Purity of monoids and characteristic-free splittings in semigroup rings
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 invariants2020 Mathematics Subject Classification:
Primary 20M32, 20M25, 13A35 ; Secondary 13C151. 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 for some , and , we study the pure -submodules of that are translations of , which algebraically corresponds to free summands of as -module. It turns out that the purity of 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 -signature of normal affine semigroup rings [ToricSingh, von2011f]. Moreover, the structure of as -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 -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 , denoted by , which is motivated by the -pure threshold in prime characteristic. This number can be described as the largest degree of a pure translation of inside the cone or, equivalently, of for some . We show that gives an upper bound for the Castelnuovo-Mumford regularity defined in terms of local cohomology, and the Castelnuovo-Mumford regularity defined in terms of graded Betti numbers of (see Section 2 for more details).
Theorem A (Theorem 5.4).
Let be a seminormal monoid with a minimal set of generators . Then, . As a consequence,
Moreover, if we present as , where and each has degree the degree of for , and is a homogeneous ideal, then
Theorem A allows us to give an upper bound for the degrees of generators of the defining ideal . We also show that is a rational if is a normal (see Proposition 4.4). Despite being inspired by -pure thresholds, these numbers do not always coincide (see Example 3.23 and Remark 3.24). In addition, is defined independently of the field and so it is a characteristic-free invariant, while the -pure threshold is only defined when has prime characteristic.
We introduce the pure prime ideal , and the pure prime face , of a seminormal monoid (see Corollary 3.28, and Definitions 3.26 and 3.29). The former emulates the splitting prime ideal of an -pure ring, while the latter is related to the quotient of a ring by its splitting prime. In fact, the submonoid is normal (see Corollary 3.28). We note that the rank of is a monoid version of the splitting dimension and so we call it the pure dimension and denote it by . It turns out that this rank is equal to the rank of if and only if is normal, and it is non-negative if and only if is seminormal (see Corollary 3.30). Therefore, in some sense, measures how far a seminormal monoid is from being normal. Furthermore, is related to the depth of as the following theorem shows.
Theorem B (Theorem 5.7).
If is a seminormal monoid, then
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 in as varies. More specifically, if is such that , we define
Theorem C (Theorem 4.6).
Let be a seminormal monoid, , and . Then,
exists and it is positive for every increasing sequence . Furthermore, if is normal, then .
We call the limit in Theorem C the pure ratio of . If the field has prime characteristic, this number coincides with the splitting ratio of [AE]. A consequence of Theorem C is that the value of the -splitting ratio depends only on the structure of , and so it is independent of the characteristic of the field as long as is -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 be a field of any characteristic and a positive integer. Let be an affine monoid, i.e., a finitely generated submonoid of . We fix a minimal set of generators of . Let denote the group generated by and the cone generated by .
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 be an affine monoid. A subset is an -module if . An -module is an ideal if it is contained in . An ideal is prime if whenever with , we must have or . The rank of is the dimension of the -vector space .
Let be the affine semigroup associated to . As a -vector space, is generated by the monomials . We note that the monomial ideals of are precisely those generated by for some ideal . Under this correspondence, prime monomial ideals of correspond to prime ideals of . For every -module we have a corresponding -module in the algebraic closure of . Moreover, we have .
Graded algebras and modules
A non-negatively graded algebra is a ring that admits a direct sum decomposition of Abelian groups such that . It follows from this that is a ring, and each is an -module. If we let , then is an ideal of , called the irrelevant ideal.
Throughout this manuscript we will make the assumption that is Noetherian or, equivalently, that there exist finitely many elements such that , which can be assumed to be homogeneous of degrees . In this case, note that is a quotient of a polynomial ring by a homogeneous ideal.
A -graded -module is an -module that admits a direct sum decomposition of Abelian groups, and such that . As a consequence, each is an -module. Moreover, if is Noetherian there exists such that for all ; on the other hand, if is Artinian there exists such that for all .
Given a -graded -module , and an integer , we define the shift as the -graded -module whose -th graded component is . In particular, is a free graded -module of rank one with generator in degree .
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 be a field and , with . Let be a finitely generated -graded -module. If we let , then the graded local cohomology modules are Artinian and -graded.
