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

Frobenius-Poincaré function and Hilbert-Kunz multiplicity

Alapan Mukhopadhyay
Abstract.

We generalize the notion of Hilbert-Kunz multiplicity of a graded triple (M,R,I)(M,R,I) in characteristic p>0p>0 by proving that for any complex number yy, the limit

limn(1pn)dim(M)j=λ((MI[pn]M)j)eiyj/pn\underset{n\to\infty}{\lim}(\frac{1}{p^{n}})^{\text{dim}(M)}\sum\limits_{j=-\infty}^{\infty}\lambda\left((\frac{M}{I^{[p^{n}]}M})_{j}\right)e^{-iyj/p^{n}}

exists. We prove that the limiting function in the complex variable yy is entire and name this function the Frobenius-Poincaré function. We establish various properties of Frobenius-Poincaré functions including its relation with the tight closure of the defining ideal II; and relate the study Frobenius-Poincaré functions to the behaviour of graded Betti numbers of RI[pn]\frac{R}{I^{[p^{n}]}} as nn varies. Our description of Frobenius-Poincaré functions in dimension one and two and other examples raises questions on the structure of Frobenius-Poincaré functions in general.

00footnotetext: Mathematics Subject Classification 2020: Primary 13D40; Secondary: 13A02, 13A35, 30D15 .

1. Introduction

In this article, we introduce the Frobenius-Poincaré function of a graded pair (R,I)(R,I), where II is a finite co-length homogeneous ideal in the standard graded domain RR over a perfect field of positive characteristic pp. This function is holomorphic everywhere on the complex plane and is roughly the limit of the Hilbert series of the graded RR-modules R1/pnIR1/pn\frac{R^{1/p^{n}}}{IR^{1/p^{n}}} as nn goes to infinity. The Frobenius-Poincaré function encodes the information of the Hilbert-Kunz multiplicity of the pair (R,I)(R,I) along with other asymptotic invariants of (R,I)(R,I).

To be precise, fix a pair (R,I)(R,I) as above. For each positive integer nn, consider the RR-module R1/pnR^{1/p^{n}}, the collection of pnp^{n}-th roots of elements of RR in a fixed algebraic closure of the fraction field of RR. There is a natural 1pn\frac{1}{p^{n}}\mathbb{Z}-grading on R1/pnR^{1/p^{n}}. So one can consider the Hilbert series of R1/pnIR1/pn\frac{R^{1/p^{n}}}{IR^{1/p^{n}}} by allowing for rational powers of the variable tt— namely ν1pnλ((R1/pnIR1/pn)ν)tν\underset{\nu\in\frac{1}{p^{n}}\mathbb{Z}}{\sum}\lambda((\frac{R^{1/p^{n}}}{IR^{1/p^{n}}})_{\nu})t^{\nu}. To study these Hilbert series as holomorphic functions on the complex plane, a natural approach is to replace tt by eiye^{-iy} 111Here ee is the complex number j=01j!\sum\limits_{j=0}^{\infty}\frac{1}{j!} and ii is a complex square root of 1-1, fixed throughout this article., which facilitates taking pnp^{n}-th roots as holomorphic functions. The process described above gives a sequence (Gn)n(G_{n})_{n} of holomorphic functions where

Gn(y)=ν1pnλ((R1/pnIR1/pn)ν)eiyν.G_{n}(y)=\underset{\nu\in\frac{1}{p^{n}}\mathbb{Z}}{\sum}\lambda((\frac{R^{1/p^{n}}}{IR^{1/p^{n}}})_{\nu})e^{-iy\nu}.

Our main result, Theorem 3.1, guarantees that the sequence of functions

Gn(y)(pdim(R))n\frac{G_{n}(y)}{(p^{\text{dim}(R)})^{n}}

converges, as nn goes to infinity, to a function F(y)F(y) which is holomorphic everywhere on the complex plane. Furthermore, this convergence is uniform on every compact subset. We call this function F(y)F(y) the Frobenius-Poincaré function associated to the pair (R,I)(R,I).

The Frobenius-Poincaré function can be viewed as a natural refinement of the Hilbert-Kunz multiplicity (see Definition 2.1). Indeed, for the pair (R,I),(R,I), the Hilbert-Kunz multiplicity is the value of the Frobenius-Poincaré function at zero. In fact, we provide an explicit formula for the coefficients of the power series expansion of the Frobenius-Poincaré entire function around zero, from which it is apparent that each of the coefficients of this power series is an invariant generalizing the Hilbert-Kunz multiplicity (see Proposition 4.1). Even when RR is a polynomial ring, there are examples of ideals with the same Hilbert-Kunz multiplicity but different Frobenius-Poincaré functions- see Example 5.10. Like the Hilbert-Kunz multiplicity itself, the Frobenius-Poincaré function of (R,I)(R,I) depends only on the tight closure of II in the ring RR, as we prove in Theorem 4.7.

The information carried by the Frobenius-Poincaré function can also be understood in terms of homological data associated to the pair (R,I)(R,I). For example, when RR is a polynomial ring (or more generally, when R/IR/I has finite projective dimension), we prove that the Frobenius-Poincaré function has the form

(1) Q(eiy)(iy)dimR\frac{Q(e^{-iy})}{(iy)^{\text{dim}R}}

where QQ is a polynomial whose coefficients are explicitly determined by the graded Betti numbers of R/IR/I; see Proposition 5.9. More generally, for an arbitrary graded pair (R,I)(R,I), the Frobenius-Poincaré function of (R,I)(R,I) can be described in terms of the sequence of graded Betti numbers of RI[pn]\frac{R}{{I}^{[{p}^{n}]}}:

Theorem A: Let SRS\hookrightarrow R be a graded Noether normalization. Set 𝔹S(j,n)=α=0(1)αλ(TorαS(RI[pn],k)j)\mathbb{B}^{S}(j,n)=\sum\limits_{\alpha=0}^{\infty}(-1)^{\alpha}\lambda(\text{Tor}_{\alpha}^{S}(\frac{R}{{I}^{[{p}^{n}]}},k)_{j}). Then the limiting function

BS(R,I)(y)=limnj𝔹S(j,n)eiyj/pnB^{S}(R,I)(y)=\underset{n\to\infty}{\lim}\underset{j\in\mathbb{N}}{\sum}\ \mathbb{B}^{S}(j,n)e^{-iyj/p^{n}}

is entire. Furthermore, BS(R,I)(y)(iy)dimR\frac{B^{S}(R,I)(y)}{(iy)^{\text{dim}R}} is the Frobenius-Poincaré function of (R,I)(R,I) – see Theorem 5.1 and Remark 5.4.

In a slightly different direction, we show that when RR is Cohen-Macaulay, the Frobenius-Poincaré entire function is a limit of a sequence of entire functions described in terms of Koszul homologies with respect to a homogeneous system of parameters for RR, or alternatively, Serre’s intersection numbers, suitably interpreted; see Theorem 5.11.

The Frobenius-Poincaré function of (R,I)(R,I) turns out to be the Fourier transform of the Hilbert-Kunz density function of (R,I)(R,I) introduced by Trivedi in [Tri18], as we show in Proposition 4.9. Using Fourier transform, Theorem A allows us to describe the higher order weak derivatives of the Hilbert-Kunz density function in terms of the sequence graded Betti numbers of RI[pn]\frac{R}{{I}^{[{p}^{n}]}} - see Remark 5.5. Such a description is not apparent in the existing theory of Hilbert-Kunz density functions. In fact, Theorem A relates the question on the order of smoothness of the Hilbert-Kunz density function raised in [Tri21] to the question asking whether BS(R,I)(y)B^{S}(R,I)(y) in Theorem A is bounded on the real line – see 5.6. Although our work on Frobenius-Poincaré functions is inspired by Trivedi’s remark that considering Fourier transforms of density functions might be useful (see [Tri18, page 3]), our proof of the existence and holomorphicity of the Frobenius-Poincaré function (Theorem 3.1) is independent of [Tri18]. When RR has dimension at least two and RR is strongly FF-regular at each point on the punctured spectrum of RR, the Hilbert-Kunz density function and hence the Frobenius-Poincaré function of (R,I)(R,I) captures the information of FF-threshold of II- see Theorem 4.9 of [TW21]. Recently Trivedi has used Hilbert-Kunz density functions to partially settle two conjectures on Hilbert-Kunz multiplicities of quadric hypersurfaces posed by Yoshida and Watanabe-Yoshida- see Theorem A and Theorem B in [Tri21].

We speculate that the entire functions that are Frobenius-Poincaré functions should have a special structure reflecting that each of these is determined by the data of a finitely generated module. Any such special structure will shed more light not only on the theory of Hilbert-Kunz multiplicities but also on the behaviour of graded Betti numbers of RI[pn]\frac{R}{{I}^{[{p}^{n}]}} as nn changes. We ask whether Frobenius-Poincaré functions always have a form generalizing the expression (1) above; see 5.13. 5.13 is answered for one dimensional rings in Proposition 4.6. When RR is two dimensional, 5.13 is answered in Theorem 6.1, where we show that the Frobenius-Poincaré function is described by the Harder-Narasimhan filtration on a sufficiently high Frobenius pullback of the syzygy bundle of II on the curve Proj(R)\text{Proj}(R) following [Tri05] and [Bre07]. The necessary background materials on vector bundles on curves and other topics are reviewed in Section 2. Also when the ideal II is generated by a homogeneous system of parameters, our computation in Proposition 4.5 answers 5.13 positively.

We develop the theory of Frobenius-Poincaré functions more generally for triples (M,R,I)(M,R,I), where MM is a finitely generated \mathbb{Z}-graded RR module; see Definition 3.2. We show that the Frobenius-Poincaré function is additive on short exact sequences in Proposition 4.4. In addition to generalizing classical additivity formulas for Hilbert series and multiplicity, Proposition 4.4 allows us to compute the Frobenius-Poincaré function of a graded ring with respect to an ideal generated by homogeneous system of parameters; see Proposition 4.5.

Notation and Convention 1.1.

In this article, kk stands for a field. By a finitely generated \mathbb{N}-graded kk-algebra, we mean an \mathbb{N}-graded commutative ring whose degree zero piece is kk and which is finitely generated over kk.

For any ring SS containing 𝔽p\mathbb{F}_{p}, the Frobenius or pp-th power endomorphism of SS is denoted by FSF_{S}. The symbol FSeF_{S}^{e} will denote the ee-times iteration of FSF_{S}. We set Spe=Fe(S)SS^{p^{e}}=F^{e}(S)\subseteq S. For an ideal JSJ\subseteq S, JSpeJS^{p^{e}} is the image of JJ in SpeS^{p^{e}} under the pep^{e}-th power map. The ideal generated by pnp^{n}-th power of elements of JJ in SS is denoted by J[pn]{J}^{[{p}^{n}]}.

For an SS-module NN, we denote the Krull dimension of NN by dimS(N)\text{dim}_{S}(N) or dim(N)\text{dim}(N) when the underlying ring SS is clear from the context. When NN has finite length, λS(N)\lambda_{S}(N) denotes the length of the SS-module NN. When S=kS=k, simply λ(N)\lambda(N) will be used to denote the length.

Recall that an entire function is a function holomorphic everywhere on the complex plane (see [Ahl79], section 2.3).

Acknowledgements: I am grateful to Karen Smith for her generosity with sharing ideas and suggestions which has improved the article substantially. I thank Mel Hochster and V. Trivedi for pointing out useful references; Jakub Witaszek, Sridhar Venkatesh for useful discussions; Ilya Smirnov and the anonymous referee for their comments. I also thank Daniel Smolkin, Janet Page, Jenny Kenkel, Swaraj Pande, Anna Brosowsky, Eamon Gallego for their questions, comments during a talk on an early version of this article. I was partially supported by NSF DMS grants # 2101075, # 1801697, NSF FRG grant # 1952399 and Rackham one term dissertation fellowship while working on this article.

2. Background material

In this section, we recall some results, adapted to our setting, for future reference.

2.1. Hilbert-Kunz multiplicity

Hilbert-Kunz multiplicity is a multiplicity theory in positive characteristic. We refer readers to [Hun13] for a survey of this theory. In this subsection, kk is a field of characteristic p>0p>0.

Definition 2.1.

Let RR be a finitely generated \mathbb{N}-graded kk-algebra; JJ be a homogeneous ideal such that R/JR/J has finite length. Given a finitely generated \mathbb{Z}-graded RR-module NN, the Hilbert-Kunz multiplicity of the triple (N,R,J)(N,R,J) is defined to be the following limit

limn(1pn)dim(N)λR(NJ[pn]N).\underset{n\to\infty}{\lim}(\frac{1}{p^{n}})^{\text{dim}(N)}\lambda_{R}(\frac{N}{{J}^{[{p}^{n}]}N}).

Similarly one can define the Hilbert-Kunz multiplicity of a triple (S,I,M)(S,I,M)- where II is any finite co-length ideal in a Noetherian local ring SS and MM is a finitely generated SS-module.

The existence of the limit in the Definition 2.1 was first established by Monsky (see [Mon83]).

The Hilbert-Kunz multiplicity of any local ring is at least one. Moreover, under mild hypothesis, it is exactly one if and only if the ring is regular; see Theorem 1.5 of [Kei00] and [Cra02]. These two facts suggest that Hilbert-Kunz multiplicity is a candidate for a multiplicity theory. In general, rings with Hilbert-Kunz multiplicity closer to one are interpreted to have better singularities; see [Man04] and [GN01].

Unlike the usual Hilbert-Samuel function, the structure of the Hilbert-Kunz function f(n)=λ(NJ[pn]N)f(n)=\lambda(\frac{N}{{J}^{[{p}^{n}]}N}) is rather elusive. We refer interested readers to [HMM04], [Tei02], [FT03].

2.2. Betti numbers

We review results on graded Betti numbers which we use in Section 5. References for most of these results are [Ser00], and [BH98]. Recall that RR is a finitely generated \mathbb{N}-graded kk-algebra (see 1.1).

Given a finitely generated \mathbb{Z}-graded RR-module MM, one can choose a minimal graded free resolution of MM: this is a free resolution (G,d)(G_{\bullet},d_{\bullet}) of MM such that each GnG_{n} is a graded free RR-module, the boundary maps preserve graded structures, and the entries of the matrices representing boundary maps are forms of positive degrees. As a consequence, GrsR(s)bMR(r,s)G_{r}\cong\underset{s\in\mathbb{Z}}{\oplus}R(-s)^{b_{M}^{R}(r,s)} where bMR(r,s)=λ(TorrR(k,M)s)b_{M}^{R}(r,s)=\lambda(\text{Tor}_{r}^{R}(k,M)_{s}).

Definition 2.2.

Let MM be a finitely generated \mathbb{Z}-graded RR-module. The rr-th Betti number of MM with respect to RR is the rank of the free module GrG_{r} at the rr-th spot in a minimal graded free resolution of MM, or equivalently, the length λ(TorrR(k,M))\lambda(\text{Tor}_{r}^{R}(k,M)).

Definition 2.3.

The \mathbb{N}-graded ring RR is a graded complete intersection over kk if Rk[X1,,Xs](f1,,fh)R\cong\frac{k[X_{1},\ldots,X_{s}]}{(f_{1},\ldots,f_{h})}- where each XjX_{j} is homogeneous of positive degree and f1,,fhf_{1},\ldots,f_{h} is a regular sequence consisting of homogeneous polynomials.

We recall a special case of a result in [GUL74].

Lemma 2.4.

Let RR be a graded complete intersection over kk. Then for any finitely generated \mathbb{Z}-graded RR-module MM, there is polynomial PM(t)[t]P_{M}(t)\in\mathbb{Z}[t] such that for all nn, λ(TornR(M,k))PM(n)\lambda(\text{Tor}_{n}^{R}(M,k))\leq P_{M}(n).

Proof.

Let mm be the homogeneous maximal ideal of RR. Then RmR_{m} is a local complete intersection as is meant in Corollary 4.2, [GUL74]. Since for all nn\in\mathbb{N}, TornR(M,k)TornRm(Mm,k)\text{Tor}_{n}^{R}(M,k)\cong\text{Tor}^{R_{m}}_{n}(M_{m},k), using Corollary 4.2, [GUL74], we have a polynomial π(t)[t]\pi(t)\in\mathbb{Z}[t] and rr\in\mathbb{N} such that, n=0λ(TornR(M,k))tn=π(t)(1t2)r\sum\limits_{n=0}^{\infty}\lambda(\text{Tor}_{n}^{R}(M,k))t^{n}=\frac{\pi(t)}{(1-t^{2})^{r}}. The assertion in Lemma 2.4 now follows by using the formal power series in tt representing 1(1t2)r\frac{1}{(1-t^{2})^{r}}. ∎

Lemma 2.5.

Let RR be a finitely generated \mathbb{N}-graded kk-algebra. Let MM be a finitely generated \mathbb{Z}-graded RR-module. There is a positive integer ll such that given any integer ss, bMR(r,s):=λ(TorrR(M,k)s)=0b^{R}_{M}(r,s):=\lambda(\text{Tor}^{R}_{r}(M,k)_{s})=0 for all rs+lr\geq s+l.

Proof.

Pick a minimal free resolution (G,d)(G_{\bullet},d_{\bullet}) of MM, then GrjR(j)bMR(r,j)G_{r}\cong\underset{j\in\mathbb{Z}}{\oplus}R(-j)^{b^{R}_{M}(r,j)}. Since the boundary maps of GG_{\bullet} are represented by matrices whose entries are positive degree forms and have non-zero columns, ϕ(r):=min{j|bMR(r,j)0}\phi(r):=\text{min}\{j\,|\,b^{R}_{M}(r,j)\neq 0\} is a strictly increasing function of rr. So we can choose an integer ll such that ϕ(l)>0\phi(l)>0. Again since ϕ(r)\phi(r) is strictly increasing, for all rj+lr\geq j+l, ϕ(r)ϕ(l)+j>j\phi(r)\geq\phi(l)+j>j. So given an integer ss, bMR(r,s)=0b^{R}_{M}(r,s)=0 for all rs+lr\geq s+l. ∎

Lemma 2.6.

Let RR be a graded complete intersection over kk and MM be a finitely generated \mathbb{Z}-graded RR-module. For a given integer ss, let 𝔹M(s)\mathbb{B}_{M}(s) denote the sum r(1)rbMR(r,s)\underset{r\in\mathbb{N}}{\sum}(-1)^{r}{b_{M}^{R}(r,s)}. Then the formal Laurent series s𝔹M(s)ts\underset{s\in\mathbb{Z}}{\sum}\mathbb{B}_{M}(s)t^{s} is absolutely convergent at every non-zero point on the open unit disk centered at the origin in \mathbb{C}.

Proof.

By Lemma 2.5, there is an ll\in\mathbb{N} such that for any integer ss, |𝔹M(s)|j=0s+lλ(TorjR(M,k))|\mathbb{B}_{M}(s)|\leq\sum\limits_{j=0}^{s+l}\lambda(\text{Tor}_{j}^{R}(M,k)). With PMP_{M} the same as in Lemma 2.4, consider the polynomial

QM(s)=j=0s+lPM(j).Q_{M}(s)=\sum_{j=0}^{s+l}P_{M}(j).

Thus for ss\in\mathbb{N}, |𝔹M(s)|QM(s)|\mathbb{B}_{M}(s)|\leq Q_{M}(s). So the radius of convergence of the power series s=0|𝔹M(s)|ts\sum\limits_{s=0}^{\infty}|\mathbb{B}_{M}(s)|t^{s} is at least one and the desired conclusion follows. ∎

2.3. Hilbert series and Hilbert-Samuel multiplicities

The references for this subsection are [BH98] and [Ser00]. Throughout, RR is a finitely generated \mathbb{N}-graded algebra over a field kk. Recall that the Hilbert series(also called the Hilbert-Poincaré series) of a finitely generated \mathbb{Z}-graded RR-module MM is the formal Laurent series HM(t):=nλ(Mn)tnH_{M}(t):=\underset{n\in\mathbb{Z}}{\sum}\lambda(M_{n})t^{n}.

Theorem 2.7.

(see Proposition 4.4.1, [BH98]) Let MM be a finitely generated \mathbb{Z}-graded RR-module.

  1. (1)

    There is a Laurent polynomial QM(t)[t,t1]Q_{M}(t)\in\mathbb{Q}[t,t^{-1}] such that,

    HM(t)=QM(t)(1tδ1)(1tδdim(M))H_{M}(t)=\frac{Q_{M}(t)}{(1-t^{\delta_{1}})\ldots(1-t^{\delta_{\text{dim(M)}}})}

    for some non-negative integers δ1,,δdim(M)\delta_{1},\ldots,\delta_{\text{dim}(M)}.

  2. (2)

    The choice of QMQ_{M} depends on the choices of δ1,,δdim(M)\delta_{1},\ldots,\delta_{\text{dim}(M)}. One can choose δ1,,δdim(M)\delta_{1},\ldots,\delta_{\text{dim}(M)} to be the degrees of elements of RAnn(M)\frac{R}{\text{Ann(M)}} forming a homogeneous system of parameters. 222A homogeneous system of parameters of a finitely generated \mathbb{N}-graded kk-algebra SS, is a collection of homogeneous elements f1,,fdim(S)f_{1},\ldots,f_{\text{dim}(S)} such that S(f1,,fdim(S))\frac{S}{(f_{1},\ldots,f_{\text{dim}(S)})} has a finite length (see [BH98], page 35).

In Proposition 2.8, we extend part of Proposition 4.1.9 of [BH98]- where RR is assumed to be standard graded- to our setting. We use Proposition 2.8 to define Hilbert-Samuel multiplicity of a finitely generated \mathbb{Z}-graded module over a graded ring- where the ring is not necessarily standard graded-in part (1), Definition 2.9.

Proposition 2.8.

Let MM be a finitely generated \mathbb{Z}-graded RR-module of Krull dimension d. Denote the Poincaré series of MM by HM(t)H_{M}(t).

  1. (1)

    The limit d!limn1nd(jnλ(Mj))d!\underset{n\to\infty}{\lim}\frac{1}{n^{d}}(\underset{j\leq n}{\sum}\lambda(M_{j})) exists . The limit is denoted by eMe_{M}.

  2. (2)

    The limit limt1(1t)dHM(t)\underset{t\to 1}{\lim}(1-t)^{d}H_{M}(t) is the same as eMe_{M}.

Proof of Proposition 2.8.

When MM has Krull dimension zero, the desired conclusion is immediate. So we assume that MM has a positive Krull dimension. We first prove (1).

Let f1,,fdf_{1},\ldots,f_{d} be a homogeneous system of parameters of RAnn(M)\frac{R}{\text{Ann}(M)} of degree δ1,,δd\delta_{1},\ldots,\delta_{d} respectively. Set δ\delta to be the product δ1δd\delta_{1}\ldots\delta_{d} and gj=fjδδjg_{j}=f_{j}^{\frac{\delta}{\delta_{j}}}. Then each of g1,,gdg_{1},\ldots,g_{d} has degree δ\delta and these form a homogeneous system of parameters of RAnn(M)\frac{R}{\text{Ann}(M)}. Denote the kk-subalgebra generated by g1,,gdg_{1},\ldots,g_{d} by SS. We endow SS with a new \mathbb{N}-grading: given a natural number nn, declare the nn-th graded piece of SS to be

Sn:=S(RAnn(M))δn.S_{n}:=S\cap(\frac{R}{\text{Ann}(M)})_{\delta n}.

From now on, by the grading on SS we refer to the grading defined above. Note that SS is a standard graded kk-algebra. Now for each rr, where 0r<δ0\leq r<\delta, set

Mr=nMnδ+r.M^{r}=\underset{n\in\mathbb{Z}}{\oplus}M_{n\delta+r}.

