[go: up one dir, main page]

Academia.eduAcademia.edu
Quantization of Hamiltonian coactions via twist Pierre Bieliavsky∗ arXiv:1804.06160v2 [math.QA] 21 Feb 2019 Faculté des Sciences Ecole de Mathématique (MATH) Institut de Recherche en Mathématique et Physique (IRMP) Chemin du Cyclotron 2 bte L7.01.02 1348 Louvain-la-Neuve Belgium Chiara Esposito†, Institut für Mathematik Universität Würzburg Campus Hubland Nord Emil-Fischer-Straße 31 97074 Würzburg Germany Ryszard Nest‡, Department of Mathematical Sciences Universitetsparken 5 DK-2100 Copenhagen Ø Denmark February 22, 2019 Abstract In this paper we introduce a notion of quantum Hamiltonian (co)action of Hopf algebras endowed with Drinfel’d twist structure (resp., 2-cocycles). First, we define a classical Hamiltonian action in the setting of Poisson Lie groups compatible with the 2-cocycle structure and we discuss a concrete example. This allows us to construct, out of the classical momentum map, a quantum momentum map in the setting of Hopf coactions and to quantize it by using Drinfel’d approach. ∗ pierre.bieliavsky@uclouvain.be chiara.esposito@mathematik.uni-wuerzburg.de ‡ rnest@math.ku.dk † 1 Contents 1 Preliminaries 3 2 Hamiltonian actions 2.1 Dressing generators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2 Hamiltonian actions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5 5 6 3 Hamiltonian Hopf algebra (co)actions 8 3.1 Quantum Hamiltonian coactions via 2-cocycles . . . . . . . . . . . . . . . . . . . . . . 10 A Dressing generators on ax + b 11 References 15 Introduction Deformation quantization has been introduced by Bayen, Flato, Fronsdal, Lichnerowicz and Sternheimer in [3] and since then many developments occurred. A (formal) star product on a Poisson manifold M is defined as a formal associative deformation of the algebra of smooth functions C ∞ (M ) on M . Existence and classification of star products on Poisson manifolds has been proved via formality theory in [17]. In the same spirit, Drinfel’d introduced the notion of quantum groups as deformations of Hopf algebras, whose semiclassical limit are the so-called Poisson Lie groups which are Lie groups with multiplicative Poisson structures (see e.g. the textbooks [8, 23] for a detailed discussion). In this paper we focus on particular classes of star products which are induced by a (formal) Drinfel’d twist by means of universal deformation formulas (UDF) as discussed e.g. in [9,10]. Roughly speaking, a Drinfel’d twist of an enveloping algebra U(g) is an element F ∈ U(g) ⊗ U(g) compatible with the Hopf algebra structure on U(g). Given a Hopf algebra action of U(g) on an associative algebra one can deform the U(g)-module algebra and the deformed product turns out to be a star product. It is important to stress that the U(g)-module algebra is automatically endowed with a Poisson bracket defined as the semiclassical limit of such star product. In recent works the UDF has been further studied, e.g. [5,7,11,13,15]. Also, a twist defines a 2-cocycle on the Hopf algebra C ∞ (G) and it can be seen that the star products induced via UDF coincide with star products induced by the 2-cocycle on C ∞ (G)-comodule algebras. Finally, a non-formal version of Drinfel’d twist and its corresponding UDF has been discussed in [6]. Given a Lie algebra action ϕ : g → Γ∞ (T M ) on a smooth manifold M , we can always obtain a Hopf algebra action U(g) × C ∞ (M ) → C ∞ (M ). Thus, Drinfel’d approach can be interpreted by saying that symmetries encoded by Lie algebra actions induce quantization. Also, this approach provides a notion of quantized action. In this paper we prove that this approach is compatible with Hamiltonian actions. In other words, given a classical Hamiltonian action our goal is to quantize it by using Drinfel’d approach and get a notion of quantum momentum map. The problem of quantizing the momentum map has been the main topic of many works, e.g. [14] and [22]. In general, the interest for the quantization of the momentum map is motivated by the fact that conserved quantities described via the momentum map lead to phase space reduction which constructs from the high-dimensional original phase space one of a smaller dimension. Thus, it is highly desirable to find an analogue in the quantum setting. A study of the compatibility of the notion of quantum action provided by Drinfeld and the notion of Hamiltonian action was so far absent. In this paper we prove that the two notions are actually compatible and we construct a quantum momentum map via twist. The content of this work is as follows. 