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

The homological spectrum and nilpotence theorems for Lie superalgebra representations

Matthew H. Hamil Department of Mathematics
University of Georgia
Athens
GA 30602, USA
[email protected]
 and  Daniel K. Nakano Department of Mathematics
University of Georgia
Athens
GA 30602, USA
[email protected]
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 π\pi-points) can be viewed as a topological space that provides an important realization of the Balmer spectrum.

Let 𝔤=𝔤0¯𝔤1¯{\mathfrak{g}}={\mathfrak{g}}_{\bar{0}}\oplus{\mathfrak{g}}_{\bar{1}} be a classical Lie superalgebra over {\mathbb{C}}. In this paper, the authors consider the tensor triangular geometry for the stable category of finite-dimensional Lie superalgebra representations: stab((𝔤,𝔤0¯))\text{stab}({\mathcal{F}}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})}), The localizing subcategories for the detecting subalgebra 𝔣{\mathfrak{f}} 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 (𝔣,𝔣0¯){\mathcal{F}}_{({\mathfrak{f}},{\mathfrak{f}}_{\bar{0}})}. The authors verify Balmer’s “Nerves of Steel” Conjecture for (𝔣,𝔣0¯){\mathcal{F}}_{({\mathfrak{f}},{\mathfrak{f}}_{\bar{0}})}.

Let FF (resp. GG) be the associated supergroup (scheme) for 𝔣{\mathfrak{f}} (resp. 𝔤{\mathfrak{g}}). Under the condition that FF is a splitting subgroup for GG, the results for the detecting subalgebra can be used to prove a nilpotence theorem for stab((𝔤,𝔤0¯))\text{stab}({\mathcal{F}}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})}), 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 stab((𝔤,𝔤0¯))\text{stab}({\mathcal{F}}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})}), and prove the Nerves of Steel Conjecture in this case.

Research of the first author was supported in part by NSF grant DMS-2101941
Research of the second author was supported in part by NSF grant DMS-2101941

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 pp-nilpotent subalgebras of the Lie algebras. Friedlander and Pevtsova developed a general theory of π\pi-points that captures both nilpotence and the realization of supports for arbitrary finite group schemes.

For a small rigid symmetric tensor triangulated category (TTC), 𝐊\bf K, Balmer introduced the concept of homological primes and homological residue fields [Bal20, BC21]. For a TTC, the collection of homological primes, Spch(𝐊)\text{Spc}^{\text{h}}(\bf K), forms a topological space that can potentially realize the Balmer spectrum, Spc(𝐊)\text{Spc}(\bf K), 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.

[NoSConj]\operatorname{[NoS\ Conj]} Let 𝐊\bf K be a small, rigid (symmetric) tensor triangulated category. Then the comparison map ϕ:Spch(𝐊)Spc(𝐊)\phi:\operatorname{Spc}^{h}(\bf K)\rightarrow\operatorname{Spc}(\bf K) 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 π\pi-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 π\pi-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 𝔤=𝔤0¯𝔤1¯{\mathfrak{g}}={\mathfrak{g}}_{\bar{0}}\oplus{\mathfrak{g}}_{\bar{1}} be a finite-dimensional classical Lie superalgebra with Lie G0¯=𝔤0¯\text{Lie }G_{\bar{0}}={\mathfrak{g}}_{\bar{0}} and G0¯G_{\bar{0}} a reductive algebraic group. Moreover, let (𝔤,𝔤0¯){\mathcal{F}}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})} be the category of finite-dimensional 𝔤{\mathfrak{g}}-supermodules that are completely reducible over 𝔤0¯{\mathfrak{g}}_{\bar{0}}. 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 G0¯G_{\bar{0}} action on 𝔤1¯{\mathfrak{g}}_{\bar{1}}. The detecting subalgebras are important because they detect the (𝔤,𝔤0¯){\mathcal{F}}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})}-cohomology, and can be regarded as the analogs of elementary abelian subgroups.

There are two families of detecting subalgebras 𝔢{\mathfrak{e}} and 𝔣{\mathfrak{f}}. The detecting subalgebras, 𝔢{\mathfrak{e}}, 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, 𝔣{\mathfrak{f}}, obtained by stable actions, were used to investigate the tensor triangular geometry and describe the Balmer spectrum for stab((𝔤,𝔤0¯))\text{stab}({\mathcal{F}}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})}) for 𝔤=𝔤𝔩(m|n){\mathfrak{g}}=\mathfrak{gl}(m|n). For any classical simple Lie superalgebra 𝔤{\mathfrak{g}}, the detecting subalgebras 𝔣{\mathfrak{f}} can also be used to form a natural triangular decomposition for 𝔤=𝔫+𝔣𝔫{\mathfrak{g}}={\mathfrak{n}}^{+}\oplus{\mathfrak{f}}\oplus{\mathfrak{n}}^{-} where 𝔟=𝔣𝔫{\mathfrak{b}}={\mathfrak{f}}\oplus{\mathfrak{n}}^{-} 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 (𝔤,𝔤0¯){\mathcal{F}}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})} where 𝔤{\mathfrak{g}} is a classical Lie superalgebra. In particular, we seek to investigate the homological and Balmer spectrum for stab((𝔤,𝔤0¯))\text{stab}({\mathcal{F}}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})}). For 𝔤=𝔤𝔩(m|n){\mathfrak{g}}=\mathfrak{gl}(m|n), 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 stab((𝔣,𝔣0¯))\text{stab}({\mathcal{F}}_{({\mathfrak{f}},{\mathfrak{f}}_{\bar{0}})}). 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 stab((𝔤,𝔤0¯))\text{stab}({\mathcal{F}}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})}).

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 𝔤=𝔤𝔩(m|n){\mathfrak{g}}=\mathfrak{gl}(m|n). 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 G0¯G_{\bar{0}}-orbit structure on 𝔤±1{\mathfrak{g}}_{\pm 1} 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 𝔷{\mathfrak{z}} 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 𝒞(𝔷,𝔷0¯){\mathcal{C}}_{({\mathfrak{z}},{\mathfrak{z}}_{\bar{0}})}.

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 𝔰𝔩(m|n)\mathfrak{sl}(m|n)) 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 𝔤{\mathfrak{g}} be a Lie superalgebra over the complex numbers {\mathbb{C}} where 𝔤=𝔤0¯𝔤1¯{\mathfrak{g}}={\mathfrak{g}}_{\bar{0}}\oplus{\mathfrak{g}}_{\bar{1}} is a 2\mathbb{Z}_{2}-graded vector space with a supercommutator [,]:𝔤𝔤𝔤[\;,\;]:\mathfrak{g}\otimes\mathfrak{g}\longrightarrow\mathfrak{g}. Let U(𝔤)U(\mathfrak{g}) be the universal enveloping superalgebra of 𝔤{\mathfrak{g}}. The representation theory entails (super)-modules which are 2\mathbb{Z}_{2}-graded left U(𝔤)U(\mathfrak{g})-modules. Given 𝔤{\mathfrak{g}}-modules MM and NN one can use the antipode and coproduct of U(𝔤)U({\mathfrak{g}}) to define a 𝔤{\mathfrak{g}}-module structure on the dual MM^{*} and the tensor product MNM\otimes N. 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 𝔤{\mathfrak{g}} is called classical if there is a connected reductive algebraic group G0¯G_{\bar{0}} such that Lie(G0¯)=𝔤0¯,\operatorname{Lie}(G_{\bar{0}})=\mathfrak{g}_{\bar{0}}, and the action of G0¯G_{\bar{0}} on 𝔤1¯\mathfrak{g}_{\bar{1}} differentiates to the adjoint action of 𝔤0¯\mathfrak{g}_{\bar{0}} on 𝔤1¯.\mathfrak{g}_{\bar{1}}.

We will be interested in studying the category of finite-dimensional and infinite-dimensional Lie superalgebra representations. Let (𝔤,𝔤0¯){\mathcal{F}}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})} (resp. 𝒞(𝔤,𝔤0¯)){\mathcal{C}}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})}) be the full subcategory of finite dimensional (resp. infinite-dimensional) 𝔤{\mathfrak{g}}-modules which are semisimple over 𝔤0¯{\mathfrak{g}}_{\bar{0}}. A 𝔤0¯{\mathfrak{g}}_{\bar{0}}-module is semisimple if it decomposes into a direct sum of finite dimensional simple 𝔤0¯{\mathfrak{g}}_{\bar{0}}-modules. The categories (𝔤,𝔤0¯){\mathcal{F}}_{(\mathfrak{g},\mathfrak{g}_{\bar{0}})} and 𝒞(𝔤,𝔤0¯){\mathcal{C}}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})} 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, stab((𝔤,𝔤0¯))\text{stab}({\mathcal{F}}_{(\mathfrak{g},\mathfrak{g}_{\bar{0}})}) and Stab(𝒞(𝔤,𝔤0¯))\text{Stab}({\mathcal{C}}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})}).

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 𝔢,{\mathfrak{e}}, arises from a polar action of G0¯G_{\bar{0}} on 𝔤1¯{\mathfrak{g}}_{\bar{1}}. These 𝔢{\mathfrak{e}}-detecting subalgebras will not be germane to the results of the paper. The other family of detecting subalgebras arises from G0¯G_{\bar{0}} having a stable action on 𝔤1¯{\mathfrak{g}}_{\bar{1}} (cf. [BKN10a, Section 3.2]). Let x0x_{0} be a generic point in 𝔤1¯{\mathfrak{g}}_{\bar{1}} (i.e., x0x_{0} is semisimple and regular). Set H=StabG0¯x0H=\text{Stab}_{G_{\bar{0}}}x_{0} and N:=NG0¯(H)N:=N_{G_{\bar{0}}}(H). In order to construct a detecting subalgebra, let 𝔣1¯=𝔤1¯H{\mathfrak{f}}_{\bar{1}}={\mathfrak{g}}_{\bar{1}}^{H}, 𝔣0¯=[𝔣1¯,𝔣1¯]{\mathfrak{f}}_{\bar{0}}=[{\mathfrak{f}}_{\bar{1}},{\mathfrak{f}}_{\bar{1}}], and set

𝔣=𝔣0¯𝔣1¯.{\mathfrak{f}}={\mathfrak{f}}_{\bar{0}}\oplus{\mathfrak{f}}_{\bar{1}}.

Note our definition is slightly modified from the original construction in [BKN10a].

Then 𝔣{\mathfrak{f}} is a classical Lie superalgebra and a sub Lie superalgebra of 𝔤{\mathfrak{g}}. The stability of the action of G0¯G_{\bar{0}} on 𝔤1¯{\mathfrak{g}}_{\bar{1}} implies the following properties.

  • (2.2.1)

    The restriction homomorphism S(𝔤1¯)S(𝔣1¯)S(\mathfrak{g}_{\bar{1}}^{*})\longrightarrow S(\mathfrak{f}_{\bar{1}}^{*}) induces an isomorphism

    res:H(𝔤,𝔤0¯,)H(𝔣,𝔣0¯,)N.\operatorname{res}:\text{H}^{\bullet}({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}},{\mathbb{C}})\rightarrow\text{H}^{\bullet}({\mathfrak{f}},{\mathfrak{f}}_{\bar{0}},{\mathbb{C}})^{N}.
  • (2.2.2)

    The set G0¯𝔣1¯G_{\bar{0}}\cdot{\mathfrak{f}}_{\bar{1}} is dense in 𝔤1¯.\mathfrak{g}_{\bar{1}}.

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 \mathbb{Z}-grading 𝔤=𝔤1𝔤0𝔤1{\mathfrak{g}}=\mathfrak{g}_{-1}\oplus{\mathfrak{g}}_{0}\oplus{\mathfrak{g}}_{1} concentrated in degrees 1,-1, 0,0, and 11 with 𝔤0¯=𝔤0{\mathfrak{g}}_{\bar{0}}={\mathfrak{g}}_{0} and 𝔤1¯=𝔤1𝔤1{\mathfrak{g}}_{\bar{1}}=\mathfrak{g}_{-1}\oplus{\mathfrak{g}}_{1}. Examples of Type I Lie superalgebras include the general linear Lie superalgebra 𝔤𝔩(m|n)\mathfrak{gl}(m|n) and simple Lie superalgebras of types A(m,n)A(m,n), C(n)C(n) and P(n)P(n). These all have stable actions of G0¯G_{\bar{0}} on 𝔤1¯{\mathfrak{g}}_{\bar{1}} 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 𝔤{\mathfrak{g}} be a Type I classical Lie superalgebra with a detecting subalgebra 𝔣=𝔣1𝔣0𝔣1{\mathfrak{f}}={\mathfrak{f}}_{-1}\oplus{\mathfrak{f}}_{0}\oplus{\mathfrak{f}}_{1}. Then 𝔣{\mathfrak{f}} is an ample detecting subalgebra if 𝔤j=G0𝔣j{\mathfrak{g}}_{j}=G_{0}\cdot{\mathfrak{f}}_{j} for j=1,1j=-1,1.

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 {\mathbb{C}}.

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 G0G_{0} on 𝔤±1{\mathfrak{g}}_{\pm 1} 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 𝔤=𝔤l(m|n){\mathfrak{g}}={\mathfrak{g}l}(m|n). As a vector space this is isomorphic to the set of m+nm+n by m+nm+n matrices. For a basis, one can take the elementary matrices Ei,jE_{i,j} where 1i,jm+n1\leq i,j\leq m+n. The degree zero component 𝔤0𝔤𝔩(m)×𝔤𝔩(n){\mathfrak{g}}_{0}\cong\mathfrak{gl}(m)\times\mathfrak{gl}(n) with corresponding reductive group G0GL(m)×GL(n)G_{0}\cong GL(m)\times GL(n).

Constructions of detecting subalgebras for classical Lie superalgebras are explicitly described in [BKN10a, Section 8]. Set r=min(m,n)r=\text{min}(m,n). A detecting subalgebra is given by 𝔣=𝔣1𝔣0¯𝔣1\mathfrak{f}=\mathfrak{f}_{-1}\oplus\mathfrak{f}_{\bar{0}}\oplus\mathfrak{f}_{1} where 𝔣1\mathfrak{f}_{-1} is the span of {Em+i,i:i=1,2,,r}\{E_{m+i,i}:\ i=1,2,\dots,r\}, 𝔣1\mathfrak{f}_{1} is the span of {Ei,m+i:i=1,2,,r}\{E_{i,m+i}:\ i=1,2,\dots,r\}, and 𝔣0¯=[𝔣1¯,𝔣1]\mathfrak{f}_{\bar{0}}=[{\mathfrak{f}}_{\bar{1}},{\mathfrak{f}}_{1}].

