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

Stable representation theory:
beyond the classical groups

Andrew Snowden Department of Mathematics, University of Michigan, Ann Arbor, MI [email protected] http://www-personal.umich.edu/~asnowden/
(Date: September 23, 2021)
Abstract.

The orthogonal groups are a series of simple Lie groups associated to symmetric bilinear forms. There is no analogous series associated to symmetric trilinear forms. We introduce an infinite dimensional group-like object that can be viewed as the limit of this non-existent series, were it to exist. We show that the representation theory of this object is well-behaved, and similar to the stable representation theory of orthogonal groups. Our theory is not specific to symmetric trilinear forms, and applies to any kind of tensorial forms. Our results can be also be viewed from the perspective of semi-linear representations of the infinite general linear group, and are closely related to twisted commutative algebras.

AS was supported by NSF DMS-1453893.

1. Introduction

Bilinear forms are remarkable objects: they have just the right amount of complexity to be tractable and yet still interesting. Their symmetry groups, the orthogonal and symplectic groups, are among the most important objects in mathematics. Trilinear forms, on the other hand, are too complicated. Their symmetry groups are diverse, but generically finite, and do not give rise to new families of simple Lie groups.

It has recently been discovered [BDE, BDES, DES] that, somewhat surprisingly, trilinear forms (and higher degree tensorial forms) in infinite dimensions are less complicated than their finite dimensional counterparts, and more like bilinear forms. In particular, up to a certain notion of equivalence, there is a unique non-degenerate form of each type (e.g., symmetric trilinear). The purpose of this paper is to introduce group-like objects (called germinal subgroups) that capture the symmetry of these forms, and to study their representation theory. We find that this representation theory is very well-behaved, and closely parallels the stable representation theory of the classical groups. Thus, while there is not a family of simple Lie groups attached to, say, symmetric trilinear forms, there is nonetheless a reasonable limiting object.

1.1. Generalized orbits and stabilizers

We explain our main ideas and results in the setting of symmetric trilinear forms over the complex numbers to keep the exposition simple. We work more generally in the body of the paper.

Let Xn=Sym3(𝐂n)X_{n}=\operatorname{Sym}^{3}(\mathbf{C}^{n})^{*} be the space of symmetric trilinear forms (i.e., cubic polynomials) in nn variables. Also, let 𝐂=n1𝐂n\mathbf{C}^{\infty}=\bigcup_{n\geq 1}\mathbf{C}^{n} and X=Sym3(𝐂)X_{\infty}=\operatorname{Sym}^{3}(\mathbf{C}^{\infty})^{*}. The set XX_{\infty} is the inverse limit of the sets XnX_{n}, and as such carries the inverse limit topology. (Each XnX_{n} is endowed with the discrete topology). Precisely, a sequence {ωi}i1\{\omega_{i}\}_{i\geq 1} in XX_{\infty} converges to ω\omega if for each nn we have ωi|𝐂n=ω|𝐂n\omega_{i}|_{\mathbf{C}^{n}}=\omega|_{\mathbf{C}^{n}} for all sufficiently large ii.

The group 𝐆𝐋n(𝐂)\mathbf{GL}_{n}(\mathbf{C}) acts on XnX_{n}, and the group 𝐆𝐋=n1𝐆𝐋n(𝐂)\mathbf{GL}=\bigcup_{n\geq 1}\mathbf{GL}_{n}(\mathbf{C}) acts on XX_{\infty}. The group 𝐆𝐋\mathbf{GL} is, in a sense, too small111The group Aut(𝐂)\operatorname{Aut}(\mathbf{C}^{\infty}) is much larger than 𝐆𝐋\mathbf{GL}, but it is also too small.. To remedy this, we employ a modification of the concept of orbit: we say that two elements of XX_{\infty} belong to the same generalized orbit if each belongs to the closure of the orbit of the other. This idea was introduced in a slightly different way in [BDES]; see §2.5 and Remark 2.3 for details.

We say that an element of XX_{\infty} is degenerate if it has the form i=1nqii\sum_{i=1}^{n}q_{i}\ell_{i} where qiSym2(𝐂)q_{i}\in\operatorname{Sym}^{2}(\mathbf{C}^{\infty})^{*} and i(𝐂)\ell_{i}\in(\mathbf{C}^{\infty})^{*}, and non-degenerate otherwise. The main theorem of [DES] asserts that the non-degnerate forms constitute a single generalized orbit. (The paper [DES] concerns only symmetric trilinear forms, but this statement was extended to other types of tensorial forms in [BDE, BDES].)

Just as the usual orbits of 𝐆𝐋\mathbf{GL} are too small, so too are the usual stabilizers. One question we sought to answer in this paper is: what is the right notion of “generalized stabilizer”? We have come to the following idea. Let ωX\omega\in X_{\infty} be given. For n1n\geq 1, define Γω(n)\Gamma_{\omega}(n) be the set of elements g𝐆𝐋g\in\mathbf{GL} such that g1ω|𝐂n=ω|𝐂ng^{-1}\omega|_{\mathbf{C}^{n}}=\omega|_{\mathbf{C}^{n}}. (The inverse here is simply to make some other definitions cleaner.) Note that Γω(n)\Gamma_{\omega}(n) is typically not a subgroup. We define the generalized stabilizer of ω\omega to be the system Γω={Γω(n)}n1\Gamma_{\omega}=\{\Gamma_{\omega}(n)\}_{n\geq 1}. One should think of Γω\Gamma_{\omega} as a kind of germ of a neighborhood of the stabilizer of ω\omega. For this reason, we refer to Γω\Gamma_{\omega} as a germinal subgroup; see Definition  7.1 for details. We view Γω\Gamma_{\omega} as an analog of the infinite orthogonal group associated to symmetric trilinear forms.

1.2. Representations of generalized stabilizers

Let ωX\omega\in X_{\infty} and Γω\Gamma_{\omega} be as above. We define a representation of Γω\Gamma_{\omega} to be a complex vector space VV such that each finite dimensional subspace WVW\subset V is endowed with an action map Γω(n)×WV\Gamma_{\omega}(n)\times W\to V, for some nn depending on WW, satisfying certain conditions. A little more precisely, the data defining a representation can be encoded as a linear map

VlimnFun(Γω(n),V).V\to\varinjlim_{n\to\infty}\operatorname{Fun}(\Gamma_{\omega}(n),V).

Thus for vVv\in V, one can regard ggvg\mapsto gv as the germ of a function on 𝐆𝐋\mathbf{GL}, with respect to the system of neighborhoods Γω\Gamma_{\omega}.

Every representation of 𝐆𝐋\mathbf{GL} restricts to a representation of Γω\Gamma_{\omega}. We say that a representation of Γω\Gamma_{\omega} is algebraic if it occurs as a subquotient of the restriction of a polynomial representation of 𝐆𝐋\mathbf{GL}. In particular, the standard representaiton 𝐂\mathbf{C}^{\infty} of 𝐆𝐋\mathbf{GL} restricts to an algebraic representation of Γω\Gamma_{\omega}, which we call the standard representation of Γω\Gamma_{\omega}. We let Repalg(Γω)\operatorname{Rep}^{\mathrm{alg}}(\Gamma_{\omega}) denote the category of algebraic representations. This is a Grothendieck abelian category equipped with a tensor product.

The primary purpose of this paper is to understand the algebraic representation theory of Γω\Gamma_{\omega} when ω\omega is non-degenerate. The following is a summary of our findings.

  • Algebraic representations enjoy several finiteness properties:

    • Every algebraic representation is the union of its finite length subrepresentations.

    • The tensor product of two finite length algebraic representations is again finite length.

    • If VV and WW are finite length algebraic representations then HomΓω(V,W)\operatorname{Hom}_{\Gamma_{\omega}}(V,W) is a finite dimensional complex vector space.

  • The simple algebraic representations are well-understood:

    • For each partition λ\lambda, there is a simple LλL_{\lambda}, and these exhaust the simples.

    • One can construct LλL_{\lambda} using a variant of Weyl’s traceless tensor construction. Let T[n]T^{[n]} be the intersection of the kernels of the maps (𝐂)n(𝐂)(n3)(\mathbf{C}^{\infty})^{\otimes n}\to(\mathbf{C}^{\infty})^{\otimes(n-3)} obtained by applying ω\omega to three tensor factors. This space carries an action of 𝔖n×Γω\mathfrak{S}_{n}\times\Gamma_{\omega}, where 𝔖n\mathfrak{S}_{n} denotes the symmetric group. The isotypic piece of T[n]T^{[n]} corresponding the Specht module SλS^{\lambda} is exactly LλL_{\lambda}.

  • Algebraic representations are well-behaved homologically:

    • The representations 𝐒λ(𝐂)\mathbf{S}_{\lambda}(\mathbf{C}^{\infty}) are exactly the indecomposable injective algebraic representations; in fact, 𝐒λ(𝐂)\mathbf{S}_{\lambda}(\mathbf{C}^{\infty}) is the injective envelope of LλL_{\lambda}.

    • Every finite length algebraic representation has finite injective dimension.

  • There is a combinatorial description of the entire category Repalg(Γω)\operatorname{Rep}^{\mathrm{alg}}(\Gamma_{\omega}): it is equivariant to a category of representations of a certain variant of the upwards Brauer category.

  • The category Repalg(Γω)\operatorname{Rep}^{\mathrm{alg}}(\Gamma_{\omega}) satisfies a universal property. Let 𝒞\mathcal{C} be a 𝐂\mathbf{C}-linear abelian category equipped with a tensor product. Then giving a left-exact symmetric monoidal 𝐂\mathbf{C}-linear functor Repalg(Γω)f𝒞\operatorname{Rep}^{\mathrm{alg}}(\Gamma_{\omega})^{\mathrm{f}}\to\mathcal{C} is equivalent to giving an object of 𝒞\mathcal{C} equipped with a symmetric trilinear form. The notation ()f(-)^{\mathrm{f}} here denotes the subcategory of finite length objects.

  • The symmetric monoidal category Repalg(Γω)\operatorname{Rep}^{\mathrm{alg}}(\Gamma_{\omega}) is independent of ω\omega, up to equivalence.

1.3. Semi-linear representations: motivation

Recall that if a group GG acts on a field KK then a semi-linear representation of GG over KK is a KK-vector space VV equipped with an additive action of GG such that the equation g(av)=(ga)(gv)g(av)=(ga)(gv) holds, for gGg\in G, aKa\in K, and vVv\in V. Semi-linear representations will be a central topic in this paper. To motivate their appearance, we first examine a familiar case.

Let Yn=Sym2(𝐂n)Y_{n}=\operatorname{Sym}^{2}(\mathbf{C}^{n})^{*} be the space of symmetric bilinear forms on 𝐂n\mathbf{C}^{n}, regarded as an algebraic variety; explicitly, Yn=Spec(Sn)Y_{n}=\operatorname{Spec}(S_{n}) where SnS_{n} is the polynomial ring Sym(Sym2(𝐂n))\operatorname{Sym}(\operatorname{Sym}^{2}(\mathbf{C}^{n})). Let YnY_{n}^{\circ} be the open subvariety of YnY_{n} consisting of non-degenerate forms. The algebraic group 𝐆𝐋n\mathbf{GL}_{n} acts transitively on YnY_{n}^{\circ}. Let yy be a closed point of YnY_{n}^{\circ}, and let 𝐎n\mathbf{O}_{n} be its stabilizer. If \mathcal{F} is a 𝐆𝐋n\mathbf{GL}_{n}-equivariant quasi-coherent sheaf on YnY_{n}^{\circ} then its fiber (y)\mathcal{F}(y) at yy is an algebraic representation of 𝐎n\mathbf{O}_{n}, and this construction gives an equivalence of categories

QCoh(Yn)𝐆𝐋nRep(𝐎n)\operatorname{QCoh}(Y_{n}^{\circ})^{\mathbf{GL}_{n}}\to\operatorname{Rep}(\mathbf{O}_{n})

In fact, we can get a similar equivalence using the generic point of YnY_{n}^{\circ}. If \mathcal{F} is a 𝐆𝐋n\mathbf{GL}_{n}-equivariant quasi-coherent sheaf on YnY_{n}^{\circ} then its generic fiber is a semi-linear representation of 𝐆𝐋n\mathbf{GL}_{n} over Frac(Sn)\operatorname{Frac}(S_{n}) that is algebraic (in the sense that it is spanned by an algebraic subrepresentation). Moreover, letting 𝒞n\mathcal{C}_{n} be the category of such semi-linear representations, this construction defines an equivalence

QCoh(Yn)𝐆𝐋n𝒞n\operatorname{QCoh}(Y_{n}^{\circ})^{\mathbf{GL}_{n}}\to\mathcal{C}_{n}

Thus, combined with the previous equivalence, we obtain an equivalence

Rep(𝐎n)=𝒞n.\operatorname{Rep}(\mathbf{O}_{n})=\mathcal{C}_{n}.

This gives us a way of studying representations of 𝐎n\mathbf{O}_{n} (or, at least, the representation category) even if we do not understand the group 𝐎n\mathbf{O}_{n} very well.

We adopt this approach in this paper to replace representations of Γω\Gamma_{\omega} with more familiar objects. Let RR be the infinite variable polynomial ring Sym(Sym3(𝐂))\operatorname{Sym}(\operatorname{Sym}^{3}(\mathbf{C}^{\infty})) and let K=Frac(R)K=\operatorname{Frac}(R). We show (Theorem 8.11) that Repalg(Γω)\operatorname{Rep}^{\mathrm{alg}}(\Gamma_{\omega}) is equivalent to a certain category of semi-linear representations of 𝐆𝐋\mathbf{GL} over KK (precisely, the category of “KK-modules” introduced below). The proof is similar to the one outlined above, but technically more involved, and relies on some non-trivial results from [BDDE] and [BDES]. We find the semi-linear perspective to be technically much easier to work with, so most of the paper is carried out in this setting.

1.4. Semi-linear representations: results

We now explain some of our results on semi-linear representations in more detail. We first introduce some fundamental definitions. A 𝐆𝐋\mathbf{GL}-algebra222In characteristic 0, 𝐆𝐋\mathbf{GL}-algebras are equivalent, under Schur–Weyl duality, to twisted commutative algebras; see [SS2, §8.1]. is an algebra object in the category of polynomial representations of 𝐆𝐋\mathbf{GL}; in other words, it is a commutative ring equipped with an action of 𝐆𝐋\mathbf{GL} under which it forms a polynomial representation. For example, the ring Sym(Sym3(𝐂))\operatorname{Sym}(\operatorname{Sym}^{3}(\mathbf{C}^{\infty})) appearing above is a 𝐆𝐋\mathbf{GL}-algebra. If RR is a 𝐆𝐋\mathbf{GL}-algebra then an RR-module is a module object; in other words, it is a 𝐆𝐋\mathbf{GL}-equivariant RR-module that forms a polynomial representation.

A 𝐆𝐋\mathbf{GL}-field is a field equipped with an action of 𝐆𝐋\mathbf{GL} that can be obtained as the fraction field of an integral 𝐆𝐋\mathbf{GL}-algebra. If KK is a 𝐆𝐋\mathbf{GL}-field then a KK-module is a semi-linear representation of 𝐆𝐋\mathbf{GL} over KK that is generated by a polynomial subrepresentation. The basic example of a KK-module is Kλ=K𝐂𝐒λ(𝐂)K^{\oplus\lambda}=K\otimes_{\mathbf{C}}\mathbf{S}_{\lambda}(\mathbf{C}^{\infty}). While KλK^{\oplus\lambda} is typically not projective, every KK-module is a quotient of a sum of ones of this form. We let ModK\operatorname{Mod}_{K} denote the category of KK-modules. This is the fundamental object of study in this paper.

We prove two main technical results about KK-modules. To state the first one, we must introduce the shift operation. Let G(n)G(n) be the subgroup of 𝐆𝐋\mathbf{GL} consisting of block matrices of the form

(100)\begin{pmatrix}1&0\\ 0&\ast\end{pmatrix}

where the top left block is n×nn\times n. This group is isomorphic to 𝐆𝐋\mathbf{GL}. If XX is a set equipped with an action of 𝐆𝐋\mathbf{GL}, we define its nnth shift, denoted Shn(X)\operatorname{Sh}_{n}(X), to be the set XX equipped with the action of 𝐆𝐋\mathbf{GL} coming from restricting the given action to G(n)𝐆𝐋G(n)\cong\mathbf{GL}. The shift operation preserves all structure introduced so far (polynomial representations, 𝐆𝐋\mathbf{GL}-fields, etc.). Our first theorem is:

Theorem 1.1 (Shift theorem).

Let KK be a 𝐆𝐋\mathbf{GL}-field and let MM be a finitely generated KK-module. Then there exists n0n\geq 0 and partitions λ1,,λr\lambda_{1},\ldots,\lambda_{r} such that Shn(M)\operatorname{Sh}_{n}(M) is isomorphic to i=1rShn(K)λi\bigoplus_{i=1}^{r}\operatorname{Sh}_{n}(K)^{\oplus\lambda_{i}} as a Shn(K)\operatorname{Sh}_{n}(K)-module.

This theorem is an instance of the general principle in representation stability that objects can be made “nice” after shifting. The first theorem of this sort was Nagpal’s shift theorem for 𝐅𝐈\mathbf{FI}-modules [Na]. The above shift theorem is closely related to the shift theorem for 𝐆𝐋\mathbf{GL}-varieties [BDES, Theorem 5.1], and follows a similar proof.

Our second main result about KK-modules is the following:

Theorem 1.2 (Embedding theorem).

Let KK be a rational 𝐆𝐋\mathbf{GL}-field, i.e., one of the form Frac(Sym(E))\operatorname{Frac}(\operatorname{Sym}(E)) where EE is a finite length polynomial representation of 𝐆𝐋\mathbf{GL}, and let MM be a finitely generated KK-module. Then there exist partitions λ1,,λr\lambda_{1},\ldots,\lambda_{r} and an injection of KK-modules Mi=1rKλiM\to\bigoplus_{i=1}^{r}K^{\oplus\lambda_{i}}.

This theorem follows rather easily from the shift theorem. It is a very important theorem for us: indeed, all the statements in §1.2 have analogs for ModK\operatorname{Mod}_{K}, and can be deduced from the embedding thoerem by comparitively easy arguments. The corresponding results for Repalg(Γω)\operatorname{Rep}^{\mathrm{alg}}(\Gamma_{\omega}) are deduced from those for ModK\operatorname{Mod}_{K}.

1.5. Summary of categories

Let 𝔘\mathfrak{U} be the analog of the upwards Brauer category for symmetric trilinear forms (see §5.2), let R=Sym(Sym3(𝐂))R=\operatorname{Sym}(\operatorname{Sym}^{3}(\mathbf{C}^{\infty})), let K=Frac(R)K=\operatorname{Frac}(R), and let ωSym3(𝐂)\omega\in\operatorname{Sym}^{3}(\mathbf{C}^{\infty})^{*} be non-degenerate. We show that the following categories are equivalent:

  1. (a)

    The category Mod𝔘lf\operatorname{Mod}_{\mathfrak{U}}^{\mathrm{lf}} of 𝔘\mathfrak{U}-modules that are locally of finite length.

  2. (b)

    The category ModRlf\operatorname{Mod}_{R}^{\mathrm{lf}} of RR-modules that are locally of finite length.

  3. (c)

    The generic category ModRgen\operatorname{Mod}_{R}^{\mathrm{gen}}, i.e., the Serre quotient of ModR\operatorname{Mod}_{R} by the subcategory of torsion modules.

  4. (d)

    The category ModK\operatorname{Mod}_{K}.

  5. (e)

    The category Repalg(Γω)\operatorname{Rep}^{\mathrm{alg}}(\Gamma_{\omega}).

The equivalence between (a) and (b) is straightforward, as is the equivalence between (c) and (d). The equivalence of (b) and (c) is much more difficult, and relies upon the embedding theorem. The equivalence between (d) and (e) is also difficult, and relies on non-trivial results from [BDDE] and [BDES].

The equivalence between (b) and (c) above has a long history: see Remark 5.9.

1.6. Fiber functors

The categories (a)–(e) above are 𝐂\mathbf{C}-linear tensor categories. However, only in (e) are the objects 𝐂\mathbf{C}-vector spaces (with extra structure), with the tensor product being the usual one on the underlying vector space. One can therefore view the equivalence ModKRepalg(Γω)\operatorname{Mod}_{K}\cong\operatorname{Rep}^{\mathrm{alg}}(\Gamma_{\omega}) as a fiber functor on ModK\operatorname{Mod}_{K}. We thus get one such fiber functor for each choice of ω\omega. We show (§6) that all fibers functors are obtained in essentially this manner.

1.7. Relation to previous work

This paper is closely related to four threads of recent work:

  • The papers [BDE, BDDE, BDES, Dr, DES] develop aspects of infinite dimensional 𝐆𝐋\mathbf{GL}-equivariant algebraic geometry. These theories are based on 𝐆𝐋\mathbf{GL}-algebras, which is the main connection to this paper. A few key arguments in this paper are in fact modeled on those from [BDES]. The work of Kazhdan–Ziegler [KaZ1, KaZ2, KaZ3, KaZ4] is closely related.

  • The papers [NSS, NSS2, NSS3, SS1, SS5, SS6] study the module theory of a handful of specific 𝐆𝐋\mathbf{GL}-algebras (and similar objects). The results of this paper generalize many of the results from those papers.

  • The papers [DPS, GS, PSe, PSt, Se, SS3] study the stable representation theory of classical (super)groups. The results summarized in §1.2 are all analogs of results from these papers (especially [SS3]).

  • The papers [NS, Ro, Ro2, Ro3] study the semi-linear representation theory of the infinite symmetric group, which is thematically similar to much of the work in this paper.

1.8. Further work

In this paper, we give a fairly complete description of ModK\operatorname{Mod}_{K} when KK is a rational 𝐆𝐋\mathbf{GL}-field. While we do prove some results for more general 𝐆𝐋\mathbf{GL}-fields (see Theorem 4.12), there is still much left to be done in this direction. We hope to treat this in a future paper.

In the study of modules over 𝐆𝐋\mathbf{GL}-algebras, it is also important to understand the generic categories ModRgen\operatorname{Mod}_{R}^{\mathrm{gen}} when RR is a “𝐆𝐋\mathbf{GL}-domain” (this means 𝔞𝔟=0\mathfrak{a}\mathfrak{b}=0 implies 𝔞=0\mathfrak{a}=0 or 𝔟=0\mathfrak{b}=0 when 𝔞\mathfrak{a} and 𝔟\mathfrak{b} are 𝐆𝐋\mathbf{GL}-ideals, which is a weaker condition than being a domain). In [Sn], we gave a useful way of understanding the 𝐆𝐋\mathbf{GL}-domain condition in terms of super mathematics, and we believe this should allow us to say something about these generic categories. We hope to return to this topic too.

1.9. Open questions

We list a few questions or problems raised by this work:

  1. (a)

    How much of standard Lie theory can be carried over to the generalized stabilizers Γω\Gamma_{\omega}? Is there a Dynkin diagram, Cartan matrix, Weyl group, etc.?

  2. (b)

    Is there a Tannakian perspective that allows one to recover the generalized stabilizer Γω\Gamma_{\omega} from the fiber functor Φω:ModKVeck\Phi_{\omega}\colon\operatorname{Mod}_{K}\to\mathrm{Vec}_{k}?

  3. (c)

    Prove Theorem 8.11 for general KK.

  4. (d)

    What are the derived specializations of simple objects of ModK\operatorname{Mod}_{K}? (See Remark 5.13.)

  5. (e)

    We introduce the concept of “germinal subgroup” to define generalized stabilizers. While our definitions work for the purposes of this paper, we are not sure if they are optimal. For instance, our conditions do not say anything about inverses. It would be good to have more clarity on this point.

  6. (f)

    In this paper, we consider generalized stabilizers for actions of 𝐆𝐋\mathbf{GL} on infinite dimensional varieties. Are there other situations where generalized stabilizers are interesting? For example, one could consider generalized stabilizers arising from actions of the infinite symmetry group on infinite dimensional varieties.

1.10. Outline

In §2 we provide background about 𝐆𝐋\mathbf{GL}-algebras and related concepts. In §3 we prove our two main technical theorems on KK-modules, the shift and embedding theorems. We apply these results in §4 to deduce our main structural results on semi-linear representations. These results are in turn used in §5 to obtain the connection to an analog of the Brauer category, which yields an analog of Weyl’s construction and a universal property for ModK\operatorname{Mod}_{K}. In §6, we classify the fiber functors of ModK\operatorname{Mod}_{K}. In §7 we introduce germinal subgroups and generalized stabilizers in the abstract. Finally, in §8, we apply these concepts to 𝐆𝐋\mathbf{GL}-varieties.

Acknowledgments

We thank Arthur Bik, Jan Draisma, Rob Eggermont, Nate Harman, Steven Sam, and David Treumann for helpful conversations. In particular, Proposition 2.2 came from an e-mail exchange with Bik, Draisma, and Eggermont, and the material in §4.1 came from unpublished notes with Sam.

2. 𝐆𝐋\mathbf{GL}-equivariant algebra and geometry

In this section, we review background material on polynomial representations, 𝐆𝐋\mathbf{GL}-algebras, 𝐆𝐋\mathbf{GL}-varieties, and related concepts. Additional details on these topics can be found in [SS2] and [BDES].

2.1. Polynomial representations

Fix, for the entirety of the paper, a field kk of characteristic 0. Put 𝐆𝐋=n1𝐆𝐋n(k)\mathbf{GL}=\bigcup_{n\geq 1}\mathbf{GL}_{n}(k), regarded as a discrete group. We let 𝐕=n1kn\mathbf{V}=\bigcup_{n\geq 1}k^{n} be the standard representation of 𝐆𝐋\mathbf{GL}. We say that a representation of 𝐆𝐋\mathbf{GL} on a kk-vector space is polynomial if it appears as a subquotient of a (possibly infinite) direct sum of tensor powers of 𝐕\mathbf{V}. We let Reppol(𝐆𝐋)\operatorname{Rep}^{\mathrm{pol}}(\mathbf{GL}) denote the category of polynomial representations. It is a semi-simple Grothendieck abelian category that is closed under tensor product.

For a partition λ\lambda, we let 𝐒λ\mathbf{S}_{\lambda} denote the corresponding Schur functor. The simple polynomial representations are exactly those of the form 𝐒λ(𝐕)\mathbf{S}_{\lambda}(\mathbf{V}). Thus every polynomial representation decomposes as a (perhaps infinite) direct sum of 𝐒λ(𝐕)\mathbf{S}_{\lambda}(\mathbf{V})’s.

Every polynomial representation of 𝐆𝐋\mathbf{GL} carries a natural grading, with 𝐒λ(𝐕)\mathbf{S}_{\lambda}(\mathbf{V}) concentrated in degree |λ||\lambda|, the size of the partition λ\lambda. This grading is compatible with tensor products: 𝐒λ(𝐕)𝐒μ(𝐕)\mathbf{S}_{\lambda}(\mathbf{V})\otimes\mathbf{S}_{\mu}(\mathbf{V}) is concentrated in degree |λ|+|μ||\lambda|+|\mu|. The degree 0 piece of a polynomial representation VV is exactly the invariant subspace V𝐆𝐋V^{\mathbf{GL}}.

We now introduce some non-standard notation that will be convenient for working with these objects. We write kλk^{\oplus\lambda} in place of 𝐒λ(𝐕)\mathbf{S}_{\lambda}(\mathbf{V}). More generally, for a kk-vector space VV we put Vλ=VkkλV^{\oplus\lambda}=V\otimes_{k}k^{\oplus\lambda}; note that if RR is a kk-algebra then RλR^{\oplus\lambda} is naturally a free RR-module. A tuple of partitions (often simply called a tuple) is a tuple λ¯=[λ1,,λr]\underline{\smash{\lambda}}=[\lambda_{1},\ldots,\lambda_{r}], where each λi\lambda_{i} is a partition. We put kλ¯=i=1rkλik^{\oplus\underline{\smash{\lambda}}}=\bigoplus_{i=1}^{r}k^{\oplus\lambda_{i}}, and define Vλ¯V^{\oplus\underline{\smash{\lambda}}} similarly. We say that λ¯\underline{\smash{\lambda}} is pure if it does not contain the empty partition. (This terminology comes from [BDES].)

