The homological spectrum and nilpotence theorems for Lie superalgebra representations
Abstract.
In the study of cohomology of finite group schemes it is well known that nilpotence theorems play a key role in determining the spectrum of the cohomology ring. Balmer recently showed that there is a more general notion of a nilpotence theorem for tensor triangulated categories through the use of homological residue fields and the connection with the homological spectrum. The homological spectrum (like the theory of -points) can be viewed as a topological space that provides an important realization of the Balmer spectrum.
Let be a classical Lie superalgebra over . In this paper, the authors consider the tensor triangular geometry for the stable category of finite-dimensional Lie superalgebra representations: , The localizing subcategories for the detecting subalgebra are classified which answers a question of Boe, Kujawa, and Nakano. As a consequence of these results, the authors prove a nilpotence theorem and determine the homological spectrum for the stable module category of . The authors verify Balmer’s “Nerves of Steel” Conjecture for .
Let (resp. ) be the associated supergroup (scheme) for (resp. ). Under the condition that is a splitting subgroup for , the results for the detecting subalgebra can be used to prove a nilpotence theorem for , and to determine the homological spectrum in this case. Now using natural assumptions in terms of realization of supports, the authors provide a method to explicitly realize the Balmer spectrum of , and prove the Nerves of Steel Conjecture in this case.
1. Introduction
1.1.
For finite group representations, the detecting nilpotence in cohomology via restriction maps and elementary abelian subgroups is an important idea that was used in the study of support varieties in the work of Quillen, Avrunin and Scott. The spectrum of the cohomology for elementary abelian groups can be described through explicit realizations with polynomial rings and this yields a concrete description of the support varieties through a rank variety description. The cohomological nilpotence theorem plays an essential role in the theory because it allows one to describe these support varieties for a finite group with the support varieties for its elementary abelian subgroups. In the case for restricted Lie algebras, Friedlander and Parshall showed that detecting nilpotence entails the use of one-dimensional -nilpotent subalgebras of the Lie algebras. Friedlander and Pevtsova developed a general theory of -points that captures both nilpotence and the realization of supports for arbitrary finite group schemes.
For a small rigid symmetric tensor triangulated category (TTC), , Balmer introduced the concept of homological primes and homological residue fields [Bal20, BC21]. For a TTC, the collection of homological primes, , forms a topological space that can potentially realize the Balmer spectrum, , and its support theory in a concrete way. The central problem in this identification for the theory of homological primes is the following “Nerves of Steel” Conjecture (cf. [Bal20]).
Conjecture 1.1.1.
Let be a small, rigid (symmetric) tensor triangulated category. Then the comparison map is a bijection.
The [NoS Conj] was verified by using a deep stratification result (see [Bal20, BIKP18]) for the stable module category for finite group schemes. In this setting, the homological primes play the role of the -points.
More recently, Balmer [Bal20] developed a nilpotence theorem for morphisms in a tensor triangulated category and Balmer and Cameron [BC21] investigated properties of homological residue fields. Balmer showed that (i) the homological primes naturally surject via the comparison map onto the categorical spectrum and (ii) if a nilpotence theorem holds for homological residue fields then the comparison map is a bijection. In the case of finite group schemes, the picture is complete: the categorical spectrum identifies with projectivization of the cohomology ring and the homological primes with the -points (in a non-trivial way). Recent work on the Nerves of Steel Conjecture in relation to stratification and nilpotence can be found in work by Barthel, Castellana, Heard, Naumann, Sanders, and Pol [BHS23a, BHS23b], [BCHNP23] [BCHS23].
1.2.
Let be a finite-dimensional classical Lie superalgebra with and a reductive algebraic group. Moreover, let be the category of finite-dimensional -supermodules that are completely reducible over . In the mid 2000s, Boe, Kujawa and Nakano [BKN10a] introduced the notion of a detecting subalgebra for classical simple Lie superalgebras by using geometric invariant theory applied to the action on . The detecting subalgebras are important because they detect the -cohomology, and can be regarded as the analogs of elementary abelian subgroups.
There are two families of detecting subalgebras and . The detecting subalgebras, , were used to provide a geometric interpretation of the well-known combinatorial notion of atypicality due to Kac and Wakimoto for basic classical Lie superalgebras. On the other hand, detecting subalgebras, , obtained by stable actions, were used to investigate the tensor triangular geometry and describe the Balmer spectrum for for . For any classical simple Lie superalgebra , the detecting subalgebras can also be used to form a natural triangular decomposition for where is a BBW parabolic subalgebra. The BBW parabolic subgroups/subalgebras have well-behaved homological properties as demonstrated in [GGNW21].
1.3.
The focus of this paper will be to study the tensor triangular geometry of the stable module category where is a classical Lie superalgebra. In particular, we seek to investigate the homological and Balmer spectrum for . For , the Balmer spectrum was computed by Boe, Kujawa and Nakano using heavy representation theoretic techniques. The approach that we use involves using a circle of ideas developed by Benson, Iyengar, and Krause involving the classification of localizing subcategories and by Balmer on homological primes and nilpotence for tensor triangulated categories.
The strategy first entails classifying localizing subcategories for the detecting subalgebras. This answers a question posed in [BKN17]. This result involving stratification by the cohomology ring leads to a nilpotence thoerem for . We then employ the recent work of Serganova and Sherman [SS22] on the existence of “splitting subgroups” where one can find a copy of the trivial module in the induction of the trivial module from the splitting subgroup to the ambient supergroup to prove a nilpotence theorem for .
The nilpotence theorem entails the use of homological residue fields that yield homological primes. In order to show that the homological spectrum identifies with the Balmer spectrum we employ natural assumptions on the realization of supports stated in [BKN17], which was earlier verified for . This new approach has the appeal that it is much more conceptual, streamlined, and applicable to a wider class of Lie superalgebras.
1.4.
The paper is organized as follows. In Section 2, a discussion about classical Lie superalgebras is presented with information about ample detecting subalgebras. Examples are provided that illustrate the -orbit structure on for Type I Lie superalgebras. In the following section (Section 3) we provide conditions for projectivity for classical Lie superalgebras using the detecting subalgebra for finite-dimensional modules and defining the concept of ample detecting subalgebras. For our purposes, stronger detection theorems will be needed to nilpotence theorems with infinitely generated modules which will be stated using the concept of “splitting subalgebras” as introduced in [SS22].
