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

A Frobenius version of Tian’s Alpha-Invariant

Swaraj Pande Department of Mathematics
University of Michigan
Ann Arbor, MI 48109-1043
USA
[email protected]
Abstract.

For a pair (X,L)(X,L) consisting of a projective variety XX over a perfect field of characteristic p>0p>0 and an ample line bundle LL on XX, we introduce and study a positive characteristic analog of Tian’s α\alpha-invariant, which we call the αF\alpha_{F}-invariant. We utilize the theory of FF-singularities in positive characteristics, and our approach is based on replacing klt singularities with the closely related notion of global FF-regularity. We show that the αF\alpha_{F}-invariant of a pair (X,L)(X,L) can be understood in terms of the global Frobenius splittings of the linear systems |mL|m0|mL|_{m\geq 0}. We establish inequalities relating the αF\alpha_{F}-invariant with the FF-signature, and use that to prove the positivity of the αF\alpha_{F}-invariant for all globally FF-regular projective varieties (with respect to any ample LL on XX). When XX is a Fano variety and L=KXL=-K_{X}, we prove that the αF\alpha_{F}-invariant of XX is always bounded above by 1/21/2 and establish tighter comparisons with the FF-signature. We also show that for toric Fano varieties, the αF\alpha_{F}-invariant matches with the usual (complex) α\alpha-invariant. Finally, we prove that the αF\alpha_{F}-invariant is lower semicontinuous in a family of globally FF-regular Fano varieties.

This work was partially supported by the NSF grants #1952399, #1801697 and #2101075

1. Introduction

The α\alpha-invariant of a complex Fano variety XX was introduced by Tian in [Tia87] to provide a sufficient criterion for KK-stability of XX, a condition that guarantees the existence of a Kähler-Einstein metric on XX. Though initially defined analytically, Demailly later reinterpreted the α\alpha-invariant in terms of the log canonical threshold [CS08], an algebraic invariant of the singularities of divisors on XX. Since then, understanding the α\alpha-invariant and KK-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 α\alpha-invariant. To do this, we replace the singularities of the MMP with singularities defined using the Frobenius map (“FF-singularities"). Though FF-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 FF-split (resp. strongly FF-regular) singularities (2.15). Since the α\alpha-invariant of a complex Fano variety XX involves the log canonicity of anti-canonical \mathbb{Q}-divisors of XX, this inspires our definition of the Frobenius version:

Definition 1.1.

Let XX be a globally FF-regular Fano variety over a perfect field of positive characteristic. Then, we define the αF\alpha_{F}-invariant of XX as

αF(X):=sup{t0|(X,tΔ)is globally F-split effective -divisor ΔKX}.\alpha_{F}(X):=\sup\{t\geq 0\,|\,(X,t\Delta)\,\text{is globally $F$-split }\forall\,\text{effective $\mathbb{Q}$-divisor }\Delta\sim_{\mathbb{Q}}-K_{X}\}.

Since we intend for the αF\alpha_{F}-invariant to capture global properties of anti-canonical \mathbb{Q}-divisors on XX, we use the notion of global FF-splitting (2.15), or equivalently, FF-splitting of the cone over the corresponding divisors on XX (with respect to KX-K_{X}). To do so, we require XX to be globally FF-regular (2.16), which is equivalent to the cone over XX being strongly FF-regular. Thus, global FF-regularity can be thought of as a Frobenius analog of the klt-condition on complex Fano varieties. Furthermore, we note that while working over \mathbb{C}, simply replacing globally FF-regular and globally FF-split in 1.1 by klt on the cone, and log canonical on the cone respectively, we obtain the minimum value between the usual α\alpha-invariant of XX and 11 (see 4.11). Thus, at least for Fano varieties with α(X)1\alpha(X)\leq 1, the αF\alpha_{F}-invariant is a “Frobenius-analog" of Tian’s α\alpha-invariant.

Our first theorem proves some surprising properties of the αF\alpha_{F}-invariant in contrast to the complex version, and establishes connections to the FF-signature of XX, another important invariant and a Frobenius version of the local volume of singularities:

Theorem 1.2.

Let XX be a globally FF-regular Fano variety over a perfect field of positive characteristic of positive dimension. Then,

  1. (1)

    The αF\alpha_{F}-invariant of XX is at most 1/2 (4.5).

  2. (2)

    Assume that XX is geometrically connected over the (perfect) base field. Then, we have αF(X)=1/2\alpha_{F}(X)=1/2 if and only if the FF-signature of XX (with respect to KX-K_{X}) equals vol(KX)2d(d+1)!\frac{\mathrm{vol}(-K_{X})}{2^{d}(d+1)!}, where dd is the dimension of XX (4.8).

  3. (3)

    More generally (and still assuming XX is geometrically connected), the FF-signature of XX is at most vol(KX)2d(d+1)!\frac{\mathrm{vol}(-K_{X})}{2^{d}(d+1)!} (4.8).

  4. (4)

    In case XX is a toric Fano variety corresponding to a fan \mathcal{F}, then αF(X)\alpha_{F}(X) is the same as the complex α\alpha-invariant of X()X_{\mathbb{C}}(\mathcal{F}), the complex toric Fano variety corresponding to \mathcal{F} (4.12).

Part (1) of 1.2 is surprising since many complex Fano varieties have α\alpha-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 α\alpha-invariant of toric Fano varieties is at most 1/2 (see [LZ22, Corollary 3.6]).

The αF\alpha_{F}-invariant (like the complex α\alpha-invariant) can be defined for any pair (X,L)(X,L), where XX is a globally FF-regular projective variety and LL is an ample line bundle on XX. So, in Section 3, we develop the theory of the αF\alpha_{F}-invariant in this more general setting. From this perspective, the αF\alpha_{F}-invariant is an asymptotic invariant of a section ring of a projective variety that shares many properties and relations with the FF-signature. In this direction, we prove:

Theorem 1.3.

Let SS denote the section ring of a globally FF-regular projective variety over a perfect field kk, with respect to some ample line bundle over XX. Then,

  1. (1)

    αF(S)\alpha_{F}(S) can be calculated as the following limit:

    αF(S)=limeme(S)pe\alpha_{F}(S)=\lim_{e\to\infty}\frac{m_{e}(S)}{p^{e}}

    where me(S)m_{e}(S) denotes that maximum integer mm such that for each non-zero homogeneous element fSf\in S of degree mm, the map SFeSS\to F^{e}_{*}S sending 11 to FefF^{e}_{*}f splits. See 3.8.

  2. (2)

    αF(S)\alpha_{F}(S) is positive (3.10).

  3. (3)

    Base-change (3.16): Assume that S0=kS_{0}=k and KK is any perfect field extension of kk. Then,

    αF(S)=αF(SkK).\alpha_{F}(S)=\alpha_{F}(S\otimes_{k}K).

Our third set of results concern the semicontinuity properties of the αF\alpha_{F}-invariant in families of globally FF-regular varieties (5.2), analogous to the results of [BL22] about the complex version. In 5.8, we prove:

Theorem 1.4.

Let f:𝒳Yf:\mathcal{X}\to Y be family of globally FF-regular Fano varieties such that K𝒳|Y-K_{\mathcal{X}|Y} is \mathbb{Q}-Cartier and ff-ample. Assume that YY is regular. Then, the map Y0Y\to\mathbb{R}_{\geq 0} given by

yαF(𝒳y)y\mapsto\alpha_{F}(\mathcal{X}_{y^{\infty}})

is lower semicontinuous, where 𝒳y\mathcal{X}_{y^{\infty}} is the perfectified-fiber over yYy\in Y.

We also prove a weaker version of 1.4 for any polarized family of globally FF-regular varieties (see 5.3). This is analogous to the corresponding result for the FF-signature proved in [CRST21] and relies on understanding the rate of convergence in Part (1) of 1.3 for the αF\alpha_{F}-invariant, which may be of independent interest (see 5.4).

Finally, we consider some examples that highlight interesting features of the αF\alpha_{F}-invariant. For instance, we see that the αF\alpha_{F}-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 pp, the αF\alpha_{F}-invariant may depend on the characteristic. Furthermore, while the limit of the αF\alpha_{F}-invariant as pp\to\infty may exist and have an interesting geometric interpretation, the limit is not necessarily the complex α\alpha-invariant. This fact raises interesting questions and obstructions to relating log canonical thresholds and FF-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, kk will denote a perfect field of characteristic pp. A variety over kk is an integral (in particular, connected), separated scheme of finite type over kk. For a point xx on a scheme XX, the residue field 𝒪X,x/𝔪x\mathcal{O}_{X,x}/\mathfrak{m}_{x} will be denoted by κ(x)\kappa(x) (where 𝒪X,x\mathcal{O}_{X,x} is the local ring at xx and 𝔪x\mathfrak{m}_{x} is the maximal ideal of the local ring).

Notation 2.2 (Divisors and Pairs).

A prime Weil-divisor on a scheme XX is a reduced and irreducible subscheme of XX of codimension one. An integral Weil-divisor is a formal \mathbb{Z}-linear combination of prime Weil-divisors. A \mathbb{Q}-divisor is a formal \mathbb{Q}-linear combination of prime Weil-divisors. By a pair (X,Δ)(X,\Delta), we mean that XX is a Noetherian, normal scheme and Δ\Delta is an effective \mathbb{Q}-divisor over XX. A projective pair is a pair (X,Δ)(X,\Delta) where XX is a projective variety over kk.

2.1. Section Rings and Modules

Definition 2.3.

Let AA be a Noetherian ring and XX be a projective scheme over AA. Given an ample invertible sheaf \mathcal{L} on XX and \mathcal{F} a coherent sheaf on XX, the \mathbb{N}-graded ring SS defined by

S=S(X,):=n0H0(X,n)S=S(X,\mathcal{L}):=\bigoplus_{n\geq 0}H^{0}(X,\mathcal{L}^{n})

is called the section ring of XX with respect to \mathcal{L}. The affine scheme Spec(S)\operatorname{\text{Spec}}(S) is called the (affine) cone over XX with respect to \mathcal{L}. The section module of \mathcal{F} with respect to \mathcal{L} is a \mathbb{Z}-graded SS-module MM defined by

M=M(X,):=nH0(X,n).M=M(X,\mathcal{L}):=\bigoplus_{n\in\mathbb{Z}}H^{0}(X,\mathcal{F}\otimes\mathcal{L}^{n}).

Similarly, the sheaf corresponding to MM on Spec(S)\operatorname{\text{Spec}}(S) is called the cone over \mathcal{F} with respect to \mathcal{L}.

Let SS be a Noetherian, \mathbb{N}-graded domain, and TT denote the set of positive degree homogeneous elements of SS. For a finitely generated, torsion-free, \mathbb{Z}-graded module MM over SS, let MM^{\prime} denote the localization M=T1MM^{\prime}=T^{-1}M. Note that MM^{\prime} is naturally a \mathbb{Z}-graded module over T1ST^{-1}S. Since MM is torsion-free, we can think of MM naturally as a subset of MM^{\prime}. In this setting, we define the saturation of MM to be the \mathbb{Z}-graded module

Msat={mM|𝔭nmMfor some n>0}M^{\text{sat}}=\{m\in M^{\prime}\,|\,\mathfrak{p}^{n}m\in M\quad\text{for some $n>0$}\}

where 𝔭\mathfrak{p} is the irrelevant ideal 𝔭=j>0Sj\mathfrak{p}=\bigoplus_{j>0}S_{j}. We say MM is saturated if M=MsatM=M^{\text{sat}}.

Lemma 2.4.

Let AA be a Noetherian domain, XX be an integral, projective scheme over AA and \mathcal{L} an ample invertible sheaf over XX.

  1. (1)

    The section ring SS of XX with respect to \mathcal{L} is a finitely generated algebra over AA and hence, is Noetherian. If XX is normal, and A=kA=k, then the section ring is also characterized as the unique normal \mathbb{N}-graded ring SS such that Proj(S)\operatorname{\text{Proj}}(S) is isomorphic to XX and the corresponding 𝒪X(1)\mathcal{O}_{X}(1) is isomorphic to \mathcal{L}.

  2. (2)

    The section module of any torsion-free coherent sheaf over XX with respect to \mathcal{L} is finitely generated over SS. It is also characterized as the unique saturated, torsion-free, \mathbb{Z}-graded SS-module MM (with respect to the irrelevant ideal I=j>0SjI=\bigoplus_{j>0}S_{j}) such that the associated coherent sheaf M~\tilde{M} on XX is isomorphic to \mathcal{F}.

  3. (3)

    For two torsion-free coherent sheaves \mathcal{F} and 𝒢\mathcal{G}, we have a natural isomorphism:

    Hom𝒪X(,𝒢)HomSgr(M(,),M(𝒢,))\operatorname{Hom}_{\mathcal{O}_{X}}(\mathcal{F},\mathcal{G})\cong\operatorname{Hom}^{\text{gr}}_{S}(M(\mathcal{F},\mathcal{L}),M(\mathcal{G},\mathcal{L}))

    where HomSgr(,)\operatorname{Hom}^{\text{gr}}_{S}(\,,\,) denotes the set of grading preserving SS-module maps between two graded SS-modules.

Proof.

See [Sta, Tag 0BXF]. ∎

2.2. Affine and Projective Cones

Given an integral, projective scheme XX over a noetherian domain AA, and an ample invertible sheaf \mathcal{L} over XX, let SS be the corresponding section ring. Assume that S0=AS_{0}=A. Let X¯\overline{X} be the projective AA-scheme defined as

X¯=Proj(S[z])\overline{X}=\operatorname{\text{Proj}}(S[z])

where S[z]S[z] is the \mathbb{N}-graded ring obtained by adjoining a new variable zz to SS in degree 11. Then, X¯\overline{X} is called the projective cone over XX with respect to \mathcal{L}. Denoting by 𝔪\mathfrak{m} the homogeneous irrelevant ideal j>0Sj\bigoplus_{j>0}S_{j} of SS, we have a map of graded AA-algebras

S[z](S/𝔪)[z]A[z]S[z]\to(S/\mathfrak{m})[z]\cong A[z]

that induces the “zero-section" map

σ:Spec(A)X¯\sigma:\operatorname{\text{Spec}}(A)\to\overline{X}

over AA. We call this map the “vertex of the cone X¯\overline{X}". We also have natural maps of graded rings

SS[z]S[z]/(z)SS\subset S[z]\to S[z]/(z)\cong S

which, via the construction in [Har77, II, Exercise 2.14 (b)] induce maps

i:XX¯i_{\infty}:X\hookrightarrow\overline{X}

called the “section at infinity" and

π:X¯σ(Spec(A))X\pi:\overline{X}\setminus\sigma(\operatorname{\text{Spec}}(A))\to X

where σ\sigma is the zero-section described above. It follows from the Proj construction that π\pi is an 𝔸1\mathbb{A}^{1}-bundle over XX. The affine cone Y=Spec(S)Y=\operatorname{\text{Spec}}(S) is isomorphic to X¯i(X)\overline{X}\setminus i_{\infty}(X), and the zero section σ\sigma actually maps into YY. Thus, π\pi restricts to a map Yσ(Spec(A))XY\setminus\sigma(\operatorname{\text{Spec}}(A))\to X that is a Spec(A[t,t1])\operatorname{\text{Spec}}(A[t,t^{-1}])-bundle over XX. See [HS04, Section 2] and [Kol13, Section 3.1] for details.

2.3. Cones over \mathbb{Q}-divisors

We follow the description of the cone over a \mathbb{Q}-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 AA be a normal domain of finite type over kk, and XX be an integral, normal, projective scheme over AA. Assume that XX is flat over AA and of positive relative dimension. Fix an ample invertible sheaf \mathcal{L} over XX and SS be the corresponding section ring. Assume further that S0=AS_{0}=A. Note that this guarantees that the codimension of the zero section σ\sigma is at least two in Y=Spec(S)Y=\operatorname{\text{Spec}}(S). Thus, it follows that SS is normal as well. In this situation, given any integral Weil-divisor D=aiDiD=\sum a_{i}D_{i} (for distinct prime Weil divisors DiD_{i}) on XX, we can construct the corresponding Weil divisor D~\tilde{D} on YY, the “cone over DD", in three equivalent ways:

  1. (1)

    Let Di~\tilde{D_{i}} be the prime Weil divisor on YY corresponding to the height one prime 𝔭iS\mathfrak{p}_{i}\subset S corresponding to DiD_{i}. Then D~=aiDi~\tilde{D}=\sum a_{i}\tilde{D_{i}}.

  2. (2)

    Let 𝒪X(D)\mathcal{O}_{X}(D) be the reflexive sheaf on XX corresponding to DD. Then, D~\tilde{D} is the divisor corresponding to the reflexive SS-module defined by

    M:=M(D,)=jH0(X,𝒪X(D)j).M:=M(D,\mathcal{L})=\bigoplus_{j\in\mathbb{Z}}H^{0}(X,\mathcal{O}_{X}(D)\otimes\mathcal{L}^{j}).

    Note that the fact that MM is reflexive can be seen by applying Part (3) of 2.4 to each twist of the graded module MM.

  3. (3)

    Let π:Yσ(Spec(A))X\pi:Y\setminus\sigma(\operatorname{\text{Spec}}(A))\to X be the A[t,t1]A[t,t^{-1}]-bundle map defined in the previous paragraph. Then, we may define D~\tilde{D} to be the pull back of DD to YY. More precisely, near the generic point of a component of DD, if DD is given by an equation ff, then is D~\tilde{D} is defined by πf\pi^{*}f. We then take closures to obtain a Weil-divisor on YY. This defines a unique divisor on YY since π\pi is flat and codimY(σ(Spec(A)))\operatorname{codim}_{Y}(\sigma(\operatorname{\text{Spec}}(A))) is at least 22.

Given a principal divisor DD on XX defined by the rational function ff, the cone over DD can be seen to be the principal divisor defined by ff again. The construction of the cone clearly preserves addition of divisors. This implies that cone construction extends to \mathbb{Q}-divisors and preserves the linear equivalence of Weil-divisors. Furthermore, by taking closures, this construction also extends to the projective cone X¯\overline{X} described in the previous paragraph. Finally, using the third description of the cone, we see that the cone over the canonical (Weil-)divisor KXK_{X} is the canonical divisor of KYK_{Y} (recall that YY is also normal).

2.4. FF-signature

Let RR be any ring of prime characteristic pp. Then for any e1e\geq 1, let Fe:RRF^{e}:R\to R sending rrper\mapsto r^{p^{e}} be the ethe^{\text{th}}-iterate of the Frobenius morphism. Since RR has characteristic pp, FeF^{e} defines a ring homomorphism, allowing us to define a new RR-module for each e1e\geq 1 obtained via restriction of scalars along FeF^{e}. We denote this new RR-module by FeRF_{*}^{e}R and its elements by FerF_{*}^{e}r (where rr is an element of RR). Concretely, FeRF_{*}^{e}R is the same as RR as an abelian group, but the RR-module action is given by:

rFes:=Fe(rpes) for rR and FesFeR.r\cdot F_{*}^{e}s:=F_{*}^{e}(r^{p^{e}}s)\textrm{\quad for $r\in R$ and $F_{*}^{e}s\in F_{*}^{e}R$}.

Now let (R,𝔪)(R,\mathfrak{m}) denote a normal local ring and XX denote the normal scheme Spec(R)\operatorname{\text{Spec}}(R). Throughout, we will assume that RR is the localization of a finitely generated kk-algebra at a maximal ideal, which also makes it FF-finite (i.e., FeRF_{*}^{e}R is a finitely generated RR-module for any e1e\geq 1), with the rank of FeRF^{e}_{*}R over RR being pedp^{ed}, where dd is the Krull dimension of RR. Let Δ\Delta be an effective \mathbb{Q}-divisor on X=Spec(R)X=\operatorname{\text{Spec}}(R). Then, note that since Δ\Delta is effective, for any e1e\geq 1, we have a natural inclusion RR((pe1)Δ)R\subset R(\lceil(p^{e}-1)\Delta\rceil) of reflexive RR-modules. Here, R((pe1)Δ)R(\lceil(p^{e}-1)\Delta\rceil) denotes the RR-module corresponding to the reflexive sheaf 𝒪X((pe1)Δ)\mathcal{O}_{X}(\lceil(p^{e}-1)\Delta\rceil). Thus, applying HomR(_,R)\operatorname{Hom}_{R}(\,\textunderscore\,,\,R) to the natural inclusion FeRFe(R((pe1)Δ))F^{e}_{*}R\subset F^{e}_{*}(R(\lceil(p^{e}-1)\Delta\rceil)), we get

HomR(FeR((pe1)Δ),R)HomR(FeR,R).\operatorname{Hom}_{R}\big{(}F^{e}_{*}R(\lceil(p^{e}-1)\Delta\rceil),R\big{)}\subset\operatorname{Hom}_{R}(F^{e}_{*}R,R).

Thus, given any element φHomR(FeR((pe1)Δ),R)\varphi\in\operatorname{Hom}_{R}\big{(}F^{e}_{*}R(\lceil(p^{e}-1)\Delta\rceil),R\big{)}, it can be naturally viewed as a map φ:FeRR\varphi:F^{e}_{*}R\to R.

Definition 2.5 (Splitting Ideals).

For any e1e\geq 1, we define the subset IeΔRI_{e}^{\Delta}\subseteq R as

IeΔ={xRφ(Fex)𝔪for every map φHomR(FeR((pe1)Δ),R)}.I_{e}^{\Delta}=\left\{x\in R\mid\varphi(F^{e}_{*}x)\in\mathfrak{m}\,\text{for every map }\varphi\in\operatorname{Hom}_{R}\big{(}F^{e}_{*}R(\lceil(p^{e}-1)\Delta\rceil),R\big{)}\,\right\}.

We observe that IeΔI_{e}^{\Delta} is an ideal of finite colength in RR and we call

aeΔ=R(R/IeΔ)a_{e}^{\Delta}=\ell_{R}(R/I_{e}^{\Delta})

the Δ\Delta-free rank of FeRF^{e}_{*}R, where R\ell_{R} denotes the length as an RR-module.

Definition 2.6.

[BST11, Theorem 3.11, Proposition 3.5] Let (R,Δ)(R,\Delta) be a pair as above, and aeΔ(R)a_{e}^{\Delta}(R) denote the Δ\Delta-free rank of FeRF^{e}_{*}R (2.5). Then the FF-signature of (R,Δ)(R,\Delta) is defined to be the limit:

𝓈(R,Δ):=limeaeΔped\mathscr{s}(R,\Delta):=\lim_{e\to\infty}\frac{a_{e}^{\Delta}}{p^{ed}}

where dd is the Krull dimension of RR. This limit exists by [BST11].

2.5. FF-signature of cones over projective varieties.

In this subsection, we describe how we can compute the FF-signature of cones over projective varieties (and pairs) using global splittings on XX. 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 AA be a regular ring of finite type over kk and XX be an integral, normal projective scheme over AA (with H0(X,𝒪X)=AH^{0}(X,\mathcal{O}_{X})=A). Assume that XX is flat over AA and of positive relative dimension. Fix an ample invertible sheaf \mathcal{L} and SS be the corresponding section ring. Fix an effective Weil divisor DD over XX and D~\tilde{D} be the cone over DD with respect to \mathcal{L}. Then, for any e1e\geq 1, the natural map

𝒪XFe(𝒪X(D))\mathcal{O}_{X}\to F^{e}_{*}(\mathcal{O}_{X}(D))

splits as a map of 𝒪X\mathcal{O}_{X}-modules if and only if the map on the cones

SFe(S(D~))S\to F^{e}_{*}(S(\tilde{D}))

splits as a map of SS-modules.

Proof.

Using the second description of the cone over a \mathbb{Q}-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 kk, we next recall a formula to compute the FF-signature of the section ring of a projective variety proved in [LP23]. Fix a normal projective variety XX over kk and Δ\Delta be an effective \mathbb{Q}-divisor over XX.

