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

Isotriviality of smooth families of varieties of general type

Chuanhao Wei Department of Mathematics, Stony Brook University, 100 Nicolls Rd, Stony Brook, NY 11794, USA [email protected]  and  Lei Wu Department of Mathematics, University of Utah, 155 S 1400 E, Salt Lake City, UT 84112, USA [email protected]
Abstract.

In this paper, we proved that a log smooth family of log general type klt pairs with a special (in the sense of Campana) quasi-projective base is isotrivial. As a consequence, we proved the generalized Kebekus-Kovács conjecture [WW19, Conjecture 1.1], for smooth families of general type varieties as well as log smooth families of log canonical pairs of log general type, assuming the existence of relative good minimal models.

1. introduction

We start with fV:(U,DU)Vf_{V}\mathrel{\mathop{\mathchar 58\relax}}(U,D^{U})\to V, a log smooth family of projective log pairs of log general type, over a smooth quasi-projective variety VV, with the coefficients of DUD^{U} are in [0,1)[0,1). Here, log smoothness of fVf_{V} means that fVf_{V} and each stratum of DUD^{U} is smooth over VV (see §3.1). Due to [BCHM], fVf_{V} has a relative canonical model, denoted by fVcf_{V}^{c}. The relative canonical model is a stable family by the invariance of log plurigenera [HMX18, Theorem 4.2], and hence it induces a moduli map μ:V𝔐\mu\mathrel{\mathop{\mathchar 58\relax}}V\to\mathfrak{M}, where 𝔐\mathfrak{M} is the coarse moduli space of the corresponding stable family (with a fixed set of coefficients) as in [KP16, §6]. Note that 𝔐\mathfrak{M} is a projective variety, [KP16, Theorem 1.1].

Following [KP16, Definition 6.16], we denote the variation of fVf_{V} by

Var(fV)=Var(fcV):=dim(μ(V)).\textup{Var}(f_{V})=\textup{Var}(f^{c}_{V})\mathrel{\mathop{\mathchar 58\relax}}=\dim(\mu(V)).

We say that fVf_{V} is of maximal variation if Var(fV)=dimV\textup{Var}(f_{V})=\dim V and that fVf_{V} is isotrivial if Var(fV)=0\textup{Var}(f_{V})=0. Since fibers of fVf_{V} are of log general type, the above definition is compatible with the variation defined in [Vie83] and [Kaw85]. We write the log Kodaira dimension of VV by κ¯(V)=κ(Y,EY)\bar{\kappa}(V)=\kappa(Y,E^{Y}), where the latter is defined to be the Iitaka dimension of the log canonical sheaf ωY(EY)\omega_{Y}(E^{Y}) and YY is a projective compactification of VV with EY=YUE^{Y}=Y\setminus U a normal crossing divisor. We call (Y,EY)(Y,E^{Y}) a log smooth compactification of VV. Note that κ¯(V)\bar{\kappa}(V) does not depend on the choice of (Y,EY)(Y,E^{Y}) and it is a birational invariant, that is, κ¯(V)\bar{\kappa}(V) remains the same, if one replaces VV by a smooth birational model. In this paper, we prove

Theorem 1.0.1.

With notations above, we have:

  1. (1)

    if κ¯(V)>\bar{\kappa}(V)>-\infty, then κ¯(V)Var(fV),\bar{\kappa}(V)\geq\textup{Var}(f_{V}),

  2. (2)

    if κ¯(V)=\bar{\kappa}(V)=-\infty, then dimV>Var(fV)\dim V>\textup{Var}(f_{V}).

When DUD^{U} is empty, we particularly have:

Theorem 1.0.2.

Let fV:UVf_{V}\mathrel{\mathop{\mathchar 58\relax}}U\to V be a smooth family of projective varieties of general type. Then,

  1. (1)

    if κ¯(V)>\bar{\kappa}(V)>-\infty, then κ¯(V)Var(fV),\bar{\kappa}(V)\geq\textup{Var}(f_{V}),

  2. (2)

    if κ¯(V)=\bar{\kappa}(V)=-\infty, then dimV>Var(fV)\dim V>\textup{Var}(f_{V}).

When fVf_{V} is a smooth family of canonical polarized varieties, the above statement is conjectured by Kebekus and Kovács [KK08, Conjecture 1.6], which is a natural extension of Viehweg’s hyperbolicity conjecture [Vie01]. Meanwhile, Campana made a related conjecture for smooth families of canonical polarized varieties [Cam, Conjecture 12.19] (see also [JK11a, Conjecture 1.4]), the Isotriviality Conjecture, which implies the Kebekus-Kovács conjecture. The Kebekus-Kovács conjecture for smooth families of canonical polarized varieties is proved by Taji [Taj16] by proving the Isotriviality Conjecture of Campana.

In the special case that Var(fV)=dimV,\textup{Var}(f_{V})=\dim V, Campana’s isotriviality conjecture is also known as Viehweg’s hyperbolicity conjecture. For smooth families of canonically polarized varieties, it was proved by Campana and Păun [CP15]. Their result has been extended to the case for smooth families of general type varieties by Popa and Schnell using Hodge modules [PS17, Theorem A]. They [PS17, §4.3] implicitly gave an extension of the Kebekus-Kovács conjecture for families with geometric generic fibers admitting good minimal models (particularly for smooth families of general type varieties). They proved in loc. cit. a special case, with additionally assuming an abundance-type conjecture of Campana–Peternell. Based on the Hodge-module construction of Popa and Schnell, we further extended Viehweg’s hyperbolicity to smooth families of log general type pairs (see [WW19, Theorem A(1)]). We then extended the Kebekus-Kovács conjecture to the case for smooth families of log general type pairs (see [WW19, Conjecture 1.1]).

Theorem 1.0.1 gives a confirmative answer to the generalized Kebekus-Kovács conjecture. It is worth mentioning that the generalized Kebekus-Kovács conjecture even for smooth families of general type varieties was not known before.

There are three reasons why we consider Theorem 1.0.1 (the pair version) instead of Theorem 1.0.2 (the non-pair version):
(1) The statement of Theorem 1.0.2 is about the log Kodaira dimension of the base, where the log Kodaira dimension is defined by embedding the base into a pair.
(2) To prove Theorem 1.0.2, we first need to compactify fVf_{V}

U{U}(X,DX){(X,D^{X})}V{V}(Y,DY),{(Y,D^{Y}),}fV\scriptstyle{f_{V}}f\scriptstyle{f}

and then do stable reductions (see §4), where ff is a morphism of log smooth pairs.
(3) The proof of Theorem 1.0.2 needs to use Campana’s fibrations, but they are naturally defined as morphisms of pairs (see §2.1).

Remark 1.0.3.

Actually, we can consider the case that fVf_{V} is a log smooth family of log smooth pairs of log general type with the coefficients of DUD^{U} in [0,1][0,1], if we assume that fVf_{V} admits a relative good minimal model over a Zariski open subset V0VV_{0}\subseteq V. Then, we have the smooth family fϵV:(U,DUϵ:=(1ϵ)DU)Vf^{\epsilon}_{V}\mathrel{\mathop{\mathchar 58\relax}}(U,D^{U}_{\epsilon}\mathrel{\mathop{\mathchar 58\relax}}=(1-\epsilon)D^{U})\to V of log smooth pairs of log general type with the coefficients of DUϵD^{U}_{\epsilon} in [0,1)[0,1) and Var(fϵV)Var(fV)\textup{Var}(f^{\epsilon}_{V})\geq\textup{Var}(f_{V}) by [WW19, Lemma 3.1]. See [WW19, §3] for the definition of variation in the log canonical case. The inequality Var(fϵV)Var(fV)\textup{Var}(f^{\epsilon}_{V})\geq\textup{Var}(f_{V}) implies that Theorem 1.0.1 still holds for smooth families of log canonical pairs of log general type, assuming the existence of relative good minimal models.

In order to prove Theorem 1.0.1, we prove Campana’s Isotriviality Conjecture for log smooth families of log general type pairs, Theorem 1.0.4 below. We continue to use the notations in Theorem 1.0.1. We now fix a log smooth compactification (Y,EY)(Y,E^{Y}) of VV, such that the moduli map μ:V𝔐\mu\mathrel{\mathop{\mathchar 58\relax}}V\to\mathfrak{M} extends to μ¯:Y𝔐\bar{\mu}\mathrel{\mathop{\mathchar 58\relax}}Y\to\mathfrak{M}. Considering the Stein factorization of μ¯\bar{\mu}, we get a morphism s:YMs\mathrel{\mathop{\mathchar 58\relax}}Y\to M with MM a normal variety so that μ¯\bar{\mu} factors through ss and ss has generically connected fibers. In particular, s:(Y,EY)Ms\mathrel{\mathop{\mathchar 58\relax}}(Y,E^{Y})\to M is a fibration of (Y,EY)(Y,E^{Y}) over MM with dimM=Var(fV)\dim M=\textup{Var}(f_{V}) (see Definition 2.1.1). Campana defined the fiberations of general type and specialty in the category of geometric orbifolds; see Definition 2.1.3 for a simplified version in our setting.

