_{X} has constant corank equal to the codimension of XX in MM. The moment map is equivariant, so XX is preserved by the GG-action, and the GG-action on XX is Hamiltonian with moment map Φ=ϕ|M\Phi=\phi|_{M}. Thus XX is a presymplectic Hamiltonian GG-manifold. Its moment map image is

Φ(X)=ϕ(ϕ1(Y))=Yϕ(M).\Phi(X)=\phi(\phi^{-1}(Y))=Y\cap\phi(M).

It is easy to choose YY in such a manner that Δ(X)=Φ(X)C\Delta(X)=\Phi(X)\cap C is not convex.

For a specific example let T=R/2πZ{\text{\bf T}}={\text{\bf R}}/2\pi{\text{\bf Z}} be the circle and let G=TdG={\text{\bf T}}^{d} be the dd-torus acting on M=CdM={\text{\bf C}}^{d} in the standard way,

tx=(t1,t2,,td)(x1,x2,,xd)=(eit1x1,eit2x2,,eitdxd).t\cdot x=(t_{1},t_{2},\dots,t_{d})\cdot(x_{1},x_{2},\dots,x_{d})=\bigl{(}e^{it_{1}}x_{1},e^{it_{2}}x_{2},\dots,e^{it_{d}}x_{d}\bigr{)}.

Then 𝔤=Rd\mathfrak{g}={\text{\bf R}}^{d} and 𝔤=(Rd)Rd\mathfrak{g}^{*}=({\text{\bf R}}^{d})^{*}\cong{\text{\bf R}}^{d}, and the map ϕ:CdRd\phi\colon{\text{\bf C}}^{d}\to{\text{\bf R}}^{d} defined by ϕ(x)=12(|x1|2,|x2|2,,|xd|2)\phi(x)=\frac{1}{2}\bigl{(}\lvert x_{1}\rvert^{2},\lvert x_{2}\rvert^{2},\dots,\lvert x_{d}\rvert^{2}\bigr{)} is a moment map for this action with respect to the standard symplectic form ω0=12ijdxjdx¯j\omega_{0}=\frac{1}{2i}\sum_{j}dx_{j}\wedge d\bar{x}_{j}. Let YY be a submanifold of Rd{\text{\bf R}}^{d} of codimension kk which is transverse to the faces of the orthant R0d{\text{\bf R}}^{d}_{\geq 0}. Then YY is transverse to ϕ\phi, so X=ϕ1(Y)X=\phi^{-1}(Y) is a submanifold of Cd{\text{\bf C}}^{d} of (real) codimension kk. Obviously YY is coisotropic with respect to the zero Poisson structure, so XX is a presymplectic Hamiltonian GG-manifold with presymplectic form of corank kk. Its moment map image is the intersection of YY with the orthant,

(4.2.2) Φ(X)=ϕ(ϕ1(Y))=YR0d,\Phi(X)=\phi(\phi^{-1}(Y))=Y\cap{\text{\bf R}}^{d}_{\geq 0},

which is a dkd-k-manifold with corners, but is of course seldom convex.

For instance, the following curve in the positive quadrant (d=2d=2, k=1k=1) is the moment map image of a T2{\text{\bf T}}^{2}-action on a presymplectic 33-sphere.

[Uncaptioned image]

This class of examples displays some other phenomena of interest, such as the existence of nontrivial deformations of presymplectic Hamiltonian actions. The equivariant Darboux theorem implies that symplectic Hamiltonian GG-manifolds cannot be continuously deformed locally near any point. The Moser stability theorem implies that the same is true globally for compact symplectic Hamiltonian GG-manifolds as long as we move the symplectic form within a fixed cohomology class. We now show that both these statements are false for presymplectic Hamiltonian manifolds. (However, we will prove in Appendix C that there is a presymplectic equivariant Darboux theorem under the assumption that the action is clean.) Take any isotopic family (Yt)0t1(Y_{t})_{0\leq t\leq 1} of compact submanifolds of Rd{\text{\bf R}}^{d}, all of which are transverse to ϕ\phi. Then the manifolds Xt=ϕ1(Yt)X_{t}=\phi^{-1}(Y_{t}) form an isotopic family of compact submanifolds of Cd{\text{\bf C}}^{d}, each of which is a presymplectic Hamiltonian GG-manifold with presymplectic form ωt=ω0|Xt\omega_{t}=\omega_{0}|_{X_{t}} and moment map Φt=ϕ|Xt\Phi_{t}=\phi|_{X_{t}}. The forms ωt\omega_{t} on XtX_{t} are exact for all tt. The fibres XtX_{t} are equivariantly diffeomorphic, but they are usually not isomorphic as presymplectic GG-manifolds. Indeed, if there existed GG-equivariant diffeomorphisms ft:X0Xtf_{t}\colon X_{0}\to X_{t} satisfying ftωt=ω0f_{t}^{*}\omega_{t}=\omega_{0}, then we would have ftΦt=Φ0+λtf_{t}^{*}\Phi_{t}=\Phi_{0}+\lambda_{t} for some λt𝔤\lambda_{t}\in\mathfrak{g}^{*}, and therefore Φt(Xt)=Φ0(X0)+λt\Phi_{t}(X_{t})=\Phi_{0}(X_{0})+\lambda_{t}, which by \oldtagform@4.2.2 would imply YtR0=Y0R0+λtY_{t}\cap{\text{\bf R}}_{\geq 0}=Y_{0}\cap{\text{\bf R}}_{\geq 0}+\lambda_{t}. So by choosing the manifolds YtY_{t} in such a way that the intersections YtR0dY_{t}\cap{\text{\bf R}}^{d}_{\geq 0} are not translates of each other we can guarantee that the presymplectic GG-manifolds XtX_{t} are not all isomorphic.

For instance, the 33-sphere represented by the picture above can be smoothly deformed to an ellipsoid, the moment map image of which is an interval. This deformation is not equivariantly presymplectically trivial.

[Uncaptioned image]

Another feature of these examples is that the components of the moment map Φ\Phi are often not Morse-Bott functions. For instance, let us take YY to be a smooth curve in the plane R2{\text{\bf R}}^{2} with the following properties: (1) YY is transverse to the coordinate axes; (2) YR02Y\cap{\text{\bf R}}^{2}_{\geq 0} is compact; and (3) the set CC consisting of all points yYy\in Y with horizontal tangent line TyYT_{y}Y is countably infinite and is contained in the open orthant R>02{\text{\bf R}}^{2}_{>0}. Then X=ϕ1(Y)X=\phi^{-1}(Y) is a compact real hypersurface in C2{\text{\bf C}}^{2}. The critical set of the second component of the moment map Φ=(Φ1,Φ2):XR2\Phi=(\Phi_{1},\Phi_{2})\colon X\to{\text{\bf R}}^{2} is equal to Φ1(C)=yCΦ1(y)\Phi^{-1}(C)=\bigcup_{y\in C}\Phi^{-1}(y), an infinite disjoint union of codimension 11 submanifolds of XX, and therefore is not a submanifold.

4.3. Prato’s toric quasifolds.  We continue the discussion of the previous subsection, but now we take the submanifold YY in \oldtagform@4.2.2 to be an affine subspace of Rd{\text{\bf R}}^{d} transverse to the orthant. Then the intersection P=Φ(X)=YR0dP=\Phi(X)=Y\cap{\text{\bf R}}^{d}_{\geq 0} is a convex polyhedron. The transversality to the orthant is equivalent to PP being simple, i.e. the link of each of its faces being a simplex. It follows from Proposition 2.11.2 that the action of G=TdG={\text{\bf T}}^{d} on the coisotropic submanifold XX of Cd{\text{\bf C}}^{d} is leafwise transitive and that its null ideal 𝔫(X)\mathfrak{n}(X) is the linear subspace of 𝔤=Rd\mathfrak{g}={\text{\bf R}}^{d} orthogonal to YY. Prato [21] calls the leaf space X/N(X)X/N(X) a toric quasifold associated with PP. It carries an action of the quotient group Td/N(X){\text{\bf T}}^{d}/N(X) and a moment map whose image is PP and whose fibres are the Td/N(X){\text{\bf T}}^{d}/N(X)-orbits. See [22] for a classification of toric quasifolds in terms of simple polyhedra.

4.4. Orbifolds.  As an illustration of our convexity theorem we present a new proof of a result due to Lerman et al. [18, Theorem 1.1].

4.4.1 Theorem.

Let (M,ωM)(M,\omega_{M}) be a symplectic orbifold equipped with a Hamiltonian GG-action and a proper moment map ΦM\Phi_{M}. Then Δ(M)=ΦM(M)C\Delta(M)=\Phi_{M}(M)\cap C is a rational convex polyhedral set.

Proof.

