1 Introduction

1.1 Semisimple quotients

The categories of finite dimensional representations \(\mathcal {T}_{m|n}\) of the general linear supergroups \(GL(m\vert n)\) over an algebraically closed field k of characteristic zero are abelian tensor categories, where representations in \(\mathcal {T}_{m|n}\) are always understood to be algebraic. However, contrary to the classical case of the general linear groups GL(n) these categories are not semisimple. Whereas the tensor product \(V \otimes V\), \(V \simeq k^{m|n}\), is completely reducible, this is no longer true for the tensor product \({{\mathbb {A}}}= V \otimes V^{\vee }\). Indeed \({{\mathbb {A}}}\) defines the indecomposable adjoint representation of \(GL(n\vert n)\) in the case \(m=n\), hence admits a trivial one dimensional subrepresentation defined by the center and a trivial one dimensional quotient representation defined by the supertrace. In contrast to the classical case the supertrace is trivial on the center, and \({{\mathbb {A}}}\) is indecomposable with three irreducible Jordan–Hoelder factors \(1,S^1,1\) with the superdimensions \(1, -2, 1\) respectively defined by the filtration \({\mathfrak {z}} \subseteq \mathfrak {sl}(n\vert n) \subseteq \mathfrak {gl}(n\vert n)\), where \({\mathfrak {z}}\) denotes the center of \(\mathfrak {gl}(n\vert n)\).

Although the irreducible representations of \(GL(m\vert n)\) can be classified by highest weights similarly to the classical case, this implies that the tensor product of irreducible representations is in general far from being completely reducible. In fact Weyl’s unitary trick fails in the superlinear setting. While the structure of \(\mathcal {T}_{m|n}\) as an abelian category is now well understood [14], its monoidal structure remains mysterious.

The perspective of this article is that, in order to restore parts of the classical picture, two finite dimensional representations M and \(M'\) of GL(m|n) should not be distinguished, if there exists an isomorphism

$$\begin{aligned} M \oplus N \cong M' \oplus N' \end{aligned}$$

where N and \(N'\) are negligible modules. Here we use the notion that a finite dimensional module is said to be negligible if it is a direct sum of indecomposable modules whose superdimensions are zero. A typical example of a negligible module is the indecomposable adjoint representation \({{\mathbb {A}}}\) in the case \(m=n\). To define this precisely we divide our category \(\mathcal {T}_{m|n}\) by the tensor ideal \(\mathcal {N}\) [2] of negligible morphisms. The quotient is a semisimple abelian tensor category. By a fundamental result of Deligne it is equivalent to the representation category of a pro-reductive supergroup \(G^{red}\) [34].

Taking the quotient of a non-semisimple tensor category by objects of categorial dimension 0 has been studied in a number of different cases. A well-known example is the quotient of the category of tilting modules by the negligible modules (of quantum dimension 0) in the representation category of the Lusztig quantum group \(U_q(\mathfrak {g})\) where \(\mathfrak {g}\) is a semisimple Lie algebra over k [1, 6]. The modular categories so obtained have been studied extensively in their applications to the 3-manifold invariants of Reshetikhin–Turaev. In [40] Jannsen proved that the category of numerical motives as defined via algebraic correspondences modulo numerical equivalence is an abelian semisimple category. It was noted by André and Kahn [2] that taking numerical equivalence amounts to taking the quotient by the negligible morphisms. Jannsen’s theorem has been generalized to a categorical setting by [2]. In particular they study quotients of tannakian categories by the ideal of negligible morphisms. Recently Etingof and Ostrik [28] studied semisimplifactions with an emphasis on finite tensor categories.

A general study of \(Rep(G)/\mathcal {N}\), where G is a supergroup scheme, was initiated in [34] where in particular the reductive group \(G^{red}\) given by \(Rep(G^{red}) \simeq Rep(GL(m|1))/\mathcal {N}\) was determined. This example is rather special since Rep(GL(m|1)) has tame representation type. One can always assume \(m\ge n\). Note for \(m\ge n \ge 2\) the problem of classifying irreducible representations of \(G^{red}\) is wild [34]. Therefore one should not study the entire quotient \(\mathcal {T}_{m|n}/\mathcal {N}\), but rather pass to a suitably small tensor subcategory in \(\mathcal {T}_{m|n}\). In our case we consider for this the Karoubi envelope of the irreducible objects Rep(G), which for convenience are only considered up to suitable parity shift. The image \(\overline{\mathcal {T}}_{m|n}\) of this subcategory in \(Rep(G)/\mathcal {N}\) defines a tannakian category, and the aim of this paper is to determine its Tannaka group \(H_n\) in the cases \(G=Gl(n\vert n)\). Indeed, as we show in [39], the computation of the corresponding Tannaka group \(H_{m|n}\) in the case \(G=Gl(m\vert n)\) can be reduced to the case \(m=n\) besides an additional factor \(GL(m-n)\) that appears in the Tannaka group \(H_{m|n}\) which arises from the Tannakian group of the Tannakian subcategory generated by the covariant tensor representations. In this sense the major complications of the constructions arise in the case \(m=n\). So, for simplicity, we restrict ourselves to the case \(m=n\) and postpone the additional combinatorial arguments that are necessary for the general case \(m>n\) to the paper [39].

1.2 The Tannaka category \(\overline{\mathcal {T}}_n\)

For convenience we now write \({\mathcal {T}}_n\) instead of \({\mathcal {T}}_{n,n}\). To define the karoubienne hull of the irreducible representations in \({\mathcal {T}}_n\), we work with the tensor subcategory \({{\mathcal {T}}}_n^+\) generated by the irreducible representations of nonnegative superdimension in the following sense. Up to parity shift the irreducible representations \(L=L(\lambda )\) of \({{\mathcal {T}}}_n\) are parametrized up to isomorphy by their highest weights \(\lambda \). We define an equivalence relation on the set of highest weights, such that \(\lambda \) and \(\lambda '\) are called equivalent if \(L(\lambda )\) or its Tannaka dual \(L(\lambda )^\vee \) is isomorphic to \(Ber^r \otimes L(\lambda ')\) for some power \(Ber^r, r\in {\mathbb {Z}}\) of the Berezin determinant representation Ber in \({{\mathcal {T}}_n}\). Let \(Y^+_0(n)\) denote the set of equivalence classes \(\lambda /\!\sim \) of maximal atypical weights. The cases where \(L(\lambda )^\vee \cong Ber^r \otimes L(\lambda )\) holds are called (SD)-cases if \(\dim (L(\lambda )) >1\). The remaining cases are called (NSD)-cases. Notice that an irreducible representation \(L(\lambda )\) of \(GL(n\vert n)\) can be replaced by a parity shift \(X_\lambda \) of \(L(\lambda )\) so that the superdimension \({{\,\textrm{sdim}\,}}(X_\lambda )\) becomes \(\ge 0\). For maximal atypical representations \(L(\lambda )=[\lambda _1,\ldots ,\lambda _n]\) this is well defined since the superdimension is not zero. But of course this is ambiguous for irreducible representations of \(GL(n\vert n)\) with \({{\,\textrm{sdim}\,}}(L)=0\), i.e. for the irreducible representations that are not maximal atypical. But these representations are negligible in the sense above. Thus we may consider only objects that are retracts of iterated tensor products of direct sums of maximal atypical irreducible representations \(X_\lambda \) of \(GL(n\vert n)\) satisfying \({{\,\textrm{sdim}\,}}(X_\lambda ) > 0\). The tensor category thus obtained will be baptized \(\mathcal {T}_n^+\). The full tensor subcategory \(\mathcal {T}_n^+\) of \({\mathcal {T}}_n\) has more amenable properties than the full category \({\mathcal {T}}_n =Rep(GL(n\vert n))\). To motivate this, let us compare it with the tensor category of finite dimensional algebraic representations Rep(G) of an arbitrary algebraic group G over k. In this situation the tensor subcategory generated by irreducible representations is semisimpleFootnote 1 and can be identified with the tensor category of the maximal reductive quotient of G. The tensor category \(\mathcal {T}_n^+\) however is not a semisimple tensor category in general. To make it semisimple we proceed as follows:

Let \(\overline{\mathcal {T}}_{n}\) denote the quotient category of \(\mathcal {T}_n^+\) obtained by killing the negligible morphisms in the tensor ideal \({{\mathcal {N}}}\) and hence in particular all neglegible objects, i.e.

$$\begin{aligned} \overline{\mathcal {T}}_{n} \cong \mathcal {T}_n^+/\mathcal {N}. \end{aligned}$$

In order to analyze these categories, we work inductively using the cohomological tensor functors \(DS: \mathcal {T}_n \rightarrow \mathcal {T}_{n-1}\) of [36]. We show in Lemma 5.5 that DS induces a tensor functor \(\eta _n: \mathcal {T}_n^+ \rightarrow \mathcal {T}_{n-1}^+\).

Theorem 1.1

The categories \(\overline{\mathcal {T}}_{n}\) are semisimple tannakian categories. A fibre functor \(\omega \) is provided by the composite of functors \(\eta _m\) for \(1\le m\le n\). Their Tannaka groups \(H_n\) are projective limits of reductive algebraic groups over k such that there is an equivalence of tensor categories

$$\begin{aligned} \overline{\mathcal {T}}_{n} \ \cong \ Rep(H_n). \end{aligned}$$

The functor \(\eta _n\) induces a closed embedding of affine group schemes \(H_{n-1} \hookrightarrow H_n\) over k such that \(\eta _n: Rep(H_n) \rightarrow Rep(H_{n-1})\), defined by DS on objects, can be identified with the restriction functor for this group scheme embedding.

The Tannaka group \(H_n\) is subgroup of the product of Tannaka groups \(H_\lambda \), where \(\lambda \) runs over the set of equivalence classes \(\lambda /\sim \) of maximal atypical highest weights \(\lambda \). By definition \(H_\lambda \) is the Tannaka group of the tannakian subcategory \({{\mathcal {T}}}_\lambda \) that is generated by the simple object \(X_\lambda \) in \(\overline{{\mathcal {T}}}_n\). For the twisted Berezin \(X_\lambda =B\) the group \(H_\lambda \) is isomorphic to the multiplicative group \({{\mathbb {G}}}_m\) whose characters correspond to the irreducible one dimensional representations in \({{\mathcal {T}}}_n^+\), the powers \(B^r\) of B. In general \(H_\lambda \) may be considered as an algebraic subgroup of the general linear group \(GL(V_\lambda )\) of the finite dimensional k-vectorspace \(V_\lambda =\omega (X_\lambda )\) defined by the fibre functor \(\omega \). Note that \(\dim (V_\lambda ) = {{\,\textrm{sdim}\,}}(X_\lambda )\) and this value is bounded by n!. The restriction of the determinant character of \(GL(V_\lambda )\) will be denoted \(\det _\lambda \). In Theorem 14.3 we show that \(\det _\lambda \) is a power \(B^{\ell (\lambda )}\) of the character defined by B.

Theorem 1.2

The character \(\det _\lambda \) of the group \(H_\lambda \) is represented by the image of \(\det (X_{\lambda }) = \Lambda ^{{{\,\textrm{sdim}\,}}(X_{\lambda })}(X_{\lambda })\) in the Tannaka category \(\overline{{\mathcal {T}}}_n\). In \({{\mathcal {T}}}_n^+\) one has

$$\begin{aligned} \det (X_{\lambda }) \, \cong \, B^{\ell (\lambda )} \, \oplus \, N \end{aligned}$$

for some negligible object N with the integer

$$\begin{aligned} \ell (\lambda ) \ =\ n^{-1} {{\,\textrm{sdim}\,}}(X_\lambda )\, D(\lambda ) \, \ \end{aligned}$$

where \(D(\lambda )\) is explicitely described by the weight \(\lambda \) in Sect. 13.

In the (SD)-cases the isomorphism \(L(\lambda )^\vee \cong Ber^r \otimes L(\lambda )\) defines a nondegenerate pairing \(X_\lambda \otimes X_\lambda \rightarrow B^r\). It induces a nondegenerate k-bilinear pairing on \(V_\lambda \) such that \(H_\lambda \) becomes a subgroup of its group of similitudes. In “Appendix C” we determine the parity of this pairing.

Theorem 1.3

In the (SD)-cases the parity \(\varepsilon (X_\lambda )\) of the invariant pairing \(\langle .,.\rangle \) on \(V_\lambda \) defining \(H_\lambda \) is

$$\begin{aligned} \varepsilon (X_\lambda ) = \varepsilon (\lambda _{basic}) \ \end{aligned}$$

where \(\varepsilon (\lambda _{basic})= (-1)^{\sum _{i=1}^n \mu _i}\) if \(L(\lambda _{basic}) \cong [\mu _1,\ldots ,\mu _n]\) is the irreducible basic representation associated to \(L(\lambda )\).

For each maximal atypical weight \(\lambda \) we define characters

$$\begin{aligned} \mu _\lambda : H_\lambda \rightarrow {{\mathbb {G}}}_m \end{aligned}$$

of the Tannaka groups \(H_\lambda \) as follows: First suppose that \(X_\lambda \) is not isomorphic to a power of B. Then, in the (NSD)-cases \(\mu _\lambda \) is defined to be \(\det _\lambda \). In the (SD)-cases \(\mu _\lambda \) is defined as the restriction of the similitude character to \(H_\lambda \). The similitude character of \(H_\lambda \) is defined by an object in \({{\mathcal {T}}}_\lambda \) that is isomorphic to the image of \(B^r\). To make these characters well defined notice the following: For the twisted Berezin the associated Tannaka group \(H_\lambda \) is isomorphic to the multiplicative group \({{\mathbb {G}}}_m\), whose characters correspond to the irreducible one-dimensional representations in \({{\mathcal {T}}}_n^+\), the powers \(B^r\) of B. Any tensor functor between tannakian categories (compatible with the fibre functors) induces a group homorphism between the Tannaka groups that is uniquely determined up to conjugacy by the functor. This observation, applied to the tannakian subcategories of \({{\mathcal {T}}}_\lambda \) that are generated by \(\det (X_\lambda )\) resp. \(B^r\), together with the fact that inner automorphisms of \({\mathbb {G}}_m\) are trivial, shows that all characters \(\mu _\lambda \) are uniquely defined once an isomorphism \(\mu _\lambda \) for \(X_\lambda \cong B\) between the Tannaka group \(H_B:=H_\lambda \) and the multiplicative group \({{\mathbb {G}}}_m\) has been chosen. We fix such an isomorphism, denoted \(\mu _B\) in the following.

This being said, a conjectural description of the structure of \(H_n\) can be given as follows:

Conjecture 1.4

(1) \(H_n\) is the subgroup of the infinite product

$$\begin{aligned} \prod _{\lambda / \sim \ \in \, Y^+_0(n)} H_\lambda \,\end{aligned}$$

defined by the elements \( h=(h_\lambda )_{\lambda /\sim }\) that satisfy \( \mu _\lambda (h_\lambda ) = \mu _B(h_B) \) for all \(\lambda \).

(2) For the (NSD)-cases the group \(H_\lambda \), considered as a subgroup of \(GL(V_\lambda )\), is equal to \(GL(V_\lambda )\) if \(\mu _\lambda \ne 1\) holds and is equal to the subgroup \(SL(V_\lambda )\) otherwise. Recall, \(\mu _\lambda =\det _\lambda \) holds in that case.

(3) Otherwise \(H_\lambda \), considered as a subgroup of the group of similitudes of the pairing \(\langle .,.\rangle \) on \(V_\lambda \), is equal to the connected component of the similitude group if \(\mu _\lambda \ne 1\) holds and is otherwise equal to the kernel of the similtude homomorphism \(\mu _\lambda \) on this connected component.

The parity shift of the twisted Berezin \(B=\Pi ^{n}(Ber)\) is an invertible object of the Tannaka category \(\overline{{\mathcal {T}}}_n\), i.e. an object I of superdimension 1 such that \(I \otimes I^\vee \cong \textbf{1}\) holds. Conjecture 1.4 implies that the group \(Pic(\overline{{\mathcal {T}}}_n)\) of isomorphism classes of invertible objects in \(\overline{{\mathcal {T}}}_n\) is generated by the twisted Berezin B. For this notice that in 1.4 (1) the kernel of the projection from \(H_n\subset \prod _{\lambda /\sim } H_\lambda \) in the product to the \(L(\lambda )=B\)-component is a connected semisimple profine groupscheme by 1.4 (2) and (3), hence only admits trivial characters. We also prove the converse.

Theorem 1.5

A necessary and sufficient condition that Conjecture 1.4 holds is that \(Pic(\overline{{\mathcal {T}}}_n)\) is generated by the twisted Berezin B.

So the main obstacle for the proof Conjecture 1.4 turns out to be the structure of \(Pic({{\mathcal {T}}}_\lambda )\). Conjecture 1.4 is also equivalent to the assertion that exceptional (SD)-cases in the sense of the next theorem do not occur.

Theorem 1.6

For all maximal atypical highest weights \(\lambda \) there is a homomorphism \(\nu : Pic({{\mathcal {T}}}_\lambda )\rightarrow {\mathbb {Z}}\) whose kernel is a two-torsion group \(\mu _2^k\) of rank \(k=k(\lambda )\) where \(0\le k\le 2\). In the (so called) regular case where \(k(\lambda )=0\), the group Tannaka group \(H_\lambda \) of \({{\mathcal {T}}}_\lambda \) is the one described in Conjecture 1.4.

Corollary 1.7

The Picard group \(Pic(\overline{{\mathcal {T}}}_n)\) is a direct product of \({\mathbb {Z}}\) and a 2-power torsion group.

1.3 The exceptional cases

The nonregular cases in the sense of the last Theorem 1.6 will be called exceptional cases. For these exceptional \(\lambda \) we show that there exists a subgroup \({\tilde{H}}_\lambda \) of \(H_\lambda \) of index two

$$\begin{aligned} 0 \rightarrow {\tilde{H}}_\lambda \rightarrow H_\lambda \rightarrow \mu _2 \rightarrow 0. \end{aligned}$$

The restriction of the irreducible representation \(V_\lambda \) of \(H_\lambda \) to the subgroup \({\tilde{H}}_\lambda \) decomposes into a direct sum \(W_\lambda \oplus W_\lambda ^\vee \) of two irreducible faithful nonisomorphic representations of \({\tilde{H}}_\lambda \) on orthogonal Lagrangian subspaces \(W_\lambda \) and \(W_\lambda ^\vee \) of the metric space \((V_\lambda ,\langle .,.\rangle )\); see Theorem 11.4 and Sect. 11.7. In this way we can view \({\tilde{H}}_\lambda \) as a subgroup of \(GL(W_\lambda )\), and we show that the following holds

$$\begin{aligned} SL(W_\lambda ) \subseteq {\tilde{H}}_\lambda \subseteq GL(W_\lambda ). \end{aligned}$$

Finally let \(G_\lambda \) denote the derived group of the connected component \(H_\lambda ^0\) of \(H_\lambda \), and let similar \(G_n\) denote the derived group of the connected component \(H_n^0\). For the derived connected subgroup \(G_n\) of \(H_n\) we prove the following result.

Theorem 1.8

(1) \(G_n\) is isomorphic to the infinite product

$$\begin{aligned} \prod _{\lambda /\sim \ \in \, Y^+_0(n)} G_\lambda . \end{aligned}$$

The groups \(G_\lambda \) are isomorphic to \(SL(V_\lambda )\), \(SO(V_\lambda )\), \(Sp(V_\lambda )\) in the (NSD) resp. the even or odd regular (SD)-cases and they are isomorphic to \(SL(W_\lambda )\), for a Lagrangian subspace \(W_\lambda \) of \(V_\lambda \), in the exceptional (SD)-cases.

(2) If \(\lambda \) is not an exceptional (SD)-case, the groups \(H_\lambda \) are as described in Conjecture 1.4. Furthermore the analog of Conjecture 1.4 holds for the Tannaka group generated by the simple objects \(X_\lambda \) for which which \(\lambda \) does not belong to an exceptional (SD)-case.

In view of Theorem 1.8 our Conjecture 1.4 is hence equivalent to the conjectural nonexistence of exceptional (SD)-cases.

Reformulating these statements for the category of representations of \(GL(n\vert n)\), what we have achieved is

  • a (partial) description of the decomposition law of tensor products of irreducible representations into indecomposable modules up to negligible indecomposable summands; and

  • a classification (in terms of the highest weights of \(H_{\lambda }\) and \(H_{\mu }\)) of the indecomposable modules of non-vanishing superdimension in iterated tensor products of \(L(\lambda )\) and \(L(\mu )\).

To determine this decomposition it suffices to know the Clebsch–Gordan coefficients for the classical simple groups of type ACD. Notice that \(\dim (V_\lambda )\) is always even in the (SD)-cases, hence simple groups of type B do not occur. Furthermore the superdimensions of the indecomposable summands are just the dimensions of the corresponding irreducible summands of the tensor products in \(Rep_k(H_n)\). Without this, to work out any such decomposition is rather elaborate. For the case \(n=2\) see [37]. In fact the knowledge of the Jordan–Hölder factors usually gives too little information on the indecomposable objects itself. In the (NSD) and the odd (SD)-case it is enough for these two applications to know the connected derived group \(G_{\lambda }\) since the restriction of any irreducible representation of \(H_{\lambda }\) to \(G_{\lambda }\) stays irreducible. Therefore these results hold unconditionally in these cases. In the even (SD)-case we need the finer (but conjectural) results of Sect. 12 to see that \(H_{\lambda }\) is connected. We refer the reader to Example 9.7 and Sect. 15 for some examples.

1.4 Relation to physics

Part of the motivation for our computations of the Tannaka groups \(H_n\) comes from the real supergroups \(G=SU(2,2\vert N)\) which are covering groups of the super conformal groups \(SO(2,4\vert N)\). The complexification \({\mathfrak {g}}\) of Lie(G) is isomorphic to the complex Lie superalgebras \(\mathfrak {sl}(4\vert N)\). The finite dimensional representations of \(\mathfrak {g}\) are hence related to those of the Lie superalgebras \(\mathfrak {gl}(n\vert n)\) for \(n\le 4\). Since the complexification \({\mathfrak {g}}\) defines complex supervector fields on four dimensional Minkowski superspace M and compactifications of it, these Lie superalgebras play a role in string theory and the AdS/CFT correspondence.

For supersymmetric fields \(\psi \) on M with values in a finite dimensional representation L of \({\mathfrak {g}}\), the Feynman integrals of conformal theories are computed from tensor contractions and superintegration. These can be considered as contractions between tensor products of fields. The computations will require the analysis of higher tensor products \(L^{\otimes r}\otimes (L^\vee )^{\otimes s}\) and generalizations of Fierz rules. The results will strongly depend on the rules of the underlying tensor categories \({{\mathcal {T}}}_n\). Since it seems reasonable to consider not only fields of superdimension zero, but besides the constant representation also such with values in maximal atypical representations L of \(\mathfrak {g}\), our study of tensor categories may be a little step into this direction. To look at examples for \(n\le 4\) we now replace the groups \(H_\lambda \) by their compact inner forms \(H_\lambda ^c\) and \({{\mathbb {G}}}_m\) by U(1), to make things look more familiar to physicists. Indeed notice that the tensor categories \(Rep_k(H_\lambda )\) and \(Rep_{{\mathbb {C}}}(H_\lambda ^c)\) are equivalent.

For \(\mathfrak {gl}(n\vert n)\) and \(n\le 4\) there only exist finitely many isomorphism classes of quotient groups \(H_\lambda \) of the tannakian groups \(H_n\) of \(\overline{{\mathcal {T}}}=Rep(H_n)\). Besides U(1) that corresponds to the twisted Berezin, the smallest such groups are SU(2) and SU(3) related to representations denoted \(L=S^1\) and \(L=S^2\). If we pass to \(\mathfrak {sl}(n\vert n)\), for \(n\le 3\) these are the only groups except for \(Sp^c(6)\) in the case \(n=3\). For the more involved discussion of the case \(n=4\) we refer to Sect. 15 and Example 15.1. One may ask whether the appearence of the groups U(1), SU(2), SU(3) here is a mere accident, or whether there does exist some connection with the symmetry groups arising in the standard model of elementary particle physics? Looking for relations between internal symmetries and supersymmetry has a long history going back to the Coleman-Mandula theorem [17], so this may be of interest. A possible relation could be the following:

If in such a theory, for some mysterious physical reasons, the tensor product contributions to the Feynman integrals from direct summands of \(L^{\otimes r}\otimes (L^\vee )^{\otimes s}\) of superdimension zero would be relatively small in a certain energy range due to supersymmetry cancellations, then to first order they would be negligible. Hence a physical observer might come up with the impression that the underlying rules of symmetry are imposed by the invariant theory of the quotient tensor category \(\overline{{\mathcal {T}}}_n\! =\! Rep(H_n)\) instead of \({\mathcal {T}}_n\), i.e. the tensor categories that are obtained by ignoring negligible indecomposable summands of superdimension zero. Thus \(H_n\) would appear as an internal symmetry group of the theory in an approximate sense. Of course, any such speculation is highly tentative for various reasons: Computations along such lines will probably be extremely involved. Fields with values in maximal atypical representations V may produce ghosts in the associated infinite dimensional representations of \({\mathfrak {g}}\). In other words, such field theories may a priori not be superunitary and it is unclear whether the passage to the cohomology groups for operators like DS or the Dirac operator \(H_D\) [36], breaking the conformal symmetry, would suffice to get rid of ghosts.

1.5 Structure of the article

Our main tool are the cohomological tensor functors \(DS: \mathcal {T}_n \rightarrow \mathcal {T}_{n-1}\) of [36]. In the main theorem of [36, Theorem 16.1] we calculate \(DS(L(\lambda ))\). In particular \(DS(L(\lambda ))\) is semisimple and multiplicity free. We show in Lemma 5.5 that DS induces a tensor functor \(DS: \mathcal {T}_n^+ \rightarrow \mathcal {T}_{n-1}^+\) and by Lemma 5.11 one can construct a tensor functor on the quotient categories

$$\begin{aligned} \eta : \mathcal {T}_n^+/\mathcal {N} \rightarrow \mathcal {T}_{n-1}^+/\mathcal {N}. \end{aligned}$$

This seemingly minor observation is one of the crucial points of the proof since it allows us to determine the groups \(H_n\) and \(G_n\) inductively. We also stress that it is not clear whether DS naturally induces a functor between the quotients \(\mathcal {T}_n^+/\mathcal {N}\) and \(\mathcal {T}_{n-1}^+/\mathcal {N}\) on the level of morphisms. In fact, if one enlarges \(\mathcal {T}^+\) to the larger category \(\mathcal {T}^{ev}\) of Sect. 14, DS does not preserve negligible morphisms. The DS functor however agrees with the functor \(\eta \) on objects. The quotient \( \mathcal {T}_n^+/\mathcal {N}\) is equivalent to the representation category \(Rep(H_n)\) of finite-dimensional representations of a pro-reductive group. By the deep and powerful Theorem 5.15 of Deligne the induced DS functor determines an embedding of algebraic groups \(H_{n-1} \hookrightarrow H_n\) and the functor DS is the restriction functor with respect to this embedding.

Hence the main theorem of [36] tells us the branching laws for the representation \(V_{\lambda }\) with respect to the embedding \(H_{n-1} \hookrightarrow H_n\). Our strategy is to determine the groups \(H_n\) or \(G_n\) inductively using the functor DS. For \(n = 2\) we need the explicit results of [37] to give us the fusion rule between two irreducible representations and we describe the corresponding Tannaka group in Lemma 9.2. Starting from the special case \(n=2\) we can proceed by induction on n. For this we use the embedding \(H_{n-1} \hookrightarrow H_n\) along with the known branching laws and the classification of small representations due to Andreev et al. [5] which allows to determine inductively the connected derived groups \(G_n = (H_n^0)_{der}\) for \(n\ge 3\); see Sect. 10. The passage to the connected derived group means that we have to deal with the possible decomposition of \(V_{\lambda }\) when restricted to \(G_n\). In order to determine \(G_n\) we first determine the connected derived groups \(G_{\lambda }\) corresponding to the tensor subcategory generated by the image of \(L(\lambda )\) in \(\overline{\mathcal {T}}_n\) in Theorem 6.2. Roughly speaking the strategy of the proof is the following: We use the inductively known situation for \(G_{n-1}\) to show that for sufficiently large n the rank and the dimension of \(G_{\lambda }\) is large compared to the dimension of \(V_{\lambda }\), i.e. \(V_{\lambda }\) or any of its irreducible constituents in the restriction to \(G_{\lambda }\) should be small in the sense of [5]. We refer to Sect. 10 for more details on the proof.

The final sections are devoted to the determination of \(H_{\lambda }\) and \(Rep(H_n)\). We determine the groups \(H_{\lambda }\) in Sect. 11. We split this determination into three cases: NSD, regular SD and exceptional SD. The crucial tool here is the determination of the determinant \(det(X_{\lambda })\). This determinant is computed in the later Sects. 13 and 14.

In Sect. 12 we conjecture a stronger structure theorem, namely that there are no exceptional SD-cases. We describe various conditions which are equivalent to this statement. We end the article with some low-rank cases and a discussion of cases of potential physical relevance.

We have outsourced a large number of technical (but necessary) results to the “Appendices A, C and D” as to not distract the reader too much from the structure of the arguments. “Appendix E” discusses some evidence for our conjectures.

1.6 Outlook

For the general \(\mathcal {T}^+_{m|n}\)-case (where \(m \ge n\)) recall that every maximal atypical block in \(\mathcal {T}_{m|n}\) is equivalent to the principal block of \(\mathcal {T}_{n|n}\). We fix the standard block equivalence due to Serganova and denote the image of an irreducible representation \(L(\lambda )\) under this equivalence by \(L(\lambda ^0)\).

Theorem 1.9

Suppose that \({{\,\textrm{sdim}\,}}(L(\lambda )) > 0\). Then \(H_{\lambda } \cong GL(m-n) \times H_{\lambda ^0}\) and \(L(\lambda )\) corresponds to the representation \(L_{\Gamma } \boxtimes V_{\lambda ^0}\) of \(H_{\lambda }\). Here \(L_{\Gamma }\) is an irreducible representation of \(GL(m-n)\) which only depends on the block \(\Gamma \) (the core of \(\Gamma \) as defined by Serganova).

We prove this in [39]. The problem of determining the semisimplification of Rep(G) can be asked for any basic supergroup. We expect that the strategy employed in this paper (induction on the rank via the DS functor) serves as a blueprint for other cases, but one can certainly not expect uniform proofs or results. In fact the semisimplicity of DS is now known in the OSp-case [32], but fails for the P(n)-case [26]. Note also that even in the OSp-case no exact analog of the structure theorem can hold since the supergroup OSp(1|2n) will appear as a Tannaka group.

