Stable representation theory:
beyond the classical groups
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.
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 be the space of symmetric trilinear forms (i.e., cubic polynomials) in variables. Also, let and . The set is the inverse limit of the sets , and as such carries the inverse limit topology. (Each is endowed with the discrete topology). Precisely, a sequence in converges to if for each we have for all sufficiently large .
The group acts on , and the group acts on . The group is, in a sense, too small111The group is much larger than , but it is also too small.. To remedy this, we employ a modification of the concept of orbit: we say that two elements of 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 is degenerate if it has the form where and , 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 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 be given. For , define be the set of elements such that . (The inverse here is simply to make some other definitions cleaner.) Note that is typically not a subgroup. We define the generalized stabilizer of to be the system . One should think of as a kind of germ of a neighborhood of the stabilizer of . For this reason, we refer to as a germinal subgroup; see Definition 7.1 for details. We view as an analog of the infinite orthogonal group associated to symmetric trilinear forms.
1.2. Representations of generalized stabilizers
Let and be as above. We define a representation of to be a complex vector space such that each finite dimensional subspace is endowed with an action map , for some depending on , satisfying certain conditions. A little more precisely, the data defining a representation can be encoded as a linear map
Thus for , one can regard as the germ of a function on , with respect to the system of neighborhoods .
Every representation of restricts to a representation of . We say that a representation of is algebraic if it occurs as a subquotient of the restriction of a polynomial representation of . In particular, the standard representaiton of restricts to an algebraic representation of , which we call the standard representation of . We let 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 when 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 and are finite length algebraic representations then is a finite dimensional complex vector space.
-
–
-
•
The simple algebraic representations are well-understood:
-
–
For each partition , there is a simple , and these exhaust the simples.
-
–
One can construct using a variant of Weyl’s traceless tensor construction. Let be the intersection of the kernels of the maps obtained by applying to three tensor factors. This space carries an action of , where denotes the symmetric group. The isotypic piece of corresponding the Specht module is exactly .
-
–
-
•
Algebraic representations are well-behaved homologically:
-
–
The representations are exactly the indecomposable injective algebraic representations; in fact, is the injective envelope of .
-
–
Every finite length algebraic representation has finite injective dimension.
-
–
-
•
There is a combinatorial description of the entire category : it is equivariant to a category of representations of a certain variant of the upwards Brauer category.
-
•
The category satisfies a universal property. Let be a -linear abelian category equipped with a tensor product. Then giving a left-exact symmetric monoidal -linear functor is equivalent to giving an object of equipped with a symmetric trilinear form. The notation here denotes the subcategory of finite length objects.
-
•
The symmetric monoidal category is independent of , up to equivalence.
1.3. Semi-linear representations: motivation
Recall that if a group acts on a field then a semi-linear representation of over is a -vector space equipped with an additive action of such that the equation holds, for , , and . Semi-linear representations will be a central topic in this paper. To motivate their appearance, we first examine a familiar case.
Let be the space of symmetric bilinear forms on , regarded as an algebraic variety; explicitly, where is the polynomial ring . Let be the open subvariety of consisting of non-degenerate forms. The algebraic group acts transitively on . Let be a closed point of , and let be its stabilizer. If is a -equivariant quasi-coherent sheaf on then its fiber at is an algebraic representation of , and this construction gives an equivalence of categories
In fact, we can get a similar equivalence using the generic point of . If is a -equivariant quasi-coherent sheaf on then its generic fiber is a semi-linear representation of over that is algebraic (in the sense that it is spanned by an algebraic subrepresentation). Moreover, letting be the category of such semi-linear representations, this construction defines an equivalence
Thus, combined with the previous equivalence, we obtain an equivalence
This gives us a way of studying representations of (or, at least, the representation category) even if we do not understand the group very well.
We adopt this approach in this paper to replace representations of with more familiar objects. Let be the infinite variable polynomial ring and let . We show (Theorem 8.11) that is equivalent to a certain category of semi-linear representations of over (precisely, the category of “-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 -algebra222In characteristic 0, -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 ; in other words, it is a commutative ring equipped with an action of under which it forms a polynomial representation. For example, the ring appearing above is a -algebra. If is a -algebra then an -module is a module object; in other words, it is a -equivariant -module that forms a polynomial representation.
A -field is a field equipped with an action of that can be obtained as the fraction field of an integral -algebra. If is a -field then a -module is a semi-linear representation of over that is generated by a polynomial subrepresentation. The basic example of a -module is . While is typically not projective, every -module is a quotient of a sum of ones of this form. We let denote the category of -modules. This is the fundamental object of study in this paper.
We prove two main technical results about -modules. To state the first one, we must introduce the shift operation. Let be the subgroup of consisting of block matrices of the form
where the top left block is . This group is isomorphic to . If is a set equipped with an action of , we define its th shift, denoted , to be the set equipped with the action of coming from restricting the given action to . The shift operation preserves all structure introduced so far (polynomial representations, -fields, etc.). Our first theorem is:
Theorem 1.1 (Shift theorem).
Let be a -field and let be a finitely generated -module. Then there exists and partitions such that is isomorphic to as a -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 -modules [Na]. The above shift theorem is closely related to the shift theorem for -varieties [BDES, Theorem 5.1], and follows a similar proof.
Our second main result about -modules is the following:
Theorem 1.2 (Embedding theorem).
Let be a rational -field, i.e., one of the form where is a finite length polynomial representation of , and let be a finitely generated -module. Then there exist partitions and an injection of -modules .
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 , and can be deduced from the embedding thoerem by comparitively easy arguments. The corresponding results for are deduced from those for .
1.5. Summary of categories
Let be the analog of the upwards Brauer category for symmetric trilinear forms (see §5.2), let , let , and let be non-degenerate. We show that the following categories are equivalent:
-
(a)
The category of -modules that are locally of finite length.
-
(b)
The category of -modules that are locally of finite length.
-
(c)
The generic category , i.e., the Serre quotient of by the subcategory of torsion modules.
-
(d)
The category .
-
(e)
The category .
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 -linear tensor categories. However, only in (e) are the objects -vector spaces (with extra structure), with the tensor product being the usual one on the underlying vector space. One can therefore view the equivalence as a fiber functor on . We thus get one such fiber functor for each choice of . 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 -equivariant algebraic geometry. These theories are based on -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.
- •
- •
- •
1.8. Further work
In this paper, we give a fairly complete description of when is a rational -field. While we do prove some results for more general -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 -algebras, it is also important to understand the generic categories when is a “-domain” (this means implies or when and are -ideals, which is a weaker condition than being a domain). In [Sn], we gave a useful way of understanding the -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:
-
(a)
How much of standard Lie theory can be carried over to the generalized stabilizers ? Is there a Dynkin diagram, Cartan matrix, Weyl group, etc.?
-
(b)
Is there a Tannakian perspective that allows one to recover the generalized stabilizer from the fiber functor ?
-
(c)
Prove Theorem 8.11 for general .
-
(d)
What are the derived specializations of simple objects of ? (See Remark 5.13.)
-
(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.
-
(f)
In this paper, we consider generalized stabilizers for actions of 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 -algebras and related concepts. In §3 we prove our two main technical theorems on -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 . In §6, we classify the fiber functors of . In §7 we introduce germinal subgroups and generalized stabilizers in the abstract. Finally, in §8, we apply these concepts to -varieties.
Acknowledgments
2. -equivariant algebra and geometry
In this section, we review background material on polynomial representations, -algebras, -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 of characteristic 0. Put , regarded as a discrete group. We let be the standard representation of . We say that a representation of on a -vector space is polynomial if it appears as a subquotient of a (possibly infinite) direct sum of tensor powers of . We let denote the category of polynomial representations. It is a semi-simple Grothendieck abelian category that is closed under tensor product.
For a partition , we let denote the corresponding Schur functor. The simple polynomial representations are exactly those of the form . Thus every polynomial representation decomposes as a (perhaps infinite) direct sum of ’s.
Every polynomial representation of carries a natural grading, with concentrated in degree , the size of the partition . This grading is compatible with tensor products: is concentrated in degree . The degree 0 piece of a polynomial representation is exactly the invariant subspace .
We now introduce some non-standard notation that will be convenient for working with these objects. We write in place of . More generally, for a -vector space we put ; note that if is a -algebra then is naturally a free -module. A tuple of partitions (often simply called a tuple) is a tuple , where each is a partition. We put , and define similarly. We say that 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 corresponding to the functor . Given a polynomial representation and a vector space , we let be the result of regarding as a polynomial functor and evaluating on . In the important special case where , we can identify with the invariant space , where is defined in §2.3. For example, if then .