Let MeffM_{\textrm{eff}} be the effective orbifold which underlies MM, as defined e.g. in [11]. The symplectic structure and the Hamiltonian action descend to MeffM_{\textrm{eff}}, and MeffM_{\textrm{eff}} has the same moment map image as MM. So we may assume without loss of generality that MM is effective. Choose a GG-invariant Riemannian metric on MM compatible with the symplectic form ωM\omega_{M}. This choice endows the tangent bundle TMTM with the structure of a Hermitian orbifold vector bundle. Let XX be the unitary frame bundle of TMTM, which is an orbifold principal bundle over MM with structure group U(n){\text{\bf U}}(n), where n=12dim(M)n=\frac{1}{2}\dim(M). Then XX is a smooth manifold (see e.g. [20, § 2.4]), every diffeomorphism of MM lifts naturally to a U(n){\text{\bf U}}(n)-equivariant diffeomorphism of XX, and the GG-action lifts to a GG-action on XX which commutes with the U(n){\text{\bf U}}(n)-action. Let p:XMp\colon X\to M be the projection, ω=pωM\omega=p^{*}\omega_{M} and Φ=pΦM\Phi=p^{*}\Phi_{M}. Then XX is a presymplectic Hamiltonian GG-manifold with proper moment map Φ\Phi. The leaves of the null foliation of XX are the U(n){\text{\bf U}}(n)-orbits. The GG-action on XX is typically not clean (cf. Example 2.6.5), but the action of G^=G×U(n)\hat{G}=G\times{\text{\bf U}}(n) is leafwise transitive and has moment map Φ^=Φ×0:X𝔤^=𝔤𝔲(n)\hat{\Phi}=\Phi\times 0\colon X\to\hat{\mathfrak{g}}^{*}=\mathfrak{g}^{*}\oplus\mathfrak{u}(n)^{*}. Since the null foliation has closed leaves, the null subgroup N^(X)\hat{N}(X) for the G^\hat{G}-action is a closed subgroup of G^\hat{G}, which contains the subgroup U(n){\text{\bf U}}(n). We conclude from Theorem 2.2 that Δ(M)=Φ(X)C=Φ^(X)C\Delta(M)=\Phi(X)\cap C=\hat{\Phi}(X)\cap C is a rational convex polyhedral set. ∎

4.5. Contact manifolds.  Let XX be a compact manifold and let α\alpha be a contact 11-form on XX. The exact 22-form ω=dα\omega=-d\alpha is nondegenerate on the contact hyperplane bundle ker(α)\ker(\alpha) and therefore is a presymplectic form of corank 11. The null foliation \mathscr{F} of ω\omega is spanned by the Reeb vector field, which is by definition the unique vector field ρ\rho with the properties ι(ρ)ω=0\iota(\rho)\omega=0 and ι(ρ)α=1\iota(\rho)\alpha=1. Suppose that GG acts on XX and leaves α\alpha invariant. Then the action is Hamiltonian with moment map Φ:X𝔤\Phi\colon X\to\mathfrak{g}^{*} defined by Φξ=ι(ξX)α\Phi^{\xi}=\iota(\xi_{X})\alpha. The action is said to be of Reeb type (see e.g. Boyer and Galicki [4, § 8.4.2]) if there exists a Lie algebra element ξ𝔤\xi\in\mathfrak{g} with the property that ξX=ρ\xi_{X}=\rho. Reeb-type actions are leafwise transitive. The next result follows immediately from the presymplectic convexity theorem.

4.5.1 Theorem (contact convexity theorem I).

Suppose that the GG-action on XX is clean (e.g. of Reeb type). Then Δ(X)=Φ(X)C\Delta(X)=\Phi(X)\cap C is a convex polytope.

There are many compact contact Hamiltonian GG-manifolds XX for which Δ(X)\Delta(X) is not convex (and hence the action is not clean). Here are two methods to produce examples of this. (1) Start with a pair (X,α)(X,\alpha) for which Δ(X)\Delta(X) is convex and then replace α\alpha by a conformally equivalent contact form efαe^{f}\alpha for some GG-invariant smooth function ff. This has the effect of multiplying the moment map Φ\Phi by efe^{f}, which will usually destroy the convexity of Δ(X)\Delta(X). (2) Follow the method of 4, starting with a hypersurface YY in Rd{\text{\bf R}}^{d} which is transverse to the faces of the orthant R0d{\text{\bf R}}^{d}_{\geq 0} as well as to the radial vector field on Rd{\text{\bf R}}^{d}. The moment map ϕ\phi for the standard Td{\text{\bf T}}^{d}-action on Cd{\text{\bf C}}^{d} is quadratic, and therefore maps the radial vector field on Cd{\text{\bf C}}^{d} to twice the radial vector field on Rd{\text{\bf R}}^{d}. It follows that the hypersurface X=ϕ1(Y)X=\phi^{-1}(Y) is transverse to the radial vector field on Cd{\text{\bf C}}^{d}, which is Liouville, and hence XX is of contact type. Unless YY is an affine hyperplane (in which case XX is a contact ellipsoid and Φ(X)\Phi(X) a simplex), the image Φ(X)=YR0d\Phi(X)=Y\cap{\text{\bf R}}^{d}_{\geq 0} is not convex.

It is instructive to compare and contrast our contact convexity theorem, Theorem 4.5.1, with a theorem of Lerman [16], as improved by Chiang and Karshon [6]. Our result regards the symplectic leaf space X/X/\mathscr{F} of XX. Their result regards the symplectization M=X×(0,)M=X\times(0,\infty) of XX, which carries the symplectic form ωM=d(tα)=tωdtα\omega_{M}=-d(t\alpha)=t\omega-dt\wedge\alpha, where tt is the coordinate on the interval (0,)(0,\infty). Letting GG act trivially on the second factor, we get a Hamiltonian GG-action on MM with moment map ΦM(x,t)=tΦ(x)\Phi_{M}(x,t)=t\Phi(x). The image of ΦM\Phi_{M} is the conical set ΦM(M)=t>0tΦ(X)\Phi_{M}(M)=\bigcup_{t>0}t\Phi(X). Intersecting the image with the chamber CC and adding the origin defines a subset Δ(M)={0}(ΦM(M)C)\Delta(M)=\{0\}\cup\bigl{(}\Phi_{M}(M)\cap C\bigr{)} of CC, which is the union of all dilatations of Δ(X)\Delta(X),

Δ(M)=t0tΔ(X).\Delta(M)=\bigcup_{t\geq 0}t\Delta(X).

The Lerman-Chiang-Karshon theorem states that, if GG is a torus of dimension 2\geq 2, then Δ(M)\Delta(M) is a rational convex polyhedral cone. They make no cleanness hypothesis on the action. In fact, by either of the methods discussed in the previous paragraph one can manufacture examples where Δ(M)\Delta(M) is convex but Δ(X)\Delta(X) is not.

Another difference between the two contact convexity theorems is that, as noted above, the image Φ(X)\Phi(X) is highly dependent on the choice of the contact form α\alpha. In contrast, the symplectic cone (M,ωM)(M,\omega_{M}) is an intrinsic invariant of the contact hyperplane bundle ker(α)\ker(\alpha), from which it follows that its moment cone Δ(M)\Delta(M) depends only on the conformal class of α\alpha.

Nevertheless, our Theorem 4.5.1 is not wholly independent from the Lerman-Chiang-Karshon theorem. To see the connection, note that the Reeb vector field ρ\rho satisfies

ι(ρ)ωM=tι(ρ)dαι(ρ)(dtα)=0+dtι(ρ)α=dt,\iota(\rho)\omega_{M}=-t\iota(\rho)d\alpha-\iota(\rho)(dt\wedge\alpha)=0+dt\wedge\iota(\rho)\alpha=dt,

in other words is the Hamiltonian vector field of the radial function tt on the symplectic cone MM. Since ρ\rho is GG-invariant, this makes MM a Hamiltonian G^\hat{G}-manifold for the product G^=G×R\hat{G}=G\times{\text{\bf R}} with moment map Φ^:M𝔤×R\hat{\Phi}\colon M\to\mathfrak{g}^{*}\times{\text{\bf R}} defined by Φ^(x,t)=(tΦ(x),t)\hat{\Phi}(x,t)=(t\Phi(x),t). Putting Δ^(M)={0}(Φ^(M)(C×R))C×R\hat{\Delta}(M)=\{0\}\cup\bigl{(}\hat{\Phi}(M)\cap(C\times{\text{\bf R}})\bigr{)}\subseteq C\times{\text{\bf R}} we find

(4.5.2) Δ^(M)=t0(tΔ(X)×{t}).\hat{\Delta}(M)=\bigcup_{t\geq 0}\bigl{(}t\Delta(X)\times\{t\}\bigr{)}.

The set Δ(X)\Delta(X) is obtained by intersecting Δ^(M)\hat{\Delta}(M) with the hyperplane t=1t=1, which is as it should be, because the symplectic quotient of MM at level 11 with respect to the R-action is the leaf space X/X/\mathscr{F}. The cone Δ(M)\Delta(M) is obtained by restricting the G^\hat{G}-action to GG, i.e. by projecting Δ^(M)\hat{\Delta}(M) along the R-axis. Thus we obtain the following nonabelian extension of the Lerman-Chiang-Karshon theorem, which appears to be new.

4.5.3 Theorem (contact convexity theorem II).

Suppose that the GG-action on XX is clean. Then Δ^(M)\hat{\Delta}(M) and Δ(M)\Delta(M) are convex polyhedral cones.

Proof.

