Topological Noetherianity of polynomial functors

By Jan Draisma

Abstract

We prove that any finite-degree polynomial functor over an infinite field is topologically Noetherian. This theorem is motivated by the recent resolution, by Ananyan-Hochster, of Stillman’s conjecture; and a recent Noetherianity proof by Derksen-Eggermont-Snowden for the space of cubics. Via work by Erman-Sam-Snowden, our theorem implies Stillman’s conjecture and indeed boundedness of a wider class of invariants of ideals in polynomial rings with a fixed number of generators of prescribed degrees.

1. Introduction

This paper is motivated by two recent developments in “asymptotic commutative algebra”. First, in Reference 4, Ananyan-Hochster prove a conjecture due to Stillman Reference 23, Problem 3.14, to the effect that the projective dimension of an ideal in a polynomial ring generated by a fixed number of homogeneous polynomials of prescribed degrees can be bounded independently of the number of variables. Second, in Reference 9, Derksen-Eggermont-Snowden prove that the inverse limit over of the space of cubic polynomials in variables is topologically Noetherian up to linear coordinate transformations. These two theorems show striking similarities in content, and in Reference 16, Erman-Sam-Snowden show that topological Noetherianity of a suitable space of tuples of homogeneous polynomials, together with Stillman’s conjecture, implies a generalisation of Stillman’s conjecture to other ideal invariants. In addition to being similar in content, the two questions have similar histories—e.g., both were first established for tuples of quadrics Reference 3Reference 15—but since Reference 4 the Noetherianity problem has been lagging behind. The goal of this paper is to make it catch up.

1.1. Polynomial functors

Let be an infinite field and let be the category of finite-dimensional vector spaces over . We consider a covariant polynomial functor of finite degree . This means that for all the map is polynomial of degree at most , with equality for at least some choice of and . The uniform upper bound rules out examples like .

Then splits as a direct sum where

see, e.g., Reference 18. For each the map is homogeneous of degree , and we have for all of sufficiently large dimension.

1.2. Topological spaces over a category

The polynomial functor will also be interpreted as a functor to the category of topological spaces. Here we equip the finite-dimensional vector space with the Zariski-topology.

A general functor from a category to is called a -topological space, or -space for short. A second -space is called a -subspace of if for each object of the space is a subset of with the induced topology, and if, moreover, for a morphism the continuous map is the restriction of to . We then write . The -subspace is called closed if is closed in for each ; we then also call a closed -subset of . The -space is called Noetherian if every descending chain of closed -subspaces stabilises.

Furthermore, a -continuous map from a -space to a -space consists of a continuous map for each object , in such a way that for any morphism we have . A -homeomorphism has the natural meaning.

1.3. The main theorem

We will establish the following fundamental theorem.

Theorem 1.

Let be the category of finite-dimensional vector spaces over an infinite field , and let be a finite-degree polynomial functor. Then is Noetherian as a -topological space.

Remark 2.

The restriction to infinite is crucial for the set-up and the proofs below—e.g., it is used in the decomposition of a polynomial functor into homogeneous parts and, implicitly, to argue that if an algebraic group preserves an ideal, then so does its Lie algebra. In future work, we will pursue versions of Theorem 1 over and possibly over finite fields.

Theorem 1 will be useful in different contexts where finiteness results are sought. In the remainder of this section we discuss several such consequences; since an earlier version of this paper appeared on the arXiv, several other ramifications have appeared, e.g., in Reference 16.

1.4. Equivariant Noetherinity of limits

The polynomial functor gives rise to a topological space , the projective limit along the linear maps where is the projection forgetting the last coordinate. By functoriality, each is acted upon by the general linear group , the map is -equivariant if we embed into via , and hence is acted upon by the direct limit , the group of all invertible -matrices which in all but finitely many entries equal the infinite identity matrix.

Given a closed -subset of , the inverse limit is a closed, -stable subset of , and using embeddings appending a zero coordinate, one finds that surjects onto each . Conversely, given a closed, -stable subset of , then for any finite-dimensional vector space and any linear isomorphism , we set , where is the image of in . One can check that is a closed -subset of and that this construction is inverse to the construction above. Thus the theorem is equivalent to the following corollary.

Corollary 3.

Let be the category of finite-dimensional vector spaces over an infinite field , let be a finite-degree polynomial functor, and equip with the inverse-limit topology of the Zariski topologies on the . Then is -Noetherian, i.e., every chain of -stable closed subsets of is eventually constant. Equivalently, every -stable closed subset of is the set of common zeros of finitely many -orbits of polynomial equations.

Example 4.

The paper Reference 9 concerns the case where , the third symmetric power. In this case, is the space of infinite cubics , and acts by linear transformations that affect only finitely many of the variables .

Remark 5.

The proofs below could have been formulated directly in this infinite-dimensional setting, rather than the finite-dimensional, functorial setting. However, one of the key techniques, namely, shifting a polynomial functor by a constant vector space, is best expressed in the functorial language. Moreover, the functorial language allows us to stay in the more familiar realm of finite-dimensional algebraic geometry.

1.5. Generalisations of Stillman’s conjecture

In Reference 16, Erman, Sam and Snowden use the following special case of Theorem 1.

Corollary 6.

Let be an infinite field, fix natural numbers , and consider the polynomial functor . Then is a Noetherian -topological space, and hence its limit is -Noetherian.

Let be a function that associates a number to any homogeneous ideal in a symmetric algebra on , in such a way that for any injective linear map with induced homomorphism , and such that is upper semicontinuous in flat families. In Reference 16 the following is proved.

Theorem 7 (Reference 16).

Corollary 6 implies that for any ideal invariant with the properties above there exists a number such that for all , any ideal generated by homogeneous polynomials of degrees either has or .

The crucial point here is that does not depend on . Stillman’s conjecture Reference 23, Problem 3.14 is this statement for equal to the projective dimension, and it is used in the proof of the generalisation just stated. However, in a follow-up paper Reference 17, the same authors give two new proofs of Stillman’s conjecture, one of which uses Corollary 6. An algorithmic variant of this latter proof is presented in Reference 13.

1.6. Twisted commutative algebras

For , the algebra of polynomial functions on , i.e., the direct limit of the symmetric algebras on , is a twisted commutative algebra in one of its incarnations Reference 31Reference 32. In this context, Theorem 1 says the following.

Corollary 8.