2.2. The maximal polynomial subrepresentation
Suppose that is an arbitrary -linear representation of . We say that an element is polynomial if the subrepresentation it generates is a polynomial representation. We let be the set of all polynomial elements in . It can be characterized as the maximal polynomial subrepresentation of . Moreover, if denotes the category of all -linear representations of then is the right adjoint of the inclusion functor . As such, is left-exact and continuous; it is not exact.
2.3. The shift operation
Recall that is the subgroup of consisting of block matrices of the form
where the top left block has size . We have a group isomorphism
Given some kind of object equipped with an action of , we define its th shift, denoted , to be the same object but with acting through the self-embedding .
One easily sees that if is a polynomial representation of then is also such a representation. From the polynomial functor point of view, we have
If has finite length then so does . It follows that if is a tuple then there is another tuple, which we denote by , such that . If consists of a single partition, we write in place of . In this case, contains exactly once, and all other partitions in it are strictly smaller.
2.4. -algebras
A -algebra (over ) is a commutative algebra object in the tensor category ; thus, it is a commutative (and associative and unital) -algebra equipped with an action of the group by algebra automorphisms, under which it forms a polynomial representation. Let be a -algebra. By an -module we mean a module object in ). Explicitly, this is an ordinary -module equipped with a compatible action of under which forms a polynomial representation. We let denote the category of modules, which is easily seen to be a Grothendieck abelian category.
We say that is -generated (over ) by a set of elements if is generated as a -algebra by the orbits of these elements. We say that is finitely -generated if it is -generated by a finite set. We similarly speak of -generation for -modules.
We say that a -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 be an integral -algebra that is finitely -generated. Then there exists , a non-zero -invariant element , and an isomorphism for some finitely generated integral -algebra (with trivial -action) and pure tuple .
Proof.
This is [BDES, Theorem 5.1], phrased in terms of coordinate rings. ∎
2.5. -varieties
An affine -scheme is an affine scheme over equipped with an action of the discrete group such that forms a polynomial representation of . Every affine -scheme has the form where is a -algebra. An affine -variety is a reduced affine -scheme such that is finitely -generated over .
For a tuple , let be the spectrum of the ring . This is an affine -variety. Moreover, every affine -variety is isomorphic to a closed -subvariety of some . Thus, in the theory of -varieties, the play the same role as the ordinary affine spaces in ordinary algebraic geometry.
Let be an affine -variety and let be a (scheme-theoretic) point of . We let be the Zariski closure of the orbit of (see [BDES, §3.1]). We say that is -generic if . Such points play a similar role to generic points in ordinary algebraic geometry. We define the generalized orbit of , denoted , to be the set of all points such that (see [BDES, §3.2]).
Write where is a -algebra. Recall that for a vector space we let be the result of treating as a polynomial functor and evaluating on ; this is a -algebra equipped with an action of . We put . The standard inclusion induces a ring homomorphism , and thus a map of -schemes . Since is the union of the , it follows that is the inverse limit of the . We define the -topology on to be the inverse limit topology, where each is given the discrete topology. The -topology is actually quite concrete: if is algebraically closed then the set of closed points of is identified with a product of ’s, and the -topology is just the usual product topology; thus a sequence of -points of converges if each coordinate is eventually constant. One easily sees that any Zariski closed set is -closed (see [NS2, Proposition 2.3]).
We require the following result that relates the Zariski and -topologies:
Proposition 2.2.
Suppose that is algebraically closed. Let be a -variety and let and be -points of . Then the following conditions are equivalent:
-
(a)
The orbits and have the same Zariski cloure.
-
(b)
The orbits and have the same -closure.
Proof.
Suppose (b) holds. Then belongs to the -closure of , which is contained in the Zariski closure of . We thus see that is contained in the Zariski closure of , and so the Zariski closure of is contained in the Zariski closure of . By symmetry, the reverse inclusion holds as well, which yields (a).
Now suppose that (a) holds. We may as well replace with the Zariski closure of , and so that and are -generic in . Let be a typical morphism (see [BDES, §8.1]), where is an irreducible variety and is a pure tuple. Let be a -point lifting , which exists by [BDES, Proposition 7.15], and let be the closure of the -orbit of . Then is dominant since its image contains , and so, by the definition of typical, . It follows that is a point and is -generic in . In what follows we ignore , and regard as a morphism satisfying .
The image of contains a non-empty open subset of by [BDES, Theorem 7.13]. Since belongs to every non-empty -subset of [BDES, Proposition 3.4], we see that . Thus, applying [BDES, Proposition 7.15] again, we can find a -point of such that .
Let be the natural map. By [BDDE, Corollary 2.6.3], the restriction of to is surjective on -points. We can thus find such that . We therefore see that the sequence converges to in the -topology. Since is -continuous, it follows that the sequence converges to in the -topology. Thus , and therefore , and therefore the -closure of , is contained in the -closure of . The reverse inclusion follows by symmetry, and so (b) holds. ∎
Remark 2.3.
Remark 2.4.
We only apply Proposition 2.2 when , 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. -fields
A -field over is a field extension equipped with an action of by -automorphisms such that every element of can be expressed in the form with . If is a -field then is an integral -algebra over , and . Thus every -field can be realized as the fraction field of an integral -algebra.
Let be a -field. A -module is a semi-linear representation of over such that every element of has the form with and . One easily sees that the category of -modules is an abelian category satisfying the (AB5) condition. Moreover, if is any -module then there is a surjection for some polynomial representation (take ), which shows that the objects form a generating set; thus is a Grothendieck abelian category.
We say that is finitely -generated over if it is generated as a field extension by the -orbits of finitely many elements. We say that is rational over if it has the form for some tuple . The invariant subfield of , denoted , is the subfield of consisting of all elements that are invariant under . It is an extension of . If is finitely -generated over then is finitely generated over ([BDES, Proposition 5.8]).
Proposition 2.5.
Let be a -field that is finitely -generated over . Then there exists such that is rational over its invariant subfield.
Proof.
One easily sees that can be -generated by finitely many polynomial elements. We can thus find a finitely -generated -subalgebra of such that . Apply Theorem 2.1 to write where is a -algebra with trivial -action and is a pure tuple. Taking fraction fields, we find where . It follows from [BDES, Proposition 5.7] that , and so is rational over its invariant subfield. ∎
2.7. Generic categories
Let be an integral -algebra. We say that an -module is torsion if every element of is annihilated by a non-zero element of . The category of torsion -modules is a Serre subcategory of . We define the generic category of , denoted , to be the Serre quotient .
The generic category can be described in terms of semi-linear representations. Let . We have a functor
We also have a functor
Indeed, if is a -module then is a polynomial representation, so its image under the natural map consists of polynomial elements, and is therefore contained in ; this shows that is stable under multiplication by , and is thus an -module.
Proposition 2.6.
We have the following:
-
(a)
The functor is exact and kills torsion modules. The induced functor is an equivalence.
-
(b)
The functors form an adjoint pair.
-
(c)
The co-unit is an isomorphism.
Proof.
See [NSS, §2.4]. ∎
We say that an -module is saturated if the natural map is an isomorphism. We will require the following result concerning this concept:
Proposition 2.7.
Let be a pure tuple, let , and let be a polynomial representation. Then is a saturated -module.
Proof.
See [NSS, Proposition 2.8]. ∎
3. The shift and embedding theorems
In this section, we prove our two main technical results on -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 be an integral -algebra, let be a partition, let and be -modules, and suppose we have a surjection of -modules
Then at least one of the following holds:
-
(a)
The given map induces an isomorphism , where is a quotient of .
-
(b)
There exists and a non-zero -invariant element such that the natural map
is surjective, where is obtained from by deleting .
We require some preparation before giving the proof. A weight of is a tuple where for all and for . For a finite subset of , we let be the weight that is 1 at the coordinates in , and 0 away from . We also write in place of when .
Suppose that is a polynomial representation and is a weight. We say that is a weight vector of weight if whenever we have
We let be the space of all weight vectors of weight ; this is the weight space. The space is the direct sum of its weight spaces over all . Moreover, if is non-zero then is non-negative in the sense that for all .
The weight space carries a representation of . Let be the subcategory of spanned by representations of degree . One formulation of Schur–Weyl duality states that the functor
is an equivalence of categories.
Lemma 3.2.
Let and be polynomial representations of degrees and , with irreducible, let be a subset of of cardinality , and let be a non-zero subrepresentation of . Then contains a vector of the form , for some , such that:
-
•
is a weight vector of of weight and is a weight vector of of weight , where and are disjoint and ;
-
•
we have , and and are non-zero;
-
•
we have for .