Section 4 focuses on Lie superalgebras whose commutator between the even and odd component is zero, and whose even component is a torus. This class of algebras includes all detecting subalgebras. Our analysis concludes with a classification of localizing subcategories for the stable module category of .
In Section 5, we present several important results on nilpotence theorems for Lie superalgebras (cf. Theorem 5.4.2 and Theorem 5.5.1). In the following section, Section 6, the homological spectrum is calculated for Lie superalgebras that contain a splitting subalgebra isomorphic to algebras considered in Section 4.
Additional conditions are presented in Section 7, that allow one to verify the Nerves of Steel Conjecture for these Lie superalgebras (with a splitting subalgebra). Finally, the paper concludes by providing computations of the homological spectrum and the Balmer spectrum for Type A classical Lie superalgebras (that include ) via our nilpotence theorem in Section 8.
Acknowledgements. We acknowledge Tobias Barthel for providing references to his work with his other coauthors with connections to this paper. Moreover, we thank David Benson and Jon Carlson for comments and discussions involving Section 5.3 .
2. Preliminaries
2.1. Classical Lie Superalgebras
Throughout this paper, let be a Lie superalgebra over the complex numbers where is a -graded vector space with a supercommutator . Let be the universal enveloping superalgebra of . The representation theory entails (super)-modules which are -graded left -modules. Given -modules and one can use the antipode and coproduct of to define a -module structure on the dual and the tensor product . These constructions are essential for applying the theory of tensor triangular geometry. For definitions and detailed background, the reader is referred to [BKN10a, BKN09, BKN10b, BKN17].
A finite dimensional Lie superalgebra is called classical if there is a connected reductive algebraic group such that and the action of on differentiates to the adjoint action of on
We will be interested in studying the category of finite-dimensional and infinite-dimensional Lie superalgebra representations. Let (resp. be the full subcategory of finite dimensional (resp. infinite-dimensional) -modules which are semisimple over . A -module is semisimple if it decomposes into a direct sum of finite dimensional simple -modules. The categories and have enough projective (and injective) objects. Furthermore, injectivity is equivalent to projectivity [BKN10b] so they are Frobenius categories. Since these categories are Frobenius, one can form the stable module categories, and .
2.2. Detecting Subalgebras
In [BKN10a, Section 3,4], Boe, Kujawa and Nakano introduced two families of detecting subalgebras that arise for classical simple Lie superalgebras. One of these families, often denoted by arises from a polar action of on . These -detecting subalgebras will not be germane to the results of the paper. The other family of detecting subalgebras arises from having a stable action on (cf. [BKN10a, Section 3.2]). Let be a generic point in (i.e., is semisimple and regular). Set and . In order to construct a detecting subalgebra, let , , and set
Note our definition is slightly modified from the original construction in [BKN10a].
Then is a classical Lie superalgebra and a sub Lie superalgebra of . The stability of the action of on implies the following properties.
-
(2.2.1)
The restriction homomorphism induces an isomorphism
-
(2.2.2)
The set is dense in
Property (2.2.1) will be discussed later in the context of our nilpotence theorem. For much of our work, a stronger version of property (2.2.2) is needed in order to prove the nilpotence theorem and compute the Balmer spectrum.
2.3. Type I Lie superalgebras:
A Lie superalgebra is Type I if it admits a -grading concentrated in degrees and with and . Examples of Type I Lie superalgebras include the general linear Lie superalgebra and simple Lie superalgebras of types , and . These all have stable actions of on which yields Type I detecting subalgebras.
For our paper, it will be important to distinguish Type I classical Lie superalgebras that contain a detecting subalgebra with favorable geometric properties.
Definition 2.3.1.
Let be a Type I classical Lie superalgebra with a detecting subalgebra . Then is an ample detecting subalgebra if for .
As we will see in next section, there is an abundance of examples of Type I Lie superalgebras with ample detecting subalgebras that encompass many cases of simple Lie superalgebras over .
2.4. Examples of Lie Superalgebras with Ample Detecting Subalgebras
In this section, we will provide examples of Type I classical Lie superalgebras that contain an ample detecting subalgebra. Many of these actions involving on arise naturally in the context of linear algebra. For a more detailed description of these actions, the reader is referred to [BKN10b, Section 3.8].
2.4.1. General Linear Superalgebra:
Let . As a vector space this is isomorphic to the set of by matrices. For a basis, one can take the elementary matrices where . The degree zero component with corresponding reductive group .
Constructions of detecting subalgebras for classical Lie superalgebras are explicitly described in [BKN10a, Section 8]. Set . A detecting subalgebra is given by where is the span of , is the span of , and .
The action of on is given by and on by . It is a well-known fact from linear algebra that the orbits representatives are the matrices of a given rank in . It follows that , and is an ample detecting subalgebra.
2.4.2. Other Type A Lie Superalgebras
The other Type A Lie superalgebras are all Type I, and have . Furthermore, one has as given above for as a subalgebra of [BKN10b, Section 3.8.2 and 3.8.3].
When , consists of the matrices of supertrace zero, and
The -orbits are the same as the -orbits, and is an ample detecting subalgebra.
For the Lie superalgebra has a one dimensional center given by scalar multiples of the identity matrix, and one has
For elements of with rank strictly less than , the -orbits coincide with the -orbits. The orbits of full rank matrices form a one parameter family with each orbit containing a unique matrix which is a scalar multiple of the identity. The orbit theory for the case is analogous to . Consequently, in both these setting the algebra is ample.
2.4.3. Type C Lie Superalgebras which are Type I
In this case with . One has , the natural module for . The action of is transitive on . One has an explicit detecting subalgebra where . The transitivity of the action of on shows that . A similar argument demonstrates that .
2.4.4. Type P Lie Superalgebras
For Type P Lie superalgebras and one has an explicit detecting subalgebra where contains matrices of all possible ranks.
Let . Then and and as -modules, where denotes the natural -module. There are a finite number of orbits given again by the condition on rank, and their closure relation forms a chain. This shows that is ample.
Now let be the simple Lie superalgebra of type . One has and are as above but . This case follows the paradigm of . The -orbits corresponding to matrices of rank less than in are also -orbits. The matrices of rank yield a one parameter family of orbits that have orbit representatives in , which demonstrates the ampleness of .
