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

The spectrum of a twisted commutative algebra

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

A twisted commutative algebra is (for us) a commutative 𝐐\mathbf{Q}-algebra equipped with an action of the infinite general linear group. In such algebras the “𝐆𝐋\mathbf{GL}-prime” ideals assume the duties fulfilled by prime ideals in ordinary commutative algebra, and so it is crucial to understand them. Unfortunately, distinct 𝐆𝐋\mathbf{GL}-primes can have the same radical, which obstructs one from studying them geometrically. We show that this problem can be eliminated by working with super vector spaces: doing so provides enough geometry to distinguish 𝐆𝐋\mathbf{GL}-primes. This yields an effective method for analyzing 𝐆𝐋\mathbf{GL}-primes.

AS was supported by NSF DMS-1453893.

1. Introduction

A twisted commutative algebra (tca) is a commutative 𝐐\mathbf{Q}-algebra equipped with an action of the infinite general linear group 𝐆𝐋\mathbf{GL}_{\infty} under which it forms a polynomial representation; at least, that will be our definition for the moment. TCA’s have been effectively used to study asymptotic problems in algebra (see, for example, [CEF, DLL, ESS, Sn]), and are closely related to many particular objects of interest (such as EFW complexes [EFW], determinantal varieties, and representations of infinite rank groups [SS3]); moreover, all evidence so far points to a rich internal theory. It is therefore sensible to study these objects in more detail. While there have been many successes for particular tca’s [CEF, LR, NSS, NSS2, SS1, SS4, SS5], there has really only been one significant result to date for general tca’s, namely, Draisma’s topological noetherianity theorem [Dr]. In this paper, we take another step towards understanding the general case: we largely solve the problem of understanding the equivariant prime ideals of tca’s.

1.1. Equivariant commutative algebra

Let AA be a tca. One can then formulate equivariant analogs of many familiar concepts from commutative algebra111In fact, one can do this for commutative algebras in any tensor category.:

  • A 𝐆𝐋\mathbf{GL}-ideal of AA is an ideal that is 𝐆𝐋\mathbf{GL}-stable.

  • A 𝐆𝐋\mathbf{GL}-prime is a 𝐆𝐋\mathbf{GL}-ideal 𝔭\mathfrak{p} such that VW𝔭VW\subset\mathfrak{p} implies V𝔭V\subset\mathfrak{p} or W𝔭W\subset\mathfrak{p}, for subrepresentations V,WAV,W\subset A. Here VWVW denotes the image of the map VWAV\otimes W\to A.

  • The 𝐆𝐋\mathbf{GL}-radical of a 𝐆𝐋\mathbf{GL}-ideal II, denoted rad𝐆𝐋I\operatorname{rad}_{\mathbf{GL}}{I}, is the sum of all subrepresentations VV of AA such that VnIV^{n}\subset I for some nn. Here VnV^{n} denotes the image of the map VnAV^{\otimes n}\to A. This is equal to the intersection of the 𝐆𝐋\mathbf{GL}-primes containing II (Proposition 2.5).

  • The 𝐆𝐋\mathbf{GL}-spectrum of AA, denoted Spec𝐆𝐋(A)\operatorname{Spec}_{\mathbf{GL}}(A), is the set of 𝐆𝐋\mathbf{GL}-primes, endowed with the usual Zariski topology. The closed subsets of Spec𝐆𝐋(A)\operatorname{Spec}_{\mathbf{GL}}(A) correspond bijectively to 𝐆𝐋\mathbf{GL}-radical ideals.

One can keep going, but this is all we need for the moment.

In ordinary commutative algebra, prime ideals are of central importance; this is no less true of 𝐆𝐋\mathbf{GL}-primes in twisted commutative algebra. For instance: the support of an equivariant module is most naturally a subset of the 𝐆𝐋\mathbf{GL}-spectrum; the 𝐆𝐋\mathbf{GL}-primes can be used to generate the Grothendieck group of equivariant modules; and, under suitable hypotheses, one has an equivariant version of primary decomposition for 𝐆𝐋\mathbf{GL}-ideals. Therefore, to understand tca’s it is of crucial importance to understand their 𝐆𝐋\mathbf{GL}-primes.

Every 𝐆𝐋\mathbf{GL}-stable prime ideal of a tca is a 𝐆𝐋\mathbf{GL}-prime. However, the converse is not true. The following is an instructive example:

Example 1.1.

Let A=n02n(𝐐)A=\bigoplus_{n\geq 0}{\textstyle\bigwedge}^{2n}(\mathbf{Q}^{\infty}). Then AA is a tca, and a nice one at that: it is finitely generated and noetherian, in the equivariant sense. Suppose that VV and WW are non-zero subrepresentations of AA. Since exterior powers are irreducible, it follows that VV contains i(𝐐){\textstyle\bigwedge}^{i}(\mathbf{Q}^{\infty}) and WW contains j(𝐐){\textstyle\bigwedge}^{j}(\mathbf{Q}^{\infty}) for some ii and jj. Thus VWVW contains i+j(𝐐){\textstyle\bigwedge}^{i+j}(\mathbf{Q}^{\infty}). We have thus shown that if VV and WW are non-zero then so is VWVW. It follows that the zero ideal of AA is 𝐆𝐋\mathbf{GL}-prime; in other words, AA is 𝐆𝐋\mathbf{GL}-integral. ∎

This example is rather shocking when one first encounters it: every positive degree element of AA is nilpotent, and yet AA is a 𝐆𝐋\mathbf{GL}-domain! This example shows that (from our current perspective) tca’s do not have enough points to “see” their 𝐆𝐋\mathbf{GL}-primes: indeed, Spec(A)\operatorname{Spec}(A) is a single point, and thus cannot distinguish the two 𝐆𝐋\mathbf{GL}-primes (0)(0) and A+A_{+} of AA. Thus 𝐆𝐋\mathbf{GL}-domains appear to be divorced from geometry, which might diminish our hopes of understanding 𝐆𝐋\mathbf{GL}-primes; fortunately, however, this appearance is deceiving.

1.2. The key principle

The category Reppol(𝐆𝐋)\operatorname{Rep}^{\mathrm{pol}}(\mathbf{GL}_{\infty}) of polynomial representations of 𝐆𝐋\mathbf{GL}_{\infty} is equivalent to the category 𝐏𝐨𝐥\mathbf{Pol} of polynomial functors of rational vector spaces; the equivalence is obtained by evaluating a functor on 𝐐\mathbf{Q}^{\infty}. We can thus view a tca AA as an algebra object of 𝐏𝐨𝐥\mathbf{Pol}. From this perspective, A(𝐐)A(\mathbf{Q}^{\infty}) can be seen as the “incarnation” of AA in the category Reppol(𝐆𝐋)\operatorname{Rep}^{\mathrm{pol}}(\mathbf{GL}_{\infty}). However, polynomial functors can be evaluated on objects in any 𝐐\mathbf{Q}-linear tensor category. We can thus form the “super incarnation” A(𝐐|)A(\mathbf{Q}^{\infty|\infty}) of AA by evaluating on the super vector space 𝐐|\mathbf{Q}^{\infty|\infty}. This is an algebra object of Reppol(𝐆𝐋|)\operatorname{Rep}^{\mathrm{pol}}(\mathbf{GL}_{\infty|\infty}). We can now succinctly express the point of this paper:


Key principle. The geometry of the super incarnation of a tca is sufficiently rich to detect its 𝐆𝐋\mathbf{GL}-primes.

This principle is borne out in the theorems stated below.

Example 1.2.

Let AA be the tca from Example 1.1, regarded in 𝐏𝐨𝐥\mathbf{Pol}. Then A(𝐐0|s)=n0Sym2n(𝐐s)A(\mathbf{Q}^{0|s})=\bigoplus_{n\geq 0}\operatorname{Sym}^{2n}(\mathbf{Q}^{s}) is the second Veronese subring of Sym(𝐐s)\operatorname{Sym}(\mathbf{Q}^{s}), which is a domain of Krull dimension ss; in particular, its spectrum has plenty of points. For r>0r>0, the algebra A(𝐐r|s)A(\mathbf{Q}^{r|s}) is a nilpotent extension of A(𝐂0|s)A(\mathbf{C}^{0|s}). ∎

1.3. Main results

We now state several precise theorems. In what follows, AA is a tca (considered as a polynomial functor) and II and JJ are 𝐆𝐋\mathbf{GL}-ideals of AA.

Theorem A.

We have rad𝐆𝐋Irad𝐆𝐋J\operatorname{rad}_{\mathbf{GL}}{I}\subset\operatorname{rad}_{\mathbf{GL}}{J} if and only if radI(𝐐|)radJ(𝐐|)\operatorname{rad}{I(\mathbf{Q}^{\infty|\infty})}\subset\operatorname{rad}{J(\mathbf{Q}^{\infty|\infty})}.

In other words, the theorem says that rad𝐆𝐋Irad𝐆𝐋J\operatorname{rad}_{\mathbf{GL}}{I}\subset\operatorname{rad}_{\mathbf{GL}}{J} if and only if 𝒱(J(𝐐|))𝒱(I(𝐐|))\mathcal{V}(J(\mathbf{Q}^{\infty|\infty}))\subset\mathcal{V}(I(\mathbf{Q}^{\infty|\infty})); the latter condition is equivalent to 𝒱(J(𝐐r|s))𝒱(I(𝐐r|s))\mathcal{V}(J(\mathbf{Q}^{r|s}))\subset\mathcal{V}(I(\mathbf{Q}^{r|s})) for all rr and ss and thus (usually) reduces to a condition about finite dimensional algebraic varieties. Note that since we are only concerned with vanishing loci here, we can pass to the reduced quotient of A(𝐐r|s)A(\mathbf{Q}^{r|s}), which is an ordinary (non-super) commutative ring. We emphasize that the theorem is false if one uses only ordinary vector spaces, as Example 1.1 shows.

Theorem B.

The ideal rad𝐆𝐋I\operatorname{rad}_{\mathbf{GL}}{I} is 𝐆𝐋\mathbf{GL}-prime if and only if the ideal radI(𝐐|)\operatorname{rad}{I(\mathbf{Q}^{\infty|\infty})} is prime.

Once again, we can verify the latter condition on finite dimensional spaces. The theorem therefore reduces the problem of showing that rad𝐆𝐋(I)\operatorname{rad}_{\mathbf{GL}}(I) is 𝐆𝐋\mathbf{GL}-prime to showing that the algebraic varieties 𝒱(I(𝐐r|s))\mathcal{V}(I(\mathbf{Q}^{r|s})) are irreducible.

In the two remaining theorems, we require a finiteness condition: we assume that A0A_{0} is noetherian and that AA is finitely generated over A0A_{0}.

Theorem C.

We have the following:

  1. (a)

    AA has finitely many minimal 𝐆𝐋\mathbf{GL}-primes, say 𝔭1,,𝔭n\mathfrak{p}_{1},\ldots,\mathfrak{p}_{n};

  2. (b)

    A(𝐐|)A(\mathbf{Q}^{\infty|\infty}) has finitely many minimal primes, say 𝔮1,,𝔮m\mathfrak{q}_{1},\ldots,\mathfrak{q}_{m};

  3. (c)

    n=mn=m, and after applying a permutation we have 𝔮i=rad𝔭i(𝐐|)\mathfrak{q}_{i}=\operatorname{rad}{\mathfrak{p}_{i}(\mathbf{Q}^{\infty|\infty})} for all ii.

This theorem gives a useful way to find the minimal 𝐆𝐋\mathbf{GL}-primes, at least up to 𝐆𝐋\mathbf{GL}-radical, and thus the irreducible components of the 𝐆𝐋\mathbf{GL}-spectrum.

Theorem D.

The 𝐆𝐋\mathbf{GL}-spectrum Spec𝐆𝐋(A)\operatorname{Spec}_{\mathbf{GL}}(A) of AA is a noetherian topological space.

This theorem is a strengthening of Draisma’s topological noetherianity theorem: indeed, Draisma’s theorem only encompasses the 𝐆𝐋\mathbf{GL}-stable prime ideals, while this theorem accommodates all 𝐆𝐋\mathbf{GL}-primes. In fact, this theorem is easily deduced by combining Draisma’s theorem with our other theorems; we do not have a new proof of Draisma’s result.

