A Frobenius version of Tian’s Alpha-Invariant
Abstract.
For a pair consisting of a projective variety over a perfect field of characteristic and an ample line bundle on , we introduce and study a positive characteristic analog of Tian’s -invariant, which we call the -invariant. We utilize the theory of -singularities in positive characteristics, and our approach is based on replacing klt singularities with the closely related notion of global -regularity. We show that the -invariant of a pair can be understood in terms of the global Frobenius splittings of the linear systems . We establish inequalities relating the -invariant with the -signature, and use that to prove the positivity of the -invariant for all globally -regular projective varieties (with respect to any ample on ). When is a Fano variety and , we prove that the -invariant of is always bounded above by and establish tighter comparisons with the -signature. We also show that for toric Fano varieties, the -invariant matches with the usual (complex) -invariant. Finally, we prove that the -invariant is lower semicontinuous in a family of globally -regular Fano varieties.
1. Introduction
The -invariant of a complex Fano variety was introduced by Tian in [Tia87] to provide a sufficient criterion for -stability of , a condition that guarantees the existence of a Kähler-Einstein metric on . Though initially defined analytically, Demailly later reinterpreted the -invariant in terms of the log canonical threshold [CS08], an algebraic invariant of the singularities of divisors on . Since then, understanding the -invariant and -stability more generally has led to many fundamental advances in our understanding of complex Fano varieties; see [OS12], [Bir21], [Xu21]. The minimal model program (MMP), and the singularities that arise therein have played a key role in these advances.
The purpose of this paper is to study a positive characteristic analog of the -invariant. To do this, we replace the singularities of the MMP with singularities defined using the Frobenius map (“-singularities"). Though -singularities have fundamentally different definitions than the singularities of the MMP, a dictionary involving many precise relationships between these classes has been established; see [Smi97], [Har98], [MS97], [HW02], [HY03], [Tak04] and [MS18]. Under this dictionary, log canonical (resp. Kawamata log-terminal (klt)) singularities correspond to -split (resp. strongly -regular) singularities (2.15). Since the -invariant of a complex Fano variety involves the log canonicity of anti-canonical -divisors of , this inspires our definition of the Frobenius version:
Definition 1.1.
Let be a globally -regular Fano variety over a perfect field of positive characteristic. Then, we define the -invariant of as
Since we intend for the -invariant to capture global properties of anti-canonical -divisors on , we use the notion of global -splitting (2.15), or equivalently, -splitting of the cone over the corresponding divisors on (with respect to ). To do so, we require to be globally -regular (2.16), which is equivalent to the cone over being strongly -regular. Thus, global -regularity can be thought of as a Frobenius analog of the klt-condition on complex Fano varieties. Furthermore, we note that while working over , simply replacing globally -regular and globally -split in 1.1 by klt on the cone, and log canonical on the cone respectively, we obtain the minimum value between the usual -invariant of and (see 4.11). Thus, at least for Fano varieties with , the -invariant is a “Frobenius-analog" of Tian’s -invariant.
Our first theorem proves some surprising properties of the -invariant in contrast to the complex version, and establishes connections to the -signature of , another important invariant and a Frobenius version of the local volume of singularities:
Theorem 1.2.
Let be a globally -regular Fano variety over a perfect field of positive characteristic of positive dimension. Then,
-
(1)
The -invariant of is at most 1/2 (4.5).
-
(2)
Assume that is geometrically connected over the (perfect) base field. Then, we have if and only if the -signature of (with respect to ) equals , where is the dimension of (4.8).
-
(3)
More generally (and still assuming is geometrically connected), the -signature of is at most (4.8).
-
(4)
In case is a toric Fano variety corresponding to a fan , then is the same as the complex -invariant of , the complex toric Fano variety corresponding to (4.12).
Part (1) of 1.2 is surprising since many complex Fano varieties have -invariants greater than 1/2 (and less than 1). Parts (1) and (4) of 1.2 together recover, and provide a positive characteristic proof of the well-known fact that the -invariant of toric Fano varieties is at most 1/2 (see [LZ22, Corollary 3.6]).
The -invariant (like the complex -invariant) can be defined for any pair , where is a globally -regular projective variety and is an ample line bundle on . So, in Section 3, we develop the theory of the -invariant in this more general setting. From this perspective, the -invariant is an asymptotic invariant of a section ring of a projective variety that shares many properties and relations with the -signature. In this direction, we prove:
Theorem 1.3.
Let denote the section ring of a globally -regular projective variety over a perfect field , with respect to some ample line bundle over . Then,
Our third set of results concern the semicontinuity properties of the -invariant in families of globally -regular varieties (5.2), analogous to the results of [BL22] about the complex version. In 5.8, we prove:
Theorem 1.4.
Let be family of globally -regular Fano varieties such that is -Cartier and -ample. Assume that is regular. Then, the map given by
is lower semicontinuous, where is the perfectified-fiber over .
We also prove a weaker version of 1.4 for any polarized family of globally -regular varieties (see 5.3). This is analogous to the corresponding result for the -signature proved in [CRST21] and relies on understanding the rate of convergence in Part (1) of 1.3 for the -invariant, which may be of independent interest (see 5.4).
Finally, we consider some examples that highlight interesting features of the -invariant. For instance, we see that the -invariant does not detect regularity of a section ring (6.1). In 6.4, we observe that when viewed through the lens of reduction modulo , the -invariant may depend on the characteristic. Furthermore, while the limit of the -invariant as may exist and have an interesting geometric interpretation, the limit is not necessarily the complex -invariant. This fact raises interesting questions and obstructions to relating log canonical thresholds and -pure thresholds of divisors on a klt variety. We hope that these observations and questions will lead to other results in the future.
Acknowledgements
I would like to thank my advisor Karen Smith for her guidance, support and many helpful discussions. I would like to thank Harold Blum and Yuchen Liu for their valuable guidance that led to some of the main theorems of this paper. I am thankful for the many useful suggestions given by Devlin Mallory, Mircea Mustaţă, Karl Schwede, Kevin Tucker, and Ziquan Zhuang. I would also like to thank Anna Brosowsky, Seungsu Lee, Linquan Ma, Shravan Patankar, Anurag Singh, David Stapleton, Vijaylaxmi Trivedi and Yueqiao Wu for helpful conversations. I thank Devlin Mallory, Mircea Mustaţă and Austyn Simpson for detailed comments on an earlier draft. Parts of this work were conducted during visits to the University of Utah, the University of Illinois at Chicago, the Tata Institute of Fundamental Research, the Indian Statistical Institute (Bangalore), and the Indian Institute of Science. I thank all these institutes for their facilities and hospitality.
2. Preliminaries
Notation 2.1.
Throughout this paper, all rings are assumed to be Noetherian and commutative with a unit. Unless specified otherwise, will denote a perfect field of characteristic . A variety over is an integral (in particular, connected), separated scheme of finite type over . For a point on a scheme , the residue field will be denoted by (where is the local ring at and is the maximal ideal of the local ring).
Notation 2.2 (Divisors and Pairs).
A prime Weil-divisor on a scheme is a reduced and irreducible subscheme of of codimension one. An integral Weil-divisor is a formal -linear combination of prime Weil-divisors. A -divisor is a formal -linear combination of prime Weil-divisors. By a pair , we mean that is a Noetherian, normal scheme and is an effective -divisor over . A projective pair is a pair where is a projective variety over .
2.1. Section Rings and Modules
Definition 2.3.
Let be a Noetherian ring and be a projective scheme over . Given an ample invertible sheaf on and a coherent sheaf on , the -graded ring defined by
is called the section ring of with respect to . The affine scheme is called the (affine) cone over with respect to . The section module of with respect to is a -graded -module defined by
Similarly, the sheaf corresponding to on is called the cone over with respect to .
Let be a Noetherian, -graded domain, and denote the set of positive degree homogeneous elements of . For a finitely generated, torsion-free, -graded module over , let denote the localization . Note that is naturally a -graded module over . Since is torsion-free, we can think of naturally as a subset of . In this setting, we define the saturation of to be the -graded module
where is the irrelevant ideal . We say is saturated if .
Lemma 2.4.
Let be a Noetherian domain, be an integral, projective scheme over and an ample invertible sheaf over .
-
(1)
The section ring of with respect to is a finitely generated algebra over and hence, is Noetherian. If is normal, and , then the section ring is also characterized as the unique normal -graded ring such that is isomorphic to and the corresponding is isomorphic to .
-
(2)
The section module of any torsion-free coherent sheaf over with respect to is finitely generated over . It is also characterized as the unique saturated, torsion-free, -graded -module (with respect to the irrelevant ideal ) such that the associated coherent sheaf on is isomorphic to .
-
(3)
For two torsion-free coherent sheaves and , we have a natural isomorphism:
where denotes the set of grading preserving -module maps between two graded -modules.
