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

Symmetric subvarieties of infinite affine space

Rohit Nagpal Department of Mathematics, University of Michigan, Ann Arbor, MI rohitna@umich.edu http://www-personal.umich.edu/~rohitna/  and  Andrew Snowden Department of Mathematics, University of Michigan, Ann Arbor, MI asnowden@umich.edu http://www-personal.umich.edu/~asnowden/
Abstract.

We classify the subvarieties of infinite dimensional affine space that are stable under the infinite symmetric group. We determine the defining equations and point sets of these varieties as well as the containments between them.

RN was partially supported by NSF DMS-1638352. AS was supported by NSF DMS-1453893.

1. Introduction

Cohen [Co, Co2] proved that ideals in the infinite variable polynomial ring R=𝐂[ξ1,ξ2,]R=\mathbf{C}[\xi_{1},\xi_{2},\ldots] that are stable under the infinite symmetric group 𝔖\mathfrak{S} satisfy the ascending chain condition; in other words, RR is 𝔖\mathfrak{S}-noetherian. This suggests that 𝔖\mathfrak{S}-equivariant commutative algebra over RR should be well-behaved. We have undertaken a detailed study of this theory, and found that this is indeed the case. In this paper, we classify the radical 𝔖\mathfrak{S}-ideals of RR. In subsequent papers [NS1, NS2], we will build on the results of this paper to determine the structure of arbitrary 𝔖\mathfrak{S}-ideals of RR, and establish results such as primary decomposition. Our work is motivated by the wide variety of applications Cohen’s theorem has found in the last decade (e.g., [AH, DE, DK, GN, HS, KLS, LNNR, NR]); see [Dr] for a good introduction.

1.1. The classification theorem

Let 𝔛=Spec(R)=𝐀\mathfrak{X}=\operatorname{Spec}(R)=\mathbf{A}^{\infty} be infinite dimensional affine space over the complex numbers (we allow arbitrary noetherian coefficient rings in the body of the paper). The goal of this paper is to classify the 𝔖\mathfrak{S}-stable Zariski closed subsets of 𝔛\mathfrak{X} and understand their structure.

We say that a 𝐂\mathbf{C}-point x=(xi)i1x=(x_{i})_{i\geq 1} of 𝔛\mathfrak{X} is finitary if the coordinates of xx assume only finitely many values, that is, {xii1}\{x_{i}\mid i\geq 1\} is a finite subset of 𝐂\mathbf{C}. This notion is closely related to discriminants. Indeed, let Δr=1i<jr(ξiξj)\Delta_{r}=\prod_{1\leq i<j\leq r}(\xi_{i}-\xi_{j}) be the rrth discriminant. Then the point xx is finitary if and only if there is some rr such that Δr(σx)=0\Delta_{r}(\sigma x)=0 for all σ𝔖\sigma\in\mathfrak{S}. It is not difficult to show that any non-zero 𝔖\mathfrak{S}-stable ideal of RR contains some discriminant (Proposition 2.6). Thus any proper 𝔖\mathfrak{S}-stable closed subset of 𝔛\mathfrak{X} is contained in the finitary locus 𝔛fin\mathfrak{X}_{\mathrm{fin}}. For this reason, finitary points play a prominent role in this paper.

Suppose that xx is a finitary point. Let a1,,ara_{1},\ldots,a_{r} be the distinct values its coordinates assume, and let λi\lambda_{i} be the number of times aia_{i} appears. Relabeling if necessary, we assume that λ1λr\lambda_{1}\geq\cdots\geq\lambda_{r}; note that λ1\lambda_{1} is necessarily \infty. We call λ=(λ1,,λr)\lambda=(\lambda_{1},\ldots,\lambda_{r}) the type of xx, and generally refer to a tuple of this sort as an \infty-parition. For example, the point

(3,3,3,5,5,6,7,6,7,,6,7,)(3,3,3,5,5,6,7,6,7,\ldots,6,7,\ldots)

has type (,,3,2)(\infty,\infty,3,2), since 6 and 7 occur infinitely, 3 occurs thrice, and 5 occurs twice. Let 𝔛[λ]\mathfrak{X}_{[\lambda]} be the set of all finitary points of type λ\lambda. We have 𝔛fin=𝔛[λ]\mathfrak{X}_{\mathrm{fin}}=\bigsqcup\mathfrak{X}_{[\lambda]}, where the union is over all \infty-partitions λ\lambda, and so we can regard the 𝔛[λ]\mathfrak{X}_{[\lambda]} as defining a stratification of 𝔛fin\mathfrak{X}_{\mathrm{fin}}.

Before stating our theorem, we need to introduce one more concept. We say that a topological space XX with an action of a group GG is GG-irreducible if it cannot be expressed as a union of two GG-stable proper closed subsets. It is an easy consequence of Cohen’s theorem that any 𝔖\mathfrak{S}-stable closed subvariety of 𝔛\mathfrak{X} is a finite union of 𝔖\mathfrak{S}-irreducible closed subvarieties. It therefore suffices to study the 𝔖\mathfrak{S}-irreducible subvarieties of 𝔛\mathfrak{X}. We note that if GG is a finite group acting on a variety XX then any GG-irreducible subvariety of XX is simply the union of the GG-translates of an irreducible subvariety.

We can now state our classification theorem:

Theorem 1.1 (Theorem 8.2).

The following two sets are in natural bijection:

  1. (a)

    The set of 𝔖\mathfrak{S}-irreducible proper closed subvarieties \mathfrak{Z} of 𝔛\mathfrak{X}.

  2. (b)

    The set of pairs (λ,Z)(\lambda,Z), where λ=(λ1,,λr)\lambda=(\lambda_{1},\ldots,\lambda_{r}) is an \infty-partition and ZZ is an Aut(λ)\operatorname{Aut}(\lambda)-irreducible closed subvariety of 𝐀[r]\mathbf{A}^{[r]}. Here Aut(λ)\operatorname{Aut}(\lambda) is the subgroup of 𝔖r\mathfrak{S}_{r} stabilizing λ\lambda, and 𝐀[r]\mathbf{A}^{[r]} is the open subvariety of 𝐀r\mathbf{A}^{r} consisting of points with distinct coordinates.

It is not difficult to describe the bijection in Theorem 1.1. To do this, it will be useful to introduce a refinement of the type stratification of 𝔛fin\mathfrak{X}_{\mathrm{fin}}. Let 𝒰={𝒰1,,𝒰r}\mathcal{U}=\{\mathcal{U}_{1},\ldots,\mathcal{U}_{r}\} be a partition of the index set []={1,2,}[\infty]=\{1,2,\ldots\} into finitely many disjoint and non-empty pieces. Define 𝔛𝒰\mathfrak{X}_{\mathcal{U}} (resp. 𝔛[𝒰]\mathfrak{X}_{[\mathcal{U}]}) be the subset of 𝔛\mathfrak{X} consisting of points xx such that xi=xjx_{i}=x_{j} if ii and jj belong to the same part of 𝒰\mathcal{U} (resp. xi=xjx_{i}=x_{j} if and only if ii and jj belong to the same part). The space 𝔛𝒰\mathfrak{X}_{\mathcal{U}} is isomorphic to 𝐀r\mathbf{A}^{r}, and under this isomorphism 𝔛[𝒰]\mathfrak{X}_{[\mathcal{U}]} corresponds to 𝐀[r]\mathbf{A}^{[r]}. Let λi=#𝒰i\lambda_{i}=\#\mathcal{U}_{i}, relabeling if necessary so that λ=(λ1,,λr)\lambda=(\lambda_{1},\ldots,\lambda_{r}) is an \infty-partition; we call this the \infty-partition associated to 𝒰\mathcal{U}, and say 𝒰\mathcal{U} is of type λ\lambda. Then any point of 𝔛[𝒰]\mathfrak{X}_{[\mathcal{U}]} has type λ\lambda, and 𝔛[λ]\mathfrak{X}_{[\lambda]} is the union of the 𝔖\mathfrak{S}-translates of 𝔛[𝒰]\mathfrak{X}_{[\mathcal{U}]}.

We now describe the bijection in Theorem 1.1. Suppose that \mathfrak{Z} is an 𝔖\mathfrak{S}-irreducible closed subvariety of 𝔛\mathfrak{X}. Let Λ\Lambda_{\mathfrak{Z}} be the set of all \infty-partitions λ\lambda such that \mathfrak{Z} has a point of type λ\lambda. We show that Λ\Lambda_{\mathfrak{Z}} contains a unique maximal element with respect to a partial order \preceq on \infty-partitions. Let λ=(λ1,,λr)\lambda=(\lambda_{1},\ldots,\lambda_{r}) be this maximal element, and pick a partition 𝒰\mathcal{U} of [][\infty] of type λ\lambda. Let Z=𝔛[𝒰]Z=\mathfrak{Z}\cap\mathfrak{X}_{[\mathcal{U}]}, where we identify 𝔛[𝒰]\mathfrak{X}_{[\mathcal{U}]} with 𝐀[r]\mathbf{A}^{[r]}. Then \mathfrak{Z} corresponds to (λ,Z)(\lambda,Z). In the reverse direction, suppose that λ=(λ1,,λr)\lambda=(\lambda_{1},\ldots,\lambda_{r}) and ZZ are given. Again, pick 𝒰\mathcal{U} of type λ\lambda. Then the corresponding 𝔖\mathfrak{S}-irreducible closed subvariety of 𝔛\mathfrak{X} is the Zariski closure of the 𝔖\mathfrak{S}-orbit of ZZ, where we regard ZZ as a subset of 𝔛\mathfrak{X} via the identification 𝐀[r]=𝔛[𝒰]\mathbf{A}^{[r]}=\mathfrak{X}_{[\mathcal{U}]}.

1.2. C3 varieties

While Theorem 1.1 does give a complete classification of 𝔖\mathfrak{S}-subvarieties of 𝔛\mathfrak{X}, it is not powerful enough to answer some finer questions, such as:

  • How does one describe the entire point set of \mathfrak{Z} from (λ,Z)(\lambda,Z)?

  • How does one determine the equations of \mathfrak{Z} from those of ZZ?

  • How does one determine a containment \mathfrak{Z}\subset\mathfrak{Z}^{\prime} in terms of the data (λ,Z)(\lambda,Z) and (λ,Z)(\lambda^{\prime},Z^{\prime})?

In fact, Theorem 1.1 is a corollory of a stronger result, that we now describe, which allows us to address these questions (and more like them).

For an \infty-partition λ=(λ1,,λr)\lambda=(\lambda_{1},\ldots,\lambda_{r}), put 𝒳λ=𝐀r\mathcal{X}_{\lambda}=\mathbf{A}^{r}. We regard the system 𝒳={𝒳λ}λ\mathcal{X}=\{\mathcal{X}_{\lambda}\}_{\lambda} as a single object. Given an 𝔖\mathfrak{S}-stable closed subset \mathfrak{Z} of 𝔛\mathfrak{X}, we define a subsystem 𝒵=Φ()\mathcal{Z}=\Phi(\mathfrak{Z}) of 𝒳\mathcal{X} as follows. For an \infty-partition λ\lambda, let 𝒰\mathcal{U} be a partition of [][\infty] of type λ\lambda, and put 𝒵λ=𝔛𝒰\mathcal{Z}_{\lambda}=\mathfrak{Z}\cap\mathfrak{X}_{\mathcal{U}}, where, as usual, we idenify 𝔛𝒰\mathfrak{X}_{\mathcal{U}} with 𝐀r=𝒳λ\mathbf{A}^{r}=\mathcal{X}_{\lambda}. This is well-defined (i.e., independent of the choice of 𝒰\mathcal{U}) since \mathfrak{Z} is 𝔖\mathfrak{S}-stable. Note that 𝒵λ\mathcal{Z}_{\lambda} is Zariski closed in 𝒳λ\mathcal{X}_{\lambda}, and so we can regard 𝒵\mathcal{Z} as a system of finite dimensional varieties.

Since Φ()\Phi(\mathfrak{Z}) records all the finitary points of \mathfrak{Z}, it follows that it completely determines \mathfrak{Z}. In other words, the construction Φ\Phi is injective. Much of the work in this paper goes into determining the image of Φ\Phi, or, in other words, determining the precise conditions that the system 𝒵\mathcal{Z} of varieties must satisfy.

We now give an example of one of the simplest conditions that 𝒵\mathcal{Z} must satisfy. Suppose that x=(b,a,a,)x=(b,a,a,\ldots) is a 𝐂\mathbf{C}-point of \mathfrak{Z}. Let xn=(a,,a,b,a,a,)x_{n}=(a,\ldots,a,b,a,a,\ldots), where all coordinates except the nnth are equal to aa. Then xnx_{n} belongs to the 𝔖\mathfrak{S}-orbit of xx, and therefore belongs to \mathfrak{Z}. The sequence {xn}n1\{x_{n}\}_{n\geq 1} converges coordinate-wise to y=(a,a,)y=(a,a,\ldots), and so yy also belongs to \mathfrak{Z}. (This is a simple compactness argument, see §2.3.) We can rephrase the above observation in terms of 𝒵\mathcal{Z} as follows. The point xx shows that 𝒵(,1)𝐀2\mathcal{Z}_{(\infty,1)}\subset\mathbf{A}^{2} contains the point (a,b)(a,b), while the point yy shows that 𝒵()𝐀1\mathcal{Z}_{(\infty)}\subset\mathbf{A}^{1} contains the point aa. We thus see that if p1:𝐀2𝐀1p_{1}\colon\mathbf{A}^{2}\to\mathbf{A}^{1} is the projection map onto the first coordinate then p1(𝒵(,1))𝒵()p_{1}(\mathcal{Z}_{(\infty,1)})\subset\mathcal{Z}_{(\infty)}. This is a non-trivial condition on 𝒵\mathcal{Z}.

By generalizing the above argument, we are led to a number of natural conditions that 𝒵\mathcal{Z} must satisfy. We define a “C3 subvariety” of 𝒳\mathcal{X} to be a subsystem satisfying these conditions. The following is the main theorem of this paper:

Theorem 1.2 (Theorem 7.5).

The construction Φ\Phi defines a bijection

{𝔖-stable closed subsets of 𝔛}{C3 subvarieties of 𝒳}\{\text{$\mathfrak{S}$-stable closed subsets of $\mathfrak{X}$}\}\to\{\text{C3 subvarieties of $\mathcal{X}$}\}

The inverse bijection Ψ\Psi is described excplitily. We use this theorem to answer the questions posed above. Additionally, we apply the theorem in §9, where we prove a result on the support of 𝔖\mathfrak{S}-equivariant RR-modules. This result will play an important role in the follow-up paper [NS1].

1.3. Some examples

We give some simple examples to illustrate Theorem 1.1.

Example 1.3.

Let λ=(,)\lambda=(\infty,\infty) and let Z𝒳[λ]Z\subset\mathcal{X}_{[\lambda]} be the 0-dimensional closed subvariety {(0,1),(1,0)}\{(0,1),(1,0)\}. Note that Aut(λ)=𝔖2\operatorname{Aut}(\lambda)=\mathfrak{S}_{2} acts on 𝒳[λ]𝐀2\mathcal{X}_{[\lambda]}\subset\mathbf{A}^{2} by permuting the two coordinates. We thus see that ZZ is stable under Aut(λ)\operatorname{Aut}(\lambda) and is Aut(λ)\operatorname{Aut}(\lambda)-irreducible. Let \mathfrak{Z} be the 𝔖\mathfrak{S}-irreducible closed subvariety of 𝔛\mathfrak{X} corresponding to (λ,Z)(\lambda,Z). Then \mathfrak{Z} consists of those points (xi)i1(x_{i})_{i\geq 1} of 𝔛\mathfrak{X} such that xi{0,1}x_{i}\in\{0,1\} for all ii. See §8.6 for details. ∎

Example 1.4.

Let λ=(,n)\lambda=(\infty,n), with nn finite, and let Z={(0,1)}Z=\{(0,1)\}. In this case, Aut(λ)\operatorname{Aut}(\lambda) is trivial, and so ZZ is clearly Aut(λ)\operatorname{Aut}(\lambda)-irreducible. Let \mathfrak{Z} be the 𝔖\mathfrak{S}-irreducible closed subvariety of 𝔛\mathfrak{X} corresponding to (λ,Z)(\lambda,Z). Then \mathfrak{Z} consists of those points (xi)i1(x_{i})_{i\geq 1} of 𝔛\mathfrak{X} such that xi{0,1}x_{i}\in\{0,1\} for all ii, and #{ixi=1}n\#\{i\mid x_{i}=1\}\leq n. ∎

Example 1.5.

Once again, take λ=(,n)\lambda=(\infty,n). Let ZZ be an irreducible curve in 𝒳[λ]\mathcal{X}_{[\lambda]}, such as y2=x3+1y^{2}=x^{3}+1. Then ZZ is Aut(λ)\operatorname{Aut}(\lambda)-irreducible, as in the previous example. Let \mathfrak{Z} correspond to (λ,Z)(\lambda,Z). Then \mathfrak{Z} consists of those points (xi)i1(x_{i})_{i\geq 1} satisfying the following condition: there is a point (a,b)Z¯(a,b)\in\overline{Z} such that xi{a,b}x_{i}\in\{a,b\} for all ii, and #{ixia}n\#\{i\mid x_{i}\neq a\}\leq n. Here Z¯\overline{Z} denotes the Zariski closure of ZZ in 𝒳λ=𝐀2\mathcal{X}_{\lambda}=\mathbf{A}^{2}. ∎

1.4. Further directions

In this paper, we classify the 𝔖\mathfrak{S}-stable subvarieties of 𝐀\mathbf{A}^{\infty}, or, equivalently, the 𝔖\mathfrak{S}-stable radical ideals of RR. We are currently pursuing two natural generalizations of this result. First, in forthcoming papers [NS1, NS2], we determine (in a certain sense) the structure of arbitrary 𝔖\mathfrak{S}-stable ideals of RR, and establish results about equivariant modules as well. And second, in ongoing joint work with Vignesh Jagathese, we are classifying the 𝔖\mathfrak{S}-stable subvarieties of XX^{\infty} for an arbitrary variety XX (the present paper treating the case X=𝐀1X=\mathbf{A}^{1}).

1.5. Notation

We list the most important notation used throughout the paper:

[n]{[n]} :

the set {1,,n}\{1,\ldots,n\} (we allow n=n=\infty)

𝔖\mathfrak{S} :

the (small) symmetric group on [][\infty]

AA :

the coefficient ring (often noetherian)

WW :

the spectrum of AA

RR :

the polynomial ring over AA in variables {ξi}i1\{\xi_{i}\}_{i\geq 1}

𝔛\mathfrak{X} :

the spectrum of RR

\mathfrak{Z} :

a subset of 𝔛\mathfrak{X}

𝒳\mathcal{X} :

the system {𝒳λ}λ\{\mathcal{X}_{\lambda}\}_{\lambda} indexed by \infty-compositions λ\lambda

𝒵\mathcal{Z} :

a subset (really subsystem) of 𝒳\mathcal{X}

2. The space 𝔛\mathfrak{X}

2.1. Definitions

Let AA be a commutative ring and let W=Spec(A)W=\operatorname{Spec}(A) be its spectrum. Let R=RA=A[ξi]i1R=R_{A}=A[\xi_{i}]_{i\geq 1} be the infinite variable polynomial ring over AA in the variables ξi\xi_{i}. Let 𝔛=𝔛W=Spec(R)\mathfrak{X}=\mathfrak{X}_{W}=\operatorname{Spec}(R). We regard 𝔛\mathfrak{X} as a topological space (under the Zariski topology), and include its scheme-theoretic (non-closed) points.

We typically work with points in 𝔛\mathfrak{X} by using KK-points, for variable fields KK. Every point of 𝔛\mathfrak{X} comes from a KK-point of 𝔛\mathfrak{X} for some field KK, and a KK-point xx and a KK^{\prime}-point xx^{\prime} define the same point of 𝔛\mathfrak{X} if and only if there is a field K′′K^{\prime\prime} and embeddings KK′′K\to K^{\prime\prime} and KK′′K^{\prime}\to K^{\prime\prime} such that xx and xx^{\prime} define the same K′′K^{\prime\prime}-points of 𝔛\mathfrak{X}. We can therefore define a condition on points of 𝔛\mathfrak{X} by defining a condition on KK-points that is invariant under field extension. For a subset \mathfrak{Z} of 𝔛\mathfrak{X}, we let (K)\mathfrak{Z}(K) be the subset of 𝔛(K)\mathfrak{X}(K) consisting of KK-points whose image lies in \mathfrak{Z}. We note that X(K)=W(K)×KX(K)=W(K)\times K^{\infty}. For readers not familiar with scheme theory, one can assume that AA is an algebraically closed field and work only with closed points without losing much.

Let 𝔖=n1𝔖n\mathfrak{S}=\bigcup_{n\geq 1}\mathfrak{S}_{n} be the “small” infinite symmetric group and let 𝔖big=Aut([])\mathfrak{S}^{\rm big}=\operatorname{Aut}([\infty]) be the “big” infinite symmetric group; of course, 𝔖\mathfrak{S} is a subgroup of 𝔖big\mathfrak{S}^{\rm big}. The group 𝔖big\mathfrak{S}^{\rm big} (and therefore 𝔖\mathfrak{S} as well) acts on RR by permuting the variables, and thus on 𝔛\mathfrak{X} as well. We prefer to work with 𝔖\mathfrak{S}, since it is more algebraic in nature; however, the difference between these two groups is largely irrelevant for algebraic questions, as we see below (e.g., Proposition 2.4).

2.2. The noetherian property

By an 𝔖\mathfrak{S}-ideal of RR, we mean an ideal of RR that is stable under the group 𝔖\mathfrak{S}. Given a subset SS of RR, we let S\langle\!\langle S\rangle\!\rangle be the ideal generated by the 𝔖\mathfrak{S}-orbit of SS; this is the minimal 𝔖\mathfrak{S}-ideal containing SS. By an 𝔖\mathfrak{S}-subset of 𝔛\mathfrak{X}, we mean a subset of 𝔛\mathfrak{X} that is stable under 𝔖\mathfrak{S}. The following result is due to Cohen [Co, Co2]:

Theorem 2.1.

Suppose that the coefficient ring AA is noetherian. Then:

  1. (a)

    The ascending chain condition holds for 𝔖\mathfrak{S}-ideals of RR.

  2. (b)

    Any 𝔖\mathfrak{S}-ideal of RR is generated by finitely many 𝔖\mathfrak{S}-orbits of elements.

  3. (c)

    The descending chain condition holds for Zariski closed 𝔖\mathfrak{S}-subsets of 𝔛\mathfrak{X}.

Corollary 2.2.

Let \mathfrak{Z} be a Zariski closed 𝔖\mathfrak{S}-subset of 𝔛\mathfrak{X}. Then =1r\mathfrak{Z}=\mathfrak{Z}_{1}\cup\cdots\cup\mathfrak{Z}_{r} where each i\mathfrak{Z}_{i} is an 𝔖\mathfrak{S}-irreducible closed subset of 𝔛\mathfrak{X}.

Proof.

The usual argument (from the non-equivariant case) applies. (In fact, by passing to the quotient space 𝔛/𝔖\mathfrak{X}/\mathfrak{S} one reduces to that case.) ∎

2.3. The Π\Pi-topology

The space 𝔛\mathfrak{X} carries the usual Zariski topology. However, it also has a second topology, what we call the Π\Pi-topology, that will be useful in this paper. The definition is as follows: a subset \mathfrak{Z} of 𝔛\mathfrak{X} is Π\Pi-closed if (K)\mathfrak{Z}(K) is closed in 𝔛(K)=W(K)×K\mathfrak{X}(K)=W(K)\times K^{\infty} for all fields KK; here W(K)W(K) and KK are given the discrete topology and 𝔛(K)\mathfrak{X}(K) is given the product topology. Concretely, a sequence (or net) {xi}\{x_{i}\} in 𝔛(K)\mathfrak{X}(K) converges to a point xx if each coordinate of xix_{i} (including the WW coordinate) is equal to the corresponding coordinate of xx for i0i\gg 0.

Proposition 2.3.

Every Zariski closed subset of 𝔛\mathfrak{X} is also Π\Pi-closed.

Proof.

Let \mathfrak{Z} be a Zariski-closed subset of 𝔛\mathfrak{X} and let {xi}\{x_{i}\} be a net in (K)\mathfrak{Z}(K) converging to a point x𝔛(K)x\in\mathfrak{X}(K) in the Π\Pi-topology. Suppose fRf\in R vanishes on \mathfrak{Z}, and uses the variables ξ1,,ξn\xi_{1},\ldots,\xi_{n}. By definition, there is some i0i_{0} such that the first nn coordinates of xix_{i} agree with those of xx for all ii0i\geq i_{0}. We thus see that f(x)=f(xi)=0f(x)=f(x_{i})=0 for any ii0i\geq i_{0}. Since all functions that vanish on \mathfrak{Z} vanish on xx, it follows that xx\in\mathfrak{Z}. Thus \mathfrak{Z} is Π\Pi-closed. ∎

Proposition 2.4.

Let \mathfrak{Z} be a Π\Pi-closed 𝔖\mathfrak{S}-subset of 𝔛\mathfrak{X}. Then \mathfrak{Z} is 𝔖big\mathfrak{S}^{\rm big}-stable.

Proof.

Let x(K)x\in\mathfrak{Z}(K), and let σ𝔖big\sigma\in\mathfrak{S}^{\rm big}. For each n1n\geq 1, pick τn𝔖\tau_{n}\in\mathfrak{S} such that τn\tau_{n} agrees with σ\sigma on [n][n]. Then τnx\tau_{n}x\in\mathfrak{Z} and τnx\tau_{n}x converges to σx\sigma x in the Π\Pi-topology. Since \mathfrak{Z} is Π\Pi-closed, we see that σx\sigma x\in\mathfrak{Z}. Thus \mathfrak{Z} is 𝔖big\mathfrak{S}^{\rm big}-stable. ∎

2.4. Discriminants

Given a commutative ring BB and elements b1,,bnBb_{1},\ldots,b_{n}\in B, we let Δ(b1,,bn)=1i<jn(bjbi)\Delta(b_{1},\ldots,b_{n})=\prod_{1\leq i<j\leq n}(b_{j}-b_{i}) be their discriminant. We put Δn=Δ(ξ1,,ξn)\Delta_{n}=\Delta(\xi_{1},\ldots,\xi_{n}), which we regard as an element of RR.

Proposition 2.5.