Any finitely generated twisted commutative algebra over is topologically Noetherian.

1.7. Functors from

Theorem 1 has the following generalisation to functors that take several distinct vector spaces as input.

Corollary 9.

Let be an infinite field, let be the category of finite-dimensional vector spaces over , let be a positive integer, and let be a functor from to such that for any the map is polynomial of uniformly bounded degree. Then is a Noetherian -topological space.

Note that the group of automorphisms of is , which when the are all equal contains a diagonal copy of . This suggests that Theorem 1 for is in fact stronger than this corollary, as we prove now.

Proof of Corollary 9 from Theorem 1.

Let be a chain of closed -subsets of . Let be the functor that sends to , and set , a closed -subset of the polynomial functor . By Theorem 1, the sequence is constant for at least some . We claim that so is . Indeed, let . Choose a with for each , and choose surjections and injections such that . Then for we have

as desired.

1.8. Slice rank

Taking , a tensor in is said to have slice rank if it is nonzero and of the form for a in one of the and an . A tensor has slice rank at most if it is a sum of at most tensors of slice rank ; in this sum the slice index may vary through . Being of slice rank at most a fixed number is a Zariski-closed condition (see Tao and Sawin’s blog post Reference 33) and preserved under tensor products of linear maps. Corollary 9 implies the following.

Corollary 10.

Let be the category of finite-dimensional vector spaces over an infinite field . For fixed and , there exists a tuple such that for all a tensor has slice rank at most if and only if for all -tuples of linear maps the tensor has slice rank at most .

Equivalently, in the space of -way -tensors, the variety of slice rank at most tensors is defined by finitely many -orbits (and even finitely many -orbits, acting diagonally) of polynomial equations. A more in-depth study of the algebraic geometry of slice rank is a forthcoming work with Oosterhof.

1.9. Related work

Theorem 1 fits in a trend at the interface between representation theory, algebraic geometry, commutative algebra, and applications, which studies algebraic structures over some base category and aims to establish stabilisation results. Recent examples, in addition to those referenced above, include the theory of modules over the category of finite sets with injective maps Reference 7; Gröbner techniques Reference 29 for modules over more general combinatorial categories that, among other things, led to a resolution of the Artinian conjecture Reference 24 and to a resolution of a conjecture by Rauh-Sullivant Reference 26 on iterated toric fibre products Reference 14; and finiteness results for secant varieties of Segre and Segre-Veronese embeddings; see Reference 21Reference 25Reference 27Reference 28 and the notion of inheritance in Reference 20. The current paper, while logically independent of these results, was very much influenced by the categorical viewpoint developed in these papers.

2. Proof of the main theorem

2.1. Overview of the proof

The proof of Theorem 1 is a double induction. The outer induction is on the polynomial functor via a (“lexicographic”) partial order on the class of polynomial functors introduced in §2.2. Using classical work by Friedlander-Suslin, we prove that this is a well-founded order. Any degree-zero polynomial functor, i.e., a constant vector space independent of the input , is smaller than all polynomial functors of positive degree, and Hilbert’s basis theorem yields the base case of the induction.

So when we want to prove the theorem for , we may assume that it holds for all polynomial functors smaller than . We then show that every closed -subset of is Noetherian, by the inner induction on the smallest degree of a nonzero equation vanishing on for some . Roughly, this works as follows (see the next paragraph for subtleties): fix an irreducible component of the highest-degree part of , and find a direction such that the directional derivative is not identically zero. Let be the largest -closed subset of on which vanishes identically. Since has a lower degree than , is Noetherian by the inner induction hypothesis. On the other hand, set and , so that ; this is discussed in §2.4. In §2.9 we show that the complement of has a closed embedding into a basic open subset of , so Noetherian by the outer induction hypothesis. Hence , the union of two Noetherian spaces, is Noetherian.

There are four subtleties: First, may not depend on the coordinates on , so that for all . We therefore need to look for in the ideal of that is nonzero even after modding out the ideal of the projection of in ; see §2.5. Second, in positive characteristic, directional derivatives (linearisations) do not necessarily behave well; we replace these by additive polynomials in §2.6. Third, the closed embedding is in fact a Zariski homeomorphism to a closed subset; see §2.7 for the relevant lemma. Fourth, it is not quite that embeds into a basic open subset of is not functorial in —but rather the locus in where (which involves only coordinates on the constant vector space ) is nonzero. The closed embedding is then just the restriction of the projection along . Smearing around by , we obtain , and in §2.9 we show that this is good enough.

Example 11.

As a running example to illustrate the proof, we assume that , where is the space of symmetric two-tensors (matrices) and is the space of skew-symmetric tensors. Let be the standard coordinates on , let be the standard coordinates on extended to by declaring them zero on , and let be the standard coordinates on , similarly extended to .

We then have

Take , the variety of rank tensors. Then is the entire space , but , so we may take and for the determinant

We take and find

In this case, for , is the subvariety of on which the -orbit of vanishes identically, i.e., is the set of rank tensors in . This is a coincidence; in the general setting of the proof, does not embed into , but it always has a lower-degree polynomial vanishing on it. We discuss the complement in Example 26.

2.2. A well-founded order on polynomial functors

We will prove Theorem 1 by induction on the polynomial functor, along a partial order that we introduce now. Define a relation on polynomial functors of finite degree by if and moreover if is the highest degree with , then is a homomorphic image of ; this is a partial order on (isomorphism classes of) polynomial functors.

Lemma 12.

The relation is a well-founded order on polynomial functors of finite degree.

Proof.

It suffices to prove that this order is well-founded when restricted to polynomial functors of degree at most a fixed number . By Reference 18, Lemma 3.4, if is any vector space of dimension at least , then the map is an equivalence of abelian categories from polynomial functors of degree at most and finite-dimensional polynomial -representations of degree at most . Hence implies that the sequence is strictly smaller than the sequence in the lexicographic order where position is more significant than position . Since this lexicographic order is a well-order, is well-founded.

2.3. -varieties and their ideals

Write for the contravariant functor from to -algebras that assigns to the coordinate ring . A closed -subset of will be called a -variety in , and denoted . Its ideal is a contravariant functor that sends to the ideal of inside .

Using scalar multiples of the identity and the fact that is infinite, one finds that the ideal of is homogeneous with respect to the -grading that assigns to the coordinates on the degree . The degree function on and its quotients by homogeneous ideals is defined relative to this grading.