2.2. Affine and Projective Cones
Given an integral, projective scheme over a noetherian domain , and an ample invertible sheaf over , let be the corresponding section ring. Assume that . Let be the projective -scheme defined as
where is the -graded ring obtained by adjoining a new variable to in degree . Then, is called the projective cone over with respect to . Denoting by the homogeneous irrelevant ideal of , we have a map of graded -algebras
that induces the “zero-section" map
over . We call this map the “vertex of the cone ". We also have natural maps of graded rings
which, via the construction in [Har77, II, Exercise 2.14 (b)] induce maps
called the “section at infinity" and
where is the zero-section described above. It follows from the Proj construction that is an -bundle over . The affine cone is isomorphic to , and the zero section actually maps into . Thus, restricts to a map that is a -bundle over . See [HS04, Section 2] and [Kol13, Section 3.1] for details.
2.3. Cones over -divisors
We follow the description of the cone over a -divisor as in [SS10, Section 5], where it is explained for a projective variety over a field. Essentially the same description holds in the following more general relative setting: Let be a normal domain of finite type over , and be an integral, normal, projective scheme over . Assume that is flat over and of positive relative dimension. Fix an ample invertible sheaf over and be the corresponding section ring. Assume further that . Note that this guarantees that the codimension of the zero section is at least two in . Thus, it follows that is normal as well. In this situation, given any integral Weil-divisor (for distinct prime Weil divisors ) on , we can construct the corresponding Weil divisor on , the “cone over ", in three equivalent ways:
-
(1)
Let be the prime Weil divisor on corresponding to the height one prime corresponding to . Then .
-
(2)
Let be the reflexive sheaf on corresponding to . Then, is the divisor corresponding to the reflexive -module defined by
Note that the fact that is reflexive can be seen by applying Part (3) of 2.4 to each twist of the graded module .
-
(3)
Let be the -bundle map defined in the previous paragraph. Then, we may define to be the pull back of to . More precisely, near the generic point of a component of , if is given by an equation , then is is defined by . We then take closures to obtain a Weil-divisor on . This defines a unique divisor on since is flat and is at least .
Given a principal divisor on defined by the rational function , the cone over can be seen to be the principal divisor defined by again. The construction of the cone clearly preserves addition of divisors. This implies that cone construction extends to -divisors and preserves the linear equivalence of Weil-divisors. Furthermore, by taking closures, this construction also extends to the projective cone described in the previous paragraph. Finally, using the third description of the cone, we see that the cone over the canonical (Weil-)divisor is the canonical divisor of (recall that is also normal).
2.4. -signature
Let be any ring of prime characteristic . Then for any , let sending be the -iterate of the Frobenius morphism. Since has characteristic , defines a ring homomorphism, allowing us to define a new -module for each obtained via restriction of scalars along . We denote this new -module by and its elements by (where is an element of ). Concretely, is the same as as an abelian group, but the -module action is given by:
Now let denote a normal local ring and denote the normal scheme . Throughout, we will assume that is the localization of a finitely generated -algebra at a maximal ideal, which also makes it -finite (i.e., is a finitely generated -module for any ), with the rank of over being , where is the Krull dimension of . Let be an effective -divisor on . Then, note that since is effective, for any , we have a natural inclusion of reflexive -modules. Here, denotes the -module corresponding to the reflexive sheaf . Thus, applying to the natural inclusion , we get
Thus, given any element , it can be naturally viewed as a map .
Definition 2.5 (Splitting Ideals).
For any , we define the subset as
We observe that is an ideal of finite colength in and we call
the -free rank of , where denotes the length as an -module.
2.5. -signature of cones over projective varieties.
In this subsection, we describe how we can compute the -signature of cones over projective varieties (and pairs) using global splittings on . We begin with a useful lemma that relates global Frobenius splitting of a divisor to splitting “on the cone". This is a slight generalization to the relative setting of [Smi00, Theorem 3.10], where it is proved over a field.
Lemma 2.7.
Let be a regular ring of finite type over and be an integral, normal projective scheme over (with ). Assume that is flat over and of positive relative dimension. Fix an ample invertible sheaf and be the corresponding section ring. Fix an effective Weil divisor over and be the cone over with respect to . Then, for any , the natural map
splits as a map of -modules if and only if the map on the cones
splits as a map of -modules.
Proof.
Using the second description of the cone over a -divisor in Section 2.3, and Part (3) of 2.4, the proof is exactly the same as that in [SS10, Proposition 5.3]. ∎
Returning to working over a perfect field , we next recall a formula to compute the -signature of the section ring of a projective variety proved in [LP23]. Fix a normal projective variety over and be an effective -divisor over .
Definition 2.8.
For any Weil-divisor on and , define the -vector subspace of as follows:
Remark 2.9.
Note that the subspace only depends on the sheaf and not on the specific divisor in its linear equivalence class.
Remark 2.10.
Lemma 2.11.
[LP23, Lemma 4.7] Let be an ample Cartier divisor on and denote the section ring of with respect to . Let denote the cone over with respect to (Section 2.3). Then, for any , if denotes the -free-rank of (2.5), then is computed by the following formula:
(2.1) |
where denotes the field . Hence, the -signature of with respect to can be computed as
2.6. Duality and the Trace Map
It will be convenient to think of the subspaces (2.8) using a pairing arising out of duality for the Frobenius map. Let be a projective pair and be any Weil divisor on . We continue to work over any perfect field of characteristic . But in this subsection, we assume that , i.e., that is geometrically connected.
Recall that by applying duality to the Frobenius map, we get the following isomorphism of reflexive -modules:
(2.2) |
See [SS10, Section 4.1] for a detailed discussion regarding duality for the Frobenius map. Furthermore, when , this gives an isomorphism
(2.3) |
Composing this isomorphism (over the global sections) with the evaluation at map, we obtain the trace map:
(2.4) |
Lemma 2.12.
The kernel of the trace map in Equation 2.4 is exactly the subspace . See 2.8 for the definition of the subspace .
Proof.
A section is contained in the corresponding -subspace if and only if for every
we have . But, is a one dimensional -vector space, and is generated by the trace map (where ). Now, the map is just the restriction of the trace map to the subspace . Thus, the lemma follows. ∎
Lemma 2.13.
Let be a normal projective pair, and be any Weil divisor on . Then, denoting and for any , we have a non-degenerate pairing
obtained from multiplication (and reflexifying) global sections. In particular,
Proof.
Using Equation 2.2, the natural multiplication map
can be identified with the evaluation map
where we identify and as subspaces of and respectively, both via the natural inclusion . Therefore, a section is contained in if and only if for all sections , the multiplication is contained in . By symmetry, a section is contained in if and only if for all sections , . This proves there is a well defined, and non-degenerate pairing as needed.
Finally, we note that since by 2.12, is the kernel of the trace map (Equation 2.4), the vector space is either one-dimensional over , or equal to . In either case, the equality of dimensions follows. ∎
The next Proposition allows us to perturb by any divisor while computing the -signature of a section ring.
Proposition 2.14.
Let be a normal projective varitey over and an ample divisor on . Assume . Fix a (not necessarily effective) Weil divisor on . Then, there exists a constant (depending only on and ) such that
for all and .
Proof.
First we prove the case when is effective: For any , by using the natural map , we will view as a subspace of . Let denote the subspace . By [LP23, Lemma 4.12], we see that . Moreover, by Equation 4.12 in [LP23, Proof of Lemma 4.14], we have
Using this and the triangle inequality, we obtain that
(2.5) |
for all . Next, to compute the second term in the above inequality, fix an and set . Then, we observe that the subspace is exactly the same as (2.8). Moreover, we also similarly have
(2.6) |
Thus, by 2.13, we have
and similarly,
By Equation 2.6, we see the natural map from to restricts to an injective map
By considering the cokernel of this map, we get that
(2.7) |
Pick an such that admits a global section that doesn’t vanish along for all (this is possible since is ample). Using the standard exact sequences to restrict to , we see that
(2.8) |
and
(2.9) |
Finally, by [LP23, Theorem 4.9], we may pick a constant such that and for . Therefore, to prove the Proposition, it is enough to consider the case when . In this case, the Proposition now follows by putting together inequalities in 2.5, 2.7, 2.8 and 2.9. This completes the proof of the Proposition when is effective.
More generally, we first pick an such that and are both effective. Then, for any and , we have
Now, we may apply the previous case of the Proposition (since both and are effective) to each of the two terms in the above inequality. Since was independent of , this completes the proof of the Proposition . ∎
2.7. -regularity:
Definition 2.15 (Sharp -splitting).
[SS10, Definition 3.1] Let be a normal variety over and be an effective -divisor. The pair is said to be globally sharply -split (resp. locally sharply -split) if there exists an integer , such that, the natural map
splits (resp. splits locally) as a map of -modules. A normal variety is said to globally -split if the pair is globally sharply -split.
