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

Purity of monoids and characteristic-free splittings in semigroup rings

Alessandro De Stefani1 Dipartimento di Matematica, Università di Genova, Via Dodecaneso 35, 16146 Genova, Italy [email protected] Jonathan Montaño2 School of Mathematical and Statistical Sciences, Arizona State University, P.O. Box 871804, Tempe, AZ 85287-18041 [email protected]  and  Luis Núñez-Betancourt3 Centro de Investigación en Matemáticas, Guanajuato, Gto., México [email protected]
Abstract.

Inspired by methods in prime characteristic in commutative algebra, we introduce and study combinatorial invariants of seminormal monoids. We relate such numbers with the singularities and homological invariants of the semigroup ring associated to the monoid. Our results are characteristic independent.

Key words and phrases:
Monoids, pure maps, seminormality, numerical invariants
2020 Mathematics Subject Classification:
Primary 20M32, 20M25, 13A35 ; Secondary 13C15
1 The first author was partially supported by the PRIN 2020 project 2020355B8Y “Squarefree Gröbner degenerations, special varieties and related topics".
2 The second author was supported by NSF Grant DMS #2001645/2303605.
3 The third author was supported by CONACyT Grant #284598.

1. Introduction

Frobenius splittings have inspired a large number of results in commutative algebra, algebraic geometry, and representation theory. In this manuscript we seek to continue this approach in the context of combinatorics of monoids. Given a monoid M0qM\subseteq\mathbb{Z}^{q}_{\geqslant 0} for some q>0q\in\mathbb{Z}_{>0}, and m>0m\in\mathbb{Z}_{>0}, we study the pure MM-submodules of 1mM\frac{1}{m}M that are translations of MM, which algebraically corresponds to free summands of 𝕜[1mM]\mathbb{k}[\frac{1}{m}M] as 𝕜[M]\mathbb{k}[M]-module. It turns out that the purity of M1mMM\subseteq\frac{1}{m}M detects both normality and seminormality (see Proposition 3.5). The study of pure submodules, or equivalently of free summands, of normal monoids was already initiated by other authors in order to compute the FF-signature of normal affine semigroup rings [ToricSingh, von2011f]. Moreover, the structure of 1mM\frac{1}{m}M as MM-module was described by Bruns and Gubeladze [BGSemigroups, BrunsDivisorNormalMonoid] for normal monoids (see [Shibuta] for a related result in prime characteristic).

In this manuscript we study combinatorial numerical invariants of a seminormal monoid. Our key motivation is that seminormality for a monoid can be seen as a characteristic-free version of FF-purity for affine semigroup rings. For more information and examples on seminormal monoids we direct the interested reader to Li’s thesis on this subject [Li].

In Definition 3.19 we introduce the notion of pure threshold of a seminormal monoid MM, denoted by mpt(M){\operatorname{mpt}}(M), which is motivated by the FF-pure threshold in prime characteristic. This number can be described as the largest degree of a pure translation of MM inside the cone 0M\mathbb{R}_{\geqslant 0}M or, equivalently, of 1mM\frac{1}{m}M for some mm. We show that mpt(M){\operatorname{mpt}}(M) gives an upper bound for the Castelnuovo-Mumford regularity reg(𝕜[M])\operatorname{reg}(\mathbb{k}[M]) defined in terms of local cohomology, and the Castelnuovo-Mumford regularity Reg(𝕜[M])\operatorname{Reg}(\mathbb{k}[M]) defined in terms of graded Betti numbers of 𝕜[M]\mathbb{k}[M] (see Section 2 for more details).

Theorem A (Theorem 5.4).

Let MM be a seminormal monoid with a minimal set of generators {γ1,,γu}\{\gamma_{1},\ldots,\gamma_{u}\}. Then, ai(𝕜[M])mpt(M)a_{i}(\mathbb{k}[M])\leqslant-{\operatorname{mpt}}(M). As a consequence,

reg(𝕜[M])=max{ai(𝕜[M])i}dim(𝕜[M])mpt(M)=rank(M)mpt(M).\operatorname{reg}(\mathbb{k}[M])=\max\{a_{i}(\mathbb{k}[M])-i\}\leqslant\dim(\mathbb{k}[M])-{\operatorname{mpt}}(M)=\operatorname{rank}(M)-{\operatorname{mpt}}(M).

Moreover, if we present RR as S/IS/I, where S=𝕜[x1,,xu]S=\mathbb{k}[x_{1},\ldots,x_{u}] and each xix_{i} has degree di:=deg(xi)=|γi|d_{i}:=\deg(x_{i})=|\gamma_{i}| the degree of γi\gamma_{i} for i=1,,ui=1,\ldots,u, and ISI\subseteq S is a homogeneous ideal, then

Reg(𝕜[M])=sup{βiS(M)ii}rank(M)+i=1u(di1)mpt(M).\operatorname{Reg}(\mathbb{k}[M])=\sup\{\beta_{i}^{S}(M)-i\mid i\in\mathbb{Z}\}\leqslant\operatorname{rank}(M)+\sum_{i=1}^{u}(d_{i}-1)-{\operatorname{mpt}}(M).

Theorem A allows us to give an upper bound for the degrees of generators of the defining ideal II. We also show that mpt(M){\operatorname{mpt}}(M) is a rational if MM is a normal (see Proposition 4.4). Despite mpt(M){\operatorname{mpt}}(M) being inspired by FF-pure thresholds, these numbers do not always coincide (see Example 3.23 and Remark 3.24). In addition, mpt(M){\operatorname{mpt}}(M) is defined independently of the field 𝕜\mathbb{k} and so it is a characteristic-free invariant, while the FF-pure threshold is only defined when 𝕜\mathbb{k} has prime characteristic.

We introduce the pure prime ideal 𝒫(M)\mathcal{P}(M), and the pure prime face M\mathcal{F}_{M}, of a seminormal monoid MM (see Corollary 3.28, and Definitions 3.26 and 3.29). The former emulates the splitting prime ideal of an FF-pure ring, while the latter is related to the quotient of a ring by its splitting prime. In fact, the submonoid MM\mathcal{F}_{M}\cap M is normal (see Corollary 3.28). We note that the rank of MM\mathcal{F}_{M}\cap M is a monoid version of the splitting dimension and so we call it the pure dimension and denote it by mpdim(M){\operatorname{mpdim}}(M). It turns out that this rank is equal to the rank of MM if and only if MM is normal, and it is non-negative if and only if MM is seminormal (see Corollary 3.30). Therefore, in some sense, mpdim(M){\operatorname{mpdim}}(M) measures how far a seminormal monoid is from being normal. Furthermore, mpdim(M){\operatorname{mpdim}}(M) is related to the depth of 𝕜[M]\mathbb{k}[M] as the following theorem shows.

Theorem B (Theorem 5.7).

If MM is a seminormal monoid, then mpdim(M)depth(𝕜[M]).{\operatorname{mpdim}}(M)\leqslant\operatorname{depth}(\mathbb{k}[M]).

We point out that Theorem B recovers Hochster’s result that normal semigroup rings are Cohen-Macaulay [HochsterToric].

Finally, we consider the growth of the number of disjoint pure translations of MM in 1mM\frac{1}{m}M as mm varies. More specifically, if m>0m\in\mathbb{Z}_{>0} is such that 1mMM=M\frac{1}{m}M\cap\mathbb{Z}M=M, we define

Vm(M):={α1mM(α+M)1mM is pure}.V_{m}(M):=\left\{\alpha\in\frac{1}{m}M\mid(\alpha+M)\subseteq\frac{1}{m}M\text{ is pure}\right\}.
Theorem C (Theorem 4.6).

Let MM be a seminormal monoid, 𝒜(M)={m>0|1mMM=M}\mathcal{A}(M)=\{m\in\mathbb{Z}_{>0}\;|\;\frac{1}{m}M\cap\mathbb{Z}M=M\}, and s=mpdim(M)s={\operatorname{mpdim}}(M). Then,

mpr(M):=limt|Vmt(M)|mts\operatorname{mpr}(M):=\lim\limits_{t\to\infty}\frac{|V_{m_{t}}(M)|}{m_{t}^{s}}

exists and it is positive for every increasing sequence mt𝒜(M)m_{t}\in\mathcal{A}(M). Furthermore, if MM is normal, then mpr(M)>0\operatorname{mpr}(M)\in\mathbb{Q}_{>0}.

We call the limit in Theorem C the pure ratio of MM. If the field has prime characteristic, this number coincides with the splitting ratio of 𝕜[M]\mathbb{k}[M] [AE]. A consequence of Theorem C is that the value of the FF-splitting ratio depends only on the structure of MM, and so it is independent of the characteristic of the field as long as 𝕜[M]\mathbb{k}[M] is FF-pure. Finally, using this result we give a monoid version of a celebrated Theorem of Kunz [Kunz, Theorem 2.1] which characterizes regularity of rings of prime characteristic in terms of Frobenius (see Theorem 5.9).

Throughout this article we adopt the following notation.

Notation 1.1.

Let 𝕜\mathbb{k} be a field of any characteristic and qq a positive integer. Let M0qM\subseteq\mathbb{Z}^{q}_{\geqslant 0} be an affine monoid, i.e., a finitely generated submonoid of q\mathbb{Z}^{q}. We fix {γ1,,γu}\{\gamma_{1},\ldots,\gamma_{u}\} a minimal set of generators of MM. Let M\mathbb{Z}M denote the group generated by MM and C(M)=0MC(M)=\mathbb{R}_{\geqslant 0}M the cone generated by MM.

2. Background

In this section we include some preliminary information that is needed in the rest of the paper.

Affine monoids and affine semigroup rings

For proofs of the claims in this subsection and further information about affine monoids we refer the reader to Bruns and Gubeladze’s book [bruns2009polytopes]. Let M0qM\subseteq\mathbb{Z}_{\geqslant 0}^{q} be an affine monoid. A subset UqU\subseteq\mathbb{Q}^{q} is an MM-module if U+MUU+M\subseteq U. An MM-module UU is an ideal if it is contained in MM. An ideal UMU\subseteq M is prime if whenever a+bUa+b\in U with a,bMa,b\in M, we must have aUa\in U or bUb\in U. The rank of MM is the dimension of the \mathbb{Q}-vector space M\mathbb{Q}\otimes_{\mathbb{Z}}\mathbb{Z}M.

Let R=𝕜[M]𝕜[𝐱]:=𝕜[x1,,xq]R=\mathbb{k}[M]\subseteq\mathbb{k}[\mathbf{x}]:=\mathbb{k}[x_{1},\ldots,x_{q}] be the affine semigroup associated to MM. As a 𝕜\mathbb{k}-vector space, RR is generated by the monomials {xααM}\{x^{\alpha}\mid\alpha\in M\}. We note that the monomial ideals of RR are precisely those generated by {xααU}\{x^{\alpha}\mid\alpha\in U\} for some ideal UMU\subseteq M. Under this correspondence, prime monomial ideals of RR correspond to prime ideals of MM. For every MM-module UqU\subseteq\mathbb{Q}^{q} we have a corresponding RR-module RU:={xα+ηαM,ηU}RU:=\{x^{\alpha+\eta}\mid\alpha\in M,\eta\in U\} in the algebraic closure of 𝕜(x1,,xq)\mathbb{k}(x_{1},\ldots,x_{q}). Moreover, we have dim(R)=rank(M)\dim(R)=\operatorname{rank}(M).

Graded algebras and modules

A non-negatively graded algebra AA is a ring that admits a direct sum decomposition A=j0AjA=\bigoplus_{j\geqslant 0}A_{j} of Abelian groups such that AiAjAi+jA_{i}\cdot A_{j}\subseteq A_{i+j}. It follows from this that A0A_{0} is a ring, and each AiA_{i} is an A0A_{0}-module. If we let A+=j>0AjA_{+}=\bigoplus_{j>0}A_{j}, then A+A_{+} is an ideal of AA, called the irrelevant ideal.

Throughout this manuscript we will make the assumption that AA is Noetherian or, equivalently, that there exist finitely many elements a1,,anA+a_{1},\ldots,a_{n}\in A_{+} such that A=A0[a1,,an]A=A_{0}[a_{1},\ldots,a_{n}], which can be assumed to be homogeneous of degrees d1,,dnd_{1},\ldots,d_{n}. In this case, note that AA is a quotient of a polynomial ring A0[x1,,xn]A_{0}[x_{1},\ldots,x_{n}] by a homogeneous ideal.