Example 13.

In our running Example 11, have degrees , respectively.

2.4. The shift functor

Fixing a , we let be the shift functor that sends and the homomorphism to the homomorphism .

Lemma 14.

For any polynomial functor of degree , is a polynomial functor of degree whose degree- homogeneous part is canonically isomorphic to that of .

Proof.

Set . For the map

given by is the composition of the affine-linear map and the polynomial map of degree at most , so is a polynomial functor of degree at most .

For let be the embedding and let be the projection . These give rise to morphisms of polynomial functors and given by and . Straightforward computations show that and map each homogeneous part into and vice versa, and that is the identity. Conversely, for in the highest-degree part we have for all . The coefficient of in the left-hand side equals , so we have and therefore

which proves that is indeed a linear isomorphism.

Example 15.

If , then , so equals plus a polynomial functor of degree .

Example 16.

In our running Example 11, , and

note that the degree-two part of is , so that .

2.5. Splitting off a term of highest degree

Assume that is a polynomial functor of degree . Let be any irreducible sub-polynomial functor of the highest-degree part of , define , and let be the natural projection. Then embeds into via the pull-back of . If is a -variety in , then let be the -variety in defined by setting equal to the Zariski-closure in of .

We will think of as a -variety over . Accordingly, we write for the contravariant functor that assigns to the ideal of in , the quotient of by the ideal in generated by the ideal of in .

In particular, we have if and only if for all we have . We write for the minimal degree of a nonzero homogeneous polynomial as runs over . Note that any polynomial in of degree zero is contained in ; here we use that . So if a degree-zero polynomial vanishes on , then it is an equation for and has already been modded out in the definition of . This explains why . Furthermore, note that if and only if .

Example 17.

In Examples 11,13, .

2.6. Additive polynomials as directional derivatives

For a finite-dimensional vector space over the infinite field , we write for the subset of consisting of polynomials such that for all . Then is a -subspace of , and equal to when . In general, if we let be the characteristic exponent of —so if and otherwise—and if we choose a basis of , then has as a basis the polynomials where runs through and through if and through if . The span of these for fixed is denoted .

Lemma 18.

Let be finite-dimensional vector spaces over the infinite field , let be a polynomial on , and consider the expression , a polynomial function of the triple . Then one of the following hold:

(1)

is independent of ; this happens if and only if factors through the projection ; or

(2)

the nonzero part of of lowest degree in is of the form for a unique (taken if ). Then for each fixed the map is in .

Proof.

For , let denote the th Hasse directional derivative in the direction . This linear map is defined in characteristic by and in arbitrary characteristic by realising that the latter expression actually has integer coefficients relative to any monomial basis of . Explicitly: let be a basis of and let with . Then

in particular, if , then this is nonzero as a polynomial over , even when is zero. A straightforward check shows that this is independent of the choice of basis, and that .

Taylor’s formula in arbitrary characteristic reads

If for all , then we see from the above that does not involve the variable , i.e., factors through . Similarly, if for all and all , then factors through .

Suppose that there exist and such that ; take such a pair with minimal. Then, in the coordinates above, contains a monomial for which is nonzero in , but in for all with . By Lucas’s theorem on binomial coefficients, is a power of , is divisible by , and in . Since in fact holds for all and all monomials in , is a multiple of , and hence for a unique polynomial . Moreover, .

More generally, let be a basis of and extend to a basis of .

We then find, by minimality of , that for a unique polynomial , and for all we have

Since for some , the right-hand side is additive in , which concludes the proof of the lemma.

For and and as in the second case of the lemma and for we write for the polynomial , and call this the directional derivative of in the direction . This polynomial has degree less than , and agrees with the usual directional derivative for . Note that depends on the choice of inside , not just on : if depends on but vanishes to a higher degree at for a specific than it does for general , then we have .

Also in the first case of the lemma we write for all . We extend the notation to rational functions with nonzero denominator by . The following lemma is immediate from Lemma 18.

Lemma 19.

For , , as in Lemma 18, , we have and for and .

Example 20.

Assume that . Let with standard coordinates , let be the span of the first two standard basis vectors, and let . Then

where the remaining terms are divisible by . Hence .

2.7. A closed embedding

Retaining the notation in the previous section, let be a basic open subset in defined by the nonvanishing of some polynomial (we allow , in which case ). Let be a Zariski-closed subset of and let be the ideal of inside . Fix a number , equal to if , and let be the set of elements such that for all (so is affine-additive in with additive part of degree ). Note that via the pull-back , is a -submodule of .

Lemma 21.

Assume that is algebraically closed and suppose that for each the map is surjective. Then restricts to a Zariski-homeomorphism from to a closed subset of .

Proof.

Fix any tuple whose restrictions to form a basis of . Then the natural map is an isomorphism by which we identify the two algebras. Under this identification, each element of can be written uniquely as for suitable elements . Let be -module generators of , and let be the coefficient of in . Let be the matrix whose entry equals . The condition in the lemma says that has rank for all .

Since is algebraically closed, the Nullstellensatz implies that lies in the ideal generated by the nonzero -subdeterminants of : for suitable .

Fix a . For each we can write , where is the th standard basis vector in and is a vector in (supported only on the positions corresponding to ) satisfying ; this is just Cramer’s rule. Now

and we conclude that contains an element of the form with .

Consider the morphism dual to the homomorphism that restricts to the identity on and sends each to . Since is algebraically closed, is a homeomorphism in the Zariski-topology, and since commutes with the projection , it suffices to show that this projection restricts to a closed embedding from into . Let be the ideal of . By the previous paragraph, contains an element of the form with for each . Hence the map is surjective, so is a closed embedding, as desired.

Remark 22.

In characteristic , the Zariski-homeomorphism from the lemma is in fact a closed embedding. In positive characteristic, it need not be.

2.8. Extending the field

Let be a finite-degree polynomial functor over the infinite field , let be an extension field of , and denote by the category of finite-dimensional vector spaces over . We construct a polynomial functor as follows. For every we fix a and an isomorphism of -vector spaces.

At the level of objects, is defined by . To define on morphisms we proceed as follows. For the polynomial map extends uniquely to a polynomial map

of the same degree; here we use that is infinite. The domain and codomain of are canonically and , respectively. Hence for we may set

A simple calculation shows that is indeed a polynomial functor .