Theorem 1.0.4.

With notations as above, the fibration ss is of general type. As a consequence, if VV is a special quasi-projective variety, then fVf_{V} is birationally isotrivial, i.e. Var(fV)=0\textup{Var}(f_{V})=0.

The first statement of Theorem 1.0.4 implies Theorem 1.0.1, thanks to the additivity of the log Kodaira dimension for fibrations of general type, [Ca11, Theorem 6.3].

When fVf_{V} is a smooth family of canonically polarized variety, the second statement of Theorem 1.0.4 is the Isotriviality Conjecture of Campana as we mentioned above. In the case that dimV3\dim V\leq 3, it is proved by Jabbusch and Kebekus [JK11a]. Their proof uses Campana’s theory of geometric orbifolds and the so-called Viehweg-Zuo sheaves constructed in [VZ02] as well as a refine result on Viehweg-Zuo sheaves in [JK11b], which roughly asserts that Viehweg-Zuo sheaves factor through the moduli. Based on the strategy of Jabbusch and Kebekus and the method of Campana and Păun in proving Viehweg’s hyperbolicity conjecture, Taji [Taj16] proved the isotriviality conjecture for smooth families of canonically polarized varieties in general.

By using Hodge modules, Popa and Schnell [PS17] constructed the Viehweg-Zuo sheaves for smooth families of general type varieties with maximal variation (or more precisely for smooth families admitting relative good minimal models). We further constructed Viehweg-Zuo sheaves for families of pairs with maximal variation in [WW19]. A key step to prove Theorem 1.0.4 is to construct the Viehweg-Zuo sheaves for families of (log) general type varieties (pairs) with arbitrary variation and prove that they factor through the “moduli” in the way analoguous to the result of Jabbusch and Kebekus.

Definition 1.0.5.

In regard to a log smooth compactification (Y,E)(Y,E) of VV and an extension μ¯:Y𝔐\bar{\mu}\mathrel{\mathop{\mathchar 58\relax}}Y\to\mathfrak{M} of the moduli map μ\mu, we define the subsheaf

(Y,EY)ΩY(logEY)\mathcal{B}_{(Y,E^{Y})}\subseteq\Omega_{Y}(\log E^{Y})

by the saturation of the image of the natural morphism μ¯Ω𝔐ΩY(logEY)\bar{\mu}^{*}\Omega_{\mathfrak{M}}\to\Omega_{Y}(\log E^{Y}) in ΩY(logEY)\Omega_{Y}(\log E^{Y}), where Ω𝔐\Omega_{\mathfrak{M}} is the sheaf of Kähler differentials (over \mathbb{C}) and ΩY(logEY)\Omega_{Y}(\log E^{Y}) the sheaf of log 1-forms with log poles along EYE^{Y}.

The above definition follows [JK11b, Notation 1.2]. However, they only considered the case when 𝔐\mathfrak{M} is the moduli space of canonically polarised manifolds in loc. cit. Note that, since μ¯\bar{\mu} factors through s:YMs\mathrel{\mathop{\mathchar 58\relax}}Y\to M, and M𝔐M\to\mathfrak{M} is quasi-finite, we have that (Y,EY)\mathcal{B}_{(Y,E^{Y})} is also the saturation of the image of sΩMΩY(logEY)s^{*}\Omega_{M}\to\Omega_{Y}(\log E^{Y}), and the rank of (Y,EY)\mathcal{B}_{(Y,E^{Y})} is equal to the dimension of MM, which is also the variation of fVf_{V}. Now, we state the main technical result of this paper.

Theorem 1.0.6.

Let fV:(U,DU)Vf_{V}\mathrel{\mathop{\mathchar 58\relax}}(U,D^{U})\to V be a log smooth family of projective log pairs of log general type, over a smooth quasi-projective variety VV, with the coefficients of DUD^{U} in [0,1)[0,1). If Var(fV)>0\textup{Var}(f_{V})>0, then after replacing VV by a further log resolution, we have a log smooth compactification (Y,EY)(Y,E^{Y}) of VV, such that there exists an invertible subsheaf 𝒜Sym[n](Y,EY)\mathcal{A}\subset\textup{Sym}^{[n]}\mathcal{B}_{(Y,E^{Y})}, for some positive integer nn, with κ(𝒜)Var(fV)\kappa(\mathcal{A})\geq\textup{Var}(f_{V}).

We call the invertible sheaf 𝒜\mathcal{A} the refined Viehweg-Zuo Sheaf of fVf_{V} on (Y,EY)(Y,E^{Y}). The proof of Theorem 1.0.6 is built upon the refinement of the stable reduction used in [WW19] (see Section LABEL:S:refined_VZ for details). Another key input of its proof is the use of Hodge modules. Roughly speaking, we essentially need Saito’s decomposition theorem for pure Hodge modules to compare Viehweg-Zuo sheaves before and after base-changes, see Theorem 3.5.1 in Section 3 for details.

Structure of the paper

In Section 2, we recall some of Campana’s definitions and results about orbifold fibrations, and concludes the proof of Theorem 1.0.4 using Theorem 1.0.6. In Section 3, we fixed the notations and show some useful results using Saito’s theory of Hodge modules. In Section 4, we prove results related to stable reductions, and using them to make the geometric constructions that are needed in Section LABEL:S:refined_VZ to construct the refined Viehweg-Zuo Sheaf in our setting.

Acknowledgement

We would like to thank Christian Schnell for useful discussions during the preparation of the paper. The first author also gets some inspiration from a workshop held in Shanghai Center for Mathematical Sciences.

2. Birationally equivalent fibrations

2.1. Birationally equivalent fibration in the sense of Campana

We now recall Campana’s birational equivalence of fibrations. We mainly restrict ourselves to the setting of Theorem 1.0.4 for our application while Campana works more generally in the category of geometric orbifolds in [Ca11]. See also [JK11a] for a more approachable introduction.

Definition 2.1.1.

We say that s:(Y,EY)Ms\mathrel{\mathop{\mathchar 58\relax}}(Y,E^{Y})\to M is a fibration of a log pair (Y,EY)(Y,E^{Y}) over MM, if ss is a dominate projective morphism with generically connected fibers, and WW a normal variety. For simplicity we always assume that YY is a smooth quasi-projective variety. Given two fibrations s:(Y,EY)Ms\mathrel{\mathop{\mathchar 58\relax}}(Y,E^{Y})\to M and s:(Y,EY)Ms^{\prime}\mathrel{\mathop{\mathchar 58\relax}}(Y^{\prime},E^{Y^{\prime}})\to M^{\prime}, we say that ss^{\prime} is dominant over ss, if we have the following commutative diagram

(Y,EY){(Y,E^{Y})}(Y,EY){(Y^{\prime},E^{Y^{\prime}})}M{M}M,{M^{\prime},}s\scriptstyle{s}s\scriptstyle{s^{\prime}}u\scriptstyle{u}v\scriptstyle{v}

with both uu and vv are birational, and uEY=EYu_{*}E^{Y^{\prime}}=E^{Y}. We say two fibrations ss and ss^{\prime} are birationally equivalent if they both can be dominated by a third fibration s:(Y,EY)Ms^{\prime\prime}\mathrel{\mathop{\mathchar 58\relax}}(Y^{\prime\prime},E^{Y^{\prime\prime}})\to M^{\prime\prime}.

Using the recipe in [Ca11, Definition 3.2] (see also [JK11a, Construction and Definition 5.3]), we obtain the 𝒞\mathcal{C}-base (M,Δs)(M,\Delta^{s}) associated to the fibration ss. We say that a fibration s:(Y,EY)(M,Δs)s\mathrel{\mathop{\mathchar 58\relax}}(Y,E^{Y})\to(M,\Delta^{s}) is a neat model if (Y,EY)(Y,E^{Y}) and (M,Δs)(M,\Delta^{s}) are log smooth, and for all of the divisors FF with codims(F)2\operatorname{codim}s(F)\geq 2, FEYF\subset E^{Y}, [JK11a, Assumption 5.4]. The name of neat model is adopted from [Taj16, Definition 4.1], which serves a similar purpose as “strictement nette et haute” model in [Ca11], in the general case. By definition, starting from an arbitrary fibration s:(Y,EY)Ms\mathrel{\mathop{\mathchar 58\relax}}(Y,E^{Y})\to M, the associated fibration s:(Y,EY)(M,Δs)s\mathrel{\mathop{\mathchar 58\relax}}(Y,E^{Y})\to(M,\Delta^{s}) over the associated 𝒞\mathcal{C}-base (M,Δs)(M,\Delta^{s}) is not always neat even when the pairs (Y,EY)(Y,E^{Y}) and (M,Δs)(M,\Delta^{s}) are log smooth. We recall the following result in [Taj16], which is essentially proved in [JK11a, Section 10].

Proposition 2.1.2.

[Taj16, Proposition 4.2] Every fibration s:(Y,EY)(M,Δs)s\mathrel{\mathop{\mathchar 58\relax}}(Y,E^{Y})\to(M,\Delta^{s}) is dominated by a neat model.

Definition 2.1.3.

