The spectrum of a twisted commutative algebra
Abstract.
A twisted commutative algebra is (for us) a commutative -algebra equipped with an action of the infinite general linear group. In such algebras the “-prime” ideals assume the duties fulfilled by prime ideals in ordinary commutative algebra, and so it is crucial to understand them. Unfortunately, distinct -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 -primes. This yields an effective method for analyzing -primes.
1. Introduction
A twisted commutative algebra (tca) is a commutative -algebra equipped with an action of the infinite general linear group 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 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 -ideal of is an ideal that is -stable.
-
•
A -prime is a -ideal such that implies or , for subrepresentations . Here denotes the image of the map .
-
•
The -radical of a -ideal , denoted , is the sum of all subrepresentations of such that for some . Here denotes the image of the map . This is equal to the intersection of the -primes containing (Proposition 2.5).
-
•
The -spectrum of , denoted , is the set of -primes, endowed with the usual Zariski topology. The closed subsets of correspond bijectively to -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 -primes in twisted commutative algebra. For instance: the support of an equivariant module is most naturally a subset of the -spectrum; the -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 -ideals. Therefore, to understand tca’s it is of crucial importance to understand their -primes.
Every -stable prime ideal of a tca is a -prime. However, the converse is not true. The following is an instructive example:
Example 1.1.
Let . Then is a tca, and a nice one at that: it is finitely generated and noetherian, in the equivariant sense. Suppose that and are non-zero subrepresentations of . Since exterior powers are irreducible, it follows that contains and contains for some and . Thus contains . We have thus shown that if and are non-zero then so is . It follows that the zero ideal of is -prime; in other words, is -integral. ∎
This example is rather shocking when one first encounters it: every positive degree element of is nilpotent, and yet is a -domain! This example shows that (from our current perspective) tca’s do not have enough points to “see” their -primes: indeed, is a single point, and thus cannot distinguish the two -primes and of . Thus -domains appear to be divorced from geometry, which might diminish our hopes of understanding -primes; fortunately, however, this appearance is deceiving.
1.2. The key principle
The category of polynomial representations of is equivalent to the category of polynomial functors of rational vector spaces; the equivalence is obtained by evaluating a functor on . We can thus view a tca as an algebra object of . From this perspective, can be seen as the “incarnation” of in the category . However, polynomial functors can be evaluated on objects in any -linear tensor category. We can thus form the “super incarnation” of by evaluating on the super vector space . This is an algebra object of . 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 -primes.
This principle is borne out in the theorems stated below.
Example 1.2.
Let be the tca from Example 1.1, regarded in . Then is the second Veronese subring of , which is a domain of Krull dimension ; in particular, its spectrum has plenty of points. For , the algebra is a nilpotent extension of . ∎
1.3. Main results
We now state several precise theorems. In what follows, is a tca (considered as a polynomial functor) and and are -ideals of .
Theorem A.
We have if and only if .
In other words, the theorem says that if and only if ; the latter condition is equivalent to for all and 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 , 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 is -prime if and only if the ideal is prime.
Once again, we can verify the latter condition on finite dimensional spaces. The theorem therefore reduces the problem of showing that is -prime to showing that the algebraic varieties are irreducible.
In the two remaining theorems, we require a finiteness condition: we assume that is noetherian and that is finitely generated over .
Theorem C.
We have the following:
-
(a)
has finitely many minimal -primes, say ;
-
(b)
has finitely many minimal primes, say ;
-
(c)
, and after applying a permutation we have for all .
This theorem gives a useful way to find the minimal -primes, at least up to -radical, and thus the irreducible components of the -spectrum.
Theorem D.
The -spectrum of is a noetherian topological space.
This theorem is a strengthening of Draisma’s topological noetherianity theorem: indeed, Draisma’s theorem only encompasses the -stable prime ideals, while this theorem accommodates all -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 . Using our main theorems, we classify the -primes of (they are the rectangular ideals) and the -radical ideals of (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 through standard geometric means. It also provides a reconceptualization of sorts for the rectangular ideals: they constitute the equivariant spectrum of . These ideals have long been of interest, as they are symbolic powers of determinantal ideals.