2 In Section 1 we discuss the well-known notions of Drinfel’d twist and its corresponding 2-cocycle and the construction of the universal deformation formula. Twist and 2-cocycle induce a quantum group structure which is briefly recalled. Section 2 contains a definition of Hamiltonian actions in the setting of Poisson Lie groups which generalizes the one contained in [19, 21]. More precisely, we need to introduce a notion of classical Hamiltonian action which is compatible with twist, which is necessary in order to quantize Hamiltonian actions by using Drinfel’d approach. It is known that the semiclassical limit of a twist gives rise to an element r ∈ g∧g, called r-matrix, satisfying the condition Jr, rK = 0 (for a detailed treatment of the relation between r-matrices and twist see [13]). It can be proved that r-matrices always induces a Lie bialgebra structure on g. Thus, the corresponding Lie group G automatically becomes a Poisson Lie group, since the Poisson tensor obtained by integrating the Lie bialgebra structure on g is multiplicative. The concept of momentum map for Poisson Lie groups acting on Poisson manifolds has been first introduced by Lu in [19, 21], in the case in which the Poisson structures of G, its dual G∗ and M are fixed. In contrast to the ordinary momentum map it takes values in G∗ and the equivariance is defined in relation to the so-called dressing action of G on G∗ . Here we introduce a slight generalization and then focus on the case in which, in the same spirit as Drinfel’d, the Poisson structure on G∗ is induced by r via the dressing action and on M via the action ϕ. In Section 3 we construct a momentum map in the setting of Hopf algebra actions and coactions and study its quantization. More precisely, given a classical Hamiltonian action ϕ : g → Γ∞ (T M ) with momentum map J : M → G∗ we construct a corresponding Hopf algebra action and we prove that J ∗ defines a momentum map for this action. This allows us to define the notion of Hamiltonian Hopf algebra action. Motivated by the significance of coactions in the theory of quantum groups in the C ∗ -algebraic framework, we give a dual version of the above result and prove that given ϕ the corresponding Hopf algebra coaction δΦ : C ∞ (M ) → C ∞ (M ) ⊗ C ∞ (G) is also Hamiltonian. Finally, using the UDF we obtain the quantized algebras C~∞ (M ) and we prove that the quantum group coaction δΦ : C~∞ (M ) → C~∞ (M ) ⊗ C~∞ (G) is again Hamiltonian. 1 Preliminaries Let g be a (finite-dimensional) Lie algebra and consider the algebra U(g)[[~]] of formal power series with coefficients in the universal enveloping algebra U(g). It can be endowed with a (topologically free) Hopf algebra structure, denoted by (U(g)[[~]], ∆, ǫ, S). Let us recall the definition of a Drinfel’d twist and its semiclassical limit, see [9, 10]. Definition 1.1 (Twist) An element F ∈ (U(g) ⊗ U(g))[[~]] is said to be a twist on U(g)[[~]] if the following three conditions are satisfied: P k i.) F = 1 ⊗ 1 + ∞ k=1 ~ Fk . ii.) (F ⊗ 1)(∆ ⊗ 1)(F) = (1 ⊗ F)(1 ⊗ ∆)(F). iii.) (ǫ ⊗ 1)F = (1 ⊗ ǫ)F = 1. We sometimes use the notation F = Fα ⊗ Fα . The semiclassical limit of a twist gives rise to a wellknown structure on the Lie algebra g called r-matrix, as proved in [10] or [16, Thm. 1.14]. In fact, we have the following claim. Proposition 1.2 Given a twist F on U(g)[[~]], the antisymmetric part of its first order is a classical r-matrix r ∈ g ∧ g. Given a twist we can obtain a deformed Hopf algebra structure on U(g)[[~]]. 3 Proposition 1.3 Let F be a twist on U(g)[[~]]. Then the algebra U(g)[[~]] endowed with coproduct given by ∆F := F∆F−1 , (1.1) α undeformed counit and antipode SF := uF S(X)u−1 F , where uF := F S(Fα ) is again a Hopf algebra denoted by UF (g). As a consequence, the twist automatically defines a Lie bialgebra structure. Given a twist on the universal enveloping algebra, we can always define a star product on any U(g)-module algebra. In particular, let us consider the algebra C ∞ (M ) of smooth functions on a manifold M with pointwise multiplication mM and a Hopf algebra action Φ : U(g) × C ∞ (M ) −→ C ∞ (M ) : (X, f ) 7→ Φ(X, f ) (1.2) This action can be immediately extended to formal power series, allowing the following result. Lemma 1.4 (Universal deformation formula) The product defined by f ⋆F g = mM (Φ(F−1 , (f ⊗ g))) (1.3) for f, g ∈ C ∞ (M )[[~]] is an associative star product quantizing the Poisson structure induced by the semiclassical limit r of F via the action. We denote the deformed algebra by CF∞ (M ). Moreover, it is important to remark that the deformed algebra CF∞ (M ) is now a left UF (g)-module algebra We can give a dual version of the above discussion by using the notions of 2-cocycles and coactions. A more detailed discussion about 2-cocycles and their duality with twist can be found in [2]. Consider the Hopf algebra C ∞ (G), where G is the Lie group corresponding to the finite-dimensional Lie algebra g. It is known that C ∞ (G) and U(g) are dually paired Hopf algebras algebras with pairing denoted by h · , · i. Thus, given a twist F there corresponds an element γ : (C ∞ (G) ⊗ C ∞ (G))[[~]] → K on C ∞ (G)[[~]] defined by γ(f ⊗ g) := hFα , f ihFα , gi, (1.4) for all f, g ∈ C ∞ (G). Roughly, from the condition ii.) mentioned in Definition 1.1 it is easy to see that γ satisfies the 2-cocycle condition γ(f(1) ⊗ g(1) )γ(f(2) g(2) ⊗ h) = γ(g(1) ⊗ h(1) )γ(f ⊗ g(2) h(2) ), (1.5) for any f, g, h ∈ C ∞ (G). Here we used the Sweedler notation. Thus, dualizing the deformed Hopf algebra UF (g) obtained in Proposition 1.3 we immediately obtain a twisted Hopf algebra denoted by Cγ∞ (G) with a new associative product mγG defined by h∆F (X), f ⊗ gi = hX, mγG (f ⊗ g)i (1.6) As it will be used in the following we resume this structure in the following defining. Definition 1.5 (Quantum group) The quantum group corresponding to G is defined to be the Hopf algebra Cγ∞ (G) given by (C ∞ (G)[[~]], mγG , ∆, ǫ, S) where the deformed product is given by (1.6) and coproduct and counit are undeformed. The deformed Hopf algebras Cγ∞ (G) and UF (g) are again dually paired via the same pairing. Finally, if C ∞ (M ) is a left U(g)-module algebra via (1.2), it is automatically a right-C ∞ (G)-comodule algebra (the coaction δ : C ∞ (M ) → C ∞ (M ) ⊗ C ∞ (G) can be easily obtained by dualizing Φ, see [23, Prop. 1.6.11]). In the same spirit of Lemma 1.4, the algebra structure of C ∞ (M ) can be equivalently deformed by considering a 2-cocycle on C ∞ (G) and pushing its deformation on C ∞ (M ) via the coaction δ. 4 2 Hamiltonian actions In this section we introduce the notion of Hamiltonian action in the setting of Poisson Lie groups. This notion has been first defined in [19, 21] in the case of a Poisson Lie group acting on a Poisson manifold with both Poisson structures fixed. In our work we are mainly interested in the case in which the Poisson structure on the manifold is the one induced by the action. This requires a slight generalization of the notion of Hamiltonian action. 2.1 Dressing generators In the same spirit of [19, 21], the notion of Hamiltonian action relies on the definition of momentum map, which provides us of a comparison tool between the dressing orbits and the orbit of the considered action. For this reason, we first focus on the dressing action and in particular on the possible descriptions of the corresponding fundamental vector fields. Let us consider a Lie bialgebra g with dual and double denoted by g∗ and d, respectively. The Lie groups G and G∗ associated to g and g∗ , respectively, turn into Poisson Lie groups. Furthermore, the Lie group D corresponding to the double Lie algebra d is called double of the Poisson Lie group G. Consider g ∈ G, u ∈ G∗ and let ug ∈ D be their product. Since d = g ⊕ g∗ , elements in D close to the unit can be decomposed in a unique way as a product of an element in G and an element in G∗ . Then, there exist elements u g ∈ G and ug ∈ G∗ such that ug = u gug . (2.1) Hence, the action of g ∈ G on u ∈ G∗ is given by (u, g) 7→ (ug)∗G (2.2) where (ug)∗G denotes the G∗ -factor of ug ∈ D. This defines a left action of G on G∗ , called dressing action. This action plays an important role in the context of Poisson actions since its orbits coincides with the symplectic leaves of G∗ and its linearization is the coadjoint action. Let us denote by ℓX the corresponding fundamental vector field for X ∈ g. In the following we introduce the notion of dressing generators, which are one-forms that give us the fundamental vector fields ℓX if contracted with the Poisson bitensor. As it will be seen in the next sections these forms are in general not globally defined, so we use the notation Ω1loc (G∗ ) to denote local forms on G∗ . Definition 2.1 (Dressing generator) The map α : g → Ω1loc (G∗ ) : X 7→ αX is said to be dressing generator with respect to the Poisson structure π on G∗ if the fundamental vector field ℓX of the dressing action can be written as ℓX = π ♯ (αX ) (2.3) and satisfies α[X,Y ] = [αX , αY ]πℓ , dαX = α ∧ α ◦ δ(X). (2.4) (2.5) Here δ denotes the Lie bialgebra structure on g. Remark 2.2 The first example of dressing generators with respect to the standard dual Poisson structure π∗ is given by the left-invariant one-forms corresponding to the element X, as proved in [18, Appendix 2, page 66]. As already mentioned, the dressing generators with respect to a generic Poisson structure on G∗ are in general not globally defined (a concrete example is computed in the next section). However, the contraction with the Poisson tensor still gives rise to a smooth vector field. 