Definition 2.16 (-regularity).
[SS10, Definition 3.1] Let be a normal variety over and be an effective -divisor. The pair is said to be globally -regular (resp. locally strongly -regular) if for any effective Weil divisor on , there exists an integer , such that, the natural map
splits (resp. splits locally) as a map of -modules. A normal variety is said to globally -regular if the pair is globally -regular. Similarly, a ring is called strongly -regular if the pair is locally (equivalently, globally) strongly -regular.
Remark 2.17.
When is an affine variety and is an effective -divisor, the pair being globally -regular (resp. globally sharply -split) is equivalent to the pair being locally strongly -regular (resp. locally sharply -split) [SS10].
Remark 2.18.
Theorem 2.19.
Remark 2.20.
(Locally) Strongly -regular varieties are normal and Cohen-Macaulay. Similarly, globally -regular varieties enjoy many of nice properties such as:
Theorem 2.21 ([Smi00], Corollary 4.3).
Let be a projective, globally -regular variety over . Suppose is a nef invertible sheaf over . Then,
We need a slight variation of 2.21 for -ample divisors that we prove here for completeness.
Proposition 2.22.
Let be a globally -split normal variety and be a -ample Weil divisor i.e., is an integral Weil divisor such that is an ample Cartier divisor for some integer . Then,
Proof.
Let be an integer such that is Cartier. Write such that is coprime to . Pick an such that divides . Then, since is a multiple of for all , using Serre vanishing theorem, we have
(2.10) |
Since the map
is split, twisting by and reflexifying, we get that
is split as well. Now the Proposition follows from the vanishing in (2.10). ∎
The next Proposition is a technical result that helps us to restrict -Cartier divisors to normal, locally complete intersection subvarieties. This is very close to [PS12, Corollary 3.3], but we will need it in the form stated below.
Proposition 2.23.
Let , where is an -finite, strongly -regular, local ring and be an integral Weil-divisor on such that is Cartier for some integer . Then, for each ,
-
(1)
There exists an such that the module is isomorphic to an -module summand of . In particular, is a Cohen-Macaulay module over .
-
(2)
Suppose that is a regular sequence on such that the ring is normal. Then, the sheaf is reflexive on . Furthermore, if we assume that the support of does not contain the subscheme , then natural map
(2.11) is an isomorphism, where denotes the restriction of to (see the description in the proof below).
Proof.
By adding a principal divisor if necessary (which leaves the module isomorphic), we assume that is effective.
-
(1)
Since is strongly -regular, there exists an such that the map splits for each . Once we have such an , choose such that is divisible by . Thus, we may assume that the map (got by twisting by and reflexifying):
is split. Since divides , is isomorphic to since is local and is Cartier. Therefore, taking global sections, we have that the map
(2.12) is split. Note that the obtained is independent of . The Cohen-Macaulayness of follows because is a Cohen-Macaulay module over , since itself is Cohen-Macaulay.
-
(2)
Firstly, we may assume since the discussion holds for an arbitrary Weil divisor and is compatible with addition of Weil-divisors. Now, since is a normal, complete intersection subscheme of , we may “restrict" the rank one reflexive sheaf on to a reflexive sheaf on as follows: Let be the regular locus of . Then there is an open subset (where denotes the regular locus of ) such that . This is possible because is a complete intersection in . Therefore, we may restrict to and then to an invertible sheaf on , since is invertible. Define to be
where is the inclusion. Then, is a rank one reflexive sheaf on because is normal and contains all the codimension one points of . Thus, we can write as for some Weil-divisor on . Furthermore, if does not contain , then since is normal, hence integral, naturally restricts to a Cartier divisor on (given by restricting the equation for ) and we may take to be the closure of . It is also clear from the description of restriction that it commutes with addition of Weil-divisors (since the restriction of Cartier divisors on the regular locus commutes with addition).
Now, since is the restriction of the sheaf (i.e., isomorphic to ), there is a natual map
which is an isomorphism if and only if is reflexive. Therefore, it is sufficient to show that the module satisfies the S2 condition on (since is normal). But since is Cohen-Macaulay by Part (1) (and clearly full dimensional), and is a regular sequence on , we get that is Cohen-Macaulay as well. This completes the proof of the Proposition. ∎
3. The -invariant of section rings
In analogy with Tian’s -invariant in complex geometry, we define the “Frobenius-" invariant (denoted by ) for any pair where is a globally -regular projective variety (2.16) and is an ample Cartier divisor on .
Throughout this section, by a section ring , we mean that is the section ring of some projective variety with respect to some ample line bundle on (see 2.3).
3.1. Definitions
Definition 3.1.
If is the principal divisor corresponding to a function , we write instead of in the notation for the -pure threshold.
Definition 3.2.
Let be a strongly -regular section ring of a projective variety. Then, we define
If is the section ring of a projective variety with respect to an ample divisor , we may also use to denote .
We have the following equivalent ways of characterizing the -invariant of a section ring.
Lemma 3.3.
Let be a globally -regular projective variety and be an ample Cartier divisor on . Let be the section ring of with respect to . Then, from 3.2 is equal to the supremum of any of the following sets:
-
(1)
The set of such that the pair is sharply -split for every and every effective divisor .
-
(2)
The set of such that the pair is strongly -regular for every and every effective divisor .
-
(3)
The set of such that the pair is globally sharply -split for every and every effective divisor .
-
(4)
The set of such that the pair is globally -regular for every and every effective divisor .
-
(5)
The set of such that the pair is globally -regular for every effective -divisor .
Proof.
Statements (1) and (2) follow immediately from the definition of the -invariant and the definition of the -pure threshold (3.1). Statements (3) and (4) follow from (1) and (2) by using 2.7. Part (5) is just a reformulation of (4) since every effective -divisor is of the form for some effective Cartier divisor . ∎
Next, we explain a more precise connection between the -invariant and Frobenius splittings in .
Proposition 3.4.
Let be a strongly -regular section ring and denote the supremum of the following set:
Then, . Moreover, belongs to the set .
Remark 3.5.
Note that in the above Proposition, it is unclear if the set contains any non-zero element. This is equivalent to the positivity of and will be addressed in 3.10.
The proof of the 3.4 is based on the following lemma which is a slight generalization of a result of Hernández:
Lemma 3.6.
Proof.
It follows immediately from the definitions that (1) implies (2), and (2) implies (3). Hence, it remains to show that (3) implies (1).
Following [Her12, Thoerem 4.9], if , we must must have that the pair is sharply -split. Thus, there is an such that the natural map
(3.1) |
splits. Since the same holds for for any natural number (see [Sch08, Proposition 3.3] for the proof), we get the map:
splits. Note that as defined in Equation 3.1 matches with the map considered in the statement of the Lemma.
Let denote the regular locus of . Since is a map between reflexive sheaves, to show that it splits, it sufficient to show that its restriction of splits. Over , we may construct the map as follows: First consider the map
obtained by twisting the -iterate of the Frobenius map by the invertible sheaf . If denotes the local equation of , then is defined by sending
Then, after restricting to , we have that
where the right hand side is the composition
Therefore, if splits then so does . This proves that part (3) implies part (1), completing the proof of the lemma. ∎
Proof of 3.4.
Set , and (which is defined in the satement of the Proposition). First we will prove that . This is clear if , so we assume that is positive. For any non-zero element , by definition of , we must have . So, if , we have
Thus, by 3.6, the map sending to splits. Since was an arbitrary non-zero element of degree , this shows that whenever . Therefore, belongs to the set , which proves that .
Next we prove that . Fix any . Then, whenever and is any non-zero element, we know that is not contained in . By 3.6, we have . In other words, if is the smallest integer such that , (equivalently, ), then Now combining this with the fact that for any integer to get:
(3.2) |
To see the right inequality, we make the following observations: Fixing and and for any , write
for some non-negative real number . Then, we have
for each . So it is sufficient to show that
This is true because given any real number , the infimum of the set is zero. This proves the inequality in (3.2) .
Since was an arbitrary non-zero homogeneous element of degree , it follows from Equation 3.2 that . Since was an arbitrary number smaller than , we must have as well. This completes the proof that . ∎
3.2. Finite-degree approximations
Now we will define finite-degree approximations to the -invariant. This establishes a limit formula for the -invariant that is analogous to the -signature (see 2.6 and the classical definition in [Tuc12]).
Definition 3.7.
Let be an -graded section ring over . For each integer , we define
and define
Theorem 3.8.
Let be a strongly -regular -graded section ring over . Then, we have
In particular, the limit exists. See 3.2 for the definition of .
Lemma 3.9.
Let be a strongly -regular -graded section ring over . For any , we have
Proof.
First note that since is strongly -regular, is a normal domain. For any , let be an element of . Then, we have that is a non-zero element of (see [Tuc12, Lemma 4.4]). This proves that
Dividing both sides by , we obtain the required inequality. ∎
Proof of 3.8.