Given an rr as above, we give MrM^{r} a \mathbb{Z}-graded structure by declaring the nn-th graded piece of MrM^{r} to be Mnδ+rM_{n\delta+r}. Then each MrM^{r} is a finitely generated \mathbb{Z}-graded module over SS. Since SS is standard graded, for each rr, 0r<δ0\leq r<\delta, the limit

limn1nd1λ(Mnr)\underset{n\to\infty}{\lim}\frac{1}{n^{d-1}}\lambda(M^{r}_{n})

exists (see Theorem 4.1.3, [BH98]). This implies the existence of a constant CC such that λ(Mn)Cnd1\lambda(M_{n})\leq Cn^{d-1} for all nn. So the sequence d!nd(jnλ(Mj))\frac{d!}{n^{d}}(\underset{j\leq n}{\sum}\lambda(M_{j})) converges if and only if the subsequence

(d!(δn+δ1)d(j(n+1)δ1λ(Mj)))n\left(\frac{d!}{(\delta n+\delta-1)^{d}}(\underset{j\leq(n+1)\delta-1}{\sum}\lambda(M_{j}))\right)_{n}

converges. Now we show that the above subsequence is convergent by computing its limit.

(2) d!limn1(δn+δ1)d(j(n+1)δ1λ(Mj))=limnnd(δn+δ1)d[r=0δ1d!nd(jnλ(Mjr))]\begin{split}&d!\underset{n\to\infty}{\lim}\frac{1}{(\delta n+\delta-1)^{d}}(\underset{j\leq(n+1)\delta-1}{\sum}\lambda(M_{j}))\\ &=\underset{n\to\infty}{\lim}\frac{n^{d}}{(\delta n+\delta-1)^{d}}[\sum\limits_{r=0}^{\delta-1}\frac{d!}{n^{d}}(\underset{j\leq n}{\sum}\lambda(M^{r}_{j}))]\end{split}

Since each MrM^{r} where 0rδ10\leq r\leq\delta-1 is a finitely generated module over the standard graded ring SS, by Proposition 4.1.9 and Remark 4.1.6 of [BH98], the last limit in (2) exists and

(3) eM=eM0++eMδ1δd.e_{M}=\frac{e_{M^{0}}+\ldots+e_{M^{\delta-1}}}{\delta^{d}}.

For (2), note that

(4) limt1(1t)dHM(t)=limt1r=0δ1(1t)dHMr(tδ)tr=limt1r=0δ1(1tδ)dHMr(tδ)tr(1+t++tδ1)d.\begin{split}\underset{t\to 1}{\lim}(1-t)^{d}H_{M}(t)&=\underset{t\to 1}{\lim}\sum\limits_{r=0}^{\delta-1}(1-t)^{d}H_{M^{r}}(t^{\delta})t^{r}\\ &=\underset{t\to 1}{\lim}\frac{\sum\limits_{r=0}^{\delta-1}(1-t^{\delta})^{d}H_{M^{r}}(t^{\delta})t^{r}}{(1+t+\ldots+t^{\delta-1})^{d}}.\end{split}

Again, since each MrM^{r} is a finitely generated module over the standard graded ring SS, by Proposition 4.1.9 and Remark 4.1.6 of [BH98]

eMr=limt1(1t)dHMr(t).e_{M^{r}}=\underset{t\to 1}{\lim}(1-t)^{d}H_{M^{r}}(t).

So from (4) and (3), we get

limt1(1t)dHM(t)=eM0++eMδ1δd=eM.\underset{t\to 1}{\lim}(1-t)^{d}H_{M}(t)=\frac{e_{M^{0}}+\ldots+e_{M^{\delta-1}}}{\delta^{d}}=e_{M}.

Definition 2.9.

Let MM be a finitely generated \mathbb{Z}-graded RR-module of Krull dimension dd.

  1. (1)

    The Hilbert-Samuel multiplicity of MM is defined to be the limit

    d!limn1nd(jnλ(Mj))d!\underset{n\to\infty}{\lim}\frac{1}{n^{d}}(\underset{j\leq n}{\sum}\lambda(M_{j}))

    and denoted by eMe_{M}. The limit exists by (1) Proposition 2.8.

  2. (2)

    Given a homogeneous ideal II of finite co-length, the Hilbert-Samuel multiplicity of MM with respect to II is defined to be the limit:

    d!limn1ndλ(MInM).d!\underset{n\to\infty}{\lim}\frac{1}{n^{d}}\lambda(\frac{M}{I^{n}M}).
Proposition 2.10.

Let f1,,fdf_{1},\ldots,f_{d} be a homogeneous system of parameters of RR of degree δ1,,δd\delta_{1},\ldots,\delta_{d} respectively. Then the Hilbert-Samuel multiplicity of RR with respect to (f1,,fd)(f_{1},\ldots,f_{d}) (see Definition 2.9) is δ1δdeR\delta_{1}\ldots\delta_{d}e_{R}.

Proof.

By Proposition 2.10 of [HTW11], the desired multiplicity is δ1δdlimt1(1t)dHR(t)\delta_{1}\ldots\delta_{d}\underset{t\to 1}{\lim}(1-t)^{d}H_{R}(t), which by Proposition 2.8 is δ1δdeR\delta_{1}\ldots\delta_{d}e_{R}. ∎

2.4. Vector bundles on curves

In this subsection, CC stands for a curve, where by a curve we mean a one dimensional, irreducible smooth projective variety over an algebraically closed field; the genus of CC is denoted by gg. A vector bundle on CC means a locally free sheaf 𝒪C\mathcal{O}_{C}-modules of finite constant rank. Morphisms of vector bundles are a priori morphisms of 𝒪C\mathcal{O}_{C}-modules. We recall some results on vector bundles on CC which we use in Section 6. For any unexplained terminology, readers are requested to turn to [Har97] or [Pot97].

Definition 2.11.

Let \mathcal{F} be a coherent sheaf on the curve C.C.

  1. (1)

    The rank of ,\mathcal{F}, denoted by rk()\text{rk}(\mathcal{F}), is the dimension of the stalk of \mathcal{F} at the generic point of CC as a vector space over the function field of CC.

  2. (2)

    The degree of ,\mathcal{F}, denoted by deg()\text{deg}(\mathcal{F}), is defined as h0(C,)h1(C,)rk()(1g)h^{0}(C,\mathcal{F})-h^{1}(C,\mathcal{F})-\text{rk}(\mathcal{F})(1-g).

  3. (3)

    The slope of \mathcal{F}, denoted by μ()\mu(\mathcal{F}), is the ratio deg()rk()\frac{\text{deg}(\mathcal{F})}{\text{rk}(\mathcal{F})}. By convention, μ()=\mu(\mathcal{F})=\infty if rk()=0\text{rk}(\mathcal{F})=0.

Definition 2.12.

A vector bundle \mathcal{E} on CC is called semistable if for any nonzero coherent subsheaf \mathcal{F} of \mathcal{E}, μ()μ()\mu(\mathcal{F})\leq\mu(\mathcal{E}).

Theorem 2.13.

(see [HN75, Prop 1.3.9]) Let \mathcal{E} be a vector bundle on CC. Then there exists a unique filtration:

0=01tt+1=0=\mathcal{E}_{0}\subset\mathcal{E}_{1}\subset\ldots\subset\mathcal{E}_{t}\subset\mathcal{E}_{t+1}=\mathcal{E}

such that,

  1. (1)

    All the quotients j+1/j\mathcal{E}_{j+1}/\mathcal{E}_{j} are non-zero, semistable vector bundles.

  2. (2)

    For all jj, μ(j/j1)>μ(j+1/j)\mu(\mathcal{E}_{j}/\mathcal{E}_{j-1})>\mu(\mathcal{E}_{j+1}/\mathcal{E}_{j}).

This filtration is called the Harder-Narasimhan filtration of \mathcal{E}.

Proposition 2.14.

Let 0=01tt+1=0=\mathcal{E}_{0}\subset\mathcal{E}_{1}\subset\ldots\subset\mathcal{E}_{t}\subset\mathcal{E}_{t+1}=\mathcal{E} be the HN filtration on \mathcal{E}. If the slope of 1\mathcal{E}_{1} is negative, \mathcal{E} cannot have a non-zero global section.

Proof.

On the contrary, assume that \mathcal{E} has a non-zero global section ss. Let λs:𝒪C\lambda_{s}:\mathcal{O}_{C}\rightarrow\mathcal{E} be the non-zero map induced by ss. Let bb be the largest integer such that the composition 𝒪Cλs/b\mathcal{O}_{C}\xrightarrow{\lambda_{s}}\mathcal{E}\rightarrow\mathcal{E}/\mathcal{E}_{b} is non-zero. Then λs\lambda_{s} induces a non-zero map from 𝒪C\mathcal{O}_{C} to b+1/b\mathcal{E}_{b+1}/\mathcal{E}_{b}, whose image \mathcal{L} is a line bundle with a non-zero global section. So the slope of \mathcal{L} is positive. On the other hand, since \mathcal{L} is a non-zero subsheaf of the semistable sheaf b+1/b\mathcal{E}_{b+1}/\mathcal{E}_{b}, μ()<μ(b+1/b)\mu({\mathcal{L}})<\mu({\mathcal{E}_{b+1}/\mathcal{E}_{b}}). Since μ(b+1/b)<μ(1)\mu(\mathcal{E}_{b+1}/\mathcal{E}_{b})<\mu(\mathcal{E}_{1}), the slope of b+1/b\mathcal{E}_{b+1}/\mathcal{E}_{b} is negative ; so \mathcal{L} cannot have a positive slope. ∎

Lemma 2.15.
  1. (1)

    For a coherent sheaf of 𝒪C\mathcal{O}_{C} modules \mathcal{F} and a line bundle \mathcal{L}, μ()=μ()+deg()\mu(\mathcal{F}\otimes\mathcal{L})=\mu(\mathcal{F})+\text{deg}(\mathcal{L}). Here we stick to the convention that the sum of \infty and a real number is \infty.

  2. (2)

    Tensor product of a semistable vector bundle and a line bundle is semistable.

  3. (3)

    Given a vector bundle \mathcal{E} and a line bundle \mathcal{L} on CC, the HN filtration on \mathcal{E}\otimes\mathcal{L} is obtained by tensoring the HN filtration on \mathcal{E} with \mathcal{L}.

Proof.

Because (3) follows from (1) and (2); and assertion (2) follows from (1), it suffices to prove (1). For (1), it is enough to show that

(5) deg()=deg()+rk()deg().\text{deg}(\mathcal{F}\otimes\mathcal{L})=\text{deg}(\mathcal{F})+\text{rk}(\mathcal{F})\text{deg}(\mathcal{L})\ .

This is clear when rk()=0\text{rk}(\mathcal{F})=0. In the general case, take the short exact sequence of sheaves 0′′00\rightarrow\mathcal{F}^{\prime}\rightarrow\mathcal{F}\rightarrow\mathcal{F}^{\prime\prime}\rightarrow 0, where \mathcal{F}^{\prime} is the torsion subsheaf of \mathcal{F} and ′′:=/\mathcal{F}^{\prime\prime}:=\mathcal{F}/\mathcal{F}^{\prime} is a vector bundle; note that the rank of \mathcal{F}^{\prime} is zero. Since degree is additive over short exact sequences (see section 2.6, [Pot97]), it suffices to show (5) when =′′\mathcal{F}=\mathcal{F}^{\prime\prime}, that is, when \mathcal{F} is locally free. In this case, deg()=deg(det())=deg(det()rk())\text{deg}(\mathcal{F}\otimes\mathcal{L})=\text{deg}(\text{det}(\mathcal{F}\otimes\mathcal{L}))=\text{deg}(\text{det}(\mathcal{F})\otimes\mathcal{L}^{\otimes{\text{rk}}(\mathcal{F})}) (for e.g. by Theorem 2.6.9 of [Pot97]), so deg()=deg()+rk()deg()\text{deg}(\mathcal{F}\otimes\mathcal{L})=\text{deg}(\mathcal{F})+\text{rk}(\mathcal{F})\text{deg}(\mathcal{L}) by Theorem 2.6.3, [Pot97]. ∎

In the next lemma, for a sheaf of 𝒪C\mathcal{O}_{C}-modules \mathcal{F}, \mathcal{F}^{\vee} denotes the dual sheaf Hom¯𝒪C(,𝒪C)\underline{\text{Hom}}_{\mathcal{O}_{C}}(\mathcal{F},\mathcal{O}_{C}).

Lemma 2.16.
  1. (1)

    The dual of a semistable vector bundle is semistable.

  2. (2)

    Let 0=01tt+1=0=\mathcal{E}_{0}\subset\mathcal{E}_{1}\ldots\subset\mathcal{E}_{t}\subset\mathcal{E}_{t+1}=\mathcal{E} be the HN filtration on a vector bundle \mathcal{E}. For jj between 0 and t+1t+1, set Kj=ker(t+1j)K_{j}=\text{ker}(\mathcal{E}^{\vee}\rightarrow\mathcal{E}_{t+1-j}^{\vee}). Then

    0=K0K1KtKt+1=0=K_{0}\subset K_{1}\subset\ldots\subset K_{t}\subset K_{t+1}=\mathcal{E}^{\vee}

    is the HN filtration on \mathcal{E}^{\vee}.

Proof.

(1) Let \mathcal{F} be semistable vector bundle and 𝒢\mathcal{G} be a non-zero subsheaf of \mathcal{F}^{\vee}. We show that μ(𝒢)μ()\mu(\mathcal{G})\leq\mu(\mathcal{F}^{\vee})- this is clear when rk(/𝒢)=0\text{rk}(\mathcal{F}^{\vee}/\mathcal{G})=0 as deg()=deg(𝒢)+deg(/𝒢)\text{deg}(\mathcal{F}^{\vee})=\text{deg}(\mathcal{G})+\text{deg}(\mathcal{F}^{\vee}/\mathcal{G}) and rk()=rk(𝒢)\text{rk}(\mathcal{F}^{\vee})=\text{rk}(\mathcal{G}). If rk(/𝒢)\text{rk}(\mathcal{F}^{\vee}/\mathcal{G}) is not zero, set 𝒢\mathcal{G}^{\prime} to be the inverse image of (/𝒢)tor(\mathcal{F}^{\vee}/\mathcal{G})_{\text{tor}}: the torsion subsheaf of /𝒢\mathcal{F}^{\vee}/\mathcal{G}, under the quotient map /𝒢\mathcal{F}^{\vee}\rightarrow\mathcal{F}^{\vee}/\mathcal{G}. Then 𝒢/𝒢(/𝒢)tor\mathcal{G}^{\prime}/\mathcal{G}\cong(\mathcal{F}^{\vee}/\mathcal{G})_{\text{tor}}. So 𝒢\mathcal{G} and 𝒢\mathcal{G}^{\prime} have the same rank. Since deg(𝒢)deg(𝒢)\text{deg}(\mathcal{G}^{\prime})\geq\text{deg}(\mathcal{G}), it is enough to show that μ(𝒢)μ()\mu(\mathcal{G}^{\prime})\leq\mu(\mathcal{F}). Since /𝒢\mathcal{F}/\mathcal{G}^{\prime} is a vector bundle, after dualizing we get an exact sequence:

0(/𝒢)(𝒢)0.0\rightarrow(\mathcal{F}/\mathcal{G}^{\prime})^{\vee}\rightarrow\mathcal{F}\rightarrow(\mathcal{G}^{\prime})^{\vee}\rightarrow 0.

Since \mathcal{F} is semistable, μ((𝒢))μ()\mu((\mathcal{G}^{\prime})^{\vee})\geq\mu(\mathcal{F})- see section 5.3, [Pot97]. Since for a vector bundle \mathcal{E}, deg()=deg()\text{deg}(\mathcal{E}^{\vee})=-\text{deg}(\mathcal{E}), we have μ(𝒢)μ()\mu(\mathcal{G}^{\prime})\leq\mu(\mathcal{F}^{\vee}).

(2) Kj+1/Kj(t+1j/tj)K_{j+1}/K_{j}\cong(\mathcal{E}_{t+1-j}/\mathcal{E}_{t-j})^{\vee}, so by (1) (t+1j/tj)(\mathcal{E}_{t+1-j}/\mathcal{E}_{t-j}) is semistable. Moreover, since μ(Kj+1/Kj)=μ(t+1j/tj)\mu(K_{j+1}/K_{j})=-\mu(\mathcal{E}_{t+1-j}/\mathcal{E}_{t-j}), slopes of Kj+1/KjK_{j+1}/K_{j} form a decreasing sequence. ∎

Let CC be a curve over an algebraically closed field of positive characteristic. Let ff be the absolute Frobenius endomorphism of CC. Since CC is smooth, ff is flat map (see Theorem 2.1, [Kun69]). So the pullback of the HN filtration on a given vector bundle gives a filtration of the pull back bundle by subbundles- in general this is not the HN filtration on the pullback bundle.

Theorem 2.17.

(see Theorem 2.7 [Lan04]) Let \mathcal{E} be a vector bundle on a curve CC. Then there is an n0n_{0}\in\mathbb{N} such that for nn0n\geq n_{0}, the HN filtration on (fn)()(f^{n})^{*}(\mathcal{E}) is the pullback of the HN filtration on (fn0)()(f^{n_{0}})^{*}(\mathcal{E}) via fnn0f^{n-n_{0}}.

3. Existence of Frobenius-Poincaré functions

In this section, we define the Frobenius-Poincaré function associated to a given triple (M,R,I)(M,R,I), where RR is a finitely generated \mathbb{N}-graded kk-algebra – kk has characteristic p>0p>0, (see 1.1), II is a homogeneous ideal of finite co-length and MM is a finitely generated \mathbb{Z}-graded RR-module. In Theorem 3.1 we prove that Frobenius-Poincaré functions are entire functions.

Given (M,R,I)(M,R,I) as above and a non-negative integer dd, define a sequence of functions Fn(M,R,I,d)F_{n}(M,R,I,d), where for a complex number yy,

(6) Fn(M,R,I,d)(y)=(1pn)dj=λ((MI[pn]M)j)eiyj/pn.F_{n}(M,R,I,d)(y)=(\frac{1}{p^{n}})^{d}\sum\limits_{j=-\infty}^{\infty}\lambda\left(({\frac{M}{{I}^{[{p}^{n}]}M}})_{j}\right)\ e^{-iyj/p^{n}}.

Since M/I[pn]MM/I^{[p^{n}]}M has only finitely many non-zero graded pieces, each Fn(M,R,I,d)F_{n}(M,R,I,d) is a polynomial in eiy/pne^{-iy/p^{n}}, hence is an entire function. When the context is clear, we suppress one or more of the parameters among M,R,I,dM,R,I,d in the notation Fn(M,R,I,d)F_{n}(M,R,I,d). Whenever there is no explicit reference to the parameter dd in Fn(M,R,I,d)F_{n}(M,R,I,d), it should be understood that d=dim(M)d=\text{dim}(M).

The goal in this section is to prove the following result:

Theorem 3.1.

Fix a triple (M,R,I)(M,R,I), where RR is a finitely generated \mathbb{N}-graded kk-algebra (see 1.1), II is a finite co-length homogeneous ideal, and MM is a finitely generated \mathbb{Z}-graded RR-module. The sequence of functions (Fn(M,R,I,dim(M)))n(F_{n}(M,R,I,\text{dim}(M)))_{n}, where

Fn(M,R,I,dim(M))(y)=(1pn)dim(M)j=λ((MI[pn]M)j)eiyj/pn,F_{n}(M,R,I,\text{dim}(M))(y)=(\frac{1}{p^{n}})^{\text{dim}(M)}\sum\limits_{j=-\infty}^{\infty}\lambda\left(({\frac{M}{{I}^{[{p}^{n}]}M}})_{j}\right)\ e^{-iyj/p^{n}},

converges for every complex number yy. Furthermore, the convergence is uniform on every compact subset of the complex plane and the limit is an entire function.

Theorem 3.1 motivates the next definition.

Definition 3.2.

The Frobenius-Poincaré function of the triple (M,R,I)(M,R,I) is the limit of the convergent sequence of functions

(Fn(M,R,I,dim(M)))n(F_{n}(M,R,I,\text{dim}(M)))_{n}

as defined in Theorem 3.1. The Frobenius-Poincaré function of the triple (M,R,I)(M,R,I) is denoted by F(M,R,I,dim(M))F(M,R,I,\text{dim}(M)) or alternately F(M,R,I)F(M,R,I) or just F(M)F(M) when the other parameters are clear from the context.

Before giving examples of Frobenius-Poincaré functions, we single out a limit computation.

Lemma 3.3.

Given a complex number aa, the sequence of functions (pn(1eiay/pn))n(p^{n}(1-e^{-iay/p^{n}}))_{n} converges to the function g(y)=iayg(y)=iay and the convergence is uniform on every compact subset of the complex plane.

Proof.

Using the power series representing eze^{z} around the origin, we get

pn(1eiay/pn)=iay+j=2(iay)jj!(pn)j1.p^{n}(1-e^{-iay/p^{n}})=iay+\sum_{j=2}^{\infty}\frac{(-iay)^{j}}{j!(p^{n})^{j-1}}.

So

|pn(1eiay/pn)iay|1pnj=2|ay|jj!1pne|ay|.|p^{n}(1-e^{-iay/p^{n}})-iay|\leq\frac{1}{p^{n}}\sum_{j=2}^{\infty}\frac{|ay|^{j}}{j!}\leq\frac{1}{p^{n}}e^{|ay|}.

Since e|ay|e^{|ay|} is bounded on any compact subset, we get the desired unifrom convergence. ∎

Example 3.4.

Consider the \mathbb{N}-graded polynomial ring in one variable R=𝔽p[X]R=\mathbb{F}_{p}[X] where XX has degree δ\delta\in\mathbb{N} and elements of 𝔽p\mathbb{F}_{p} have degree zero. Take II to be the ideal generated by XX. Then, for any nonzero yy\in\mathbb{C},

Fn(R)(y)=1pnj=0pn1eiyδj/pn=1pn1eiδy1eiδy/pn.F_{n}(R)(y)=\frac{1}{p^{n}}\sum\limits_{j=0}^{p^{n}-1}e^{-iy\delta j/p^{n}}=\frac{1}{p^{n}}\frac{1-e^{-i\delta y}}{1-e^{-i\delta y/p^{n}}}.

Taking limit as nn goes to infinity in the above equation and using Lemma 3.3, we get that for a non-zero complex number yy, F(R)(y)=1eiδyiδyF(R)(y)=\frac{1-e^{-i\delta y}}{i\delta y}. Note that 1eiδyiδy\frac{1-e^{-i\delta y}}{i\delta y} can be extended to an analytic function with value one at the origin. Since Fn(0)=1F_{n}(0)=1 for all nn, F(R)F(R) and the analytic extension of 1eiδyiδy\frac{1-e^{-i\delta y}}{i\delta y} are the same function.

Similar computation shows that the Frobenius-Poincaré function of the triple (R,I,R)(R,I,R) where II is the ideal generated by XtX^{t} is 1eiδtyiδy\frac{1-e^{-i\delta ty}}{i\delta y}.

Example 3.5.

