1 Introduction
Singularities of the Minimal Model Program (MMP) have poor cohomological properties in positive characteristic, unlike in characteristic
$0$
. Striking examples are canonical and terminal singularities that are not Cohen–Macaulay [Reference Kovács22, Reference Bernasconi7, Reference Yasuda34, Reference Totaro32, Reference Totaro33]. In this article, we construct several new examples of non-
$S_3$
singularities of the MMP in every positive characteristic
$p>0$
.
The examples constructed in the aforementioned works are of two types: they stem either from a cone construction on a variety failing Kodaira vanishing or from a wild
$\mathbb {Z}/p\mathbb {Z}$
-quotient. Our examples are
$\alpha _p$
-quotients: to understand their singularities we use the perspective of
$1$
-foliations and the tools developed in [Reference Posva27].
Our first series of examples gives the following:
Theorem 1.1 (Theorem 3.2)
Let k be an algebraically closed field of characteristic
$p>0$
. Then there exists an isolated
$\mathbb {Q}$
-factorial singularity
$(\mathbf {0}\in Y)$
over k that is canonical (resp. terminal) of dimension
$\max \{p, 3\}$
(resp.
$p+1$
) and non-
$S_3$
.
In particular, we obtain a three-dimensional terminal singularity in characteristic
$p=2$
that is not
$S_3$
: the other known examples of this dimension were all constructed by Totaro in positive and mixed characteristics
$2,3$
, and
$5$
[Reference Totaro32, Reference Totaro33]. For an arbitrary characteristic
$p>2$
, Totaro also constructed terminal non-
$S_3$
cone singularities of dimension
$2p+2$
[Reference Totaro32]. In this regard, we achieve an improvement on the asymptotic dimensional lower bound.
Note that Yasuda exhibits in [Reference Yasuda34] a series of terminal and canonical singularities, starting in dimension
$4$
, which have depth
$\geq 3$
but are not Cohen–Macaulay: their optimal asymptotic dimension, as a function of p, is roughly given by
$\sqrt {2p}$
.
We establish several other properties of the singularities
$(\textbf {0}\in Y)$
in Proposition 3.4, Proposition 3.5, Corollary 3.6, and Proposition 3.7. Notably, they are not F-injective, and they are rational up to Frobenius nilpotence.
We use the same method to produce another series of pathological examples. In characteristic
$0$
, fibers of locally stable one-parameter families (in the sense of Definition 2.1) are
$S_2$
[Reference Kollár16, §7.3]. This is true for one-parameter families of surfaces with residue characteristics different from
$2$
,
$3$
, and
$5$
[Reference Arvidsson, Bernasconi and Patakfalvi3], but recent examples of Kollár [Reference Kollár17] show that this property can fail dramatically for families of threefolds in every positive characteristic. Using
$\alpha _p$
-quotients, we construct new examples of this kind. First we obtain:
Theorem 1.2 (Theorem 4.3)
Let k be an algebraically closed field of characteristic
$p>0$
. Then there exists a locally stable family
$Y\to \mathbb {A}^1$
over k of relative dimension
$\max \{p,3\}$
such that
$Y_t$
is smooth for
$t\neq 0$
and
$Y_0$
is reduced but non-
$S_2$
.
Then we compactify the families of Theorem 1.2 to obtain pathological stable families (in the sense of Definition 2.1):
Theorem 1.3 (Theorem 4.7)
Let k be an algebraically closed field of characteristic
$p>0$
. Then there exists a projective flat family of pairs
$(\mathcal {Y},\mathcal {B})\to \mathbb {A}^1$
over k of relative dimension
$\max \{p,3\}$
such that:
-
1.
$\mathcal {Y}$
is normal and
$(\mathcal {Y},\mathcal {B})\to \mathbb {A}^1$
is stable, -
2.
$\mathcal {Y}_t$
has only
$\mu _p$
-quotient singularities (or is regular if
$p=2$
) for
$t\neq 0$
, -
3.
$\mathcal {Y}_0$
is reduced but non-
$S_2$
.
Moreover, if C is a normal curve and if
$C\to \mathbb {A}^1$
is a finite flat morphism, then
$\mathcal {Y}_C$
is normal and
$(\mathcal {Y}_C,\mathcal {B}_C)\to C$
is stable.
Our method produces boundary divisors
$\mathcal {B}\neq 0$
whose coefficients are quite small, but the number of irreducible components of
$\mathcal {B}$
and the volume of
$K_{\mathcal {Y}_t}+\mathcal {B}_t$
are up to our liking: see Remark 4.10 and the proof of Theorem A.12.
The examples given by Theorem 1.3 differ from Kollár’s ones [Reference Kollár17] in two important ways: their dimension is constrained by the characteristic (while Kollár constructs pathological families of threefold pairs in every characteristic, and indicates how to generalize the construction to any higher dimension), but their general fibers are klt, Cohen–Macaulay, and F-injective by [Reference Posva27, Theorems 1 and 2] (while the general fibers in Kollár’s examples are not klt nor Cohen–Macaulay: see Remark A.11). Boundary divisors appear in both situations. We discuss the implications of these examples for moduli theory in Section 1.1 below.
Let me sketch the construction of these examples. We start with a
$1$
-foliation
$\mathcal {F}$
of rank
$1$
on an affine space
$\mathbb {A}^n$
, and we wish to understand the singularities of
$\mathbb {A}^n/\mathcal {F}$
. There is a simple criterion guaranteeing that
$\mathbb {A}^n/\mathcal {F}$
is only
$S_2$
(see [Reference Posva27, Lemma 2.5.4]), so we only need to study its singularities in the sense of the MMP. If F is a divisor over
$\mathbb {A}^n/\mathcal {F}$
, then by [Reference Posva27, §4.2] we can compute the discrepancy
$a(F;\mathbb {A}^n/\mathcal {F})$
in terms of
$a(E;\mathbb {A}^n)$
and
$a(E;\mathcal {F})$
where E is the closure of F in the function field of
$\mathbb {A}^n$
(this is recalled in Theorem 2.10). The problem is that we cannot estimate
$a(E;\mathcal {F})$
easily. So we blow-up
$\mathbb {A}^n$
until the singularities of
$\mathcal {F}$
become log canonical. This leads to a commutative diagram

where the horizontal arrows are the quotient morphisms, and both morphisms b and
$\mathfrak {b}$
are birational. Writing
where the
$F_i$
’s are the
$\mathfrak {b}$
-exceptional divisors, we reduce to computing the discrepancies over the sub-pair
$(Z, -\sum _ia_i F_i)$
. The advantage is now that
$b^*\mathcal {F}$
is log canonical, so the tools given by Theorem 2.10 are much more efficient. The appearance of the characteristic in the computations is due to the adjunction formula of Proposition 2.9.
For Theorem 1.2 and Theorem 1.3 we use a relative version of this construction. Proving that the central fiber of the quotient is not
$S_2$
is more involved (as the criterion used before does not apply): we use the tools developed in [Reference Posva27, §5] (see Theorem 2.12). To ensure that the fibers of the quotient family are regular in codimension one, we need to work in relative dimension d at least
$3$
. When
$d=2=p$
the output (Example 4.6) is quite different: we obtain a locally stable family of surfaces
$\mathcal {S}\to \mathbb {A}^1$
where
$\mathcal {S}$
is Cohen–Macaulay and
$\mathcal {S}_0$
has an inseparable node in the sense of [Reference Posva28].
In our examples, the
$1$
-foliation
$\mathcal {F}$
is generated by (some variation of) the quadratic derivation
$\partial =\sum _{i=1}^nx_i^2\partial _{x_i}$
. The latter satisfies
$\partial ^{[p]}=0$
in every characteristic
$p>0$
, and thus defines an
$\alpha _p$
-action on
$\mathbb {A}^n$
such that
$\mathbb {A}^n/\alpha _p=\mathbb {A}^n/\partial $
[Reference Posva27, Proposition 2.3.7]. If
$n=2=p$
then
$\mathbb {A}^2/\partial $
is the RDP
$D^0_4$
. The quadratic derivation has the nice property that a single blow-up simplifies its singularities into lc ones, making the necessary computations straightforward.
In principle, other choices of
$1$
-foliations
$\mathcal {F}$
can lead to similar examples, possibly with better dimensional bounds. The difficult part is the computation of the blow-up sequence
$b\colon X\to \mathbb {A}^n$
needed to simplify the singularities of
$\mathcal {F}$
. The results of [Reference Posva29] suggest that such a sequence usually exists, but the number of necessary blow-ups might be quite large. To illustrate this, in Section 4.2 we consider the derivation
$y^3\partial _x+x\partial _y +t\partial _z$
on
$\mathbb {A}^3_{x,y,z}\times \mathbb {A}^1_t$
: this choice was motivated by the fact that the quotient of
$\mathbb {A}^2$
by
$y^3\partial _x+x\partial _y$
gives the RDP
$E^0_6$
in characteristic
$3$
. The quotient gives a pathological family of affine threefolds in characteristic
$3$
(Theorem 4.14). This family is more difficult to compactify, as explained in Remark 4.15: we shall not pursue this task here.
Remark 1.4. The underlying families
$Y\to \mathbb {A}^1$
(or
$\mathcal {Y}\to \mathbb {A}^1$
) appearing in Theorem 1.2, Theorem 1.3, and Theorem 4.14 share the following properties: the central fibers
$Y_0$
do not have liftable cohomology (in the sense of [Reference Kollár and Kovács20]), the pair
$(Y,Y_0)$
is not F-pure along
$Y_0$
, and Y is not F-injective at the non-
$S_2$
-points of
$Y_0$
. All the other fibers
$Y_t (t\neq 0)$
have liftable cohomology, and
$(Y\setminus Y_0,Y_t)$
is F-pure (Footnote
1
).
1.1 Consequences for KSBA moduli theory of stable pairs
The existence of the pathological stable families of Theorem 1.3 has the following consequence, which extends the main result of [Reference Kollár17]:
Theorem 1.5 (Theorem A.12)
Fix an integer
$n\geq 3$
. Then in characteristics
$p\leq n$
the stable families of n-dimensional Cohen–Macaulay pairs (resp. of pairs with F-injective underlying varieties) usually do not compactify to a proper KSBA moduli stack.
The formulation of that theorem requires an explanation. In positive characteristic, no precise definition of the KSBA moduli stack (Footnote
2
) of stable pairs has been proposed so far in the literature: defining families of pairs over general bases is the delicate part. It is easier to define boundary-free families: to wit, the moduli stacks of stable surfaces are defined in [Reference Patakfalvi25, §1.3] and studied in [Reference Patakfalvi25, Reference Arvidsson, Bernasconi and Patakfalvi3, Reference Posva28]. That definition can be extended to any dimension d, but then only few properties of the resulting stacks are known: for example, when
$d\geq 3$
we do not know if they are algebraic. In any case, the boundary-free case is not our setting.
There is however no ambiguity about what the values of the KSBA moduli stacks should be over regular curves: stable families of pairs with demi-normal underlying fibers (see Definition A.2). If the moduli stacks would satisfy the valuative criterion of properness, then a stable family with demi-normal fibers over a punctured curve could be extended with a demi-normal fiber, possibly after replacing the base with a finite base-change. The examples of Theorem 1.3 show that this is usually not possible when the characteristic is small compared to the dimension, even if the generic fiber has only klt, F-injective and Cohen–Macaulay singularities.
This discussion and the content of Theorem 1.5 are made precise in Appendix A.
Let us close this introduction with a few open questions related to KSBA moduli theory in positive characteristic:
Question 1.6. Let k be an algebraically closed field of characteristic
$p>0$
.
-
1. Are the poor properties of stable families of Cohen–Macaulay/F-injective pairs related to the ratio between the dimension and the characteristic? More precisely, are there families of n-dimensional pairs such as those of Theorem 1.3 if
$p\gg n$
? -
2. Do such pathologies also appear for stable families of n-dimensional varieties? More precisely, are there stable families of n-dimensional varieties (without boundaries) over regular curves with pathologies such as those of [Reference Kollár17] and Theorem 1.3? What if we assume that
$p\gg n$
, or that the generic fiber is Cohen–Macaulay or F-injective? -
3. Are there pathological families of stable surface pairs in characteristics
$\leq 5$
?
2 Preliminaries
2.1 Conventions and notations
We work over an algebraically closed field k of positive characteristic
$p>0$
.
-
1. A variety (over k) is an integral quasi-projective k-scheme of finite type. A curve (resp. surface, threefold) is k-variety of dimension one (resp. two, three).
-
2. The conditions
$S_i$
are the Serre’s conditions, see [1, 0341]. -
3. Given a variety X, a divisor E over X is a prime divisor appearing on a (proper) birational model
$\mu \colon Y\to X$
. Its center
$c_X(E)\subset X$
is the closure of
$\mu (E)$
. -
4. Let us recall some standard MMP terminology from [Reference Kollár and Mori21, Reference Kollár16]. Let X be a normal variety and
$\Delta $
be a
$\mathbb {Q}$
-divisor such that
$K_X+\Delta $
is
$\mathbb {Q}$
-Cartier. For a proper birational morphism
$\mu \colon Y\to X$
with Y normal we write We define
$$ \begin{align*}K_Y = \mu^*(K_X+\Delta)+\sum_{E} a(E;X,\Delta)\cdot E.\end{align*} $$
We say that
$$ \begin{align*}\operatorname{\mathrm{discrep}}(X,\Delta)=\inf\{a(E;X,\Delta) \mid E\text{ exceptional over }X\}.\end{align*} $$
$(X,\Delta )$
is sub-lc (resp. sub-klt) if for every
$\mu $
and E we have
$a(E;X,\Delta )\geq -1$
(resp.
$a(E;X,\Delta )>-1$
). If
$\Delta \geq 0$
and
$(X,\Delta )$
is sub-lc (resp. sub-klt), we say that
$(X,\Delta )$
is lc (resp. klt). We say that
$(X,\Delta )$
is canonical (resp. terminal) if
$\Delta $
is effective, and for every
$\mu $
and
$\mu $
-exceptional E we have
$a(E;X,\Delta )\geq 0$
(resp.
$a(E;X,\Delta )>0$
). In case
$(X,\operatorname {\mathrm {Supp}}(\Delta ))$
is log smooth, sufficient and necessary conditions on the coefficients of
$\Delta $
for these conditions to hold are given in [Reference Kollár16, Corollary 2.11]. In particular, a regular variety is terminal.
-
5. We will several times use tacitly the following fact [Reference Kollár16, Lemma 2.5]: if
$K_X+\Delta $
is
$\mathbb {Q}$
-Cartier, if
$\mu \colon Y\to X$
is birational with Y normal, and if then
$$ \begin{align*}K_Y+\mu^{-1}_*\Delta+F=\mu^*(K_X+\Delta) \quad \text{where }F\text{ is }\mu\text{-exceptional},\end{align*} $$
$a(E;X;\Delta )=a(E;Y,\mu _*^{-1}\Delta +F)$
for every divisor E over X.