A \mathbb{Z}-graded AA-module is an AA-module NN that admits a direct sum decomposition N=jNjN=\bigoplus_{j\in\mathbb{Z}}N_{j} of Abelian groups, and such that AiNjNi+jA_{i}\cdot N_{j}\subseteq N_{i+j}. As a consequence, each NiN_{i} is an A0A_{0}-module. Moreover, if NN is Noetherian there exists i0i_{0}\in\mathbb{Z} such that Ni=0N_{i}=0 for all i<i0i<i_{0}; on the other hand, if NN is Artinian there exists j0j_{0}\in\mathbb{Z} such that Nj=0N_{j}=0 for all j>j0j>j_{0}.

Given a \mathbb{Z}-graded AA-module NN, and an integer jj\in\mathbb{Z}, we define the shift N(j)N(j) as the \mathbb{Z}-graded AA-module whose ii-th graded component is N(j)i=Ni+jN(j)_{i}=N_{i+j}. In particular, A(j)A(-j) is a free graded AA-module of rank one with generator in degree jj.

Graded local cohomology and Castelnuovo-Mumford regularity

In this subsection we recall general properties of local cohomology. We refer the interested reader to Brodmann and Sharp’s book on this subject [BrodmannSharp]. Let 𝕜\mathbb{k} be a field and S=𝕜[x1,,xn]S=\mathbb{k}[x_{1},\ldots,x_{n}], with deg(xi)=di>0\deg(x_{i})=d_{i}>0. Let NN be a finitely generated \mathbb{Z}-graded SS-module. If we let 𝔪=(x1,,xn)\mathfrak{m}=(x_{1},\ldots,x_{n}), then the graded local cohomology modules H𝔪i(N)H^{i}_{\mathfrak{m}}(N) are Artinian and \mathbb{Z}-graded.

Definition 2.1.

Let ai(N)=sup{jH𝔪i(N)j0}a_{i}(N)=\sup\{j\in\mathbb{Z}\mid H^{i}_{\mathfrak{m}}(N)_{j}\neq 0\} be the ii-th aa-invariant of NN. If N0N\neq 0 we define the Castelnuovo-Mumford regularity of NN as reg(N)=sup{ai(N)+ii=0,,n}\operatorname{reg}(N)=\sup\{a_{i}(N)+i\mid i=0,\ldots,n\}. On the other hand, if N=0N=0 we let reg(N)=\operatorname{reg}(N)=-\infty.

In the standard graded case, reg(N)\operatorname{reg}(N) has a well-known interpretation in terms of graded Betti numbers of NN. In our setup this is still the case, but the degrees of the algebra generators of SS must be taken into account. For i0i\in\mathbb{Z}_{\geqslant 0}, let βiS(N)=a0(ToriS(N,𝕜)){}\beta_{i}^{S}(N)=a_{0}(\operatorname{Tor}_{i}^{S}(N,\mathbb{k}))\in\mathbb{Z}\cup\{-\infty\}. As another way to see this, for a non-zero graded SS-module TT let β(T)\beta(T) be the maximum degree of an element in a minimal homogeneous generating set of TT. If

F:0\textstyle{F_{\bullet}:0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Fc\textstyle{F_{c}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Fc1\textstyle{F_{c-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\ldots\ldots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}F1\textstyle{F_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}F0\textstyle{F_{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}N\textstyle{N\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0,\textstyle{0,}

is a minimal graded free resolution of NN where c:=pd(N)c:=\operatorname{pd}(N) is the projective dimension of NN, then βiS(N)=β(Fi)\beta_{i}^{S}(N)=\beta(F_{i}).

Definition 2.2.

For N0N\neq 0 we let Reg(N)=sup{βiS(N)ii}\operatorname{Reg}(N)=\sup\{\beta_{i}^{S}(N)-i\mid i\in\mathbb{Z}\}, while for N=0N=0 we let Reg(N)=\operatorname{Reg}(N)=-\infty.

In the standard graded case, that is, when d1==dn=1d_{1}=\ldots=d_{n}=1, then reg(N)=Reg(N)\operatorname{reg}(N)=\operatorname{Reg}(N). In our more general scenario, we still have the following relation between the two notions of regularity.

Lemma 2.3.

With the above notation we have that

Reg(N)=reg(N)+i=1n(di1).\operatorname{Reg}(N)=\operatorname{reg}(N)+\sum_{i=1}^{n}(d_{i}-1).
Proof.

We may assume that n>0n>0, otherwise the claim is trivial. We prove the statement by induction on c=pdS(N)c=\operatorname{pd}_{S}(N). If c=0c=0 then NN is free, and it is clear that Reg(N)=β(N)\operatorname{Reg}(N)=\beta(N). On the other hand, ai(N)=0a_{i}(N)=0 for all ini\neq n, while an(N)=an(S)+β(N)=i=1ndi+β(N)a_{n}(N)=a_{n}(S)+\beta(N)=-\sum_{i=1}^{n}d_{i}+\beta(N). It follows that reg(N)=n+an(N)=β(N)+i=1n(1di)=Reg(N)+i=1n(1di)\operatorname{reg}(N)=n+a_{n}(N)=\beta(N)+\sum_{i=1}^{n}(1-d_{i})=\operatorname{Reg}(N)+\sum_{i=1}^{n}(1-d_{i}), and the base case follows.

Now assume that c1c\geqslant 1. We have a graded short exact sequence 0ΩF0N00\to\Omega\to F_{0}\to N\to 0, where F0F_{0} is the first free module in a minimal free resolution of NN, and pdS(Ω)=c1\operatorname{pd}_{S}(\Omega)=c-1. By induction we have that Reg(Ω)=reg(Ω)+i=1n(di1)\operatorname{Reg}(\Omega)=\operatorname{reg}(\Omega)+\sum_{i=1}^{n}(d_{i}-1). Moreover, it is clear from the definitions that Reg(N)=max{β(N),Reg(Ω)1}=max{β(N),reg(Ω)+i=1n(di1)1}\operatorname{Reg}(N)=\max\{\beta(N),\operatorname{Reg}(\Omega)-1\}=\max\{\beta(N),\operatorname{reg}(\Omega)+\sum_{i=1}^{n}(d_{i}-1)-1\}. We have that β(N)reg(N)+i=1n(di1)\beta(N)\leqslant\operatorname{reg}(N)+\sum_{i=1}^{n}(d_{i}-1) [DSMNB, Proposition 3.1], and therefore the above equality gives that

(2.1) Reg(N)max{reg(N),reg(Ω)1}+i=1n(di1).\operatorname{Reg}(N)\leqslant\max\{\operatorname{reg}(N),\operatorname{reg}(\Omega)-1\}+\sum_{i=1}^{n}(d_{i}-1).

We now show that max{reg(N),reg(Ω)1}=reg(N)\max\{\operatorname{reg}(N),\operatorname{reg}(\Omega)-1\}=\operatorname{reg}(N). The short exact sequence 0ΩF0N00\to\Omega\to F_{0}\to N\to 0 yields graded isomorphisms H𝔪i(N)H𝔪i+1(Ω)H^{i}_{\mathfrak{m}}(N)\cong H^{i+1}_{\mathfrak{m}}(\Omega) for all i<n1i<n-1, and a graded exact sequence 0H𝔪n1(N)H𝔪n(Ω)H𝔪n(F0)H𝔪n(N)00\to H^{n-1}_{\mathfrak{m}}(N)\to H^{n}_{\mathfrak{m}}(\Omega)\to H^{n}_{\mathfrak{m}}(F_{0})\to H^{n}_{\mathfrak{m}}(N)\to 0. If we had reg(Ω)1>reg(N)\operatorname{reg}(\Omega)-1>\operatorname{reg}(N), then necessarily reg(Ω)=an(Ω)+n\operatorname{reg}(\Omega)=a_{n}(\Omega)+n, and looking at top degrees in the above exact sequence we also conclude that an(Ω)=an(F0)a_{n}(\Omega)=a_{n}(F_{0}). On the other hand, an(F0)=β(N)+an(S)=β(N)i=1ndireg(N)na_{n}(F_{0})=\beta(N)+a_{n}(S)=\beta(N)-\sum_{i=1}^{n}d_{i}\leqslant\operatorname{reg}(N)-n [DSMNB, Proposition 3.1], and so, reg(Ω)reg(N)\operatorname{reg}(\Omega)\leqslant\operatorname{reg}(N), a contradiction. Thus we always have that reg(Ω)1reg(N)\operatorname{reg}(\Omega)-1\leqslant\operatorname{reg}(N), and by (2.1) the inequality Reg(N)reg(N)+i=1n(di1)\operatorname{Reg}(N)\leqslant\operatorname{reg}(N)+\sum_{i=1}^{n}(d_{i}-1) is proved.

For the reverse inequality, first observe that the above isomorphisms give that ai(N)=ai+1(Ω)a_{i}(N)=a_{i+1}(\Omega) for all i<n1i<n-1, while the exact sequence yields that an1(N)an(Ω)a_{n-1}(N)\leqslant a_{n}(\Omega). Since n>0n>0 and Ω\Omega is a submodule of a free module, it has positive depth, and thus a0(Ω)=0a_{0}(\Omega)=0. It follows that max{ai(N)+ii=0,,n1}max{ai+1(Ω)+ii=0,,n1}=reg(Ω)1\max\{a_{i}(N)+i\mid i=0,\ldots,n-1\}\leqslant\max\{a_{i+1}(\Omega)+i\mid i=0,\ldots,n-1\}=\operatorname{reg}(\Omega)-1. By induction we have that reg(Ω)=Reg(Ω)+i=1n(1di)\operatorname{reg}(\Omega)=\operatorname{Reg}(\Omega)+\sum_{i=1}^{n}(1-d_{i}), and thus max{ai(N)+ii=0,,n1}Reg(Ω)1+i=1n(1di)Reg(N)+i=1n(1di)\max\{a_{i}(N)+i\mid i=0,\ldots,n-1\}\leqslant\operatorname{Reg}(\Omega)-1+\sum_{i=1}^{n}(1-d_{i})\leqslant\operatorname{Reg}(N)+\sum_{i=1}^{n}(1-d_{i}) since the inequality Reg(Ω)1Reg(N)\operatorname{Reg}(\Omega)-1\leqslant\operatorname{Reg}(N) always holds. Now, the above exact sequence on local cohomology also gives that an(N)an(F0)=β(N)+an(S)Reg(N)i=1ndia_{n}(N)\leqslant a_{n}(F_{0})=\beta(N)+a_{n}(S)\leqslant\operatorname{Reg}(N)-\sum_{i=1}^{n}d_{i}, and thus an(N)+nReg(N)+i=1n(1di)a_{n}(N)+n\leqslant\operatorname{Reg}(N)+\sum_{i=1}^{n}(1-d_{i}). In conclusion, we have that reg(N)=sup{ai(N)+ii=0,,n}Reg(N)+i=1n(1di)\operatorname{reg}(N)=\sup\{a_{i}(N)+i\mid i=0,\ldots,n\}\leqslant\operatorname{Reg}(N)+\sum_{i=1}^{n}(1-d_{i}), and the proof is complete. ∎

3. Purity of MM-modules and (semi)normal affine monoids

Definition 3.1.

Let UVqU\subseteq V\subseteq\mathbb{Q}^{q} be MM-modules. We say that the inclusion UVU\subseteq V is pure if VUV\setminus U is also an MM-module.

Example 3.2.

Let M=0(2,1)+0(1,2)M=\mathbb{Z}_{\geqslant 0}(2,1)+\mathbb{Z}_{\geqslant 0}(1,2) be the monoid generated by {(2,1),(1,2)}\{(2,1),(1,2)\}. We have that the inclusion (32,32)+M12M2(\frac{3}{2},\frac{3}{2})+M\subset\frac{1}{2}M\subset\mathbb{Q}^{2} is pure. In Figure 1 we represent the elements of MM with circles and the ones from (32,32)+M(\frac{3}{2},\frac{3}{2})+M with multiplication signs. The shaded region is included to illustrate that (32,32)+M(\frac{3}{2},\frac{3}{2})+M is obtained as a translation of MM.

Refer to caption
Figure 1. For M=0(2,1)+0(1,2)M=\mathbb{Z}_{\geqslant 0}(2,1)+\mathbb{Z}_{\geqslant 0}(1,2), the inclusion (32,32)+M12M2(\frac{3}{2},\frac{3}{2})+M\subset\frac{1}{2}M\subset\mathbb{Q}^{2} is pure.

