1 Introduction

Let \((R,\mathfrak {m})\) be a Noetherian local ring of dimension d such that \(\widehat{R}\) is reduced and let \(I\subseteq R\) be an \(\mathfrak {m}\)-primary ideal. Then for \(n\gg 0\), \(\ell (R/\overline{I^{n+1}})\) agrees with a polynomial in n of degree d, and we have integers \(\overline{e}_0(I),\dots , \overline{e}_d(I)\) such that

$$\ell (R/\overline{I^{n+1}})=\overline{e}_0(I)\left( {\begin{array}{c}n+d\\ d\end{array}}\right) - \overline{e}_1(I)\left( {\begin{array}{c}n+d-1\\ d-1\end{array}}\right) + \cdots + (-1)^d\overline{e}_d(I).$$

These integers \(\overline{e}_i(I)\) are called the normal Hilbert coefficients of I.

It is well-known that \(\overline{e}_0(I)\) is the Hilbert-Samuel multiplicity of I, which is always a positive integer. In this paper, we are interested in the first coefficient \(\overline{e}_1(I)\). It was proved by Goto-Hong-Mandal [10] that when \(\widehat{R}\) is unmixed, \(\overline{e}_1(I)\ge 0\) for all \(\mathfrak {m}\)-primary ideals \(I\subseteq R\) (which answers a question posed by Vasconcelos [30]). They proposed a further problem in [10, Section 3] regarding the vanishing of \(\overline{e}_1(I)\) and the regularity of the normalization of R. Since any \(\mathfrak {m}\)-primary ideal I is integral over a parameter ideal when the residue field is infinite, to study \(\overline{e}_1(I)\) we may assume that \(I=Q\) is a parameter ideal (i.e., it is generated by a system of parameters). In this paper, we prove the following main result which will lead to an affirmative answer to the question proposed in [10]. This theorem is also a generalization of the main result of [27].

Theorem 1.1

(Theorem 3.7) Let \((R,\mathfrak {m})\) be a Noetherian local ring such that \(\widehat{R}\) is reduced and \(S_2\). If \(\overline{e}_1(Q)=0\) for some parameter ideal \(Q\subseteq R\), then R is regular and \(\nu (\mathfrak {m}/Q)\le 1\).

In [7], it was shown that when R has characteristic \(p>0\), for \(n\gg 0\), \(\ell (R/(I^{n+1})^*)\) also agrees with a polynomial of degree d and one can define the tight Hilbert coefficients \(e_0^*(I),\dots , e_d^*(I)\) in a similar way (see Section 2 for more details). It is easy to see that \(\overline{e}_1(I)\ge e_1^*(I)\). We strengthen the main result of [10] in characteristic \(p>0\) by showing that \(e_1^*(Q)\ge 0\) for any parameter ideal \(Q\subseteq R\) under mild assumptions.

Theorem 1.2

(Corollary 3.3) Let \((R,\mathfrak {m})\) be an excellent local ring of characteristic \(p>0\) such that \(\widehat{R}\) is reduced and equidimensional. Then we have \(e^*_1(Q)\ge 0\) for all parameter ideals \(Q\subseteq R\).

Our proofs of both theorems rely on the existence of big Cohen-Macaulay algebras. In fact, we show that the tight Hilbert coefficients \(e_i^*(I)\) is a special case of what we call the BCM Hilbert coefficients \(e_i^B(I)\) associated to a big Cohen-Macaulay algebra B, and the latter can be defined in arbitrary characteristic. In this context, we will show in Theorem 3.1 that \(\overline{e}_1(Q)\ge e_1^B(Q)\ge 0\) for all parameter ideals \(Q\subseteq R\) when B satisfies some mild assumptions. This recovers and extends the main result of [10] in arbitrary characteristic.

Throughout this article, all rings are commutative with multiplicative identity 1. We will use \((R,\mathfrak {m})\) to denote a Noetherian local ring with unique maximal ideal \(\mathfrak {m}\). We refer the reader to [4, Chapter 1-4] for basic notions such as Cohen-Macaulay rings, regular sequence, Euler characteristic, integral closure, and the Hilbert-Samuel multiplicity. We refer the reader to [29, Section 07QS] for the definition and basic properties of excellent rings. The paper is organized as follows. In Section 2 we collect the definitions and some basic results on big Cohen-Macaulay algebras and variants of Hilbert coefficients. In Section 3 we prove our main results and we propose some further questions.

2 Preliminaries

Recall that an element x in a ring R is integral over an ideal \(I\subseteq R\) if it satisfies an equation of the form \(x^n+ a_1x^{n-1}+\cdots + a_{n-1}x + a_n=0\), where \(a_k\in I^k\). The set of all elements integral over I forms an ideal and is denoted by \(\overline{I}\), called the integral closure of I. An ideal \(I\subseteq R\) is called integrally closed if \(I=\overline{I}\). It is well-known that an element \(x\in R\) is integral over I if and only if the image of x in \(R/\mathfrak {p}\) is integral over \(I(R/\mathfrak {p})\) for all minimal primes \(\mathfrak {p}\), see [21, Proposition 1.1.5].

Suppose that R is a Noetherian ring of prime characteristic \(p>0\). The tight closure of an ideal \(I\subseteq R\), introduced by Hochster–Huneke, is defined as follows:

$$\begin{aligned}I^* := \{x\in R \, | \, \text {there exists}\, c\in R-\cup _{\mathfrak {p}\in \text {Min}(R)}\mathfrak {p}\, \text { such that }\, cx^{p^e}\in I^{[p^e]} \text { for all }\, e\gg 0\}.\end{aligned}$$

An ideal \(I\subseteq R\) is called tightly closed if \(I=I^*\). In general, tight closure is always contained in the integral closure, that is, \(I^*\subseteq \overline{I}\) (see [15, Proposition on page 58]). Similar to integral closure, an element \(x\in R\) is in the tight closure of I if and only if the image of x in \(R/\mathfrak {p}\) is in the tight closure of \(I(R/\mathfrak {p})\) for all minimal primes \(\mathfrak {p}\), see [15, Theorem on page 49].

Let R be a Noetherian complete local domain and let \(I\subseteq R\) be an ideal. The solid closure of I, denoted by \(I^\bigstar \), consists of those element \(x\in R\) such that there exists an R-algebra S such that \({\text {Hom}}_R(S, R)\ne 0\) and such that \(x\in IS\). One can define solid closure of ideals in more general rings, see [16, Definition 1.2], but we will only need this notion for complete local domains. It was shown in [16, Theorem 5.10] that solid closure is contained in the integral closure, i.e., \(I^\bigstar \subseteq \overline{I}\). If R has prime characteristic \(p>0\), then solid closure agrees with tight closure \(I^\bigstar = I^*\), see [16, Theorem 8.6].

2.1 Big Cohen-Macaulay Algebras

Let \((R,\mathfrak {m})\) be a Noetherian local ring. An R-algebra B, not necessarily Noetherian, is called balanced big Cohen-Macaulay over R if every system of parameters of R is a regular sequence on B and \(\mathfrak {m} B\ne B\). Balanced big Cohen-Macaulay algebras exist, in equal characteristic, this is due to Hochster-Huneke [18], and in mixed characteristic, this is proved by André [1] (see also [2, 3, 12]). In this article, we need to compare the closure operation induced by a balanced big Cohen-Macaulay algebra with integral closure. We begin with the following result.