1.7 About the Arxiv version

The present version is identical to the version posted on the Arxiv except that the latter contains further appendices with sample calculations. In particular the theorem numbering is identical.

2 The superlinear groups

Let k be an algebraically closed field of characteristic zero. We adopt the notations of [36]. With GL(m|n) we denote the general linear supergroup and by \(\mathfrak {g} = \mathfrak {gl}(m|n)\) its Lie superalgebra. A representation \(\rho \) of GL(m|n) is a representation of \(\mathfrak {g}\) such that its restriction to \({{\mathfrak {g}}}_{\bar{0}}\) comes from an algebraic representation of \(G_{\bar{0}} = GL(m) \times GL(n)\). We denote by \(\mathcal {T} = \mathcal {T}_{m|n}\) the category of all finite dimensional representations with parity preserving morphisms.

2.1 The category \({{{\mathcal {R}}}}\)

Fix the morphism \(\varepsilon : {\mathbb {Z}}/2{\mathbb {Z}} \rightarrow G_{\overline{0}}=GL(n)\times GL(n)\) which maps \(-1\) to the element \(diag(E_n,-E_n)\in GL(n)\times GL(n)\) denoted \(\epsilon _n\). Notice that \(Ad(\epsilon _n)\) induces the parity morphism on the Lie superalgebra \(\mathfrak {gl}(n|n)\) of G. We define the abelian subcategory \({{\mathcal {R}}}= sRep(G,\varepsilon )\) of \(\mathcal {T}\) as the full subcategory of all objects \((V,\rho )\) in \(\mathcal {T}\) with the property \( p_V = \rho (\epsilon _n)\); here \(p_V\) denotes the parity morphism of V and \(\rho \) denotes the underlying homomorphism \(\rho : GL(n)\times GL(n) \rightarrow GL(V)\) of algebraic groups over k. The subcategory \({{{\mathcal {R}}}}\) is stable under the dualities \({}^\vee \) and \(^*\). For \(G=GL(n\vert n)\) we usually write \(\mathcal {T}_n\) instead of \(\mathcal {T}\), and \({{{\mathcal {R}}}}_n\) instead of \({{\mathcal {R}}}\). The irreducible representations in \({{\mathcal {R}}}_n\) are parametrized by their highest weight with respect to the Borel subalgebra of upper triangular matrices. A weight \(\lambda =(\lambda _1,\ldots ,\lambda _n \ | \ \lambda _{n+1}, \cdots , \lambda _{2n})\) of an irreducible representation in \({{\mathcal {R}}}_n\) satisfies \(\lambda _1 \ge \ldots \lambda _n\), \(\lambda _{n+1} \ge \ldots \lambda _{2n}\) with integer entries. The Berezin determinant of the supergroup \(G=G_n\) defines a one dimensional representation Ber. Its weight is is given by \(\lambda _i=1\) and \(\lambda _{n+i}=-1\) for \(i=1,\ldots ,n\). For each representation \(M \in \mathcal {R}_n\) we also have its parity shifted version \(\Pi (M)\) in \(\mathcal {T}_n\). Since we only consider parity preserving morphisms, these two are not isomorphic. In particular the irreducible representations in \(\mathcal {T}_{n}\) are given by the \(\{L(\lambda ), \Pi L(\lambda ) \ | \ \lambda \in X^+ \}\). The whole category \(\mathcal {T}_n\) decomposes as \(\mathcal {T}_{n} = {{\mathcal {R}}}_{n} \oplus \Pi {{\mathcal {R}}}_{n}\) [11, Corollary 4.44]. For maximal atypical \(\lambda \) exactly one of \(L(\lambda ), \Pi L(\lambda ))\) has positive superdimension. We call this irreducible module \(X_{\lambda }\) and \(B = \Pi ^n(Ber)\) in \(\mathcal {T}_n^+\), for \(Ber=[1,\ldots ,1]\), the twisted Berezin.

2.2 Kac objects

We put \(\mathfrak {p}_{\pm } = {\mathfrak {g}}_{(0)} \oplus {\mathfrak {g}}_{(\pm 1)}\) for the usual \({\mathbb {Z}}\)-grading \({\mathfrak {g}}= {\mathfrak {g}}_{(-1)} \oplus {\mathfrak {g}}_{(0)} \oplus {\mathfrak {g}}_{(1)}\). We consider a simple \({\mathfrak {g}}_{(0)}\)-module as a \(\mathfrak {p}_{\pm }\)-module in which \({\mathfrak {g}}_{(1)}\) respectively \({\mathfrak {g}}_{(-1)}\) acts trivially. We then define the Kac module \(V(\lambda )\) and the anti-Kac module \(V'(\lambda )\) via

$$\begin{aligned} V(\lambda ) = Ind_{\mathfrak {p}_+}^{{\mathfrak {g}}} L_0(\lambda ) \, \ V'(\lambda ) = Ind_{\mathfrak {p}_-}^{{\mathfrak {g}}} L_0(\lambda ) \end{aligned}$$

where \(L_0(\lambda )\) is the simple \({\mathfrak {g}}_{(0)}\)-module with highest weight \(\lambda \). The Kac modules are universal highest weight modules. \(V(\lambda )\) has a unique maximal submodule \(I(\lambda )\) and \(L(\lambda ) = V(\lambda )/I(\lambda )\) [41, Proposition 2.4]. We denote by \({{\mathcal {C}}}^+\) the tensor ideal of modules with a filtration by Kac modules in \({{\mathcal {R}}}_n\) and by \({{\mathcal {C}}}^-\) the tensor ideal of modules with a filtration by anti-Kac modules in \({{\mathcal {R}}}_n\).

2.3 Equivalence classes of weights

Two irreducible representations M, N in \(\mathcal {T}\) are said to be equivalent \(M \sim N\), if either \(M \cong Ber^r \otimes N\) or \(M^\vee \cong Ber^r \otimes N\) holds for some \(r\in {\mathbb {Z}}\). This obviously defines an equivalence relation on the set of isomorphism classes of irreducible representations of T. A self-equivalence of M is given by an isomorphism \(f: M \cong Ber^r \otimes M\) (which implies \(r=0\) and f to be a scalar multiple of the identity) respectively an isomorphism \(f: M^\vee \cong Ber^r \otimes M\). If it exists, such an isomorphism uniquely determines r and is unique up to a scalar and we say M is of type (SD). Otherwise we say M is of type (NSD). The isomorphism f can be viewed as a nondegenerate G-equivariant bilinear form

$$\begin{aligned} M \otimes M \rightarrow Ber^r \, \end{aligned}$$

which is either symmetric or alternating. So we distinguish between the cases (\(\text {SD}_{\pm }\)).

3 Weight and cup diagrams

3.1 Weight diagrams and cups

Consider a weight

$$\begin{aligned} \lambda =(\lambda _1,\ldots ,\lambda _n | \lambda _{n+1}, \ldots , \lambda _{2n}).\end{aligned}$$

Then \(\lambda _1 \ge \cdots \ge \lambda _n\) and \(\lambda _{n+1} \ge \cdots \ge \lambda _{2n}\) are integers, and every \(\lambda \in {{\mathbb {Z}}}^{2n}\) satisfying these inequalities occurs as the highest weight of an irreducible representation \(L(\lambda )\). The set of highest weights will be denoted by \(X^+=X^+(n)\). Following [14] to each highest weight \(\lambda \in X^+(n)\) we associate two subsets of cardinality n of the numberline \({\mathbb {Z}}\)

$$\begin{aligned} I_\times (\lambda )\ {}&=\ \{ \lambda _1 , \lambda _2 - 1, \ldots , \lambda _n - n +1 \} \\ I_\circ (\lambda )\ {}&= \ \{ 1 - n - \lambda _{n+1} , 2 - n - \lambda _{n+2} , \ldots , - \lambda _{2n} \}. \end{aligned}$$

We now define a labeling of the numberline \({\mathbb {Z}}\). The integers in \( I_\times (\lambda ) \cap I_\circ (\lambda ) \) are labeled by \(\vee \), the remaining ones in \(I_\times (\lambda )\) resp. \(I_\circ (\lambda )\) are labeled by \(\times \) respectively \(\circ \). All other integers are labeled by \(\wedge \). This labeling of the numberline uniquely characterizes the weight vector \(\lambda \). If the label \(\vee \) occurs r times in the labeling, then \(r=atyp(\lambda )\) is called the degree of atypicality of \(\lambda \). Notice \(0 \le r \le n\), and for \(r=n\) the weight \(\lambda \) is called maximal atypical. A weight is maximally atypical if and only if \(\lambda _i = - \lambda _{2n-i+1}\) for \(i=1,\ldots ,n\) in which case we write

$$\begin{aligned}L(\lambda ) = [\lambda _1,\ldots ,\lambda _n].\end{aligned}$$

To each weight diagram we associate a cup diagram as in [13, 36]. The outer cups in a cup diagram define the sectors of the weight as in [36]. We number the sectors from left to right \(S_1\), \(S_2\), \(\ldots \), \(S_k\).

Example 3.1

Consider the (maximal atypical) irreducible representation [7, 7, 4, 2, 2, 2] of GL(6|6). Its associated weight and cup diagram have two sectors:

figure a

3.2 Important invariants

The segment and sector structure of a weight diagram is completely encoded by the positions of the \(\vee \)’s. Hence any finite subset of \({{\mathbb {Z}}}\) defines a unique weight diagram in a given block. We associate to a maximal atypical highest weight the following invariants:

  • the type (SD) resp. (NSD),

  • the number \(k=k(\lambda )\) of sectors of \(\lambda \),

  • the sectors \(S_\nu =(I_\nu ,K_\nu )\) from left to right (for \(\nu =1,\ldots ,k\)),

  • the ranks \(r_\nu = r(S_\nu )\), so that \(\# I_\nu = 2r_\nu \),

  • the distances \(d_\nu \) between the sectors (for \(\nu =1,\ldots ,k-1\)),

  • and the total shift factor \(d_0=\lambda _n + n-1\).

If convenient, k sometimes may also denote the number of segments, but hopefully no confusion will arise from this.

A maximally atypical weight \([\lambda ]\) is called basic if \((\lambda _1,\ldots ,\lambda _n)\) defines a decreasing sequence \(\lambda _1 \ge \cdots \ge \lambda _{n-1} \ge \lambda _n=0\) with the property \(n-i \ge \lambda _i\) for all \(i=1,\ldots ,n\). The total number of such basic weights in \(X^+(n)\) is the Catalan number \(C_n\). Reflecting the graph of such a sequence \([\lambda ]\) at the diagonal, one obtains another basic weight \([\lambda ]^*\). By [36, Lemma 21.4] a basic weight \(\lambda \) is of type (SD) if and only if \([\lambda ]^* = [\lambda ]\) holds. To every maximal atypical highest weight \(\lambda \) is attached a unique maximal atypical highest weight \(\lambda _{basic}\)

$$\begin{aligned} \lambda \mapsto \lambda _{basic} \ \end{aligned}$$

having the same invariants as \(\lambda \), except that \(d_1=\cdots = d_{k-1}=0\) holds for \(\lambda _{basic}\) and the leftmost \(\vee \) is at the vertex \(-n+1\).

Example 3.2

In Example 3.1 the weight [7, 7, 4, 2, 2, 2] is of type (NSD). It has two \(k=2\) sectors of rank \(r_1 = 4\) and \(r_2=2\) with shift factor \(d_0=7\) and \(d_1 = 1\). Its associated basic weight is

figure b

4 Cohomological tensor functors

4.1 The Duflo–Serganova functor

We attach to every irreducible representation a sign. If \(L(\lambda )\) is maximally atypical in \({{\mathcal {R}}}_n\) we put \(\varepsilon (L(\lambda )) = (-1)^{p(\lambda )}\) for the parity \(p(\lambda ) = \sum _{i=1}^n \lambda _i\). For the general case see [36]. Now for \(\varepsilon \) define the full subcategories \({{\mathcal {R}}}_n(\varepsilon )\). These consists of all objects in \({{\mathcal {R}}}_n\) whose irreducible constituents L have sign \(\varepsilon (L) = \varepsilon \). Then by [36, Corollary 15.1] the categories \({{\mathcal {R}}}_n(\varepsilon )\) are semisimple categories.

Note that \({{\,\textrm{sdim}\,}}(X)\ge 0\) holds for all irreducible objects \(X\in {{\mathcal {R}}}_n(\varepsilon )\) in case \(\varepsilon (X)=1\) and also for all irreducible objects \(X\in \Pi {{\mathcal {R}}}_n(\varepsilon )\) in case \(\varepsilon (X)=-1\). For each irreducible representation \(L(\lambda )\) with \(sdim(X_\lambda ) \ne 0\) let

$$\begin{aligned} X_\lambda = \Pi ^{p(\lambda )}( L(\lambda )) \end{aligned}$$

denote the parity shift of \(L(\lambda )\) that satisfies \(sdim(X_\lambda )\ge 0\). In the case of the Berezin representation \(Ber =[1,\ldots ,1]\) we also write B for this parity shift. Notice, \(B=Ber\) if n is even and \(B=\Pi (Ber)\) if n is odd.

We recall some constructions from the article [36]. Fix the following element \(x\in \mathfrak {g}_1\),

$$\begin{aligned} x = \begin{pmatrix} 0 &{}\quad y \\ 0 &{}\quad 0 \end{pmatrix} \text { for } \ y = \begin{pmatrix} 0 &{}\quad 0 &{}\quad \ldots &{}\quad 0 \\ 0 &{}\quad 0 &{}\quad \ldots &{}\quad 0 \\ \ldots &{} &{}\quad \ldots &{} \\ 1 &{}\quad 0 &{}\quad 0 &{}\quad 0 \\ \end{pmatrix}. \end{aligned}$$

Since x is an odd element with \([x,x]=0\), we get

$$\begin{aligned} 2 \cdot \rho (x)^2 =[\rho (x),\rho (x)] =\rho ([x,x]) =0 \end{aligned}$$

for any representation \((V,\rho )\) of GL(n|n) in \({{{\mathcal {R}}}}_n\). Notice \(d= \rho (x)\) supercommutes with \(\rho (GL(n-1|n-1))\). Then we define the cohomological tensor functor DS as

$$\begin{aligned} DS = DS_{n,n-1}: \mathcal {T}_n \rightarrow \mathcal {T}_{n-1} \end{aligned}$$

via \(DS_{n,n-1}(V,\rho )= V_x:=Kern(\rho (x))/Im(\rho (x))\).

In fact DS(V) has a natural \({\mathbb {Z}}\)-grading and decomposes into a direct sum of \(GL(n-1|n-1)\)-modules

$$\begin{aligned} DS(V,\rho ) \ = \ \bigoplus _{\ell \in {\mathbb {Z}}} \ \Pi ^\ell (H^\ell (V)) \, \end{aligned}$$

for certain cohomology groups \(H^{\ell }(V)\). If we want to emphasize the \({\mathbb {Z}}\)-grading, we also write this in the form

$$\begin{aligned} DS(V,\rho ) \ = \ \bigoplus _{\ell \in {\mathbb {Z}}} \ H^\ell (V)[-\ell ].\end{aligned}$$

Theorem 4.1

[36, Theorem 16.1] Suppose \(L(\lambda )\in {{\mathcal {R}}}_n\) is an irreducible atypical representation, so that \(\lambda \) corresponds to a cup diagram

$$\begin{aligned} \bigcup _{j=1}^r \ \ [a_j,b_j] \end{aligned}$$

with r sectors \([a_j,b_j]\) for \(j=1,\ldots ,r\). Then

$$\begin{aligned}DS(L(\lambda )) \ \cong \ \bigoplus _{i=1}^r \ \Pi ^{n_i} L(\lambda _i)\end{aligned}$$

is the direct sum of irreducible atypical representations \(L(\lambda _i)\) in \({{\mathcal {R}}}_{n-1}\) with shift \(n_i \equiv p(\lambda ) - p(\lambda _i)\) modulo 2. The representation \(L(\lambda _i)\) is uniquely defined by the property that its cup diagram is

$$\begin{aligned}{}[a_i +1, b_i-1] \ \ \ \cup \ \ \bigcup _{j=1, j\ne i}^r \ \ [a_j,b_j] \,\end{aligned}$$

the union of the sectors \([a_j,b_j]\) for \(1\le j\ne i \le r\) and (the sectors occuring in) the segment \([a_i+1,b_i-1]\).

In particular \(DS(L(\lambda ))\) is semisimple and multiplicity free.

Example 4.2

Consider again the (maximal atypical) irreducible representation [7, 7, 4, 2, 2, 2] of GL(6|6) of Example 3.1. The parity is \(\varepsilon (\lambda ) = 1\). Applying DS gives 2 irreducible representations. The representation \([\lambda _1] = [7,7,4,2,2]\) is associated to the derivative of the first sector

figure c

Then the parity is \(\varepsilon (\lambda _1) = 1 = \varepsilon (\lambda )\). The second irreducible representation is \(\Pi [7,3,1,1,1]\) (note the parity shift since \(\varepsilon (\lambda _2) \ne \varepsilon (\lambda )\)) with cup diagram

figure d

All in all \(DS[7,7,4,2,2,2] \cong [7,7,4,2,2] \oplus \Pi [7,3,1,1,1]\).

4.2 The Hilbert polynomial

Similarly to DS we can define the tensor functors \(DS_{n,n-m}: \mathcal {T}_n \rightarrow T_{n-m}\) by replacing the x in the definition of DS by an x with m 1’s on the antidiagonal. These functors admit again a \({\mathbb {Z}}\)-grading. In particular we can consider the functor \(DS_{n,0}: \mathcal {T}_n \rightarrow T_0=svec_k\) with its decomposition \(DS_{n,0}(X) = \bigoplus _{\ell \in {\mathbb {Z}}} D_{n,0}^\ell (X)[-\ell ]\) for objects X in \(\mathcal {T}_n\) and objects \(D_{n,0}^\ell (X)\) in \(svec_k\) where \(D_{n,0}^\ell (X)[-\ell ]\) is the object \(\Pi ^{\ell }D_{n,0}^\ell (X)\) concentrated in degree \(\ell \) with respect to the \({\mathbb {Z}}\)-gradation of \(DS_{n,0}(X)\). For \(X\in \mathcal {T}_n\) we define the Laurent polynomial

$$\begin{aligned} \omega (X,t) = \sum _{\ell \in {\mathbb {Z}}} {{\,\textrm{sdim}\,}}(DS_{n,0}^\ell (X)) \cdot t^\ell \ \end{aligned}$$

as the Hilbert polynomial of the graded module \(DS^\bullet _{n,0}(X)= \bigoplus _{\ell \in {\mathbb {Z}}} DS_{n,0}^\ell (X)\). Since \({{\,\textrm{sdim}\,}}(W[-\ell ])=(-1)^\ell {{\,\textrm{sdim}\,}}(W)\) and \(X= \bigoplus DS_{n,0}^\ell (X)[-\ell ]\) holds, the formula

$$\begin{aligned}{{\,\textrm{sdim}\,}}(X) = \omega (X,-1)\end{aligned}$$

follows. For \(X= Ber_n^i\)

$$\begin{aligned} \omega \big (Ber_n^i,t\big ) \ = \ t^{ni}. \end{aligned}$$

For more details we refer the reader to [36, section 25].

4.3 The Dirac functor

In [36, Section 5] we also consider the Dirac operator

$$\begin{aligned} D = \partial + {\overline{\partial }}. \end{aligned}$$

Here \({\overline{x}} = x^T\) denotes the supertranspose of x and \({\overline{\partial }}= i \rho ({\overline{x}})\). Let \(H=diag(0_{n-1},1,1,0_{n-1})\) and \(M:= V^H\). Then we show that

$$\begin{aligned} H_D(V) = Kern(D: M \rightarrow M)/Im(D: M \rightarrow M) \ \end{aligned}$$

defines a symmetric monoidal functor \(\mathcal {T}_n \rightarrow \mathcal {T}_{n-s}\) where s is the rank of x. It follows from Lemma 5.8 that \(H_D\) agrees with DS on the subcategory \(\mathcal {T}_n^+\).

5 Tannakian arguments

5.1 The category \(\mathcal {T}_n^+\)

Let \(\mathcal {T}_n^+\) denote the Karoubian envelope of the simple nonnegative representations, i.e the full subcategory of \(\mathcal {T}_n\), whose objects consist of all retracts of iterated tensor products of irreducible representations in \(\mathcal {T}_n\) that are not maximal atypical and of maximal atypical irreducible representations \(X_\lambda \) in \({{\mathcal {R}}}_n(+1) \oplus \Pi {{\mathcal {R}}}_n(-1)\), defined as at the begining of Sect. 4.1. Obviously \(\mathcal {T}_n^+\) is a symmetric monoidal idempotent complete k-linear category closed under the \(*\)-involution. It contains all irreducible objects of \(\mathcal {T}_n\) up to a parity shift. It contains the standard representation V and its dual \(V^\vee \), and hence contains all mixed tensors [35]. Furthermore all objects X in \(\mathcal {T}_n^+\) satisfy condition \(\mathtt T\) (see section 6 in [36]) and \(\mathcal {T}_n^+\) is rigid. For this it suffices for irreducible \(X\in \mathcal {T}_n^+\) that \(X^\vee \in \mathcal {T}_n^+\). This is obvious since \(X^\vee \) is irreducible with \({{\,\textrm{sdim}\,}}(X^\vee ) = {{\,\textrm{sdim}\,}}(X) \ge 0\), and hence \(X^\vee \in \mathcal {T}_n^+\).

5.2 Conventions on tensor categories

We use the same definition of a tensor category as is used in [27] except that we do not require the category to be abelian. Tensor functors are additive as in [27] but need not be exact. Our definition of a tensor functor therefore agrees with the one used in [22].

5.3 The ideal of negligible morphisms

An ideal in a k-linear category \({{\mathcal {A}}}\) is for any two objects XY the specification of a k-submodule \(\mathcal {I}(X,Y)\) of \(Hom_{{{\mathcal {A}}}}(X,Y)\), such that \(g \mathcal {I}(X',Y)f \ \subseteq \mathcal {I}(X,Y')\) holds for all pairs of morphisms \(f \in Hom_{{{\mathcal {A}}}}(X,X')\), \( g \in Hom_{{{\mathcal {A}}}}(Y,Y')\). Let \(\mathcal {I}\) be an ideal in \({{\mathcal {A}}}\). By definition \({{\mathcal {A}}}/\mathcal {I}\) is the category with the same objects as \({{\mathcal {A}}}\) and with

$$\begin{aligned} Hom_{{{\mathcal {A}}}/\mathcal {I}} (X,Y ) = Hom_{{{\mathcal {A}}}}(X,Y)/\mathcal {I}(X,Y).\end{aligned}$$

An ideal in a tensor category is a tensor ideal if it is stable under \({\textbf{1}}_C \otimes -\) and \(- \otimes {\textbf{1}}_C\) for all \(C \in {{\mathcal {A}}}\). Let Tr be the trace. For any two objects AB we define \(\mathcal {N}(A,B) \subset Hom(A,B)\) by

$$\begin{aligned} \mathcal {N}(A,B) = \{ f \in Hom(A,B) \ | \ \forall g \in Hom(B,A), \ Tr(g \circ f ) = 0 \}. \end{aligned}$$

The collection of all \(\mathcal {N}(A,B)\) defines a tensor ideal \(\mathcal {N}\) of \({{\mathcal {A}}}\) [2].

Let \({{\mathcal {A}}}\) be a super tannakian category. An indecomposable object will be called negligible, if its image in \({{\mathcal {A}}}/{{\mathcal {N}}}\) is the zero object. By [34] an object is negligible if and only if its categorial dimension is zero.

Example 5.1

An irreducible representation has superdimension zero if and only if it is not maximal atypical, see Sect. 3. The standard representation \(V \simeq k^{n|n}\) has superdimension zero and therefore also the indecomposable adjoint representation \(\mathbb {A} = V \otimes V^{\vee }\).

Any super tannakian category is equivalent (over an algebraically closed field) to the representation category of a supergroup scheme by [20]. In that case the categorial dimension is the superdimension of a module. If \({{\mathcal {A}}}\) is a super tannakian category over k, the quotient of \({{\mathcal {A}}}\) by the ideal \({{\mathcal {N}}}\) of negligible morphisms is again a super tannakian category by [2, 34]. More generally, for any pseudo-abelian full subcategory \({\tilde{{{\mathcal {A}}}}}\) in \({{\mathcal {A}}}\) closed under tensor products, duals and containing the identity element the following holds:

Lemma 5.2

The quotient category \({\tilde{{{\mathcal {A}}}}}/{{\mathcal {N}}}_{\tilde{\mathcal {A}}}\) is a semisimple super tannakian category.

Proof

The quotient is a k-linear semisimple rigid tensor category by [3, Theorem 1 a)]. The quotient is idempotent complete by lifting of idempotents (or see [2, 2.3.4 b)] and by [2, 2.1.2] a k-linear pseudoabelian category is abelian. The Schur finiteness [20, 34] is inherited from \({{\mathcal {A}}}\) to \({\tilde{{{\mathcal {A}}}}}/{{\mathcal {N}}}\). \(\square \)

This in particular applies to the situation where \({\tilde{{{\mathcal {A}}}}}\) is the full subcategory of objects which are retracts of iterated tensor products of a fixed set of objects in \({{\mathcal {A}}}\). In particular for \({\tilde{{{\mathcal {A}}}}} = \mathcal {T}_n^+\) and \({{\mathcal {A}}}=\mathcal {T}_n\) this implies

Corollary 5.3

The tensor functor \(\mathcal {T}_n^+ \rightarrow \mathcal {T}_n^+/{{\mathcal {N}}}\) maps \(\mathcal {T}_n^+\) to a semisimple super tannakian category \(\overline{\mathcal {T}}_{n}:=\mathcal {T}_n^+/{{\mathcal {N}}}\).

Proposition 5.4

The category \(\overline{\mathcal {T}}_{n}\) is a tannakian category, i.e. there exists a pro-reductive algebraic k-groups \(H_n\) such that the category \(\overline{\mathcal {T}}_{n}\) is equivalent as a tensor category to the category \(Rep_k(H_n)\) of finite dimensional k-representations of \(H_n\)

$$\begin{aligned} \overline{\mathcal {T}}_{n} \sim Rep_k(H_n). \end{aligned}$$

Proof

By a result of Deligne [21, Theorem 7.1] it suffices to show that for all objects X in \(\mathcal {T}_n^+\) we have \({{\,\textrm{sdim}\,}}(X)\ge 0\). We prove this by induction on n. Suppose we know this assertion for \(\mathcal {T}_{n-1}\) already. Then all objects of \( \mathcal {T}_{n-1}^+\) have superdimension \(\ge 0\) (for the induction start \(n=0\) our assertion is obvious). Since the tensor functor \(DS: \mathcal {T}_n \rightarrow \mathcal {T}_{n-1}\) preserves superdimensions, it suffices for the induction step that DS maps \(\mathcal {T}_n^+\) to \(\mathcal {T}_{n-1}^+\). \(\square \)

Lemma 5.5

The functors \(DS_{n,n-m}: \mathcal {T}_n \rightarrow \mathcal {T}_{n-m}\) and \(\omega _{n,n-m}: \mathcal {T}_n \rightarrow \mathcal {T}_{n-m}\) restrict to functors from \(\mathcal {T}_n^+\) to \(\mathcal {T}_{n-m}^+\). In particular

$$\begin{aligned} DS: \mathcal {T}_n^+ \rightarrow \mathcal {T}_{n-1}^+. \end{aligned}$$

Proof

Since \(DS_{n,n-m}\) and \(\omega _{n-m}\) preserve tensor products and idempotents, it suffices by the definition of \(\mathcal {T}_n^+\) that \(DS_{n-m}(X), \omega _{n-m}(X) \in \mathcal {T}_{n-m}^+\) holds for all irreducible objects X in \(\mathcal {T}_n^+\). Now Theorem 4.1 implies \(DS(X) \in \mathcal {T}_{n-1}^+\) since any irreducible representation X maps to a semisimple representation DS(X) and for maximal atypical \(X \in \mathcal {T}_{n-1}^+\) all summands of DS(X) are in \(\mathcal {T}_{n-1}^+\). This proves the claim for DS(X), X irreducible. But then also for \(DS_{n,n-m}(X)\), X irreducible, since then again \(DS_{n,n-m}(X)\) is semisimple by proposition 8.1 in [36]. The same then also holds for \(\omega _{n,n-m}(X) = H_{{\overline{\partial }}}(DS_{n-m}(X))\) by loc.cit. \(\square \)

Corollary 5.6

Under DS negligible objects in \(\mathcal {T}_n^+\) map to negligible objects in \(\mathcal {T}_{n-1}^+\).

Proof

We have shown \({{\,\textrm{sdim}\,}}(Y)\ge 0\) for all objects Y in \(\mathcal {T}_{n-1}^+\). Therefore \({{\,\textrm{sdim}\,}}(DS(X))= {{\,\textrm{sdim}\,}}(X)=0\) implies \(sdim(Y_i)=0\) for all indecomposable summands \(Y_i\) of \(Y=DS(X)\), since \(sdim(Y_i)\ge 0\). \(\square \)

Remark 5.7

Since irreducible objects L satisfy condition T in the sense that \({\overline{\partial }}\) is trivial on \(DS_{n,n-m}(L)\) [36, proposition 8.5], and since condition T is inherited by tensor products and retracts, all objects in \(\mathcal {T}_n^+\) satisfy condition T. Hence [36, proposition 8.5] implies the following lemma.

Lemma 5.8

On the category \(\mathcal {T}^+_n\) the functor \(H_D(.)\) is naturally equivalent to the functor \(DS: \mathcal {T}_n^+ \rightarrow \mathcal {T}_{n-1}^+\). Similarly the functors \(\omega _{n,n-m}(.): \mathcal {T}_n^+ \rightarrow \mathcal {T}_{n-1}^+\) are naturally equivalent to \(DS_{n,n-m}(.)\).

Corollary 5.9

\(DS(X)=0\) in \(\mathcal {T}_{n-1}^+\) if and only if X is a projective object in \(\mathcal {T}_n\).

Proof

Any negligible maximal atypical object in \(\mathcal {T}_n^+\) (i.e. a negligible object in the principal block) maps under DS to a negligible maximal atypical object in \(\mathcal {T}_{n-1}^+\). Furthermore \(DS(X)=0\) for X in \(\mathcal {T}_n^+\) implies that X is an anti-Kac object. If \(X\ne 0\), then \(X^*\) is a Kac object in \(\mathcal {T}_n^+\). Hence \(H_D(X^*)=0\). Since \(X^*\in \mathcal {T}_n^+\) satisfies condition T, this implies \(DS(X^*)=0\) and hence \(X^*\) is a Kac and anti-Kac object. The corollary follows since \(\mathcal {C}^+ \cap \mathcal {C}^- = Proj\). \(\square \)

Corollary 5.10

If \(X \in \mathcal {T}_n^+\) and X is a Kac or anti-Kac object, then \(X \in Proj\).