The action of G0G_{0} on 𝔤1\mathfrak{g}_{-1} is given by (A,B).X=BXA1(A,B).X=BXA^{-1} and on 𝔤1\mathfrak{g}_{1} by (A,B).X=AXB1(A,B).X=AXB^{-1}. It is a well-known fact from linear algebra that the orbits representatives are the matrices of a given rank in 𝔤±1\mathfrak{g}_{\pm 1}. It follows that 𝔤±1=G0𝔣±1{\mathfrak{g}}_{\pm 1}=G_{0}\cdot{\mathfrak{f}}_{\pm 1}, and 𝔣{\mathfrak{f}} is an ample detecting subalgebra.

2.4.2. Other Type A Lie Superalgebras

The other Type A Lie superalgebras 𝔤{\mathfrak{g}} are all Type I, and have 𝔤±1𝔤l(m|n)±1{\mathfrak{g}}_{\pm 1}\cong{\mathfrak{g}l}(m|n)_{\pm 1}. Furthermore, one has 𝔣{\mathfrak{f}} as given above for 𝔤l(m|n){\mathfrak{g}l}(m|n) as a subalgebra of 𝔤{\mathfrak{g}} [BKN10b, Section 3.8.2 and 3.8.3].

When mnm\neq n, 𝔤=𝔰𝔩(m|n)𝔤𝔩(m|n)\mathfrak{g}=\mathfrak{sl}(m|n)\subseteq\mathfrak{gl}(m|n) consists of the matrices of supertrace zero, and

G0={(A,B)GL(m)×GL(n)det(A)det(B)1=1},G_{0}=\left\{(A,B)\in GL(m)\times GL(n)\mid\det(A)\det(B)^{-1}=1\right\},

The G0¯G_{\bar{0}}-orbits are the same as the GL(m)×GL(n)GL(m)\times GL(n)-orbits, and 𝔣{\mathfrak{f}} is an ample detecting subalgebra.

For the Lie superalgebra 𝔰𝔩(n|n)\mathfrak{sl}(n|n) has a one dimensional center given by scalar multiples of the identity matrix, and one has

G0{(A,B)GL(n)×GL(n)det(A)det(B)1=1}.G_{0}\cong\left\{(A,B)\in GL(n)\times GL(n)\mid\det(A)\det(B)^{-1}=1\right\}.

For elements of 𝔤±1\mathfrak{g}_{\pm 1} with rank strictly less than nn, the G0G_{0}-orbits coincide with the GL(n)×GL(n)GL(n)\times GL(n)-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 𝔤=𝔭𝔰𝔩(n|n){\mathfrak{g}}=\mathfrak{psl}(n|n) case is analogous to 𝔰𝔩(n|n)\mathfrak{sl}(n|n). Consequently, in both these setting the algebra 𝔣{\mathfrak{f}} is ample.

2.4.3. Type C Lie Superalgebras which are Type I

In this case 𝔤=𝔬𝔰𝔭(2|2n)\mathfrak{g}=\mathfrak{osp}(2|2n) with G0¯××Sp(2n)G_{\bar{0}}\cong\mathbb{C}^{\times}\times Sp(2n). One has 𝔤1V2n\mathfrak{g}_{1}\cong V_{2n}, the natural module for Sp(2n)Sp(2n). The action of Sp(2n)Sp(2n) is transitive on V2n{0}V_{2n}\smallsetminus\{0\}. One has an explicit detecting subalgebra 𝔣=𝔣1𝔣0¯𝔣1{\mathfrak{f}}=\mathfrak{f}_{-1}\oplus\mathfrak{f}_{\bar{0}}\oplus\mathfrak{f}_{1} where dim𝔣±1=1\dim{\mathfrak{f}}_{\pm 1}=1. The transitivity of the action of G0¯G_{\bar{0}} on 𝔤1{\mathfrak{g}}_{1} shows that 𝔤1=G0𝔣1{\mathfrak{g}}_{1}=G_{0}\cdot{\mathfrak{f}}_{1}. A similar argument demonstrates that 𝔤1=G0𝔣1{\mathfrak{g}}_{-1}=G_{0}\cdot{\mathfrak{f}}_{-1}.

2.4.4. Type P Lie Superalgebras

For Type P Lie superalgebras 𝔤=𝔭~(n){\mathfrak{g}}=\tilde{\mathfrak{p}}(n) and 𝔤=𝔭~(n){\mathfrak{g}}=\tilde{\mathfrak{p}}(n) one has an explicit detecting subalgebra 𝔣=𝔣1𝔣0¯𝔣1{\mathfrak{f}}=\mathfrak{f}_{-1}\oplus\mathfrak{f}_{\bar{0}}\oplus\mathfrak{f}_{1} where 𝔣±1{\mathfrak{f}}_{\pm 1} contains matrices of all possible ranks.

Let 𝔤=𝔭~(n)\mathfrak{g}=\tilde{\mathfrak{p}}(n). Then G0¯GL(n)G_{\bar{0}}\cong GL(n) and g1Λ2(V)g_{-1}\cong\Lambda^{2}(V^{*}) and g1S2(V)g_{1}\cong S^{2}(V) as G0¯G_{\bar{0}}-modules, where VV denotes the natural GL(n)GL(n)-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 𝔣{\mathfrak{f}} is ample.

Now let 𝔤=𝔭(n)=[𝔭~(n),𝔭~(n)]\mathfrak{g}=\mathfrak{p}(n)=[\tilde{\mathfrak{p}}(n),\tilde{\mathfrak{p}}(n)] be the simple Lie superalgebra of type P(n1)P(n-1). One has 𝔤1\mathfrak{g}_{-1} and 𝔤1\mathfrak{g}_{1} are as above but G0¯SL(n)G_{\bar{0}}\cong SL(n). This case follows the paradigm of 𝔰𝔩(n|n)\mathfrak{sl}(n|n). The GL(n)GL(n)-orbits corresponding to matrices of rank less than nn in 𝔤±1{\mathfrak{g}}_{\pm 1} are also G0¯G_{\bar{0}}-orbits. The matrices of rank nn yield a one parameter family of orbits that have orbit representatives in 𝔣±1{\mathfrak{f}}_{\pm 1}, which demonstrates the ampleness of 𝔣{\mathfrak{f}}.

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 𝔤=𝔤1𝔤0¯𝔤1{\mathfrak{g}}=\mathfrak{g}_{-1}\oplus{\mathfrak{g}}_{\bar{0}}\oplus\mathfrak{g}_{1} be a Type I Lie superalgebra. Then 𝔤±1\mathfrak{g}_{\pm 1} are abelian Lie superalgebras. Consequently, U(𝔤±1)U({\mathfrak{g}}_{\pm 1}) identifies with an exterior algebra, and the cohomology ring for these superalgebras identifies with the symmetric algebra on the dual of 𝔤±1\mathfrak{g}_{\pm 1}. Set R±1=H(𝔤±1,)S(𝔤±1)R_{\pm 1}=\operatorname{H}^{\bullet}(\mathfrak{g}_{\pm 1},\mathbb{C})\cong S^{\bullet}(\mathfrak{g}_{\pm 1}^{*}). Let MM be a finite-dimensional U(𝔤±1))U({\mathfrak{g}}_{\pm 1}))-module and let

JM={rR±1r.m=0 for all mExtU(𝔤±1)(M,M)}J_{M}=\left\{r\in R_{\pm 1}\mid r.m=0\text{ for all }m\in\operatorname{Ext}_{U({\mathfrak{g}}_{\pm 1})}^{\bullet}(M,M)\right\}

The (cohomological) support variety of MM is defined as

𝒱𝔤±1(M)=MaxSpec(R±1/JM).\mathcal{V}_{\mathfrak{g}_{\pm 1}}(M)=\operatorname{MaxSpec}\left(R_{\pm 1}/J_{M}\right).

Moreover, the support variety 𝒱𝔤±1(M)\mathcal{V}_{\mathfrak{g}_{\pm 1}}(M) is canonically isomorphic to the following rank variety:

𝒱𝔤±1rank(M)={x𝔤±1M is not projective as a U(x)-module}{0}.\mathcal{V}_{\mathfrak{g}_{\pm 1}}^{\text{rank}}(M)=\left\{x\in\mathfrak{g}_{\pm 1}\mid M\text{ is not projective as a $U(\langle x\rangle)$-module}\right\}\cup\{0\}.

These varieties satisfy many of the important properties of support theory that include (i) the detection of projectivity over U(𝔤±1)U({\mathfrak{g}}_{\pm 1}) and (ii) the tensor product property.

For a detecting subalgebra 𝔣=𝔣1𝔣0𝔣1{\mathfrak{f}}={\mathfrak{f}}_{-1}\oplus{\mathfrak{f}}_{0}\oplus{\mathfrak{f}}_{1}, one can apply the prior construction to obtain support varieties for M(𝔣,𝔣0)M\in{\mathcal{F}}_{({\mathfrak{f}},{\mathfrak{f}}_{0})}, namely 𝒱𝔣±1(M)\mathcal{V}_{{\mathfrak{f}}_{\pm 1}}(M) and 𝒱𝔣±1rank(M)\mathcal{V}^{\text{rank}}_{{\mathfrak{f}}_{\pm 1}}(M).

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 (𝔤,𝔤0){\mathcal{F}}_{({\mathfrak{g}},{\mathfrak{g}}_{0})} 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 (𝔤,𝔤0){\mathcal{F}}_{({\mathfrak{g}},{\mathfrak{g}}_{0})} (see [BKN10b, Section 3]).

Theorem 3.2.1.

Let 𝔤{\mathfrak{g}} be a Type I classical Lie superalgebra and M(𝔤,𝔤0)M\in{\mathcal{F}}_{({\mathfrak{g}},{\mathfrak{g}}_{0})}. The following are equivalent.

  • (a)

    MM is a projective module in (𝔤,𝔤0){\mathcal{F}}_{({\mathfrak{g}},{\mathfrak{g}}_{0})}

  • (b)

    MM is a tilting module

  • (c)

    𝒱𝔤1(M)={0}\mathcal{V}_{\mathfrak{g}_{1}}(M)=\{0\} and 𝒱𝔤1(M)={0}\mathcal{V}_{\mathfrak{g}_{-1}}(M)=\{0\}

  • (d)

    𝒱𝔤1rank(M)={0}\mathcal{V}^{\operatorname{rank}}_{\mathfrak{g}_{1}}(M)=\{0\} and 𝒱𝔤1rank(M)={0}\mathcal{V}^{\operatorname{rank}}_{\mathfrak{g}_{-1}}(M)=\{0\}

It should be noted that for an arbitrary (infinitely generated) module M𝒞(𝔤,𝔤0)M\in{\mathcal{C}}_{({\mathfrak{g}},{\mathfrak{g}}_{0})}, one can have projectivity over U(𝔤±1)U({\mathfrak{g}}_{\pm 1}), but MM may not be projective in 𝒞(𝔤,𝔤0){\mathcal{C}}_{({\mathfrak{g}},{\mathfrak{g}}_{0})}. For example, if 𝔤=𝔤𝔩(1|1){\mathfrak{g}}=\mathfrak{gl}(1|1), one can take an infinite coproduct of projective modules in the principal block P=mP(m|m)P=\oplus_{m\in{\mathbb{Z}}}P(m|-m). By making suitable identifications, one can form an (infinite) “zigzag module” (of radical length 22) that has a Kac and dual Kac filtration, which is projective upon restriction to U(𝔤±1)U({\mathfrak{g}}_{\pm 1}). The zigzag module is not projective in 𝒞(𝔤,𝔤0){\mathcal{C}}_{({\mathfrak{g}},{\mathfrak{g}}_{0})} because it has radical length less than 44. This construction can also be performed for projective modules in the principal block for the restricted enveloping algebra of 𝔰l2{\mathfrak{s}l}_{2} (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 (𝔤,𝔤0){\mathcal{F}}_{({\mathfrak{g}},{\mathfrak{g}}_{0})} to projectivity when restricting the module to the detecting subalgebra.

Theorem 3.3.1.

Let 𝔤{\mathfrak{g}} be a Type I classical Lie superalgebra with an ample detecting subalgebra 𝔣{\mathfrak{f}} and M(𝔤,𝔤0)M\in{\mathcal{F}}_{({\mathfrak{g}},{\mathfrak{g}}_{0})}. Then MM is projective in (𝔤,𝔤0){\mathcal{F}}_{({\mathfrak{g}},{\mathfrak{g}}_{0})} if and only if MM is projective in (𝔣,𝔣0){\mathcal{F}}_{({\mathfrak{f}},{\mathfrak{f}}_{0})}.

Proof.

Let MM be projective in (𝔤,𝔤0){\mathcal{F}}_{({\mathfrak{g}},{\mathfrak{g}}_{0})}. Then by Theorem 3.2.1, 𝒱𝔤±1(M)={0}\mathcal{V}_{\mathfrak{g}_{\pm 1}}(M)=\{0\}. It follows that 𝒱𝔣±1(M)={0}\mathcal{V}_{{\mathfrak{f}}_{\pm 1}}(M)=\{0\} and by Theorem 3.2.1, MM is a projective module in (𝔣,𝔣0){\mathcal{F}}_{({\mathfrak{f}},{\mathfrak{f}}_{0})}.