3. Supports and Projectivity
3.1. Cohomological and Rank Varieties
We review the constructions in [BKN10b, Section 3.2] for Type I Lie superalgebras. Let be a Type I Lie superalgebra. Then are abelian Lie superalgebras. Consequently, identifies with an exterior algebra, and the cohomology ring for these superalgebras identifies with the symmetric algebra on the dual of . Set . Let be a finite-dimensional -module and let
The (cohomological) support variety of is defined as
Moreover, the support variety is canonically isomorphic to the following rank variety:
These varieties satisfy many of the important properties of support theory that include (i) the detection of projectivity over and (ii) the tensor product property.
For a detecting subalgebra , one can apply the prior construction to obtain support varieties for , namely and .
3.2. Projectivity for Type I Lie Superalgebras
For Type I classical Lie superalgebras, one can construct Kac and dual Kac modules (cf. [BKN10b, Section 3.1]). A module in is tilting if and only if it has both a Kac and a dual Kac filtration. The use of these filtrations was a key idea in proving the following criteria for projectivity in the category (see [BKN10b, Section 3]).
Theorem 3.2.1.
Let be a Type I classical Lie superalgebra and . The following are equivalent.
-
(a)
is a projective module in
-
(b)
is a tilting module
-
(c)
and
-
(d)
and
It should be noted that for an arbitrary (infinitely generated) module , one can have projectivity over , but may not be projective in . For example, if , one can take an infinite coproduct of projective modules in the principal block . By making suitable identifications, one can form an (infinite) “zigzag module” (of radical length ) that has a Kac and dual Kac filtration, which is projective upon restriction to . The zigzag module is not projective in because it has radical length less than . This construction can also be performed for projective modules in the principal block for the restricted enveloping algebra of (cf. [Po68]), and has been observed in other situations by Cline, Parshall and Scott [CPS85, (3.2) Example].
3.3. Projectivity via Ample Detecting Subalgebras
The following theorem allows us connect projectivity of a module in to projectivity when restricting the module to the detecting subalgebra.
Theorem 3.3.1.
Let be a Type I classical Lie superalgebra with an ample detecting subalgebra and . Then is projective in if and only if is projective in .
Proof.
Let be projective in . Then by Theorem 3.2.1, . It follows that and by Theorem 3.2.1, is a projective module in .
Conversely, assume that is a projective module in . Then . Let . Then where and since is an ample detecting subalgebra. Since is a rational -module, is -stable. This implies that and , and . Consequently, , and by the same reasoning . One can now conclude by Theorem 3.2.1 that is a projective module in . ∎
3.4. Splitting Subalgebras
For our purposes we will need a stronger projectivity criterion than the one given in Theorem 3.3.1 that can be applied to infinite-dimensional modules. In order to state such a projectivity criterion, one need to employ the concept of splitting subgroups/subalgebras as introduced by Serganova and Sherman [SS22]. The following definition which fits in our context is equivalent to their definition.
Definition 3.4.1.
Let be a classical Lie superalgebra and be a Lie supergroup scheme with . Moreover, let be a subgroup with being classical and . Then is a splitting subalgebra if and only if the trivial module is a direct summand of .
The following theorem summarizes results in [SS22, Section 2]. We present a slightly different approach using the work for BBW parabolic subgroups by D. Grantcharov, N. Grantcharov, Nakano and Wu [GGNW21].
Theorem 3.4.2.
Let be a classical Lie superalgebra and be a splitting subalgebra. Let , be modules in .
-
(a)
for .
-
(b)
is projective in if and only if when restricted to is projective in .
-
(c)
For all , .
-
(d)
For all , the restriction map is injective.
Proof.
First observe that the category (resp. ) is equivalent to the rational category (resp. ). The categories are Frobenius which means that projectivity is equivalent to injectivity. Furthermore, the quotient is affine so the induction functor is exact (cf. [SS22, Lemma 2.1]). Thus (a) holds.
For (b), since the quotient is affine, if is projective/injective then upon restriction to is projective/injective. Conversely, suppose that is a -module that is injective upon restriction to . Then is injective, and by the tensor identity . Now since is a splitting subgroup of , it follows that is a summand of , and is thus injective.
For (c), there exists a spectral sequence,
By using (a), the spectral sequence collapses and yields the result. For (d), one can use the isomorphism in (c), and note that the restriction map is realized by taking a summand of . ∎
Serganova and Sherman proved that the detecting subalgebra for classical Lie algebras of Type A are splitting subalgebras. They also present other examples of splitting subalgebras for Type Q classical “simple” Lie superalgebras that are not detecting subalgebras. See [SS22, Theorem 1.1]. An interesting question would be to determine for Type I classical Lie superalgebras with an ample detecting subalgebra whether the detecting subalgebra is a splitting subalgebra.
4. Tensor Triangular Geometry
4.1. Triangulated Categories
We will review the basic notions of triangulated categories, as these ideas lie at the foundation of the rest of the paper. Let be a triangulated category. Recall that this means is an additive category equipped with an auto-equivalence called the shift, and a set of distinguished triangles:
all subject to a list of axioms the reader can find in [Nee01, Ch. 1].
A non-empty, full, additive subcategory of is called a triangulated subcategory if (i) implies that for all and (ii) if is a distinguished triangle in and two of are objects in then the third object is also in A triangulated subcategory of is called thick if is closed under taking direct summands.
Assume that the triangulated category admits set-indexed coproducts. A triangulated subcategory of is called a localizing subcategory if is closed under taking set-indexed coproducts. It follows from the Eilenberg swindle that localizing subcategories are necessarily thick.
An object is called compact if commutes with set-indexed coproducts. The full subcategory of compact objects in is denoted by and the triangulated category is said to be compactly generated if the isomorphism classes of compact objects form a set, and if for each non-zero object there is an object such that
4.2. Central Ring Actions
Let be a triangulated category with suspension and let be a graded-commutative ring. Recall that being graded-commutative means that admits a grading and that for homogeneous elements In this section we recall what it means for to act on
Let and be objects in and set
Then is a graded abelian group, and is a graded ring where the multiplication is “shift and compose”. Notice that is a right and a left -bimodule.