Proof.
We may as well assume . By Schur–Weyl duality, the -weight space of is non-zero. We can thus find a non-zero element of of the form satisfying the first condition, and with the and linearly independent. Applying an element of the symmetric group , we can assume that has weight . Relabeling, we can assume that have weight , and that the remaining have weight .
Now, by Schur–Weyl duality, the weight space of is an irreducible representation of (acting through the standard inclusion ). Since are linearly independent elements, we can find such that and for , Since has weight with , the group acts trivially on it, and so . For the element is a sum of weight vectors having weight of the form with . We thus see that is an element of of the required form. ∎
Proof of Proposition 3.1.
Let be the kernel of , and let be the projection of to . If then is contained in , and case (a) holds with . Suppose now that . Let and let be such that has a non-zero element of degree . Recall that , where . Applying Lemma 3.2, we can find an element of of the form , where:
-
•
is a weight vector of of weight and is a weight vector of of weight , where and are disjoint and ;
-
•
, and and are non-zero.
-
•
for .
Say that a weight is big if for , and small otherwise. Let and be the sum of the big and small weight spaces in . Then we have a decomposition of -representations
Identifying with , this becomes the decomposition
We thus see that is irreducible as a -representation. Note that is a non-zero element of (and thus generates it as a -representation), and that is -invariant (as it has weight ).
Let be such that . Let be the image of in . Since maps to 0 in , we see that the image of in belongs to , and so the image of belongs to . Since is -stable, is -invariant, and generates as a -representation, we see that any element of maps into . Thus , 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 be an integral -algebra and let be a finitely generated -module. Then there exists , a tuple , and a non-zero -invariant element such that we have an isomorphism of -modules.
Proof.
Say that an -module is good if the conclusion of the theorem holds for it. Consider the following statement, for a tuple :
-
If is an integral -algebra and is a quotient module of then is good.
It suffices to prove for all tuples . The magnitude of a tuple , denoted , is the tuple where is the number of partitions of size in . We order magnitudes lexicographically; this is a well-order. We can thus prove by induction on . Thus let be given, and suppose holds for all with . We prove . If is empty the statement is vacuous, so suppose this is not the case.
Let be an integral -algebra and let be a quotient of . Let be a partition in of maximal size, and let be the tuple obtained from by deleting . We thus have a surjection . We apply Proposition 3.1 with . We consider the two cases separately.
Suppose case (a) holds. Then where is a quotient of . Since has smaller magnitude than , statement holds, and so is good. It is clear then that is good as well.
Now suppose case (b) holds. Then there is some and a -invariant function such that the natural map
is a surjection, where . Now, the left side above has the form where . The tuple has smaller magnitude than , and so statement holds. We thus see that is good as a -module, from which it easily follows that is good as an -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 be an integral -algebra and let be a finitely generated -module. Then there exists , a tuple , and a non-zero -invariant element such that there is an injection of -modules with cokernel annihilated by .
Proof.
Let , , and be as in the shift theorem, so that we have an isomorphism of -modules . Let be the image of in . Scaling our isomorphism by an appropriate power of , we can assume that maps into . Since is projective, we can find a lift of our map, which is necessarily injective. Since this map is an isomorphism after inverting , every element in the cokernel is annihilated by a power of . But the cokernel is finitely -generated and is -invariant, so there is some power of that annihilates the entire cokernel. Replace by this power. ∎
The shift theorem for -algebras implies an analogous result for -fields:
Corollary 3.5.
Let be a -field and let be a finitely generated -module. Then there exists and a tuple such that we have an isomorphism of -modules.
Proof.
Let and let be a finitely -generated -module that spans over . By Theorem 3.3, we have an isomorphism for some , , and . Tensoring up to , we obtain the stated result. ∎
3.3. Some consequences
We now give a few consequences of the shift theorem.
Proposition 3.6.
Let be an integral -algebra and let be a finitely generated -module. Then there exists a non-empty open -stable subset of such that is free over for all .
Proof.
The shift theorem shows that is free as an -module, for some non-zero . Since freeness does not depend on the -actions, it follows that is free as an -module. Thus is free over for all , where is the distinguished open defined by . Since the free locus is obvious -stable, we can take . ∎
Corollary 3.7.
Let be an integral -algebra, let be an -module, and let be a -generic prime of . Then is flat over .
Proof.
First suppose that is finitely generated. By Proposition 3.6, is flat at an non-empty -stable open subset of . Such a subset contains all -generic points [BDES, Proposition 3.4]. Thus is flat at . In general, write with each finitely generated. Then 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 be an integral -algebra, let be a finitely generated -module, and let be the torsion submodule of . Then there exists a non-zero such that .
Proof.
Applying Corollary 3.4, let be an injection of -modules with cokernel annihilated by . Since is torsion, it cannot intersect , and so it injects into . Since is annihilated by , so is . ∎
3.4. The embedding theorem
We now prove the second main result of this section.
Theorem 3.9 (Embedding theorem).
Let be an integral -algebra, let be a tuple, and let . Let be a finitely -generated torsion-free -module. Then there is a tuple and an injection of -modules.
We require some discussion before giving the proof. If is a polynomial representation of then is identified with and is identified with . The natural inclusion thus induces a map , which is injective. If is a -algebra then the map is one of -algebras, and if is an -module then the map is one of -modules. We say that is shift-free if for each the -module has the form for some polynomial representation . Theorem 3.9 thus follows from the following two lemmas.
Lemma 3.10.
Let be a -algebra, let be a tuple, and let . Then is shift-free.
Proof.
We have . Write for some tuple , and let . Then , as -algebras, and, in particular, as -modules. Thus is shift-free. ∎
Lemma 3.11.
Let be an integral shift-free -algebra and let be a finitely -generated torsion-free -module. Then there is a tuple and an injection of -modules.
Proof.
Applying the shift theorem (Theorem 3.3), we have an isomorphism for some , , and . Since is torsion-free, the natural map is injective. Scaling by a power of , we can assume it maps into . Composing with the natural map , we obtain an injection of -modules . As -modules, we have for some polynomial representation . Thus we can identify the target of with where . Since is finitely generated, the image of is contained in for some finite length subrepresentation of . Writing for some tuple yields the result. ∎
There is also an embedding theorem for rational -fields:
Corollary 3.12.
Let be a pure tuple, let , and let be a finitely generated -module. Then there exists an injection of -modules for some tuple .
Proof.
Let and let be a finitely generated -module with . Since is contained in , it is torsion-free. By Theorem 3.9, there is an injection of -modules for some tuple . Tensoring up to gives the stated result. ∎
Remark 3.13.
Remark 3.14.
We do not know if there are any examples of shift-free -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 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 , when is a rational -field. Finally, in §4.3, we prove some results for more general -fields.
4.1. Some category theory
Let be a -linear Grothendieck abelian category, let be a set of non-zero objects in , and let be a function. Suppose that the following conditions hold:
-
(A1)
Every object of is the union of its finitely generated subobjects.
-
(A2)
For any , there are only finitely many with .
-
(A3)
The object is finitely generated, for all .
-
(A4)
The space is finite dimensional over for all .
-
(A5)
The ring is a division ring for all .
-
(A6)
We have only if or .
-
(A7)
Let and let be a direct sum of objects of the form with . Then there is no injection .
-
(A8)
Every finitely generated object of injects into a finite direct sum of the ’s.
We introduce one more piece of notation: for , we let be the intersection of the kernels of all maps with . The object is non-zero by (A7).
Proposition 4.1.
In the above situation, we have the following:
-
(a)
Every object of is locally of finite length.
-
(b)
The ’s are exactly the indecomposable injectives of .
-
(c)
Every finite length object of has finite injective dimension.
-
(d)
The object is simple, and is equal to the socle of . Every simple object is isomorphic to a unique .
-
(e)
The simple object occurs in with multiplicity one; the remaining simple constituents of have the form with .
We break the proof up into a series of lemmas. We assume that satisfies (A1)–(A8) in the following.
Lemma 4.2.
Let be a finitely generated object of . Suppose that every injection , with finitely generated, splits. Then is injective.
Proof.
Let be a finitely generated object of , let be a subobject of , and let be a given morphism. Consider the map , where is embedded diagonally, which is easily seen to be injective. Since is finitely generated, this map splits by hypothesis. This yields a map extending the given map . A variant of Baer’s criterion (see [Stacks, Tag 079G]) now shows that is injective. (The key point here is that is generated by its finitely generated objects, due to (A1).) ∎
Lemma 4.3.
The object is an indecomposable injective, for all .
Proof.