The category of polynomial representations is equivalent to the category of polynomial functors, with the representation 𝐒λ(𝐕)\mathbf{S}_{\lambda}(\mathbf{V}) corresponding to the functor 𝐒λ\mathbf{S}_{\lambda}. Given a polynomial representation VV and a vector space UU, we let V{U}V\{U\} be the result of regarding VV as a polynomial functor and evaluating on UU. In the important special case where U=knU=k^{n}, we can identify V{U}V\{U\} with the invariant space VG(n)V^{G(n)}, where G(n)G(n) is defined in §2.3. For example, if V=kλV=k^{\oplus\lambda} then V{kn}=𝐒λ(kn)V\{k^{n}\}=\mathbf{S}_{\lambda}(k^{n}).

2.2. The maximal polynomial subrepresentation

Suppose that VV is an arbitrary kk-linear representation of 𝐆𝐋\mathbf{GL}. We say that an element xVx\in V is polynomial if the subrepresentation it generates is a polynomial representation. We let VpolV^{\mathrm{pol}} be the set of all polynomial elements in VV. It can be characterized as the maximal polynomial subrepresentation of VV. Moreover, if REP(𝐆𝐋)\operatorname{REP}(\mathbf{GL}) denotes the category of all kk-linear representations of 𝐆𝐋\mathbf{GL} then VVpolV\mapsto V^{\mathrm{pol}} is the right adjoint of the inclusion functor Reppol(𝐆𝐋)REP(𝐆𝐋)\operatorname{Rep}^{\mathrm{pol}}(\mathbf{GL})\to\operatorname{REP}(\mathbf{GL}). As such, ()pol(-)^{\mathrm{pol}} is left-exact and continuous; it is not exact.

2.3. The shift operation

Recall that G(n)G(n) is the subgroup of 𝐆𝐋\mathbf{GL} consisting of block matrices of the form

(100),\begin{pmatrix}1&0\\ 0&\ast\end{pmatrix},

where the top left block has size n×nn\times n. We have a group isomorphism

𝐆𝐋G(n),A(100A).\mathbf{GL}\to G(n),\qquad A\mapsto\begin{pmatrix}1&0\\ 0&A\end{pmatrix}.

Given some kind of object XX equipped with an action of 𝐆𝐋\mathbf{GL}, we define its nnth shift, denoted Shn(X)\operatorname{Sh}_{n}(X), to be the same object XX but with 𝐆𝐋\mathbf{GL} acting through the self-embedding 𝐆𝐋G(n)𝐆𝐋\mathbf{GL}\cong G(n)\subset\mathbf{GL}.

One easily sees that if VV is a polynomial representation of 𝐆𝐋\mathbf{GL} then Shn(V)\operatorname{Sh}_{n}(V) is also such a representation. From the polynomial functor point of view, we have

(ShnV){U}=V{knU}.(\operatorname{Sh}_{n}{V})\{U\}=V\{k^{n}\oplus U\}.

If VV has finite length then so does Shn(V)\operatorname{Sh}_{n}(V). It follows that if λ¯\underline{\smash{\lambda}} is a tuple then there is another tuple, which we denote by shn(λ¯)\operatorname{sh}_{n}(\underline{\smash{\lambda}}), such that Shn(kλ¯)=kshn(λ¯)\operatorname{Sh}_{n}(k^{\oplus\underline{\smash{\lambda}}})=k^{\oplus\operatorname{sh}_{n}(\underline{\smash{\lambda}})}. If λ¯=[λ]\underline{\smash{\lambda}}=[\lambda] consists of a single partition, we write shn(λ)\operatorname{sh}_{n}(\lambda) in place of shn(λ¯)\operatorname{sh}_{n}(\underline{\smash{\lambda}}). In this case, shn(λ)\operatorname{sh}_{n}(\lambda) contains λ\lambda exactly once, and all other partitions in it are strictly smaller.

2.4. 𝐆𝐋\mathbf{GL}-algebras

A 𝐆𝐋\mathbf{GL}-algebra (over kk) is a commutative algebra object in the tensor category Reppol(𝐆𝐋)\operatorname{Rep}^{\mathrm{pol}}(\mathbf{GL}); thus, it is a commutative (and associative and unital) kk-algebra equipped with an action of the group 𝐆𝐋\mathbf{GL} by algebra automorphisms, under which it forms a polynomial representation. Let RR be a 𝐆𝐋\mathbf{GL}-algebra. By an RR-module we mean a module object in Reppol(𝐆𝐋\operatorname{Rep}^{\mathrm{pol}}(\mathbf{GL}). Explicitly, this is an ordinary RR-module MM equipped with a compatible action of 𝐆𝐋\mathbf{GL} under which MM forms a polynomial representation. We let ModR\operatorname{Mod}_{R} denote the category of modules, which is easily seen to be a Grothendieck abelian category.

We say that RR is 𝐆𝐋\mathbf{GL}-generated (over kk) by a set of elements if RR is generated as a kk-algebra by the orbits of these elements. We say that RR is finitely 𝐆𝐋\mathbf{GL}-generated if it is 𝐆𝐋\mathbf{GL}-generated by a finite set. We similarly speak of 𝐆𝐋\mathbf{GL}-generation for RR-modules.

We say that a 𝐆𝐋\mathbf{GL}-algebra is integral if it is integral in the usual sense (i.e., it is a domain). We will require the following important shift theorem from [BDES].

Theorem 2.1.

Let RR be an integral 𝐆𝐋\mathbf{GL}-algebra that is finitely 𝐆𝐋\mathbf{GL}-generated. Then there exists n0n\geq 0, a non-zero 𝐆𝐋\mathbf{GL}-invariant element fShn(R)f\in\operatorname{Sh}_{n}(R), and an isomorphism Shn(R)[1/f]ASym(kσ¯)\operatorname{Sh}_{n}(R)[1/f]\cong A\otimes\operatorname{Sym}(k^{\oplus\underline{\smash{\sigma}}}) for some finitely generated integral kk-algebra AA (with trivial 𝐆𝐋\mathbf{GL}-action) and pure tuple σ¯\underline{\smash{\sigma}}.

Proof.

This is [BDES, Theorem 5.1], phrased in terms of coordinate rings. ∎

2.5. 𝐆𝐋\mathbf{GL}-varieties

An affine 𝐆𝐋\mathbf{GL}-scheme is an affine scheme XX over kk equipped with an action of the discrete group 𝐆𝐋\mathbf{GL} such that Γ(X,𝒪X)\Gamma(X,\mathcal{O}_{X}) forms a polynomial representation of 𝐆𝐋\mathbf{GL}. Every affine 𝐆𝐋\mathbf{GL}-scheme has the form Spec(R)\operatorname{Spec}(R) where RR is a 𝐆𝐋\mathbf{GL}-algebra. An affine 𝐆𝐋\mathbf{GL}-variety is a reduced affine 𝐆𝐋\mathbf{GL}-scheme XX such that Γ(X,𝒪X)\Gamma(X,\mathcal{O}_{X}) is finitely 𝐆𝐋\mathbf{GL}-generated over kk.

For a tuple λ¯\underline{\smash{\lambda}}, let 𝐀λ¯\mathbf{A}^{\underline{\smash{\lambda}}} be the spectrum of the ring Sym(kλ¯)\operatorname{Sym}(k^{\oplus\underline{\smash{\lambda}}}). This is an affine 𝐆𝐋\mathbf{GL}-variety. Moreover, every affine 𝐆𝐋\mathbf{GL}-variety is isomorphic to a closed 𝐆𝐋\mathbf{GL}-subvariety of some 𝐀λ¯\mathbf{A}^{\underline{\smash{\lambda}}}. Thus, in the theory of 𝐆𝐋\mathbf{GL}-varieties, the 𝐀λ¯\mathbf{A}^{\underline{\smash{\lambda}}} play the same role as the ordinary affine spaces 𝐀n\mathbf{A}^{n} in ordinary algebraic geometry.

Let XX be an affine 𝐆𝐋\mathbf{GL}-variety and let xx be a (scheme-theoretic) point of XX. We let O¯x\overline{O}_{x} be the Zariski closure of the orbit 𝐆𝐋x\mathbf{GL}\cdot x of xx (see [BDES, §3.1]). We say that xx is 𝐆𝐋\mathbf{GL}-generic if O¯x=X\overline{O}_{x}=X. Such points play a similar role to generic points in ordinary algebraic geometry. We define the generalized orbit of xx, denoted OxO_{x}, to be the set of all points yy such that O¯x=O¯y\overline{O}_{x}=\overline{O}_{y} (see [BDES, §3.2]).

Write X=Spec(R)X=\operatorname{Spec}(R) where RR is a 𝐆𝐋\mathbf{GL}-algebra. Recall that for a vector space UU we let R{U}R\{U\} be the result of treating RR as a polynomial functor and evaluating on UU; this is a kk-algebra equipped with an action of 𝐆𝐋(U)\mathbf{GL}(U). We put X{U}=Spec(R{U})X\{U\}=\operatorname{Spec}(R\{U\}). The standard inclusion kn𝐕k^{n}\to\mathbf{V} induces a ring homomorphism R{kn}RR\{k^{n}\}\to R, and thus a map of kk-schemes πn:XX{kn}\pi_{n}\colon X\to X\{k^{n}\}. Since RR is the union of the R{kn}R\{k^{n}\}, it follows that XX is the inverse limit of the X{kn}X\{k^{n}\}. We define the Π\Pi-topology on XX to be the inverse limit topology, where each X{kn}X\{k^{n}\} is given the discrete topology. The Π\Pi-topology is actually quite concrete: if kk is algebraically closed then the set of closed points of 𝐀λ¯\mathbf{A}^{\underline{\smash{\lambda}}} is identified with a product of kk’s, and the Π\Pi-topology is just the usual product topology; thus a sequence of kk-points of 𝐀λ¯\mathbf{A}^{\underline{\smash{\lambda}}} converges if each coordinate is eventually constant. One easily sees that any Zariski closed set is Π\Pi-closed (see [NS2, Proposition 2.3]).

We require the following result that relates the Zariski and Π\Pi-topologies:

Proposition 2.2.

Suppose that kk is algebraically closed. Let XX be a 𝐆𝐋\mathbf{GL}-variety and let xx and yy be kk-points of XX. Then the following conditions are equivalent:

  1. (a)

    The orbits 𝐆𝐋x\mathbf{GL}\cdot x and 𝐆𝐋y\mathbf{GL}\cdot y have the same Zariski cloure.

  2. (b)

    The orbits 𝐆𝐋x\mathbf{GL}\cdot x and 𝐆𝐋y\mathbf{GL}\cdot y have the same Π\Pi-closure.

Proof.

Suppose (b) holds. Then yy belongs to the Π\Pi-closure of 𝐆𝐋x\mathbf{GL}\cdot x, which is contained in the Zariski closure of 𝐆𝐋x\mathbf{GL}\cdot x. We thus see that 𝐆𝐋y\mathbf{GL}\cdot y is contained in the Zariski closure of 𝐆𝐋x\mathbf{GL}\cdot x, and so the Zariski closure of 𝐆𝐋y\mathbf{GL}\cdot y is contained in the Zariski closure of 𝐆𝐋x\mathbf{GL}\cdot x. By symmetry, the reverse inclusion holds as well, which yields (a).

Now suppose that (a) holds. We may as well replace XX with the Zariski closure of 𝐆𝐋x\mathbf{GL}\cdot x, and so that xx and yy are 𝐆𝐋\mathbf{GL}-generic in XX. Let φ:B×𝐀λ¯X\varphi\colon B\times\mathbf{A}^{\underline{\smash{\lambda}}}\to X be a typical morphism (see [BDES, §8.1]), where BB is an irreducible variety and λ¯\underline{\smash{\lambda}} is a pure tuple. Let (b,x~)B×𝐀λ¯(b,\tilde{x})\in B\times\mathbf{A}^{\underline{\smash{\lambda}}} be a kk-point lifting xx, which exists by [BDES, Proposition 7.15], and let ZZ be the closure of the 𝐆𝐋\mathbf{GL}-orbit of (b,x~)(b,\tilde{x}). Then φ|Z\varphi|_{Z} is dominant since its image contains xx, and so, by the definition of typical, Z=B×𝐀λ¯Z=B\times\mathbf{A}^{\underline{\smash{\lambda}}}. It follows that B={b}B=\{b\} is a point and x~\tilde{x} is 𝐆𝐋\mathbf{GL}-generic in 𝐀λ¯\mathbf{A}^{\underline{\smash{\lambda}}}. In what follows we ignore BB, and regard φ\varphi as a morphism φ:𝐀λ¯X\varphi\colon\mathbf{A}^{\underline{\smash{\lambda}}}\to X satisfying φ(x~)=x\varphi(\tilde{x})=x.

The image of φ\varphi contains a non-empty open subset of XX by [BDES, Theorem 7.13]. Since yy belongs to every non-empty 𝐆𝐋\mathbf{GL}-subset of XX [BDES, Proposition 3.4], we see that yim(φ)y\in\operatorname{im}(\varphi). Thus, applying [BDES, Proposition 7.15] again, we can find a kk-point y~\tilde{y} of 𝐀λ¯\mathbf{A}^{\underline{\smash{\lambda}}} such that φ(y~)=y\varphi(\tilde{y})=y.

Let πn:𝐀λ¯𝐀λ¯{Kn}\pi_{n}\colon\mathbf{A}^{\underline{\smash{\lambda}}}\to\mathbf{A}^{\underline{\smash{\lambda}}}\{K^{n}\} be the natural map. By [BDDE, Corollary 2.6.3], the restriction of πn\pi_{n} to 𝐆𝐋x~\mathbf{GL}\cdot\tilde{x} is surjective on kk-points. We can thus find gn𝐆𝐋g_{n}\in\mathbf{GL} such that πn(gnx~)=y~\pi_{n}(g_{n}\tilde{x})=\tilde{y}. We therefore see that the sequence (gnx~)n1(g_{n}\tilde{x})_{n\geq 1} converges to y~\tilde{y} in the Π\Pi-topology. Since φ\varphi is Π\Pi-continuous, it follows that the sequence (gnx)n1(g_{n}x)_{n\geq 1} converges to yy in the Π\Pi-topology. Thus yy, and therefore 𝐆𝐋y\mathbf{GL}\cdot y, and therefore the Π\Pi-closure of 𝐆𝐋y\mathbf{GL}\cdot y, is contained in the Π\Pi-closure of 𝐆𝐋x\mathbf{GL}\cdot x. The reverse inclusion follows by symmetry, and so (b) holds. ∎

Remark 2.3.

Proposition 2.2 shows that, when working with closed points over an algebraically closed field, one can define the generalized orbit of xx using the Π\Pi-topology (as we did in §1.1): that is, a kk-point yy belongs to OxO_{x} if and only if one can find sequences (gn)n1(g_{n})_{n\geq 1} and (hn)n1(h_{n})_{n\geq 1} in 𝐆𝐋\mathbf{GL} such that gnxyg_{n}x\to y and hnyxh_{n}y\to x in the Π\Pi-topology. ∎

Remark 2.4.

We only apply Proposition 2.2 when X=𝐀λ¯X=\mathbf{A}^{\underline{\smash{\lambda}}}, in which case the proof simplifies some. However, we feel that the general statement is important enough that it is worth recording here. ∎

2.6. 𝐆𝐋\mathbf{GL}-fields

A 𝐆𝐋\mathbf{GL}-field over kk is a field extension K/kK/k equipped with an action of 𝐆𝐋\mathbf{GL} by kk-automorphisms such that every element of KK can be expressed in the form a/ba/b with a,bKpola,b\in K^{\mathrm{pol}}. If KK is a 𝐆𝐋\mathbf{GL}-field then KpolK^{\mathrm{pol}} is an integral 𝐆𝐋\mathbf{GL}-algebra over kk, and K=Frac(Kpol)K=\operatorname{Frac}(K^{\mathrm{pol}}). Thus every 𝐆𝐋\mathbf{GL}-field can be realized as the fraction field of an integral 𝐆𝐋\mathbf{GL}-algebra.

Let KK be a 𝐆𝐋\mathbf{GL}-field. A KK-module is a semi-linear representation MM of 𝐆𝐋\mathbf{GL} over KK such that every element of MM has the form axax with aKa\in K and xMpolx\in M^{\mathrm{pol}}. One easily sees that the category ModK\operatorname{Mod}_{K} of KK-modules is an abelian category satisfying the (AB5) condition. Moreover, if MM is any KK-module then there is a surjection KVMK\otimes V\to M for some polynomial representation VV (take V=MpolV=M^{\mathrm{pol}}), which shows that the objects KλK^{\oplus\lambda} form a generating set; thus ModK\operatorname{Mod}_{K} is a Grothendieck abelian category.

We say that KK is finitely 𝐆𝐋\mathbf{GL}-generated over kk if it is generated as a field extension by the 𝐆𝐋\mathbf{GL}-orbits of finitely many elements. We say that KK is rational over kk if it has the form Frac(Sym(kσ¯))\operatorname{Frac}(\operatorname{Sym}(k^{\oplus\underline{\smash{\sigma}}})) for some tuple σ¯\underline{\smash{\sigma}}. The invariant subfield of KK, denoted K𝐆𝐋K^{\mathbf{GL}}, is the subfield of KK consisting of all elements that are invariant under 𝐆𝐋\mathbf{GL}. It is an extension of kk. If KK is finitely 𝐆𝐋\mathbf{GL}-generated over kk then K𝐆𝐋K^{\mathbf{GL}} is finitely generated over kk ([BDES, Proposition 5.8]).

Proposition 2.5.

Let KK be a 𝐆𝐋\mathbf{GL}-field that is finitely 𝐆𝐋\mathbf{GL}-generated over kk. Then there exists n0n\geq 0 such that Shn(K)\operatorname{Sh}_{n}(K) is rational over its invariant subfield.

Proof.

One easily sees that KK can be 𝐆𝐋\mathbf{GL}-generated by finitely many polynomial elements. We can thus find a finitely 𝐆𝐋\mathbf{GL}-generated kk-subalgebra RR of KK such that K=Frac(R)K=\operatorname{Frac}(R). Apply Theorem 2.1 to write Shn(R)[1/f]ASym(kσ¯)\operatorname{Sh}_{n}(R)[1/f]\cong A\otimes\operatorname{Sym}(k^{\oplus\underline{\smash{\sigma}}}) where AA is a kk-algebra with trivial 𝐆𝐋\mathbf{GL}-action and σ¯\underline{\smash{\sigma}} is a pure tuple. Taking fraction fields, we find Shn(K)Frac(Sym(σ¯))\operatorname{Sh}_{n}(K)\cong\operatorname{Frac}(\operatorname{Sym}(\ell^{\oplus\underline{\smash{\sigma}}})) where =Frac(A)\ell=\operatorname{Frac}(A). It follows from [BDES, Proposition 5.7] that K𝐆𝐋K^{\mathbf{GL}}\cong\ell, and so Shn(K)\operatorname{Sh}_{n}(K) is rational over its invariant subfield. ∎

2.7. Generic categories

Let RR be an integral 𝐆𝐋\mathbf{GL}-algebra. We say that an RR-module MM is torsion if every element of MM is annihilated by a non-zero element of RR. The category ModRtors\operatorname{Mod}_{R}^{\mathrm{tors}} of torsion RR-modules is a Serre subcategory of ModR\operatorname{Mod}_{R}. We define the generic category of RR, denoted ModRgen\operatorname{Mod}_{R}^{\mathrm{gen}}, to be the Serre quotient ModR/ModRtors\operatorname{Mod}_{R}/\operatorname{Mod}_{R}^{\mathrm{tors}}.

The generic category can be described in terms of semi-linear representations. Let K=Frac(R)K=\operatorname{Frac}(R). We have a functor

T:ModRModK,T(M)=KRM.T\colon\operatorname{Mod}_{R}\to\operatorname{Mod}_{K},\qquad T(M)=K\otimes_{R}M.

We also have a functor

S:ModKModR,T(N)=Npol.S\colon\operatorname{Mod}_{K}\to\operatorname{Mod}_{R},\qquad T(N)=N^{\mathrm{pol}}.

Indeed, if NN is a KK-module then RNpolR\otimes N^{\mathrm{pol}} is a polynomial representation, so its image under the natural map RNpolNR\otimes N^{\mathrm{pol}}\to N consists of polynomial elements, and is therefore contained in NpolN^{\mathrm{pol}}; this shows that NpolN^{\mathrm{pol}} is stable under multiplication by RR, and is thus an RR-module.

Proposition 2.6.

We have the following:

  1. (a)

    The functor TT is exact and kills torsion modules. The induced functor ModRgenModK\operatorname{Mod}_{R}^{\mathrm{gen}}\to\operatorname{Mod}_{K} is an equivalence.

  2. (b)

    The functors (T,S)(T,S) form an adjoint pair.

  3. (c)

    The co-unit TSidTS\to\mathrm{id} is an isomorphism.

Proof.

See [NSS, §2.4]. ∎

We say that an RR-module MM is saturated if the natural map MS(T(M))M\to S(T(M)) is an isomorphism. We will require the following result concerning this concept:

Proposition 2.7.

Let σ¯\underline{\smash{\sigma}} be a pure tuple, let R=Sym(kσ¯)R=\operatorname{Sym}(k^{\oplus\underline{\smash{\sigma}}}), and let VV be a polynomial representation. Then RVR\otimes V is a saturated RR-module.

Proof.

See [NSS, Proposition 2.8]. ∎

3. The shift and embedding theorems

In this section, we prove our two main technical results on KK-modules: the shift theorem (Theorem 3.3) and the embedding theorem (Theorem 3.9).

3.1. A preliminary result

The following proposition is the key input needed for the shift theorem proven in the subsequent subsection. It is a linear analog of [BDES, Theorem 4.2], a result that was essentially taken from arguments in [Dr].

Proposition 3.1.

Let RR be an integral 𝐆𝐋\mathbf{GL}-algebra, let λ\lambda be a partition, let FF and MM be RR-modules, and suppose we have a surjection of RR-modules

RλFM.R^{\oplus\lambda}\oplus F\to M.

Then at least one of the following holds:

  1. (a)

    The given map induces an isomorphism RλNMR^{\oplus\lambda}\oplus N\to M, where NN is a quotient of FF.

  2. (b)

    There exists n0n\geq 0 and a non-zero 𝐆𝐋\mathbf{GL}-invariant element fShn(R)f\in\operatorname{Sh}_{n}(R) such that the natural map

    Shn(R)[1/f]μ¯Shn(F)[1/f]Shn(M)[1/f]\operatorname{Sh}_{n}(R)[1/f]^{\oplus\underline{\smash{\mu}}}\oplus\operatorname{Sh}_{n}(F)[1/f]\to\operatorname{Sh}_{n}(M)[1/f]

    is surjective, where μ¯\underline{\smash{\mu}} is obtained from shn(λ)\operatorname{sh}_{n}(\lambda) by deleting λ\lambda.

We require some preparation before giving the proof. A weight of 𝐆𝐋\mathbf{GL} is a tuple λ=(λ1,λ2,)\lambda=(\lambda_{1},\lambda_{2},\ldots) where λi𝐙\lambda_{i}\in\mathbf{Z} for all ii and λi=0\lambda_{i}=0 for i0i\gg 0. For a finite subset AA of []={1,2,}[\infty]=\{1,2,\ldots\}, we let 1A1^{A} be the weight that is 1 at the coordinates in AA, and 0 away from AA. We also write 1n1^{n} in place of 1A1^{A} when A=[n]A=[n].

Suppose that VV is a polynomial representation and λ\lambda is a weight. We say that vVv\in V is a weight vector of weight λ\lambda if whenever g=diag(a1,a2,)g=\operatorname{diag}(a_{1},a_{2},\ldots) we have

gv=(i1aiλi)v.gv=\big{(}\prod_{i\geq 1}a_{i}^{\lambda_{i}}\big{)}\cdot v.

We let VλV_{\lambda} be the space of all weight vectors of weight λ\lambda; this is the λ\lambda weight space. The space VV is the direct sum of its weight spaces VλV_{\lambda} over all λ\lambda. Moreover, if VλV_{\lambda} is non-zero then λ\lambda is non-negative in the sense that λi0\lambda_{i}\geq 0 for all ii.

The weight space V1nV_{1^{n}} carries a representation of 𝔖n𝐆𝐋\mathfrak{S}_{n}\subset\mathbf{GL}. Let Reppol,n(𝐆𝐋)\operatorname{Rep}^{\mathrm{pol},n}(\mathbf{GL}) be the subcategory of Reppol(𝐆𝐋)\operatorname{Rep}^{\mathrm{pol}}(\mathbf{GL}) spanned by representations of degree nn. One formulation of Schur–Weyl duality states that the functor

Reppol,n(𝐆𝐋)\displaystyle\operatorname{Rep}^{\mathrm{pol},n}(\mathbf{GL}) Rep(𝔖n)\displaystyle\to\operatorname{Rep}(\mathfrak{S}_{n})
V\displaystyle V V1n\displaystyle\mapsto V_{1^{n}}

is an equivalence of categories.

Lemma 3.2.

Let VV and WW be polynomial representations of degrees nn and mm, with VV irreducible, let SS be a subset of [n+m][n+m] of cardinality nn, and let UU be a non-zero subrepresentation of VWV\otimes W. Then UU contains a vector of the form x=i=1rviwix=\sum_{i=1}^{r}v_{i}\otimes w_{i}, for some r1r\geq 1, such that:

  • viv_{i} is a weight vector of VV of weight 1Ai1^{A_{i}} and wiw_{i} is a weight vector of WW of weight 1Bi1^{B_{i}}, where AiA_{i} and BiB_{i} are disjoint and AiBi=[n+m]A_{i}\cup B_{i}=[n+m];

  • we have A1=SA_{1}=S, and v1v_{1} and w1w_{1} are non-zero;

  • we have AiSA_{i}\neq S for i>1i>1.

Proof.

We may as well assume S=[n]S=[n]. By Schur–Weyl duality, the 1n+m1^{n+m}-weight space of UU is non-zero. We can thus find a non-zero element xx of UU of the form x=i=1rviwix=\sum_{i=1}^{r}v_{i}\otimes w_{i} satisfying the first condition, and with the viv_{i} and wiw_{i} linearly independent. Applying an element of the symmetric group 𝔖n+m𝐆𝐋\mathfrak{S}_{n+m}\subset\mathbf{GL}, we can assume that v1v_{1} has weight 1n1^{n}. Relabeling, we can assume that v1,,vkv_{1},\ldots,v_{k} have weight 1n1^{n}, and that the remaining viv_{i} have weight 1n\neq 1^{n}.

Now, by Schur–Weyl duality, the 1n1^{n} weight space of VV is an irreducible representation of 𝔖n\mathfrak{S}_{n} (acting through the standard inclusion 𝔖n𝐆𝐋\mathfrak{S}_{n}\subset\mathbf{GL}). Since v1,,vkv_{1},\ldots,v_{k} are linearly independent elements, we can find a𝐂[𝔖n]a\in\mathbf{C}[\mathfrak{S}_{n}] such that av1=v1av_{1}=v_{1} and avi=0av_{i}=0 for 2ik2\leq i\leq k, Since w1w_{1} has weight 1B11^{B_{1}} with B1={n+1,,n+m}B_{1}=\{n+1,\ldots,n+m\}, the group 𝔖n\mathfrak{S}_{n} acts trivially on it, and so a(v1w1)=v1w1a(v_{1}\otimes w_{1})=v_{1}\otimes w_{1}. For k<ik<i the element aviav_{i} is a sum of weight vectors having weight of the form 1A1^{A} with A[n]A\neq[n]. We thus see that axax is an element of UU of the required form. ∎

Proof of Proposition 3.1.