We have

σ𝔖n/𝔖n1sgn(σ)σ(ξnkΔn1)={Δnif k=n10if k<n1\sum_{\sigma\in\mathfrak{S}_{n}/\mathfrak{S}_{n-1}}\operatorname{sgn}(\sigma)\sigma\left(\xi_{n}^{k}\cdot\Delta_{n-1}\right)=\begin{cases}\Delta_{n}&\text{if $k=n-1$}\\ 0&\text{if $k<n-1$}\end{cases}
Proof.

It suffices to treat the case where A=𝐙A=\mathbf{Z}, so we assume this is the case. Call the sum FF. Since FF is a skew-invariant polynomial in ξ1,,ξn\xi_{1},\ldots,\xi_{n} it is therefore divisible by Δn\Delta_{n}. (Here we have used that 2 is a non-zerodivisor in AA.) If k<n1k<n-1 then deg(F)<deg(Δn)\deg(F)<\deg(\Delta_{n}), and so F=0F=0. Now suppose k=n1k=n-1. No monomial appearing in Δn1\Delta_{n-1} has an n1n-1 power in it: indeed, in the product description for Δn1\Delta_{n-1}, the variable ξi\xi_{i} appears in only n2n-2 factors. It follows that the coefficient of ξnn1\xi_{n}^{n-1} in FF is Δn1\Delta_{n-1}, and, in particular, non-zero. Since deg(F)=deg(Δn)\deg(F)=\deg(\Delta_{n}), it follows that FF is a non-zero scalar multiple of Δn\Delta_{n}. Comparing the coefficients of ξnn1\xi_{n}^{n-1}, we see that the scalar is 1. ∎

Proposition 2.6.

Let 𝔞\mathfrak{a} be a non-zero 𝔖\mathfrak{S}-ideal of RR. Then 𝔞\mathfrak{a} contains cΔnc\cdot\Delta_{n} for some nn and some non-zero cAc\in A. In fact, if f𝔞f\in\mathfrak{a} is non-zero then there is some non-zero coefficient cc of ff such that cΔnc\cdot\Delta_{n} belongs to 𝔞\mathfrak{a}.

Proof.

Let f𝔞f\in\mathfrak{a} be a non-zero element. Suppose ff uses the variables ξ1,,ξr\xi_{1},\ldots,\xi_{r}, and write

f=i=0dgi(ξ1,,ξr1)ξrif=\sum_{i=0}^{d}g_{i}(\xi_{1},\ldots,\xi_{r-1})\xi_{r}^{i}

with gd0g_{d}\neq 0. By induction on the number of variables, the 𝔖\mathfrak{S}-ideal generated by gd(ξ1,,ξr1)g_{d}(\xi_{1},\ldots,\xi_{r-1}) contains cΔmc\Delta_{m} for some mm and some non-zero c𝐤c\in\mathbf{k}. Let XA[𝔖]X\in A[\mathfrak{S}] be such that

Xgd(ξ1,,ξr1)=cΔ(ξa1,,ξam)Xg_{d}(\xi_{1},\ldots,\xi_{r-1})=c\Delta(\xi_{a_{1}},\ldots,\xi_{a_{m}})

for some indices a1,,ama_{1},\ldots,a_{m}. We assume, without loss of generality, that XX avoids the index rr, that is, all polynomials appearing in XX do not use ξr\xi_{r}, and all permutations appearing in XX fix rr.

Now, let {b1,,bd}[]\{b_{1},\ldots,b_{d}\}\subset[\infty] be disjoint from {1,,r}\{1,\ldots,r\} and any index appearing in XX. Let G𝔖G\subset\mathfrak{S} be the symmetric group on {r,b1,,bd}\{r,b_{1},\ldots,b_{d}\}, and let GGG^{\prime}\subset G be the stabilizer of rr. Put Y=σG/Gsgn(σ)σY=\sum_{\sigma\in G/G^{\prime}}\operatorname{sgn}(\sigma)\sigma, so that, by the previous lemma, we have

Yf=gd(ξ1,,ξr1)Δ(ξr,ξb1,,ξbd).Yf=g_{d}(\xi_{1},\ldots,\xi_{r-1})\Delta(\xi_{r},\xi_{b_{1}},\ldots,\xi_{b_{d}}).

Therefore,

XY(f)=cΔ(ξa1,,ξam)Δ(ξr,ξb1,,ξbd).XY(f)=c\Delta(\xi_{a_{1}},\ldots,\xi_{a_{m}})\Delta(\xi_{r},\xi_{b_{1}},\ldots,\xi_{b_{d}}).

(Note that Δ(ξr,ξb1,,ξbd)\Delta(\xi_{r},\xi_{b_{1}},\ldots,\xi_{b_{d}}) commutes with XX by our choices.) Let hh be the product of all linear forms of the form ξaiξbj\xi_{a_{i}}-\xi_{b_{j}} and ξaiξr\xi_{a_{i}}-\xi_{r}. Then

hXY(f)=cΔ(ξa1,,ξam,ξr,ξb1,,ξbd).hXY(f)=c\Delta(\xi_{a_{1}},\ldots,\xi_{a_{m}},\xi_{r},\xi_{b_{1}},\ldots,\xi_{b_{d}}).

We thus see (after applying an element of 𝔖\mathfrak{S}) that cΔn𝔞c\Delta_{n}\in\mathfrak{a}, with n=m+d+1n=m+d+1. ∎

Proposition 2.7.

Let S=R[(ξiξj)1]ijS=R[(\xi_{i}-\xi_{j})^{-1}]_{i\neq j}. Then extension and contraction give mutually inverse bijections

{ideals of A}{𝔖-ideals of S}.\{\text{ideals of $A$}\}\leftrightarrow\{\text{$\mathfrak{S}$-ideals of $S$}\}.
Proof.

It is clear that every ideal of AA is the contraction of its extension. Conversely, let 𝔟\mathfrak{b} be an 𝔖\mathfrak{S}-ideal of SS and let 𝔞\mathfrak{a} be its contraction to AA. Of course, 𝔞e𝔟\mathfrak{a}^{e}\subset\mathfrak{b}, where ()e(-)^{e} denotes extension. Suppose 𝔟\mathfrak{b} is strictly larger than 𝔞\mathfrak{a}, and let f𝔟𝔞ef\in\mathfrak{b}\setminus\mathfrak{a}^{e}. Clearing denominators, we can assume that ff belongs to RR. Furthermore, modifying ff by an element of 𝔞e\mathfrak{a}^{e}, we can assume that no coefficient of ff belongs to 𝔞\mathfrak{a}. Proposition 2.6 then implies that cΔnc\Delta_{n} belongs to 𝔟\mathfrak{b} for some nn where cc is some non-zero coefficient of ff. But Δn\Delta_{n} is a unit of SS, and so c𝔟A=𝔞c\in\mathfrak{b}\cap A=\mathfrak{a}, a contradiction. ∎

2.5. Finitary points

Let x=(w,(xi)i[])x=(w,(x_{i})_{i\in[\infty]}) be a KK-point of 𝔛\mathfrak{X}. We define the width of xx to be the cardinality of the subset {xii[]}\{x_{i}\mid i\in[\infty]\} of KK, and we say that xx is finitary if it has finite width. It is clear that width is invariant under field extension, and can thus be defined for points of 𝔛\mathfrak{X}. We let 𝔛n\mathfrak{X}_{\leq n} be the set of points of width n\leq n, and 𝔛fin\mathfrak{X}_{\mathrm{fin}} the set of points of finitary points. We say that a subset of 𝔛\mathfrak{X} is bounded if it is contained in 𝔛n\mathfrak{X}_{\leq n} for some nn. Let 𝔡n=Δn\mathfrak{d}_{n}=\langle\!\langle\Delta_{n}\rangle\!\rangle be the ideal of RR generated by the 𝔖\mathfrak{S}-orbit of Δn\Delta_{n}.

Proposition 2.8.

We have 𝔛n=V(𝔡n+1)\mathfrak{X}_{\leq n}=V(\mathfrak{d}_{n+1}). In particular, 𝔛n\mathfrak{X}_{\leq n} is Zariski closed.

Proof.

It is clear that a KK-point xx has width n\leq n if and only if Δn+1(σx)=0\Delta_{n+1}(\sigma x)=0 for all σ𝔖\sigma\in\mathfrak{S}, and so the result follows. ∎

Proposition 2.9.

Suppose that AA is noetherian. Then every Zariski closed 𝔖\mathfrak{S}-subset \mathfrak{Z} of 𝔛=𝔛W\mathfrak{X}=\mathfrak{X}_{W} has the form 12\mathfrak{Z}_{1}\cup\mathfrak{Z}_{2} where 1=𝔛W\mathfrak{Z}_{1}=\mathfrak{X}_{W^{\prime}} for some closed subset WW^{\prime} of WW and 2\mathfrak{Z}_{2} is a closed and bounded 𝔖\mathfrak{S}-subset of 𝔛W\mathfrak{X}_{W}.

Proof.

We proceed by noetherian induction on WW; thus we assume the result holds whenever WW is replaced by a proper closed subset. Let \mathfrak{Z} be given, and let IRI\subset R be its radical ideal which we may assume is not nilpotent. Then, by Proposition 2.6, II contains aΔna\Delta_{n} for some non-nilpotent aAa\in A and some nn. We thus have

=(V(a))(𝔛n).\mathfrak{Z}=(\mathfrak{Z}\cap V(a))\cup(\mathfrak{Z}\cap\mathfrak{X}_{\leq n}).

Now, V(a)V(a) is a proper closed subset of WW. Thus, by the inductive hypothesis, we have V(a)=12\mathfrak{Z}\cap V(a)=\mathfrak{Z}_{1}\cup\mathfrak{Z}_{2}^{\prime} where 1=𝔛W\mathfrak{Z}_{1}=\mathfrak{X}_{W^{\prime}} for some closed subset WW^{\prime} of WW, and 2\mathfrak{Z}_{2}^{\prime} is a closed and bounded 𝔖\mathfrak{S}-subset of 𝔛\mathfrak{X}. Of course, 2′′=𝔛Wn\mathfrak{Z}_{2}^{\prime\prime}=\mathfrak{Z}\cap\mathfrak{X}_{W}^{\leq n} is also a closed and bounded 𝔖\mathfrak{S}-subset of 𝔛\mathfrak{X}. We thus have =12\mathfrak{Z}=\mathfrak{Z}_{1}\cup\mathfrak{Z}_{2}, where 2=22′′\mathfrak{Z}_{2}=\mathfrak{Z}_{2}^{\prime}\cup\mathfrak{Z}_{2}^{\prime\prime}, as required. ∎

2.6. The type of a point

We now introduce an invariant that is finer than width. Recall that an \infty-partition is a sequence λ=(λ1,λ2,)\lambda=(\lambda_{1},\lambda_{2},\ldots) such that λ1=\lambda_{1}=\infty, λiλi+1\lambda_{i}\geq\lambda_{i+1}, and λi=0\lambda_{i}=0 for i0i\gg 0. Let x=(w,(xi)i[])x=(w,(x_{i})_{i\in[\infty]}) be a finitary KK-point of 𝔛\mathfrak{X} of width nn. Let a1,,ana_{1},\ldots,a_{n} be the nn values that the xix_{i} assume, and let 𝒰k[]\mathcal{U}_{k}\subset[\infty] be the set of indices ii such that xi=akx_{i}=a_{k}. Permuting the aa’s if necessary, we assume that #𝒰k#𝒰k+1\#\mathcal{U}_{k}\geq\#\mathcal{U}_{k+1}. Then putting λk=#𝒰k\lambda_{k}=\#\mathcal{U}_{k} for 1kn1\leq k\leq n, and λk=0\lambda_{k}=0 for k>nk>n, we see that λ\lambda is an \infty-partition. We define the type of xx to be λ\lambda. This quantity is clear invariant under field extension, and can thus be associated to fintary points of 𝔛\mathfrak{X}. It is also clearly invariant under the action of the big symmetric group.

3. The space 𝒳\mathcal{X}

3.1. Affine spaces

For a finite set SS, we let 𝐀WS\mathbf{A}^{S}_{W} be the affine space over the base scheme WW with coordinates indexed by the set SS. We let 𝐀W[S]\mathbf{A}^{[S]}_{W} be the open subscheme of 𝐀WS\mathbf{A}^{S}_{W} where the coordinates are all distinct. Given a function f:STf\colon S\to T of finite sets, we let αf:𝐀WT𝐀WS\alpha_{f}\colon\mathbf{A}_{W}^{T}\to\mathbf{A}_{W}^{S} be the map that takes (xj)jT(x_{j})_{j\in T} to (xf(i))iS(x_{f(i)})_{i\in S}. We note that if ff is an injection then αf\alpha_{f} is a projection, while if ff is a surjection then αf\alpha_{f} is a multi-diagonal map (and thus a closed immersion).

3.2. Generalized partitions and compositions

A generalized partition is a tuple λ=(λ1,λ2,)\lambda=(\lambda_{1},\lambda_{2},\ldots) where λi𝐍{}\lambda_{i}\in\mathbf{N}\cup\{\infty\} such that λiλi+1\lambda_{i}\geq\lambda_{i+1} for all ii and λi=0\lambda_{i}=0 for i0i\gg 0. We define length of λ\lambda, denoted (λ)\ell(\lambda), to be the largest ii such that λi>0\lambda_{i}>0. A generalized partition λ\lambda is an \infty-partition if λ1=\lambda_{1}=\infty, otherwise λ\lambda is called a finite partition. We denote the set of generalized partition by Λ~\widetilde{\Lambda}, and the set of partitions of infinity by Λ\Lambda.

A generalized composition is a pair (S,λ)(S,\lambda) consisting of a finite set SS and a function λ:S𝐙>0{}\lambda\colon S\to\mathbf{Z}_{>0}\cup\{\infty\}. We typically simply write λ\lambda for the generalized composition, and denote the index set SS by λ\langle\lambda\rangle. Given a generalized composition λ\lambda, we let |λ|=iλλi|\lambda|=\sum_{i\in\langle\lambda\rangle}\lambda_{i} and (λ)=#λ\ell(\lambda)=\#\langle\lambda\rangle. We say that λ\lambda is a composition of \infty or an \infty-composition if |λ|=|\lambda|=\infty, or equivalently, if λi=\lambda_{i}=\infty for some iλi\in\langle\lambda\rangle. We regard any generalized partition λ\lambda as a generalized composition on the index set [(λ)][\ell(\lambda)].

Let λ\lambda and μ\mu be generalized compositions. A map f:λμf\colon\lambda\to\mu is a function f:λμf\colon\langle\lambda\rangle\to\langle\mu\rangle such that f(λ)μf_{*}(\lambda)\leq\mu, i.e., for all jμj\in\langle\mu\rangle we have if1(j)λiμj\sum_{i\in f^{-1}(j)}\lambda_{i}\leq\mu_{j}. We say that ff is a principal surjection if f(λ)=μf_{*}(\lambda)=\mu, which implies that ff is a surjection. We write f:λμf\colon\lambda\rightarrowtail\mu to indicate that ff is a principal surjection. Every map ff can be factored as ghg\circ h where hh is a principal surjection and gg is an injection. The composition of two maps is again a map, and so we have a category of generalized compositions.

Let λ\lambda and μ\mu be generalized partitions. We define λμ\lambda\leq\mu if λ\lambda can be obtained from μ\mu by decreasing (or removing) parts, and we define λμ\lambda\preceq\mu if λ\lambda can be obtained from μ\mu by combining and decreasing (or removing) parts. We note here that if f:λμf\colon\lambda\to\mu is an injection then λμ\lambda\leq\mu and if ff is a principal surjection then μλ\mu\preceq\lambda. A poset is called a well-quasi-order (wqo) if it satisfies the descending chain condition and has no infinite antichains. We refer to [SS, §2] for general facts about wqo’s, though note there the term “noetherian” is used in place of “well-quasi-order.”

Proposition 3.1.

The posets (Λ~,)(\widetilde{\Lambda},\leq) and (Λ~,)(\widetilde{\Lambda},\preceq) are wqo’s.

Proof.

Since 𝐍{}\mathbf{N}\cup\{\infty\} is a wqo, we conclude that the poset of sequences on alphabet 𝐍{}\mathbf{N}\cup\{\infty\} is also a wqo (Higman’s lemma, or see [SS, Theorem 2.5]). Thus (Λ~,)(\widetilde{\Lambda},\leq), being a subposet of this poset of sequences, is also a wqo. Since λμλμ\lambda\leq\mu\implies\lambda\preceq\mu, the poset (Λ~,)(\widetilde{\Lambda},\preceq) does not have infinite antichains. Now suppose λ1λ2\lambda_{1}\succeq\lambda_{2}\succeq\cdots is a decreasing chain. Then (λi+1)(λi)\ell(\lambda_{i+1})\leq\ell(\lambda_{i}) for each ii. So we may as well assume that (λi)\ell(\lambda_{i}) is independent of ii. But then λiλi+1\lambda_{i}\geq\lambda_{i+1} for all ii. Since (Λ~,)(\widetilde{\Lambda},\leq) is a wqo, this chain stabilizes. This completes the proof. ∎

3.3. The space 𝒳\mathcal{X} and its subspaces

For an \infty-composition λ\lambda, we put 𝒳W,λ=𝐀Wλ\mathcal{X}_{W,\lambda}=\mathbf{A}^{\langle\lambda\rangle}_{W} and 𝒳W,[λ]=𝐀W[λ]\mathcal{X}_{W,[\lambda]}=\mathbf{A}^{[\langle\lambda\rangle]}_{W}. We regard 𝒳W={𝒳W,λ}λ\mathcal{X}_{W}=\{\mathcal{X}_{W,\lambda}\}_{\lambda} as a single object. We omit the WW from the notation when possible.

By a “subset” of 𝒳\mathcal{X} we mean a rule 𝒵\mathcal{Z} assigning to every \infty-composition λ\lambda a subset 𝒵λ\mathcal{Z}_{\lambda} of 𝒳λ\mathcal{X}_{\lambda}; here we simply regard 𝒳λ\mathcal{X}_{\lambda} as set of points in its underlying topological space (including its non-closed points). We consider several conditions on such a subset 𝒵\mathcal{Z}:

  • We say that 𝒵\mathcal{Z} is a C1 subset if for every principal surjection f:λμf\colon\lambda\rightarrowtail\mu of \infty-compositions we have αf1(𝒵λ)=𝒵μ\alpha_{f}^{-1}(\mathcal{Z}_{\lambda})=\mathcal{Z}_{\mu}.

  • We say that a point x𝒳λx\in\mathcal{X}_{\lambda} is NN-approximable by 𝒵\mathcal{Z}, for a positive integer NN, if there exists an \infty-composition μ\mu and an injective function f:λμf\colon\langle\lambda\rangle\to\langle\mu\rangle such that μf(i)min(λi,N)\mu_{f(i)}\geq\min(\lambda_{i},N) for all iλi\in\langle\lambda\rangle and xx belongs to the image of 𝒵μ\mathcal{Z}_{\mu} under the projection map αf:𝒳μ𝒳λ\alpha_{f}\colon\mathcal{X}_{\mu}\to\mathcal{X}_{\lambda}. (Note: ff need not define a map of \infty-compositions λμ\lambda\to\mu.) We say that xx is approximable by 𝒵\mathcal{Z} if it is NN-approximable for all NN. We say that 𝒵\mathcal{Z} is a C2 subset if it is C1 and contains all of its approximable points (i.e., whenever x𝒳λx\in\mathcal{X}_{\lambda} is approximable by 𝒵\mathcal{Z} we have x𝒵λx\in\mathcal{Z}_{\lambda}).

  • We say that 𝒵\mathcal{Z} is a C3 subvariety if it is C2 and 𝒵λ\mathcal{Z}_{\lambda} is Zariski closed in 𝒳λ\mathcal{X}_{\lambda} for all λ\lambda.

We note that if 𝒵\mathcal{Z} is a C1 subset then for any isomorphism f:λμf\colon\lambda\to\mu of \infty-compositions we have αf(𝒵μ)=𝒵λ\alpha_{f}(\mathcal{Z}_{\mu})=\mathcal{Z}_{\lambda}, since an isomorphism is a principal surjection. Thus 𝒵\mathcal{Z} is determined by 𝒵λ\mathcal{Z}_{\lambda} with λ\lambda an \infty-partition. Ultimately, we are most interested in C3 subvarieties, but C1 and C2 subsets will be helpful intermediate objects.

Remark 3.2.

Let us try to motivate the above conditions. Suppose that \mathfrak{Z} is a Zariski closed 𝔖\mathfrak{S}-stable subset of 𝔛\mathfrak{X}, and let 𝒵=Φ()\mathcal{Z}=\Phi(\mathfrak{Z}) as defined in §1.1. Then 𝒵()𝐀1\mathcal{Z}_{(\infty)}\subset\mathbf{A}^{1} consists of all points aa such that (a,a,a,)(a,a,a,\ldots)\in\mathfrak{Z}, while 𝒵(,1)𝐀2\mathcal{Z}_{(\infty,1)}\subset\mathbf{A}^{2} consists of all points (a,b)(a,b) such that (b,a,a,a,)(b,a,a,a,\ldots)\in\mathfrak{Z}.

It is clear that 𝒵()\mathcal{Z}_{(\infty)} is simply the pullback of 𝒵(,1)\mathcal{Z}_{(\infty,1)} along the diagonal map 𝐀1𝐀2\mathbf{A}^{1}\to\mathbf{A}^{2}. This is a special case of the C1 condition. The general case is not much different (see Proposition 7.1), and so 𝒵\mathcal{Z} is a C1 subset.

Suppose now that (a,b)𝒵(,1)(a,b)\in\mathcal{Z}_{(\infty,1)}. Then a𝒳()a\in\mathcal{X}_{(\infty)} is approximable by 𝒵\mathcal{Z}: indeed, if f:()(,1)f\colon(\infty)\to(\infty,1) denotes the natural inclusion then a=αf(b,a)a=\alpha_{f}(b,a). We have already seen (in §1.1) that aa belongs to 𝒵()\mathcal{Z}_{(\infty)} in this case. This verifies the C2 condition in a special case. Again, the general case is not much more difficult (see Proposition 7.4), and so 𝒵\mathcal{Z} is a C2 subset.

By definition, 𝒵λ\mathcal{Z}_{\lambda} is in the inverse image of \mathfrak{Z} under the multi-diagonal map 𝒳λ𝔛\mathcal{X}_{\lambda}\to\mathfrak{X} corresponding to some parition 𝒰\mathcal{U} of [][\infty] with #𝒰i=λi\#\mathcal{U}_{i}=\lambda_{i}. Since the multi-diagonal map is continuous and \mathfrak{Z} is Zariski closed, it follows that 𝒵λ\mathcal{Z}_{\lambda} is Zariski closed in 𝒳λ\mathcal{X}_{\lambda}. Thus 𝒵\mathcal{Z} is a C3 subvariety. ∎

3.4. First properties of subspaces of 𝒳\mathcal{X}

We now establish some simple properties of C1, C2, and C3 subsets of 𝒳\mathcal{X}. For a subset 𝒵\mathcal{Z} of 𝒳\mathcal{X}, we let 𝒵[λ]=𝒵λ𝒳[λ]\mathcal{Z}_{[\lambda]}=\mathcal{Z}_{\lambda}\cap\mathcal{X}_{[\lambda]}.

Proposition 3.3.

Let 𝒵\mathcal{Z} and 𝒵\mathcal{Z}^{\prime} be two C1 subsets of 𝒳\mathcal{X}. Then 𝒵𝒵\mathcal{Z}\subset\mathcal{Z}^{\prime} if and only if 𝒵[λ]𝒵[λ]\mathcal{Z}_{[\lambda]}\subset\mathcal{Z}^{\prime}_{[\lambda]} for all \infty-compositions λ\lambda.

Proof.

If 𝒵𝒵\mathcal{Z}\subset\mathcal{Z}^{\prime} then obviously 𝒵[λ]𝒵[λ]\mathcal{Z}_{[\lambda]}\subset\mathcal{Z}^{\prime}_{[\lambda]} for all λ\lambda. Thus suppose 𝒵[λ]𝒵[λ]\mathcal{Z}_{[\lambda]}\subset\mathcal{Z}^{\prime}_{[\lambda]} for all λ\lambda. Let λ\lambda be an \infty-composition and let x𝒵λx\in\mathcal{Z}_{\lambda}. Define an equivalence relation \sim on λ\langle\lambda\rangle by iji\sim j if xi=xjx_{i}=x_{j}. Let μ=λ/\langle\mu\rangle=\langle\lambda\rangle/\sim, let f:λμf\colon\langle\lambda\rangle\to\langle\mu\rangle be the quotient map, and let μ=f(λ)\mu=f_{*}(\lambda); thus f:λμf\colon\lambda\rightarrowtail\mu is a principal surjection. Then x=αf(y)x=\alpha_{f}(y) for some y𝒳[μ]y\in\mathcal{X}_{[\mu]}. Since yαf1(𝒵λ)y\in\alpha_{f}^{-1}(\mathcal{Z}_{\lambda}) and 𝒵\mathcal{Z} is C1, we have y𝒵[μ]y\in\mathcal{Z}_{[\mu]}. Thus y𝒵[μ]y\in\mathcal{Z}^{\prime}_{[\mu]} by our hypothesis. Therefore xαf(𝒵μ)𝒵λx\in\alpha_{f}(\mathcal{Z}^{\prime}_{\mu})\subset\mathcal{Z}^{\prime}_{\lambda} since 𝒵\mathcal{Z}^{\prime} is C1, and so 𝒵𝒵\mathcal{Z}\subset\mathcal{Z}^{\prime}. ∎

Proposition 3.4.

Let 𝒵\mathcal{Z} be a C2 subset of 𝒳\mathcal{X} and let f:λμf\colon\lambda\to\mu be a map of \infty-compositions. Then αf(𝒵μ)𝒵λ\alpha_{f}(\mathcal{Z}_{\mu})\subset\mathcal{Z}_{\lambda}.

Proof.

First suppose that ff is an injective map, and let x𝒵μx\in\mathcal{Z}_{\mu}. Then αf(x)𝒳λ\alpha_{f}(x)\in\mathcal{X}_{\lambda} is approximable by 𝒵\mathcal{Z}: indeed, this is witnessed by the point x𝒵μx\in\mathcal{Z}_{\mu}. Since 𝒵\mathcal{Z} is C2, it follows that αf(x)𝒵λ\alpha_{f}(x)\in\mathcal{Z}_{\lambda}. If ff is a principal surjection, then αf(𝒵μ)𝒵λ\alpha_{f}(\mathcal{Z}_{\mu})\subset\mathcal{Z}_{\lambda} by the C1 property. The general case now follows. ∎

Proposition 3.5.