We also use the following terminology (see Appendix A for more details):
Definition 2.1 [Reference Kollár18, (2.3, 2.44)]
Let C be a normal k-variety of dimension
$\leq 1$
, let
$f\colon (X,\Delta )\to C$
be a family of pairs [Reference Kollár18, Definition 2.2]–in particular, the fibers
$X_c$
are geometrically reduced–, and assume that X is normal. Then we say that f is locally stable if
$(X,\Delta +X_\alpha )$
is lc for every
$\alpha \in C(k)$
. We say that f is stable if it is proper, locally stable and
$K_X+\Delta $
is f-ample.
We record the following well-known result.
Lemma 2.2. Let X be a regular k-variety and G be a divisor over X with nonempty center
$\mathfrak {c}=c_X(G)\subset X$
. Then
$a(G;X)\geq \operatorname {\mathrm {codim}}_X(\mathfrak {c})-1$
.
Proof. By a result of Zariski [Reference Kollár and Mori21, Lemma 2.45], there is a finite sequence of birational proper morphisms
such that each
$f_i$
is the blow-up of
$X_{i-1}$
along the center
$\mathfrak {c}_{i-1}$
of G on
$X_{i-1}$
, and G appears on
$X_n$
as an
$f_n$
-exceptional divisor dominating
$\mathfrak {c}_{n-1}$
. To prove the statement we may localize each
$X_i$
in a neighbourhood of the generic point of
$\mathfrak {c}_{i}$
, and so we can assume that each
$\mathfrak {c}_i$
is a regular subvariety and that each
$X_i$
is regular. Let
$E_i\subset X_i$
be the unique
$f_i$
-exceptional divisor: since
$E_i=f_i^{-1}(\mathfrak {c}_{i-1})$
it holds that
$\mathfrak {c}_i\subseteq E_i$
for each i, with equality if and only if
$i=n$
. By abuse of notation, the strict transform of
$E_i$
on each
$X_{i+1},\dots ,X_n$
will also be denoted by
$E_i$
.
Let
$f\colon X_n\to X$
be the composition of all blow-ups. We write
$$ \begin{align*}f^*K_{X}=K_{X_n}-\sum_{i=1}^{n}a_iE_i.\end{align*} $$
We claim that
$a_{i+1}>a_i$
for all i: since
$a_1=\operatorname {\mathrm {codim}}(\mathfrak {c})-1$
and
$E_n=G$
, the result will follow. These inequalities are easily proved by induction: we have
$$ \begin{align*}f_i^*\left(K_{X_{i-1}}-\sum_{j=1}^{i-1}a_jE_j\right) = K_{X_i}-\sum_{j=1}^{i-1}a_jE_j -\left(\operatorname{\mathrm{codim}}_{X_{i-1}}(\mathfrak{c}_{i-1})-1 +\operatorname{\mathrm{mult}}_{\mathfrak{c}_{i-1}}\sum_{j}a_jE_j\right) \cdot E_i.\end{align*} $$
Since
$\mathfrak {c}_{i-1}\subsetneq E_{i-1}$
we have
This proves the lemma.
2.2 1-Foliations
For basics on derivations and
$1$
-foliations, we refer to [Reference Posva27]. We recall below the most important facts for us. For derivations we need:
Lemma 2.3. Let
$\partial $
be a p-closed derivation on an integral scheme
$\operatorname {\mathrm {Spec}}(A)$
. If there exists
$a\in A$
such that
$\partial (a)\neq 0$
and
$\partial ^{[p]}(a)=h\partial (a)$
for some
$h\in A$
, then
$\partial ^{[p]}=h\partial $
.
For
$1$
-foliations, let us recall:
Definition 2.4. Let
$\mathcal {F}$
be a
$1$
-foliation on a normal variety X. Assume that the divisorial sheaf
$\mathcal {O}_X(K_{\mathcal {F}})=(\bigwedge ^{\operatorname {\mathrm {rk}}\mathcal {F}}\mathcal {F})^{[-1]}$
is
$\mathbb {Q}$
-Cartier. Then
$\mathcal {F}$
is log canonical (lc) if for every proper birational morphism
$\mu \colon Y\to X$
we can write
$$ \begin{align*}K_{\mu^*\mathcal{F}}=\mu^*K_{\mathcal{F}}+\sum_{E: \ \mu\text{-exc.}} a(E;\mathcal{F}) \cdot E\end{align*} $$
with
$$ \begin{align*}a(E;\mathcal{F})\geq \begin{cases} 0 & \text{if }E\text{ is }\mu^*\mathcal{F}\text{-invariant,}\\ -1 & \text{otherwise.} \end{cases}\end{align*} $$
Definition 2.5. Let
$\mathcal {F}$
be a
$1$
-foliation of rank
$1$
on a normal variety X. Then
$\mathcal {F}$
has only multiplicative singularities (Footnote
3
) if for every
$x\in \operatorname {\mathrm {Sing}}(\mathcal {F})$
, the restriction
$\mathcal {F}\otimes \widehat {\mathcal {O}}_{X,x}$
is generated up-to-saturation by a continuous derivation
$\partial \in \operatorname {\mathrm {Der}}^{\text {cont}}_k \left (\widehat {\mathcal {O}}_{X,x}\right )$
satisfying
$\partial ^{[p]}=u\partial $
where
$u\in \widehat {\mathcal {O}}_{X,x}^\times $
.
Proposition 2.6 [Reference Posva27, Corollary 1]
Let X be a regular variety. If
$\mathcal {F}$
is a
$1$
-foliation of rank
$1$
, then
$\mathcal {F}$
has only multiplicative singularities if and only if
$\mathcal {F}$
is lc.
Pullbacks of
$1$
-foliations along smooth blow-ups can be computed as follows:
Example 2.7. Let
$\pi \colon X\to \mathbb {A}^n_{x_1,\dots ,x_n}$
be the blow-up of the ideal
$(x_1,\dots ,x_r)$
for
$r\leq n$
. The
$x_1$
-chart
$U_1$
of the blow-up
$\pi ^*\colon k[x_1,\dots ,x_n]\to k[y_1,\dots ,y_n]$
is given by
The pullbacks of derivations on
$\mathbb {A}^n$
by
$\pi |_{U_1}$
can be computed using the following equalities:
$$ \begin{align*}\pi^*\partial_{y_1}|_{U_1}=\partial_{y_1}-\sum_{i=2}^r\frac{y_i}{y_1}\partial_{y_i}, \quad \pi^*\partial_{x_i}|_{U_1}=\frac{1}{y_1}\partial_{y_i} \ (1<i\leq r), \quad \pi^*\partial_{x_j}|_{U_1}=\partial_{y_j} \ (j>r).\end{align*} $$
Let us recall some properties of quotients by
$1$
-foliations.
Definition 2.8 [Reference Posva27, Definition 4.2.1]
Let X be a normal variety,
$\mathcal {F}$
a
$1$
-foliation on X and
$q\colon X\to X/\mathcal {F}$
the quotient morphism. If
$\Delta =\sum _i a_i\Delta _i$
is a
$\mathbb {Q}$
-divisor on X, where
$\Delta _i$
are prime divisors, then we let
where
$$ \begin{align*}\varepsilon_i= \begin{cases} 1 & \text{if }\Delta_i\text{ is }\mathcal{F}\text{-invariant,}\\ \frac{1}{p} & \text{otherwise.} \end{cases}\end{align*} $$
Proposition 2.9 [Reference Posva27, Proposition 4.2.3]
Let X be a normal variety,
$\Delta $
a
$\mathbb {Q}$
-divisor on X,
$\mathcal {F}$
a
$1$
-foliation on X and
$q\colon X\to X/\mathcal {F}$
the quotient morphism. Then we have an equality of
$\mathbb {Q}$
-divisors
The following theorem is our main tool to compute the discrepancies over quotient singularities.
Theorem 2.10. Let
$(X,\Delta )$
be a normal sub-pair,
$\mathcal {F}$
be a
$1$
-foliation on X, and
$q\colon (X,\Delta )\to (Y=X/\mathcal {F},\Delta _Y)$
be the associated quotient morphism.
-
1. If
$\mathcal {F}$
is lc and
$(X,\Delta )$
is sub-lc (respectively sub-klt), then
$(Y,\Delta _Y)$
is sub-lc (respectively sub-klt).
Let
$\mu \colon Y'\to Y$
be a birational proper morphism. Then:
-
2. If
$X'$
is the normalization of
$Y'$
in
$K(X)$
, then we have a commutative diagram where
$\mu '$
is a proper birational morphism, and
$q'$
is the quotient by
$(\mu ')^*\mathcal {F}$
.
-
3. If
$E\subset X'$
is a prime divisor with image the prime divisor
$q'(E)=F\subset Y'$
, then
$$ \begin{align*}a(F; Y,\Delta_Y)= \begin{cases} a(E; X,\Delta)+(p-1)\cdot a(E; \mathcal{F}) & \text{if }E\text{ is }(\mu')^*\mathcal{F}\text{-invariant,} \\ \frac{1}{p}\left[ a(E; X,\Delta)+(p-1)\cdot a(E; \mathcal{F})\right] & \text{otherwise}. \end{cases}\end{align*} $$
Proof. This is proved in [Reference Posva27, Theorem 4.2.5], under the additional assumption that
$\Delta $
is effective. However, that assumption is not needed in the proof.
Let us also recall the notion of family of
$1$
-foliations.
Definition 2.11 [Reference Posva27, §5.1]
Let
$X\to B$
be a flat morphism between normal varieties with geometrically normal fibers. A relative
$1$
-foliation is a
$1$
-foliation on X that is contained in the sub-module
$T_{X/B}$
of
$T_{X/k}$
. If
$X\to B$
is smooth, we say that a relative
$1$
-foliation
$\mathcal {F}\subset T_{X/B}$
is a family of
$1$
-foliations if
$T_{X/B}/\mathcal {F}$
is flat over B and for every
$b\in B$
the sub-module
$\mathcal {F}\otimes \mathcal {O}_{X_b}\hookrightarrow T_{X_b/k(b)}$
is a
$1$
-foliation.
Theorem 2.12 [Reference Posva27, Proposition 5.2.4, Corollary 5.2.5]
Let
$X\to B$
be a smooth morphism over a normal curve B, and
$\mathcal {F}\subset T_{X/B}$
be a family of
$1$
-foliations. Then:
-
1. for every
$b\in B$
there is a natural morphism
$\varphi _b\colon X_b/\mathcal {F}|_{X_b}\to (X/\mathcal {F})_b$
, -
2.
$\varphi _b$
is an isomorphism if and only if
$(X/\mathcal {F})_b$
is
$S_2$
, and -
3. if
$\mathcal {F}$
has only multiplicative singularities in a neighborhood of
$X_b$
then
$\varphi _b$
is an isomorphism.
Finally, the following lemma will be useful for computations related to base-changes:
Lemma 2.13. Let
$f\colon X\to B$
be a smooth morphism, with B a normal variety. Let
$g\colon B'\to B$
be a flat morphism from a normal variety, inducing the cartesian diagram

Let
$\mathcal {F}\subset T_{X/B}$
be a relative
$1$
-foliation, and
$\mathcal {F}'=(g')^*\mathcal {F}\hookrightarrow (g')^*T_{X/B}\cong T_{X'/B'}$
. Then
$\mathcal {F}'$
is a relative
$1$
-foliation, and the natural morphism
$X'\to (X/\mathcal {F})\times _BB'$
is isomorphic over
$B'$
to the quotient
$X'\to X'/\mathcal {F}'$
.
Proof. This can be checked Zariski-locally on
$X, B$
, and
$B'$
, so we may assume that
$X=\operatorname {\mathrm {Spec}}(R)$
,
$B=\operatorname {\mathrm {Spec}}(A)$
,
$B'=\operatorname {\mathrm {Spec}}(A')$
and that
$\mathcal {F}=\tilde {F}$
is generated by
$\partial _1,\dots ,\partial _n\in \operatorname {\mathrm {Der}}_A(R)$
. Let
$R'=R\otimes _A A'$
. We have a natural map
and the
$A'$
-sub-module
$F'$
of
$\operatorname {\mathrm {Der}}_{A'}(R')$
generated by
$\mathfrak {g}'(F)$
sheafifies into
$\mathcal {F}'$
. We check below that
$F'$
is closed under Lie brackets and p-powers. As the
$\partial _i'=\mathfrak {g}'(\partial _i)$
generate
$F'$
we see that
Similarly we have
Since
$A'$
is flat over A we therefore obtain that
$R^F\otimes _A A'=(R')^{F'}$
, and the lemma follows.
It remains to check that
$F'$
is closed under brackets and p-powers. Let
$\xi ,\psi \in \mathfrak {g}'(F)$
and
$r,s\in R'$
. As
$\mathfrak {g}'$
commutes with Lie brackets we have
$[r\xi , s\psi ]=rs\cdot [\xi ,\psi ]\in F'$
. By Hochschild’s formula [Reference Posva27, (2.3.0.a)] we see that
$(r\xi )^{[p]}\in F'$
. Finally, a formula of Jacobson [Reference Jacobson13, p. 209] shows that
$(r\xi +s\psi )^{[p]}$
and
$(r\xi )^{[p]}+(s\psi )^{[p]}$
differ by an
$R'$
-linear combination of multifold Lie brackets of
$r\xi $
and
$s\psi $
. Thus
$(r\xi +s\psi )^{[p]}\in F'$
, and we are done.
3 Non-
$S_3$
isolated MMP singularities
Consider the affine space
$\mathbb {A}^n$
over the field k with coordinates
$x_1,\dots ,x_n$
where
$n\geq 2$
. We introduce the derivation
$$ \begin{align*} \partial = \sum_{i=1}^n x_i^2\partial_{x_i} \quad \text{on } \mathbb{A}^n. \end{align*} $$
First we observe that:
Claim 3.1.
$\partial ^{[p]}=0$
.
Proof. Since the summands
$x_i^2\partial _{x_i}$
commute with each other, we have
$$ \begin{align*}\partial^{[p]}=\sum_{i=1}^n \left( x_i^2\partial_{x_i}\right)^{[p]}.\end{align*} $$
Since
$\partial _{x_i}$
is p-closed, by Hochschild’s formula [Reference Posva27, (2.3.0.a)] the scaling
$x_i^2\partial _{x_i}$
is also p-closed. It is easily seen that
$(x_i^2\partial _{x_i})^{[p]}(x_i)=0$
, and by Lemma 2.3 it follows that
$(x_i^2\partial _{x_i})^{[p]}=0$
. Therefore
$\partial ^{[p]}=0$
.
In particular
$\mathcal {F}=\mathcal {O}_{\mathbb {A}^n}\cdot \partial $
is a
$1$
-foliation on
$\mathbb {A}^n$
.
Theorem 3.2. With the notations as above, let
$Y=\mathbb {A}^n/\mathcal {F}$
(with
$n\geq 2$
). Then:
-
1. Y is a normal
$\mathbb {Q}$
-factorial variety, with a unique singular closed point
$\mathbf {0}$
; -
2.
$\mathcal {O}_{Y,\textbf {0}}$
is not
$S_3$
if
$n\geq 3$
; -
3. It holds that:
-
• If
$n\leq p-2$
, Y is not lc. -
• If
$n\geq p-1$
, Y is lc; -
• If
$n\geq p$
, Y is canonical; -
• If
$n\geq p+1$
, Y is terminal.