5 Here we are interested to the case in which g is endowed with an r-matrix and we consider the Poisson structure πℓ induced by the infinitesimal dressing action ℓ : g → Γ∞ (T G∗ ) via (2.6) πℓ = rij ℓXi ∧ ℓXj . This is a natural candidate since the contraction of πℓ with one-forms satisfying (2.4)-(2.5) gives rise automatically to an infinitesimal Poisson action, as proved in the following Lemma. Lemma 2.3 Given a map α : g → Ω1loc (G∗ ) satisfying (2.4)-(2.5) then we have: i.) The map g ∋ X 7→ πℓ♯ (αX ) ∈ Γ∞ (T G∗ ) is a Lie algebra morphism ii.) The map g ∋ X 7→ πℓ♯ (αX ) ∈ Γ∞ (T G∗ ) is an (infinitesimal) Poisson action. Proof: Let us compute: (2.4) πℓ♯ (α[X,Y ] ) = πℓ♯ ([αX , αY ]πℓ ) (∗) = [πℓ♯ (αX ), πℓ♯ (αY )]. In (∗) we used the fact that πℓ♯ is a Lie algebra morphism with respect to the Lie bracket of one-forms [a, b]πℓ = Lπ♯ (a) b − Lπ♯ (b) a − dπℓ (a, b). Furthermore, we have: ℓ ℓ (2.5) ∧2 πℓ♯ (α ∧ α ◦ δ(X)) = ∧2 πℓ♯ (dαX ) (∗) = dπℓ πℓ♯ (αX ). In (∗) we used dπ (∧p π ♯ )(ξ)) = (∧p+1 π ♯ )(dξ).  Example 2.4 (Dressing generators on ax + b) Let us denote by s the Lie algebra with basis H, E and commutation relation [H, E] = 2E, (2.7) also known as ax + b. The corresponding group is denoted by S and we consider the dressing action S × S ∗ → S. Then we have that The dressing generators with respect to πℓ are given by the local forms 1 1 dy. (2.8) αH = dx and αE = y 2y The complete discussion of this example can be found in the Appendix A. 2.2 Hamiltonian actions Using the notion of dressing generator we give a new definition of Hamiltonian action in this context. Definition 2.5 (Momentum map) Let Φ : G × M → M be an action of (G, πG ) on (M, π) and αX the dressing generator with respect to a Poisson structure πG∗ on G∗ . i.) A momentum map for Φ is a map J : M → G∗ such that ϕ(X) = π ♯ (J ∗ (αX )), (2.9) where ϕ(X) is the fundamental vector field of Φ. In other words, J is defined by the commutativity of the following diagram: ϕ g Γ∞ (T M ) ♯ πM α Γ∞ (T ∗ G∗ ) J∗ 6 Γ∞ (T ∗ M ) (2.10) ii.) A map J : M → G∗ is said to be ℓ-equivariant if it intertwines the fundamental vector field ϕ(X) and the dressing action ℓX for any X. Lemma 2.6 The momentum map J defined above is ℓ-equivariant if and only if is Poisson. Proof: Let us consider generic Poisson structures π on M and πG∗ on G∗ . Thus, J is a Poisson map if and only if ♯ J∗ (π ♯ (J ∗ (α))) = πG ∗ (α). ♯ ♯ ∗ Let α be the dressing generator corresponding to πG∗ . Thus πG ∗ (αX ) = ℓX and π (J (αX )) = ϕ(X) and the equation above coincides with the ℓ-equivariance.  Now the notion of Hamiltonian follows naturally: Definition 2.7 (Hamiltonian action) An action Φ of (G, πG ) on (M, πM ) is said to be Hamiltonian if it is Poisson and is generated by a ℓ-equivariant momentum map J : M → G∗ . Since in the following we mainly use the infinitesimal action ϕ, we say that it is Hamiltonian whenever the corresponding Φ is Hamiltonian. Remark 2.8 i.) If we choose the standard dual Poisson structure on G∗ , the dressing generators are the left-invariant one-forms and the above definition boils down to the definition of momentum map and Hamiltonian action given by Lu in [19, 21]. ii.) Let g be a triangular Lie algebra with r-matrix r, acting on a manifold M by ϕ : g → Γ∞ (T M ). We denote by πr the Poisson structure induced by r via πr = rij ϕ(Xi ) ∧ ϕ(Xj ). (2.11) In this case the action ϕ and its global corresponding are automatically Poisson. (The proof is the same as the one given in Lemma 2.3). Example 2.9 (Dressing action) The easiest example is given by the dressing action. Here the momentum map is just the identity. Example 2.10 (Coadjoint action) Let us consider the Poisson structure πr induced by the coadjoint action. Notice that πr does not coincide with the linear one. As proved in [1, Section 3.3] one can define a map j : g∗ → d by j(ξ) = ξ − r(ξ, · ). Thus, the modified exponential is given by Exp : g∗ → G∗ : Exp(ξ) := prG∗ (exp(j(ξ))). In contrast to the usual exponential map it intertwines the coadjoint action with the dressing action, hence it takes symplectic leaves to symplectic leaves. In other words, we have ℓX = Exp∗ ϕ(X). If G is compact with the Lu-Weinstein Poisson structure [20], Exp is a global diffeomorphism (see [1, Remarks 3.5]). An easy computation shows that Exp is a momentum map for the coadjoint action. Remark 2.11 From the above example we can construct other Hamiltonian actions. Given a standard momentum map µ : M → g∗ which is ad∗ -equivariant we can always construct a momentum map J : M → G∗ by composing µ and Exp. For instance, observing that r♯ : g∗ → g intertwines adjoint and coadjoint actions we can conclude that the adjoint action is Hamiltonian with momentum map given by the composition of r♯ with Exp. Remark 2.12 The reduction can been obtained with various techniques (see e.g. [12]). We here remark that the preimage C = J −1 ({0}) of a ℓ-invariant momentum map is a coisotropic submanifold and IC the corresponding vanishing ideal. Thus the reduced algebra can be easily obtained by the quotient BC /IC where BC = {f ∈ C ∞ (M )|{f, IC } ⊆ IC }. 