The sequence is decreasing, by 3.9. Since it is a decreasing sequence of non-negative real numbers, the sequence converges to its infimum. Moreover, since the sequence converges to zero, the sequence also converges and
It remains to show that the limit is equal to . Using the definition of , we have that
for each . This is because we know that belongs to the set from 3.4. Taking a limit over , we obtain
For the reverse inequality, setting , we note that
By the definition of , the subspace is equal to zero for each . Thus, belongs to the set defined in 3.4. Since is the supremum of , we get that . This completes the proof of 3.8. ∎
3.3. Positivity and comparison to the -signature.
Next we will show that the -invariant is positive by comparing it to the -signature (2.6). Recall that for a section ring of any globally -regular projective variety, there exists a positive constant such that for any and any , we have This follows from [LP23, Theorem 4.9].
Theorem 3.10.
The -invariant of a strongly -regular section ring is positive. Moreover, setting and fixing a constant as discussed above (so that for any and any , we have ), we have the following comparisons:
(3.3) |
where denotes the Hilbert-Samuel multiplicity of .
Lemma 3.11.
Given a non-zero homogeneous element in of degree , let be a real number. Then,
(3.4) |
Proof.
Since we have assumed that , there exist integers such that
Replacing by , by 3.6 we may assume that the map defined by does not split (since . We may also assume that . Now, since belongs to the ideal , we have for any yielding the inequality
(3.5) |
for all . Further, setting for any integer , we have that , and so belongs to . Therefore, we similarly have
(3.6) |
for all . Then, using the 2.11 we may compute the -signature as follows:
where is the field . Here we have used Equation 3.6 and the defining property of the constant . Finally, calculating the dimensions in the above inequality using the formula
we obtain
The proof of the lemma is now complete by using the fact that . ∎
Proof of 3.10.
We note that if , then the rightmost inequality of Equation 3.3 implies that the -signature is zero. But this is a contradiction since was assumed to be strongly -regular (see [AL03, Theorem 0.2]). So the positivity of follows from Equation 3.3, which we will now prove.
The rightmost inequality follows from 3.11 by taking a limit as , since the Lemma applies to each such that for some non-zero .
To prove the leftmost inequality, we use 2.11 again to compute the -signature of and we observe that for any ,
(3.7) |
Recall that for any , is the largest such that , which justifies Equation 3.7. But the right hand side is equal to
We conclude the proof by using 3.8, which says that . This concludes the proof of Equation 3.3 and thus of 3.10. ∎
3.4. Behaviour under certain ring extensions.
In this subsection, we record some useful results on the behaviour of the -invariant under suitably nice extensions of section rings.
Proposition 3.13.
Let and be two -graded section rings (of possibly different varieties) and be an inclusion such that for a fixed integer and any other , all degree elements of are mapped to degree elements of . Further, assume that the inclusion splits as a map of -modules. Then, we have
(3.8) |
Moreover, equality holds in (3.8) if is the -Veronese subring of .
Proof.
The first part follows immediately from 3.8 and the fact that a homogeneous element of splits from if it splits from . In other words, we have
for any and . For the statement about Veronese subrings, the key observation is that since is a section ring (of , say), then
Thus, the equality again follows from 3.8. ∎
Remark 3.14.
Proposition 3.15.
Let be a degree preserving map of -graded, strongly -regular section rings (of possibly different varieties and over possibly different perfect fields). Assume that both and are generated in degree one , is flat over , and that the ring is regular. Then, we have
Proof.
Recall that was assumed to be a perfect field of characteristic .
Corollary 3.16.
Let be a strongly -regular section ring over and be an arbitrary perfect field extension of . Then, the base-change is isomorphic to a product of strongly -regular section rings over finite extensions of and for each , we have
Proof.
Firstly, we may assume that is generated in degree one by using 3.13. Set . Since is perfect and is a finite separable extension of , we see that
is a finite product of perfect fields . Thus, if was the section ring of , then where is the section ring of . Note that is isomorphic to , and hence each is strongly -regular by [CRST21, Theorem 3.6]. The corollary now follows from 3.15 by applying it to each inclusion . ∎
Remark 3.17.
Remark 3.18.
The results of this section naturally extend to the more general setting of globally -regular pairs such that is an integral Weil-divisor. This will be addressed in a future version of the paper.
4. The -invariant of globally -regular Fano varieties.
In this section, we specialize the study of the -invariant to the case of globally -regular Fano varieties (and when the ample divisor is a multiple of ). We begin by defining what we mean by a -Fano variety in positive characteristic. Recall that denotes a perfect field of characteristic .
Definition 4.1.
A -Fano variety is a projective variety over such that
-
(1)
is locally strongly -regular (2.16).
-
(2)
is a -Cartier divisor.
-
(3)
is ample.
Note that since has only strongly -regular singularities, is automatically normal and Cohen-Macaulay. In particular, we may define the canonical Weil-divisor by extending a canonical divisor from the smooth locus. In fact, is a dualizing sheaf over . In particular, we have
(4.1) |
where is the dimension of . Moreover, the second and third conditions in Definition 4.1 guarantee that there is a positive integer such that is Cartier and is ample. The smallest such is called the index of .
Definition 4.2.
Let be a globally -regular -Fano variety over and be a positive integer divisible by the index of . Let
denote the section ring of with respect to . Then, the -invariant of is defined to be
where denotes the -invariant of the strongly -regular ring , as defined in Defintion 3.2.
By taking the affine cone over a -Fano variety, we also define the global -signature of the a -Fano variety.
Definition 4.3 (-signature).
Let be a -Fano variety over and denote a positive integer divisible by the index of . Let
denote the section ring of with respect to . Then, the -signature of is defined to be
where denotes the -signature of , as defined in Defintion 2.6.
Remark 4.4.
Theorem 4.5.
Let be a globally -regular -Fano variety of positive dimension. Then, is at most .
Proof of 4.5:.
Let denote the dimension of , and be an integer divisible by the index of and such that for all . First, we claim that there is an integer such that
for all . This is clear if is a Cartier divisor (and we may take in this case), since is a polynomial in of degree and a positive leading term (because is ample). More generally, by the asymptotic Riemann-Roch formula ([Laz04, Example 1.2.19]), for each , there exists polynomials of degree such that for all
Moreover, setting , each has the form
for polynomials of degree at most . In this situation, the existence of an integer as required is guaranteed by 4.6 stated and proved below.
Assume, for the sake of contradiction, that for some small . Then, note that by 3.4, for all , we have for all .
Now, for , we can find an integer satisfying the following properties:
-
•
,
-
•
, and,
-
•
.
This is equivalent to finding an integer such that
which is possible since is fixed and as . The third condition on guarantees that
(4.2) |
The second condition guarantees that there exists a non-zero effective Weil divisor that induces an injective map
for some . Therefore, we know that as noted above. By [LP23, Lemma 4.12], this implies that as well. Finally, note that 2.13 applied to tells us that
But since both the -subspaces in the above equation are zero, this is in contradiction to Equation 4.2. This proves that is at most 1/2. ∎
The following lemma was used in the proof of 4.5.
Lemma 4.6.
For each , let be a polynomial with real coefficients. Moreover, assume that all the ’s have the same positive degree and the same positive leading term. In other words, for each , there exists a real polynomial of degree at most such that
for some real number (independent of ). Then, there is an integer such that for all integers , and all pairs , we have
Proof.
Since each is a real polynomial of degree at most , we may find positive constants such that
for each and each . Now, it is sufficient to find a constant such that
(4.3) |
for all . For this, expanding using the binomial theorem, Equation 4.3 is equivalent to
First, we note that we may choose such that is larger than both and . This implies that for all . Furthermore, having chosen the as before, we have for all . This proves the lemma. ∎
For the rest of this section, we assume that , i.e., that is geometrically connected. However, see 3.17 for ways to extend the results to more general cases. In the case of Fano varieties, we have a stronger version of comparison of the -invariant to the -signature than the formula in 3.10:
Theorem 4.7.
Let be a -dimensional globally -regular -Fano variety over . Assume that and that is positive. Set . Then, we have the following inequalities relating the -signature and the -invariant:
(4.4) |
Here, denotes the volume of the -Cartier divisor .
Corollary 4.8.
Let be a globally -regular -Fano variety of dimension . Then,
-
(1)
We have
-
(2)
Moreover, is equal to if and only if the value of the -signature is equal to
For -Fano varieties, the -signature has a more refined formula than 2.11, which we prove next.
Proposition 4.9.
Let be a globally -regular -Fano variety over . Assume that . Let be an integer such that is Cartier. Then, the -signature of can be computed as
Lemma 4.10.
Let be a globally -regular -Fano variety that is geometrically connected over . Then, for any , we have whenever .
Proof.
We will prove this lemma in two different ways, since both ideas may be useful in other situations.