Definition 2.1.
Let be the -th -invariant of . If we define the Castelnuovo-Mumford regularity of as . On the other hand, if we let .
In the standard graded case, has a well-known interpretation in terms of graded Betti numbers of . In our setup this is still the case, but the degrees of the algebra generators of must be taken into account. For , let . As another way to see this, for a non-zero graded -module let be the maximum degree of an element in a minimal homogeneous generating set of . If
is a minimal graded free resolution of where is the projective dimension of , then .
Definition 2.2.
For we let , while for we let .
In the standard graded case, that is, when , then . 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
Proof.
We may assume that , otherwise the claim is trivial. We prove the statement by induction on . If then is free, and it is clear that . On the other hand, for all , while . It follows that , and the base case follows.
Now assume that . We have a graded short exact sequence , where is the first free module in a minimal free resolution of , and . By induction we have that . Moreover, it is clear from the definitions that . We have that [DSMNB, Proposition 3.1], and therefore the above equality gives that
(2.1) |
We now show that . The short exact sequence yields graded isomorphisms for all , and a graded exact sequence . If we had , then necessarily , and looking at top degrees in the above exact sequence we also conclude that . On the other hand, [DSMNB, Proposition 3.1], and so, , a contradiction. Thus we always have that , and by (2.1) the inequality is proved.
For the reverse inequality, first observe that the above isomorphisms give that for all , while the exact sequence yields that . Since and is a submodule of a free module, it has positive depth, and thus . It follows that . By induction we have that , and thus since the inequality always holds. Now, the above exact sequence on local cohomology also gives that , and thus . In conclusion, we have that , and the proof is complete. ∎
3. Purity of -modules and (semi)normal affine monoids
Definition 3.1.
Let be -modules. We say that the inclusion is pure if is also an -module.
Example 3.2.
Let be the monoid generated by . We have that the inclusion is pure. In Figure 1 we represent the elements of with circles and the ones from with multiplication signs. The shaded region is included to illustrate that is obtained as a translation of .

In the following proposition, we provide equivalent statements for Definition 3.1 in a particular case.
Proposition 3.3.
Let be an -module and . The following statements are equivalent:
-
(i)
The inclusion is pure.
-
(ii)
For every and we have
-
(iii)
Proof.
First, assume (i) and let and with . Thus, and then (ii) follows.
Now, assume (ii) and let be such that . Write with . Then implies . Thus, (iii) follows.
Finally, assume (iii). Let and . Assume by means of contradiction that . Then for some , which implies which contradicts (iii). Thus, 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 is called seminormal if, whenever is such that and , then . The monoid is called normal if .
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 be an affine monoid.
-
(1)
is seminormal if and only if there exists such that .
-
(2)
is normal if and only if for every .
Proof.
We note that the containment holds trivially for any .
For (1), assume that for some . Let be such that and . We can find nonnegative integers and such that , and thus . It follows by our assumption that , and thus is seminormal. Conversely, let be the set of all faces of (of any dimension); we note that is a finite set. For any we consider the finitely generated Abelian group . Let be a prime number such that the ideal is not associated to as a -module for any . We claim that . Let , then the interior of some , and also . By the choice of it follows that and then [BrunsLiRomer, Theorem 2.1].
For (2), if is normal then . Conversely, let , then , with and . If we let , it then follows that , as desired. ∎
Motivated by the previous result, we consider the following definition.
Definition 3.6.
We set .
Remark 3.7.
As a consequence of Proposition 3.5, we deduce that
-
(1)
if and only if is seminormal;
-
(2)
is normal if and only if .
We now see that is a multiplicative set.
Lemma 3.8.
Let . Then if and only if both and .
Proof.
First assume that . Then because . Likewise . For the converse, assume that is such that . Note that is such that , and thus because . But then is such that , and thus because . It follows that . ∎
Now, we consider a set that records the pure translations of in . In Section 5 we see that this set corresponds to the free summands of as a -module.
Definition 3.9.
Let . We set
Moreover, we set
Remark 3.10.
We note that, because of Proposition 3.3, is precisely the set of such that is pure.
Remark 3.11.
We note that if and only if . As a consequence we have
Indeed, if then . Conversely, fix . From the containments , it follows that .