Furthermore, if is a -closed subset of , then we obtain a -closed subset of by letting be the Zariski closure of in . The following lemma is straightforward.

Lemma 23.

The map from -closed subsets of to -closed subsets of is inclusion-preserving and injective. Consequently, Noetherianity of implies that of .

2.9. Proof of Theorem 1

If , then is a finite-dimensional space independent of , and for every linear map the map is the identity, so the theorem is the topological corollary to Hilbert’s basis theorem. We therefore may and will assume that . Furthermore, by Lemma 23 we may assume that is algebraically closed, so that we can use Lemma 21.

We proceed by induction, assuming that the theorem holds for all polynomial functors in the well-founded order from §2.2; this is our outer induction hypothesis. From §2.5 we recall the definition of . We now order -varieties inside the fixed by if either or else and . Since takes values in a well-ordered set, for any strictly decreasing sequence , must become strictly smaller infinitely often; but this is impossible since is Noetherian by the outer induction hypothesis. Hence is a well-founded order on -subvarieties of .

We set out to prove, by induction along this well-founded order, that each -variety is Noetherian as a -topological space. Our inner induction hypothesis states that this holds for each -variety inside .

First assume that , which means that is the pre-image of its projection , and let be any proper closed -subset. Then either or else and . Hence , so that is Noetherian by the inner induction hypothesis. Since any inclusion-wise strictly decreasing chain of closed -subsets of must contain such a as its first or second element, is Noetherian, as well.

So we may assume that . Take of minimal dimension for which contains a nonzero homogeneous element of degree , and let be a homogeneous polynomial representing this element. Regarding as a polynomial with coefficients from in coordinates that restrict to a basis of , we may remove from all terms with coefficients that vanish identically on , and then at least one nonconstant term survives.

By Lemma 18 applied to with and , this implies that there exists an such that the directional derivative in the sense of §2.6 also has at least some coefficient in that does not vanish on . Let be the exponent in the Lemma 18; so the map lies in for each . Since coordinate functions on were assigned degree 2.3), we have and in particular . By minimality of the degree of , the polynomial does not vanish identically on .

Let be the largest closed -subset of such that does vanish identically on , i.e., . Then either or else and . Hence , so is Noetherian by the inner induction hypothesis.

Define by . This is typically not a -subset of , since for the map might map points of into , i.e., outside . Indeed, since we chose of minimal dimension, when we pull back to a for , we obtain a polynomial that is zero modulo the ideal of . This implies that the pull-back of is identically zero on , so that . To remedy this, we now construct a -variety closely related to , by shifting our polynomial functor over as in §2.4.

Set and and consider the open -subset of defined by . Here we regard as a polynomial on via the map coming from the projection along . As the maps with are Zariski-dense in , and are related by

Recall that is an irreducible subfunctor of . Write where is homogeneous of degree and by Lemma 14. Define and note that since has degree at most and the degree- part of is equal to by Lemma 14. In particular, is a Noetherian -topological space by the outer induction hypothesis.

Remark 24.

Note that for , typically will have dimension larger than ! Also, is not equal to : there is a surjection with kernel .

The map has in its kernel, so we can regard as a polynomial on , as well. Consider the basic open -subset of defined by As is Noetherian, so is with its induced topology.

Lemma 25.

For every -vector space the projection is a Zariski-homeomorphism with a closed subset of .

Before proving this lemma, we use it to complete the proof of Theorem 1. First, since is Noetherian, Lemma 25 implies that so is .

Then suppose that is a sequence of closed -subsets of . By Noetherianity of there exists an such that for all the sequence is constant for . By Noetherianity of there exists an such that for all the sequence is constant for . Using Equation * we find

where in the last step we used that is -stable. We find that, for each of dimension at least , the sequence is constant for . Since for of dimension less than , we find that is constant for . This proves Noetherianity of and concludes the proof of the inner induction step.

Example 26.

We pause and see what Lemma 25 says in the running Examples 11,13,16. Here and we take . Then is the variety of -matrices

of rank such that the upper-left -submatrix has a nonzero coordinate (recall that ).

The lemma says that forgetting the skew-symmetric part of is a closed embedding of into the open subset of where is nonzero. We prove this by showing that, on , each entry of the skew-symmetric part of can be expressed as a rational function in the entries of and the entries of the symmetric part of , with a denominator equal to .

We assume and consider a surjective linear map . Thinking of as a -matrix, we have

Since is a -closed subset, for all , the -determinant vanishes on the latter -matrix for all choices of and of . Hence expanding as a polynomial in , the coefficient of also vanishes for all . Take in and specialise to the linear map sending the th standard basis vector to , the th standard basis vector to , and all other standard basis vectors to zero. Then the matrix above reads

The coefficient of in the determinant of this matrix is

where the remaining terms do not involve . Using, as in Example 11, the variables for the symmetric and skew-symmetric parts of , and using the variables for the symmetric and skew-symmetric parts of , this reads

where the dots in the last two expressions contain only variables that we do not discard in the projection . This shows that, on , the coordinate can be expressed in the entries of and the coordinates on the symmetric part of , as desired.

Proof of Lemma 25.

First assume that . We want to apply Lemma 21 with equal to , equal to , equal to , and equal to . Let be the ideal of in the coordinate ring of the pre-image of inside , and let be as in the text preceding Lemma 21.

Fix any surjective linear map . For and consider the linear map . This map depends on as a polynomial of degree at most , hence decomposes as for unique linear maps . Note that and . On the other hand, decompose where is a homogeneous polynomial functor of degree . Then is zero on except when . To see this, take a , compute

and observe that the right-hand side is homogeneous of degree in .

The together form the map . Since is a -subset of and vanishes on , for all and . This implies that the coefficient of in also vanishes identically on .

To determine how depends on , consider and with , and for variables compute

where in the second equality we have used that and in the last step we have used Lemma 18. We see that is the lowest power of to whose coefficient contributes, and that this contribution is additive in . More specifically, for each we have , and equals the value of the directional derivative at the point . In particular, .

From now on, assume that the image of in lies in . Since is surjective, so is . In particular, there exists an such that . For such an we have , so is not zero.

Keeping fixed but replacing by for , transforms into the additive function . Hence by varying we find that the image of in under the map from Lemma 21 contains a nonzero -submodule of . Since is irreducible and , is an irreducible -module Reference 18, Lemma 3.4, and this implies the irreducibility of and of —indeed, raising to the power gives a bijection from -submodules to -submodules of the latter.