Take R=𝔽p[X1,,Xn]R=\mathbb{F}_{p}[X_{1},\ldots,X_{n}] with the grading assigning degree δj\delta_{j} to XjX_{j} and degree zero to the elements of 𝔽p\mathbb{F}_{p}. Since as graded rings,

𝔽p[X1,,Xd](X1pn,X2pn,,Xdpn)𝔽p[X1]/(X1pn)𝔽p𝔽p[X2]/(X2pn)𝔽p𝔽p𝔽p[Xd]/(Xdpn),\frac{\mathbb{F}_{p}[X_{1},\ldots,X_{d}]}{(X_{1}^{p^{n}},X_{2}^{p^{n}},\ldots,X_{d}^{p^{n}})}\cong\mathbb{F}_{p}[X_{1}]/(X_{1}^{p^{n}})\otimes_{\mathbb{F}_{p}}\mathbb{F}_{p}[X_{2}]/(X_{2}^{p^{n}})\otimes_{\mathbb{F}_{p}}\ldots\otimes_{\mathbb{F}_{p}}\mathbb{F}_{p}[X_{d}]/(X_{d}^{p^{n}}),

we have Fn(R,R,(X1,,Xn))(y)=Fn(𝔽p[X1],𝔽p[X1],(X1)).Fn(𝔽p[Xd],𝔽p[Xd],(Xd))F_{n}(R,R,(X_{1},\ldots,X_{n}))(y)=F_{n}(\mathbb{F}_{p}[X_{1}],\mathbb{F}_{p}[X_{1}],(X_{1})).\ldots F_{n}(\mathbb{F}_{p}[X_{d}],\mathbb{F}_{p}[X_{d}],(X_{d})). So from Example 3.4, it follows that F(R,R,(X1,,Xd))(y)=j=1d(1eiδjyiδjy)F(R,R,(X_{1},\ldots,X_{d}))(y)=\prod\limits_{j=1}^{d}(\frac{1-e^{-i\delta_{j}y}}{i\delta_{j}y}).

Remark 3.6.

The fractional exponents of exponentials in the definition of Frobenius-Poincaré function of a triple (M,R,I)(M,R,I) comes from the natural 1pn\frac{1}{p^{n}}\mathbb{Z} grading on Fn(M)F^{n}_{*}(M). Here Fn(M)F^{n}_{*}(M) is the abelian group MM enowed with the RR-module structure coming from the restriction of scalars via the nn-th iteration of Frobenius FRn:RRF^{n}_{R}:R\rightarrow R. The 1pn\frac{1}{p^{n}}\mathbb{Z}-graded structure on Fn(M)F^{n}_{*}(M) is as follows: for an integer mm, Fn(M)m/pn=MmF^{n}_{*}(M)_{m/p^{n}}=M_{m}. For example, when RR is a domain, the 1pn\frac{1}{p^{n}}\mathbb{Z}-grading on Fn(R)F^{n}_{*}(R) described above is the one obtained by importing the natural 1pn\frac{1}{p^{n}}\mathbb{Z}-grading on R1/pnR^{1/p^{n}} via the RR-module isomorphism R1/pnFn(R)R^{1/p^{n}}\rightarrow F^{n}_{*}(R) given by the pnp^{n}-th power map.

Note that as 1pn\frac{1}{p^{n}}\mathbb{Z}-graded modules Fn(MI[pn]M)Fn(M)RR/IF^{n}_{*}(\frac{M}{{I}^{[{p}^{n}]}M})\cong F^{n}_{*}(M)\otimes_{R}R/I. Set \ell to be the degree of the field extension kpkk^{p}\subseteq k and [1/p]\mathbb{Z}[1/p] to be the ring \mathbb{Z} with pp inverted. Alternatively FnF_{n} can be expressed as,

Fn(M,R,I)(y)=(1pdim(M))nt[1/p]λk((Fn(M)R/I)t)eity.F_{n}(M,R,I)(y)=(\frac{1}{\ell p^{\text{dim}(M)}})^{n}\underset{t\in\mathbb{Z}[1/p]}{\sum}\lambda_{k}\left((F^{n}_{*}(M)\otimes R/I)_{t}\right)e^{-ity}.

That is, FnF_{n} is the Hilbert series (see Section 2.3) of Fn(M)R/IF^{n}_{*}(M)\otimes R/I normalized by (1pdim(M))n(\frac{1}{\ell p^{\text{dim(M)}}})^{n} in the ‘variable’ eiye^{-iy}. The associated Frobenius-Poincaré function is the limit of these normalized Hilbert series.

Remark 3.7.

Given a field extension kkk\subseteq k^{\prime}, RkkR\otimes_{k}k^{\prime} is a finitely generated \mathbb{N}-graded kk^{\prime}-algebra. Note that Fn(M,R,I,d)=Fn(Mkk,Rkk,Ikk,d)F_{n}(M,R,I,d)=F_{n}(M\otimes_{k}k^{\prime},R\otimes_{k}k^{\prime},I\otimes_{k}k^{\prime},d). Thus for any complex number yy, (Fn(M,R,I,d)(y))n(F_{n}(M,R,I,d)(y))_{n} converges if and only if (Fn(Mkk,Rkk,Ikk,d)(y))n(F_{n}(M\otimes_{k}k^{\prime},R\otimes_{k}k^{\prime},I\otimes_{k}k^{\prime},d)(y))_{n} converges. So in the proof of Theorem 3.1 without loss of generality we can assume that kpkk^{p}\subseteq k is a finite extension or even algebraically closed. thus we can assume k=kpk=k^{p}.

The remainder of the section is dedicated to the proof of Theorem 3.1. The proof has two main steps. First, we reduce the problem to the case where RR is an \mathbb{N}-graded domain and M=RM=R as a graded module- this reduction step is achieved in Theorem 3.16. Then we show that when RR is a domain, Fn(R,R,I)(y)F_{n}(R,R,I)(y) is uniformly Cauchy on every compact subset of \mathbb{C}. Thus Fn(R,R,I)F_{n}(R,R,I) converges uniformly on every compact subset of the complex plane. The analyticity of the limiting function then follows from Theorem 1 in Chapter 5 of [Ahl79]: a sequence of holomorphic functions on an open subset UU\subseteq\mathbb{C}, which converges uniformly on every compact subset of UU has a holomorphic limiting function.

One of the purposes of the next result is to show that in the definition of FnF_{n} in (6), it is enough to take the sum over indices jj, where |j|/pn|j|/p^{n} is bounded by a constant which is independent of nn.

Lemma 3.8.

Let MM be a finitely generated \mathbb{Z}-graded RR-module. Given ii\in\mathbb{N}, there exists a positive integer CC such that for all nn whenever |j|Cpn|j|\geq Cp^{n}, (ToriR(M,RI[pn]R))j(\text{Tor}^{R}_{i}(M,\frac{R}{{I}^{[{p}^{n}]}R}))_{j} is zero.

Proof.

We claim that it is sufficient to prove existence of some CC such that (RI[pn]R)j=0(\frac{R}{{I}^{[{p}^{n}]}R})_{j}=0 for |j|Cpn|j|\geq Cp^{n}. To see this, consider a graded minimal free resolution (see section 2.2) of the RR-module MM:

Fi{F_{i}}{\ldots}F1{F_{1}}F0{F_{0}}M{M}0.{0.}i+1\scriptstyle{\partial_{i+1}}i\scriptstyle{\partial_{i}}2\scriptstyle{\partial_{2}}1\scriptstyle{\partial_{1}}0\scriptstyle{\partial_{0}}

As a graded module, ToriR(M,RI[pn]R)\text{Tor}^{R}_{i}(M,\frac{R}{{I}^{[{p}^{n}]}R}) is a subquotient of Tor0R(Fi,RI[pn])Fi/I[pn]Fi\text{Tor}^{R}_{0}(F_{i},\frac{R}{{I}^{[{p}^{n}]}})\cong F_{i}/{I}^{[{p}^{n}]}F_{i} and FiF_{i} is a direct sum of finitely many graded modules of the form R(l)R(l), so our claim follows.

Now choose homogeneous elements of positive degrees r1,r2,,rsr_{1},r_{2},\ldots,r_{s} of RR such that R=k[r1,,rs]R=k[r_{1},\ldots,r_{s}]. Let Δ=max{deg(ri)}\Delta=\text{max}\{\text{deg}(r_{i})\} and δ=min{deg(ri)}\delta=\text{min}\{\text{deg}(r_{i})\}. Denote the ideal of generated by homogeneous elements of degree at least tt by RtR_{\geq t}. Then note that for every nn\in\mathbb{N},

(7) RnΔ(Rδ)n.R_{\geq n\Delta}\subseteq(R_{\geq\delta})^{n}.

Indeed, for each rRtr\in R_{t} where tnΔt\geq n\Delta, we can choose homogeneous elements λ1,,λs\lambda_{1},\ldots,\lambda_{s} of RR such that

r=λ1r1++λsrs.r=\lambda_{1}r_{1}+\ldots+\lambda_{s}r_{s}.

Then each λiRtΔR(n1)Δ\lambda_{i}\in R_{\geq t-\Delta}\subseteq R_{(n-1)\Delta}. Since each rir_{i} is in RδR_{\geq\delta}, the claimed assertion in (7) follows by induction on nn.

Pick m0m_{0} such that (Rδ)m0I(R_{\geq\delta})^{m_{0}}\subseteq I. Suppose the minimal number of homogeneous generators of II is μ\mu. Using (7), we get

Rm0μpnΔ(Rδ)m0μpnIμpnI[pn].R_{\geq m_{0}\mu p^{n}\Delta}\subseteq(R_{\geq\delta})^{m_{0}\mu p^{n}}\subseteq I^{\mu p^{n}}\subseteq{I}^{[{p}^{n}]}.

Therefore, if we set C=m0μΔC=m_{0}\mu\Delta, for mCpnm\geq Cp^{n}, (RI[pn]R)m=0(\frac{R}{{I}^{[{p}^{n}]}R})_{m}=0. ∎

We now bound the asymptotic growths of two length functions.

Lemma 3.9.

Let (S,m)(S,m) be Noetherian local ring containing a field of positive characteristic pp, JJ be an mm-primary ideal. For any finitely generated SS-module NN, there exist positive constants C1,C2C_{1},C_{2} such that for all nn\in\mathbb{N},

λS(N/J[pn]N)C1(pn)dim(N)andλS(Tor1S(N,SJ[pn]S))C2(pn)dim(N).\lambda_{S}(N/{J}^{[{p}^{n}]}N)\leq C_{1}(p^{n})^{\text{dim}(N)}\,\,\,\,\text{and}\,\,\hskip 14.22636pt\lambda_{S}(\text{Tor}^{S}_{1}(N,\frac{S}{{J}^{[{p}^{n}]}S}))\leq C_{2}(p^{n})^{\text{dim}(N)}.
Proof.

The assertion on the growth of λS(N/J[pn]N)\lambda_{S}(N/{J}^{[{p}^{n}]}N) is standard, for example see Lemma 3.5 of [Hun13] for a proof. For the other assertion, we present a simplified version of the argument in Lemma 7.2 of [Hun13]. Suppose that JJ is generated by f1,,fμf_{1},\ldots,f_{\mu}. Let KK_{\bullet} be the Koszul complex of f1,f2,,fμf_{1},f_{2},\ldots,f_{\mu}. Recall that the Frobenius functor 𝔉\mathfrak{F}, from the category of SS-modules to itself, is the scalar extension via the Frobenius FS:SSF_{S}:S\rightarrow S (see page 7, [HH93]). Let 𝔉n(K)\mathfrak{F}^{n}(K_{\bullet}) stand for the complex of SS-modules obtained by applying nn-th iteration of 𝔉\mathfrak{F} to the terms and the boundary maps of KK_{\bullet}. Then the part 𝔉n(K1)𝔉n(K0)\mathfrak{F}^{n}(K_{1})\rightarrow\mathfrak{F}^{n}(K_{0}) of 𝔉n(K)\mathfrak{F}^{n}(K_{\bullet}) can be extended to a free resolution of the SS-module S/J[pn]SS/{J}^{[{p}^{n}]}S. So Tor1S(N,SJ[pn]S)\text{Tor}^{S}_{1}(N,\frac{S}{{J}^{[{p}^{n}]}S}) is isomorphic to a quotient of H1(NS𝔉n(K))H_{1}(N\otimes_{S}\mathfrak{F}^{n}(K_{\bullet})). Hence λ(Tor1S(N,SJ[pn]S))λS(H1(NS𝔉n(K)))\lambda(\text{Tor}^{S}_{1}(N,\frac{S}{{J}^{[{p}^{n}]}S}))\leq\lambda_{S}(H_{1}(N\otimes_{S}\mathfrak{F}^{n}(K_{\bullet}))). The conclusion follows from Theorem 6.6 of [HH93], which guarantees that there is a constant C2C_{2} such that for all nn, λS(H1(NS𝔉n(K)))C2(pn)dim(N)\lambda_{S}(H_{1}(N\otimes_{S}\mathfrak{F}^{n}(K_{\bullet})))\leq C_{2}(p^{n})^{\text{dim}(N)}.∎

The next result bounds the growth of the sequence (Fn(M,R,I,d))n(F_{n}(M,R,I,d))_{n} on a given compact subset of the complex plane.

Proposition 3.10.

Let MM be a finitely generated \mathbb{Z}-graded RR-module.

  1. (1)

    Given a compact subset AA\subseteq\mathbb{C}, there exists a constant DD such that for all yAy\in A and nn\in\mathbb{N},

    |Fn(M,R,I,d)(y)|(1pn)ddim(M)D.|F_{n}(M,R,I,d)(y)|\leq(\frac{1}{p^{n}})^{d-\text{dim}(M)}D.
  2. (2)

    On a given compact subset of \mathbb{C}, when ddim(M)d\geq\text{dim}(M), the sequence (Fn(M,R,I,d))n(F_{n}(M,R,I,d))_{n} is uniformly bounded and when d>dim(M)d>\text{dim}(M), the sequence (Fn(M,R,I,d))n(F_{n}(M,R,I,d))_{n} uniformly converges to the constant function taking value zero.

Proof.

Assertion (2) is immediate from (1).

We now prove (1). Let CC^{\prime} be a positive constant such that for yAy\in A, |y|C|y|\leq C^{\prime}. According to Lemma 3.8, we can choose a positive constant CC such that for all nn\in\mathbb{N} and |j|>Cpn|j|>Cp^{n}, (M/I[pn]M)j=0(M/{I}^{[{p}^{n}]}M)_{j}=0. For yAy\in A and |j|Cpn|j|\leq Cp^{n},

(8) |eiyj/pn|2=eiyj/pn.eiy¯j/pn=e2Re(iyj/pn)e2|iyj/pn|=e2|y||j|/pne2CC.|e^{-iyj/p^{n}}|^{2}=e^{-iyj/p^{n}}.e^{i\bar{y}j/p^{n}}=e^{2\text{Re}(-iyj/p^{n})}\leq e^{2|-iyj/p^{n}|}=e^{2|y||j|/p^{n}}\leq e^{2CC^{\prime}}\ .

Now note that,

|Fn(M,d)(y)|=|(1pn)d|j|Cpnλ((MI[pn]M)j)eiyj/pn|(1pn)d|j|Cpnλ((MI[pn]M)j)|eiyj/pn|,|F_{n}(M,d)(y)|=|(\frac{1}{p^{n}})^{d}\underset{|j|\leq Cp^{n}}{\sum}\lambda\left((\frac{M}{{I}^{[{p}^{n}]}M})_{j}\right)e^{-iyj/p^{n}}|\leq(\frac{1}{p^{n}})^{d}\underset{|j|\leq Cp^{n}}{\sum}\lambda\left((\frac{M}{{I}^{[{p}^{n}]}M})_{j}\right)|e^{-iyj/p^{n}}|,

So using (8) first and then Lemma 3.9, we get,

|Fn(M,d)(y)|(1pn)dλ(MI[pn]M)eCC(1pn)ddim(M)C1eCC.|F_{n}(M,d)(y)|\leq(\frac{1}{p^{n}})^{d}\lambda(\frac{M}{{I}^{[{p}^{n}]}M})e^{CC^{\prime}}\leq(\frac{1}{p^{n}})^{d-\text{dim}(M)}C_{1}e^{CC^{\prime}}.

for some constant C1C_{1}. ∎

The next few results are aimed towards Theorem 3.16.

Lemma 3.11.

Let 0{0}K{K}M1{M_{1}}M2{M_{2}}C{C}0{0}ϕ1\scriptstyle{\phi_{1}}ϕ\scriptstyle{\phi}ϕ2\scriptstyle{\phi_{2}} be an exact sequence of finitely generated \mathbb{Z}-graded RR-modules (i.e. assume the boundary maps preserve the respective gradings). Let dd be an integer greater than both dim(K)anddim(C)\text{dim}(K)\,\text{and}\,\,\text{dim}(C).

  1. (1)

    Given a compact subset AA\subseteq\mathbb{C}, there exists a constant DD, such that for all yAy\in A and nn\in\mathbb{N},

    |Fn(M2,R,I,d)(y)Fn(M1,R,I,d)(y)|Dpn.|F_{n}(M_{2},R,I,d)(y)-F_{n}(M_{1},R,I,d)(y)|\leq\frac{D}{p^{n}}.
  2. (2)

    The sequence of functions (Fn(M2,R,I,d)Fn(M1,R,I,d))n(F_{n}(M_{2},R,I,d)-F_{n}(M_{1},R,I,d))_{n} converges to the constant function zero and the convergence is uniform on every compact subset of \mathbb{C}.

Proof.

We prove assertion (1) below, assertion (2) is immediate from assertion (1).

Break the given exact sequence into two short exact sequences:

(*) 0KM1Im(ϕ)0ϕ1ϕ,\leavevmode\hbox to204.55pt{\vbox to17.42pt{\pgfpicture\makeatletter\hbox{\hskip 102.27605pt\lower-8.75955pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{}{}{}{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{\offinterlineskip{}{}{{{}}{{}}{{}}{{}}{{}}}{{{}}}{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-102.27605pt}{-8.65971pt}\pgfsys@invoke{ }\hbox{\vbox{\halign{\pgf@matrix@init@row\pgf@matrix@step@column{\pgf@matrix@startcell#\pgf@matrix@endcell}&#\pgf@matrix@padding&&\pgf@matrix@step@column{\pgf@matrix@startcell#\pgf@matrix@endcell}&#\pgf@matrix@padding\cr\hfil\hskip 6.80554pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-2.5pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${0}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}}}&\hskip 6.80554pt\hfil&\hfil\hskip 32.90967pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-4.60416pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${K}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}}}&\hskip 8.9097pt\hfil&\hfil\hskip 35.10133pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-6.79582pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${M_{1}}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}}}&\hskip 11.10136pt\hfil&\hfil\hskip 42.15392pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-13.8484pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${Im(\phi)}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}}}&\hskip 18.15395pt\hfil&\hfil\hskip 33.30553pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-5.00002pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${0}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}&\hskip 9.30556pt\hfil\cr}}}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}}{{{{}}}{{}}{{}}{{}}{{}}{{}}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} { {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{-88.46497pt}{-6.15971pt}\pgfsys@lineto{-65.26495pt}{-6.15971pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-65.06497pt}{-6.15971pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{{{}{}}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{-46.6456pt}{-6.15971pt}\pgfsys@lineto{-23.44559pt}{-6.15971pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-23.2456pt}{-6.15971pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-37.93103pt}{-2.44585pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{$\scriptstyle{\phi_{1}}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{{{}{}}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{-0.4429pt}{-6.15971pt}\pgfsys@lineto{22.75711pt}{-6.15971pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{22.95709pt}{-6.15971pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{9.27167pt}{-2.44585pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{$\scriptstyle{\phi}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{59.86496pt}{-6.15971pt}\pgfsys@lineto{83.06497pt}{-6.15971pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{83.26495pt}{-6.15971pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{{ {}{}{}}}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}\ ,
(**) 0{0}Im(ϕ){Im(\phi)}M2{M_{2}}C{C}0.{0\ .}ϕ2\scriptstyle{\phi_{2}}

Now apply RRI[pn]R\otimes_{R}\frac{R}{{I}^{[{p}^{n}]}R} to (*3) and (**3); the corresponding long exact sequences of Tor modules give the following two exact sequences of graded modules for each nn:

(n*_{n}) KI[pn]K{\frac{K}{{I}^{[{p}^{n}]}K}}M1I[pn]M1{\frac{M_{1}}{{I}^{[{p}^{n}]}M_{1}}}Im(ϕ)I[pn]Im(ϕ){\frac{Im(\phi)}{{I}^{[{p}^{n}]}Im(\phi)}}0,{0\ ,}ϕ1,n\scriptstyle{\phi_{1,n}}
(n**_{n}) Tor1R(C,RI[pn]R){\text{Tor}^{R}_{1}(C,\frac{R}{{I}^{[{p}^{n}]}R})}Im(ϕ)I[pn]Im(ϕ){\frac{Im(\phi)}{{I}^{[{p}^{n}]}Im(\phi)}}M2I[pn]M2{\frac{M_{2}}{{I}^{[{p}^{n}]}M_{2}}}CI[pn]C{\frac{C}{{I}^{[{p}^{n}]}C}}0.{0\ .}τn\scriptstyle{\tau_{n}}

Using (n*_{n}) and (n**_{n}), for each jj\in\mathbb{Z}, we get

λ((M1I[pn]M1)j)=λ((ϕ1,n(KI[pn]K))j)+λ((Im(ϕ)I[pn]Im(ϕ))j),\lambda\left((\frac{M_{1}}{{I}^{[{p}^{n}]}M_{1}})_{j}\right)=\lambda\left((\phi_{1,n}(\frac{K}{{I}^{[{p}^{n}]}K}))_{j}\right)+\lambda\left((\frac{Im(\phi)}{{I}^{[{p}^{n}]}Im(\phi)})_{j}\right),
λ((M2I[pn]M2)j)=λ((Im(ϕ)I[pn]Im(ϕ))j)λ(τn(Tor1R(C,RI[pn]R))j)+λ((CI[pn]C)j).\lambda\left((\frac{M_{2}}{{I}^{[{p}^{n}]}M_{2}})_{j}\right)=\lambda\left((\frac{Im(\phi)}{{I}^{[{p}^{n}]}Im(\phi)})_{j}\right)-\lambda\left(\tau_{n}(\text{Tor}^{R}_{1}(C,\frac{R}{{I}^{[{p}^{n}]}R}))_{j}\right)+\lambda\left((\frac{C}{{I}^{[{p}^{n}]}C})_{j}\right).

Therefore,

(9) Fn(M2,R,I,d)(y)Fn(M1,R,I,d)(y)=Fn(C,R,I,d)(y)(1pn)dj=λ((τn(Tor1R(C,RI[pn]R))j)eiyj/pn(1pn)dj=λ((ϕ1,n(KI[pn]K))j)eiyj/pn.\begin{split}&F_{n}(M_{2},R,I,d)(y)-F_{n}(M_{1},R,I,d)(y)\\ &=F_{n}(C,R,I,d)(y)-(\frac{1}{p^{n}})^{d}\sum\limits_{j=-\infty}^{\infty}\lambda\left(({\tau_{n}(\text{Tor}^{R}_{1}(C,\frac{R}{{I}^{[{p}^{n}]}R})})_{j}\right)\ e^{-iyj/p^{n}}\\ &-(\frac{1}{p^{n}})^{d}\sum\limits_{j=-\infty}^{\infty}\lambda\left(({\phi_{1,n}(\frac{K}{{I}^{[{p}^{n}]}K})})_{j}\right)\ e^{-iyj/p^{n}}\ .\end{split}

By Lemma 3.8, one can choose a positive integer C1C_{1} such that given any nn and all mm such that |m|>C1pn|m|>C_{1}p^{n},

(CI[pn]C)m=(Tor1R(C,RI[pn]R))m=(KI[pn]K)m=0.(\frac{C}{{I}^{[{p}^{n}]}C})_{m}=(\text{Tor}^{R}_{1}(C,\frac{R}{{I}^{[{p}^{n}]}R}))_{m}=(\frac{K}{{I}^{[{p}^{n}]}K})_{m}=0.