Remark 3.12.
If is seminormal, then . Indeed, if , then for every .
In the following remarks we observe that is compatible with projections onto faces of and with isomorphisms of monoids.
Remark 3.13.
For every face of and we have . Indeed, let , then
Remark 3.14.
Since every isomorphism of monoids extends to an isomorphism of groups , it follows that . Furthermore, for every we have .
We now describe basic properties . In particular, we see that this set is finite.
Proposition 3.15.
Let and be such that . If , then .
Proof.
Let be such that . Thus, from
it follows that . This argument implies that . Likewise, , finishing the proof. ∎
Corollary 3.16.
For every the set is an ideal of .
Proof.
Let and . Suppose , then . By Proposition 3.15 this implies which is a contradiction. ∎
The following lemma provides useful facts about the sets . We recall that the Minkowski sum of two subsets is defined as .
Lemma 3.17.
Let . Then,
-
(1)
.
-
(2)
;
-
(3)
.
Proof.
We begin with the containment in (1). Let , then . Consider , then , Thus, and the conclusion follows. Now we prove the containment . Let and let be such that . Then .
We continue with (2). Let and . Clearly we have . Let and be such that , then . Since , Proposition 3.3 applied to implies , i.e, . Thus, Proposition 3.3 applied to implies . Therefore, by Proposition 3.3 we have .
We now show (3). Let . We first show that for , we proceed by contradiction. Set , and note that splits since
is an -module (see Remark 3.10). Let . If and , then but also , which is not possible. We conclude , and so . Thus, . By symmetry, , an then . It follows that by Proposition 3.15 and Remark 3.12, which is a contradiction. Therefore, the union is disjoint, and so, is also disjoint. By applying the same argument, the union
is disjoint. Hence,
is a set of of cardinality Finally, the inequality follows from Part (2). ∎
Proposition 3.18.
Let and . Write with . If , then . Furthermore, .
Proof.
By assumption we have that for some . By way of contradiction suppose that . Since , it follows by Proposition 3.15 that . However, this contradicts Remark 3.12, and therefore .
For the second claim, recall that is a disjoint union of -modules (see proof of Lemma 3.17 (3)), and thus is pure. As a consequence, the -module contains as a free direct summand, and thus , where the latter is the rank of as a -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 -pure threshold of a ring [TW2004]. This is because the -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 [DSNB]. However, the -pure threshold of can be different than the pure threshold of (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 as
If , i.e., if is not seminormal, we set .
Remark 3.20.
Let . By Proposition 3.18 we have for every . Therefore, .
We now discuss how the pure threshold of a monoid can be obtained from any increasing sequence in .
Proposition 3.21.
Let be the elements of ordered increasingly. Then,
In particular,
for any .
Proof.
We now compute some examples of pure thresholds. We note that this invariant depends on the grading given by the embedding
Example 3.22.
Let be generated by , where and is the canonical basis in . Then, , and so, .
Example 3.23.
Let , and . Then is the Veronese subring of order of a polynomial ring with the grading . We have that
and therefore . We point out that, if has prime characteristic, then [HWY, Example 6.1].
Remark 3.24.
It follows from Example 3.23 that may differ from even when is normal. This is not surprising since is independent of the presentation of as a quotient of a polynomial ring, while we have already observed that heavily depends on the degrees of the generators and on the embedding of .
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 by
Proposition 3.26.
Let be an affine monoid. Then is a prime ideal of .
Proof.
We obtain the following theorem that relates with the normality of .
Theorem 3.27.
Let be an affine monoid. Then is normal if and only
Proof.
We begin with the forward direction. Since is normal, by Remark 3.14 we can assume that is a submonoid of for some and such that [bruns2009polytopes, Theorem 2.29]. Fix and chose bigger than every entry in . By Remark 3.7 we have and . By the choice of , it is clear that
(3.1) |
Thus, the left hand side expression in (3.1) is equal to . Intersecting this with we obtain,
Therefore, which shows .
We continue with the backward direction. Let be such that for some . By Proposition 3.5 it suffices to show . Write with , then . Thus, , where denotes the ideal of generated by the set . It follows that for every . Hence,
By assumption there exists such that . Therefore,
On the other hand, , then , which finishes the proof. ∎
Corollary 3.28.