-
Proof. By construction Y is normal and
$\mathbb {Q}$
-factorial: in particular
$K_Y$
is
$\mathbb {Q}$
-Cartier. Since
$\operatorname {\mathrm {Sing}}(\mathcal {F})$
is the origin
$0\in \mathbb {A}^n$
, we see that Y is regular away from the image
$\textbf {0}$
of the origin [Reference Posva27, Lemma 2.5.10]. As the ideal generated by the image of
$\partial $
in the local ring
$\mathcal {O}_{\mathbb {A}^n,0}$
is
$\mathfrak {m}_{\mathbb {A}^n,0}$
-primary, by Claim 3.1 and [Reference Posva27, Lemma 2.5.4] we obtain that
$\mathcal {O}_{Y,\textbf {0}}=\mathcal {O}_{\mathbb {A}^n,0}^\partial $
is not
$S_3$
.
It remains to analyze the singularity of
$(\textbf {0}\in Y)$
through the glasses of the MMP. To begin with, we blow-up the origin of
$\mathbb {A}^n$
to simplify the singularities of
$\mathcal {F}$
:
Consider the
$x_1$
-patch
$U_1$
of this blow-up:
Using the blow-up computations of Example 2.7, we see that on this chart we have
$$ \begin{align*} b^*\partial|_{U_1} &= u_1 \cdot \left(u_1\partial_{u_1}-\sum_{i=2}^nu_i\partial_{u_i}\right) + \sum_{i=2}^n u_1^2u_i^2 \cdot \frac{1}{u_1}\partial_{u_i} \\&= u_1 \cdot \underbrace{\left[ u_1\partial_{u_1} + \sum_{i=2}^n (-u_i+u_i^2)\partial_{u_i} \right]}_{\psi} \end{align*} $$
Since
$b^*\partial $
is p-closed, its scaling
$\psi $
is also p-closed. As
$\psi (u_1)=u_1$
we obtain by Lemma 2.3 that
$\psi ^{[p]}=\psi $
. Therefore by Proposition 2.6 we obtain that
The situation is similar on the other charts, and therefore we find that
$b^*\mathcal {F}$
is an lc
$1$
-foliation. As
$\psi (u_1)\in (u_1)$
we also see that the exceptional divisor
$E\subset X$
of b is
$b^*\mathcal {F}$
-invariant. Finally, as
$b^*\partial |_{U_1}=u_1\cdot \psi $
we see that
$a(E;\mathcal {F})=-1$
.
Now let
$Z=X/b^*\mathcal {F}$
and consider the induced commutative diagram

where the horizontal arrows are the quotient morphisms and where
$\mathfrak {b}\colon Z\to Y$
is birational. Let
$F\subset Z$
be the unique prime
$\mathfrak {b}$
-exceptional divisor. By Theorem 2.10 we have
If
$n\leq p-2$
then
$a(F;Y)\leq -2$
and Y is not lc. From now on assume that
$n\geq p-1$
. We have:
This crepant relation shows that Y is lc if and only if
$\operatorname {\mathrm {discrep}}(Z,(p-n)F)\geq -1$
. Since
$b^*\mathcal {F}$
is lc, Theorem 2.10 shows that
$\operatorname {\mathrm {discrep}}(Z,(p-n)F)\geq -1$
as soon as
$\operatorname {\mathrm {discrep}}(X,(p-n)E)\geq -1$
. As
$(X,E)$
is log smooth, this holds when
$p-n\leq 1$
by [Reference Kollár16, Corollary 2.11]. This settles the lc case of the theorem.
It remains to study when Y is canonical (resp. terminal): for this we assume that
$n\geq p$
. Consider an exceptional divisor
$F'$
over Y, appearing on a birational model
$Z'\to Y$
. Replacing
$Z'$
by some blow-up, we may assume that
$Z'\to Y$
factors through a birational morphism
$\mathfrak {f}\colon Z'\to Z$
. Let
$f\colon X'\to X$
be the normalization of
$\mathfrak {f}$
in
$K(X)$
, and
$E'\subset X'$
be the prime divisor dominating
$F'\subset Z'$
. If
$f(E')\not \subset E$
then
$c_{F'}(Y)\neq \textbf {0}$
, and therefore
$a(F';Y)\geq 1$
as Y is regular away from
$\textbf {0}$
. Thus we may assume that
$f(E')\subset E$
. By Theorem 2.10 we have
$$ \begin{align*} a(F';Y) &= a(F';Z,(p-n)F) \\&= \delta_{E'} \cdot \left[ a(E'; X, (p-n)E) + (p-1)\cdot a(E';b^*\mathcal{F})\right] \end{align*} $$
where
$\delta _{E'}\in \{1, p^{-1}\}$
. We must find conditions that guarantee that
$a(F';Y)\geq 0$
(resp.
$>0$
), hence the factor
$\delta _{E'}$
will not play any role. As
$(X,E)$
is log smooth and
$p-n\leq 0$
, we have
$a(E';X,(p-n)E)\geq 1$
. Since regular
$1$
-foliations are canonical [Reference Posva27, Lemma 3.0.3], it follows that
$a(F';Y)>0$
unless the center of
$E'$
is contained in the singular locus of
$b^*\mathcal {F}$
: so from now on we assume that
$c_X(E')$
is contained in
$\operatorname {\mathrm {Sing}}(b^*\mathcal {F})$
. It is easily seen, using the computations above, that the singular locus of
$b^*\mathcal {F}$
is a finite collection of isolated points. Therefore, by Lemma 2.2 we see that
$a(E';X)\geq n-1$
. Moreover, as
$b^*\mathcal {F}$
is lc we have
$a(E';b^*\mathcal {F})\geq -1$
. Putting everything together, we get
$$ \begin{align*} \frac{1}{\delta_{E'}}a(F';Y) &= a(E';X)+(n-p)\cdot \operatorname{\mathrm{mult}}_{E'}(f^*E) +(p-1)\cdot a(E';b^*\mathcal{F}) \\&\geq (n-1) + (n-p) + (p-1)\cdot (-1) \\&= 2(n-p). \end{align*} $$
So if
$n\geq p$
(resp. if
$n>p$
) it holds that
$a(F';Y)\geq 0$
(resp.
$a(F';Y)>0$
). This completes the proof.
Remarks 3.3.
-
1. The case
$n=3\geq p$
gives a canonical non-Cohen–Macaulay isolated threefold singularity. It was shown in [Reference Arvidsson, Bernasconi and Lacini2] that klt threefold singularities (with perfect residue fields) are Cohen–Macaulay in characteristic
$p>5$
. This is optimal, as there also exist non-Cohen–Macaulay terminal threefold singularities in characteristic
$5$
[Reference Totaro33]. -
2. In characteristic
$0$
, strictly lc singularities may not be
$S_3$
, but this can only happen if they are log canonical centers [Reference Kollár16, Corollary 7.21]. If
$n=p-1$
in Theorem 3.2 then
$\textbf {0}$
is an lc center of Y, as the proof shows.
We end this section with some further remarks on the singularities
$(\textbf {0}\in Y)$
constructed in Theorem 3.2. Unless we specify otherwise, there will be no constraint on the characteristic p nor on
$n=\dim Y$
.
Proposition 3.4. If
$p>2$
, then the singularity
$(\mathbf {0}\in Y)$
is never
$1$
-Gorenstein (i.e.,
$K_Y$
is never Cartier).
This is false if
$p=2$
: for example, in dimension two we obtain the RDP
$D^0_4$
, which is Gorenstein.
Proof. Let us use the notation of the proof of Theorem 3.2. We perform a weighted blow-up of the closed point defined by
$(u_1,\dots ,u_n)$
in
$U_1\subset X$
with weights
$(1,p-1,\dots ,p-1)$
(see [Reference Posva29, Example 2.4.3]). A (schematic) affine chart of the weighted blow-up is given by
The (unique) exceptional divisor
$E'$
is given by
$(v_1=0)$
. We have
and thus
$a(E';X)=(n-1)(p-1)$
. Since
$$ \begin{align*}u_1\partial_{u_1}=v_1\partial_{v_1}-\sum_{i=2}^n(p-1)v_i\partial_{v_i}, \quad u_i\partial_{u_i}=v_i\partial_{v_i} \ (i\geq 2),\end{align*} $$
we find
$$ \begin{align*}\psi=v_1\cdot \left[ \partial_{v_1}+ \sum_{i=2}^n v_1^{p-2}v_i^2\partial_{v_i}\right].\end{align*} $$
Therefore
$a(E';b^*\mathcal {F})=-1$
and
$E'$
is not invariant with respect to the pullback of
$b^*\mathcal {F}$
. So if
$F'$
is the divisor over Y induced by
$E'$
, we find:
$$ \begin{align*} a(F';Y) &= a(F';Z, (p-n)F) \\&= \frac{1}{p}\cdot\left[ a(E'; X, (p-n)E)+(p-1)\cdot a(E';b^*\mathcal{F})\right] \\&= \frac{1}{p}\cdot \left[ (n-1)(p-1)-(p-n)+(p-1)(-1)\right] \\&= n-3 + \frac{2}{p}. \end{align*} $$
Thus when
$p>2$
we see that
$a(F';Y)\notin \mathbb {Z}$
. This shows that Y is not
$1$
-Gorenstein.
For the definition of
$W\mathcal {O}$
-rationality and Cohen–Macaulayness up to (Frobenius) nilpotence used in the following proposition, and the notions that appear in its proof, we refer to [Reference Baudin, Bernasconi and Kawakami6].
Proposition 3.5. The singularities
$(\mathbf {0}\in Y)$
are
$W\mathcal {O}$
-rational and Cohen–Macaulay up to nilpotence.
Proof. The property of
$W\mathcal {O}$
-rationality descends along finite universal homeomorphisms by [Reference Chatzistamatiou and Rülling9, Proposition 4.4.9]. It evidently holds for
$\mathbb {A}^n$
, and so it holds for Y.
Let us show that Y is CM up to nilpotence. By [Reference Baudin, Bernasconi and Kawakami6, Lemma 2.3] the natural morphism
$\mathcal {O}_Y\to Rq_*\mathcal {O}_{\mathbb {A}^n}$
is an isomorphism in
$D^b(\operatorname {\mathrm {Crys}}^F_Y)$
. Let
$\omega _Y^\bullet $
be a dualizing complex on Y: by [Reference Baudin5, Theorem 4.3.5], by applying
$\operatorname {\mathrm {RHom}}(-,\omega _Y^\bullet )$
to the previous isomorphism we obtain
Now we have
$$ \begin{align*} \operatorname{\mathrm{RHom}}(Rq_*\mathcal{O}_{\mathbb{A}^n},\omega_Y^\bullet) &\cong Rq_*\operatorname{\mathrm{RHom}}(\mathcal{O}_{\mathbb{A}^n},\omega_{\mathbb{A}^n}) \\&\cong Rq_*\omega_{\mathbb{A}^n} \\&\cong q_*\omega_{\mathbb{A}^n} \end{align*} $$
in
$D^b(\operatorname {\mathrm {Crys}}^C_Y)$
, where the first isomorphism holds by [Reference Baudin5, Corollary 5.1.7] and the third one holds as q is finite. This shows that
$\operatorname {\mathrm {RHom}}(\mathcal {O}_Y,\omega _Y^\bullet )$
is supported in a single degree as an element of the category
$\operatorname {\mathrm {Crys}}^C_Y$
. By [Reference Baudin, Bernasconi and Kawakami6, Lemma 3.2] it follows that
$\mathcal {O}_Y$
is CM up to nilpotence as an element of
$\operatorname {\mathrm {Crys}}_Y^F$
.
Corollary 3.6. The singularities
$(\mathbf {0}\in Y)$
are never F-injective.
Proof. This follows from [Reference Posva27, Lemma 2.5.4], but here is another proof that utilizes the notions introduced above. Since Y is CM up to nilpotence, the action of Frobenius on the local cohomology groups
$H^i_{\mathfrak {m}}(\mathcal {O}_{Y,\textbf {0}})$
, where
$\mathfrak {m}$
is the maximal ideal and
$0\leq i< \dim \mathcal {O}_{Y,\textbf {0}}$
, is nilpotent (see [Reference Baudin, Bernasconi and Kawakami6, Lemma 3.2]). Since
$\mathcal {O}_{Y,\textbf {0}}$
is
$S_2$
but not
$S_3$
, we have
$H^2_{\mathfrak {m}}(\mathcal {O}_{Y,\textbf {0}})\neq 0$
. Thus the Frobenius action on
$H^2_{\mathfrak {m}}(\mathcal {O}_{Y,\textbf {0}})$
is not injective. This shows that
$\mathcal {O}_{Y,\textbf {0}}$
is not F-injective.
Proposition 3.7. There exist proper birational morphisms
$\pi \colon V\to Y$
where V is a regular variety. Moreover, for any such:
-
1. the Frobenius action on
$R^i\pi _*\mathcal {O}_V$
is nilpotent for
$i>0$
(Footnote
4
), and -
2. the Frobenius trace action on
$R^i\pi _*\omega _V$
is nilpotent for
$i>0$
.
Proof. We begin by constructing a specific resolution of singularities and checking the nilpotence statements along it. Consider the diagram (3.1) in the proof of Theorem 3.2. The
$1$
-foliation
$b^*\mathcal {F}$
is locally generated by multiplicative derivations with isolated singularities. Thus the quotient
$Z=X/b^*\mathcal {F}$
has only isolated
$\mu _p$
-singularities [Reference Posva27, Proposition 4.1.1]. These toroidal singularities can be resolved using [Reference Kempf, Knudsen, Mumford and Saint-Donat15, Theorem 11* p.94]: let
$\mu \colon V'\to Z$
be the induced resolution of singularities. The composition
$\pi '=\mathfrak {b}\circ \mu \colon V'\to Y$
is also a resolution of singularities.
We claim that the sheaves
$R^i\mu _*\mathcal {O}_{V'}$
and
$R^i\mu _*\omega _{V'}$
vanish for
$i>0$
. This can be checked formally on Z, and thus we reduce to the toric situation. Then vanishing holds for
$i>0$
by [Reference Kempf, Knudsen, Mumford and Saint-Donat15, Theorem 14 p. 52]. From the Leray spectral sequence it follows that
By naturality of the Frobenius, these equalities reduce the nilpotence statements along
$\pi '$
to the nilpotence statements along the morphism
$\mathfrak {b}\colon Z\to Y$
. By construction the absolute Frobenius morphism
$F_Z\colon \mathcal {O}_Z\to F_*\mathcal {O}_Z$
factors through
$q^{\prime }_*\mathcal {O}_X$
. Therefore
$R^i\mathfrak {b}_*(F_Z)$
factors through
$R^i\mathfrak {b}_*(q^{\prime }_*\mathcal {O}_X)=q_*R^ib_*\mathcal {O}_X$
. Since b is the blow-up of a point of
$\mathbb {A}^n$
we have
$R^ib_*\mathcal {O}_X=0$
for
$i>0$
, which shows that
$R^i\mathfrak {b}_*(F_Y)$
is the zero map for
$i>0$
. By applying (derived) Hom into
$\omega _Z$
to the map
$F_Z$
, one finds dually that the Frobenius trace
$\operatorname {\mathrm {Tr}}_Z\colon F_*\omega _Z\to \omega _Z$
factors through
$q^{\prime }_*\omega _X$
. As before we have
$R^i\mathfrak {b}_*(q_*'\omega _X)=q_*R^ib_*\omega _X=0$
for
$i>0$
, which shows that
$R^i\mathfrak {b}_*(\operatorname {\mathrm {Tr}}_Z)$
is the zero map for
$i>0$
. This proves the nilpotence statements along
$\mathfrak {b}$
and
$\pi '$
.
