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

Relative big polynomial rings

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

Let KK be the field of Laurent series with complex coefficients, let \mathcal{R} be the inverse limit of the standard-graded polynomial rings K[x1,,xn]K[x_{1},\ldots,x_{n}], and let \mathcal{R}^{\flat} be the subring of \mathcal{R} consisting of elements with bounded denominators. In previous joint work with Erman and Sam, we showed that \mathcal{R} and \mathcal{R}^{\flat} (and many similarly defined rings) are abstractly polynomial rings, and used this to give new proofs of Stillman’s conjecture. In this paper, we prove the complementary result that \mathcal{R} is a polynomial algebra over \mathcal{R}^{\flat}.

The author was supported by NSF DMS-1453893.

1. Introduction

1.1. Statement of results

Let A=𝐂tA=\mathbf{C}\llbracket t\rrbracket be the ring of power series over the complex numbers and let K=Frac(A)K=\operatorname{Frac}(A) be the field of Laurent series. The following rings are the main players in this paper:

  • Let \mathcal{R} be the inverse limit of the standard-graded polynomial rings K[x1,,xn]K[x_{1},\ldots,x_{n}] in the category of graded rings. Thus \mathcal{R} is a graded ring, and a degree dd element of \mathcal{R} is a formal KK-linear combination of degree dd monomials in the variables {xi}i1\{x_{i}\}_{i\geq 1}.

  • Let +\mathcal{R}^{+} be the subring of \mathcal{R} with coefficients in AA.

  • Let 0\mathcal{R}^{0} be the subring of \mathcal{R} with coefficients in 𝐂\mathbf{C}.

  • Let \mathcal{R}^{\flat} be the subring of \mathcal{R} where the coefficients have bounded denominators, i.e., ff\in\mathcal{R}^{\flat} if and only if there is some nn such that tnf+t^{n}f\in\mathcal{R}^{+}. Thus =+[1/t]\mathcal{R}^{\flat}=\mathcal{R}^{+}[1/t].

In [ESS], we showed that both \mathcal{R} and \mathcal{R}^{\flat} are abstractly polynomial algebras over KK, and used these results to give new proofs of Stillman’s conjecture. Given these results, it is natural to ask if \mathcal{R} is isomorphic to a polynomial algebra over its subring \mathcal{R}^{\flat}. Our main result is that this is indeed the case:

Theorem 1.1.

The ring \mathcal{R} is a polynomial algebra over \mathcal{R}^{\flat}. More precisely, the map

(1.2) +/(+)2+/+2\mathcal{R}^{\flat}_{+}/(\mathcal{R}^{\flat}_{+})^{2}\to\mathcal{R}_{+}/\mathcal{R}_{+}^{2}

is injective. Suppose that {ξi}iI\{\xi_{i}\}_{i\in I} are homogeneous elements of +\mathcal{R}_{+} whose images form a basis of +/(+++2)\mathcal{R}_{+}/(\mathcal{R}^{\flat}_{+}+\mathcal{R}^{2}_{+}). Then the \mathcal{R}^{\flat}-algebra homomorphism [Xi]iI\mathcal{R}^{\flat}[X_{i}]_{i\in I}\to\mathcal{R} mapping XiX_{i} to ξi\xi_{i} is an isomorphism of graded \mathcal{R}^{\flat}-algebras.

Recall from [AH] that a homogeneous element ff of a graded ring RR has strength n\leq n if there is an expression f=i=1ngihif=\sum_{i=1}^{n}g_{i}h_{i} where the gig_{i} and hih_{i} are homogeneous elements of positive degree. If no such expression exists, we say that ff has strength \infty. The ideal R+2R_{+}^{2} is exactly the ideal of finite strength elements. Thus the injectivity of (1.2) (which is the only non-trivial part of the theorem) is equivalent to the following, which is what we actually prove:

Theorem 1.3.

Let ff be an element of \mathcal{R}^{\flat} that has finite strength in \mathcal{R}. Then ff has finite strength in \mathcal{R}^{\flat}.

Here is the idea of the proof. Let ff be a given element of \mathcal{R}^{\flat} that has finite strength in \mathcal{R}. Scaling by a power of tt, we can assume that f+f\in\mathcal{R}^{+}. Let II be the ideal of +\mathcal{R}^{+} generated by the partial derivatives of ff. We show that the extension of II to \mathcal{R}^{\flat} is contained in the extension of some finitely generated ideal J+J\subset\mathcal{R}^{+}. This follows from elementary arguments involving heights, combined with the polynomiality of \mathcal{R}^{\flat} and \mathcal{R}. It follows that II is contained in the tt-adic saturation of JJ. The main technical result of this paper (Theorem 5.1) shows that such a saturation is (close enough to) finitely generated. From here, another elementary argument shows that ff has finite strength in +\mathcal{R}^{+}.

1.2. Additional results on +\mathcal{R}^{+}

As mentioned, the rings \mathcal{R} and \mathcal{R}^{\flat} are polynomial KK-algebras. It is therefore easy to prove all sorts of results about heights in these rings. The ring +\mathcal{R}^{+}, on other hand, is not a polynomial AA-algebra: indeed, if it were then its graded pieces would be free AA-modules, but its graded pieces are infinite products of AA, which are not free. It is therefore not obvious how heights behave in +\mathcal{R}^{+}.

In the course of this work, we discovered a number of results about heights in +\mathcal{R}^{+}, such as a version of the Hauptidealsatz (Proposition 7.10) and a form of the catenary property (Proposition 7.4). Although these results are not needed to prove the main theorem, we have included them in §7 as they use closely related methods.

1.3. Motivation

There are two sources of motivation for this work. One comes from commutative algebra. As mentioned, our polynomiality results for rings like \mathcal{R} and \mathcal{R}^{\flat} were used in [ESS] to give two new proofs of Stillman’s conjectures, following the original proof in [AH]. Shortly thereafter, our polynomiality results were used in [DLL] to give a fourth proof of Stillman’s conjecture. In [ESS2], we strengthened our polynomiality results, which allowed us to strengthen the results of [AH] on small subalgebras. Due to these applications, we believe it is worthwhile to try to better understand the precise nature and extent of the polynomiality phenomena. Theorem 1.1 is a step in this direction.

The second source of motivation comes from infinite dimensional algebraic geometry. More precisely, in [BDES], we study certain infinite dimensional algebraic varieties equipped with an action of 𝐆𝐋\mathbf{GL}_{\infty}, and establish a number of nice properties in this situation (such as an analog of Chevalley’s theorem). An important open problem remaining in [BDES] concerns the precise structure of image closures. Theorem 1.1 was proved with this problem in mind. To see the connection, suppose that F(X1,,Xr)F(X_{1},\ldots,X_{r}) is a polynomial in rr variables that is homogeneous of degree dd, where XiX_{i} has degree did_{i}. Then FF defines a function

F:d10××dr0d0.F\colon\mathcal{R}^{0}_{d_{1}}\times\cdots\times\mathcal{R}^{0}_{d_{r}}\to\mathcal{R}^{0}_{d}.

The space d0\mathcal{R}^{0}_{d} is an example of an infinite dimensional variety with 𝐆𝐋\mathbf{GL}_{\infty}-action, as it can be identified with the dual of Symd(𝐂)\operatorname{Sym}^{d}(\mathbf{C}^{\infty}). Suppose fd0f\in\mathcal{R}^{0}_{d}. Theorem 1.1 implies that if ff can be realized in the form limt0F(g1,,gr)\lim_{t\to 0}F(g_{1},\ldots,g_{r}) with gidig_{i}\in\mathcal{R}_{d_{i}}, then it can be realized in this form with gidig_{i}\in\mathcal{R}^{\flat}_{d_{i}}. In other words, if ff can be realized as a certain kind of “wild” limit in the image of FF then it can also be realized by a much nicer “tame” kind of limit. We had hoped to use this to resolve the open question in [BDES]. Unfortunately, it does not appear to be quite enough. However, we believe that Theorem 1.1 could still be useful in studying similar problems.

1.4. Open problems

Here are some open problems raised by our work:

  • In the setting of §5, is it true that the saturation of a finitely generated ideal is finitely generated?