Definition 2.8.

For any Weil-divisor DD on XX and e1e\geq 1, define the kk-vector subspace IeΔ(D)I_{e}^{\Delta}(D) of H0(X,𝒪X(D))H^{0}(X,\mathcal{O}_{X}(D)) as follows:

IeΔ(D):={fH0(X,D)|φ(Fef)=0 for all φHom𝒪X(Fe𝒪X((pe1)Δ+D),𝒪X)}.I_{e}^{\Delta}(D):=\{f\in H^{0}(X,D)\ |\ \varphi(F^{e}_{*}f)=0\text{ for all }\varphi\in\operatorname{Hom}_{\mathcal{O}_{X}}(F^{e}_{*}\mathcal{O}_{X}(\lceil(p^{e}-1)\Delta\rceil+D),\mathcal{O}_{X})\,\}.
Remark 2.9.

Note that the subspace IeΔ(D)I_{e}^{\Delta}(D) only depends on the sheaf 𝒪X(D)\mathcal{O}_{X}(D) and not on the specific divisor DD in its linear equivalence class.

Remark 2.10.

Let LL be an ample divisor on XX. Then, it follows from 2.7 that IeΔ(mL)I_{e}^{\Delta}(mL) is the degree mm component of the Δ\Delta-splitting ideal of the section ring of SS with respect to LL (2.5).

Lemma 2.11.

[LP23, Lemma 4.7] Let LL be an ample Cartier divisor on XX and SS denote the section ring of XX with respect to LL. Let ΔS\Delta_{S} denote the cone over Δ\Delta with respect to LL (Section 2.3). Then, for any e1e\geq 1, if aeΔ(L)a_{e}^{\Delta}(L) denotes the ΔS\Delta_{S}-free-rank of FeSF^{e}_{*}S (2.5), then aeΔ(L)a_{e}^{\Delta}(L) is computed by the following formula:

aeΔ(L)=1[k:k]m=0dimkH0(X,mL)IeΔ(mL)a_{e}^{\Delta}(L)=\frac{1}{[k^{\prime}:k]}\,\sum_{m=0}^{\infty}\dim_{k}\frac{H^{0}(X,mL)}{I_{e}^{\Delta}(mL)} (2.1)

where kk^{\prime} denotes the field H0(X,𝒪X)H^{0}(X,\mathcal{O}_{X}). Hence, the FF-signature of (X,Δ)(X,\Delta) with respect to LL can be computed as

𝓈(X,Δ)(L):=𝓈(S(X,L),ΔS)=1[k:k]limem=0dimkH0(X,mL)IeΔ(mL)pe(dim(X)+1)\mathscr{s}_{(X,\Delta)}(L):=\mathscr{s}(S(X,L),\Delta_{S})=\frac{1}{[k^{\prime}:k]}\,\lim_{e\to\infty}\frac{\sum\limits_{m=0}^{\infty}\dim_{k}\frac{H^{0}(X,mL)}{I_{e}^{\Delta}(mL)}}{p^{e(\dim(X)+1)}}
Proof.

Using 2.7, we have that

IeΔS=m0IeΔ(mL).I_{e}^{\Delta_{S}}=\bigoplus_{m\geq 0}I_{e}^{\Delta}(mL).

See 2.5 for the definition of IeΔSI_{e}^{\Delta_{S}}. Therefore, we have

S(S/IeΔS)=m0dimkH0(mL)IeΔ(mL)=1[k:k]m=0dimkH0(X,mL)IeΔ(mL)\ell_{S}(S/I_{e}^{\Delta_{S}})=\sum_{m\geq 0}\dim_{k^{\prime}}\frac{H^{0}(mL)}{I_{e}^{\Delta}(mL)}=\frac{1}{[k^{\prime}:k]}\,\sum_{m=0}^{\infty}\dim_{k}\frac{H^{0}(X,mL)}{I_{e}^{\Delta}(mL)}

where S\ell_{S} denotes the length as an SS-module. This completes the proof of the lemma. ∎

2.6. Duality and the Trace Map

It will be convenient to think of the subspaces IeΔI_{e}^{\Delta} (2.8) using a pairing arising out of duality for the Frobenius map. Let (X,Δ)(X,\Delta) be a projective pair and DD be any Weil divisor on XX. We continue to work over any perfect field kk of characteristic p>0p>0. But in this subsection, we assume that H0(X,𝒪X)=kH^{0}(X,\mathcal{O}_{X})=k, i.e., that XX is geometrically connected.

Recall that by applying duality to the Frobenius map, we get the following isomorphism of reflexive 𝒪X\mathcal{O}_{X}-modules:

om𝒪X(Fe𝒪X((pe1)Δ+D),𝒪X)Fe𝒪X((pe1)KX(pe1)ΔD).\mathscr{H}om_{\mathcal{O}_{X}}(F^{e}_{*}\mathcal{O}_{X}(\lceil(p^{e}-1)\Delta\rceil+D),\mathcal{O}_{X})\cong F^{e}_{*}\mathcal{O}_{X}(-(p^{e}-1)K_{X}-\lceil(p^{e}-1)\Delta\rceil-D). (2.2)

See [SS10, Section 4.1] for a detailed discussion regarding duality for the Frobenius map. Furthermore, when D=0D=0, this gives an isomorphism

om𝒪X(Fe𝒪X((pe1)Δ,𝒪X)Fe𝒪X((pe1)KX(pe1)Δ).\mathscr{H}om_{\mathcal{O}_{X}}(F^{e}_{*}\mathcal{O}_{X}(\lceil(p^{e}-1)\Delta\rceil,\mathcal{O}_{X})\cong F^{e}_{*}\mathcal{O}_{X}(-(p^{e}-1)K_{X}-\lceil(p^{e}-1)\Delta\rceil). (2.3)

Composing this isomorphism (over the global sections) with the evaluation at Fe1F^{e}_{*}1 map, we obtain the trace map:

TrΔe:H0(X,Fe(𝒪X((1pe)KX(pe1)Δ)))k=H0(X,𝒪X).\text{Tr}^{e}_{\Delta}:H^{0}\Big{(}X,F^{e}_{*}\big{(}\mathcal{O}_{X}((1-p^{e})K_{X}-\lceil(p^{e}-1)\Delta\rceil)\big{)}\Big{)}\to k=H^{0}(X,\mathcal{O}_{X}). (2.4)
Lemma 2.12.

The kernel of the trace map TrΔe\text{Tr}^{e}_{\Delta} in Equation 2.4 is exactly the subspace FeIeΔ((1pe)KX(pe1)Δ)F^{e}_{*}I_{e}^{\Delta}((1-p^{e})K_{X}-\lceil(p^{e}-1)\Delta\rceil). See 2.8 for the definition of the subspace IeΔI_{e}^{\Delta}.

Proof.

A section fH0(X,𝒪X((1pe)KX(pe1)Δ))f\in H^{0}(X,\mathcal{O}_{X}((1-p^{e})K_{X}-\lceil(p^{e}-1)\Delta\rceil)) is contained in the corresponding IeΔI_{e}^{\Delta}-subspace if and only if for every

φHom𝒪X(Fe(𝒪X((1pe)KX)),𝒪X)H0(X,Fe𝒪X)Fek,\varphi\in\operatorname{Hom}_{\mathcal{O}_{X}}(F^{e}_{*}(\mathcal{O}_{X}((1-p^{e})K_{X})),\mathcal{O}_{X})\cong H^{0}(X,F^{e}_{*}\mathcal{O}_{X})\cong F^{e}_{*}k,

we have φ(Fef)=0\varphi(F^{e}_{*}f)=0. But, Hom𝒪X(Fe(𝒪X((1pe)KX)),𝒪X)\operatorname{Hom}_{\mathcal{O}_{X}}(F^{e}_{*}(\mathcal{O}_{X}((1-p^{e})K_{X})),\mathcal{O}_{X}) is a one dimensional kk-vector space, and is generated by the trace map Tre\text{Tr}^{e} (where Δ=0\Delta=0). Now, the map TrΔe\text{Tr}^{e}_{\Delta} is just the restriction of the trace map Tre\text{Tr}^{e} to the subspace H0(X,Fe(𝒪X((1pe)KX(pe1)Δ)))H^{0}\Big{(}X,F^{e}_{*}\big{(}\mathcal{O}_{X}((1-p^{e})K_{X}-\lceil(p^{e}-1)\Delta\rceil)\big{)}\Big{)}. Thus, the lemma follows. ∎

Lemma 2.13.

Let (X,Δ)(X,\Delta) be a normal projective pair, and DD be any Weil divisor on XX. Then, denoting D1=(1pe)KX(pe1)ΔDD_{1}=(1-p^{e})K_{X}-\lceil(p^{e}-1)\Delta\rceil-D and D2=(1pe)KX(pe1)ΔD_{2}=(1-p^{e})K_{X}-\lceil(p^{e}-1)\Delta\rceil for any e1e\geq 1, we have a non-degenerate pairing

H0(D)IeΔ(D)×H0(D1)IeΔ(D1)H0(D2)IeΔ(D2)\frac{H^{0}(D)}{I_{e}^{\Delta}(D)}\times\frac{H^{0}(D_{1})}{I_{e}^{\Delta}(D_{1})}\to\frac{H^{0}(D_{2})}{I_{e}^{\Delta}(D_{2})}

obtained from multiplication (and reflexifying) global sections. In particular,

dimkH0(D)IeΔ(D)=dimkH0(D1)IeΔ(D1).\dim_{k}\frac{H^{0}(D)}{I_{e}^{\Delta}(D)}=\dim_{k}\frac{H^{0}(D_{1})}{I_{e}^{\Delta}(D_{1})}.
Proof.

Using Equation 2.2, the natural multiplication map

H0(D)×H0(D1)H0(D2)H^{0}(D)\times H^{0}(D_{1})\to H^{0}(D_{2})

can be identified with the evaluation map

H0(Fe(𝒪X(D)))×Hom𝒪X(Fe(𝒪X((pe1)Δ+D)),𝒪X)Hom𝒪X(Fe𝒪X((pe1)Δ),𝒪X)H^{0}\Big{(}F^{e}_{*}\big{(}\mathcal{O}_{X}(D)\big{)}\Big{)}\times\operatorname{Hom}_{\mathcal{O}_{X}}\Big{(}F^{e}_{*}\big{(}\mathcal{O}_{X}(\lceil(p^{e}-1)\Delta\rceil+D)\big{)},\mathcal{O}_{X}\Big{)}\to\operatorname{Hom}_{\mathcal{O}_{X}}(F^{e}_{*}\mathcal{O}_{X}(\lceil(p^{e}-1)\Delta\rceil),\mathcal{O}_{X})

where we identify Hom𝒪X(Fe𝒪X((pe1)Δ+D),𝒪X)\operatorname{Hom}_{\mathcal{O}_{X}}(F^{e}_{*}\mathcal{O}_{X}(\lceil(p^{e}-1)\Delta\rceil+D),\mathcal{O}_{X}) and Hom𝒪X(Fe𝒪X((pe1)Δ),𝒪X)\operatorname{Hom}_{\mathcal{O}_{X}}(F^{e}_{*}\mathcal{O}_{X}(\lceil(p^{e}-1)\Delta\rceil),\mathcal{O}_{X}) as subspaces of Hom𝒪X(Fe𝒪X(D),𝒪X)\operatorname{Hom}_{\mathcal{O}_{X}}(F^{e}_{*}\mathcal{O}_{X}(D),\mathcal{O}_{X}) and Hom𝒪X(Fe𝒪X,𝒪X)\operatorname{Hom}_{\mathcal{O}_{X}}(F^{e}_{*}\mathcal{O}_{X},\mathcal{O}_{X}) respectively, both via the natural inclusion 𝒪X𝒪X((pe1)Δ)\mathcal{O}_{X}\to\mathcal{O}_{X}(\lceil(p^{e}-1)\Delta\rceil). Therefore, a section fH0(D)f\in H^{0}(D) is contained in IeΔ(D)I_{e}^{\Delta}(D) if and only if for all sections gH0(D1)g\in H^{0}(D_{1}), the multiplication fgf\,g is contained in IeΔ(D2)I_{e}^{\Delta}(D_{2}). By symmetry, a section gH0(D1)g\in H^{0}(D_{1}) is contained in IeΔ(D1)I_{e}^{\Delta}(D_{1}) if and only if for all sections fH0(D)f\in H^{0}(D), gfIeΔ(D2)g\,f\in I_{e}^{\Delta}(D_{2}). This proves there is a well defined, and non-degenerate pairing as needed.

Finally, we note that since by 2.12, IeΔ(D2)I_{e}^{\Delta}(D_{2}) is the kernel of the trace map (Equation 2.4), the vector space H0(D2)/IeΔ(D2)H^{0}(D_{2})/I_{e}^{\Delta}(D_{2}) is either one-dimensional over kk, or equal to 0. In either case, the equality of dimensions follows. ∎

The next Proposition allows us to perturb by any divisor while computing the FF-signature of a section ring.

Proposition 2.14.

Let XX be a normal projective varitey over kk and LL an ample divisor on XX. Assume H0(X,𝒪X)=kH^{0}(X,\mathcal{O}_{X})=k. Fix a (not necessarily effective) Weil divisor DD on XX. Then, there exists a constant C>0C>0 (depending only on DD and LL) such that

|dimkH0(mL)Ie(mL)dimkH0(mL+D)Ie(mL+D)|Cpe(dim(X)1)\left|\dim_{k}\frac{H^{0}(mL)}{I_{e}(mL)}-\dim_{k}\frac{H^{0}(mL+D)}{I_{e}(mL+D)}\right|\leq Cp^{e(\dim(X)-1)}

for all m>0m>0 and e>0e>0.

Proof.

First we prove the case when DD is effective: For any m1m\geq 1, by using the natural map 𝒪X(mL)𝒪X(mL+D)\mathcal{O}_{X}(mL)\to\mathcal{O}_{X}(mL+D), we will view H0(mL)H^{0}(mL) as a subspace of H0(mL+D)H^{0}(mL+D). Let Je(mL)J_{e}(mL) denote the subspace H0(mL)Ie(mL+D)H^{0}(mL)\cap I_{e}(mL+D). By [LP23, Lemma 4.12], we see that Ie(mL)Je(mL)I_{e}(mL)\subset J_{e}(mL). Moreover, by Equation 4.12 in [LP23, Proof of Lemma 4.14], we have

dimkH0(mL+D)Ie(mL+D)=dimkH0(mL)Je(mL)+dimkH0(mL+D)H0(mL)+Ie(mL+D).\dim_{k}\frac{H^{0}(mL+D)}{I_{e}(mL+D)}=\dim_{k}\frac{H^{0}(mL)}{J_{e}(mL)}+\dim_{k}\frac{H^{0}(mL+D)}{H^{0}(mL)+I_{e}(mL+D)}.

Using this and the triangle inequality, we obtain that

|dimkH0(mL)Ie(mL)dimkH0(mL+D)Ie(mL+D)|dimkH0(mL+D)H0(mL)+dimkJe(mL)Ie(mL)\left|\dim_{k}\frac{H^{0}(mL)}{I_{e}(mL)}-\dim_{k}\frac{H^{0}(mL+D)}{I_{e}(mL+D)}\right|\leq\dim_{k}\frac{H^{0}(mL+D)}{H^{0}(mL)}+\dim_{k}\frac{J_{e}(mL)}{I_{e}(mL)} (2.5)

for all m,e>0m,e>0. Next, to compute the second term in the above inequality, fix an e>0e>0 and set Δe=1pe1D\Delta_{e}=\frac{1}{p^{e}-1}D. Then, we observe that the subspace Je(mL)J_{e}(mL) is exactly the same as IeΔe(mL)I_{e}^{\Delta_{e}}(mL) (2.8). Moreover, we also similarly have

IeΔe((1pe)KXmLD)=H0((1pe)KXmLD)Ie((1pe)KXmL).I_{e}^{\Delta_{e}}((1-p^{e})K_{X}-mL-D)=H^{0}((1-p^{e})K_{X}-mL-D)\cap I_{e}((1-p^{e})K_{X}-mL). (2.6)

Thus, by 2.13, we have

dimkH0(mL)Ie(mL)=dimkH0((1pe)KXmL)Ie((1pe)KXmL)\dim_{k}\frac{H^{0}(mL)}{I_{e}(mL)}=\dim_{k}\frac{H^{0}((1-p^{e})K_{X}-mL)}{I_{e}((1-p^{e})K_{X}-mL)}

and similarly,

dimkH0(mL)Je(mL)=dimkH0((1pe)KXmLD)IeΔe((1pe)KXmLD).\dim_{k}\frac{H^{0}(mL)}{J_{e}(mL)}=\dim_{k}\frac{H^{0}((1-p^{e})K_{X}-mL-D)}{I_{e}^{\Delta_{e}}((1-p^{e})K_{X}-mL-D)}.

By Equation 2.6, we see the natural map from H0((1pe)KXmLD)H^{0}((1-p^{e})K_{X}-mL-D) to H0((1pe)KXmL)H^{0}((1-p^{e})K_{X}-mL) restricts to an injective map

H0((1pe)KXmLD)IeΔe((1pe)KXmLD)H0((1pe)KXmL)Ie((1pe)KXmL).\frac{H^{0}((1-p^{e})K_{X}-mL-D)}{I_{e}^{\Delta_{e}}((1-p^{e})K_{X}-mL-D)}\hookrightarrow\frac{H^{0}((1-p^{e})K_{X}-mL)}{I_{e}((1-p^{e})K_{X}-mL)}.

By considering the cokernel of this map, we get that

dimkJe(mL)Ie(mL)=dimkH0(mL)Ie(mL)dimkH0(mL)Je(mL)dimkH0((1pe)KXmL)H0((1pe)KXmLD).\dim_{k}\frac{J_{e}(mL)}{I_{e}(mL)}=\dim_{k}\frac{H^{0}(mL)}{I_{e}(mL)}-\dim_{k}\frac{H^{0}(mL)}{J_{e}(mL)}\leq\dim_{k}\frac{H^{0}((1-p^{e})K_{X}-mL)}{H^{0}((1-p^{e})K_{X}-mL-D)}. (2.7)

Pick an M0M\gg 0 such that mLmL admits a global section that doesn’t vanish along DD for all mMm\geq M (this is possible since LL is ample). Using the standard exact sequences to restrict to DD, we see that

dimkH0((1pe)KXmL)H0((1pe)KXmLD)vol(KX|D)(d1)!pe(d1)+o(pe(d2))\dim_{k}\frac{H^{0}((1-p^{e})K_{X}-mL)}{H^{0}((1-p^{e})K_{X}-mL-D)}\leq\frac{\mathrm{vol}(-K_{X}|_{D})}{(d-1)!}p^{e(d-1)}+o(p^{e(d-2)}) (2.8)

and

dimkH0(mL+D)H0(mL)vol(L|D)(d1)!md1+o(md2).\dim_{k}\frac{H^{0}(mL+D)}{H^{0}(mL)}\leq\frac{\mathrm{vol}(L|_{D})}{(d-1)!}m^{d-1}+o(m^{d-2}). (2.9)

Finally, by [LP23, Theorem 4.9], we may pick a constant C2>0C_{2}>0 such that H0(mL)=Ie(mL)H^{0}(mL)=I_{e}(mL) and H0(mL+D)=Ie(mL+D)H^{0}(mL+D)=I_{e}(mL+D) for m>C2pem>C_{2}p^{e}. Therefore, to prove the Proposition, it is enough to consider the case when mC2pem\leq C_{2}p^{e}. 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 DD is effective.

More generally, we first pick an r0r\gg 0 such that rLrL and D+rLD+rL are both effective. Then, for any e1e\geq 1 and m>rm>r, we have

|dimkH0(mL)Ie(mL)dimkH0(mL+D)Ie(mL+D)||dimkH0(mL)Ie(mL)dimkH0((mr)L)Ie((mr)L)|+|dimkH0((mr)L)Ie((mr)L)dimkH0((mr)L+rL+D)Ie((mr)L+rL+D)|.\begin{split}\left|\dim_{k}\frac{H^{0}(mL)}{I_{e}(mL)}-\dim_{k}\frac{H^{0}(mL+D)}{I_{e}(mL+D)}\right|\leq&\left|\dim_{k}\frac{H^{0}(mL)}{I_{e}(mL)}-\dim_{k}\frac{H^{0}((m-r)L)}{I_{e}((m-r)L)}\right|\\ +&\left|\dim_{k}\frac{H^{0}((m-r)L)}{I_{e}((m-r)L)}-\dim_{k}\frac{H^{0}((m-r)L+rL+D)}{I_{e}((m-r)L+rL+D)}\right|.\end{split}

Now, we may apply the previous case of the Proposition (since both D+rLD+rL and rLrL are effective) to each of the two terms in the above inequality. Since rr was independent of ee, this completes the proof of the Proposition . ∎

2.7. FF-regularity:

Definition 2.15 (Sharp FF-splitting).

[SS10, Definition 3.1] Let XX be a normal variety over kk and Δ0\Delta\geq 0 be an effective \mathbb{Q}-divisor. The pair (X,Δ)(X,\Delta) is said to be globally sharply FF-split (resp. locally sharply FF-split) if there exists an integer e0e\gg 0, such that, the natural map

𝒪XFe𝒪X((pe1)Δ)\mathcal{O}_{X}\to F^{e}_{*}\mathcal{O}_{X}(\lceil(p^{e}-1)\Delta\rceil)

splits (resp. splits locally) as a map of 𝒪X\mathcal{O}_{X}-modules. A normal variety XX is said to globally FF-split if the pair (X,0)(X,0) is globally sharply FF-split.

Definition 2.16 (FF-regularity).

[SS10, Definition 3.1] Let XX be a normal variety over kk and Δ0\Delta\geq 0 be an effective \mathbb{Q}-divisor. The pair (X,Δ)(X,\Delta) is said to be globally FF-regular (resp. locally strongly FF-regular) if for any effective Weil divisor DD on XX, there exists an integer e0e\gg 0, such that, the natural map

𝒪XFe𝒪X((pe1)Δ+D)\mathcal{O}_{X}\to F^{e}_{*}\mathcal{O}_{X}(\lceil(p^{e}-1)\Delta\rceil+D)

splits (resp. splits locally) as a map of 𝒪X\mathcal{O}_{X}-modules. A normal variety XX is said to globally FF-regular if the pair (X,0)(X,0) is globally FF-regular. Similarly, a ring RR is called strongly FF-regular if the pair (Spec(R),0)(\operatorname{\text{Spec}}(R),0) is locally (equivalently, globally) strongly FF-regular.

Remark 2.17.

When X=Spec(R)X=\operatorname{\text{Spec}}(R) is an affine variety and Δ\Delta is an effective \mathbb{Q}-divisor, the pair (X,Δ)(X,\Delta) being globally FF-regular (resp. globally sharply FF-split) is equivalent to the pair (R,Δ)(R,\Delta) being locally strongly FF-regular (resp. locally sharply FF-split) [SS10].

Remark 2.18.

A local domain RR is strongly FF-regular if and only if its FF-signature 𝓈(R)\mathscr{s}(R) is positive [AL03]. More generally, a pair (R,Δ)(R,\Delta) (where RR is normal, local) is strongly FF-regular if and only if the FF-signature s(R,Δ)s(R,\Delta) (2.6) is positive [BST11, Theorem 3.18].

Theorem 2.19.

[Smi00, Theorem 3.10] Let XX be a projective variety over kk. Then, XX is globally FF-regular if and only if the section ring S(X,L)S(X,L) (2.3) with respect to some (equivalently, every) ample invertible sheaf LL is strongly FF-regular.

Remark 2.20.

(Locally) Strongly FF-regular varieties are normal and Cohen-Macaulay. Similarly, globally FF-regular varieties enjoy many of nice properties such as:

  • As proved in [SS10, Theorem 4.3], they are log-Fano type. More precisely, there exists an effective divisor Δ0\Delta\geq 0 such that the pair (X,Δ)(X,\Delta) is globally FF-regular and KXΔ-K_{X}-\Delta is ample.