Conversely, assume that MM is a projective module in (𝔣,𝔣0){\mathcal{F}}_{({\mathfrak{f}},{\mathfrak{f}}_{0})}. Then 𝒱𝔣±1(M)={0}\mathcal{V}_{{\mathfrak{f}}_{\pm 1}}(M)=\{0\}. Let y𝒱𝔤1(M)y\in\mathcal{V}_{{\mathfrak{g}}_{1}}(M). Then y=gxy=g\cdot x where gG0g\in G_{0} and x𝔣1x\in{\mathfrak{f}}_{1} since 𝔣{\mathfrak{f}} is an ample detecting subalgebra. Since MM is a rational G0G_{0}-module, 𝒱𝔤1(M)\mathcal{V}_{{\mathfrak{g}}_{1}}(M) is G0G_{0}-stable. This implies that x𝒱𝔤1(M)x\in\mathcal{V}_{{\mathfrak{g}}_{1}}(M) and x𝒱𝔣1(M)x\in\mathcal{V}_{{\mathfrak{f}}_{1}}(M), and x=0x=0. Consequently, 𝒱𝔤1(M)={0}\mathcal{V}_{{\mathfrak{g}}_{1}}(M)=\{0\}, and by the same reasoning 𝒱𝔤1(M)={0}\mathcal{V}_{{\mathfrak{g}}_{-1}}(M)=\{0\}. One can now conclude by Theorem 3.2.1 that MM is a projective module in (𝔤,𝔤0){\mathcal{F}}_{({\mathfrak{g}},{\mathfrak{g}}_{0})}. ∎

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 𝔤=𝔤0¯𝔤1¯{\mathfrak{g}}={\mathfrak{g}}_{\bar{0}}\oplus{\mathfrak{g}}_{\bar{1}} be a classical Lie superalgebra and GG be a Lie supergroup scheme with LieG=𝔤\operatorname{Lie}G={\mathfrak{g}}. Moreover, let ZGZ\leq G be a subgroup with 𝔷=𝔷0¯𝔷1¯{\mathfrak{z}}={\mathfrak{z}}_{\bar{0}}\oplus{\mathfrak{z}}_{\bar{1}} being classical and LieZ=𝔷\operatorname{Lie}Z={\mathfrak{z}}. Then 𝔷{\mathfrak{z}} is a splitting subalgebra if and only if the trivial module {\mathbb{C}} is a direct summand of indZG\operatorname{ind}_{Z}^{G}{\mathbb{C}}.

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 𝔤{\mathfrak{g}} be a classical Lie superalgebra and 𝔷{\mathfrak{z}} be a splitting subalgebra. Let MM, NN be modules in 𝒞(𝔤,𝔤0¯){\mathcal{C}}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})}.

  • (a)

    RjindZG=0R^{j}\operatorname{ind}_{Z}^{G}{\mathbb{C}}=0 for j>0j>0.

  • (b)

    MM is projective in 𝒞(𝔤,𝔤0¯){\mathcal{C}}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})} if and only if MM when restricted to 𝔷{\mathfrak{z}} is projective in 𝒞(𝔷,𝔷0¯){\mathcal{C}}_{({\mathfrak{z}},{\mathfrak{z}}_{\bar{0}})}.

  • (c)

    For all n0n\geq 0, Ext(𝔤,𝔤0¯)n(M,NindZG)Ext(𝔷,𝔷0¯)n(M,N)\operatorname{Ext}^{n}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})}(M,N\otimes\operatorname{ind}_{Z}^{G}{\mathbb{C}})\cong\operatorname{Ext}^{n}_{({\mathfrak{z}},{\mathfrak{z}}_{\bar{0}})}(M,N).

  • (d)

    For all n0n\geq 0, the restriction map res:Ext(𝔤,𝔤0¯)n(M,N)Ext(𝔷,𝔷0¯)n(M,N)\operatorname{res}:\operatorname{Ext}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})}^{n}(M,N)\rightarrow\operatorname{Ext}^{n}_{({\mathfrak{z}},{\mathfrak{z}}_{\bar{0}})}(M,N) is injective.

Proof.

First observe that the category 𝒞(𝔤,𝔤0¯){\mathcal{C}}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})} (resp. 𝒞(𝔷,𝔷0¯){\mathcal{C}}_{({\mathfrak{z}},{\mathfrak{z}}_{\bar{0}})}) is equivalent to the rational category Mod(G)\text{Mod}(G) (resp. Mod(Z)\text{Mod}(Z)). The categories are Frobenius which means that projectivity is equivalent to injectivity. Furthermore, the quotient G/ZG/Z is affine so the induction functor is exact (cf. [SS22, Lemma 2.1]). Thus (a) holds.

For (b), since the quotient G/ZG/Z is affine, if MM is projective/injective then MM upon restriction to ZZ is projective/injective. Conversely, suppose that MM is a GG-module that is injective upon restriction to ZZ. Then indZGM\text{ind}_{Z}^{G}M is injective, and by the tensor identity MindZGindZGMM\cong\text{ind}_{Z}^{G}{\mathbb{C}}\cong\text{ind}_{Z}^{G}M. Now since ZZ is a splitting subgroup of GG, it follows that MM is a summand of indZGM\text{ind}_{Z}^{G}M, and is thus injective.

For (c), there exists a spectral sequence,

E2i,j=Ext(𝔤,𝔤0¯)i(M,NRjindZG)Ext(𝔷,𝔷0¯)i+j(M,N).E_{2}^{i,j}=\operatorname{Ext}^{i}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})}(M,N\otimes R^{j}\operatorname{ind}_{Z}^{G}{\mathbb{C}})\cong\operatorname{Ext}^{i+j}_{({\mathfrak{z}},{\mathfrak{z}}_{\bar{0}})}(M,N).

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 Ext(𝔤,𝔤0¯)n(M,NindZG)\operatorname{Ext}^{n}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})}(M,N\otimes\operatorname{ind}_{Z}^{G}{\mathbb{C}}). ∎

Serganova and Sherman proved that the detecting subalgebra 𝔣{\mathfrak{f}} 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 𝐓\mathbf{T} be a triangulated category. Recall that this means 𝐓\mathbf{T} is an additive category equipped with an auto-equivalence Σ:𝐓𝐓\Sigma:\mathbf{T}\longrightarrow\mathbf{T} called the shift, and a set of distinguished triangles:

MNQΣMM\longrightarrow N\longrightarrow Q\longrightarrow\Sigma M

all subject to a list of axioms the reader can find in [Nee01, Ch. 1].

A non-empty, full, additive subcategory 𝐒\mathbf{S} of 𝐓\mathbf{T} is called a triangulated subcategory if (i) M𝐒M\in\mathbf{S} implies that ΣnM𝐒\Sigma^{n}M\in\mathbf{S} for all nn\in\mathbb{Z} and (ii) if MNQΣMM\longrightarrow N\longrightarrow Q\longrightarrow\Sigma M is a distinguished triangle in 𝐓,\mathbf{T}, and two of {M,N,Q}\{M,N,Q\} are objects in 𝐒,\mathbf{S}, then the third object is also in 𝐒.\mathbf{S}. A triangulated subcategory 𝐒\mathbf{S} of 𝐓\mathbf{T} is called thick if 𝐒\mathbf{S} is closed under taking direct summands.

Assume that the triangulated category 𝐓\mathbf{T} admits set-indexed coproducts. A triangulated subcategory 𝐒\mathbf{S} of 𝐓\mathbf{T} is called a localizing subcategory if 𝐒\mathbf{S} is closed under taking set-indexed coproducts. It follows from the Eilenberg swindle that localizing subcategories are necessarily thick.

An object C𝐓C\in\mathbf{T} is called compact if Hom𝐓(C,)\operatorname{Hom}_{\mathbf{T}}(C,-) commutes with set-indexed coproducts. The full subcategory of compact objects in 𝐓\mathbf{T} is denoted by 𝐓c,\mathbf{T}^{c}, and the triangulated category 𝐓\mathbf{T} is said to be compactly generated if the isomorphism classes of compact objects form a set, and if for each non-zero object M𝐓M\in\mathbf{T} there is an object C𝐓cC\in\mathbf{T}^{c} such that Hom𝐓(C,M)0.\operatorname{Hom}_{\mathbf{T}}(C,M)\neq 0.

4.2. Central Ring Actions

Let 𝐓\mathbf{T} be a triangulated category with suspension Σ,\Sigma, and let RR be a graded-commutative ring. Recall that RR being graded-commutative means that RR admits a \mathbb{Z} grading and that for homogeneous elements x,yR,x,y\in R, xy=(1)|x||y|yx.xy=(-1)^{|x||y|}yx. In this section we recall what it means for RR to act on 𝐓.\mathbf{T}.

Let MM and NN be objects in 𝐓,\mathbf{T}, and set

Hom𝐓(M,N):=iHom𝐓(M,ΣiN).\operatorname{Hom}^{\bullet}_{\mathbf{T}}(M,N):=\bigoplus_{i\in\mathbb{Z}}\operatorname{Hom}_{\mathbf{T}}(M,\Sigma^{i}N).

Then Hom𝐓(M,N)\operatorname{Hom}^{\bullet}_{\mathbf{T}}(M,N) is a graded abelian group, and End𝐓(M):=Hom𝐓(M,M)\operatorname{End}^{\bullet}_{\mathbf{T}}(M):=\operatorname{Hom}^{\bullet}_{\mathbf{T}}(M,M) is a graded ring where the multiplication is “shift and compose”. Notice that Hom𝐓(M,N)\operatorname{Hom}^{\bullet}_{\mathbf{T}}(M,N) is a right End𝐓(M)\operatorname{End}^{\bullet}_{\mathbf{T}}(M) and a left End𝐓(N)\operatorname{End}^{\bullet}_{\mathbf{T}}(N)-bimodule.

The graded center of 𝐓\mathbf{T} is the graded-commutative ring denoted by Z(𝐓)Z^{\bullet}(\mathbf{T}) whose degree nn component is given by

Zn(𝐓)={η:Id𝐓Σn|ηΣ=(1)nΣη}.Z^{n}(\mathbf{T})=\{\eta:\text{Id}_{\mathbf{T}}\longrightarrow\Sigma^{n}\ |\ \eta\Sigma=(-1)^{n}\Sigma\eta\}.

With this setup, an action of RR on 𝐓\mathbf{T} is a homomorphism ϕ:RZ(𝐓)\phi:R\longrightarrow Z^{\bullet}(\mathbf{T}) of graded rings, and if 𝐓\mathbf{T} admits such an action, 𝐓\mathbf{T} is called an RR-linear triangulated category. It follows that if 𝐓\mathbf{T} is RR-linear, then for each object MM in 𝐓\mathbf{T} there is an induced homomorphism of graded rings ϕM:REnd𝐓(M)\phi_{M}:R\longrightarrow\operatorname{End}^{\bullet}_{\mathbf{T}}(M) such that the induced RR-module structures on Hom𝐓(M,N)\operatorname{Hom}^{\bullet}_{\mathbf{T}}(M,N) by ϕM\phi_{M} and ϕN\phi_{N} 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 (𝐊,,𝟏)(\mathbf{K},\otimes,\mathbf{1}) where 𝐊\mathbf{K} is a triangulated category, :𝐊×𝐊𝐊\otimes:\mathbf{K}\times\mathbf{K}\longrightarrow\mathbf{K} is a symmetric monoidal (tensor) product which is exact in each variable, and a monoidal unit 𝟏.\mathbf{1}.

Let 𝐊\mathbf{K} be a tensor triangulated category. With the additional structure of the tensor product on 𝐊\mathbf{K} one can define analogues to the usual notions in commutative algebra like prime ideal and spectrum. A tensor ideal in 𝐊\mathbf{K} is a triangulated subcategory 𝐈\mathbf{I} of 𝐊\mathbf{K} such that MN𝐈M\otimes N\in\mathbf{I} for all M𝐈M\in\mathbf{I} and N𝐊.N\in\mathbf{K}. A proper, thick tensor ideal 𝐏\mathbf{P} of 𝐊\mathbf{K} is said to be a prime ideal if MN𝐏M\otimes N\in\mathbf{P} implies that either M𝐏M\in\mathbf{P} or N𝐏.N\in\mathbf{P}. The Balmer spectrum [Bal05, Definition 2.1] is defined as

Spc(𝐊)={𝐏𝐊|𝐏 is a prime ideal}.\operatorname{Spc}(\mathbf{K})=\{\mathbf{P}\subsetneq\mathbf{K}\ |\ \mathbf{P}\text{ is a prime ideal}\}.

The Zariski topology on Spc(𝐊)\operatorname{Spc}(\mathbf{K}) has as closed sets

Z(𝒞)={𝐏Spc(𝐊)|𝒞𝐏},Z(\mathcal{C})=\{\mathbf{P}\in\operatorname{Spc}(\mathbf{K})\ |\ \mathcal{C}\cap\mathbf{P}\neq\emptyset\},

where 𝒞\mathcal{C} is any family of objects in 𝐊.\mathbf{K}.

The tensor triangulated category 𝐊\mathbf{K} is said to be a compactly generated TTC if 𝐊\mathbf{K} is closed under set indexed coproducts, the tensor product preserves set indexed coproducts, 𝐊\mathbf{K} is compactly generated as a triangulated category, the tensor product of compact objects is compact, 𝟏\mathbf{1} 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 𝐊,\mathbf{K}, 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 𝐊\mathbf{K} 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 RR-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 𝐊\mathbf{K} be a tensor triangulated category. Furthermore, assume that the unit object 𝟏\mathbf{1} is a compact generator for 𝐊\mathbf{K}. Then End𝐊(𝟏)\operatorname{End}_{\mathbf{K}}^{\bullet}(\mathbf{1}) is a graded commutative ring which induces an action on 𝐊\mathbf{K} by the homomorphisms

End𝐊(𝟏)MEnd𝐊(M)\operatorname{End}_{\mathbf{K}}^{\bullet}(\mathbf{1})\xrightarrow{M\otimes-}\operatorname{End}_{\mathbf{K}}^{\bullet}(M)

for each object M𝐊.M\in\mathbf{K}. Let RR be a graded-commutative ring. A canonical action of RR on 𝐊\mathbf{K} is an action of RR on 𝐊\mathbf{K} which is induced by a homomorphism of graded-commutative rings REnd𝐊(𝟏)R\longrightarrow\operatorname{End}_{\mathbf{K}}^{\bullet}(\mathbf{1}).

Let Proj(R)\operatorname{Proj}(R) denote the set of homogeneous prime ideals in R.R. One can construct a notion of support for objects of 𝐊\mathbf{K} 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 UProj(R)U\subseteq\operatorname{Proj}(R) is the set of all 𝔭Proj(R)\mathfrak{p}\in\operatorname{Proj}(R) for which there exists 𝔮U\mathfrak{q}\in U with 𝔮𝔭.\mathfrak{q}\subseteq\mathfrak{p}. A subset VProj(R)V\subseteq\operatorname{Proj}(R) is specialization closed if it equals its specialization closure. Given a specialization closed subset VProj(R)V\subseteq\operatorname{Proj}(R), the full subcategory of VV-torsion objects 𝐓V\mathbf{T}_{V} is defined as

𝐓V={M𝐓|Hom𝐓(C,M)𝔭=0for allC𝐓c,𝔭Proj(R)V}.\mathbf{T}_{V}=\{M\in\mathbf{T}\ |\ \operatorname{Hom}_{\mathbf{T}}^{\bullet}(C,M)_{\mathfrak{p}}=0\ \text{for all}\ C\in\mathbf{T}^{c},\ \mathfrak{p}\in\operatorname{Proj}(R)\setminus V\}.