7 3 Hamiltonian Hopf algebra (co)actions In this section we aim to give a definition of Hamiltonian (co)action in the setting of Hopf algebra (co)actions and a possible quantization procedure. In the same spirit of Definition 2.7, given an Hopf algebra action Φ, a momentum map has to be an intertwiner between dressing action and Φ. In order to introduce this notion we first prove that given a classical Hamiltonian action we can always associate a Hopf algebra action and construct, out of the classical momentum map, the desired intertwiner. First, we observe that any Lie algebra action gives rise to a Hopf algebra action. Lemma 3.1 Consider the infinitesimal action ϕ : g → Γ∞ (T M ). This is equivalent to a Hopf algebra action Φ : U(g) × C ∞ (M ) → C ∞ (M ) by setting Φ(X, f ) := LϕX f, (3.1) where L denotes the Lie derivative. Equivalently, it defines a Hopf algebra coaction δΦ : C ∞ (M ) → C ∞ (M ) ⊗ C ∞ (G) Proof: The Lie algebra elements act as derivations of C ∞ (M ), thus Φ defines a Lie algebra action ϕ : g → Γ∞ (T M ). Since the elements of g generate U(g), the action Φ is given by differential operators with order determined by the natural filtration of the universal enveloping algebra. Conversely, every Lie algebra action ϕ of g on M determines via the fundamental vector fields ϕX ∈ Γ∞ (T M ) a representation of g on C ∞ (M ) by derivations which therefore extends to a Hopf algebra action Φ as above. The action Φ and the coaction δΦ are always equivalent.  In particular, given the infinitesimal dressing action ℓ : g → Γ∞ (T G∗ ) we obtain the Hopf algebra action Λ : U(g) × C ∞ (G∗ ) → C ∞ (G∗ ) by setting: Λ(X, f ) := LℓX f. (3.2) We denote by δΛ the corresponding Hopf algebra coaction. As a next step we lift the notion of dressing generator to the setting of Hopf algebra actions. We observe that, given the Lie algebra representation α : g → Ω1loc (G∗ ), we can define another Hopf algebra action by using the Lie derivative in the direction of a one-form Lα which has been defined by Bhaskara and Viswanath [4]. In particular, for f ∈ C ∞ (G∗ ) Lα f = Lπ♯ (α) f. (3.3) More precisely, we have: Lemma 3.2 Given a dressing generator α : g → Ω1loc (G∗ ), the corresponding map given by U(g) × C ∞ (G∗ ) → C ∞ (G∗ ) : (X, f ) 7→ LαX f is a Hopf algebra action. Furthemore we have Λ(X, f ) = LαX f, (3.4) where Λ(X, f ) is given by (3.2). Proof: First, as in Lemma 3.1 the map U(g) × C ∞ (G∗ ) → C ∞ (G∗ ) : (X, f ) 7→ LαX immediately satisfies the condition to be a Hopf algebra action. Also, from the definition of dressing generator we have ℓX = πℓ♯ (αX ). Thus Λ(X, f ) = LℓX f = Lπ♯ (α ) f ℓ X = LαX f. 8  Now, let us consider a Hamiltonian action ϕ : g → Γ∞ (T M ) with momentum map J : M → G∗ . Notice that its pullback of functions J ∗ : C ∞ (G∗ ) → C ∞ (M ) is an algebra morphism. With an abuse of notation, we also refer to J ∗ as the pullback of forms. Since the latter is always defined, we can extend J to a map J ∗ acting on Lα by J ∗ Lα := LJ ∗ α ◦ J ∗ . (3.5) Theorem 3.3 Let ϕ : g → Γ∞ (T M ) be an Hamiltonian action with momentum map J : M → G∗ and consider the corresponding Hopf algebra action Φ : U(g)×C ∞ (M ) → C ∞ (M ) given by Φ(X) = Lϕ(X) . Then we have: i.) The pullback J ∗ : C ∞ (G∗ ) → C ∞ (M ) of J intertwines Φ and the Hopf algebra action Λ corresponding to the dressing action via (3.2). ii.) The pullback J ∗ : C ∞ (G∗ ) → C ∞ (M ) of J intertwines the corresponding Hopf algebra coaction δΦ and the Hopf algebra coaction δΛ corresponding to the dressing action. Proof: The two claims above can be rephrased by saying that J ∗ defines a U(g)-module algebra morphism and C ∞ (G)-comodule algebra morphism. i.) We already observed that J ∗ : C ∞ (G∗ ) → C ∞ (M ) is an algebra morphism. Thus, we only need to prove that it is a module morphism, i.e. the commutativity of the following diagram: U(g) × C ∞ (G∗ ) Λ C ∞ (G∗ ) id ×J ∗ J∗ U(g) × C ∞ (M ) Φ (3.6) C ∞ (M ) In other words, we need to prove Φ(X, J ∗ f ) = J ∗ (Λ(X, f )). (3.7) Using (3.5) we can easily compute: J ∗ (Λ(X, f )) = J ∗ (LαX f ) = L J ∗ αX J ∗ f = Lπ♯ (J ∗ (αX )) J ∗ f = Lϕ(X) J ∗ f = Φ(X, J ∗ f ). Here we used the fact that, from Definition 2.7, we have ϕ(X) = π ♯ (J ∗ (αX )). ii.) Given the Hopf algebra action Φ we can always find the corresponding Hopf algebra coaction δΦ , as discussed in Section 1. Thus we can immediately state the dual version of the above claim. In fact, dualizing the commutative diagram (3.6) we immediately get the following commutative diagram C ∞ (G∗ ) δΛ C ∞ (G∗ ) ⊗ C ∞ (G) J ∗ ⊗ id J∗ C ∞ (M ) (3.8) C ∞ (M ) ⊗ C ∞ (G) δΦ which gives the comodule morphism condition δΦ ◦ J ∗ = (J ∗ ⊗ id) ◦ δΛ . Since J ∗ is an algebra morphism the claim is proved.  