Let KK be the kernel of RλFMR^{\oplus\lambda}\oplus F\to M, and let K¯\overline{K} be the projection of KK to RλR^{\oplus\lambda}. If K¯=0\overline{K}=0 then KK is contained in FF, and case (a) holds with N=F/KN=F/K. Suppose now that K¯0\overline{K}\neq 0. Let n=|λ|n=|\lambda| and let m0m\geq 0 be such that K¯\overline{K} has a non-zero element of degree n+mn+m. Recall that Rλ=ERR^{\oplus\lambda}=E\otimes R, where E=kλE=k^{\oplus\lambda}. Applying Lemma 3.2, we can find an element xx of K¯\overline{K} of the form x=i=1rfieix=\sum_{i=1}^{r}f_{i}e_{i}, where:

  • eie_{i} is a weight vector of EE of weight 1Ai1^{A_{i}} and fif_{i} is a weight vector of RR of weight 1Bi1^{B_{i}}, where AiA_{i} and BiB_{i} are disjoint and AiBi=[n+m]A_{i}\cup B_{i}=[n+m];

  • A1={m+1,,n+m}A_{1}=\{m+1,\ldots,n+m\}, and f1f_{1} and e1e_{1} are non-zero.

  • Ai{m+1,,n+m}A_{i}\neq\{m+1,\ldots,n+m\} for i>1i>1.

Say that a weight λ\lambda is big if λi=0\lambda_{i}=0 for i[m]i\in[m], and small otherwise. Let EbigE^{\rm big} and EsmallE^{\rm small} be the sum of the big and small weight spaces in EE. Then we have a decomposition of G(m)G(m)-representations

E=EbigEsmall.E=E^{\rm big}\oplus E^{\rm small}.

Identifying G(m)G(m) with 𝐆𝐋\mathbf{GL}, this becomes the decomposition

Shm(E)=kλkμ¯.\operatorname{Sh}_{m}(E)=k^{\oplus\lambda}\oplus k^{\oplus\underline{\smash{\mu}}}.

We thus see that EbigE^{\rm big} is irreducible as a G(m)G(m)-representation. Note that e1e_{1} is a non-zero element of EbigE^{\rm big} (and thus generates it as a G(m)G(m)-representation), and that f1f_{1} is G(m)G(m)-invariant (as it has weight 1m1^{m}).

Let yFy\in F be such that x+yKx+y\in K. Let MM^{\prime} be the image of (EsmallR)F(E^{\rm small}\otimes R)\oplus F in MM. Since x+yx+y maps to 0 in MM, we see that the image of f1e1f_{1}e_{1} in MM belongs to MM^{\prime}, and so the image of e1e_{1} belongs to M[1/f1]M^{\prime}[1/f_{1}]. Since MM^{\prime} is G(m)G(m)-stable, f1f_{1} is G(m)G(m)-invariant, and e1e_{1} generates EbigE^{\rm big} as a G(m)G(m)-representation, we see that any element of EbigRE^{\rm big}\otimes R maps into M[1/f1]M^{\prime}[1/f_{1}]. Thus M[1/f1]=M[1/f1]M^{\prime}[1/f_{1}]=M[1/f_{1}], and the result follows. ∎

3.2. The shift theorem

We now prove the first main result of this section. It is an analog of [BDES, Theorem 5.1].

Theorem 3.3 (Shift theorem).

Let RR be an integral 𝐆𝐋\mathbf{GL}-algebra and let MM be a finitely generated RR-module. Then there exists n0n\geq 0, a tuple λ¯\underline{\smash{\lambda}}, and a non-zero 𝐆𝐋\mathbf{GL}-invariant element fShn(R)f\in\operatorname{Sh}_{n}(R) such that we have an isomorphism Shn(M)[1/f]Shn(R)[1/f]λ¯\operatorname{Sh}_{n}(M)[1/f]\cong\operatorname{Sh}_{n}(R)[1/f]^{\oplus\underline{\smash{\lambda}}} of Shn(R)[1/f]\operatorname{Sh}_{n}(R)[1/f]-modules.

Proof.

Say that an RR-module is good if the conclusion of the theorem holds for it. Consider the following statement, for a tuple μ¯\underline{\smash{\mu}}:

  • S(μ¯)S(\underline{\smash{\mu}})

    If RR is an integral 𝐆𝐋\mathbf{GL}-algebra and MM is a quotient module of Rμ¯R^{\oplus\underline{\smash{\mu}}} then MM is good.

It suffices to prove S(μ¯)S(\underline{\smash{\mu}}) for all tuples μ¯\underline{\smash{\mu}}. The magnitude of a tuple μ¯\underline{\smash{\mu}}, denoted magn(μ¯)\operatorname{magn}(\underline{\smash{\mu}}), is the tuple (n0,n1,)(n_{0},n_{1},\ldots) where nin_{i} is the number of partitions of size ii in μ\mu. We order magnitudes lexicographically; this is a well-order. We can thus prove S(μ¯)S(\underline{\smash{\mu}}) by induction on magn(μ¯)\operatorname{magn}(\underline{\smash{\mu}}). Thus let μ¯\underline{\smash{\mu}} be given, and suppose S(ν¯)S(\underline{\smash{\nu}}) holds for all ν¯\underline{\smash{\nu}} with magn(ν¯)<magn(μ¯)\operatorname{magn}(\underline{\smash{\nu}})<\operatorname{magn}(\underline{\smash{\mu}}). We prove S(μ¯)S(\underline{\smash{\mu}}). If μ¯\underline{\smash{\mu}} is empty the statement is vacuous, so suppose this is not the case.

Let RR be an integral 𝐆𝐋\mathbf{GL}-algebra and let MM be a quotient of Rμ¯R^{\oplus\underline{\smash{\mu}}}. Let κ\kappa be a partition in μ¯\underline{\smash{\mu}} of maximal size, and let ν¯\underline{\smash{\nu}} be the tuple obtained from μ¯\underline{\smash{\mu}} by deleting κ\kappa. We thus have a surjection RκRν¯MR^{\oplus\kappa}\oplus R^{\oplus\underline{\smash{\nu}}}\to M. We apply Proposition 3.1 with F=Rν¯F=R^{\oplus\underline{\smash{\nu}}}. We consider the two cases separately.

Suppose case (a) holds. Then M=RκNM=R^{\oplus\kappa}\oplus N where NN is a quotient of Rν¯R^{\oplus\underline{\smash{\nu}}}. Since ν¯\underline{\smash{\nu}} has smaller magnitude than μ¯\underline{\smash{\mu}}, statement S(ν¯)S(\underline{\smash{\nu}}) holds, and so NN is good. It is clear then that MM is good as well.

Now suppose case (b) holds. Then there is some n0n\geq 0 and a 𝐆𝐋\mathbf{GL}-invariant function fShn(R)f\in\operatorname{Sh}_{n}(R) such that the natural map

Shn(R)[1/f]ρ¯Shn(Rν¯)[1/f]Shn(M)[1/f]\operatorname{Sh}_{n}(R)[1/f]^{\oplus\underline{\rho}}\oplus\operatorname{Sh}_{n}(R^{\oplus\underline{\smash{\nu}}})[1/f]\to\operatorname{Sh}_{n}(M)[1/f]

is a surjection, where shn(κ)=[κ]ρ¯\operatorname{sh}_{n}(\kappa)=[\kappa]\cup\underline{\rho}. Now, the left side above has the form Shn(R)[1/f]σ¯\operatorname{Sh}_{n}(R)[1/f]^{\oplus\underline{\sigma}} where σ¯=ρ¯shn(ν¯)\underline{\sigma}=\underline{\rho}\cup\operatorname{sh}_{n}(\underline{\smash{\nu}}). The tuple σ¯\underline{\sigma} has smaller magnitude than μ¯\underline{\smash{\mu}}, and so statement S(σ¯)S(\underline{\sigma}) holds. We thus see that Shn(M)[1/f]\operatorname{Sh}_{n}(M)[1/f] is good as a Shn(R)[1/f]\operatorname{Sh}_{n}(R)[1/f]-module, from which it easily follows that MM is good as an RR-module. This completes the proof. ∎

We also have the following statement, which appears to be slightly stronger, but in fact follows easily from the theorem:

Corollary 3.4.

Let RR be an integral 𝐆𝐋\mathbf{GL}-algebra and let MM be a finitely generated RR-module. Then there exists n0n\geq 0, a tuple λ¯\underline{\smash{\lambda}}, and a non-zero 𝐆𝐋\mathbf{GL}-invariant element fShn(R)f\in\operatorname{Sh}_{n}(R) such that there is an injection Shn(R)λ¯Shn(M)\operatorname{Sh}_{n}(R)^{\oplus\underline{\smash{\lambda}}}\to\operatorname{Sh}_{n}(M) of Shn(R)\operatorname{Sh}_{n}(R)-modules with cokernel annihilated by ff.

Proof.

Let nn, λ¯\underline{\smash{\lambda}}, and ff be as in the shift theorem, so that we have an isomorphism of Shn(R)[1/f]\operatorname{Sh}_{n}(R)[1/f]-modules Shn(R)[1/f]λ¯Shn(M)[1/f]\operatorname{Sh}_{n}(R)[1/f]^{\oplus\underline{\smash{\lambda}}}\to\operatorname{Sh}_{n}(M)[1/f]. Let MM^{\prime} be the image of Shn(M)\operatorname{Sh}_{n}(M) in Shn(M)[1/f]\operatorname{Sh}_{n}(M)[1/f]. Scaling our isomorphism by an appropriate power of ff, we can assume that Shn(R)λ¯\operatorname{Sh}_{n}(R)^{\oplus\underline{\smash{\lambda}}} maps into MM^{\prime}. Since Shn(R)λ¯\operatorname{Sh}_{n}(R)^{\oplus\underline{\smash{\lambda}}} is projective, we can find a lift Shn(R)λ¯Shn(M)\operatorname{Sh}_{n}(R)^{\oplus\underline{\smash{\lambda}}}\to\operatorname{Sh}_{n}(M) of our map, which is necessarily injective. Since this map is an isomorphism after inverting ff, every element in the cokernel is annihilated by a power of ff. But the cokernel is finitely 𝐆𝐋\mathbf{GL}-generated and ff is 𝐆𝐋\mathbf{GL}-invariant, so there is some power of ff that annihilates the entire cokernel. Replace ff by this power. ∎

The shift theorem for 𝐆𝐋\mathbf{GL}-algebras implies an analogous result for 𝐆𝐋\mathbf{GL}-fields:

Corollary 3.5.

Let KK be a 𝐆𝐋\mathbf{GL}-field and let MM be a finitely generated KK-module. Then there exists n0n\geq 0 and a tuple λ¯\underline{\smash{\lambda}} such that we have an isomorphism Shn(M)Shn(K)λ¯\operatorname{Sh}_{n}(M)\cong\operatorname{Sh}_{n}(K)^{\oplus\underline{\smash{\lambda}}} of Shn(K)\operatorname{Sh}_{n}(K)-modules.

Proof.

Let R=KpolR=K^{\mathrm{pol}} and let NMpolN\subset M^{\mathrm{pol}} be a finitely 𝐆𝐋\mathbf{GL}-generated RR-module that spans MM over KK. By Theorem 3.3, we have an isomorphism Shn(N)[1/f]Shn(R)[1/f]λ¯\operatorname{Sh}_{n}(N)[1/f]\cong\operatorname{Sh}_{n}(R)[1/f]^{\oplus\underline{\smash{\lambda}}} for some nn, ff, and λ¯\underline{\smash{\lambda}}. Tensoring up to Shn(K)\operatorname{Sh}_{n}(K), we obtain the stated result. ∎

3.3. Some consequences

We now give a few consequences of the shift theorem.

Proposition 3.6.

Let RR be an integral 𝐆𝐋\mathbf{GL}-algebra and let MM be a finitely generated RR-module. Then there exists a non-empty open 𝐆𝐋\mathbf{GL}-stable subset UU of Spec(R)\operatorname{Spec}(R) such that M𝔭M_{\mathfrak{p}} is free over R𝔭R_{\mathfrak{p}} for all 𝔭U\mathfrak{p}\in U.

Proof.

The shift theorem shows that Shn(M)[1/f]\operatorname{Sh}_{n}(M)[1/f] is free as an Shn(R)[1/f]\operatorname{Sh}_{n}(R)[1/f]-module, for some non-zero ff. Since freeness does not depend on the 𝐆𝐋\mathbf{GL}-actions, it follows that M[1/f]M[1/f] is free as an R[1/f]R[1/f]-module. Thus M𝔭M_{\mathfrak{p}} is free over R𝔭R_{\mathfrak{p}} for all 𝔭D(f)\mathfrak{p}\in D(f), where D(f)Spec(A)D(f)\subset\operatorname{Spec}(A) is the distinguished open defined by ff. Since the free locus is obvious 𝐆𝐋\mathbf{GL}-stable, we can take U=g𝐆𝐋gD(f)U=\bigcup_{g\in\mathbf{GL}}gD(f). ∎

Corollary 3.7.

Let RR be an integral 𝐆𝐋\mathbf{GL}-algebra, let MM be an RR-module, and let 𝔭\mathfrak{p} be a 𝐆𝐋\mathbf{GL}-generic prime of RR. Then M𝔭M_{\mathfrak{p}} is flat over R𝔭R_{\mathfrak{p}}.

Proof.

First suppose that MM is finitely generated. By Proposition 3.6, MM is flat at an non-empty 𝐆𝐋\mathbf{GL}-stable open subset of Spec(R)\operatorname{Spec}(R). Such a subset contains all 𝐆𝐋\mathbf{GL}-generic points [BDES, Proposition 3.4]. Thus MM is flat at 𝔭\mathfrak{p}. In general, write M=limMiM=\varinjlim M_{i} with each MiM_{i} finitely generated. Then M𝔭=lim(Mi)𝔭M_{\mathfrak{p}}=\varinjlim(M_{i})_{\mathfrak{p}} is a direct limit of flat modules, and thus flat. ∎

The following result shows that finitely generated modules have “bounded torsion” in an appropriate sense:

Proposition 3.8.

Let RR be an integral 𝐆𝐋\mathbf{GL}-algebra, let MM be a finitely generated RR-module, and let MtorsM_{\rm tors} be the torsion submodule of MM. Then there exists a non-zero fRf\in R such that fMtors=0fM_{\rm tors}=0.

Proof.

Applying Corollary 3.4, let i:Shn(R)λ¯Shn(M)i\colon\operatorname{Sh}_{n}(R)^{\oplus\underline{\smash{\lambda}}}\to\operatorname{Sh}_{n}(M) be an injection of Shn(R)\operatorname{Sh}_{n}(R)-modules with cokernel annihilated by ff. Since Shn(Mtors)\operatorname{Sh}_{n}(M_{\rm tors}) is torsion, it cannot intersect im(i)\operatorname{im}(i), and so it injects into coker(i)\operatorname{coker}(i). Since coker(i)\operatorname{coker}(i) is annihilated by ff, so is MtorsM_{\rm tors}. ∎

3.4. The embedding theorem

We now prove the second main result of this section.

Theorem 3.9 (Embedding theorem).

Let AA be an integral kk-algebra, let σ¯\underline{\smash{\sigma}} be a tuple, and let R=ASym(kσ¯)R=A\otimes\operatorname{Sym}(k^{\oplus\underline{\smash{\sigma}}}). Let MM be a finitely 𝐆𝐋\mathbf{GL}-generated torsion-free RR-module. Then there is a tuple μ¯\underline{\smash{\mu}} and an injection MRμ¯M\to R^{\oplus\underline{\smash{\mu}}} of RR-modules.

We require some discussion before giving the proof. If VV is a polynomial representation of 𝐆𝐋\mathbf{GL} then VV is identified with V{𝐕}V\{\mathbf{V}\} and Shn(V)\operatorname{Sh}_{n}(V) is identified with V{kn𝐕}V\{k^{n}\oplus\mathbf{V}\}. The natural inclusion 𝐕kn𝐕\mathbf{V}\to k^{n}\oplus\mathbf{V} thus induces a map VShn(V)V\to\operatorname{Sh}_{n}(V), which is injective. If RR is a 𝐆𝐋\mathbf{GL}-algebra then the map RShn(R)R\to\operatorname{Sh}_{n}(R) is one of 𝐆𝐋\mathbf{GL}-algebras, and if MM is an RR-module then the map MShn(M)M\to\operatorname{Sh}_{n}(M) is one of RR-modules. We say that RR is shift-free if for each nn the RR-module Shn(R)\operatorname{Sh}_{n}(R) has the form EnRE_{n}\otimes R for some polynomial representation EnE_{n}. Theorem 3.9 thus follows from the following two lemmas.

Lemma 3.10.

Let AA be a kk-algebra, let σ¯\underline{\smash{\sigma}} be a tuple, and let R=ASym(kσ¯)R=A\otimes\operatorname{Sym}(k^{\oplus\underline{\smash{\sigma}}}). Then RR is shift-free.

Proof.

We have Shn(R)=ASym(Shn(kσ¯))\operatorname{Sh}_{n}(R)=A\otimes\operatorname{Sym}(\operatorname{Sh}_{n}(k^{\oplus\underline{\smash{\sigma}}})). Write Shn(kσ¯)=kσ¯kτ¯(n)\operatorname{Sh}_{n}(k^{\oplus\underline{\smash{\sigma}}})=k^{\oplus\underline{\smash{\sigma}}}\oplus k^{\oplus\underline{\smash{\tau}}(n)} for some tuple τ¯(n)\underline{\smash{\tau}}(n), and let En=Sym(kτ¯(n))E_{n}=\operatorname{Sym}(k^{\oplus\underline{\smash{\tau}}(n)}). Then Shn(R)EnR\operatorname{Sh}_{n}(R)\cong E_{n}\otimes R, as 𝐆𝐋\mathbf{GL}-algebras, and, in particular, as RR-modules. Thus RR is shift-free. ∎

Lemma 3.11.

Let RR be an integral shift-free 𝐆𝐋\mathbf{GL}-algebra and let MM be a finitely 𝐆𝐋\mathbf{GL}-generated torsion-free RR-module. Then there is a tuple μ¯\underline{\smash{\mu}} and an injection MRμ¯M\to R^{\oplus\underline{\smash{\mu}}} of RR-modules.

Proof.

Applying the shift theorem (Theorem 3.3), we have an isomorphism i:Shn(M)[1/f]=Shn(R)[1/f]λ¯i\colon\operatorname{Sh}_{n}(M)[1/f]=\operatorname{Sh}_{n}(R)[1/f]^{\oplus\underline{\smash{\lambda}}} for some nn, ff, and λ¯\underline{\smash{\lambda}}. Since MM is torsion-free, the natural map Shn(M)Shn(M)[1/f]\operatorname{Sh}_{n}(M)\to\operatorname{Sh}_{n}(M)[1/f] is injective. Scaling ii by a power of ff, we can assume it maps Shn(M)\operatorname{Sh}_{n}(M) into Shn(R)λ¯\operatorname{Sh}_{n}(R)^{\oplus\underline{\smash{\lambda}}}. Composing with the natural map MShn(M)M\to\operatorname{Sh}_{n}(M), we obtain an injection of RR-modules j:MShn(R)λ¯j\colon M\to\operatorname{Sh}_{n}(R)^{\oplus\underline{\smash{\lambda}}}. As RR-modules, we have Shn(R)EnR\operatorname{Sh}_{n}(R)\cong E_{n}\otimes R for some polynomial representation EnE_{n}. Thus we can identify the target of jj with FRF\otimes R where F=kλ¯EnF=k^{\oplus\underline{\smash{\lambda}}}\otimes E_{n}. Since MM is finitely generated, the image of jj is contained in F0RF_{0}\otimes R for some finite length subrepresentation F0F_{0} of FF. Writing F0kμ¯F_{0}\cong k^{\oplus\underline{\smash{\mu}}} for some tuple μ¯\underline{\smash{\mu}} yields the result. ∎

There is also an embedding theorem for rational 𝐆𝐋\mathbf{GL}-fields:

Corollary 3.12.

Let σ¯\underline{\smash{\sigma}} be a pure tuple, let K=Frac(Sym(kσ¯))K=\operatorname{Frac}(\operatorname{Sym}(k^{\oplus\underline{\smash{\sigma}}})), and let MM be a finitely generated KK-module. Then there exists an injection MKλ¯M\to K^{\oplus\underline{\smash{\lambda}}} of KK-modules for some tuple λ¯\underline{\smash{\lambda}}.

Proof.

Let R=Sym(kσ¯)R=\operatorname{Sym}(k^{\oplus\underline{\smash{\sigma}}}) and let M0MpolM_{0}\subset M^{\mathrm{pol}} be a finitely generated RR-module with M=KRM0M=K\otimes_{R}M_{0}. Since M0M_{0} is contained in MM, it is torsion-free. By Theorem 3.9, there is an injection of RR-modules M0Rλ¯M_{0}\to R^{\oplus\underline{\smash{\lambda}}} for some tuple λ¯\underline{\smash{\lambda}}. Tensoring up to KK gives the stated result. ∎

Remark 3.13.

Theorem 3.9 is a linear analog of [BDES, Theorem 5.4]. That result is stated for 𝐆𝐋\mathbf{GL}-varieties, but if formulated in terms of 𝐆𝐋\mathbf{GL}-algebras it states that certain 𝐆𝐋\mathbf{GL}-algebras can be embedded into polynomial 𝐆𝐋\mathbf{GL}-algebras, which is analogous to embedding modules into free modules. ∎

Remark 3.14.

We do not know if there are any examples of shift-free 𝐆𝐋\mathbf{GL}-algebras besides the ones appearing in Lemma 3.10. ∎

4. The main structural results for semi-linear representations

In this section, we prove many of the results stated in §1.2 in the ModK\operatorname{Mod}_{K} setting. These results essentially follow in a formal manner from the embedding theorem. In §4.1, we give an axiomatization of the formal arguments. In §4.2, we apply this axiomatization to prove the results on ModK\operatorname{Mod}_{K}, when KK is a rational 𝐆𝐋\mathbf{GL}-field. Finally, in §4.3, we prove some results for more general 𝐆𝐋\mathbf{GL}-fields.

4.1. Some category theory

Let 𝒜\mathcal{A} be a kk-linear Grothendieck abelian category, let {Iλ}λΛ\{I_{\lambda}\}_{\lambda\in\Lambda} be a set of non-zero objects in 𝒜\mathcal{A}, and let ||:Λ𝐙0|\cdot|\colon\Lambda\to\mathbf{Z}_{\geq 0} be a function. Suppose that the following conditions hold:

  • (A1)

    Every object of 𝒜\mathcal{A} is the union of its finitely generated subobjects.

  • (A2)

    For any nn, there are only finitely many λΛ\lambda\in\Lambda with |λ|n|\lambda|\leq n.

  • (A3)

    The object IλI_{\lambda} is finitely generated, for all λΛ\lambda\in\Lambda.

  • (A4)

    The space Hom(Iλ,Iμ)\operatorname{Hom}(I_{\lambda},I_{\mu}) is finite dimensional over kk for all λ,μΛ\lambda,\mu\in\Lambda.

  • (A5)

    The ring End(Iλ)\operatorname{End}(I_{\lambda}) is a division ring for all λΛ\lambda\in\Lambda.

  • (A6)

    We have Hom(Iλ,Iμ)0\operatorname{Hom}(I_{\lambda},I_{\mu})\neq 0 only if |μ|<|λ||\mu|<|\lambda| or λ=μ\lambda=\mu.

  • (A7)

    Let λΛ\lambda\in\Lambda and let JJ be a direct sum of objects of the form IμI_{\mu} with |μ|<|λ||\mu|<|\lambda|. Then there is no injection IλJI_{\lambda}\to J.

  • (A8)

    Every finitely generated object of 𝒜\mathcal{A} injects into a finite direct sum of the IλI_{\lambda}’s.

We introduce one more piece of notation: for λΛ\lambda\in\Lambda, we let LλL_{\lambda} be the intersection of the kernels of all maps IλIμI_{\lambda}\to I_{\mu} with |μ|<|λ||\mu|<|\lambda|. The object LλL_{\lambda} is non-zero by (A7).

Proposition 4.1.

In the above situation, we have the following:

  1. (a)

    Every object of 𝒜\mathcal{A} is locally of finite length.

  2. (b)

    The IλI_{\lambda}’s are exactly the indecomposable injectives of 𝒜\mathcal{A}.

  3. (c)

    Every finite length object of 𝒜\mathcal{A} has finite injective dimension.

  4. (d)

    The object LλL_{\lambda} is simple, and is equal to the socle of IλI_{\lambda}. Every simple object is isomorphic to a unique LλL_{\lambda}.

  5. (e)

    The simple object LλL_{\lambda} occurs in IλI_{\lambda} with multiplicity one; the remaining simple constituents of IλI_{\lambda} have the form LμL_{\mu} with |μ|<|λ||\mu|<|\lambda|.

We break the proof up into a series of lemmas. We assume that 𝒜\mathcal{A} satisfies (A1)–(A8) in the following.

Lemma 4.2.

Let II be a finitely generated object of 𝒜\mathcal{A}. Suppose that every injection IMI\to M, with MM finitely generated, splits. Then II is injective.

Proof.

Let MM be a finitely generated object of 𝒜\mathcal{A}, let NN be a subobject of MM, and let NIN\to I be a given morphism. Consider the map I(IM)/NI\to(I\oplus M)/N, where NN is embedded diagonally, which is easily seen to be injective. Since (IM)/N(I\oplus M)/N is finitely generated, this map splits by hypothesis. This yields a map MIM\to I extending the given map NIN\to I. A variant of Baer’s criterion (see [Stacks, Tag 079G]) now shows that II is injective. (The key point here is that 𝒜\mathcal{A} is generated by its finitely generated objects, due to (A1).) ∎

Lemma 4.3.

The object IλI_{\lambda} is an indecomposable injective, for all λΛ\lambda\in\Lambda.

Proof.

Since End(Iλ)\operatorname{End}(I_{\lambda}) has no non-trivial idempotents, it follows that IλI_{\lambda} is indecomposable. It is clear that IλI_{\lambda} is injective if |λ|<0|\lambda|<0, since this hypothesis is void. Assume now that IμI_{\mu} is injective for all μ\mu with |μ|<n|\mu|<n and let λ\lambda satisfy |λ|=n|\lambda|=n. Suppose we have an injection f:IλMf\colon I_{\lambda}\to M, with MM finitely generated. Choose an injection g:MIg\colon M\to I, where II is a finite direct sum of IμI_{\mu}’s, which is possible by (A8). Write I=I1I2I3I=I_{1}\oplus I_{2}\oplus I_{3}, where I1I_{1} is a sum of IμI_{\mu}’s with |μ|<n|\mu|<n, I2I_{2} is a sum of IλI_{\lambda}’s and I3I_{3} is a sum of IμI_{\mu}’s with |μ|n|\mu|\geq n and μλ\mu\neq\lambda. Let pip_{i} be the projection of II onto IiI_{i}. Then p3gf=0p_{3}gf=0 by (A6) and p1gfp_{1}gf is not injective by (A7). Since gfgf is injective, it follows that p2gfp_{2}gf is non-zero. Thus p2gp_{2}g, followed by a further projection, provides a map h:MIλh\colon M\to I_{\lambda} such that hfhf is non-zero. Since End(Iλ)\operatorname{End}(I_{\lambda}) is a division algebra by (A5), we can find hEnd(Iλ)h^{\prime}\in\operatorname{End}(I_{\lambda}) such that hhf=idh^{\prime}hf=\mathrm{id}, and so ff is split. It follows from Lemma 4.2 that IλI_{\lambda} is injective. The result now follows by induction. ∎

Lemma 4.4.

The object LλL_{\lambda} is simple, and is the socle of IλI_{\lambda}.

Proof.

Consider the natural map

f:IλJ,J=|μ|<nHom(Iλ,Iμ)Iμ.f\colon I_{\lambda}\to J,\qquad J=\bigoplus_{|\mu|<n}\operatorname{Hom}(I_{\lambda},I_{\mu})^{*}\otimes I_{\mu}.

This is the universal map from IλI_{\lambda} to a sum of IμI_{\mu}’s with |μ|<|λ||\mu|<|\lambda|. Thus Lλ=ker(f)L_{\lambda}=\ker(f).

