Moduli of -Gorenstein pairs and applications
Abstract.
We develop a framework to construct moduli spaces of -Gorenstein pairs. To do so, we fix certain invariants; these choices are encoded in the notion of -stable pair. We show that these choices give a proper moduli space with projective coarse moduli space and they prevent some pathologies of the moduli space of stable pairs when the coefficients are smaller than . Lastly, we apply this machinery to provide an alternative proof of the projectivity of the moduli space of stable pairs.
2020 Mathematics Subject Classification:
Primary 14J10, 14D22, Secondary 14E30.1. Introduction
Since the seminal work of Mumford on the moduli space of smooth curves with and its compactification , significant progress has been made in understanding its higher-dimensional analog, namely the moduli space of stable varieties. On a first approximation, those are -Gorenstein varieties with relatively mild singularities, and such that is ample. A complete and satisfactory moduli theory for stable varieties of any dimension has been settled due to the work of several mathematicians (see [KSB, Kol90, Ale94, Vie95, Kol13-mod, Kol13, HK04, AH11, kol08, HX13, kol_new, HMX]). Furthermore, in this setup, it is well understood what families should be parametrized by the moduli functor.
A natural generalization of is given by the moduli space of -pointed curves and its compactification, denoted by and , respectively. A first attempt to generalize the notion of stable pointed curve is to consider mildly singular pairs (specifically, semi-log canonical pairs) where is a reduced divisor, such that is ample. This approach has been worked out Alexeev in dimension 2 [Ale94, Ale96], and combining the efforts of several mathematicians (see [Ale94, Kol13, kol08, HX13, KP17, kol_new, HMX, kol19s]), it has been generalized in higher dimensions.
While a stable variety is -Gorenstein, a stable pair might not be, so if one is interested in -Gorenstein pairs, the aforementioned formalism might not give the desired moduli space. In fact, one cannot simply consider the -Gorenstein locus of the moduli of stable pairs, as this could be not proper. Therefore, the main goal of this paper is to construct a proper moduli space parametrizing -Gorenstein pairs . We observe that, in [Ale15], Alexeev already considered certain moduli spaces of -Gorenstein stable pairs, see also [kol_new]*§ 6.4. As in the classical case, we require our pairs to have semi-log canonical singularities, but we relax the ampleness condition on , requiring only that is ample for . We also impose a numerical condition on the intersection number , analogous to the genus condition for curves, which we encode via a polynomial function . This condition guarantees the boundedness of our moduli problem. These choices are encoded in the notion of -stable pair, a pair on which both and are -Cartier, (Defintion 2.7) and lead to the following:
Theorem 1.1.
Fix an integer and a polynomial . Then, there is a proper Deligne–Mumford stack , with projective coarse moduli space, parametrizing -stable (-Gorenstein) pairs of dimension and with polynomial and reduced boundary.
In the statement of Theorem 1.1, the subscript 1 in means that is a reduced divisor.
While it is very natural to consider the divisor with its reduced structure, the framework of the Minimal Model Program highlights the importance of the use of fractional coefficients for the divisor . For example, in the case of curves, one can replace with a weighted version, where the markings can attain any fractional value in . This was accomplished by Hassett in [Has03], and this approach leads to different compactifications of .
It turns out, however, that the construction of the higher dimensional analogs of these weighted moduli spaces is very delicate. Among the many difficulties, the definition of a suitable notion of family of boundary divisors represents a major problem, see [kol_new]*Ch. 4. Nevertheless, over the past decade, there has been significant progress in the development of a moduli theory for higher dimensional stable pairs (see [Ale94, Kol13, kol08, HX13, KP17, kol_new, HMX]), and this last missing piece has finally been settled by Kollár in [kol19s]. In loc. cit. a rather subtle refinement of the flatness condition is introduced, leading to the ultimate notion of family of divisors, that in turn gives a satisfactory treatment of a moduli functor of families of stable pairs, admitting divisors with arbitrary coefficients.
The difficulty in defining a suitable notion of family of boundary divisors is due to the fact that, in general, a deformation of a pair cannot be reduced to a deformation of the total space and a deformation of the divisor . This naïve expectation is only true if the coefficients of are all strictly greater than (see [Kol14]), in which case a deformation of over a base curve induces a flat deformation of . On the other hand, if smaller coefficients are allowed, the situation becomes much more subtle. For instance, as showed by Hassett (see § 2.11 for details), by allowing the coefficient , it is possible to define a family of stable surface pairs such that the flat limit of acquires an embedded point along the special fiber.
One advantage of our formalism is that it allows overcoming the aforementioned difficulties. That is, our setup easily generalizes to the case of any fractional coefficients, preventing the above-mentioned pathologies of families of divisors, and leading to a more transparent definition of the moduli functor in Definition 2.21. Indeed, we generalize to an analogous moduli functor in Definition 2.21, allowing arbitrary coefficients on the divisor. The main feature of Definition 2.21 is that the family of boundaries is characterized as a flat, proper, and relatively morphism over the base, thus significantly simplifying the definition of the classical moduli functor. The condition prevents the existence of embedded points, while the polynomial controls the relevant invariants of the varieties, preventing them from jumping as in Hassett’s example. This leads to the following generalization of Theorem 1.1:
Theorem 1.2.
Fix an integer , a finite subset , and a polynomial . Assume that is closed under sum: that is, if and , then . Then there is a proper Deligne–Mumford stack with projective coarse space, parametrizing -stable pairs of dimension and coefficients in .
After the completion of this work, we learnt that Kollár has also developed an analogous machinery allowing pairs where the boundary divisors have coefficients that can be independently perturbed; see [kol_new]*§ 8.3.
In § 6 we relate our moduli functor with the moduli functor of stable pairs, and as the main application of the theory of -stable pairs, we obtain a simpler proof of the projectivity of the moduli space of stable pairs, originally proved by Kovács and Patakfalvi [KP17].
Theorem 1.3 (Corollary 7.7).
Consider a proper DM stack that satisfies the following two conditions:
-
(1)
for every normal scheme , the data of a morphism is equivalent to a stable family of pairs with fibers of dimension , volume and coefficients in ; and
-
(2)
there is such that, for every , there is a line bundle on such that, for every morphism as above, .
Then, the coarse moduli space of is projective.
As observed in [KP17]*§ 1.1, the approach of [Has03] for proving projectivity of these moduli spaces cannot be adapted to higher dimensions, as certain sheaves are no longer functorial with respect to base change. For this reason, Kovács and Patakfalvi develop a refinement of Kollár’s ampleness lemma. On the other hand, in the setup of -stable pairs, all the needed sheaves remain functorial with respect to base change. Indeed, flatness and the condition guarantee that all notions of pull-back agree. Thus, we can follow Hassett’s strategy and directly apply [Kol90]. In this way, the projectivity part of Theorem 1.2 is established; then, to deduce Theorem 1.3, it suffices to show that the moduli space of -stable pairs naturally admits a finite morphism to the moduli space of stable pairs, and the needed polarization descends with this morphism.
1.1. Structure of the paper
The first part of this work is devoted to developing the notion of -stable pair and extending several statements from pairs to -stable pairs. In particular, in § 2 we set the key definitions and properties, while in § 3 we extend the boundedness results of [HMX18] to the context of -stable pairs.
In § 4 and § 5, we analyze the moduli functor . In particular, in § 4 we show that , which a priori is only a category fibered in groupoids over , is a Deligne–Mumford stack. Then, in § 5 we show that is proper. Thus, § 4 and § 5 settle Theorem 1.2, except for the projectivity part.
In § 6, we analyze the notion of family of -stable pairs when the base is reduced. Under these assumptions, we show that the existence of a relative good minimal model is determined by a fiberwise condition.
1.2. Connections with other works
One of the main difficulties to extend the moduli theory from stable varieties to stable pairs lays in the behavior of the boundary divisor in families. Indeed, given a pair , and need not be -Cartier. In this case, we cannot expect a deformation of to induce a flat deformation of , so one cannot define the moduli problem by simply requiring that the divisor varies flatly in families. A final and satisfactory notion for defining families of divisors has been achieved in [kol19s] with the notion of K-flatness.
Over the years, there have been attempts to overcome these difficulties in some special situations where the deformations of do induce flat deformations of . In [Ale06, Ale15], Alexeev considered pairs , where each is prime and the coefficients are irrational and -linearly independent. This assumption can be thought as a “general choice” of the coefficients of , as opposed to having rational coefficients. Indeed, many of the examples, it can be showed that, as we vary the coefficients of the boundary, the behavior of the moduli space is locally constant, leading to a chamber decomposition of some appropriate polytope contained in . These chambers have rational vertices, determined by the conditions of being semi-log canonical and of being ample. For more details on this topic, we refer to [ABIP], and some examples of this phenomena are illustrated in [Ale15, inc20, AB21].
In a different direction, Kollár showed in [Kol14] that, if the coefficients of are strictly greater than , deformations of lead to flat deformations of . Thus, even though is not necessarily -Cartier, this allows for a satisfactory moduli theory that does not rely on the latest developments in [kol19s].
In this paper, we work with rational coefficients and impose that the boundary is -Cartier. As we do not require the coefficients to be irrational, in order to retain the properness of the moduli problem, we trade the ampleness of for the flatness of the deformations of . Thus, while the deformations of the boundary are easier to understand than the general framework of [kol19s], is in general only big and semi-ample.
In [kol_new]*§ 8.3, Kollár presents a more general version of the approach pursued in this work. In particular, he considers arbitrary coefficients of and allows for independent perturbations of different components of . In this way, he reconciles Alexeev’s work with our work, thus showing that the case of “ with rational coefficients and big and semi-ample” we consider can be thought of as a limit case of the “general coefficients with ample” considered by Alexeev.
Similarly, we remark that many of the ideas in the current paper fit in the general formalism of stable minimal models, developed by Birkar (see for example [birkar2021boundedness]), culminating in [birkar2022moduli], where he develops a moduli theory of varieties and pairs of non-negative Kodaira dimension. We refer the reader to Definition 2.7 and [birkar2021boundedness]*Definition 1.8 for the similarities between the definitions of -stable pair and stable minimal models. In particular, we observe that both definitions entail a stability condition prescribed by a polynomial, which guarantees the boundedness of the moduli problem, thus taking inspiration from ideas of Viehweg [Vie95].