9 Finally, the above discussion motivates the following definition. Let C ∞ (M ) be a U(g)-module algebra where the module structure is given by a generic Hopf algebra action Φ : U(g) × C ∞ (M ) → C ∞ (M ). Equivalently, C ∞ (M ) is endowed with a C ∞ (G)-comodule algebra structure. Furthermore, given the dressing action ℓ we showed that C ∞ (G∗ ) automatically turns into a U(g)-module algebra where the Hopf algebra action Λ is given by (3.2) (and equivalently into a C ∞ (G)-comodule algebra). Definition 3.4 (Hamiltonian (co)action) i.) A Hopf algebra action Φ : U(g) × C ∞ (M ) → C ∞ (M ) is said to be Hamiltonian if there exist a U(g)-module algebra morphism, called momentum map, J : C ∞ (G∗ ) → C ∞ (M ). In other words, Φ is Hamiltonian if it allows a map J satisfying the following condition: Φ(X, Jf ) = J(Λ(X, f )). (3.9) ii.) A Hopf algebra coaction δΦ : C ∞ (M ) → C ∞ (M ) ⊗ C ∞ (G) is said to be Hamiltonian if there exist C ∞ (G)-module algebra morphism J, called momentum map, which intertwines it with the Hopf algebra coaction δΛ corresponding to the dressing action. 3.1 Quantum Hamiltonian coactions via 2-cocycles In this section we prove that, using Drinfeld approach, we obtain a quantization of the Hamiltonian coactions as in Definition 3.4. Since actions and coactions are completely equivalent we here prefer to focus only on the coaction case. Let us consider a twist F on U(g) with corresponding 2-cocycle γ on C ∞ (G). As seen in Definition 1.5, the 2-cocycle γ induces a deformed product ⋆γ and we denote by C~∞ (G) the corresponding quantum group. Furthermore, we obtain a deformed product on the comodule algebras C ∞ (M ) and C ∞ (G∗ ). More precisely, the action ϕ : g → Γ∞ (T M ) induces a star product ⋆ϕ on M whose semiclassical limit is the Poisson structure πr induced by r via ϕ. Similarly, the dressing action ℓ : g → Γ∞ (T G∗ ) induces a star product ⋆ℓ on G∗ . Let us denote by C~∞ (G∗ ) the deformed algebra given by the pair (C ∞ (G∗ )[[~]], ⋆ℓ ) and by C~∞ (M ) the pair (C ∞ (M )[[~]], ⋆ϕ ). Notice that C~∞ (G∗ ) and C~∞ (M ) are now C~∞ (G)-comodule algebras. In other words, the coactions δΦ : C~∞ (M ) → C~∞ (M ) ⊗ C~∞ (G) and δΛ : C~∞ (G∗ ) → C~∞ (G∗ ) ⊗ C~∞ (G) (3.10) are morphisms of algebras. Thus we can state our main result. Theorem 3.5 Let ϕ : g → Γ∞ (T M ) be an Hamiltonian action with momentum map J : M → G∗ . Then the corresponding quantum group coaction δΦ : C~∞ (M ) → C~∞ (M ) ⊗ C~∞ (G) is Hamiltonian in the sense of Definition 3.4. Proof: Since in the Drinfeld approach the coactions do not change but they only intertwine different algebraic structures, the classical momentum map is still a comodule morphism as in Lemma ??. More explicitly, the diagram C~∞ (G∗ ) δΛ C~∞ (G∗ ) ⊗ C~∞ (G) J ∗ ⊗ id J∗ C~∞ (M ) δΦ (3.11) C~∞ (M ) ⊗ C~∞ (G) commutes. Thus, we only need to prove that J ∗ : C~∞ (G∗ ) → C~∞ (M ) is a morphism of algebras. This can be immediately checked by using the UDF (1.3) and Lemma ??. We can extend the action (3.2) by (3.12) Λ(F, f ⊗ g) = Λ(Fα , f ) ⊗ Λ(Fα , g). 10 As a consequence, we have: J ∗ (f ⋆ℓ g) = J ∗ (m(Λ(F, f ⊗ g))) = J ∗ (m(Λ(Fα , f ), Λ(Fα , g))) = (J ∗ Λ(Fα , f ))(J ∗ Λ(Fα , g)) (3.7) = Φ(Fα , J ∗ f )Φ(Fα , J ∗ g) = m(Φ(F−1 , J ∗ f ⊗ J ∗ g)) = J ∗ f ⋆ϕ J ∗ g. A  Dressing generators on ax + b In this appendix we discuss a concrete example of dressing generators. Let s be the Lie algebra with basis H, E and commutation relation [H, E] = 2E, (A.1) also known as the Lie algebra ax + b. Consider the triangular r-matrix r = H ∧ E. This induces the Lie bialgebra structure on g∗ : δ(H) = [r, H ⊗ 1 + 1 ⊗ H] = H ⊗ [E, H] − [E, H] ⊗ H = −2H ∧ E, δ(E) = [r, E ⊗ 1 + 1 ⊗ E] = 0. As a consequence, the dual basis H ∗ , E ∗ satisfies the following commutation relation: [H ∗ , E ∗ ] = −2H ∗ . (A.2) Note that the element r corresponds to the Poisson structure associated to the bilinear symplectic structure ω on s defined by ω(H, E) := 1. Within this set up the Lie algebra structure (A.2) on s∗ is simply obtained by transporting the Lie bracket on s to s∗ under the linear musical isomorphism ♭ : s → s∗ : X 7→ ♭X := ι ω i.e. X [ ♭X, ♭Y ]s∗ := ♭[X, Y ]. (A.3) In our case we have: ♭ ♭ H = E∗ E = −H ∗ The double g := D(s) is given by the vector space s ⊕ (using the notation induced by musical isomorphism) [H, E] = 2E, [H, ♭E] = 2E, [ ♭H, ♭E] = 2 ♭E, [E, ♭E] = 0, s∗ (A.4) equipped with the following Lie brackets [ ♭H, H] = 2( ♭H − H), [E, ♭H] = −2 ♭E. (A.5) We observe that the the first derivative g′ := [g, g] is spanned by E, ♭E and F := ♭H − H and admits the table: [E, F ] = 2E − 2 ♭E =: Z, [E, Z] = [F, Z] = 0. Thus, g′ is isomorphic to the Heisenberg algebra h1 := V ⊕ RZ associated to the symplectic plane (V, Ω) spanned by E and F and structured by [v + zZ, v ′ + z ′ Z] = Ω(v, v ′ )Z with v, v ′ ∈ V 11 and Ω(E, F ) := 1. In this setting, the double D(s) can be viewed as the semidirect product of the Lie algebra h1 with the abelian Lie algebra RH: D(s) ≃ R ⋉ρ h1 whose Lie algebra homomorphism ρ : RH → Der(h1 ) is defined in the basis E, F, Z by   2 0 0 ρ(H) := 0 −2 0. 