In what follows, when \(R\rightarrow S\) is a (not necessarily injective) homomorphism of rings, \(IS\cap R\) should be interpreted as the contraction of IS to R. That is, those elements of R whose image in S are contained in IS.

Lemma 2.1

Let \((R,\mathfrak {m})\) be a Noetherian local ring. Then the following conditions are equivalent:

  1. (1)

    \(\widehat{R}\) is equidimensional.

  2. (2)

    There exists a balanced big Cohen-Macaulay R-algebra B such that

    figure a
  3. (3)

    There exists a balanced big Cohen-Macaulay R-algebra B such that \(I^B\subseteq \overline{I}\) for all \(I\subseteq R\).

Proof

Since \((3)\Rightarrow (2)\) is obvious, we only need to show \((1)\Rightarrow (3)\) and \((2)\Rightarrow (1)\). Suppose that \(\widehat{R}\) is equidimensional and let \(P_1,\dots ,P_n\) be the minimal primes of \(\widehat{R}\). Let \(B_i\) be any balanced big Cohen-Macaulay algebra over \(\widehat{R}/P_i\). Since \(\widehat{R}\) is equidimensional, each system of parameters of \(\widehat{R}\) is also a system of parameters of \(\widehat{R}/P_i\) and thus \(B_i\) is a balanced big Cohen-Macaulay algebra over \(\widehat{R}\). It follows that \(B:= \prod _{i=1}^{n}B_i\) is a balanced big Cohen-Macaulay algebra over \(\widehat{R}\).

Claim 2.2

\((I\widehat{R})^B= IB\cap \widehat{R} \subseteq \overline{I\widehat{R}}\).

Proof of the Claim

Since integral closure can be checked after modulo each minimal prime, it suffices to show that \((I\widehat{R})^B\cdot (\widehat{R}/P_i)\subseteq \overline{I(\widehat{R}/P_i)}\). It is easy to see (by our construction of B) that

$$(I\widehat{R})^B\cdot (\widehat{R}/P_i) = (I(\widehat{R}/P_i))^{B_i}.$$

Since \(B_i\) is a solid algebra over the complete local domain \(\widehat{R}/P_i\) by [16, Corollary 2.4], we have

$$(I(\widehat{R}/P_i))^{B_i} \subseteq (I(\widehat{R}/P_i))^\bigstar \subseteq \overline{I(\widehat{R}/P_i)},$$

where the second inclusion follows from [16, Theorem 5.10].\(\square \)

By the claim above, we have

$$I^B\subseteq (I\widehat{R})^B \cap R \subseteq \overline{I\widehat{R}} \cap R=\overline{I},$$

where the last equality follows from [21, Proposition 1.6.2].

We next assume there exists a balanced big Cohen-Macaulay R-algebra B that satisfies (\(\dagger \)). We first note that \(\widehat{B}\) (the \(\mathfrak {m}\)-adic completion of B) is still a balanced big Cohen-Macaulay algebra over \(\widehat{R}\) by [4, Corollary 8.5.3]. If I is an \(\mathfrak {m}\)-primary ideal, then we have \(R/I \cong \widehat{R}/I\widehat{R} \) and \(B/IB \cong \widehat{B}/I\widehat{B}\) (see [29, Tag 05GG]). It follows that \((I\widehat{R})^{\widehat{B}} = (I^B)\widehat{R} \subseteq \overline{I} \widehat{R} = \overline{I\widehat{R}}\) (where the last equality follows from [21, Lemma 9.1.1]). Thus without loss of generality, we may replace R by \(\widehat{R}\) and B by \(\widehat{B}\) to assume that R is complete. Suppose that R is not equidimensional. Let \(P_1, \ldots , P_n\) be all the minimal primes of R such that \(\dim (R/P_i) = \dim (R)\), and \(Q_1, \ldots , Q_m\) be all the minimal primes of R such that \(\dim (R/Q_j) < d\). We pick \(y \in Q_1 \cap \cdots \cap Q_m \setminus P_1\cup \cdots \cup P_n\). Then y is a parameter element in R, and thus y is a nonzerodivisor on B, since B is balanced big Cohen-Macaulay. Since \(y \cdot (P_1 \cap \cdots \cap P_n) \subseteq \sqrt{0}\), there exists t such that \(y^t \cdot (P_1 \cap \cdots \cap P_n)^t=0\). It follows that \((P_1 \cap \cdots \cap P_n)^tB = 0\). Hence

$$(P_1 \cap \cdots \cap P_n)^t \subseteq \mathfrak {m}^k B \cap R \subseteq \overline{\mathfrak {m}^k}$$

for all k by (\(\dagger \)). Thus \((P_1 \cap \cdots \cap P_n)^t\subseteq \cap _k \overline{\mathfrak {m}^k} = \sqrt{0}\) by [21, Exercise 5.14], which is a contradiction. \(\square \)

Remark 2.3

In the proof of Lemma 2.1, we have proved the fact that when \((R,\mathfrak {m})\) is a Noetherian complete local domain, then every balanced big Cohen-Macaulay algebra B satisfies (\(\dagger \)). We suspect that when \((R,\mathfrak {m})\) is Noetherian, complete, reduced and equidimensional, then every balanced big Cohen-Macaulay algebra B such that \({\text {Supp}}(\widehat{B})={\text {Spec}}(R)\) satisfies (\(\dagger \)).

2.2 Hilbert Coefficients

Let \((R, \mathfrak {m})\) be a Noetherian local ring of dimension d and let \(I\subseteq R\) be an \(\mathfrak {m}\)-primary ideal. Then for all \(n \gg 0\) we have

$$\ell (R/I^{n+1}) = e_0(I) \left( {\begin{array}{c}n+d\\ d\end{array}}\right) - e_1(I) \left( {\begin{array}{c}n+d-1\\ d-1\end{array}}\right) + \cdots + (-1)^d e_d(I),$$

where \(e_0(I), \dots , e_d(I)\) are all integers, and are called the Hilbert coefficients of I.

Now suppose that \(R \oplus \overline{I}t \oplus \overline{I^2}t^2 \oplus \cdots \) is module-finite over the Rees algebra R[It]. For instance, by a famous result of Rees (see [21, Corollary 9.2.1]), this is the case when \(\widehat{R}\) is reduced. Then one can show that for all \(n\gg 0\), \(\ell (R/\overline{I^{n+1}})\) agrees with a polynomial in n and one can write

$$\ell (R/\overline{I^{n+1}}) = \overline{e}_0(I) \left( {\begin{array}{c}n+d\\ d\end{array}}\right) - \overline{e}_1(I) \left( {\begin{array}{c}n+d-1\\ d-1\end{array}}\right) + \cdots + (-1)^d \overline{e}_d(I),$$

where the integers \(\overline{e}_0(Q), \dots , \overline{e}_d(Q)\) are called the normal Hilbert coefficients. It is well-known that \(e_0(I)=\overline{e}_0(I)\) agrees with the Hilbert-Samuel multiplicity e(IR) of I.

We also recall the tight Hilbert coefficients studied in [7]. Again, we suppose that \(\widehat{R}\) is reduced and R has characteristic \(p>0\). Then we have

$$\ell (R/(I^{n+1})^*) = e_0^*(I) \left( {\begin{array}{c}n+d\\ d\end{array}}\right) - e_1^*(I) \left( {\begin{array}{c}n+d-1\\ d-1\end{array}}\right) + \cdots + (-1)^d e_d^*(I)$$

for all \(n\gg 0\), and the integers \(e_0^*(I),\dots ,e_d^*(I)\) are called the tight Hilbert coefficients, see [7] for more details.