Even though negligible objects map to negligible objects by Corollary 5.6, it is highly non-trivial negligible morphisms map to negligible morphism and that we therefore get an inducted functor between the quotient categories.

Lemma 5.11

The functor \(DS: \mathcal {T}_n^+ \rightarrow \mathcal {T}_{n-1}^+\) gives rise to a k-linear exact tensor functor between the quotient categories

$$\begin{aligned} \eta : \overline{\mathcal {T}}_{n} \rightarrow \overline{\mathcal {T}}_{n-1}.\end{aligned}$$

In particular the iterated tensor functor \(\omega = \eta \circ \ldots \eta :\overline{\mathcal {T}}_n \rightarrow vec_k\) defines a fibre functor for \(\overline{\mathcal {T}}_n\).

Proof

We define the ideal \(\mathcal {I}^0\) via

$$\begin{aligned} \mathcal {I}^0(X,Y) = \{ f:X \rightarrow Y \ | \ f \text { factorizes over a negligible object.} \} \end{aligned}$$

Obviously \(\mathcal {I}^0\) is a tensor ideal for \(\mathcal {T}_n^+\). As for any tensor ideal \(\mathcal {I}^0 \subset \mathcal {N}\) the quotient \(\mathcal {T}_n^+/\mathcal {I}^0 =: \mathcal {A}_n^+\) becomes a rigid tensor category and \(\mathcal {T}_n^+\rightarrow \mathcal {T}_n^+/\mathcal {I}^0 = \mathcal {A}_n^+\) a tensor functor. Under this tensor functor an indecomposable object X in \(\mathcal {T}_n^+\) maps to zero in the quotient \(\mathcal {A}_n^+\) if and only if \({{\,\textrm{sdim}\,}}(X) = 0\). Furthermore, since the tensor functor DS maps negligible objects of \(\mathcal {T}_n^+\) to negligible objects of \(\mathcal {T}_{n-1}^+\), the functor DS induces a k-linear tensor functor \(DS': \mathcal {A}_n^+ \rightarrow \mathcal {A}_{n-1}^+\). The category \(\mathcal {A}_n^+\) is pseudoabelian since we have idempotent lifting in the sense of [44, Theorem 5.2] due to the finite dimensionality of the Hom spaces. By the definition of \(\mathcal {A}_n^+\) and \(\mathcal {T}_n^+\), the dimension of each object in \(\mathcal {A}_n^+\) is a natural number and, contrary to \(\mathcal {T}_n^+\), it does not contain any nonzero object that maps to an element isomorphic to zero under the quotient functor \(\mathcal {A}_n^+ \rightarrow \mathcal {A}_n^+/\mathcal {N}\). Therefore \(\mathcal {A}_n^+\) satisfies conditions d) and g) in [2, Theorem 8.2.4]. By [2, Theorem 8.2.4 (i),(ii)] this implies that \(\mathcal {N}( \mathcal {A}_n^+)\) equals the radical \(\mathcal {R}( \mathcal {A}_n^+)\) of \( \mathcal {A}_n^+\); note that \(\mathcal {N}( \mathcal {A}_n^+) = \mathcal {N}( \mathcal {T}_n^+)/\mathcal {I}^0\) and that \({{\mathcal {N}}}(A,A)\) is a nilpotent ideal in End(A) for any A in \(\mathcal {A}_n^+\) by assertion b) of [2, Theorem 8.2.4 (i),(ii)]. Since \(\mathcal {N}\) always is a tensor ideal, \(\mathcal {R}( \mathcal {A}_n^+)\) in particular is a tensor ideal. This allows to apply [2, Theorem 13.2.1] to construct a monoidal section \(s_n: \mathcal {A}^+_n/\mathcal {N}(\mathcal {A}_n^+) \rightarrow \mathcal {A}^+_n\) for the tensor functor \(\pi _n: \mathcal {A}^+_n \rightarrow \mathcal {A}^+_n/\mathcal {N}(\mathcal {A}_n^+)\). The composite tensor functor

$$\begin{aligned} \eta :=\pi _{n-1}\circ DS' \circ s_n \end{aligned}$$

defines a k-linear tensor functor

$$\begin{aligned} \eta : \overline{\mathcal {T}}_{n} \rightarrow \overline{\mathcal {T}}_{n-1}. \end{aligned}$$

Since \(DS'\) is additive and \(\overline{\mathcal {T}}_{n}\) is semisimple, \(\eta \) is additive and hence exact. \(\square \)

Remark 5.12

  1. (1)

    The k-linear tensor functor \(\pi _{n-1}\circ DS': \mathcal {A}_n^+ \rightarrow \overline{\mathcal {T}}_{n-1}\) defines the tensor ideal \({{\mathcal {K}}}_n\) of \(\mathcal {A}_n^+\) of morphisms annihilated by \(\pi _{n-1}\circ DS'\). Obviously \( {{\mathcal {K}}}_n \subseteq {{\mathcal {N}}}\).

  2. (2)

    Let S be the image of a simple object in \(\mathcal {A}_n^+\). Since \( {{\mathcal {N}}}(\mathcal {A}_n^+)= {{\mathcal {R}}}(\mathcal {A}_{n}^+)\), some given morphism \(f\in Hom_{\mathcal {A}_{n}^+}(S,A)\) is in \( {{\mathcal {N}}}(\mathcal {A}_{n}^+)(S,A)\) if and only if for all \( g \in Hom_{\mathcal {A}_{n}^+}(S,A)\) the composite \(g\circ f\) is zero [2, Lemma 1.4.9] (note that the endomorphisms of S in \(\mathcal {A}_{n}^+\) are in \(k\cdot id\), hence [2, Lemma 1.4.9] can be applied.).

  3. (3)

    By [2, Theorem 13.2.1] the section \(s_n\) is unique up to isomorphism. Therefore the functor \(\eta \) so constructed is unique up to isomorphism.

Remark 5.13

We do not know whether \(DS({{\mathcal {N}}}(\mathcal {T}_n^+)) \subseteq {{\mathcal {N}}}(\mathcal {T}_{n-1}^+)\) holds. If this were true for all n, then also \(DS_{n,n-i}({{\mathcal {N}}}(\mathcal {T}_n^+)) \subseteq {{\mathcal {N}}}(\mathcal {T}_{n-i}^+)\) would hold. We consider this a fundamental question in the theory. For \(n=1\) observe that \(\mathcal {A}_1^+ = \mathcal {T}_1^+/\mathcal {N}\). Indeed \(\mathcal {T}_1^+\) has only one proper tensor ideal \(\mathcal {N} = \mathcal {I}^0\) as can be easily seen by looking at the maximal atypical objects \(Ber^i\) and \(P(Ber^j)\) in \(\mathcal {T}_1^+\). The tensor ideal \(\mathcal {I}^0\) could be different from \(\mathcal {N}\) for \(n \ge 2\). With respect to the partial ordering on the set of tensor ideals given by inclusion, \(\mathcal {I}^0\) is the minimal element in the fibre of the decategorification map of the thick ideal of indecomposable objects of superdimension 0 [19, Theorem 4.1.3]. The negligible morphisms are the largest tensor ideal in this fibre.

Example 5.14

Note that it is important here to work in \(\mathcal {T}_n^+\) since for example \(DS(K({\textbf{1}}))\) (where \(K({\textbf{1}})\) is the Kac-module of the trivial representation) splits into a direct sum of maximal atypical irreducible modules (see [36, Section 10]). Hence the identity morphism of \(K({\textbf{1}})\) does not map to a negligible morphism. There are even counter examples in the smaller category \(\mathcal {T}^{ev}\) of Sect. 14. In \(\mathcal {T}_{1}\) consider the indecomposable ZigZag module (see [34]) with socle \(Ber^{-1}\) and Ber and top \({\textbf{1}}\). The inclusion of \(Ber^{-1}\) induces an isomorphism when taking DS-cohomology. On the other hand the inclusion is negligible. Note that \(Ber^{-1}\) is odd. The ZigZag module can be obtained as \({\textbf{1}}[1]\) in the stable category \(\mathcal {K}\). Since \({\textbf{1}}\) is even, \({\textbf{1}}[1]\) is odd. Hence their parity shifts define even objects in \(\mathcal {T}_1^{ev}\).

5.4 DS as a restriction functor

Recall from [21, Theorem 8.17] the following fundamental theorem on k-linear tensor categories: Suppose \({{\mathcal {A}}}_1, {{\mathcal {A}}}_2\) are k-linear abelian rigid symmetric monoidal tensor categories with \(k \cong End_{{{\mathcal {A}}}_i}(\textbf{1})\) as in loc. cit. Assume that all objects of \({{\mathcal {A}}}_i\) have finite length and all Hom-groups have finite k-dimension. Assume that k is a perfect field so that \({{\mathcal {A}}}_1 \otimes {{\mathcal {A}}}_2\) is again k-linear abelian rigid symmetric monoidal tensor categories with \(k \cong End_{{{\mathcal {A}}}_i}(\textbf{1})\) as in [21, 8.1]. Suppose

$$\begin{aligned} \eta : {{\mathcal {A}}}_1 \rightarrow {{\mathcal {A}}}_2 \end{aligned}$$

is an exact tensor functor. Then \(\eta \) is faithful [22, Proposition 1.19].

Theorem 5.15

[21, Theorem 8.17] Under the assumptions above there exists a morphism

$$\begin{aligned} \pi ({{\mathcal {A}}}_2) \rightarrow \eta (\pi ({{\mathcal {A}}}_1)) \end{aligned}$$

as in [21, 8.15.2] such that \(\eta \) induces a tensor equivalence between the category \({{\mathcal {A}}}_1\) and the tensor category of objects in \({{\mathcal {A}}}_2\) equipped with an action of \(\eta (\pi ({{\mathcal {A}}}_1))\), so that the natural action of \(\pi ({{\mathcal {A}}}_2))\) is obtained via the morphism \( \pi ({{\mathcal {A}}}_2) \rightarrow \eta (\pi ({{\mathcal {A}}}_1)) \).

Suppose \(\omega : {{\mathcal {A}}}_2 \rightarrow Vec_k\) is fiber functor of \({{\mathcal {A}}}_2\), i.e. \(\omega \) is an exact faithful tensor functor. Then \({{\mathcal {A}}}_2\) is a Tannakian category and \({{\mathcal {A}}}_2 \cong Rep_k(H)\) as a tensor category. If \({{\mathcal {A}}}_2 = Rep_k(H)\) is a Tannakian category for some affine group H over k, then \(\pi ({{\mathcal {A}}}_2)= H\) by [21, Example 8.14 (ii)]. More precisely, an \({{\mathcal {A}}}_2\)-group is the same as an affine k-group equipped with an H-action, and here H acts on itself by conjugation. The forgetful functor \(\omega \) of \(Rep_k(G)\) to \(Vec_k\) is a fiber functor. By applying this fiber functor we obtain a fiber functor \(\omega \circ \eta : {{\mathcal {A}}}_1 \rightarrow Vec_k\) for the tensor category \({{\mathcal {A}}}_1\). In particular \({{\mathcal {A}}}_1\) becomes a Tannakian category with Tannaka group \(H'= \omega \circ \eta (\pi ({{\mathcal {A}}}_1))\). Furthermore, by applying \(\eta \) to the morphism \( \pi ({{\mathcal {A}}}_2) \rightarrow \eta (\pi ({{\mathcal {A}}}_2)) \) in \({{\mathcal {A}}}_2\), we get a morphism \(\omega (\pi ({{\mathcal {A}}}_2)) \rightarrow (\omega \circ \eta )(\pi ({{\mathcal {A}}}_1))\) in the category of k-vectorspaces, which defines a group homomorphism

$$\begin{aligned} f: H' \rightarrow H \end{aligned}$$

of affine k-groups inducing a pullback functor

$$\begin{aligned}Rep(H') \rightarrow Rep(H) \,\end{aligned}$$

that gives back the functor \(\eta : {{\mathcal {A}}}_1 \rightarrow {{\mathcal {A}}}_2\) via the equivalences \({{\mathcal {A}}}_1=Rep_k(H')\) and \({{\mathcal {A}}}_2 =Rep_k(H)\) obtained from the fiber functors.

Lemma 5.16

[22, Proposition 2.21(b)] The morphism \(f:H' \rightarrow H\) thus obtained is a closed immersion if and only if every object Y of \({{\mathcal {A}}}_2\) is isomorphic to a subquotient of an object of the form \(\eta (X), X \in {{\mathcal {A}}}_1\).

The statements above will now be applied for the tensor functor

$$\begin{aligned} \eta : {{\mathcal {A}}}_1 \rightarrow {{\mathcal {A}}}_2 \end{aligned}$$

obtained from DS between the quotient categories \({{\mathcal {A}}}_1= \mathcal {T}_n^+/{{\mathcal {N}}}\) and \({{\mathcal {A}}}_2= \mathcal {T}_{n-1}^+/{{\mathcal {N}}}\). Notice that the assumptions above on k and \({{\mathcal {A}}}_i\) are satisfied so that \({{\mathcal {A}}}_2\) is a tannakian category with fiber functor \(\omega \) giving an equivalence of tensor categories \({{\mathcal {A}}}_2 = Rep_k(H_{n-1})\). Obviously \(\eta \) induces an exact tensor functor between the quotient categories, since DS is additive, maps negligible objects of \(\mathcal {T}_n^+\) into negligible objects of \(\mathcal {T}_{n-1}^+\) and since the categories \({{\mathcal {A}}}_i\) are semisimple. As in our case k is algebraically closed, we know that up to an isomorphism the group \(H_n\) only depends on \({{\mathcal {A}}}_1\) but not on the choice of a fiber functor. As explained above, this defines a homomorphism of affine k-groups (uniquely defined up to conjugacy) \( f: H_{n-1} \longrightarrow H_n \).

Theorem 5.17

The homomorphism \(f:H_{n-1} \rightarrow H_n\) is injective and the functor \(\eta : Rep_k(H_{n})\rightarrow Rep_k(H_{n-1})\) induced by \(DS: \mathcal {T}_n^+ \rightarrow \mathcal {T}_{n-1}^+\) can be identified with the restriction functor for the homomorphism f.

Proof

By Lemma 5.16 it suffices that every indecomposable Y in \(\mathcal {T}_{n-1}^+\) with \({{\,\textrm{sdim}\,}}(Y) >0\) is a subobject of an object \(DS(X), X\in \mathcal {T}_{n}^+\). By assumption Y is a retract of a tensor product of irreducible modules \(L_i \in \mathcal {T}_{n-1}^+\). So it suffices that each \(L_i\) is a subobject of some object \(DS(X_i), X_i\in \mathcal {T}_{n}^+\). We can assume that Y is not negligible and irreducible, hence maximal atypical and \(Y=\Pi ^{r}L(\lambda )\) for some r. Then \(L(\lambda ) = [\lambda ] =[\lambda _1,\ldots ,\lambda _{n-1}]\). By a twist with Berezin we may assume that \(\lambda _{n-1}\ge 0\). Then we define \([{\tilde{\lambda }}] = [\lambda _1,\ldots ,\lambda _{n-1},0]\) so that for \(X = \Pi ^rL({\tilde{\lambda }})\) we get by Theorem 4.1 and [36, Lemma 10.2] the assertion \(DS(X) = Y \oplus \) other summands. Notice that by construction \(X = \Pi ^rL({\tilde{\lambda }})\) is in \(\mathcal {T}_n^+\). But this proves our claim. \(\square \)

In other words, the description of the functor DS on irreducible objects in \(\mathcal {T}_n\) given by Theorem 4.1 can be interpreted as branching rules for the inclusion

$$\begin{aligned} f: \ H_{n-1} \hookrightarrow H_n. \end{aligned}$$

We will show later how this fact gives information on the groups \(H_n\).

5.5 Enriched morphism

Using the \({\mathbb {Z}}\)-grading of DS (see Sect. 4.1), we can define an extra structure on the tower of Tannaka groups. This extra structure will not be used in the later determination of the Tannaka groups. The collection of cohomology functors \(H^i: {{\mathcal {R}}}_n \rightarrow {{\mathcal {R}}}_{n-1}\) for \(i\in {\mathbb {Z}}\) defines a tensor functor

$$\begin{aligned} H^\bullet : {{\mathcal {R}}}_n \rightarrow Gr^\bullet ({{\mathcal {R}}}_{n-1}) \ \end{aligned}$$

to the category of \({\mathbb {Z}}\)-graded objects in \({{\mathcal {R}}}_{n-1}\). Using the parity shift functor \(\Pi \), this functor can be extended to a tensor functor

$$\begin{aligned} H^\bullet : \mathcal {T}_n^+ \rightarrow Gr^\bullet \big (\mathcal {T}_{n-1}^+\big ) \, \end{aligned}$$

which induces a corresponding tensor functor on the level of the quotient categories

$$\begin{aligned} H^\bullet : \overline{\mathcal {T}}_{n}=\mathcal {T}_n^+/{{\mathcal {N}}}\rightarrow Gr^\bullet \big (\mathcal {T}_{n-1}^+/{{\mathcal {N}}}\big )\big ) = Gr^\bullet (\overline{\mathcal {T}}_{n-1}).\end{aligned}$$

Using the language of tannakian categories this induces an ’enriched’ group homomorphism

$$\begin{aligned} f^\bullet :\ H_{n-1} \times {\mathbb {G}}_m \rightarrow H_{n}. \end{aligned}$$

Its restriction to the subgroup \(1 \times H_{n-1}\) is the homomorphism f from above.

5.6 The involution \(\tau \)

Note that the category \(\mathcal {T}_n^+\) is closed under \(\vee \) and \(*\) and hence is equipped with the tensor equivalence \(\tau : X \mapsto (X^\vee )^*\). This tensor equivalence induces a tensor equivalence of \(\overline{\mathcal {T}}_{n}= \mathcal {T} _n^+/{{\mathcal {N}}}\) and hence an automorphism \(\tau =\tau _n\) (unique up to conjugacy) of the group \(H_n\). Since all objects of \(\mathcal {T}_n^+\) satisfy property \(\mathtt T\) [36, Section 6], the involution \(*\) commutes with DS. Since this also holds for the Tannaka duality, we get a compatibility

$$\begin{aligned} (H_{n-1},\tau _{n-1}) \hookrightarrow (H_n,\tau _n). \end{aligned}$$

5.7 Characteristic polynomial

By iteration the morphisms \(f^\bullet \) successively define homomorphisms \( H_{n-i}\times ({\mathbb {G}}_m)^{i} \rightarrow H_n\) and therefore we get a homomorphism in the case \(i=n\)

$$\begin{aligned}h: ({\mathbb {G}}_m)^n \rightarrow H_n.\end{aligned}$$

This allows to define a characteristic polynomial, defined by the restriction \(h^*(V_X)\) of the representation \(V_X=\omega (X)\) of H to the torus \(({\mathbb {G}}_m)^n\)

$$\begin{aligned} h_X(t) = \sum _{\chi } \dim (h^*(V_X)_\chi )\cdot t^\chi \end{aligned}$$

where \(\chi \) runs over the characters \(\chi =(\nu _1,\ldots ,\nu _n) \in {\mathbb {Z}}^n = {\mathbb {X}}^*(({\mathbb {G}}_m)^n)\). It is easy to see that \(\omega (X,t)=h_X(t,\ldots ,t)\) (see Sect. 4.2).

6 The structure of the derived connected groups \(G_n\)

6.1 Setup and notations

The Tannaka group generated by the object \(X_\lambda =\Pi ^{p(\lambda )}L(\lambda )\) for \(p(\lambda ) = \sum _{i=1}^{n} \lambda _i\) will be denoted \(H_\lambda \) and we define

$$\begin{aligned} G_\lambda := \big (H_\lambda ^0\big )_{der} \subseteq H_{\lambda }^0 \subseteq H_\lambda . \end{aligned}$$

Finally define \(V_\lambda \in Rep(H_\lambda )\) as the irreducible finite dimensional faithful representation (or the underlying vector space) of \(H_\lambda \) corresponding to \(X_{\lambda }\), i.e. the representation \(\omega (X_{\lambda })\) for the fibre functor \(\omega \) defined in Sect. 5.

(SD)-Types. Now assume for \(L(\lambda )\) that

$$\begin{aligned}\varphi : L(\lambda )^{\vee } \cong L(\lambda ) \otimes Ber^{-r}\end{aligned}$$

holds for some \(r \in \mathbb {Z}\) and some isomorphism \(\varphi \) in \(\mathcal {T}_{\lambda }\). The evaluation morphism \(eval: L(\lambda )^{\vee } \otimes L(\lambda ) \rightarrow {\textbf{1}}\) and the isomorphism \(L(\lambda )^{\vee } \cong L(\lambda ) \otimes Ber^{-r}\) gives rise to a nondegenerate pairing

$$\begin{aligned} L(\lambda ) \otimes L(\lambda ) \rightarrow Ber^r \ \end{aligned}$$

As a nondegenerate pairing of the simple object \(L(\lambda )\) it is even either or odd, where the parity \(\varepsilon _\lambda \in \{\pm 1\} \) is given by \(\varphi ^\vee = \varepsilon _\lambda \cdot \varphi \) (see “Appendix C”, where this sign will be computed). If we replace \(L_\lambda \) by the parity shift \(X_\lambda = \Pi ^{p(\lambda )}(L_\lambda )\), our pairing on \(L(\lambda )\) induces a pairing \( X_\lambda \otimes X_\lambda \rightarrow Ber^r \) on \(X_\lambda \). For this notice that \(\Pi ^2(L\otimes L) \cong \Pi (L) \otimes \Pi (L)\) holds and the underlying vectorspaces of \(L(\lambda )\) and \(X_\lambda \) coincide. Notice that nr is even by Lemma C.7 so that \(B^r = Ber^r\) always holds. The resulting pairing

$$\begin{aligned} X_\lambda \otimes X_\lambda \rightarrow B^r \end{aligned}$$

will be denoted \(<\cdot ,\cdot>\). Since a parity shift switches symmetric and antisymmetric pairings [46], the parity \(\varepsilon (X_\lambda )\) of the induced pairing \(<\cdot ,\cdot>\) for \(X_\lambda \) is

$$\begin{aligned} \varepsilon (X_\lambda ) = (-1)^{p(\lambda )} \varepsilon _\lambda . \end{aligned}$$

This sign \(\varepsilon (X_\lambda )\) will only depend on \(\lambda _{basic}\) by Lemma 6.1. Since \(<\cdot ,\cdot>\) is uniquely defined up to a nonvanishing constant, we will fix the pairing once for all.

The nondegenerate pairing \(<\cdot ,\cdot>\) on \(X_\lambda \) induces a nondegenerate pairing on \(V_\lambda \) of the same parity \(\varepsilon (X_\lambda )\), and will also be denoted \(<\cdot ,\cdot>\) by abuse of notation. Indeed, since \(<\cdot ,\cdot>\) is defined in terms of Tannaka duality, the evaluation morphism eval and the isomorphism \(\varphi \), this follows by functoriality.

The Tannaka group \(H_{\lambda }\) of \(\mathcal {T}_{\lambda } = {<} X_\lambda {>}\) acts faithfully on \(V_\lambda \) such that \( {<}h v_1,h v_2{>} \ = \ \mu (h) \cdot {<}v_1,v_2{>} \) holds for all \(h\in H_\lambda \) and the similitude character \(\mu : GSp(V_\lambda )\rightarrow \mathbb {G}_m\) of the pairing \(<\cdot ,\cdot>\) on \(V_\lambda \).

For a vector space \(V_{\lambda }\) over an algebraically closed field with a nondegenerate symmetric or antisymmetric pairing \(\langle .,.\rangle \) let

$$\begin{aligned} G(V_{\lambda },\langle .,.\rangle ) = \{ g \in GL(V_{\lambda }) | {<}gv,gw{>} = \mu (g){<}v,w{>}, \forall v,w \in V_{\lambda } \} \end{aligned}$$

be the similitude group with its similitude character \(\mu : G(V_{\lambda },\langle .,.\rangle ) \rightarrow k^*\). The sign character \(sgn: G(V_{\lambda },\langle .,.\rangle ) \rightarrow \mu _2\) is defined by

$$\begin{aligned} sgn(g) = \frac{det(g)}{\mu (g)^m}. \end{aligned}$$

Notice that \(\dim (V_{\lambda }) = 2m\) by Lemma D.4 will always be even unless \(X(\lambda )\) has dimension one and hence is a power of B. In the symmetric resp. antisymmetric case this above similtude group defines the orthogonal similitude group \(GO(V_{\lambda })\) resp. the symplectic similitude group \(GSp(V_{\lambda })\).

In the GSp-case sgn is trivial and the kernel of \(\mu \) is the connected symplectic group \(Sp(V_{\lambda })\). In the GO-case the kernel of \(\mu \) is the orthogonal group \(O(V_{\lambda })\), and the kernel of sgn on \(O(V_{\lambda })\) is the connected group \(SO(V_{\lambda })\). The kernel of sgn on \(GO(V_{\lambda })\) is the connected subgroup \(GSO(V_{\lambda })\).

The Tannaka group \(H_{\lambda }\) of the Tannaka category \(\mathcal {T}_{\lambda } = {<} X_\lambda {>}\) generated by \(X_\lambda \) acts faithfully on \(V_\lambda = \omega (X_\lambda )\) such that \( {<}h v_1,h v_2{>} \ = \ \mu (h) \cdot {<}v_1,v_2{>} \) holds for all \(h\in H_\lambda \). Hence \(H_\lambda \) is a subgroup of \(G\bigl (V_{\lambda },\langle .,.\rangle \bigr )\).

A priori bounds. We distinguish two cases: Either \(X_\lambda \) is a weakly selfdual object (SD), i.e. \(X_\lambda ^\vee \cong B^r \otimes X_\lambda \) for some r; or alternatively \(X_\lambda \) is not weakly selfdual (NSD). Summarizing we obtain the following bounds for the groups \(H_{\lambda }\). We have \( H_\lambda \subseteq GL(V_\lambda ) \) in the case (NSD) and \( H_\lambda \subseteq GO(V_\lambda )\) resp. \( H_\lambda \subseteq GSp(V_\lambda ) \) for even resp. odd \(\varepsilon (X_\lambda )\) in the (SD)-cases. If \(X_\lambda \) is properly self dual in the sense \(X_\lambda ^\vee \cong X_\lambda \), these simlitude groups can be replaced by their subgroups \(O(V_\lambda )\) resp. \(Sp(V_\lambda )\).

6.2 The structure of \(G_{\lambda }\)

Recall that two maximal atypical weights \(\lambda , \ \mu \) are equivalent \(\lambda \sim \mu \) if there exists \(r\in {\mathbb {Z}}\) such that \(L(\lambda ) \cong Ber^r \otimes L(\mu )\) or \(L(\lambda )^\vee \cong Ber^r \otimes L(\mu )\) holds. Another way to express this is to consider the restriction of the representations \(L(\lambda )\) and \(L(\mu )\) to the Lie superalgebra \(\mathfrak {sl}(n|n)\). These restrictions remain irreducible and \(\lambda \sim \mu \) holds if and only if \(L(\lambda ) \cong L(\mu )\) or \(L(\lambda )\cong L(\mu )^\vee \) as representations of \(\mathfrak {sl}(n|n)\). Let \(X^+(n)\) be the set of dominant weights and let \(Y^+(n)\) be the set of equivalence classes of dominant weights. Similarly let \(X^+_0(n)\) denote the class of maximal atypical dominant weights and \(Y^+_0(n)\) the set of corresponding equivalence classes. If we write \(\lambda \in Y^+_0(n)\), we mean that \(\lambda \in X^+_0(n)\) is some representative of the class in \(Y^+_0(n)\) defined by \(\lambda \).

Let \(\lambda \) be of SD-type. Then there exists \(r\in {\mathbb {Z}}\) such that \(L(\lambda ) \cong Ber^r \otimes L(\lambda )^\vee \). Hence there exists an equivariant nondegenerate pairing

$$\begin{aligned} X_\lambda \times X_\lambda \longrightarrow B^r. \end{aligned}$$

This pairing is either symmetric (even) or antisymmetric (odd). We then say \(X_{\lambda }\) is even and put \(\varepsilon (X_\lambda )=1\), or odd and \( \varepsilon (X_\lambda )=-1\). The next lemma is proven in “Appendix C”.

Lemma 6.1

For all irreducible objects \(X_\lambda \) of SD-type in \(\mathcal {T}_n^+\) we have

$$\begin{aligned} \varepsilon (X_\lambda ) = \varepsilon (X_{\lambda _{basic}}) = (-1)^{p(\lambda _{basic})}. \end{aligned}$$

Theorem 6.2

\(G_\lambda = SL(V_\lambda )\) if \(X_{\lambda }\) is (NSD). If \(X_{\lambda }\) is (SD) and \(V_{\lambda }|_{G_{\lambda '}}\) is irreducible, then \(G_\lambda = SO(V_\lambda )\) respectively \(G_\lambda = Sp(V_\lambda )\) according to whether \(X_\lambda \) is even respectively odd. If \(X_{\lambda }\) is (SD) and \(V_{\lambda }|_{G_{\lambda '}}\) is not irreducible, then \(G_\lambda \cong SL(W)\) for \(V_{\lambda }|_{G_{\lambda '}} \cong W \oplus W^{\vee }\).

This theorem is proven in Sects. 710. Many examples can be found in Sect. 9. We conjecture that a stronger version is true: \(V_{\lambda }\) should always stay irreducible. We refer to Sect. 11 for a discussion of this case.

Remark 6.3

The (NSD) case is the generic case for \(n \ge 4\). Since \(SL(V_{\lambda }) \cong G_{\lambda } \subset GL(V_{\lambda })\), all representations of \(H_{\lambda }\) stay irreducible upon restriction to \(G_{\lambda }\). Hence the derived group sees already the entire tensor product decomposition into indecomposable representations up to superdimension zero. The same remark is true for a selfdual weight of symplectic type. In the orthogonal case we could have a decomposition of an irreducible representation of \(H_{\lambda }\) into two irreducible representations of \(G_{\lambda }\) since \(O(V_{\lambda })\) and \(GO(V_{\lambda })\) have two connected components.

Example 6.4

The smallest case for which \(V_{\lambda }\) could decompose when restricted to \(G_{\lambda }\) is the case \([\lambda ] = [3,2,1,0] \in \mathcal {T}_4^+\) with sector structure

figure e

Then DS[3, 2, 1, 0] decomposes into four irreducible representations

$$\begin{aligned} L_1 = [3,2,1], L_2 = [3,2,-1], L_3 = [3,0,-1], L_4 = [1,0,-1 ]. \end{aligned}$$

Since \(L_1 = Ber^{2}L_4\) and \(L_2 \cong L_3^{\vee }\) we have two equivalence classes

$$\begin{aligned}\{ L_1,L_4 \}, \{ L_2,L_3 \}.\end{aligned}$$

In fact \( V_{\lambda _1} \cong V_{\lambda _4} \cong st(SO(6))\) and \( V_{\lambda _2} \cong st(SL(6)), V_{\lambda _3} \cong st(SL(6))^{\vee }\).

