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

Graded covering of a supermanifold I.
The case of a Lie supergroup

Mikolaj Rotkiewicz and Elizaveta Vishnyakova
Abstract.

We generalize the Donagi and Witten construction of a first obstruction class for splitting of a supermanifold via differential operators using the theory of nn-fold vector bundles and graded manifolds. Applying the generalized Donagi–Witten construction we obtain a family of embeddings of the category of supermanifolds into the category of nn-fold vector bundles and into the category of graded manifolds. This leads to a realization of any non-split supermanifold in terms of a collection of vector bundles and some morphism between them. Further we study the images of these embeddings into the category of graded manifolds in the case of a Lie supergroup and a Lie superalgebra. We show that these images satisfy universal property of a graded covering or a graded semicovering.

1. Introduction

This is the first part of a series of papers about \mathbb{Z}-graded coverings of a supermanifold. In this paper we investigate the case of a Lie supergroup and its Lie superalgebra. In [DW2, Section 2] Donagi and Witten gave a description of the first obstruction class ω2\omega_{2} for a supermanifold to be split via differential operators. More precisely, Donagi and Witten constructed an exact sequence of locally free sheaves with Atiyah class ω2\omega_{2}. This is a very interesting and important observation that pure even vector bundles can keep certain information about the original (non-split) supermanifold. In the case a supermanifold has odd dimension 22, the exact sequence of Donagi and Witten keeps the whole information of this supermanifold.

In this paper we generalize this idea. We show that the Donagi and Witten exact sequence corresponds to a double vector bundle with some additional structure, the whole information of which can be reduced to a graded manifold of degree 22. Further we suggest to use a dual approach which leads to iterated differential forms, i.e. functions on the iterated antitangent bundle (as it is explained below) instead of differential operators, which leads to a simplification of the Donagi and Witten construction and suggests a natural generalization. For any (non-split) supermanifold \mathcal{M} of odd dimension n\leq n we construct a pure even geometric object, an nn-fold vector bundle Vbn()\mathrm{Vb}_{n}(\mathcal{M}) which keeps the whole information about the original supermanifold. Moreover the obtained nn-fold vector bundle possesses additional symmetries. We interpret this in the following way: the nn-fold vector bundle arising from a supermanifold, can be reduced to a graded manifold of degree nn.

Summing up we construct a family Fn\mathrm{F}_{n}, where n=2,3,n=2,3,\ldots\infty, of functors from the category of supermanifolds of odd dimension nn to the category of nn-fold vector bundles and to the category of graded manifolds of degree nn, where the first functor F2\mathrm{F}_{2} is a modification of the Donagi–Witten construction. Each functor Fn\mathrm{F}_{n} is a composition of four functors: the (n1)(n-1)-iteratation of the antitangent functor 𝐓\mathbf{T}, the functor split gr\mathrm{gr} (gr\mathrm{gr}\mathcal{M} denotes the retract of the supermanifold \mathcal{M}) , which is defined in the category of supermanifolds, the functor parity change π\pi, which is defined in the category of nn-fold vector bundles and another functor ι\iota, which we call inverse that was discovered independently in [BGR] (under the name the diagonalization functor) and [Vi1]. We show that the functor Fn\mathrm{F}_{n} determines an embedding of the category of supermanifolds of odd dimension nn (or smaller) into the category of graded manifolds of degree nn. The functor F\mathrm{F}_{\infty} is a limit of functors Fn\mathrm{F}_{n}. This functor defines an embedding of the category of supermanifolds into the category of graded manifolds of any degree. This means that the graded manifold F()\mathrm{F}_{\infty}(\mathcal{M}), the image of a (non-split) supermanifold \mathcal{M}, contains the whole information about the original supermanofold \mathcal{M}.

Moreover, for a Lie supergroup 𝒢\mathcal{G} we prove that F(𝒢)\mathrm{F}_{\infty}(\mathcal{G}) satisfies universal properties. These properties of F(𝒢)\mathrm{F}_{\infty}(\mathcal{G}) led use to introduce the notion of a graded covering of 𝒢\mathcal{G}. That is: the covering space is unique up to isomorphism and every homomorphism from a graded Lie supergroup to the Lie supergroup 𝒢\mathcal{G} can be lifted to F(𝒢)\mathrm{F}_{\infty}(\mathcal{G}). We also introduce a notion of a graded semicovering of 𝒢\mathcal{G}. This explains the meaning of other functors Fn\mathrm{F}_{n}. Our covering (semicovering) is in some sense a ”local diffeomorphism” of Lie supergroups. Further we show that the Lie superalgebra of F(𝔤)\mathrm{F}^{\prime}_{\infty}(\mathfrak{g}) is an example of a loop algebra construction by Allison, Berman, Faulkner, and Pianzola [ABFP], see also [Eld]. The loop algebra was used by several authors to investigate graded-simple (Lie) algebras. Therefore results of our paper establish a connection between Donagi–Witten observation and pure algebraic results. Our method leads to a notion of a graded covering and semicovering of any supermanifold as well. However due to technical difficulties of all proofs in this case we present the general theory of covering and semicovering spaces for any supermanifold in the second part of this research [RV].

Our results are especially interesting in the complex-analytic (and algebraic) category. The reason is the following. According to the Batchelor–Gawedzki Theorem any smooth supermanifold is (non-canonically) split, that is its structure sheaf is isomorphic to the wedge power of a certain vector bundle. Therefore very often we can study geometry of a split supermanifold using geometry of vector bundles. This is not the case in the complex-analytic situation. The study of non-split supermanifolds was initiated in [Ber, Gr], where the first non-split supermanifold was described. Significant advances in this direction were achieved in the work of A.L. Onishchik. Interest in this problem arose again after Donagi and Witten’s papers [DW1, DW2], where they proved that the moduli space of super Riemann surfaces is non-split (or more generally not projected). To each (non-split) supermanifold \mathcal{M} our functor Fn\mathrm{F}_{n} associates a graded manifold Fn()\mathrm{F}_{n}(\mathcal{M}) and the image Fn()\mathrm{F}_{n}(\mathcal{M}) keeps the whole information about the (non-split) supermanifold \mathcal{M} for sufficiently large nn. Therefore we can study a non-split supermanifold in the category of graded manifolds using the tools of classical complex geometry due the fact that geometrically a graded manifold is a family of vector bundles.

Our method studying a supermanifold is interesting in the case of a split supermanifold as well, for instance if a split supermanifold possesses an additional structure. Any Lie supergroup 𝒢\mathcal{G} is totally split, i.e. as a supermanifold it is a cartesian product of its underlying space 𝒢0\mathcal{G}_{0} with a fixed odd vector space 𝔤1¯\mathfrak{g}_{\bar{1}} — the odd part of the Lie superalgebra of 𝒢\mathcal{G}. However the supermanifold 𝒢\mathcal{G} has an additional structure which can be non-trivial: the Lie supergroup structure. We show that any graded Lie supergroup Fn(𝒢)\mathrm{F}_{n}(\mathcal{G}) contains the whole information about the original supergroup 𝒢\mathcal{G} for any nn, non only for sufficiently large nn. Moreover our result is the first step toward a graded covering of a homogeneous supermanifold, which is very often non-split. For instance all (except of few exceptional cases) super-grassmannians and flag supermanifolds are non-split.

Acknowledgements: M. R. was partially supported by Capes/PrInt, UFMG. E. V. was partially supported by Coordenação de Aperfeiçoamento de Pessoal de Nível Superior - Brasil (CAPES) – Finance Code 001, (Capes-Humboldt Research Fellowship) and by Tomsk State University, Competitiveness Improvement Program. M. R. thanks UFMG Department of Mathematics for its kind hospitality during the realization of this work. The authors thank Peter Littelmann for his suggestion to investigate the case of a Lie supergroup and a Lie superalgebra. This helped us to understand the meaning of the functor F\mathrm{F}_{\infty}. We also thank Alexey Sharapov and Mikhail Borovoi for useful comments.

2. Preliminaries

2.1. Lie superalgebras, graded Lie superalgebras and graded Lie superalgebras of type Δ\Delta

Throughout the paper we work over the field 𝕂=\mathbb{K}=\mathbb{C} or \mathbb{R}. However the algebraic part of the paper holds true over any field of characteristic 0. We shall work in a supergeometry context. Our vector spaces are 2\mathbb{Z}_{2}-graded, V=V0¯V1¯V=V_{\bar{0}}\oplus V_{\bar{1}}, elements of Vi¯{0}V_{\bar{i}}\setminus\{0\} are called homogenous of parity i¯2\bar{i}\in\mathbb{Z}_{2}. For vVi¯v\in V_{\bar{i}} we write |v|=i¯|v|=\bar{i}. Our main references on supergeomtery, in particular on Lie superalgebras, are [Kac, BLMS, Ber, L] and [Var].

Definition 1.

A Lie superalgebra is a 2\mathbb{Z}_{2}-graded vector superspace 𝔤=𝔤0¯𝔤1¯\mathfrak{g}=\mathfrak{g}_{\bar{0}}\oplus\mathfrak{g}_{\bar{1}} with a graded bilinear operation [,][\,,] satisfying the following conditions:

(1) [x,y]=(1)|x||y|[y,x],[x,y]=-(-1)^{|x||y|}[y,x],
(2) [x,[y,z]](1)|x||y|[y,[x,z]]=[[x,y],z],[x,[y,z]]-(-1)^{|x||y|}[y,[x,z]]=[[x,y],z],

where xx, yy, and zz are homogeneous elements.

In more general setting, if AA is an abelian group, an AA-graded Lie superalgebra is a vector space 𝔤\mathfrak{g} graded by an abelian group AA, 𝔤=aA𝔤a\mathfrak{g}=\oplus_{a\in A}\mathfrak{g}_{a}, the bracket on 𝔤\mathfrak{g} is AA-graded, i.e. [𝔤a,𝔤b]𝔤a+b[\mathfrak{g}_{a},\mathfrak{g}_{b}]\subset\mathfrak{g}_{a+b}, and satisfies a generalized Jacobi identity in which the sign (1)|x||y|(-1)^{|x||y|} is replaced with (1)δ(a)δ(b)(-1)^{\delta(a)\delta(b)}, where δ:A2\delta:A\to\mathbb{Z}_{2} is a group homomorphism.

Definition 2.

Let 𝔤\mathfrak{g} be an AA-graded Lie superalgebra. Then

supp(𝔤)={αA|𝔤α{0}}.supp(\mathfrak{g})=\{\alpha\in A\,\,|\,\,\mathfrak{g}_{\alpha}\neq\{0\}\}.

In this paper most of the time we will work with r\mathbb{Z}^{r}-graded Lie superalgebras. Let α1,,αr\alpha_{1},\ldots,\alpha_{r} are formal variables.

Definition 3.

A weight system is a pair (Δ,χ)(\Delta,\chi), where Δ\Delta is a subset in α1αr\mathbb{Z}\alpha_{1}\oplus\cdots\oplus\mathbb{Z}\alpha_{r} satisfying the following properties

  1. (1)

    0Δ0\in\Delta and αiΔ\alpha_{i}\in\Delta, where i=1,,ri=1,\ldots,r;

  2. (2)

    if δΔ\delta\in\Delta and δ=iaiαi\delta=\sum_{i}a_{i}\alpha_{i}, where aia_{i}\in\mathbb{Z}, then ai0a_{i}\geqslant 0,

and χ:A2\chi:A\to\mathbb{Z}_{2}, δ|δ|\delta\mapsto|\delta|, is a group homomorphism. We call |δ||\delta| the parity of δ\delta. A weight system (Δ,χ)(\Delta,\chi) is called multiplicity free, if δ=aiαiΔ\delta=\sum a_{i}\alpha_{i}\in\Delta implies ai{0,1}a_{i}\in\{0,1\}.

Let (Δ,χ)(\Delta,\chi) be a weight system. Then Δ=Δ0¯Δ1¯\Delta=\Delta_{\bar{0}}\cup\Delta_{\bar{1}}, where Δ0¯\Delta_{\bar{0}} contains all even weights, while Δ1¯\Delta_{\bar{1}} contains all odd ones. Note that 0 is always in Δ0¯\Delta_{\bar{0}}. A Lie superalgebra 𝔤=𝔤0¯𝔤1¯\mathfrak{g}=\mathfrak{g}_{\bar{0}}\oplus\mathfrak{g}_{\bar{1}} is called a graded Lie superalgebra of type (Δ,χ)(\Delta,\chi), if 𝔤\mathfrak{g} is a r\mathbb{Z}^{r}-graded Lie superalgebra of type Δ\Delta and

𝔤0¯=δΔ0¯𝔤δ,𝔤1¯=δΔ1¯𝔤δ.\mathfrak{g}_{\bar{0}}=\bigoplus\limits_{\delta\in\Delta_{\bar{0}}}\mathfrak{g}_{\delta},\quad\mathfrak{g}_{\bar{1}}=\bigoplus\limits_{\delta\in\Delta_{\bar{1}}}\mathfrak{g}_{\delta}.

Very often we will omit χ\chi and write a Lie superalgebra of type Δ\Delta meaning a Lie superalgebra of type (Δ,χ)(\Delta,\chi).

2.2. Supermanifolds, graded manifolds and nn-fold vector bundles

2.2.1. Supermanifolds

We consider a complex-analytic supermanifold in the sense of Berezin and Leites [Ber, L], see also [BLMS]. Thus, a supermanifold =(0,𝒪)\mathcal{M}=(\mathcal{M}_{0},\mathcal{O}_{\mathcal{M}}) of dimension n|mn|m is a 2\mathbb{Z}_{2}-graded ringed space that is locally isomorphic to a superdomain in n|m\mathbb{C}^{n|m}. Here the underlying space 0\mathcal{M}_{0} is a complex-analytic manifold. The dimension nn of the underlying manifold 0\mathcal{M}_{0} is called even dimension of \mathcal{M}, while mm is called odd dimension of \mathcal{M}. A morphism 𝒩\mathcal{M}\to\mathcal{N} between two supermanifolds is a morphism between 2\mathbb{Z}_{2}-graded ringed spaces, this is, a pair F=(F0,F)F=(F_{0},F^{*}), where F0:0𝒩0F_{0}:\mathcal{M}_{0}\to\mathcal{N}_{0} is a holomorphic mapping and F:𝒪𝒩(F0)(𝒪)F^{*}:\mathcal{O}_{\mathcal{N}}\to(F_{0})_{*}(\mathcal{O}_{\mathcal{M}}) is a homomorphism of sheaves of 2\mathbb{Z}_{2}-graded ringed spaces. A morphism FF is called an isomorphism if FF is invertible. A supermanifold \mathcal{M} is called split, if its structure sheaf is isomorphic to \bigwedge\mathcal{E}^{*}, where \mathcal{E} is a sheaf of sections of a certain vector bundle 𝔼[1¯]\mathbb{E}[\bar{1}]. (Here 𝔼[1¯]\mathbb{E}[\bar{1}] means a usual vector bundle 𝔼\mathbb{E} additionally assumed that its local sections are odd.) We see that in this case the structure sheaf is \mathbb{Z}-graded. A morphism of split supermanifolds is called split if it preserves the fixed \mathbb{Z}-gradings. Split supermanifolds with split morphisms form a category of split supermanifolds.

According to the Batchelor–Gawedzki Theorem any smooth supermanifold is split. This is not true in the complex-analytic case, see [Ber, Gr]. However we can define a functor from the category of supermanifols to the category of split supermanifolds. Let =(0,𝒪)\mathcal{M}=(\mathcal{M}_{0},\mathcal{O}) be a supermanifold. (Sometimes we will omit \mathcal{M} in 𝒪\mathcal{O}_{\mathcal{M}} if the meaning of 𝒪\mathcal{O} is clear from the context.) Then its structure sheaf possesses the following filtration

(3) 𝒪=𝒥0𝒥𝒥2𝒥p,\mathcal{O}=\mathcal{J}^{0}\supset\mathcal{J}\supset\mathcal{J}^{2}\supset\cdots\supset\mathcal{J}^{p}\supset\cdots,

where 𝒥\mathcal{J} is the sheaf of ideals generated by odd elements in 𝒪\mathcal{O}. We define

gr():=(0,gr𝒪),\mathrm{gr}(\mathcal{M}):=(\mathcal{M}_{0},\mathrm{gr}\mathcal{O}),

where

gr𝒪=p0𝒥p/𝒥p+1.\mathrm{gr}\mathcal{O}=\bigoplus_{p\geq 0}\mathcal{J}^{p}/\mathcal{J}^{p+1}.

The supermanifold gr()\mathrm{gr}(\mathcal{M}) is split, that is its structure sheaf is isomorphic to \bigwedge\mathcal{E}^{*}, where =𝒥/𝒥2\mathcal{E}^{*}=\mathcal{J}/\mathcal{J}^{2} is a locally free sheaf. Since any morphism FF of supermanifolds preserves the filtration (3), the morphism gr(F)\mathrm{gr}(F) is defined. Summing up, the functor gr\mathrm{gr} is a functor from the category of supermanifolds to the category of split supermanifolds, see for example [Vi2, Section 3.1] for details. We can apply the functor gr\mathrm{gr} to a Lie supergroup 𝒢\mathcal{G} and we will get a split Lie supergroup gr(𝒢)\mathrm{gr}(\mathcal{G}). Later we also will define a corresponding functor gr\mathrm{gr}^{\prime} for Lie superalgebras so that Liegr(𝒢)=grLie(𝒢)\operatorname{Lie}\circ\mathrm{gr}(\mathcal{G})=\mathrm{gr}^{\prime}\circ\operatorname{Lie}(\mathcal{G}) for any Lie supergroup 𝒢\mathcal{G}.

2.2.2. Graded manifolds of type Δ\Delta, graded manifolds of degree nn and rr-fold vector bundles

Let us start with a notion of a graded manifold of type (Δ,χ)(\Delta,\chi), where (Δ,χ)(\Delta,\chi) is as in Definition 3. Again very often we will omit χ\chi, and write a graded manifold of type Δ\Delta. Let

V=δΔVδ,V0¯=δΔ0¯Vδ,V1¯=δΔ1¯Vδ.V=\bigoplus_{\delta\in\Delta}V_{\delta},\quad V_{\bar{0}}=\bigoplus_{\delta\in\Delta_{\bar{0}}}V_{\delta},\quad V_{\bar{1}}=\bigoplus_{\delta\in\Delta_{\bar{1}}}V_{\delta}.

Consider a r\mathbb{Z}^{r}-graded ringed space 𝒰=(𝒰0,𝒪𝒰)\mathcal{U}=(\mathcal{U}_{0},\mathcal{O}_{\mathcal{U}}), where 𝒰0V0\mathcal{U}_{0}\subset V^{*}_{0} is an open set, 𝒪𝒰:=𝒰0𝕂S(V)\mathcal{O}_{\mathcal{U}}:=\mathcal{F}_{\mathcal{U}_{0}}\otimes_{\mathbb{K}}S^{*}(V) and 𝒰0\mathcal{F}_{\mathcal{U}_{0}} is the sheaf of smooth or holomorphic functions on 𝒰0\mathcal{U}_{0}.

Definition 4.

We call the ringed space 𝒰\mathcal{U} a graded domain of type (Δ,χ)(\Delta,\chi) and of dimension {nδ}δΔ\{n_{\delta}\}_{\delta\in\Delta}, where nδ=dimVδn_{\delta}=\dim V_{\delta}.

Let us choose a basis (xi)(x_{i}) in V0V_{0} and a basis (ξjδ)(\xi_{j}^{\delta}) in any VδV_{\delta}. Then the set (xi,ξjδ)δΔ{0}(x_{i},\xi_{j}^{\delta})_{\delta\in\Delta\setminus\{0\}} is a system of local coordinates on 𝒰\mathcal{U}. We assign the weight 0 and the parity 0¯\bar{0} to xix_{i} and the weight δ\delta and the parity |δ||\delta| to ξjδ\xi_{j}^{\delta}. Such coordinates (xi,ξjδ)(x_{i},\xi_{j}^{\delta}) are called graded.

Definition 5.

A graded manifold of type (Δ,χ)(\Delta,\chi) and of dimension {nδ}δΔ\{n_{\delta}\}_{\delta\in\Delta} is a r\mathbb{Z}^{r}-graded ringed space 𝒩=(𝒩0,𝒪𝒩)\mathcal{N}=(\mathcal{N}_{0},\mathcal{O}_{\mathcal{N}}), that is locally isomorphic to a graded domain of type (Δ,χ)(\Delta,\chi) and of dimension {nδ}δΔ.\{n_{\delta}\}_{\delta\in\Delta}.

A morphism of graded manifolds of type (Δ,χ)(\Delta,\chi) is a morphism of the corresponding r\mathbb{Z}^{r}-graded ringed spaces.

If Δ={0,1,,n}\Delta=\{0,1,\ldots,n\} and χ(1)=1¯\chi(1)=\bar{1}, a graded manifold of type Δ\Delta is also called a graded manifold of degree nn. (See [R] for more information about graded manifolds of degree nn.) If Δ\Delta is multiplicity free, this is if δ=iaiαiΔ\delta=\sum_{i}a_{i}\alpha_{i}\in\Delta then ai{0,1}a_{i}\in\{0,1\}, a graded manifold of type Δ\Delta is called an rr-fold vector bundle of type Δ\Delta. This definition of an rr-fold vector bundle is equivalent to a classical one as it was shown in [GR, Theorem 4.1].

Remark 6.

Now let Δ\Delta be as in Definition 3 and ΔΔ\Delta^{\prime}\subset\Delta satisfies the following property: if δΔ\delta\in\Delta^{\prime} and δ=iδi\delta=\sum_{i}\delta_{i} for δiΔ\delta_{i}\in\Delta, then δi\delta_{i} are also in Δ\Delta^{\prime}. In this case to any graded manifold 𝒩\mathcal{N} of type Δ\Delta we can assign a graded manifold 𝒩\mathcal{N}^{\prime} of type Δ\Delta^{\prime}. A detailed description of this well-known construction can be found for example in [Vi1, Section 4.1]. For instance to any graded manifold of degree nn we can assign a graded manifold of degree 0n<n0\leq n^{\prime}<n. In this case we have a natural morphism 𝒩𝒩\mathcal{N}\to\mathcal{N}^{\prime}, which is called projection.

2.3. Lie supergroups and super Harish-Chandra pairs

2.3.1. Lie supergroups

A Lie supergroup is a group object in the category of supermanifolds. In other words a Lie supergroup is a supermanifold 𝒢=(𝒢0,𝒪𝒢)\mathcal{G}=(\mathcal{G}_{0},\mathcal{O}_{\mathcal{G}}) with three morphisms: μ:𝒢×𝒢𝒢\mu:\mathcal{G}\times\mathcal{G}\to\mathcal{G} (the multiplication morphism), κ:𝒢𝒢\kappa:\mathcal{G}\to\mathcal{G} (the inversion morphism), ε:(pt,)𝒢\varepsilon:(\mathrm{pt},\mathbb{C})\to\mathcal{G} (the identity morphism). Moreover, these morphisms have to satisfy the usual conditions, modeling the group axioms. The underlying manifold 𝒢0\mathcal{G}_{0} of 𝒢\mathcal{G} is a smooth or complex-analytic Lie group. As in the theory of Lie groups and Lie algebras, we can define the Lie superalgebra 𝔤=Lie(𝒢)\mathfrak{g}=\operatorname{Lie}(\mathcal{G}) of a Lie group 𝒢\mathcal{G}. By definition, 𝔤\mathfrak{g} is the subalgebra of the Lie superalgebra of vector fields on 𝒢\mathcal{G} consisting of all right invariant vector fields on 𝒢\mathcal{G}. Any right invariant vector field Y𝔤Y\in\mathfrak{g} has the following form

(4) Y=(Yeid)μ,Y=(Y_{e}\otimes\mathrm{id})\circ\mu^{*},

where Ye(𝔪e/𝔪e2)Y_{e}\in(\mathfrak{m}_{e}/\mathfrak{m}^{2}_{e})^{*} and 𝔪e\mathfrak{m}_{e} is the maximal ideal of (𝒪𝒢)e(\mathcal{O}_{\mathcal{G}})_{e}. Note that as the case of classical geometry the vector superspace (𝔪e/𝔪e2)(\mathfrak{m}_{e}/\mathfrak{m}^{2}_{e})^{*} may be identified with the vector superspace of all maps D:(𝒪𝒢)eD:(\mathcal{O}_{\mathcal{G}})_{e}\to\mathbb{C} satisfying the Leibniz rule. Further the map YeY=(Yeid)μY_{e}\mapsto Y=(Y_{e}\otimes\mathrm{id})\circ\mu^{*} is an isomorphism of (𝔪e/𝔪e2)(\mathfrak{m}_{e}/\mathfrak{m}^{2}_{e})^{*} onto 𝔤\mathfrak{g}.

Similarly we can define a graded Lie supergroup of type Δ\Delta as a group object in the category of graded manifolds of type Δ\Delta or a graded Lie supergroup of degree nn as a group object in the category of graded manifolds of degree nn. Let us consider some examples of Lie supergroups with an additional gradation in the structure sheaf.

Example 7 (Lie supergroup GL(V)GL(V)).

The space of endomorphisms of a vector superspace V=V0¯V1¯V=V_{\bar{0}}\oplus V_{\bar{1}} has a natural gradation by integer numbers 1,0,1-1,0,1:

End(V)=i=1,0,1Endi(V),\operatorname{End}(V)=\bigoplus\limits_{i=-1,0,1}\operatorname{End}^{i}(V),

where

End(V)0=End(V0¯)\displaystyle\operatorname{End}(V)^{0}=\operatorname{End}(V_{\bar{0}}) End(V1¯),End(V)1=Hom(V0¯,V1¯),\displaystyle\oplus\operatorname{End}(V_{\bar{1}}),\quad\operatorname{End}(V)^{1}=\operatorname{Hom}(V_{\bar{0}},V_{\bar{1}}),
End(V)1=Hom(V1¯,V0¯)\displaystyle\operatorname{End}(V)^{-1}=\operatorname{Hom}(V_{\bar{1}},V_{\bar{0}})

giving rise to a \mathbb{Z}-graded Lie superalgebra 𝔤𝔩(V)\mathfrak{gl}(V). It integrates to the \mathbb{Z}-graded Lie supergroup GL(V)\mathrm{GL}(V). Consider the case dimV=(1|1)\dim V=(1|1). The entries of a matrix

A=[aαβb]GL(𝕂1|1),A=\begin{bmatrix}a&\alpha\\ \beta&b\end{bmatrix}\in\mathrm{GL}(\mathbb{K}^{1|1}),

constitute a graded coordinate domain, the subdomain of 𝕂2|2\mathbb{K}^{2|2}, the coordinates aa, bb are of degree 0, and α\alpha, β\beta are of degrees 1-1 and 11, respectively. The matrix multiplication respects the assigned grading, e.g. m(α)=aα′′+αb′′m^{*}(\alpha)=a^{\prime}\alpha^{\prime\prime}+\alpha^{\prime}b^{\prime\prime} is of degree 0=0+0=1+(1)0=0+0=1+(-1). Note that GL(V)\mathrm{GL}(V) with this gradation is not a graded manifold of degree nn, since this supermanifold possesses graded coordinates of negative degree.

Remark 8.

In this paper we denote by 𝐓\mathbf{T} the antitangent functor. This means that 𝐓\mathbf{T}\mathcal{M} is the vector bundle obtained from the tangent bundle T\mathrm{T}\mathcal{M} by means of the parity functor. If (xA)(x^{A}) are even/odd coordinates on \mathcal{M} then the fiber coordinates on 𝐓\mathbf{T}\mathcal{M} will be denoted by (dxA)(\mathrm{d}x^{A}) and the parity of dxA\mathrm{d}x^{A} is the parity of xAx^{A} plus one modulo two.

A canonical example of graded Lie supergroups of degree nn comes from the tangent lifts.

Example 9.

Higher order tangent bundles, not to be confused with the related iterated tangent bundles, are the natural geometric home of higher derivative Lagrangian mechanics [B]. A higher order (kk-order) tangent bundle Tk\mathrm{T}^{k}\mathcal{M} of a supermanifold \mathcal{M} can be defined in a natural way. By applying the functor Tk\mathrm{T}^{k} to the Lie supergroup 𝒢\mathcal{G} and the structure morphisms of 𝒢\mathcal{G} we get a graded Lie supergroup of type (Δ,χ)(\Delta,\chi), where Δ={0,1,,k}\Delta=\{0,1,\ldots,k\} and χ(1)=0¯\chi(1)=\bar{0}.

Another interesting example of \mathbb{Z}-gradation on a Lie supergroup was noticed by V. Kac:

Example 10.

Let V=V0¯V1¯V=V_{\bar{0}}\oplus V_{\bar{1}} be a vector superspace equipped with a non-degenerate even symmetric bilinear form QQ, i.e. Q(x,y)=(1)|x||y|Q(y,x)Q(x,y)=(-1)^{|x||y|}Q(y,x) for homogeneous x,yVx,y\in V and Q:V×V𝕂Q:V\times V\to\mathbb{K} is an even map. It follows that V0¯V1¯V_{\bar{0}}\perp V_{\bar{1}} and QQ restricted to V0¯V_{\bar{0}} (resp. V1¯V_{\bar{1}}) is symmetric (resp. skew-symmetric). Thus the dimension of V1¯V_{\bar{1}} is even and one can find lagrangian subspaces L,LV1¯L,L^{\prime}\subset V_{\bar{1}} such that V1¯=LLV_{\bar{1}}=L\oplus L^{\prime}, where LL^{\prime} is naturally identified with the dual to LL and Q|V1¯Q|_{V_{\bar{1}}} has the following natural form with respect to this decomposition:

Q((l1,l1),(l2,l2))=l1(l2)l2(l1),Q((l_{1},l_{1}^{\prime}),(l_{2},l_{2}^{\prime}))=l_{1}^{\prime}(l_{2})-l_{2}^{\prime}(l_{1}),

where liLl_{i}\in L and liLl^{\prime}_{i}\in L^{\prime}. The orthosymplectic Lie superalgebra 𝔬𝔰𝔭(V,Q)\mathfrak{osp}(V,Q) is spanned by homogeneous endomorphism T:VVT:V\to V such that

Q(Tx,y)+(1)|T||x|Q(x,Ty)=0Q(Tx,y)+(-1)^{|T||x|}Q(x,Ty)=0

for any homogeneous x,yVx,y\in V. It follows that 𝔬𝔰𝔭(V,Q)0¯=𝔰𝔬(V0¯)𝔰𝔭(V1¯)\mathfrak{osp}(V,Q)_{\bar{0}}=\mathfrak{so}(V_{\bar{0}})\oplus\mathfrak{sp}(V_{\bar{1}}) and 𝔬𝔰𝔭(V,Q)1¯=V0¯V1¯\mathfrak{osp}(V,Q)_{\bar{1}}=V_{\bar{0}}\otimes V_{\bar{1}}. There are well known isomorphisms 𝔰𝔬(W)=WW\mathfrak{so}(W)=W\wedge W and 𝔰𝔭(W)=S2W\mathfrak{sp}(W)=S^{2}W. Assigning the weight 11 to LL and the weight 1-1 to LL^{\prime} one find a canonical \mathbb{Z}-grading on 𝔬𝔰𝔭(V)\mathfrak{osp}(V) with non-zero parts in degrees ±2\pm 2, ±1\pm 1 and 0:

𝔬𝔰𝔭(V)=S2(L)(V0¯L)(LLV0¯V0¯)(V0¯L)S2L.\mathfrak{osp}(V)=S^{2}(L^{\prime})\oplus(V_{\bar{0}}\otimes L^{\prime})\oplus(L^{\prime}\otimes L\oplus\mathrm{V}_{\bar{0}}\wedge V_{\bar{0}})\oplus(V_{\bar{0}}\otimes L)\oplus S^{2}L.

Thus the corresponding Lie supergroup, the orthosymplectic Lie supergroup, is also naturally \mathbb{Z}-graded.

2.3.2. Super Harish-Chandra pairs

To study a Lie supergroup one uses the theory of super Harish-Chandra pairs, see [Ber] and also [BCC, CCF, Vi3].

Definition 11.

