1 Introduction

The automorphism group of a compact complex manifold M carries the structure of a complex Lie group which acts holomorphically on M and whose Lie algebra consists of the holomorphic vector fields on M (see [6]). In this article, we investigate how this result can be extended to the category of compact complex supermanifolds.

Let \({\mathcal {M}}\) be a compact complex supermanifold, i.e. a complex supermanifold whose underlying manifold is compact. An automorphism of \({\mathcal {M}}\) is a biholomorphic morphism \({\mathcal {M}}\rightarrow {\mathcal {M}}\). A first candidate for the automorphism group of such a supermanifold is the set of automorphisms, which we denote by \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\). However, every automorphism \(\varphi \) of a supermanifold \({\mathcal {M}}\) (with structure sheaf \({\mathcal {O}}_{\mathcal {M}}\)) is “even” in the sense that its pullback \(\varphi ^*:{\mathcal {O}}_{\mathcal {M}}\rightarrow {{\tilde{\varphi }}}_* ({\mathcal {O}}_{\mathcal {M}})\) is a parity-preserving morphism. Therefore, we can (at most) expect this set of automorphisms of \({\mathcal {M}}\) to carry the structure of a classical Lie group if we require its action on \({\mathcal {M}}\) to be smooth or holomorphic. We cannot obtain a Lie supergroup of positive odd dimension.

We prove that the group \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\), endowed with an analogue of the compact-open topology, carries the structure of a complex Lie group such that the action on \({\mathcal {M}}\) is holomorphic and its Lie algebra is the Lie algebra of even holomorphic super vector fields on \({\mathcal {M}}\). It should be noted that the group \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) is in general different from the group \({{\mathrm{Aut}}}(M)\) of automorphisms of the underlying manifold M. There is a group homomorphism \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}}) \rightarrow {{\mathrm{Aut}}}(M)\) given by assigning the underlying map to an automorphism of the supermanifold; this group homomorphism is in general neither injective nor surjective.

We define the automorphism group of a compact complex supermanifold  \({\mathcal {M}}\) to be a complex Lie supergroup which acts holomorphically on \({\mathcal {M}}\) and satisfies a universal property. In analogy to the classical case, its Lie superalgebra is the Lie superalgebra of holomorphic super vector fields on \({\mathcal {M}}\), and the underlying Lie group is \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\), the group of automorphisms of \({\mathcal {M}}\). Using the equivalence of complex Harish-Chandra pairs and complex Lie supergroups (see [24]), we construct the appropriate automorphism Lie supergroup of \({\mathcal {M}}\).

More precisely, the outline of this article is the following: First, we introduce a topology on the set \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) of automorphisms on a compact complex supermanifold \({\mathcal {M}}\) (cf. Sect. 3). This topology is an analogue of the compact-open topology in the classical case, which coincides in the case of a compact complex manifold with the topology of uniform convergence. We prove that the topological space \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) with composition and inversion of automorphisms as group operations is a locally compact topological group which satisfies the second axiom of countability.

In Sect. 4, the non-existence of small subgroups of \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) is proven, which means that there exists a neighbourhood of the identity in \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) with the property that this neighbourhood does not contain any non-trivial subgroup. A result on the existence of Lie group structures on locally compact topological groups without small subgroups (see [25]) then implies that \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) carries the structure of a real Lie group.

In the case of a split compact complex supermanifold \({\mathcal {M}}\), the fact that \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) carries the structure of a Lie group follows more easily as described in Remark 8. In this case it can be proven that \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) is the semi-direct product of a finite-dimensional vector space and the automorphism group of the vector bundle corresponding to \({\mathcal {M}}\), which is by [17] a complex Lie group.

Then, continuous one-parameter subgroups of \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) and their action on the supermanifold \({\mathcal {M}}\) are studied (see Sect. 5). This is done in order to obtain results on the regularity of the \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\)-action on \({\mathcal {M}}\) and characterize the Lie algebra of \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\). We prove that the action of each continuous one-parameter subgroup of \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) on \({\mathcal {M}}\) is analytic. As a corollary we get that the Lie algebra of \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) is isomorphic to the Lie algebra \(\mathrm {Vec}_{{{\bar{0}}}}({\mathcal {M}})\) of even holomorphic super vector fields on \({\mathcal {M}}\), and \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) carries the structure of a complex Lie group so that its natural action on \({\mathcal {M}}\) is holomorphic.

Next, we show that the Lie superalgebra \(\mathrm {Vec}({\mathcal {M}})\) of holomorphic super vector fields on a compact complex supermanifold \({\mathcal {M}}\) is finite-dimensional (see Sect. 6). Since \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) carries the structure of a complex Lie group, we already know that \(\mathrm {Vec}_{{{\bar{0}}}}({\mathcal {M}})\), the even part of \(\mathrm {Vec}({\mathcal {M}})\), is finite-dimensional. The key point in the proof in the case of a split supermanifold \({\mathcal {M}}\) is that the tangent sheaf of \({\mathcal {M}}\) is a coherent sheaf of \({\mathcal {O}}_M\)-modules on the compact complex manifold M, where \({\mathcal {O}}_M\) is the sheaf of holomorphic functions on M.

Let \(\alpha \) denote the action of \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) on the Lie superalgebra \(\mathrm {Vec}({\mathcal {M}})\) by conjugation: \(\alpha (\varphi )(X)=\varphi _*(X)=(\varphi ^{-1})^*\circ X\circ \varphi ^*\) for \(\varphi \in {{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\), \(X\in \mathrm {Vec}({\mathcal {M}})\). The restriction of this representation \(\alpha \) to \(\mathrm {Vec}_{{{\bar{0}}}} ({\mathcal {M}})\), the even part of the Lie superalgebra \(\mathrm {Vec}({\mathcal {M}})\), coincides with the adjoint action of the Lie group \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) on its Lie algebra, which is isomorphic to \(\mathrm {Vec}_{{{\bar{0}}}}({\mathcal {M}})\). Hence \(\alpha \) defines a Harish-Chandra pair \(({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}}), \mathrm {Vec}({\mathcal {M}}))\). The equivalence between Harish-Chandra pairs and complex Lie supergroups allows us to define the automorphism Lie supergroup of a compact complex supermanifold as follows (see Definition 2):

Definition

(Automorphism Lie supergroup) Define the automorphism group \({{\mathrm{Aut}}}({\mathcal {M}})\) of a compact complex supermanifold to be the unique complex Lie supergroup associated with the Harish-Chandra pair \(({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}}), \mathrm {Vec}({\mathcal {M}}))\) with representation \(\alpha \).

The natural action of the automorphism Lie supergroup \({{\mathrm{Aut}}}({\mathcal {M}})\) on \({\mathcal {M}}\) is holomorphic, i.e. we have a morphism \(\varPsi :{{\mathrm{Aut}}}({\mathcal {M}})\times {\mathcal {M}}\rightarrow {\mathcal {M}}\) of complex supermanifolds. The automorphism Lie supergroup \({{\mathrm{Aut}}}({\mathcal {M}})\) satisfies the following universal property (see Theorem 22):

Theorem

If \({\mathcal {G}}\) is a complex Lie supergroup with a holomorphic action \(\varPsi _{{\mathcal {G}}}:{\mathcal {G}}\times {\mathcal {M}}\rightarrow {\mathcal {M}}\) on \({\mathcal {M}}\), then there is a unique morphism \(\sigma :{\mathcal {G}}\rightarrow {{\mathrm{Aut}}}({\mathcal {M}})\) of Lie supergroups such that the diagram

is commutative.

The automorphism Lie supergroup of a compact complex supermanifold is the unique complex Lie supergroup satisfying the preceding universal property.

Using the “functor of points” approach to supermanifolds, an alternative definition of the automorphism group as a functor in analogy to [20, 22] is possible, which is studied in Sect. 8. If \({\mathcal {M}}\) is a compact complex supermanifold, this functor from the category of supermanifolds to the category of sets can be defined by the assignment

$$\begin{aligned}{\mathcal {N}}\mapsto \{\varphi :{\mathcal {N}}\times {\mathcal {M}}\rightarrow {\mathcal {N}}\times {\mathcal {M}}\,|\, \varphi \text { is invertible, and } \mathrm {pr}_{{\mathcal {N}}} \circ \varphi =\mathrm {pr}_{{\mathcal {N}}}\}, \end{aligned}$$

where \(\mathrm {pr}_{{\mathcal {N}}}:{\mathcal {N}}\times {\mathcal {M}} \rightarrow {\mathcal {N}}\) denotes the projection onto the first component. The two approaches to the automorphism group are equivalent and the constructed automorphism group \({{\mathrm{Aut}}}({\mathcal {M}})\) represents the just defined functor.

In the classical case, another class of complex manifolds where the automorphism group carries the structure of a Lie group is given by the bounded domains in \({\mathbb {C}}^m\) (see [8]). An analogue statement is false in the case of supermanifolds. In Sect. 9, we give an example showing that in the case of a complex supermanifold \({\mathcal {M}}\) whose underlying manifold is a bounded domain in \({\mathbb {C}}^m\) there does in general not exist a Lie supergroup acting on \({\mathcal {M}}\) and satisfying the universal property of the preceding theorem.

In Sect. 10, the automorphism group \({{\mathrm{Aut}}}({\mathcal {M}})\) or its underlying Lie group \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) are determined for some supermanifolds \({\mathcal {M}}\) with underlying manifold \(M={\mathbb {P}}_1{\mathbb {C}}\).

2 Preliminaries and notation

Throughout, we work with the “Berezin-Leĭtes-Kostant-approach” to supermanifolds (cf. [1, 15, 16]). If a supermanifold is denoted by a calligraphic letter \({\mathcal {M}}\), then we denote the underlying manifold by the corresponding uppercase standard letter M, and the structure sheaf by \({\mathcal {O}}_{\mathcal {M}}\). We call a supermanifold \({\mathcal {M}}\) compact if its underlying manifold M is compact. By a complex supermanifold we mean a supermanifold \({\mathcal {M}}\) with structure sheaf \({\mathcal {O}}_{\mathcal {M}}\) which is locally, on small enough open subsets \(U\subset M\), isomorphic to \({\mathcal {O}}_U\otimes \bigwedge {\mathbb {C}}^n\), where \({\mathcal {O}}_U\) denotes the sheaf of holomorphic functions on U. For a morphism \(\varphi :{\mathcal {M}}\rightarrow {\mathcal {N}}\) between supermanifolds \({\mathcal {M}}\) and \({\mathcal {N}}\), the underlying map \(M\rightarrow N\) is denoted by \({{\tilde{\varphi }}}\) and its pullback by \(\varphi ^*:{\mathcal {O}}_{\mathcal {N}} \rightarrow {{\tilde{\varphi }}}_*{\mathcal {O}}_{\mathcal {M}}\). An automorphism of a complex supermanifold \({\mathcal {M}}\) is a biholomorphic morphism \({\mathcal {M}}\rightarrow {\mathcal {M}}\), i.e. an invertible morphism in the category of complex supermanifolds.

Let E be a vector bundle on a complex manifold M and \({\mathcal {E}}\) its sheaf of sections. Then we can associate a supermanifold \({\mathcal {M}}=(M,{\mathcal {O}}_{{\mathcal {M}}})\) by setting \({\mathcal {O}}_{\mathcal {M}}=\bigwedge {\mathcal {E}}\), which has a natural \({\mathbb {Z}}\)-grading (and hence a \({\mathbb {Z}}/{2{\mathbb {Z}}}\)-grading). Split supermanifolds are supermanifolds \({\mathcal {M}}\) such that there is a vector bundle on M with sheaf of sections \({\mathcal {E}}\) such that \({\mathcal {M}}\cong (M,\bigwedge {\mathcal {E}})\). If E is e.g. the trivial bundle of rank n on \(M={\mathbb {C}}^m\), then we get the supermanifold \({\mathbb {C}}^{m|n}=({\mathbb {C}}^m,\bigwedge {\mathcal {E}})=({\mathbb {C}}^m,{\mathcal {O}}_{{\mathbb {C}}^m}\otimes \bigwedge {\mathbb {C}}^n)\).

For a complex supermanifold \({\mathcal {M}}\), let \({\mathcal {T}}_{\mathcal {M}}\) denote the tangent sheaf of \({\mathcal {M}}\). The Lie superalgebra of holomorphic vector fields on \({\mathcal {M}}\) is \(\mathrm {Vec}({\mathcal {M}})={\mathcal {T}}_{\mathcal {M}}(M)\), it consists of the subspace \(\mathrm {Vec}_{{{\bar{0}}}}({\mathcal {M}})\) of even and the subspace \(\mathrm {Vec}_{{{\bar{1}}}}({\mathcal {M}})\) of odd super vector fields on \({\mathcal {M}}\).

Let \({\mathcal {M}}\) be a complex supermanifold of dimension (m|n), and let \({\mathcal {I}}_{\mathcal {M}}\) be the subsheaf of ideals generated by the odd elements in the structure sheaf \({\mathcal {O}}_{\mathcal {M}}\) of a supermanifold \({\mathcal {M}}\). As described in [19], we have the filtration

$$\begin{aligned} {\mathcal {O}}_{\mathcal {M}}=({\mathcal {I}}_{\mathcal {M}})^0\supset ({\mathcal {I}}_{\mathcal {M}})^1\supset ({\mathcal {I}}_{\mathcal {M}})^2\supset \cdots \supset ({\mathcal {I}}_{\mathcal {M}})^{n+1}=0 \end{aligned}$$

of the structure sheaf \({\mathcal {O}}_{\mathcal {M}}\) by the powers of \({\mathcal {I}}_{\mathcal {M}}\). Define the quotient sheaves \(\text {gr}_k({\mathcal {O}}_{\mathcal {M}})=({\mathcal {I}}_{\mathcal {M}})^k/ ({\mathcal {I}}_{\mathcal {M}})^{k+1}\). This gives rise to the \({\mathbb {Z}}\)-graded sheaf \(\text {gr}\,{\mathcal {O}}_{\mathcal {M}} ={\textstyle \bigoplus _k} \text {gr}_k({\mathcal {O}}_{\mathcal {M}})\). Furthermore, \(\text {gr}\,{\mathcal {M}}= (M,\text {gr}\,{\mathcal {O}}_{\mathcal {M}})\) is a split complex supermanifold of the same dimension as \({\mathcal {M}}\).

Note that \({\mathcal {E}}:=\text {gr}_1({\mathcal {O}}_{\mathcal {M}})\) defines a vector bundle E on M. An automorphism \(\varphi \) of \({\mathcal {M}}\) yields a pullback \(\varphi ^*\) on \({\mathcal {O}}_{\mathcal {M}}\). Following [10], its reduction to the \({\mathcal {O}}_M\)-module E yields a morphism of vector bundles \(\varphi _0\in {{\mathrm{Aut}}}(E)\) over the reduction \({{\tilde{\varphi }}}\in {{\mathrm{Aut}}}(M)\). By [17] the automorphism group of a principal fibre bundle over a compact complex manifold carries the structure of a complex Lie group. Since every automorphism of a vector bundle canonically induces an automorphism of the associated principal fibre bundle and vice versa, the automorphism group of the associated principal fibre bundle and \({{\mathrm{Aut}}}(E)\) may be identified. Moreover, this identification also respects the topology of compact convergence on both groups. Hence, the group \({{\mathrm{Aut}}}(E)\) also carries the structure of a complex Lie group. On local coordinate domains UV with \({{\tilde{\varphi }}}(U)\subset V\) we can identify \({\mathcal {O}}_{\mathcal {M}}|_V\cong \varGamma _{\varLambda E}|_V\) and \({\mathcal {O}}_{\mathcal {M}}|_U\cong \varGamma _{\varLambda E}|_U\) and following [21] decompose \(\varphi ^*=\varphi _0^*\exp (Y)\) with \({\mathbb {Z}}\)-degree preserving automorphism \(\varphi _0^*:\varGamma _{\varLambda E}|_V \rightarrow \varGamma _{\varLambda E}|_U\) induced by \(\varphi _0\) and where Y is an even super derivation on \(\varGamma _{\varLambda E}|_V\) increasing the \({\mathbb {Z}}\)-degree by 2 or more. Note that the exponential series \(\exp (Y)\) is finite since Y is nilpotent.

More generally, there is a relation between nilpotent even super vector fields on a supermanifold and morphisms of this supermanifold satisfying a certain nilpotency condition. This is a direct consequence of a technical result on the relation of algebra homomorphisms and derivations (cf. [23], Proposition 2.1.3 and Lemma 2.1.4). If \(\varphi :{\mathcal {M}}\rightarrow {\mathcal {M}}\) is a morphism of supermanifolds with underlying map \({{\tilde{\varphi }}}=\mathrm {id}_M\) and such that \(\varphi ^*-\mathrm {id}_{\mathcal {M}}^*:{\mathcal {O}}_{\mathcal {M}}\rightarrow {\mathcal {O}}_{\mathcal {M}}\) is nilpotent, i.e. there is \(N\in {\mathbb {N}}\) with \((\varphi ^*-\mathrm {id}_{\mathcal {M}}^*)^N=0\), then

$$\begin{aligned} X=\log (\varphi ^*)=\sum _{n=1}^N\frac{(-1)^{n+1}}{n} (\varphi ^*-\mathrm {id}_{\mathcal {M}}^*)^n \end{aligned}$$

is a nilpotent even super vector field on \({\mathcal {M}}\) and we have

$$\begin{aligned} \varphi ^*=\exp (X)=\sum _{n\ge 0}\frac{1}{n!} X^n. \end{aligned}$$

Furthermore, for any nilpotent even super vector fifeld X on \({\mathcal {M}}\), the (finite) sum \(\exp (X)\) defines a map \({\mathcal {O}}_{\mathcal {M}}\rightarrow {\mathcal {O}}_{\mathcal {M}}\) which is the pullback of an invertible morphism \({\mathcal {M}}\rightarrow {\mathcal {M}}\) with the identity as underlying map, and the pullback of the inverse is \(\exp (-X)\).