([Ca11, Definition 4.10, 4.16 and 4.17]) Let s:(Y,EY)Ms\mathrel{\mathop{\mathchar 58\relax}}(Y,E^{Y})\to M be a fibration with (Y,EY)(Y,E^{Y}) log smooth and MM projective. Let s~:(Y~,EY~)(M~,Δ)\tilde{s}\mathrel{\mathop{\mathchar 58\relax}}(\tilde{Y},E^{\tilde{Y}})\to(\tilde{M},\Delta) be neat model dominant over ss, with (M~,Δ)(\tilde{M},\Delta) the induced 𝒞\mathcal{C}-base of s~\tilde{s}. We define the canonical dimension of ss by κ(M~,Δs~)\kappa(\tilde{M},\Delta^{\tilde{s}}). It does not depend on the choice of the neat model s~\tilde{s} dominant over ss (see [Ca11, Corollaire 4.11]). Then we define:

  1. (1)

    a fibration ss is of general type if its canonical dimension is the same as the dimension of the base,

  2. (2)

    (Y,EY)(Y,E^{Y}) is special if there exists no fibration s:(Y,EY)Ms\mathrel{\mathop{\mathchar 58\relax}}(Y,E^{Y})\to M of general type with dim(M)>0\dim(M)>0.

We also say the smooth quasi-projective variety V=YEYV=Y\setminus E^{Y} is special if (Y,EY)(Y,E^{Y}) is so. One can easily check that the specialty of VV does not depend on the choice of the compactification (Y,EY)(Y,E^{Y}) and that specialty is a birational invariant, that is, VV is special if and only if VV^{\prime} is so, where VV^{\prime} is a smooth quasi-projective variety properly birational to VV.

2.2. Using the refined Viehweg-Zuo sheaf to prove the main theorem

We use Theorem 1.0.6 to give a proof of Theorem 1.0.4. The proof follows the strategy in [JK11a], which is also used by Taji in proving the Isotriviality conjecture of Campana [Taj16, Theorem 1.5].

Proof of Theorem 1.0.4.

When fVf_{V} is birationally isotrivial, the first statement of Theorem 1.0.4 is obvious. We hence assume Var(fV)>0\textup{Var}(f_{V})>0. Using Proposition 2.1.2, we fix a neat model dominating ss, s~:(Y~,EY~)(M~,Δs~)\tilde{s}\mathrel{\mathop{\mathchar 58\relax}}(\tilde{Y},E^{\tilde{Y}})\to(\tilde{M},\Delta^{\tilde{s}}), . Let μ:(Y~,EY~)(Y,EY)\mu\mathrel{\mathop{\mathchar 58\relax}}(\tilde{Y},E^{\tilde{Y}})\to(Y,E^{Y}) be the induced birational morphism. Without loss of generality, we further assume that EY~μ1EYE^{\tilde{Y}}\supset\mu^{-1}E^{Y} by adding more components to EY~E^{\tilde{Y}} (by doing this the 𝒞\mathcal{C}-base (M~,Δ)(\tilde{M},\Delta) stays the same).

Recall that (Y,EY)\mathcal{B}_{(Y,E^{Y})} is defined to be the saturation of the image of the natural morphism s~:ΩMΩY(logEY)\tilde{s}^{*}\mathrel{\mathop{\mathchar 58\relax}}\Omega_{M}\to\Omega_{Y}(\log E^{Y}). We similarly define (Y~,EY~)\mathcal{B}_{(\tilde{Y},E^{\tilde{Y}})} to be the saturation of the image of the natural morphism s~:ΩM~ΩY~(logEY~)\tilde{s}^{*}\mathrel{\mathop{\mathchar 58\relax}}\Omega_{\tilde{M}}\to\Omega_{\tilde{Y}}(\log E^{\tilde{Y}}). By [JK11a, Proposition 3.3], we have that Sym[n](Y~,EY~)\textup{Sym}^{[n]}\mathcal{B}_{(\tilde{Y},E^{\tilde{Y}})} is also the saturation of the image of the composed natural morphisms

μSym[n](Y,EY)μSym[n]ΩY(logEY)Sym[n]ΩY~(logEY~).\mu^{*}\textup{Sym}^{[n]}\mathcal{B}_{(Y,E^{Y})}\to\mu^{*}\textup{Sym}^{[n]}\Omega_{Y}(\log E^{Y})\to\textup{Sym}^{[n]}\Omega_{\tilde{Y}}(\log E^{\tilde{Y}}).

Hence, the refined Viehweg-Zuo sheaf 𝒜Sym[n](Y,EY)\mathcal{A}\subset\textup{Sym}^{[n]}\mathcal{B}_{(Y,E^{Y})} on (Y,EY)(Y,E^{Y}) lifted to μ𝒜Sym[n](Y~,EY~)\mu^{*}\mathcal{A}\subset\textup{Sym}^{[n]}\mathcal{B}_{(\tilde{Y},E^{\tilde{Y}})}. Hence, to make notations simple, we can assume that s:(Y,EY)(M,Δs)s\mathrel{\mathop{\mathchar 58\relax}}(Y,E^{Y})\to(M,\Delta^{s}) itself is neat.

By definition, to prove the fibration s:(Y,EY)Ms\mathrel{\mathop{\mathchar 58\relax}}(Y,E^{Y})\to M is of general type, one only needs to show that (M,Δs)(M,\Delta^{s}) is of log general type. Since (Y,EY)\mathcal{B}_{(Y,E^{Y})} is saturated, Sym[n](Y,EY)Sym[n]ΩY(logEY)\textup{Sym}^{[n]}\mathcal{B}_{(Y,E^{Y})}\subseteq\textup{Sym}^{[n]}\Omega_{Y}(\log E^{Y}) is also a saturated subsheaf by [JK11a, Proposition 3.3]. Due to [JK11a, Proposition 5.7], see [JK11a, Notation 4.1] for the notations, we have a natural isomorphism

ι:Sym[n]𝒞ΩM(logΔs)sSym[n](Y,EY)\iota\mathrel{\mathop{\mathchar 58\relax}}\textup{Sym}^{[n]}_{\mathcal{C}}\Omega_{M}(\log\Delta^{s})\to s_{*}\textup{Sym}^{[n]}\mathcal{B}_{(Y,E^{Y})}

for all nn. Using Theorem 1.0.6, we have the refined Viehweg-Zuo sheaf 𝒜Sym[n](Y,EY)\mathcal{A}\subset\textup{Sym}^{[n]}\mathcal{B}_{(Y,E^{Y})}, and we can further assume that 𝒜\mathcal{A} is saturated. Let 𝒜M\mathcal{A}_{M} be the saturation of s𝒜s_{*}\mathcal{A} in Sym[n]𝒞ΩM(logΔs)\textup{Sym}^{[n]}_{\mathcal{C}}\Omega_{M}(\log\Delta^{s}), and in particular, it is a line bundle with κ𝒞(𝒜M)=κ(𝒜)=dimM\kappa_{\mathcal{C}}(\mathcal{A}_{M})=\kappa(\mathcal{A})=\dim M, by [JK11a, Proposition 5.7, Corollary 5.8]. By applying [Taj16, Theorem 5.2], we get that (M,Δs)(M,\Delta^{s}) is of log general type. ∎

3. Construction of Higgs sheaves

In this section, we use Seito’s theory of Hodge modules to show some results about Hodge bundles that will be used to construct the Viehweg-Zuo Sheaves. Some constructions are inspired by [PS17].

3.1. Notations and remarks on log smooth morphism

For a log pair (X,DX)(X,D^{X}), we mean that XX is a normal variety with DXD^{X} a \mathbb{Q}-divisor and the log canonical divisor KX+DXK_{X}+D^{X} is \mathbb{Q}-Cartier. We also write ωX(DX)\omega_{X}(D^{X}) the \mathbb{Q}-line bundle given by the \mathbb{Q}-Cartier divisor KX+DXK_{X}+D^{X}. We follow the terminology of singularities of pairs as in [KM98, §2.3].

We say that the pair (X,DX)(X,D^{X}) is log smooth if XX is smooth and the support of DXD^{X}, denoted by DXredD^{X}_{\operatorname{red}}, is normal crossing. We denote ΩX(logDX):=ΩX(logDXred)\Omega_{X}(\log D^{X})\mathrel{\mathop{\mathchar 58\relax}}=\Omega_{X}(\log D^{X}_{\operatorname{red}}), the sheaf of log 1-forms with logarithmic poles along DXredD^{X}_{\operatorname{red}}. Notice that we have used ΩX(logDX)\Omega_{X}(\log D^{X}) to denote the sheaf of 𝒞\mathcal{C}-log forms in the sense of Campana in the proof of Theorem 1.0.4. However, Campana’s 𝒞\mathcal{C}-sheaves only make their appearance in the proof of Theorem 1.0.4 in Section 2 but not in the rest part of this paper.

Definition 3.1.1.