We conclude that , and since was arbitrary in the pre-image of , the conditions of Lemma 21 are fulfilled. Hence the projection is a Zariski-homeomorphism with a closed subset of , as desired.

Finally, if , then take any embedding where does have sufficiently high dimension. Then we have a commuting diagram

where all arrows except, a priori, the left-most one are homeomorphisms with closed subsets of the target space. But then so is the left-most one.

2.10. Comments on the proof

The idea to do induction on is not completely new: it is also used, in special cases, in Reference 4Reference 9Reference 11Reference 15. In these papers, more information than just Noetherianity is extracted from the proof, e.g., that the tuple rank of a tuple of matrices is bounded in a proper closed subvariety of a polynomial functor capturing matrices Reference 11Reference 15, or that the -rank of a cubic is bounded Reference 9, or that the strength of a homogeneous form is bounded Reference 4. Our proof above does not directly yield such qualitative information. However, in Reference 6 we repair this defect for symmetric, alternating, and ordinary tensors and characteristic or sufficiently large.

If has characteristic , then symmetric powers are irreducible -modules, and one can prove Noetherianity for direct sums of these without the need for more general polynomial functors—though also without the proof becoming any easier. But in general characteristic, symmetric powers need not be irreducible, and polynomial functors need not be completely reducible into irreducible summands, so reducing modulo an irreducible subfunctor is the only natural thing to do.

The idea further to do induction on and to use directional derivatives is new, but inspired by techniques used earlier in Reference 10Reference 11Reference 12, where a determinant is regarded as an affine-linear polynomial in one matrix entry, whose coefficient is a determinant of lower order, and induction is done over that order.

2.11. An open problem

The most tantalising open problem in this area is the following.

Question 27.

Let be a finite-degree polynomial functor over an infinite field . Does any sequence of ideals in eventually become constant?

For of degree at most , the answer is yes, and it follows from the stronger statement that the ring , acted upon by via , is -Noetherian for any Noetherian ground ring Reference 5Reference 8Reference 19. For and in characteristic , the answer is also yes, since we know all -stable ideals from Reference 2 and Reference 1, respectively. In Reference 22 a much stronger result than this is established for and , namely, that finitely generated modules over are also Noetherian. These questions were first raised, in the setting of twisted commutative algebras, in Reference 30. They remain widely open even for more general degree-two functors, and also for, say, in positive characteristic.

Acknowledgments

The author thanks Arthur Bik, Michał Lasoń, Florian Oosterhof, and Andrew Snowden for useful discussions and comments on an earlier version of this paper. The author also thanks the organisers of the April 2016 Banff workshop on Free Resolutions, Representations, and Asymptotic Algebra for bringing together participants with a wide variety of backgrounds—they have strongly influenced my understanding of these infinite algebraic structures. Finally, the author thanks the referees for several valuable suggestions, including the running Example 11.

Mathematical Fragments

Theorem 1.

Let be the category of finite-dimensional vector spaces over an infinite field , and let be a finite-degree polynomial functor. Then is Noetherian as a -topological space.

Corollary 6.

Let be an infinite field, fix natural numbers , and consider the polynomial functor . Then is a Noetherian -topological space, and hence its limit is -Noetherian.

Corollary 9.

Let be an infinite field, let be the category of finite-dimensional vector spaces over , let be a positive integer, and let be a functor from to such that for any the map is polynomial of uniformly bounded degree. Then is a Noetherian -topological space.

Example 11.

As a running example to illustrate the proof, we assume that , where is the space of symmetric two-tensors (matrices) and is the space of skew-symmetric tensors. Let be the standard coordinates on , let be the standard coordinates on extended to by declaring them zero on , and let be the standard coordinates on , similarly extended to .

We then have

Take , the variety of rank tensors. Then is the entire space , but , so we may take and for the determinant

We take and find

In this case, for , is the subvariety of on which the -orbit of vanishes identically, i.e., is the set of rank tensors in . This is a coincidence; in the general setting of the proof, does not embed into , but it always has a lower-degree polynomial vanishing on it. We discuss the complement in Example 26.

Example 13.

In our running Example 11, have degrees , respectively.

Lemma 14.

For any polynomial functor of degree , is a polynomial functor of degree whose degree- homogeneous part is canonically isomorphic to that of .

Example 16.

In our running Example 11, , and

note that the degree-two part of is , so that .

Lemma 18.

Let be finite-dimensional vector spaces over the infinite field , let be a polynomial on , and consider the expression , a polynomial function of the triple . Then one of the following hold:

(1)

is independent of ; this happens if and only if factors through the projection ; or

(2)

the nonzero part of of lowest degree in is of the form for a unique (taken if ). Then for each fixed the map is in .

Lemma 21.

Assume that is algebraically closed and suppose that for each the map is surjective. Then restricts to a Zariski-homeomorphism from to a closed subset of .

Lemma 23.

The map from -closed subsets of to -closed subsets of is inclusion-preserving and injective. Consequently, Noetherianity of implies that of .

Equation (*)
Lemma 25.

For every -vector space the projection is a Zariski-homeomorphism with a closed subset of .

Example 26.

We pause and see what Lemma 25 says in the running Examples 11,13,16. Here and we take . Then is the variety of -matrices

of rank such that the upper-left -submatrix has a nonzero coordinate (recall that ).

The lemma says that forgetting the skew-symmetric part of is a closed embedding of into the open subset of where is nonzero. We prove this by showing that, on , each entry of the skew-symmetric part of can be expressed as a rational function in the entries of and the entries of the symmetric part of , with a denominator equal to .

We assume and consider a surjective linear map . Thinking of as a -matrix, we have

Since is a -closed subset, for all , the -determinant vanishes on the latter -matrix for all choices of and of . Hence expanding as a polynomial in , the coefficient of also vanishes for all . Take in and specialise to the linear map sending the th standard basis vector to , the th standard basis vector to , and all other standard basis vectors to zero. Then the matrix above reads

The coefficient of in the determinant of this matrix is

where the remaining terms do not involve . Using, as in Example 11, the variables for the symmetric and skew-symmetric parts of , and using the variables for the symmetric and skew-symmetric parts of , this reads

where the dots in the last two expressions contain only variables that we do not discard in the projection . This shows that, on , the coordinate can be expressed in the entries of and the coordinates on the symmetric part of , as desired.