A super Harish-Chandra pair (G,𝔤)(G,\mathfrak{g}) consists of a Lie superalgebra 𝔤=𝔤0¯𝔤1¯\mathfrak{g}=\mathfrak{g}_{\bar{0}}\oplus\mathfrak{g}_{\bar{1}}, a Lie group GG such that 𝔤0¯=Lie(G)\mathfrak{g}_{\bar{0}}=\mathrm{Lie}(G), and an action α\alpha of GG on 𝔤\mathfrak{g} by authomorphisms such that

  • for any gGg\in G the action α(g)\alpha(g) restricted to 𝔤0¯\mathfrak{g}_{\bar{0}} coincides with the adjoint action Ad(g)Ad(g);

  • the differential (dα)e:𝔤0TeAut(𝔤)=Der(𝔤)(d\alpha)_{e}:\mathfrak{g}_{0}\to\mathrm{T}_{e}\operatorname{Aut}(\mathfrak{g})=\operatorname{Der}(\mathfrak{g}) at the identity eGe\in G coincides with the adjoint representation ad\operatorname{ad} of 𝔤0¯\mathfrak{g}_{\bar{0}} in 𝔤\mathfrak{g}.

Super Harish-Chandra pairs form a category. The following theorem was proved in [Vi3] for complex-analytic Lie supergroups.

Theorem 12.

The category of complex-analytic Lie supergroups is equivalent to the category of complex-analytic super Harish-Chandra pairs.

A similar result in the real case was obtained in [Kos], the algebraic case was treated by several authors, see for example [Gav, Mas, MS] and references therein. Theorem 12 implies that any Lie supergroup is globally split, that is, its structure sheaf is isomorphic to a wedge product of a certain trivial vector bundle. More explanation of this fact will be given later.

The notion of super Harish-Chandra pair has an obvious AA-graded generalization, where AA is an abelian group. Namely, we assume that 𝔤\mathfrak{g} is an AA-graded Lie superalgebra, the Lie algebra of GG is 𝔤0\mathfrak{g}_{0}, where 𝔤0\mathfrak{g}_{0} is the part of 𝔤\mathfrak{g} in degree 0, α(g):𝔤𝔤\alpha(g):\mathfrak{g}\to\mathfrak{g}, gGg\in G, is an automorphism of 𝔤\mathfrak{g}, in particular it preserves the AA-gradation and deα(X)=[X,]𝔤\mathrm{d}_{e}\alpha(X)=[X,\cdot]_{\mathfrak{g}} for any X𝔤0X\in\mathfrak{g}_{0}. In [JKPS, Theorem 5.6] it was noticed that an analogue of Theorem 12 holds true for graded Lie supergroups of degree nn. We will also work with super Harish-Chandra pairs of type Δ\Delta. A super Harish-Chandra pairs of type Δ\Delta is a r\mathbb{Z}^{r}-graded super Harish-Chandra pair (G,𝔤)(G,\mathfrak{g}), where 𝔤\mathfrak{g} is a r\mathbb{Z}^{r}-graded Lie superalgebra of type Δ\Delta. Recall that in this case r\mathbb{Z}^{r} is generated by the set Δ\Delta, with additional agreement about parities of generators of Δ\Delta, see Definition 3. In this case repeating argument [JKPS, Theorem 5.6], we get an analogue of Theorem 12 as well.

Example 13.

Let V=k=0nVkV=\oplus_{k=0}^{n}V_{k} be a non-negatively \mathbb{Z}-graded vector space with dimV<\dim V<\infty. Denote by Endq(V)\operatorname{End}_{q}(V), where qZq\in Z, the space of endomorphisms T:VVT:V\to V increasing the degree by qq, i.e. T:VVT:V\to V such that T|Vi:ViVi+qT|_{V_{i}}:V_{i}\to V_{i+q}. Thus End¯(V)=q=nnEndq(V)\underline{\operatorname{End}}(V)=\oplus_{q=-n}^{n}\operatorname{End}_{q}(V) is a \mathbb{Z}-graded vector space. It will be considered as a \mathbb{Z}-graded Lie superalgebra with the bracket given by the graded commutator. The corresponding \mathbb{Z}-graded Lie supergroup GL(V)\mathrm{GL}(V) has the body G0=×i=0nGL(Vi)G_{0}=\times_{i=0}^{n}GL(V_{i}) and its sheaf of functions is given by the super symmetric algebra over F=0ijnViVjF=\oplus_{0\leq i\neq j\leq n}V_{i}\otimes V_{j}^{*}, as ViVj=Hom(Vi,Vj)V_{i}\otimes V_{j}^{*}=\operatorname{Hom}(V_{i},V_{j})^{*}, i.e. 𝒪(U)=𝒞(U)SymF\mathcal{O}(U)=\mathcal{C}^{\infty}(U)\otimes\operatorname{Sym}F. The gradation in FF is obvious, |ViVj|=ij|V_{i}\otimes V_{j}^{*}|=i-j and the group multiplication preserve this gradation. The graded Harish-Chandra pair corresponding to GL(V)GL(V) is (G0,End¯(V))(G_{0},\underline{\operatorname{End}}(V)) equipped with the adjoint action of G0G_{0} on End¯(V)\underline{\operatorname{End}}(V). The language of Harish-Chandra pairs allows us to study some infinite dimensional generalizations of Lie supergroups in a convienient way. We simply drop the assumption that 𝔤\mathfrak{g} has finite dimension. For example, if V=k=0VkV=\oplus_{k=0}^{\infty}V_{k} is a \mathbb{Z}-graded vector space such that dimVk<\dim V_{k}<\infty, but the dimension of VV can be infinite, then GL(V)GL(V) can be understood as a pair consisting of the direct sum of GL(Vi)GL(V_{i}), G0=i=0GL(Vi)G_{0}=\oplus_{i=0}^{\infty}GL(V_{i}) and the natural action of G0G_{0} on End¯(V)=i,j0Hom(Vi,Vj)\underline{\operatorname{End}}(V)=\oplus_{i,j\geq 0}\operatorname{Hom}(V_{i},V_{j}).

2.3.3. Construction of the Lie supergroup corresponding to a super Harish-Chandra pair

Let us remind a reader how to construct a Lie supergroup 𝒢\mathcal{G} (or a graded Lie supergroup of type Δ\Delta) using a given super (or of type Δ\Delta) Harish-Chandra pair (G,𝔤)(G,\mathfrak{g}). Denote by 𝒰(𝔥)\mathcal{U}(\mathfrak{h}) the universal enveloping algebra of Lie superalgebra 𝔥\mathfrak{h}. We need to define a structure sheaf 𝒪\mathcal{O} of 𝒢\mathcal{G}. In the super and graded cases we, respectively, put

(5) 𝒪=Hom𝒰(𝔤0¯)(𝒰(𝔤),G),𝒪=Hom𝒰(𝔤0)(𝒰(𝔤),G).\begin{split}\mathcal{O}=\mathrm{Hom}_{\mathcal{U}(\mathfrak{g}_{\bar{0}})}(\mathcal{U}(\mathfrak{g}),\mathcal{F}_{G}),\quad\mathcal{O}=\mathrm{Hom}_{\mathcal{U}(\mathfrak{g}_{0})}(\mathcal{U}(\mathfrak{g}),\mathcal{F}_{G}).\end{split}

Here G\mathcal{F}_{G} is the sheaf of (holomorphic) functions on GG. Using the Hopf algebra structure on 𝒰(𝔤)\mathcal{U}(\mathfrak{g}) we can define explicitly the multiplication morphism μ\mu, the inversion morphism κ\kappa and the identity ε\varepsilon, see for instance [Vi3]. Indeed, assume that a super or graded of type Δ\Delta Harish-Chandra pair (G,𝔤)(G,\mathfrak{g}) is given. Let us define the supergroup structure of the corresponding Lie supergroup or graded Lie supergroup 𝒢\mathcal{G}. Let XY𝒰(𝔤𝔤)𝒰(𝔤)𝒰(𝔤)X\cdot Y\in\mathcal{U}(\mathfrak{g}\oplus\mathfrak{g})\simeq\mathcal{U}(\mathfrak{g})\otimes\mathcal{U}(\mathfrak{g}), where XX is from the first copy of 𝒰(𝔤)\mathcal{U}(\mathfrak{g}) and YY from the second one, f𝒪f\in\mathcal{O}, see (5), and g,hGg,\,h\in G. The following formulas define a multiplication morphism, an inverse morphism and an identity morphism respectively:

(6) μ(f)(XY)(g,h)=f(α(h1)(X)Y)(gh);κ(f)(X)(g1)=f(α(g1)(S(X)))(g);ε(f)=f(1)(e).\begin{split}\mu^{*}(f)(X\cdot Y)(g,h)&=f(\alpha(h^{-1})(X)\cdot Y)(gh);\\ \kappa^{*}(f)(X)(g^{-1})&=f(\alpha(g^{-1})(S(X)))(g);\\ \varepsilon^{*}(f)&=f(1)(e).\end{split}

Here SS is the antipode map in 𝒰(𝔤)\mathcal{U}(\mathfrak{g}) considered as a Hopf algebra and α\alpha is as in the definition of a super Harish-Chandra pair.

2.4. (Skew-symmetric) double vector bundles

A double vector bundle (DVB , in short) is a graded manifold of type Δ={0,α,β,α+β}\Delta=\{0,\alpha,\beta,\alpha+\beta\}. In this subsection we assume that the weights α\alpha and β\beta are even, however later on we shall drop this assumption. Geometrically we can see a double vector bundle as a quadruple (D;A,B;M)(D;A,B;M) with the following diagram of morphisms

(7) DABM,\leavevmode\hbox to60.57pt{\vbox to46.41pt{\pgfpicture\makeatletter\hbox{\hskip 30.28537pt\lower-23.2526pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{}{}{}{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{\offinterlineskip{}{}{{{}}{{}}{{}}{{}}}{{{}}}{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-30.28537pt}{-23.15276pt}\pgfsys@invoke{ }\hbox{\vbox{\halign{\pgf@matrix@init@row\pgf@matrix@step@column{\pgf@matrix@startcell#\pgf@matrix@endcell}&#\pgf@matrix@padding&&\pgf@matrix@step@column{\pgf@matrix@startcell#\pgf@matrix@endcell}&#\pgf@matrix@padding\cr\hfil\hskip 8.58401pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-4.27847pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${D}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}}}&\hskip 8.58401pt\hfil&\hfil\hskip 32.05551pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-3.75pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${A}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}&\hskip 8.05554pt\hfil\cr\vskip 18.00005pt\cr\hfil\hskip 8.34894pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-4.0434pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${B}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}&\hskip 8.34894pt\hfil&\hfil\hskip 33.70134pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-5.39583pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${M}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}&\hskip 9.70137pt\hfil\cr}}}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}}{{{{}}}{{}}{{}}{{}}{{}}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} { {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{{ {\pgfsys@beginscope \pgfsys@setdash{}{0.0pt}\pgfsys@roundcap\pgfsys@roundjoin{} {}{}{} {}{}{} \pgfsys@moveto{-2.07988pt}{2.39986pt}\pgfsys@curveto{-1.69989pt}{0.95992pt}{-0.85313pt}{0.27998pt}{0.0pt}{0.0pt}\pgfsys@curveto{-0.85313pt}{-0.27998pt}{-1.69989pt}{-0.95992pt}{-2.07988pt}{-2.39986pt}\pgfsys@stroke\pgfsys@endscope}} }{}{}{{}}{}{}{{}}\pgfsys@moveto{-12.91734pt}{11.50003pt}\pgfsys@lineto{11.9285pt}{11.50003pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{12.12848pt}{11.50003pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{-21.70135pt}{5.14032pt}\pgfsys@lineto{-21.70135pt}{-12.05977pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{0.0}{-1.0}{1.0}{0.0}{-21.70135pt}{-12.25975pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{20.584pt}{5.14032pt}\pgfsys@lineto{20.584pt}{-12.05977pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{0.0}{-1.0}{1.0}{0.0}{20.584pt}{-12.25975pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{-13.15242pt}{-20.65276pt}\pgfsys@lineto{10.28267pt}{-20.65276pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{10.48265pt}{-20.65276pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}},

where all maps are bundle projections and there are imposed some natural compatibility conditions, see [P, M].

In particular, DD has two vector bundle structures, one over the manifold AA and the second over BB. Moreover, AA, BB are vector bundles over MM called the side bundles of DD. The compatibility condition can be easily expressed using Euler vector fields: [𝔼A,𝔼B]=0[\mathbbm{E}_{A},\mathbbm{E}_{B}]=0 [GR]. Essentially, a DVB can be consider as a manifold equipped with two Euler vector fields, D=(D;𝔼A,𝔼B)D=(D;\mathbbm{E}_{A},\mathbbm{E}_{B}), describing the vector bundle structure on the legs DAD\to A, DBD\to B of DD. This lead to many simplifications in the theory of DVBs.

The core CC of DVB is defined as intersection of the kernels of two bundle projections DAD\to A and DBD\to B. It is a vector bundle over MM with the Euler vector field defined by the restriction, 𝔼A|C=𝔼B|C\mathbbm{E}_{A}|_{C}=\mathbbm{E}_{B}|_{C}.

The flip of a DVB is obtained by interchanging the legs of DD, ie. the flip of D=(D;𝔼A,𝔼B)D=(D;\mathbbm{E}_{A},\mathbbm{E}_{B}) is (D;𝔼B,𝔼A)(D;\mathbbm{E}_{B},\mathbbm{E}_{A}).

Symmetric and skew-symmetric DVBs are DVBs (D;A,B;M)(D;A,B;M) with an involution σ:DD\sigma:D\to D satisfying a certain addition condition. They were introduced in [BGR] in order to recognize the image of the linearization functor which associates with a graded manifold of type Δ={0,α,2α}\Delta=\{0,\alpha,2\alpha\} a DVB. Under this association graded manifolds of type Δ\Delta with even α\alpha (i.e., purely even graded bundles of degree 2) are in one-to-one correspondence with symetric DVBs, while graded manifolds of type Δ\Delta with odd α\alpha (i.e., NN-manifolds of degree 2) — with skew-symmetric DVBs. The results are extended to any order. For the purposes of this manuscript it is enough to recall a definition of skew-symmetric DVBs.

Definition 14.

A skew-symmetric DVB is a pair (D,σ)(D,\sigma) consisting of a DVB DD and an involution σ:DD\sigma:D\to D (i.e. σσ=id\sigma\circ\sigma=\mathrm{id}) such that

  1. (i)

    σ\sigma exchanges the legs of DD, i.e. σ\sigma is a DVB morphism from DD to the flip of DD,

  2. (ii)

    the restriction of σ\sigma to the core is minus the identity.

A morphism between skew-symmetric DVB s is assumed to intertwines with their involutions.

It follows immediately that the side bundles of a skew-symmetric DVB are isomorphic and that any skew-symmetric DVB admits an atlas with graded coordinates (xa;yαi,Yβj;zα+βμ)(x^{a};y^{i}_{\alpha},Y^{j}_{\beta};z^{\mu}_{\alpha+\beta}) such that

  • the transition functions for yiy^{i} and YiY^{i} are the same and

    zijμ=Qνμzν+12QijμyiYjz^{\mu}_{ij}=Q^{\mu}_{\nu}z^{\nu}+\frac{1}{2}Q^{\mu}_{ij}y^{i}Y^{j}

    with Qijμ=QjiμQ^{\mu}_{ij}=-Q^{\mu}_{ji},

  • the involution σ\sigma has a form: σ(yi)=Yi\sigma^{*}(y^{i})=Y^{i} (hence σ(Yi)=yi\sigma^{*}(Y^{i})=y^{i}) and σ(zμ)=zμ+σijμ(x)yiYj\sigma^{*}(z^{\mu})=-z^{\mu}+\sigma^{\mu}_{ij}(x)y^{i}Y^{j} with σijμ=σjiμ\sigma^{\mu}_{ij}=\sigma^{\mu}_{ji}.

To any double vector bundle DD (with the core CC and side bundles AA, BB) we can assign a short exact sequence

(8) 0CD^AB0,0\to C\to\widehat{D}\to A\otimes B\to 0,

which is obtained by dualizing the short exact sequence

0kerprC𝒜α+β(D)prCC00\to\ker\mathrm{pr}_{C}\to\mathcal{A}^{\alpha+\beta}(D)\xrightarrow{\mathrm{pr}_{C}}C^{*}\to 0

where 𝒜α+β(D)\mathcal{A}^{\alpha+\beta}(D) is the (M)\mathcal{F}(M)-module of homogeneaus functions on DD of weight α+β\alpha+\beta and prC\mathrm{pr}_{C} denotes the restriction of a function ff to the core CC, ie. prC(f)=f|C\mathrm{pr}_{C}(f)=f|_{C}. Locally, 𝒜α+β(D)\mathcal{A}^{\alpha+\beta}(D) is generated by the functions yαiy_{\alpha}^{i}, YβjY^{j}_{\beta} and zα+βμz^{\mu}_{\alpha+\beta}. In other words, the dual to D^\widehat{D} is the vector bundle whose space of sections is 𝒜α+β(D)\mathcal{A}^{\alpha+\beta}(D).

If (D,σ)(D,\sigma) is skew-symmetric then we can decompose

𝒜α+β(D)=𝒜+α+β(D)𝒜α+β(D),\mathcal{A}^{\alpha+\beta}(D)=\mathcal{A}^{\alpha+\beta}_{+}(D)\oplus\mathcal{A}^{\alpha+\beta}_{-}(D),

where 𝒜εα+β(D)\mathcal{A}^{\alpha+\beta}_{\varepsilon}(D) consists of those f𝒜α+β(D)f\in\mathcal{A}^{\alpha+\beta}(D) such that σ(f)=εf\sigma^{*}(f)=\varepsilon f and ε\varepsilon is plus or minus.

Theorem 15.

Let (D,σ)(D,\sigma) be a skew-symmetric DVB with side bundle EME\to M. Then 𝒜+α+β(D)\mathcal{A}^{\alpha+\beta}_{+}(D) coincides with the space of sections of S2ES^{2}E^{*} while 𝒜α+β(D)\mathcal{A}^{\alpha+\beta}_{-}(D) gives rise to a short exact sequence

(9) 0Γ(2E)𝒜α+β(D)Γ(C)0.0\to\Gamma(\bigwedge^{2}E^{*})\to\mathcal{A}^{\alpha+\beta}_{-}(D)\to\Gamma(C^{*})\to 0.

The skew-symmetric DVB (D,σ)(D,\sigma) can be reconstructed completely from the above sequence.

Proof.

Let (xa,yi,Yj,zμ)(x^{a},y^{i},Y^{j},z^{\mu}) be adopted coordinates for a skew-symmetric DVB (D,σ)(D,\sigma). In these coordinates the decomposition of 𝒜α+β(D)\mathcal{A}^{\alpha+\beta}(D) is clear: 𝒜+α+β(D)\mathcal{A}^{\alpha+\beta}_{+}(D) is locally generated by functions yiYj+yjYiy^{i}Y^{j}+y^{j}Y^{i} and 𝒜α+β(D)\mathcal{A}^{\alpha+\beta}_{-}(D) by zμz^{\mu} and yiYjyjYiy^{i}Y^{j}-y^{j}Y^{i}. The projection on Γ(C)\Gamma(C^{*}) is the same, it is given as the restriction to the core bundle of DD. ∎

2.5. Inverse limit of supermanifolds and Lie supergroups

We shall work with some types of infinite dimensional supermanifolds and Lie supergroups. There is no need to present a general theory as all examples we shall work with have a form of an inverse limit of supermanifolds (even with the same body 0\mathcal{M}_{0}):

1pr122pr233\mathcal{M}_{1}\xleftarrow{\mathrm{pr}^{2}_{1}}\mathcal{M}_{2}\xleftarrow{\mathrm{pr}^{3}_{2}}\mathcal{M}_{3}\xleftarrow{}\ldots

where i=(0,𝒪i)\mathcal{M}_{i}=(\mathcal{M}_{0},\mathcal{O}_{\mathcal{M}_{i}}) and prkk+1:k+1k\mathrm{pr}^{k+1}_{k}:\mathcal{M}_{k+1}\rightarrow\mathcal{M}_{k} is the projection of the graded manifold k+1\mathcal{M}_{k+1} of degree k+1k+1 to the graded manifold k\mathcal{M}_{k} of degree kk, see Remark 6. We define

=(0,𝒪), where 𝒪=k=1𝒪k,\mathcal{M}_{\infty}=(\mathcal{M}_{0},\mathcal{O}_{\mathcal{M}_{\infty}}),\quad\text{ where }\mathcal{O}_{\mathcal{M}_{\infty}}=\bigcup_{k=1}^{\infty}\mathcal{O}_{\mathcal{M}_{k}},

i.e. f𝒪f\in\mathcal{O}_{\mathcal{M}_{\infty}} if and only if there exist kk such that f𝒪kf\in\mathcal{O}_{\mathcal{M}_{k}}. A morphism f:𝒩f:\mathcal{M}_{\infty}\to\mathcal{N}_{\infty} is a family of morphisms f=(fk:k𝒩ak)f=(f_{k}:\mathcal{M}_{k}\to\mathcal{N}_{a_{k}}) where (ak)(a_{k}) is a non-decreasing sequence of positive integers and the family (fk)(f_{k}) is compatible with the projections:

k+1\textstyle{\mathcal{M}_{k+1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}fk+1\scriptstyle{f_{k+1}}prkk+1\scriptstyle{\mathrm{pr}^{k+1}_{k}}𝒩ak+1\textstyle{\mathcal{N}_{a_{k+1}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}prakak+1\scriptstyle{\mathrm{pr}^{a_{k+1}}_{a_{k}}}k\textstyle{\mathcal{M}_{k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}fk\scriptstyle{f_{k}}𝒩ak\textstyle{\mathcal{N}_{a_{k}}}

where prakak+1\mathrm{pr}^{a_{k+1}}_{a_{k}} is the composition of projections prii+1\mathrm{pr}^{i+1}_{i} for i=ak+11i=a_{k+1}-1 to aka_{k}.

The inverse limit of Lie supergroups has a Lie supergroup structure. We additionally have to assume that the projections prii+1\mathrm{pr}^{i+1}_{i} are Lie supergroup homomorphisms. Also the inverse limit of Lie (super)algebras has a Lie (super)algebra structure thus a Lie functor makes sense for inverse limit of Lie supergroups. For example a \mathbb{Z}-graded Lie superalgebra 𝔤=k=0𝔤k\mathfrak{g}=\bigoplus_{k=0}^{\infty}\mathfrak{g}_{k} is an inverse limit of Lie algebras 𝔤/Ik\mathfrak{g}/I_{k} where Ik=𝔤k+1𝔤k+2I_{k}=\mathfrak{g}_{k+1}\oplus\mathfrak{g}_{k+2}\oplus\ldots. There is one-to-one correspondence of the inverse limit of graded Lie supergroups and the inverse limit of their graded Harish-Chandra pairs.

3. The Donagi–Witten construction, supermanifolds, double vector bundles and graded manifolds of degree 22

In [DW2, Section 2.1] Donagi and Witten gave a description of the first obstruction class ω=ω2\omega=\omega_{2} via differential operators. In this section we remind the definition of ω2\omega_{2} and the Donagi–Witten construction. Further we give an interpretation of the Donagi–Witten construction using the language of double vector bundles and graded manifolds of degree 22. At the end using differential forms instead of differential operators we simplify this construction, which allows us to find its higher analogue.

3.1. First obstruction class ω2\omega_{2}

Let us describe the first obstruction class to splitting a supermanifold using results [Ber, Gr, Oni, R]. We follow the exposition of [Oni]. First of all consider a split supermanifold =(0,𝒪)\mathcal{M}=(\mathcal{M}_{0},\mathcal{O}), where 𝒪=\mathcal{O}=\bigwedge\mathcal{E}^{*} and \mathcal{E} is the sheaf of sections of 𝔼[1¯]\mathbb{E}[\bar{1}]. Denote by 𝒟er𝒪\mathcal{D}er\mathcal{O} the sheaf of vector fields on \mathcal{M} and by 𝒟er\mathcal{D}er\mathcal{F} the sheaf of vector fields on the underlying space 0\mathcal{M}_{0}. The sheaf 𝒟er𝒪=p1𝒟erp𝒪\mathcal{D}er\mathcal{O}=\bigoplus_{p\geq-1}\mathcal{D}er_{p}\mathcal{O} is naturally \mathbb{Z}-graded since \bigwedge\mathcal{E}^{*} is \mathbb{Z}-graded. We have the following exact sequence

(10) 03𝒟er2𝒪2𝒟er0,0\to\bigwedge^{3}\mathcal{E}^{*}\otimes\mathcal{E}\to\mathcal{D}er_{2}\mathcal{O}\to\bigwedge^{2}\mathcal{E}^{*}\otimes\mathcal{D}er\mathcal{F}\to 0,

see [Oni, Formula (5)].

According Green [Gr] we can describe all non-split supermanifolds 𝒩\mathcal{N} such that gr(𝒩)\mathrm{gr}(\mathcal{N})\simeq\mathcal{M} using the sheaf of automorphisms 𝒜ut𝒪\mathcal{A}ut\mathcal{O} of 𝒪\mathcal{O}. More precisely consider the following subsheaf of 𝒜ut𝒪\mathcal{A}ut\mathcal{O}:

𝒜ut(2)𝒪={a𝒜ut𝒪|a(u)u𝒥2for anyu𝒪},\mathcal{A}ut_{(2)}\mathcal{O}=\{a\in\mathcal{A}ut\mathcal{O}\,\,|\,\,a(u)-u\in\mathcal{J}^{2}\,\,\text{for any}\,\,u\in\mathcal{O}\},

see also [Oni, Formula (17)]. Recall that 𝒥\mathcal{J} is the sheaf of ideals generated by odd elements in 𝒪\mathcal{O}. Denote by Aut(𝔼)\mathrm{Aut}(\mathbb{E}^{*}) the group of automorthisms of 𝔼\mathbb{E}^{*}. There is a natural action of Aut(𝔼)\mathrm{Aut}(\mathbb{E}^{*}) on the sheaf 𝒜ut(2)𝒪\mathcal{A}ut_{(2)}\mathcal{O}, see [Oni, Section 1.4]. This action induces an action of Aut(𝔼)\mathrm{Aut}(\mathbb{E}^{*}) on the set H1(0,𝒜ut(2)𝒪)H^{1}(\mathcal{M}_{0},\mathcal{A}ut_{(2)}\mathcal{O}). By Green [Gr] points of the set of orbits H1(0,𝒜ut(2)𝒪)/Aut(𝔼)H^{1}(\mathcal{M}_{0},\mathcal{A}ut_{(2)}\mathcal{O})/\mathrm{Aut}(\mathbb{E}^{*}) are in one-to-one correspondence with isomorphism classes of supermanifolds 𝒩\mathcal{N} such that gr(𝒩)\mathrm{gr}(\mathcal{N})\simeq\mathcal{M}. More precisely to any supermanifold 𝒩\mathcal{N} such that gr(𝒩)\mathrm{gr}(\mathcal{N})\simeq\mathcal{M} we can assign a class γH1(0,𝒜ut(2)𝒪)\gamma\in H^{1}(\mathcal{M}_{0},\mathcal{A}ut_{(2)}\mathcal{O}). If γi\gamma_{i} is the class corresponding to a supermanifold 𝒩i\mathcal{N}_{i} such that gr(𝒩i)\mathrm{gr}(\mathcal{N}_{i})\simeq\mathcal{M}, where i=1,2i=1,2. Then 𝒩1𝒩2\mathcal{N}_{1}\simeq\mathcal{N}_{2} if and only if γ1\gamma_{1} and γ2\gamma_{2} are in the same orbit of Aut(𝔼)\mathrm{Aut}(\mathbb{E}^{*}).

In [R] the following map of sheaves was defined

(11) 𝒜ut(2)𝒪𝒟er2𝒪,\mathcal{A}ut_{(2)}\mathcal{O}\to\mathcal{D}er_{2}\mathcal{O},

see also [Oni, Formula (19)]. Combining the map (11) and the map 𝒟er2𝒪3𝒟er\mathcal{D}er_{2}\mathcal{O}\to\bigwedge^{3}\mathcal{E}^{*}\otimes\mathcal{D}er\mathcal{F} from (10) we get the following map of cohomology sets

H1(0,𝒜ut(2)𝒪)H1(0,2𝒟er)H^{1}(\mathcal{M}_{0},\mathcal{A}ut_{(2)}\mathcal{O})\to H^{1}(\mathcal{M}_{0},\bigwedge^{2}\mathcal{E}^{*}\otimes\mathcal{D}er\mathcal{F})

and the corresponding map of Aut(𝔼)\mathrm{Aut}(\mathbb{E}^{*})-orbits

H1(0,𝒜ut(2)𝒪)/Aut(𝔼)H1(0,2𝒟er)/Aut(𝔼).H^{1}(\mathcal{M}_{0},\mathcal{A}ut_{(2)}\mathcal{O})/\mathrm{Aut}(\mathbb{E}^{*})\to H^{1}(\mathcal{M}_{0},\bigwedge^{2}\mathcal{E}^{*}\otimes\mathcal{D}er\mathcal{F})/\mathrm{Aut}(\mathbb{E}^{*}).

If γH1(0,𝒜ut(2)𝒪)\gamma\in H^{1}(\mathcal{M}_{0},\mathcal{A}ut_{(2)}\mathcal{O}) corresponds to a non-split supermanifold 𝒩\mathcal{N}, the image of γ\gamma in H1(0,2𝒟er)H^{1}(\mathcal{M}_{0},\bigwedge^{2}\mathcal{E}^{*}\otimes\mathcal{D}er\mathcal{F}) is called the first obstruction class to splitting of 𝒩\mathcal{N} and following [DW2, Section 2.1] we denote this class by ω2\omega_{2}.

Consider the case when \mathcal{M} has odd dimension 22 in details. Since 3={0}\bigwedge^{3}\mathcal{E}^{*}=\{0\}, from (10) it follows that

𝒟er2𝒪2𝒟er.\mathcal{D}er_{2}\mathcal{O}\simeq\bigwedge^{2}\mathcal{E}^{*}\otimes\mathcal{D}er\mathcal{F}.

In this case the map (11) is an isomorphism. Therefore we have the following set bijection

H1(0,𝒜ut(2)𝒪)H1(0,2𝒟er)H^{1}(\mathcal{M}_{0},\mathcal{A}ut_{(2)}\mathcal{O})\simeq H^{1}(\mathcal{M}_{0},\bigwedge^{2}\mathcal{E}^{*}\otimes\mathcal{D}er\mathcal{F})

and the corresponding bijection of the sets of orbits. Now Green’s result [Gr] implies that ω2\omega_{2} is the only obstruction for a supermanifold to be split in this case. In other words a supermanifold 𝒩\mathcal{N} of odd dimension 22 such that gr(𝒩)\mathrm{gr}(\mathcal{N})\simeq\mathcal{M} is split if and only if ω2=0\omega_{2}=0. Note that the notion of a split and a projectable, see [DW1] for details, supermanifold coincide in this case.

3.2. The Donagi and Witten construction

In [DW2, Section 2.1] the first obstruction class ω2\omega_{2} to splitting a supermanifold =(0,𝒪)\mathcal{M}=(\mathcal{M}_{0},\mathcal{O}) was interpreted in terms of a certain sheaf of differential operators on \mathcal{M}. Namely, the obstruction class ω\omega defined above is the Atiyah class of the extension

0TMDω2E00\to\mathrm{T}M\to D_{\omega}\to\bigwedge^{2}E\to 0

where sections of DωD_{\omega} are identified with some factor of 𝒟2|0\mathcal{D}^{2}_{-}|_{\mathcal{M}_{0}} where the meaning of the latter is explained below.

Let us remind this construction using charts and local coordinates. Consider two charts 𝒰1\mathcal{U}_{1} and 𝒰2\mathcal{U}_{2} on \mathcal{M} with non-empty intersection and with local coordinates (xi,ξa)(x_{i},\xi_{a}) and (yj,ηb)(y_{j},\eta_{b}), where i,j=1,,ni,j=1,\ldots,n and a,b=1,,ma,b=1,\ldots,m. Let in 𝒰1𝒰2\mathcal{U}_{1}\cap\mathcal{U}_{2} we have the following transition functions

(12) yj=Fj+12Gja1a2ξa1ξa2+j=1,,n;ηb=Hbaξa+,b=1,,m,\begin{split}&y_{j}=F_{j}+\frac{1}{2}G_{j}^{a_{1}a_{2}}\xi_{a_{1}}\xi_{a_{2}}+\cdots\,\,\,j=1,\ldots,n;\\ &\eta_{b}=H_{b}^{a}\xi_{a}+\cdots,\,\,\,b=1,\ldots,m,\end{split}

where Fj,Gja1a2,HibF_{j},G^{a_{1}a_{2}}_{j},H_{i}^{b} are (holomorphic) functions depending only on even coordinates (xi)(x_{i}).