Finally, let
$\pi \colon V\to Y$
be another proper birational morphism with V regular. By [Reference Chatzistamatiou and Rülling8, Theorem 1] we have
$R^i\pi _*\mathcal {O}_V\cong R^i\pi ^{\prime }_*\mathcal {O}_{V'}$
and
$R^i\pi _*\omega _V\cong R^i\pi ^{\prime }_*\omega _{V'}$
for every
$i\geq 0$
. So the nilpotence statements along
$\pi '$
imply the corresponding nilpotence statements along
$\pi $
, which concludes the proof.
4 Locally stable families with non-
$S_2$
special fibers
4.1 First example:
$\mathbf {\dim \geq \max \{p,3\}.}$
Consider the affine space
$\mathbb {A}^{n+1}_{x_1,\dots ,x_n,y}\times \mathbb {A}^1_t$
and the derivation
$$ \begin{align} \partial_m=\sum_{i=1}^{n}x_i^2\partial_{x_i} + t^m\mu(y)\partial_{y} \quad \text{on }\mathbb{A}^{n+1}\times\mathbb{A}^1, \end{align} $$
where
$m\geq 0$
and
$\mu (y)\in k[y]$
satisfies
$(\mu \partial _{y})^{[p]}=0$
and becomes a unit in a neighborhood of the origin, that is,
$\mu (0)\neq 0$
.
We let
$\mathbb {A}=\mathbb {A}^n_{\textbf {x}}\times D(\mu (y))\subset \mathbb {A}^{n+1}$
be the affine open subset where
$\mu (y)$
is invertible. We let
$\mathcal {F}_m\subset T_{(\mathbb {A}\times \mathbb {A}^1)/k}$
be the sub-module generated by
$\partial _m$
.
(We can of course take
$m=1$
,
$\mu =1$
and
$\mathbb {A}=\mathbb {A}^{n+1}$
; but this extra generality will be useful for Theorem 4.7 below.)
Claim 4.1.
$\mathcal {F}_m$
is a family of
$1$
-foliations over
$\mathbb {A}^1_t$
(Definition 2.11) if and only if
$n\geq 2$
.
Proof. Clearly
$\partial _m\in T_{(\mathbb {A}\times \mathbb {A}^1)/\mathbb {A}^1}$
, and we see that
$\partial _m^{[p]}=0$
as in Claim 3.1 (and this survives along any specialization of t). Thus
$\mathcal {F}_m$
is a relative
$1$
-foliation over
$\mathbb {A}^1_t$
, for any
$n\geq 1$
. It is easily seen that the cokernel of
$\mathcal {F}_m\hookrightarrow T_{\mathbb {A}\times \mathbb {A}^1/\mathbb {A}^1}$
is torsion-free, hence flat, over
$\mathbb {A}^1$
.
The restriction of
$\mathcal {F}_m$
to the generic fiber of the projection
$\mathbb {A}\times \mathbb {A}^1_t\to \mathbb {A}^1_t$
is the restriction to
$\mathbb {A}\times \operatorname {\mathrm {Spec}}(k(t))$
of the module generated by
$$ \begin{align*}\sum_{i=1}^nx_i^2\partial_{x_i}+t^m\mu(y)\partial_y\in \operatorname{\mathrm{Der}}_{k(t)}k(t)[x_1,\dots,x_n,y],\end{align*} $$
which is clearly a
$1$
-foliation for every
$n\geq 1$
. A similar computation shows that
$\mathcal {F}_m|_{(t=\alpha )}$
is a
$1$
-foliation for every
$0\neq \alpha \in k$
and
$n\geq 1$
. The only condition is given by the restriction to the central fiber
$(t=0)$
: indeed,
$\mathcal {F}_m|_{(t=0)}$
is generated by
$$ \begin{align*}\sum_{i=1}^nx_i^2\partial_{x_i}\in \operatorname{\mathrm{Der}}_{k}k[x_1,\dots,x_n,y],\end{align*} $$
so
$\mathcal {F}_m|_{(t=0)}$
is a
$1$
-foliation if and only
$n\geq 2$
. (For
$n=1$
, observe that
$x_1^2\partial _{x_1}$
does not generate a saturated sub-module of
$\operatorname {\mathrm {Der}}_{k}k[x_1,\dots ,x_n,y]$
.) The statement follows.
The following lemma explains how to simplify the singularities of
$\mathcal {F}_m$
.
Lemma 4.2. We have
$\operatorname {\mathrm {Sing}}(\mathcal {F}_m)=V(x_1,\dots ,x_n,t)$
for
$m\geq 1$
, and
$\mathcal {F}_0$
is regular. If
$m\geq 1$
and
$b\colon X=\operatorname {\mathrm {Bl}}_{\operatorname {\mathrm {Sing}}(\mathcal {F}_m)}\to \mathbb {A}\times \mathbb {A}^1$
is the blow-up of that singular locus with exceptional divisor E, then:
-
1.
$a(E;\mathcal {F}_m)=-1$
; -
2. E is
$b^*\mathcal {F}_m$
-invariant; -
3.
$b^*\mathcal {F}_m$
is lc on the
$x_i$
-patches, and isomorphic over
$\mathbb {A}^1_t$
to
$\mathcal {F}_{m-1}$
on the t-patch; -
4. if
$m\geq 2$
, the locus of non-lc singularities S of
$b^*\mathcal {F}_m$
is closed in X and contained on the t-patch; -
5.
$\operatorname {\mathrm {mult}}_SE=1$
,
$\operatorname {\mathrm {codim}}_XS=n+1$
and S is disjoint from the strict transform of
$(t=0)\subset \mathbb {A}\times \mathbb {A}^1_t$
.
Proof. The statement about the singular locus is clear. On the
$x_1$
-patch
$U_1$
the blow-up b is given by
$$ \begin{align*} b^*\colon & k[x_1,\dots,x_n,y,t][\mu(y)^{-1}]\longrightarrow k[u_1,\dots,u_n,y,s][\mu(y)^{-1}], \\& x_1\mapsto u_1, \quad x_i\mapsto u_1u_i \ (1<i\leq n), \quad y\mapsto y,\quad t\mapsto u_1s. \end{align*} $$
Therefore we have
$$ \begin{align*} b^*\partial_m|_{U_1} &= u_1\cdot \left(u_1\partial_{u_1} -\sum_{i=2}^{n} u_i\partial_{u_i} - s\partial_s\right) + \sum_{i=2}^{n} u_1^2u_i^2 \cdot \frac{1}{u_1}\partial_{u_i} + u_1^ms^m \mu(y) \cdot \partial_{y} \\&= u_1 \cdot \underbrace{\left[ u_1\partial_{u_1} +\sum_{i=2}^{n} (-u_i+u_i^2)\partial_{u_i} +u_1^{m-1}s^m\mu(y)\partial_{y} -s\partial_s \right]}_{\psi} \end{align*} $$
Since
$b^*\partial _m|_{U_1}$
is p-closed, so is
$\psi $
. As
$\psi (u_1)=u_1$
, by Lemma 2.3 we obtain that
$\psi ^{[p]}=\psi $
. Thus
$b^*\mathcal {F}_m|_{U_1}=\mathcal {O}_{U_1}\cdot \psi $
is lc by Proposition 2.6. Moreover, the above computations show that E is
$b^*\mathcal {F}_m$
-invariant and that
$a(E;\mathcal {F}_m)=-1$
. The situation on the other
$x_i$
-patches
$U_i$
is similar.
On the t-patch V the blow-up b is given by
$$ \begin{align*} b^*\colon & k[x_1,\dots,x_n,y,t][\mu(y)^{-1}]\longrightarrow k[v_1,\dots,v_n,y,s][\mu(y)^{-1}], \\& x_i\mapsto sv_i \ (1\leq i\leq n), \quad y\mapsto y, \quad t\mapsto s. \end{align*} $$
Therefore we have:
$$ \begin{align*} b^*\partial_m|_V &= \sum_{i=1}^n v_i^2s \partial_{v_i} + s^m\mu(y)\partial_{y} \\&= s\cdot \underbrace{\left[ \sum_{i=1}^{n}v_i^2\partial_{v_i}+s^{m-1}\mu(y)\partial_{y} \right]}_{\xi} \end{align*} $$
Under the description
$V\cong D_{\mathbb {A}^{n+1}_{v_1,\dots ,v_n,y}}(\mu (y))\times \mathbb {A}^1_s$
, we see that
$\xi $
corresponds to
$\partial _{m-1}$
and thus
$b^*\mathcal {F}_m|_V\cong \mathcal {F}_{m-1}$
. To finish the proof, it remains to observe that
$\operatorname {\mathrm {Sing}}(\xi )=V(v_1,\dots ,v_n,s)\subset V$
is in fact a closed subset of X, and that the strict transform of
$(t=0)$
is disjoint from the patch V.
Theorem 4.3. With the notations as above, let
$Y=(\mathbb {A}\times \mathbb {A}^1)/\mathcal {F}_m$
with
$n\geq 2$
and
$m\geq 1$
. Then:
-
1. The projection
$\mathbb {A}\times \mathbb {A}^1_t\to \mathbb {A}^1_t$
factors through a flat morphism
$Y\to \mathbb {A}^1_t$
of relative dimension
$n+1$
; -
2.
$Y_t$
is smooth for
$t\neq 0$
, while
$Y_0$
is reduced and non-
$S_2$
; -
3. If
$n+1\geq \max \{p,3\}$
, the family
$Y\to \mathbb {A}^1$
is locally stable.
The hypothesis
$n\geq 2$
ensures that the
$1$
-foliation
$\mathcal {F}_m$
is a family of
$1$
-foliations over
$\mathbb {A}^1_t$
, see Claim 4.1. The case
$(p=2,n=1)$
is considered in Example 4.6 below.
Proof. Since
$\partial _m(t)=0$
we have
$k[t]\subset k[x_1,\dots ,x_n,y,t]^{\partial _m}$
: this proves the first point. On the principal open set
$D(t)\subset \mathbb {A}\times \mathbb {A}^1$
the
$1$
-foliation
$\mathcal {F}_m|_{D(t)}$
is regular. Thus for
$0\neq \alpha \in k$
, by Theorem 2.12 we have
$$ \begin{align*}Y_\alpha=\mathbb{A}/\mathcal{F}_{m,\alpha} \quad\text{for}\quad \mathcal{F}_{m,\alpha} = \mathcal{O}_{\mathbb{A}}\cdot \left( \sum_{i=1}^{n}x_i^2\partial_{x_i} + \alpha^m\mu(\alpha)\partial_{y} \right).\end{align*} $$
As
$\mathcal {F}_{m,\alpha }$
is regular for
$\alpha \neq 0$
we obtain that
$Y_\alpha $
is smooth [Reference Posva27, Lemma 2.5.10].
Let us prove that
$Y_0$
is not
$S_2$
. For this we reduce to
$m=1$
. Indeed, consider the cartesian diagram

where the vertical arrows are the projections and
$u\colon \mathbb {A}^1_t\to \mathbb {A}^1_{t'}$
is given by
$u^*(t')=t^m$
. By Lemma 2.13 we have
$v^*\mathcal {F}_1=\mathcal {F}_m$
(where
$\mathcal {F}_1$
is defined as in (4.1) using
$t'$
instead of t) and an
$\mathbb {A}^1_t$
-isomorphism
Hence it suffices to show that
$Y_0$
is non-
$S_2$
in the case
$m=1$
. As
$\mathcal {F}_1$
is a family of
$1$
-foliations over
$\mathbb {A}^1_t$
, by Theorem 2.12 it is sufficient to prove that
$$ \begin{align*}\mathcal{O}_{\mathbb{A}\times \mathbb{A}^1,\textbf{0}}^{\partial_1} \otimes_{k[t]} k[t]/(t) \subsetneqq \left(\mathcal{O}_{\mathbb{A},\textbf{0}}/(t)\right)^{\bar{\partial}} \quad \text{with}\quad \bar{\partial}=\sum_{i=1}^{n}x_i^2\partial_{x_i},\end{align*} $$
where
$\textbf {0}=V(x_1,\dots ,x_n,y,t)$
. Clearly
$\bar {\partial }(y)=0$
, so it suffices to show that there is no
$f\in \mathcal {O}_{\mathbb {A}\times \mathbb {A}^1,\textbf {0}}$
such that
$\partial _1(y+tf)=0$
. If there was such an f, we would have
$$ \begin{align*}0=\partial_1(y+tf)=t\mu +\sum_{i=1}^n tx_i^2\partial_{x_i}(f)+t^2\mu \partial_y(f).\end{align*} $$
Taking into account that t is not a zero-divisor in
$\mathcal {O}_{\mathbb {A}\times \mathbb {A}^1,\textbf {0}}$
, we would obtain
$$ \begin{align*}-\mu = \sum_{i=1}^nx_i^2\partial_{x_i}(f)+t\mu \partial_y(f).\end{align*} $$
However, the right-hand side belongs to the prime ideal
$(x_1,\dots ,x_{n},t)$
, so this equality cannot hold in the local ring
$\mathcal {O}_{\mathbb {A}\times \mathbb {A}^1,\textbf {0}}$
by our choice of
$\mu $
. We have reached a contradiction, and hence
$Y_0$
is not
$S_2$
at the image of
$\textbf {0}$
.
On the other hand, I claim that
$Y_0$
is (geometrically) reduced. Since Y is
$S_2$
, the Cartier divisor
$Y_0$
is
$S_1$
and it is sufficient to show that
$Y_0$
is generically reduced. Observe that
$Y_0$
is irreducible. The
$1$
-foliation
$\mathcal {F}_m$
is regular on the open set
$U=D(x_1)$
, which intersects
$Y_0$
nontrivially. By Theorem 2.12 it follows that the open subset
$(U/{\mathcal {F}_m}|_U)_0$
of
$Y_0$
is the quotient of
$U_0$
by (the restriction of)
$\mathcal {F}_{m,0}$
. In particular it is normal, so
$Y_0$
is generically reduced.