The graded center of is the graded-commutative ring denoted by whose degree component is given by
With this setup, an action of on is a homomorphism of graded rings, and if admits such an action, is called an -linear triangulated category. It follows that if is -linear, then for each object in there is an induced homomorphism of graded rings such that the induced -module structures on by and agree up to sign.
4.3. Tensor Triangulated Categories
Our paper will concern triangulated categories that have an additional monoidal structure. Namely, we work with tensor triangulated categories as defined in [Bal05]. A tensor triangulated category, or TTC, is a triple where is a triangulated category, is a symmetric monoidal (tensor) product which is exact in each variable, and a monoidal unit
Let be a tensor triangulated category. With the additional structure of the tensor product on one can define analogues to the usual notions in commutative algebra like prime ideal and spectrum. A tensor ideal in is a triangulated subcategory of such that for all and A proper, thick tensor ideal of is said to be a prime ideal if implies that either or The Balmer spectrum [Bal05, Definition 2.1] is defined as
The Zariski topology on has as closed sets
where is any family of objects in
The tensor triangulated category is said to be a compactly generated TTC if is closed under set indexed coproducts, the tensor product preserves set indexed coproducts, is compactly generated as a triangulated category, the tensor product of compact objects is compact, is a compact object, and every compact object is rigid (i.e., strongly dualizable).
4.4. Stratifying Tensor Triangulated Categories
Given a tensor triangulated category it is a difficult problem to classify its localizing subcategories and its thick tensor ideals. However, there are some circumstances in which such classifications are known, particularly in the case when is some category of representations.
Stratification was introduced for triangulated categories in [BIK11a], and for tensor triangulated categories in [BIK11b]. The property of stratification allows for the classification of the localizing subcategories in terms of the -support. Importantly, stratification is also easily tracked under change-of-categories, which plays a key role in these arguments. Since our work deals with tensor triangulated categories, we review that setting in this section. For a more detailed exposition the reader is referred to the original sources.
Let be a tensor triangulated category. Furthermore, assume that the unit object is a compact generator for . Then is a graded commutative ring which induces an action on by the homomorphisms
for each object Let be a graded-commutative ring. A canonical action of on is an action of on which is induced by a homomorphism of graded-commutative rings .
Let denote the set of homogeneous prime ideals in One can construct a notion of support for objects of by considering localization and colocalization functors. This support is described in detail in [BIK11a, Section 3.1]. Recall that the specialization closure of a subset is the set of all for which there exists with A subset is specialization closed if it equals its specialization closure. Given a specialization closed subset , the full subcategory of -torsion objects is defined as
It can be checked that is a localizing subcategory of so there exists a localization functor and an induced colocalization functor called the local cohomology functors of which for each object give a functorial triangle
For each , let and be specialization closed subsets of such that . There is a local-cohomology functor independent of the choice of and . These functors give rise to a notion of support for objects and a notion of support for itself. In particular, set
and
For each , the full subcategory is a tensor ideal and localizing. The main definition of this section is the concept of stratification.
Definition 4.4.1.
The -linear tensor triangulated category is said to be stratified by if is either zero or a minimal tensor ideal localizing subcategory of for all .
4.5.
For each localizing subcategory of set
and for each subset of , set
The following theorems [BIK11a, Theorem 3.8] classify localizing subcategories and thick tensor ideals via the aforementioned maps.
Theorem 4.5.1.
Let be a tensor triangulated category stratified by the action of Then the maps and are mutually inverse bijections between the tensor ideal localizing subcategories of and the subsets of
In the case when is stratified by a noetherian ring , the maps and also give a classification of thick tensor ideals in
Theorem 4.5.2.
Let be a compactly generated tensor triangulated category which is stratified by the noetherian ring . Then the maps and are mutually inverse bijections between the thick tensor ideal subcategories of and the specialization closed subsets of :
4.6.
The following result, whose proof of which can be found in [Nee01], demonstrates a close relationship between stratification of the derived category of a commutative ring with its canonical action, and is needed for our classifications.
Theorem 4.6.1.
Let be a commutative Noetherian ring and let be the derived category of differential graded -modules. Then is stratified by the canonical action of
4.7. Lie Superalgebras of the Form
Throughout this section denotes a Lie superalgebra where is a torus, and . This is a classical Lie superalgebra. As an example, could be any of the detecting subalgebras introduced in [BKN10a, Section 4.4]. Let be the stable module category whose objects are all -modules which are finitely semi-simple as -modules. Then is a compactly generated tensor-triangulated category with full subcategory of compact objects . Note that the objects of consist precisely of the finite-dimensional objects in . Let denote the cohomology ring . According to [BKN17, Section 4.3],
and thus is a polynomial algebra. The following theorem was established in [BKN17, Theorem 4.5.4].
Theorem 4.7.1.
The thick tensor ideals in are in bijective correspondence with specialization closed subsets of Furthermore, is homeomorphic to .
In [BKN17] the authors point out that is an -linear triangulated category, and that the local-global principle holds. It was conjectured that stratifies a result which would recover the theorem. The goal of this section is to pursue the stratification avenue, and to prove the following theorem.
Theorem 4.7.2.
The tensor-triangulated category is stratified by the action of the cohomology ring .
Proof.
The following argument follows closely that of [BIK11a, Section 5.2] where the category is considered, where is an algebraically closed field of characteristic two, and is an elementary abelian -group. The key observation that makes these situations similar is that the cohomology rings in both cases are polynomial rings, and both the group algebra and the universal enveloping superalgebra of are exterior algebras.
Consider the universal enveloping superalgebra of the quotient . Then there is an isomorphism of -algebras Therefore, there is an isomorphism of rings . Set . Choose a basis of so that is an isomorphism of rings, and view as a differential graded algebra with zero differential and for each .
The -algebra is graded with degree component and with multiplication defined by Consider as a differential graded algebra with zero differential. The degree one element defined as
satisfied Let denote the differential graded module over with graded module and differential given by
Since is a differential graded module over for each differential graded module over there is an induced structure of a differential graded -module on Then the functor
is an equivalence of triangulated categories.
To see this first observe that as a complex, consists of injective -modules. This follows from the fact that is self-injective. That the actions coincide follows from [BIK11a, Theorem 5.4], since the R-action on is also canonical. Finally, the result follows from the equivalence of categories and the recollement from [BIK11a, Theorem 3.19].