  • A version of the Kawamata-Viehweg vanishing theorem holds on all globally FF-regular varieties [SS10, Theorem 6.8].

Theorem 2.21 ([Smi00], Corollary 4.3).

Let XX be a projective, globally FF-regular variety over kk. Suppose LL is a nef invertible sheaf over XX. Then,

Hi(X,L)=0for all i>0.H^{i}(X,L)=0\quad\text{for all $i>0$}.

We need a slight variation of 2.21 for \mathbb{Q}-ample divisors that we prove here for completeness.

Proposition 2.22.

Let XX be a globally FF-split normal variety and LL be a \mathbb{Q}-ample Weil divisor i.e., LL is an integral Weil divisor such that rLrL is an ample Cartier divisor for some integer r>0r>0. Then,

Hi(X,𝒪X(L))=0for i>0.H^{i}(X,\mathcal{O}_{X}(L))=0\quad\text{for $i>0$.}
Proof.

Let rr be an integer such that rLrL is Cartier. Write r=pe0sr=p^{e_{0}}s such that ss is coprime to pp. Pick an e>0e>0 such that ss divides pe1p^{e}-1. Then, since pe0(pne1)p^{e_{0}}(p^{ne}-1) is a multiple of rr for all n>0n>0, using Serre vanishing theorem, we have

Hi(X,𝒪X(pne+e0L))=Hi(X,𝒪X(pe0L+(pne1)pe0L)=0for all i>0 and n0.H^{i}(X,\mathcal{O}_{X}(p^{ne+e_{0}}L))=H^{i}(X,\mathcal{O}_{X}(p^{e_{0}}L+(p^{ne}-1)p^{e_{0}}L)=0\quad\text{for all $i>0$ and $n\gg 0$.} (2.10)

Since the map

𝒪XFne+e0𝒪X\mathcal{O}_{X}\to F_{*}^{ne+e_{0}}\mathcal{O}_{X}

is split, twisting by 𝒪X(L)\mathcal{O}_{X}(L) and reflexifying, we get that

𝒪X(L)Fne+e0𝒪X(pne+e0L)\mathcal{O}_{X}(L)\to F_{*}^{ne+e_{0}}\mathcal{O}_{X}(p^{ne+e_{0}}L)

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 \mathbb{Q}-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 X=Spec(R)X=\operatorname{\text{Spec}}(R), where (R,𝔪)(R,\mathfrak{m}) is an FF-finite, strongly FF-regular, local ring and DD be an integral Weil-divisor on XX such that rDrD is Cartier for some integer rr. Then, for each m0m\geq 0,

  1. (1)

    There exists an e0e\gg 0 such that the module R(mD)=H0(X,𝒪X(mD))R(mD)=H^{0}(X,\mathcal{O}_{X}(mD)) is isomorphic to an RR-module summand of FeRF^{e}_{*}R. In particular, R(mD)R(mD) is a Cohen-Macaulay module over RR.

  2. (2)

    Suppose that x1,,xtx_{1},\dots,x_{t} is a regular sequence on RR such that the ring R/(x1,,xt)RR/(x_{1},\dots,x_{t})R is normal. Then, the sheaf 𝒪X(mD)R𝒪Y\mathcal{O}_{X}(mD)\otimes_{R}\mathcal{O}_{Y} is reflexive on Y=Spec(R/(x1,,xt)R)Y=\operatorname{\text{Spec}}(R/(x_{1},\dots,x_{t})R). Furthermore, if we assume that the support of DD does not contain the subscheme YY, then natural map

    𝒪X(mD)R𝒪Y𝒪Y(mDY)\mathcal{O}_{X}(mD)\otimes_{R}\mathcal{O}_{Y}\to\mathcal{O}_{Y}(mD_{Y}) (2.11)

    is an isomorphism, where DYD_{Y} denotes the restriction of DD to YY (see the description in the proof below).

Proof.

By adding a principal divisor if necessary (which leaves the module R(mD)R(mD) isomorphic), we assume that DD is effective.

  1. (1)

    Since RR is strongly FF-regular, there exists an e0e\gg 0 such that the map 𝒪XFe(𝒪X(sD))\mathcal{O}_{X}\to F^{e}_{*}(\mathcal{O}_{X}(sD)) splits for each 0sr0\leq s\leq r. Once we have such an ee, choose srs\leq r such that pe+sp^{e}+s is divisible by rr. Thus, we may assume that the map (got by twisting by DD and reflexifying):

    𝒪X(D)Fe𝒪X((pe+s)D)\mathcal{O}_{X}(D)\to F^{e}_{*}\mathcal{O}_{X}((p^{e}+s)D)

    is split. Since rr divides pe+sp^{e}+s, 𝒪X((pe+s)D)\mathcal{O}_{X}((p^{e}+s)D) is isomorphic to 𝒪X\mathcal{O}_{X} since RR is local and rDrD is Cartier. Therefore, taking global sections, we have that the map

    R(mD)FeRR(mD)\to F^{e}_{*}R (2.12)

    is split. Note that the ee obtained is independent of mm. The Cohen-Macaulayness of R(mD)R(mD) follows because FeRF^{e}_{*}R is a Cohen-Macaulay module over RR, since RR itself is Cohen-Macaulay.

  2. (2)

    Firstly, we may assume m=1m=1 since the discussion holds for an arbitrary Weil divisor and is compatible with addition of Weil-divisors. Now, since YY is a normal, complete intersection subscheme of XX, we may “restrict" the rank one reflexive sheaf :=𝒪X(D)\mathcal{F}:=\mathcal{O}_{X}(D) on XX to a reflexive sheaf Y\mathcal{F}_{Y} on YY as follows: Let UU be the regular locus of YY. Then there is an open subset VXregV\subset X_{\text{reg}} (where XregX_{\text{reg}} denotes the regular locus of XX) such that VY=UV\cap Y=U. This is possible because YY is a complete intersection in XX. Therefore, we may restrict \mathcal{F} to VV and then to an invertible sheaf on UU, since |V\mathcal{F}|_{V} is invertible. Define Y\mathcal{F}_{Y} to be

    Y:=i(|U)\mathcal{F}_{Y}:=i_{*}(\mathcal{F}|_{U})

    where i:UYi:U\to Y is the inclusion. Then, Y\mathcal{F}_{Y} is a rank one reflexive sheaf on YY because YY is normal and UU contains all the codimension one points of YY. Thus, we can write Y\mathcal{F}_{Y} as 𝒪Y(DY)\mathcal{O}_{Y}(D_{Y}) for some Weil-divisor DYD_{Y} on YY. Furthermore, if Supp(D)\text{Supp}(D) does not contain YY, then since YY is normal, hence integral, DD naturally restricts to a Cartier divisor DUD_{U} on UU (given by restricting the equation for DD) and we may take DYD_{Y} to be the closure of DUD_{U}. 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 |U\mathcal{F}|_{U} is the restriction of the sheaf \mathcal{F} (i.e., isomorphic to 𝒪X𝒪U\mathcal{F}\otimes_{\mathcal{O}_{X}}\mathcal{O}_{U}), there is a natual map

    𝒪X𝒪YY\mathcal{F}\otimes_{\mathcal{O}_{X}}\mathcal{O}_{Y}\to\mathcal{F}_{Y}

    which is an isomorphism if and only if 𝒪X𝒪Y\mathcal{F}\otimes_{\mathcal{O}_{X}}\mathcal{O}_{Y} is reflexive. Therefore, it is sufficient to show that the module R(D)RR/(x1,,xt)RR(D)\otimes_{R}R/(x_{1},\dots,x_{t})R satisfies the S2 condition on R/(x1,,xt)RR/(x_{1},\dots,x_{t})R (since R/(x1,,xt)RR/(x_{1},\dots,x_{t})R is normal). But since R(mD)R(mD) is Cohen-Macaulay by Part (1) (and clearly full dimensional), and x1,,xtx_{1},\dots,x_{t} is a regular sequence on RR, we get that R(mD)R/(x1,,xt)RR(mD)\otimes R/(x_{1},\dots,x_{t})R is Cohen-Macaulay as well. This completes the proof of the Proposition. ∎

3. The αF\alpha_{F}-invariant of section rings

In analogy with Tian’s α\alpha-invariant in complex geometry, we define the “Frobenius-α\alpha" invariant (denoted by αF\alpha_{F}) for any pair (X,L)(X,L) where XX is a globally FF-regular projective variety (2.16) and LL is an ample Cartier divisor on LL.

Throughout this section, by a section ring SS, we mean that SS is the section ring S(X,L)S(X,L) of some projective variety X(Proj(S))X\,(\cong\operatorname{\text{Proj}}(S)) with respect to some ample line bundle L(𝒪X(1))L\,(\cong\mathcal{O}_{X}(1)) on XX (see 2.3).

3.1. Definitions

Definition 3.1.

Let (R,𝔪)(R,\mathfrak{m}) be an FF-finite normal local ring. Assume that RR is strongly FF-regular (2.16). Then, we define the FF-pure threshold of an effective \mathbb{Q}-divisor D0D\geq 0 on XX to be:

fpt𝔪(R,D):=sup{λ|(R,λD) is sharply F-split}.\text{fpt}_{\mathfrak{m}}(R,D):=\sup\{\,\lambda\,|\,(R,\lambda D)\text{ is sharply $F$-split}\,\}.

Moreover, by [SS10, Lemma 4.9], it follows that the FF-pure threshold of (R,D)(R,D) is equivalently characterized as the supremum of the set {λ|(R,λD) is strongly F-regular}\{\,\lambda\,|\,(R,\lambda D)\text{ is strongly $F$-regular}\}.

If DD is the principal divisor corresponding to a function fRf\in R, we write ff instead of DD in the notation for the FF-pure threshold.

Definition 3.2.

Let (S,𝔪)(S,\mathfrak{m}) be a strongly FF-regular section ring of a projective variety. Then, we define

αF(S)=inf{fpt𝔪(S,f)deg(f)| 0fS homogeneous element}.\alpha_{F}(S)=\inf\{\,\text{fpt}_{\mathfrak{m}}(S,f)\,\deg(f)\,|\,0\neq f\in S\text{ homogeneous element}\,\}.

If SS is the section ring of a projective variety XX with respect to an ample divisor LL, we may also use αF(X,L)\alpha_{F}(X,L) to denote αF(S)\alpha_{F}(S).

We have the following equivalent ways of characterizing the αF\alpha_{F}-invariant of a section ring.

Lemma 3.3.

Let XX be a globally FF-regular projective variety and LL be an ample Cartier divisor on XX. Let S=S(X,L)S=S(X,L) be the section ring of XX with respect to LL. Then, αF(S)\alpha_{F}(S) from 3.2 is equal to the supremum of any of the following sets:

  1. (1)

    The set of λ0\lambda\geq 0 such that the pair (S,λnD)(S,\frac{\lambda}{n}D) is sharply FF-split for every nn\in\mathbb{N} and every effective divisor DnLD\sim nL.

  2. (2)

    The set of λ0\lambda\geq 0 such that the pair (S,λnD)(S,\frac{\lambda}{n}D) is strongly FF-regular for every nn\in\mathbb{N} and every effective divisor DnLD\sim nL.

  3. (3)

    The set of λ0\lambda\geq 0 such that the pair (X,λnD)(X,\frac{\lambda}{n}D) is globally sharply FF-split for every nn\in\mathbb{N} and every effective divisor DnLD\sim nL.

  4. (4)

    The set of λ0\lambda\geq 0 such that the pair (X,λnD)(X,\frac{\lambda}{n}D) is globally FF-regular for every nn\in\mathbb{N} and every effective divisor DnLD\sim nL.

  5. (5)

    The set of λ0\lambda\geq 0 such that the pair (X,λD)(X,\lambda D) is globally FF-regular for every effective \mathbb{Q}-divisor DLD\sim_{\mathbb{Q}}L.

Proof.

Statements (1) and (2) follow immediately from the definition of the αF\alpha_{F}-invariant and the definition of the FF-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 \mathbb{Q}-divisor DLD\sim_{\mathbb{Q}}L is of the form 1nnD\frac{1}{n}nD for some effective Cartier divisor nDnLnD\sim nL. ∎

Next, we explain a more precise connection between the αF\alpha_{F}-invariant and Frobenius splittings in SS.

Proposition 3.4.

Let SS be a strongly FF-regular section ring and α(S)\alpha^{\prime}(S) denote the supremum of the following set:

𝒜(S):={λ0| for any integers e1 and mλ(pe1), we have Ie(m)=0}.\mathscr{A}(S):=\{\lambda\in\mathbb{R}_{\geq 0}\,|\,\text{ for any integers $e\geq 1$ and $m\leq\lambda(p^{e}-1)$, we have }I_{e}(m)=0\}.

Then, α(S)=αF(S)\alpha^{\prime}(S)=\alpha_{F}(S). Moreover, αF(S)\alpha_{F}(S) belongs to the set 𝒜(S)\mathscr{A}(S).

Remark 3.5.

Note that in the above Proposition, it is unclear if the set 𝒜(S)\mathscr{A}(S) contains any non-zero element. This is equivalent to the positivity of αF(S)\alpha_{F}(S) 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.

Let (S,𝔪)(S,\mathfrak{m}) be an FF-finite, strongly FF-regular local ring. Fix an effective Weil-divisor DD on X=Spec(S)X=\textrm{Spec}(S). Then, for any fixed e0>0e_{0}>0, let ψe0\psi_{e_{0}} denote the natural map

ψe0:𝒪XFe0(𝒪X(D)).\psi_{e_{0}}:\mathcal{O}_{X}\to F_{*}^{e_{0}}(\mathcal{O}_{X}(D)).

Then, the following are equivalent:

  1. (1)

    The map ψe0\psi_{e_{0}} splits as a map of 𝒪X\mathcal{O}_{X}-modules.

  2. (2)

    The pair (X,1pe01D)(X,\frac{1}{p^{e_{0}}-1}D) is sharply FF-split (2.15).

  3. (3)

    The FF-pure threshold of (X,D)(X,D) is at least 1pe01\frac{1}{p^{e_{0}}-1} (3.1).

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 fpt𝔪(X,D)1pe01\text{fpt}_{\mathfrak{m}}(X,D)\geq\frac{1}{p^{e_{0}}-1}, we must must have that the pair (X,1pe0D)(X,\frac{1}{p^{e_{0}}}D) is sharply FF-split. Thus, there is an e>0e>0 such that the natural map

ψe(D):𝒪XFe(𝒪X(pe1pe0D)\psi_{e}(D):\mathcal{O}_{X}\to F^{e}_{*}(\mathcal{O}_{X}(\lceil\frac{p^{e}-1}{p^{e_{0}}}\rceil D) (3.1)

splits. Since the same holds for ψne(D)\psi_{ne}(D) for any natural number n1n\geq 1 (see [Sch08, Proposition 3.3] for the proof), we get the map:

ψee0(D):𝒪XFee0(𝒪X(pee01pe0D))=Fee0(𝒪X(p(e1)e0D))\psi_{ee_{0}}(D):\mathcal{O}_{X}\to F_{*}^{ee_{0}}(\mathcal{O}_{X}(\lceil\frac{p^{ee_{0}}-1}{p^{e_{0}}}\rceil D))=F_{*}^{ee_{0}}(\mathcal{O}_{X}(p^{(e-1)e_{0}}D))

splits. Note that ψe0\psi_{e_{0}} as defined in Equation 3.1 matches with the map considered in the statement of the Lemma.

Let UXU\subset X denote the regular locus of XX. Since ϕe0\phi_{e_{0}} is a map between reflexive sheaves, to show that it splits, it sufficient to show that its restriction of UU splits. Over UU, we may construct the map ψee0\psi_{ee_{0}} as follows: First consider the map

ϕ(e1)e0:𝒪U(D)F(e1)e0𝒪U(p(e1)e0D)\phi_{(e-1)e_{0}}:\mathcal{O}_{U}(D)\to F_{*}^{(e-1)e_{0}}\mathcal{O}_{U}(p^{(e-1)e_{0}}D)

obtained by twisting the (e1)e0th(e-1)e_{0}^{\text{th}}-iterate of the Frobenius map by the invertible sheaf 𝒪U(D)\mathcal{O}_{U}(D). If ff denotes the local equation of DD, then ϕ(e1)e0\phi_{(e-1)e_{0}} is defined by sending

fF(e1)e0fp(e1)e0.f\mapsto F_{*}^{(e-1)e_{0}}f^{p^{(e-1)e_{0}}}.

Then, after restricting to UU, we have that

ψee0=Fe0ϕ(e1)e0(ψe0|U)\psi_{ee_{0}}=F_{*}^{e_{0}}\phi_{(e-1)e_{0}}\circ(\psi_{e_{0}}|_{U})

where the right hand side is the composition

Fe0ϕ(e1)e0(ψe0|U):𝒪UFe0𝒪U(D)Fee0(𝒪X(p(e1)e0D)).F_{*}^{e_{0}}\phi_{(e-1)e_{0}}\circ(\psi_{e_{0}}|_{U}):\mathcal{O}_{U}\to F_{*}^{e_{0}}\mathcal{O}_{U}(D)\to F_{*}^{ee_{0}}(\mathcal{O}_{X}(p^{(e-1)e_{0}}D)).

Therefore, if ψee0\psi_{ee_{0}} splits then so does ψe0\psi_{e_{0}}. This proves that part (3) implies part (1), completing the proof of the lemma. ∎

Proof of 3.4.

Set α=αF(S)\alpha=\alpha_{F}(S), and α=α(S)\alpha^{\prime}=\alpha^{\prime}(S) (which is defined in the satement of the Proposition). First we will prove that αα\alpha\leq\alpha^{\prime}. This is clear if α=0\alpha=0, so we assume that α\alpha is positive. For any non-zero element fSmf\in S_{m}, by definition of α\alpha, we must have fpt𝔪(S,f)αm\text{fpt}_{\mathfrak{m}}(S,f)\geq\frac{\alpha}{m}. So, if mα(pe1)m\leq\alpha(p^{e}-1), we have

fpt𝔪(S,f)αmαα(pe1)=1pe1.\text{fpt}_{\mathfrak{m}}(S,f)\geq\frac{\alpha}{m}\geq\frac{\alpha}{\alpha(p^{e}-1)}=\frac{1}{p^{e}-1}.

Thus, by 3.6, the map RFeRR\to F^{e}_{*}R sending 11 to FefF^{e}_{*}f splits. Since ff was an arbitrary non-zero element of degree mm, this shows that Ie(m)=0I_{e}(m)=0 whenever mpe1m\leq p^{e}-1. Therefore, α\alpha belongs to the set 𝒜(S)\mathscr{A}(S), which proves that αα\alpha\geq\alpha^{\prime}.

Next we prove that αα\alpha\geq\alpha^{\prime}. Fix any λ<α\lambda<\alpha^{\prime}. Then, whenever mλ(pe1)m\leq\lambda(p^{e}-1) and fSmf\in S_{m} is any non-zero element, we know that ff is not contained in Ie(m)I_{e}(m). By 3.6, we have fpt𝔪(S,f)1pe1\text{fpt}_{\mathfrak{m}}(S,f)\geq\frac{1}{p^{e}-1} . In other words, if ee is the smallest integer such that mλ(pe1)m\leq\lambda(p^{e}-1), (equivalently, e=logp(mλ)e=\lceil\log_{p}(\frac{m}{\lambda})\rceil), then fpt𝔪(S,f)1pe1.\text{fpt}_{\mathfrak{m}}(S,f)\geq\frac{1}{p^{e}-1}. Now combining this with the fact that fpt𝔪(S,fa)=fpt𝔪(S,f)a\text{fpt}_{\mathfrak{m}}(S,f^{a})=\frac{\text{fpt}_{\mathfrak{m}}(S,f)}{a} for any integer aa to get:

fpt𝔪(S,f)supa1aplogp(amλ)1λm.\text{fpt}_{\mathfrak{m}}(S,f)\geq\sup_{a\geq 1}\frac{a}{p^{\lceil\log_{p}(\frac{am}{\lambda})\rceil}-1}\geq\frac{\lambda}{m}. (3.2)

To see the right inequality, we make the following observations: Fixing mm and λ\lambda and for any aa, write

logp(a)+logp(mλ)=logp(a)+logp(mλ)+ε(a)\lceil\log_{p}(a)+\log_{p}(\frac{m}{\lambda})\rceil=\log_{p}(a)+\log_{p}(\frac{m}{\lambda})+\varepsilon(a)

for some non-negative real number ε(a)\varepsilon(a). Then, we have

infa1plogp(a)+logp(mλ)+ε(a)1a=infa1mλpε(a)1ainfa1mλpε(a)\inf_{a\geq 1}\frac{p^{\log_{p}(a)+\log_{p}(\frac{m}{\lambda})+\varepsilon(a)}-1}{a}=\inf_{a\geq 1}\frac{m}{\lambda}p^{\varepsilon(a)}-\frac{1}{a}\leq\inf_{a\geq 1}\frac{m}{\lambda}p^{\varepsilon(a)}

for each a1a\geq 1. So it is sufficient to show that

infa1pε(a)=1.\inf_{a\geq 1}p^{\varepsilon(a)}=1.

This is true because given any real number γ\gamma, the infimum of the set {γ+logp(a)γlogp(a)|a}\{\lceil\gamma+\log_{p}(a)\rceil-\gamma-\log_{p}(a)\,|\,a\in\mathbb{N}\} is zero. This proves the inequality in (3.2) .

Since ff was an arbitrary non-zero homogeneous element of degree mm, it follows from Equation 3.2 that αλ\alpha\geq\lambda. Since λ\lambda was an arbitrary number smaller than α\alpha^{\prime}, we must have αα\alpha\geq\alpha^{\prime} as well. This completes the proof that α=α\alpha=\alpha^{\prime}. ∎

3.2. Finite-degree approximations

Now we will define finite-degree approximations to the αF\alpha_{F}-invariant. This establishes a limit formula for the αF\alpha_{F}-invariant that is analogous to the FF-signature (see 2.6 and the classical definition in [Tuc12]).

Definition 3.7.

Let SS be an \mathbb{N}-graded section ring over kk. For each integer e1e\geq 1, we define

me(S):=max{m0|Ie(m)=0}m_{e}(S):=\max\{m\geq 0\,|\,I_{e}(m)=0\}

and define

αe(S):=me(S)pe.\alpha_{e}(S):=\frac{m_{e}(S)}{p^{e}}.
Theorem 3.8.

Let SS be a strongly FF-regular \mathbb{N}-graded section ring over kk. Then, we have

limeαe(S)=αF(S).\lim_{e\to\infty}\alpha_{e}(S)=\alpha_{F}(S).

In particular, the limit exists. See 3.2 for the definition of αF(S)\alpha_{F}(S).

Lemma 3.9.

Let SS be a strongly FF-regular \mathbb{N}-graded section ring over kk. For any e1e\geq 1, we have

αe(S)+1peαe+1(S)+1pe+1.\alpha_{e}(S)+\frac{1}{p^{e}}\geq\alpha_{e+1}(S)+\frac{1}{p^{e+1}}.
Proof.

First note that since SS is strongly FF-regular, SS is a normal domain. For any e1e\geq 1, let 0f0\neq f be an element of Ie(me+1)I_{e}(m_{e}+1). Then, we have that fpf^{p} is a non-zero element of Ie+1(p(me+1))I_{e+1}(p(m_{e}+1)) (see [Tuc12, Lemma 4.4]). This proves that

p(me+1)me+1+1.p(m_{e}+1)\geq m_{e+1}+1.

Dividing both sides by pe+1p^{e+1}, we obtain the required inequality. ∎

Proof of 3.8.