3 The topology on the group of automorphisms

Let \({\mathcal {M}}\) be a compact complex supermanifold. An automorphism of \({\mathcal {M}}\) is a biholomorphic morphism \(\varphi :{\mathcal {M}}\rightarrow {\mathcal {M}}\). Denote by \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) the set of automorphisms of \({\mathcal {M}}\).

In this section, a topology on \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) is introduced, which generalizes the compact-open topology and topology of compact convergence of the classical case. Then we show that \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) is a locally compact topological group with respect to this topology.

Let \(K\subseteq M\) be a compact subset such that there are local odd coordinates \(\theta _1,\ldots , \theta _n\) for \({\mathcal {M}}\) on an open neighbourhood of K. Moreover, let \(U\subseteq M\) be open and \(f\in {\mathcal {O}}_{\mathcal {M}}(U)\), and let \(U_{\nu }\) be open subsets of \({\mathbb {C}}\) for \(\nu \in ({\mathbb {Z}}_2)^n\). Let \(\varphi :{\mathcal {M}}\rightarrow {\mathcal {M}}\) be an automorphism with \({{\tilde{\varphi }}}(K)\subseteq U\). Then there are holomorphic functions \(\varphi _{f,\nu }\) on a neighbourhood of K such that

$$\begin{aligned} \varphi ^*(f)=\sum _{\nu \in ({\mathbb {Z}}_2)^n}\varphi _{f,\nu } \theta ^\nu . \end{aligned}$$

Let

$$\begin{aligned} \varDelta (K, U, f,\theta _j, U_{\nu }) =\{\varphi \in {{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})|\,{\tilde{\varphi }}(K)\subseteq U, \, \varphi _{f,\nu }(K)\subseteq U_{\nu }\}, \end{aligned}$$

and endow \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) with the topology generated by sets of this form, i.e. the sets of the form \(\varDelta (K, U,f,\theta _j, U_{\nu })\) form a subbase of the topology.

For any open subset \(U\subseteq M\) such that there exist coordinates for \({\mathcal {M}}\) on U, fix a set of coordinates functions \(f_1^U,\ldots , f_{m+n}^U \in {\mathcal {O}}_{{\mathcal {M}}}(U)\). Using Taylor expansion one can show that the sets of the form \(\varDelta (K, U, f_l^U,\theta _j, U_{\nu })\) then also form a subbase of the topology.

Remark 1

In particular, the subsets of the form

$$\begin{aligned} \varDelta (K,U)=\{\varphi \in {{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}}) |\,{\tilde{\varphi }}(K)\subseteq U\} \end{aligned}$$

are open for \(K\subseteq M\) compact and \(U\subseteq M\) open. Hence the map \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\rightarrow {{\mathrm{Aut}}}(M)\), associating with an automorphism \(\varphi \) of \({\mathcal {M}}\) the underlying automorphism \({\tilde{\varphi }}\) of M, is continuous.

Remark 2

The group \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) endowed with the above topology is a second-countable Hausdorff space since M is second-countable.

Let \(U\subseteq M\) be open. Then we can define a topology on \({\mathcal {O}}_{\mathcal {M}}(U)\) as follows: If \(K\subseteq U\) is compact such that there exist odd coordinates \(\theta _1,\ldots , \theta _n\) on a neighbourhood of K, write \(f\in {\mathcal {O}}_{\mathcal {M}}(U)\) on K as \(f=\sum _{\nu }f_\nu \theta ^\nu \). Let \(U_\nu \subseteq {\mathbb {C}}\) be open subsets. Then define a topology on \({\mathcal {O}}_{\mathcal {M}}(U)\) by requiring that the sets of the form \(\{f\in {\mathcal {O}}_{\mathcal {M}}(U)|\,f_\nu (K)\subseteq U_\nu \}\) are a subbase of the topology. A sequence of functions \(f_k\) converges to f if and only if in all local coordinate domains with odd coordinates \(\theta _1,\ldots ,\theta _n\) and \(f_k=\sum _{\nu }f_{k,\nu }\theta ^\nu \), \(f=\sum _{\nu }f_\nu \theta ^\nu \), the coefficient functions \(f_{k,\nu }\) converge uniformly to \(f_\nu \) on compact subsets. Note that for any open subsets \(U_1, U_2\subseteq M\) with \(U_1\subset U_2\) the restriction map \({\mathcal {O}}_{\mathcal {M}}(U_2)\rightarrow {\mathcal {O}}_{\mathcal {M}}(U_1)\), \(f\mapsto f|_{U_1}\), is continuous.

Using Taylor expansion (in local coordinates) of automorphisms of \({\mathcal {M}}\) we can deduce the following lemma:

Lemma 3

A sequence of automorphisms \(\varphi _k:{\mathcal {M}}\rightarrow {\mathcal {M}}\) converges to an automorphism \(\varphi :{\mathcal {M}}\rightarrow {\mathcal {M}}\) with respect to the topology of \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) if and only if the following condition is satisfied: For all \(U,V\subseteq M\) open subsets of M such that V contains the closure of \({\tilde{\varphi }}(U)\), there is an \(N\in {\mathbb {N}}\) such that \(\tilde{\varphi _k}(U)\subseteq V\) for all \(k\ge N\). Furthermore, for any \(f\in {\mathcal {O}}_{\mathcal {M}}(V)\) the sequence \((\varphi _k)^*(f)\) converges to \(\varphi ^*(f)\) on U in the topology of \({\mathcal {O}}_{\mathcal {M}}(U)\).

Lemma 4

If \(U, V\subseteq M\) are open subsets, \(K\subseteq M\) is compact with \(V\subseteq K\), then the map

$$\begin{aligned} \varDelta (K,U)\times {\mathcal {O}}_{\mathcal {M}}(U) \rightarrow {\mathcal {O}}_{\mathcal {M}}(V),\, (\varphi ,f)\mapsto \varphi ^*(f) \end{aligned}$$

is continuous.

Proof

Let \(\varphi _k\in \varDelta (K,U)\) be a sequence of automorphisms of \({\mathcal {M}}\) converging to \(\varphi \in \varDelta (K,U)\), and \(f_l\in {\mathcal {O}}_{\mathcal {M}}(U)\) a sequence converging to \(f\in {\mathcal {O}}_{\mathcal {M}}(U)\). Choosing appropriate local coordinates and using Taylor expansion of the pullbacks \((\varphi _k)^*(f_l)\), it can be shown that \((\varphi _k)^*(f_l)\) converges to \(\varphi ^*(f)\) as \(k,l\rightarrow \infty \). This uses that the derivatives of a sequence of uniformly converging holomorphic functions also uniformly converge. \(\square \)

Lemma 5

The topological space \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) is locally compact.

The following remark about invertible morphisms is useful for the proof of this lemma.

Remark 6

(See e.g. Proposition 2.15.1 in [15] or Corollary 2.3.3 in [16]) Let \({\mathcal {M}}\) be a complex supermanifold and \(\varphi :{\mathcal {M}}\rightarrow {\mathcal {M}}\) any morphism. Let \(\xi _1,\ldots , \xi _n\) and \(\theta _1,\ldots ,\theta _n\) be local odd coordinates for \({\mathcal {M}}\), and superfunctions \(\varphi _{j,k}\), \(\varphi _{j,\nu }\) such that \(\varphi ^*(\xi _j)=\sum _{k=1}^n \varphi _{j,k} \theta _k +\sum _{||\nu ||\ge 3} \varphi _{j,\nu }\theta ^\nu ,\) where \(||\nu ||=||(\nu _1,\ldots ,\nu _n)||=\nu _1+\cdots +\nu _n\ge 3\). Then \(\varphi \) is locally biholomorphic if and only if the underlying map \({{\tilde{\varphi }}}\) is locally biholomorphic and \(\det \left( (\varphi _{j,k}(y))_{1\le j,k\le n}\right) \ne 0\). The morphism \(\varphi \) is hence invertible if it is everywhere locally biholomorphic and \({{\tilde{\varphi }}}\) is biholomorphic.

Proof (of Lemma 5)

Let \(\psi \in {{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\). For each fixed \(x\in M\) there are open neighbourhoods \(V_x\) and \(U_x\) of x and \({\tilde{\psi }}(x)\) respectively such that \({\tilde{\psi }}(K_x)\subseteq U_x\) for \(K_x:={\overline{V}}_x\). We may additionally assume that there are local odd coordinates \(\xi _1,\ldots , \xi _n\) for \({\mathcal {M}}\) on \(U_x\), and \(\theta _1,\ldots , \theta _n\) local odd coordinates on an open neighbourhood of \(K_x\). For any automorphism \(\varphi :{\mathcal {M}}\rightarrow {\mathcal {M}}\) with \({\tilde{\varphi }}(K_x)\subseteq U_x\), let \(\varphi _{j,k}\), \(\varphi _{j,\nu }\) (for \(||\nu ||=||(\nu _1,\ldots ,\nu _n)||=\nu _1+\cdots +\nu _n\ge 3\)) be local holomorphic functions such that

$$\begin{aligned} \varphi ^*(\xi _j)=\sum _{k=1}^n \varphi _{j,k} \theta _k +\sum _{||\nu ||\ge 3} \varphi _{j,\nu }\theta ^\nu . \end{aligned}$$

Choose bounded open subsets \(U_{j,k}, U_{j,\nu }\subset {\mathbb {C}}\), such that \(\psi _{j,k}(x)\in U_{j,k}\) and \(\psi _{j,\nu }(x)\in U_{j,\nu }\). Since \(\psi \) is an automorphism, we have

$$\begin{aligned} \det \left( (\psi _{j,k}(y))_{1\le j,k\le n}\right) \ne 0 \end{aligned}$$

for all \(y\in K_x\) by Remark 6. For later considerations shrink \(U_{j,k}\) such that \(\det (C)\ne 0\) for all \(C=(c_{j,k})_{1\le j,k\le n}\) with \(c_{j,k}\in U_{j,k}\). After shrinking \(V_x\) we may assume \(\psi _{j,k}(K_x)\subseteq U_{j,k}\) and \(\psi _{j,\nu }(K_x)\subseteq U_{j,\nu }\). Hence \(\psi \) is contained in the set \(\varTheta (x)=\{\varphi \in {{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\,|\,{\tilde{\varphi }}(K_x) \subseteq {\overline{U}}_x,\,\varphi _{j,k}(K_x)\subseteq {\overline{U}}_{j,k}, \varphi _{j,\nu }(K_x) \subseteq {\overline{U}}_{j,\nu }\}\), which contains an open neighbourhood of \(\psi \). Since M is compact, M is covered by finitely many of the sets \(V_x\), say \(V_{x_1},\ldots ,V_{x_l}\). Then \(\psi \) is contained in \(\varTheta =\varTheta (x_1)\cap \cdots \cap \varTheta (x_l)\). We will now prove that \(\varTheta \) is sequentially compact:

Let \(\varphi _k\) be any sequence of automorphisms contained in \(\varTheta \). Then, using Montel’s theorem and passing to a subsequence, the sequence \(\varphi _k\) converges to a morphism \(\varphi :{\mathcal {M}}\rightarrow {\mathcal {M}}\). It remains to show that \(\varphi \) is an automorphism of \({\mathcal {M}}\).

The underlying map \({\tilde{\varphi }}:M\rightarrow M\) is surjective since if \(p\notin {\tilde{\varphi }}(M)\), then \(\varphi \in \varDelta (M,M\setminus \{p\})\) and therefore \(\varphi _k\in \varDelta (M,M\setminus \{p\})\) for k large enough which contradicts the assumption that \(\varphi _k\) is an automorphism. This also implies that there is an \(x\in M\) such that the differential \(D{\tilde{\varphi }}(x)\) is invertible. Using Hurwitz’s theorem (see e.g. [18], p. 80) it follows \(\det (D{\tilde{\varphi }}(x))\ne 0\) for all \(x\in M\). Thus \({\tilde{\varphi }}\) is locally biholomorphic. Moreover, \(\varphi \) is locally invertible due to the special form of the sets \(\varTheta (x_i)\).

In order check that \({\tilde{\varphi }}\) is injective, let \(p_1,p_2\in M\), \(p_1\ne p_2\), such that \(q={\tilde{\varphi }}(p_1)={\tilde{\varphi }}(p_2)\). Let \(\varOmega _j\), \(j=1,2\), be open neighbourhoods of \(p_j\) with \(\varOmega _1\cap \varOmega _2= \emptyset \). By [18], p. 79, Proposition 5, there exists \(k_0\) with the property that \(q\in {\tilde{\varphi }}_k(\varOmega _1)\) and \(q\in {\tilde{\varphi }}_k(\varOmega _2)\) for all \(k\ge k_0\). The bijectivity of the \(\varphi _k\)’s now yields a contradiction to \(\varOmega _1\cap \varOmega _2=\emptyset \). \(\square \)

Proposition 7

The set \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) is a topological group with respect to composition and inversion of automorphisms.

Proof

Let \(\varphi _k\) and \(\psi _l\) be two sequences of automorphisms of \({\mathcal {M}}\) converging to \(\varphi \) and \(\psi \) respectively. By the classical theory, \(\tilde{\varphi _k}\circ \tilde{\psi _l}\) converges to \({\tilde{\varphi }}\circ {\tilde{\psi }}\), and \(\tilde{\varphi _k}^{-1}\) to \({\tilde{\varphi }}^{-1}\).

Let \(U, V, W\subseteq M\) be open subsets with \({\tilde{\varphi }}(V)\subseteq W\), \(\tilde{\varphi _k}(V)\subseteq W\), \({\tilde{\psi }}(U)\subseteq V\), \(\tilde{\psi _l}(U)\subseteq V\), for k and l sufficiently large and let \(f\in {\mathcal {O}}_{\mathcal {M}}(W)\). Then the sequence \((\varphi _k)^*(f)\in {\mathcal {O}}_{\mathcal {M}}(V)\) converges to \(\varphi ^*(f)\) on V, and by Lemma 4 \((\varphi _k\circ \psi _l)^*(f)=(\psi _l)^*\left( (\varphi _k)^*(f)\right) \) converges to \(\psi ^*(\varphi ^*(f))=(\varphi \circ \psi )^*(f)\) on U as \(k,l\rightarrow \infty \) , which shows that the multiplication is continuous.

Consider now the inversion map \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\rightarrow {{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\), \(\varphi \mapsto \varphi ^{-1}\). Let \(\varphi _k\) be a sequence in \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) converging to \(\varphi \in {{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\). Note that since the automorphism group \({{\mathrm{Aut}}}(M)\) of the underlying manifold M is a topological group, the inversion map \({{\mathrm{Aut}}}(M)\rightarrow {{\mathrm{Aut}}}(M)\) is continuous. For any choice of local coordinate charts on \(U, V\subseteq M\) such that the closure of \({\tilde{\varphi }}^{-1}(U)\) is contained in V we can conclude: Since \({\tilde{\varphi }}_k^{-1}\) converges to \({\tilde{\varphi }}^{-1}\), we have \(\tilde{\varphi _k}^{-1}(U)\subseteq V\) for k sufficiently large. Identify \({\mathcal {O}}_{\mathcal {M}}(U) \cong \varGamma _{\varLambda E}(U)\), resp. \({\mathcal {O}}_{\mathcal {M}}(V) \cong \varGamma _{\varLambda E}(V)\) and decompose \(\varphi ^*=\varphi ^*_0\exp (Y)\), \(\varphi _k^*=\varphi _{k,0}^*\exp (Y_k)\) as in Section 2. Note that \(\varphi _0^*\) is induced by an automorphism \(\varphi _0\) of the vector bundle E. We can verify by an observation in local coordinates that the map \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\rightarrow {{\mathrm{Aut}}}(E)\), \(\varphi \mapsto \varphi _0\), is continuous. Hence, the sequence \(\varphi _{k,0}\) converges to \(\varphi _0\) and \(\varphi _{k,0}^*\) converges to \(\varphi _{0}^*\). By [17] the inversion on \({{\mathrm{Aut}}}(E)\) is continuous. Therefore, \((\varphi _{k,0}^{-1})^*\) converges to  \((\varphi _{0}^{-1})^*\). Due to the finiteness of the logarithm and exponential series on nilpotent elements, \(Y_k\) converges to Y. Hence, \((\varphi ^{-1}_k)^*=\exp ({-Y_k})(\varphi ^{*}_{k,0})^{-1}\) converges to \(\exp ({-Y})(\varphi ^{*}_{0})^{-1}=(\varphi ^{*})^{-1}\). \(\square \)