It can be checked that 𝐓V\mathbf{T}_{V} is a localizing subcategory of 𝐓,\mathbf{T}, so there exists a localization functor LV:𝐓𝐓,L_{V}:\mathbf{T}\longrightarrow\mathbf{T}, and an induced colocalization functor ΓV,\mathit{\Gamma_{V}}, called the local cohomology functors of V,V, which for each object M𝐓M\in\mathbf{T} give a functorial triangle

ΓVMMLVM.\mathit{\Gamma}_{V}M\longrightarrow M\longrightarrow L_{V}M\longrightarrow.

For each 𝔭Proj(R)\mathfrak{p}\in\operatorname{Proj}(R), let VV and WW be specialization closed subsets of Proj(R)\operatorname{Proj}(R) such that {𝔭}=VW\{\mathfrak{p}\}=V\setminus W. There is a local-cohomology functor Γ𝔭:=ΓVLW\mathit{\Gamma_{\mathfrak{p}}}:=\mathit{\Gamma_{V}}L_{W} independent of the choice of VV and WW. These functors give rise to a notion of support for objects x𝐓,x\in\mathbf{T}, and a notion of support for 𝐓\mathbf{T} itself. In particular, set

suppR(x):={𝔭Proj(R)|Γ𝔭(x)0},\operatorname{supp}_{R}(x):=\{\mathfrak{p}\in\operatorname{Proj}(R)\ |\ \mathit{\Gamma_{\mathfrak{p}}}(x)\neq 0\},

and

suppR(𝐓):={𝔭Proj(R)|Γ𝔭(𝐓)0}.\operatorname{supp}_{R}(\mathbf{T}):=\{\mathfrak{p}\in\operatorname{Proj}(R)\ |\ \mathit{\Gamma_{\mathfrak{p}}}(\mathbf{T})\neq 0\}.

For each 𝔭Proj(R)\mathfrak{p}\in\operatorname{Proj}(R), the full subcategory Γ𝔭(𝐓)\mathit{\Gamma_{\mathfrak{p}}}(\mathbf{T}) is a tensor ideal and localizing. The main definition of this section is the concept of stratification.

Definition 4.4.1.

The RR-linear tensor triangulated category 𝐓\mathbf{T} is said to be stratified by RR if Γ𝔭(𝐓)\mathit{\Gamma_{\mathfrak{p}}}(\bf T) is either zero or a minimal tensor ideal localizing subcategory of 𝐓\mathbf{T} for all 𝔭Proj(R)\mathfrak{p}\in\operatorname{Proj}(R).

4.5.

For each localizing subcategory 𝐂\mathbf{C} of 𝐓,\mathbf{T}, set

σ(𝐂)=suppR(𝐂)={𝔭Proj(R)|Γ𝔭(𝐂)0},\sigma(\mathbf{C})=\operatorname{supp}_{R}(\mathbf{C})=\{\mathfrak{p}\in\operatorname{Proj}(R)\ |\ \mathit{\Gamma_{\mathfrak{p}}}(\mathbf{C})\neq 0\},

and for each subset VV of Proj(R)\operatorname{Proj}(R), set

τ(V)={X𝐓|suppRXV}.\tau(V)=\{X\in\mathbf{T}\ |\ \operatorname{supp}_{R}X\subseteq V\}.

The following theorems [BIK11a, Theorem 3.8] classify localizing subcategories and thick tensor ideals via the aforementioned maps.

Theorem 4.5.1.

Let 𝐊\mathbf{K} be a tensor triangulated category stratified by the action of R.R. Then the maps σ\sigma and τ\tau are mutually inverse bijections between the tensor ideal localizing subcategories of 𝐊\mathbf{K} and the subsets of suppR𝐊:\operatorname{supp}_{R}\mathbf{K}:

{tensor ideal localizing subcategories of 𝐊}\xleftrightarrow[τ]σ{subsets of suppR𝐊}.\{\text{tensor ideal localizing subcategories of }\mathbf{K}\}\xleftrightarrow[\tau]{\sigma}\{\text{subsets of }\operatorname{supp}_{R}\mathbf{K}\}.

In the case when 𝐊\mathbf{K} is stratified by a noetherian ring RR, the maps σ\sigma and τ\tau also give a classification of thick tensor ideals in 𝐊c:\mathbf{K}^{c}:

Theorem 4.5.2.

Let 𝐊\mathbf{K} be a compactly generated tensor triangulated category which is stratified by the noetherian ring RR. Then the maps σ\sigma and τ\tau are mutually inverse bijections between the thick tensor ideal subcategories of 𝐊c\mathbf{K}^{c} and the specialization closed subsets of suppR𝐊\operatorname{supp}_{R}\mathbf{K}:

{thick tensor ideal subcategories of 𝐊c}\xleftrightarrow[τ]σ{specialization closed subsets of suppR𝐊}.\{\text{thick tensor ideal subcategories of }\mathbf{K}^{c}\}\xleftrightarrow[\tau]{\sigma}\{\text{specialization closed subsets of }\operatorname{supp}_{R}\mathbf{K}\}.

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 AA be a commutative Noetherian ring and let D(A)D(A) be the derived category of differential graded AA-modules. Then D(A)D(A) is stratified by the canonical action of A.A.

4.7. Lie Superalgebras of the Form 𝔷=𝔷0¯𝔷1¯\mathfrak{z}=\mathfrak{z}_{\overline{0}}\oplus\mathfrak{z}_{\overline{1}}

Throughout this section 𝔷=𝔷0¯𝔷1¯\mathfrak{z}=\mathfrak{z}_{\overline{0}}\oplus\mathfrak{z}_{\overline{1}} denotes a Lie superalgebra where 𝔷0¯=𝔱\mathfrak{z}_{\overline{0}}=\mathfrak{t} is a torus, and [𝔷0¯,𝔷1¯]=0[\mathfrak{z}_{\overline{0}},\mathfrak{z}_{\overline{1}}]=0. This is a classical Lie superalgebra. As an example, 𝔷\mathfrak{z} could be any of the detecting subalgebras introduced in [BKN10a, Section 4.4]. Let 𝐊=Stab(𝒞(𝔷,𝔷0¯))\mathbf{K}=\operatorname{Stab}(\mathcal{C}_{(\mathfrak{z},\mathfrak{z}_{\overline{0}})}) be the stable module category whose objects are all 𝔷\mathfrak{z}-modules which are finitely semi-simple as 𝔷0¯\mathfrak{z}_{\overline{0}}-modules. Then 𝐊\mathbf{K} is a compactly generated tensor-triangulated category with full subcategory of compact objects 𝐊c=stab((𝔷,𝔷0¯))\mathbf{K}^{c}=\operatorname{stab}(\mathcal{F}_{(\mathfrak{z},\mathfrak{z}_{\overline{0}})}). Note that the objects of 𝐊c\mathbf{K}^{c} consist precisely of the finite-dimensional objects in 𝐊\mathbf{K}. Let RR denote the cohomology ring H(𝔷,𝔷0¯;)\operatorname{H}^{*}(\mathfrak{z},\mathfrak{z}_{\overline{0}};\mathbb{C}). According to [BKN17, Section 4.3],

H(𝔷,𝔷0¯;)S(𝔷1¯)\operatorname{H}^{\bullet}(\mathfrak{z},\mathfrak{z}_{\overline{0}};\mathbb{C})\cong S^{\bullet}({\mathfrak{z}_{\overline{1}}}^{*})

and thus RR is a polynomial algebra. The following theorem was established in [BKN17, Theorem 4.5.4].

Theorem 4.7.1.

The thick tensor ideals in stab((𝔤,𝔤0¯))\operatorname{stab}({\mathcal{F}}_{(\mathfrak{g},\mathfrak{g}_{\bar{0}})}) are in bijective correspondence with specialization closed subsets of Proj(S(𝔷1¯)).\operatorname{Proj}(S^{\bullet}({\mathfrak{z}_{\overline{1}}}^{*})). Furthermore, Spc(stab((𝔤,𝔤0¯)))\operatorname{Spc}(\operatorname{stab}({\mathcal{F}}_{(\mathfrak{g},\mathfrak{g}_{\bar{0}})})) is homeomorphic to Proj(S(𝔷1¯))\operatorname{Proj}(S^{\bullet}({\mathfrak{z}_{\overline{1}}}^{*})).

In [BKN17] the authors point out that 𝐊\mathbf{K} is an RR-linear triangulated category, and that the local-global principle holds. It was conjectured that RR stratifies 𝐊,\mathbf{K}, 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 Stab(𝒞(𝔷,𝔷0¯))\operatorname{Stab}(\mathcal{C}_{(\mathfrak{z},\mathfrak{z}_{\overline{0}})}) is stratified by the action of the cohomology ring H(𝔷,𝔷0¯;)S(𝔷1¯)\operatorname{H}^{\bullet}(\mathfrak{z},\mathfrak{z}_{\overline{0}};\mathbb{C})\cong S^{\bullet}({\mathfrak{z}_{\overline{1}}}^{*}).

Proof.

The following argument follows closely that of [BIK11a, Section 5.2] where the category Stab(kE)\text{Stab}(kE) is considered, where kk is an algebraically closed field of characteristic two, and EE is an elementary abelian 22-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 𝔷/𝔷0¯{\mathfrak{z}}/{\mathfrak{z}}_{\bar{0}} are exterior algebras.

Consider the universal enveloping superalgebra of the quotient U(𝔷/𝔷0¯)U(\mathfrak{z}/\mathfrak{z}_{\overline{0}}). Then there is an isomorphism of \mathbb{C}-algebras U(𝔷/𝔷0¯)Λ(𝔷1¯).U(\mathfrak{z}/\mathfrak{z}_{\overline{0}})\cong\Lambda^{\bullet}(\mathfrak{z}^{*}_{\overline{1}}). Therefore, there is an isomorphism of rings Λ(𝔷1¯)[z1,,zr]/(zi2)\Lambda^{\bullet}(\mathfrak{z}^{*}_{\overline{1}})\cong\mathbb{C}[z_{1},\dots,z_{r}]/(z_{i}^{2}). Set R=H(𝔷,𝔷0¯,)R=\operatorname{H}^{\bullet}({\mathfrak{z}},{\mathfrak{z}}_{\bar{0}},\mathbb{C}). Choose a basis {y1,,yr}\{y_{1},\dots,y_{r}\} of 𝔷1¯\mathfrak{z}^{*}_{\overline{1}} so that R[y1,,yr]R\cong\mathbb{C}[y_{1},\dots,y_{r}] is an isomorphism of rings, and view RR as a differential graded algebra with zero differential and |yi|=1|y_{i}|=1 for each ii.

The \mathbb{C}-algebra U(𝔷/𝔷0¯)RU(\mathfrak{z}/\mathfrak{z}_{\overline{0}})\otimes_{\mathbb{C}}R is graded with degree ii component U(𝔷/𝔷0¯)RiU(\mathfrak{z}/\mathfrak{z}_{\overline{0}})\otimes_{\mathbb{C}}R^{i} and with multiplication defined by (as)(bt)=abst.(a\otimes s)(b\otimes t)=ab\otimes st. Consider U(𝔷/𝔷0¯)RU(\mathfrak{z}/\mathfrak{z}_{\overline{0}})\otimes_{\mathbb{C}}R as a differential graded algebra with zero differential. The degree one element δ\delta defined as

δ=i=1rziyi.\delta=\sum_{i=1}^{r}z_{i}\otimes_{\mathbb{C}}y_{i}.

satisfied δ2=0.\delta^{2}=0. Let JJ denote the differential graded module over U(𝔷/𝔷0¯)RU(\mathfrak{z}/\mathfrak{z}_{\overline{0}})\otimes_{\mathbb{C}}R with graded module and differential given by

J=U(𝔷/𝔷0¯)R,d(e)=δe.J=U(\mathfrak{z}/\mathfrak{z}_{\overline{0}})\otimes_{\mathbb{C}}R,\ d(e)=\delta e.

Since JJ is a differential graded module over U(𝔷/𝔷0¯)R,U(\mathfrak{z}/\mathfrak{z}_{\overline{0}})\otimes_{\mathbb{C}}R, for each differential graded module MM over U(𝔷/𝔷0¯)U(\mathfrak{z}/\mathfrak{z}_{\overline{0}}) there is an induced structure of a differential graded RR-module on HomU(𝔷/𝔷0¯)(J,M).\operatorname{Hom}_{{U}(\mathfrak{z}/\mathfrak{z}_{\overline{0}})}(J,M). Then the functor

HomU(𝔷/𝔷0¯)(J,):K(Inj𝒞(𝔷,𝔷0¯))D(R)\operatorname{Hom}_{U(\mathfrak{z}/\mathfrak{z}_{\overline{0}})}(J,-):K(\operatorname{Inj}\mathcal{C}_{(\mathfrak{z},\mathfrak{z}_{\overline{0}})})\longrightarrow D(R)

is an equivalence of triangulated categories.

To see this first observe that as a complex, JJ consists of injective U(𝔷/𝔷0¯)U(\mathfrak{z}/\mathfrak{z}_{\overline{0}})-modules. This follows from the fact that U(𝔷/𝔷0¯)U(\mathfrak{z}/\mathfrak{z}_{\overline{0}}) is self-injective. That the RR actions coincide follows from [BIK11a, Theorem 5.4], since the R-action on K(Inj𝒞(𝔷,𝔷0¯))K(\operatorname{Inj}\mathcal{C}_{(\mathfrak{z},\mathfrak{z}_{\overline{0}})}) is also canonical. Finally, the result follows from the equivalence of categories Stab(𝒞(𝔷,𝔷0¯))Kac(Inj𝒞(𝔷,𝔷0¯))\operatorname{Stab}(\mathcal{C}_{(\mathfrak{z},\mathfrak{z}_{\overline{0}})})\cong K_{\text{ac}}(\operatorname{Inj}\mathcal{C}_{(\mathfrak{z},\mathfrak{z}_{\overline{0}})}) 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 𝔷{\mathfrak{z}}.

Theorem 4.8.1.

Let 𝔷=𝔷0¯𝔷1¯\mathfrak{z}=\mathfrak{z}_{\overline{0}}\oplus\mathfrak{z}_{\overline{1}} denotes a Lie superalgebra where 𝔷0¯=𝔱\mathfrak{z}_{\overline{0}}=\mathfrak{t} is a torus, and [𝔷0¯,𝔷1¯]=0[\mathfrak{z}_{\overline{0}},\mathfrak{z}_{\overline{1}}]=0.