Theorem 4.5.1 gives that Δ(X)\Delta(X) is a convex polytope. Let η1\eta_{1}, η2,\eta_{2},\dots, ηk𝔱\eta_{k}\in\mathfrak{t} be inward-pointing normal vectors to the facets of Δ(X)\Delta(X); then points vΔ(X)v\in\Delta(X) are determined by inequalities of the form ηi,vai\langle\eta_{i},v\rangle\geq a_{i} with aiRa_{i}\in{\text{\bf R}}. It then follows from \oldtagform@4.5.2 that points v^=(tv,t)Δ^(M)\hat{v}=(tv,t)\in\hat{\Delta}(M) are determined by the homogeneous inequalities ηi,tvait0\langle\eta_{i},tv\rangle-a_{i}t\geq 0, i.e. η^i,v^0\langle\hat{\eta}_{i},\hat{v}\rangle\geq 0, where η^i=(ηi,ai)𝔱×R\hat{\eta}_{i}=(\eta_{i},-a_{i})\in\mathfrak{t}\times{\text{\bf R}}. Hence Δ^(M)\hat{\Delta}(M) is a convex polyhedral cone. Hence its projection Δ(M)\Delta(M) onto 𝔱\mathfrak{t}^{*} is likewise a convex polyhedral cone. ∎

Unfortunately our proof does not enable us to show that the cone Δ(M)\Delta(M) is rational, nor that it is convex if the action is not clean.

Appendix A Immersed normal subgroups

Let GG be a connected compact Lie group. What immersed connected normal Lie subgroups NN does GG have? “Not very many” is the answer. There are two basic types of such immersions: (1) GG is semisimple and simply connected and NN is a product of simple factors of GG; (2) GG is a torus and NN is a product of a torus and a vector space immersing into GG. Type (1) is a closed embedding, but type (2) may not be. We have the following straightforward result.

A.1 Lemma.

Every immersed connected normal Lie subgroup NN of GG is, up to finite covering groups, a product of types (1) and (2).

Proof.

The Lie algebra 𝔤=Lie(G)\mathfrak{g}=\operatorname{\mathrm{Lie}}(G) is the direct sum of the derived subalgebra 𝔤1=[𝔤,𝔤]\mathfrak{g}_{1}=[\mathfrak{g},\mathfrak{g}] and the centre 𝔤2=𝔷(𝔤)\mathfrak{g}_{2}=\mathfrak{z}(\mathfrak{g}). The ideal 𝔫=Lie(N)\mathfrak{n}=\operatorname{\mathrm{Lie}}(N) is the direct sum of the ideals 𝔫1=𝔫𝔤1\mathfrak{n}_{1}=\mathfrak{n}\cap\mathfrak{g}_{1} and 𝔫2=𝔫𝔤2\mathfrak{n}_{2}=\mathfrak{n}\cap\mathfrak{g}_{2}. The ideal 𝔫1\mathfrak{n}_{1} is a direct sum of simple ideals of the semisimple Lie algebra 𝔤1\mathfrak{g}_{1}. (See e.g. [3, § I.6].) Letting G1G_{1} and N1N_{1} be the corresponding simply connected groups, we have an embedding N1G1N_{1}\to G_{1} of type (1). Letting G2G_{2} be the identity component of the centre Z(G)Z(G) and N2=exp(𝔫2)N_{2}=\exp(\mathfrak{n}_{2}), we have an immersion N2G2N_{2}\to G_{2} of type (2). The product G1×G2G_{1}\times G_{2} is a finite covering group of GG and N1×N2N_{1}\times N_{2} is a finite covering group of the identity component of NN. ∎

The following consequence is used in the proof of Theorem 2.2\oldtagform@iii. As usual, by a rational subspace of the Lie algebra of a torus we mean a subspace that is rational with respect to the Q-structure defined by the character lattice of the torus.

A.2 Corollary.

Let NN be an immersed normal Lie subgroup of GG and 𝔱\mathfrak{t} a Cartan subalgebra of 𝔤\mathfrak{g}. The following statements are equivalent:

  1. (i)

    NN is closed;

  2. (ii)

    𝔫𝔷(𝔤)\mathfrak{n}\cap\mathfrak{z}(\mathfrak{g}) is a rational subspace of 𝔷(𝔤)\mathfrak{z}(\mathfrak{g});

  3. (iii)

    𝔫𝔱\mathfrak{n}\cap\mathfrak{t} is a rational subspace of 𝔱\mathfrak{t}.

Appendix B Some presymplectic linear algebra

This appendix lists a few elementary facts referred to in the proof of the convexity theorem. Let GG be a fixed connected compact Lie group. Let EE be a finite-dimensional real vector space and FF a linear subspace. We denote by FEF^{\circ}\subseteq E^{*} the annihilator of FF. Let σ\sigma be a presymplectic form on EE. We denote by

Fσ={uEσ(u,v)=0 for all vF}F^{\sigma}=\{\,u\in E\mid\text{$\sigma(u,v)=0$ for all $v\in F$}\,\}

the subspace of EE orthogonal to FF with respect to σ\sigma. We call EE a presymplectic GG-module if GG acts smoothly and linearly on EE and the GG-action preserves σ\sigma. We state the following simple result without proof.

B.1 Lemma.

A presymplectic GG-module (E,σ)(E,\sigma) is a presymplectic Hamiltonian GG-manifold with moment map ΦE\Phi_{E} given by ΦEξ(e)=12σ(ξ(e),e)\Phi_{E}^{\xi}(e)=\frac{1}{2}\sigma(\xi(e),e), where eξ(e)e\mapsto\xi(e) denotes the action of ξ𝔤\xi\in\mathfrak{g} on EE.

We require the following simple lemma concerning linear GG-actions.

B.2 Lemma.

Let EE be finite-dimensional real GG-module. Let FF be a GG-submodule and eEe\in E. Then the following conditions are equivalent.

  1. (i)

    e=e0+e1e=e_{0}+e_{1} for some e0EGe_{0}\in E^{G} and some e1Fe_{1}\in F;

  2. (ii)

    Gee+FG\cdot e\subseteq e+F;

  3. (iii)

    the affine subspace e+Fe+F is preserved by the GG-action;

  4. (iv)

    Te(Ge)FT_{e}(G\cdot e)\subseteq F.

Proof.

First we prove \oldtagform@ii \implies \oldtagform@i (which is the only implication that requires the compactness of GG). Let dgdg be the normalized Haar measure on GG and put e0=Gge𝑑ge_{0}=\int_{G}g\cdot e\,dg. Then e0e_{0} is fixed under the action. Since gee+Fg\cdot e\in e+F for all gGg\in G, we have ϕ(ge)=ϕ(e)\phi(g\cdot e)=\phi(e) for all ϕF\phi\in F^{\circ}. Therefore

ϕ(e0)=Gϕ(ge)𝑑g=Gϕ(e)𝑑g=ϕ(e)\phi(e_{0})=\int_{G}\phi(g\cdot e)\,dg=\int_{G}\phi(e)\,dg=\phi(e)

for all ϕF\phi\in F^{\circ}. This shows that ee0Fe-e_{0}\in F, which proves \oldtagform@i. Next we prove \oldtagform@iv \implies \oldtagform@ii (which is the only implication that requires the connectedness of GG). If Te(Ge)FT_{e}(G\cdot e)\subseteq F, then Tf(Ge)FT_{f}(G\cdot e)\subseteq F for all fGef\in G\cdot e, because the action preserves FF. Hence the orbit GeG\cdot e is everywhere tangent to the foliation of EE given by the affine subspaces parallel to FF. Therefore GeG\cdot e is contained in the leaf e+Fe+F. The other implications are straightforward. ∎

B.3 Lemma.

Let 𝔞\mathfrak{a} be an ideal of 𝔤\mathfrak{g} and λ𝔤\lambda\in\mathfrak{g}^{*}. The coadjoint orbit GλG\cdot\lambda is contained in the affine subspace λ+𝔞\lambda+\mathfrak{a}^{\circ} if and only if 𝔞\mathfrak{a} is contained in the centralizer 𝔤λ\mathfrak{g}_{\lambda} of λ\lambda.

Proof.

Apply Lemma B.2 to the module E=𝔤E=\mathfrak{g}^{*} and the submodule F=𝔞F=\mathfrak{a}^{\circ} and use the fact that Tλ(Gλ)=𝔤λT_{\lambda}(G\cdot\lambda)=\mathfrak{g}_{\lambda}^{\circ}. ∎

In the next statements XX denotes a presymplectic Hamiltonian GG-manifold with moment map Φ\Phi and null foliation \mathscr{F}. The result holds regardless of any cleanness assumptions on the GG-action on XX. We denote by x¯\bar{x} the leaf of a point xXx\in X, considered as a point in the leaf space X/X/\mathscr{F}, and by 𝔤x¯\mathfrak{g}_{\bar{x}} its stabilizer subalgebra, which consists of all ξ𝔤\xi\in\mathfrak{g} satisfying ξX(x)Tx\xi_{X}(x)\in T_{x}\mathscr{F}.

B.4 Lemma.