∎
4.8.
Theorems 4.4.2 and 4.4.3 yield the following result on localizing subcategories of thick tensor ideals for the stable module categories associated with .
Theorem 4.8.1.
Let denotes a Lie superalgebra where is a torus, and .
-
(a)
There is an equivalence of triangulated categories .
-
(b)
The localizing subcategories of are in bijective correspondence with subsets of .
-
(c)
The thick tensor ideals in are in bijective correspondence with specialization closed subsets of .
5. Nilpotence Theorems
In this section, we first recall the important ideas involving the homological spectrum, homological residue fields, and abstract nilpotence theorems in tensor triangulated categories. These concepts were developed by Balmer in [Bal20]. For the ease of exposition it is convenient to modify notation slightly. Unless otherwise stated, will denote a tensor triangulated category satisfying the assumptions from [Bal20, 2.2], namely is essentially small and rigid. At times for our purposes we view as sitting inside of a “large” compactly tensor triangulated category with This is in anticipation of applying the “Balmer machine” to the categories and which are the main point of study for this article.
5.1. The Homological Spectrum
The Grothendieck abelian category of right -modules denoted - has as objects the additive functors from the opposite category of to the category of abelian groups, and has as morphisms the natural transformations between functors. Let h denote the Yoneda embedding:
For details about -modules, the reader is referred to [Bal20, Section 2.6] and [BKS19, Appendix A]. The main facts needed in what follows are that (i) the category - admits a tensor product -- which is right-exact in each variable and which makes h a tensor functor, (ii) h preserves rigidity, (iii) is compact, projective, and -flat in and (iv) the tensor subcategory - of compact objects is abelian and is the Freyd envelope of (i.e., - is the universal homological functor out of .) Recall that a functor from a triangulated category to an abelian category is homological if it maps distinguished triangles to exact sequences. Also, recall that a subcategory of an abelian category is called a Serre subcategory if it is closed under subobjects, quotients, and extensions. The main definitions of this subsection, which are given in [Bal20, Definition 3.1] are the following.
Definition 5.1.1.
-
(a)
A coherent homological prime for is a maximal proper Serre -ideal subcategory of - of the Freyd envelope of
-
(b)
The homological spectrum of denoted as a set consists of all the homological primes of
and has as topology that generated by the basis of closed subsets for all where
It should be noted that one can check that is a support data on Therefore, there exists a unique continuous map
such that for all See [Bal05, Thm. 3.2]. This map is often called the comparison map, and is surjective as long as is rigid. When is rigid, there are many examples where the comparison map is a bijection. At the time of the writing of this article, all known examples have the property that is bijective. See [Bal20, Section 5].
5.2. Homological Residue Fields
In this section, we recall Balmer’s construction [Bal20] of homological residue fields. One of the main questions in tensor triangular geometry is to find the appropriate tensor triangular analogue to ordinary fields in commutative algebra. In particular, given how does one construct functors to its “residue fields”? This question is explored in [BKS19], and some major takeaways are that there are several important properties one would like the notion of field to have. Moreover, there are many examples of tensor triangulated categories that should be considered as tensor triangulated fields. However, it is not clear exactly what the definition should be. The following definition was proposed in [BKS19, Definition 1.1], and will be the running definition in this work.
Definition 5.2.1.
A non-trivial (big) tensor triangulated category is a tensor triangulated field if every object of is a coproduct of compact-rigid objects of and if every non-zero object in is tensor-faithful.
While this definition encapsulates many of the desired properties of fields, there is not yet a purely tensor triangular construction of them analogous to extracting residue fields in commutative algebra. Instead, Balmer uses the homological spectrum to construct homological tensor functors to abelian categories:
Definition 5.2.2.
Given a coherent homological prime the homological residue field corresponding to is the functor
composed of the Yoneda embedding followed by the Gabriel quotient.
A natural question at this point is whether or not homological residue fields are related to the tensor triangular fields of Definition 5.2.2. The answer is yes, and an explicit connection useful for the computation of homological residue fields in examples is the content of the following theorem stated in [BC21, Lemma 2.2].
Theorem 5.2.3.
Given a big tensor-triangulated category a tensor-triangulated field and a monoidal exact functor with right adjoint one has the following diagram:
where is the exact cocontinuous functor induced by the functor is the Gabriel quotient with respect to and the functor is induced by the universal property, hence and is exact and faithful.
The adjunctions and are depicted with and . Moreover, is a homological prime and and is a homological residue field of .
5.3. Nilpotence and Colimits
In this section we clarify the notions of nilpotence in the stable categories of Lie superalgebra representations and relate them to colimit constructions in module categories and homotopy colimits in the stable categories. We first discuss the concept of nilpotence.
Definition 5.3.1.
Let and be modules in .
-
(a)
A map is called null if in ; i.e., is null if and only if factors through a projective module.
-
(b)
A map is called tensor nilpotent if there exists some such that the tensor power is null.
In the case when is compact, one can transform the condition of the nilpotence of the map to the adjoint map.
Lemma 5.3.2.
Let be a compact object. A map is tensor nilpotent if and only if the adjoint map is tensor nilpotent.
Proof.
By adjointness, we have an isomorphism . Since is tensor nilpotent, there exists some such that factors through a projective. But since tensor products of projective modules are projective, this implies that factors through a projective; i.e., that is tensor nilpotent. ∎
Next we need to recall the definition of a colimit in the category and a homotopy colimit in its stable module category.
Definition 5.3.3.
Let be a classical Lie superalgebra.
-
(a)
Let
be a system of modules and homomorphisms in . Let be defined by whenever . The colimit of the system, if ti exists, is the module given by .
-
(b)
Let
be a system of modules and homomorphisms in . The homotopy colimit of the system is the module obtained by completing the map
to a triangle:
The following lemmas are given in [Ric97] and are modified here for Lie superalgebra representations.
Lemma 5.3.4.
Let be a sequence of maps in a triangulated category with countable direct sums. If for each there exists such that , then .
The next lemma clarifies the relationship between homotopy colimits in the stable category with colimits in the ordinary module category.
Lemma 5.3.5.
Let be a sequence of modules and homomorphisms in . The colimit in is isomorphic in to the homotopy colimit .