  • (a)

    There is an equivalence of triangulated categories K(Inj𝒞(𝔷,𝔷0¯))D(S(𝔷1¯))K(\operatorname{Inj}\mathcal{C}_{(\mathfrak{z},\mathfrak{z}_{\overline{0}})})\cong D(S^{\bullet}({\mathfrak{z}_{\overline{1}}}^{*})).

  • (b)

    The localizing subcategories of Stab(𝒞(𝔷,𝔷0¯))\operatorname{Stab}(\mathcal{C}_{(\mathfrak{z},\mathfrak{z}_{\overline{0}})}) are in bijective correspondence with subsets of Proj(S(𝔷1¯))\operatorname{Proj}(S^{\bullet}({\mathfrak{z}_{\overline{1}}}^{*})).

  • (c)

    The thick tensor ideals in stab((𝔤,𝔤0¯))\operatorname{stab}({\mathcal{F}}_{(\mathfrak{g},\mathfrak{g}_{\bar{0}})}) are in bijective correspondence with specialization closed subsets of Proj(S(𝔷1¯))\operatorname{Proj}(S^{\bullet}({\mathfrak{z}_{\overline{1}}}^{*})).

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, 𝐊\mathbf{K} will denote a tensor triangulated category satisfying the assumptions from [Bal20, 2.2], namely 𝐊\mathbf{K} is essentially small and rigid. At times for our purposes we view 𝐊\mathbf{K} as sitting inside of a “large” compactly tensor triangulated category 𝐓\mathbf{T} with 𝐊=𝐓c.\mathbf{K}=\mathbf{T}^{c}. This is in anticipation of applying the “Balmer machine” to the categories stab((𝔤,𝔤0¯))\operatorname{stab}(\mathcal{F}_{(\mathfrak{g},\mathfrak{g}_{\overline{0}})}) and Stab(𝒞(𝔤,𝔤0¯))\operatorname{Stab}(\mathcal{C}_{(\mathfrak{g},\mathfrak{g}_{\overline{0}})}) which are the main point of study for this article.

5.1. The Homological Spectrum

The Grothendieck abelian category of right 𝐊\mathbf{K}-modules denoted 𝐀=Mod\mathbf{A}=\operatorname{Mod}-𝐊\mathbf{K} has as objects the additive functors M:𝐊opAbM:\mathbf{K}^{\text{op}}\longrightarrow\text{Ab} from the opposite category of 𝐊\mathbf{K} to the category of abelian groups, and has as morphisms the natural transformations between functors. Let h denote the Yoneda embedding:

h:𝐊\displaystyle\text{h}:\ \mathbf{K} Mod-𝐊=Add(𝐊op,Ab)\displaystyle\hookrightarrow\operatorname{Mod}\text{-}\mathbf{K}=\operatorname{Add}(\mathbf{K}^{\text{op}},\text{Ab})
x\displaystyle x x^:=Hom𝐊(,x)\displaystyle\mapsto\hat{x}:=\operatorname{Hom}_{\mathbf{K}}(-,x)
f\displaystyle f f^.\displaystyle\mapsto\hat{f}.

For details about 𝐊\mathbf{K}-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 Mod\operatorname{Mod}-𝐊\mathbf{K} admits a tensor product :Mod\otimes:\operatorname{Mod}-𝐊Mod\mathbf{K}\longrightarrow\operatorname{Mod}-𝐊\mathbf{K} which is right-exact in each variable and which makes h a tensor functor, (ii) h preserves rigidity, (iii) x^\hat{x} is compact, projective, and \otimes-flat in 𝐀,\mathbf{A}, and (iv) the tensor subcategory 𝐀c=mod\mathbf{A}^{c}=\operatorname{mod}-𝐊\mathbf{K} of compact objects is abelian and is the Freyd envelope of 𝐊\mathbf{K} (i.e., h:𝐊mod\text{h}:\mathbf{K}\longrightarrow\operatorname{mod}-𝐊\mathbf{K} is the universal homological functor out of 𝐊\mathbf{K}.) 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 𝐁\mathbf{B} 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 𝐊\mathbf{K} is a maximal proper Serre \otimes-ideal subcategory 𝐁\mathbf{B} of 𝐀c=mod\mathbf{A}^{c}=\operatorname{mod}-𝐊\mathbf{K} of the Freyd envelope of 𝐊.\mathbf{K}.

  • (b)

    The homological spectrum of 𝐊,\mathbf{K}, denoted Spch(𝐊)\operatorname{Spc}^{\text{h}}(\mathbf{K}) as a set consists of all the homological primes of 𝐊:\mathbf{K}:

    Spch(𝐊)={𝐁mod-𝐊|𝐁 is maximal proper Serre ideal},\operatorname{Spc}^{\text{h}}(\mathbf{K})=\{\mathbf{B}\subsetneq\operatorname{mod}\text{-}\mathbf{K}\ |\ \mathbf{B}\text{ is maximal proper Serre ideal}\},

    and has as topology that generated by the basis of closed subsets supph(x)\operatorname{supp}^{\text{h}}(x) for all x𝐊,x\in\mathbf{K}, where

    supph(x)={𝐁Spch(𝐊)|x^𝐁}.\operatorname{supp}^{\text{h}}(x)=\{\mathbf{B}\in\operatorname{Spc}^{\text{h}}(\mathbf{K})\ |\ \hat{x}\notin\mathbf{B}\}.

It should be noted that one can check that (Spch,supph)(\operatorname{Spc}^{\text{h}},\operatorname{supp}^{\text{h}}) is a support data on 𝐊.\mathbf{K}. Therefore, there exists a unique continuous map

ϕ:Spch(𝐊)Spc(𝐊)\phi:\operatorname{Spc}^{\text{h}}(\mathbf{K})\longrightarrow\operatorname{Spc}(\mathbf{K})

such that supph(x)=ϕ1(supp(x))\operatorname{supp}^{\text{h}}(x)=\phi^{-1}(\operatorname{supp}(x)) for all x𝐊.x\in\mathbf{K}. See [Bal05, Thm. 3.2]. This map ϕ\phi is often called the comparison map, and is surjective as long as 𝐊\mathbf{K} is rigid. When 𝐊\mathbf{K} 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 ϕ\phi 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 𝐊,\mathbf{K}, how does one construct functors F:𝐊𝐅F:\mathbf{K}\longrightarrow\mathbf{F} 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 𝐅\mathbf{F} is a tensor triangulated field if every object of 𝐅\mathbf{F} is a coproduct of compact-rigid objects of 𝐅c,\mathbf{F}^{c}, and if every non-zero object in 𝐅\mathbf{F} 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 𝐁Spch(𝐊),\mathbf{B}\in\operatorname{Spc}^{\text{h}}(\mathbf{K}), the homological residue field corresponding to 𝐁\mathbf{B} is the functor

h¯𝐁=Q𝐁h:𝐊𝐀=Mod-𝐊𝐀¯(𝐊,𝐁):=Mod-𝐊𝐁\overline{\text{h}}_{\mathbf{B}}=Q_{\mathbf{B}}\circ\text{h}:\mathbf{K}\hookrightarrow\mathbf{A}=\operatorname{Mod}\text{-}\mathbf{K}\twoheadrightarrow\overline{\mathbf{A}}(\mathbf{K},\mathbf{B}):=\frac{\operatorname{Mod}\text{-}\mathbf{K}}{\langle\mathbf{B}\rangle}

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 𝐓,\mathbf{T}, a tensor-triangulated field 𝐅,\mathbf{F}, and a monoidal exact functor F:𝐓𝐅F:\mathbf{T}\longrightarrow\mathbf{F} with right adjoint U,U, one has the following diagram:

𝐓{{\mathbf{T}}}Mod-𝐓c{{\text{Mod-}\mathbf{T}^{c}}}Mod-𝐓c/Ker(F^)=𝐀¯𝐁{{\text{Mod-}\mathbf{T}^{c}/\operatorname{Ker}(\hat{F})=\overline{\mathbf{A}}_{\mathbf{B}}}}𝐅{{\mathbf{F}}}Mod-𝐅c{{\text{Mod-}\mathbf{F}^{c}}}F\scriptstyle{F}U\scriptstyle{U}hF¯\scriptstyle{\overline{F}}hF^\scriptstyle{\hat{F}}U^\scriptstyle{\hat{U}}U¯\scriptstyle{\overline{U}}Q\scriptstyle{Q}R\scriptstyle{R}

where F^\hat{F} is the exact cocontinuous functor induced by F,F, the functor QQ is the Gabriel quotient with respect to Ker(F^)\operatorname{Ker}(\hat{F}) and the functor F¯\overline{F} is induced by the universal property, hence F^=F¯Q\hat{F}=\overline{F}Q and F¯\overline{F} is exact and faithful.

The adjunctions FU,F^U^,F¯U¯,F\dashv U,\hat{F}\dashv\hat{U},\overline{F}\dashv\overline{U}, and QR,Q\dashv R, are depicted with F^h=hF\hat{F}h=hF and U^h=hU\hat{U}h=hU. Moreover, 𝐁:=Ker(F^)𝐀fp\mathbf{B}:=\operatorname{Ker}(\hat{F})\cap\mathbf{A}^{fp} is a homological prime and Ker(F^)=𝐁\operatorname{Ker}(\hat{F})=\langle\mathbf{B}\rangle and h¯𝐁=Qh:𝐓𝐀¯𝐁\overline{\text{h}}_{\mathbf{B}}=Q\circ\text{h}:\mathbf{T}\longrightarrow\overline{\mathbf{A}}_{\mathbf{B}} is a homological residue field of 𝐓\mathbf{T}.

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 MM and NN be modules in 𝒞(𝔤,𝔤0¯)\mathcal{C}_{(\mathfrak{g},\mathfrak{g}_{\overline{0}})}.

  • (a)

    A map f:MNf:M\longrightarrow N is called null if f=0f=0 in Stab(𝒞(𝔤,𝔤0¯))\operatorname{Stab}(\mathcal{C}_{(\mathfrak{g},\mathfrak{g}_{\overline{0}})}); i.e., ff is null if and only if ff factors through a projective module.

  • (b)

    A map f:MNf:M\longrightarrow N is called tensor nilpotent if there exists some n0n\in\mathbb{Z}_{\geq 0} such that the tensor power fn:MnNnf^{\otimes n}:M^{\otimes n}\longrightarrow N^{\otimes n} is null.

In the case when MM is compact, one can transform the condition of the nilpotence of the map ff to the adjoint map.

Lemma 5.3.2.

Let MM be a compact object. A map f:MNf:M\longrightarrow N is tensor nilpotent if and only if the adjoint map f^:kMN\hat{f}:k\longrightarrow M^{*}\otimes N is tensor nilpotent.

Proof.

By adjointness, we have an isomorphism Hom(M,N)Hom(k,MN)\operatorname{Hom}(M,N)\cong\operatorname{Hom}(k,M^{*}\otimes N). Since ff is tensor nilpotent, there exists some nn such that fn:MnNnf^{\otimes n}:M^{\otimes n}\longrightarrow N^{\otimes n} factors through a projective. But since tensor products of projective modules are projective, this implies that f^n\hat{f}^{\otimes n} factors through a projective; i.e., that f^\hat{f} is tensor nilpotent. ∎

Next we need to recall the definition of a colimit in the category 𝒞(𝔤,𝔤0¯)\mathcal{C}_{(\mathfrak{g},\mathfrak{g}_{\bar{0}})} and a homotopy colimit in its stable module category.

Definition 5.3.3.

Let 𝔤{\mathfrak{g}} be a classical Lie superalgebra.

  • (a)

    Let

    θ:N1f1N2f2N3f3\theta:N_{1}\overset{f_{1}}{\longrightarrow}N_{2}\overset{f_{2}}{\longrightarrow}N_{3}\overset{f_{3}}{\longrightarrow}\dots

    be a system of modules and homomorphisms in 𝒞(𝔤,𝔤0¯)\mathcal{C}_{(\mathfrak{g},\mathfrak{g}_{\bar{0}})}. Let γ:i=1Nii=1Ni\gamma:\bigoplus_{i=1}^{\infty}N_{i}\longrightarrow\bigoplus_{i=1}^{\infty}N_{i} be defined by γ(m)=mfi(m)\gamma(m)=m-f_{i}(m) whenever mNim\in N_{i}. The colimit of the system, if ti exists, is the module given by coker γ\text{coker }\gamma.

  • (b)

    Let

    θ:X1f1X2f2X3f3\theta:X_{1}\overset{f_{1}}{\longrightarrow}X_{2}\overset{f_{2}}{\longrightarrow}X_{3}\overset{f_{3}}{\longrightarrow}\dots

    be a system of modules and homomorphisms in Stab(𝒞(𝔤,𝔤0¯))\operatorname{Stab}(\mathcal{C}_{(\mathfrak{g},\mathfrak{g}_{\overline{0}})}). The homotopy colimit of the system is the module obtained by completing the map

    Xi1fXi\bigoplus X_{i}\overset{1-f}{\longrightarrow}\bigoplus X_{i}

    to a triangle:

    Xi1fXihocolim(Xi).\bigoplus X_{i}\overset{1-f}{\longrightarrow}\bigoplus X_{i}\longrightarrow\text{hocolim}(X_{i})\longrightarrow.

The following lemmas are given in [Ric97] and are modified here for Lie superalgebra representations.

Lemma 5.3.4.

Let X1α1X2α2X_{1}\overset{\alpha_{1}}{\longrightarrow}X_{2}\overset{\alpha_{2}}{\longrightarrow}\cdots be a sequence of maps in a triangulated category with countable direct sums. If for each i>0i>0 there exists k>ik>i such that α1αk=0\alpha_{1}\dots\alpha_{k}=0, then hocolim(Xi)0\text{hocolim}(X_{i})\cong 0.

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 X1α1X2α2X_{1}\overset{\alpha_{1}}{\longrightarrow}X_{2}\overset{\alpha_{2}}{\longrightarrow}\cdots be a sequence of modules and homomorphisms in Stab(𝒞(𝔤,𝔤0¯))\operatorname{Stab}(\mathcal{C}_{(\mathfrak{g},\mathfrak{g}_{\overline{0}})}). The colimit colim(Xi)\text{colim}(X_{i}) in 𝒞(𝔤,𝔤0¯)\mathcal{C}_{(\mathfrak{g},\mathfrak{g}_{\overline{0}})} is isomorphic in Stab(𝒞(𝔤,𝔤0¯))\operatorname{Stab}(\mathcal{C}_{(\mathfrak{g},\mathfrak{g}_{\overline{0}})}) to the homotopy colimit hocolim(Xi)\text{hocolim}(X_{i}).