Now if B is a balanced big Cohen-Macaulay R-algebra that satisfies (\(\dagger \)), then we know that \(R\oplus I^Bt \oplus (I^2)^Bt^2 \oplus \cdots \) is an R-algebra that is also module-finite over R[It]: the fact that it is an R-algebra follows from the fact that \((I^a)^B(I^b)^B\subseteq (I^{a+b})^B\) for all ab (i.e., \(\{(I^n)^B\}_n\) form a graded family of ideals), and that it is module-finite over R[It] follows because by (\(\dagger \)), it is an R[It]-submodule of \(R \oplus \overline{I}t \oplus \overline{I^2}t^2 \oplus \cdots \), and the latter is module-finite over R[It] (note that R[It] is Noetherian). Based on the discussion above, one can show that for all \(n\gg 0\), \(\ell (R/(I^{n+1})^B)\) also agrees with a polynomial in n, and we write

$$\ell (R/(I^{n+1})^B) = e_0^B(I) \left( {\begin{array}{c}n+d\\ d\end{array}}\right) - e_1^B(I) \left( {\begin{array}{c}n+d-1\\ d-1\end{array}}\right) + \cdots + (-1)^d e_d^B(I)$$

for all \(n\gg 0\) (see [19] for more general results). We call the integers \(e_0^B(I),\dots , e_d^B(I)\) the BCM Hilbert coefficients with respect to B. It is easy to see that \(e_0^B(I)=e(I, R)\) is still the Hilbert-Samuel multiplicity of I, and that we always have \(\overline{e}_1(I)\ge e_1^B(I)\ge e_1(I)\) by comparing the coefficients of \(n^{d-1}\) and noting that \(I^n\subseteq (I^n)^B\subseteq \overline{I^n}\) for all n by (\(\dagger \)).

Remark 2.4

We point out that when \((R,\mathfrak {m})\) is excellent and \(\widehat{R}\) is reduced and equidimensional of characteristic \(p>0\), the tight Hilbert coefficient is a particular case of BCM Hilbert coefficient. This follows from the fact that under these assumptions, there exists a balanced big Cohen-Macaulay algebra B such that \(I^*=I^B\) for all \(I\subseteq R\) (and any such B will satisfy (\(\dagger \)), since tight closure is contained in the integral closure [20, Theorem 1.3]). When R is a complete local domain this is proved in [15, Theorem on page 250]. In general, one can take such a \(B_i\) for each complete local domain \(\widehat{R}/P_i\), where \(P_i\) is a minimal prime of \(\widehat{R}\), and let \(B=\prod B_i\). Since R is excellent, \(I^*\widehat{R}=(I\widehat{R})^*\) (see [20, Proposition 1.5]) and as tight closure can be checked after modulo each minimal prime, it follows that \(I^*\widehat{R}=(I\widehat{R})^B\) and thus \(I^*=I^B\).

Throughout the rest of this article, we will be mainly working with parameter ideals, i.e., ideals generated by a system of parameters. As we mentioned in the introduction, this will not affect the study of \(\overline{e}_1(I)\), since we can often enlarge the residue field and replace I by its minimal reduction.

3 The Main Results

In this section we prove our main results that \(e_1^B(Q)\) (and hence \(\overline{e}_1(Q)\)) is always nonnegative for a parameter ideal Q, and that \(\overline{e}_1(Q)=0\) for some parameter ideal Q implies R is regular.

3.1 Non-negativity of \(\bar{e}_1(Q)\) and \(e_1^B(Q)\)

Theorem 3.1

Let \((R,\mathfrak {m})\) be a Noetherian local ring such that \(\widehat{R}\) is reduced and equidimensional. Let B be any balanced big Cohen-Macaulay R-algebra that satisfies (\(\dagger \)). Then for all parameter ideals \(Q\subseteq R\) we have

$$\overline{e}_1(Q) \ge e_1^B(Q) \ge 0 \ge e_1(Q).$$

Remark 3.2

\(\overline{e}_1(Q) \ge 0\) was the main theorem of [10, Theorem 1.1], and \(0 \ge e_1(Q)\) was first proved in full generality in [26, Theorem 3.6]. Our method gives alternative proofs, and is inspired by some work of Goto [9] (in fact the proof that \(e_1(Q)\le 0\) via this method is due to Goto [9], see also [13, Theorem 1.1] for a generalization).

Corollary 3.3

Let \((R,\mathfrak {m})\) be an excellent local ring of characteristic \(p>0\) such that \(\widehat{R}\) is reduced and equidimensional. Then we have \(e^*_1(Q)\ge 0\) for all parameter ideals \(Q\subseteq R\).

Proof

This follows from Theorem 3.1 and Remark 2.4.\(\square \)

Proof of Theorem 3.1

Let \(Q = (x_1, \ldots , x_d) \subseteq R\). Set \(S = R[[y_1, \ldots , y_d]]\) and \(\mathfrak {q} = (y_1 - x_1, \ldots ,\) \( y_d - x_d) \subseteq S\). For all \(n \ge 0\) we have \(y_1, \ldots , y_d\) is a system of parameters on \(S/\mathfrak {q}^{n+1}\), and that

$$R/Q^{n+1} = S/(\mathfrak {q}^{n+1} + (y_1, \ldots , y_d)).$$

We next note that

$$\begin{aligned} e_0(Q) = e(Q, R)= & \chi (x_1, \ldots , x_d; R)\\= & \chi (x_1, \ldots , x_d, y_1, \ldots , y_d;S)\\= & \chi (y_1, \ldots , y_d, y_1-x_1, \ldots , y_d - x_d;S)\\= & \chi (y_1, \ldots , y_d; S/\mathfrak {q})\\= & e(y_1, \ldots , y_d; S/\mathfrak {q}), \end{aligned}$$

where the equalities on the second and the fourth lines follow from the fact that \(y_1,\dots ,y_d\) and \(y_1 - x_1, \ldots , y_d - x_d\) are both regular sequences on S. Now since \(S/\mathfrak {q}^{n+1}\) has a filtration by \(\left( {\begin{array}{c}n+d\\ d\end{array}}\right) \) copies of \(S/\mathfrak {q}\), by the additivity formula for multiplicity (see [21, Theorem 11.2.3]) we have

$$e(y_1, \ldots , y_d; S/\mathfrak {q}^{n+1}) = \left( {\begin{array}{c}n+d\\ d\end{array}}\right) e(y_1, \ldots , y_d; S/\mathfrak {q}).$$

Putting these together, we have

$$\left( {\begin{array}{c}n+d\\ d\end{array}}\right) e_0(Q) = \left( {\begin{array}{c}n+d\\ d\end{array}}\right) e(y_1, \ldots , y_d; S/\mathfrak {q}) = e(y_1, \ldots , y_d; S/\mathfrak {q}^{n+1}), $$

and

$$\ell (R/Q^{n+1}) = \ell \left( \frac{S/\mathfrak {q}^{n+1}}{(y_1, \ldots , y_d)S/\mathfrak {q}^{n+1}} \right) .$$

Since \(y_1,\dots , y_d\) is a system of parameters of \(S/\mathfrak {q}^{n+1}\), we have

$$\ell \left( \frac{S/\mathfrak {q}^{n+1}}{(y_1, \ldots , y_d)S/\mathfrak {q}^{n+1}} \right) \ge e(y_1, \ldots , y_d; S/\mathfrak {q}^{n+1}).$$

It follows that

$$\ell (R/Q^{n+1}) \ge \left( {\begin{array}{c}n+d\\ d\end{array}}\right) e_0(Q),$$

and thus \(e_1(Q) \le 0\) (note that this does not require any assumption on \(\widehat{R}\)).