In the following proposition, we provide equivalent statements for Definition 3.1 in a particular case.

Proposition 3.3.

Let VqV\subseteq\mathbb{Q}^{q} be an MM-module and αV\alpha\in V. The following statements are equivalent:

  1. (i)

    The inclusion (α+M)V(\alpha+M)\subseteq V is pure.

  2. (ii)

    For every γM\gamma\in M and βV\beta\in V we have γ+βαM implies βαM.\gamma+\beta-\alpha\in M\text{ implies }\beta-\alpha\in M.

  3. (iii)

    (Vα)MM\left(V-\alpha\right)\cap\mathbb{Z}M\subseteq M

Proof.

First, assume (i) and let γM\gamma\in M and βV\beta\in V with γ+βα+M\gamma+\beta\in\alpha+M. Thus, βα+M\beta\in\alpha+M and then (ii) follows.

Now, assume (ii) and let βV\beta\in V be such that βαM\beta-\alpha\in\mathbb{Z}M. Write βα=σγ\beta-\alpha=\sigma-\gamma with σ,γM\sigma,\gamma\in M. Then γ+βα=σM\gamma+\beta-\alpha=\sigma\in M implies βαM\beta-\alpha\in M. Thus, (iii) follows.

Finally, assume (iii). Let βV(α+M)\beta\in V\setminus(\alpha+M) and γM\gamma\in M. Assume by means of contradiction that β+γα+M\beta+\gamma\in\alpha+M. Then β+γ=α+σ\beta+\gamma=\alpha+\sigma for some σM\sigma\in M, which implies βα(Vα)MM\beta-\alpha\in\left(V-\alpha\right)\cap\mathbb{Z}M\setminus M which contradicts (iii). Thus, β+γV(α+M)\beta+\gamma\in V\setminus(\alpha+M) and then (i) follows. ∎

We now discuss seminormality and normality, which are the main subjects of study in this manuscript. We refer to the work of Bruns, Li and Römer [BrunsLiRomer] to reader in seminormal rings.

Definition 3.4.

A monoid MM is called seminormal if, whenever αM\alpha\in\mathbb{Z}M is such that 2αM2\alpha\in M and 3αM3\alpha\in M, then αM\alpha\in M. The monoid is called normal if C(M)M=MC(M)\cap\mathbb{Z}M=M.

The following alternative characterization of seminormality and normality is useful for the proof of our main results. While it might be already known to experts, we record it here with a proof for convenience of the reader. For a related result in prime characteristic we refer to the work of Bruns, Li, and Römer [BrunsLiRomer, Section 6].

Proposition 3.5.

Let MM be an affine monoid.

  1. (1)

    MM is seminormal if and only if there exists m>1m\in\mathbb{Z}_{>1} such that 1mMM=M\frac{1}{m}M\cap\mathbb{Z}M=M.

  2. (2)

    MM is normal if and only if 1mMM=M\frac{1}{m}M\cap\mathbb{Z}M=M for every m>1m\in\mathbb{Z}_{>1}.

Proof.

We note that the containment 1mMMM\frac{1}{m}M\cap\mathbb{Z}M\supseteq M holds trivially for any m>1m\in\mathbb{Z}_{>1}.

For (1), assume that 1mMMM\frac{1}{m}M\cap\mathbb{Z}M\subseteq M for some m>1m\in\mathbb{Z}_{>1}. Let αM\alpha\in\mathbb{Z}M be such that 2αM2\alpha\in M and 3αM3\alpha\in M. We can find nonnegative integers aa and bb such that m=2a+3bm=2a+3b, and thus mα=a(2α)+b(3α)Mm\alpha=a(2\alpha)+b(3\alpha)\in M. It follows by our assumption that αM\alpha\in M, and thus MM is seminormal. Conversely, let \mathcal{F} be the set of all faces of C(M)C(M) (of any dimension); we note that \mathcal{F} is a finite set. For any FF\in\mathcal{F} we consider the finitely generated Abelian group GF=(MF)/((MF))G_{F}=(\mathbb{Z}M\cap\mathbb{Q}F)/(\mathbb{Z}(M\cap F)). Let p0p\gg 0 be a prime number such that the ideal (p)(p) is not associated to GFG_{F} as a \mathbb{Z}-module for any FF\in\mathcal{F}. We claim that 1pMMM\frac{1}{p}M\cap\mathbb{Z}M\subseteq M. Let α1pMMC(M)M\alpha\in\frac{1}{p}M\cap\mathbb{Z}M\subseteq C(M)\cap\mathbb{Z}M, then αint(F)\alpha\in\operatorname{int}(F^{\prime}) the interior of some FF^{\prime}\in\mathcal{F}, and also pαMFp\alpha\in M\cap F^{\prime}. By the choice of pp it follows that α(MF)int(F)\alpha\in\mathbb{Z}(M\cap F^{\prime})\cap\operatorname{int}(F^{\prime}) and then αM\alpha\in M [BrunsLiRomer, Theorem 2.1].

For (2), if MM is normal then 1mMMC(M)M=M\frac{1}{m}M\cap\mathbb{Z}M\subseteq C(M)\cap\mathbb{Z}M=M. Conversely, let αC(M)M\alpha\in C(M)\cap\mathbb{Z}M, then α=i=1rtimiαi\alpha=\sum_{i=1}^{r}\frac{t_{i}}{m_{i}}\alpha_{i}, with αiM\alpha_{i}\in M and timi0\frac{t_{i}}{m_{i}}\in\mathbb{Q}_{\geqslant 0}. If we let m=i=1rmim=\prod_{i=1}^{r}m_{i}, it then follows that α1mMM=M\alpha\in\frac{1}{m}M\cap\mathbb{Z}M=M, as desired. ∎

Motivated by the previous result, we consider the following definition.

Definition 3.6.

We set 𝒜(M)={m>1|1mMM=M}\mathcal{A}(M)=\{m\in\mathbb{Z}_{>1}\;|\;\frac{1}{m}M\cap\mathbb{Z}M=M\}.

Remark 3.7.

As a consequence of Proposition 3.5, we deduce that

  1. (1)

    𝒜(M)\mathcal{A}(M)\neq\emptyset if and only if MM is seminormal;

  2. (2)

    MM is normal if and only if 𝒜(M)=>1\mathcal{A}(M)=\mathbb{Z}_{>1}.

We now see that 𝒜(M)\mathcal{A}(M) is a multiplicative set.

Lemma 3.8.

Let m,n>1m,n\in\mathbb{Z}_{>1}. Then mn𝒜(M)mn\in\mathcal{A}(M) if and only if both m𝒜(M)m\in\mathcal{A}(M) and n𝒜(M)n\in\mathcal{A}(M).

Proof.

First assume that mn𝒜(M)mn\in\mathcal{A}(M). Then m𝒜(M)m\in\mathcal{A}(M) because 1mMM1mnMM=M\frac{1}{m}M\cap\mathbb{Z}M\subseteq\frac{1}{mn}M\cap\mathbb{Z}M=M. Likewise n𝒜(M)n\in\mathcal{A}(M). For the converse, assume that αM\alpha\in\mathbb{Z}M is such that mnαMmn\alpha\in M. Note that β=nαM\beta=n\alpha\in\mathbb{Z}M is such that mβMm\beta\in M, and thus βM\beta\in M because m𝒜(M)m\in\mathcal{A}(M). But then αM\alpha\in\mathbb{Z}M is such that nαMn\alpha\in M, and thus αM\alpha\in M because n𝒜(M)n\in\mathcal{A}(M). It follows that mn𝒜(M)mn\in\mathcal{A}(M). ∎

Now, we consider a set that records the pure translations of MM in 1mM\frac{1}{m}M. In Section 5 we see that this set corresponds to the free summands of 𝕜[1mM]\mathbb{k}\left[\frac{1}{m}M\right] as a 𝕜[M]\mathbb{k}[M]-module.

Definition 3.9.

Let m>1m\in\mathbb{Z}_{>1}. We set

Vm(M)={α1mM|(1mMα)MM}.V_{m}(M)=\left\{\alpha\in\frac{1}{m}M\;\bigg{|}\;\left(\frac{1}{m}M-\alpha\right)\cap\mathbb{Z}M\subseteq M\right\}.

Moreover, we set

V(M)=m>1Vm(M).V(M)=\bigcup_{m\in\mathbb{Z}_{>1}}V_{m}(M).
Remark 3.10.

We note that, because of Proposition 3.3, Vm(M)V_{m}(M) is precisely the set of α1mM\alpha\in\frac{1}{m}M such that (α+M)1mM(\alpha+M)\subseteq\frac{1}{m}M is pure.

Remark 3.11.

We note that Vm(M)V_{m}(M)\neq\emptyset if and only if m𝒜(M)m\in\mathcal{A}(M). As a consequence we have

V(M)=m𝒜(M)Vm(M).V(M)=\bigcup_{m\in\mathcal{A}(M)}V_{m}(M).

Indeed, if m𝒜(M)m\in\mathcal{A}(M) then 0Vm(M)0\in V_{m}(M). Conversely, fix αVm(M)\alpha\in V_{m}(M). From the containments 1mMM=((1mM+α)α)M(1mMα)MM\frac{1}{m}M\cap\mathbb{Z}M=\left(\left(\frac{1}{m}M+\alpha\right)-\alpha\right)\cap\mathbb{Z}M\subseteq\left(\frac{1}{m}M-\alpha\right)\cap\mathbb{Z}M\subseteq M, it follows that m𝒜(M)m\in\mathcal{A}(M).

Remark 3.12.

If MM is seminormal, then MV(M)={0}M\cap V(M)=\{0\}. Indeed, if 0γM0\neq\gamma\in M, then γ((1mMγ)M)M-\gamma\in\left(\left(\frac{1}{m}M-\gamma\right)\cap\mathbb{Z}M\right)\setminus M for every m𝒜(M)m\in\mathcal{A}(M).

In the following remarks we observe that Vm(M)V_{m}(M) is compatible with projections onto faces of C(M)C(M) and with isomorphisms of monoids.

Remark 3.13.

For every face FF of C(M)C(M) and m1m\in\mathbb{Z}_{\geqslant 1} we have Vm(M)FVm(MF)V_{m}(M)\cap F\subseteq V_{m}(M\cap F). Indeed, let αVm(M)F\alpha\in V_{m}(M)\cap F, then

(1m(MF)α)(MF)((1mMα)M)F=MF.\left(\frac{1}{m}(M\cap F)-\alpha\right)\cap\mathbb{Z}(M\cap F)\subseteq\left(\left(\frac{1}{m}M-\alpha\right)\cap\mathbb{Z}M\right)\cap F=M\cap F.
Remark 3.14.

Since every isomorphism of monoids φ:MM\varphi:M\to M^{\prime} extends to an isomorphism of groups φ:MM\varphi:\mathbb{Z}M\to\mathbb{Z}M^{\prime}, it follows that 𝒜(M)=𝒜(M)\mathcal{A}(M^{\prime})=\mathcal{A}(M). Furthermore, for every m𝒜(M)m\in\mathcal{A}(M) we have mVm(M)=φ(mVm(M))mV_{m}(M^{\prime})=\varphi(mV_{m}(M)).

We now describe basic properties Vm(M)V_{m}(M). In particular, we see that this set is finite.

Proposition 3.15.

Let m>1m\in\mathbb{Z}_{>1} and α,β,η1mM\alpha,\beta,\eta\in\frac{1}{m}M be such that α=β+η\alpha=\beta+\eta. If αVm(M)\alpha\in V_{m}(M), then β,ηVm(M)\beta,\eta\in V_{m}(M).

Proof.

Let w1mMw\in\frac{1}{m}M be such that wβMw-\beta\in\mathbb{Z}M. Thus, from

wβ=(w+η)α(1mMα)MMw-\beta=(w+\eta)-\alpha\in\left(\frac{1}{m}M-\alpha\right)\cap\mathbb{Z}M\subseteq M

it follows that wβMw-\beta\in M. This argument implies that βVm(M)\beta\in V_{m}(M). Likewise, ηVm(M)\eta\in V_{m}(M), finishing the proof. ∎

Corollary 3.16.

For every m>1m\in\mathbb{Z}_{>1} the set MmVm(M)M\setminus mV_{m}(M) is an ideal of MM.

Proof.