For all xXx\in X we have

  1. (i)

    ker(TxΦ)=Tx(Gx)ωx\ker(T_{x}\Phi)=T_{x}(G\cdot x)^{\omega_{x}};

  2. (ii)

    im(TxΦ)=𝔤x¯\operatorname{\mathrm{im}}(T_{x}\Phi)=\mathfrak{g}_{\bar{x}}^{\circ} and coker(TxΦ)=𝔤x¯\operatorname{\mathrm{coker}}(T_{x}\Phi)=\mathfrak{g}_{\bar{x}}^{*}.

Proof.
\oldtagform@

i The moment map condition dΦξ=ι(ξX)ωd\Phi^{\xi}=\iota(\xi_{X})\omega implies that vker(TxΦ)v\in\ker(T_{x}\Phi) if and only if ωx(ξX(x),v)=0\omega_{x}(\xi_{X}(x),v)=0, i.e. vTx(Gx)ωxv\in T_{x}(G\cdot x)^{\omega_{x}}.

\oldtagform@

ii Similarly, a vector ξ𝔤\xi\in\mathfrak{g} is in im(TxΦ)\operatorname{\mathrm{im}}(T_{x}\Phi)^{\circ} if and only if ωx(ξX(x),v)=0\omega_{x}(\xi_{X}(x),v)=0 for all vTxXv\in T_{x}X, i.e. ξX(x)ker(ωx)=Tx\xi_{X}(x)\in\ker(\omega_{x})=T_{x}\mathscr{F}. This is equivalent to ξX(x)Tx(Gx)Tx=Tx(Gx¯x)\xi_{X}(x)\in T_{x}(G\cdot x)\cap T_{x}\mathscr{F}=T_{x}(G_{\bar{x}}\cdot x), where we used \oldtagform@2.6.2. Thus im(TxΦ)=𝔤x¯\operatorname{\mathrm{im}}(T_{x}\Phi)^{\circ}=\mathfrak{g}_{\bar{x}}. In other words, im(TxΦ)=𝔤x¯\operatorname{\mathrm{im}}(T_{x}\Phi)=\mathfrak{g}_{\bar{x}}^{\circ} and coker(TxΦ)=𝔤/𝔤x¯=𝔤x¯\operatorname{\mathrm{coker}}(T_{x}\Phi)=\mathfrak{g}^{*}/\mathfrak{g}_{\bar{x}}^{\circ}=\mathfrak{g}_{\bar{x}}^{*}. ∎

Recall that any smooth map of manifolds f:ABf\colon A\to B has an intrinsically defined Hessian or second derivative Ta2f:ker(Taf)coker(Taf)T^{2}_{a}f\colon\ker(T_{a}f)\to\operatorname{\mathrm{coker}}(T_{a}f) at every aAa\in A.

B.5 Corollary.

For every xXx\in X the second derivative of the moment map at xx is a linear map Tx2Φ:Tx(Gx)ωx𝔤x¯T^{2}_{x}\Phi\colon T_{x}(G\cdot x)^{\omega_{x}}\to\mathfrak{g}_{\bar{x}}^{*}. In particular, if the leaf x¯\bar{x} is GG-fixed, the second derivative is a linear map Tx2Φ:TxX𝔤T^{2}_{x}\Phi\colon T_{x}X\to\mathfrak{g}^{*}.

Appendix C The local normal form

C.1. This appendix, the results of which are not used in the main body of the paper, but which develops a theme touched upon in Section 4, contains a local normal form theorem for clean presymplectic Hamiltonian Lie group actions, which is a refinement of the slice theorem, Theorem 2.10.1. It is a type of equivariant Darboux-Weinstein theorem, which extends results established in the symplectic case by Guillemin and Sternberg [9, § 41] and Marle [19]. It says that up to isomorphism an invariant neighbourhood of a point xx in a presymplectic Hamiltonian GG-manifold is entirely determined by infinitesimal data, namely the stabilizer subgroup GxG_{x}, the image of xx under the moment map, and two finite-dimensional representations of GxG_{x}, which describe the relevant information “orthogonal” to the orbit of xx. These GxG_{x}-modules, which are subquotients of the tangent space at xx and which we call the symplectic slice and the null slice, capture respectively the symplectic directions and the null directions complementary to the orbit. As shown in Section 4, the equivariant Darboux-Weinstein theorem is false without a cleanness assumption on the point xx. For a fixed point xx the theorem reduces to Corollary 2.10.4.

As in Section 2 we denote by XX a manifold with presymplectic form ω\omega and by GG a connected compact Lie group acting on MM in a Hamiltonian fashion with moment map Φ\Phi. (See 2.) We denote the null foliation of ω\omega by =X\mathscr{F}=\mathscr{F}_{X}, the null ideal sheaf by 𝔫~=𝔫~X\tilde{\mathfrak{n}}=\tilde{\mathfrak{n}}_{X} and the null subgroup sheaf by N~=N~X\tilde{N}=\tilde{N}_{X}. (See 2.2.) In CC.2.2 we introduce the slice modules; in C.3.2 we describe the local model and in C.4.3 we prove the equivariant Darboux-Weinstein theorem.

C.2.  Let EE be a finite-dimensional real vector space and σ\sigma a presymplectic form on EE. We denote by EE^{\natural} the largest symplectic quotient space of EE, i.e. E=E/ker(σ)E^{\natural}=E/\ker(\sigma), and by σ\sigma^{\natural} the symplectic form on EE^{\natural} induced by σ\sigma. If FF is a linear subspace of EE equipped with the presymplectic form σF=σ|F\sigma_{F}=\sigma|_{F}, then ker(σF)=FFσ\ker(\sigma_{F})=F\cap F^{\sigma}, where FσF^{\sigma} denotes the subspace of EE orthogonal to FF with respect to σ\sigma. Therefore the largest symplectic quotient space of FF is

F=F/(FFσ)(F+Fσ)/Fσ.F^{\natural}=F/(F\cap F^{\sigma})\cong(F+F^{\sigma})/F^{\sigma}.

We have (Fσ)σ=F+ker(σ)(F^{\sigma})^{\sigma}=F+\ker(\sigma) and so

ker(σFσ)=Fσ(F+ker(σ))=FσF+ker(σ).\ker(\sigma_{F^{\sigma}})=F^{\sigma}\cap(F+\ker(\sigma))=F^{\sigma}\cap F+\ker(\sigma).

Hence the largest symplectic quotient space of FσF^{\sigma} is

(C.2.1) (Fσ)=Fσ/(FσF+ker(σ))(F+Fσ)/(F+ker(σ)).(F^{\sigma})^{\natural}=F^{\sigma}/(F^{\sigma}\cap F+\ker(\sigma))\cong(F+F^{\sigma})/(F+\ker(\sigma)).

Let HH be a Lie group and suppose that EE is a presymplectic HH-module. Suppose also that FF is an HH-submodule of EE. Then the vector spaces EE^{\natural}, FF^{\natural} and (Fσ)(F^{\sigma})^{\natural} are symplectic HH-modules in a natural way. We omit the proof of the following elementary assertion.

C.2.2 Lemma.

Let (E1,σ1)(E_{1},\sigma_{1}) and (E2,σ2)(E_{2},\sigma_{2}) be presymplectic HH-modules.

  1. (i)

    Let f:E1E2f\colon E_{1}\to E_{2} be an HH-equivariant surjective linear map satisfying fσ2=σ1f^{*}\sigma_{2}=\sigma_{1}. Then ff descends to an isomorphism of symplectic HH-modules

    f:E1E2.f^{\natural}\colon E_{1}^{\natural}\overset{\cong}{\longrightarrow}E_{2}^{\natural}.

    Let F1F_{1} be an HH-submodule of E1E_{1} and put F2=f(F1)F_{2}=f(F_{1}). Then F2σ2=f(F1σ1)F_{2}^{\sigma_{2}}=f(F_{1}^{\sigma_{1}}). Hence ff descends to isomorphisms of symplectic HH-modules

    f:F1F2,f:(F1σ1)(F2σ2).f^{\natural}\colon F_{1}^{\natural}\overset{\cong}{\longrightarrow}F_{2}^{\natural},\qquad f^{\natural}\colon\bigl{(}F_{1}^{\sigma_{1}}\bigr{)}^{\natural}\overset{\cong}{\longrightarrow}\bigl{(}F_{2}^{\sigma_{2}}\bigr{)}^{\natural}.
  2. (ii)

    Let F1F_{1} be an HH-submodule of E1E_{1} and F2F_{2} an HH-submodule of E2E_{2}. Let E=E1E2E=E_{1}\oplus E_{2} and F=F1F2F=F_{1}\oplus F_{2}. Then EE1E2E^{\natural}\cong E_{1}^{\natural}\oplus E_{2}^{\natural}, FF1F2F^{\natural}\cong F_{1}^{\natural}\oplus F_{2}^{\natural}, and (Fσ)(F1σ1)(F2σ2)(F^{\sigma})^{\natural}\cong(F_{1}^{\sigma_{1}})^{\natural}\oplus(F_{2}^{\sigma_{2}})^{\natural}.