Acknowledgements
We thank János Kollár for pointing to us a mistake in an earlier version of this work, for sharing the latest draft of his book on moduli, for providing helpful comments to improve our work, and for providing feedback on an eralier version of this work. We thank Jarod Alper, Dori Bejleri, Christopher Hacon, Sándor Kovács, Joaquín Moraga, and Zsolt Patakfalvi for helpful discussions. We thank the anonymous referees for helpful suggestions.
2. Preliminaries
2.1. Terminology and conventions
Throughout this paper, we will work over the field of complex numbers. For the standard notions in the Minimal Model Program (MMP) that are not addressed explicitly, we direct the reader to the terminology and the conventions of [KM98]. Similarly, for the relevant notions regarding non-normal varieties, we direct the reader to [Kol13]. A variety will be an integral separated scheme of finite type over . A birational morphism between non-normal schemes will be a morphism with a dense open subset such that is dense and is an isomorphism.
2.2. Contractions
A contraction is a projective morphism of quasi-projective varieties with . If is normal, then so is .
2.3. Divisors
Let denote , , or . We say that is a -divisor on a variety if we can write where , and the are prime Weil divisors on for all . We say that is -Cartier if it can be written as a -linear combination of -divisors that are Cartier. The support of a -divisor is the union of the prime divisors appearing in the formal sum .
In all of the above, if , we will systematically drop it from the notation.
Given a prime divisor in the support of , we will denote by the coefficient of in . Given a divisor on a normal variety , and a morphism , we define
Let and be divisors on . We write if there is a -Cartier divisor on such that . Equivalently, we may also write , or over . If , we omit it from the notation. Similarly, if , where is the ground field, we omit from the notation.
Let be a projective morphism of normal varieties. Let and be two -divisors on . We say that and are numerically equivalent over , and write , if for every curve such that is a point. In case the setup is clear, we just write , omitting the notation .
2.4. Non-normal varieties and pairs
There are two important generalizations of the notion of normal variety.
Definition 2.1.
An scheme is called demi-normal if its codimension 1 points are either regular or nodal.
Roughly speaking, the notion of demi-normal schemes allows extending the notion of log canonical singularities to non-normal varieties, allowing for a generalization of the notion of stable curve to higher dimensions. We refer to [kol_new]*§ 10.8 for more details.
Definition 2.2.
A finite morphism of schemes is called a partial seminormalization if is reduced and, for every point , the induced map is an isomorphism. There exists a unique maximal partial seminormalization, which is called the seminormalization of . A scheme is called seminormal if the seminormalization is an isomorphism.
This is an auxiliary lemma, it is probably well known. We include it for completeness. We refer to [kol_new]*§ 10.8 for the details about seminormality.
Lemma 2.3.
Let be a proper surjective morphism with connected fibers, with seminormal and reduced. Then .
Remark 2.4.
Observe that, in situation of Lemma 2.3, by the projection formula it follows that, for any line bundle on , we have .
Proof.
We can take the Stein factorization . Since is reduced, then is reduced. Observe that , so the desired result follows if we can show that is an isomorphism.
Since the composition has connected fibers and is surjective, then has connected fibers. Since is finite, it is injective. It is also surjective since is surjective, so is a bijection. Then, we can take the seminormalization and consider the composition . This is a bijective morphism since it is a composition of bijections, and it is proper since is proper and is proper. Then, the composition is an isomorphism, since the source and the target are seminormal.
Now, on the topological space given by we have the following morphsims of sheaves:
The composition is surjective, so is surjective. Moreover, since is reduced, the map is also injective. Then it is an isomorphism, so . ∎
2.5. Divisorial sheaves
Throughout this section, will be and reduced. We begin this subsection with the following definition:
Definition 2.5.
Let be a scheme. A sheaf on is called divisorial sheaf if it is and there is a closed subscheme of codimension at least 2 such that is locally free of rank 1.
Definition 2.6 ([HK04]*§ 3).
Let a flat morphism of schemes, and a coherent sheaf on . We say that is relatively if it is flat over and its restriction to each fiber is
Now, let be a demi-normal scheme, and let denote the subgroup of generated by the prime divisors that are not contained in the conductor of . Then, there is an identification between and the group of divisorial sheaves, where denotes linear equivalence. The identification is defined as follows. By the definition of and the demi-normality of , for every element , there is a closed subset of codimension at least 2 such that is a Cartier divisor. Then, the corresponding divisorial sheaf is defined as , where we have .
Consider a flat family with fibers, and assume that there is an open subset such that is big for every . Then, for every locally free sheaf on , we can consider on . From [HK04]*Corollary 3.7, this is a reflexive sheaf on (a priori, it is not relatively to , see § 2.11).
If is demi-normal, its canonical sheaf is a divisorial sheaf, as is Gorenstein in codimension 1. By the above identification, we can then choose a canonical divisor such that . Observe that this construction can be carried out in families. Indeed, if is a flat morphism with demi-normal fibers of dimension , there is an open locus that has codimension 2 along each fiber, on which is Gorenstein. Then we can define . Observe that this agrees with the -th cohomology of the relative dualizing complex. Indeed, the latter is (see [LN18]*§ 5).
Let be demi-normal, and consider two divisorial sheaves and . Then, their reflexive tensor product is defined as and it is a divisorial sheaf itself. If we have and , we have . The -fold reflexive power is defined as the -fold self reflexive tensor product.
For more details, we refer to [kol_new]*§ 3.3, [Kol13]*5.6 and [HK04].
2.6. Boundedness
Let be a set of projective pairs. Then, we say that is log bounded (resp. log birationally bounded) if there exist a variety , a reduced divisor on , and a projective morphism , where is of finite type, such that does not contain any fiber of , and, for every , there are a closed point and a morphism (resp. a birational map) inducing an isomorphism (resp. such that contains the strict transform of and all the exceptional divisors).
A set of projective pairs is said to be strongly log bounded if there is a quasi-projective pair and a projective morphism , where is of finite type, such that does not contain any fiber of , and for every , there is a closed point and an isomorphism such that .
A set of projective pairs is effectively log bounded if it is strongly log bounded and we may choose a bounding pair such that, for every closed point , we have .
2.7. Index of a set
Given a finite subset , we define the index of to be the smallest positive rational number such that .
2.8. Stable pairs and -stable pairs
Let denote a projective semi-log canonical pair, where has coefficients in . We say that is a stable pair if is ample.
Definition 2.7.
Consider a polynomial and a set . A -pair with polynomial and coefficients in is the datum of a semi-log canonical pair and a -Cartier -divisor on satisfying the following properties:
-
(1)
is semi-log canonical;
-
(2)
;
-
(3)
there is a stable pair with a birational contraction such that , and ; and
-
(4)
the coefficients of and are in .
If moreover there is such that for every the pair is stable, the -pair is called a -stable pair. Finally, we will call the canonical model of .
Remark 2.8.
For brevity, we denote the datum of a -pair by , where the polynomial and the set of coefficients are omitted in the notation. In case , we then write .
Remark 2.9.
For example, if is a semi-log canonical -stable pair, and is the normalization of with conductor , then is a -stable pair.
Notation 2.10.
If is a finite set and is its index, given a -pair with polynomial and coefficients in we denote by as the subscheme of defined by the reflexive sheaf of ideals .
Remark 2.11.
Observe that if is a -stable pair, the -divisor is nef since it is limit of ample -divisors. Furthermore, is big, as it is the sum of an ample -divisor and an effective -divisor.
Lemma 2.12.
Let be a projective variety, and let and be two nef -Cartier -divisors. Assume that for some , the divisor is ample. Then, is ample for every .
Proof.
Since ampleness is an open condition, we may assume that . Then, using convex combinations, the claim follows from Kleiman’s criterion and the denseness of in . ∎
Lemma 2.13.
Let be a -stable pair, and assume that is its canonical model. Then, .
Proof.
By assumption, there exists such that is a stable pair. Thus, if is an irreducible curve that is not contained in , we have that
since the first summand is positive by ampleness and the second is non-negative as is not contained in . Thus, the curves contracted by are contained in and the claim follows. ∎
Lemma 2.14.
Let and be natural numbers. Then, there exists , only depending on and , such that the following holds. Let be a normal -stable pair of dimension , and let be its canonical model . Further assume that the Cartier index of is less than . Then, for every , the pair is stable.
Proof.
By definition of -stable pair and of canonical model, for every -exceptional curve , we have and for . For every such curve , the function is linear and not identically 0, so it has at most one zero. In particular, for every we have
| (1) |
Hence, if we choose , we have
| (2) |
for every -exceptional curve.
By the same argument, for every curve such that , it follows that
| (3) |
for every . Now, define
Since is ample with Cartier index bounded by , for every curve that is not -exceptional, by the projection formula, we have
| (4) |
From [Fuj11]*Theorem 1.4, for every -negative extremal ray , we may find a curve generating such that
| (5) |
By (2), any such curve is not -exceptional. Then, by (4) and (5), for any such , we have
| (6) |
Now, consider the cone of curves and its decomposition given by the cone theorem associated to the pair [Fuj11]*Theorem 1.4. Then, by (3), we have that is positive on . Thus, every -negative extremal ray is also a -negative extremal ray. Then, by (6), we have that is positive on . Thus, by the cone theorem [Fuj11]*Theorem 1.4, the log canonical pair has no negative extremal rays; thus, is nef.
Now, by the definition of -stable pair, some convex combination of and is ample. Then, the claim follows by Lemma 2.12. ∎
Lemma 2.15.
Fix an integer , a volume , and a finite subset . There is , only depending only on and , such that, for every -stable pair of dimension , coefficients in , and polynomial with , the pair is stable for every .
Proof.
Without loss of generality, we may assume from now on that . Let be the normalization of . Let be the pair induced by , and let denote the pull-back of to . Here, consists of the sum of the divisorial part of the preimage of and the conductor. Notice that possibly has more than one connected component. Then, we have:
-
(1)
has still coefficients in ;
-
(2)
is semi-ample and big; and
-
(3)
is a -stable pair.