  • Let AA be an integral domain such that K=Frac(A)K=\operatorname{Frac}(A) is perfect, let RR be the inverse limit of the graded rings K[x1,,xn]K[x_{1},\ldots,x_{n}], and let RR^{\flat} be the subring where the denominators are bounded (i.e., fRf\in R^{\flat} if afaf has coefficients in AA for some non-zero aAa\in A). In [ESS], we showed that RR and RR^{\flat} are polynomial KK-algebras. Is RR a polynomial algebra over RR^{\flat}? This paper only addresses the special case where A=𝐂tA=\mathbf{C}\llbracket t\rrbracket.

  • As mentioned, +\mathcal{R}^{+} is not a polynomial ring. However, the results of §7 show that in some ways it behaves like a polynomial ring. Can this observation be sharpened, or made more precise?

1.5. Outline

In §2, we give some general background on heights. In §3, we prove a comparison result for heights in \mathcal{R}^{\flat} and \mathcal{R}. In §4, we prove a Nakayama-like lemma that will be used in our analysis of saturation. In §5, we prove the main technical result of the paper (Theorem 5.1) on saturations. Using this, we prove our main theorem in §6. Finally, in §7, we prove some additional results about +\mathcal{R}^{+}.

Acknowledgments

We thank Dan Erman for comments on a draft of this paper.

2. Background on heights

Let RR be a ring. Recall that the height of a prime ideal 𝔭\mathfrak{p}, denoted htR(𝔭)\operatorname{ht}_{R}(\mathfrak{p}), is the maximal value of nn for which there exists a strict chain of primes 𝔭0𝔭n=𝔭\mathfrak{p}_{0}\subset\cdots\subset\mathfrak{p}_{n}=\mathfrak{p}, or \infty if there exist arbitrarily long such chains. The height of an ideal II, denoted htR(I)\operatorname{ht}_{R}(I), is defined as the minimum of htR(𝔭)\operatorname{ht}_{R}(\mathfrak{p}) over primes 𝔭\mathfrak{p} containing II; by convention, the height of the unit ideal is infinity. If II is an ideal in a finite variable polynomial ring R=F[x1,,xn]R=F[x_{1},\ldots,x_{n}], with FF a field, then htR(𝔭)\operatorname{ht}_{R}(\mathfrak{p}) is the codimension of the locus V(I)𝐀FnV(I)\subset\mathbf{A}^{n}_{F}. We note that if IJI\subset J then htR(I)htR(J)\operatorname{ht}_{R}(I)\leq\operatorname{ht}_{R}(J). In what follows, polynomial rings can have infinitely many variables.

Proposition 2.1.

Let RR be a polynomial ring over a field. Then any finite height prime ideal is finitely generated.

Proof.

See [ESS, Proposition 3.2]

Proposition 2.2.

Let RR be a polynomial ring over a field. Then an ideal has finite height if and only if it is contained in a finitely generated non-unital ideal.

Proof.

Suppose II has finite height. Then, by definition, II is contained in a prime of finite height, which is finitely generated by Proposition 2.1.

Now suppose II is contained in a finitely generated non-unital ideal, say (f1,,fr)(f_{1},\ldots,f_{r}). Each fif_{i} uses only finitely many variables, and so (f1,,fr)(f_{1},\ldots,f_{r}) is extended from a finite variable subring. Performing such an extension does not change height [ESS, Proposition 3.3]. ∎

Proposition 2.3.

Let RR be a finite variable polynomial ring over a field FF, let E/FE/F be a field extension, and let S=EFRS=E\otimes_{F}R. Let 𝔭\mathfrak{p} be a prime of SS and let 𝔮\mathfrak{q} be its contraction to RR. Then htR(𝔮)htS(𝔭)\operatorname{ht}_{R}(\mathfrak{q})\leq\operatorname{ht}_{S}(\mathfrak{p}).

Proof.

The natural homomorphism EFR/𝔭S/𝔮E\otimes_{F}R/\mathfrak{p}\to S/\mathfrak{q} is surjective. We thus find

dim(R/𝔭)=dim(EFR/𝔭)dim(S/𝔮),\dim(R/\mathfrak{p})=\dim(E\otimes_{F}R/\mathfrak{p})\geq\dim(S/\mathfrak{q}),

where dim\dim denotes Krull dimension. Since dim(R/𝔭)=nhtR(𝔭)\dim(R/\mathfrak{p})=n-\operatorname{ht}_{R}(\mathfrak{p}) and dim(S/𝔮)=nhtS(𝔮)\dim(S/\mathfrak{q})=n-\operatorname{ht}_{S}(\mathfrak{q}), the result follows. ∎

Remark 2.4.

(a) We can actually have htR(𝔮)<htS(𝔭)\operatorname{ht}_{R}(\mathfrak{q})<\operatorname{ht}_{S}(\mathfrak{p}). For instance, let F=𝐂F=\mathbf{C} and E=𝐂(t)E=\mathbf{C}(t), and take 𝔭\mathfrak{p} to be the ideal generated by x1tx2x_{1}-tx_{2}. Then 𝔭\mathfrak{p} has height 1 but 𝔮=0\mathfrak{q}=0 has height 0. (b) The proposition holds in the infinite variable case too. ∎

Proposition 2.5.

Let RR be a ring. Suppose the following condition holds:

  • ()(\ast)

    If II is a finitely generated ideal of height c<c<\infty, then there are only finitely many primes 𝔭\mathfrak{p} of height cc that contain II.

Let II be an ideal of RR and let I=αJαI=\bigcup_{\alpha\in\mathcal{I}}J_{\alpha} be a directed union, where JαJ_{\alpha} are ideals contained in II. Then htR(I)=supαhtR(Jα)\operatorname{ht}_{R}(I)=\sup_{\alpha\in\mathcal{I}}\operatorname{ht}_{R}(J_{\alpha}).

Proof.

For any JIJ\subset I we have htR(J)htR(I)\operatorname{ht}_{R}(J)\leq\operatorname{ht}_{R}(I), and so supαhtR(Jα)htR(I)\sup_{\alpha\in\mathcal{I}}\operatorname{ht}_{R}(J_{\alpha})\leq\operatorname{ht}_{R}(I). We now prove the reverse inequality. First, suppose that the JαJ_{\alpha} are finitely generated. If supαhtR(Jα)\sup_{\alpha}\operatorname{ht}_{R}(J_{\alpha}) is infinite then there is nothing to prove, so suppose it is a finite number cc. Passing to a cofinal subset, we may as well suppose that htR(Jα)=c\operatorname{ht}_{R}(J_{\alpha})=c for all α\alpha. For each α\alpha, let 𝒫α\mathcal{P}_{\alpha} be the set of prime ideals of height cc containing JαJ_{\alpha}; this set is finite by hypothesis. Of course, if αβ\alpha\leq\beta then 𝒫β𝒫α\mathcal{P}_{\beta}\subset\mathcal{P}_{\alpha}. By a standard compactness result, we have α𝒫α\bigcap_{\alpha\in\mathcal{I}}\mathcal{P}_{\alpha}\neq\varnothing. We can thus find a prime 𝔭\mathfrak{p} of height cc such that Jα𝔭J_{\alpha}\subset\mathfrak{p} for all α\alpha. Since I=αJαI=\bigcup_{\alpha\in\mathcal{I}}J_{\alpha}, we thus find I𝔭I\subset\mathfrak{p}, and so htR(I)c\operatorname{ht}_{R}(I)\leq c.

We now treat the general case. For each α\alpha, let {Jα,β}β𝒦α\{J_{\alpha,\beta}\}_{\beta\in\mathcal{K}_{\alpha}} be the finitely generated ideals contained in JαJ_{\alpha}. Then

htR(I)=supα,βJα,β=supαsupβht(Jα,β)=supαhtR(Jα),\operatorname{ht}_{R}(I)=\sup_{\alpha,\beta}J_{\alpha,\beta}=\sup_{\alpha}\sup_{\beta}\operatorname{ht}(J_{\alpha,\beta})=\sup_{\alpha}\operatorname{ht}_{R}(J_{\alpha}),

where in the first and last step we used the previous case. ∎

Proposition 2.6.