The sequence {αe+1pe}e1\{\alpha_{e}+\frac{1}{p^{e}}\}_{e\geq 1} 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 1pe\frac{1}{p^{e}} converges to zero, the sequence {αe}e1\{\alpha_{e}\}_{e\geq 1} also converges and

limeαe=infe1{αe+1pe}.\lim_{e\to\infty}\alpha_{e}=\inf_{e\geq 1}\{\alpha_{e}+\frac{1}{p^{e}}\}.

It remains to show that the limit is equal to αF(S)\alpha_{F}(S). Using the definition of αe\alpha_{e}, we have that

αF(S)pepe1(αe+1pe)=me+1pe1\alpha_{F}(S)\leq\frac{p^{e}}{p^{e}-1}(\alpha_{e}+\frac{1}{p^{e}})=\frac{m_{e}+1}{p^{e}-1}

for each e1e\geq 1. This is because we know that αF(S)\alpha_{F}(S) belongs to the set 𝒜(S)\mathscr{A}(S) from 3.4. Taking a limit over ee, we obtain

αF(S)limeαe.\alpha_{F}(S)\leq\lim_{e\to\infty}\alpha_{e}.

For the reverse inequality, setting α:=limαe\alpha:=\lim\alpha_{e}, we note that

α(pe1)(αe+1pe)(pe1)<peαe+1.\alpha(p^{e}-1)\leq(\alpha_{e}+\frac{1}{p^{e}})(p^{e}-1)<p^{e}\alpha_{e}+1.

By the definition of me=peαem_{e}=p^{e}\alpha_{e}, the subspace Ie(m)I_{e}(m) is equal to zero for each mα(pe1)mem\leq\alpha(p^{e}-1)\leq m_{e}. Thus, α\alpha belongs to the set 𝒜(S)\mathscr{A}(S) defined in 3.4. Since αF(S)\alpha_{F}(S) is the supremum of 𝒜(S)\mathscr{A}(S), we get that αF(S)α\alpha_{F}(S)\geq\alpha. This completes the proof of 3.8. ∎

3.3. Positivity and comparison to the FF-signature.

Next we will show that the αF\alpha_{F}-invariant is positive by comparing it to the FF-signature (2.6). Recall that for a section ring SS of any globally FF-regular projective variety, there exists a positive constant CC such that for any e>0e>0 and any mCpem\leq C\,p^{e}, we have Ie(m)=Sm.I_{e}(m)=S_{m}. This follows from [LP23, Theorem 4.9].

Theorem 3.10.

The αF\alpha_{F}-invariant of a strongly FF-regular section ring is positive. Moreover, setting α=αF(S)\alpha=\alpha_{F}(S) and fixing a constant CC as discussed above (so that for any e>0e>0 and any mCpem\leq C\,p^{e}, we have Ie(m)=SmI_{e}(m)=S_{m}), we have the following comparisons:

e(S)αdim(S)dim(S)!𝓈(S)e(S)dim(S)!(Cdim(S)(Cα)dim(S))\frac{e(S)\,\alpha^{\dim(S)}}{\dim(S)!}\,\leq\,\mathscr{s}(S)\,\leq\frac{e(S)}{\dim(S)!}\,\big{(}C^{\dim(S)}-(C-\alpha)^{\dim(S)}\big{)} (3.3)

where e(S)e(S) denotes the Hilbert-Samuel multiplicity of SS.

Lemma 3.11.

Given a non-zero homogeneous element ff in SS of degree nn, let λ>nfpt𝔪(S,f)\lambda>n\,\text{fpt}_{\mathfrak{m}}(S,f) be a real number. Then,

𝓈(S)e(S)dim(S)!(Cdim(S)(Cλ)dim(S)).\mathscr{s}(S)\leq\frac{e(S)}{\dim(S)!}\,\big{(}C^{\dim(S)}-(C-\lambda)^{\dim(S)}\big{)}. (3.4)
Proof.

Since we have assumed that λ>nfpt𝔪(S,f)\lambda>n\,\text{fpt}_{\mathfrak{m}}(S,f), there exist integers a,e0>0a,e_{0}>0 such that

fpt𝔪(S,f)<ape01<λd.\text{fpt}_{\mathfrak{m}}(S,f)<\frac{a}{p^{e_{0}}-1}<\frac{\lambda}{d}.

Replacing ff by faf^{a}, by 3.6 we may assume that the map SFe0SS\to F_{*}^{e_{0}}S defined by 1Fe0f1\to F_{*}^{e_{0}}f does not split (since fpt𝔪(S,fa)<1pe01)\text{fpt}_{\mathfrak{m}}(S,f^{a})<\frac{1}{p^{e_{0}}-1}). We may also assume that nλ(pe01)n\leq\lambda\,(p^{e_{0}}-1). Now, since ff belongs to the ideal Ie0I_{e_{0}}, we have SmnfIe0(m)S_{m-n}\cdot f\subset I_{e_{0}}(m) for any mdm\geq d yielding the inequality

dimkSmIe0(m)dimkSmdimkSmn.\dim_{k}\frac{S_{m}}{I_{e_{0}}(m)}\leq\dim_{k}S_{m}-\dim_{k}S_{m-n}. (3.5)

for all mnm\geq n. Further, setting vr=pre01pe01v_{r}=\frac{p^{re_{0}}-1}{p^{e_{0}}-1} for any integer rr, we have that fpt𝔪(S,fvr)<1pre01\text{fpt}_{\mathfrak{m}}(S,f^{v_{r}})<\frac{1}{p^{re_{0}}-1}, and so fvrf^{v_{r}} belongs to Ire0(nvr)I_{re_{0}}(nv_{r}). Therefore, we similarly have

dimkSmIne0(m)dimkSmdimkSmnvr\dim_{k}\frac{S_{m}}{I_{ne_{0}}(m)}\leq\dim_{k}S_{m}-\dim_{k}S_{m-nv_{r}} (3.6)

for all mnvrm\geq nv_{r}. Then, using the 2.11 we may compute the FF-signature 𝓈(S)\mathscr{s}(S) as follows:

𝓈(S)=1[k:k]limrm=0m=Cpre0dimkSmIre0(m)pre0dim(S)1[k:k](limrm=0m=Cpre0dimkSmpre0dim(S)m=dvrm=Cpre0dimkSmnvrpre0dim(S)),\mathscr{s}(S)=\frac{1}{[k^{\prime}:k]}\,\lim_{r\to\infty}\frac{\sum\limits_{m=0}^{m=Cp^{re_{0}}}\dim_{k}\frac{S_{m}}{I_{re_{0}}(m)}}{p^{re_{0}\dim(S)}}\leq\frac{1}{[k^{\prime}:k]}\,\Bigg{(}\lim_{r\to\infty}\frac{\sum\limits_{m=0}^{m=Cp^{re_{0}}}\dim_{k}S_{m}}{p^{re_{0}\dim(S)}}-\frac{\sum\limits_{m=dv_{r}}^{m=Cp^{re_{0}}}\dim_{k}S_{m-nv_{r}}}{p^{re_{0}\dim(S)}}\Bigg{)},

where kk^{\prime} is the field S0S_{0}. Here we have used Equation 3.6 and the defining property of the constant CC. Finally, calculating the dimensions in the above inequality using the formula

dimkSm=[k:k]e(S)(dimS1)!mdimS1+o(mdimS2),\dim_{k}S_{m}=[k^{\prime}:k]\,\frac{e(S)}{(\dim S-1)!}m^{\dim S-1}+o(m^{\dim S-2}),

we obtain

𝓈(S)e(S)dim(S)!(Cdim(S)(Cnpe01)dim(S)).\mathscr{s}(S)\leq\frac{e(S)}{\dim(S)!}\big{(}C^{\dim(S)}-(C-\frac{n}{p^{e_{0}}-1})^{\dim(S)}\big{)}.

The proof of the lemma is now complete by using the fact that λnpe01\lambda\geq\frac{n}{p^{e_{0}}-1}. ∎

Proof of 3.10.

We note that if αF(S)=0\alpha_{F}(S)=0, then the rightmost inequality of Equation 3.3 implies that the FF-signature 𝓈(S)\mathscr{s}(S) is zero. But this is a contradiction since SS was assumed to be strongly FF-regular (see [AL03, Theorem 0.2]). So the positivity of αF(S)\alpha_{F}(S) follows from Equation 3.3, which we will now prove.

The rightmost inequality follows from 3.11 by taking a limit as λαF(S)\lambda\to\alpha_{F}(S), since the Lemma applies to each λ\lambda such that λ>fpt𝔪(S,f)deg(f)\lambda>\text{fpt}_{\mathfrak{m}}(S,f)\,\deg(f) for some non-zero ff.

To prove the leftmost inequality, we use 2.11 again to compute the FF-signature of SS and we observe that for any e1e\geq 1,

𝓈(S)=limem=0m=CpedimkSmIe(m)pedim(S)limem=0m=medimkSmpedim(S).\mathscr{s}(S)=\lim_{e\to\infty}\frac{\sum\limits_{m=0}^{m=Cp^{e}}\dim_{k}\frac{S_{m}}{I_{e}(m)}}{p^{e\dim(S)}}\geq\lim_{e\to\infty}\frac{\sum\limits_{m=0}^{m=m_{e}}\dim_{k}S_{m}}{p^{e\dim(S)}}. (3.7)

Recall that for any e1e\geq 1, mem_{e} is the largest mm such that Ie(m)=0I_{e}(m)=0, which justifies Equation 3.7. But the right hand side is equal to

limee(S)dim(S)!(mepe)dim(S).\lim_{e\to\infty}\frac{e(S)}{\dim(S)!}\big{(}\frac{m_{e}}{p^{e}}\big{)}^{\dim(S)}.

We conclude the proof by using 3.8, which says that limemepe=αF(S)\lim_{e\to\infty}\frac{m_{e}}{p^{e}}=\alpha_{F}(S). This concludes the proof of Equation 3.3 and thus of 3.10. ∎

Remark 3.12.

The positivity of the αF\alpha_{F}-invariant proved in 3.10 can also be deduced from the main theorem of [Sat18].

3.4. Behaviour under certain ring extensions.

In this subsection, we record some useful results on the behaviour of the αF\alpha_{F}-invariant under suitably nice extensions of section rings.

Proposition 3.13.

Let SS and SS^{\prime} be two \mathbb{N}-graded section rings (of possibly different varieties) and SSS\subset S^{\prime} be an inclusion such that for a fixed integer n1n\geq 1 and any other mm, all degree mm elements of SS are mapped to degree nmnm elements of SS^{\prime}. Further, assume that the inclusion SSS\subset S^{\prime} splits as a map of SS-modules. Then, we have

αF(S)αF(S)n.\alpha_{F}(S)\geq\frac{\alpha_{F}(S^{\prime})}{n}. (3.8)

Moreover, equality holds in (3.8) if SS is the nthn^{\text{th}}-Veronese subring of SS^{\prime}.

Proof.

The first part follows immediately from 3.8 and the fact that a homogeneous element ff of SS splits from FeSF^{e}_{*}S if it splits from FeSF^{e}_{*}S^{\prime}. In other words, we have

Ie(Sm)Ie(Smn)I_{e}(S_{m})\subset I_{e}(S^{\prime}_{mn})

for any ee and mm. For the statement about Veronese subrings, the key observation is that since SS^{\prime} is a section ring (of (X,L)(X,L), say), then

Ie(Smn)=Ie(Sm)=Ie(X,Lmn).I_{e}(S^{\prime}_{mn})=I_{e}(S_{m})=I_{e}(X,L^{mn}).

Thus, the equality again follows from 3.8. ∎

Remark 3.14.

The FF-signature of section rings also transforms in a similar manner to the αF\alpha_{F}-invariant above. Indeed, see [VK12, Theorem 2.6.2] and its generalization in [CR17, Theorem 4.8]. A simple proof of this transformation rule for section rings can also be obtained by using 2.14.

Proposition 3.15.

Let (S,𝔪)(S,𝔫)(S,\mathfrak{m})\subset(S^{\prime},\mathfrak{n}) be a degree preserving map of \mathbb{N}-graded, strongly FF-regular section rings (of possibly different varieties and over possibly different perfect fields). Assume that both SS and SS^{\prime} are generated in degree one , SS^{\prime} is flat over SS, and that the ring S/𝔪SS^{\prime}/\mathfrak{m}S^{\prime} is regular. Then, we have

αF(S)=αF(S).\alpha_{F}(S^{\prime})=\alpha_{F}(S).
Proof.

First note that since SS and SS^{\prime} are generated in degree 11, the αF\alpha_{F}-invariant must be at most 11 for both of them. Next, for any e1e\geq 1 and mpe1m\leq p^{e}-1, we may apply Claim 3.4 in [CRST21, Proof of Theorem 3.1] to get that

Ie(S,m)S=Ie(S,m).I_{e}(S,m)\,S^{\prime}=I_{e}(S^{\prime},m).

This implies that me(S)=me(S)m_{e}(S)=m_{e}(S^{\prime}) for any e1e\geq 1. Now, the Proposition follows immediately by using 3.8. ∎

Recall that kk was assumed to be a perfect field of characteristic p>0p>0.

Corollary 3.16.

Let SS be a strongly FF-regular section ring over kk and KK be an arbitrary perfect field extension of kk. Then, the base-change SkKS\otimes_{k}K is isomorphic to a product of strongly FF-regular section rings SiS^{i} over finite extensions of KK and for each ii, we have

αF(S)=αF(Si).\alpha_{F}(S)=\alpha_{F}(S^{i}).
Proof.

Firstly, we may assume that SS is generated in degree one by using 3.13. Set S=SkKS^{\prime}=S\otimes_{k}K. Since kk is perfect and S/𝔪S/\mathfrak{m} is a finite separable extension of kk, we see that

S/𝔪SS/𝔪kKiLiS^{\prime}/\mathfrak{m}S^{\prime}\cong S/\mathfrak{m}\otimes_{k}K\cong\prod_{i}L_{i}

is a finite product of perfect fields LiL_{i}. Thus, if SS was the section ring of XX, then SiSiS^{\prime}\cong\prod_{i}S^{i} where SiS^{i} is the section ring of X×S/𝔪LiX\times_{S/\mathfrak{m}}L_{i}. Note that SiS^{i} is isomorphic to SS/𝔪LiS\otimes_{S/\mathfrak{m}}L_{i}, and hence each SiS^{i} is strongly FF-regular by [CRST21, Theorem 3.6]. The corollary now follows from 3.15 by applying it to each inclusion SSiS\subset S^{i}. ∎

Remark 3.17.

In the setting of 3.16, the same results also holds for the FF-signature of SS instead of the αF\alpha_{F}-invariant by using [Yao06, Theorem 5.6] in place of 3.15. This allows to reduce computing both the FF-signature and the αF\alpha_{F}-invariant of a section ring to the geometrically connected case, i.e., when S0=kS_{0}=k.

Remark 3.18.

The results of this section naturally extend to the more general setting of globally FF-regular pairs (X,Δ)(X,\Delta) such that (pe1)Δ(p^{e}-1)\Delta is an integral Weil-divisor. This will be addressed in a future version of the paper.

4. The αF\alpha_{F}-invariant of globally FF-regular Fano varieties.

In this section, we specialize the study of the αF\alpha_{F}-invariant to the case of globally FF-regular Fano varieties (and when the ample divisor is a multiple of KX-K_{X}). We begin by defining what we mean by a \mathbb{Q}-Fano variety in positive characteristic. Recall that kk denotes a perfect field of characteristic p>0p>0.

Definition 4.1.

A \mathbb{Q}-Fano variety XX is a projective variety over kk such that

  1. (1)

    XX is locally strongly FF-regular (2.16).

  2. (2)

    KXK_{X} is a \mathbb{Q}-Cartier divisor.

  3. (3)

    KX-K_{X} is ample.

Note that since XX has only strongly FF-regular singularities, XX is automatically normal and Cohen-Macaulay. In particular, we may define the canonical Weil-divisor KXK_{X} by extending a canonical divisor from the smooth locus. In fact, ωX=𝒪X(KX)\omega_{X}=\mathcal{O}_{X}(K_{X}) is a dualizing sheaf over XX. In particular, we have

Hd(X,ωX)kH^{d}(X,\omega_{X})\cong k (4.1)

where dd is the dimension of XX. Moreover, the second and third conditions in Definition 4.1 guarantee that there is a positive integer rr such that rKXrK_{X} is Cartier and rKX-rK_{X} is ample. The smallest such rr is called the index of KXK_{X}.

Definition 4.2.

Let XX be a globally FF-regular \mathbb{Q}-Fano variety over kk and rr be a positive integer divisible by the index of KXK_{X}. Let

S:=S(X,rKX)=m0H0(X,𝒪X(mrKX))S:=S(X,-rK_{X})=\bigoplus_{m\geq 0}H^{0}(X,\mathcal{O}_{X}(-mrK_{X}))

denote the section ring of XX with respect to rKX-rK_{X}. Then, the αF\alpha_{F}-invariant of XX is defined to be

αF(X):=rαF(S)\alpha_{F}(X):=r\,\alpha_{F}(S)

where αF(S)\alpha_{F}(S) denotes the αF\alpha_{F}-invariant of the strongly FF-regular ring SS, as defined in Defintion 3.2.

By taking the affine cone over a \mathbb{Q}-Fano variety, we also define the global FF-signature of the a \mathbb{Q}-Fano variety.

Definition 4.3 (FF-signature).

Let XX be a \mathbb{Q}-Fano variety over kk and rr denote a positive integer divisible by the index of KXK_{X}. Let

S:=S(X,rKX)=m0H0(X,𝒪X(mrKX))S:=S(X,-rK_{X})=\bigoplus_{m\geq 0}H^{0}(X,\mathcal{O}_{X}(-mrK_{X}))

denote the section ring of XX with respect to rKX-rK_{X}. Then, the FF-signature of XX is defined to be

𝓈(X):=r𝓈(S)\mathscr{s}(X):=r\,\mathscr{s}(S)

where 𝓈(S)\mathscr{s}(S) denotes the FF-signature of SS, as defined in Defintion 2.6.

Remark 4.4.

Though the definitions of the αF\alpha_{F}-invariant and the FF-signature involve making a choice of a multiple of the index of KXK_{X}, both invariants are well-defined thanks to 3.13 (for the αF\alpha_{F}-invariant) and [VK12, Theorem 2.6.2] (for the FF-signature).

Theorem 4.5.

Let XX be a globally FF-regular \mathbb{Q}-Fano variety of positive dimension. Then, αF(X)\alpha_{F}(X) is at most 1/21/2.

Proof of 4.5:.

Let dd denote the dimension of XX, and r0r\gg 0 be an integer divisible by the index of KXK_{X} and such that H0(X,𝒪X(mKX))0H^{0}(X,\mathcal{O}_{X}(-mK_{X}))\neq 0 for all mrm\geq r. First, we claim that there is an integer n>0n>0 such that

dimkH0(X,𝒪X(mKX))<dimkH0(X,𝒪X((m+n)KX))\dim_{k}H^{0}(X,\mathcal{O}_{X}(-mK_{X}))<\dim_{k}H^{0}(X,\mathcal{O}_{X}(-(m+n)K_{X}))

for all m0m\gg 0. This is clear if KX-K_{X} is a Cartier divisor (and we may take n=1n=1 in this case), since dimkH0(X,𝒪X(mKX))\dim_{k}H^{0}(X,\mathcal{O}_{X}(-mK_{X})) is a polynomial in mm of degree d>0d>0 and a positive leading term (because KX-K_{X} is ample). More generally, by the asymptotic Riemann-Roch formula ([Laz04, Example 1.2.19]), for each 0ir10\leq i\leq r-1, there exists polynomials PiP_{i} of degree dd such that for all m0m\gg 0

dimkH0(X,𝒪X((i+mr)KX))=Pi(m).\dim_{k}H^{0}(X,\mathcal{O}_{X}(-(i+mr)K_{X}))=P_{i}(m).

Moreover, setting V=(rKX)dV=(-rK_{X})^{d}, each PiP_{i} has the form

Pi(m)=Vd!md+Qi(m)P_{i}(m)=\frac{V}{d!}m^{d}+Q_{i}(m)

for polynomials QiQ_{i} of degree at most d1d-1. In this situation, the existence of an integer nn as required is guaranteed by 4.6 stated and proved below.

Assume, for the sake of contradiction, that αF(X)>12+ε\alpha_{F}(X)>\frac{1}{2}+\varepsilon for some small ε>0\varepsilon>0. Then, note that by 3.4, for all e0e\gg 0, we have Ie(mrKX)=0I_{e}(-mrK_{X})=0 for all mpe12r+εr(pe1)m\leq\frac{p^{e}-1}{2r}+\frac{\varepsilon}{r}(p^{e}-1).

Now, for e0e\gg 0, we can find an integer mm satisfying the following properties:

  • mr<pe12mr<\frac{p^{e}-1}{2},

  • pe1mr+2r<pe12+ε(pe1)p^{e}-1-mr+2r<\frac{p^{e}-1}{2}+\varepsilon(p^{e}-1), and,

  • npe12mrn\leq p^{e}-1-2mr.

This is equivalent to finding an integer mm such that

pe12r+2εr(pe1)mpe12rn2r,\frac{p^{e}-1}{2r}+2-\frac{\varepsilon}{r}(p^{e}-1)\leq m\leq\frac{p^{e}-1}{2r}-\frac{n}{2r},

which is possible since nn is fixed and ε2(pe1)\frac{\varepsilon}{2}(p^{e}-1)\to\infty as ee\to\infty. The third condition on mm guarantees that

dimkH0(X,mrKX)<dimkH0(X,(pe1mr)KX).\dim_{k}H^{0}(X,-mrK_{X})<\dim_{k}H^{0}(X,-(p^{e}-1-mr)K_{X}). (4.2)

The second condition guarantees that there exists a non-zero effective Weil divisor E0E\geq 0 that induces an injective map

𝒪X((pe1mr)KX)𝒪X(mrKX)\mathcal{O}_{X}(-(p^{e}-1-mr)K_{X})\hookrightarrow\mathcal{O}_{X}(-m^{\prime}rK_{X})

for some m<mpe12r+εr(pe1)m<m^{\prime}\leq\frac{p^{e}-1}{2r}+\frac{\varepsilon}{r}(p^{e}-1). Therefore, we know that Ie(mrKX)=Ie(mrKX)=0I_{e}(-mrK_{X})=I_{e}(-m^{\prime}rK_{X})=0 as noted above. By [LP23, Lemma 4.12], this implies that Ie((pe1mr)KX)=0I_{e}(-(p^{e}-1-mr)K_{X})=0 as well. Finally, note that 2.13 applied to D=mrKXD=-mrK_{X} tells us that

dimkH0(X,𝒪X(mrKX))Ie(mrKX)=dimkH0(X,𝒪X((pe1mr)KX))Ie((pe1mr)KX)\dim_{k}\frac{H^{0}(X,\mathcal{O}_{X}(-mrK_{X}))}{I_{e}(-mrK_{X})}=\dim_{k}\frac{H^{0}(X,\mathcal{O}_{X}(-(p^{e}-1-mr)K_{X}))}{I_{e}(-(p^{e}-1-mr)K_{X})}

But since both the IeI_{e}-subspaces in the above equation are zero, this is in contradiction to Equation 4.2. This proves that αF(X)\alpha_{F}(X) is at most 1/2. ∎

The following lemma was used in the proof of 4.5.

Lemma 4.6.

For each 0i<r0\leq i<r, let PiP_{i} be a polynomial with real coefficients. Moreover, assume that all the PiP_{i}’s have the same positive degree and the same positive leading term. In other words, for each 0ir10\leq i\leq r-1, there exists a real polynomial QiQ_{i} of degree at most d1d-1 such that

Pi(m)=admd+Qi(m)P_{i}(m)=a_{d}m^{d}+Q_{i}(m)

for some real number ad>0a_{d}>0 (independent of ii). Then, there is an integer n0n\gg 0 such that for all integers m0m\gg 0, and all pairs 0i,jr10\leq i,j\leq r-1, we have

Pi(m)Pj(m+n).P_{i}(m)\leq P_{j}(m+n).
Proof.