Remark 8

Let \({\mathcal {M}}\) be a split supermanifold and let \(E\rightarrow M\) be a vector bundle with associated sheaf of sections \({\mathcal {E}}\) such that the structure sheaf \({\mathcal {O}}_{\mathcal {M}}\) is isomorphic to \(\bigwedge {\mathcal {E}}\). By [17] the group of automorphisms \({{\mathrm{Aut}}}(E)\) of the vector bundle E is a complex Lie group. Each automorphism \(\varphi \) of the supermanifold \({\mathcal {M}}\) induces an automorphism \(\varphi _0\) of the vector bundle E over the underlying map \({{\tilde{\varphi }}}\) of \(\varphi \), and the map \(\pi :{{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\rightarrow {{\mathrm{Aut}}}(E)\), \(\varphi \mapsto \varphi _0\), is continuous. An automorphism of the bundle E lifts to an automorphism of the supermanifold \({\mathcal {M}}\) if we fix a splitting \({\mathcal {O}}_{\mathcal {M}}\cong \bigwedge {\mathcal {E}}\). If \(\chi :E\rightarrow E\) is an automorphism with pullback \(\chi ^*\) we define an automorphism of \({\mathcal {M}}\) by the pullback \(f_1\wedge \ldots \wedge f_k\mapsto \chi ^*(f_1)\wedge \ldots \wedge \chi ^*(f_k)\) for \(f_1\wedge \ldots \wedge f_k\in \bigwedge ^k{\mathcal {E}}\). This assignment defines a section of \(\pi \). In particular, \(\pi \) is surjective and we have an exact sequence

$$\begin{aligned} 0\rightarrow \ker \pi \rightarrow {{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\rightarrow {{\mathrm{Aut}}}(E)\rightarrow 0, \end{aligned}$$

which splits. Consequently, the topological group \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) is a semidirect product

$$\begin{aligned} {{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\cong \ker \pi \rtimes {{\mathrm{Aut}}}(E). \end{aligned}$$

The kernel of \(\pi \) consists of those automorphisms \(\varphi \) of \({\mathcal {M}}\) whose underlying map \({{\tilde{\varphi }}}\) is the identity on M and whose pullback \(\varphi ^*\) satisfies

$$\begin{aligned} (\varphi ^*-\mathrm {id}^*)({\mathcal {E}})\subseteq \bigoplus _{k\ge 2} \left( \bigwedge \!{}^k\, {\mathcal {E}}\right) .\\ \end{aligned}$$

In this case \((\varphi ^*-\mathrm {id}^*)\) is nilpotent and there is an even super vector field X on \({\mathcal {M}}\) with \(\exp (X)=\varphi ^*\) as mentioned in Sect. 2. The super vector field X is nilpotent and fulfills

$$\begin{aligned} X\left( \bigwedge \!{}^k\,{\mathcal {E}}\right) \subseteq \bigoplus _{l\ge k+2}\left( \bigwedge \!{}^l\,{\mathcal {E}}\right) \end{aligned}$$

for all k. More generally, the map

$$\begin{aligned} \left\{ X\in \mathrm {Vec}_{{{\bar{0}}}}({\mathcal {M}})\left| \, X\left( \bigwedge \!{}^k\,{\mathcal {E}}\right) \subseteq \bigoplus _{l\ge k+2}\left( \bigwedge \!{}^l\,{\mathcal {E}}\right) \text { for all } k \right\} \right.&\ \longrightarrow \ \ \ker \pi ,\ \\ X&\mapsto \exp (X), \end{aligned}$$

which assigns to a super vector field X the automorphism of \({\mathcal {M}}\) with pullback \(\exp (X)\), is bijective. In Sect. 6, we will prove that the Lie superalgebra \(\mathrm {Vec}({\mathcal {M}})\) of super vector fields on \({\mathcal {M}}\) and thus subspaces of \(\mathrm {Vec}({\mathcal {M}})\) are finite-dimensional. Therefore, the topological group \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\cong \ker \pi \rtimes {{\mathrm{Aut}}}(E)\) carries the structure of a complex Lie group.

In the general case of a not necessarily split supermanifold \({\mathcal {M}}\), the proof that \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) can be endowed with the structure of a complex Lie group is more difficult. In order to prove the corresponding result also for non-split supermanifolds, the structure of \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) is further studied in the next two sections.

4 Non-existence of small subgroups of \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\)

In this section, we prove that \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) does not contain small subgroups, i.e. that there exists an open neighbourhood of the identity in \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) such that each subgroup contained in this neighbourhood consists only of the identity. As a consequence, the topological group \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) carries the structure of a real Lie group by a result of Yamabe (cf. [25]).

Before proving the non-existence of small subgroups, a few technical preparations are needed: Consider \({\mathbb {C}}^{m|n}\) and let \(z_1,\ldots ,z_m,\xi _1,\ldots ,\xi _n\) denote coordinates on \({\mathbb {C}}^{m|n}\). Let \(U\subseteq {\mathbb {C}}^m\) be an open subset. For \(f=\sum _\nu f_{\nu }\xi ^{\nu } \in {\mathcal {O}}_{{\mathbb {C}}^{m|n}}(U)\) define

$$\begin{aligned} \left| \left| f\right| \right| _U =\left| \left| \sum _{\nu }f_\nu \xi ^\nu \right| \right| _U :=\sum _\nu \left| \left| f_\nu \right| \right| _U, \end{aligned}$$

where \(||f_\nu ||_U\) denotes the supremum norm of the holomorphic function \(f_\nu \) on U. For any morphism \(\varphi :{\mathcal {U}}=(U,{\mathcal {O}}_{{\mathbb {C}}^{m|n}}|_U) \rightarrow {\mathbb {C}}^{m|n}\) define

$$\begin{aligned} ||\varphi ||_U:=\sum _{i=1}^m ||\varphi ^*(z_i)||_U +\sum _{j=1}^n ||\varphi ^*(\xi _j)||_U. \end{aligned}$$

Lemma 9

Let \({\mathcal {U}}=(U,{\mathcal {O}}_{{\mathbb {C}}^{m|n}}|_{U})\) be a superdomain in \({\mathbb {C}}^{m|n}\). For any relatively compact open subset \(U'\) of U there exists \(\varepsilon >0\) such that any morphism \(\psi :{\mathcal {U}}\rightarrow {\mathbb {C}}^{m|n}\) with the property \(||\psi -\mathrm {id}||_U<\varepsilon \) is biholomorphic as a morphism from \({\mathcal {U}}'=(U',{\mathcal {O}}_{{\mathbb {C}}^{m|n}}|_{U'})\) onto its image.

Proof

Let \(r>0\) such that the closure of the polydisc

$$\begin{aligned} \varDelta ^n_r(z)=\{(w_1,\ldots , w_m)|\,|w_j-z_j|< r\} \end{aligned}$$

is contained in U for any \(z=(z_1,\ldots , z_m)\in U'\). Let \(v\in {\mathbb {C}}^m\) be any non-zero vector. Then we have \(z+\zeta v\in U\) for any \(z\in U'\) and \(\zeta \) in the closure of \(\varDelta _{\frac{r}{||v||}}(0)=\{t\in {\mathbb {C}}|\, |t|<\frac{r}{||v||}\}\). If for given \(\varepsilon >0\) it is \(||\psi -\mathrm {id}||_U<\varepsilon \) then we have in particular \(||{{\tilde{\psi }}}-\mathrm {id}||_U<\varepsilon \) for the supremum norm of the underlying maps \({{\tilde{\psi }}},\mathrm {id}:U\rightarrow {\mathbb {C}}^m\). Then, for the differential \(D{{\tilde{\psi }}}\) of \({{\tilde{\psi }}}\) and any non-zero vector \(v\in {\mathbb {C}}^m\) and any \(z \in U^\prime \) we have

$$\begin{aligned} \left| \left| D{{\tilde{\psi }}} (z)(v)-v\right| \right|&=\left| \left| \frac{d}{dt}\left( {{\tilde{\psi }}}(z+tv)-(z+tv)\right) \right| \right| \\&=\frac{1}{2\pi }\left| \left| \int _{\partial \varDelta _{\frac{r}{||v||}}(0)} \frac{{{\tilde{\psi }}}(z+\zeta v)-(z+\zeta v)}{\zeta ^2} d\zeta \right| \right| \\&\le \frac{1}{2\pi }\int _{\partial \varDelta _{\frac{r}{||v||}}(0)}\left| \left| \frac{{{\tilde{\psi }}}(z+\zeta v)-(z+\zeta v)}{\zeta ^2} \right| \right| d\zeta \\&<\frac{\varepsilon ||v||}{r}. \end{aligned}$$

This implies \(|| D{{\tilde{\psi }}}(z)-\mathrm {id}||< \frac{\varepsilon }{r}\) with respect to the operator norm, for any \(z\in U'\). Thus \({{\tilde{\psi }}}\) is locally biholomorphic on \(U'\) if \(\varepsilon \) is small enough. Moreover, \(\varepsilon \) might now be chosen such that \({{\tilde{\psi }}}\) is injective (see e.g. [13], Chapter 2, Lemma 1.3).

Let \(\psi _{j,k},\psi _{j,\nu }\) be holomorphic functions on U such that \(\psi ^*(\xi _j)=\sum _{k=1}^n \psi _{j,k}\xi _k +\sum _{||\nu ||\ge 3} \psi _{j,\nu } \xi ^\nu .\) By Remark 6 it is now enough to show

$$\begin{aligned} \det ((\psi _{j,k})_{1\le j,k\le n}(z))\ne 0 \end{aligned}$$

for all \(z\in U'\) and \(\varepsilon \) small enough in order to prove that \(\psi \) is a biholomorphism form \({\mathcal {U}}'\) onto its image. This follows from the fact that we assumed, via \(||\psi -\mathrm {id}||_U<\varepsilon \), that \(||\psi _{j,k}||_U< \varepsilon \) if \(j\ne k\) and \(||\psi _{j,j}-1||_U<\varepsilon \). \(\square \)

This lemma now allows us to prove that \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) contains no small subgroups; for a similar result in the classical case see [5], Theorem 1.

Proposition 10

The topological group \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) has no small subgroups, i.e. there is a neighbourhood of the identity which contains no non-trivial subgroup.

Proof

Let \(U\subset V\subset W\) be open subsets of M such that U is relatively compact in V and V is relatively compact in W. Suppose that \({\mathcal {W}}=(W, {\mathcal {O}}_{{\mathcal {M}}}|_W)\) is isomorphic to a superdomain in \({\mathbb {C}}^{m|n}\) and let \(z_1,\ldots , z_m,\xi _1,\ldots ,\xi _n\) be local coordinates on \({\mathcal {W}}\). By definition \(\varDelta ({\overline{V}},W)=\{\varphi \in {{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})|\,{\tilde{\varphi }} ({\overline{V}})\subseteq W\}\) and \(\varDelta ({\overline{U}},V)\) are open neighbourhoods of the identity in \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\). Choose \(\varepsilon >0\) as in the preceding lemma such that any morphism \(\chi :{\mathcal {V}}\rightarrow {\mathbb {C}}^{m|n}\) with \(||\chi -\mathrm {id}||_V<\varepsilon \) is biholomorphic as a morphism from \({\mathcal {U}}\) onto its image. Let \(\varOmega \subseteq \varDelta ({\overline{V}},W)\cap \varDelta ({\overline{U}},V)\) be the subset whose elements \(\varphi \) satisfy \(||\varphi -\mathrm {id}||_V<\varepsilon \). The set \(\varOmega \) is open and contains the identity. Since \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) is locally compact by Lemma 5, it is enough to show that each compact subgroup \(Q\subseteq \varOmega \) is trivial. Otherwise for non-compact Q, let \(\varOmega '\) be an open neighbourhood of the identity with compact closure \({\overline{\varOmega }}'\) which is contained in \(\varOmega \), and suppose \(Q\subseteq \varOmega '\). Then \({\overline{Q}}\subseteq {\overline{\varOmega }}'\subset \varOmega \) is a compact subgroup, and Q is trivial if \({\overline{Q}}\) is trivial.

Define a morphism \(\psi :{\mathcal {V}}\rightarrow {\mathbb {C}}^{m|n}\) by setting

$$\begin{aligned} \psi ^*(z_i)=\int _Q q^*(z_i)\,dq\ \ \text { and }\ \ \psi ^*(\xi _j)=\int _Q q^*(\xi _j)\,dq, \end{aligned}$$

where the integral is taken with respect to the normalized Haar measure on Q. This yields a holomorphic morphism \(\psi :{\mathcal {V}}\rightarrow {\mathbb {C}}^{m|n}\) since each \(q\in Q\) defines a holomorphic morphism \({\mathcal {V}}\rightarrow {\mathcal {W}}\subseteq {\mathbb {C}}^{m|n}\). Its underlying map is \({\tilde{\psi }}(z)=\int _Q {{\tilde{q}}} (z) \,dq\). The morphism \(\psi \) satisfies

$$\begin{aligned} ||\psi ^*(z_i)-z_i||_V =\left| \left| \int _Q (q^*(z_i)-z_i)\,dq\right| \right| _V \le \int _Q ||q^*(z_i)-z_i||_V\, dq \end{aligned}$$

and similarly

$$\begin{aligned} ||\psi ^*(\xi _j)-\xi _j||_V \le \int _Q ||q^*(\xi _j)-\xi _j||_V \, dq. \end{aligned}$$

Consequently, we have

$$\begin{aligned} ||\psi -\mathrm {id}||_V&= \sum _{i=1}^m ||\psi ^*(z_i)-z_i||_V+\sum _{j=1}^n || \psi ^*(\xi _j)-\xi _j||_V\\&\le \int _Q\left( \sum _{i=1}^m||q^*(z_i)-z_i||_V\, + \sum _{j=1}^n ||q^*(\xi _j)-\xi _j||_V \right) \, dq \\&=\int _Q ||q-\mathrm {id}||_V\, dq < \varepsilon . \end{aligned}$$

Thus by the preceding lemma, \(\psi |_U\) is a biholomorphic morphism onto its image. Furthermore, on U we have \(\psi \circ q'=\psi \) for any \(q'\in Q\) since

$$\begin{aligned} (\psi \circ q')^*(z_i)&=(q')^*(\psi ^*(z_i)) =(q')^*\left( \int _Q q^*(z_i)\,dq\right) =\int _Q (q')^*(q^*(z_i))\,dq\\&=\int _Q(q\circ q')^*(z_i)\,dq =\int _Q q^*(z_i)\,dq =\psi ^*(z_i) \end{aligned}$$

due to the invariance of the Haar measure, and also

$$\begin{aligned} (\psi \circ q')^*(\xi _j)=\psi ^*(\xi _j). \end{aligned}$$

The equality \(\psi \circ q'=\psi \) on U implies \(q'|_U=\mathrm {id}_{\mathcal {U}}\) because of the invertibility of \(\psi \). By the identity principle it follows that \(q'=\mathrm {id}_{{\mathcal {M}}}\) if M is connected, and hence \(Q=\{\mathrm {id}_{{\mathcal {M}}}\}\).

In general, M has only finitely many connected components since M is compact. Therefore, a repetition of the preceding argument yields the existence of a neighbourhood of the identity of \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) without any non-trivial subgroups. \(\square \)

By Theorem 3 in [25], the preceding proposition implies the following:

Corollary 11

The topological group \(Aut_{{{\bar{0}}}}({\mathcal {M}})\) can be endowed with the structure of a real Lie group.

5 One-parameter subgroups of \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\)

In order to obtain results on the regularity of the action of \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) on the compact complex supermanifold \({\mathcal {M}}\) and to characterize the Lie algebra of \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\), we study continuous one-parameter subgroups of \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\). Each continuous one-parameter subgroup \({\mathbb {R}}\rightarrow {{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) is an analytic map between the Lie groups \({\mathbb {R}}\) and \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\).

We prove that the action of each continuous one-parameter subgroup of \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) on \({\mathcal {M}}\) is analytic and induces an even holomorphic super vector field on \({\mathcal {M}}\). Consequently, the Lie algebra of \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) may be identified with the Lie algebra \(\mathrm {Vec}_{{{\bar{0}}}}({\mathcal {M}})\) of even holomorphic super vector fields on \({\mathcal {M}}\), and \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) carries the structure of a complex Lie group whose action on the supermanifold \({\mathcal {M}}\) is holomorphic.

Definition 1

A continuous one-parameter subgroup \(\varphi \) of automorphisms of \({\mathcal {M}}\) is a family of automorphisms \(\varphi _t:{\mathcal {M}}\rightarrow {\mathcal {M}}\), \(t\in {\mathbb {R}}\), such that the map \(\varphi :{\mathbb {R}}\rightarrow {{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\), \(t\mapsto \varphi _t\), is a continuous group homomorphism.

Remark 12

Let \(\varphi _t:{\mathcal {M}}\rightarrow {\mathcal {M}}\), \(t\in {\mathbb {R}}\), be a family of automorphisms satisfying \(\varphi _{s+t}=\varphi _s\circ \varphi _t\) for all \(s,t\in {\mathbb {R}}\), and such that \({{\tilde{\varphi }}}:{\mathbb {R}}\times M\rightarrow M\), \({{\tilde{\varphi }}}(t,p)= {\tilde{\varphi }}_t(p)\) is continuous. Then \(\varphi _t\) is a continuous one-parameter subgroup if and only if the following condition is satisfied: Let \(U, V\subset M\) be open subsets, and \([a,b]\subset {\mathbb {R}}\) such that \({{\tilde{\varphi }}}([a,b]\times U)\subseteq V\). Assume moreover that there are local coordinates \(z_1,\ldots , z_m,\xi _1,\ldots ,\xi _n\) for \({\mathcal {M}}\) on U. Then for any \(f\in {\mathcal {O}}_{{\mathcal {M}}}(V)\) there are continuous functions \(f_\nu :[a,b]\times U\rightarrow {\mathbb {C}}\) with \((f_\nu )_t=f_\nu (t,\cdot )\in {\mathcal {O}}_{{\mathcal {M}}}(U)\) for fixed \(t\in [a,b]\) such that

$$\begin{aligned} (\varphi _t)^*(f)=\sum _\nu f_\nu (t,z)\xi ^\nu . \end{aligned}$$

We say that the action of the one-parameter subgroup \(\varphi \) on \({\mathcal {M}}\) is analytic if each \(f_\nu (t,z)\) is analytic in both components.