Since AA is compact, there is a constant C2C_{2} such that for all jj, where |j|C1pn|j|\leq C_{1}p^{n} and for yAy\in A, |eiyj/pn|C2|e^{-iyj/p^{n}}|\leq C_{2}- the argument is similar to that in (8). Using (9), we conclude that for yAy\in A,

|Fn(M2,R,I,d)(y)Fn(M1,R,I,d)(y)|(1pn)d|j|C1pn[λ((CI[pn]C)j)+λ((Tor1R(C,RI[pn]R))j)+λ((KI[pn]K)j)]|eiyj/pn|C2(1pn)d[λ(CI[pn]C)+λ(Tor1R(C,RI[pn]R))+λ(KI[pn]K)].\begin{split}&|F_{n}(M_{2},R,I,d)(y)-F_{n}(M_{1},R,I,d)(y)|\\ &\leq(\frac{1}{p^{n}})^{d}\underset{|j|\leq C_{1}p^{n}}{\sum}[\lambda((\frac{C}{{I}^{[{p}^{n}]}C})_{j})+\lambda((\text{Tor}^{R}_{1}(C,\frac{R}{{I}^{[{p}^{n}]}R}))_{j})+\lambda((\frac{K}{{I}^{[{p}^{n}]}K})_{j})]|e^{-iyj/p^{n}}|\\ &\leq C_{2}(\frac{1}{p^{n}})^{d}[\lambda(\frac{C}{{I}^{[{p}^{n}]}C})+\lambda(\text{Tor}^{R}_{1}(C,\frac{R}{{I}^{[{p}^{n}]}R}))+\lambda(\frac{K}{{I}^{[{p}^{n}]}K})]\ .\end{split}

Since both dim(C)\text{dim}(C) and dim(K)\text{dim}(K) are less than dd, the desired result follows from Lemma 3.9. ∎

Recall that for an integer hh, M(h)M(h) denotes the RR-module MM but with a different \mathbb{Z}-grading: the nn-th graded piece of M(h)M(h) is Mn+hM_{n+h}. From now on, we use the terminology set in the next definition.

Definition 3.12.

Whenever the sequence of complex numbers (Fn(M,R,I,d)(y))n(F_{n}(M,R,I,d)(y))_{n} (see (6)) converges, we set

F(M,R,I,d)(y)=limnFn(M,R,I,d)(y).F(M,R,I,d)(y)=\underset{n\to\infty}{\lim}F_{n}(M,R,I,d)(y).

In the case d=dim(M)d=\text{dim}(M), we set F(M,R,I)(y)=F(M,R,I,dim(M))(y)F(M,R,I)(y)=F(M,R,I,\text{dim}(M))(y). Analogously we use F(M)(y)F(M)(y) when R,IR,I are clear from the context.

Proposition 3.13.

Let RR be a finitely generated \mathbb{N}-graded kk-algebra, II be a homogeneous ideal of finite co-length and MM be a finitely generated \mathbb{Z}-graded RR-module. Fix an integer hh.

  1. (1)

    Given a compact subset AA\subseteq\mathbb{C}, there exists a constant DD such that for all yAy\in A and all nn,

    |Fn(M(h),R,I,d)(y)Fn(M,R,I,d)(y)|Dpn|Fn(M,R,I,d)(y)|.|F_{n}(M(h),R,I,d)(y)-F_{n}(M,R,I,d)(y)|\leq\frac{D}{p^{n}}|F_{n}(M,R,I,d)(y)|.
  2. (2)

    For any complex number yy, (Fn(M,R,I,d)(y))n(F_{n}(M,R,I,d)(y))_{n} converges if and only if
    (Fn(M(h),R,I,d)(y))n(F_{n}(M(h),R,I,d)(y))_{n} converges. When both of these converge, their limits are equal.

Proof.

(1) Note that

Fn(M(h),R,I,d)(y)=(1pn)dj=λ((MI[pn]M)j+h)eiy(j+h)/pneiyh/pn=eiyh/pnFn(M,R,I,d)(y).F_{n}(M(h),R,I,d)(y)=(\frac{1}{p^{n}})^{d}\sum\limits_{j=-\infty}^{\infty}\lambda((\frac{M}{{I}^{[{p}^{n}]}M})_{j+h})e^{-iy(j+h)/p^{n}}e^{iyh/p^{n}}=e^{iyh/p^{n}}F_{n}(M,R,I,d)(y).

Thus,

|Fn(M(h),R,I,d)(y)Fn(M,R,I,d)(y)|=|eiyh/pn1||Fn(M,R,I,d)(y)|.|F_{n}(M(h),R,I,d)(y)-F_{n}(M,R,I,d)(y)|=|e^{iyh/p^{n}}-1||F_{n}(M,R,I,d)(y)|.

Since AA is bounded, it follows from Lemma 3.3, there is a constant DD such that for yAy\in A, |1eihy/pn|Dpn|1-e^{-ihy/p^{n}}|\leq\frac{D}{p^{n}}.

(2) It follows from the first assertion that whenever (Fn(M,d)(y))n(F_{n}(M,d)(y))_{n} converges, the sequence Fn(M(h),d)(y)F_{n}(M(h),d)(y) also converges and the two limits coincide. The other direction follows from the observation that as a graded module MM is isomorphic to (M(h))(h)(M(h))(-h). ∎

Lemma 3.14.

Let RSR\rightarrow S be a degree preserving finite homomorphism of finitely generated \mathbb{N}-graded kk-algebras. For any finitely generated \mathbb{Z}-graded SS-module NN and any complex number yy, (Fn(N,R,I,d)(y))n(F_{n}(N,R,I,d)(y))_{n} converges if and only if (Fn(N,S,IS,d)(y))n(F_{n}(N,S,IS,d)(y))_{n} converges. When both of these converge F(N,R,I,d)(y)=F(N,S,IS,d)(y)F(N,R,I,d)(y)=F(N,S,IS,d)(y).

Proof.

Since the RR-module structure on NN comes via the restriction of scalars, for each nn, the two kk-vector spaces (NI[pn]N)j(\frac{N}{{I}^{[{p}^{n}]}N})_{j} and (N(IS)[pn]N)j(\frac{N}{{(IS)}^{[{p}^{n}]}N})_{j} are isomorphic. Thus Fn(N,R,I,d)(y)=Fn(N,S,IS,d)(y)F_{n}(N,R,I,d)(y)=F_{n}(N,S,IS,d)(y) and the conclusion follows. ∎

Note that, for a finitely generated \mathbb{N}-graded kk-algebra RR, Rpe={rpe|rR}R^{p^{e}}=\{r^{p^{e}}\,|\,r\in R\} is an \mathbb{N}-graded subring of RR- the \mathbb{N}-grading on RpeR^{p^{e}} will refer to this grading.

Proposition 3.15.

Let kk be a field of characteristic p>0p>0 such that kpkk^{p}\subseteq k is finite. Let RR be a finitely generated \mathbb{N}-graded kk-algebra, II be a homogeneous ideal of finite co-length and MM be a finitely generated \mathbb{Z}-graded RR-module. Given two non-negative integers d,md,m and a complex number yy,

  1. (1)

    Denote the image of II in RpmR^{p^{m}} under the pmp^{m}-th power map by IRpmIR^{p^{m}}. Then

    Fn+m(M,R,I,d)(y)=1pmd[k:kpm]Fn(M,Rpm,IRpm,d)(y/pm).\begin{split}F_{n+m}(M,R,I,d)(y)&=\frac{1}{p^{md}[k:k^{p^{m}}]}F_{n}(M,R^{p^{m}},IR^{p^{m}},d)(y/p^{m})\ .\end{split}
  2. (2)

    When (Fn(M,R,I,d)(y))n(F_{n}(M,R,I,d)(y))_{n} converges,

    F(M,R,I,d)(y)=1pmd[k:kpm]F(M,Rpm,IRpm,d)(y/pm).F(M,R,I,d)(y)=\frac{1}{p^{md}[k:k^{p^{m}}]}F(M,R^{p^{m}},IR^{p^{m}},d)(y/p^{m}).
  3. (3)

    If RR is reduced, for all nn

    Fn(Rp,Rp,IRp,d)(y/p)=Fn(R,R,I,d)(y).F_{n}(R^{p},R^{p},IR^{p},d)(y/p)=F_{n}(R,R,I,d)(y).
Proof.

Given nn\in\mathbb{N},

Fn(M,Rpm,IRpm,d)(y/pm)=(1pn)dj=λkpm((M(IRpm)[pn]M)j)eiyj/pn+m=(1pn)dj=λkpm((MI[pn+m]M)j)eiyj/pn+m=pmd[k:kpm](1pn+m)dj=λk((MI[pn+m]M)j)eiyj/pn+m=pmd[k:kpm]Fn+m(M,R,I,d)(y).\begin{split}F_{n}(M,R^{p^{m}},IR^{p^{m}},d)(y/p^{m})&=(\frac{1}{p^{n}})^{d}\sum\limits_{j=-\infty}^{\infty}\lambda_{k^{p^{m}}}((\frac{M}{{(IR^{p^{m}})}^{[{p}^{n}]}M})_{j})e^{-iyj/p^{n+m}}\\ &=(\frac{1}{p^{n}})^{d}\sum\limits_{j=-\infty}^{\infty}\lambda_{k^{p^{m}}}((\frac{M}{{I}^{[{p}^{n+m}]}M})_{j})e^{-iyj/p^{n+m}}\\ &=p^{md}[k:k^{p^{m}}](\frac{1}{p^{n+m}})^{d}\sum\limits_{j=-\infty}^{\infty}\lambda_{k}((\frac{M}{{I}^{[{p}^{n+m}]}M})_{j})e^{-iyj/p^{n+m}}\\ &=p^{md}[k:k^{p^{m}}]F_{n+m}(M,R,I,d)(y)\ .\end{split}

(1) and (2) follows directly from the calculation above.

Now, we verify (3). Since RR is reduced, the Frobenius FR:RRF_{R}:R\rightarrow R induces an isomorphism onto RpR^{p}; it takes RjR_{j} to (Rp)jp(R^{p})_{jp}. Thus for each n,jn,j\in\mathbb{N}, FRF_{R} induces an isomorphism of abelian groups from (RI[pn])j(\frac{R}{{I}^{[{p}^{n}]}})_{j} to (Rp(IRp)[pn]Rp)jp(\frac{R^{p}}{{(IR^{p})}^{[{p}^{n}]}R^{p}})_{jp}. So, λk((RI[pn])j)=λkp((Rp(IRp)[pn]Rp)jp)\lambda_{k}((\frac{R}{{I}^{[{p}^{n}]}})_{j})=\lambda_{k^{p}}((\frac{R^{p}}{{(IR^{p})}^{[{p}^{n}]}R^{p}})_{jp}). Now,

Fn(Rp,Rp,IRp,d)(y/p)=(1pn)dj=0λkp((Rp(IRp)[pn]Rp)jp)eiyj/pn=(1pn)dj=0λk((RI[pn]R)j)eiyj/pn.\begin{split}F_{n}(R^{p},R^{p},IR^{p},d)(y/p)&=(\frac{1}{p^{n}})^{d}\sum\limits_{j=0}^{\infty}\lambda_{k^{p}}((\frac{R^{p}}{{(IR^{p})}^{[{p}^{n}]}R^{p}})_{jp})e^{-iyj/p^{n}}\\ &=(\frac{1}{p^{n}})^{d}\sum\limits_{j=0}^{\infty}\lambda_{k}((\frac{R}{{I}^{[{p}^{n}]}R})_{j})e^{-iyj/p^{n}}.\end{split}

The rightmost quantity on the above equality is Fn(R,R,IR,d)(y)F_{n}(R,R,IR,d)(y). ∎

Theorem 3.16.

Let RR be a finitely generated \mathbb{N}-graded kk-algebra, II be a homogeneous ideal of finite co-length. Let MM be a finitely generated \mathbb{Z}-graded RR-module of Krull dimension dd and Q1,,QlQ_{1},\ldots,Q_{l} be the dd dimensional minimal prime ideals in the support of MM. Let mm be such that nil(R)[pm]{\text{nil}(R)}^{[{p}^{m}]} is zero, where nil(R)\text{nil}(R) is the nilradical of RR.

  1. (1)

    Given a compact subset AA\subseteq\mathbb{C}, there exists a constant DD such that for all yAy\in A,

    |Fn+m(M,R,I,d)(y)j=1lλRQj(MQj)Fn(RQj,RQj,IRQj,d)(y)|Dpn.|F_{n+m}(M,R,I,d)(y)-\sum\limits_{j=1}^{l}\lambda_{R_{Q_{j}}}(M_{Q_{j}})F_{n}(\frac{R}{Q_{j}},\frac{R}{Q_{j}},I\frac{R}{Q_{j}},d)(y)|\leq\frac{D}{p^{n}}\ .
  2. (2)

    Given yy\in\mathbb{C}, whenever (Fn(RQj,RQj,IRQj,d)(y))n(F_{n}(\frac{R}{Q_{j}},\frac{R}{Q_{j}},I\frac{R}{Q_{j}},d)(y))_{n} is convergent for every jj, (Fn(M,R,I,d)(y))n(F_{n}(M,R,I,d)(y))_{n} is also convergent and

    (10) F(M,R,I,d)(y)=j=1lλRQj(MQj)F(RQj,RQj,IRQj,d)(y).F(M,R,I,d)(y)=\sum\limits_{j=1}^{l}\lambda_{R_{Q_{j}}}(M_{Q_{j}})F(\frac{R}{Q_{j}},\frac{R}{Q_{j}},I\frac{R}{Q_{j}},d)(y)\ .

To prove 3.16, we first establish several lemmas to handle the reduced case. The main algebraic input into the proof is the following.

Lemma 3.17.

Let RR be a reduced finitely generated \mathbb{N}-graded kk-algebra and let QQ be a minimal prime ideal of RR. Let UU be the multiplicative set of homogeneous elements in RQR-Q. Then,

  1. (1)

    The ideal QU1RQU^{-1}R is zero. Moreover, there is a field kk such that U1RU^{-1}R is isomorphic to either kk or k[t,t1]k[t,t^{-1}], where tt is an indeterminate over kk.

  2. (2)

    Set r=λRQ(NQ)r=\lambda_{R_{Q}}(N_{Q}). Then there exist integers h1,,hrh_{1},\ldots,h_{r} and a grading preserving RR-linear morphism

    ϕQ:j=1rRQ(hj)N,\phi_{Q}:\bigoplus\limits_{j=1}^{r}\frac{R}{Q}(-h_{j})\longrightarrow N,

    such that the map induced by ϕQ\phi_{Q} after localizing at QQ is an isomorphism.

Proof.

(1)Any non-zero homogeneous prime ideal of U1RU^{-1}R is the extension of a homogeneous prime ideal of RR contained in RUR\setminus U; so is contained in QQ. As QQ is minimal, we conclude that U1RU^{-1}R has a unique prime ideal namely QU1RQU^{-1}R. Since RR is reduced, so is U1RU^{-1}R. So QU1R=0QU^{-1}R=0. Since U1RU^{-1}R does not have any non-zero homogeneous prime ideal, every non-zero homogeneous element of RR is a unit. Therefore U1RU^{-1}R is isomorphic to either kk or k[t,t1]k[t,t^{-1}] for some field kk; see [BH98, Lemma 1.5.7].

(2) Since RR is reduced and QQ is a minimal prime, RQR_{Q} is a field. We produce rr homogeneous elements of NN, each of which is annihilated by QQ and their images in NQN_{Q} form an RQR_{Q}-basis of NQN_{Q}. For that, start with rr homogeneous elements m1,,mrm_{1}^{\prime},\ldots,m_{r}^{\prime} such that {m11,,mr1}\{\frac{m_{1}^{\prime}}{1},\ldots,\frac{m_{r}^{\prime}}{1}\} is an RQR_{Q}-basis of NQN_{Q}. Since by part (1) QU1RQU^{-1}R is the zero ideal and QQ is finitely generated, we can pick an element ss in UU such that ss annihilates QQ. Now set mj=smjm_{j}=sm_{j}^{\prime} for each jj. Each mjm_{j} is annihilated by QQ. Since ss is not in QQ, the images of m1,,mrm_{1},\ldots,m_{r} in NQN_{Q} form an RQR_{Q}-basis of NQN_{Q}.

Now, set hj=deg(mj)h_{j}=\text{deg}(m_{j}). Let

ϕQ:j=1rRQ(hj)N\phi_{Q}:\bigoplus\limits_{j=1}^{r}\frac{R}{Q}(-h_{j})\longrightarrow N

be the RR-linear map sending 1RQ(hj)1\in\frac{R}{Q}(-h_{j}) to mjm_{j}. Clearly ϕQ\phi_{Q} preserves gradings. Since the images of m1,,mrm_{1},\ldots,m_{r} form an RQR_{Q}-basis of MQM_{Q}, the map induced by ϕQ\phi_{Q} after localizing at QQ is an isomorphism, so our desired conclusion in Lemma 3.18 follows. ∎

Lemma 3.18.

Suppose that RR is reduced and let P1,P2,,PtP_{1},P_{2},\ldots,P_{t} be those among the minimal prime ideals of RR such that dim(R)=dim(RPj)\text{dim}(R)=\text{dim}(\frac{R}{P_{j}}). Let NN be a finitely generated \mathbb{Z}-graded RR-module. For each jj, where 1jt1\leq j\leq t, let rj=λRPj(NPj)r_{j}=\lambda_{R_{P_{j}}}(N_{P_{j}}). Then there exist integers hj,njh_{j,n_{j}} where 1jt,1njrj1\leq j\leq t,1\leq n_{j}\leq r_{j} and a degree preserving RR-linear map,

ϕ:j=1tnj=1rjRPj(hj,nj)N,\phi:\bigoplus\limits_{j=1}^{t}\bigoplus\limits_{n_{j}=1}^{r_{j}}\frac{R}{P_{j}}(-h_{j,n_{j}})\longrightarrow N,

such that the dimR(ker(ϕ))<dim(R)\text{dim}_{R}(\text{ker}(\phi))<\text{dim}(R), dimR(coker(ϕ))<dim(R)\text{dim}_{R}(\text{coker}(\phi))<\text{dim}(R).

Proof.

Consider for each jj, 1jt1\leq j\leq t, a ϕPj\phi_{P_{j}} as in assertion (2) of Lemma 3.17. Let ϕ\phi be the map induced by these ϕPj\phi_{P_{j}}’s. Since P1,,PtP_{1},\ldots,P_{t} are all distinct minimal primes, after localizing at any PjP_{j}, the maps induced by ϕ\phi and ϕPj\phi_{P_{j}} coincide. So the map induced by ϕ\phi after localizing at each PjP_{j} is an isomorphism. Hence, none of the supports of kernel and cokernel of ϕ\phi include any of P1,,PtP_{1},\ldots,P_{t}. Since P1,,PtP_{1},\ldots,P_{t} are precisely the minimal primes of RR of maximal dimension, Lemma 3.18 is proved. ∎

Proof of Theorem 3.16:.

The second assertion follows from the first one; so we just prove the first assertion below.

By Remark 3.7 we can assume that kpkk^{p}\subseteq k is a finite extension. Using Lemma 3.14 we can replace (M,R,I)(M,R,I) by (M,RAnn(M),IRAnn(M))(M,\frac{R}{\text{Ann}(M)},I\frac{R}{\text{Ann}(M)}). So we assume that d=dim(R)d=\text{dim}(R).

First we additionally assume that RR is reduced and show that taking m=0m=0 works in assertion (1). By Lemma 3.14, for all jj and nn,

Fn(RQj,RQj,IRQj,d)(y)=Fn(RQj,R,I,d)(y).F_{n}(\frac{R}{Q_{j}},\frac{R}{Q_{j}},I\frac{R}{Q_{j}},d)(y)=F_{n}(\frac{R}{Q_{j}},R,I,d)(y).

Assertion (1) follows from direct applications of Lemma 3.18, assertion (1) of Lemma 3.11 and assertion (1) of Proposition 3.13.

We now prove assertion (1) of Theorem 3.16 without assuming RR is reduced. We use the Frobenius endomorphism to pass to the reduced case. Pick an mm such that nil(R)[pm]=0{\text{nil}(R)}^{[{p}^{m}]}=0. Then the kernel of the mm-th iteration of the Frobenius FRm:RRF_{R}^{m}:R\rightarrow R is nil(R)\text{nil}(R); thus RpmR^{p^{m}}- the image of FRmF_{R}^{m}- is reduced. Recall RpmR^{p^{m}} inherits the graded structure of RR. The dd-dimensional minimal primes of RpmR^{p^{m}} in the support of the RpmR^{p^{m}} module MM are precisely of Q1Rpm,,QlRpmQ_{1}R^{p^{m}},\ldots,Q_{l}R^{p^{m}} the respective images under the pmp^{m}-th power map. Since RpmR^{p^{m}} is reduced and 1pmA:={z/pm|zA}\frac{1}{p^{m}}A:=\{z/p^{m}\,|\,z\in A\} is compact, we can find a DD such that for each yAy\in A and all nn,

(11) |Fn(M,Rpm,IRpm)(y/pm)j=1lλRpmQjRpm((M)QjRpm)Fn(RpmQjRpm,RpmQjRpm,IRpmQjRpm)(y/pm)|\displaystyle|F_{n}(M,R^{p^{m}},IR^{p^{m}})(y/p^{m})-\sum\limits_{j=1}^{l}\lambda_{R^{p^{m}}_{Q_{j}R^{p^{m}}}}\left((M)_{Q_{j}R^{p^{m}}}\right)F_{n}(\frac{R^{p^{m}}}{Q_{j}R^{p^{m}}},\frac{R^{p^{m}}}{Q_{j}R^{p^{m}}},I\frac{R^{p^{m}}}{Q_{j}R^{p^{m}}})(y/p^{m})|
(12) Dpn.\displaystyle\leq\frac{D}{p^{n}}\ .

For each jj, the graded ring RpmQjRpm\frac{R^{p^{m}}}{Q_{j}R^{p^{m}}} is isomorphic to the graded subring (RQj)pmRQj(\frac{R}{Q_{j}})^{p^{m}}\subset\frac{R}{Q_{j}}; so for all nn and yy\in\mathbb{C},

Fn((RQj)pm,(RQj)pm,I(RQj)pm,d)(y)=Fn(RpmQjRpm,RpmQjRpm,IRpmQjRpm,d)(y).F_{n}((\frac{R}{Q_{j}})^{p^{m}},(\frac{R}{Q_{j}})^{p^{m}},I(\frac{R}{Q_{j}})^{p^{m}},d)(y)=F_{n}(\frac{R^{p^{m}}}{Q_{j}R^{p^{m}}},\frac{R^{p^{m}}}{Q_{j}R^{p^{m}}},I\frac{R^{p^{m}}}{Q_{j}R^{p^{m}}},d)(y)\ .

Since RQj\frac{R}{Q_{j}} is reduced,

Fn((RQj)pm,(RQj)pm,I(RQj)pm,d)(y/pm)=Fn(RQj,RQj,IRQj,d)(y)F_{n}((\frac{R}{Q_{j}})^{p^{m}},(\frac{R}{Q_{j}})^{p^{m}},I(\frac{R}{Q_{j}})^{p^{m}},d)(y/p^{m})=F_{n}(\frac{R}{Q_{j}},\frac{R}{Q_{j}},I\frac{R}{Q_{j}},d)(y)

by Proposition 3.15, (3).

Using Proposition 3.15, assertion (1), we have

pmd[k:kpm]Fn+m(M,R,I,d)(y)=Fn(M,Rpm,IRpm,d)(y/pm).p^{md}[k:k^{p^{m}}]F_{n+m}(M,R,I,d)(y)=F_{n}(M,R^{p^{m}},IR^{p^{m}},d)(y/p^{m})\ .