We say that f:(X,DX)(Y,DY)f\mathrel{\mathop{\mathchar 58\relax}}(X,D^{X})\to(Y,D^{Y}) is a morphism of log smooth pairs, if both (X,DX)(X,D^{X}) and (Y,DY)(Y,D^{Y}) are log smooth, and f1(DY):=f(DY)redDXredf^{-1}(D^{Y})\mathrel{\mathop{\mathchar 58\relax}}=f^{*}(D^{Y})_{\operatorname{red}}\subset D^{X}_{\operatorname{red}}. We say that ff is strict if f1(DY)=DXredf^{-1}(D^{Y})=D^{X}_{\operatorname{red}}.

We say that ff (as a morphism of log smooth pairs) is log smooth if we further have that ff is dominant as a morphism of schemes and the cokernal of the log differential map

df:fΩY(logDY)ΩX(logDX)df\mathrel{\mathop{\mathchar 58\relax}}f^{*}\Omega_{Y}(\log D^{Y})\to\Omega_{X}(\log D^{X})

is locally free. In the case that DYD^{Y} is empty, we have the following well-known result. For completeness, we give a brief proof here.

Lemma 3.1.2.

f:(X,DX)Yf\mathrel{\mathop{\mathchar 58\relax}}(X,D^{X})\to Y being log smooth is equivalent to each stratum (including XX) of (X,DX)(X,D^{X}) being smooth over YY.

Proof.

Fix one component DX1D^{X}_{1} of DXD^{X}. Consider the following commutative diagram, and define 𝒜\mathcal{A} and \mathcal{B} with all horizontal and vertical complexes forming short exact sequences

fΩY(DX1){f^{*}\Omega_{Y}(-D^{X}_{1})}fΩY{f^{*}\Omega_{Y}}(fΩY)|DX1{(f^{*}\Omega_{Y})|_{D^{X}_{1}}}ΩX(logDX)(DX1){\Omega_{X}(\log D^{X})(-D^{X}_{1})}ΩX(logDXDX1){\Omega_{X}(\log D^{X}-D^{X}_{1})}ΩDX1(log(DXDX1)|DX1){\Omega_{D^{X}_{1}}(\log(D^{X}-D^{X}_{1})|_{D^{X}_{1}})}Ωf(DX1){\Omega_{f}(-D^{X}_{1})}𝒜{\mathcal{A}}.{\mathcal{B}.}

Working locally on XX, if each stratum of (X,DX)(X,D^{X}) being smooth over YY, then by induction on the number of components of DXD^{X}, we can assume that the right two vertical complexes split, and hence the left complex splits which implies that ff is log smooth.

On the other hand, consider the following commutative diagram, with all horizontal and vertical complexes forming short exact sequences

fΩY{f^{*}\Omega_{Y}}fΩY{f^{*}\Omega_{Y}}0{0}ΩX(logDXDX1){\Omega_{X}(\log D^{X}-D^{X}_{1})}ΩX(logDX){\Omega_{X}(\log D^{X})}𝒪DX1{\mathscr{O}_{D^{X}_{1}}}𝒜{\mathcal{A}}Ωf{\Omega_{f}}𝒪DX1.{\mathscr{O}_{D^{X}_{1}}.}

If ff is log smooth, then locally on XX the middle vertical short exact complex splits, which implies the left one splits. This implies the right vertical complex in the first commutative diagram splits. Hence, by induction, each stratum of (X,DX)(X,D^{X}) is smooth over YY. ∎

In the case that the morphism f:(X,DX)(Y,DY)f\mathrel{\mathop{\mathchar 58\relax}}(X,D^{X})\to(Y,D^{Y}) is dominant, we further denote by DhXD_{h}^{X}, the horizontal part of DXD^{X}, which means it contains all components that are dominant over YY, and by DvX:=DXDXhD_{v}^{X}\mathrel{\mathop{\mathchar 58\relax}}=D^{X}-D^{X}_{h}, the vertical part of DXD^{X}. In this case, we are always making the following assumption in this pape:
(Assumption. 0) DYD^{Y}, the boundary divisor of the log smooth base, and DvXD_{v}^{X}, the vertical boundary divisor, are reduced, that is, their coefficients are 11, and the coefficients of DhXD_{h}^{X}, the horizontal boundary divisors are always in (0,1](0,1].

Notice that (Assumption. 0) is enough for the proof of Theorem 1.0.1 as the log smooth morphism fVf_{V} has no boundary on the base and DUD^{U} is dominant over VV.

Moreover we write by f:(X,DX)(Y,DY=DY)\lceil f\rceil\mathrel{\mathop{\mathchar 58\relax}}(X,\lceil D^{X}\rceil)\to(Y,\lceil D^{Y}\rceil=D^{Y}) the morphism with rounding up the pairs. Hence ff is log smooth if and only if f\lceil f\rceil is.

When we write Symi\textup{Sym}^{i}\mathscr{F} (or Sym[i]\textup{Sym}^{[i]}\mathscr{F} to be constent with notations from references), we always consider the reflexive hull of the ii-th symmetric power of the coherent sheaf \mathscr{F}; similarly, for det\det\mathscr{F}, we also consider the reflexive hull of the determinant of \mathscr{F}. Actually, due to the following important remark, taking the reflexive hull is also not necessary, since it only modifies the sheaf over a closed subset of codimension 2\geq 2. The next remark will be frequently used in later sections.

Remark 3.1.3.

To prove Theorem 1.0.6 we only need to prove it over a big open subset of YY, a Zariski open subset with its complement of codimension 2\geq 2, so, we can feel free to ignore a small closed subset of YY, a Zariski closed subset of codimension 2\geq 2. In particular, we can always assume the divisors on YY are smooth. Furthermore, if we have the morphism of log smooth pairs, f:(X,DX)(Y,DY)f\mathrel{\mathop{\mathchar 58\relax}}(X,D^{X})\to(Y,D^{Y}), is log smooth outside of the vertical boundary, i.e. ff is log smooth over YDYY\setminus D^{Y}, then by ignoring a small closed subset of YY, we have that ff itself is log smooth by the following lemma.

Lemma 3.1.4.

Fix a morphism of log smooth pairs, f:(X,DX)(Y,DY)f\mathrel{\mathop{\mathchar 58\relax}}(X,D^{X})\to(Y,D^{Y}). Assume that f|Xf1DYf|_{X\setminus f^{-1}D^{Y}} is log smooth over YDYY\setminus D^{Y}, then it is log smooth over a big open subset of YY.

Proof.

Since we can ignore all components in DXD^{X} that maps to a small closed subset of YY, we only need to show that, for any closed point pEp\in E, where EE is any smooth stratum of DXD^{X} that dominates a smooth component FF of DYD^{Y}, the cokernal of the natural morphism df:fΩY(logDY)ΩX(logDX)df\mathrel{\mathop{\mathchar 58\relax}}f^{*}\Omega_{Y}(\log D^{Y})\to\Omega_{X}(\log D^{X}) is locally free, on an open neighborhood UU of pp. However, restricted on UU, the local sections of ΩX(logDX)|U\Omega_{X}(\log D^{X})|_{U} are generated by the sections from ΩE(logDE)\Omega_{E}(\log D^{E}) and {dx1/x1,,dxk/xk}\{dx_{1}/x_{1},...,dx_{k}/x_{k}\}, where DE:=DXh|ED^{E}\mathrel{\mathop{\mathchar 58\relax}}=D^{X}_{h}|_{E}, (noting that EE is contained in DXvD^{X}_{v},) and xix_{i} are local regular functions that defines EE. Similarly, locally around f(p)f(p), local sections of ΩY(logDY)\Omega_{Y}(\log D^{Y}) are generated by the local sections of ΩF\Omega_{F} and dy/ydy/y, where yy is a local function that defines FF. However, on UU, we have fy=uΠixiaif^{*}y=u\Pi_{i}x_{i}^{a_{i}}, where uu is a local unit, and aia_{i}’s are positive integers. Hence we have, df(fdy/y)=iaidxi/xidf(f^{*}dy/y)=\sum_{i}a_{i}dx_{i}/x_{i}, which is a local section of ΩX(logDX)|U\Omega_{X}(\log D^{X})|_{U} without zero locus. Now we are only left to show that the cokernal of the natural morphism f|EΩFΩE(logDE)f|_{E}^{*}\Omega_{F}\to\Omega_{E}(\log D^{E}) is locally free, over a Zariski open subset of FF, which is true by generic smoothness. ∎

Definition 3.1.5.

Given a log smooth pair (X,D)(X,D), we denote

𝒯(X,D):\displaystyle\mathcal{T}_{(X,D)}\mathrel{\mathop{\mathchar 58\relax}} =ΩX(logD)\displaystyle=\Omega_{X}(\log D)^{\vee}
𝒜(X,D):\displaystyle\mathscr{A}^{\bullet}_{(X,D)}\mathrel{\mathop{\mathchar 58\relax}} =iSymi𝒯(X,D),\displaystyle=\bigoplus_{i}\textup{Sym}^{i}\mathcal{T}_{(X,D)},

that is, 𝒯(X,D)\mathcal{T}_{(X,D)} is the sheaf of (algebraic) vector fields with logarithmic zeroes along DXD^{X} and 𝒜(X,D)\mathscr{A}^{\bullet}_{(X,D)} is the associated graded algebra of symmetric powers. The sheaves 𝒯(X,D)\mathcal{T}_{(X,D)} and 𝒪X\mathscr{O}_{X} generate a subalgebra 𝒟X,D\mathscr{D}_{X,D} of 𝒟X\mathscr{D}_{X}, the sheaf of (algebraic) differential operators. We call 𝒟(X,D)\mathscr{D}_{(X,D)} the sheaf of (algebraic) log differential operators. The order filtration FF_{\bullet} on 𝒟X\mathscr{D}_{X} induces the order filtration on 𝒟(X,D)\mathscr{D}_{(X,D)}. With this filtration, we have a canonical isomorphism

𝒜(X,D)=GrF𝒟(X,D)\mathscr{A}^{\bullet}_{(X,D)}=\textup{Gr}^{F}_{\bullet}\mathscr{D}_{(X,D)}

where GrF𝒟(X,D)\textup{Gr}^{F}_{\bullet}\mathscr{D}_{(X,D)} denotes the associated graded algebra.