0 0 0 Lemma A.1 Let s = ax + b. Then we have: i.) The connected simply connected Lie group G := D(s), with Lie algebra given by the vector space g := D(s) := s ⊕ s∗ with Lie algebra structure given by (A.5), is diffeomorphic to the product manifold: G = R × V × R. (A.6) ii.) Within this model, the group law is given by 1 (a, v, z) · (a′ , v ′ , z ′ ) = (a + a′ , v + e2aB v ′ , z + z ′ + Ω(v, e2aB v ′ )) 2 where 1 B := ρ(H) 2 V =  1 0 0 −1  in basis {E, F }. (A.7) (A.8) iii.) Realizing the Lie algebra g as g = RH ⊕ V ⊕ RZ = {(a0 , v0 , z0 )}, (A.9) the exponential mapping is given by   1 2a0 B 1 1 2a0 B exp(a0 , v0 , z0 ) = a0 , (e − I)Bv0 , z0 + Ω(Bv0 , v0 ) + 2 Ω(v0 , e v0 ) . (A.10) 2a0 4a0 8a0 Proof: The connected simply connected Lie group H1 corresponding to h1 can be modelled on V × RZ with group law given by 1 (v, z) · (v ′ , z ′ ) = (v + v ′ , z + z ′ + Ω(v, v ′ )). 2 (A.11) Within this setting, we observe that the symplectic group Sp(V, Ω) (which in our two-dimensional case just coincides with the group SL2 (R)) acts by centre-fixing group-automorphisms on H1 under: R : Sp(V, Ω) × H1 → H1 : (a, (v, z)) 7→ Ra (v, z) := (a(v), z). (A.12) Every sub-group A of Sp(V, Ω) therefore determines the semi-direct product group G := A ⋉R H1 (A.13) modelled on the Cartesian product G = A × H1 with group law defined by (a, a′ ∈ A): 1 (a, v, z) · (a′ , v ′ , z ′ ) := (a · a′ , (v, z) · Ra (v ′ , z ′ )) = (a · a′ , v + a(v ′ ), z + z ′ + Ω(v, a(v ′ ))). 2 12 (A.14) In the case A :=  exp(2aB) =  e2a 0 0 e−2a  (A.15) a∈R the semi-direct product is therefore the Lie group (A.16) G = R × H1 with group law given by (A.7). One then readily verifies that the given expression in (A.10) satisfies the condition exp t(a0 , v0 , z0 ) · exp s(a0 , v0 , z0 ) = exp(t + s)(a0 , v0 , z0 ) for all s, t ∈ R. The fact that d B 2 = I then implies dt exp t(a0 , v0 , z0 ) = (a0 , v0 , z0 ). The computation of the Lie algebra of G is t=0 then performed using the expression of the above exponential mapping (A.10). It identifies with the one of g.  We now pass to realize s and s∗ in the double G. For this we start from expressing the generators at the Lie algebra level: 1 H ∗ = −( Z + E) and E ∗ = H + F. (A.17) 2 The coordinates on s∗ are given by (ν, κ)∗ := exp νE ∗ exp κH ∗ where   1 κ 1 exp κH ∗ = exp κ(−E − Z) = (0, −κE, − ) and exp νE ∗ = ν, (e−2ν − 1)F, 0 . (A.18) 2 2 2 Using the group law (A.7) we get (ν, κ)∗ = Similarly, we have   κ 1 −2ν 2ν − 1)F, − (1 + e ) . ν, −κe E, (e 2 4 2ν  (a, n) := exp(aH) exp(nE) = a, e2a nE, 0 . (A.19) (A.20) Lemma A.2 Let us consider the dressing action S × S ∗ → S. Then we have i.) The dressing generators with respect to the standard dual Poisson structure π∗ are given by the left-invariant forms 1 1 αH = − dx and αE = dy (A.21) y+1 2(y + 1) ii.) The dressing generators with respect to πℓ are given by the local forms αH = 1 dx y and αE = 1 dy 2y (A.22) Proof: The first step consists in computing the fundamental vector field of the dressing action by using the realization obtained above of s and s∗ in terms of the double. More explicitely, using the coordinates (A.19)-(A.20) and the group law (A.7) we have that   e−2a −2ν 1 2a 2ν 2ν −2ν (a, n)(ν, κ)∗ = a + ν, e (n − κe )E, (e − 1)F, − (κ + e κ − ne + n) . (A.23) 2 4 Similarly, we have    1 −2ν 1 2ν 2a 2ν 2ν 2a (ν, κ)∗ (a, n) = ν + a, e (e n − κ)E , (e − 1)F, − κ(1 + e ) + (1 − e )e n . 2 4 13 (A.24) The dressing action S ⋆ × S → S ⋆ therefore amounts to solve the equation (a, n)(ν, κ)∗ = (ν, κ)∗ (a, n) for (ν, κ)∗ as a function of a, n, κ, ν. From an easy computation it follows that the solution is given by  = κ − nη(ν) κ (A.25) η(ν) = e−2a η(ν) where η is the diffeomorphism defined by η : R →] − 1, ∞[ : x 7→ η(x) := e−2x − 1. Considering the coordinate system S ⋆ ֒→ R2 : ξ := (ν, κ)∗ 7→ (x, y) := (κ, η(ν)), the local right dressing action then reads: (x, y) · (a, n) := (x − ny, e−2a y). (A.26) Indeed, the multiplication map S ⋆ × S → G : (ξ, x) 7→ ξ · s (A.27) is an open embedding. Hence locally one may set: s · ξ = ξ s · sξ with sξ ∈ S One then notes that for all s1 , s2 ∈ S and ξ ∈ and ξ s ∈ S ⋆ . (A.28) S⋆: s2 ξ s1 s2 (s1 s2 )ξ = s1 s2 ξ = s1 ξ s2 xξ2 = (ξ s2 )s1 s1ξ sξ2 which implies (A.29) ξ s1 s2 = (ξ s2 )s1 . S⋆ S⋆ ξs (A.30) S⋆ Hence the map ×S → : (ξ, s) 7→ which given elements s ∈ S and ξ ∈ expresses the ⋆ S -component (local) of the product s · ξ in terms of the decomposition (A.27) is a right action of S on S ⋆ . The latter globalizes under the usual matrix left-action of the affine group on the plane as    2a x e 0 2 2 −1 . (A.31) S × R → R : (s = (a, n), v = (x, y)) 7→ s.v := v.s := 2a y ne 1 Now we express the group multiplication in S ⋆ within the above coordinate system:   (x, y).