1.4. An example

In §5, we examine the tca A=Sym(Sym2(𝐂))A=\operatorname{Sym}(\operatorname{Sym}^{2}(\mathbf{C}^{\infty})). Using our main theorems, we classify the 𝐆𝐋\mathbf{GL}-primes of AA (they are the rectangular ideals) and the 𝐆𝐋\mathbf{GL}-radical ideals of AA (they are the ideals generated by a single irreducible representation). This example provides a good illustration of how our main theorems allow one to understand the equivariant commutative algebra of AA through standard geometric means. It also provides a reconceptualization of sorts for the rectangular ideals: they constitute the equivariant spectrum of AA. These ideals have long been of interest, as they are symbolic powers of determinantal ideals.

1.5. Connection to other work

The notions of 𝐆𝐋\mathbf{GL}-prime and the 𝐆𝐋\mathbf{GL}-spectrum were discussed in [SS4, §3] (with slightly different terminology). However, that paper only works with so-called bounded tca’s, and for these 𝐆𝐋\mathbf{GL}-primes are the same as 𝐆𝐋\mathbf{GL}-stable primes, so the chief difficulties disapper. (Note that the tca in Example 1.1 is not bounded.)

In forthcoming joint work with Rohit Nagpal [NS1, NS2], we study the 𝔖\mathfrak{S}_{\infty}-primes in the ring 𝐂[x1,x2,]\mathbf{C}[x_{1},x_{2},\ldots] and manage to completely classify them. (Here 𝔖\mathfrak{S}_{\infty} denotes the infinite symmetric group.) However, there is no general theory of 𝔖\mathfrak{S}_{\infty}-primes yet.

One can, and usually does, define twisted commutative algebras without any reference to super vector spaces. However, our results show that one is essentially forced to use super vector spaces to fully understand tca’s. This is reminscent of Deligne’s theorem [De, Theorem 0.6], in which a natural class of tensor categories (defined without any reference to super vector spaces) is characterized using super objects. It would be interesting to find a direct connection between these results.

1.6. Outline

In §2, we review relevant background material. In §3, we prove the key theorem, which is a certain special case of Theorem A. In §4, we deduce the main theorems stated above. Finally, in §5, we work out the Sym(Sym2)\operatorname{Sym}(\operatorname{Sym}^{2}) case in detail.

Acknowledgments

We thank Steven Sam for helpful discussions, and allowing us to include some jointly conceived ideas in §5.

2. Background

2.1. Polynomial representations

We recall some background material on polynomial representations. We refer to [SS2] and [NSS2, §2.2] for more details.

Regard 𝐆𝐋n\mathbf{GL}_{n} as an algebraic group over 𝐐\mathbf{Q}. By an algebraic representation of 𝐆𝐋n\mathbf{GL}_{n}, we mean a comodule over 𝐐[𝐆𝐋n]\mathbf{Q}[\mathbf{GL}_{n}]; the dimension is not required to be finite. We regard 𝐆𝐋\mathbf{GL}_{\infty} as the inductive system of algebraic groups n1𝐆𝐋n\bigcup_{n\geq 1}\mathbf{GL}_{n}. An algebraic representation of 𝐆𝐋\mathbf{GL}_{\infty} is a vector space equipped with compatible algebraic representations of 𝐆𝐋n\mathbf{GL}_{n} for all nn. The basic example of such a representation is the standard representation 𝐐=n1𝐐n\mathbf{Q}^{\infty}=\bigcup_{n\geq 1}\mathbf{Q}^{n}. A polynomial representation of 𝐆𝐋\mathbf{GL}_{\infty} is one that occurs as a subquotient of a (possibly infinite) direct sum of tensor powers of the standard representation. We denote the category of such representations by Reppol(𝐆𝐋)\operatorname{Rep}^{\mathrm{pol}}(\mathbf{GL}_{\infty}). It is a 𝐐\mathbf{Q}-linear abelian category and closed under tensor products. The structure of this category is well-understood: it is semi-simple, and the simple objects are given by 𝐒λ(𝐐)\mathbf{S}_{\lambda}(\mathbf{Q}^{\infty}), where 𝐒λ\mathbf{S}_{\lambda} denotes the Schur functor associated to the partition λ\lambda. All simple objects are absolutely simple. Every polynomial representation VV admits a canonical decomposition V=λVλ𝐒λ(𝐐)V=\bigoplus_{\lambda}V_{\lambda}\otimes\mathbf{S}_{\lambda}(\mathbf{Q}^{\infty}) where the VλV_{\lambda} are multiplicity spaces. We endow VV with a grading by declaring the elements of 𝐒λ(𝐐)\mathbf{S}_{\lambda}(\mathbf{Q}^{\infty}) to have degree |λ||\lambda|.

By an algebraic representation of 𝐆𝐋r|s\mathbf{GL}_{r|s}, we mean a comodule over the Hopf superalgebra 𝐐[𝐆𝐋r|s]\mathbf{Q}[\mathbf{GL}_{r|s}]. We then define polynomial representations of 𝐆𝐋|\mathbf{GL}_{\infty|\infty} just as before, and denote the category by Reppol(𝐆𝐋|)\operatorname{Rep}^{\mathrm{pol}}(\mathbf{GL}_{\infty|\infty}). It is again semi-simple abelian and closed under tensor product, and the simple objects have the form 𝐒λ(𝐐|)\mathbf{S}_{\lambda}(\mathbf{Q}^{\infty|\infty}). Warning: if VV is a polynomial representation of 𝐆𝐋|\mathbf{GL}_{\infty|\infty} then V[1]V[1] (shift in super grading) is typically not a polynomial representation, according to our definition. This disagrees with the convention used in [NSS2].

Consider the category Fun(Vec,Vec)\operatorname{Fun}(\mathrm{Vec},\mathrm{Vec}) of endofunctors of the category Vec\mathrm{Vec} of 𝐐\mathbf{Q}-vector spaces. Let TnT_{n} be the functor given by Tn(V)=VnT_{n}(V)=V^{\otimes n}. We say that an object of Fun(Vec,Vec)\operatorname{Fun}(\mathrm{Vec},\mathrm{Vec}) is polynomial if it is a subquotient of a (possibly infinite) direct sum of TnT_{n}’s. We let 𝐏𝐨𝐥\mathbf{Pol} be the category of polynomial functors. It is semi-simple abelian and closed under tensor products. The simple objects are the Schur functors 𝐒λ\mathbf{S}_{\lambda}. We have equivalences of categories

Reppol(𝐆𝐋)\textstyle{\operatorname{Rep}^{\mathrm{pol}}(\mathbf{GL}_{\infty})}𝐏𝐨𝐥\textstyle{\mathbf{Pol}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Reppol(𝐆𝐋|)\textstyle{\operatorname{Rep}^{\mathrm{pol}}(\mathbf{GL}_{\infty|\infty})}F(𝐐)\textstyle{F(\mathbf{Q}^{\infty})}F\textstyle{F\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}F(𝐐|)\textstyle{F(\mathbf{Q}^{\infty|\infty})}

These equivalences are compatible with tensor products, and respect algebras, modules, and ideals within the categories.

We will at times need to evaluate polynomial functors on finite dimensional spaces. We recall the relevant result.

Proposition 2.1.

Let λ\lambda be a partition, and let r,s𝐍r,s\in\mathbf{N}. If λr+1s\lambda_{r+1}\leq s then 𝐒λ(𝐐r|s)\mathbf{S}_{\lambda}(\mathbf{Q}^{r|s}) is an absolutely irreducible representation of 𝐆𝐋r|s\mathbf{GL}_{r|s}; otherwise it vanishes. Moreover, the irreducible representations of 𝐆𝐋r|s\mathbf{GL}_{r|s} obtained in this way are mutually non-isomorphic.

Proof.

See [CW, §3.2.2]. ∎

We will require one additional simple result on polynomial representations. Define the width of a partition λ\lambda, denote w(λ)w(\lambda), to be λ1\lambda_{1}. Define the width of a polynomial functor FF, denoted w(F)w(F), to be the supremum of w(λ)w(\lambda) over those λ\lambda for which 𝐒λ\mathbf{S}_{\lambda} occurs in FF.

Proposition 2.2.

Let FF be a polynomial functor. The following are equivalent:

  1. (a)

    w(F)Nw(F)\leq N.

  2. (b)

    For every nn, every weight μ\mu appearing in F(𝐐n)F(\mathbf{Q}^{n}) satisfies μ1N\mu_{1}\leq N.

Proof.

It suffices to treat the case where F=𝐒λF=\mathbf{S}_{\lambda} is a simple object; we then have w(F)=λ1w(F)=\lambda_{1}. For n0n\gg 0, the 𝐆𝐋n\mathbf{GL}_{n} representation F(𝐐n)F(\mathbf{Q}^{n}) is irreducible with highest weight λ\lambda. Thus there is a weight μ\mu in F(𝐐n)F(\mathbf{Q}^{n}) with μ1=w(F)\mu_{1}=w(F) (namely, μ=λ\mu=\lambda). Furthermore, since λ\lambda is a highest weight, we have μ1λ1=w(F)\mu_{1}\leq\lambda_{1}=w(F) for any other weight μ\mu. This proves the result. ∎

2.2. Twisted commutative algebras

A twisted commutative algebra (tca) is a commutative algebra object in one of the three equivalent categories Reppol(𝐆𝐋)\operatorname{Rep}^{\mathrm{pol}}(\mathbf{GL}_{\infty}), Reppol(𝐆𝐋|)\operatorname{Rep}^{\mathrm{pol}}(\mathbf{GL}_{\infty|\infty}), or 𝐏𝐨𝐥\mathbf{Pol}. For the moment, we work in Reppol(𝐆𝐋)\operatorname{Rep}^{\mathrm{pol}}(\mathbf{GL}_{\infty}) to be definite.

Let AA be a tca in Reppol(𝐆𝐋)\operatorname{Rep}^{\mathrm{pol}}(\mathbf{GL}_{\infty}). Every polynomial representation carries a natural grading, and so AA is canonically graded. This is compatible with the ring structure, i.e., AA is a graded ring. In particular, we can regard AA as an algebra over is degree 0 piece A0A_{0}, which is an ordinary commutative ring.

By a “subrepresentation” of AA, we mean a 𝐐\mathbf{Q}-subspace that is stable by 𝐆𝐋\mathbf{GL}_{\infty}. In practice, AA will often be a 𝐂\mathbf{C}-algebra, but we use 𝐐\mathbf{Q}-subrepresentations nonetheless. One should think of a finite length subrepresentations of AA as providing an equivariant substitute for the concept of element (or perhaps finite sets of elements).

Suppose ABA\to B is a homomorphism of tca’s. We say that BB is finitely 𝐆𝐋\mathbf{GL}-generated over AA if there is some finite length subrepresentation EE of BB such that the natural map ASym(E)BA\otimes\operatorname{Sym}(E)\to B is surjective. We typically apply this in the case A=B0A=B_{0}.

2.3. Ideals in tca’s

Let AA be as above, i.e., a tca in Reppol(𝐆𝐋)\operatorname{Rep}^{\mathrm{pol}}(\mathbf{GL}_{\infty}). A 𝐆𝐋\mathbf{GL}-ideal of AA is a 𝐆𝐋\mathbf{GL}-stable ideal of AA. We say that a 𝐆𝐋\mathbf{GL}-ideal II is finitely 𝐆𝐋\mathbf{GL}-generated if there is a finite length subrepresentation EE of AA that generates II as an ideal. The sum and product of two finitely 𝐆𝐋\mathbf{GL}-generated ideals is again finitely 𝐆𝐋\mathbf{GL}-generated; for products, this relies on the fact that the tensor product of two finite length polynomial representations is again finite length.