This equivalent characterization of continuous one-parameter subgroups of automorphisms also allows us to define this notion for non-compact complex supermanifolds.

Proposition 13

Let \(\varphi \) be a continuous one-parameter subgroup of automorphisms on \({\mathcal {M}}\). Then the action of \(\varphi \) on \({\mathcal {M}}\) is analytic.

Remark 14

The statement of Proposition 13 also holds true for complex supermanifolds \({\mathcal {M}}\) with non-compact underlying manifold M as compactness of M is not needed for the proof.

For the proof of the proposition the following technical lemma is needed:

Lemma 15

Let \(U\subseteq V\subseteq {\mathbb {C}}^m\) be open subsets, \(p\in U\), \(\varOmega \subseteq {\mathbb {R}}\) an open connected neighbourhood of 0, and let \(\alpha :\varOmega \times U\rightarrow V\) be a continuous map satisfying

$$\begin{aligned} \alpha (t,z)=\alpha (t+s,z)-f(t,s,z) \end{aligned}$$

for (tsz) in a neighbourhood of (0, 0, p) and for some continuous function f which is analytic in (tz). If \(\alpha \) is holomorphic in the second component, then it is analytic on a neighbourhood of (0, p).

Proof

For small t, \(h>0\), z near p, we have

$$\begin{aligned} h\cdot \alpha (t,z)&=\int _0^h\alpha (t+s,z)ds - \int _0^h f(t,s,z) ds\\&=\int _t^{h+t}\alpha (s,z)ds - \int _0^h \alpha (s,z)ds -\int _0^h (f(t,s,z)-\alpha (s,z))ds\\&=\int _h^{h+t}\alpha (s,z)ds-\int _0^t\alpha (s,z)ds -\int _0^h(f(t,s,z)-\alpha (s,z))ds\\&=\int _0^t(\alpha (s+h,z)-\alpha (s,z))ds -\int _0^h(f(t,s,z)-\alpha (s,z))ds\\&=\int _0^t f(s,h,z)ds -\int _0^h(f(t,s,z)-\alpha (s,z))ds. \end{aligned}$$

The assumption that f is a continuous function which is analytic in the first and third component therefore implies that \(\alpha \) is analytic. \(\square \)

Proof (of Proposition 13)

Due to the action property \(\varphi _{s+t}=\varphi _s\circ \varphi _t\) it is enough to show the statement for the restriction of \(\varphi \) to \((-\varepsilon ,\varepsilon )\times {\mathcal {M}}\) for some \(\varepsilon >0\). Let \(U,V\subseteq M\) be open subsets such that U is relatively compact in V, and such that there are local coordinates \(z_1,\ldots , z_m,\xi _1,\ldots , \xi _n\) on V for \({\mathcal {M}}\). Choose \(\varepsilon >0\) such that \({{\tilde{\varphi }}}_t(U)\subseteq V\) for any \(t\in (-\varepsilon ,\varepsilon )\). Let \(\alpha _{i,\nu }\), \(\beta _{j,\nu }\) be continuous functions on \((-\varepsilon ,\varepsilon )\times U\) with

$$\begin{aligned} (\varphi _t)^*(z_i)=\sum _{|\nu |=0} \alpha _{i,\nu }(t,z) \xi ^\nu \end{aligned}$$

and

$$\begin{aligned} (\varphi _t)^*(\xi _j)=\sum _{|\nu |=1} \beta _{j,\nu }(t,z) \xi ^\nu , \end{aligned}$$

where \(|\nu |=|(\nu _1,\ldots ,\nu _n)|=(\nu _1+\ldots +\nu _n)\!\!\mod 2 \in {\mathbb {Z}}_2\). We have to show that \(\alpha \) and \(\beta \) are analytic in (tz). The induced map \(\psi ':(-\varepsilon ,\varepsilon )\times U\times {\mathbb {C}}^n\rightarrow V\times {\mathbb {C}}^n\) on the underlying vector bundle is given by

$$\begin{aligned} \left( t,\left( \begin{matrix} z_1\\ \vdots \\ z_m \\ v_1\\ \vdots \\ v_n \end{matrix}\right) \right) \mapsto \left( \begin{matrix} \alpha _{1,0}(t,z)\\ \vdots \\ \alpha _{m,0}(t, z)\\ \sum _{k=1}^n \beta _{1,k}(t,z)v_k\\ \vdots \\ \sum _{k=1}^n \beta _{n,k}(t,z)v_k \end{matrix}\right) , \end{aligned}$$

where \(\beta _{j,k}=\beta _{j,e_k}\) if \(e_k=(0,\ldots ,0,1,0,\ldots , 0)\) denotes the k-th unit vector. The map \(\psi '\) is a local continuous one-parameter subgroup on \(U\times {\mathbb {C}}^n\) because \(\varphi \) is a continuous one-parameter subgroup. By a result of Bochner and Montgomery the map \(\psi '\) is analytic in (tzv) (see [4], Theorem 4). Hence, the map \(\psi :(-\varepsilon ,\varepsilon )\times {\mathcal {U}}\rightarrow {\mathcal {V}}\) given by \((\psi _t)^*(z_i)=\alpha _i(t,z)\), \((\psi _t)^*(\xi _j)=\sum _{k=1}^n \beta _{j,k}(t,z)\xi _k\) is analytic. Let X be the local vector field on \({\mathcal {U}}\) induced by \(\psi \), i.e.

$$\begin{aligned} X(f)=\left. \frac{\partial }{\partial t}\right| _0 (\psi _t)^*(f). \end{aligned}$$

We may assume that X is non-degenerate, i.e. the evaluation of X in p, X(p), does not vanish for all \(p\in U\). Otherwise, consider, instead of \(\varphi \), the diagonal action on \({\mathbb {C}}\times {\mathcal {M}}\) acting by addition of t in the first component and \(\varphi _t\) in the second, and note that this action is analytic precisely if \(\varphi \) is analytic. For the differential \(d\psi \) of \(\psi \) in (0, p) we have

$$\begin{aligned} d\psi \left( \left. \frac{\partial }{\partial t}\right| _{(0,p)}\right) =\left. \frac{\partial }{\partial t}\right| _{(0,p)} \circ \psi ^* =X(p)\ne 0. \end{aligned}$$

Therefore, the restricted map \(\psi |_{(-\varepsilon ,\varepsilon )\times \{p\}}\) is an immersion and its image \(\psi ((-\varepsilon ,\varepsilon )\times \{p\})\) is a subsupermanifold of \({\mathcal {V}}\). Let \({\mathcal {S}}\) be a subsupermanifold of \({\mathcal {U}}\) transversal to \(\psi ((-\varepsilon ,\varepsilon )\times \{p\})\) in p. The map \(\psi |_{(-\varepsilon ,\varepsilon )\times {\mathcal {S}}}\) is a submersion in (0, p) since \(d\psi (T_{(0,p)}(-\varepsilon ,\varepsilon )\times \{p\})) = T_p \psi ((-\varepsilon ,\varepsilon )\times \{p\})\) and \(d\psi (T_{(0,p)}\{0\}\times {\mathcal {S}})=T_p{\mathcal {S}}\) because \(\psi |_{\{0\}\times {\mathcal {U}}}=\mathrm {id}\). Hence \(\chi :=\psi |_{(-\varepsilon ,\varepsilon )\times {\mathcal {S}}}\) is locally invertible around (0, p), and thus invertible as a map onto its image after possibly shrinking U and \(\varepsilon \), and

$$\begin{aligned} \chi _*\left( \frac{\partial }{\partial t}\right) =(\chi ^{-1})^*\circ \frac{\partial }{\partial t}\circ \chi ^* =(\chi ^{-1})^*\circ \chi ^*\circ X=X. \end{aligned}$$

Therefore, after defining new coordinates \(w_1,\ldots , w_m,\theta _1,\ldots , \theta _n\) for \({\mathcal {M}}\) on U via \(\chi \), we have \(X=\frac{\partial }{\partial w_1}\) and \((\varphi _t)^*\) is of the form

$$\begin{aligned} (\varphi _t)^*(w_1)&=w_1+t+\sum _{|\nu |=0, \nu \ne 0} \alpha _{1,\nu }(t,w)\theta ^\nu ,\\ (\varphi _t)^*(w_i)&=w_i+\sum _{|\nu |=0,\nu \ne 0} \alpha _{i,\nu }(t,w)\theta ^\nu \ \ \ \text { for } i\ne 1,\\ (\varphi _t)^*(\theta _j)&=\theta _j+\sum _{|\nu |=1,||\nu ||\ne 1} \beta _{j,\nu }(t,w)\theta ^\nu , \end{aligned}$$

for appropriate \(\alpha _{i,\nu }\), \(\beta _{j,\nu }\), where \(||\nu ||=||(\nu _1,\ldots ,\nu _n)||=\nu _1+\cdots +\nu _n\).

For small s and t we have

$$\begin{aligned} \varphi _t^*\left( \varphi _s^*(w_i)\right)&=\varphi _t^*\left( w_i+\delta _{1,i}s+ \sum _{|\nu |=0,||\nu ||\ne 0} \alpha _{i,\nu }(s,w)\theta ^\nu \right) \nonumber \\&= w_i+\delta _{i,1}(t+s) +\sum _{|\nu |=0,||\nu ||\ne 0} \alpha _{i,\nu }(t,w)\theta ^\nu +\sum _{|\nu |=0,||\nu ||\ne 0} \varphi _t^*(\alpha _{i,\nu }(s,w)\theta ^\nu ). \end{aligned}$$
(1)

Let \(f_{i,\nu }(t,s,w)\) be such that

$$\begin{aligned} \sum _{|\nu |=0,||\nu ||\ne 0} \varphi _t^*(\alpha _{i,\nu }(s,w)\theta ^\nu ) =\sum _{|\nu |=0,||\nu ||\ne 0} f_{i,\nu }(t,s,w) \theta ^\nu . \end{aligned}$$
(2)

For fixed \(\nu _0\) the coefficient \(f_{i,{\nu _0}}(t,s,w)\) of \(\theta ^{\nu _0}\) depends only on \(\alpha _{i,\nu _0}(s,w+t e_1)\), \(\beta _{j,\mu }(t,w)\) for \(\mu \) with \(||\mu ||\le ||\nu _0||-1\), and \(\alpha _{j,\nu }(t,w)\) and its partial derivatives in the second component for \(\nu \) with \(||\nu ||\le ||\nu _0||-2\). This can be shown by a calculation using the special form of \(\varphi _t^*(w_j)\) and \(\varphi _t^*(\theta _j)\) and general properties of the pullback of a morphism of supermanifolds. Assume now that the analyticity near (0, p) of \(\alpha _{i,\nu }\), \(\beta _{j,\mu }\) is shown for \(||\nu ||, ||\mu ||< 2k\) and all ij. Let \(\nu _0\) be such that \(||\nu _0||=2k\). Then \(f_{i,\nu _0}(t,s,w)\) is a continuous function which is analytic in (tw) near (0, p) for fixed s. Since \(\varphi _t^*(\varphi _s^*(w_i)) =\varphi _{t+s}^*(w_i)\), using (1) and (2) we get

$$\begin{aligned} \alpha _{i,\nu _0}(t,w)+f_{i,\nu _0}(t,s,w) =\alpha _{i,\nu _0}(t+s,w), \end{aligned}$$

and thus \(\alpha _{i,\nu _0}(t,w)\) is analytic near (0, p) by Lemma 15. Similarly, it can be shown that \(\beta _{j,\mu _0}\) is analytic for \(||\mu _0||=2k+1\) if \(\alpha _{i,\nu }\), \(\beta _{j,\mu }\) for \(||\nu ||\), \(||\mu ||< 2k+1\). \(\square \)

Corollary 16

The Lie algebra of \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) is isomorphic to the Lie algebra \(\mathrm {Vec}_{{{\bar{0}}}}({\mathcal {M}})\) of even super vector fields on \({\mathcal {M}}\), and \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) is a complex Lie group.

Proof

If \(\gamma :{\mathbb {R}}\rightarrow {{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\), \(t\mapsto \gamma _t\) is a continuous one-parameter subgroup, then by Proposition 13 the action of \(\varphi \) on \({\mathcal {M}}\) is analytic. Therefore, \(\gamma \) induces an even holomorphic super vector field \(X(\gamma )\) on \({\mathcal {M}}\) by setting

$$\begin{aligned} X(\gamma )=\left. \frac{\partial }{\partial t}\right| _0 (\gamma _t)^*, \end{aligned}$$

and \(\gamma \) is the flow map of \(X(\gamma )\). On the other hand, since M is compact, the underlying vector field of each \(X\in \mathrm {Vec}_{{{\bar{0}}}}({\mathcal {M}})\) is globally integrable and the proof of Theorem 5.4 in [12] then shows that X is also globally integrable. Its flow defines a one-parameter subgroup \(\gamma ^X\) of \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\), which is continuous. This yields an isomorphism of Lie algebras

$$\begin{aligned} \mathrm {Lie}({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})) \rightarrow \mathrm {Vec}_{{{\bar{0}}}}({\mathcal {M}}). \end{aligned}$$

Consequently, we have \(\mathrm {Lie}({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})) \cong \mathrm {Vec}_{{{\bar{0}}}}({\mathcal {M}})\) and since \(\mathrm {Vec}_{{{\bar{0}}}}({\mathcal {M}})\) is a complex Lie algebra, \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) carries the structure of a complex Lie group. \(\square \)

The Lie group \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) naturally acts on \({\mathcal {M}}\); this action \(\psi :{{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\times {\mathcal {M}}\rightarrow {\mathcal {M}}\) is given by \(\mathrm {ev}_g\circ \psi ^*=g^*\) where \(\mathrm {ev}_g \) denotes the evaluation in \(g\in {{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) in the first component.

Corollary 17

The natural action of \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) on \({\mathcal {M}}\) defines a holomorphic morphism of supermanifolds \(\psi :{{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\times {\mathcal {M}} \rightarrow {\mathcal {M}}\).

Proof

Since the action of each continuous one-parameter subgroup of \( {{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) on \({\mathcal {M}}\) is holomorphic by the preceding considerations, and each \(g\in {{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) is a biholomorphic morphism \(g:{\mathcal {M}}\rightarrow {\mathcal {M}}\), the action \(\psi \) is a holomorphic. \(\square \)

If a Lie supergroup \({\mathcal {G}}\) (with Lie superalgebra \({\mathfrak {g}}\) of right-invariant super vector fields) acts on a supermanifold \({\mathcal {M}}\) via \(\psi :{\mathcal {G}}\times {\mathcal {M}}\rightarrow {\mathcal {M}}\), this action \(\psi \) induces an infinitesimal action \(d\psi :{\mathfrak {g}}\rightarrow \mathrm {Vec}({\mathcal {M}})\) defined by \(d\psi (X)=(X(e)\otimes \mathrm {id}_{\mathcal {M}}^*)\circ \psi ^*\) for any \(X\in {\mathfrak {g}}\), where \(X\otimes \mathrm {id}_{\mathcal {M}}^*\) denotes the canonical extension of the vector field X on \({\mathcal {G}}\) to a vector field on \({\mathcal {G}}\times {\mathcal {M}}\), and \((X(e)\otimes \mathrm {id}_{\mathcal {M}}^*)\) is its evaluation in the neutral element e of \({\mathcal {G}}\).

Corollary 18

Identifying the Lie algebra of \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) with \(\mathrm {Vec}_{{{\bar{0}}}}({\mathcal {M}})\) as in Corollary 16, the induced infinitesimal action of the action \(\psi :{{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\times {\mathcal {M}}\rightarrow {\mathcal {M}}\) in Corollary 17 is the inclusion \(\mathrm {Vec}_{{{\bar{0}}}}({\mathcal {M}})\hookrightarrow \mathrm {Vec}({\mathcal {M}})\).

6 The Lie superalgebra of vector fields

In this section, we prove that the Lie superalgebra \(\mathrm {Vec}({\mathcal {M}})\) of holomorphic super vector fields on a compact complex supermanifold \({\mathcal {M}}\) is finite-dimensional.

First, we prove that \(\mathrm {Vec}({\mathcal {M}})\) is finite-dimensional if \({\mathcal {M}}\) is a split supermanifold using that its tangent sheaf \({\mathcal {T}}_{{\mathcal {M}}}\) is a coherent sheaf of \({\mathcal {O}}_M\)-modules, where \({\mathcal {O}}_M\) denotes again the sheaf of holomorphic functions on the underlying manifold M. Then the statement in the general case is deduced using a filtration of the tangent sheaf.

Remark that since \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) is a complex Lie group with Lie algebra isomorphic to the Lie algebra \(\mathrm {Vec}_{{{\bar{0}}}}({\mathcal {M}})\) of even holomorphic super vector fields on \({\mathcal {M}}\) (see Corollary 16), we already know that the even part of \(\mathrm {Vec}({\mathcal {M}}) =\mathrm {Vec}_{{{\bar{0}}}}({\mathcal {M}})\oplus \mathrm {Vec}_{{{\bar{1}}}}({\mathcal {M}})\) is finite-dimensional.

Lemma 19

Let \({\mathcal {M}}\) be a split complex supermanifold. Then its tangent sheaf \({\mathcal {T}}_{\mathcal {M}}\) is a coherent sheaf of \({\mathcal {O}}_M\)-modules.