If D:𝒪𝒪D:\mathcal{O}\to\mathcal{O} is a differential operator let [D]:𝒪\left[{D}\right]:\mathcal{O}\to\mathcal{F} denotes the composition of DD with the projection 𝒪𝒪/𝒥=\mathcal{O}\to\mathcal{O}/\mathcal{J}=\mathcal{F}.

Following Donagi and Witten, see [DW2, Section 2.1], we define the sheaf 𝒟2|0\mathcal{D}^{2}_{-}|_{\mathcal{M}_{0}} on the underlying manifold 0\mathcal{M}_{0} as a sheaf locally generated over \mathcal{F} by

1,[xi],[ξj],[2ξaξb].\langle 1,\left[{\frac{\partial}{\partial x_{i}}}\right],\left[{\frac{\partial}{\partial\xi_{j}}}\right],\left[{\frac{\partial^{2}}{\partial\xi_{a}\partial\xi_{b}}}\right]\rangle.

In [DW2, Theorem 2.5] it was shown that this definition does not depend on local coordinates. Note that the operators ξa\frac{\partial}{\partial\xi_{a}} anticommute, ie. 2ξaξb=2ξbξa\frac{\partial^{2}}{\partial\xi_{a}\partial\xi_{b}}=-\frac{\partial^{2}}{\partial\xi_{b}\partial\xi_{a}}. We can get transition function for generators of 𝒟2|0\mathcal{D}^{2}_{-}|_{\mathcal{M}_{0}} writing down the transition function for xi\frac{\partial}{\partial x_{i}}, ξj\frac{\partial}{\partial\xi_{j}}, 2ξaξb\frac{\partial^{2}}{\partial\xi_{a}\partial\xi_{b}} and then factorizing them by 𝒥\mathcal{J}. Indeed,

xi=yjxiyj+ηαxiηα;ξk=yjξkyj+ηαξkηα.\displaystyle\frac{\partial}{\partial x_{i}}=\frac{\partial y_{j}}{\partial x_{i}}\frac{\partial}{\partial y_{j}}+\frac{\partial\eta_{\alpha}}{\partial x_{i}}\frac{\partial}{\partial\eta_{\alpha}};\quad\frac{\partial}{\partial\xi_{k}}=\frac{\partial y_{j}}{\partial\xi_{k}}\frac{\partial}{\partial y_{j}}+\frac{\partial\eta_{\alpha}}{\partial\xi_{k}}\frac{\partial}{\partial\eta_{\alpha}}.

Therefore modulo 𝒥\mathcal{J} we have:

[2ξa1ξa2]=(ηαξa12(yj)ξa1ξa2)red[yj](ηαξa1ηβξa2)red[2ηαηβ].\displaystyle\left[{\frac{\partial^{2}}{\partial\xi_{a_{1}}\partial\xi_{a_{2}}}}\right]=\Big{(}\frac{\partial\eta_{\alpha}}{\partial\xi_{a_{1}}}\frac{\partial^{2}(y_{j})}{\partial\xi_{a_{1}}\partial\xi_{a_{2}}}\Big{)}_{\mathrm{red}}\left[{\frac{\partial}{\partial y_{j}}}\right]-\Big{(}\frac{\partial\eta_{\alpha}}{\partial\xi_{a_{1}}}\frac{\partial\eta_{\beta}}{\partial\xi_{a_{2}}}\Big{)}_{\mathrm{red}}\left[{\frac{\partial^{2}}{\partial\eta_{\alpha}\partial\eta_{\beta}}}\right].

Compare with [DW2, Formula 2.13]. Using (12) we can write the transition functions for 𝒟2|0\mathcal{D}^{2}_{-}|_{\mathcal{M}_{0}} explicitly.

(13) xi=Fi1(yj);[xi]=Fjxi[yj];[ξa]=Hab[ηb];[2ξb1ξb2]=Gjb1b2[yj]Hαb1Hβb2[2ηαηβ].\begin{split}&x_{i}=F^{-1}_{i}(y_{j});\\ &\left[{\frac{\partial}{\partial x_{i}}}\right]=\frac{\partial F_{j}}{\partial x_{i}}\left[{\frac{\partial}{\partial y_{j}}}\right];\quad\left[{\frac{\partial}{\partial\xi_{a}}}\right]=H_{a}^{b}\left[{\frac{\partial}{\partial\eta_{b}}}\right];\\ &\left[{\frac{\partial^{2}}{\partial\xi_{b_{1}}\partial\xi_{b_{2}}}}\right]=-G_{j}^{b_{1}b_{2}}\left[{\frac{\partial}{\partial y_{j}}}\right]-H_{\alpha}^{b_{1}}H_{\beta}^{b_{2}}\left[{\frac{\partial^{2}}{\partial\eta_{\alpha}\partial\eta_{\beta}}}\right].\end{split}

Note that in (13) it is more correct to use the index redred for even coordinates, for example (xi)red(x_{i})_{\mathrm{red}}. However we omit red\mathrm{red} for simplicity of notations. We conclude this subsection with the following important remark.

Remark 16.

Comparing Formulas (12) and (13), we see that Formulas (13) contain the whole information about Formulas (12) modulo 𝒥3\mathcal{J}^{3}. In other words using Formulas (13) we can reconstruct Formulas (12) modulo 𝒥3\mathcal{J}^{3}.

One of purposes of this paper is to develop this observation. This leads to an idea to use the theory of nn-fold vector bundles and the theory of graded manifolds of degree nn to recover a supermanifolds of odd dimension greater than 22.

3.3. A geometric interpretation of the Donagi and Witten construction

In this subsection we give a description of the geometric object with transition functions (13). We use the theory of double vector bundles and graded manifolds of degree 22.

Lemma 17.

Let 𝔻\mathbb{D} be a vector bundle over MM characterized by the space of sections, Γ(𝔻)=𝒟2|0\mathrm{\Gamma}(\mathbb{D})=\mathcal{D}^{2}_{-}|_{\mathcal{M}_{0}}. Then the dual bundle 𝔻\mathbb{D}^{*} has transition functions of a skew-symmetric DVB . More precisely, there exist a skew-symmetric DVB (D,σ)(D,\sigma) such that the space 𝒜α+β(D)\mathcal{A}^{\alpha+\beta}_{-}(D) coincides with the space of sections of the vector bundle 𝔻\mathbb{D}^{*}.

Proof.

We simply transpose and reverse the formulas (13). ∎

The skew-symmetric DVB arising from the above lemma will be denoted by Vb2()\mathrm{Vb}_{2}(\mathcal{M}).

The Atiyah class of the exact sequence (8), that is the obstruction class of splitting of this sequence, is an element in H1(M;ABC)H^{1}(M;A^{*}\otimes B^{*}\otimes\ C). It will be also called the Atiyah class associated with the DVB DD and denoted by At(D)\mathrm{At}(D). Note that the Atiyah class associated with any dual (vertical or horizontal) of DD coincides with At(D)\mathrm{At}(D). Conversely if we have the sequence (8) we can reconstruct the double vector bundle DD.

Now let \mathcal{M} be a supermanifold with transition function (12) and let E=Vb1()E=\mathrm{Vb}_{1}(\mathcal{M}). The Atiyah sequence associated with D=Vb2()D=\mathrm{Vb}_{2}(\mathcal{M}) (which is a double vector bundle with the same side bundles EE and the core T0\mathrm{T}\mathcal{M}_{0}) is

0T0D^EE0.0\to\mathrm{T}\mathcal{M}_{0}\to\widehat{D}\to E\otimes E\to 0.

The obstruction for splitting of this exact sequence is

At(Vb2())H1(0,T(0)EE).\mathrm{At}(\mathrm{Vb}_{2}(\mathcal{M}))\in H^{1}(\mathcal{M}_{0},\mathrm{T}(\mathcal{M}_{0})\otimes E^{*}\otimes E^{*}).

Due to the decomposition EE=Sym2E2EE\otimes E=\operatorname{Sym}^{2}E\oplus\bigwedge^{2}E, the Atiyah class At(D)\mathrm{At}(D), where D=Vb2()D=\mathrm{Vb}_{2}(\mathcal{M}), decomposes to At+(D)+At(D)\mathrm{At}_{+}(D)+\mathrm{At}_{-}(D), where At+(D)H1(0,T(0)Sym2E)\mathrm{At}_{+}(D)\in H^{1}(\mathcal{M}_{0},\mathrm{T}(\mathcal{M}_{0})\otimes\operatorname{Sym}^{2}E^{*}) and At(D)H1(M;TM2E)\mathrm{At}_{-}(D)\in H^{1}(M;\mathrm{T}M\otimes\bigwedge^{2}E^{*}). We are only interested with At(D)\mathrm{At}_{-}(D) because our DVB Vb2()\mathrm{Vb}_{2}(\mathcal{M}) is a skew-symmetric DVB , hence At+(D)=0\mathrm{At}_{+}(D)=0.

The class At()\mathrm{At}_{-}(\mathcal{M}) is here the same as the Atiyah class of the sequence (9) associated with the skew-symmetric DVB (D,σ)(D,\sigma). It also coincides with the first obstruction class to splitting of the supermanifold \mathcal{M} as in [DW2, Section 2].

Later we will give a geometric interpretation of this fact.

3.4. A modification of the Donagi–Witten construction

In this subsection we suggest a different way how to obtain a double vector bundle with obstruction the class ω2\omega_{2}. First of all to write Formulas (13) we need the transition functions (12) and their inverse. To avoid this inconvenience we can use differential forms instead of differential operators.

Let \mathcal{M} be a supermanifold as above. Consider as above two charts 𝒰1\mathcal{U}_{1} and 𝒰2\mathcal{U}_{2} with non-empty intersection on \mathcal{M} with local coordinates (xi,ξa)(x_{i},\xi_{a}) and (yj,ηb)(y_{j},\eta_{b}) and transition functions (12). Further consider the antitangent bundles 𝐓()\mathbf{T}(\mathcal{M}) of \mathcal{M} and two charts 𝐓(𝒰1)\mathbf{T}(\mathcal{U}_{1}) and 𝐓(𝒰2)\mathbf{T}(\mathcal{U}_{2}) with standard coordinates (xi,ξa,dxi,dξa)(x_{i},\xi_{a},\mathrm{d}x_{i},\mathrm{d}\xi_{a}) and (yj,ηb,dyj,dηb)(y_{j},\eta_{b},\mathrm{d}y_{j},\mathrm{d}\eta_{b}), respectively. Thus xix_{i}, dξa\mathrm{d}\xi_{a} are even local coordinates in 𝐓(𝒰1)\mathbf{T}(\mathcal{U}_{1}), while ξa\xi_{a},dxi)\mathrm{d}x_{i}) are odd ones. In 𝐓(𝒰1)𝐓(𝒰2)\mathbf{T}(\mathcal{U}_{1})\cap\mathbf{T}(\mathcal{U}_{2}) we get the following transition functions.

yj=Fj+Gja1a2ξa1ξa2+,ηa=Habξb+;\displaystyle y_{j}=F_{j}+G_{j}^{a_{1}a_{2}}\xi_{a_{1}}\xi_{a_{2}}+\cdots,\quad\eta_{a}=H_{a}^{b}\xi_{b}+\cdots;
dyj=b=1n(Fj)xbdxb+b=1n(Gja1a2)xbdxbξ1ξ2+Gja1a2(d(ξa1)ξa2ξa1d(ξa2))+;\displaystyle\mathrm{d}y_{j}=\sum_{b=1}^{n}(F_{j})_{x_{b}}\mathrm{d}x_{b}+\sum_{b=1}^{n}(G_{j}^{a_{1}a_{2}})_{x_{b}}\mathrm{d}x_{b}\xi_{1}\xi_{2}+G_{j}^{a_{1}a_{2}}(\mathrm{d}(\xi_{a_{1}})\xi_{a_{2}}-\xi_{a_{1}}\mathrm{d}(\xi_{a_{2}}))+\cdots;
dηa=b=1n(Hai)xbdxbξj+Habdξb+.\displaystyle\mathrm{d}\eta_{a}=\sum_{b=1}^{n}(H_{a}^{i})_{x_{b}}\mathrm{d}x_{b}\xi_{j}+H_{a}^{b}\mathrm{d}\xi_{b}+\cdots.

Here we denoted by (Fj)xb(F_{j})_{x_{b}} the derivation of FjF_{j} by xbx_{b}. Now we apply the functor split gr\mathrm{gr} to 𝐓()\mathbf{T}(\mathcal{M}). In local coordinates this means that we factorize our transition functions by all terms that contain more than one odd variable. Then gr𝐓()\mathrm{gr}\mathbf{T}(\mathcal{M}) has the following transition functions

(14) yj=Fj(x);ηa=Habξb;dyj=b=1n(Fj)xbdxb+Gja1a2(d(ξa1)ξa2ξa1d(ξa2));dηa=Habdξb.\begin{split}&y_{j}=F_{j}(x);\\ &\eta_{a}=H_{a}^{b}\xi_{b};\\ &\mathrm{d}y_{j}=\sum_{b=1}^{n}(F_{j})_{x_{b}}\mathrm{d}x_{b}+G_{j}^{a_{1}a_{2}}(\mathrm{d}(\xi_{a_{1}})\xi_{a_{2}}-\xi_{a_{1}}\mathrm{d}(\xi_{a_{2}}));\\ &\mathrm{d}\eta_{a}=H_{a}^{b}\mathrm{d}\xi_{b}.\end{split}

If we compare Formulas (14) with [Vo, Section 2.2, Formulas (9)-(12)], we see that a manifold with such transition functions is a double vector bundle, which we denote by 𝔻\mathbb{D}. We can see gr𝐓()\mathrm{gr}\mathbf{T}(\mathcal{M}) as a graded manifold of type Δ={0,α,β,α+β}\Delta=\{0,\alpha,\beta,\alpha+\beta\}, where α\alpha is odd (the weight of ξa\xi_{a}) and β\beta is even (the weight of dξ\mathrm{d}\xi). Formulas (14) shows that all weights are well-defined. The double vector bundle

(15) 𝔻{\mathbb{D}^{\prime}}E[1¯]{E[\bar{1}]}E{E}0{\mathcal{M}_{0}}

is only slightly different from the double vector bundle Vb2()\mathrm{Vb}_{2}(\mathcal{M}). In our case one side bundle is pure even.

Remark 18.

By definition the composition of functors gr𝐓\mathrm{gr}\circ\mathbf{T} is a functor from the category of supermanifolds to the category of (split) supermanifolds. However above implies that we can see the image of gr𝐓\mathrm{gr}\circ\mathbf{T} as the category of double vector bundles (with some additional structure). For simplicity we will use the same notation gr𝐓\mathrm{gr}\circ\mathbf{T} meaning a functor from the category of supermanifolds to the category of double vector bundles. Similarly in next sections we will consider the functor gr𝐓(n)\mathrm{gr}\circ\mathbf{T}^{(n)} as a functor from the category of supermanifolds to the category of nn-fold vector bundles.

Now we can go further and give an interpretation of ω2\omega_{2} as the obstruction class of splitting of a graded manifold of degree 22. First of all let us change that parity of the side bundle 𝔼\mathbb{E}. (In other words we apply the functor parity change, see [Vo].) To do this we need to rewrite (14) in the following form

(16) yj=Fj;ηa=Habξb;dyj=b=1n(Fj)xbdxb+Gja1a2(ξa2d(ξa1)ξa1d(ξa2));dηa=Habdξb.\begin{split}&y_{j}=F_{j};\\ &\eta_{a}=H_{a}^{b}\xi_{b};\\ &\mathrm{d}y_{j}=\sum_{b=1}^{n}(F_{j})_{x_{b}}\mathrm{d}x_{b}+G_{j}^{a_{1}a_{2}}(\xi_{a_{2}}\mathrm{d}(\xi_{a_{1}})-\xi_{a_{1}}\mathrm{d}(\xi_{a_{2}}));\\ &\mathrm{d}\eta_{a}=H_{a}^{b}\mathrm{d}\xi_{b}.\end{split}

and to change the parity of the weight β\beta.

Now we use a result obtained by [JL, CM] for double vector bundles and by [BGR, Vi1] for nn-fold vector bundles. Later we will call this step ”to apply the functor inverse”. In more details in [JL, CM] a functor was constructed from the category of graded manifolds of degree 22 to the category of double vector bundles with some additional structures. In [BGR, Vi1] an analogue of this result was obtained for graded manifolds of degree nn and for graded manifolds of type Δ\Delta. (Note that in all these papers [JL, CM, BGR, Vi1] the categories of double vector bundles with additional structures are different. In this paper we follow approaches of [BGR, Vi1].) In [JL, CM] it was shown that this functor is an equivalence of the category of double vector bundles with some additional structures and the category of graded manifolds of degree 22. In [BGR, Vi1] it was shown that this functor is an equivalence of the category of nn-fold vector bundles with some additional structures and the category of graded manifolds of degree nn or of type Δ\Delta. The inversion of this functor, that is the functor from the category of nn-fold vector bundles with some additional structures to the category of graded manifolds of degree nn, we call the functor inverse. We will denote the functor inverse by ι\iota.

In terms of local coordinates ”to apply the functor inverse” means that we identify ξa\xi_{a} with dξa\mathrm{d}\xi_{a} in (16). We get

(17) yj=Fj;θa=Haiζi;tj=b=1n(Fj)xbzb+2Gja1a2ζa1ζa2.\begin{split}&y_{j}=F_{j};\\ &\theta_{a}=H_{a}^{i}\zeta_{i};\\ &t_{j}=\sum_{b=1}^{n}(F_{j})_{x_{b}}z_{b}+2G_{j}^{a_{1}a_{2}}\zeta_{a_{1}}\zeta_{a_{2}}.\end{split}

We obtained transition functions of a graded manifold 𝒩\mathcal{N} of degree 22, which we will denote also by F2()\mathrm{F}_{2}(\mathcal{M}). We assign the following weights to our local coordinates: xix_{i} (weight 0); ζj\zeta_{j} (weight α\alpha); zsz_{s} (weight 2α2\alpha). In other words, the transition functions (17) defines a graded manifold of type Δ={0,α,2α}\Delta=\{0,\alpha,2\alpha\}, where α\alpha is odd. Note that we can remove the coefficient 22 in (17). Indeed, it is enough to replace ζi\zeta_{i} by 12ζi\frac{1}{\sqrt{2}}\zeta^{\prime}_{i} in any chart.

Remark 19.

Let us give another explanation of the procedure ”to apply the functor inverse”. Consider two graded domain 𝒱1\mathcal{V}_{1} and 𝒱2\mathcal{V}_{2} with graded coordinates (xi,ζj,zi)(x_{i},\zeta_{j},z_{i}) and (yi,θj,ti)(y_{i},\theta_{j},t_{i}), respectively, and with weights as above. Define transition functions 𝒱1𝒱2\mathcal{V}_{1}\to\mathcal{V}_{2} by (17). (Let us omit the coefficient 22.) Now we apply the tangent functor 𝐓\mathbf{T} to 𝒱1\mathcal{V}_{1} and 𝒱2\mathcal{V}_{2} and to the morphism (17) and factorize the result by the sheaf of ideals locally generated by dxi\mathrm{d}x_{i} (or by dyi\mathrm{d}y_{i}). We get

yj=Fj;\displaystyle y_{j}=F_{j};
θa=Haiζi;dθa=Haidζi;\displaystyle\theta_{a}=H_{a}^{i}\zeta_{i};\quad\mathrm{d}\theta_{a}=H_{a}^{i}\mathrm{d}\zeta_{i};
dtj=b=1n(Fj)xbdzb+Gja1a2(ζa2d(ζa1)ζa1d(ζa2)).\displaystyle\mathrm{d}t_{j}=\sum_{b=1}^{n}(F_{j})_{x_{b}}\mathrm{d}z_{b}+G_{j}^{a_{1}a_{2}}(\zeta_{a_{2}}\mathrm{d}(\zeta_{a_{1}})-\zeta_{a_{1}}\mathrm{d}(\zeta_{a_{2}})).

We obtain Formulas (16) up to appropriate change of variables. The inversion of this procedure is called ”to apply the functor inverse”.

The structure sheaf 𝒪𝒩=q0(𝒪𝒩)q\mathcal{O}_{\mathcal{N}}=\bigoplus_{q\geq 0}(\mathcal{O}_{\mathcal{N}})_{q} of 𝒩\mathcal{N} is \mathbb{Z}-graded. Clearly (𝒪𝒩)1(𝒪𝒩)1(𝒪𝒩)2(\mathcal{O}_{\mathcal{N}})_{1}\cdot(\mathcal{O}_{\mathcal{N}})_{1}\subset(\mathcal{O}_{\mathcal{N}})_{2}. Therefore we can assign to 𝒩\mathcal{N} the following exact sequence

0(𝒪𝒩)1(𝒪𝒩)1(𝒪𝒩)2(𝒪𝒩)2/(𝒪𝒩)1(𝒪𝒩)10.\displaystyle 0\to(\mathcal{O}_{\mathcal{N}})_{1}\cdot(\mathcal{O}_{\mathcal{N}})_{1}\to(\mathcal{O}_{\mathcal{N}})_{2}\to(\mathcal{O}_{\mathcal{N}})_{2}/(\mathcal{O}_{\mathcal{N}})_{1}\cdot(\mathcal{O}_{\mathcal{N}})_{1}\to 0.

The Atiyah class of this sequence is represented by a cocycle (GJa1a2)(G_{J}^{a_{1}a_{2}}) and coincides with the obstruction class of the double vector bundle (16). It is called the obstruction class of the graded manifold F2()\mathrm{F}_{2}(\mathcal{M}) of degree 22.

Let us summarize our results in the following theorem.

Theorem 20 (Main theorem about the first obstruction class).

The first obstruction class w2w_{2} for a supermanifold \mathcal{M} coincides with the obstruction class At(Vb2())\mathrm{At}_{-}(\mathrm{Vb}_{2}(\mathcal{M})) of the skew-symetric double vector bundle Vb2()\mathrm{Vb}_{2}(\mathcal{M}) and with the obstruction class of the graded manifold F2()\mathrm{F}_{2}(\mathcal{M}) of degree 22.

Note that the map F2()\mathcal{M}\mapsto\mathrm{F}_{2}(\mathcal{M}) is a functor from the category of supermanifolds of odd dimension 22 to the category of graded manifolds of degree 22. Indeed, F2\mathrm{F}_{2} is a composition of four functors: 𝐓\mathbf{T}, gr\mathrm{gr}, π\pi and ι\iota. This functor defines an equivalence of the category of supermanifolds of odd dimension 22 and the category of graded manifolds of degree 22 with the following additional condition. If 𝒩=(𝒩0,𝒪𝒩)\mathcal{N}=(\mathcal{N}_{0},\mathcal{O}_{\mathcal{N}}) is a graded manifold of degree 22, we additionally assume that the locally free sheaf :=(𝒪𝒩)2/(𝒪𝒩)1(𝒪𝒩)1\mathcal{E}:=(\mathcal{O}_{\mathcal{N}})_{2}/(\mathcal{O}_{\mathcal{N}})_{1}\cdot(\mathcal{O}_{\mathcal{N}})_{1} is isomorphic to the sheaf of sections of T[1¯](𝒩0)\mathrm{T}[\bar{1}](\mathcal{N}_{0}). In this case we can choose local coordinates in the form (17). Using (17) we can write transition function for a supermanifold of odd dimension 22

yj=Fj+Gja1a2ζa1ζa2;\displaystyle y_{j}=F_{j}+G_{j}^{a_{1}a_{2}}\zeta_{a_{1}}\zeta_{a_{2}};
θa=Haiζi.\displaystyle\theta_{a}=H_{a}^{i}\zeta_{i}.

These transition functions satisfy the cocycle condition. Indeed, consider three charts 𝒰1,𝒰2\mathcal{U}_{1},\mathcal{U}_{2} and 𝒰3\mathcal{U}_{3} with 𝒰1𝒰2𝒰3\mathcal{U}_{1}\cap\mathcal{U}_{2}\cap\mathcal{U}_{3}\neq\emptyset. Denote by TijT_{ij} the transition function Tij:𝒰i𝒰jT_{ij}:\mathcal{U}_{i}\to\mathcal{U}_{j}. Consider the following composition of maps

T31T23T12=:R.T_{31}\circ T_{23}\circ T_{12}=:R.

Since F2\mathrm{F}_{2} is a functor, we get

F2(T31)F2(T23)F2(T12)=F2(R).\mathrm{F}_{2}(T_{31})\circ\mathrm{F}_{2}(T_{23})\circ\mathrm{F}_{2}(T_{12})=\mathrm{F}_{2}(R).

The composition F2(T31)F2(T23)F2(T12)\mathrm{F}_{2}(T_{31})\circ\mathrm{F}_{2}(T_{23})\circ\mathrm{F}_{2}(T_{12}) is equal to idid, since the cocycle condition for the graded manifold 𝒩\mathcal{N} holds true. Therefore F2(R)=id\mathrm{F}_{2}(R)=id. The morphism R:𝒰1𝒰1R:\mathcal{U}_{1}\to\mathcal{U}_{1} is completely defined by its image F2(R)=id\mathrm{F}_{2}(R)=id. Therefore, R=idR=id.

Denote the category of supermanifold of odd dimension 22 by Smf2\mathrm{Smf_{2}} and the category of graded manifolds of degree 22 and of odd dimension 22 with the additional condition for (𝒪𝒩)2/(𝒪𝒩)1(𝒪𝒩)1(\mathcal{O}_{\mathcal{N}})_{2}/(\mathcal{O}_{\mathcal{N}})_{1}\cdot(\mathcal{O}_{\mathcal{N}})_{1} as above, by GrMan2T\mathrm{GrMan_{2}T}. Now we can summarize our results in the following theorem.

Theorem 21.

The categories Smf2\mathrm{Smf_{2}} and GrMan2T\mathrm{GrMan_{2}T} are equivalent.

In [RV] we generalize this theorem to the category of supermanifolds of any odd dimension.

Remark 22.

If our supermanifold \mathcal{M} has the odd dimension m>2m>2, we still can repeat the procedure above. Therefore the functor F2\mathrm{F}_{2} is a functor from the category of supermanifolds to the category of graded manifold of degree 22. However for any dimension the functor F2\mathrm{F}_{2} is not an embedding. In this case from a graded manifold of degree 22 we can recover a supermanifold module 𝒥3\mathcal{J}^{3}.

Summing up, in this section we showed that the first obstruction class to splitting of the supermanifold in the sense of [DW2, Section 2] coincides with the obstruction class of the splitting of the double vector bundle gr𝐓()\mathrm{gr}\mathbf{T}(\mathcal{M}) and with the obstruction class of the splitting of the graded manifold F2()\mathrm{F}_{2}(\mathcal{M}). A general splitting theory for supermanifolds of odd dimension mm based on splitting of the corresponding mm-fold vector bundles and graded manifolds of degree mm will be developed in our oncoming paper.

4. A generalization of the Donagi–Witten construction

In this section we give a construction of functors Fn\mathrm{F}_{n}, where n=2,3,,n=2,3,\ldots,\infty, from the category of supermanifolds to the category of graded manifolds of degree nn. Due to the size of this paper and technical difficulty of the proof of the main result, the existence of the functor inverse ι\iota for any supermanifold of the form gr𝐓()\mathrm{gr}\circ\mathbf{T}(\mathcal{M}), where \mathcal{M} is a supermanifold, we leave details of this proof to Part IIII of this paper. Here we give only the idea of the proof.

The functor Fn\mathrm{F}_{n} is again a composition of four functors: the (n1)(n-1)-times (or infinity many times for n=n=\infty) iterated antitangent functor 𝐓(n1):=𝐓𝐓\mathbf{T}^{(n-1)}:=\mathbf{T}\circ\cdots\circ\mathbf{T}, the functor split gr\mathrm{gr}, the functor parity change π\pi and the functor inverse ι\iota. If \mathcal{M} is a supermanifold, then 𝐓(n1)()\mathbf{T}^{(n-1)}(\mathcal{M}), gr𝐓(n1)()\mathrm{gr}\circ\mathbf{T}^{(n-1)}(\mathcal{M}) are already defined. Denote by Δ\Delta the maximal multiplicity free weight system generated by an odd weight α\alpha and by even weights β1,,βn1\beta_{1},\ldots,\beta_{n-1}, see Definition 3. We need the following propositions.

Proposition 23.

The supermanifold gr𝐓(n1)()\mathrm{gr}\circ\mathbf{T}^{(n-1)}(\mathcal{M}) is an nn-fold vector bundle of type Δ\Delta, where Δ\Delta is the maximal multiplicity free system generated by an odd weight α\alpha and by even weights β1,,βn1\beta_{1},\ldots,\beta_{n-1}.

Proof.

Let di\mathrm{d}_{i}, where i=1,,n1i=1,\cdots,n-1, be the iterated de Rham differentials. They are vector fields in the structure sheaf of 𝐓(n1)()\mathbf{T}^{(n-1)}(\mathcal{M}). Denote by dI\mathrm{d}_{I}, where I=(i1,,ik)I=(i_{1},\ldots,i_{k}), the composition di1dik\mathrm{d}_{i_{1}}\circ\ldots\circ\mathrm{d}_{i_{k}}. Let \mathcal{I} be the ideal generated by odd elements in 𝒪gr𝐓(n1)()\mathcal{O}_{\mathrm{gr}\circ\mathbf{T}^{(n-1)}(\mathcal{M})} and let us choose a chart 𝒰\mathcal{U} on \mathcal{M} with local coordinates (xa,ξb)(x_{a},\xi_{b}). Then the corresponding standard local coordinates in the chart 𝐓(n1)(𝒰)\mathbf{T}^{(n-1)}(\mathcal{U}) are (dI(xa),dJ(ξb))C(I),C(J)n1(\mathrm{d}_{I}(x_{a}),\mathrm{d}_{J}(\xi_{b}))_{C(I),C(J)\leq n-1}, where C(I)C(I) is the cardinality of II. Hence the corresponding local coordinates in the chart gr𝐓(n1)(𝒰)\mathrm{gr}\circ\mathbf{T}^{(n-1)}(\mathcal{U}) are (dI(xa)+2,dJ(ξb)+2)C(I),C(J)n1(\mathrm{d}_{I}(x_{a})+\mathcal{I}^{2},\mathrm{d}_{J}(\xi_{b})+\mathcal{I}^{2})_{C(I),C(J)\leq n-1}. We assign the weight α+βii++βik\alpha+\beta_{i_{i}}+\cdots+\beta_{i_{k}} to dI(Z)+2\mathrm{d}_{I}(Z)+\mathcal{I}^{2}, where Z{xa,ξb}Z\in\{x_{a},\xi_{b}\}, if dI(Z)\mathrm{d}_{I}(Z) is odd and the weight βii++βik\beta_{i_{i}}+\cdots+\beta_{i_{k}} to dI(Z)+2\mathrm{d}_{I}(Z)+\mathcal{I}^{2} if dI(Z)\mathrm{d}_{I}(Z) is even.

It remains to prove that transition functions in gr𝐓(n1)()\mathrm{gr}\circ\mathbf{T}^{(n-1)}(\mathcal{M}) preserve our weights. Let 𝒰\mathcal{U}^{\prime} be another chart on \mathcal{M} with coordinates (xa,ξb)(x^{\prime}_{a},\xi^{\prime}_{b}) such that 𝒰𝒰\mathcal{U}\cap\mathcal{U}^{\prime}\neq\emptyset. Let

xa=C(K)=2kFK(x)ξKx_{a}^{\prime}=\sum_{C(K)=2k}F_{K}(x)\xi_{K}

be the expression of xax_{a}^{\prime} in coordinates of 𝒰\mathcal{U}. Then dI(xa)+2\mathrm{d}_{I}(x^{\prime}_{a})+\mathcal{I}^{2} in coordinates of gr𝐓(n1)(𝒰)\mathrm{gr}\circ\mathbf{T}^{(n-1)}(\mathcal{U}) is a sum of monomials in the following form.