These two lemmas together allow one to derive an analog of [BenC18, Lemma 2.3]

Theorem 5.3.6.

A map f:kNf:k\longrightarrow N is \otimes-nilpotent if and only if the colimit of

k𝑓Nf1NNf11NNNk\overset{f}{\longrightarrow}N\overset{f\otimes 1}{\longrightarrow}N\otimes N\overset{f\otimes 1\otimes 1}{\longrightarrow}N\otimes N\otimes N\longrightarrow\cdots

is projective.

Proof.

First suppose that f:kNf:k\longrightarrow N is \otimes-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 ff. ∎

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 KK-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, π\pi-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 𝐊\mathbf{K} sits inside of a big rigidly compactly generated tensor triangulated category 𝐓\mathbf{T} with 𝐊=𝐓c,\mathbf{K}=\mathbf{T}^{c}, 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 𝐓\mathbf{T} be a rigidly-compactly generated “big” tensor-triangulated category with 𝐊=𝐓c.\mathbf{K}=\mathbf{T}^{c}. Let f:xYf:x\longrightarrow Y be a morphism in 𝐓\mathbf{T} with x𝐊x\in\mathbf{K} compact and YY arbitrary. If h¯(f)=0\overline{\emph{h}}(f)=0 in 𝐀¯(𝐊;𝐁)\overline{\mathbf{A}}(\mathbf{K};\mathbf{B}) for every homological residue field h¯𝐁\overline{\emph{h}}_{\mathbf{B}} for every homological prime 𝐁mod\mathbf{B}\subsetneq\operatorname{mod}-𝐊,\mathbf{K}, then there exists n1n\geq 1 such that fn=0f^{\otimes n}=0 in 𝐓.\mathbf{T}.

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 𝔤=𝔤0¯𝔤1¯\mathfrak{g}=\mathfrak{g}_{\bar{0}}\oplus\mathfrak{g}_{\bar{1}} be a classical Lie superalgebra, and let f:MNf:M\longrightarrow N be a morphism in Stab(𝒞(𝔤,𝔤0¯))\operatorname{Stab}(\mathcal{C}_{(\mathfrak{g},\mathfrak{g}_{\overline{0}})}) with Mstab((𝔤,𝔤0¯)).M\in\operatorname{stab}(\mathcal{F}_{(\mathfrak{g},\mathfrak{g}_{\overline{0}})}). Suppose that h¯𝐁(f)=0\overline{\emph{h}}_{\mathbf{B}}(f)=0 for all 𝐁Spch(stab((𝔤,𝔤0¯))).\mathbf{B}\in\operatorname{Spc}^{\emph{h}}(\operatorname{stab}(\mathcal{F}_{(\mathfrak{g},\mathfrak{g}_{\overline{0}})})). Then there exists n1n\geq 1 such that fn=0f^{\otimes n}=0 in Stab(𝒞(𝔤,𝔤0¯)).\operatorname{Stab}(\mathcal{C}_{(\mathfrak{g},\mathfrak{g}_{\overline{0}})}).

Proof.

The first observation is that stab((𝔤,𝔤0¯))\operatorname{stab}(\mathcal{F}_{(\mathfrak{g},\mathfrak{g}_{\overline{0}})}) sits inside of Stab(𝒞(𝔤,𝔤0¯))\operatorname{Stab}(\mathcal{C}_{(\mathfrak{g},\mathfrak{g}_{\overline{0}})}) as the compact objects: stab((𝔤,𝔤0¯))=(Stab(𝒞(𝔤,𝔤0¯)))c.\operatorname{stab}(\mathcal{F}_{(\mathfrak{g},\mathfrak{g}_{\overline{0}})})=(\operatorname{Stab}(\mathcal{C}_{(\mathfrak{g},\mathfrak{g}_{\overline{0}})}))^{c}. Moreover, the compact objects and the rigid objects coincide and generate Stab(𝒞(𝔤,𝔤0¯))\operatorname{Stab}(\mathcal{C}_{(\mathfrak{g},\mathfrak{g}_{\overline{0}})}) 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 f:MNf:M\longrightarrow N is nilpotent in the big stable module category where MM 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 𝔤=𝔤0¯𝔤1¯\mathfrak{g}=\mathfrak{g}_{\bar{0}}\oplus\mathfrak{g}_{\bar{1}} be a classical Lie subalgebra with a splitting subalgebra 𝔷=𝔷0¯𝔷1¯𝔤.\mathfrak{z}=\mathfrak{z}_{\bar{0}}\oplus\mathfrak{z}_{\bar{1}}\subseteq\mathfrak{g}. Let f:MNf:M\longrightarrow N be a morphism in Stab(𝒞(𝔤,𝔤0¯))\operatorname{Stab}(\mathcal{C}_{(\mathfrak{g},\mathfrak{g}_{\overline{0}})}) with Mstab((𝔤,𝔤0¯))M\in\operatorname{stab}(\mathcal{F}_{(\mathfrak{g},\mathfrak{g}_{\overline{0}})}). If h¯𝐁(res(f))=0\overline{\emph{h}}_{\mathbf{B}}(\operatorname{res}(f))=0 for all 𝐁Spch(stab((𝔷,𝔷0¯))\mathbf{B}\in\operatorname{Spc}^{\emph{h}}(\operatorname{stab}(\mathcal{F}_{(\mathfrak{z},\mathfrak{z}_{\overline{0}})}) then there exists n1n\geq 1 such that fn=0f^{\otimes n}=0 in Stab(𝒞(𝔤,𝔤0¯)).\operatorname{Stab}(\mathcal{C}_{(\mathfrak{g},\mathfrak{g}_{\overline{0}})}).

Proof.

Let res:Stab(𝒞(𝔤,𝔤0¯))Stab(𝒞(𝔷,𝔷0¯))\text{res}:\operatorname{Stab}(\mathcal{C}_{(\mathfrak{g},\mathfrak{g}_{\overline{0}})})\longrightarrow\operatorname{Stab}(\mathcal{C}_{(\mathfrak{z},\mathfrak{z}_{\overline{0}})}) be the usual restriction functor. By our hypothesis, h¯𝐁(res(f))=0\overline{\text{h}}_{\mathbf{B}}(\operatorname{res}(f))=0 for all 𝐁Spch(stab((𝔷,𝔷0¯))\mathbf{B}\in\operatorname{Spc}^{\emph{h}}(\operatorname{stab}(\mathcal{F}_{(\mathfrak{z},\mathfrak{z}_{\overline{0}})}), Theorem 5.4.2 implies that res(f)\operatorname{res}(f) is tensor nilpotent in Stab(𝒞(𝔷,𝔷0¯))\operatorname{Stab}(\mathcal{C}_{(\mathfrak{z},\mathfrak{z}_{\overline{0}})}).

It follows that res(f^):MN\text{res}(\widehat{f}):{\mathbb{C}}\rightarrow M^{*}\otimes N is tensor nilpotent in Stab(𝒞(𝔷,𝔷0¯))\operatorname{Stab}(\mathcal{C}_{(\mathfrak{z},\mathfrak{z}_{\overline{0}})}), and by Theorem 5.3.6 its associated colimit is projective in 𝒞(𝔷,𝔷0¯)\mathcal{C}_{({\mathfrak{z}},{\mathfrak{z}}_{\bar{0}})}. Therefore, the colimit as an object in 𝒞(𝔤,𝔤0¯)\mathcal{C}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})} is projective by Theorem 3.4.2(b). Invoking Theorem 5.3.6 again implies that f^\widehat{f} is tensor nilpotent, thus ff is tensor nilpotent. ∎

6. Identifying the Homological Spectrum

The goals of this section are to determine the homological spectrum for stab((𝔷,𝔷0¯))\operatorname{stab}(\mathcal{F}_{(\mathfrak{z},\mathfrak{z}_{\overline{0})}}) and stab((𝔤,𝔤0¯)),\operatorname{stab}(\mathcal{F}_{(\mathfrak{g},\mathfrak{g}_{\overline{0}})}), where 𝔤\mathfrak{g} is a classical Lie superalgebra with splitting subalgebra 𝔷.\mathfrak{z}. 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 𝔷=𝔷0¯𝔷1¯\mathfrak{z}=\mathfrak{z}_{\bar{0}}\oplus\mathfrak{z}_{\bar{1}} be a Type I classical Lie superalgebra with 𝔷0¯\mathfrak{z}_{\bar{0}} a torus and [𝔷0¯,𝔷1¯]=0.[\mathfrak{z}_{\bar{0}},\mathfrak{z}_{\bar{1}}]=0. Then the comparison map

ϕ:Spch(stab((𝔷,𝔷0¯))Spc(stab((𝔷,𝔷0¯)))\phi:\operatorname{Spc}^{h}(\operatorname{stab}(\mathcal{F}_{(\mathfrak{z},\mathfrak{z}_{\overline{0}})})\longrightarrow\operatorname{Spc}(\operatorname{stab}(\mathcal{F}_{(\mathfrak{z},\mathfrak{z}_{\overline{0}})}))

is a bijection.

Proof.

Since (𝔷,𝔷0¯)\mathcal{F}_{({\mathfrak{z}},{\mathfrak{z}}_{\bar{0}})} is rigid, the map ϕ\phi is surjective. In order to prove that the map is injective one can use the argument outlined in [Bal20, Example 5.13]. The main point is to use the classification of localizing subcategories of 𝒞(𝔷,𝔷0¯)\mathcal{C}_{({\mathfrak{z}},{\mathfrak{z}}_{\bar{0}})} and the existence of pure injective objects. See also [BKS19, Corollary 4.26]. ∎

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 𝔤{\mathfrak{g}} be a classical Lie superalgebra and 𝔷{\mathfrak{z}} be a detecting subalgebra in 𝔤{\mathfrak{g}}. We will need to work with a field extension KK of {\mathbb{C}} such that the transcendence degree is larger than the dimension of 𝔷{\mathfrak{z}}. 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 KK. Let 𝒫x{\mathcal{P}}_{x} be the prime ideal in Proj(S(𝔷1¯))\text{Proj}(S^{\bullet}({\mathfrak{z}}_{\bar{1}})) associated with the “generic point” xx (cf. [BCR96, Sections 2 and 3] for an explanation).

For x𝔷1¯x\in{\mathfrak{z}}_{\bar{1}} with 𝔷1¯{\mathfrak{z}}_{\bar{1}} viewed as a vector space over KK, one has U(x.)U(\langle x.\rangle) is either 𝕂[x]/(x2){\mathbb{K}}[x]/(x^{2}) or U(𝔮(1))U({\mathfrak{q}}(1)). In either case, the blocks are either semisimple or have finite representation type. One can verify that Stab(C(x,x0¯))\text{Stab}(C_{(\langle x\rangle,\langle x\rangle_{\bar{0}})}) is a tensor triangular field. For x𝔷1¯x\in{\mathfrak{z}}_{\bar{1}}, one have two monoidal exact functors (given by restriction):

πx𝔤:Stab(C(𝔤,𝔤0¯))Stab(C(x,x0¯))\pi_{x}^{\mathfrak{g}}:\text{Stab}(C_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})})\rightarrow\text{Stab}(C_{(\langle x\rangle,\langle x\rangle_{\bar{0}})}) (6.2.1)
πx𝔷:Stab(C(𝔷,𝔷0¯))Stab(C(x,x0¯))\pi_{x}^{\mathfrak{z}}:\text{Stab}(C_{({\mathfrak{z}},{\mathfrak{z}}_{\bar{0}})})\rightarrow\text{Stab}(C_{(\langle x\rangle,\langle x\rangle_{\bar{0}})}) (6.2.2)

Let res:Stab(C(𝔤,𝔤0¯))Stab(C(𝔷,𝔷0¯))\text{res}:\text{Stab}(C_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})})\rightarrow\text{Stab}(C_{({\mathfrak{z}},{\mathfrak{z}}_{\bar{0}})}) be the natural functor obtained by restricting 𝔤{\mathfrak{g}}-modules to 𝔷{\mathfrak{z}}-modules. Then πx𝔤=πx𝔷res\pi_{x}^{\mathfrak{g}}=\pi_{x}^{\mathfrak{z}}\circ\text{res} for all x𝔷1¯x\in{\mathfrak{z}}_{\bar{1}}.

Now one can apply Theorem 5.2.3 (where F=πx𝔤F=\pi_{x}^{\mathfrak{g}} and πx𝔷\pi_{x}^{\mathfrak{z}}), to obtain x{\mathcal{B}}_{x} a homological prime (resp. x{\mathcal{B}}_{x}^{\prime}) associated to πx𝔤\pi_{x}^{\mathfrak{g}} (resp. πx𝔷\pi_{x}^{\mathfrak{z}}). Similarly, let h¯x\overline{h}_{{\mathcal{B}}_{x}} (resp. h¯x\overline{h}_{{\mathcal{B}}^{\prime}_{x}}) be the homological residue field corresponding to x{\mathcal{B}}_{x} (resp. x{\mathcal{B}}^{\prime}_{x}).

6.3.

Let 𝔤=𝔤0¯𝔤1¯\mathfrak{g}=\mathfrak{g}_{\bar{0}}\oplus\mathfrak{g}_{\bar{1}} be a Type I classical Lie superalgebra with detecting subalgebra 𝔷=𝔷0¯𝔷1¯𝔤.\mathfrak{z}=\mathfrak{z}_{\bar{0}}\oplus\mathfrak{z}_{\bar{1}}\subseteq\mathfrak{g}. Let x𝔷1¯,x\in\mathfrak{z}_{\bar{1}}, and let πx:x𝔤\pi_{x}:\langle x\rangle\longrightarrow\mathfrak{g} be the usual inclusion of the Lie subsuperalgebra generated by xx into 𝔤.\mathfrak{g}. This yields, for each x𝔷1¯,x\in\mathfrak{z}_{\bar{1}}, an induced functor πx:Stab(𝒞(𝔤,𝔤0¯))Stab(𝒞(x,x0¯)).\pi_{x}^{*}:\operatorname{Stab}(\mathcal{C}_{(\mathfrak{g},\mathfrak{g}_{\overline{0}})})\longrightarrow\operatorname{Stab}(\mathcal{C}_{(\langle x\rangle,{\langle x\rangle}_{\overline{0}})}). These are monoidal functors, each of which with right adjoint given by induction. Moreover, Stab(𝒞(x,x0¯))\operatorname{Stab}(\mathcal{C}_{(\langle x\rangle,{\langle x\rangle}_{\bar{0}})}) is a tensor-triangulated field, so Theorem 5.2.3 produces a homological residue field. Denote the corresponding homological prime of stab((𝔤,𝔤0¯))\operatorname{stab}(\mathcal{F}_{(\mathfrak{g},\mathfrak{g}_{\overline{0})}}) by 𝐁x.\mathbf{B}_{x}. The goal now is the show that {𝐁x}x𝔷1¯\{\mathbf{B}_{x}\}_{x\in\mathfrak{z}_{\bar{1}}} contains all of the homological primes. Recall the following result in [Bal20, Theorem 5.4].