There exists a face of such that . Moreover, the monoid 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 is normal it suffices to show Moreover, we may assume is seminormal. We note that
where the last inclusion follows from Remark 3.13. Since we always have the other inclusion , 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 in Proposition 3.28 is called the pure prime face of . We define the pure dimension of by If , we set .
Corollary 3.30.
Let be an affine monoid. Then
-
(1)
.
-
(2)
if and only if is seminormal.
-
(3)
if and only if is normal.
Proof.
We finish this section with the following example.
Example 3.31.

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 . Throughout we adopt the same notation from Section 3.
Definition 4.1.
Let be a seminormal affine monoid, and let be its pure prime face. For every we define
Moreover, we set
When is normal, there is a simple description of as the region in Notation 4.2. We prove that these regions coincide in Lemma 4.3, which also includes important properties of .
Notation 4.2.
Let be a normal affine monoid. Let be the supporting hyperplanes of so that Let be rational vectors such that if and only if for . We can further assume that for every and that [bruns2009polytopes, Remark 1.72]. We define by
Lemma 4.3.
Let be a seminormal affine monoid. Then
-
(1)
is a bounded set and it has volume, i.e., its boundary has measure zero in the -dimensional Lebesgue measure on .
-
(2)
There exists an increasing sequence such that for every and
-
(3)
for every .
-
(4)
If is normal and is as in Notation 4.2, then .
Proof.
We begin with (1). Let be a minimal set of generators of and consider the region . Let , then for some and . Since is normal by Corollary 3.28, it follows that . If , then Proposition 3.15 implies which contradicts Remark 3.12. We conclude , and then , is contained in which is bounded.
Now, let and ∘ denote boundary and interior on , respectively. Let denote the -dimensional Lebesgue measure on . We note that for any we have ; indeed, if for some , then for some , which would imply Therefore, for any and any we have , where denotes the ball in with radius and center . Therefore, there exists a real such that for any such and we have By Lebesgue’s density theorem [mattila, Corollary 2.14], we conclude .
We continue with (2). Let be the elements of ordered increasingly. For each set and notice by Lemma 3.8. The conclusion now follows from Lemma 3.12 (1).
Now we prove (3). Let and , it suffices to show . By (2), we have for some . We may assume and then divides . Therefore, there exists and such that . Since is normal by Corollary 3.28, it follows that . Thus, by Proposition 3.15 and Lemma 3.17, which finishes the proof.
We finish with (4). If is normal we have [von2011f, Lemma 3.11]. Thus, the equality follows as the set is dense in . ∎
From Lemma 4.3 (4) we obtain that the pure threshold of normal monoids is rational.
Proposition 4.4.
If is normal, then .
Proof.
The statement follows readily from Lemma 4.3 (4) and the equality . ∎
We now turn our focus to asymptotic growth of the number of elements in the sets . We define the following limit, which we prove exists in Theorem 4.6
Definition 4.5.
Let be a seminormal affine monoid. Set and let be the elements of ordered increasingly. We define the pure ratio of as
We define the pure signature of as
In the following theorem we show that exists as a limit, and that it equals the relative volume of . Here, by relative volume with respect to a lattice of rank in an -dimensional hyperplane , denoted by , we mean the -dimensional volume in normalized such that any fundamental domain of has volume one.
Theorem 4.6.
Let be a seminormal affine monoid. We have that
In particular, is normal if and only if . Furthermore, in this case .
Proof.
By Lemma 4.3 (1), the characteristic function is Riemann integrable. Now, by Lemma 4.3 (3) and Corollary 3.28 we have . Thus, is a Riemann sum for with normalized volumes of the cells and mesh the diameter of a fundamental domain for . Therefore, by taking the limit we obtain that the limit exists and is equal to . We note that is positive since has interior points of (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 -signature of normal semigroup rings [ToricSingh, von2011f].
Example 4.7.
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 be a seminormal affine monoid. Is 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 be the affine semigroup ring of . Given , we set the -algebra . Given and , we consider the -linear map given by if and zero otherwise. For an ideal , we denote by the corresponding -homogeneous -ideal,
For an -homogeneous element , we denote by the corresponding element in .