Proof 1:
Let . By definition of the subspace (2.8), it is sufficient to show that there are no non-zero maps . We have
By assumption, we have . Since is ample, this means that . Hence, there are no non-zero maps , which proves the lemma.
Proof 2:
Let and suppose that there is a non-zero global section that is not in . Then, by definition of , we have a map
such that . Thus, we have the splitting
(4.5) |
where the first map is got by sending and the second map is . Twisting equation 4.5 by and reflexifying, we obtain a splitting:
(4.6) |
By assumption, is non-negative. Hence, by 2.22, in turn implying that (using the splitting in equation 4.6). This is a contradiction, since is the canonical sheaf of . This completes the proof of 4.10. ∎
Proof of 4.9.
Let denote the section ring of with respect to . Fix an and let denote the free rank of as an -module (2.5). Recall that by 2.11, we have
(4.7) |
so that
Then, 4.10 shows that the terms of the sum in Equation 4.7 are zero for . Furthermore, using 2.13, we have
Let be an integer between and such that . Hence, we have
Thus, we have
Moreover, using 2.14, we see that there is a constant such that
Thus, we have that
The proof is now complete since the right hand side limits to zero when divided by and as . ∎
Proof of 4.7.
The proof of this Theorem is exactly the same proof as the proof of Equation 3.3 in 3.10 (see the proof of 3.11), once we replace the formula from 2.11 with the formula from 4.9 to compute the -signature of . ∎
Proof of 4.8.
Part (1) follows immediately from the right-hand inequality in 4.7, since we know that by 4.5. We also see that if , we must have
Thus, Part (2) also follows from Equation 4.4 once we note that when , both sides of the inequality in Equation 4.4 are equal to . ∎
Remark 4.11.
Let be a -Fano variety over . This means that is a normal variety, is -Cartier and ample and has only klt singularities. Let be such that is Cartier, and . Then, for any effective -divisor on with , we have
where lct denotes the log canonical threshold and denotes the cone over . This follows from [Kol13, Lemma 3.1]. Thus, if we let
then, we have that
Therefore, for any -Fano variety with , the -invariant from 4.2 is a “Frobenius analog" of the complex -invariant.
The -invariant of toric Fano varieties
Let denote an algebraically closed field of prime characteristic . Fix a lattice and let be the dual lattice (where is some positive integer).
Theorem 4.12.
Let be a -Fano toric variety over defined by a fan in . Let be the corresponding complex toric variety (which is also automatically -Fano). Then, we have
Proof.
Let denote the primitive generators for the one dimensional cones in and write for rational numbers .
First we choose an such that for each and the section ring is generated in degree one. Let denote the polytope associated to , and defined by:
Since we are assuming that is generated in degree , the vertices of are lattice points of . For any , let be the corresponding effective divisor in the linear system . By [BJ20, Corollary 7.16], we have that
(4.8) |
where denotes the log canonical threshold of a divisor on . Note that since the vertices of are lattice points, just the vertices are sufficient to compute .
Let denote the polytope . Then, the section ring is the semigroup ring associated to the cone over in . Note that is -Gorenstein. Therefore, by [Bli04, Theorem 3], we see that for any , we have
(4.9) |
Next, note that since is a normal toric variety, it is automatically globally -regular. Now we prove that the -invariant of can also be computed by only considering the torus invariant divisors. To see this, let and let be a non-zero homogeneous element. Then, following the discussion in [BJ20, Section 7.4] and [Eis95, Theorem 15.17], there exists an integral weight vector with such that . Here denotes the weight monomial order with respect to and denotes the graded lexicographic monomial order on . Then, we have a flat degeneration of to its initial term. In other words, if for monomials , then setting , the element
satisfies the following properties:
-
•
Viewing as a -algebra, the ring is a flat -module.
-
•
The image of modulo is equal to , the initial term of with respect to the graded lex monomial order on .
-
•
For any point , the image of in satisfies
With this construction in place, we conclude the proof of the theorem with the following lemma:
Lemma 4.13.
For any non-zero homogeneous element of , we have
Assuming this lemma for a moment, we see that
Furthermore, by Equation 4.9, we have
(4.10) |
Since by 4.5 we have , we must have for some . Note that corresponds to a torus-invariant divisor on linearly equivalent to . Therefore, by 4.11, we have
for any such that . Putting this together with Equation 4.10 and Equation 4.8, we get that
as required. ∎
Finally, it remains to prove 4.13.
Proof of 4.13.
By 3.6, it is sufficient to show that for all rational numbers of the form such that
the map sending to splits. Equivalently, for all such , it suffices to show that . Since the pair is strongly -regular, in particular it is sharply -split. By 3.6 again, we know that . Now, since is a toric ring, is a monomial ideal of . Therefore, if , we also have as required. ∎
Remark 4.14.
Remark 4.15.
The results of this section also extend naturally to the case of globally -regular log-Fano pairs such that is an integral Weil-divisor. This will be addressed in a future version of the paper.
5. Semicontinuity properties.
In this section we will examine the behaviour of the -invariant in geometric families, analogous to the case of the -signature which was proved in [CRST21] and the case of complex -invariant discussed in [BL22]. Throughout this section, we fix to be an algebraically closed field of characteristic .
5.1. Weak semicontinuity
First, we will prove a result for a family of arbitrary globally -regular varieties.
Notation 5.1.
Recall that the perfection of a field (of positive characteristic) is the union
of all the -th roots of elements of . For a map of varieties over , and a point (not necessarily closed), we denote the fiber of over as and denote the perfectified fiber over as
Similar for a coherent sheaf on , we have the fiber and the perfectified fiber . Furthermore , if and is a finitely generated -algebra, then for any , we denote the perfectified fiber by .
Notation 5.2.
Recall that as in 3.2, we can consider the -invariant of a pair where is a globally -regular projective variety (over a perfect field) and is an ample line bundle over .
Theorem 5.3.
Let be a flat family of globally -regular varieties, where is an ample line bundle over . Let denote the fraction field of and denote the -invariant of , the perfectified generic fiber. Then, for each real number , there exists a dense open subset such that
Recall that in 3.8, we defined the sequence that converges to the -invariant. To prove 5.3, we need to understand the rate of convergence in 3.8. For this we will use “degree-lowering operators" as below.
Theorem 5.4.
Let be a projective globally -regular variety over and be an ample invertible sheaf over . Suppose we have integers and such that the sheaf
is generically globally generated for each . Then, if denotes the section ring , we have
Proof.
Set . First, we claim that for each , we can find an injective map of -modules
(5.1) |
To see this, let denote the generic point of and consider the following restriction map to the generic stalk:
(5.2) |
The assumption that is generically globally generated means that the image of the map in Equation 5.2 generates as an -module. Recall that is just the fraction field of . Since is a free -vector space of rank , we can choose maps in such that their images under the map in Equation 5.2 forms a basis of over . Thus, defining to be the product map
we see that is generically an isomorphism by construction. Furthermore, since is a torsion-free sheaf of rank over , is injective since it is generically an isomorphism. This completes that proof of the claim that maps as in Equation 5.1 exist.
Now, fix a map for each as in Equation 5.1. Then, by taking section modules with respect to (2.3), we get a corresponding map of graded -modules
and let
denote the direct sum of the ’s. Here, for any , denotes the -module
and we naturally have an -graded -module decomposition
Note that is injective because the ’s were injective. Furthermore, the key property of that we will use is the following: for each non-zero homogeneous element of degree in , is a non-zero homogeneous element of degree
(5.3) |
Thus, we may use this map as a “degree-lowering operator".
From 3.8, recall that where is defined to be the number . The condition that means that for all non-zero elements , there exists a splitting that sends to .
Claim: For any , let , and be an integer such that . Then .
We will prove the claim by induction on . If , note that the sum in the claim is empty, and hence we have . In this case, by the definition of .
Now let and be any non-zero element of . We need to show that splits from . To see this, note that by Equation 5.3, is a non-zero element of degree at most
Thus, the inductive hypothesis applies to , implying that is not contained in . Let denote a splitting of . Then, we see that (forgetting the degrees) defines a splitting of , as required. This completes the proof of the claim.
To complete the proof of the Theorem, we note that the claim above implies that for any ,
Letting and using 3.8, we get
(5.4) |
Lastly, note that by 3.9, we already have
which, together with Equation 5.4 completes the proof of the theorem. ∎
Lemma 5.5.
Let be a globally -regular variety of dimension and be an ample and globally generated invertible sheaf. Suppose is such that is linearly equivalent to an effective Weil divisor for all . Then, for all , and each , the sheaf
is generically globally generated.
Proof.
First, by an application of Castelnuovo-Mumford regularity, we prove the following statement:
Claim: For all and all , the sheaf is -regular with respect to , and hence globally generated.