It remains to show that \(e_1^B(Q) \ge 0\) (since \(\overline{e}_1(Q)\ge e_1^B(Q)\) always holds, see the discussion in Section 2.2). Since \(e_0^B(Q)=e_0(Q)\), it is enough to show that

$$\begin{aligned} \ell (R/(Q^{n+1})^B) \le \left( {\begin{array}{c}n+d\\ d\end{array}}\right) e_0(Q) \end{aligned}$$
(1)

for any balanced big Cohen-Macaulay algebra B. Below we will prove a slightly stronger result. Recall that for a parameter ideal \((z_1,\dots ,z_d)\) of R, the limit closure is defined as \((z_1,\dots ,z_d)^{\lim _R}:= \bigcup _t (z_1^{t+1},\dots ,z_d^{t+1}): (z_1z_2\cdots z_d)^{t}\). The limit closure does not depend on the choice of the elements \(z_1,\dots , z_d\) (i.e., it only depends on the ideal \((z_1,\dots ,z_d)\)). This is because \((z_1,\dots ,z_d)^{\lim _R}/(z_1,\dots ,z_d)\) is the kernel of the natural map \(R/(z_1,\dots , z_d)\rightarrow H_{\mathfrak {m}}^d(R)\).

Claim 3.4

Set \(\varLambda _{n+1}= \{(\alpha _1,\dots , \alpha _d)\in \mathbb {N}^d \, | \, \alpha _i\ge 1 \text { and } \sum _{i=1}^{d} \alpha _i = 1+n \}\) and for each \(\alpha =(\alpha _1,\dots ,\alpha _d) \in \varLambda _{n+1}\), set \(Q(\alpha )=(x_1^{\alpha _1},\dots ,x_d^{\alpha _d})\). Then we have

$$\ell \left( R/(\bigcap _{\alpha \in \varLambda _{n+1}} Q(\alpha )^{\lim _R})\right) \le \left( {\begin{array}{c}n+d\\ d\end{array}}\right) e_0(Q).$$

Proof of the Claim

Recall that we have already proved that

$$\left( {\begin{array}{c}n+d\\ d\end{array}}\right) e_0(Q) = e(y_1, \ldots , y_d; S/\mathfrak {q}^{n+1}).$$

Moreover, we always have (for example, see [23, Theorem 9])

$$e(y_1, \ldots , y_d; S/\mathfrak {q}^{n+1}) \ge \ell \left( \frac{S/\mathfrak {q}^{n+1}}{(y_1, \ldots , y_d)^{\lim _{S/\mathfrak {q}^{n+1}}}} \right) .$$

Therefore it is enough to prove that

$$\begin{aligned} \ell \left( \frac{S/\mathfrak {q}^{n+1}}{(y_1, \ldots , y_d)^{\lim _{S/\mathfrak {q}^{n+1}}}} \right) \ge \ell \left( R/(\bigcap _{\alpha \in \varLambda _{n+1}} Q(\alpha )^{\lim _R})\right) . \end{aligned}$$
(2)

Consider \(z \in S\) whose image in \(S/\mathfrak {q}^{n+1}\) is contained in \((y_1, \ldots , y_d)^{\lim _{S/\mathfrak {q}^{n+1}}}\). This means there exists some \(t \ge 1\) such that

$$\begin{aligned} (y_1 y_2\cdots y_d)^{t} z\in & (y_1^{t+1}, \ldots , y_d^{t+1}, (y_1 - x_1, \ldots , y_d - x_d)^{n+1})\\\subseteq & (y_1^{t+1}, \ldots , y_d^{t+1}, (y_1 - x_1)^{\alpha _1}, \ldots , (y_d - x_d)^{\alpha _d}) \end{aligned}$$

for each \(\alpha =(\alpha _1,\dots ,\alpha _d)\in \varLambda _{n+1}\). This implies

$$z \in (y_1, \ldots , y_d, (y_1 - x_1)^{\alpha _1}, \ldots , (y_d - x_d)^{\alpha _d})^{\lim _S} = (y_1, \ldots , y_d, x_1^{\alpha _1}, \ldots , x_d^{\alpha _d})^{\lim _S}.$$

But since \(S = R[[y_1, \ldots , y_d]]\), it is straightforward to check that

$$(y_1, \ldots , y_d, x_1^{\alpha _1}, \ldots , x_d^{\alpha _d})^{\lim _S} = (x_1^{\alpha _1}, \ldots , x_d^{\alpha _d})^{\lim _R}S + (y_1, \ldots , y_d)S.$$

Thus if the image of z is contained in \((y_1, \ldots , y_d)^{\lim _{S/\mathfrak {q}^{n+1}}}\), then after modulo \((y_1, \ldots , y_d)S\), \(\overline{z} \in (x_1^{\alpha _1}, \ldots , x_d^{\alpha _d})^{\lim _R}\) for each \((\alpha _1,\dots ,\alpha _d)\in \varLambda _{n+1}\), i.e., \(\overline{z} \in \bigcap _{\alpha \in \varLambda _{n+1}} Q(\alpha )^{\lim _R}\). It follows that the natural surjection

$$S/\mathfrak {q}^{n+1} \xrightarrow {\mod (y_1, \ldots , y_d)S} R/Q^{n+1}$$

induces a surjection

$$\frac{S/\mathfrak {q}^{n+1}}{(y_1, \ldots , y_d)^{\lim _{S/\mathfrak {q}^{n+1}}}} \twoheadrightarrow \frac{R}{\bigcap _{\alpha \in \varLambda _{n+1}} Q(\alpha )^{\lim _R}}.$$

This clearly establishes (2) and completes the proof of the claim.\(\square \)

Finally, since \(x_1,\dots , x_d\) is a regular sequence on B, we have \(Q(\alpha )^{\lim _R}\subseteq Q(\alpha )^B\) for each \(\alpha \). It follows that \(\bigcap _{\alpha \in \varLambda _{n+1}} Q(\alpha )^{\lim _R} \subseteq \bigcap _{\alpha \in \varLambda _{n+1}} Q(\alpha )^B\). Now if \(x \in \bigcap _{\alpha \in \varLambda _{n+1}} Q(\alpha )^B\), then we have \(x \in \big (\bigcap _{\alpha \in \varLambda _{n+1}} Q(\alpha )B \big ) \cap R\). But since \(x_1,\dots ,x_d\) is a regular sequence on B, it is not hard to check that \(\bigcap _{\alpha \in \varLambda _{n+1}} Q(\alpha )B = Q^{n+1}B\) (see [28, Remark 3.3] or [11]) and thus \(x\in Q^{n+1}B \cap R = (Q^{n+1})^B\). Therefore we have \(\bigcap _{\alpha \in \varLambda _{n+1}} Q(\alpha )^B = (Q^{n+1})^B\). Putting these together, we have

$$\bigcap _{\alpha \in \varLambda _{n+1}} Q(\alpha )^{\lim _R} \subseteq \bigcap _{\alpha \in \varLambda _{n+1}} Q(\alpha )^B = (Q^{n+1})^B.$$