Let aMmVm(M)a\in M\setminus mV_{m}(M) and gMg\in M. Suppose a+gmVm(M)a+g\in mV_{m}(M), then a+gmVm(M)\frac{a+g}{m}\in V_{m}(M). By Proposition 3.15 this implies amVm(M)\frac{a}{m}\in V_{m}(M) which is a contradiction. ∎

The following lemma provides useful facts about the sets Vm(M)V_{m}(M). We recall that the Minkowski sum of two subsets A,BqA,B\subseteq\mathbb{R}^{q} is defined as AB={a+baA,bB}A\oplus B=\{a+b\mid a\in A,b\in B\}.

Lemma 3.17.

Let m,n𝒜(M)m,n\in\mathcal{A}(M). Then,

  1. (1)

    1mMVmn(M)=Vm(M)\frac{1}{m}M\cap V_{mn}(M)=V_{m}(M).

  2. (2)

    1nVm(M)Vn(M)Vmn(M)\frac{1}{n}V_{m}(M)\oplus V_{n}(M)\subseteq V_{mn}(M);

  3. (3)

    |1nVm(M)Vn(M)|=|Vm(M)||Vn(M)||Vmn(M)||\frac{1}{n}V_{m}(M)\oplus V_{n}(M)|=|V_{m}(M)|\cdot|V_{n}(M)|\leqslant|V_{mn}(M)|.

Proof.

We begin with the containment \supseteq in (1). Let αVm(M)\alpha\in V_{m}(M), then α1mM1mnM\alpha\in\frac{1}{m}M\subseteq\frac{1}{mn}M. Consider y(1mnMα)My\in\left(\frac{1}{mn}M-\alpha\right)\cap\mathbb{Z}M, then my+mα1nMM=Mmy+m\alpha\in\frac{1}{n}M\cap\mathbb{Z}M=M, Thus, y(1mMα)MMy\in\left(\frac{1}{m}M-\alpha\right)\cap\mathbb{Z}M\subseteq M and the conclusion follows. Now we prove the containment \subseteq. Let α1mMVmn(M)\alpha\in\frac{1}{m}M\cap V_{mn}(M) and let β1mM\beta\in\frac{1}{m}M be such that βαM\beta-\alpha\in\mathbb{Z}M. Then βα(1mMα)M(1mnMα)MM\beta-\alpha\in\left(\frac{1}{m}M-\alpha\right)\cap\mathbb{Z}M\subseteq\left(\frac{1}{mn}M-\alpha\right)\cap\mathbb{Z}M\subseteq M.

We continue with (2). Let αVm(M)\alpha\in V_{m}(M) and βVn(M)\beta\in V_{n}(M). Clearly we have 1nα+β1mnM\frac{1}{n}\alpha+\beta\in\frac{1}{mn}M. Let γM\gamma\in M and η1mnM\eta\in\frac{1}{mn}M be such that γ+η(1nα+β)M\gamma+\eta-(\frac{1}{n}\alpha+\beta)\in M, then nγ+nηαMn\gamma+n\eta-\alpha\in M. Since nη1mMn\eta\in\frac{1}{m}M, Proposition 3.3 applied to α\alpha implies nηαMn\eta-\alpha\in M, i.e, η1nα1nM\eta-\frac{1}{n}\alpha\in\frac{1}{n}M. Thus, Proposition 3.3 applied to β\beta implies η(1nα+β)=(η1nα)βM\eta-(\frac{1}{n}\alpha+\beta)=(\eta-\frac{1}{n}\alpha)-\beta\in M. Therefore, by Proposition 3.3 we have 1nα+βVmn(M)\frac{1}{n}\alpha+\beta\in V_{mn}(M).

We now show (3). Let α,βVm(M)\alpha,\beta\in V_{m}(M). We first show that (α+M)(β+M)=(\alpha+M)\cap(\beta+M)=\emptyset for αβ\alpha\neq\beta, we proceed by contradiction. Set W=(α+M)(β+M)W=(\alpha+M)\cap(\beta+M), and note that W(α+M)W\subseteq(\alpha+M) splits since

(α+M)W=(α+M)(1mM(β+M))(\alpha+M)\setminus W=(\alpha+M)\cap\left(\frac{1}{m}M\setminus(\beta+M)\right)

is an MM-module (see Remark 3.10). Let T=(α+M)WT=(\alpha+M)\setminus W. If γT\gamma\in T and zWαMz\in W-\alpha\subseteq M, then γ+zT\gamma+z\in T but also γ+zγ+(Wα)W+M=W\gamma+z\in\gamma+(W-\alpha)\subseteq W+M=W, which is not possible. We conclude T=T=\emptyset, and so W=(α+M)W=(\alpha+M). Thus, α+Mβ+M\alpha+M\subseteq\beta+M. By symmetry, β+Mα+M\beta+M\subseteq\alpha+M, an then β+M=α+M\beta+M=\alpha+M. It follows that αβMVm(M)={0}\alpha-\beta\in M\cap V_{m}(M)=\{0\} by Proposition 3.15 and Remark 3.12, which is a contradiction. Therefore, the union αVm(M)(α+M)1mM\bigcup_{\alpha\in V_{m}(M)}(\alpha+M)\subseteq\frac{1}{m}M is disjoint, and so, αVm(M)(αn+1nM)1mnM\bigcup_{\alpha\in V_{m}(M)}\left(\frac{\alpha}{n}+\frac{1}{n}M\right)\subset\frac{1}{mn}M is also disjoint. By applying the same argument, the union

αVm(M)βVn(M)(αn+β+M)1mnM\bigcup_{\alpha\in V_{m}(M)}\bigcup_{\beta\in V_{n}(M)}\left(\frac{\alpha}{n}+\beta+M\right)\subset\frac{1}{mn}M

is disjoint. Hence,

1nVm(M)Vn(M)={αn+βαVm(M),βVn(M)}1mnM\frac{1}{n}V_{m}(M)\oplus V_{n}(M)=\left\{\frac{\alpha}{n}+\beta\mid{\alpha\in V_{m}(M)},{\beta\in V_{n}(M)}\right\}\subseteq\frac{1}{mn}M

is a set of of cardinality |Vm(M)||Vn(M)|.|V_{m}(M)|\cdot|V_{n}(M)|. Finally, the inequality follows from Part (2). ∎

Proposition 3.18.

Let m𝒜(M)m\in\mathcal{A}(M) and α1mM\alpha\in\frac{1}{m}M. Write α=c1mγ1++cumγu\alpha=\frac{c_{1}}{m}\gamma_{1}+\cdots+\frac{c_{u}}{m}\gamma_{u} with c1,,cu0c_{1},\ldots,c_{u}\in\mathbb{Z}_{\geqslant 0}. If c1++cu(m1)u+1c_{1}+\ldots+c_{u}\geqslant(m-1)u+1, then αVm(M)\alpha\not\in V_{m}(M). Furthermore, |Vm(M)|mrank(M)|V_{m}(M)|\leqslant m^{\operatorname{rank}(M)}.

Proof.

By assumption we have that αγi+1mM\alpha\in\gamma_{i}+\frac{1}{m}M for some 1iu1\leqslant i\leqslant u. By way of contradiction suppose that αVm(M)\alpha\in V_{m}(M). Since αγi1mM\alpha-\gamma_{i}\in\frac{1}{m}M, it follows by Proposition 3.15 that γiVm(M)\gamma_{i}\in V_{m}(M). However, this contradicts Remark 3.12, and therefore αVm(M)\alpha\notin V_{m}(M).

For the second claim, recall that αVm(M)(α+M)\bigcup_{\alpha\in V_{m}(M)}(\alpha+M) is a disjoint union of MM-modules (see proof of Lemma 3.17 (3)), and thus αVm(M)(α+M)1mM\bigcup_{\alpha\in V_{m}(M)}(\alpha+M)\subseteq\frac{1}{m}M is pure. As a consequence, the M\mathbb{Z}M-module (1mM)\mathbb{Z}\left(\frac{1}{m}M\right) contains αVm(M)(α+M)\bigoplus_{\alpha\in V_{m}(M)}\mathbb{Z}(\alpha+M) as a free direct summand, and thus |Vm(M)|mrank(M)|V_{m}(M)|\leqslant m^{\operatorname{rank}(M)}, where the latter is the rank of (1mM)\mathbb{Z}\left(\frac{1}{m}M\right) as a M\mathbb{Z}M-module. ∎

We now define a new numerical invariant for seminormal monoid. This number plays an important role in our main results. This invariant is inspired by the FF-pure threshold of a ring [TW2004]. This is because the FF-pure threshold of a standard graded algebra can be described as the supremum among the degrees of a minimal generator of a free summand of R1/peR^{1/p^{e}} [DSNB]. However, the FF-pure threshold of R=𝕜[M]R=\mathbb{k}[M] can be different than the pure threshold of MM (see Example 3.23 and Remark 3.24). In Proposition 4.4 we prove that for normal monoids this invariant is rational.

Definition 3.19.

We define the pure threshold of MM as

mpt(M)=sup{|α||αV(M)}.{\operatorname{mpt}}(M)=\sup\{|\alpha|\;|\;\alpha\in V(M)\}.

If V(M)=V(M)=\emptyset, i.e., if MM is not seminormal, we set mpt(M)={\operatorname{mpt}}(M)=-\infty.

Remark 3.20.

Let b=max{|γ1|,,|γu|}b=\max\{|\gamma_{1}|,\ldots,|\gamma_{u}|\}. By Proposition 3.18 we have |α|<bu|\alpha|<bu for every αV(M)\alpha\in V(M). Therefore, mpt(M)<{\operatorname{mpt}}(M)<\infty.

We now discuss how the pure threshold of a monoid MM can be obtained from any increasing sequence in 𝒜(M)\mathcal{A}(M).

Proposition 3.21.

Let {mt}t1\{m_{t}\}_{t\in\mathbb{Z}_{\geqslant 1}} be the elements of 𝒜(M)\mathcal{A}(M) ordered increasingly. Then,

limtmax{|α||αVmt(M)}=mpt(M),\lim\limits_{t\to\infty}\max\{|\alpha|\;|\;\alpha\in V_{m_{t}}(M)\}={\operatorname{mpt}}(M),

In particular,

limtmax{|α||αVmt(M)}=mpt(M)\lim\limits_{t\to\infty}\max\{|\alpha|\;|\;\alpha\in V_{m^{t}}(M)\}={\operatorname{mpt}}(M)

for any m𝒜(M)m\in\mathcal{A}(M).

Proof.

If mpt(M)=0{\operatorname{mpt}}(M)=0, the result follows. We assume mpt(M)>0{\operatorname{mpt}}(M)>0. Let b=max{|γ1|,,|γu|}b=\max\{|\gamma_{1}|,\ldots,|\gamma_{u}|\} and for any n𝒜(M)n\in\mathcal{A}(M) set an=max{|α||αVn(M)}a_{n}=\max\{|\alpha|\;|\;\alpha\in V_{n}(M)\}. Fix ϵ>0\epsilon>0 and N0N\in\mathbb{Z}_{\geqslant 0} such that bumN<ϵ2\frac{bu}{m_{N}}<\frac{\epsilon}{2}. Let m𝒜(M)m^{\prime}\in\mathcal{A}(M) be such that bm>mpt(M)ϵ2b_{m^{\prime}}>{\operatorname{mpt}}(M)-\frac{\epsilon}{2} and fix tNt\geqslant N. By Lemma 3.17 we have bmtm>mpt(M)ϵ2b_{m_{t}m^{\prime}}>{\operatorname{mpt}}(M)-\frac{\epsilon}{2}. Consider α=c1mtmγ1++cumtmγuVmtm(M)\alpha=\frac{c_{1}}{m_{t}m^{\prime}}\gamma_{1}+\cdots+\frac{c_{u}}{m_{t}m^{\prime}}\gamma_{u}\in V_{m_{t}m^{\prime}}(M) with ci0c_{i}\in\mathbb{Z}_{\geqslant 0} and |α|=bmtm|\alpha|=b_{m_{t}m^{\prime}}. For each 1iu1\leqslant i\leqslant u let 0ri<m0\leqslant r_{i}<m^{\prime} be such that ciri(modm).c_{i}\equiv r_{i}\pmod{m^{\prime}}. By Proposition 3.15 and Lemma 3.17 (1) we have

α:=c1r1mtmγ1++curumtmγu1mtMVmtm(M)=Vmt(M).\alpha^{\prime}:=\frac{c_{1}-r_{1}}{m_{t}m^{\prime}}\gamma_{1}+\cdots+\frac{c_{u}-r_{u}}{m_{t}m^{\prime}}\gamma_{u}\in\frac{1}{m_{t}}M\cap V_{m_{t}m^{\prime}}(M)=V_{m_{t}}(M).