(18) HI,K(x)dI1(xa1)dIp(xap)dIp+1(ξb1)dIq(ξbq)+2,H_{I,K}(x)\mathrm{d}_{I_{1}}(x_{a_{1}})\cdots\mathrm{d}_{I_{p}}(x_{a_{p}})\mathrm{d}_{I_{p+1}}(\xi_{b_{1}})\cdots\mathrm{d}_{I_{q}}(\xi_{b_{q}})+\mathcal{I}^{2},

where HI,K(x)H_{I,K}(x) is a certain derivative of FK(x)F_{K}(x), which has weight 0, and (I1,,Iq)(I_{1},\ldots,I_{q}) is a decomposition of the sequence II into qq parts. We see that the weights of dI(xa)+2\mathrm{d}_{I}(x^{\prime}_{a})+\mathcal{I}^{2} and of (18) coincide. Similarly for dI(ξb)+2\mathrm{d}_{I}(\xi^{\prime}_{b})+\mathcal{I}^{2}. This completes the proof. ∎

Proposition 24.

The functor gr𝐓(n1)\mathrm{gr}\circ\mathbf{T}^{(n-1)} is an embedding of the category of supermanifolds of odd dimension nn into the category of nn-fold vector bundles of type Δ\Delta.

Proof.

Let 𝒩\mathcal{N} be an nn-fold vector bundle. Let us show that if a preimage \mathcal{M} of 𝒩\mathcal{N} exists, it is unique. The idea of the proof of similar to the idea of the proof of Theorem 21. Assume that there exists another supermanifold \mathcal{M}^{\prime} with gr𝐓(n1)()=gr𝐓(n1)()\mathrm{gr}\circ\mathbf{T}^{(n-1)}(\mathcal{M})=\mathrm{gr}\circ\mathbf{T}^{(n-1)}(\mathcal{M}^{\prime}). By construction we have 0=0=𝒩0\mathcal{M}_{0}=\mathcal{M}^{\prime}_{0}=\mathcal{N}_{0}. Let us choose an atlas {𝒰i}\{\mathcal{U}_{i}\} on \mathcal{M} and an atlas {𝒰i}\{\mathcal{U}^{\prime}_{i}\} on \mathcal{M}^{\prime} such that (𝒰i)0=(𝒰i)0(\mathcal{U}_{i})_{0}=(\mathcal{U}^{\prime}_{i})_{0}. Clearly such atlases there exist, since it is sufficient to choose any atlas on 0\mathcal{M}_{0} of Stein domains. Now consider two charts 𝒰1\mathcal{U}^{\prime}_{1} and 𝒰2\mathcal{U}^{\prime}_{2} with 𝒰1𝒰2\mathcal{U}^{\prime}_{1}\cap\mathcal{U}^{\prime}_{2}\neq\emptyset on \mathcal{M}^{\prime}. Denote by TT^{\prime} the transition function T:𝒰1𝒰2T^{\prime}:\mathcal{U}^{\prime}_{1}\to\mathcal{U}^{\prime}_{2} and by TT the transition function T:𝒰1𝒰2T:\mathcal{U}_{1}\to\mathcal{U}_{2}. By our assumption gr𝐓(n1)(T)=gr𝐓(n1)(T)\mathrm{gr}\circ\mathbf{T}^{(n-1)}(T)=\mathrm{gr}\circ\mathbf{T}^{(n-1)}(T^{\prime}). But this implies that T=TT=T^{\prime}. The proof is complete. ∎

Remark 25.

The question when the functor gr𝐓(n1)\mathrm{gr}\circ\mathbf{T}^{(n-1)} possesses a preimage is treated in [RV].

Now we can use results of [BGR] and [Vi1] to define the functor inverse. Note that these two results lead to two different approaches to the problem. Let us start with a description of results [BGR] or [Vi1]. In [BGR] a functor Gn\mathrm{G}_{n} was constructed from the category of graded manifolds of degree nn to the category of nn-fold symmetric vector bundles. In [Vi1] a functor Vn\mathrm{V}_{n} was constructed from the category of graded manifolds of degree nn to the category of nn-fold vector bundles with n1n-1 odd commutative homological vector fields. In both cases it was proven that these functors are equivalence of categories. In Section 3 we saw that due to the fact that the first obstruction class ω2\omega_{2} is an element in

H1(0,T[1¯](0)2𝔼)H^{1}(\mathcal{M}_{0},\mathrm{T}[\bar{1}](\mathcal{M}_{0})\otimes\bigwedge^{2}\mathbb{E}^{*})

we can construct a graded manifold of degree 22 with transition functions (17) from a double vector bundle given by (16). In terms of [BGR] this means that the double vector bundle with the obstruction class ω2\omega_{2} is symmetric and πgr𝐓()\pi\circ\mathrm{gr}\circ\mathbf{T}(\mathcal{M}) is in the image of the functor G2\mathrm{G}_{2} constructed in [BGR]. Our functor inverse ι\iota is the inversion of the equivalence G2\mathrm{G}_{2}. More generally we have

Theorem 26.

For a supermanifold \mathcal{M} of odd dimension nn the image πgr𝐓(n1)()\pi\circ\mathrm{gr}\circ\mathbf{T}^{(n-1)}(\mathcal{M}) is a symmetric nn-fold vector bundle. Therefore there exists a graded manifold 𝒩\mathcal{N} of degree nn such that Gn(𝒩)πgr𝐓(n1)()\mathrm{G}_{n}(\mathcal{N})\simeq\pi\circ\mathrm{gr}\circ\mathbf{T}^{(n-1)}(\mathcal{M}). In other words the functor inverse ι=Gn1\iota=\mathrm{G}_{n}^{-1} is defined on the image of the functor πgr𝐓(n1)\pi\circ\mathrm{gr}\circ\mathbf{T}^{(n-1)}.

Let us describe another approach based on results of [Vi1]. The graded supermanifold gr𝐓()\mathrm{gr}\circ\mathbf{T}(\mathcal{M}) possesses a natural odd homological vector field. Indeed, let dR\mathrm{d}_{R} be the exterior derivative (de Rham defferential) of the supermanifold \mathcal{M}. Clearly dR\mathrm{d}_{R} is an odd homological vector field in the structure sheaf 𝒪𝐓()\mathcal{O}_{\mathbf{T}(\mathcal{M})} of 𝐓()\mathbf{T}(\mathcal{M}). Denote by \mathcal{I} the sheaf of ideals locally generated by all odd variables of 𝒪𝐓()\mathcal{O}_{\mathbf{T}(\mathcal{M})}. Then we have

dR()𝒪𝐓()anddR(2).\mathrm{d}_{R}(\mathcal{I})\subset\mathcal{O}_{\mathbf{T}(\mathcal{M})}\quad\text{and}\quad\mathrm{d}_{R}(\mathcal{I}^{2})\subset\mathcal{I}.

Therefore we have an induced operator in the structure sheaf of gr𝐓()\mathrm{gr}\circ\mathbf{T}(\mathcal{M})

𝒪gr𝐓()=gr𝒪𝐓()=p0p/p+1.\mathcal{O}_{\mathrm{gr}\circ\mathbf{T}(\mathcal{M})}=\mathrm{gr}\mathcal{O}_{\mathbf{T}(\mathcal{M})}=\bigoplus_{p\geq 0}\mathcal{I}^{p}/\mathcal{I}^{p+1}.

Clearly this induced operator is odd and homological. Hence the double vector bundle gr𝐓()\mathrm{gr}\circ\mathbf{T}(\mathcal{M}) is a double vector bundle with a homological vector field. Therefore by [Vi1] there exists a graded manifold V21(gr𝐓())\mathrm{V}_{2}^{-1}(\mathrm{gr}\circ\mathbf{T}(\mathcal{M})) of degree 22. This procedute can be generalized. Now we can formulate a general result.

Theorem 27.

For a supermanifold \mathcal{M} of odd dimension nn the image gr𝐓(n1)()\mathrm{gr}\circ\mathbf{T}^{(n-1)}(\mathcal{M}) is an nn-fold vector bundle with n1n-1 commuting odd homological vector fields. Therefore there exists a graded manifold 𝒩\mathcal{N} of degree nn such that Vn(𝒩)gr𝐓(n1)()\mathrm{V}_{n}(\mathcal{N})\simeq\mathrm{gr}\circ\mathbf{T}^{(n-1)}(\mathcal{M}). In other words the functor ι=Vn1\iota^{\prime}=\mathrm{V}_{n}^{-1} is defined on the image of gr𝐓(n1)\mathrm{gr}\circ\mathbf{T}^{(n-1)}.

The idea of the proof is simple: the (n1)(n-1)-times iterated tangent functor leads to n1n-1 commuting de Rham differentials, which induce (n1)(n-1) commuting odd homological vector fields on gr𝐓(n1)()\mathrm{gr}\circ\mathbf{T}^{(n-1)}(\mathcal{M}). We leave details of the proof to Part II of this paper. Note that in this case the functor parity change π\pi is included in the functor Vn\mathrm{V}_{n}, therefore we do not need to apply it explicitly. We conclude this section with the following proposition

Proposition 28.

The functors ιπgr𝐓(n1)\iota\circ\pi\circ\mathrm{gr}\circ\mathbf{T}^{(n-1)} and ιgr𝐓(n1)\iota^{\prime}\circ\mathrm{gr}\circ\mathbf{T}^{(n-1)} are embeddings of the category of supermanifolds of odd dimension nn to the category of graded manifolds of degree nn.

Proof.

The proposition follows from Proposition 24, from the fact that π\pi is an equivalence of categories of nn-fold vector bundles and from Theorems 26 and 27. ∎

The inverse limit of functors Fn\mathrm{F}_{n} is denoted by F\mathrm{F}_{\infty}. That is if \mathcal{M} is a supermanifold, we put F():=limFn()\mathrm{F}_{\infty}(\mathcal{M}):=\varprojlim\mathrm{F}_{n}(\mathcal{M}) and the same for morphisms.

5. Donagi–Witten functor for Lie superalgebras and
Lie supergroups

In this section, we adapt the results of Section 3 to the case of Lie superagebras and Lie supergroups. We show that there is an analogue of the functor F2\mathrm{F}_{2} in the category of Lie superalgebras. We denote by sLieAlg\mathrm{sLieAlg} the category of Lie superalgebras, by grLieAlgn\mathrm{grLieAlg}_{n} the category of \mathbb{Z}-graded Lie superalgebras of type {0,1,,n}\{0,1,\ldots,n\} with |1|=1¯|1|=\bar{1} and by grLieAlg\mathrm{grLieAlg}_{\infty} the category of non-negatively \mathbb{Z}-graded Lie superalgebras.

5.1. Donagi and Witten construction for Lie superalgebras

In this section we will construct a functor F2:sLieAlggrLieAlg2\mathrm{F}^{\prime}_{2}:\mathrm{sLieAlg}\to\mathrm{grLieAlg}_{2}. In next sections we will use this construction to obtain a functor from the category of Lie superalgebras sLieAlg\mathrm{sLieAlg} to the category grLieAlg\mathrm{grLieAlg}_{\infty}. As in 3 the functor F2\mathrm{F}^{\prime}_{2} is a composition of four functors, which we denote by 𝐓\mathbf{T}^{\prime} (tangent), gr\mathrm{gr}^{\prime} (split), π\pi^{\prime} (parity change) and ι\iota^{\prime} (inverse). (To distinguish the category of Lie superalgebras in the case we will use superscript. That is 𝐓\mathbf{T}^{\prime} instead of 𝐓\mathbf{T} and so on.) Let 𝔤=𝔤0¯𝔤1¯\mathfrak{g}=\mathfrak{g}_{\bar{0}}\oplus\mathfrak{g}_{\bar{1}} be a Lie superalgebra. Recall that to define a Lie superalgebra we need to define a Lie algebra 𝔤0¯\mathfrak{g}_{\bar{0}}, a 𝔤0¯\mathfrak{g}_{\bar{0}}-module 𝔤1¯\mathfrak{g}_{\bar{1}} and a symmetric 𝔤0¯\mathfrak{g}_{\bar{0}}-module map [,]:𝔤1¯𝔤1¯𝔤0¯[\cdot,\cdot]:\mathfrak{g}_{\bar{1}}\otimes\mathfrak{g}_{\bar{1}}\to\mathfrak{g}_{\bar{0}} such that [[Y,Y],Y]=0[[Y,Y],Y]=0, for any Y𝔤1¯Y\in\mathfrak{g}_{\bar{1}}.

Tangent functor 𝐓\mathbf{T}^{\prime}. Since a Lie superalgebra 𝔤=𝔤0¯𝔤1¯\mathfrak{g}=\mathfrak{g}_{\bar{0}}\oplus\mathfrak{g}_{\bar{1}} is a vector superspace, its antitangent bundle is the following linear superspace

𝐓(𝔤)=𝔤d(𝔤).\mathbf{T}^{\prime}(\mathfrak{g})=\mathfrak{g}\oplus\mathrm{d}(\mathfrak{g}).

Here d𝔤\mathrm{d}\mathfrak{g} denote a copy of 𝔤\mathfrak{g} with reversed parity. In more details, V=𝔤d(𝔤)V=\mathfrak{g}\oplus\mathrm{d}(\mathfrak{g}) is a linear superspace with the underlying space V0¯=𝔤0¯d𝔤1¯V_{\bar{0}}=\mathfrak{g}_{\bar{0}}\oplus\mathrm{d}\mathfrak{g}_{\bar{1}} and with the structure sheaf V0¯S(𝔤1¯d𝔤0¯)\mathcal{F}_{V_{\bar{0}}}\otimes S^{*}(\mathfrak{g}^{*}_{\bar{1}}\oplus\mathrm{d}\mathfrak{g}^{*}_{\bar{0}}). If X𝔤X\in\mathfrak{g}, sometimes we will denote by d(X)\mathrm{d}(X) the corresponding element in d(𝔤)\mathrm{d}(\mathfrak{g}). Further the Lie superalgebra structure on 𝐓(𝔤)\mathbf{T}^{\prime}(\mathfrak{g}) is defined as follows. The Lie bracket on the vector subspece 𝔤{0}\mathfrak{g}\oplus\{0\} coincides with the Lie bracket of 𝔤\mathfrak{g}, further d(𝔤)\mathrm{d}(\mathfrak{g}) is a 𝔤\mathfrak{g}-module, where the action of 𝔤\mathfrak{g} coincide with the adjoint action up to sign

[X,dY]:=(1)|X|d([X,Y]).[X,\mathrm{d}Y]:=(-1)^{|X|}\mathrm{d}([X,Y]).

and the product [d(𝔤),d(𝔤)]=0[\mathrm{d}(\mathfrak{g}),\mathrm{d}(\mathfrak{g})]=0 is trivial.

Remark 29.

The Lie superalgebra 𝐓(𝔤)\mathbf{T}^{\prime}(\mathfrak{g}) sometimes is called in the literature a Takiff superalgebra, named after the author of [T]. This Lie superalgebra is a Lie superalgebra in the form 𝔤𝕂𝕂[τ]\mathfrak{g}\otimes_{\mathbb{K}}\mathbb{K}[\tau], where τ\tau is an odd element and 𝕂[τ]\mathbb{K}[\tau] are all polynomials in τ\tau. Summing up, in our paper 𝐓\mathbf{T}^{\prime} is derived from both “Takiff” and “antitangent”.

Functor split gr\mathrm{gr}^{\prime}. The functor gr\mathrm{gr}^{\prime} is defined as follows. For a Lie superalgebra 𝔤\mathfrak{g} we put gr(𝔤):=𝔤\mathrm{gr}^{\prime}(\mathfrak{g}):=\mathfrak{g}^{\prime}, where 𝔤\mathfrak{g}^{\prime} is obtained from the Lie superalgebra 𝔤\mathfrak{g} putting [𝔤1¯,𝔤1¯]=0.[\mathfrak{g}^{\prime}_{\bar{1}},\mathfrak{g}^{\prime}_{\bar{1}}]=0. In [Vi2, Theorem 3], see also Therem 36 below, it was shown that Lie(gr𝒢)=gr(Lie𝒢)\operatorname{Lie}(\mathrm{gr}\mathcal{G})=\mathrm{gr}^{\prime}(\operatorname{Lie}\mathcal{G}) for any Lie supergroup 𝒢\mathcal{G}. In other words 𝔤\mathfrak{g}^{\prime} is the Lie superalgebra of gr(𝒢)\mathrm{gr}(\mathcal{G}), where 𝒢\mathcal{G} is a Lie supergroup with the Lie superalgebra 𝔤\mathfrak{g}. Clearly gr\mathrm{gr}^{\prime} is defined on morphisms as well. Let us compute the composition of the functors gr𝐓\mathrm{gr}^{\prime}\circ\mathbf{T}^{\prime}.

Lemma 30.

Let 𝔤\mathfrak{g} be a Lie superalgebra. Then the Lie superalgebra gr(𝐓(𝔤))\mathrm{gr}^{\prime}(\mathbf{T}^{\prime}(\mathfrak{g})) is equal to 𝐓(𝔤)\mathbf{T}^{\prime}(\mathfrak{g}) as 𝔤0¯\mathfrak{g}_{\bar{0}}-modules. And the following holds true for the Lie superalgebra multiplication

[𝔤1¯,𝔤1¯]=0,[𝔤1¯,d(𝔤0¯)]=0,[d(𝔤),d(𝔤)]=0.\displaystyle[\mathfrak{g}_{\bar{1}},\mathfrak{g}_{\bar{1}}]=0,\quad[\mathfrak{g}_{\bar{1}},\mathrm{d}(\mathfrak{g}_{\bar{0}})]=0,\quad[\mathrm{d}(\mathfrak{g}),\mathrm{d}(\mathfrak{g})]=0.

and [Y1,d(Y2)]=d([Y1,Y2])[Y_{1},\mathrm{d}(Y_{2})]=-\mathrm{d}([Y_{1},Y_{2}]) for Y1,Y2𝔤1¯Y_{1},Y_{2}\in\mathfrak{g}_{\bar{1}}.

Remark 31.

Note that the Lie algebra gr𝐓(𝔤)\mathrm{gr}\mathbf{T}(\mathfrak{g}) keeps all the information on the bracket on 𝔤\mathfrak{g}. We can see the Lie superalgebra 𝔞=gr(𝐓(𝔤))\mathfrak{a}=\mathrm{gr}^{\prime}(\mathbf{T}^{\prime}(\mathfrak{g})) as a ×\mathbb{Z}\times\mathbb{Z}-graded Lie superalgebra of type {0,α,β,α+β}\{0,\alpha,\beta,\alpha+\beta\}, where α\alpha is odd and β\beta is even. Indeed, we put

𝔞0:=𝔤0¯,𝔞α:=𝔤1¯,𝔞β:=d𝔤1¯,𝔞α+β:=d𝔤0¯.\mathfrak{a}_{0}:=\mathfrak{g}_{\bar{0}},\quad\mathfrak{a}_{\alpha}:=\mathfrak{g}_{\bar{1}},\quad\mathfrak{a}_{\beta}:=\mathrm{d}\mathfrak{g}_{\bar{1}},\quad\mathfrak{a}_{\alpha+\beta}:=\mathrm{d}\mathfrak{g}_{\bar{0}}.

Lemma 30 implies that the multiplication in 𝔞\mathfrak{a} is ×\mathbb{Z}\times\mathbb{Z}-graded.

Functor parity change π\pi^{\prime}. In previous sections we reminded how to define the functor π\pi for double vector bundles. In this section we show that on the parity reversed double vector bundle πgr𝐓(𝒢)\pi\circ\mathrm{gr}\circ\mathbf{T}(\mathcal{G}), where 𝒢\mathcal{G} is a Lie supergroup, a Lie supergroup structure can be defined. We start with the case of Lie superalgebras.

We define the functor π\pi^{\prime} on the image of the composition of functors gr𝐓\mathrm{gr}^{\prime}\circ\mathbf{T}^{\prime} as follows. The Lie superalgebra gr(𝐓(𝔤))\mathrm{gr}^{\prime}(\mathbf{T}^{\prime}(\mathfrak{g})) possesses a parity reversion of the weight β\beta. Indeed, by definition the Lie superalgebra π(gr(𝐓(𝔤)))\pi^{\prime}(\mathrm{gr}^{\prime}(\mathbf{T}^{\prime}(\mathfrak{g}))) is equal to gr(𝐓(𝔤))\mathrm{gr}^{\prime}(\mathbf{T}^{\prime}(\mathfrak{g})) as 𝔤0¯\mathfrak{g}_{\bar{0}}-modules, but now we assume that elements XX and d(X)\mathrm{d}(X) have the same parities. In other words we assume that d\mathrm{d} is even. More precisely, let 𝔞=gr(𝐓(𝔤))\mathfrak{a}=\mathrm{gr}^{\prime}(\mathbf{T}^{\prime}(\mathfrak{g})). Denote by 𝔥\mathfrak{h} the vector superspace π(gr(𝐓(𝔤)))\pi^{\prime}(\mathrm{gr}^{\prime}(\mathbf{T}^{\prime}(\mathfrak{g}))). We have

𝔥=δΔ𝔥δ,\mathfrak{h}=\bigoplus_{\delta\in\Delta}\mathfrak{h}_{\delta},

where Δ={0,α,β,α+β}\Delta=\{0,\alpha,\beta,\alpha+\beta\} with |α|=|β|=1¯|\alpha|=|\beta|=\bar{1}. Let us define a Lie superalgebra structure on 𝔥\mathfrak{h} of type Δ\Delta. We denote by [,]𝔞[\,,]_{\mathfrak{a}} the Lie bracket on 𝔞\mathfrak{a} and by [,]𝔥[\,,]_{\mathfrak{h}} the Lie bracket on 𝔥\mathfrak{h}.

Proposition 32.

The Lie superalgebra structure on 𝔥\mathfrak{h} is defines by the following data.

  • We set [X,Y]𝔥=[X,Y]𝔞[X,Y]_{\mathfrak{h}}=[X,Y]_{\mathfrak{a}} for all homogeneous X,YX,Y, except for the case X𝔥αX\in\mathfrak{h}_{\alpha} and Y𝔥βY\in\mathfrak{h}_{\beta}.

  • For X𝔥αX\in\mathfrak{h}_{\alpha} and Y𝔥βY\in\mathfrak{h}_{\beta} we set [X,Y]𝔥:=[X,Y]𝔞=[Y,X]𝔞=:[Y,X]𝔥[X,Y]_{\mathfrak{h}}:=[X,Y]_{\mathfrak{a}}=-[Y,X]_{\mathfrak{a}}=:[Y,X]_{\mathfrak{h}}.

Proof.

Step 1. Clearly 𝔥0¯=𝔥0𝔥α+β\mathfrak{h}_{\bar{0}}=\mathfrak{h}_{0}\oplus\mathfrak{h}_{\alpha+\beta} is a Lie algebra, namely it is a semidirect product of the Lie algebra 𝔞0\mathfrak{a}_{0} and 𝔞0\mathfrak{a}_{0}-module 𝔞α+β\mathfrak{a}_{\alpha+\beta}.

Step 2. 𝔥1¯\mathfrak{h}_{\bar{1}} is a 𝔥0¯\mathfrak{h}_{\bar{0}}-module. Clearly, 𝔥1¯\mathfrak{h}_{\bar{1}} is a 𝔥0\mathfrak{h}_{0}-module and 𝔥α+β\mathfrak{h}_{\alpha+\beta} acts trivially on 𝔥1¯\mathfrak{h}_{\bar{1}}.

Step 3. Note that the map 𝔥1¯𝔥1¯𝔥0¯\mathfrak{h}_{\bar{1}}\otimes\mathfrak{h}_{\bar{1}}\to\mathfrak{h}_{\bar{0}} induced by [,]𝔥[\,,]_{\mathfrak{h}} is an 𝔥0¯\mathfrak{h}_{\bar{0}}-module map. It remains only to check Jacobi identity for homogeneous X,Y,Z𝔥1¯=𝔥α𝔥βX,Y,Z\in\mathfrak{h}_{\bar{1}}=\mathfrak{h}_{\alpha}\oplus\mathfrak{h}_{\beta}. However the product of any three elements of this form is 0. ∎

Later we will give another proof of this result. It is easy to check that π\pi^{\prime} is defined on the morphisms of the form gr𝐓(ϕ)\mathrm{gr}^{\prime}\circ\mathbf{T}^{\prime}(\phi), where ϕ:𝔤𝔤1\phi:\mathfrak{g}\to\mathfrak{g}_{1} is a morphism of Lie superalgebras.

Functor inverse ι\iota^{\prime}. Above we explained the meaning of the functor inverse ι\iota for supermanifolds. Now we show that an analogue of this functor can be defined in the category of Lie superalgebras. Let 𝔤\mathfrak{g} be a Lie superalgebra. Denote by 𝔭:=ιπgr𝐓(𝔤)\mathfrak{p}:=\iota^{\prime}\circ\pi^{\prime}\circ\mathrm{gr}^{\prime}\circ\mathbf{T}^{\prime}(\mathfrak{g}) a \mathbb{Z}-graded subsuperspace of 𝔥:=π(gr(𝐓(𝔤)))\mathfrak{h}:=\pi^{\prime}(\mathrm{gr}^{\prime}(\mathbf{T}^{\prime}(\mathfrak{g}))) with support {0,1,2}\{0,1,2\}, where |1|=1¯|1|=\bar{1}, given by

𝔭0:=𝔥0¯,𝔭1={Y+dY|Y𝔤1¯}𝔥α𝔥β,𝔭2:=𝔥α+β.\displaystyle\mathfrak{p}_{0}:=\mathfrak{h}_{\bar{0}},\quad\mathfrak{p}_{1}=\{Y+\mathrm{d}Y\,\,|\,\,Y\in\mathfrak{g}_{\bar{1}}\}\subset\mathfrak{h}_{\alpha}\oplus\mathfrak{h}_{\beta},\quad\mathfrak{p}_{2}:=\mathfrak{h}_{\alpha+\beta}.
Proposition 33.

The superspace 𝔭\mathfrak{p} is a \mathbb{Z}-graded Lie superalgebra with support {0,1,2}\{0,1,2\} and with |1|=1¯|1|=\bar{1}.

Proof.

The proposition follows from a direct calculation. For example let us show that [𝔭1,𝔭1]𝔭2[\mathfrak{p}_{1},\mathfrak{p}_{1}]\subset\mathfrak{p}_{2}. For Y1,Y2𝔤1Y_{1},Y_{2}\in\mathfrak{g}_{1} we have

[Y1+dY1,Y2+dY2]=\displaystyle[Y_{1}+\mathrm{d}Y_{1},Y_{2}+\mathrm{d}Y_{2}]= [Y1,Y2]+[dY1,Y2]+[Y1,dY2]+[dY1,dY2]=\displaystyle[Y_{1},Y_{2}]+[\mathrm{d}Y_{1},Y_{2}]+[Y_{1},\mathrm{d}Y_{2}]+[\mathrm{d}Y_{1},\mathrm{d}Y_{2}]=
d[Y1,Y2]+d[Y1,Y2]=2d[Y1,Y2]𝔭2.\displaystyle d[Y_{1},Y_{2}]+d[Y_{1},Y_{2}]=2d[Y_{1},Y_{2}]\in\mathfrak{p}_{2}.

Note that the functor ι\iota^{\prime} is defined only for Lie superalgebras of the form πgr𝐓(𝔤)\pi^{\prime}\circ\mathrm{gr}^{\prime}\circ\mathbf{T}^{\prime}(\mathfrak{g}). Further, if f:𝔤𝔤1f:\mathfrak{g}\to\mathfrak{g}_{1} is a morphism of Lie superalgebras, then the morphism πgr𝐓(f)\pi^{\prime}\circ\mathrm{gr}^{\prime}\circ\mathbf{T}^{\prime}(f) can be restricted to the subalgebras ιπgr𝐓(𝔤)ιπgr𝐓(𝔤1)\iota^{\prime}\circ\pi^{\prime}\circ\mathrm{gr}^{\prime}\circ\mathbf{T}^{\prime}(\mathfrak{g})\to\iota^{\prime}\circ\pi^{\prime}\circ\mathrm{gr}^{\prime}\circ\mathbf{T}^{\prime}(\mathfrak{g}_{1}). Summing up, we constructed the following functor

F2:=ιπgr𝐓:sLieAlggrLieAlg2.\mathrm{F}^{\prime}_{2}:=\iota^{\prime}\circ\pi^{\prime}\circ\mathrm{gr}^{\prime}\circ\mathbf{T}^{\prime}:\mathrm{sLieAlg}\to\mathrm{grLieAlg}_{2}.

This functor is an embedding of category sLieAlg\mathrm{sLieAlg} into grLieAlg2\mathrm{grLieAlg}_{2}.

Remark 34.

(1) It is necessary to apply the functor π\pi^{\prime} before the functor inverse. Indeed, if YY and dY\mathrm{d}Y, where Y𝔭1Y\in\mathfrak{p}_{1}, have different parities, we will get

(19) [Y1+dY1,Y2+dY2]=[Y1,Y2]+[dY1,Y2]+[Y1,dY2]+[dY1,dY2]=d[Y1,Y2]d[Y1,Y2]=0.\begin{split}[Y_{1}+\mathrm{d}Y_{1},Y_{2}+\mathrm{d}Y_{2}]=&[Y_{1},Y_{2}]+[\mathrm{d}Y_{1},Y_{2}]+[Y_{1},\mathrm{d}Y_{2}]+[\mathrm{d}Y_{1},\mathrm{d}Y_{2}]=\\ &\mathrm{d}[Y_{1},Y_{2}]-\mathrm{d}[Y_{1},Y_{2}]=0.\end{split}

(2) Another observation is the following. Lie superalgebras of the form gr(𝔨)\mathrm{gr}^{\prime}(\mathfrak{k}), where 𝔨\mathfrak{k} is a Lie superalgebra possesses another parity reversion

𝔨0¯𝔨1¯𝔨0¯𝔨1¯[1¯],\mathfrak{k}_{\bar{0}}\oplus\mathfrak{k}_{\bar{1}}\longmapsto\mathfrak{k}_{\bar{0}}\oplus\mathfrak{k}_{\bar{1}}[\bar{1}],

where [1¯][\bar{1}] is the shift of parity be 1¯\bar{1}. In other words we assume that all odd elements are even. Clearly 𝔨0¯𝔨1¯[1¯]\mathfrak{k}_{\bar{0}}\oplus\mathfrak{k}_{\bar{1}}[\bar{1}] is a Lie algebra (not superalgebra). In this case the same argument, see (19), shows that the functor inverse loses information about original Lie superalgebra.

5.2. Donagi and Witten construction for Lie supergroups

In this section we develop the construction of Section 3 for Lie supergroups. In details, we apply the functor F2\mathrm{F}_{2} to a Lie supergroup 𝒢\mathcal{G}. Again F2\mathrm{F}_{2} is a composition of four functors: 𝐓\mathbf{T} (tangent), gr\mathrm{gr} (split), π\pi (parity change) and ι\iota (inverse).

Tangent functor 𝐓\mathbf{T}. Let 𝒢\mathcal{G} be a Lie supergroup with multiplication μ\mu, inversion κ\kappa and identity ee. Clearly, 𝐓(𝒢)\mathbf{T}(\mathcal{G}) is a Lie supergroup as the functor 𝐓\mathbf{T} preserves products. Indeed, since 𝐓\mathbf{T} is a functor the morthisms 𝐓(μ)\mathbf{T}(\mu), 𝐓(κ)\mathbf{T}(\kappa) and 𝐓(e)\mathbf{T}(e) satisfy the Lie supergroup axioms.

Proposition 35.

The Lie superalgebra of the Lie supergroup 𝐓(𝒢)\mathbf{T}(\mathcal{G}) is equal to 𝐓(𝔤)\mathbf{T}^{\prime}(\mathfrak{g}), where 𝔤=Lie(𝒢)\mathfrak{g}=\mathrm{Lie}(\mathcal{G}).

Proof.

We shall follow the definition of the Lie functor of a Lie supergroup given in [BLMS]. Let X1,X2Te𝒢X_{1},X_{2}\in\mathrm{T}_{e}\mathcal{G} be homogeneous. Any homogeneous vector XTe𝒢X\in\mathrm{T}_{e}\mathcal{G} can be represented by a curve g:𝒢g:\mathcal{R}\to\mathcal{G} where \mathcal{R} is 1|0\mathbb{R}^{1|0} or 0|1\mathbb{R}^{0|1} depending on the parity of XX, so that

g(f)=f(e)+tX(f)modIt2,g^{\ast}(f)=f(e)+tX(f)\bmod I_{t}^{2},

for any f𝒪𝒢(U)f\in\mathcal{O}_{\mathcal{G}}(U) defined in a neighbourhood U𝒢0U\subset\mathcal{G}_{0} of ee, where tt is the distinguished coordinate on \mathcal{R} of parity 0¯\bar{0} or 1¯\bar{1}, and It=tI_{t}=\langle t\rangle is the ideal generated by tt. Say X1,X2Te𝒢X_{1},X_{2}\in\mathrm{T}_{e}\mathcal{G} are represented by curves gi:i𝒢g_{i}:\mathcal{R}_{i}\to\mathcal{G}, where i=1,2i=1,2 and s,ts,t be the distinguished coordinates on 1\mathcal{\mathbb{R}}_{1}, 2\mathcal{\mathbb{R}}_{2}, respectively. Then there exist ZTe𝒢Z\in\mathrm{T}_{e}\mathcal{G} such that for Φ:=g1g2(g1)1(g2)1:1×2𝒢\Phi:=g_{1}g_{2}(g_{1})^{-1}(g_{2})^{-1}:\mathcal{R}_{1}\times\mathcal{R}_{2}\to\mathcal{G} we have

Φ(f)=f(e)+(1)|X||Y|tsZ(f)modIt2+Is2.\Phi^{\ast}(f)=f(e)+(-1)^{|X||Y|}tsZ(f)\,\bmod{I_{t}^{2}+I_{s}^{2}}.