It remains to show that
$(Y,Y_0)$
is lc (in the general case
$m\geq 1$
). For this, as in the proof of Theorem 3.2, we blow-up
$\mathbb {A}\times \mathbb {A}^1$
until the pullback of
$\mathcal {F}_m$
has only lc singularities. By Lemma 4.2, this produces a sequence of m smooth blow-ups
such that, if
$b_m\colon X_m\to W$
is the composition of the blow-ups, the
$1$
-foliation
$b_m^*\mathcal {F}_m$
is lc. Let
$E_j\subset X_m$
be the strict transform of the exceptional divisor of
$X_j\to X_{j-1}$
(for
$1\leq j\leq m$
), and denote
$(t=0)\subset W$
by
$W_0$
. We have
$a(E_1;W,W_0)=n-1$
, and using Lemma 4.2 we compute
$$ \begin{align*} a(E_2;W,W_0)=&a(E_2;X_1, (X_1\to W)^{-1}_*W_0 + (1-n)E_1)\\ =& 2n-1. \end{align*} $$
Continuing by induction, we find
Now let
$Z=X_m/b_m^*\mathcal {F}_m$
and consider the induced commutative diagram

where the horizontal arrows are the quotient morphisms and where
$\mathfrak {b}\colon Z\to Y$
is birational. Let
$F_j\subset Z$
be the prime divisor which is the image of
$E_j$
. Using Lemma 4.2 we get
Now we use Theorem 2.10: since
$W_0$
is
$\mathcal {F}_m$
-invariant we find
$$ \begin{align*} a(F_j;Y,Y_0) =& a(E_j; W,W_0)+(p-1)\cdot a(E_j; \mathcal{F}_m) \\ =& j(n-p+1)-1. \end{align*} $$
Therefore we have the crepant equation
$$ \begin{align*}K_Z+\mathfrak{b}^{-1}_*Y_0-\sum_{j=1}^m(j(n-p+1)-1)F_j =\mathfrak{b}^*(K_Y+Y_0).\end{align*} $$
To show that
$(Y,Y_0)$
is lc it suffices to show that
$$ \begin{align*}\operatorname{\mathrm{discrep}}\left(Z,\mathfrak{b}^{-1}_*Y_0-\sum_{j=1}^m(j(n-p+1)-1)F_j\right)\geq -1.\end{align*} $$
Since
$b_m^*\mathcal {F}_m$
is lc, by Theorem 2.10 it suffices to show that
$$ \begin{align*}\operatorname{\mathrm{discrep}}\left(X_m,(b_m)^{-1}_*W_0-\sum_{j=1}^m(j(n-p+1)-1)E_j\right)\geq -1.\end{align*} $$
As
$(X_m,(b_m^{-1})_*W_0+\sum _{j=1}^mE_j)$
is log smooth, this holds by [Reference Kollár16, Corollary 2.11] as soon as
${1-j(n-p+1)\leq 1}$
for every
$j\geq 1$
. This is the case when
$n+1\geq p$
. The proof is complete.
Remark 4.4. The normalization
$Y_0^\nu $
of the
$n+1$
-dimensional central fiber
$Y_0$
in Theorem 4.3 is isomorphic to the product of
$\mathbb {A}^1$
with the n-dimensional quotient singularity of Theorem 3.2. In particular, it is
$\mathbb {Q}$
-factorial,
$S_3$
but non-
$S_4$
, non-F-injective, with a singular locus of dimension one. Since
$n+1\geq \max \{p,3\}$
we have
$n\geq p-1$
and so
$Y_0^\nu $
is lc. If
$n\geq p$
(resp.
$n>p$
), it is even canonical (resp. terminal).
Corollary 4.5. Let
$Y=\mathbb {A}^{n+2}_{x_1,\dots ,x_{n},y,t}/\langle \sum _{i=1}^nx_i^2\partial _{x_i}+t\partial _y\rangle $
with
$n\geq 2$
. If
$\dim Y=n+2\geq p+1$
, then Y has non-
$S_3$
canonical singularities along the image of
$V(x_1,\dots ,x_n,t)$
.
Proof. The computations in the proof of Theorem 4.3 show that the central fiber
$Y_0$
of
$Y\to \mathbb {A}^1_t$
is not
$S_2$
along the image of
$V(x_1,\dots ,x_n,t)$
. Thus Y is non-
$S_3$
along that curve. To show that Y is canonical, as the open set
$Y\setminus Y_0$
is regular and thus terminal, it suffices to show that
$a(E;Y)\geq 0$
for every divisor E over Y with
$c_Y(E)\subseteq Y_0$
. As
$Y_0$
is an effective Cartier divisor and
$(Y,Y_0)$
is lc we have
and we are done.
Example 4.6. Suppose that k has characteristic
$p=2$
, and consider the derivation
$\partial =x^2\partial _x+t\partial _y$
on
$\mathbb {A}^3_{x,y,t}$
. Then
$\partial ^{[2]}=0$
and that
$X=\mathbb {A}^3/\partial $
is a flat
$\mathbb {A}^1_t$
-scheme given by the spectrum of
$k[x^2,y^2,t,tx+x^2y]$
: indeed, the latter ring is clearly
$\partial $
-invariant, is normal and has degree
$4$
over
$k[x^2,y^2,t^2]$
.
It is easily seen that X is isomorphic to an hypersurface in
$\mathbb {A}^4$
, and thus X is Cohen–Macaulay. Its central fiber
$X_0=\operatorname {\mathrm {Spec}} k[x^2,y^2,x^2y]\cong V(W^2-X^2Y)\subset \mathbb {A}^3_{X,Y,W}$
is demi-normal but not normal: it has a nodal singularity along the curve
$(x=0)$
. In fact, this is an inseparable node [Reference Posva28, §3.1], and thus
$X_0$
is semi-normal but not weakly normal [Reference Posva30, Lemma 2.3.7]. Simple computations in the spirit of the proof of Theorem 4.3 show that the pair
$(X,X_0)$
is lc but not plt, as the image of the line
$(x=0=t)$
in X is the singular locus of
$X_0$
and an lc center of
$(X,X_0)$
. Compare also with the general semi-normality results of [Reference Arvidsson and Posva4].
Theorem 4.3 gives a series of local examples. We can compactify them as explained in the next theorem to obtain global ones.
Theorem 4.7. For each
$p>0$
, there exists a projective family
$(\mathcal {Y},\mathcal {B}=\frac {1}{p}\mathcal {H})\to \mathbb {A}^1$
of relative dimension
$N\geq \max \{p,3\}$
such that:
-
1.
$(\mathcal {Y},\frac {1}{p}\mathcal {H})\to \mathbb {A}^1$
is a stable family, -
2.
$\mathcal {Y}_t$
has only
$\mu _p$
-quotient singularities for
$t\neq 0$
, and -
3.
$\mathcal {Y}_0$
is not
$S_2$
.
Moreover:
-
• There exists a finite purely inseparable morphism
$\mathbb {P}^{N-1}\times E\times \mathbb {A}^1\to \mathcal {Y}$
of degree p over
$\mathbb {A}^1$
, where E is a supersingular elliptic curve, and
$\mathcal {H}$
is the image of a
$\mathbb {Q}$
-divisor
$H_0\times \mathbb {A}^1\subset \mathbb {P}^{N-1}\times E\times \mathbb {A}^1$
; -
• If
$p=2$
then we can modify the construction so that
$\mathcal {Y}_t$
is regular for every
$t\neq 0$
; -
• If C is a normal curve and
$C\to \mathbb {A}^1$
is a finite flat morphism, then
$\mathcal {Y}_C=\mathcal {Y}\times _{\mathbb {A}^1}C$
is normal and
$(\mathcal {Y}_C, \frac {1}{p}\mathcal {H}_C)\to C$
is stable.
The following lemma will be useful for the proof.
Lemma 4.8. Consider E a supersingular elliptic curve,
$\omega \in H^0(E,T_E)$
a global generator,
$\mathbb {A}^{N-1}_{x_1,\dots ,x_{N-1}}\subset \mathbb {P}^{N-1}$
a standard chart (with
$N\geq 3$
), A an integral Noetherian k-algebra and
$a\in A$
. Then the derivation
$$ \begin{align*}\partial_{(0)}=\sum_{i=1}^{N-1}x_i^2\partial_{x_i} +a\omega \quad \text{on} \quad \mathbb{A}^{N-1}\times E\times \operatorname{\mathrm{Spec}}(A)\end{align*} $$
induces a
$1$
-foliation
$\mathcal {G}$
of rank
$1$
on
$\mathbb {P}^{N-1}\times E\times \operatorname {\mathrm {Spec}}(A)$
such that:
-
1.
$K_{\mathcal {G}}=\mathcal {O}_{\mathbb {P}^{N-1}}(1)\boxtimes \mathcal {O}_{E\times \operatorname {\mathrm {Spec}}(A)}$
, -
2.
$\mathcal {G}$
has only multiplicative singularities away from
$\{[1:0:\dots :0]\}\times E \times V(a)$
, -
3. If
$A=k[t]$
and
$a=t^m$
with
$m\geq 1$
, then
$\mathcal {G}\subset T_{(\mathbb {P}^{N-1}\times E\times \mathbb {A}^1)/\mathbb {A}^1}$
is a family of
$1$
-foliations.
Proof. Since E is supersingular we have
$\omega ^{[p]}=0$
[Reference Katz and Mazur14, 12.4.1.3]. An argument similar to Claim 3.1 shows that
$\partial _{(0)}$
is p-closed. Therefore it induces a
$1$
-foliation
$\mathcal {G}$
of rank
$1$
on
$\mathbb {P}^{N-1}\times E\times \operatorname {\mathrm {Spec}}(A)$
. We compute the expression of
$\partial _{(0)}$
on the other charts. Let
$X_0,\dots ,X_{N-1}$
be homogeneous coordinates of
$\mathbb {P}^{N-1}$
: we may assume that
$x_i=X_i/X_0$
for
$1\leq i\leq N-1$
. If
$u_i=X_i/X_{N-1}$
for
$0\leq i<N-1$
, then we have
$$ \begin{align*} x_i&=\frac{u_i}{u_0} \ (1\leq i<N-1), \quad x_{N-1}=\frac{1}{u_0},\\u_i&=\frac{x_i}{x_{N-1}} \ (1\leq i< N-1), \quad u_0=\frac{1}{x_{N-1}}. \end{align*} $$
Therefore
$$ \begin{align*}\partial_{x_i}=u_0\partial_{u_i} \ (1\leq i<N-1),\quad \partial_{x_{N-1}}=-u_0^2\partial_{u_0}-\sum_{i=1}^{N-2}u_0u_i\partial_{u_i}\end{align*} $$
and so
$$ \begin{align} \partial_{(0)}=\frac{-1}{u_0}\cdot \underbrace{\left[ u_0\partial_{u_0}+\sum_{i=1}^{N-2}(u_i-u_i^2)\partial_{u_i} - au_0\omega \right]}_{\partial_{(N-1)}}. \end{align} $$
This shows that
$\mathcal {G}|_{D_+(X_{N-1})\times E\times \operatorname {\mathrm {Spec}}(A)}$
is generated by
$\partial _{(N-1)}$
and that
$K_{\mathcal {G}}^{-1}=\mathcal {G}\cong \mathcal {O}_{\mathbb {P}^{N-1}}(-1)\boxtimes \mathcal {O}_{E\times \operatorname {\mathrm {Spec}}(A)}$
. This proves the first assertion.
On
$D_+(X_{N-1})\times E\times \mathbb {A}^1$
, notice that
$\partial _{(N-1)}$
is p-closed and that
$\partial _{(N-1)}(u_0)=u_0$
. Therefore
$\partial _{(N-1)}^{[p]}=\partial _{(N-1)}$
by Lemma 2.3, which shows that
$\mathcal {G}$
has only multiplicative singularities on
$D_+(X_{N-1})\times E\times \mathbb {A}^1$
. The computations on the patches
$D_+(X_i)\times E\times \mathbb {A}^1$
for
$1\leq i<N-1$
are similar. On
$D_+(X_0)\times E\times \mathbb {A}^1$
, we see that the singular locus of
$\mathcal {G}$
is
$\{[1:0:\dots : 0]\}\times E\times V(a)$
. This proves the second assertion. In case
$A=k[t]$
and
$a=t^m$
, we proceed as in Claim 4.1 over each chart to see that
$\mathcal {G}$
is a family of
$1$
-foliations if and only if
$N\geq 3$
, which holds by assumption.
Remark 4.9. The multiplicative singularities of
$\mathcal {G}$
cannot be resolved by smooth blow-ups (unless
$p=2$
). Indeed, the singular locus of
$\partial _{(N-1)}$
is the union of the closed subsets
where
$\varepsilon _1,\dots ,\varepsilon _{N-2}\in \{0,1\}$
. It follows from [Reference Posva27, Proposition 4.1.1] that along
$Z_{\varepsilon _\bullet }$
, and formal-locally on its first factor, the
$1$
-foliation
$\mathcal {G}$
is generated by
$$ \begin{align*}\sum_{i: \ \varepsilon_i=0} \bar{u}_i\partial_{\bar{u}_i} -\sum_{j: \ \varepsilon_j=1} \bar{u}_j\partial_{\bar{u}_j} -a\bar{u}_0\omega\end{align*} $$
where
$\bar {u}_\bullet $
are formal parameters of
$\mathcal {O}_{\mathbb {P}^{N-1},[\varepsilon _\bullet :1]}$
with
$\bar {u}_0=u_0$
. To resolve the singularity of
$\mathcal {G}$
along
$Z_{\varepsilon _\bullet }$
, we need to take the weighted blow-up given by the Rees algebra
$$ \begin{align*}\sum_{i: \ \varepsilon_i=0} (\bar{u}_i,1)+ \sum_{j: \ \varepsilon_j=1} (\bar{u}_j, p-1).\end{align*} $$
As soon as some weight
$p-1\neq 1$
appears, this blow-up produces a nonschematic tame regular Deligne–Mumford stack whose coarse moduli space has quotient singularities: see [Reference Posva29] for details. So the singularities of
$\mathcal {G}$
along
$Z_{\varepsilon _\bullet }$
cannot be resolved by usual blow-ups, unless
$\varepsilon _\bullet =(0,\dots ,0)$
or
$p=2$
. The situation on the other patches
$D_+(X_i)\times E\times \operatorname {\mathrm {Spec}}(A)$
, for
$1\leq i<N-1$
, is similar.
Proof of Theorem 4.7
For clarity, we divide the proof into several steps. If
$p=2$
we can use a slightly different construction, which we describe in Step 5.
Step 1: Construction of
$\mathcal {Y}\to \mathbb {A}^1$
. Let E be a supersingular elliptic curve and
$\omega \in H^0(E,T_E)$
be a global generator. If
$X_0,\dots ,X_{N-1}$
are the homogeneous coordinates on
$\mathbb {P}^{N-1}$
, consider on
$D_+(X_0)\times E\times \mathbb {A}^1_t$
the derivation
$$ \begin{align*}\partial_{(0)}=\sum_{i=1}^{N-1}x_i^2\partial_{x_i}+t\omega, \quad x_i=X_i/X_0.\end{align*} $$
By Lemma 4.8 it induces a
$1$
-foliation
$\mathcal {G}$
on
$W=\mathbb {P}^{N-1}\times E\times \mathbb {A}^1$
. We let
$\mathcal {Y}=(\mathbb {P}^{N-1}\times E\times \mathbb {A}^1)/\mathcal {G}$
, with quotient morphism
$q\colon W\to \mathcal {Y}$
. By construction the projection
$W\to \mathbb {A}^1$
factors through a flat projective morphism
$\mathcal {Y}\to \mathbb {A}^1$
.
Step 2: Construction of the divisor
$\mathcal {H}$
. The divisor
$\mathcal {H}$
is chosen as follows. Take a
$\mathbb {Q}$
-divisor H on W such that:
-
•
$H=H_0\times \mathbb {A}^1$
for some
$\mathbb {Q}$
-divisor
$H_0$
on
$\mathbb {P}^{N-1}\times E$
; -
•
$K_W+(p-1)K_{\mathcal {G}}+H$
is ample over
$\mathbb {A}^1$
, -
•
$(W,W_\alpha +H)$
is lc for every
$\alpha \in \mathbb {A}^1\setminus \{0\}$
, -
• the support of H lies in general position with respect to the curve
$\{[1:0:\dots : 0]\}\times E\times \{0\}$
, and -
• H is not
$\mathcal {G}$
-invariant.