Proof

Since \({\mathcal {M}}\) is split, its structure sheaf \({\mathcal {O}}_{\mathcal {M}}\) is isomorphic to \(\bigwedge {\mathcal {E}}\) as an \({\mathcal {O}}_M\)-module, where \({\mathcal {E}}\) is the sheaf of sections of a holomorphic vector bundle on the underlying manifold M. Thus, the structure sheaf \({\mathcal {O}}_{\mathcal {M}}\), and hence also the tangent sheaf \({\mathcal {T}}_{\mathcal {M}}\), carry the structure of a sheaf of \({\mathcal {O}}_M\)-modules. Let \(U\subset M\) be an open subset such that there exist even coordinates \(z_1,\ldots , z_m\) and odd coordinates \(\xi _1,\ldots , \xi _n\). Any derivation \(D\in {\mathcal {T}}_{\mathcal {M}}(U)\) on U can uniquely be written as

$$\begin{aligned} D=\sum _{\nu \in ({\mathbb {Z}}_2)^n}\left( \sum _{i=1}^m f_{i,\nu } (z) \xi ^\nu \frac{\partial }{\partial z_i} +\sum _{j=1}^n g_{j,\nu }(z)\xi ^\nu \frac{\partial }{\partial \xi _j}\right) \end{aligned}$$

where \(f_{i,\nu }\), \(g_{j,\nu }\) are holomorphic functions on U. Therefore, the restricted sheaf \({\mathcal {T}}_{\mathcal {M}}|_U\) is isomorphic to \(({\mathcal {O}}_M|_U)^{2^n (m+n)}\) and \({\mathcal {T}}_{\mathcal {M}}\) is coherent over \({\mathcal {O}}_M\). \(\square \)

Proposition 20

The Lie superalgebra \(\mathrm {Vec}({\mathcal {M}})\) of holomorphic super vector fields on a compact complex supermanifold \({\mathcal {M}}\) is finite-dimensional.

Proof

First, assume that \({\mathcal {M}}\) is split. Then the tangent sheaf \({\mathcal {T}}_{\mathcal {M}}\) is a coherent sheaf of \({\mathcal {O}}_M\)-modules. Thus, the space of global sections of \({\mathcal {T}}_{\mathcal {M}}\), \(\mathrm {Vec}({\mathcal {M}}) ={\mathcal {T}}_{\mathcal {M}}(M)\), is finite-dimensional since M is compact (cf. [9]).

Now, let \({\mathcal {M}}\) be an arbitrary compact complex supermanifold. We associate the split complex supermanifold \(\text {gr}\,{\mathcal {M}}=(M, \text {gr}\, {\mathcal {O}}_{\mathcal {M}})\) as described in Section 2. Let \({\mathcal {I}}_{\mathcal {M}}\) denote as before the subsheaf of ideal in \({\mathcal {O}}_{\mathcal {M}}\) generated by the odd elements. Define the filtration of sheaves of Lie superalgebras

$$\begin{aligned} {\mathcal {T}}_{\mathcal {M}}=:({\mathcal {T}}_{\mathcal {M}})_{(-1)}\supset ({\mathcal {T}}_{\mathcal {M}})_{(0)}\supset ({\mathcal {T}}_{\mathcal {M}})_{(1)}\supset \cdots \supset ({\mathcal {T}}_{\mathcal {M}})_{(n+1)}=0 \end{aligned}$$

of the tangent sheaf \({\mathcal {T}}_{\mathcal {M}}\) by setting

$$\begin{aligned} ({\mathcal {T}}_{\mathcal {M}})_{(k)}=\{D\in {\mathcal {T}}_{\mathcal {M}}|\, D({\mathcal {O}}_{\mathcal {M}})\subset ({\mathcal {I}}_{\mathcal {M}})^k, \,D({\mathcal {I}}_{\mathcal {M}})\subset ({\mathcal {I}}_{\mathcal {M}})^{k+1}\} \end{aligned}$$

for \(k\ge 0\). Moreover, define \(\text {gr}_k({\mathcal {T}}_{\mathcal {M}})= ({\mathcal {T}}_{\mathcal {M}})_{(k)}/({\mathcal {T}}_{\mathcal {M}})_{(k+1)}\) and set

$$\begin{aligned} \text {gr}({\mathcal {T}}_{\mathcal {M}})=\bigoplus _{k\ge -1} \text {gr}_k({\mathcal {T}}_{\mathcal {M}}). \end{aligned}$$

By [19], Proposition 1, the sheaf \(\text {gr}({\mathcal {T}}_{\mathcal {M}})\) is isomorphic to the tangent sheaf of the associated split supermanifold \(\text {gr}\, {\mathcal {M}}\). By the preceding considerations, the space of holomorphic super vector fields on \(\text {gr}\,{\mathcal {M}}\),

$$\begin{aligned} \text {Vec}(\text {gr}\,{\mathcal {M}})=\text {gr}({\mathcal {T}}_{\mathcal {M}})(M)= \bigoplus _{k\ge -1} \text {gr}_k({\mathcal {T}}_{\mathcal {M}})(M), \end{aligned}$$

is of finite dimension. The projection onto the quotient yields

$$\begin{aligned} \dim ({\mathcal {T}}_{\mathcal {M}})_{(k)}(M)-\dim ({\mathcal {T}}_{\mathcal {M}})_{(k+1)}(M) \le \dim (\text {gr}_k({\mathcal {T}}_{\mathcal {M}})(M)) \end{aligned}$$

and \(\dim ({\mathcal {T}}_{\mathcal {M}})_{(n)}(M) =\dim (\text {gr}_n({\mathcal {T}}_{\mathcal {M}})(M))\) and hence by induction

$$\begin{aligned} \dim ({\mathcal {T}}_{\mathcal {M}})_{(k)}(M) \le \sum _{j\ge k}\dim (\text {gr}_j({\mathcal {T}}_{\mathcal {M}})(M)), \end{aligned}$$

which gives

$$\begin{aligned} \dim ({\mathcal {T}}_{\mathcal {M}}(M)) =\dim \left( ({\mathcal {T}}_{\mathcal {M}})_{(-1)}(M)\right) \le \dim \left( \text {gr}({\mathcal {T}}_{\mathcal {M}})(M)\right) . \end{aligned}$$

In particular, \(\dim ({\mathcal {T}}_{\mathcal {M}}(M))\) is finite. \(\square \)

Remark 21

The proof of the preceding proposition also shows the following inequality:

$$\begin{aligned} \dim (\mathrm {Vec}({\mathcal {M}}))\le \dim (\mathrm {Vec}(\text {gr}\,{\mathcal {M}})) \end{aligned}$$

7 The automorphism group

In this section, the automorphism group of a compact complex supermanifold is defined. This is done via the formalism of Harish-Chandra pairs for complex Lie supergroups (cf. [24]). The underlying classical Lie group is \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) and the Lie superalgebra is \(\mathrm {Vec}({\mathcal {M}})\), the Lie superalgebra of super vector fields on \({\mathcal {M}}\). Moreover, we prove that the automorphism group satisfies a universal property.

Consider the representation \(\alpha \) of \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) on \(\mathrm {Vec}({\mathcal {M}})\) given by

$$\begin{aligned} \alpha (g)(X)=g_*(X)=(g^{-1})^*\circ X\circ g^*\ \ \ \text { for } \ \ \ g\in {{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}}), \, X\in \mathrm {Vec}({\mathcal {M}}). \end{aligned}$$

This representation \(\alpha \) preserves the parity on \(\mathrm {Vec}({\mathcal {M}})\), and its restriction to \(\mathrm {Vec}_{{{\bar{0}}}}({\mathcal {M}})\) coincides with the adjoint action of \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) on its Lie algebra \(\mathrm {Lie}({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}}))\cong \mathrm {Vec}_{{{\bar{0}}}} ({\mathcal {M}})\). Moreover, the differential \((d\alpha )_{\mathrm {id}}\) at the identity \(\mathrm {id}\in {{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) is the adjoint representation of \(\mathrm {Vec}_{{{\bar{0}}}}({\mathcal {M}})\) on \(\mathrm {Vec}({\mathcal {M}})\):

Let X and Y be super vector fields on \({\mathcal {M}}\). Assume that X is even and let \(\varphi ^X\) denote the corresponding one-parameter subgroup. Then we have

$$\begin{aligned} (d\alpha )_{\mathrm {id}}(X)(Y)=\left. \frac{\partial }{\partial t}\right| _0 (\varphi ^X_t)_*(Y)=[X,Y]; \end{aligned}$$

see e.g. [2], Corollary 3.8. Therefore, the pair \(({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}}),\mathrm {Vec}({\mathcal {M}}))\) together with the representation \(\alpha \) is a complex Harish-Chandra pair, and using the equivalence between the category of complex Harish-Chandra pairs and complex Lie supergroups (cf. [24], \(\S \) 2), we can define the automorphism group of a compact complex supermanifold \({\mathcal {M}}\) as follows:

Definition 2

Define the automorphism group \({{\mathrm{Aut}}}({\mathcal {M}})\) of a compact complex supermanifold to be the unique complex Lie supergroup associated with the Harish-Chandra pair \(({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}}), \mathrm {Vec}({\mathcal {M}}))\) with adjoint representation \(\alpha \).

Since the action \(\psi :{{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\times {\mathcal {M}}\rightarrow {\mathcal {M}}\) induces the inclusion \(\mathrm {Vec}_{{{\bar{0}}}}({\mathcal {M}}) \hookrightarrow \mathrm {Vec}({\mathcal {M}})\) as infinitesimal action (see Corollary 18), there exists a Lie supergroup action \(\varPsi :{{\mathrm{Aut}}}({\mathcal {M}})\times {\mathcal {M}} \rightarrow {\mathcal {M}}\) with the identity \(\mathrm {Vec}({\mathcal {M}}) \rightarrow \mathrm {Vec}({\mathcal {M}})\) as induced infinitesimal action and \(\varPsi |_{{{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\times {\mathcal {M}}}=\psi \) (cf. Theorem 5.35 in [2]).

The automorphism group together with \(\varPsi \) satisfies a universal property:

Theorem 22

Let \({\mathcal {G}}\) be a complex Lie supergroup with a holomorphic action \(\varPsi _{{\mathcal {G}}}:{\mathcal {G}}\times {\mathcal {M}}\rightarrow {\mathcal {M}}\). Then there is a unique morphism \(\sigma :{\mathcal {G}}\rightarrow {{\mathrm{Aut}}}({\mathcal {M}})\) of Lie supergroups such that the diagram

is commutative.

Proof

Let G be the underlying Lie group of \({\mathcal {G}}\). For each \(g\in G\), we have a morphism \(\varPsi _{\mathcal {G}}(g):{\mathcal {M}}\rightarrow {\mathcal {M}}\) by setting \((\varPsi _{\mathcal {G}}(g))^*=\mathrm {ev}_g\circ (\varPsi _{\mathcal {G}})^*\). This morphism \(\varPsi _{\mathcal {G}}(g)\) is an automorphism of \({\mathcal {M}}\) with inverse \(\varPsi _{\mathcal {G}}(g^{-1})\) and gives rise to a group homomorphism \({{\tilde{\sigma }}}:G\rightarrow {{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\), \(g\mapsto \varPsi _{\mathcal {G}}(g)\).

Let \({\mathfrak {g}}\) denote the Lie superalgebra (of right-invariant super vector fields) of \({\mathcal {G}}\), and \(d\varPsi _{\mathcal {G}}:{\mathfrak {g}}\rightarrow \mathrm {Vec}({\mathcal {M}})\) the infinitesimal action induced by \(\varPsi _{\mathcal {G}}\). The restriction of \(d\varPsi _{\mathcal {G}}\) to the even part \({\mathfrak {g}}_{{{\bar{0}}}}=\mathrm {Lie}(G)\) of \({\mathfrak {g}}\) coincides with the differential \((d{{\tilde{\sigma }}})_e\) of \({{\tilde{\sigma }}}\) at the identity \(e\in G\).

Moreover, if \(\alpha _{\mathcal {G}}\) denotes the adjoint action of G on \({\mathfrak {g}}\), and \(\alpha \) denotes, as before, the adjoint action of \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) on \(\mathrm {Vec}({\mathcal {M}})\), we have

$$\begin{aligned} d\varPsi _{\mathcal {G}}(\alpha _{\mathcal {G}}(g)(X))&=(\varPsi _{\mathcal {G}}(g^{-1}))^*\circ d\varPsi _{\mathcal {G}}(X) \circ (\varPsi _{\mathcal {G}}(g))^*\\&=({{\tilde{\sigma }}}(g^{-1}))^*\circ d\varPsi _{\mathcal {G}}(X)\circ ({{\tilde{\sigma }}}(g))^*\\&=\alpha ({{\tilde{\sigma }}}(g))( d\varPsi _{\mathcal {G}}(X)) \end{aligned}$$

for any \(g\in G\), \(X\in {\mathfrak {g}}\). Using the correspondence between Lie supergroups and Harish-Chandra pairs, it follows that there is a unique morphism \(\sigma :{\mathcal {G}} \rightarrow {{\mathrm{Aut}}}({\mathcal {M}})\) of Lie supergroups with underlying map \({{\tilde{\sigma }}}\) and derivative \(d\varPsi _{\mathcal {G}}: {\mathfrak {g}}\rightarrow \mathrm {Vec}({\mathcal {M}})\) (see e.g. [24], \(\S \) 2), and \(\sigma \) satisfies \(\varPsi \circ (\sigma \times \mathrm {id}_{\mathcal {M}})=\varPsi _{\mathcal {G}}\).

The uniqueness of \(\sigma \) follows from the fact that each morphism \(\tau :{\mathcal {G}}\rightarrow {{\mathrm{Aut}}}({\mathcal {M}})\) of Lie supergroups fulfilling the same properties as \(\sigma \) necessarily induces the map \(d\varPsi _{\mathcal {G}}: {\mathfrak {g}}\rightarrow \mathrm {Vec}({\mathcal {M}})\) on the level of Lie superalgebras and its underlying map \({{\tilde{\tau }}}\) has to satisfy \({{\tilde{\tau }}}(g)=\varPsi _{\mathcal {G}}(g)={{\tilde{\sigma }}}(g)\). \(\square \)

Remark 23

Since the morphism \(\sigma \) in Theorem 22 is unique, the automorphism group of a compact complex supermanifold \({\mathcal {M}}\) is the unique Lie supergroup satisfying the universal property formulated in Theorem 22.

Remark 24

We say that a real Lie supergroup \({\mathcal {G}}\) acts on \({\mathcal {M}}\) by holomorphic transformations if the underlying Lie group G acts on the complex manifold M by holomorphic transformations and if there is a homomorphism of Lie superalgebras \({\mathfrak {g}}\rightarrow \mathrm {Vec}({\mathcal {M}})\) which is compatible with the action of G on M. Using the theory of Harish-Chandra pairs, we also have the Lie supergroup \({\mathcal {G}}^{\mathbb {C}}\), the universal complexification of \({\mathcal {G}}\); see [14]. The underlying Lie group of \({\mathcal {G}}^{\mathbb {C}}\) is the universal complexification \(G^{\mathbb {C}}\) of the Lie group G. Let \({\mathfrak {g}}={\mathfrak {g}}_{{{\bar{0}}}}\oplus {\mathfrak {g}}_{{{\bar{1}}}}\) denote the Lie superalgebra of \({\mathcal {G}}\), \({\mathfrak {g}}_{{{\bar{0}}}}\) the Lie algebra of G. Then the Lie algebra \({\mathfrak {g}}_{{{\bar{0}}}}^{\mathbb {C}}\) of \(G^{\mathbb {C}}\) is a quotient of \({\mathfrak {g}}_{{{\bar{0}}}}\otimes {\mathbb {C}}\), and the Lie superalgebra of \({\mathcal {G}}^{\mathbb {C}}\) can be realized as \({\mathfrak {g}}_{{{\bar{0}}}}^{\mathbb {C}}\oplus ({\mathfrak {g}}_{{{\bar{1}}}}\otimes {\mathbb {C}})\). The action of G on \({\mathcal {M}}\) extends to a holomorphic \(G^{\mathbb {C}}\)-action on \({\mathcal {M}}\), and the homomorphism \({\mathfrak {g}}\rightarrow \mathrm {Vec}({\mathcal {M}})\) extends to a homomorphism \({\mathfrak {g}}_{{{\bar{0}}}}^{\mathbb {C}}\oplus ({\mathfrak {g}}_{{{\bar{1}}}}\otimes {\mathbb {C}}) \rightarrow \mathrm {Vec}({\mathcal {M}})\) of complex Lie superalgebras, which is compatible with the \(G^{\mathbb {C}}\)-action on \({\mathcal {M}}\). Thus, we have a holomorphic \({\mathcal {G}}^{\mathbb {C}}\)-action on \({\mathcal {M}}\) extending the \({\mathcal {G}}\)-action. Moreover, there is a morphism \(\sigma :{\mathcal {G}}^{\mathbb {C}}\rightarrow {{\mathrm{Aut}}}({\mathcal {M}})\) of Lie supergroups as in Theorem 22.