To see this, we check that
Here, we have used the projection formula to see that and the cohomology vanishing follows from (2.21), since is nef for all . Thus, is -regular with respect to , and hence is globally generated (see [Laz04, Theorem 1.8.5] for the details regarding Castelnuovo-Mumford regularity).
Next, for any and , write where is an effective Weil-divisor. Note that it is always possible to find such an thanks to [SS10, Theorem 4.3]. Let denote the map obtained by twisting the defining map for by and pushing forward under :
Note that for any point , restricts to an isomorphism in an open neighbourhood around . For any sheaf , let denote the stalk of at .
Applying duality for the Frobenius map (Equation 2.2), we have (for any ):
Therefore, for any and , set and consider the diagram
where is any point not contained in . Since the horizontal arrows are injective and the bottom horizontal arrow is an isomorphism, any set of global sections generating , viewed as global sections of , will also generate . Since by the claim above is globally generated, we have that is globally generated at any (and hence generically globally generated). This completes the proof of the lemma. ∎
Lemma 5.6.
Let be a flat family of globally -regular varieties over where is regular. Let be an integral Weil-divisor such that is Cartier for some integer . We also assume that is ample.
-
(1)
Then, for any integer , the sheaf is locally free on and for any point (not necessarily closed), the natural “restriction to the fiber" map
is an isomorphism. Moreover, for any , there exists an affine open neighbourhood of such that if the closure of is defined by a regular sequence and for any , the natural map
is an isomorphism where is the ideal .
-
(2)
Suppose, in addition that is a locally strongly -regular variety (2.16). Then, for any , setting , we have
-
(a)
is flat over .
-
(b)
For any , the restriction is reflexive.
-
(c)
is locally free on and for any , the natural map
is an isomorphism.
-
(d)
Assume that does not contain any fiber of . Then,
is an isomorphism as well. Here, we restrict the Weil-divisor to as explained in 2.23.
-
(a)
Proof.
Note that since is an invertible sheaf on , it is flat over for any .
-
(1)
Since is flat over , the claim follows from Grauert’s Theorem [Har77, Chapter III, Corollary 12.9] once we note that the function
is constant on . To see this, we note that if denotes the Euler-characteristic for sheaf-cohomology, we have
where denotes the generic point of . Here, we are using the higher cohomology vanishing (2.21) for nef invertible sheaves on the fibers (which are globally -regular varieties by assumption), and the fact that the Euler-characteristic is constant for all fibers of a flat map [Har77, III, Theorem 9.9]. Since the restriction of to each fiber of has vanishing higher cohomology, Grauert’s Theorem also implies that is zero for any .
Fix a and choose an open neighbourhood of such that is generated by a regular sequence and is regular (where is the prime ideal of corresponding to ). This is possible because is regular by assumption. Then, note that the following exact sequence of sheaves on :
remains exact after applying since . This tells us that
Now, since is also regular, we may proceed inductively to complete the proof of Part (1).
-
(2)
Fix any and let be the local ring at and be the local ring of any point mapping to .
-
(a)
Since is strongly -regular by assumption, we may apply Part (1) of 2.23 to conclude that is isomorphic to a summand of . Since is regular, we have is flat over and by assumption is flat over . Thus, we see that is flat over and consequently, is flat over because it is a direct summand of .
-
(b)
Fix a regular sequence on generating the maximal ideal (this is possible because is regular). Because is flat over , is also a regular sequence on . Now, it is sufficient to show that is reflexive over . Note that is normal by [Mat89, Theorem 23.9] since all the fibers of are assumed to be globally -regular, which in particular implies that they are normal. Now the fact that is reflexive over is guaranteed by Part (2) of 2.23.
- (c)
-
(d)
This is immediate by combining Part (c) with the last part of 2.23. ∎
-
(a)
Remark 5.7.
Though 5.6 is stated for restriction to the fibers of , all parts of the lemma hold if we further base-change to the perfectified-fibers by flat base-change [Har77, III, Proposition 9.3] and using the fact that the perfectified-fibers of are also normal by assumption (since they are globally -regular).
Proof of 5.3.
We divide the proof into several steps, since some of the steps will be used again in the next subsection.
Step 1:
By 3.13, we may replace by a multiple of if necessary and assume that is globally generated on . In that case, the restriction of to each perfectified-fiber of is also automatically globally generated.
Step 2:
Recall the sequence converging to the -invariant introduced in 3.8. Putting together 5.5 and 5.4, we get that for each , and each ,
(5.5) |
Here, denotes the section ring of the perfect fiber of over with respect to the restriction of , and denotes the dimension of every fiber of (which is well-defined because is flat).
Now, given any , let . Fix an such that . Then, applying Equation 5.5 to the (perfectified) generic fiber of , we have
(5.6) |
Claim:
There exists a dense open set (depending on ) such that
for each . Assuming the claim, Equation 5.5, and Equation 5.6 together imply that
for every as required.
Step 3:
Fix an as in Step 2. We now proceed to prove the claim used above, i.e., that there exists a dense open set (depending on the choice of ) such that for every . Working locally, we may assume that and let denote the section ring . Set . By 5.6, is a locally free -module for any . So by shrinking around any point if necessary, we may assume that is a free -module with a basis for each . By 5.6 again, restricts to a basis of for any . Let . Note that is at most since for any non-zero section (which exists because is assumed to be globally generated), we have .
Step 4:
By definition of (3.7), for any and any non-zero , the map sending to splits. Thus, using [LP23, Lemma 2.7 (a)] repeatedly on the set (fixed in Step 3), we can construct a surjective map such that if is a basis element, then , where is the standard basis element corresponding to of . Considering the induced map on the section modules and putting together the maps for all , along with the zero map for , we get a surjective -module map
(5.7) |
which satisfies the following property: if and , then is a basis element of . In other words, simultaneously splits all the non-zero sections of degree at most on . By [CRST21, Lemma 4.8], there is an integer , a non-zero element such that if , we have a map
which satisfies . Here, denotes the -relative Frobenius over the base change . Note also that after base-changing to , we are identifying the relative and absolute -Frobenius over . Since every is mapped to a basis element of after tensoring to , we may assume that the same is true for after inverting another element of if necessary. Finally, for any , base changing to , we see that
simultaneously splits each non-zero element . Note that we are again identifying the absolute and relative Frobenius over the perfect field . This shows that . This completes the proof of 5.3. ∎
5.2. Semicontinuity for a family of -Fano varieties
In this subsection, we will prove that the -invariant is lower semicontinuous in a family of globally -regular -Fano varieties (see 4.1 for the definition of -Fano varieties):
Theorem 5.8.
Let be flat family of globally -regular -Fano varieties over , i.e., is a family of globally -regular varieties (with regular) such that is -Cartier and -ample. Then, the map from given by
is lower semicontinuous, where is the perfectified-fiber over . See 4.2 for the definition of the -invariant of a Fano variety.
Remark 5.9.
Idea of the proof:
Roughly, the proof of 5.8 involves combining 5.3 with the inversion of adjunction for strong -regularity as proved in [PSZ18]. The main technical difficulty arises when divides the index of . In this situation, we use a standard perturbation trick similar to [Pat14, Lemma 3.15] and [HX15, Lemma 2.13]. But to do this in a family, we need to be able to restrict -Cartier, ample Weil-divisors to the fibers of the family. So we begin by observing that this can indeed be done in our situation.
Setup and Notation:
Let be a regular -algebra of finite type. Suppose we have a flat family of globally -regular Fano varieties (see 5.2) such that is Cartier and ample for some integer . Then, we can form the section ring as described in 2.3. Since is normal and is ample, is a normal, finitely generated , -graded algebra over . For any -algebra , let denote the section ring , where denotes the base change .
Lemma 5.10.
With notation as above, the construction of satisfies the following properties:
-
(1)
is flat over and for any prime ideal we have that
-
(2)
For each prime , there is an affine open neighbourhood containing such that the restriction of to (as explained in 2.23) is linearly equivalent to . In particular, for any , is Cartier. Moreover, and are both -Gorenstein.
-
(3)
For each , there is an affine open neighbourhood containing and a regular sequence on generating such that , and is strongly -regular for each . In particular, and are both globally -regular.
-
(4)
For any Weil-divisor on such that is Cartier and ample for some , the section module
is flat over , and compatible with base change to fibers. In other words, for any , the natural map
is an isomorphism. Moreover, if does not contain any fiber of then the natural map
is an isomorphism as well.
Proof.
Since all parts of the lemma can be proved locally on , we may shrink if necessary to assume that is a free -module.
-
(1)
This is immediate from Part (1) of 5.6.
-
(2)
By inverting an element , and setting , we may assume that is generated by a regular sequence on (this is possible since is regular). Fix any and let . By [Mat89, Theorem 23.9], since all fibers of are normal, we in particular know that each (and hence, ) is normal. By shrinking further if necessary, using Part (1) of 5.6, we may also assume .