From [HMX14]*Theorem 1.3, there are finitely many possibilities to write as , where is the volume of a log canonical pair of general type with coefficients in . So from (1), there is a finite set such that the volume of each connected component of is in .
For a log canonical pair of general type , the canonical model is such that . In particular, it follows from (2) that the irreducible components of admit a canonical model, and the volumes of these models are contained in as well. From [HMX]*Theorem 1.1, there is an such that every stable pair with coefficients in , dimension , and volume in the finite set has Cartier index less than . Since we can check ampleness after passing to the normalization, and the Cartier indexes of each irreducible component of the normalization are bounded, the thesis follows from Lemma 2.14. ∎
Lemma 2.16.
Let be a stable pair. Then, there exists a -stable pair having as canonical model.
Observe in particular that from Lemma 2.13, the morphism induces an isomorphism at the generic point of the codimension one singular locus of and , as those are points not contained in , and is an isomorphism away from the exceptional locus. So is normal if and only if its codimension one singular locus is empty and if and only if the one of is empty.
Proof.
We consider the semi-canonical modification of , in the sense of [Fuj15]. We observe that, in order to consider a semi-canonical modification, does not need to be -Cartier. Such modification exists by [Fuj15]*Theorem 1.1 and the fact that is demi-normal. By [Fuj15]*Definition 2.6, has the following properties:
-
•
is an isomorphism around every generic point of the double locus of ;
-
•
this procedure is compatible with taking the normalizations of and , see [Fuj15]*Lemma 3.7.(2). In particular, is normal if so is , and, in general, establishes a bijection between irreducible components of and ; and
-
•
is -Cartier and -ample.
Now, set . Since is ample, for , we have that is ample. Since both and are -Cartier, then so is . Lastly, as and are semi-log canonical, to conclude it suffices to show that .
Now, let and denote the respective normalizations, where and denote the double loci. Thus, it suffices to show that . Since this can be checked by considering one irreducible component of at the time, by abusing notation, we may assume that and are irreducible. By construction, we have
and is relatively ample over . Then, since we have , by the negativity lemma [KM98]*Lemma 3.39, it follows that . This concludes the proof. ∎
2.9. Families of pairs
We recall the main definitions of families of pairs from [kol_new]*Ch. 4 and [kol19s]. A family of pairs over a reduced base is the datum of a morphism and an effective -divisor on , such that the following conditions hold:
-
•
is flat with reduced fibers of pure dimension ;
-
•
the fibers of are either empty or of pure dimension ; and
-
•
is smooth at the generic points of for every .
Furthermore, we say that a family of pairs is well defined if it also satisfies the following property:
-
•
is Cartier locally around the generic point of each irreducible component of for every , where is a sufficiently divisible natural number clearing the denominators of .
This latter condition guarantees that is Cartier on a big open set with the property that is a big open set of for every . This guarantees that we have a well-defined notion of pull-back of under any possible base change , as we can pull back , take its closure in , and then divide the coefficients by .
There is a more general definition of families of divisors, over possibly non-reduced bases due to Kollár in [kol19s]. We will not report it here since we will not need it, we refer the interested reader to loc. cit.
A well defined family of pairs over a reduced base is called locally stable if, for every base change where is the spectrum of a DVR with closed point , is a semi-log canonical pair. Then, a family is called stable if it is locally stable, is proper, and is -ample.
2.10. Families of -pairs
Let us fix a positive integer , a polynomial , and a finite set of coefficients . Let be the index of , see § 2.7. Before we introduce our main functor, we discuss some applications of the abundance conjecture that are relevant for this paper.
Notation 2.17.
Throughout the paper, when we say “assume that assumption (A) holds” we mean “assume that the following condition holds”:
if is a log canonical pair such that is ample for , then is semi-ample.
Remark 2.18.
We remark that assumption (A) is a weakening of the Abundance Conjecture, which is known up to dimension 3. Moreover, assumption (A) is known to hold for klt pairs (it is the basepoint-free theorem).
Definition 2.19.
Let be the category fibered in groupoids over whose fibers over a scheme consists of:
-
•
a flat and proper morphism of relative dimension ;
-
•
a flat and proper morphism of relative dimension and relatively ;
-
•
a closed embedding over ; and
-
•
for every point , the fiber is -stable with polynomial and coefficients in , where is the index of , see § 2.7.
For every morphism , we denote by the first projection, and by the fiber product . Similarly, for every point , we denote by . If we denote by the ideal sheaf of , we require that for every and every , , the natural map
| (K) |
is an isomorphism. We will denote by an object of over , and we will call it a weak family of -stable pairs.
Definition 2.20.
We will call weak -stable morphism the datum of a flat and proper morphisms and a closed embedding that satisfies the four bullet points of Definition 2.19. If there is no ambiguity we will still denote it with . Lastly, if is relatively semi-ample, we say it is a -stable morphism.
The condition (K) on commutativity with base change in Definition 2.19 is usually referred to as Kollár’s condition.
We need an additional condition to prove that is representable in full generality. Indeed, while proving representability if one restricts itself to the category of reduced schemes (i.e., if one is only interested in families over a reduced base) essentially follows from [HX13], to prove that is representable in general we will need some version of assumption (A) to hold in families. If one works with moduli of surfaces or threefolds, then the Abundance Conjecture is known and there are no issues. Otherwise one possible solution, which was suggested to us by Kollár, is to only consider families such that “assumption (A) holds in families”. The advantage of this approach is that the functor can be defined in any dimension unconditionally to the Abundance Conjecture.
Definition 2.21 (Kollár).
Let be the functor representing families in such that there is a family of stable pairs together with a morphism such that for every , the restriction is the canonical model of . We will denote these families as families of -stable pairs.
The main advantage of dealing with assumption (A) in this way is proved [kol_new]*Proposition 8.36, which he kindly shared with us, where he proves that imposing assumption (A) in families is a constructible condition:
Theorem 2.22 (Kollár).
Given a proper locally stable morphism there is a locally closed partial decomposition such that for any , the pull-back has a simultaneous, canonical, crepant, birational contraction with stable iff factors via .
We will use this result for constructing , and to show that it is bounded.
Summary of notations.
As it might be confusing to remember all the different definitions of families, we recall the essential differences here:
-
(1)
in weak family of -stable pairs, we require condition (K) but not condition (A);
-
(2)
in weak -stable morphism, we require neither condition (K) nor condition (A);
-
(3)
in -stable morphism, we require condition (A) but not condition (K); and
-
(4)
in family of -stable pairs, we require both condition (A) and condition (K).
In particular, we antepone “weak” if we are not requiring condition (A), and we write “family” if we require condition (K). This choice will be maintained when introducing the notion of “constant part” in § 2.12.
Now, we add a series of remarks and technical statements that are relevant in this context. We keep the notation of Definition 2.19.
Remark 2.24.
The ideal sheaf is flat over . This follows from the fact that and are flat, by considering the associated long exact sequence of .
Remark 2.25.
Since is flat and by condition (3) its fibers are , is relatively . Then, since is flat and relatively , it follows from [Kol13]*Corollary 2.61 that is relatively .
Remark 2.26.
By condition (3), there exists an open subset whose restriction to any fiber is a big open subset, such that is a Cartier divisor on for every . Then, by flatness, we may apply [stacks-project]*Tag 062Y, and we conclude that is Cartier along . Then, by Remark 2.24 and [HK04]*Proposition 3.5, and are reflexive. In particular, we have , and satisfies the conditions in [kol_new]*Definition 3.28.
Remark 2.27.
Remark 2.28.
The sheaves are ideal sheaves of . Indeed, we can again consider an open subset whose restriction to any fiber is a big open subset and such that is a Cartier divisor on . Then, if we denote by the inclusion of , by [HK04]*Corollary 3.7 we have
Then, the inclusion can be pushed forward via , to have an inclusion , where the last equality follows from [HK04]*Proposition 3.5 since is .
Notation 2.29.
If is a weak family of -stable pairs, we denote by the closed subscheme of with ideal sheaf .
Remark 2.30.
By [AH11]*Proposition 5.1.4, the sheaves and are flat over for every .
Lemma 2.31.
The morphism is flat with fibers (i.e., the fibers have no embedded points).
Proof.
To check that is flat it suffices to check that for every closed point we have . We pull back the exact sequence
via , and we obtain
However, , so in particular it is a torsion free sheaf of rank 1. Then, the map is injective, so as desired.
Finally, is and is from the commutativity with base change (K) in Definition 2.19, so is from [Kol13]*Corollary 2.61.∎
Remark 2.32.
There is an , which does not depend on , such that and are locally free on . Indeed, we will prove in Theorem 3.2 that there is an such that for every fiber of , the sheaves and are Cartier. Since our family is bounded (see Theorem 3.2), we can choose such an that does not depend on the basis . Then, by condition (K) in Definition 2.19 and Remark 2.30, we may apply [stacks-project]*Tag 00MH, which implies that and are locally free since they restrict to locally free sheaves along each fiber.
Now, we specify the morphisms in the fibered category over a morphism Let be an element of , and let be an element of . An arrow is the datum of two morphisms that fit in a diagram like the one below, where all the squares are fibered diagrams:
Observation 2.33.
The only morphisms over the identity are isomorphisms. Thus, is fibered in groupoids.
2.11. Hassett’s example
In this subsection, we present a well-known example due to Hassett that is helpful to keep in mind to navigate the rest of the paper. See also [KP17]*§ 1.2 or [kol_new].
Consider the DVR , let (resp. ) be the generic (resp. closed) point of , and let . Consider a smooth member of , and let be the blow-up of in the special fiber of . Then, if we compose the blow-down with the projection , we get a family of surfaces where the generic fiber is a copy of , while the special fiber is a surface with two irreducible components. One irreducible component of the special fiber is isomorphic to (the proper transform of the special fiber of ), and the other one is the exceptional divisor . The surface is the projectivization of the normal bundle of . Since and the normal bundle of in is isomorphic to , we have that is isomorphic to the Hirzebruch surface . We denote by the preimage of the double locus of the central fiber on .