The last three properties are satisfied for a general choice of divisor of the form
$H_0\times \mathbb {A}^1$
, thus by choosing
$H_0$
general in a sufficiently ample linear system on
$\mathbb {P}^{N-1}$
, we see there exists an H with all these properties. Notice that
$K_W+(p-1)K_{\mathcal {G}}\cong \mathcal {O}_{\mathbb {P}^{N-1}}(-N+p-1)\boxtimes \mathcal {O}_{E\times \mathbb {A}^1}$
by Lemma 4.8, so because of the second property and on the assumption
$N\geq \max \{p,3\}$
we cannot take
$H=0$
.
We let
$\mathcal {H}\subset \mathcal {Y}$
be the prime divisor with support
$q(H)$
.
Step 3: Behaviour under base-change. Let us immediately observe what happens when we base-change
$\mathcal {Y}\to \mathbb {A}^1_t$
along a flat finite morphism
$g\colon C\to \mathbb {A}^1_t$
. By Lemma 2.13 we have a natural isomorphism
where
$f\colon \mathbb {P}^{N-1}\times E\times C\to \mathbb {P}^{N-1}\times E\times \mathbb {A}^1$
is the base-change of g. This shows that
$\mathcal {Y}_C$
is normal. Observe that
$H_C=f^*H$
still satisfies the five properties listed in Step 2. The pullback of
$K_{\mathcal {Y}}+\frac {1}{p}\mathcal {H}$
along
$\mathcal {Y}_C\to \mathcal {Y}$
is equal to
$K_{\mathcal {Y}_C}+\frac {1}{p}\mathcal {H}_C$
: therefore the latter is ample over C as soon as
$K_{\mathcal {Y}}+\frac {1}{p}\mathcal {H}$
is ample over
$\mathbb {A}^1$
. By Proposition 2.9 and the non-
$\mathcal {G}$
-invariance of H we have
$$ \begin{align} q^*\left(K_{\mathcal{Y}}+\frac{1}{p}\mathcal{H}\right)=K_W+(p-1)K_{\mathcal{G}}+H \end{align} $$
which, by choice of H, is ample over
$\mathbb {A}^1$
.
To complete the proof, it remains to check properties (2) and (3), and that
$(\mathcal {Y}_C, \frac {1}{p}\mathcal {H}_C)\to C$
is locally stable. Properties (2) and (3) can be checked after an arbitrary surjective base-change. For local stability, by [Reference Kollár18, 2.15.5] and [Reference Hu and Zong12] it is enough to check it in case
$g=F^r_{\mathbb {A}^1/k}$
is an iterate of the k-linear Frobenius morphism of
$\mathbb {A}^1$
(Footnote
5
).
Step 4: Local stability and singularities of fibers. To summarize, we have reduced the proof to the following: let
$q'\colon W=\mathbb {P}^{N-1}\times E\times \mathbb {A}^1_t\to \mathcal {Y}'$
be the quotient by the
$1$
-foliation
$\mathcal {G}'$
induced by the derivation
$$ \begin{align*}\partial_{(0)}'=\sum_{i=1}^{N-1}x_i^2\partial_{x_i}+t^{p^r}\omega, \quad \text{on}\quad D_+(X_0)\times E\times \mathbb{A}^1_t.\end{align*} $$
It is endowed with a flat proper morphism
$\mathcal {Y}'\to \mathbb {A}^1_t$
. Let
$H'$
be a
$\mathbb {Q}$
-divisor on W satisfying the properties listed in Step 2: we let
$\mathcal {H}'=q(H')$
. We must show that
$(\mathcal {Y}',\frac {1}{p}\mathcal {H}')\to \mathbb {A}^1$
is locally stable, that
$\mathcal {Y}^{\prime }_t$
has only
$\mu _p$
-quotient singularities for
$t\neq 0$
, and that
$\mathcal {Y}_0'$
is not
$S_2$
. As we will see, everything is a consequence of Theorem 4.3 and Lemma 4.8.
Step 4.1: Singularity of the central fiber. First, let us prove that
$\mathcal {Y}_0'$
is not
$S_2$
along the image through
$q'$
of the proper curve
$\{[1:0:\dots :0]\}\times E\times \{0\}$
. Indeed, let
$z=\{[1:0:\dots :0]\}\times \{e\}\times \{0\}$
where e is an arbitrary closed point of E. As
$\mathcal {Y}_0'$
is a Cartier divisor in
$\mathcal {Y}$
, it is equivalent to showing that
$\mathcal {O}_{\mathcal {Y}',q'(z)}$
is not
$S_3$
. I claim that this is étale-local over z. For if
$h\colon \operatorname {\mathrm {Spec}}(\mathcal {O}_{W,z}^h)\to \operatorname {\mathrm {Spec}}(\mathcal {O}_{W,z})$
is the henselization, by [Reference Posva29, Lemma 3.2.1] we have a cartesian diagram

where g is an étale morphism. Therefore it suffices to prove that
$\operatorname {\mathrm {Spec}}(\mathcal {O}_{W,z}^h)/h^*\mathcal {G}'$
is not
$S_3$
. If
$y\in \mathcal {O}_{E,e}^h$
is an étale coordinate at
$e\in E$
, then an étale-local generator for
$\mathcal {G}'$
at z is given by
$$ \begin{align*}\partial^h=\sum_{i=1}^{p}x_i^2\partial_{x_i} + t^{p^r} \mu(y) \partial_y\end{align*} $$
where
$\mu (y)\in \mathcal {O}_{E,e}^h$
is a unit such that
$\omega \otimes \mathcal {O}_{E,e}^h=\mu (y)\partial _y$
. Then we apply Theorem 4.3 (or rather its formal-local version at the origin, which is proved similarly) to see that the
$\partial ^h$
-invariant sub-ring of
$\mathcal {O}_{W,z}^h$
is not
$S_3$
.
Since
$\mathcal {Y}'$
is
$S_2$
, the irreducible Cartier divisor
$\mathcal {Y}_0'$
is nonetheless
$S_1$
. As it is generically reduced by Theorem 4.3, we obtain that
$\mathcal {Y}_0'$
is (geometrically) reduced.
Step 4.2: Local stability. Now we prove that
$(\mathcal {Y}',\frac {1}{p}\mathcal {H}')\to \mathbb {A}^1$
is locally stable. First, let
$0\neq \alpha \in k$
. As
$\mathcal {G}'|_{D(t)}$
is lc and
$(W,W_\alpha +H')$
is lc, by Theorem 2.10 we have that
$(\mathcal {Y}',\mathcal {Y}^{\prime }_\alpha +\frac {1}{p}\mathcal {H}')$
is lc along
$\mathcal {Y}_\alpha $
. Moreover, as
$\mathcal {G}'|_{D(t)}$
has only multiplicative singularities by Lemma 4.8, by Theorem 2.12 we obtain that
where
$\mathcal {G}^{\prime }_\alpha $
is the
$1$
-foliation induced by
$\sum _{i=1}^{N-1}x_i^2\partial _{x_i}+\alpha ^{p^r}\omega $
. By Lemma 4.8 we see that
$\mathcal {G}^{\prime }_\alpha $
has only multiplicative singularities. Therefore
$\mathcal {Y}_\alpha $
has only
$\mu _p$
-quotient singularities by [Reference Posva27, Proposition 4.1.1].
It remains to prove that
$(\mathcal {Y}',\mathcal {Y}^{\prime }_0+\frac {1}{p}\mathcal {H}')$
is lc. To this end, we blow-up W to simplify the singularities of
$\mathcal {G}'$
. By Lemma 4.8 the non-lc singularities of
$\mathcal {G}'$
are contained on the chart
$D_+(X_0)\times E\times \mathbb {A}^1$
, where (étale-locally on the factor E) the
$1$
-foliation
$\mathcal {G}'$
is generated by a derivation of the form (4.1) with
$m=p^r$
. Therefore, the proofs of Lemma 4.2 and Theorem 4.3 show that if we blow-up repeatedly the locus of non-lc singularities of
$\mathcal {G}'$
, we obtain a birational proper morphism
with
$p^r$
exceptional divisors
$E_1,\dots ,E_{p^r}\subset X$
, such that
$b^*\mathcal {G}'$
is lc,
$(X,b_*^{-1}W_0+\sum _j E_j)$
is log smooth, each
$E_j$
is
$b^*\mathcal {G}'$
-invariant,
$a(E_j;\mathcal {G}')=-1$
and
$a(E_j;W,W_0)=j(N-1)-1$
. By our choice of
$H'$
, we also have that
$(X,b_*^{-1}(W_0+H')+\sum _jE_j)$
is log smooth and
The arguments are now the same as the ones at the end of the proof of Theorem 4.3. Consider the commutative diagram

If
$F_1,\dots ,F_{p^r}\subset Z$
denote the
$\mathfrak {b}$
-exceptional divisors, then by Theorem 2.10 we have
$$ \begin{align*} a\left(F_j; \mathcal{Y}', \mathcal{Y}_0'+\frac{1}{p}\mathcal{H}'\right) &= a(E_j; W, W_0+H')+(p-1)\cdot a(E_j;\mathcal{G}') \\&= a(E_j; W, W_0) +(p-1)(-j) \\&= j(N-p)-1 \end{align*} $$
Hence we have the crepant equation
$$ \begin{align*}K_Z+\mathfrak{b}^{-1}_*\left(\mathcal{Y}_0'+\frac{1}{p}\mathcal{H}'\right)+\sum_{j=1}^{p^r}(1-j(N-p))F_j =\mathfrak{b}^*\left(K_{\mathcal{Y}'}+\mathcal{Y}^{\prime}_0+\frac{1}{p}\mathcal{H}'\right).\end{align*} $$
Notice that
$1-j(N-p)\leq 1$
for every
$j\geq 1$
is equivalent to
$N\geq p$
, which holds as
$N=\max \{p,3\}$
. Therefore we have that
$$ \begin{align*}\operatorname{\mathrm{discrep}}\left(X,b^{-1}_*(W_0+H)+\sum_{j=1}^{p^r}(1-j(N-p))E_j\right)\geq -1.\end{align*} $$
Hence by Theorem 2.10 we obtain that
$$ \begin{align*}\operatorname{\mathrm{discrep}}\left(Z,\mathfrak{b}^{-1}_*(\mathcal{Y}_0+\frac{1}{p}\mathcal{H})+\sum_{j=1}^{p^r}(1-j(N-p))F_j\right)\geq -1,\end{align*} $$
and therefore
$(\mathcal {Y}',\mathcal {Y}^{\prime }_0+\frac {1}{p}\mathcal {H}')$
is lc.
Step 5: the case
$p=2$
. It follows from Remark 4.9 that there exists a smooth blow-up
$Z\to \mathbb {P}^{N-1}$
whose (reducible) center does not contain
$[1:0:\dots :0]$
, and such that if
$b\colon Z\times E\times \mathbb {A}^1\to W$
is the induced morphism, the birational pullback
$b^*\mathcal {G}$
is regular away from the preimage of
$\{[1:0: \dots : 0]\}\times E\times \{0\}$
. We take
as the example. By [Reference Posva27, Proposition 5.2.4, Lemma 2.5.10] we have that
$\mathcal {Z}_\alpha $
is regular for
$\alpha \neq 0$
. The rest of the properties are proved exactly as above. This completes the proof.
Let us make some observations about the preceding proof.
Remark 4.10. The coefficients of H are
$\leq 1$
, so the coefficients of
$\frac {1}{p}\mathcal {H}$
are
$\leq 1/p$
. The extra factor
$1/p$
appears because H is chosen to be non-
$\mathcal {G}$
-invariant. I do not know whether there exists a
$\mathcal {G}$
-invariant divisor H that satisfies the other properties. On the other hand, the number of irreducible components of H can be any (in particular,
$\mathcal {H}$
can be irreducible).
Remark 4.11. If
$e\in E$
an arbitrary point then the line bundle
$\mathcal {L}=\mathcal {O}_{\mathbb {P}^{N-1}}(N+p)\boxtimes \mathcal {O}_E(e)$
is ample on
$\mathbb {P}^{N-1}\times E$
, and
$\mathcal {L}^{\otimes n}$
is very ample for every
$n\geq 3$
. Moreover, for any integer
$m\geq 1$
, if
$H_0$
is a general element of the
$\mathbb {Q}$
-linear system
$\frac {1}{m}|\mathcal {L}^{\otimes n}|$
then
$H_0\times \mathbb {A}^1$
satisfies the conditions listed in Step 2.
4.2 Second example:
$\mathbf {\dim =3, p=3}.$
Consider the affine space
$\mathbb {A}^3_{x,y,z}\times \mathbb {A}^1_t$
and the derivation
Using an argument as in Claim 3.1 together with [Reference Mitsui and Sato24, Corollary 6], one checks that
$\partial _m^{[p]} =0$
as soon as
$p>2$
. In that case
$\partial _m$
generates a
$1$
-foliation
$\mathcal {F}_m$
on
$\mathbb {A}^3\times \mathbb {A}^1$
. As in Claim 4.1, one sees that
$\mathcal {F}_m$
is a family of
$1$
-foliations over
$\mathbb {A}^1_t$
.
Let us show how to simplify the singularities of
$\mathcal {F}_m$
. As in Lemma 4.2, we are able to reduce
$\mathcal {F}_m$
to
$\mathcal {F}_{m-1}$
, but we need two blow-ups to achieve this.
Claim 4.12. For
$p>2$
, we have
$((1-2x^2)\partial _x)^{[p]}=2^{(5p-3)/2} (1-2x^2)\partial _x$
.
Proof. After the change of variable
$u=1+\sqrt {2}x$
, we reduce to consider the p-th power of the derivation
$\partial =\sqrt {2}u(2-u)\partial _u$
on
$k[u]$
. By a formula of Jacobson [Reference Jacobson13, p. 209], we have
where S is a k-linear combination of
$(p-1)$
-fold commutators of
$-u^2\partial _x$
and
$2u\partial _u$
. Now we make several observations. First, by Hochschild’s formula
$\partial $
is p-closed and so
$\partial ^{[p]}$
is a
$k[u]$
-scaling of
$\partial $
. As
$(-u^2\partial _u)^{[p]}=0$
and
$(2u\partial _u)^{[p]}=2^{p}u\partial _u$
, it follows that
$S\neq 0$
. As
we see that
$S\neq 0$
can only be of the form
$-\alpha u^2\partial _u$
for some
$\alpha \in k^\times $
. Hence
$\partial ^{[p]}=2^{p/2}(2^pu\partial _u-\alpha u^2\partial _u)$
is a
$k[u]$
-scaling of
$\partial =2^{1/2}(2u\partial _u-u^2\partial _u)$
, which forces
$\alpha =2^{(5p-3)/2}$
.
Lemma 4.13. Suppose that
$p>2$
. We have
$\operatorname {\mathrm {Sing}}(\mathcal {F}_m)=V(x,y,t)$
for
$m\geq 1$
, and
$\mathcal {F}_0$
is regular. If
$m\geq 1$
and
with exceptional divisor E, then:
-
1.