Definition 3.1.6.

Given a log smooth morphism of log smooth pairs f:(X,DX)(Y,DY)f\mathrel{\mathop{\mathchar 58\relax}}(X,D^{X})\to(Y,D^{Y}),

Ωf\displaystyle\Omega_{f} :=coker(fΩY(logDY)ΩX(logDX)),\displaystyle\mathrel{\mathop{\mathchar 58\relax}}=\text{coker}(f^{*}\Omega_{Y}(\log D^{Y})\to\Omega_{X}(\log D^{X})),
Ωif\displaystyle\Omega^{i}_{f} :=iΩf,\displaystyle\mathrel{\mathop{\mathchar 58\relax}}=\wedge^{i}\Omega_{f},
DRf\displaystyle\textup{DR}_{f} :=[𝒪XΩ1fΩfk],\displaystyle\mathrel{\mathop{\mathchar 58\relax}}=[\mathscr{O}_{X}\to\Omega^{1}_{f}\to...\to\Omega_{f}^{k}],

with cohomology degrees at k,,0-k,...,0, k=dimXdimYk=\dim X-\dim Y. We call DRf\textup{DR}_{f} the relative log de Rham complex of ff.

We set FpΩif=ΩifF_{p}\Omega^{i}_{f}=\Omega^{i}_{f} for p0p\geq 0 and FpΩif=0F_{p}\Omega^{i}_{f}=0 for p<0p<0. Then we have the induced filtration on the complex DRf\textup{DR}_{f}:

FDRf:=[Fk𝒪XFk+1Ω1fFΩfk].F_{\bullet}\textup{DR}_{f}\mathrel{\mathop{\mathchar 58\relax}}=[F_{\bullet-k}\mathscr{O}_{X}\to F_{\bullet-k+1}\Omega^{1}_{f}\to...\to F_{\bullet}\Omega_{f}^{k}].

The associated graded complexes are GrFiDRf=Ωkif[i]\textup{Gr}^{F}_{i}\textup{DR}_{f}=\Omega^{k-i}_{f}[i] for all ii.

Taking the push-forward functor RfRf_{*} on (DRf,FDRf)(\textup{DR}_{f},F_{\bullet}\textup{DR}_{f}), we then have the relative log Hodge-to-de Rham spectral squence. See Theorem 3.3.1 for further disucssions.

Since DXD^{X} and DYD^{Y} are possibly \mathbb{Q}-divisors, we write the \mathbb{Q}-line bundle by

ωf:=ωX(DX)f(ω1Y(DY)).\omega_{f}\mathrel{\mathop{\mathchar 58\relax}}=\omega_{X}(D^{X})\otimes f^{*}(\omega^{-1}_{Y}(-D^{Y})).

Using f\lceil f\rceil, we particularly have ωf=Ωfk\omega_{\lceil f\rceil}=\Omega_{f}^{k}. When DXD^{X} and DYD^{Y} are empty, ωf=ωXf(ω1Y)\omega_{f}=\omega_{X}\otimes f^{*}(\omega^{-1}_{Y}), the relative canonical sheaf of ff.

Everything defined as above except ωf\omega_{f}, only depends on the the information of f\lceil f\rceil. We still use the sub-index ff to make notations simpler.

3.2. Direct image of graded 𝒜\mathscr{A}^{\bullet}-modules

Given a morphism of log smooth pairs f:(X,DX)(Y,DY)f\mathrel{\mathop{\mathchar 58\relax}}(X,D^{X})\to(Y,D^{Y}), fixing a line bundle \mathcal{L} on XX, we denote

if,,:=𝐑if(1ωX(DX)𝐋𝒜(X,DX)f(𝒜(Y,DY)ω1Y(DY)))\mathcal{M}^{i}_{f,\mathcal{L},\bullet}\mathrel{\mathop{\mathchar 58\relax}}=\mathbf{R}^{i}f_{*}(\mathcal{L}^{-1}\otimes\omega_{X}(\lceil D^{X}\rceil)\otimes^{\mathbf{L}}_{\mathscr{A}^{\bullet}_{(X,D^{X})}}f^{*}(\mathscr{A}^{\bullet}_{(Y,D^{Y})}\otimes\omega^{-1}_{Y}(-\lceil D^{Y}\rceil)))

for each ii. By definition, if,,\mathcal{M}^{i}_{f,\mathcal{L},\bullet} is a graded 𝒜(Y,DY)\mathscr{A}^{\bullet}_{(Y,D^{Y})}-modules. Hence it possesses a Higgs-type morphism

θf,i:f,,if,,+1iΩ(Y,DY).\theta_{f,\mathcal{L}}^{i}\mathrel{\mathop{\mathchar 58\relax}}\mathcal{M}_{f,\mathcal{L},\bullet}^{i}\to\mathcal{M}_{f,\mathcal{L},\bullet+1}^{i}\otimes\Omega_{(Y,D^{Y})}.

Denote 𝒦if,,:=Ker(θif,)\mathcal{K}^{i}_{f,\mathcal{L},\bullet}\mathrel{\mathop{\mathchar 58\relax}}=\textup{Ker}(\theta^{i}_{f,\mathcal{L}}). By repeatedly composing θf,iId\theta_{f,\mathcal{L}}^{i}\otimes\textup{Id} with itself kk times, we have the induced morphism

θf,,i,k:f,,if,,+kiSymkΩ(Y,DY).\theta_{f,\mathcal{L},\bullet}^{i,k}\mathrel{\mathop{\mathchar 58\relax}}\mathcal{M}_{f,\mathcal{L},\bullet}^{i}\to\mathcal{M}_{f,\mathcal{L},\bullet+k}^{i}\otimes\textup{Sym}^{k}\Omega_{(Y,D^{Y})}.

The following proposition is the pair analogue of [WW19, Lemma 6.2].

Proposition 3.2.1.

If ff is log smooth, we have the following quasi-isomorphism of graded 𝒜(Y,DY)\mathscr{A}^{\bullet}_{(Y,D^{Y})}-modules

f,,i\displaystyle\mathcal{M}_{f,\mathcal{L},\bullet}^{i} q.i.𝐑if(1GrDRf).\displaystyle\stackrel{{\scriptstyle q.i.}}{{\simeq}}\mathbf{R}^{i}f_{*}(\mathcal{L}^{-1}\otimes\textup{Gr}_{\bullet}\textup{DR}_{f}).

In particular, we have that if,,\mathcal{M}^{i}_{f,\mathcal{L},\bullet} are coherent over 𝒪Y\mathscr{O}_{Y}.

Proof.

First, we have the relative log Spencer complex

𝒜(X,DX)𝒯kf𝒜(X,DX)𝒯k1f𝒜(X,DX),\mathscr{A}^{\bullet}_{(X,D^{X})}\otimes\mathcal{T}^{k}_{f}\to\mathscr{A}^{\bullet}_{(X,D^{X})}\otimes\mathcal{T}^{k-1}_{f}\to\dots\to\mathscr{A}^{\bullet}_{(X,D^{X})},

given by

Pv1vpiPviv1v^ivp,P\otimes v_{1}\wedge\cdots\wedge v_{p}\mapsto\sum_{i}Pv_{i}\otimes v_{1}\wedge\cdots\wedge\hat{v}_{i}\wedge\cdots\wedge v_{p},

for viv_{i} sections of 𝒯f\mathcal{T}_{f}, where 𝒯fi\mathcal{T}_{f}^{i} is the 𝒪X\mathscr{O}_{X}-dual of Ωfi\Omega_{f}^{i}. We locally choose a free basis of 𝒯kf\mathcal{T}^{k}_{f}, denoted by (ξ1,,ξk)(\xi_{1},\dots,\xi_{k}). Then by construction, the relative log Spencer complex locally is the Koszul complex of 𝒜(X,DX)\mathscr{A}^{\bullet}_{(X,D^{X})}, with actions given by multiplications of ξi\xi_{i}, for all ii. Since ff is log smooth, we know that (ξ1,,ξk)(\xi_{1},\dots,\xi_{k}) is a regular sequence in 𝒜(X,DX)\mathscr{A}^{\bullet}_{(X,D^{X})} and the 0-th cohomology sheaf is f𝒜(Y,DY)f^{*}\mathscr{A}^{\bullet}_{(Y,D^{Y})}. Therefore, the relative log Spencer complex is a locally free resolution of f𝒜(Y,DY)f^{*}\mathscr{A}^{\bullet}_{(Y,D^{Y})}. Using the dual-pair (Ωfi,𝒯fi)(\Omega_{f}^{i},\mathcal{T}_{f}^{i}), we then see that f,,\mathcal{M}_{f,\mathcal{L},\bullet} is quasi-isomorphic to

𝐑if(1GrDRf).\mathbf{R}^{i}f_{*}(\mathcal{L}^{-1}\otimes\textup{Gr}_{\bullet}\textup{DR}_{f}).