Theorem 6.3.1.

Let 𝐓\mathbf{T} be a big tensor-triangulated category with 𝐊=𝐓c.\mathbf{K}=\mathbf{T}^{c}. Consider a family Spch(𝐊){\mathcal{E}}\subseteq\operatorname{Spc}^{h}(\mathbf{K}) of points in the homological spectrum. Suppose that the corresponding functors h¯𝐁:𝐓𝐀¯(𝐊;𝐁)\overline{h}_{\mathbf{B}}:\mathbf{T}\longrightarrow\overline{\mathbf{A}}(\mathbf{K};\mathbf{B}) collectively detect \otimes-nilpotence in the following sense: If f:xYf:x\longrightarrow Y in 𝐓\mathbf{T} is such that x𝐓cx\in\mathbf{T}^{c} and h¯𝐁(f)=0\overline{h}_{\mathbf{B}}(f)=0 for all 𝐁\mathbf{B}\in{\mathcal{E}} then fn=0f^{\otimes n}=0 for some n1.n\geq 1. Then we have =Spch(𝐊).{\mathcal{E}}=\operatorname{Spc}^{h}(\mathbf{K}).

6.4.

We are now ready to provide conditions on when one can identify a collection of homological primes that detect nilpotence on stab((𝔤,𝔤0¯))\text{stab}({\mathcal{F}}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})}).

Theorem 6.4.1.

Let 𝔤=𝔤0¯𝔤1¯{\mathfrak{g}}={\mathfrak{g}}_{\bar{0}}\oplus{\mathfrak{g}}_{\bar{1}} be a classical Lie superalgebra and 𝔷𝔤{\mathfrak{z}}\leq{\mathfrak{g}} be a sub Lie superalgebras. Denote by GG, G0¯G_{\bar{0}} and ZZ the associated supergroup (schemes) such that 𝔤=LieG{\mathfrak{g}}=\operatorname{Lie}G, 𝔤0¯=LieG0¯{\mathfrak{g}}_{\bar{0}}=\operatorname{Lie}G_{\bar{0}} and 𝔷=LieZ{\mathfrak{z}}=\operatorname{Lie}Z. Set N=NG0¯(𝔷1¯)N=N_{G_{\bar{0}}}({\mathfrak{z}}_{\bar{1}}). Assume that

  • (a)

    𝔷=𝔷0¯𝔷1¯{\mathfrak{z}}={\mathfrak{z}}_{\bar{0}}\oplus{\mathfrak{z}}_{\bar{1}} with [𝔷0¯,𝔷1¯]=0[{\mathfrak{z}}_{\bar{0}},{\mathfrak{z}}_{\bar{1}}]=0;

  • (b)

    ZZ is a splitting subgroup of GG.

Then /N={x:x𝔷1¯}/N{\mathcal{E}}/N=\{{\mathcal{B}}_{x}:x\in{\mathfrak{z}}_{\bar{1}}\}/N (i.e., a set of NN-orbit representatives) detects nilpotence in stab((𝔤,𝔤0¯))\operatorname{stab}({\mathcal{F}}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})}).

Proof.

The idea of the proof is to find a set of homological primes {\mathcal{E}} that detects nilpotence in Stab(𝒞(𝔤,𝔤0¯))\text{Stab}(\mathcal{C}_{(\mathfrak{g},\mathfrak{g}_{\overline{0}})}). Then one can apply Theorem 6.3.1 (e.g., [Bal20, Theorem 5.4].)

The first step is to compare homological residue fields for 𝔤{\mathfrak{g}} and 𝔷{\mathfrak{z}}. If f:MNf:M\rightarrow N is in Stab(C(𝔤,𝔤0¯))\text{Stab}(C_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})}) with MM compact then one can compare the diagrams for h¯x\overline{h}_{{\mathcal{B}}_{x}} and h¯x\overline{h}_{{\mathcal{B}}^{\prime}_{x}} to conclude the following. (1) If h¯x(f)=0\overline{h}_{{\mathcal{B}}_{x}}(f)=0 then h¯x(res(f))=0\overline{h}_{{\mathcal{B}}^{\prime}_{x}}(\operatorname{res}(f))=0 for x𝔷1¯x\in{\mathfrak{z}}_{\bar{1}}. Now one can apply the stratification result, Theorem 6.1.1, to conclude that {x:x𝔷1¯}\{{\mathcal{B}}^{\prime}_{x}:\ x\in{\mathfrak{z}}_{\bar{1}}\} are the homological primes for Stab((𝔷,𝔷0¯))\operatorname{Stab}(\mathcal{F}_{(\mathfrak{z},\mathfrak{z}_{\overline{0}})}). Therefore, by (1) and Theorem 5.4.2, one has (2) If h¯x(res(f))=0\overline{h}_{{\mathcal{B}}^{\prime}_{x}}(\operatorname{res}(f))=0 for all x𝔷1¯x\in{\mathfrak{z}}_{\bar{1}} then res(f):MN\operatorname{res}(f):M\rightarrow N is \otimes-nilpotent in Stab(𝒞(𝔷,𝔷0¯))\operatorname{Stab}(\mathcal{C}_{(\mathfrak{z},\mathfrak{z}_{\overline{0}})}). Applying Theorem 5.5.1 since 𝔷{\mathfrak{z}} is a splitting subalgebra of 𝔤{\mathfrak{g}}, one can conclude that f:MNf:M\rightarrow N is \otimes-nilpotent in Stab(𝒞(𝔤,𝔤0¯))\text{Stab}(\mathcal{C}_{(\mathfrak{g},\mathfrak{g}_{\overline{0}})}). Let ={x:x𝔷1¯}/N{\mathcal{E}}=\{{\mathcal{B}}_{x}:x\in{\mathfrak{z}}_{\bar{1}}\}/N. Since MM is a G0¯G_{\bar{0}}-module, it follows that the functors πx𝔤\pi^{\mathfrak{g}}_{x} (resp. πnx𝔤\pi^{\mathfrak{g}}_{nx}) will provide the same decomposition of MM in Stab(C(x,x0¯))\text{Stab}(C_{(\langle x\rangle,\langle x\rangle_{\bar{0}})}) (resp. Stab(C(nx,nx0¯))\text{Stab}(C_{(\langle nx\rangle,\langle nx\rangle_{\bar{0}})})). By considering Theorem 5.2.3, it follows that h¯x(f)=0\overline{h}_{{\mathcal{B}}_{x}}(f)=0 if and only if h¯nx(f)=0\overline{h}_{{\mathcal{B}}_{nx}}(f)=0. Therefore, /N{\mathcal{E}}/N detects nilpotence.

In the previous theorem, one can state that /N=Spch((𝔤,𝔤0¯)){\mathcal{E}}/N=\text{Spc}^{h}({\mathcal{F}}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})}). However, with the definition of /N{\mathcal{E}}/N there are certain homological primes that might be identified in the set. We will show in the following section that different NN-orbit representatives yield different elements in Spch(stab((𝔤,𝔤0¯)))\operatorname{Spc}^{h}(\text{stab}({\mathcal{F}}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})})).

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:

Spc(stab((𝔤,𝔤0¯)))Proj(H(𝔤,𝔤0¯,))\operatorname{Spc}(\text{stab}({\mathcal{F}}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})}))\rightarrow\operatorname{Proj}(\operatorname{H}^{\bullet}({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}},{\mathbb{C}})) (7.1.1)

is not always homeomorphism (e.g., when 𝔤=𝔤𝔩(m|n){\mathfrak{g}}=\mathfrak{gl}(m|n)). In general the cohomology ring H(𝔤,𝔤0¯,)\operatorname{H}^{\bullet}({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}},{\mathbb{C}}) does not stratify Stab(𝒞(𝔤,𝔤0¯))\text{Stab}({\mathcal{C}}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})}). 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 𝔤=𝔤𝔩(m|n){\mathfrak{g}}=\mathfrak{gl}(m|n), one has a homeomorphism:

Spc(stab((𝔤,𝔤0¯)))N-Proj(H(𝔣,𝔣0¯,))\operatorname{Spc}(\text{stab}({\mathcal{F}}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})}))\cong\text{$N$-}\operatorname{Proj}(\operatorname{H}^{\bullet}({\mathfrak{f}},{\mathfrak{f}}_{\bar{0}},{\mathbb{C}})) (7.1.2)

where 𝔣{\mathfrak{f}} is a detecting (splitting) subalgebra of 𝔤{\mathfrak{g}} and NN is the normalizer of 𝔣1¯{\mathfrak{f}}_{\bar{1}} in G0¯G_{\bar{0}}. 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 𝔷{\mathfrak{z}} of 𝔤{\mathfrak{g}}, one can compute the homological spectrum and show there is a surjection:

Spch(stab((𝔤,𝔤0¯)))N-Proj(H(𝔷,𝔷0¯,))\operatorname{Spc}^{h}(\text{stab}({\mathcal{F}}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})}))\rightarrow\text{$N$-}\operatorname{Proj}(\operatorname{H}^{\bullet}({\mathfrak{z}},{\mathfrak{z}}_{\bar{0}},{\mathbb{C}})) (7.1.3)

Since (𝔤,𝔤0¯){\mathcal{F}}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})} is rigid, the comparison map

ϕ:Spch(stab((𝔤,𝔤0¯)))Spc(stab((𝔤,𝔤0¯)))\phi:\operatorname{Spc}^{h}(\text{stab}({\mathcal{F}}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})}))\rightarrow\operatorname{Spc}(\text{stab}({\mathcal{F}}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})})) (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 ϕ\phi 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 𝔤{\mathfrak{g}} be a classical Lie superalgebra with a splitting subalgebra 𝔷𝔷0¯𝔷1¯{\mathfrak{z}}\cong{\mathfrak{z}}_{\bar{0}}\oplus{\mathfrak{z}}_{\bar{1}}. Assume that

  • (i)

    𝔷=𝔷0¯𝔷1¯{\mathfrak{z}}={\mathfrak{z}}_{\bar{0}}\oplus{\mathfrak{z}}_{\bar{1}} where 𝔷0¯{\mathfrak{z}}_{\bar{0}} is a torus and [𝔷0¯,𝔷1¯]=0[{\mathfrak{z}}_{\bar{0}},{\mathfrak{z}}_{\bar{1}}]=0.

  • (ii)

    Given WW an NN-invariant closed subvariety of Proj(S(𝔷1¯))\operatorname{Proj}(S^{\bullet}(\mathfrak{z}_{\bar{1}}^{*})), there exists Mstab((𝔤,𝔤0¯))M\in\operatorname{stab}(\mathcal{F}_{(\mathfrak{g},\mathfrak{g}_{\bar{0}})}) with V(𝔷,𝔷0¯)(M)=WV_{(\mathfrak{z},\mathfrak{z}_{\bar{0}})}(M)=W.

Then

  • (a)

    There exists a 1-1 correspondence

    {thick tensor ideals of stab((𝔤,𝔤0¯))}𝒳sp\{\text{thick tensor ideals of $\operatorname{stab}({\mathcal{F}}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})})$}\}\begin{array}[]{c}{}\atop{\longrightarrow}\\ {\longleftarrow}\atop{}\end{array}\mathcal{X}_{sp}

    where X=N-Proj(S(𝔷1¯))X=N\text{-}\operatorname{Proj}(S^{\bullet}({\mathfrak{z}}_{\bar{1}})) and 𝒳sp{\mathcal{X}}_{sp} are specialization closed sets of XX.

  • (b)

    There exists a homeomorphism η:N\eta:N-Proj(S(𝔷1¯))Spc(stab((𝔤,𝔤0¯)))\operatorname{Proj}(S^{\bullet}({\mathfrak{z}}_{\bar{1}}))\longrightarrow\operatorname{Spc}(\operatorname{stab}{(\mathcal{F}}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})})).

  • (c)

    The comparison map ϕ:Spch(stab((𝔤,𝔤0¯)))Spc(stab((𝔤,𝔤0¯)))\phi:\operatorname{Spc}^{h}(\operatorname{stab}({\mathcal{F}}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})}))\rightarrow\operatorname{Spc}(\operatorname{stab}({\mathcal{F}}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})})) is bijective.

Proof.

(a) and (b) follow by [BKN17, Theorems 3.4.1, 3.5.1]. For part (c), let ρ=η1\rho=\eta^{-1} which is given by a concrete description in [NVY, Corollary 6.2.4]. Consider the following diagram of topological spaces:

Spch(stab((𝔷,𝔷0¯))){\operatorname{Spc}^{h}(\operatorname{stab}({\mathcal{F}}_{({\mathfrak{z}},{\mathfrak{z}}_{\bar{0}})}))}Spc(stab((𝔷,𝔷0¯))){\operatorname{Spc}(\operatorname{stab}({\mathcal{F}}_{({\mathfrak{z}},{\mathfrak{z}}_{\bar{0}})}))}Proj(S(𝔷1¯)){\operatorname{Proj}(S^{\bullet}({\mathfrak{z}}_{\bar{1}}))}Spch(stab((𝔤,𝔤0¯))){\operatorname{Spc}^{h}(\operatorname{stab}({\mathcal{F}}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})}))}Spc(stab((𝔤,𝔤0¯))){\operatorname{Spc}(\operatorname{stab}({\mathcal{F}}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})}))}N-Proj(S(𝔷1¯)){N\text{-}\operatorname{Proj}(S^{\bullet}({\mathfrak{z}}_{\bar{1}}))}ϕ\scriptstyle{\phi^{\prime}}θ\scriptstyle{\theta}ρ\scriptstyle{\rho^{\prime}}π^\scriptstyle{\hat{\pi}}π\scriptstyle{\pi}ϕ\scriptstyle{\phi}ρ\scriptstyle{\rho} (7.2.1)

One has that ρ\rho^{\prime} is a homeomorphism and ϕ\phi^{\prime} is a bijection for 𝔷{\mathfrak{z}}. From part (b), the map ρ\rho is a homomorphism. The maps π\pi and π^\hat{\pi} are surjections. The map θ\theta sends x{\mathcal{B}}_{x} to x{\mathcal{B}}^{\prime}_{x} in /N{\mathcal{E}}/N. Suppose that ϕ(x1)=ϕ(x2)\phi({\mathcal{B}}_{x_{1}})=\phi({\mathcal{B}}_{x_{2}}). The using the commutativity, one has 𝒫x1=𝒫x2{\mathcal{P}}_{x_{1}}={\mathcal{P}}_{x_{2}} in NN-Proj(S(𝔷1¯))\operatorname{Proj}(S^{\bullet}({\mathfrak{z}}_{\bar{1}})) which means that x1x_{1} and x2x_{2} are NN-conjugate. This proves that x1=x2{\mathcal{B}}_{x_{1}}={\mathcal{B}}_{x_{2}} in /N{\mathcal{E}}/N.