If \(V_{[3,2,1,0]}\) does not decompose under restriction to \(G_{[3,2,1,0]}\), then \(G_{\lambda } \cong SO(24)\) and \(V_{\lambda } \cong st(SO(24))\). If it decomposes \(V_{\lambda } = W \oplus W^{\vee }\), then \(G_{\lambda } \cong SL(12)\) and \(W \cong st(SL(12))\). Since \(W \not \sim W^{\vee }\) this implies that the embedding \(SO(6) \times SL(6) \rightarrow SL(12)\) gives the branching rules

$$\begin{aligned} W&\mapsto st(SL(6)) \oplus st(SO(6)) \\ W^{\vee }&\mapsto st(SL(6))^{\vee } \oplus st(SO(6)). \end{aligned}$$

6.3 The structure theorem on \(G_n\)

We now determine \(G_n\).

Lemma 6.5

Suppose a tannakian category \({{\mathcal {R}}}\) with Tannaka group H is \(\otimes \)-generated as a tannakian category by the union of two subsets \(V'\) and \(V''\). Let \(H'\) and \(H''\) be the Tannaka groups of the tannakian subcategories generated by \(V'\) respectively \(V''\). Then there exists an embedding \(H \hookrightarrow H' \times H''\) so that the composition with the projections is surjective.

Proof

For arbitrary \(V', \ V''\) (not necessarily finite) we get \(\mathcal {T}(V') \hookrightarrow T(V' \cup V'')\), \(\mathcal {T}(V'') \hookrightarrow \mathcal {T}(V \cup V")\) for the tensor categories \(\mathcal {T}\) generated by \(V'\), V", \(V' \cup V"\) respectively. This gives natural epimorphisms \(\pi ': H \rightarrow H'\) and \(\pi '': H\rightarrow H''\) which induce a morphism \(i: H \rightarrow H'\times H''\) so that the composition with the projections are \(\pi '\) and \(\pi ''\). It remains to show that i is injective. The morphism i is injective since \(g \in H\) is trivial if it acts trivially on all generators in \(V' \cup V''\) of \(\mathcal {T}(V' \cup V'')\). \(\square \)

We remark that the inclusion \(H \hookrightarrow H' \times H''\) induces an inclusion \(H^0 \hookrightarrow (H')^0 \times (H'')^0\) of the Zariski connected components and hence an inclusion of the corresponding adjoint groups \(H^0_{ad}:= (H^0)_{ad}\)

$$\begin{aligned} H^0_{ad} \hookrightarrow (H')^0_{ad} \times (H'')^0_{ad} \, \end{aligned}$$

and, abbreviating \( H^0_{der} \hookrightarrow (H')^0_{der} \times (H'')^0_{der}\), similarly for the derived groups \(G:=H^0_{der}:= (H^0)_{der}\)

$$\begin{aligned} G \hookrightarrow G' \times G''. \end{aligned}$$

We also need the following variant of Goursat’s lemma.

Lemma 6.6

Suppose H is a connected reductive subgroup of the product \(A\times B\) of two semisimple affine algebraic k-groups A and B, so that the projections to A and B are surjective. Then

  1. (1)

    If A and B are connected simple k-groups, then either \(H_{ad}= A_{ad}\times B_{ad}\) or \(H_{ad}\cong A_{ad} \cong B_{ad}\).

  2. (2)

    \(H\cong A\times B\), if A and B are of adjoint type without common factor.

  3. (3)

    If A and B are connected, \(H \cong A \times B\) if and only if \(H_{ad} \cong A_{ad} \times B_{ad}\).

  4. (4)

    Suppose A is a connected semisimple group and B is a connected simple group. Let H be a proper subgroup H of \(A\times B\), that surjects onto A and B for the projections. Then there exists a simple normal subgroup C of A, such that the image H/C of H in \((A/C) \times B\) is a proper subgroup of \((A/C) \times B\), if A is not a simple group.

Proof

(1)–(3) are obvious. Part (4) can be reduced to the case of adjoint groups by part (3). So we may assume that B and A are groups of adjoint type. We now use the following fact. Any semisimple A group of adjoint type is isomorphic to the product \(\prod _{i=1}^r A_i\) of its simple subgroups \(A_i\). Its factors are the normal simple subgroups of A. These factors and hence this product decomposition is unique up to a permutation of the factors. Any nontrivial algebraic homomorphism of A to a simple group B is obtained as projection of A onto some factor \(A_i\) of the product decomposition composed with an injective homomorphism \(A_i \rightarrow B\). Since \(H \subseteq A \times B\) projects onto the first factor A and B is simple, and since H is a proper subgroup of the connected semisimple group \(A\times B\), the kernel of the projection \(p_A: H\rightarrow A\) is a finite normal and hence central subgroup of H. It injects into the center of B, hence is trivial. Thus \(p_A:H\rightarrow A\) is an isomorphism so that H defines the graph of a group homomorphism \(A \rightarrow B\). Since A is of adjoint type and therefore a product of simple groups \(A \cong \prod _{i=1}^r A_i\), the kernel of the homomorphism \(A\rightarrow B\) must be of the form \(\prod _{i\ne j} A_i\). Unless A is simple, for \(C=A_i\) and any \(i\ne j\) assertion (4) becomes obvious. \(\square \)

Corollary 6.7

Let \(\lambda \) and \(\mu \) be two maximal atypical weights and denote by \(G_{\lambda ,\mu }\) the connected derived group of the Tannaka group \(H_{\lambda ,\mu }\) corresponding to the subcategory in \(\overline{\mathcal {T}}_n\) generated by \(L(\lambda )\) and \(L(\mu )\). If \(\lambda \) is not equivalent to \(\mu \),

$$\begin{aligned} G_{\lambda ,\mu } \cong G_{\lambda } \times G_{\mu }.\end{aligned}$$

Proof

If \(G_{\lambda }\) and \(G_{\mu }\) are not isomorphic, Lemma 6.6 implies the claim. Otherwise \(G_{\lambda ,\mu } \cong G_{\lambda } \cong G_{\mu }\) (special case of Lemma 6.6.1). We assume by induction that the statement holds for smaller n. By Theorem B.3 there exists either \(L(\lambda _i)\) which is not equivalent to any \(L(\mu _j)\) or there exists \(L(\mu _j)\) which is not equivalent to any \(L(\lambda _i)\)—a contradiction since then the branching of \(V_{\lambda }\) and \(V_{\mu }\) to \(G_{n-1}\) would not be the same. For \(n=2\) we give an adhoc argument in Sect. 9. \(\square \)

Theorem 6.8

Structure Theorem for \(G_n\). The connected derived group \(G_n\) of the Tannaka group \(H_n\) of the category \(\mathcal {T}_n^+\) is isomorphic to the product

$$\begin{aligned} G_n \ \cong \ \prod _{\lambda \in Y^+_0(n)} G_\lambda . \end{aligned}$$

Proof

This follows essentially from Theorem 6.2, where the structure of the individual groups \(G_\lambda \) was determined. Using Lemma 6.6, one reduces the statement of the theorem to a situation that involves only two inequivalent weights \(\lambda \) and \(\mu \): By part (3) of Lemma 6.6 we may replace the derived groups by the adjoint groups. Then the assertion follows from part (4) of the lemma by induction on the number of factors reducing the assertion to the case of two groups \(G_{\lambda }\), \(G_{\mu }\) dealt with in Corollary 6.7. \(\square \)

Example 6.9

Consider the tensor product of two inequivalent representations \(L(\lambda )\) and \(L(\mu )\) of non-vanishing superdimension. Then

$$\begin{aligned} L(\lambda ) \otimes L(\mu ) = I \ \mod \ {{\mathcal {N}}}\end{aligned}$$

for an indecomposable representation I. Indeed \(L(\lambda )\) and \(L(\mu )\) correspond to representations of the derived connected Tannaka groups \(G_{\lambda }\) and \(G_{\mu }\). Since \(G_{\lambda }\) and \(G_{\mu }\) are disjoint groups in \(G_n\), tensoring with \(L(\lambda )\) and \(L(\mu )\) corresponds to taking the external tensor product of these representations.

7 Proof of the structure theorem: overview

We now determine \(G_{\lambda }\) inductively using the k-linear exact tensor functor between the quotient categories of the representation categories

$$\begin{aligned} \eta : \overline{\mathcal {T}}_{n} \rightarrow \overline{\mathcal {T}}_{n-1} \end{aligned}$$

constructed in Lemma 5.11 with the help of \(DS: \mathcal {T}_n^+ \rightarrow \mathcal {T}_{n-1}^+\). By the main theorem on DS (Theorem 4.1), the restriction of \(V_{\lambda } = \omega (X_{\lambda })\) to the subgroup \(H_{n-1}\) is a multiplicity free representation. We assume by induction that Theorems 6.2 and 6.8 hold for \(H_{n-1}\) and \(G_{n-1}\).

We have inclusions

$$\begin{aligned} G_{\lambda '} \hookrightarrow G_\lambda \hookrightarrow H_\lambda ^0 \hookrightarrow H_\lambda \end{aligned}$$

where \(G_{\lambda '}\) denotes the image of the natural map \((H_{n-1}^0)_{der} \rightarrow G_\lambda = (H_\lambda ^0)_{der}\). The restriction of \(V_\lambda \) to \(G_{\lambda '}\) decomposes

$$\begin{aligned} V_\lambda \ \cong \ \bigoplus _{i=1}^k \ V_{\lambda _i} \end{aligned}$$

where the \(V_{\lambda _i}\) are the irreducible representations in the category \(\overline{\mathcal {T}}_{n-1}^+\) corresponding to the irreducible constituents \(L(\lambda _i), i=1,\ldots ,k\), of \(DS(L(\lambda ))\). By induction we obtain

$$\begin{aligned} G_{\lambda '} \cong \prod _{\lambda _i /\sim } G_{\lambda _i} \end{aligned}$$

where the \(G_{\lambda _i}\) are described in Theorem 6.2.

In a first step we discuss the situation in the \(n=2\) and the \(n=3\) case as well as the Tannaka groups \(G_{\lambda }\) for \(L(\lambda ) =Ber^r \otimes [i,0,\ldots ,0]\), \(r,i \in {\mathbb {Z}}\). The \(n=2\)-case is needed for the start of the inductive determination of \(G_n\). In this case we can use the known tensor product decomposition between irreducible modules in \(\mathcal {T}_2\) to determine \(G_2\) and \(H_2\). In order to get a clear induction scheme in the proof of the structure theorem, we need to rule out certain exceptional cases which can only occur for \(n \le 3\) and for the modules \(Ber^r \otimes [i,0,\ldots ,0]\). This will allow us to assume \(n \ge 4\) in Sect. 10.

In the next step we show that \(G_{\lambda }\) is simple. By induction all the \(V_{\lambda _i}\) are standard representations for simple groups of type ABCD or \(V_{\lambda _i}|_{G_{\lambda _i}} = W \oplus W^{\vee }\) for \(G_{\lambda _i} \cong SL(W)\). The representation \(V_{\lambda }\) decomposes under restriction to \(G_{\lambda }\) in the form \(W_1 \oplus \ldots \oplus W_s\) (we later show that s is at most 2). If we restrict these \(W_{\nu }\) to \(G_{\lambda '}\), they are meager representation of \(G_{\lambda '}\) in the sense of Definition 10.2. The crucial Lemma 10.3 shows then that \(G_{\lambda }\) is simple. This allows us to use the classification of small representations due to Andreev–Elashvili–Vinberg.

Our aim is then to show that the dimension of the subgroup \(G_{\lambda '}\) is large compared to the dimension of \(V_{\lambda }\) (given by the superdimension formula for \(L(\lambda )\) in [36]) as in Lemma 8.1 or Corollary 8.2. A large rank and a large dimension of \(G_{\lambda '}\) implies that the rank and the dimension of \(G_{\lambda }\) must be large, forcing \(V_{\lambda }\) to be a small representation of \(G_{\lambda }\) in the sense of Lemma 8.1 and Corollary 8.2. If we additionally know that \(G_{\lambda }\) is simple and that also \(r(G_{\lambda }) \ge \frac{1}{2}(\dim (V_{\lambda })-1)\), Corollary 8.2 will immediately imply that \(G_{\lambda }\) is of type \(SL(V_{\lambda }), \ SO(V_{\lambda })\) or \(Sp(V_{\lambda })\). However the strong rank estimate will not always hold and we will be in the less restrictive situation of Lemma 8.1.

Here the (NSD) and the (SD) case differ considerably. In the (NSD) case each irreducible representation \(V_{\lambda _i}\) (corresponding to \(L(\lambda _i)\) in \(DS(L(\lambda ))\)) gives a distinct direct factor in the product \(G_{\lambda '} \cong \prod _{\lambda _i /\sim } G_{\lambda _i}\) since all irreducible representations of \(DS(L(\lambda ))\) are inequivalent in the (NSD) case by Lemma A.4. The dimension estimate for \(G_{\lambda }\) so obtained then implies that \(V_{\lambda }\) is a small representation. In the (SD) case however two representations \(V_{\lambda _i}, V_{\lambda _j}\) will contribute the same direct factor \(G_{\lambda _i} \simeq G_{\lambda _j}\) if \(\lambda _i \sim \lambda _j\). This decreases the dimension and rank estimate of the subgroup \(G_{\lambda '}\) in \(G_{\lambda }\) and therefore of \(G_{\lambda }\).

To finish the proof we need to understand the restriction of \(V_{\lambda }\) to \(G_{\lambda }\). The group of connected components acts transitively on the irreducible constituents \(V_{\lambda } = W_1 \oplus \cdots \oplus W_s\) of the restriction to \(H_{\lambda }^0\) and \(G_{\lambda }\). Using that the decomposition of \(V_{\lambda }\) to \(H_{n-1}\) is multiplicity free in a weak sense (obtained from an analysis of the derivatives of \(L(\lambda )\) in “Appendix A”), we show finally in Sect. 10.3, using Clifford–Mackey theory, that \(V_{\lambda }\) can decompose into at most \(s=2\) irreducible representations of \(G_{\lambda }\).

8 Small representations

Our aim is to understand the Tannaka groups associated to an irreducible representation by means of the restriction functor \(DS: \mathcal {T}_n^+ \rightarrow \mathcal {T}_{n-1}^+\). We have a formula for the superdimension of an irreducible representation [36] and we know inductively the ranks and dimensions of the groups arising for \(k <n\). This gives strong restrictions about the groups in the \(\mathcal {T}_n^+\)-case due to the following list of small representations.

List of small representations. For a simple connected algebraic group H and a nontrivial irreducible representation V of H the following holds [5]

Lemma 8.1

\(\dim (V) = \dim (H)\) implies that V is isomorphic to the adjoint representation of H. Furthermore, except for a finite number of exceptional cases, \(\dim (V) < \dim (H)\) implies that V belongs to the regular cases

R.1:

\(V\cong st, S^2(st), \Lambda ^2(st)\) or their duals in the \(A_r\)-case,

R.2:

\(V=st\) (the standard representation) in the \(B_r,D_r\)-case,

R.3:

\(V\cong st\) in the \(C_r\)-case,

R.4:

\(V \hookrightarrow \Lambda ^2(st)\) in the \(C_r\)-case

where the list of exceptional cases is

E.1:

\(\dim (V)=20,35,56\) for \(V=\Lambda ^3(st)\) and \(A_r\) in the cases \(r=5,6,7\).

E.2:

\(\dim (V)=4,8,16,32,64\) for the spin representations of \(B_r\) in the cases \(r=2,3,4,5,6\).

E.3:

\(\dim (V)=8,8,16,16,32,32,64,64\) for the two spin representations of \(D_r\) in the cases \(r=4,5,6,7\).

E.4:

\(\dim (V)=27,27\) for \(E_6\) with \(\dim (E_6) =78\) (standard representation and its dual).

E.5:

\(\dim (V)=56\) for \(E_7\) with \(\dim (E_7)=133\).

E.6:

\(\dim (V)=7\) for \(G_2\) with \(\dim (G_2)=14\).

E.7:

\(\dim (V)=26\) for \(F_4\) with \(\dim (F_4)=52\).

In particular \(\dim (V) \ge r+2\) holds, except for \(G=A_r\) in the cases \(V\cong st\) or \(V\cong st^\vee \).

Corollary 8.2

Let V be an irreducible representation of a simple connected group H such that \(4 \le \dim (V) < \dim (H)\) and

$$\begin{aligned}2r(H) \ge \dim (V) -1\end{aligned}$$

holds. Then H is of type \(A_r, B_r, C_r,D_r\) and \(V=st\) the standard representation of this group of dimension \(r+1,\ 2r+1, \ 2r, \ 2r\) for \(r\ge 3,\ 2,\ 2,\ 2\) respectively, or \(H=D_4\) and V is one of the two 8-dimensional spin representations.

From the classification in Lemma 8.1 one also obtains

Lemma 8.3

For a simple connected grous H with an irreducible root system of rank r we have \(\dim (H) \ge r(2r-1)\) except for \(H\cong SL(n)\) with \(\dim (H) = r(r+2)\). Furthermore \(r\le \dim (V)\) holds for any nontrivial irreducible representation V of H.

9 The cases \(n=2, 3\) and the \(S^i\)-case

In the next sections we determine the group \(G_n\) and the groups \(G_{\lambda }\). Since we will determine these groups inductively starting from \(n=2\), we need to start with this case. We also discuss the \(n=3\) case separately since we have to rule out some exceptional low rank examples in the classification of [5] in Sect. 8.

Warm-up. Suppose \(n=1\). Then \(H_1\) is the multiplicative group \({\mathbb {G}}_m\). Indeed the irreducible representations of it correspond to the irreducible modules \(\Pi ^i Ber^i\) for \(i\in {\mathbb {Z}}\).

9.1 The case \(n=2\)

For \(S^i = L([i,0])\) and \(i\ge 1\) let denote

$$\begin{aligned} X_i:=\Pi ^i([i,0]). \end{aligned}$$

Then \(X_i^\vee \cong B^{1-i} \otimes X_i\), hence \(X_1^{\vee } \cong X_1\). We use from [37] the fusion rule

$$\begin{aligned}{}[i,0] \otimes [j,0] \ = \ \text {indecomposable}\ \oplus \ \delta _i^j \cdot Ber^{i-1} \oplus \text {negligible} \end{aligned}$$

for \(1\le i \le j\) together with \(Ber^r \otimes [i,0] \cong [r+i,r]\) for all \(r\in {\mathbb {Z}}\).

Lemma 9.1

If \(H_{X_i}\) denotes the Tannaka group of \(X_i\), then

$$\begin{aligned} H_{X_i} \simeq {\left\{ \begin{array}{ll} SL(2) &{}\quad i = 1\\ GL(2) &{}\quad i \ge 2.\end{array}\right. } \end{aligned}$$

Proof

Since \(H_1 \hookrightarrow H_2\twoheadrightarrow H_{X_i} \) can be computed from DS we see that \(H_1\) injects into \(H = H_{X_i}\) and the two dimensional irreducible representation \(V = V_{X_i}\) of \(H_{X_i}\) attached to \(X_i\) decomposes into

$$\begin{aligned} V\vert _{H_1} \ = \ det^{-1} \oplus det^i. \end{aligned}$$

corresponding to \(DS(X_i) \ = \ Ber^{-1} \oplus Ber^i\). If \(H^0_{X_i} \cong {\mathbb {G}}_m\), the finite group \(\pi _0(H)\) acts on \(H^0\). By Mackey’s theorem the stabilizer of the character \(Ber^{-1}\) has index two in \(H_{X_i}\) and acts by a character on V. Since the only automorphisms of \({\mathbb {G}}_m\) are the identity and the inversion, this would imply \(i=1\). Hence \(V\otimes V\) would restrict to \({\mathbb {G}}_m\) with at least three irreducible constituents \(det^{-2} \oplus det^2\) (corresponding to \(Ber^{-2} \oplus Ber^2\)) and a two dimensional module W with an action of \(\pi _0(H)\) such that a subgroup of index two acts by a character. But \(X^\vee _1 \cong X_1\) implies that V is self dual, and hence W contains the trivial representation. This contradicts the fusion rule from above. Hence \(H^0 \ne {\mathbb {G}}_m\) and the same argument as above shows that \(H^0\) can not be a torus. Hence the rank r of each irreducible component of the Dynkin diagram of \((H^0_{der})_{sc}\) is \(r\ge 1\) and hence \(\dim (H)\ge 3\). By Lemma 8.3 we know \(r\le \dim (V)=2\) and accordingly \(\dim (H)=3\) by Lemma 8.1. Therefore \((H^0_{der})=SL(2)\) and \(V\vert _{H_{der}^0}\) is the irreducible standard representation. Since H acts faithful on V

$$\begin{aligned} SL(2) \subseteq H \subseteq GL(2). \end{aligned}$$

Now we use \(V^\vee \cong Ber^{i-1} \otimes V\), which implies \(H = GL(2)\) for \(i>1\). Indeed \(\Lambda ^2(V)\) is the character \(Ber^{i-1}\) by the fusion rules above. For \(i=1\) the isomorphism \(V^\vee \cong V\) implies that det(V) is trivial on H, hence

$$\begin{aligned} H=SL(2) \end{aligned}$$

in the case \(i=1\). \(\square \)

9.2 The \(H_2\)-case

We discuss the Tannaka group generated by all irreducible representations. First consider the Tannaka group H of \(\langle X_i, X_j \rangle _\otimes \) for some pair \(j > i\). The derived groups of the Tannaka groups \(H'\) resp. \(H''\) of \(\langle X_i \rangle _\otimes \) and \(\langle X_j \rangle _\otimes \) are SL(2).

We claim that \(H_{der} \cong H'_{der} \times H''_{der}\). If this were not the case, then \(H_{der} \cong SL(2)\) (special case of Lemma 6.6.1). But then the tensor product \(X_i \otimes X_j\) considered as a representation of H corresponds to the tensor product of two standard representation and hence is a reducible representation with two irreducible factors. However this contradicts the fusion rules stated above. This implies \(H_{der} \cong SL(2) \times SL(2)\) and hence \(H_{ad} \cong H'_{ad} \times H''_{ad}\).

Now consider the Tannaka group H of \(\langle X_{i_1},\ldots ,X_{i_k}\rangle _\otimes \) for \(k > 2\). We claim that H is connected and that it is the product

$$\begin{aligned} H_{der} \ \cong \ \prod _{\nu =1}^k H_{der}(X_{i_\nu }) \end{aligned}$$

of the derived Tannaka groups of the \(\langle X_{i_\nu }\rangle _\otimes \). This is an immediate consequence of Lemma 6.6

The fusion rule \(S^i \otimes S^i \cong Ber^{i-1} \oplus \, \text {indecomposable} \, \oplus \, \text {negligible}\) implies \(\Lambda ^2(X_i) \cong B^{i-1} \oplus \text {negligible}\). In particular the image of \(B^{i-1}\) is contained in \(Rep(H_{X_i})\) and generates a subgroup of form \(\mathbb {G}_m\). So the Tannaka group \(H_2\) of the category \(\mathcal {T}_2^+/\mathcal {N}\) sits in an exact sequence

$$\begin{aligned} 0 \rightarrow \lim _{k} \prod _{\nu =0}^{k-1} SL(2) \rightarrow H_2 \rightarrow {\mathbb {G}}_m \rightarrow 0. \end{aligned}$$

The derived group of \(H_2\) is the projective limit of groups SL(2) with a copy for each irreducible object \(X_{\nu +1}\) for \(\nu =0,1,2,3,\ldots \). The structure of the extension is now easily recovered from the following decription:

Lemma 9.2

\( H_2 \ \subset \ \prod _{\nu =0}^\infty GL(2) \) is the subgroup defined by all elements \(g=\prod _{\nu =0}^\infty g_\nu \) in the product with the property \(det(g_\nu ) = det(g_1)^\nu \). The automorphism \(\tau _2\) is inner.

We usually write \(GL(2)_\nu \) for the \(\nu \)-th factor of the product \(\prod _{\nu =0}^\infty GL(2) \). Using the description of the last lemma, the torus \(H_1\cong {\mathbb {G}}_m\) embeds into \(H_2\) as follows

$$\begin{aligned} H_1 \ni t \mapsto \prod _{\nu =0}^\infty diag(t^{\nu +1},t^{-1}) \in H_2 \subset \prod _{\nu =0}^\infty GL(2)_\nu . \end{aligned}$$

Defining \(det(g)=det(g_1)\) for \(g=\prod _{\nu =0}^\infty g_\nu \) in \(H_2\), the representation of the quotient group \({\mathbb {G}}_m\) of \(H_2\) defined by the Berezin determinant \(Ber \in \mathcal {T}_2\), corresponds to the character det(g) of the group \(H_2\).

We continue with two special cases: The \(S^i\)-case for any n, and the case \(G_3\).

9.3 The \(S^i\)-case

Consider the modules \(X_i = \Pi ^i([i,0,0])\) in \(\mathcal {T}_3^+\). They have super dimension 3 for \(i\ge 2\). Let H (or sometimes \(H_{X_i}\)) denote the associated Tannaka group and V the associated irreducible representation of H.

Lemma 9.3

We have \(H_{X_1} = SL(2)\) and \(G_{X_i} \simeq SL(3)\) for any \(i \ge 2\) and \(H_{X_i} \simeq GL(3)\) for any \(i \ge 3\).

Proof

The natural map \(H_2 \rightarrow H_3 \rightarrow H\) allows to consider V as a representation of \(H_2\), and as such we get

$$\begin{aligned} V\vert _{H_2} \ \cong \ {\det }^{-1}\ \oplus \ X_i \ \end{aligned}$$

for \(i\ge 2\) (here \(X_i\) on the right is the irreducible 2-dimensional standard representation of \(GL(2)_{i-1}\), restricted to \(H_2\)). Hence \(\dim (A) \ge 3\) for at least one simple factor A of \(H^0\) and every irreducible summand W of \(V\vert _A\) has dimension \(\le \dim (A)\). By Lemma 8.1 therefore W either has dimension 3 and \(A_{sc}=SL(3)\), \(W=st\) or \(W=st^\vee \), or \(A_{sc} =SL(2)\) and \(W=S^2(st)\). If \(H^0_{der}\) is not simple, we replace it by its simply connected cover and write \((H^0_{der})_{sc} = A_{sc} \times A'\) (where \(A'\) is a product of simple groups). The representation V is then an external tensor product

$$\begin{aligned} V = W \boxtimes W' \end{aligned}$$

of irreducible representations \(W, W'\) of \(A_{sc}\) and \(A'\). Since V is a faithful representation of H, the lift of V (again denoted V) to \((H^0_{der})_{sc}\) has finite kernel. Since it has finite kernel, \(\dim (W)>1, \ \dim (W') > 1\) holds. Hence \(\dim (W) = 3\) implies \((H^0_{der})=A\) and \(V\vert _{H^0}\) and \(V\vert _{H^0_{der}}\) remain irreducible by dimension reasons. If \(A_{sc} =SL(2)\) and \(W=S^2(st)\), the image of \(H_2\) surjects onto \(H_{der}\). This contradicts the fact that V is irreducible but \(V\vert _{H_2}\) decomposes, and excludes the case \(A_{sc}=SL(2)\). Hence

$$\begin{aligned} H^0_{der} \ \cong \ SL(3). \end{aligned}$$

Since H acts faithfully on V, we also have \( H \subseteq GL(V)=GL(3)\). The restriction of V to \(H_2\) has determinant \(det^{-1} \cdot \det (X_i) \cong det^{-1}det^{i-1} = det^{i-2}\). Hence

$$\begin{aligned} H \ \cong \ GL(3) \end{aligned}$$

for all \(i\ge 3\). \(\square \)

For \(j > i \ge 2\) let H denote the Tannaka group of \(\langle X_i, X_j \rangle _\otimes \) and \(H',H''\) the connected components of the Tannaka groups of \(\langle X_i \rangle _\otimes \) resp. \(\langle X_j \rangle _\otimes \). Then we claim

$$\begin{aligned} H^0_{der} \ \cong \ H'_{der} \times H''_{der} \, \end{aligned}$$

since otherwise \(H'_{der} \cong H''_{der}\) by Lemma 6.6.1. But this is impossible since then the morphisms \(H_2 \rightarrow H_3 \rightarrow H\) would induce the same morphisms \((H_2)_{der} \rightarrow H_{der} \rightarrow H'_{der}\) and \((H_2)_{der} \rightarrow H_{der} \rightarrow H''_{der}\), which contradicts Theorem 4.1. Indeed the factor \(SL(2)_{i-1}\) maps nontrivially to \(H'_{der}\) but trivially to \(H''_{der}\). Since H acts faithfully on the representation associated to the object \(X_i \oplus X_j\) on the other hand \(H \subseteq GL(\omega (X_i)) \times GL(\omega (X_j))\).

The same arguments enable us to determine the connected derived groups for any \(n \ge 3\):

Lemma 9.4

The Tannaka group H of the modules \(X_i=\Pi ^i([i,0,\ldots ,0])\) in \(\mathcal {T}_n^+\) satisfies \(H^0_{der} \cong SL(n)\) and \(H \subseteq GL(n)\) for all \(i\ge n-1\), and \(H=GL(n)\) for all \(i\ge n\). For \(i < n-1\) we get \(H^0_{der} \cong SL({{\,\textrm{sdim}\,}}(X_i))\).

Proof

Indeed we have in \(H^0_{der}\) a simple component A of semisimple rank \(r \ge n-1\) by induction. Obviously A contains \(SL(n-1)\) and cannot be of Dynkin type \(A_r\) unless \(A=SL(n)\) by Lemma 8.1.

Notice that \(\dim (A) \ge r(2r-1) \ge (n-1)(2n-3) > n\) or \(\dim (A) \ge r(r+2) \ge (n-1)(2n) > n\), for \(n\ge 3\) by Lemma 8.3. The restriction of V decomposes into irreducible summands \(W,W',\ldots \) of dimension \(\dim (W)\le n\), and the dimension of all these representations is \(\le r\). So the possible representations are listed in Lemma 8.1. None of them has dimension \(\le r+1\) except for the case where A is of type \(A_r\) and \(V\cong st\) or \(V\cong st^\vee \). \(\square \)

9.4 The \(n=3\)-case

We analyse the remaining \(n=3\)-cases.