$a(E;\mathcal {F}_m)=0$
, -
2. E is
$b_1^*\mathcal {F}_m$
-invariant, -
3. the locus S of (non-lc) singularities of
$b_1^*\mathcal {F}_m$
is closed,
$\operatorname {\mathrm {mult}}_{S}E=1$
,
$\operatorname {\mathrm {codim}}_YS=2$
and S is not contained in
$(b_1)^{-1}_*(t=0)$
.
Consider
$b_2\colon \operatorname {\mathrm {Bl}}_SY\longrightarrow Y$
with exceptional divisor F and
$b=b_1\circ b_2$
. Then:
-
1.
$a(F;\mathcal {F}_m)=-1$
, -
2. F is
$b^*\mathcal {F}_m$
-invariant, -
3. if
$m\geq 2$
, the non-lc locus of
$b^*\mathcal {F}_m$
is closed, contained in a blow-up chart V of
$b_2$
, not contained in the support of
$b_*^{-1}(t=0)+(b_2)^{-1}_*E$
, and -
4. we have an
$\mathbb {A}^1_t$
-isomorphism
$V\cong \mathbb {A}^3\times \mathbb {A}^1_t$
under which
$b^*\mathcal {F}_m|_V$
is isomorphic to
$\mathcal {F}_{m-1}$
, and the
$b_2$
-exceptional divisor corresponds to
$(t=0)$
.
Proof. Let us first consider the blow-up
$b_1$
. We have three blow-up charts, which we denote by
$U_{A}$
,
$U_B$
, and
$U_C$
:
-
• Chart
$U_A$
, given by
$x\mapsto u, \ y\mapsto uv, \ z\mapsto z, \ t\mapsto us$
. Then which shows that
$$ \begin{align*}b_1^*\partial_m|_{U_A}= u^3v^3\partial_u+(1-u^2v^4)\partial_v-u^2v^3s\partial_s+u^ms^m\partial_z\end{align*} $$
$b_1^*\mathcal {F}_m|_{U_A}$
is a regular
$1$
-foliation. Notice that
$b_1^*\partial _m(u)\in (u)$
, so E is
$b_1^*\mathcal {F}_m$
-invariant. Since
$b_1^*\partial _m|_{U_A}$
generates
$b^*\mathcal {F}_m|_{U_A}$
, without having to saturate, we have
$a(E;\mathcal {F}_m)=0$
.
-
• Chart
$U_B$
, given by
$x\mapsto uv, \ y\mapsto v, \ z\mapsto z, \ t\mapsto vs$
. Then which is singular along the plane
$$ \begin{align*}b_1^*\partial_m|_{U_B}= (v^2-u^2)\partial_u+uv\partial_v-us\partial_s+v^ms^m\partial_z\end{align*} $$
$u=v=0$
. These singularities are non-lc.
-
• Chart
$U_C$
, given by
$x\mapsto us, \ y\mapsto vs, \ z\mapsto z, \ t\mapsto s$
. Then which is singular along the plane
$$ \begin{align*}b_1^*\partial_m|_{U_C}= v^3s^2\partial_u+u\partial_v+s^m\partial_z\end{align*} $$
$u=s=0$
. These singularities are non-lc.
We see that the locus
$S\subset Y$
of singularities of
$b_1^*\mathcal {F}_m$
is smooth, closed, and irreducible of codimension
$2$
. As can be checked on chart
$U_B$
, that singular locus is not contained in the strict transform
$b_*^{-1}(t=0)$
.
Next let us blow-up S. We have two blow-up charts above
$U_B$
(resp. above
$U_C$
), which we denote by
$U_{BA}$
and
$U_{BB}$
(resp. by
$U_{CA}$
and
$U_{CB}$
):
-
• Chart
$U_{BA}$
, given by
$u\mapsto \bar {u},\ v\mapsto \bar {u}\bar {v}, \ z\mapsto z, \ s\mapsto s$
. Then As
$$ \begin{align*}b^*\partial_m|_{U_{BA}}= \bar{u} \cdot \underbrace{\left[ \bar{u}(\bar{v}^2-1)\partial_{\bar{u}} +\bar{v}(2-\bar{v}^2)\partial_{\bar{v}} -s\partial_s + \bar{u}^{m-1}\bar{v}^ms^m\partial_z\right]}_{\psi_{BA}}.\end{align*} $$
$\psi _{BA}(s)=-s$
, by Lemma 2.3 we see that
$\psi _{BA}^{[p]}=-\psi _{BA}$
. Thus
$b^*\mathcal {F}_m$
is lc on the chart
$U_{BA}$
. As
$\psi _{BA}(\bar {u})\in (\bar {u})$
we see that F is
$b^*\mathcal {F}_m$
-invariant. As
$b^*\partial _m=\bar {u}\cdot \psi _{BA}$
we find that
$a(F;\mathcal {F}_m)=-1$
.
-
• Chart
$U_{BB}$
, given by
$u\mapsto \bar {u}\bar {v},\ v\mapsto \bar {v}, \ z\mapsto z, \ s\mapsto s$
. Then By Claim 4.12 we see that
$$ \begin{align*}b^*\partial_m|_{U_{BB}}= \bar{v} \cdot \underbrace{\left[ (1-2\bar{u}^2)\partial_{\bar{u}} +\bar{u}\bar{v}\partial_{\bar{v}} -\bar{u}s\partial_s + \bar{v}^{m-1}s^,\partial_z\right]}_{\psi_{BB}}.\end{align*} $$
$\psi _{BB}(u)=2^{(5p-3)/2}\psi _{BB}(u)$
. Thus by Proposition 2.6 we see that
$b^*\mathcal {F}_m$
is lc on the chart
$U_{BB}$
.
-
• Chart
$U_{CA}$
, given by
$u\mapsto \bar {u},\ v\mapsto v, \ z\mapsto z, \ s\mapsto \bar {u}\bar {s}$
. Then As
$$ \begin{align*}b^*\partial_m|_{U_{CA}}= \bar{u} \cdot \underbrace{\left[ \bar{u}v^3\bar{s}^2\partial_{\bar{u}} +\partial_{v} +\bar{u}^{m-1}\bar{s}^m\partial_z - v^3\bar{s}^3\partial_{\bar{s}}\right]}_{\psi_{CA}}.\end{align*} $$
$\psi _{CA}$
generates a regular
$1$
-foliation, we see that
$b^*\mathcal {F}_m$
is regular on the chart
$U_{CA}$
.
-
• Chart
$U_{CB}$
, given by
$u\mapsto \bar {u}\bar {s},\ v\mapsto v, \ z\mapsto z, \ s\mapsto \bar {s}$
. Then Under the
$$ \begin{align*}b^*\partial_m|_{U_{CB}}= \bar{s} \cdot \underbrace{\left[ v^3\partial_{\bar{u}} +\bar{u}\partial_{v} +\bar{s}^{m-1}\partial_z\right]}_{\psi_{CB}}.\end{align*} $$
$\mathbb {A}^1_t$
-isomorphism
$U_{CB}\cong \mathbb {A}^3_{x,y,z}\times \mathbb {A}^1_t$
given by
$(\bar {u},\bar {v},z,\bar {s})\mapsto (x,y,z,t)$
, we see that
$\psi _{CB}$
corresponds to
$\partial _{m-1}$
, and thus
$b^*\mathcal {F}_m|_{U_{CB}}$
corresponds to
$\mathcal {F}_{m-1}$
.
The above computations show that the non-lc locus of
$b^*\mathcal {F}_m$
is contained in
$U_{CB}$
. It is elementary (but tedious) to check that this locus does not intersect the other patches
$U_{CA}$
,
$U_{BA}$
, and
$U_{BB}$
. It remains to observe that the strict transforms of
$E\subset Y$
and of
$(t=0)\subset \mathbb {A}^3\times \mathbb {A}^1_t$
are disjoint from the chart
$U_{CB}$
.
Theorem 4.14. With the notations as above, assume that
$p=3$
and let
$Y=(\mathbb {A}^3\times \mathbb {A}^1)/\mathcal {F}_1$
. Then:
-
1. The projection
$\mathbb {A}^3\times \mathbb {A}^1_t\to \mathbb {A}^1_t$
factors through a flat morphism
$Y\to \mathbb {A}^1_t$
of relative dimension
$3$
; -
2.
$Y_t$
is smooth for
$t\neq 0$
, while
$Y_0$
is reduced and non-
$S_2$
; -
3.
$Y\to \mathbb {A}^1$
is locally stable, and Y has canonical non-
$S_3$
singularities along
$Y_0$
.
Moreover, if C is a normal curve and
$C\to \mathbb {A}^1_t$
is a finite flat morphism, then
$Y\times _{\mathbb {A}^1_t}C$
is normal and
$Y_C\to C$
is locally stable.
Proof. The first item is clear. For the remaining ones, as in the proof of Corollary 4.5 and Theorem 4.7, we reduce the proof to the following: if
$Y'=(\mathbb {A}^3\times \mathbb {A}^1)/\mathcal {F}_{p^r}$
then
$Y'\to \mathbb {A}^1_t$
is locally stable, has smooth fibers above
$t\neq 0$
and a non-
$S_2$
reduced central fiber
$Y^{\prime }_0$
. The statements about the singularities of the fibers are proved as in Theorem 4.3 (and are true whenever
$p>2$
). It remains to check that
$Y'\to \mathbb {A}^1$
is locally stable above
$t=0$
.
As usual by now, we blow-up
$W=\mathbb {A}^3\times \mathbb {A}^1_t$
. By Lemma 4.13, there is a sequence of smooth blow-ups
with exceptional divisors
$E_i,F_i\subset X$
for
$i=1,\dots ,p^r$
such that:
$(X,\sum _i E_i+F_i)$
is log smooth,
$b^*\mathcal {F}_{p^r}$
has only lc singularities, the
$E_i$
’s and the
$F_i$
’s are
$b^*\mathcal {F}_{p^r}$
-invariant, and
Using Lemma 4.13 and proceeding by induction, one computes
Now let
$Z=X/b^*\mathcal {F}_{p^r}$
and denote by
$E^{\prime }_i\subset Z$
(resp. by
$F^{\prime }_i\subset Z$
) the image of
$E_i$
(resp. of
$F_i$
). Then by Lemma 4.13 and Theorem 2.10 we have
and
As
$b^*\mathcal {F}_{p^r}$
is lc, by Theorem 2.10 it suffices to ensure that
$$ \begin{align*}\operatorname{\mathrm{discrep}}\left(Z, -\sum_i a(E_i';Y',Y_0')E_i'-\sum_i a(F_i';Y',Y_0')F_i'\right)\geq -1\end{align*} $$
to obtain that
$(Y',Y_0')$
is lc. This holds as soon as
$i(4-p)\geq 0$
for every
$i=1,\dots ,p^r$
, which happens if (and only if)
$p=3$
, so the proof is complete.
Remark 4.15. One can try to compactify the examples of Theorem 4.14 using the method of Theorem 4.7. This is quite delicate, as the
$1$
-foliation
$\psi =y^3\partial _x+x\partial _y$
acquires complicated singularities when we compactify. For example, if we regard
$\mathbb {A}^2_{x,y}$
as a standard chart of
$\mathbb {P}^2$
, then on the chart with coordinates
$u=1/x, v=y/x$
the
$1$
-foliation induced by
$\psi $
is generated by
$uv^3\partial _u+(v^4-u^2)\partial _v$
. Similarly, if we regard
$\mathbb {A}^2_{x,y}$
has a standard chart of the Hirzebruch surface
$F_n$
, then on the chart with coordinates
$u=1/y, v=1/(xy^n)$
the
$1$
-foliation induced by
$\psi $
is generated by
$u^{2n+5}\partial _u+(v^3+nu^{2n+4}v)\partial _v$
.
A Appendix: KSBA moduli stacks in positive characteristics
In this appendix we explore the consequences of Theorem 1.3 for KSBA moduli theory. The point is to show that the most straightforward adaptations of the characteristic
$0$
definitions do not yield satisfactory theories.
As we will see, the point is the unavoidable appearance of non-
$S_2$
fibers as limits of stable families over punctured curves. To formulate this in a precise way we need to define which moduli stacks we are working with (or, equivalently, which families we allow in our theory), and we face two difficulties:
-
1. Our examples involve boundary divisors with small coefficients, and defining stable families of pairs with such coefficients over general bases is already extremely delicate in characteristic
$0$
; -
2. The algebraicity of the stacks we could write down is an open question (Footnote 6 ), except in the surface case where it should follow from [Reference Hacon and Kovács11].
We go around the first difficulty by using a minimalistic, underdetermined, working definition of the KSBA moduli stacks. We will ignore the second difficulty by imposing algebraicity in our definition: it is natural to expect from the point of view of moduli theory (since we expect boundedness to hold), and on a technical level it allows us to state generic conditions for the objects parametrized by the stacks. In any case, (non-)algebraicity is irrelevant for the phenomenon that we will exhibit.
We work over an algebraically closed field k of characteristic
$p>0$
. First, we define one-parameter stable pairs and stable families of pairs, generalizing slightly Definition 2.1.
Definition A.1. A proper pair
$(X,\Delta )$
over k is stable if it is semi-log canonical (slc) (Footnote
7
) and if
$K_X+\Delta $
is ample.
Definition A.2. Let T be regular k-curve. Let
$f\colon X\to T$
be a flat proper pure-dimensional morphism with geometrically reduced fibers. Let
$\Delta $
be a
$\mathbb {Q}$
-divisor on X such that
$f\colon (X,\Delta )\to T$
is a family of pairs [Reference Kollár18, Definition 2.2]. We say that
$f\colon (X,\Delta )\to T$
is a stable family if:
-
1.
$(X,X_t+\Delta )$
is slc for every closed point
$t\in T$
, and -
2.
$K_X+\Delta $
is f-ample.
Lemma A.3 (cf. [Reference Patakfalvi25, Lemma 9.4])
Let T be an affine regular k-curve. Let
$(X_i,\Delta _i)\to T (i=1,2)$
be two stable families over T whose total spaces
$X_i$
are normal. Assume that for a fixed closed point
$0\in T$
, with complement
$T^*=T\setminus \{0\}$
, there is a
$T^*$
-isomorphism of pairs
Then
$\phi ^*$
extends to a T-isomorphism
$\phi \colon (X_1,\Delta _1)\cong (X_2,\Delta _2)$
.
Proof. If Z is the normalization of the closure of the graph of
$\phi ^*$
, with projections
$\alpha _i\colon Z\to X_i$
, let us write for each i
$$ \begin{align*}\alpha_i^*(K_{X_i}+\Delta_i)=K_Z+\underbrace{\Gamma_{i,-}}_{\leq 0}+\underbrace{\Gamma_{i,+}}_{\geq 0}\end{align*} $$
where
$\Gamma _{i,-}$
and
$\Gamma _{i,+}$
have no components in common. We have
Now since
$(X_i,\Delta _i+X_{i,0})$
is lc and
$X_{i,0}$
is Cartier, if E is an exceptional divisor over
$X_i$
with
${a(E;X_i,\Delta _i)<0}$
then E dominates T. In particular, every component of
$\Gamma _{i,+}$
dominates T. By assumption on
$\phi ^*$
it follows that
$\Gamma _{1,+}=\Gamma _{2,+}$
and therefore the above
$\operatorname {\mathrm {Proj}}$
description of the
$X_i$
yields the extension
$\phi \colon (X_1,\Delta _1)\cong (X_2,\Delta _2)$
.