Since has no non-trivial idempotents, it follows that is indecomposable. It is clear that is injective if , since this hypothesis is void. Assume now that is injective for all with and let satisfy . Suppose we have an injection , with finitely generated. Choose an injection , where is a finite direct sum of ’s, which is possible by (A8). Write , where is a sum of ’s with , is a sum of ’s and is a sum of ’s with and . Let be the projection of onto . Then by (A6) and is not injective by (A7). Since is injective, it follows that is non-zero. Thus , followed by a further projection, provides a map such that is non-zero. Since is a division algebra by (A5), we can find such that , and so is split. It follows from Lemma 4.2 that is injective. The result now follows by induction. ∎
Lemma 4.4.
The object is simple, and is the socle of .
Proof.
Consider the natural map
This is the universal map from to a sum of ’s with . Thus .
Suppose is a non-zero subobject of . The object is finitely generated by (A3). Thus, by (A8), we have an injection where is a finite sum of ’s. Any map with and is automatically zero by (A6); similarly, any map is zero, since any non-zero map is injective by (A5). It follows that can be taken to be a finite sum of ’s with . Let be the composition . By the universality of , we have for some , and so . Since , this shows that , and so is simple.
Since is indecomposable, it follows that it is the injective envelope of . Since is simple, it is therefore the socle of . ∎
Lemma 4.5.
Every simple object of is isomorphic to , for a unique .
Proof.
Let be a simple object of . Then is necessarily finitely generated, and so by (A8) we have an injection , where is a finite sum of ’s. Since is simple, it follows that must inject into one of the factors, and land in the socle. This gives an isomorphism .
Suppose now that . Then the injective envelopes of and would be isomorphic, i.e., . By (A6), this implies that . ∎
Lemma 4.6.
Every object of is locally of finite length.
Proof.
By (A1), it suffices to show that every finitely generated object of is finite length. By (A8), it suffices to show that each has finite length. We proceed by induction on . Thus suppose has finite length for and let be given with . Using notation as in Lemma 4.4, we have an exact sequence
By Lemma 4.4, the object is simple. By induction, each appearing in the sum on the right has finite length. By (A2), the sum is finite, and by (A4) each space is finite dimensional. Thus the rightmost term above has finite length. It follows that has finite length, as required. ∎
Lemma 4.7.
Every indecomposable injective object of is isomorphic to for a unique .
Proof.
Let be an indecomposable injective. Since is the union of its finite length subobjects b Lemma 4.6, it follows that the socle of is simple, and that is its injective envelope. Thus for some . This is unique, as implies by (A6). ∎
Lemma 4.8.
The simple occurs in with multiplicity one. The remaining simple constituents of have the form with .
Proof.
We proceed by induction on . Using notation as in Lemma 4.4, we have an exact sequence
The result now follows. ∎
Lemma 4.9.
Every finite length object of has finite injective dimension.
Proof.
It suffices to prove that each has finite injective dimension. We proceed by induction on . Thus suppose that has finite injective dimension for . By Lemma 4.8, it thus follows that has finite injective dimension, and so does as well. ∎
4.2. Applications to rational -fields
Fix a pure tuple , let , and let . For a partition , we let be the intersection of the kernels of all maps with . The following is our main result on the structure of -modules.
Theorem 4.10.
We have the following:
-
(a)
Every finitely generated -module has finite length.
-
(b)
The indecomposable injective -modules are exactly the , with a partition.
-
(c)
Every finite lengh -module has finite injective dimension.
-
(d)
The -module is simple, and is the socle of . Every simple -module is isomorphic to a unique .
-
(e)
The simple occurs in with multiplicity one; the remaining simple constituents have the form with .
Proof.
We apply Proposition 4.1. We take , take to be the set of partitions, and take to have its usual meaning (the size of ). For , we let . We verify the conditions (A1)–(A8). The first three conditions are clear.
Now, recall from §2.7 that we have a functor given by , which has a right adjoint given by . Moreover, for any tuple (Proposition 2.7). In particular, we have
This is finite dimensional over since occurs in with finite multiplicity; this proves (A4). If then we find that the above space is isomorphic to , which proves (A5). Finally, if , or if but , then the above space is 0, which proves (A6).
We now handle (A7). Since is finitely generated, it suffices to consider the case where is a finite direct sum in (A7). Thus, suppose by way of contraction that we have an injection , where is a tuple composed of partitions that are strictly smaller than . Applying the functor, this gives an injection of -modules . Let be such that . This is possible since is a polynomial in of degree , while is a polynomial of degree . Evaluating our injection on , we obtain an injection
of -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).
Corollary 4.11.
All projective -modules are injective.
Proof.
Let and be as in the above proof. Since is exact, the its right adjoint takes injectives to injectives. In particular, we see that is an injective -module for any polynomial representation . As (Proposition 2.7), and every projective -module has this form, the result follows. ∎
4.3. Applications to other -fields
By leveraging Theorem 4.10, we are able to deduce the following fundamental result for more general -fields:
Theorem 4.12.
Let be a -field that is finitely generated over its invariant subfield .
-
(a)
Any finitely generated -module has finite length.
-
(b)
If and are finitely generated -modules then is a finite dimensional -vector space.
The first statement is reasonably straightforward:
Proof of Theorem 4.12(a).
Applying Proposition 2.5, let be such that is a rational -field over its invariant subfield. Let be a finitely generated -module. Then is a finitely generated -module, and therefore of finite length by Theorem 4.10(a). It follows that has finite length. In fact, if has length then has length , for if is any chain of -submodules of then is a chain of -submodules of , and so for some , and so . ∎
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 is a rational -field over .
Proof.
Choose a surjection for some tuple , which is possible in general, and an injection for some tuple , which is possible by the embedding theorem (Corollary 3.12) since is rational. We thus obtain an injection
We have seen (in the proof of Theorem 4.10) that this is finite dimensional over . The result follows. ∎
Lemma 4.14.
Let be a finitely generated -module and let be an endomorphism of . Then satisfies a non-zero polynomial with coefficients in .
Proof.
Applying Proposition 2.5, let be such that is a rational -field over its invariant subfield; in other words, this means is rational over as a -field. Let be the space of all -linear -equivariant maps ; this is identified with . By Lemma 4.13, is a finite dimensional vector space over the field . Thus the elements of are linearly dependent, which gives the requisite polynomial. ∎
For as above, the set of all polynomials that satisfies forms an ideal in the univariate polynomial ring . We define the minimal polynomial of to be the unique monic generator of this ideal. In other words, the minimal polynomial of is the unique monic polynomial that satisfies of minimal degree.
Lemma 4.15.
Let be a finitely generated -module and let be an endomorphism of . Then the minimal polyomial of has coefficients in the invariant field .
Proof.
Suppose that is the equation given by the minimal polynomial. If then we also have . By uniquness of the minimal polynomial, we therefore have . Since this holds for all , it follows that , as required. ∎
The following lemma is a version of Schur’s lemma:
Lemma 4.16.
Suppose that is algebraically closed and is a simple -module. Then .
Proof.
Since is simple, it follows that is a division ring. We know that contains in its center. By the Lemma 4.15, every element of is algebraic over . (Note that is necessarily finitely generated since it is simple.) Since is algebraically closed, it follows that . ∎
Lemma 4.17.
Suppose that is algebraically closed. Then Theorem 4.12(b) holds.
Proof.
It follows from the previous lemma that is finite dimensional over if and are simple. As and have finite length by Theorem 4.12(a), the general case follows from dévissage. ∎
We now deduce the general case from the case with algebraically closed using a base change argument. For this, we require two more lemmas.
Lemma 4.18.
Any element of that is algebraic over belongs to , i.e., is algebraically closed within .
Proof.
Let be algebraic over , and let be its minimal polynomial. Since acts on by field homomorphisms, it permutes the roots of in . This action corresponds to a homomorphism where is the number of roots of in . Since any group homomorphism is trivial, it follows that is trivial on each group of elementary matrices in . Since these groups generate , it follows that . We thus see that is fixed by . However, is also fixed by for . It follows that is fixed by , i.e., . ∎
Suppose is an algebraic extension of . Then the above lemma implies that is a field. Letting act on by acting trivially on , one easily sees that is a -field, its invariant field is , and it is finitely -generated over .
Lemma 4.19.
Let be an algebraic extension of and put . Let and be -modules, with finitely generated. Then the natural map
is an isomorphism.
Proof.
By adjunction, we have
Now, for any -vector space , we have a natural map
Picking a -basis for , we find that the above map is isomorphic to the map
This map is an isomorphism since is finitely generated. Applying this with gives the result. ∎
Proof of Theorem 4.12(b).