1.5. Connection to other work
The notions of -prime and the -spectrum were discussed in [SS4, §3] (with slightly different terminology). However, that paper only works with so-called bounded tca’s, and for these -primes are the same as -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 -primes in the ring and manage to completely classify them. (Here denotes the infinite symmetric group.) However, there is no general theory of -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
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 as an algebraic group over . By an algebraic representation of , we mean a comodule over ; the dimension is not required to be finite. We regard as the inductive system of algebraic groups . An algebraic representation of is a vector space equipped with compatible algebraic representations of for all . The basic example of such a representation is the standard representation . A polynomial representation of 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 . It is a -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 , where denotes the Schur functor associated to the partition . All simple objects are absolutely simple. Every polynomial representation admits a canonical decomposition where the are multiplicity spaces. We endow with a grading by declaring the elements of to have degree .
By an algebraic representation of , we mean a comodule over the Hopf superalgebra . We then define polynomial representations of just as before, and denote the category by . It is again semi-simple abelian and closed under tensor product, and the simple objects have the form . Warning: if is a polynomial representation of then (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 of endofunctors of the category of -vector spaces. Let be the functor given by . We say that an object of is polynomial if it is a subquotient of a (possibly infinite) direct sum of ’s. We let be the category of polynomial functors. It is semi-simple abelian and closed under tensor products. The simple objects are the Schur functors . We have equivalences of categories
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 be a partition, and let . If then is an absolutely irreducible representation of ; otherwise it vanishes. Moreover, the irreducible representations of 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 , denote , to be . Define the width of a polynomial functor , denoted , to be the supremum of over those for which occurs in .
Proposition 2.2.
Let be a polynomial functor. The following are equivalent:
-
(a)
.
-
(b)
For every , every weight appearing in satisfies .
Proof.
It suffices to treat the case where is a simple object; we then have . For , the representation is irreducible with highest weight . Thus there is a weight in with (namely, ). Furthermore, since is a highest weight, we have for any other weight . 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 , , or . For the moment, we work in to be definite.
Let be a tca in . Every polynomial representation carries a natural grading, and so is canonically graded. This is compatible with the ring structure, i.e., is a graded ring. In particular, we can regard as an algebra over is degree 0 piece , which is an ordinary commutative ring.
By a “subrepresentation” of , we mean a -subspace that is stable by . In practice, will often be a -algebra, but we use -subrepresentations nonetheless. One should think of a finite length subrepresentations of as providing an equivariant substitute for the concept of element (or perhaps finite sets of elements).
Suppose is a homomorphism of tca’s. We say that is finitely -generated over if there is some finite length subrepresentation of such that the natural map is surjective. We typically apply this in the case .
2.3. Ideals in tca’s
Let be as above, i.e., a tca in . A -ideal of is a -stable ideal of . We say that a -ideal is finitely -generated if there is a finite length subrepresentation of that generates as an ideal. The sum and product of two finitely -generated ideals is again finitely -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 -ideal is -prime if implies or , for subrepresentations and of ; here denotes the image of the map . It is equivalent to ask the same condition with and finite length representations, or cyclic representations, or -idelas, or finitely generated -ideals. We define the -radical of a -ideal , denoted to be the sum of all subrepresentations or such that for some ; again, one can use ideals in place of subrepresentations. We say that is -radical if . We note that every -prime is -radical.
Remark 2.3.
A “prime -ideal” of is a -ideal of that is prime. This is potentially very different from a “-prime ideal” of . Similarly, “radical -ideal” and “-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 be a -ideal of and let be a finite length subrepresentation of . Then for some . Similarly, if is a finitely -generated ideal contained in then for some .
Proof.