We say that a 𝐆𝐋\mathbf{GL}-ideal 𝔭\mathfrak{p} is 𝐆𝐋\mathbf{GL}-prime if VW𝔭VW\subset\mathfrak{p} implies V𝔭V\subset\mathfrak{p} or W𝔭W\subset\mathfrak{p}, for subrepresentations VV and WW of AA; here VWVW denotes the image of the map VWAV\otimes W\to A. It is equivalent to ask the same condition with VV and WW finite length representations, or cyclic representations, or 𝐆𝐋\mathbf{GL}-idelas, or finitely generated 𝐆𝐋\mathbf{GL}-ideals. We define the 𝐆𝐋\mathbf{GL}-radical of a 𝐆𝐋\mathbf{GL}-ideal II, denoted rad𝐆𝐋I\operatorname{rad}_{\mathbf{GL}}{I} to be the sum of all subrepresentations VV or AA such that VnIV^{n}\subset I for some nn; again, one can use ideals in place of subrepresentations. We say that II is 𝐆𝐋\mathbf{GL}-radical if I=rad𝐆𝐋II=\operatorname{rad}_{\mathbf{GL}}{I}. We note that every 𝐆𝐋\mathbf{GL}-prime is 𝐆𝐋\mathbf{GL}-radical.

Remark 2.3.

A “prime 𝐆𝐋\mathbf{GL}-ideal” of AA is a 𝐆𝐋\mathbf{GL}-ideal of AA that is prime. This is potentially very different from a “𝐆𝐋\mathbf{GL}-prime ideal” of AA. Similarly, “radical 𝐆𝐋\mathbf{GL}-ideal” and “𝐆𝐋\mathbf{GL}-radical ideal” are potentially very different. ∎

We now establish some properties of the above definitions that are analogous to the classical situation.

Proposition 2.4.

Let II be a 𝐆𝐋\mathbf{GL}-ideal of AA and let EE be a finite length subrepresentation of rad𝐆𝐋(I)\operatorname{rad}_{\mathbf{GL}}(I). Then EnIE^{n}\subset I for some nn. Similarly, if JJ is a finitely 𝐆𝐋\mathbf{GL}-generated ideal contained in rad𝐆𝐋(I)\operatorname{rad}_{\mathbf{GL}}(I) then JnIJ^{n}\subset I for some nn.

Proof.

By definition, we can write rad𝐆𝐋(I)=iWi\operatorname{rad}_{\mathbf{GL}}(I)=\sum_{i\in\mathcal{I}}W_{i} where WiW_{i} is a subrepresentation of AA such that Win(i)IW_{i}^{n(i)}\subset I for some n(i)n(i). Since EE is contained in rad(I)\operatorname{rad}(I) and of finite length, there is some finite subset 𝒥\mathcal{J} of \mathcal{I} such that Ei𝒥WiE\subset\sum_{i\in\mathcal{J}}W_{i}. We thus have EnIE^{n}\subset I where n=#𝒥maxi𝒥n(i)n=\#\mathcal{J}\cdot\max_{i\in\mathcal{J}}n(i). For the ideal case, simply pick a finite length subrepresentation that generates and appeal to the previous argument. ∎

Proposition 2.5.

Let II be an ideal of AA. Then rad𝐆𝐋(I)\operatorname{rad}_{\mathbf{GL}}(I) is the intersection of the 𝐆𝐋\mathbf{GL}-primes containing II.

Proof.

Let 𝒫\mathcal{P} be the set of 𝐆𝐋\mathbf{GL}-primes containing II. Suppose 𝔭𝒫\mathfrak{p}\in\mathcal{P}. Let 𝔞\mathfrak{a} be a finitely 𝐆𝐋\mathbf{GL}-generated 𝐆𝐋\mathbf{GL}-ideal contained in rad𝐆𝐋(I)\operatorname{rad}_{\mathbf{GL}}(I). Then 𝔞nI𝔭\mathfrak{a}^{n}\subset I\subset\mathfrak{p} for some nn, and so 𝔞𝔭\mathfrak{a}\subset\mathfrak{p} since 𝔭\mathfrak{p} is 𝐆𝐋\mathbf{GL}-prime. Since this holds for all 𝔞\mathfrak{a}, we have rad𝐆𝐋(I)𝔭\operatorname{rad}_{\mathbf{GL}}(I)\subset\mathfrak{p}. Since this holds for all 𝔭\mathfrak{p}, we have rad𝐆𝐋(I)𝔭𝒫𝔭\operatorname{rad}_{\mathbf{GL}}(I)\subset\bigcap_{\mathfrak{p}\in\mathcal{P}}\mathfrak{p}.

We now prove the reverse inclusion. Let 𝔠\mathfrak{c} be a finitely 𝐆𝐋\mathbf{GL}-generated 𝐆𝐋\mathbf{GL}-ideal of AA not contained in rad(I)\operatorname{rad}(I). Let SS be the set of 𝐆𝐋\mathbf{GL}-ideals 𝔞\mathfrak{a} of AA such that no power of 𝔠\mathfrak{c} is contained in 𝔞\mathfrak{a}. Suppose that 𝔞1𝔞2\mathfrak{a}_{1}\subset\mathfrak{a}_{2}\subset\cdots is a chain in SS, and let 𝔞\mathfrak{a} be its union. Then 𝔞\mathfrak{a} belongs to SS too. Indeed, if 𝔠n\mathfrak{c}^{n} belongs to 𝔞\mathfrak{a} then, because it is finitely generated, it belongs to some 𝔞i\mathfrak{a}_{i}, a contradiction. Let 𝔭\mathfrak{p} be a maximal element of SS, which exists by Zorn’s lemma. We claim that 𝔭\mathfrak{p} is prime. Indeed, suppose 𝔞𝔟𝔭\mathfrak{a}\mathfrak{b}\subset\mathfrak{p}, but 𝔞,𝔟𝔭\mathfrak{a},\mathfrak{b}\not\subset\mathfrak{p}. Then 𝔭+𝔞\mathfrak{p}+\mathfrak{a} and 𝔭+𝔟\mathfrak{p}+\mathfrak{b} strictly contain 𝔭\mathfrak{p}, and therefore do not belong to SS. Thus 𝔠n𝔭+𝔞\mathfrak{c}^{n}\subset\mathfrak{p}+\mathfrak{a} and 𝔠m𝔭+𝔟\mathfrak{c}^{m}\subset\mathfrak{p}+\mathfrak{b} for some nn and mm. Thus 𝔠n+m(𝔭+𝔞)(𝔭+𝔟)𝔭+𝔞𝔟=𝔭\mathfrak{c}^{n+m}\subset(\mathfrak{p}+\mathfrak{a})(\mathfrak{p}+\mathfrak{b})\subset\mathfrak{p}+\mathfrak{a}\mathfrak{b}=\mathfrak{p}, a contradiction. It follows that 𝔠𝔭\mathfrak{c}\not\subset\mathfrak{p}, which completes the proof. Indeed, if 𝔭𝒫𝔭\bigcap_{\mathfrak{p}\in\mathcal{P}}\mathfrak{p} were strictly larger than rad𝐆𝐋I\operatorname{rad}_{\mathbf{GL}}{I}, then we could find a finitely 𝐆𝐋\mathbf{GL}-generated 𝔠\mathfrak{c} contained in 𝔭𝒫𝔭\bigcap_{\mathfrak{p}\in\mathcal{P}}\mathfrak{p} but not contained rad𝐆𝐋I\operatorname{rad}_{\mathbf{GL}}{I}, and the above argument would yield a contradiction. ∎

Proposition 2.6.

Every 𝐆𝐋\mathbf{GL}-prime of AA contains some minimal 𝐆𝐋\mathbf{GL}-prime of AA.

Proof.

An intersection of a descending chain of 𝐆𝐋\mathbf{GL}-primes is clearly 𝐆𝐋\mathbf{GL}-prime, so the claim follows from Zorn’s lemma. ∎

The above concepts (𝐆𝐋\mathbf{GL}-prime, 𝐆𝐋\mathbf{GL}-radical, etc.) are defined using only the language of the tensor category Reppol(𝐆𝐋)\operatorname{Rep}^{\mathrm{pol}}(\mathbf{GL}). It follows that the same definitions can be made in Reppol(𝐆𝐋|)\operatorname{Rep}^{\mathrm{pol}}(\mathbf{GL}_{\infty|\infty}) and 𝐏𝐨𝐥\mathbf{Pol}, and that the definitions agree on objects that correspond under the equivalences. Thus the above propositions also hold in all three settings. In fact, one can formulate and prove these results for commutative algebra objects in quite general tensor categories.

One important construction that cannot be formulated using only the language of the tensor category is the ordinary radical. Suppose AA is a tca in 𝐏𝐨𝐥\mathbf{Pol} and II is a 𝐆𝐋\mathbf{GL}-ideal in it. We can then consider their incarnations I(𝐐)A(𝐐)I(\mathbf{Q}^{\infty})\subset A(\mathbf{Q}^{\infty}) in Reppol(𝐆𝐋)\operatorname{Rep}^{\mathrm{pol}}(\mathbf{GL}_{\infty}), and form rad(I(𝐐))\operatorname{rad}(I(\mathbf{Q}^{\infty})). Similarly, we can consider their incarnations in Reppol(𝐆𝐋|)\operatorname{Rep}^{\mathrm{pol}}(\mathbf{GL}_{\infty|\infty}) and form rad(I(𝐐|))\operatorname{rad}(I(\mathbf{Q}^{\infty|\infty})). There is no reason to expect these two radicals to be comparable in any way (except in tautological ways, e.g., both contain II). In fact, the point of this paper is that they really are not comparable, and the construction is better behaved on the super side.

2.4. Minimal primes

We require the following result.

Proposition 2.7.

Let AA be a tca in Reppol(𝐆𝐋)\operatorname{Rep}^{\mathrm{pol}}(\mathbf{GL}_{\infty}) and let 𝔭\mathfrak{p} be a minimal prime of AA. Then 𝔭\mathfrak{p} is 𝐆𝐋\mathbf{GL}-stable.

Proof.

Consider the maps

A\textstyle{A\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Δ\scriptstyle{\Delta}A𝐐[𝐆𝐋n]\textstyle{A\otimes\mathbf{Q}[\mathbf{GL}_{n}]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}πid\scriptstyle{\pi\otimes\mathrm{id}}A/𝔭𝐐[𝐆𝐋n]\textstyle{A/\mathfrak{p}\otimes\mathbf{Q}[\mathbf{GL}_{n}]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}idϵ\scriptstyle{\mathrm{id}\otimes\epsilon}A/𝔭\textstyle{A/\mathfrak{p}}

where Δ\Delta is comultiplication, π:AA/𝔭\pi\colon A\to A/\mathfrak{p} is the quotient map, and ϵ:𝐐[𝐆𝐋n]𝐐\epsilon\colon\mathbf{Q}[\mathbf{GL}_{n}]\to\mathbf{Q} is the counit. The composition is equal to π\pi by the axioms for a comodule, and thus has kernel 𝔭\mathfrak{p}. We thus see that 𝔮=ker((πid)Δ)𝔭\mathfrak{q}=\ker((\pi\otimes\mathrm{id})\circ\Delta)\subset\mathfrak{p}. However, 𝔭\mathfrak{p} is prime and 𝐐[𝐆𝐋n]\mathbf{Q}[\mathbf{GL}_{n}] is a localization of a polynomial algebra over 𝐐\mathbf{Q}, and so A/𝔭𝐐[𝐆𝐋n]A/\mathfrak{p}\otimes\mathbf{Q}[\mathbf{GL}_{n}] is a domain. Thus 𝔮\mathfrak{q} is prime. Since 𝔭\mathfrak{p} is minimal, we must have 𝔭=𝔮\mathfrak{p}=\mathfrak{q}. Now, let xx be an element of 𝔭\mathfrak{p}, and write Δ(x)=i=1naibi\Delta(x)=\sum_{i=1}^{n}a_{i}\otimes b_{i} where aiAa_{i}\in A and bi𝐐[𝐆𝐋n]b_{i}\in\mathbf{Q}[\mathbf{GL}_{n}] are 𝐐\mathbf{Q}-linearly independent elements. Then 0=(πid)(Δ(x))=i=1nπ(ai)bi0=(\pi\otimes\mathrm{id})(\Delta(x))=\sum_{i=1}^{n}\pi(a_{i})\otimes b_{i}. Since the bib_{i} are linearly independent, it follows that π(ai)=0\pi(a_{i})=0 for all ii, and so ai𝔭a_{i}\in\mathfrak{p} for all ii. Thus Δ(𝔭)𝔭𝐐[𝐆𝐋n]\Delta(\mathfrak{p})\subset\mathfrak{p}\otimes\mathbf{Q}[\mathbf{GL}_{n}], and so 𝔭\mathfrak{p} is 𝐆𝐋n\mathbf{GL}_{n}-stable. Since this holds for all nn, it follows that 𝔭\mathfrak{p} is 𝐆𝐋\mathbf{GL}_{\infty}-stable. ∎

Remark 2.8.