Suppose NN is a non-zero subobject of LλL_{\lambda}. The object Iλ/NI_{\lambda}/N is finitely generated by (A3). Thus, by (A8), we have an injection Iλ/NII_{\lambda}/N\to I where II is a finite sum of IμI_{\mu}’s. Any map Iλ/NIμI_{\lambda}/N\to I_{\mu} with |μ||λ||\mu|\geq|\lambda| and μλ\mu\neq\lambda is automatically zero by (A6); similarly, any map Iλ/NIλI_{\lambda}/N\to I_{\lambda} is zero, since any non-zero map IλIλI_{\lambda}\to I_{\lambda} is injective by (A5). It follows that II can be taken to be a finite sum of IμI_{\mu}’s with |μ|<|λ||\mu|<|\lambda|. Let h:IλIh\colon I_{\lambda}\to I be the composition IλIλ/NII_{\lambda}\to I_{\lambda}/N\to I. By the universality of ff, we have h=gfh=gf for some g:JIg\colon J\to I, and so ker(f)ker(h)\ker(f)\subset\ker(h). Since ker(h)=N\ker(h)=N, this shows that N=LλN=L_{\lambda}, and so LλL_{\lambda} is simple.

Since IλI_{\lambda} is indecomposable, it follows that it is the injective envelope of LλL_{\lambda}. Since LλL_{\lambda} is simple, it is therefore the socle of IλI_{\lambda}. ∎

Lemma 4.5.

Every simple object of 𝒜\mathcal{A} is isomorphic to LλL_{\lambda}, for a unique λ\lambda.

Proof.

Let LL be a simple object of 𝒜\mathcal{A}. Then LL is necessarily finitely generated, and so by (A8) we have an injection LIL\to I, where II is a finite sum of IλI_{\lambda}’s. Since LL is simple, it follows that LL must inject into one of the factors, and land in the socle. This gives an isomorphism LLλL\cong L_{\lambda}.

Suppose now that LλLμL_{\lambda}\cong L_{\mu}. Then the injective envelopes of LλL_{\lambda} and LμL_{\mu} would be isomorphic, i.e., IλIμI_{\lambda}\cong I_{\mu}. By (A6), this implies that λ=μ\lambda=\mu. ∎

Lemma 4.6.

Every object of 𝒜\mathcal{A} is locally of finite length.

Proof.

By (A1), it suffices to show that every finitely generated object of 𝒜\mathcal{A} is finite length. By (A8), it suffices to show that each IλI_{\lambda} has finite length. We proceed by induction on |λ||\lambda|. Thus suppose IμI_{\mu} has finite length for |μ|<n|\mu|<n and let λ\lambda be given with |λ|=n|\lambda|=n. Using notation as in Lemma 4.4, we have an exact sequence

0LλIλ|μ|<nHom(Iλ,Iμ)Iμ.0\to L_{\lambda}\to I_{\lambda}\to\bigoplus_{|\mu|<n}\operatorname{Hom}(I_{\lambda},I_{\mu})^{*}\otimes I_{\mu}.

By Lemma 4.4, the object LλL_{\lambda} is simple. By induction, each IμI_{\mu} appearing in the sum on the right has finite length. By (A2), the sum is finite, and by (A4) each Hom\operatorname{Hom} space is finite dimensional. Thus the rightmost term above has finite length. It follows that IλI_{\lambda} has finite length, as required. ∎

Lemma 4.7.

Every indecomposable injective object of 𝒜\mathcal{A} is isomorphic to IλI_{\lambda} for a unique λ\lambda.

Proof.

Let II be an indecomposable injective. Since II is the union of its finite length subobjects b Lemma 4.6, it follows that the socle of II is simple, and that II is its injective envelope. Thus IIλI\cong I_{\lambda} for some λ\lambda. This λ\lambda is unique, as IλIμI_{\lambda}\cong I_{\mu} implies λ=μ\lambda=\mu by (A6). ∎

Lemma 4.8.

The simple LλL_{\lambda} occurs in IλI_{\lambda} with multiplicity one. The remaining simple constituents of IλI_{\lambda} have the form LμL_{\mu} with |μ|<|λ||\mu|<|\lambda|.

Proof.

We proceed by induction on |λ||\lambda|. Using notation as in Lemma 4.4, we have an exact sequence

0LλIλ|μ|<nHom(Iλ,Iμ)Iμ.0\to L_{\lambda}\to I_{\lambda}\to\bigoplus_{|\mu|<n}\operatorname{Hom}(I_{\lambda},I_{\mu})^{*}\otimes I_{\mu}.

The result now follows. ∎

Lemma 4.9.

Every finite length object of 𝒜\mathcal{A} has finite injective dimension.

Proof.

It suffices to prove that each LλL_{\lambda} has finite injective dimension. We proceed by induction on λ\lambda. Thus suppose that LμL_{\mu} has finite injective dimension for |μ|<|λ||\mu|<|\lambda|. By Lemma 4.8, it thus follows that Iλ/LλI_{\lambda}/L_{\lambda} has finite injective dimension, and so LλL_{\lambda} does as well. ∎

4.2. Applications to rational 𝐆𝐋\mathbf{GL}-fields

Fix a pure tuple σ¯\underline{\smash{\sigma}}, let R=Sym(kσ)R=\operatorname{Sym}(k^{\oplus\sigma}), and let K=Frac(R)K=\operatorname{Frac}(R). For a partition λ\lambda, we let LλL_{\lambda} be the intersection of the kernels of all maps KλKμK^{\oplus\lambda}\to K^{\oplus\mu} with |μ|<|λ||\mu|<|\lambda|. The following is our main result on the structure of KK-modules.

Theorem 4.10.

We have the following:

  1. (a)

    Every finitely generated KK-module has finite length.

  2. (b)

    The indecomposable injective KK-modules are exactly the KλK^{\oplus\lambda}, with λ\lambda a partition.

  3. (c)

    Every finite lengh KK-module has finite injective dimension.

  4. (d)

    The KK-module LλL_{\lambda} is simple, and is the socle of KλK^{\oplus\lambda}. Every simple KK-module is isomorphic to a unique LλL_{\lambda}.

  5. (e)

    The simple LλL_{\lambda} occurs in KλK^{\oplus\lambda} with multiplicity one; the remaining simple constituents have the form LμL_{\mu} with |μ|<|λ||\mu|<|\lambda|.

Proof.

We apply Proposition 4.1. We take 𝒜=ModK\mathcal{A}=\operatorname{Mod}_{K}, take Λ\Lambda to be the set of partitions, and take |λ||\lambda| to have its usual meaning (the size of λ\lambda). For λΛ\lambda\in\Lambda, we let Iλ=KλI_{\lambda}=K^{\oplus\lambda}. We verify the conditions (A1)–(A8). The first three conditions are clear.

Now, recall from §2.7 that we have a functor T:ModRModKT\colon\operatorname{Mod}_{R}\to\operatorname{Mod}_{K} given by MKRMM\mapsto K\otimes_{R}M, which has a right adjoint S:ModKModRS\colon\operatorname{Mod}_{K}\to\operatorname{Mod}_{R} given by S(N)=NpolS(N)=N^{\mathrm{pol}}. Moreover, S(Kλ¯)=Rλ¯S(K^{\oplus\underline{\smash{\lambda}}})=R^{\oplus\underline{\smash{\lambda}}} for any tuple λ¯\underline{\smash{\lambda}} (Proposition 2.7). In particular, we have

HomK(Kλ,Kμ)=HomR(Rλ,Rμ)=Hom𝐆𝐋(kλ,Rμ).\operatorname{Hom}_{K}(K^{\oplus\lambda},K^{\oplus\mu})=\operatorname{Hom}_{R}(R^{\oplus\lambda},R^{\oplus\mu})=\operatorname{Hom}_{\mathbf{GL}}(k^{\oplus\lambda},R^{\oplus\mu}).

This is finite dimensional over kk since kλk^{\oplus\lambda} occurs in RμR^{\oplus\mu} with finite multiplicity; this proves (A4). If λ=μ\lambda=\mu then we find that the above space is isomorphic to kk, which proves (A5). Finally, if |λ|<|μ||\lambda|<|\mu|, or if |λ|=|μ||\lambda|=|\mu| but λμ\lambda\neq\mu, then the above space is 0, which proves (A6).

We now handle (A7). Since IλI_{\lambda} is finitely generated, it suffices to consider the case where JJ is a finite direct sum in (A7). Thus, suppose by way of contraction that we have an injection KλKμ¯K^{\oplus\lambda}\to K^{\oplus\underline{\smash{\mu}}}, where μ¯\underline{\smash{\mu}} is a tuple composed of partitions that are strictly smaller than λ\lambda. Applying the SS functor, this gives an injection of RR-modules RλRμ¯R^{\oplus\lambda}\to R^{\oplus\underline{\smash{\mu}}}. Let nn be such that dim𝐒λ(kn)>dim𝐒μ¯(kn)\dim\mathbf{S}_{\lambda}(k^{n})>\dim\mathbf{S}_{\underline{\smash{\mu}}}(k^{n}). This is possible since dim𝐒λ(kn)\dim\mathbf{S}_{\lambda}(k^{n}) is a polynomial in nn of degree |λ||\lambda|, while dim𝐒μ¯(kn)\dim\mathbf{S}_{\underline{\smash{\mu}}}(k^{n}) is a polynomial of degree <|λ|<|\lambda|. Evaluating our injection on knk^{n}, we obtain an injection

R{kn}𝐒λ(kn)R{kn}𝐒μ¯(kn)R\{k^{n}\}\otimes\mathbf{S}_{\lambda}(k^{n})\to R\{k^{n}\}\otimes\mathbf{S}_{\underline{\smash{\mu}}}(k^{n})

of R{kn}R\{k^{n}\}-modules. This is impossible, as the two modules above are free of finite rank, and the domain has greater rank. We thus have a contradiction, which proves (A7).

Finally, (A8) is exactly Corollary 3.12. This completes the verification of (A1)–(A8). Thus Proposition 4.1 applies, which completes the proof. ∎

Corollary 4.11.

All projective RR-modules are injective.

Proof.

Let SS and TT be as in the above proof. Since TT is exact, the its right adjoint SS takes injectives to injectives. In particular, we see that S(KV)S(K\otimes V) is an injective RR-module for any polynomial representation VV. As S(KV)=RVS(K\otimes V)=R\otimes V (Proposition 2.7), and every projective RR-module has this form, the result follows. ∎

4.3. Applications to other 𝐆𝐋\mathbf{GL}-fields

By leveraging Theorem 4.10, we are able to deduce the following fundamental result for more general 𝐆𝐋\mathbf{GL}-fields:

Theorem 4.12.

Let KK be a 𝐆𝐋\mathbf{GL}-field that is finitely generated over its invariant subfield kk.

  1. (a)

    Any finitely generated KK-module has finite length.

  2. (b)

    If MM and NN are finitely generated KK-modules then HomK(M,N)\operatorname{Hom}_{K}(M,N) is a finite dimensional kk-vector space.

The first statement is reasonably straightforward:

Proof of Theorem 4.12(a).

Applying Proposition 2.5, let nn be such that Shn(K)\operatorname{Sh}_{n}(K) is a rational 𝐆𝐋\mathbf{GL}-field over its invariant subfield. Let VV be a finitely generated KK-module. Then Shn(V)\operatorname{Sh}_{n}(V) is a finitely generated Shn(K)\operatorname{Sh}_{n}(K)-module, and therefore of finite length by Theorem 4.10(a). It follows that VV has finite length. In fact, if Shn(V)\operatorname{Sh}_{n}(V) has length \ell then VV has length \leq\ell, for if U0U+1U_{0}\subset\cdots\subset U_{\ell+1} is any chain of KK-submodules of VV then Shn(U0)Shn(U+1)\operatorname{Sh}_{n}(U_{0})\subset\cdots\subset\operatorname{Sh}_{n}(U_{\ell+1}) is a chain of Shn(K)\operatorname{Sh}_{n}(K)-submodules of Shn(V)\operatorname{Sh}_{n}(V), and so Shn(Ui)=Shn(Ui+1)\operatorname{Sh}_{n}(U_{i})=\operatorname{Sh}_{n}(U_{i+1}) for some ii, and so Ui=Ui+1U_{i}=U_{i+1}. ∎

The second part of the theorem will take the remainder of the section. We require a number of lemmas.

Lemma 4.13.

Theorem 4.12(b) holds if KK is a rational 𝐆𝐋\mathbf{GL}-field over kk.

Proof.

Choose a surjection Kλ¯MK^{\oplus\underline{\smash{\lambda}}}\to M for some tuple λ¯\underline{\smash{\lambda}}, which is possible in general, and an injection NKμ¯N\to K^{\oplus\underline{\smash{\mu}}} for some tuple μ¯\underline{\smash{\mu}}, which is possible by the embedding theorem (Corollary 3.12) since KK is rational. We thus obtain an injection

HomK(M,N)HomK(Kλ¯,Kμ¯).\operatorname{Hom}_{K}(M,N)\to\operatorname{Hom}_{K}(K^{\oplus\underline{\smash{\lambda}}},K^{\oplus\underline{\smash{\mu}}}).

We have seen (in the proof of Theorem 4.10) that this is finite dimensional over kk. The result follows. ∎

Lemma 4.14.

Let MM be a finitely generated KK-module and let φ\varphi be an endomorphism of MM. Then φ\varphi satisfies a non-zero polynomial with coefficients in KK.

Proof.

Applying Proposition 2.5, let nn be such that Shn(K)\operatorname{Sh}_{n}(K) is a rational 𝐆𝐋\mathbf{GL}-field over its invariant subfield; in other words, this means KK is rational over KG(n)K^{G(n)} as a G(n)G(n)-field. Let EE be the space of all KK-linear G(n)G(n)-equivariant maps MMM\to M; this is identified with EndShn(K)(Shn(M))\operatorname{End}_{\operatorname{Sh}_{n}(K)}(\operatorname{Sh}_{n}(M)). By Lemma 4.13, EE is a finite dimensional vector space over the field KG(n)K^{G(n)}. Thus the elements {φi}i0\{\varphi^{i}\}_{i\geq 0} of EE are linearly dependent, which gives the requisite polynomial. ∎

For φ\varphi as above, the set of all polynomials that φ\varphi satisfies forms an ideal in the univariate polynomial ring K[T]K[T]. We define the minimal polynomial of φ\varphi to be the unique monic generator of this ideal. In other words, the minimal polynomial of φ\varphi is the unique monic polynomial that φ\varphi satisfies of minimal degree.

Lemma 4.15.

Let MM be a finitely generated KK-module and let φ\varphi be an endomorphism of MM. Then the minimal polyomial of φ\varphi has coefficients in the invariant field kk.

Proof.

Suppose that i=0dciφi=0\sum_{i=0}^{d}c_{i}\varphi^{i}=0 is the equation given by the minimal polynomial. If g𝐆𝐋g\in\mathbf{GL} then we also have i=0d(gci)φi=0\sum_{i=0}^{d}(gc_{i})\varphi^{i}=0. By uniquness of the minimal polynomial, we therefore have gci=cigc_{i}=c_{i}. Since this holds for all gg, it follows that cikc_{i}\in k, as required. ∎

The following lemma is a version of Schur’s lemma:

Lemma 4.16.

Suppose that kk is algebraically closed and MM is a simple KK-module. Then EndK(M)=k\operatorname{End}_{K}(M)=k.

Proof.

Since MM is simple, it follows that D=EndK(M)D=\operatorname{End}_{K}(M) is a division ring. We know that DD contains kk in its center. By the Lemma 4.15, every element of DD is algebraic over kk. (Note that MM is necessarily finitely generated since it is simple.) Since kk is algebraically closed, it follows that D=kD=k. ∎

Lemma 4.17.

Suppose that kk is algebraically closed. Then Theorem 4.12(b) holds.

Proof.

It follows from the previous lemma that HomK(M,N)\operatorname{Hom}_{K}(M,N) is finite dimensional over kk if MM and NN are simple. As MM and NN have finite length by Theorem 4.12(a), the general case follows from dévissage. ∎

We now deduce the general case from the case with kk algebraically closed using a base change argument. For this, we require two more lemmas.

Lemma 4.18.

Any element of KK that is algebraic over kk belongs to kk, i.e., kk is algebraically closed within KK.

Proof.

Let aKa\in K be algebraic over kk, and let f(T)k[T]f(T)\in k[T] be its minimal polynomial. Since 𝐆𝐋\mathbf{GL} acts on KK by field homomorphisms, it permutes the roots of ff in KK. This action corresponds to a homomorphism φ:𝐆𝐋𝔖n\varphi\colon\mathbf{GL}\to\mathfrak{S}_{n} where nn is the number of roots of ff in KK. Since any group homomorphism 𝐐𝔖n\mathbf{Q}\to\mathfrak{S}_{n} is trivial, it follows that φ\varphi is trivial on each group of elementary matrices in 𝐆𝐋\mathbf{GL}. Since these groups generate 𝐒𝐋\mathbf{SL}, it follows that φ(𝐒𝐋)=1\varphi(\mathbf{SL})=1. We thus see that aa is fixed by 𝐒𝐋\mathbf{SL}. However, aa is also fixed by G(m)G(m) for m0m\gg 0. It follows that aa is fixed by 𝐆𝐋=𝐒𝐋G(m)\mathbf{GL}=\mathbf{SL}\cdot G(m), i.e., aka\in k. ∎

Suppose kk^{\prime} is an algebraic extension of kk. Then the above lemma implies that K=kkKK^{\prime}=k^{\prime}\otimes_{k}K is a field. Letting 𝐆𝐋\mathbf{GL} act on KK^{\prime} by acting trivially on kk^{\prime}, one easily sees that KK^{\prime} is a 𝐆𝐋\mathbf{GL}-field, its invariant field is kk^{\prime}, and it is finitely 𝐆𝐋\mathbf{GL}-generated over kk^{\prime}.

Lemma 4.19.

Let kk^{\prime} be an algebraic extension of kk and put K=kkKK^{\prime}=k^{\prime}\otimes_{k}K. Let MM and NN be KK-modules, with MM finitely generated. Then the natural map

kkHomK(M,N)HomK(kkM,kkN)k^{\prime}\otimes_{k}\operatorname{Hom}_{K}(M,N)\to\operatorname{Hom}_{K^{\prime}}(k^{\prime}\otimes_{k}M,k^{\prime}\otimes_{k}N)

is an isomorphism.

Proof.

By adjunction, we have

HomK(kkM,kkN)=HomK(M,kkN).\operatorname{Hom}_{K^{\prime}}(k^{\prime}\otimes_{k}M,k^{\prime}\otimes_{k}N)=\operatorname{Hom}_{K}(M,k^{\prime}\otimes_{k}N).

Now, for any kk-vector space EE, we have a natural map

EkHomK(M,N)HomK(M,EkN).E\otimes_{k}\operatorname{Hom}_{K}(M,N)\to\operatorname{Hom}_{K}(M,E\otimes_{k}N).

Picking a kk-basis {ei}iI\{e_{i}\}_{i\in I} for EE, we find that the above map is isomorphic to the map

HomK(M,N)IHomK(M,NI).\operatorname{Hom}_{K}(M,N)^{\oplus I}\to\operatorname{Hom}_{K}(M,N^{\oplus I}).

This map is an isomorphism since MM is finitely generated. Applying this with E=kE=k^{\prime} gives the result. ∎

Proof of Theorem 4.12(b).

Let kk^{\prime} be an algebraic closure of kk, and let K=kkKK^{\prime}=k^{\prime}\otimes_{k}K. Let MM and NN be finitely generated KK-modules. By Lemma 4.19, the map

kkHomK(M,N)HomK(kkM,kkN)k^{\prime}\otimes_{k}\operatorname{Hom}_{K}(M,N)\to\operatorname{Hom}_{K^{\prime}}(k^{\prime}\otimes_{k}M,k^{\prime}\otimes_{k}N)

is an isomorphism. As kkMk^{\prime}\otimes_{k}M and kkNk^{\prime}\otimes_{k}N are finitely generated KK^{\prime}-modules, it follows from Lemma 4.17 that the right side above is a finite dimensional kk^{\prime}-vector space. It thus follows that HomK(M,N)\operatorname{Hom}_{K}(M,N) is a finite dimensional kk-vector space, which completes the proof. ∎

5. Brauer categories, Weyl’s construction, universal properties

The purpose of this section is to describe ModK\operatorname{Mod}_{K}, when KK is a rational 𝐆𝐋\mathbf{GL}-field, in terms of a combinatorial category, the upwards σ¯\underline{\smash{\sigma}}-Brauer category 𝔘(σ¯)\mathfrak{U}(\underline{\smash{\sigma}}). We begin in §5.1 by reviewing generalities on representations of categories. We introduce 𝔘(σ¯)\mathfrak{U}(\underline{\smash{\sigma}}) in §5.2. The main equivalences are established in §5.3. Finally, in §5.4 and §5.5, we give applications of these equivalences: we establish a version of Weyl’s traceless tensor construction for ModK\operatorname{Mod}_{K}, and give a universal property for ModK\operatorname{Mod}_{K}.

5.1. Representations of categories

We now review a bit of material on representations of categories. See [SS7, §3] for more detail.

Let 𝔊\mathfrak{G} be an essentially small kk-linear category. A representation of 𝔊\mathfrak{G}, or a 𝔊\mathfrak{G}-module, is a functor 𝔊Veck\mathfrak{G}\to\mathrm{Vec}_{k}, and a map of 𝔊\mathfrak{G}-modules is a natural transformation. We let Mod𝔊\operatorname{Mod}_{\mathfrak{G}} be the category of 𝔊\mathfrak{G}-modules. For 𝔊\mathfrak{G}-modules MM and NN, we write Hom𝔊(M,N)\operatorname{Hom}_{\mathfrak{G}}(M,N) for the set of maps of 𝔊\mathfrak{G}-modules MNM\to N.

Let xx be an object of 𝔊\mathfrak{G}. We define the principal projective 𝔊\mathfrak{G}-module at xx, denoted 𝐏x\mathbf{P}_{x}, by 𝐏x(y)=Hom𝔊(x,y)\mathbf{P}_{x}(y)=\operatorname{Hom}_{\mathfrak{G}}(x,y). If MM is an arbitary 𝔊\mathfrak{G}-module then we have an identification

Hom𝔊(𝐏x,M)=M(x)\operatorname{Hom}_{\mathfrak{G}}(\mathbf{P}_{x},M)=M(x)

by Yoneda’s lemma, which shows that 𝐏x\mathbf{P}_{x} is projective. The above identity also shows that MM can be realized as a quotient of a direct sum of principal projectives.

We similarly define the principal injective 𝔊\mathfrak{G}-module at xx, denoted 𝐈x\mathbf{I}_{x}, by 𝐈x(y)=Hom𝔊(y,x)\mathbf{I}_{x}(y)=\operatorname{Hom}_{\mathfrak{G}}(y,x)^{*}. If MM is an arbitary 𝔊\mathfrak{G}-module then we have an identification

Hom𝔊(M,𝐈x)=M(x)\operatorname{Hom}_{\mathfrak{G}}(M,\mathbf{I}_{x})=M(x)^{*}

(see [SS7, Proposition 3.2]), which shows that 𝐈x\mathbf{I}_{x} is injective.

Proposition 5.1.

Suppose that the Hom\operatorname{Hom} sets in 𝔊\mathfrak{G} are finite dimensional. Then the following categories are equivalent:

  1. (a)

    The category 𝔊op\mathfrak{G}^{\mathrm{op}}.

  2. (b)

    The full subcategory of Mod𝔊\operatorname{Mod}_{\mathfrak{G}} spanned by the principal projectives.

  3. (c)

    The full subcategory of Mod𝔊\operatorname{Mod}_{\mathfrak{G}} spanned by the principal injectives.

Proof.

Let 𝒞\mathcal{C} be the category in (b). We have a functor 𝔊op𝒞\mathfrak{G}^{\mathrm{op}}\to\mathcal{C} given by x𝐏xx\mapsto\mathbf{P}_{x}. It is obviously essentially surjective, and is fully faithful by Yoneda’s lemma. Similarly, let 𝒞\mathcal{C}^{\prime} be the category in (c). Then we have a functor 𝔊op𝒞\mathfrak{G}^{\mathrm{op}}\to\mathcal{C}^{\prime} given by x𝐈xx\mapsto\mathbf{I}_{x}. We have

Hom𝔊(𝐈x,𝐈y)=𝐈x(y)=Hom𝔊(y,x)=Hom𝔊(y,x)=Hom𝔊op(x,y).\operatorname{Hom}_{\mathfrak{G}}(\mathbf{I}_{x},\mathbf{I}_{y})=\mathbf{I}_{x}(y)^{*}=\operatorname{Hom}_{\mathfrak{G}}(y,x)^{**}=\operatorname{Hom}_{\mathfrak{G}}(y,x)=\operatorname{Hom}_{\mathfrak{G}^{\mathrm{op}}}(x,y).

One easily sees that this identification is induced by the functor under consideration, which shows that it is fully faithful. ∎

Suppose now that the isomorphism classes of 𝔊\mathfrak{G} are in bijection with the set 𝐍\mathbf{N} of natural numbers; for n𝐍n\in\mathbf{N}, we let [n][n] be a representative of the nnth isomorphism class. We say that 𝔊\mathfrak{G} is upwards if Hom𝔊([n],[m])0\operatorname{Hom}_{\mathfrak{G}}([n],[m])\neq 0 implies nmn\leq m.

Proposition 5.2.

Suppose 𝔊\mathfrak{G} is upwards and all Hom\operatorname{Hom} sets are finite dimensional. Then the principal injectives are of finite length, and every finite length 𝔊\mathfrak{G}-module embeds into a finite sum of principal injectives.

Proof.

Let MM be a 𝔊\mathfrak{G}-module, and write MnM_{n} in place of M([n])M([n]). Define the support of MM to be the set of natural numbers nn for which Mn0M_{n}\neq 0. Define the nnth truncation of MM, denoted τn(M)\tau_{\geq n}(M), to be the 𝔊\mathfrak{G}-module given by

τn(M)m={Mmif mn0if m<n;\tau_{\geq n}(M)_{m}=\begin{cases}M_{m}&\text{if $m\geq n$}\\ 0&\text{if $m<n$}\end{cases};

one easily sees that this is a 𝔊\mathfrak{G}-submodule of MM since 𝔊\mathfrak{G} is upwards. From the above structure, one easily verifies the following two statements:

  1. (a)

    A 𝔊\mathfrak{G}-module MM is simple if and only if it is supported in a single degree nn and MnM_{n} is a simple module over the ring End𝔊([n])\operatorname{End}_{\mathfrak{G}}([n]).

  2. (b)

    A 𝔊\mathfrak{G}-module MM has finite length if and only if it has finite support and MnM_{n} is finite dimensional for all nn.

It follows from (b) that the principal injective 𝐈n\mathbf{I}_{n} is of finite length. It follows from (a) that if MM is a simple supported in degree nn then MM embeds into 𝐈n\mathbf{I}_{n}. One now easily sees that any finite length objects embeds into a sum of 𝐈n\mathbf{I}_{n}’s. ∎

5.2. A variant of the Brauer category

The upwards and downwards Brauer categories were introduced in [SS3, §4.2.5] as a means to describe the category of algebraic representations of the infinite orthogonal group. We now introduce a generalization that will similarly allow us to describe the category of KK-modules.

For a partition λ\lambda of nn, recall that SλS^{\lambda} is the irreducible representation of 𝔖n\mathfrak{S}_{n} associated to λ\lambda (the Specht module). For a finite set AA of cardinality nn, we let SAλS^{\lambda}_{A} be the associated representation of Aut(A)𝔖n\operatorname{Aut}(A)\cong\mathfrak{S}_{n}. One can define this in a canonical manner by mimicking the construction of SλS^{\lambda}, but using elements of AA in place of the integers 1,,n1,\ldots,n.

Fix a pure tuple σ¯=[σ1,,σr]\underline{\smash{\sigma}}=[\sigma_{1},\ldots,\sigma_{r}]. A σ¯\underline{\smash{\sigma}}-block on a set SS is a triple (A,p,x)(A,p,x) where

  • pp is an element of [r][r],

  • AA is a subset of SS of cardinality σp\sigma_{p} (called the support of the block),

  • xx is an element of the Specht module SAσpS^{\sigma_{p}}_{A}.