Lemma 9.5

The derived connected group \(G_3=(H_3)^0_{der}\) of \(H_3\) is

$$\begin{aligned} G_3 \ \cong \ \prod _{\lambda } G_\lambda \, \end{aligned}$$

where \(\lambda \) runs over all \(\lambda =[\lambda _1,\lambda _2,0]\) with integers \(\lambda _1,\lambda _2\) such that

$$\begin{aligned}0\le 2\lambda _2 \le \lambda _1 \ \end{aligned}$$

and \(G_\lambda \cong 1, SL(2),SL(3),Sp(6),SL(6)\) according to whether \(\lambda \) is 0, [1, 0, 0] or \([2+\nu ,0,0]\), for \(\nu \ge 0\), or \(\lambda = [2\lambda _2,\lambda _2,0]\), for \(\lambda _2> 0\), or \(0<2 \lambda _2 < \lambda _1\).

Remark 9.6

We discuss the general case in the next section assuming \(n \ge 4\). The assumption \(n \ge 4\) is only relevant because we want to have a uniform behaviour regarding derivatives. Essentially all the arguments regarding simplicity of \(G_{\lambda }\) and Clifford–Mackey theory apply to the \(n=3\) case at hand. In the proof we discuss [2, 1, 0] in detail and sketch the key inputs for the other cases.

Proof

Let us consider \(X=X_\lambda \) for \(L(\lambda )=[210]\). The associated irreducible representation Tannaka group \(H = H_X\) admits an alternating pairing, hence \(H_X\) is contained in the symplectic group of this pairing

$$\begin{aligned} H_X \ \subseteq \ Sp(6). \end{aligned}$$

We claim that \(H^0_{der}\) is simple. If not, we replace it by its simply connected cover and write it as a product

$$\begin{aligned} \big (H^0_{der}\big )_{sc} = G_1 \times G_2.\end{aligned}$$

The faithful representation \(V_X\) of \(H_X\) has finite kernel when seen lifted to a representation of \((H^0_{der})_{sc}\). Therefore \(V_{\lambda }\) as a representation of \((H^0_{der})_{sc}\) is of the form \(V_1 \boxtimes V_2\) with \(\dim (V_i) > 1\). The representation \(V_{\lambda }\) restricts to the subgroup \(SL(2) \times SL(2) = G_{\lambda '}\) as

$$\begin{aligned} V_{\lambda }|_{G_{\lambda '}} \cong 2 \cdot (st \boxtimes {\textbf{1}}) \oplus ({\textbf{1}}\boxtimes st).\end{aligned}$$

This is easily seen using

$$\begin{aligned} DS(\Pi [2,1,0]) \cong \Pi [2,1] \oplus \Pi [2,-1] \oplus \Pi [0,-1].\end{aligned}$$

Since \(\Pi [2,1] \cong Ber^{-2} \otimes [0,-1]\) they both give a copy of the standard representation of the same SL(2). Hence the restriction of \(V_{\lambda }\) to the first SL(2)-factor is of the form

$$\begin{aligned} V_{\lambda }|_{SL(2)} \cong 2 \cdot st\oplus 2 \cdot {\textbf{1}}\end{aligned}$$

and

$$\begin{aligned} V_{\lambda }|_{SL(2)} \cong st\oplus 4 \cdot {\textbf{1}}\end{aligned}$$

for the second SL(2)-factor. Now consider the restriction to any of the two SL(2)-factors

$$\begin{aligned} V\vert _{SL(2)} = V_1\vert _{SL(2)} \otimes V_2\vert _{SL(2)}. \end{aligned}$$

Since \(dim(V_1) =2\) and \(dim(V_2) = 3\), their restriction to SL(2) is either st or \(2 \cdot {\textbf{1}}\) for \(V_1\) and \(st \oplus {\textbf{1}}\) or \(3 \cdot {\textbf{1}}\) for \(V_2\). The Clebsch–Gordan rule for SL(2) shows that \(V\vert _{SL(2)} = V_1\vert _{SL(2)} \otimes V_2\vert _{SL(2)}\) is not possible, hence \(H^0_{der}\) must be simple. The image of \(H_2\) in H contains two copies of SL(2). Since \(H^0_{der}\) is not \(SL(2) \times SL(2)\), we get \(\dim (H^0_{der}) \ge 7\) and the representation V is small. Since \(V_{\lambda }\) restricted to the subgroup \(SL(2) \times SL(2)\) has 3 summands of dimension 2 each, the restriction to \(H^0_{der}\) can decompose into at most 3 summands: either \(V_{\lambda }\) stays irreducible, or decomposes in the form \(W \oplus W^{\vee }\) or in the form \(W_1 \oplus W_2 \oplus W_3\) with \(\dim (W_i) = 2\). But the latter implies \(W_i \cong st\) for the standard representation of SL(2). This would mean \(\dim (H^0_{der}) \le 6\), a contradiction. The case \(W \oplus W^{\vee }\) cannot happen either since the restriction of \(W \oplus W^{\vee }\) to \(SL(2) \times SL(2)\) would have an even number of summands. Therefore \(V_{\lambda }|_{H^0_{der}}\) is irreducible. Since it is selfdual irreducible of dimension 6 and carries a symplectic pairing, we conclude from Lemma 8.1 or Lemma 8.2 that \(H^0_{der}=Sp(6)\) and V is the standard representation. But then

$$\begin{aligned} H_X \ \cong \ Sp(6). \end{aligned}$$

Similarly consider \(X=\Pi (Ber^{1-b}\otimes [2b,b,0])\) for \(b>1\). Then \(X^\vee \cong X\). Then either \(H \subseteq O(6)\) or \(H\subseteq Sp(6)\) for \(H=H_X\). The image of \(H_2\) in H contains \(SL(2)^2\). Hence \(\dim (H_{der}^0) \ge 6\) and \(r\ge 2\). Furthermore \(H_{der}^0 \not \cong SL(3)\). If \(r=2\), then we get a contradiction by Mackey’s lemma. Hence \(r\ge 3\) and the restriction of the 6-dimensional representation \(V=\omega (X)\) of H to \(H_{der}^0\) remains irreducible. By the upper bound obtained from duality therefore the semisimple rank is \(r=3\). Hence V is a small irreducible representation of \(H_{der}^0\) of dimension 6. Hence by Lemma 8.1 we get \(H_{der}^0 = SO(V)\) resp. Sp(V), since \(H_{der}^0 \not \cong SL(3)\). In the second case then \(H=Sp(6)\). In the first case it remains to determine whether \(H = SO(6)\) or \(H=O(6)\).

Finally the case \(X=X_\lambda \) where \(L(\lambda )=[a,b,0]\) for \(a>b>0\) and \(a\ne 2b\). In this case \(X^\vee \not \cong Ber^\nu \otimes X\) for all \(\nu \in {\mathbb {Z}}\). The image of \(H_2\) in \(H=H_X\) contains \(SL(2)^3\), hence the restriction of \(V=\omega (X)\) to \(H_{der}^0\) remains again irreducible and defines a small representation of dimension 6. This now implies \(H_{der}^0 = SL(6)\), since X is not weakly selfdual which excludes the cases Sp(6) and SO(6). On the other hand we know that det(V) is nontrivial on the image of \(H_1\), and hence

$$\begin{aligned} H_X \ \cong \ GL(6). \end{aligned}$$

The structure of \(G_3\) follws from Theorem 6.8. \(\square \)

Example 9.7

For \(\Pi [2,1,0]\) the associated Tannaka group is \(H_X = Sp(6)\). Furthermore X corresponds to the standard representation of Sp(6) and decomposes accordingly. Hence

$$\begin{aligned} X \otimes X \ = \ I_1 \oplus I_2 \oplus I_3 \ \ \mod {{\mathcal {N}}}\end{aligned}$$

with the indecomposable representations \(I_i \in {{\mathcal {R}}}_3\) corresponding to the irreducible Sp(6) representations L(2, 0, 0), L(1, 1, 0) and L(0, 0, 0). Now consider the tensor product \(I_1 \otimes I_1\). For \(I_1\) corresponding to L(2, 0, 0) it decomposes as

$$\begin{aligned} I_1 \otimes I_1 \ = \ \bigoplus _{i=1}^{6}\ J_i \mod {{\mathcal {N}}}\end{aligned}$$

with the 6 indecomposable representations \(J_i\) corresponding to the 6 irreducible Sp(6)-representations in the decomposition

$$\begin{aligned} L(2,0,0)^{\otimes 2} \ = \ L(4,0,0) \oplus L(3,1,0) \oplus L(2,2,0) \oplus L(2,0,0) \oplus L(1,1,0) \oplus {\textbf{1}}.\end{aligned}$$

In this way we obtain the tensor product decomposition up to superdimension 0 for any summand of nonvanishing superdimension in such an iterated tensor product. Furthermore these indecomposable summands are parametrized by the irreducible representation of Sp(6). Although \(n = 3\) and the weight [2, 1, 0] are small, we found it hardly possible to achieve this result by a brute force calculation.

10 Tannakian induction: proof of the structure theorem

10.1 Restriction to the connected derived group

Recall that \(H_\lambda \) denotes the Tannaka group of the tensor category generated by \(X_\lambda \) and \(V_\lambda =\omega (X_\lambda )\) is a faithful representation of \(H_\lambda \). We have inclusions

$$\begin{aligned} G_{\lambda '} \hookrightarrow G_\lambda \hookrightarrow H_\lambda ^0 \hookrightarrow H_\lambda \end{aligned}$$

where \(G_{\lambda '}\) denotes the image of the natural map \((H_{n-1}^0)_{der} \rightarrow G_\lambda = (H_\lambda ^0)_{der}\). Similarly we denote by \(H_{\lambda '}\) the image of \(H_n\) in \(H_{\lambda }\). The restriction of \(V_\lambda \) to \(H_{n-1}\) (or \(H_{\lambda '}\)) decomposes

$$\begin{aligned} V_\lambda \ \cong \ \bigoplus _{i=1}^k \ V_{\lambda _i} \end{aligned}$$

where \(V_{\lambda _i}\) are the irreducible representations in the category \(Rep(H_{n-1})\) corresponding to the irreducible constituents \(L(\lambda _i), i=1,\ldots ,k\) of \(DS(L(\lambda ))\). To describe \(G_{\lambda '}\) we use the structure theorem for \(\mathcal {T}_{n-1}^+\) (induction assumption). Therefore it suffices to group the highest weight \(\lambda _i\) for \(i=1,\ldots ,k\) into equivalence classes. Using the structure theorem for the category \(\mathcal {T}_{n-1}^+\) and Theorem 4.1, we then obtain

$$\begin{aligned} G_{\lambda '} \cong \prod _{\lambda _i /\sim } G_{\lambda _i} \end{aligned}$$

Again using the structure theorem for \(G_{n-1}\), each \(V_{\lambda _i}\) is either irreducible on \(G_{\lambda _i}\) or it decomposes in the form \(W_i \oplus W_i^{\vee }\) and \(G_{\lambda _i} \cong SL(W)\). The groups \(G_{\lambda _i}\) are independent in case (NSD). For (SD) the only dependencies between them come from the equalities \(G_{\lambda _{k+1-i}} = G_{\lambda _i}\) for \(i=1,\ldots ,k\) by “Appendix A”. Using these strong conditions let us consider \(V_\lambda \) as a representation of \(H_\lambda ^0\). Since an irreducible representation of \(H_\lambda ^0\) is an irreducible representation of its derived group \(G_\lambda \), the decomposition of \(V_\lambda \) into irreducible representation for the restriction to \(H_\lambda ^0\) resp. \(G_\lambda \) coincide. Let

$$\begin{aligned} V_\lambda \ = \ \bigoplus _{\nu =1}^s \ W_\nu \end{aligned}$$

denote this decomposition. We then restrict each \(W_{\nu }\) to \(G_{\lambda '}\).

By induction each \(W'_l\) can be seen as the standard representation or its dual of a simple group of type ABCD.

10.2 Meager representations

If we use by induction the structure theorem for \(G_{n-1}\), we see that the representations \(W_i\) in \(V_{\lambda }|_{G_{\lambda }}\) are meager in the sense below. We analyze in this section the implications of \(W_i\) to be meager.

Definition 10.1

A finite dimensional representation V of a reductive group H will be called small if \(\dim (V) < \dim (H)\) holds.

Definition 10.2

A representation V of a semisimple connected group G will be called meager, if every irreducible constituent W of V factorizes over a simple quotient group of G and is isomorphic to the standard representation of this simple quotient group or isomorphic to the dual of the standard representation for a simple quotient group of Dynkin type ABCD.

If a representation V of H is small resp. meager, any subrepresentation of V is small resp. meager.

We now relax the notation and write G instead of \((H^0_{\lambda })_{der}\) and \(G'\) instead of \((H_{\lambda '})^0_{sc}\), the simply connected cover of \(G_{\lambda '} = (H_{\lambda '}^0)_{der}\). Then there exists a homomorphism \(\varphi :G' \rightarrow G\) with finite kernel. We show later in Theorem 10.7 that except for some special cases (\(n \le 3\) or Berezin twists of \(S^i\)) the situation will be as in the assumptions of the next Lemma 10.3.

So suppose \(G'\) is a semisimple connected simply connected group and V is a faithful meager representation of G. Each irreducible constituent of V then factorizes over one of the projections \(p_\mu : G' \rightarrow G'_\mu \) where \(G' \cong \prod G'_{\mu }\). We then say that the corresponding constituent is of type \(\mu \).

Lemma 10.3

Suppose V is an irreducible faithful representation of the semisimple connected group G of dimension \(\ge 2\). Suppose \(G'\) is a connected semisimple group and \(\varphi : G'\rightarrow G\) is a homomorphism with finite kernel such that

  1. (1)

    The restriction \(\varphi ^*(V)\) of V to \(G'\) is meager and for fixed \(\mu \) every (nontrivial) irreducible constituents of type \(\mu \) in the restriction of V to \(G'\) has multiplicity at most 2.

  2. (2)

    If an irreducible constituent \(W'\) of \(V|_{G'}\) occurs with multiplicity 2 for a type \(\mu \) in \(V\vert _{G'}\) (such a \(\mu \) is called an exceptional type), then either

    1. (i)

      \(W'\) is the standard representation of \(G_{\mu } \cong SL(2)\), or

    2. (ii)

      there is a unique type, say \(\mu = \mu _2\), such that the restriction of V to \(G_{\mu }'\) is equal to either \( 2W\oplus 2W^\vee \) as a representation of the quotient SL(W) of \(G'\) for \(\dim (W) \ge 3\) or equal to \(W \oplus W^{\vee }\) for \(\dim (W) = 2\) or

    3. (iii)

      there is a unique type, say \(\mu =\mu _0\), with \(G'_\mu \cong Sp(W')\) or \((G'_\mu )_{sc} = Spin(W)\) such that the standard representation st of \(G'_\mu \) occurs twice.

  3. (3)

    No irreducible constituent of the restriction of \(V\vert _{G'}\) is a trivial representation of \(G'\).

  4. (4)

    The semisimple group \(G'\) has at most one simple factor isomorphic to SL(2). The index, if it occurs, will be denoted \(\mu _1\).

Under these assumptions G is a simple group or \(G'\) is a product of exceptional types in the sense of (2).

Remark 10.4

For the connection to our case see Theorem 10.7. The cases (2)(ii) and (2)(iii) can appear for weakly selfdual weights, see “Appendix A.5” for the possible \(\lambda \). It is crucial here that we can assume \(n \ge 4\).

Proof

We may replace G and \(G'\) by their simply connected coverings without changing our assumptions, so that we can assume that G and \(G'=\prod _\mu G'_\mu \) decompose into a product of simple groups. Then V is not faithful any more, but has finite kernel. The restriction of the meager representation V to \(G'\) decomposes into the sum \(\bigoplus _\mu J_\mu \) of representations \(J_\mu \) such that \(J_\mu \) is trivial on \(\prod _{\lambda \ne \mu } G'_\lambda \)

$$\begin{aligned} V\vert _{G'} = \bigoplus _\mu J_\mu \end{aligned}$$

hence \(J_\mu \) can be considered as a representation of the factor \(G'_\mu \) of \(G'\). Furthermore \(J_\mu \) is either an irreducible representation of \(G'_\mu \), or the direct sum \(J_\mu \cong W\oplus W^\vee \) (as a representation of \(G'_{\mu }\cong SL(W)\)) by the assumption (1) and (2) or there exists a unique type \(\mu \) of Dynkin type BCD where \(J_\mu = st \oplus st\) for the standard representation st of this group \(G'_\mu \).

If the semisimple connected G is not simple, \(G=G_1 \times G_2\) is a product of semisimple groups and the irreducible representation V is an external tensor product

$$\begin{aligned} V = V_1 \boxtimes V_2 \end{aligned}$$

of irreducible representations \(V_1, V_2\) of \(G_1\) resp. \(G_2\). Since V has finite kernel and G is connected, \(\dim (V_i)>1\) holds. For each factor \(G'_\mu \hookrightarrow G'=\prod _\mu G'_\mu \) consider the composed map

$$\begin{aligned} G'_\mu \rightarrow G_1 \times G_2. \end{aligned}$$

This map has finite kernel.

We claim that there exists at least one index \(\mu \) such that both compositions \(G'_\mu \rightarrow G_i\) with the projections \(G\rightarrow G_i\) (\(i=1,2\)) are nontrivial except when \(G'\) has only exceptional types. To prove the claim, suppose \(G'_\mu \rightarrow G_2\) would be the trivial map. Then the restriction of V to \(G'_\mu \subseteq G'\) is \(V\vert G'_\mu = \dim (V_2) \cdot V_1\vert _{G'_\mu }\). Hence \(\dim (V_2) \le 2\), since otherwise we get a contradiction to assumption (1) of the lemma. \(V_1\vert _{G'_\mu }\) also contains at least one nontrivial irreducible constituent by assumption (3), and this constituent can occur by assumption (1) at most with multiplicity two in \(V\vert _{G'}\). If then \(\dim (V_2)=2\), then there must exist a nontrivial irreducible constituent \(I_\mu \subseteq V_1\vert _{G'_\mu }\) of \(G'_\mu \) by assumption (3). Hence if \(\dim (V_2) = 2\), \(V\vert _{G'_\mu } \) contains \(I_\mu \oplus I_\mu \) both of some type \(\mu \) and we are in an exceptional type [see (2)].

We assume now that \(\{\mu \}\) is not an exceptional type. We may therefore choose \(\mu \) so that both \(G'_\mu \rightarrow G_i\) are both nontrivial. Then

$$\begin{aligned} V\vert _{G'_\mu } = V_1\vert _{G'_\mu } \otimes V_2\vert _{G'_\mu } \end{aligned}$$

is the tensor product of two nontrivial representations \(V_1\vert _{G'_\mu } \) and \(V_2\vert _{G'_\mu }\) of \(G'_\mu \). Since \(V\vert _{G'}\) is a meager representation of \(G'\), all irreducible constituents of the restriction of \(V\vert _{G'}\) to \(G'_\mu \) are trivial representations of \(G'_\mu \) except for at most two of them [see assumption (1)], which are standard representations up to duality. Since \(V_i\) are irreducible representations of G (recall \(V \cong V_1 \boxtimes V_2\)) and V has finite kernel, the restriction of V to \(G'_\nu \) has finite kernel. Hence both of the representations \(V_i\vert _{G'_\mu }\) have finite kernel, hence contain an irreducible nontrivial representation of \(G'_\mu \). Otherwise the restriction \(V\vert _{G'_\mu }\) would be trivial contradicting that \(G'_\mu \rightarrow G_i\) have finite kernel for both \(i=1,2\) and \(V_i\) both have finite kernel on \(G_i\). For every nontrivial irreducible representations \(I_1 \subseteq V_1\vert _{G'_\mu }\) and \(I_2 \subseteq V_2\vert _{G'_\mu }\) of \(G'_\mu \) the representation

$$\begin{aligned} I_1 \otimes I_2 \end{aligned}$$

only contains trivial representations and standard representations st up to duality by assumption (2). Since the trivial representation occurs at most once in the tensor product of two irreducible representations, this implies \(I_1\otimes I_2 \subseteq J_\mu \oplus 1 \subseteq st \oplus st^\vee \oplus 1\) (note that \(\mu \) is not exceptional). Hence \(\dim (I_1)\dim (I_2)\le 1 + 2\cdot \dim (st) < 1 + 2\cdot \dim (st) + \dim (st)^2\). Hence \(\min (\dim (I_\nu )) < 1 + \dim (st)\). In particular, the corresponding representation with minimal dimension, say \(I_1\), has dimension \(\le \dim (st)\) and hence \(I_1\) is a small representation of \(G'_\mu \). Since it is small, it belongs to the list of Lemma 8.2. Therefore \(I_1\) is the standard representation of \(G'_\mu \) or its dual, unless the group \(G'_\mu \) is of Dynkin type \(D_4\) and \(I_1\) is a spin representation. In the first case, considering highest weights it is clear that \(st \otimes I_2 \subseteq st\oplus st^\vee \oplus 1\) is impossible. In the remaining orthogonal case \(G'_\mu \) of Dynkin type \(D_4\), the representation \(I_1 \otimes I_2\) must have dimension \(\ge 8^2\). But this contradicts \(\dim (I_1)\dim (I_2)\le 1 + 2\cdot \dim (st) = 1 + 8 + 8 = 17\), and finally proves our assertion. \(\square \)

Corollary 10.5

In the situation of Lemma 10.3, the restriction of the representation V to the group \(G'\) is multiplicity free unless \(G'\) contains an exceptional type (in which case the irreducible constituent has multiplicity 2). If \(G'\) has at least one non-exceptional type, then the restriction contains at least one constituent with multiplicity 1.

Proof

If the restriction of V to \(G'\) contains an irreducible summand I of \(G'\) with multiplicity \(\ge 2\), then the restriction of I at least under one map \(G'_\mu \rightarrow G\) contains a nontrivial constituent of \(G'_\mu \) with multiplicity \(>1\). Hence the restriction of I contains \(J_\mu \) by the assumption (1) and (2) of the main lemma such that \(J_\mu \cong I_\mu \oplus I_\mu \) and we are in an exceptional type. \(\square \)

Definition 10.6

Let \(G, \ G'\) be semisimple connected groups and \(\varphi : G'\rightarrow G\) a homomorphism with finite kernel. The restriction of the irreducible representation V of G to \(G'\) is called weakly multiplicity-free if at least one irreducible constituent has multiplicity 1.

10.3 Mackey–Clifford theory

Let H be a reductive group and \(H^0\) its connected component. We assume that G is the connected derived group of \(H^0\). Let V be a finite dimensional irreducible faithful representation of H and let

$$\begin{aligned} V\vert _{H^0} = W_1 \oplus \cdots \oplus W_s \end{aligned}$$

be the decomposition of V into irreducible summands \((W_\nu ,\rho _\nu )\) after restriction to \(H^0\). The restriction of each \(W_\nu \) to G remains irreducible (this follows from Schur’s lemma and the fact that the image of \(H^0\) in \(GL(W_\nu )\) is generated by the image of G in \(GL(W_\nu )\) and the image of the connected component of the center of \(H^0\), whose image is in the center of GL(W)). By Clifford theory [16] \(\pi _0(H)=H/H^0\) acts on the isotypic components \(m_{\mu } W_\nu \) permuting them transitively; i.e. \(\rho _\nu (g) = \rho _1(h g h^{-1})\) for certain \(h \in H\). Here we define the isotypic part of an irreducible \(W_{\nu }\) to be the sum of all subrepresentations of \(V|_{H^0}\) which are isomorphic to \(W_{\nu }\). Since \(\pi _0(H)\) acts transitively on these isotypic components, the multiplicity \(m = m_{\mu }\) of each isotypic part is the same. Let us write

$$\begin{aligned} V\vert _{H^0} = m \cdot (W_1 \oplus \cdots \oplus W_{\tilde{s}}).\end{aligned}$$

Representations \((W_\nu ,\rho _\nu )\) from different isotypic parts are pairwise nonisomorphic representations of \(H^0\) (in our application later this also remains true for the restriction to G by the \(G'\)-multiplicity arguments). But \(\rho _1(h_1 g h_1^{-1})\cong \rho _1(h_2 g h_2^{-1})\) as representations of \(g\in H^0\) (or \(g\in G\)) holds if \(h_1^{-1}h_2 \in H^0\) (resp. \(h=h_1^{-1}h_2 \in H^0\)). Therefore the automorphism \(int_h: H^0 \rightarrow H^0\) acts trivially and the isotypic components \(m W_{\nu }\) are permuted transitively by \(Out(H^0) = Aut(H^0)/Inn(H^0)\). If a finite group acts transitvely on a set X, this implies that the cardinality of the set divides the order of the group. Therefore \(\tilde{s} \le | Out(H^0) |\). Furthermore \(\dim W_i = \frac{1}{m}\frac{1}{\tilde{s}} \dim V\).

If \(H = H_{\lambda }\) is the Tannaka group of an irreducible maximal atypical module \(L(\lambda ) \in \mathcal {T}_n^+\) and \(V = V_{\lambda } =\omega (L(\lambda ))\) is the associated irreducible representation of H and \(W_1,\ldots ,W_s\) are the irreducible constituents of the restriction of V to \(H^0\), then the following theorem holds.

Theorem 10.7

Suppose that \(L(\lambda )\) is not a Berezin twist of \(S^i\) for some i or its dual, and suppose \(n\ge 4\). Then for \(G=(H^0_{\lambda })_{der}\) and \(G' = G_{\lambda '}\) the irreducible representations \(W_1,\ldots ,W_s\) of G satisfy the conditions of Lemma 10.3 and \(G'\) has at least one non-exceptional type \(\mu \). In particular G is a connected simple algebraic group and V is a weakly multiplicity free representation of \(H^0\).

Remark 10.8

See also “Appendix A.5” for an overview.

Proof

The irreducibility and faithfulness is a tannakian consequence of the definitions. We claim that condition (1) and (2) follow from induction on n and the classification of similar and selfdual derivatives \(\lambda _i\) of \(\lambda \) in “Appendix A”.

If we restrict \(V_{\lambda _i}\) to \(G_{\lambda '}\), the induction assumption implies that the restriction is either irreducible (the regular case) or \(V_{\lambda _i}\) decomposes as \(W \oplus W^{\vee }\) for the group SL(W) (the exceptional case). The exceptional case can only happen if \(\lambda _i\) is of type (SD).

If \(\lambda \) is (NSD), then equivalence classes of its derivatives consist of one element by Proposition A.4. At most one \(\lambda _i\) is of (SD) type. If \(\lambda \) is (SD), equivalence classes can consist of one or two elements by Corollary A.6. At most two derivatives \(\lambda _{\nu }, \lambda _{\mu }\) can be of type (SD) by Lemma A.9. In case \(\lambda _i\) is (NSD), the restriction of \(V_{\lambda _i}\) to \(G_{\lambda '}\) remains irreducible.

Let us then assume that we are in the case where \(\lambda _{\nu }\) is not equivalent to any other derivative. Then \(\lambda _{\nu }\) belongs to one the three cases (1), (2), (3) treated in the proof of Lemma A.9. In the regular cases \(V_{\lambda _i}\) remains irreducible. If \(\lambda _i\) is exceptional, it decomposes as \(W \oplus W^{\vee }\) for the group SL(W). Then the restriction of \(W_{\nu }\) to \(G_{\lambda '}\) has multiplicity 1 unless \(\dim (W) = 2\) or \(\dim (W) = 1\) since \(W^{\vee } \ncong W\) for \(\dim (W) \ge 3\). These two cases would lead to an irreducible constituent of multiplicity 2 (trivial representation or standard representation of SL(2)).

Assume therefore that \(\lambda _{\nu } \sim \lambda _{\mu }\) for \(\nu \ne \mu \). Then \(\lambda \) is of ladder type and \(\{\nu ,\mu \} = \{1,k\}\) such that \(G_{\nu } = G_{\mu }\) is of (SD)-type and either symplectic or orthogonal regular or exceptional (see Lemma A.9). Here we use that \(n \ge 4\).

In the exceptional case the standard representation W and its dual \(W^{\vee }\) of SL(W) appear with multiplicity 2 in the restriction of V to \(G_{\lambda '}\). We don’t have \(W \cong W^{\vee }\) unless \(\dim (W_{\mu }) = 2\) which is impossible for \(n \ge 4\) (it would mean \(\dim (V_{\lambda _1}) = 4\), but \(\dim (V_{\lambda _1}) = (n-1)!\) since \(\lambda \) is of ladder type).

In the regular case (SD)-case we have \(G_1 = G_k = SO(V_{\lambda _1})\) or \(Sp(V_{\lambda _1})\). Both \(V_{\lambda _1}\) and \(V_{\lambda _k}\) remain irreducible after restriction, hence the multiplicity is again 2.

The uniqueness assertion about the types in (2)(ii) and (2)(iii) follows since case (iii) can only occur for \(\lambda _{\nu } \sim \lambda _{\mu }\) of (SD) type, and there are at most two such derivatives. The multiplicity 2 assertion for (2)(ii) holds since at most one selfdual derivative can have dimension 4 or 2 (see proof of condition 3 and 4).

Condition (3) is seen as follows: The trivial representation of \(G'\) is attached to a derivative \(\lambda _\mu \) of \(\lambda \) only if \(L(\lambda )\) isomorphic to \(S^i \otimes Ber^j\) for some \(i\ge 1\) and some \(j\in {\mathbb {Z}}\) by Lemma D.3. Concerning condition 4): A factor \(G'_\mu \) of \(G'\) of rank 1 (i.e. with derived group SL(2)) is attached to some derivative \(\lambda _\mu \) of \(\lambda \) only if \(L(\lambda )=S^1\) or \(\lambda \) has only two sectors, one sector S of rank 1 and the other sector \(S'\) corresponds to \(S^1\) on the level \(n-1\). In other words \(\partial S S'\) resp. \(S' \partial S\) gives \(S^1\) and the corresponding group SL(2), but not the other derivative unless \(n\le 3\).

figure f

Hence by our assumptions, the group \(G'\) has at most one simple factor SL(2). If an irreducible constituent of the restriction of V to \(G'\) has multiplicity 2, it comes from a derivative of type (SD). Hence if all types of \(G'\) are exceptional, all derivatives of \(L(\lambda )\) would have to be selfdual. This can only happen for \(n \le 3\) by the analysis in “Appendix A”. Hence Lemma 10.3 and Corollary 10.5 imply the last statement. \(\square \)

Theorem 10.9

The simple group G is of type ABCD and \(W_1\vert _G\) is either the standard representation of G or its dual.