The bracket [X1,X2][X_{1},X_{2}] is defined as ZZ.

We are ready to describe the Lie superalgebra Lie(𝐓G)\operatorname{Lie}(\mathbf{T}G). Let (U,xA)(U,x^{A}) be a chart around eU𝒢0e\in U\subset\mathcal{G}_{0} with local coordinates (xA)(x^{A}). Without loss of generality we may assume that xA(e)=0x^{A}(e)=0. The group law assures that the Taylor expansion of the multiplication map m:𝒢×𝒢𝒢m:\mathcal{G}\times\mathcal{G}\to\mathcal{G} has the form

m((xA),(x¯A))=(xA+x¯A+12QBCAxCx¯BmodI),m((x^{A}),(\underline{x}^{A}))=(x^{A}+\underline{x}^{A}+\frac{1}{2}Q^{A}_{BC}x^{C}\underline{x}^{B}\bmod I),

where I=x2,x¯2I=\langle x^{2},\underline{x}^{2}\rangle. Moreover, QBCA=(1)B~C~QBCAQ^{A}_{BC}=-(-1)^{\tilde{B}\tilde{C}}Q^{A}_{BC}. The inverse (xA)1(x^{A})^{-1} is given by (xA)(modI)(-x^{A})(\bmod I). The multiplication on 𝐓(𝒢)\mathbf{T}(\mathcal{G}) is given by 𝐓(m)\mathbf{T}(m), ie.

(xA,x˙A)(x¯A,x¯˙A)=(xA+x¯A+12QBCAxCx¯B,x˙A+x¯˙A+12QBCA(x˙Cx¯B+xCx¯˙B)).(x^{A},\dot{x}^{A})\cdot(\underline{x}^{A},\dot{\underline{x}}^{A})=\left(x^{A}+\underline{x}^{A}+\frac{1}{2}Q^{A}_{BC}x^{C}\underline{x}^{B},\dot{x}^{A}+\dot{\underline{x}}^{A}+\frac{1}{2}Q^{A}_{BC}(\dot{x}^{C}\underline{x}^{B}+x^{C}\dot{\underline{x}}^{B})\right).

The inverse of (xA,x˙A)(x^{A},\dot{x}^{A}) is (xA,x˙A)(-x^{A},-\dot{x}^{A}) modulo II, hence

(xA,x˙A)(x¯A,x¯˙A)(xA,x˙A)(x¯A,x¯˙A)\displaystyle(x^{A},\dot{x}^{A})\cdot(\underline{x}^{A},\underline{\dot{x}}^{A})\cdot(-x^{A},-\dot{x}^{A})\cdot(-\underline{x}^{A},-\underline{\dot{x}}^{A})\equiv
(xA+x¯A+12QBCAxCx¯B,x˙A+x¯˙A+12QBCA(x˙Cx¯B+xCx¯˙B)\displaystyle\left(x^{A}+\underline{x}^{A}+\frac{1}{2}Q^{A}_{BC}x^{C}\underline{x}^{B},\dot{x}^{A}+\dot{\underline{x}}^{A}+\frac{1}{2}Q^{A}_{BC}(\dot{x}^{C}\underline{x}^{B}+x^{C}\dot{\underline{x}}^{B}\right)\cdot
(xAx¯A+12QBCAxCx¯B,x˙Ax¯˙A+12QBCA(x˙Cx¯B+xCx¯˙B))\displaystyle\cdot\left(-x^{A}-\underline{x}^{A}+\frac{1}{2}Q^{A}_{BC}x^{C}\underline{x}^{B},-\dot{x}^{A}-\dot{\underline{x}}^{A}+\frac{1}{2}Q^{A}_{BC}(\dot{x}^{C}\underline{x}^{B}+x^{C}\dot{\underline{x}}^{B})\right)
(QBCAxCx¯B,QBCA(x˙Cx¯B+xCx¯˙B))(modI).\displaystyle\equiv(Q^{A}_{BC}x^{C}\underline{x}^{B},Q^{A}_{BC}(\dot{x}^{C}\underline{x}^{B}+x^{C}\underline{\dot{x}}^{B}))\,(\bmod\,I).

We read off from the last formula that [xC,xB]=QBCAxA[\partial_{x^{C}},\partial_{x^{B}}]=Q^{A}_{BC}\partial_{x^{A}}, [x˙C,xB]=QBCAx˙A[\partial_{\dot{x}^{C}},\partial_{x^{B}}]=Q^{A}_{BC}\partial_{\dot{x}^{A}}, and [x˙A,x˙B]=0[\partial_{\dot{x}^{A}},\partial_{\dot{x}^{B}}]=0. This completes the proof. ∎

The following theorem was proved [Vi2].

Theorem 36.

[Vi2, Theorem 3] Let 𝒢\mathcal{G} be a Lie supergroup corresponding to the super Harish-Chandra pair (𝒢0,𝔤)(\mathcal{G}_{0},\mathfrak{g}), where 𝒢0\mathcal{G}_{0} is the underlying space of 𝒢\mathcal{G} and 𝔤=Lie(𝒢)\mathfrak{g}=\mathrm{Lie}(\mathcal{G}). Then gr(𝒢)\mathrm{gr}(\mathcal{G}) is a Lie supergroup corresponding to the following super Harish-Chandra pair (𝒢0,𝔤)(\mathcal{G}_{0},\mathfrak{g}^{\prime}), where 𝔤=gr(𝔤)\mathfrak{g}^{\prime}=\mathrm{gr}^{\prime}(\mathfrak{g}).

Proposition 35 and Theorem 36 imply the following corollary.

Corollary 37.

Let 𝒢\mathcal{G} be a Lie supergroup, then

gr𝐓(Lie𝒢)=Lie(gr𝐓(𝒢)).\mathrm{gr}^{\prime}\circ\mathbf{T}^{\prime}(\operatorname{Lie}\mathcal{G})=\mathrm{Lie}(\mathrm{gr}\circ\mathbf{T}(\mathcal{G})).

As a consequence we get that the Lie supergroup gr𝐓(𝒢)\mathrm{gr}\circ\mathbf{T}(\mathcal{G}) is a graded Lie supergroup of type {0,α,β,α+β}\{0,\alpha,\beta,\alpha+\beta\}, where |α|=1¯|\alpha|=\bar{1} and |β|=0¯|\beta|=\bar{0}, with the graded Harish-Chandra pair (𝒢0,gr𝐓(𝔤))(\mathcal{G}_{0},\mathrm{gr}^{\prime}\circ\mathbf{T}^{\prime}(\mathfrak{g})) of the same type {0,α,β,α+β}\{0,\alpha,\beta,\alpha+\beta\}. (Compare also with Proposition 23.)

Functor parity change π\pi. Let us describe the functor π\pi using graded Harish-Chandra pairs. Let (𝒢0,gr𝐓(𝔤))(\mathcal{G}_{0},\mathrm{gr}^{\prime}\circ\mathbf{T}^{\prime}(\mathfrak{g})) be the graded Harish-Chandra pair of type {0,α,β,α+β}\{0,\alpha,\beta,\alpha+\beta\} of the Lie supergroup gr𝐓(𝒢)\mathrm{gr}\circ\mathbf{T}(\mathcal{G}) as above. In this case gr𝐓(𝒢)\mathrm{gr}\circ\mathbf{T}(\mathcal{G}) is a group object in the category of double vector bundles. As we have seen in Section 3, the double vector bundle gr𝐓(𝒢)\mathrm{gr}\circ\mathbf{T}(\mathcal{G}) possesses the following parity reversion: we change the parity of the weight β\beta from even to odd. We denote this new vector bundle by πgr𝐓(𝒢)\pi\circ\mathrm{gr}\circ\mathbf{T}(\mathcal{G}). Further we can define a Lie supergroup structure on πgr𝐓(𝒢)\pi\circ\mathrm{gr}\circ\mathbf{T}(\mathcal{G}) using the graded Harish-Chandra pair (𝒢0,πgr𝐓(𝔤))(\mathcal{G}_{0},\pi^{\prime}\circ\mathrm{gr}^{\prime}\circ\mathbf{T}(\mathfrak{g})) of type Δ={0,α,β,α+β}\Delta=\{0,\alpha,\beta,\alpha+\beta\}, where |α|=|β|=1¯|\alpha|=|\beta|=\bar{1}. Now the Lie supergroup πgr𝐓(𝒢)\pi\circ\mathrm{gr}\circ\mathbf{T}^{\prime}(\mathcal{G}) is defined.

Let us describe the Lie supergroup morphisms of πgr𝐓(𝒢)\pi\circ\mathrm{gr}\circ\mathbf{T}^{\prime}(\mathcal{G}) using the language of double vector bundles. If μ\mu, κ\kappa and ee are group morphisms of the Lie supergroup 𝒢\mathcal{G}, then gr𝐓(𝒢)\mathrm{gr}\circ\mathbf{T}(\mathcal{G}) is a group object in the category of double vector bundles with the structure morphisms gr𝐓(μ)\mathrm{gr}\circ\mathbf{T}(\mu), gr𝐓(κ)\mathrm{gr}\circ\mathbf{T}(\kappa) and gr𝐓(e)\mathrm{gr}\circ\mathbf{T}(e). Since the category of double vector bundles possesses the parity change: β\beta even to β\beta odd, we denote by πgr𝐓(𝒢)\pi\circ\mathrm{gr}\circ\mathbf{T}(\mathcal{G}) with πgr𝐓(μ)\pi\circ\mathrm{gr}\circ\mathbf{T}(\mu), πgr𝐓(κ)\pi\circ\mathrm{gr}\circ\mathbf{T}(\kappa) and πgr𝐓(e)\pi\circ\mathrm{gr}\circ\mathbf{T}(e) the result of this parity change. Formulas (6) tells us that this definition coincides with the definition in terms of graded Harish-Chandra pairs.

Functor inverse ι\iota. In Proposition 33 we saw that the Lie superalgebra πgr𝐓(𝔤)\pi^{\prime}\circ\mathrm{gr}^{\prime}\circ\mathbf{T}^{\prime}(\mathfrak{g}) possesses a \mathbb{Z}-graded Lie subsuperalgebra 𝔭=ιπgr𝐓(𝔤)\mathfrak{p}=\iota^{\prime}\circ\pi^{\prime}\circ\mathrm{gr}^{\prime}\circ\mathbf{T}^{\prime}(\mathfrak{g}). We define by ιπgr𝐓(𝒢)\iota\circ\pi\circ\mathrm{gr}\circ\mathbf{T}(\mathcal{G}) the corresponding Lie subsupergroup. More precisely, we define ιπgr𝐓(𝒢)\iota\circ\pi\circ\mathrm{gr}\circ\mathbf{T}(\mathcal{G}) using the graded Harish-Chandra pairs (𝒢0,ιπgr𝐓(𝔤)).(\mathcal{G}_{0},\iota^{\prime}\circ\pi^{\prime}\circ\mathrm{gr}^{\prime}\circ\mathbf{T}^{\prime}(\mathfrak{g})).

6. Generalized Donagi–Witten construction for
Lie superalgebras

In Section 4 we constructed an injective functor F:=F\mathrm{F}:=\mathrm{F}_{\infty} from the category of supermanifolds to the category of graded manifolds. In this section we show that this functor can be defined in the category of Lie supergroups. We start with Lie superalgebras. In more details we will construct a functor F\mathrm{F}^{\prime} from the category of Lie superalgebras sLieAlg\mathrm{sLieAlg} to the category of non-negatively \mathbb{Z}-graded Lie algebras grLieAlg\mathrm{grLieAlg}_{\infty}. Further we will use these results for the category of Lie supergroups. Again the functor F\mathrm{F}^{\prime} is a composition of four functors: the iterated antitangent functor 𝐓\mathbf{T}^{\prime\infty}, the functor split gr\mathrm{gr}^{\prime}, the functor parity change π\pi^{\prime} and the functor inverse ι\iota^{\prime}.

6.1. Iterated tangent functor 𝐓\mathbf{T}^{\prime\infty}

Let 𝔤\mathfrak{g} be a Lie superalgebra. Let us describe the superalgebra 𝐓(𝔤)\mathbf{T}^{\prime\infty}(\mathfrak{g}), where 𝐓:=𝐓𝐓\mathbf{T}^{\prime\infty}:=\mathbf{T}^{\prime}\circ\mathbf{T}^{\prime}\circ\cdots is the infinitely many times iterated antitangent functor. Let us consider first twice iterated tangent functor T2(𝔤)\mathrm{T^{\prime}}^{2}(\mathfrak{g}). By definition

T2(𝔤)=𝐓(𝐓(𝔤))=𝐓(𝔤d1(𝔤))=𝔤d1(𝔤)d2(𝔤d1(𝔤)).\mathrm{T^{\prime}}^{2}(\mathfrak{g})=\mathbf{T}^{\prime}(\mathbf{T}^{\prime}(\mathfrak{g}))=\mathbf{T}^{\prime}(\mathfrak{g}\oplus\mathrm{d}_{1}(\mathfrak{g}))=\mathfrak{g}\oplus\mathrm{d}_{1}(\mathfrak{g})\oplus\mathrm{d}_{2}(\mathfrak{g}\oplus\mathrm{d}_{1}(\mathfrak{g})).

We replaced d\mathrm{d} from Section 5.1 by d1\mathrm{d}_{1} and d2\mathrm{d}_{2} is the second de Rham differential. The multiplication in TT(𝔤)\mathrm{T^{\prime}\circ T^{\prime}}(\mathfrak{g}) is defined in a natural way. Moreover we can easily verify the following lemma.

Lemma 38.

To obtain the multiplication in T2(𝔤)\mathrm{T^{\prime}}^{2}(\mathfrak{g}) we can use the following rule: [diX,djY]=(1)|X|didj([X,Y])[\mathrm{d}_{i}X,\mathrm{d}_{j}Y]=(-1)^{|X|}\mathrm{d}_{i}\mathrm{d}_{j}([X,Y]) for any X,YT2(𝔤)X,Y\in\mathrm{T^{\prime}}^{2}(\mathfrak{g}).

Corollary 39.

Since the operators di\mathrm{d}_{i} assumed to be odd, we have

didi([X,Y])=0\mathrm{d}_{i}\mathrm{d}_{i}([X,Y])=0

for any X,YT2(𝔤)X,Y\in\mathrm{T^{\prime}}^{2}(\mathfrak{g}).

To define the functor T\mathrm{T^{\prime}}^{\infty} we use de Rham differentials d1,d2,,dn,.\mathrm{d}_{1},\mathrm{d}_{2},\ldots,\mathrm{d}_{n},\ldots. Lemma 38 and Corrolary 39 imply the following lemma.

Lemma 40.

The Lie superalgebra T(𝔤)\mathrm{T^{\prime}}^{\infty}(\mathfrak{g}) is an infinite dimensional Lie superalgebra with the underlying vector space

p0i1<<ipdi1dip(𝔤)\bigoplus_{p\geq 0}\bigoplus_{i_{1}<\cdots<i_{p}}\mathrm{d}_{i_{1}}\cdots\mathrm{d}_{i_{p}}(\mathfrak{g})

and multiplication defined by the following formula

[diX,djY]=(1)|X|didj([X,Y])[\mathrm{d}_{i}X,\mathrm{d}_{j}Y]=(-1)^{|X|}\mathrm{d}_{i}\mathrm{d}_{j}([X,Y])

for any X,YT(𝔤)X,Y\in\mathrm{T^{\prime}}^{\infty}(\mathfrak{g}).

6.2. A connection with the functor of points for Lie superalgebras

For details about the functor of points for Lie superalgebras we refer for example to [Gav, Section 2.2.4]. Let us recall this construction. To any Lie superalgebra 𝔤\mathfrak{g} we can associate a functor L𝔤L_{\mathfrak{g}} from the category of supercommutative algebras to the category of Lie algebras. It is defined as follows

L𝔤(A)=(A𝔤)0¯,L_{\mathfrak{g}}(A)=(A\otimes\mathfrak{g})_{\bar{0}},

where AA is a super-commutative algebra. The product A𝔤A\otimes\mathfrak{g} is a Lie superalgebra with the following multiplication

(20) [aX,aX]:=(1)|X||a|aa[X,X].[a\otimes X,a^{\prime}\otimes X^{\prime}]:=(-1)^{|X||a^{\prime}|}aa^{\prime}\otimes[X,X^{\prime}].

Comparing with Lemma 40 we see that the Lie superalgebra Tk(𝔤)\mathrm{T^{\prime}}^{k}(\mathfrak{g}) is isomorphic to the Lie superalgebra (ξ1,,ξk)𝔤\bigwedge(\xi_{1},\ldots,\xi_{k})\otimes\mathfrak{g}. Clearly, T(𝔤)\mathrm{T^{\prime}}^{\infty}(\mathfrak{g}) is isomorphic to (ξ1,ξ2,)𝔤\bigwedge(\xi_{1},\xi_{2},\ldots)\otimes\mathfrak{g}, where (ξ1,ξ2,)\bigwedge(\xi_{1},\xi_{2},\ldots) is the Grassmann algebra with infinitely many variables ξ1,ξ2,\xi_{1},\xi_{2},\ldots. Now we see that

L𝔤((ξ1,ξ2,))=T(𝔤)0¯.L_{\mathfrak{g}}(\bigwedge(\xi_{1},\xi_{2},\ldots))=\mathrm{T^{\prime}}^{\infty}(\mathfrak{g})_{\bar{0}}.

6.3. Functors split gr\mathrm{gr}^{\prime} and parity change π\pi^{\prime}

The functor split is defined as above.

Remark 41.

The Lie superalgebra grT(𝔤)\mathrm{gr}^{\prime}\circ\mathrm{T^{\prime}}^{\infty}(\mathfrak{g}) is a graded Lie superalgebra with support Δ\Delta, where Δ\Delta is the maximal multiplicity free weight system generated by α,β1,,βk,\alpha,\beta_{1},\ldots,\beta_{k},\ldots. Here α\alpha is odd, while β1,,βk,\beta_{1},\ldots,\beta_{k},\ldots are even. The grading is defined as follows. We assign the weight α+βii++βip\alpha+\beta_{i_{i}}+\cdots+\beta_{i_{p}} to di1dip(Z)\mathrm{d}_{i_{1}}\cdots\mathrm{d}_{i_{p}}(Z), if di1dip(Z)\mathrm{d}_{i_{1}}\cdots\mathrm{d}_{i_{p}}(Z) is odd, and we assign the weight βii++βip\beta_{i_{i}}+\cdots+\beta_{i_{p}} to di1dip(Z)\mathrm{d}_{i_{1}}\cdots\mathrm{d}_{i_{p}}(Z) if di1dip(Z)\mathrm{d}_{i_{1}}\cdots\mathrm{d}_{i_{p}}(Z) is even.

For example elements of the subspace 𝔤0¯\mathfrak{g}_{\bar{0}} have weight 0 and elements of 𝔤1¯\mathfrak{g}_{\bar{1}} have weight α\alpha.

Let us define the parity change functor π\pi^{\prime}. Let us take the Lie superalgebra gr(T(𝔤))\mathrm{gr}^{\prime}(\mathrm{T^{\prime}}^{\infty}(\mathfrak{g})), where 𝔤\mathfrak{g} is a Lie superalgebra. The functor parity change π\pi^{\prime} is defined as follows

𝔥:=π(gr(T(𝔤)))=gr(T(𝔤))\mathfrak{h}:=\pi^{\prime}(\mathrm{gr^{\prime}}(\mathrm{T^{\prime}}^{\infty}(\mathfrak{g})))=\mathrm{gr^{\prime}}(\mathrm{T^{\prime}}^{\infty}(\mathfrak{g}))

as 𝔤0¯\mathfrak{g}_{\bar{0}}-modules. Further we assume that all operators di\mathrm{d}_{i} are even and again didi=0\mathrm{d}_{i}\circ\mathrm{d}_{i}=0. In other words, this means that

𝔥0¯:=p0i1<<ipdi1dip(𝔤0¯),𝔥1¯:=p0i1<<ipdi1dip(𝔤1¯).\displaystyle\mathfrak{h}_{\bar{0}}:=\bigoplus_{p\geq 0}\bigoplus_{i_{1}<\cdots<i_{p}}\mathrm{d}_{i_{1}}\cdots\mathrm{d}_{i_{p}}(\mathfrak{g}_{\bar{0}}),\quad\mathfrak{h}_{\bar{1}}:=\bigoplus_{p\geq 0}\bigoplus_{i_{1}<\cdots<i_{p}}\mathrm{d}_{i_{1}}\cdots\mathrm{d}_{i_{p}}(\mathfrak{g}_{\bar{1}}).

To simplify our presentation we will use the following notations

I={i1,,ip},J={j1,,jq},K={k1,,kr},\displaystyle I=\{i_{1},\ldots,i_{p}\},\quad J=\{j_{1},\ldots,j_{q}\},\quad K=\{k_{1},\ldots,k_{r}\},

and dI\mathrm{d}_{I} for di1dip\mathrm{d}_{i_{1}}\cdots\mathrm{d}_{i_{p}}. We denote by I:=C(I)mod  2=p¯\sharp I:=C(I)\,\,mod\,\,2=\bar{p} the cardinality of II modulo 22. The multiplication in 𝔥=π(gr(T(𝔤)))\mathfrak{h}=\pi^{\prime}(\mathrm{gr}^{\prime}(\mathrm{T^{\prime}}^{\infty}(\mathfrak{g}))) is defined by the following rules

  1. (Rule 1)

    If IJI\cap J\neq\emptyset, we have [dI(𝔤),dJ(𝔤)]={0}[\mathrm{d}_{I}(\mathfrak{g}),\mathrm{d}_{J}(\mathfrak{g})]=\{0\}.

  2. (Rule 2)

    If I+i¯\sharp I+\bar{i} and J+j¯\sharp J+\bar{j} are odd, we have [dI(𝔤i¯),dI(𝔤j¯)]={0}[\mathrm{d}_{I}(\mathfrak{g}_{\bar{i}}),\mathrm{d}_{I}(\mathfrak{g}_{\bar{j}})]=\{0\}.

  3. (Rule 3)

    In other cases we have [dI(𝔤i¯),dI(𝔤j¯)]=dIJ([𝔤i¯,𝔤j¯])[\mathrm{d}_{I}(\mathfrak{g}_{\bar{i}}),\mathrm{d}_{I}(\mathfrak{g}_{\bar{j}})]=\mathrm{d}_{I\cup J}([\mathfrak{g}_{\bar{i}},\mathfrak{g}_{\bar{j}}]).

Theorem 42.

The superspace 𝔥=π(gr(T(𝔤)))\mathfrak{h}=\pi^{\prime}(\mathrm{gr}^{\prime}(\mathrm{T^{\prime}}^{\infty}(\mathfrak{g}))) is a Lie superalgebra.

Proof.

Step 1. Let us prove first that 𝔥0¯\mathfrak{h}_{\bar{0}} is a Lie algebra. Consider the following subspace

𝔯:=p0i1<<ipdi1dip(𝔤0¯)gr(T(𝔤)).\mathfrak{r}:=\bigoplus_{p\geq 0}\bigoplus_{i_{1}<\cdots<i_{p}}\mathrm{d}_{i_{1}}\cdots\mathrm{d}_{i_{p}}(\mathfrak{g}_{\bar{0}})\subset\mathrm{gr}^{\prime}(\mathrm{T^{\prime}}^{\infty}(\mathfrak{g})).

Clearly, 𝔯\mathfrak{r} is a Lie subsuperalgebra in gr(T(𝔤))\mathrm{gr^{\prime}}(\mathrm{T^{\prime}}^{\infty}(\mathfrak{g})) and this Lie superalgebra is split. This is 𝔯=𝔯0¯𝔯1¯\mathfrak{r}=\mathfrak{r}_{\bar{0}}\oplus\mathfrak{r}_{\bar{1}} with [𝔯1¯,𝔯1¯]={0}[\mathfrak{r}_{\bar{1}},\mathfrak{r}_{\bar{1}}]=\{0\}. Such Lie superalgebras possesses a parity reversion, see Remark 34 part (2). We can see that in notations of Remark 34 we have 𝔯0¯𝔯1¯[1¯]=𝔥0¯\mathfrak{r}_{\bar{0}}\oplus\mathfrak{r}_{\bar{1}}[\bar{1}]=\mathfrak{h}_{\bar{0}}.

Step 2. Let us prove that 𝔥1¯\mathfrak{h}_{\bar{1}} is an 𝔥0¯\mathfrak{h}_{\bar{0}}-module. Let us take dI(X)dI(𝔤0¯)\mathrm{d}_{I}(X)\in\mathrm{d}_{I}(\mathfrak{g}_{\bar{0}}), dJ(Y)dJ(𝔤0¯)\mathrm{d}_{J}(Y)\in\mathrm{d}_{J}(\mathfrak{g}_{\bar{0}}) and dK(Z)dK(𝔤1¯)\mathrm{d}_{K}(Z)\in\mathrm{d}_{K}(\mathfrak{g}_{\bar{1}}) with IJ=I\cap J=\emptyset, IK=I\cap K=\emptyset and JK=J\cap K=\emptyset. We consider the following cases

  1. (1)

    Let I=J=0¯\sharp I=\sharp J=\bar{0}, any KK. Then

    [dI(X),[dJ(Y),dK(Z)]]+[dJ(Y),[dK(Z),dI(X)]]+[dK(Z),[dI(X),dJ(Y)]]=\displaystyle[\mathrm{d}_{I}(X),[\mathrm{d}_{J}(Y),\mathrm{d}_{K}(Z)]]+[\mathrm{d}_{J}(Y),[\mathrm{d}_{K}(Z),\mathrm{d}_{I}(X)]]+[\mathrm{d}_{K}(Z),[\mathrm{d}_{I}(X),\mathrm{d}_{J}(Y)]]=
    dIdJdK([X,[Y,Z]])+dJdKdI([Y,[Z,X]])+dKdIdJ([Z,[X,Y]])=\displaystyle\mathrm{d}_{I}\circ\mathrm{d}_{J}\circ\mathrm{d}_{K}([X,[Y,Z]])+\mathrm{d}_{J}\circ\mathrm{d}_{K}\circ\mathrm{d}_{I}([Y,[Z,X]])+\mathrm{d}_{K}\circ\mathrm{d}_{I}\circ\mathrm{d}_{J}([Z,[X,Y]])=
    dIdJdK([X,[Y,Z]]+[Y,[Z,X]]+[Z,[X,Y]])=0.\displaystyle\mathrm{d}_{I}\circ\mathrm{d}_{J}\circ\mathrm{d}_{K}([X,[Y,Z]]+[Y,[Z,X]]+[Z,[X,Y]])=0.
  2. (2)

    Let I=1¯\sharp I=\bar{1}, J=0¯\sharp J=\bar{0} and K=0¯\sharp K=\bar{0}. Then

    [dI(X),[dJ(Y),dK(Z)]]=[dJ(Y),[dK(Z),dI(X)]]=[dK(Z),[dI(X),dJ(Y)]]=0.\displaystyle[\mathrm{d}_{I}(X),[\mathrm{d}_{J}(Y),\mathrm{d}_{K}(Z)]]=[\mathrm{d}_{J}(Y),[\mathrm{d}_{K}(Z),\mathrm{d}_{I}(X)]]=[\mathrm{d}_{K}(Z),[\mathrm{d}_{I}(X),\mathrm{d}_{J}(Y)]]=0.
  3. (3)

    Let I=1¯\sharp I=\bar{1}, J=0¯\sharp J=\bar{0} and K=1¯\sharp K=\bar{1}. Then

    [dI(X),[dJ(Y),dK(Z)]]+[dJ(Y),[dK(Z),dI(X)]]+[dK(Z),[dI(X),dJ(Y)]]=\displaystyle[\mathrm{d}_{I}(X),[\mathrm{d}_{J}(Y),\mathrm{d}_{K}(Z)]]+[\mathrm{d}_{J}(Y),[\mathrm{d}_{K}(Z),\mathrm{d}_{I}(X)]]+[\mathrm{d}_{K}(Z),[\mathrm{d}_{I}(X),\mathrm{d}_{J}(Y)]]=
    dIdJdK([X,[Y,Z]]+[Y,[Z,X]]+[Z,[X,Y]])=0.\displaystyle\mathrm{d}_{I}\circ\mathrm{d}_{J}\circ\mathrm{d}_{K}([X,[Y,Z]]+[Y,[Z,X]]+[Z,[X,Y]])=0.
  4. (4)

    Let I=J=1¯\sharp I=\sharp J=\bar{1} and K=0¯\sharp K=\bar{0}. Then

    [dI(X),[dJ(Y),dK(Z)]]=[dJ(Y),[dK(Z),dI(X)]]=[dK(Z),[dI(X),dJ(Y)]]=0.\displaystyle[\mathrm{d}_{I}(X),[\mathrm{d}_{J}(Y),\mathrm{d}_{K}(Z)]]=[\mathrm{d}_{J}(Y),[\mathrm{d}_{K}(Z),\mathrm{d}_{I}(X)]]=[\mathrm{d}_{K}(Z),[\mathrm{d}_{I}(X),\mathrm{d}_{J}(Y)]]=0.
  5. (5)

    Let I=J=K=1¯\sharp I=\sharp J=\sharp K=\bar{1}. Then

    [dI(X),[dJ(Y),dK(Z)]]=[dJ(Y),[dK(Z),dI(X)]]=[dK(Z),[dI(X),dJ(Y)]]=0.\displaystyle[\mathrm{d}_{I}(X),[\mathrm{d}_{J}(Y),\mathrm{d}_{K}(Z)]]=[\mathrm{d}_{J}(Y),[\mathrm{d}_{K}(Z),\mathrm{d}_{I}(X)]]=[\mathrm{d}_{K}(Z),[\mathrm{d}_{I}(X),\mathrm{d}_{J}(Y)]]=0.

Step 3. Let us check Jacobi identity for elements from 𝔥1¯\mathfrak{h}_{\bar{1}}. Let us take dI(X)dI(𝔤1¯)\mathrm{d}_{I}(X)\in\mathrm{d}_{I}(\mathfrak{g}_{\bar{1}}), dJ(Y)dJ(𝔤1¯)\mathrm{d}_{J}(Y)\in\mathrm{d}_{J}(\mathfrak{g}_{\bar{1}}) and dK(Z)dK(𝔤1¯)\mathrm{d}_{K}(Z)\in\mathrm{d}_{K}(\mathfrak{g}_{\bar{1}}) with IJ=I\cap J=\emptyset, IK=I\cap K=\emptyset and JK=J\cap K=\emptyset and consider the following cases

  1. (1)

    Let I=J=0¯\sharp I=\sharp J=\bar{0}, any KK. Then

    [dI(X),[dJ(Y),dK(Z)]]=[dJ(Y),[dK(Z),dI(X)]]=[dK(Z),[dI(X),dJ(Y)]]=0.\displaystyle[\mathrm{d}_{I}(X),[\mathrm{d}_{J}(Y),\mathrm{d}_{K}(Z)]]=[\mathrm{d}_{J}(Y),[\mathrm{d}_{K}(Z),\mathrm{d}_{I}(X)]]=[\mathrm{d}_{K}(Z),[\mathrm{d}_{I}(X),\mathrm{d}_{J}(Y)]]=0.
  2. (2)

    Let I=0¯\sharp I=\bar{0}, J=1¯\sharp J=\bar{1} and K=1¯\sharp K=\bar{1}. Then

    [dI(X),[dJ(Y),dK(Z)]]+[dJ(Y),[dK(Z),dI(X)]]+[dK(Z),[dI(X),dJ(Y)]]=\displaystyle[\mathrm{d}_{I}(X),[\mathrm{d}_{J}(Y),\mathrm{d}_{K}(Z)]]+[\mathrm{d}_{J}(Y),[\mathrm{d}_{K}(Z),\mathrm{d}_{I}(X)]]+[\mathrm{d}_{K}(Z),[\mathrm{d}_{I}(X),\mathrm{d}_{J}(Y)]]=
    dIdJdK([X,[Y,Z]]+[Y,[Z,X]]+[Z,[X,Y]])=0.\displaystyle\mathrm{d}_{I}\circ\mathrm{d}_{J}\circ\mathrm{d}_{K}([X,[Y,Z]]+[Y,[Z,X]]+[Z,[X,Y]])=0.
  3. (3)

    Let I=J=K=1¯\sharp I=\sharp J=\sharp K=\bar{1}. Then again

    [dI(X),[dJ(Y),dK(Z)]]+[dJ(Y),[dK(Z),dI(X)]]+[dK(Z),[dI(X),dJ(Y)]]=\displaystyle[\mathrm{d}_{I}(X),[\mathrm{d}_{J}(Y),\mathrm{d}_{K}(Z)]]+[\mathrm{d}_{J}(Y),[\mathrm{d}_{K}(Z),\mathrm{d}_{I}(X)]]+[\mathrm{d}_{K}(Z),[\mathrm{d}_{I}(X),\mathrm{d}_{J}(Y)]]=
    dIdJdK([X,[Y,Z]]+[Y,[Z,X]]+[Z,[X,Y]])=0.\displaystyle\mathrm{d}_{I}\circ\mathrm{d}_{J}\circ\mathrm{d}_{K}([X,[Y,Z]]+[Y,[Z,X]]+[Z,[X,Y]])=0.