Therefore by Claim 3.4, we have

$$\ell (R/(Q^{n+1})^B) \le \left( {\begin{array}{c}n+d\\ d\end{array}}\right) e_0(Q)$$

as wanted.\(\square \)

Remark 3.5

With notation as in Theorem 3.1, we do not know whether we have

$$\ell \left( \frac{S/\mathfrak {q}^{n+1}}{(y_1, \ldots , y_d)^{\lim _{S/\mathfrak {q}^{n+1}}}}\right) = \ell \left( \frac{R}{\bigcap _{\alpha \in \varLambda _{n+1}} Q(\alpha )^{\lim _R}}\right) .$$

Remark 3.6

With notation as in Claim 3.4, fix a generating set \((x_1,\dots ,x_d)\) of Q, one may try to define \((Q^n)^{\lim } := \bigcap _{\alpha \in \varLambda _{n}} Q(\alpha )^{\lim }\) and call this the limit closure of \(Q^n\). However, it is not clear to us whether this is independent of the choice of the generators \(x_1,\dots ,x_d\). It is also not clear to us (even when fixing the generators \((x_1,\dots ,x_d)\) of Q) whether \(\{(Q^n)^{\lim }\}_n\) form a graded family of ideals, i.e., we do not know whether \((Q^a)^{\lim } (Q^b)^{\lim } \subseteq (Q^{a+b})^{\lim }\) for all ab.

3.2 Vanishing of \(\bar{e}_1(Q)\)

In this subsection we prove our main result. Recall that for a finitely generated R-module M, we use the notation \(\nu (M)\) to denote its minimal number of generators.

Theorem 3.7

Let \((R,\mathfrak {m})\) be a Noetherian local ring such that \(\widehat{R}\) is reduced and \(S_2\). If \(\overline{e}_1(Q)=0\) for some parameter ideal \(Q\subseteq R\), then R is regular and \(\nu (\mathfrak {m}/Q)\le 1\).

Proof

We first note that if R is Cohen-Macaulay, then by [19, Corollary 4.9], Q is integrally closed.Footnote 1 But then by the main result of [8], R is regular and \(\nu (\mathfrak {m}/Q)\le 1\).

We may assume that R is complete. We use induction on \(d:=\dim (R)\). If \(d\le 2\), then R is Cohen-Macaulay and we are done by the previous paragraph. Now suppose that \(d\ge 3\) and we have established the theorem in dimension \(<d\). Let \(Q=(x_1,\dots ,x_d)\), \(R^\prime =\) \(R[t_1,\dots ,t_d]_{\mathfrak {m} R[t_1,\dots ,t_d]}\), and \(x=t_1x_1+ \cdots +t_dx_d\).

Claim 3.8

We have \(R^{\prime \prime }:=\widehat{R^\prime }/x\widehat{R^\prime }\) is reduced, equidimensional, and \(S_2\) on the punctured spectrum. Moreover, we have \(\overline{e}_1(QR^{\prime \prime })=0\).

Proof

This is essentially contained in [10, Proof of Theorem 1.1] under the assumption that R is (complete and) normal. The key ingredient is [22, Theorem 2.1]. Since [22] does not require the normal assumption, the same proof as in [10] works in our setup. For the ease of the reader (and also because the \(S_2\) on the punctured spectrum conclusion is not stated in [10]), we give a complete and self-contained argument here.

First of all, since \(R^\prime \) is \(S_2\) and \(R_0\), we know that \(R^\prime /xR^\prime \) is \(S_1\) and \(R_0\) (see [25, Lemma 10]), so \(R^\prime /xR^\prime \) and thus \(R^{\prime \prime }\) is reduced (as \(R^\prime /xR^\prime \) is excellent). \(R^{\prime \prime }\) is clearly equidimensional since \(\widehat{R^\prime }\) is so and x is a parameter in \(\widehat{R^\prime }\). To see \(R^{\prime \prime }\) is \(S_2\) on the punctured spectrum, it is enough to show \(R^\prime /xR^\prime \) is \(S_2\) on the punctured spectrum (as \(R^\prime /xR^\prime \) is excellent). Now we use a similar argument as in [25, Lemma 10] (the idea follows from [14]): every non-maximal \(P^\prime \in {\text {Spec}}(R^\prime /xR^\prime )\) corresponds to a prime ideal of \(R^\prime \) that contracts to a non-maximal \(P\in {\text {Spec}}(R)\), thus \((R^\prime /xR^\prime )_{P^\prime }\) is a localization of \(R_P[t_1,\dots ,t_d]/(t_1x_1+\cdots + t_dx_d)\), but at least one \(x_i\) is invertible in \(R_P\) (say \(x_1\) is invertible) so the latter is isomorphic to \(R_P[t_2,\dots , t_d]\), which is \(S_2\) as \(R_P\) is \(S_2\), thus \(R^\prime /xR^\prime \) is \(S_2\) on the punctured spectrum as wanted.

It remains to show that \(\overline{e}_1(QR^{\prime \prime })=0\). By [21, Corollary 6.8.13], we have a short exact sequence

$$0\rightarrow R^\prime /\overline{Q^n} \xrightarrow {\cdot x} R^\prime /\overline{Q^{n+1}} \rightarrow R^\prime /(x, \overline{Q^{n+1}})\rightarrow 0.$$

Since \(\overline{e}_1(Q)=0\), for \(n\gg 0\) we have

$$\ell (R^\prime /\overline{Q^{n+1}})=\overline{e}_0(Q)\cdot \left( {\begin{array}{c}n+d\\ d\end{array}}\right) + \overline{e}_2(Q)\cdot \left( {\begin{array}{c}n+d-2\\ d-2\end{array}}\right) + o(n^{d-2}),$$
$$\ell (R^\prime /\overline{Q^n})=\overline{e}_0(Q)\cdot \left( {\begin{array}{c}n+d-1\\ d\end{array}}\right) + \overline{e}_2(Q)\cdot \left( {\begin{array}{c}n+d-3\\ d-2\end{array}}\right) + o(n^{d-2}).$$

It follows that

$$\begin{aligned} \ell (R^\prime /(x, \overline{Q^{n+1}})) = \overline{e}_0(Q)\cdot \left( {\begin{array}{c}n+d-1\\ d-1\end{array}}\right) + o(n^{d-2}). \end{aligned}$$
(3)

We next show that for all \(n\gg 0\), \(\overline{Q^n}(R^\prime /xR^\prime )= \overline{Q^n(R^\prime /xR^\prime )}\). Once this is proved, we will have \(\overline{Q^n}R^{\prime \prime }= \overline{Q^nR^{\prime \prime }}\) for all \(n\gg 0\) by [21, Lemma 9.1.1] and thus (3) will tell us that

$$\ell (R^{\prime \prime }/\overline{Q^{n+1}R^{\prime \prime }})= \overline{e}_0(Q)\left( {\begin{array}{c}n+d-1\\ d-1\end{array}}\right) +o(n^{d-2}).$$

Since x is a general element of Q, we have \(\overline{e}_0(Q)=e(Q, R^\prime )=e(QR^{\prime \prime }, R^{\prime \prime })=\overline{e}_0(QR^{\prime \prime })\) and so the above equation implies that \(\overline{e}_1(QR'')=0\) as wanted.