Moreover,

mpt(M)|α|=mpt(M)|α|+(|α||α|)<ϵ2+bumt<ϵ.{\operatorname{mpt}}(M)-|\alpha^{\prime}|={\operatorname{mpt}}(M)-|\alpha|+(|\alpha|-|\alpha^{\prime}|)<\frac{\epsilon}{2}+\frac{bu}{m_{t}}<\epsilon.

Since ϵ\epsilon was chosen arbitrarily, the result follows. ∎

We now compute some examples of pure thresholds. We note that this invariant depends on the grading given by the embedding Mq.M\subseteq\mathbb{Z}^{q}.

Example 3.22.

Let MM be generated by d1e1,,dqeqqd_{1}e_{1},\ldots,d_{q}e_{q}\in\mathbb{Z}^{q}, where di>0d_{i}\in\mathbb{Z}_{>0} and {e1,,eq}\{e_{1},\ldots,e_{q}\} is the canonical basis in q\mathbb{Z}^{q}. Then, Vm(M)={(d1α1m,,dqαqm)1mq| 0αim1}V_{m}(M)=\{(d_{1}\frac{\alpha_{1}}{m},\ldots,d_{q}\frac{\alpha_{q}}{m})\in\frac{1}{m}\mathbb{Z}^{q}\;|\;0\leqslant\alpha_{i}\leqslant m-1\}, and so, mpt(M)=d1++dq{\operatorname{mpt}}(M)=d_{1}+\ldots+d_{q}.

Example 3.23.

Let q>1q\in\mathbb{Z}_{>1}, t>0t\in\mathbb{Z}_{>0} and M={α0q||α|t>0}M=\{\alpha\in\mathbb{Z}^{q}_{\geqslant 0}\;|\;|\alpha|\in t\mathbb{Z}_{>0}\}. Then 𝕜[M]\mathbb{k}[M] is the Veronese subring of order tt of a polynomial ring 𝕜[x1,,xq]\mathbb{k}[x_{1},\ldots,x_{q}] with the grading deg(xi)=1\deg(x_{i})=1. We have that

Vm(M)={(α1m,,αqm)1mq| 0αim1 and |α|t>0},V_{m}(M)=\left\{\left(\frac{\alpha_{1}}{m},\ldots,\frac{\alpha_{q}}{m}\right)\in\frac{1}{m}\mathbb{Z}^{q}\;|\;0\leqslant\alpha_{i}\leqslant m-1\hbox{ and }|\alpha|\in t\mathbb{Z}_{>0}\right\},

and therefore mpt(M)=q{\operatorname{mpt}}(M)=q. We point out that, if 𝕜\mathbb{k} has prime characteristic, then fpt(𝕜[M])=qt{\operatorname{fpt}}(\mathbb{k}[M])=\frac{q}{t} [HWY, Example 6.1].

Remark 3.24.

It follows from Example 3.23 that mpt(M){\operatorname{mpt}}(M) may differ from fpt(𝕜[M]){\operatorname{fpt}}(\mathbb{k}[M]) even when MM is normal. This is not surprising since fpt(𝕜[M]){\operatorname{fpt}}(\mathbb{k}[M]) is independent of the presentation of 𝕜[M]\mathbb{k}[M] as a quotient of a polynomial ring, while we have already observed that mpt(M){\operatorname{mpt}}(M) heavily depends on the degrees of the generators and on the embedding of MM.

The following construction allows us to provide bounds for depths of affine semigroup rings (see Section 5). In Proposition 3.26 we justify the terminology used in the definition.

Definition 3.25.

We define the pure prime of MM by

𝒫(M)=Mm>1mVm(M).\mathcal{P}(M)=M\setminus\displaystyle\bigcup_{m\in\mathbb{Z}_{>1}}mV_{m}(M).
Proposition 3.26.

Let MM be an affine monoid. Then 𝒫(M)\mathcal{P}(M) is a prime ideal of MM.

Proof.

Since 𝒫(M)=mm>1(MmVm(M))\mathcal{P}(M)=\bigcap_{m\in\mathbb{Z}_{m>1}}\left(M\setminus mV_{m}(M)\right), it follows from Corollary 3.16 that 𝒫(M)\mathcal{P}(M) is an ideal of MM. Now, let a,bM𝒫(M)a,b\in M\setminus\mathcal{P}(M) and m,n𝒜(M)m,n\in\mathcal{A}(M) be such that α:=amVm(M)\alpha:=\frac{a}{m}\in V_{m}(M) and β:=bmVn(M)\beta:=\frac{b}{m}\in V_{n}(M). We claim that a+bmnVmn(M)\frac{a+b}{mn}\in V_{mn}(M) which implies a+b𝒫(M)a+b\not\in\mathcal{P}(M), finishing the proof. Indeed, suppose a+bmnVmn(M)\frac{a+b}{mn}\not\in V_{mn}(M), then from Proposition 3.15 it follows that

1nα+β=amn+mbmn=a+bmn+(m1)bmnVmn(M),\frac{1}{n}\alpha+\beta=\frac{a}{mn}+\frac{mb}{mn}=\frac{a+b}{mn}+\frac{(m-1)b}{mn}\not\in V_{mn}(M),

which contradicts Lemma 3.17 (2). ∎

We obtain the following theorem that relates 𝒫(M)\mathcal{P}(M) with the normality of MM.

Theorem 3.27.

Let MM be an affine monoid. Then MM is normal if and only 𝒫(M)=.\mathcal{P}(M)=\emptyset.

Proof.

We begin with the forward direction. Since MM is normal, by Remark 3.14 we can assume that MM is a submonoid of 0n\mathbb{Z}_{\geqslant 0}^{n} for some n>0n\in\mathbb{Z}_{>0} and such that M=M0nM=\mathbb{Z}M\cap\mathbb{Z}_{\geqslant 0}^{n} [bruns2009polytopes, Theorem 2.29]. Fix aM0na\in M\subseteq\mathbb{Z}_{\geqslant 0}^{n} and chose m>1m\in\mathbb{Z}_{>1} bigger than every entry in aa. By Remark 3.7 we have m𝒜(M)m\in\mathcal{A}(M) and m𝒜(0n)m\in\mathcal{A}(\mathbb{Z}_{\geqslant 0}^{n}). By the choice of mm, it is clear that

(3.1) (1m0nam)n0n.\left(\frac{1}{m}\mathbb{Z}_{\geqslant 0}^{n}-\frac{a}{m}\right)\cap\mathbb{Z}^{n}\subseteq\mathbb{Z}_{\geqslant 0}^{n}.

Thus, the left hand side expression in (3.1) is equal to (1m0nam)0n\left(\frac{1}{m}\mathbb{Z}_{\geqslant 0}^{n}-\frac{a}{m}\right)\cap\mathbb{Z}_{\geqslant 0}^{n}. Intersecting this with M\mathbb{Z}M we obtain,

(1m0nam)0nM\displaystyle\left(\frac{1}{m}\mathbb{Z}_{\geqslant 0}^{n}-\frac{a}{m}\right)\cap\mathbb{Z}_{\geqslant 0}^{n}\cap\mathbb{Z}M =(1m0nam)M=(1mMam)M.\displaystyle=\left(\frac{1}{m}\mathbb{Z}_{\geqslant 0}^{n}-\frac{a}{m}\right)\cap M=\left(\frac{1}{m}M-\frac{a}{m}\right)\cap M.

Therefore, (1mMam)M0nM=M,\left(\frac{1}{m}M-\frac{a}{m}\right)\cap M\subseteq\mathbb{Z}_{\geqslant 0}^{n}\cap\mathbb{Z}M=M, which shows amVm(M)a\in mV_{m}(M).

We continue with the backward direction. Let fMf\in\mathbb{Z}M be such that nfMnf\in M for some n>1n\in\mathbb{Z}_{>1}. By Proposition 3.5 it suffices to show fMf\in M. Write f=abf=a-b with a,bMa,b\in M, then nanb+Mna\in nb+M. Thus, n(a,b)=b+(n1)(a,b)n(a,b)=b+(n-1)(a,b), where (a,b)(a,b) denotes the ideal of MM generated by the set {a,b}\{a,b\}. It follows that (n+r)(a,b)=rb+n(a,b)(n+r)(a,b)=rb+n(a,b) for every r>0r\in\mathbb{Z}_{>0}. Hence,

ra+nb(n+r)(a,b)=rb+n(a,b)rb+M for every r>0.ra+nb\in(n+r)(a,b)=rb+n(a,b)\subseteq rb+M\quad\text{ for every }r\in\mathbb{Z}_{>0}.

By assumption there exists m𝒜(M)m\in\mathcal{A}(M) such that nbmVm(M)nb\in mV_{m}(M). Therefore,

f=ab=(ma+nbm)nbmb(b+1mM)nbmb=1mMnbm.f=a-b=\left(\frac{ma+nb}{m}\right)-\frac{nb}{m}-b\in\left(b+\frac{1}{m}M\right)-\frac{nb}{m}-b=\frac{1}{m}M-\frac{nb}{m}.

On the other hand, fMf\in\mathbb{Z}M, then f(1mMnbm)M=Mf\in\left(\frac{1}{m}M-\frac{nb}{m}\right)\cap\mathbb{Z}M=M, which finishes the proof. ∎

Corollary 3.28.

There exists a face M\mathcal{F}_{M} of C(M)C(M) such that M𝒫(M)=MMM\setminus\mathcal{P}(M)=M\cap\mathcal{F}_{M}. Moreover, the monoid MMM\cap\mathcal{F}_{M} is normal.

Proof.

The first part follows from Proposition 3.26 and the correspondence between prime ideals of monoids and faces of their cones [bruns2009polytopes, Proposition 2.36]. By Theorem 3.27 to show that MMM\cap\mathcal{F}_{M} is normal it suffices to show MM=m>1mVm(MM).M\cap\mathcal{F}_{M}=\bigcup_{m\in\mathbb{Z}_{>1}}mV_{m}(M\cap\mathcal{F}_{M}). Moreover, we may assume MM is seminormal. We note that

MM=m>1mVm(M)m>1mVm(MM),M\cap\mathcal{F}_{M}=\bigcup_{m\in\mathbb{Z}_{>1}}mV_{m}(M)\subseteq\bigcup_{m\in\mathbb{Z}_{>1}}mV_{m}(M\cap\mathcal{F}_{M}),

where the last inclusion follows from Remark 3.13. Since we always have the other inclusion m>1mVm(MM)MM\bigcup_{m\in\mathbb{Z}_{>1}}mV_{m}(M\cap\mathcal{F}_{M})\subseteq M\cap\mathcal{F}_{M}, the proof is complete. ∎

The previous proposition allows us to define the following invariant of affine monoids, the pure dimension. As we see in Corollary 3.30, this new notion measures how far a monoid is from being normal. In Theorem 5.7 we use this invariant to provide lower bounds for the depth of affine semigoup rings.

Definition 3.29.

The face M\mathcal{F}_{M} in Proposition 3.28 is called the pure prime face of MM. We define the pure dimension of MM by mpdim(M):=rank(MM).{\operatorname{mpdim}}(M):=\operatorname{rank}(M\cap\mathcal{F}_{M}). If M=\mathcal{F}_{M}=\emptyset, we set mpdim(M)={\operatorname{mpdim}}(M)=-\infty.

Corollary 3.30.

Let MM be an affine monoid. Then

  1. (1)

    mpdim(M)rank(M){\operatorname{mpdim}}(M)\leqslant\operatorname{rank}(M).

  2. (2)

    mpdim(M)0{\operatorname{mpdim}}(M)\geqslant 0 if and only if MM is seminormal.

  3. (3)

    mpdim(M)=rank(M){\operatorname{mpdim}}(M)=\operatorname{rank}(M) if and only if MM is normal.

Proof.

Part (1) follows directly from the definition. For part (2) we note that mpdim(M)0{\operatorname{mpdim}}(M)\geqslant 0 if and only if m>1mVm(M)\displaystyle\bigcup_{m\in\mathbb{Z}_{>1}}mV_{m}(M)\neq\emptyset, and by Remarks 3.27 and 3.11 this is equivalent to MM being seminormal. Part (3) follows from Theorem 3.27. ∎

We finish this section with the following example.

Example 3.31.