Let 𝒵\mathcal{Z} be a C2 subset of 𝒳\mathcal{X}, and let λ\lambda and μ\mu be \infty-compositions with λ=μ\langle\lambda\rangle=\langle\mu\rangle and λiμi\lambda_{i}\leq\mu_{i} for all iλi\in\langle\lambda\rangle. Then 𝒵μ𝒵λ\mathcal{Z}_{\mu}\subset\mathcal{Z}_{\lambda}.

Proof.

The identity map f:λμf\colon\lambda\to\mu is a map of \infty-compositions, and so 𝒵μ=αf(𝒵μ)𝒵λ\mathcal{Z}_{\mu}=\alpha_{f}(\mathcal{Z}_{\mu})\subset\mathcal{Z}_{\lambda} by Proposition 3.4. ∎

Proposition 3.6.

Let 𝒵\mathcal{Z} be a C2 subset of 𝒳\mathcal{X}, and let λ1λ2\lambda^{1}\leq\lambda^{2}\leq\ldots be \infty-partitions, all of length rr, with limit λ\lambda. Then 𝒵λ=i1𝒵λi\mathcal{Z}_{\lambda}=\bigcap_{i\geq 1}\mathcal{Z}_{\lambda^{i}}.

Proof.

By the previous proposition, we have 𝒵λi1𝒵λi\mathcal{Z}_{\lambda}\subset\bigcap_{i\geq 1}\mathcal{Z}_{\lambda^{i}}. Conversely, let x𝒵λix\in\mathcal{Z}_{\lambda^{i}} for each ii. Then x𝒳λx\in\mathcal{X}_{\lambda}, and for each NN we may find an nn large enough such that λinmin(λi,N)\lambda^{n}_{i}\geq\min(\lambda_{i},N) for all i[r]i\in[r]. Thus xx is NN-approximable for each NN, and so x𝒵λx\in\mathcal{Z}_{\lambda}. ∎

4. Constructing subvarieties of 𝒳\mathcal{X}

4.1. Correspondences

Let λ\lambda and μ\mu be generalized compositions. A correspondence f:λμf\colon\lambda\dashrightarrow\mu is a pair (f1,f2)(f_{1},f_{2}) where f1:ρλf_{1}\colon\rho\rightarrowtail\lambda is a principal surjection and f2:ρμf_{2}\colon\rho\to\mu is an arbitrary map; here ρ\rho denotes another generalized composition. Suppose that λ\lambda and μ\mu are \infty-compositions; note that ρ\rho is then as well. For S𝒳μS\subset\mathcal{X}_{\mu}, we let αf(S)=αf11(αf2(S))\alpha_{f}(S)=\alpha_{f_{1}}^{-1}(\alpha_{f_{2}}(S)), a subset of 𝒳λ\mathcal{X}_{\lambda}. Note that xαf(S)x\in\alpha_{f}(S) if and only if αf1(x)αf2(S)\alpha_{f_{1}}(x)\in\alpha_{f_{2}}(S). We say that a subset 𝒵\mathcal{Z} of 𝒳\mathcal{X} is closed under correspondences if αf(𝒵μ)𝒵λ\alpha_{f}(\mathcal{Z}_{\mu})\subset\mathcal{Z}_{\lambda} for all correspondences f:λμf\colon\lambda\dashrightarrow\mu.

Proposition 4.1.

We have the following implications for subsets of 𝒳\mathcal{X}:

C2closed under correspondencesC1.\text{C2}\implies\text{closed under correspondences}\implies\text{C1}.
Proof.

Suppose 𝒵\mathcal{Z} is C2. Let f:λμf\colon\lambda\dashrightarrow\mu be a correspondence of \infty-compositions, given by f=(f1:ρλ,f2:ρμ)f=(f_{1}\colon\rho\to\lambda,f_{2}\colon\rho\to\mu). Then αf(𝒵μ)=αf11(αf2(𝒵μ))\alpha_{f}(\mathcal{Z}_{\mu})=\alpha_{f_{1}}^{-1}(\alpha_{f_{2}}(\mathcal{Z}_{\mu})). By Proposition 3.4, we have αf2(𝒵μ)𝒵ρ\alpha_{f_{2}}(\mathcal{Z}_{\mu})\subset\mathcal{Z}_{\rho}, and by the C1 property we have αf11(𝒵ρ)𝒵λ\alpha_{f_{1}}^{-1}(\mathcal{Z}_{\rho})\subset\mathcal{Z}_{\lambda}. Thus 𝒵\mathcal{Z} is closed under correspondences.

Now suppose that 𝒵\mathcal{Z} is closed under correspondences, and let f:λμf\colon\lambda\rightarrowtail\mu be a principal surjection. Then αf(𝒵μ)𝒵λ\alpha_{f}(\mathcal{Z}_{\mu})\subset\mathcal{Z}_{\lambda} by assumption. Let g:μλg\colon\mu\dashrightarrow\lambda be the correspondence (f,idλ)(f,\mathrm{id}_{\lambda}). Again αf1(𝒵λ)=αg(𝒵λ)𝒵μ\alpha_{f}^{-1}(\mathcal{Z}_{\lambda})=\alpha_{g}(\mathcal{Z}_{\lambda})\subset\mathcal{Z}_{\mu} by assumption. Thus αf1(𝒵λ)=𝒵μ\alpha_{f}^{-1}(\mathcal{Z}_{\lambda})=\mathcal{Z}{\mu}, and so 𝒵\mathcal{Z} is C1. ∎

4.2. Composing correspondences

In general, composition of correspondences is carried out using fiber products. Unfortunately, the category of \infty-compositions does not have fiber products (see Example 4.4). However, the following proposition provides us with an approximate fiber product that will be good enough.

Proposition 4.2.

For i=1,2i=1,2, let fi:μiμf_{i}\colon\mu^{i}\to\mu be a map of \infty-weightings. Then there exists an \infty-composition μ~\widetilde{\mu} and maps gi:μ~μig_{i}\colon\widetilde{\mu}\to\mu^{i} of \infty-compositions such that the following diagram commutes:

μ~\textstyle{\widetilde{\mu}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}g2\scriptstyle{g_{2}}g1\scriptstyle{g_{1}}μ2\textstyle{\mu^{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f2\scriptstyle{f_{2}}μ1\textstyle{\mu^{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f1\scriptstyle{f_{1}}μ.\textstyle{\mu.}

Moreover, the following holds:

  1. (a)

    If f2f_{2} is a principal surjection then so is g1g_{1}.

  2. (b)

    If f2f_{2} is an injection then so is g1g_{1}.

Proof.

Note that we can prove the result fiberwise, in other words, it is enough to deal with the case when μ\mu is a singleton (but possibly with finite nonzero weight). So let μ=(μ1)\mu=(\mu_{1}). Suppose μi=(μ1i,μ2i,,μrii)\mu^{i}=(\mu^{i}_{1},\mu^{i}_{2},\ldots,\mu^{i}_{r_{i}}) where μji\mu^{i}_{j} is a non-increasing sequence with values in 𝐍{}\mathbf{N}\cup\{\infty\}. We now construct μ~\widetilde{\mu}, g1g_{1} and g2g_{2} by induction on the lengths r1,r2r_{1},r_{2} of the \infty-compositions μ1,μ2\mu^{1},\mu^{2}. Set μ~1=min{μ11,μ12}\widetilde{\mu}_{1}=\min\{\mu^{1}_{1},\mu^{2}_{1}\}. The base case is when μ~1=μ1i\widetilde{\mu}_{1}=\mu^{i}_{1} for some i{1,2}i\in\{1,2\} and μi\mu^{i} has only one part. In this case, we take μ~=μi\widetilde{\mu}=\mu^{i}. In general, by induction, there exists a diagram

λ~\textstyle{\widetilde{\lambda}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}g2\scriptstyle{g^{\prime}_{2}}g1\scriptstyle{g^{\prime}_{1}}λ2\textstyle{\lambda^{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f2\scriptstyle{f^{\prime}_{2}}λ1\textstyle{\lambda^{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f1\scriptstyle{f^{\prime}_{1}}λ,\textstyle{\lambda,}

where λ=(μ1μ~1)\lambda=(\mu_{1}-\widetilde{\mu}_{1}), λi=(μ1iμ~1,μ2i,,μrii)\lambda^{i}=(\mu^{i}_{1}-\widetilde{\mu}_{1},\mu^{i}_{2},\ldots,\mu^{i}_{r_{i}}) (where the elements may need to be rearranged to make them non-increasing), and f1,f2f_{1}^{\prime},f_{2}^{\prime} are obtained from restricting f1,f2f_{1},f_{2}. Now set μ~=(μ~1,λ~)\widetilde{\mu}=(\widetilde{\mu}_{1},\widetilde{\lambda}). Then gig_{i}^{\prime} can be extended to gi:μ~μig_{i}\colon\widetilde{\mu}\to\mu^{i} by sending μ~1\widetilde{\mu}_{1} to μ1i\mu^{i}_{1}. It is clear that f2g2=f1g1f_{2}g_{2}=f_{1}g_{1}, which finishes the construction of μ~\widetilde{\mu}, g1g_{1} and g2g_{2}.

For (a), assume that f2f_{2} is a principal surjection. In the base case, we must have μ~=μ1\widetilde{\mu}=\mu^{1} and g1=idg_{1}=\mathrm{id}, so the result follows. In the inductive case, we see that f2f_{2}^{\prime} is a principal surjection. By induction g1g_{1}^{\prime} is a principal surjection. It follows that g2g_{2} is a principal surjection.

For (b), assume that f2f_{2} is an injection. In the base case, either μ~=μ1\widetilde{\mu}=\mu^{1} and μ1\mu^{1} has only one part, or μ~=μ2\widetilde{\mu}=\mu^{2} and μ11μ12\mu^{1}_{1}\geq\mu^{2}_{1}. In both cases g1g_{1} is an injection. In the inductive case, note that f2f_{2}^{\prime} is an injection. By induction, g1g_{1}^{\prime} is an injection. Moreover, in this case we must have μ~1=μ11\widetilde{\mu}_{1}=\mu^{1}_{1} (as we are not in the base case), and so λ1\lambda^{1} has r11r_{1}-1 parts. It follows that g1g_{1} is an injection. This finishes the proof. ∎

Let f:λμf\colon\lambda\dashrightarrow\mu and g:μνg\colon\mu\dashrightarrow\nu be correspondences, given by data (f1:ρλ,f2:ρμ)(f_{1}\colon\rho\rightarrowtail\lambda,f_{2}\colon\rho\to\mu) and (g1:σμ,g2:σν)(g_{1}\colon\sigma\rightarrowtail\mu,g_{2}\colon\sigma\to\nu). We say that a correspondence h:λνh\colon\lambda\dashrightarrow\nu, given by data (h1:τλ,τν)(h_{1}\colon\tau\to\lambda,\tau\to\nu) is a composition of ff and gg if there exists a commutative diagram

τ\textstyle{\tau\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}h1\scriptstyle{h_{1}^{\prime}}h2\scriptstyle{h_{2}^{\prime}}ρ\textstyle{\rho\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f1\scriptstyle{f_{1}}f2\scriptstyle{f_{2}}σ\textstyle{\sigma\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}g1\scriptstyle{g_{1}}g2\scriptstyle{g_{2}}λ\textstyle{\lambda}μ\textstyle{\mu}ν\textstyle{\nu}

such that h1h_{1}^{\prime} is a principal surjection, h1=f1h1h_{1}=f_{1}\circ h_{1}^{\prime}, and h2=h2h2h_{2}=h_{2}\circ h_{2}^{\prime}. We note that a composition need not be unique.

Proposition 4.3.

Let f:λμf\colon\lambda\dashrightarrow\mu and g:μνg\colon\mu\dashrightarrow\nu be correspondences. Then a composition exists. Moreover, for any composition hh and any S𝒳νS\subset\mathcal{X}_{\nu}, we have αf(αg(S))αh(S)\alpha_{f}(\alpha_{g}(S))\subset\alpha_{h}(S).

Proof.

Existence follows from Proposition 4.2. Now suppose that hh is any composition of ff and gg. Use notation as in the previous paragraph. Suppose xαf(αg(S))x\in\alpha_{f}(\alpha_{g}(S)). By definition, this means αf1(x)=αf2(y)\alpha_{f_{1}}(x)=\alpha_{f_{2}}(y) for some yαg(S)y\in\alpha_{g}(S); again, by definition, we have αg1(y)αg2(S)\alpha_{g_{1}}(y)\in\alpha_{g_{2}}(S). We thus see that

αh1(x)=αh1(αf1(x))=αh1(αf2(y))=αh2(αg1(y))αh2(αg2(S))=αh2(S),\alpha_{h_{1}}(x)=\alpha_{h_{1}^{\prime}}(\alpha_{f_{1}}(x))=\alpha_{h_{1}^{\prime}}(\alpha_{f_{2}}(y))=\alpha_{h_{2}^{\prime}}(\alpha_{g_{1}}(y))\in\alpha_{h_{2}^{\prime}}(\alpha_{g_{2}}(S))=\alpha_{h_{2}}(S),

and so xαh(S)x\in\alpha_{h}(S). ∎

Example 4.4.

Fiber products do not exist in the category of \infty-compositions, in general. For example, let μ=(,8),μ~1=(,6,2),μ~2=(,4,4)\mu=(\infty,8),\widetilde{\mu}^{1}=(\infty,6,2),\widetilde{\mu}^{2}=(\infty,4,4), and let πi:μ~iμ\pi_{i}\colon\widetilde{\mu}^{i}\rightarrowtail\mu be the unique principal surjections. Then the natural candidate for μ~1×μμ~2\widetilde{\mu}^{1}\times_{\mu}\widetilde{\mu}^{2} is (,4,2,1,1)(\infty,4,2,1,1). But there are two maps (,4,2,1,1)μ~1(\infty,4,2,1,1)\to\widetilde{\mu}^{1} – one that combines 44 and 22 together, and the second one that combines 44 and the two 11s together. These two maps do not factor through each other and so the fiber product μ~1×μμ~2\widetilde{\mu}^{1}\times_{\mu}\widetilde{\mu}^{2} doesn’t exist. ∎

4.3. Construction of C2 subsets

Let λ\lambda be a \infty-composition and let ZZ be a subset of 𝒳λ\mathcal{X}_{\lambda}. Define a subset Γλ(Z)\Gamma_{\lambda}^{\circ}(Z) of 𝒳\mathcal{X} by

Γλ(Z)μ=f:μλαf(Z),\Gamma^{\circ}_{\lambda}(Z)_{\mu}=\bigcup_{f\colon\mu\dashrightarrow\lambda}\alpha_{f}(Z),

where the union is over all correspondences ff.

Theorem 4.5.

Let ZZ be a subset of 𝒳λ\mathcal{X}_{\lambda}. Then 𝒵=Γλ(Z)\mathcal{Z}=\Gamma^{\circ}_{\lambda}(Z) is a C2 subset of 𝒳\mathcal{X}. Moreover, if 𝒵\mathcal{Z}^{\prime} is any C2 subset of 𝒳\mathcal{X} such that 𝒵λ\mathcal{Z}^{\prime}_{\lambda} contains ZZ then 𝒵\mathcal{Z}^{\prime} contains 𝒵\mathcal{Z}.

Proof.

We first claim that 𝒵\mathcal{Z} is closed under the action of correspondences. Indeed, suppose x𝒵νx\in\mathcal{Z}_{\nu} and f:μνf\colon\mu\dashrightarrow\nu is a correspondence. Then, by definition of 𝒵ν\mathcal{Z}_{\nu}, there exists a correspondence g:νλg\colon\nu\dashrightarrow\lambda such that xαg(Z)x\in\alpha_{g}(Z). Let h:μλh\colon\mu\dashrightarrow\lambda be a composition of ff and gg. Then

αf(x)αf(αg(Z))αh(Z)𝒵μ,\alpha_{f}(x)\in\alpha_{f}(\alpha_{g}(Z))\subset\alpha_{h}(Z)\subset\mathcal{Z}_{\mu},

where we have used Proposition 4.3. This proves the claim. In particular, 𝒵\mathcal{Z} is C1 by Proposition 4.1.

We now claim that 𝒵\mathcal{Z} is a C2 subset of 𝒳\mathcal{X}. Without loss of generality, suppose λ=[n]\langle\lambda\rangle=[n] and that λ1,,λr\lambda_{1},\ldots,\lambda_{r} are infinite and λr+1,,λn\lambda_{r+1},\ldots,\lambda_{n} are finite. Suppose x𝒳μx\in\mathcal{X}_{\mu} is approximable by 𝒵\mathcal{Z}. Let NN be larger than λr+1++λn\lambda_{r+1}+\cdots+\lambda_{n} and all finite parts of μ\mu. Since xx is NN-approximable by 𝒵\mathcal{Z}, we can find an \infty-composition ν\nu with μν\langle\mu\rangle\subset\langle\nu\rangle and νimin(N,μi)\nu_{i}\geq\min(N,\mu_{i}) for all iμi\in\langle\mu\rangle such that xx is the image of some point y𝒵νy\in\mathcal{Z}_{\nu} under the projection map 𝒳ν𝒳μ\mathcal{X}_{\nu}\to\mathcal{X}_{\mu}. Now, since y𝒵νy\in\mathcal{Z}_{\nu}, there is some correspondence f:νλf\colon\nu\dashrightarrow\lambda such that yαf(Z)y\in\alpha_{f}(Z). Let f=(f1:ρν,f2:ρλ)f=(f_{1}\colon\rho\rightarrowtail\nu,f_{2}\colon\rho\to\lambda) where f1f_{1} is a principal surjection.

Suppose νiN\nu_{i}\geq N. We claim that there is some jf11(i)j\in f_{1}^{-1}(i) such that f2(j)rf_{2}(j)\leq r. Indeed, suppose this were not the case. Then f2f_{2} would map f11(i)f_{1}^{-1}(i) into {r+1,,n}\{r+1,\ldots,n\}, and so f11(i)f_{1}^{-1}(i) would be contained in k=r+1nf21(k)\bigcup_{k=r+1}^{n}f_{2}^{-1}(k). But the total weight of the set f21(k)f_{2}^{-1}(k) is at most λk\lambda_{k}, and so the total weight of the set k=r+1nf21(k)\bigcup_{k=r+1}^{n}f_{2}^{-1}(k) is λr+1++λn<N\lambda_{r+1}+\cdots+\lambda_{n}<N. However, the total weight of the set f11(i)f_{1}^{-1}(i) is νi\nu_{i}, which is N\geq N, a contradiction.

For each iνi\in\langle\nu\rangle with νiN\nu_{i}\geq N pick a(i)f11(i)a(i)\in f_{1}^{-1}(i) such that f2(a(i))rf_{2}(a(i))\leq r. Let ν¯\overline{\nu} be the \infty-composition obtained from ν\nu by changing any entry that is N\geq N to to \infty; similarly, let ρ¯\overline{\rho} be the \infty-composition obtained from ρ\rho by changing the entries at a(1),,a(r)a(1),\ldots,a(r) to \infty. Then the function f1:ρνf_{1}\colon\langle\rho\rangle\to\langle\nu\rangle defines a principal surjection g1:ρ¯ν¯g_{1}\colon\overline{\rho}\rightarrowtail\overline{\nu} and the function f2:ρ[n]f_{2}\colon\langle\rho\rangle\to[n] defines a map of \infty-compositions g2:ρ¯λg_{2}\colon\overline{\rho}\to\lambda. Let g:ν¯λg\colon\overline{\nu}\dashrightarrow\lambda be the correspondence (g1,g2)(g_{1},g_{2}). Then yαg(Z)y\in\alpha_{g}(Z), and so y𝒵ν¯y\in\mathcal{Z}_{\overline{\nu}}. (Note that αg1\alpha_{g_{1}} and αg2\alpha_{g_{2}} are the same maps between affines spaces as αf1\alpha_{f_{1}} and αf2\alpha_{f_{2}}.)

The standard inclusion μν\langle\mu\rangle\to\langle\nu\rangle defines a map of \infty-compositions μν¯\mu\to\overline{\nu}. Thus by the first paragraph, the projection map 𝒳ν¯𝒳μ\mathcal{X}_{\overline{\nu}}\to\mathcal{X}_{\mu} maps 𝒵ν¯\mathcal{Z}_{\overline{\nu}} into 𝒵λ\mathcal{Z}_{\lambda}. Since the image of y𝒵ν¯y\in\mathcal{Z}_{\overline{\nu}} under this map is xx, we see that x𝒵μx\in\mathcal{Z}_{\mu}. Thus 𝒵\mathcal{Z} is C2, as claimed.

Finally, suppose 𝒵\mathcal{Z}^{\prime} is a second C2 subset of 𝒳\mathcal{X} such that 𝒵λ\mathcal{Z}^{\prime}_{\lambda} contains ZZ. Since 𝒵\mathcal{Z}^{\prime} is closed under the action of correspondences (Proposition 4.1), we thus have 𝒵𝒵\mathcal{Z}\subset\mathcal{Z}^{\prime}. ∎

We need two additional properties of the Γλ\Gamma_{\lambda}^{\circ} construction. We say that a correspondence f:μλf\colon\mu\dashrightarrow\lambda given by f=(f1:ρμ,f2:ρλ)f=(f_{1}\colon\rho\rightarrowtail\mu,f_{2}\colon\rho\to\lambda) is good if for each iμi\in\langle\mu\rangle the size of f11(i)f_{1}^{-1}(i) is bounded by the number of parts in λ\lambda, and if μi\mu_{i} is larger than the sum of finite parts of λ\lambda then the fiber f11(i)f_{1}^{-1}(i) is a singleton. We note that the number of good correspondences with fixed domain and target is finite.

Proposition 4.6.

Let ZZ be a subset of 𝒳λ\mathcal{X}_{\lambda}. Then

Γλ(Z)μ=f:μλf is goodαf(Z).\Gamma_{\lambda}^{\circ}(Z)_{\mu}=\bigcup_{\begin{subarray}{c}f\colon\mu\dashrightarrow\lambda\\ \text{$f$ is good}\end{subarray}}\alpha_{f}(Z).
Proof.

Let f:μλf\colon\mu\dashrightarrow\lambda be an arbitrary correspondence, given by f=(f1:ρμ,f2:ρλ)f=(f_{1}\colon\rho\rightarrowtail\mu,f_{2}\colon\rho\to\lambda) with f1f_{1} a principal surjection. Define an equivalence relation \sim on ρ\langle\rho\rangle by iji\sim j if f1(i)=f1(j)f_{1}(i)=f_{1}(j) and f2(i)=f2(j)f_{2}(i)=f_{2}(j). Let τ=ρ/\langle\tau\rangle=\langle\rho\rangle/\sim, let π:ρτ\pi\colon\langle\rho\rangle\to\langle\tau\rangle be the quotient map, and let τ=π(ρ)\tau=\pi_{*}(\rho), so that π:ρτ\pi\colon\rho\rightarrowtail\tau is a principal surjection. Write f1=g1πf_{1}=g_{1}\circ\pi and f2=g2πf_{2}=g_{2}\circ\pi. Then we have a correspondence g:μλg\colon\mu\dashrightarrow\lambda given by (g1:τμ,g2:τλ)(g_{1}\colon\tau\rightarrowtail\mu,g_{2}\colon\tau\to\lambda). One easily sees that αf(Z)=αg(Z)\alpha_{f}(Z)=\alpha_{g}(Z). Note that #g11(i)(λ)\#g_{1}^{-1}(i)\leq\ell(\lambda) for all i[(μ)]i\in[\ell(\mu)].

Let ee be the sum of finite parts of λ\lambda. For each iμi\in\langle\mu\rangle with μi>e\mu_{i}>e, choose a(i)g11(i)a(i)\in g_{1}^{-1}(i) such that λg2(a(i))=\lambda_{g_{2}(a(i))}=\infty, which must exist. Let σ\sigma be the composition on the underlying set σ{a(i)μ:μi>e}{iμ:μie}\langle\sigma\rangle\coloneq\{a(i)\in\langle\mu\rangle\colon\mu_{i}>e\}\sqcup\{i\in\langle\mu\rangle\colon\mu_{i}\leq e\} where σa(i)=jg11(i)τj\sigma_{a(i)}=\sum_{j\in g_{1}^{-1}(i)}\tau_{j} for each ii with μi>e\mu_{i}>e. Let h1:σμh_{1}\colon\sigma\rightarrowtail\mu and h2:σρh_{2}\colon\sigma\to\rho be the induced maps. Then we have a correspondence h:μλh\colon\mu\dashrightarrow\lambda given by (h1,h2)(h_{1},h_{2}), which is good. Furthermore, it is clear that αg(Z)αh(Z)\alpha_{g}(Z)\subset\alpha_{h}(Z).

We have thus shown that for every correspondence f:μλf\colon\mu\dashrightarrow\lambda there is a good correspondence h:μλh\colon\mu\dashrightarrow\lambda such that αf(Z)αh(Z)\alpha_{f}(Z)\subset\alpha_{h}(Z). The result follows. ∎

Proposition 4.7.

Let Z𝒳λZ\subset\mathcal{X}_{\lambda} and 𝒵=Γλ(Z)\mathcal{Z}=\Gamma^{\circ}_{\lambda}(Z). Then

𝒵[λ]=fAut(λ)αf(Z𝒳[λ]).\mathcal{Z}_{[\lambda]}=\bigcup_{f\in\operatorname{Aut}(\lambda)}\alpha_{f}(Z\cap\mathcal{X}_{[\lambda]}).
Proof.