Example 25

Let \({\mathcal {M}}={\mathbb {C}}^{0|1}\). Denoting the odd coordinate on \({\mathbb {C}}^{0|1}\) by \(\xi \), each super vector field on \({\mathbb {C}}^{0|1}\) is of the form \(X=a \xi \frac{\partial }{\partial \xi }+ b\frac{\partial }{\partial \xi }\) for \(a,b\in {\mathbb {C}}\). The flow \(\varphi :{\mathbb {C}}\times {\mathcal {M}}\rightarrow {\mathcal {M}}\) of \(a \xi \frac{\partial }{\partial \xi }\) is given by \((\varphi _t)^*( \xi )=e^{at}\xi \), and the flow \(\psi :{\mathbb {C}}^{0|1}\times {\mathcal {M}}\rightarrow {\mathcal {M}}\) of \(b\frac{\partial }{\partial \xi }\) by \(\psi ^*(\xi )=b\tau +\xi \). Let \(X_0=\xi \frac{\partial }{\partial \xi }\) and \(X_1=\frac{\partial }{\partial \xi }\). Then \(\mathrm {Vec}({\mathbb {C}}^{0|1})={\mathbb {C}} X_0\oplus {\mathbb {C}} X_1= {\mathbb {C}}^{1|1}\), where the Lie algebra structure on \({\mathbb {C}}^{1|1}\) is given by \([X_0,X_1]= -X_1\) and \([X_1,X_1]=0\). Note that this Lie superalgebra is isomorphic to the Lie superalgebra of right-invariant vector fields on the Lie supergroup \(({\mathbb {C}}^{1|1}, \mu _{0,1})\), where the multiplication \(\mu =\mu _{0,1}\) is given by \(\mu ^*(t)=t_1+t_2\) and \(\mu ^*(\tau )=\tau _1+e^{t_1} \tau _2\); for the Lie supergroup structures on \({\mathbb {C}}^{1|1}\) see e.g. [12], Lemma 3.1. In particular, the Lie superalgebra \(\mathrm {Vec}({\mathbb {C}}^{0|1})\) is not abelian.

Since each automorphism \(\varphi \) of \({\mathbb {C}}^{0|1}\) is given by \(\varphi ^*(\xi )=c\cdot \xi \) for some \(c\in {\mathbb {C}}\), \(c\ne 0\), we have \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathbb {C}}^{0|1})\cong {\mathbb {C}}^*\).

8 The functor of points of the automorphism group

In [22], the diffeomorphism supergroup of a real compact supermanifold is proven to carry the structure of a Fréchet Lie supergroup. This diffeomorphism supergroup is defined using the “functor of points” approach to supermanifolds, i.e. a supermanifold is a representable contravariant functor from the category of supermanifolds to the category of sets. Starting with a supermanifold \({\mathcal {M}}\) we define the corresponding functor \(\mathrm {Hom}(-,{\mathcal {M}})\) by the assignment \({\mathcal {N}}\mapsto \mathrm {Hom}({\mathcal {N}},{\mathcal {M}})\), where \(\mathrm {Hom}({\mathcal {N}},{\mathcal {M}})\) denotes the set of morphisms of supermanifolds \({\mathcal {N}}\rightarrow {\mathcal {M}}\), and for morphisms \(\alpha :{\mathcal {N}}_1\rightarrow {\mathcal {N}}_2\) between supermanifolds \({\mathcal {N}}_1\) and \({\mathcal {N}}_2\) we define \(\mathrm {Hom}(-,{\mathcal {M}})(\alpha ): \mathrm {Hom}({\mathcal {N}}_2,{\mathcal {M}}) \rightarrow \mathrm {Hom}({\mathcal {N}}_1,{\mathcal {M}})\) by \(\varphi \mapsto \varphi \circ \alpha \).

In analogy to the definition in [22] for the diffeomorphism supergroup, a functor \({\overline{{{\mathrm{Aut}}}}}({\mathcal {M}})\) associated with a complex supermanifold \({\mathcal {M}}\) can be defined. In the case of a compact complex supermanifold \({\mathcal {M}}\), the automorphism Lie supergroup as defined in Section 7 represents the functor \({\overline{{{\mathrm{Aut}}}}}({\mathcal {M}})\), i.e. the functors \({\overline{{{\mathrm{Aut}}}}}({\mathcal {M}})\) and \(\mathrm {Hom}(-,{{\mathrm{Aut}}}({\mathcal {M}}))\) are isomorphic. This is proven in [3], Section 5.4. Here we give an outline of the main steps in the proof.

Definition 3

Let \({\mathcal {M}}\) be a complex supermanifold. We define the functor \({\overline{{{\mathrm{Aut}}}}}({\mathcal {M}})\) from the category of supermanifolds to the category of groups as follows:

On objects, we define \({\overline{{{\mathrm{Aut}}}}}({\mathcal {M}})\) by the assignment

$$\begin{aligned} {\mathcal {N}}\mapsto \{\varphi :{\mathcal {N}}\times {\mathcal {M}}\rightarrow {\mathcal {N}}\times {\mathcal {M}}\,|\, \varphi \text { is invertible, and } \mathrm {pr}_{{\mathcal {N}}}\circ \varphi =\mathrm {pr}_{{\mathcal {N}}}\}, \end{aligned}$$

where \(\mathrm {pr}_{{\mathcal {N}}}:{\mathcal {N}}\times {\mathcal {M}}\rightarrow {\mathcal {N}}\) is the projection. For morphisms \(\alpha :{\mathcal {N}}_1\rightarrow {\mathcal {N}}_2\), we set \({\overline{{{\mathrm{Aut}}}}}({\mathcal {M}})(\alpha ): {\overline{{{\mathrm{Aut}}}}}({\mathcal {M}})({\mathcal {N}}_2)\rightarrow {\overline{{{\mathrm{Aut}}}}}({\mathcal {M}})({\mathcal {N}}_1)\),

$$\begin{aligned} \varphi \mapsto (\mathrm {id}_{{\mathcal {N}}_1}\times (\mathrm {pr}_{\mathcal {M}}\circ \varphi \circ (\alpha \times \mathrm {id}_{\mathcal {M}}))) \circ (\mathrm {diag}\times \mathrm {id}_{\mathcal {M}}), \end{aligned}$$

denoting by \(\mathrm {diag}:{\mathcal {N}}_1\rightarrow {\mathcal {N}}_1\times {\mathcal {N}}_1\) the diagonal map and by \(\mathrm {pr}_{\mathcal {M}}\) the projection onto \({\mathcal {M}}\). Thus \({\overline{{{\mathrm{Aut}}}}}({\mathcal {M}})(\alpha )(\varphi )\) is the unique automorphism \(\psi :{\mathcal {N}}_1\times {\mathcal {M}}\rightarrow {\mathcal {N}}_1\times {\mathcal {M}}\) with \(\mathrm {pr}_{{\mathcal {N}}_1}\circ \psi =\mathrm {pr}_{{\mathcal {N}}_1}\) and \(\mathrm {pr}_{{\mathcal {M}}}\circ \psi =\mathrm {pr}_{\mathcal {M}}\circ \varphi \circ (\alpha \times \mathrm {id}_{\mathcal {M}})\).

The group structure on \({\overline{{{\mathrm{Aut}}}}}({\mathcal {M}})({\mathcal {N}})\) is defined by the composition and inversion of automorphisms \({\mathcal {N}}\times {\mathcal {M}}\rightarrow {\mathcal {N}}\times {\mathcal {M}}\), and the neutral element is the identity map \({\mathcal {N}}\times {\mathcal {M}} \rightarrow {\mathcal {N}}\times {\mathcal {M}}\).

Let \(\chi :{\mathcal {N}}\rightarrow {{\mathrm{Aut}}}({\mathcal {M}})\) be an arbitrary morphism of complex supermanifolds and let \(\varPsi :{{\mathrm{Aut}}}({\mathcal {M}})\times {\mathcal {M}}\rightarrow {\mathcal {M}}\) denote the natural action of \({{\mathrm{Aut}}}({\mathcal {M}})\) on \({\mathcal {M}}\). Then the composition

$$\begin{aligned} \varphi _\chi =(\mathrm {id}_{{\mathcal {N}}}\times (\varPsi \circ (\chi \times \mathrm {id}_{\mathcal {M}})))\circ (\mathrm {diag}\times \mathrm {id}_{\mathcal {M}}) \end{aligned}$$

is an invertible map \({\mathcal {N}}\times {\mathcal {M}}\rightarrow {\mathcal {N}} \times {\mathcal {M}}\) with \(\mathrm {pr}_{{\mathcal {N}}}=\mathrm {pr}_{{\mathcal {N}}}\circ \varphi _\chi \). This defines a natural transformation:

Lemma 26

The assignments \(\mathrm {Hom}({\mathcal {N}},{{\mathrm{Aut}}}({\mathcal {M}})) \rightarrow {\overline{{{\mathrm{Aut}}}}}({\mathcal {M}})({\mathcal {N}})\), \(\chi \mapsto \varphi _\chi \), define a natural transformation \(\mathrm {Hom}(-,{{\mathrm{Aut}}}({\mathcal {M}}))\rightarrow {\overline{{{\mathrm{Aut}}}}}({\mathcal {M}})\).

This statement of the lemma can be verified by direct calculations; see also Lemma 5.4.2 in [3].

The natural transformation between \(\mathrm {Hom}(-,{{\mathrm{Aut}}}({\mathcal {M}}))\) and \({\overline{{{\mathrm{Aut}}}}}({\mathcal {M}})\) is actually an isomorphism of functors. The injectivity of the assignment \(\chi \mapsto \varphi _\chi \) follows from the fact that the \({{\mathrm{Aut}}}({\mathcal {M}})\)-action on \({\mathcal {M}}\) is effective. As a generalization of the classical definition of effectiveness, we call an action \(\varPsi \) of a Lie supergroup \({\mathcal {G}}\) on a supermanifold \({\mathcal {M}}\) effective if for arbitrary morphisms \(\chi _1,\chi _2:{\mathcal {N}}\rightarrow {\mathcal {G}}\) of supermanifolds the equality

$$\begin{aligned} \varPsi \circ (\chi _1\times \mathrm {id}_{\mathcal {M}})=\varPsi \circ (\chi _2\times \mathrm {id}_{\mathcal {M}}) \end{aligned}$$

implies \(\chi _1=\chi _2\); cf. Section 2.5 in [3].

In the proof of the surjectivity a “normal form” of the pullback of automorphisms \(\varphi :{\mathbb {C}}^{0|k}\times {\mathcal {M}}\rightarrow {\mathbb {C}}^{0|k}\times {\mathcal {M}}\) with \(\mathrm {pr}_{{\mathbb {C}}^{0|k}}\circ \varphi =\mathrm {pr}_{{\mathbb {C}}^{0|k}}\) is used. Let \({\mathcal {M}}\) be a complex supermanifold and \(\varphi :{\mathbb {C}}^{0|k}\times {\mathcal {M}}\rightarrow {\mathbb {C}}^{0|k}\times {\mathcal {M}}\) be an invertible morphism with \(\mathrm {pr}_{{\mathbb {C}}^{0|k}}\circ \varphi =\mathrm {pr}_{{\mathbb {C}}^{0|k}}\). Let \(\iota :{\mathcal {M}}\hookrightarrow \{0\}\times {\mathcal {M}} \subset {\mathbb {C}}^{0|k}\times {\mathcal {M}}\) denote the canonical inclusion. The composition \({{\bar{\varphi }}}=\mathrm {pr}_{\mathcal {M}}\circ \varphi \circ \iota \) is an automorphism of \({\mathcal {M}}\). Then \(\varphi \) is uniquely determined by \({{\bar{\varphi }}}\) and a set of super vector fields on \({\mathcal {M}}\):

Lemma 27

Let \(\varphi :{\mathbb {C}}^{0|k}\times {\mathcal {M}}\rightarrow {\mathbb {C}}^{0|k}\times {\mathcal {M}}\) be an invertible morphism with \(\mathrm {pr}_{ {\mathbb {C}}^{0|k}}\circ \varphi =\mathrm {pr}_{{\mathbb {C}}^{0|k}}\). Let \(\tau _1,\ldots , \tau _k\) denote coordinates on \({\mathbb {C}}^{0|k}\subset {\mathbb {C}}^{0|k}\times {\mathcal {M}}\). Then there are super vector fields \(X_{\nu }\) on \({\mathcal {M}}\), of parity \(|\nu |\) for \(\nu \in ({\mathbb {Z}}_2)^k\), \(\nu \ne 0\), such that

$$\begin{aligned} \varphi ^*=(\mathrm {id}_{{\mathbb {C}}^{0|k}}\times {{\bar{\varphi }}})^* \exp \left( \sum _{\nu \ne 0} \tau ^\nu X_{\nu }\right) , \end{aligned}$$

By \(\tau ^\nu X_{\nu }\) we mean the super vector field on \({\mathbb {C}}^{0|k}\times {\mathcal {M}}\) which is induced by the extension of the super vector field \(X_{\nu }\) on \({\mathcal {M}}\) to a super vector field on the product \({\mathbb {C}}^{0|k}\times {\mathcal {M}}\) followed by the multiplication with \(\tau ^\nu =\tau _1^{\nu _1}\ldots \tau _k^{\nu _k}\). In other words for \(U\subseteq M\) open we have \(\tau ^\nu X_\nu (f)=0\) for \(f\in {\mathcal {O}}_{{\mathbb {C}}^{0|k}}(\{0\}) \subset {\mathcal {O}}_{{\mathbb {C}}^{0|k}\times {\mathcal {M}}}(\{0\}\times U)\) and \((\tau ^\nu X_\nu )(g)=\tau ^\nu X_{\nu }(g)\) for \(g\in {\mathcal {O}}_{{\mathcal {M}}}(U)\subset {\mathcal {O}}_{{\mathbb {C}}^{0|k}\times {\mathcal {M}}}(\{0\}\times U)\) considering \(X_\nu (g)\) as a function on the product.

Moreover,

$$\begin{aligned} \exp \left( \sum _{\nu \ne 0} \tau ^\nu X_\nu \right) =\sum _{n \ge 0}\frac{1}{n!} \left( \sum _{\nu \ne 0} \tau ^\nu X_\nu \right) ^n \end{aligned}$$

is a finite sum since \(\left( \sum _{\nu \ne 0} \tau ^\nu X_\nu \right) ^{k+1}=0\).

A version of this lemma is also proven in [22], Theorem 5.1. A different proof using the relation between nilpotent even super vector fields on a supermanifold and morphisms of this supermanifold satisfying a certain nilpotency condition as formulated in Sect. 2 is also possible; for details see also [3], Lemma 5.4.3.

Using the normal form of the lemma, we can prove that the assignment \(\chi \mapsto \varphi _\chi \) defines a surjective map by directly constructing a morphism \(\chi \) with \(\varphi _\chi =\varphi \) for any \(\varphi :{\mathcal {N}}\times {\mathcal {M}}\rightarrow {\mathcal {N}}\times {\mathcal {M}}\) with \(\mathrm {pr}_{{\mathcal {N}}}\circ \varphi =\mathrm {pr}_{{\mathcal {N}}}\). It is here enough to prove this statement locally (in \({\mathcal {N}}\)) and thus to consider the case where \({\mathcal {N}}=N\times {\mathbb {C}}^{0|k}\) for a classical complex manifold N. In the following we indicate how such a morphism \(\chi \) can be defined; for the proof that \(\chi \) fulfills the desired property \(\varphi _\chi =\varphi \) see Proposition 5.4.4 in [3].

Let \(\varphi :N\times {\mathbb {C}}^{0|k}\times {\mathcal {M}}\rightarrow N\times {\mathbb {C}}^{0|k}\times {\mathcal {M}}\) be an invertible morphism with \(\mathrm {pr}_{N\times {\mathbb {C}}^{0|k}}\circ \varphi =\mathrm {pr}_{N\times {\mathbb {C}}^{0|k}}\). Each \(z\in N\) induces an invertible morphism \(\varphi _z:{\mathbb {C}}^{0|k}\times {\mathcal {M}}\rightarrow {\mathbb {C}}^{0|k}\times {\mathcal {M}}\) with \(\mathrm {pr}_{{\mathbb {C}}^{0|k}}\circ \varphi _z=\mathrm {pr}_{{\mathbb {C}}^{0|k}}\), and the family \(\varphi _z\), \(z\in N\), uniquely determines \(\varphi \).

Let \(X_{\nu , z}\) be super vector fields on \({\mathcal {M}}\) of parity \(|\nu |\), \(\nu \in ({\mathbb {Z}}_2)^k\), \(\nu \ne 0\), and \({{\bar{\varphi }}}_z:{\mathcal {M}}\rightarrow {\mathcal {M}}\) automorphisms such that \(\varphi _z^*=(\mathrm {id}_{{\mathbb {C}}^{0|k}}\times {{\bar{\varphi }}}_z)^* \exp \left( \sum _{\nu \ne 0} \tau ^\nu X_{\nu ,z}\right) \) as in Lemma 27. Since \(\varphi \) is holomorphic, the coefficients of the super vector fields \(X_{\nu ,z}\) and the pullbacks \({{\bar{\varphi }}}_z^*\) in local coordiantes depend holomorphically on \(z\in N\). Each \({{\bar{\varphi }}}_z\) is the automorphism of \({\mathcal {M}}\) induced by the evalutation in \((z,0)\in N\times {\mathbb {C}}^{0|k}\) and an element of \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) by definition. Let \(\mathrm {ev}_{{{\bar{\varphi }}}_z}\) denote the evaluation in \({{\bar{\varphi }}}_z\), i.e. \(\mathrm {ev}_{{{\bar{\varphi }}}_z}\) is the pullback of the canonical inclusion \(\{{{\bar{\varphi }}}_z\}\hookrightarrow {{\mathrm{Aut}}}({\mathcal {M}})\), and let \(\mathrm {pr}_{{{\mathrm{Aut}}}({\mathcal {M}})}:N\times {\mathbb {C}}^{0|k}\times {{\mathrm{Aut}}}({\mathcal {M}}) \rightarrow {{\mathrm{Aut}}}({\mathcal {M}})\) be the projection. We define \(\chi :N\times {\mathbb {C}}^{0|k}\rightarrow {{\mathrm{Aut}}}({\mathcal {M}})\) as the morphism whose underlying map is \(\{z\}\hookrightarrow \{{{\bar{\varphi }}}_z\}\subset {{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) and whose pullback evaluated in \(z\in N\) is

$$\begin{aligned} \chi _z^*=(\mathrm {id}_{{\mathbb {C}}^{0|k}}^*\otimes \mathrm {ev}_{{{\bar{\varphi }}}_z})\circ \exp \left( \sum _{\nu \ne 0}\tau ^ \nu (X_{\nu ,z})_R\right) \circ \mathrm {pr}_{{{\mathrm{Aut}}}({\mathcal {M}})}^*, \end{aligned}$$

where \((X_{\nu ,z})_R\) denotes the right-invariant super vector field on \({{\mathrm{Aut}}}({\mathcal {M}})\) corresponding to the super vector field \(X_{\nu ,z}\) on \({\mathcal {M}}\) which is an element of the Lie superalgebra \(\mathrm {Vec}({\mathcal {M}})\) of \({{\mathrm{Aut}}}({\mathcal {M}})\).