To show \(\overline{Q^n}(R^\prime /xR^\prime )= \overline{Q^n(R^\prime /xR^\prime )}\) for \(n\gg 0\), let \(\mathcal {R}^\prime \) denote the integral closure of \(R^\prime [Qt, t^{-1}]\) inside \(R^\prime [t, t^{-1}]\). Concretely, \(\mathcal {R}^\prime \) is the \(\mathbb {Z}\)-graded ring such that \(\mathcal {R}_n= \overline{Q^n}t^n\) for \(n>0\) and \(\mathcal {R}^\prime _n=R^\prime t^{n}\) for \(n\le 0\). Consider the map

$$\begin{aligned} \mathcal {R}^\prime /(xt)\mathcal {R}^\prime \rightarrow R^\prime [t, t^{-1}]/(xt)R^\prime [t, t^{-1}]. \end{aligned}$$

If we localize at any prime ideal \(\mathcal {P}\) of \(R^\prime [Qt, t^{-1}]\) that does not contain \((Qt, t^{-1})\), then we note that \((\mathcal {R}^\prime /(xt)\mathcal {R}^\prime )_{\mathcal {P}}\) is integrally closed inside \((R^\prime [t, t^{-1}]/(xt)R^\prime [t, t^{-1}])_{\mathcal {P}}\). To see this, one can “unlocalize" the ring \(R^\prime \), and consider the integral closure of \(R[t_1,\dots , t_d][Qt, t^{-1}]\) inside \(R[t_1,\dots ,t_d][t, t^{-1}]\), call this ring \(\mathcal {R}\). If one localizes the map

$$\mathcal {R}/(xt)\mathcal {R}\rightarrow R[t_1,\dots ,t_d][t, t^{-1}]/(xt)R[t_1,\dots ,t_d][t, t^{-1}]$$

at any prime ideal that does not contain \((Qt, t^{-1})\) (say it does not contain \(x_1t\)), then the resulting map is a localization of \(R[t_2,\dots ,t_d][\overline{Q^n}t^n, t^{-1}][\frac{1}{x_1t}] \rightarrow R[t_2,\dots , t_d][t, t^{-1}][\frac{1}{x_1t}]\), and the former is already integrally closed in the latter.

Since the radical of \((Qt, t^{-1})\) is the unique homogeneous maximal ideal of \(R[Qt, t^{-1}]\), it follows that \(\mathcal {R}^\prime /(xt)\mathcal {R}^\prime \) and the integral closure of \(R^\prime [Qt, t^{-1}]/(xt)R^\prime [Qt, t^{-1}]\) inside\(R^\prime [t, t^{-1}]/(xt)R^\prime [t, t^{-1}]\) agree in large degree. But note that for \(n>0\),

$$[\mathcal {R}^\prime /(xt)\mathcal {R}^\prime ]_n\cong \frac{\overline{Q^n}}{x\overline{Q^{n-1}}}\cdot t^n\cong \frac{\overline{Q^n}}{x(\overline{Q^n}: x)}\cdot t^n\cong \frac{\overline{Q^n}}{(xR^\prime )\cap \overline{Q^n}}\cdot t^n\cong \overline{Q^n}(R^\prime /xR^\prime )\cdot t^n,$$

where we have used [21, Corollary 6.8.13] again, while the degree n part of the integral closure of \(R'[Qt, t^{-1}]/(xt)R^\prime [Qt, t^{-1}]\) inside \(R^\prime [t, t^{-1}]/(xt)R^\prime [t, t^{-1}]\) is \(\overline{Q^n(R^\prime /xR^\prime )}\cdot t^n\). Thus the fact that they agree in degree \(n\gg 0\) is precisely saying that \(\overline{Q^n}(R^\prime /xR^\prime )= \overline{Q^n(R^\prime /xR^\prime )}\) for \(n\gg 0\).\(\square \)

Now we come back to the proof of the theorem. Let S be the \(S_2\)-ification of \(R^{\prime \prime }\). We have a short exact sequence

$$0\rightarrow R^{\prime \prime } \rightarrow S \rightarrow S/R^{\prime \prime } \rightarrow 0$$

such that \(S/R^{\prime \prime }\) has finite length (since \(R^{\prime \prime }\) is \(S_2\) on the punctured spectrum). Also note that \((S,\mathfrak {n})\) is (complete) local by [17, Proposition (3.9)] and that S is reduced (since S is a subring of the total quotient ring of \(R^{\prime \prime }\)). Since \(R^{\prime \prime }\rightarrow S\) is an integral extension, we have \(\overline{IS}\cap R^{\prime \prime } =\overline{I}\) for every ideal \(I\subseteq R^{\prime \prime }\) by [21, Proposition 1.6.1]. It follows that \(\ell _{R^{\prime \prime }}(S/\overline{Q^nS})\ge \ell _{R^{\prime \prime }}(R^{\prime \prime }/\overline{Q^nR^{\prime \prime }})\) for all \(n\ge 0\). Thus for \(n\gg 0\) we have

$$\begin{aligned} & \overline{e}_0(QR^{\prime \prime })\left( {\begin{array}{c}n+d\\ d\end{array}}\right) -\overline{e}_1(QR^{\prime \prime })\left( {\begin{array}{c}n+d-1\\ d-1\end{array}}\right) + o(n^{d-1})\\= & \ell _{R^{\prime \prime }}(R^{\prime \prime }/\overline{Q^{n+1}R^{\prime \prime }}) \\\le & \ell _{R^{\prime \prime }}(S/\overline{Q^{n+1}S}) \\= & [S/\mathfrak {n}: R/\mathfrak {m}]\cdot \ell _S(S/\overline{Q^{n+1}S})\\= & [S/\mathfrak {n}: R/\mathfrak {m}]\cdot \left( \overline{e}_0(QS)\left( {\begin{array}{c}n+d\\ d\end{array}}\right) -\overline{e}_1(QS)\left( {\begin{array}{c}n+d-1\\ d-1\end{array}}\right) +o(n^{d-1})\right) . \end{aligned}$$

Since S is a rank one module over \(R^{\prime \prime }\), we also know that

$$\overline{e}_0(QR^{\prime \prime })=e(QR^{\prime \prime }, R^{\prime \prime }) = [S/\mathfrak {n}: R/\mathfrak {m}] \cdot e(QS, S)=[S/\mathfrak {n}: R/\mathfrak {m}] \cdot \overline{e}_0(QS),$$

where the second equality is the projection formula for the Hilbert-Samuel multiplicity (which can be seen by combining [21, Theorems 11.2.4 and 11.2.7]). Putting these together we have

$$ [S/\mathfrak {n}: R/\mathfrak {m}] \cdot \overline{e}_1(QS)\le \overline{e}_1(QR^{\prime \prime })=0.$$

But since \(\overline{e}_1(QS)\ge 0\) by [10, Theorem 1.1] (see Theorem 3.1), we must have \(\overline{e}_1(QS)=0\). Now \((S,\mathfrak {n})\) is a reduced complete local ring that is \(S_2\) and \(\dim (S)=d-1\), such that \(\overline{e}_1(QS)=0\). By our inductive hypothesis, we know that S is regular. But since \(S/R^{\prime \prime }\) has finite length, by the long exact sequence of local cohomology induced by \(0\rightarrow R^{\prime \prime } \rightarrow S \rightarrow S/R^{\prime \prime } \rightarrow 0\), we obtain that

$$\begin{aligned}H_{\mathfrak {m}}^i(R^{\prime \prime })=0\, \text {for all}\, i<\dim (R^{\prime \prime })\, \text {and}\, i\ne 1, \text {and} \, H_{\mathfrak {m}}^1(R^{\prime \prime })\cong S/R^{\prime \prime }.\end{aligned}$$