The analog of this statement for tca’s in Reppol(𝐆𝐋|)\operatorname{Rep}^{\mathrm{pol}}(\mathbf{GL}_{\infty|\infty}) does not hold: the above proof fails since 𝐐[𝐆𝐋r|s]\mathbf{Q}[\mathbf{GL}_{r|s}] is not a domain. ∎

2.5. Radicals of 𝐆𝐋\mathbf{GL}-primes

We require the following result on 𝐆𝐋\mathbf{GL}-primes. See [SS2, §8.6] for some similar results.

Proposition 2.9.

Let AA be a tca in Reppol(𝐆𝐋|)\operatorname{Rep}^{\mathrm{pol}}(\mathbf{GL}_{\infty|\infty}) and let 𝔭\mathfrak{p} be a 𝐆𝐋\mathbf{GL}-prime of AA. Then rad(𝔭)\operatorname{rad}(\mathfrak{p}) is prime.

We require some preliminary work before proving the proposition. Let {ei}i\{e_{i}\}_{i\in\mathcal{I}} be a homogeneous basis for 𝐕=𝐐|\mathbf{V}=\mathbf{Q}^{\infty|\infty}. Given an element xx in a polynomial representation VV of 𝐆𝐋|\mathbf{GL}_{\infty|\infty} and a subset SS of \mathcal{I}, we say that xx has support contained in SS if VV can be embedded into a direct sum of tensor powers of 𝐐|\mathbf{Q}^{\infty|\infty} such that xx can be expressed using the basis elements in \mathcal{I}. One can define this more canonically by looking at the weight decomposition of xx. We define the support of an element XX of 𝔤𝔩|\mathfrak{gl}_{\infty|\infty} to be the set of indices ii such that XX has a non-zero entry in row ii or column ii. We say that an element of 𝒰(𝔤𝔩|)\mathcal{U}(\mathfrak{gl}_{\infty|\infty}) has support contained in SS if it can be expressed in terms of elements of 𝔤𝔩|\mathfrak{gl}_{\infty|\infty} having this property. We say that elements of representations or 𝒰(𝔤𝔩|)\mathcal{U}(\mathfrak{gl}_{\infty|\infty}) are disjoint if they have disjoint supports (or if they have supports contained in disjoint sets). For an element xAx\in A, we let x\langle x\rangle be the 𝐆𝐋\mathbf{GL}-ideal it generates.

Lemma 2.10.

Suppose x,yAx,y\in A are disjoint elements such that xy=0xy=0. Then xy=0\langle x\rangle\cdot\langle y\rangle=0.

Proof.

Let yy^{\prime} be an element of y\langle y\rangle that is disjoint from xx. Write y=ayy^{\prime}=ay where a𝒰(𝔤𝔩|)a\in\mathcal{U}(\mathfrak{gl}_{\infty|\infty}). Now, the support of aa may overlap with that of xx, that is, we may use auxiliary basis vectors in the process of building yy^{\prime} from yy. However, it does not matter which auxiliary basis vectors we use, so we can modify aa if necessary so that it is disjoint from xx. More rigorously, choose a permutation σ\sigma of \mathcal{I} that fixes the supports of yy and yy^{\prime}, and such that σaσ1\sigma a\sigma^{-1} is disjoint from xx. Then y=σaσ1yy^{\prime}=\sigma a\sigma^{-1}y and σaσ1\sigma a\sigma^{-1} is disjoint from xx. Now, applying aa to the expression xy=0xy=0, and using the fact that aa commutes with xx since it is disjoint from xx, we find x(ay)=0x(ay)=0, that is, xy=0xy^{\prime}=0.

Now let yy^{\prime} be an arbitrary element of VV. We can then write y=σy′′y^{\prime}=\sigma y^{\prime\prime} where σ\sigma is a permutation of \mathcal{I} and y′′y^{\prime\prime} is disjoint from xx. Let E𝔤𝔩×𝔤𝔩E\in\mathfrak{gl}_{\infty}\times\mathfrak{gl}_{\infty} act by σ\sigma on the support of y′′y^{\prime\prime} and 0 on the remaining basis vectors. Then Ex=0Ex=0 and Ey′′=yEy^{\prime\prime}=y^{\prime}. Since y′′y^{\prime\prime} is disjoint from xx, we have xy′′=0xy^{\prime\prime}=0 by the previous paragraph. Applying EE to this equation gives xy=0xy^{\prime}=0. This completes the proof. ∎

Lemma 2.11.

Let x,yAx,y\in A be super homogeneous elements satisfying xy=0xy=0. Then there exists n0n\geq 0 such that xny=0\langle x^{n}\rangle\cdot\langle y\rangle=0.

Proof.

We claim that for any a𝒰(𝔤𝔩|)a\in\mathcal{U}(\mathfrak{gl}_{\infty|\infty}) there exists n0n\geq 0 such that xnay=0x^{n}\cdot ay=0. This is clear for a=1a=1. Suppose now it is true for aa, and let us prove it for EaEa, with E𝔤𝔩|E\in\mathfrak{gl}_{\infty|\infty}. It suffices to treat the case where EE is super homogeneous. Let nn be such that xnay=0x^{n}\cdot ay=0. Applying EE, we find nxn1Exay±xnEay=0nx^{n-1}Ex\cdot ay\pm x^{n}\cdot Eay=0, where the sign depends on super degrees. Multiplying by xx and using the fact that xnay=0x^{n}\cdot ay=0, we find xn+1Eay=0x^{n+1}\cdot Eay=0. The claim now follows.

Now let yyy^{\prime}\in\langle y\rangle be disjoint from xx and generate y\langle y\rangle; for instance, one could take y=σyy^{\prime}=\sigma y for an appropriate permutation σ\sigma of \mathcal{I}. Since y=ayy^{\prime}=ay for some a𝒰(𝔤𝔩|)a\in\mathcal{U}(\mathfrak{gl}_{\infty|\infty}), the previous paragraph gives xny=0x^{n}y^{\prime}=0 for some nn. By Lemma 2.10, we find xny=0\langle x^{n}\rangle\cdot\langle y\rangle=0, and so the result follows. ∎

Proof of Proposition 2.9.

Passing to A/𝔭A/\mathfrak{p}, we assume 𝔭=0\mathfrak{p}=0 is 𝐆𝐋\mathbf{GL}-prime. We must show that rad(A)\operatorname{rad}(A) is prime. Since all odd elements of AA are nilpotent, we have A1rad(A)A_{1}\subset\operatorname{rad}(A). It thus suffices to show that if xyrad(A)xy\in\operatorname{rad}(A) with xx and yy even then xrad(A)x\in\operatorname{rad}(A) or yrad(A)y\in\operatorname{rad}(A). Thus let even elements xx and yy be given such that xyxy is nilpotent, say (xy)k=0(xy)^{k}=0. Since xx and yy are even, they commute, and so xkyk=0x^{k}y^{k}=0. By Lemma 2.11, there exists n0n\geq 0 such that xnkyk=0\langle x^{nk}\rangle\cdot\langle y^{k}\rangle=0. Since (0)(0) is 𝐆𝐋\mathbf{GL}-prime, it follows that xnk=0x^{nk}=0 or yk=0y^{k}=0. Thus either xx or yy is nilpotent, which completes the proof. ∎

2.6. Draisma’s theorem

Suppose that a group GG acts on a topological space XX. We say that XX is GG-noetherian if every descending chain of closed GG-stable subsets stabilizes. Draisma [Dr, Corollary 3] proved the following important theorem in this context:

Theorem 2.12.

Let AA be a tca in Reppol(𝐆𝐋)\operatorname{Rep}^{\mathrm{pol}}(\mathbf{GL}_{\infty}) such that A0A_{0} is noetherian and AA is finitely generated over A0A_{0}. Then Spec(A)\operatorname{Spec}(A) is 𝐆𝐋\mathbf{GL}_{\infty}-noetherian.

In fact, Draisma only states this theorem when A0A_{0} is finitely generated over a field, but a slight modification in his proof yields the above statement. Since it is not critical for this paper, we do not include details. We give a few corollaries of the theorem.

Corollary 2.13.

Let AA be as in Theorem 2.12. Then every ascending chain of radical 𝐆𝐋\mathbf{GL}-ideals in AA stabilizes.

Proof.

This follows since radical 𝐆𝐋\mathbf{GL}-ideals of AA correspond bijectively to 𝐆𝐋\mathbf{GL}-stable closed subsets of Spec(A)\operatorname{Spec}(A). ∎

Corollary 2.14.

Let AA be as in Theorem 2.12. Let Y=Spec(A)Y=\operatorname{Spec}(A) and let X=Y𝐆𝐋X=Y^{\mathbf{GL}} be the subset consisting of 𝐆𝐋\mathbf{GL}_{\infty}-stable prime ideals, endowed with the subspace topology. Then XX is a noetherian spectral space.

Proof.

Suppose that ZZ is a closed subset of XX. Then ZZ has the form WXW\cap X for some closed subset WW of YY. Since every point in ZZ is 𝐆𝐋\mathbf{GL}_{\infty}-invariant, it follows that Z=(gW)XZ=(gW)\cap X for any g𝐆𝐋g\in\mathbf{GL}_{\infty}. Thus Z=WXZ=W^{\prime}\cap X where W=g𝐆𝐋gWW^{\prime}=\bigcap_{g\in\mathbf{GL}_{\infty}}gW is a 𝐆𝐋\mathbf{GL}-stable closed subset of YY. It follows that Z=Z¯XZ=\overline{Z}\cap X, where Z¯\overline{Z} is the closure of ZZ in YY, since ZZ¯WZ\subset\overline{Z}\subset W^{\prime}.

Now suppose that Z2Z1\cdots\subset Z_{2}\subset Z_{1} is a descending chain of closed sets in XX. Then Z¯2Z¯1\cdots\subset\overline{Z}_{2}\subset\overline{Z}_{1} is a descending chain of 𝐆𝐋\mathbf{GL}-stable closed subsets of YY, and thus stabilizes by Theorem 2.12. Since Zi=Z¯iXZ_{i}=\overline{Z}_{i}\cap X, it follows that the original chain stabilizes too. Thus XX is noetherian.

Finally, let ZZ be an irreducible closed subset of XX. Then Z¯\overline{Z} is a 𝐆𝐋\mathbf{GL}-stable irreducible closed subset of YY. Its generic point is thus 𝐆𝐋\mathbf{GL}-stable, and therefore belongs to XX. One easily sees that it is the unique generic point for ZZ. Thus XX is sober. Since it is also noetherian, it is spectral. ∎

Corollary 2.15.

Let AA be as in Theorem 2.12. Then AA has finitely many minimal primes.

Proof.

Let XX be as in Corollary 2.14. Since XX is noetherian, it has finitely many irreducible components. Since it is sober, these components correspond to the minimal 𝐆𝐋\mathbf{GL}-stable prime ideals of XX. There are thus finitely many of these. However, if 𝔭\mathfrak{p} is any minimal prime then it is 𝐆𝐋\mathbf{GL}-stable by Proposition 2.7, and thus obviously a minimal 𝐆𝐋\mathbf{GL}-stable prime. The result follows. ∎

3. The key result

The following is the key theorem of this paper: it is the bridge that connects equivariant concepts to ordinary ones.

Theorem 3.1.

Let AA be a tca in 𝐏𝐨𝐥\mathbf{Pol} that is finitely generated over 𝐐\mathbf{Q}. The following are equivalent:

  1. (a)

    Every positive degree homogeneous element of A(𝐐r|s)A(\mathbf{Q}^{r|s}) is nilpotent, for all rr, ss.

  2. (b)

    The ideal A+A_{+} is nilpotent.

Proof.

It is clear that (b) implies (a). We prove the converse. We proceed by induction on the degree of generation of AA. Thus suppose AA is generated in degrees d\leq d and satisfies (a), and that the theorem is true for tca’s generated in degrees <d<d.