The next proposition is then a consequence of Lemma 26 and the surjectivity of the assignment \(\chi \mapsto \varphi _\chi \).

Proposition 28

(See [3], Corollary 5.4.5) The functors \({\overline{{{\mathrm{Aut}}}}}({\mathcal {M}})\) and \(\mathrm {Hom}(-,{{\mathrm{Aut}}}({\mathcal {M}}))\) are isomorphic. This isomorphism is realized by the natural transformation introduced in Lemma 26.

9 The case of a superdomain with bounded underlying domain

In the classical case, the automorphism group of a bounded domain \(U\subset {\mathbb {C}}^m\) is a (real) Lie group (see Theorem 13 in “Sur les groupes de transformations analytiques” in [8]). If \({\mathcal {U}}\subset {\mathbb {C}}^{m|n}\) is a superdomain whose underlying set U is a bounded domain in \({\mathbb {C}}^m\), it is in general not possible to endow its set of automorphisms with the structure of a Lie group such that the action on \({\mathcal {U}}\) is smooth, as will be illustrated in an example. In particular, there is no Lie supergroup satisfying the universal property as the automorphism group of a compact complex supermanifold \({\mathcal {M}}\) does as formulated in Theorem 22.

Example 29

Consider a superdomain \({\mathcal {U}}\) of dimension (1|2) whose underlying set is a bounded domain \(U\subset {\mathbb {C}}\). Let \(z,\theta _1,\theta _2\) denote coordinates for \({\mathcal {M}}\). For any holomorphic function f on U, define the even super vector field \(X_f=f(z)\theta _1\theta _2\frac{\partial }{\partial z}\). The reduced vector field \({\tilde{X}}_f=0\) is completely integrable and thus the flow of \(X_f\) can be defined on \({\mathbb {C}}\times {\mathcal {U}}\) (cf. [12] Lemma 5.2). The flow is given by \((\varphi _t)^*(z)=z+t\cdot f(z)\theta _1\theta _2\) and \((\varphi _t)^*(\theta _j)=\theta _j\). For all holomorphic functions f and g we have \([X_f,X_g]=0\), and thus their flows locally commute (cf. [2], Corollary 3.8). Therefore, \(\{X_f|\,f\in {\mathcal {O}}(U)\}\cong {\mathcal {O}}(U)\) is an uncountably infinite-dimensional abelian Lie algebra. If the set of automorphisms of \({\mathcal {U}}\) carried the structure of a Lie group such that its action on \({\mathcal {U}}\) was smooth, its Lie algebra would necessarily contain \(\{X_f|\,f\in {\mathcal {O}}(U)\}\cong {\mathcal {O}}(U)\) as a Lie subalgebra, which is not possible.

10 Examples

In this section, we determine the automorphism group \({{\mathrm{Aut}}}({\mathcal {M}})\) for some complex supermanifolds \({\mathcal {M}}\) with underlying manifold \(M={\mathbb {P}}_1{\mathbb {C}}\).

Let \(L_1\) denote the hyperplane bundle on \(M={\mathbb {P}}_1{\mathbb {C}}\) with sheaf of sections \({\mathcal {O}}(1)\), and \(L_k=(L_1)^{\otimes k}\) the line bundle of degree k, \(k\in {\mathbb {Z}}\), on \({\mathbb {P}}_1{\mathbb {C}}\), and sheaf of sections \({\mathcal {O}}(k)\). Each holomorphic vector bundle on \({\mathbb {P}}_1{\mathbb {C}}\) is isomorphic to a direct sum of line bundles \(L_{k_1}\oplus \ldots \oplus L_{k_n}\) (see [11]). Therefore, if \({\mathcal {M}}\) is a split supermanifold with \(M={\mathbb {P}}_1{\mathbb {C}}\) and \(\dim {\mathcal {M}}=(1|n)\), there exist \(k_1,\ldots , k_n\in {\mathbb {Z}}\) such that the structure sheaf \({\mathcal {O}}_{{\mathcal {M}}}\) of \({\mathcal {M}}\) is isomorphic to

$$\begin{aligned} \bigwedge ({\mathcal {O}}(k_1)\oplus \ldots \oplus {\mathcal {O}}(k_n)). \end{aligned}$$

Let \(U_j=\{[z_0:z_1]\in {\mathbb {P}}_1{\mathbb {C}}\,|\, z_j\ne 0\}\), \(j=1,2\), and \({\mathcal {U}}_j=(U_j,{\mathcal {O}}_{{\mathcal {M}}}|_{U_j})\). Moreover, define \({U_0}^*=U_0\setminus \{[1:0]\}\) and \({U_1}^*=U_1\setminus \{[0:1]\}\), and let \({{\mathcal {U}}_j}^*=({U_j}^*,{\mathcal {O}}_{{\mathcal {M}}}|_{{U_j}^*})\). We can now choose local coordinates \(z,\theta _1,\ldots , \theta _n\) for \({\mathcal {M}}\) on \(U_0\), and local coordinates \(w,\eta _1,\ldots ,\eta _n\) on \(U_1\) so that the transition map \(\chi :{{\mathcal {U}}_0}^*\rightarrow {{\mathcal {U}}_1}^*\), which determines the supermanifold structure of \({\mathcal {M}}\), is given by

$$\begin{aligned} \chi ^*(w)=\frac{1}{z} \ \ \text { and }\ \ \chi ^*(\eta _j)={z^{k_j}}\theta _j. \end{aligned}$$

Example 30

Let \({\mathcal {M}}=({\mathbb {P}}_1{\mathbb {C}}, {\mathcal {O}}_{\mathcal {M}})\) be a complex supermanifold of dimension (1|1). Since the odd dimension is 1, the supermanifold \({\mathcal {M}}\) has to be split. Let \(-k\in {\mathbb {Z}}\) be the degree of the associated line bundle. Choose local coordinates \(z,\theta \) for \({\mathcal {M}}\) on \(U_0\) and \(w,\eta \) on \(U_1\) as above so that the transition map \(\chi :{{\mathcal {U}}_0}^*\rightarrow {{\mathcal {U}}_1}^*\) is given by \(\chi ^*(w)=\frac{1}{z}\) and \(\chi ^*(\eta )=\frac{1}{z^k}\theta \).

We first want to determine the Lie superalgebra \(\mathrm {Vec}({\mathcal {M}})\) of super vector fields on \({\mathcal {M}}\). A calculation in local coordinates verifying the compatibility condition with the transition map \(\chi \) yields that the restriction to \(U_0\) of any super vector field on \({\mathcal {M}}\) is of the form

$$\begin{aligned} \left( (\alpha _0+\alpha _1z+\alpha _2z^2)\frac{\partial }{\partial z} +(\beta +k\alpha _2 z)\theta \frac{\partial }{\partial \theta }\right) +\left( p(z)\frac{\partial }{\partial \theta } +q(z)\theta \frac{\partial }{\partial z}\right) , \end{aligned}$$

where \(\alpha _0,\alpha _1,\alpha _2,\beta \in {\mathbb {C}}\), p is a polynomial of degree at most k, and q is a polynomial of degree at most \(2-k\). If \(k<0\) (respectively \(2-k<0\)), the polynomial p (respectively q) is 0. The Lie algebra \(\mathrm {Vec}_{{{\bar{0}}}}({\mathcal {M}})\) of even super vector fields is isomorphic to \({{\mathfrak {s}}}{{\mathfrak {l}}}_2({\mathbb {C}})\oplus {\mathbb {C}}\), where an isomorphism \({{\mathfrak {s}}}{{\mathfrak {l}}}_2({\mathbb {C}})\oplus {\mathbb {C}} \rightarrow \mathrm {Vec}_{{{\bar{0}}}}({\mathcal {M}})\) is given by

$$\begin{aligned} \left( \left( \begin{matrix} a&{}b\\ c&{}-a \end{matrix}\right) , d\right) \mapsto (-b-2az+cz^2)\frac{\partial }{\partial z} +((d-ka)+kcz)\theta \frac{\partial }{\partial \theta }. \end{aligned}$$

Note that since the odd dimension of \({\mathcal {M}}\) is 1 each automorphism \(\varphi :{\mathcal {M}}\rightarrow {\mathcal {M}}\) gives rise to an automorphism of the line bundle \(L_{-k}\) and vice versa. Hence, the automorphism group \({{\mathrm{Aut}}}(L_{-k})\) of the line bundle \(L_{-k}\) and \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) coincide.

A calculation yields that the group \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) of automorphisms \({\mathcal {M}}\rightarrow {\mathcal {M}}\) can be identified with \(\mathrm {PSL}_2({\mathbb {C}})\times {\mathbb {C}}^*\) if k is even and with \(\mathrm {SL}_2({\mathbb {C}})\times {\mathbb {C}}^*\) if k is odd. Consider the element \(\left( \left( {\begin{matrix} a &{} b\\ c&{} d\end{matrix}}\right) ,s\right) \), where \(s\in {\mathbb {C}}^*\) and \(\left( {\begin{matrix} a &{} b\\ c&{} d\end{matrix}}\right) \) is either an element of \(\mathrm {SL}_2({\mathbb {C}})\) or the representative of the corresponding class in \(\mathrm {PSL}_2({\mathbb {C}})\). The action of the corresponding element \(\varphi \in {{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) on \({\mathcal {M}}\) is then given by

$$\begin{aligned} \varphi ^*(z)=\frac{c+dz}{a+bz}\ \ \text { and } \ \ \varphi ^*(\theta )=\left( \frac{1}{(a+bz)^k}+s\right) \theta \end{aligned}$$

as a morphism over appropriate subsets of \(U_0\) and by

$$\begin{aligned} \varphi ^*(w)=\frac{aw+b}{cw+d}\ \ \text { and } \ \ \varphi ^*(\eta )=\left( \frac{1}{(cw+d)^k}+s\right) \eta \end{aligned}$$

over appropriate subsets of \(U_1\).

The Lie supergroup structure on \({{\mathrm{Aut}}}({\mathcal {M}})\) is now uniquely determined by \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\), \(\mathrm {Vec}({\mathcal {M}})\), and the adjoint action of \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) on \(\mathrm {Vec}({\mathcal {M}})\). Since \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) is a connected Lie group, it is enough to calculate the adjoint action of \(\mathrm {Vec}_{{{\bar{0}}}}({\mathcal {M}}) \cong {{\mathfrak {s}}}{{\mathfrak {l}}}_2{{\mathbb {C}}} \oplus {\mathbb {C}}\) on \(\mathrm {Vec}_{{{\bar{1}}}}({\mathcal {M}})\).

Let \(P_l\) denote the space of polynomials of degree at most l, and set \(P_l=\{0\}\) for \(l<0\). The space of odd super vector fields \(\mathrm {Vec}_{{{\bar{1}}}}({\mathcal {M}})\) is isomorphic to \(P_{k}\oplus P_{2-k}\) via \(\left( p(z)\frac{\partial }{\partial \theta } +q(z)\theta \frac{\partial }{\partial z}\right) \mapsto (p(z),q(z))\).

The element \(H=\left( {\begin{matrix} 1&{}0\\ 0&{}-1 \end{matrix}}\right) \in {{\mathfrak {s}}}{{\mathfrak {l}}}_2({\mathbb {C}})\subset {{\mathfrak {s}}}{{\mathfrak {l}}}_2({\mathbb {C}})\oplus {\mathbb {C}} \cong \mathrm {Vec}_{{{\bar{0}}}}({\mathcal {M}})\) corresponds to \(-2z\frac{\partial }{\partial z}-k\theta \frac{\partial }{\partial \theta }\). The adjoint action of this super vector field on the first factor \(P_k\) of \(\mathrm {Vec}_{{{\bar{1}}}}({\mathcal {M}})\) is given by by \(-2z\frac{\partial }{\partial z}+k\cdot \mathrm {Id}\), and on the second factor \(P_{2-k}\) by \(-2z\frac{\partial }{\partial z}+ (2-k)\cdot \mathrm {Id}\). Calculating the weights of the \({{\mathfrak {s}}}{{\mathfrak {l}}}_2({\mathbb {C}})\)-representation on \(P_k\) and \(P_{2-k}\), we get that \(P_k\) is the unique irreducible \((k+1)\)-dimensional representation and \(P_{2-k}\) the unique irreducible \((3-k)\)-dimensional representation. Moreover, a calculation yields that \(d\in {\mathbb {C}}\) corresponding to \(d\cdot \theta \frac{\partial }{\partial \theta }\in \mathrm {Vec}_{{{\bar{0}}}}({\mathcal {M}})\) acts on \(P_k\) by multiplication with \(-d\) and on \(P_{2-k}\) by multiplication with d.

If \(k<0\) or \(k>2\), we have

$$\begin{aligned} \left[ \mathrm {Vec}_{{{\bar{1}}}}({\mathcal {M}}), \mathrm {Vec}_{{{\bar{1}}}}({\mathcal {M}})\right] =0. \end{aligned}$$

In the case \(k=0\), we have \(P_k\cong {\mathbb {C}}\). Since \([\frac{\partial }{\partial \theta },q(z)\theta \frac{\partial }{\partial z}] =q(z)\frac{\partial }{\partial z}\) for any \(q\in P_2\), we get

$$\begin{aligned} \left[ \mathrm {Vec}_{{{\bar{1}}}}({\mathcal {M}}), \mathrm {Vec}_{{{\bar{1}}}}({\mathcal {M}})\right] =\left. \left\{ a(z)\frac{\partial }{\partial z}\,\right| \,a\in P_2\right\} \cong {{\mathfrak {s}}}{{\mathfrak {l}}}_2({\mathbb {C}}), \end{aligned}$$

and the map \(P_0\times P_2\rightarrow \mathrm {Vec}_{{{\bar{0}}}}({\mathcal {M}})\), \((X,Y)\mapsto [X,Y]\), corresponds to \({\mathbb {C}}\times P_2\rightarrow \mathrm {Vec}_{{{\bar{0}}}}({\mathcal {M}})\), \((p,q(z))\mapsto p\cdot q(z)\frac{\partial }{\partial z}\).

Similarly, if \(k=2\), we have \(P_{2-k}\cong {\mathbb {C}}\), and

$$\begin{aligned}{}[\mathrm {Vec}_{{{\bar{1}}}}({\mathcal {M}}),&\mathrm {Vec}_{{{\bar{1}}}}({\mathcal {M}})] =\left\{ \left. (\alpha _0+\alpha _1z+\alpha _2 z^2)\frac{\partial }{\partial z} +(\alpha _1+2\alpha _2 z)\theta \frac{\partial }{\partial \theta }\,\right| \, \alpha _j\in {\mathbb {C}}\right\} \\&\cong {{\mathfrak {s}}}{{\mathfrak {l}}}_2({\mathbb {C}}) \end{aligned}$$

since \([p(z)\frac{\partial }{\partial \theta }, \theta \frac{\partial }{\partial z}] =p(z)\frac{\partial }{\partial z}+ p'(z)\theta \frac{\partial }{\partial \theta }\), and the map \(P_2\times P_0\rightarrow \mathrm {Vec}_{{{\bar{0}}}}({\mathcal {M}})\), \((X,Y)\mapsto [X,Y]\), corresponds to \(P_2\times {\mathbb {C}}\rightarrow \mathrm {Vec}_{{{\bar{0}}}}({\mathcal {M}})\), \((p(z),q)\mapsto q\cdot p(z)\frac{\partial }{\partial z}+q\cdot p'(z)\theta \frac{\partial }{\partial \theta }\).

If \(k=1\), then \(P_k\oplus P_{2-k}\cong {\mathbb {C}}^2\oplus {\mathbb {C}}^2\). We have

$$\begin{aligned} \left[ \frac{\partial }{\partial \theta },\theta \frac{\partial }{\partial z}\right]= & {} \frac{\partial }{\partial z}, \left[ z\frac{\partial }{\partial \theta },\theta \frac{\partial }{\partial z}\right] =z\frac{\partial }{\partial z}+\theta \frac{\partial }{\partial \theta },\\ \left[ \frac{\partial }{\partial \theta },z\theta \frac{\partial }{\partial z}\right]= & {} z\frac{\partial }{\partial z}, \left[ z\frac{\partial }{\partial \theta },z\theta \frac{\partial }{\partial z}\right] =z^2\frac{\partial }{\partial z}+z\theta \frac{\partial }{\partial \theta }, \end{aligned}$$

and consequently \([\mathrm {Vec}_{{{\bar{1}}}}({\mathcal {M}}),\mathrm {Vec}_{{{\bar{1}}}}({\mathcal {M}})] =\mathrm {Vec}_{{{\bar{0}}}}({\mathcal {M}})\).