Suppose RR is a polynomial ring over a field. Then condition ()(\ast) holds.

Proof.

Let II be a finitely generated ideal of RR of height cc. Then II is extended from an ideal II^{\prime} of some finite variable subring R0R_{0} of RR. Since such extensions do not affect height [ESS, Proposition 3.3], we see that II^{\prime} has height cc also. Let 𝔮1,,𝔮r\mathfrak{q}_{1},\ldots,\mathfrak{q}_{r} be the minimal primes above II^{\prime} in R0R_{0}. Now, let 𝔭\mathfrak{p} be height cc prime of RR containing II. Then 𝔭\mathfrak{p} is finitely generated by Proposition 2.1, and thus extended from a prime 𝔭\mathfrak{p}^{\prime} of a finite variable subring R1R_{1} of RR, which we can assume contains R0R_{0}. The ideal IR1I^{\prime}R_{1} has height cc and 𝔮1R1,,𝔮rR1\mathfrak{q}_{1}R_{1},\ldots,\mathfrak{q}_{r}R_{1} are the minimal primes over it. Since 𝔭\mathfrak{p}^{\prime} contains IR1I^{\prime}R_{1} and has the same height, it follows that 𝔭=𝔮iR1\mathfrak{p}^{\prime}=\mathfrak{q}_{i}R_{1} for some ii, and thus 𝔭=𝔮iR\mathfrak{p}=\mathfrak{q}_{i}R. We thus see that there are only rr choices for 𝔭\mathfrak{p}. ∎

Proposition 2.7.

Let AA be a polynomial ring over a field and let II be a finite height ideal of A[x]A[x]. Then htA(IA)htA[x](I)\operatorname{ht}_{A}(I\cap A)\leq\operatorname{ht}_{A[x]}(I).

Proof.

If II is prime then it is finitely generated (Proposition 2.1) and the result follows from the corresponding result for finite variable polynomial rings and the fact that extending to larger polynomial rings does not change height [ESS, Proposition 3.3]. Now suppose that II is a general ideal of finite height cc. Let 𝔭\mathfrak{p} be a height cc prime containing II. Then IA𝔭AI\cap A\subset\mathfrak{p}\cap A, and the latter has height c\leq c. Thus the result follows. ∎

Let RR be a ring and let SS be a multiplicative subset. A basic theorem states that the primes of S1RS^{-1}R correspond bijectively (via extension and contraction) to the primes of RR disjoint from SS. The following result shows that this correspondence preserves height.

Proposition 2.8.

Let 𝔭\mathfrak{p} be a prime of S1RS^{-1}R and let 𝔮\mathfrak{q} be its contraction to RR. Then htR(𝔮)=htS1R(𝔭)\operatorname{ht}_{R}(\mathfrak{q})=\operatorname{ht}_{S^{-1}R}(\mathfrak{p}).

Proof.

Let 𝔭0𝔭r=𝔭\mathfrak{p}_{0}\subset\cdots\subset\mathfrak{p}_{r}=\mathfrak{p} be a strict chain of primes in S1RS^{-1}R. Contracting gives a strict chain of primes in RR. Thus htR(𝔮)htS1R(𝔭)\operatorname{ht}_{R}(\mathfrak{q})\geq\operatorname{ht}_{S^{-1}R}(\mathfrak{p}).

Now let 𝔮0𝔮r=𝔮\mathfrak{q}_{0}\subset\cdots\subset\mathfrak{q}_{r}=\mathfrak{q} be a strict chain of primes in RR. Since 𝔮\mathfrak{q} is the contraction of an ideal from S1RS^{-1}R, it is disjoint from SS, and so all the ideals 𝔮i\mathfrak{q}_{i} are as well. Thus extending gives a strict chain, and so htS1R(𝔭)htR(𝔮)\operatorname{ht}_{S^{-1}R}(\mathfrak{p})\geq\operatorname{ht}_{R}(\mathfrak{q}). ∎

Corollary 2.9.

Let II be an ideal of RR. Then htR(I)htS1R(S1I)\operatorname{ht}_{R}(I)\leq\operatorname{ht}_{S^{-1}R}(S^{-1}I).

Proof.

Let c=htS1R(S1I)c=\operatorname{ht}_{S^{-1}R}(S^{-1}I). If c=c=\infty there is nothing to prove, so suppose cc is finite. Let 𝔭\mathfrak{p} be a height cc prime of S1RS^{-1}R containing S1IS^{-1}I. Let 𝔮\mathfrak{q} be the contraction of 𝔭\mathfrak{p}. Then htR(𝔮)=c\operatorname{ht}_{R}(\mathfrak{q})=c by the proposition. Since I𝔮I\subset\mathfrak{q}, we thus have htR(I)c\operatorname{ht}_{R}(I)\leq c. ∎

3. Comparing heights in \mathcal{R}^{\flat} and \mathcal{R}

Let >n\mathcal{R}_{>n} be defined like \mathcal{R}, but only using the variables xix_{i} with i>ni>n. Similarly define >n\mathcal{R}^{\flat}_{>n}, and so on. We have a natural isomorphism =>n[x1,,xn]\mathcal{R}=\mathcal{R}_{>n}[x_{1},\ldots,x_{n}], and similarly for the other variants.

Proposition 3.1.

Let II be an ideal of \mathcal{R} of finite height. Then I>n=0I\cap\mathcal{R}_{>n}=0 for n0n\gg 0. Similarly for \mathcal{R}^{\flat} and 0\mathcal{R}^{0}.

Proof.

Let ff be a non-zero element of II. As in [ESS, Lemma 4.9], we can find a linear change of variables γ\gamma in finitely many of the xx’s so that γ(f)\gamma(f) is monic in x1x_{1}. We then have ht>1(γ(I)>1)=ht(I)1\operatorname{ht}_{\mathcal{R}_{>1}}(\gamma(I)\cap\mathcal{R}_{>1})=\operatorname{ht}_{\mathcal{R}}(I)-1. Indeed, if II is fintiely generated, this is [ESS, Corollary 3.8] (which easily reduces to the finite variable case), while Propositions 2.5 and 2.6 allow us to reduce to this case. It follows from Proposition 2.7 that ht>n(γ(I)>n)<ht(I)\operatorname{ht}_{\mathcal{R}_{>n}}(\gamma(I)\cap\mathcal{R}_{>n})<\operatorname{ht}_{\mathcal{R}}(I) for all n1n\geq 1. Taking nn larger than the variables used in γ\gamma, we thus have ht>n(I>n)<ht(I)\operatorname{ht}_{\mathcal{R}_{>n}}(I\cap\mathcal{R}_{>n})<\operatorname{ht}_{\mathcal{R}}(I). Thus, by induction on height, the result follows. ∎

Remark 3.2.

The proposition does not hold for +\mathcal{R}^{+}: indeed, the ideal (t)(t) is a counterexample. We will formulate a version for +\mathcal{R}^{+} in Proposition 7.1. ∎

Corollary 3.3.

Let 𝔭\mathfrak{p} be a prime ideal of \mathcal{R} of finite height. Then 𝔭\mathfrak{p} is the contraction of a prime ideal of Frac(>n)[x1,,xn]\operatorname{Frac}(\mathcal{R}_{>n})[x_{1},\ldots,x_{n}] for all sufficiently large nn.

Proposition 3.4.

Let II be an ideal of \mathcal{R}^{\flat}. Then ht(I)ht(I)\operatorname{ht}_{\mathcal{R}^{\flat}}(I)\leq\operatorname{ht}_{\mathcal{R}}(I\mathcal{R}).

Proof.

