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

Moduli of \mathbb{Q}-Gorenstein pairs and applications

Stefano Filipazzi EPFL, SB MATH CAG, MA C3 625 (Bâtiment MA), Station 8, CH-1015 Lausanne, Switzerland [email protected]  and  Giovanni Inchiostro University of Washington, Department of Mathematics, Box 354350, Seattle, WA 98195 USA [email protected]
Abstract.

We develop a framework to construct moduli spaces of \mathbb{Q}-Gorenstein pairs. To do so, we fix certain invariants; these choices are encoded in the notion of \mathbb{Q}-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 12\frac{1}{2}. 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.
SF was partially supported by ERC starting grant #804334.

1. Introduction

Since the seminal work of Mumford on the moduli space g\mathcal{M}_{g} of smooth curves with g2g\geq 2 and its compactification ¯g\overline{\mathcal{M}}_{g}, significant progress has been made in understanding its higher-dimensional analog, namely the moduli space of stable varieties. On a first approximation, those are \mathbb{Q}-Gorenstein varieties with relatively mild singularities, and such that KXK_{X} 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 g\mathcal{M}_{g} is given by the moduli space of nn-pointed curves and its compactification, denoted by g,n\mathcal{M}_{g,n} and ¯g,n\overline{\mathcal{M}}_{g,n}, respectively. A first attempt to generalize the notion of stable pointed curve is to consider mildly singular pairs (X,D)(X,D) (specifically, semi-log canonical pairs) where DD is a reduced divisor, such that KX+DK_{X}+D 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 \mathbb{Q}-Gorenstein, a stable pair might not be, so if one is interested in \mathbb{Q}-Gorenstein pairs, the aforementioned formalism might not give the desired moduli space. In fact, one cannot simply consider the \mathbb{Q}-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 \mathbb{Q}-Gorenstein pairs (X,D)(X,D). We observe that, in [Ale15], Alexeev already considered certain moduli spaces of \mathbb{Q}-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 KX+DK_{X}+D, requiring only that KX+(1ϵ)DK_{X}+(1-\epsilon)D is ample for 0<ϵ10<\epsilon\ll 1. We also impose a numerical condition on the intersection number (KX+tD)dim(X)(K_{X}+tD)^{\dim(X)}, analogous to the genus condition for curves, which we encode via a polynomial function p(t)p(t). This condition guarantees the boundedness of our moduli problem. These choices are encoded in the notion of \mathbb{Q}-stable pair, a pair on which both KXK_{X} and DD are \mathbb{Q}-Cartier, (Defintion 2.7) and lead to the following:

Theorem 1.1.

Fix an integer nn\in\mathbb{N} and a polynomial p(t)[t]p(t)\in\mathbb{Q}[t]. Then, there is a proper Deligne–Mumford stack n,p,1\mathscr{F}_{n,p,1}, with projective coarse moduli space, parametrizing \mathbb{Q}-stable (\mathbb{Q}-Gorenstein) pairs (X;D)(X;D) of dimension nn and with polynomial p(t)p(t) and reduced boundary.

In the statement of Theorem 1.1, the subscript 1 in n,p,1\mathscr{F}_{n,p,1} means that DD is a reduced divisor.

While it is very natural to consider the divisor DD with its reduced structure, the framework of the Minimal Model Program highlights the importance of the use of fractional coefficients for the divisor DD. For example, in the case of curves, one can replace ¯g,n\overline{\mathcal{M}}_{g,n} with a weighted version, where the markings can attain any fractional value in (0,1](0,1]. This was accomplished by Hassett in [Has03], and this approach leads to different compactifications of g,n\mathcal{M}_{g,n}.

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 (X,D)(X,D) cannot be reduced to a deformation of the total space XX and a deformation of the divisor DD. This naïve expectation is only true if the coefficients of DD are all strictly greater than 12\frac{1}{2} (see [Kol14]), in which case a deformation of (X,D)(X,D) over a base curve induces a flat deformation of DD. 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 12\frac{1}{2}, it is possible to define a family of stable surface pairs (𝒳,𝒟)𝔸1(\mathcal{X},\mathcal{D})\rightarrow\mathbb{A}^{1} such that the flat limit of 𝒟\mathcal{D} 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 n,p,1\mathscr{F}_{n,p,1} to an analogous moduli functor n,p,I\mathscr{F}_{n,p,I} 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 S1S_{1} morphism over the base, thus significantly simplifying the definition of the classical moduli functor. The S1S_{1} condition prevents the existence of embedded points, while the polynomial p(t)p(t) 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 nn\in\mathbb{N}, a finite subset I(0,1]I\subseteq(0,1]\cap\mathbb{Q}, and a polynomial p(t)[t]p(t)\in\mathbb{Q}[t]. Assume that II is closed under sum: that is, if a,bIa,b\in I and a+b1a+b\leq 1, then a+bIa+b\in I. Then there is a proper Deligne–Mumford stack n,p,I\mathscr{F}_{n,p,I} with projective coarse space, parametrizing \mathbb{Q}-stable pairs of dimension nn and coefficients in II.

After the completion of this work, we learnt that Kollár has also developed an analogous machinery allowing pairs (X;D1,,Dk)(X;D_{1},...,D_{k}) where the kk 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 \mathbb{Q}-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 𝒦n,v,I\mathcal{K}_{n,v,I} that satisfies the following two conditions:

  1. (1)

    for every normal scheme SS, the data of a morphism f:S𝒦n,v,If\colon S\to\mathcal{K}_{n,v,I} is equivalent to a stable family of pairs q:(𝒴,𝒟)Bq\colon(\mathcal{Y},\mathcal{D})\to B with fibers of dimension nn, volume vv and coefficients in II; and

  2. (2)

    there is m0m_{0}\in\mathbb{N} such that, for every kk\in\mathbb{N}, there is a line bundle k\mathcal{L}_{k} on 𝒦n,v,I\mathcal{K}_{n,v,I} such that, for every morphism ff as above, fkdet(q(ω𝒴/B[km0](km0𝒟)))f^{*}\mathcal{L}_{k}\cong\det(q_{*}(\omega_{\mathcal{Y}/B}^{[km_{0}]}(km_{0}\mathcal{D}))).

Then, the coarse moduli space of 𝒦n,v,I\mathcal{K}_{n,v,I} 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 \mathbb{Q}-stable pairs, all the needed sheaves remain functorial with respect to base change. Indeed, flatness and the S1S_{1} 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 \mathbb{Q}-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 \mathbb{Q}-stable pair and extending several statements from pairs to \mathbb{Q}-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 \mathbb{Q}-stable pairs.

In § 4 and § 5, we analyze the moduli functor n,p,I\mathscr{F}_{n,p,I}. In particular, in § 4 we show that n,p,I\mathscr{F}_{n,p,I}, which a priori is only a category fibered in groupoids over Sch/k\mathrm{Sch}/k, is a Deligne–Mumford stack. Then, in § 5 we show that n,p,I\mathscr{F}_{n,p,I} is proper. Thus, § 4 and §  5 settle Theorem 1.2, except for the projectivity part.

In § 6, we analyze the notion of family of \mathbb{Q}-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.

In § 7 we conclude our work. We use the ampleness lemma to show that the coarse space of n,p,I\mathscr{F}_{n,p,I} is projective, and we use the results from § 6 to descend the analogous statement to the moduli space of stable pairs.

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 (X,D)(X,D), KXK_{X} and DD need not be \mathbb{Q}-Cartier. In this case, we cannot expect a deformation of (X,D)(X,D) to induce a flat deformation of DD, 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 (X,D)(X,D) do induce flat deformations of DD. In [Ale06, Ale15], Alexeev considered pairs (X,Δ=i=1laiDi)(X,\Delta=\sum_{i=1}^{l}a_{i}D_{i}), where each DiD_{i} is prime and the coefficients aia_{i} are irrational and \mathbb{Q}-linearly independent. This assumption can be thought as a “general choice” of the coefficients of DD, 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 [0,1]l[0,1]^{l}. These chambers have rational vertices, determined by the conditions of (X,D)(X,D) being semi-log canonical and of KX+DK_{X}+D 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 DD are strictly greater than 12\frac{1}{2}, deformations of (X,D)(X,D) lead to flat deformations of DD. Thus, even though DD is not necessarily \mathbb{Q}-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 DD is \mathbb{Q}-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 KX+DK_{X}+D for the flatness of the deformations of DD. Thus, while the deformations of the boundary are easier to understand than the general framework of [kol19s], KX+DK_{X}+D 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 DD and allows for independent perturbations of different components of DD. In this way, he reconciles Alexeev’s work with our work, thus showing that the case of “DD with rational coefficients and KX+DK_{X}+D big and semi-ample” we consider can be thought of as a limit case of the “general coefficients with KX+DK_{X}+D 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 \mathbb{Q}-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 \mathbb{C}. A birational morphism between non-normal schemes will be a morphism f:XYf\colon X\to Y with a dense open subset UYU\subseteq Y such that f1(U)f^{-1}(U) is dense and f1(U)Uf^{-1}(U)\to U is an isomorphism.

2.2. Contractions

A contraction is a projective morphism f:XZf\colon X\rightarrow Z of quasi-projective varieties with f𝒪X=𝒪Zf_{\ast}\mathcal{O}_{X}=\mathcal{O}_{Z}. If XX is normal, then so is ZZ.

2.3. Divisors

Let 𝕂\mathbb{K} denote \mathbb{Z}, \mathbb{Q}, or \mathbb{R}. We say that DD is a 𝕂\mathbb{K}-divisor on a variety XX if we can write D=i=1ndiPiD=\sum_{i=1}^{n}d_{i}P_{i} where di𝕂{0}d_{i}\in\mathbb{K}\setminus\{0\}, nn\in\mathbb{N} and the PiP_{i} are prime Weil divisors on XX for all i=1,,ni=1,\ldots,n. We say that DD is 𝕂\mathbb{K}-Cartier if it can be written as a 𝕂\mathbb{K}-linear combination of \mathbb{Z}-divisors that are Cartier. The support of a 𝕂\mathbb{K}-divisor D=i=1ndiPiD=\sum_{i=1}^{n}d_{i}P_{i} is the union of the prime divisors appearing in the formal sum Supp(D)=i=1nPi\mathrm{Supp}(D)=\sum_{i=1}^{n}P_{i}.

In all of the above, if 𝕂=\mathbb{K}=\mathbb{Z}, we will systematically drop it from the notation.

Given a prime divisor PP in the support of DD, we will denote by μP(D)\mu_{P}(D) the coefficient of PP in DD. Given a divisor D=μPi(D)PiD=\sum\mu_{P_{i}}(D)P_{i} on a normal variety XX, and a morphism π:XZ\pi\colon X\to Z, we define

Dvπ(Pi)ZμPi(D)Pi,Dhπ(Pi)=ZμPi(D)Pi.D^{v}\coloneqq\sum_{\pi(P_{i})\subsetneqq Z}\mu_{P_{i}}(D)P_{i},\ D^{h}\coloneqq\sum_{\pi(P_{i})=Z}\mu_{P_{i}}(D)P_{i}.

Let D1D_{1} and D2D_{2} be divisors on XX. We write D1𝕂,ZD2D_{1}\sim_{\mathbb{K},Z}D_{2} if there is a 𝕂\mathbb{K}-Cartier divisor LL on ZZ such that D1D2𝕂fLD_{1}-D_{2}\sim_{\mathbb{K}}f^{\ast}L. Equivalently, we may also write D1𝕂D2/ZD_{1}\sim_{\mathbb{K}}D_{2}/Z, or D1𝕂D2D_{1}\sim_{\mathbb{K}}D_{2} over ZZ. If 𝕂=\mathbb{K}=\mathbb{Z}, we omit it from the notation. Similarly, if Z=Spec(k)Z=\mathrm{Spec}(k), where kk is the ground field, we omit ZZ from the notation.

Let π:XZ\pi\colon X\rightarrow Z be a projective morphism of normal varieties. Let D1D_{1} and D2D_{2} be two 𝕂\mathbb{K}-divisors on XX. We say that D1D_{1} and D2D_{2} are numerically equivalent over ZZ, and write D1D2/ZD_{1}\equiv D_{2}/Z, if D1.C=D2.CD_{1}.C=D_{2}.C for every curve CXC\subset X such that π(C)\pi(C) is a point. In case the setup is clear, we just write D1D2D_{1}\equiv D_{2}, omitting the notation /Z/Z.

2.4. Non-normal varieties and pairs

There are two important generalizations of the notion of normal variety.

Definition 2.1.

An S2S_{2} 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 XXX^{\prime}\rightarrow X is called a partial seminormalization if XX^{\prime} is reduced and, for every point xXx\in X, the induced map k(x)k(red(g1(x)))k(x)\rightarrow k(\mathrm{red}(g^{-1}(x))) is an isomorphism. There exists a unique maximal partial seminormalization, which is called the seminormalization of XX. 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 p:XYp\colon X\to Y be a proper surjective morphism with connected fibers, with YY seminormal and XX reduced. Then p𝒪X=𝒪Yp_{*}\mathcal{O}_{X}=\mathcal{O}_{Y}.

Remark 2.4.

Observe that, in situation of Lemma 2.3, by the projection formula it follows that, for any line bundle LL on YY, we have H0(Y,L)=H0(X,pL)H^{0}(Y,L)=H^{0}(X,p^{*}L).

Proof.

We can take the Stein factorization X𝑞Z𝑎YX\xrightarrow{q}Z\xrightarrow{a}Y. Since XX is reduced, then ZZ is reduced. Observe that q𝒪X=𝒪Zq_{*}\mathcal{O}_{X}=\mathcal{O}_{Z}, so the desired result follows if we can show that aa is an isomorphism.

Since the composition p=aqp=a\circ q has connected fibers and q:XZq\colon X\to Z is surjective, then a:ZYa\colon Z\to Y has connected fibers. Since aa is finite, it is injective. It is also surjective since pp is surjective, so aa is a bijection. Then, we can take the seminormalization ZsnZZ^{sn}\to Z and consider the composition ZsnZYZ^{sn}\to Z\to Y. This is a bijective morphism since it is a composition of bijections, and it is proper since aa is proper and ZsnZZ^{sn}\to Z is proper. Then, the composition ZsnYZ^{sn}\to Y is an isomorphism, since the source and the target are seminormal.

Now, on the topological space given by ZZ we have the following morphsims of sheaves:

𝒪Ya#𝒪Z𝑏𝒪Zsn.\mathcal{O}_{Y}\xrightarrow{a^{\#}}\mathcal{O}_{Z}\xrightarrow{b}\mathcal{O}_{Z^{sn}}.

The composition is surjective, so bb is surjective. Moreover, since ZZ is reduced, the map 𝒪Z𝑏𝒪Zsn\mathcal{O}_{Z}\xrightarrow{b}\mathcal{O}_{Z^{sn}} is also injective. Then it is an isomorphism, so ZZsnYZ\cong Z^{sn}\cong Y. ∎

2.5. Divisorial sheaves

Throughout this section, XX will be S2S_{2} and reduced. We begin this subsection with the following definition:

Definition 2.5.

Let XX be a scheme. A sheaf 𝔉\mathfrak{F} on XX is called divisorial sheaf if it is S2S_{2} and there is a closed subscheme ZXZ\subset X of codimension at least 2 such that 𝔉|XZ\mathfrak{F}|_{X\setminus Z} is locally free of rank 1.

Definition 2.6 ([HK04]*§ 3).

Let f:XSf\colon X\to S a flat morphism of schemes, and \mathcal{E} a coherent sheaf on XX. We say that \mathcal{E} is relatively SnS_{n} if it is flat over SS and its restriction to each fiber is Sn.S_{n}.

Now, let XX be a demi-normal scheme, and let Weil(X)\mathrm{Weil}^{*}(X) denote the subgroup of Weil(X)\mathrm{Weil}(X) generated by the prime divisors that are not contained in the conductor of XX. Then, there is an identification between Weil(X)/\mathrm{Weil}^{*}(X)/\sim and the group of divisorial sheaves, where \sim denotes linear equivalence. The identification is defined as follows. By the definition of Weil(X)\mathrm{Weil}^{*}(X) and the demi-normality of XX, for every element BWeil(X)B\in\mathrm{Weil}^{*}(X), there is a closed subset ZXZ\subset X of codimension at least 2 such that B|XZB|_{X\setminus Z} is a Cartier divisor. Then, the corresponding divisorial sheaf is defined as j𝒪XZ(B|XZ)j_{*}\mathcal{O}_{X\setminus Z}(B|_{X\setminus Z}), where we have j:XZXj\colon X\setminus Z\rightarrow X.

Consider a flat family f:XBf\colon X\to B with S2S_{2} fibers, and assume that there is an open subset i:UXi\colon U\subseteq X such that UbU_{b} is big for every bBb\in B. Then, for every locally free sheaf 𝒢\mathcal{G} on UU, we can consider i𝒢i_{*}\mathcal{G} on XX. From [HK04]*Corollary 3.7, this is a reflexive sheaf on XX (a priori, it is not S2S_{2} relatively to BB, see § 2.11).

If XX is demi-normal, its canonical sheaf is a divisorial sheaf, as XX is Gorenstein in codimension 1. By the above identification, we can then choose a canonical divisor KXK_{X} such that 𝒪X(KX)ωX\mathcal{O}_{X}(K_{X})\cong\omega_{X}. Observe that this construction can be carried out in families. Indeed, if f:XBf\colon X\to B is a flat morphism with demi-normal fibers of dimension nn, there is an open locus i:UXi\colon U\subseteq X that has codimension 2 along each fiber, on which ff is Gorenstein. Then we can define ωX/BiωU/B\omega_{X/B}\coloneqq i_{*}\omega_{U/B}. Observe that this agrees with the (n)(-n)-th cohomology of the relative dualizing complex. Indeed, the latter is S2S_{2} (see [LN18]*§ 5).

Let XX be demi-normal, and consider two divisorial sheaves L1L_{1} and L2L_{2}. Then, their reflexive tensor product is defined as L1^L2(L1L2)L_{1}\hat{\otimes}L_{2}\coloneqq(L_{1}\otimes L_{2})^{**} and it is a divisorial sheaf itself. If we have L1𝒪X(D1)L_{1}\cong\mathcal{O}_{X}(D_{1}) and L2𝒪X(D2)L_{2}\cong\mathcal{O}_{X}(D_{2}), we have L1^L2𝒪X(D1+D2)L_{1}\hat{\otimes}L_{2}\cong\mathcal{O}_{X}(D_{1}+D_{2}). The mm-fold reflexive power L[m]L^{[m]} is defined as the mm-fold self reflexive tensor product.

For more details, we refer to [kol_new]*§ 3.3, [Kol13]*5.6 and [HK04].

2.6. Boundedness

Let 𝔇\mathfrak{D} be a set of projective pairs. Then, we say that 𝔇\mathfrak{D} is log bounded (resp. log birationally bounded) if there exist a variety 𝒳\mathcal{X}, a reduced divisor \mathcal{B} on 𝒳\mathcal{X}, and a projective morphism π:𝒳T\pi\colon\mathcal{X}\rightarrow T, where TT is of finite type, such that \mathcal{B} does not contain any fiber of π\pi, and, for every (X,B)𝔇(X,B)\in\mathfrak{D}, there are a closed point tTt\in T and a morphism (resp. a birational map) ft:𝒳tXf_{t}\colon\mathcal{X}_{t}\rightarrow X inducing an isomorphism (X,Supp(B))(𝒳t,t)(X,\operatorname{Supp}(B))\cong(\mathcal{X}_{t},\mathcal{B}_{t}) (resp. such that Supp(t)\operatorname{Supp}(\mathcal{B}_{t}) contains the strict transform of Supp(B)\operatorname{Supp}(B) and all the ftf_{t} exceptional divisors).

A set of projective pairs 𝔇\mathfrak{D} is said to be strongly log bounded if there is a quasi-projective pair (𝒳,)(\mathcal{X},\mathcal{B}) and a projective morphism π:𝒳T\pi\colon\mathcal{X}\rightarrow T, where TT is of finite type, such that Supp()\operatorname{Supp}(\mathcal{B}) does not contain any fiber of π\pi, and for every (X,B)𝔇(X,B)\in\mathfrak{D}, there is a closed point tTt\in T and an isomorphism f:X𝒳tf\colon X\rightarrow\mathcal{X}_{t} such that fB=tf_{*}B=\mathcal{B}_{t}.

A set of projective pairs 𝔇\mathfrak{D} is effectively log bounded if it is strongly log bounded and we may choose a bounding pair π:(𝒳,)T\pi\colon(\mathcal{X},\mathcal{B})\rightarrow T such that, for every closed point tTt\in T, we have (𝒳t,t)𝔇(\mathcal{X}_{t},\mathcal{B}_{t})\in\mathfrak{D}.

2.7. Index of a set

Given a finite subset InI\subseteq\mathbb{Q}^{n}, we define the index of II to be the smallest positive rational number rr such that rInrI\subseteq\mathbb{Z}^{n}.

2.8. Stable pairs and \mathbb{Q}-stable pairs

Let (X,D)(X,D) denote a projective semi-log canonical pair, where DD has coefficients in \mathbb{Q}. We say that (X,D)(X,D) is a stable pair if KX+DK_{X}+D is ample.

Definition 2.7.

Consider a polynomial p(t)[t]p(t)\in\mathbb{Q}[t] and a set I(0,1]I\subseteq(0,1]. A \mathbb{Q}-pair with polynomial p(t)p(t) and coefficients in II is the datum of a semi-log canonical pair (X,Δ)(X,\Delta) and a \mathbb{Q}-Cartier \mathbb{Q}-divisor DD on XX satisfying the following properties:

  1. (1)

    (X,D+Δ)(X,D+\Delta) is semi-log canonical;

  2. (2)

    p(t)=(KX+tD+Δ)dim(X)p(t)=(K_{X}+tD+\Delta)^{\dim(X)};

  3. (3)

    there is a stable pair (Xc,Δc+Dc)(X^{c},\Delta^{c}+D^{c}) with a birational contraction π:XXc\pi\colon X\to X^{c} such that π(KXc+Δc+Dc)=KX+Δ+D\pi^{*}(K_{X^{c}}+\Delta^{c}+D^{c})=K_{X}+\Delta+D, πD=Dc\pi_{*}D=D^{c} and πΔ=Δc\pi_{*}\Delta=\Delta^{c}; and

  4. (4)

    the coefficients of DD and Δ\Delta are in II.

If moreover there is ϵ0>0\epsilon_{0}>0 such that for every 0<ϵϵ00<\epsilon\leq\epsilon_{0} the pair (X,(1ϵ)D+Δ)(X,(1-\epsilon)D+\Delta) is stable, the \mathbb{Q}-pair (X,Δ;D)(X,\Delta;D) is called a \mathbb{Q}-stable pair. Finally, we will call (Xc,Δc+Dc)(X^{c},\Delta^{c}+D^{c}) the canonical model of (X,Δ;D)(X,\Delta;D).

Remark 2.8.

If (X,Δ;D)(X,\Delta;D) is a pair that satisfies points (1), (2) and (4) of Definition 2.7 and such that there is ϵ0>0\epsilon_{0}>0 such that for every 0<ϵϵ00<\epsilon\leq\epsilon_{0} the pair (X,(1ϵ)D+Δ)(X,(1-\epsilon)D+\Delta) is stable, then one can prove that point (3) of Definition 2.7 is equivalent to KX+D+ΔK_{X}+D+\Delta being basepoint free.

For brevity, we denote the datum of a \mathbb{Q}-pair by (X,Δ;D)(X,\Delta;D), where the polynomial p(t)p(t) and the set of coefficients II are omitted in the notation. In case Δ=0\Delta=0, we then write (X;D)(X;D).

Remark 2.9.

For example, if (X,0;D)(X,0;D) is a semi-log canonical \mathbb{Q}-stable pair, and ν:XnX\nu\colon X^{n}\to X is the normalization of XX with conductor Δ\Delta, then (Xn,Δ;ν1D)(X^{n},\Delta;\nu_{*}^{-1}D) is a \mathbb{Q}-stable pair.

Notation 2.10.

If II is a finite set and rr is its index, given a \mathbb{Q}-pair (X,Δ;D)(X,\Delta;D) with polynomial p(t)p(t) and coefficients in II we denote by DscD^{sc} as the subscheme of XX defined by the reflexive sheaf of ideals 𝒪X(rD)\mathcal{O}_{X}(-rD).

Remark 2.11.

Observe that if (X,Δ;D)(X,\Delta;D) is a \mathbb{Q}-stable pair, the \mathbb{Q}-divisor KX+D+ΔK_{X}+D+\Delta is nef since it is limit of ample \mathbb{Q}-divisors. Furthermore, KX+D+ΔK_{X}+D+\Delta is big, as it is the sum of an ample \mathbb{Q}-divisor and an effective \mathbb{Q}-divisor.

Lemma 2.12.

Let XX be a projective variety, and let D1D_{1} and D2D_{2} be two nef \mathbb{Q}-Cartier \mathbb{Q}-divisors. Assume that for some t(0,1)t\in(0,1), the divisor tD1+(1t)D2tD_{1}+(1-t)D_{2} is ample. Then, cD1+(1c)D2cD_{1}+(1-c)D_{2} is ample for every c(0,1)c\in(0,1).

Proof.

Since ampleness is an open condition, we may assume that tt\in\mathbb{Q}. Then, using convex combinations, the claim follows from Kleiman’s criterion and the denseness of \mathbb{Q} in \mathbb{R}. ∎

Lemma 2.13.

Let (X,Δ;D)(X,\Delta;D) be a \mathbb{Q}-stable pair, and assume that f:XYf\colon X\rightarrow Y is its canonical model. Then, Ex(f)Supp(D)\mathrm{Ex}(f)\subset\operatorname{Supp}(D).

Proof.

By assumption, there exists ϵ>0\epsilon>0 such that (X,(1ϵ)D+Δ)(X,(1-\epsilon)D+\Delta) is a stable pair. Thus, if CC is an irreducible curve that is not contained in Supp(D)\operatorname{Supp}(D), we have that

(KX+D+Δ).C=(KX+(1ϵ)D+Δ).C+ϵD.C>0,(K_{X}+D+\Delta).C=(K_{X}+(1-\epsilon)D+\Delta).C+\epsilon D.C>0,

since the first summand is positive by ampleness and the second is non-negative as CC is not contained in Supp(D)\operatorname{Supp}(D). Thus, the curves contracted by ff are contained in Supp(D)\operatorname{Supp}(D) and the claim follows. ∎

The following lemma is the main technical tool that we will use in § 3 and § 5:

Lemma 2.14.

Let nn and MM be natural numbers. Then, there exists ϵ0>0\epsilon_{0}>0, only depending on nn and MM, such that the following holds. Let (Y,ΔY;DY)(Y,\Delta_{Y};D_{Y}) be a normal \mathbb{Q}-stable pair of dimension nn, and let π:YX\pi\colon Y\rightarrow X be its canonical model (X,DX+ΔX)(X,D_{X}+\Delta_{X}). Further assume that the Cartier index of KX+DX+ΔXK_{X}+D_{X}+\Delta_{X} is less than MM. Then, for every 0<ϵ<ϵ00<\epsilon<\epsilon_{0}, the pair (Y,(1ϵ)DY+ΔY)(Y,(1-\epsilon)D_{Y}+\Delta_{Y}) is stable.

Proof.

By definition of \mathbb{Q}-stable pair and of canonical model, for every π\pi-exceptional curve CC, we have (KY+DY+ΔY).C=0(K_{Y}+D_{Y}+\Delta_{Y}).C=0 and (KY+(1ϵ)DY+ΔY).C>0(K_{Y}+(1-\epsilon)D_{Y}+\Delta_{Y}).C>0 for 0<ϵ10<\epsilon\ll 1. For every such curve CC, the function t(KY+tDY+ΔY).Ct\mapsto(K_{Y}+tD_{Y}+\Delta_{Y}).C is linear and not identically 0, so it has at most one zero. In particular, for every ϵ>0\epsilon>0 we have

(1) (KY+(1ϵ)DY+ΔY).C>0.(K_{Y}+(1-\epsilon)D_{Y}+\Delta_{Y}).C>0.

Hence, if we choose ϵ=1\epsilon=1, we have

(2) (KY+ΔY).C>0(K_{Y}+\Delta_{Y}).C>0

for every π\pi-exceptional curve.