Simplification of Notations.

To simplify notations, we write

f,=𝐑0f(1GrDRf),\mathcal{M}_{f,\mathcal{L}}=\mathbf{R}^{0}f_{*}(\mathcal{L}^{-1}\otimes\textup{Gr}_{\bullet}\textup{DR}_{f}),

for the rest of this paper. In the case that \mathcal{L} is trivial, we omit \mathcal{L} from the lower-index, that is, we write

if=𝐑if(GrDRf) and f=0f.\mathcal{M}^{i}_{f}=\mathbf{R}^{i}f_{*}(\textup{Gr}_{\bullet}\textup{DR}_{f})\textup{ and }\mathcal{M}_{f}=\mathcal{M}^{0}_{f}.

Similarly, when \mathcal{L} is trivial, 𝒦0f,,\mathcal{K}^{0}_{f,\mathcal{L},\bullet} (resp. θ0,kf,,\theta^{0,k}_{f,\mathcal{L},\bullet}) is simplified to 𝒦f,\mathcal{K}_{f,\bullet} (resp. θkf,\theta^{k}_{f,\bullet}) or 𝒦\mathcal{K}_{\bullet} (resp. θk\theta^{k}_{\bullet}) when there is no ambiguity of ff.

3.3. Direct image of filtered log 𝒟\mathscr{D}-modules and canonical extensions of variation of mixed Hodge structures

We now discuss variations of mixed Hodge structures (VMHS) associated to log smooth morphisms. Let us refer to [KasVMHS] for the definition of VMHS and admissible VMHS. In contrast to the definition in loc. cit., we assume the Hodge filtrations and the weight filtrations are both increasing filtrations, to be consistent with the good filtration for 𝒟\mathscr{D}-modules underlying Hodge modules.

The following theorem is well-known; see for instance [FF] and [KawH]. We give it an alternative proof by using Saito’s mixed Hodge modules.

Theorem 3.3.1.

Suppose that f:(X,DX)(Y,DY)f\mathrel{\mathop{\mathchar 58\relax}}(X,D^{X})\to(Y,D^{Y}) is a projective log smooth morphism between log smooth pairs with DYD^{Y} smooth (but not necessarily irreducible). We then have:

  1. (1)

    The relative log Hodge-to-de Rham spectral sequence degenerate at E1E_{1}.

  2. (2)

    We have an admissible VMHS VV underlying the local system 𝐑ifoUX\mathbf{R}^{i}f^{o}_{*}\mathbb{Q}_{U^{X}}, where fof^{o} denotes the morphism fo:UXXDXUYYDYf^{o}\mathrel{\mathop{\mathchar 58\relax}}U^{X}\coloneqq X\setminus D^{X}\to U^{Y}\coloneqq Y\setminus D^{Y}. Moreover, if ff is strict, then VV is a variation of Hodge structure (VHS).

  3. (3)

    The associated graded module (with respect to the Hodge filtration) of the upper-canonical extension V0V^{\geq 0}, that is, the logarithmic extension of VV with eigenvalues of the monodromy of 𝐑ifoUX\mathbf{R}^{i}f^{o}_{*}\mathbb{Q}_{U^{X}} along DYD^{Y} in [0,1)[0,1)), is if\mathcal{M}^{i}_{f}.

Proof.

We first deal with the case that DYD^{Y} is empty. We consider the mixed Hodge module MM underlying RjUX[n]Rj_{*}\mathbb{Q}_{U^{X}}[n], where j:UXXj\mathrel{\mathop{\mathchar 58\relax}}U^{X}\hookrightarrow X. The underlying filtered 𝒟X\mathscr{D}_{X}-module of MM is 𝒪X(DX)\mathscr{O}_{X}(*D^{X}) with

𝒪X(DX)=k𝒪X(kDX)\mathscr{O}_{X}(*D^{X})=\bigcup_{k\in\mathbb{Z}}\mathscr{O}_{X}(kD^{X})

and the filtration

Fp𝒪X(DX)={𝒪X((p+1)DX),q00,q<0F_{p}\mathscr{O}_{X}(*D^{X})=\left\{\begin{array}[]{lr}\mathscr{O}_{X}((p+1)D^{X}),&q\geq 0\\ 0,&q<0\end{array}\right.

By [BPW17, Proposition 2.3], we obtained a filtered quasi-isomorphism

(3.3.1) f+(𝒪X(DX),F)Rf(DRf,FDRf)f_{+}(\mathscr{O}_{X}(*D^{X}),F_{\bullet})\simeq Rf_{*}(\textup{DR}_{f},F_{\bullet}DR_{f})

where f+f_{+} denotes the filtered 𝒟\mathscr{D}-module direct image functor. By the strictness of the Hodge filtration in [Sa88, 3.3.17] and [Sa90, 2.15], the filtered complex f+(𝒪X(DX),F)f_{+}(\mathscr{O}_{X}(*D^{X}),F_{\bullet}) is strict or equivalently the relative log Hodge-to-de Rham spectral sequence degenerate at E1E_{1} by the filtered quasi-isomorphism (3.3.1). Hence Part (1) follows in this case.

Since RifDRfR^{i}f_{*}\textup{DR}_{f} is 𝒪\mathscr{O}-coherent and hence it is locally free over 𝒪Y\mathscr{O}_{Y} as

RifDRfif+𝒪X(DX)R^{i}f_{*}\textup{DR}_{f}\simeq\mathcal{H}^{i}f_{+}\mathscr{O}_{X}(*D^{X})

are both 𝒟Y\mathscr{D}_{Y}-modules. Hence, the underlying perverse sheaf 𝐑ifoUX[n]\mathbf{R}^{i}f^{o}_{*}\mathbb{Q}_{U^{X}}[n] is locally constant (upto a shift). By [Sa90, 3.27], we hence obtain Part (2) in this case.

If DYD^{Y} is smooth but not empty, then one can use the double-strictness of direct image functor for mixed Hodge modules of normal crossing type (see [Sa90, §3] and also [W17a]). More precisely, one applies for instance [W17a, Theorem 15] and Part (1), (2) and (3) follow in general. ∎

We also need the following weak negativity of Kodaira-Spencer kernals proved in [PW].

Theorem 3.3.2.

[PW, Theorem 4.8] In the situation of Theorem 3.3.1, if ff is strict, then 𝒦p\mathcal{K}^{\vee}_{p} is weakly positive for each pp\in\mathbb{Z}, where \bullet^{\vee} denotes the 𝒪\mathscr{O}-dual of \bullet.

3.4. Adding extra boundary divisors

We assume that f:(X,DX)(Y,DY)f\colon(X,D^{X})\to(Y,D^{Y}) is a log smooth morphism of log smooth pairs. If we add a divisor DD^{\prime} to DYD^{Y} (assuming DD^{\prime} is not supported on DYD^{Y}) and obtain another normal crossing divisor DYD^{\prime Y} over a big open subset of YY (by getting rid of the singular locus of DD^{\prime} and the intersection of DD^{\prime} and DYD^{Y}), then we define DXD^{\prime X} by DXh=DXhD^{\prime X}_{h}=D^{X}_{h} and DXv=f1DYD^{\prime X}_{v}=f^{-1}D^{\prime Y}, and denote f:(X,DX)(Y,DY)f^{\prime}\mathrel{\mathop{\mathchar 58\relax}}(X,D^{\prime X})\to(Y,D^{\prime Y}) the new log smooth morphism (by Lemma 3.1.4). Since ff is smooth away from DYD^{Y}, by a local computation one easily obtains that

(3.4.1) ΩfΩf,\Omega_{f}\simeq\Omega_{f^{\prime}},

over a big open subset of YY, which further implies

(3.4.2) if,if,,\mathcal{M}^{i}_{f,\mathcal{L}}\simeq\mathcal{M}^{i}_{f^{\prime},\mathcal{L}},

as graded 𝒜(Y,DY)\mathscr{A}_{(Y,D^{Y})}-modules over a big open subset of YY.

More generally, we have the following.

Proposition 3.4.1.