The proof is complete. ∎

Remark 43.

It is unexpected that such a parity change can give a well-defined Lie superalgebra.

6.4. Functor inverse ι\iota^{\prime}

Our goal now is to define a \mathbb{Z}-graded subsuperalgebra 𝔭:=ιπgrT(𝔤)\mathfrak{p}:=\iota^{\prime}\circ\pi^{\prime}\circ\mathrm{gr}^{\prime}\circ\mathrm{T^{\prime}}^{\infty}(\mathfrak{g}) in πgrT(𝔤)\pi^{\prime}\circ\mathrm{gr}^{\prime}\circ\mathrm{T^{\prime}}^{\infty}(\mathfrak{g}). We put

𝔭0:=𝔤0¯;\displaystyle\mathfrak{p}_{0}:=\mathfrak{g}_{\bar{0}};
𝔭1:=diag(𝔤1¯idi(𝔤1¯));\displaystyle\mathfrak{p}_{1}:=\mathrm{diag}(\mathfrak{g}_{\bar{1}}\oplus\bigoplus_{i}\mathrm{d}_{i}(\mathfrak{g}_{\bar{1}}));
𝔭2:=diag(idi(𝔤0¯)i<jdidj(𝔤0¯));\displaystyle\mathfrak{p}_{2}:=\mathrm{diag}(\bigoplus_{i}\mathrm{d}_{i}(\mathfrak{g}_{\bar{0}})\oplus\bigoplus_{i<j}\mathrm{d}_{i}\mathrm{d}_{j}(\mathfrak{g}_{\bar{0}}));
𝔭3:=diag(i<jdidj(𝔤1¯)i<j<kdidjdk(𝔤1¯));\displaystyle\mathfrak{p}_{3}:=\mathrm{diag}(\bigoplus_{i<j}\mathrm{d}_{i}\mathrm{d}_{j}(\mathfrak{g}_{\bar{1}})\oplus\bigoplus_{i<j<k}\mathrm{d}_{i}\mathrm{d}_{j}\mathrm{d}_{k}(\mathfrak{g}_{\bar{1}}));
\displaystyle\cdots
𝔭n:=diag(C(I)=n1dI(𝔤n¯)C(J)=ndJ(𝔤n¯));\displaystyle\mathfrak{p}_{n}:=\mathrm{diag}(\bigoplus_{C(I)=n-1}\mathrm{d}_{I}(\mathfrak{g}_{\bar{n}})\oplus\bigoplus_{C(J)=n}\mathrm{d}_{J}(\mathfrak{g}_{\bar{n}}));
\displaystyle\cdots

Here C(I)C(I) is the cardinality of II.

Proposition 44.

The subsuperspace 𝔭\mathfrak{p} is a \mathbb{Z}-graded Lie subsuperalgebra.

Proof.

Let us prove that [𝔭i,𝔭j]𝔭i+j[\mathfrak{p}_{i},\mathfrak{p}_{j}]\subset\mathfrak{p}_{i+j}. Let us take X𝔤i¯X\in\mathfrak{g}_{\bar{i}} and Y𝔤j¯Y\in\mathfrak{g}_{\bar{j}}. Consider

[C(I)=i1dI(X)+C(J)=idJ(X),C(I)=j1dI(Y)+C(J)=jdJ(Y)].\displaystyle\left[\sum_{C(I)=i-1}\mathrm{d}_{I}(X)+\sum_{C(J)=i}\mathrm{d}_{J}(X),\sum_{C(I)^{\prime}=j-1}\mathrm{d}_{I^{\prime}}(Y)+\sum_{C(J^{\prime})=j}\mathrm{d}_{J^{\prime}}(Y)\right].

Using Rules (2) we get

[C(I)=i1dI(X),C(I)=j1dI(Y)]=0.\displaystyle\left[\sum_{C(I)=i-1}\mathrm{d}_{I}(X),\sum_{C(I)^{\prime}=j-1}\mathrm{d}_{I^{\prime}}(Y)\right]=0.

Further,

[C(I)=i1dI(X),C(J)=jdJ(Y)]+[C(J)=idJ(X),C(I)=j1dI(Y)]+\displaystyle\left[\sum_{C(I)=i-1}\mathrm{d}_{I}(X),\sum_{C(J^{\prime})=j}\mathrm{d}_{J^{\prime}}(Y)\right]+\left[\sum_{C(J)=i}\mathrm{d}_{J}(X),\sum_{C(I^{\prime})=j-1}\mathrm{d}_{I^{\prime}}(Y)\right]+
[C(J)=idJ(X),C(J)=jdJ(Y)]=\displaystyle\left[\sum_{C(J)=i}\mathrm{d}_{J}(X),\sum_{C(J^{\prime})=j}\mathrm{d}_{J^{\prime}}(Y)\right]=
((i+j1j)+(i+j1j1))(i+ji)C(K)=i+j1dK([X,Y])+\displaystyle\underbrace{\left(\binom{i+j-1}{j}+\binom{i+j-1}{j-1}\right)}_{\binom{i+j}{i}}\sum_{C(K)=i+j-1}\mathrm{d}_{K}([X,Y])+
(i+ji)C(K)=i+jdK([X,Y]).\displaystyle\binom{i+j}{i}\sum_{C(K)=i+j}\mathrm{d}_{K}([X,Y]).

The proof is complete. ∎

If we have a morphism ψ\psi of Lie superalgebras, then πgrT(ψ)\pi^{\prime}\circ\mathrm{gr}^{\prime}\circ\mathrm{T^{\prime}}^{\infty}(\psi) preserves subalgebras 𝔭\mathfrak{p}’s. Therefore we can define ι\iota^{\prime} on morphisms. Therefore the functor ι\iota^{\prime} is defined on the image of the functor πgrT\pi^{\prime}\circ\mathrm{gr}^{\prime}\circ\mathrm{T^{\prime}}^{\infty}. Summing up, we constructed the following functor

F:=ιπgrT:sLieAlggrLieAlg.\mathrm{F}^{\prime}:=\iota^{\prime}\circ\pi^{\prime}\circ\mathrm{gr}^{\prime}\circ\mathrm{T^{\prime}}^{\infty}:\mathrm{sLieAlg}\to\mathrm{grLieAlg}.
Remark 45.

The Lie superalgebra 𝔭\mathfrak{p} is “locally isomorphic” to 𝔤\mathfrak{g} in the following sense. We have 𝔭0𝔤0¯\mathfrak{p}_{0}\simeq\mathfrak{g}_{\bar{0}} as Lie superalgebras and 𝔭n𝔤n¯\mathfrak{p}_{n}\simeq\mathfrak{g}_{\bar{n}} as 𝔤0¯\mathfrak{g}_{\bar{0}}-modules for any nn.

6.5. The functor F\mathrm{F}_{\infty} as a inverse limit

We can define the Lie superalgebras 𝐓(𝔤)\mathbf{T}^{\prime\infty}(\mathfrak{g}), gr(𝐓(𝔤))\mathrm{gr}^{\prime}(\mathbf{T}^{\prime\infty}(\mathfrak{g})), π(gr(𝐓(𝔤)))\pi^{\prime}(\mathrm{gr}^{\prime}(\mathbf{T}^{\prime\infty}(\mathfrak{g}))) and ι(π(gr(𝐓(𝔤))))\iota^{\prime}(\pi^{\prime}(\mathrm{gr}^{\prime}(\mathbf{T}^{\prime\infty}(\mathfrak{g})))) using the inverse limit. Indeed,

(21) 𝐓(𝔤)\displaystyle\mathbf{T}^{\prime\infty}(\mathfrak{g}) =lim𝐓n(𝔤);\displaystyle=\varprojlim\mathbf{T}^{\prime n}(\mathfrak{g});
(22) gr(𝐓(𝔤))\displaystyle\mathrm{gr}^{\prime}(\mathbf{T}^{\prime\infty}(\mathfrak{g})) =limgr(𝐓n(𝔤));\displaystyle=\varprojlim\mathrm{gr}^{\prime}(\mathbf{T}^{\prime n}(\mathfrak{g}));
(23) π(gr(𝐓(𝔤)))\displaystyle\pi^{\prime}(\mathrm{gr}^{\prime}(\mathbf{T}^{\prime\infty}(\mathfrak{g}))) =limπ(gr(𝐓n(𝔤)));\displaystyle=\varprojlim\pi^{\prime}(\mathrm{gr}^{\prime}(\mathbf{T}^{\prime n}(\mathfrak{g})));
(24) ι(π(gr(𝐓(𝔤))))\displaystyle\iota^{\prime}(\pi^{\prime}(\mathrm{gr}^{\prime}(\mathbf{T}^{\prime\infty}(\mathfrak{g})))) =limι(π(gr(𝐓n(𝔤)))).\displaystyle=\varprojlim\iota^{\prime}(\pi^{\prime}(\mathrm{gr}^{\prime}(\mathbf{T}^{\prime n}(\mathfrak{g})))).

Lemma 40 implies (21). Further, Equality (22) follows from definition of the functor gr\mathrm{gr}^{\prime}. To obtain (23) we can repeat Theorem 42 assuming that we have only nn de Rham differentials di\mathrm{d}_{i}. Finally (24) follows from Proposition 44 again assuming that we have only nn de Rham differentials di\mathrm{d}_{i}. Summing up,

F(𝔤)=F(𝔤)=limFn(𝔤).\mathrm{F}^{\prime}(\mathfrak{g})=\mathrm{F}^{\prime}_{\infty}(\mathfrak{g})=\varprojlim\mathrm{F}^{\prime}_{n}(\mathfrak{g}).

7. Generalized Donagi–Witten construction for
Lie supergroups

Again the functor F\mathrm{F} is a composition of four functors: the iterated tangent functor 𝐓\mathbf{T}^{\infty}, the functor split gr\mathrm{gr}, the functor parity change π\pi and the functor inverse ι\iota. Let 𝒢\mathcal{G} be a Lie supergroup with the supergroup morphisms μ\mu, κ\kappa and ee.

7.1. Iterated antitangent functor 𝐓\mathbf{T}^{\infty}

Above we considered the tangent functor 𝐓\mathbf{T} applying to a Lie supergroup 𝒢\mathcal{G}. We saw that 𝐓(𝒢)\mathbf{T}(\mathcal{G}) is a Lie supergroup again. Therefore we can iterate this procedure and get the Lie supergroup

𝐓2(𝒢):=𝐓(𝐓(𝒢)).\mathbf{T}^{2}(\mathcal{G}):=\mathbf{T}(\mathbf{T}(\mathcal{G})).

By definition we put

𝐓n(𝒢):=𝐓𝐓(𝒢),\mathbf{T}^{n}(\mathcal{G}):=\mathbf{T}\circ\cdots\circ\mathbf{T}(\mathcal{G}),

where on the left hand side the tangent functor 𝐓\mathbf{T} is iterated nn times. Now we define

𝐓(𝒢)=lim𝐓n(𝒢),\displaystyle\mathbf{T}^{\infty}(\mathcal{G})=\varprojlim\mathbf{T}^{n}(\mathcal{G}),
𝐓(μ)=lim𝐓n(μ),\displaystyle\mathbf{T}^{\infty}(\mu)=\varprojlim\mathbf{T}^{n}(\mu),\quad 𝐓(κ)=lim𝐓n(κ),𝐓(e)=lim𝐓n(e),\displaystyle\mathbf{T}^{\infty}(\kappa)=\varprojlim\mathbf{T}^{n}(\kappa),\quad\mathbf{T}^{\infty}(e)=\varprojlim\mathbf{T}^{n}(e),

see Section 2.5. Clearly the infinite dimensional supermanifold 𝐓(𝒢)\mathbf{T}^{\infty}(\mathcal{G}) is a Lie supergroup, since the morphisms 𝐓(μ)\mathbf{T}^{\infty}(\mu), 𝐓(κ)\mathbf{T}^{\infty}(\kappa) and 𝐓(e)\mathbf{T}^{\infty}(e) satisfy the group axioms.

Let us describe 𝐓(𝒢)\mathbf{T}^{\infty}(\mathcal{G}) in terms of graded Harish-Chandra pairs. Above we saw that

Tn(𝔤)(ξ1,ξ2,ξn)𝔤andT(𝔤)(ξ1,ξ2,)𝔤,\mathrm{T^{\prime}}^{n}(\mathfrak{g})\simeq\bigwedge(\xi_{1},\xi_{2},\ldots\xi_{n})\otimes\mathfrak{g}\quad\text{and}\quad\mathrm{T^{\prime}}^{\infty}(\mathfrak{g})\simeq\bigwedge(\xi_{1},\xi_{2},\ldots)\otimes\mathfrak{g},

where (ξ1,ξ2,)\bigwedge(\xi_{1},\xi_{2},\ldots) is the Grassmann algebra with infinitely many variables ξ1,ξ2,\xi_{1},\xi_{2},\ldots labeled by natural numbers. We can identify the structure sheaf 𝒪𝐓(𝒢)\mathcal{O}_{\mathbf{T}^{\infty}(\mathcal{G})} of 𝐓(𝒢)\mathbf{T}^{\infty}(\mathcal{G}) with

Hom𝒰(𝔤0¯)(𝒰(𝐓(𝔤)),𝒢0)Hom𝒰(𝔤0¯)(𝒰(𝐓(𝔤)),𝒢0),\mathrm{Hom}^{\prime}_{\mathcal{U}(\mathfrak{g}_{\bar{0}})}(\mathcal{U}(\mathbf{T}^{\infty}(\mathfrak{g})),\mathcal{F}_{\mathcal{G}_{0}})\subset\mathrm{Hom}_{\mathcal{U}(\mathfrak{g}_{\bar{0}})}(\mathcal{U}(\mathbf{T}^{\infty}(\mathfrak{g})),\mathcal{F}_{\mathcal{G}_{0}}),

where Hom\mathrm{Hom}^{\prime} are all 𝒰(𝔤0¯)\mathcal{U}(\mathfrak{g}_{\bar{0}})-homomorphisms that are zero on an ideal generated by

q>0q(ξn,ξn+1,)𝔤𝒰(𝐓(𝔤))\bigoplus_{q>0}\bigwedge^{q}(\xi_{n},\xi_{n+1},\ldots)\otimes\mathfrak{g}\subset\mathcal{U}(\mathbf{T}^{\infty}(\mathfrak{g}))

for some n1n\geq 1.

7.2. Functor split gr\mathrm{gr}, functor parity change π\pi and functor inverse ι\iota

The functor gr\mathrm{gr} for 𝐓(𝒢)\mathbf{T}^{\infty}(\mathcal{G}) is defined as above. More precisely, let 𝒥\mathcal{J} be the sheaf of ideals in 𝒪𝐓(𝒢)\mathcal{O}_{\mathbf{T}^{\infty}(\mathcal{G})} generated by odd elements. We get a filtration in 𝒪𝐓(𝒢)\mathcal{O}_{\mathbf{T}^{\infty}(\mathcal{G})} by the subsheaves 𝒥p\mathcal{J}^{p}, where p0p\geq 0. The corresponded graded sheaf we denote by gr𝒪𝐓(𝒢)\mathrm{gr}\mathcal{O}_{\mathbf{T}^{\infty}(\mathcal{G})}. By definition gr𝒪𝐓(𝒢)\mathrm{gr}\mathcal{O}_{\mathbf{T}^{\infty}(\mathcal{G})} is the structure sheaf of gr(𝐓(𝒢))\mathrm{gr}(\mathbf{T}^{\infty}(\mathcal{G})). In terms of graded Harish-Chandra pairs we have

𝒪gr(𝐓(𝒢))=Hom𝒰(𝔤0¯)(𝒰(grT(𝔤)),𝒢0),\mathcal{O}_{\mathrm{gr}(\mathbf{T}^{\infty}(\mathcal{G}))}=\mathrm{Hom}^{\prime}_{\mathcal{U}(\mathfrak{g}_{\bar{0}})}(\mathcal{U}(\mathrm{gr^{\prime}}\circ\mathrm{T^{\prime}}^{\infty}(\mathfrak{g})),\mathcal{F}_{\mathcal{G}_{0}}),

where Hom\mathrm{Hom}^{\prime} are all 𝒰(𝔤0¯)\mathcal{U}(\mathfrak{g}_{\bar{0}})-homomorphisms that are zero on some ideal generated by

gr(q>0q(ξn,ξn+1,)𝔤)𝒰(grT(𝔤)),n1.\mathrm{gr^{\prime}}\Big{(}\bigoplus_{q>0}\bigwedge^{q}(\xi_{n},\xi_{n+1},\ldots)\otimes\mathfrak{g}\Big{)}\subset\mathcal{U}(\mathrm{gr^{\prime}}\circ\mathrm{T^{\prime}}^{\infty}(\mathfrak{g})),\,\,n\geq 1.

The supermanifold gr(𝐓(𝒢))\mathrm{gr}(\mathbf{T}^{\infty}(\mathcal{G})) is an \infty-fold vector bundle of type Δ\Delta, where Δ\Delta is generated by an odd weight α\alpha and even weights β1,β2,\beta_{1},\beta_{2},\ldots, see Proposition 23. Since grT(𝔤))\mathrm{gr^{\prime}}\circ\mathrm{T^{\prime}}^{\infty}(\mathfrak{g})) is a graded Lie superalgebra of type Δ\Delta, the universal enveloping algebra

𝒰(grT(𝔤))=δα×β1×𝒰(grT(𝔤))δ\mathcal{U}(\mathrm{gr^{\prime}}\circ\mathrm{T^{\prime}}^{\infty}(\mathfrak{g}))=\bigoplus_{\delta\in\mathbb{Z}\alpha\times\mathbb{Z}\beta_{1}\times\cdots}\mathcal{U}(\mathrm{gr^{\prime}}\circ\mathrm{T^{\prime}}^{\infty}(\mathfrak{g}))_{\delta}

is ××\mathbb{Z}\times\mathbb{Z}\times\cdots-graded. We have

(25) 𝒪gr(𝐓(𝒢))=δα×β1×(𝒪gr(𝐓(𝒢)))δ,\mathcal{O}_{\mathrm{gr}(\mathbf{T}^{\infty}(\mathcal{G}))}=\bigoplus_{\delta\in\mathbb{Z}\alpha\times\mathbb{Z}\beta_{1}\times\cdots}(\mathcal{O}_{\mathrm{gr}(\mathbf{T}^{\infty}(\mathcal{G}))})_{\delta},

where

(𝒪gr(𝐓(𝒢)))δ=Hom𝒰(𝔤0¯)(𝒰(grT(𝔤))δ,𝒢0).(\mathcal{O}_{\mathrm{gr}(\mathbf{T}^{\infty}(\mathcal{G}))})_{\delta}=\mathrm{Hom}_{\mathcal{U}(\mathfrak{g}_{\bar{0}})}(\mathcal{U}(\mathrm{gr^{\prime}}\circ\mathrm{T^{\prime}}^{\infty}(\mathfrak{g}))_{\delta},\mathcal{F}_{\mathcal{G}_{0}}).

Note that in this formula we can omit and write simply Hom\mathrm{Hom}. We can use the equality (25) as a definition of the structure sheaf 𝒪gr(𝐓(𝒢))\mathcal{O}_{\mathrm{gr}(\mathbf{T}^{\infty}(\mathcal{G}))}. Since gr\mathrm{gr} is a functor, the structure morphisms in gr(𝐓(𝒢))\mathrm{gr}(\mathbf{T}^{\infty}(\mathcal{G})) are graded as well. In terms of inverse limit we have gr(𝐓(𝒢)=limgr(𝐓n(𝒢))\mathrm{gr}(\mathbf{T}^{\infty}(\mathcal{G})=\varprojlim\mathrm{gr}(\mathbf{T}^{n}(\mathcal{G})).

To define the functor π\pi we change parities in any gr𝐓n(𝒢)\mathrm{gr}\circ\mathbf{T}^{n}(\mathcal{G}). In terms of graded Harish-Chandra pairs we get

𝒪πgr𝐓(𝒢)=Hom𝒰(𝔤0¯)(𝒰(πgrT(𝔤)),𝒢0),\mathcal{O}_{\pi\circ\mathrm{gr}\circ\mathbf{T}^{\infty}(\mathcal{G})}=\mathrm{Hom}^{\prime}_{\mathcal{U}(\mathfrak{g}_{\bar{0}})}(\mathcal{U}(\pi^{\prime}\circ\mathrm{gr^{\prime}}\circ\mathrm{T^{\prime}}^{\infty}(\mathfrak{g})),\mathcal{F}_{\mathcal{G}_{0}}),

where Hom\mathrm{Hom}^{\prime} are all 𝒰(𝔤0¯)\mathcal{U}(\mathfrak{g}_{\bar{0}})-homomorphisms that are zero on some ideal generated by

πgr(q>0q(ξn,ξn+1,)𝔤)𝒰(πgrT(𝔤)),n1.\pi^{\prime}\circ\mathrm{gr^{\prime}}\Big{(}\bigoplus_{q>0}\bigwedge^{q}(\xi_{n},\xi_{n+1},\ldots)\otimes\mathfrak{g}\Big{)}\subset\mathcal{U}(\pi^{\prime}\circ\mathrm{gr^{\prime}}\circ\mathrm{T^{\prime}}^{\infty}(\mathfrak{g})),\,\,n\geq 1.

In terms of inverse limit again we have πgr(𝐓(𝒢)=limπgr(𝐓n(𝒢))\pi\circ\mathrm{gr}(\mathbf{T}^{\infty}(\mathcal{G})=\varprojlim\pi\circ\mathrm{gr}(\mathbf{T}^{n}(\mathcal{G})). The Lie supergroup πgr𝐓(𝒢)\pi\circ\mathrm{gr}\circ\mathbf{T}^{\infty}(\mathcal{G}) is a graded Lie supergroup of type π(Δ)\pi(\Delta), where π(Δ)\pi(\Delta) is the maximal multiplicity free system generated by odd weights α,β1,β2,\alpha,\beta_{1},\beta_{2},\ldots.

A similar idea we use for functor ι\iota. Recall that F:=ιπgr𝐓\mathrm{F}:=\iota\circ\pi\circ\mathrm{gr}\circ\mathbf{T}^{\infty} and we denote Fn:=ιπgr𝐓n1\mathrm{F}_{n}:=\iota\circ\pi\circ\mathrm{gr}\circ\mathbf{T}^{n-1}. First of all we define the graded Lie supergroup Fn(𝒢)\mathrm{F}_{n}(\mathcal{G}) of degree nn (that is of type {0,1,,n}\{0,1,\ldots,n\} with |1|=1¯|1|=\bar{1}) as the \mathbb{Z}-graded Lie supergroup corresponding to the \mathbb{Z}-graded Harish-Chandra pair (𝒢0,Fn(𝔤)).(\mathcal{G}_{0},\mathrm{F}^{\prime}_{n}(\mathfrak{g})). For any nn we have a natural homomorphism Fn+1(𝔤)Fn(𝔤).\mathrm{F}^{\prime}_{n+1}(\mathfrak{g})\to\mathrm{F}^{\prime}_{n}(\mathfrak{g}). This induces a homomorphism of enveloping algebras

Φn:𝒰(Fn+1(𝔤))𝒰(Fn(𝔤)).\Phi_{n}:\mathcal{U}(\mathrm{F}^{\prime}_{n+1}(\mathfrak{g}))\to\mathcal{U}(\mathrm{F}^{\prime}_{n}(\mathfrak{g})).

Therefore we have a natural map of sheaves

𝒪Fn(𝒢)=Hom𝒰(𝔤0¯)(𝒰(Fn(𝔤)),𝒢0)𝒪Fn+1(𝒢)=Hom𝒰(𝔤0¯)(𝒰(Fn+1(𝔤)),𝒢0)\mathcal{O}_{\mathrm{F}_{n}(\mathcal{G})}=\mathrm{Hom}_{\mathcal{U}(\mathfrak{g}_{\bar{0}})}(\mathcal{U}(\mathrm{F}^{\prime}_{n}(\mathfrak{g})),\mathcal{F}_{\mathcal{G}_{0}})\hookrightarrow\mathcal{O}_{\mathrm{F}_{n_{+}1}(\mathcal{G})}=\mathrm{Hom}_{\mathcal{U}(\mathfrak{g}_{\bar{0}})}(\mathcal{U}(\mathrm{F}^{\prime}_{n+1}(\mathfrak{g})),\mathcal{F}_{\mathcal{G}_{0}})

Now we can identify the structure sheaf of F(𝒢)\mathrm{F}(\mathcal{G}) with

𝒪F(𝒢)=Hom𝒰(𝔤0¯)(𝒰(F(𝔤)),𝒢0),\mathcal{O}_{\mathrm{F}(\mathcal{G})}=\mathrm{Hom}^{\prime}_{\mathcal{U}(\mathfrak{g}_{\bar{0}})}(\mathcal{U}(\mathrm{F}^{\prime}(\mathfrak{g})),\mathcal{F}_{\mathcal{G}_{0}}),

where Hom\mathrm{Hom}^{\prime} are all 𝒰(𝔤0¯)\mathcal{U}(\mathfrak{g}_{\bar{0}})-homomorphisms that are zero on some Ker(Φn)\mathrm{Ker}(\Phi_{n}). Further we see that 𝒪F(𝒢)\mathcal{O}_{\mathrm{F}(\mathcal{G})} is \mathbb{Z}-graded. Indeed,

𝒪F(𝒢)=n(𝒪F(𝒢))n=nHom𝒰(𝔤0¯)(𝒰(F(𝔤))n,𝒢0).\mathcal{O}_{\mathrm{F}(\mathcal{G})}=\bigoplus_{n\in\mathbb{Z}}(\mathcal{O}_{\mathrm{F}(\mathcal{G})})_{n}=\bigoplus_{n\in\mathbb{Z}}\mathrm{Hom}_{\mathcal{U}(\mathfrak{g}_{\bar{0}})}(\mathcal{U}(\mathrm{F}^{\prime}(\mathfrak{g}))_{n},\mathcal{F}_{\mathcal{G}_{0}}).

The graded Lie supergroup morphisms can be defined by formulas (6) or using inverse limit. In terms of inverse limit again we have F(𝒢)=limFn(𝒢))\mathrm{F}(\mathcal{G})=\varprojlim\mathrm{F}_{n}(\mathcal{G})).

8. Coverings and semicoverings of a Lie superalgebra and
a Lie supergroup

In this section we give a definition of a covering and a semicovering of a Lie superalgebra and a Lie supergroup. Further we show that the generalized Donagi–Witten construction leads to a construction of a covering and semicovering spaces of a Lie superalgebra and a Lie supergroup. The case of any supermanifold will be considered in [RV].

8.1. Coverings and semicoverings of a Lie superalgebra

We start with Lie superalgebras. Throughout this subsection we fix a surjective homomorphism ϕ:AB\phi:A\to B of abelian groups.

8.1.1. A ϕ\phi-covering of a BB-graded Lie superalgebra along a homomorphism ϕ:AB\phi:A\to B

Definition 46.

A ϕ\phi-covering of a BB-graded Lie superalgebra 𝔤\mathfrak{g} along a surjective homomorphism ϕ:AB\phi:A\to B of abelian groups is an AA-graded superalgebra 𝔭=αA𝔭α\mathfrak{p}=\oplus_{\alpha\in A}\mathfrak{p}_{\alpha} together with a homomorphism Π:𝔭𝔤\Pi^{\prime}:\mathfrak{p}\to\mathfrak{g} such that Π|𝔭a:𝔭a𝔤ϕ(a)\Pi^{\prime}|_{\mathfrak{p}_{a}}:\mathfrak{p}_{a}\to\mathfrak{g}_{\phi(a)} is a linear bijection for any αA\alpha\in A.

Note that the bracket on 𝔭\mathfrak{p} is fully determined by the bracket on 𝔤\mathfrak{g}. Indeed, for X𝔭αX\in\mathfrak{p}_{\alpha}, X𝔭αX^{\prime}\in\mathfrak{p}_{\alpha^{\prime}} we have Π([X,X]𝔭)=[Π(X),Π(X)]𝔤𝔤ϕ(α+α)\Pi^{\prime}([X,X^{\prime}]_{\mathfrak{p}})=[\Pi^{\prime}(X),\Pi^{\prime}(X^{\prime})]_{\mathfrak{g}}\in\mathfrak{g}_{\phi(\alpha+\alpha^{\prime})}, hence

[X,Y]𝔭=(Πα+α|𝔭α+α)1([Π(X),Π(X)]𝔤).[X,Y]_{\mathfrak{p}}=(\Pi^{\prime}_{\alpha+\alpha^{\prime}}|_{\mathfrak{p}_{\alpha+\alpha^{\prime}}})^{-1}([\Pi^{\prime}(X),\Pi^{\prime}(X^{\prime})]_{\mathfrak{g}}).

Also Πα=Π|𝔭α:𝔭α𝔤ϕ(α)\Pi^{\prime}_{\alpha}=\Pi^{\prime}|_{\mathfrak{p}_{\alpha}}:\mathfrak{p}_{\alpha}\to\mathfrak{g}_{\phi(\alpha)} is a 𝔤0\mathfrak{g}_{0}-module map, where we identify the Lie superalgebras 𝔭0\mathfrak{p}_{0} and 𝔤0\mathfrak{g}_{0} via Π0\Pi_{0}.

Proposition 47.

Let 𝔞\mathfrak{a}, 𝔤\mathfrak{g} be an AA- and a BB-graded Lie superalgebras, respectively. Let ψ:𝔞𝔤\psi:\mathfrak{a}\to\mathfrak{g} be a BB-graded homomorphism of Lie superalgebras111Any AA-graded Lie superalgebra is automatically BB-graded., and let 𝔭\mathfrak{p} be a covering of 𝔤\mathfrak{g} along ϕ:AB\phi:A\to B. Then there exists a unique AA-graded homomorphism Ψ:𝔞𝔭\Psi:\mathfrak{a}\to\mathfrak{p} such that the following diagram is commutative

𝔭{\mathfrak{p}}𝔞{\mathfrak{a}}𝔤{\mathfrak{g}}Π\scriptstyle{\Pi^{\prime}}!Ψ\scriptstyle{\exists!\Psi}ψ\scriptstyle{\psi}
Proof.

We define Ψ\Psi as a linear map such that Ψ(𝔞s)𝔭s\Psi(\mathfrak{a}_{s})\subset\mathfrak{p}_{s} for any sAs\in A and such that ΠΨ=ψ\Pi^{\prime}\circ\Psi=\psi. Let us check that Ψ\Psi is a homomorphism. Indeed, let us take X𝔞sX\in\mathfrak{a}_{s} and Y𝔞tY\in\mathfrak{a}_{t}. Then

Π([Ψ(X),Ψ(Y)])=[ΠΨ(X),ΠΨ(Y)]=[ψ(X),ψ(Y)]=\displaystyle\Pi^{\prime}([\Psi(X),\Psi(Y)])=[\Pi^{\prime}\circ\Psi(X),\Pi^{\prime}\circ\Psi(Y)]=[\psi(X),\psi(Y)]=
ψ([X,Y])=ΠΨ([X,Y]).\displaystyle\psi([X,Y])=\Pi^{\prime}\circ\Psi([X,Y]).