Let be an algebraic closure of , and let . Let and be finitely generated -modules. By Lemma 4.19, the map
is an isomorphism. As and are finitely generated -modules, it follows from Lemma 4.17 that the right side above is a finite dimensional -vector space. It thus follows that is a finite dimensional -vector space, which completes the proof. ∎
5. Brauer categories, Weyl’s construction, universal properties
The purpose of this section is to describe , when is a rational -field, in terms of a combinatorial category, the upwards -Brauer category . We begin in §5.1 by reviewing generalities on representations of categories. We introduce 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 , and give a universal property for .
5.1. Representations of categories
We now review a bit of material on representations of categories. See [SS7, §3] for more detail.
Let be an essentially small -linear category. A representation of , or a -module, is a functor , and a map of -modules is a natural transformation. We let be the category of -modules. For -modules and , we write for the set of maps of -modules .
Let be an object of . We define the principal projective -module at , denoted , by . If is an arbitary -module then we have an identification
by Yoneda’s lemma, which shows that is projective. The above identity also shows that can be realized as a quotient of a direct sum of principal projectives.
We similarly define the principal injective -module at , denoted , by . If is an arbitary -module then we have an identification
(see [SS7, Proposition 3.2]), which shows that is injective.
Proposition 5.1.
Suppose that the sets in are finite dimensional. Then the following categories are equivalent:
-
(a)
The category .
-
(b)
The full subcategory of spanned by the principal projectives.
-
(c)
The full subcategory of spanned by the principal injectives.
Proof.
Let be the category in (b). We have a functor given by . It is obviously essentially surjective, and is fully faithful by Yoneda’s lemma. Similarly, let be the category in (c). Then we have a functor given by . We have
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 are in bijection with the set of natural numbers; for , we let be a representative of the th isomorphism class. We say that is upwards if implies .
Proposition 5.2.
Suppose is upwards and all sets are finite dimensional. Then the principal injectives are of finite length, and every finite length -module embeds into a finite sum of principal injectives.
Proof.
Let be a -module, and write in place of . Define the support of to be the set of natural numbers for which . Define the th truncation of , denoted , to be the -module given by
one easily sees that this is a -submodule of since is upwards. From the above structure, one easily verifies the following two statements:
-
(a)
A -module is simple if and only if it is supported in a single degree and is a simple module over the ring .
-
(b)
A -module has finite length if and only if it has finite support and is finite dimensional for all .
It follows from (b) that the principal injective is of finite length. It follows from (a) that if is a simple supported in degree then embeds into . One now easily sees that any finite length objects embeds into a sum of ’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 -modules.
For a partition of , recall that is the irreducible representation of associated to (the Specht module). For a finite set of cardinality , we let be the associated representation of . One can define this in a canonical manner by mimicking the construction of , but using elements of in place of the integers .
Fix a pure tuple . A -block on a set is a triple where
-
•
is an element of ,
-
•
is a subset of of cardinality (called the support of the block),
-
•
is an element of the Specht module .
Let and be a finite sets. A downwards -diagram from to is a pair where is a collection of -blocks on with disjoint supports and is a bijection, where is the union of the supports of the blocks in . The space of downwards -diagrams is the vector space spanned by elements , with an downwards -diagram, with the following relation imposed:
-
•
Suppose that contains a block , and let be a linear combination in the Specht module. Let be the diagram obtained by replacing this block with , and let be defined similarly but using . Then .
We now come to the main definition:
Definition 5.3.
The downwards -Brauer category, denoted , is the -linear category described as follows.
-
•
The objects of are finite sets.
-
•
Given finite sets and , the space of morphisms is the space of downwards -diagrams from to .
-
•
Composition is defined as follows. Let be a diagram from to , and let be a diagram from to . Let and let , where denotes the result of transporting along the bijection . Then . ∎
Example 5.4.
Example 5.5.
The category carries a natural symmetric monoidal structure given by disjoint union. Precisely, for two objects and , the object is simply the disjoint union of the sets and . Given two morphisms and , the morphism is defined to be . Note that is a -linear functor in each of its arguments.
The category admits a universal property, which we now describe. Let be an -linear symmetric monoidal category. Let be the category whose objects are pairs , where is an object of and is a morphism in the Karoubian–additive envelope of , where is the unit object of ; of course, if is additive and Karoubian (e.g., abelian) then one does not need to take the envelope here. Morphisms in are defined in the obvious manner.
Proposition 5.6.
Notation as above, we have a natural equivalence of categories
Here denotes the category of symmetric monoidal -linear functors.
Proof.
We first define the functor . Thus suppose given a symmetric monoidal -linear functor . Let . We define a map . It suffices to define maps for each . Thus fix such . Put . Since is a symmetric monoidal functor, it induces an -equivariant map
Now, . Inside of one has the subspace spanned by diagrams that consist of a single block of type . This subspace is isomorphic to as a representation of . We thus obtain a canonical -equivariant map , which yields a map , as required. We have thus defined . We define on objects by . The definition on morphisms is clear.
To show that is an equivalence, we construct a quasi-inverse functor
Thus let in be given. We define a symmetric monoidal -linear functor . On objects, we define by . Now, consider a -block . We have
The summand on the right side contains the element . We say that the corresponding morphism is associated to this block. Note that this construction is linear in the element . Now, consider a morphism in represented by a diagram. Suppose this diagram corresponds to a pair , where is a collection of disjoint blocks on and is a bijection. We define a morphism as follows. Write . We have a map by tensoring together the maps associated to individual blocks. We also have a map from the bijection . The map is the tensor product of these two maps. The construction extends to a -linear map
One easily verifies that is compatible with composition and is naturally a symmetric monoidal functor. We define on objects by . The definition of on morphisms is clear.
One easily verifies that and are naturally quasi-inverse. This completes the proof. ∎
There is also an upwards -Brauer category , defined in the same manner, but where now blocks are only allowed on the target of a morphism. In other words, is simply the opposite category of . The category admits a natural symmetric monoidal structure, and has a similar universal property to .
5.3. Equivalences
We now establish a number of equivalences between categories associated to , , and the -Brauer categories..
Proposition 5.7.
The following symmetric monoidal -linear categories are equivalent:
-
(a)
The downwards -Brauer category .
-
(b)
The full subcategory of spanned by the objects for .
-
(c)
The full subcategory of spanned by the objects for .
As -linear categories (ignoring the monoidal structure), these categories are also equivalent to
-
(d)
The full subcategory of spanned by the principal projective objects.
-
(e)
The full subcategory of spanned by the principal injective objects.
Proof.
We break the proof into three steps.
Step 1: equivalence of (a) and (b). Let be the category in (b) and let . Then , where on the left side is formed with respect to . Since contains as a subrepresentation, there is a natural map of -modules . We thus have a natural map . Since is the unit object for , the universal property of (Proposition 5.6) furnishes a symmetric monoidal -linear functor
This functor has the property that . It is clear that is essentially surjective. To complete this step, it suffices to show that is fully faithful.
Before doing this, we introduce some notation. Identify the weight lattice of the diagonal torus in with . For a finite subset , let denote the weight that is 1 in the coordinates and 0 elsewhere; also, write in place of . Given a weight and a polynomial representation , let be the weight space of .
Now, we have
By Schur–Weyl duality, for any polynomial representation ; explicitly, a map corresponds to . We must therefore understand the weight space of .
Let . The -weight space of is canonically isomorphic to the Specht module if , and vanishes for other values of . More generally, let be a subset of of size . Then we have a canonical isomorphism . Fix a basis for . For , let be the image of under this isomorphism. We refer to as the support of the element . Let be the set of all elements of the form for all choices of , , and , and let be the set of all monomials where the ’s belong to and their supports form a partition of . We thus see that is a basis for .
From the above discussion, we see that the -weight space of has for a basis all elements of the form
where are distinct elements of and with . We associate to the above element the -diagram given by the pair , where is the collection of blocks corresponding to (each corresponds to a block ), and is the bijection taking to . We have thus constructed a natural linear isomorphism
As we have already seen, the left side above is identified with . One easily sees that the resulting isomorphism
is induced by . This shows that is fully faithful. This completes the first step of the proof.
Step 2: equivalence of (b) and (c). Let be the category in (c). The functor given by induces a functor . 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 -linear categories. This follows from Proposition 5.1. ∎
For an abelian category , we let 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 -linear abelian categories:
-
(a)
-
(b)
-
(c)
.
Proof.