Before getting into the argument, we introduce a piece of notation. For a polynomial functor FF, we let FF^{\prime} be the polynomial functor defined by F(V)=F(𝐐V)F^{\prime}(V)=F(\mathbf{Q}\oplus V). We note that FF^{\prime} carries an action of 𝐆m\mathbf{G}_{m}, through its action on 𝐐\mathbf{Q}. If xF(𝐐r|s)=F(𝐐r+1|s)x\in F^{\prime}(\mathbf{Q}^{r|s})=F(\mathbf{Q}^{r+1|s}) is a weight vector for 𝐆𝐋r|s\mathbf{GL}_{r|s} of weight (a1,,ar;b1,,bs)(a_{1},\ldots,a_{r};b_{1},\ldots,b_{s}) and simultaneously a weight vector for 𝐆m\mathbf{G}_{m} of weight kk then it is also a weight vector for 𝐆𝐋r+1|s\mathbf{GL}_{r+1|s} of weight (k,a1,,ar;b1,,bs)(k,a_{1},\ldots,a_{r};b_{1},\ldots,b_{s}).

Let EA0AdE\subset A_{0}\oplus\cdots\oplus A_{d} be a subrepresentation generating AA. Let E=E0EdE^{\prime}=E^{\prime}_{0}\oplus\cdots\oplus E^{\prime}_{d} be the weight space decomposition for EE^{\prime} with respect to the 𝐆m\mathbf{G}_{m} action; note that EiE^{\prime}_{i} is a polynomial functor of degree di\leq d-i. Let BB be the subalgebra of AA^{\prime} generated by E1,,EdE^{\prime}_{1},\ldots,E^{\prime}_{d}, and let CC be the subalgebra generated by E0E^{\prime}_{0}. For i>0i>0, every weight vector of Ei(𝐐r|s)E^{\prime}_{i}(\mathbf{Q}^{r|s}) is a weight vector of E(𝐐r+1|s)E(\mathbf{Q}^{r+1|s}) of non-zero weight (since the first component of the weight is ii), and thus a positive degree homogenous element of A(𝐐r+1|s)A(\mathbf{Q}^{r+1|s}), and thus nilpotent. We thus see that the generators of B(𝐐r|s)B(\mathbf{Q}^{r|s}) are nilpotent, and so BB satisfies (a). Since BB is finitely generated in degrees <d<d, we can apply the inductive hypothesis to conclude that Bn=0B_{n}=0 for n0n\gg 0. Since the degree 0 generators of BB are also nilpotent, it follows that BB has finite length as a polynomial functor. In particular, only finitely many 𝐆m\mathbf{G}_{m} weights appear in BB; say that the largest one is NN.

Let TT be the maximal torus of 𝐆𝐋n+1\mathbf{GL}_{n+1}. Suppose that λ=(λ1,,λn+1)\lambda=(\lambda_{1},\ldots,\lambda_{n+1}) is a weight of TT that appears in A(𝐂n+1)=A(𝐐n)A(\mathbf{C}^{n+1})=A^{\prime}(\mathbf{Q}^{n}); we note that λ1\lambda_{1} records the action of 𝐆m\mathbf{G}_{m}. Since B(𝐐n)B(\mathbf{Q}^{n}) and C(𝐐n)C(\mathbf{Q}^{n}) generate A(𝐐n)A^{\prime}(\mathbf{Q}^{n}) are are TT-stable, it follows that λ\lambda can be written in the form μ+ν\mu+\nu where μ\mu is a weight of TT appearing in B(𝐐n)B(\mathbf{Q}^{n}) and ν\nu is one in C(𝐂n)C(\mathbf{C}^{n}). We have μ1N\mu_{1}\leq N by the definition of NN. Since CC is generated by E0E^{\prime}_{0}, on which 𝐆m\mathbf{G}_{m} acts trivially, we see that ν1=0\nu_{1}=0. Thus λ1N\lambda_{1}\leq N. Since this holds for all weights in A(𝐐n+1)A(\mathbf{Q}^{n+1}) for any nn, it follows that AA has width N\leq N by Proposition 2.2.

Decompose AA as λAλ𝐒λ\bigoplus_{\lambda}A_{\lambda}\otimes\mathbf{S}_{\lambda} where AλA_{\lambda} is a multiplicity space. We have just shown that AλA_{\lambda} is only non-zero when λ1N\lambda_{1}\leq N. Consider the superalgebra A(𝐐0|N)=λAλ𝐒λ(𝐐0|N)A(\mathbf{Q}^{0|N})=\bigoplus_{\lambda}A_{\lambda}\otimes\mathbf{S}_{\lambda}(\mathbf{Q}^{0|N}). This is finitely generated, and every positive degree element is nilpotent by assumption. Thus A(𝐐0|N)n=0A(\mathbf{Q}^{0|N})_{n}=0 for nn sufficiently large, say n>Mn>M. We thus have Aλ𝐒λ(𝐐0|N)=0A_{\lambda}\otimes\mathbf{S}_{\lambda}(\mathbf{Q}^{0|N})=0 for |λ|>M|\lambda|>M, and so Aλ=0A_{\lambda}=0: indeed, we know this already if λ1>N\lambda_{1}>N, and otherwise 𝐒λ(𝐐0|N)\mathbf{S}_{\lambda}(\mathbf{Q}^{0|N}) is non-zero by Proposition 2.1. It follows that An=0A_{n}=0 for n>Mn>M as well, and so AA satisfies (b). This proves the theorem. ∎

Example 3.2.

Suppose AA is generated over 𝐐\mathbf{Q} by E=Sym2E=\operatorname{Sym}^{2}. The space E(𝐐n)=E(𝐐n+1)E^{\prime}(\mathbf{Q}^{n})=E(\mathbf{Q}^{n+1}) has for a basis elements xi,jx_{i,j} with 1ijn+11\leq i\leq j\leq n+1. The degree of xi,jx_{i,j} under the 𝐆m\mathbf{G}_{m} action is simply the number of indices equal to 1. Thus E2(𝐐n)E^{\prime}_{2}(\mathbf{Q}^{n}) is spanned by x1,1x_{1,1}, while E1(𝐐n)E^{\prime}_{1}(\mathbf{Q}^{n}) is spanned by the x1,jx_{1,j} with 2j2\leq j, and E0(𝐐n)E^{\prime}_{0}(\mathbf{Q}^{n}) is spanned by the xi,jx_{i,j} with 2i,j2\leq i,j. Thus B(𝐐n)B(\mathbf{Q}^{n}) is generated by x1,1x_{1,1}, which is 𝐆𝐋n\mathbf{GL}_{n} invariant, and the x1,jx_{1,j} with 2jn+12\leq j\leq n+1, which generate a copy of the standard representation of 𝐆𝐋n\mathbf{GL}_{n}. We thus see that BB is generated by Sym0Sym1\operatorname{Sym}^{0}\oplus\operatorname{Sym}^{1}, which has degree 1\leq 1 as a polynomial functor. This shows that BB can have degree 0 generators even if AA does not. The algebra C(𝐐n)C(\mathbf{Q}^{n}) is generated by the xi,jx_{i,j} with 2i,j2\leq i,j, and so 𝐆m\mathbf{G}_{m} acts trivially on it. ∎

Corollary 3.3.

Let AA be an arbitrary tca in 𝐏𝐨𝐥\mathbf{Pol} and let EE be a finite length subrepresentation of AA. The following are equivalent:

  1. (a)

    Every element of E(𝐐r|s)E(\mathbf{Q}^{r|s}) is nilpotent, for all rr and ss.

  2. (b)

    The space EE is nilpotent, i.e., the map EnAE^{\otimes n}\to A is zero for some nn.

Proof.

Obviously (b) implies (a); we prove the converse. First suppose that EE is generated in positive degrees. Let BB be the sub tca of AA generated over 𝐐\mathbf{Q} by EE. Then B(𝐐r|s)B(\mathbf{Q}^{r|s}) is generated as a subalgebra of A(𝐐r|s)A(\mathbf{Q}^{r|s}) by E(𝐐r|s)E(\mathbf{Q}^{r|s}), and so every positive degree element is nilpotent. Thus, by Theorem 3.1, we see that B+B_{+} is nilpotent. Since EB+E\subset B_{+}, we see that it too is nilpotent.

In general, write E=E0E+E=E_{0}\oplus E_{+} where E0E_{0} is the degree 0 piece of EE and E+E_{+} is the sum of the positive degree pieces of EE. Then E0E_{0} is finite dimensional and every element is nilpotent, so E0E_{0} is nilpotent, and E+E_{+} is nilpotent by the previous paragraph. Thus EE is nilpotent. ∎

The proof of Theorem 3.1 is effective, in the following sense. For a finite length polynomial functor EE, let 𝒮k(E)\mathcal{S}_{k}(E) be the class of tca’s AA such that (i) AA contains a copy of EE that generates it over 𝐐\mathbf{Q}; and (ii) each weight space of E(𝐂r|s)E(\mathbf{C}^{r|s}) of non-zero weight admits a basis consisting of kk-nilpotent elements, for any rr and ss. Let ηk(E)\eta_{k}(E) be the supremum of maxdeg(A)\operatorname{maxdeg}(A) over A𝒮k(E)A\in\mathcal{S}_{k}(E), where maxdeg\operatorname{maxdeg} denotes the maximum non-zero degree. Theorem 3.1 simply states that maxdeg(A)\operatorname{maxdeg}(A) is finite for A𝒮k(E)A\in\mathcal{S}_{k}(E). In fact, the proof yields a bound on ηk(E)\eta_{k}(E). Let dd be the degree of EE, let EiE^{\prime}_{i} be as in the proof of the theorem, and let PP be the polynomial defined by P(n)=dimE(𝐐0|n)P(n)=\dim{E(\mathbf{Q}^{0|n})}. Examining the proof, one finds

ηk(E)kdP(dηk(E1Ed)+kddimE(𝐂)).\eta_{k}(E)\leq kd\cdot P(d\eta_{k}(E^{\prime}_{1}\oplus\cdots\oplus E^{\prime}_{d})+kd\dim{E(\mathbf{C})}).

This allows one to inductively obtain a bound on ηk(E)\eta_{k}(E) since the argument to ηk\eta_{k} on the right has smaller degree than EE. Making the rough approximation P(x)xdP(x)\approx x^{d}, one finds

ηk(E)i=1d(ik)d!/i!.\eta_{k}(E)\lessapprox\prod_{i=1}^{d}(ik)^{d!/i!}.

This upper bound is quite large; e.g., it is substantially larger than kd!22dk^{d!}2^{2^{d}}. We do not know how close it is to the true behavior of ηk(E)\eta_{k}(E).

4. The main theorems

We fix a tca AA in Reppol(𝐆𝐋|)\operatorname{Rep}^{\mathrm{pol}}(\mathbf{GL}_{\infty|\infty}) for this section. Consider the following condition:

  • (\ast)

    A0A_{0} is noetherian and AA is finitely generated over A0A_{0}.

We will sometimes require this condition, and sometimes not. Our goal now is to prove the main theorems stated in §1.3.

Proposition 4.1 (Theorem A).

Let II and JJ be 𝐆𝐋\mathbf{GL}-ideals of AA. Then Irad𝐆𝐋JI\subset\operatorname{rad}_{\mathbf{GL}}{J} if and only if IradJI\subset\operatorname{rad}{J}.

Note that Irad𝐆𝐋JI\subset\operatorname{rad}_{\mathbf{GL}}{J} if and only if rad𝐆𝐋Irad𝐆𝐋J\operatorname{rad}_{\mathbf{GL}}{I}\subset\operatorname{rad}_{\mathbf{GL}}{J}, and similarly for ordinary radicals, so this proposition is indeed equivalent to Theorem A.

Proof.

We may replace II with I+JI+J without changing either condition. We may then check the conditions after passing to A/JA/J. Thus we may simply assume from the outset that J=0J=0.