Since by definition of Ψ\Psi both Ψ([X,Y])\Psi([X,Y]) and [Ψ(X),Ψ(Y)][\Psi(X),\Psi(Y)] are in 𝔭s+t\mathfrak{p}_{s+t} and Π\Pi^{\prime} is locally bijective, we get the equality [Ψ(X),Ψ(Y)]=Ψ([X,Y])[\Psi(X),\Psi(Y)]=\Psi([X,Y]). ∎

Proposition 48.

Let f:𝔤𝔤~f:\mathfrak{g}\to\tilde{\mathfrak{g}} be a homomorphism of BB-graded Lie superalgebras. Then there exists unique homomorphism FF of ϕ\phi-coverings 𝔭\mathfrak{p} and 𝔭~\tilde{\mathfrak{p}} such that the following diagram is commutative

𝔭{\mathfrak{p}}𝔭~{\tilde{\mathfrak{p}}}𝔤{\mathfrak{g}}𝔤~{\tilde{\mathfrak{g}}}F\scriptstyle{F}Π\scriptstyle{\Pi^{\prime}}Π~\scriptstyle{\tilde{\Pi}^{\prime}}f\scriptstyle{f}
Proof.

It follows immediately from Proposition 47, just take ψ=fΠ\psi=f\circ\Pi^{\prime}. ∎

From Proposition 48 it follows that ϕ\phi-coverings are unique up to isomorphism.

8.1.2. A ϕ\phi-covering and ϕ\phi-semicovering with support CC

Let AA and BB be abelian groups, ϕ:AB\phi:A\to B be a surjective homomorphism and CAC\subset A be a subset.

Definition 49.

A ϕ\phi-covering with support CC of a BB-graded Lie superalgebra 𝔤\mathfrak{g} along a surjective homomorphism ϕ:AB\phi:A\to B is an AA-graded superalgebra 𝔭=αA𝔭α\mathfrak{p}=\oplus_{\alpha\in A}\mathfrak{p}_{\alpha} with supp(𝔭)=Csupp(\mathfrak{p})=C such that ϕ(C)=B\phi(C)=B together with a surjective homomorphism Π:𝔭𝔤\Pi^{\prime}:\mathfrak{p}\to\mathfrak{g} such that Π|𝔭a:𝔭a𝔤ϕ(a)\Pi^{\prime}|_{\mathfrak{p}_{a}}:\mathfrak{p}_{a}\to\mathfrak{g}_{\phi(a)} is a linear bijection for any αC\alpha\in C.

Let AA, BB, CC and ϕ\phi be as above, 𝔤\mathfrak{g} and 𝔤\mathfrak{g}^{\prime} be a AA-graded and BB-graded Lie superalgebra, respectively.

Definition 50.

A map Ψ:𝔤𝔤\Psi:\mathfrak{g}\to\mathfrak{g}^{\prime} is called a partial homomorphism if Ψ([X,Y])=[Ψ(Y),Ψ(Y)]\Psi([X,Y])=[\Psi(Y),\Psi(Y)] for any X𝔤αX\in\mathfrak{g}_{\alpha}, Y𝔤βY\in\mathfrak{g}_{\beta} such that α\alpha, β\beta and α+β\alpha+\beta are in CC.

Definition 51.

A ϕ\phi-semicovering with support CC of a BB-graded Lie superalgebra 𝔤\mathfrak{g} along a surjective homomorphism ϕ:AB\phi:A\to B is an AA-graded Lie superalgebra 𝔭=αC𝔭α\mathfrak{p}=\oplus_{\alpha\in C}\mathfrak{p}_{\alpha} with supp(𝔭)=Csupp(\mathfrak{p})=C such that ϕ(C)=B\phi(C)=B together with a surjective partial homomorphism Π:𝔭𝔤\Pi^{\prime}:\mathfrak{p}\to\mathfrak{g} such that Π|𝔭a:𝔭a𝔤ϕ(a)\Pi^{\prime}|_{\mathfrak{p}_{a}}:\mathfrak{p}_{a}\to\mathfrak{g}_{\phi(a)} is a linear bijection for any αC\alpha\in C.

For a ϕ\phi-covering and ϕ\phi-semicovering with support CC of a BB-graded Lie superalgebra 𝔤\mathfrak{g} we can prove analogues of Propositions 47 and 48.

Proposition 52.

(1) Let 𝔞\mathfrak{a}, 𝔤\mathfrak{g} be an AA- and BB-graded Lie superalgebras, respectively, and supp(𝔞)=Csupp(\mathfrak{a})=C. Let ψ:𝔞𝔤\psi:\mathfrak{a}\to\mathfrak{g} be a BB-graded homomorphism of Lie superalgebras, and let 𝔭\mathfrak{p} be a ϕ\phi-covering (or ϕ\phi-semicovering) with support CC of 𝔤\mathfrak{g}. Then there exists a unique AA-graded homomorphism Ψ:𝔞𝔭\Psi:\mathfrak{a}\to\mathfrak{p} such that ψ=ΠΨ\psi=\Pi^{\prime}\circ\Psi.

(2) Let f:𝔤𝔤~f:\mathfrak{g}\to\tilde{\mathfrak{g}} be a homomorphism of BB-graded Lie superalgebras. Then there exists unique homomorphism FF of ϕ\phi-coverings (or ϕ\phi-semicoverings) 𝔭\mathfrak{p} and 𝔭~\tilde{\mathfrak{p}} with support CC such that fΠ=Π~Ff\circ\Pi^{\prime}=\tilde{\Pi}^{\prime}\circ F.

Proof.

We define Ψ\Psi as a linear map such that Ψ(𝔞s)𝔭s\Psi(\mathfrak{a}_{s})\subset\mathfrak{p}_{s} for any sCs\in C and such that ΠΨ=ψ\Pi^{\prime}\circ\Psi=\psi. Now we just repeat arguments of the proofs of Propositions 47 and 48. One non-trivial point is the proof that Ψ\Psi is a homomorphism in the case of a semicovering. We have for X𝔞sX\in\mathfrak{a}_{s} and Y𝔞tY\in\mathfrak{a}_{t}, where s,t,s+tCs,t,s+t\in C,

Π([Ψ(X),Ψ(Y)])=[ΠΨ(X),ΠΨ(Y)]=[ψ(X),ψ(Y)]=\displaystyle\Pi^{\prime}([\Psi(X),\Psi(Y)])=[\Pi^{\prime}\circ\Psi(X),\Pi^{\prime}\circ\Psi(Y)]=[\psi(X),\psi(Y)]=
ψ([X,Y])=ΠΨ([X,Y]).\displaystyle\psi([X,Y])=\Pi^{\prime}\circ\Psi([X,Y]).

Since by definition of Ψ\Psi both Ψ([X,Y])\Psi([X,Y]) and [Ψ(X),Ψ(Y)][\Psi(X),\Psi(Y)] are in 𝔭s+t\mathfrak{p}_{s+t} and Π\Pi^{\prime} is locally bijective, we get the equality [Ψ(X),Ψ(Y)]=Ψ([X,Y])[\Psi(X),\Psi(Y)]=\Psi([X,Y]). In the case s,tCs,t\in C, but s+tCs+t\notin C, we have

[Ψ(X),Ψ(Y)]=Ψ([X,Y])=0.[\Psi(X),\Psi(Y)]=\Psi([X,Y])=0.

From Proposition 52 it follows that ϕ\phi-coverings (and ϕ\phi-semicoverings) with support CC are unique up to isomorphism.

8.2. A covering and a semicovering of a Lie supergroup

Unlike the notion of a covering of a Lie superalgebra, as far as we know, the notion of a covering of a Lie supergroup was never considered in the literature before. One possible way to give a definition of a covering of a Lie supergroup is to use the graded covering of the corresponding Lie superalgebra. However we suggest a different way that is closer to the notion of a topological covering space and this approach can be used to define a covering for any supermanifold, see [RV].

We start with a definition of a semicovering spaces for a Lie supergroup. In Section 9 we will show that for any Lie supergroup 𝒢\mathcal{G} the image F(𝒢)\mathrm{F}(\mathcal{G}) is a (2¯)(\mathbb{Z}\to\mathbb{Z}_{\bar{2}})-covering of 𝒢\mathcal{G} with support {0,1,2,}\{0,1,2,\ldots\}. In Section 10 we will give a simple explicit construction of a covering of any matrix Lie supergroup. Note that in general such a simple construction for a Lie supergroup is not applicable for the case of a supermanifold. Recall that if \mathcal{M} is a supermanifold, we denote by 𝒪\mathcal{O}_{\mathcal{M}} its structure sheaf and by 0\mathcal{F}_{\mathcal{M}_{0}} the structure sheaf of the underlying space 0\mathcal{M}_{0}.

8.2.1. A semicovering of a Lie supergroup

Let 𝒫\mathcal{P} be a graded Lie supergroup of degree nn (or equivalently of type C={0,1,,n}C=\{0,1,\ldots,n\} with |1|=1¯|1|=\bar{1}) with multiplication morphism ν\nu and 𝒢\mathcal{G} be a Lie supergroup with multiplication morphism μ\mu. Assume in addition that 𝒫0=𝒢0\mathcal{P}_{0}=\mathcal{G}_{0}.

Definition 53.

A sum of nn morphisms Πn=c=0nΠc\Pi^{n}=\bigoplus\limits_{c=0}^{n}\Pi_{c}, where

Πc=(id,Πc):(𝒢0,(𝒪𝒫)c)(𝒢0,𝒪𝒢),cC,\Pi_{c}=(id,\Pi^{*}_{c}):(\mathcal{G}_{0},(\mathcal{O}_{\mathcal{P}})_{c})\to(\mathcal{G}_{0},\mathcal{O}_{\mathcal{G}}),\,\,c\in C,

and Πc\Pi^{*}_{c} is a morphism of sheaves of vector spaces, is called a partial homomorphism of 𝒫\mathcal{P} to 𝒢\mathcal{G} with support CC if

Πc(fg)=c1+c2=sΠc1(f)Πc2(g),f,g𝒪𝒢,c,ciC,\Pi^{*}_{c}(f\cdot g)=\sum_{c_{1}+c_{2}=s}\Pi^{*}_{c_{1}}(f)\cdot\Pi^{*}_{c_{2}}(g),\quad f,g\in\mathcal{O}_{\mathcal{G}},\,\,c,c_{i}\in C,

and

νΠc=c1+c2=c(Πc1×Πc2)μfor any c,ciC.\nu^{*}\circ\Pi^{*}_{c}=\sum_{c_{1}+c_{2}=c}(\Pi^{*}_{c_{1}}\times\Pi^{*}_{c_{2}})\circ\mu^{*}\quad\text{for any $c,c_{i}\in C$.}

Sometimes we will write a partial homomorphism Πn\Pi^{n} of 𝒫\mathcal{P} to 𝒢\mathcal{G} in the form Πn:𝒫𝒢\Pi^{n}:\mathcal{P}\to\mathcal{G}. Note that the sheaf morphism (Πn):𝒪𝒢𝒪𝒫(\Pi^{n})^{*}:\mathcal{O}_{\mathcal{G}}\to\mathcal{O}_{\mathcal{P}} is in general not defined.

Definition 54.

A ϕ:2\phi:\mathbb{Z}\to\mathbb{Z}_{2}-semicovering with support C={0,1,2,,n}C=\{0,1,2,\ldots,n\} of a Lie supergroup 𝒢\mathcal{G} is a graded Lie supergroup 𝒫\mathcal{P} of degree nn with 𝒫0=𝒢0\mathcal{P}_{0}=\mathcal{G}_{0} together with a partial homomorphism Πn:𝒫𝒢\Pi^{n}:\mathcal{P}\to\mathcal{G} such that we can choose atlases {𝒰i}\{\mathcal{U}_{i}\} and {𝒱i}\{\mathcal{V}_{i}\} on 𝒢\mathcal{G} and 𝒫\mathcal{P}, respectively, with the same base space (𝒰i)0=(𝒱i)0(\mathcal{U}_{i})_{0}=(\mathcal{V}_{i})_{0}, with even and odd coordinates (xa,ξb)(x_{a},\xi_{b}) in 𝒰i\mathcal{U}_{i} and with graded coordinates (yas,ηbt)(y_{a}^{s},\eta^{t}_{b}), where sCs\in C is an even integer and tCt\in C is an odd integer, in 𝒱i\mathcal{V}_{i} such that

Πs(xa)=yas,Πt(ξb)=ηbt.\Pi^{*}_{s}(x_{a})=y_{a}^{s},\quad\Pi^{*}_{t}(\xi_{b})=\eta^{t}_{b}.

8.2.2. A covering of a Lie supergroup

Denote by 𝒫=lim𝒫n\mathcal{P}=\varprojlim\mathcal{P}_{n} the inverse limit of graded Lie supergroups 𝒫n\mathcal{P}_{n} of degree nn with the same underlying space (𝒫n)0=𝒢0(\mathcal{P}_{n})_{0}=\mathcal{G}_{0}.

Definition 55.

A ϕ:2\phi:\mathbb{Z}\to\mathbb{Z}_{2}-covering with support C={0,1,2,,}C=\{0,1,2,\ldots,\} of a Lie supergroup 𝒢\mathcal{G} is a Lie supergroup 𝒫=lim𝒫n\mathcal{P}=\varprojlim\mathcal{P}_{n} together with a Lie supergroup homomorphism Π=lim(Πk):𝒫𝒢\Pi=\varprojlim(\Pi^{k}):\mathcal{P}\to\mathcal{G} such that Πk:𝒫k𝒢\Pi^{k}:\mathcal{P}_{k}\to\mathcal{G} is a semicovering with support C={0,1,2,,k}C=\{0,1,2,\ldots,k\} for any k0k\geq 0.

Remark 56.

Definition 55 implies that in the case of a ϕ:2\phi:\mathbb{Z}\to\mathbb{Z}_{2}-covering Π:𝒫𝒢\Pi:\mathcal{P}\to\mathcal{G} with support C={0,1,2,,}C=\{0,1,2,\ldots,\} we can choose ”atlases” {𝒰i}\{\mathcal{U}_{i}\} and {𝒱i}\{\mathcal{V}_{i}\} on 𝒢\mathcal{G} and 𝒫\mathcal{P}, respectively, with the same base space (𝒰i)0=(𝒱i)0(\mathcal{U}_{i})_{0}=(\mathcal{V}_{i})_{0}, with even and odd coordinates (xa,ξb)(x_{a},\xi_{b}) in 𝒰i\mathcal{U}_{i} and with graded coordinates (yas,ηbt)(y_{a}^{s},\eta^{t}_{b}), where ss is an even integer and tt is an odd integer, in 𝒱i\mathcal{V}_{i} such that

prsΠ(xa)=yas,prtΠ(ξb)=ηbt,pr_{s}\circ\Pi^{*}(x_{a})=y_{a}^{s},\quad pr_{t}\circ\Pi^{*}(\xi_{b})=\eta^{t}_{b},

where prq:𝒪𝒫(𝒪𝒫)qpr_{q}:\mathcal{O}_{\mathcal{P}}\to(\mathcal{O}_{\mathcal{P}})_{q}, qq\in\mathbb{Z}, is the natural projection. In this case each 𝒱i\mathcal{V}_{i} is an ”infinite graded domain”, which we understand as inverse limit of the corresponding finite graded domains.

Sometimes we will call a ϕ:2\phi:\mathbb{Z}\to\mathbb{Z}_{2}-covering with support C={0,1,2,,}C=\{0,1,2,\ldots,\} of a Lie supergroup 𝒢\mathcal{G} simply a 0\mathbb{Z}^{\geq 0}-covering of 𝒢\mathcal{G}.

Remark 57.

It looks more natural to give a definition of a covering space with support \mathbb{Z}. The study of \mathbb{Z}-graded Lie supergroups (not necessary non-negatively \mathbb{Z}-graded) was announced in [KPS]. In the case of any \mathbb{Z}-graduation graded coordinates may be not nilpotent, this leads to significant difficulties, see [CGP]. Therefore a definition of a covering space with support \mathbb{Z} we leave for a future research.

Remark 58.

Our definition implies that in some sense 𝒫\mathcal{P} is locally diffeomorphic to 𝒢\mathcal{G}. Indeed, let us choose qq\in\mathbb{Z} assuming for example that qq is even. Then the sheaf morphism prsΠpr_{s}\circ\Pi^{*} determines a sheaf isomorphism of S(xa)|(𝒰i)0𝒢0S^{*}(x_{a})|_{(\mathcal{U}_{i})_{0}}\subset\mathcal{F}_{\mathcal{G}_{0}} and S(yas)|(𝒰i)0S^{*}(y^{s}_{a})|_{(\mathcal{U}_{i})_{0}}. Here S(za)S^{*}(z_{a}) is the supersymmetric algebra generated by zaz_{a}. And similarly in the case of odd coordinates. In other words we have the following isomorphism of superdomains

((𝒰i)0,S(xa))((𝒱i)0,S(yas)),((𝒰i)0,S(ξb))((𝒱i)0,S(ηbt)).\displaystyle((\mathcal{U}_{i})_{0},S^{*}(x_{a}))\simeq((\mathcal{V}_{i})_{0},S^{*}(y^{s}_{a})),\quad((\mathcal{U}_{i})_{0},S^{*}(\xi_{b}))\simeq((\mathcal{V}_{i})_{0},S^{*}(\eta^{t}_{b})).

We say that a Lie supergroup 𝒢\mathcal{G} possesses an additional non-negative \mathbb{Z}-grading if its structure sheaf possesses a \mathbb{Z}-grading 𝒪𝒢=q0(𝒪𝒢)q\mathcal{O}_{\mathcal{G}}=\bigoplus\limits_{q\geq 0}(\mathcal{O}_{\mathcal{G}})_{q} and all Lie supergroup morphisms are \mathbb{Z}-graded. In this paper we assume in addition that (𝒪𝒢)q(𝒪𝒢)q¯(\mathcal{O}_{\mathcal{G}})_{q}\subset(\mathcal{O}_{\mathcal{G}})_{\bar{q}}. Note that a Lie supergroup with a \mathbb{Z}-grading is not the same as a graded Lie supergroup of degree nn. Now we can prove that our covering satisfies the same universal properties as a covering for a Lie superalgebra.

Theorem 59.

Let 𝒢\mathcal{G} be a Lie supergroup with an additional non-negative \mathbb{Z}-grading or a graded Lie supergroup of degree nn and 𝒢\mathcal{G}^{\prime} be a Lie supergroup. Let ψ:𝒢𝒢\psi:\mathcal{G}\to\mathcal{G}^{\prime} be a Lie supergroup homomorphism and let 𝒫\mathcal{P}^{\prime} be a 0\mathbb{Z}^{\geq 0}-covering of 𝒢\mathcal{G}^{\prime}. Then there exists a unique homomorphism of Lie supergroups Ψ:𝒢𝒫\Psi:\mathcal{G}\to\mathcal{P}^{\prime}, which preserves the \mathbb{Z}-gradings, such that the following diagram is commutative

𝒫{\mathcal{P}^{\prime}}𝒢{\mathcal{G}}𝒢{\mathcal{G}^{\prime}}Π\scriptstyle{\Pi}!Ψ\scriptstyle{\exists!\Psi}ψ\scriptstyle{\psi}
Proof.

We split the proof into steps.

Step 1. We put Wi:=ψ1((𝒰i)0)W_{i}:=\psi^{-1}((\mathcal{U}_{i})_{0}). Let us define first (ΨWi)(\Psi_{W_{i}})^{*} for any ii using coordinates from Definition 54. We put

(ΨWi)(yas):=prsψ(xa)(𝒪𝒢)s,(ΨWi)(ηbt):=prtψ(ξb)(𝒪𝒢)t,(\Psi_{W_{i}})^{*}(y_{a}^{s}):=pr_{s}\circ\psi^{*}(x_{a})\in(\mathcal{O}_{\mathcal{G}})_{s},\quad(\Psi_{W_{i}})^{*}(\eta_{b}^{t}):=pr_{t}\circ\psi^{*}(\xi_{b})\in(\mathcal{O}_{\mathcal{G}})_{t},

where prq:𝒪𝒢(𝒪𝒢)qpr_{q}:\mathcal{O}_{\mathcal{G}}\to(\mathcal{O}_{\mathcal{G}})_{q}. By [L, Section 2.1.7, Theorema] we defined a morphism ΨWi:(Wi,𝒪𝒫)𝒱i\Psi_{W_{i}}:(W_{i},\mathcal{O}_{\mathcal{P}^{\prime}})\to\mathcal{V}_{i} of superdomains.

Step 2. The morphism ΨWi\Psi_{W_{i}} satisfies the following equality

(26) (ΨWi)(prsΠ(F))=prsψ(F),(\Psi_{W_{i}})^{*}(pr_{s}\circ\Pi^{*}(F))=pr_{s}\circ\psi^{*}(F),

where ss\in\mathbb{Z} and F𝒪𝒢|(𝒰i)0F\in\mathcal{O}_{\mathcal{G}^{\prime}}|_{(\mathcal{U}_{i})_{0}}. First of all assume that FF is a polynomial in (xa,ξb)(x_{a},\xi_{b}). It is sufficient to assume that

F=xa1xakξb1ξblF=x_{a_{1}}\cdots x_{a_{k}}\xi_{b_{1}}\cdots\xi_{b_{l}}

is a monomial. By Definition 55 we have

Π(xa)=s=2qyas,Π(ξb)=t=2q+1ηbt.\Pi^{*}(x_{a})=\sum_{s=2q}y^{s}_{a},\quad\Pi^{*}(\xi_{b})=\sum_{t=2q+1}\eta^{t}_{b}.

(Here for simplicity of notations we write infinite sums. We understand this sum as an inverse limit.) Further, we get

prsΠ(xa1xakξb1ξbl)=isi+jtj=sya1s1yakskηb1t1ηbltl.\displaystyle pr_{s}\circ\Pi^{*}(x_{a_{1}}\cdots x_{a_{k}}\xi_{b_{1}}\cdots\xi_{b_{l}})=\sum_{\sum_{i}s_{i}+\sum_{j}t_{j}=s}y^{s_{1}}_{a_{1}}\cdots y^{s_{k}}_{a_{k}}\eta^{t_{1}}_{b_{1}}\cdots\eta^{t_{l}}_{b_{l}}.

Therefore,

(ΨWi)(prsΠ(F))\displaystyle(\Psi_{W_{i}})^{*}(pr_{s}\circ\Pi^{*}(F)) =\displaystyle=
isi+jtj=s\displaystyle\sum_{\sum_{i}s_{i}+\sum_{j}t_{j}=s} (ΨWi)(ya1s1)(ΨWi)(yaksk)(ΨWi)(ηb1t1)(ΨWi)(ηbltl).\displaystyle(\Psi_{W_{i}})^{*}(y^{s_{1}}_{a_{1}})\cdots(\Psi_{W_{i}})^{*}(y^{s_{k}}_{a_{k}})(\Psi_{W_{i}})^{*}(\eta^{t_{1}}_{b_{1}})\cdots(\Psi_{W_{i}})^{*}(\eta^{t_{l}}_{b_{l}}).

On the other hand,

prsψ(F)\displaystyle pr_{s}\circ\psi^{*}(F) =prsψ(xa1xakξb1ξbl)=\displaystyle=pr_{s}\circ\psi^{*}(x_{a_{1}}\cdots x_{a_{k}}\xi_{b_{1}}\cdots\xi_{b_{l}})=
isi+jtj=s\displaystyle\sum_{\sum_{i}s_{i}+\sum_{j}t_{j}=s} prs1ψ(xa1)prskψ(xak)prt1ψ(ξb1)prtlψ(ξbl).\displaystyle pr_{s_{1}}\circ\psi^{*}(x_{a_{1}})\cdots pr_{s_{k}}\circ\psi^{*}(x_{a_{k}})pr_{t_{1}}\circ\psi^{*}(\xi_{b_{1}})\cdots pr_{t_{l}}\circ\psi^{*}(\xi_{b_{l}}).

Now the result follows from the definition of ΨWi\Psi_{W_{i}}. If morphisms coincide on all polynomials, a standard argument, see [L], implies the result for any functions F𝒪𝒢F\in\mathcal{O}_{\mathcal{G}^{\prime}}.

Step 3. We have to show that ΨWi=ΨWj\Psi_{W_{i}}=\Psi_{W_{j}} in WiWjW_{i}\cap W_{j}. Let x~a=H(xa,ξb)\tilde{x}_{a}=H(x_{a},\xi_{b}) in 𝒰i𝒰j\mathcal{U}_{i}\cap\mathcal{U}_{j} and y~as=prsψ(xa)\tilde{y}^{s}_{a}=pr_{s}\circ\psi^{*}(x_{a}). We have by Step 2

(ΨWj)(y~as)=(ΨWj)(prsΠ(x~a))=(ΨWj)(prsΠ(H(xa,ξb)))=\displaystyle(\Psi_{W_{j}})^{*}(\tilde{y}^{s}_{a})=(\Psi_{W_{j}})^{*}(pr_{s}\circ\Pi^{*}(\tilde{x}_{a}))=(\Psi_{W_{j}})^{*}(pr_{s}\circ\Pi^{*}(H(x_{a},\xi_{b})))=
prsψ(H(xa,ξb))=(ΨWi)(prsΠ(H(xa,ξb))).\displaystyle pr_{s}\circ\psi^{*}(H(x_{a},\xi_{b}))=(\Psi_{W_{i}})^{*}(pr_{s}\circ\Pi^{*}(H(x_{a},\xi_{b}))).

Now we can define the morphism Ψ\Psi by Ψ|Wi:=ΨWi\Psi|_{W_{i}}:=\Psi_{W_{i}}.

Step 4. Let us prove that Ψ\Psi is a homomorphism of Lie supergroups. Denote by μ𝒢\mu_{\mathcal{G}}, μ𝒢\mu_{\mathcal{G}^{\prime}} and by μ𝒫\mu_{\mathcal{P}} the multiplication morphism in 𝒢\mathcal{G}, 𝒢\mathcal{G} and 𝒫\mathcal{P}, respectively. We have to show that

(27) (Ψ×Ψ)μ𝒫=μ𝒢Ψ.(\Psi^{*}\times\Psi^{*})\circ\mu_{\mathcal{P}}^{*}=\mu_{\mathcal{G}}^{*}\circ\Psi^{*}.

Clearly it is sufficient to show (27) only for coordinates (yas,ηbt)(y_{a}^{s},\eta_{b}^{t}). Step 22 implies that

ΨprsΠ=prsψ.\Psi^{*}\circ pr_{s}\circ\Pi^{*}=pr_{s}\circ\psi^{*}.

We have

μ𝒢Ψ(yas)=μ𝒢ΨprsΠ(xa)=μ𝒢prsψ(xa)=prsμ𝒢ψ(xa)=\displaystyle\mu_{\mathcal{G}}^{*}\circ\Psi^{*}(y^{s}_{a})=\mu_{\mathcal{G}}^{*}\circ\Psi^{*}\circ pr_{s}\circ\Pi^{*}(x_{a})=\mu_{\mathcal{G}}^{*}\circ pr_{s}\circ\psi^{*}(x_{a})=pr_{s}\circ\mu_{\mathcal{G}}^{*}\circ\psi^{*}(x_{a})=
prs(ψ×ψ)μ𝒢(xa)=(Ψ×Ψ)prs(Π×Π)μ𝒢(xa)=\displaystyle pr_{s}\circ(\psi^{*}\times\psi^{*})\circ\mu_{\mathcal{G}^{\prime}}^{*}(x_{a})=(\Psi^{*}\times\Psi^{*})\circ pr_{s}\circ(\Pi^{*}\times\Pi^{*})\circ\mu_{\mathcal{G}^{\prime}}^{*}(x_{a})=
(Ψ×Ψ)prsμ𝒢Π(xa)=(Ψ×Ψ)μ𝒢prsΠ(xa)=(Ψ×Ψ)μ𝒢(yas).\displaystyle(\Psi^{*}\times\Psi^{*})\circ pr_{s}\circ\mu_{\mathcal{G}}^{*}\circ\Pi^{*}(x_{a})=(\Psi^{*}\times\Psi^{*})\circ\mu_{\mathcal{G}}^{*}\circ pr_{s}\circ\Pi^{*}(x_{a})=(\Psi^{*}\times\Psi^{*})\circ\mu_{\mathcal{G}}^{*}(y^{s}_{a}).

For ηbt\eta_{b}^{t} the proof is similar. The proof is complete. ∎

Corollary 60.

If a 0\mathbb{Z}^{\geq 0}-covering of a Lie supergroup 𝒢\mathcal{G} exists, it is unique up to isomorphism.

Proof.

In Theorem 59 we assume that 𝒢\mathcal{G} is finite dimensional. However the same argument as in the proof of Theorem 59 implies that we can replace 𝒢\mathcal{G} by a 0\mathbb{Z}^{\geq 0}-covering of 𝒢\mathcal{G}. Therefore we get the result. ∎

If there exists a 0\mathbb{Z}^{\geq 0}-covering Π:𝒫𝒢\Pi:\mathcal{P}\to\mathcal{G} of a Lie supergroup 𝒢\mathcal{G}, then from Definition 55 it follows that there exists a ϕ:2\phi:\mathbb{Z}\to\mathbb{Z}_{2}-semicovering of 𝒢\mathcal{G} with support C={0,1,2,,n}C=\{0,1,2,\ldots,n\} for any nn. Indeed, let 𝒫n\mathcal{P}_{n} be the graded Lie supergroup of degree nn as in Definition 55. Then by Definition 55, Πn\Pi^{n} is a partial homomorphism. Further, for any Lie supergroups homomorphism ψ:𝒢𝒢\psi:\mathcal{G}^{\prime}\to\mathcal{G}, where 𝒢\mathcal{G}^{\prime} is a graded Lie supergroup of degree nn, there exists unique Lie supergroup homomorpism (not a partial homomorphism!) Ψn:𝒢𝒫n\Psi_{n}:\mathcal{G}^{\prime}\to\mathcal{P}_{n} such that prsψ=ΨΠspr_{s}\circ\psi^{*}=\Psi^{*}\circ\Pi_{s}.

We finalize this section with the following proposition.

Proposition 61.

Let 𝒢\mathcal{G}, 𝒢~\tilde{\mathcal{G}} be Lie supergroups and 𝒫\mathcal{P}, 𝒫~\tilde{\mathcal{P}} be their 0\mathbb{Z}^{\geq 0}-coverings, respectively. Let ψ:𝒢𝒢~\psi:\mathcal{G}\to\tilde{\mathcal{G}} be a homomorphism of Lie supergroups. Then there exists unique Lie supergroup homomorphism Ψ\Psi of 0\mathbb{Z}^{\geq 0}-coverings 𝒫\mathcal{P} to 𝒫~\tilde{\mathcal{P}} such that the following diagram is commutative

𝒫{\mathcal{P}}𝒫~{\tilde{\mathcal{P}}}𝒢{\mathcal{G}}𝒢~{\tilde{\mathcal{G}}}Ψ\scriptstyle{\Psi}Π\scriptstyle{\Pi}Π\scriptstyle{\Pi}ψ\scriptstyle{\psi}
Proof.

Locally we put

Ψ(y~as):=prsΠψ(x~a),Ψ(η~bt):=prtΠψ(ξ~b).\displaystyle\Psi^{*}(\tilde{y}^{s}_{a}):=pr_{s}\circ\Pi^{*}\circ\psi^{*}(\tilde{x}_{a}),\quad\Psi^{*}(\tilde{\eta}^{t}_{b}):=pr_{t}\circ\Pi^{*}\circ\psi^{*}(\tilde{\xi}_{b}).

The rest of the proof is similar to the proof of Theorem (59). ∎

Later we will prove that such a covering exists for any Lie supergroup. More precisely we will show that F(𝒢)\mathrm{F}(\mathcal{G}) is a covering of a Lie supergroup 𝒢\mathcal{G}.

9. Existence of a 0\mathbb{Z}^{\geq 0}-covering of a Lie superalgebra and
a Lie supergroup

In this section we will show that the F\mathrm{F}^{\prime}-image of a Lie superalgebra 𝔤\mathfrak{g} and the F\mathrm{F}-image a Lie supergroup 𝒢\mathcal{G} are 0\mathbb{Z}^{\geq 0}-coverings of 𝔤\mathfrak{g} and 𝒢\mathcal{G}, respectively. We start with a Lie superalgebara.