Let and be Grothendieck abelian categories, and let and be full subcategories of and consisting of projective objects. Suppose that and are enough projectives (i.e., they form generating families). Then any equivalence extends uniquely to an equivalence . 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 and . 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 follows from Theorem 4.10, while the fact that category (d) gives enough injectives in is Proposition 5.2 (note that is an upwards category, as defined before Proposition 5.2). Statement (c) follows from statements (a) and (b). ∎
Remark 5.9.
For a partition , recall that denote the simple object of indexed by . Using the above proposition, we can compute the groups between these objects:
Corollary 5.10.
We have
Proof.
Let and let be the equivalence constructed above. Tracing through the definition, we see that takes to the principal injective . We thus see that is the -isotypic piece of , with respect to its natural -action. Taking socles, we see that is the simple -module with . We thus have
The on the right side can be computed by taking an injective resolution of in . As we have seen (Proposition 5.2), this can be accomplished using principal injectives. As these objects are injective in the larger category , we find
We now appeal to the equivalence . One easily sees that corresponds to the simple -module (with positive degree elements of acting by 0). We thus have
The right group can be computed using the projective resolution of provided by the Koszul complex. This yields the stated result. ∎
5.4. Weyl’s construction
We recall Weyl’s classical traceless tensor construction. Equip with a non-degenerate symmetric bilinear form. Let . Given , let be the map obtained by applying the form to the and tensor factors. Let be the intersection of the kernels of , over all choices of and ; this is the space of traceless tensors. The space is a -subrepresentation of . Weyl proved that the isotypic piece of is either 0 or the irreducible of with highest weight .
We now establish an analog of this construction for . Recall that . For each , let be the natural map (coming from the inclusion ). Given an element of the Specht module , let be the composition
where the first map comes from the projection provided by , and the second map is . Let . Given , , and a subset of of size , we let be the map obtained by applying to the tensor factors indexed by . Let be the intersection of the kernels of the over all choices of , , and . This is a -module equipped with an action of . The following is our analog of Weyl’s construction:
Proposition 5.11.
Let be a partition of . Then the isotypic piece of is the simple -module .
Proof.
Under the equivalence , the -module corresponds to the th principal injective -module. Thinking in terms of -diagrams, we see that any map , with , is a linear combination of maps of the form , where is some map. It follows that is the intersection of the kernels of all maps with . From this, we see that the -isotypic piece of is the intersection of the kernels of alls maps with . This is the simple object (see §4.2). ∎
5.5. Universal properties
We can now give the universal property for the category . This is analogous to the universal property for given in [SS3, §4.4]. For symmetric monoidal -linear abelian categories and , we let be the category of left-exact symmetric monoidal -linear functors . Also, recall the category defined before Proposition 5.6.
Theorem 5.12.
Let be a symmetric monoidal -linear abelian category with exact. Then we have a natural equivalence of categories
In other words, to give a -linear left-exact symmetric monoidal functor is the same as to give an object of equipped with a -form.
Proof.
Let be the full subcategory of spanned by the objects for . This category is stable under tensor products. As a -linear symmetric monoidal category, it is equivalent to by Proposition 5.7. Thus by the universal property for (Proposition 5.6), we have a natural equivalence
Now, every object of is injective in (Theorem 4.10(b)), and every object of embeds into a finite direct sum of objects in (Theorem 3.9). It follows that any functor extends uniquely to a left-exact functor . Since 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 be a finite dimensional -vector space equipped with a form . From the universal property, get a left-exact cocontinuous symmetric monoidal functor
that we call the specialization functor with respect to and . Since is left-exact, one can consider its right derived functors , which we call the derived specialization functors. Is it possible to compute the values of these functors on simple objects for a generic form ? When the category 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 , and give a complete classification of them.
6.1. Definitions
Fix, for the duration of §6, a -field that is finitely generated over its invariant subfield , and a -algebra finitely generated over with . Furthermore, let be the -variety associated to . The following is the main object of study in this section:
Definition 6.1.
A fiber functor on is a symmetric monoidal functor that is exact, faithful, cocontinuous, and -linear. ∎
The goal of this section is to classify the fiber functors on . This is accomplished in Theorem 6.5 below.
6.2. Examples of fiber functors
Let be a -generic -point of and let be the corresponding maximal ideal of . Define a functor
Since every -module is flat at (Corollary 3.7), it follows that is exact. Moreover, it is clear that kills torsion -modules. It follows that factors through the generic category . Identifying this with , we thus obtain a functor
We now have:
Proposition 6.2.
The functor is a fiber functor (in a natural manner).
Proof.
The functor is clearly exact, cocontinuous, and -linear, and also admits a natural symmetric monoidal structure; it follows that inherits these properties. To complete the proof, we must show that is faithful.
We first claim that if is a torsion-free -module such that then . To see this, first suppose that is finitely generated. Then is free over (Proposition 3.6). Thus the vanishing of implies that of , and thus of since is torsion-free. We now treat the general case. Let be a finitely generated submodule of . Since is flat at (Corollary 3.7), the map is injective, and so . Thus by the previous case. Since was arbitrary, it follows that as well.
Now, to prove faithfulness, it suffices to show that if is a map of torsion-free -modules such that the induced map vanishes then . Thus let such an be given. Let be the image of . Since is flat at (Corollary 3.7), it follows that is the image of , and thus vanishes. Hence by the previous paragraph, and so as well. ∎
Remark 6.3.
Proposition 6.2 was proven for (and a specific choice of ) 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 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 such that, putting , the category admits a fiber functor.
6.3. More examples of fiber functors
Let be an infinite dimensional -vector space. Recall that , where is obtained by treating as a polynomial functor and evaluating on . Suppose that is a -generic -point of , corresponding to the maximal ideal of . (By -generic here, we mean there is no proper closed -subvariety of with .) Define a functor
Once again, this functor is exact and kills torsion modules, and thus induces a functor
The same argument as in Proposition 6.2 shows that it too is a fiber functor. If then is the funtor introduced above. We note that , and so and can only be isomorphic if (as cardinal numbers). In particular, if then will not be isomorphic to a fiber functor of the form .
6.4. The main theorem
In the remainder of this section, a pair will always stand for an infinite dimensional -vector space and a -generic -point of . If is a second such pair, then an isomorphism is a linear isomorphism such that the induced map carries to . The following theorem classifies fiber functors:
Theorem 6.5.
We have the following:
-
(a)
Any fiber functor on is isomorphic to one of the form .
-
(b)
Given two pairs and , we have a natural bijection
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 is equivalent to the groupoid of pairs . ∎
Remark 6.7.
It follows from the theorem that the automorphism group of the fiber functor is the stabilizer of in the group . In most cases, this group will be finite, and so 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 , there are potentially many choices of . The theorem implies that any two choices of have the same set of -generic points (up to natural bijection). In fact, this can be seen directly. Suppose is a second -algebra that is finitely -generated over and has , and let . One can show that and are birational, in the sense that there are open -subsets and and an isomorphism of -varieties. Every -generic point of is contained in , and similarly every -generic point of is contained in (see [BDES, Proposition 3.4]). Clearly, these points are mapped bijectively to one another via . ∎
6.5. Proof of Theorem 6.5
Let be a vector space, let be a -point of , and let be the corresponding maximal ideal of . We define as above; that is, for an -module , we put
Previously, we had only used this when is infinite dimensional and is -generic, but we now consider it more generally.
Lemma 6.9.
Let and be as above. Suppose that there is a fiber functor on such that . Then is infinite dimensional and is -generic.
Proof.
Suppose, by way of contradiction, that is not -generic (which is automatic if is finite dimensional). There is then a non-zero -ideal of such that belongs to the vanishing locus of . Then is non-zero. On the other hand, , and so . This is a contradiction, which completes the proof. ∎
Lemma 6.10.
Let be a fiber functor on . Then is isomorphic to some .
Proof.
Let . Suppose is a polynomial representation of . Then we have , where on the right side we treat as a polynomial functor and apply it to the object of . We thus find
where in the second step we used that commutes with the action of poylnomial functors, as is symmetric monoidal.
We have a natural surjective map of algebra objects in , given by multiplication. Applying , and appealing to the above, this yields a surjective -algebra homomorphism map . Let , a maximal ideal of , and let is the associated point.
Now, let be an -module. Choose a presentation
where and are polynomial representations. We obtain a commutative diagram
with exact rows. Applying , we obtain a commutative diagram
It follows that the right vertical map induces an isomorphism
By Lemma 6.9, we see that is infinite dimensional and is -generic. The above isomorphism thus induces an isomorphism . ∎
Lemma 6.11.
Let and be given. Then we have a natural bijection
that is compatible with composition of isomorphisms.
Proof.