Let Z=Z𝒳[λ]Z^{\prime}=Z\cap\mathcal{X}_{[\lambda]}. It is clear that 𝒵\mathcal{Z} contains αf(Z)\alpha_{f}(Z^{\prime}) for each automorphism ff of λ\lambda. Conversely, suppose that x𝒵[λ]x\in\mathcal{Z}_{[\lambda]}. Then xαg(Z)x\in\alpha_{g}(Z) for some correspondence g:λλg\colon\lambda\dashrightarrow\lambda. Write g=(g1:ρλ,g2:ρλ)g=(g_{1}\colon\rho\rightarrowtail\lambda,g_{2}\colon\rho\to\lambda). Then αg1(x)=αg2(y)\alpha_{g_{1}}(x)=\alpha_{g_{2}}(y) for some yZy\in Z. Each coordinate of xx must appear as a coordinate of yy. Since xx has distinct coordinates, the same is true for yy. Thus yZy\in Z^{\prime} and there is a permutation ff of λ\langle\lambda\rangle such that xi=yf(i)x_{i}=y_{f(i)} for each ii. We claim that ff is an automorphism of the \infty-composition λ\lambda. Let iλi\in\langle\lambda\rangle. Given any jg11(i)j\in g_{1}^{-1}(i), we have xi=yg2(i)x_{i}=y_{g_{2}(i)}, and so g2(i)=f(i)g_{2}(i)=f(i). It follows that g2g_{2} maps all of g11(i)g_{1}^{-1}(i) to f(i)f(i). Since g11(i)g_{1}^{-1}(i) has total weight λi\lambda_{i} (as g1g_{1} is a principal surjection), it follows that λiλf(i)\lambda_{i}\leq\lambda_{f(i)} (since g2g_{2} is a map of \infty-compositions). Since ff is bijective, we must have λi=λf(i)\lambda_{i}=\lambda_{f(i)} for all ii, and so ff is an automorphism of λ\lambda. This completes the proof. ∎

4.4. End\operatorname{End}-stable subsets

Let λ\lambda be an \infty-composition. We say that a set Z𝒳λZ\subset\mathcal{X}_{\lambda} is End(λ)\operatorname{End}(\lambda)-stable if αf(Z)Z\alpha_{f}(Z)\subset Z for every map f:λλf\colon\lambda\to\lambda of \infty-compositions. We define ZeZ^{e} to be the Zariski closure of the set fEnd(λ)αf(Z)\bigcup_{f\in\operatorname{End}(\lambda)}\alpha_{f}(Z). It is clear that ZeZ^{e} is End(λ)\operatorname{End}(\lambda)-stable. Moreover, if ZZ is End(λ)\operatorname{End}(\lambda)-stable and Zariski closed then Ze=ZZ^{e}=Z. In particular, we see that (Ze)e=Ze(Z^{e})^{e}=Z^{e} for any ZZ.

Proposition 4.8.

Suppose ZZ is End(λ)\operatorname{End}(\lambda)-stable Zariski closed subset of 𝒳λ\mathcal{X}_{\lambda}. Let f:μλf\colon\mu\dashrightarrow\lambda be a correspondence of \infty-compositions. Then αf(Z)\alpha_{f}(Z) is Zariski closed in 𝒳μ\mathcal{X}_{\mu}.

Proof.

First suppose that ff is an injective map of \infty-compositions. Choose iim(f)i\in\operatorname{im}(f) such that λi=\lambda_{i}=\infty. Define g:λμg\colon\langle\lambda\rangle\to\langle\mu\rangle by

g(j)={f1(j)if jim(f)f1(i)if jim(f).g(j)=\begin{cases}f^{-1}(j)&\text{if $j\in\operatorname{im}(f)$}\\ f^{-1}(i)&\text{if $j\not\in\operatorname{im}(f)$.}\end{cases}

Note that gf=id[n]g\circ f=\mathrm{id}_{[n]}. Let h:λλh\colon\langle\lambda\rangle\to\langle\lambda\rangle be the composition fgf\circ g; explicitly,

h(j)={jif jim(f)iif jim(f).h(j)=\begin{cases}j&\text{if $j\in\operatorname{im}(f)$}\\ i&\text{if $j\not\in\operatorname{im}(f)$.}\end{cases}

Note that hh defines a map of \infty-compositions λλ\lambda\to\lambda.

We claim that αf(Z)=αg1(Z)\alpha_{f}(Z)=\alpha_{g}^{-1}(Z). Indeed, suppose xαf(Z)x\in\alpha_{f}(Z). Then x=αf(y)x=\alpha_{f}(y) for some yZy\in Z. We thus have αg(x)=αg(αf(y))=αh(y)Z\alpha_{g}(x)=\alpha_{g}(\alpha_{f}(y))=\alpha_{h}(y)\in Z since ZZ is stable under αh\alpha_{h}. Thus xαg1(Z)x\in\alpha_{g}^{-1}(Z). Conversely, suppose xαg1(Z)x\in\alpha_{g}^{-1}(Z). Since gf=id[n]g\circ f=\mathrm{id}_{[n]}, it follows that αfαg\alpha_{f}\circ\alpha_{g} is the identity. Thus x=αf(αg(z))αf(Z)x=\alpha_{f}(\alpha_{g}(z))\in\alpha_{f}(Z), as αg(z)Z\alpha_{g}(z)\in Z. This proves the claim. It follows that αf(Z)\alpha_{f}(Z) is closed.

Now suppose that ff is an arbitrary correspondence. Let f=(f1:ρμ,f2:ρλ)f=(f_{1}\colon\rho\rightarrowtail\mu,f_{2}\colon\rho\to\lambda) where f1f_{1} is a principal projection. Factor f2f_{2} as ghg\circ h where hh is a principal projection and gg is an injection. Thus αf(Z)=αf11(αh(αg(Z)))\alpha_{f}(Z)=\alpha_{f_{1}}^{-1}(\alpha_{h}(\alpha_{g}(Z))). By the first paragraph, αg(Z)\alpha_{g}(Z) is closed. Since αh\alpha_{h} is a closed immersion, it follows that αh(αg(Z))\alpha_{h}(\alpha_{g}(Z)) is closed. Thus αf(Z)\alpha_{f}(Z) is closed. ∎

4.5. Construction of C3 subvarieties

For Z𝒳λZ\subset\mathcal{X}_{\lambda} define Γλ(Z)=Γλ(Ze)\Gamma_{\lambda}(Z)=\Gamma_{\lambda}^{\circ}(Z^{e}). The following theorem describes the most important properties of this construction.

Theorem 4.9.

Let Z𝒳λZ\subset\mathcal{X}_{\lambda} be an arbitrary subset. Then 𝒵=Γλ(Z)\mathcal{Z}=\Gamma_{\lambda}(Z) is a C3 subvariety of 𝒳\mathcal{X}. Furthermore, if 𝒵\mathcal{Z}^{\prime} is any C3 subvariety such that 𝒵λ\mathcal{Z}^{\prime}_{\lambda} contains ZZ then 𝒵\mathcal{Z}^{\prime} contains 𝒵\mathcal{Z}.

Proof.

By Theorem 4.5, we know that 𝒵\mathcal{Z} is a C2 subset of 𝒳\mathcal{X}. It thus suffices to show that 𝒵μ\mathcal{Z}_{\mu} is Zariski closed for all μ\mu. By Proposition 4.6, we have

𝒵μ=Γλ(Ze)μ=f:μλf is goodαf(Ze).\mathcal{Z}_{\mu}=\Gamma^{\circ}_{\lambda}(Z^{e})_{\mu}=\bigcup_{\begin{subarray}{c}f\colon\mu\dashrightarrow\lambda\\ \text{$f$ is good}\end{subarray}}\alpha_{f}(Z^{e}).

Since each αf(Ze)\alpha_{f}(Z^{e}) is Zariski closed (Proposition 4.8) and the union is finite, we see that 𝒵μ\mathcal{Z}_{\mu} is Zariski closed. Thus 𝒵\mathcal{Z} is a C3 subvariety.

Now, suppose 𝒵\mathcal{Z}^{\prime} is a C3 subvariety such that 𝒵λ\mathcal{Z}^{\prime}_{\lambda} contains ZZ. Since 𝒵\mathcal{Z}^{\prime} is closed under correspondences (Proposition 4.1), we see that 𝒵λ\mathcal{Z}^{\prime}_{\lambda} contains fEnd(λ)αf(Z)\bigcup_{f\in\operatorname{End}(\lambda)}\alpha_{f}(Z). Since 𝒵λ\mathcal{Z}^{\prime}_{\lambda} is also Zariski closed, it follows that it contains ZeZ^{e}. Since 𝒵\mathcal{Z}^{\prime} is a C2 subset such that 𝒵λ\mathcal{Z}^{\prime}_{\lambda} contains ZeZ^{e}, Theorem 4.5 shows that 𝒵\mathcal{Z}^{\prime} contains Γλ(Ze)=𝒵\Gamma^{\circ}_{\lambda}(Z^{e})=\mathcal{Z}. This completes the proof. ∎

We will require some additional properties of the Γλ\Gamma_{\lambda} construction:

Proposition 4.10.

Let Z𝒳[λ]Z\subset\mathcal{X}_{[\lambda]} and put 𝒵=Γλ(Z)\mathcal{Z}=\Gamma_{\lambda}(Z). Then 𝒵[λ]\mathcal{Z}_{[\lambda]} is the Zariski closure of fAut(λ)αf(Z)\bigcup_{f\in\operatorname{Aut}(\lambda)}\alpha_{f}(Z) in 𝒳[λ]\mathcal{X}_{[\lambda]}.

Proof.

Let E=𝒳λ𝒳[λ]E=\mathcal{X}_{\lambda}\setminus\mathcal{X}_{[\lambda]}; this is a Zariski closure subset of 𝒳λ\mathcal{X}_{\lambda}. Also, let ZZ^{\prime} be the Zariski closure of fAut(λ)αf(Z)\bigcup_{f\in\operatorname{Aut}(\lambda)}\alpha_{f}(Z). If f:λλf\colon\lambda\to\lambda is any endomorphism that is not an automorphism then αf(𝒳λ)E\alpha_{f}(\mathcal{X}_{\lambda})\subset E. It follows that ZZeZEZ^{\prime}\subset Z^{e}\subset Z^{\prime}\cup E, and so Ze𝒳[λ]=ZZ^{e}\cap\mathcal{X}_{[\lambda]}=Z^{\prime}. Since 𝒵=Γλ(Ze)\mathcal{Z}=\Gamma_{\lambda}^{\circ}(Z^{e}), the result now follows from Proposition 4.7. ∎

Proposition 4.11.

Let ZZ be an Aut(λ)\operatorname{Aut}(\lambda)-irreducible closed subvariety of 𝒳[λ]\mathcal{X}_{[\lambda]}. Then 𝒵=Γλ(Z)\mathcal{Z}=\Gamma_{\lambda}(Z) is an irreducible C3-subvariety of 𝒳\mathcal{X}.

Proof.

Suppose 𝒵\mathcal{Z} can be written as a union 𝒵1𝒵2\mathcal{Z}^{1}\cup\mathcal{Z}^{2} of C3 subvarieties of 𝒳\mathcal{X}. By Proposition 4.10, we have 𝒵[λ]=Z\mathcal{Z}_{[\lambda]}=Z; on the other hand, 𝒵[λ]=𝒵[λ]1𝒵[λ]2\mathcal{Z}_{[\lambda]}=\mathcal{Z}^{1}_{[\lambda]}\cup\mathcal{Z}^{2}_{[\lambda]}. Since ZZ is Aut(λ)\operatorname{Aut}(\lambda)-irreducible, we have 𝒵1=Z\mathcal{Z}^{1}=Z or 𝒵2=Z\mathcal{Z}^{2}=Z, say the former. Since 𝒵1\mathcal{Z}^{1} is a C3 subvariety containing ZZ it must contain 𝒵\mathcal{Z} by Theorem 4.9, and so 𝒵=𝒵1\mathcal{Z}=\mathcal{Z}^{1}, which completes the proof. ∎

Remark 4.12.

We note that if ZZ is End(λ)\operatorname{End}(\lambda)-stable closed subvariety of 𝒳λ\mathcal{X}_{\lambda}, then Γλ(Z)λ\Gamma_{\lambda}(Z)_{\lambda} may strictly contain ZZ. To see this, let W=Spec(𝐂)W=\operatorname{Spec}(\mathbf{C}), λ=(,2,1,1)\lambda=(\infty,2,1,1), and let ZZ be the End(λ)\operatorname{End}(\lambda)-closure of {(1,2,3,3)}\{(1,2,3,3)\}. Then it is easy to see that (1,3,2,2)(1,3,2,2) is not in ZZ. On the other hand, let f=(f1,f2)f=(f_{1},f_{2}) be the correspondence where f1:(,1,1,1,1)λf_{1}\colon(\infty,1,1,1,1)\to\lambda is the principal surjection which combines the second and the third coordinates and f2:(,1,1,1,1)λf_{2}\colon(\infty,1,1,1,1)\to\lambda is the principal surjection which combines the third and the fourth coordinates. Now it is easy to see that (1,3,2,2)αf({(1,2,3,3)})(1,3,2,2)\in\alpha_{f}(\{(1,2,3,3)\}). Thus Γλ(Z)λ\Gamma_{\lambda}(Z)_{\lambda} strictly contain ZZ. ∎

5. Classification of subvarieties of 𝒳\mathcal{X}

We assume AA is noetherian throughout this section

5.1. Bounded subvarieties of 𝒳\mathcal{X}

Let 𝒵\mathcal{Z} be a subset of 𝒳\mathcal{X}. We say that 𝒵\mathcal{Z} is bounded if there is some nn such that (λ)n\ell(\lambda)\leq n for all λ\lambda with 𝒵[λ]\mathcal{Z}_{[\lambda]} non-empty. For an \infty-composition λ\lambda, let 𝒳λn𝒳λ\mathcal{X}^{\leq n}_{\lambda}\subset\mathcal{X}_{\lambda} be the subset where the coordinates take at most nn distinct values. This is a Zariski closed subset, as it is defined by the nn-variable discriminants. One easily sees that 𝒳n\mathcal{X}^{\leq n} is a C3-subvariety of 𝒳\mathcal{X}. Essentially by definition, a subset of 𝒳\mathcal{X} is bounded if and only if it is contained in 𝒳n\mathcal{X}^{\leq n} for some nn.

The purpose of this section is to establish the following result:

Proposition 5.1.

Let 𝒵\mathcal{Z} be a C3 subvariety of 𝒳\mathcal{X}. Then we have a decomposition 𝒵=𝒵1𝒵2\mathcal{Z}=\mathcal{Z}_{1}\cup\mathcal{Z}_{2} where 𝒵1=𝒳W\mathcal{Z}_{1}=\mathcal{X}_{W^{\prime}} for some closed subset WW^{\prime} of WW and 𝒵2\mathcal{Z}_{2} is a bounded C3 subvariety.

Here 𝒳W\mathcal{X}_{W^{\prime}} denotes the subset of 𝒳\mathcal{X} given by 𝒳W,λ=𝐀Wλ\mathcal{X}_{W^{\prime},\lambda}=\mathbf{A}_{W^{\prime}}^{\langle\lambda\rangle}. We require several lemmas. We say that a KK-point wW(K)w\in W(K) is ever-present in 𝒵\mathcal{Z} if 𝒵\mathcal{Z} contains 𝒳{w}\mathcal{X}_{\{w\}}. This notion is invariant under field extension, and thus we can speak of ever-present points of 𝒵\mathcal{Z}.

Lemma 5.2.

Let KK be a field, and let ww be a KK-point of WW. Let 𝒵\mathcal{Z} be a C3 subvariety of 𝒳\mathcal{X}. Suppose that there is an element aKa\in K such that 𝒵(,1n)(K)\mathcal{Z}_{(\infty,1^{n})}(K) contains the KK-points in {(w,a)}×𝐀n\{(w,a)\}\times\mathbf{A}^{n} for all nn. Then ww is ever-present in 𝒵\mathcal{Z}.

Proof.

Suppose λ=(λ1,λ2,,λr)\lambda=(\lambda_{1},\lambda_{2},\ldots,\lambda_{r}) is an \infty-partition with λ1=\lambda_{1}=\infty and λi<\lambda_{i}<\infty for i>1i>1. We can then find a principal surjection (,1n)λ(\infty,1^{n})\to\lambda with n=λ2++λrn=\lambda_{2}+\cdots+\lambda_{r}. Since 𝒵\mathcal{Z} is C1, we have 𝒵λ=𝒵(,1n)𝒳λ\mathcal{Z}_{\lambda}=\mathcal{Z}_{(\infty,1^{n})}\cap\mathcal{X}_{\lambda}, and so 𝒵λ(K)\mathcal{Z}_{\lambda}(K) contains the KK-points in {(w,a)}×𝐀r1\{(w,a)\}\times\mathbf{A}^{r-1}.

Now suppose that λ=(λ1,,λr)\lambda=(\lambda_{1},\ldots,\lambda_{r}) is an \infty-partition with λ1==λs=\lambda_{1}=\cdots=\lambda_{s}=\infty and λs+1<\lambda_{s+1}<\infty. Let λn=(,n,,n,λs+1,,λr)\lambda^{n}=(\infty,n,\ldots,n,\lambda_{s+1},\ldots,\lambda_{r}), so that λ\lambda is the limit of the λn\lambda^{n}’s. Since 𝒵\mathcal{Z} is C2, we have 𝒵λ=n1𝒵λn\mathcal{Z}_{\lambda}=\bigcap_{n\geq 1}\mathcal{Z}_{\lambda^{n}} (Proposition 3.6). Since each 𝒵λn(K)\mathcal{Z}_{\lambda^{n}}(K) contains the KK-points in {(w,a)}×𝐀r1\{(w,a)\}\times\mathbf{A}^{r-1}, so does 𝒵λ\mathcal{Z}_{\lambda}.

Keeping λ\lambda as above, let μ=(,λ1,,λr)\mu=(\infty,\lambda_{1},\ldots,\lambda_{r}). We have a natural injection λμ\lambda\to\mu. By Proposition 3.4, we see that 𝒵λ\mathcal{Z}_{\lambda} contains the image of 𝒵μ\mathcal{Z}_{\mu} under the projection 𝒳μ𝒳λ\mathcal{X}_{\mu}\to\mathcal{X}_{\lambda}. However, any KK-point of the form (w,b1,,br)𝒳λ(K)(w,b_{1},\ldots,b_{r})\in\mathcal{X}_{\lambda}(K) is the image of the point (w,a,b1,,br)𝒳μ(K)(w,a,b_{1},\ldots,b_{r})\in\mathcal{X}_{\mu}(K), which belongs to 𝒵μ(K)\mathcal{Z}_{\mu}(K). We thus see that 𝒵λ(K)\mathcal{Z}_{\lambda}(K) contains the KK-points in {w}×𝐀(λ)\{w\}\times\mathbf{A}^{\ell(\lambda)}, as claimed. ∎

Lemma 5.3.

Let 𝒵\mathcal{Z} be a C3 subvariety of 𝒳\mathcal{X} with no ever-present point. Then 𝒵\mathcal{Z} is bounded.

Proof.

We prove the contrapositive. Thus suppose that 𝒵\mathcal{Z} is unbounded, and we will produce an ever-present point. It suffices to treat the case where WW is affine, so say W=Spec(A)W=\operatorname{Spec}(A).

We first claim that 𝒵[,1n]\mathcal{Z}_{[\infty,1^{n}]} is non-empty for all n0n\geq 0. Indeed, let nn be given. Since 𝒵\mathcal{Z} is unbounded, there is an \infty-composition λ\lambda with (λ)n\ell(\lambda)\geq n such that 𝒵[λ]\mathcal{Z}_{[\lambda]} is non-empty. Pick an injection f:(,1n)λf\colon(\infty,1^{n})\to\lambda, which is possible. Then αf\alpha_{f} maps 𝒵[λ]\mathcal{Z}_{[\lambda]} into 𝒵[,1n]\mathcal{Z}_{[\infty,1^{n}]}, and so the latter is non-empty, as claimed.

Put W=W×𝐀1W^{\prime}=W\times\mathbf{A}^{1}. Let II be a finite subset of [][\infty], and let λI\lambda_{I} be the \infty-composition on I{}I\cup\{\ast\} that is 1 on II and \infty on \ast. Put 𝒳I=𝒳λI\mathcal{X}_{I}=\mathcal{X}_{\lambda_{I}}, which we identify with W×𝐀IW^{\prime}\times\mathbf{A}^{I}, and let 𝒳I\mathcal{X}^{\prime}_{I} be the open subvariety W×𝐀[I]W^{\prime}\times\mathbf{A}^{[I]}. Put 𝒵I=𝒵λI\mathcal{Z}_{I}=\mathcal{Z}_{\lambda_{I}} and 𝒵I=𝒵I𝒳I\mathcal{Z}^{\prime}_{I}=\mathcal{Z}_{I}\cap\mathcal{X}^{\prime}_{I}. Let RI=A[u,ξi]iIR_{I}=A[u,\xi_{i}]_{i\in I} and SI=RI[(ξiξj)1]S_{I}=R_{I}[(\xi_{i}-\xi_{j})^{-1}]. We identify A[u]A[u], RIR_{I}, and SIS_{I} with the coordinate rings of WW^{\prime}, 𝒳I\mathcal{X}_{I}, and 𝒳I\mathcal{X}^{\prime}_{I}. Let 𝔞ISI\mathfrak{a}_{I}\subset S_{I} be the ideal corresponding to 𝒵I\mathcal{Z}^{\prime}_{I}. Since 𝒵I\mathcal{Z}^{\prime}_{I} is non-empty, the ideal 𝔞I\mathfrak{a}_{I} is not the unit ideal.

Suppose that JJ is a second finite subset of [][\infty] containing II. We have a natural injection of \infty-compositions λIλJ\lambda_{I}\to\lambda_{J} that induces a projection map 𝒳J𝒳I\mathcal{X}_{J}\to\mathcal{X}_{I} which carries 𝒳J\mathcal{X}^{\prime}_{J} into 𝒳I\mathcal{X}^{\prime}_{I}, and hence 𝒵J\mathcal{Z}^{\prime}_{J} into 𝒵I\mathcal{Z}^{\prime}_{I}. The map 𝒳J𝒳I\mathcal{X}_{J}\to\mathcal{X}_{I} corresponds to the canonical inclusion of rings RIRJR_{I}\to R_{J}. We thus see that, under this inclusion, 𝔞I\mathfrak{a}_{I} is contained in 𝔞J\mathfrak{a}_{J}.

Let R=A[u,ξi]i[]R=A[u,\xi_{i}]_{i\in[\infty]} and let S=R[(ξiξj)1]ijS=R[(\xi_{i}-\xi_{j})^{-1}]_{i\neq j}. We identify RR and SS with the direct limits of RIR_{I} and SIS_{I}, taken over all finite subsets II of [][\infty]. Let 𝔞S\mathfrak{a}\subset S be the direct limit of the 𝔞I\mathfrak{a}_{I}’s. This is not the unit ideal: indeed, if 𝔞\mathfrak{a} contained 1 then some 𝔞I\mathfrak{a}_{I} would contain 1, which it does not. By Proposition 2.7, we see that 𝔞\mathfrak{a} is the extension of the ideal 𝔟=𝔞c\mathfrak{b}=\mathfrak{a}^{c} of A[u]A[u].

Let (w,a)(w,a) be a KK-point of V(𝔟)W×𝐀1V(\mathfrak{b})\subset W\times\mathbf{A}^{1}. Then (w,a)(w,a) determines a map A[u]KA[u]\to K whose kernel contains 𝔟\mathfrak{b}. Let b=(b1,b2,)b=(b_{1},b_{2},\ldots) be any sequence of distinct elements of KK. We then have a homomorphism φb:S𝐤\varphi_{b}\colon S\to\mathbf{k} given by mapping an element of A[u]A[u] to its image in KK and ξi\xi_{i} to bib_{i}, and the kernel of φb\varphi_{b} contains 𝔞\mathfrak{a}. It follows that ker(φb|SI)\ker(\varphi_{b}|_{S_{I}}) contains 𝔞I\mathfrak{a}_{I}, and so (a,b1,,bn)(a,b_{1},\ldots,b_{n}) is a KK-point of 𝒵I\mathcal{Z}^{\prime}_{I}. Since this holds for all choices of distinct b1,,bnb_{1},\ldots,b_{n}, we see that the set of KK-points in 𝒵I\mathcal{Z}_{I} contains {(w,a)}×𝐀[I]\{(w,a)\}\times\mathbf{A}^{[I]}, and thus contains its closure {(w,a)}×𝐀I\{(w,a)\}\times\mathbf{A}^{I}. By Lemma 5.2, we see that ww is an ever-present KK-point in 𝒵\mathcal{Z}. Thus 𝒵\mathcal{Z} has an ever-present point. ∎

Lemma 5.4.

Let ZZ be a closed subvariety of W×𝐀nW\times\mathbf{A}^{n}. Let WW^{\prime} be the set of points wWw\in W such that ZZ contains {w}×𝐀n\{w\}\times\mathbf{A}^{n}. Then WW^{\prime} is closed.

Proof.

It suffices to treat the case where WW is affine, so say W𝐀mW\subset\mathbf{A}^{m}. Thus ZZ is a closed subvariety of 𝐀m×𝐀n\mathbf{A}^{m}\times\mathbf{A}^{n}. Let fi(ξ1,,ξm,η1,,ηn)=0f_{i}(\xi_{1},\ldots,\xi_{m},\eta_{1},\ldots,\eta_{n})=0, for 1ir1\leq i\leq r, be equations definining ZZ. If x=(x1,,xm)x=(x_{1},\ldots,x_{m}) is a point of 𝐀m\mathbf{A}^{m}, then ZZ contains {x}×𝐀n\{x\}\times\mathbf{A}^{n} if and only if the polynomials fi(x1,,xm,η1,,ηn)f_{i}(x_{1},\ldots,x_{m},\eta_{1},\ldots,\eta_{n}) vanish identically, i.e., all coefficients vanish. Thus this locus is defined by a system of polynomial equations, and thus closed. ∎

Lemma 5.5.

Let 𝒵\mathcal{Z} be a C3 subvariety of 𝒳\mathcal{X} and let WWW^{\prime}\subset W be the set of ever-present points. Then WW^{\prime} is closed.

Proof.

For an \infty-composition λ\lambda, let WλWW_{\lambda}\subset W be the set of points ww such that 𝒵λ\mathcal{Z}_{\lambda} contains 𝒳{w},λ\mathcal{X}_{\{w\},\lambda}. By Lemma 5.4, we see that WλW_{\lambda} is closed. Since W=WλW^{\prime}=\bigcap W_{\lambda}, with the intersection over all \infty-compositions λ\lambda, we see that WW^{\prime} is also closed. ∎

Proof of Proposition 5.1.