References

Reference [1]
S. Abeasis and A. Del Fra, Young diagrams and ideals of Pfaffians, Adv. in Math. 35 (1980), no. 2, 158–178, DOI 10.1016/0001-8708(80)90046-8. MR560133,
Show rawAMSref \bib{Abeasis80b}{article}{ author={Abeasis, S.}, author={Del Fra, A.}, title={Young diagrams and ideals of Pfaffians}, journal={Adv. in Math.}, volume={35}, date={1980}, number={2}, pages={158--178}, issn={0001-8708}, review={\MR {560133}}, doi={10.1016/0001-8708(80)90046-8}, }
Reference [2]
Silvana Abeasis, The -invariant ideals in (Italian, with English summary), Rend. Mat. (6) 13 (1980), no. 2, 235–262. MR602662,
Show rawAMSref \bib{Abeasis80a}{article}{ author={Abeasis, Silvana}, title={The $\operatorname {GL}(V)$-invariant ideals in $S(S^{2}V)$}, language={Italian, with English summary}, journal={Rend. Mat. (6)}, volume={13}, date={1980}, number={2}, pages={235--262}, issn={0034-4427}, review={\MR {602662}}, }
Reference [3]
Tigran Ananyan and Melvin Hochster, Ideals generated by quadratic polynomials, Math. Res. Lett. 19 (2012), no. 1, 233–244, DOI 10.4310/MRL.2012.v19.n1.a18. MR2923188,
Show rawAMSref \bib{Ananyan12}{article}{ author={Ananyan, Tigran}, author={Hochster, Melvin}, title={Ideals generated by quadratic polynomials}, journal={Math. Res. Lett.}, volume={19}, date={2012}, number={1}, pages={233--244}, issn={1073-2780}, review={\MR {2923188}}, doi={10.4310/MRL.2012.v19.n1.a18}, }
Reference [4]
Tigran Ananyan and Melvin Hochster, Small subalgebras of polynomial rings and Stillman’s conjecture, (2016), Preprint, arXiv:1610.09268.
Reference [5]
Matthias Aschenbrenner and Christopher J. Hillar, Finite generation of symmetric ideals, Trans. Amer. Math. Soc. 359 (2007), no. 11, 5171–5192, DOI 10.1090/S0002-9947-07-04116-5. MR2327026,
Show rawAMSref \bib{Aschenbrenner07}{article}{ author={Aschenbrenner, Matthias}, author={Hillar, Christopher J.}, title={Finite generation of symmetric ideals}, journal={Trans. Amer. Math. Soc.}, volume={359}, date={2007}, number={11}, pages={5171--5192}, issn={0002-9947}, review={\MR {2327026}}, doi={10.1090/S0002-9947-07-04116-5}, }
Reference [6]
Arthur Bik, Jan Draisma, and Rob H. Eggermont, Polynomials and tensors of bounded strength, Commun. Contemp. Math. (2018), To appear, arXiv:1805.01816.
Reference [7]
Thomas Church, Jordan S. Ellenberg, and Benson Farb, FI-modules and stability for representations of symmetric groups, Duke Math. J. 164 (2015), no. 9, 1833–1910, DOI 10.1215/00127094-3120274. MR3357185,
Show rawAMSref \bib{Church12}{article}{ author={Church, Thomas}, author={Ellenberg, Jordan S.}, author={Farb, Benson}, title={FI-modules and stability for representations of symmetric groups}, journal={Duke Math. J.}, volume={164}, date={2015}, number={9}, pages={1833--1910}, issn={0012-7094}, review={\MR {3357185}}, doi={10.1215/00127094-3120274}, }
Reference [8]
Daniel E. Cohen, Closure relations. Buchberger’s algorithm, and polynomials in infinitely many variables, Computation theory and logic, Lecture Notes in Comput. Sci., vol. 270, Springer, Berlin, 1987, pp. 78–87, DOI 10.1007/3-540-18170-9_156. MR907514,
Show rawAMSref \bib{Cohen87}{article}{ author={Cohen, Daniel E.}, title={Closure relations. Buchberger's algorithm, and polynomials in infinitely many variables}, conference={ title={Computation theory and logic}, }, book={ series={Lecture Notes in Comput. Sci.}, volume={270}, publisher={Springer, Berlin}, }, date={1987}, pages={78--87}, review={\MR {907514}}, doi={10.1007/3-540-18170-9\_156}, }
Reference [9]
Harm Derksen, Rob H. Eggermont, and Andrew Snowden, Topological noetherianity for cubic polynomials, Algebra Number Theory 11 (2017), no. 9, 2197–2212, DOI 10.2140/ant.2017.11.2197. MR3735467,
Show rawAMSref \bib{Derksen17}{article}{ author={Derksen, Harm}, author={Eggermont, Rob H.}, author={Snowden, Andrew}, title={Topological noetherianity for cubic polynomials}, journal={Algebra Number Theory}, volume={11}, date={2017}, number={9}, pages={2197--2212}, issn={1937-0652}, review={\MR {3735467}}, doi={10.2140/ant.2017.11.2197}, }
Reference [10]
Jan Draisma, Finiteness for the -factor model and chirality varieties, Adv. Math. 223 (2010), no. 1, 243–256, DOI 10.1016/j.aim.2009.08.008. MR2563217,
Show rawAMSref \bib{Draisma08b}{article}{ author={Draisma, Jan}, title={Finiteness for the $k$-factor model and chirality varieties}, journal={Adv. Math.}, volume={223}, date={2010}, number={1}, pages={243--256}, issn={0001-8708}, review={\MR {2563217}}, doi={10.1016/j.aim.2009.08.008}, }
Reference [11]
Jan Draisma and Rob H. Eggermont, Plücker varieties and higher secants of Sato’s Grassmannian, J. Reine Angew. Math. 737 (2018), 189–215, DOI 10.1515/crelle-2015-0035. MR3781335,
Show rawAMSref \bib{Draisma14a}{article}{ author={Draisma, Jan}, author={Eggermont, Rob H.}, title={Pl\"{u}cker varieties and higher secants of Sato's Grassmannian}, journal={J. Reine Angew. Math.}, volume={737}, date={2018}, pages={189--215}, issn={0075-4102}, review={\MR {3781335}}, doi={10.1515/crelle-2015-0035}, }
Reference [12]
Jan Draisma and Jochen Kuttler, Bounded-rank tensors are defined in bounded degree, Duke Math. J. 163 (2014), no. 1, 35–63, DOI 10.1215/00127094-2405170. MR3161311,
Show rawAMSref \bib{Draisma11d}{article}{ author={Draisma, Jan}, author={Kuttler, Jochen}, title={Bounded-rank tensors are defined in bounded degree}, journal={Duke Math. J.}, volume={163}, date={2014}, number={1}, pages={35--63}, issn={0012-7094}, review={\MR {3161311}}, doi={10.1215/00127094-2405170}, }
Reference [13]
Jan Draisma, Michał Lasoń, and Anton Leykin, Stillman’s conjecture via generic initial ideals, Commun. Algebra (2019), Special issue in honour of Gennady Lyubeznik, DOI 10.1080/00927872.2019.1574806.
Reference [14]
Jan Draisma and Florian M. Oosterhof, Markov random fields and iterated toric fibre products, Adv. in Appl. Math. 97 (2018), 64–79, DOI 10.1016/j.aam.2018.02.003. MR3777348,
Show rawAMSref \bib{Draisma16b}{article}{ author={Draisma, Jan}, author={Oosterhof, Florian M.}, title={Markov random fields and iterated toric fibre products}, journal={Adv. in Appl. Math.}, volume={97}, date={2018}, pages={64--79}, issn={0196-8858}, review={\MR {3777348}}, doi={10.1016/j.aam.2018.02.003}, }
Reference [15]
Rob H. Eggermont, Finiteness properties of congruence classes of infinite-by-infinite matrices, Linear Algebra Appl. 484 (2015), 290–303, DOI 10.1016/j.laa.2015.06.035. MR3385063,
Show rawAMSref \bib{Eggermont14}{article}{ author={Eggermont, Rob H.}, title={Finiteness properties of congruence classes of infinite-by-infinite matrices}, journal={Linear Algebra Appl.}, volume={484}, date={2015}, pages={290--303}, issn={0024-3795}, review={\MR {3385063}}, doi={10.1016/j.laa.2015.06.035}, }
Reference [16]
Daniel Erman, Steven Sam, and Andrew Snowden, Generalizations of Stillman’s conjecture via twisted commutative algebra, (2017), Preprint, arXiv:1804.09807.
Reference [17]
Daniel Erman, Steven V. Sam, and Andrew Snowden, Big polynomial rings and Stillman’s conjecture, (2018), Preprint, arXiv:1801.09852.