By the same argument, for every curve CC^{\prime} such that (KY+ΔY).C0(K_{Y}+\Delta_{Y}).C^{\prime}\geq 0, it follows that

(3) (KY+(1η)DY+ΔY).C>0(K_{Y}+(1-\eta)D_{Y}+\Delta_{Y}).C^{\prime}>0

for every η(0,1)\eta\in(0,1). Now, define

a12M,ϵ0a2n+a+1.a\coloneqq\frac{1}{2M},\quad\epsilon_{0}\coloneqq\displaystyle{\frac{a}{2n+a+1}}.

Since KX+DX+ΔXK_{X}+D_{X}+\Delta_{X} is ample with Cartier index bounded by MM, for every curve Γ\Gamma that is not π\pi-exceptional, by the projection formula, we have

(4) 12M=a<(KX+DX+ΔX).pΓ=(KY+DY+ΔY).Γ.\frac{1}{2M}=a<(K_{X}+D_{X}+\Delta_{X}).p_{*}\Gamma=(K_{Y}+D_{Y}+\Delta_{Y}).\Gamma.

From [Fuj11]*Theorem 1.4, for every (KY+ΔY)(K_{Y}+\Delta_{Y})-negative extremal ray RR, we may find a curve Γ\Gamma generating RR such that

(5) 2n(KY+ΔY).Γ<0.-2n\leq(K_{Y}+\Delta_{Y}).\Gamma<0.

By (2), any such curve is not π\pi-exceptional. Then, by (4) and (5), for any such Γ\Gamma, we have

(6) (KY+(1ϵ0)DY+ΔY).Γ=(1ϵ0)(KY+DY+ΔY).Γ+ϵ0(KY+ΔY).Γ>(1ϵ0)a2nϵ0>0.(K_{Y}+(1-\epsilon_{0})D_{Y}+\Delta_{Y}).\Gamma=(1-\epsilon_{0})(K_{Y}+D_{Y}+\Delta_{Y}).\Gamma+\epsilon_{0}(K_{Y}+\Delta_{Y}).\Gamma>(1-\epsilon_{0})a-2n\epsilon_{0}>0.

Now, consider the cone of curves NE¯(Y)\overline{NE}(Y) and its decomposition given by the cone theorem associated to the pair (Y,ΔY)(Y,\Delta_{Y}) [Fuj11]*Theorem 1.4. Then, by (3), we have that KY+(1ϵ0)DY+ΔYK_{Y}+(1-\epsilon_{0})D_{Y}+\Delta_{Y} is positive on NE¯(Y)KY+ΔY0\overline{NE}(Y)_{K_{Y}+\Delta_{Y}\geq 0}. Thus, every (KY+(1ϵ0)DY+ΔY)(K_{Y}+(1-\epsilon_{0})D_{Y}+\Delta_{Y})-negative extremal ray is also a (KY+ΔY)(K_{Y}+\Delta_{Y})-negative extremal ray. Then, by (6), we have that (KY+(1ϵ0)DY+ΔY)(K_{Y}+(1-\epsilon_{0})D_{Y}+\Delta_{Y}) is positive on RR. Thus, by the cone theorem [Fuj11]*Theorem 1.4, the log canonical pair (Y,(1ϵ0)DY+ΔY)(Y,(1-\epsilon_{0})D_{Y}+\Delta_{Y}) has no negative extremal rays; thus, KY+(1ϵ0)DY+ΔYK_{Y}+(1-\epsilon_{0})D_{Y}+\Delta_{Y} is nef.

Now, by the definition of \mathbb{Q}-stable pair, some convex combination of KY+(1ϵ0)DY+ΔYK_{Y}+(1-\epsilon_{0})D_{Y}+\Delta_{Y} and KY+DY+ΔYK_{Y}+D_{Y}+\Delta_{Y} is ample. Then, the claim follows by Lemma 2.12. ∎

Lemma 2.15.

Fix an integer nn\in\mathbb{N}, a volume v>0v\in\mathbb{Q}_{>0}, and a finite subset I(0,1]I\subseteq(0,1]\cap\mathbb{Q}. There is 0<ϵ00<\epsilon_{0}, only depending only on nn and vv, such that, for every \mathbb{Q}-stable pair (Y,ΔY;DY)(Y,\Delta_{Y};D_{Y}) of dimension nn, coefficients in II, and polynomial p(t)p(t) with p(1)=vp(1)=v, the pair (Y,(1ϵ)DY+ΔY)(Y,(1-\epsilon)D_{Y}+\Delta_{Y}) is stable for every 0<ϵ<ϵ00<\epsilon<\epsilon_{0}.

Proof.

Without loss of generality, we may assume from now on that 1I1\in I. Let YnYY^{n}\to Y be the normalization of YY. Let (Yn,ΔYn)(Y^{n},\Delta_{Y}^{n}) be the pair induced by (Y,ΔY)(Y,\Delta_{Y}), and let DYnD_{Y}^{n} denote the pull-back of DYD_{Y} to YnY^{n}. Here, ΔYn\Delta_{Y}^{n} consists of the sum of the divisorial part of the preimage of ΔY\Delta_{Y} and the conductor. Notice that YnY^{n} possibly has more than one connected component. Then, we have:

  1. (1)

    (Yn,DYn+ΔYn)(Y^{n},D_{Y}^{n}+\Delta_{Y}^{n}) has still coefficients in II;

  2. (2)

    KYn+DYn+ΔYnK_{Y^{n}}+D_{Y}^{n}+\Delta_{Y}^{n} is semi-ample and big; and

  3. (3)

    (Yn,ΔYn;DYn)(Y^{n},\Delta_{Y}^{n};D_{Y}^{n}) is a \mathbb{Q}-stable pair.

From [HMX14]*Theorem 1.3, there are finitely many possibilities to write vv as v=viv=\sum v_{i}, where vi>0v_{i}>0 is the volume of a log canonical pair of general type with coefficients in II. So from (1), there is a finite set 𝒮={v1,,vk}\mathscr{S}=\{v_{1},...,v_{k}\} such that the volume of each connected component of YnY^{n} is in 𝒮\mathscr{S}.

For a log canonical pair of general type (Z,D)(Z,D), the canonical model (Zc,Dc)(Z^{c},D^{c}) is such that vol(Z,D)=vol(Zc,Dc)\operatorname{vol}(Z,D)=\operatorname{vol}(Z^{c},D^{c}). In particular, it follows from (2) that the irreducible components of (Yn,DYn+ΔYn)(Y^{n},D_{Y}^{n}+\Delta_{Y}^{n}) admit a canonical model, and the volumes of these models are contained in 𝒮\mathcal{S} as well. From [HMX]*Theorem 1.1, there is an M>0M>0 such that every stable pair with coefficients in II, dimension nn, and volume in the finite set 𝒮\mathscr{S} has Cartier index less than MM. 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 (Y,DY)(Y,D_{Y}) be a stable pair. Then, there exists a \mathbb{Q}-stable pair (X;D)(X;D) having (Y,DY)(Y,D_{Y}) as canonical model.

Observe in particular that from Lemma 2.13, the morphism π:XY\pi\colon X\to Y induces an isomorphism at the generic point of the codimension one singular locus of XX and YY, as those are points not contained in DD, and π\pi is an isomorphism away from the exceptional locus. So XX is normal if and only if its codimension one singular locus is empty and if and only if the one of YY is empty.

Proof.

We consider the semi-canonical modification π:XY\pi\colon X\rightarrow Y of (Y,0)(Y,0), in the sense of [Fuj15]. We observe that, in order to consider a semi-canonical modification, KYK_{Y} does not need to be \mathbb{Q}-Cartier. Such modification exists by [Fuj15]*Theorem 1.1 and the fact that YY is demi-normal. By [Fuj15]*Definition 2.6, π:XY\pi\colon X\rightarrow Y has the following properties:

  • π\pi is an isomorphism around every generic point of the double locus of XX;

  • this procedure is compatible with taking the normalizations of XX and YY, see [Fuj15]*Lemma 3.7.(2). In particular, XX is normal if so is YY, and, in general, π\pi establishes a bijection between irreducible components of XX and YY; and

  • KXK_{X} is \mathbb{Q}-Cartier and π\pi-ample.

Now, set KX+D=π(KY+D)K_{X}+D=\pi^{*}(K_{Y}+D). Since KY+DK_{Y}+D is ample, for 0<ϵ10<\epsilon\ll 1, we have that ϵKX+(1ϵ)(KX+D)=KX+(1ϵ)D\epsilon K_{X}+(1-\epsilon)(K_{X}+D)=K_{X}+(1-\epsilon)D is ample. Since both KXK_{X} and KX+DK_{X}+D are \mathbb{Q}-Cartier, then so is DD. Lastly, as XX and (Y,DY)(Y,D_{Y}) are semi-log canonical, to conclude it suffices to show that D0D\geq 0.

Now, let (Xν,Dν+Δν)(X^{\nu},D^{\nu}+\Delta^{\nu}) and (Yν,DYν+ΔYν)(Y^{\nu},D_{Y}^{\nu}+\Delta^{\nu}_{Y}) denote the respective normalizations, where Δν\Delta^{\nu} and ΔYν\Delta_{Y}^{\nu} denote the double loci. Thus, it suffices to show that Dν0D^{\nu}\geq 0. Since this can be checked by considering one irreducible component of XνX^{\nu} at the time, by abusing notation, we may assume that XνX^{\nu} and YνY^{\nu} are irreducible. By construction, we have

KXν+(1ϵ)Dν+ΔνϵDνϵ(KXν+Δν)/Yν,K_{X^{\nu}}+(1-\epsilon)D^{\nu}+\Delta^{\nu}\sim_{\mathbb{Q}}-\epsilon D^{\nu}\sim_{\mathbb{Q}}\epsilon(K_{X^{\nu}}+\Delta^{\nu})/Y^{\nu},

and KXν+ΔνK_{X^{\nu}}+\Delta^{\nu} is relatively ample over YνY^{\nu}. Then, since we have DYν0D^{\nu}_{Y}\geq 0, by the negativity lemma [KM98]*Lemma 3.39, it follows that Dν0D^{\nu}\geq 0. 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 f:(X,D)Sf\colon(X,D)\rightarrow S over a reduced base SS is the datum of a morphism f:XSf\colon X\rightarrow S and an effective \mathbb{Q}-divisor DD on XX, such that the following conditions hold:

  • ff is flat with reduced fibers of pure dimension nn;

  • the fibers of Supp(D)S\operatorname{Supp}(D)\rightarrow S are either empty or of pure dimension n1n-1; and

  • ff is smooth at the generic points of XsSupp(D)X_{s}\cap\operatorname{Supp}(D) for every sSs\in S.

Furthermore, we say that a family of pairs is well defined if it also satisfies the following property:

  • mDmD is Cartier locally around the generic point of each irreducible component of XsSupp(D)X_{s}\cap\operatorname{Supp}(D) for every sSs\in S, where mm\in\mathbb{N} is a sufficiently divisible natural number clearing the denominators of DD.

This latter condition guarantees that mDmD is Cartier on a big open set UXU\subset X with the property that UXsU\cap X_{s} is a big open set of XsX_{s} for every sSs\in S. This guarantees that we have a well-defined notion of pull-back of DD under any possible base change SSS^{\prime}\rightarrow S, as we can pull back mD|UmD|_{U}, take its closure in X×SSX\times_{S}S^{\prime}, and then divide the coefficients by mm.

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 f:(X,D)Sf\colon(X,D)\rightarrow S over a reduced base SS is called locally stable if, for every base change g:(XC,DC)Cg\colon(X_{C},D_{C})\rightarrow C where CC is the spectrum of a DVR with closed point 0, (XC,DC+X0)(X_{C},D_{C}+X_{0}) is a semi-log canonical pair. Then, a family f:(X,D)Sf\colon(X,D)\rightarrow S is called stable if it is locally stable, ff is proper, and KX/S+DK_{X/S}+D is ff-ample.

2.10. Families of \mathbb{Q}-pairs

Let us fix a positive integer nn, a polynomial p(t)[t]p(t)\in\mathbb{Q}[t], and a finite set of coefficients I(0,1]I\subset(0,1]\cap\mathbb{Q}. Let rr be the index of II, 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 (X,D)(X,D) is a log canonical pair such that KX+(1ϵ)DK_{X}+(1-\epsilon)D is ample for 0<ϵ10<\epsilon\ll 1, then KX+DK_{X}+D 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 (X,D)(X,D) (it is the basepoint-free theorem).

Definition 2.19.

Let n,p,I\mathscr{F}_{n,p,I}^{\prime} be the category fibered in groupoids over Sch/\operatorname{Sch}/\mathbb{C} whose fibers over a scheme BB consists of:

  • a flat and proper morphism f:𝒳Bf\colon\mathcal{X}\to B of relative dimension nn;

  • a flat and proper morphism 𝒟B\mathcal{D}\to B of relative dimension n1n-1 and relatively S1S_{1};

  • a closed embedding i:𝒟𝒳i\colon\mathcal{D}\to\mathcal{X} over BB; and

  • for every point bBb\in B, the fiber (𝒳b;1r𝒟b)(\mathcal{X}_{b};\frac{1}{r}\mathcal{D}_{b}) is \mathbb{Q}-stable with polynomial p(t)p(t) and coefficients in II, where rr is the index of II, see § 2.7.

For every morphism BBB^{\prime}\to B, we denote by jB:𝒳×BB𝒳j_{B^{\prime}}\colon\mathcal{X}\times_{B}B^{\prime}\to\mathcal{X} the first projection, and by 𝒳\mathcal{X}^{\prime} the fiber product 𝒳×BB\mathcal{X}\times_{B}B^{\prime}. Similarly, for every point bBb\in B, we denote by 𝒟bSpec(k(b))×B𝒟\mathcal{D}_{b}\coloneqq\operatorname{Spec}(k(b))\times_{B}\mathcal{D}. If we denote by 𝒟\mathcal{I}_{\mathcal{D}} the ideal sheaf of 𝒟\mathcal{D}, we require that for every BBB^{\prime}\to B and every mm, mm^{\prime}, the natural map

(K) jB((ω𝒳/BmDm)[1])(ω𝒳/BmjB(Dm))[1]j_{B^{\prime}}^{*}((\omega_{\mathcal{X}/B}^{\otimes m^{\prime}}\otimes\mathcal{I}_{D}^{\otimes m})^{[1]})\to(\omega_{\mathcal{X}^{\prime}/B^{\prime}}^{\otimes m^{\prime}}\otimes j^{*}_{B^{\prime}}(\mathcal{I}_{D}^{\otimes m}))^{[1]}

is an isomorphism. We will denote by (𝒳;𝒟)B(\mathcal{X};\mathcal{D})\to B an object of n,p,I\mathscr{F}_{n,p,I}^{\prime} over BB, and we will call it a weak family of \mathbb{Q}-stable pairs.

Definition 2.20.

We will call weak \mathbb{Q}-stable morphism the datum of a flat and proper morphisms f:𝒳Bf\colon\mathcal{X}\to B and a closed embedding i:𝒟𝒳i\colon\mathcal{D}\to\mathcal{X} that satisfies the four bullet points of Definition 2.19. If there is no ambiguity we will still denote it with (𝒳;𝒟)B(\mathcal{X};\mathcal{D})\to B. Lastly, if K𝒳/B+1r𝒟K_{\mathcal{X}/B}+\frac{1}{r}\mathcal{D} is relatively semi-ample, we say it is a \mathbb{Q}-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 n,p,I\mathscr{F}_{n,p,I}^{\prime} 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 n,p,I\mathscr{F}_{n,p,I}^{\prime} 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 (𝒳;𝒟)B(\mathcal{X};\mathcal{D})\to B 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 n,p,I\mathscr{F}_{n,p,I} be the functor representing families (𝒳;𝒟)B(\mathcal{X};\mathcal{D})\to B in n,p,I\mathscr{F}_{n,p,I}^{\prime} such that there is a family of stable pairs (𝒳s,1r𝒟s)B(\mathcal{X}^{s},\frac{1}{r}\mathcal{D}^{s})\to B together with a morphism 𝒳𝒳s\mathcal{X}\to\mathcal{X}^{s} such that for every bBb\in B, the restriction (𝒳b,1r𝒟b)(𝒳bs;1r𝒟bs)(\mathcal{X}_{b},\frac{1}{r}\mathcal{D}_{b})\to(\mathcal{X}^{s}_{b};\frac{1}{r}\mathcal{D}^{s}_{b}) is the canonical model of (𝒳b,1r𝒟b)(\mathcal{X}_{b},\frac{1}{r}\mathcal{D}_{b}). We will denote these families as families of \mathbb{Q}-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 (X,Δ)S(X,\Delta)\to S there is a locally closed partial decomposition SSS^{\prime}\to S such that for any TST\to S, the pull-back (XT,ΔT)S(X_{T},\Delta_{T})\to S has a simultaneous, canonical, crepant, birational contraction (XT,ΔT)(XTs,ΔTs)T(X_{T},\Delta_{T})\to(X_{T}^{s},\Delta_{T}^{s})\to T with (XTs,ΔTs)T(X_{T}^{s},\Delta_{T}^{s})\to T stable iff TST\to S factors via TST\to S^{\prime}.

We will use this result for constructing n,p,I\mathscr{F}_{n,p,I}, 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. (1)

    in weak family of \mathbb{Q}-stable pairs, we require condition (K) but not condition (A);

  2. (2)

    in weak \mathbb{Q}-stable morphism, we require neither condition (K) nor condition (A);

  3. (3)

    in \mathbb{Q}-stable morphism, we require condition (A) but not condition (K); and

  4. (4)

    in family of \mathbb{Q}-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 2.23.

Observe that, with the notation of Definition 2.19 and Notation 2.10, (1r𝒟b)sc=𝒟b(\frac{1}{r}\mathcal{D}_{b})^{sc}=\mathcal{D}_{b}.

Remark 2.24.

The ideal sheaf 𝒟\mathcal{I}_{\mathcal{D}} is flat over BB. This follows from the fact that 𝒪𝒳\mathcal{O}_{\mathcal{X}} and 𝒪𝒟\mathcal{O}_{\mathcal{D}} are flat, by considering the associated long exact sequence of Tor\operatorname{Tor}.

Remark 2.25.

Since ff is flat and by condition (3) its fibers are S2S_{2}, ff is relatively S2S_{2}. Then, since 𝒟B\mathcal{D}\to B is flat and relatively S1S_{1}, it follows from [Kol13]*Corollary 2.61 that 𝒟\mathcal{I}_{\mathcal{D}} is relatively S2S_{2}.

Remark 2.26.

By condition (3), there exists an open subset V𝒳V\subseteq\mathcal{X} whose restriction to any fiber is a big open subset, such that 𝒟b\mathcal{D}_{b} is a Cartier divisor on VbV_{b} for every bBb\in B. Then, by flatness, we may apply [stacks-project]*Tag 062Y, and we conclude that 𝒟\mathcal{D} is Cartier along VV. Then, by Remark 2.24 and [HK04]*Proposition 3.5, 𝒪𝒳\mathcal{O}_{\mathcal{X}} and 𝒟\mathcal{I}_{\mathcal{D}} are reflexive. In particular, we have (𝒟)[1]=𝒟(\mathcal{I}_{\mathcal{D}})^{[1]}=\mathcal{I}_{\mathcal{D}}, and 𝒟\mathcal{I}_{\mathcal{D}} satisfies the conditions in [kol_new]*Definition 3.28.

Remark 2.27.

Observe that in Definition 2.19, the divisor 𝒟b\mathcal{D}_{b} in point (3) was defined via a fiber product. However, according to our conventions (see § 2.5) it should also correspond to a Weil divisor on 𝒳b\mathcal{X}_{b}. This is true since 𝒟B\mathcal{D}\to B is relatively S1S_{1}, so 𝒟b\mathcal{D}_{b} is S1S_{1} so its ideal sheaf in 𝒳b\mathcal{X}_{b} is S2S_{2} from [Kol13]*Corollary 2.61.

Remark 2.28.

The sheaves 𝒟[m]\mathcal{I}_{\mathcal{D}}^{[m]} are ideal sheaves of 𝒪𝒳\mathcal{O}_{\mathcal{X}}. Indeed, we can again consider an open subset V𝒳V\subseteq\mathcal{X} whose restriction to any fiber is a big open subset and such that 𝒟\mathcal{D} is a Cartier divisor on VV. Then, if we denote by i:V𝒳i\colon V\to\mathcal{X} the inclusion of VV, by [HK04]*Corollary 3.7 we have

𝒟[m]=i(𝒪V(m𝒟|V)).\mathcal{I}_{\mathcal{D}}^{[m]}=i_{*}(\mathcal{O}_{V}(-m\mathcal{D}|_{V})).

Then, the inclusion 𝒪V(m𝒟|V)𝒪V\mathcal{O}_{V}(-m\mathcal{D}|_{V})\hookrightarrow\mathcal{O}_{V} can be pushed forward via ii, to have an inclusion 𝒟[m]i(𝒪V)=𝒪𝒳\mathcal{I}_{\mathcal{D}}^{[m]}\hookrightarrow i_{*}(\mathcal{O}_{V})=\mathcal{O}_{\mathcal{X}}, where the last equality follows from [HK04]*Proposition 3.5 since 𝒳B\mathcal{X}\to B is S2S_{2}.

Notation 2.29.

If f:(𝒳;𝒟)Bf\colon(\mathcal{X};\mathcal{D})\to B is a weak family of \mathbb{Q}-stable pairs, we denote by m𝒟m\mathcal{D} the closed subscheme of 𝒳\mathcal{X} with ideal sheaf 𝒟[m]\mathcal{I}^{[m]}_{\mathcal{D}}.

Remark 2.30.

By [AH11]*Proposition 5.1.4, the sheaves 𝒟[m]\mathcal{I}_{\mathcal{D}}^{[m]} and ω𝒳/B[m]\omega_{\mathcal{X}/B}^{[m]} are flat over BB for every mm\in\mathbb{Z}.

Lemma 2.31.

The morphism m𝒟Bm\mathcal{D}\to B is flat with S1S_{1} fibers (i.e., the fibers have no embedded points).

Proof.

To check that m𝒟Bm\mathcal{D}\to B is flat it suffices to check that for every closed point Spec(k(b))B\operatorname{Spec}(k(b))\to B we have Tor1(k(b),𝒪m𝒟)=0\operatorname{Tor}^{1}(k(b),\mathcal{O}_{m\mathcal{D}})=0. We pull back the exact sequence

0𝒟[m]𝒪𝒳𝒪m𝒟00\to\mathcal{I}_{\mathcal{D}}^{[m]}\to\mathcal{O}_{\mathcal{X}}\to\mathcal{O}_{m\mathcal{D}}\to 0

via jb:𝒳b𝒳j_{b}\colon\mathcal{X}_{b}\to\mathcal{X}, and we obtain

0=Tor1(k(b),𝒪𝒳)Tor1(k(b),𝒪m𝒟)jb𝒟[m]𝒪𝒳b𝒪(m𝒟)b0.0=\operatorname{Tor}^{1}(k(b),\mathcal{O}_{\mathcal{X}})\to\operatorname{Tor}^{1}(k(b),\mathcal{O}_{m\mathcal{D}})\to j_{b}^{*}\mathcal{I}_{\mathcal{D}}^{[m]}\to\mathcal{O}_{\mathcal{X}_{b}}\to\mathcal{O}_{(m\mathcal{D})_{b}}\to 0.

However, jb𝒟[m](jb𝒟)[m]j_{b}^{*}\mathcal{I}_{\mathcal{D}}^{[m]}\cong(j_{b}^{*}\mathcal{I}_{\mathcal{D}})^{[m]}, so in particular it is a torsion free sheaf of rank 1. Then, the map jb𝒟[m]𝒪𝒳bj_{b}^{*}\mathcal{I}_{\mathcal{D}}^{[m]}\to\mathcal{O}_{\mathcal{X}_{b}} is injective, so Tor1(k(b),𝒪m𝒟)=0\operatorname{Tor}^{1}(k(b),\mathcal{O}_{m\mathcal{D}})=0 as desired.

Finally, 𝒟p[m]\mathcal{I}_{\mathcal{D}_{p}}^{[m]} is S2S_{2} and 𝒪𝒳p\mathcal{O}_{\mathcal{X}_{p}} is S2S_{2} from the commutativity with base change (K) in Definition 2.19, so 𝒪(m𝒟)p\mathcal{O}_{(m\mathcal{D})_{p}} is S1S_{1} from [Kol13]*Corollary 2.61.∎

Remark 2.32.

There is an m>0m>0, which does not depend on BB, such that ω𝒳/B[m]\omega_{\mathcal{X}/B}^{[m]} and 𝒟[m]\mathcal{I}_{\mathcal{D}}^{[m]} are locally free on 𝒳\mathcal{X}. Indeed, we will prove in Theorem 3.2 that there is an mm such that for every fiber (X;D)(X;D) of ff, the sheaves ωX[m]\omega_{X}^{[m]} and 𝒪X(mD)\mathcal{O}_{X}(-mD) are Cartier. Since our family is bounded (see Theorem 3.2), we can choose such an mm that does not depend on the basis BB. Then, by condition (K) in Definition 2.19 and Remark 2.30, we may apply [stacks-project]*Tag 00MH, which implies that ω𝒳/B[m]\omega_{\mathcal{X}/B}^{[m]} and 𝒟[m]\mathcal{I}_{\mathcal{D}}^{[m]} are locally free since they restrict to locally free sheaves along each fiber.

Now, we specify the morphisms in the fibered category n,p,I\mathscr{F}_{n,p,I} over a morphism f:TB.f\colon T\to B. Let α=((𝒴,𝒟𝒴)T)\alpha=((\mathcal{Y},\mathcal{D}_{\mathcal{Y}})\to T) be an element of n,p,I(T)\mathscr{F}_{n,p,I}(T), and let β=((𝒳;𝒟𝒳)B)\beta=((\mathcal{X};\mathcal{D}_{\mathcal{X}})\to B) be an element of n,p,I(B)\mathscr{F}_{n,p,I}(B). An arrow αβ\alpha\to\beta is the datum of two morphisms (g,h)(g,h) that fit in a diagram like the one below, where all the squares are fibered diagrams:

𝒟𝒴\textstyle{\mathcal{D}_{\mathcal{Y}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}g\scriptstyle{g}𝒟\textstyle{\mathcal{D}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒴\textstyle{\mathcal{Y}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}h\scriptstyle{h}𝒳\textstyle{\mathcal{X}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}T\textstyle{T\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f\scriptstyle{f}B.\textstyle{B.}
Observation 2.33.

The only morphisms over the identity Id:BB\operatorname{Id}\colon B\to B are isomorphisms. Thus, n,p,I\mathscr{F}_{n,p,I} 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 R=Spec(k[t](t))R=\operatorname{Spec}(k[t]_{(t)}), let η\eta (resp. pp) be the generic (resp. closed) point of Spec(R)\operatorname{Spec}(R), and let 𝒳=1×1×Spec(R)\mathcal{X}=\mathbb{P}^{1}\times\mathbb{P}^{1}\times\operatorname{Spec}(R). Consider CC a smooth member of 𝒪1×1(1,2)\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,2), and let 𝒴\mathcal{Y} be the blow-up of CC in the special fiber of 𝒳\mathcal{X}. Then, if we compose the blow-down 𝒴𝒳\mathcal{Y}\to\mathcal{X} with the projection 𝒳Spec(R)\mathcal{X}\to\operatorname{Spec}(R), we get a family of surfaces 𝒴Spec(R)\mathcal{Y}\to\operatorname{Spec}(R) where the generic fiber is a copy of 1×1\mathbb{P}^{1}\times\mathbb{P}^{1}, while the special fiber is a surface with two irreducible components. One irreducible component of the special fiber is isomorphic to 1×1\mathbb{P}^{1}\times\mathbb{P}^{1} (the proper transform of the special fiber of 𝒳Spec(R)\mathcal{X}\to\operatorname{Spec}(R)), and the other one is the exceptional divisor FF. The surface FF is the projectivization of the normal bundle of CC. Since C1C\cong\mathbb{P}^{1} and the normal bundle of CC in 𝒳\mathcal{X} is isomorphic to 𝒪1(0)𝒪1(4)\mathcal{O}_{\mathbb{P}^{1}}(0)\oplus\mathcal{O}_{\mathbb{P}^{1}}(4), we have that FF is isomorphic to the Hirzebruch surface 𝔽4\mathbb{F}_{4}. We denote by Δ𝔽4\Delta\subseteq\mathbb{F}_{4} the preimage of the double locus of the central fiber on 𝔽4\mathbb{F}_{4}.