Let 𝒵𝒳\mathcal{Z}\subset\mathcal{X} be a given C3-subvariety. Let WWW^{\prime}\subset W be the set of ever-present points, which is closed by Lemma 5.5, and put 𝒵1=𝒳W𝒵\mathcal{Z}_{1}=\mathcal{X}_{W^{\prime}}\subset\mathcal{Z}. Let U=WWU=W\setminus W^{\prime}, and write U=U1UrU=U_{1}\cup\cdots\cup U_{r} with each UiU_{i} affine, which is possible since WW is noetherian. Let 𝒵i=𝒵𝒳Ui\mathcal{Z}^{\prime}_{i}=\mathcal{Z}\cap\mathcal{X}_{U_{i}}. It is clear that 𝒵i\mathcal{Z}^{\prime}_{i} is a C3-subvariety of 𝒳Ui\mathcal{X}_{U_{i}} that contains no ever-present points, and is therefore bounded by Lemma 5.3. Thus there is some nn such that 𝒵i\mathcal{Z}^{\prime}_{i} is contained in 𝒳Uin\mathcal{X}^{\leq n}_{U_{i}} for all 1ir1\leq i\leq r. Let 𝒵2=𝒵𝒳n\mathcal{Z}_{2}=\mathcal{Z}\cap\mathcal{X}^{\leq n}, which is a bounded C3-subvariety of 𝒵\mathcal{Z}. Since 𝒵2\mathcal{Z}_{2} contains 𝒵i\mathcal{Z}^{\prime}_{i} for all ii, we have 𝒵=𝒵1𝒵2\mathcal{Z}=\mathcal{Z}_{1}\cup\mathcal{Z}_{2}, as required. ∎

5.2. The noetherian property for 𝒳\mathcal{X}

The purpose of this section is to prove that 𝒳\mathcal{X} is noetherian, in the following sense:

Proposition 5.6.

The descending chain condition holds for C3 subvarieties of 𝒳\mathcal{X}.

Proof.

Recall that an ideal II of a poset (P,)(P,\preceq) is a subset such that xI,xyyIx\in I,x\preceq y\implies y\in I. Let Λ\Lambda be the poset \infty-partitions with respect to \preceq. Being a subposet of Λ~\widetilde{\Lambda}, we see that Λ\Lambda is a wqo (Proposition 3.1). It follows that the poset of ideals of Λ\Lambda under inclusion satisfies ACC; see [SS, §2].

We first deal with the bounded case. Let 𝒵1𝒵2\mathcal{Z}^{1}\supset\mathcal{Z}^{2}\supset\cdots be a decreasing chain of bounded C3-subvarieties of 𝒳\mathcal{X}. Let Λn\Lambda_{n} be the set given by

Λn={λΛ:μλ,𝒵[μ]k stabilizes for kn}.\Lambda_{n}=\{\lambda\in\Lambda\colon\forall\mu\succeq\lambda,\mathcal{Z}^{k}_{[\mu]}\text{ stabilizes for }k\geq n\}.

Then Λ1Λ2\Lambda_{1}\subset\Lambda_{2}\subset\cdots is an increasing chain of ideals of Λ\Lambda, and so it must stabilize. Let NN be large enough such that Λn=ΛN\Lambda_{n}=\Lambda_{N} for all nNn\geq N. Suppose, if possible, ΛΛN\Lambda\setminus\Lambda_{N}\neq\varnothing. We claim that ΛΛN\Lambda\setminus\Lambda_{N} has a maximal element:

To see this let Λ~n\widetilde{\Lambda}_{n} be the set given by

Λ~n={λΛ:𝒵[λ]k stabilizes for kn}.\widetilde{\Lambda}_{n}=\{\lambda\in\Lambda\colon\mathcal{Z}^{k}_{[\lambda]}\text{ stabilizes for }k\geq n\}.

Note that Λ~n\widetilde{\Lambda}_{n} may not be an ideal of Λ\Lambda. Moreover, the set ΛΛ~N\Lambda\setminus\widetilde{\Lambda}_{N} is nonempty as ΛΛN\Lambda\setminus\Lambda_{N} is nonempty. We first show that ΛΛ~N\Lambda\setminus\widetilde{\Lambda}_{N} has a maximal element. By Zorn’s lemma, it suffices to show that every increasing chain λ1λ2\lambda^{1}\preceq\lambda^{2}\preceq\cdots in ΛΛ~N\Lambda\setminus\widetilde{\Lambda}_{N} has an upper bound in ΛΛ~N\Lambda\setminus\widetilde{\Lambda}_{N}. Since 𝒵1\mathcal{Z}^{1} is bounded, there exists a dd such that (λ)>d\ell(\lambda)>d implies λΛ1ΛNΛ~N\lambda\in\Lambda^{1}\subset\Lambda_{N}\subset\widetilde{\Lambda}_{N}. This shows that (λi)d\ell(\lambda^{i})\leq d for all ii. Suppose that the limiting value of (λi)\ell(\lambda^{i}) is some ddd^{\prime}\leq d. We may as well assume that (λi)=d\ell(\lambda^{i})=d^{\prime} for all ii. Then we can regard each λi\lambda^{i} as an \infty-weighting on the set [d][d^{\prime}], and so we have λ1λ2\lambda^{1}\leq\lambda^{2}\leq\cdots. Let λΛ\lambda\in\Lambda be the limit of these \infty-compositions. By noetherianity, 𝒵λ1N𝒵λ2N\mathcal{Z}^{N}_{\lambda^{1}}\supset\mathcal{Z}^{N}_{\lambda^{2}}\supset\cdots stabilizes for each kk. So by the continuity of 𝒵N\mathcal{Z}^{N}, we see that 𝒵[λ]N=𝒵[λi]N\mathcal{Z}^{N}_{[\lambda]}=\mathcal{Z}^{N}_{[\lambda^{i}]} for i0i\gg 0. If λ\lambda were in Λ~N\widetilde{\Lambda}_{N}, then for such an ii and all nNn\geq N we would have

𝒵[λi]n𝒵[λ]n=𝒵[λ]N=𝒵[λi]N𝒵[λi]n,\mathcal{Z}^{n}_{[\lambda^{i}]}\supset\mathcal{Z}^{n}_{[\lambda]}=\mathcal{Z}^{N}_{[\lambda]}=\mathcal{Z}^{N}_{[\lambda^{i}]}\supset\mathcal{Z}^{n}_{[\lambda^{i}]},

where the first equality comes from λΛ~N\lambda\in\widetilde{\Lambda}_{N}, and the second equality is true as i0i\gg 0. This shows that 𝒵λin=𝒵λiN\mathcal{Z}^{n}_{\lambda^{i}}=\mathcal{Z}^{N}_{\lambda^{i}} for all nNn\geq N (by Proposition 3.3), which contradicts λiΛΛ~N\lambda^{i}\in\Lambda\setminus\widetilde{\Lambda}_{N}. Thus λΛΛ~N\lambda\in\Lambda\setminus\widetilde{\Lambda}_{N}. By Zorn’s lemma, there exists a maximal element, say λ\lambda, in ΛΛ~N\Lambda\setminus\widetilde{\Lambda}_{N}. It is clear that λ\lambda is also a maximal element in ΛΛN\Lambda\setminus\Lambda_{N}. This proves the claim.

Now let λ\lambda be a maximal element in ΛΛN\Lambda\setminus\Lambda_{N}. By noetherianity, there exists an NN^{\prime} such that 𝒵λ1𝒵λ2\mathcal{Z}^{1}_{\lambda}\supset\mathcal{Z}^{2}_{\lambda}\supset\cdots stabilizes for nNn\geq N^{\prime}. Since λΛN\lambda\notin\Lambda_{N}, we must have N>NN^{\prime}>N. But then λΛNΛN\lambda^{\prime}\in\Lambda_{N^{\prime}}\setminus\Lambda_{N}, contradicting the definition of NN. Thus Λ=ΛN\Lambda=\Lambda_{N}. By the previous lemma, the increasing chain 𝒵1𝒵2\mathcal{Z}^{1}\supset\mathcal{Z}^{2}\supset\cdots stabilizes for nNn\geq N. This completes the proof in the bounded case.

In general, let 𝒵1𝒵2\mathcal{Z}^{1}\supset\mathcal{Z}^{2}\supset\cdots be a decreasing chain of (non necessarily bounded) C3-subvarieties of 𝒳=𝒳W\mathcal{X}=\mathcal{X}_{W}. Let 𝒳Wi\mathcal{X}_{W_{i}} be the ever-present part of 𝒵i\mathcal{Z}^{i} as in Proposition 5.1. Then W1W2W_{1}\supset W_{2}\supset\cdots is a decreasing chain of closed subvarities of WW. By noetherianity of WW, this decreasing chain stabilizes. So we may as well assume that Wi=WW_{i}=W^{\prime} for all i1i\geq 1. Set U=WWU=W\setminus W^{\prime}, and let 𝒴i=𝒵i𝒳U\mathcal{Y}^{i}=\mathcal{Z}^{i}\cap\mathcal{X}_{U}. By Lemma 5.3, 𝒴i\mathcal{Y}^{i} is bounded for each ii. By the bounded case, the decreasing chain 𝒴1𝒴2\mathcal{Y}^{1}\supset\mathcal{Y}^{2}\supset\cdots stabilizes. Since 𝒵i=𝒳W𝒴i\mathcal{Z}^{i}=\mathcal{X}_{W^{\prime}}\cup\mathcal{Y}^{i}, the decreasing chain 𝒵1𝒵2\mathcal{Z}^{1}\supset\mathcal{Z}^{2}\supset\cdots also stabilizes, completing the proof. ∎

The following corollary follows immediately from the proposition:

Corollary 5.7.

Every C3 subvariety is a finite union of irreducible C3 subvarieties.

5.3. The classification theorem

The following is the classification theorem for irreducible C3 subvarieties of 𝒳\mathcal{X}.

Theorem 5.8.

Let 𝒵\mathcal{Z} be an irreducible C3 subvariety of 𝒳\mathcal{X}.

  1. (a)

    If 𝒵\mathcal{Z} is bounded then 𝒵=Γλ(Z)\mathcal{Z}=\Gamma_{\lambda}(Z) for a unique \infty-partition λ\lambda and a unique Aut(λ)\operatorname{Aut}(\lambda)-irreducible subvariety ZZ of 𝒳[λ]\mathcal{X}_{[\lambda]}.

  2. (b)

    If 𝒵\mathcal{Z} is unbounded, then 𝒵=𝒳W\mathcal{Z}=\mathcal{X}_{W^{\prime}} for a unique irreducible closed subset WWW^{\prime}\subset W.

Proof.

Part (b) follows immediately from Proposition 5.1. For Part (a), let Λ𝒵\Lambda_{\mathcal{Z}} be the set of partitions λ\lambda such that 𝒵[λ]\mathcal{Z}_{[\lambda]} is nonempty. Since 𝒵\mathcal{Z} is bounded, there exists an NN such that the number of parts in any λΛ𝒵\lambda\in\Lambda_{\mathcal{Z}} is bounded by NN. It follows that if λ1λ2\lambda^{1}\preceq\lambda^{2}\preceq\cdots is a chain in Λ𝒵\Lambda_{\mathcal{Z}}, then the number of parts in λn\lambda^{n} is eventually independent of nn. Ignoring the first few entries, we can think of each of λn\lambda^{n} to be an \infty-composition on the same underlying set. By Proposition 3.6, we see that the limit λ\lambda of λn\lambda^{n} is also in Λ𝒵\Lambda_{\mathcal{Z}}. By Zorn’s lemma, every element in Λ𝒵\Lambda_{\mathcal{Z}} is contained in a maximal element. By Proposition 3.1, there are no infinite antichains in (Λ𝒵,)(\Lambda_{\mathcal{Z}},\preceq). So there are finitely many maximal elements in Λ𝒵\Lambda_{\mathcal{Z}}, call them μ1,,μn\mu^{1},\ldots,\mu^{n}. Let 𝒵i\mathcal{Z}^{i} be the minimal C3-subvariety of 𝒳\mathcal{X} such that 𝒵μii=𝒳μi\mathcal{Z}^{i}_{\mu^{i}}=\mathcal{X}_{\mu^{i}}. By Theorem 4.9, we see that 𝒵λi=𝒳λ\mathcal{Z}^{i}_{\lambda}=\mathcal{X}_{\lambda} for each λμi\lambda\preceq\mu_{i}. So 𝒵1in𝒵i\mathcal{Z}\subset\bigcup_{1\leq i\leq n}\mathcal{Z}^{i}. By irreducibility, there exists an ii such that 𝒵𝒵i\mathcal{Z}\subset\mathcal{Z}^{i}. Since 𝒵[λ]i\mathcal{Z}^{i}_{[\lambda]} is nonempty if and only if λμi\lambda\preceq\mu^{i}, it follows that μi\mu^{i} is the unique maximal element of Λ𝒵\Lambda_{\mathcal{Z}}. Set μμi,Z𝒵[μ]\mu\coloneq\mu_{i},Z\coloneq\mathcal{Z}_{[\mu]}, and let 𝒵\mathcal{Z}^{\prime} be the minimal C3-subvariety of 𝒳\mathcal{X} such that 𝒵μZ\mathcal{Z}^{\prime}_{\mu}\supset Z (which is as described in the Theorem 4.9). By minimality, we have 𝒵𝒵\mathcal{Z}^{\prime}\subset\mathcal{Z}. We claim that 𝒵=𝒵\mathcal{Z}^{\prime}=\mathcal{Z}.

To see this, let Λ={λΛ𝒵:𝒵[λ]=𝒵[λ]}\Lambda^{\prime}=\{\lambda\in\Lambda_{\mathcal{Z}}\colon\mathcal{Z}^{\prime}_{[\lambda]}=\mathcal{Z}_{[\lambda]}\}. Let λ1λ2\lambda^{1}\preceq\lambda^{2}\preceq\cdots be a chain in Λ𝒵Λ\Lambda_{\mathcal{Z}}\setminus\Lambda^{\prime}. By boundedness of 𝒵\mathcal{Z}, the number of parts in λn\lambda^{n} is eventually constant in nn. Ignoring the first few entries, we can think of each of λn\lambda^{n} to be an \infty-composition on the same underlying set. By noetherianity (Proposition 5.6), we see that the chains 𝒵λ1𝒵λ2\mathcal{Z}_{\lambda^{1}}\supset\mathcal{Z}_{\lambda^{2}}\supset\cdots and 𝒵λ1𝒵λ2\mathcal{Z}^{\prime}_{\lambda^{1}}\supset\mathcal{Z}^{\prime}_{\lambda^{2}}\supset\cdots stabilize. So Proposition 3.6 applied to 𝒵\mathcal{Z} and 𝒵\mathcal{Z}^{\prime}, we see that for large enough NN, we have 𝒵[λN]=𝒵[λ]\mathcal{Z}_{[\lambda^{N}]}=\mathcal{Z}_{[\lambda]} and 𝒵[λN]=𝒵[λ]\mathcal{Z}^{\prime}_{[\lambda^{N}]}=\mathcal{Z}^{\prime}_{[\lambda]}. It follows that 𝒵[λ]=𝒵[λN]𝒵[λN]=𝒵[λ]\mathcal{Z}^{\prime}_{[\lambda]}=\mathcal{Z}^{\prime}_{[\lambda^{N}]}\subsetneq\mathcal{Z}_{[\lambda^{N}]}=\mathcal{Z}_{[\lambda]}, and so λΛ𝒵Λ\lambda\in\Lambda_{\mathcal{Z}}\setminus\Lambda^{\prime}. By Zorn’s lemma, every element in Λ𝒵Λ\Lambda_{\mathcal{Z}}\setminus\Lambda^{\prime} is contained in a maximal element. By Proposition 3.1, there are no infinite antichains in (Λ,)(\Lambda,\preceq). So there are finitely many maximal elements in Λ𝒵Λ\Lambda_{\mathcal{Z}}\setminus\Lambda^{\prime}, call them μ1,,μn\mu^{1},\ldots,\mu^{n}. As in the previous paragraph, let 𝒵i\mathcal{Z}^{i} be the minimal C3-subvariety of 𝒳\mathcal{X} such that 𝒵μii=𝒳μi\mathcal{Z}^{i}_{\mu^{i}}=\mathcal{X}_{\mu^{i}}. Let 𝒵′′=𝒵𝒵i=1n𝒵i\mathcal{Z}^{\prime\prime}=\mathcal{Z}\subset\mathcal{Z}^{\prime}\cup\bigcup_{i=1}^{n}\mathcal{Z}^{i}. Our construction implies that 𝒵[λ]𝒵[λ]′′\mathcal{Z}_{[\lambda]}\subset\mathcal{Z}^{\prime\prime}_{[\lambda]} for each λ\lambda. By Proposition 3.3, we see that 𝒵𝒵′′=𝒵i=1n𝒵i\mathcal{Z}\subset\mathcal{Z}^{\prime\prime}=\mathcal{Z}^{\prime}\cup\bigcup_{i=1}^{n}\mathcal{Z}^{i}. Since 𝒵[μ]i=\mathcal{Z}^{i}_{[\mu]}=\varnothing, the irreducibility of 𝒵\mathcal{Z} implies that 𝒵𝒵\mathcal{Z}\subset\mathcal{Z}^{\prime}. This establishes the claim.

Finally, suppose Z=𝒵[μ]Z=\mathcal{Z}_{[\mu]} can be written as Z=i=1nZnZ=\bigcup_{i=1}^{n}Z_{n} where each ZnZ_{n} is Aut(μ)\operatorname{Aut}(\mu)-stable Zariski closed Aut(μ)\operatorname{Aut}(\mu)-irreducible subset of 𝒳[μ]\mathcal{X}_{[\mu]}. Let 𝒵i\mathcal{Z}^{i} be the irreducible C3-subvariety corresponding to ZiZ_{i}. Then by Theorem 4.9, we see that 𝒵=i=1n𝒵i\mathcal{Z}=\bigcup_{i=1}^{n}\mathcal{Z}^{i}. By irreducibility of 𝒵\mathcal{Z} again, we see that 𝒵=𝒵i\mathcal{Z}=\mathcal{Z}^{i} for some ii. This finishes the proof. ∎

6. Extending from 𝒳\mathcal{X} to 𝒳+\mathcal{X}^{+}

In the next section, we study the correspondence between subvarieties of 𝔛\mathfrak{X} and subvarieties of 𝒳\mathcal{X}. To carry this out, we extend 𝒳\mathcal{X} to also include finite compositions. We now study this construction.

Given a non-empty generalized composition λ\lambda, we let 𝒳λ+=𝐀W[λ]\mathcal{X}^{+}_{\lambda}=\mathbf{A}^{[\lambda]}_{W}. If λ\lambda is infinite then 𝒳λ+\mathcal{X}^{+}_{\lambda} is simply 𝒳λ\mathcal{X}_{\lambda}. We now consider subsets of 𝒳+\mathcal{X}^{+}. We define the conditions C1, C2, and C3 exactly as for 𝒳\mathcal{X}. The elementary results about subsets of 𝒳\mathcal{X} extend to subsets of 𝒳+\mathcal{X}^{+} without difficulty. Given a subset 𝒵\mathcal{Z} of 𝒳+\mathcal{X}^{+}, we let 𝒵𝒳\mathcal{Z}\cap\mathcal{X} be the subset of 𝒳\mathcal{X} it induces, i.e., (𝒵𝒳)λ=𝒵λ(\mathcal{Z}\cap\mathcal{X})_{\lambda}=\mathcal{Z}_{\lambda} for \infty-compositions λ\lambda. We now define a construction in the opposite direction.

Let 𝒵\mathcal{Z} be a subset of 𝒳\mathcal{X}. Define a subset 𝒵+\mathcal{Z}^{+} of 𝒳+\mathcal{X}^{+} as follows. For an \infty-composition λ\lambda we put 𝒵λ+=𝒵λ\mathcal{Z}^{+}_{\lambda}=\mathcal{Z}_{\lambda}. Now suppose λ\lambda is a finite non-empty composition. Let λ+=λ{}\lambda^{+}=\lambda\cup\{\infty\}, and let f:λλ+f\colon\lambda\to\lambda^{+} be the standard inclusion. We define 𝒵λ+=αf(𝒵λ+)\mathcal{Z}^{+}_{\lambda}=\alpha_{f}(\mathcal{Z}_{\lambda^{+}}). Note that 𝒵+\mathcal{Z}^{+} extends 𝒵\mathcal{Z}, that is, 𝒵+𝒳=𝒵\mathcal{Z}^{+}\cap\mathcal{X}=\mathcal{Z}. We now study this construction in more detail.

Lemma 6.1.

Let 𝒵\mathcal{Z} be a C2 subset of 𝒳\mathcal{X} and let f:λμf\colon\lambda\to\mu be a map of finite partitions. Then αf(𝒵μ+)𝒵λ+\alpha_{f}(\mathcal{Z}^{+}_{\mu})\subset\mathcal{Z}^{+}_{\lambda}. If ff is a principal surjection then 𝒵μ+=αf1(𝒵λ+)\mathcal{Z}^{+}_{\mu}=\alpha_{f}^{-1}(\mathcal{Z}^{+}_{\lambda}).

Proof.

Let f+:λ+μ+f^{+}\colon\lambda^{+}\to\mu^{+} be the induced map. Note that 𝒳λ+=𝒳λ×𝐀1\mathcal{X}_{\lambda^{+}}=\mathcal{X}_{\lambda}\times\mathbf{A}^{1} and similarly for 𝒳μ+\mathcal{X}_{\mu^{+}}, and αf+=αf×id\alpha_{f^{+}}=\alpha_{f}\times\mathrm{id}. We have a commutative diagram

𝒳μ×𝐀1\textstyle{\mathcal{X}_{\mu}\times\mathbf{A}^{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}αf×id\scriptstyle{\alpha_{f}\times\mathrm{id}}q\scriptstyle{q}𝒳λ×𝐀1\textstyle{\mathcal{X}_{\lambda}\times\mathbf{A}^{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p\scriptstyle{p}𝒳μ\textstyle{\mathcal{X}_{\mu}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}αf\scriptstyle{\alpha_{f}}𝒳λ\textstyle{\mathcal{X}_{\lambda}}

where pp and qq are the projection maps. Now, suppose x𝒵μ+x\in\mathcal{Z}^{+}_{\mu}. Since 𝒵μ+=q(𝒵μ+)\mathcal{Z}^{+}_{\mu}=q(\mathcal{Z}_{\mu^{+}}), it follows that there is some y𝐀1y\in\mathbf{A}^{1} such that (x,y)𝒵μ+(x,y)\in\mathcal{Z}_{\mu^{+}}. By Proposition 3.4, we see that αf+(x,y)=(αf(x),y)\alpha_{f^{+}}(x,y)=(\alpha_{f}(x),y) belongs to 𝒵λ+\mathcal{Z}_{\lambda^{+}}. Thus p(αf(x),y)=αf(x)p(\alpha_{f}(x),y)=\alpha_{f}(x) belongs to 𝒵λ+=p(𝒵λ+)\mathcal{Z}^{+}_{\lambda}=p(\mathcal{Z}_{\lambda^{+}}).

Now suppose that ff is a principal surjection. Then f+f^{+} is as well. Suppose x𝒳μ+x\in\mathcal{X}^{+}_{\mu} and αf(x)𝒵λ+\alpha_{f}(x)\in\mathcal{Z}^{+}_{\lambda}. Then there is some y𝐀1y\in\mathbf{A}^{1} such that (αf(x),y)𝒵λ+(\alpha_{f}(x),y)\in\mathcal{Z}_{\lambda^{+}}. Thus αf+(x,y)𝒵λ+\alpha_{f^{+}}(x,y)\in\mathcal{Z}_{\lambda^{+}} and so (x,y)𝒵μ+(x,y)\in\mathcal{Z}_{\mu^{+}} since 𝒵\mathcal{Z} is C1. Hence x=q(x,y)x=q(x,y) belongs to 𝒵μ+\mathcal{Z}^{+}_{\mu}. This completes the proof. ∎

Proposition 6.2.

Let 𝒵\mathcal{Z} be a C2 subset of 𝒳\mathcal{X}. Then 𝒵+\mathcal{Z}^{+} is a C2 subset of 𝒳+\mathcal{X}^{+}. Moreover, if 𝒵\mathcal{Z}^{\prime} is any C2 subset of 𝒳+\mathcal{X}^{+} such that 𝒵𝒳\mathcal{Z}^{\prime}\cap\mathcal{X} contains 𝒵\mathcal{Z} then 𝒵\mathcal{Z}^{\prime} contains 𝒵+\mathcal{Z}^{+}.

Proof.

We first note that 𝒵\mathcal{Z} is C1. Indeed, suppose that f:λμf\colon\lambda\to\mu is a principal surjection. If λ\lambda is infinite, so is μ\mu, and then αf1(𝒵λ+)=𝒵μ+\alpha_{f}^{-1}(\mathcal{Z}^{+}_{\lambda})=\mathcal{Z}^{+}_{\mu} since 𝒵\mathcal{Z} is C1. If λ\lambda is finite, so is μ\mu, and then αf1(𝒵λ+)=𝒵μ+\alpha_{f}^{-1}(\mathcal{Z}^{+}_{\lambda})=\mathcal{Z}^{+}_{\mu} by Lemma 6.1.

Now suppose that f:λμf\colon\lambda\to\mu is an injection of generalized partitions. We claim that αf(𝒵μ+)𝒵λ+\alpha_{f}(\mathcal{Z}^{+}_{\mu})\subset\mathcal{Z}^{+}_{\lambda}. If λ\lambda is infinite then so is μ\mu, and this follows from Proposition 3.4. Thus suppose λ\lambda is finite. If μ\mu is finite then the claim follows from Lemma 6.1. Thus suppose that μ\mu is infinite. We can then factor ff as ghg\circ h where h:λλ+h\colon\lambda\to\lambda^{+} is the standard inclusion and g:λ+μg\colon\lambda^{+}\to\mu is any extension of ff (simply map λ+\infty\in\lambda^{+} to any element of μ1()\mu^{-1}(\infty)). We have

αf(𝒵μ+)=αh(αg(𝒵μ))αh(𝒵λ+)=𝒵λ+\alpha_{f}(\mathcal{Z}^{+}_{\mu})=\alpha_{h}(\alpha_{g}(\mathcal{Z}_{\mu}))\subset\alpha_{h}(\mathcal{Z}_{\lambda^{+}})=\mathcal{Z}^{+}_{\lambda}

where in the second step we used Proposition 3.4. This establishes the claim. Combined with the first paragraph, we see that 𝒵+\mathcal{Z}^{+} is closed under correspondences.