Reference [18]
Eric M. Friedlander and Andrei Suslin, Cohomology of finite group schemes over a field, Invent. Math. 127 (1997), no. 2, 209–270, DOI 10.1007/s002220050119. MR1427618,
Show rawAMSref \bib{Friedlander97}{article}{ author={Friedlander, Eric M.}, author={Suslin, Andrei}, title={Cohomology of finite group schemes over a field}, journal={Invent. Math.}, volume={127}, date={1997}, number={2}, pages={209--270}, issn={0020-9910}, review={\MR {1427618}}, doi={10.1007/s002220050119}, }
Reference [19]
Christopher J. Hillar and Seth Sullivant, Finiteness theorems in infinite dimensional polynomial rings and applications, personal communication, 2008.
Reference [20]
J. M. Landsberg and Giorgio Ottaviani, Equations for secant varieties of Veronese and other varieties, Ann. Mat. Pura Appl. (4) 192 (2013), no. 4, 569–606, DOI 10.1007/s10231-011-0238-6. MR3081636,
Show rawAMSref \bib{Landsberg11}{article}{ author={Landsberg, J. M.}, author={Ottaviani, Giorgio}, title={Equations for secant varieties of Veronese and other varieties}, journal={Ann. Mat. Pura Appl. (4)}, volume={192}, date={2013}, number={4}, pages={569--606}, issn={0373-3114}, review={\MR {3081636}}, doi={10.1007/s10231-011-0238-6}, }
Reference [21]
J. M. Landsberg and L. Manivel, On the ideals of secant varieties of Segre varieties, Found. Comput. Math. 4 (2004), no. 4, 397–422, DOI 10.1007/s10208-003-0115-9. MR2097214,
Show rawAMSref \bib{Landsberg04}{article}{ author={Landsberg, J. M.}, author={Manivel, L.}, title={On the ideals of secant varieties of Segre varieties}, journal={Found. Comput. Math.}, volume={4}, date={2004}, number={4}, pages={397--422}, issn={1615-3375}, review={\MR {2097214}}, doi={10.1007/s10208-003-0115-9}, }
Reference [22]
Rohit Nagpal, Steven V. Sam, and Andrew Snowden, Noetherianity of some degree two twisted commutative algebras, Selecta Math. (N.S.) 22 (2016), no. 2, 913–937, DOI 10.1007/s00029-015-0205-y. MR3477338,
Show rawAMSref \bib{Nagpal15}{article}{ author={Nagpal, Rohit}, author={Sam, Steven V.}, author={Snowden, Andrew}, title={Noetherianity of some degree two twisted commutative algebras}, journal={Selecta Math. (N.S.)}, volume={22}, date={2016}, number={2}, pages={913--937}, issn={1022-1824}, review={\MR {3477338}}, doi={10.1007/s00029-015-0205-y}, }
Reference [23]
Irena Peeva and Mike Stillman, Open problems on syzygies and Hilbert functions, J. Commut. Algebra 1 (2009), no. 1, 159–195, DOI 10.1216/JCA-2009-1-1-159. MR2462384,
Show rawAMSref \bib{Peeva09}{article}{ author={Peeva, Irena}, author={Stillman, Mike}, title={Open problems on syzygies and Hilbert functions}, journal={J. Commut. Algebra}, volume={1}, date={2009}, number={1}, pages={159--195}, issn={1939-0807}, review={\MR {2462384}}, doi={10.1216/JCA-2009-1-1-159}, }
Reference [24]
Andrew Putman and Steven V. Sam, Representation stability and finite linear groups, Duke Math. J. 166 (2017), no. 13, 2521–2598, DOI 10.1215/00127094-2017-0008. MR3703435,
Show rawAMSref \bib{Putman14}{article}{ author={Putman, Andrew}, author={Sam, Steven V.}, title={Representation stability and finite linear groups}, journal={Duke Math. J.}, volume={166}, date={2017}, number={13}, pages={2521--2598}, issn={0012-7094}, review={\MR {3703435}}, doi={10.1215/00127094-2017-0008}, }
Reference [25]
Claudiu Raicu, Secant varieties of Segre-Veronese varieties, Algebra Number Theory 6 (2012), no. 8, 1817–1868, DOI 10.2140/ant.2012.6.1817. MR3033528,
Show rawAMSref \bib{Raicu10}{article}{ author={Raicu, Claudiu}, title={Secant varieties of Segre-Veronese varieties}, journal={Algebra Number Theory}, volume={6}, date={2012}, number={8}, pages={1817--1868}, issn={1937-0652}, review={\MR {3033528}}, doi={10.2140/ant.2012.6.1817}, }
Reference [26]
Johannes Rauh and Seth Sullivant, Lifting Markov bases and higher codimension toric fiber products, J. Symbolic Comput. 74 (2016), 276–307, DOI 10.1016/j.jsc.2015.07.003. MR3424043,
Show rawAMSref \bib{Rauh14}{article}{ author={Rauh, Johannes}, author={Sullivant, Seth}, title={Lifting Markov bases and higher codimension toric fiber products}, journal={J. Symbolic Comput.}, volume={74}, date={2016}, pages={276--307}, issn={0747-7171}, review={\MR {3424043}}, doi={10.1016/j.jsc.2015.07.003}, }
Reference [27]
Steven V. Sam, Ideals of bounded rank symmetric tensors are generated in bounded degree, Invent. Math. 207 (2017), no. 1, 1–21, DOI 10.1007/s00222-016-0668-2. MR3592755,
Show rawAMSref \bib{Sam15b}{article}{ author={Sam, Steven V.}, title={Ideals of bounded rank symmetric tensors are generated in bounded degree}, journal={Invent. Math.}, volume={207}, date={2017}, number={1}, pages={1--21}, issn={0020-9910}, review={\MR {3592755}}, doi={10.1007/s00222-016-0668-2}, }
Reference [28]
Steven V. Sam, Syzygies of bounded rank symmetric tensors are generated in bounded degree, Math. Ann. 368 (2017), no. 3-4, 1095–1108, DOI 10.1007/s00208-016-1509-8. MR3673648,
Show rawAMSref \bib{Sam16}{article}{ author={Sam, Steven V.}, title={Syzygies of bounded rank symmetric tensors are generated in bounded degree}, journal={Math. Ann.}, volume={368}, date={2017}, number={3-4}, pages={1095--1108}, issn={0025-5831}, review={\MR {3673648}}, doi={10.1007/s00208-016-1509-8}, }
Reference [29]
Steven V. Sam and Andrew Snowden, Gröbner methods for representations of combinatorial categories, J. Amer. Math. Soc. 30 (2017), no. 1, 159–203, DOI 10.1090/jams/859. MR3556290,
Show rawAMSref \bib{Sam14}{article}{ author={Sam, Steven V.}, author={Snowden, Andrew}, title={Gr\"{o}bner methods for representations of combinatorial categories}, journal={J. Amer. Math. Soc.}, volume={30}, date={2017}, number={1}, pages={159--203}, issn={0894-0347}, review={\MR {3556290}}, doi={10.1090/jams/859}, }
Reference [30]
Andrew Snowden, Syzygies of Segre embeddings and -modules, Duke Math. J. 162 (2013), no. 2, 225–277, DOI 10.1215/00127094-1962767. MR3018955,
Show rawAMSref \bib{Snowden10}{article}{ author={Snowden, Andrew}, title={Syzygies of Segre embeddings and $\Delta $-modules}, journal={Duke Math. J.}, volume={162}, date={2013}, number={2}, pages={225--277}, issn={0012-7094}, review={\MR {3018955}}, doi={10.1215/00127094-1962767}, }
Reference [31]
Steven V. Sam and Andrew Snowden, Introduction to twisted commutative algebras, (2012), Preprint, arXiv:1209.5122.
Reference [32]
Steven V. Sam and Andrew Snowden, GL-equivariant modules over polynomial rings in infinitely many variables, Trans. Amer. Math. Soc. 368 (2016), no. 2, 1097–1158, DOI 10.1090/tran/6355. MR3430359,
Show rawAMSref \bib{Sam12}{article}{ author={Sam, Steven V.}, author={Snowden, Andrew}, title={GL-equivariant modules over polynomial rings in infinitely many variables}, journal={Trans. Amer. Math. Soc.}, volume={368}, date={2016}, number={2}, pages={1097--1158}, issn={0002-9947}, review={\MR {3430359}}, doi={10.1090/tran/6355}, }
Reference [33]
Terence Tao and Will Sawin, Notes on the “slice rank” of tensors, (2016), https://terrytao.wordpress.com/2016/08/24/notes-on-the-slice-rank-of-tensors/.