We consider a divisor on consisting of five irreducible components, three general members of (which we denote by ), , and a smooth member of . We consider a deformation of in given by the trivial deformation of and , and we deform to for every . We denote by the total space of this deformation, and by its proper transform in . More explicitly, if is the zero locus of a global section , is the zero locus of a global section , and are generic sections of , the deformation we consider is .
Figure 1: the family
Now, we construct the canonical model of , with the coefficient in a neighbourhood of . First, we introduce some notation. We denote by , the irreducible component of the central fiber isomorphic to by , and the two rulings of by and , with the convention that . Since the generic fiber is stable for every with , we just need to control the intersection pairings on the special fiber.
Let , where .
Figure 2: special fiber of the family .
One can check that:
-
(1)
when , the divisor is nef. It is ample on , positive on and 0 on ;
-
(2)
when , the divisor is nef. It is 0 on and on ; and
-
(3)
when , the divisor is not nef. On , it is negative on and 0 on , while on it is negative on .
Therefore, we can explicitly describe the special fiber of the canonical model of over . We denote by (resp. , ) the canonical model of when (resp. , ):
-
(1)
when , to construct the canonical model we contract the ruling . Since via this contraction the map is a ramified cover of , the special fiber is the push-out of the following diagram:
-
(2)
when , to construct the canonical model, we contract and . The special fiber is isomorphic to with the section contracted, which is the projectivization of the cone over a rational quartic curve; and
-
(3)
When , we perform a divisorial contraction to make the divisor nef. We contract the ruling , and the special fiber is .
In particular, there are morphisms , and from to the special fibers of , and . The divisors , and can be described as follows. First, recall that is the projectivization of the normal bundle of inside , so it has a relative . We denote a generic section of by . Then, the following holds:
-
(1)
when , the divisor is the image via of four general fibers of , together with a divisor linearly equivalent to . All the components have coefficient ;
-
(2)
when , the divisor is the image via of four general fibers with a divisor linearly equivalent to . All the components have coefficient ; and
-
(3)
when , the divisor consists of four general fibers with a divisor linearly equivalent to , four generic fibers, and . All the components have coefficient , with the exception of , which has coefficient .
Recall now that the flat limit of in is not , since it has an embedded point (see [KP17]*§ 1.2). However, is a -stable morphism with coefficients (see Proposition 5.1), so in particular the flat limit of in does not have an embedded point on the special fiber.
2.12. -stable morphisms with constant part
For proving that is bounded, it will be useful to introduce the following definition.
Definition 2.34.
Assume is reduced. A locally weak -stable morphism with constant part over and with coefficients in is the datum of a proper morphism , a closed subscheme on , and an effective -divisor on , such that the following conditions hold:
-
•
is a proper, locally stable family of pairs, where is a family of Mumford divisors [kol19s];
-
•
is flat and relatively (namely, with no embedded points) with fibers of pure dimension ; and
-
•
for every , there is a -pair with coefficients in (i.e., both and have coefficients in ) such that , , and .
Furthermore, if we have a polynomial and for every , has polynomial , we say that it is a locally weak -stable morphism with constant part with polynomial . Lastly, if is relatively semi-ample, we drop the “weak” from the notation.
We remark that we call the constant part since, contrary to , the divisor might not be -Cartier on .
Notation 2.35.
We say that a locally (weak) -stable morphism with constant part (with coefficients in and polynomial ) is a (weak) -stable morphism with constant part (with coefficients in and polynomial ) if, for every , the fiber over is a -stable pair.
Proposition 2.36.
Let us fix a set of coefficients , a polynomial and an integer . Let denote the index of . Let be a weak -stable morphism with constant part over a reduced base . Then is a well defined family of pairs.
Proof.
By definition is flat with fibers of pure dimension . Thus, it follows that the fibers of are either empty or of pure dimension . Thus, to show that is a family of pairs, we are left with showing that is smooth at the generic points of for every .
By assumption, this is the case for all the generic points of arising from . Thus, we may focus on the contribution of . But then, since each fiber is a -pair, it follows that the generic points of are contained in the smooth locus of by the semi-log canonical condition. Thus, is a family of pairs.
To conclude, we need to show that the family is well defined. To this end, it suffices we focus on , as satisfied the needed conditions by definition. As argued in [kol_new]*Definition 3.35, there is a big open subset such that every point of is either smooth or nodal, and has codimension at least 2 in for every . For every , does not contain any irreducible component of the double locus. Thus, the intersection between and has codimension at least 2 in every fiber. Let denote the open set obtained by removing this intersection from . Then, as the scheme theoretic restriction is an integral Weil divisor, it is a Carter divisor along . Then, by [stacks-project]*Tag 062Y, is a Cartier divisor along . Thus, the claim follows. ∎
Proposition 2.37.
Fix an integer , a polynomial , and a finite subset . Let be as in Lemma 2.15, and let be a weak -stable morphism with constant part, coefficients in , and polynomial , over a reduced base . Then is a stable family of pairs for every rational. In particular, is -Cartier.
Proof.
Fix as in the statement. Then, by Proposition 2.36, is a well defined family of pairs. By assumption, the self-intersection is independent of , as it is . Since is a well defined family of pairs, we may find a big open subset such that has codimension at least 2 for every and is -Cartier along . Then, is -Cartier by [kol_new]*Theorem 5.8. The claim follows by [kol_new]*Definition-Theorem 4.7 and the fact that the argument was independent of . ∎
2.13. Existence of good minimal models
Let be a log canonical pair, and let be a projective morphism over a normal variety such that is -pseudo-effective. Then, it is expected that admits a good minimal model over . That is, admits a birational contraction over to a log canonical pair , such that is the push-forward of to , is -negative, and is semi-ample over .
Here, we collect a technical statement that shows the existence of relative good minimal models under certain assumptions. In particular, this statement is crucial to show that, under suitable hypotheses, a weak -stable morphism (resp. weak family of -stable pairs) is actually a -stable morphism (resp. family of -stable pairs).
Lemma 2.38.
Let be a log canonical pair, and let be a projective morphism to a normal variety such that is -pseudo-effective. Assume that the general fiber of has a good minimal model. Then, admits a relative good minimal model over . Furthermore, if is nef over , then it is semi-ample.
Proof.
Let be a log resolution of , and let . Here denotes the strict transform of , the divisors , , and are effective, -exceptional, and share no common components. Furthermore, is reduced, while the coefficients of are in . Let denote the reduced -exceptional divisor, and fix a rational number . Then, is dlt, and it has the same pluricanonical ring as . Furthermore, by the addition of , every -exceptional divisor that is not in is in the relative stable base locus of . Finally, by assumption and the choice of , every log canonical center of dominates .
By assumption, the general fiber has a good minimal model. Thus, by [HMX18]*Theorem 1.9.1, it follows that has a relative good minimal model over a non-empty smooth affine open subset . Then, as by assumption there are no vertical log canonical centers, it follows from [HX13]*Theorem 1.1 that has a relative good minimal model over . Since every -exceptional divisor that is not in is in the relative stable base locus, any such divisor is contracted on the minimal model. This shows that the achieved model is a relative good minimal model of over in the sense of Birkar–Shokurov (see [LT22]*Definition 2.8); that it, it is a good minimal model where we allow to extract some log canonical places. But then, by [LT22]*Lemma 2.9, the existence of such model also implies the existence of a minimal model in the usual sense. In turn, this latter model is also good by [HMX]*Lemma 2.9.1, and the first part of the claim follows.
Now, assume that is relatively nef. Then, is a relative weak log canonical model for in the sense of [HMX]. Then, we conclude by [HMX]*Lemma 2.9.1 that is a relatively semi-ample model.
In the course of the proof, we used [HMX]*Lemma 2.9.1 in the relative setting. Notice that [HMX]*Lemma 2.9.1 is phrased for projective pairs. On the other hand, by [HX13]*Corollary 1.2, one can first take a projective closure of over a compactification of , and take a projective relative good minimal model. Then, by adding the pull-back of some divisor on , we can regard the relative good minimal model as a projective minimal model. Thus, it follows that is relatively semi-ample over , and it defines a morphism to the relative canonical model. ∎
3. Boundedness
The goal of this section is to prove that, if we fix a set of coefficients , a polynomial and a dimension , the corresponding set of -stable pairs is effectively log-bounded.
Proposition 3.1.
Fix an integer . Consider a locally stable family of relative dimension over a reduced scheme , with being -Cartier. Assume that there is an such that is stable. Then, the set of points such that is semi-ample is constructible.
Proof.
This follows immediately from Theorem 2.22. ∎
Theorem 3.2.
Fix an integer , a finite subset , and a polynomial . Then, the set of -stable pairs of dimension , polynomial and coefficients in is effectively log bounded.
Proof.
We proceed in several steps.
Step 1: In this step, we show that the -stable pairs of interest are log bounded.
From Lemma 2.15, there is an such that, for every -stable pair as in the statement, is a stable pair. Then, from [HMX], there is a bounding family of stable pairs of volume , coefficients in the finite set and dimension .
Step 2: In this step, we show that the -stable of pairs of interest are strongly log bounded. Furthermore, we may choose the family to be locally stable.
Since the set of coefficients involved is finite, up to taking finitely many copies of the family in order to assign coefficients to , we may find divisors and supported on such that restricts to fiberwise, and restricts to fiberwise. By [kol_new]*Lemma 4.44, up to replacing with a finite disjoint union of locally closed subsets, we can further assume that both and are locally stable. In particular, is -Cartier for any . Furthermore, by flatness, up to disregarding some irreducible components of , we can assume that for every . Finally, up to stratifying , we may assume that each irreducible component of is smooth; in particular, is well defined, and it follows that the pairs are strongly log bounded.
Step 3: In this step we finish the proof.
By construction, for the choice of , is ample on the general fibers. Thus, up to removing some proper closed subset of , we may assume that is a stable family. Thus, to conclude the proof it suffices to use Proposition 3.1, which guarantees that the set is constructible. ∎
4. The moduli functor
The goal of this section is to prove that is an algebraic stack. We begin by the following proposition:
Proposition 4.1.
The fibered category is a stack.
Proof.