We consider a divisor on 1×1\mathbb{P}^{1}\times\mathbb{P}^{1} consisting of five irreducible components, three general members of 𝒪1×1(1,2)\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(1,2) (which we denote by C1,C2,C3C_{1},C_{2},C_{3}), CC, and a smooth member GG of 𝒪1×1(2,0)\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(2,0). We consider a deformation of C1+C2+C3+C+GC_{1}+C_{2}+C_{3}+C+G in 𝒳\mathcal{X} given by the trivial deformation of CC and GG, and we deform CiC_{i} to CC for every ii. We denote by 𝒟𝒳\mathcal{D}_{\mathcal{X}} the total space of this deformation, and by 𝒟\mathcal{D} its proper transform in 𝒴\mathcal{Y}. More explicitly, if CC is the zero locus of a global section hH0(𝒪1×1(C))h\in H^{0}(\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(C)), GG is the zero locus of a global section gH0(𝒪1×1(G))g\in H^{0}(\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(G)), and φ1,φ2,φ3\varphi_{1},\varphi_{2},\varphi_{3} are generic sections of H0(𝒪1×1(C))H^{0}(\mathcal{O}_{\mathbb{P}^{1}\times\mathbb{P}^{1}}(C)), the deformation we consider is 𝒟𝒳=V(h(tφ1h)(tφ2h)(tφ3h)g)\mathcal{D}_{\mathcal{X}}=V(h(t\varphi_{1}-h)(t\varphi_{2}-h)(t\varphi_{3}-h)g).

[Uncaptioned image]

Figure 1: the family (𝒴,c𝒟)Spec(R).(\mathcal{Y},c\mathcal{D})\to\operatorname{Spec}(R).

Now, we construct the canonical model of (𝒴,c𝒟)Spec(R)(\mathcal{Y},c\mathcal{D})\to\operatorname{Spec}(R), with the coefficient cc in a neighbourhood of 12\frac{1}{2}. First, we introduce some notation. We denote K𝒴/Spec(R)+c𝒟K_{\mathcal{Y}/\operatorname{Spec}(R)}+c\mathcal{D} by (c)\mathcal{L}(c), the irreducible component of the central fiber isomorphic to 1×1\mathbb{P}^{1}\times\mathbb{P}^{1} by 𝒫\mathscr{P}, and the two rulings of 𝒫\mathscr{P} by f1f_{1} and f2f_{2}, with the convention that Cf1+2f2C\equiv f_{1}+2f_{2}. Since the generic fiber (𝒴,c𝒟)η(\mathcal{Y},c\mathcal{D})_{\eta} is stable for every cc with |c12|1|c-\frac{1}{2}|\ll 1, we just need to control the intersection pairings on the special fiber.

Let c=12+ϵc=\frac{1}{2}+\epsilon, where |ϵ|1|\epsilon|\ll 1.

[Uncaptioned image]

Figure 2: special fiber of the family (𝒴,c𝒟)Spec(R)(\mathcal{Y},c\mathcal{D})\to\operatorname{Spec}(R).

One can check that:

  1. (1)

    when ϵ>0\epsilon>0, the divisor (12+ϵ)\mathcal{L}(\frac{1}{2}+\epsilon) is nef. It is ample on FF, positive on f2f_{2} and 0 on f1f_{1};

  2. (2)

    when ϵ=0\epsilon=0, the divisor (12)\mathcal{L}(\frac{1}{2}) is nef. It is 0 on 𝒫\mathscr{P} and on Δ\Delta; and

  3. (3)

    when ϵ<0\epsilon<0, the divisor (12+ϵ)\mathcal{L}(\frac{1}{2}+\epsilon) is not nef. On 𝒫\mathscr{P}, it is negative on f2f_{2} and 0 on f1f_{1}, while on FF it is negative on Δ\Delta.

Therefore, we can explicitly describe the special fiber of the canonical model of (𝒴,c𝒟)(\mathcal{Y},c\mathcal{D}) over Spec(R)\operatorname{Spec}(R). We denote by (𝒵+,c𝒟+)(\mathcal{Z}^{+},c\mathcal{D}^{+}) (resp. (𝒵0,c𝒟0)(\mathcal{Z}^{0},c\mathcal{D}^{0}), (𝒵,c𝒟)(\mathcal{Z}^{-},c\mathcal{D}^{-})) the canonical model of (𝒴,c𝒟)Spec(R)(\mathcal{Y},c\mathcal{D})\to\operatorname{Spec}(R) when 0<ϵ10<\epsilon\ll 1 (resp. ϵ=0\epsilon=0, 1ϵ<0-1\ll\epsilon<0):

  1. (1)

    when ϵ>0\epsilon>0, to construct the canonical model we contract the ruling f1f_{1}. Since via this contraction the map C1×11C\subseteq\mathbb{P}^{1}\times\mathbb{P}^{1}\to\mathbb{P}^{1} is a 2:12\colon 1 ramified cover of 1\mathbb{P}^{1}, the special fiber is the push-out of the following diagram:

    1Δ\textstyle{\mathbb{P}^{1}\cong\Delta\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}2:1\scriptstyle{2\colon 1}F\textstyle{F\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π\scriptstyle{\pi}1\textstyle{\mathbb{P}^{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒵p+\textstyle{\mathcal{Z}^{+}_{p}}
  2. (2)

    when ϵ=0\epsilon=0, to construct the canonical model, we contract 𝒫\mathscr{P} and Δ\Delta. The special fiber is isomorphic to 𝔽4\mathbb{F}_{4} with the section Δ\Delta contracted, which is the projectivization of the cone over a rational quartic curve; and

  3. (3)

    When ϵ<0\epsilon<0, we perform a divisorial contraction to make the divisor nef. We contract the ruling f2f_{2}, and the special fiber is 𝔽4\mathbb{F}_{4}.

In particular, there are morphisms π+:𝔽4𝒵p+\pi^{+}\colon\mathbb{F}_{4}\to\mathcal{Z}^{+}_{p}, π0:𝔽4𝒵p0\pi^{0}\colon\mathbb{F}_{4}\to\mathcal{Z}^{0}_{p} and π:𝔽4𝒵p\pi^{-}\colon\mathbb{F}_{4}\to\mathcal{Z}^{-}_{p} from 𝔽4\mathbb{F}_{4} to the special fibers of 𝒵+\mathcal{Z}^{+}, 𝒵0\mathcal{Z}^{0} and 𝒵\mathcal{Z}^{-}. The divisors 𝒟p+\mathcal{D}^{+}_{p}, 𝒟p0\mathcal{D}^{0}_{p} and 𝒟p\mathcal{D}^{-}_{p} can be described as follows. First, recall that 𝔽4\mathbb{F}_{4} is the projectivization of the normal bundle of 1\mathbb{P}^{1} inside 𝒳\mathcal{X}, so it has a relative 𝒪𝔽4(1)\mathcal{O}_{\mathbb{F}_{4}}(1). We denote a generic section of 𝒪𝔽4(1)\mathcal{O}_{\mathbb{F}_{4}}(1) by hh. Then, the following holds:

  1. (1)

    when ϵ>0\epsilon>0, the divisor 𝒟p+\mathcal{D}^{+}_{p} is the image via π+\pi^{+} of four general fibers of FF, together with a divisor linearly equivalent to 4h4h. All the components have coefficient 12+ϵ\frac{1}{2}+\epsilon;

  2. (2)

    when ϵ=0\epsilon=0, the divisor 𝒟p0\mathcal{D}^{0}_{p} is the image via π0\pi^{0} of four general fibers with a divisor linearly equivalent to 4h4h. All the components have coefficient 12\frac{1}{2}; and

  3. (3)

    when ϵ<0\epsilon<0, the divisor 𝒟p\mathcal{D}^{-}_{p} consists of four general fibers with a divisor linearly equivalent to 4h4h, four generic fibers, and Δ\Delta. All the components have coefficient 12+ϵ\frac{1}{2}+\epsilon, with the exception of Δ\Delta, which has coefficient 1+2ϵ1+2\epsilon.

Recall now that the flat limit of 𝒟η𝒵η0\mathcal{D}_{\eta}\subseteq\mathcal{Z}^{0}_{\eta} in 𝒵p0\mathcal{Z}^{0}_{p} is not S1S_{1}, since it has an embedded point (see [KP17]*§ 1.2). However, (𝒵;𝒟)Spec(R)(\mathcal{Z}^{-};\mathcal{D}^{-})\to\operatorname{Spec}(R) is a \mathbb{Q}-stable morphism with coefficients I={12,1}I=\{\frac{1}{2},1\} (see Proposition 5.1), so in particular the flat limit of 𝒟η\mathcal{D}_{\eta} in 𝒵0\mathcal{Z}_{0}^{-} does not have an embedded point on the special fiber.

2.12. \mathbb{Q}-stable morphisms with constant part

For proving that n,p,I\mathscr{F}_{n,p,I} is bounded, it will be useful to introduce the following definition.

Definition 2.34.

Assume SS is reduced. A locally weak \mathbb{Q}-stable morphism with constant part f:(𝒳,Ω;𝒟)Sf\colon(\mathcal{X},\Omega;\mathcal{D})\rightarrow S over SS and with coefficients in II is the datum of a proper morphism f:𝒳Sf\colon\mathcal{X}\rightarrow S, a closed subscheme 𝒟\mathcal{D} on 𝒳\mathcal{X}, and an effective \mathbb{Q}-divisor Ω\Omega on XX, such that the following conditions hold:

  • f:(X,Ω)Sf\colon(X,\Omega)\rightarrow S is a proper, locally stable family of pairs, where Ω\Omega is a family of Mumford divisors [kol19s];

  • 𝒟S\mathcal{D}\rightarrow S is flat and relatively S1S_{1} (namely, with no embedded points) with fibers of pure dimension n1n-1; and

  • for every sSs\in S, there is a \mathbb{Q}-pair (X,Δ;D)(X,\Delta;D) with coefficients in II (i.e., both Δ\Delta and DD have coefficients in II) such that X=𝒳sX=\mathcal{X}_{s}, Δ=Ωs\Delta=\Omega_{s}, and Dsc=𝒟sD^{sc}=\mathcal{D}_{s}.

Furthermore, if we have a polynomial p(t)[t]p(t)\in\mathbb{Q}[t] and for every sSs\in S, (𝒳s,Ωs;1r𝒟s)(\mathcal{X}_{s},\Omega_{s};\frac{1}{r}\mathcal{D}_{s}) has polynomial p(t)p(t), we say that it is a locally weak \mathbb{Q}-stable morphism with constant part with polynomial p(t)p(t). Lastly, if K𝒳/S+1r𝒟+ΩK_{\mathcal{X}/S}+\frac{1}{r}\mathcal{D}+\Omega is relatively semi-ample, we drop the “weak” from the notation.

We remark that we call Ω\Omega the constant part since, contrary to 𝒟\mathcal{D}, the divisor Ω\Omega might not be \mathbb{Q}-Cartier on XX.

Notation 2.35.

We say that a locally (weak) \mathbb{Q}-stable morphism with constant part (with coefficients in II and polynomial p(t)p(t)) is a (weak) \mathbb{Q}-stable morphism with constant part (with coefficients in II and polynomial p(t)p(t)) if, for every sSs\in S, the fiber over ss is a \mathbb{Q}-stable pair.

Proposition 2.36.

Let us fix a set of coefficients I(0,1]I\subset(0,1]\cap\mathbb{Q}, a polynomial pp and an integer nn. Let rr denote the index of II. Let f:(𝒳,Ω;𝒟)Sf\colon(\mathcal{X},\Omega;\mathcal{D})\rightarrow S be a weak \mathbb{Q}-stable morphism with constant part Ω\Omega over a reduced base SS. Then f:(𝒳,1r𝒟+Ω)Sf\colon(\mathcal{X},\frac{1}{r}\mathcal{D}+\Omega)\rightarrow S is a well defined family of pairs.

Proof.

By definition 𝒟S\mathcal{D}\rightarrow S is flat with fibers of pure dimension n1n-1. Thus, it follows that the fibers of Supp(𝒟+Ω)S\mathrm{Supp}(\mathcal{D}+\Omega)\rightarrow S are either empty or of pure dimension n1n-1. Thus, to show that f:(𝒳,1r𝒟+Ω)Sf\colon(\mathcal{X},\frac{1}{r}\mathcal{D}+\Omega)\rightarrow S is a family of pairs, we are left with showing that ff is smooth at the generic points of XsSupp(𝒟+Ω)X_{s}\cap\operatorname{Supp}(\mathcal{D}+\Omega) for every sSs\in S.

By assumption, this is the case for all the generic points of XsSupp(𝒟+Ω)X_{s}\cap\operatorname{Supp}(\mathcal{D}+\Omega) arising from Ω\Omega. Thus, we may focus on the contribution of 𝒟\mathcal{D}. But then, since each fiber (𝒳s,Ωs;𝒟s)(\mathcal{X}_{s},\Omega_{s};\mathcal{D}_{s}) is a \mathbb{Q}-pair, it follows that the generic points of 𝒟s\mathcal{D}_{s} are contained in the smooth locus of XsX_{s} by the semi-log canonical condition. Thus, f:(𝒳,1r𝒟+Ω)Sf\colon(\mathcal{X},\frac{1}{r}\mathcal{D}+\Omega)\rightarrow S 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 𝒟\mathcal{D}, as Ω\Omega satisfied the needed conditions by definition. As argued in [kol_new]*Definition 3.35, there is a big open subset U𝒳U\subset\mathcal{X} such that every point of UU is either smooth or nodal, and U𝒳sU\cap\mathcal{X}_{s} has codimension at least 2 in 𝒳s\mathcal{X}_{s} for every sSs\in S. For every sSs\in S, 𝒟s\mathcal{D}_{s} does not contain any irreducible component of the double locus. Thus, the intersection between 𝒟\mathcal{D} and UU has codimension at least 2 in every fiber. Let VV denote the open set obtained by removing this intersection from UU. Then, as the scheme theoretic restriction 𝒟s\mathcal{D}_{s} is an integral Weil divisor, it is a Carter divisor along VsV_{s}. Then, by [stacks-project]*Tag 062Y, 𝒟\mathcal{D} is a Cartier divisor along VV. Thus, the claim follows. ∎

Proposition 2.37.

Fix an integer nn\in\mathbb{N}, a polynomial p(t)[t]p(t)\in\mathbb{Q}[t], and a finite subset I(0,1]I\subseteq(0,1]\cap\mathbb{Q}. Let ϵ0>0\epsilon_{0}>0 be as in Lemma 2.15, and let f:(𝒳,Ω;𝒟)Sf\colon(\mathcal{X},\Omega;\mathcal{D})\rightarrow S be a weak \mathbb{Q}-stable morphism with constant part, coefficients in II, and polynomial p(t)p(t), over a reduced base SS. Then f:(𝒳,1ϵr𝒟+Ω)Sf\colon(\mathcal{X},\frac{1-\epsilon}{r}\mathcal{D}+\Omega)\rightarrow S is a stable family of pairs for every 0<ϵ<ϵ00<\epsilon<\epsilon_{0} rational. In particular, 𝒟\mathcal{D} is \mathbb{Q}-Cartier.

Proof.

Fix ϵ\epsilon as in the statement. Then, by Proposition 2.36, f:(𝒳,1ϵr𝒟+Ω)Sf\colon(\mathcal{X},\frac{1-\epsilon}{r}\mathcal{D}+\Omega)\rightarrow S is a well defined family of pairs. By assumption, the self-intersection (K𝒳s+1ϵr𝒟s+Ωs)n(K_{\mathcal{X}_{s}}+\frac{1-\epsilon}{r}\mathcal{D}_{s}+\Omega_{s})^{n} is independent of sSs\in S, as it is p(1ϵ)p(1-\epsilon). Since f:(𝒳,1ϵr𝒟+Ω)Sf\colon(\mathcal{X},\frac{1-\epsilon}{r}\mathcal{D}+\Omega)\rightarrow S is a well defined family of pairs, we may find a big open subset U𝒳U\subset\mathcal{X} such that UsU_{s} has codimension at least 2 for every sSs\in S and K𝒳/S+1ϵr𝒟+ΩK_{\mathcal{X}/S}+\frac{1-\epsilon}{r}\mathcal{D}+\Omega is \mathbb{Q}-Cartier along UU. Then, K𝒳/S+1ϵr𝒟+ΩK_{\mathcal{X}/S}+\frac{1-\epsilon}{r}\mathcal{D}+\Omega is \mathbb{Q}-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 ϵ(0,ϵ0)\epsilon\in(0,\epsilon_{0})\cap\mathbb{Q}. ∎

2.13. Existence of good minimal models

Let (X,Δ)(X,\Delta) be a log canonical pair, and let f:XSf\colon X\rightarrow S be a projective morphism over a normal variety SS such that KX+ΔK_{X}+\Delta is ff-pseudo-effective. Then, it is expected that XX admits a good minimal model over SS. That is, XX admits a birational contraction ϕ:(X,Δ)(X,Δ)\phi\colon(X,\Delta)\dashrightarrow(X^{\prime},\Delta^{\prime}) over SS to a log canonical pair (X,Δ)(X^{\prime},\Delta^{\prime}), such that Δ\Delta^{\prime} is the push-forward of Δ\Delta to XX^{\prime}, ϕ\phi is (KX+Δ)(K_{X}+\Delta)-negative, and KX+ΔK_{X^{\prime}}+\Delta^{\prime} is semi-ample over SS.

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 \mathbb{Q}-stable morphism (resp. weak family of \mathbb{Q}-stable pairs) is actually a \mathbb{Q}-stable morphism (resp. family of \mathbb{Q}-stable pairs).

Lemma 2.38.

Let (X,Δ)(X,\Delta) be a log canonical pair, and let f:XSf\colon X\rightarrow S be a projective morphism to a normal variety such that KX+ΔK_{X}+\Delta is ff-pseudo-effective. Assume that the general fiber of ff has a good minimal model. Then, (X,Δ)(X,\Delta) admits a relative good minimal model over SS. Furthermore, if KX+ΔK_{X}+\Delta is nef over SS, then it is semi-ample.

Proof.

Let π:XX\pi\colon X^{\prime}\rightarrow X be a log resolution of (X,Δ)(X,\Delta), and let π(KX+Δ)=KX+Δ+E+ΓF\pi^{*}(K_{X}+\Delta)=K_{X^{\prime}}+\Delta^{\prime}+E^{\prime}+\Gamma^{\prime}-F^{\prime}. Here Δ\Delta^{\prime} denotes the strict transform of Δ\Delta, the divisors EE^{\prime}, Γ\Gamma^{\prime}, and FF^{\prime} are effective, π\pi-exceptional, and share no common components. Furthermore, Γ\Gamma^{\prime} is reduced, while the coefficients of EE^{\prime} are in (0,1)(0,1). Let Ξ\Xi^{\prime} denote the reduced π\pi-exceptional divisor, and fix a rational number 0<ϵ10<\epsilon\ll 1. Then, (X,Δ+E+Γ+ϵ(ΞΓ))(X^{\prime},\Delta^{\prime}+E^{\prime}+\Gamma^{\prime}+\epsilon(\Xi^{\prime}-\Gamma^{\prime})) is dlt, and it has the same pluricanonical ring as (X,Δ)(X,\Delta). Furthermore, by the addition of ϵ(ΞΓ)\epsilon(\Xi^{\prime}-\Gamma^{\prime}), every π\pi-exceptional divisor that is not in Γ\Gamma is in the relative stable base locus of KX+Δ+E+Γ+ϵ(ΞΓ)K_{X^{\prime}}+\Delta^{\prime}+E^{\prime}+\Gamma^{\prime}+\epsilon(\Xi^{\prime}-\Gamma^{\prime}). Finally, by assumption and the choice of ϵ\epsilon, every log canonical center of (X,Δ+E+Γ+ϵ(ΞΓ))(X^{\prime},\Delta^{\prime}+E^{\prime}+\Gamma^{\prime}+\epsilon(\Xi^{\prime}-\Gamma^{\prime})) dominates SS.

By assumption, the general fiber has a good minimal model. Thus, by [HMX18]*Theorem 1.9.1, it follows that (X,Δ+E+Γ+ϵ(ΞΓ))(X^{\prime},\Delta^{\prime}+E^{\prime}+\Gamma^{\prime}+\epsilon(\Xi^{\prime}-\Gamma^{\prime})) has a relative good minimal model over a non-empty smooth affine open subset USU\subseteq S. Then, as by assumption there are no vertical log canonical centers, it follows from [HX13]*Theorem 1.1 that (X,Δ+E+Γ+ϵ(ΞΓ))(X^{\prime},\Delta^{\prime}+E^{\prime}+\Gamma^{\prime}+\epsilon(\Xi^{\prime}-\Gamma^{\prime})) has a relative good minimal model over SS. Since every π\pi-exceptional divisor that is not in Γ\Gamma^{\prime} 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 (X,Δ)(X,\Delta) over SS 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 KX+ΔK_{X}+\Delta is relatively nef. Then, (X,Δ)(X,\Delta) is a relative weak log canonical model for (X,Δ+E+Γ+ϵ(ΞΓ))(X^{\prime},\Delta^{\prime}+E^{\prime}+\Gamma^{\prime}+\epsilon(\Xi^{\prime}-\Gamma^{\prime})) in the sense of [HMX]. Then, we conclude by [HMX]*Lemma 2.9.1 that (X,Δ)(X,\Delta) 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 X~\tilde{X} over a compactification S¯\overline{S} of SS, and take a projective relative good minimal model. Then, by adding the pull-back of some divisor on S¯\overline{S}, we can regard the relative good minimal model as a projective minimal model. Thus, it follows that KX+ΔK_{X}+\Delta is relatively semi-ample over SS, 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 II, a polynomial pp and a dimension nn, the corresponding set of \mathbb{Q}-stable pairs is effectively log-bounded.

Proposition 3.1.

Fix an integer nn\in\mathbb{N}. Consider a locally stable family of relative dimension nn (𝒳,𝒟)B(\mathcal{X},\mathcal{D})\to B over a reduced scheme BB, with 𝒟\mathcal{D} being \mathbb{Q}-Cartier. Assume that there is an 0<ϵ0<10<\epsilon_{0}<1 such that (𝒳,(1ϵ0)𝒟)B(\mathcal{X},(1-\epsilon_{0})\mathcal{D})\to B is stable. Then, the set of points bBb\in B such that K𝒳b+𝒟bK_{\mathcal{X}_{b}}+\mathcal{D}_{b} is semi-ample is constructible.

Proof.

This follows immediately from Theorem 2.22. ∎

Theorem 3.2.

Fix an integer nn\in\mathbb{N}, a finite subset I(0,1]I\subseteq(0,1]\cap\mathbb{Q}, and a polynomial p(t)[t]p(t)\in\mathbb{Q}[t]. Then, the set of \mathbb{Q}-stable pairs (X,Δ;D)(X,\Delta;D) of dimension nn, polynomial p(t)p(t) and coefficients in II is effectively log bounded.

Proof.

We proceed in several steps.

Step 1: In this step, we show that the \mathbb{Q}-stable pairs of interest are log bounded.

From Lemma 2.15, there is an ϵ0>0\epsilon_{0}>0 such that, for every \mathbb{Q}-stable pair (X,Δ;D)(X,\Delta;D) as in the statement, (X,(1ϵ0)D+Δ)(X,(1-\epsilon_{0})D+\Delta) is a stable pair. Then, from [HMX], there is a bounding family (𝒳,)B(\mathcal{X},\mathcal{E})\to B of stable pairs of volume p(1ϵ0)p(1-\epsilon_{0}), coefficients in the finite set (1ϵ0)II{1}(1-\epsilon_{0})I\cup I\cup\{1\} and dimension nn.

Step 2: In this step, we show that the \mathbb{Q}-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 \mathcal{E}, we may find divisors 𝒟\mathcal{D} and Ω\Omega supported on \mathcal{E} such that (1ϵ0)𝒟(1-\epsilon_{0})\mathcal{D} restricts to (1ϵ0)D(1-\epsilon_{0})D fiberwise, and Ω\Omega restricts to Δ\Delta fiberwise. By [kol_new]*Lemma 4.44, up to replacing BB with a finite disjoint union of locally closed subsets, we can further assume that both (𝒳,(1ϵ)𝒟+Ω)B(\mathcal{X},(1-\epsilon)\mathcal{D}+\Omega)\to B and (𝒳,𝒟+Ω)B(\mathcal{X},\mathcal{D}+\Omega)\to B are locally stable. In particular, K𝒳/B+t𝒟+ΩK_{\mathcal{X}/B}+t\mathcal{D}+\Omega is \mathbb{Q}-Cartier for any tt. Furthermore, by flatness, up to disregarding some irreducible components of BB, we can assume that for every bB,b\in B, p(t)=(K𝒳/B+t𝒟+Ω)dim(X)p(t)=(K_{\mathcal{X}/B}+t\mathcal{D}+\Omega)^{\dim(X)}. Finally, up to stratifying BB, we may assume that each irreducible component of BB is smooth; in particular, KBK_{B} 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 t=1ϵ0t=1-\epsilon_{0}, K𝒳/B+(1ϵ0)𝒟+ΩK_{\mathcal{X}/B}+(1-\epsilon_{0})\mathcal{D}+\Omega is ample on the general fibers. Thus, up to removing some proper closed subset of BB, we may assume that (𝒳,(1ϵ0)𝒟+Ω)B(\mathcal{X},(1-\epsilon_{0})\mathcal{D}+\Omega)\to B is a stable family. Thus, to conclude the proof it suffices to use Proposition 3.1, which guarantees that the set {pB:(K𝒳/B+𝒟+Ω)|𝒳p is semi-ample}\{p\in B\colon(K_{\mathcal{X}/B}+\mathcal{D}+\Omega)|_{{\mathcal{X}}_{p}}\text{ is semi-ample}\} is constructible. ∎

4. The moduli functor

The goal of this section is to prove that n,p,I\mathscr{F}_{n,p,I} is an algebraic stack. We begin by the following proposition:

Proposition 4.1.

The fibered category n,p,I\mathscr{F}_{n,p,I} 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 ω𝒞3\omega_{\mathscr{C}}^{\otimes 3}, for us is L𝒪X(m(KX+(1ϵ)D))L\coloneqq\mathcal{O}_{X}(m(K_{X}+(1-\epsilon)D)), where ϵ\epsilon and mm are chosen such that LL is very ample with hi(X,L)=0h^{i}(X,L)=0 for i1i\geq 1 and such that H0(X,L)H0(Dsc,L|Dsc)H^{0}(X,L)\to H^{0}(D^{sc},L|_{D^{sc}}) is surjective. These mm and ϵ\epsilon can be chosen uniformly by Theorem 3.2.

The fact that isomorphisms are a sheaf in the étale topology of n,p,I\mathscr{F}_{n,p,I} follows from descent as in [Alp21]*Proposition 1.4.6.

For proving that n,p,I\mathscr{F}_{n,p,I} satisfies descent, we begin by the following observation. Consider an object f:(𝒳,𝒟)Bf\colon(\mathcal{X},\mathcal{D})\to B of n,p,I(B)\mathscr{F}_{n,p,I}(B), and pick mm and ϵ\epsilon as before. Then, up to replacing mm with some uniform multiple, from Remark 2.32 and from cohomology and base change, we can assume that 𝒢ω𝒳/B[m]𝒟[m(1ϵ)r]\mathcal{G}\coloneqq\omega_{\mathcal{X}/B}^{[m]}\otimes\mathcal{I}_{\mathcal{D}}^{[-\frac{m(1-\epsilon)}{r}]} is relatively very ample, and f𝒢f_{*}\mathcal{G} is a vector bundle over BB. Indeed, by the deformation invariance of χ(𝒳b,𝒢b)\chi(\mathcal{X}_{b},\mathcal{G}_{b}) and the vanishing of Hi(X,L)H^{i}(X,L) for i1i\geq 1, it follows that the sections of H0(𝒳b,𝒢b)H^{0}(\mathcal{X}_{b},\mathcal{G}_{b}) are deformation invariant, and hence f𝒢f_{*}\mathcal{G} is a vector bundle. Then, on every affine open trivializing f𝒢f_{*}\mathcal{G}, the morphism 𝒳(f𝒢)\mathcal{X}\rightarrow\mathbb{P}(f_{*}\mathcal{G}) can be identified with 𝒳B×(H0(𝒳b,𝒢b))\mathcal{X}\rightarrow B\times\mathbb{P}(H^{0}(\mathcal{X}_{b},\mathcal{G}_{b})), and the latter is an embedding as it is an embedding over BB fiber by fiber, by Nakayama’s lemma. This gives an embedding 𝒳(f𝒢)\mathcal{X}\hookrightarrow\mathbb{P}(f_{*}\mathcal{G}), and composing it with 𝒟𝒳\mathcal{D}\hookrightarrow\mathcal{X} an embedding 𝒟(f𝒢)\mathcal{D}\hookrightarrow\mathbb{P}(f_{*}\mathcal{G}). Now the proof is analogous to the one in [Alp21]*Proposition 1.4.6. ∎

Theorem 4.2.