Let c=ht(I)c=\operatorname{ht}_{\mathcal{R}}(I\mathcal{R}). If c=c=\infty there is nothing to prove, so suppose cc is finite. Let 𝔭\mathfrak{p} be a height cc prime of \mathcal{R} containing II\mathcal{R}. Let nn be such that 𝔭\mathfrak{p} is contracted from a prime 𝔭\mathfrak{p}^{\prime} of Frac(>n)[x1,,xn]\operatorname{Frac}(\mathcal{R}_{>n})[x_{1},\ldots,x_{n}], necessarily of height cc (by Proposition 2.8). Let 𝔮\mathfrak{q}^{\prime} be the contraction of 𝔭\mathfrak{p}^{\prime} to Frac(>n)[x1,,xn]\operatorname{Frac}(\mathcal{R}^{\flat}_{>n})[x_{1},\ldots,x_{n}]. Then 𝔮\mathfrak{q}^{\prime} has height c\leq c by Proposition 2.3. Let 𝔮\mathfrak{q} be the contraction of 𝔮\mathfrak{q}^{\prime} to \mathcal{R}^{\flat}. Then 𝔮\mathfrak{q} has the same height as 𝔮\mathfrak{q}^{\prime} by Proposition 2.8, which is c\leq c. Since II is contained in 𝔮\mathfrak{q}, it too has height c\leq c. ∎

4. A Nakayama-like lemma

Let Π\Pi be an infinite product of copies of A=𝐂tA=\mathbf{C}\llbracket t\rrbracket, and suppose that MM is an AA-submodule of Π\Pi. We recall several concepts:

  • MM is (tt-adically) complete if the following holds: given elements xitiMx_{i}\in t^{i}M for i0i\geq 0 the sum i0xi\sum_{i\geq 0}x_{i} belongs to MM.

  • MM is (tt-adically) closed is the following holds: given elements xiMtiΠx_{i}\in M\cap t^{i}\Pi for i0i\geq 0 the sum i0xi\sum_{i\geq 0}x_{i} belongs to MM. Obviously, closed implies complete.

  • We let Σn(M)\Sigma_{n}(M) be the set of elements xΠx\in\Pi such that tnxMt^{n}x\in M. It is an AA-submodule of Π\Pi. We let Σ(M)=n1Σn(M)\Sigma(M)=\bigcup_{n\geq 1}\Sigma_{n}(M), which we call the saturation of MM.

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

Proposition 4.1.

Let MM be an AA-submodule of Π\Pi. Suppose:

  1. (a)

    MM is complete.

  2. (b)

    Σn(M)/M\Sigma_{n}(M)/M is finite dimensional over 𝐂\mathbf{C} for all nn.

  3. (c)

    Σ(M)\Sigma(M) is contained in the closure of MM.

Then M=Σ(M)M=\Sigma(M), that is, MM is saturated.

Proof.

Let δ(n)\delta(n) be the dimension of Σn(M)/M\Sigma_{n}(M)/M. Let x1,x2,x_{1},x_{2},\ldots be elements of Σ(M)\Sigma(M) such that x1,x2,,xδ(n)x_{1},x_{2},\ldots,x_{\delta(n)} is a basis of Σn(M)/M\Sigma_{n}(M)/M for each nn. Note that x1,x2,x_{1},x_{2},\ldots is a basis for Σ(M)/M\Sigma(M)/M.

Let yy be an element of the tt-adic closure of MM. We can thus write y=j=0zjy=\sum_{j=0}^{\infty}z_{j}, where zjMtjΠz_{j}\in M\cap t^{j}\Pi. Since tjzjΣj(M)t^{-j}z_{j}\in\Sigma_{j}(M), we can write tjzj=bj+c1,jx1++ca(j),jxa(j)t^{-j}z_{j}=b_{j}+c_{1,j}x_{1}+\cdots+c_{a(j),j}x_{a(j)} with bjMb_{j}\in M and ci,j𝐂c_{i,j}\in\mathbf{C}. We thus find

y=j0tj(bj+c1,jx1++ca(j),jxa(j))=j0tjbj+i1δ(j)ici,jtjxiy=\sum_{j\geq 0}t^{j}(b_{j}+c_{1,j}x_{1}+\cdots+c_{a(j),j}x_{a(j)})=\sum_{j\geq 0}t^{j}b_{j}+\sum_{i\geq 1}\sum_{\delta(j)\geq i}c_{i,j}t^{j}x_{i}

Let ϵ(i)\epsilon(i) be the minimal value of jj such that δ(j)i\delta(j)\geq i, with the convention ϵ(i)=\epsilon(i)=\infty if δ(j)<i\delta(j)<i for all jj. Thus in the inner sum above, we can write jϵ(i)j\geq\epsilon(i). Note that ϵ(i)1\epsilon(i)\geq 1 for all ii and ϵ(i)\epsilon(i)\to\infty as ii\to\infty. Now, the first sum above belongs to MM by (a). And the inner sum in the second sum converges to an element of AA that is dividible by ϵ(i)\epsilon(i). We conclude that for any yy in the tt-adic closure of MM, we can write

y=b+i0fi(t)xiy=b+\sum_{i\geq 0}f_{i}(t)x_{i}

where bMb\in M and fi(t)tϵ(i)Af_{i}(t)\in t^{\epsilon(i)}A.

We now apply this to xix_{i}, which belongs to the tt-adic closure of MM by (c). We can thus write

xi=bi+j0fi,jxjx_{i}=b_{i}+\sum_{j\geq 0}f_{i,j}x_{j}

where biMb_{i}\in M and fi,jtϵ(j)Af_{i,j}\in t^{\epsilon(j)}A. Let x¯\underline{x} and b¯\underline{b} be the infinite column vectors (x1,x2,)(x_{1},x_{2},\ldots) and (b1,b2,)(b_{1},b_{2},\ldots), and let AA be the infinite square matrix given by Ai,j=fi,jA_{i,j}=f_{i,j}. We then get the linear equation

(1A)x¯=b¯.(1-A)\underline{x}=\underline{b}.

The entries of AA all belong to the maximal ideal of AA. In fact, for any nn there exists an mm such that all columns to the right of the mmth column of AA are divisible by tnt^{n}. It follows that 1A1-A is invertible, and so the entries of x¯=(1A)1b¯\underline{x}=(1-A)^{-1}\underline{b} belongs to MM. Thus xiMx_{i}\in M for all ii, which completes the proof. ∎

5. A result on saturation

Let SS be a graded polynomial 𝐂\mathbf{C}-algebra, where each variable is homogeneous of positive degree; the interesting case is where there are infinitely many variables. Let RR be the graded version of StS\llbracket t\rrbracket; thus RR is a graded ring and a degree dd element of RR is a power series in tt with coefficients in SdS_{d}. Note that elements of RR can use infinitely many of the variables in SS, which is why RR can be hard to work with.

Given a homogeneous ideal II of RR, we define its saturation, denoted Sat(I)\operatorname{Sat}(I), to be the set of all elements fRf\in R such that tnfIt^{n}f\in I for some nn. Thus Sat(I)\operatorname{Sat}(I) is a homogeneous ideal with Sat(I)n=Σ(In)\operatorname{Sat}(I)_{n}=\Sigma(I_{n}), where Σ\Sigma is as in the previous section. (We note that each graded piece of RR is isomorphic to a product of copies of 𝐂t\mathbf{C}\llbracket t\rrbracket.) We would like to know that the saturation of a finitely generated ideal is finitely generated. Unfortunately, we have not been able to prove this. However, we prove a weaker statement that is sufficient for our applications. For an integer d0d\geq 0, define Satd(I)\operatorname{Sat}_{\leq d}(I) to be the ideal generated by II and Sat(I)k\operatorname{Sat}(I)_{k} for 0kd0\leq k\leq d. The result is:

Theorem 5.1.

If II is finitely generated then so is Satd(I)\operatorname{Sat}_{\leq d}(I), for any dd.

Fix II and dd as in the theorem. We may assume, by induction, that the theorem holds for smaller values of dd, and so we may replace II by Satd1(I)\operatorname{Sat}_{\leq d-1}(I); this is still finitely generated by the inductive hypothesis. Thus if fSat(I)f\in\operatorname{Sat}(I) has degree <d<d then ff already belongs to II.