These two lemmas together allow one to derive an analog of [BenC18, Lemma 2.3]
Theorem 5.3.6.
A map is -nilpotent if and only if the colimit of
is projective.
Proof.
First suppose that is -nilpotent. Then Lemma 5.3.4 implies that the homotopy colimit of this system viewed in the stable category is isomorphic to zero, which is to say the colimit of the system is zero by Lemma 5.3.5. Now suppose that the colimit of the system is projective. Again, when viewed in the stable category this implies that the homotopy colimit is zero, which gives the tensor nilpotence of . ∎
5.4. Nilpotence Theorems
Nilpotence theorems have played an important role in cohomology and representation theory. Devinatz, Hopkins, and Smith showed in [DHS88] that a map between finite spectra which gets annihilated by all Morava -theories must be tensor-nilpotent. Neeman [Nee92] and Thompson [Tho97] proved nilpotence theorems for maps in derived categories using ordinary residue fields, and Benson, Carlson, and Rickard [BCR97] proved nilpotence theorems in modular representation theory, where the residue fields are given by cyclic shifted subgroups, or, in the case of finite group schemes, -points [FP07]. With these examples in mind, Balmer in [Bal20] using homological residue fields.presented a unified treatment that applies to all tensor triangulated categories. In the case where sits inside of a big rigidly compactly generated tensor triangulated category with one can make a connection to the homological spectrum. In particular, he proved the following theorem [Bal20, Corollary 4.7]:
Theorem 5.4.1.
Let be a rigidly-compactly generated “big” tensor-triangulated category with Let be a morphism in with compact and arbitrary. If in for every homological residue field for every homological prime - then there exists such that in
The nilpotence theorem stated above can combined with the theory of detecting subalgebras developed by Boe, Kujawa, and Nakano, to the study of nilpotence in the stable categories of Lie superalgebra representations. The following nilpotence theorem via homological residue fields is a direct translation of Theorem 5.4.1 in the context of superalgebra representatiions.
Theorem 5.4.2.
Let be a classical Lie superalgebra, and let be a morphism in with Suppose that for all Then there exists such that in
Proof.
The first observation is that sits inside of as the compact objects: Moreover, the compact objects and the rigid objects coincide and generate as a tensor-triangulated category. This is the setup of Theorem 5.4.1. ∎
5.5. A Nilpotence Theorem via Detecting Subalgebra
The salient feature first discovered about detecting subalgebras was that these subalgebras detect nilpotence in cohomology. We will now show that a remarkable feature for classical Lie subalgebras with a splitting subalgebras is that nilpotence of arbitrary maps in the stable module category is governed by nilpotence when restricting the the map to a splitting subalgebra. In particular, to show that a morphism is nilpotent in the big stable module category where is compact, it is enough to check vanishing on those homological residue fields constructed via homological primes from the stable categories of modules over the splitting subalgebra.
Theorem 5.5.1.
Let be a classical Lie subalgebra with a splitting subalgebra Let be a morphism in with . If for all then there exists such that in
Proof.
Let be the usual restriction functor. By our hypothesis, for all , Theorem 5.4.2 implies that is tensor nilpotent in .
6. Identifying the Homological Spectrum
The goals of this section are to determine the homological spectrum for and where is a classical Lie superalgebra with splitting subalgebra We also consider the comparison map first defined in Section 5.1.
6.1. Stratification and the Comparison Map
The first observation in this direction is that classification of localizing subcategories via stratification enables one to show that the comparison map is a bijection:
Theorem 6.1.1.
Let be a Type I classical Lie superalgebra with a torus and Then the comparison map
is a bijection.
Proof.
A more general argument that shows that stratification implies the Nerves of Steel Conjecture can be found in [BHS23a, Theorem 4.7].
6.2.
Let be a classical Lie superalgebra and be a detecting subalgebra in . We will need to work with a field extension of such that the transcendence degree is larger than the dimension of . Note that this is the analogous setup as in [BC21, Example 3.9]. The stable module categories involved will be viewed over the field extension . Let be the prime ideal in associated with the “generic point” (cf. [BCR96, Sections 2 and 3] for an explanation).
For with viewed as a vector space over , one has is either or . In either case, the blocks are either semisimple or have finite representation type. One can verify that is a tensor triangular field. For , one have two monoidal exact functors (given by restriction):
(6.2.1) |
(6.2.2) |
Let be the natural functor obtained by restricting -modules to -modules. Then for all .
Now one can apply Theorem 5.2.3 (where and ), to obtain a homological prime (resp. ) associated to (resp. ). Similarly, let (resp. ) be the homological residue field corresponding to (resp. ).
6.3.
Let be a Type I classical Lie superalgebra with detecting subalgebra Let and let be the usual inclusion of the Lie subsuperalgebra generated by into This yields, for each an induced functor These are monoidal functors, each of which with right adjoint given by induction. Moreover, is a tensor-triangulated field, so Theorem 5.2.3 produces a homological residue field. Denote the corresponding homological prime of by The goal now is the show that contains all of the homological primes. Recall the following result in [Bal20, Theorem 5.4].
Theorem 6.3.1.
Let be a big tensor-triangulated category with Consider a family of points in the homological spectrum. Suppose that the corresponding functors collectively detect -nilpotence in the following sense: If in is such that and for all then for some Then we have
6.4.
We are now ready to provide conditions on when one can identify a collection of homological primes that detect nilpotence on .
Theorem 6.4.1.
Let be a classical Lie superalgebra and be a sub Lie superalgebras. Denote by , and the associated supergroup (schemes) such that , and . Set . Assume that
-
(a)
with ;
-
(b)
is a splitting subgroup of .
Then (i.e., a set of -orbit representatives) detects nilpotence in .
Proof.
The idea of the proof is to find a set of homological primes that detects nilpotence in . Then one can apply Theorem 6.3.1 (e.g., [Bal20, Theorem 5.4].)
The first step is to compare homological residue fields for and . If is in with compact then one can compare the diagrams for and to conclude the following. (1) If then for . Now one can apply the stratification result, Theorem 6.1.1, to conclude that are the homological primes for . Therefore, by (1) and Theorem 5.4.2, one has (2) If for all then is -nilpotent in . Applying Theorem 5.5.1 since is a splitting subalgebra of , one can conclude that is -nilpotent in . Let . Since is a -module, it follows that the functors (resp. ) will provide the same decomposition of in (resp. ). By considering Theorem 5.2.3, it follows that if and only if . Therefore, detects nilpotence.