Proof

We suppose that \(L(\lambda )\) is not a Berezin twist of \(S^i\) for some i and suppose \(n\ge 4\). We distinguish the cases NSD and SD. In the NSD-case we claim that we have

$$\begin{aligned} r(G_\lambda ) \ge (\dim (V_\lambda )-1)/2 \ \end{aligned}$$

and that for \(n\ge 4\) and \(\dim (V_\lambda ) \ge 4\)

$$\begin{aligned} \dim (G_\lambda ) > \dim (V_\lambda ) \end{aligned}$$

holds (note that \(\dim (V_\lambda ) \le 3\) for \(n\ge 4\) implies \(k=1\) and \(\dim (V_\lambda )=\dim (V_{\lambda _1})\)). For all \(i=1,\ldots ,k\) the superdimension formula of [51] [36, Section 16] implies by Lemma D.5 that

$$\begin{aligned} \dim (V_\lambda ) \le n \cdot \dim (V_{\lambda _i})/r_i \ \end{aligned}$$

where \(r_i =r(V_{\lambda _i})\ge 1\) is the rank of \(\lambda _i\). Obviously \( \dim (G_{\lambda _i}) \le \dim (G_{\lambda })\).

Since we excluded the \(S^i\)-case, no \(V_{\lambda _i}\) has dimension 1 by Lemma D.3. At most one of the representations \(V_{\lambda _i}\) is selfdual by Lemma A.9. We make a case distinction on whether there exists one \(V_{\lambda _i}\) that splits in the form \(W'_i \oplus (W'_i)^{\vee }\) upon restriction to \(G_{\lambda '}\) or not. In the latter case we know \(r(G_{\lambda _i}) \ge \frac{1}{2} \dim (V_{\lambda _i}) \) by Theorem 6.2 and the induction assumption. Now by Proposition A.4 and the assumption (NSD) all \(\lambda _i\) in the derivative of \(\lambda \) are inequivalent for \(i\ne j\). Hence we get

$$\begin{aligned} r(G_\lambda ) \ge \sum _{i} r(G_{\lambda _i}) \ge \sum _i \frac{1}{2}\dim (V_{\lambda _i}) \ge \frac{1}{2}(\dim (V_\lambda )).\end{aligned}$$

Since \(\dim (G_{\lambda _i}) \ge 3 r(G_{\lambda _i})\), this implies \( \dim (G_\lambda ) \ge \frac{3}{2} (\dim (V_\lambda ) - 1) \) and hence \( \dim (G_\lambda ) > \dim (V_\lambda ) \) (note that we have at least one SL factor \(G_{\lambda _i}\) for which \(r(G_{\lambda _i}) > \frac{1}{2} \dim (V_{\lambda _i})\)). If \(V_{\lambda }\) splits \(V_{\lambda } = W_1 \oplus \cdots \oplus W_s\) we may replace \(V_{\lambda }\) by any \(W_\nu \) for an even better estimate. Therefore Lemma 8.2 implies that \(V_{\lambda }\) (or \(W_{\nu }\)) is the standard representation or its dual of a simple group of type ABCD. If \(V_{\lambda }\) stays irreducible, then we obtain \(G_{\lambda } \cong SL(V_{\lambda })\) since \(V_{\lambda }\) is not selfdual.

If \(V_{\lambda _i}\) splits, \(G_{\lambda _i} \cong SL(W_i)\) for \(V_{\lambda } \cong W_i \oplus W_i^{\vee }\) by induction assumption. If the dimension of \(V_{\lambda _i}\) is \(2d_i\), we then have \(r(G_{\lambda _i}) = d_i - 1\) and therefore have to replace the estimate \(r(G_{\lambda _i}) \ge \frac{1}{2} \dim (V_{\lambda _i})\) by the estimate \(r(G_{\lambda _i}) \ge \frac{1}{2} (\dim (V_{\lambda _i} -2))\). Since \(V_{\lambda _i}\) can only decompose if it is of type SD, \(L(\lambda )\) has more than one sector. All the other \(k-1 \ge 1\) derivatives \(L(\lambda _i)\) are of type NSD and define inequivalent \(SL(V_{\lambda _i})\). For each of these we obtain \(r(G_{\lambda _j}) = \dim V_{\lambda _j} - 1\). Summing up we obtain

$$\begin{aligned} r(G_\lambda ) \ge \sum _{i} r(G_{\lambda _i}) \ge \frac{1}{2}(\dim (V_{\lambda _i}) - 2) + \sum _{j \ne i} \dim (V_{\lambda _j}) - 1.\end{aligned}$$

This implies again the necessary estimates to apply Lemma 8.2.

We now consider the SD-case. If \(V_{\lambda }\) decomposes

$$\begin{aligned} V_{\lambda }|_{G_{\lambda }} \cong W_1 \oplus \cdots \oplus W_s \end{aligned}$$

then we can assume by reindexing that \(\dim (W_i) = \frac{1}{s}\dim (V_{\lambda })\). Note that \(\dim (W_1) > 1\) follows from the induction assumption.

In the SD case we proceed as follows: We first show that \(V_{\lambda }\) or \(W_1\) is small. Since we cannot prove the strong rank estimates for \(r(G_{\lambda })\) as in the NSD case, we work through the list of exceptional cases in Lemma 8.1.

The list of superdimensions in the n \(=\) 4 and n \(=\) 5 case in Sects. 15 and D.2 along with the induction assumption shows in these cases that \(V_{\lambda }\) is small. Therefore we can assume \(n \ge 5\). We use the known formulas \(\dim (SL(n)) = n^2 - 1\), \(\dim SO(n) = \frac{n(n-1)}{2}\) and \(\dim (Sp(2n)) = n (2n+1)\).

We recall from the analysis in Lemma A.9 that \(L(\lambda )\) can only have more than one selfdual derivative if it is completely unnested, i.e. it has n sectors of cardinality 2. In this case it has 2 selfdual derivatives coming from the left and rightmost sectors and, if n is odd, another derivative coming from the middle sectors. If \(\lambda \) is not of this form, then the unique weakly selfdual derivative comes from the middle sector (of arbitrary rank).

We want to show \(\dim (G_{\lambda }) > \dim (V_{\lambda })\). By induction \(G_{\lambda _i}\) is either \(SO(V_{\lambda _i})\), \(Sp(V_{\lambda _i})\), \(SL(V_{\lambda _i})\) or \(SL(W_i)\) for \(V_{\lambda _i} = W'_i \oplus (W'_i)^{\vee }\). We estimate the dimension of \(G_{\lambda _i}\) via \(\sum \dim (G_{\lambda _i})\). We claim that we can assume that we have more than one sector because otherwise \(\dim (V_{\lambda }) = \dim (V_{\lambda _1})\) implies that \(V_{\lambda }\) is small using the induction assumption. If \(V_{\lambda _1}\) is an irreducible representation of \(G_{\lambda '}\) the claim is clear by induction assumption. If it splits \(V_{\lambda _1} = W'_1 \oplus (W'_1)^{\vee }\), then \(\dim ( V_{\lambda _1}) < \dim (SL(W'_1))\) provided \({{\,\textrm{sdim}\,}}(L(\lambda _1)) \ge 3\). Now \({{\,\textrm{sdim}\,}}(L(\lambda _1)) = 2\) can only happen for \(L(\lambda _i) \cong Ber^{r} \otimes S^1\) for some r (and then \(V_{\lambda _1}\) is an irreducible representation of \(G_{\lambda '}\)). We therefore assume \(k >1\). The worst estimate for the dimension is obtained if all \(V_{\lambda _i}\) split as \(W'_i \oplus (W'_i)^{\vee }\) and therefore \(G_{\lambda _i} \cong SL(W_i)\). This case can only happen if either \(n=2\) or \(n=3\). For \(n \ge 4\) the lowest estimate for the dimension of \(G_{\lambda }\) occurs if \(\lambda \) is completely unnested with 2 selfdual derivatives coming from the left and right sector and we have n/2 equivalence classes of derivatives (or \(\left\lfloor {n/2}\right\rfloor + 1\) for odd n). The left and right sector then contribute a single \(SL(W'_1) = SL(W'_k)\) and if n is even for all other derivatives \(G_{\lambda _i} \cong SL(V_{\lambda _i})\) with \(V_{\lambda _i} \sim V_{\lambda _{k-i}}\) and therefore same connected derived Tannaka group. If \(n =2l+1\) is odd the middle sector can contribute another derivative of type SD with Tannaka group \(SL(W'_{l+1})\). The dimension estimate works as in the case above and we therefore ignore this case.

We show now that \(\dim ( G_{\lambda }) > \dim (V_{\lambda })\) provided we have two SD derivatives coming from the left- and rightmost sector. Denote by \(d_i\) the dimension of \(V_{\lambda _i}\). For \(i=1,k\) it is even \(d_1 = 2d'_1 = 2d'_k\) by Lemma D.4. We then obtain for the dimension of \(G_{\lambda '}\)

$$\begin{aligned} \dim ( G_{\lambda '}) = \frac{1}{2} \big (\big (d'_1\big )^2 -1 + \big (d'_k\big )^2 - 1\big ) + \frac{1}{2} \sum _{j \ne 1,k} d_j^2 - 1.\end{aligned}$$

It is enough to show \(2 \dim V_{\lambda _i} < \dim G_{\lambda _i}\) for each i. The smallest possible superdimensions for a selfdual irreducible representation are \(2,4,12,\ldots \). The \(\dim = 2\) case can only happen for \(L(\lambda _i) \cong Ber^{r} \otimes S^1\) which is not possible by assumption. Hence \(d'_1 \ge 3\). This case occurs for [2, 1, 0] for \(n=3\), [2, 2, 0, 0] for \(n=4\) and all their counterparts for larger n by appending zeros to the weight (e.g. [2, 1, 0, 0]). These are not derivatives of a selfdual representation \(L(\lambda )\) unless \(L(\lambda )\) has one sector (which we excluded). Therefore we can assume \(d'_1 \ge 6\). Then

$$\begin{aligned}2 \ \dim (V_{\lambda _1}) = 4 d'_1 < \big (d'_1\big )^2 -1 = \dim (G_{\lambda _1}).\end{aligned}$$

For the NSD derivatives we can exclude the case \(d_i = 2\) since this only happens for \(L(\lambda _i) \cong Ber^{\ldots } \otimes S^1\). For \(d_i \ge 3\) we obtain \(2 d_i < d_i^2 - 1\), hence again \(2 \dim (V_{\lambda _i}) < \dim (G_{\lambda _i})\). Clearly this estimates also hold if we have more than n/2 equivalence classes of weights or if we have \(SO(V_{\lambda _i})\) or \(Sp(V_{\lambda _i})\) in case of \(SL(W_i)\).

Hence \(\dim (V_{\lambda }) < \dim (G_{\lambda })\). If \(V_{\lambda }\) is an irreducible representation of \(G_{\lambda }\), it is a small representation of \(G_{\lambda }\) and Lemma 8.1 applies. If it decomposes \(V_{\lambda } \cong W_1 \oplus \cdots \oplus W_s\), then each \(W_{\nu }\) is an irreducible small representation of \(G_{\lambda }\).

Assume first that \(V_{\lambda } \cong W_1 \oplus \cdots \oplus W_s\) with \(s \ge 3\) and \(\dim (W_1) \le \frac{1}{s}\dim (V_{\lambda })\). Again the smallest rank estimate for the subgroup \(G_{\lambda '}\) occurs for \(n \ge 4\) if \(\lambda \) is completely unnested with 2 selfdual derivatives coming from the left and right sector and we have n/2 equivalence classes of derivatives (we assume here n even. In the odd case we can have another derivative from the middle sector. The estimate below still holds). Then

$$\begin{aligned} r(G_{\lambda }) \ge r(G_{\lambda '})&\ge \frac{1}{2}\left( d_1/2 -1 + d_k/2 - 1 + \sum _{j \ne 1,k} d_j -1\right) \\ {}&= \frac{1}{2}(\dim (V_{\lambda }) - k - d_1/2 - d_k/2). \end{aligned}$$

In the completely unnested case this equals

$$\begin{aligned} \frac{1}{2}(n! - n - (n-1)!).\end{aligned}$$

We need \(r(G_{\lambda }) \ge \frac{1}{2}(\dim (V_{\lambda }) - 1)\) to apply Lemma 8.2. We replace now \(V_{\lambda }\) by \(W_1\) with \(\dim (W_1) \le 1/s \ \dim (V_{\lambda })\). For \(n\ge 4\) and \(s \ge 2\) we obtain \((n!/s) -1 \le n! - n - (n-1)!\), hence Lemma 8.2 can be applied to the irreducible representation \(W_1\).

If \(\lambda \) is not completely unnested, it can have at most one SD derivative coming from the middle sector for \(k = 2l+1\) odd. Then we obtain

$$\begin{aligned} r(G_{\lambda }) \ge r(G_{\lambda '})&\ge \frac{1}{2}\left( d_{l+1}/2 -1 + \sum _{j \ne l+1} d_j -1\right) \\ {}&= \frac{1}{2}(\dim (V_{\lambda }) - k - d_{l+1}/2).\end{aligned}$$

As above we replace \(V_{\lambda }\) with \(W_1\) with \(\dim (W_1) \le \frac{1}{s}V_{\lambda }\) and show \(\dim (V_{\lambda }/s - 1) \le \dim (V_{\lambda }) - k - d_{l+1}/2\). For \(s=2\) this is equivalent to \(\dim (V_{\lambda }) \ge d_{l+1} + 2(k-1)\). This follows easily from \(\dim (V_{\lambda }) = \dim (V_{\lambda _{l+1}}) \frac{n}{r_{l+1}}\) (Lemma D.5). For \(s>2\) the estimates are even stronger. The cases where the SD derivative occurs and contributes \(SO(V_{\lambda _{l+1}})\) or \(Sp(V_{\lambda _{l+1}})\), or the case in which no SD derivative occurs, can be treated the same way.

We can therefore assume that either (a) \(V_{\lambda }\) is an irreducible representation of \(G_{\lambda }\) or it splits in the form \(V_{\lambda } = W \oplus W^{\vee }\). The analysis of small superdimensions in Sect. D.2 shows that the possible superdimensions of weakly selfdual irreducible representations less than 129 are

$$\begin{aligned} 1, \ 2, \ 6, \ 12, \ 20, \ 24, \ 30, \ 42, \ 56, \ 70, \ 72, \ 80, \ 90, \ 110, \ 112.\end{aligned}$$

Except for the numbers 20 and 56 none of the exceptional dimensions in Lemma 8.1 is equal to either the superdimension or half the superdimension of an irreducible weakly selfdual representation in \(\mathcal {T}_n^+\). It is easy to exclude these two cases (see Sect. D.2) since in this case \(V_{\lambda }\) or W would be either a symmetric or alternating square of a standard representation (which would give a contradiction to the induction assumption) or the irreducible representation of minimal dimension of \(E_7\) which is impossible by rank estimates. \(\square \)

Theorem 10.10

Either the restriction of \(V_{\lambda }\) to \(H^0_{\lambda }\) and \(G_{\lambda }\) is irreducible, or \(G\cong SL(W)\) and \(V\vert _G \cong W\oplus W^\vee \) for a vectorspace W of dimension \(\ge 3\). If \(V\vert _G \cong W\oplus W^\vee \), then

$$\begin{aligned} V_{\lambda } \cong Ind_{H_1}^{H}( W) \end{aligned}$$

for a subgroup \(H_1\) of index 2 between \(H^0\) and H. In particular \(V_{\lambda }\) is an irreducible representation of \(G_{\lambda }\) if \(L(\lambda )\) is not weakly selfdual.

Proof

As in the statement of Theorem 10.7 we can assume that \(n \ge 4\) and that \(L(\lambda )\) is not a Berezin twist of \(S^i\) (or its dual) since these cases were already treated in Sect. 9.

We claim that the representation \(V\vert _{H^0} = W_1\oplus \cdots \oplus W_s\) is multiplicity free. Since the restriction of V to \(G'\) is weakly multiplicity free, at least one irreducible constituent occurs only with multiplicity 1 for some (non-exceptional) \(\mu \). By Clifford theory the multiplicity of each isotypic part in the restriction of V to \(H^0\) is the same (since \(\pi _0\) acts transitively). If the multiplicity of each isotypic part would be bigger than 1, the restriction of V to \(G'\) could not be weakly multiplicity free. Therefore the multiplicity of each isotypic part is 1. Any \(W_{\nu }\) restricted to \(G_{\lambda }\) is irreducible (restriction to the derived group). Since \(G_{\lambda }\) is a normal subgroup of H, H still operates transitively on the set \(\{ W_{\nu }|_{G_{\lambda }} \}\). Fix any \(W_{\nu }|_{G_{\lambda }}\). Its H-orbit has \(s'\) elements where \(s'\) divides s and \(s/s'\) is the multiplicity of each \(W_{\nu }|_{G_{\lambda }}\) in \(V_{\lambda }|_{G_{\lambda }}\). Hence the argument from Clifford theory explained preceding Theorem 10.7 shows

$$\begin{aligned} s' \le | Out(G)|. \end{aligned}$$

But a nontrivial outer automorphism of G that does not fix the isomorphism class of the standard representation \(W_1\) of G exists only for the groups G of the Dynkin type \(A_r\) for \(r\ge 2\) (note that we can ignore the \(D_4\)-case since no weakly selfdual irreducible representation has superdimension 8). For the special linear groups \(G=SL({\mathbb {C}}^{r+1})\) the nontrivial representative in Out(G) it is given by \(g\mapsto g^{-t}\). The twist of the standard representation by this automorphism gives the isomorphism class of the dual standard representation \(W_1^\vee \). This implies \(s'=1\) or \(s'=2\). If \(s' = 2\), then \(V_{\lambda }|_{G_{\lambda }} \cong W \oplus W^{\vee }\) where W is the standard representation of SL and \(G_{\lambda } \cong SL(W)\). Since \(V_{\lambda }|_{G_{\lambda '}}\) is weakly multiplicity free and \(G_{\lambda '} \subset G_{\lambda }\), \(V_{\lambda }|_{G_{\lambda }}\) is weakly multiplicity free as well. Accordingly \(s/s' = 1\) and we also obtain \(s=1\) or 2. If \(s=2\), Clifford theory further implies that

$$\begin{aligned} V_{\lambda } \cong Ind_{H_1}^{H}(W) \end{aligned}$$

for a subgroup \(H_1\) of index 2 between \(H^0\) and H. \(\square \)

Remark 10.11

Since \(W \oplus W^{\vee }\) is selfdual, this implies in particular that \(V_{\lambda }\) can only decompose if \(L(\lambda )\) is weakly selfdual. If \(V_{\lambda }\) decomposes, its restriction to \(G_{\lambda '}\) is of the form \(\bigoplus _i W_i \oplus W_i^{\vee }\). This leads to some restrictions on SD weights \(\lambda \) such that \(V_{\lambda }\) decomposes in the form \(W \oplus W^{\vee }\). Consider for an instance the weakly selfdual weight \([n-1,n-2,\ldots ,1,0]\) for odd \(n = 2l+1\). Then \(V_{\lambda }\) can only decompose if the irreducible representation \(V_{\lambda _l+1}\) associated to the middle derivative \(L(\lambda _{l+1})\) decomposes upon restriction to \(G_{\lambda '}\) in the form \(W'_{l+1} \oplus (W'_{l+1})^{\vee }\).

11 The structure theorem on the full Tannaka groups

We discuss the full Tannaka groups \(H_{\lambda }\) in this section. To this end we analyze the invertible elements in \(Rep(H_n)\), i.e. \(Pic(H_n)\), or in down-to-earth terms the character group of \(H_n\).

11.1 Invertible elements

For a rigid symmetric k-linear tensor category \({\mathcal {C}}\) an object I of \({\mathcal {C}}\) is called invertible if \(I \otimes I^\vee \cong \textbf{1}\) holds. The tensor product of two invertible objects of \({{\mathcal {C}}}\) is an invertible object of \({{\mathcal {C}}}\). Let \(Pic({{\mathcal {C}}})\) denote the set of isomorphism classes of invertible objects of \({{\mathcal {C}}}\). The tensor product canonically turns \((Pic({{\mathcal {C}}}, \otimes ))\) into an abelian group with unit object \(\textbf{1}\), the Picard group of \({{\mathcal {C}}}\).

Suppose that the categorial dimension \(\dim \) is an integer \(\ge 0\) for all indecomposable objects of \({\mathcal {C}}\). An indecomposable object I of \({\mathcal {C}}\) is an invertible object in \(\overline{{\mathcal {C}}} = \mathcal {C}/\mathcal {N}\) if and only if \({{\,\textrm{sdim}\,}}(I) =1\) holds. In fact \(\dim (I)=1\) implies \(\dim (I^\vee )=1\) and hence \(\dim (I \otimes I^\vee )=1\). Hence \(I \otimes I^\vee \cong {\textbf{1}}\oplus N\) for some negligible object N. Note that the evaluation morphisms \(eval: I \otimes I^\vee \rightarrow {\textbf{1}}\) splits since \(\dim (I)\ne 0\).

11.2 \(Pic(\overline{\mathcal {T}}_{n})\) and its generators

Since \({\overline{{\mathcal {T}}}}_n \sim Rep_k(H_n)\), to determine the Picard group \(Pic({\overline{{\mathcal {T}}}}_n)\) is tantamount to determine the character group of \(H_n\). It coincides with the character group of the factor commutator group \(H^{ab}_n\) of \(H_n\). Hence \(H_n^{ab}=H_n/G_n\) is determined by \(Pic({\overline{{\mathcal {T}}}}_n)\).

The Picard group \(Pic({\overline{{\mathcal {T}}}}_n)\) can be determined from the individual \(Pic(H_{\lambda })\), but it is preferable to choose different generators.

For a Tannakian category \({\mathcal {T}}\) over an algebraically closed field k of characteristic zero, generated by finitely many irreducible objects, let \(H({\mathcal {T}})\) denote the Tannaka group of \({\mathcal {T}}\). In our situation the Tannaka group \(H_n\) is a projective limit of certain algebraic Tannaka groups \(H({\mathcal {T}})\) as above, so that \( Pic(H_n)\) correspondingly is an inductive limit

$$\begin{aligned} Pic(H_n) = \lim _{\rightarrow } Pic(H({\mathcal {T}})) \end{aligned}$$

of the Picard groups \(Pic(H({\mathcal {T}}))\). Let us consider generators of this inductive limit. Such generators are the Picard groups \(Pic(H(\overline{\mathcal {T}}_\lambda ))\), attached to the Tannakian categories \(\overline{\mathcal {T}}_\lambda \) that are generated by \(X_\lambda \) and the normalized Berezin B. By definition, the Tannakian category \(\overline{\mathcal {T}}_\lambda \) only depends on the equivalence class \(\lambda /\!\sim \) of \(\lambda \) since \(\overline{\mathcal {T}}_{\lambda '} = \overline{\mathcal {T}}_{\lambda }\) for \(X_{\lambda '} = X_{\lambda } \otimes B\).

In the limit, the passage from \(H_\lambda \) to \(H(\overline{\mathcal {T}}_\lambda )\) allows a slicker description of the structure of the projective limit \(H_n\): Obviously there exists a canonical splitting

$$\begin{aligned} Pic(H(\overline{\mathcal {T}}_\lambda )) \cong Pic^0(H(\overline{\mathcal {T}}_\lambda )) \times {\mathbb {Z}} \, \end{aligned}$$

compatible with the splitting \( Pic(H_n) \cong Pic^0(H_n) \times {\mathbb {Z}} \) given in 13, in such a way that \(Pic^0(H_n)\) is generated by the images of the groups \(Pic^0(H(\overline{\mathcal {T}}_\lambda ))\) for \(\lambda \) ranging over the equivalence classes \(\lambda /\!\sim \).

11.3 \(Pic(\overline{\mathcal {T}}_{n})\) and the determinant

The elements of \(Pic({\overline{{\mathcal {T}}}}_n)\) are represented by indecomposable objects \(I \in \mathcal {T}_n^+\) with the property

$$\begin{aligned} I \otimes I^{\vee } \cong {\textbf{1}}\oplus \text { negligible}.\end{aligned}$$

Since \({{\,\textrm{sdim}\,}}(X_\lambda ) \ge 0\), we can define \(det(X_{\lambda }) = \Lambda ^{{{\,\textrm{sdim}\,}}(X_\lambda )}(X_\lambda ) \). Notice

$$\begin{aligned} \det (X_\lambda )\ = \ I_\lambda \ \oplus \ \text {negligible} \end{aligned}$$

is the sum of a unique indecomposable module \(I_\lambda \) in \(\mathcal {T}_n^+\) and a direct sum of negligible indecomposable modules in \(\mathcal {T}_n^+\). Furthermore \( I_\lambda ^* \cong I_\lambda \) and \( {{\,\textrm{sdim}\,}}(I_\lambda ) = 1\) holds, and if \(X_\lambda \) is selfdual, then \(I_\lambda \) is selfdual. In particular, \(det(X_\lambda )\) in \({ \mathcal {T}_n^+}\) has superdimension one, hence its image defines an invertible object of the representation category \({\overline{{\mathcal {T}}}}_n \sim Rep_k(H_n)\). By abuse of notation we also write \(det(X_\lambda ) \ \in \ Rep_k(H_n)\).

11.4 The NSD-case

According to the structure theorem Theorem 10.10 in the (NSD)-cases the group \(H_\lambda \) satisfies

$$\begin{aligned} SL(V_\lambda ) \subseteq H_\lambda \subseteq GL(V_\lambda ) \ldots \end{aligned}$$

Hence to determine \(H_\lambda \) it suffices to show that the restriction of the determinant \(\det : GL(V_\lambda )\rightarrow k^*\) to the \(H_\lambda \) is either trivial or surjective. In Theorem 14.3 we later show that \(\det : H_\lambda \rightarrow {{\mathbb {G}}}_m\) a a representation is represented by a power \(B^{\ell (\lambda )}\) of the twisted Berezin object B in \({{\mathcal {T}}_n}^+\). Hence \(H_\lambda \cong SL(V_\lambda )\) if and only if the integer \(\ell (\lambda )\) is zero, and \(H_\lambda \cong GL(V_\lambda )\) holds otherwise. In particular \(H_\lambda ^{ab}\) is the Tannaka group of the Tannaka category generated by \(B^{\ell (\lambda )}\) in the (NSD)-cases.

11.5 The SD-cases

We distinguish between the exceptional cases where \(V_{\lambda }|_{G_{\lambda }} \cong W \oplus W^{\vee }\) and the remaining cases where this does not happen. We call these the regular (SD)-cases since we conjecture that the exceptional (SD)-case do not occur.

In all (SD)-cases the group \(H_\lambda \) is a genuine subgroup of the similitude group \(G(V_\lambda ,\langle .,.\rangle )\). As such it inherits the similitude character \(\mu \) resp. the determinant character \(\det \) of \(G(V_\lambda ,\langle .,.\rangle )\) that are represented by \(B^r\) resp. \(B^{\ell (\lambda )}\). Recall from Sect. 6 that \(sign= \det /\mu ^n\) is a character of order two of the similitude group \(G(V_\lambda ,\langle .,.\rangle )\). Since the restriction of sign to \(H_\lambda \) is represented by \(B^{\ell (\lambda )- r}\), and the latter is either trivial or non-torsion. This implies that sign is trivial on \(H_\lambda \), hence \(H_\lambda \subset GSO(V_\lambda )\) resp. \(H_\lambda \subset GSp(V_\lambda )\) holds according to the parity \(\varepsilon (X_\lambda )\) of the pairing \(\langle .,.\rangle \).

11.6 The regular SD-cases

For the regular (SD)-cases the group \(H_\lambda \) satisfies \(H_\lambda \subseteq GO(V_\lambda )\) resp. \(H_\lambda \subseteq GSp(V_\lambda )\) according to the parity of the pairing \(\langle .,. \rangle \). This follows from 6 as well as the fact that the kernel of the characters \(\det \) and the similtude character \(\mu \) are the subgroup \(SO(V_\lambda )\) resp. \(Sp(V_\lambda )\). Both \(\mu =B^r\) and \(\det = B^{\ell (\lambda )}\) are represented by tensor powers powers of the twisted Berezin by theorem 14.3. Hence in the regular (SD)-cases Theorem 10.10 implies the following structure result: The group \(H_\lambda \) is isomorphic to \(SO(V_\lambda )\) resp. \(Sp(V_\lambda )\) if and only if \(\ell (\lambda )=0\), and it is isomorphic to \(GSO(V_\lambda )\) resp. \(GSp(V_\lambda )\) otherwise.