Let f1,,frf_{1},\ldots,f_{r} be homogeneous generators of II. Let XkX_{k} be the set of all tuples (g1,,gr)Rr(g_{1},\ldots,g_{r})\in R^{r} such that deg(gifi)=d\deg(g_{i}f_{i})=d for all ii and tkg1f1++grfrt^{k}\mid g_{1}f_{1}+\cdots+g_{r}f_{r}. Note that Xk+1XkX_{k+1}\subset X_{k}. Define

πk:XkRd,(g1,,gr)g1f1++grfrtk.\pi_{k}\colon X_{k}\to R_{d},\qquad(g_{1},\ldots,g_{r})\mapsto\frac{g_{1}f_{1}+\cdots+g_{r}f_{r}}{t^{k}}.

and put Yk=im(πk)Y_{k}=\operatorname{im}(\pi_{k}). Note that YkY_{k} is exactly Σk(Id)\Sigma_{k}(I_{d}), that is, the set of gSdg\in S_{d} such that tkgIt^{k}g\in I.

Let EE be the set of all tuples (g1,,gr)Sr(g_{1},\ldots,g_{r})\in S^{r} such that g1f1(0)+grfr(0)=0g_{1}f_{1}(0)+\cdots g_{r}f_{r}(0)=0. Here fi(0)Sf_{i}(0)\in S is the result of substituting 0 for tt in fif_{i}. Thus EE is just the syzygy module for (f1(0),,fr(0))(f_{1}(0),\ldots,f_{r}(0)). Let E¯=Ed/(EdS+E)\overline{E}=E_{d}/(E_{d}\cap S_{+}E), i.e., the space of degree dd generators for EE. Here EdE_{d} consists of tuples (g1,,gr)E(g_{1},\ldots,g_{r})\in E such that deg(gifi)=d\deg(g_{i}f_{i})=d.

Lemma 5.2.

E¯\overline{E} is a finite dimensional vector space.

Proof.

Let AA be the set of variables appearing in f1(0),,fr(0)f_{1}(0),\ldots,f_{r}(0), which is finite. Let SSS^{\prime}\subset S be the polynomial ring in the AA variables, and define EE^{\prime} like EE, but using SS^{\prime} instead of SS. Since SS^{\prime} is noetherian, it follows that EE^{\prime} is a finitely generated module, and so E¯=Ed/(EdS+E)\overline{E}^{\prime}=E^{\prime}_{d}/(E^{\prime}_{d}\cap S_{+}^{\prime}E^{\prime}) is finite dimensional.

We have a natural inclusion EEE^{\prime}\to E. We claim that the induced map E¯E¯\overline{E}^{\prime}\to\overline{E} is surjective, which will complete the proof. Thus suppose (g1,,gr)(g_{1},\ldots,g_{r}) is a degree dd element of EE. Write gi=egi,exeg_{i}=\sum_{e}g_{i,e}x^{e}, where the sum is over all monomials in variables not in AA, and gi,eSg_{i,e}\in S^{\prime}. Each tuple (g1,e,,gr,e)(g_{1,e},\ldots,g_{r,e}) belongs to EE and so if e0e\neq 0 then (xeg1,e,,xegr,e)(x^{e}g_{1,e},\ldots,x^{e}g_{r,e}) belongs to S+ES_{+}E. Hence (g1,,gr)(g_{1},\ldots,g_{r}) and (g1,0,,gr,0)(g_{1,0},\ldots,g_{r,0}) are equal in E¯\overline{E}. Since the latter belongs to EE^{\prime}, the result follows. ∎

If (g1,,gr)(g_{1},\ldots,g_{r}) belongs to X1X_{1} then g1f1++grfrg_{1}f_{1}+\cdots+g_{r}f_{r} is divisible by tt, which exactly means that g1(0)f1(0)++gr(0)fr(0)=0g_{1}(0)f_{1}(0)+\cdots+g_{r}(0)f_{r}(0)=0, i.e., (g1(0),,gr(0))(g_{1}(0),\ldots,g_{r}(0)) belongs to EdE_{d}. Define ρ:X1E¯\rho\colon X_{1}\to\overline{E} by taking (g1,,gr)(g_{1},\ldots,g_{r}) to the element represented by (g1(0),,gr(0))(g_{1}(0),\ldots,g_{r}(0)). Put Zk=ρ(Xk)Z_{k}=\rho(X_{k}). Since Xk+1XkX_{k+1}\subset X_{k}, we have Zk+1ZkZ_{k+1}\subset Z_{k}.

Lemma 5.3 (Key Lemma).

Let (g1,,gr)Xk(g_{1},\ldots,g_{r})\in X_{k} with k1k\geq 1. Suppose that ρ(g1,,gr)=0\rho(g_{1},\ldots,g_{r})=0. Then πk(g1,,gr)Yk1\pi_{k}(g_{1},\ldots,g_{r})\in Y_{k-1}.

Proof.

Since (g1(0),,gr(0))(g_{1}(0),\ldots,g_{r}(0)) maps to 0 in E¯\overline{E}, we can find elements (a1,i,,ar,i)E(a_{1,i},\ldots,a_{r,i})\in E for 1in1\leq i\leq n of degree <d<d such that

(g1(0),,gr(0))=i=1nbi(a1,i,,ar,i)(g_{1}(0),\ldots,g_{r}(0))=\sum_{i=1}^{n}b_{i}\cdot(a_{1,i},\ldots,a_{r,i})

for some biSb_{i}\in S. Now, we have

g1(0)f1++gr(0)frt=i=1nbia1,if1++ar,ifrt.\frac{g_{1}(0)f_{1}+\cdots+g_{r}(0)f_{r}}{t}=\sum_{i=1}^{n}b_{i}\cdot\frac{a_{1,i}f_{1}+\cdots+a_{r,i}f_{r}}{t}.

Since a1,if1++ar,ifra_{1,i}f_{1}+\cdots+a_{r,i}f_{r} is divisible by tt, its quotient by tt belongs to Sat(I)\operatorname{Sat}(I). It also has degree <d<d, and so it belongs to II by our initial setup. Thus we can write

bia1,if1++ar,ifrt=c1,if1++cr,ifrb_{i}\cdot\frac{a_{1,i}f_{1}+\cdots+a_{r,i}f_{r}}{t}=c_{1,i}f_{1}+\cdots+c_{r,i}f_{r}

for elements ci,jRc_{i,j}\in R. Letting ci=ci,1++ci,rc_{i}=c_{i,1}+\cdots+c_{i,r}, we thus have

g1(0)f1++gr(0)t=c1f1++crfr.\frac{g_{1}(0)f_{1}+\cdots+g_{r}(0)}{t}=c_{1}f_{1}+\cdots+c_{r}f_{r}.

Now, write gi=gi(0)+tgig_{i}=g_{i}(0)+tg_{i}^{\prime}. Then

πk(g1,,gr)\displaystyle\pi_{k}(g_{1},\ldots,g_{r}) =g1f1++grfrtk\displaystyle=\frac{g_{1}f_{1}+\cdots+g_{r}f_{r}}{t^{k}}
=g1(0)f1++gr(0)frt+g1f1++grfrtk1\displaystyle=\frac{\frac{g_{1}(0)f_{1}+\cdots+g_{r}(0)f_{r}}{t}+g_{1}^{\prime}f_{1}+\cdots+g_{r}^{\prime}f_{r}}{t^{k-1}}
=(g1+c1)f1++(gr+cr)frtk1\displaystyle=\frac{(g_{1}^{\prime}+c_{1})f_{1}+\cdots+(g_{r}^{\prime}+c_{r})f_{r}}{t^{k-1}}
=πk1(g1+c1,,gr+cr).\displaystyle=\pi_{k-1}(g_{1}^{\prime}+c_{1},\ldots,g_{r}^{\prime}+c_{r}).

This completes the proof. ∎

Lemma 5.4.

Yk/Yk1Y_{k}/Y_{k-1} is finite dimensional over 𝐂\mathbf{C}.

Proof.

The key lemma gives a surjection ZkXk/ker(ρ|Xk)Yk/Yk1Z_{k}\cong X_{k}/\ker(\rho|_{X_{k}})\to Y_{k}/Y_{k-1}, and ZkZ_{k} is finite dimensional. ∎

Lemma 5.5.

Let (g1,,gr)Xk(g_{1},\ldots,g_{r})\in X_{k} with k1k\geq 1. Suppose that ρ(g1,,gr)Zk+m\rho(g_{1},\ldots,g_{r})\in Z_{k+m}. Then πk(g1,,gr)Yk1+tmYk+m\pi_{k}(g_{1},\ldots,g_{r})\in Y_{k-1}+t^{m}Y_{k+m}.