Fix an integer nn\in\mathbb{N}, a finite subset I(0,1]I\subseteq(0,1]\cap\mathbb{Q}, and a polynomial p(t)[t]p(t)\in\mathbb{Q}[t]. Then n,p,I\mathscr{F}_{n,p,I} is an algebraic stack.

Proof.

Let rr denote the index of II. 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 \mathbb{Q}-pairs of interest.

By Lemma 2.15, we may find a rational number ϵ(0,1)\epsilon\in(0,1) such that KX+(1ϵ)DK_{X}+(1-\epsilon)D is ample for every \mathbb{Q}-stable pair (X;D)(X;D) with polynomial p(t)p(t) and coefficients in II. Without loss of generality, we may assume that ϵ=1k\epsilon=\frac{1}{k} for a suitable kk\in\mathbb{N}. By Theorem 3.2, we may consider a weak \mathbb{Q}-stable morphism π:(𝒳;𝒟)B\pi\colon(\mathcal{X};\mathcal{D})\to B with coefficients in II, polynomial p(t)p(t), and of relative dimension nn that is effectively log bounding for our moduli problem. By stratification of BB, we may assume that BB is smooth, π\pi and π|𝒟\pi|_{\mathcal{D}} are flat, and that π|𝒟\pi|_{\mathcal{D}} has S1S_{1} fibers. Furthermore, by Proposition 2.37, π\pi also induces a family of stable pairs. In particular, we can also regard 𝒟\mathcal{D} as a divisor, not only as a subscheme, and we will be free to take its multiples. Similarly, any natural multiple k𝒟k\mathcal{D} can be regarded as a subscheme, by considering the vanishing locus of 𝒪𝒳(k𝒟)\mathcal{O}_{\mathcal{X}}(-k\mathcal{D}). Then, there is m>0m>0 such that, for every \mathbb{Q}-stable pair (X;D)(X;D) in our moduli problem, we have:

  1. (1)

    mDmD and m(1ϵ)Dm(1-\epsilon)D are Cartier;

  2. (2)

    mKXmK_{X} is Cartier; and

  3. (3)

    m(KX+(1ϵ)D)m(K_{X}+(1-\epsilon)D) is a very ample line bundle with vanishing higher cohomologies that embeds XX into (H0(X,𝒪X(m(KX+(1ϵ)D)))\mathbb{P}(H^{0}(X,\mathcal{O}_{X}(m(K_{X}+(1-\epsilon)D))).

Furthermore, up to taking a multiple, we may assume that mr\frac{m}{r} and m(1ϵ)r\frac{m(1-\epsilon)}{r} are integers, and

  1. (1)

    mr𝒟\frac{m}{r}\mathcal{D} and m(1ϵ)r𝒟\frac{m(1-\epsilon)}{r}\mathcal{D} are Cartier;

  2. (2)

    mK𝒳/BmK_{\mathcal{X}/B} is Cartier; and

  3. (3)

    m(K𝒳/B+1ϵr𝒟)m(K_{\mathcal{X}/B}+\frac{1-\epsilon}{r}\mathcal{D}) is a relatively very ample line bundle.

By boundedness and upper semi-continuity of the space of global sections, h0(X,𝒪X(m(KX+(1ϵ)D))h^{0}(X,\mathcal{O}_{X}(m(K_{X}+(1-\epsilon)D)) attains finitely many values. Then, we have finitely many polynomials q1,,qlq_{1},\ldots,q_{l} such that the fibers of π\pi have Hilbert polynomial qiq_{i} for some ii, for the relatively ample line bundle 𝒪𝒳(m(K𝒳/B+1ϵr𝒟))\mathcal{O}_{\mathcal{X}}(m(K_{\mathcal{X}/B}+\frac{1-\epsilon}{r}\mathcal{D})). We consider a union of Hilbert schemes for the polynomials qiq_{i}, and we denote such a union with 0\mathscr{H}_{0}. Over 0\mathscr{H}_{0}, we have a universal family f0:𝒳0×N0f_{0}\colon\mathscr{X}\subseteq\mathscr{H}_{0}\times\mathbb{P}^{N}\to\mathscr{H}_{0}, where the fibers are closed subschemes of N\mathbb{P}^{N} with Hilbert polynomial qiq_{i} for some ii. Here, we observe that NN may actually attain finitely many distinct values, as we are assuming that each XX 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 NN.

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 PGLN+1\mathrm{PGL}_{N+1}. For this reason, we will shrink 0\mathscr{H}_{0} 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 PGLN+1\mathrm{PGL}_{N+1}.

Step 3: In this step, we cut the locus parametrizing demi-normal schemes.

Since being S2S_{2} 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 0\mathscr{H}_{0} we may assume that the fibers of f0f_{0} are S2S_{2} 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 (X;D)(X;D) be a \mathbb{Q}-pair in our moduli problem, and let IXI_{X} denote its ideal sheaf in N\mathbb{P}^{N}. Then, by assumption we have 𝒪N(1)|X𝒪X(m(KX+(1ϵ)D))\mathcal{O}_{\mathbb{P}^{N}}(1)|_{X}\cong\mathcal{O}_{X}(m(K_{X}+(1-\epsilon)D)) and the higher cohomologies of both sheaves vanish. Thus, if we consider the short exact sequence

0IX𝒪N(1)𝒪N(1)𝒪N(1)|X0,0\rightarrow I_{X}\otimes\mathcal{O}_{\mathbb{P}^{N}}(1)\rightarrow\mathcal{O}_{\mathbb{P}^{N}}(1)\rightarrow\mathcal{O}_{\mathbb{P}^{N}}(1)|_{X}\rightarrow 0,

it provides the following long exact sequence of cohomology groups

0H0(N,IX𝒪N(1))H0(N,𝒪N(1))H0(X,𝒪N(1)|X)H1(N,IX𝒪N(1))=0,0\rightarrow H^{0}(\mathbb{P}^{N},I_{X}\otimes\mathcal{O}_{\mathbb{P}^{N}}(1))\rightarrow H^{0}(\mathbb{P}^{N},\mathcal{O}_{\mathbb{P}^{N}}(1))\rightarrow H^{0}(X,\mathcal{O}_{\mathbb{P}^{N}}(1)|_{X})\rightarrow H^{1}(\mathbb{P}^{N},I_{X}\otimes\mathcal{O}_{\mathbb{P}^{N}}(1))=0,

where the vanishing of H1(N,IX𝒪N(1))H^{1}(\mathbb{P}^{N},I_{X}\otimes\mathcal{O}_{\mathbb{P}^{N}}(1)) follows from the surjectivity of the map H0(N,𝒪N(1))H0(X,𝒪N(1)|X)H^{0}(\mathbb{P}^{N},\mathcal{O}_{\mathbb{P}^{N}}(1))\rightarrow H^{0}(X,\mathcal{O}_{\mathbb{P}^{N}}(1)|_{X}) and the vanishing of the higher order cohomologies of 𝒪N(1)|X\mathcal{O}_{\mathbb{P}^{N}}(1)|_{X}. By definition of Hilbert scheme, 𝒳0\mathscr{X}\rightarrow\mathscr{H}_{0} is flat, and the subschemes parametrized correspond to a flat quotient sheaf of 𝒪𝒳\mathcal{O}_{\mathscr{X}}. 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 0\mathscr{H}_{0} to the open locus parametrizing varieties YY with H0(N,IY𝒪N(1))=0H^{0}(\mathbb{P}^{N},I_{Y}\otimes\mathcal{O}_{\mathbb{P}^{N}}(1))=0. This guarantees that, for every such variety YY, any automorphism of YY preserving 𝒪N(1)|Y\mathcal{O}_{\mathbb{P}^{N}}(1)|_{Y} is induced by an automorphism of N\mathbb{P}^{N}.

Step 5: In this step, we introduce a relative Hilbert scheme, in order to parametrize the boundaries of the \mathbb{Q}-pairs of interest.

Proceeding as in Step 1, for every \mathbb{Q}-pair (X;D)(X;D) in our moduli problem, we may consider the Hilbert polynomial of rDrD with respect to 𝒪X(m(KX+(1ϵ)D))\mathcal{O}_{X}(m(K_{X}+(1-\epsilon)D)). Here, recall that 𝒟\mathcal{D} corresponds to rDrD fiberwise, hence the choice of Hilbert polynomial for rDrD rather than for DD. By effective log boundedness and generic flatness of 𝒟\mathcal{D}, 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, ff is projective over 0\mathscr{H}_{0}. In particular, if we pull back the ample line bundle 𝒪0×N(1)\mathcal{O}_{\mathscr{H}_{0}\times\mathbb{P}^{N}}(1) to get a relatively very ample line bundle 𝒢\mathscr{G} on 𝒳\mathscr{X}, we can take the relative Hilbert scheme for the morphism ff, the line bundle 𝒢\mathscr{G}, and the polynomial determined by 𝒟\mathcal{D} (see [ACH11]*Ch. IX). This gives a scheme 10\mathscr{H}_{1}\to\mathscr{H}_{0}, together with an universal family 𝒟𝒳1𝒳×01\mathscr{D}\subseteq\mathscr{X}_{1}\coloneqq\mathscr{X}\times_{\mathscr{H}_{0}}\mathscr{H}_{1}. Then, as 𝒢\mathscr{G} corresponds to 𝒪X(m(KX+(1ϵ)D))\mathcal{O}_{X}(m(K_{X}+(1-\epsilon)D)) on the elements of our moduli problem, every \mathbb{Q}-pair of interest appears as a fiber of this family.

Step 6: In this step, we shrink 1\mathscr{H}_{1} to an open subset such that 𝒟Sing(𝒳11)\mathscr{D}\cap\mathrm{Sing}(\mathscr{X}_{1}\rightarrow\mathscr{H}_{1}) has codimension at least 2 along each fiber and such that the ideal sheaf 𝒟\mathscr{I}_{\mathscr{D}} of 𝒟\mathscr{D} is relatively S2S_{2}.

For every \mathbb{Q}-pair (X;D)(X;D), Supp(D)\operatorname{Supp}(D) does not contain any component of the double locus of XX. Thus, by upper semi-continuity of the dimension of the fibers of a morphism, we may shrink 1\mathscr{H}_{1} to an open subset such that 𝒟Sing(𝒳11)\mathscr{D}\cap\mathrm{Sing}(\mathscr{X}_{1}\rightarrow\mathscr{H}_{1}) has codimension at least 2 along each fiber.

Now, by Step 3, all the fibers of 𝒳11\mathscr{X}_{1}\rightarrow\mathscr{H}_{1} are demi-normal. Thus, we may find an open subset V𝒳1V\subset\mathscr{X}_{1} such that the following properties hold:

  1. (1)

    for every s1s\in\mathscr{H}_{1}, VsV_{s} is a big open set in 𝒳1,p\mathscr{X}_{1,p};

  2. (2)

    𝒟Sing(𝒳11)𝒳1V\mathscr{D}\cap\mathrm{Sing}(\mathscr{X}_{1}\rightarrow\mathscr{H}_{1})\subset\mathscr{X}_{1}\setminus V; and

  3. (3)

    the fibers of V1V\rightarrow\mathscr{H}_{1} have at worst nodal singularities.

Then, V1V\rightarrow\mathscr{H}_{1} is a Gorenstein morphism, so, by [stacks-project]*Tag 0C08, ω𝒳1/1\omega_{\mathscr{X}_{1}/\mathscr{H}_{1}} is an invertible sheaf along VV. Furthermore, by [stacks-project]*Tag 062Y, the ideal sheaf 𝒟\mathscr{I}_{\mathscr{D}} is locally free along VV. Finally, up to shrinking 1\mathscr{H}_{1}, we may assume that 𝒟\mathscr{I}_{\mathscr{D}} is relatively S2S_{2}.

Indeed, 𝒳1\mathscr{X}_{1} is flat over 1\mathscr{H}_{1}, as 𝒳\mathscr{X} is flat over 0\mathscr{H}_{0}, and 𝒟\mathscr{D} is flat over 1\mathscr{H}_{1} by definition of relative Hilbert scheme. Thus, 𝒟\mathscr{I}_{\mathscr{D}} is flat over 1\mathscr{H}_{1}, as it is the kernel of a surjection of flat sheaves. Then, we conclude, as being S2S_{2} is an open condition for flat and proper families [EGAIV]*Theorem 12.2.1. Notice that, by [HK04]*Proposition 3.5, we have 𝒟=𝒟[1]\mathscr{I}_{\mathscr{D}}=\mathscr{I}_{\mathscr{D}}^{[1]}.

Notice that in this step we shrank 1\mathscr{H}_{1} twice, and both times the process is invariant under the natural action of PGLN+1\mathrm{PGL}_{N+1}, as the locus is characterized by properties of the fibers.

Step 7: In this step, we cut 1\mathscr{H}_{1} to a locally closed subset 2\mathscr{H}_{2} to ensure that (ω𝒳1/1[μ]𝒟[μr])[1](\omega_{\mathscr{X}_{1}/\mathscr{H}_{1}}^{[\mu^{\prime}]}\otimes\mathscr{I}_{\mathscr{D}}^{[\frac{\mu}{r}]})^{[1]} is flat and S2S_{2} over 1\mathscr{H}_{1}, commutes with base change for every μ,μ\mu,\mu^{\prime}\in\mathbb{Z}, and that 𝒟[m(1ϵ)r]\mathscr{I}_{\mathscr{D}}^{[\frac{m(1-\epsilon)}{r}]}, 𝒟[mr]\mathscr{I}_{\mathscr{D}}^{[\frac{m}{r}]}, and ω𝒳1/sH1[m]\omega_{\mathscr{X}_{1}/sH_{1}}^{[m]} are invertible sheaves. Here, if μr\frac{\mu}{r} is not an integer, we set 𝒟[μr](𝒟μr)[1]\mathscr{I}_{\mathscr{D}}^{[\frac{\mu}{r}]}\coloneqq\left(\mathscr{I}_{\mathscr{D}}^{\lfloor\frac{\mu}{r}\rfloor}\right)^{[1]}.

The first claim follows immediately from [kol08] applied to the sheaves 𝒟[μr]\mathscr{I}_{\mathscr{D}}^{[\frac{\mu}{r}]} and ω𝒳1/1[μ]\omega_{\mathscr{X}_{1}/\mathscr{H}_{1}}^{[\mu^{\prime}]} for 0μr,μm0\leq\frac{\mu}{r},\mu^{\prime}\leq m. Thus, there is a stratification into functorial locally closed subsets of 1\mathscr{H}_{1}, which we denote by Ci1C_{i}\subseteq\mathscr{H}_{1}, where the above sheaves are flat over CiC_{i}, they are S2S_{2}, and they commute with base change. Then, by Remark 2.32, the sheaves 𝒟[m(1ϵ)r]\mathscr{I}_{\mathscr{D}}^{[\frac{m(1-\epsilon)}{r}]}, 𝒟[mr]\mathscr{I}_{\mathscr{D}}^{[\frac{m}{r}]}, and ω𝒳1/1[m]\omega_{\mathscr{X}_{1}/\mathscr{H}_{1}}^{[m]} are invertible over each component CiC_{i}.

Considering the union of the CiC_{i}’s produces another base 2\mathscr{H}_{2}, with a family f2:𝒳22f_{2}\colon\mathscr{X}_{2}\to\mathscr{H}_{2} and a closed subset 𝒟2𝒳2\mathscr{D}_{2}\subset\mathscr{X}_{2} as the one over 1\mathscr{H}_{1}, but such that 𝒟2[m(1ϵ)r]\mathscr{I}_{\mathscr{D}_{2}}^{[\frac{m(1-\epsilon)}{r}]}, 𝒟2[mr]\mathscr{I}_{\mathscr{D}_{2}}^{[\frac{m}{r}]}, and ω𝒳2/2[m]\omega_{\mathscr{X}_{2}/\mathscr{H}_{2}}^{[m]} are Cartier and the formation of (ω𝒳1/1[μ]𝒟[μr])[1](\omega_{\mathscr{X}_{1}/\mathscr{H}_{1}}^{[\mu^{\prime}]}\otimes\mathscr{I}_{\mathscr{D}}^{[\frac{\mu}{r}]})^{[1]} commutes with base change for every 0μr,μm0\leq\frac{\mu}{r},\mu^{\prime}\leq m. Then, the formation of (ω𝒳1/1[μ]𝒟[μr])[1](\omega_{\mathscr{X}_{1}/\mathscr{H}_{1}}^{[\mu^{\prime}]}\otimes\mathscr{I}_{\mathscr{D}}^{[\frac{\mu}{r}]})^{[1]} commutes with base change for every μ,μ\mu,\mu^{\prime}, since we can write μ=km+b\mu=km+b and μr=km+b\lfloor\frac{\mu}{r}\rfloor=k^{\prime}m+b^{\prime} for 0b,b<m0\leq b,b^{\prime}<m, so we can write (ω𝒳1/1[μ]𝒟[μr])[1](\omega_{\mathscr{X}_{1}/\mathscr{H}_{1}}^{[\mu^{\prime}]}\otimes\mathscr{I}_{\mathscr{D}}^{[\frac{\mu}{r}]})^{[1]} as a tensor product of a line bundle (namely (ω𝒳1/1[km]𝒟[kmr])[1](\omega_{\mathscr{X}_{1}/\mathscr{H}_{1}}^{[k^{\prime}m]}\otimes\mathscr{I}_{\mathscr{D}}^{[k\frac{m}{r}]})^{[1]}) and a relatively S2S_{2} sheaf which commutes with base change (namely (ω𝒳1/1[b]𝒟[b])[1](\omega_{\mathscr{X}_{1}/\mathscr{H}_{1}}^{[b^{\prime}]}\otimes\mathscr{I}_{\mathscr{D}}^{[b]})^{[1]}).

Step 8: In this step, we shrink 2\mathscr{H}_{2} to an open subset parametrizing semi-log canonical pairs.

By the reductions in the previous steps, for every k1k\geq 1, the ideal sheaf 𝒟2[k]\mathscr{I}_{\mathscr{D}_{2}}^{[k]} determines a family of generically Cartier divisors, in the sense of [kol_new]*Definition 4.24. Indeed, this is obtained by dualizing the inclusion 𝒟2[k]𝒪𝒳2\mathscr{I}_{\mathscr{D}_{2}}^{[k]}\rightarrow\mathcal{O}_{\mathscr{X}_{2}}, 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 𝒟2\mathscr{D}_{2} the family of divisors corresponding to 𝒟2\mathscr{I}_{\mathscr{D}_{2}}.

Now, recall that taking the reduced structure does not change the topology. Thus, in the remainder of this step, we may assume that 2\mathscr{H}_{2} is reduced. Then, (𝒳2,1r𝒟2)2(\mathscr{X}_{2},\frac{1}{r}\mathscr{D}_{2})\rightarrow\mathscr{H}_{2} is a well defined family of pairs, and by [kol_new]*Corollary 4.45, there is an open locus where the fibers of (𝒳2,1r𝒟2)2(\mathscr{X}_{2},\frac{1}{r}\mathscr{D}_{2})\rightarrow\mathscr{H}_{2} are semi-log canonical. Also, since we took an open subset of 2\mathscr{H}_{2}, this choice is not affected by considering the reduced structure of 2\mathscr{H}_{2}.

Step 9: In this step, we use Kollár’s theorem 2.22 to shrink 2\mathscr{H}_{2} to a locally closed and functorial subscheme parametrizing the \mathbb{Q}-stable pairs in our moduli problem.

Using Theorem 2.22, up to replacing 2\mathcal{H}_{2} with such a (functorial) locally closed subscheme of it, we can assume that ω𝒳2/2[m]𝒟2[mr]\omega_{\mathscr{X}_{2}/\mathscr{H}_{2}}^{[m]}\otimes\mathscr{I}_{\mathscr{D}_{2}}^{[-\frac{m}{r}]} is relatively semi-ample. Similarly, since being ample is an open condition, we can assume that ω𝒳2/2[m]𝒟2[m(1ϵ)r]\omega_{\mathscr{X}_{2}/\mathscr{H}_{2}}^{[m]}\otimes\mathscr{I}_{\mathscr{D}_{2}}^{[-\frac{m(1-\epsilon)}{r}]} is relatively ample. Now, recall that we have chosen ϵ=1k\epsilon=\frac{1}{k} for a chosen fixed kk\in\mathbb{N}. Notice that, the two former conditions guarantee that, for every jkj\geq k, we have that ω𝒳2/2[rjm]𝒟2[m(j1)]\omega_{\mathscr{X}_{2}/\mathscr{H}_{2}}^{[rjm]}\otimes\mathscr{I}_{\mathscr{D}_{2}}^{[-m(j-1)]} is relatively ample, as we have mj>m(j1)mj(11k)mj>m(j-1)\geq mj(1-\frac{1}{k}). Then, we may fix n+1n+1 such values j1,,jn+1j_{1},\ldots,j_{n+1} and disregard all components of 2\mathscr{H}_{2} but the ones where, over s2s\in\mathscr{H}_{2}, (ω𝒳2/2[rjim]𝒟2[m(ji1)])n(\omega_{\mathscr{X}_{2}/\mathscr{H}_{2}}^{[rj_{i}m]}\otimes\mathscr{I}_{\mathscr{D}_{2}}^{[-m(j_{i}-1)]})^{n} has prescribed value. By flatness, these self-intersections are locally constant, and thus this condition is open. For each ii, the self-intersection is prescribed by p(11ji)p(1-\frac{1}{j_{i}}), up to the rescaling factor given by rjimrj_{i}m. Since we are prescribing n+1n+1 values of a polynomial of degree nn, this guarantees that all the fibers correspond to \mathbb{Q}-stable pairs (X;D)(X;D) with (KX+tD)n=p(t)(K_{X}+tD)^{n}=p(t).

Step 10: In this step, we cut 2\mathscr{H}_{2} to a closed subset 3\mathscr{H}_{3} to ensure that the natural polarization 𝒪3×N(1)\mathcal{O}_{\mathscr{H}_{3}\times\mathbb{P}^{N}}(1) coincides with ω𝒳3/3[m]𝒟3[m(1ϵ)r]\omega_{\mathscr{X}_{3}/\mathscr{H}_{3}}^{[m]}\otimes\mathscr{I}_{\mathscr{D}_{3}}^{[-\frac{m(1-\epsilon)}{r}]}.

By construction, for every \mathbb{Q}-stable pair (X;D)(X;D) in our moduli problem, we have ω𝒳3/3[m]3[m(1ϵ)r]|X𝒪X(m(KX+(1ϵ)D))𝒪N(1)|X\omega_{\mathscr{X}_{3}/\mathscr{H}_{3}}^{[m]}\otimes\mathscr{I}_{3}^{-[\frac{m(1-\epsilon)}{r}]}|_{X}\sim\mathcal{O}_{X}(m(K_{X}+(1-\epsilon)D))\sim\mathcal{O}_{\mathbb{P}^{N}}(1)|_{X}. Since ω𝒳3/3[m]3[m(1ϵ)r]\omega_{\mathscr{X}_{3}/\mathscr{H}_{3}}^{[m]}\otimes\mathscr{I}_{3}^{-[\frac{m(1-\epsilon)}{r}]} is a line bundle and the natural polarization of 2\mathscr{H}_{2} coming from the original choice of Hilbert scheme restricts to 𝒪N(1)\mathcal{O}_{\mathbb{P}^{N}}(1) fiberwise, by [Vie95]*Lemma 1.19 there is a locally closed subscheme 3\mathscr{H}_{3} where ω𝒳3/3[m]3[m(1ϵ)r]\omega_{\mathscr{X}_{3}/\mathscr{H}_{3}}^{[m]}\otimes\mathscr{I}_{3}^{-[\frac{m(1-\epsilon)}{r}]} is linearly equivalent to the natural polarization of 𝒳22\mathscr{X}_{2}\rightarrow\mathscr{H}_{2} over 3\mathscr{H}_{3}.

Step 11: In this step, we show that there is an isomorphism n,p,I[3/PGLN+1]\mathscr{F}_{n,p,I}\cong[\mathscr{H}_{3}/\operatorname{PGL}_{N+1}].

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 PGLN+1\operatorname{PGL}_{N+1}, thus the natural action of PGLN+1\operatorname{PGL}_{N+1} descends onto 3\mathscr{H}_{3}. Then, observe that from its construction, over 3\mathscr{H}_{3} there is a stable family of \mathbb{Q}-stable pairs. This gives a morphism 3n,p,I\mathscr{H}_{3}\to\mathscr{F}_{n,p,I}, and if we forget the embedding into N\mathbb{P}^{N}, this descends to a morphism Φpre:[3/PGLN+1]pren,p,I\Phi^{pre}\colon[\mathscr{H}_{3}/\operatorname{PGL}_{N+1}]^{pre}\to\mathscr{F}_{n,p,I}, where the superscript pre stands for prestack (see [Alp21]*Definition 1.3.12). This induces a map Φ:[3/PGLN+1]n,p,I\Phi\colon[\mathscr{H}_{3}/\operatorname{PGL}_{N+1}]\to\mathscr{F}_{n,p,I}, which we now show is an isomorphism.

To show it is fully faithful, as in [Alp21], it suffices to check that Φpre\Phi^{pre} is fully faithful. But Φpre\Phi^{pre} is fully faithful since any isomorphism between two families of \mathbb{Q}-stable pairs π:(𝒴;𝒟)B\pi\colon(\mathscr{Y};\mathscr{D})\to B and π:(𝒴,𝒟)B\pi\colon(\mathscr{Y}^{\prime},\mathscr{D}^{\prime})\to B over BB sends ω𝒴/B[m]𝒴[m(1ϵ)r]\mathscr{L}\coloneqq\omega_{\mathscr{Y}/B}^{[m]}\otimes\mathscr{I}_{\mathscr{Y}}^{-[\frac{m(1-\epsilon)}{r}]} to ω𝒴/B[m]𝒴[m(1ϵ)r]\mathscr{L}^{\prime}\coloneqq\omega_{\mathscr{Y}^{\prime}/B}^{[m]}\otimes\mathscr{I}_{\mathscr{Y}^{\prime}}^{-[\frac{m(1-\epsilon)}{r}]}, where we denoted by 𝒴\mathscr{I}_{\mathscr{Y}} (resp. 𝒴\mathscr{I}_{\mathscr{Y}^{\prime}}) the ideal sheaves of 𝒟\mathscr{D} (res. 𝒟\mathscr{D}^{\prime}) in 𝒴\mathscr{Y} (resp. 𝒴\mathscr{Y}^{\prime}). This induces an unique isomorphism (π)(π)\mathbb{P}(\pi_{*}\mathscr{L})\cong\mathbb{P}(\pi^{\prime}_{*}\mathscr{L}^{\prime}) which sends 𝒴\mathscr{Y} to 𝒴\mathscr{Y}^{\prime}.

Since Φ\Phi is a morphism of stacks, also essential surjectivity can be checked locally on BB. In particular, it suffices to check that if π:(𝒴;𝒟)B\pi\colon(\mathscr{Y};\mathscr{D})\to B is a family of \mathbb{Q}-stable pairs such that π\pi_{*}\mathscr{L} is free, then the morphism Bn,p,IB\to\mathscr{F}_{n,p,I} lifts to a morphism B3B\to\mathscr{H}_{3}. This follows since if we pick an isomorphism (π)N×B\mathbb{P}(\pi_{*}\mathscr{L})\cong\mathbb{P}^{N}\times B then 𝒴,𝒟(π)=N×B\mathscr{Y},\mathscr{D}\subseteq\mathbb{P}(\pi_{*}\mathscr{L})=\mathbb{P}^{N}\times B, and then from the functorial properties of 3\mathscr{H}_{3} it induces a morphism B3B\to\mathscr{H}_{3}. ∎

Remark 4.3.