Fix a projective log smooth morphism f:(X,DX)(Y,DY)f\mathrel{\mathop{\mathchar 58\relax}}(X,D^{X})\to(Y,D^{Y}), and a morphism of log smooth pairs η:(Y,DY)(Y,DY)\eta\mathrel{\mathop{\mathchar 58\relax}}(Y^{\prime},D^{Y^{\prime}})\to(Y,D^{Y}), with the underlying morphism of schemes YYY^{\prime}\to Y being smooth. Let X=X×YYX^{\prime}=X\times_{Y}Y^{\prime}, DYD^{Y^{\prime}} be a SNC divisor containing η1DY\eta^{-1}D^{Y} and f:(X,DX)(Y,DY)f^{\prime}\mathrel{\mathop{\mathchar 58\relax}}(X^{\prime},D^{X^{\prime}})\to(Y^{\prime},D^{Y^{\prime}}) be the induced morphism of log smooth pairs with DX=f1DYη¯1DXD^{X^{\prime}}=f^{\prime-1}D^{Y^{\prime}}\cup\bar{\eta}^{-1}D^{X} in the following diagram:

(X,DX){(X^{\prime},D^{X^{\prime}})}(X,DX){(X,D^{X})}(Y,DY){(Y^{\prime},D^{Y^{\prime}})}(Y,DY){(Y,D^{Y})}η¯\scriptstyle{\bar{\eta}}f\scriptstyle{f^{\prime}}f\scriptstyle{f}η\scriptstyle{\eta}

Then, over a big open subset of YY^{\prime}, we have an natural isomorphism of graded 𝒜(Y,DY)\mathscr{A}^{\bullet}_{(Y^{\prime},D^{Y^{\prime}})}-modules

ηff.\eta^{*}\mathcal{M}_{f}\simeq\mathcal{M}_{f^{\prime}}.
Proof.

We first note that, using the identity (3.4.2), we only need to show the case when DY=η1DYD^{Y^{\prime}}=\eta^{-1}D^{Y}. In this case we also have DX=η¯1DXD^{X^{\prime}}=\bar{\eta}^{-1}D^{X}. The required statement then follows from Proposition 3.2.1 and the smooth base-change.

Proposition 3.4.2.

Assume that f:(X,DX)(Y,DY)f\mathrel{\mathop{\mathchar 58\relax}}(X,D^{X})\to(Y,D^{Y}) is a projective log smooth morphism with DX=f1DYD^{X}=f^{-1}D^{Y} (that is, ff is strict), and ψ:YY\psi\mathrel{\mathop{\mathchar 58\relax}}Y^{\prime}\to Y is a finite morphism branched over DYD^{Y} with ψ1DY\psi^{-1}D^{Y} normal crossing. Let (X,DX)(X^{\prime},D^{X^{\prime}}) be a log resolution of the normalization of the main component of X×YYX\times_{Y}Y^{\prime} with DYD^{Y^{\prime}} a SNC divisor containing ψ1DY\psi^{-1}D^{Y} and DX=f1DYD^{X^{\prime}}=f^{\prime-1}D^{Y^{\prime}} (we can assume that DXD^{X^{\prime}} is also normal crossing after taking further log resolutions), and f:(X,DX)(Y,DY)f^{\prime}\mathrel{\mathop{\mathchar 58\relax}}(X^{\prime},D^{X^{\prime}})\to(Y^{\prime},D^{Y^{\prime}}) be the induced projective morphism of log smooth pairs in the following diagram

(X,DX){(X^{\prime},D^{X^{\prime}})}(X,DX){(X,D^{X})}(Y,DY){(Y^{\prime},D^{Y^{\prime}})}(Y,DY){(Y,D^{Y})}ψ¯\scriptstyle{\bar{\psi}}f\scriptstyle{f^{\prime}}f\scriptstyle{f}ψ\scriptstyle{\psi}

Then, over a big open subset of YY^{\prime}, we have a natural inclusion of 𝒜(Y,DY)\mathscr{A}^{\bullet}_{(Y^{\prime},D^{Y^{\prime}})}-modules

ψff.\psi^{*}\mathcal{M}_{f}\to\mathcal{M}_{f^{\prime}}.
Proof.

Since over YDYY^{\prime}\setminus D^{Y^{\prime}} ψ\psi is étale, we have ψf\psi^{*}\mathcal{M}_{f} is identical with f\mathcal{M}_{f^{\prime}} over YDYY^{\prime}\setminus D^{Y^{\prime}}, by Proposition 3.4.2. Then by Theorem 3.3.1 (3), we further have that f\mathcal{M}_{f} (resp. f\mathcal{M}_{f^{\prime}}) is the associated graded module of the upper-canonical extension V0fV^{\geq 0}_{f} (resp. V0fV^{\geq 0}_{f^{\prime}} ) of the VHS of the 0-th cohomology of the smooth fibers of ff (resp. ff^{\prime}). By [PW, Proposition 4.4], we see that

ψV0fV0f.\psi^{*}V^{\geq 0}_{f}\subseteq V^{\geq 0}_{f^{\prime}}.

Then we apply [PW, Corollary 4.7] for each filtrant of V0fV^{\geq 0}_{f} and V0fV^{\geq 0}_{f^{\prime}} in their Hodge filtrations and obtain the required inclusion. ∎

3.5. Birationally invariant Hodge modules

Theorem 3.5.1.

Given a projective dominant morphism f:XYf\mathrel{\mathop{\mathchar 58\relax}}X\to Y with XX and YY being smooth, there exists a pure Hodge module MfM_{f} on YY, which corresponds to a generically defined VHS (by using the equivalence in [Sa90, 3.21]), such that for any projective morphism f:XYf^{\prime}\mathrel{\mathop{\mathchar 58\relax}}X^{\prime}\to Y with XX^{\prime} being smooth and birational to XX, MfM_{f} is isomorphic to a direct summand of 0f+HX\mathcal{H}^{0}f^{\prime}_{+}\mathbb{Q}^{H}_{X^{\prime}}, and its lowest filtered piece in the Hodge filtration is isomorphic to fωf=fωff_{*}\omega_{f}=f^{\prime}_{*}\omega_{f^{\prime}}, where HX\mathbb{Q}^{H}_{X^{\prime}} is the trivial pure Hodge module on XX^{\prime} and f+f^{\prime}_{+} denotes the (derived) direct image functor for Hodge modules.

Proof.

We first fix X~\tilde{X}, a common resolution of XX and XX^{\prime}.

X{X}X~{\tilde{X}}X{X^{\prime}}Y{Y}f\scriptstyle{f}ψ\scriptstyle{\psi}f~\scriptstyle{\tilde{f}}ϕ\scriptstyle{\phi}f\scriptstyle{f^{\prime}}

Note that fωf=fωf=f~ωf~f^{\prime}_{*}\omega_{f^{\prime}}=f_{*}\omega_{f}=\tilde{f}_{*}\omega_{\tilde{f}} as XX and XX^{\prime} are both smooth. Now we define MfM_{f} the smallest sub-Hodge module of 0f~+HX~\mathcal{H}^{0}\tilde{f}_{+}\mathbb{Q}^{H}_{\tilde{X}} that contains f~ωf~\tilde{f}_{*}\omega_{\tilde{f}}. More precisely, due to the semi-simplicity of pure Hodge modules (see [Sa88, §5]), we define \mathcal{M} the sub-pure Hodge module of 0f~+HX~\mathcal{H}^{0}\tilde{f}_{+}\mathbb{Q}^{H}_{\tilde{X}} that consists of all the simple sub-pure Hodge modules that intersect f~ωf~\tilde{f}_{*}\omega_{\tilde{f}} non-trivially. Thanks to the decomposition theorem of the direct image of pure Hodge modules under projective morphism [Sa88, Théorème 1 and Corollaire 3] (see also [Sc14, §16]) and the equivalence in [Sa90, 3.21], we have that both 0f+HX\mathcal{H}^{0}f_{+}\mathbb{Q}^{H}_{X} and 0f+HX\mathcal{H}^{0}f^{\prime}_{+}\mathbb{Q}^{H}_{X^{\prime}} are direct summands of 0f~+HX~\mathcal{H}^{0}\tilde{f}_{+}\mathbb{Q}^{H}_{\tilde{X}}, and all three pure Hodge modules share the same lowest filtered piece. As a consequence, we conclude that MfM_{f} satisfies the conditions. ∎

Proposition 3.5.2.

Fix two projective morphisms of log smooth pairs

fi:(Xi,DXi)(Y,DY), i=1,2,f_{i}\mathrel{\mathop{\mathchar 58\relax}}(X_{i},D^{X_{i}})\to(Y,D^{Y}),\text{ }i=1,2,

with DYD^{Y} a smooth divisor (but not necessarily irreducible) and X1X_{1} and X2X_{2} birational. Assume that DXi=f1iDYD^{X_{i}}=f^{-1}_{i}D^{Y} (in particular, DXiD^{X_{i}} are SNC divisors) and fif_{i} are smooth over YDYY\setminus D^{Y}, for i=1,2i=1,2. Recalling the notations in §3.2, we can find a graded 𝒜(Y,DY)\mathscr{A}^{\bullet}_{(Y,D^{Y})}-module ^\hat{\mathcal{M}}, which is a direct summand of both fi\mathcal{M}_{f_{i}} for i=1,2i=1,2. Furthermore, all of ^,fi\hat{\mathcal{M}},\mathcal{M}_{f_{i}} share the same lowest graded piece.

Proof.