Let M=0(2,0)+0(1,1)+0(0,1)M=\mathbb{Z}_{\geqslant 0}(2,0)+\mathbb{Z}_{\geqslant 0}(1,1)+\mathbb{Z}_{\geqslant 0}(0,1) be the monoid with set of generators {(2,0),(1,1),(0,1)}\{(2,0),(1,1),(0,1)\}. It is easy to see that M=02{(2a+1,0)a0}M=\mathbb{Z}_{\geqslant 0}^{2}\setminus\{(2a+1,0)\mid a\in\mathbb{Z}_{\geqslant 0}\} (see Figure 2). We also have that 𝒜(M)={m>1m is odd}\mathcal{A}(M)=\{m\in\mathbb{Z}_{>1}\mid m\text{ is odd}\}. Then, for every m𝒜(M)m\in\mathcal{A}(M) we have mVm(M)={(a,0)a is even and a<m}mV_{m}(M)=\{(a,0)\mid a\text{ is even and }a<m\}. Therefore, mpt(M)=1{\operatorname{mpt}}(M)=1, M=0(1,0)\mathcal{F}_{M}=\mathbb{R}_{\geqslant 0}(1,0), and mpdim(M)=1{\operatorname{mpdim}}(M)=1. In particular, by Corollary 3.30, MM is seminormal but not normal; this is consistent with [Li, Example 1.0.3].

Refer to caption
Figure 2. The monoid M=0(2,0)+0(1,1)+0(0,1)M=\mathbb{Z}_{\geqslant 0}(2,0)+\mathbb{Z}_{\geqslant 0}(1,1)+\mathbb{Z}_{\geqslant 0}(0,1).

4. Asymptotic growth of number of pure translations

In the short section, we study the asymptotic behavior of the number of elements in the sets Vm(M)V_{m}(M). Throughout we adopt the same notation from Section 3.

Definition 4.1.

Let MM be a seminormal affine monoid, and let M\mathcal{F}_{M} be its pure prime face. For every m𝒜(M)m\in\mathcal{A}(M) we define

Bm(M)=αVm(M)((αM)M).B_{m}(M)=\bigcup_{\alpha\in V_{m}(M)}\left((\alpha-\mathcal{F}_{M})\cap\mathcal{F}_{M}\right).

Moreover, we set

B(M)=m𝒜(M)Bm(M).B(M)=\bigcup_{m\in\mathcal{A}(M)}B_{m}(M).

When MM is normal, there is a simple description of B(M)B(M) as the region in Notation 4.2. We prove that these regions coincide in Lemma 4.3, which also includes important properties of B(M)B(M).

Notation 4.2.

Let MM be a normal affine monoid. Let {H1,,Hs}\{H_{1},\ldots,H_{s}\} be the supporting hyperplanes of C(M)C(M) so that C(M)=H1+Hs+.C(M)=H_{1}^{+}\cap\cdots\cap H_{s}^{+}. Let viqv_{i}\in\mathbb{Q}^{q} be rational vectors such that xHi+x\in H_{i}^{+} if and only if x,vi0\langle x,v_{i}\rangle\geqslant 0 for 1is1\leqslant i\leqslant s. We can further assume that x,vi\langle x,v_{i}\rangle\in\mathbb{Z} for every xMx\in\mathbb{Z}M and that min{α,viαM}=1\min\{\langle\alpha,v_{i}\rangle\mid\alpha\in M\}=1 [bruns2009polytopes, Remark 1.72]. We define Δ\Delta by

Δ={xq0x,vi<1 for 1is}.\Delta=\{x\in\mathbb{R}^{q}\mid 0\leqslant\langle x,v_{i}\rangle<1\text{ for }1\leqslant i\leqslant s\}.
Lemma 4.3.

Let MM be a seminormal affine monoid. Then

  1. (1)

    B(M)B(M) is a bounded set and it has volume, i.e., its boundary has measure zero in the dim(M)\dim(\mathbb{R}\mathcal{F}_{M})-dimensional Lebesgue measure on M\mathbb{R}\mathcal{F}_{M}.

  2. (2)

    There exists an increasing sequence {pt}t1𝒜(M)\{p_{t}\}_{t\in\mathbb{Z}_{\geqslant 1}}\subseteq\mathcal{A}(M) such that BptBpt+1B_{p_{t}}\subseteq B_{p_{t+1}} for every t1t\in\mathbb{Z}_{\geqslant 1} and B(M)=t1Bpt(M).B(M)=\bigcup_{t\in\mathbb{Z}_{\geqslant 1}}B_{p_{t}}(M).

  3. (3)

    Vm(M)=1mMB(M)V_{m}(M)=\frac{1}{m}M\cap B(M) for every m𝒜(M)m\in\mathcal{A}(M).

  4. (4)

    If MM is normal and Δ\Delta is as in Notation 4.2, then B(M)=ΔB(M)=\Delta.

Proof.

We begin with (1). Let {g1,,gl}\{g_{1},\ldots,g_{l}\} be a minimal set of generators of MMM\cap\mathcal{F}_{M} and consider the region Γ=i=1l(gi+M)\Gamma=\cup_{i=1}^{l}(g_{i}+\mathcal{F}_{M}). Let α1mMΓ\alpha\in\frac{1}{m}M\cap\Gamma, then α=gi+η\alpha=g_{i}+\eta for some ii and ηM\eta\in\mathcal{F}_{M}. Since MMM\cap\mathcal{F}_{M} is normal by Corollary 3.28, it follows that ηm(MM)MmM\eta\in\frac{\mathbb{Z}}{m}(M\cap\mathcal{F}_{M})\cap\mathcal{F}_{M}\subseteq\frac{\mathbb{Z}}{m}M. If αVm(M)\alpha\in V_{m}(M), then Proposition 3.15 implies giVm(M)g_{i}\in V_{m}(M) which contradicts Remark 3.12. We conclude Vm(M)V_{m}(M), and then Bm(M)B_{m}(M), is contained in MΓ\mathcal{F}_{M}\setminus\Gamma which is bounded.

Now, let \partial and denote boundary and interior on M\mathbb{R}\mathcal{F}_{M}, respectively. Let μ\mu denote the dim(M)\dim(\mathbb{R}\mathcal{F}_{M})-dimensional Lebesgue measure on M\mathbb{R}\mathcal{F}_{M}. We note that for any xB(M)Mx\in\partial B(M)\setminus\partial\mathcal{F}_{M} we have x+MMB(M)¯x+\mathcal{F}_{M}^{\circ}\subseteq\mathcal{F}_{M}\setminus\overline{B(M)}; indeed, if x+yB(M)¯x+y\in\overline{B(M)} for some yMy\in\mathcal{F}_{M}^{\circ}, then x+yB(M)x+y^{\prime}\in B(M) for some yMy^{\prime}\in\mathcal{F}_{M}^{\circ}, which would imply xB(M).x\in B(M)^{\circ}. Therefore, for any r>0r>0 and any xB(M)Mx\in\partial B(M)\setminus\partial\mathcal{F}_{M} we have B(M)B(r,x)(x+M)=\partial B(M)\cap B(r,x)\cap(x+\mathcal{F}_{M}^{\circ})=\emptyset, where B(r,x)B(r,x) denotes the ball in M\mathbb{R}\mathcal{F}_{M} with radius rr and center xx. Therefore, there exists a real c<1c<1 such that for any such rr and xx we have μ(B(M)B(r,x))μ(B(r,x))<c.\frac{\mu(\partial B(M)\cap B(r,x))}{\mu(B(r,x))}<c. By Lebesgue’s density theorem [mattila, Corollary 2.14], we conclude μ(B(M))=0\mu(\partial B(M))=0.

We continue with (2). Let {mt}t1\{m_{t}\}_{t\in\mathbb{Z}_{\geqslant 1}} be the elements of 𝒜(M)\mathcal{A}(M) ordered increasingly. For each t1t\in\mathbb{Z}_{\geqslant 1} set pt=m1mtp_{t}=m_{1}\cdots m_{t} and notice pt𝒜(M)p_{t}\in\mathcal{A}(M) by Lemma 3.8. The conclusion now follows from Lemma 3.12 (1).

Now we prove (3). Let mt𝒜(M)m_{t}\in\mathcal{A}(M) and α1mtMB(M)\alpha\in\frac{1}{m_{t}}M\cap B(M), it suffices to show αVmt(M)\alpha\in V_{m_{t}}(M). By (2), we have αBpi(M)\alpha\in B_{p_{i}}(M) for some ii. We may assume iti\geqslant t and then mtm_{t} divides pip_{i}. Therefore, there exists ηM\eta\in\mathcal{F}_{M} and βVpi(M)\beta\in V_{p_{i}}(M) such that α+η=β\alpha+\eta=\beta. Since MMM\cap\mathcal{F}_{M} is normal by Corollary 3.28, it follows that ηpi(MM)MpiM\eta\in\frac{\mathbb{Z}}{p_{i}}(M\cap\mathcal{F}_{M})\cap\mathcal{F}_{M}\subseteq\frac{\mathbb{Z}}{p_{i}}M. Thus, α1mtMVpi(M)=Vmt(M)\alpha\in\frac{1}{m_{t}}M\cap V_{p_{i}}(M)=V_{m_{t}}(M) by Proposition 3.15 and Lemma 3.17, which finishes the proof.

We finish with (4). If MM is normal we have Vm(M)=mMΔV_{m}(M)=\frac{\mathbb{Z}}{m}M\cap\Delta [von2011f, Lemma 3.11]. Thus, the equality B(M)=ΔB(M)=\Delta follows as the set m1mM\cup_{m\in\mathbb{Z}_{\geqslant 1}}\frac{\mathbb{Z}}{m}M is dense in q\mathbb{R}^{q}. ∎

From Lemma 4.3 (4) we obtain that the pure threshold of normal monoids is rational.

Proposition 4.4.

If MM is normal, then mpt(M)>0{\operatorname{mpt}}(M)\in\mathbb{Q}_{>0}.

Proof.

The statement follows readily from Lemma 4.3 (4) and the equality mpt(M)=sup{|α|αB(M)}{\operatorname{mpt}}(M)=\sup\{|\alpha|\mid\alpha\in B(M)\}. ∎

We now turn our focus to asymptotic growth of the number of elements in the sets Vm(M)V_{m}(M). We define the following limit, which we prove exists in Theorem 4.6

Definition 4.5.

Let MM be a seminormal affine monoid. Set s=mpdim(M)s={\operatorname{mpdim}}(M) and let {mt}t1\{m_{t}\}_{t\in\mathbb{Z}_{\geqslant 1}} be the elements of 𝒜(M)\mathcal{A}(M) ordered increasingly. We define the pure ratio of MM as

mpr(M)=limt|Vmt(M)|mts.\operatorname{mpr}(M)=\lim\limits_{t\to\infty}\frac{|V_{m_{t}}(M)|}{m_{t}^{s}}.

We define the pure signature of MM as

mps(M)=limt|Vmt(M)|mtrank(M).\operatorname{mps}(M)=\lim\limits_{t\to\infty}\frac{|V_{m_{t}}(M)|}{m_{t}^{\operatorname{rank}(M)}}.

In the following theorem we show that mpr(M)\operatorname{mpr}(M) exists as a limit, and that it equals the relative volume of B(M)B(M). Here, by relative volume with respect to a lattice LHL\subseteq H of rank rr in an rr-dimensional hyperplane HqH\subseteq\mathbb{R}^{q}, denoted by volL\operatorname{vol}_{L}, we mean the rr-dimensional volume in HH normalized such that any fundamental domain of LL has volume one.

Theorem 4.6.

Let MM be a seminormal affine monoid. We have that

mpr(M)=vol(MM)(B(M))>0.\operatorname{mpr}(M)=\operatorname{vol}_{\mathbb{Z}(M\cap\mathcal{F}_{M})}(B(M))>0.

In particular, MM is normal if and only if mps(M)>0\operatorname{mps}(M)>0. Furthermore, in this case mps(M)>0\operatorname{mps}(M)\in\mathbb{Q}_{>0}.

Proof.