Let SS and TT be a finite sets. A downwards σ¯\underline{\smash{\sigma}}-diagram from SS to TT is a pair (i,Γ)(i,\Gamma) where Γ\Gamma is a collection of σ¯\underline{\smash{\sigma}}-blocks on SS with disjoint supports and i:S|Γ|Ti\colon S\setminus|\Gamma|\to T is a bijection, where |Γ||\Gamma| is the union of the supports of the blocks in Γ\Gamma. The space of downwards σ¯\underline{\smash{\sigma}}-diagrams is the vector space spanned by elements [i,Γ][i,\Gamma], with (i,Γ)(i,\Gamma) an downwards σ¯\underline{\smash{\sigma}}-diagram, with the following relation imposed:

  • Suppose that Γ\Gamma contains a block (A,p,x)(A,p,x), and let x=αy+βzx=\alpha y+\beta z be a linear combination in the Specht module. Let Γ\Gamma^{\prime} be the diagram obtained by replacing this block with (A,p,y)(A,p,y), and let Γ\Gamma^{\prime} be defined similarly but using zz. Then [i,Γ]=α[i,Γ]+β[i,Γ′′][i,\Gamma]=\alpha[i,\Gamma^{\prime}]+\beta[i,\Gamma^{\prime\prime}].

We now come to the main definition:

Definition 5.3.

The downwards σ¯\underline{\smash{\sigma}}-Brauer category, denoted 𝔇(σ¯)\mathfrak{D}(\underline{\smash{\sigma}}), is the kk-linear category described as follows.

  • The objects of 𝔇(σ¯)\mathfrak{D}(\underline{\smash{\sigma}}) are finite sets.

  • Given finite sets SS and TT, the space of morphisms Hom𝔇(σ¯)(S,T)\operatorname{Hom}_{\mathfrak{D}(\underline{\smash{\sigma}})}(S,T) is the space of downwards σ¯\underline{\smash{\sigma}}-diagrams from SS to TT.

  • Composition is defined as follows. Let (i,Γ)(i,\Gamma) be a diagram from SS to TT, and let (i,Γ)(i^{\prime},\Gamma^{\prime}) be a diagram from TT to UU. Let j=iij=i^{\prime}\circ i and let Δ=Γi1(Γ)\Delta=\Gamma\sqcup i^{-1}(\Gamma^{\prime}), where i1(Γ)i^{-1}(\Gamma^{\prime}) denotes the result of transporting Γ\Gamma^{\prime} along the bijection i1:|Γ|i1(|Γ|)i^{-1}\colon|\Gamma^{\prime}|\to i^{-1}(|\Gamma^{\prime}|). Then [i,Γ][i,Γ]=[j,Δ][i^{\prime},\Gamma^{\prime}]\circ[i,\Gamma]=[j,\Delta]. ∎

Example 5.4.

If σ¯=[(2)]\underline{\smash{\sigma}}=[(2)] then 𝔇(σ¯)\mathfrak{D}(\underline{\smash{\sigma}}) is the downwards Brauer category from [SS3, §4.2.5]. Similarly, if σ=[(1,1)]\sigma=[(1,1)] then 𝔇(σ¯)\mathfrak{D}(\underline{\smash{\sigma}}) is the signed downwards Brauer category, also discussed in [SS3, §4.2.11]. ∎

Example 5.5.

Suppose that σ¯=[σ1,,σd]\underline{\smash{\sigma}}=[\sigma_{1},\ldots,\sigma_{d}] where σi=(1)\sigma_{i}=(1) for all ii. Then a downwards σ¯\underline{\smash{\sigma}}-diagram from SS to TT is an injection f:TSf\colon T\to S together with a dd-coloring on Sim(f)S\setminus\operatorname{im}(f). We thus see that 𝔇(σ¯)\mathfrak{D}(\underline{\smash{\sigma}}) is the opposite of the category 𝐅𝐈d\mathbf{FI}_{d} introducing in [SS4, §7]. (The category 𝐅𝐈1\mathbf{FI}_{1} is just the category 𝐅𝐈\mathbf{FI} of finite sets and injections, as in [CEF].) ∎

The category 𝔇(σ¯)\mathfrak{D}(\underline{\smash{\sigma}}) carries a natural symmetric monoidal structure \amalg given by disjoint union. Precisely, for two objects SS and TT, the object STS\amalg T is simply the disjoint union of the sets SS and TT. Given two morphisms [i,Γ]:ST[i,\Gamma]\colon S\to T and [i,Γ]:ST[i^{\prime},\Gamma^{\prime}]\colon S^{\prime}\to T^{\prime}, the morphism [i,Γ][i,Γ][i,\Gamma]\amalg[i^{\prime},\Gamma^{\prime}] is defined to be [ii,ΓΓ][i\amalg i^{\prime},\Gamma\amalg\Gamma^{\prime}]. Note that \amalg is a kk-linear functor in each of its arguments.

The category 𝔇(σ¯)\mathfrak{D}(\underline{\smash{\sigma}}) admits a universal property, which we now describe. Let 𝒞\mathcal{C} be an kk-linear symmetric monoidal category. Let Tσ¯(𝒞)T_{\underline{\smash{\sigma}}}(\mathcal{C}) be the category whose objects are pairs (X,ω)(X,\omega), where XX is an object of 𝒞\mathcal{C} and ω:𝐒σ¯(X)𝟏\omega\colon\mathbf{S}_{\underline{\smash{\sigma}}}(X)\to\mathbf{1} is a morphism in the Karoubian–additive envelope of 𝒞\mathcal{C}, where 𝟏\mathbf{1} is the unit object of 𝒞\mathcal{C}; of course, if 𝒞\mathcal{C} is additive and Karoubian (e.g., abelian) then one does not need to take the envelope here. Morphisms in Tσ¯(𝒞)T_{\underline{\smash{\sigma}}}(\mathcal{C}) are defined in the obvious manner.

Proposition 5.6.

Notation as above, we have a natural equivalence of categories

Φ:Funk(𝔇(σ¯),𝒞)Tσ¯(𝒞).\Phi\colon\operatorname{Fun}^{\otimes}_{k}(\mathfrak{D}(\underline{\smash{\sigma}}),\mathcal{C})\to T_{\underline{\smash{\sigma}}}(\mathcal{C}).

Here Funk(,)\operatorname{Fun}^{\otimes}_{k}(-,-) denotes the category of symmetric monoidal kk-linear functors.

Proof.

We first define the functor Φ\Phi. Thus suppose given a symmetric monoidal kk-linear functor θ:𝔇(σ¯)𝒞\theta\colon\mathfrak{D}(\underline{\smash{\sigma}})\to\mathcal{C}. Let X=θ([1])X=\theta([1]). We define a map ω:𝐒σ¯(X)𝟏\omega\colon\mathbf{S}_{\underline{\smash{\sigma}}}(X)\to\mathbf{1}. It suffices to define maps ωp:𝐒σp(X)𝟏\omega_{p}\colon\mathbf{S}_{\sigma_{p}}(X)\to\mathbf{1} for each p[r]p\in[r]. Thus fix such pp. Put np=|σp|n_{p}=|\sigma_{p}|. Since θ\theta is a symmetric monoidal functor, it induces an 𝔖np\mathfrak{S}_{n_{p}}-equivariant map

Hom𝔇(σ¯)([1]np,[0])Hom𝒞(Xnp,𝟏).\operatorname{Hom}_{\mathfrak{D}(\underline{\smash{\sigma}})}([1]^{\otimes n_{p}},[0])\to\operatorname{Hom}_{\mathcal{C}}(X^{\otimes n_{p}},\mathbf{1}).

Now, [1]np=[np][1]^{\otimes n_{p}}=[n_{p}]. Inside of Hom𝔇(σ¯)([np],[0])\operatorname{Hom}_{\mathfrak{D}(\underline{\smash{\sigma}})}([n_{p}],[0]) one has the subspace spanned by diagrams that consist of a single block of type pp. This subspace is isomorphic to SσpS^{\sigma_{p}} as a representation of 𝔖np\mathfrak{S}_{n_{p}}. We thus obtain a canonical 𝔖np\mathfrak{S}_{n_{p}}-equivariant map SσpHom𝒞(Xnp,𝟏)S^{\sigma_{p}}\to\operatorname{Hom}_{\mathcal{C}}(X^{\otimes n_{p}},\mathbf{1}), which yields a map 𝐒σp(X)𝟏\mathbf{S}_{\sigma_{p}}(X)\to\mathbf{1}, as required. We have thus defined ω\omega. We define Φ\Phi on objects by Φ(θ)=(X,ω)\Phi(\theta)=(X,\omega). The definition on morphisms is clear.

To show that Φ\Phi is an equivalence, we construct a quasi-inverse functor

Ψ:Tσ¯(𝒞)Funk(𝔇(σ¯),𝒞).\Psi\colon T_{\underline{\smash{\sigma}}}(\mathcal{C})\to\operatorname{Fun}^{\otimes}_{k}(\mathfrak{D}(\underline{\smash{\sigma}}),\mathcal{C}).

Thus let (X,ω)(X,\omega) in Tσ¯(𝒞)T_{\underline{\smash{\sigma}}}(\mathcal{C}) be given. We define a symmetric monoidal kk-linear functor θ:𝔇(σ¯)𝒞\theta\colon\mathfrak{D}(\underline{\smash{\sigma}})\to\mathcal{C}. On objects, we define θ\theta by θ(S)=XS\theta(S)=X^{\otimes S}. Now, consider a σ¯\underline{\smash{\sigma}}-block (A,p,x)(A,p,x). We have

Hom𝒞(XA,𝟏)=μnpSAμHom𝒞(𝐒μ(X),𝟏).\operatorname{Hom}_{\mathcal{C}}(X^{\otimes A},\mathbf{1})=\bigoplus_{\mu\vdash n_{p}}S^{\mu}_{A}\otimes\operatorname{Hom}_{\mathcal{C}}(\mathbf{S}_{\mu}(X),\mathbf{1}).

The μ=σp\mu=\sigma_{p} summand on the right side contains the element xωpx\otimes\omega_{p}. We say that the corresponding morphism XA𝟏X^{\otimes A}\to\mathbf{1} is associated to this block. Note that this construction is linear in the element xx. Now, consider a morphism f:STf\colon S\to T in 𝔇(σ¯)\mathfrak{D}(\underline{\smash{\sigma}}) represented by a diagram. Suppose this diagram corresponds to a pair (i,Γ)(i,\Gamma), where Γ\Gamma is a collection of disjoint blocks on SS and i:S|Γ|Ti\colon S\setminus|\Gamma|\to T is a bijection. We define a morphism θ(f):XSXT\theta(f)\colon X^{\otimes S}\to X^{\otimes T} as follows. Write XS=X|Γ|XS|Γ|X^{\otimes S}=X^{\otimes|\Gamma|}\otimes X^{\otimes S\setminus|\Gamma|}. We have a map X|Γ|𝟏X^{\otimes|\Gamma|}\to\mathbf{1} by tensoring together the maps associated to individual blocks. We also have a map XS|Γ|XTX^{\otimes S\setminus|\Gamma|}\to X^{\otimes T} from the bijection ii. The map θ(f)\theta(f) is the tensor product of these two maps. The construction θ\theta extends to a kk-linear map

θ:Hom𝔇(σ¯)(S,T)Hom𝒞(XS,XT).\theta\colon\operatorname{Hom}_{\mathfrak{D}(\underline{\smash{\sigma}})}(S,T)\to\operatorname{Hom}_{\mathcal{C}}(X^{\otimes S},X^{\otimes T}).

One easily verifies that θ\theta is compatible with composition and is naturally a symmetric monoidal functor. We define Ψ\Psi on objects by Ψ(X,ω)=θ\Psi(X,\omega)=\theta. The definition of Ψ\Psi on morphisms is clear.

One easily verifies that Φ\Phi and Ψ\Psi are naturally quasi-inverse. This completes the proof. ∎

There is also an upwards σ¯\underline{\smash{\sigma}}-Brauer category 𝔘(σ¯)\mathfrak{U}(\underline{\smash{\sigma}}), defined in the same manner, but where now blocks are only allowed on the target of a morphism. In other words, 𝔘(σ¯)\mathfrak{U}(\underline{\smash{\sigma}}) is simply the opposite category of 𝔇(σ¯)\mathfrak{D}(\underline{\smash{\sigma}}). The category 𝔘(σ¯)\mathfrak{U}(\underline{\smash{\sigma}}) admits a natural symmetric monoidal structure, and has a similar universal property to 𝔇(σ¯)\mathfrak{D}(\underline{\smash{\sigma}}).

5.3. Equivalences

We now establish a number of equivalences between categories associated to RR, KK, and the σ¯\underline{\smash{\sigma}}-Brauer categories..

Proposition 5.7.

The following symmetric monoidal kk-linear categories are equivalent:

  1. (a)

    The downwards σ¯\underline{\smash{\sigma}}-Brauer category 𝔇(σ¯)\mathfrak{D}(\underline{\smash{\sigma}}).

  2. (b)

    The full subcategory of ModR\operatorname{Mod}_{R} spanned by the objects R𝐕nR\otimes\mathbf{V}^{\otimes n} for n0n\geq 0.

  3. (c)

    The full subcategory of ModK\operatorname{Mod}_{K} spanned by the objects K𝐕nK\otimes\mathbf{V}^{\otimes n} for n0n\geq 0.

As kk-linear categories (ignoring the monoidal structure), these categories are also equivalent to

  1. (d)

    The full subcategory of Mod𝔘(σ¯)\operatorname{Mod}_{\mathfrak{U}(\underline{\smash{\sigma}})} spanned by the principal projective objects.

  2. (e)

    The full subcategory of Mod𝔘(σ¯)\operatorname{Mod}_{\mathfrak{U}(\underline{\smash{\sigma}})} spanned by the principal injective objects.

Proof.

We break the proof into three steps.

Step 1: equivalence of (a) and (b). Let 𝒞\mathcal{C} be the category in (b) and let X=R𝐕X=R\otimes\mathbf{V}. Then 𝐒σ¯(X)=R𝐒σ¯(𝐕)\mathbf{S}_{\underline{\smash{\sigma}}}(X)=R\otimes\mathbf{S}_{\underline{\smash{\sigma}}}(\mathbf{V}), where on the left side 𝐒σ¯\mathbf{S}_{\underline{\smash{\sigma}}} is formed with respect to R\otimes_{R}. Since RR contains 𝐒σ¯(𝐕)\mathbf{S}_{\underline{\smash{\sigma}}}(\mathbf{V}) as a subrepresentation, there is a natural map of RR-modules R𝐒σ¯(𝐕)RR\otimes\mathbf{S}_{\underline{\smash{\sigma}}}(\mathbf{V})\to R. We thus have a natural map ω:𝐒σ¯(X)R\omega\colon\mathbf{S}_{\underline{\smash{\sigma}}}(X)\to R. Since RR is the unit object for R\otimes_{R}, the universal property of 𝔇(σ¯)\mathfrak{D}(\underline{\smash{\sigma}}) (Proposition 5.6) furnishes a symmetric monoidal kk-linear functor

θ:𝔇(σ¯)𝒞.\theta\colon\mathfrak{D}(\underline{\smash{\sigma}})\to\mathcal{C}.

This functor has the property that θ([n])=Xn=R𝐕n\theta([n])=X^{\otimes n}=R\otimes\mathbf{V}^{\otimes n}. It is clear that θ\theta is essentially surjective. To complete this step, it suffices to show that θ\theta is fully faithful.

Before doing this, we introduce some notation. Identify the weight lattice of the diagonal torus in 𝐆𝐋\mathbf{GL} with 𝐙\mathbf{Z}^{\oplus\infty}. For a finite subset A[]A\subset[\infty], let 1A1^{A} denote the weight that is 1 in the AA coordinates and 0 elsewhere; also, write 1n1^{n} in place of 1[n]1^{[n]}. Given a weight λ\lambda and a polynomial representation VV, let VλV_{\lambda} be the λ\lambda weight space of VV.

Now, we have

HomR(Xn,Xm)=HomR(R𝐕n,R𝐕m)=Hom𝐆𝐋(𝐕n,R𝐕m).\operatorname{Hom}_{R}(X^{\otimes n},X^{\otimes m})=\operatorname{Hom}_{R}(R\otimes\mathbf{V}^{\otimes n},R\otimes\mathbf{V}^{\otimes m})=\operatorname{Hom}_{\mathbf{GL}}(\mathbf{V}^{\otimes n},R\otimes\mathbf{V}^{\otimes m}).

By Schur–Weyl duality, Hom𝐆𝐋(𝐕n,W)=W1n\operatorname{Hom}_{\mathbf{GL}}(\mathbf{V}^{\otimes n},W)=W_{1^{n}} for any polynomial representation WW; explicitly, a map φ:𝐕nW\varphi\colon\mathbf{V}^{\otimes n}\to W corresponds to φ(e1en)W1n\varphi(e_{1}\otimes\cdots\otimes e_{n})\in W_{1^{n}}. We must therefore understand the 1n1^{n} weight space of R𝐕mR\otimes\mathbf{V}^{\otimes m}.

Let p[r]p\in[r]. The 1n1^{n}-weight space of 𝐒σp(𝐕)\mathbf{S}_{\sigma_{p}}(\mathbf{V}) is canonically isomorphic to the Specht module SσpS^{\sigma_{p}} if n=#σpn=\#\sigma_{p}, and vanishes for other values of nn. More generally, let AA be a subset of [][\infty] of size #σp\#\sigma_{p}. Then we have a canonical isomorphism (𝐒σp(𝐕))1A=SAσp(\mathbf{S}_{\sigma_{p}}(\mathbf{V}))_{1^{A}}=S^{\sigma_{p}}_{A}. Fix a basis 𝒮Aσp\mathcal{S}^{\sigma_{p}}_{A} for SAσpS^{\sigma_{p}}_{A}. For x𝒮Aσpx\in\mathcal{S}^{\sigma_{p}}_{A}, let tA,p,x(𝐒σp(𝐕))1At_{A,p,x}\in(\mathbf{S}_{\sigma_{p}}(\mathbf{V}))_{1^{A}} be the image of xx under this isomorphism. We refer to AA as the support of the element tA,p,xt_{A,p,x}. Let 𝒯R\mathcal{T}\subset R be the set of all elements of the form tA,p,xt_{A,p,x} for all choices of AA, pp, and xx, and let A\mathcal{M}_{A} be the set of all monomials t1tst_{1}\cdots t_{s} where the tit_{i}’s belong to 𝒯\mathcal{T} and their supports form a partition of AA. We thus see that A\mathcal{M}_{A} is a basis for R1AR_{1^{A}}.

From the above discussion, we see that the 1n1^{n}-weight space of R𝐕mR\otimes\mathbf{V}^{\otimes m} has for a basis all elements of the form

Tes1smT\otimes e_{s_{1}}\otimes\cdots\otimes_{s_{m}}

where s1,,sms_{1},\ldots,s_{m} are distinct elements of [m][m] and TAT\in\mathcal{M}_{A} with A=[n]{s1,,sm}A=[n]\setminus\{s_{1},\ldots,s_{m}\}. We associate to the above element the σ¯\underline{\smash{\sigma}}-diagram given by the pair (i,Γ)(i,\Gamma), where Γ\Gamma is the collection of blocks corresponding to TT (each tA,p,xt_{A,p,x} corresponds to a block (A,p,x)(A,p,x)), and i:[n]A[m]i\colon[n]\setminus A\to[m] is the bijection taking sjs_{j} to jj. We have thus constructed a natural linear isomorphism

(R𝐕m)1n=Hom𝔇(σ¯)([n],[m]).(R\otimes\mathbf{V}^{\otimes m})_{1^{n}}=\operatorname{Hom}_{\mathfrak{D}(\underline{\smash{\sigma}})}([n],[m]).

As we have already seen, the left side above is identified with HomR(Xn,Xm)\operatorname{Hom}_{R}(X^{\otimes n},X^{\otimes m}). One easily sees that the resulting isomorphism

Hom𝔇(σ¯)([n],[m])=HomR(Xn,Xm)\operatorname{Hom}_{\mathfrak{D}(\underline{\smash{\sigma}})}([n],[m])=\operatorname{Hom}_{R}(X^{\otimes n},X^{\otimes m})

is induced by θ\theta. This shows that θ\theta is fully faithful. This completes the first step of the proof.

Step 2: equivalence of (b) and (c). Let 𝒞\mathcal{C}^{\prime} be the category in (c). The functor ModRModK\operatorname{Mod}_{R}\to\operatorname{Mod}_{K} given by MKRMM\mapsto K\otimes_{R}M induces a functor 𝒞𝒞\mathcal{C}\to\mathcal{C}^{\prime}. This functor is clearly symmetric monoidal, faithful, and essentially surjective. It is full by Proposition 2.7. Thus it is an equivalence.

Step 3: the remainder. To complete the proof, it suffices to show that the categories in (a), (d), and (e) are equivalent, as kk-linear categories. This follows from Proposition 5.1. ∎

For an abelian category 𝒜\mathcal{A}, we let 𝒜lf\mathcal{A}^{\mathrm{lf}} be the full subcategory spanned by objects that are locally of finite length (i.e., the union of their finite length subobjects).

Proposition 5.8.

We have the following equivalences of kk-linear abelian categories:

  1. (a)

    ModRMod𝔘(σ¯)\operatorname{Mod}_{R}\cong\operatorname{Mod}_{\mathfrak{U}(\underline{\smash{\sigma}})}

  2. (b)

    ModKMod𝔘(σ¯)lf\operatorname{Mod}_{K}\cong\operatorname{Mod}_{\mathfrak{U}(\underline{\smash{\sigma}})}^{\mathrm{lf}}

  3. (c)

    ModKModRlf\operatorname{Mod}_{K}\cong\operatorname{Mod}_{R}^{\mathrm{lf}}.

Proof.

Let 𝒞\mathcal{C} and 𝒞\mathcal{C}^{\prime} be Grothendieck abelian categories, and let 𝒫\mathcal{P} and 𝒫\mathcal{P}^{\prime} be full subcategories of 𝒞\mathcal{C} and 𝒞\mathcal{C}^{\prime} consisting of projective objects. Suppose that 𝒫\mathcal{P} and 𝒫\mathcal{P}^{\prime} are enough projectives (i.e., they form generating families). Then any equivalence 𝒫𝒫\mathcal{P}\to\mathcal{P}^{\prime} extends uniquely to an equivalence 𝒞𝒞\mathcal{C}\to\mathcal{C}^{\prime}. A similar statement holds for categories of injective objects.

Statement (a) now follows from the equivalence between the categories (b) and (d) in Proposition 5.7; it is clear that the categories in (b) and (d) are enough projectives in ModR\operatorname{Mod}_{R} and Mod𝔘(σ¯)\operatorname{Mod}_{\mathfrak{U}(\underline{\smash{\sigma}})}. Statement (b) follows from the equivalence between the categories (c) and (e) in Proposition 5.7; the fact that category (c) gives enough injectives in ModK\operatorname{Mod}_{K} follows from Theorem 4.10, while the fact that category (d) gives enough injectives in Mod𝔘(σ¯)lf\operatorname{Mod}_{\mathfrak{U}(\underline{\smash{\sigma}}})^{\mathrm{lf}} is Proposition 5.2 (note that 𝔘(σ¯)\mathfrak{U}(\underline{\smash{\sigma}}) is an upwards category, as defined before Proposition 5.2). Statement (c) follows from statements (a) and (b). ∎

Remark 5.9.

The equivalence ModKModRlf\operatorname{Mod}_{K}\cong\operatorname{Mod}_{R}^{\mathrm{lf}} has previously been established for a few values of σ¯\underline{\smash{\sigma}}: for [(1)][(1)] in [SS1], for [(1),,(1)][(1),\ldots,(1)] in [SS5], for [(2)][(2)] and [(1,1)][(1,1)] in [NSS], and for [(1,1),(1)][(1,1),(1)] in [SS6]. Related results also appear in [NSS2] and [NSS3]. ∎

For a partition λ\lambda, recall that LλL_{\lambda} denote the simple object of ModK\operatorname{Mod}_{K} indexed by λ\lambda. Using the above proposition, we can compute the Ext\operatorname{Ext} groups between these objects:

Corollary 5.10.

We have

ExtKi(Lλ,Lμ)=Hom𝐆𝐋(i(kσ¯)kλ,kμ).\operatorname{Ext}^{i}_{K}(L_{\lambda},L_{\mu})=\operatorname{Hom}_{\mathbf{GL}}({\textstyle\bigwedge}^{i}(k^{\oplus\underline{\smash{\sigma}}})\otimes k^{\oplus\lambda},k^{\oplus\mu}).
Proof.

Let 𝒜=Mod𝔘(σ¯)\mathcal{A}=\operatorname{Mod}_{\mathfrak{U}(\underline{\smash{\sigma}})} and let Φ:ModK𝒜lf\Phi\colon\operatorname{Mod}_{K}\to\mathcal{A}^{\mathrm{lf}} be the equivalence constructed above. Tracing through the definition, we see that Φ\Phi takes K𝐕nK\otimes\mathbf{V}^{\otimes n} to the principal injective 𝐈n\mathbf{I}_{n}. We thus see that Φ(Kλ)\Phi(K^{\oplus\lambda}) is the SλS^{\lambda}-isotypic piece of 𝐈n\mathbf{I}_{n}, with respect to its natural 𝔖n\mathfrak{S}_{n}-action. Taking socles, we see that Lλ=Φ(Lλ)L^{\prime}_{\lambda}=\Phi(L_{\lambda}) is the simple 𝔘(σ)\mathfrak{U}(\sigma)-module with (Lλ)n=Sλ(L^{\prime}_{\lambda})_{n}=S^{\lambda}. We thus have

ExtKi(Lλ,Lμ)=Ext𝒜lfi(Lλ,Lμ).\operatorname{Ext}^{i}_{K}(L_{\lambda},L_{\mu})=\operatorname{Ext}^{i}_{\mathcal{A}^{\mathrm{lf}}}(L^{\prime}_{\lambda},L^{\prime}_{\mu}).

The Ext\operatorname{Ext} on the right side can be computed by taking an injective resolution of LμL^{\prime}_{\mu} in 𝒜lf\mathcal{A}^{\mathrm{lf}}. As we have seen (Proposition 5.2), this can be accomplished using principal injectives. As these objects are injective in the larger category 𝒜\mathcal{A}, we find

Ext𝒜lfi(Lλ,Lμ)=Ext𝒜i(Lλ,Lμ).\operatorname{Ext}^{i}_{\mathcal{A}^{\mathrm{lf}}}(L^{\prime}_{\lambda},L^{\prime}_{\mu})=\operatorname{Ext}^{i}_{\mathcal{A}}(L^{\prime}_{\lambda},L^{\prime}_{\mu}).

We now appeal to the equivalence 𝒜=ModR\mathcal{A}=\operatorname{Mod}_{R}. One easily sees that LλL^{\prime}_{\lambda} corresponds to the simple RR-module kλk^{\oplus\lambda} (with positive degree elements of RR acting by 0). We thus have

Ext𝒜i(Lλ,Lμ)=ExtRi(kλ,kμ).\operatorname{Ext}^{i}_{\mathcal{A}}(L^{\prime}_{\lambda},L^{\prime}_{\mu})=\operatorname{Ext}^{i}_{R}(k^{\oplus\lambda},k^{\oplus\mu}).

The right group can be computed using the projective resolution of kλk^{\oplus\lambda} provided by the Koszul complex. This yields the stated result. ∎

5.4. Weyl’s construction

We recall Weyl’s classical traceless tensor construction. Equip 𝐂r\mathbf{C}^{r} with a non-degenerate symmetric bilinear form. Let Tn=(𝐂r)nT^{n}=(\mathbf{C}^{r})^{\otimes n}. Given 1i<jn1\leq i<j\leq n, let φi,j:TnTn2\varphi_{i,j}\colon T^{n}\to T^{n-2} be the map obtained by applying the form to the ii and jj tensor factors. Let T[n]T^{[n]} be the intersection of the kernels of φi,j\varphi_{i,j}, over all choices of ii and jj; this is the space of traceless tensors. The space T[n]T^{[n]} is a (𝔖n×𝐎r)(\mathfrak{S}_{n}\times\mathbf{O}_{r})-subrepresentation of TnT^{n}. Weyl proved that the SλS^{\lambda} isotypic piece of T[n]T^{[n]} is either 0 or the irreducible of 𝐎r\mathbf{O}_{r} with highest weight λ\lambda.