C.3.  Let xXx\in X. Applying the observations of C to the Lie group H=GxH=G_{x}, the GxG_{x}-module E=TxXE=T_{x}X, the presymplectic form σ=ωx\sigma=\omega_{x}, and the submodule F=Tx(Gx)F=T_{x}(G\cdot x), we arrive at a symplectic GxG_{x}-module

Sx(X)=(Tx(Gx)ωx)=Tx(Gx)ωx/(Tx(Gx)ωxTx(Gx)+Tx)),S_{x}(X)=\bigl{(}T_{x}(G\cdot x)^{\omega_{x}}\bigr{)}^{\natural}=T_{x}(G\cdot x)^{\omega_{x}}\big{/}\bigl{(}T_{x}(G\cdot x)^{\omega_{x}}\cap T_{x}(G\cdot x)+T_{x}\mathscr{F})\bigr{)},

which we call the symplectic slice of XX at xx.

C.3.1 Lemma.

The symplectic slice Sx(X)S_{x}(X) is naturally isomorphic to a submodule of the module V1=TxX/(Tx(Gx)+Tx)V_{1}=T_{x}X\big{/}\bigl{(}T_{x}(G\cdot x)+T_{x}\mathscr{F}\bigr{)} of Theorem 2.10.1. The ideal 𝔤x𝔫x\mathfrak{g}_{x}\cap\mathfrak{n}_{x} of 𝔤x\mathfrak{g}_{x} acts trivially on Sx(X)S_{x}(X).

Proof.

It follows from \oldtagform@C.2.1 that

Sx(X)(Tx(Gx)+Tx(Gx)ωx)/(Tx(Gx)+Tx),S_{x}(X)\cong\bigl{(}T_{x}(G\cdot x)+T_{x}(G\cdot x)^{\omega_{x}}\bigr{)}\big{/}\bigl{(}T_{x}(G\cdot x)+T_{x}\mathscr{F}\bigr{)},

which is a submodule of V1V_{1}. Let η𝔤x𝔫x\eta\in\mathfrak{g}_{x}\cap\mathfrak{n}_{x}. Then the function Φη\Phi^{\eta} is constant near xx, and therefore its Hessian Tx2Φη:TxXRT_{x}^{2}\Phi^{\eta}\colon T_{x}X\to{\text{\bf R}} is 0. By the equivariant Darboux theorem, Corollary 2.10.4, applied to the fixed point xx of the GxG_{x}-action, the Hessian Tx2ΦηT_{x}^{2}\Phi^{\eta} is the η\eta-component of the moment map of the linear GxG_{x}-action on TxXT_{x}X. Therefore the linearization at xx of the vector field ηX\eta_{X} is tangent to the leaves of the constant presymplectic form ωx\omega_{x} on TxXT_{x}X. It follows that η\eta acts trivially on the quotient module (TxX)=TxX/Tx(T_{x}X)^{\natural}=T_{x}X/T_{x}\mathscr{F}. Hence η\eta acts trivially on the subquotient Sx(X)S_{x}(X) of (TxX)(T_{x}X)^{\natural}. ∎

Let x¯=(x)\bar{x}=\mathscr{F}(x) be the leaf of xx, considered as a point in the leaf space X/X/\mathscr{F}, and let Gx¯G_{\bar{x}} the stabilizer of x¯\bar{x}. The null slice of XX at xx is the GxG_{x}-module

Vx(X)=(Tx(Gx)+Tx)/Tx(Gx)Tx/(Tx(Gx)Tx)=Tx/Tx(Gx¯x),V_{x}(X)=\bigl{(}T_{x}(G\cdot x)+T_{x}\mathscr{F}\bigr{)}\big{/}T_{x}(G\cdot x)\cong T_{x}\mathscr{F}\big{/}\bigl{(}T_{x}(G\cdot x)\cap T_{x}\mathscr{F}\bigr{)}=T_{x}\mathscr{F}/T_{x}(G_{\bar{x}}\cdot x),

where the last equality follows from \oldtagform@2.6.2. This is the module denoted by V0V_{0} in Theorem 2.10.1. Note that Vx(X)=0V_{x}(X)=0 if and only if the leaf (x)\mathscr{F}(x) is contained in the GG-orbit of xx. The next result describes how the symplectic slice and the null slice behave under presymplectic submersions and under symplectization.

C.3.2 Lemma.

Let xXx\in X.

  1. (i)

    Let (Y,ωY)(Y,\omega_{Y}) be a presymplectic Hamiltonian GG-manifold and let p:XYp\colon X\to Y be an equivariant submersion with the property pωY=ωp^{*}\omega_{Y}=\omega. Let y=p(x)y=p(x) and assume Gy=GxG_{y}=G_{x}. Then pp induces an isomorphism of symplectic GxG_{x}-modules

    p:Sx(X)Sx(Y)p^{\natural}\colon S_{x}(X)\overset{\cong}{\longrightarrow}S_{x}(Y)

    and a short exact sequence of GxG_{x}-modules

    0ker(Txp)Vx(X)TxpVy(Y)0.0\longrightarrow\ker(T_{x}p)\longrightarrow V_{x}(X)\overset{T_{x}p}{\longrightarrow}V_{y}(Y)\longrightarrow 0.
  2. (ii)

    Let M=TM=T^{*}\mathscr{F} be the symplectization of XX. Then Vx(M)=0V_{x}(M)=0 and there is an isomorphism of symplectic GxG_{x}-modules

    Sx(M)Sx(X)Vx(X)Vx(X).S_{x}(M)\cong S_{x}(X)\oplus V_{x}(X)\oplus V_{x}(X)^{*}.

    In particular, Sx(M)Sx(X)S_{x}(M)\cong S_{x}(X) if (x)Gx\mathscr{F}(x)\subseteq G\cdot x.

Proof.
\oldtagform@

i Let H=Gx=GyH=G_{x}=G_{y}, E1=TxXE_{1}=T_{x}X, and E2=TyYE_{2}=T_{y}Y. Then E1E_{1} and E2E_{2} are presymplectic HH-modules with presymplectic forms σ1=ωx\sigma_{1}=\omega_{x}, resp. σ2=ωY,y\sigma_{2}=\omega_{Y,y}. The tangent spaces to the orbits F1=Tx(Gx)F_{1}=T_{x}(G\cdot x) and F2=Ty(Gy)F_{2}=T_{y}(G\cdot y) are submodules of E1E_{1}, resp. E2E_{2}. The tangent map p=Txp:E1E2p_{*}=T_{x}p\colon E_{1}\to E_{2} satisfies pσ2=σ1p^{*}\sigma_{2}=\sigma_{1} and p(F1)=F2p_{*}(F_{1})=F_{2}. By definition the symplectic slices are Sx(X)=(Fσ1)S_{x}(X)=(F^{\sigma_{1}})^{\natural} and Sy(Y)=(Fσ2)S_{y}(Y)=(F^{\sigma_{2}})^{\natural}. The statement that the two are isomorphic now follows from the third isomorphism in Lemma C.2.2\oldtagform@i. The restriction of pp_{*} to the subspace F1+TxF_{1}+T_{x}\mathscr{F} has image F2+TyYF_{2}+T_{y}\mathscr{F}_{Y}. Hence pp_{*} descends to a surjection from Vx(X)=(F1+Tx)/F1V_{x}(X)=(F_{1}+T_{x}\mathscr{F})/F_{1} to Vy(Y)=(F2+TyY)/F2V_{y}(Y)=(F_{2}+T_{y}\mathscr{F}_{Y})/F_{2} with kernel ker(p)\ker(p_{*}).

\oldtagform@

ii Since MM is symplectic, we have Vx(M)=0V_{x}(M)=0. Let E1=TxXE_{1}=T_{x}X and E2=E1E_{2}=E_{1}^{\natural}. On E1E_{1} we have the presymplectic form σ1=ωx\sigma_{1}=\omega_{x} and on E2E_{2} we have the symplectic form σ2=σ1\sigma_{2}=\sigma_{1}^{\natural}. Let π:E1E2\pi\colon E_{1}\to E_{2} be the quotient map and E0=ker(π)E_{0}=\ker(\pi) its kernel. Then E0=ker(σ1)E_{0}=\ker(\sigma_{1}) and πσ2=σ1\pi^{*}\sigma_{2}=\sigma_{1}. Let F1=Tx(Gx)𝔤/𝔤xF_{1}=T_{x}(G\cdot x)\cong\mathfrak{g}/\mathfrak{g}_{x}, let F2=π(F1)E2F_{2}=\pi(F_{1})\subseteq E_{2} be the image of F1F_{1}, and let F0=F1E0F_{0}=F_{1}\cap E_{0} be the kernel of π:F1F2\pi\colon F_{1}\to F_{2}. It follows from \oldtagform@2.6.2 that

F0=Tx(Gx)Tx=Tx(Gx¯x)𝔤x¯/𝔤x.F_{0}=T_{x}(G\cdot x)\cap T_{x}\mathscr{F}=T_{x}(G_{\bar{x}}\cdot x)\cong\mathfrak{g}_{\bar{x}}/\mathfrak{g}_{x}.