By definition, we can write where is a subrepresentation of such that for some . Since is contained in and of finite length, there is some finite subset of such that . We thus have where . For the ideal case, simply pick a finite length subrepresentation that generates and appeal to the previous argument. ∎
Proposition 2.5.
Let be an ideal of . Then is the intersection of the -primes containing .
Proof.
Let be the set of -primes containing . Suppose . Let be a finitely -generated -ideal contained in . Then for some , and so since is -prime. Since this holds for all , we have . Since this holds for all , we have .
We now prove the reverse inclusion. Let be a finitely -generated -ideal of not contained in . Let be the set of -ideals of such that no power of is contained in . Suppose that is a chain in , and let be its union. Then belongs to too. Indeed, if belongs to then, because it is finitely generated, it belongs to some , a contradiction. Let be a maximal element of , which exists by Zorn’s lemma. We claim that is prime. Indeed, suppose , but . Then and strictly contain , and therefore do not belong to . Thus and for some and . Thus , a contradiction. It follows that , which completes the proof. Indeed, if were strictly larger than , then we could find a finitely -generated contained in but not contained , and the above argument would yield a contradiction. ∎
Proposition 2.6.
Every -prime of contains some minimal -prime of .
Proof.
An intersection of a descending chain of -primes is clearly -prime, so the claim follows from Zorn’s lemma. ∎
The above concepts (-prime, -radical, etc.) are defined using only the language of the tensor category . It follows that the same definitions can be made in and , 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 is a tca in and is a -ideal in it. We can then consider their incarnations in , and form . Similarly, we can consider their incarnations in and form . There is no reason to expect these two radicals to be comparable in any way (except in tautological ways, e.g., both contain ). 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 be a tca in and let be a minimal prime of . Then is -stable.
Proof.
Consider the maps
where is comultiplication, is the quotient map, and is the counit. The composition is equal to by the axioms for a comodule, and thus has kernel . We thus see that . However, is prime and is a localization of a polynomial algebra over , and so is a domain. Thus is prime. Since is minimal, we must have . Now, let be an element of , and write where and are -linearly independent elements. Then . Since the are linearly independent, it follows that for all , and so for all . Thus , and so is -stable. Since this holds for all , it follows that is -stable. ∎
Remark 2.8.
The analog of this statement for tca’s in does not hold: the above proof fails since is not a domain. ∎
2.5. Radicals of -primes
We require the following result on -primes. See [SS2, §8.6] for some similar results.
Proposition 2.9.
Let be a tca in and let be a -prime of . Then is prime.
We require some preliminary work before proving the proposition. Let be a homogeneous basis for . Given an element in a polynomial representation of and a subset of , we say that has support contained in if can be embedded into a direct sum of tensor powers of such that can be expressed using the basis elements in . One can define this more canonically by looking at the weight decomposition of . We define the support of an element of to be the set of indices such that has a non-zero entry in row or column . We say that an element of has support contained in if it can be expressed in terms of elements of having this property. We say that elements of representations or are disjoint if they have disjoint supports (or if they have supports contained in disjoint sets). For an element , we let be the -ideal it generates.
Lemma 2.10.
Suppose are disjoint elements such that . Then .
Proof.
Let be an element of that is disjoint from . Write where . Now, the support of may overlap with that of , that is, we may use auxiliary basis vectors in the process of building from . However, it does not matter which auxiliary basis vectors we use, so we can modify if necessary so that it is disjoint from . More rigorously, choose a permutation of that fixes the supports of and , and such that is disjoint from . Then and is disjoint from . Now, applying to the expression , and using the fact that commutes with since it is disjoint from , we find , that is, .
Now let be an arbitrary element of . We can then write where is a permutation of and is disjoint from . Let act by on the support of and 0 on the remaining basis vectors. Then and . Since is disjoint from , we have by the previous paragraph. Applying to this equation gives . This completes the proof. ∎
Lemma 2.11.
Let be super homogeneous elements satisfying . Then there exists such that .