Now we will show that restricted to is linearly equivalent to the divisor . Let denote the smooth locus of , and be an open set such that . This is possible because is a complete intersection in . Then, applying the adjunction formula for the complete intersection , we have . Since is the smooth locus and is normal, it contains all the codimension one points of . Thus, taking closures we get that
This proves that restricted to is linearly equivalent to . In particular is Cartier. Furthermore, it follows from the discussion in [SS10, Section 5.2]) that the canonical divisor is the cone over the canonical divisor (as Weil-divisor). Thus, we get that as a graded module over . Also note that for any maximal ideal , the fiber is strongly -regular. In particular, all fibers over closed points of are Cohen-Macaulay, we see that is Cohen-Macaulay as well. Therefore, (and similarly ) is -Gorenstein.
-
(3)
From part (2), we may assume that is generated by a regular sequence on and is Cartier for each , where . Next, note that by assumption (and Part (1)), we have is strongly -regular. So, there exists a non-zero element such that is strongly -regular where . Here, we are using the fact that the non-strongly -regular locus of is closed, compatible with localization and homogeneous with respect to the -grading on . Let . Then, since is -Cartier and is a non-zero divisor on , by [Das15, Theorem A], we conclude that is strongly -regular in a neighbourhood of . Since the locus of points where is not strongly -regular is defined by a homogeneous ideal (and is disjoint from ), we may pick a non-zero element such that if we set , is strongly -regular. Replacing with , we may proceed inductively to get a localization of (at finitely many elements), and an open neighbourhood of such that and is strongly -regular for each , as required.
Applying this to maximal ideals in , we see that is strongly -regular. Moreover, let for some be an effective divisor and denote the corresponding degree element of . Then is isomorphic to the product of and . Since is affine and strongly -regular (because we have shown that is strongly -regular), to show that is globally -regular, using [SS10, Theorem 3.9], it is sufficient to show that for some , the map splits. But this again follows from the fact that is strongly -regular by using 2.7. Hence, is globally -regular.
-
(4)
Using Part (3) above, is in particular, locally strongly -regular. Now the claim is an immediate consequence of Part (2) of 5.6. ∎
Lemma 5.11.
Let be as above and assume is regular. Suppose is a -divisor on satisfying the following two properties:
-
(1)
there is some for which is an integral Weil divisor linearly equivalent to as Weil divisors, and
-
(2)
does not contain any fiber of .
Then, if is a point such that the pair is globally -regular for some , then there is an open neighbourhood of such that is also globally -regular for all .
Proof.
The proof is divided into several steps, but the strategy is to apply [PSZ18, Corollary 4.19] carefully. See Section 2.3 for a detailed discussion of the process of taking cones over divisors in family.
Step 1:
By shrinking to a neighbourhood of if necessary, we may also assume that is isomorphic to . Fix an such that is an ample Cartier divisor, and set and be the corresponding section ring and is the cone over . By 5.6, taking the cone over commutes with base change to fibers of . For any integral Weil divisor , let denote the corresponding section module over .
Step 2:
For any sufficiently divisible, let , which we may assume is an integral Weil-divisor. Since is -Cartier, so is . Therefore, by Part (4) of 5.10, the section module
is compatible with base change to fibers. Thus, we may consider the cone over as a Weil-divisor on defined by the reflexive sheaf . The compatibility with base changing to fibers guarantees that taking the cone over commutes with restricting to the fibers of .
Step 3:
Since the pair is globally -regular for , it remains so for all for some small . Now we set to be a rational number such that for positive integers and with the following properties:
-
(1)
is divisible by (the Cartier index of ).
-
(2)
is not divisible by .
Furthermore, choose such that divides and and are both integral Weil-divisors. Such an exists by our assumptions on . Lastly, set . With this notation, we note that the following properties are satisfied:
-
•
, where is a Weil-divisor and does not divide .
-
•
is linearly equivalent to
(5.8) And since divides , we get that is an integral Cartier divisor. Also note that clearly divides and is not divisible by .
-
•
The fibers of are geometrically normal since the perfect fibers are globally -regular. They are also geometrically connected by assumption. Thus, our assumption that does not contain any fibers of guarantees that does not contain any generic point of any geometric fiber of .
Step 4:
In this context, we may use [PSZ18, Corollary 4.19] applied to the projective cone of with respect to to conclude the proof (see Section 2.2 for details about the projective cone construction). More precisely, consider the map
where is just another variable adjoined to in degree . Note that is Cartier on (since this is true at the zero-section by construction, and away from the zero section we know that is an -bundle over ).
By 5.10, the construction of the projective cone with respect to is compatible with base change to fibers. In other words, for any , the fiber is the map
Let denote the -divisor obtained as the closure in of the cone over . For any such that is integral, the section module corresponding to is by construction. Thus, by Step 2,the construction of the projective cone over is compatible with restricting to fibers. Thus, by Equation 5.8 and the fact that away from the zero-section, is an -bundle over , we have that is -Cartier with index not divisible by .
Step 5:
Finally, we observe that for any , the local strong -regularity of is equivalent to the global -regularity of , since both correspond to the strong -regularity of the pair . With these observations in place, [PSZ18, Corollary 4.19] gives us an open neighbourhood of such that is globally -regular for all . Since, and , the same is true for as required. ∎
Proof of 5.8.
Recall that to prove that the given map is lower semicontinuous, we need to show that given any point and such that , there exists an open neighbourhood such that for all . The idea of the proof is similar to the proof of 5.3, but we need a slight variation since we need an open neighbourhood of instead of just any open subset of .
Firstly, by shrinking to a neighbourhood of , we assume is affine and is a free -module. Next, using 3.13, it is sufficient to prove the lower semicontinuity of the function
for any . So we pick an such that is a globally generated ample divisor on . In particular, is Cartier. Therefore, Part (2) of 5.10, we have that is a globally generated ample Cartier divisor on for any . Additionally, fix an integer such that for all .
Let be the relative dimension of , let denote , and . Recall that for any -algebra , denotes the section ring , where denotes the base change and . By the argument in Step 2 of the proof of 5.3 (replacing the generic point with ), it is sufficient to show that there exists an such that and an open neighbourhood of such that
for all . To prove this, choose an such that , and
(5.9) |
Let be an integer such that be an integer that is not divisible by . By Part 2c of 5.6, we may assume (by shrinking if necessary) that is a free -module for each with a basis . Let be any element of and be any rational number. Since is a basis element using Part 2c of 5.6 again, we see that does not contain any fibers of , since restricts to a non-zero global section on each fiber. Then, we apply 5.11, to the -divisor to get an open neighbourhood such that for each , the pair is globally -regular. This is because since , the pair is globally -regular by the definition of . Since by construction, 3.6 tells us that for each , and each in , the map
(5.10) |
splits. Furthermore, we may pick an integer satisfying: divides , , and . This is possible by our choice of . Since , we can pick a non-zero that restricts to a non-zero Weil-divisor on each fiber (by base-change). Thus, for any element in , since the corresponding map for splits for each by Equation 5.10, the corresponding map for also splits for each . Finally, we apply [LP23, Lemma 2.7 (a)] repeatedly to the basis to conclude that for for each and thus
for all as required. This completes the proof of 5.8. ∎
6. Examples
In this section, we compute some examples of the -invariant for non-toric varieties and highlight some interesting features.
6.1. Quadric hypersurfaces
Fix any algebraically closed field of characteristic and let be the -dimensional smooth quadric hypersurface over . Note that by the adjunction formula, where denotes a hyperplane section.
Example 6.1.
Then, . Equivalently, if denotes the section ring
then . This follows from a description of the structure of the sheaves proved in [Lan08] and [Ach12]. More precisely, for any and , [Ach12, Theorem 2] tells us that is a direct sum of and for , where is an ACM bundle that sits in an exact sequence of the form
for suitable positive integers and . Here is the inclusion. See [Ach12, Section 1.3] for the details. Since , we deduce from the exact sequence above that has no global sections for any . Therefore, all global sections of appear in the trivial summands. In other words, for any and any . Moreover, since , we know that . Therefore, by 3.8, we have
Remark 6.2.
This example shows that the -invariant does not characterize regularity of section rings, since the -invariant of a polynomial ring is also equal to .
Remark 6.3.
Another interesting feature of this example is that the -invariant of smooth quadrics is independent of the characteristic (for ). This is far from true in general (as seen in the next example). Furthermore, for any , the -signature of is known to depend on in a rather complicated way (see [Tri23]).
6.2. Comparison to the complex -invariant
Let for some prime number and be the diagonal cubic surface defined by over .
Example 6.4.
For each , we have . However,
To see this, we recall the following result proved by Shideler (see [Shi, Example 4.2.2 and Section 5.1]), building on the techniques of Han and Monsky: Let denote the -signature of , equivalently, of the ring . Then for any , we have . Moreover,
Using this, our claims about the -invariant of follow from 4.7 and 4.8, once we observe that
Remark 6.5.