Since each QiQ_{i} is a real polynomial of degree at most d1d-1, we may find positive constants C1,C2C_{1},C_{2} such that

C2md1Qi(m)C1md1-C_{2}m^{d-1}\leq Q_{i}(m)\leq C_{1}m^{d-1}

for each m0m\geq 0 and each 0ir10\leq i\leq r-1. Now, it is sufficient to find a constant nn such that

admd+C1md1ad(m+n)dC2(m+n)d1a_{d}m^{d}+C_{1}m^{d-1}\leq a_{d}(m+n)^{d}-C_{2}(m+n)^{d-1} (4.3)

for all m0m\gg 0. For this, expanding (m+n)d(m+n)^{d} using the binomial theorem, Equation 4.3 is equivalent to

C1md1ad(dnmd1++dmnd1+nd)C2(m+n)d1C_{1}m^{d-1}\leq a_{d}(dnm^{d-1}+\dots+dmn^{d-1}+n^{d})-C_{2}(m+n)^{d-1}

First, we note that we may choose n0n\gg 0 such that addn2\frac{a_{d}dn}{2} is larger than both C1C_{1} and C2C_{2}. This implies that addn2md1>C1md1\frac{a_{d}dn}{2}m^{d-1}>C_{1}m^{d-1} for all m0m\geq 0. Furthermore, having chosen the n0n\gg 0 as before, we have addn2md1C2(m+n)d1\frac{a_{d}dn}{2}m^{d-1}\geq C_{2}(m+n)^{d-1} for all m0m\gg 0. This proves the lemma. ∎

For the rest of this section, we assume that H0(X,𝒪X)=kH^{0}(X,\mathcal{O}_{X})=k, i.e., that XX 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 αF\alpha_{F}-invariant to the FF-signature than the formula in 3.10:

Theorem 4.7.

Let XX be a dd-dimensional globally FF-regular \mathbb{Q}-Fano variety over kk. Assume that H0(X,𝒪X)=kH^{0}(X,\mathcal{O}_{X})=k and that dd is positive. Set α=αF(X)\alpha=\alpha_{F}(X). Then, we have the following inequalities relating the FF-signature and the αF\alpha_{F}-invariant:

2αd+1vol(X)(d+1)!𝓈(X)2((12)d+1(12α)d+1)vol(X)(d+1)!\frac{2\,\alpha^{d+1}\,\mathrm{vol}(X)}{(d+1)!}\leq\mathscr{s}(X)\leq\frac{2\,\big{(}(\frac{1}{2})^{d+1}-(\frac{1}{2}-\alpha)^{d+1}\big{)}\,\mathrm{vol}(X)}{(d+1)!} (4.4)

Here, vol(X)\mathrm{vol}(X) denotes the volume of the \mathbb{Q}-Cartier divisor KX-K_{X}.

Corollary 4.8.

Let XX be a globally FF-regular \mathbb{Q}-Fano variety of dimension d>0d>0. Then,

  1. (1)

    We have

    𝓈(X)vol(X)2d(d+1)!.\mathscr{s}(X)\leq\frac{\mathrm{vol}(X)}{2^{d}(d+1)!}.
  2. (2)

    Moreover, αF(X)\alpha_{F}(X) is equal to 1/21/2 if and only if the value of the FF-signature 𝓈(X)\mathscr{s}(X) is equal to

    vol(X)2d(d+1)!.\frac{\mathrm{vol}(X)}{2^{d}(d+1)!}.

For \mathbb{Q}-Fano varieties, the FF-signature has a more refined formula than 2.11, which we prove next.

Proposition 4.9.

Let XX be a globally FF-regular \mathbb{Q}-Fano variety over kk. Assume that H0(X,𝒪X)=kH^{0}(X,\mathcal{O}_{X})=k. Let rr be an integer such that rKXrK_{X} is Cartier. Then, the FF-signature of XX can be computed as

𝓈(X)=lime2rm=0pe12rdimkH0(mrKX)Ie(mrKX)pe(dim(X)+1)\mathscr{s}(X)=\lim_{e\to\infty}\frac{2r\,\sum\limits_{m=0}^{\lfloor\frac{p^{e}-1}{2r}\rfloor}\dim_{k}\frac{H^{0}(-mrK_{X})}{I_{e}(-mrK_{X})}}{p^{e(\dim(X)+1)}}
Lemma 4.10.

Let XX be a globally FF-regular \mathbb{Q}-Fano variety that is geometrically connected over kk. Then, for any r>0r>0, we have H0(mrKX)=Ie(mrKX)H^{0}(-mrK_{X})=I_{e}(-mrK_{X}) whenever m>pe1rm>\frac{p^{e}-1}{r}.

Proof.

We will prove this lemma in two different ways, since both ideas may be useful in other situations.

Proof 1:

Let m>pe1rm>\frac{p^{e}-1}{r}. By definition of the subspace IeI_{e} (2.8), it is sufficient to show that there are no non-zero maps ϕ:Fe𝒪X(mrKX)𝒪X\phi:F^{e}_{*}\mathcal{O}_{X}(-mrK_{X})\to\mathcal{O}_{X}. We have

Hom𝒪X(Fe𝒪X(mrKX,𝒪X)H0(X,(1pe+mr)KX).\operatorname{Hom}_{\mathcal{O}_{X}}(F^{e}_{*}\mathcal{O}_{X}(-mrK_{X},\mathcal{O}_{X})\cong H^{0}(X,(1-p^{e}+mr)K_{X}).

By assumption, we have 1pe+mr>01-p^{e}+mr>0. Since rKX-rK_{X} is ample, this means that H0(X,(1pe+mr)KX)=0H^{0}(X,(1-p^{e}+mr)K_{X})=0. Hence, there are no non-zero maps ϕ:Fe𝒪X(mrKX)𝒪X\phi:F^{e}_{*}\mathcal{O}_{X}(-mrK_{X})\to\mathcal{O}_{X}, which proves the lemma.

Proof 2:

Let m>pe1rm>\frac{p^{e}-1}{r} and suppose that there is a non-zero global section fH0(mrKX)f\in H^{0}(-mrK_{X}) that is not in Ie(mrKX)I_{e}(-mrK_{X}). Then, by definition of IeI_{e}, we have a map

ϕHom𝒪X(Fe𝒪X(mrKX),𝒪X)\phi\in\operatorname{Hom}_{\mathcal{O}_{X}}(F^{e}_{*}\mathcal{O}_{X}(-mrK_{X}),\mathcal{O}_{X})

such that ϕ(Fef)=1\phi(F^{e}_{*}f)=1. Thus, we have the splitting

𝒪XFe𝒪X(mrKX)𝒪X\mathcal{O}_{X}\hookrightarrow F^{e}_{*}\mathcal{O}_{X}(-mrK_{X})\twoheadrightarrow\mathcal{O}_{X} (4.5)

where the first map is got by sending 1Fef1\to F^{e}_{*}f and the second map is ϕ\phi. Twisting equation 4.5 by 𝒪X(KX)\mathcal{O}_{X}(K_{X}) and reflexifying, we obtain a splitting:

𝒪X(KX)Fe𝒪X((mrpe)KX)𝒪X(KX)\mathcal{O}_{X}(K_{X})\hookrightarrow F^{e}_{*}\mathcal{O}_{X}(-(mr-p^{e})K_{X})\twoheadrightarrow\mathcal{O}_{X}(K_{X}) (4.6)

By assumption, mrpemr-p^{e} is non-negative. Hence, Hd(X,Fe𝒪X((mrpe)KX))=0H^{d}(X,F^{e}_{*}\mathcal{O}_{X}(-(mr-p^{e})K_{X}))=0 by 2.22, in turn implying that Hd(X,𝒪X(KX))=0H^{d}(X,\mathcal{O}_{X}(K_{X}))=0 (using the splitting in equation 4.6). This is a contradiction, since 𝒪X(KX)\mathcal{O}_{X}(K_{X}) is the canonical sheaf of XX. This completes the proof of 4.10. ∎

Proof of 4.9.

Let SS denote the section ring of XX with respect to rKX-rK_{X}. Fix an e>0e>0 and let aea_{e} denote the free rank of FeSF^{e}_{*}S as an SS-module (2.5). Recall that by 2.11, we have

ae=m=0dimkH0(mrKX)Ie(mrKX)a_{e}=\sum_{m=0}^{\infty}\dim_{k}\frac{H^{0}(-mrK_{X})}{I_{e}(-mrK_{X})} (4.7)

so that

𝓈(X)=rlimeaepe(dim(X)+1).\mathscr{s}(X)=r\,\lim_{e\to\infty}\frac{a_{e}}{p^{e(\dim(X)+1)}}.

Then, 4.10 shows that the terms of the sum in Equation 4.7 are zero for m>pe1rm>\frac{p^{e}-1}{r}. Furthermore, using 2.13, we have

m=pe12rpe1rdimkH0(mrKX)Ie(mrKX)=m=pe12rpe1rdimkH0((pe1mr)KX)Ie((pe1mr)KX).\sum_{m=\lceil\frac{p^{e}-1}{2r}\rceil}^{\lfloor\frac{p^{e}-1}{r}\rfloor}\dim_{k}\frac{H^{0}(-mrK_{X})}{I_{e}(-mrK_{X})}=\sum_{m=\lceil\frac{p^{e}-1}{2r}\rceil}^{\lfloor\frac{p^{e}-1}{r}\rfloor}\dim_{k}\frac{H^{0}(-(p^{e}-1-mr)K_{X})}{I_{e}(-(p^{e}-1-mr)K_{X})}.

Let aa be an integer between 0 and rr such that pe1amodrp^{e}-1\equiv a\mod r. Hence, we have

m=pe12rpe1rdimkH0((pe1mr)KX)Ie((pe1mr)KX)=m=0pe12rdimkH0((a+mr)KX)Ie((a+mr)KX).\sum_{m=\lceil\frac{p^{e}-1}{2r}\rceil}^{\lfloor\frac{p^{e}-1}{r}\rfloor}\dim_{k}\frac{H^{0}(-(p^{e}-1-mr)K_{X})}{I_{e}(-(p^{e}-1-mr)K_{X})}=\sum_{m=0}^{\lfloor\frac{p^{e}-1}{2r}\rfloor}\dim_{k}\frac{H^{0}(-(a+mr)K_{X})}{I_{e}(-(a+mr)K_{X})}.

Thus, we have

ae=m=0pe12rdimkH0(mrKX)Ie(mrKX)+m=0pe12rdimkH0((a+mr)KX)Ie((a+mr)KX).a_{e}=\sum_{m=0}^{\lfloor\frac{p^{e}-1}{2r}\rfloor}\dim_{k}\frac{H^{0}(-mrK_{X})}{I_{e}(-mrK_{X})}+\sum_{m=0}^{\lfloor\frac{p^{e}-1}{2r}\rfloor}\dim_{k}\frac{H^{0}(-(a+mr)K_{X})}{I_{e}(-(a+mr)K_{X})}.

Moreover, using 2.14, we see that there is a constant C>0C>0 such that

|H0((a+mr)KX)Ie((a+mr)KX)H0(mrKX)Ie(mrKX)|<Cpe(dim(X)1).\left|\frac{H^{0}(-(a+mr)K_{X})}{I_{e}(-(a+mr)K_{X})}-\frac{H^{0}(-mrK_{X})}{I_{e}(-mrK_{X})}\right|<Cp^{e(\dim(X)-1)}.

Thus, we have that

|ae2m=0pe12rdimkH0(mrKX)Ie(mrKX)|<Cpedim(X).\left|a_{e}-2\sum\limits_{m=0}^{\lfloor\frac{p^{e}-1}{2r}\rfloor}\dim_{k}\frac{H^{0}(-mrK_{X})}{I_{e}(-mrK_{X})}\right|<Cp^{e\dim(X)}.

The proof is now complete since the right hand side limits to zero when divided by pe(dim(X)+1)p^{e(\dim(X)+1)} and as ee\to\infty. ∎

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 FF-signature of S=S(X,rKX)S=S(X,-rK_{X}). ∎

Proof of 4.8.

Part (1) follows immediately from the right-hand inequality in 4.7, since we know that αF(X)12\alpha_{F}(X)\leq\frac{1}{2} by 4.5. We also see that if αF(X)<12\alpha_{F}(X)<\frac{1}{2}, we must have

𝓈(X)<vol(X)2d(d+1)!.\mathscr{s}(X)<\frac{\mathrm{vol}(X)}{2^{d}(d+1)!}.

Thus, Part (2) also follows from Equation 4.4 once we note that when αF(X)=12\alpha_{F}(X)=\frac{1}{2}, both sides of the inequality in Equation 4.4 are equal to vol(X)2d(d+1)!\frac{\mathrm{vol}(X)}{2^{d}(d+1)!}. ∎

Remark 4.11.

Let XX be a \mathbb{Q}-Fano variety over \mathbb{C}. This means that XX is a normal variety, KX-K_{X} is \mathbb{Q}-Cartier and ample and XX has only klt singularities. Let rr be such that rKX-rK_{X} is Cartier, and S=S(X,rKX)S=S(X,-rK_{X}). Then, for any effective \mathbb{Q}-divisor Δ\Delta on XX with ΔrKX\Delta\sim_{\mathbb{Q}}-rK_{X}, we have

lct𝔪(S,ΔS)=min{lct(X,Δ),1r}\text{lct}_{\mathfrak{m}}(S,\Delta_{S})=\min\,\{\,\text{lct}(X,\Delta),\,\frac{1}{r}\,\}

where lct denotes the log canonical threshold and ΔS\Delta_{S} denotes the cone over Δ\Delta. This follows from [Kol13, Lemma 3.1]. Thus, if we let

α~(X)=rinf{lct𝔪(S,ΔS)|Δ0is a -divisor on X such that ΔKX},\tilde{\alpha}(X)=r\,\inf\,\{\text{lct}_{\mathfrak{m}}(S,\Delta_{S})\,|\,\Delta\geq 0\,\text{is a $\mathbb{Q}$-divisor on $X$ such that $\Delta\sim_{\mathbb{Q}}-K_{X}$}\,\},

then, we have that

α~(X)=min{α(X), 1}.\tilde{\alpha}(X)=\min\,\{\alpha(X),\,1\}.

Therefore, for any \mathbb{Q}-Fano variety with α(X)1\alpha(X)\leq 1, the αF\alpha_{F}-invariant from 4.2 is a “Frobenius analog" of the complex α\alpha-invariant.

The αF\alpha_{F}-invariant of toric Fano varieties

Let kk denote an algebraically closed field of prime characteristic p>0p>0. Fix a lattice NdN\cong\mathbb{Z}^{d} and let MM be the dual lattice (where dd is some positive integer).

Theorem 4.12.

Let XpX_{p} be a \mathbb{Q}-Fano toric variety over kk defined by a fan \mathcal{F} in NN. Let XX_{\mathbb{C}} be the corresponding complex toric variety (which is also automatically \mathbb{Q}-Fano). Then, we have

αF(Xp)=α(X).\alpha_{F}(X_{p})=\alpha(X_{\mathbb{C}}).
Proof.

Let v1,,vnv_{1},\dots,v_{n} denote the primitive generators for the one dimensional cones in \mathcal{F} and write KX=ibivi-K_{X}=\sum_{i}b_{i}v_{i} for rational numbers bib_{i}.

First we choose an r>0r>0 such that rbirb_{i}\in\mathbb{Z} for each ii and the section ring S(X,rKX)S(X,-rK_{X}) is generated in degree one. Let PMP\subset M denote the polytope associated to rKX-rK_{X}, and defined by:

P={uM=M|u,vibifor all 1in}.P=\{u\in M_{\mathbb{R}}=M\otimes_{\mathbb{Z}}\mathbb{R}\,|\,\langle u,v_{i}\rangle\geq-b_{i}\,\,\text{for all }1\leq i\leq n\}.

Since we are assuming that S(X,rKX)S(X,-rK_{X}) is generated in degree 11, the vertices of PP are lattice points of MM. For any uPMu\in P\cap M, let DuD_{u} be the corresponding effective divisor in the linear system |rKX||-rK_{X}|. By [BJ20, Corollary 7.16], we have that

α(X)=minuPMrlct(X,Du)\alpha(X_{\mathbb{C}})=\min_{u\in P\cap M}\,r\,\text{lct}(X_{\mathbb{C}},D_{u}) (4.8)

where lct(X,)\text{lct}(X_{\mathbb{C}},-) denotes the log canonical threshold of a divisor on XX_{\mathbb{C}}. Note that since the vertices of PP are lattice points, just the vertices are sufficient to compute α(X)\alpha(X_{\mathbb{C}}).

Let P~\tilde{P} denote the polytope P×{1}M×P\times\{1\}\subset M\times\mathbb{Z}. Then, the section ring S(X,rKX)S(X,-rK_{X}) is the semigroup ring associated to the cone over P~\tilde{P} in (M×)(M\times\mathbb{Z})\otimes_{\mathbb{Z}}\mathbb{R}. Note that SS is \mathbb{Q}-Gorenstein. Therefore, by [Bli04, Theorem 3], we see that for any u~P~(M×)\tilde{u}\in\tilde{P}\cap(M\times\mathbb{Z}), we have

fpt𝔪(S(Xp,KXp),Du~)=lct𝔪(S(X,KX),Du~).\text{fpt}_{\mathfrak{m}}(S(X_{p},-K_{X_{p}}),D_{\tilde{u}})=\text{lct}_{\mathfrak{m}}(S(X_{\mathbb{C}},-K_{X_{\mathbb{C}}}),D_{\tilde{u}}). (4.9)

Next, note that since XX is a normal toric variety, it is automatically globally FF-regular. Now we prove that the αF\alpha_{F}-invariant of XpX_{p} can also be computed by only considering the torus invariant divisors. To see this, let S=S(Xp,rKXp)S=S(X_{p},-rK_{X_{p}}) and let fSf\in S 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 μ=(μ1,,μd+1)\mu=(\mu_{1},\dots,\mu_{d+1}) with μi>0\mu_{i}\in\mathbb{Z}_{>0} such that in>μ(f)=in>(f)\text{in}_{>_{\mu}}(f)=\text{in}_{>}(f). Here >μ>_{\mu} denotes the weight monomial order with respect to μ\mu and >> denotes the graded lexicographic monomial order on SS. Then, we have a flat degeneration of ff to its initial term. In other words, if f=uβuχuf=\sum_{u}\beta_{u}\chi^{u} for monomials χuS\chi^{u}\in S, then setting w=max{μ,u|βu0}w=\max\{\langle\mu,u\rangle\,|\,\beta_{u}\neq 0\}, the element

f~=twuβutμ,uχuS[t]\tilde{f}=t^{w}\,\sum_{u}\beta_{u}t^{-\langle\mu,u\rangle}\chi^{u}\in S[t]

satisfies the following properties:

  • Viewing S[t]S[t] as a k[t]k[t]-algebra, the ring S[t]/(f~)S[t]/(\tilde{f}) is a flat k[t]k[t]-module.

  • The image of f~\tilde{f} modulo tt is equal to in>(f)\text{in}_{>}(f), the initial term of ff with respect to the graded lex monomial order on SS.

  • For any point 0λk0\neq\lambda\in k, the image fλf_{\lambda} of f~\tilde{f} in S[t]/(tλ)S[t]/(t-\lambda) satisfies

    fpt𝔪(S,fλ)=fpt𝔪(S,f).\text{fpt}_{\mathfrak{m}}(S,f_{\lambda})=\text{fpt}_{\mathfrak{m}}(S,f).

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 ff of SS, we have

fpt𝔪(S,f)fpt𝔪(S,in>(f)).\text{fpt}_{\mathfrak{m}}(S,f)\geq\text{fpt}_{\mathfrak{m}}(S,\text{in}_{>}(f)).

Assuming this lemma for a moment, we see that

αF(Xp)=infu~P~(M×)fpt𝔪(S,Du~).\alpha_{F}(X_{p})=\inf_{\tilde{u}\in\tilde{P}\cap(M\times\mathbb{Z})}\text{fpt}_{\mathfrak{m}}(S,D_{\tilde{u}}).

Furthermore, by Equation 4.9, we have

αF(Xp)=infu~P~(M×)lct𝔪(S(X,rKX),Du~).\alpha_{F}(X_{p})=\inf_{\tilde{u}\in\tilde{P}\cap(M\times\mathbb{Z})}\text{lct}_{\mathfrak{m}}(S(X_{\mathbb{C}},-rK_{X_{\mathbb{C}}}),D_{\tilde{u}}). (4.10)

Since by 4.5 we have αF(Xp)1/2\alpha_{F}(X_{p})\leq 1/2, we must have lct(S(X,rKX),Du~)<1nr\text{lct}(S(X_{\mathbb{C}},-rK_{X_{\mathbb{C}}}),D_{\tilde{u}})<\frac{1}{nr} for some u~=(u,n)P~(M×)\tilde{u}=(u,n)\in\tilde{P}\cap(M\times\mathbb{Z}). Note that DuD_{u} corresponds to a torus-invariant divisor on XX_{\mathbb{C}} linearly equivalent to nrKX-nrK_{X_{\mathbb{C}}}. Therefore, by 4.11, we have

lct(X,Du)=lct𝔪(S(X,rKX),Du~)\text{lct}(X_{\mathbb{C}},D_{u})=\text{lct}_{\mathfrak{m}}(S(X_{\mathbb{C}},-rK_{X_{\mathbb{C}}}),D_{\tilde{u}})

for any u~=(u,n)\tilde{u}=(u,n) such that lct(S(X,rKX),Du~)<1nr\text{lct}(S(X_{\mathbb{C}},-rK_{X_{\mathbb{C}}}),D_{\tilde{u}})<\frac{1}{nr}. Putting this together with Equation 4.10 and Equation 4.8, we get that

αF(Xp)=α(X)\alpha_{F}(X_{p})=\alpha(X_{\mathbb{C}})

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 ape1\frac{a}{p^{e}-1} such that

ape1<fpt𝔪(S,in>(f)),\frac{a}{p^{e}-1}<\text{fpt}_{\mathfrak{m}}(S,\text{in}_{>}(f)),

the map SFeSS\to F^{e}_{*}S sending 11 to FefaF^{e}_{*}f^{a} splits. Equivalently, for all such ape1\frac{a}{p^{e}-1}, it suffices to show that faIe(S)f^{a}\notin I_{e}(S). Since the pair (S,in>(f)ape1)(S,\text{in}_{>}(f)^{\frac{a}{p^{e}-1}}) is strongly FF-regular, in particular it is sharply FF-split. By 3.6 again, we know that in>(fa)=(in>(f))aIe(S)\text{in}_{>}(f^{a})=(\text{in}_{>}(f))^{a}\notin I_{e}(S). Now, since SS is a toric ring, Ie(S)I_{e}(S) is a monomial ideal of SS. Therefore, if in>(fa)Ie(S)\text{in}_{>}(f^{a})\notin I_{e}(S), we also have faIe(S)f^{a}\notin I_{e}(S) as required. ∎

Remark 4.14.

Combining 4.12 with 4.5, we recover the well-known fact that the α\alpha-invariant of a toric Fano variety is at most 1/2 (see [LZ22, Corollary 3.6]).

Remark 4.15.

The results of this section also extend naturally to the case of globally FF-regular log-Fano pairs (X,Δ)(X,\Delta) such that (pe1)Δ(p^{e}-1)\Delta 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 αF\alpha_{F}-invariant in geometric families, analogous to the case of the FF-signature which was proved in [CRST21] and the case of complex α\alpha-invariant discussed in [BL22]. Throughout this section, we fix kk to be an algebraically closed field of characteristic p>0p>0.

5.1. Weak semicontinuity

First, we will prove a result for a family of arbitrary globally FF-regular varieties.