Proof.
We claim that for any there exists such that . This is clear for . Suppose now it is true for , and let us prove it for , with . It suffices to treat the case where is super homogeneous. Let be such that . Applying , we find , where the sign depends on super degrees. Multiplying by and using the fact that , we find . The claim now follows.
Now let be disjoint from and generate ; for instance, one could take for an appropriate permutation of . Since for some , the previous paragraph gives for some . By Lemma 2.10, we find , and so the result follows. ∎
Proof of Proposition 2.9.
Passing to , we assume is -prime. We must show that is prime. Since all odd elements of are nilpotent, we have . It thus suffices to show that if with and even then or . Thus let even elements and be given such that is nilpotent, say . Since and are even, they commute, and so . By Lemma 2.11, there exists such that . Since is -prime, it follows that or . Thus either or is nilpotent, which completes the proof. ∎
2.6. Draisma’s theorem
Suppose that a group acts on a topological space . We say that is -noetherian if every descending chain of closed -stable subsets stabilizes. Draisma [Dr, Corollary 3] proved the following important theorem in this context:
Theorem 2.12.
Let be a tca in such that is noetherian and is finitely generated over . Then is -noetherian.
In fact, Draisma only states this theorem when 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 be as in Theorem 2.12. Then every ascending chain of radical -ideals in stabilizes.
Proof.
This follows since radical -ideals of correspond bijectively to -stable closed subsets of . ∎
Corollary 2.14.
Let be as in Theorem 2.12. Let and let be the subset consisting of -stable prime ideals, endowed with the subspace topology. Then is a noetherian spectral space.
Proof.
Suppose that is a closed subset of . Then has the form for some closed subset of . Since every point in is -invariant, it follows that for any . Thus where is a -stable closed subset of . It follows that , where is the closure of in , since .
Now suppose that is a descending chain of closed sets in . Then is a descending chain of -stable closed subsets of , and thus stabilizes by Theorem 2.12. Since , it follows that the original chain stabilizes too. Thus is noetherian.
Finally, let be an irreducible closed subset of . Then is a -stable irreducible closed subset of . Its generic point is thus -stable, and therefore belongs to . One easily sees that it is the unique generic point for . Thus is sober. Since it is also noetherian, it is spectral. ∎
Corollary 2.15.
Let be as in Theorem 2.12. Then has finitely many minimal primes.
Proof.
Let be as in Corollary 2.14. Since is noetherian, it has finitely many irreducible components. Since it is sober, these components correspond to the minimal -stable prime ideals of . There are thus finitely many of these. However, if is any minimal prime then it is -stable by Proposition 2.7, and thus obviously a minimal -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 be a tca in that is finitely generated over . The following are equivalent:
-
(a)
Every positive degree homogeneous element of is nilpotent, for all , .
-
(b)
The ideal is nilpotent.
Proof.
It is clear that (b) implies (a). We prove the converse. We proceed by induction on the degree of generation of . Thus suppose is generated in degrees and satisfies (a), and that the theorem is true for tca’s generated in degrees .
Before getting into the argument, we introduce a piece of notation. For a polynomial functor , we let be the polynomial functor defined by . We note that carries an action of , through its action on . If is a weight vector for of weight and simultaneously a weight vector for of weight then it is also a weight vector for of weight .
Let be a subrepresentation generating . Let be the weight space decomposition for with respect to the action; note that is a polynomial functor of degree . Let be the subalgebra of generated by , and let be the subalgebra generated by . For , every weight vector of is a weight vector of of non-zero weight (since the first component of the weight is ), and thus a positive degree homogenous element of , and thus nilpotent. We thus see that the generators of are nilpotent, and so satisfies (a). Since is finitely generated in degrees , we can apply the inductive hypothesis to conclude that for . Since the degree 0 generators of are also nilpotent, it follows that has finite length as a polynomial functor. In particular, only finitely many weights appear in ; say that the largest one is .