Since for each jj, RQj\frac{R}{Q_{j}} has krull dimension dd,

λRpmQjRpm((M)QjRpm)=[k:kpm]pdmλRQj(MQj)\lambda_{R^{p^{m}}_{Q_{j}R^{p^{m}}}}((M)_{Q_{j}R^{p^{m}}})=[k:k^{p^{m}}]p^{dm}\lambda_{R_{Q_{j}}}(M_{Q_{j}})

So Equation 11 yields

|Fn+m(M,R,I,d)(y)j=1lλRQj(MQj)Fn(RQj,RQj,IRQj,d)(y)|Dpdm[k:kpm]1pn,|F_{n+m}(M,R,I,d)(y)-\sum\limits_{j=1}^{l}\lambda_{R_{Q_{j}}}(M_{Q_{j}})F_{n}(\frac{R}{Q_{j}},\frac{R}{Q_{j}},I\frac{R}{Q_{j}},d)(y)|\leq\frac{D}{p^{dm}[k:k^{p^{m}}]}\frac{1}{p^{n}},

proving assertion (1).

Proof of Theorem 3.1:.

Using Remark 3.7 we can assume that kpkk^{p}\subseteq k is a finite extension. We argue that the sequence (Fn(M,R,I,dim(M)))n(F_{n}(M,R,I,\text{dim}(M)))_{n} is uniformly Cauchy on every compact subset. By Theorem 3.16, assertion (1), we can assume that RR is a domain and M=RM=R. Fix a compact subset AA\subseteq\mathbb{C}. Since the torsion free rank of RR as an RpR^{p} module is pd[k:kp]p^{d}[k:k^{p}], we have an exact sequence of finitely generated graded RpR^{p} modules (see Lemma 3.18):

0{0}K{K}j=1pd[k:kp]Rp(hj){\bigoplus\limits_{j=1}^{p^{d}[k:k^{p}]}R^{p}(h_{j})}R{R}C{C}0{0}

for some integers hih_{i} such that both dimRp(K)anddimRp(C)\text{dim}_{R^{p}}(K)\,\text{and}\,\text{dim}_{R^{p}}(C) are less than dd. Hence there exist constants D,DD,D^{\prime} such that for all nn and for any yAy\in A,

|Fn+1(R,R,I)(y)Fn(R,R,I)(y)|=|1pd[k:kp]Fn(R,Rp,IRp)(y/p)Fn(R,R,I)(y)||1pd[k:kp]j=1pd[k:kp]Fn(Rp(hj),Rp,IRp)(y/p)Fn(R,R,I)(y)|+Dpn|Fn(Rp,Rp,IRp)(y/p)Fn(R,R,I)(y)|+Dpn+Dpn=D+Dpn.\begin{split}|F_{n+1}(R,R,I)(y)-F_{n}(R,R,I)(y)|&=|\frac{1}{p^{d}[k:k^{p}]}F_{n}(R,R^{p},IR^{p})(y/p)-F_{n}(R,R,I)(y)|\\ &\leq|\frac{1}{p^{d}[k:k^{p}]}\sum\limits_{j=1}^{p^{d}[k:k^{p}]}F_{n}(R^{p}(h_{j}),R^{p},IR^{p})(y/p)\\ &-F_{n}(R,R,I)(y)|+\frac{D}{p^{n}}\\ &\leq|F_{n}(R^{p},R^{p},IR^{p})(y/p)-F_{n}(R,R,I)(y)|+\frac{D^{\prime}}{p^{n}}+\frac{D}{p^{n}}\\ &=\frac{D+D^{\prime}}{p^{n}}\ .\end{split}

The first equality comes from assertion 1 of Proposition 3.15. The first inequality is a consequence of assertion (1) of Lemma 3.11. The second inequality is obtained by applying assertion (1) of Proposition 3.13 and assertion (1) of Proposition 3.10. The last equality follows from Proposition 3.15, assertion (3). Hence for m,nm,n\in\mathbb{N} and for any yAy\in A,

|Fn+m(R,R,I)(y)Fn(R,R,I)(y)|(D+D)(j=n1pj)=D+Dpnpp1.|F_{n+m}(R,R,I)(y)-F_{n}(R,R,I)(y)|\leq(D+D^{\prime})(\sum\limits_{j=n}^{\infty}\frac{1}{p^{j}})=\frac{D+D^{\prime}}{p^{n}}\frac{p}{p-1}.

Thus the sequence of entire functions (Fn(R,R,I)(y))n(F_{n}(R,R,I)(y))_{n} is uniformly Cauchy on AA.

A sequence of entire functions which is uniformly Cauchy on every compact subset of \mathbb{C} converges to a entire function and the convergence is uniform on every compact subset; see Theorem 1 in Chapter 5 of [Ahl79]. This finishes the proof of Theorem 3.1. ∎

4. Properties of Frobenius-Poincaré functions

This section is devoted to developing general properties of Frobenius-Poincaré functions. Some of these are analogues of properties of Hilbert-Kunz multiplicities. In Proposition 4.5 and Proposition 4.6, we use these general properties to compute Frobenius-Poincaré functions in some special cases.

Proposition 4.1.

Let MM be a finitely generated \mathbb{Z}-graded RR-module of Krull dimension dd. Then the power series expansion of F(M,R,I)(y)F(M,R,I)(y) around the origin in the complex plane is given by

F(M,R,I)(y)=m=0amym,F(M,R,I)(y)=\sum\limits_{m=0}^{\infty}a_{m}y^{m},

where for each mm,

am=(i)m1m!limn(1pn)d+mj=jmλ((MI[pn]M)j).a_{m}=(-i)^{m}\frac{1}{m!}\,\underset{n\to\infty}{\lim}(\frac{1}{p^{n}})^{d+m}\sum\limits_{j=-\infty}^{\infty}j^{m}\lambda{((\frac{M}{{I}^{[{p}^{n}]}M})_{j})}.
Proof.

Since the sequence (Fn(M))n(F_{n}(M))_{n} converges uniformly to F(M)F(M) on the closed unit disc around zero, it follows from Lemma 3, Chapter 4 of [Ahl79] that for each mm, the sequence

dmdym(Fn)(0)=(i)m(1pn)d+mj=jmλ((MI[pn]M)j)\frac{d^{m}}{dy^{m}}(F_{n})(0)=(-i)^{m}(\frac{1}{p^{n}})^{d+m}\sum\limits_{j=-\infty}^{\infty}j^{m}\lambda{((\frac{M}{{I}^{[{p}^{n}]}M})_{j})}

converges to dmdym(F)(0)\frac{d^{m}}{dy^{m}}(F)(0). Since am=1m!dmdym(F)(0)a_{m}=\frac{1}{m!}\frac{d^{m}}{dy^{m}}(F)(0), we get the result. ∎

Corollary 4.2.

The Hilbert-Kunz multiplicity of the triple (M,R,I)(M,R,I) is F(M,R,I)(0)F(M,R,I)(0).

The next result provides a associativity formula for Frobenius-Poincaré functions.

Theorem 4.3.

Let MM be a finitely generated \mathbb{Z}-graded RR-module of Krull dimension dd. Let P1,,PtP_{1},\ldots,P_{t} be the dimension dd minimal prime ideals in the support of MM. Then

F(M,R,I,d)(y)=j=1tλRPj(MPj)F(RPj,RPj,IRPj,d)(y).F(M,R,I,d)(y)=\sum\limits_{j=1}^{t}\lambda_{R_{P_{j}}}(M_{P_{j}})F(\frac{R}{P_{j}},\frac{R}{P_{j}},I\frac{R}{P_{j}},d)(y).
Proof.

Follows from Theorem 3.16. ∎

As a consequence of Theorem 4.3, we prove that Frobenius-Poincaré functions are additive over a short exact sequence.

Proposition 4.4.

Consider a short exact sequence of finitely generated \mathbb{Z}-graded RR-modules where the boundary maps preserve gradings,

0{0}M{M^{\prime}}M{M}M{M^{\prime\prime}}0.{0.}

Let dd be the Krull dimension of MM. Then F(M,R,I,d)=F(M,R,I,d)+F(M,R,I,d)F(M,R,I,d)=F(M^{\prime},R,I,d)+F(M^{\prime\prime},R,I,d).

Proof.

The support of MM is the union of supports of MM^{\prime} and MM^{\prime\prime}. Since for a dd dimensional minimal prime QQ in the support of MM, λRQ(MQ)=λRQ(MQ)+λRQ(MQ)\lambda_{R_{Q}}(M_{Q})=\lambda_{R_{Q}}(M^{\prime}_{Q})+\lambda_{R_{Q}}(M^{\prime\prime}_{Q}), the desired result follows from Theorem 4.3. ∎

In Proposition 4.5 we apply Theorem 4.3 to compute the Frobenius-Poincaré function with respect to an ideal generated by a homogeneous system of parameters.

Proposition 4.5.

Let RR be an \mathbb{N}-graded, Noetherian ring such that R0=kR_{0}=k. Let II be an ideal generated by a homogeneous system of parameters of degrees δ1,δ2,,δd\delta_{1},\delta_{2},\ldots,\delta_{d}. Denote the Hilbert-Samuel multiplicity of RR by eRe_{R}- see Definition 2.9. Then

F(R,I)(y)=eRj=1d(1eiδjyiy).F(R,I)(y)=e_{R}\prod\limits_{j=1}^{d}(\frac{1-e^{-i\delta_{j}y}}{iy}).
Proof.

Suppose f1,,fdf_{1},\ldots,f_{d} be a homogeneous system of parameters of degrees δ1,,δd\delta_{1},\ldots,\delta_{d} respectively, such that I=(f1,,fd)I=(f_{1},\ldots,f_{d}). Then the extension of rings k[f¯]:=k[f1,,fd]Rk[\underline{f}]:=k[f_{1},\ldots,f_{d}]\hookrightarrow R is finite (see Theorem 1.5.17, [BH98]). Suppose that the generic rank of RR as an k[f¯]k[\underline{f}] module is rr. Since k[f¯]k[\underline{f}] is isomorphic to the graded polynomial ring in dd variables where the degrees of the variables are δ1,,δd\delta_{1},\ldots,\delta_{d}, from Example 3.5 and Theorem 4.3, we have

(13) F(R,I)(y)=F(R,k[f¯],(f1,,fd)k[f¯])=rj=1d(1eiδjyiδjy).F(R,I)(y)=F(R,k[\underline{f}],(f_{1},\ldots,f_{d})k[\underline{f}])=r\prod\limits_{j=1}^{d}(\frac{1-e^{-i\delta_{j}y}}{i\delta_{j}y})\ .

Taking limit as yy tends to zero in (13) and proposition 4.1, we conclude that rr is the Hilbert-Kunz multiplicity of the pair (R,I)(R,I). The Hilbert-Kunz and the Hilbert-Samuel multiplicities are the same with respect to a given ideal generated by a system of parameters (see Theorem 11.2.10, [HS06]). So using Proposition 2.10 we get that r=δ1δjeRr=\delta_{1}\ldots\delta_{j}e_{R}. ∎

Now we compute the Frobenius-Poincaré function of a one dimensional graded domain whose degree zero piece is an algebraically closed field. This will indeed allow us to compute the Frobenius-Poincaré function of any one dimensional graded ring by using Remark 3.7 and Theorem 4.3.

Proposition 4.6.

Let RR be a one dimensional finitely generated \mathbb{N}-graded kk-algebra , where kk is algebraically closed and RR is a domain. Let II be a finite co-length homogeneous ideal. Let hh be the smallest integer such that II contains a non-zero homogeneous element of degree hh. Then

F(R,R,I)(y)=eR(1eihyiy),F(R,R,I)(y)=e_{R}(\frac{1-e^{-ihy}}{iy}),

where eRe_{R} is the Hilbert-Samuel multiplicity of RR (see Definition 2.9).

Proof.

Let SS be the normalization of RR. By Theorem 11, chapter VII, [ZS60], SS is an \mathbb{N}-graded RR-module and by Theorem 9, chapter 5, [ZS65] SS is finitely generated over kk. The generic rank of SS as an RR-module is one, hence F(S,R,I)=F(R,R,I)F(S,R,I)=F(R,R,I). Since kk is algebraically closed S0=kS_{0}=k. So by Lemma 3.14 F(S,R,I)F(S,R,I) is the same as F(S,S,IS)F(S,S,IS). So we compute F(S,S,IS)F(S,S,IS). Since SS is an \mathbb{N}-graded normal kk-algebra, by Theorem 1, section 3, Appendix III of [Ser00], SS is isomorphic to a graded polynomial ring in one variable. So the ideal ISIS is a homogeneous principal ideal. By our assumption, ISIS is generated by a degree hh homogeneous element ff; note that ff is a homogeneous system of parameter of SS. Thus by Proposition 4.5, F(S,S,IS)(y)=eS(1eihyiy)F(S,S,IS)(y)=e_{S}(\frac{1-e^{-ihy}}{iy}). Since SS has generic rank one as an RR module eR=eSe_{R}=e_{S}. ∎

Let SS be a ring containing a field of characteristic p>0p>0. Recall that the tight closure of an ideal JSJ\subseteq S is the ideal consisting of all xSx\in S such that there is a cSc\in S, not in any minimal prime of SS such that, cxpnJ[pn]cx^{p^{n}}\in{J}^{[{p}^{n}]} for all large nn –see Definition 3.1, [HH90]). The tight closure of JJ is denoted by JJ^{*}. The theory of Hilbert-Kunz multiplicity is related to the theory of tight closure: for a Noetherian local domain SS whose completion is also a domain and ideals J1J2J_{1}\subseteq J_{2}, the corresponding Hilbert-Kunz multiplicities are the same if and only if J1=J2J_{1}^{*}=J_{2}^{*} - see Proposition 5.4, Theorem 5.5 of [Hun13] and Theorem 8.17, [HH90]. A similar relation between tight closure of an ideal and the corresponding Frobenius-Poincaré function is the content of the next result.

Theorem 4.7.

Let RR is a finitely generated \mathbb{N}- graded kk-algebra. Let IJI\subseteq J be two finite colength homogeneous ideals of RR

  1. (1)

    If JJ is contained in II^{*}-the tight closure of II, F(R,R,I)=F(R,R,J)F(R,R,I)=F(R,R,J).

  2. (2)

    Suppose that all of the minimal primes of RR have the same dimension. If F(R,R,I)=F(R,R,J)F(R,R,I)=F(R,R,J), JIJ\subseteq I^{*}.

Proof.

1) Denote the Krull dimension of RR by dd. For (1), first we argue that there is a constant DD such that λ(J[pn]I[pn])\lambda(\frac{{J}^{[{p}^{n}]}}{{I}^{[{p}^{n}]}}) is bounded above by D(pn)d1D(p^{n})^{d-1} for all large n. Since JIJ\subseteq I^{*}, there exists a cRc\in R- not in any minimal primes of RR such that cJ[pn]I[pn]c{J}^{[{p}^{n}]}\subseteq{I}^{[{p}^{n}]}, for all large nn333cc can be chosen to be homogeneous. Pick a set of homogeneous generators of g1,g2,,grg_{1},g_{2},\ldots,g_{r} of JJ. Since the images of g1pn,,grpng_{1}^{p^{n}},\ldots,g_{r}^{p^{n}} generate J[pn]I[pn]\frac{{J}^{[{p}^{n}]}}{{I}^{[{p}^{n}]}}, we get a surjection for each nn:

j=1rR(c,I[pn])J[pn]I[pn].\bigoplus\limits_{j=1}^{r}\frac{R}{(c,{I}^{[{p}^{n}]})}\twoheadrightarrow\frac{{J}^{[{p}^{n}]}}{{I}^{[{p}^{n}]}}.

So the length of J[pn]I[pn]\frac{{J}^{[{p}^{n}]}}{{I}^{[{p}^{n}]}} is bounded above by rλ(R(c,I[pn]))r\lambda(\frac{R}{(c,{I}^{[{p}^{n}]})}). Since cc is not in any minimal prime of RR, dim(RcR)\text{dim}(\frac{R}{cR}) is at most d1d-1. The existence of the desired DD is apparent once we use Lemma 3.9 to bound the growth of λ(R(c,I[pn]))\lambda(\frac{R}{(c,{I}^{[{p}^{n}]})}).

Now, pick N0N_{0}\in\mathbb{N} such that for jN0pnj\geq N_{0}p^{n}, (RI[pn]R)j=0(\frac{R}{{I}^{[{p}^{n}]}R})_{j}=0 for all nn. Given yy\in\mathbb{C},

|Fn(R,R,I)(y)Fn(R,R,J)(y)|=(1pn)d|j=0λ((J[pn]I[pn])j)eiyj/pn|(1pn)dj=0|λ((J[pn]I[pn])j)||eiyj/pn|(1pn)dλ(J[pn]I[pn])eN0|y|.\begin{split}|F_{n}(R,R,I)(y)-F_{n}(R,R,J)(y)|&=(\frac{1}{p^{n}})^{d}|\sum\limits_{j=0}^{\infty}\lambda((\frac{{J}^{[{p}^{n}]}}{{I}^{[{p}^{n}]}})_{j})e^{-iyj/p^{n}}|\\ &\leq(\frac{1}{p^{n}})^{d}\sum\limits_{j=0}^{\infty}|\lambda((\frac{{J}^{[{p}^{n}]}}{{I}^{[{p}^{n}]}})_{j})||e^{-iyj/p^{n}}|\\ &\leq(\frac{1}{p^{n}})^{d}\lambda(\frac{{J}^{[{p}^{n}]}}{{I}^{[{p}^{n}]}})e^{N_{0}|y|}\ .\end{split}

To get the last inequality, we have used that for jN0pnj\leq N_{0}p^{n}, |eiyj/pn|eN0|y||e^{-iyj/p^{n}}|\leq e^{N_{0}|y|}. Since

limn(1pn)dλ(J[pn]I[pn])limn1pnD=0,\underset{n\to\infty}{\lim}(\frac{1}{p^{n}})^{d}\lambda(\frac{{J}^{[{p}^{n}]}}{{I}^{[{p}^{n}]}})\leq\underset{n\to\infty}{\lim}\frac{1}{p^{n}}D=0,

we get F(R,R,I)(y)=F(R,R,J)(y)F(R,R,I)(y)=F(R,R,J)(y).

(2) Let P1,,PtP_{1},\ldots,P_{t} be the minimal primes of RR. For a finite co-length homogeneous ideal 𝔞\mathfrak{a}, denote the Hilbert-Kunz multiplicity of the triple (R,R,𝔞)(R,R,\mathfrak{a}) (see Definition 2.1) by eHK(R,𝔞)e_{\text{HK}}(R,\mathfrak{a}). Since all the minimal primes of RR have the same dimension, evaluating the equality in Theorem 4.3 at y=0y=0 and using Corollary 4.2, we get

(14) eHK(R,𝔞)=j=1teHK(RPj,𝔞RPj).e_{HK}(R,\mathfrak{a})=\sum\limits_{j=1}^{t}e_{HK}(\frac{R}{P_{j}},\mathfrak{a}\frac{R}{P_{j}})\ .

Since for each jj, where 1jt1\leq j\leq t, eHK(RPj,IRPj)eHK(RPj,JRPj)e_{HK}(\frac{R}{P_{j}},I\frac{R}{P_{j}})\geq e_{HK}(\frac{R}{P_{j}},J\frac{R}{P_{j}}) and eHK(R,I)=F(R,R,I)(0)=F(R,R,J)(0)=eHK(R,J)e_{HK}(R,I)=F(R,R,I)(0)=F(R,R,J)(0)=e_{HK}(R,J), using (14), we conclude that for each minimal prime PjP_{j}, eHK(RPj,IRPj)=eHK(RPj,JRPj)e_{HK}(\frac{R}{P_{j}},I\frac{R}{P_{j}})=e_{HK}(\frac{R}{P_{j}},J\frac{R}{P_{j}}). From here we show that the tight closure of IRPjI\frac{R}{P_{j}} and JRPjJ\frac{R}{P_{j}} in RPj\frac{R}{P_{j}} are the same for any jj; this coupled with Theorem 1.3, (c) of [Hun96] establishes that I=JI^{*}=J^{*}. To this end, fix a minimal prime PjP_{j}. First note that RPj^\hat{\frac{R}{P_{j}}}: the completion of RPj\frac{R}{P_{j}} at the homogeneous maximal ideal is a domain. To see this, set InRPjI_{n}\subseteq\frac{R}{P_{j}} to be the ideal generated by forms of degree at least nn. Then the associated graded ring of RPj^\hat{\frac{R}{P_{j}}} with respect to the filtration (InRPj^)n(I_{n}\hat{\frac{R}{P_{j}}})_{n} is isomorphic to the domain RPj\frac{R}{P_{j}}- so by Theorem 4.5.8, [BH98] RPj^\hat{\frac{R}{P_{j}}} is a domain. Set mjm_{j} to be the maximal ideal of RPj\frac{R}{P_{j}}. Since RPj^\hat{\frac{R}{P_{j}}} is a domain, by Theorem 5.5, [Hun13] (I(RPj)mj)=(J(RPj)mj)(I(\frac{R}{P_{j}})_{m_{j}})^{*}=(J(\frac{R}{P_{j}})_{m_{j}})^{*}. Since both II and JJ are mjm_{j}-primary, by Theorem 1.5, [Hun96], we conclude that the tight closures of IRPjI\frac{R}{P_{j}} and JRPjJ\frac{R}{P_{j}} in RPj\frac{R}{P_{j}} are the same. ∎

Next, we set to show that over a standard graded ring, our Frobenius-Poincaré functions are holomorphic Fourier transforms of Hilbert-Kunz density functions introduced in [Tri18]. We first recall a part of a result in [Tri18] that implies the existence of Hilbert-Kunz density functions.

Theorem 4.8.

(see Theorem 1.1 and Theorem 2.19 [Tri18]) Let kk be a field of characteristic p>0p>0, RR be a standard graded kk-algebra of Krull dimension d1d\geq 1, II be a homogeneous ideal of finite co-length. Given a finitely generated \mathbb{N}-graded RR-module MM, consider the sequence (gn)n(g_{n})_{n} of real valued functions defined on the real line where

(15) gn(x)=(1pn)d1λk((MI[pn]M)xpn).g_{n}(x)=(\frac{1}{p^{n}})^{d-1}\lambda_{k}((\frac{M}{{I}^{[{p}^{n}]}M})_{\lfloor xp^{n}\rfloor}).

Then

  1. (1)

    There is a compact subset of the non-negative real line containing the support of gng_{n} for all nn.

  2. (2)

    The sequence (gn)(g_{n}) converges pointwise to a compactly supported function gg. Furthermore, when d2d\geq 2, the convergence is uniform and gg is continuous.

The function gg in Theorem 4.8 is called the Hilbert-Kunz density function associated to the triple (M,R,I)(M,R,I).

Recall that the holomorphic Fourier transform of a compactly supported Lebesgue integrable function hh defined on the real line is the holomorphic function h^\hat{h} given by

h^(y)=h(x)eiyxdx,\hat{h}(y)=\underset{\mathbb{R}}{\int}h(x)e^{-iyx}dx,

where the integral is a Lebesgue integral (see Chapter 2, [Rud87]).

Proposition 4.9.

The holomorphic Fourier transform of the Hilbert-Kunz density function associated to a triple (M,R,I)(M,R,I) as in Theorem 4.8 is the Frobenius-Poincaré function F(M,R,I,d)F(M,R,I,d).

Proof.