Notation 5.1.

Recall that the perfection of a field KK (of positive characteristic) is the union

K:=e=1K1/peK^{\infty}:=\bigcup_{e=1}^{\infty}K^{1/p^{e}}

of all the pep^{e}-th roots of elements of KK. For a map f:𝒳Yf:\mathcal{X}\to Y of varieties over kk, and a point yYy\in Y (not necessarily closed), we denote the fiber of ff over yy as 𝒳y\mathcal{X}_{y} and denote the perfectified fiber over yy as

𝒳y:=𝒳×YSpec(κ(y)).\mathcal{X}_{y^{\infty}}:=\mathcal{X}\times_{Y}\operatorname{\text{Spec}}(\kappa(y)^{\infty}).

Similar for a coherent sheaf \mathcal{F} on 𝒳\mathcal{X}, we have the fiber y\mathcal{F}_{y} and the perfectified fiber y\mathcal{F}_{y^{\infty}}. Furthermore , if Y=Spec(A)Y=\operatorname{\text{Spec}}(A) and SS is a finitely generated AA-algebra, then for any yYy\in Y, we denote the perfectified fiber SAκ(y)S\otimes_{A}\kappa(y)^{\infty} by SyS_{y^{\infty}}.

Notation 5.2.

By a family of globally FF-regular varieties, we mean that

  1. (1)

    We have a flat and projective morphism f:𝒳Yf:\mathcal{X}\to Y where 𝒳\mathcal{X} is normal and YY is smooth over kk.

  2. (2)

    We assume additionally that ff has connected fibers (i.e., f𝒪𝒳=𝒪Yf_{*}\mathcal{O}_{\mathcal{X}}=\mathcal{O}_{Y}).

  3. (3)

    For each point yYy\in Y (not necessarily closed), the fiber XyX_{y^{\infty}} (5.1) is globally FF-regular (2.16).

Recall that as in 3.2, we can consider the αF\alpha_{F}-invariant of a pair (X,L)(X,L) where XX is a globally FF-regular projective variety (over a perfect field) and LL is an ample line bundle over XX.

Theorem 5.3.

Let f:(𝒳,)Yf:(\mathcal{X},\mathcal{L})\to Y be a flat family of globally FF-regular varieties, where \mathcal{L} is an ample line bundle over 𝒳\mathcal{X}. Let KK denote the fraction field of YY and αgen\alpha_{\text{gen}} denote the αF\alpha_{F}-invariant of (XK,|XK)(X_{K^{\infty}},\mathcal{L}|_{X_{K^{\infty}}}), the perfectified generic fiber. Then, for each real number 0<α<αgen0<\alpha<\alpha_{\text{gen}}, there exists a dense open subset UαYU_{\alpha}\subset Y such that

αF(Xy,|Xy)>αfor every point y in Uα.\alpha_{F}(X_{y^{\infty}},\mathcal{L}|_{X_{y^{\infty}}})>\alpha\quad\text{for every point $y$ in $U_{\alpha}$.}

Recall that in 3.8, we defined the sequence (αe)e1(\alpha_{e})_{e\geq 1} that converges to the αF\alpha_{F}-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 XX be a projective globally FF-regular variety over kk and LL be an ample invertible sheaf over XX. Suppose we have integers N1N\geq 1 and e>0e>0 such that the sheaf

om𝒪X((Fe(Lm),LN)\mathscr{H}om_{\mathcal{O}_{X}}\big{(}\,(F^{e}_{*}(L^{m}),L^{N}\,\big{)}

is generically globally generated for each 0mpe10\leq m\leq p^{e}-1. Then, if SS denotes the section ring S(X,L)S(X,L), we have

|αF(S)αe(S)|Npe1.\left|\alpha_{F}(S)-\alpha_{e}(S)\right|\leq\frac{N}{p^{e}-1}.
Proof.

Set d=dim(X)d=\dim(X). First, we claim that for each 0mpe10\leq m\leq p^{e}-1, we can find an injective map of 𝒪X\mathcal{O}_{X}-modules

ιe,m:Fe(Lm)(LN)ped.\iota_{e,m}:F^{e}_{*}(L^{m})\hookrightarrow(L^{N})^{\oplus p^{ed}}. (5.1)

To see this, let η\eta denote the generic point of XX and consider the following restriction map to the generic stalk:

Hom𝒪X(Fe(Lm),LN)om𝒪X(Fe(Lm),LN)ηHom𝒪X,η(Fe(Lηm),LηN).\operatorname{Hom}_{\mathcal{O}_{X}}(F^{e}_{*}(L^{m}),L^{N})\to\mathscr{H}om_{\mathcal{O}_{X}}(F^{e}_{*}(L^{m}),L^{N})_{\eta}\cong\operatorname{Hom}_{\mathcal{O}_{X,\eta}}(F^{e}_{*}(L^{m}_{\eta}),L^{N}_{\eta}). (5.2)

The assumption that om𝒪X((Fe(Lm),LN)\mathscr{H}om_{\mathcal{O}_{X}}\big{(}\,(F^{e}_{*}(L^{m}),L^{N}\,\big{)} is generically globally generated means that the image of the map in Equation 5.2 generates Hom𝒪X,η(Fe(Lηm),LηN)\operatorname{Hom}_{\mathcal{O}_{X,\eta}}(F^{e}_{*}(L^{m}_{\eta}),L^{N}_{\eta}) as an 𝒪X,η\mathcal{O}_{X,\eta}-module. Recall that 𝒪X,η\mathcal{O}_{X,\eta} is just the fraction field of XX. Since Fe(Lηm)F^{e}_{*}(L^{m}_{\eta}) is a free 𝒪X,η\mathcal{O}_{X,\eta}-vector space of rank pedp^{ed}, we can choose pedp^{ed} maps ϕ1,,ϕped\phi_{1},\dots,\phi_{p^{ed}} in Hom𝒪X(Fe(Lm),LN)\operatorname{Hom}_{\mathcal{O}_{X}}(F^{e}_{*}(L^{m}),L^{N}) such that their images under the map in Equation 5.2 forms a basis of Hom𝒪X,η(Fe(Lηm),LηN)\operatorname{Hom}_{\mathcal{O}_{X,\eta}}(F^{e}_{*}(L^{m}_{\eta}),L^{N}_{\eta}) over 𝒪X,η\mathcal{O}_{X,\eta}. Thus, defining ιe,m\iota_{e,m} to be the product map

ιe,m:=ϕ1××ϕped:Fe(Lm)(LN)ped,\iota_{e,m}:=\phi_{1}\times\dots\times\phi_{p^{ed}}:F^{e}_{*}(L^{m})\rightarrow(L^{N})^{\oplus p^{ed}},

we see that ιe,m\iota_{e,m} is generically an isomorphism by construction. Furthermore, since Fe(Lm)F^{e}_{*}(L^{m}) is a torsion-free sheaf of rank pedp^{ed} over 𝒪X\mathcal{O}_{X}, ιe,m\iota_{e,m} 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 ιe,m\iota_{e,m} for each 0mpe10\leq m\leq p^{e}-1 as in Equation 5.1. Then, by taking section modules with respect to LL (2.3), we get a corresponding map of graded SS-modules

ιe,m:(FeS)mmodpeS(N)ped\iota_{e,m}:(F^{e}_{*}S)_{m\bmod p^{e}}\hookrightarrow S(N)^{\oplus p^{ed}}

and let

ιe:FeSS(N)pe(d+1)\iota_{e}:F^{e}_{*}S\hookrightarrow S(N)^{\oplus p^{e(d+1)}}

denote the direct sum of the ιe,m\iota_{e,m}’s. Here, for any mm\in\mathbb{Z}, (FeS)mmodpe(F^{e}_{*}S)_{m\bmod p^{e}} denotes the SS-module

(FeS)mmodpe:=jFe(Sm+jpe),(F^{e}_{*}S)_{m\bmod p^{e}}:=\bigoplus_{j\in\mathbb{Z}}F^{e}_{*}(S_{m+jp^{e}}),

and we naturally have an \mathbb{N}-graded SS-module decomposition

FeS0mpe1(FeS)mmodpe.F^{e}_{*}S\cong\bigoplus_{0\leq m\leq p^{e}-1}(F^{e}_{*}S)_{m\bmod p^{e}}.

Note that ιe\iota_{e} is injective because the ιe,m\iota_{e,m}’s were injective. Furthermore, the key property of ιe\iota_{e} that we will use is the following: for each non-zero homogeneous element ff of degree mm in SS, ιe(Fe(f))\iota_{e}(F^{e}_{*}(f)) is a non-zero homogeneous element of degree

mpe+N.\lfloor\frac{m}{p^{e}}\rfloor+N. (5.3)

Thus, we may use this map as a “degree-lowering operator".

From 3.8, recall that αe:=αe(S)=mepe\alpha_{e}:=\alpha_{e}(S)=\frac{m_{e}}{p^{e}} where mem_{e} is defined to be the number max{m|Ie(m)=0}\max\{m\,|\,I_{e}(m)=0\}. The condition that Ie(m)=0I_{e}(m)=0 means that for all non-zero elements fSmf\in S_{m}, there exists a splitting FeSSF^{e}_{*}S\to S that sends FefF^{e}_{*}f to 11.

Claim: For any 1\ell\geq 1, let e=ee^{\prime}=\ell e, and mm be an integer such that m(αet=11Npte)pem\leq(\alpha_{e}-\sum_{t=1}^{\ell-1}\frac{N}{p^{te}})p^{\ell e}. Then Ie(m)=0I_{e^{\prime}}(m)=0.

We will prove the claim by induction on \ell. If =1\ell=1, note that the sum in the claim is empty, and hence we have mme=αepem\leq m_{e}=\alpha_{e}p^{e}. In this case, Ie(m)=0I_{e}(m)=0 by the definition of mem_{e}.

Now let >1\ell>1 and ff be any non-zero element of SmS_{m}. We need to show that FefF^{e}_{*}f splits from FeSF^{e}_{*}S. To see this, note that by Equation 5.3, ιe(Fe(f))\iota_{e}(F^{e}_{*}(f)) is a non-zero element of degree at most

1pe(αet=11Npte)pe+N=mep(2)eNp(2)eN+N=(αet=12Npte)p(1)e.\lfloor\frac{1}{p^{e}}\,(\alpha_{e}-\sum_{t=1}^{\ell-1}\frac{N}{p^{te}})\,p^{\ell e}\rfloor+N=m_{e}\,p^{(\ell-2)e}-N\,p^{(\ell-2)e}-\dots-N+N=(\alpha_{e}-\sum_{t=1}^{\ell-2}\frac{N}{p^{te}})\,p^{(\ell-1)e}.

Thus, the inductive hypothesis applies to ιe(Fe(f))\iota_{e}(F^{e}_{*}(f)), implying that ιe(Fe(f))\iota_{e}(F^{e}_{*}(f)) is not contained in IeeI_{e^{\prime}-e}. Let φ:FeeSS\varphi:F_{*}^{e^{\prime}-e}S\to S denote a splitting of ιe(Fe(f))\iota_{e}(F^{e}_{*}(f)). Then, we see that (forgetting the degrees) φFeeιe:FeSS\varphi\circ F_{*}^{e^{\prime}-e}\iota_{e}:F_{*}^{e^{\prime}}S\to S defines a splitting of Fe(f)F_{*}^{e^{\prime}}(f), 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 1\ell\geq 1,

αe=mepeαet=11Npte.\alpha_{\ell e}=\frac{m_{\ell e}}{p^{\ell e}}\geq\alpha_{e}-\sum_{t=1}^{\ell-1}\frac{N}{p^{te}}.

Letting \ell\to\infty and using 3.8, we get

αe(S)αF(S)Npe1.\alpha_{e}(S)-\alpha_{F}(S)\leq\frac{N}{p^{e}-1}. (5.4)

Lastly, note that by 3.9, we already have

αF(S)αe(S)1pe,\alpha_{F}(S)-\alpha_{e}(S)\leq\frac{1}{p^{e}},

which, together with Equation 5.4 completes the proof of the theorem. ∎

Lemma 5.5.

Let XX be a globally FF-regular variety of dimension dd and LL be an ample and globally generated invertible sheaf. Suppose e0>0e_{0}>0 is such that (1pe)KX(1-p^{e})K_{X} is linearly equivalent to an effective Weil divisor for all ee0e\geq e_{0}. Then, for all ee0e\geq e_{0}, and each 0mpe10\leq m\leq p^{e}-1, the sheaf

om𝒪X(Fe(Lm),Ld+1)\mathscr{H}om_{\mathcal{O}_{X}}(F^{e}_{*}(L^{m}),L^{d+1})

is generically globally generated.

Proof.

First, by an application of Castelnuovo-Mumford regularity, we prove the following statement:

Claim: For all e1e\geq 1 and all 0npe10\leq n\leq p^{e}-1, the sheaf Fe(Ln+dpe)F^{e}_{*}(L^{n+dp^{e}}) is 0-regular with respect to LL, and hence globally generated.

To see this, we check that

Hi(X,Fe(Ln+dpe)Li)=Hi(X,Ln+(di)pe)=0for all i>0.H^{i}(X,F^{e}_{*}(L^{n+dp^{e}})\otimes L^{-i})=H^{i}(X,L^{n+(d-i)p^{e}})=0\quad\text{for all $i>0$.}

Here, we have used the projection formula to see that Fe(Ln+dpe)LiFe(Ln+(di)pe)F^{e}_{*}(L^{n+dp^{e}})\otimes L^{-i}\cong F^{e}_{*}(L^{n+(d-i)p^{e}}) and the cohomology vanishing follows from (2.21), since Ln+(di)peL^{n+(d-i)p^{e}} is nef for all idi\leq d. Thus, Fe(Ln+dpe)F^{e}_{*}(L^{n+dp^{e}}) is 0-regular with respect to LL, and hence is globally generated (see [Laz04, Theorem 1.8.5] for the details regarding Castelnuovo-Mumford regularity).

Next, for any ee0e\geq e_{0} and n0n\geq 0, write (1pe)KXDe(1-p^{e})K_{X}\sim D_{e} where DeD_{e} is an effective Weil-divisor. Note that it is always possible to find such an e0e_{0} thanks to [SS10, Theorem 4.3]. Let ϕe(Ln)\phi_{e}(L^{n}) denote the map obtained by twisting the defining map for DeD_{e} by LnL^{n} and pushing forward under FeF^{e}:

ϕe(Ln):Fe(Ln)Fe(𝒪X(De)Ln).\phi_{e}(L^{n}):F^{e}_{*}(L^{n})\hookrightarrow F^{e}_{*}(\mathcal{O}_{X}(D_{e})\otimes L^{n}).

Note that for any point xSupp(De)x\notin\text{Supp}(D_{e}), ϕe\phi_{e} restricts to an isomorphism in an open neighbourhood around xx. For any sheaf \mathcal{F}, let x\mathcal{F}_{x} denote the stalk of \mathcal{F} at xx.

Applying duality for the Frobenius map (Equation 2.2), we have (for any mm\in\mathbb{Z}):

om𝒪X(Fe(Lm),Ld+1)Fe(𝒪X(De)Ldpe+pem).\mathscr{H}om_{\mathcal{O}_{X}}(F^{e}_{*}(L^{m}),L^{d+1})\cong F^{e}_{*}(\mathcal{O}_{X}(D_{e})\otimes L^{dp^{e}+p^{e}-m}).

Therefore, for any ee0e\geq e_{0} and 0mpe10\leq m\leq p^{e}-1, set n=dpe+pemn=dp^{e}+p^{e}-m and consider the diagram

H0(X,Fe(Ln)){H^{0}(X,F^{e}_{*}(L^{n}))}H0(X,Fe(𝒪X(De)Ln)){H^{0}(X,F^{e}_{*}(\mathcal{O}_{X}(D_{e})\otimes L^{n}))}Fe(Ln)x{F^{e}_{*}(L^{n})_{x}}Fe(𝒪X(De)Ln)x{F^{e}_{*}(\mathcal{O}_{X}(D_{e})\otimes L^{n})_{x}}H0(ϕe(Ln))\scriptstyle{H^{0}(\phi_{e}(L^{n}))}ϕe(Ln)x\scriptstyle{\phi_{e}(L^{n})_{x}}\scriptstyle{\sim}

where xx is any point not contained in Supp(De)\text{Supp}(D_{e}). Since the horizontal arrows are injective and the bottom horizontal arrow is an isomorphism, any set of global sections generating Fe(Ln)xF^{e}_{*}(L^{n})_{x}, viewed as global sections of Fe(𝒪X(De)Ln)F^{e}_{*}(\mathcal{O}_{X}(D_{e})\otimes L^{n}), will also generate Fe(𝒪X(De)Ln)xF^{e}_{*}(\mathcal{O}_{X}(D_{e})\otimes L^{n})_{x}. Since by the claim above Fe(Ln)F^{e}_{*}(L^{n}) is globally generated, we have that om𝒪X(Fe(Lm),Ld+1)Fe(𝒪X(De)Ln)\mathscr{H}om_{\mathcal{O}_{X}}(F^{e}_{*}(L^{m}),L^{d+1})\cong F^{e}_{*}(\mathcal{O}_{X}(D_{e})\otimes L^{n}) is globally generated at any xSupp(De)x\notin\text{Supp}(D_{e}) (and hence generically globally generated). This completes the proof of the lemma. ∎

Lemma 5.6.

Let f:𝒳Yf:\mathcal{X}\to Y be a flat family of globally FF-regular varieties over kk where YY is regular. Let 𝒟\mathcal{D} be an integral Weil-divisor such that =𝒪𝒳(r𝒟)\mathcal{L}=\mathcal{O}_{\mathcal{X}}(r\mathcal{D}) is Cartier for some integer r1r\geq 1 . We also assume that \mathcal{L} is ample.

  1. (1)

    Then, for any integer m0m\geq 0, the sheaf f(m)f_{*}(\mathcal{L}^{m}) is locally free on YY and for any point yYy\in Y (not necessarily closed), the natural “restriction to the fiber" map

    f(m)𝒪Yκ(y)H0(𝒳y,m|𝒳y)f_{*}(\mathcal{L}^{m})\otimes_{\mathcal{O}_{Y}}\kappa(y)\to H^{0}(\mathcal{X}_{y},\mathcal{L}^{m}|_{\mathcal{X}_{y}})

    is an isomorphism. Moreover, for any yYy\in Y, there exists an affine open neighbourhood Spec(B)=UY\operatorname{\text{Spec}}(B)=U\subset Y of yy such that if the closure of yy is defined by a regular sequence b1,btBb_{1},\dots b_{t}\in B and for any m0m\geq 0, the natural map

    H0(f1(U),m)BB/𝔭BH0(𝒳B/𝔭B,m|𝒳B/𝔭B)H^{0}(f^{-1}(U),\mathcal{L}^{m})\otimes_{B}B/\mathfrak{p}B\to H^{0}(\mathcal{X}_{B/\mathfrak{p}B},\mathcal{L}^{m}|_{\mathcal{X}_{B/\mathfrak{p}B}})

    is an isomorphism where 𝔭\mathfrak{p} is the ideal (b1,,bt)(b_{1},\dots,b_{t}).

  2. (2)

    Suppose, in addition that 𝒳\mathcal{X} is a locally strongly FF-regular variety (2.16). Then, for any m0m\geq 0, setting :=𝒪𝒳(m𝒟)\mathcal{F}:=\mathcal{O}_{\mathcal{X}}(m\mathcal{D}), we have

    1. (a)

      \mathcal{F} is flat over YY.

    2. (b)

      For any yYy\in Y, the restriction y:=𝒪𝒳𝒪𝒳y\mathcal{F}_{y}:=\mathcal{F}\otimes_{\mathcal{O}_{\mathcal{X}}}\mathcal{O}_{\mathcal{X}_{y}} is reflexive.

    3. (c)

      ff_{*}\mathcal{F} is locally free on YY and for any yYy\in Y, the natural map

      f()𝒪Yκ(y)H0(𝒳y,y)f_{*}(\mathcal{F})\otimes_{\mathcal{O}_{Y}}\kappa(y)\to H^{0}(\mathcal{X}_{y},\mathcal{F}_{y})

      is an isomorphism.

    4. (d)

      Assume that Supp(𝒟)\text{Supp}(\mathcal{D}) does not contain any fiber of ff. Then,

      f(𝒪𝒳(m𝒟))𝒪Yκ(y)H0(𝒳y,𝒪𝒳y(m𝒟|𝒳y))f_{*}(\mathcal{O}_{\mathcal{X}}(m\mathcal{D}))\otimes_{\mathcal{O}_{Y}}\kappa(y)\to H^{0}(\mathcal{X}_{y},\mathcal{O}_{\mathcal{X}_{y}}(m\mathcal{D}|_{\mathcal{X}_{y}}))

      is an isomorphism as well. Here, we restrict the Weil-divisor m𝒟m\mathcal{D} to YY as explained in 2.23.

Proof.

Note that since m\mathcal{L}^{m} is an invertible sheaf on 𝒳\mathcal{X}, it is flat over YY for any mm.

  1. (1)

    Since 𝒪𝒳(m)\mathcal{O}_{\mathcal{X}}(m\mathcal{L}) is flat over YY, the claim follows from Grauert’s Theorem [Har77, Chapter III, Corollary 12.9] once we note that the function

    ydimκ(y)H0(Xy,y)y\mapsto\dim_{\kappa(y)}H^{0}(X_{y},\mathcal{L}_{y})

    is constant on YY. To see this, we note that if χ\chi denotes the Euler-characteristic for sheaf-cohomology, we have

    dimκ(y)H0(Xy,ym)=χ(ym)=χ(ηm)=H0(Xη,ηm)\dim_{\kappa(y)}H^{0}(X_{y},\mathcal{L}_{y}^{m})=\chi(\mathcal{L}^{m}_{y})=\chi(\mathcal{L}^{m}_{\eta})=H^{0}(X_{\eta},\mathcal{L}^{m}_{\eta})

    where η\eta denotes the generic point of YY. Here, we are using the higher cohomology vanishing (2.21) for nef invertible sheaves on the fibers (which are globally FF-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 m\mathcal{L}^{m} to each fiber of ff has vanishing higher cohomology, Grauert’s Theorem also implies that Rif(m)R^{i}f_{*}(\mathcal{L}^{m}) is zero for any m0m\geq 0.

    Fix a yYy\in Y and choose an open neighbourhood Spec(B)=UY\operatorname{\text{Spec}}(B)=U\subset Y of yy such that 𝔭\mathfrak{p} is generated by a regular sequence (b1,,bt)(b_{1},\dots,b_{t}) and B/𝔭BB/\mathfrak{p}B is regular (where 𝔭\mathfrak{p} is the prime ideal of BB corresponding to yy). This is possible because YY is regular by assumption. Then, note that the following exact sequence of sheaves on f1(U)f^{-1}(U):

    0{0}m{\mathcal{L}^{m}}m{\mathcal{L}^{m}}m𝒪𝒳𝒪𝒳B/x1B{\mathcal{L}^{m}\otimes_{\mathcal{O}_{\mathcal{X}}}\mathcal{O}_{\mathcal{X}_{B/x_{1}B}}}0{0}b1\scriptstyle{\cdot b_{1}}

    remains exact after applying H0(f1(U),)H^{0}(f^{-1}(U),\,-\,) since H1(f1(U),m)=0H^{1}(f^{-1}(U),\mathcal{L}^{m})=0. This tells us that

    H0(f1(U),m)BB/b1BH0(𝒳B/x1B,m|𝒳B/x1B).H^{0}(f^{-1}(U),\mathcal{L}^{m})\otimes_{B}B/b_{1}B\cong H^{0}(\mathcal{X}_{B/x_{1}B},\mathcal{L}^{m}|_{\mathcal{X}_{B/x_{1}B}}).

    Now, since B/x1BB/x_{1}B is also regular, we may proceed inductively to complete the proof of Part (1).