We first construct a map
Thus let be a -linear isomorphism such that the induced map takes to . It follows that under the induced ring homomorphism the ideal contracts to the ideal . Let be an -module. Then induces an isomorphism , which further induces an isomorphism on the quotients by and . This yields an isomorphism which, in turn, leads to an isomorphism . We define .
We now define a map
Let be an isomorphism of fiber functors. As , and similarly for , we see that induces a -linear isomorphism . Let be the kernel of the map in . As we have seen, , and similarly for . We thus see that under the ring isomorphism induces by , the ideal is taken to . Thus defines an isomorphism . We put .
We leave to the reader the verification that and 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 -varieties.
7.1. Germinal subgroups
Fix a group . The following definition introduces the main concept studied in this section:
Definition 7.1.
A germinal subgroup of is a family , where is a directed set and each is a subset of , satisfying the following conditions:
-
(a)
If then .
-
(b)
Each contains the identity element.
-
(c)
Given there is some such that . ∎
The generalized stabilizer of a point on a -variety will be a germinal subgroup. In this case, the intersection of the sets will be the usual stabilizer, which is typically “too small.” Each of the sets , on the other hand, is “too big.” One can think of the germinal subgroup as a kind of filter on that is attempting to pick out a hypothetical subset that is bigger than the intersection but smaller than each . As this picture suggests, one should always be allowed to pass to a cofinal subset of when working in the setting of germinal subgroups.
7.2. Representations
We fix a germinal subgroup of for §7.2.
Definition 7.2.
A pre-representation of over a field consists of a -vector space and a linear function
Suppose that and are pre-representationss of over . A map of pre-representations is a -linear map such that the obvious diagram commutes. ∎
Suppose is a pre-representation. Given , its image in is represented by a function for some . Given an element of this , we denote its image in under this function by . We thus think of a pre-representation as a kind of partially defined action map .
Definition 7.3.
A representation of is a pre-representation such that the following two conditions hold:
-
•
We have for all .
-
•
Given there exists such that for each there exists some such that for all .
A map of representations is simply a map of pre-representations. We let be the category of representations of over . ∎
We make a number of remarks concerning this definition.
-
•
Let be a representation of and let be a subspace of . Then is a subrepresentation of if and only if for every there exists such that .
-
•
Let and be representations of and let be a linear map. Then is a map of representations if and only if for each there exists such that for all .
-
•
Let be a representation of . Then naturally carries the structure of a -representation. A similar comments applies to maps of representations. We thus have a restriction functor .
-
•
The category 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 be a representation of . Extend the partially defined action map to a function in any manner. This gives the structure of a module over the non-commutative polynomial ring with variables indexed by . Let be the dimension of as a -vector space. One easily sees that any -submodule of is a -subrepresentation. It follows that every is contained in a -subrepresentation of dimension at most , namely, . Thus, taking one representation from each isomorphism class of representations of dimension at most , one obtains a generating set for . It follows that is a Grothendieck abelian category. In particular, it is complete.
-
•
From the above, we see that 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 -representation. This intersection may not coincide with the usual intersection of vector subspaces.
-
•
Let and be representations of . We give the vector space the structure of a representation in the usual manner: that is, we define
provided and are defined for all . One easily verifies that this is indeed a representation. This construction endows with a symmetric monoial structure.
7.3. Weak subrepresentations
Given any vector space , the dual space carries a natural topology: namely, a sequence (or net) in converges to if for every vector there is some such that for all . We call this the -topology. For a subspace of , we let be its annihilator, i.e., the set of functionals such that for all . One easily sees that is -closed, and that is a bijection between subspaces of and closed subspaces of .
Let be a representation of and let be a subspace. A -sequence is a sequence in , indexed by some directed set , such that for each there exists such that for all . We say that is a weak -subrepresentation of if it satisfies the following condition: given and a -sequence such that converges in to an element , we have .
Proposition 7.4.
Let be a representation of and let be -subrepresentation of . Then is a weak -subrepresentation of .
Proof.
Let and let be a -sequence such that converges in to some element . Let . Since is a -subrepresentation, there exists such that for all . Since converges to there is some such that for all . Let be such that for all . Then for we have since and vanishes on . Thus vanishes on , and so . This shows that is a weak subrepresentation. ∎
7.4. Generalized stabilizers
Let be a directed set and let be an inverse system of sets; for , let be the transistion map. Let be the inverse limit of the system. For , we let be the natural map. We suppose that a group acts on , and that the action satisfies the following condition: given and there exists such that factors through ; in other words, one can complete the following commutative diagram:
Equivalently, this means that each acts uniformly continuously on , when is endowed with the inverse limit uniform structure (and each with the discrete uniform structure).
We now come to a fundamental definition:
Definition 7.5.
Let . For , let be the set of elements such that . The generalized stabilizer of is the system . ∎
Proposition 7.6.
The generalized stabilizer is a germinal subgroup of .
Proof.
We verify the three conditions of Definition 7.1. It is clear that for all , which verifies condition (a). If and then taking the given identity and applying the transition map , we find that , and so . This shows that , which verifies condition (b).
Finally, we come to condition (c). Suppose . Let be such that we have a factorization for some . Suppose . Then . Applying , we find , which shows that . Thus , as required. ∎
Proposition 7.7.
The intersection is the usual stabilizer of , i.e., the set of all such that .
Proof.
It is clear that if stabilizes then for all . Conversely, if for all then we have for all , and so , which shows that stabilizes . ∎
7.5. Representations from equivariant bundles
Maintain the notation from §7.4. We now describe how to produce representations of from certain kinds of equivariant vector bundles on . This discussion is included simply to offer some intuition for germinal subgroups, and is not used in what follows.
For each , let be a vector bundle on ; since is discrete, this simply amounts to giving a vector space for each . To keep this discussion less technical, we assume that each is finite dimensional. Suppose that the dual bundles have the structure of an inverse system of vector bundles, and let be the inverse limit, which is a vector bundle on (in a loose sense; it may not be locally trivial). For a point of , the fiber is the inverse limit of the vector spaces . Define to be the corresponding direct limit; note that is the dual space of . We say that is good if there exists such that the transition map is injective for all .
Suppose now that is endowed with a -equivariant structure. Thus for and we have linear isomorphisms and that satisfy the cocycle conditions. As in the previous section, we assume the map is uniformly continuous. Let be a good point. We claim that is naturally a representation of the generalized stabilizer . Indeed, suppose , so that . We have a (likely non-commutative) diagram
Assuming is large enough, is an inclusion. For , we define to be the element . One easily verifies that this is independent of , and defines the structure of a -representation on .
8. Generalized stabilizers on -varieties
In this final section, we study the generalized stabilizer of a point on a -variety . Our main result provides an equivalence between the category of polynomial representations of and the category of -modules when is a -generic point on . This yields the statements in §1.2, as the corresponding statements for have already been established.
8.1. Generalized stabilizers on -varieties
Let be an irreducible affine -variety over the field . Let be the ring obtained by evaluating on and let , a finite dimensional variety over . Then is the inverse limit of the in the category of sets. Let be the natural map. Given and , we see that can be obtained from the image of in by applying . Thus if is such that , then one can recover from . This shows that the action of is uniformly continuous, as described in §7.4.
Fix a point . Let be its generalized stabilizer for the action of on . Thus is the set of elements such that and have the same image in . Letting be the defining ideal of , we see that can also be described as the set of elements such that .
We say that a representation of is polynomial if there is a polynomial representation of such that is isomorphic to a subquotient of (regarded as a representation of ). We write for the category of polynomial representations of . 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 by using restrictions of algebraic representations of (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 and be polynomial representations of and let be a map of -modules. Then the linear map obtained by reducing modulo is a map of -representations.
Proof.
Let be given. Let be such that is invariant under . We claim that for , which will complete the proof. Thus let be given. Write with and . Then we have
so it is enough to show that for each . Since is -invariant, so is ; in other words, . We thus see that belongs to . By definition of , we have , and so belongs to . This exactly means that vanishes at , i.e., . This verifies the claim. ∎
We now suppose that is -generic; if it is not, one can simply replace with the orbit closure of . The following proposition is our main construction of -representations:
Proposition 8.3.
There exists a unique right exact functor
satisfying the following two conditions:
-
(a)
We have as vector spaces (and similarly for morphisms).
-
(b)
If is a polynomial representation then the -action on is the restriction of the action.
The functor is exact and kills the torsion category, and thus induces a functor
The functor is exact, cocontinuous, faithful, -linear, and naturally symmetric monoidal.
Proof.