Remark 5.2.
We note that via the -algebra map given by .
Proposition 5.3.
Let be an affine monoid, and let . Then, is a map of -modules if and only if .
Proof.
We note that is a map of -modules if and only if for every and we have . By the definition of these are equivalent to implies , or equivalently to,
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 -invariants and regularity.
Theorem 5.4.
Let and be as in Notation 5.1. Then, . As a consequence,
Moreover, if we present as , where and each has degree the degree of for , and is a homogeneous ideal, then
Proof.
We can assume that is seminormal. Let be such that , which exists by Proposition 3.5 (1). Fix and . From Proposition 5.3 it follows that gives a splitting of the homogeneous injective map defined as multiplication by . Thus, for each , the induced map
also splits. By comparing the highest degrees of these modules we obtain
By taking the maximum value of over all and letting , by Proposition 3.21 we obtain that 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 follow at once from the fact that 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 -pure threshold [DSNB, Theorem B], which was motivated by a conjecture posted by Hirose, Watanabe and Yoshida [HWY].
Theorem 5.5.
Assume that is normal of rank . If is Gorenstein, then .
Proof.
Since is normal, the Gorenstein property of is independent of the field [bruns2009polytopes, Remark 6.34]. The pure threshold is also independent of . If has characteristic zero, then for all (see for instance by [DDSM, Lemma 4.3] adapted to the positively graded case). If is any field extension of , and is the homogeneous maximal ideal of , then we have graded isomorphisms . Thus, we may assume that is a perfect field of characteristic . We can write , where , each maps to a generator of and . Since is Gorenstein, we have that [FedderFputityFsing]. If is a minimal free resolution of over , then and , where . The minimal free resolution of of is such that . The comparison map induced by the natural surjection in homological degree is . Furthermore, it is given, up to an invertible element, by multiplication by [VraciuGorTightClosure, Lemma 1]. Since such a map is homogeneous of degree zero, we conclude that . Let . As by Fedder’s criterion [FedderFputityFsing], there is a monomial in its support with for all . This implies that the map sending splits. Via the isomorphism , this means that the map sending splits, and so, . Note that
We have that 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 implies that is Gorenstein if has a structure of standard graded -algebra [STV]. This motivates the following question.
Question 5.6.
Assume that is normal of rank . If , is is Gorenstein?
We now provide a bound for the depth of , which recovers Hochster’s result that normal semigroup rings are Cohen-Macaulay [HochsterToric, Theorem 1].
Theorem 5.7.
Let and be as in Notation 5.1. Then,
Proof.
We can assume that is seminormal. Let endowed with the -grading given by . We set a surjection of -algebras by . Let and . We note that Set , so that .
Set . We first show that ; we proceed by contradiction. Suppose that there exists an -homogeneous element such that . Let be such that . Since the multiplication map is the zero map, we have that is the zero map by Matlis duality [Matlis]. Thus, is the zero map as well. Since the composition of is the identity, we have the same for the composition
Since the middle map is zero, we have that , which is not possible because
We now relate the pure ratio of a monoid to the splitting ratio of [AE] and -signature [SmithVdB, HunekeLeuschke, Tucker].
Proposition 5.8.
Let and be as in Notation 5.1. If , then is the -splitting ratio of . As a consequence, if is a prime number and is normal, then equal to the -signature of .
Proof.
Let and be the maximal homogeneous ideal in . Let
We note that , and that is the splitting prime of [AE]. It follows that , and the result follows for the ratios.
We now discuss the claim about -signature. We have that for normal monoids by Proposition 3.5. The result follows because the -signature coincides with the -splitting ratio for strongly -regular rings, and is strongly F-regular if and only if is normal. ∎
We end this section with a monoid version of Kunz’s characterization of regularity [Kunz].
Theorem 5.9.
Let be an affine monoid. Then, for some if and only if for some .
Proof.
Since , we have that by Lemma 3.17 (3). Then,
Hence, , and so, is a normal monoid by Corollary 3.30. We have that . Then, is a regular graded -algebra, because by Proposition 5.8 and the characterization of regular rings via -signature [HunekeLeuschke, Corollary 16]. Moreover, has a set of minimal generators. Hence, . ∎
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.