Observe that the stack n,p,I\mathscr{F}_{n,p,I} 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 \mathbb{Q}-pair (X;D)(X;D) induces an automorphism of the stable pair (X,(1ϵ)D)(X,(1-\epsilon)D), and those are finite from [KP17]*Proposition 5.5.

5. Properness of n,p,I\mathscr{F}_{n,p,I}

The goal of this section is to prove that n,p,I\mathscr{F}_{n,p,I} is proper. In particular, since in the definition of a \mathbb{Q}-stable pair (X;D)(X;D) there are prescribed conditions on the scheme-theoretic structure of DD. Then, when proving that a moduli functor for \mathbb{Q}-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 DscD^{sc} (recall that DscD^{sc} was introduced in Notation 2.10) is S1S_{1}. This will be the content of the next proposition.

Proposition 5.1.

Let Spec(R)\operatorname{Spec}(R) be the spectrum of a DVR with generic point η\eta and closed point pp. Consider a locally stable family (X,D)Spec(R)(X,D)\to\operatorname{Spec}(R) such that DD is \mathbb{Q}-Cartier. Then, for every m,mm,m^{\prime}\in\mathbb{N}, the sheaf ωX/B[m](mD)\omega_{X/B}^{[m^{\prime}]}(-mD) is S3S_{3} on every point xXpx\in X_{p}. In particular:

  1. (1)

    the restriction ωX/B[m](mD)|Xp\omega_{X/B}^{[m^{\prime}]}(-mD)|_{X_{p}} is S2S_{2}; and

  2. (2)

    if we denote by mDmD the closed subscheme of XX with ideal sheaf 𝒪X(mD)\mathcal{O}_{X}(-mD), then (mD)|Xp(mD)|_{X_{p}} is S1S_{1}.

Proof.

First, observe that since XSpec(R)X\to\operatorname{Spec}(R) is locally stable, xXpx\in X_{p} cannot be a log canonical center for XX (see [kol_new]*Proposition 2.15). The statement is local, so up to shrinking XX, we can assume that XX is affine, and, since both ωX/B\omega_{X/B} and DD are \mathbb{Q}-Cartier, ωX/B[m](mD)𝒪X\omega_{X/B}^{[m^{\prime}]}(-mD)\cong\mathcal{O}_{X} for a certain m0m_{0}\in\mathbb{N}. Then, for every m,mm,m^{\prime}\in\mathbb{N}, we have ωX/B[m0m](m0mD)0\omega_{X/B}^{[m_{0}m^{\prime}]}(-m_{0}mD)\sim_{\mathbb{Q}}0. In particular, if we apply [Kol13]*Theorem 7.20 where, with the notations of loc. cit., we take Δ=0\Delta^{\prime}=0 and if we denote by DKD_{K} the divisor DD in loc. cit., DK=mKX/B+mDD_{K}=-m^{\prime}K_{X/B}+mD, we conclude that ωX/B[m](mD)\omega_{X/B}^{[m^{\prime}]}(-mD) is S3S_{3}.

Now, we denote by π\pi the pull-back to XX of a uniformizer on Spec(R)\operatorname{Spec}(R). Since ωX/B[m](mD)\omega_{X/B}^{[m^{\prime}]}(-mD) is S2S_{2} and π\pi is not a zero divisor on 𝒪X\mathcal{O}_{X}, it is not a zero divisor for ωX/B[m](mD)\omega_{X/B}^{[m^{\prime}]}(-mD). Then,

(ωX/B[m](mD))|Xp(\omega_{X/B}^{[m^{\prime}]}(-mD))|_{X_{p}} is S2S_{2} by [KM98]*Proposition 5.3.

Notice that 𝒪X\mathcal{O}_{X} is S3S_{3} along the special fiber, since it is a flat and proper family of S2S_{2} schemes over a smooth base (see [KM98]*Proposition 5.3). Then, by [Kol13]*Corollary 2.61, 𝒪mD\mathcal{O}_{mD} is S2S_{2} along the special fiber. In particular, it is S1S_{1}, so its generic points are the only associated points. Hence, if we denote by mDmD the closed subscheme of XX with ideal sheaf 𝒪X(mD)\mathcal{O}_{X}(-mD), then mDSpec(R)mD\to\operatorname{Spec}(R) is flat. Then, if we pull back the exact sequence

0𝒪X(mD)𝒪X𝒪mD00\to\mathcal{O}_{X}(-mD)\to\mathcal{O}_{X}\to\mathcal{O}_{mD}\to 0

to XpX_{p}, the sequence remains exact and we get

0(𝒪X(mD))|Xp(𝒪X)|Xp(𝒪mD)|Xp0.0\to(\mathcal{O}_{X}(-mD))|_{X_{p}}\to(\mathcal{O}_{X})|_{X_{p}}\to(\mathcal{O}_{mD})|_{X_{p}}\to 0.

The desired result follows again by [Kol13]*Corollary 2.61. ∎

Proposition 5.2.

Fix an integer nn\in\mathbb{N}, a finite subset I(0,1]I\subseteq(0,1]\cap\mathbb{Q}, and a polynomial p(t)[t]p(t)\in\mathbb{Q}[t]. Assume that II is closed under sum: that is, if a,bIa,b\in I and a+b1a+b\leq 1, then a+bIa+b\in I. Let CC be a smooth affine curve, let 0C0\in C be a distinguished closed point, and let UC{0}U\coloneqq C\setminus\{0\}. Let (𝒳U,ΔU;𝒟U)U(\mathcal{X}_{U},\Delta_{U};\mathcal{D}_{U})\rightarrow U be a weak \mathbb{Q}-stable morphism of relative dimension nn, polynomial p(t)p(t), coefficients in II, and constant part ΔU\Delta_{U}. Further, assume that the geometric generic fiber is normal. Then, (𝒳U,ΔU;𝒟U)U(\mathcal{X}_{U},\Delta_{U};\mathcal{D}_{U})\rightarrow U is a \mathbb{Q}-stable morphism, and, up to a finite base change BCB\rightarrow C, the fibration can be filled with a unique \mathbb{Q}-stable pair of dimension nn, with polynomial p(t)p(t) and coefficients in II.

Proof.

We proceed in several steps. In the following, rr will denote the index of II.

Step 1: In this step, we show that the family of pairs (𝒳U,1r𝒟U+ΔU)U(\mathcal{X}_{U},\frac{1}{r}\mathcal{D}_{U}+\Delta_{U})\rightarrow U admits a relative canonical model (𝒴U,1r𝒟U,𝒴+ΔU,𝒴)(\mathcal{Y}_{U},\frac{1}{r}\mathcal{D}_{U,\mathcal{Y}}+\Delta_{U,\mathcal{Y}}) over UU.

By Proposition 2.37, the fact that CC is smooth and inversion of adjunction, it follows that (𝒳U,1r𝒟U+ΔU)(\mathcal{X}_{U},\frac{1}{r}\mathcal{D}_{U}+\Delta_{U}) 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 (𝒳U,1r𝒟U+ΔU)(\mathcal{X}_{U},\frac{1}{r}\mathcal{D}_{U}+\Delta_{U}) admits a morphism to its relative canonical model (𝒴U,1r𝒟U,𝒴+ΔU,𝒴)(\mathcal{Y}_{U},\frac{1}{r}\mathcal{D}_{U,\mathcal{Y}}+\Delta_{U,\mathcal{Y}}) over UU.

Step 2: In this step, we show that, up to a base change, (𝒴U,1r𝒟U,𝒴+ΔU,𝒴)U(\mathcal{Y}_{U},\frac{1}{r}\mathcal{D}_{U,\mathcal{Y}}+\Delta_{U,\mathcal{Y}})\rightarrow U can be compactified to a family of stable pairs (𝒴,1r𝒟𝒴+Δ𝒴)(\mathcal{Y},\frac{1}{r}\mathcal{D}_{\mathcal{Y}}+\Delta_{\mathcal{Y}}).

This is accomplished in [kol_new]*Theorem 4.59.

Step 3: In this step, we choose a suitable \mathbb{Q}-factorial dlt modification (𝒳,1r𝒟+Δ+𝒳0)(𝒴,1r𝒟𝒴+Δ𝒴+𝒴0)(\mathcal{X}^{\prime},\frac{1}{r}\mathcal{D}^{\prime}+\Delta^{\prime}+\mathcal{X}^{\prime}_{0})\rightarrow(\mathcal{Y},\frac{1}{r}\mathcal{D}_{\mathcal{Y}}+\Delta_{\mathcal{Y}}+\mathcal{Y}_{0}) so that 𝒳U𝒳U\mathcal{X}^{\prime}_{U}\dashrightarrow\mathcal{X}_{U} is a rational contraction over UU.

Since K𝒳U+1ϵr𝒟U+ΔUK_{\mathcal{X}_{U}}+\frac{1-\epsilon}{r}\mathcal{D}_{U}+\Delta_{U} is ample over UU for 0<ϵ10<\epsilon\ll 1, the divisors contracted by 𝒳U𝒴U\mathcal{X}_{U}\rightarrow\mathcal{Y}_{U} are contained in the support of 𝒟\mathcal{D}. Then, by [Mor19]*Theorem 1, we can extract all of these divisors. Finally, we take a \mathbb{Q}-factorial dlt modification of this model just constructed. We denote this model by (𝒳,1r𝒟+Δ+𝒳0)(𝒴,1r𝒟𝒴+Δ𝒴+𝒴0)(\mathcal{X}^{\prime},\frac{1}{r}\mathcal{D}^{\prime}+\Delta^{\prime}+\mathcal{X}^{\prime}_{0})\rightarrow(\mathcal{Y},\frac{1}{r}\mathcal{D}_{\mathcal{Y}}+\Delta_{\mathcal{Y}}+\mathcal{Y}_{0}). In writing (𝒳,1r𝒟+Δ+𝒳0)(\mathcal{X}^{\prime},\frac{1}{r}\mathcal{D}^{\prime}+\Delta^{\prime}+\mathcal{X}^{\prime}_{0}), we define 𝒟\mathcal{D}^{\prime} by first pulling back 𝒟U\mathcal{D}_{U} to a common resolution of 𝒳U\mathcal{X}^{\prime}_{U} and 𝒳U\mathcal{X}_{U}, then pushing it forward to 𝒳U\mathcal{X}^{\prime}_{U}, and finally taking its closure in 𝒳\mathcal{X}^{\prime}. In particular, 𝒟\mathcal{D}^{\prime} is horizontal over CC.

Step 4: In this step, we construct the compactification (𝒳,Δ;1r𝒟)(\mathcal{X},\Delta;\frac{1}{r}\mathcal{D}).

We consider the pair (𝒳,1ϵr𝒟+(Δ)h)(\mathcal{X}^{\prime},\frac{1-\epsilon}{r}\mathcal{D}^{\prime}+(\Delta^{\prime})^{h}) for 0<ϵ10<\epsilon\ll 1, where the notation (Δ)h(\Delta^{\prime})^{h} stands for horizontal over CC. Then, (𝒳U,1ϵr𝒟U+ΔU)(\mathcal{X}_{U},\frac{1-\epsilon}{r}\mathcal{D}_{U}+\Delta_{U}) is a relative good minimal model for (𝒳U,1ϵr𝒟U+(ΔU)h)(\mathcal{X}^{\prime}_{U},\frac{1-\epsilon}{r}\mathcal{D}^{\prime}_{U}+(\Delta^{\prime}_{U})^{h}) over 𝒴U\mathcal{Y}_{U}. Then, by [HX13], (𝒳,1ϵr𝒟+(Δ)h)(\mathcal{X}^{\prime},\frac{1-\epsilon}{r}\mathcal{D}^{\prime}+(\Delta^{\prime})^{h}) admits a relative good minimal model over 𝒴\mathcal{Y}. We denote its relative canonical model over 𝒴\mathcal{Y} by (𝒳,1ϵr𝒟+Δh)(\mathcal{X},\frac{1-\epsilon}{r}\mathcal{D}+\Delta^{h}).

First, we observe that Δh=Δ\Delta^{h}=\Delta. Notice that ΔU\Delta_{U} is horizontal over UU by assumption, so Δv\Delta^{v}, which denotes the vertical components of Δ\Delta, has to be supported on 𝒳0\mathcal{X}_{0}. Then, the claim follows, since (𝒴,1r𝒟𝒴+Δ𝒴+𝒴0)(\mathcal{Y},\frac{1}{r}\mathcal{D}_{\mathcal{Y}}+\Delta_{\mathcal{Y}}+\mathcal{Y}_{0}) is log canonical.

Now, we observe that 𝒟\mathcal{D} is \mathbb{Q}-Cartier. Indeed, K𝒳+1r𝒟+ΔK_{\mathcal{X}}+\frac{1}{r}\mathcal{D}+\Delta is \mathbb{Q}-Cartier, as it is the pull-back of K𝒴+1r𝒟𝒴+Δ𝒴K_{\mathcal{Y}}+\frac{1}{r}\mathcal{D}_{\mathcal{Y}}+\Delta_{\mathcal{Y}}. By construction, we have that K𝒳+1ϵr𝒟+ΔK_{\mathcal{X}}+\frac{1-\epsilon}{r}\mathcal{D}+\Delta is ample over 𝒴\mathcal{Y}, and in particular it is \mathbb{Q}-Cartier. Then, 𝒟\mathcal{D} is the difference of two \mathbb{Q}-Cartier divisors.

Since 𝒟\mathcal{D} is \mathbb{Q}-Cartier, it follows that this canonical model is independent of 0<ϵ10<\epsilon\ll 1, as K𝒳+1ϵr𝒟+Δ,𝒴ϵr𝒟K_{\mathcal{X}}+\frac{1-\epsilon}{r}\mathcal{D}+\Delta\sim_{\mathbb{Q},\mathcal{Y}}-\frac{\epsilon}{r}\mathcal{D}. Furthermore, if 0<ϵ10<\epsilon\ll 1, we have that K𝒳+1ϵr𝒟+ΔK_{\mathcal{X}}+\frac{1-\epsilon}{r}\mathcal{D}+\Delta is ample over CC, as it is ample over 𝒴\mathcal{Y} and K𝒴+1r𝒟𝒴+Δ𝒴K_{\mathcal{Y}}+\frac{1}{r}\mathcal{D}_{\mathcal{Y}}+\Delta_{\mathcal{Y}} is ample over CC.

Step 5: In this step, we show that (𝒳,Δ;1r𝒟)(\mathcal{X},\Delta;\frac{1}{r}\mathcal{D}) is a \mathbb{Q}-pair and that the central fiber of (𝒳,Δ;1r𝒟)(\mathcal{X},\Delta;\frac{1}{r}\mathcal{D}) is a \mathbb{Q}-stable pair with polynomial p(t)p(t) and coefficients in II.

Recall that 𝒳\mathcal{X} is independent of ϵ\epsilon, and that 𝒟\mathcal{D} is \mathbb{Q}-Cartier. As (𝒳,1r𝒟+Δ)(\mathcal{X},\frac{1}{r}\mathcal{D}+\Delta) is log canonical by construction, it follows that (𝒳,Δ;1r𝒟)(\mathcal{X},\Delta;\frac{1}{r}\mathcal{D}) is a \mathbb{Q}-pair.

By construction and by adjunction, (𝒳0,1r𝒟0+Δ0)(\mathcal{X}_{0},\frac{1}{r}\mathcal{D}_{0}+\Delta_{0}) is semi-log canonical. Furthermore, K𝒳0+1ϵr𝒟0+Δ0K_{\mathcal{X}_{0}}+\frac{1-\epsilon}{r}\mathcal{D}_{0}+\Delta_{0} is ample for 0<ϵ10<\epsilon\ll 1. Since 𝒟\mathcal{D} and K𝒳+1r𝒟+ΔK_{\mathcal{X}}+\frac{1}{r}\mathcal{D}+\Delta are \mathbb{Q}-Cartier, the self-intersection (K𝒳c+1r𝒟c+Δcϵ𝒟c)dim(𝒳0)(K_{\mathcal{X}_{c}}+\frac{1}{r}\mathcal{D}_{c}+\Delta_{c}-\epsilon\mathcal{D}_{c})^{{\rm dim}(\mathcal{X}_{0})} is well-defined for every ϵ\epsilon and cCc\in C, and does not depend on cCc\in C. Since the general fiber is \mathbb{Q}-stable with polynomial p(t)p(t), we have (K𝒳0+1r𝒟0+Δ0ϵ𝒟0)dim(𝒳0)=p(1ϵ)(K_{\mathcal{X}_{0}}+\frac{1}{r}\mathcal{D}_{0}+\Delta_{0}-\epsilon\mathcal{D}_{0})^{{\rm dim}(\mathcal{X}_{0})}=p(1-\epsilon). The coefficients of 1r𝒟0\frac{1}{r}\mathcal{D}_{0} are still in II, since II is closed under addition.

Step 6: In this step, we show that (𝒳,Δ;𝒟)C(\mathcal{X},\Delta;\mathcal{D})\rightarrow C is a \mathbb{Q}-stable morphism with constant part Δ\Delta.

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 \mathbb{Q}-stable pair. Then, as 𝒟\mathcal{D} is \mathbb{Q}-Cartier, every fiber of (𝒳,Δ)C(\mathcal{X},\Delta)\rightarrow C is semi-log canonical. Thus, by [kol_new]*Definition-Theorem 4.7, the morphism (𝒳,Δ)C(\mathcal{X},\Delta)\rightarrow C is locally stable.

Step 7: In this step, we show that the limit is unique.

From Theorem 3.2, there exists ϵ0>0\epsilon_{0}>0 such that, for every \mathbb{Q}-stable pair (Z,B,Γ)(Z,B,\Gamma) with coefficients in II and polynomial p(t)p(t), (Z,(1ϵ0)B+Γ)(Z,(1-\epsilon_{0})B+\Gamma) is a stable pair. Then, the claim follows from the separatedness of stable morphisms [kol_new]*2.49. ∎

Theorem 5.3.

Fix an integer nn\in\mathbb{N}, a finite subset I(0,1]I\subseteq(0,1]\cap\mathbb{Q}, and a polynomial p(t)[t]p(t)\in\mathbb{Q}[t]. Assume that II is closed under sum: that is, if a,bIa,b\in I and a+b1a+b\leq 1, then a+bIa+b\in I. Let CC be a smooth affine curve, let 0C0\in C be a distinguished closed point, and let UC{0}U\coloneqq C\setminus\{0\}. Let (𝒳U,ΔU;𝒟U)U(\mathcal{X}_{U},\Delta_{U};\mathcal{D}_{U})\rightarrow U be a weak \mathbb{Q}-stable morphism of dimension nn, polynomial p(t)p(t), coefficients in II, and constant part ΔU\Delta_{U}. Then, (𝒳U,ΔU;𝒟U)U(\mathcal{X}_{U},\Delta_{U};\mathcal{D}_{U})\rightarrow U is a \mathbb{Q}-stable morphism, and, up to a finite base change BCB\rightarrow C, the fibration can be filled with a unique \mathbb{Q}-stable pair of dimension nn, with polynomial p(t)p(t) and coefficients in II.

Proof.

By Proposition 5.2, we may assume that the geometric generic fiber is not normal. In the following, rr will denote the index of II.

Let (𝒳¯U,1r𝒟¯U+Δ¯U)(\overline{\mathcal{X}}_{U},\frac{1}{r}\overline{\mathcal{D}}_{U}+\overline{\Delta}_{U}) denote the normalization of (𝒳U,1r𝒟U+ΔU)(\mathcal{X}_{U},\frac{1}{r}\mathcal{D}_{U}+\Delta_{U}), where Δ¯U\overline{\Delta}_{U} also includes the conductor with coefficient 1. By assumption, 𝒟U\mathcal{D}_{U} is \mathbb{Q}-Cartier. Then, by [Kol13]*Corollary 5.39, 𝒟¯U\overline{\mathcal{D}}_{U} is \mathbb{Q}-Cartier. Then, since K𝒳U+1ϵr𝒟U+ΔUK_{\mathcal{X}_{U}}+\frac{1-\epsilon}{r}\mathcal{D}_{U}+\Delta_{U} is ample over UU, it follows that (𝒳¯U,Δ¯U;𝒟¯U)(\overline{\mathcal{X}}_{U},\overline{\Delta}_{U};\overline{\mathcal{D}}_{U}) is \mathbb{Q}-stable morphism with constant part Δ¯U\overline{\Delta}_{U}, where the polynomial on each connected component depends on the original choice of pp.

Then, as 𝒳¯U\overline{\mathcal{X}}_{U} has finitely many connected components, by Proposition 5.2, there is a finite base change BCB\rightarrow C such that the family can be filled with a unique \mathbb{Q}-stable pair. To simplify the notation, we omit the base change BCB\rightarrow C, and we assume that the filling is realized over CC itself. We denote this family by (𝒳¯,Δ¯;𝒟¯)C(\overline{\mathcal{X}},\overline{\Delta};\overline{\mathcal{D}})\rightarrow C. Then, we may find 0<ϵ10<\epsilon\ll 1 such that (𝒳¯,1ϵr𝒟¯+Δ¯)C(\overline{\mathcal{X}},\frac{1-\epsilon}{r}\overline{\mathcal{D}}+\overline{\Delta})\rightarrow C is a stable morphism. By [kol_new]*Lemma 2.54, also (𝒳U,1ϵr𝒟U+ΔU)(\mathcal{X}_{U},\frac{1-\epsilon}{r}\mathcal{D}_{U}+\Delta_{U}) admits a compactification (𝒳,1ϵr𝒟+Δ)(\mathcal{X},\frac{1-\epsilon}{r}\mathcal{D}+\Delta) over CC obtained by gluing 𝒳¯\overline{\mathcal{X}} along some components of Δ¯\overline{\Delta}. By [Kol13]*Corollary 5.39, the divisor 𝒟\mathcal{D} is \mathbb{Q}-Cartier. Thus, we have that 𝒟0\mathcal{D}_{0} is \mathbb{Q}-Cartier, as needed. Similarly, the coefficients of 1r𝒟0+Δ0\frac{1}{r}\mathcal{D}_{0}+\Delta_{0} are in II, by construction.

Now, we prove that the special fiber (𝒳0,Δ0;1r𝒟0)(\mathcal{X}_{0},\Delta_{0};\frac{1}{r}\mathcal{D}_{0}) is a \mathbb{Q}-stable pair. We already verified that 𝒟0\mathcal{D}_{0} is \mathbb{Q}-Cartier. Now, we verify that (𝒳0,Δ0;1r𝒟0)(\mathcal{X}_{0},\Delta_{0};\frac{1}{r}\mathcal{D}_{0}) is semi-log canonical. By construction, (𝒳0,1ϵr𝒟0+Δ0)(\mathcal{X}_{0},\frac{1-\epsilon}{r}\mathcal{D}_{0}+\Delta_{0}) is semi-log canonical. Thus, by [Kol13]*Definition-Lemma 5.10, it suffices to show that the normalization of (𝒳0,1r𝒟0+Δ0)(\mathcal{X}_{0},\frac{1}{r}\mathcal{D}_{0}+\Delta_{0}) is log canonical. But then, this holds by construction, as its normalization coincides with the normalization of the (possibly disconnected) semi-log canonical pair (𝒳¯0,1r𝒟¯0+Δ¯0)(\overline{\mathcal{X}}_{0},\frac{1}{r}\overline{\mathcal{D}}_{0}+\overline{\Delta}_{0}). The same argument applied to the total space shows that (X¯,1r𝒟+Δ)(\overline{X},\frac{1}{r}\mathcal{D}+\Delta) is semi-log canonical.

Lastly, we need to check that K𝒳0+1r𝒟0+Δ0K_{\mathcal{X}_{0}}+\frac{1}{r}\mathcal{D}_{0}+\Delta_{0} is semi-ample. To this end, by adjunction, it suffices to show the stronger statement that K𝒳+1r𝒟+ΔK_{\mathcal{X}}+\frac{1}{r}\mathcal{D}+\Delta is semi-ample over CC. By Proposition 5.2, this is true for K𝒳¯+1r𝒟¯+Δ¯K_{\overline{\mathcal{X}}}+\frac{1}{r}\overline{\mathcal{D}}+\overline{\Delta}. Then, the claim follows by [HX16]*Theorem 2. Then, since 𝒟\mathcal{D} does not contain any irreducible component of the double locus of 𝒳\mathcal{X}, it is immediate that the relative canonical model of (𝒳,1r𝒟+Δ)(\mathcal{X},\frac{1}{r}\mathcal{D}+\Delta) coincides with the gluing of the relative canonical models of the irreducible components of (𝒳¯,1r𝒟¯+Δ¯)(\overline{\mathcal{X}},\frac{1}{r}\overline{\mathcal{D}}+\overline{\Delta}), 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 𝒳¯\overline{\mathcal{X}} to the canonical model is contained in 𝒟¯\overline{\mathcal{D}} and 𝒟¯\overline{\mathcal{D}} does not contain any irreducible component of the conductor, the involution defined on the conductor via the normalization of (𝒳,1r𝒟+Δ)(\mathcal{X},\frac{1}{r}\mathcal{D}+\Delta) naturally descends to the canonical model of (𝒳¯,1r𝒟¯+Δ¯)(\overline{\mathcal{X}},\frac{1}{r}\overline{\mathcal{D}}+\overline{\Delta}). We observe that the involution defined on the conductor via the normalization of (𝒳,1r𝒟+Δ)(\mathcal{X},\frac{1}{r}\mathcal{D}+\Delta) 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 (K𝒳0+tr𝒟0+Δ0)n=p(t)(K_{\mathcal{X}_{0}}+\frac{t}{r}\mathcal{D}_{0}+\Delta_{0})^{n}=p(t). This follows from flatness over the base CC, as we have

(K𝒳0+tr𝒟0+Δ0)n=(K𝒳c+tr𝒟c+Δc)n=p(t),(K_{\mathcal{X}_{0}}+\frac{t}{r}\mathcal{D}_{0}+\Delta_{0})^{n}=(K_{\mathcal{X}_{c}}+\frac{t}{r}\mathcal{D}_{c}+\Delta_{c})^{n}=p(t),

where cC{0}c\in C\setminus\{0\}. This concludes the proof. ∎

Corollary 5.4.

Fix an integer nn\in\mathbb{N}, a finite subset I(0,1]I\subseteq(0,1]\cap\mathbb{Q}, and a polynomial p(t)[t]p(t)\in\mathbb{Q}[t]. Assume that II is closed under sum: that is, if a,bIa,b\in I and a+b1a+b\leq 1, then a+bIa+b\in I. Then the algebraic stack n,p,I\mathscr{F}_{n,p,I} is proper.

Proof.

It suffices to check that it satisfies the valuative criterion for properness. So assume that we have a family of \mathbb{Q}-stable pairs fη:(X;D)ηf_{\eta}\colon(X;D)\to\eta over the generic point η\eta of the spectrum of a DVR RR, and we need to fill in the central fiber up to replacing Spec(R)\operatorname{Spec}(R) with a ramified cover of it. Theorem 5.3 guarantees the existence and uniqueness of a \mathbb{Q}-stable morphism f:(𝒳;𝒟)Spec(R)f\colon(\mathcal{X};\mathcal{D})\to\operatorname{Spec}(R) extending fηf_{\eta}, up to a ramified cover of Spec(R)\operatorname{Spec}(R). To check that ff satisfies condition (K) of Definition 2.19 we use Proposition 5.1 point (1). ∎

6. Relative canonical models over reduced base

Given a pair (X,D)(X,D) with KX+DK_{X}+D semi-ample and big, one can take its canonical model. Similarly, from Definition 2.21, if one starts with a family of \mathbb{Q}-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 \mathbb{Q}-stable pairs are actually families of \mathbb{Q}-stable pairs.

Lemma 6.1.

Fix an integer nn\in\mathbb{N}, a finite subset I(0,1]I\subseteq(0,1]\cap\mathbb{Q}, and a polynomial p(t)[t]p(t)\in\mathbb{Q}[t]. Let rr denote the index of II. Consider a weak \mathbb{Q}-stable morphism with coefficients in II and polynomial p(t)p(t) over a smooth scheme UU: p:(𝒳;𝒟)Up\colon(\mathcal{X};\mathcal{D})\to U. Then, there is a stable family of pairs (𝒴,1r𝒟𝒴)U(\mathcal{Y},\frac{1}{r}\mathcal{D}_{\mathcal{Y}})\to U such that:

  1. (1)

    there is a contraction π:𝒳𝒴\pi\colon\mathcal{X}\to\mathcal{Y}; and

  2. (2)

    we have π(K𝒴/U+1r𝒟𝒴)=K𝒳/U+1r𝒟\pi^{*}(K_{\mathcal{Y}/U}+\frac{1}{r}\mathcal{D}_{\mathcal{Y}})=K_{\mathcal{X}/U}+\frac{1}{r}\mathcal{D}.