11.7 The exceptional (SD)-cases

For this exceptional situation we recall the following facts:

In the exceptional case we have shown that W is not isomorphic to \(W^\vee \) as a representation of \(G_\lambda \) (in other words we have \(m>2\)). By Schur’s lemma this implies that the restriction of the pairing \(<\cdot ,\cdot>\) must be trivial on the subspaces \(W \subset V_\lambda \) and \(W^\vee \subset V_\lambda \), and that these two subspaces are orthogonal to each other for the nondegenerate pairing \(\langle .,.\rangle \) on \(V_\lambda \). Hence W and \(W^\vee \) define an orthogonal pair of Lagrangian subspaces of \(V_\lambda \). In the exceptional cases the representation of \(G_\lambda \) on the vectorspace

$$\begin{aligned} V_{\lambda }|_{G_{\lambda }} \cong W \oplus W^{\vee } \end{aligned}$$

decomposes into two faithful nonisomorphic irreducible representations on the subspaces W and the subspace \(W^\vee \) such that the image of \(G_\lambda \) in GL(W) contains the perfect group SL(W). Furthermore \(H=H_\lambda \) (as well as \(H=H({{\mathcal {T}}}_\lambda )\)) preserves the unordered pair \(\{W,W^\vee \}\) of disjoint Lagrangian subspaces W and \(W^\vee \) and the pairing on \(V_\lambda \) (up to a similitude factor). Since they fix the Lagrangian decomposition \(V_\lambda = W \oplus W^\vee \) up to a permutation of the two subspaces, this induces a permutation character \(\chi _\lambda : H_\lambda \rightarrow \mu _2\) and similarly for \(H({{\mathcal {T}}}_\lambda )\). In the exceptional cases, by definition there exist elements w in H that \(\chi _\lambda (w)\ne 1\); let us fix such w. For exceptional \(\lambda \) we thus obtain an exact sequence

$$\begin{aligned} 0 \rightarrow {\tilde{H}} \rightarrow H \rightarrow \mathbb \mu _2 \rightarrow 0.\end{aligned}$$

The kernel \( {\tilde{H}}\) is a subgroup of the group G(W) of similitudes of \(V_\lambda \) that individually preserve the subspaces W and \(W^\vee \), and \({\tilde{H}}\) contains SL(W). In terms of a basis of W and a dual basis of \(W^\vee \) the elements of G(W) are of the blockdiagonal form \(g=diag(A,\lambda \cdot A^{-t})\) for \(A\in GL(W)\) and \(\lambda \in k^*\). In fact \(g\mapsto (A,\lambda )\) induces a group isomorphism \(G(W) \cong GL(W) \times {\mathbb {G}}_m\), such that the projection onto the second factor \({\mathbb {G}}_m\) induces the similitude character \(\mu : G(W)\rightarrow {\mathbb {G}}_m\). The determinant \(\det _W: GL(W)\rightarrow {\mathbb {G}}_m\) on the first factor induces an isomorphism

$$\begin{aligned} ({\det }_W,\mu ):\ G(W)/SL(W) \ \cong \ {\mathbb {G}}_m \times {\mathbb {G}}_m. \end{aligned}$$

The action of w by conjugation on G(W) preserves the subgroups SL(W) and \({\tilde{H}}\) of G(W) for either \(H=H_\lambda \) or \(H={\tilde{H}}(\overline{\mathcal {T}}_\lambda )\) such that the induced action on \({\mathbb {G}}_m \times {\mathbb {G}}_m \cong G(W)/SL(W)\) is inversion on the first factor and the identity on the second factor. This follows from \(\det _W(w g w^{-1}) = {\det }_W(g)^{-1}\) and \(\mu (w g w^{-1}) = \mu (g)\). The algebraic group H/SL(W), as a subgroup of G(W)/SL(W), therefore is a closed subgroup of \({\mathbb {G}}_m \times {\mathbb {G}}_m\). Since \({\tilde{H}}/SL(W)\) is contained in the first factor \({\mathbb {G}}_m\), the quotient \(Q=\tilde{H}/SL(W)\) is either \({\mathbb {G}}_m\) or a finite cyclic group. In the first case Q is the full commutator group of H/SL(W). But in both cases, the factor commutator group \(H^{ab}\) as a diagonalizable group is isomorphic to a direct product of a finite torsion group and a torus of rank \(\le 1\). Its rank is nonzero if and only if the similitude \(\mu : H\rightarrow {\mathbb {G}}_m\) is nontrivial and thus surjective, i.e. for the cases \(H=H(\overline{\mathcal {T}}_\lambda )\) resp. \(r(\lambda )\ne 0\) for \(H=H_\lambda \). Recall that the subgroup \(Pic^0(H)\) of \(Pic(H)=X^*(H^{ab})\) is the annihilator of a cocharacter \({\mathbb {G}}_m \rightarrow H\) that is induced by the embedding \(i: {\mathbb {G}}_m \cong H_1 \rightarrow H_n\) composed with the surjection \(H_n \twoheadrightarrow H\). Since \( \mu \circ i \) is surjective unless \(\mu : H\rightarrow {\mathbb {G}}_m\) is trivial, it easily follows that \(Pic^0(H)\) is the torsion subgroup \(Pic(H)_{tor}\) of Pic(H), and this group \(Pic^0(H)\) only depends on the equivalence \(\lambda /\!\sim \) of \(\lambda \) and is the same for \(H=H_\lambda \) and \(H(\overline{\mathcal {T}}_\lambda )\). We make this more explicit in the next section.

To study the equivalence class \(\lambda /\!\sim \), we consider normalised representatives \(\lambda \). Twisting \(\lambda \) by the ath power of the Berezin B, in the pairing \(\mu : X_\lambda \times X_\lambda \rightarrow B^r\) the character \(\mu =B^r\) changes to \(\mu \otimes B^{2a}=B^{r+2a}\). Hence one can choose a normalised representative \(\lambda \) in its class so that \(r=r(\lambda )\) is one or zero. In the first case \(r=1\) the Tannakian category \(\overline{\mathcal {T}}_\lambda \) is generated by \(X_\lambda \), hence \(H_\lambda = H (\overline{\mathcal {T}}_\lambda )\) contains the group Z of scalar homotheties of \(V_\lambda \). In the second case \(r=0\), the similitude character \(\mu \) of \(H_\lambda \) is trivial. Then it is easy to see that \(H (\overline{\mathcal {T}}_\lambda ) = Z \cdot H_\lambda \), and \(Z\cap H_\lambda = \{ \pm id_{V_\lambda }\}\). Hence \(H (\overline{\mathcal {T}}_\lambda ) = Z \cdot H_\lambda \) holds for any exceptional \(\lambda \). Since therefore \(H (\overline{\mathcal {T}}_\lambda )\) contains the group Z of all diagonal matrices, it is easy to see that we can modify w by an element of \({\tilde{H}}(\overline{\mathcal {T}}_\lambda )\) (without changing its conjugation action on S, but possibly the sign of \(\chi _\lambda (w)\)) such that \(w^2=id_{V_\lambda }\) holds. As a consequence, in the exceptional cases for \(Q=Q_\lambda \subseteq {\mathbb {G}}_m\) this implies

$$\begin{aligned} H^{ab}(\overline{\mathcal {T}}_\lambda ) \cong (Q/Q^2) \times \mu _2 \times {\mathbb {G}}_m. \end{aligned}$$

Notice \(Q/Q^2\) is trivial if Q is finite of odd order or \(Q\cong {\mathbb {G}}_m\), and it is isomorphic to \({\mathbb {Z}}/2{\mathbb {Z}}\) otherwise.

For \(H(\langle B \rangle )= {\mathbb {G}}_m\) the Tannakian subcategory \(\langle B^r \rangle \) generated by \(B^r\) gives rise to an r-fold covering \(H(\langle B \rangle ) \rightarrow H(\langle B^r \rangle )\), i.e. \(H(\langle B^r \rangle )\) is the quotient of \(H(\langle B \rangle )\cong {\mathbb {G}}_m\) by the unique cyclic subgroup of order r. In a similar vein we can recover \( H_\lambda \) from \(H(\overline{\mathcal {T}}_\lambda )\). This is seen as follows: The inclusion

$$\begin{aligned}H(\overline{\mathcal {T}}_\lambda )\ \hookrightarrow \ H_\lambda \times H(\langle B \rangle ) \,\end{aligned}$$

composed with both projections, defines surjective maps. Furthermore (ht) for \(h\in H_\lambda \) and \(t\in k^*\) is in \(H(\overline{\mathcal {T}}_\lambda )\) if and only if \(t^r=\mu (h)\) holds, where \(\mu =\mu _\lambda \) is the similitude character defined on \(H_\lambda \) and \(r=r(\lambda )\) is defined by \(\mu _\lambda = B^{r}\). Thus \(H(\overline{\mathcal {T}}_\lambda )\) is a fibre product of \(H_\lambda \) and \({\mathbb {G}}_m\) by \(\mu \) and the r-th power map, and \(H_\lambda \) is obtained from its r-fold covering group \(H(\overline{\mathcal {T}}_\lambda )\) as the quotient by the cyclic subgroup of order r that is contained in the second factor \( H(\langle B \rangle ) \cong {\mathbb {G}}_m\).

From the discussion above of the exceptional case we immediately derive the next two Lemmas 11.1 and 11.2.

Lemma 11.1

Suppose that the class \(\lambda /\! \sim \) is exceptional. Then either for the normalized representative \(\lambda \) the Tannaka group \(H_\lambda \) has trivial similitude character (if r is even), or (understood nonexclusively) for the normalized representative \(\lambda \) the group \(H(\overline{\mathcal {T}}_\lambda )\) is isomorphic as an algebraic group to \((GL(\dim (V_\lambda ))\cdot \mu _2) \times {\mathbb {G}}_m\), so it has two connected components and \(Pic^0(H(\overline{\mathcal {T}}_\lambda ))\) is isomorphic to \(\mu _2\).

Lemma 11.2

For exceptional classes \(\lambda /\!\sim \) the group \(Pic^0(H(\overline{\mathcal {T}}_\lambda )) = Pic^0(H_{\lambda })\) is a nontrivial two-torsion group of rank \(\le 2\).

Remark 11.3

The object I realizing the surjective projection \(\chi _\lambda : H_\lambda \rightarrow \mu _2 =\langle w\rangle \) corresponds to an element with the properties in Lemma E.1. Indeed, I appears as an indecomposable constituent of superdimension 1 in \(L(\lambda )\otimes L(\lambda )^{\vee }\) that is not isomorphic to the trivial representation. This follows immediately from the description of \(V_{\lambda } \cong Ind_{H_1}^{H}(W)\) as an induced representation.

We summarize the results of this section in the following theorem.

Theorem 11.4

In the (NSD) resp. in the regular (SD)-cases the groups \(H_\lambda \) are isomorphic to \(GL(V_\lambda )\), resp. to \(Kern(sign: G(V_\lambda ,\langle .,.\rangle ) \rightarrow \mu _2)\), i.e. \(GSO(V_\lambda )\) or \(GSp(V_\lambda )\), if \(\ell (\lambda )\ne 0\), and to \(SL(V_\lambda )\), resp. to \(SO(V_\lambda )\) or \(Sp(V_\lambda )\) if \(\ell (\lambda )\ne 0\). So in these cases \(H_\lambda \) is connected and \(H_\lambda /G_\lambda \cong H_\lambda ^{ab}\). In the exceptional (SD)-cases the groups \(H_\lambda \) are not connected and there exists a nontrivial homomorphism \(\chi _\lambda : H_\lambda \rightarrow \mu _2\). The kernel \({\tilde{H}}_\lambda \) is a subgroup that contains SL(W) such that \(S={\tilde{H}}_\lambda /SL(W)\) becomes an algebraic subgroup of \(\mathbb {G}_m^2\) of dimension one or two, depending on whether \(\ell (\lambda )=0\) or \(\ell (\lambda )\ne 0\) (or, equivalently, \(r = 0\) or \(r \ne 0\)).

11.8 The full Tannaka group

Now consider the full tannakian category \(\bar{\mathcal {T}}_n\). Note that

$$\begin{aligned} H_n \hookrightarrow \prod _{\lambda /\sim } H_{\lambda }.\end{aligned}$$

As in the introduction we fix an isomorphism \(\mu _B\) between the Tannaka group \(H_B:=H_\lambda \) of the twisted Berezin and the multiplicative group \({{\mathbb {G}}}_m\). Recall further that \(\mu _{\lambda }\) was given by \(\det _{\lambda }\) in the (NSD)-case and the similitude character in the (SD)-case. Then \(H_n\) is a subgroup of the infinite fibre product defined by the elements \(h=(h_\lambda )_{\lambda /\sim }\) in \( \prod _{\lambda / \sim \ \in \, Y^+_0(n)} H_\lambda \) that satisfy \( \mu _\lambda (h_\lambda ) = \mu _B(h_B) \) for all \(\lambda \). The induced fibre homomorphism \(\mu : H_n \rightarrow {\mathbb {G}}_m\) defined by \(\mu _B\) is surjective. Its kernel \({\tilde{H}}_n= Ker(\mu : H_n \rightarrow {\mathbb {G}}_m)\) contains the projective limit \(G_n\) of the derived groups of the connected component of the \(H_\lambda \) as a normal subgroup. For \(H_{n}^{ab}\) determined by \(X^*(H_n^{ab})= Pic(H_n) = Pic^0(H_n) \times {\mathbb {Z}} \) and \(X^*({\tilde{H}}_n^{ab}) \cong Pic^0(H_n)\), as explained above, our computations imply

Corollary 11.5

\(H_n^{ab}\) is the factor commutator group of \(H_n/G_n\), and the commutator subgroup M of \(H_n/G_n\) is a pro-diagonalizable group.

Proof

Although only left exact, projective limits preserve exactness if the so called Mittag-Leffler conditions are satisfied. For a projective limit of algebraic groups as in our case, these conditions hold. Thus the projective limit \(H_n\) contains the projective limit \(G_n\) as a normal subgroup, and the quotient group \(H_n/G_n\) admits the projective limit \(H_n^{ab}\) as quotient group such that the remaining kernel M in \(H_n/G_n\) is the projective limit of the diagonalizable algebraic groups \(\prod _{\lambda /\!\sim \in I} (Q_\lambda )^2\) extended over all finite subsets I of the set of exceptional classes \(\lambda /\!\sim \). Here the \(Q_\lambda \) are the diagonalizable groups from Sect. 11.7. \(\square \)

Remark If we knew that the characters \(\chi _\lambda \) of \(H_\lambda \) (attached to the exceptional classes \(\lambda /\!\sim \)) were linear independent on \(H_n\), this would easily imply that \(Pic^0(H_n)\) is a two torsion group. To prove linear independency of the \(\chi _\lambda \), by Schur’s lemma amounts to show that every finite tensor product \(\bigotimes _\lambda X_\lambda \) of simple objects attached to exceptional classes \(\lambda /\!\sim \) is a simple object. However, in absence of such stronger results, we have to be content with the following weaker statement.

Corollary 11.6

\(Pic^0(H_n)\) is a 2-power torsion group (possibly infinite), and this group is nontrivial if and only if exceptional classes \(\lambda /\!\sim \) exist.

Proof

By our previous computations it suffices to consider the contribution of the exceptional classes. Let \(\lambda _i, i\in I\), denote a finite set I of exceptional weights \(\lambda _i\). Let \(H_I\) denote the Tannaka group attached to the Tannakian category \(\overline{\mathcal {T}}_I\) that is generated by B and the objects \(X_{\lambda _i}\) for \(i\in I\). Notice \(H_I \subset \prod _{i\in I} H(\overline{\mathcal {T}}_{\lambda _i})\), and \(\prod _{i\in T}\mu _i \) defines a character \(\mu _I\) on \(H_I\), factorizing over the quotient group \(H_I^{ab}\). To prove our assertion it is enough to show that the kernel of \(\mu _I: H_I^{ab}\rightarrow {\mathbb {G}}_m\) is a finite torsion group annihilated by \(2^{\# I}\). For each exceptional class \(\lambda /\!\sim \), the group \(H(\overline{\mathcal {T}}_\lambda )\) contains in its commutator group the perfect group \(SL(W_{\lambda _i})\). Hence we may replace \(H_I\) by its image D in \(\prod _{i\in I} D_i \), for \(D_i =Kern(\mu _i: H_{\lambda _i} \rightarrow {\mathbb {G}}_m)/SL(W_{\lambda _i})\), as is easy to see. Each group \(D_i\) for \(i\in I\) is a generalized dihedral group, i.e. a semidirect product

$$\begin{aligned} D_i = S_i \cdot \langle w_i \rangle , \quad w_i^2 = 1 \end{aligned}$$

for an algebraic group \(S_i\) of multiplicative type, such that the involution \(w_i\) acts on \(S_i\) by conjugation: \(w_i s_i w_i = s_i^{-1}\). As a subgroup of \(\prod _{i\in I} D_i\) the group D admits an exact sequence

$$\begin{aligned} 0 \rightarrow S \rightarrow D \rightarrow \prod _{i\in I} \ \langle w_i \rangle \, \end{aligned}$$

where S is a subgroup of \(\prod _{i\in I} S_i\). The image W of D is a subvectorspace of \(\prod _{i\in I} \langle w_i \rangle \), considered as a vector space over the prime field of characteristic 2. The latter may be identified with the power set of I, so that the elements of W may be considered as subsets \(J\subseteq I\) of I. Let \(\chi _J: I \rightarrow {\mathbb {Z}}\) thus denote the characteristic function of \(J\subset I\) in this sense. Any \({\mathbb {Z}}\)-valued function f on I can be considered as an endomorphism of \(\prod _{i\in I} S_i\) via \(f((s_i)_{i\in I})= (f(i)\cdot s_i)_{i\in I}\). For \(J\in W\) with representative \(w\in D\), the commutator of w with elements of S induces an endomorphism of S. Considered as a subgroup of \(\prod _{i\in I} S_i\), this is the endomorphism attached to the integer-valued function \( f= 2 \chi _J\) on I. By definition this endomorphism of \(\prod _{i\in I} S_i\) preserves the subgroup S. Therefore any element of S in the image of the endomorphism \(\sum _{J\in W} 2 \chi _J\) of S is contained in the commutator group of D. By Lemma 11.7, the factor commutator group of D thus is a finite group annihilated by \(2^{\dim (W)}\), so also by \(2^{\# I}\). Notice, Lemma 11.7 can be applied since \(\bigcup _{J\in W} = I\) is satisfied, as follows from the fact that all projections \(H_I \mapsto H_{\lambda _i}\), \(i\in I\) are surjective. \(\square \)

Lemma 11.7

If \(\ \bigcup _{J\in W} J = I\) holds, then as endomorphism of S we obtain:

$$\begin{aligned} \sum _{J\in W} 2 \chi _J \ =\ 2^{\dim (W)} \cdot id_S. \end{aligned}$$

Proof

For a basis \(J_1,\ldots ,J_d\) of the \({{\mathbb {F}}}_2\)-vectorspace W any \(J\in W\) can be uniquely written in the form \(J= \sum _{\nu =1}^d a_\nu J_\nu \) for \(a_\nu \in \{0,1\}\). For \(i\in I\) fixed, we may reorder the basis such that \(i\in J_1,\ldots ,J_k\) and \(i\notin J_{k+1},\ldots ,J_d\). Notice \(k=k(i)\ge 1\) holds for all \(i\in I\) by \(\ \bigcup _{J\in W} J = I\). Hence \(\chi _J(i) =1\) if and only if \(a_1 +\cdots + a_k\) is odd. Since \(k\ge 1\), therefore there exist precisely \(\frac{1}{2} 2^{k} \cdot 2^{d-k} = 2^{\dim (W)-1}\) vectors J in W with \(\chi _J(i) =1\). The sum of the endomorphisms \(2\chi _J\) of S, for \(J\in W\), therefore gives \(2^{\dim (W)}\) times the identity of S. \(\square \)

12 A conjectural picture

12.1 Equivalent conjectures

We conjecture that all SD-cases are regular:

Conjecture 12.1

\(G_\lambda = SL(V_\lambda )\) resp. \(G_\lambda = SO(V_\lambda )\) resp. \(G_\lambda = Sp(V_\lambda )\) according to whether \(X_\lambda \) satisfies (NSD) respectively (SD) with either \(X_\lambda \) being even respectively odd.

Lemma 12.2

The following are equivalent:

  1. (1)

    Conjecture 12.1 holds.

  2. (2)

    The module \(I = I_{\lambda }\) is trivial.

  3. (3)

    Any invertible object I in \(\overline{\mathcal {T}}_{n}\) is represented in \(\mathcal {T}_n^+\) by a power of the Berezin determinant.

  4. (4)

    Conjecture 12.8 holds, i.e. every special module is trivial.

Proof

By Theorem 11.4 the Picard group is generated by Berezin powers if there are no exceptional SD-cases, i.e. \(I \cong {\textbf{1}}\). \(\square \)

We discuss a possible approach to proving \(I \cong {\textbf{1}}\) in “Appendix E”.

12.2 An element of the Picard group of \({{\mathcal {T}}}_{2\vert 2}/{{\mathcal {N}}}\)

The assertion (3) from Lemma 12.2 cannot hold for \(\mathcal {T}_n\) instead of \(\mathcal {T}_n^+\). In the category \({{\mathcal {T}}}_{2}\) consider the indecomposable representation I, defined uniquely up to isomorphism by the nontrivial extension

This extension is realized in the Kac module of the trivial representation

$$\begin{aligned} K({\textbf{1}}) = \begin{pmatrix} {\textbf{1}}\\ [0,-1] \\ Ber^{-2} \end{pmatrix} \end{aligned}$$

Since \({{\,\textrm{sdim}\,}}(K({\textbf{1}})) = 0\), I has superdimension \(-1\), so its parity shift has \({{\,\textrm{sdim}\,}}(\Pi I) = 1\). Clearly \(I\otimes I^{\vee } \cong {\textbf{1}}\oplus N\) with \({{\,\textrm{sdim}\,}}(N) = 0\). We show that N is negligible. For this we can pass to the homotopy category \(\mathcal {H}oT\) attached to the lower parabolic \(\mathfrak {p}^-\) [38]. The kernel of the homotopy functor \(\pi : \mathcal {T}_{2} \rightarrow \mathcal {H}oT\) consists of the Kac objects, i.e. those modules with a filtration by Kac-modules. The short exact sequence for \(K({\textbf{1}})\) induces in the tensor triangulated category \(\mathcal {H}oT\) an exact triangle

Since \(K({\textbf{1}})\) is zero in \(\mathcal {H}oT\), this gives the identification \(I \cong Ber^{-2}[1]\). In particular

$$\begin{aligned} I \otimes I^{\vee } \cong Ber^{-2}[1] \otimes (Ber^{-2})^{\vee }[1] \cong {\textbf{1}}[2].\end{aligned}$$

Now suppose N would not be negligible. Then \(N \cong N_1 \oplus N_2\) with \({{\,\textrm{sdim}\,}}(N_1) \ne 0\). In particular \(\pi (N_1) \ne 0\), a contradiction since \(\pi \) is a symmetric monoidal functor and \({\textbf{1}}[2]\) is indecomposable.

Corollary 12.3

\(I\otimes I^\vee =\textbf{1}\oplus N\) and N is negligible. Therefore I and its parity shift \(\Pi I\) define elements of the Picard group of \({{\mathcal {T}}}_{2}/{{\mathcal {N}}}\).

12.3 An application

The conjectural structure theorem would have the following consequences.

Corollary 12.4

For given \(L=L(\lambda )\) in \({{\mathcal {T}}}_n\) and \(r\in {\mathbb {Z}}\) there can exist at most one summand M in \(L \otimes (Ber^r \otimes L^\vee )\) with the property \({{\,\textrm{sdim}\,}}(M)=\pm 1\). If it exists then \(M \cong Ber^r\).

Proof of the corollary

We can assume that L is maximal atypical. Then \(\textbf{1}\) is a direct summand of \(L\otimes L^\vee \) and hence \(Ber^r\) is a direct summand of \(L \otimes (Ber^r \otimes L^\vee )\). Hence it suffices to show that \(\textbf{1}\) is the unique summand M of \(L \otimes L^\vee \) with \({{\,\textrm{sdim}\,}}(M)=\pm 1\). Equivently it suffices to show that \(V_\lambda \otimes V_\lambda ^\vee \) contains no one-dimensional summand except \(\textbf{1}\). This now follows from conjecture 12.1 using the well known fact that \(st \otimes st^\vee \) for the standard representation st of SL(V), SO(V), Sp(V) contains only one summand of dimension 1. \(\square \)

12.4 The Tannaka groups \(H_{\lambda }\) revisited

The following theorem is an immediate consequence of theorem 11.4.

Theorem 12.5

Assuming Conjecture 12.1, the Tannaka groups \(H_{\lambda }\) of \(X_{\lambda }\) are the following:

  1. (1)

    NSD non-basic: \(H_{\lambda } = GL(V_{\lambda })\).

  2. (2)

    NSD basic: \(H_{\lambda } = SL(V_{\lambda })\).

  3. (3)

    SD, proper selfdual, \({{\,\textrm{sdim}\,}}(L(\lambda _{basic})) > 0\): \(H_{\lambda } = SO(V_{\lambda })\).

  4. (4)

    SD, proper selfdual, \({{\,\textrm{sdim}\,}}(L(\lambda _{basic})) < 0\): \(H_{\lambda } = Sp(V_{\lambda })\).

  5. (5)

    SD, weakly selfdual, \({{\,\textrm{sdim}\,}}(L(\lambda _{basic})) > 0\): \(H_{\lambda } = GSO(V_{\lambda })\).

  6. (6)

    SD, weakly selfdual, \({{\,\textrm{sdim}\,}}(L(\lambda _{basic})) < 0\): \(H_{\lambda } = GSp(V_{\lambda })\).

In each case the representation \(V_{\lambda }\) of \(H_{\lambda }\) coming from \(X_{\lambda }\) corresponds to the standard representation. In the GL case the determinant comes from a (nontrivial) Berezin power.

Note that a basic representation of SD type always satisfies \(L \cong L^{\vee }\). In the (SD) case \(\ell (\lambda ) = 0\) if and only if \(L(\lambda ) \simeq L(\lambda )^{\vee }\). In the (NSD)-case \(\ell (\lambda ) = 0\) if and only if \(\lambda \) is basic.

12.5 Special modules

We discuss a conjecture which would show \(I \cong {\textbf{1}}\).

Definition 12.6

An indecomposable module V in \(\mathcal {T}_n^+\) with \({{\,\textrm{sdim}\,}}(V)=1\) will be called special, if \(V^*\cong V\) and \(H^0(V)\) contains \(\textbf{1}\) as a direct summand.

For special modules

$$\begin{aligned} DS(V) \cong \textbf{1} \oplus N \end{aligned}$$

holds for some negligible module N, since \({{\,\textrm{sdim}\,}}(DS(V))={{\,\textrm{sdim}\,}}(V)\). This also implies

$$\begin{aligned} H_D(V) = \textbf{1} \oplus N. \end{aligned}$$

Lemma 12.7

Suppose \(V\cong V^* \cong V^\vee \) and \(DS(V) \cong \textbf{1} \oplus N\) holds for some negligible module N. Then V is special.

Proof

The assumptions imply that there exists a unique integer \(\nu \) for which \(H^\nu (V)\) is not a negligible module. Since \(H^\nu (V)^\vee \cong H^{-\nu }(V^\vee )\), the assumption \(V \cong V^\vee \) implies \(\nu =0\). Hence \(H^0(V)=\textbf{1} \oplus N\) for some negligible N. \(\square \)

Conjecture 12.8

Up to a parity shift, any special module V in \(\mathcal {T}_n^+\) is isomorphic to the trivial module \(\textbf{1}\).

13 The Picard group of \(\overline{\mathcal {T}}_{n}\)

We study the determinant \(\det (X_{\lambda })\) in this section.

13.1 The invariant \(\ell (\lambda )\)

As one easily shows, for any object X of \({{\mathcal {T}}}_n\)

$$\begin{aligned} \det (B^m \otimes X) = B^{m \cdot {{\,\textrm{sdim}\,}}(X)}\otimes \det (X). \end{aligned}$$

Hence to determine \(I_\lambda \) we may assume \(\lambda _n=0\). So let us assume this for the moment. Then, for a maximal atypical weight \(\lambda \) with the property \(\lambda _n=0\), let \(S_1,\ldots ,S_k\) denote its corresponding sectors, from left to right. If \(i=1,\ldots ,k-1\) let \(d_i= dist(S_i, S_{i+1})\) denote the distances between these sectors and \(r(S_i)\) denotes the rank of \(S_i\), then \(\sum _{i=1}^k r(S_i)= n\). Furthermore \(d=\sum _{i=1}^k d_i =0\) holds if and only if the weight \(\lambda \) is a basic weight. Recall, if we translate \(S_2\) by shifting it \(d_1\) times to the left, then shift \(S_3\) translating it \(d_1+d_2\) to left and so on, we obtain a basic weight. This basic weight is called the basic weight associated to \(\lambda \). The weighted total number of shifts necessary to obtain this associated basic weight by definition is the integer

$$\begin{aligned} \ell (\lambda ):= \sum _{i=1}^k {{\,\textrm{sdim}\,}}(X_{\lambda _i}) \cdot \left( \sum _{j<i} d_j\right) \ \end{aligned}$$

where \(L(\lambda _i)\in {{\mathcal {R}}}_{n-1}\) denote the irreducible representations associated to the derivatives \(S_1\ldots \partial S_i\ldots S_k\). By [51] [36, Section 16] \({{\,\textrm{sdim}\,}}(X_{\lambda _i}) = \frac{r_i}{n} \cdot {{\,\textrm{sdim}\,}}(X_\lambda )\) holds for \(r_\nu =r(S_\nu )\), which allows to rewrite this in the form

$$\begin{aligned} \ell (\lambda ) = n^{-1} {{\,\textrm{sdim}\,}}(X_{\lambda }) D(\lambda ) \, \end{aligned}$$

where \(D(\lambda )\) is the total number of left moves needed to shift the support of the plot \(\lambda \) into the support of the associated basic plot \(\lambda _{basic}\), i.e. the integer

$$\begin{aligned} D(\lambda ) \,= \ \sum _{\nu =1}^k \ r_\nu \cdot \left( \sum _{\mu <\nu } d_\mu \right) . \end{aligned}$$

Now, to remove our temporary assumption \(\lambda _n=0\) and hence to make the formulas above true unconditionally, we have to introduce the additional terms \(d_0= \lambda _n\) (for \(\mu =0\)) in the formulas above. For further details on this see [36, Section 25]. We remark that in the following we also write D(L) instead of \(D(\lambda )\) for the irreducible representations \(L = L(\lambda )\) and similarly \(\ell (L)\) instead of \(L(\lambda )\).

13.2 \(Pic^0\)

We return to indecomposable objects \(I \in \mathcal {T}_n^+\) representing invertible objects of \({\overline{{\mathcal {T}}}}_n\).

Since \(I \otimes I^\vee \cong \textbf{1} \oplus \) negligible objects, we obtain

$$\begin{aligned} \omega (I,t) \omega (I^\vee ,t) = \omega (I \otimes I^\vee ,t) = 1.\end{aligned}$$

Indeed, the functor \(\omega \) annihilates negligible objects. For the Laurent polynamial \(\omega (I,t)\) this now implies

$$\begin{aligned} \omega (I,t) = t^{\nu }\end{aligned}$$

for some integer \(\nu \in {\mathbb {Z}}\) which defines the degree \(\nu (I) = \nu \). Obviously this degree \(\nu (I)\) induces a homomorphism \(Pic({{\mathcal {R}}}_n) \rightarrow {\mathbb {Z}}\) of groups by \(I \mapsto \nu = \nu (I) \in {\mathbb {Z}}\) and gives an exact sequence

with kernel \(Pic^0({\overline{{\mathcal {T}}}}_n)\). Clearly \(\nu (B) = n\), hence the next lemma follows.

Lemma 13.1

The intersection of \(Pic^0({\overline{{\mathcal {T}}}}_n)\) with the subgroup generated by the normalized Berezin B is trivial.

Lemma 13.2

For any irreducible object X in \(\mathcal {T}_n^+\) the invertible element \(\det (X) \in {{\mathcal {T}}}_n\) has the property

$$\begin{aligned} \nu (\det (X)) = {{\,\textrm{sdim}\,}}(X)\cdot D(X) = \ell (X) \cdot n.\end{aligned}$$

In particular, the image of the homomorphism \(\nu \) contains \(n \cdot {\mathbb {Z}}\).