We now establish an analog of this construction for ModK\operatorname{Mod}_{K}. Recall that σ¯=[σ1,,σr]\underline{\smash{\sigma}}=[\sigma_{1},\ldots,\sigma_{r}]. For each 1ir1\leq i\leq r, let φi:KσiK\varphi_{i}\colon K^{\oplus\sigma_{i}}\to K be the natural map (coming from the inclusion kσiKk^{\oplus\sigma_{i}}\subset K). Given an element xx of the Specht module SσiS^{\sigma_{i}}, let φi,x\varphi_{i,x} be the composition

K𝐕|σi|K𝐒σi(𝐕)KK\otimes\mathbf{V}^{\otimes|\sigma_{i}|}\to K\otimes\mathbf{S}_{\sigma_{i}}(\mathbf{V})\to K

where the first map comes from the projection 𝐕|σi|𝐒σi(𝐕)\mathbf{V}^{\otimes|\sigma_{i}|}\to\mathbf{S}_{\sigma_{i}}(\mathbf{V}) provided by xx, and the second map is φi\varphi_{i}. Let Tn=K𝐕nT^{n}=K\otimes\mathbf{V}^{\otimes n}. Given 1in1\leq i\leq n, xSσix\in S^{\sigma_{i}}, and a subset SS of [n][n] of size |σi||\sigma_{i}|, we let φi,x,S:TnTn|σi|\varphi_{i,x,S}\colon T^{n}\to T^{n-|\sigma_{i}|} be the map obtained by applying φi,x\varphi_{i,x} to the tensor factors indexed by SS. Let T[n]T^{[n]} be the intersection of the kernels of the φi,x,S\varphi_{i,x,S} over all choices of ii, xx, and SS. This is a KK-module equipped with an action of 𝔖n\mathfrak{S}_{n}. The following is our analog of Weyl’s construction:

Proposition 5.11.

Let λ\lambda be a partition of nn. Then the SλS^{\lambda} isotypic piece of T[n]T^{[n]} is the simple KK-module LλL_{\lambda}.

Proof.

Under the equivalence ModK=Mod𝔘(λ¯)lf\operatorname{Mod}_{K}=\operatorname{Mod}_{\mathfrak{U}(\underline{\smash{\lambda}})}^{\mathrm{lf}}, the KK-module TnT^{n} corresponds to the nnth principal injective 𝔘(λ¯)\mathfrak{U}(\underline{\smash{\lambda}})-module. Thinking in terms of λ¯\underline{\smash{\lambda}}-diagrams, we see that any map TnTmT^{n}\to T^{m}, with m<nm<n, is a linear combination of maps of the form fφi,x,Sf\circ\varphi_{i,x,S}, where ff is some map. It follows that T[n]T^{[n]} is the intersection of the kernels of all maps TnTmT^{n}\to T^{m} with m<nm<n. From this, we see that the SλS^{\lambda}-isotypic piece of T[n]T^{[n]} is the intersection of the kernels of alls maps KλKμK^{\oplus\lambda}\to K^{\oplus\mu} with |μ|<|λ||\mu|<|\lambda|. This is the simple object LλL_{\lambda} (see §4.2). ∎

5.5. Universal properties

We can now give the universal property for the category ModK\operatorname{Mod}_{K}. This is analogous to the universal property for Rep(𝐎)\operatorname{Rep}(\mathbf{O}) given in [SS3, §4.4]. For symmetric monoidal kk-linear abelian categories 𝒞\mathcal{C} and 𝒟\mathcal{D}, we let LExk(𝒞,𝒟)\operatorname{LEx}^{\otimes}_{k}(\mathcal{C},\mathcal{D}) be the category of left-exact symmetric monoidal kk-linear functors 𝒞𝒟\mathcal{C}\to\mathcal{D}. Also, recall the category Tσ¯(𝒞)T_{\underline{\smash{\sigma}}}(\mathcal{C}) defined before Proposition 5.6.

Theorem 5.12.

Let (𝒞,)(\mathcal{C},\otimes) be a symmetric monoidal kk-linear abelian category with \otimes exact. Then we have a natural equivalence of categories

LExk(ModKf,𝒞)Tσ¯(𝒞).\operatorname{LEx}^{\otimes}_{k}(\operatorname{Mod}_{K}^{\mathrm{f}},\mathcal{C})\cong T_{\underline{\smash{\sigma}}}(\mathcal{C}).

In other words, to give a kk-linear left-exact symmetric monoidal functor ModKf𝒞\operatorname{Mod}_{K}^{\mathrm{f}}\to\mathcal{C} is the same as to give an object of 𝒞\mathcal{C} equipped with a σ¯\underline{\smash{\sigma}}-form.

Proof.

Let \mathcal{I} be the full subcategory of ModKf\operatorname{Mod}_{K}^{\mathrm{f}} spanned by the objects K𝐕nK\otimes\mathbf{V}^{\otimes n} for n0n\geq 0. This category is stable under tensor products. As a kk-linear symmetric monoidal category, it is equivalent to 𝔇(σ¯)\mathfrak{D}(\underline{\smash{\sigma}}) by Proposition 5.7. Thus by the universal property for 𝔇(σ¯)\mathfrak{D}(\underline{\smash{\sigma}}) (Proposition 5.6), we have a natural equivalence

Funk(,𝒞)Tσ¯(𝒞).\operatorname{Fun}^{\otimes}_{k}(\mathcal{I},\mathcal{C})\cong T_{\underline{\smash{\sigma}}}(\mathcal{C}).

Now, every object of \mathcal{I} is injective in ModK\operatorname{Mod}_{K} (Theorem 4.10(b)), and every object of ModKf\operatorname{Mod}_{K}^{\mathrm{f}} embeds into a finite direct sum of objects in \mathcal{I} (Theorem 3.9). It follows that any functor 𝒞\mathcal{I}\to\mathcal{C} extends uniquely to a left-exact functor ModKf𝒞\operatorname{Mod}_{K}^{\mathrm{f}}\to\mathcal{C}. Since \mathcal{I} is stable under tensor products, and all tensor products are exact, it follows that this extended functor is symmetric monoidal if the original functor is. This completes the proof. ∎

Remark 5.13.

Let VV be a finite dimensional kk-vector space equipped with a form ω:𝐒σ¯(V)k\omega\colon\mathbf{S}_{\underline{\smash{\sigma}}}(V)\to k. From the universal property, get a left-exact cocontinuous symmetric monoidal functor

Γ:ModKVeck\Gamma\colon\operatorname{Mod}_{K}\to\mathrm{Vec}_{k}

that we call the specialization functor with respect to VV and ω\omega. Since Γ\Gamma is left-exact, one can consider its right derived functors RiΓ\mathrm{R}^{i}\Gamma, which we call the derived specialization functors. Is it possible to compute the values of these functors on simple objects for a generic form ω\omega? When σ¯=[(2)]\underline{\smash{\sigma}}=[(2)] the category ModK\operatorname{Mod}_{K} is equivalent to the category of algebraic representations of the infinite orthogonal group (see [NSS, Theorem 3.1]), as studied in [SS3], and the derived specialization of simple objects was computed in [SSW]. ∎

6. Classification of fiber functors

In this section, we introduce the notion of a fiber functor on ModK\operatorname{Mod}_{K}, and give a complete classification of them.

6.1. Definitions

Fix, for the duration of §6, a 𝐆𝐋\mathbf{GL}-field KK that is finitely generated over its invariant subfield kk, and a 𝐆𝐋\mathbf{GL}-algebra RR finitely generated over R0=kR_{0}=k with Frac(R)=K\operatorname{Frac}(R)=K. Furthermore, let X=Spec(R)X=\operatorname{Spec}(R) be the 𝐆𝐋\mathbf{GL}-variety associated to RR. The following is the main object of study in this section:

Definition 6.1.

A fiber functor on ModK\operatorname{Mod}_{K} is a symmetric monoidal functor Φ:ModKVeck\Phi\colon\operatorname{Mod}_{K}\to\mathrm{Vec}_{k} that is exact, faithful, cocontinuous, and kk-linear. ∎

The goal of this section is to classify the fiber functors on ModK\operatorname{Mod}_{K}. This is accomplished in Theorem 6.5 below.

6.2. Examples of fiber functors

Let xx be a 𝐆𝐋\mathbf{GL}-generic kk-point of XX and let 𝔪\mathfrak{m} be the corresponding maximal ideal of RR. Define a functor

Φ~x:ModRVeck,Φ~x(M)=M/𝔪M.\tilde{\Phi}_{x}\colon\operatorname{Mod}_{R}\to\mathrm{Vec}_{k},\qquad\tilde{\Phi}_{x}(M)=M/\mathfrak{m}M.

Since every RR-module is flat at 𝔪\mathfrak{m} (Corollary 3.7), it follows that Φ~x\tilde{\Phi}_{x} is exact. Moreover, it is clear that Φ~x\tilde{\Phi}_{x} kills torsion RR-modules. It follows that Φ~x\tilde{\Phi}_{x} factors through the generic category ModRgen\operatorname{Mod}_{R}^{\mathrm{gen}}. Identifying this with ModK\operatorname{Mod}_{K}, we thus obtain a functor

Φx:ModKVeck.\Phi_{x}\colon\operatorname{Mod}_{K}\to\mathrm{Vec}_{k}.

We now have:

Proposition 6.2.

The functor Φx\Phi_{x} is a fiber functor (in a natural manner).

Proof.

The functor Φ~x\tilde{\Phi}_{x} is clearly exact, cocontinuous, and kk-linear, and also admits a natural symmetric monoidal structure; it follows that Φx\Phi_{x} inherits these properties. To complete the proof, we must show that Φx\Phi_{x} is faithful.

We first claim that if MM is a torsion-free RR-module such that M/𝔪M=0M/\mathfrak{m}M=0 then M=0M=0. To see this, first suppose that MM is finitely generated. Then M𝔪M_{\mathfrak{m}} is free over R𝔪R_{\mathfrak{m}} (Proposition 3.6). Thus the vanishing of M/𝔪M=M𝔪/𝔪M𝔪M/\mathfrak{m}M=M_{\mathfrak{m}}/\mathfrak{m}M_{\mathfrak{m}} implies that of M𝔪M_{\mathfrak{m}}, and thus of MM since MM is torsion-free. We now treat the general case. Let NN be a finitely generated submodule of MM. Since M/NM/N is flat at 𝔪\mathfrak{m} (Corollary 3.7), the map N/𝔪NM/𝔪MN/\mathfrak{m}N\to M/\mathfrak{m}M is injective, and so N/𝔪N=0N/\mathfrak{m}N=0. Thus N=0N=0 by the previous case. Since NN was arbitrary, it follows that M=0M=0 as well.

Now, to prove faithfulness, it suffices to show that if f:MNf\colon M\to N is a map of torsion-free RR-modules such that the induced map f¯:M/𝔪MN/𝔪N\overline{f}\colon M/\mathfrak{m}M\to N/\mathfrak{m}N vanishes then f=0f=0. Thus let such an ff be given. Let II be the image of ff. Since N/IN/I is flat at 𝔪\mathfrak{m} (Corollary 3.7), it follows that I/𝔪II/\mathfrak{m}I is the image of f¯\overline{f}, and thus vanishes. Hence I=0I=0 by the previous paragraph, and so f=0f=0 as well. ∎

Remark 6.3.

Proposition 6.2 was proven for R=Sym(Sym2(𝐂))R=\operatorname{Sym}(\operatorname{Sym}^{2}(\mathbf{C}^{\infty})) (and a specific choice of xx) in [NSS, §3]. Similar results were also proved in [NSS2, §6], [NSS3, §5], [SS6, §5]. However, these papers did not have the benefit of the shift theorem and its corollaries, such as Corollary 3.7, and as a result the arguments given there are much more involved. ∎

It is possible that ModK\operatorname{Mod}_{K} does not admit a fiber functor. However, this can be fixed by passing to a finite extension:

Proposition 6.4.

There exists a finite extension k/kk^{\prime}/k such that, putting K=kkKK^{\prime}=k^{\prime}\otimes_{k}K, the category ModK\operatorname{Mod}_{K^{\prime}} admits a fiber functor.

Proof.

There is a finite extension k/kk^{\prime}/k such that XX contains a 𝐆𝐋\mathbf{GL}-generic kk^{\prime}-point xx [BDES, Theorem 8.8]. As we have seen (Lemma 4.18 and following discussion), K=kkKK^{\prime}=k^{\prime}\otimes_{k}K is then a 𝐆𝐋\mathbf{GL}-field that is finitely generated over its invariant field kk^{\prime}. It follows that Φx\Phi_{x} is a fiber functor for ModK\operatorname{Mod}_{K^{\prime}}. ∎

6.3. More examples of fiber functors

Let VV be an infinite dimensional kk-vector space. Recall that X{V}=Spec(R{V})X\{V\}=\operatorname{Spec}(R\{V\}), where R{V}R\{V\} is obtained by treating RR as a polynomial functor and evaluating on VV. Suppose that xx is a 𝐆𝐋\mathbf{GL}-generic kk-point of X{V}X\{V\}, corresponding to the maximal ideal 𝔪\mathfrak{m} of R{V}R\{V\}. (By 𝐆𝐋\mathbf{GL}-generic here, we mean there is no proper closed 𝐆𝐋\mathbf{GL}-subvariety ZZ of XX with xZ{V}x\in Z\{V\}.) Define a functor

Φ~V,x:ModRModk,Φ~V,x(M)=M{V}/𝔪M{V}.\tilde{\Phi}_{V,x}\colon\operatorname{Mod}_{R}\to\operatorname{Mod}_{k},\qquad\tilde{\Phi}_{V,x}(M)=M\{V\}/\mathfrak{m}M\{V\}.

Once again, this functor is exact and kills torsion modules, and thus induces a functor

ΦV,x:ModKModk.\Phi_{V,x}\colon\operatorname{Mod}_{K}\to\operatorname{Mod}_{k}.

The same argument as in Proposition 6.2 shows that it too is a fiber functor. If V=k𝐕V=k\otimes\mathbf{V} then ΦV,x\Phi_{V,x} is the funtor Φx\Phi_{x} introduced above. We note that ΦV,x(K𝐕)=V\Phi_{V,x}(K\otimes\mathbf{V})=V, and so ΦV,x\Phi_{V,x} and ΦV,x\Phi_{V^{\prime},x^{\prime}} can only be isomorphic if dimV=dimV\dim{V}=\dim{V^{\prime}} (as cardinal numbers). In particular, if dim(V)dim(𝐕)\dim(V)\neq\dim(\mathbf{V}) then ΦV,x\Phi_{V,x} will not be isomorphic to a fiber functor of the form Φx\Phi_{x^{\prime}}.

6.4. The main theorem

In the remainder of this section, a pair (V,x)(V,x) will always stand for an infinite dimensional kk-vector space VV and a 𝐆𝐋\mathbf{GL}-generic kk-point xx of X{V}X\{V\}. If (V,x)(V^{\prime},x^{\prime}) is a second such pair, then an isomorphism (V,x)(V,x)(V,x)\to(V^{\prime},x^{\prime}) is a linear isomorphism VVV\to V^{\prime} such that the induced map X{V}X{V}X\{V^{\prime}\}\to X\{V\} carries xx^{\prime} to xx. The following theorem classifies fiber functors:

Theorem 6.5.

We have the following:

  1. (a)

    Any fiber functor on ModK\operatorname{Mod}_{K} is isomorphic to one of the form ΦV,x\Phi_{V,x}.

  2. (b)

    Given two pairs (V,x)(V,x) and (V,x)(V^{\prime},x^{\prime}), we have a natural bijection

    Isom((V,x),(V,x))=Isom(ΦV,x,ΦV,x).\operatorname{Isom}((V,x),(V^{\prime},x^{\prime}))=\operatorname{Isom}(\Phi_{V,x},\Phi_{V^{\prime},x^{\prime}}).

    These bijections are compatible with composition of isomorphisms.

The theorem is proved in §6.5 below. We make a few remarks here.

Remark 6.6.

The theorem can be stated more concisely as: the groupoid of fiber functors on ModK\operatorname{Mod}_{K} is equivalent to the groupoid of pairs (V,x)(V,x). ∎

Remark 6.7.

It follows from the theorem that the automorphism group of the fiber functor ΦV,x\Phi_{V,x} is the stabilizer of xx in the group Autk(V)\operatorname{Aut}_{k}(V). In most cases, this group will be finite, and so ModK\operatorname{Mod}_{K} cannot be recovered as its representation category. This issue is addressed in §7 and §8 by introducing the notion of “generalized stabilizers.” ∎

Remark 6.8.

Given KK, there are potentially many choices of XX. The theorem implies that any two choices of XX have the same set of 𝐆𝐋\mathbf{GL}-generic points (up to natural bijection). In fact, this can be seen directly. Suppose RR^{\prime} is a second 𝐆𝐋\mathbf{GL}-algebra that is finitely 𝐆𝐋\mathbf{GL}-generated over kk and has Frac(R)=K\operatorname{Frac}(R^{\prime})=K, and let X=Spec(R)X^{\prime}=\operatorname{Spec}(R^{\prime}). One can show that XX and XX^{\prime} are birational, in the sense that there are open 𝐆𝐋\mathbf{GL}-subsets UXU\subset X and UXU^{\prime}\subset X^{\prime} and an isomorphism i:UUi\colon U\to U^{\prime} of 𝐆𝐋\mathbf{GL}-varieties. Every 𝐆𝐋\mathbf{GL}-generic point of XX is contained in UU, and similarly every 𝐆𝐋\mathbf{GL}-generic point of XX^{\prime} is contained in UU^{\prime} (see [BDES, Proposition 3.4]). Clearly, these points are mapped bijectively to one another via ii. ∎

6.5. Proof of Theorem 6.5

Let VV be a vector space, let xx be a kk-point of X{V}X\{V\}, and let 𝔪\mathfrak{m} be the corresponding maximal ideal of R{V}R\{V\}. We define Φ~V,x\tilde{\Phi}_{V,x} as above; that is, for an RR-module MM, we put

Φ~V,x(M)=M{V}/𝔪M{V}.\tilde{\Phi}_{V,x}(M)=M\{V\}/\mathfrak{m}M\{V\}.

Previously, we had only used this when VV is infinite dimensional and xx is 𝐆𝐋\mathbf{GL}-generic, but we now consider it more generally.

Lemma 6.9.

Let VV and xx be as above. Suppose that there is a fiber functor Φ\Phi on ModK\operatorname{Mod}_{K} such that Φ~V,x(M)=Φ(KRM)\tilde{\Phi}_{V,x}(M)=\Phi(K\otimes_{R}M). Then VV is infinite dimensional and xx is 𝐆𝐋\mathbf{GL}-generic.

Proof.

Suppose, by way of contradiction, that xx is not 𝐆𝐋\mathbf{GL}-generic (which is automatic if VV is finite dimensional). There is then a non-zero 𝐆𝐋\mathbf{GL}-ideal II of RR such that xx belongs to the vanishing locus of I{V}I\{V\}. Then Φ~V,x(R/I)=R{V}/𝔪\tilde{\Phi}_{V,x}(R/I)=R\{V\}/\mathfrak{m} is non-zero. On the other hand, KRR/I=0K\otimes_{R}R/I=0, and so Φ(KRR/I)=0\Phi(K\otimes_{R}R/I)=0. This is a contradiction, which completes the proof. ∎

Lemma 6.10.

Let Φ\Phi be a fiber functor on ModK\operatorname{Mod}_{K}. Then Φ\Phi is isomorphic to some ΦV,x\Phi_{V,x}.

Proof.

Let V=Φ(𝐕K)V=\Phi(\mathbf{V}\otimes K). Suppose UU is a polynomial representation of 𝐆𝐋\mathbf{GL}. Then we have UK=U(𝐕K)U\otimes K=U(\mathbf{V}\otimes K), where on the right side we treat UU as a polynomial functor and apply it to the object 𝐕K\mathbf{V}\otimes K of ModK\operatorname{Mod}_{K}. We thus find

Φ(UK)=Φ(U(𝐕K))=U(Φ(𝐕K))=U{V},\Phi(U\otimes K)=\Phi(U(\mathbf{V}\otimes K))=U(\Phi(\mathbf{V}\otimes K))=U\{V\},

where in the second step we used that Φ\Phi commutes with the action of poylnomial functors, as Φ\Phi is symmetric monoidal.

We have a natural surjective map α:RkKK\alpha\colon R\otimes_{k}K\to K of algebra objects in ModK\operatorname{Mod}_{K}, given by multiplication. Applying Φ\Phi, and appealing to the above, this yields a surjective kk-algebra homomorphism map β:R{V}k\beta\colon R\{V\}\to k. Let 𝔪=ker(β)\mathfrak{m}=\ker(\beta), a maximal ideal of R{V}R\{V\}, and let xX{V}x\in X\{V\} is the associated point.

Now, let MM be an RR-module. Choose a presentation

U1kRU0kRM0U_{1}\otimes_{k}R\to U_{0}\otimes_{k}R\to M\to 0

where U0U_{0} and U1U_{1} are polynomial representations. We obtain a commutative diagram

U1kRkK\textstyle{U_{1}\otimes_{k}R\otimes_{k}K\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}idα\scriptstyle{\mathrm{id}\otimes\alpha}U0kRkK\textstyle{U_{0}\otimes_{k}R\otimes_{k}K\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}idα\scriptstyle{\mathrm{id}\otimes\alpha}MkK\textstyle{M\otimes_{k}K\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}U1kK\textstyle{U_{1}\otimes_{k}K\ignorespaces\ignorespaces\ignorespaces\ignorespaces}U0kK\textstyle{U_{0}\otimes_{k}K\ignorespaces\ignorespaces\ignorespaces\ignorespaces}MRK\textstyle{M\otimes_{R}K\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

with exact rows. Applying Φ\Phi, we obtain a commutative diagram

U1{V}kR{V}\textstyle{U_{1}\{V\}\otimes_{k}R\{V\}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}idβ\scriptstyle{\mathrm{id}\otimes\beta}U0{V}kR{V}\textstyle{U_{0}\{V\}\otimes_{k}R\{V\}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}idβ\scriptstyle{\mathrm{id}\otimes\beta}M{V}\textstyle{M\{V\}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}U1{V}\textstyle{U_{1}\{V\}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}U0{V}\textstyle{U_{0}\{V\}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Φ(MRK)\textstyle{\Phi(M\otimes_{R}K)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

It follows that the right vertical map induces an isomorphism

Φ~V,x(M)=M{V}/𝔪M{V}Φ(MRK).\tilde{\Phi}_{V,x}(M)=M\{V\}/\mathfrak{m}M\{V\}\to\Phi(M\otimes_{R}K).

By Lemma 6.9, we see that VV is infinite dimensional and xx is 𝐆𝐋\mathbf{GL}-generic. The above isomorphism thus induces an isomorphism ΦΦV,x\Phi\cong\Phi_{V,x}. ∎

Lemma 6.11.

Let (V,x)(V,x) and (V,x)(V^{\prime},x^{\prime}) be given. Then we have a natural bijection

Isom((V,x),(V,x))=Isom(ΦV,x,ΦV,x)\operatorname{Isom}((V,x),(V^{\prime},x^{\prime}))=\operatorname{Isom}(\Phi_{V,x},\Phi_{V^{\prime},x^{\prime}})

that is compatible with composition of isomorphisms.

Proof.

We first construct a map

α:Isom((V,x),(V,x))Isom(ΦV,x,ΦV,x).\alpha\colon\operatorname{Isom}((V,x),(V^{\prime},x^{\prime}))\to\operatorname{Isom}(\Phi_{V,x},\Phi_{V^{\prime},x^{\prime}}).

Thus let f:VVf\colon V\to V^{\prime} be a kk-linear isomorphism such that the induced map X{V}X{V}X\{V^{\prime}\}\to X\{V\} takes xx^{\prime} to xx. It follows that under the induced ring homomorphism R{V}R{V}R\{V\}\to R\{V^{\prime}\} the ideal 𝔪\mathfrak{m}^{\prime} contracts to the ideal 𝔪\mathfrak{m}. Let MM be an RR-module. Then ff induces an isomorphism M{V}M{V}M\{V\}\to M\{V^{\prime}\}, which further induces an isomorphism on the quotients by 𝔪\mathfrak{m} and 𝔪\mathfrak{m}^{\prime}. This yields an isomorphism Φ~V,xΦ~V,x\tilde{\Phi}_{V,x}\cong\tilde{\Phi}_{V^{\prime},x^{\prime}} which, in turn, leads to an isomorphism g:ΦV,xΦV,xg\colon\Phi_{V,x}\cong\Phi_{V^{\prime},x^{\prime}}. We define α(f)=g\alpha(f)=g.

We now define a map

β:Isom(ΦV,x,ΦV,x)Isom((V,x),(V,x)).\beta\colon\operatorname{Isom}(\Phi_{V,x},\Phi_{V^{\prime},x^{\prime}})\to\operatorname{Isom}((V,x),(V^{\prime},x^{\prime})).

Let g:ΦV,xΦV,xg\colon\Phi_{V,x}\to\Phi_{V^{\prime},x^{\prime}} be an isomorphism of fiber functors. As ΦV,x(𝐕kK)=V\Phi_{V,x}(\mathbf{V}\otimes_{k}K)=V, and similarly for ΦV,x\Phi_{V^{\prime},x^{\prime}}, we see that gg induces a kk-linear isomorphism f:VVf\colon V\to V^{\prime}. Let II be the kernel of the map RkKKR\otimes_{k}K\to K in ModK\operatorname{Mod}_{K}. As we have seen, ΦV,x=𝔪\Phi_{V,x}=\mathfrak{m}, and similarly for ΦV,x\Phi_{V^{\prime},x^{\prime}}. We thus see that under the ring isomorphism R{V}R{V}R\{V\}\to R\{V^{\prime}\} induces by ff, the ideal 𝔪\mathfrak{m} is taken to 𝔪\mathfrak{m}^{\prime}. Thus ff defines an isomorphism (V,x)(V,x)(V,x)\to(V^{\prime},x^{\prime}). We put β(g)=f\beta(g)=f.

We leave to the reader the verification that α\alpha and β\beta are mutually inverse, and that these bijections are compatible with composition of isomorphisms. ∎

7. Germinal subgroups and their representations

In this section, we introduce germinal subgroups (§7.1), their representation theory (§7.2), and generalized stabilizers (§7.4) in the abstract. We also describe a general procedure for construction representations of generalized stabilizers (§7.5). This theory is applied in the next section when we study generalized stabilizers on 𝐆𝐋\mathbf{GL}-varieties.

7.1. Germinal subgroups

Fix a group GG. The following definition introduces the main concept studied in this section:

Definition 7.1.

A germinal subgroup of GG is a family Γ={Γ(i)}iI\Gamma=\{\Gamma(i)\}_{i\in I}, where II is a directed set and each Γ(i)\Gamma(i) is a subset of GG, satisfying the following conditions:

  1. (a)

    If iji\leq j then Γ(j)Γ(i)\Gamma(j)\subset\Gamma(i).

  2. (b)

    Each Γ(i)\Gamma(i) contains the identity element.

  3. (c)

    Given gΓ(i)g\in\Gamma(i) there is some jIj\in I such that Γ(j)gΓ(i)\Gamma(j)g\subset\Gamma(i). ∎

The generalized stabilizer of a point on a 𝐆𝐋\mathbf{GL}-variety will be a germinal subgroup. In this case, the intersection of the sets Γ(i)\Gamma(i) will be the usual stabilizer, which is typically “too small.” Each of the sets Γ(i)\Gamma(i), on the other hand, is “too big.” One can think of the germinal subgroup Γ\Gamma as a kind of filter on GG that is attempting to pick out a hypothetical subset that is bigger than the intersection but smaller than each Γ(i)\Gamma(i). As this picture suggests, one should always be allowed to pass to a cofinal subset of II when working in the setting of germinal subgroups.

7.2. Representations

We fix a germinal subgroup Γ={Γ(i)}iI\Gamma=\{\Gamma(i)\}_{i\in I} of GG for §7.2.

Definition 7.2.

A pre-representation of Γ\Gamma over a field kk consists of a kk-vector space VV and a linear function

VlimiIFun(Γ(i),V).V\to\varinjlim_{i\in I}\operatorname{Fun}(\Gamma(i),V).