Let gng_{n} and gg be as in Theorem 4.8. We first establish the claim that there is a constant CC, such that for any real number xx and all nn, gn(x)Cg_{n}(x)\leq C. We can assume that there is compact subset [0,N][0,N] containing the support of gng_{n} for all nn (see (1), Theorem 4.8). Now given xx where 1pnxN\frac{1}{p^{n}}\leq x\leq N,

gn(x)(1pn)d1λ(Mxpn)=(xpnpn)d1λ(Mxpn)(xpn)d1Nd1λ(Mxpn)(xpn)d1.g_{n}(x)\leq(\frac{1}{p^{n}})^{d-1}\lambda(M_{\lfloor xp^{n}\rfloor})=(\frac{{\lfloor xp^{n}\rfloor}}{p^{n}})^{d-1}\frac{\lambda(M_{\lfloor xp^{n}\rfloor})}{({\lfloor xp^{n}\rfloor})^{d-1}}\leq N^{d-1}\frac{\lambda(M_{\lfloor xp^{n}\rfloor})}{({\lfloor xp^{n}\rfloor})^{d-1}}.

Since the function λ(Mm)md1\frac{\lambda(M_{m})}{m^{d-1}} is bounded above by a constant (see Proposition 4.4.1 and Exercise 4.4.11 of [BH98]), the claim follows.

The bound on gng_{n} allows us to use dominated convergence theorem to the sequence (gn)n(g_{n})_{n}, which implies that the sequence of functions (gn^)n(\hat{g_{n}})_{n} converges to g^\hat{g} pointwise. Now we claim that the sequence (gn^)(\hat{g_{n}}) in fact converges to F(M,R,I,d)F(M,R,I,d) pointwise; this would imply g^=F(M,R,I,d)\hat{g}=F(M,R,I,d).
Now for a non-zero complex number yy,

(16) gn^(y)=(1pn)d10λk((MI[pn]M)xpn)eiyxdx=(1pn)d1j=0jpnj+1pnλk((MI[pn]M)j)eiyxdx=(1pn)d1j=0λk((MI[pn]M)j)(eiy(j+1)/pneiyj/pniy)=(1pn)d1j=0λk((MI[pn]M)j)eiyj/pn(eiy/pn1iy)=(1pn)dj=0λk((MI[pn]M)j)eiyj/pn(eiy/pn1iy/pn).\begin{split}\hat{g_{n}}(y)&=(\frac{1}{p^{n}})^{d-1}\int_{0}^{\infty}\lambda_{k}((\frac{M}{{I}^{[{p}^{n}]}M})_{\lfloor xp^{n}\rfloor})e^{-iyx}dx\\ &=(\frac{1}{p^{n}})^{d-1}\sum\limits_{j=0}^{\infty}\int_{\frac{j}{p^{n}}}^{\frac{j+1}{p^{n}}}\lambda_{k}((\frac{M}{{I}^{[{p}^{n}]}M})_{j})e^{-iyx}dx\\ &=(\frac{1}{p^{n}})^{d-1}\sum\limits_{j=0}^{\infty}\lambda_{k}((\frac{M}{{I}^{[{p}^{n}]}M})_{j})\left(\frac{e^{-iy(j+1)/p^{n}}-e^{-iyj/p^{n}}}{-iy}\right)\\ &=(\frac{1}{p^{n}})^{d-1}\sum\limits_{j=0}^{\infty}\lambda_{k}((\frac{M}{{I}^{[{p}^{n}]}M})_{j})e^{-iyj/p^{n}}\left(\frac{e^{-iy/p^{n}}-1}{-iy}\right)\\ &=(\frac{1}{p^{n}})^{d}\sum\limits_{j=0}^{\infty}\lambda_{k}((\frac{M}{{I}^{[{p}^{n}]}M})_{j})e^{-iyj/p^{n}}\left(\frac{e^{-iy/p^{n}}-1}{-iy/p^{n}}\right)\ .\end{split}

So using the last line of (16) and Lemma 3.3, we get that for a non-zero complex number yy,

g^(y)=limngn^(y)=limn(1pn)dj=0λk((MI[pn]M)j)eiyj/pn=F(M,R,I,d)(y).\hat{g}(y)=\underset{n\to\infty}{\lim}\hat{g_{n}}(y)=\underset{n\to\infty}{\lim}(\frac{1}{p^{n}})^{d}\sum\limits_{j=0}^{\infty}\lambda_{k}((\frac{M}{{I}^{[{p}^{n}]}M})_{j})e^{-iyj/p^{n}}=F(M,R,I,d)(y).

Note that,

(17) gn^(0)=(1pn)dj=0λ((MI[pn]M)j)=(1pn)dλ(MI[pn]M)=Fn(M,d)(0).\hat{g_{n}}(0)=(\frac{1}{p^{n}})^{d}\sum\limits_{j=0}^{\infty}\lambda((\frac{M}{{I}^{[{p}^{n}]}M})_{j})=(\frac{1}{p^{n}})^{d}\lambda(\frac{M}{{I}^{[{p}^{n}]}M})=F_{n}(M,d)(0)\ .

Taking limit as nn approaches infinity in (17) gives g^(0)=F(M,R,I,d)(0).\hat{g}(0)=F(M,R,I,d)(0).

Remark 4.10.
  1. (1)

    Since a compactly supported continuous function can be recovered from its holomorphic Fourier transform (see Theorem 1.7.3, [Hor65]), the existence of Frobenius-Poincaré functions gives an alternate proof of the existence of Hilbert-Kunz density functions in dimension d2d\geq 2.

  2. (2)

    One way to incorporate zero dimensional ambient rings into the theory of Hilbert-Kunz density functions could be to realize the functions gng_{n} in (15) and the resulting Hilbert-Kunz density function as compactly supported distributions (see Definition 1.3.2, [Hor65]). Here by a distribution, we mean a \mathbb{C}-linear map from the space of complex valued smooth functions on \mathbb{R} to \mathbb{C}. In our case, the distribution defined by each gng_{n} sends the function ff to f(x)gn(x)dx\underset{\mathbb{R}}{\int}f(x)g_{n}(x)dx. When the ambient ring has dimension at least one, the sequence of distributions defined (gn)n(g_{n})_{n} converges to the distribution defined by the corresponding Hilbert-Kunz density function; see the Remark on page 7 of [Hor65] for a precise meaning of convergence of distributions. Now suppose that RR has dimension zero and MM is a finitely generated \mathbb{Z}-graded RR-module; let (gn)n(g_{n})_{n} be the corresponding sequence of functions given by (15) with d=0d=0. Direct calculation shows that for a complex valued smooth function ff, the sequence of numbers f(x)gn(x)dx\underset{\mathbb{R}}{\int}f(x)g_{n}(x)dx converges to λk(M)f(0)\lambda_{k}(M)f(0). This means that the sequence of distributions defined by (gn)n(g_{n})_{n} converges to the distribution λk(M)δ0\lambda_{k}(M)\delta_{0}- where δ0\delta_{0} is the distribution such that δ0(f)=f(0)\delta_{0}(f)=f(0). So it is reasonable to define the Hilbert-Kunz density function g(M,R,I)g(M,R,I) to be the distribution λk(M)δ0\lambda_{k}(M)\delta_{0}. In fact, incorporating the language of Fourier transform of distributions (see section 1.7, [Hor65]), it follows that the Fourier transform of the Hilbert-Kunz density function (or distribution) is our Frobenius-Poincaré function irrespective of the dimension of the ambient ring. Going in the reverse direction, Hilbert-Kunz density function of a triple can be defined to be the unique compactly supported distribution whose Fourier transform is the corresponding Frobenius-Poincaré function.

5. Descriptions using Homological Information

In this section, we give alternate descriptions of Frobenius-Poincaré functions of (R,R,I)(R,R,I) in terms of the sequence of graded Betti numbers of R/I[pn]R/{I}^{[{p}^{n}]}. Moreover when RR is Cohen-Macaulay, the Frobenius-Poincaré functions are described using the Koszul homologies of RI[pn]\frac{R}{{I}^{[{p}^{n}]}} with respect to a homogeneous system of parameters of R. Some background material for this section on Hilbert-Samuel multiplicity, Hilbert series and graded Betti numbers is reviewed in Section 2.2 and Section 2.3.

Theorem 5.1.

Let SS be a graded complete intersection over kk of Krull dimension dd and Hilbert-Samuel multiplicity eSe_{S} (see Definition 2.9). Let SRS\rightarrow R be a module finite kk-algebra map to a finitely generated \mathbb{N}-graded kk-algebra. Let IRI\subseteq R be a homogeneous ideal of finite co-length and MM be a finitely generated \mathbb{Z}-graded RR-module. Set

(18) 𝔹S(j,n)=α=0(1)αλ((TorαS(k,M/I[pn]M)j).\mathbb{B}^{S}(j,n)=\sum_{\alpha=0}^{\infty}(-1)^{\alpha}\,\lambda((\text{Tor}_{\alpha}^{S}(k,M/{I}^{[{p}^{n}]}M)_{j}).

Then

  1. (1)

    limn(pn)ddimMj=0𝔹S(j,n)eiyj/pn{\underset{n\to\infty}{\lim}(p^{n})^{d-\text{dim}\,M}\sum\limits_{j=0}^{\infty}\,\mathbb{B}^{S}(j,n)\,e^{-iyj/p^{n}}} admits an analytic extension to the complex plane.

  2. (2)

    The Frobenius-Poincaré function F(M,R,I)(y)F(M,R,I)(y) is the same as the analytic extension of the function
    eS(iy)dlimn(pn)ddimMj=0𝔹S(j,n)eiyj/pn\frac{e_{S}}{(iy)^{d}}\,\,{\underset{n\to\infty}{\lim}(p^{n})^{d-\text{dim}\,M}\sum\limits_{j=0}^{\infty}\,\mathbb{B}^{S}(j,n)\,e^{-iyj/p^{n}}} to the complex plane.

Note that for fixed integers j,nj,n the sum in (18) is finite- see Lemma 2.5. We record some remarks and consequences related to Theorem 5.1 before proving the result.

Corollary 5.2.

For a graded complete intersection RR over kk and a homogeneous ideal IRI\subseteq R of finite colength, the function limnj=0𝔹R(j,n)eiyj/pn\underset{n\to\infty}{\lim}\sum\limits_{j=0}^{\infty}\mathbb{B}^{R}(j,n)e^{-iyj/p^{n}} extends to the entire function 1eR(iy)dim(R)F(R,R,I)(y)\frac{1}{e_{R}}(iy)^{\text{dim}(R)}F(R,R,I)(y).

Remark 5.3.

When applied to the triple (R,R,m)(R,R,m), where mm is the homogeneous maximal ideal of a graded complete intersection RR over kk, Theorem 5.1 applied to the case S=R=MS=R=M, gives a way to compare the Hilbert-Kunz multiplicity of (R,m)(R,m) to the Hilbert-Samuel multiplicity eRe_{R}.

Remark 5.4.

One way to apply Theorem 5.1 to describe the Frobenius-Poincaré function of a graded triple (M,R,I)(M,R,I) is to take SS to be a subring of RR generated by a homogeneous system of parameters. Since such an SS is regular, for any integer nn, the sum defining 𝔹S(j,n)\mathbb{B}^{S}(j,n) in (18) is finite and the function j=0𝔹S(j,n)eiyj/pn\sum\limits_{j=0}^{\infty}\,\mathbb{B}^{S}(j,n)\,e^{-iyj/p^{n}} appearing in Theorem 5.1 is a polynomial in eiy/pne^{-iy/p^{n}}.

Remark 5.5.

In [Tri21, page 7] V. Trivedi asks whether the Hilbert-Kunz density function (see Theorem 4.8) of a d(2)d(\geq 2) dimensional standard graded pair (R,I)(R,I) is always (d2)(d-2) times differentiable and the (d2)(d-2)-th order derivative is continuous. We use Theorem 5.1 to reformulate Trivedi’s question and produce the candidate function for the (d2)(d-2)-th order derivative. Denote the restriction of (iy)d2F(R,R,I)(y)(iy)^{d-2}F(R,R,I)(y) to the real line by hh. The Fourier transform of the temperate distribution (see Definition 1.7.2, 1.7.3, [Hor65]) defined by hh determines the (d2)(d-2)-th order derivative of the distribution (see Definition 1.4.1, [Hor65]) defined by the Hilbert-Kunz density function of (R,I)(R,I)- see Remark 4.10, Theorem 1.7.3, [Hor65]. In fact using tools from analysis, one can show that Trivedi’s question has an affirmative answer if the integral |h(x)|dx\underset{\mathbb{R}}{\int}|h(x)|dx is finite444The argument is recorded in [Muk23, Theorem 8.3.8].. When |h(x)|dx\underset{\mathbb{R}}{\int}|h(x)|dx is finite, the Fourier transform of hh is in fact given by the actual function h^(y)=h(x)eiyxdx\hat{h}(y)=\underset{\mathbb{R}}{\int}h(x)e^{-iyx}dx for yy\in\mathbb{R} and the (d2)(d-2)-th order derivative is the function 12πh^(y)\frac{1}{2\pi}\hat{h}(-y). Fix a subring SRS\subseteq R generated by a homogeneous system of parameters of RR. Since by Theorem 5.1 applied to the case M=RM=R (also see Remark 5.4) h(y)=eSlimnj=0𝔹S(j,n)eiyj/pn(iy)2h(y)=e_{S}\underset{n\to\infty}{\lim}\frac{\sum\limits_{j=0}^{\infty}\mathbb{B}^{S}(j,n)e^{-iyj/p^{n}}}{(iy)^{2}}, it is natural to ask

Question 5.6.
  1. (1)

    Is the function h(y)=limnj=0𝔹S(j,n)eiyj/pn(iy)2h(y)=\frac{\underset{n\to\infty}{\lim}\sum\limits_{j=0}^{\infty}\mathbb{B}^{S}(j,n)e^{-iyj/p^{n}}}{(iy)^{2}} integrable on \mathbb{R}?

  2. (2)

    Is function limnj=0𝔹S(j,n)eiyj/pn\underset{n\to\infty}{\lim}\sum\limits_{j=0}^{\infty}\mathbb{B}^{S}(j,n)e^{-iyj/p^{n}} restricted to the real line bounded?

Note that an affirmative answer to part (2) implies an affirmative answer to part (1) of the 5.6.

We use a consequence of a result from [AB93]-where it is cited as a folklore- in the proof of Theorem 5.1 below.

Proposition 5.7.

(see Lemma 7.ii, [AB93]) Let RR be a finitely generated \mathbb{N}-graded kk-algebra and M,NM,N be two finitely generated \mathbb{Z}-graded RR-modules. Denote the formal Laurent series i(1)iHToriR(M,N)(t)\underset{i\in\mathbb{N}}{\sum}(-1)^{i}H_{\text{Tor}_{i}^{R}(M,N)}(t) by χR(M,N)(t)\chi^{R}(M,N)(t). Then

χR(M,N)(t)=HM(t)HN(t)HR(t),\chi^{R}(M,N)(t)=\frac{H_{M}(t)H_{N}(t)}{H_{R}(t)},

where for a finitely generated \mathbb{Z}-graded RR-module NN^{\prime}, HN(t)H_{N^{\prime}}(t) is the Hilbert series of NN^{\prime}.

Proof of Theorem 5.1:.

Let \mathfrak{H} be the set of complex numbers with a negative imaginary part. We shall prove that on the connected open subset \mathfrak{H} of the complex plane

eS(iy)dlimn(pn)ddimMj=0𝔹S(j,n)eiyj/pn\frac{e_{S}}{(iy)^{d}}\,\,{\underset{n\to\infty}{\lim}(p^{n})^{d-\text{dim}\,M}\sum\limits_{j=0}^{\infty}\,\mathbb{B}^{S}(j,n)\,e^{-iyj/p^{n}}}

defines a holomorphic function and is the same as the restriction of F(R,R,I)F(R,R,I) to \mathfrak{H}. Since F(R,R,I)F(R,R,I) is an entire function, the analytic continuity in assertion 1 and the desired equality in assertion 2 follows.

Given an integer nn, χS(MI[pn]M,k)(t)=j=𝔹S(j,n)tj\chi^{S}(\frac{M}{{I}^{[{p}^{n}]}M},k)(t)=\sum\limits_{j=-\infty}^{\infty}\mathbb{B}^{S}(j,n)t^{j}. So using Proposition 5.7 we get

(19) HMI[pn]M(t)=HS(t)(j=𝔹S(j,n)tj).H_{\frac{M}{{I}^{[{p}^{n}]}M}}(t)=H_{S}(t)(\sum\limits_{j=-\infty}^{\infty}\mathbb{B}^{S}(j,n)t^{j}).

Now for any yy\in\mathfrak{H}, |eiy/pn|<1|e^{-iy/p^{n}}|<1; so by Lemma 2.6, the series j𝔹S(j,n)(eiy/pn)j\underset{j\in\mathbb{Z}}{\sum}\mathbb{B}^{S}(j,n)(e^{-iy/p^{n}})^{j} converges absolutely. For yy\in\mathfrak{H}, plugging in t=eiy/pnt=e^{-iy/p^{n}} in (19), we get

(20) (1pn)dim(M)HMI[pn]M(eiy/pn)=(1pn)dHS(eiy/pn)(pn)ddim(M)(j=𝔹S(j,n)eiyj/pn)=(1eiy/pn)d(pn(1eiy/pn))dHS(eiy/pn)(pn)ddim(M)(j=𝔹S(j,n)eiyj/pn).\begin{split}(\frac{1}{p^{n}})^{\text{dim}(M)}H_{\frac{M}{{I}^{[{p}^{n}]}M}}(e^{-iy/p^{n}})&=(\frac{1}{p^{n}})^{d}H_{S}(e^{-iy/p^{n}})(p^{n})^{d-\text{dim}(M)}(\sum\limits_{j=-\infty}^{\infty}\mathbb{B}^{S}(j,n)e^{-iyj/p^{n}})\\ &=\frac{(1-e^{-iy/p^{n}})^{d}}{(p^{n}(1-e^{-iy/p^{n}}))^{d}}H_{S}(e^{-iy/p^{n}})(p^{n})^{d-\text{dim}(M)}(\sum\limits_{j=-\infty}^{\infty}\mathbb{B}^{S}(j,n)e^{-iyj/p^{n}})\ .\end{split}

For a fixed yy\in\mathfrak{H}, as nn approaches infinity, (1eiy/pn)dHS(eiy/pn)(1-e^{-iy/p^{n}})^{d}H_{S}(e^{-iy/p^{n}}) approaches eSe_{S} (see Proposition 2.8) and (pn(1eiy/pn))d(p^{n}(1-e^{-iy/p^{n}}))^{d} approaches (iy)d(iy)^{d}(see Lemma 3.3). Now taking limit as nn approaches infinity in (20) gives the following equality on \mathfrak{H}:

F(R,R,I)(y)=eS(iy)dlimn(pn)ddimMj=0𝔹S(j,n)eiyj/pn.F(R,R,I)(y)=\frac{e_{S}}{(iy)^{d}}\,\,{\underset{n\to\infty}{\lim}(p^{n})^{d-\text{dim}\,M}\sum\limits_{j=0}^{\infty}\,\mathbb{B}^{S}(j,n)\,e^{-iyj/p^{n}}}.

Since the left hand side of the last equation is holomorphic on \mathfrak{H}, so is the right hand side; this finishes the proof. ∎

Remark 5.8.

Take S=R=MS=R=M in Theorem 5.1 and let \mathfrak{H} be the same as in the proof of Theorem 5.1. Although the analyticity of j=0𝔹(j,n)eiyj/pn\sum\limits_{j=0}^{\infty}\mathbb{B}(j,n)e^{-iyj/p^{n}} on \mathfrak{H}, for each nn, follows from Lemma 2.6, the existence of the analytic extension of their limit crucially depends on Theorem 5.1 and that Frobenius-Poincaré functions are entire.

When the RR-module R/IR/I has finite projective dimension, the line of argument in Theorem 5.1 (also see [TW22]) allows to describe F(R,R,I)F(R,R,I) in terms of the graded Betti numbers of R/IR/I.

Proposition 5.9.

Let II be a homogeneous ideal of the dd dimensional ring RR, such that the projective dimension of the RR-module R/IR/I is finite. Set

bα,j=λ(TorαR(k,R/I)j),𝔹(j)=α=0(1)αbα,j,eR=Hilbert-Samuel multiplicity ofR.b_{\alpha,j}=\lambda(\text{Tor}_{\alpha}^{R}(k,R/I)_{j}),\hskip 7.11317pt\mathbb{B}(j)=\sum\limits_{\alpha=0}^{\infty}(-1)^{\alpha}b_{\alpha,j},\hskip 7.11317pte_{R}=\text{Hilbert-Samuel multiplicity of}\,R.

Let bb be the smallest integer such that 𝔹(j)=0\mathbb{B}(j)=0 for all j>bj>b. Then for a non-zero complex number yy, we have:

F(R,R,I)(y)=eRj=0b𝔹(j)eiyj(iy)d.F(R,R,I)(y)=e_{R}\ \frac{\sum\limits_{j=0}^{b}\mathbb{B}(j)e^{-iyj}}{(iy)^{d}}.
Proof.

Take a minimal graded free resolution of the RR-module R/IR/I:

0{0}jR(j)bd,j{\underset{j\in\mathbb{N}}{\oplus}R(-j)^{\oplus b_{d,j}}}{\ldots}jR(j)b1,j{\underset{j\in\mathbb{N}}{\oplus}R(-j)^{\oplus b_{1,j}}}jR(j)b0,j{\underset{j\in\mathbb{N}}{\oplus}R(-j)^{\oplus b_{0,j}}}R/I{R/I}0.{0.}

Then we get a minimal graded free resolution of R/I[pn]R/{I}^{[{p}^{n}]} by applying nn-th iteration of the Frobenius functor to the chosen minimal graded resolution of R/IR/I (see Theorem 1.13 of [PS73]),

0{0}jR(pnj)bd,j{\underset{j\in\mathbb{N}}{\oplus}R(-p^{n}j)^{\oplus b_{d,j}}}{\ldots}jR(pnj)b0,j{\underset{j\in\mathbb{N}}{\oplus}R(-p^{n}j)^{\oplus b_{0,j}}}R/I[pn]{{R/{I}^{[{p}^{n}]}}}0.{0.}

So using the notation set in (18) in the case S=R=MS=R=M and the ideal II, we have that for any positive integer nn, 𝔹R(jpn,n)=𝔹(j)\mathbb{B}^{R}(jp^{n},n)=\mathbb{B}(j) and 𝔹(m,n)=0\mathbb{B}(m,n)=0 if pnp^{n} does not divide mm. So for all nn\in\mathbb{N}, χR(RI[pn],k)(t)=j𝔹R(j,n)tj=j=0b𝔹(j)tjpn\chi^{R}(\frac{R}{{I}^{[{p}^{n}]}},k)(t)=\underset{j\in\mathbb{N}}{\sum}\mathbb{B}^{R}(j,n)t^{j}=\sum\limits_{j=0}^{b}\mathbb{B}(j)t^{jp^{n}} is a polynomial in tt. So for any nn\in\mathbb{N}, using Proposition 5.7 for M=R/I[pn]M=R/{I}^{[{p}^{n}]}, N=kN=k we have for all yy\in\mathbb{C},

(21) (1pn)dHRI[pn](eiy/pn)=(1eiy/pn)dHR(eiy/pn)(pn(1eiy/pn))d(j=0b𝔹(j)eiyj).(\frac{1}{p^{n}})^{d}H_{\frac{R}{{I}^{[{p}^{n}]}}}(e^{-iy/p^{n}})=\frac{(1-e^{-iy/p^{n}})^{d}H_{R}(e^{-iy/p^{n}})}{(p^{n}(1-e^{-iy/p^{n}}))^{d}}(\sum\limits_{j=0}^{b}\mathbb{B}(j)e^{-iyj}).