  2. (2)

    Fix any yYy\in Y and let AA be the local ring at yYy\in Y and RR be the local ring of any point xXx\in X mapping to yy.

    1. (a)

      Since RR is strongly FF-regular by assumption, we may apply Part (1) of 2.23 to conclude that R(m𝒟)R(m\mathcal{D}) is isomorphic to a summand of FeRF^{e}_{*}R. Since AA is regular, we have FeAF^{e}_{*}A is flat over AA and by assumption FeRF^{e}_{*}R is flat over FeAF^{e}_{*}A. Thus, we see that FeRF^{e}_{*}R is flat over AA and consequently, R(m𝒟)R(m\mathcal{D}) is flat over AA because it is a direct summand of FeRF^{e}_{*}R.

    2. (b)

      Fix a regular sequence b1,,btb_{1},\dots,b_{t} on AA generating the maximal ideal (this is possible because AA is regular). Because RR is flat over AA, b1,,btb_{1},\dots,b_{t} is also a regular sequence on RR. Now, it is sufficient to show that R(m𝒟)RRR(m\mathcal{D})\otimes_{R}R^{\prime} is reflexive over R=R/(b1,,bt)RR^{\prime}=R/(b_{1},\dots,b_{t})R. Note that RR^{\prime} is normal by [Mat89, Theorem 23.9] since all the fibers of ff are assumed to be globally FF-regular, which in particular implies that they are normal. Now the fact that R(m𝒟)RRR(m\mathcal{D})\otimes_{R}R^{\prime} is reflexive over RR^{\prime} is guaranteed by Part (2) of 2.23.

    3. (c)

      Since we have proved that \mathcal{F} is flat over YY in Part (a), we can repeat the argument from Part (1) of the Lemma using Grauert’s Theorem by replacing the vanishing theorem for \mathbb{Q}-Cartier ample divisors proved in 2.22 instead of 2.21.

    4. (d)

      This is immediate by combining Part (c) with the last part of 2.23. ∎

Remark 5.7.

Though 5.6 is stated for restriction to the fibers of ff, 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 ff are also normal by assumption (since they are globally FF-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 \mathcal{L} by a multiple of \mathcal{L} if necessary and assume that \mathcal{L} is globally generated on 𝒳\mathcal{X}. In that case, the restriction of \mathcal{L} to each perfectified-fiber of ff is also automatically globally generated.

Step 2:

Recall the sequence αe\alpha_{e} converging to the αF\alpha_{F}-invariant introduced in 3.8. Putting together 5.5 and 5.4, we get that for each yYy\in Y, and each e1e\geq 1,

|αe(Sy)αF(Sy)|d+1pe.|\alpha_{e}(S_{y^{\infty}})-\alpha_{F}(S_{y^{\infty}})|\leq\frac{d+1}{p^{e}}. (5.5)

Here, SyS_{y^{\infty}} denotes the section ring of the perfect fiber of ff over yy with respect to the restriction of \mathcal{L}, and dd denotes the dimension of every fiber of ff (which is well-defined because ff is flat).

Now, given any α<αgen\alpha<\alpha_{\text{gen}}, let ε=αgenα\varepsilon=\alpha_{\text{gen}}-\alpha. Fix an e0e\gg 0 such that d+1pe<ε/2\frac{d+1}{p^{e}}<\varepsilon/2. Then, applying Equation 5.5 to the (perfectified) generic fiber of ff, we have

αgenε/2<αe(SK).\alpha_{\text{gen}}-\varepsilon/2<\alpha_{e}(S_{K^{\infty}}). (5.6)

Claim:

There exists a dense open set UαYU_{\alpha}\subset Y (depending on ee) such that

αe(Sy)αe(SK)\alpha_{e}(S_{y^{\infty}})\geq\alpha_{e}(S_{K^{\infty}})

for each yUαy\in U_{\alpha}. Assuming the claim, Equation 5.5, and Equation 5.6 together imply that

αF(Sy)>αe(Sy)ε/2αe(SK)ε/2>αgenε=α\alpha_{F}(S_{y^{\infty}})>\alpha_{e}(S_{y^{\infty}})-\varepsilon/2\geq\alpha_{e}(S_{K^{\infty}})-\varepsilon/2>\alpha_{\text{gen}}-\varepsilon=\alpha

for every yUαy\in U_{\alpha} as required.

Step 3:

Fix an e0e\gg 0 as in Step 2. We now proceed to prove the claim used above, i.e., that there exists a dense open set UαYU_{\alpha}\subset Y (depending on the choice of e>0e>0) such that αe(Sy)αe(SK)\alpha_{e}(S_{y^{\infty}})\geq\alpha_{e}(S_{K^{\infty}}) for every yUαy\in U_{\alpha}. Working locally, we may assume that Y=Spec(A)Y=\operatorname{\text{Spec}}(A) and let SAS_{A} denote the section ring S(𝒳,)S(\mathcal{X},\mathcal{L}). Set me=me(SK)m_{e}=m_{e}(S_{K^{\infty}}). By 5.6, H0(𝒳,m)H^{0}(\mathcal{X},\mathcal{L}^{m}) is a locally free AA-module for any m0m\geq 0. So by shrinking YY around any point if necessary, we may assume that H0(𝒳,m)H^{0}(\mathcal{X},\mathcal{L}^{m}) is a free AA-module with a basis m\mathscr{B}_{m} for each 0mme0\leq m\leq m_{e}. By 5.6 again, m\mathscr{B}_{m} restricts to a basis of H0(Xy,y)H^{0}(X_{y},\mathcal{L}_{y}) for any yYy\in Y. Let =m=0mem\mathscr{B}=\sqcup_{m=0}^{m_{e}}\mathscr{B}_{m}. Note that mem_{e} is at most pe1p^{e}-1 since for any non-zero section fH0(XK,K)f\in H^{0}(X_{K^{\infty}},\mathcal{L}_{K^{\infty}}) (which exists because \mathcal{L} is assumed to be globally generated), we have 0fpeIe(pe)0\neq f^{p^{e}}\in I_{e}(p^{e}).

Step 4:

By definition of mem_{e} (3.7), for any mmem\leq m_{e} and any non-zero fH0(XK,Km)f\in H^{0}(X_{K^{\infty}},\mathcal{L}^{m}_{K^{\infty}}), the map SKFeSKS_{K^{\infty}}\to F^{e}_{*}S_{K^{\infty}} sending 11 to FefF^{e}_{*}f splits. Thus, using [LP23, Lemma 2.7 (a)] repeatedly on the set m\mathscr{B}_{m} (fixed in Step 3), we can construct a surjective map ψm:Fe(Km)𝒪XKm\psi_{m}:F^{e}_{*}(\mathcal{L}_{K^{\infty}}^{m})\to\mathcal{O}_{X_{K^{\infty}}}^{\oplus\mathscr{B}_{m}} such that if fmf\in\mathscr{B}_{m} is a basis element, then ψm(Fef)=𝟏f\psi_{m}(F^{e}_{*}f)=\mathbf{1}_{f}, where 𝟏f\mathbf{1}_{f} is the standard basis element corresponding to ff of 𝒪XKm\mathcal{O}_{X_{K^{\infty}}}^{\oplus\mathscr{B}_{m}} . Considering the induced map on the section modules and putting together the maps ψm\psi_{m} for all 0mme0\leq m\leq m_{e}, along with the zero map for me<mpe1m_{e}<m\leq p^{e}-1, we get a surjective SKS_{K^{\infty}}-module map

ψK:FeSKSK\psi_{K^{\infty}}:F^{e}_{*}S_{K^{\infty}}\to S_{K^{\infty}}^{\oplus\mathscr{B}} (5.7)

which satisfies the following property: if 0fH0(XK,Km)0\neq f\in H^{0}(X_{K^{\infty}},\mathcal{L}^{m}_{K^{\infty}}) and mmem\leq m_{e}, then ψ(Fef)\psi(F^{e}_{*}f) is a basis element of SKS^{\oplus\mathscr{B}}_{K^{\infty}}. In other words, ψK\psi_{K^{\infty}} simultaneously splits all the non-zero sections of degree at most mem_{e} on XKX_{K^{\infty}}. By [CRST21, Lemma 4.8], there is an integer de>0d_{e}>0, a non-zero element gAg\in A such that if B=A[g1]B=A[g^{-1}], we have a map

ψB1/pe+de:SB1/pe+de1/peSB1/pe+de\psi_{B^{1/p^{e+d_{e}}}}:S_{B^{1/p^{e+d_{e}}}}^{1/p^{e}}\to S_{B^{1/p^{e+d_{e}}}}^{\oplus\mathscr{B}}

which satisfies ψB1/pe+deB1/pe+deK=ψK\psi_{B^{1/p^{e+d_{e}}}}\otimes_{B^{1/p^{e+d_{e}}}}K^{\infty}=\psi_{K^{\infty}}. Here, SB1/pe+de1/peS_{B^{1/p^{e+d_{e}}}}^{1/p^{e}} denotes the ethe^{\text{th}}-relative Frobenius over the base change SB1/pe+de:=SAAB1/pe+deS_{B^{1/p^{e+d_{e}}}}:=S_{A}\otimes_{A}B^{1/p^{e+d_{e}}}. Note also that after base-changing to KK^{\infty}, we are identifying the relative and absolute ethe^{\text{th}}-Frobenius over KK^{\infty}. Since every ff\in\mathscr{B} is mapped to a basis element of SB1/pe+deS_{B^{1/p^{e+d_{e}}}}^{\oplus\mathscr{B}} after tensoring to KK^{\infty}, we may assume that the same is true for ψB1/pe+de\psi_{B^{1/p^{e+d_{e}}}} after inverting another element of AA if necessary. Finally, for any ySpec(B)y\in\operatorname{\text{Spec}}(B), base changing to κ(y)\kappa(y)^{\infty}, we see that

ψκ(y)=(ψB1/pe+deB1/pe+deκ(y)1/pe+de)κ(y)1/pe+deκ(y).\psi_{\kappa(y)^{\infty}}=\big{(}\psi_{B^{1/p^{e+d_{e}}}}\otimes_{B^{1/p^{e+d_{e}}}}\kappa(y)^{1/p^{e+d_{e}}}\big{)}\otimes_{\kappa(y)^{1/p^{e+d_{e}}}}\kappa(y)^{\infty}.

simultaneously splits each non-zero element 0fH0(Xy,ym)0\neq f\in H^{0}(X_{y^{\infty}},\mathcal{L}^{m}_{y^{\infty}}). Note that we are again identifying the absolute and relative Frobenius over the perfect field κ(y)\kappa(y)^{\infty}. This shows that αe(Sy)αe(SK)\alpha_{e}(S_{y^{\infty}})\geq\alpha_{e}(S_{K^{\infty}}). This completes the proof of 5.3. ∎

5.2. Semicontinuity for a family of \mathbb{Q}-Fano varieties

In this subsection, we will prove that the αF\alpha_{F}-invariant is lower semicontinuous in a family of globally FF-regular \mathbb{Q}-Fano varieties (see 4.1 for the definition of \mathbb{Q}-Fano varieties):

Theorem 5.8.

Let f:𝒳Yf:\mathcal{X}\to Y be flat family of globally FF-regular \mathbb{Q}-Fano varieties over kk, i.e., ff is a family of globally FF-regular varieties (with YY regular) such that K𝒳|Y-K_{\mathcal{X}|Y} is \mathbb{Q}-Cartier and ff-ample. Then, the map from Y0Y\to\mathbb{R}_{\geq 0} given by

yαF(𝒳y)y\mapsto\alpha_{F}(\mathcal{X}_{y^{\infty}})

is lower semicontinuous, where 𝒳y\mathcal{X}_{y^{\infty}} is the perfectified-fiber over yYy\in Y. See 4.2 for the definition of the αF\alpha_{F}-invariant of a Fano variety.

Remark 5.9.

For related semicontinuity results for the FF-signature and the Hilbert-Kunz multiplicity, see [Pol18], [PT18], [Smi16] and [Smi20]. Similarly, for the corresponding lower semicontinuity result for the complex α\alpha-invariant, see [BL22].

Idea of the proof:

Roughly, the proof of 5.8 involves combining 5.3 with the inversion of adjunction for strong FF-regularity as proved in [PSZ18]. The main technical difficulty arises when pp divides the index of K𝒳|YK_{\mathcal{X}|Y}. 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 \mathbb{Q}-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 AA be a regular kk-algebra of finite type. Suppose we have a flat family f:𝒳Y=Spec(A)f:\mathcal{X}\to Y=\operatorname{\text{Spec}}(A) of globally FF-regular Fano varieties (see 5.2) such that =𝒪𝒳(rK𝒳|Y)\mathcal{L}=\mathcal{O}_{\mathcal{X}}(-rK_{\mathcal{X}|Y}) is Cartier and ample for some integer r>0r>0. Then, we can form the section ring SA:=S(𝒳,)S_{A}:=S(\mathcal{X},\mathcal{L}) as described in 2.3. Since 𝒳\mathcal{X} is normal and \mathcal{L} is ample, SAS_{A} is a normal, finitely generated , \mathbb{N}-graded algebra over AA. For any AA-algebra BB, let SBS_{B} denote the section ring S(𝒳B,|𝒳B)S(\mathcal{X}_{B},\mathcal{L}|_{\mathcal{X}_{B}}), where 𝒳B\mathcal{X}_{B} denotes the base change 𝒳×ASpec(B)\mathcal{X}\times_{A}\operatorname{\text{Spec}}(B).

Lemma 5.10.

With notation as above, the construction of SAS_{A} satisfies the following properties:

  1. (1)

    SAS_{A} is flat over AA and for any prime ideal 𝔭A\mathfrak{p}\subset A we have that

    SAAκ(𝔭)Sκ(𝔭).S_{A}\otimes_{A}\kappa(\mathfrak{p})\cong S_{\kappa(\mathfrak{p})}.
  2. (2)

    For each prime 𝔭Spec(A)\mathfrak{p}\in\operatorname{\text{Spec}}(A), there is an affine open neighbourhood Spec(B)=UY\operatorname{\text{Spec}}(B)=U\subset Y containing 𝔭\mathfrak{p} such that the restriction of K𝒳K_{\mathcal{X}} to Spec(B/𝔭B)\operatorname{\text{Spec}}(B/\mathfrak{p}B) (as explained in 2.23) is linearly equivalent to K𝒳B/𝔭BK_{\mathcal{X}_{B/\mathfrak{p}B}}. In particular, for any 𝔭Spec(A)\mathfrak{p}\in\operatorname{\text{Spec}}(A), rK𝒳𝔭-rK_{\mathcal{X}_{\mathfrak{p}}} is Cartier. Moreover, 𝒳\mathcal{X} and SAS_{A} are both \mathbb{Q}-Gorenstein.

  3. (3)

    For each 𝔭Spec(A)\mathfrak{p}\in\operatorname{\text{Spec}}(A), there is an affine open neighbourhood Spec(B)=UY\operatorname{\text{Spec}}(B)=U\subset Y containing 𝔭\mathfrak{p} and a regular sequence b1,,btb_{1},\dots,b_{t} on BB generating 𝔭\mathfrak{p} such that SBS_{B}, and SB/(b1,,bi)BS_{B/(b_{1},\dots,b_{i})B} is strongly FF-regular for each 1it1\leq i\leq t. In particular, 𝒳\mathcal{X} and SAS_{A} are both globally FF-regular.

  4. (4)

    For any Weil-divisor 𝒟\mathcal{D} on 𝒳\mathcal{X} such that r𝒟r\mathcal{D} is Cartier and ample for some r>0r>0, the section module

    MA(𝒟)=m0H0(𝒳,𝒪X(𝒟)m)M_{A}(\mathcal{D})=\bigoplus_{m\geq 0}H^{0}(\mathcal{X},\mathcal{O}_{X}(\mathcal{D})\otimes\mathcal{L}^{m})

    is flat over AA, and compatible with base change to fibers. In other words, for any 𝔭Spec(A)\mathfrak{p}\in\operatorname{\text{Spec}}(A), the natural map

    MA(𝒟)Aκ(𝔭)Mκ(𝔭)(𝒪𝒳(𝒟)|𝒳𝔭)M_{A}(\mathcal{D})\otimes_{A}\kappa(\mathfrak{p})\to M_{\kappa(\mathfrak{p})}(\mathcal{O}_{\mathcal{X}}(\mathcal{D})|_{\mathcal{X}_{\mathfrak{p}}})

    is an isomorphism. Moreover, if Supp(𝒟)\text{Supp}(\mathcal{D}) does not contain any fiber of ff then the natural map

    MA(𝒟)Aκ(𝔭)Mκ(𝔭)(𝒟|𝔭)M_{A}(\mathcal{D})\otimes_{A}\kappa(\mathfrak{p})\to M_{\kappa(\mathfrak{p})}(\mathcal{D}|_{\mathfrak{p}})

    is an isomorphism as well.

Proof.

Since all parts of the lemma can be proved locally on YY, we may shrink YY if necessary to assume that ωY\omega_{Y} is a free AA-module.

  1. (1)

    This is immediate from Part (1) of 5.6.

  2. (2)

    By inverting an element gA𝔭g\in A\setminus\mathfrak{p}, and setting B=A[g1]B=A[g^{-1}], we may assume that 𝔭\mathfrak{p} is generated by a regular sequence b1,,btb_{1},\dots,b_{t} on BB (this is possible since A𝔭A_{\mathfrak{p}} is regular). Fix any 1it1\leq i\leq t and let Bi=B/(b1,,bi)BB_{i}=B/(b_{1},\dots,b_{i})B. By [Mat89, Theorem 23.9], since all fibers of ff are normal, we in particular know that each 𝒳Bi\mathcal{X}_{B_{i}} (and hence, SBiS_{B_{i}}) is normal. By shrinking UU further if necessary, using Part (1) of 5.6, we may also assume SBiSBBBiS_{B_{i}}\cong S_{B}\otimes_{B}B_{i}.

    Now we will show that K𝒳K_{\mathcal{X}} restricted to 𝒳Bi{\mathcal{X}_{B_{i}}} is linearly equivalent to the divisor K𝒳BiK_{\mathcal{X}_{B_{i}}}. Let V𝒳BiV^{\prime}\subset\mathcal{X}_{B_{i}} denote the smooth locus of 𝒳Bi\mathcal{X}_{B_{i}}, and V𝒳smV\subset\mathcal{X}_{\text{sm}} be an open set such that V𝒳Bi=VV\cap\mathcal{X}_{B_{i}}=V^{\prime}. This is possible because 𝒳Bi\mathcal{X}_{B_{i}} is a complete intersection in 𝒳B\mathcal{X}_{B}. Then, applying the adjunction formula for the complete intersection VVV^{\prime}\subset V, we have KVKV|VK_{V^{\prime}}\sim K_{V}|_{V^{\prime}}. Since V𝒳BiV^{\prime}\subset\mathcal{X}_{B_{i}} is the smooth locus and 𝒳Bi\mathcal{X}_{B_{i}} is normal, it contains all the codimension one points of 𝒳Bi\mathcal{X}_{B_{i}}. Thus, taking closures we get that

    rK𝒳|𝒳Bi=rK𝒳|V¯=rKV|V¯rK𝒳Bi.-rK_{\mathcal{X}}|_{\mathcal{X}_{B_{i}}}=\overline{-rK_{\mathcal{X}}|_{V^{\prime}}}=\overline{-rK_{V}|_{V^{\prime}}}\sim-rK_{\mathcal{X}_{B_{i}}}.

    This proves that \mathcal{L} restricted to 𝒳Bi\mathcal{X}_{B_{i}} is linearly equivalent to rK𝒳Bi-rK_{\mathcal{X}_{B_{i}}}. In particular rK𝒳Bi-rK_{\mathcal{X}_{B_{i}}} is Cartier. Furthermore, it follows from the discussion in [SS10, Section 5.2]) that the canonical divisor KSBiK_{S_{B_{i}}} is the cone over the canonical divisor K𝒳BiK_{\mathcal{X}_{B_{i}}} (as Weil-divisor). Thus, we get that 𝒪SBi(rKSBi)𝒪SBi(1)\mathcal{O}_{S_{B_{i}}}(-rK_{S_{B_{i}}})\cong\mathcal{O}_{S_{B_{i}}}(1) as a graded module over SBiS_{B_{i}}. Also note that for any maximal ideal 𝔪A\mathfrak{m}\subset A, the fiber Sκ(𝔪)S_{\kappa(\mathfrak{m})} is strongly FF-regular. In particular, all fibers over closed points of Spec(A)\operatorname{\text{Spec}}(A) are Cohen-Macaulay, we see that SAS_{A} is Cohen-Macaulay as well. Therefore, SAS_{A} (and similarly SBiS_{B_{i}}) is \mathbb{Q}-Gorenstein.

  3. (3)

    From part (2), we may assume that 𝔭\mathfrak{p} is generated by a regular sequence a1,ata_{1},\dots a_{t} on AA and rK𝒳Ai-rK_{\mathcal{X}_{A_{i}}} is Cartier for each 1it1\leq i\leq t, where Ai=A/(a1,,ai)AA_{i}=A/(a_{1},\dots,a_{i})A. Next, note that by assumption (and Part (1)), we have Sκ(𝔭)S_{\kappa(\mathfrak{p})} is strongly FF-regular. So, there exists a non-zero element gA𝔭g\in A\setminus\mathfrak{p} such that SBtS_{B_{t}} is strongly FF-regular where Bt=At[g1]B_{t}=A_{t}[g^{-1}]. Here, we are using the fact that the non-strongly FF-regular locus of SBtS_{B_{t}} is closed, compatible with localization and homogeneous with respect to the \mathbb{N}-grading on SBtS_{B_{t}}. Let Bt1=At1[g1]B_{t-1}=A_{t-1}\,[g^{-1}]. Then, since KSBt1-K_{S_{B_{t-1}}} is \mathbb{Q}-Cartier and ata_{t} is a non-zero divisor on Bt1B_{t-1}, by [Das15, Theorem A], we conclude that SBt1S_{B_{t-1}} is strongly FF-regular in a neighbourhood of 𝕍(bt)\mathbb{V}(b_{t}). Since the locus of points where SBt1S_{B_{t-1}} is not strongly FF-regular is defined by a homogeneous ideal (and is disjoint from 𝕍(at)\mathbb{V}(a_{t})), we may pick a non-zero element hA[g1](a1,,at)h\in A[g^{-1}]\setminus(a_{1},\dots,a_{t}) such that if we set Bt1=Bt1[h1]B_{t-1}^{\prime}=B_{t-1}[h^{-1}], SBt1S_{B^{\prime}_{t-1}} is strongly FF-regular. Replacing 𝔭\mathfrak{p} with (a1,,at1)(a_{1},\dots,a_{t-1}), we may proceed inductively to get a localization BB of AA (at finitely many elements), and an open neighbourhood U=Spec(B)Spec(A)U=\operatorname{\text{Spec}}(B)\subset\operatorname{\text{Spec}}(A) of 𝔭\mathfrak{p} such that SBS_{B} and SB/(a1,,ai)BS_{B/(a_{1},\dots,a_{i})B} is strongly FF-regular for each 1it1\leq i\leq t, as required.

    Applying this to maximal ideals in AA, we see that SAS_{A} is strongly FF-regular. Moreover, let 𝒟n\mathcal{D}\sim\mathcal{L}^{n} for some n0n\gg 0 be an effective divisor and ff denote the corresponding degree nn element of SAS_{A}. Then Spec(SA[1f])\operatorname{\text{Spec}}(S_{A}[\frac{1}{f}]) is isomorphic to the product of Spec(A[t,t1])\operatorname{\text{Spec}}(A[t,t^{-1}]) and 𝒳Supp(𝒟)\mathcal{X}\setminus\text{Supp}(\mathcal{D}). Since 𝒳Supp(𝒟)\mathcal{X}\setminus\text{Supp}(\mathcal{D}) is affine and strongly FF-regular (because we have shown that SAS_{A} is strongly FF-regular), to show that 𝒳\mathcal{X} is globally FF-regular, using [SS10, Theorem 3.9], it is sufficient to show that for some e0e\gg 0, the map 𝒪𝒳Fe𝒪𝒳(𝒟)\mathcal{O}_{\mathcal{X}}\to F^{e}_{*}\mathcal{O}_{\mathcal{X}}(\mathcal{D}) splits. But this again follows from the fact that SAS_{A} is strongly FF-regular by using 2.7. Hence, 𝒳\mathcal{X} is globally FF-regular.