Since our argument follows the same strategy in [Alp21]*Proposition 1.4.6, we only sketch the salient steps here. The role that in loc. cit. is the one of , for us is , where and are chosen such that is very ample with for and such that is surjective. These and can be chosen uniformly by Theorem 3.2.
The fact that isomorphisms are a sheaf in the étale topology of follows from descent as in [Alp21]*Proposition 1.4.6.
For proving that satisfies descent, we begin by the following observation. Consider an object of , and pick and as before. Then, up to replacing with some uniform multiple, from Remark 2.32 and from cohomology and base change, we can assume that is relatively very ample, and is a vector bundle over . Indeed, by the deformation invariance of and the vanishing of for , it follows that the sections of are deformation invariant, and hence is a vector bundle. Then, on every affine open trivializing , the morphism can be identified with , and the latter is an embedding as it is an embedding over fiber by fiber, by Nakayama’s lemma. This gives an embedding , and composing it with an embedding . Now the proof is analogous to the one in [Alp21]*Proposition 1.4.6. ∎
Theorem 4.2.
Fix an integer , a finite subset , and a polynomial . Then is an algebraic stack.
Proof.
Let denote the index of . We will proceed in several steps.
Step 1: In this step, we fix some invariants and consider a suitable Hilbert scheme parametrizing (among others) the total spaces of the -pairs of interest.
By Lemma 2.15, we may find a rational number such that is ample for every -stable pair with polynomial and coefficients in . Without loss of generality, we may assume that for a suitable . By Theorem 3.2, we may consider a weak -stable morphism with coefficients in , polynomial , and of relative dimension that is effectively log bounding for our moduli problem. By stratification of , we may assume that is smooth, and are flat, and that has fibers. Furthermore, by Proposition 2.37, also induces a family of stable pairs. In particular, we can also regard as a divisor, not only as a subscheme, and we will be free to take its multiples. Similarly, any natural multiple can be regarded as a subscheme, by considering the vanishing locus of . Then, there is such that, for every -stable pair in our moduli problem, we have:
-
(1)
and are Cartier;
-
(2)
is Cartier; and
-
(3)
is a very ample line bundle with vanishing higher cohomologies that embeds into .
Furthermore, up to taking a multiple, we may assume that and are integers, and
-
(1)
and are Cartier;
-
(2)
is Cartier; and
-
(3)
is a relatively very ample line bundle.
By boundedness and upper semi-continuity of the space of global sections, attains finitely many values. Then, we have finitely many polynomials such that the fibers of have Hilbert polynomial for some , for the relatively ample line bundle . We consider a union of Hilbert schemes for the polynomials , and we denote such a union with . Over , we have a universal family , where the fibers are closed subschemes of with Hilbert polynomial for some . Here, we observe that may actually attain finitely many distinct values, as we are assuming that each is embedded with a full linear series; on the other hand, we will work on one Hilbert scheme at the time, thus, by abusing notation, we will simply write .
Step 2: In this step, we highlight the strategy for the construction of the moduli functor.
We will construct our moduli functor as a subfunctor of a suitable relative Hilbert scheme, modulo the action of . For this reason, we will shrink to cut the locus of interest for our moduli problem. In doing so, we have to guarantee that this locus is locally closed and has a well-defined scheme structure. If we shrink to an open subset, there is no ambiguity in the scheme structure. On the other hand, if we need to consider a closed or locally closed subset, we need to show this choice has a well-defined scheme structure, which will be functorial in nature. Finally, we need to guarantee that, at each step, the locus we consider is invariant under the action of .
Step 3: In this step, we cut the locus parametrizing demi-normal schemes.
Since being is an open condition for flat and proper families [EGAIV]*Theorem 12.2.1, and since small deformations of nodes are either nodes or regular points, up to shrinking we may assume that the fibers of are and nodal in codimension one. That is, the fibers are demi-normal.
Step 4: In this step, we cut the locus parametrizing varieties embedded with a full linear series.
Let be a -pair in our moduli problem, and let denote its ideal sheaf in . Then, by assumption we have and the higher cohomologies of both sheaves vanish. Thus, if we consider the short exact sequence
it provides the following long exact sequence of cohomology groups
where the vanishing of follows from the surjectivity of the map and the vanishing of the higher order cohomologies of . By definition of Hilbert scheme, is flat, and the subschemes parametrized correspond to a flat quotient sheaf of . Thus, as we have a short exact sequence of sheaves where the last two terms are flat over the base, then so is the first term, which is the family of ideal sheaves of the subschemes of interest. Thus, upper semi-continuity of the dimension of cohomology groups applies, and we may shrink to the open locus parametrizing varieties with . This guarantees that, for every such variety , any automorphism of preserving is induced by an automorphism of .
Step 5: In this step, we introduce a relative Hilbert scheme, in order to parametrize the boundaries of the -pairs of interest.
Proceeding as in Step 1, for every -pair in our moduli problem, we may consider the Hilbert polynomial of with respect to . Here, recall that corresponds to fiberwise, hence the choice of Hilbert polynomial for rather than for . By effective log boundedness and generic flatness of , there exist finitely many such polynomials. As before, we will deal with one Hilbert polynomial at the time, and omit this choice from the notation.
Now, is projective over . In particular, if we pull back the ample line bundle to get a relatively very ample line bundle on , we can take the relative Hilbert scheme for the morphism , the line bundle , and the polynomial determined by (see [ACH11]*Ch. IX). This gives a scheme , together with an universal family . Then, as corresponds to on the elements of our moduli problem, every -pair of interest appears as a fiber of this family.
Step 6: In this step, we shrink to an open subset such that has codimension at least 2 along each fiber and such that the ideal sheaf of is relatively .
For every -pair , does not contain any component of the double locus of . Thus, by upper semi-continuity of the dimension of the fibers of a morphism, we may shrink to an open subset such that has codimension at least 2 along each fiber.
Now, by Step 3, all the fibers of are demi-normal. Thus, we may find an open subset such that the following properties hold:
-
(1)
for every , is a big open set in ;
-
(2)
; and
-
(3)
the fibers of have at worst nodal singularities.
Then, is a Gorenstein morphism, so, by [stacks-project]*Tag 0C08, is an invertible sheaf along . Furthermore, by [stacks-project]*Tag 062Y, the ideal sheaf is locally free along . Finally, up to shrinking , we may assume that is relatively .
Indeed, is flat over , as is flat over , and is flat over by definition of relative Hilbert scheme. Thus, is flat over , as it is the kernel of a surjection of flat sheaves. Then, we conclude, as being is an open condition for flat and proper families [EGAIV]*Theorem 12.2.1. Notice that, by [HK04]*Proposition 3.5, we have .
Notice that in this step we shrank twice, and both times the process is invariant under the natural action of , as the locus is characterized by properties of the fibers.
Step 7: In this step, we cut to a locally closed subset to ensure that is flat and over , commutes with base change for every , and that , , and are invertible sheaves. Here, if is not an integer, we set .
The first claim follows immediately from [kol08] applied to the sheaves and for . Thus, there is a stratification into functorial locally closed subsets of , which we denote by , where the above sheaves are flat over , they are , and they commute with base change. Then, by Remark 2.32, the sheaves , , and are invertible over each component .
Considering the union of the ’s produces another base , with a family and a closed subset as the one over , but such that , , and are Cartier and the formation of commutes with base change for every . Then, the formation of commutes with base change for every , since we can write and for , so we can write as a tensor product of a line bundle (namely ) and a relatively sheaf which commutes with base change (namely ).
Step 8: In this step, we shrink to an open subset parametrizing semi-log canonical pairs.
By the reductions in the previous steps, for every , the ideal sheaf determines a family of generically Cartier divisors, in the sense of [kol_new]*Definition 4.24. Indeed, this is obtained by dualizing the inclusion , as observed in [kol_new]*Definition 4.24. Then, as observed in [kol_new]*4.25, a family of generically Cartier divisors induces a well-defined family of divisors. Abusing notation, we denote by the family of divisors corresponding to .
Now, recall that taking the reduced structure does not change the topology. Thus, in the remainder of this step, we may assume that is reduced. Then, is a well defined family of pairs, and by [kol_new]*Corollary 4.45, there is an open locus where the fibers of are semi-log canonical. Also, since we took an open subset of , this choice is not affected by considering the reduced structure of .
Step 9: In this step, we use Kollár’s theorem 2.22 to shrink to a locally closed and functorial subscheme parametrizing the -stable pairs in our moduli problem.
Using Theorem 2.22, up to replacing with such a (functorial) locally closed subscheme of it, we can assume that is relatively semi-ample. Similarly, since being ample is an open condition, we can assume that is relatively ample. Now, recall that we have chosen for a chosen fixed . Notice that, the two former conditions guarantee that, for every , we have that is relatively ample, as we have . Then, we may fix such values and disregard all components of but the ones where, over , has prescribed value. By flatness, these self-intersections are locally constant, and thus this condition is open. For each , the self-intersection is prescribed by , up to the rescaling factor given by . Since we are prescribing values of a polynomial of degree , this guarantees that all the fibers correspond to -stable pairs with .
Step 10: In this step, we cut to a closed subset to ensure that the natural polarization coincides with .
By construction, for every -stable pair in our moduli problem, we have . Since is a line bundle and the natural polarization of coming from the original choice of Hilbert scheme restricts to fiberwise, by [Vie95]*Lemma 1.19 there is a locally closed subscheme where is linearly equivalent to the natural polarization of over .
Step 11: In this step, we show that there is an isomorphism .
Our argument follows closely [Alp21]*Theorem 2.1.11. First, we observe that all the cuts performed in the previous steps depend on properties that are invariant under the action of , thus the natural action of descends onto . Then, observe that from its construction, over there is a stable family of -stable pairs. This gives a morphism , and if we forget the embedding into , this descends to a morphism , where the superscript pre stands for prestack (see [Alp21]*Definition 1.3.12). This induces a map , which we now show is an isomorphism.
To show it is fully faithful, as in [Alp21], it suffices to check that is fully faithful. But is fully faithful since any isomorphism between two families of -stable pairs and over sends to , where we denoted by (resp. ) the ideal sheaves of (res. ) in (resp. ). This induces an unique isomorphism which sends to .