If Irad𝐆𝐋(A)I\subset\operatorname{rad}_{\mathbf{GL}}(A) then Irad(A)I\subset\operatorname{rad}(A), since we have a containment rad𝐆𝐋(A)rad(A)\operatorname{rad}_{\mathbf{GL}}(A)\subset\operatorname{rad}(A). Conversely, suppose that Irad(A)I\subset\operatorname{rad}(A). Let EE be a finite length subrepresentation of II. Then every element of EE is nilpotent, and so EE is nilpotent by Corollary 3.3. Thus Erad𝐆𝐋(A)E\subset\operatorname{rad}_{\mathbf{GL}}(A). Since this holds for all EE, it follows that Irad𝐆𝐋(A)I\subset\operatorname{rad}_{\mathbf{GL}}(A). ∎

We now introduce an auxiliary algebra that we will be helpful in what follows. Let B=A/rad(A)B=A/\operatorname{rad}(A). The ideal rad(B)\operatorname{rad}(B) is typically not 𝐆𝐋|\mathbf{GL}_{\infty|\infty} stable, but is clearly stable by 𝐆𝐋×𝐆𝐋𝐆𝐋|\mathbf{GL}_{\infty}\times\mathbf{GL}_{\infty}\subset\mathbf{GL}_{\infty|\infty}. It is therefore also stable by the diagonal subgroup 𝐆𝐋𝐆𝐋×𝐆𝐋\mathbf{GL}_{\infty}\subset\mathbf{GL}_{\infty}\times\mathbf{GL}_{\infty}, and so this acts on BB. Choose a surjection A0Sym(V)AA_{0}\otimes\operatorname{Sym}(V)\to A, for some representation VAV\subset A. The restriction of VV to the diagonal 𝐆𝐋\mathbf{GL}_{\infty} has the form W0W1[1]W_{0}\oplus W_{1}[1], where W0W_{0} and W1W_{1} are polynomial representations of 𝐆𝐋\mathbf{GL}_{\infty}, and [1][1] indicates the super grading. It follows that AA is a quotient of A0Sym(W0W1[1])A_{0}\otimes\operatorname{Sym}(W_{0}\oplus W_{1}[1]). Since BB has no odd part, we see that it is a quotient of A0Sym(W0)A_{0}\otimes\operatorname{Sym}(W_{0}). In other words, BB is a twisted commutative algebra in the category Reppol(𝐆𝐋)\operatorname{Rep}^{\mathrm{pol}}(\mathbf{GL}_{\infty}). If AA satisfies ()(\ast) then we can take VV to be a finite length representation. It follows from basic properties of Schur functors that W0W_{0} and W1W_{1} are then of finite length as well, and so BB is finitely generated over B0=A0B_{0}=A_{0}.

Proposition 4.2 (Theorem D).

Suppose ()(\ast) holds. Then Spec𝐆𝐋(A)\operatorname{Spec}_{\mathbf{GL}}(A) is noetherian.

Proof.

It suffices to show that every ascending chain of 𝐆𝐋\mathbf{GL}-radical ideals in AA stabilizes. Thus let I1I2I_{1}\subset I_{2}\subset\cdots be such a chain. Then rad(I1)rad(I2)\operatorname{rad}(I_{1})\subset\operatorname{rad}(I_{2}) is an ascending chain of 𝐆𝐋\mathbf{GL}_{\infty}-stable radical ideals of AA, and thus corresponds to an ascending chain of 𝐆𝐋\mathbf{GL}_{\infty}-stable radical ideals of BB. It therefore stabilizes by Corollary 2.13. By Proposition 4.1, it follows that the original chain stabilizes. ∎

Proposition 4.3 (Theorem C).

Suppose ()(\ast) holds.

  1. (a)

    AA has finitely many minimal 𝐆𝐋\mathbf{GL}-primes, say 𝔭1,,𝔭n\mathfrak{p}_{1},\ldots,\mathfrak{p}_{n};

  2. (b)

    AA has finitely many minimal primes, say 𝔮1,,𝔮m\mathfrak{q}_{1},\ldots,\mathfrak{q}_{m};

  3. (c)

    n=mn=m, and after applying a permutation we have 𝔮i=rad(𝔭i)\mathfrak{q}_{i}=\operatorname{rad}(\mathfrak{p}_{i}) for all ii.

Proof.

(a) This is an immediate consequence of Proposition 4.2.

(b) The minimal primes of AA correspond bijectively to those of BB, and there are finitely many of these by Corollary 2.15.

(c) Consider a minimal prime 𝔮i\mathfrak{q}_{i}. We have

𝔭1𝔭n=rad𝐆𝐋(A)rad(A)𝔮i.\mathfrak{p}_{1}\cap\cdots\cap\mathfrak{p}_{n}=\operatorname{rad}_{\mathbf{GL}}(A)\subset\operatorname{rad}(A)\subset\mathfrak{q}_{i}.

Since 𝔮i\mathfrak{q}_{i} is prime, it follows that 𝔭j𝔮i\mathfrak{p}_{j}\subset\mathfrak{q}_{i} for some jj. Thus rad(𝔭j)𝔮i\operatorname{rad}(\mathfrak{p}_{j})\subset\mathfrak{q}_{i}. Since rad(𝔭j)\operatorname{rad}(\mathfrak{p}_{j}) is prime (Proposition 2.9) and 𝔮i\mathfrak{q}_{i} is a minimal prime, we have rad(𝔭j)=𝔮i\operatorname{rad}(\mathfrak{p}_{j})=\mathfrak{q}_{i}. By Proposition 4.1, it follows that jj is unique: indeed, if rad(𝔭j)=rad(𝔭k)\operatorname{rad}(\mathfrak{p}_{j})=\operatorname{rad}(\mathfrak{p}_{k}) then 𝔭j=𝔭k\mathfrak{p}_{j}=\mathfrak{p}_{k} and so j=kj=k.

To complete the proof, it suffices to show that rad(𝔭j)\operatorname{rad}(\mathfrak{p}_{j}) is a minimal prime for all jj. We know that rad(𝔭j)\operatorname{rad}(\mathfrak{p}_{j}) is prime. It therefore contains some minimal prime 𝔮i\mathfrak{q}_{i}. We have shown that 𝔮i=rad(𝔭k)\mathfrak{q}_{i}=\operatorname{rad}(\mathfrak{p}_{k}) for some kk. Thus 𝔭krad(𝔭j)\mathfrak{p}_{k}\subset\operatorname{rad}(\mathfrak{p}_{j}), and so 𝔭krad𝐆𝐋(𝔭j)=𝔭j\mathfrak{p}_{k}\subset\operatorname{rad}_{\mathbf{GL}}(\mathfrak{p}_{j})=\mathfrak{p}_{j} by Proposition 4.1. Since 𝔭j\mathfrak{p}_{j} is a minimal 𝐆𝐋\mathbf{GL}-prime, it follows that 𝔭j=𝔭k\mathfrak{p}_{j}=\mathfrak{p}_{k}. Hence rad(𝔭j)=𝔮i\operatorname{rad}(\mathfrak{p}_{j})=\mathfrak{q}_{i} is a minimal prime. ∎

Proposition 4.4 (Theorem B).

Let II be a 𝐆𝐋\mathbf{GL}-ideal of AA. Then rad𝐆𝐋(I)\operatorname{rad}_{\mathbf{GL}}(I) is 𝐆𝐋\mathbf{GL}-prime if and only if rad(I)\operatorname{rad}(I) is prime.

Proof.

If II is 𝐆𝐋\mathbf{GL}-prime then rad(I)\operatorname{rad}(I) is prime by Proposition 2.9. Now suppose that rad(I)\operatorname{rad}(I) is prime and AA satisfies ()(\ast). Then A/IA/I has a unique minimal prime, and therefore a unique minimal 𝐆𝐋\mathbf{GL}-prime by Proposition 4.3. Thus there is a unique minimal 𝐆𝐋\mathbf{GL}-prime 𝔭\mathfrak{p} over II, and so rad𝐆𝐋(I)=𝔭\operatorname{rad}_{\mathbf{GL}}(I)=\mathfrak{p} is 𝐆𝐋\mathbf{GL}-prime. Finally, suppose that rad(I)\operatorname{rad}(I) is prime and AA is arbitrary. Write A=AiA=\bigcup A_{i} where {Ai}\{A_{i}\} is a directed family of sub tca’s satisfying ()(\ast). Then rad(IAi)=rad(I)Ai\operatorname{rad}(I\cap A_{i})=\operatorname{rad}(I)\cap A_{i} is prime, and so rad𝐆𝐋(IAi)=rad𝐆𝐋(I)Ai\operatorname{rad}_{\mathbf{GL}}(I\cap A_{i})=\operatorname{rad}_{\mathbf{GL}}(I)\cap A_{i} is 𝐆𝐋\mathbf{GL}-prime by the previous case. Since rad𝐆𝐋(I)Ai\operatorname{rad}_{\mathbf{GL}}(I)\cap A_{i} is prime for all ii, it follows that rad𝐆𝐋(I)\operatorname{rad}_{\mathbf{GL}}(I) is prime by Lemma 4.5 below. ∎

Lemma 4.5.

Let AA be a tca, and suppose that A=AiA=\bigcup A_{i} for some directed family {Ai}\{A_{i}\} of sub tca’s. Let 𝔭\mathfrak{p} be a 𝐆𝐋\mathbf{GL}-ideal of AA. If 𝔭Ai\mathfrak{p}\cap A_{i} is 𝐆𝐋\mathbf{GL}-prime for all ii then 𝔭\mathfrak{p} is 𝐆𝐋\mathbf{GL}-prime.

Proof.

Let VV and WW be finite length subrepresentations of AA such that VW𝔭VW\subset\mathfrak{p}. Since VV adn WW are finite length, there is some ii such that VV and WW are contained in AiA_{i}. Thus VW𝔭AiVW\subset\mathfrak{p}\cap A_{i}. Since 𝔭Ai\mathfrak{p}\cap A_{i} is 𝐆𝐋\mathbf{GL}-prime, it follows that V𝔭AiV\subset\mathfrak{p}\cap A_{i} or W𝔭AiW\subset\mathfrak{p}\cap A_{i}. Thus V𝔭V\subset\mathfrak{p} or W𝔭W\subset\mathfrak{p}, and so 𝔭\mathfrak{p} is 𝐆𝐋\mathbf{GL}-prime. ∎

5. An example

Let AA be the tca in 𝐏𝐨𝐥\mathbf{Pol} given by A(V)=𝐂Sym(Sym2(V))A(V)=\mathbf{C}\otimes\operatorname{Sym}(\operatorname{Sym}^{2}(V)). Our goal is to classify the 𝐆𝐋\mathbf{GL}-prime and 𝐆𝐋\mathbf{GL}-radical ideals of AA.

5.1. The ideal lattice of AA

The decomposition of AA into irreducibles is well-known:

A=𝐒2λ,A=\bigoplus\mathbf{S}_{2\lambda},

where the sum is over all partitions λ\lambda, and 2λ=(2λ1,2λ2,)2\lambda=(2\lambda_{1},2\lambda_{2},\ldots). See, for example, [M, §I.5, Example 5]. For a partition λ\lambda, let IλI_{\lambda} be the ideal of AA generated by 𝐒2λ\mathbf{S}_{2\lambda}. The following result determines the ideal structure of AA:

Proposition 5.1.

The ideal IλI_{\lambda} is the sum of those 𝐒2μ\mathbf{S}_{2\mu} for which λμ\lambda\subset\mu. In particular, IλIμI_{\lambda}\subset I_{\mu} if and only if μλ\mu\subset\lambda.

Proof.

This was originally proved in [Ab], but that is a difficult reference to obtain. The analogous result for Sym(2)\operatorname{Sym}({\textstyle\bigwedge}^{2}) is proved in [AdF, Theorem 3.1]. That case actually implies this one, since Sym(Sym2(V))=Sym(2(V[1]))\operatorname{Sym}(\operatorname{Sym}^{2}(V))=\operatorname{Sym}({\textstyle\bigwedge}^{2}(V[1])). A complete proof in the case where λ\lambda is a rectangle also appears in [NSS2, Corollary 2.8]. A closely related result appears in [CEP, Theorem 4.1]. ∎

The ideals generated by rectangular shapes will be particularly important, so we introduce some notation for them. We let ρ(r,s)\rho(r,s) be the partition with rr rows each of length ss; thus the Young diagram for ρ(r,s)\rho(r,s) is an r×sr\times s rectangle. We let Ir,s=Iρ(r,s)I_{r,s}=I_{\rho(r,s)}. If r=0r=0 or s=0s=0 then ρ(r,s)\rho(r,s) is an empty partition and Ir,sI_{r,s} is the unit ideal.