  4. (4)

    Using Part (3) above, 𝒳\mathcal{X} is in particular, locally strongly FF-regular. Now the claim is an immediate consequence of Part (2) of 5.6. ∎

Lemma 5.11.

Let f:𝒳Y=Spec(A)f:\mathcal{X}\to Y=\operatorname{\text{Spec}}(A) be as above and assume AA is regular. Suppose 𝒟\mathcal{D} is a \mathbb{Q}-divisor on 𝒳\mathcal{X} satisfying the following two properties:

  1. (1)

    there is some e>0e>0 for which (pe1)𝒟(p^{e}-1)\mathcal{D} is an integral Weil divisor linearly equivalent to (1pe)K𝒳|Y(1-p^{e})K_{\mathcal{X}|Y} as Weil divisors, and

  2. (2)

    Supp(𝒟)\text{Supp}(\mathcal{D}) does not contain any fiber of ff.

Then, if y0Yy_{0}\in Y is a point such that the pair (𝒳y0,λ𝒟y0)(\mathcal{X}_{y_{0}^{\infty}},\lambda\,\mathcal{D}_{y_{0}^{\infty}}) is globally FF-regular for some λ0\lambda\in\mathbb{Q}_{\geq 0}, then there is an open neighbourhood UYU\subset Y of y0y_{0} such that (𝒳y,λ𝒟y)(\mathcal{X}_{y^{\infty}},\lambda\mathcal{D}_{y^{\infty}}) is also globally FF-regular for all yUy\in U.

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 YY to a neighbourhood of y0y_{0} if necessary, we may also assume that ωY\omega_{Y} is isomorphic to AA. Fix an r>0r>0 such that rK𝒳-rK_{\mathcal{X}} is an ample Cartier divisor, and set :=rK𝒳\mathcal{L}:=-rK_{\mathcal{X}} and SA=S(𝒳,)S_{A}=S(\mathcal{X},\mathcal{L}) be the corresponding section ring and Spec(SA)\operatorname{\text{Spec}}(S_{A}) is the cone over 𝒳\mathcal{X}. By 5.6, taking the cone over 𝒳\mathcal{X} commutes with base change to fibers of ff. For any integral Weil divisor \mathcal{E}, let MA()=m0H0(𝒳,𝒪X()m)M_{A}(\mathcal{E})=\bigoplus_{m\geq 0}H^{0}(\mathcal{X},\mathcal{O}_{X}(\mathcal{E})\otimes\mathcal{L}^{m}) denote the corresponding section module over SAS_{A}.

Step 2:

For any ee sufficiently divisible, let 𝒟e=(pe1)𝒟\mathcal{D}_{e}=(p^{e}-1)\mathcal{D}, which we may assume is an integral Weil-divisor. Since K𝒳-K_{\mathcal{X}} is \mathbb{Q}-Cartier, so is 𝒟e\mathcal{D}_{e}. Therefore, by Part (4) of 5.10, the section module MA(𝒟e)M_{A}(\mathcal{D}_{e}) is compatible with base change to fibers. Thus, we may consider the cone over 𝒟e\mathcal{D}_{e} as a Weil-divisor on Spec(SA)\operatorname{\text{Spec}}(S_{A}) defined by the reflexive sheaf MA(𝒟e)M_{A}(\mathcal{D}_{e}). The compatibility with base changing to fibers guarantees that taking the cone over 𝒟e\mathcal{D}_{e} commutes with restricting 𝒟e\mathcal{D}_{e} to the fibers of ff.

Step 3:

Since the pair (𝒳κ(y0),μ𝒟κ(y0))(\mathcal{X}_{\kappa(y_{0})^{\infty}},\mu\,\mathcal{D}_{\kappa(y_{0})^{\infty}}) is globally FF-regular for μ=λ\mu=\lambda, it remains so for all μ<λ+ε\mu<\lambda+\varepsilon for some small ε>0\varepsilon>0. Now we set μ:=12\mu:=\frac{\ell_{1}}{\ell_{2}} to be a rational number such that λμ<λ+ε\lambda\leq\mu<\lambda+\varepsilon for positive integers 1\ell_{1} and 2\ell_{2} with the following properties:

  1. (1)

    21\ell_{2}-\ell_{1} is divisible by rr (the Cartier index of K𝒳K_{\mathcal{X}}).

  2. (2)

    2\ell_{2} is not divisible by pp.

Furthermore, choose e0>0e_{0}>0 such that 2\ell_{2} divides pe01p^{e_{0}}-1 and (pe01)𝒟(p^{e_{0}}-1)\mathcal{D} and (pe01)K𝒳-(p^{e_{0}}-1)K_{\mathcal{X}} are both integral Weil-divisors. Such an e0e_{0} exists by our assumptions on 𝒟\mathcal{D}. Lastly, set Δ=μ𝒟\Delta=\mu\,\mathcal{D}. With this notation, we note that the following properties are satisfied:

  • Δ=1(pe01)2(pe01)𝒟\Delta=\frac{\ell_{1}}{(p^{e_{0}}-1)\ell_{2}}\,(p^{e_{0}}-1)\mathcal{D}, where (pe01)𝒟(p^{e_{0}}-1)\mathcal{D} is a Weil-divisor and pp does not divide (pe01)2(p^{e_{0}}-1)\ell_{2}.

  • (pe01)2(K𝒳+Δ)(p^{e_{0}}-1)^{2}\,(K_{\mathcal{X}}+\Delta) is linearly equivalent to

    (pe01)2(K𝒳+Δ)(pe01)2(112)K𝒳=(pe01)2(21)2K𝒳.(p^{e_{0}}-1)^{2}\,(K_{\mathcal{X}}+\Delta)\sim(p^{e_{0}}-1)^{2}(1-\frac{\ell_{1}}{\ell_{2}})\,K_{\mathcal{X}}=\frac{(p^{e_{0}}-1)^{2}(\ell_{2}-\ell_{1})}{\ell_{2}}\,K_{\mathcal{X}}. (5.8)

    And since rr divides 21\ell_{2}-\ell_{1}, we get that (pe01)2(K𝒳+Δ)(p^{e_{0}}-1)^{2}(K_{\mathcal{X}}+\Delta) is an integral Cartier divisor. Also note that (pe01)2(p^{e_{0}}-1)^{2} clearly divides (pe01)2(p^{e_{0}}-1)\ell_{2} and is not divisible by pp.

  • The fibers of ff are geometrically normal since the perfect fibers are globally FF-regular. They are also geometrically connected by assumption. Thus, our assumption that Supp(𝒟)\text{Supp}(\mathcal{D}) does not contain any fibers of ff guarantees that Supp(𝒟)\text{Supp}(\mathcal{D}) does not contain any generic point of any geometric fiber of ff.

Step 4:

In this context, we may use [PSZ18, Corollary 4.19] applied to the projective cone of ff with respect to \mathcal{L} to conclude the proof (see Section 2.2 for details about the projective cone construction). More precisely, consider the map

f¯:𝒳¯:=Proj(SA[z])Spec(A)\overline{f}:\overline{\mathcal{X}}:=\operatorname{\text{Proj}}(S_{A}[z])\to\operatorname{\text{Spec}}(A)

where zz is just another variable adjoined to SAS_{A} in degree 11. Note that rK𝒳¯-rK_{\overline{\mathcal{X}}} is Cartier on 𝒳¯\overline{\mathcal{X}} (since this is true at the zero-section by construction, and away from the zero section we know that 𝒳¯\overline{\mathcal{X}} is an 𝔸1\mathbb{A}^{1}-bundle over 𝒳\mathcal{X}).

By 5.10, the construction of the projective cone with respect to \mathcal{L} is compatible with base change to fibers. In other words, for any yYy\in Y, the fiber f¯y:𝒳¯ySpec(κ(y))\overline{f}_{y}:\overline{\mathcal{X}}_{y}\to\operatorname{\text{Spec}}(\kappa(y)) is the map

f¯y:𝒳¯y=Proj(Sκ(y)[z])Spec(κ(y)).\overline{f}_{y}:\overline{\mathcal{X}}_{y}=\operatorname{\text{Proj}}(S_{\kappa(y)}[z])\to\operatorname{\text{Spec}}(\kappa(y)).

Let Δ¯\overline{\Delta} denote the \mathbb{Q}-divisor obtained as the closure in 𝒳¯\overline{\mathcal{X}} of the cone over Δ\Delta. For any r>0r>0 such that rΔr\Delta is integral, the section module corresponding to rΔ¯r\overline{\Delta} is MA(rΔ)[z]M_{A}(r\Delta)[z] by construction. Thus, by Step 2,the construction of the projective cone over Δ\Delta is compatible with restricting to fibers. Thus, by Equation 5.8 and the fact that away from the zero-section, 𝒳¯\overline{\mathcal{X}} is an 𝔸1\mathbb{A}^{1}-bundle over 𝒳\mathcal{X}, we have that K𝒳¯+Δ¯K_{\overline{\mathcal{X}}}+\overline{\Delta} is \mathbb{Q}-Cartier with index not divisible by pp.

Step 5:

Finally, we observe that for any yYy\in Y, the local strong FF-regularity of (𝒳¯κ(y),Δ¯κ(y))(\overline{\mathcal{X}}_{\kappa(y)^{\infty}},\overline{\Delta}_{\kappa(y)^{\infty}}) is equivalent to the global FF-regularity of (𝒳y,Δy)(\mathcal{X}_{y^{\infty}},\Delta_{y^{\infty}}), since both correspond to the strong FF-regularity of the pair (SA,Δ)(S_{A},\Delta) . With these observations in place, [PSZ18, Corollary 4.19] gives us an open neighbourhood UYU\subset Y of y0y_{0} such that (𝒳y,Δy)(\mathcal{X}_{y^{\infty}},\Delta_{y^{\infty}}) is globally FF-regular for all yUy\in U. Since, Δ=μ𝒟\Delta=\mu\mathcal{D} and λμ\lambda\leq\mu, the same is true for (𝒳y,λ𝒟y)(\mathcal{X}_{y^{\infty}},\lambda\mathcal{D}_{y^{\infty}}) 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 y0Yy_{0}\in Y and α>0\alpha>0 such that αF(𝒳y0)>α\alpha_{F}(\mathcal{X}_{y_{0}^{\infty}})>\alpha, there exists an open neighbourhood U(y0,α)YU(y_{0},\alpha)\subset Y such that αF(Xy)>α\alpha_{F}(X_{y^{\infty}})>\alpha for all yU(y0,α)y\in U(y_{0},\alpha). 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 y0y_{0} instead of just any open subset of YY.

Firstly, by shrinking YY to a neighbourhood of y0y_{0}, we assume Y=Spec(A)Y=\operatorname{\text{Spec}}(A) is affine and ωY\omega_{Y} is a free AA-module. Next, using 3.13, it is sufficient to prove the lower semicontinuity of the function

yαF(S(𝒳y,rK𝒳y))y\mapsto\alpha_{F}(S(\mathcal{X}_{y^{\infty}},-rK_{\mathcal{X}_{y^{\infty}}}))

for any r0r\gg 0. So we pick an r0r\gg 0 such that rK𝒳-rK_{\mathcal{X}} is a globally generated ample divisor on 𝒳\mathcal{X}. In particular, rK𝒳-rK_{\mathcal{X}} is Cartier. Therefore, Part (2) of 5.10, we have that rK𝒳y-rK_{\mathcal{X}_{y^{\infty}}} is a globally generated ample Cartier divisor on 𝒳y\mathcal{X}_{y^{\infty}} for any yYy\in Y. Additionally, fix an integer t>0t>0 such that H0(𝒳,𝒪𝒳(mK𝒳))0H^{0}(\mathcal{X},\mathcal{O}_{\mathcal{X}}(-mK_{\mathcal{X}}))\neq 0 for all mtm\geq t.

Let dd be the relative dimension of ff, let α0\alpha_{0} denote αF(Sy0)\alpha_{F}(S_{y_{0}^{\infty}}), and ε:=α0α\varepsilon:=\alpha_{0}-\alpha. Recall that for any AA-algebra BB, SBS_{B} denotes the section ring S(𝒳B,|𝒳B)S(\mathcal{X}_{B},\mathcal{L}|_{\mathcal{X}_{B}}), where 𝒳B\mathcal{X}_{B} denotes the base change 𝒳×ASpec(B)\mathcal{X}\times_{A}\operatorname{\text{Spec}}(B) and =𝒪X(rK𝒳)\mathcal{L}=\mathcal{O}_{X}(-rK_{\mathcal{X}}). By the argument in Step 2 of the proof of 5.3 (replacing the generic point with y0y_{0}), it is sufficient to show that there exists an e0e\gg 0 such that d+1pe<ε/2\frac{d+1}{p^{e}}<\varepsilon/2 and an open neighbourhood U(y0,α)U(y_{0},\alpha) of y0y_{0} such that

αe(Sy)α0ε2\alpha_{e}(S_{y^{\infty}})\geq\alpha_{0}-\frac{\varepsilon}{2}

for all yU(y0,α)y\in U(y_{0},\alpha). To prove this, choose an e0e\gg 0 such that d+1pe<ε/2\frac{d+1}{p^{e}}<\varepsilon/2, and

α0(pe1)>pe(α0ε/2)+t.\alpha_{0}\,(p^{e}-1)>p^{e}\,(\alpha_{0}-\varepsilon/2)+t. (5.9)

Let nn be an integer such that r(pe(α0ε/2)+t)<n<rα0(pe1)r(p^{e}(\alpha_{0}-\varepsilon/2)+t)<n<r\,\alpha_{0}\,(p^{e}-1) be an integer that is not divisible by pp. By Part 2c of 5.6, we may assume (by shrinking YY if necessary) that H0(𝒳,𝒪𝒳(mK𝒳))H^{0}(\mathcal{X},\mathcal{O}_{\mathcal{X}}(-mK_{\mathcal{X}})) is a free AA-module for each 0mn0\leq m\leq n with a basis m\mathscr{B}_{m}. Let 𝒟\mathcal{D} be any element of n\mathscr{B}_{n} and npe1<λ<rα0\frac{n}{p^{e}-1}<\lambda<r\,\alpha_{0} be any rational number. Since 𝒟\mathcal{D} is a basis element using Part 2c of 5.6 again, we see that 𝒟\mathcal{D} does not contain any fibers of ff, since 𝒟\mathcal{D} restricts to a non-zero global section on each fiber. Then, we apply 5.11, to the \mathbb{Q}-divisor 1n𝒟\frac{1}{n}\mathcal{D} to get an open neighbourhood U=U(y0,α)U=U(y_{0},\alpha) such that for each yUy\in U, the pair (𝒳y,λn𝒟|𝒳y)(\mathcal{X}_{y^{\infty}},\frac{\lambda}{n}\mathcal{D}|_{\mathcal{X}_{y^{\infty}}}) is globally FF-regular. This is because since λ<rα0\lambda<r\alpha_{0}, the pair (𝒳y0,λn𝒟|y0)(\mathcal{X}_{y_{0}^{\infty}},\frac{\lambda}{n}\mathcal{D}|_{y_{0}^{\infty}}) is globally FF-regular by the definition of α0\alpha_{0}. Since λn>1pe1\frac{\lambda}{n}>\frac{1}{p^{e}-1} by construction, 3.6 tells us that for each yYy\in Y, and each 𝒟\mathcal{D} in n\mathscr{B}_{n}, the map

SyFe(Sy(𝒟y))S_{y^{\infty}}\to F^{e}_{*}(S_{y^{\infty}}(\mathcal{D}_{y^{\infty}})) (5.10)

splits. Furthermore, we may pick an integer mnm\leq n satisfying: rr divides mm, nmtn-m\geq t, and m>rpe(α0ε/2)m>rp^{e}\,(\alpha_{0}-\varepsilon/2). This is possible by our choice of nn. Since nmtn-m\geq t, we can pick a non-zero nm\mathcal{E}\in\mathscr{B}_{n-m} that restricts to a non-zero Weil-divisor on each fiber (by base-change). Thus, for any element 𝒟\mathcal{D} in m\mathscr{B}_{m}, since the corresponding map for (𝒟+)y(\mathcal{D}+\mathcal{E})_{y^{\infty}} splits for each yUy\in U by Equation 5.10, the corresponding map for 𝒟y\mathcal{D}_{y^{\infty}} also splits for each yUy\in U. Finally, we apply [LP23, Lemma 2.7 (a)] repeatedly to the basis m\mathscr{B}_{m} to conclude that Ie(m)=0I_{e}(m)=0 for SyS_{y^{\infty}} for each yUy\in U and thus

αe(Sy)mrpe>α0ε/2\alpha_{e}(S_{y^{\infty}})\geq\frac{m}{rp^{e}}>\alpha_{0}-\varepsilon/2

for all yUy\in U as required. This completes the proof of 5.8. ∎

6. Examples

In this section, we compute some examples of the αF\alpha_{F}-invariant for non-toric varieties and highlight some interesting features.

6.1. Quadric hypersurfaces

Fix any algebraically closed field kk of characteristic p2p\neq 2 and let Qdd+1Q_{d}\subset\mathbb{P}^{d+1} be the dd-dimensional smooth quadric hypersurface over kk. Note that by the adjunction formula, KQd=dH-K_{Q_{d}}=dH where HH denotes a hyperplane section.

Example 6.1.

Then, αF(Qd)=1d\alpha_{F}(Q_{d})=\frac{1}{d}. Equivalently, if SS denotes the section ring

S:=S(Qd,𝒪X(1))k[x0,,xd+1]/(x02+xd+12),S:=S(Q_{d},\mathcal{O}_{X}(1))\cong k[x_{0},\dots,x_{d+1}]/(x_{0}^{2}+\dots x_{d+1}^{2}),

then αF(S)=1\alpha_{F}(S)=1. This follows from a description of the structure of the sheaves Fe(𝒪Qd(m))F^{e}_{*}(\mathcal{O}_{Q_{d}}(m)) proved in [Lan08] and [Ach12]. More precisely, for any e1e\geq 1 and 0mpe10\leq m\leq p^{e}-1, [Ach12, Theorem 2] tells us that Fe(𝒪Qd(m))F^{e}_{*}(\mathcal{O}_{Q_{d}}(m)) is a direct sum of 𝒪Qd(t)\mathcal{O}_{Q_{d}}(-t) and 𝒮(t)\mathcal{S}(-t) for t0t\geq 0, where 𝒮\mathcal{S} is an ACM bundle that sits in an exact sequence of the form

0𝒪d+1(2)a𝒪d+1(1)bi𝒮00\to\mathcal{O}_{\mathbb{P}^{d+1}}(-2)^{\oplus a}\to\mathcal{O}_{\mathbb{P}^{d+1}}(-1)^{\oplus b}\to i_{*}\mathcal{S}\to 0

for suitable positive integers aa and bb. Here i:Qdd+1i:Q_{d}\hookrightarrow\mathbb{P}^{d+1} is the inclusion. See [Ach12, Section 1.3] for the details. Since H1(d+1,𝒪d+1(2))=0H^{1}(\mathbb{P}^{d+1},\mathcal{O}_{\mathbb{P}^{d+1}}(-2))=0, we deduce from the exact sequence above that 𝒮(t)\mathcal{S}(-t) has no global sections for any t0t\geq 0. Therefore, all global sections of Fe(𝒪Qd(m))F^{e}_{*}(\mathcal{O}_{Q_{d}}(m)) appear in the trivial summands. In other words, Ie(S(m))=0I_{e}(S(m))=0 for any e1e\geq 1 and any mpe1m\leq p^{e}-1. Moreover, since S10S_{1}\neq 0, we know that me=pe1m_{e}=p^{e}-1. Therefore, by 3.8, we have

αF(S)=limepe1pe=1.\alpha_{F}(S)=\lim_{e\to\infty}\frac{p^{e}-1}{p^{e}}=1.
Remark 6.2.

This example shows that the αF\alpha_{F}-invariant does not characterize regularity of section rings, since the αF\alpha_{F}-invariant of a polynomial ring is also equal to 11.

Remark 6.3.

Another interesting feature of this example is that the αF\alpha_{F}-invariant of smooth quadrics is independent of the characteristic pp (for p2p\neq 2). This is far from true in general (as seen in the next example). Furthermore, for any d>2d>2, the FF-signature of QdQ_{d} is known to depend on pp in a rather complicated way (see [Tri23]).

6.2. Comparison to the complex α\alpha-invariant

Let k=𝔽pk=\mathbb{F}_{p} for some prime number p5p\geq 5 and Xp3X_{p}\subset\mathbb{P}^{3} be the diagonal cubic surface defined by x3+y3+z3+w3=0x^{3}+y^{3}+z^{3}+w^{3}=0 over kk.

Example 6.4.

For each p5p\geq 5, we have αF(Xp)<12\alpha_{F}(X_{p})<\frac{1}{2}. However,

limpαF(Xp)=12.\lim_{p\to\infty}\alpha_{F}(X_{p})=\frac{1}{2}.

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 𝓈p\mathscr{s}_{p} denote the FF-signature of XpX_{p}, equivalently, of the ring 𝔽p[x,y,z,w]/(x3+y3+z3+w3)\mathbb{F}_{p}[x,y,z,w]/(x^{3}+y^{3}+z^{3}+w^{3}). Then for any p5p\geq 5, we have 𝓈p<18\mathscr{s}_{p}<\frac{1}{8} . Moreover,

limp𝓈p=18.\lim_{p\to\infty}\mathscr{s}_{p}=\frac{1}{8}.

Using this, our claims about the αF\alpha_{F}-invariant of XpX_{p} follow from 4.7 and 4.8, once we observe that

vol(KXp)22 3!=18.\frac{\mathrm{vol}(-K_{X_{p}})}{2^{2}\,3!}=\frac{1}{8}.
Remark 6.5.

The complex α\alpha-invariant of the cubic surface defined by x3+y3+z3+w3x^{3}+y^{3}+z^{3}+w^{3} is equal to 2/32/3 (see [Che08, Theorem 1.7]). Note that by [HY03], we know that for a fixed divisor DD on a variety XX

limpfpt(Xp,Dp)=lct(X,D)\lim_{p\to\infty}\text{fpt}(X_{p},D_{p})=\text{lct}(X,D)

where XpX_{p} and DpD_{p} denote the reduction to characteristic pp of XX and DD respectively. 6.4 points to limitations of approximating the log canonical threshold by FF-pure threshold for an unbounded family of divisors on XX.

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 FF-signature and strong FF-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 \mathbb{Q}-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. FF-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 FF-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 F{F}-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. FF-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. DD-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 \mathbb{Q}-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 FF-singularities and base change of relative canonical sheaves. 2012. arXiv:1207.1910.
  • [PSZ18] Zsolt Patakfalvi, Karl Schwede, and Wenliang Zhang. FF-singularities in families. Algebr. Geom., 5(3):264–327, 2018.
  • [PT18] Thomas Polstra and Kevin Tucker. FF-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 FF-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. FF-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 FF-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 C1(M)>0C_{1}(M)>0. Invent. Math., 89(2):225–246, 1987.
  • [Tri23] Vijaylaxmi Trivedi. The hilbert-kunz density functions of quadric hypersurfaces, 2023.
  • [Tuc12] Kevin Tucker. FF-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 FF-signature of local rings of characteristic pp. J. Algebra, 299(1):198–218, 2006.