Remark that \({{\mathrm{Aut}}}({\mathcal {M}})\) carries the structure of a split Lie supergroup if and only if \(k<0\) or \(k>2\) (cf. Proposition 4 in [24]).

Example 31

Let \({\mathcal {M}}=({\mathbb {P}}_1{\mathbb {C}},{\mathcal {O}}_{{\mathcal {M}}})\) be a split complex supermanifold of dimension \(\dim {\mathcal {M}}=(1|2)\) associated with \({\mathcal {O}}(-k_1)\oplus {\mathcal {O}}(-k_2)\), \(k_1,k_2\in {\mathbb {Z}}\). We will determine the group \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) of automorphisms \({\mathcal {M}}\rightarrow {\mathcal {M}}\).

We choose coordinates \(z,\theta _1,\theta _2\) for \({\mathcal {U}}_0\) and \(w,\eta _1,\eta _2\) for \({\mathcal {U}}_1\) as described above such that the transition map \(\chi \) is given by \(\chi ^*(w)=z^{-1}\) and \(\chi ^*(\eta _j)={z^{-k_j}}\theta _j\).

The action of \(\mathrm {PSL}_2({\mathbb {C}})\) on \({\mathbb {P}}_1{\mathbb {C}}\) by Möbius transformations lifts to an action of \(\mathrm {SL}_2({\mathbb {C}})\) on \({\mathcal {M}}\) by letting \(A=\left( {\begin{matrix} a&{}b\\ c&{}d \end{matrix}}\right) \in \mathrm {SL}_2({\mathbb {C}})\) act by the automorphism \(\varphi _A:{\mathcal {M}}\rightarrow {\mathcal {M}}\) with pullback

$$\begin{aligned} \varphi _A^*(z)=\frac{c+dz}{a+bz}\ \ \text { and }\ \ \varphi _A^*(\theta _j)=(a+bz)^{-k_j}\theta _j \end{aligned}$$

as a morphism over appropriate subsets of \(U_0\), and

$$\begin{aligned} \varphi _A^*(w)=\frac{aw+b}{cw+d}\ \ \text { and }\ \ \varphi _A^*(\eta _j)=(cw+d)^{-k_j}\eta _j \end{aligned}$$

over appropriate subsets of \(U_1\). Using the transition map \(\chi \) one might also calculate the representation of \(\varphi \) in coordinates as a morphism over subsets \(U_0\rightarrow U_1\) and \(U_1\rightarrow U_0\).

If \(k_1\) and \(k_2\) are both even, we have \(\varphi _A=\mathrm {Id}_{\mathcal {M}}\) for \(A=\left( {\begin{matrix} -1&{}0\\ 0&{}-1 \end{matrix}}\right) \) and thus we get an action of \(\mathrm {PSL}_2({\mathbb {C}})\) on \({\mathcal {M}}\).

Consider the homomorphism of Lie groups \(\varPsi :{{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}}) \rightarrow {{\mathrm{Aut}}}({\mathbb {P}}_1{\mathbb {C}})\) assigning to each automorphism \(\varphi :{\mathcal {M}}\rightarrow {\mathcal {M}}\) the underlying biholomorphic map \({{\tilde{\varphi }}}:{\mathbb {P}}_1{\mathbb {C}}\rightarrow {\mathbb {P}}_1{\mathbb {C}}\). This homomorphism \(\varPsi \) is surjective since \({{\mathrm{Aut}}}({\mathbb {P}}_1{\mathbb {C}}) \cong \mathrm {PSL}_2({\mathbb {C}})\) and since the \(\mathrm {PSL}_2({\mathbb {C}})\)-action on \({\mathbb {P}}_1{\mathbb {C}}\) lifts to an action (of \(\mathrm {SL}_2({\mathbb {C}})\)) on the supermanifold \({\mathcal {M}}\). The kernel \(\ker \varPsi \) of the homomorphism \(\varPsi \) consists of those automorphisms \(\varphi :{\mathcal {M}}\rightarrow {\mathcal {M}}\) whose underlying map \({{\tilde{\varphi }}}\) is the identity \({\mathbb {P}}_1{\mathbb {C}}\rightarrow {\mathbb {P}}_1{\mathbb {C}}\). This kernel \(\ker \varPsi \) is a normal subgroup, \(\mathrm {SL}_2({\mathbb {C}})\) acts on \(\ker \varPsi \), and we have

$$\begin{aligned} {{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\cong \ker \varPsi \rtimes \mathrm {SL}_2({\mathbb {C}}) \end{aligned}$$

if \(k_1\) and \(k_2\) are not both even, and \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\cong \ker \varPsi \rtimes \mathrm {PSL}_2({\mathbb {C}})\) if \(k_1\) and \(k_2\) are even. Thus, it remains to determine \(\ker \varPsi \).

Let \(\varphi :{\mathcal {M}}\rightarrow {\mathcal {M}}\) be an automorphism with \({{\tilde{\varphi }}}=\mathrm {Id}\). Let f and \(b_{jk}\), \(j,k=1,2\), be holomorphic functions on \(U_0\cong {\mathbb {C}}\) such that the pullback of \(\varphi \) over \(U_0\) is given by

$$\begin{aligned} \varphi ^*(z)=z+f(z)\theta _1\theta _2\ \ \text { and }\ \ \varphi ^*(\theta )=B(z)\theta , \end{aligned}$$

where \(B(z)=\left( {\begin{matrix} b_{11}(z)&{}b_{12}(z)\\ b_{21}(z)&{}b_{22}(z) \end{matrix}}\right) \) and \(\varphi ^*(\theta )=B(z)\theta \) is an abbreviation for

$$\begin{aligned} \varphi ^*(\theta _j)=b_{j1}(z)\theta _1+b_{j2}(z)\theta _2\ \text {for}\ j=1,2. \end{aligned}$$

Similarly, let g and \(c_{jk}\) be holomorphic functions on \(U_1\cong {\mathbb {C}}\) such that the pullback of \(\varphi \) over \(U_1\) is given by

$$\begin{aligned} \varphi ^*(w)=w+g(w)\eta _1\eta _2\ \ \text { and }\ \ \varphi ^*(\eta )=C(z)\eta , \end{aligned}$$

where \(C(z)=\left( {\begin{matrix} c_{11}(z)&{}c_{12}(z)\\ c_{21}(z)&{}c_{22}(z) \end{matrix}}\right) \). The compatibility condition with the transition map \(\chi \) gives now the relation

$$\begin{aligned} f(z)=-z^{2-(k_1+k_2)}g\left( \frac{1}{z}\right) \ \text { for } z\in {\mathbb {C}}^*. \end{aligned}$$

Therefore, f and g are both polynomials of degree at most \(2-(k_1+k_2)\), and they are 0 in the case \(k_1+k_2>2\). For the matrices B and C we get the relation

$$\begin{aligned} B(z)=\begin{pmatrix} z^{k_1}&{}0\\ 0&{}z^{k_2} \end{pmatrix}C\left( \frac{1}{z}\right) \begin{pmatrix} z^{-k_1}&{}0\\ 0&{}z^{-k_2} \end{pmatrix}\ \text { for } z\in {\mathbb {C}}^*. \end{aligned}$$

If \(k_1=k_2\), this implies \(B(z)=C\left( \frac{1}{z}\right) \) for all \(z\in {\mathbb {C}}^*\). Thus, \(B(z)=B\) and \(C(w)=C\) are constant matrices, and \(B=C\in \mathrm {GL}_2({\mathbb {C}})\) since \(\varphi \) was assumed to be invertible. Consequently, we have

$$\begin{aligned} \ker \varPsi \cong P_{2-(k_1+k_2)}\rtimes \mathrm {GL}_2({\mathbb {C}}) \end{aligned}$$

in the case \(k_1=k_2\), where \(P_{2-(k_1+k_2)}\) denotes the space of polynomials of degree at most \(2-(k_1+k_2)\) if \(k_1+k_2< 2\) and \(P_{2-(k_1+k_2)}=\{0\}\) otherwise. The group structure on the semidirect product is given by \((f_1(z),B_1)\cdot (f_2(z),B_2)=(\det B_1 f_1(z)+f_2(z), B_1B_2)\).

Let now \(k_1\ne k_2\). After possibly changing coordinates we may assume \(k_1>k_2\). Then we have

$$\begin{aligned} B(z) =\begin{pmatrix} z^{k_1}&{}0\\ 0&{}z^{k_2} \end{pmatrix}C\left( \frac{1}{z}\right) \begin{pmatrix} z^{-k_1}&{}0\\ 0&{}z^{-k_2} \end{pmatrix} =\begin{pmatrix} c_{11}\left( \frac{1}{z}\right) &{} z^{k_1-k_2}c_{12}\left( \frac{1}{z}\right) \\ z^{k_2-k_1}c_{21}\left( \frac{1}{z}\right) &{} c_{22}\left( \frac{1}{z}\right) \end{pmatrix} \end{aligned}$$

for all \(z\in {\mathbb {C}}^*\). This implies that \(b_{11}=c_{11}\) and \(b_{22}=c_{22}\) are constants. Since we assume \(k_1>k_2\), we also get \(b_{21}=c_{21}=0\) and \(b_{12}\) and \(c_{12}\) are polynomials of degree at most \(k_1-k_2\). Therefore,

$$\begin{aligned} \ker \varPsi \cong P_{2-(k_1+k_2)}\rtimes \left\{ \left. \begin{pmatrix} \lambda &{} p(z)\\ 0&{}\mu \end{pmatrix} \,\right| \, \lambda ,\mu \in {\mathbb {C}}^*,\, p\in P_{k_1-k_2}\right\} , \end{aligned}$$

and the group structure is again given by

$$\begin{aligned} (f_1(z),B_1)\cdot (f_2(z),B_2)=(\det B_1 f_1(z)+f_2(z), B_1B_2) \end{aligned}$$

for \(f_1,f_2\in P_{2-(k_1+k_2)}\), \(B_1,B_2\in \left\{ \left. \left( {\begin{matrix} \lambda &{} p(z)\\ 0&{}\mu \end{matrix}}\right) \,\right| \, \lambda ,\mu \in {\mathbb {C}}^*,\, p\in P_{k_1-k_2}\right\} \).

The semidirect product \(\ker \varPsi \rtimes \mathrm {SL}_2({\mathbb {C}})\) (or \(\ker \varPsi \rtimes \mathrm {PSL}_2({\mathbb {C}})\)) is a direct product if and only if \(k_1=k_2\) and \(k_1+k_2\ge 2\).

Example 32

Let \({\mathcal {M}}=({\mathbb {P}}_1{\mathbb {C}}, {\mathcal {O}}_{\mathcal {M}})\) be the complex supermanifold of dimension \(\dim {\mathcal {M}}=(1|2)\) given by the transition map \(\chi :{{\mathcal {U}}_0}^*\rightarrow {{\mathcal {U}}_1}^*\) with pullback

$$\begin{aligned} \chi ^*(w)=\frac{1}{z} +\frac{1}{z^3}\theta _1\theta _2\ \ \text { and }\ \ \chi ^*(\eta _j)=\frac{1}{z^2} \theta _j. \end{aligned}$$

The supermanifold \({\mathcal {M}}\) is not split and the associated split supermanifold corresponds to \({\mathcal {O}}(-2)\oplus {\mathcal {O}}(-2)\); see e.g. [7].

As in the previous example, the action of \(\mathrm {PSL}_2({\mathbb {C}})\) on \({\mathbb {P}}_1{\mathbb {C}}\) by Möbius transformations lifts to an action of \(\mathrm {PSL}_2({\mathbb {C}})\) on \({\mathcal {M}}\). Let A denote the class of \(\left( {\begin{matrix} a&{}b\\ c&{}d \end{matrix}}\right) \in \mathrm {SL}_2({\mathbb {C}})\) in \(\mathrm {PSL}_2({\mathbb {C}})\). Then A acts by the morphism \(\varphi _A:{\mathcal {M}}\rightarrow {\mathcal {M}}\) whose pullback as a morphism over appropriate subsets of \(U_0\) is given by

$$\begin{aligned} \varphi _A^*(z)=\frac{c+dz}{a+bz}-\frac{b}{(a+bz)^3}\theta _1\theta _2 \ \ \text { and }\ \ \varphi _A^*(\theta _j)=\frac{1}{(a+bz)^2}\theta _j. \end{aligned}$$

Let \(\varPsi :{{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\rightarrow {{\mathrm{Aut}}}({\mathbb {P}}_1{\mathbb {C}}) \cong \mathrm {PSL}_2({\mathbb {C}})\) denote again the Lie group homomorphism which assigns to an automorphism of \({\mathcal {M}}\) the underlying automorphism of \({\mathbb {P}}_1{\mathbb {C}}\). The assignment \(A\mapsto \varphi _A\in {{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) defines a section \(\mathrm {PSL}_2({\mathbb {C}})\rightarrow {{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) of \(\varPsi \), and we have

$$\begin{aligned} {{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\cong \ker \varPsi \rtimes \mathrm {PSL}_2({\mathbb {C}}). \end{aligned}$$

The section \(\mathrm {PSL}_2({\mathbb {C}})\rightarrow {{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\) induces on the level of Lie algebras the morphism \(\sigma :{{\mathfrak {s}}}{{\mathfrak {l}}}_2({\mathbb {C}})\hookrightarrow \mathrm {Vec}_{{{\bar{0}}}}({\mathcal {M}})\), which maps an element \(\left( {\begin{matrix} a&{}b\\ c&{}-a \end{matrix}}\right) \in {{\mathfrak {s}}}{{\mathfrak {l}}}_2({\mathbb {C}})\) to the super vector field on \({\mathcal {M}}\) whose restriction to \({\mathcal {U}}_0\) is

$$\begin{aligned} \left( c-2az-bz^2-b\theta _1\theta _2\right) \frac{\partial }{\partial z} -2(a+bz)\left( \theta _1\frac{\partial }{\partial \theta _1} +\theta _2\frac{\partial }{\partial \theta _2}\right) . \end{aligned}$$

We now calculate the kernel \(\ker \varPsi \). Let \(\varphi \in \ker \varPsi \). Its underlying map \({{\tilde{\varphi }}}\) is the identity and we thus have

$$\begin{aligned} \varphi ^*(z)=z+a_0(z)\theta _1\theta _2\ \ \text { and }\ \ \varphi ^*(\theta )=A_0(z)\theta \end{aligned}$$

on \(U_0\) and

$$\begin{aligned} \varphi ^*(w)=w+a_1(w)\eta _1\eta _2\ \ \text { and }\ \ \varphi ^*(\eta )=A_1(w)\eta \end{aligned}$$

on \(U_1\) for holomorphic functions \(a_0\) and \(a_1\) and invertible matrices \(A_0\) and \(A_1\) whose entries are holomorphic functions. The notation \(\varphi ^*(\theta )=A_0(z)\theta \) (and similarly \(\varphi ^*(\eta )=A_1(w)\eta \)) is again an abbreviation for \(\varphi ^*(\theta _j)=(A_0(z))_{j1}\theta _1+(A_0(z))_{j2}\theta _2\), where \(A_0(z)=\left( (A_0(z))_{jk}\right) _{1\le j,k\le 2}\). A calculation with the transition map \(\chi \) then yields the relations

$$\begin{aligned} A_1(w)=A_0\left( \frac{1}{w}\right) \ \ \text { and }\ \ a_1(w)=\frac{1}{w} \left( \left( \det A_0\left( \frac{1}{w} \right) -1\right) -\frac{1}{w} a_0\left( \frac{1}{w} \right) \right) \end{aligned}$$

for any \(w\in {\mathbb {C}}^*\). Since \(a_0\), \(a_1\), \(A_0\), and \(A_1\) are holomorphic on \({\mathbb {C}}\), we get that \(A_0=A_1\) are constant matrices, \(\det A_0=1\), and \(a_0=a_1=0\). Therefore, \(\ker \varPsi \cong \mathrm {SL}_2({\mathbb {C}})\), and its Lie algebra is

$$\begin{aligned} \left\{ \left. \left( a_{11}\theta _1+a_{12}\theta _2\right) \frac{\partial }{\partial \theta _1} +\left( a_{21}\theta _1+a_{22}\theta _2\right) \frac{\partial }{\partial \theta _2}\,\right| \, \begin{pmatrix} a_{11}&{}a_{12}\\ a_{21}&{}a_{22} \end{pmatrix}\in {{\mathfrak {s}}}{{\mathfrak {l}}}_2({\mathbb {C}})\right\} . \end{aligned}$$

Since \(\mathrm {Lie}(\ker \varPsi )\) and \(\sigma (\mathrm {Lie}(\mathrm {PSL}_2({\mathbb {C}}))\) commute, the semidirect product \(\ker \varPsi \rtimes \mathrm {PSL}_2({\mathbb {C}})\) is direct and we have

$$\begin{aligned} {{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {M}})\cong \mathrm {SL}_2({\mathbb {C}})\times \mathrm {PSL}_2({\mathbb {C}}). \end{aligned}$$

Remark in particular that this group is different from the automorphism group of the corresponding split supermanifold \({\mathcal {N}}\), which is associated with \({\mathcal {O}}(-2)\oplus {\mathcal {O}}(-2)\), with \({{\mathrm{Aut}}}_{{{\bar{0}}}}({\mathcal {N}}) \cong \mathrm {GL}_2({\mathbb {C}})\times \mathrm {PSL}_2({\mathbb {C}})\).