Now suppose x𝒳λ+x\in\mathcal{X}^{+}_{\lambda} is approximable by 𝒵+\mathcal{Z}^{+}. We show x𝒵λ+x\in\mathcal{Z}^{+}_{\lambda}. First suppose that λ\lambda is finite. Let NN be larger than all parts of λ\lambda. Let μ\mu be a generalized composition with λμ\langle\lambda\rangle\subset\langle\mu\rangle and μimin(N,λi)\mu_{i}\geq\min(N,\lambda_{i}) for all iλi\in\langle\lambda\rangle such that xαf(𝒵μ+)x\in\alpha_{f}(\mathcal{Z}^{+}_{\mu}) where f:λμf\colon\langle\lambda\rangle\to\langle\mu\rangle is the inclusion. By our choice of NN, we have μiλi\mu_{i}\geq\lambda_{i} for all ii, and so f:λμf\colon\lambda\to\mu is a map of generalizaed partitions. Thus αf\alpha_{f} carries 𝒵μ+\mathcal{Z}^{+}_{\mu} into 𝒵λ+\mathcal{Z}^{+}_{\lambda}, and so x𝒵λ+x\in\mathcal{Z}^{+}_{\lambda}, as required.

Now suppose λ\lambda is infinite. We claim that xx is in fact approximable by 𝒵\mathcal{Z}. Let NN be given. Since xx is NN-approximable by 𝒵+\mathcal{Z}^{+} we can find a generalized composition μ\mu and an inclusion f:λμf\colon\langle\lambda\rangle\to\langle\mu\rangle such that μf(i)min(N,λi)\mu_{f(i)}\geq\min(N,\lambda_{i}) for all iλi\in\langle\lambda\rangle and xαf(𝒵μ)x\in\alpha_{f}(\mathcal{Z}_{\mu}). If μ\mu is infinite then this shows that xx is NN-approximable by 𝒵\mathcal{Z}. Thus suppose μ\mu is finite. Let ν=μ{}\nu=\mu\cup\{\infty\} and let g:μνg\colon\mu\to\nu be the standard inclusion. By definition, 𝒵μ=αg(𝒵ν)\mathcal{Z}_{\mu}=\alpha_{g}(\mathcal{Z}_{\nu}), and so xαgf(𝒵ν)x\in\alpha_{g\circ f}(\mathcal{Z}_{\nu}). Since gfg\circ f is injective and νg(f(i))=μf(i)λi\nu_{g(f(i))}=\mu_{f(i)}\geq\lambda_{i} for all iλi\in\langle\lambda\rangle, this shows that xx is NN-approximable by 𝒵\mathcal{Z}. As this holds for all NN, the claim follows. Since 𝒵\mathcal{Z} is C2, we conclude that x𝒵λx\in\mathcal{Z}_{\lambda}, and thus 𝒵+\mathcal{Z}^{+} is C2.

Finally, suppose that 𝒵\mathcal{Z}^{\prime} is a C2 subset of 𝒳+\mathcal{X}^{+} such that 𝒵𝒳\mathcal{Z}^{\prime}\cap\mathcal{X} contains 𝒵+\mathcal{Z}^{+}. Let λ\lambda be a non-empty composition, and let f:λλ+f\colon\lambda\to\lambda^{+} be the standard inclusion. Then 𝒵λ+\mathcal{Z}^{\prime}_{\lambda^{+}} contains 𝒵λ+\mathcal{Z}_{\lambda^{+}} by assumption. Thus αf(𝒵λ+)\alpha_{f}(\mathcal{Z}^{\prime}_{\lambda^{+}}) contains αf(𝒵λ+)=𝒵λ+\alpha_{f}(\mathcal{Z}_{\lambda^{+}})=\mathcal{Z}^{+}_{\lambda}. Since 𝒵\mathcal{Z}^{\prime} is C2, we have αf(𝒵λ+)𝒵λ\alpha_{f}(\mathcal{Z}^{\prime}_{\lambda^{+}})\subset\mathcal{Z}^{\prime}_{\lambda}. Thus 𝒵λ\mathcal{Z}^{\prime}_{\lambda} contains 𝒵λ+\mathcal{Z}^{+}_{\lambda}. If λ\lambda is an \infty-composition then 𝒵λ\mathcal{Z}^{\prime}_{\lambda} contains 𝒵λ+=𝒵λ\mathcal{Z}^{+}_{\lambda}=\mathcal{Z}_{\lambda} by assumption. Thus 𝒵\mathcal{Z}^{\prime} contains 𝒵+\mathcal{Z}^{+}. This completes the proof. ∎

Proposition 6.3.

Let ZZ be a subset of 𝒳λ\mathcal{X}_{\lambda} for some \infty-composition λ\lambda and let 𝒵=Γλ(Z)\mathcal{Z}=\Gamma^{\circ}_{\lambda}(Z). Then for any non-empty generalized composition μ\mu, we have

𝒵μ+=f:μλαf(Z).\mathcal{Z}^{+}_{\mu}=\bigcup_{f\colon\mu\dashrightarrow\lambda}\alpha_{f}(Z).
Proof.

If μ\mu is infinite, this is simply the definition of Γλ\Gamma_{\lambda}^{\circ}. Suppose μ\mu is finite. Since 𝒵+\mathcal{Z}^{+} is closed under compositions and Z𝒵λ+Z\subset\mathcal{Z}^{+}_{\lambda}, we have αf(Z)𝒵μ\alpha_{f}(Z)\subset\mathcal{Z}_{\mu} for any f:μλf\colon\mu\dashrightarrow\lambda. Let g:μμ+g\colon\mu\to\mu^{+} be the standard inclusion. Then

𝒵μ+=αg(𝒵μ+)=f:μ+λαg(αf(Z))h:μλαh(Z).\mathcal{Z}^{+}_{\mu}=\alpha_{g}(\mathcal{Z}_{\mu^{+}})=\bigcup_{f\colon\mu^{+}\dashrightarrow\lambda}\alpha_{g}(\alpha_{f}(Z))\subset\bigcup_{h\colon\mu\dashrightarrow\lambda}\alpha_{h}(Z).

The first equality is the definition of 𝒵μ+\mathcal{Z}^{+}_{\mu}, the second follows from the definition of 𝒵μ+\mathcal{Z}_{\mu^{+}}, and the third follows by taking hh to be a composition fgf\circ g and applying Proposition 4.3. This completes the proof. ∎

Proposition 6.4.

Let λ\lambda be an \infty-composition and let ZZ be an End(λ)\operatorname{End}(\lambda)-stable Zariski closed subset of 𝒳λ\mathcal{X}_{\lambda}. Let f:μλf\colon\mu\dashrightarrow\lambda be a correspondence of generalized compositions. Suppose |μ||\mu| is larger than the sum of all finite parts of λ\lambda. Then αf(Z)\alpha_{f}(Z) is Zariski closed in 𝒳μ+\mathcal{X}^{+}_{\mu}.

Proof.

If ff is an injective map then there must be some iim(f)i\in\operatorname{im}(f) with λi=\lambda_{i}=\infty by our assumption on |μ||\mu|. With this observation in hand, the proof of Proposition 4.8 can be carried over verbatim. ∎

Proposition 6.5.

Suppose AA is noetherian, and let 𝒵\mathcal{Z} be a C3 subvariety of 𝒳\mathcal{X}. Then 𝒵μ+\mathcal{Z}^{+}_{\mu} is Zariski closed in 𝒳μ+\mathcal{X}^{+}_{\mu} for all generalized compositions μ\mu with |μ|0|\mu|\gg 0.

Proof.

Since 𝒵𝒵+\mathcal{Z}\mapsto\mathcal{Z}^{+} is compatible with finite unions, it suffices to treat the case where 𝒵\mathcal{Z} is irreducible. If 𝒵\mathcal{Z} is unbounded then 𝒵=𝒳W\mathcal{Z}=\mathcal{X}_{W^{\prime}} for some WWW^{\prime}\subset W, and one finds that 𝒵+=𝒳W+\mathcal{Z}^{+}=\mathcal{X}^{+}_{W^{\prime}}. In this case, 𝒵+\mathcal{Z}^{+} is a C3 subvariety of 𝒳+\mathcal{X}^{+}. Now suppose that 𝒵\mathcal{Z} is bounded. Then 𝒵=Γλ(Z)\mathcal{Z}=\Gamma_{\lambda}(Z) for some \infty-composition λ\lambda and some End(λ)\operatorname{End}(\lambda)-stable closed subvariety ZZ of 𝒳λ\mathcal{X}_{\lambda}. By Proposition 6.3, we have

𝒵μ+=f:μλαf(Z).\mathcal{Z}^{+}_{\mu}=\bigcup_{f\colon\mu\dashrightarrow\lambda}\alpha_{f}(Z).

Since μ\mu is finite, there are (up to isomorphism) only finitely many principal surjections ρλ\rho\to\lambda, and thus only finitely many choices for ff. Thus the above union is finite. If |μ||\mu| is sufficiently large, then αf(Z)\alpha_{f}(Z) is Zariski closed by Proposition 6.4. The result follows. ∎

7. The correspondence between 𝔛\mathfrak{X} and 𝒳\mathcal{X}

7.1. Setup

Regard 11^{\infty} as a composition on the index set 1=[]\langle 1^{\infty}\rangle=[\infty] taking the value 1 on each point. Suppose that λ\lambda is a generalized composition. A map f:1λf\colon 1^{\infty}\to\lambda is a function f:[]λf\colon[\infty]\to\langle\lambda\rangle such that #f1(i)λi\#f^{-1}(i)\leq\lambda_{i} for all iλi\in\langle\lambda\rangle. We say that such a map ff is a principal surjection if λi=#f1(i)\lambda_{i}=\#f^{-1}(i) for all iλi\in\langle\lambda\rangle. Given a map ff, we let αf:𝒳λ𝔛\alpha_{f}\colon\mathcal{X}_{\lambda}\to\mathfrak{X} be the associated map of schemes. If ff is a principal surjection then αf\alpha_{f} is a multi-diagonal map, and thus a closed immersion.

Let \mathfrak{Z} be a 𝔖big\mathfrak{S}^{\rm big}-subset of 𝔛fin\mathfrak{X}_{\mathrm{fin}}. We define a subset 𝒵=Φ()\mathcal{Z}=\Phi(\mathfrak{Z}) of 𝒳\mathcal{X} as follows. Given an \infty-composition λ\lambda, choose a principal surjection f:1λf\colon 1^{\infty}\rightarrowtail\lambda, and put 𝒵λ=αf1()\mathcal{Z}_{\lambda}=\alpha_{f}^{-1}(\mathfrak{Z}). Note that if g:1λg\colon 1^{\infty}\rightarrowtail\lambda is a second principal surjection then f=gσf=g\circ\sigma for some σ𝔖big\sigma\in\mathfrak{S}^{\rm big}, and so αf1()=αg1()\alpha_{f}^{-1}(\mathfrak{Z})=\alpha_{g}^{-1}(\mathfrak{Z}) since \mathfrak{Z} is 𝔖big\mathfrak{S}^{\rm big}-stable. Thus 𝒵\mathcal{Z} is well-defined.

Now let 𝒵\mathcal{Z} be a subset of 𝒳\mathcal{X}. We define a subset =Ψ(𝒵)\mathfrak{Z}=\Psi(\mathcal{Z}) of 𝔛fin\mathfrak{X}_{\mathrm{fin}} by =f:1λαf(𝒵λ)\mathfrak{Z}=\bigcup_{f\colon 1^{\infty}\rightarrowtail\lambda}\alpha_{f}(\mathcal{Z}_{\lambda}). Here the union is taken over all \infty-compositions λ\lambda and all principal surjections f:1λf\colon 1^{\infty}\rightarrowtail\lambda. It is clear that \mathfrak{Z} is 𝔖big\mathfrak{S}^{\rm big}-stable.

In the remainder of this section, we see that Φ\Phi and Ψ\Psi give bijective correspondence between certain classes of subsets of 𝒳\mathcal{X} and 𝔛\mathfrak{X}.

7.2. The correspondence on C1 subsets

The following result is the most basic of our correspondences between 𝔛\mathfrak{X} and 𝒳\mathcal{X}.

Proposition 7.1.

The constructions Φ\Phi and Ψ\Psi define mutually inverse bijections

{𝔖big-subsets of 𝔛fin}\textstyle{\{\text{$\mathfrak{S}^{\rm big}$-subsets of $\mathfrak{X}_{\mathrm{fin}}$}\}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}{C1 subsets of 𝒳}\textstyle{\{\text{C1 subsets of $\mathcal{X}$}\}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}
Proof.

Let \mathfrak{Z} be a 𝔖big\mathfrak{S}^{\rm big}-subset of 𝔛\mathfrak{X} and let 𝒵=Φ()\mathcal{Z}=\Phi(\mathfrak{Z}). We start by verifying that 𝒵\mathcal{Z} is indeed a C1 subset. Let f:λμf\colon\lambda\rightarrowtail\mu be a principal surjection. By definition of Φ\Phi, we have 𝒵μ=αg1()\mathcal{Z}_{\mu}=\alpha_{g}^{-1}(\mathfrak{Z}) for a principal surjection g:1μg\colon 1^{\infty}\rightarrowtail\mu. Clearly, gg factors through hh, so we can write g=fhg=fh for a principal surjection h:1λh\colon 1^{\infty}\rightarrowtail\lambda. Thus 𝒵μ=αg1()=αf1(αh1())=αf1(𝒵λ)\mathcal{Z}_{\mu}=\alpha_{g}^{-1}(\mathfrak{Z})=\alpha_{f}^{-1}(\alpha_{h}^{-1}(\mathfrak{Z}))=\alpha_{f}^{-1}(\mathcal{Z}_{\lambda}), as required.

We now verify that =Ψ(𝒵)\mathfrak{Z}=\Psi(\mathcal{Z}). Consider the map π:λ,f:1λ𝒳λ𝔛fin\pi\colon\coprod_{\lambda,f\colon 1^{\infty}\rightarrowtail\lambda}\mathcal{X}_{\lambda}\to\mathfrak{X}_{\mathrm{fin}}, where the disjoint union is taken over all isomorphism classes of \infty-compositions λ\lambda and all principal surjections f:1λf\colon 1^{\infty}\rightarrowtail\lambda (the map π\pi on the (λ,f)(\lambda,f)-coordinate is given by αf\alpha_{f}). Putting Z=λ,f:1λ𝒵λZ=\coprod_{\lambda,f\colon 1^{\infty}\rightarrowtail\lambda}\mathcal{Z}_{\lambda}, we have Z=π1()Z=\pi^{-1}(\mathfrak{Z}) and π(Z)=Ψ(𝒵)\pi(Z)=\Psi(\mathcal{Z}), by definition, and π(π1())=\pi(\pi^{-1}(\mathfrak{Z}))=\mathfrak{Z} since π\pi is surjective. Thus =π(π1())=π(Z)=Ψ(𝒵)\mathfrak{Z}=\pi(\pi^{-1}(\mathfrak{Z}))=\pi(Z)=\Psi(\mathcal{Z}).

Now let 𝒵\mathcal{Z} be a C1 subset of 𝒳\mathcal{X}, and put 𝒵=Φ(Ψ(𝒵))\mathcal{Z}^{\prime}=\Phi(\Psi(\mathcal{Z})). We must show 𝒵=𝒵\mathcal{Z}=\mathcal{Z}^{\prime}. Again putting Z=λ,f:1λ𝒵λZ=\coprod_{\lambda,f\colon 1^{\infty}\rightarrowtail\lambda}\mathcal{Z}_{\lambda} and reasoning as above, we see that Zπ1(π(Z))=π1(Ψ(𝒵))Z\subset\pi^{-1}(\pi(Z))=\pi^{-1}(\Psi(\mathcal{Z})). Moreover, π1(Ψ(𝒵))(λ,f)=Φ(Ψ(𝒵))λ\pi^{-1}(\Psi(\mathcal{Z}))_{(\lambda,f)}=\Phi(\Psi(\mathcal{Z}))_{\lambda}, by definition. Thus 𝒵𝒵\mathcal{Z}\subset\mathcal{Z}^{\prime}, so we simply have to establish the reverse inclusion. Let λ\lambda be an \infty-composition and let z𝒵[λ]z\in\mathcal{Z}^{\prime}_{[\lambda]} be given. Let f:1λf\colon 1^{\infty}\rightarrowtail\lambda be a principal surjection. Then αf(z)\alpha_{f}(z) belongs to Ψ(𝒵)\Psi(\mathcal{Z}) and has type λ\lambda (see §2.6). Thus, by the following lemma, we have αf(z)=αf(z)\alpha_{f}(z)=\alpha_{f^{\prime}}(z^{\prime}) for some z𝒵[λ]z^{\prime}\in\mathcal{Z}_{[\lambda]} and some principal surjection f:1λf^{\prime}\colon 1^{\infty}\rightarrowtail\lambda. Since zz and zz^{\prime} have distinct coordinates, we must have z=σzz=\sigma z^{\prime} for some automorphism of λ\lambda. Since 𝒵[λ]\mathcal{Z}_{[\lambda]} is stable under σ\sigma, we thus have z𝒵[λ]z\in\mathcal{Z}_{[\lambda]}. Thus 𝒵𝒵\mathcal{Z}^{\prime}\subset\mathcal{Z} by Proposition 3.3. ∎

Lemma 7.2.

Let 𝒵\mathcal{Z} be a C1-subset of 𝒳\mathcal{X}, let yΨ(𝒵)y\in\Psi(\mathcal{Z}), and let λ\lambda be an \infty-weighting. Suppose yy has type λ\lambda. Then there exists z𝒵[λ]z\in\mathcal{Z}_{[\lambda]} and a principal surjection f:1λf\colon 1^{\infty}\rightarrowtail\lambda such that y=αf(z)y=\alpha_{f}(z).

Proof.

By definition of Ψ\Psi, there is a principal surjection g:1μg\colon 1^{\infty}\rightarrowtail\mu such that y=αg(z)y=\alpha_{g}(z^{\prime}) for some z𝒵μz^{\prime}\in\mathcal{Z}_{\mu}. Since yy has type λ\lambda, we can find a principal surjection f:1λf\colon 1^{\infty}\rightarrowtail\lambda such that y=αf(z)y=\alpha_{f}(z) for some z𝒳[λ]z\in\mathcal{X}_{[\lambda]}. Since αf(z)=αg(z)\alpha_{f}(z)=\alpha_{g}(z^{\prime}) and zz has distinct coordinates, there is a unique principal surjection h:μλh\colon\mu\rightarrowtail\lambda such that z=αh(z)z^{\prime}=\alpha_{h}(z). Since 𝒵\mathcal{Z} is C1, we see that z𝒵[λ]z\in\mathcal{Z}_{[\lambda]}, which completes the proof. ∎

7.3. The correspondence on C2 subsets

For a subset 𝒵\mathcal{Z} of 𝒳\mathcal{X}, the subset Ψ(𝒵)\Psi(\mathcal{Z}) of 𝔛\mathfrak{X} is defined as a union. If 𝒵\mathcal{Z} is C2 it can also be described as an intersection:

Proposition 7.3.

Let 𝒵\mathcal{Z} be a C2 subset of 𝒳\mathcal{X}. Then

Ψ(𝒵)=𝔛finf:[n][]αf1(𝒵1n+),\Psi(\mathcal{Z})=\mathfrak{X}_{\mathrm{fin}}\cap\bigcap_{f\colon[n]\to[\infty]}\alpha_{f}^{-1}(\mathcal{Z}^{+}_{1^{n}}),

where the intersection is over all n𝐍n\in\mathbf{N} and all injections ff. In fact, for fixed NN, it suffices to intersect over nNn\geq N.

Proof.

Write \mathfrak{Z} for the right side in the above formula. Let xΨ(𝒵)x\in\Psi(\mathcal{Z}). Then x=αg(y)x=\alpha_{g}(y) for some \infty-partition λ\lambda, some y𝒵λy\in\mathcal{Z}_{\lambda}, and some principal surjection g:1λg\colon 1^{\infty}\rightarrowtail\lambda. Let f:[n][]f\colon[n]\to[\infty] be an injection. Then αf(x)=αf(αg(y))=αgf(y)\alpha_{f}(x)=\alpha_{f}(\alpha_{g}(y))=\alpha_{g\circ f}(y). Since gfg\circ f defines a map of generalized partitions 1nλ1^{n}\to\lambda, we see that αgf(y)𝒵1n\alpha_{g\circ f}(y)\in\mathcal{Z}_{1^{n}}. Thus xx\in\mathfrak{Z}.

Now suppose that xx\in\mathfrak{Z}. Since xx is finitary, we can write x=αg(y)x=\alpha_{g}(y) for some principal surjection g:1λg\colon 1^{\infty}\rightarrowtail\lambda and some y𝒳λy\in\mathcal{X}_{\lambda}. Now, for any injection f:[n][]f\colon[n]\to[\infty], we have αf(x)=αgf(y)𝒵1n\alpha_{f}(x)=\alpha_{g\circ f}(y)\in\mathcal{Z}_{1^{n}}. It is easy to see that any map of generalized partitions h:1nλh\colon 1^{n}\to\lambda can be factored as gfg\circ f for some injection f:[n][]f\colon[n]\to[\infty], and so we see that αh(y)𝒵1n\alpha_{h}(y)\in\mathcal{Z}_{1^{n}} for all such hh.

We claim that yy is approximable by 𝒵\mathcal{Z}. Indeed, let NN be given; assume NN is greater than all the finite parts of λ\lambda. Let μ\mu be the generalized partition obtained by replacing all infinite parts of λ\lambda with NN. Choose a principal surjection h:1nμh\colon 1^{n}\rightarrowtail\mu. Then the function hh defines a map of partitions h:1nλh\colon 1^{n}\to\lambda, and so αh(y)𝒵1n\alpha_{h}(y)\in\mathcal{Z}_{1^{n}} by the previous paragraph. Since hh is a principal surjection to μ\mu and 𝒵\mathcal{Z} is C1, it follows that y𝒵μy\in\mathcal{Z}_{\mu}. This verifies the claim. Since 𝒵\mathcal{Z} is C2, it follows that y𝒵λy\in\mathcal{Z}_{\lambda}. Thus x=αg(y)x=\alpha_{g}(y) belongs to Ψ(Z)\Psi(Z). This completes the proof. ∎

The following proposition is the next level of our correspondence between 𝔛\mathfrak{X} and 𝒳\mathcal{X}.

Proposition 7.4.

The constructions Φ\Phi and Ψ\Psi define mutually inverse bijections

{Π-closed 𝔖-subsets of 𝔛fin}\textstyle{\{\text{$\Pi$-closed $\mathfrak{S}$-subsets of $\mathfrak{X}_{\mathrm{fin}}$}\}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}{C2 subsets of 𝒳}\textstyle{\{\text{C2 subsets of $\mathcal{X}$}\}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}
Proof.

Suppose 𝒵\mathcal{Z} is a C2 subset of 𝒳\mathcal{X}. Given an injection f:[n][]f\colon[n]\to[\infty], the map αf:𝔛𝒳1n\alpha_{f}\colon\mathfrak{X}\to\mathcal{X}_{1^{n}} is continuous for the Π\Pi-topology on 𝔛\mathfrak{X} and the discrete topology on 𝒳1n\mathcal{X}_{1^{n}}. It follows that αf1(𝒵1n+)\alpha_{f}^{-1}(\mathcal{Z}^{+}_{1^{n}}) is Π\Pi-closed. Thus Ψ(𝒵)\Psi(\mathcal{Z}) is Π\Pi-closed by Proposition 7.3.

Conversely, suppose that \mathfrak{Z} is a Π\Pi-closed 𝔖\mathfrak{S}-subset of 𝔛fin\mathfrak{X}_{\mathrm{fin}}. Then \mathfrak{Z} is 𝔖big\mathfrak{S}^{\rm big}-stable (Proposition 2.4), and so 𝒵=Φ()\mathcal{Z}=\Phi(\mathfrak{Z}) is a well-defined C1 subset of 𝒳\mathcal{X} by Proposition 7.1. We now show that 𝒵\mathcal{Z} is C2. By Proposition 7.1, 𝒵\mathcal{Z} is C1. Now suppose that x𝒳λx\in\mathcal{X}_{\lambda} is approximable by 𝒵\mathcal{Z}, and let g:1λg\colon 1^{\infty}\rightarrowtail\lambda be a principal surjection. Then for a positive integer NN, there exists an \infty-composition μ\mu and an injective function f:λμf\colon\langle\lambda\rangle\to\langle\mu\rangle such that μf(i)min(λi,N)\mu_{f(i)}\geq\min(\lambda_{i},N) for all iλi\in\langle\lambda\rangle and y𝒵μy\in\mathcal{Z}_{\mu} such that z=αf(y)z=\alpha_{f}(y). Thus we may choose a principal projection h:1μh\colon 1^{\infty}\rightarrowtail\mu such that αh(y)\alpha_{h}(y) agrees with αg(x)\alpha_{g}(x) on [N][N]. Since NN is arbitrary and αh(y)\alpha_{h}(y)\in\mathfrak{Z}, we see that αg(x)\alpha_{g}(x) is in the Π\Pi-closure of \mathfrak{Z}. Since \mathfrak{Z} is Π\Pi-closed, αg(x)\alpha_{g}(x)\in\mathfrak{Z}. It follows that x𝒵x\in\mathcal{Z}, proving that 𝒵\mathcal{Z} is C2. ∎

7.4. The correspondence on C3 subvarieties

The following theorem is the final level of our correspondence between 𝔛\mathfrak{X} and 𝒳\mathcal{X}, and the most important.

Theorem 7.5.

Suppose AA is noetherian. Then the constructions Φ\Phi and Ψ\Psi induce mutually inverse bijections

{Zariski closed 𝔖-subsets of 𝔛fin}\textstyle{\{\text{Zariski closed $\mathfrak{S}$-subsets of $\mathfrak{X}_{\mathrm{fin}}$}\}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}{C3 subvarieties of 𝒳}\textstyle{\{\text{C3 subvarieties of $\mathcal{X}$}\}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}
Proof.

Suppose \mathfrak{Z} is a Zariski closed 𝔖\mathfrak{S}-subset of 𝔛fin\mathfrak{X}_{\mathrm{fin}}, and let 𝒵=Φ()\mathcal{Z}=\Phi(\mathfrak{Z}). Since \mathfrak{Z} is Π\Pi-closed (Proposition 2.3), it follows that 𝒵\mathcal{Z} is a C2 subset of 𝒳\mathcal{X} (Proposition 7.4). By definition, 𝒵λ=αf()\mathcal{Z}_{\lambda}=\alpha_{f}(\mathfrak{Z}) where f:1λf\colon 1^{\infty}\rightarrowtail\lambda is a principal surjection. Since αf\alpha_{f} is a map of schemes, it is continuous for the Zariski topology, and so 𝒵λ\mathcal{Z}_{\lambda} is a closed subset of 𝒳λ\mathcal{X}_{\lambda}. Thus 𝒵\mathcal{Z} is a C3 subvariety of 𝒳\mathcal{X}.