Now taking limit as nn approaches infinity in (21) and using Proposition 2.8 and Lemma 3.3, we get

F(R,R,I)(y)=eRj=0b𝔹(j)eiyj(iy)d.F(R,R,I)(y)=e_{R}\frac{\sum\limits_{j=0}^{b}\mathbb{B}(j)e^{-iyj}}{(iy)^{d}}.

Example 5.10.

Let R=k[X,Y]R=k[X,Y] be the standard graded polynomial ring in two variables, I=(f,g)RI=(f,g)R where ff and gg have degree d1d_{1} and d2d_{2} respectively. Using Proposition 5.9, we can compute F(R,I)(y)F(R,I)(y). A minimal free resolution of R/IR/I is given by the Koszul complex of (f,g)(f,g):

0R(d1d2)R(d1)R(d2)R0.0\rightarrow R(-d_{1}-d_{2})\rightarrow R(-d_{1})\oplus R(-d_{2})\rightarrow R\rightarrow 0.

Hence we get,

(22) F(R,I)(y)=𝔹(0)+𝔹(d1)eiyd1+𝔹(d2)eiyd2+𝔹(d1+d2)eiy(d1+d2)(iy)2=1eiyd1eiyd2+eiy(d1+d2)(iy)2.\begin{split}F(R,I)(y)&=\frac{\mathbb{B}(0)+\mathbb{B}(d_{1})e^{-iyd_{1}}+\mathbb{B}(d_{2})e^{-iyd_{2}}+\mathbb{B}(d_{1}+d_{2})e^{-iy(d_{1}+d_{2})}}{(iy)^{2}}\\ &=\frac{1-e^{-iyd_{1}}-e^{-iyd_{2}}+e^{-iy(d_{1}+d_{2})}}{(iy)^{2}}.\end{split}

The Hilbert-Kunz multiplicity eHK(R,I)=d1d2e_{HK}(R,I)=d_{1}d_{2} (see for example Theorem 11.2.10 of [HS06]). Using this observation, we can construct finite co-length ideals II and JJ in RR such that, eHK(R,I)=eHK(R,J)e_{HK}(R,I)=e_{HK}(R,J) but F(R,R,I)F(R,R,I) and F(R,R,J)F(R,R,J) are different.

In the next result, we show that the Frobenius-Poincaré function of a Cohen-Macaulay ring can be described in terms of the sequence of Koszul homologies of RI[pn]\frac{R}{{I}^{[{p}^{n}]}} with respect to a homogeneous system of parameters.

Theorem 5.11.

Let RR be a Cohen-Macaulay \mathbb{N}-graded ring of dimension dd, II be a homogeneous ideal of finite co-length of RR. Let x1,x2,,xdx_{1},x_{2},\ldots,x_{d} be a homogeneous system of parameters of RR of degree δ1,,δd\delta_{1},\ldots,\delta_{d} respectively. Then

F(R,I)(y)=1δ1δ2δd(iy)dlimnχR(R(x1,,xd)R,RI[pn]R)(eiy/pn),F(R,I)(y)=\frac{1}{\delta_{1}\delta_{2}\ldots\delta_{d}(iy)^{d}}\lim_{n\to\infty}\chi^{R}(\frac{R}{(x_{1},\ldots,x_{d})R},\frac{R}{{I}^{[{p}^{n}]}R})(e^{-iy/p^{n}}),

where χR(R(x1,,xd)R,RI[pn]R)(t)\chi^{R}(\frac{R}{(x_{1},\ldots,x_{d})R},\frac{R}{{I}^{[{p}^{n}]}R})(t) has the same meaning as in Proposition 5.7.

Remark 5.12.

The Laurent series χR(M,N)(t)\chi^{R}(M,N)(t) for a pair of graded modules as defined in Proposition 5.7 has been used before to define multiplicity or intersection multiplicity in different contexts; see for example [Ser00] Chapter IV, A, Theorem 1 and [Erm17]. The assertion in Theorem 5.11 should be thought of as an analogue of the these results since here the Frobenius-Poincaré function and hence the Hilbert-Kunz multiplicity is expressed in terms of the limit of power series χR(RI[pn],R(x¯)R)(t)\chi^{R}(\frac{R}{{I}^{[{p}^{n}]}},\frac{R}{(\underline{x})R})(t).

Proof of Theorem 5.11:.

Using Proposition 5.7 we have,

(23) HRI[pn](eiy/pn)HR(x1,,xd)(eiy/pn)=HR(eiy/pn)χR(R(x1,,xd)R,RI[pn]R)(eiy/pn).H_{\frac{R}{{I}^{[{p}^{n}]}}}(e^{-iy/p^{n}})H_{\frac{R}{(x_{1},\ldots,x_{d})}}(e^{-iy/p^{n}})=H_{R}(e^{-iy/p^{n}})\chi^{R}(\frac{R}{(x_{1},\ldots,x_{d})R},\frac{R}{{I}^{[{p}^{n}]}R})(e^{-iy/p^{n}})\ .

Since RR is Cohen-Macaulay, x1,,xdx_{1},\ldots,x_{d} is a regular sequence. Inducing on dd, one can show that,

(24) HR(x1,,xd)(t)=(1tδ1)(1tδ2)(1tδd)HR(t).H_{\frac{R}{(x_{1},\ldots,x_{d})}}(t)=(1-t^{\delta_{1}})(1-t^{\delta_{2}})\ldots(1-t^{\delta_{d}})H_{R}(t)\ .

Using (24) in (23) we get

(1pn)dHR/I[pn](eiy/pn)=χR(R(x1,,xd)R,RI[pn]R)(eiy/pn)(pn)d(1eiyδ1/pn)(1eiyδd/pn).(\frac{1}{p^{n}})^{d}H_{R/{I}^{[{p}^{n}]}}(e^{-iy/p^{n}})=\frac{\chi^{R}(\frac{R}{(x_{1},\ldots,x_{d})R},\frac{R}{{I}^{[{p}^{n}]}R})(e^{-iy/p^{n}})}{(p^{n})^{d}(1-e^{-iy\delta_{1}/p^{n}})\ldots(1-e^{-iy\delta_{d}/p^{n}})}.

The desired assertion now follows from taking limit as nn approaching infinity in the last equation and using Lemma 3.3. ∎

In each of Theorem 5.1, Proposition 5.9, Theorem 5.11, the Frobenius-Poincaré function is described as a quotient: the denominator is a power of iyiy and the numerator is a limit of a sequence of power series or polynomials in eiy/pne^{-iy/p^{n}}. In particular, in Theorem 5.11, the maximum value of j/pnj/p^{n} that appears in eiyj/pne^{-iyj/p^{n}} in the sequence of functions is bounded above by a constant independent of nn- see Lemma 3.8. So we ask

Question 5.13.

Let RR be a Cohen-Macaulay \mathbb{N}-graded ring of dimension dd, II is a homogeneous ideal of finite co-length. Does there exist a real number rr and a polynomial Q[X]Q\in\mathbb{R}[X] such that

F(R,R,I)(y)=Q(eiry)(iy)d?F(R,R,I)(y)=\frac{Q(e^{-iry})}{(iy)^{d}}?

6. Frobenius-Poincaré functions in dimension two

We compute the Frobenius-Poincaré function of two dimensional graded rings following the work of [Bre07] and [Tri05]. In this section, RR stands for a normal, two dimensional, standard graded domain. We assume that R0=kR_{0}=k is an algebraically closed field of prime characteristic pp. The smooth embedded curve Proj(R)\text{Proj}(R) is denoted by CC, δR\delta_{R} stands for the Hilbert-Samuel multiplicity of RR; alternatively δR\delta_{R} is the degree of the line bundle 𝒪C(1)\mathcal{O}_{C}(1). The genus of CC is denoted by gg. For a sheaf of 𝒪C\mathcal{O}_{C}-modules \mathcal{F}, (j)\mathcal{F}(j) stands for the sheaf 𝒪C(j)\mathcal{F}\otimes\mathcal{O}_{C}(j). The absolute Frobenius endomorphism of CC is denoted by ff. Some background materials for this section are reviewed in Section 2.4.

Theorem 6.1.

With notation as in the paragraph above, let II be an ideal of finite colength in RR generated by degree one elements h1,h2,,hrh_{1},h_{2},\ldots,h_{r}. Consider the short exact sequence of vector bundles on C=Proj(R)C=Proj(R)

(25) 0𝒮𝑟𝒪C(h1,,hr)𝒪C(1)0.0\longrightarrow\mathcal{S}\longrightarrow\overset{r}{\bigoplus}\mathcal{O}_{C}\xrightarrow{(h_{1},\ldots,h_{r})}\mathcal{O}_{C}(1)\longrightarrow 0\ .

Choose n0n_{0} such that the Harder-Narasimhan filtration on fn0(𝒮)f^{n_{0}*}(\mathcal{S}) given by

(26) 0=01tt+1=fn0𝒮0=\mathcal{E}_{0}\subset\mathcal{E}_{1}\subset\ldots\subset\mathcal{E}_{t}\subset\mathcal{E}_{t+1}=f^{n_{0}*}\mathcal{S}\

is strong 555that is the pull back of the Harder-Narasimhan filtration on (fn0)𝒮(f^{n_{0}})^{*}\mathcal{S} via fnn0f^{n-n_{0}} gives the the Harder-Narasimhan filtration on (fn)𝒮(f^{n})^{*}\mathcal{S}- see Theorem 2.13, Theorem 2.17.. For any 1st+11\leq s\leq t+1, set μs\mu_{s} to be the normalized slope μ(s/s1)pn0\frac{\mu(\mathcal{E}_{s}/\mathcal{E}_{s-1})}{p^{n_{0}}} of the factor s/s1\mathcal{E}_{s}/\mathcal{E}_{s-1} and set rsr_{s} to be its rank rk(s/s1)\text{rk}(\mathcal{E}_{s}/\mathcal{E}_{s-1}) (see Definition 2.11). Then

(27) F(R,I)(y)=δR1(1+rk(𝒮))eiy+j=1t+1rjeiy(1μjδR)(iy)2.F(R,I)(y)=\delta_{R}\frac{1-(1+\text{rk}(\mathcal{S}))e^{-iy}+\sum\limits_{j=1}^{t+1}r_{j}e^{-iy(1-\frac{\mu_{j}}{\delta_{R}})}}{(iy)^{2}}\ .
Remark 6.2.

The two relations rk(𝒮)=j=1t+1rj\text{rk}(\mathcal{S})=\sum\limits_{j=1}^{t+1}r_{j} and j=1t+1μjrj=δR\sum\limits_{j=1}^{t+1}\mu_{j}r_{j}=-\delta_{R} imply that the numerator of the right hand side of (27) has a zero of order two at the origin. So the right hand side of (27) is holomorphic at the origin. Conversely, the holomorphicity of the Frobenius-Poincaré function and the equality (27) for non-zero complex numbers reveal the two relations.

The key steps in the proof of Theorem 6.1 are Lemma 6.3 and Lemma 6.4. The standard reference for results on sheaf cohomology used here is [Har97].

Lemma 6.3.

For j>2g2j>2g-2 and for all nn, we have

(28) λ((RI[pn])pn+j)=h1(C,(fn𝒮)(j)).\lambda((\frac{R}{{I}^{[{p}^{n}]}})_{p^{n}+j})=h^{1}(C,(f^{n*}\mathcal{S})(j))\ .
Proof.

Given natural numbers nn and jj, first pulling back (25) via fnf^{n} and then tensoring with 𝒪C(j)\mathcal{O}_{C}(j), we get a short exact sequence:

(29) 0(fn𝒮)(j)𝑟𝒪C(j)(h1pn,,hrpn)𝒪C(pn+j)0.0\longrightarrow(f^{n*}\mathcal{S})(j)\longrightarrow\overset{r}{\bigoplus}\mathcal{O}_{C}(j)\xrightarrow{(h_{1}^{p^{n}},\ldots,h_{r}^{p^{n}})}\mathcal{O}_{C}(p^{n}+j)\longrightarrow 0\ .

Note that, since RR is normal, for each jj, the canonical inclusion Rjh0(𝒪C(j))R_{j}\subset h^{0}(\mathcal{O}_{C}(j)) is an isomorphism - see Exercise 5.14 of [Har97]. Also for j>2g2j>2g-2, h1(𝒪C(j))=0h^{1}(\mathcal{O}_{C}(j))=0 (see Example 1.3.4 of [Har97]). So the long exact sequence of sheaf cohomologies corresponding to (29) gives (28). ∎

Lemma 6.4.

Fix ss such that 1st1\leq s\leq t. For all large nn, if an index jj satisfies

(30) pnμsδR+2g2δR<j<pnμs+1δR,-p^{n}\frac{{\mu}_{s}}{\delta_{R}}+\frac{2g-2}{\delta_{R}}<j<-p^{n}\frac{{\mu}_{s+1}}{\delta_{R}}\ ,

then

(31) h1(fn(𝒮)(j))=pn(b=stμb+1rb+1)+(b=strb+1)(g1jδR).h^{1}(f^{n*}(\mathcal{S})(j))=-p^{n}(\sum\limits_{b=s}^{t}{\mu}_{b+1}r_{b+1})+(\sum\limits_{b=s}^{t}r_{b+1})(g-1-j\delta_{R})\,.

And for j>pnμt+1δRj>-p^{n}\frac{\mu_{t+1}}{\delta_{R}},

h1(fn(𝒮)(j))=0.h^{1}(f^{n*}(\mathcal{S})(j))=0.

The only reason for choosing large nn is to ensure that for all ss

pnμsδR+2g2δR<pnμs+1δR.-p^{n}\frac{{\mu}_{s}}{\delta_{R}}+\frac{2g-2}{\delta_{R}}<-p^{n}\frac{{\mu}_{s+1}}{\delta_{R}}.
Proof.

We introduce some notation below which are used in this section.

Set Kj=ker((fn0𝒮)t+1j)K_{j}=\text{ker}((f^{n_{0}*}\mathcal{S})^{\vee}\rightarrow\mathcal{E}_{t+1-j}^{\vee}). Then by Lemma 2.16, there is a HN filtration-

(32) 0=K0K1KtKt+1=(fn0𝒮)0=K_{0}\subset K_{1}\subset\ldots K_{t}\subset K_{t+1}=(f^{n_{0}*}\mathcal{S})^{\vee}

.

Claim.

Denote the sheaf of differentials of CC by ωC\omega_{C}. For jj as in (30), we have

(33) (1)h1(fn(𝒮)(j))=h0(fnn0(Kt+1s)ωC(j))and(2)h1(fnn0(Kt+1s)ωC(j))=0.(1)\ h^{1}(f^{n*}(\mathcal{S})(j))=h^{0}(f^{n-n_{0}*}(K_{t+1-s})\otimes\omega_{C}(-j))\hskip 14.22636pt\text{and}\hskip 14.22636pt(2)\ h^{1}(f^{n-n_{0}*}(K_{t+1-s})\otimes\omega_{C}(-j))=0\ .

We defer proving the above claim until deriving Lemma 6.4 from it.

Combining (33) and the Riemann-Roch theorem on curves (see Theorem 2.6.9, [Pot97]), we get that for the range of values as in (30),

(34) h1(fn(S)(j))=h0(fnn0(Kt+1s)ωC(j))h1(fnn0(Kt+1s)ωC(j))=deg(fnn0(Kt+1s)ωC(j))+(1g)rk(fnn0(Kt+1s)ωC(j))=deg(fnn0Kt+1s)+rk(Kt+1s)deg(ωC(j))+(1g)rk(Kt+1s)(using,(1),Lemma 2.15).\begin{split}h^{1}(f^{n*}(S)(j))&=h^{0}(f^{n-n_{0}*}(K_{t+1-s})\otimes\omega_{C}(-j))-h^{1}(f^{n-n_{0}*}(K_{t+1-s})\otimes\omega_{C}(-j))\\ &=\text{deg}(f^{n-n_{0}*}(K_{t+1-s})\otimes\omega_{C}(-j))+(1-g)\text{rk}(f^{n-n_{0}*}(K_{t+1-s})\otimes\omega_{C}(-j))\\ &=\text{deg}(f^{n-n_{0}*}K_{t+1-s})+\text{rk}(K_{t+1-s})\cdot\text{deg}(\omega_{C}(-j))+(1-g)\text{rk}(K_{t+1-s})\ (\text{using},(1),\lx@cref{creftypecap~refnum}{properties of HN filtration})\ .\end{split}

Since Kt+1sK_{t+1-s} is the kernel of a surjection (fn0S)s(f^{n_{0}*}S)^{\vee}\rightarrow\mathcal{E}_{s}^{\vee}, we have,

(35) deg(fnn0Kt+1s)=pnn0deg(Kt+1s)=pnn0[deg(fn0S))deg(s)]=pnn0[deg(fn0S)+deg(s)]=pnn0b=stdeg(b+1/b)=pnb=stμb+1rb+1.\begin{split}\text{deg}(f^{n-n_{0}*}K_{t+1-s})&=p^{n-n_{0}}\text{deg}(K_{t+1-s})=p^{n-n_{0}}[\text{deg}(f^{n_{0}*}S)^{\vee})-\text{deg}(\mathcal{E}_{s}^{\vee})]\\ &=p^{n-n_{0}}[-\text{deg}(f^{n_{0}*}S)+\text{deg}(\mathcal{E}_{s})]\\ &=-p^{n-n_{0}}\sum\limits_{b=s}^{t}\text{deg}(\mathcal{E}_{b+1}/\mathcal{E}_{b})=-p^{n}\sum\limits_{b=s}^{t}{\mu}_{b+1}r_{b+1}.\end{split}

Similarly one can compute the rank of Kt+1sK_{t+1-s}.

(36) rk(Kt+1s)=rk((fn0S))rk(s)=rk(fn0S)rk(s)=b=strk(b+1/b)=b=strb+1.\begin{split}\text{rk}(K_{t+1-s})&=\text{rk}((f^{n_{0}*}S)^{\vee})-\text{rk}(\mathcal{E}_{s}^{\vee})=\text{rk}(f^{n_{0}*}S)-\text{rk}(\mathcal{E}_{s})\\ &=\sum\limits_{b=s}^{t}\text{rk}(\mathcal{E}_{b+1}/\mathcal{E}_{b})=\sum\limits_{b=s}^{t}r_{b+1}.\end{split}

Now the desired conclusion follows from combining (34), (35), (36) and noting that deg(ωC(j))=2g2jδR\text{deg}(\omega_{C}(-j))=2g-2-j\delta_{R}.

Proof of Claim:.

By Serre Duality (see Corollary 7.7, Chapter III, [Har97]), h1(fn(𝒮)(j))=h0((fn𝒮)ωC(j))h^{1}(f^{n*}(\mathcal{S})(j))=h^{0}((f^{n*}\mathcal{S})^{\vee}\otimes\omega_{C}(-j)). We prove (1) by showing that if the left most inequality in (30) holds, the cokernel of the inclusion

H0(fnn0(Kt+1s)ωC(j))H0((fn𝒮)ωC(j))H^{0}(f^{n-n_{0}*}(K_{t+1-s})\otimes\omega_{C}(-j))\subseteq H^{0}((f^{n*}\mathcal{S})^{\vee}\otimes\omega_{C}(-j))

is zero. For this, first note that by Lemma 2.15,(3), there is a HN filtration -

(37) 0fnn0(Kt+2s)ωC(j)fnn0(Kt+1s)ωC(j)fnn0(Kt)ωC(j)fnn0(Kt+1s)ωC(j)(fn𝒮)ωC(j)fnn0(Kt+1s)ωC(j).0\subset\frac{f^{n-n_{0}*}(K_{t+2-s})\otimes\omega_{C}(-j)}{f^{n-n_{0}*}(K_{t+1-s})\otimes\omega_{C}(-j)}\subset\ldots\subset\frac{f^{n-n_{0}*}(K_{t})\otimes\omega_{C}(-j)}{f^{n-n_{0}*}(K_{t+1-s})\otimes\omega_{C}(-j)}\subset\frac{(f^{n*}\mathcal{S})^{\vee}\otimes\omega_{C}(-j)}{f^{n-n_{0}*}(K_{t+1-s})\otimes\omega_{C}(-j)}.

The slope of the first non-zero term in the HN filtration in (37) is μspn+2g2jδR-\mu_{s}p^{n}+2g-2-j\delta_{R}, which is negative by (30). The desired conclusion now follows from Proposition 2.14.

Now we show that if jj satisfies the right most inequality in (30), then assertion (2) in the claim holds. By Serre duality h1(fnn0(Kt+1s)ωC(j))=h0((fnSfnn0s)(j))h^{1}(f^{n-n_{0}*}(K_{t+1-s})\otimes\omega_{C}(-j))=h^{0}((\frac{f^{n*}S}{f^{n-n_{0}*}\mathcal{E}_{s}})(j)). By Lemma 2.15, (3), the HN filtration on (fnSfnn0s)(j)(\frac{f^{n*}S}{f^{n-n_{0}*}\mathcal{E}_{s}})(j) is as given below-

(38) 0fnn0s+1fnn0s(j)fnn0tfnn0s(j)fnSfnn0s(j).0\subset\frac{f^{n-n_{0}*}\mathcal{E}_{s+1}}{f^{n-n_{0}*}\mathcal{E}_{s}}(j)\subset\ldots\subset\frac{f^{n-n_{0}*}\mathcal{E}_{t}}{f^{n-n_{0}*}\mathcal{E}_{s}}(j)\subset\frac{f^{n*}S}{f^{n-n_{0}*}\mathcal{E}_{s}}(j).

Since the slope of the first non-zero term in the above filtration is pnμs+1+jδRp^{n}\mu_{s+1}+j\delta_{R}, which negative by (30), using Proposition 2.14, we get the desired conclusion. ∎

Proof of Theorem 6.1:.

We shall show that (27) holds for all non-zero yy. Then by the principle of analytic continuation (see page 127, [Ahl79]), we get (27) at all points.

Fix an open subset UU of the complex plane whose closure is compact and the closure does not contain the origin. We fix some notations below which we use in the ongoing proof. For 1st+11\leq s\leq t+1, set,

ls(n)=μs.pnδRandus(n)=μs.pnδR+2g2+1.l_{s}(n)=\lfloor-\mu_{s}.\frac{p^{n}}{\delta_{R}}\rfloor\,\,\,\text{and}\,\,\,u_{s}(n)=\lceil-\mu_{s}.\frac{p^{n}}{\delta_{R}}+2g-2\rceil+1.

Note that

(39) limnls(n)pn=us(n)pn=μsδR.\underset{n\to\infty}{\lim}\frac{l_{s}(n)}{p^{n}}=\frac{u_{s}(n)}{p^{n}}=\frac{-\mu_{s}}{\delta_{R}}\ .

There is a sequence of functions (gn)n(g_{n})_{n} such that for yUy\in U, we have

(40) j=0λ((RI[pn])j)eiyj/pn=j<pnλ((RI[pn])j)eiyj/pn+j=2g1l1(n)1λ((RI[pn])j+pn)eiy(j+pn)/pn+s=1tj=us(n)ls+1(n)1λ((RI[pn])j+pn)eiy(j+pn)/pn+gn(y).\begin{split}\sum\limits_{j=0}^{\infty}\lambda((\frac{R}{{I}^{[{p}^{n}]}})_{j})e^{-iyj/p^{n}}&=\underset{j<p^{n}}{\sum}\lambda((\frac{R}{{I}^{[{p}^{n}]}})_{j})e^{-iyj/p^{n}}+\sum\limits_{j=2g-1}^{l_{1}(n)-1}\lambda((\frac{R}{{I}^{[{p}^{n}]}})_{j+p^{n}})e^{-iy(j+p^{n})/p^{n}}\\ &+\sum\limits_{s=1}^{t}\sum\limits_{j=u_{s}(n)}^{l_{s+1}(n)-1}\lambda((\frac{R}{{I}^{[{p}^{n}]}})_{j+p^{n}})e^{-iy(j+p^{n})/p^{n}}+g_{n}(y)\ .\end{split}

In Lemma 6.5, we compute limits of the terms appearing on the right hand side of (40) normalized by (1pn)2(\frac{1}{p^{n}})^{2}.