Article Information

MSC 2010
Primary: 13E05 (Noetherian rings and modules), 14M12 (Determinantal varieties), 15A69 (Multilinear algebra, tensor products), 20G05 (Representation theory)
Author Information
Jan Draisma
Mathematisches Institut, Universität Bern, Sidlerstrasse 5, 3012 Bern; and Eindhoven University of Technology, P.O. Box 513, 5600 MB, Eindhoven, The Netherlands
jan.draisma@math.unibe.ch
ORCID
MathSciNet
Additional Notes

The author was partially supported by the NWO Vici grant entitled Stabilisation in Algebra and Geometry.

Journal Information
Journal of the American Mathematical Society, Volume 32, Issue 3, ISSN 1088-6834, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on , revised on , and published on .
Copyright Information
Copyright 2019 American Mathematical Society
Article References
  • Permalink
  • Permalink (PDF)
  • DOI 10.1090/jams/923
  • MathSciNet Review: 3981986
  • Show rawAMSref \bib{3981986}{article}{ author={Draisma, Jan}, title={Topological Noetherianity of polynomial functors}, journal={J. Amer. Math. Soc.}, volume={32}, number={3}, date={2019-07}, pages={691-707}, issn={0894-0347}, review={3981986}, doi={10.1090/jams/923}, }

Settings

Change font size
Resize article panel
Enable equation enrichment

Note. To explore an equation, focus it (e.g., by clicking on it) and use the arrow keys to navigate its structure. Screenreader users should be advised that enabling speech synthesis will lead to duplicate aural rendering.

For more information please visit the AMS MathViewer documentation.