Conversely, suppose that 𝒵\mathcal{Z} is a C3 subvariety of 𝒳\mathcal{X}. Then Ψ(𝒵)\Psi(\mathcal{Z}) is C2 by Proposition 7.3. By Proposition 6.5, 𝒵1n+\mathcal{Z}^{+}_{1^{n}} is Zariski closed for n0n\gg 0. Thus by Proposition 7.3, we see that Ψ(𝒵)\Psi(\mathcal{Z}) is Zariski closed, as each αf1(𝒵1n+)\alpha_{f}^{-1}(\mathcal{Z}^{+}_{1^{n}}) is Zariski closed. This completes the proof. ∎

8. Symmetric subvarieties of 𝔛\mathfrak{X}

We assume AA is noetherian throughout this section

8.1. Classification of symmetric subvarieties of 𝔛\mathfrak{X}

For a subset ZZ of 𝒳μ\mathcal{X}_{\mu} define

Θμ(Z)=f:1μαf(Z),\Theta^{\circ}_{\mu}(Z)=\bigcup_{f\colon 1^{\infty}\to\mu}\alpha_{f}(Z),

where the union is over all maps; this is a subset of 𝔛\mathfrak{X}. It is clear that Θμ(Z)\Theta^{\circ}_{\mu}(Z) is stable under 𝔖big\mathfrak{S}^{\rm big}. We also put Θμ(Z)=Θμ(Ze)\Theta_{\mu}(Z)=\Theta^{\circ}_{\mu}(Z^{e}).

Proposition 8.1.

We have Θμ(Z)=Ψ(Γμ(Z))\Theta^{\circ}_{\mu}(Z)=\Psi(\Gamma_{\mu}^{\circ}(Z)).

Proof.

Applying the definitions, we have

Θμ(Z)=f:1μαf(Z),Ψ(Γμ(Z))=λg:1λh:λμαg(αh(Z)).\Theta^{\circ}_{\mu}(Z)=\bigcup_{f\colon 1^{\infty}\to\mu}\alpha_{f}(Z),\qquad\Psi(\Gamma_{\mu}^{\circ}(Z))=\bigcup_{\lambda}\bigcup_{g\colon 1^{\infty}\rightarrowtail\lambda}\bigcup_{h\colon\lambda\dashrightarrow\mu}\alpha_{g}(\alpha_{h}(Z)).

Since any map f:1μf\colon 1^{\infty}\to\mu can be written as a composite hgh\circ g with g:1λg\colon 1^{\infty}\rightarrowtail\lambda and h:λμh\colon\lambda\to\mu, we see that Θμ(Z)Ψ(Γμ(Z))\Theta^{\circ}_{\mu}(Z)\subset\Psi(\Gamma_{\mu}^{\circ}(Z)). Conversely, suppose xΨ(Γμ(Z))x\in\Psi(\Gamma_{\mu}^{\circ}(Z)), and let λ\lambda, gg, and hh be such that xαg(αh(Z))x\in\alpha_{g}(\alpha_{h}(Z)). Let h=(h1:ρλ,h2:ρμ)h=(h_{1}\colon\rho\rightarrowtail\lambda,h_{2}\colon\rho\to\mu). We have αh(Z)=αh11(αh2(Z))\alpha_{h}(Z)=\alpha_{h_{1}}^{-1}(\alpha_{h_{2}}(Z)) by definition. Thus we have x=αg(y)x=\alpha_{g}(y) for some y𝒳λy\in\mathcal{X}_{\lambda} with αh1(y)αh2(Z)\alpha_{h_{1}}(y)\in\alpha_{h_{2}}(Z), i.e., αh1(y)=αh2(z)\alpha_{h_{1}}(y)=\alpha_{h_{2}}(z) for some zZz\in Z. Consider the following diagram:

1\textstyle{1^{\infty}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}e\scriptstyle{e}g\scriptstyle{g}ρ\textstyle{\rho\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}h1\scriptstyle{h_{1}}h2\scriptstyle{h_{2}}λ\textstyle{\lambda}μ\textstyle{\mu}

One easily sees that one can find a principal surjection ee making the triangle commute. Let f=h2ef=h_{2}\circ e. We have

x=αg(y)=αe(αh1(y))=αe(αh2(z))=αf(z)x=\alpha_{g}(y)=\alpha_{e}(\alpha_{h_{1}}(y))=\alpha_{e}(\alpha_{h_{2}}(z))=\alpha_{f}(z)

and so xΘμ(Z)x\in\Theta^{\circ}_{\mu}(Z). This completes the proof. ∎

It follows from the above proposition that Θμ(Z)=Ψ(Γμ(Z))\Theta_{\mu}(Z)=\Psi(\Gamma_{\mu}(Z)). We can now classify the 𝔖\mathfrak{S}-irreducible subvarieties of 𝔛\mathfrak{X}:

Theorem 8.2.

We have a bijection

{pairs (μ,Z) with μ an -partition and Z an Aut(μ)-irreducible closed subset of 𝒳[μ]}{𝔖-irreducible Zariski closed bounded subsets of 𝔛}\left\{\parbox{180.00027pt}{pairs $(\mu,Z)$ with $\mu$ an $\infty$-partition and $Z$ an $\operatorname{Aut}(\mu)$-irreducible closed subset of $\mathcal{X}_{[\mu]}$}\right\}\leftrightarrow\left\{\parbox{120.00018pt}{$\mathfrak{S}$-irreducible Zariski closed bounded subsets of $\mathfrak{X}$}\right\}

given by (μ,Z)Θμ(Z)(\mu,Z)\mapsto\Theta_{\mu}(Z). Moreover, Θμ(Z)\Theta_{\mu}(Z) is the Zariski closure of

f:1μαf(Z),\bigcup_{f\colon 1^{\infty}\rightarrowtail\mu}\alpha_{f}(Z),

where ff varies over principal surjections.

Proof.

The first statement follows from Theorem 5.8, Theorem 7.5 and the previous proposition. For the second statement, let \mathfrak{Z} be the Zariski closure of

f:1μαf(Z),\bigcup_{f\colon 1^{\infty}\rightarrowtail\mu}\alpha_{f}(Z),

where ff varies over principal surjections. It is clear that Θμ(Z)\Theta_{\mu}(Z) contains \mathfrak{Z}. By Proposition 7.1, we see that Φ()λ\Phi(\mathfrak{Z})_{\lambda} contains ZZ. So by Theorem 4.9, we see that Φ()\Phi(\mathfrak{Z}) contains Γλ(Z)\Gamma_{\lambda}(Z). It follows that \mathfrak{Z} contains Θμ(Z)\Theta_{\mu}(Z) (Theorem 7.5), completing the proof. ∎

The inverse to the correspondence in the above theorem is described in the introduction. From the above theorem, we obtain a complete classification of closed 𝔖\mathfrak{S}-subsets of 𝔛\mathfrak{X}:

Corollary 8.3.

Let \mathfrak{Z} be a Zariski closed 𝔖\mathfrak{S}-subset of 𝔛=𝔛W\mathfrak{X}=\mathfrak{X}_{W}. Then there exists a closed subset WW^{\prime} of WW, finitely many \infty partitions μ1,μ2,,μn\mu^{1},\mu^{2},\ldots,\mu^{n}, and Aut(μi)\operatorname{Aut}(\mu^{i})-irreducible closed subvarieties ZiZ_{i} of 𝒳[μi]\mathcal{X}_{[\mu^{i}]} for 1in1\leq i\leq n such that

=𝒳W1inΘμi(Zi).\mathfrak{Z}=\mathcal{X}_{W^{\prime}}\cup\bigcup_{1\leq i\leq n}\Theta_{\mu^{i}}(Z_{i}).
Proof.

This follows from the theorem, Proposition 2.9, and Corollary 2.2

8.2. The locus 𝔛λ\mathfrak{X}_{\lambda}

Let λ\lambda be an \infty-partition. We let 𝔛[λ]\mathfrak{X}_{[\lambda]} be the set of all points of type λ\lambda (see §2.6). This locus can also be described as follows:

𝔛[λ]=f:1λαf(𝒳[λ]),\mathfrak{X}_{[\lambda]}=\bigcup_{f\colon 1^{\infty}\rightarrowtail\lambda}\alpha_{f}(\mathcal{X}_{[\lambda]}),

where the union is taken over all principal surjections ff. We define 𝔛λ\mathfrak{X}_{\lambda} to be the Zariski closure of 𝔛[λ]\mathfrak{X}_{[\lambda]}. These sets are perhaps the most fundamental examples of Zariski closed 𝔖\mathfrak{S}-subsets of 𝔛\mathfrak{X}. We now study them in more detail.

Proposition 8.4.

Let λ\lambda be an \infty-partition. Then

𝔛λ=Θλ(𝒳[λ])=μλ𝔛[μ].\mathfrak{X}_{\lambda}=\Theta_{\lambda}(\mathcal{X}_{[\lambda]})=\bigcup_{\mu\preceq\lambda}\mathfrak{X}_{[\mu]}.
Proof.

Clearly, 𝒳[λ]e=𝒳λ\mathcal{X}_{[\lambda]}^{e}=\mathcal{X}_{\lambda}. So by Proposition 8.1, we see that Θλ(𝒳[λ])=μλ𝔛[μ]\Theta_{\lambda}(\mathcal{X}_{[\lambda]})=\bigcup_{\mu\preceq\lambda}\mathfrak{X}_{[\mu]}, where 𝔛[μ]𝔛\mathfrak{X}_{[\mu]}\subset\mathfrak{X} denote the set of all points of type μ\mu. By Theorem 8.2, it follows that Θλ(𝒳[λ])\Theta_{\lambda}(\mathcal{X}_{[\lambda]}) is Zariski closed, and is the Zariski closure of 𝔛[λ]\mathfrak{X}_{[\lambda]}. Thus Θλ(𝒳[λ])=𝔛λ\Theta_{\lambda}(\mathcal{X}_{[\lambda]})=\mathfrak{X}_{\lambda}. ∎

Corollary 8.5.

Suppose WW is irreducible. Then 𝔛λ\mathfrak{X}_{\lambda} is 𝔖\mathfrak{S}-irreducible and corresponds to (λ,𝒳[λ])(\lambda,\mathcal{X}_{[\lambda]}) under the bijection in Theorem 8.2

We now investigate the defining equations for 𝔛λ\mathfrak{X}_{\lambda}. Let α\alpha be a (finite) partition and let TT be a tableau of shape α\alpha with distinct entries in [][\infty]. We define hTh_{T} to be the product of ξiξj\xi_{i}-\xi_{j} over those i,j[]i,j\in[\infty] that appear in TT in distinct rows. We write hαh_{\alpha} for any element hTh_{T} with TT of shape α\alpha. Note that if TT and TT^{\prime} have shape α\alpha then hTh_{T} and hTh_{T^{\prime}} are in the same 𝔖\mathfrak{S}-orbit, and so hαh_{\alpha} is well-defined up to the action of 𝔖\mathfrak{S}; in particular, hα\langle\!\langle h_{\alpha}\rangle\!\rangle is well-defined. If αβ\alpha\preceq\beta then hβhα\langle\!\langle h_{\beta}\rangle\!\rangle\subset\langle\!\langle h_{\alpha}\rangle\!\rangle. (Recall from §2.2 that S\langle\!\langle S\rangle\!\rangle denotes the 𝔖\mathfrak{S}-ideal of RR generated by SS.)

We define IλI_{\lambda} to be the ideal generated by the elements hαh_{\alpha} with αλ\alpha\npreceq\lambda. To be precise, let SS be the set of partitions α\alpha such that αλ\alpha\npreceq\lambda. Then Iλ=hααSI_{\lambda}=\langle\!\langle h_{\alpha}\rangle\!\rangle_{\alpha\in S}. Note that if αβ\alpha\preceq\beta are elements of SS then hβh_{\beta} already belongs to hα\langle\!\langle h_{\alpha}\rangle\!\rangle, and so can be omitted from the list of generators of IλI_{\lambda}. In other words, IλI_{\lambda} is generated by the hαh_{\alpha} with αS\alpha\in S a minimal element. By Proposition 3.1 and [SS, Proposition 2.1], there are finitely many minimal elements in SS. Suppose μ\mu is such a minimal element. Let \ell be the number of parts in λ\lambda, and let ee be the sum of the finite parts of λ\lambda. Then it is clear that μ\mu has at most +1\ell+1 parts, and each part is at most e+1e+1. In particular, we can find explicit upper bounds on the degree of generation and the number of generators for IλI_{\lambda}.

Theorem 8.6.

We have 𝔛λ=V(Iλ)\mathfrak{X}_{\lambda}=V(I_{\lambda}).

We need a few preliminary results. Let λ\lambda and μ\mu be generalized partitions. A λ\lambda-filling of μ\mu is a tableau of shape μ\mu with entries in positive integers such that the number ii appears at most λi\lambda_{i} times. We say that a filling is good if each number appears in at most one row.

Proposition 8.7.

μ\mu admits a good λ\lambda-filling if and only if μλ\mu\preceq\lambda.

Proof.

First suppose μλ\mu\preceq\lambda. Clearly, λ\lambda admits a good λ\lambda-filling TT, namely, the one obtained by filling the iith row of the λ\lambda-shaped Young diagram with ii. By definition, μ\mu is obtained from λ\lambda by combining some of the parts and then decreasing some of the parts. By doing the same operations on TT, we obtain a good λ\lambda-filling of μ\mu.

Conversely, suppose TT is a good λ\lambda-filling of μ\mu. Suppose ii appears λi\lambda^{\prime}_{i} times in TT. Then (λ1,λ2,)(\lambda^{\prime}_{1},\lambda^{\prime}_{2},\ldots) define a generalized partition (possibly after rearranging to make sure λiλi+1\lambda^{\prime}_{i}\geq\lambda^{\prime}_{i+1}). Since TT is good, we see that μ\mu can be obtained by combining some of the parts of λ\lambda^{\prime}. So μλ\mu\preceq\lambda^{\prime}. Since TT is a λ\lambda-filling, we have λiλi\lambda^{\prime}_{i}\leq\lambda_{i} for all ii. So λ\lambda^{\prime} can be obtained from λ\lambda by decreasing some of the parts. Thus λλ\lambda^{\prime}\preceq\lambda. We conclude that μλ\mu\preceq\lambda, finishing the proof. ∎

Proposition 8.8.

Let μ\mu and λ\lambda be generalized partitions. The following are equivalent:

  1. (a)

    μλ\mu\preceq\lambda

  2. (b)

    αμ\alpha\preceq\mu implies αλ\alpha\preceq\lambda for all finite partitions α\alpha.

Proof.

Clearly, (a) \Rightarrow (b). Conversely, suppose (b) holds. Pick a number ee larger than the sum of the finite parts of μ\mu and λ\lambda. Let d1,d2d_{1},d_{2} be the number of infinite parts of μ,λ\mu,\lambda. Let α\alpha be the finite partition obtained from μ\mu by decreasing all infinite parts of μ\mu to ee. Clearly, αμ\alpha\preceq\mu. By (b), we have αλ\alpha\preceq\lambda. By the previous proposition there exists a λ\lambda-good filling TT of α\alpha. Since the size of the iith row of TT is larger than the sum of finite parts of λ\lambda, for each id1i\leq d_{1}, we can find an element jij_{i} in the iith row of TT such that λji=\lambda_{j_{i}}=\infty. By goodness of TT, there is a well-defined injective correspondence ι:[d1][d2]\iota\colon[d_{1}]\to[d_{2}] given by ijii\mapsto j_{i}. Now define a λ\lambda filling TT^{\prime} of the Young diagram of shape μ\mu as follows:

  • For each id1i\leq d_{1}, fill row ii with ι(i)\iota(i).

  • For each i>d1i>d_{1} fill row ii of TT^{\prime} exactly as in TT.

It is clear that TT^{\prime} is a good λ\lambda-filling of μ\mu. Thus by the previous proposition, we have μλ\mu\preceq\lambda. This completes the proof of the first assertion. ∎

Proposition 8.9.

We have V(hα)=αλ𝔛[λ]V(\langle\!\langle h_{\alpha}\rangle\!\rangle)=\bigcup_{\alpha\npreceq\lambda}\mathfrak{X}_{[\lambda]}.

Proof.

Let x𝔛x\in\mathfrak{X} be a finitary point of type λ\lambda, and let 𝒰\mathcal{U} be the partition of [][\infty] induced by xx, as in §2.6. Let UiU_{i} be the part of 𝒰\mathcal{U} of size λi\lambda_{i}. We claim that xV(hα)x\in V(\langle\!\langle h_{\alpha}\rangle\!\rangle) if and only if α\alpha does not have a good λ\lambda-filling. To see this, first suppose that xV(hα)x\notin V(\langle\!\langle h_{\alpha}\rangle\!\rangle). Then there exists a tableau TT of shape α\alpha, and a KK-point yy above xx such that hT(y)0h_{T}(y)\neq 0. Let Ti,jT_{i,j} be the jjth entry in row ii of TT, and let 1ni,j(λ)1\leq n_{i,j}\leq\ell(\lambda) be the unique number such that Ti,jUni,jT_{i,j}\in U_{n_{i,j}}. Let TT^{\prime} be the tableau of shape α\alpha whose jjth entry in row ii is ni,jn_{i,j}. Since hT(y)0h_{T}(y)\neq 0, we conclude that TT^{\prime} is a good λ\lambda-filling of α\alpha. Conversely, suppose that there exists a good λ\lambda-filling TT^{\prime} of α\alpha. Suppose ii appears μi\mu_{i} times in TT^{\prime}. We know that μiλi\mu_{i}\leq\lambda_{i}, and so we can choose a subset UiU^{\prime}_{i} of UiU_{i} of size μi\mu_{i}. For each ii, we replace instances of ii with distinct elements of UiU^{\prime}_{i} in TT^{\prime} to obtain a tableau TT. Goodness of TT shows that hT(y)0h_{T}(y)\neq 0 for any KK-point yy above xx. This proves the claim.

By the claim and Proposition 8.7, we conclude that 𝔛[λ]V(hα)αλ\mathfrak{X}_{[\lambda]}\subset V(\langle\!\langle h_{\alpha}\rangle\!\rangle)\iff\alpha\npreceq\lambda. The result now follows immediately. ∎

Proof of Theorem 8.6.

We have already seen that 𝔛λ=μλ𝔛[μ]\mathfrak{X}_{\lambda}=\bigcup_{\mu\preceq\lambda}\mathfrak{X}_{[\mu]}. Thus it suffices to show that V(Iλ)=μλ𝔛[μ]V(I_{\lambda})=\bigcup_{\mu\preceq\lambda}\mathfrak{X}_{[\mu]}. We have

V(Iλ)=αλV(hα)=αλαμ𝔛[μ].V(I_{\lambda})=\bigcap_{\alpha\npreceq\lambda}V(\langle\!\langle h_{\alpha}\rangle\!\rangle)=\bigcap_{\alpha\npreceq\lambda}\bigcup_{\alpha\npreceq\mu}\mathfrak{X}_{[\mu]}.

We thus see that V(Iλ)=μS𝔛[μ]V(I_{\lambda})=\bigcup_{\mu\in S}\mathfrak{X}_{[\mu]} where SS is the set of partitions μ\mu for which αλ\alpha\npreceq\lambda implies αμ\alpha\npreceq\mu. Taking the contrapositive, we see that μS\mu\in S if and only if αμ\alpha\preceq\mu implies αλ\alpha\preceq\lambda. By Proposition 8.8, this condition is equivalent to μλ\mu\preceq\lambda. Thus SS consists of those μ\mu for which μλ\mu\preceq\lambda, and so V(Iλ)=μλ𝔛[μ]V(I_{\lambda})=\bigcup_{\mu\preceq\lambda}\mathfrak{X}_{[\mu]}. ∎

8.3. Example A

Suppose λ=(,n)\lambda=(\infty,n) with 1n<1\leq n<\infty. Then the set of minimal partitions which are not below λ\lambda is given by {(1,1,1),(n+1,n+1)}\{(1,1,1),(n+1,n+1)\}. Thus the ideal IλI_{\lambda} is generated by the 𝔖\mathfrak{S}-orbits of the polynomials:

h1\displaystyle h_{1} =(ξ1ξ2)(ξ2ξ3)(ξ3ξ1)\displaystyle=(\xi_{1}-\xi_{2})(\xi_{2}-\xi_{3})(\xi_{3}-\xi_{1})
h2\displaystyle h_{2} =0k,ln(ξn+1kξ2n+2l),\displaystyle=\prod_{0\leq k,l\leq n}(\xi_{n+1-k}-\xi_{2n+2-l}),

and V(Iλ)V(I_{\lambda}) consists of points of types (,a)(\infty,a) satisfying 0an0\leq a\leq n. In the case when n=1n=1, we see that the degrees of h1,h2h_{1},h_{2} are strictly larger than 2, but the ideal generated by the orbits of h1h_{1} and (ξ1ξ2)(ξ3ξ4)(\xi_{1}-\xi_{2})(\xi_{3}-\xi_{4}) also cuts out points of types ()(\infty) and (,1)(\infty,1). Since IλI_{\lambda} does not contain (ξ1ξ2)(ξ3ξ4)(\xi_{1}-\xi_{2})(\xi_{3}-\xi_{4}) for degree reasons, we conclude that I(,1)I_{(\infty,1)} is not radical.

8.4. Example B

Suppose λ=(,,2,1)\lambda=(\infty,\infty,2,1). Then the set of minimal partitions which are not below λ\lambda is given by

{(1,1,1,1,1),(2,2,2,2),(3,3,3,1),(4,4,4)}.\{(1,1,1,1,1),(2,2,2,2),(3,3,3,1),(4,4,4)\}.

Thus the ideal IλI_{\lambda} is generated by the 𝔖\mathfrak{S}-orbits of the following polynomials:

h1\displaystyle h_{1} =1i<j5(ξiξj)\displaystyle=\prod_{1\leq i<j\leq 5}(\xi_{i}-\xi_{j})
h2\displaystyle h_{2} =1i<j40k,l1(ξ2ikξ2jl)\displaystyle=\prod_{1\leq i<j\leq 4}\prod_{0\leq k,l\leq 1}(\xi_{2i-k}-\xi_{2j-l})
h3\displaystyle h_{3} =(1i9(ξiξ10))1i<j30k,l2(ξ3ikξ3jl)\displaystyle=\left(\prod_{1\leq i\leq 9}(\xi_{i}-\xi_{10})\right)\prod_{1\leq i<j\leq 3}\prod_{0\leq k,l\leq 2}(\xi_{3i-k}-\xi_{3j-l})
h4\displaystyle h_{4} =1i<j30k,l3(ξ4ikξ4jl).\displaystyle=\prod_{1\leq i<j\leq 3}\prod_{0\leq k,l\leq 3}(\xi_{4i-k}-\xi_{4j-l}).

Moreover, V(Iλ)V(I_{\lambda}) consists of points of types (,a,b,c)(\infty,a,b,c) satisfying b+c3b+c\leq 3 and 0cba0\leq c\leq b\leq a.

8.5. Equations for general irreducible subvarieties

We now determine equations for general 𝔖\mathfrak{S}-irreducible subvarieties of 𝔛\mathfrak{X}. Fix an \infty-partition λ\lambda and an Aut(λ)\operatorname{Aut}(\lambda)-irreducible closed subvariety ZZ of 𝒳[λ]\mathcal{X}_{[\lambda]}. Let ee be the sum of the finite parts of λ\lambda. For a generalized partition μ\mu, we let μ\mu^{-} to be the finite partition obtained by reducing all parts larger than e+1e+1 to e+1e+1, and we let μs\mu^{s} to be the unique maximal partition such that (μs)=μ(\mu^{s})^{-}=\mu^{-}. Then Λλ{μ:μλ}\Lambda_{\leq\lambda}^{-}\coloneq\{\mu^{-}\colon\mu\leq\lambda\} is a finite set. For each μΛλ\mu\in\Lambda_{\leq\lambda}^{-}, fix a tableau TμT_{\mu} of shape μ\mu with distinct entries in [][\infty] and a map ιμ:A[t1,,t(μs)]R\iota_{\mu}\colon A[t_{1},\ldots,t_{\ell(\mu^{s})}]\to R such that for each ii, ιμ(ti)ξj\iota_{\mu}(t_{i})\in\xi_{j} for some jj in row ii of TμT_{\mu}. We set

Iλ(Z)=Iλ+μΛλhTμιμ(I(𝒵μ))I_{\lambda}(Z)=I_{\lambda}+\sum_{\mu\in\Lambda_{\leq\lambda}^{-}}\langle\!\langle h_{T_{\mu}}\iota_{\mu}(I(\mathcal{Z}_{\mu}))\rangle\!\rangle

where 𝒵=Γλ(Z)+\mathcal{Z}=\Gamma_{\lambda}(Z)^{+}. Our main result is the following:

Theorem 8.10.

We have Θλ(Z)=V(Iλ(Z))\Theta_{\lambda}(Z)=V(I_{\lambda}(Z)).

Lemma 8.11.

Let μλ\mu\preceq\lambda be an \infty-partition. The map αg:Γλ(Z)μΓλ(Z)μ+\alpha_{g}\colon\Gamma_{\lambda}(Z)_{\mu}\to\Gamma_{\lambda}(Z)^{+}_{\mu^{-}} induced by the natural inclusion g:μμg\colon\mu^{-}\to\mu is an isomorphism.

Proof.

Since ff is an isomorphism on the underlying set, we see that αg\alpha_{g} is injective. To prove surjectivity, it suffices to treat the case when μ=μs\mu=\mu^{s}. For that suppose zΓλ(Z)μ+z\in\Gamma_{\lambda}(Z)^{+}_{\mu^{-}}, then there exists a good correspondence f=(f1:ρμ,f2:ρλ)f=(f_{1}\colon\rho\rightarrowtail\mu^{-},f_{2}\colon\rho\to\lambda) such that zαf(Ze)z\in\alpha_{f}(Z^{e}) (We note here that Proposition 4.6 is valid if μ\mu is finite). But since ff is good and μ=μs\mu=\mu^{s}, it induces a correspondence fs=(f1s:ρsμ,f2s:ρsλ)f^{s}=(f^{s}_{1}\colon\rho^{s}\rightarrowtail\mu,f^{s}_{2}\colon\rho^{s}\to\lambda). Since the underlying maps for fif_{i} is same as that of fisf^{s}_{i} for i=1,2i=1,2, we see that zαfs(Ze)z\in\alpha_{f^{s}}(Z^{e}). This shows that zΓλ(Z)μz\in\Gamma_{\lambda}(Z)_{\mu}, completing the proof. ∎

Proof of Theorem 8.10.