Lemma 6.5.

For yUy\in U, we have

  1. (1)

    limn(1pn)2gn(y)=0.\underset{n\to\infty}{\lim}(\frac{1}{p^{n}})^{2}g_{n}(y)=0.

  2. (2)

    limn(1pn)2j<pnλ((RI[pn])j)eiyj/pn=δR(eiy1(iy)2)δReiyiy.\underset{n\to\infty}{\lim}(\frac{1}{p^{n}})^{2}\underset{j<p^{n}}{\sum}\lambda((\frac{R}{{I}^{[{p}^{n}]}})_{j})e^{-iyj/p^{n}}=-\delta_{R}(\frac{e^{-iy}-1}{(iy)^{2}})-\delta_{R}\frac{e^{-iy}}{iy}.

  3. (3)

    limn(1pn)2j=2g1l1(n)1λ((RI[pn])j+pn)eiy(j+pn)/pn.\underset{n\to\infty}{\lim}(\frac{1}{p^{n}})^{2}\sum\limits_{j=2g-1}^{l_{1}(n)-1}\lambda((\frac{R}{{I}^{[{p}^{n}]}})_{j+p^{n}})e^{-iy(j+p^{n})/p^{n}}.
    =(b=0tμb+1rb+1)(1eiyμ1δRiy)eiyiδR(b=0trb+1)ddy(1eiyμ1δRiy)eiy.=-(\sum\limits_{b=0}^{t}\mu_{b+1}r_{b+1})(\frac{1-e^{\frac{iy\mu_{1}}{\delta_{R}}}}{iy})e^{-iy}-i\delta_{R}(\sum\limits_{b=0}^{t}r_{b+1})\dfrac{d}{dy}(\frac{1-e^{\frac{iy\mu_{1}}{\delta_{R}}}}{iy})e^{-iy}.

  4. (4)

    For any ss, 1st1\leq s\leq t, limn(1pn)2j=us(n)ls+1(n)1λ((RI[pn])j+pn)eiy(j+pn)/pn.\underset{n\to\infty}{\lim}(\frac{1}{p^{n}})^{2}\sum\limits_{j=u_{s}(n)}^{l_{s+1}(n)-1}\lambda((\frac{R}{{I}^{[{p}^{n}]}})_{j+p^{n}})e^{-iy(j+p^{n})/p^{n}}.
    =(b=stμb+1rb+1)(eiyμsδReiyμs+1δRiy)eiyiδR(b=strb+1)ddy(eiyμsδReiyμs+1δRiy)eiy.=(-\sum\limits_{b=s}^{t}\mu_{b+1}r_{b+1})(\frac{e^{\frac{iy\mu_{s}}{\delta_{R}}}-e^{\frac{iy\mu_{s+1}}{\delta_{R}}}}{iy})e^{-iy}-i\delta_{R}(\sum\limits_{b=s}^{t}r_{b+1})\dfrac{d}{dy}(\frac{e^{\frac{iy\mu_{s}}{\delta_{R}}}-e^{\frac{iy\mu_{s+1}}{\delta_{R}}}}{iy})e^{-iy}.

Continuation of proof of Theorem 6.1: We establish the statement Theorem 6.1 using Lemma 6.5 before verifying Lemma 6.5. When we use Lemma 6.5 to compute limn(1pn)2jλ((RI[pn])j)eiyj/pn\underset{n\to\infty}{\lim}(\frac{1}{p^{n}})^{2}\underset{j\in\mathbb{N}}{\sum}\lambda((\frac{R}{{I}^{[{p}^{n}]}})_{j})e^{-iyj/p^{n}}, some terms on the right hand side of 3., Lemma 6.5 cancel some terms on the right hand side of 4., Lemma 6.5. After cancelling appropriate terms we get, for yUy\in U

limn(1pn)2jλ((RI[pn])j)eiyj/pn\underset{n\to\infty}{\lim}(\frac{1}{p^{n}})^{2}\underset{j\in\mathbb{N}}{\sum}\lambda((\frac{R}{{I}^{[{p}^{n}]}})_{j})e^{-iyj/p^{n}}

=δR(eiy1(iy)2)δReiyiy(b=0tμb+1rb+1)eiyiy+b=0tμb+1rb+1eiy(1μb+1δR)iy=-\delta_{R}(\frac{e^{-iy}-1}{(iy)^{2}})-\delta_{R}\frac{e^{-iy}}{iy}-\frac{(\sum\limits_{b=0}^{t}\mu_{b+1}r_{b+1})e^{-iy}}{iy}+\frac{\sum\limits_{b=0}^{t}\mu_{b+1}r_{b+1}e^{-iy(1-\frac{\mu_{b+1}}{\delta_{R}})}}{iy}

iδR(b=0trb+1)ddy(1iy)eiy+iδRb=0trb+1ddy(eiyμb+1δRiy)eiy-i\delta_{R}(\sum\limits_{b=0}^{t}r_{b+1})\frac{d}{dy}(\frac{1}{iy})e^{-iy}+i\delta_{R}\sum\limits_{b=0}^{t}r_{b+1}\frac{d}{dy}(\frac{e^{iy\frac{\mu_{b+1}}{\delta_{R}}}}{iy})e^{-iy}

=δR(eiy1(iy)2)δReiyiy+δReiyiy+b=0tμb+1rb+1eiy(1μb+1δR)iy=-\delta_{R}(\frac{e^{-iy}-1}{(iy)^{2}})-\delta_{R}\frac{e^{-iy}}{iy}+\frac{\delta_{R}e^{-iy}}{iy}+\frac{\sum\limits_{b=0}^{t}\mu_{b+1}r_{b+1}e^{-iy(1-\frac{\mu_{b+1}}{\delta_{R}})}}{iy}

δR(rk(𝒮))eiy(iy)2b=0tμb+1rb+1eiy(1μb+1δR)iy+δRb=0trb+1eiy(1μb+1δR)(iy)2-\delta_{R}(\text{rk}(\mathcal{S}))\frac{e^{-iy}}{(iy)^{2}}-\frac{\sum\limits_{b=0}^{t}\mu_{b+1}r_{b+1}e^{-iy(1-\frac{\mu_{b+1}}{\delta_{R}})}}{iy}+\delta_{R}\frac{\sum\limits_{b=0}^{t}r_{b+1}e^{-iy(1-\frac{\mu_{b+1}}{\delta_{R}})}}{(iy)^{2}}  .

The last line is indeed equal to the right hand side of (27).

Proof of Lemma 6.5:.

1) We show that there is a constant CC such that |gn(y)|Cpn|g_{n}(y)|\leq Cp^{n} on UU. By Lemma 6.3 and Lemma 6.4, λ((RI[pn])l)=0\lambda((\frac{R}{{I}^{[{p}^{n}]}})_{l})=0 for l>μt+1pnδR+pn+2g2l>-\mu_{t+1}\frac{p^{n}}{\delta_{R}}+p^{n}+2g-2. So using (40) we get an integer NN, such that each gng_{n} is a sum of at most NN functions of the form λ((RI[pn])l)eiyl/pn\lambda((\frac{R}{{I}^{[{p}^{n}]}})_{l})e^{-iyl/p^{n}}, where ll is at most μt+1pnδR+pn+2g2-\mu_{t+1}\frac{p^{n}}{\delta_{R}}+p^{n}+2g-2. We prove (1) by showing that there is a CC^{\prime} such that for each of these functions λ((RI[pn])l)eiyl/pn\lambda((\frac{R}{{I}^{[{p}^{n}]}})_{l})e^{-iyl/p^{n}} appearing in gng_{n}, |λ((RI[pn])l)eiyl/pn|Cpn|\lambda((\frac{R}{{I}^{[{p}^{n}]}})_{l})e^{-iyl/p^{n}}|\leq C^{\prime}p^{n} on UU. For that, note since UU has a compact closure, there is a constant C1C_{1} such that, for all lμt+1pnδR+pn+2g2l\leq-\mu_{t+1}\frac{p^{n}}{\delta_{R}}+p^{n}+2g-2, |eiyl/pn||e^{-iyl/p^{n}}| is bounded above by C1C_{1} on UU. Since λ((RI[pn])l)λ(Rl)\lambda((\frac{R}{{I}^{[{p}^{n}]}})_{l})\leq\lambda(R_{l}) and there is a constant C2C_{2} such that for all ll, λ(Rl)C2.l\lambda(R_{l})\leq C_{2}.l, we are done.

(2)

(41) limn(1pn)2j<pnλ((RI[pn])j)eiyj/pn=limn(1pn)2j<pnλ(Rj)eiyj/pn=limn(1pn)2j=0pn1λ(Rj)eiyj/pnpn(1eiy/pn)iy=limn1pnj=0pn1λ(Rj)j/pn(j+1)/pneiyxdx.\begin{split}\underset{n\to\infty}{\lim}(\frac{1}{p^{n}})^{2}\underset{j<p^{n}}{\sum}\lambda((\frac{R}{{I}^{[{p}^{n}]}})_{j})e^{-iyj/p^{n}}&=\underset{n\to\infty}{\lim}(\frac{1}{p^{n}})^{2}\underset{j<p^{n}}{\sum}\lambda(R_{j})e^{-iyj/p^{n}}\\ &=\underset{n\to\infty}{\lim}(\frac{1}{p^{n}})^{2}\sum\limits_{j=0}^{p^{n}-1}\lambda(R_{j})e^{-iyj/p^{n}}\frac{p^{n}(1-e^{-iy/p^{n}})}{iy}\\ &=\underset{n\to\infty}{\lim}\frac{1}{p^{n}}\sum\limits_{j=0}^{p^{n}-1}\lambda(R_{j})\int_{j/p^{n}}^{(j+1)/p^{n}}e^{-iyx}dx.\end{split}

Now consider (hn)n(h_{n})_{n} the sequence of real valued functions on [0,1][0,1] defined by hn(x)=1pnλ(Rxpn)h_{n}(x)=\frac{1}{p^{n}}\lambda(R_{\lfloor xp^{n}\rfloor}). The last line of (41) is then 01hn(x)eiyxdx\int_{0}^{1}h_{n}(x)e^{-iyx}dx. Since hn(x)h_{n}(x) converges to the function δRx\delta_{R}x uniformly on [0,1][0,1], we have

(42) limn(1pn)2j<pnλ((RI[pn])j)eiyj/pn=δR01xeiyxdx=δR(eiy1(iy)2)δReiyiy.\begin{split}\underset{n\to\infty}{\lim}(\frac{1}{p^{n}})^{2}\underset{j<p^{n}}{\sum}\lambda((\frac{R}{{I}^{[{p}^{n}]}})_{j})e^{-iyj/p^{n}}&=\delta_{R}\int_{0}^{1}xe^{-iyx}dx\\ &=-\delta_{R}(\frac{e^{-iy}-1}{(iy)^{2}})-\delta_{R}\frac{e^{-iy}}{iy}\ .\end{split}

(3) Using Lemma 6.3 and Lemma 6.4, we get,

(43) (1pn)2j=2g1l1(n)1λ((RI[pn])j+pn)eiy(j+pn)/pn=[eiypn(b=0tμb+1rb+1)+eiy(1pn)2(b=0trb+1)(g1)](j=2g1l1(n)1eiyj/pn)eiy(1pn)2(b=0trb+1)δRj=2g1l1(n)1jeiyj/pn=[1pnb=0tμb+1rb+1+(1pn)2(b=0trb+1)(g1)]1eiy(l1(n)2g+1)/pn1eiy/pneiy(1+2g1pn)ipn(b=0trb+1)δRddy(eiy(2g1)/pneiyl1(n)/pn1eiy/pn)eiy(usingddy(eiyj/pn)=ijpneiyj/pn).\begin{split}&(\frac{1}{p^{n}})^{2}\sum\limits_{j=2g-1}^{l_{1}(n)-1}\lambda((\frac{R}{{I}^{[{p}^{n}]}})_{j+p^{n}})e^{-iy(j+p^{n})/p^{n}}\\ &=[-\frac{e^{-iy}}{p^{n}}(\sum\limits_{b=0}^{t}\mu_{b+1}r_{b+1})+e^{-iy}(\frac{1}{p^{n}})^{2}(\sum\limits_{b=0}^{t}r_{b+1})(g-1)](\sum\limits_{j=2g-1}^{l_{1}(n)-1}e^{-iyj/p^{n}})\\ &-e^{-iy}(\frac{1}{p^{n}})^{2}(\sum\limits_{b=0}^{t}r_{b+1})\delta_{R}\sum\limits_{j=2g-1}^{l_{1}(n)-1}je^{-iyj/p^{n}}\\ &=[-\frac{1}{p^{n}}\sum\limits_{b=0}^{t}\mu_{b+1}r_{b+1}+(\frac{1}{p^{n}})^{2}(\sum\limits_{b=0}^{t}r_{b+1})(g-1)]\frac{1-e^{-iy(l_{1}(n)-2g+1)/p^{n}}}{1-e^{-iy/p^{n}}}e^{-iy(1+\frac{2g-1}{p^{n}})}\\ &-\frac{i}{p^{n}}(\sum\limits_{b=0}^{t}r_{b+1})\delta_{R}\dfrac{d}{dy}(\frac{e^{-iy(2g-1)/p^{n}}-e^{-iyl_{1}(n)/p^{n}}}{1-e^{-iy/p^{n}}})e^{-iy}\,\,\,\,(\text{using}\ \frac{d}{dy}(e^{-iyj/p^{n}})=\frac{-ij}{p^{n}}e^{-iyj/p^{n}})\ .\end{split}

While taking limit as nn approaches infinity of the last line in (43), the order of differentiation and taking limit can be exchanged (see [Ahl79], Chapter 5, Theorem 1). So taking limit as nn approaches infinity in (43) and using (39), Lemma 3.3 we get (3).

(4) Lemma 6.3, Lemma 6.4 and a computation as in the proof of (3) shows that for 1st1\leq s\leq t,

(44) j=us(n)ls+1(n)1λ((RI[pn])j+pn)eiy(j+pn)/pn=[1pnb=stμb+1rb+1+(1pn)2(b=strb+1)(g1)]eiyus(n)/pneiyls+1(n)/pn1eiy/pneiyipn(b=strb+1)δRddy(eiyus(n)/pneiyls+1(n)/pn1eiy/pn)eiy.\begin{split}&\sum\limits_{j=u_{s}(n)}^{l_{s+1}(n)-1}\lambda((\frac{R}{{I}^{[{p}^{n}]}})_{j+p^{n}})e^{-iy(j+p^{n})/p^{n}}\\ &=[-\frac{1}{p^{n}}\sum\limits_{b=s}^{t}\mu_{b+1}r_{b+1}+(\frac{1}{p^{n}})^{2}(\sum\limits_{b=s}^{t}r_{b+1})(g-1)]\frac{e^{-iyu_{s}(n)/p^{n}}-e^{-iyl_{s+1}(n)/p^{n}}}{1-e^{-iy/p^{n}}}e^{-iy}\\ &-\frac{i}{p^{n}}(\sum\limits_{b=s}^{t}r_{b+1})\delta_{R}\dfrac{d}{dy}(\frac{e^{-iyu_{s}(n)/p^{n}}-e^{-iyl_{s+1}(n)/p^{n}}}{1-e^{-iy/p^{n}}})e^{-iy}\ .\end{split}

Now (4) follows from taking limit as nn approaches infinity and arguing as in the proof of (3). ∎

References

  • [Ahl79] Lars V. Ahlfors “Complex Analysis” McGraw-Hill Book Company, 1979
  • [AB93] Luchezar L. Avramov and Ragnar-Olaf Buchweitz “Lower bounds for Betti numbers” In Compositio Mathematica 86.2 Kluwer Academic Publishers, 1993, pp. 147–158 URL: http://www.numdam.org/item/CM_1993__86_2_147_0/
  • [Bre07] Holger Brenner “The Hilbert–Kunz Function in Graded Dimension Two” In Communications in Algebra 35.10 Taylor & Francis, 2007, pp. 3199–3213 URL: https://doi.org/10.1080/00914030701410203
  • [BH98] Winfried Bruns and H. Herzog “Cohen-Macaulay Rings”, Cambridge Studies in Advanced Mathematics Cambridge University Press, 1998 DOI: 10.1017/CBO9780511608681
  • [Cra02] Yongwei Yao Craig Huneke “Unmixed local rings with minimal Hilbert-Kunz multiplicity are regular” In Proc. of the Amer. Math. Soc. 130.4, 2002, pp. 661–665 DOI: https://doi.org/10.1090/S0002-9939-01-06113-5
  • [Erm17] Daniel Erman “Divergent series and Serre’s intersection formula for graded rings” In Advances in Mathematics 314, 2017, pp. 573–582 DOI: https://doi.org/10.1016/j.aim.2017.05.010
  • [FT03] N. Fakhruddin and V. Trivedi “Hilbert–Kunz functions and multiplicities for full flag varieties and elliptic curves” In Journal of Pure and Applied Algebra 181.1, 2003, pp. 23–52 DOI: https://doi.org/10.1016/S0022-4049(02)00304-3
  • [GN01] Shiro Goto and Yukio Nakamura “Multiplicity and Tight Closures of Parameters” In Journal of Algebra 244.1, 2001, pp. 302–311 DOI: https://doi.org/10.1006/jabr.2001.8907
  • [GUL74] TOR H.A GULLIKSEN “A Change of Ring Theorem with Applications to Poincare Series and Intersection Multiplicity” In Mathematica Scandinavica 34.2, 1974, pp. 167–183
  • [HN75] G Harder and M.S. Narasimhan “On the cohomology groups of moduli spaces of vector bundles on curves.” In Math. Ann. 212, 1975, pp. 215–248 DOI: https://doi.org/10.1007/BF01357141
  • [Har97] Robin Hartshorne “Algebraic Geometry”, Graduate Texts in Mathematics Springer-Verlag New York, 1997 DOI: 10.1007/978-1-4757-3849-0
  • [HH90] Melvin Hochster and Craig Huneke “Tight Closure, Invariant Theory, and the Briancon-Skoda Theorem” In Journal of the American Mathematical Society 3.1, 1990, pp. 31–116 URL: https://www.jstor.org/stable/1990984
  • [Hor65] Lars Hormander “LINEAR PARTIAL DIFFERENTIAL OPERATORS” Springer-Verlag, 1965
  • [Hun13] Craig Huneke “Hilbert–Kunz Multiplicity and the F-Signature” In Commutative Algebra: Expository Papers Dedicated to David Eisenbud on the Occasion of His 65th Birthday Springer, 2013, pp. 485–525 URL: https://arxiv.org/abs/1409.0467
  • [Hun96] Craig Huneke “Tight Closure and its Applications”, Regional Conference Series in Mathematics 88 AMSCMBS, 1996
  • [HH93] Craig Huneke and Melvin Hochster “Phantom Homology” In Memoirs of the Amer. Math. Soc 103.490, 1993 DOI: https://doi.org/http://dx.doi.org/10.1090/memo/0490
  • [HMM04] Craig Huneke, Moira McDermott and Paul Monsky “Hilbert-Kunz functions for normal rings” In Math. Res. Letters 11.4, 2004, pp. 539–546 DOI: DOI: https://dx.doi.org/10.4310/MRL.2004.v11.n4.a11
  • [HS06] Craig Huneke and Irena Swanson “Integral Closure of Ideals, Rings, and Modules”, Lecture Note Series 336, London Mathematical Society Cambridge University Press, 2006
  • [HTW11] Craig Huneke, Shunsuke Takagi and Kei-ichi Watanabe “Multiplicity bounds in graded rings” In Kyoto J. Math 51, 2011, pp. 127–147 DOI: https://doi.org/10.1215/0023608X-2010-022
  • [Kei00] Ken-ichi Yoshida Kei-ichi Watanabe “Hilbert–Kunz Multiplicity and an Inequality between Multiplicity and Colength” In Journal of Algebra 230, 2000, pp. 295–317
  • [Kun69] Ernst Kunz “Characterizations of regular local rings for characteristic p” In Amer. J. Math 91, 1969, pp. 772–784 DOI: https://doi.org/10.2307/2373351
  • [Lan04] Adrian Langer “Semistable sheaves in positive characteristic” In Ann. of Math. 159, 2004, pp. 251–276 DOI: https://doi.org/10.4007/annals.2004.159.251
  • [Man04] Florian Enescu Manuel Blickle “On rings with small Hilbert-Kunz multiplicity” In Proc. of the Amer. Math. Soc. 132, 2004, pp. 2505–2509 DOI: https://doi.org/10.1090/S0002-9939-04-07469-6
  • [Mon83] Paul Monsky “The Hilbert-Kunz Function” In Math. Ann. 263 Springer-Verlag, 1983, pp. 43–49 DOI: https://doi.org/10.1007/BF01457082
  • [Muk23] Alapan Mukhopadhyay “The Frobenius-Poincaré Function and Hilbert-Kunz Multiplicity” Available at https://dx.doi.org/10.7302/8449, 2023
  • [PS73] Christian Peskine and Lucien Szpiro “Dimension projective finie et cohomologie locale” In Publications Mathématiques de l’IHÉS 42, 1973, pp. 47–119 URL: http://www.numdam.org/item/PMIHES_1973__42__47_0/
  • [Pot97] Joseph Le Potier “Lectures on vector bundles” Cambridge University Press, 1997
  • [Rud87] Walter Rudin “Real and Complex Analysis”, McGraw-Hill Series in Higher Mathematics McGraw-Hill Book Company, 1987
  • [Ser00] J.P Serre “Local Algebra” Springer-Verlag Berlin Heidelberg, 2000 DOI: 10.1007/978-3-662-04203-8
  • [Tei02] Pedro Teixeira “p-Fractals and Hilbert-Kunz Series”, 2002 URL: http://www.math.union.edu/people/faculty/publications/teixeirp/teixeira-thesis.pdf
  • [Tri18] V. Trivedi “Hilbert-Kunz Density Function and Hilbert-Kunz Multiplicity” In Transactions of the American Mathematical Society 370.12, 2018, pp. 8403–8428 DOI: https://doi.org/10.1090/tran/7268
  • [Tri05] Vijaylaxmi Trivedi “Semistability and Hilbert–Kunz multiplicities for curves” In Journal of Algebra 284, 2005, pp. 627–644
  • [Tri21] Vijaylaxmi Trivedi “The Hilbert-Kunz density functions of quadric hypersurfaces”, 2021 URL: arXiv:2109.11784v1
  • [TW22] Vijaylaxmi Trivedi and Kei-Ichi Watanabe “Hilbert-Kunz density function for graded domains” In Journal of Pure and Applied Algebra 226.2, 2022, pp. 106835 DOI: https://doi.org/10.1016/j.jpaa.2021.106835
  • [TW21] Vijaylaxmi Trivedi and Kei-Ichi Watanabe “Hilbert-Kunz density functions and F-thresholds” In Journal of Algebra 567, 2021, pp. 533–563 DOI: https://doi.org/10.1016/j.jalgebra.2020.09.025
  • [ZS60] Oscar Zariski and Pierre Samuel “Commutative Algebra” D. VAN NOSTRAND COMPANY, INC., 1960
  • [ZS65] Oscar Zariski and Pierre Samuel “Commutative Algebra” D. VAN NOSTRAND COMPANY, INC., 1965