In particular, p:(𝒳;𝒟)Up\colon(\mathcal{X};\mathcal{D})\to U is a \mathbb{Q}-stable morphism.

Proof.

By Proposition 2.37 and the fact that UU is smooth, it follows that (𝒳,1r𝒟)(\mathcal{X},\frac{1}{r}\mathcal{D}) is a pair. Furthermore, by inversion of adjunction, this pair is semi-log canonical.

First, we assume that 𝒳\mathcal{X} is normal. Then, by Lemma 2.38 and our assumptions, the canonical model of (𝒳,1r𝒟)(\mathcal{X},\frac{1}{r}\mathcal{D}) over UU exists. Since UU 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 𝒳\mathcal{X} is not normal. First, we normalize n𝒳:𝒳n𝒳n_{\mathcal{X}}\colon\mathcal{X}^{n}\to\mathcal{X} to get (𝒳n,Δ;𝒟)(\mathcal{X}^{n},\Delta;\mathcal{D}^{\prime}). As argued in the previous case, we can construct the canonical model of (𝒳n,Δ;𝒟)(\mathcal{X}^{n},\Delta;\mathcal{D}^{\prime}) over UU. As above, this gives a family of stable pairs (𝒴,1r𝒟𝒴+Δ𝒴)U(\mathcal{Y}^{\prime},\frac{1}{r}\mathcal{D}_{\mathcal{Y}}^{\prime}+\Delta_{\mathcal{Y}})\to U. From Kollár’s gluing theory (see [Kol13]*Ch. 5), there is an involution τ:ΔnΔn\tau\colon\Delta^{n}\to\Delta^{n} that fixes the different. We first show that this involution descends onto Δ𝒴n\Delta_{\mathcal{Y}}^{n}.

Recall that, by Lemma 2.13, the map 𝒳n𝒴\mathcal{X}^{n}\to\mathcal{Y}^{\prime} does not contract any component of Δ\Delta. In particular, for every irreducible component FΔnF\subseteq\Delta^{n}, there is an irreducible component FYΔ𝒴nF_{Y}\subseteq\Delta_{\mathcal{Y}}^{n} birational to it, and we have the following diagram:

Δn\textstyle{\Delta^{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f\scriptstyle{f}Δ𝒴n\textstyle{\Delta^{n}_{\mathcal{Y}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒳n\textstyle{\mathcal{X}^{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}πn\scriptstyle{\pi_{n}}𝒴.\textstyle{\mathcal{Y}^{\prime}.}

Observe now that

(\ast) a curve CΔnC\subset\Delta^{n} gets contracted by ff if and only if (KΔn+DiffΔn(1r𝒟+Δ)).C=0(K_{\Delta^{n}}+\operatorname{Diff}_{\Delta^{n}}(\frac{1}{r}\mathcal{D}^{\prime}+\Delta)).C=0.

In particular, since the involution τ\tau preserves the different, it preserves all the curves that are contracted by ff. Hence, the involution descends to an involution τ𝒴\tau_{\mathcal{Y}} on Δ𝒴n\Delta_{\mathcal{Y}}^{n}.

We prove that τ𝒴\tau_{\mathcal{Y}} preserves the different. Indeed, by Lemma 2.13, the only divisors contracted by ff are contained in Supp(𝒟)\operatorname{Supp}(\mathcal{D}^{\prime}), so the morphism ff is an isomorphism generically around each divisor not contained in Supp(𝒟)\operatorname{Supp}(\mathcal{D}). In particular, since the computation of the different is local,

f(DiffΔn(1r𝒟+Δ))=DiffΔ𝒴n(1r𝒟𝒴+Δ𝒴).f_{*}(\operatorname{Diff}_{\Delta^{n}}(\frac{1}{r}\mathcal{D}^{\prime}+\Delta))=\operatorname{Diff}_{\Delta^{n}_{\mathcal{Y}}}(\frac{1}{r}\mathcal{D}^{\prime}_{\mathcal{Y}}+\Delta_{\mathcal{Y}}).

Since τ\tau preserves the different on Δn\Delta^{n}, τ𝒴\tau_{\mathcal{Y}} preserves the different on Δ𝒴n\Delta_{\mathcal{Y}}^{n}.

Then from [Kol13]*Ch. 5, we can glue 𝒴\mathcal{Y}^{\prime} to get an semi-log canonical pair (𝒴,1r𝒟𝒴)(\mathcal{Y},\frac{1}{r}\mathcal{D}_{\mathcal{Y}}), and let n𝒴:𝒴𝒴n_{\mathcal{Y}}\colon\mathcal{Y}^{\prime}\to\mathcal{Y} be the normalization. Now, recall that 𝒴\mathcal{Y} is a geometric quotient (see [Kol13]*Theorem 5.32 and the proof of [Kol13]*Corollary 5.33), so for any morphism ϕ:𝒴nZ\phi\colon\mathcal{Y}^{n}\to Z such that ϕ|Δ𝒴n=τϕ|Δ𝒴n\phi|_{\Delta_{\mathcal{Y}}^{n}}=\tau\circ\phi|_{\Delta_{\mathcal{Y}}^{n}}, there is a unique morphism 𝒴Z\mathcal{Y}\to Z which fits in the obvious commutative diagram. Therefore, we have a morphism 𝒴U\mathcal{Y}\to U, and applying this result to 𝒳n\mathcal{X}^{n} and 𝒳\mathcal{X}, we obtain a morphism π:𝒳𝒴\pi\colon\mathcal{X}\to\mathcal{Y}.

To show (1) it suffices to observe the following two exact sequence:

(7) 0𝒪𝒳(n𝒳)𝒪𝒳ng𝒳(n𝒳)𝒪Δn.0\to\mathcal{O}_{\mathcal{X}}\to(n_{\mathcal{X}})_{*}\mathcal{O}_{\mathcal{X}^{n}}\xrightarrow{g_{\mathcal{X}}}(n_{\mathcal{X}})_{*}\mathcal{O}_{\Delta^{n}}.

Observe that πn𝒳=n𝒴πn\pi\circ n_{\mathcal{X}}=n_{\mathcal{Y}}\circ\pi_{n} and both πn\pi_{n} and ΔnΔ𝒴n\Delta^{n}\to\Delta^{n}_{\mathcal{Y}} is birational. So (πn)𝒪𝒳n=𝒪𝒴(\pi_{n})_{*}\mathcal{O}_{\mathcal{X}^{n}}=\mathcal{O}_{\mathcal{Y}^{\prime}} and (πn𝒳)𝒪Δn=𝒪Δ𝒴n.(\pi\circ n_{\mathcal{X}})_{*}\mathcal{O}_{\Delta^{n}}=\mathcal{O}_{\Delta^{n}_{\mathcal{Y}}}. Therefore pushing forward the sequence (7) via π\pi we have

0π𝒪𝒳(n𝒴)𝒪𝒴g𝒳(n𝒳)𝒪Δ𝒴n.0\to\pi_{*}\mathcal{O}_{\mathcal{X}}\to(n_{\mathcal{Y}})_{*}\mathcal{O}_{\mathcal{Y}^{\prime}}\xrightarrow{g_{\mathcal{X}}}(n_{\mathcal{X}})_{*}\mathcal{O}_{\Delta^{n}_{\mathcal{Y}}}.

In particular, π𝒪𝒳=𝒪𝒴\pi_{*}\mathcal{O}_{\mathcal{X}}=\mathcal{O}_{\mathcal{Y}}. Similarly we can tensor the sequence above by π(K𝒴/U+1r𝒟)\pi^{*}(K_{\mathcal{Y}/U}+\frac{1}{r}\mathcal{D}) to deduce (2). ∎

Theorem 6.2.

Fix an integer nn\in\mathbb{N} and a finite subset I(0,1]I\subseteq(0,1]\cap\mathbb{Q}. Let rr denote the index of II. Assume that the fibers of pp have an canonical model. Let (𝒳;𝒟)B(\mathcal{X};\mathcal{D})\to B be a weak \mathbb{Q}-stable morphism of dimension nn, with coefficients in II, and polynomial p(t)p(t), over a reduced connected base BB, and assume that there is an open dense subscheme UBU\subseteq B with:

  1. (1)

    a stable family of pairs (𝒴U,1r𝒟𝒴,U)U(\mathcal{Y}_{U},\frac{1}{r}\mathcal{D}_{\mathcal{Y},U})\to U of relative dimension nn; and

  2. (2)

    a contraction πU:𝒳U𝒴U\pi_{U}\colon\mathcal{X}_{U}\to\mathcal{Y}_{U} such that π(K𝒴U/U+1r𝒟𝒴,U)=K𝒳U/U+1r𝒟U\pi^{*}(K_{\mathcal{Y}_{U}/U}+\frac{1}{r}\mathcal{D}_{\mathcal{Y},U})=K_{\mathcal{X}_{U}/U}+\frac{1}{r}\mathcal{D}_{U}.

Then, there there is an m>0m>0 such that for every dd and every bBb\in B we have

p(𝒪𝒳(md(K𝒳/B+1r𝒟)))k(b)=H0(𝒳b,𝒪𝒳b(md(K𝒳b+1r𝒟b))).p_{*}(\mathcal{O}_{\mathcal{X}}(md(K_{\mathcal{X}/B}+\frac{1}{r}\mathcal{D})))\otimes k(b)=H^{0}(\mathcal{X}_{b},\mathcal{O}_{\mathcal{X}_{b}}(md(K_{\mathcal{X}_{b}}+\frac{1}{r}\mathcal{D}_{b}))).

Moreover, if we define 𝒴Proj(dp𝒪𝒳(md(K𝒳/B+1r𝒟)))\mathcal{Y}\coloneqq\operatorname{Proj}(\bigoplus_{d}p_{*}\mathcal{O}_{\mathcal{X}}(md(K_{\mathcal{X}/B}+\frac{1}{r}\mathcal{D}))), then:

  • there is a family of divisors 𝒟𝒴\mathcal{D}_{\mathcal{Y}} such that the pair q:(𝒴,1r𝒟𝒴)Bq\colon(\mathcal{Y},\frac{1}{r}\mathcal{D}_{\mathcal{Y}})\to B is a stable family extending qUq_{U};

  • there is a contraction π:𝒳𝒴\pi\colon\mathcal{X}\to\mathcal{Y} over BB that extends πU\pi_{U}; and

  • π(K𝒴/B+1r𝒟𝒴)=K𝒳/B+1r𝒟\pi^{*}(K_{\mathcal{Y}/B}+\frac{1}{r}\mathcal{D}_{\mathcal{Y}})=K_{\mathcal{X}/B}+\frac{1}{r}\mathcal{D}.

In particular, (𝒳,𝒟)B(\mathcal{X},\mathcal{D})\rightarrow B is a \mathbb{Q}-stable morphism.

Remark 6.3.

Observe that since K𝒴U/U+1r𝒟𝒴,UK_{\mathcal{Y}_{U}/U}+\frac{1}{r}\mathcal{D}_{\mathcal{Y},U} is \mathbb{Q}-Cartier, we can define its pull-back as a \mathbb{Q}-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 K𝒳/B+1r𝒟K_{\mathcal{X}/B}+\frac{1}{r}\mathcal{D} and 𝒟\mathcal{D} are \mathbb{Q}-Cartier. Then, observe that for every s,tUs,t\in U, the volumes of the pairs (𝒴s,1r𝒟𝒴,s)(\mathcal{Y}_{s},\frac{1}{r}\mathcal{D}_{\mathcal{Y},s}) and (𝒴t,1r𝒟𝒴,t)(\mathcal{Y}_{t},\frac{1}{r}\mathcal{D}_{\mathcal{Y},t}) agree. Indeed, since (𝒳;𝒟)B(\mathcal{X};\mathcal{D})\to B is a weak \mathbb{Q}-stable morphism, K𝒳s+1r𝒟sK_{\mathcal{X}_{s}}+\frac{1}{r}\mathcal{D}_{s} and K𝒳t+1r𝒟tK_{\mathcal{X}_{t}}+\frac{1}{r}\mathcal{D}_{t} are nef. Thus, their volumes are computed by the nn-fold self-intersection, which is independent of s,tSs,t\in S. But from condition (2) the morphisms πs:𝒳s𝒴s\pi_{s}\colon\mathcal{X}_{s}\to\mathcal{Y}_{s} and πt:𝒳t𝒴t\pi_{t}\colon\mathcal{X}_{t}\to\mathcal{Y}_{t} have connected fibers, and we have πs(K𝒴s+1r𝒟𝒴,s)=K𝒳s+1r𝒟s\pi_{s}^{*}(K_{\mathcal{Y}_{s}}+\frac{1}{r}\mathcal{D}_{\mathcal{Y},s})=K_{\mathcal{X}_{s}}+\frac{1}{r}\mathcal{D}_{s} and πt(K𝒴t+1r𝒟𝒴,t)=K𝒳t+1r𝒟t\pi_{t}^{*}(K_{\mathcal{Y}_{t}}+\frac{1}{r}\mathcal{D}_{\mathcal{Y},t})=K_{\mathcal{X}_{t}}+\frac{1}{r}\mathcal{D}_{t}. Then, by Remark 2.4, the volumes of K𝒴s+1r𝒟𝒴,sK_{\mathcal{Y}_{s}}+\frac{1}{r}\mathcal{D}_{\mathcal{Y},s} (resp. K𝒴t+1r𝒟𝒴,tK_{\mathcal{Y}_{t}}+\frac{1}{r}\mathcal{D}_{\mathcal{Y},t}) and K𝒳s+1r𝒟sK_{\mathcal{X}_{s}}+\frac{1}{r}\mathcal{D}_{s} (resp. K𝒳t+1r𝒟tK_{\mathcal{X}_{t}}+\frac{1}{r}\mathcal{D}_{t}) agree. Let then vv be the volume of any fiber of qUq_{U} and let kk be a natural number such that rr divides kk and, for every stable pair (Y,D)(Y,D) of dimension nn, volume vv and coefficients in II, the line bundle 𝒪Y(k(KY+D))\mathcal{O}_{Y}(k(K_{Y}+D)) is very ample and the higher cohomologies of all of its natural multiples vanish. Notice that kk exists by [HMX18]*Theorem 1.2.2. Then, we set 𝒪𝒳(k(K𝒳/B+1r𝒟))\mathcal{L}\coloneqq\mathcal{O}_{\mathcal{X}}(k(K_{\mathcal{X}/B}+\frac{1}{r}\mathcal{D})). Up to replacing kk with a multiple, we may further assume that \mathcal{L} is Cartier.

Step 2: In this step, we show that the theorem holds if BB is a smooth curve.

From Lemma 6.1, we can construct a family of stable pairs (𝒵,1r𝒟𝒵)B(\mathcal{Z},\frac{1}{r}\mathcal{D}_{\mathcal{Z}})\to B with a contraction ϕ:𝒳𝒵\phi\colon\mathcal{X}\to\mathcal{Z}. First, observe that (𝒵U,1r𝒟𝒵,U)(𝒴U,1r𝒟𝒴,U)(\mathcal{Z}_{U},\frac{1}{r}\mathcal{D}_{\mathcal{Z},U})\cong(\mathcal{Y}_{U},\frac{1}{r}\mathcal{D}_{\mathcal{Y},U}) over UU. Indeed, consider the reflexive sheaves 𝒵𝒪𝒵(k(K𝒵+1r𝒟𝒵))\mathcal{L}_{\mathcal{Z}}\coloneqq\mathcal{O}_{\mathcal{Z}}(k(K_{\mathcal{Z}}+\frac{1}{r}\mathcal{D}_{\mathcal{Z}})) and 𝒴,U𝒪𝒴U(k(K𝒴U+1r𝒟𝒴,U))\mathcal{L}_{\mathcal{Y},U}\coloneqq\mathcal{O}_{\mathcal{Y}_{U}}(k(K_{\mathcal{Y}_{U}}+\frac{1}{r}\mathcal{D}_{\mathcal{Y},U})). By construction, the fibers of (𝒵,1r𝒟𝒵)B(\mathcal{Z},\frac{1}{r}\mathcal{D}_{\mathcal{Z}})\rightarrow B and (𝒴U,1r𝒟𝒴,U)U(\mathcal{Y}_{U},\frac{1}{r}\mathcal{D}_{\mathcal{Y},U})\rightarrow U belong to the moduli problem of stable pairs with volume vv and coefficients in the finite set II. Thus, by [kol_new]*Theorem 5.8.(4) and the choice of kk, 𝒵\mathcal{L}_{\mathcal{Z}} and 𝒴,U\mathcal{L}_{\mathcal{Y},U} are line bundles. In order to apply [kol_new]*Theorem 5.8.(4), notice that we know that 𝒵\mathcal{L}_{\mathcal{Z}} and 𝒴,U\mathcal{L}_{\mathcal{Y},U} are line bundles away from the exceptional locus of ϕ\phi and πU\pi_{U}, which are big open subsets restricting to big open subsets fiberwise.

Since πU\pi_{U} and ϕU\phi_{U} have connected fibers, we have

πU𝒴,UUϕU𝒵,U.\pi_{U}^{*}\mathcal{L}_{\mathcal{Y},U}\cong\mathcal{L}_{U}\cong\phi^{*}_{U}\mathcal{L}_{\mathcal{Z},U}.

But both πU\pi_{U} and ϕU\phi_{U} have connected fibers so, by the projection formula, for every m1m\geq 1, we have

H0(𝒴U,𝒴,Um)=H0(𝒳U,πU𝒴,Um)=H0(𝒳U,Um)=H0(𝒳U,ϕU𝒵,Um)=H0(𝒵U,𝒵,Um).H^{0}(\mathcal{Y}_{U},\mathcal{L}_{\mathcal{Y},U}^{\otimes m})=H^{0}(\mathcal{X}_{U},\pi^{*}_{U}\mathcal{L}_{\mathcal{Y},U}^{\otimes m})=H^{0}(\mathcal{X}_{U},\mathcal{L}_{U}^{\otimes m})=H^{0}(\mathcal{X}_{U},\phi^{*}_{U}\mathcal{L}_{\mathcal{Z},U}^{\otimes m})=H^{0}(\mathcal{Z}_{U},\mathcal{L}_{\mathcal{Z},U}^{\otimes m}).

But 𝒵,U\mathcal{L}_{\mathcal{Z},U} and 𝒴,U\mathcal{L}_{\mathcal{Y},U} are ample over UU, so 𝒵U\mathcal{Z}_{U} and 𝒴U\mathcal{Y}_{U} are isomorphic as UU-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 (𝒵,1r𝒟𝒵)(\mathcal{Z},\frac{1}{r}\mathcal{D}_{\mathcal{Z}}).

Thus, we are left with proving that for every dd and every bBb\in B, we have p(d)k(b)=H0(𝒳b,bd)p_{*}(\mathcal{L}^{\otimes d})\otimes k(b)=H^{0}(\mathcal{X}_{b},\mathcal{L}^{\otimes d}_{b}). But since ϕ𝒵=\phi^{*}\mathcal{L}_{\mathcal{Z}}=\mathcal{L}, we have ϕb(𝒵,b)=b\phi_{b}^{*}(\mathcal{L}_{\mathcal{Z},b})=\mathcal{L}_{b}. Thus, by Remark 2.4, we have h0(𝒳b,bm)=h0(𝒵b,𝒵,bm)h^{0}(\mathcal{X}_{b},\mathcal{L}_{b}^{\otimes m})=h^{0}(\mathcal{Z}_{b},\mathcal{L}_{\mathcal{Z},b}^{\otimes m}) for all m1m\geq 1. Then, the latter is locally constant from the assumptions on kk, as by the vanishing of the higher cohomologies we have h0(𝒵b,𝒵,bm)=χ(𝒵b,𝒵,bm)h^{0}(\mathcal{Z}_{b},\mathcal{L}_{\mathcal{Z},b}^{\otimes m})=\chi(\mathcal{Z}_{b},\mathcal{L}_{\mathcal{Z},b}^{\otimes m}), and the Euler characteristic of 𝒵,bm\mathcal{L}_{\mathcal{Z},b}^{\otimes m} is independent of bBb\in B. 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 mm, the morphism Bbh0(𝒳b,|𝒳bm)B\ni b\mapsto h^{0}(\mathcal{X}_{b},\mathcal{L}|_{\mathcal{X}_{b}}^{\otimes m}) is constant and the algebra mH0(𝒳b,|𝒳bm)\bigoplus_{m}H^{0}(\mathcal{X}_{b},\mathcal{L}|_{\mathcal{X}_{b}}^{\otimes m}) is finitely generated.

Observe that the claim holds for every pUp\in U. Indeed, the following diagram commutes:

𝒳b\textstyle{\mathcal{X}_{b}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}πb\scriptstyle{\pi_{b}}𝒳U\textstyle{\mathcal{X}_{U}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒴b\textstyle{\mathcal{Y}_{b}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒴U\textstyle{\mathcal{Y}_{U}}

which from point (2) guarantees πb(𝒴,b)=b\pi_{b}^{*}(\mathcal{L}_{\mathcal{Y},b})=\mathcal{L}_{b}. Then, in this case, we can conclude as at the end of Step 2. Furthermore, the finite generation follows from the fact that b\mathcal{L}_{b} is very ample and 𝒴b\mathcal{Y}_{b} is the Proj of its associated graded ring.

Now, we treat the case when bUb\not\in U. Consider a smooth curve CC with a map CBC\to B. Assume that the generic point of CC maps into UU. Notice that any point bBb\in B is contained in the image of such a curve. Then, for any point sCs\in C, we have the following diagram:

𝒳s\textstyle{\mathcal{X}_{s}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒳C\textstyle{\mathcal{X}_{C}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒳\textstyle{\mathcal{X}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}{s}\textstyle{\{s\}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}C\textstyle{C\ignorespaces\ignorespaces\ignorespaces\ignorespaces}B.\textstyle{B.}

Since both squares are fibered squares, the big rectangle is a fibered square. In particular, since we want to prove that Bbh0(𝒳b,|𝒳bm)B\ni b\mapsto h^{0}(\mathcal{X}_{b},\mathcal{L}|_{\mathcal{X}_{b}}^{\otimes m}) is constant, and since we know it is constant as long as bUb\in U, it suffices to check that for every such CC the functions Csh0(𝒳s,|𝒳sm)C\ni s\mapsto h^{0}(\mathcal{X}_{s},\mathcal{L}|_{\mathcal{X}_{s}}^{\otimes m}) are constant. Now, this follows by Step 2. Similarly, the finite generation of mH0(𝒳s,|𝒳sm)\bigoplus_{m}H^{0}(\mathcal{X}_{s},\mathcal{L}|_{\mathcal{X}_{s}}^{\otimes m}) follows from Step 2.

Step 4: In this step, we construct the model 𝒴\mathcal{Y} and the morphism π\pi.

By Step 3 and cohomology and base change (see [Mum74]), the sheaves p(m)p_{*}(\mathcal{L}^{\otimes m}) commute with the restriction to points. In particular, the algebra 𝒜mp(m)\mathcal{A}\coloneqq\bigoplus_{m}p_{*}(\mathcal{L}^{\otimes m}) is finitely generated since it is finitely generated when restricted to every point bBb\in B. So we can consider 𝒴ProjB(𝒜)\mathcal{Y}\coloneqq\operatorname{Proj}_{B}(\mathcal{A}).

As observed in Step 3, the pluri-sections of |𝒳b\mathcal{L}|_{\mathcal{X}b} are deformation invariant. As |𝒳b\mathcal{L}|_{\mathcal{X}b} is semi-ample for every bBb\in B by our assumptions, it follows that \mathcal{L} is relatively semi-ample. In particular, we have a morphism 𝒳𝒴\mathcal{X}\rightarrow\mathcal{Y}. Furthermore, this implies the equality 𝒴b=Proj(mH0(𝒳b,bm))\mathcal{Y}_{b}=\mathrm{Proj}(\bigoplus_{m}H^{0}(\mathcal{X}_{b},\mathcal{L}_{b}^{\otimes m})) for every bBb\in B.

As already discussed, this construction commutes with base change. In particular, for every bBb\in B, for checking properties of 𝒴b\mathcal{Y}_{b} we can consider a smooth curve CBC\to B which sends the generic point to UU and the special one to bb, and first pull back 𝒴\mathcal{Y} to CC and then restrict it to bb:

𝒴b\textstyle{\mathcal{Y}_{b}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒴C\textstyle{\mathcal{Y}_{C}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒴\textstyle{\mathcal{Y}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}{b}\textstyle{\{b\}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}C\textstyle{C\ignorespaces\ignorespaces\ignorespaces\ignorespaces}B.\textstyle{B.}

The advantage is that now we can apply the results of Step 2 to 𝒴C\mathcal{Y}_{C}.

Since we have UπU𝒴,U\mathcal{L}_{U}\cong\pi_{U}^{*}\mathcal{L}_{\mathcal{Y},U} and 𝒴,U\mathcal{L}_{\mathcal{Y},U} is relatively ample over UU, it follows that 𝒴×BU=𝒴U\mathcal{Y}\times_{B}U=\mathcal{Y}_{U} and that π\pi extends πU\pi_{U}. Furthermore, since BB is reduced, the construction commutes with base change, and each fiber is reduced, it follows that 𝒴\mathcal{Y} is reduced.

Step 5: In this step, we show that π\pi is a contraction and we construct the divisor 𝒟𝒴\mathcal{D}_{\mathcal{Y}}.

We first prove that π\pi is a contraction. We denote by V𝒴V\subseteq\mathcal{Y} the locus where π1(V)V\pi^{-1}(V)\to V is an isomorphism. Then it follows from Lemma 2.13:

(\ast\ast) for every fiber 𝒴b\mathcal{Y}_{b}, the complement of VbV𝒴bV_{b}\coloneqq V\cap\mathcal{Y}_{b} has codimension at least 2 in 𝒴b\mathcal{Y}_{b} and it does not contain any irreducible component of the conductor of 𝒴b\mathcal{Y}_{b}.

Consider now the inclusion i:π1(V)𝒳i\colon\pi^{-1}(V)\to\mathcal{X}, which induces the injective map 0𝒪𝒳i𝒪π1(V)0\to\mathcal{O}_{\mathcal{X}}\to i_{*}\mathcal{O}_{\pi^{-1}(V)}. We can push this sequence forward via π\pi and we obtain 0π𝒪𝒳πi𝒪π1(V)0\to\pi_{*}\mathcal{O}_{\mathcal{X}}\to\pi_{*}i_{*}\mathcal{O}_{\pi^{-1}(V)}. But πi:π1(V)𝒴\pi\circ i\colon\pi^{-1}(V)\to\mathcal{Y} is the inclusion j:VYj\colon V\hookrightarrow Y, so πi𝒪π1(V)=jj𝒪𝒴\pi_{*}i_{*}\mathcal{O}_{\pi^{-1}(V)}=j_{*}j^{*}\mathcal{O}_{\mathcal{Y}} is reflexive from [HK04]*Corollary 3.7, and it is isomorphic to 𝒪𝒴\mathcal{O}_{\mathcal{Y}} from [HK04]*Proposition 3.6.2. In particular, this gives an injective map π𝒪𝒳𝒪𝒴\pi_{*}\mathcal{O}_{\mathcal{X}}\to\mathcal{O}_{\mathcal{Y}}. One can check that this is the inverse of the canonical morphism 𝒪𝒴π𝒪𝒳\mathcal{O}_{\mathcal{Y}}\to\pi_{*}\mathcal{O}_{\mathcal{X}}, so in particular the latter is an isomorphism.

Consider the ideal sheaf \mathcal{I} of 𝒟\mathcal{D}, and consider the inclusion

0𝒪𝒳.0\to\mathcal{I}\to\mathcal{O}_{\mathcal{X}}.

Then, as π\pi is a contraction, if we push it forward via π\pi, we get

0π𝒪𝒴.0\to\pi_{*}\mathcal{I}\to\mathcal{O}_{\mathcal{Y}}.

In particular, π\pi_{*}\mathcal{I} is an ideal sheaf on 𝒴\mathcal{Y}. We denote by 𝒮𝒴\mathcal{S}\subseteq\mathcal{Y} the closed subscheme with ideal sheaf π\pi_{*}\mathcal{I}.