By Lemma 4.3 (1), the characteristic function χB(M)\chi_{B(M)} is Riemann integrable. Now, by Lemma 4.3 (3) and Corollary 3.28 we have Vmt(M)=mt(MM)B(M)V_{m_{t}}(M)=\frac{\mathbb{Z}}{m_{t}}(M\cap\mathcal{F}_{M})\cap B(M). Thus, |Vmt(M)|mts\frac{|V_{m_{t}}(M)|}{m_{t}^{s}} is a Riemann sum for χB(M)\chi_{B(M)} with normalized volumes of the cells and mesh the diameter of a fundamental domain for (MM)mt\frac{\mathbb{Z}(M\cap\mathcal{F}_{M})}{m_{t}}. Therefore, by taking the limit tt\to\infty we obtain that the limit exists and is equal to vol(MM)(B(M))\operatorname{vol}_{\mathbb{Z}(M\cap\mathcal{F}_{M})}(B(M)). We note that vol(MM)(B(M))\operatorname{vol}_{\mathbb{Z}(M\cap\mathcal{F}_{M})}(B(M)) is positive since Vm(M)V_{m}(M) has interior points of M\mathcal{F}_{M} (see Corollary 3.28). The last statements follow from Corollary 3.30 and Lemma 4.3 (4). ∎

Theorem 4.6 is related to previous computations done for the FF-signature of normal semigroup rings [ToricSingh, von2011f].

Example 4.7.

Let MM be as in Example 3.31. We observe that |Vm(M)|=m2|V_{m}(M)|=\lceil\frac{m}{2}\rceil for every m𝒜(M)m\in\mathcal{A}(M). Then, mpr(M)=12\operatorname{mpr}(M)=\frac{1}{2}. We also have B(M)=[0,1)B(M)=[0,1), therefore vol(MM)(B(M))=12\operatorname{vol}_{\mathbb{Z}(M\cap\mathcal{F}_{M})}(B(M))=\frac{1}{2}, which is consistent with Theorem 4.6.

We end this section with a question motivated by Proposition 4.4. This question is open, to the best of our knowledge, for seminormal monoids that are not normal.

Question 4.8.

Let MM be a seminormal affine monoid. Is mpr(M)\operatorname{mpr}(M) a rational number?

5. Applications to affine semigroup rings

Throughout this section we adopt the following notation.

Notation 5.1.

Given an affine monoid as in Notation 1.1, we let R=𝕜[M]=K[𝐱ααM]𝕜[𝐱]:=𝕜[x1,xq]R=\mathbb{k}[M]=K[\mathbf{x}^{\mathbf{\alpha}}\mid\mathbf{\alpha}\in M]\subseteq\mathbb{k}[\mathbf{x}]:=\mathbb{k}[x_{1}\ldots,x_{q}] be the affine semigroup ring of MM. Given m>0m\in\mathbb{Z}_{>0}, we set R1/m=𝕜[1mM]R^{1/m}=\mathbb{k}\left[\frac{1}{m}M\right] the 𝕜\mathbb{k}-algebra 𝕜[xα|α1mM]\mathbb{k}\left[x^{\mathbf{\alpha}}|\mathbf{\alpha}\in\frac{1}{m}M\right]. Given m>0m\in\mathbb{Z}_{>0} and α1mM\alpha\in\frac{1}{m}M, we consider ϕαm:R1/mR\phi_{\alpha}^{m}:R^{1/m}\to R the 𝕜\mathbb{k}-linear map given by ϕαm(𝐱β)=𝐱βα\phi_{\alpha}^{m}(\mathbf{x}^{\beta})=\mathbf{x}^{\beta-\alpha} if βαM\beta-\alpha\in M and zero otherwise. For an ideal IMI\subseteq M, we denote by 𝐱I\mathbf{x}^{I} the corresponding MM-homogeneous RR-ideal,

𝐱I=(𝐱ααI).\mathbf{x}^{I}=(\mathbf{x}^{\alpha}\mid\alpha\in I).

For an MM-homogeneous element f=𝐱αRf=\mathbf{x}^{\alpha}\in R, we denote by log(f)=αM\log(f)=\alpha\in M the corresponding element in MM.

Remark 5.2.

We note that RR1/mR\cong R^{1/m} via the 𝕜\mathbb{k}-algebra map given by 𝐱α𝐱α/m\mathbf{x}^{\alpha}\mapsto\mathbf{x}^{\alpha/m}.

Proposition 5.3.

Let MM be an affine monoid, and let α1mM\alpha\in\frac{1}{m}M. Then, ϕαm\phi_{\alpha}^{m} is a map of RR-modules if and only if αVm(M)\alpha\in V_{m}(M).

Proof.

We note that ϕαm\phi_{\alpha}^{m} is a map of RR-modules if and only if for every γM\gamma\in M and β1mM\beta\in\frac{1}{m}M we have ϕαm(𝐱γ𝐱β)=𝐱γϕαm(𝐱β)\phi_{\alpha}^{m}(\mathbf{x}^{\gamma}\mathbf{x}^{\beta})=\mathbf{x}^{\gamma}\phi_{\alpha}^{m}(\mathbf{x}^{\beta}). By the definition of ϕαm\phi_{\alpha}^{m} these are equivalent to ϕαm(𝐱γ𝐱β)0\phi_{\alpha}^{m}(\mathbf{x}^{\gamma}\mathbf{x}^{\beta})\neq 0 implies 𝐱γϕαm(𝐱β)0\mathbf{x}^{\gamma}\phi_{\alpha}^{m}(\mathbf{x}^{\beta})\neq 0, or equivalently to,

γ+βαM implies βαM.\gamma+\beta-\alpha\in M\text{ implies }\beta-\alpha\in M.

The conclusion now follows from Proposition 3.3. ∎

In the next result, we use the semigroup splitting threshold to provide a bound for the Castelnuovo-Mumford regularity of affine semigoup rings. We refer the reader to Section 2 for information about aa-invariants and regularity.

Theorem 5.4.

Let MM and RR be as in Notation 5.1. Then, ai(R)mpt(M)a_{i}(R)\leqslant-{\operatorname{mpt}}(M). As a consequence,

reg(R)dim(R)mpt(M)=rank(M)mpt(M).\operatorname{reg}(R)\leqslant\dim(R)-{\operatorname{mpt}}(M)=\operatorname{rank}(M)-{\operatorname{mpt}}(M).

Moreover, if we present RR as S/IS/I, where S=𝕜[x1,,xu]S=\mathbb{k}[x_{1},\ldots,x_{u}] and each xix_{i} has degree di:=deg(xi)=|γi|d_{i}:=\deg(x_{i})=|\gamma_{i}| the degree of γi\gamma_{i} for i=1,,ui=1,\ldots,u, and ISI\subseteq S is a homogeneous ideal, then

β(I)dim(R)+i=1u(di1)mpt(M)1=rank(M)+i=1u(di1)mpt(M)1.\beta(I)\leqslant\dim(R)+\sum_{i=1}^{u}(d_{i}-1)-{\operatorname{mpt}}(M)-1=\operatorname{rank}(M)+\sum_{i=1}^{u}(d_{i}-1)-{\operatorname{mpt}}(M)-1.
Proof.

We can assume that MM is seminormal. Let m𝒜(M)m\in\mathcal{A}(M) be such that m>1m>1, which exists by Proposition 3.5 (1). Fix t0t\in\mathbb{Z}_{\geqslant 0} and αVmt(M)\alpha\in V_{m^{t}}(M). From Proposition 5.3 it follows that ϕαmt\phi^{m^{t}}_{\alpha} gives a splitting of the homogeneous injective map R(|α|)R1/mtR(-|\alpha|)\hookrightarrow R^{1/m^{t}} defined as multiplication by 𝐱α\mathbf{x}^{\alpha}. Thus, for each ii, the induced map

H𝔪i(R)(|α|)=H𝔪i(R(|α|))H𝔪i(R1/mt)=H𝔪i(R)1/mtH_{\mathfrak{m}}^{i}(R)(-|\alpha|)=H_{\mathfrak{m}}^{i}(R(-|\alpha|))\to H_{\mathfrak{m}}^{i}(R^{1/m^{t}})=H_{\mathfrak{m}}^{i}(R)^{1/m^{t}}

also splits. By comparing the highest degrees of these modules we obtain

ai(R)+|α|ai(R)mt.a_{i}(R)+|\alpha|\leqslant\frac{a_{i}(R)}{m^{t}}.

By taking the maximum value of |α||\alpha| over all αVmt(M)\alpha\in V_{m^{t}}(M) and letting tt\to\infty, by Proposition 3.21 we obtain that ai(R)mpt(M)a_{i}(R)\leqslant-{\operatorname{mpt}}(M) as desired. The inequality for regularity follows by definition, and the last equality by the relation between the rank of semigroups and dimension of semigroup rings (see e.g. [BrHe, p.257]).

Finally, the inequalities involving β(I)\beta(I) follow at once from the fact that β(I)Reg(I)Reg(R)1=reg(R)+i=1u(di1)1\beta(I)\leqslant\operatorname{Reg}(I)\leqslant\operatorname{Reg}(R)-1=\operatorname{reg}(R)+\sum_{i=1}^{u}(d_{i}-1)-1 by Lemma 2.3 and the previous inequalities. ∎

We now compute the pure threshold for a normal monoid that is Gorenstein. This follows previous work done for the FF-pure threshold [DSNB, Theorem B], which was motivated by a conjecture posted by Hirose, Watanabe and Yoshida [HWY].

Theorem 5.5.

Assume that MM is normal of rank dd. If RR is Gorenstein, then mpt(M)=ad(R){\operatorname{mpt}}(M)=-a_{d}(R).

Proof.

Since MM is normal, the Gorenstein property of R=𝕜[M]R=\mathbb{k}[M] is independent of the field 𝕜\mathbb{k} [bruns2009polytopes, Remark 6.34]. The pure threshold mpt(M){\operatorname{mpt}}(M) is also independent of 𝕜\mathbb{k}. If 𝕜\mathbb{k} has characteristic zero, then ad(𝕜[M])=ad([M])=ad(𝔽p[M])a_{d}(\mathbb{k}[M])=a_{d}(\mathbb{Q}[M])=a_{d}(\mathbb{F}_{p}[M]) for all p0p\gg 0 (see for instance by [DDSM, Lemma 4.3] adapted to the positively graded case). If 𝕃\mathbb{L} is any field extension of 𝕜\mathbb{k}, and 𝔪\mathfrak{m} is the homogeneous maximal ideal of RR, then we have graded isomorphisms H𝔪d(R)𝕜𝕃H𝔪d(R𝕜𝕃)=H𝔪d(𝕃[M])H^{d}_{\mathfrak{m}}(R)\otimes_{\mathbb{k}}\mathbb{L}\cong H^{d}_{\mathfrak{m}}(R\otimes_{\mathbb{k}}\mathbb{L})=H^{d}_{\mathfrak{m}}(\mathbb{L}[M]). Thus, we may assume that 𝕜\mathbb{k} is a perfect field of characteristic p>0p>0. We can write R=S/IR=S/I, where S=𝕜[T1,,Tu]S=\mathbb{k}[T_{1},\ldots,T_{u}], each TiT_{i} maps to a generator 𝐱γi\mathbf{x}^{\gamma_{i}} of RR and deg(Ti)=|γi|=di>0\deg(T_{i})=|\gamma_{i}|=d_{i}>0. Since RR is Gorenstein, we have that HomR(R1/pe,R)(I[pe]:SI)/I[pe]=(fe+I[pe])/I[pe]\operatorname{Hom}_{R}(R^{1/p^{e}},R)\cong(I^{[p^{e}]}:_{S}I)/I^{[p^{e}]}=(f_{e}+I^{[p^{e}]})/I^{[p^{e}]} [FedderFputityFsing]. If 𝔽:0FcF0=SR0\mathbb{F}_{\bullet}:0\to F_{c}\to\ldots\to F_{0}=S\to R\to 0 is a minimal free resolution of RR over SS, then c=ht(I)c=\operatorname{ht}(I) and Fc=S(Dad(R))F_{c}=S(-D-a_{d}(R)), where D=i=1udiD=\sum_{i=1}^{u}d_{i}. The minimal free resolution of 𝔽e:0FceF0e=SS/I[pe]0\mathbb{F}_{\bullet}^{e}:0\to F_{c}^{e}\to\ldots\to F_{0}^{e}=S\to S/I^{[p^{e}]}\to 0 of S/I[pe]S/I^{[p^{e}]} is such that Fce=S(pe(Dad(R)))F_{c}^{e}=S(p^{e}(-D-a_{d}(R))). The comparison map 𝔽e𝔽\mathbb{F}_{\bullet}^{e}\to\mathbb{F}_{\bullet} induced by the natural surjection S/I[pe]RS/I^{[p^{e}]}\to R in homological degree cc is S(pe(Dad(R)))S(Dad(R))S(p^{e}(-D-a_{d}(R)))\to S(-D-a_{d}(R)). Furthermore, it is given, up to an invertible element, by multiplication by fef_{e} [VraciuGorTightClosure, Lemma 1]. Since such a map is homogeneous of degree zero, we conclude that deg(fe)=(pe1)(D+ad(R))\deg(f_{e})=(p^{e}-1)(D+a_{d}(R)). Let 𝔫=(T1,,Tu)\mathfrak{n}=(T_{1},\ldots,T_{u}). As fe𝔫[pe]f_{e}\notin\mathfrak{n}^{[p^{e}]} by Fedder’s criterion [FedderFputityFsing], there is a monomial T1n1TunuT_{1}^{n_{1}}\cdots T_{u}^{n_{u}} in its support with 0nipe10\leqslant n_{i}\leqslant p^{e}-1 for all ii. This implies that the map S/I(S/I)1/peS/I\to(S/I)^{1/p^{e}} sending 1(T1pe1n1Tupe1nu)1/pe1\mapsto(T_{1}^{p^{e}-1-n_{1}}\cdots T_{u}^{p^{e}-1-n_{u}})^{1/p^{e}} splits. Via the isomorphism S/IRS/I\cong R, this means that the map RR1/pe=𝕜[M1/pe]R\to R^{1/p^{e}}=\mathbb{k}[M^{1/p^{e}}] sending 1(𝐱γ1(pe1n1)𝐱γu(pe1nu))1/pe=𝐱β(e)1\mapsto\left(\mathbf{x}^{\gamma_{1}(p^{e}-1-n_{1})}\cdots\mathbf{x}^{\gamma_{u}(p^{e}-1-n_{u})}\right)^{1/p^{e}}=\mathbf{x}^{\beta(e)} splits, and so, β(e)Vpe(M)\beta(e)\in V_{p^{e}}(M). Note that