Now we want to give our working definition of KSBA moduli stacks. We will only specify their values on points and regular curves, using the two definitions above. But at least for psychological comfort, it is better to specify what kind of values our stacks take in general. We follow the first step of the approach of [Reference Kollár18] (see in particular Chapter 7 there): the objects of interest lie amongst relative effective Mumford divisors.
We refer to [Reference Kollár18, 4.29, 4.78] for the definition of a relative effective Mumford divisor D supported on a morphism
$X\to T$
. To simplify the terminology, from now on a relative effective Mumford divisor will refer to the
$2$
-tuple
$(X\to T, D)$
. As explained in [Reference Kollár18, 4.29], given any
$h\colon T'\to T$
we can form the pullback
$h_X^{[*]}D$
which is again a relative effective Mumford divisor supported on
$X_{T'}\to T'$
, and this pullback operation is functorial. Thus we can make the following definition:
Definition A.4 (Fibered category of relative Mumford divisors)
Fix a (possibly empty) vector of positive rational numbers
$\textbf {c}=(c_1,\dots ,c_r)\in (\mathbb {Q}_{>0})^{\oplus r}$
. We define the category
${\mathcal {M}um}_{\textbf {c}}$
as follows:
-
• its objects are
$2$
-tuples
$\mathcal {D}=(f\colon X\to T,\sum _{i=1}^rc_iD_i)$
where f is a morphism of k-schemes, and each
$(f\colon X\to T,D_i)$
is a relative effective Mumford divisor supported on f; -
• the class of arrows is generated by the following ones:
-
1. if
$\mathcal {D}=(X\to T,\sum _{i=1}^rc_iD_i)$
and
$\mathcal {D}'=(X'\to T,\sum _{i=1}^rc_iD_i')$
are two relative Mumford divisors over a common base T, then every T-isomorphism
$\phi \colon X\cong X'$
sending
$D_i$
to
$D_i'$
for each i defines an arrow
$\phi \colon \mathcal {D}\cong \mathcal {D}'$
; -
2. if
$\mathcal {D}$
is a relative Mumford divisor over T and
$h\colon T'\to T$
is a k-morphism, then there is a pullback morphism
$h^{[*]}\mathcal {D}\to \mathcal {D}$
.
-
We have a forgetful functor
${\mathcal {M}um}_{\textbf {c}}\to \operatorname {\mathrm {Sch}}_k$
sending
$(X\to T,\sum _{i=1}^rc_iD_i)$
to T, the isomorphism
$\phi $
to
$\operatorname {\mathrm {id}}_T$
, and
$h^{[*]}\mathcal {D}\to \mathcal {D}$
to h. By functoriality of pullbacks, this makes
${\mathcal {M}um}_{\textbf {c}}$
into a fibered category over
$\operatorname {\mathrm {Sch}}_k$
.
With these definitions in place, we can at last introduce our working definition of KSBA moduli stacks.
Terminology A.5. A stack over k means a stack in groupoids over the big étale site of k in the sense of [1, 02ZH]. Algebraic stacks are understood in the sense of [1, 026N].
Definition A.6 (Potential KSBA moduli stacks)
Let
$\mathcal {M}$
be an algebraic stack over k. We say that
$\mathcal {M}$
is a potential KSBA moduli stack if there exist
$n\in \mathbb {N}$
,
$v\in \mathbb {Q}_{>0}$
, and
$\textbf {c}=(c_1,\dots ,c_r)\in (\mathbb {Q}_{>0})^{\oplus r}$
with the following properties:
-
1. There exists a fully faithful functor
$\mathcal {M}\hookrightarrow {\mathcal {M}um}_{\textbf {c}}$
of fibered categories over
$\operatorname {\mathrm {Sch}}_k$
; -
2. if K is an algebraically closed field extension of k, then
$\mathcal {M}(K)$
is the groupoid of all stable pairs
$(X\to \operatorname {\mathrm {Spec}}(K),\Delta =\sum _{i=1}^rc_i\Delta _i)\in {\mathcal {M}um}_{\textbf {c}}(K)$
with
$\dim X=n$
and
$(K_X+\Delta )^n=v$
; -
3. if T is a regular (germ of) k-curve then
$\mathcal {M}(T)$
is the set of
$(X\to T,\Delta =\sum _{i=1}^rc_i\Delta _i)\in {\mathcal {M}um}_{\textbf {c}}(T)$
such that
$(X,\Delta )\to T$
is a stable family.
We refer to the array
$(n,v,\textbf {c})$
as the numerical constants of
$\mathcal {M}$
.
Remark A.7. Let us stress that if
$\mathcal {M}$
is a potential KSBA moduli stack and T a regular k-curve, then the families in
$\mathcal {M}(T)$
are stable families of pairs whose underlying fibers are demi-normal (in particular reduced and
$S_2$
). This is forced by the way we compute pullbacks in
${\mathcal {M}um}_{\textbf {c}}$
and by the prescription of the values of
$\mathcal {M}$
on points.
We introduce two variants of the above definition.
Variant A.8 (Potential KSBA-CM moduli stacks)
Let
$\mathcal {M}$
be a potential KSBA moduli stack over k. We say that
$\mathcal {M}$
is a potential KSBA-CM moduli stack if there exists a dense open algebraic sub-stack
$\mathcal {M}^{\text {CM}}\subset \mathcal {M}$
such that, whenever
$(X,\Delta )\in \mathcal {M}^{\text {CM}}(K)$
for an algebraically closed field K, the variety X is Cohen–Macaulay.
Remark A.9. Because of [Reference Grothendieck10, 12.2.1], we think of a potential KSBA-CM moduli stack as an open sub-stack of a potential KSBA moduli stack. In characteristic
$0$
we have the following striking picture: if
$\mathfrak {M}=\mathfrak {M}^{\text {KSBA}}_{\mathbb {C},n,v,\textbf {c}}$
is the KSBA moduli stack over
$\mathbb {C}$
with numerical constants
$(n,v,\textbf {c})$
, defined as in [Reference Kollár18, §8.2], then we have a decomposition into open and closed sub-stacks
$$ \begin{align*}\mathfrak{M}=\bigsqcup_{i=2}^n\mathfrak{M}(i)\end{align*} $$
where
$\mathfrak {M}(i)$
parametrizes pairs whose underlying varieties are
$S_i$
but not
$S_{i+1}$
. This follows from [Reference Kollár and Kovács20] (see also [Reference Kollár and Kovács19, Corollary 1.3]). In particular
$\mathfrak {M}^{\text {CM}}=\mathfrak {M}(n)$
is a connected component of
$\mathfrak {M}$
.
Variant A.10 (Potential KSBA-F-injective moduli stacks)
Let
$\mathcal {M}$
be a potential KSBA moduli stack over k. We say that
$\mathcal {M}$
is a potential KSBA-F-injective moduli stack if there exists a dense open algebraic sub-stack
$\mathcal {M}^{F\text {-inj}}\subset \mathcal {M}$
such that, whenever
$(X,\Delta )\in \mathcal {M}^{F\text {-inj}}(K)$
for an algebraically closed field K, the variety X is F-injective.
F-injectivity is often described as a positive characteristic analogue of the du Bois condition in characteristic
$0$
: see, for example, [Reference Schwede31] and [Reference Kovács and Schwede23].
Remark A.11 (Generic singularities in [Reference Kollár17])
We briefly comment on the singularities of a general fiber
$Y^c_1$
of the examples of [Reference Kollár17], using the notations of that paper. As noted in [Reference Kollár17, Lemma 17],
$Y_1^c$
is not CM. It is easily seen from the construction that
$Y_1^c$
is not klt: a resolution is given by
$\tau _Y\colon Y_1\to Y_1^c$
, and the exceptional locus is the divisor
$Z_0$
which is a log canonical place. We check whether
$Y_1^c$
is F-injective along
$\tau _Y(Z_0)$
. Recall that
$Z_0\to \tau _Y(Z_0)$
is the elliptic fibration
$\tau \colon S_1\to \mathbb {P}^1$
[Reference Kollár17, Proof of Lemma 17]. Arguing as in [Reference Arvidsson, Bernasconi and Patakfalvi3, Lemma 3.10, Proposition 3.12], we see that for any
$y\in \tau _Y(Z_0)$
we have
The fibration
$\tau $
has one multiple fiber
$pD_1$
, and the other fibers are smooth elliptic curves [Reference Kollár17, §9]. The torsion of
$R^1\tau _*\mathcal {O}_{S_1}$
is concentrated at the image
$\mathfrak {y}$
of
$pD_1$
, and we have
$\dim H^0_{\mathfrak {y}}(\mathbb {P}^1, R^1\tau _*\mathcal {O}_{S_1})=1$
[Reference Kollár17, Proof of Lemma 17]. By cohomology and base-change, this torsion module embeds into
$H^1(pD_1,\mathcal {O}_{pD_1})$
, and to check F-injectivity it remains to understand the action of Frobenius on this cohomology group. By [Reference Kollár17, Proof of Lemma 17] we have
$H^1(pD_1,\mathcal {O}_{pD_1})=H^1(E,F_{p-1}\oplus \mathcal {O}_E)$
where E is an elliptic curve and
$F_{p-1}$
is the unique unipotent indecomposable bundle of rank
$p-1$
on E. By induction on the rank of such bundles, we see that the action of Frobenius on
$H^1(E,F_{p-1}\oplus {\mathcal {O}}_E)$
is bijective if and only if E is ordinary. In particular, we may arrange
$Y_1^c$
to be F-injective.
With all these preparations, we can finally formulate:
Theorem A.12. Let
$n\geq \max \{p,3\}$
. Then there is a dense subset I of
$(0;+\infty )$
such that for every
$v\in I$
there exists
$\textbf {c}=\textbf {c}(v)\neq \emptyset $
with the following property: there is no proper potential KSBA, nor KSBA-CM/F-injective, moduli stack over k with numerical constants
$(n,v,\textbf {c})$
.
Proof. Let
$T=\mathbb {A}^1_t$
and consider some family
$f\colon (\mathcal {Y},\mathcal {B})\to T$
constructed in Theorem 4.7 of relative dimension n. There are many choices for
$\mathcal {B}$
, and we exploit this later. For now let
$T^*=T\setminus \{0\}$
and
$f^*\colon (\mathcal {Y}^*,\mathcal {B}^*)\to T^*$
denote the family restricted to the punctured curve. A geometric fiber of
$f^*$
has only
$\mu _p$
-quotient singularities: in particular it is Cohen–Macaulay and F-injective [Reference Posva27, Theorem 1]. So if
$\mathcal {M}$
is a potential KSBA (or KSBA-CM/F-injective) moduli stack whose numerical constants match those of
$(\mathcal {Y}^*,\mathcal {B}^*)\to T^*$
, we must have
$[(\mathcal {Y}^*,\mathcal {B}^*)\to T^*]\in \mathcal {M}(T^*)$
. Restricting over the generic point of
$T^*$
, we obtain
$[f]\colon \operatorname {\mathrm {Spec}}(k(t))\to \mathcal {M}$
. I claim that there is no finite extension of DVRs
$k[t]_{(t)}\hookrightarrow R$
such that we have a commutative diagram over k

where
$\iota $
is induced by the DVR extension. This implies nonproperness of
$\mathcal {M}$
by [1, 0CLZ].
We proceed by contradiction. Assume that there exists such an extension of DVRs and morphism
$[g]\colon \operatorname {\mathrm {Spec}}(R)\to \mathcal {M}$
. This implies that there exists a stable family
$g\colon (\mathcal {Y}',\mathcal {B}')\to \operatorname {\mathrm {Spec}}(R)$
, belonging to
$\mathcal {M}(R)$
, whose generic fiber is the base-change of the generic fiber of
$(\mathcal {Y},\mathcal {B})\to T$
along
$\iota $
. The scheme
$\mathcal {Y}'$
is also normal, since its generic fiber is normal and its special fiber is reduced. Moreover, by Theorem 4.7 the base-change
$\mathcal {Y}_R$
is normal. Therefore by Lemma A.3 we actually have an R-isomorphism
$(\mathcal {Y}_R,\mathcal {B}_R)\cong (\mathcal {Y}',\mathcal {B}')$
. So
$(\mathcal {Y}_R,\mathcal {B}_R)\in \mathcal {M}(R)$
: but the central fiber of
$\mathcal {Y}_R\to \operatorname {\mathrm {Spec}}(R)$
is not
$S_2$
by Theorem 4.7, and we obtain a contradiction with the definition of the elements of
$\mathcal {M}(R)$
(see Remark A.7).
It remains to prove that as we change the boundary
$\mathcal {B}$
, the set of volumes
$(K_{\mathcal {Y}_t}+\mathcal {B}_t)^n$
, for t an arbitrary point in
$T^*$
, is dense in the interval
$(0;+\infty )$
. Recall that the support of
$\mathcal {B}_t$
is the image through a degree p quotient map
of a divisor
$H_0$
. By Remark 4.11 it suffices to consider the case where
$H_0$
belongs to the
$\mathbb {Q}$
-linear systems
where
$r\geq 3$
and
$s\geq 1$
are integers. From (4.3) and Lemma 4.8 it follows that
$$ \begin{align*} (K_{\mathcal{Y}_t}+\mathcal{B}_t)^n &= p\cdot (K_{\mathbb{P}^{n-1}\times E}+(p-1)K_{\mathcal{G}_t}+H_0)^n \\&= p\cdot \left[\mathcal{O}_{\mathbb{P}^{n-1}}\left( p-n-1+\frac{r}{s}(n+p) \right)\boxtimes \mathcal{O}_E\left(\frac{r}{s}e\right)\right]^n \\&= np\cdot \frac{r}{s}\cdot \left( p-n-1+\frac{r}{s}(n+p) \right)^n. \end{align*} $$
We consider this expression as a function
$\nu $
in
$r/s\in \mathbb {R}_{>0}$
. Then
$\nu $
is continuous, increases to infinity with
$r/s$
, and is bounded below by
By continuity the image
$I=\nu (\mathbb {Q}_{>0})$
is dense in
$(0;+\infty )$
. By construction, for every
$v=\nu (r/s)\in I$
there is a family
$(\mathcal {Y}^*,\mathcal {B}^*)\to T^*$
constructed by Theorem 4.7 which belongs to any potential KSBA or KSBA-CM/F-injective moduli stacks with numerical constants
$(n,v,\textbf {c}\neq \emptyset )$
(Footnote
8
). This completes the proof.
Acknowledgments
I thank the Institute of Mathematics of the HHU Düsseldorf and Stefan Schröer’s research group for their hospitality. I am also thankful to Zs. Patakfalvi and B. Totaro for several comments on early drafts, to J. Baudin for discussions related to [Reference Baudin, Bernasconi and Kawakami6], and to the reviewer for several corrections.
Competing interests
The authors have no competing interest to declare.
Funding statement
This research was supported by the PostDoc Mobility grant P500PT-210980 from the Swiss National Science Foundation.