9.1. A 0\mathbb{Z}^{\geq 0}-covering of a Lie superalgebra

In this section we will show that the Lie superalgebra 𝔭:=F(𝔤)\mathfrak{p}:=\mathrm{F}^{\prime}(\mathfrak{g}) is a 0\mathbb{Z}^{\geq 0}-covering of 𝔤\mathfrak{g}, see Definition 49. We are ready to prove the following theorem.

Theorem 62.

For any Lie superalgebra 𝔤\mathfrak{g} the Lie superalgebra 𝔭=F(𝔤)\mathfrak{p}=\mathrm{F}^{\prime}(\mathfrak{g}) is a 0\mathbb{Z}^{\geq 0}-covering of 𝔤\mathfrak{g}.

Proof.

In notations of Proposition 44 the homomorphism Π:𝔭𝔤\Pi^{\prime}:\mathfrak{p}\to\mathfrak{g} is defined as follows

Π(I=i1dI(X)+J=idJ(X))=1i!X𝔤i¯.\Pi^{\prime}(\sum_{\sharp I=i-1}\mathrm{d}_{I}(X)+\sum_{\sharp J=i}\mathrm{d}_{J}(X))=\frac{1}{i!}X\in\mathfrak{g}_{\bar{i}}.

Denote Z:=I=i1dI(Z)+J=idJ(Z)𝔭iZ^{\prime}:=\sum\limits_{\sharp I=i-1}\mathrm{d}_{I}(Z)+\sum\limits_{\sharp J=i}\mathrm{d}_{J}(Z)\in\mathfrak{p}_{i} for any Z𝔤i¯Z\in\mathfrak{g}_{\bar{i}}. Using arguments from the proof of Proposition 44 we see that

Π([X,Y])=1i!j![X,Y]=[Π(X),Π(Y)]\Pi^{\prime}([X^{\prime},Y^{\prime}])=\frac{1}{i!j!}[X,Y]=[\Pi^{\prime}(X^{\prime}),\Pi^{\prime}(Y^{\prime})]

for any X𝔭iX^{\prime}\in\mathfrak{p}_{i} and Y𝔭jY^{\prime}\in\mathfrak{p}_{j}. Clearly Π|𝔭n:𝔭n𝔤n¯\Pi^{\prime}|_{\mathfrak{p}_{n}}:\mathfrak{p}_{n}\to\mathfrak{g}_{\bar{n}} is a bijective linear map for any n0n\in\mathbb{Z}^{\geq 0}. ∎

In Proposition 52 we proved that if f:𝔤𝔤f:\mathfrak{g}\to\mathfrak{g}^{\prime} is a homomorphism of Lie algebras, then there exists unique Lie algebra homomorphism f¯:𝔭𝔭\bar{f}:\mathfrak{p}\to\mathfrak{p}^{\prime} such that fΠ=Πf¯f\circ\Pi^{\prime}=\Pi^{\prime}\circ\bar{f}. Now we can construct f¯\bar{f} explicitly. We put f¯=F(f)\bar{f}=\mathrm{F}^{\prime}(f).

We conclude this section with the following observation.

Proposition 63.

The Lie superalgebra Fn(𝔤)\mathrm{F}^{\prime}_{n}(\mathfrak{g}) is a semicovering with support {0,,n}\{0,\ldots,n\} of the Lie superalgebra 𝔤\mathfrak{g} along the homomorphism 2\mathbb{Z}\to\mathbb{Z}_{2}.

9.2. A 0\mathbb{Z}^{\geq 0}-covering of a Lie supergroup

Let 𝒢\mathcal{G} be a Lie supergroup with the Lie superalgebra 𝔤\mathfrak{g} and 𝒫\mathcal{P} be a graded Lie supergroup with the graded Lie superalgebra 𝔭\mathfrak{p} and with 𝒫0=𝒢0\mathcal{P}_{0}=\mathcal{G}_{0}. We put 𝔭(1):=𝔭1𝔭2\mathfrak{p}_{(1)}:=\mathfrak{p}_{1}\oplus\mathfrak{p}_{2}\oplus\cdots. By the Poincaré–Birkhoff–Witt theorem we have the following sheaf isomorphisms

(28) 𝒪𝒢=Hom𝒰(𝔤0¯)(𝒰(𝔤),𝒢0)Hom(S(𝔤1¯),𝒢0)S(𝔤1¯)𝒢0,(𝒪𝒫)q=Hom𝒰(𝔭0)(𝒰(𝔭)q,𝒫0)Hom(Sq(𝔭(1)),𝒫0)(Sq(𝔭(1)))𝒢0,\begin{split}\mathcal{O}_{\mathcal{G}}&=\mathrm{Hom}_{\mathcal{U}(\mathfrak{g}_{\bar{0}})}(\mathcal{U}(\mathfrak{g}),\mathcal{F}_{\mathcal{G}_{0}})\simeq\mathrm{Hom}_{\mathbb{C}}(S^{*}(\mathfrak{g}_{\bar{1}}),\mathcal{F}_{\mathcal{G}_{0}})\simeq S^{*}(\mathfrak{g}^{*}_{\bar{1}})\otimes_{\mathbb{C}}\mathcal{F}_{\mathcal{G}_{0}},\\ (\mathcal{O}_{\mathcal{P}})_{q}&=\mathrm{Hom}_{\mathcal{U}(\mathfrak{p}_{0})}(\mathcal{U}(\mathfrak{p})_{q},\mathcal{F}_{\mathcal{P}_{0}})\simeq\mathrm{Hom}_{\mathbb{C}}(S^{q}(\mathfrak{p}_{(1)}),\mathcal{F}_{\mathcal{P}_{0}})\simeq\\ &\simeq(S^{q}(\mathfrak{p}_{(1)}))^{*}\otimes_{\mathbb{C}}\mathcal{F}_{\mathcal{G}_{0}},\end{split}

where q0q\in\mathbb{Z}^{\geq 0}. The isomorphism (28) leads to a certain choice of global odd and global graded coordinates on 𝒢\mathcal{G} and 𝒫\mathcal{P}, respectively. Indeed, let (ξi)(\xi_{i}) be a basis in 𝔤1¯\mathfrak{g}_{\bar{1}}^{*} and let (ζbq)(\zeta^{q}_{b}) be a basis in 𝔭q\mathfrak{p}_{q}^{*}, where q1q\geq 1. Since 𝔤1¯\mathfrak{g}_{\bar{1}}^{*} is a direct term in S(𝔤1¯)S^{*}(\mathfrak{g}^{*}_{\bar{1}}), we assume that ξiS(𝔤1¯)\xi_{i}\in S^{*}(\mathfrak{g}^{*}_{\bar{1}}), and similarly ζbqSq(𝔭(1)))\zeta^{q}_{b}\in S^{q}(\mathfrak{p}_{(1)}))^{*}. Then the system (ξi)(\xi_{i}) is a system of global odd coordinates on 𝒢\mathcal{G}, and (ζbq)(\zeta^{q}_{b}), where q1q\geq 1 is the degree of the coordinate ζbq\zeta^{q}_{b}, is a system of global graded coordinates of 𝒫\mathcal{P}. We will call such coordinates of a (graded) Lie supergroup standard. Note that in general we cannot choose global coordinates of degree 0.

We need the following two lemmas.

Lemma 64.

Let us fix a graded chart 𝒰\mathcal{U} on a graded Lie supergroup 𝒫\mathcal{P} with standard graded coordinates (ζbq)(\zeta^{q}_{b}), where q1q\geq 1. Denote by θcq\theta^{q}_{c} the basis in 𝔭q\mathfrak{p}_{q} dual to the basis (ζbq)(\zeta^{q}_{b}) of 𝔭q\mathfrak{p}_{q}^{*}. Let f(𝒪𝒰)pf\in(\mathcal{O}_{\mathcal{U}})_{p} be a graded function of degree pp. If f(θcp)=0f(\theta^{p}_{c})=0 for any cc, then ff is a sum of products of coordinates of degree p<pp^{\prime}<p.

Proof.

By definition of standard coordinates, ff is a sum of products of coordinates of degree p<pp^{\prime}<p with respect to the coordinates (ζbq)(\zeta^{q}_{b}). Therefore the same holds for any other coordinates. ∎

Lemma 65.

Let 𝒰\mathcal{U} be a graded superdomain with graded coordinates (θbq)(\theta^{q}_{b}), where qq\in\mathbb{Z} is the degree of a coordinate. The system (θ~bq)(\tilde{\theta}^{q}_{b}), where

θ~bq=θbq+q1+qk=qfi1ikθi1q1θikqk,fi1ik=fi1ik(θa0),\displaystyle\tilde{\theta}^{q}_{b}=\theta^{q}_{b}+\sum_{q_{1}+\cdots q_{k}=q}f_{i_{1}\ldots i_{k}}\theta^{q_{1}}_{i_{1}}\cdots\theta^{q_{k}}_{i_{k}},\quad f_{i_{1}\ldots i_{k}}=f_{i_{1}\ldots i_{k}}(\theta^{0}_{a}),

is a new graded coordinate system in 𝒰\mathcal{U}.

Proof.

By induction. ∎

Now we are ready to prove the main result of this section.

Theorem 66.

For any Lie supergroup 𝒢\mathcal{G} the image F(𝒢)\mathrm{F}(\mathcal{G}) is a 0\mathbb{Z}^{\geq 0}-covering of 𝒢\mathcal{G}.

Proof.

Denote 𝒫:=F(𝒢)\mathcal{P}:=\mathrm{F}(\mathcal{G}). We have a natural homomorphism of Lie supergroups Π:𝒫𝒢\Pi:\mathcal{P}\to\mathcal{G} given by

Π:Hom𝒰(𝔤0¯)(𝒰(𝔤),𝒢0)Hom𝒰(𝔭0)(𝒰(𝔭),𝒢0),\displaystyle\Pi^{*}:\mathrm{Hom}_{\mathcal{U}(\mathfrak{g}_{\bar{0}})}(\mathcal{U}(\mathfrak{g}),\mathcal{F}_{\mathcal{G}_{0}})\to\mathrm{Hom}_{\mathcal{U}(\mathfrak{p}_{0})}(\mathcal{U}(\mathfrak{p}),\mathcal{F}_{\mathcal{G}_{0}}),
Π(f)(X)=f(Π(X)),fHom𝒰(𝔤0¯)(𝒰(𝔤),𝒢0),X𝒰(𝔭),\displaystyle\Pi^{*}(f)(X)=f(\Pi^{\prime}(X)),\quad f\in\mathrm{Hom}_{\mathcal{U}(\mathfrak{g}_{\bar{0}})}(\mathcal{U}(\mathfrak{g}),\mathcal{F}_{\mathcal{G}_{0}}),\,\,X\in\mathcal{U}(\mathfrak{p}),

where Π:𝔭𝔤\Pi^{\prime}:\mathfrak{p}\to\mathfrak{g} is the differential of Π\Pi. (With the same letter we denote the corresponding homomorphism of the universal enveloping algebras.) Let us choose an atlas 𝒜\mathcal{A} on 𝒢\mathcal{G} and let us fix 𝒰𝒜\mathcal{U}\in\mathcal{A} with coordinates (xa,ξb)(x_{a},\xi_{b}), where ξb\xi_{b} are standard global odd coordinates.

Let us choose a basis Z1,,ZnZ_{1},\ldots,Z_{n} in 𝔤0¯\mathfrak{g}_{\bar{0}} and a basis T1,,TmT_{1},\ldots,T_{m} in 𝔤1¯\mathfrak{g}_{\bar{1}}, dual to the basis (ξb)(\xi_{b}). Let Z1s,,ZnsZ^{s}_{1},\ldots,Z^{s}_{n} is the corresponding basis in 𝔭s\mathfrak{p}_{s}, where ss is even, and T1t,,TmtT^{t}_{1},\ldots,T^{t}_{m} is the corresponding basis in 𝔭s\mathfrak{p}_{s}, where tt is odd. Further let us identify xax_{a} with elements in Xa𝒪𝒢X_{a}\in\mathcal{O}_{\mathcal{G}}, where

Xa(1)=xa,Xa(Tj)=0,Xa(Zj)=Z~j(xa)red.X_{a}(1)=x_{a},\quad X_{a}(T_{j})=0,\quad X_{a}(Z_{j})=\tilde{Z}_{j}(x_{a})_{red}.

Here Z~j\tilde{Z}_{j} is the fundamental vector field on 𝒢\mathcal{G} corresponding to ZjZ_{j} and Z~j(xa)red\tilde{Z}_{j}(x_{a})_{red} is the image of the function Z~j(xa)\tilde{Z}_{j}(x_{a}) via the natural projection 𝒪𝒢𝒢0\mathcal{O}_{\mathcal{G}}\to\mathcal{F}_{\mathcal{G}_{0}}.

Let us define an atlas {𝒱i}\{\mathcal{V}_{i}\} as in Definition 55 and 54. We put

(29) yas:=prsΠ(xa),ηbt:=prtΠ(ξb)y^{s}_{a}:=pr_{s}\circ\Pi^{*}(x_{a}),\quad\eta^{t}_{b}:=pr_{t}\circ\Pi^{*}(\xi_{b})

for ss even and tt odd. Let us check that (xa,yas,ηbt)(x_{a},y^{s}_{a},\eta^{t}_{b}) are graded coordinates in 𝒫\mathcal{P}. According above we may choose standard graded coordinates (xa,y~as,η~bt)(x_{a},\tilde{y}^{s}_{a},\tilde{\eta}^{t}_{b}), where xa𝒢0x_{a}\in\mathcal{F}_{\mathcal{G}_{0}} are as above, and

y~is(Zjs)=δij,y~is(Tjt)=0,η~it(Tjt)=δij,η~it(Zjs)=0.\tilde{y}^{s}_{i}(Z^{s}_{j})=\delta_{ij},\quad\tilde{y}^{s}_{i}(T^{t}_{j})=0,\quad\tilde{\eta}^{t}_{i}(T^{t}_{j})=\delta_{ij},\quad\tilde{\eta}^{t}_{i}(Z^{s}_{j})=0.

And we denote by the same letter y~is\tilde{y}^{s}_{i} the corresponding elements in (Ss(𝔭/𝔭0))(S^{s}(\mathfrak{p}/\mathfrak{p}_{0}))^{*} (or by η~it\tilde{\eta}^{t}_{i} the corresponding elements in (St(𝔭/𝔭0))(S^{t}(\mathfrak{p}/\mathfrak{p}_{0}))^{*}).

Let us show that the coordinates (xa,y~as,η~bt)(x_{a},\tilde{y}^{s}_{a},\tilde{\eta}^{t}_{b}) may be expressed via (xa,yas,ηbt)(x_{a},y^{s}_{a},\eta^{t}_{b}). This will imply that the system (xa,yas,ηbt)(x_{a},y^{s}_{a},\eta^{t}_{b}) is a system of local coordinates as well. We have

ηbt(Tjt)=Π(ξb)(Tjt)=ξb(Π(Tjt))=ξb(Tj)=δbj.\eta^{t}_{b}(T^{t}_{j})=\Pi^{*}(\xi_{b})(T^{t}_{j})=\xi_{b}(\Pi^{\prime}(T^{t}_{j}))=\xi_{b}(T_{j})=\delta_{bj}.

Similarly, checking other values of ηbt(A)\eta^{t}_{b}(A) and η~bt(A)\tilde{\eta}^{t}_{b}(A), where A𝒰(𝔭)tA\in\mathcal{U}(\mathfrak{p})_{t}, we get ηbt=η~bt\eta^{t}_{b}=\tilde{\eta}^{t}_{b}. Further, the systems of vector fields (xa)(\frac{\partial}{\partial x_{a}}) and (Z~a)(\tilde{Z}_{a}) are both local bases of the sheaf of vector fields on 𝒢\mathcal{G}. Therefore the matrix A=(Z~i(xj))redA=(\tilde{Z}_{i}(x_{j}))_{red} is invertible. Let B=A1B=A^{-1}. Consider aBaiyas\sum_{a}B^{i}_{a}{y}^{s}_{a}. Then

aBaiyas(Zjs)=aBaiZ~j(xa)=δij=y~is(Zjs).\sum_{a}B^{i}_{a}{y}^{s}_{a}(Z^{s}_{j})=\sum_{a}B^{i}_{a}\tilde{Z}_{j}(x_{a})=\delta_{ij}=\tilde{y}^{s}_{i}(Z^{s}_{j}).

By Lemma 64 the difference aBaiyasy~is\sum_{a}B^{i}_{a}{y}^{s}_{a}-\tilde{y}^{s}_{i} is a sum of products of coordinates of degrees smaller than ss. By Lemma 65 and induction of ss we conclude that y~is\tilde{y}^{s}_{i} may be expressed via yas{y}^{s^{\prime}}_{a}. The proof is complete. ∎

9.3. A semicovering of a Lie supergroup with support {0,1,,n}\{0,1,\ldots,n\}

As we have seen after Definition 54, we can construct a semicovering of a Lie supergroup 𝒢\mathcal{G} with support {0,1,,n}\{0,1,\ldots,n\} using a 0\mathbb{Z}^{\geq 0}-covering of 𝒢\mathcal{G}. The definition of the functor F\mathrm{F} implies that Fn(𝒢)\mathrm{F}_{n}(\mathcal{G}) is a semicovering of 𝒢\mathcal{G} with support {0,1,,n}\{0,1,\ldots,n\}.

Remark 67.

The functor Fn\mathrm{F}_{n} is an embedding of the category of Lie supergroups to the category of \mathbb{Z}-graded Lie supergroups with support {0,1,,n}\{0,1,\ldots,n\}. Clearly we can recover the Lie supergroup structure on the supermanifold 𝒢\mathcal{G} using the Lie supergroup structure of Fn(𝒢)\mathrm{F}_{n}(\mathcal{G}).

10. Loop superalgebras

In this section we give an explicit matrix realization of a 0\mathbb{Z}^{\geq 0}-covering for a finite dimensional Lie superalgebra and for a matrix Lie supergroup. Let 𝔤\mathfrak{g} be a Lie superalgebra and 𝔭:=F(𝔤)\mathfrak{p}:=\mathrm{F}^{\prime}(\mathfrak{g}). We can easily see that 𝔭\mathfrak{p} is a subalgebra of a loop algebra in the sense [ABFP, Definition 3.1.1], see also [Eld]. Indeed, clearly we have

𝔭i0𝔤i¯ti,𝔭i𝔤i¯ti,\mathfrak{p}\simeq\bigoplus_{i\geq 0}\mathfrak{g}_{\bar{i}}\otimes t^{i},\quad\mathfrak{p}_{i}\simeq\mathfrak{g}_{\bar{i}}\otimes t^{i},

where tt is a formal even variable. (We will get a definition of a loop algebra if we replace i0i\geq 0 by ii\in\mathbb{Z}.)

Now consider the case 𝔤=𝔤𝔩m|n(𝕂)\mathfrak{g}=\mathfrak{gl}_{m|n}(\mathbb{K}). The Lie superalgebra 𝔭=F(𝔤)\mathfrak{p}=\mathrm{F}^{\prime}(\mathfrak{g}) possesses a simple matrix realization. This implies by Ado’s Theorem that we have such a matrix realization for any finite dimensional Lie superalgebra 𝔥𝔤𝔩m|n(𝕂)\mathfrak{h}\subset\mathfrak{gl}_{m|n}(\mathbb{K}). Recall that the general linear Lie superalgebra 𝔤𝔩m|n(𝕂)\mathfrak{gl}_{m|n}(\mathbb{K}) contains all matrices in the following form

(ABCD),\left(\begin{array}[]{cc}A&B\\ C&D\\ \end{array}\right),

where AA is m×mm\times m-matrix and DD is n×nn\times n-matrix over 𝕂\mathbb{K}. Then the Lie superalgebra 𝔭\mathfrak{p} contains all matrices in the following form

(A1000C1D100A2B1A10C2D2C1D1).\left(\begin{array}[]{ccccc}A_{1}&0&0&0&\cdots\\ C_{1}&D_{1}&0&0&\cdots\\ A_{2}&B_{1}&A_{1}&0&\cdots\\ C_{2}&D_{2}&C_{1}&D_{1}&\cdots\\ \cdots&\cdots&\cdots&\cdots&\cdots\\ \end{array}\right).

Here AiA_{i} are m×mm\times m-matrices, DiD_{i} are n×nn\times n-matrices, BiB_{i} are m×nm\times n-matrices, CiC_{i} are n×mn\times m-matrices over 𝕂\mathbb{K} and so on. We have

𝔭=n0𝔭n,𝔭=𝔭0¯𝔭1¯,\mathfrak{p}=\bigoplus_{n\geq 0}\mathfrak{p}_{n},\quad\mathfrak{p}=\mathfrak{p}_{\bar{0}}\oplus\mathfrak{p}_{\bar{1}},

where 𝔭0¯\mathfrak{p}_{\bar{0}} contains all matrices with Bi=0B_{i}=0 and Cj=0C_{j}=0, 𝔭1¯\mathfrak{p}_{\bar{1}} contains all matrices with Ai=0A_{i}=0 and Dj=0D_{j}=0. The \mathbb{Z}-grading of 𝔭\mathfrak{p} has natural meaning: 𝔭0\mathfrak{p}_{0} contains all matrices with Bi=0B_{i}=0, Ci=0C_{i}=0 for any ii, Aj=0A_{j}=0 and Dj=0D_{j}=0 for j>1j>1; 𝔭1\mathfrak{p}_{1} contains all matrices with Ai=0A_{i}=0, Di=0D_{i}=0 for any ii, Cj=0C_{j}=0 and Bj=0B_{j}=0 for j>1j>1.

Let 𝒢\mathcal{G} be a Lie subsupergroup in the general linear Lie supergroup GLm|n(𝕂)\mathrm{GL}_{m|n}(\mathbb{K}). Using our results for the Lie superalgebra 𝔤𝔩m|n(𝕂)\mathfrak{gl}_{m|n}(\mathbb{K}) we can explicitly construct F(GLm|n(𝕂))\mathrm{F}(\mathrm{GL}_{m|n}(\mathbb{K})). Indeed, let GLm|n(𝕂)\mathrm{GL}_{m|n}(\mathbb{K}) has the following coordinate superdomain 𝒰\mathcal{U}

(XΞHY),\left(\begin{array}[]{cc}X&\Xi\\ \mathrm{H}&Y\\ \end{array}\right),

where XGLm(𝕂)X\in\mathrm{GL}_{m}(\mathbb{K}), YGLn(𝕂)Y\in\mathrm{GL}_{n}(\mathbb{K}) are matrices of even coordinates of 𝒰\mathcal{U} and Ξ\Xi, H\mathrm{H} are matrices of odd coordinates. Now F(GLm|n(𝕂))\mathrm{F}(\mathrm{GL}_{m|n}(\mathbb{K})) has the following coordinate superdomain

(X1000H1Y100X2Ξ1X10Ξ2Y2H1Y1).\left(\begin{array}[]{ccccc}X_{1}&0&0&0&\cdots\\ \mathrm{H}_{1}&Y_{1}&0&0&\cdots\\ X_{2}&\Xi_{1}&X_{1}&0&\cdots\\ \Xi_{2}&Y_{2}&\mathrm{H}_{1}&Y_{1}&\cdots\\ \cdots&\cdots&\cdots&\cdots&\cdots\\ \end{array}\right).

XiGLm(𝕂)X_{i}\in\mathrm{GL}_{m}(\mathbb{K}), YjGLn(𝕂)Y_{j}\in\mathrm{GL}_{n}(\mathbb{K}) are matrices of even coordinates of F(GLm|n(𝕂))\mathrm{F}(\mathrm{GL}_{m|n}(\mathbb{K})) and Ξs\Xi_{s}, Ht\mathrm{H}_{t} are matrices of odd coordinates. The Lie supergroup multiplication is given by usual matrix multiplication. A similar idea can be used to construct explicitly F(𝒢)\mathrm{F}(\mathcal{G}) for any matrix Lie supergoup 𝒢\mathcal{G}.

We can give a matrix realization of the Lie superalgebras Fn(𝔤)\mathrm{F}^{\prime}_{n}(\mathfrak{g}), n2n\geq 2. The Lie superalgebra Fn(𝔤)\mathrm{F}^{\prime}_{n}(\mathfrak{g}) contains all matrices in the following form

(A1000C1D100A2B1A10ClDlCl1D1),(A1000C1D100A2B1A10Ar+1BrArA1),\left(\begin{array}[]{ccccc}A_{1}&0&0&\cdots&0\\ C_{1}&D_{1}&0&\cdots&0\\ A_{2}&B_{1}&A_{1}&\cdots&0\\ \cdots&\cdots&\cdots&\cdots&\cdots\\ C_{l}&D_{l}&C_{l-1}&\cdots&D_{1}\\ \end{array}\right),\quad\left(\begin{array}[]{ccccc}A_{1}&0&0&\cdots&0\\ C_{1}&D_{1}&0&\cdots&0\\ A_{2}&B_{1}&A_{1}&\cdots&0\\ \cdots&\cdots&\cdots&\cdots&\cdots\\ A_{r+1}&B_{r}&A_{r}&\cdots&A_{1}\\ \end{array}\right),

for n=2l1n=2l-1 and for n=2rn=2r, respectively. Similarly for matrix Lie supergroups. Note that in general these constructions are not applicable for supermanifolds.

References

  • [ABFP] Bruce Allison, Stephen Berman, John Faulkner, and Arturo Pianzola, Realization of graded-simple algebras as loop algebras, Forum Math. 20 (2008), no. 3, 395–432.
  • [BCC] Balduzzi L., Carmeli C., Cassinelli G. Super GG-spaces. Symmetry in mathematics and physics, 159176, Contemp. Math., 490, AMS Providence, RI, 2009.
  • [Ber] F. A. Berezin. Introduction to algebra and analysis with anticommuting variables. V. P. Palamodov, ed., Moscow State University Press, Moscow, 1983. Expanded transl. into English: Introduction to superanalysis. A. A. Kirillov, ed., D. Reidel, Dordrecht, 1987
  • [BLMS] J. Bernstein, D. Leites, V. Molotkov, V. Shander, Seminar of Supersymmetries: volume 1 (edited by D. Leites) Moscow, Russia: Moscow Center of Continuous Mathematical Education (MCCME)
  • [BGR] A. Bruce, J. Grabowski and M. Rotkiewicz, Polarisation of graded bundles. SIGMA 12 (2016), 106.
  • [B] A. Bruce, On curves and jets on supermanifolds, Arch. Math. (2), 2014.
  • [CCF] C. Carmeli, L. Caston, R. Fioresi, Mathematical foundations of supersymmetry, EMS Series of Lectures in Mathematics, European Math. Soc., Zurich, 2011.
  • [CM] del Carpio-Marek. F. Geometric structure on degree 22 manifolds. PhD-thesis, IMPA, Rio de Janeiro, 2015.
  • [CW] S.-J. Cheng, W. Wang. Dualities and representations of Lie superalgebras. Graduate Studies in Mathematics, 144. American Mathematical Society, Providence, RI, 2012.
  • [CGP] Tiffany Covolo, Janusz Grabowski, Norbert Poncin The category of Z2nZ^{n}_{2}-supermanifolds. J. Math. Phys. 57 (2016), 073503 (16pp).
  • [Ber] Deligne P., Morgan J.W. Notes on supersymmetry (following Joseph Bernstein), Quantum Fields and Strings: A Course for Mathematicians, Vols. 1,2 (Princeton, NJ, 1996/1997), 41-97. American Mathematical Society. Providence, R.I. 1999.
  • [Eld] Alberto Elduque, Graded-simple algebras and cocycle twisted loop algebras. Proc. Amer. Math. Soc. 147 (2019), 2821-2833.
  • [Gav] F. Gavarini, Global splittings and super Harish-Chandra pairs for affine supergroups. Trans. Amer. Math. Soc. 368 (2016), 3973-4026.
  • [Gr] P. Green, On holomorphic graded manifolds, Proc. Amer. Math. Soc. 85 (1982), 587-590.
  • [DW1] R. Donagi, E. Witten Supermoduli Space Is Not Projected. Proc. Symp. Pure Math. 90 (2015) 19-72.
  • [DW2] R. Donagi, E. Witten Super Atiyah classes and obstructions to splitting of supermoduli space. Pure and Applied Mathematics Quarterly Volume 9 (2013) Number 4, Pages: 739 – 788.
  • [GR] Grabowski, J., Rotkiewicz, M. Higher vector bundles and multi-graded symplectic manifolds, J. Geom. Phys., Volume 59, Issue 9, 2009, Pages 1285-1305.
  • [JKPS] B. Jubin, A. Kotov, N. Poncin, V. Salnikov, Differential graded Lie groups and their differential graded Lie algebras, arXiv:1906.09630.
  • [Kac] V. Kac. Lie superalgebras, Advances in Mathematics, Volume 26, Issue 1, October 1977, Pages 8-96.
  • [Kos] Kostant B. Graded manifolds, graded Lie theory, and prequantization. Lecture Notes in Mathematics 570. Berlin e.a.: Springer-Verlag, 1977. P. 177-306.
  • [KPS] Alexei Kotov, Norbert Poncin, and Vladimir Salnikov, On the structure of graded Lie groups. In preparation.
  • [JL] Jotz Lean, M. N-manifolds of degree 2 and metric double vector bundles, arXiv:1504.00880.
  • [L] Leites D.A. Introduction to the theory of supermanifolds. Uspekhi Mat. Nauk, 1980, Volume 35, Issue 1(211), 3-57
  • [M] K. C. H. MackenzieGeneral theory of Lie groupoids and Lie algebroids, Cambridge University Press, 2005
  • [Mas] A. Masuoka, Harish-Chandra pairs for algebraic affine supergroup schemes over an arbitrary field, Transform. Groups 17 (2012), no. 4, 1085–1121.
  • [MS] A. Masuoka, T. Shibata, Algebraic supergroups and Harish-Chandra pairs over a commutative ring, Trans. Amer. Math. Soc. 369 (2017), 3443–3481.
  • [Oni] A.L. Onishchik. A Construction of Non-Split Supermanifolds. Annals of Global Analysis and Geometry 16(4): P. 309-333.
  • [P] Pradines J., Représentation des jets non holonomes par des morphismes vectoriels doubles soudés, C. R. Acad. Sci. Paris Sér. A, 278 (1974), 1523–1526.
  • [R] M.J. Rothstein. “Deformations of Complex Supermanifolds.” Proceedings of the American Mathematical Society, vol. 95, no. 2, 1985, pp. 255–260.
  • [R] Roytenberg, D. On the structure of graded symplectic supermanifolds and Courant algebroids. Contemp. Math., Vol. 315, Amer. Math. Soc., Providence, RI, (2002).
  • [RV] M. Rotkiewicz and E. Vishnyakova. Graded covering of a supermanifold II. The general case.
  • [Sch] M. Scheunert. Generalized Lie algebras. Journal of Mathematical Physics 20, 712 (1979).
  • [T] S.J. Takiff. Rings of invariant polynomials for a class of Lie algebras. Trans. Amer. Math. Soc. 160 (1971), 249-262.
  • [Var] V. S. Varadarajan. Supersymmetry for Mathematicians. Courant Lecture Notes, Volume: 11; 2004; 300 pp.
  • [Vi1] E. Vishnyakova. Graded manifolds of type Δ\Delta and nn-fold vector bundles, Letters in Mathematical Physics 109 (2), 243-293.
  • [Vi2] E. Vishnyakova. On holomorphic functions on a compact complex homogeneous supermanifold. Journal of Algebra 350 (1), 2012, 174-196.
  • [Vi3] E. Vishnyakova. On complex Lie supergroups and split homogeneous supermanifolds. Transformation Groups 16 (2011), no. 1, 265-285.
  • [Vo] Th.Th. Voronov, Q-Manifolds and Mackenzie Theory. Commun. Math. Phys. 315, 279-310 (2012).

E. V.: Departamento de Matemática, Instituto de Ciências Exatas, Universidade Federal de Minas Gerais, Av. Antônio Carlos, 6627, CEP: 31270-901, Belo Horizonte, Minas Gerais, BRAZIL, Institute of Mathematics, University of Cologne, Weyertal 86-90, 50931 Cologne, GERMANY, and Laboratory of Theoretical and Mathematical Physics, Tomsk State University, Tomsk 634050, RUSSIA.

M.  Rotkiewicz: Department of Mathematics, Warsaw University, Warsaw, Poland.