Since is a morphism of stacks, also essential surjectivity can be checked locally on . In particular, it suffices to check that if is a family of -stable pairs such that is free, then the morphism lifts to a morphism . This follows since if we pick an isomorphism then , and then from the functorial properties of it induces a morphism . ∎
Remark 4.3.
Observe that the stack is in fact Deligne–Mumford. Indeed, since we are working over a field of characteristic 0, it suffices to show that the automorphisms of the objects over the points are finite. But this follows since an automorphism of a -pair induces an automorphism of the stable pair , and those are finite from [KP17]*Proposition 5.5.
5. Properness of
The goal of this section is to prove that is proper. In particular, since in the definition of a -stable pair there are prescribed conditions on the scheme-theoretic structure of . Then, when proving that a moduli functor for -pairs satisfies the valuative criterion for properness, one needs to check that these scheme-theoretic properties are preserved. It is convenient to check that the flat limit of (recall that was introduced in Notation 2.10) is . This will be the content of the next proposition.
Proposition 5.1.
Let be the spectrum of a DVR with generic point and closed point . Consider a locally stable family such that is -Cartier. Then, for every , the sheaf is on every point . In particular:
-
(1)
the restriction is ; and
-
(2)
if we denote by the closed subscheme of with ideal sheaf , then is .
Proof.
First, observe that since is locally stable, cannot be a log canonical center for (see [kol_new]*Proposition 2.15). The statement is local, so up to shrinking , we can assume that is affine, and, since both and are -Cartier, for a certain . Then, for every , we have . In particular, if we apply [Kol13]*Theorem 7.20 where, with the notations of loc. cit., we take and if we denote by the divisor in loc. cit., , we conclude that is .
Now, we denote by the pull-back to of a uniformizer on . Since is and is not a zero divisor on , it is not a zero divisor for . Then,
is by [KM98]*Proposition 5.3.
Notice that is along the special fiber, since it is a flat and proper family of schemes over a smooth base (see [KM98]*Proposition 5.3). Then, by [Kol13]*Corollary 2.61, is along the special fiber. In particular, it is , so its generic points are the only associated points. Hence, if we denote by the closed subscheme of with ideal sheaf , then is flat. Then, if we pull back the exact sequence
to , the sequence remains exact and we get
The desired result follows again by [Kol13]*Corollary 2.61. ∎
Proposition 5.2.
Fix an integer , a finite subset , and a polynomial . Assume that is closed under sum: that is, if and , then . Let be a smooth affine curve, let be a distinguished closed point, and let . Let be a weak -stable morphism of relative dimension , polynomial , coefficients in , and constant part . Further, assume that the geometric generic fiber is normal. Then, is a -stable morphism, and, up to a finite base change , the fibration can be filled with a unique -stable pair of dimension , with polynomial and coefficients in .
Proof.
We proceed in several steps. In the following, will denote the index of .
Step 1: In this step, we show that the family of pairs admits a relative canonical model over .
By Proposition 2.37, the fact that is smooth and inversion of adjunction, it follows that is a log canonical pair. By Lemma 2.38 and the fact that, by definition, all the fibers admit a good minimal model, the pair admits a morphism to its relative canonical model over .
Step 2: In this step, we show that, up to a base change, can be compactified to a family of stable pairs .
This is accomplished in [kol_new]*Theorem 4.59.
Step 3: In this step, we choose a suitable -factorial dlt modification so that is a rational contraction over .
Since is ample over for , the divisors contracted by are contained in the support of . Then, by [Mor19]*Theorem 1, we can extract all of these divisors. Finally, we take a -factorial dlt modification of this model just constructed. We denote this model by . In writing , we define by first pulling back to a common resolution of and , then pushing it forward to , and finally taking its closure in . In particular, is horizontal over .
Step 4: In this step, we construct the compactification .
We consider the pair for , where the notation stands for horizontal over . Then, is a relative good minimal model for over . Then, by [HX13], admits a relative good minimal model over . We denote its relative canonical model over by .
First, we observe that . Notice that is horizontal over by assumption, so , which denotes the vertical components of , has to be supported on . Then, the claim follows, since is log canonical.
Now, we observe that is -Cartier. Indeed, is -Cartier, as it is the pull-back of . By construction, we have that is ample over , and in particular it is -Cartier. Then, is the difference of two -Cartier divisors.
Since is -Cartier, it follows that this canonical model is independent of , as . Furthermore, if , we have that is ample over , as it is ample over and is ample over .
Step 5: In this step, we show that is a -pair and that the central fiber of is a -stable pair with polynomial and coefficients in .
Recall that is independent of , and that is -Cartier. As is log canonical by construction, it follows that is a -pair.
By construction and by adjunction, is semi-log canonical. Furthermore, is ample for . Since and are -Cartier, the self-intersection is well-defined for every and , and does not depend on . Since the general fiber is -stable with polynomial , we have . The coefficients of are still in , since is closed under addition.
Step 6: In this step, we show that is a -stable morphism with constant part .
By construction, the fibers are proper. Since the base is a curve and every divisor is horizontal, all the morphisms are flat of the appropriate relative dimension. By Step 7, every fiber is a -stable pair. Then, as is -Cartier, every fiber of is semi-log canonical. Thus, by [kol_new]*Definition-Theorem 4.7, the morphism is locally stable.
Step 7: In this step, we show that the limit is unique.
From Theorem 3.2, there exists such that, for every -stable pair with coefficients in and polynomial , is a stable pair. Then, the claim follows from the separatedness of stable morphisms [kol_new]*2.49. ∎
Theorem 5.3.
Fix an integer , a finite subset , and a polynomial . Assume that is closed under sum: that is, if and , then . Let be a smooth affine curve, let be a distinguished closed point, and let . Let be a weak -stable morphism of dimension , polynomial , coefficients in , and constant part . Then, is a -stable morphism, and, up to a finite base change , the fibration can be filled with a unique -stable pair of dimension , with polynomial and coefficients in .
Proof.
By Proposition 5.2, we may assume that the geometric generic fiber is not normal. In the following, will denote the index of .
Let denote the normalization of , where also includes the conductor with coefficient 1. By assumption, is -Cartier. Then, by [Kol13]*Corollary 5.39, is -Cartier. Then, since is ample over , it follows that is -stable morphism with constant part , where the polynomial on each connected component depends on the original choice of .
Then, as has finitely many connected components, by Proposition 5.2, there is a finite base change such that the family can be filled with a unique -stable pair. To simplify the notation, we omit the base change , and we assume that the filling is realized over itself. We denote this family by . Then, we may find such that is a stable morphism. By [kol_new]*Lemma 2.54, also admits a compactification over obtained by gluing along some components of . By [Kol13]*Corollary 5.39, the divisor is -Cartier. Thus, we have that is -Cartier, as needed. Similarly, the coefficients of are in , by construction.
Now, we prove that the special fiber is a -stable pair. We already verified that is -Cartier. Now, we verify that is semi-log canonical. By construction, is semi-log canonical. Thus, by [Kol13]*Definition-Lemma 5.10, it suffices to show that the normalization of is log canonical. But then, this holds by construction, as its normalization coincides with the normalization of the (possibly disconnected) semi-log canonical pair . The same argument applied to the total space shows that is semi-log canonical.
Lastly, we need to check that is semi-ample. To this end, by adjunction, it suffices to show the stronger statement that is semi-ample over . By Proposition 5.2, this is true for . Then, the claim follows by [HX16]*Theorem 2. Then, since does not contain any irreducible component of the double locus of , it is immediate that the relative canonical model of coincides with the gluing of the relative canonical models of the irreducible components of , which in turn provides a stable family. We observe that the gluing of said ample models of is possible by [HX13]*§ 7 and the following fact: as the exceptional locus of the morphism from to the canonical model is contained in and does not contain any irreducible component of the conductor, the involution defined on the conductor via the normalization of naturally descends to the canonical model of . We observe that the involution defined on the conductor via the normalization of preserves the different, which is a necessary condition for the gluing of the ample models, by [Kol13]*Proposition 5.12.
To conclude, we need to show that . This follows from flatness over the base , as we have
where . This concludes the proof. ∎
Corollary 5.4.
Fix an integer , a finite subset , and a polynomial . Assume that is closed under sum: that is, if and , then . Then the algebraic stack is proper.
Proof.
It suffices to check that it satisfies the valuative criterion for properness. So assume that we have a family of -stable pairs over the generic point of the spectrum of a DVR , and we need to fill in the central fiber up to replacing with a ramified cover of it. Theorem 5.3 guarantees the existence and uniqueness of a -stable morphism extending , up to a ramified cover of . To check that satisfies condition (K) of Definition 2.19 we use Proposition 5.1 point (1). ∎
6. Relative canonical models over reduced base
Given a pair with semi-ample and big, one can take its canonical model. Similarly, from Definition 2.21, if one starts with a family of -stable pairs, one can take this canonical model in families. The goal of this section is to show that, over a reduced base, the condition in Definition 2.21 is a fiberwise condition: weak families of -stable pairs are actually families of -stable pairs.
Lemma 6.1.
Fix an integer , a finite subset , and a polynomial . Let denote the index of . Consider a weak -stable morphism with coefficients in and polynomial over a smooth scheme : . Then, there is a stable family of pairs such that:
-
(1)
there is a contraction ; and
-
(2)
we have .
In particular, is a -stable morphism.
Proof.
By Proposition 2.37 and the fact that is smooth, it follows that is a pair. Furthermore, by inversion of adjunction, this pair is semi-log canonical.
First, we assume that is normal. Then, by Lemma 2.38 and our assumptions, the canonical model of over exists. Since is smooth, this canonical model gives a family of stable pairs by [kol_new]*Corollary 4.57, and the result follows from how canonical models are constructed.
We now treat the case where is not normal. First, we normalize to get . As argued in the previous case, we can construct the canonical model of over . As above, this gives a family of stable pairs . From Kollár’s gluing theory (see [Kol13]*Ch. 5), there is an involution that fixes the different. We first show that this involution descends onto .
Recall that, by Lemma 2.13, the map does not contract any component of . In particular, for every irreducible component , there is an irreducible component birational to it, and we have the following diagram:
Observe now that
() a curve gets contracted by if and only if .