Therefore F2=F1/F0𝔤/𝔤x¯F_{2}=F_{1}/F_{0}\cong\mathfrak{g}/\mathfrak{g}_{\bar{x}}. From the third isomorphism in Lemma C.2.2\oldtagform@i we obtain

(C.3.3) Sx(X)=(F1σ1)(F2σ2)=F2σ2/(F2σ2F2).S_{x}(X)=\bigl{(}F_{1}^{\sigma_{1}}\bigr{)}^{\natural}\cong\bigl{(}F_{2}^{\sigma_{2}}\bigr{)}^{\natural}=F_{2}^{\sigma_{2}}\big{/}\bigl{(}F_{2}^{\sigma_{2}}\cap F_{2}\bigr{)}.

We can express the relationships among the various presymplectic GxG_{x}-modules as a commutative diagram with exact rows:

(C.3.4) 0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔤x¯/𝔤x\textstyle{\mathfrak{g}_{\bar{x}}/\mathfrak{g}_{x}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}𝔤/𝔤x\textstyle{\mathfrak{g}/\mathfrak{g}_{x}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}𝔤/𝔤x¯\textstyle{\mathfrak{g}/\mathfrak{g}_{\bar{x}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}0\textstyle{0}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}F0\textstyle{F_{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}F1\textstyle{F_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}F2\textstyle{F_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}E0\textstyle{E_{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}E1\textstyle{E_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}E2\textstyle{E_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

Recall from 2.10.4 that the symplectic form Ω\Omega on M=TM=T^{*}\mathscr{F} depends on a choice of a GG-invariant Riemannian metric on XX. Given such a choice, we obtain compatible splittings of the rows of the diagram \oldtagform@C.3.4, namely by identifying the middle terms with orthogonal direct sums E1E2E0E_{1}\cong E_{2}\oplus E_{0} and F1F2F0F_{1}\cong F_{2}\oplus F_{0}. Let E=TxME=T_{x}M and σ=Ωx\sigma=\Omega_{x}. Then EE is a direct sum, E=E2E0E0E=E_{2}\oplus E_{0}\oplus E_{0}^{*}, and the symplectic form on EE is σ=σ2σ0\sigma=\sigma_{2}\oplus\sigma_{0}, where σ0\sigma_{0} is the standard symplectic form on E0E0E_{0}\oplus E_{0}^{*}. Applying Lemma C.2.2\oldtagform@ii to the subspace F1=F2F0F_{1}=F_{2}\oplus F_{0} of EE we obtain

(C.3.5) Sx(M)=F2σ2/(F2σ2F2)F0σ0/(F0σ0F0).S_{x}(M)=F_{2}^{\sigma_{2}}\big{/}\bigl{(}F_{2}^{\sigma_{2}}\cap F_{2}\bigr{)}\oplus F_{0}^{\sigma_{0}}\big{/}\bigl{(}F_{0}^{\sigma_{0}}\cap F_{0}\bigr{)}.

Now F0σ0=E0F0F_{0}^{\sigma_{0}}=E_{0}\oplus F_{0}^{\circ}, where F0F_{0}^{\circ} is the annihilator of F0F_{0} in E0E_{0}^{*}, so F0σ0F0=F0F_{0}^{\sigma_{0}}\cap F_{0}=F_{0} and

F0σ0/(F0σ0F0)=E0/F0F0=VV,F_{0}^{\sigma_{0}}\big{/}\bigl{(}F_{0}^{\sigma_{0}}\cap F_{0}\bigr{)}=E_{0}/F_{0}\oplus F_{0}^{\circ}=V\oplus V^{*},

where V=Vx(X)V=V_{x}(X). Substituting this and \oldtagform@C.3.3 into \oldtagform@C.3.5 gives Sx(M)Sx(X)VVS_{x}(M)\cong S_{x}(X)\oplus V\oplus V^{*}. If (x)Gx\mathscr{F}(x)\subseteq G\cdot x, then V=0V=0, so Sx(M)Sx(X)S_{x}(M)\cong S_{x}(X). ∎

C.4.  We now describe the local model for clean presymplectic Hamiltonian actions. First a quick review of the symplectic case. (See [9, § 41] or [19] for a complete exposition.) The symplectic local model has a list of four ingredients (λ,H,θ,S)(\lambda,H,\theta,S), consisting of

  1. (1)

    a covector λ𝔤\lambda\in\mathfrak{g}^{*},

  2. (2)

    a closed subgroup HH of the coadjoint stabilizer GλG_{\lambda} of λ\lambda,

  3. (3)

    an HH-equivariant splitting θ:𝔤λ/𝔥𝔤λ\theta\colon\mathfrak{g}_{\lambda}/\mathfrak{h}\to\mathfrak{g}_{\lambda} of the quotient map 𝔤λ𝔤λ/𝔥\mathfrak{g}_{\lambda}\to\mathfrak{g}_{\lambda}/\mathfrak{h},

  4. (4)

    a finite-dimensional symplectic HH-module SS.

We denote the HH-module 𝔤λ/𝔥\mathfrak{g}_{\lambda}/\mathfrak{h} by 𝔪\mathfrak{m}. We use the splitting θ\theta to identify the HH-module 𝔪\mathfrak{m} with a direct summand of 𝔤λ\mathfrak{g}_{\lambda} and the HH-module 𝔥\mathfrak{h}^{*} with a direct summand of 𝔤λ\mathfrak{g}_{\lambda}^{*}. The homogeneous bundle

(C.4.1) 𝔐=𝔐(λ,H,θ,S)=G×H(𝔪×S)\mathfrak{M}=\mathfrak{M}(\lambda,H,\theta,S)=G\times^{H}(\mathfrak{m}^{*}\times S)

carries a closed 22-form ω𝔐\omega_{\mathfrak{M}} which is nondegenerate in a neighbourhood of the zero section. The left multiplication action of GG on 𝔐\mathfrak{M} is Hamiltonian with moment map Φ𝔐\Phi_{\mathfrak{M}} given by

Φ𝔐([g,a,s])=Adg(λ+a+ΦS(s)),\Phi_{\mathfrak{M}}([g,a,s])=\operatorname{\mathrm{Ad}}_{g}^{*}\bigl{(}\lambda+a+\Phi_{S}(s)\bigr{)},

where ΦSη(s)=12ωS(ηS(s),s)\Phi_{S}^{\eta}(s)=\frac{1}{2}\omega_{S}(\eta_{S}(s),s) for η𝔥\eta\in\mathfrak{h}. The formula for Φ𝔐\Phi_{\mathfrak{M}} is to be interpreted as follows. The inclusion 𝔤λ𝔤\mathfrak{g}_{\lambda}\to\mathfrak{g} has a unique GλG_{\lambda}-equivariant left inverse, which we use to identify 𝔤λ\mathfrak{g}_{\lambda}^{*} with a direct summand of 𝔤\mathfrak{g}^{*}. This allows us to identify 𝔤\mathfrak{g}^{*} with a product

(C.4.2) 𝔤𝔤λ×𝔤λ𝔤λ×𝔪×𝔥,\mathfrak{g}^{*}\cong\mathfrak{g}_{\lambda}^{\circ}\times\mathfrak{g}_{\lambda}^{*}\cong\mathfrak{g}_{\lambda}^{\circ}\times\mathfrak{m}^{*}\times\mathfrak{h}^{*},

and to regard a𝔪a\in\mathfrak{m}^{*} and ΦS(s)𝔥\Phi_{S}(s)\in\mathfrak{h}^{*} as elements of 𝔤\mathfrak{g}^{*}. Then for gGg\in G we let Adg\operatorname{\mathrm{Ad}}_{g}^{*} act on the element λ+a+ΦS(s)𝔤\lambda+a+\Phi_{S}(s)\in\mathfrak{g}^{*}.

The presymplectic local model requires six ingredients (λ,H,θ,S,V,𝔞)(\lambda,H,\theta,S,V,\mathfrak{a}), where λ\lambda, HH, θ\theta, SS are as in \oldtagform@1\oldtagform@4 and in addition we have

  1. (5)

    an HH-module VV,

  2. (6)

    an ideal 𝔞\mathfrak{a} of 𝔤\mathfrak{g} with the properties that 𝔨𝔞𝔤λ\mathfrak{k}\subseteq\mathfrak{a}\subseteq\mathfrak{g}_{\lambda} and that the ideal 𝔞𝔥\mathfrak{a}\cap\mathfrak{h} of 𝔥\mathfrak{h} acts trivially on SS.

Here 𝔨\mathfrak{k} denotes the kernel of the infinitesimal GG-action 𝔤Γ(T𝔐)\mathfrak{g}\to\Gamma(T\mathfrak{M}) on the symplectic Hamiltonian GG-manifold 𝔐\mathfrak{M} defined in \oldtagform@C.4.1. The ideal 𝔨\mathfrak{k} of 𝔤\mathfrak{g} is determined by the data λ\lambda, HH and SS. The quotient 𝔭=(𝔞+𝔥)/𝔥𝔞/(𝔞𝔥)\mathfrak{p}=(\mathfrak{a}+\mathfrak{h})/\mathfrak{h}\cong\mathfrak{a}/(\mathfrak{a}\cap\mathfrak{h}) is an HH-submodule of 𝔪=𝔤λ/𝔥\mathfrak{m}=\mathfrak{g}_{\lambda}/\mathfrak{h}. We require the splitting θ:𝔪𝔤λ\theta\colon\mathfrak{m}\to\mathfrak{g}_{\lambda} to be compatible with the ideal 𝔞\mathfrak{a} in the sense that θ(𝔭)\theta(\mathfrak{p}) is contained in 𝔞\mathfrak{a}. Because the HH-action on 𝔤λ\mathfrak{g}_{\lambda} preserves the ideal 𝔞\mathfrak{a}, such a splitting always exists. We form the quotient module 𝔮=𝔪/𝔭\mathfrak{q}=\mathfrak{m}/\mathfrak{p} and the homogeneous vector bundle

𝔛=𝔛(λ,H,θ,S,V,𝔞)=G×H(𝔮×S×V).\mathfrak{X}=\mathfrak{X}(\lambda,H,\theta,S,V,\mathfrak{a})=G\times^{H}(\mathfrak{q}^{*}\times S\times V).