Proof.

Let (h1,,hr)Xk+m(h_{1},\ldots,h_{r})\in X_{k+m} be such that ρ(g1,,gr)=ρ(h1,,hr)\rho(g_{1},\ldots,g_{r})=\rho(h_{1},\ldots,h_{r}). We have

πk(h1,,hr)=tmπk+m(h1,,hr)tmYk+m.\pi_{k}(h_{1},\ldots,h_{r})=t^{m}\pi_{k+m}(h_{1},\ldots,h_{r})\in t^{m}Y_{k+m}.

We have

πk(g1,,gr)=πk(g1h1,,grhr)+πk(h1,,hr).\pi_{k}(g_{1},\ldots,g_{r})=\pi_{k}(g_{1}-h_{1},\ldots,g_{r}-h_{r})+\pi_{k}(h_{1},\ldots,h_{r}).

The key lemma shows that the first term belongs to Yk1Y_{k-1}, while we have just seen that the second belongs to tmYk+mt^{m}Y_{k+m}. ∎

Proof of Theorem 5.1.

The sets ZkZ_{k} form a descending chain in the finite dimensional space E¯\overline{E}, and so they stabilize. Let \ell be the point at which they stabilize, so that Z+n=ZZ_{\ell+n}=Z_{\ell} for all nn. We claim that Yk=YY_{k}=Y_{\ell} for all kk\geq\ell, which will establish the result: indeed, Satd(I)\operatorname{Sat}_{\leq d}(I) will then be generated by II and YkY_{k}, and since Yk/Y0=Yk/IdY_{k}/Y_{0}=Y_{k}/I_{d} is finite dimensional, we are only adding finitely many generators to II. We prove this by applying Proposition 4.1 with M=YM=Y_{\ell}. We check the three axioms:

(a) Suppose x0,x1,Yx_{0},x_{1},\ldots\in Y_{\ell}. We can thus write xi=t(g1,if1++gr,ifr)x_{i}=t^{-\ell}(g_{1,i}f_{1}+\cdots+g_{r,i}f_{r}), and so i0xiti=t(g1f1++grfr)\sum_{i\geq 0}x_{i}t^{i}=t^{-\ell}(g_{1}f_{1}+\cdots+g_{r}f_{r}) where gj=i0gj,itig_{j}=\sum_{i\geq 0}g_{j,i}t^{i}. This clearly belongs to YY_{\ell}.

(b) We have Σk(Y)=Yk+\Sigma_{k}(Y_{\ell})=Y_{k+\ell}, and we have already seen that Yk+/YY_{k+\ell}/Y_{\ell} is finite dimensional for all kk.

(c) Let (g1,,gr)Xk(g_{1},\ldots,g_{r})\in X_{k} with k>k>\ell. Then ρ(g1,,gr)Zk+m\rho(g_{1},\ldots,g_{r})\in Z_{k+m} for all mm. Thus, by the previous lemma, we have πk(g1,,gr)Yk1+tmYk+m\pi_{k}(g_{1},\ldots,g_{r})\in Y_{k-1}+t^{m}Y_{k+m}. We thus see that YkYk1+tmYk+mYk1+tmRY_{k}\subset Y_{k-1}+t^{m}Y_{k+m}\subset Y_{k-1}+t^{m}R. Applying this inductively, we find YkY+tmRY_{k}\subset Y_{\ell}+t^{m}R, for any m0m\geq 0. Thus YkY_{k} is contained in the tt-adic closure of YY_{\ell}, for all kk\geq\ell.

Proposition 4.1 now applies, and shows that YY_{\ell} is saturated. Thus Yk=YY_{k}=Y_{\ell} for all kk\geq\ell, which establishes the result. ∎

6. Proof of the main theorem

Before proving the main theorem, we need the following simple result on strength.

Proposition 6.1.

Let ff\in\mathcal{R} be homogeneous. Then the following are equivalent:

  1. (a)

    ff has finite strength.

  2. (b)

    The ideal of \mathcal{R} generated by the partial derivatives of ff is contained in an ideal generated by finitely many homogeneous elements of positive degrees.

The same statement holds for +\mathcal{R}^{+} and 0\mathcal{R}^{0}.

Proof.

Suppose (a) holds, and write f=j=1ngjhjf=\sum_{j=1}^{n}g_{j}h_{j}, where each gjg_{j} and hjh_{j} is homogeneous of positive degree. Letting i=xi\partial_{i}=\frac{\partial}{\partial x_{i}}, we have

i(f)=j=1n(i(gj)hj+gji(hj))(g1,,gn,h1,,hn).\partial_{i}(f)=\sum_{j=1}^{n}(\partial_{i}(g_{j})h_{j}+g_{j}\partial_{i}(h_{j}))\in(g_{1},\ldots,g_{n},h_{1},\ldots,h_{n}).

Thus (b) holds.

Now suppose (b) holds. Let g1,,gng_{1},\ldots,g_{n} be positive degree homogeneous elements such that if(g1,,gn)\partial_{i}f\in(g_{1},\ldots,g_{n}) for all ii. Write if=j=1nhi,jgj\partial_{i}f=\sum_{j=1}^{n}h_{i,j}g_{j}. Then by Euler’s identity, we have

f=1di1xiif=j=1ngjhj,f=\frac{1}{d}\sum_{i\geq 1}x_{i}\partial_{i}f=\sum_{j=1}^{n}g_{j}h_{j},

where hj=1di1xihi,jh_{j}=\tfrac{1}{d}\sum_{i\geq 1}x_{i}h_{i,j} and d=deg(f)d=\deg(f), and so ff has strength n\leq n. ∎

Remark 6.2.

The proof given above for (a) \implies (b) is valid in \mathcal{R}^{\flat}. The proof of the converse is not valid in \mathcal{R}^{\flat}, though: the problem is that there is no apparent reason for i1xihi,j\sum_{i\geq 1}x_{i}h_{i,j} to have bounded coefficients. However, the converse direction still holds in \mathcal{R}^{\flat}, and can be proved using a variant of our next argument. ∎

Proof of Theorem 1.3.

Let ff\in\mathcal{R}^{\flat} be given such that ff has finite strength in \mathcal{R}. We show that ff has finite strength in \mathcal{R}^{\flat}. We may as well scale ff by a power of tt and assume that f+f\in\mathcal{R}^{+}. We will in fact show that ff has finite strength in +\mathcal{R}^{+}.

Let I+I\subset\mathcal{R}^{+} be the ideal generated by the partial derivatives of ff. Then II\mathcal{R} is the ideal of \mathcal{R} generated by the partial derivatives of ff. Since ff has finite strength in \mathcal{R}, Proposition 6.1 implies that II\mathcal{R} is contained in a finitely generated non-unital ideal. Since \mathcal{R} is a polynomial ring, it follows that II\mathcal{R} has finite height (Proposition 2.2). By Proposition 3.4, we see that II\mathcal{R}^{\flat} has finite height. Since \mathcal{R}^{\flat} is a polynomial ring, it follows from Proposition 2.2 that II\mathcal{R}^{\flat} is contained in a finitely generated non-unital ideal J=(g1,,gr)J^{\prime}=(g_{1},\ldots,g_{r}). Since II\mathcal{R}^{\flat} is homogeneous, we can assume that the gig_{i}’s are homogeneous of positive degree. Scale each gig_{i} by a power of tt if necessary so that gi+g_{i}\in\mathcal{R}^{+}.

Now, II is contained in the contraction of JJ^{\prime} to +\mathcal{R}^{+}, which is exactly the saturation of the ideal JJ of +\mathcal{R}^{+} generated by g1,,grg_{1},\ldots,g_{r}. Thus ISat(J)I\subset\operatorname{Sat}(J). If ff has degree dd then II is generated by elements of degree d1d-1, and so we have ISatd1(J)I\subset\operatorname{Sat}_{\leq d-1}(J). This is finitely generated by Theorem 5.1; note that, in the notation of §5, if S=0S=\mathcal{R}^{0} then R+R\cong\mathcal{R}^{+}. Since II has no non-zero degree 0 elements, the same is true for Satd1(J)\operatorname{Sat}_{\leq d-1}(J). Thus ff has finite strength in +\mathcal{R}^{+} by Proposition 6.1. ∎

7. Heights in +\mathcal{R}^{+}

We now prove some additional results about heights in the ring +\mathcal{R}^{+}. In what follows, we let Bn=(>n+)(t)B_{n}=(\mathcal{R}^{+}_{>n})_{(t)} be the localization of >n+\mathcal{R}^{+}_{>n} at the prime ideal (t)(t). We note that these rings are all isomorphic to each other.