∎
In the previous theorem, one can state that . However, with the definition of there are certain homological primes that might be identified in the set. We will show in the following section that different -orbit representatives yield different elements in .
7. Nerves of Steel Conjecture
7.1.
There are noticeable differences between the stable module category for finite group schemes versus the stable module category for Lie superalgebras. For example, the comparison map:
(7.1.1) |
is not always homeomorphism (e.g., when ). In general the cohomology ring does not stratify . There are many examples where the support theory does not detect projectivity. This is the main reason one needs to use the cohomology of the detecting subalgebra to realize the homological spectrum and the Balmer spectrum.
Boe, Kujawa and Nakano [BKN17] showed that for , one has a homeomorphism:
(7.1.2) |
where is a detecting (splitting) subalgebra of and is the normalizer of in . From this example, it is clear that in order to compute the Balmer spectrum for Lie superalgebras one needs to find a suitable replacement for the cohomology ring.
From Section 6, when one has a splitting subalgebra of , one can compute the homological spectrum and show there is a surjection:
(7.1.3) |
Since is rigid, the comparison map
(7.1.4) |
is surjective. Our goal is to use the prior calculation of the homological spectrum to give conditions on when the Nerves of Steel Conjecture holds (i.e., when is bijective).
7.2.
We can now identify the homological spectrum and the Balmer spectrum for classical Lie superalgebras with a splitting subalgebra under a suitable realization condition.
Theorem 7.2.1.
Let be a classical Lie superalgebra with a splitting subalgebra . Assume that
-
(i)
where is a torus and .
-
(ii)
Given an -invariant closed subvariety of , there exists with .
Then
-
(a)
There exists a 1-1 correspondence
where and are specialization closed sets of .
-
(b)
There exists a homeomorphism -.
-
(c)
The comparison map is bijective.
Proof.
(a) and (b) follow by [BKN17, Theorems 3.4.1, 3.5.1]. For part (c), let which is given by a concrete description in [NVY, Corollary 6.2.4]. Consider the following diagram of topological spaces:
(7.2.1) |
One has that is a homeomorphism and is a bijection for . From part (b), the map is a homomorphism. The maps and are surjections. The map sends to in . Suppose that . The using the commutativity, one has in - which means that and are -conjugate. This proves that in .
∎
We remark that the verification of the Nerves of Steel Conjecture in the previous theorem uses stratification results only for , unlike the the case for finite group schemes where the stratification is needed for (see [Bal20, 5.13 Example]). A general stratification result for will be addressed in future work.
8. Applications
8.1.
We need to verify the realization condition in Theorem 7.2.1(ii) for other Lie superalgebras of Type A, namely . The main ideas have been established in the -case (cf [BKN17, Section 7 ]). We will use the same notation to show how the arguments need to modified to handle the -case.
Let and write for the detecting subalgebra of . Without loss of generality one can assume that . The subalgebra has a basis given by the matrix units
Let denote the torus of consisting of diagonal matrices. The torus acts on by the adjoint action and the aforementioned basis for consists of weight vectors. Let
where and are defined by letting be the linear functional given on our basis of matrix units for by and otherwise zero. The functionals are defined similarly.
For our purposes it will suffice to consider the maximal ideal spectrum version of the support varieties. One has
If then where fixes and acts trivially on . Here denote the symmetric group on letters embedded diagonally in . Let be such a variety. Since is a finite group we may write
for some -invariant closed subvariety of .
Let be the closed subvariety of determined by a homogeneous ideal of . It follows that must be of the form
(8.1.1) |
where are homogeneous polynomials of weight zero for .
For the analysis in [BKN17], one employed the following conventions that will also be useful here. Let be non-negative integers with , and let
(8.1.2) |
This is the conical variety given by the vanishing of “ coordinates”, “ coordinates”, with overlapping pairs.
8.2. Examples:
In this section, we will use examples to illustrate the concepts from the prior sections.
8.2.1.
In this case is one-dimensional and consists of matrices which are non-zero scalar multiples of the identity matrix. The group acts trivally on . Therefore, the homogeneous polynomials in all have weight zero. This is in stark comparison to the -case where the weight zero polynomials are the polynomials in the variable .
8.2.2.
This case is more like the case for The torus has dimension 2, and the vectors and are not fixed by . The the weight zero polynomial are polynomials in the variable .
8.2.3.
The maximal torus does not act trivially on . A direct computation via the determinant condition shows that is generated by , , , and . One still needs to consider the varieties is this case to describe the -invariant subvarieties of .
8.3.
The aim of this section is to outline how to realize every -invariant closed subvariety of as for some in the case of . The verification will use the results and the proofs in [BKN17] which accomplish this for .
The first proposition can be proved in the same way as in in [BKN17, Proposition 7.2.1] via the construction of -modules from elements in the cohomology ring.
Proposition 8.3.1.
Let be homogeneous polynomials in of weight zero with respect to . Then there exists a module in such that .
The next series of steps involves realization involving certain subvarieties using modules for . These varieties can be realized for modules in by simply restricting the action of the modules realized for . Finally, one can use the proof in [BKN17, Theorem 7.12.1], to obtain the following result.
Theorem 8.3.2.
Let and let be homogeneous weight zero polynomials. Set . Then there exists a finite dimensional -module such that
8.4.
For Lie superalgebras of Type A, we can now compute the Balmer spectrum, and verify the Nerves of Steel Conjecture.
Theorem 8.4.1.
Let be or , and be a detecting subalgebra of . Then
-
(a)
There exists a 1-1 correspondence
where .
-
(b)
There exists a homeomorphism -.
-
(c)
The comparison map is bijective.
Proof.
The statement of the theorem follows from Theorem 7.2.1 once Conditions (i) and (ii) are verified. For (i), in the cases when or , the detecting subalgebras are splitting subalgebras.
We remark that the Lie superalgebra is a quotient of . One cannot simply restrict modules like in the case for to . This means that new techniques need to be developed to verify Condition (ii) for .
References
- [Ba09] I. Bagci, Cohomology and Support Varieties for Lie Superalgebras, Ph.D Thesis, University of Georgia, 2009.