In particular, since the involution preserves the different, it preserves all the curves that are contracted by . Hence, the involution descends to an involution on .
We prove that preserves the different. Indeed, by Lemma 2.13, the only divisors contracted by are contained in , so the morphism is an isomorphism generically around each divisor not contained in . In particular, since the computation of the different is local,
Since preserves the different on , preserves the different on .
Then from [Kol13]*Ch. 5, we can glue to get an semi-log canonical pair , and let be the normalization. Now, recall that is a geometric quotient (see [Kol13]*Theorem 5.32 and the proof of [Kol13]*Corollary 5.33), so for any morphism such that , there is a unique morphism which fits in the obvious commutative diagram. Therefore, we have a morphism , and applying this result to and , we obtain a morphism .
To show (1) it suffices to observe the following two exact sequence:
| (7) |
Observe that and both and is birational. So and Therefore pushing forward the sequence (7) via we have
In particular, . Similarly we can tensor the sequence above by to deduce (2). ∎
Theorem 6.2.
Fix an integer and a finite subset . Let denote the index of . Assume that the fibers of have an canonical model. Let be a weak -stable morphism of dimension , with coefficients in , and polynomial , over a reduced connected base , and assume that there is an open dense subscheme with:
-
(1)
a stable family of pairs of relative dimension ; and
-
(2)
a contraction such that .
Then, there there is an such that for every and every we have
Moreover, if we define , then:
-
•
there is a family of divisors such that the pair is a stable family extending ;
-
•
there is a contraction over that extends ; and
-
•
.
In particular, is a -stable morphism.
Remark 6.3.
Observe that since is -Cartier, we can define its pull-back as a -Cartier divisor.
Proof.
We proceed in several steps.
Step 1: In this step, we make some preliminary considerations and set some notation.
By Proposition 2.37, we have that and are -Cartier. Then, observe that for every , the volumes of the pairs and agree. Indeed, since is a weak -stable morphism, and are nef. Thus, their volumes are computed by the -fold self-intersection, which is independent of . But from condition (2) the morphisms and have connected fibers, and we have and . Then, by Remark 2.4, the volumes of (resp. ) and (resp. ) agree. Let then be the volume of any fiber of and let be a natural number such that divides and, for every stable pair of dimension , volume and coefficients in , the line bundle is very ample and the higher cohomologies of all of its natural multiples vanish. Notice that exists by [HMX18]*Theorem 1.2.2. Then, we set . Up to replacing with a multiple, we may further assume that is Cartier.
Step 2: In this step, we show that the theorem holds if is a smooth curve.
From Lemma 6.1, we can construct a family of stable pairs with a contraction . First, observe that over . Indeed, consider the reflexive sheaves and . By construction, the fibers of and belong to the moduli problem of stable pairs with volume and coefficients in the finite set . Thus, by [kol_new]*Theorem 5.8.(4) and the choice of , and are line bundles. In order to apply [kol_new]*Theorem 5.8.(4), notice that we know that and are line bundles away from the exceptional locus of and , which are big open subsets restricting to big open subsets fiberwise.
Since and have connected fibers, we have
But both and have connected fibers so, by the projection formula, for every , we have
But and are ample over , so and are isomorphic as -schemes, as they are the relative Proj of the same sheaf of graded algebras. In particular, the three final claims of the theorem hold if we consider the family .
Thus, we are left with proving that for every and every , we have . But since , we have . Thus, by Remark 2.4, we have for all . Then, the latter is locally constant from the assumptions on , as by the vanishing of the higher cohomologies we have , and the Euler characteristic of is independent of . Now the desired statement follows from [Mum74]*Corollary 2, page 50.
Step 3: In this step, we return to the general case and we show that for every , the morphism is constant and the algebra is finitely generated.
Observe that the claim holds for every . Indeed, the following diagram commutes:
which from point (2) guarantees . Then, in this case, we can conclude as at the end of Step 2. Furthermore, the finite generation follows from the fact that is very ample and is the Proj of its associated graded ring.
Now, we treat the case when . Consider a smooth curve with a map . Assume that the generic point of maps into . Notice that any point is contained in the image of such a curve. Then, for any point , we have the following diagram:
Since both squares are fibered squares, the big rectangle is a fibered square. In particular, since we want to prove that is constant, and since we know it is constant as long as , it suffices to check that for every such the functions are constant. Now, this follows by Step 2. Similarly, the finite generation of follows from Step 2.
Step 4: In this step, we construct the model and the morphism .
By Step 3 and cohomology and base change (see [Mum74]), the sheaves commute with the restriction to points. In particular, the algebra is finitely generated since it is finitely generated when restricted to every point . So we can consider .
As observed in Step 3, the pluri-sections of are deformation invariant. As is semi-ample for every by our assumptions, it follows that is relatively semi-ample. In particular, we have a morphism . Furthermore, this implies the equality for every .
As already discussed, this construction commutes with base change. In particular, for every , for checking properties of we can consider a smooth curve which sends the generic point to and the special one to , and first pull back to and then restrict it to :
The advantage is that now we can apply the results of Step 2 to .
Since we have and is relatively ample over , it follows that and that extends . Furthermore, since is reduced, the construction commutes with base change, and each fiber is reduced, it follows that is reduced.
Step 5: In this step, we show that is a contraction and we construct the divisor .
We first prove that is a contraction. We denote by the locus where is an isomorphism. Then it follows from Lemma 2.13:
() for every fiber , the complement of has codimension at least 2 in and it does not contain any irreducible component of the conductor of .
Consider now the inclusion , which induces the injective map . We can push this sequence forward via and we obtain . But is the inclusion , so is reflexive from [HK04]*Corollary 3.7, and it is isomorphic to from [HK04]*Proposition 3.6.2. In particular, this gives an injective map . One can check that this is the inverse of the canonical morphism , so in particular the latter is an isomorphism.
Consider the ideal sheaf of , and consider the inclusion
Then, as is a contraction, if we push it forward via , we get
In particular, is an ideal sheaf on . We denote by the closed subscheme with ideal sheaf .
A priori, may not be pure dimensional, however, consider the intersection between and the locus in where is -Cartier: we denote this locus with . Observe that, since is an isomorphism, we can identify with a subset of . Moreover, since is Cartier on codimension one point of and is a big open subset, the locus contains all the codimension one points of for every . Then, we consider , and we define to be the closure of in .
Step 6: In this step, we show that is a well defined family of pairs.
By construction, it is immediate that is a relative Mumford divisor in the sense of [kol19s]*Definition 1. Indeed, the three conditions of [kol19s]*Definition 1 are now clear, since they hold on .
Thus, we just need to check that is flat. By pulling back along a smooth curve through , it follows from Step 1 that all the fibers of are reduced and equidimensional. Thus, is an equidimensional morphism with reduced fibers over a reduced base, so [kol_new]*Lemma 10.58 applies.
Step 7: In this step, we show that is a stable family of pairs and .
Since is a well defined family of pairs and its fibers belong to a prescribed moduli problem for stable pairs, we can argue as in the proof of Proposition 2.37 to conclude that is -Cartier. In particular, is a stable family of pairs.
By construction, we have that . Furthermore, we have that the equality holds over . Then, by construction, all the exceptional divisors of dominate , as they are contained in the support of . Thus, as is -Cartier and so is well defined, it follows that . ∎
Lemma 6.4.
With the notation and assumptions of Theorem 6.2, assume that is an affine curve and that there is a stable pair such that . Then there are finitely many isomorphism classes of -stable pairs in the fibers of .
Proof.
The proof is analogous to the proof of [ABIP]*Claim 6.2, we summarize here the most salient steps of the argument.
Step 1: Using Kollár’s gluing theory and the fact that stable pairs have finitely many isomorphisms, up to normalizing and disregarding finitely many points on , we can assume that (and therefore also ) is normal. This is achieved in [ABIP]*Lemma 6.5. In particular, is the canonical model of .
Step 2: We observe that the divisors contracted by have negative discrepancies and can be extracted by a log resolution of the form , where is a log resolution. This is achieved in [ABIP]*Proposition 6.13.
Step 3: To conclude, we observe that all the fibers of are isomorphic in codimension 2.
But two stable pairs and which are isomorphic in codimension 2 must be isomorphic. Indeed, if is the open subset where they agree, and (resp. ) is the log-canonical divisor (resp. ) then
Therefore and are of the same graded algebra, so they are all isomorphic. ∎
7. Projectivity of the moduli of stable pairs
The goal of this section is to provide a different proof of the projectivity of the moduli of stable pairs, established in [KP17], using -pairs. As a consequence, we also deduce the projectivity of the coarse moduli space of .
We first construct a suitable polarization on the base of families of -stable pairs with additional technical assumptions, see Theorem 7.6. Then, we use this result to deduce the projectivity of the moduli of stable pairs, see Corollary 7.7. Lastly, we deduce the projectivity of the coarse moduli space of from Corollary 7.7 and Lemma 2.15, see Corollary 7.9. In particular, one could alternatively deduce Corollary 7.9 from Lemma 2.15 and [KP17].
Lemma 7.1.
With the notation of Theorem 6.2, we will denote by the -stable morphism, and by the resulting stable family of Theorem 6.2. Assume this -stable morphism is a family of -stable pairs, and let be the smallest positive integer such that is Cartier. Then, there is divisible by such that, for every positive multiple of , the sheaf satisfies the following properties:
-
(a)
is Cartier;
-
(b)
for every and ;
-
(c)
gives an embedding ;
-
(d)
is relatively ample; and
-
(e)
for every , there is a section such that has codimension 1 in each irreducible component of and the scheme-theoretic image of restricted to is contained in .
Remark 7.2.
Observe that the definition of is given in Notation 2.29. In our case, we can still define even if was a -stable morphism instead, since the base is reduced, and Lemma 7.1 would go through verbatim. Since we will use Lemma 7.1 only in the case in which is a family of -stable pairs, for simplicity we stick with the family case.
Proof.
By construction, is relatively ample. Thus, for a sufficiently divisible , properties (a) to (d) are satisfied for .