Let λ\lambda be a partition. By a corner of λ\lambda we mean a pair (r,s)(r,s) such that λ\lambda has a box in the rrth row and ssth column, but no box below or to the right of this one. For example, in the following Young diagram the corners have been shaded:

\ydiagram[(white)]6,5,4,2,1,1[(gray)]6,6,5,3,2,1,1\ydiagram[*(white)]{6,5,4,2,1,1}*[*(gray)]{6,6,5,3,2,1,1}

The following observation illustrates the importance of the rectangular ideals.

Proposition 5.2.

Let 𝒞\mathcal{C} be the set of corners of λ\lambda. Then Iλ=(r,s)𝒞Ir,sI_{\lambda}=\bigcap_{(r,s)\in\mathcal{C}}I_{r,s}.

Proof.

Let μ\mu be a partition. We have

𝐒2μ(r,s)𝒞Ir,s\displaystyle\mathbf{S}_{2\mu}\subset\bigcap\nolimits_{(r,s)\in\mathcal{C}}I_{r,s} ρ(r,s)μfor all (r,s)𝒞\displaystyle\iff\rho(r,s)\subset\mu\ \text{for all $(r,s)\in\mathcal{C}$}
(r,s)𝒞ρ(r,s)μ\displaystyle\iff\bigcup\nolimits_{(r,s)\in\mathcal{C}}\rho(r,s)\subset\mu
λμ\displaystyle\iff\lambda\subset\mu
𝐒2μIμ.\displaystyle\iff\mathbf{S}_{2\mu}\subset I_{\mu}.

The first and last step follow from Proposition 5.1, the second is trivial, and the third is simply the observation that λ=(r,s)𝒞ρ(r,s)\lambda=\bigcup_{(r,s)\in\mathcal{C}}\rho(r,s). The result thus follows. ∎

5.2. The variety XX

For V=EF[1]V=E\oplus F[1], let B(V)=Sym(Sym2(E)2(F))B(V)=\operatorname{Sym}(\operatorname{Sym}^{2}(E)\oplus{\textstyle\bigwedge}^{2}(F)). Note that B(V)=A(V)/rad(A(V))B(V)=A(V)/\operatorname{rad}(A(V)). We regard BB as a 2-variable tca (in the variables EE and FF). Let X(V)=Spec(B(V))X(V)=\operatorname{Spec}(B(V)), which we identify with Sym2(E)×2(F)\operatorname{Sym}^{2}(E)^{*}\times{\textstyle\bigwedge}^{2}(F)^{*}. By a “closed subvariety” of XX, we mean a subfunctor YY of XX such that Y(V)Y(V) is a closed subvariety of X(V)X(V) for all finite dimensional VV. Closed subvarieties of XX correspond bijectively to 𝐆𝐋×𝐆𝐋\mathbf{GL}\times\mathbf{GL} stable radical ideals of BB. For r,s𝐍{}r,s\in\mathbf{N}\cup\{\infty\}, let Xr,s(V)X(V)X_{r,s}(V)\subset X(V) be the locus of pairs (ω,η)(\omega,\eta) such that rank(ω)r\operatorname{rank}(\omega)\leq r and rank(η)2s\operatorname{rank}(\eta)\leq 2s. Then Xr,sX_{r,s} is a closed subvariety of XX in the above sense. We now show that these account for essentially all examples:

Proposition 5.3.

Let YY be a closed subvariety of XX. Then there is a finite subset 𝒞\mathcal{C} of (𝐍{})2(\mathbf{N}\cup\{\infty\})^{2} such that Y=(r,s)𝒞Xr,sY=\bigcup_{(r,s)\in\mathcal{C}}X_{r,s}.

Proof.

By the rank of a point (ω,η)X(E,F)(\omega,\eta)\in X(E,F), we mean the pair (rank(ω),12rank(η))(\operatorname{rank}(\omega),\tfrac{1}{2}\operatorname{rank}(\eta)). Let S𝐍2S\subset\mathbf{N}^{2} be the set of pairs (r,s)(r,s) such that Y(E,F)Y(E,F) has a point of rank (r,s)(r,s) for some EE and FF. We claim that (r,s)S(r,s)\in S if and only if Xr,sYX_{r,s}\subset Y. It is clear that Xr,sYX_{r,s}\subset Y implies (r,s)S(r,s)\in S. Conversely, suppose that (r,s)S(r,s)\in S. Then there exists some (ω0,η0)Y(E0,F0)(\omega_{0},\eta_{0})\in Y(E_{0},F_{0}) of rank (r,s)(r,s) for some E0E_{0} and F0F_{0}. Let (ω,η)Xr,s(E,F)(\omega,\eta)\in X_{r,s}(E,F) be given. By basic linear algebra, there are linear maps φ:EE0\varphi\colon E\to E_{0} and ψ:FF0\psi\colon F\to F_{0} such that ω=φ(ω0)\omega=\varphi^{*}(\omega_{0}) and η=ψ(η0)\eta=\psi^{*}(\eta_{0}). Thus the map X(E0,F0)X(E,F)X(E_{0},F_{0})\to X(E,F) defined by (φ,ψ)(\varphi,\psi) carries (ω0,η0)(\omega_{0},\eta_{0}) to (ω,η)(\omega,\eta). Since (ω0,η0)Y(E0,F0)(\omega_{0},\eta_{0})\in Y(E_{0},F_{0}) and YY is a subfunctor of XX, it follows that (ω,η)Y(E,F)(\omega,\eta)\in Y(E,F). This proves the claim.

It now follows that if (r,s)S(r,s)\in S and (r,s)(r,s)(r^{\prime},s^{\prime})\leq(r,s) then (r,s)S(r^{\prime},s^{\prime})\in S, where here (r,s)(r,s)(r^{\prime},s^{\prime})\leq(r,s) means rrr^{\prime}\leq r and sss^{\prime}\leq s. A simple combinatorial argument now shows that there is a finite subset 𝒞\mathcal{C} of (𝐍{})2(\mathbf{N}\cup\{\infty\})^{2} such that (r,s)S(r^{\prime},s^{\prime})\in S if and only if (r,s)(r,s)(r^{\prime},s^{\prime})\leq(r,s) for some (r,s)𝒞(r,s)\in\mathcal{C}. It follows that YY is the union of the Xr,sX_{r,s} with (r,s)𝒞(r,s)\in\mathcal{C}, which proves the proposition. ∎

Corollary 5.4.

Any irreducible closed subvariety of XX is one of the Xr,sX_{r,s}.

5.3. The vanishing locus of Ir,sI_{r,s}

The goal of this section is to prove the following:

Proposition 5.5.

We have 𝒱(Ir+1,s+1(V))=Xr,s(V)\mathcal{V}(I_{r+1,s+1}(V))=X_{r,s}(V) for r,s0r,s\geq 0.

We break the proof into two lemmas.

Lemma 5.6.

We have the following:

  1. (a)

    We have rad(Ir+1,s+1(E))=Ir+1,1(E)\operatorname{rad}(I_{r+1,s+1}(E))=I_{r+1,1}(E) in Sym(Sym2(E))\operatorname{Sym}(\operatorname{Sym}^{2}(E)).

  2. (b)

    We have rad(Ir+1,s+1(F[1]))=I1,s+1(F)\operatorname{rad}(I_{r+1,s+1}(F[1]))=I_{1,s+1}(F) in Sym(2(F))\operatorname{Sym}({\textstyle\bigwedge}^{2}(F)).

  3. (c)

    We have 𝒱(Ir+1,s+1(V))Xr,s(V)\mathcal{V}(I_{r+1,s+1}(V))\subset X_{r,s}(V).

Proof.

(i) This is proved in [Ab], but we include an argument (due to Steven Sam) to be self-contained. By Proposition 5.1, we have Ir+1,s+1(E)Ir+1,1(E)I_{r+1,s+1}(E)\subset I_{r+1,1}(E). We thus have a surjection π:A(E)/Ir+1,s+1(E)A(E)/Ir+1,1(E)\pi\colon A(E)/I_{r+1,s+1}(E)\to A(E)/I_{r+1,1}(E). By Proposition 5.1, we have

A(E)/Ir+1,1(E)\displaystyle A(E)/I_{r+1,1}(E) =(λ)r𝐒2λ(E),\displaystyle=\bigoplus_{\ell(\lambda)\leq r}\mathbf{S}_{2\lambda}(E),
A(E)/Ir+1,s+1(E)\displaystyle A(E)/I_{r+1,s+1}(E) =(λ)r𝐒2λ(E)+w(λ)s𝐒2λ(E)\displaystyle=\bigoplus_{\ell(\lambda)\leq r}\mathbf{S}_{2\lambda}(E)+\bigoplus_{w(\lambda)\leq s}\mathbf{S}_{2\lambda}(E)

We thus see that ker(π)\ker(\pi) is the sum of those 𝐒λ(E)\mathbf{S}_{\lambda}(E)’s with w(λ)sw(\lambda)\leq s and (λ)>r\ell(\lambda)>r. However, 𝐒λ(E)=0\mathbf{S}_{\lambda}(E)=0 if (λ)>dim(E)\ell(\lambda)>\dim(E), and so there are only finitely many relevant such λ\lambda. Thus ker(π)\ker(\pi) is finite dimensional, and therefore nilpotent (since it is homogeneous and consists of positive degree elements), and so the claim follows.

(ii) This is proved in [AdF, Theorem 5.1]. We can also argue analogously to the above.

(iii) Since Ir+1,s+1(V)I_{r+1,s+1}(V) contains both Ir+1,s+1(E)I_{r+1,s+1}(E) and Ir+1,s+1(F[1])I_{r+1,s+1}(F[1]), we see that its radical contains both Ir+1,1(E)I_{r+1,1}(E) and I1,s+1(F[1])I_{1,s+1}(F[1]) by (i) and (ii), and thus the extensions of these ideals to A(V)A(V). It follows that 𝒱(Ir+1,s+1(V))\mathcal{V}(I_{r+1,s+1}(V)) is contained in the intersection of 𝒱(Ir+1,1(E)e)\mathcal{V}(I_{r+1,1}(E)^{e}) and 𝒱(I1,s+1(F[1])e)\mathcal{V}(I_{1,s+1}(F[1])^{e}). Now, Ir+1,1(E)I_{r+1,1}(E) is the classical determinantal ideal: its vanishing locus in Sym2(E)\operatorname{Sym}^{2}(E)^{*} consists of those forms rank r\leq r. The vanishing locus of Ir+1,1(E)eI_{r+1,1}(E)^{e} is thus Xr,(V)X_{r,\infty}(V). Similarly, the vanishing locus of I1,s+1(F[1])eI_{1,s+1}(F[1])^{e} is X,s(V)X_{\infty,s}(V). We thus find that 𝒱(Ir+1,s+1(V))\mathcal{V}(I_{r+1,s+1}(V)) is contained in Xr,(V)X,s(V)=Xr,s(V)X_{r,\infty}(V)\cap X_{\infty,s}(V)=X_{r,s}(V). ∎

Lemma 5.7.

We have Xr,s(V)V(Ir+1,s+1(V))X_{r,s}(V)\subset V(I_{r+1,s+1}(V)).

Proof.

Let α\alpha be a non-degenerate symmetric bilinear form on 𝐂r\mathbf{C}^{r}. We obtain a natural map

Sym2(E)Sym2(𝐂r)Sym2(E)Sym2(𝐂rE),\operatorname{Sym}^{2}(E)\to\operatorname{Sym}^{2}(\mathbf{C}^{r})\otimes\operatorname{Sym}^{2}(E)\to\operatorname{Sym}^{2}(\mathbf{C}^{r}\otimes E),

where the first map comes from the inclusion 𝐂Sym2(𝐂r)\mathbf{C}\to\operatorname{Sym}^{2}(\mathbf{C}^{r}) of the 𝐎r\mathbf{O}_{r} invariant provided by α\alpha, and the second map comes from the Cauchy decomposition. The above map induces an algebra homomorphism f:Sym(Sym2(E))Sym(𝐂rE)f\colon\operatorname{Sym}(\operatorname{Sym}^{2}(E))\to\operatorname{Sym}(\mathbf{C}^{r}\otimes E). The induced map on spectra f:Hom(E,𝐂r)Sym2(E)f^{*}\colon\operatorname{Hom}(E,\mathbf{C}^{r})\to\operatorname{Sym}^{2}(E)^{*} takes a linear map φ:E𝐂r\varphi\colon E\to\mathbf{C}^{r} to the form φ(α)\varphi^{*}(\alpha) on EE, and thus surjects onto the locus of forms of rank r\leq r.