Let be the maximal torus of . Suppose that is a weight of that appears in ; we note that records the action of . Since and generate are are -stable, it follows that can be written in the form where is a weight of appearing in and is one in . We have by the definition of . Since is generated by , on which acts trivially, we see that . Thus . Since this holds for all weights in for any , it follows that has width by Proposition 2.2.
Decompose as where is a multiplicity space. We have just shown that is only non-zero when . Consider the superalgebra . This is finitely generated, and every positive degree element is nilpotent by assumption. Thus for sufficiently large, say . We thus have for , and so : indeed, we know this already if , and otherwise is non-zero by Proposition 2.1. It follows that for as well, and so satisfies (b). This proves the theorem. ∎
Example 3.2.
Suppose is generated over by . The space has for a basis elements with . The degree of under the action is simply the number of indices equal to 1. Thus is spanned by , while is spanned by the with , and is spanned by the with . Thus is generated by , which is invariant, and the with , which generate a copy of the standard representation of . We thus see that is generated by , which has degree as a polynomial functor. This shows that can have degree 0 generators even if does not. The algebra is generated by the with , and so acts trivially on it. ∎
Corollary 3.3.
Let be an arbitrary tca in and let be a finite length subrepresentation of . The following are equivalent:
-
(a)
Every element of is nilpotent, for all and .
-
(b)
The space is nilpotent, i.e., the map is zero for some .
Proof.
Obviously (b) implies (a); we prove the converse. First suppose that is generated in positive degrees. Let be the sub tca of generated over by . Then is generated as a subalgebra of by , and so every positive degree element is nilpotent. Thus, by Theorem 3.1, we see that is nilpotent. Since , we see that it too is nilpotent.
In general, write where is the degree 0 piece of and is the sum of the positive degree pieces of . Then is finite dimensional and every element is nilpotent, so is nilpotent, and is nilpotent by the previous paragraph. Thus is nilpotent. ∎
The proof of Theorem 3.1 is effective, in the following sense. For a finite length polynomial functor , let be the class of tca’s such that (i) contains a copy of that generates it over ; and (ii) each weight space of of non-zero weight admits a basis consisting of -nilpotent elements, for any and . Let be the supremum of over , where denotes the maximum non-zero degree. Theorem 3.1 simply states that is finite for . In fact, the proof yields a bound on . Let be the degree of , let be as in the proof of the theorem, and let be the polynomial defined by . Examining the proof, one finds
This allows one to inductively obtain a bound on since the argument to on the right has smaller degree than . Making the rough approximation , one finds
This upper bound is quite large; e.g., it is substantially larger than . We do not know how close it is to the true behavior of .
4. The main theorems
We fix a tca in for this section. Consider the following condition:
-
()
is noetherian and is finitely generated over .
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 and be -ideals of . Then if and only if .
Note that if and only if , and similarly for ordinary radicals, so this proposition is indeed equivalent to Theorem A.
Proof.
We may replace with without changing either condition. We may then check the conditions after passing to . Thus we may simply assume from the outset that .
If then , since we have a containment . Conversely, suppose that . Let be a finite length subrepresentation of . Then every element of is nilpotent, and so is nilpotent by Corollary 3.3. Thus . Since this holds for all , it follows that . ∎
We now introduce an auxiliary algebra that we will be helpful in what follows. Let . The ideal is typically not stable, but is clearly stable by . It is therefore also stable by the diagonal subgroup , and so this acts on . Choose a surjection , for some representation . The restriction of to the diagonal has the form , where and are polynomial representations of , and indicates the super grading. It follows that is a quotient of . Since has no odd part, we see that it is a quotient of . In other words, is a twisted commutative algebra in the category . If satisfies then we can take to be a finite length representation. It follows from basic properties of Schur functors that and are then of finite length as well, and so is finitely generated over .
Proposition 4.2 (Theorem D).
Suppose holds. Then is noetherian.
Proof.