A priori, 𝒮\mathcal{S} may not be pure dimensional, however, consider the intersection between π1(V)\pi^{-1}(V) and the locus in 𝒳\mathcal{X} where 𝒟\mathcal{D} is \mathbb{Q}-Cartier: we denote this locus with WW. Observe that, since π1(V)V\pi^{-1}(V)\to V is an isomorphism, we can identify WW with a subset of 𝒴\mathcal{Y}. Moreover, since 𝒟\mathcal{D} is Cartier on codimension one point of 𝒳\mathcal{X} and VV is a big open subset, the locus WW contains all the codimension one points of 𝒴b\mathcal{Y}_{b} for every bb. Then, we consider 𝒮W\mathcal{S}\cap W, and we define 𝒟𝒴\mathcal{D}_{\mathcal{Y}} to be the closure of 𝒮W\mathcal{S}\cap W in 𝒴\mathcal{Y}.

Step 6: In this step, we show that q:(𝒴,1r𝒟𝒴)Bq\colon(\mathcal{Y},\frac{1}{r}\mathcal{D}_{\mathcal{Y}})\rightarrow B is a well defined family of pairs.

By construction, it is immediate that 𝒟𝒴\mathcal{D}_{\mathcal{Y}} 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 𝒳\mathcal{X}.

Thus, we just need to check that q:𝒴Bq\colon\mathcal{Y}\to B is flat. By pulling back 𝒴B\mathcal{Y}\to B along a smooth curve through UU, it follows from Step 1 that all the fibers of qq are reduced and equidimensional. Thus, qq 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 q:(𝒴,1r𝒟𝒴)Bq\colon(\mathcal{Y},\frac{1}{r}\mathcal{D}_{\mathcal{Y}})\rightarrow B is a stable family of pairs and π(K𝒴/B+1r𝒟𝒴)=K𝒳/B+1r𝒟\pi^{*}(K_{\mathcal{Y}/B}+\frac{1}{r}\mathcal{D}_{\mathcal{Y}})=K_{\mathcal{X}/B}+\frac{1}{r}\mathcal{D}.

Since q:(𝒴,1r𝒟𝒴)Bq\colon(\mathcal{Y},\frac{1}{r}\mathcal{D}_{\mathcal{Y}})\rightarrow B 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 K𝒴/B+1r𝒟𝒴K_{\mathcal{Y}/B}+\frac{1}{r}\mathcal{D}_{\mathcal{Y}} is \mathbb{Q}-Cartier. In particular, q:(𝒴,1r𝒟𝒴)Bq\colon(\mathcal{Y},\frac{1}{r}\mathcal{D}_{\mathcal{Y}})\rightarrow B is a stable family of pairs.

By construction, we have that K𝒴/B+1r𝒟𝒴=π(K𝒳/B+1r𝒟)K_{\mathcal{Y}/B}+\frac{1}{r}\mathcal{D}_{\mathcal{Y}}=\pi_{*}(K_{\mathcal{X}/B}+\frac{1}{r}\mathcal{D}). Furthermore, we have that the equality π(K𝒴/B+1r𝒟𝒴)=K𝒳/B+1r𝒟\pi^{*}(K_{\mathcal{Y}/B}+\frac{1}{r}\mathcal{D}_{\mathcal{Y}})=K_{\mathcal{X}/B}+\frac{1}{r}\mathcal{D} holds over UU. Then, by construction, all the exceptional divisors of π\pi dominate BB, as they are contained in the support of 𝒟\mathcal{D}. Thus, as K𝒴/B+1r𝒟𝒴K_{\mathcal{Y}/B}+\frac{1}{r}\mathcal{D}_{\mathcal{Y}} is \mathbb{Q}-Cartier and so π(K𝒴/B+1r𝒟𝒴)\pi^{*}(K_{\mathcal{Y}/B}+\frac{1}{r}\mathcal{D}_{\mathcal{Y}}) is well defined, it follows that π(K𝒴/B+1r𝒟𝒴)=K𝒳/B+1r𝒟\pi^{*}(K_{\mathcal{Y}/B}+\frac{1}{r}\mathcal{D}_{\mathcal{Y}})=K_{\mathcal{X}/B}+\frac{1}{r}\mathcal{D}. ∎

Lemma 6.4.

With the notation and assumptions of Theorem 6.2, assume that BB is an affine curve and that there is a stable pair (Y,DY)(Y,D_{Y}) such that (𝒴,1r𝒟𝒴)(Y×B,DY×B)(\mathcal{Y},\frac{1}{r}\mathcal{D}_{\mathcal{Y}})\cong(Y\times B,D_{Y}\times B). Then there are finitely many isomorphism classes of \mathbb{Q}-stable pairs in the fibers of (𝒳;𝒟)B(\mathcal{X};\mathcal{D})\to B.

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 BB, we can assume that 𝒳\mathcal{X} (and therefore also 𝒴\mathcal{Y}) is normal. This is achieved in [ABIP]*Lemma 6.5. In particular, (Y×B,DY×B)(Y\times B,D_{Y}\times B) is the canonical model of (𝒳,1r𝒟)(\mathcal{X},\frac{1}{r}\mathcal{D}).

Step 2: We observe that the divisors contracted by 𝒳Y×B\mathcal{X}\to Y\times B have negative discrepancies and can be extracted by a log resolution of the form Z×BY×BZ\times B\to Y\times B, where ZYZ\to Y is a log resolution. This is achieved in [ABIP]*Proposition 6.13.

Step 3: To conclude, we observe that all the fibers of (𝒳;𝒟)B(\mathcal{X};\mathcal{D})\to B are isomorphic in codimension 2.

But two stable pairs (X1,D1)(X_{1},D_{1}) and (X2,D2)(X_{2},D_{2}) which are isomorphic in codimension 2 must be isomorphic. Indeed, if UU is the open subset where they agree, and L1L_{1} (resp. L2L_{2}) is the log-canonical divisor KX1+D1K_{X_{1}}+D_{1} (resp. KX2+D2K_{X_{2}}+D_{2}) then

H0(X1,L1[m])=H0(U,L1[m])=H0(U,L2[m])=H0(X2,L2[m]).H^{0}(X_{1},L_{1}^{[m]})=H^{0}(U,L_{1}^{[m]})=H^{0}(U,L_{2}^{[m]})=H^{0}(X_{2},L_{2}^{[m]}).

Therefore X1X_{1} and X2X_{2} are Proj\mathrm{Proj} 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 \mathbb{Q}-pairs. As a consequence, we also deduce the projectivity of the coarse moduli space of n,p,I\mathscr{F}_{n,p,I}.

We first construct a suitable polarization on the base of families of \mathbb{Q}-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 n,p,I\mathscr{F}_{n,p,I} 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 p:(𝒳;𝒟)Bp\colon(\mathcal{X};\mathcal{D})\to B the \mathbb{Q}-stable morphism, and by q:(𝒴,1r𝒟𝒴)Bq\colon(\mathcal{Y},\frac{1}{r}\mathcal{D}_{\mathcal{Y}})\to B the resulting stable family of Theorem 6.2. Assume this \mathbb{Q}-stable morphism is a family of \mathbb{Q}-stable pairs, and let m0m_{0} be the smallest positive integer such that m01r𝒟m_{0}\frac{1}{r}\mathcal{D} is Cartier. Then, there is k0k_{0} divisible by m0m_{0} such that, for every k=k0k=\ell k_{0} positive multiple of k0k_{0}, the sheaf 𝒴𝒪𝒴(k(K𝒴/B+1r𝒟𝒴))\mathcal{L}_{\mathcal{Y}}\coloneqq\mathcal{O}_{\mathcal{Y}}(k(K_{\mathcal{Y}/B}+\frac{1}{r}\mathcal{D}_{\mathcal{Y}})) satisfies the following properties:

  1. (a)

    𝒴\mathcal{L}_{\mathcal{Y}} is Cartier;

  2. (b)

    Riq(𝒴j)=0R^{i}q_{*}(\mathcal{L}_{\mathcal{Y}}^{\otimes j})=0 for every i>0i>0 and j>0j>0;

  3. (c)

    𝒴\mathcal{L}_{\mathcal{Y}} gives an embedding 𝒴(q𝒴)\mathcal{Y}\hookrightarrow\mathbb{P}(q_{*}\mathcal{L}_{\mathcal{Y}});

  4. (d)

    𝒪𝒳(k(K𝒳/B+1r𝒟𝒴)m0r𝒟)\mathcal{O}_{\mathcal{X}}(k(K_{\mathcal{X}/B}+\frac{1}{r}\mathcal{D}_{\mathcal{Y}})-\frac{m_{0}}{r}\mathcal{D}) is relatively ample; and

  5. (e)

    for every bBb\in B, there is a section sH0(𝒴b,𝒴,b)s\in H^{0}(\mathcal{Y}_{b},\mathcal{L}_{\mathcal{Y},b}) such that V(s)V(s) has codimension 1 in each irreducible component of 𝒴b\mathcal{Y}_{b} and the scheme-theoretic image of m0𝒟𝒴m_{0}\mathcal{D}\to\mathcal{Y} restricted to 𝒴b\mathcal{Y}_{b} is contained in V(s)V(s).

Remark 7.2.

Observe that the definition of m01r𝒟m_{0}\frac{1}{r}\mathcal{D} is given in Notation 2.29. In our case, we can still define m01r𝒟m_{0}\frac{1}{r}\mathcal{D} even if (𝒳;𝒟)B(\mathcal{X};\mathcal{D})\to B was a \mathbb{Q}-stable morphism instead, since the base BB is reduced, and Lemma 7.1 would go through verbatim. Since we will use Lemma 7.1 only in the case in which (𝒳;𝒟)B(\mathcal{X};\mathcal{D})\to B is a family of \mathbb{Q}-stable pairs, for simplicity we stick with the family case.

Proof.

By construction, K𝒴/B+1r𝒟𝒴K_{\mathcal{Y}/B}+\frac{1}{r}\mathcal{D}_{\mathcal{Y}} is relatively ample. Thus, for a sufficiently divisible k0k_{0}, properties (a) to (d) are satisfied for k=k0k=k_{0}.

Then, we can achieve (a) to (d) since if 𝒪𝒴(k0(K𝒴/B+1r𝒟𝒴))\mathcal{O}_{\mathcal{Y}}(k_{0}(K_{\mathcal{Y}/B}+\frac{1}{r}\mathcal{D}_{\mathcal{Y}})) satisfies any condition between (a) to (e), then for every m>0m>0 also 𝒪𝒴(mk0(K𝒴/B+1r𝒟𝒴))\mathcal{O}_{\mathcal{Y}}(mk_{0}(K_{\mathcal{Y}/B}+\frac{1}{r}\mathcal{D}_{\mathcal{Y}})) satisfies the same condition. We only need to check that up to choosing k0k_{0} divisible enough, also (e) holds.

Let 𝒢𝒪𝒴(k0(K𝒴/B+1r𝒟𝒴))\mathcal{G}\coloneqq\mathcal{O}_{\mathcal{Y}}(k_{0}(K_{\mathcal{Y}/B}+\frac{1}{r}\mathcal{D}_{\mathcal{Y}})) for a k0k_{0} which satisfies (a)-(d), and we need to show that 𝒢m\mathcal{G}^{\otimes m} satisfies (a)-(e) for some m0m\gg 0. First observe that from cohomology and base change and point (b), for every bBb\in B we have q(𝒢m)k(b)=H0(𝒴b,𝒢|𝒴bm)q_{*}(\mathcal{G}^{\otimes m})\otimes k(b)=H^{0}(\mathcal{Y}_{b},\mathcal{G}_{|\mathcal{Y}_{b}}^{\otimes m}).

Let 𝒵𝒴\mathcal{Z}\subseteq\mathcal{Y} be the scheme-theoretic image of π|m0r𝒟\pi_{|\frac{m_{0}}{r}\mathcal{D}}. Up to replacing BB with a locally closed stratification BBB^{\prime}\to B, we can assume that 𝒵𝒵×BBB\mathcal{Z}^{\prime}\coloneqq\mathcal{Z}\times_{B}B^{\prime}\to B^{\prime} is flat, and q𝒢q_{*}\mathcal{G} is free. Then, we define 𝒴𝒴×BB\mathcal{Y}^{\prime}\coloneqq\mathcal{Y}\times_{B}B^{\prime}, the first projection π1:𝒴𝒴\pi_{1}\colon\mathcal{Y}^{\prime}\to\mathcal{Y}, the second projection q:𝒴Bq^{\prime}\colon\mathcal{Y}^{\prime}\to B^{\prime}, and 𝒢π1𝒢\mathcal{G}^{\prime}\coloneqq\pi_{1}^{*}\mathcal{G}. Then, consider the exact sequence

0𝒪𝒴𝒪𝒵0.0\to\mathcal{I}\to\mathcal{O}_{\mathcal{Y}^{\prime}}\to\mathcal{O}_{\mathcal{Z}^{\prime}}\to 0.

We twist the exact sequence above by 𝒢m\mathcal{G}^{\prime\otimes m} and we obtain

0𝒢m𝒢m𝒪𝒵𝒢m0.0\to\mathcal{I}\otimes\mathcal{G}^{\prime\otimes m}\to\mathcal{G}^{\prime\otimes m}\to\mathcal{O}_{\mathcal{Z}^{\prime}}\otimes\mathcal{G}^{\prime\otimes m}\to 0.

Since 𝒵B\mathcal{Z}^{\prime}\to B^{\prime} is flat, also \mathcal{I} is flat over BB^{\prime}. Therefore since 𝒢\mathcal{G}^{\prime} is relatively ample, up to choosing mm big enough, the following is an exact sequence on BB^{\prime}:

0q(𝒢m)q(𝒢m)q(𝒪𝒵𝒢m)0.0\to q^{\prime}_{*}(\mathcal{I}\otimes\mathcal{G}^{\prime\otimes m})\to q^{\prime}_{*}(\mathcal{G}^{\prime\otimes m})\to q^{\prime}_{*}(\mathcal{O}_{\mathcal{Z}^{\prime}}\otimes\mathcal{G}^{\prime\otimes m})\to 0.

By [ACH11]*Corollary 4.5 (iii), there is n0n_{0} such that for every nn0n\geq n_{0} and any a1a\geq 1, the multiplication map

Syma(q𝒢)q(𝒢n)q(𝒢n+a)\operatorname{Sym}^{a}(q^{\prime}_{*}\mathcal{G}^{\prime})\otimes q^{\prime}_{*}(\mathcal{I}\otimes\mathcal{G}^{\prime\otimes n})\to q^{\prime}_{*}(\mathcal{I}\otimes\mathcal{G}^{\prime\otimes n+a})

is surjective. Then for every bBb\in B^{\prime}, also

Syma(q𝒢)q(𝒢n)k(b)(Syma(q𝒢)k(b))k(b)(q(𝒢n)k(b))q(𝒢n+a)k(b)\operatorname{Sym}^{a}(q^{\prime}_{*}\mathcal{G}^{\prime})\otimes q^{\prime}_{*}(\mathcal{I}\otimes\mathcal{G}^{\prime\otimes n})\otimes k(b)\cong(\operatorname{Sym}^{a}(q^{\prime}_{*}\mathcal{G}^{\prime})\otimes k(b))\otimes_{k(b)}(q^{\prime}_{*}(\mathcal{I}\otimes\mathcal{G}^{\prime\otimes n})\otimes k(b))\to q^{\prime}_{*}(\mathcal{I}\otimes\mathcal{G}^{\prime\otimes n+a})\otimes k(b)

is surjective. From [ACH11]*Corollary 4.5 (iv) (or cohomology and base change), for nn0n\geq n_{0} and a,s1a,s\geq 1, we have:

Syma(q𝒢)k(b)H0(𝒪N(a)),\operatorname{Sym}^{a}(q^{\prime}_{*}\mathcal{G}^{\prime})\otimes k(b)\cong H^{0}(\mathcal{O}_{\mathbb{P}^{N}}(a)),
q(𝒢n)k(b)H0(𝒴b,(𝒢n)|𝒴b)=H0(𝒴b,|𝒴b(n)),   and q^{\prime}_{*}(\mathcal{I}\otimes\mathcal{G}^{\prime\otimes n})\otimes k(b)\cong H^{0}(\mathcal{Y}_{b},(\mathcal{I}\otimes\mathcal{G}^{\prime\otimes n})_{|\mathcal{Y}_{b}})=H^{0}(\mathcal{Y}_{b},\mathcal{I}_{|\mathcal{Y}_{b}}(n)),\text{ }\text{ }\text{ and }
q(𝒢n+s)k(b)H0(𝒴b,(𝒢n+s)|𝒴b)=H0(𝒴b,|𝒴b(n+s)).q^{\prime}_{*}(\mathcal{I}\otimes\mathcal{G}^{\prime\otimes n+s})\otimes k(b)\cong H^{0}(\mathcal{Y}_{b},(\mathcal{I}\otimes\mathcal{G}^{\prime\otimes n+s})_{|\mathcal{Y}_{b}})=H^{0}(\mathcal{Y}_{b},\mathcal{I}_{|\mathcal{Y}_{b}}(n+s)).

In other terms, for every bBb\in B the ideal 𝒴b\mathcal{I}_{\mathcal{Y}_{b}} is generated in degree n0n_{0}. Then for every bBb\in B there is a section sH0(𝒴b,𝒴,b)=H0(𝒴b,𝒢𝒴,bn0)s\in H^{0}(\mathcal{Y}_{b},\mathcal{L}_{\mathcal{Y},b})=H^{0}(\mathcal{Y}_{b},\mathcal{G}^{\otimes n_{0}}_{\mathcal{Y},b}) (and therefore also a section smH0(𝒴b,𝒴,bm)s^{\otimes m}\in H^{0}(\mathcal{Y}_{b},\mathcal{L}_{\mathcal{Y},b}^{\otimes m}) if mn0m\geq n_{0}) such that V(s)V(s) does not contain any irreducible component of 𝒴b\mathcal{Y}_{b} and the scheme-theoretic image of m0r𝒟𝒴\frac{m_{0}}{r}\mathcal{D}\to\mathcal{Y} restricted to 𝒴b\mathcal{Y}_{b} is contained in V(s)V(s).

Therefore, for the locally closed subset BB^{\prime}, we may choose k0=k0n0k_{0}^{\prime}=k_{0}n_{0}, where k0k_{0} was chosen at the beginning of the proof to satisfy (a)-(d) and thus define 𝒢\mathcal{G}. Then, by noetherianity, there are finitely many locally closed subsets BB^{\prime} in the decomposition of BB, and we may thus choose k0′′k_{0}^{\prime\prime} as the least common multiple of the k0k_{0}^{\prime} defined on each BB^{\prime}. ∎

Corollary 7.3.

With the notation and assumptions of Lemma 7.1, for every bBb\in B and every a>0a>0, there is a global section tt of |𝒳ba\mathcal{L}_{|\mathcal{X}_{b}}^{\otimes a} that is not zero on the generic points of 𝒳b\mathcal{X}_{b} but its maps to 0 via the restriction map H0(𝒳b,|𝒳ba)H0(am0r𝒟b,|am0r𝒟ba)H^{0}(\mathcal{X}_{b},\mathcal{L}_{|\mathcal{X}_{b}}^{\otimes a})\to H^{0}(\frac{am_{0}}{r}\mathcal{D}_{b},\mathcal{L}^{\otimes a}_{|\frac{am_{0}}{r}\mathcal{D}_{b}}).

Proof.

From Lemma 7.1, for every bBb\in B, there is a section ss that does not vanish on the generic points of 𝒴b\mathcal{Y}_{b} but vanishes on the restriction to 𝒴b\mathcal{Y}_{b} of the scheme theoretic image of m0r𝒟\frac{m_{0}}{r}\mathcal{D}. From point (b) of Lemma 7.1, q𝒴q_{*}\mathcal{L}_{\mathcal{Y}} is a vector bundle on BB, so there is an open subset bUBb\in U\subseteq B such that (q𝒴)|U(q_{*}\mathcal{L}_{\mathcal{Y}})_{|U} is free. Then, we can extend the section ss to a global section sH0(U,q(𝒴)|U)=H0(U,(qU)(𝒴,U))s^{\prime}\in H^{0}(U,q_{*}(\mathcal{L}_{\mathcal{Y}})_{|U})=H^{0}(U,(q_{U})_{*}(\mathcal{L}_{\mathcal{Y},U})). But π\pi is a contraction, so πU:𝒳U𝒴U\pi_{U}\colon\mathcal{X}_{U}\to\mathcal{Y}_{U} is still a contraction, and πU𝒴,U=U\pi_{U}^{*}\mathcal{L}_{\mathcal{Y},U}=\mathcal{L}_{U}. In particular, consider the section tH0(𝒳U,U)t\in H^{0}(\mathcal{X}_{U},\mathcal{L}_{U}) that is the pull-back of ss^{\prime}. Then, tt does not vanish on the generic points of 𝒳b\mathcal{X}_{b}, as by Lemma 2.13 those are in bijection with the generic points of 𝒴b\mathcal{Y}_{b}. Moreover, it vanishes along m0r𝒟b\frac{m_{0}}{r}\mathcal{D}_{b}: this is a local computation. Indeed, if we replace 𝒴\mathcal{Y} and 𝒳\mathcal{X} with appropriate open subsets Spec(A)\operatorname{Spec}(A), Spec(B)\operatorname{Spec}(B); and ff is the generator for the ideal sheaf of m0r𝒟\frac{m_{0}}{r}\mathcal{D}, we have the following commutative diagram:

Bk(b)/f1\textstyle{B\otimes k(b)/f\otimes 1}B/f\textstyle{B/f\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Bk(b)\textstyle{B\otimes k(b)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}tB\textstyle{t\in B\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ak(b)\textstyle{A\otimes k(b)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}sA.\textstyle{s\in A.\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}

Since the image of ss vanishes in Bk(b)/f1B\otimes k(b)/f\otimes 1, so does the image of tt.

Therefore for every a>0a>0, the section taH0(𝒳U,Ua)t^{\otimes a}\in H^{0}(\mathcal{X}_{U},\mathcal{L}_{U}^{\otimes a}) will vanish along am0r𝒟b\frac{am_{0}}{r}\mathcal{D}_{b}. In other terms, if we look at the exact sequence

0H0(𝒳b,|𝒳ba𝒟b[am0r])H0(𝒳b,|𝒳ba)𝜓H0(am0r𝒟b,|am0r𝒟ba),0\to H^{0}(\mathcal{X}_{b},\mathcal{L}_{|\mathcal{X}_{b}}^{\otimes a}\otimes\mathcal{I}_{\mathcal{D}_{b}}^{[\frac{am_{0}}{r}]})\to H^{0}(\mathcal{X}_{b},\mathcal{L}_{|\mathcal{X}_{b}}^{\otimes a})\xrightarrow{\psi}H^{0}(\frac{am_{0}}{r}\mathcal{D}_{b},\mathcal{L}^{\otimes a}_{|\frac{am_{0}}{r}\mathcal{D}_{b}}),

we have an element of H0(𝒳b,|𝒳ba)H^{0}(\mathcal{X}_{b},\mathcal{L}_{|\mathcal{X}_{b}}^{\otimes a}) (namely, the restriction of tat^{\otimes a} to 𝒳b\mathcal{X}_{b}) that does not vanish along the generic points of 𝒳b\mathcal{X}_{b} but maps to 0 via ψ\psi. ∎

Lemma 7.4.

With the notation and assumptions of Lemma 7.1, we will denote by p:(𝒳;𝒟)Bp\colon(\mathcal{X};\mathcal{D})\to B the family of \mathbb{Q}-stable pairs, and by q:(𝒴,1r𝒟𝒴)Bq\colon(\mathcal{Y},\frac{1}{r}\mathcal{D}_{\mathcal{Y}})\to B the resulting stable family of Theorem 6.2. Let kk be a multiple of k0k_{0}, where k0k_{0} is as in Lemma 7.1. Also, set 𝒪𝒳(k(K𝒳/B+1r𝒟))\mathcal{L}\coloneqq\mathcal{O}_{\mathcal{X}}(k(K_{\mathcal{X}/B}+\frac{1}{r}\mathcal{D})).

Then, there is an a0a_{0} such that, for every aa0a\geq a_{0}, the sheaf p(|am0r𝒟a)p_{*}(\mathcal{L}_{|\frac{am_{0}}{r}\mathcal{D}}^{\otimes a}) is a vector bundle, and fits in an exact sequence as follows:

0p(𝒪𝒳(ka(K𝒳/B+1r𝒟)am0r𝒟))p(a)p(|am0r𝒟a)0.0\to p_{*}(\mathcal{O}_{\mathcal{X}}(ka(K_{\mathcal{X}/B}+\frac{1}{r}\mathcal{D})-\frac{am_{0}}{r}\mathcal{D}))\to p_{*}(\mathcal{L}^{\otimes a})\to p_{*}(\mathcal{L}_{|\frac{am_{0}}{r}\mathcal{D}}^{\otimes a})\to 0.
Proof.

First, from relative Serre’s vanishing and Lemma 7.1 (b) and (d), there is an integer a0a_{0} such that for every bBb\in B and aa0a\geq a_{0}, we have

  • Hi(𝒪𝒳b(ak(K𝒳b+1r𝒟b)am0r𝒟b)=0H^{i}(\mathcal{O}_{\mathcal{X}_{b}}(ak(K_{\mathcal{X}_{b}}+\frac{1}{r}\mathcal{D}_{b})-\frac{am_{0}}{r}\mathcal{D}_{b})=0 for i>0i>0; and

  • Hi(𝒪𝒴b(ak(K𝒴b+1r𝒟𝒴,b)))=0H^{i}(\mathcal{O}_{\mathcal{Y}_{b}}(ak(K_{\mathcal{Y}_{b}}+\frac{1}{r}\mathcal{D}_{\mathcal{Y},b})))=0 for i>0i>0.

Fix a positive integer a>a0a>a_{0}. We begin by considering the following exact sequence, where 𝒥\mathcal{J} is the locally free ideal sheaf 𝒪𝒳(m0r𝒟)\mathcal{O}_{\mathcal{X}}\left(-\frac{m_{0}}{r}\mathcal{D}\right):

0a𝒥aa|am0r𝒟a0.0\to\mathcal{L}^{\otimes a}\otimes\mathcal{J}^{\otimes a}\to\mathcal{L}^{\otimes a}\to\mathcal{L}_{|\frac{am_{0}}{r}\mathcal{D}}^{\otimes a}\to 0.

We push it forward via pp, and from the first bullet point and from cohomology and base change, R1p(a𝒥a)=0R^{1}p_{*}(\mathcal{L}^{\otimes a}\otimes\mathcal{J}^{\otimes a})=0, so we have:

(8) 0p(a𝒥a)p(a)p(|m0ar𝒟a)0.0\to p_{*}(\mathcal{L}^{\otimes a}\otimes\mathcal{J}^{\otimes a})\to p_{*}(\mathcal{L}^{\otimes a})\to p_{*}(\mathcal{L}_{|\frac{m_{0}a}{r}\mathcal{D}}^{\otimes a})\to 0.

Recall that there is a contraction π:𝒳𝒴\pi\colon\mathcal{X}\to\mathcal{Y} such that, if we denote 𝒴𝒪𝒴(k(K𝒴/B+1r𝒟𝒴))\mathcal{L}_{\mathcal{Y}}\coloneqq\mathcal{O}_{\mathcal{Y}}(k(K_{\mathcal{Y}/B}+\frac{1}{r}\mathcal{D}_{\mathcal{Y}})), then π(𝒴)=\pi^{*}(\mathcal{L}_{\mathcal{Y}})=\mathcal{L}. In particular

(9) p(a)=qπ(a)=q(𝒴a).p_{*}(\mathcal{L}^{\otimes a})=q_{*}\pi_{*}(\mathcal{L}^{\otimes a})=q_{*}(\mathcal{L}_{\mathcal{Y}}^{\otimes a}).