Suppose that VV and WW are pre-representationss of Γ\Gamma over kk. A map of pre-representations is a kk-linear map VWV\to W such that the obvious diagram commutes. ∎

Suppose VV is a pre-representation. Given vVv\in V, its image in limiIFun(Γ(i),V)\varinjlim_{i\in I}\operatorname{Fun}(\Gamma(i),V) is represented by a function Γ(i)V\Gamma(i)\to V for some ii. Given an element gg of this Γ(i)\Gamma(i), we denote its image in VV under this function by gvgv. We thus think of a pre-representation as a kind of partially defined action map G×VVG\times V\dashrightarrow V.

Definition 7.3.

A representation of Γ\Gamma is a pre-representation VV such that the following two conditions hold:

  • We have 1v=v1v=v for all vVv\in V.

  • Given vVv\in V there exists iIi\in I such that for each gΓ(i)g\in\Gamma(i) there exists some jij\geq i such that h(gv)=(hg)vh(gv)=(hg)v for all hΓ(j)h\in\Gamma(j).

A map of representations is simply a map of pre-representations. We let Rep(Γ)\operatorname{Rep}(\Gamma) be the category of representations of Γ\Gamma over kk. ∎

We make a number of remarks concerning this definition.

  • Let VV be a representation of Γ\Gamma and let WW be a subspace of VV. Then WW is a subrepresentation of VV if and only if for every wWw\in W there exists iIi\in I such that Γ(i)wW\Gamma(i)w\subset W.

  • Let VV and WW be representations of Γ\Gamma and let f:VWf\colon V\to W be a linear map. Then ff is a map of representations if and only if for each vVv\in V there exists iIi\in I such that f(gv)=gf(v)f(gv)=gf(v) for all gΓ(i)g\in\Gamma(i).

  • Let VV be a representation of GG. Then VV naturally carries the structure of a Γ\Gamma-representation. A similar comments applies to maps of representations. We thus have a restriction functor Rep(G)Rep(Γ)\operatorname{Rep}(G)\to\operatorname{Rep}(\Gamma).

  • The category Rep(Γ)\operatorname{Rep}(\Gamma) is abelian. Kernels, cokernels, images, (arbitrary) direct sums, and direct limits are given in the usual manner on the underlying vector spaces. It follows that axiom (AB5) holds.

  • Let VV be a representation of Γ\Gamma. Extend the partially defined action map to a function G×VVG\times V\to V in any manner. This gives VV the structure of a module over the non-commutative polynomial RR ring with variables indexed by gg. Let κ\kappa be the dimension of RR as a kk-vector space. One easily sees that any RR-submodule of VV is a Γ\Gamma-subrepresentation. It follows that every vVv\in V is contained in a Γ\Gamma-subrepresentation of dimension at most κ\kappa, namely, RvRv. Thus, taking one Γ\Gamma representation from each isomorphism class of representations of dimension at most κ\kappa, one obtains a generating set for Rep(Γ)\operatorname{Rep}(\Gamma). It follows that Rep(Γ)\operatorname{Rep}(\Gamma) is a Grothendieck abelian category. In particular, it is complete.

  • From the above, we see that Rep(Γ)\operatorname{Rep}(\Gamma) has arbitrary products. These are not necessarily computed in the usual manner on the underlying vector space.

  • Similarly, we see that there is a notion of intersection for an arbitrary family of subrepresentations of a Γ\Gamma-representation. This intersection may not coincide with the usual intersection of vector subspaces.

  • Let VV and WW be representations of Γ\Gamma. We give the vector space VWV\otimes W the structure of a representation in the usual manner: that is, we define

    g(i=1nviwi)=i=1ngvigwi,g\cdot\big{(}\sum_{i=1}^{n}v_{i}\otimes w_{i}\big{)}=\sum_{i=1}^{n}gv_{i}\otimes gw_{i},

    provided gvigv_{i} and gwigw_{i} are defined for all ii. One easily verifies that this is indeed a representation. This construction endows Rep(Γ)\operatorname{Rep}(\Gamma) with a symmetric monoial structure.

7.3. Weak subrepresentations

Given any vector space VV, the dual space VV^{*} carries a natural topology: namely, a sequence (or net) {λj}jJ\{\lambda_{j}\}_{j\in J} in VV^{*} converges to λ\lambda if for every vector vVv\in V there is some j0Jj_{0}\in J such that λj(v)=λ(v)\lambda_{j}(v)=\lambda(v) for all jj0j\geq j_{0}. We call this the Π\Pi-topology. For a subspace WW of VV, we let WVW^{\perp}\subset V^{*} be its annihilator, i.e., the set of functionals λV\lambda\in V^{*} such that λ(w)=0\lambda(w)=0 for all wWw\in W. One easily sees that WW^{\perp} is Π\Pi-closed, and that WWW\mapsto W^{\perp} is a bijection between subspaces of VV and closed subspaces of VV^{*}.

Let VV be a representation of GG and let WW be a subspace. A Γ\Gamma-sequence is a sequence {gj}jJ\{g_{j}\}_{j\in J} in GG, indexed by some directed set JJ, such that for each iIi\in I there exists j0Jj_{0}\in J such that gjΓ(i)g_{j}\in\Gamma(i) for all jj0j\geq j_{0}. We say that WW is a weak Γ\Gamma-subrepresentation of VV if it satisfies the following condition: given λW\lambda\in W^{\perp} and a Γ\Gamma-sequence {gj}\{g_{j}\} such that gj1λg_{j}^{-1}\lambda converges in VV^{*} to an element μ\mu, we have μW\mu\in W^{\perp}.

Proposition 7.4.

Let VV be a representation of GG and let WW be Γ\Gamma-subrepresentation of VV. Then WW is a weak Γ\Gamma-subrepresentation of VV.

Proof.

Let λW\lambda\in W^{\perp} and let {gj}jJ\{g_{j}\}_{j\in J} be a Γ\Gamma-sequence such that gj1λg_{j}^{-1}\lambda converges in VV^{*} to some element μ\mu. Let wWw\in W. Since WW is a Γ\Gamma-subrepresentation, there exists iIi\in I such that gwWgw\in W for all gΓ(i)g\in\Gamma(i). Since gj1λg_{j}^{-1}\lambda converges to μ\mu there is some j0Jj_{0}\in J such that μ(w)=λ(gjw)\mu(w)=\lambda(g_{j}w) for all jj0j\geq j_{0}. Let j1j0j_{1}\geq j_{0} be such that gjΓ(i)g_{j}\in\Gamma(i) for all jj1j\geq j_{1}. Then for jj1j\geq j_{1} we have μ(w)=λ(gjw)=0\mu(w)=\lambda(g_{j}w)=0 since giwWg_{i}w\in W and λ\lambda vanishes on VV. Thus μ\mu vanishes on WW, and so μW\mu\in W^{\perp}. This shows that WW is a weak subrepresentation. ∎

7.4. Generalized stabilizers

Let II be a directed set and let {Xi}iI\{X_{i}\}_{i\in I} be an inverse system of sets; for iji\leq j, let πj,i:XjXi\pi_{j,i}\colon X_{j}\to X_{i} be the transistion map. Let XX be the inverse limit of the system. For iIi\in I, we let πi:XXi\pi_{i}\colon X\to X_{i} be the natural map. We suppose that a group GG acts on XX, and that the action satisfies the following condition: given gGg\in G and iIi\in I there exists jIj\in I such that πig\pi_{i}\circ g factors through πj\pi_{j}; in other words, one can complete the following commutative diagram:

X\textstyle{X\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}πj\scriptstyle{\pi_{j}}g\scriptstyle{g}X\textstyle{X\ignorespaces\ignorespaces\ignorespaces\ignorespaces}πi\scriptstyle{\pi_{i}}Xj\textstyle{X_{j}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Xi\textstyle{X_{i}}

Equivalently, this means that each gGg\in G acts uniformly continuously on XX, when XX is endowed with the inverse limit uniform structure (and each XiX_{i} with the discrete uniform structure).

We now come to a fundamental definition:

Definition 7.5.

Let xXx\in X. For iIi\in I, let Γx(i)\Gamma_{x}(i) be the set of elements gGg\in G such that πi(g1x)=πi(x)\pi_{i}(g^{-1}x)=\pi_{i}(x). The generalized stabilizer of xx is the system Γx={Γx(i)}iI\Gamma_{x}=\{\Gamma_{x}(i)\}_{i\in I}. ∎

Proposition 7.6.

The generalized stabilizer Γx\Gamma_{x} is a germinal subgroup of GG.

Proof.

We verify the three conditions of Definition 7.1. It is clear that 1Γx(i)1\in\Gamma_{x}(i) for all ii, which verifies condition (a). If gΓx(j)g\in\Gamma_{x}(j) and iji\leq j then taking the given identity πj(g1x)=πj(x)\pi_{j}(g^{-1}x)=\pi_{j}(x) and applying the transition map πj,i\pi_{j,i}, we find that πi(g1x)=πi(x)\pi_{i}(g^{-1}x)=\pi_{i}(x), and so gΓx(i)g\in\Gamma_{x}(i). This shows that Γx(j)Γx(i)\Gamma_{x}(j)\subset\Gamma_{x}(i), which verifies condition (b).

Finally, we come to condition (c). Suppose gΓig\in\Gamma_{i}. Let jj be such that we have a factorization πig1=φπj\pi_{i}\circ g^{-1}=\varphi\circ\pi_{j} for some φ:XjXi\varphi\colon X_{j}\to X_{i}. Suppose hΓx(j)h\in\Gamma_{x}(j). Then πj(h1x)=πj(x)\pi_{j}(h^{-1}x)=\pi_{j}(x). Applying φ\varphi, we find πi(g1h1x)=πi(x)\pi_{i}(g^{-1}h^{-1}x)=\pi_{i}(x), which shows that hgΓx(i)hg\in\Gamma_{x}(i). Thus Γx(j)gΓx(i)\Gamma_{x}(j)g\subset\Gamma_{x}(i), as required. ∎

Proposition 7.7.

The intersection iIΓx(i)\bigcap_{i\in I}\Gamma_{x}(i) is the usual stabilizer of xx, i.e., the set of all gGg\in G such that gx=xgx=x.

Proof.

It is clear that if gg stabilizes xx then gΓx(i)g\in\Gamma_{x}(i) for all ii. Conversely, if gΓx(i)g\in\Gamma_{x}(i) for all ii then we have πi(g1x)=πi(x)\pi_{i}(g^{-1}x)=\pi_{i}(x) for all ii, and so g1x=xg^{-1}x=x, which shows that gg stabilizes xx. ∎

7.5. Representations from equivariant bundles

Maintain the notation from §7.4. We now describe how to produce representations of Γx\Gamma_{x} from certain kinds of equivariant vector bundles on XX. This discussion is included simply to offer some intuition for germinal subgroups, and is not used in what follows.

For each iIi\in I, let EiE_{i} be a vector bundle on XiX_{i}; since XiX_{i} is discrete, this simply amounts to giving a vector space Ei(x)E_{i}(x) for each xXix\in X_{i}. To keep this discussion less technical, we assume that each Ei(x)E_{i}(x) is finite dimensional. Suppose that the dual bundles {Ei}iI\{E_{i}^{*}\}_{i\in I} have the structure of an inverse system of vector bundles, and let EE^{*} be the inverse limit, which is a vector bundle on XX (in a loose sense; it may not be locally trivial). For a point x={xi}iIx=\{x_{i}\}_{i\in I} of XX, the fiber E(x)E^{*}(x) is the inverse limit of the vector spaces Ei(xi)E^{*}_{i}(x_{i}). Define E(x)E(x) to be the corresponding direct limit; note that E(x)E^{*}(x) is the dual space of E(x)E(x). We say that xXx\in X is good if there exists i0Ii_{0}\in I such that the transition map Ei(xi)Ej(xj)E_{i}(x_{i})\to E_{j}(x_{j}) is injective for all i0iji_{0}\leq i\leq j.

Suppose now that EE^{*} is endowed with a GG-equivariant structure. Thus for gGg\in G and xXx\in X we have linear isomorphisms g:E(x)E(gx)g\colon E^{*}(x)\to E^{*}(gx) and g:E(x)E(gx)g\colon E(x)\to E(gx) that satisfy the cocycle conditions. As in the previous section, we assume the map g:Eg(E)g\colon E^{*}\to g^{*}(E^{*}) is uniformly continuous. Let x={xi}iIx=\{x_{i}\}_{i\in I} be a good point. We claim that E(x)E(x) is naturally a representation of the generalized stabilizer Γx\Gamma_{x}. Indeed, suppose gΓx(i)g\in\Gamma_{x}(i), so that πi(g1x)=πi(x)\pi_{i}(g^{-1}x)=\pi_{i}(x). We have a (likely non-commutative) diagram

Ei(xi)\textstyle{E_{i}(x_{i})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}α\scriptstyle{\alpha}β\scriptstyle{\beta}E(x)\textstyle{E(x)}E(g1x)\textstyle{E(g^{-1}x)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}g\scriptstyle{g}

Assuming ii is large enough, α\alpha is an inclusion. For x=α(y)x=\alpha(y), we define gxgx to be the element gβ(y)g\beta(y). One easily verifies that this is independent of ii, and defines the structure of a Γ\Gamma-representation on E(x)E(x).

8. Generalized stabilizers on 𝐆𝐋\mathbf{GL}-varieties

In this final section, we study the generalized stabilizer Γx\Gamma_{x} of a point xx on a 𝐆𝐋\mathbf{GL}-variety XX. Our main result provides an equivalence between the category of polynomial representations of Γx\Gamma_{x} and the category of KK-modules when xx is a 𝐆𝐋\mathbf{GL}-generic point on X=𝐀λ¯X=\mathbf{A}^{\underline{\smash{\lambda}}}. This yields the statements in §1.2, as the corresponding statements for ModK\operatorname{Mod}_{K} have already been established.

8.1. Generalized stabilizers on 𝐆𝐋\mathbf{GL}-varieties

Let X=Spec(R)X=\operatorname{Spec}(R) be an irreducible affine 𝐆𝐋\mathbf{GL}-variety over the field kk. Let Rn=R{kn}R_{n}=R\{k^{n}\} be the ring obtained by evaluating RR on knk^{n} and let Xn=Spec(Rn)X_{n}=\operatorname{Spec}(R_{n}), a finite dimensional variety over kk. Then X(k)X(k) is the inverse limit of the Xn(k)X_{n}(k) in the category of sets. Let πn:X(k)Xn(k)\pi_{n}\colon X(k)\to X_{n}(k) be the natural map. Given g𝐆𝐋g\in\mathbf{GL} and xX(k)x\in X(k), we see that πn(gx)\pi_{n}(gx) can be obtained from the image of xx in X{g1kn}X\{g^{-1}k^{n}\} by applying gg. Thus if mnm\geq n is such that g1knkmg^{-1}k^{n}\subset k^{m}, then one can recover πn(gx)\pi_{n}(gx) from πm(x)\pi_{m}(x). This shows that the action of gg is uniformly continuous, as described in §7.4.

Fix a point xX(k)x\in X(k). Let Γx\Gamma_{x} be its generalized stabilizer for the action of 𝐆𝐋\mathbf{GL} on X(k)X(k). Thus Γx(n)\Gamma_{x}(n) is the set of elements g𝐆𝐋g\in\mathbf{GL} such that g1xg^{-1}x and xx have the same image in Xn(k)X_{n}(k). Letting 𝔪R\mathfrak{m}\subset R be the defining ideal of xx, we see that Γx(n)\Gamma_{x}(n) can also be described as the set of elements g𝐆𝐋g\in\mathbf{GL} such that g1𝔪Rn=𝔪Rng^{-1}\mathfrak{m}\cap R_{n}=\mathfrak{m}\cap R_{n}.

We say that a representation VV of Γx\Gamma_{x} is polynomial if there is a polynomial representation WW of 𝐆𝐋\mathbf{GL} such that VV is isomorphic to a subquotient of WW (regarded as a representation of Γx\Gamma_{x}). We write Reppol(Γx)\operatorname{Rep}^{\mathrm{pol}}(\Gamma_{x}) for the category of polynomial representations of Γx\Gamma_{x}. It is a Grothendieck abelian category that is closed under tensor products.

Remark 8.1.

One can also define a notion of algebraic representation of Γx\Gamma_{x} by using restrictions of algebraic representations of 𝐆𝐋\mathbf{GL} (as defined in, e.g., [SS3, §3.1.1]). In many cases, polynomial and algebraic representations coincide. We therefore confine our attention to the polynomial case. ∎

8.2. From modules to representations

Maintain the above setup. The following proposition is the key result that justifies our definitions:

Proposition 8.2.

Let VV and WW be polynomial representations of 𝐆𝐋\mathbf{GL} and let φ:RVRW\varphi\colon R\otimes V\to R\otimes W be a map of RR-modules. Then the linear map φx:VW\varphi_{x}\colon V\to W obtained by reducing φ\varphi modulo 𝔪\mathfrak{m} is a map of Γx\Gamma_{x}-representations.

Proof.

Let vVv\in V be given. Let nn be such that vv is invariant under G(n)G(n). We claim that φx(gv)=gφx(v)\varphi_{x}(gv)=g\varphi_{x}(v) for gΓx(n)g\in\Gamma_{x}(n), which will complete the proof. Thus let gΓx(n)g\in\Gamma_{x}(n) be given. Write φ(1v)=i=1rfiwi\varphi(1\otimes v)=\sum_{i=1}^{r}f_{i}\otimes w_{i} with fiRf_{i}\in R and wiWw_{i}\in W. Then we have

φx(gv)=i=1rfi(g1x)gwi,gφx(v)=i=1rfi(x)gwi,\varphi_{x}(gv)=\sum_{i=1}^{r}f_{i}(g^{-1}x)gw_{i},\qquad g\varphi_{x}(v)=\sum_{i=1}^{r}f_{i}(x)gw_{i},

so it is enough to show that fi(g1x)=fi(x)f_{i}(g^{-1}x)=f_{i}(x) for each 1ir1\leq i\leq r. Since vv is G(n)G(n)-invariant, so is fif_{i}; in other words, fiRnf_{i}\in R_{n}. We thus see that fifi(x)f_{i}-f_{i}(x) belongs to 𝔪Rn\mathfrak{m}\cap R_{n}. By definition of Γx\Gamma_{x}, we have g1𝔪Rn=𝔪Rng^{-1}\mathfrak{m}\cap R_{n}=\mathfrak{m}\cap R_{n}, and so fifi(x)f_{i}-f_{i}(x) belongs to g1𝔪g^{-1}\mathfrak{m}. This exactly means that fifi(x)f_{i}-f_{i}(x) vanishes at g1xg^{-1}x, i.e., fi(g1x)=fi(x)f_{i}(g^{-1}x)=f_{i}(x). This verifies the claim. ∎

We now suppose that xx is 𝐆𝐋\mathbf{GL}-generic; if it is not, one can simply replace XX with the orbit closure of xx. The following proposition is our main construction of Γx\Gamma_{x}-representations:

Proposition 8.3.

There exists a unique right exact functor

Ψ~x:ModRReppol(Γx)\tilde{\Psi}_{x}\colon\operatorname{Mod}_{R}\to\operatorname{Rep}^{\mathrm{pol}}(\Gamma_{x})

satisfying the following two conditions:

  1. (a)

    We have Ψ~x(M)=M/𝔪M\tilde{\Psi}_{x}(M)=M/\mathfrak{m}M as vector spaces (and similarly for morphisms).

  2. (b)

    If VV is a polynomial representation then the Γx\Gamma_{x}-action on Ψ~x(RV)V\tilde{\Psi}_{x}(R\otimes V)\cong V is the restriction of the 𝐆𝐋\mathbf{GL} action.

The functor Ψ~x\tilde{\Psi}_{x} is exact and kills the torsion category, and thus induces a functor

Ψx:ModKReppol(Γx).\Psi_{x}\colon\operatorname{Mod}_{K}\to\operatorname{Rep}^{\mathrm{pol}}(\Gamma_{x}).

The functor Ψx\Psi_{x} is exact, cocontinuous, faithful, kk-linear, and naturally symmetric monoidal.

Proof.

Let MM be an RR-module. Choose a presentation

RV\textstyle{R\otimes V\ignorespaces\ignorespaces\ignorespaces\ignorespaces}φ\scriptstyle{\varphi}RW\textstyle{R\otimes W\ignorespaces\ignorespaces\ignorespaces\ignorespaces}M\textstyle{M\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

where VV and WW are polynomial representations. Applying RR/𝔪-\otimes_{R}R/\mathfrak{m}, we obtain a sequence

V\textstyle{V\ignorespaces\ignorespaces\ignorespaces\ignorespaces}φx\scriptstyle{\varphi_{x}}W\textstyle{W\ignorespaces\ignorespaces\ignorespaces\ignorespaces}M/𝔪M\textstyle{M/\mathfrak{m}M\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0.\textstyle{0.}

By Proposition 8.2, the first map is one of Γx\Gamma_{x}-representations. It follows that M/𝔪MM/\mathfrak{m}M inherits the structure of a Γx\Gamma_{x}-representation, which is easily seen to be independent of the choice of presentation. This representation is polynomial since it is a quotient of WW. One easily sees that this construction defines a right-exact functor

Ψ~x:ModRReppol(Γx),MM/𝔪M.\tilde{\Psi}_{x}\colon\operatorname{Mod}_{R}\to\operatorname{Rep}^{\mathrm{pol}}(\Gamma_{x}),\qquad M\mapsto M/\mathfrak{m}M.

It is clear that (a) and (b) hold. The uniqueness of Ψ~x\tilde{\Psi}_{x} follows from the fact that it is right-exact and determined on the category of projective RR-modules by (a) and (b).

Since xx is 𝐆𝐋\mathbf{GL}-generic, MM is flat at 𝔪\mathfrak{m} (Corollary 3.7), and so Ψ~x\tilde{\Psi}_{x} is exact. It is clear that Ψ~x\tilde{\Psi}_{x} kills the torsion subcategory. It thus factors through the generic category, which is equivalent to ModK\operatorname{Mod}_{K}. We therefore obtain a functor Ψx\Psi_{x} as in the statement of the proposition. Of course, ignoring the representation structure, Ψx\Psi_{x} is just the fiber functor Φx\Phi_{x} we constructed in §6.2. In other words, the diagram

ModK\textstyle{\operatorname{Mod}_{K}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ψx\scriptstyle{\Psi_{x}}Φx\scriptstyle{\Phi_{x}}Reppol(Γx)\textstyle{\operatorname{Rep}^{\mathrm{pol}}(\Gamma_{x})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Veck\textstyle{\mathrm{Vec}_{k}}

commutes, where the vertical arrow is the forgetful functor. It follows that Ψx\Psi_{x} is exact, cocontinuous, faithful, and kk-linear; moreover, one easily sees that the symmetric monoidal structure on Φx\Phi_{x} respects the Γx\Gamma_{x}-structure, and so Ψx\Psi_{x} is naturally symmetric monoidal as well. ∎

We expect that Ψx\Psi_{x} is an equivalence in general. In the remainder of this section, we prove this when KK is a rational 𝐆𝐋\mathbf{GL}-field (Theorem 8.11) and kk is algebraically clsoed.

Remark 8.4.

The above construction is essentially a special case of the one from §7.5, as we can regard Spec(Sym(M))\operatorname{Spec}(\operatorname{Sym}(M)) as a vector bundle (loosely interpreted) over XX. ∎

Remark 8.5.

Let MM be a submodule of Rλ¯R^{\oplus\underline{\smash{\lambda}}}, and let V=M/𝔪Mkλ¯V=M/\mathfrak{m}M\subset k^{\oplus\underline{\smash{\lambda}}}. It is easy to see that VV is a weak subrepresentation of kλ¯k^{\oplus\underline{\smash{\lambda}}}. Indeed, let =Spec(Sym(Rλ¯/M))\mathcal{E}=\operatorname{Spec}(\operatorname{Sym}(R^{\oplus\underline{\smash{\lambda}}}/M)), a closed 𝐆𝐋\mathbf{GL}-subscheme of the vector bundle X×(kλ¯)X\times(k^{\oplus\underline{\smash{\lambda}}})^{*}, and (x)=V\mathcal{E}(x)=V^{\perp}. Suppose α(x)\alpha\in\mathcal{E}(x) and {gi}\{g_{i}\} is a Γx\Gamma_{x}-sequence such that giαg_{i}\alpha converges to β\beta in (kλ¯)(k^{\oplus\underline{\smash{\lambda}}})^{*}. Since gixg_{i}x converges to xx, it follows that gi(x,α)g_{i}(x,\alpha) converges in X×(kλ¯)X\times(k^{\oplus\underline{\smash{\lambda}}})^{*} to (x,β)(x,\beta). Since each gi(x,α)g_{i}(x,\alpha) belongs to \mathcal{E} and \mathcal{E} is closed, we see that β(x)\beta\in\mathcal{E}(x). This verifies the claim.

We had originally defined a Γx\Gamma_{x}-representation to be a pair (V,kλ¯)(V,k^{\oplus\underline{\smash{\lambda}}}) consisting of a polynomial representation kλ¯k^{\oplus\underline{\smash{\lambda}}} and a weak subrepresentation VV. This can be made to work, thanks to the above proposition. However, it is not a good definition since we really just want the space VV; the ambient representation kλ¯k^{\oplus\underline{\smash{\lambda}}} is extrinsic. (Also, it is not immediately clear that this definition yields an abelian category.) It took some time for us to realize that the data intrinsic to VV is that of a Γx\Gamma_{x}-pre-representation, as in Definition 7.2. ∎

8.3. From representations to modules

We assume for the remainder of §8 that kk is algebraically closed. Fix a pure tuple σ¯\underline{\smash{\sigma}}, put R=Sym(kσ¯)R=\operatorname{Sym}(k^{\oplus\underline{\smash{\sigma}}}), put X=Spec(R)=𝐀σ¯X=\operatorname{Spec}(R)=\mathbf{A}^{\underline{\smash{\sigma}}}, and let K=Frac(R)K=\operatorname{Frac}(R). Fix a 𝐆𝐋\mathbf{GL}-generic kk-point xx of XX, and let 𝔪R\mathfrak{m}\subset R be its defining ideal. The goal of this subsection is to prove the following proposition, which is the key to the proof of Theorem 8.11.

Proposition 8.6.

Let μ¯\underline{\smash{\mu}} be a tuple and let VV be a subspace of kμ¯k^{\oplus\underline{\smash{\mu}}}. The following are equivalent:

  1. (a)

    There is an RR-submodule MM of Rμ¯R^{\oplus\underline{\smash{\mu}}} such that V=M/𝔪MV=M/\mathfrak{m}M.

  2. (b)

    The space VV is a Γx\Gamma_{x}-subrepresentation of kμ¯k^{\oplus\underline{\smash{\mu}}}.

  3. (c)

    The space VV is a weak Γx\Gamma_{x}-subrepresentation of kμ¯k^{\oplus\underline{\smash{\mu}}}.

We have already seen that (a) implies (b) (Proposition 8.3), and that (b) implies (c) (Proposition 7.4), so it suffices to prove that (c) implies (a). This will take the remainder of the subsection.

We use the theory of systems of variables from [BDES, §9.1]. We say that a kk-point of 𝐀λ¯\mathbf{A}^{\underline{\smash{\lambda}}} is degenerate if it is not 𝐆𝐋\mathbf{GL}-generic, and non-degenerate otherwise. For a single partition λ\lambda, the degenerate points in 𝐀λ(k)\mathbf{A}^{\lambda}(k) form a kk-subspace [BDES, Proposition 9.2]. A system of λ\lambda-variables is a set of points in 𝐀λ(k)\mathbf{A}^{\lambda}(k) that forms a basis modulo the subspace of degenerate elements. A system of variables is a choice of system of λ\lambda-variables for all λ\lambda.

Lemma 8.7.

Let μ¯\underline{\smash{\mu}} and ν¯\underline{\smash{\nu}} be pure tuples, let p𝐀μ¯(k)p\in\mathbf{A}^{\underline{\smash{\mu}}}(k) be 𝐆𝐋\mathbf{GL}-generic, and let E𝐀ν¯E\subset\mathbf{A}^{\underline{\smash{\nu}}} be the set of kk-points qq such that (q,p)𝐀ν¯×𝐀μ¯(q,p)\in\mathbf{A}^{\underline{\smash{\nu}}}\times\mathbf{A}^{\underline{\smash{\mu}}} is 𝐆𝐋\mathbf{GL}-generic. Then EE is Zariski dense in 𝐀ν¯\mathbf{A}^{\underline{\smash{\nu}}}.