The complex -invariant of the cubic surface defined by is equal to (see [Che08, Theorem 1.7]). Note that by [HY03], we know that for a fixed divisor on a variety
where and denote the reduction to characteristic of and respectively. 6.4 points to limitations of approximating the log canonical threshold by -pure threshold for an unbounded family of divisors on .
References
- [Ach12] Piotr Achinger. Frobenius push-forwards on quadrics. Comm. Algebra, 40(8):2732–2748, 2012.
- [AL03] Ian M. Aberbach and Graham J. Leuschke. The -signature and strong -regularity. Math. Res. Lett., 10(1):51–56, 2003.
- [Bir21] Caucher Birkar. Singularities of linear systems and boundedness of Fano varieties. Ann. of Math. (2), 193(2):347–405, 2021.
- [BJ20] Harold Blum and Mattias Jonsson. Thresholds, valuations, and K-stability. Adv. Math., 365:107062, 57, 2020.
- [BL22] Harold Blum and Yuchen Liu. Openness of uniform K-stability in families of -Fano varieties. Ann. Sci. Éc. Norm. Supér. (4), 55(1):1–41, 2022.
- [Bli04] Manuel Blickle. Multiplier ideals and modules on toric varieties. Math. Z., 248(1):113–121, 2004.
- [BST11] Manuel Blickle, Karl Schwede, and Kevin Tucker. -signature of pairs and the asymptotic behavior of Frobenius splittings. 2011. arXiv:1107.1082.
- [Che08] Ivan Cheltsov. Log canonical thresholds of del Pezzo surfaces. Geom. Funct. Anal., 18(4):1118–1144, 2008.
- [CR17] Javier Carvajal-Rojas. Finite torsors over strongly F-regular singularities. 2017. arXiv:1710.06887.
- [CRST21] Javier Carvajal-Rojas, Karl Schwede, and Kevin Tucker. Bertini theorems for -signature and Hilbert-Kunz multiplicity. Math. Z., 299(1-2):1131–1153, 2021.
- [CS08] I. A. Cheltsov and K. A. Shramov. Log-canonical thresholds for nonsingular Fano threefolds. Uspekhi Mat. Nauk, 63(5(383)):73–180, 2008. With an appendix by J.-P. Demailly.
- [Das15] Omprokash Das. On strongly -regular inversion of adjunction. J. Algebra, 434:207–226, 2015.
- [Eis95] David Eisenbud. Commutative algebra, volume 150 of Graduate Texts in Mathematics. Springer-Verlag, New York, 1995. With a view toward algebraic geometry.
- [Har77] Robin Hartshorne. Algebraic geometry. Springer-Verlag, New York, 1977. Graduate Texts in Mathematics, No. 52.
- [Har98] Nobuo Hara. A characterization of rational singularities in terms of injectivity of Frobenius maps. Amer. J. Math., 120(5):981–996, 1998.
- [Her12] Daniel J. Hernández. -purity of hypersurfaces. Math. Res. Lett., 19(2):389–401, 2012.
- [HS04] Eero Hyry and Karen E. Smith. Core versus graded core, and global sections of line bundles. Trans. Amer. Math. Soc., 356(8):3143–3166 (electronic), 2004.
- [HW02] Nobuo Hara and Kei-Ichi Watanabe. F-regular and F-pure rings vs. log terminal and log canonical singularities. J. Algebraic Geom., 11(2):363–392, 2002.
- [HX15] Christopher D. Hacon and Chenyang Xu. On the three dimensional minimal model program in positive characteristic. J. Amer. Math. Soc., 28(3):711–744, 2015.
- [HY03] Nobuo Hara and Ken-Ichi Yoshida. A generalization of tight closure and multiplier ideals. Trans. Amer. Math. Soc., 355(8):3143–3174 (electronic), 2003.
- [Kol13] János Kollár. Singularities of the minimal model program, volume 200 of Cambridge Tracts in Mathematics. Cambridge University Press, Cambridge, 2013. With the collaboration of Sándor Kovács.
- [Lan08] Adrian Langer. -affinity and Frobenius morphism on quadrics. Int. Math. Res. Not. IMRN, (1):Art. ID rnm 145, 26, 2008.
- [Laz04] Robert Lazarsfeld. Positivity in algebraic geometry. I, volume 48 of Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge. A Series of Modern Surveys in Mathematics [Results in Mathematics and Related Areas. 3rd Series. A Series of Modern Surveys in Mathematics]. Springer-Verlag, Berlin, 2004. Classical setting: line bundles and linear series.
- [LP23] Seungsu Lee and Swaraj Pande. The F-Signature Function on the Ample Cone. International Mathematics Research Notices, page rnad174, 07 2023.
- [LZ22] Yuchen Liu and Ziquan Zhuang. On the sharpness of Tian’s criterion for K-stability. Nagoya Math. J., 245:41–73, 2022.
- [Mat89] Hideyuki Matsumura. Commutative ring theory, volume 8 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, second edition, 1989. Translated from the Japanese by M. Reid.
- [MS97] V. B. Mehta and V. Srinivas. A characterization of rational singularities. Asian J. Math., 1(2):249–271, 1997.
- [MS18] Linquan Ma and Karl Schwede. Singularities in mixed characteristic via perfectoid big Cohen–Macaulay algebras. 2018. arXiv:1806.09567, to appear in Duke Math. J.
- [OS12] Yuji Odaka and Yuji Sano. Alpha invariant and K-stability of -Fano varieties. Adv. Math., 229(5):2818–2834, 2012.
- [Pat14] Zsolt Patakfalvi. Semi-positivity in positive characteristics. Ann. Sci. Éc. Norm. Supér. (4), 47(5):991–1025, 2014.
- [Pol18] Thomas Polstra. Uniform bounds in F-finite rings and lower semi-continuity of the F-signature. Trans. Amer. Math. Soc., 370(5):3147–3169, 2018.
- [PS12] Zsolt Patakfalvi and Karl Schwede. Depth of -singularities and base change of relative canonical sheaves. 2012. arXiv:1207.1910.
- [PSZ18] Zsolt Patakfalvi, Karl Schwede, and Wenliang Zhang. -singularities in families. Algebr. Geom., 5(3):264–327, 2018.
- [PT18] Thomas Polstra and Kevin Tucker. -signature and Hilbert-Kunz multiplicity: a combined approach and comparison. Algebra Number Theory, 12(1):61–97, 2018.
- [Sat18] Kenta Sato. Stability of test ideals of divisors with small multiplicity. Math. Z., 288(3-4):783–802, 2018.
- [Sch08] Karl Schwede. Generalized test ideals, sharp -purity, and sharp test elements. Math. Res. Lett., 15(6):1251–1261, 2008.
- [Shi] Samuel Joseph Shideler. The f-signature of diagonal hypersurfaces. Available at http://arks.princeton.edu/ark:/88435/dsp01ns0646128.
- [Smi97] Karen E. Smith. -rational rings have rational singularities. Amer. J. Math., 119(1):159–180, 1997.
- [Smi00] Karen E. Smith. Globally F-regular varieties: applications to vanishing theorems for quotients of Fano varieties. Michigan Math. J., 48:553–572, 2000. Dedicated to William Fulton on the occasion of his 60th birthday.
- [Smi16] Ilya Smirnov. Upper semi-continuity of the Hilbert-Kunz multiplicity. Compos. Math., 152(3):477–488, 2016.
- [Smi20] Ilya Smirnov. On semicontinuity of multiplicities in families. Doc. Math., 25:381–400, 2020.
- [SS10] Karl Schwede and Karen E. Smith. Globally -regular and log Fano varieties. Adv. Math., 224(3):863–894, 2010.
- [Sta] The Stacks Project Authors. Stacks Project. http://stacks.math.columbia.edu.
- [Tak04] Shunsuke Takagi. An interpretation of multiplier ideals via tight closure. J. Algebraic Geom., 13(2):393–415, 2004.
- [Tia87] Gang Tian. On Kähler-Einstein metrics on certain Kähler manifolds with . Invent. Math., 89(2):225–246, 1987.
- [Tri23] Vijaylaxmi Trivedi. The hilbert-kunz density functions of quadric hypersurfaces, 2023.
- [Tuc12] Kevin Tucker. -signature exists. Invent. Math., 190(3):743–765, 2012.
- [VK12] Michael R. Von Korff. The F-Signature of Toric Varieties. ProQuest LLC, Ann Arbor, MI, 2012. Thesis (Ph.D.)–University of Michigan.
- [Xu21] Chenyang Xu. K-stability of Fano varieties: an algebro-geometric approach. EMS Surv. Math. Sci., 8(1-2):265–354, 2021.
- [Yao06] Yongwei Yao. Observations on the -signature of local rings of characteristic . J. Algebra, 299(1):198–218, 2006.