(x′ , y ′ ) := Φ Φ−1 (x, y).Φ−1 (x′ , y ′ ) = (y ′ + 1)x + x′ , (y ′ + 1)y + y ′ . (A.32) The unit consists in the vector origin (0, 0) and the inverse (which is only local at the level of the 1 entire ambient space R2 ) is given by (x, y)−1 = y+1 (−x, −y). It is useful to rewrite the dressing action using musical notation; in this case we consider the coordinate system S ⋆ ֒→ R2 : ξ := (ν, κ)∗ 7→ (x, y) := (κ, η(ν)), (A.33) where η(x) = 1 − e−2x and the local right dressing action ξ.(a, n) := (κ, ν) then reads: (x, y) · (a, n) = (x + ny, e−2a y). (A.34) This implies that the dressing action is infinitesimally generated by the following fields: d ♭b b(x,y) := d (x, y)(t, 0) = −2y∂y and ♭ E (x, y)(0, t) = y∂x . (A.35) H(x,y) := t=0 dt dt t=0 The next step consists in computing explicitely the dressing generators. Note that there are 3 Poisson structures involved here on the image U of S ⋆ ֒→ R2 , the dual Poisson Lie group structure π∗ = 2y(y + 1)∂x ∧ ∂y , (A.36) the Poisson structure πℓ induced by the action πℓ = 2y 2 ∂x ∧ ∂y (A.37) and the linear one πs⋆ . It is easy to see that π∗ = πℓ + πs⋆ . (A.38) Finally, imposing the condition (2.3) we obtain that the dressing generators with respect to to π∗ and πℓ we get the expressions (A.21) and (A.22), resp.  14 References [1] Alekseev, A., Meinrenken, E.: Linearization of Poisson Lie group structures. J. Symplectic Geom. 14 (2016), 227–267. 7 [2] Aschieri, P, Bieliavsky, P., Pagani, C., Schenkel, A.: Noncommutative principal bundles through twist deformation. Comm. Math. Phys. 352 (2017), 287–344. 4 [3] Bayen, F., Flato, M., Frønsdal, C., Lichnerowicz, A., Sternheimer, D.: Deformation Theory and Quantization. Ann. Phys. 111 (1978), 61–151. 2 [4] Bhaskara, K., Viswanath, K.: Calculus on Poisson manifolds. Bull. London Math. Soc. 20 (1988), 68–72. 8 [5] Bieliavsky, P., Bonneau, Ph., Maeda, Y.: Universal deformations formulae, symplectic Lie groups and symmetric spaces; Pacific J. Math. 230 (2007), 41–57. 2 [6] Bieliavsky, P., Gayral, V.: Deformation Quantization for Actions of Kählerian Lie Groups. Mem. Amer. Math. Soc. 236 (2015). 2 [7] Bieliavsky, P., Esposito, C., Waldmann, S., Weber, T.: Products. Lett. Math. Phys. 108 (2018), 1341–1350. 2 Obstructions for Twist Star [8] Chari, V., Pressley, A.: A Guide to Quantum Groups. Cambridge University Press, Cambridge, 1994. 2 [9] Drinfeld, V. G.: Hamiltonian structures on Lie groups, Lie bialgebras, and the geometric meaning of the classical Yang-Baxter equations. Soviet Math. Dokl. 27 (1983), 285–287. 2, 3 [10] Drinfeld, V. G.: Quantum groups. J. Sov. Math. 41 (1988), 898–918. 2, 3 [11] D’Andrea, F., Weber, T.: Twist star products and Morita equivalence. C.R. Acad. Sci. Paris, Ser. I 355 (2017), 1178–1184. 2 [12] Esposito, C.: Poisson Reduction. Trends in Mathematics, Geom. Methods Phys. (2014), 131– 142. 7 [13] Esposito, C., Schnitzer, J., Waldmann, S.: A Universal Construction of Universal Deformation Formulas, Drinfel’d Twists and their Positivity. Pacific J. Math. 291 (2017), 319–358. 2, 3 [14] Fedosov, B.:Non Abelian Reduction in deformation quantization. Lett. Math. Phys. 43 (1998), 137–154. 2 [15] Gerstenhaber, M.: New Universal Deformation Formulas for deformation quantization. Preprint arXiv:1802.04919. 2 [16] Giaquinto, A., Zhang, J. J.: Bialgebra actions, twists, and universal deformation formulas. J. Pure Appl. Algebra 128.2 (1998), 133–152. 3 [17] Kontsevich, M.: (2003), 157–216. 2 Deformation quantization of Poisson manifolds. Lett. Math. Phys. 66.3 [18] Kosmann-Schwarzbach, Y.:Lie bialgebras, Poisson Lie groups and dressing transformations. In: Integrability of Nonlinear Systems, vol. 638 in Lecture Notes in Physics, 107—173. SpringerVerlag, second. edition, 2004. 5 15 [19] Lu, J.-H.: Multiplicative and Affine Poisson structure on Lie groups. PhD thesis, University of California (Berkeley), 1990. 3, 5, 7 [20] Lu, J.-H., Weinstein, A.:Poisson Lie groups, dressing transformations, and Bruhat decompositions. J. Differential Geom. 31.2 (1990), 501–526. 7 [21] Lu, J.-H.: Momentum mappings and reduction of Poisson actions. In: Symplectic geometry, groupoids, and integrable systems, Math. Sci. Res. Inst. Publ., Berkeley, California, 1991. 3, 5, 7 [22] Lu, J.-H.: Moment Maps at the Quantum level. Comm. Math. Phys. 157.2 (1993), 389–404. 2 [23] Majid, S.: Foundations of Quantum Group Theory. Cambridge University Press (1995). 2, 4 16