|β(e)|=i=1u|γi|(pe1ni)pe=(pe1)Di=1unidipe=(pe1)Ddeg(fe)pe=ad(R)pe1pe.|\beta(e)|=\frac{\sum_{i=1}^{u}|\gamma_{i}|(p^{e}-1-n_{i})}{p^{e}}=\frac{(p^{e}-1)D-\sum_{i=1}^{u}n_{i}d_{i}}{p^{e}}=\frac{(p^{e}-1)D-\deg(f_{e})}{p^{e}}=-a_{d}(R)\frac{p^{e}-1}{p^{e}}.

We have that mpt(M)lime|β(e)|=ad(R){\operatorname{mpt}}(M)\geqslant\lim\limits_{e\to\infty}|\beta(e)|=-a_{d}(R) by Proposition 3.21. As the other inequality always holds by Theorem 5.4, we have equality. ∎

From the previous result, one may wonder if the converse is true. In particular, as fpt(𝕜[M])=ad(𝕜[M]){\operatorname{fpt}}(\mathbb{k}[M])=a_{d}(\mathbb{k}[M]) implies that 𝕜[M]\mathbb{k}[M] is Gorenstein if 𝕜[M]\mathbb{k}[M] has a structure of standard graded 𝕜\mathbb{k}-algebra [STV]. This motivates the following question.

Question 5.6.

Assume that MM is normal of rank dd. If mpt(M)=ad(R){\operatorname{mpt}}(M)=-a_{d}(R), is RR is Gorenstein?

We now provide a bound for the depth of RR, which recovers Hochster’s result that normal semigroup rings are Cohen-Macaulay [HochsterToric, Theorem 1].

Theorem 5.7.

Let MM and RR be as in Notation 5.1. Then, mpdim(M)depth(R).{\operatorname{mpdim}}(M)\leqslant\operatorname{depth}(R).

Proof.

We can assume that MM is seminormal. Let S=𝕜[y1,,yu]S=\mathbb{k}[y_{1},\ldots,y_{u}] endowed with the MM-grading given by deg(yi)=γi\deg(y_{i})=\gamma_{i}. We set a surjection of 𝕜\mathbb{k}-algebras ρ:SR\rho:S\to R by yi𝐱γiy_{i}\mapsto\mathbf{x}^{\gamma_{i}}. Let 𝔪=(𝐱γ1,,𝐱γu)R\mathfrak{m}=(\mathbf{x}^{\gamma_{1}},\ldots,\mathbf{x}^{\gamma_{u}})\subseteq R and η=(y1,,yu)S\eta=(y_{1},\ldots,y_{u})\subseteq S. We note that ρ(η)=𝔪.\rho(\eta)=\mathfrak{m}. Set J=Ker(ρ)J=\operatorname{Ker}(\rho), so that R=S/JR=S/J.

Set t=depth(R)=min{i|ExtSui(R,S)0}t=\operatorname{depth}(R)=\min\{i\;|\;\operatorname{Ext}^{u-i}_{S}(R,S)\neq 0\}. We first show that AnnR(ExtSut(R,S))𝐱𝒫(M)\operatorname{Ann}_{R}\left(\operatorname{Ext}^{u-t}_{S}(R,S)\right)\subseteq\mathbf{x}^{\mathcal{P}(M)}; we proceed by contradiction. Suppose that there exists an MM-homogeneous element fRf\in R such that fAnnR(ExtSut(R,S))𝐱𝒫(M)f\in\operatorname{Ann}_{R}\left(\operatorname{Ext}_{S}^{u-t}(R,S)\right)\setminus\mathbf{x}^{\mathcal{P}(M)}. Let m0m\in\mathbb{Z}_{\geqslant 0} be such that log(f)mVm(M)\log(f)\in mV_{m}(M). Since the multiplication map ExtSut(R,S)fExtSut(R,S)\operatorname{Ext}_{S}^{u-t}(R,S)\stackrel{{\scriptstyle f}}{{\to}}\operatorname{Ext}_{S}^{u-t}(R,S) is the zero map, we have that H𝔪t(R)fH𝔪t(R)H^{t}_{\mathfrak{m}}(R)\stackrel{{\scriptstyle f}}{{\to}}H^{t}_{\mathfrak{m}}(R) is the zero map by Matlis duality [Matlis]. Thus, H𝔪t(R1/m)f1/mH𝔪t(R1/m)H^{t}_{\mathfrak{m}}(R^{1/m})\stackrel{{\scriptstyle f^{1/m}}}{{\to}}H^{t}_{\mathfrak{m}}(R^{1/m}) is the zero map as well. Since the composition of RιR1/mf1/mR1/mϕlog(f)mRR\stackrel{{\scriptstyle\iota}}{{\to}}R^{1/m}\stackrel{{\scriptstyle f^{1/m}}}{{\to}}R^{1/m}\stackrel{{\scriptstyle\phi_{\log(f)}^{m}}}{{\to}}R is the identity, we have the same for the composition

H𝔪t(R)ιH𝔪t(R1/m)f1/mH𝔪t(R1/m)ϕlog(f)mH𝔪t(R).H^{t}_{\mathfrak{m}}(R)\stackrel{{\scriptstyle\iota}}{{\to}}H^{t}_{\mathfrak{m}}(R^{1/m})\stackrel{{\scriptstyle f^{1/m}}}{{\to}}H^{t}_{\mathfrak{m}}(R^{1/m})\stackrel{{\scriptstyle\phi_{\log(f)}^{m}}}{{\to}}H^{t}_{\mathfrak{m}}(R).

Since the middle map is zero, we have that H𝔪t(R)=0H^{t}_{\mathfrak{m}}(R)=0, which is not possible because t=depth(R).t=\operatorname{depth}(R).

Since AnnR(ExtSut(R,S))𝐱𝒫(M)\operatorname{Ann}_{R}\left(\operatorname{Ext}^{u-t}_{S}(R,S)\right)\subseteq\mathbf{x}^{\mathcal{P}(M)}, we have that

(5.1) dim(ExtSut(R,S))dim(R/𝐱𝒫(M))=mpdim(M).\dim\left(\operatorname{Ext}^{u-t}_{S}(R,S)\right)\geqslant\dim\left(R/\mathbf{x}^{\mathcal{P}(M)}\right)={\operatorname{mpdim}}(M).

Since SS is a Gorenstein ring, its injective resolutions as SS-module is given by

0SE0E2Eu,0\to S\to E^{0}\to E^{2}\to\ldots E^{u},

where Ej=ht(𝔭)=jE(S/𝔭)E^{j}=\bigoplus_{\operatorname{ht}(\mathfrak{p})=j}E(S/\mathfrak{p}) is the direct sum of the injective hulls of all the prime ideals in SS of height jj [Bass]. Therefore,

(5.2) dim(ExtSut(R,S))t.\dim\left(\operatorname{Ext}^{u-t}_{S}(R,S)\right)\leqslant t.

Combining Inequalities (5.1) and (5.2) we obtain the desired result. ∎

We now relate the pure ratio of a monoid MM to the splitting ratio of RR [AE] and FF-signature [SmithVdB, HunekeLeuschke, Tucker].

Proposition 5.8.

Let MM and RR be as in Notation 5.1. If char(𝕜)𝒜(M)\operatorname{char}(\mathbb{k})\in\mathcal{A}(M), then mpr(M)\operatorname{mpr}(M) is the FF-splitting ratio of RR. As a consequence, if char(𝕜)\operatorname{char}(\mathbb{k}) is a prime number and MM is normal, then mps(M)\operatorname{mps}(M) equal to the FF-signature of RR.

Proof.

Let p=char(𝕜)p=\operatorname{char}(\mathbb{k}) and 𝔪\mathfrak{m} be the maximal homogeneous ideal in RR. Let

Ie={fR|ϕ(f1/pe)𝔪ϕHom(R1/pe,R)}.I_{e}=\{f\in R\;|\;\phi(f^{1/p^{e}})\in\mathfrak{m}\;\;\forall\phi\in\operatorname{Hom}(R^{1/p^{e}},R)\}.

We note that dim𝕜(R/Ie)=|Vpe(M)|\dim_{\mathbb{k}}(R/I_{e})=|V_{p^{e}}(M)|, and that 𝐱𝒫(M)\mathbf{x}^{\mathcal{P}(M)} is the splitting prime of RR [AE]. It follows that mpdim(M)=sdim(R){\operatorname{mpdim}}(M)=\operatorname{sdim}(R), and the result follows for the ratios.

We now discuss the claim about FF-signature. We have that char(𝕜)𝒜(M)\operatorname{char}(\mathbb{k})\in\mathcal{A}(M) for normal monoids by Proposition 3.5. The result follows because the FF-signature coincides with the FF-splitting ratio for strongly FF-regular rings, and RR is strongly F-regular if and only if MM is normal. ∎

We end this section with a monoid version of Kunz’s characterization of regularity [Kunz].

Theorem 5.9.

Let MM be an affine monoid. Then, |Vm(M)|=mrank(M)|V_{m}(M)|=m^{\operatorname{rank}(M)} for some m>0m\in\mathbb{Z}_{>0} if and only if M>0tM\cong\mathbb{Z}^{t}_{>0} for some tt.

Proof.

Since |Vm(M)|=mrank(M)|V_{m}(M)|=m^{\operatorname{rank}(M)}, we have that |Vmt(M)|=mtrank(M)|V_{m^{t}}(M)|=m^{t\operatorname{rank}(M)} by Lemma 3.17 (3). Then,

mps(R)=limt|Vm(M)|mrank(M)=1.\operatorname{mps}(R)=\lim\limits_{t\to\infty}\frac{|V_{m}(M)|}{m^{\operatorname{rank}(M)}}=1.

Hence, mpdim(M)=rank(M){\operatorname{mpdim}}(M)=\operatorname{rank}(M), and so, MM is a normal monoid by Corollary 3.30. We have that mps(R)=1\operatorname{mps}(R)=1. Then, 𝔽p[M]\mathbb{F}_{p}[M] is a regular graded 𝔽p\mathbb{F}_{p}-algebra, because s(R)=1s(R)=1 by Proposition 5.8 and the characterization of regular rings via FF-signature [HunekeLeuschke, Corollary 16]. Moreover, MM has a set of rank(M)\operatorname{rank}(M) minimal generators. Hence, M>0tM\cong\mathbb{Z}^{t}_{>0}. ∎

Acknowledgments

We thank the reviewer for the careful reading of our paper and for the suggested improvements. The first author was partially supported by the PRIN 2020 project 2020355B8Y “Squarefree Gröbner degenerations, special varieties and related topics". The second author was supported by NSF Grant DMS #2001645/2303605. The third author was supported by CONACyT Grant #284598.

References