First suppose xV(Iλ(Z))x\in V(I_{\lambda}(Z)). Since Iλ(Z)I_{\lambda}(Z) contains IλI_{\lambda}, we see by Theorem 8.6 that xx is of type μ\mu for some μλ\mu\preceq\lambda. Let yy be a KK-point above xx, we claim that yy is a KK-point of Θλ(Z)\Theta_{\lambda}(Z). Since yy is of type μ\mu, by replacing yy by σy\sigma y for some σ𝔖\sigma\in\mathfrak{S}, we can assume that hTμ(y)0h_{T_{\mu^{-}}}(y)\neq 0. Thus we see that ιμ(f)(y)=0\iota_{\mu^{-}}(f)(y)=0 for each fI(𝒵μ)f\in I(\mathcal{Z}_{\mu^{-}}). By the previous lemma, we have 𝒵μ=𝒵μ\mathcal{Z}_{\mu^{-}}=\mathcal{Z}_{\mu}. Thus ιμ(f)(y)=0\iota_{\mu^{-}}(f)(y)=0 for each fI(𝒵μ)f\in I(\mathcal{Z}_{\mu}). We conclude that yy is a KK-point of αg(𝒵μ)\alpha_{g}(\mathcal{Z}_{\mu}) where g:1μg\colon 1^{\infty}\to\mu is any principal surjection such that g(j)=ig(j)=i for each jj in row ii of TμT_{\mu^{-}}. Since αg(𝒵μ)Ψ(𝒵)=Θλ(Z)\alpha_{g}(\mathcal{Z}_{\mu})\subset\Psi(\mathcal{Z})=\Theta_{\lambda}(Z), it follows that yy is a KK-point of Θλ(Z)\Theta_{\lambda}(Z). Thus xΘλ(Z)x\in\Theta_{\lambda}(Z).

Conversely, suppose xΘλ(Z)x\in\Theta_{\lambda}(Z), and let yy be a KK-point above xx. Since Θλ(Z)\Theta_{\lambda}(Z) is contained in 𝒳λ\mathcal{X}_{\lambda}, we see by Theorem 8.6 that f(y)=0f(y)=0 for all yIλy\in I_{\lambda}. Now suppose hTμ(σy)h_{T_{\mu}}(\sigma y) is nonzero for some μΛλ\mu\in\Lambda_{\leq\lambda}^{-} and σ𝔖\sigma\in\mathfrak{S}. Then σy\sigma y is of type ρ\rho such that ρ=μ\rho^{-}=\mu, and there is a principal surjection g:1ρg\colon 1^{\infty}\rightarrowtail\rho such that g(j)=ig(j)=i for each jj in row ii of TμT_{\mu} and yαg(𝒵ρ)y\in\alpha_{g}(\mathcal{Z}_{\rho}). It follows that ιμ(f)(y)=0\iota_{\mu}(f)(y)=0 for each fI(𝒵ρ)f\in I(\mathcal{Z}_{\rho}). Since I(𝒵ρ)I(𝒵μ)I(\mathcal{Z}_{\rho})\supset I(\mathcal{Z}_{\mu}), we see that yy is a KK-point of V(Iλ(Z))V(I_{\lambda}(Z)). Thus xx is in V(Iλ(Z))V(I_{\lambda}(Z)), completing the proof. ∎

8.6. Example C

Suppose λ=(,)\lambda=(\infty,\infty), and let Z𝒳[λ]Z\subset\mathcal{X}_{[\lambda]} be the 0-dimensional closed subvariety {(0,1),(1,0)}\{(0,1),(1,0)\}. Clearly, Aut(λ)=𝔖2\operatorname{Aut}(\lambda)=\mathfrak{S}_{2} and ZZ is Aut(λ)\operatorname{Aut}(\lambda)-irreducible. We have e=0e=0, and so Λλ={(1,1),(1)}\Lambda_{\leq\lambda}^{-}=\{(1,1),(1)\}. It is easy to check that 𝒵(1,1)={(0,1),(1,0),(0,0),(1,1)}\mathcal{Z}_{(1,1)}=\{(0,1),(1,0),(0,0),(1,1)\} and 𝒵(1)={(0),(1)}\mathcal{Z}_{(1)}=\{(0),(1)\}. These are cut out by t1(t11),t2(t21)\langle t_{1}(t_{1}-1),t_{2}(t_{2}-1)\rangle and t1(t11)\langle t_{1}(t_{1}-1)\rangle respectively. Fixing T(1,1)T_{(1,1)} with labels 1,2 and T(1)T_{(1)} with label 1, we conclude that Iλ(Z)I_{\lambda(Z)} is generated by the 𝔖\mathfrak{S}-orbits of

(ξ1ξ2)(ξ2ξ3)(ξ3ξ1)\displaystyle(\xi_{1}-\xi_{2})(\xi_{2}-\xi_{3})(\xi_{3}-\xi_{1})
(ξ1ξ2)ξ1(ξ11)\displaystyle(\xi_{1}-\xi_{2})\xi_{1}(\xi_{1}-1)
(ξ1ξ2)ξ2(ξ21)\displaystyle(\xi_{1}-\xi_{2})\xi_{2}(\xi_{2}-1)
ξ1(ξ11).\displaystyle\xi_{1}(\xi_{1}-1).

In fact, the first three elements are redundant, as they belong to the 𝔖\mathfrak{S}-ideal generated by the fourth element. These equations implies that Θλ(Z)\Theta_{\lambda}(Z) consists of points whose coordinates take the values 0 or 1.

8.7. Containments

We have parametrized the 𝔖\mathfrak{S}-irreducible closed subsets of 𝔛\mathfrak{X} in terms of pairs (λ,Z)(\lambda,Z). We now show how to determine containment on the 𝔛\mathfrak{X} side in terms of this parametrization.

Proposition 8.12.

Let Z1Z_{1} be an Aut(μ)\operatorname{Aut}(\mu)-irreducible closed subvariety of 𝒳[μ]\mathcal{X}_{[\mu]}, and let Z2Z_{2} be an Aut(λ)\operatorname{Aut}(\lambda)-irreducible closed subvariety of 𝒳[λ]\mathcal{X}_{[\lambda]}. Then Θμ(Z1)Θλ(Z2)\Theta_{\mu}(Z_{1})\subset\Theta_{\lambda}(Z_{2}) if and only if

Z1f:μλf is goodαf(Z2e).Z_{1}\subset\bigcup_{\begin{subarray}{c}f\colon\mu\dashrightarrow\lambda\\ \text{$f$ is good}\end{subarray}}\alpha_{f}(Z_{2}^{e}).
Proof.

By Theorem 7.5, the containment Θμ(Z1)Θλ(Z2)\Theta_{\mu}(Z_{1})\subset\Theta_{\lambda}(Z_{2}) is equivalent to the containment Γμ(Z1e)Γλ(Z2e)\Gamma^{\circ}_{\mu}(Z_{1}^{e})\subset\Gamma^{\circ}_{\lambda}(Z_{2}^{e}); by Theorem 4.9, this is equivalent to the containment Z1eΓλ(Z2e)μZ_{1}^{e}\subset\Gamma^{\circ}_{\lambda}(Z_{2}^{e})_{\mu}. By Proposition 4.6, this is equivalent to the containment in the statement of the proposition. ∎

9. The support of a module

We assume AA is noetherian throughout this section

9.1. Induction of subvarieties

Given a subset \mathfrak{Z} of 𝔛\mathfrak{X}, we let I()=σ𝔖σ\mathrm{I}(\mathfrak{Z})=\bigcup_{\sigma\in\mathfrak{S}}\sigma\mathfrak{Z} and let I¯()\overline{\mathrm{I}}(\mathfrak{Z}) be the Zariski closure of I()\mathrm{I}(\mathfrak{Z}). The goal of §9.1 is to show that I()\mathrm{I}(\mathfrak{Z}) is not so far from I¯()\overline{\mathrm{I}}(\mathfrak{Z}) in many cases.

To make this precise, we introduce some notation. Suppose \mathfrak{Z} is an 𝔖\mathfrak{S}-irreducible bounded closed 𝔖\mathfrak{S}-subvariety of 𝔛\mathfrak{X} of type λ\lambda. We define top=𝔛[λ]\mathfrak{Z}^{\mathrm{top}}=\mathfrak{Z}\cap\mathfrak{X}_{[\lambda]} to be the subset of \mathfrak{Z} consisting of points of type λ\lambda. It is Zariski dense in \mathfrak{Z}. For a general bounded closed 𝔖\mathfrak{S}-subvariety \mathfrak{Z} of 𝔛\mathfrak{X}, write =1r\mathfrak{Z}=\mathfrak{Z}_{1}\cup\cdots\cup\mathfrak{Z}_{r} where the i\mathfrak{Z}_{i} are the 𝔖\mathfrak{S}-irreducible components of \mathfrak{Z}, and put top=1toprtop\mathfrak{Z}^{\mathrm{top}}=\mathfrak{Z}_{1}^{\mathrm{top}}\cup\cdots\cup\mathfrak{Z}_{r}^{\mathrm{top}}. Again, this is Zariski dense in \mathfrak{Z}.

Proposition 9.1.

Let 𝒰\mathcal{U} be a partition of [][\infty] and let ZZ be a Zariski closed subset of 𝔛𝒰\mathfrak{X}_{\mathcal{U}}. Then I¯(Z)topI(Z)\overline{\mathrm{I}}(Z)^{\mathrm{top}}\subset\mathrm{I}(Z).

Proof.

It suffices to treat the case where ZZ is irreducible. Let 𝒱\mathcal{V} be the partition of [][\infty] defined as follows: ii and jj belong to the same part if and only if xi=xjx_{i}=x_{j} for all xZx\in Z. Then we have Z𝔛𝒱𝔛𝒰Z\subset\mathfrak{X}_{\mathcal{V}}\subset\mathfrak{X}_{\mathcal{U}}, and furthermore, Z𝔛[𝒱]Z\cap\mathfrak{X}_{[\mathcal{V}]} is non-empty. Let λ\lambda be the partition of \infty associated to 𝒱\mathcal{V}, and identify 𝒳λ\mathcal{X}_{\lambda} with 𝔛𝒱\mathfrak{X}_{\mathcal{V}} so that we can regard ZZ as a closed subvariety of 𝒳λ\mathcal{X}_{\lambda}. Let 𝒵=Γλ(Z)\mathcal{Z}=\Gamma_{\lambda}(Z) be the C3 subvariety generated by ZZ and let =Ψ(𝒵)\mathfrak{Z}=\Psi(\mathcal{Z}) be the corresponding 𝔖\mathfrak{S}-stable Zariski closed subset of 𝔛\mathfrak{X}. Then, by the nature of the correspondence, we have [𝒱]=𝒵[λ]\mathfrak{Z}_{[\mathcal{V}]}=\mathcal{Z}_{[\lambda]}, which by Proposition 4.7, is identified with Ze𝔛[𝒱]=σAut(𝒱)σ(Z𝔛[𝒱])Z^{e}\cap\mathfrak{X}_{[\mathcal{V}]}=\bigcup_{\sigma\in\operatorname{Aut}(\mathcal{V})}\sigma(Z\cap\mathfrak{X}_{[\mathcal{V}]}). We have

top=𝔛[λ]=σ𝔖σ(𝔛[𝒱])\mathfrak{Z}^{\mathrm{top}}=\mathfrak{Z}\cap\mathfrak{X}_{[\lambda]}=\bigcup_{\sigma\in\mathfrak{S}}\sigma(\mathfrak{Z}\cap\mathfrak{X}_{[\mathcal{V}]})

from which it follows that top=I(Z𝔛[𝒱])I(Z)\mathfrak{Z}^{\mathrm{top}}=\mathrm{I}(Z\cap\mathfrak{X}_{[\mathcal{V}]})\subset\mathrm{I}(Z). Since \mathfrak{Z} is 𝔖\mathfrak{S}-stable and Zariski closed and contains ZZ, we have I¯(Z)\overline{\mathrm{I}}(Z)\subset\mathfrak{Z}. On the other hand, the left side contains top\mathfrak{Z}^{\mathrm{top}}, which is dense in \mathfrak{Z}, and so we have equality. We thus conclude that I¯(Z)topI(Z)\overline{\mathrm{I}}(Z)^{\mathrm{top}}\subset\mathrm{I}(Z). ∎

Let 𝔖n\mathfrak{S}_{\geq n} be the subgroup of 𝔖\mathfrak{S} fixing each of 1,,n11,\ldots,n-1. Given an 𝔖n\mathfrak{S}_{\geq n}-stable subset \mathfrak{Z} of 𝔛\mathfrak{X}, we think of I()\mathrm{I}(\mathfrak{Z}) as a kind of induction of this set to 𝔖\mathfrak{S}. The following is the main result of §9.1.

Proposition 9.2.

Let \mathfrak{Z} be a bounded 𝔖n\mathfrak{S}_{\geq n}-stable Zariski closed subset of 𝔛\mathfrak{X}. Then I¯()\overline{\mathrm{I}}(\mathfrak{Z}) is bounded and I¯()topI()\overline{\mathrm{I}}(\mathfrak{Z})^{\mathrm{top}}\subset\mathrm{I}(\mathfrak{Z}).

Proof.

It suffices to treat the case where \mathfrak{Z} is irreducible. We can identify 𝔖n\mathfrak{S}_{\geq n} with 𝔖\mathfrak{S}, under which 𝔛W\mathfrak{X}_{W} is identified with 𝔛W×𝐀n1\mathfrak{X}_{W\times\mathbf{A}^{n-1}}. Applying our theory, we see that there is some 𝒰\mathcal{U} (a partition of [][\infty]) such that, putting Z=𝔛𝒰Z=\mathfrak{Z}\cap\mathfrak{X}_{\mathcal{U}}, we have that \mathfrak{Z} is the Zariski closure of σ𝔖nσZ\bigcup_{\sigma\in\mathfrak{S}_{\geq n}}\sigma Z. It follows that I¯()=I¯(Z)\overline{\mathrm{I}}(\mathfrak{Z})=\overline{\mathrm{I}}(Z); as this is contained in 𝔛λ\mathfrak{X}_{\lambda}, where λ\lambda is the partition of \infty associated to 𝒰\mathcal{U}, it is bounded. By the previous proposition, we have I¯()topI(Z)\overline{\mathrm{I}}(\mathfrak{Z})^{\mathrm{top}}\subset\mathrm{I}(Z). Since ZZ\subset\mathfrak{Z}, we have I(Z)I()\mathrm{I}(Z)\subset\mathrm{I}(\mathfrak{Z}), which completes the proof. ∎

We now give some examples to illustrate some different phenomena that can occur.

Example 9.3.

Let xx be the point (1,0,0,)(1,0,0,\ldots), and let ={x}\mathfrak{Z}=\{x\}. Then \mathfrak{Z} is a closed 𝔖2\mathfrak{S}_{\geq 2}-subvariety of 𝔛\mathfrak{X}. The set I()\mathrm{I}(\mathfrak{Z}) is just the 𝔖\mathfrak{S}-orbit of \mathfrak{Z}. The Zariski closure of this set contains one additional point, namely the origin; see Example 1.4. Thus I()\mathrm{I}(\mathfrak{Z}) is not Zariski closed, but I¯()\overline{\mathrm{I}}(\mathfrak{Z}) is the Π\Pi-closure of I()\mathrm{I}(\mathfrak{Z}). ∎

Example 9.4.

Let \mathfrak{Z} be the closed 𝔖2\mathfrak{S}_{\geq 2}-subvariety of 𝔛\mathfrak{X} defined by the equations ξ1ξi=1\xi_{1}\xi_{i}=1 for i2i\geq 2. Thus \mathfrak{Z} consists of all points of the form (a1,a,a,)(a^{-1},a,a,\ldots) with aa non-zero. The Π\Pi-closure of I()\mathrm{I}(\mathfrak{Z}) contains all points of the form (a,a,)(a,a,\ldots) with aa non-zero. We thus see that I¯()\overline{\mathrm{I}}(\mathfrak{Z}) contains the origin. But this does not belong to the Π\Pi-closure of I()\mathrm{I}(\mathfrak{Z}), since no point in I()\mathrm{I}(\mathfrak{Z}) has zero as a coordinate. Thus I¯()\overline{\mathrm{I}}(\mathfrak{Z}) is strictly larger than the Π\Pi-closure of I()\mathrm{I}(\mathfrak{Z}). ∎

Example 9.5.

Take W=𝐀1W=\mathbf{A}^{1}, and let η\eta be the coordinate on WW. Let \mathfrak{Z} be the (unbounded) closed 𝔖2\mathfrak{S}_{\geq 2}-subvariety of 𝔛\mathfrak{X} defined by ηξ1=1\eta\xi_{1}=1. Then 𝔜\mathfrak{Y} contains all points of the form (a|a1,b2,b3,)(a|a^{-1},b_{2},b_{3},\ldots), where the bar separates the η\eta and ξ\xi coordinates. We thus see that the Π\Pi-closure of I()\mathrm{I}(\mathfrak{Z}) consists of all points of the form (a|b1,b2,)(a|b_{1},b_{2},\ldots) with aa invertible, which is a dense open subset of 𝔛\mathfrak{X}, and so I¯()=𝔛\overline{\mathrm{I}}(\mathfrak{Z})=\mathfrak{X}. Seemingly the best we can say about I()\mathrm{I}(\mathfrak{Z}) is that it has finite codimension in I¯()\overline{\mathrm{I}}(\mathfrak{Z}). ∎

9.2. Supports

We say that an action of 𝔖\mathfrak{S} on a set SS is smooth if each xSx\in S is stabilized by some 𝔖>n\mathfrak{S}_{>n}. Let MM be a smooth 𝔖\mathfrak{S}-equivariant RR-module. We say that MM is finitely 𝔖\mathfrak{S}-generated if it is generated, as an RR-module, by the 𝔖\mathfrak{S}-orbits of finitely many elements. Recall that the support of MM, denoted Supp(M)\operatorname{Supp}(M), is the subset of 𝔛\mathfrak{X} consisting of those prime ideals 𝔭\mathfrak{p} of RR for which the localization M𝔭M_{\mathfrak{p}} is non-zero. Since MM is 𝔖\mathfrak{S}-equivariant, this set is 𝔖\mathfrak{S}-stable. The following is our main theorem on this invariant.

Theorem 9.6.

Let MM be a smooth 𝔖\mathfrak{S}-equivariant RR-module that is finitely 𝔖\mathfrak{S}-generated.

  1. (a)

    There exists a closed 𝔖n\mathfrak{S}_{\geq n}-subvariety \mathfrak{Z} of 𝔛\mathfrak{X}, for some nn, such that Supp(M)=I()\operatorname{Supp}(M)=\mathrm{I}(\mathfrak{Z}) and V(Ann(M))=I¯()V(\operatorname{Ann}(M))=\overline{\mathrm{I}}(\mathfrak{Z}).

  2. (b)

    Supp(M)\operatorname{Supp}(M) is Zariski dense in V(Ann(M))V(\operatorname{Ann}(M)).

  3. (c)

    If Supp(M)\operatorname{Supp}(M) is bounded then V(Ann(M))topSupp(M)V(\operatorname{Ann}(M))^{\mathrm{top}}\subset\operatorname{Supp}(M).

Proof.

Let MM be generated by the 𝔖\mathfrak{S}-orbits of x1,,xrx_{1},\ldots,x_{r}, and let NN be the submodule generated by just x1,,xrx_{1},\ldots,x_{r}. Let nn be such that each xix_{i} is 𝔖n\mathfrak{S}_{\geq n}-invariant; thus NN is stable under 𝔖n\mathfrak{S}_{\geq n}. Let =Supp(N)\mathfrak{Z}=\operatorname{Supp}(N). Since NN is finitely generated as an RR-module, we have =V(Ann(N))\mathfrak{Z}=V(\operatorname{Ann}(N)), and so \mathfrak{Z} is a closed 𝔖n\mathfrak{S}_{\geq n}-subvariety of 𝔛\mathfrak{X}. Since M=σ𝔖σNM=\sum_{\sigma\in\mathfrak{S}}\sigma N, we have Supp(M)=σ𝔖σ=I()\operatorname{Supp}(M)=\bigcup_{\sigma\in\mathfrak{S}}\sigma\mathfrak{Z}=\mathrm{I}(\mathfrak{Z}). Since Supp(M)V(Ann(M))\operatorname{Supp}(M)\subset V(\operatorname{Ann}(M)) and V(Ann(M))V(\operatorname{Ann}(M)) is closed, we have I¯()V(Ann(M))\overline{\mathrm{I}}(\mathfrak{Z})\subset V(\operatorname{Ann}(M)).

Write I¯()=V(𝔞)\overline{\mathrm{I}}(\mathfrak{Z})=V(\mathfrak{a}) for some radical 𝔖\mathfrak{S}-ideal 𝔞\mathfrak{a} of RR. Since =V(Ann(N))V(𝔞)\mathfrak{Z}=V(\operatorname{Ann}(N))\subset V(\mathfrak{a}), we have 𝔞rad(Ann(N))\mathfrak{a}\subset\operatorname{rad}(\operatorname{Ann}(N)). Write 𝔞=f1,,fs\mathfrak{a}=\langle\!\langle f_{1},\ldots,f_{s}\rangle\!\rangle and let mm be such that each fjf_{j} is 𝔖m\mathfrak{S}_{\geq m}-invariant. For 1ir1\leq i\leq r and 1js1\leq j\leq s and σ𝔖\sigma\in\mathfrak{S}, let ki,j(σ)k_{i,j}(\sigma) be the minimal kk such that (σfj)kxi=0(\sigma f_{j})^{k}x_{i}=0. Then for fixed ii and jj, the quantity ki,j(σ)k_{i,j}(\sigma) only depends on the double coset 𝔖nσ𝔖m\mathfrak{S}_{\geq n}\cdot\sigma\cdot\mathfrak{S}_{\geq m}. Since there are finitely many such double cosets, there is some kk such that ki,j(σ)kk_{i,j}(\sigma)\leq k for all ii, jj, and σ\sigma. We thus have fjkσxi=0f_{j}^{k}\cdot\sigma x_{i}=0 for all ii, jj, and σ\sigma, and so 𝔟=f1k,,fsk\mathfrak{b}=\langle\!\langle f_{1}^{k},\ldots,f_{s}^{k}\rangle\!\rangle annihilates MM. We therefore have V(Ann(M))V(𝔟)V(\operatorname{Ann}(M))\subset V(\mathfrak{b}). Since V(𝔟)=V(𝔞)=I¯()V(\mathfrak{b})=V(\mathfrak{a})=\overline{\mathrm{I}}(\mathfrak{Z}), we see that V(Ann(M))=I¯()V(\operatorname{Ann}(M))=\overline{\mathrm{I}}(\mathfrak{Z}).

We have thus proved (a). Statement (b) follows immediately from (a), while (c) follows from (a) and Proposition 9.2. ∎

Example 9.7.

We note that Supp(M)\operatorname{Supp}(M) can be a proper subset of V(Ann(M))V(\operatorname{Ann}(M)). Let JiJ_{i} be the ideal of RR generated by ξi1\xi_{i}-1 and ξj\xi_{j} for jij\neq i, and consider the module M=i[]Rei/(Jiei)M=\bigoplus_{i\in[\infty]}Re_{i}/(J_{i}e_{i}) with obvious 𝔖\mathfrak{S}-action. Then 𝔜=V(J1)\mathfrak{Y}=V(J_{1}) is the set considered in Example 9.3, and Supp(M)=I(𝔜)\operatorname{Supp}(M)=\mathrm{I}(\mathfrak{Y}) is not Zariski closed; in fact, Supp(M)\operatorname{Supp}(M) is not even Π\Pi-closed. ∎

References

  • [AH] Matthias Aschenbrenner, Christopher J. Hillar. Finite generation of symmetric ideals. Trans. Amer. Math. Soc., 359 (2007), 5171–5192; erratum, ibid. 361 (2009), 5627–5627. arXiv:math/0411514
  • [Co] D. E. Cohen. On the laws of a metabelian variety. J. Algebra 5 (1967), 267–273.
  • [Co2] D. E. Cohen. Closure relations, Buchberger’s algorithm, and polynomials in infinitely many variables. In Computation theory and logic, volume 270 of Lect. Notes Comput. Sci., pp. 78–87, 1987.
  • [Dr] Jan Draisma. Noetherianity up to symmetry. Combinatorial algebraic geometry, Lecture Notes in Math. 2108, Springer, 2014. arXiv:1310.1705v2
  • [DE] Jan Draisma, Rob H. Eggermont. Plücker varieties and higher secants of Sato’s Grassmannian. J. Reine Angew. Math., to appear. arXiv:1402.1667v2
  • [DK] Jan Draisma, Jochen Kuttler. Bounded-rank tensors are defined in bounded degree. Duke Math. J. 163 (2014), no. 1, 35–63. arXiv:1103.5336v2
  • [GN] Sema Güntürkün, Uwe Nagel. Equivariant Hilbert Series of Monomial Orbits. arXiv:1608.06372
  • [HS] Christopher J. Hillar, Seth Sullivant. Finite Gröbner bases in infinite dimensional polynomial rings and applications. Adv. Math., 229 (2012), no. 1, 1–25. arXiv:0908.1777
  • [KLS] Robert Krone, Anton Leykin and Andrew Snowden. Hilbert series of symmetric ideals in infinite polynomial rings via formal languages. J. Algebra 485 (2017), 353–362. arXiv:1606.07956
  • [LNNR] Dinh Van Le, Uwe Nagel, Hop D. Nguyen, Tim Roemer. Castelnuovo–Mumford regularity up to symmetry. arXiv:1806.00457
  • [NR] Uwe Nagel, Tim Römer. Equivariant Hilbert series in non-Noetherian Polynomial Rings. J. Algebra 486 (2017), 204–245. arXiv:1510.02757
  • [NS1] Rohit Nagpal, Andrew Snowden. Symmetric ideals of the infinite polynomial ring. In preparation.
  • [NS2] Rohit Nagpal, Andrew Snowden. Symmetric modules over the infinite polynomial ring. In preparation.
  • [SS] Steven V Sam, Andrew Snowden. Gröbner methods for representations of combinatorial categories. J. Amer. Math. Soc. 30 (2017), 159–203. arXiv:1409.1670v3