At this point, we consider the long exact sequence of local cohomology induced by \(0\rightarrow \widehat{R^\prime } \xrightarrow {\cdot x} \widehat{R^\prime } \rightarrow R^{\prime \prime }\rightarrow 0\), we get

$$0=H_{\mathfrak {m}}^1(R^\prime )\rightarrow H_{\mathfrak {m}}^1(R^{\prime \prime }) \rightarrow H_{\mathfrak {m}}^2(R^\prime )\xrightarrow {\cdot x} H_{\mathfrak {m}}^2(R^\prime ) \rightarrow H_{\mathfrak {m}}^2(R^{\prime \prime }) \rightarrow \cdots .$$

If \(d\ge 4\), then \(\dim (R^{\prime \prime })\ge 3\) and thus \(H_{\mathfrak {m}}^2(R^{\prime \prime })=0\). Since \(\widehat{R^\prime }\) is \(S_2\), \(H_{\mathfrak {m}}^2(R^\prime )\) has finite length and the above exact sequence tells us that \(H_{\mathfrak {m}}^2(R^\prime )=0\) by Nakayama’s lemma. But then by the above exact sequence again, we have \(H_{\mathfrak {m}}^1(R^{\prime \prime })=0\) and hence \(S/R^{\prime \prime }=0\). Thus \(R^{\prime \prime }\cong S\) is regular. But then \(\widehat{R^\prime }\) and hence R is regular as wanted.

Finally, suppose \(d=3\). Let B be a balanced big Cohen-Macaulay algebra of \(\widehat{R^\prime }\) that is \(\mathfrak {m}\)-adic complete, then B/xB is a balanced big Cohen-Macaulay algebra of \(R^{\prime \prime }\). It follows that the canonical map \(R^{\prime \prime }\rightarrow B/xB\) factors through S.

Claim 3.9

B/xB is a balanced big Cohen-Macaulay algebra over S.

Proof of the Claim

It is clear that some system of parameters of S (namely those coming from \(R^{\prime \prime }\)) are regular sequences on B/xB. To see that every system of parameters of S is a regular sequence on B/xB, we first note that B/xB is \(\mathfrak {m}\)-adically complete: since B is \(\mathfrak {m}\)-adic complete, B/xB is derived \(\mathfrak {m}\)-complete by [29, Tag 091U], take (yz) that is a system of parameters of \(R^{\prime \prime }\), then as yz is a regular sequence on B/xB, the derived completion with respect to (yz), which is B/xB itself, agrees with the usual completion with respect to (yz) by [29, Tag 0920] (equivalently, with respect to \(\mathfrak {m}\) as \(\sqrt{(y,z)}=\mathfrak {m}\)). Hence by [4, Corollary 8.5.3], every system of parameters of S is a regular sequence on \(\widehat{B/xB}\cong B/xB\).\(\square \)

Note that \(\dim (R^{\prime \prime })=\dim (S)=2\) and S is regular, thus the long exact sequence of local cohomology induced by \(0\rightarrow R^{\prime \prime } \rightarrow S \rightarrow S/R^{\prime \prime } \rightarrow 0\) implies that \(H_{\mathfrak {m}}^2(R^{\prime \prime })\cong H_{\mathfrak {m}}^2(S)\). Hence we have the following commutative algebra:

where the injectivity of \(\phi \) follows from the fact that B/xB is a balanced big Cohen-Macaulay algebra over S and thus faithfully flat over S (as S is regular). Chasing this diagram we find that the map \(H_{\mathfrak {m}}^2(R^\prime )\xrightarrow {\cdot x} H_{\mathfrak {m}}^2(R^\prime )\) is surjective. But since \(\widehat{R^\prime }\) is \(S_2\), \(H_{\mathfrak {m}}^2(R^\prime )\) has finite length, thus \(H_{\mathfrak {m}}^2(R^\prime )=0\) by Nakayama’s lemma. Hence \(\widehat{R^\prime }\) is Cohen-Macaulay and thus \(R^{\prime \prime }\) is also Cohen-Macaulay. But then \(R^{\prime \prime }\cong S\) and so \(R^{\prime \prime }\) is regular and thus \(\widehat{R^\prime }\) is regular. Thus R is regular as wanted.

Now we have established that R is regular, we can repeat the argument in the first paragraph of the proof to show that \(\nu (\mathfrak {m}/Q)\le 1\) (essentially, this follows from the main result of [8]). \(\square \)

As a consequence, we answer the problem raised in [10, Section 3] for excellent rings.

Corollary 3.10

Let R be an excellent local ring such that \(\widehat{R}\) is reduced and equidimensional. Suppose that \(I\subseteq R\) is an \(\mathfrak {m}\)-primary ideal such that \(\overline{e}_1(I)=0\). Then \(R^{N }\), the normalization of R, is regular and \(IR^N \) is normal (i.e., all powers of \(IR^N \) are integrally closed in \(R^N \)).

Proof

Replacing R by \(R[t]_{\mathfrak {m} R[t]}\), we may assume that the residue field of R is infinite (we leave it to the readers to check that the hypotheses and conclusions are stable under such a base change). Let S be the \(S_2\)-ification of R. We will show that the \(\mathfrak {m}\)-adic completion of \(\widehat{S}\) is regular. Since R is excellent, \(\widehat{S}\) agrees with the \(S_2\)-ification of \(\widehat{R}\) by [17, Proposition 3.8]. Thus \(\widehat{S}\) is semilocal, reduced, and \(S_2\). Since \(\overline{J\widehat{S}}\cap \widehat{R}=\overline{J}\) for every \(\mathfrak {m}\)-primary ideal \(J\subseteq \widehat{R}\) by [21, Proposition 1.6.1], we have \(\ell _{\widehat{R}}(\widehat{R}/\overline{J})\le \ell _{\widehat{R}}(\widehat{S}/\overline{J\widehat{S}})\).

Let \(\mathfrak {n}_1,\dots , \mathfrak {n}_s\) be the maximal ideals of \(\widehat{S}\) and let \(S_i:= (\widehat{S})_{\mathfrak {n}_i}\) (in fact, since \(\widehat{S}\) is complete, we have \(\widehat{S}\cong \prod _{i=1}^{s}S_i\), and each \(S_i\) is complete local, reduced, and \(S_2\)). Then we have

$$\begin{aligned} & \overline{e}_0(I)\left( {\begin{array}{c}n+d\\ d\end{array}}\right) -\overline{e}_1(I)\left( {\begin{array}{c}n+d-1\\ d-1\end{array}}\right) + o(n^{d-1})\\= & \ell _R(R/\overline{I^{n+1}})=\ell _{\widehat{R}}(\widehat{R}/\overline{I^{n+1}\widehat{R}}) \\\le & \ell _{\widehat{R}}(\widehat{S}/\overline{I^{n+1}\widehat{S}}) \\= & \sum _{i=1}^{s}[S_i/\mathfrak {n}_i: R/\mathfrak {m}]\cdot \ell _{S_i}(S_i/\overline{I^{n+1}S_i})\\= & \sum _{i=1}^{s}[S_i/\mathfrak {n}_i: R/\mathfrak {m}]\cdot \left( \overline{e}_0(IS_i)\left( {\begin{array}{c}n+d\\ d\end{array}}\right) -\overline{e}_1(IS_i)\left( {\begin{array}{c}n+d-1\\ d-1\end{array}}\right) +o(n^{d-1})\right) , \end{aligned}$$