Now let β\beta be a non-degenerate symplectic form on 𝐂2s\mathbf{C}^{2s}. A similar construction yields a homomorphism g:Sym(2(F))Sym(𝐂2sF)g\colon\operatorname{Sym}({\textstyle\bigwedge}^{2}(F))\to\operatorname{Sym}(\mathbf{C}^{2s}\otimes F) such that gg^{*} surjects onto the locus of forms in 2(F){\textstyle\bigwedge}^{2}(F)^{*} of rank 2s\leq 2s.

Finally, let γ\gamma be the non-degenerate orthosymplectic form on 𝐂r|2s\mathbf{C}^{r|2s} that restricts to α\alpha and β\beta on the even and odd pieces. Once again, we get a natural algebra homomorphism h:Sym(Sym2(V))Sym(𝐂r|2sV)h\colon\operatorname{Sym}(\operatorname{Sym}^{2}(V))\to\operatorname{Sym}(\mathbf{C}^{r|2s}\otimes V). One easily verifies that the following square commutes:

Sym(Sym2(V))\textstyle{\operatorname{Sym}(\operatorname{Sym}^{2}(V))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}h\scriptstyle{h}Sym(Sym2(E))Sym(2(F))\textstyle{\operatorname{Sym}(\operatorname{Sym}^{2}(E))\otimes\operatorname{Sym}({\textstyle\bigwedge}^{2}(F))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}fg\scriptstyle{f\otimes g}Sym(𝐂r|2sV)\textstyle{\operatorname{Sym}(\mathbf{C}^{r|2s}\otimes V)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Sym(𝐂rE)Sym(𝐂2sF)\textstyle{\operatorname{Sym}(\mathbf{C}^{r}\otimes E)\otimes\operatorname{Sym}(\mathbf{C}^{2s}\otimes F)}

Here the horizontal maps are the surjection of the left ring onto the quotient by its nilradical. It follows that hh^{*} surjects onto Xr,s(V)X_{r,s}(V).

Finally, observe that the multiplicity space of 𝐒ρ(r+1,s+1)(V)\mathbf{S}_{\rho(r+1,s+1)}(V) in the algebra Sym(𝐂r|2sV)\operatorname{Sym}(\mathbf{C}^{r|2s}\otimes V) is 𝐒ρ(r+1,s+1)(𝐂r|2s)\mathbf{S}_{\rho(r+1,s+1)}(\mathbf{C}^{r|2s}) by the Cauchy decomposition, which vanishes by Proposition 2.1. Thus ker(h)\ker(h) contains Ir+1,s+1(V)I_{r+1,s+1}(V), and so 𝒱(Ir+1,s+1(V))\mathcal{V}(I_{r+1,s+1}(V)) contains the image of hh^{*}. This proves the lemma. ∎

5.4. The main theorem

We now come to our main result:

Theorem 5.8.

The 𝐆𝐋\mathbf{GL}-primes of AA are exactly the ideals Ir,sI_{r,s} with r,s1r,s\geq 1 and the zero ideal. The 𝐆𝐋\mathbf{GL}-radical ideals of AA are exactly the ideals IλI_{\lambda}, and the zero and unit ideal.

Corollary 5.9.

Let S=𝐍2{}S=\mathbf{N}^{2}\cup\{\infty\} equipped with the partial order described as follows: (r,s)<(r,s)<\infty for all (r,s)(r,s); and (r,s)(r,s)(r,s)\leq(r^{\prime},s^{\prime}) if rrr\leq r^{\prime} and sss\leq s^{\prime}. Endow SS with the unique sober topology for which \leq is the generalization order on points. Then Spec𝐆𝐋(A)\operatorname{Spec}_{\mathbf{GL}}(A) is homeomorphic to SS.

Before proving the theorem, we require a lemma.

Lemma 5.10.

Suppose dim(E)r\dim(E)\geq r and dim(F)2s\dim(F)\geq 2s. Let λ\lambda be a partition such that ρ(r,s)λ\rho(r,s)\not\subset\lambda. Then 𝒱(Ir,s(V))𝒱(Iλ(V))\mathcal{V}(I_{r,s}(V))\not\subset\mathcal{V}(I_{\lambda}(V)).

Proof.

By Proposition 5.2, we have 𝒱(Iλ(V))=(p,q)𝒞𝒱(Ip,q(V))\mathcal{V}(I_{\lambda}(V))=\bigcup_{(p,q)\in\mathcal{C}}\mathcal{V}(I_{p,q}(V)) where 𝒞\mathcal{C} is the set of corners of λ\lambda. For any (p,q)𝒞(p,q)\in\mathcal{C} we have ρ(r,s)ρ(p,q)\rho(r,s)\not\subset\rho(p,q), that is, r>pr>p or s>qs>q. It follows that V(Ip,q(V))V(I_{p,q}(V)) does not contain any pair (ω,η)(\omega,\eta) with rank(ω)=r1\operatorname{rank}(\omega)=r-1 and rank(η)=2s2\operatorname{rank}(\eta)=2s-2. However, 𝒱(Ir,s(V))\mathcal{V}(I_{r,s}(V)) does contain such pairs. ∎

Proof of Theorem 5.8.

Since 𝒱(Ir,s(V))=Xr1,s1(V)\mathcal{V}(I_{r,s}(V))=X_{r-1,s-1}(V) is irreducible for all super vector spaces VV, we see that rad𝐆𝐋(Ir,s)\operatorname{rad}_{\mathbf{GL}}(I_{r,s}) is 𝐆𝐋\mathbf{GL}-prime by Theorem B. If II is a 𝐆𝐋\mathbf{GL}-ideal of AA that properly contains Ir,sI_{r,s} then it contains some 𝐒2λ\mathbf{S}_{2\lambda} with ρ(r,s)λ\rho(r,s)\not\subset\lambda, and then 𝒱(Ir,s(V))𝒱(I(V))\mathcal{V}(I_{r,s}(V))\not\subset\mathcal{V}(I(V)) for large VV by the lemma. Since 𝒱(Ir,s(V))=𝒱(rad𝐆𝐋(Ir,s)(V))\mathcal{V}(I_{r,s}(V))=\mathcal{V}(\operatorname{rad}_{\mathbf{GL}}(I_{r,s})(V)) for all VV by Theorem A, it follows that Ir,s=rad𝐆𝐋(Ir,s)I_{r,s}=\operatorname{rad}_{\mathbf{GL}}(I_{r,s}). Thus Ir,sI_{r,s} is 𝐆𝐋\mathbf{GL}-prime.

We now show that the Ir,sI_{r,s} account for all the non-zero 𝐆𝐋\mathbf{GL}-primes of AA. Thus let II be some non-zero 𝐆𝐋\mathbf{GL}-prime ideal of AA. Then 𝒱(I(V))\mathcal{V}(I(V)) is irreducible for all VV by Theorem A. The rule V𝒱(I(V))V\mapsto\mathcal{V}(I(V)) defines an irreducible closed subvariety of XX, and therefore coincides with Xr,sX_{r,s} for some r,s𝐍{}r,s\in\mathbf{N}\cup\{\infty\} by Corollary 5.4. Since II contains Iρ(r,s)I_{\rho(r^{\prime},s^{\prime})} for some rr^{\prime} and ss^{\prime}, it follows that r,s<r,s<\infty. Thus rad(I(V))=rad(Ir+1,s+1(V))\operatorname{rad}(I(V))=\operatorname{rad}(I_{r+1,s+1}(V)) for all VV, and so I=Ir+1,s+1I=I_{r+1,s+1} by Theorem A, since both II and Ir+1,s+1I_{r+1,s+1} are 𝐆𝐋\mathbf{GL}-radical.

Since IλI_{\lambda} is an intersection of rectangular ideals (Proposition 5.2), it is therefore 𝐆𝐋\mathbf{GL}-radical. An argument similar to the one in the proof of Proposition 5.2 shows that any intersection of rectangular ideals is equal to IλI_{\lambda} for some λ\lambda, or the zero or unit ideal. Since any 𝐆𝐋\mathbf{GL}-radical ideal is an intersection of 𝐆𝐋\mathbf{GL}-primes, the result follows. ∎

Remark 5.11.

The idea that Theorem 5.8 should be true came out of joint work with Steven Sam. ∎

Remark 5.12.

There is an alternate method for proving that Ir,sI_{r,s} is 𝐆𝐋\mathbf{GL}-prime: explicitly compute the product ideal IλIμI_{\lambda}I_{\mu} in AA, for all λ\lambda and μ\mu, and verify the primality condition directly. As far as we know, the computation of IλIμI_{\lambda}I_{\mu} does not appear in the literature in this case. However, a closely related case (namely, that of Sym(VW)\operatorname{Sym}(V\otimes W) with 𝐆𝐋(V)×𝐆𝐋(W)\mathbf{GL}(V)\times\mathbf{GL}(W) acting) is treated in [W]. ∎

References

  • [Ab] Silvana Abeasis. The GL(V){\rm GL}(V)-invariant ideals in S(S2V)S(S^{2}V). Rend. Mat. (6) 13 (1980), no. 2, 235–262.
  • [AdF] S. Abeasis, A. Del Fra. Young diagrams and ideals of Pfaffians. Adv. in Math. 35 (1980), no. 2, 158–178.
  • [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
  • [CEP] C. de Concini, David Eisenbud, C. Procesi. Young diagrams and determinantal varieties. Invent. Math. 56 (1980), no. 2, 129–165.
  • [CW] Shun-Jen Cheng, Weiqiang Wang. Dualities and Representations of Lie Superalgebras, Graduate Studies in Mathematics 144, American Mathematical Society, Providence, RI, 2012.
  • [De] P. Deligne. Catégories tensorielles. Mosc. Math. J. 2 (2002), no. 2, pp. 227–248.
    https://www.math.ias.edu/files/deligne/Tensorielles.pdf
  • [Dr] Jan Draisma. Topological Noetherianity of polynomial functors. J. Amer. Math. Soc. 32(3) (2019), pp. 691–707. arXiv:1705.01419
  • [DLL] Jan Draisma, Michal Lason, Anton Leykin. Stillman’s conjecture via generic initial ideals. Comm. Alg. 47 (2019), no. 6, 2384–2395. arXiv:1802.10139
  • [EFW] David Eisenbud, Gunnar Fløystad, Jerzy Weyman. The existence of equivariant pure free resolutions. Ann. Inst. Fourier (Grenoble) 61 (2011), no. 3, 905–926. arXiv:0709.1529v5
  • [ESS] Daniel Erman, Steven V Sam, Andrew Snowden. Big polynomial rings and Stillman’s conjecture. Invert. Math., to appear. arXiv:1801.09852v4
  • [LR] Liping Li, Eric Ramos. Depth and the local cohomology of 𝐅𝐈G\mathbf{FI}_{G}-modules. Adv. Math. 329 (2018), 704–741. arXiv:1602.04405
  • [M] I. G. Macdonald. Symmetric Functions and Hall Polynomials, second edition, Oxford Mathematical Monographs, Oxford, 1995.
  • [NS1] Rohit Nagpal, Andrew Snowden. Symmetric subvarieties of infinite affine space. In preparation.
  • [NS2] Rohit Nagpal, Andrew Snowden. The structure of symmetric ideals. In preparation.
  • [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
  • [Sn] Andrew Snowden. Syzygies of Segre embeddings and Δ\Delta-modules. Duke Math. J. 162 (2013), no. 2, 225–277, arXiv:1006.5248v4.
  • [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. GL-equivariant modules over polynomial rings in infinitely many variables. II. Forum Math., Sigma 7 (2019), e5, 71 pp. arXiv:1703.04516v1
  • [SS5] Steven V Sam, Andrew Snowden. Sp-equivariant modules over polynomial rings in infinitely many variables. In preparation.
  • [W] Karen Louise Whitehead. Products of generalized determinantal ideals and decompositions under the action of general linear groups. Ph.D. dissertation, University of Minnesota, 1982.