Proposition 7.1.

Let II be an ideal of +\mathcal{R}^{+} of finite height. Then I>n+t>n+I\cap\mathcal{R}^{+}_{>n}\subset t\mathcal{R}^{+}_{>n} for n0n\gg 0.

Proof.

Since II is contained in a finite height prime, it suffices to treat the case where I=𝔭I=\mathfrak{p} is itself prime. First suppose t𝔭t\in\mathfrak{p}, and let 𝔭¯\overline{\mathfrak{p}} be the extension of 𝔭\mathfrak{p} to 0=+/t+\mathcal{R}^{0}=\mathcal{R}^{+}/t\mathcal{R}^{+}. Then 𝔭¯\overline{\mathfrak{p}} is prime and has finite height, as a chain of primes below 𝔭¯\overline{\mathfrak{p}} would give one below 𝔭\mathfrak{p}, and thus has length bounded by the height of 𝔭\mathfrak{p}. Thus by Proposition 3.1, we have 𝔭¯>n0=0\overline{\mathfrak{p}}\cap\mathcal{R}^{0}_{>n}=0 for n0n\gg 0. This gives 𝔭>n+t>n+\mathfrak{p}\cap\mathcal{R}^{+}_{>n}\subset t\mathcal{R}^{+}_{>n}, as required.

Next suppose t𝔭t\not\in\mathfrak{p}. Then 𝔭\mathfrak{p} is the contraction of a prime 𝔮\mathfrak{q} of =+[1/t]\mathcal{R}^{\flat}=\mathcal{R}^{+}[1/t], necessarily of finite height by Proposition 2.8. Appealing to Proposition 3.1 again, we have 𝔮>n=0\mathfrak{q}\cap\mathcal{R}^{\flat}_{>n}=0 for n0n\gg 0. This implies 𝔭>n+=0\mathfrak{p}\cap\mathcal{R}^{+}_{>n}=0 for n0n\gg 0. ∎

Corollary 7.2.

Let 𝔭\mathfrak{p} be a finite height prime of +\mathcal{R}^{+}. Then 𝔭\mathfrak{p} is the contraction of a prime of Bn[x1,,xn]B_{n}[x_{1},\ldots,x_{n}] for all sufficiently large nn.

Proposition 7.3.

The ring BnB_{n} is a DVR containing 𝐂t\mathbf{C}\llbracket t\rrbracket, and has tt for a uniformizer.

Proof.

Let ff be a non-zero element of BB. Then we can write f=a/bf=a/b where a,b>n+a,b\in\mathcal{R}^{+}_{>n} and bt+b\not\in t\mathcal{R}^{+}. Write a=tka0a=t^{k}a_{0} where a0t>n+a_{0}\not\in t\mathcal{R}^{+}_{>n}; note that kk is the minimal non-negative integer such that tkt^{k} divides all coefficients of aa. Then f=tk(a0/b)f=t^{k}(a_{0}/b), and a0/ba_{0}/b is a unit of BnB_{n}. Thus every non-zero element of BnB_{n} has the form utnut^{n} for uu a unit, which proves the claim. ∎

Proposition 7.4.

Let 𝔭𝔮\mathfrak{p}\subset\mathfrak{q} be finite height primes of +\mathcal{R}^{+}. Then any two maximal chains of primes between 𝔭\mathfrak{p} and 𝔮\mathfrak{q} have the same length.

Proof.

Let 𝔭=𝔞0𝔞r=𝔮\mathfrak{p}=\mathfrak{a}_{0}\subset\cdots\subset\mathfrak{a}_{r}=\mathfrak{q} and 𝔭=𝔟0𝔟s=𝔮\mathfrak{p}=\mathfrak{b}_{0}\subset\cdots\subset\mathfrak{b}_{s}=\mathfrak{q} be two maximal chains. Note that rr and ss are finite since they are bounded by ht+(𝔮)\operatorname{ht}_{\mathcal{R}^{+}}(\mathfrak{q}), which is finite. Let nn be sufficiently large so that the 𝔞i\mathfrak{a}_{i} and 𝔟j\mathfrak{b}_{j} are all contracted from Bn[x1,,xn]B_{n}[x_{1},\ldots,x_{n}]. The extensions of these two chains are both maximal chains between the extensions of 𝔭\mathfrak{p} and 𝔮\mathfrak{q} in Bn[x1,,xn]B_{n}[x_{1},\ldots,x_{n}]. They thus have the same length since Bn[x1,,xn]B_{n}[x_{1},\ldots,x_{n}] is a catenary ring. (Any DVR is universally catenary, see [Stacks, Tag 00NM].) ∎

Corollary 7.5.

Let 𝔭\mathfrak{p} be a prime of +\mathcal{R}^{+} of finite height. Then any maximal chain 𝔭0𝔭r=𝔭\mathfrak{p}_{0}\subset\cdots\subset\mathfrak{p}_{r}=\mathfrak{p} has length r=ht+(𝔭)r=\operatorname{ht}_{\mathcal{R}^{+}}(\mathfrak{p}).

Proposition 7.6.

Let II be an ideal of +\mathcal{R}^{+} of finite height that contains tt, and let I¯=I0\overline{I}=I\mathcal{R}^{0}. Then

ht+(I)=ht0(I¯)+1.\operatorname{ht}_{\mathcal{R}^{+}}(I)=\operatorname{ht}_{\mathcal{R}^{0}}(\overline{I})+1.
Proof.

First suppose that I=𝔭I=\mathfrak{p} is prime. Let c=ht0(𝔭¯)c=\operatorname{ht}_{\mathcal{R}^{0}}(\overline{\mathfrak{p}}), and let 0=𝔭¯0𝔭¯c=𝔭¯0=\overline{\mathfrak{p}}_{0}\subset\cdots\subset\overline{\mathfrak{p}}_{c}=\overline{\mathfrak{p}} be a maximal chain of primes. Let 𝔭i+1\mathfrak{p}_{i+1} be the inverse image of 𝔭i\mathfrak{p}_{i} in +\mathcal{R}^{+}; note that 𝔭1=(t)\mathfrak{p}_{1}=(t). Put 𝔭0=0\mathfrak{p}_{0}=0. Then 𝔭0𝔭c+1=𝔭\mathfrak{p}_{0}\subset\cdots\subset\mathfrak{p}_{c+1}=\mathfrak{p} is a maximal chain of primes in +\mathcal{R}^{+}, and so ht+(𝔭)=c+1\operatorname{ht}_{\mathcal{R}^{+}}(\mathfrak{p})=c+1 by Corollary 7.5.

Now let II be an arbitrary ideal containing tt of height c<c<\infty, and let d=ht0(I¯)d=\operatorname{ht}_{\mathcal{R}^{0}}(\overline{I}). Let 𝔭\mathfrak{p} be a height cc prime of +\mathcal{R}^{+} containing II. Then 𝔭\mathfrak{p} contains tt, and so ht0(𝔭¯)=c1\operatorname{ht}_{\mathcal{R}^{0}}(\overline{\mathfrak{p}})=c-1. Since I¯𝔭¯\overline{I}\subset\overline{\mathfrak{p}}, we find dc1d\leq c-1. Conversely, suppose that 𝔭¯\overline{\mathfrak{p}} is a height dd of 0\mathcal{R}^{0} containing I¯\overline{I}. Then its inverse image 𝔭\mathfrak{p} has height d+1d+1 and contains II, and so cd+1c\leq d+1. This completes the proof. ∎

Proposition 7.7.

The ring +\mathcal{R}^{+} satisfies condition ()(\ast) of Proposition 2.5.