Proof.

A point is non-degenerate if and only if each homogeneous piece of it is non-degenerate [BDES, Proposition 9.3]. It thus suffices to prove the lemma when μ¯\underline{\smash{\mu}} and ν¯\underline{\smash{\nu}} are composed of partitions of some constant size dd. First suppose that d=1d=1. Then a point is non-degenerate if its components are linearly independent. We can clearly choose qq such that the components of (q,p)(q,p) are independent while at the same time realizing arbitary values at finitely many coordinates of qq. Since any non-zero function ff on 𝐀ν¯\mathbf{A}^{\underline{\smash{\nu}}} uses only finitely many coordinates, it follows that we can choose qEq\in E such that f(q)0f(q)\neq 0. Thus EE is Zariski dense.

The case when d>1d>1 is similar. The set EE is non-empty: we can choose a system of variables that includes the components of pp, and then take the components of qq to be other elements from the system. Let qEq\in E. Then we can find a degenerate kk-point rr of 𝐀ν¯\mathbf{A}^{\underline{\smash{\nu}}} realizing arbitary values at finitely many coordiantes. It follows that q+rEq+r\in E also realizes arbitrary values at these coordinates, and so again EE is Zariski dense. ∎

Lemma 8.8.

Let μ¯\underline{\smash{\mu}} be a tuple and let pp be a kk-point of 𝐀μ¯\mathbf{A}^{\underline{\smash{\mu}}}. Then there exists a pure tuple ν¯\underline{\smash{\nu}}, a kk-point qq of 𝐀ν¯\mathbf{A}^{\underline{\smash{\nu}}} such that (q,x)𝐀ν¯×X(q,x)\in\mathbf{A}^{\underline{\smash{\nu}}}\times X is 𝐆𝐋\mathbf{GL}-generic, and a map of 𝐆𝐋\mathbf{GL}-varieties f:𝐀ν¯×X𝐀μ¯×Xf\colon\mathbf{A}^{\underline{\smash{\nu}}}\times X\to\mathbf{A}^{\underline{\smash{\mu}}}\times X over XX such that f(q,x)=(y,x)f(q,x)=(y,x).

Proof.

Write σ¯=[σ1,,σr]\underline{\smash{\sigma}}=[\sigma_{1},\ldots,\sigma_{r}] and let x=(x1,,xr)x=(x_{1},\ldots,x_{r}) be the components of xx. Pick a system of variables including x1,,xrx_{1},\ldots,x_{r}. By [BDES, Theorem 9.5], there exists a pure tuple λ¯=[λ1,,λs]\underline{\smash{\lambda}}=[\lambda_{1},\ldots,\lambda_{s}] and a map of 𝐆𝐋\mathbf{GL}-varieties g:𝐀λ¯𝐀μ¯g\colon\mathbf{A}^{\underline{\smash{\lambda}}}\to\mathbf{A}^{\underline{\smash{\mu}}} such that p=g(ξ1,,ξs)p=g(\xi_{1},\ldots,\xi_{s}), where ξ1,,ξs\xi_{1},\ldots,\xi_{s} are distinct elements from the system of variables. Now, after applying a permutation, we can assume that ξi=xi\xi_{i}=x_{i} for 1it1\leq i\leq t and the remaining ξi\xi_{i} and xjx_{j} are distinct. Let ν¯=[λt+1,,λs]\underline{\smash{\nu}}=[\lambda_{t+1},\ldots,\lambda_{s}] and q=(ξt+1,,ξs)𝐀ν¯q=(\xi_{t+1},\ldots,\xi_{s})\in\mathbf{A}^{\underline{\smash{\nu}}}. Now, let ff be the composition

𝐀ν¯×X\textstyle{\mathbf{A}^{\underline{\smash{\nu}}}\times X\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Δ\scriptstyle{\Delta}𝐀λ¯×X\textstyle{\mathbf{A}^{\underline{\smash{\lambda}}}\times X\ignorespaces\ignorespaces\ignorespaces\ignorespaces}g×idX\scriptstyle{g\times\mathrm{id}_{X}}𝐀μ¯×X\textstyle{\mathbf{A}^{\underline{\smash{\mu}}}\times X}

where Δ\Delta is the diagonal map that copies the first tt coordinates of XX into those of 𝐀λ¯\mathbf{A}^{\underline{\smash{\lambda}}}. Then Δ(q,x)=(ξ1,,ξs,x1,,xr)\Delta(q,x)=(\xi_{1},\ldots,\xi_{s},x_{1},\ldots,x_{r}), and so f(q,x)=(p,x)f(q,x)=(p,x). By construction (q,x)(q,x) is 𝐆𝐋\mathbf{GL}-generic. ∎

Given vector spaces VUV\subset U, we let VV^{\perp} be the annihilator of VV in the dual space UU^{*}.

Lemma 8.9.

Let μ¯\underline{\smash{\mu}} be a tuple, let VV be a weak Γx\Gamma_{x}-subrepresentation of kμ¯k^{\oplus\underline{\smash{\mu}}}, and let pVp\in V^{\perp}. Then there exists a tuple ν¯\underline{\smash{\nu}} and a map of 𝐆𝐋\mathbf{GL}-varieties f:𝐀ν¯×X𝐀μ¯×Xf\colon\mathbf{A}^{\underline{\smash{\nu}}}\times X\to\mathbf{A}^{\underline{\smash{\mu}}}\times X over XX such that im(fx)\operatorname{im}(f_{x}) contains pp and is contained in VV^{\perp}.

Proof.

Note that VV^{\perp} is a subspace of (kμ¯)=𝐀μ¯(k^{\oplus\underline{\smash{\mu}}})^{*}=\mathbf{A}^{\underline{\smash{\mu}}}. Applying Lemma 8.8, there exists a pure tuple ν¯\underline{\smash{\nu}}, a kk-point qq of 𝐀ν¯\mathbf{A}^{\underline{\smash{\nu}}} such that (q,x)(q,x) is 𝐆𝐋\mathbf{GL}-generic in 𝐀ν¯×X\mathbf{A}^{\underline{\smash{\nu}}}\times X, and a map of 𝐆𝐋\mathbf{GL}-varieties f:𝐀ν¯×X𝐀μ¯×Xf\colon\mathbf{A}^{\underline{\smash{\nu}}}\times X\to\mathbf{A}^{\underline{\smash{\mu}}}\times X over XX such that f(q,x)=(p,x)f(q,x)=(p,x). Thus pim(fx)p\in\operatorname{im}(f_{x}).

Now, let qq^{\prime} be a kk-point of 𝐀ν¯\mathbf{A}^{\underline{\smash{\nu}}} such that (q,x)(q^{\prime},x) is 𝐆𝐋\mathbf{GL}-generic. We claim that fx(q)Vf_{x}(q^{\prime})\in V^{\perp}. Since (q,x)(q,x) is 𝐆𝐋\mathbf{GL}-generic there is a sequence {gi}i1\{g_{i}\}_{i\geq 1} in 𝐆𝐋\mathbf{GL} such that gi(q,x)g_{i}(q,x) converges to (q,x)(q^{\prime},x) in the Π\Pi-topology (Proposition 2.2). We thus see that gixg_{i}x converges to xx in the Π\Pi-topology, and so {gi}\{g_{i}\} is a Γx\Gamma_{x}-sequence. Applying ff, we see that gipg_{i}p converges to f(q)f(q^{\prime}). Since VV is a weak subrepresentation, this implies that f(q)Vf(q^{\prime})\in V^{\perp}, as claimed.

Now, let EE be the set of kk-points q𝐀ν¯q^{\prime}\in\mathbf{A}^{\underline{\smash{\nu}}} such that (q,x)(q^{\prime},x) is 𝐆𝐋\mathbf{GL}-generic. By the previous paragraph, we see that fx(E)Vf_{x}(E)\subset V^{\perp}. Since EE is Zariski dense in 𝐀ν¯\mathbf{A}^{\underline{\smash{\nu}}} by Lemma 8.7 and VV^{\perp} is a Zariski closed subset of 𝐀μ¯\mathbf{A}^{\underline{\smash{\mu}}}, it follows that im(fx)V\operatorname{im}(f_{x})\subset V^{\perp}, as required. ∎

Lemma 8.10.

Let μ¯\underline{\smash{\mu}} and ν¯\underline{\smash{\nu}} be tuples, and let f:𝐀ν¯×X𝐀μ¯×Xf\colon\mathbf{A}^{\underline{\smash{\nu}}}\times X\to\mathbf{A}^{\underline{\smash{\mu}}}\times X be a map of 𝐆𝐋\mathbf{GL}-varieties over XX. Then there exists a closed 𝐆𝐋\mathbf{GL}-subvariety YY of 𝐀μ¯×X\mathbf{A}^{\underline{\smash{\mu}}}\times X such that the following two conditions hold:

  1. (a)

    YY is defined by fiberwise linear equations, that is, Y=Spec(Sym(M))Y=\operatorname{Spec}(\operatorname{Sym}(M)) for some RR-module quotient MM of Rμ¯R^{\oplus\underline{\smash{\mu}}}

  2. (b)

    the kk-subspace YxY_{x} of 𝐀μ¯\mathbf{A}^{\underline{\smash{\mu}}} is exactly the Π\Pi-closure of the kk-span of im(fx)\operatorname{im}(f_{x}).

Proof.

First suppose that ff is fiberwise linear. This means that ff is induced from a map of RR-modules g:Rμ¯Rν¯g\colon R^{\oplus\underline{\smash{\mu}}}\to R^{\oplus\underline{\smash{\nu}}}. Let MM be the image of gg, and let Y=Spec(Sym(M))Y=\operatorname{Spec}(\operatorname{Sym}(M)). Let gx:kμ¯kν¯g_{x}\colon k^{\oplus\underline{\smash{\mu}}}\to k^{\oplus\underline{\smash{\nu}}} be the map obtained by reducing gg modulo the maximal ideal 𝔪\mathfrak{m}. Since coker(g)\operatorname{coker}(g) is flat at xx (Corollary 3.7), it follows that the image of gxg_{x} is M/𝔪MM/\mathfrak{m}M. As fxf_{x} is the dual of gxg_{x}, we see that its image is the dual of M/𝔪MM/\mathfrak{m}M, which is exactly YxY_{x}. This completes the proof in the linear case. (In this case, taking the Π\Pi-closure is not necessary.)

We now treat the general case. The map ff corresponds to a map of RR-algebras g:RSym(kμ¯)RSym(kν¯)g\colon R\otimes\operatorname{Sym}(k^{\oplus\underline{\smash{\mu}}})\to R\otimes\operatorname{Sym}(k^{\oplus\underline{\smash{\nu}}}). The image of kμ¯k^{\oplus\underline{\smash{\mu}}} under this map is contained in RSymd(kν¯)R\otimes\operatorname{Sym}^{\leq d}(k^{\oplus\underline{\smash{\nu}}}) for some dd, where Symd=i=0dSymi\operatorname{Sym}^{\leq d}=\bigoplus_{i=0}^{d}\operatorname{Sym}^{i}. The map gg then factors as

RSym(kμ¯)\textstyle{R\otimes\operatorname{Sym}(k^{\oplus\underline{\smash{\mu}}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}g1\scriptstyle{g_{1}}RSym(Symd(kν¯))\textstyle{R\otimes\operatorname{Sym}(\operatorname{Sym}^{\leq d}(k^{\oplus\underline{\smash{\nu}}}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}g2\scriptstyle{g_{2}}RSym(kν¯),\textstyle{R\otimes\operatorname{Sym}(k^{\otimes\underline{\smash{\nu}}}),}

where g1g_{1} is linear (i.e., induced from a map of RR-modules). Let f=f1f2f=f_{1}\circ f_{2} be the corresponding factorization of ff. Let Y𝐀μ¯×XY\subset\mathbf{A}^{\underline{\smash{\mu}}}\times X be the subvariety provided by the linear case, applied to f1f_{1}. The map fxf_{x} factors as

(kν¯)\textstyle{(k^{\oplus\underline{\smash{\nu}}})^{*}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f2,x\scriptstyle{f_{2,x}}(Symd(kν¯))\textstyle{(\operatorname{Sym}^{\leq d}(k^{\oplus\underline{\smash{\nu}}}))^{*}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f1,x\scriptstyle{f_{1,x}}(kν¯)\textstyle{(k^{\oplus\underline{\smash{\nu}}})^{*}}

We know that the image of f1,xf_{1,x} is exactly YxY_{x}. The map f2,xf_{2,x} is the canonical map, taking aa to (1,a,,ad)(1,a,\ldots,a^{d}). One easily sees that the kk-span of the image of f2,xf_{2,x} is Π\Pi-dense. Since f1,xf_{1,x} is Π\Pi-continuous, the result follows. ∎

Proof of Proposition 8.6.

Let VV be a weak Γx\Gamma_{x}-subrepresentation of kμ¯k^{\oplus\underline{\smash{\mu}}}, and fix an element vVv\in V^{\perp}. By Lemma 8.9, we can find a tuple ν¯\underline{\smash{\nu}} and a map of 𝐆𝐋\mathbf{GL}-varieties f:𝐀ν¯×X𝐀μ¯×Xf\colon\mathbf{A}^{\underline{\smash{\nu}}}\times X\to\mathbf{A}^{\underline{\smash{\mu}}}\times X over XX such that im(fx)\operatorname{im}(f_{x}) contains vv and is contained in VV. By Lemma 8.10, there is an RR-module M=Rμ¯/NM=R^{\oplus\underline{\smash{\mu}}}/N such that (M/𝔪M)(M/\mathfrak{m}M)^{*} is the Π\Pi-closure of the span of im(fx)\operatorname{im}(f_{x}). Since VV^{\perp} is Π\Pi-closed, it follows that (M/𝔪M)(M/\mathfrak{m}M)^{*} is contained in VV^{\perp}; of course, it also contains vv. We thus see that N/𝔪NN/\mathfrak{m}N contains VV and is contained in ker(v)\ker(v).

Now, let {vi}iU\{v_{i}\}_{i\in U} be a basis for VV^{\perp}, and for each ii pick a submodule NiN_{i} of Rμ¯R^{\oplus\underline{\smash{\mu}}} as in the previous paragraph, so that Ni/𝔪NiN_{i}/\mathfrak{m}N_{i} contains VV and is contained in ker(vi)\ker(v_{i}). For a finite subset II of UU, let NI=iINiN_{I}=\bigcap_{i\in I}N_{i}. The NIN_{I} form a descending family of submodules of Rμ¯R^{\oplus\underline{\smash{\mu}}}. Since Rμ¯R^{\oplus\underline{\smash{\mu}}} is an artinian object in the generic category (Theorem 4.10(a)), it follows that there is some finite subset JJ such that NJ/NIN_{J}/N_{I} is torsion for all JIJ\subset I. We thus have NJ/𝔪NJ=NI/𝔪NIN_{J}/\mathfrak{m}N_{J}=N_{I}/\mathfrak{m}N_{I} for all such II. It follows that NJ/𝔪NJN_{J}/\mathfrak{m}N_{J} is contains VV and is contained in iUker(vi)=V\bigcap_{i\in U}\ker(v_{i})=V. This completes the proof. ∎

8.4. The main theorem

Maintain the setup from §8.3. The following is our main theorem on representations of Γx\Gamma_{x}:

Theorem 8.11.

The functor Ψx:ModKReppol(Γx)\Psi_{x}\colon\operatorname{Mod}_{K}\to\operatorname{Rep}^{\mathrm{pol}}(\Gamma_{x}) is an equivalence.

From the theorem, we see that all properties of ModK\operatorname{Mod}_{K} transer to Reppol(Γx)\operatorname{Rep}^{\mathrm{pol}}(\Gamma_{x}). This yields the statements of §1.2. (We note that in the setting of §1.2, there is no distinction between algebraic and polynomial representation.) Before proving the theorem, we require a lemma.

Lemma 8.12.

Let VV be a KK-module. Then the map

α:{K-submodules of V}{Γx-subrepresentations of Ψx(V)}\alpha\colon\{\text{$K$-submodules of $V$}\}\to\{\text{$\Gamma_{x}$-subrepresentations of $\Psi_{x}(V)$}\}

induced by Ψx\Psi_{x} is an isomorphism of partially ordered sets.

Proof.

We first show that α\alpha is injective. First suppose that UWU\subset W are KK-submodules of VV and α(U)=α(W)\alpha(U)=\alpha(W). Then the containment of RR-modules UpolWpolU^{\mathrm{pol}}\subset W^{\mathrm{pol}} induces an isomorphism modulo 𝔪\mathfrak{m}. It follows that Wpol/UpolW^{\mathrm{pol}}/U^{\mathrm{pol}} has vanishing fiber at 𝔪\mathfrak{m}, and thus vanishes (see the proof of Proposition 6.2). Hence Upol=WpolU^{\mathrm{pol}}=W^{\mathrm{pol}}, and so U=WU=W. Now suppose that UU and WW are arbitary and α(U)=α(W)\alpha(U)=\alpha(W). Then α(U+W)=α(U)+α(W)=α(U)\alpha(U+W)=\alpha(U)+\alpha(W)=\alpha(U). Since UU+WU\subset U+W, the previous case shows that U=U+WU=U+W, and so WUW\subset U. By symmetry, we have UWU\subset W. Thus α\alpha is injective.

We now see that α\alpha is strictly order-preserving. Indeed, let UU and WW be KK-submodules of VV. If UWU\subset W then it is clear that α(U)α(W)\alpha(U)\subset\alpha(W). Conversely, if α(U)α(W)\alpha(U)\subset\alpha(W) then α(U+W)=α(U)+α(W)=α(W)\alpha(U+W)=\alpha(U)+\alpha(W)=\alpha(W), and so U+W=WU+W=W since α\alpha is injective, whence UWU\subset W.

To complete the proof, we must show that α\alpha is surjective. If V=KV0V=K\otimes V_{0} for a finite length polynomial representation V0V_{0}, then this follows from Proposition 8.6. Suppose now that V=KV0V=K\otimes V_{0} for an arbitrary polynomial representation V0V_{0}. Write V0=jJV0,jV_{0}=\bigcup_{j\in J}V_{0,j} where JJ is a directed set and V0,jV_{0,j} has finite length, and put Vj=KV0,jV_{j}=K\otimes V_{0,j}. Since Ψx\Psi_{x} is cocontinuous, we have Ψx(V)=jJΨx(Vj)\Psi_{x}(V)=\bigcup_{j\in J}\Psi_{x}(V_{j}). Let EE be a Γx\Gamma_{x}-subrepresentation of Ψx(V)\Psi_{x}(V), and put Ej=EΨx(Vj)E_{j}=E\cap\Psi_{x}(V_{j}). Since Reppol(Γx)\operatorname{Rep}^{\mathrm{pol}}(\Gamma_{x}) is a Grothendieck category, we have E=jJEjE=\bigcup_{j\in J}E_{j}. By the finite length case, we have Ej=α(Wj)E_{j}=\alpha(W_{j}) for a unique KK-submodule WjW_{j} of VV. Since α\alpha is strictly order-preserving, it follows that WjWkW_{j}\subset W_{k} if jkj\leq k. Thus the WjW_{j}’s form a directed system. Let W=jJWjW=\bigcup_{j\in J}W_{j}. Again, by the cocontinuity of Ψx\Psi_{x}, we have α(W)=jJα(Wj)=E\alpha(W)=\bigcup_{j\in J}\alpha(W_{j})=E.

Finally, suppose that VV is an arbitrary KK-module. Since ModK\operatorname{Mod}_{K} is a Grothendieck abelian category, VV embeds into an injective object II. Since ModK\operatorname{Mod}_{K} is locally noetherian (Theorem 4.10(a)), II is a direct sum of indecomposable injectives. Thus II has the form KV0K\otimes V_{0} for a polynomial representation V0V_{0} (Theorem 4.10(b)). Now, suppose that EE is a Γx\Gamma_{x}-subrepresentation of Ψx(V)\Psi_{x}(V). Since Ψx(V)Ψx(I)\Psi_{x}(V)\subset\Psi_{x}(I), the previous paragraph shows that E=α(W)E=\alpha(W) for some KK-submodule WW of II. Since α\alpha is strictly order preserving, it follows that WVW\subset V, which completes the proof. ∎

Proof of Theorem 8.11.

We first show that Ψx\Psi_{x} is essentially surjective. Thus let EE be a given polynomial representation of Γx\Gamma_{x}. By definition, there is some polynomial representation VV of 𝐆𝐋\mathbf{GL} and Γx\Gamma_{x}-subrepresentations E2E1VE_{2}\subset E_{1}\subset V such that EE1/E2E\cong E_{1}/E_{2}. By Lemma 8.12, there exist KK-submodules W2W1KVW_{2}\subset W_{1}\subset K\otimes V such that Ei=Ψx(Wi)E_{i}=\Psi_{x}(W_{i}). Thus EΨx(W2/W1)E\cong\Psi_{x}(W_{2}/W_{1}), and so Ψx\Psi_{x} is essentially surjective.

We now prove that Ψx\Psi_{x} is full. Let VV and WW be KK-modules and let f:Ψx(V)Ψx(W)f\colon\Psi_{x}(V)\to\Psi_{x}(W) be a map of Γx\Gamma_{x}-subrepresentations. Let EΨx(V)Ψx(W)E\subset\Psi_{x}(V)\oplus\Psi_{x}(W) be the graph of ff. By Lemma 8.12, we have E=Ψx(U)E=\Psi_{x}(U) for a unique KK-submodule UVWU\subset V\oplus W. The projection map UVU\to V becomes an isomorphism after applying Ψx\Psi_{x}, and is therefore an isomorphism since Ψx\Psi_{x} is exact and faithful. Thus UU is the graph of a morphism g:VWg\colon V\to W of KK-modules, and clearly f=Ψx(g)f=\Psi_{x}(g).

We have already seen that Ψx\Psi_{x} is faithful, and so it is an equivalence. ∎

References

  • [BDE] Arthur Bik, Jan Draisma, Rob H. Eggermont. Polynomials and tensors of bounded strength. Commun. Contemp. Math. 21(7) (2019), paper number 1850062. arXiv:1805.01816
  • [BDDE] Arthur Bik, Alessandro Danelon, Jan Draisma, Rob H. Eggermont. Universality of high-strength tensors. arXiv:2105.00016
  • [BDES] Arthur Bik, Rob H. Eggermont, Jan Draisma, Andrew Snowden. The geometry of polynomial representations. arXiv:2105.12621
  • [CEF] Thomas Church, Jordan S. Ellenberg, Benson Farb. FI-modules and stability for representations of symmetric groups. Duke Math. J. 164 (2015), no. 9, 1833–1910. arXiv:1204.4533v4
  • [Dr] Jan Draisma. Topological Noetherianity of polynomial functors. J. Amer. Math. Soc. 32(3) (2019), pp. 691–707. arXiv:1705.01419
  • [DES] Harm Derksen, Rob H. Eggermont, Andrew Snowden. Topological noetherianity for cubic polynomials. Algebra Number Theory 11(9) (2017), pp. 2197–2212. arXiv:1701.01849
  • [DPS] Elizabeth Dan-Cohen, Ivan Penkov, Vera Serganova. A Koszul category of representations of finitary Lie algebras. Adv. Math. 289 (2016), 250–278. arXiv:1105.3407v2
  • [KaZ1] David Kazhdan, Tamar Ziegler. On ranks of polynomials. Algebr. Represent. Theory 21 (2018), no. 5, 1017–1021. arXiv:1802.04984
  • [KaZ2] David Kazhdan, Tamar Ziegler. Properties of high rank subvarieties of affine spaces. Geom. Funct. Analysis 30 (2020), 1063–1096. arXiv:1902.00767
  • [KaZ3] David Kazhdan, Tamar Ziegler. On the codimension of the singular locus. arXiv:1907.11750
  • [KaZ4] David Kazhdan, Tamar Ziegler. Applications of Algebraic Combinatorics to Algebraic Geometry. arXiv:2005.12542
  • [GS] Dimitar Grantcharov, Vera Serganova. Tensor representations of 𝔮()\mathfrak{q}(\infty). Algebr. Represent. Theory 22 (2019), 867–893. arXiv:1605.02389v1
  • [Na] Rohit Nagpal. FI-modules and the cohomology of modular SnS_{n}-representations. Ph.D. thesis, University of Wisconsin, 2015. arXiv:1505.04294v1
  • [NS] Rohit Nagpal, Andrew Snowden. Semi-linear representations of the infinite symmetric group. arXiv:1909.08753
  • [NS2] Rohit Nagpal, Andrew Snowden. Symmetric subvarieties of infinite affine space. arXiv:2011.09009
  • [NSS] Rohit Nagpal, Steven V Sam, Andrew Snowden. Noetherianity of some degree two twisted commutative algebras. Selecta Math. (N.S.) 22 (2016), no. 2, 913–937. arXiv:1501.06925v2
  • [NSS2] Rohit Nagpal, Steven V Sam, Andrew Snowden. Noetherianity of some degree two twisted skew-commutative algebras. Selecta Math. (N.S.) 25 (2019), no. 1. arXiv:1610.01078
  • [NSS3] Rohit Nagpal, Steven V Sam, Andrew Snowden. On the geometry and representation theory of isomeric matrices. arXiv:2101.10422
  • [Ro] M. Rovinsky. On semilinear representations of the infinite symmetric group. Preprint, 2015. arXiv:1405.3265
  • [Ro2] M. Rovinsky. An analogue of Hilbert’s Theorem 90 for infinite symmetric groups. Preprint, 2017. arXiv:1508.02267
  • [Ro3] M. Rovinsky. Semilinear representations of symmetric groups and of automorphism groups of universal domains. Sel. Math. 24 (2018), no. 3, 2319–2349.
  • [PSe] Ivan Penkov, Vera Serganova. Categories of integrable sl()sl(\infty)-, o()o(\infty)-, sp()sp(\infty)-modules. Representation Theory and Mathematical Physics, Contemp. Math. 557, AMS 2011, pp. 335–357. arXiv:1006.2749v1
  • [PSt] Ivan Penkov, Konstantin Styrkas. Tensor representations of classical locally finite Lie algebras. Developments and trends in infinite-dimensional Lie theory, Progr. Math. 288, Birkhäuser Boston, Inc., Boston, MA, 2011, pp. 127–150. arXiv:0709.1525v1
  • [Se] Vera Serganova. Classical Lie superalgebras at infinity. Advances in Lie Superalgebras, Springer INdAM Series 7, (2014), 181–201.
  • [Sn] Andrew Snowden. The spectrum of a twisted commutative algebra. arXiv:2002.01152
  • [SS1] Steven V Sam, Andrew Snowden. GL-equivariant modules over polynomial rings in infinitely many variables. Trans. Amer. Math. Soc. 368 (2016), 1097–1158. arXiv:1206.2233v3
  • [SS2] Steven V Sam, Andrew Snowden. Introduction to twisted commutative algebras. Preprint. arXiv:1209.5122v1
  • [SS3] Steven V Sam, Andrew Snowden. Stability patterns in representation theory. Forum Math. Sigma 3 (2015), e11, 108 pp. arXiv:1302.5859v2
  • [SS4] Steven V Sam, Andrew Snowden. Gröbner methods for representations of combinatorial categories. J. Amer. Math. Soc. 30 (2017), 159–203. arXiv:1409.1670v3
  • [SS5] Steven V Sam, Andrew Snowden. GL-equivariant modules over polynomial rings in infinitely many variables. II. Forum Math., Sigma 7 (2019), e5, 71 pp. arXiv:1703.04516v1
  • [SS6] Steven V Sam, Andrew Snowden. Sp-equivariant modules over polynomial rings in infinitely many variables. arXiv:2002.03243
  • [SS7] Steven V Sam, Andrew Snowden. The representation theory of Brauer categories I: triangular categories. arXiv:2006.04328
  • [SSW] Steven V Sam, Andrew Snowden, Jerzy Weyman. Homology of Littlewood complexes. Selecta Math. (N.S.) 19 (2013), no. 3, 655–698. arXiv:1209.3509v2
  • [Stacks] Stacks Project, http://stacks.math.columbia.edu, 2021.