where we have used [21, Lemma 9.1.1] for the equality in the second line. Since \(\widehat{S}\) is a rank one module over \(\widehat{R}\), we also know that

$$\overline{e}_0(I)=e(I\widehat{R}, \widehat{R}) = \sum _{i=1}^{s}[S_i/\mathfrak {n}_i: R/\mathfrak {m}] \cdot e(IS_i, S_i)=\sum _{i=1}^{s}[S_i/\mathfrak {n}_i: R/\mathfrak {m}] \cdot \overline{e}_0(IS_i),$$

where we have used the projection formula for the Hilbert-Samuel multiplicity (see [21, Theorems 11.2.4 and 11.2.7]). The above inequality implies that

$$\sum _{i=1}^{s}[S_i/\mathfrak {n}_i: R/\mathfrak {m}] \cdot \overline{e}_1(IS_i) \le \overline{e}_1(I)=0.$$

But since \(\overline{e}_1(IS_i)\ge 0\) by [10, Theorem 1.1], we must have \(\overline{e}_1(IS_i)=0\) for all i. Let Q be a minimal reduction of I (note that Q is a parameter ideal of R, since we have reduced to the case that R has an infinite residue field). It follows that \(\overline{e}_1(QS_i)=0\) and thus by Theorem 3.7, \(S_i\) is regular and \(\nu (\mathfrak {n}_i/Q)\le 1\). But then \(QS_i\) is normal in \(S_i\). It follows that \(\widehat{S}\cong \prod _{i=1}^{s}S_i\) is regular, \(Q\widehat{S}\) is normal in \(\widehat{S}\) and in particular, \(Q\widehat{S}=I\widehat{S}\).

Since \(S\rightarrow \widehat{S}\cong \widehat{R}\otimes _RS\) is faithfully flat with geometrically regular fibers (as R is excellent). We have S is regular and \(QS=IS\) is normal in S by [21, Theorem 19.2.1]. Finally, since S is regular, S agrees with the normalization \(R^\text {N}\) of R.

Remark 3.11

The condition \(\widehat{R}\) is \(S_2\) cannot be dropped in Theorem 3.7. This was already observed in [10, Section 3]. We give a different example that is a complete local domain. Let \(R=k[[x,xy,y^2, y^3]]\), where k is a field. Then the \(S_2\)-ification of R is \(S=k[[x,y]]\) and we have \(0\rightarrow R\rightarrow S\rightarrow S/R\cong k\cdot \overline{y}\rightarrow 0\). Let \(Q=(x, y^2)\subseteq R\) and we claim that \(\overline{e}_1(Q)=0\). To see this, note that \(QS=(x,y^2)\subseteq S\) is normal and \(\ell (S/Q^{n+1}S)=2\cdot \left( {\begin{array}{c}n+2\\ 2\end{array}}\right) \). It follows from the short exact sequence

$$0\rightarrow R/\overline{Q^{n+1}}\rightarrow S/Q^{n+1}S \rightarrow k\rightarrow 0$$

that \(\ell (R/\overline{Q^{n+1}})=2\cdot \left( {\begin{array}{c}n+2\\ 2\end{array}}\right) -1\). In particular, \(\overline{e}_1(Q)=0\).

Recall that a Noetherian local ring \((R,\mathfrak {m})\) of prime characteristic \(p>0\) is called F-rational if every ideal generated by a system of parameters is tightly closed. It was mentioned in [5] that Huneke asked that when \(\widehat{R}\) is reduced and equidimensional of prime characteristic \(p>0\), whether \(e_1^*(Q)=0\) for some system of parameters \(Q\subseteq R\) implies R is F-rational. In general, counter-examples to the question were constructed in [5, Examples 5.4 and 5.5] (in fact, the example in Remark 3.11 is a counter-example that is a complete local domain). However, all these examples do not satisfy Serre’s \(S_2\) condition.

Let \((R,\mathfrak {m})\) be a Noetherian local ring and let B be a big Cohen-Macaulay R-algebra. Recall that R is called \(\mathrm{{BCM}}_B\)-rational if R is Cohen-Macaulay and the natural map \(H_{\mathfrak {m}}^d(R)\rightarrow H_{\mathfrak {m}}^d(B)\) is injective, where \(d=\dim (R)\). If R is an excellent local ring of prime characteristic \(p>0\), then R is F-rational if and only if R is \(\mathrm{{BCM}}_B\)-rational for all big Cohen-Macaulay algebra B, see [24, Proposition 3.5].

We propose the following conjecture relating the vanishing of \(e_1^B(Q)\) and \(\mathrm{{BCM}}_B\)-rational singularities, which modifies Huneke’s question and makes sense in all characteristics.

Conjecture 3.12

Let \((R,\mathfrak {m})\) be a Noetherian local ring such that \(\widehat{R}\) is reduced and \(S_2\). Let B be a balanced big Cohen-Macaulay R-algebra that satisfies (\(\dagger \)). If \(e_1^B(Q) = 0\) for some parameter ideal \(Q\subseteq R\), then R is \(\mathrm{{BCM}}_B\)-rational.

In particular, if R is excellent and has characteristic \(p>0\) (such that \(\widehat{R}\) is reduced and \(S_2\)), and \(e_1^*(Q) = 0\) for some parameter ideal \(Q\subseteq R\), then R is F-rational.

We have the following partial result towards Conjecture 3.12, which is an analog of the main result of [27].

Proposition 3.13

Let \((R, \mathfrak {m})\) be a Noetherian local ring such that \(\widehat{R}\) is reduced and equidimensional. Let B be a balanced big Cohen-Macaulay R-algebra that satisfies (\(\dagger \)). If \(e_1^B(Q) = e_1(Q)\) for some parameter ideal \(Q\subseteq R\), then R is \(\mathrm{{BCM}}_B\)-rational.

In particular, if R is excellent and has characteristic \(p>0\), and \(e_1^*(Q) = e_1(Q)\) for some parameter ideal \(Q\subseteq R\), then R is F-rational.

Proof

By Theorem 3.1, we know that \(e_1^B(Q)=e_1(Q)=0\). By the main result of [6], \(e_1(Q)=0\) implies that R is Cohen-Macaulay. By [19, Corollary 4.9], we have \(Q^B = Q\). Now we consider the commutative diagram:

where the injectivity of the top row follows from \(Q^B=Q\), the injectivity of the left column is because R is Cohen-Macaulay, and the injectivity of the right column is because B is balanced big Cohen-Macaulay. Since R is Cohen-Macaulay, we know that \(\textrm{Soc}(R/Q) \cong \textrm{Soc}(H^d_{\mathfrak {m}}(R))\). Chasing the commutative diagram we find that the map \(H^d_{\mathfrak {m}}(R) \rightarrow H^d_{\mathfrak {m}}(B)\) is injective. Therefore, R is \(\textrm{BCM}_B\)-rational.\(\square \)

Remark 3.14

It is clear from the proof of Proposition 3.13 that Conjecture 3.12 holds when R is Cohen-Macaulay, and this essentially follows from [19, Corollary 4.9].