- [BaKN08] I. Bagci, J.R. Kujawa, D.K. Nakano, Cohomology and support varieties for the Lie superalgebra , International Math. Research Notices, doi:10.1093/imrn/rnn115, (2008).
- [Bal05] P. Balmer, The spectrum of prime ideals in tensor triangulated categories, J. Reine Angew. Math., 588, (2005), 149–168.
- [Bal20] P. Balmer, Nilpotence theorems via homological residue fields, Tunisian J. Math, 2, (2020), 359–378.
- [BC21] P. Balmer, J.C. Cameron, Computing homological residue fields in algebra and topology, Proceedings of the AMS, 149, (2021), 3177–3185.
- [BKS19] P. Balmer, H. Krause, G. Stevenson., Tensor-triangular fields: ruminations, Selecta Mathematica, 25, (2019), 1–36.
- [BHS23a] T. Barthel, D. Heard, B. Sanders, Stratification and the comparison between homological and tensor triangular support, Q.J. Math., 74, (2023), 747-766.
- [BHS23b] T. Barthel, D. Heard, B. Sanders, Stratification in tensor triangular geometry with applications to spectral Mackey functors, Camb. J. Math., 11, (2023), 829-915.
- [BCHNP23] T. Barthel, N. Castellana, D. Heard, N. Naumann, L. Pol, Quillen stratification in equivariant homotopy theory, ArXiv:2301.02212.
- [BCHS23] T. Barthel, N. Castellana, D. Heard, B. Sanders, On surjectivity in tensor triangular geometry, ArXiv:2305.05604.
- [Ben98] D.J. Benson, Representations and Cohomology. II, second ed., Cambridge Studies in Advanced Mathematics, vol. 31, Cambridge University Press, Cambridge, 1998.
- [BenC18] D.J. Benson, J.F. Carlson, Nilpotence and generation in the stable module category, J. Algebra, 222, (2018), 3566-3584.
- [BCR96] D.J. Benson, J.F. Carlson, J. Rickard, Complexity and varieties for infinitely generated modules, II, Math. Proc. Camb. Phil. Soc., 120, (1996), 597-615.
- [BCR97] D.J. Benson, J.F. Carlson, J. Rickard, Thick subcategories of the stable module category, Fund. Math., 153, (1997), 59–80.
- [BIK11a] Benson, D.J., S. Iyengar, and H. Krause., Representations of finite groups: Local cohomology and support, Vol. 43. Springer Science and Business Media, (2011)
- [BIK11b] D.J. Benson, S. Iyengar, H. Krause., Stratifying Triangulated Categories, Journal of Topology 4, (2011), 641-666
- [BIKP18] D.J., Benson, S. Iyengar, H. Krause, J. Pevtsova, Stratification for module categories of finite group schemes, Journal of the AMS, 31, (2018), 265–302.
- [BKN10a] B.D. Boe, J.R. Kujawa, D.K. Nakano, Cohomology and support varieties for Lie superalgebras, Transactions of the AMS, 362, (2010), 6551-6590.
- [BKN09] B.D. Boe, J.R. Kujawa, D.K. Nakano, Cohomology and support varieties for Lie superalgebras II, Proc. London Math. Soc., 98 (2009), no. 1, 19–44.
- [BKN10b] B.D. Boe, J.R. Kujawa, D.K. Nakano, Complexity and module varieties for classical Lie superalgebras, International Math. Research Notices, doi:10.1093/imrn/rnq090, (2010).
- [BKN17] B.D. Boe, J.R. Kujawa, D.K. Nakano, Tensor triangular geometry for classical Lie superalgebras, Advances in Math., 314, (2017), 228–277.
- [Car83] J.F. Carlson, The varieties and the cohomology ring of a module, J. Algebra, 85, (1983), no. 1, 104–143.
- [Car84] J.F. Carlson, The variety of an indecomposable module is connected, Invent. Math., 77, (1984), no. 2, 291-299.
- [CPS85] E.T. Cline, B.J. Parshall, L.L. Scott, On injective modules for infinitesimal algebraic groups I, J. London Math. Soc., 31, (1985), 277-291.
- [CPS88] E.T. Cline, B.J. Parshall, L.L. Scott, Finite-dimensional algebras and highest weight categories, J. Reine Angew. Math., 391, (1988), 85–99.
- [DK85] J. Dadok, V. G. Kac, Polar representations, J. Algebra, 92, (1985), no. 2, 504–524.
- [DHS88] E.S. Devinatz, M.J. Hopkins, J.H. Smith, Nilpotence and stable homotopy theory I, Annals of Math., 128, (1988), 207–241.
- [FP07] E.M. Friedlander, J. Pevtsova, -supports for modules for finite group schemes, Duke Math. J., 139, (2007), 317–368.
- [GGNW21] D. Grantcharov, N. Grantcharov, D.K. Nakano, J. Wu, On BBW parabolics for simple classical Lie superalgebras, Advances in Math., 381, (2021), 107647.
- [NVY] D.K. Nakano, K.B. Vashaw, M.T. Yakimov, On the spectrum and support theory for finite tensor categories, Math. Annalen, (2023) (published online), https://doi.org/10.1007/s00208-023-02759-8
- [Nee92] A. Neeman, The chromatic tower for . Topology, 31, (1992), 519–532.
- [Nee01] A. Neeman, Triangulated Categories, Annals of Mathematics Studies, vol. 148, Princeton University Press, Princeton, NJ, 2001.
- [LR79] D. Luna, R.W. Richardson, A generalization of the Chevalley restriction theorem, Duke Math. J., 46, (1979), 487–496.
- [Ric97] J. Rickard, Idempotent modules in the stable categories, J. London Math. Soc., 56, (1997), 149–170.
- [Po68] R.D. Pollack, Restricted Lie algebras of bounded type, Bulletin of the AMS, 74, (1968), 326-331.
- [Se05] V. Serganova, On representations of Cartan type Lie superalgebras, Amer. Math. Soc. Transl., 213, (2005), no. 2, 223–239.
- [SS22] V. Serganova, A. Sherman, Spitting quasireductive supergroups and volumes of supergrassmannians, arXiv: 2206.07693.
- [Tho97] R.W. Thomason, The classification of triangulated subcategories, Compositio Math., 105, (1997), 1–27.