We remark that the verification of the Nerves of Steel Conjecture in the previous theorem uses stratification results only for Stab(𝒞(𝔷,𝔷0¯))\text{Stab}({\mathcal{C}}_{({\mathfrak{z}},{\mathfrak{z}}_{\bar{0}})}), unlike the the case for finite group schemes where the stratification is needed for Stab(G)\text{Stab}(G) (see [Bal20, 5.13 Example]). A general stratification result for Stab(𝒞(𝔤,𝔤0¯))\text{Stab}({\mathcal{C}}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})}) 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 𝔤=𝔰𝔩(m|n){\mathfrak{g}}=\mathfrak{sl}(m|n). The main ideas have been established in the 𝔤𝔩(m|n)\mathfrak{gl}(m|n)-case (cf [BKN17, Section 7 ]). We will use the same notation to show how the arguments need to modified to handle the 𝔰𝔩(m|n)\mathfrak{sl}(m|n)-case.

Let 𝔤=𝔰𝔩(m|n)\mathfrak{g}=\mathfrak{sl}(m|n) and write 𝔣\mathfrak{f} for the detecting subalgebra of 𝔤\mathfrak{g}. Without loss of generality one can assume that mnm\leq n. The subalgebra 𝔣1¯\mathfrak{f}_{\bar{1}} has a basis given by the matrix units

e1,2m,e2,2m1,,em,m+1,em+1,m,,e2m,1.e_{1,2m},e_{2,2m-1},\dotsc,e_{m,m+1},e_{m+1,m},\dotsc,e_{2m,1}.

Let TT denote the torus of G0¯G_{\bar{0}} consisting of diagonal matrices. The torus TT acts on 𝔣{\mathfrak{f}} by the adjoint action and the aforementioned basis for 𝔣1¯\mathfrak{f}_{\bar{1}} consists of weight vectors. Let

R:=H(𝔣,𝔣0¯;)S(𝔣1¯)=[𝔣1¯][X1,X2,,Xm,Y1,Y2,,Ym],R:=\operatorname{H}^{\bullet}(\mathfrak{f},\mathfrak{f}_{\bar{0}};\mathbb{C})\cong S^{\bullet}(\mathfrak{f}_{\bar{1}}^{*})=\mathbb{C}[\mathfrak{f}_{\bar{1}}]\cong\mathbb{C}[X_{1},X_{2},\dots,X_{m},Y_{1},Y_{2},\dots,Y_{m}],

where XjX_{j} and YjY_{j} are defined by letting Xj:𝔣1¯X_{j}:\mathfrak{f}_{\bar{1}}\longrightarrow\mathbb{C} be the linear functional given on our basis of matrix units for 𝔣1¯\mathfrak{f}_{\bar{1}} by Xj(emj+1,m+j)=1X_{j}\left({e_{m-j+1,m+j}}\right)=1 and otherwise zero. The functionals Yj:𝔣1¯Y_{j}:\mathfrak{f}_{\bar{1}}\longrightarrow\mathbb{C} are defined similarly.

For our purposes it will suffice to consider the maximal ideal spectrum version of the support varieties. One has

V(𝔣,𝔣0¯)()maxV𝔣1¯r()=Proj(MaxSpec(R))=Proj(𝔣1¯).V_{({\mathfrak{f}},{\mathfrak{f}}_{\bar{0}})}(\mathbb{C})_{\max}\cong V^{r}_{\mathfrak{f}_{\bar{1}}}(\mathbb{C})=\operatorname{Proj}\left(\operatorname{MaxSpec}(R)\right)=\operatorname{Proj}(\mathfrak{f}_{\bar{1}}).

If N=NormG0¯(𝔣1¯)N=\operatorname{Norm}_{G_{\bar{0}}}(\mathfrak{f}_{\bar{1}}) then N=ΣmTLN=\Sigma_{m}TL where LL fixes RR and acts trivially on V(𝔣,𝔣0¯)()maxV_{({\mathfrak{f}},{\mathfrak{f}}_{\bar{0}})}(\mathbb{C})_{\max}. Here Σm\Sigma_{m} denote the symmetric group on mm letters embedded diagonally in G0¯G_{\bar{0}}. Let WW be such a variety. Since Σm\Sigma_{m} is a finite group we may write

W=ΣmVW=\Sigma_{m}V

for some TT-invariant closed subvariety VV of Proj(𝔣1¯)\operatorname{Proj}(\mathfrak{f}_{\bar{1}}).

Let Z(I)Z(I) be the closed subvariety of Proj(𝔣1¯)\operatorname{Proj}(\mathfrak{f}_{\bar{1}}) determined by a homogeneous ideal II of RR. It follows that VV must be of the form

V=Z(Xa1,,Xas,Yb1,,Ybt,g1,,gr)V=Z(X_{a_{1}},\dotsc,X_{a_{s}},Y_{b_{1}},\dotsc,Y_{b_{t}},g_{1},\dotsc,g_{r}) (8.1.1)

where g1,,grg_{1},\dotsc,g_{r} are homogeneous polynomials of weight zero for TT.

For the analysis in [BKN17], one employed the following conventions that will also be useful here. Let s,t,ps,t,p be non-negative integers with ms,tp0m\geq s,t\geq p\geq 0, and let

V(s,t,p)=Z(X1,,Xs,Ysp+1,,Ysp+t).V(s,t,p)=Z(X_{1},\dots,X_{s},Y_{s-p+1},\dots,Y_{s-p+t}). (8.1.2)

This is the conical variety given by the vanishing of ssXX coordinates”, ttYY coordinates”, with pp overlapping pairs.

8.2. Examples:

In this section, we will use examples to illustrate the concepts from the prior sections.

8.2.1. 𝔤=𝔰𝔩(1|1)\mathfrak{g}=\mathfrak{sl}(1|1)

In this case G0¯=TG_{\bar{0}}=T is one-dimensional and consists of 2×22\times 2 matrices which are non-zero scalar multiples of the identity matrix. The group TT acts trivally on R=S(𝔣1¯)R=S^{\bullet}({\mathfrak{f}}_{\bar{1}}^{*}). Therefore, the homogeneous polynomials in RR all have weight zero. This is in stark comparison to the 𝔤𝔩(1|1)\mathfrak{gl}(1|1)-case where the weight zero polynomials are the polynomials in the variable Z1=X1Y1Z_{1}=X_{1}Y_{1}.

8.2.2. 𝔤=𝔰𝔩(1|2)\mathfrak{g}=\mathfrak{sl}(1|2)

This case is more like the case for 𝔤𝔩(1|2)\mathfrak{gl}(1|2) The torus TT has dimension 2, and the vectors X1X_{1} and Y1Y_{1} are not fixed by TT. The the weight zero polynomial are polynomials in the variable Z1=X1Y1Z_{1}=X_{1}Y_{1}.

8.2.3. 𝔤=𝔰𝔩(2|2)\mathfrak{g}=\mathfrak{sl}(2|2)

The maximal torus TT does not act trivially on RR. A direct computation via the determinant condition shows that RTR^{T} is generated by X1Y1X_{1}Y_{1}, X1X2X_{1}X_{2}, X2Y2X_{2}Y_{2}, and Y1Y2Y_{1}Y_{2}. One still needs to consider the varieties V(s,t,p)V(s,t,p) is this case to describe the NN-invariant subvarieties of 𝔣1¯{\mathfrak{f}}_{\bar{1}}.

8.3.

The aim of this section is to outline how to realize every NN-invariant closed subvariety of V𝔣1¯r()Proj(𝔣1¯)V^{r}_{\mathfrak{f}_{\bar{1}}}(\mathbb{C})\cong\operatorname{Proj}(\mathfrak{f}_{\bar{1}}) as V𝔣1¯r(M)V^{r}_{\mathfrak{f}_{\bar{1}}}(M) for some M(𝔤,𝔤0¯)M\in\mathcal{F}_{(\mathfrak{g},\mathfrak{g}_{\bar{0}})} in the case of 𝔰𝔩(m|n)\mathfrak{sl}(m|n). The verification will use the results and the proofs in [BKN17] which accomplish this for 𝔤𝔩(m|n)\mathfrak{gl}(m|n).

The first proposition can be proved in the same way as in in [BKN17, Proposition 7.2.1] via the construction of LζL_{\zeta}-modules from elements ζ\zeta in the cohomology ring.

Proposition 8.3.1.

Let g1,g2,,grg_{1},g_{2},\dots,g_{r} be homogeneous polynomials in RR of weight zero with respect to TT. Then there exists a module MM in \mathcal{F} such that V𝔣1¯r(M)=ΣmZ(g1,g2,,gr)V^{r}_{\mathfrak{f}_{\bar{1}}}(M)=\Sigma_{m}Z(g_{1},g_{2},\dots,g_{r}).

The next series of steps involves realization involving certain subvarieties ΣrV(s,t,p)\Sigma_{r}V(s,t,p) using modules for 𝔤𝔩(m|n)\mathfrak{gl}(m|n). These varieties can be realized for modules in 𝔰𝔩(m|n)\mathfrak{sl}(m|n) by simply restricting the action of the modules realized for 𝔤𝔩(m|n)\mathfrak{gl}(m|n). Finally, one can use the proof in [BKN17, Theorem 7.12.1], to obtain the following result.

Theorem 8.3.2.

Let ms,tp0m\geq s,t\geq p\geq 0 and let g1,,grg_{1},\dotsc,g_{r} be homogeneous weight zero polynomials. Set V=V(s,t,p)Z(g1,,gr)V=V(s,t,p)\cap Z(g_{1},\dotsc,g_{r}). Then there exists a finite dimensional 𝔤\mathfrak{g}-module MM such that

V𝔣1¯r(M)=ΣmV.V^{r}_{\mathfrak{f}_{\bar{1}}}(M)=\Sigma_{m}V.

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 𝔤{\mathfrak{g}} be 𝔤𝔩(m|n)\mathfrak{gl}(m|n) or 𝔰𝔩(m|n)\mathfrak{sl}(m|n), and 𝔣𝔣0¯𝔣1¯{\mathfrak{f}}\cong{\mathfrak{f}}_{\bar{0}}\oplus{\mathfrak{f}}_{\bar{1}} be a detecting subalgebra of 𝔤{\mathfrak{g}}. Then

  • (a)

    There exists a 1-1 correspondence

    {thick tensor ideals of stab((𝔤,𝔤0¯))}𝒳sp\{\text{thick tensor ideals of $\operatorname{stab}({\mathcal{F}}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})})$}\}\begin{array}[]{c}{}\atop{\longrightarrow}\\ {\longleftarrow}\atop{}\end{array}\mathcal{X}_{sp}

    where X=N-Proj(S(𝔣1¯))X=N\text{-}\operatorname{Proj}(S^{\bullet}({\mathfrak{f}}_{\bar{1}})).

  • (b)

    There exists a homeomorphism η:N\eta:N-Proj(S(𝔣1¯))Spc(stab((𝔤,𝔤0¯)))\operatorname{Proj}(S^{\bullet}({\mathfrak{f}}_{\bar{1}}))\longrightarrow\operatorname{Spc}(\operatorname{stab}{(\mathcal{F}}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})})).

  • (c)

    The comparison map ϕ:Spch(stab((𝔤,𝔤0¯)))Spc(stab((𝔤,𝔤0¯)))\phi:\operatorname{Spc}^{h}(\operatorname{stab}({\mathcal{F}}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})}))\rightarrow\operatorname{Spc}(\operatorname{stab}({\mathcal{F}}_{({\mathfrak{g}},{\mathfrak{g}}_{\bar{0}})})) 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 𝔤𝔤𝔩(m|n){\mathfrak{g}}\cong\mathfrak{gl}(m|n) or 𝔰𝔩(m|n)\mathfrak{sl}(m|n), the detecting subalgebras 𝔣{\mathfrak{f}} are splitting subalgebras.

For (ii), the realization property can be reduced to the case of the maximal ideal spectrum, i.e., one can realize every NN-invariant closed subvariety of V𝔣1¯r()Proj(𝔣1¯)V^{r}_{\mathfrak{f}_{\bar{1}}}(\mathbb{C})\cong\operatorname{Proj}(\mathfrak{f}_{\bar{1}}) as V𝔣1¯r(M)V^{r}_{\mathfrak{f}_{\bar{1}}}(M) for some M(𝔤,𝔤0¯)M\in\mathcal{F}_{(\mathfrak{g},\mathfrak{g}_{\bar{0}})} (cf. [BKN17, Section 2.4]). For 𝔤=𝔤𝔩(m|n){\mathfrak{g}}=\mathfrak{gl}(m|n), this holds by [BKN17, Theorem 7.21.1], and for 𝔤=𝔰𝔩(m|n){\mathfrak{g}}=\mathfrak{sl}(m|n) by Theorem‘8.3.2. ∎

We remark that the Lie superalgebra 𝔤=𝔭𝔰𝔩(n|n)\mathfrak{g}=\mathfrak{psl}(n|n) is a quotient of 𝔰𝔩(n|n)\mathfrak{sl}(n|n). One cannot simply restrict modules like in the case for 𝔤𝔩(m|n)\mathfrak{gl}(m|n) to 𝔰𝔩(m|n)\mathfrak{sl}(m|n). This means that new techniques need to be developed to verify Condition (ii) for 𝔭𝔰𝔩(n|n)\mathfrak{psl}(n|n).

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 W(n)W(n), 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, Π\Pi-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 D(R)D(R). 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.