We define an equivariant vector bundle map f:𝔛𝔐f\colon\mathfrak{X}\to\mathfrak{M} by

f([g,b,s,v])=[g,i(b),s],f([g,b,s,v])=[g,i(b),s],

where i:𝔮𝔪i\colon\mathfrak{q}^{*}\to\mathfrak{m}^{*} is the natural inclusion. Let 𝔜\mathfrak{Y} be the direct summand

𝔜=𝔛(λ,H,θ,S,0,𝔞)=G×H(𝔮×S)\mathfrak{Y}=\mathfrak{X}(\lambda,H,\theta,S,0,\mathfrak{a})=G\times^{H}(\mathfrak{q}^{*}\times S)

of the vector bundle 𝔛\mathfrak{X}. Then f=jpf=j\circ p, where

𝔛𝑝𝔜𝑗𝔐\mathfrak{X}\overset{p}{\longrightarrow}\mathfrak{Y}\overset{j}{\longrightarrow}\mathfrak{M}

are defined by p([g,b,s,v])=[g,b,s]p([g,b,s,v])=[g,b,s] and j([g,b,s])=[g,i(b),s]j([g,b,s])=[g,i(b),s]. We write ω𝔛=fω𝔐\omega_{\mathfrak{X}}=f^{*}\omega_{\mathfrak{M}}, Φ𝔛=fΦ𝔐\Phi_{\mathfrak{X}}=f^{*}\Phi_{\mathfrak{M}}, ω𝔜=jω𝔐\omega_{\mathfrak{Y}}=j^{*}\omega_{\mathfrak{M}}, and Φ𝔜=jΦ𝔐\Phi_{\mathfrak{Y}}=j^{*}\Phi_{\mathfrak{M}}. Note that jj is an equivariant embedding of the vector bundle 𝔜\mathfrak{Y}. We identify 𝔜\mathfrak{Y} with the subbundle j(𝔜)j(\mathfrak{Y}) of 𝔐\mathfrak{M}. Let x0𝔛x_{0}\in\mathfrak{X} denote the basepoint [1,0,0,0][1,0,0,0] and let AA be the connected immersed normal subgroup of GG generated by the ideal 𝔞\mathfrak{a}. Here are the relevant properties of the model 𝔛\mathfrak{X}.

C.4.3 Lemma.
  1. (i)

    Φ𝔐:𝔐𝔤\Phi_{\mathfrak{M}}\colon\mathfrak{M}\to\mathfrak{g}^{*} intersects the affine subspace λ+𝔞\lambda+\mathfrak{a}^{\circ} cleanly and 𝔜=Φ𝔐1(λ+𝔞)\mathfrak{Y}=\Phi_{\mathfrak{M}}^{-1}(\lambda+\mathfrak{a}^{\circ}). Hence near the zero section 𝔜\mathfrak{Y} is a coisotropic submanifold of 𝔐\mathfrak{M} and 𝔐\mathfrak{M} is the symplectization of 𝔜\mathfrak{Y}. The leaves of the null foliation of ω𝔜\omega_{\mathfrak{Y}} are the orbits of the AA-action on 𝔜\mathfrak{Y}.

  2. (ii)

    Near the zero section 𝔛\mathfrak{X} is a presymplectic Hamiltonian GG-manifold with presymplectic form ω𝔛\omega_{\mathfrak{X}} and moment map Φ𝔛\Phi_{\mathfrak{X}}. The GG-action on 𝔛\mathfrak{X} is clean at x0x_{0}. Near the zero section the null ideal sheaf 𝔫~𝔛\tilde{\mathfrak{n}}_{\mathfrak{X}} is constant with stalk 𝔞\mathfrak{a}.

  3. (iii)

    The stabilizer of the basepoint is Gx0=HG_{x_{0}}=H, its moment map value is Φ𝔛(x0)=λ\Phi_{\mathfrak{X}}(x_{0})=\lambda, the symplectic slice is Sx0(𝔛)SS_{x_{0}}(\mathfrak{X})\cong S, and the null slice is Vx0(𝔛)VV_{x_{0}}(\mathfrak{X})\cong V.

Proof.
\oldtagform@

i Let [g,a,s]𝔐[g,a,s]\in\mathfrak{M}. Writing Φ𝔐([g,a,s])=Adg(λ+ϕ(a,s))\Phi_{\mathfrak{M}}([g,a,s])=\operatorname{\mathrm{Ad}}_{g}^{*}(\lambda+\phi(a,s)), where ϕ\phi is the HH-equivariant map 𝔪×S𝔤\mathfrak{m}^{*}\times S\to\mathfrak{g}^{*} defined by (a,s)a+ΦS(s)(a,s)\mapsto a+\Phi_{S}(s), we have

Φ𝔐([g,a,s])λ+𝔞ϕ(a,s)𝔞.\Phi_{\mathfrak{M}}([g,a,s])\in\lambda+\mathfrak{a}^{\circ}\iff\phi(a,s)\in\mathfrak{a}^{\circ}.

Under the identification \oldtagform@C.4.2 we have 𝔞𝔤λ×𝔮×(𝔞𝔥)\mathfrak{a}^{\circ}\cong\mathfrak{g}_{\lambda}^{\circ}\times\mathfrak{q}^{*}\times(\mathfrak{a}\cap\mathfrak{h})^{\circ}, where 𝔤λ\mathfrak{g}_{\lambda}^{\circ} denotes the annihilator of 𝔤λ\mathfrak{g}_{\lambda} in 𝔤\mathfrak{g}^{*} and (𝔞𝔥)(\mathfrak{a}\cap\mathfrak{h})^{\circ} the annihilator of 𝔞𝔥\mathfrak{a}\cap\mathfrak{h} in 𝔥\mathfrak{h}^{*}. By assumption 𝔞𝔥\mathfrak{a}\cap\mathfrak{h} acts trivially on SS, so the HH-moment map ΦS\Phi_{S} maps SS into (𝔞𝔥)(\mathfrak{a}\cap\mathfrak{h})^{\circ}. Therefore ϕ(a,s)𝔞\phi(a,s)\in\mathfrak{a}^{\circ} is equivalent to a𝔮a\in\mathfrak{q}^{*}, i.e. ϕ1(𝔞)=𝔮×S\phi^{-1}(\mathfrak{a}^{\circ})=\mathfrak{q}^{*}\times S. Therefore 𝔜=Gϕ1(𝔞)=Φ𝔐1(λ+𝔞)\mathfrak{Y}=G\cdot\phi^{-1}(\mathfrak{a}^{\circ})=\Phi_{\mathfrak{M}}^{-1}(\lambda+\mathfrak{a}^{\circ}). Moreover, ϕ\phi intersects the linear subspace 𝔞\mathfrak{a}^{\circ} cleanly, which implies that Φ𝔐\Phi_{\mathfrak{M}} intersects λ+𝔞\lambda+\mathfrak{a}^{\circ} cleanly. The remaining assertions now follow from Proposition 2.11.2.

\oldtagform@

ii The first assertion follows from \oldtagform@i and the fact that p:𝔛𝔜p\colon\mathfrak{X}\to\mathfrak{Y} is an equivariant surjective submersion. By Proposition 2.11.2 the action on 𝔜\mathfrak{Y} is leafwise transitive and the null ideal is 𝔫(𝔜)=𝔞\mathfrak{n}(\mathfrak{Y})=\mathfrak{a}. Therefore the action on 𝔛\mathfrak{X} is clean at x0x_{0} and, by Corollary 2.10.3, the sheaf 𝔫𝔛\mathfrak{n}_{\mathfrak{X}} is constant near x0x_{0} with stalk 𝔞\mathfrak{a}.

\oldtagform@

iii The space 𝔛\mathfrak{X} is a homogeneous bundle over G/HG/H and the basepoint x0x_{0} is the identity coset in G/HG/H, so its stabilizer is Gx0=HG_{x_{0}}=H. We have Φ𝔛(x0)=Φ𝔐([1,0,0])=λ\Phi_{\mathfrak{X}}(x_{0})=\Phi_{\mathfrak{M}}([1,0,0])=\lambda. The equivariant surjection pp induces an isomorphism Sx0(𝔛)Sx0(𝔜)S_{x_{0}}(\mathfrak{X})\cong S_{x_{0}}(\mathfrak{Y}) by Lemma C.3.2\oldtagform@i. The fact that the action on 𝔜\mathfrak{Y} is leafwise transitive implies Vx0(𝔜)=0V_{x_{0}}(\mathfrak{Y})=0, and hence Sx0(𝔜)Sx0(𝔐)SS_{x_{0}}(\mathfrak{Y})\cong S_{x_{0}}(\mathfrak{M})\cong S by Lemma C.3.2\oldtagform@ii. This shows Sx0(𝔛)SS_{x_{0}}(\mathfrak{X})\cong S. Moreover, Vx0(𝔛)ker(Tx0p)=VV_{x_{0}}(\mathfrak{X})\cong\ker(T_{x_{0}}p)=V by Lemma C.3.2\oldtagform@i. ∎

C.5.  The local normal form theorem is as follows.