It suffices to show that every ascending chain of -radical ideals in stabilizes. Thus let be such a chain. Then is an ascending chain of -stable radical ideals of , and thus corresponds to an ascending chain of -stable radical ideals of . It therefore stabilizes by Corollary 2.13. By Proposition 4.1, it follows that the original chain stabilizes. ∎
Proposition 4.3 (Theorem C).
Suppose holds.
-
(a)
has finitely many minimal -primes, say ;
-
(b)
has finitely many minimal primes, say ;
-
(c)
, and after applying a permutation we have for all .
Proof.
(a) This is an immediate consequence of Proposition 4.2.
(b) The minimal primes of correspond bijectively to those of , and there are finitely many of these by Corollary 2.15.
(c) Consider a minimal prime . We have
Since is prime, it follows that for some . Thus . Since is prime (Proposition 2.9) and is a minimal prime, we have . By Proposition 4.1, it follows that is unique: indeed, if then and so .
To complete the proof, it suffices to show that is a minimal prime for all . We know that is prime. It therefore contains some minimal prime . We have shown that for some . Thus , and so by Proposition 4.1. Since is a minimal -prime, it follows that . Hence is a minimal prime. ∎
Proposition 4.4 (Theorem B).
Let be a -ideal of . Then is -prime if and only if is prime.
Proof.
If is -prime then is prime by Proposition 2.9. Now suppose that is prime and satisfies . Then has a unique minimal prime, and therefore a unique minimal -prime by Proposition 4.3. Thus there is a unique minimal -prime over , and so is -prime. Finally, suppose that is prime and is arbitrary. Write where is a directed family of sub tca’s satisfying . Then is prime, and so is -prime by the previous case. Since is prime for all , it follows that is prime by Lemma 4.5 below. ∎
Lemma 4.5.
Let be a tca, and suppose that for some directed family of sub tca’s. Let be a -ideal of . If is -prime for all then is -prime.
Proof.
Let and be finite length subrepresentations of such that . Since adn are finite length, there is some such that and are contained in . Thus . Since is -prime, it follows that or . Thus or , and so is -prime. ∎
5. An example
Let be the tca in given by . Our goal is to classify the -prime and -radical ideals of .
5.1. The ideal lattice of
The decomposition of into irreducibles is well-known:
where the sum is over all partitions , and . See, for example, [M, §I.5, Example 5]. For a partition , let be the ideal of generated by . The following result determines the ideal structure of :
Proposition 5.1.
The ideal is the sum of those for which . In particular, if and only if .
Proof.
This was originally proved in [Ab], but that is a difficult reference to obtain. The analogous result for is proved in [AdF, Theorem 3.1]. That case actually implies this one, since . A complete proof in the case where 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 be the partition with rows each of length ; thus the Young diagram for is an rectangle. We let . If or then is an empty partition and is the unit ideal.
Let be a partition. By a corner of we mean a pair such that has a box in the th row and th column, but no box below or to the right of this one. For example, in the following Young diagram the corners have been shaded:
The following observation illustrates the importance of the rectangular ideals.
Proposition 5.2.
Let be the set of corners of . Then .
Proof.
Let be a partition. We have
The first and last step follow from Proposition 5.1, the second is trivial, and the third is simply the observation that . The result thus follows. ∎
5.2. The variety
For , let . Note that . We regard as a 2-variable tca (in the variables and ). Let , which we identify with . By a “closed subvariety” of , we mean a subfunctor of such that is a closed subvariety of for all finite dimensional . Closed subvarieties of correspond bijectively to stable radical ideals of . For , let be the locus of pairs such that and . Then is a closed subvariety of in the above sense. We now show that these account for essentially all examples:
Proposition 5.3.
Let be a closed subvariety of . Then there is a finite subset of such that .
Proof.