Let be an -module. Choose a presentation
where and are polynomial representations. Applying , we obtain a sequence
By Proposition 8.2, the first map is one of -representations. It follows that inherits the structure of a -representation, which is easily seen to be independent of the choice of presentation. This representation is polynomial since it is a quotient of . One easily sees that this construction defines a right-exact functor
It is clear that (a) and (b) hold. The uniqueness of follows from the fact that it is right-exact and determined on the category of projective -modules by (a) and (b).
Since is -generic, is flat at (Corollary 3.7), and so is exact. It is clear that kills the torsion subcategory. It thus factors through the generic category, which is equivalent to . We therefore obtain a functor as in the statement of the proposition. Of course, ignoring the representation structure, is just the fiber functor we constructed in §6.2. In other words, the diagram
commutes, where the vertical arrow is the forgetful functor. It follows that is exact, cocontinuous, faithful, and -linear; moreover, one easily sees that the symmetric monoidal structure on respects the -structure, and so is naturally symmetric monoidal as well. ∎
We expect that is an equivalence in general. In the remainder of this section, we prove this when is a rational -field (Theorem 8.11) and is algebraically clsoed.
Remark 8.4.
The above construction is essentially a special case of the one from §7.5, as we can regard as a vector bundle (loosely interpreted) over . ∎
Remark 8.5.
Let be a submodule of , and let . It is easy to see that is a weak subrepresentation of . Indeed, let , a closed -subscheme of the vector bundle , and . Suppose and is a -sequence such that converges to in . Since converges to , it follows that converges in to . Since each belongs to and is closed, we see that . This verifies the claim.
We had originally defined a -representation to be a pair consisting of a polynomial representation and a weak subrepresentation . 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 ; the ambient representation 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 is that of a -pre-representation, as in Definition 7.2. ∎
8.3. From representations to modules
We assume for the remainder of §8 that is algebraically closed. Fix a pure tuple , put , put , and let . Fix a -generic -point of , and let 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 be a tuple and let be a subspace of . The following are equivalent:
-
(a)
There is an -submodule of such that .
-
(b)
The space is a -subrepresentation of .
-
(c)
The space is a weak -subrepresentation of .
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 -point of is degenerate if it is not -generic, and non-degenerate otherwise. For a single partition , the degenerate points in form a -subspace [BDES, Proposition 9.2]. A system of -variables is a set of points in that forms a basis modulo the subspace of degenerate elements. A system of variables is a choice of system of -variables for all .
Lemma 8.7.
Let and be pure tuples, let be -generic, and let be the set of -points such that is -generic. Then is Zariski dense in .
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 and are composed of partitions of some constant size . First suppose that . Then a point is non-degenerate if its components are linearly independent. We can clearly choose such that the components of are independent while at the same time realizing arbitary values at finitely many coordinates of . Since any non-zero function on uses only finitely many coordinates, it follows that we can choose such that . Thus is Zariski dense.
The case when is similar. The set is non-empty: we can choose a system of variables that includes the components of , and then take the components of to be other elements from the system. Let . Then we can find a degenerate -point of realizing arbitary values at finitely many coordiantes. It follows that also realizes arbitrary values at these coordinates, and so again is Zariski dense. ∎
Lemma 8.8.
Let be a tuple and let be a -point of . Then there exists a pure tuple , a -point of such that is -generic, and a map of -varieties over such that .
Proof.
Write and let be the components of . Pick a system of variables including . By [BDES, Theorem 9.5], there exists a pure tuple and a map of -varieties such that , where are distinct elements from the system of variables. Now, after applying a permutation, we can assume that for and the remaining and are distinct. Let and . Now, let be the composition
where is the diagonal map that copies the first coordinates of into those of . Then , and so . By construction is -generic. ∎
Given vector spaces , we let be the annihilator of in the dual space .
Lemma 8.9.
Let be a tuple, let be a weak -subrepresentation of , and let . Then there exists a tuple and a map of -varieties over such that contains and is contained in .
Proof.
Note that is a subspace of . Applying Lemma 8.8, there exists a pure tuple , a -point of such that is -generic in , and a map of -varieties over such that . Thus .
Now, let be a -point of such that is -generic. We claim that . Since is -generic there is a sequence in such that converges to in the -topology (Proposition 2.2). We thus see that converges to in the -topology, and so is a -sequence. Applying , we see that converges to . Since is a weak subrepresentation, this implies that , as claimed.
Now, let be the set of -points such that is -generic. By the previous paragraph, we see that . Since is Zariski dense in by Lemma 8.7 and is a Zariski closed subset of , it follows that , as required. ∎
Lemma 8.10.
Let and be tuples, and let be a map of -varieties over . Then there exists a closed -subvariety of such that the following two conditions hold:
-
(a)
is defined by fiberwise linear equations, that is, for some -module quotient of
-
(b)
the -subspace of is exactly the -closure of the -span of .
Proof.
First suppose that is fiberwise linear. This means that is induced from a map of -modules . Let be the image of , and let . Let be the map obtained by reducing modulo the maximal ideal . Since is flat at (Corollary 3.7), it follows that the image of is . As is the dual of , we see that its image is the dual of , which is exactly . This completes the proof in the linear case. (In this case, taking the -closure is not necessary.)
We now treat the general case. The map corresponds to a map of -algebras . The image of under this map is contained in for some , where . The map then factors as
where is linear (i.e., induced from a map of -modules). Let be the corresponding factorization of . Let be the subvariety provided by the linear case, applied to . The map factors as
We know that the image of is exactly . The map is the canonical map, taking to . One easily sees that the -span of the image of is -dense. Since is -continuous, the result follows. ∎
Proof of Proposition 8.6.
Let be a weak -subrepresentation of , and fix an element . By Lemma 8.9, we can find a tuple and a map of -varieties over such that contains and is contained in . By Lemma 8.10, there is an -module such that is the -closure of the span of . Since is -closed, it follows that is contained in ; of course, it also contains . We thus see that contains and is contained in .
Now, let be a basis for , and for each pick a submodule of as in the previous paragraph, so that contains and is contained in . For a finite subset of , let . The form a descending family of submodules of . Since is an artinian object in the generic category (Theorem 4.10(a)), it follows that there is some finite subset such that is torsion for all . We thus have for all such . It follows that is contains and is contained in . This completes the proof. ∎
8.4. The main theorem
Maintain the setup from §8.3. The following is our main theorem on representations of :
Theorem 8.11.
The functor is an equivalence.
From the theorem, we see that all properties of transer to . 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 be a -module. Then the map
induced by is an isomorphism of partially ordered sets.
Proof.
We first show that is injective. First suppose that are -submodules of and . Then the containment of -modules induces an isomorphism modulo . It follows that has vanishing fiber at , and thus vanishes (see the proof of Proposition 6.2). Hence , and so . Now suppose that and are arbitary and . Then . Since , the previous case shows that , and so . By symmetry, we have . Thus is injective.
We now see that is strictly order-preserving. Indeed, let and be -submodules of . If then it is clear that . Conversely, if then , and so since is injective, whence .
To complete the proof, we must show that is surjective. If for a finite length polynomial representation , then this follows from Proposition 8.6. Suppose now that for an arbitrary polynomial representation . Write where is a directed set and has finite length, and put . Since is cocontinuous, we have . Let be a -subrepresentation of , and put . Since is a Grothendieck category, we have . By the finite length case, we have for a unique -submodule of . Since is strictly order-preserving, it follows that if . Thus the ’s form a directed system. Let . Again, by the cocontinuity of , we have .
Finally, suppose that is an arbitrary -module. Since is a Grothendieck abelian category, embeds into an injective object . Since is locally noetherian (Theorem 4.10(a)), is a direct sum of indecomposable injectives. Thus has the form for a polynomial representation (Theorem 4.10(b)). Now, suppose that is a -subrepresentation of . Since , the previous paragraph shows that for some -submodule of . Since is strictly order preserving, it follows that , which completes the proof. ∎
Proof of Theorem 8.11.
We first show that is essentially surjective. Thus let be a given polynomial representation of . By definition, there is some polynomial representation of and -subrepresentations such that . By Lemma 8.12, there exist -submodules such that . Thus , and so is essentially surjective.
We now prove that is full. Let and be -modules and let be a map of -subrepresentations. Let be the graph of . By Lemma 8.12, we have for a unique -submodule . The projection map becomes an isomorphism after applying , and is therefore an isomorphism since is exact and faithful. Thus is the graph of a morphism of -modules, and clearly .
We have already seen that 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 . Algebr. Represent. Theory 22 (2019), 867–893. arXiv:1605.02389v1
- [Na] Rohit Nagpal. FI-modules and the cohomology of modular -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 -, -, -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.