Then, we can achieve (a) to (d) since if satisfies any condition between (a) to (e), then for every also satisfies the same condition. We only need to check that up to choosing divisible enough, also (e) holds.
Let for a which satisfies (a)-(d), and we need to show that satisfies (a)-(e) for some . First observe that from cohomology and base change and point (b), for every we have .
Let be the scheme-theoretic image of . Up to replacing with a locally closed stratification , we can assume that is flat, and is free. Then, we define , the first projection , the second projection , and . Then, consider the exact sequence
We twist the exact sequence above by and we obtain
Since is flat, also is flat over . Therefore since is relatively ample, up to choosing big enough, the following is an exact sequence on :
By [ACH11]*Corollary 4.5 (iii), there is such that for every and any , the multiplication map
is surjective. Then for every , also
is surjective. From [ACH11]*Corollary 4.5 (iv) (or cohomology and base change), for and , we have:
In other terms, for every the ideal is generated in degree . Then for every there is a section (and therefore also a section if ) such that does not contain any irreducible component of and the scheme-theoretic image of restricted to is contained in .
Therefore, for the locally closed subset , we may choose , where was chosen at the beginning of the proof to satisfy (a)-(d) and thus define . Then, by noetherianity, there are finitely many locally closed subsets in the decomposition of , and we may thus choose as the least common multiple of the defined on each . ∎
Corollary 7.3.
With the notation and assumptions of Lemma 7.1, for every and every , there is a global section of that is not zero on the generic points of but its maps to 0 via the restriction map .
Proof.
From Lemma 7.1, for every , there is a section that does not vanish on the generic points of but vanishes on the restriction to of the scheme theoretic image of . From point (b) of Lemma 7.1, is a vector bundle on , so there is an open subset such that is free. Then, we can extend the section to a global section . But is a contraction, so is still a contraction, and . In particular, consider the section that is the pull-back of . Then, does not vanish on the generic points of , as by Lemma 2.13 those are in bijection with the generic points of . Moreover, it vanishes along : this is a local computation. Indeed, if we replace and with appropriate open subsets , ; and is the generator for the ideal sheaf of , we have the following commutative diagram:
Since the image of vanishes in , so does the image of .
Therefore for every , the section will vanish along . In other terms, if we look at the exact sequence
we have an element of (namely, the restriction of to ) that does not vanish along the generic points of but maps to 0 via . ∎
Lemma 7.4.
With the notation and assumptions of Lemma 7.1, we will denote by the family of -stable pairs, and by the resulting stable family of Theorem 6.2. Let be a multiple of , where is as in Lemma 7.1. Also, set .
Then, there is an such that, for every , the sheaf is a vector bundle, and fits in an exact sequence as follows:
Proof.
First, from relative Serre’s vanishing and Lemma 7.1 (b) and (d), there is an integer such that for every and , we have
-
•
for ; and
-
•
for .
Fix a positive integer . We begin by considering the following exact sequence, where is the locally free ideal sheaf :
We push it forward via , and from the first bullet point and from cohomology and base change, , so we have:
| (8) |
Recall that there is a contraction such that, if we denote , then . In particular
| (9) |
Now, pick . From the first bullet point and from cohomology and base-change, we have
| (10) |
and from the second bullet point, equation (9) and the fact that is a contraction,
| (11) |
Now, we tensor the sequence (8) by , and we obtain
From the natural adjunction between pull-back and push-forward, we have the following commutative diagram:
Observe that the exactness of the map follows from the first bullet point. Then, from diagram chasing, also is injective. In particular,
However the right-hand side does not depend on . Therefore also the left-hand side does not depend on , so is a vector bundle. ∎
Corollary 7.5.
With the notation and assumptions of Lemma 7.4, there is an isomorphism
Theorem 7.6.
With the notation and assumptions of Lemma 7.4, further assume that the map given by the family of -stable pairs is finite. Then, for divisible enough, the line bundle is ample.
Proof.
We plan on using Kollár’s Ampleness Lemma, see [Kol90]. By Corollary 7.5, it suffices to show that is ample.
Consider the vector bundles
for appropriate choices of , , and .
When , there is a morphism given by the sum of the multiplication morphisms
and the composition of with the surjection in the exact sequence of Lemma 7.4, denoted by
First, observe that for divisible enough, the vector bundles and are nef. Indeed, from the assumptions of Lemma 7.4, the formation of and commutes with base change, so we can assume that the base is a smooth curve. Then the statement follows from [Fuj18]*Theorem 1.11. Therefore, their symmetric powers and sum are nef, so is nef for divisible enough.
Since is relatively ample and , there is a such that, for every , the map is surjective. For the same reason, if we choose large enough so that is relatively very ample, also is surjective for sufficiently large (see [ACH11]*Corollary 4.5). Then, also is surjective since it is the composition of surjective morphisms. Therefore we have a surjection . We denote by the structure group of . This gives a map of sets
where (resp. ) is the rank of (resp. ), and where for a stack we denote by its associated topological space. If we show that has finite fibers, then the theorem follows from [Kol90]*Ampleness Lemma, 3.9.
Consider two points , that map to the same point via . Over the point we have the surjection , and similarly we have one denoted by over . Choose two isomorphisms
Similarly we define and for the same isomorphisms over . This gives an isomorphism
Since and map to the same point via , using the identifications above, there is an element such that .
In particular, we can choose a basis for by first choosing a basis for each summand of the left-hand side, choosing the same basis for the first and third summand. Then, since is a direct sum of vector bundles, in this basis the linear transformation is block diagonal:
In particular, will send (resp. ) to (resp. ). But the kernel of corresponds to the symmetric functions of degree that vanish on a subvariety of isomorphic to , which we will still denote by . From [ACH11]*Corollary 4.5, up to choosing big enough, this kernel generates a graded ideal that corresponds to a subvariety isomorphic to . The same conclusion holds for , since there is a linear transformation (given by a block of the matrix ) that induces an isomorphism . So, in particular, this linear transformation induces a map of projective spaces, that gives an isomorphism .
By Corollary 7.3, contains a function that does not vanish along the generic points of the irreducible components of but, if pulled back via , it vanishes along . Therefore, the zero locus of the polynomials in has codimension 1 in and the union of its irreducible components of codimension 1, which we denote by , contains . Since has finitely many irreducible components of codimension one and the coefficients of are in the finite set , the divisor is determined up to finitely many possible choices of prime divisors and coefficients.
A similar conclusion holds by replacing with . Since in the description of the block at position (1,1) is the same as the block at position (3,3), we also know that .
Therefore the fiber of corresponds to stable pairs in our moduli problem where , . But there are finitely many such subvarieties, and since is finite by our assumptions, the fiber has to be finite. ∎
Corollary 7.7.
Consider a proper DM stack which satisfies the following two conditions:
-
(1)
for every normal scheme , the data of a morphism is equivalent to a stable family of pairs with fibers of dimension , volume and coefficients in ; and
-
(2)
there is such that, for every , there is a line bundle on such that, for every morphism as above, .
Then, for divisible enough, descends to an ample line bundle on the coarse moduli space of for every . In particular, the coarse moduli space of is projective.
Remark 7.8.
As a concrete example of Corollary 7.7, one can consider to be any moduli space of KSB-stable pairs , such that, if we denote by its universal family, then is -ample for divisible enough. Indeed, in this case, by cohomology and base change is a vector bundle for divisible enough, its formation commutes with base change from cohomology and base change, and the formation of the determinant commutes with base change as well.
Proof.
The argument is divided into two steps. We denote by the coarse moduli space of , and let be an irreducible component of it.
Step 1. There is a -stable morphism which satisfies the following conditions:
-
(1)
is normal and projective,
-
(2)
The map is finite,
-
(3)
There is a dense open subset and a stable family which satisfies the assumptions of Theorem 6.2, and
-
(4)
The family induces a map which dominates .
Consider the generic point of , and consider the stable pair it corresponds to. By Lemma 2.16, there is the spectrum of a field and a -stable family over it such that the canonical model of is . The family induces a morphism , and we can take to be the closure of its image. This is still a proper DM stack, it admits a finite and surjective cover by a scheme with a normal and projective scheme using [LMB18]*Theorem 16.6, Chow’s lemma and potentially normalizing. Up to replacing with a finite cover of it, we can lift to , and consider an irreducible component of containing the image of .
The morphism induces a -stable family , and by construction its generic fiber admits an canonical model. Then such an canonical model can be spread out: there is an open subset and a family as in Theorem 6.2, and the generic fiber is isomorphic to . Therefore the image of the corresponding map contains the generic point of , and since both and are irreducible, is dominant.
Step 2. The map extends to a finite map .
The extension follows from Theorem 6.2. To prove that is finite, we can use Lemma 6.4. Indeed, if it was not finite, there would be a curve contained in a fiber of . But then, up to replacing with an open subset of it, there would be a stable pair and a -stable family as in Lemma 6.4. Therefore there would be finitely many isomorphism types of -stable pairs in the fibers of : this contradicts point (2) above.
End of the proof. From Theorem 7.6, up to replacing with a multiple (which depends only on ), the line bundle is ample. But a multiple of descends to a line bundle on . In other terms, there is a line bundle on whose pull-back is for a certain . Therefore is ample since it is ample once pulled-back via the finite map . But if a line bundle is ample when restricted to each irreducible component, it is ample. ∎
Corollary 7.9.
The stack admits a projective coarse moduli space.
Proof.
We will denote by the coarse moduli space of , which exists from Keel–Mori’s theorem, and with the one of the normalization of .
Consider such that is ample for every -pair parametrized by . Recall that such an exists from Lemma 2.15 (or from boundedness). In particular, if we denote with the universal family, for divisibie enough the formation of commutes with base change and an high enough power of descends to a line bundle on . Therefore to prove that is ample, it suffices to replace (resp. ) with (resp. the pull-back of via ).
Consider a proper DM stack as in Corollary 7.7. When has coefficient the pairs parametrized by are stable of volume , so we have a map which is finite, as different points of parametrize different -pairs, and the normalization is a finite morphism. From Corollary 7.7, the formation of commutes with base change, so pulls back to . Then is ample as it is the pull-back of an ample line bundle via a finite morphism.∎