C.5.1 Theorem.

Let xXx\in X and assume that the GG-action on XX is clean at xx. Let

λ=Φ(x),H=Gx,S=Sx(X),V=Vx(X),𝔞=𝔫x.\lambda=\Phi(x),\quad H=G_{x},\quad S=S_{x}(X),\quad V=V_{x}(X),\quad\mathfrak{a}=\mathfrak{n}_{x}.

Choose an HH-equivariant splitting θ:𝔤λ/𝔥𝔤λ\theta\colon\mathfrak{g}_{\lambda}/\mathfrak{h}\to\mathfrak{g}_{\lambda} of the quotient map 𝔤λ𝔤λ/𝔥\mathfrak{g}_{\lambda}\to\mathfrak{g}_{\lambda}/\mathfrak{h} which is compatible with 𝔞\mathfrak{a}. Then a GG-invariant neighbourhood of xx in XX is isomorphic as a presymplectic Hamiltonian GG-manifold to a GG-invariant neighbourhood of x0x_{0} in the local model 𝔛=𝔛(λ,H,θ,S,V,𝔞)\mathfrak{X}=\mathfrak{X}(\lambda,H,\theta,S,V,\mathfrak{a}).

Proof.

First we verify that the list (λ,H,θ,S,V,𝔞)(\lambda,H,\theta,S,V,\mathfrak{a}) satisfies the conditions imposed in C.3.2\oldtagform@1\oldtagform@6. That the subgroup HH is contained in GλG_{\lambda} follows from the equivariance of the moment map Φ\Phi. That the null ideal 𝔞\mathfrak{a} contains 𝔨\mathfrak{k} follows from \oldtagform@2.6.3. That 𝔞\mathfrak{a} is contained in 𝔤λ\mathfrak{g}_{\lambda} follows from Lemma B.3. That 𝔞𝔥\mathfrak{a}\cap\mathfrak{h} acts trivially on the symplectic slice module SS is Lemma C.3.1. Now choose a leafwise transitive transversal YY at xx as in Theorem 2.10.1 and let MM be the symplectization of YY. Let us denote by γ(X)\gamma(X) and γ(M)\gamma(M) the germs of XX and MM at the orbit GxG\cdot x. Similarly, let us denote by γ(𝔛)\gamma(\mathfrak{X}) and γ(𝔐)\gamma(\mathfrak{M}) the germs of 𝔛\mathfrak{X} and 𝔐=𝔐(λ,H,S)\mathfrak{M}=\mathfrak{M}(\lambda,H,S) at the orbit Gx0G\cdot x_{0}. Then

Sx(M)Sx(Y)Sx(X)=SSx0(𝔐),S_{x}(M)\cong S_{x}(Y)\cong S_{x}(X)=S\cong S_{x_{0}}(\mathfrak{M}),

where the first two isomorphisms follow from Lemma C.3.2 and the last from Lemma C.4.3. It now follows from the symplectic local normal form theorem (see [9, § 41] or [19]) that γ(M)\gamma(M) and γ(𝔐)\gamma(\mathfrak{M}) are isomorphic as germs of symplectic Hamiltonian GG-manifolds. Isomorphisms intertwine moment maps, so from Proposition 2.11.1 we get that γ(Y)=γ(Ψ1(λ+𝔞))\gamma(Y)=\gamma\bigl{(}\Psi^{-1}(\lambda+\mathfrak{a}^{\circ})\bigr{)} is isomorphic to

γ(Φ𝔐1(λ+𝔞))=γ(𝔜).\gamma\bigl{(}\Phi_{\mathfrak{M}}^{-1}(\lambda+\mathfrak{a}^{\circ})\bigr{)}=\gamma(\mathfrak{Y}).

Theorem 2.10.1 states that γ(X)\gamma(X) is isomorphic to the equivariant bundle over γ(Y)\gamma(Y) with fibre VV, that is to say the bundle γ(G×H(𝔮×S×V))=γ(𝔛)\gamma\bigl{(}G\times^{H}(\mathfrak{q}^{*}\times S\times V)\bigr{)}=\gamma(\mathfrak{X}). ∎

References

  • [1] M. Atiyah, Convexity and commuting Hamiltonians, Bull. London Math. Soc. 14 (1982), no. 1, 1–15.
  • [2] J. Block and E. Getzler, Quantization of foliations, Proceedings of the XXth International Conference on Differential Geometric Methods in Theoretical Physics (New York, 1991) (S. Catto and A. Rocha, eds.), World Scientific Publishing Co., River Edge, NJ, 1992, pp. 471–487.
  • [3] N. Bourbaki, Groupes et algèbres de Lie, Éléments de mathématique, Springer-Verlag, 2007, réimpression inchangée des éditions de 1968, 1982.
  • [4] C. Boyer and K. Galicki, Sasakian geometry, Oxford Mathematical Monographs, Oxford University Press, Oxford, 2008.
  • [5] H. Bursztyn, A brief introduction to Dirac manifolds, Geometric and topological methods for quantum field theory (Villa de Leyva, 2009) (A. Cardona et al., ed.), Cambridge Univ. Press, Cambridge, 2013, pp. 4–38.
  • [6] R. Chiang and Y. Karshon, Convexity package for momentum maps on contact manifolds, Algebr. Geom. Topol. 10 (2010), no. 2, 925–977.
  • [7] M. Gotay, On coisotropic imbeddings of presymplectic manifolds, Proc. Amer. Math. Soc. 84 (1982), no. 1, 111–114.
  • [8] V. Guillemin and S. Sternberg, Convexity properties of the moment mapping, Invent. Math. 67 (1982), no. 3, 491–513.
  • [9] by same author, Symplectic techniques in physics, second ed., Cambridge University Press, Cambridge, 1990.
  • [10] Z. He, Odd-dimensional symplectic manifolds, Ph.D. thesis, Massachusetts Institute of Technology, 2010, http://hdl.handle.net/1721.1/60189.
  • [11] A. Henriques and D. Metzler, Presentations of noneffective orbifolds, Trans. Amer. Math. Soc. 356 (2004), no. 6, 2481–2499.
  • [12] J. Hilgert, K.-H. Neeb, and W. Plank, Symplectic convexity theorems and coadjoint orbits, Compositio Math. 94 (1994), no. 2, 129–180.
  • [13] B. Hoffman, Y. Lin, and R. Sjamaar, Stacky hamiltonian actions and their moment polytopes, in preparation, Cornell University and Georgia Southern University.
  • [14] H. Ishida, Torus invariant transverse Kähler foliations, eprint, 2015, arXiv:1505.06035.
  • [15] F. Kirwan, Convexity properties of the moment mapping, III, Invent. Math. 77 (1984), no. 3, 547–552.
  • [16] E. Lerman, A convexity theorem for torus actions on contact manifolds, Illinois J. Math. 46 (2002), no. 1, 171–184.
  • [17] E. Lerman and A. Malkin, Hamiltonian group actions on symplectic Deligne-Mumford stacks and toric orbifolds, Adv. Math. 229 (2012), no. 2, 984–1000.
  • [18] E. Lerman, E. Meinrenken, S. Tolman, and C. Woodward, Nonabelian convexity by symplectic cuts, Topology 37 (1998), no. 2, 245–259.
  • [19] C.-M. Marle, Modèle d’action hamiltonienne d’un groupe de Lie sur une variété symplectique, Rend. Sem. Mat. Univ. Politec. Torino 43 (1985), no. 2, 227–251.
  • [20] I. Moerdijk and J. Mrčun, Introduction to foliations and Lie groupoids, Cambridge Studies in Advanced Mathematics, vol. 91, Cambridge University Press, Cambridge, 2003.
  • [21] E. Prato, Simple non-rational convex polytopes via symplectic geometry, Topology 40 (2001), no. 5, 961–975.
  • [22] T. Ratiu and N. T. Zung, Presymplectic convexity and (ir)rational polytopes, eprint, Shanghai Jiao Tong University, 2017, arXiv:1705.11110.
  • [23] P. Rukimbira, On KK-contact manifolds with minimal number of closed characteristics, Proc. Amer. Math. Soc. 127 (1999), no. 11, 3345–3351.
  • [24] R. Sjamaar, Convexity properties of the moment mapping re-examined, Adv. Math. 138 (1998), no. 1, 46–91.
  • [25] J.-M. Souriau, Structure of dynamical systems, Progress in Mathematics, vol. 149, Birkhäuser, Boston, 1997, translated from the French by C. H. Cushman-de Vries, translation edited and with a preface by R. H. Cushman and G. M. Tuynman.