Proof.

Let II be an ideal of +\mathcal{R}^{+} of height c<c<\infty. Let SS be the set of primes of +\mathcal{R}^{+} of height cc that contain II. We must show that SS is finite. Let S1S_{1} be the set of 𝔭S\mathfrak{p}\in S such that t𝔭t\not\in\mathfrak{p}, and let S2S_{2} be the complement. We show that S1S_{1} and S2S_{2} are each finite.

Suppose S1S_{1} is non-empty. We first claim that ht(I)=c\operatorname{ht}_{\mathcal{R}^{\flat}}(I\mathcal{R}^{\flat})=c. We have cht(I)c\leq\operatorname{ht}_{\mathcal{R}^{\flat}}(I\mathcal{R}^{\flat}) by Corollary 2.9. For 𝔭S1\mathfrak{p}\in S_{1} we have I𝔭I\mathcal{R}^{\flat}\subset\mathfrak{p}\mathcal{R}^{\flat} and 𝔭\mathfrak{p}\mathcal{R}^{\flat} has height cc by Proposition 2.8. Thus ht(I)c\operatorname{ht}_{\mathcal{R}^{\flat}}(I\mathcal{R}^{\flat})\leq c, which proves the claim. The same reasoning shows that S1S_{1} is in bijection with the set of height cc primes of \mathcal{R}^{\flat} containing II\mathcal{R}^{\flat}. Since condition ()(\ast) holds for \mathcal{R}^{\flat} (Proposition 2.6), it follows that S1S_{1} is finite.

Suppose S2S_{2} is non-empty. Let J=I+(t)J=I+(t). Then IJ𝔭I\subset J\subset\mathfrak{p} for any 𝔭S2\mathfrak{p}\in S_{2}. Since II and 𝔭\mathfrak{p} have height cc, it follows that JJ has height cc. Let J¯=J0\overline{J}=J\mathcal{R}^{0}, which has height c1c-1 by the Proposition 7.6. By Proposition 7.6, we see that S2S_{2} is in bijection with the height c1c-1 primes of 0\mathcal{R}^{0} containing J¯\overline{J}. Since ()(\ast) holds for 0\mathcal{R}^{0} (Proposition 2.6), it follows that S2S_{2} is finite. ∎

Corollary 7.8.

Let II be an ideal of +\mathcal{R}^{+} and let I=αJαI=\bigcup_{\alpha\in\mathcal{I}}J_{\alpha} be a directed union. Then ht+(I)=supαht+(Jα)\operatorname{ht}_{\mathcal{R}^{+}}(I)=\sup_{\alpha\in\mathcal{I}}\operatorname{ht}_{\mathcal{R}^{+}}(J_{\alpha}).

Proof.

This follows from Proposition 2.5. ∎

Proposition 7.9.

Let II be a finitely generated non-unital ideal of +\mathcal{R}^{+}. Then II has finite height.

Proof.

First suppose that J=I+(t)J=I+(t) is not the unit ideal. It is finitely generated, and so J¯=J0\overline{J}=J\mathcal{R}^{0} is finitely generated. By Proposition 7.6, we have ht+(J)=ht0(J¯)+1\operatorname{ht}_{\mathcal{R}^{+}}(J)=\operatorname{ht}_{\mathcal{R}^{0}}(\overline{J})+1. Since J¯\overline{J} is a finitely generated ideal in the polynomial ring 0\mathcal{R}^{0}, it has finite height (Proposition 2.2). Thus JJ has finite height, and so II does as well.

Now suppose that I+(t)I+(t) is the unit ideal. Then tt is a unit in +/I\mathcal{R}^{+}/I, and so II does not contain any power of tt. It follows that II\mathcal{R}^{\flat} is a finitely generated non-unital ideal. It is thus finite height (Proposition 7.6), and so is contained in a finite height prime 𝔭\mathfrak{p}. Thus II is contained in the contraction of 𝔭\mathfrak{p}, which has finite height (Proposition 2.8), and so II has finite height. ∎

Proposition 7.10 (Hauptidealsatz).

Let II be an ideal of +\mathcal{R}^{+} of finite height cc and let f+f\in\mathcal{R}^{+}. Suppose that I+(f)I+(f) is not the unit ideal. Then I+(f)I+(f) has height c+1\leq c+1.

Proof.

First suppose that II is finitely generated. Then J=I+(f)J=I+(f) is also finitely generated, and thus has finite height by the previous proposition. Let 𝔭\mathfrak{p} be a height cc prime containing II and let 𝔮\mathfrak{q} be a finite height prime containing JJ. Let nn be such that 𝔭\mathfrak{p} and 𝔮\mathfrak{q} are contracted from Bn[x1,,xn]B_{n}[x_{1},\ldots,x_{n}]. The extension II^{\prime} of II to Bn[x1,,xn]B_{n}[x_{1},\ldots,x_{n}] has height cc: indeed, it has height at least cc (Corollary 2.9) and is contained in the extension of 𝔭\mathfrak{p}, which has height cc (Proposition 2.8). Furthermore, I+(f)I^{\prime}+(f) is not the unit ideal of Bn[x1,,xn]B_{n}[x_{1},\ldots,x_{n}], as it is contained in the extension of the prime 𝔮\mathfrak{q}. Thus, by the classical Hauptidealsatz, I+(f)I^{\prime}+(f) has height c+1\leq c+1. Let 𝔓\mathfrak{P} be a prime of Bn[x1,,xn]B_{n}[x_{1},\ldots,x_{n}] of height c+1\leq c+1 containing I+(f)I^{\prime}+(f). Then the contraction of 𝔓\mathfrak{P} to +\mathcal{R}^{+} has height c+1\leq c+1 (Proposition 2.8) and contains J=I+(f)J=I+(f). Thus ht+(J)c+1\operatorname{ht}_{\mathcal{R}^{+}}(J)\leq c+1.

We now treat the general case. Write I=αJαI=\bigcup_{\alpha\in\mathcal{I}}J_{\alpha} (directed union) with JαJ_{\alpha} a finitely generated ideal contained in II. By Corollary 7.8, ht+(Jα)=c\operatorname{ht}_{\mathcal{R}^{+}}(J_{\alpha})=c for α\alpha sufficiently large; we may as well assume it holds for all α\alpha by passing to a cofinal subset. We have I+(f)=α(Jα+(f))I+(f)=\bigcup_{\alpha\in\mathcal{I}}(J_{\alpha}+(f)). Thus, applying Corollary 7.8 again, we have ht+(I+(f))=supαht+(Jα+(f))\operatorname{ht}_{\mathcal{R}^{+}}(I+(f))=\sup_{\alpha\in\mathcal{I}}\operatorname{ht}_{\mathcal{R}^{+}}(J_{\alpha}+(f)). By the first paragraph, we have ht+(Jα+(f))c+1\operatorname{ht}_{\mathcal{R}^{+}}(J_{\alpha}+(f))\leq c+1 for all α\alpha. It follows that ht+(I+(f))c+1\operatorname{ht}_{\mathcal{R}^{+}}(I+(f))\leq c+1, which completes the proof. ∎

References

  • [AH] Tigran Ananyan, Melvin Hochster. Small subalgebras of polynomial rings and Stillman’s conjecture. J. Amer. Math. Soc. 33 (2019), 291–309. arXiv:1610.09268v1
  • [BDES] Arhtur Bik, Jan Draisma, Rob Eggermont, Andrew Snowden. The geometry of polynomial representations. In preparation.
  • [DLL] Jan Draisma, Michal Lason, Anton Leykin. Stillman’s conjecture via generic initial ideals. Commun. Alg. 47 (2019), no. 6, 2384–2395. arXiv:1802.10139
  • [ESS] Daniel Erman, Steven V Sam, Andrew Snowden. Big polynomial rings and Stillman’s conjecture. Invent. Math. 218 (2019), no. 2, 413–439 arXiv:1801.09852v4
  • [ESS2] Daniel Erman, Steven V Sam, Andrew Snowden. Big polynomial rings with imperfect coefficient fields. arXiv:1806.04208
  • [Stacks] Stacks Project. http://stacks.math.columbia.edu, 2020.