By the rank of a point , we mean the pair . Let be the set of pairs such that has a point of rank for some and . We claim that if and only if . It is clear that implies . Conversely, suppose that . Then there exists some of rank for some and . Let be given. By basic linear algebra, there are linear maps and such that and . Thus the map defined by carries to . Since and is a subfunctor of , it follows that . This proves the claim.
It now follows that if and then , where here means and . A simple combinatorial argument now shows that there is a finite subset of such that if and only if for some . It follows that is the union of the with , which proves the proposition. ∎
Corollary 5.4.
Any irreducible closed subvariety of is one of the .
5.3. The vanishing locus of
The goal of this section is to prove the following:
Proposition 5.5.
We have for .
We break the proof into two lemmas.
Lemma 5.6.
We have the following:
-
(a)
We have in .
-
(b)
We have in .
-
(c)
We have .
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 . We thus have a surjection . By Proposition 5.1, we have
We thus see that is the sum of those ’s with and . However, if , and so there are only finitely many relevant such . Thus 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 contains both and , we see that its radical contains both and by (i) and (ii), and thus the extensions of these ideals to . It follows that is contained in the intersection of and . Now, is the classical determinantal ideal: its vanishing locus in consists of those forms rank . The vanishing locus of is thus . Similarly, the vanishing locus of is . We thus find that is contained in . ∎
Lemma 5.7.
We have .
Proof.
Let be a non-degenerate symmetric bilinear form on . We obtain a natural map
where the first map comes from the inclusion of the invariant provided by , and the second map comes from the Cauchy decomposition. The above map induces an algebra homomorphism . The induced map on spectra takes a linear map to the form on , and thus surjects onto the locus of forms of rank .
Now let be a non-degenerate symplectic form on . A similar construction yields a homomorphism such that surjects onto the locus of forms in of rank .
Finally, let be the non-degenerate orthosymplectic form on that restricts to and on the even and odd pieces. Once again, we get a natural algebra homomorphism . One easily verifies that the following square commutes:
Here the horizontal maps are the surjection of the left ring onto the quotient by its nilradical. It follows that surjects onto .
Finally, observe that the multiplicity space of in the algebra is by the Cauchy decomposition, which vanishes by Proposition 2.1. Thus contains , and so contains the image of . This proves the lemma. ∎
5.4. The main theorem
We now come to our main result:
Theorem 5.8.
The -primes of are exactly the ideals with and the zero ideal. The -radical ideals of are exactly the ideals , and the zero and unit ideal.
Corollary 5.9.
Let equipped with the partial order described as follows: for all ; and if and . Endow with the unique sober topology for which is the generalization order on points. Then is homeomorphic to .
Before proving the theorem, we require a lemma.
Lemma 5.10.
Suppose and . Let be a partition such that . Then .
Proof.
By Proposition 5.2, we have where is the set of corners of . For any we have , that is, or . It follows that does not contain any pair with and . However, does contain such pairs. ∎
Proof of Theorem 5.8.
Since is irreducible for all super vector spaces , we see that is -prime by Theorem B. If is a -ideal of that properly contains then it contains some with , and then for large by the lemma. Since for all by Theorem A, it follows that . Thus is -prime.
We now show that the account for all the non-zero -primes of . Thus let be some non-zero -prime ideal of . Then is irreducible for all by Theorem A. The rule defines an irreducible closed subvariety of , and therefore coincides with for some by Corollary 5.4. Since contains for some and , it follows that . Thus for all , and so by Theorem A, since both and are -radical.
Since is an intersection of rectangular ideals (Proposition 5.2), it is therefore -radical. An argument similar to the one in the proof of Proposition 5.2 shows that any intersection of rectangular ideals is equal to for some , or the zero or unit ideal. Since any -radical ideal is an intersection of -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 is -prime: explicitly compute the product ideal in , for all and , and verify the primality condition directly. As far as we know, the computation of does not appear in the literature in this case. However, a closely related case (namely, that of with acting) is treated in [W]. ∎
References
- [Ab] Silvana Abeasis. The -invariant ideals in . 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 -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 -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.