Now, pick bBb\in B. From the first bullet point and from cohomology and base-change, we have

(10) p(a𝒥a)k(b)H0(𝒳b,(a𝒥a)|𝒳b)H0(𝒳b,𝒪𝒳b(ak(K𝒳b+1r𝒟b)am0r𝒟b))p_{*}(\mathcal{L}^{\otimes a}\otimes\mathcal{J}^{\otimes a})\otimes k(b)\cong H^{0}(\mathcal{X}_{b},(\mathcal{L}^{\otimes a}\otimes\mathcal{J}^{\otimes a})_{|\mathcal{X}_{b}})\cong H^{0}(\mathcal{X}_{b},\mathcal{O}_{\mathcal{X}_{b}}(ak(K_{\mathcal{X}_{b}}+\frac{1}{r}\mathcal{D}_{b})-\frac{am_{0}}{r}\mathcal{D}_{b}))

and from the second bullet point, equation (9) and the fact that πb:𝒳b𝒴b\pi_{b}\colon\mathcal{X}_{b}\to\mathcal{Y}_{b} is a contraction,

(11) p(a)k(b)H0(𝒴b,𝒴,ba)H0(𝒴b,𝒪𝒴b(ak(K𝒴b+1r𝒟b)))H0(𝒳b,|𝒳ba).p_{*}(\mathcal{L}^{\otimes a})\otimes k(b)\cong H^{0}(\mathcal{Y}_{b},\mathcal{L}^{\otimes a}_{\mathcal{Y},b})\cong H^{0}(\mathcal{Y}_{b},\mathcal{O}_{\mathcal{Y}_{b}}(ak(K_{\mathcal{Y}_{b}}+\frac{1}{r}\mathcal{D}_{b})))\cong H^{0}(\mathcal{X}_{b},\mathcal{L}^{\otimes a}_{|\mathcal{X}_{b}}).

Now, we tensor the sequence (8) by k(b)k(b), and we obtain

p(a𝒥a)k(b)p(a)k(b)p(|m0ar𝒟a)k(b)0.p_{*}(\mathcal{L}^{\otimes a}\otimes\mathcal{J}^{\otimes a})\otimes k(b)\to p_{*}(\mathcal{L}^{\otimes a})\otimes k(b)\to p_{*}(\mathcal{L}_{|\frac{m_{0}a}{r}\mathcal{D}}^{\otimes a})\otimes k(b)\to 0.

From the natural adjunction between pull-back and push-forward, we have the following commutative diagram:

p(a𝒥a)k(b)\textstyle{p_{*}(\mathcal{L}^{\otimes a}\otimes\mathcal{J}^{\otimes a})\otimes k(b)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces} from (10)\scriptstyle{\cong\text{ from \eqref{eq_13}}}ϕ\scriptstyle{\phi}p(a)k(b)\textstyle{p_{*}(\mathcal{L}^{\otimes a})\otimes k(b)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces} from (11)\scriptstyle{\cong\text{ from \eqref{eq_14}}}p(|m0ar𝒟a)k(b)\textstyle{p_{*}(\mathcal{L}_{|\frac{m_{0}a}{r}\mathcal{D}}^{\otimes a})\otimes k(b)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H0(𝒳b,(a𝒥a)|𝒳b)\textstyle{H^{0}(\mathcal{X}_{b},(\mathcal{L}^{\otimes a}\otimes\mathcal{J}^{\otimes a})_{|\mathcal{X}_{b}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H0(𝒳b,|𝒳ba)\textstyle{H^{0}(\mathcal{X}_{b},\mathcal{L}^{\otimes a}_{|\mathcal{X}_{b}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H0(m0ar𝒟,(a)|m0ar𝒟b)\textstyle{H^{0}(\frac{m_{0}a}{r}\mathcal{D},(\mathcal{L}^{\otimes a})_{|\frac{m_{0}a}{r}\mathcal{D}_{b}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}()\scriptstyle{(\ast)}0\textstyle{0}

Observe that the exactness of the map ()(\ast) follows from the first bullet point. Then, from diagram chasing, also ϕ\phi is injective. In particular,

dimk(b)(p(|m0ar𝒟a)k(b))=dimk(b)(H0(𝒳b,|𝒳ba))dimk(b)(H0(𝒳b,(a𝒥a)|𝒳b)).\dim_{k(b)}(p_{*}(\mathcal{L}_{|\frac{m_{0}a}{r}\mathcal{D}}^{\otimes a})\otimes k(b))=\dim_{k(b)}(H^{0}(\mathcal{X}_{b},\mathcal{L}^{\otimes a}_{|\mathcal{X}_{b}}))-\dim_{k(b)}(H^{0}(\mathcal{X}_{b},(\mathcal{L}^{\otimes a}\otimes\mathcal{J}^{\otimes a})_{|\mathcal{X}_{b}})).

However the right-hand side does not depend on bBb\in B. Therefore also the left-hand side does not depend on bBb\in B, so p(|m0ar𝒟a)p_{*}(\mathcal{L}_{|\frac{m_{0}a}{r}\mathcal{D}}^{\otimes a}) is a vector bundle. ∎

Corollary 7.5.

With the notation and assumptions of Lemma 7.4, there is an isomorphism

detp(a)det(p(𝒪𝒳(ak(K𝒳/B+1r𝒟)am0r𝒟)))det(p(|m0ar𝒟a)).\det p_{*}(\mathcal{L}^{\otimes a})\cong\det(p_{*}(\mathcal{O}_{\mathcal{X}}(ak(K_{\mathcal{X}/B}+\frac{1}{r}\mathcal{D})-\frac{am_{0}}{r}\mathcal{D})))\otimes\det(p_{*}(\mathcal{L}_{|\frac{m_{0}a}{r}\mathcal{D}}^{\otimes a})).
Theorem 7.6.

With the notation and assumptions of Lemma 7.4, further assume that the map Bn,p,IB\to\mathscr{F}_{n,p,I} given by the family of \mathbb{Q}-stable pairs (𝒳;𝒟)B(\mathcal{X};\mathcal{D})\to B is finite. Then, for aa divisible enough, the line bundle detp(a)\det p_{*}(\mathcal{L}^{\otimes a}) is ample.

Proof.

We plan on using Kollár’s Ampleness Lemma, see [Kol90]. By Corollary 7.5, it suffices to show that detp(a)det(p(𝒪𝒳(ak(K𝒳/B+1r(1m0k)𝒟))))det(p(|m0ar𝒟a))\det p_{*}(\mathcal{L}^{\otimes a})\otimes\det(p_{*}(\mathcal{O}_{\mathcal{X}}(ak(K_{\mathcal{X}/B}+\frac{1}{r}(1-\frac{m_{0}}{k})\mathcal{D}))))\otimes\det(p_{*}(\mathcal{L}_{|\frac{m_{0}a}{r}\mathcal{D}}^{\otimes a})) is ample.

Consider the vector bundles

𝒬ap(a)p(𝒪𝒳(ak(K𝒳/B+1r(1m0k)𝒟))))p(|m0ar𝒟a)    and \mathcal{Q}_{a}\coloneqq p_{*}(\mathcal{L}^{\otimes a})\oplus p_{*}(\mathcal{O}_{\mathcal{X}}(ak(K_{\mathcal{X}/B}+\frac{1}{r}(1-\frac{m_{0}}{k})\mathcal{D}))))\oplus p_{*}(\mathcal{L}_{|\frac{m_{0}a}{r}\mathcal{D}}^{\otimes a})\text{ }\text{ }\text{ }\text{ and }
𝒲b,cSymb(p(c))Symb(𝒪𝒳(ck(K𝒳/B+1r(1m0k)𝒟))Symb(p(c))\mathcal{W}_{b,c}\coloneqq\operatorname{Sym}^{b}(p_{*}(\mathcal{L}^{\otimes c}))\oplus\operatorname{Sym}^{b}(\mathcal{O}_{\mathcal{X}}(ck(K_{\mathcal{X}/B}+\frac{1}{r}(1-\frac{m_{0}}{k})\mathcal{D}))\oplus\operatorname{Sym}^{b}(p_{*}(\mathcal{L}^{\otimes c}))

for appropriate choices of aa, bb, and cc.

When a=bca=bc, there is a morphism 𝒲b,c𝒬bc\mathcal{W}_{b,c}\to\mathcal{Q}_{bc} given by the sum of the multiplication morphisms

Symb(p(c))𝛼p(bc),Symb(p(𝒪𝒳(ck(K𝒳/B+1r(1m0k)𝒟)))𝛽p(𝒪𝒳(bck(K𝒳/B+1r(1m0k)𝒟)),\operatorname{Sym}^{b}(p_{*}(\mathcal{L}^{\otimes c}))\xrightarrow{\alpha}p_{*}(\mathcal{L}^{\otimes bc}),\quad\operatorname{Sym}^{b}(p_{*}(\mathcal{O}_{\mathcal{X}}(ck(K_{\mathcal{X}/B}+\frac{1}{r}(1-\frac{m_{0}}{k})\mathcal{D})))\xrightarrow{\beta}p_{*}(\mathcal{O}_{\mathcal{X}}(bck(K_{\mathcal{X}/B}+\frac{1}{r}(1-\frac{m_{0}}{k})\mathcal{D})),

and the composition of α\alpha with the surjection in the exact sequence of Lemma 7.4, denoted by

Symb(p(c))𝛾p(|m0bcr𝒟bc).\operatorname{Sym}^{b}(p_{*}(\mathcal{L}^{\otimes c}))\xrightarrow{\gamma}p_{*}(\mathcal{L}_{|\frac{m_{0}bc}{r}\mathcal{D}}^{\otimes bc}).

First, observe that for cc divisible enough, the vector bundles p(c)p_{*}(\mathcal{L}^{\otimes c}) and p(𝒪𝒳(ck(K𝒳/B+1r(1m0k)𝒟)))p_{*}(\mathcal{O}_{\mathcal{X}}(ck(K_{\mathcal{X}/B}+\frac{1}{r}(1-\frac{m_{0}}{k})\mathcal{D}))) are nef. Indeed, from the assumptions of Lemma 7.4, the formation of p(c)p_{*}(\mathcal{L}^{\otimes c}) and p(𝒪𝒳(ck(K𝒳/B+1r(1m0k)𝒟)))p_{*}(\mathcal{O}_{\mathcal{X}}(ck(K_{\mathcal{X}/B}+\frac{1}{r}(1-\frac{m_{0}}{k})\mathcal{D}))) commutes with base change, so we can assume that the base BB is a smooth curve. Then the statement follows from [Fuj18]*Theorem 1.11. Therefore, their symmetric powers and sum are nef, so 𝒲b,c\mathcal{W}_{b,c} is nef for cc divisible enough.

Since 𝒴\mathcal{L}_{\mathcal{Y}} is relatively ample and q(𝒴m)p(m)q_{*}(\mathcal{L}_{\mathcal{Y}}^{\otimes m})\cong p_{*}(\mathcal{L}^{\otimes m}), there is a b0b_{0} such that, for every bb0b\geq b_{0}, the map α\alpha is surjective. For the same reason, if we choose cc large enough so that ck(K𝒳/B+1r(1m0k)𝒟)ck(K_{\mathcal{X}/B}+\frac{1}{r}(1-\frac{m_{0}}{k})\mathcal{D}) is relatively very ample, also β\beta is surjective for b0=b0(c)b_{0}=b_{0}(c) sufficiently large (see [ACH11]*Corollary 4.5). Then, also γ\gamma is surjective since it is the composition of surjective morphisms. Therefore we have a surjection Φ:𝒲b,c𝒬bc\Phi\colon\mathcal{W}_{b,c}\to\mathcal{Q}_{bc}. We denote by GG the structure group of 𝒲b,c\mathcal{W}_{b,c}. This gives a map of sets

Ψ:|B||[Gr(w,q)/G]|,\Psi\colon|B|\to|[\operatorname{Gr}(w,q)/G]|,

where ww (resp. qq) is the rank of 𝒲b,c\mathcal{W}_{b,c} (resp. 𝒬bc\mathcal{Q}_{bc}), and where for a stack 𝒵\mathcal{Z} we denote by |𝒵||\mathcal{Z}| its associated topological space. If we show that Ψ\Psi has finite fibers, then the theorem follows from [Kol90]*Ampleness Lemma, 3.9.

Consider two points x1x_{1}, x2x_{2} that map to the same point via Ψ\Psi. Over the point x1x_{1} we have the surjection Φx1:𝒲b,ck(x1)𝒬bck(x1)\Phi_{x_{1}}\colon\mathcal{W}_{b,c}\otimes k(x_{1})\to\mathcal{Q}_{bc}\otimes k(x_{1}), and similarly we have one denoted by Φx2\Phi_{x_{2}} over x2x_{2}. Choose two isomorphisms

τ1:H0(N,𝒪N(1))p(c)k(x1), and \tau_{1}\colon H^{0}(\mathbb{P}^{N},\mathcal{O}_{\mathbb{P}^{N}}(1))\to p_{*}(\mathcal{L}^{\otimes c})\otimes k(x_{1}),\text{ and }
τ1:H0(M,𝒪M(1))p(𝒪𝒳(ck(K𝒳/B+1r(1Mk)𝒟)))k(x1).\tau_{1}^{\prime}\colon H^{0}(\mathbb{P}^{M},\mathcal{O}_{\mathbb{P}^{M}}(1))\to p_{*}(\mathcal{O}_{\mathcal{X}}(ck(K_{\mathcal{X}/B}+\frac{1}{r}(1-\frac{M}{k})\mathcal{D})))\otimes k(x_{1}).

Similarly we define τ2\tau_{2} and τ2\tau_{2}^{\prime} for the same isomorphisms over x2x_{2}. This gives an isomorphism

H0(N,𝒪N(b))H0(M,𝒪M(b))H0(N,𝒪N(b))Symb(τ1)Symb(τ1)Symb(τ1)𝒲b,ck(x1).H^{0}(\mathbb{P}^{N},\mathcal{O}_{\mathbb{P}^{N}}(b))\oplus H^{0}(\mathbb{P}^{M},\mathcal{O}_{\mathbb{P}^{M}}(b))\oplus H^{0}(\mathbb{P}^{N},\mathcal{O}_{\mathbb{P}^{N}}(b))\xrightarrow{\operatorname{Sym}^{b}(\tau_{1})\oplus\operatorname{Sym}^{b}(\tau_{1}^{\prime})\oplus\operatorname{Sym}^{b}(\tau_{1})}\mathcal{W}_{b,c}\otimes k(x_{1}).

Since x1x_{1} and x2x_{2} map to the same point via Ψ\Psi, using the identifications above, there is an element gGg\in G such that gKer(Φx1)=Ker(Φx2)g\operatorname{Ker}(\Phi_{x_{1}})=\operatorname{Ker}(\Phi_{x_{2}}).

In particular, we can choose a basis for 𝒲b,ck(x1)\mathcal{W}_{b,c}\otimes k(x_{1}) 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 𝒲b,c\mathcal{W}_{b,c} is a direct sum of vector bundles, in this basis the linear transformation gg is block diagonal:

g=[A000B000A].g=\begin{bmatrix}A&0&0\\ 0&B&0\\ 0&0&A\end{bmatrix}.

In particular, gg will send Ker(αx1)\operatorname{Ker}(\alpha_{x_{1}}) (resp. Ker(γx1)\operatorname{Ker}(\gamma_{x_{1}})) to Ker(αx2)\operatorname{Ker}(\alpha_{x_{2}}) (resp. Ker(γx2)\operatorname{Ker}(\gamma_{x_{2}})). But the kernel of αx1\alpha_{x_{1}} corresponds to the symmetric functions of degree bb that vanish on a subvariety of n1\mathbb{P}^{n_{1}} isomorphic to 𝒴x1\mathcal{Y}_{x_{1}}, which we will still denote by 𝒴x1\mathcal{Y}_{x_{1}}. From [ACH11]*Corollary 4.5, up to choosing b0b_{0} big enough, this kernel generates a graded ideal that corresponds to a subvariety isomorphic to 𝒴x1\mathcal{Y}_{x_{1}}. The same conclusion holds for Ker(αx2)\operatorname{Ker}(\alpha_{x_{2}}), since there is a linear transformation (given by a block of the matrix AA) that induces an isomorphism Ker(αx1)Ker(αx2)\operatorname{Ker}(\alpha_{x_{1}})\to\operatorname{Ker}(\alpha_{x_{2}}). So, in particular, this linear transformation induces a map of projective spaces, that gives an isomorphism h:𝒴x1𝒴x2h\colon\mathcal{Y}_{x_{1}}\to\mathcal{Y}_{x_{2}}.

By Corollary 7.3, Ker(γx1)\operatorname{Ker}(\gamma_{x_{1}}) contains a function that does not vanish along the generic points of the irreducible components of 𝒴x1\mathcal{Y}_{x_{1}} but, if pulled back via πx1:𝒳x1𝒴x1\pi_{x_{1}}\colon\mathcal{X}_{x_{1}}\to\mathcal{Y}_{x_{1}}, it vanishes along m0ar𝒟x1\frac{m_{0}a}{r}\mathcal{D}_{x_{1}}. Therefore, the zero locus of the polynomials in Ker(γx1)\operatorname{Ker}(\gamma_{x_{1}}) has codimension 1 in 𝒴x1\mathcal{Y}_{x_{1}} and the union of its irreducible components of codimension 1, which we denote by Γx1\Gamma_{x_{1}}, contains Supp(1r𝒟𝒴,x1)\operatorname{Supp}(\frac{1}{r}\mathcal{D}_{\mathcal{Y},x_{1}}). Since Γx1\Gamma_{x_{1}} has finitely many irreducible components of codimension one and the coefficients of 𝒟𝒴,x1\mathcal{D}_{\mathcal{Y},x_{1}} are in the finite set rIrI, the divisor 𝒟𝒴,x1\mathcal{D}_{\mathcal{Y},x_{1}} is determined up to finitely many possible choices of prime divisors and coefficients.

A similar conclusion holds by replacing x1x_{1} with x2x_{2}. Since in the description of gg the block at position (1,1) is the same as the block at position (3,3), we also know that h(Γx1)=Γx2h(\Gamma_{x_{1}})=\Gamma_{x_{2}}.

Therefore the fiber of Ψ(x1)\Psi(x_{1}) corresponds to stable pairs (Y,D)(Y,D) in our moduli problem where Y𝒴x1Y\cong\mathcal{Y}_{x_{1}}, Supp(D)Supp(Γx1)\operatorname{Supp}(D)\subseteq\operatorname{Supp}(\Gamma_{x_{1}}). But there are finitely many such subvarieties, and since Bn,p,IB\to\mathscr{F}_{n,p,I} is finite by our assumptions, the fiber Ψ(x1)\Psi(x_{1}) has to be finite. ∎

Corollary 7.7.

Consider a proper DM stack 𝒦n,v,I\mathcal{K}_{n,v,I} which satisfies the following two conditions:

  1. (1)

    for every normal scheme SS, the data of a morphism f:S𝒦n,v,If\colon S\to\mathcal{K}_{n,v,I} is equivalent to a stable family of pairs q:(𝒴,𝒟)Bq\colon(\mathcal{Y},\mathcal{D})\to B with fibers of dimension nn, volume vv and coefficients in II; and

  2. (2)

    there is m0m_{0}\in\mathbb{N} such that, for every kk\in\mathbb{N}, there is a line bundle k\mathcal{L}_{k} on 𝒦n,v,I\mathcal{K}_{n,v,I} such that, for every morphism ff as above, fkdet(q(ω𝒴/B[km0](km0𝒟)))f^{*}\mathcal{L}_{k}\cong\det(q_{*}(\omega_{\mathcal{Y}/B}^{[km_{0}]}(km_{0}\mathcal{D}))).

Then, for m0m_{0} divisible enough, k\mathcal{L}_{k} descends to an ample line bundle on the coarse moduli space of 𝒦n,v,I\mathcal{K}_{n,v,I} for every k1k\geq 1. In particular, the coarse moduli space of 𝒦n,v,I\mathcal{K}_{n,v,I} is projective.

Remark 7.8.

As a concrete example of Corollary 7.7, one can consider 𝒦n,v,I\mathcal{K}_{n,v,I} to be any moduli space of KSB-stable pairs 𝒦\mathscr{K}, such that, if we denote by π:(𝒴,𝒟)𝒦\pi\colon(\mathscr{Y},\mathscr{D})\to\mathscr{K} its universal family, then ω𝒴/𝒦[m0](m0𝒟)\omega_{\mathscr{Y}/\mathscr{K}}^{[m_{0}]}(m_{0}\mathscr{D}) is π\pi-ample for m0m_{0} divisible enough. Indeed, in this case, by cohomology and base change πω𝒴/𝒦[m0](m0𝒟)\pi_{*}\omega_{\mathscr{Y}/\mathscr{K}}^{[m_{0}]}(m_{0}\mathscr{D}) is a vector bundle for m0m_{0} 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 Kn,v,IK_{n,v,I} the coarse moduli space of 𝒦n,v,I\mathcal{K}_{n,v,I}, and let \mathcal{M} be an irreducible component of it.

Step 1. There is a \mathbb{Q}-stable morphism (𝒳;𝒟)B(\mathcal{X};\mathcal{D})\to B which satisfies the following conditions:

  1. (1)

    BB is normal and projective,

  2. (2)

    The map Bn,p,IB\to\mathscr{F}_{n,p,I} is finite,

  3. (3)

    There is a dense open subset UBU\subseteq B and a stable family (𝒴U,𝒟𝒴,U)(\mathcal{Y}_{U},\mathcal{D}_{\mathcal{Y},U}) which satisfies the assumptions of Theorem 6.2, and

  4. (4)

    The family q:(𝒴U,𝒟𝒴,U)Uq\colon(\mathcal{Y}_{U},\mathcal{D}_{\mathcal{Y},U})\to U induces a map U𝒦n,v,IKn,v,IU\to\mathcal{K}_{n,v,I}\to K_{n,v,I} which dominates XX.

Consider the generic point of \mathcal{M}, and consider the stable pair (𝒵η,𝒟𝒵,η)(\mathcal{Z}_{\eta},\mathcal{D}_{\mathcal{Z},\eta}) it corresponds to. By Lemma 2.16, there is the spectrum of a field Spec(𝔽)\operatorname{Spec}(\mathbb{F}) and a \mathbb{Q}-stable family over it f:(𝒳η;𝒟η)Spec(𝔽)f\colon(\mathcal{X}_{\eta};\mathcal{D}_{\eta})\to\operatorname{Spec}(\mathbb{F}) such that the canonical model of (𝒳η;𝒟η)(\mathcal{X}_{\eta};\mathcal{D}_{\eta}) is (𝒵η,𝒟𝒵,η)(\mathcal{Z}_{\eta},\mathcal{D}_{\mathcal{Z},\eta}). The family ff induces a morphism Spec(𝔽)n,p,I\operatorname{Spec}(\mathbb{F})\to\mathscr{F}_{n,p,I}, and we can take \mathcal{F} to be the closure of its image. This is still a proper DM stack, it admits a finite and surjective cover by a scheme BB^{\prime}\to\mathcal{F} with BB^{\prime} a normal and projective scheme using [LMB18]*Theorem 16.6, Chow’s lemma and potentially normalizing. Up to replacing 𝔽\mathbb{F} with a finite cover of it, we can lift Spec(𝔽)\operatorname{Spec}(\mathbb{F})\to\mathcal{F} to Spec(𝔽)B\operatorname{Spec}(\mathbb{F})\to B^{\prime}, and consider BB an irreducible component of BB containing the image of Spec(𝔽)B\operatorname{Spec}(\mathbb{F})\to B^{\prime}.

The morphism Bn,p,IB\to\mathscr{F}_{n,p,I} induces a \mathbb{Q}-stable family (𝒳;𝒟)B(\mathcal{X};\mathcal{D})\to B, and by construction its generic fiber admits an canonical model. Then such an canonical model can be spread out: there is an open subset UBU\subseteq B and a family (𝒴U,𝒟𝒴,U)(\mathcal{Y}_{U},\mathcal{D}_{\mathcal{Y},U}) as in Theorem 6.2, and the generic fiber (𝒴U,𝒟𝒴,U)U(\mathcal{Y}_{U},\mathcal{D}_{\mathcal{Y},U})\to U is isomorphic to (𝒵η,𝒟𝒵,η)(\mathcal{Z}_{\eta},\mathcal{D}_{\mathcal{Z},\eta}). Therefore the image of the corresponding map U𝒦n,v,IKn,v,IU\to\mathcal{K}_{n,v,I}\to K_{n,v,I} contains the generic point of \mathcal{M}, and since both UU and \mathcal{M} are irreducible, UXU\to X is dominant.

Step 2. The map U𝒦n,v,IU\to\mathcal{K}_{n,v,I} extends to a finite map Φ:B𝒦n,v,I\Phi\colon B\to\mathcal{K}_{n,v,I}.

The extension follows from Theorem 6.2. To prove that Φ\Phi is finite, we can use Lemma 6.4. Indeed, if it was not finite, there would be a curve contained in a fiber of Φ\Phi. But then, up to replacing CC with an open subset of it, there would be a stable pair (Y,DY)(Y,D_{Y}) and a \mathbb{Q}-stable family g:(𝒳;𝒟)Cg\colon(\mathcal{X};\mathcal{D})\to C as in Lemma 6.4. Therefore there would be finitely many isomorphism types of \mathbb{Q}-stable pairs in the fibers of gg: this contradicts point (2) above.

End of the proof. From Theorem 7.6, up to replacing m0m_{0} with a multiple (which depends only on \mathcal{M}), the line bundle Φk\Phi^{*}\mathcal{L}_{k} is ample. But a multiple of k\mathcal{L}_{k} descends to a line bundle on Kn,v,IK_{n,v,I}. In other terms, there is a line bundle GG on Kn,v,IK_{n,v,I} whose pull-back is kc\mathcal{L}_{k}^{\otimes c} for a certain c>0c>0. Therefore G|G|_{\mathcal{M}} is ample since it is ample once pulled-back via the finite map BB\to\mathcal{M}. But if a line bundle is ample when restricted to each irreducible component, it is ample. ∎

Corollary 7.9.

The stack n,p,I\mathscr{F}_{n,p,I} admits a projective coarse moduli space.

Proof.

We will denote by Fn,p,IF_{n,p,I} the coarse moduli space of n,p,I\mathscr{F}_{n,p,I}, which exists from Keel–Mori’s theorem, and with Fn,p,InF_{n,p,I}^{n} the one of the normalization n,p,In\mathscr{F}_{n,p,I}^{n} of n,p,I\mathscr{F}_{n,p,I}.

Consider ϵ0\epsilon_{0} such that KX+(1ϵ0)DK_{X}+(1-\epsilon_{0})D is ample for every \mathbb{Q}-pair (X;D)(X;D) parametrized by n,p,I\mathscr{F}_{n,p,I}. Recall that such an ϵ0\epsilon_{0} exists from Lemma 2.15 (or from boundedness). In particular, if we denote with (𝒳;𝒟)n,p,I(\mathscr{X};\mathscr{D})\to\mathscr{F}_{n,p,I} the universal family, for mm divisibie enough the formation of 𝒢det(p(𝒪𝒳(m(K𝒳/n,p,I+1ϵr𝒟))))\mathscr{G}\coloneqq\det(p_{*}(\mathcal{O}_{\mathscr{X}}(m(K_{\mathscr{X}/\mathscr{F}_{n,p,I}}+\frac{1-\epsilon}{r}\mathscr{D})))) commutes with base change and an high enough power of 𝒢\mathscr{G} descends to a line bundle GG on Fn,p,IF_{n,p,I}. Therefore to prove that GG is ample, it suffices to replace Fn,p,IF_{n,p,I} (resp. GG) with Fn,p,InF_{n,p,I}^{n} (resp. the pull-back GnG^{n} of GG via Fn,p,InFn,p,IF_{n,p,I}^{n}\to F_{n,p,I}).

Consider a proper DM stack 𝒦n,p(1ϵ0),I\mathcal{K}_{n,p(1-\epsilon_{0}),I} as in Corollary 7.7. When DD has coefficient 1ϵ01-\epsilon_{0} the pairs parametrized by n,p,In\mathscr{F}_{n,p,I}^{n} are stable of volume p(1ϵ0)p(1-\epsilon_{0}), so we have a map n,p,In𝒦n,p(1ϵ0),I\mathscr{F}_{n,p,I}^{n}\to\mathcal{K}_{n,p(1-\epsilon_{0}),I} which is finite, as different points of n,p,I\mathscr{F}_{n,p,I} parametrize different \mathbb{Q}-pairs, and the normalization is a finite morphism. From Corollary 7.7, the formation of k\mathcal{L}_{k} commutes with base change, so k\mathcal{L}_{k} pulls back to GnG^{n}. Then GnG^{n} is ample as it is the pull-back of an ample line bundle via a finite morphism.∎

References