Proof

Fix some \(X=X_\lambda \). We can assume \(\lambda \) to be maximally atypical. The functor \(\omega : {{\mathcal {T}}}_n^+ \rightarrow gr{-}vec_k\) is a tensor functor. Hence \(\nu (\det (X)) = \nu (\det (\omega (X))\). Hence

$$\begin{aligned} \nu (\det (X)) = \sum _i i \cdot a_i \ \ \ \ \ \ \ \ (*) \end{aligned}$$

for \(\omega (X,t) = \sum _i a_i t^i\). By [36, Lemma 25.2] \(\omega (X,t^{-1}) = t^{-2D(\lambda )} \omega (X,t)\) and hence \(\omega (X_{basic},t) = \omega (X_{basic},t^{-1})\), the latter because of \(D(X_{basic}) = 0\). So \(a_i =a_{-i}\) holds for basic X, and formula (*) implies \(\nu (\det (X_{basic})) = 0\). From \(\omega (X,t) = t^{D(X)} \omega (X_{basic},t)\) and \({{\,\textrm{sdim}\,}}(X_{basic}) = {{\,\textrm{sdim}\,}}(X)\), again by (*) we therefore obtain

$$\begin{aligned} \sum _i i \, a_i = \sum _i i \, a_{basic,i} + D(\lambda )\cdot \sum _i a_i = D(\lambda ) {{\,\textrm{sdim}\,}}(X).\end{aligned}$$

Note that \(X\in \mathcal {T}_n^+\) has superdimension \(\ge 0\), hence \(\omega (X,1) = \sum a_i \) is the superdimension of X (not only up to a sign). \(\square \)

Since \(\omega (L(\lambda ),t)t^{-D(\lambda )}\) is invariant under \(t\mapsto t^{-1}\) for irreducible \(L=L(\lambda )\), we also obtain

Corollary 13.3

\(d \ log(\omega (L,t))|_{t=1} = D(L)\).

Corollary 13.4

We have \(\det (X) \otimes B^{-\ell (X)} \in Pic^0(\overline{\mathcal {T}}_n^+)\), i.e.

$$\begin{aligned} \det (X) \in Pic^0(\overline{\mathcal {T}}_n) \times B^{{\mathbb {Z}}} \end{aligned}$$

for irreducible \(X \in \mathcal {T}_n^+\).

Example 13.5

For GL(2|2) we obtained (up to parity shifts) in [37] the formula \(S^i \otimes S^i = Ber^{i-1} \oplus M\) for some module M of superdimension 3. Since \({{\,\textrm{sdim}\,}}(S^i) = 2\), \(\det (S^i) = Ber^{i-1} \oplus \text { negligible}\). Indeed for \(S^i\) we obtain \(\ell ([i,0]) = r_1 d_0 + r_2 d_1\) where \(r_i\) denotes the rank of the i-th sector. Clearly \(r_1 = r_2 = 1\) and \(d_0 = 0\) and \(d_1 = i-1\), hence \(\ell ([i,0]) = i-1\).

14 The determinant of an irreducible representation

We now compute the determinant of irreducible representations. The computation uses the passage to the stable category with its triangulated structure. This determinant calculation determines in all regular cases (along with the results of Sect. 11) the full Picard group of \(H_{\lambda }\).

14.1 The full even category \({\mathcal {T}}^{ev}\)

Since a svectorspace is the sum of an even and odd subspace, we have \(svec_k = vec_k \oplus \Pi (vec_k)\) as a decomposition of abelian categories. We say, an object X of \({\mathcal {T}}\) is even resp. odd if \(\omega (X)\) is in \(vec_k\) resp. \(\Pi (vec_k)\). In terms of the Hilbert polynomial \(\omega (X,t)\) defined in Sect. 4.2 even means that all t-powers are even. Let \({\mathcal {T}}^{ev}\) and \(\mathcal {T}^{odd}\) denote the corresponding full subcategories. In [36, Section 24] it is shown that simple objects in \({\mathcal {T}}\) are always even or odd, hence

$$\begin{aligned} \mathcal {T}^+ \subset \mathcal {T}^{ev}.\end{aligned}$$

Although \(D^n\) is not an exact functor, exact sequences in \({\mathcal {T}}\) become exact hexagons in \({{\mathcal {T}}}_0\). [Sometimes it is useful that in a certain sense we need not distinguish between \(D^{n-1}\) and \(D^n\) since \(D: \overline{{\mathcal {T}}}_1 \rightarrow {{\mathcal {T}}}_0\) is faithful. This refines the notion of even/odd if the \(D^{n-1}\)-image is semisimple and even resp. odd.] Extensions of even (odd) objects in \({\mathcal {T}}\) are even (odd) objects in \({\mathcal {T}}\). Obviously \({\mathcal {T}}^{ev}\), as a full karoubian subcategory of \({\mathcal {T}}\), is closed under extensions, retracts and tensor products and Tannaka duals. In particular we have stability with respect to Schur functors.

We consider now the semisimplification of \(\mathcal {T}^{ev}\) with the same method as in Sect.  5. Here we however use the Dirac tensor functor D (see Sect. 4.3) instead of DS (note they agree on \(\mathcal {T}^+\)). Iterated n times it factorizes over the additive quotient category \({{\mathcal {T}}}^{ev} \rightarrow {{\mathcal {A}}}\) defined by dividing through the ideal of all morphisms that factorize over null objects (objects whose indecomposable summands have superdimension zero). i.e. \(D^n: {{\mathcal {T}}}^{ev}_n \rightarrow {{\mathcal {T}}}_0^{ev}\). We define

$$\begin{aligned} \omega = D^n \circ s \ \end{aligned}$$

via a section s of the semisimplification functor \(\nu : {{\mathcal {A}}} \rightarrow \overline{{\mathcal {T}}}^{ev}\), i.e. \(\nu \circ s = id_{\overline{{\mathcal {T}}}}\) whose existence follows from [2]. The k-linear tensor functor \(\omega \) is an exact functor since \(\overline{{\mathcal {T}}}^{ev} \) is semisimple. Thus it defines a superfibre functor \(\omega : \overline{{\mathcal {T}}}^{ev} \rightarrow svec_k\) with values in \(vec_k\).

Since for indecomposable objects X of \({\mathcal {T}}^{ev}\) the space \(\omega (X)\) has dimension \(sdim(X)>0\) (unless \(sdim(X)=0\) and X is negligible), the quotient \(\overline{{\mathcal {T}}}^{ev}\) defines a Tannakian category. Hence, as a tensor category \(\overline{{\mathcal {T}}}^{ev}\) is equivalent to \(Rep_k(H^{ev},\varepsilon )\) for some affine (pro)reductive group scheme \(H^{ev}\) over k.

Remark 14.1

All in all we attached to the category \(\mathcal {T}\) three different semisimple supertannakian categories:

$$\begin{aligned} \mathcal {T}^+/\mathcal {N}, \ \mathcal {T}^{ev}/\mathcal {N}, \ \mathcal {T}/\mathcal {N}.\end{aligned}$$

The inclusion \(\mathcal {T}^{+} \subset \mathcal {T}^{ev}\) is strict. Already for GL(1|1), \(\mathcal {T}^{ev}\) contains ZigZag modules of length \(2\,m+1\) for \(m \in \mathbb {N}\) [34] which have nonvanishing superdimension. In fact for GL(1|1) the quotients \(\mathcal {T}^{ev}/\mathcal {N}\) and \(\mathcal {T}/\mathcal {N}\) are isomorphic. The relationship betweeen \(\mathcal {T}\) and \(\mathcal {T}^{ev}\) and their corresponding semisimple quotients is unclear as it is not obvious how to find indecomposable objects of nonvanishing superdimension which are mixed (i.e. neither even nor odd).

14.2 The stable category

Recall that \(\mathcal {T}_n = {{\mathcal {T}}}\) is a k-linear tensor category and as an abelian category it is a Frobenius category. Associated to a Frobenius category \({{\mathcal {T}}}\) one defines its stable category \({{\mathcal {K}}}\) as a quotient category [33]. For the quotient functor

$$\begin{aligned} \alpha : {{\mathcal {T}}}\rightarrow {{\mathcal {K}}} \end{aligned}$$

the objects of \({{\mathcal {K}}}\) are those of \({{\mathcal {T}}}\), but morphisms are equivalence classes of morphisms in \({{\mathcal {T}}}\). Two morphisms become equivalent if their difference is a morphism that factorizes over a projective module in \({{\mathcal {T}}}\). \({{\mathcal {K}}}\) is a triangulated category with a suspension functor \(S(X)=X[1]\) such that \(Ext^i_{{\mathcal {T}}}(X,Y)\cong Hom_{{\mathcal {K}}}(X,Y[i])\) holds for all \(i\ge 0\) and \(\alpha \) is a tensor functor. The \(\alpha \)-image of an exact sequence in \({\mathcal {T}}\) induces a distinguished triangle in \({\mathcal {K}}\). Any distinguished triangle in \({\mathcal {K}}\) is isomorphic as a triangle to the \(\alpha \)-image of an exact sequence in \({{\mathcal {T}}}\) [33].

Let \({\mathcal {K}}^{ev}, \mathcal {K}^{odd}\) denote the corresponding full subcategories of \(\mathcal {K}\) corresponding to \(\mathcal {T}^{ev}, \mathcal {T}^{odd}\). Similarly to the \(\mathcal {T}\)-case under the Dirac functor D exact triangles in \({\mathcal {K}}\) become exact triangles in \(svec_k\). In \({\mathcal {K}}\) the following holds: If \(X\rightarrow Y\rightarrow Z\rightarrow \) is a distiguished triangle and X and Z are in \({\mathcal {K}}^{ev}\), then also Y. This follows since \(\omega =D^n\) induces an exact hexagon from each exact sequence in \({\mathcal {T}}\) as in [36, Lemma 2.1] via \(\omega =\omega ^+\oplus \omega ^-\).

14.3 Determinants

Recall \({{\,\textrm{sdim}\,}}(X) \ge 0\) for \(X \in \mathcal {T}^{ev}\). Hence \(\det (X) = \Lambda ^{{{\,\textrm{sdim}\,}}(X)}(X)\) is defined so that \( \Lambda ^{{{\,\textrm{sdim}\,}}(X)+1}(X)\) is negligible. Here \(L=\det (X)\) by definition is \({{\textbf{1}}}\) if \(sdim(X)=0\). The image of \(\det (X)\) under the functor \(\omega =D^n\) defines an invertible object in the Tannakian subcategory generated by \(\omega (X)\) in \(\overline{{\mathcal {T}}}^{ev}\).

For exact sequences \(0\rightarrow X\rightarrow Y\rightarrow Z \rightarrow 0\) in \({{\mathcal {T}}}^{ev}\) we have \(\det (Y)\cong \det (X)\otimes \det (Z)\) in \({{\mathcal {T}}}^{ev}\). The analogous assertion holds for a distinguished triangle \(0\rightarrow X\rightarrow Y\rightarrow Z \rightarrow \) with \(X,Y,Z\in {{\mathcal {K}}}^{ev}\), simply by lifting this to an exact sequence \(0\rightarrow X'\rightarrow Y'\rightarrow Z'\rightarrow 0\) in \({\mathcal {T}}^{ev}\). For this notice, if \(X,X'\) in \({\mathcal {T}}^{ev}\) become isomorphic in \({\mathcal {K}}^{ev}\), then \(\det (X)\) and \(\det (X')\) become isomorphic on \({\mathcal {K}}^{ev}\). In particular, their images in the Tannakian representation category \(\overline{{\mathcal {T}}}^{ev}\) of the reductive group \(H^{ev}\) become isomorphic.

Now consider the following special situation, where I in \({\mathcal {T}}\) has a filtration of length 3 whose graded pieces are a submodule S, a middle layer M and the quotient \(T=I/M\) on top.

Lemma 14.2

Suppose ST are in \({\mathcal {T}}^{odd}\) and I is in \({\mathcal {T}}^{ev}\). Then MS[1] and \(T[-1]\) are in \({\mathcal {K}}^{ev}\) and the following holds in \({\mathcal {K}}\) up to negligible summands:

$$\begin{aligned} \det (M) \cong \det (I)\otimes \det (S[1])\otimes \det (T[-1]). \end{aligned}$$

Proof

If S is odd/even in the stable category, then S[1] is even/odd in the stable category, and conversely. Indeed S[1] is represented by a quotient P/S in the representation category for a suitable projective/injective module P. To the exact sequence \(0\rightarrow S \rightarrow P\rightarrow P/S\rightarrow 0\) the functor \(\omega =D^n\) attaches an exact hexagon for \(\omega = \omega ^+ \oplus \omega ^-\) which immediately implies \(\omega ^\pm (S)\cong \omega ^{\mp }(P/S)\) in \({\mathcal {T}}\) and then \({\mathcal {K}}\). From this the above assertion follows. We now have two distinguished triangles \(I \rightarrow Y \rightarrow S[1] \rightarrow \) for suitable Y and \(T[-1] \rightarrow M \rightarrow Y \rightarrow \) in \({\mathcal {K}}\). Here S[1], \(T[-1]\) and I are in \({{\mathcal {K}}}^{ev}\) by our assumptions. Therefore Y and then also M are in \({{\mathcal {K}}}^{ev}\). All objects S[1], \(T[-1]\), M, I and Y are represented by objects U, V, \(M'\), \(I'\), and \(Y''\) in \({\mathcal {T}}^{ev}\) so that there are exact sequences \(0 \rightarrow I' \rightarrow Y' \rightarrow U\rightarrow 0\) and \(0\rightarrow V \rightarrow M' \rightarrow Y'' \rightarrow 0\) in \({\mathcal {T}}\). Since all the superdimensions are \(\ge 0\), we conclude \(\det (Y') \cong \det (I')\otimes \det (U)\) and \(\det (M')\cong \det (V)\otimes \det (Y'')\). Since \(\det (Y')\cong \det (Y'')\), \(\det (I')\cong \det (I)\) and \(\det (M')\cong \det (M)\) hold in \({\mathcal {K}}^{ev}\), in the stable category \({\mathcal {K}}^{ev}\) we obtain \( \det (M) \cong \det (I)\otimes \det (S[1])\otimes \det (T[-1])\). \(\square \)

A symbolic way of writing.

For \(Y\in {{\mathcal {K}}}^{ev}\) define \( Y \langle \pm 1 \rangle = \Pi (Y)[\pm 1]\) in \( {{\mathcal {K}}}^{ev}\). Then \(Y\langle \pm 1 \rangle \cong Y \otimes {{\textbf{1}}}\langle \pm \rangle \). For the iterated tensor powers \(({{\textbf{1}}}\langle \pm m \rangle )\) of \(({{\textbf{1}}}\langle \pm \rangle )\) one has \({{\textbf{1}}}\langle n_1 \rangle \otimes {{\textbf{1}}}\langle n_2 \rangle \cong {{\textbf{1}}}\langle n_1 + n_2 \rangle \) for all \(n_1,n_2\in {\mathbb {Z}}\). For \(X\in {{\mathcal {K}}}^{odd}\) put \(\det (X):=\det (\Pi (X))^\vee \in {{\mathcal {K}}}^{ev} \). Using this definition, up to negligible objects the formula in the lemma above becomes

$$\begin{aligned} \det (I)\cong \det (S)\otimes \det (M)\otimes \det (T)\langle {{\,\textrm{sdim}\,}}(T)-{{\,\textrm{sdim}\,}}(S) \rangle . \end{aligned}$$

For this notice \(\det (\Pi (X)\langle \pm 1\rangle )\cong \det (\Pi (X)) \langle \pm {{\,\textrm{sdim}\,}}(X)\rangle \). This formula follows from [20, Proposition 1.11]

$$\begin{aligned} \det (X \otimes Y) \cong det(X)^{{{\,\textrm{sdim}\,}}(Y)} \otimes \det (Y)^{{{\,\textrm{sdim}\,}}(X)} \end{aligned}$$

applied for \(Y = {\textbf{1}}\) using \({{\,\textrm{sdim}\,}}({\textbf{1}}\langle 1 \rangle ) =1\) and \(X \langle 1 \rangle \cong X \otimes {\textbf{1}}\langle 1 \rangle \).

14.4 Calculation of determinants

Theorem 14.3

For any maximal atypical weight \(\lambda \) defining \(X_\lambda \) in \(\mathcal {T}_n^+\), for \(\lambda _n=0\) the module \(\det (X_\lambda )\) satisfies

$$\begin{aligned} \det (X_\lambda ) \ = \ B^{\ell (\lambda )} \ \oplus \ \text {negligible}.\end{aligned}$$

In particular, for \(\lambda _n=0\) we have \(\det (X_\lambda )=\textbf{1}\) if (and only if) the maximal atypical weight weight \(\lambda \) is a basic weight.

Proof

We prove this claim in \({\mathcal {K}}\) by a kind of induction, using the method of [36]. This requires a certain ordering of the maximal atypical simple representations, described in the section on the algorithms I, II, and III in [51] [36, Section 20]. For that we define an order on the set of cup diagrams for a fixed block such that the representations with completely nested cup diagrams (in our case the Ber powers) are the minimal elements.

In [51] [36, Section 20] it is also shown that for every maximal atypical irreducible module X there exists a negligible indecomposable object I of Lowey length 3 in \({\mathcal {T}}\) such that the socle \(S\cong A\) and cosocle \(T\cong A\) are isomorphic simple objects A, and the middle layer \(M= X \oplus M'\) is a direct sum of simple objects such that A and all simple summands of \(M'\) are smaller than X with respect to the ordering. More precisely, this indecomposable module is one of the translation functors \(F_i L^{\times \circ }\) of [36, Section 18]. Furthermore, it was shown that all simple objects in M have the same parity [36, Section 20]. Without restriction of generality we may therefore assume that \(M\in {\mathcal {T}}^{ev}\) and \(A\in {\mathcal {T}}^{odd}\) holds. Hence Lemma 14.2 implies that \(\omega (\det (X))\) is a power of \( \omega (B)\), by induction on X with respect to the mentioned ordering.

Concerning the start of this induction: the claim holds for groundstates X in the sense of [36, 51]. In the present situation for \(GL(n\vert n)\) these are the powers \(B^k\) of B for \(k\le 0\). Since the groundstates are the start of the induction above, the determinant is a Ber-power. The specific power \(\ell (\lambda )\) follows then from Corollary 13.4. \(\square \)

Remark 14.4

In [36] we showed that if i is chosen correctly, one can find for given maximal atypical L an irreducible module \(L^{\times \circ }\) such that the translation functor \(F_i (L^{\times \circ })\) satisfies the conditions of the proof. In particular it contains L in the middle layer such that all other composition factors of \(F_i (L^{\times \circ })\) are of lower order. For L with more then one segment we can choose i and \(L^{\times \circ }\) in such a way that all composition factors have one segment less then L. We can now apply the same procedure to all the composition factors of \(F_i(L^{\times \circ })\) with more then one segment. Iterating this we finally end up with a finite number of indecomposable modules where all composition factors have weight diagrams with only one segment. This procedure is called Algorithm I. In Algorithm II we decrease the number of sectors in the same way. Iterating we finally relate L to a finite number of maximal atypical representations with only one sector. Hence after finitely many iterations we have reduced everything to irreducible modules with one segment and one sector. This sector might not be completely nested. In this case we can apply Algorithm II to the internal cup diagram having one segment enclosed by the outer cup. If we iterate this procedure we will finally end up in a collection of Berezin powers.

15 The conformal group and low rank cases

In view of the relation with the conformal group G of the Lorentz metric we discuss ceretain cases of rank \(\le 4\). The complexified Lie algebra \(Lie(G)\otimes _{{\mathbb {R}}} {{\mathbb {C}}}\) of the conformal group is isomorphic to the complex Lie algebra \(\mathfrak {sl}(4)\). So the Lie superalgebras \(\mathfrak {gl}(4\vert N)\) are of potential interest as supersymmetry algebras of conformal field theories and the finite dimensional representations L of these Lie superalgebras may serve as targets of fields \(\psi : M\rightarrow L\) on certain supermanifolds M related to Minkowski space such that Lie(G) acts on M by supervector fields. A covering of the Poincare group can be embedded into G, and in particular the universal covering \(SL(2,{{\mathbb {C}}})\) of the Lorentz group SO(1, 3). The restriction of the representation L to the Lie subalgebra \(Lie(SL(2,{{\mathbb {C}}}))\) decomposes into irreducible representations of the complex Lie algebra \(\mathfrak {sl}(2)\) and their highest weights defines the underlying classical spin values of the L-valued fields. For physical reasons it seems relevant that these spins s are contained in the set \(\{0,\frac{1}{2},1,\frac{3}{2},2\}\). In other words, the highest weights should not exceed 5. We refer to this as the spin condition.

The structure of tensor products of irreducible representation of \(GL(4\vert N)\) resp. \(SL(4\vert N)\) is controlled by the number \(m= min(N,4)\). For \(m < 4\) the information is encoded in the tensor products of irreducible representations of the reduced group \(GL(m\vert m) \times GL(4-m)\). For \(m=4\), this reduced group has to be replaced by \(GL(4\vert 4) \times GL(N-4)\). So these leads us to consider \(GL(n\vert n)\) for \(n=4\) and 3. The case \(n=2\) was completely discussed in Sect. 9. In the following we therefore list some interesting candidates for irreducible superrepresentations L where the spin condition is satisfied. In fact there exist only finitely many isomorphism classes of irreducible representation where the above spin condition is satisfied. The most prominent example is given by \(L=S^1\) where only spin \(s=0\) and \(s=\frac{1}{2}\) shows up in the restriction to the Lie algebra of the Lorentz group of this representation of dimension \(\dim (L)= n^2-2\). In the case \(n=4\) the largest and most interesting example we give is probably the irreducible representation \(L=[3,2,1,0]\) of \(\mathfrak {gl}(4\vert 4)\) of dimension \(\dim (L)= 11{,}163{,}160\). Here all spins s of the restriction are in \(\{0,\frac{1}{2},1,\frac{3}{2},2\}\) and all these numbers occur. As already explained, the Tannaka groups \(H_\lambda \) related to the irreducible representations \(L=L(\lambda )\) may perhaps show up in such theories as hidden approximate symmetry groups. \(L(\lambda )=[3,2,1,0]\) defines a symmetric (SD)-case. The underlying group \(H_\lambda \) should be the group SO(24) if this case is regular (if not \(G_\lambda \) would be SL(12), but we could not exclude this). This case is the only case of our example where we could not exclude exceptional (SD)-case.

One remark for this section. For the convenience of physicists we replace here the groups \(H_\lambda \) by their compact inner forms \(H_\lambda ^c\). So we write U(1) instead of \({\mathbb {G}}_m\) and SU(k) instead of SL(k), \(Sp^c(2k)\) of Sp(2k) etc. In fact, the tensor categories \({{\mathcal {T}}}_\lambda \) of the complex algebraic groups \(H_\lambda \) are isomorphic to the tensor categories of their compact inner forms \(H_\lambda ^c\).

The expected behaviour of the groups \(H_{\lambda }\) was summarized in Theorem 12.5 (all of which is proven except for the exceptional SD-case!). Here we discuss the GL(3|3) and GL(4|4)-case.

Example 15.1

The GL(3|3)-case. For \(n=3\) the structure theorem on the \(G_{\lambda }\) holds unconditionally and therefore also the results on the \(H_{\lambda }\). Here is a list of the nontrivial basic representations and their Tannaka groups. We automatically consider the possible parity shifted representation with positive superdimension here.

  1. (1)

    [2, 1, 0], \({{\,\textrm{sdim}\,}}= 6\), \(H_{\lambda } = Sp^c(6)\).

  2. (2)

    [1, 1, 0], \({{\,\textrm{sdim}\,}}= 3\), \(H_{\lambda } = SU(3)\).

  3. (3)

    [2, 0, 0], \({{\,\textrm{sdim}\,}}= 3\), \(H_{\lambda } = SU(3)\).

  4. (4)

    [1, 0, 0], \({{\,\textrm{sdim}\,}}= 2\), \(H_{\lambda } = SU(2)\).

Twisting any of these with a nontrivial Berezin power gives the GL, GSO or GSp version. The appearing groups exhaust all possible Tannaka groups arising from an \(L(\lambda )\).

Example 15.2

The GL(4|4)-case. Here the structure theorem for \(G_{\lambda }\) (and therefore the determination of \(H_{\lambda }\)) holds unconditionally for basic weights except for the case where \(L(\lambda )\) is weakly selfdual with \([\lambda ] \ne [3,2,1,0]\) by the following lemma:

Lemma 15.3

The basic representations of (SD) type

$$\begin{aligned}{}[3,1,1,0], \ [2,1,0,0], \ [2,2,0,0] \end{aligned}$$

are regular (i.e. \(I \cong {\textbf{1}}\)).

Proof

For [2, 2, 0, 0] this follows from “Appendix E” and example E.8. It is enough to verify that DS([2, 2, 0, 0]) does not contain a summand \(L(\lambda _i)\) with \((\lambda _i)_{basic} = [2,1,0]\). The irreducible representations [3, 1, 1, 0] and [2, 1, 0, 0] have \(k=3\) sectors each. However \(V_{\lambda }\) can only decompose under the restriction to \(G_{\lambda }\) if k is even. Alternatively note that we have embedded subgroups Sp(6) and \(Sp(6) \times SL(3)\) in \(G_{[2,1,0,0]}\) and \(G_{[3,1,1,0]}\) respectively which implies that \(G_{\lambda }\) cannot be SL(3) or SL(6). \(\square \)

For \(n=4\) there are 14 maximal atypical basic irreducible representations in \({{\mathcal {R}}}_4\), the self dual representations

$$\begin{aligned} {\textbf{1}}= [0,0,0,0], S^1=[1,0,0,0], [2,1,0,0], [2,2,0,0], [3,1,1,0], [3,2,1,0] \end{aligned}$$

of superdimension \(1,-2,-6,6,-12,24\) and the representations

$$\begin{aligned} S^2=[2,0,0,0], S^3=[3,0,0,0], [3,1,0,0], [3,2,0,0] \end{aligned}$$

of superdimension \(3,-4, 8, -12\) and their duals

$$\begin{aligned}{}[1,1,0,0], [1,1,1,0], [2,1,1,0], [2,2,1,0]. \end{aligned}$$

Here is a list of the nontrivial basic representations and their Tannaka groups. We automatically consider the possible parity shifted representation with positive superdimension here. Note that the result for the first example [3, 2, 1, 0] assumes that \(G_{\lambda } \cong SO(24)\) (a consequence of the conjectural structure Theorem 12.1).

  1. (1)

    [3, 2, 1, 0], \({{\,\textrm{sdim}\,}}= 24\), \(H_{\lambda } = SO(24)\) (conjecturally).

  2. (2)

    [3, 2, 0, 0], \({{\,\textrm{sdim}\,}}= 12\), \(H_{\lambda } = SU(12)\).

  3. (3)

    [3, 1, 1, 0], \({{\,\textrm{sdim}\,}}= 12\), \(H_{\lambda } = Sp^c(12)\).

  4. (4)

    [3, 1, 0, 0], \({{\,\textrm{sdim}\,}}= 8\), \(H_{\lambda } = SU(8)\).

  5. (5)

    [3, 0, 0, 0], \({{\,\textrm{sdim}\,}}= 4\), \(H_{\lambda } = SU(4)\).

  6. (6)

    [2, 2, 1, 0], \({{\,\textrm{sdim}\,}}= 12\), \(H_{\lambda } = SU(12)\).

  7. (7)

    [2, 2, 0, 0], \({{\,\textrm{sdim}\,}}= 6\), \(H_{\lambda } = SO(6)\).

  8. (8)

    [2, 1, 1, 0], \({{\,\textrm{sdim}\,}}= 8\), \(H_{\lambda } = SU(8)\).

  9. (9)

    [2, 1, 0, 0], \({{\,\textrm{sdim}\,}}= 6\), \(H_{\lambda } = Sp^c(6)\).

  10. (10)

    [2, 0, 0, 0], \({{\,\textrm{sdim}\,}}= 3\), \(H_{\lambda } = SU(3)\).

  11. (11)

    [1, 1, 1, 0], \({{\,\textrm{sdim}\,}}= 4\), \(H_{\lambda } = SU(4)\).

  12. (12)

    [1, 1, 0, 0], \({{\,\textrm{sdim}\,}}= 3\), \(H_{\lambda } = SU(3)\).

  13. (13)

    [1, 0, 0, 0], \({{\,\textrm{sdim}\,}}= 2\), \(H_{\lambda } = SU(2)\).

In addition there is the normalised Berezin representation B, with [1, 1, 1, 1] and \({{\,\textrm{sdim}\,}}= 1\) and in the notation above

$$\begin{aligned} H_{\lambda } = U(1).\end{aligned}$$

Twisting any of the basic representations above with a nontrivial Berezin power gives the GL, GSO or GSp versions. For \(n=4\) the appearing groups exhaust all possible Tannaka groups arising from an \(L(\lambda )\).

Theorem 4.1 implies the following branching rules (the lower index indicates the superdimensions up to a sign):

  1. (1)

    \(DS([3,2,1,0]_{24}) \cong \ [3,2,1]_6 \oplus [1,0,-1]_6 \oplus [3,0,-1]_6 \oplus [3,2,-1]_6\)

  2. (2)

    \(DS([3,2,0,0]_{12}) \cong \ [3,2,0]_6 \oplus [1,-1,-1 ]_3 \oplus [3,-1,-1]_3\)

  3. (3)

    \(DS([3,1,1,0]_{12}) \cong \ [3,1,1]_3 \oplus [3,1,-1]_6 \oplus [0,0,-1]_3\)

  4. (4)

    \(DS([3,1,0,0]_8) \ \cong \ [3,1,0]_6 \oplus [0,-1,-1]_2\)

  5. (5)

    \(DS([3,0,0,0]_4) \ \cong \ [3,0,0]_3 \oplus [-1,-1,-1]_1\)

  6. (6)

    \(DS([2,2,0,0]_6) \ \cong \ [2,2,0]_3 \oplus [2,-1,-1]_3\)

  7. (7)

    \(DS([2,1,0,0]_6) \ \cong \ [2,1,0]_6\)

  8. (8)

    \(DS([2,0,0,0]_3) \ \cong \ [2,0,0]_3\)

  9. (9)

    \(DS([1,0,0,0]_2) \ \cong \ [1,0,0]_2\)

  10. (10)

    \(DS([1,1,1,1]_1) \ \cong \ [1,1,1]_1\)

and \(DS([n,0,0,0]_4) \cong [n,0,0]_3 \oplus [-1,-1,-1]_1\) for all \(n\ge 4\). We also have to consider the dual representations in the cases (2), (4), (5) and (8). We remark that even while most of the derivatives are not basic, they also give examples for \(n=3\) of representations which satisfy the spin condition.

Example 15.4

Consider \(L(\lambda ) = [6,6,1,1]\). It is weakly selfdual with the dual representation \([1,1,-4,-4] = Ber^{-5} \otimes [6,6,1,1,]\). Its superdimension is 6. Since \(\ell (\lambda ) \ne 0\) and its basic weight [2, 2, 0, 0] carries an even pairing, the associated Tannakagroup is therefore \(H_{\lambda } = GSO(V_{\lambda }) \simeq GSO(6)\). This does not depend on the conjecture \(I \simeq {\textbf{1}}\). Indeed DS([6, 6, 1, 1]) does not contain an irreducible summand \(L(\lambda _i)\) with \((\lambda _i)_{basic} = [2,1,0]\) and one can argue as in Lemma 15.3.