Recall that fi\mathcal{M}_{f_{i}} are isomorphic to the associated graded modules of the upper-canonical extension along DYD^{Y} of the VHS given by the restriciton of 0fi+HXi\mathcal{H}^{0}f_{i+}\mathbb{Q}^{H}_{X_{i}} on YDYY\setminus D^{Y} (see Theorem 3.3.1(2) and (3)). By the previous theorem, we get a pure Hodge module Mf1M_{f_{1}} that corresponds to a VHS on YDYY\setminus D^{Y}, called VV. Now we set ^\hat{\mathcal{M}} to be the associated graded module of the canonical extension along DYD^{Y} of VV, with the real part of the eigenvalues of the residues in [0,1)[0,1). ∎

3.6. Construction of Higgs sheaves from a section

Suppose that we have the following diagram of projective morphisms of log smooth pairs:

(Z,DZ){(Z,D^{Z})}(X,DX){(X,D^{X})}(Y,DY),{(Y,D^{Y}),}π\scriptstyle{\pi}g\scriptstyle{g}f\scriptstyle{f}

with π\pi generically finite, and both ff and gg being log smooth. By forgetting the horizontal part of DZD^{Z} respect to gg, we denote g0:(Z,DZv)(Y,DY)g^{0}\mathrel{\mathop{\mathchar 58\relax}}(Z,D^{Z}_{v})\to(Y,D^{Y}). Note that, by Lemma 3.1.2 and Lemma 3.1.4, g0g^{0} is log smooth at least over a big open subset of YY. For simplicity, in the rest of this subsection, we assume g0g^{0} is log smooth by removal a small closed subset of YY.

We have a natural morphism πΩX(logDX)ΩZ(logDZ),\pi^{*}\Omega_{X}(\log D^{X})\to\Omega_{Z}(\log D^{Z}), which induces a natural graded morphism of graded complexes

(3.6.1) πGrDRfGrDRg.\pi^{*}\textup{Gr}_{\bullet}\textup{DR}_{f}\to\textup{Gr}_{\bullet}\textup{DR}_{g}.

For invertible sheaves \mathcal{L} and 𝒜\mathcal{A} on XX and YY respectively, we assume the following two assumptions:
(Assumption. 1) H0(Z,πg𝒜1)0.H^{0}(Z,\pi^{*}\mathcal{L}\otimes g^{*}\mathcal{A}^{-1})\neq 0.
(Assumption. 2) the support of the effective Cartier divisor (θ)(\theta) contains DZhD^{Z}_{h}.

By (Assumption. 1), fixing one non-trivial section

θH0(Z,πg𝒜1)0,\theta\in H^{0}(Z,\pi^{*}\mathcal{L}\otimes g^{*}\mathcal{A}^{-1})\neq 0,

we have an induced inclusion

π1g𝒜1.\pi^{*}\mathcal{L}^{-1}\to g^{*}\mathcal{A}^{-1}.

We then have an induced graded morphism

(3.6.2) π(GrDRf1)GrDRgg𝒜1.\pi^{*}(\textup{Gr}_{\bullet}\textup{DR}_{f}\otimes\mathcal{L}^{-1})\to\textup{Gr}_{\bullet}\textup{DR}_{g}\otimes g^{*}\mathcal{A}^{-1}.

as the composition of the natural morphisms

π(GrDRf𝒪1)GrDRg𝒪π1GrDRg𝒪g𝒜1.\pi^{*}(\textup{Gr}_{\bullet}\textup{DR}_{f}\otimes_{\mathscr{O}}\mathcal{L}^{-1})\to\textup{Gr}_{\bullet}\textup{DR}_{g}\otimes_{\mathscr{O}}\pi^{*}\mathcal{L}^{-1}\to\textup{Gr}_{\bullet}\textup{DR}_{g}\otimes_{\mathscr{O}}g^{*}\mathcal{A}^{-1}.

By (Assumption. 2), the morphism (3.6.2) naturally factors through

π(GrDRf1)GrDRg(DZh)g(𝒜1).\pi^{*}(\textup{Gr}_{\bullet}\textup{DR}_{f}\otimes\mathcal{L}^{-1})\to\textup{Gr}_{\bullet}\textup{DR}_{g}(-D^{Z}_{h})\otimes g^{*}(\mathcal{A}^{-1}).

Note that the natural inclusion ΩZ(logDZ)(DZh)ΩZ(logDZv)\Omega_{Z}(\log D^{Z})(-D^{Z}_{h})\to\Omega_{Z}(\log D^{Z}_{v}) induces a natural inclusion of complexes:

GrDRg(DZh)GrDRg0.\textup{Gr}_{\bullet}\textup{DR}_{g}(-D^{Z}_{h})\to\textup{Gr}_{\bullet}\textup{DR}_{g^{0}}.

We then have an induced morphism

π(GrDRf1)GrDRg0g𝒜1.\pi^{*}(\textup{Gr}_{\bullet}\textup{DR}_{f}\otimes\mathcal{L}^{-1})\to\textup{Gr}_{\bullet}\textup{DR}_{g^{0}}\otimes g^{*}\mathcal{A}^{-1}.

This further induces a morphism

ηθ:f,g0,g𝒜,\eta_{\theta}\mathrel{\mathop{\mathchar 58\relax}}\mathcal{M}_{f,\mathcal{L}}\to\mathcal{M}_{g^{0},g^{*}\mathcal{A}},

as graded 𝒜(Y,DY)\mathscr{A}^{\bullet}_{(Y,D^{Y})}-modules by Proposition 3.2.1. Furthermore, the lowest graded piece of ηθ\eta_{\theta}:

gπ(ωf1)gωg0𝒜1,g_{*}\pi^{*}(\omega_{f}\otimes\mathcal{L}^{-1})\to g_{*}\omega_{g^{0}}\otimes\mathcal{A}^{-1},

is induced by the morphism π(ωf1)ωg0g𝒜1\pi^{*}(\omega_{f}\otimes\mathcal{L}^{-1})\to\omega_{g^{0}}\otimes g^{*}\mathcal{A}^{-1}. Since π(ωf1)ωg0g𝒜1\pi^{*}(\omega_{f}\otimes\mathcal{L}^{-1})\to\omega_{g^{0}}\otimes g^{*}\mathcal{A}^{-1} is induced by θ\theta and hence injective, by the left exactness of gg_{*}, we see that

gπ(ωf1)gωg0𝒜1,g_{*}\pi^{*}(\omega_{f}\otimes\mathcal{L}^{-1})\to g_{*}\omega_{g^{0}}\otimes\mathcal{A}^{-1},

is also injective.

4. Geometric construction

4.1. Stable reduction for families with arbitrary variation

In [WW19, §4], we introduced stable reduction for log smooth families with maximal variation. In this section, we discuss stable reduction for log smooth families with arbitrary variation.

Suppose that fV:(U,DU)Vf_{V}\colon(U,D^{U})\to V is a log smooth family of projective log smooth pairs of log general type, with (U,DU)(U,D^{U}) being klt and VV smooth. Consider the relative canonical model

fV,c:(Uc,DUc)V.f_{V,c}\colon(U_{c},D^{U_{c}})\to V.

We write vv the volume of KUc,y+DUcyK_{U_{c,y}}+D^{U_{c}}_{y} for yVy\in V (by invariance of pluri-genera, vv is constant over VV). Let II be a finite coefficient set II closed under addition and containing the coefficients of DUcD^{U_{c}}. We consider the coarse moduli space, denoted by 𝔐\mathfrak{M}, for stable log varieties of a fixed dimension, volume vv and the coefficient set II. Let us refer to [KP16, §6] for the construction of 𝔐\mathfrak{M} and related properties. It is proved in loc. cit. that 𝔐\mathfrak{M} is a projective reduced scheme and the corresponding moduli stack is an Deligne-Mumford stack. The relative canonical model fV,cf_{V,c} (, which is a stable family thanks to invariance of pluri-genera in the log smooth case [HMX18, Theorem 4.2],) induces a moduli map V𝔐.V\longrightarrow\mathfrak{M}. Since 𝔐\mathfrak{M} is projective, we take a projective compactification YY of VV so that EY=YVE^{Y}=Y\setminus V is normal crossing and the moduli map extends to a projective map Y𝔐Y\to\mathfrak{M} (we probably need to replace fVf_{V} by a birational model).

Proposition 4.1.1.

Under the above construction, replacing YY by a further resolution and removing a small closed subset, we can construct the following commutative diagram

(4.1.1) (Xc,DXc){(X_{c},D^{X_{c}})}(X1,c,DX1,c){(X^{\prime}_{1,c},D^{X^{\prime}_{1,c}})}(X2,c,DX2,c){(X^{\prime}_{2,c},D^{X^{\prime}_{2,c}})}(Xc,DXc){(X^{\sharp}_{c},D^{X^{\sharp}_{c}})}Y{Y}Y{Y^{\prime}}Y{Y^{\sharp}}fc\scriptstyle{f_{c}}1{\square_{1}}f1,c\scriptstyle{f^{\prime}_{1,c}}ψc\scriptstyle{\psi_{c}}ρ\scriptstyle{\rho}2{\square_{2}}f2,c\scriptstyle{f^{\prime}_{2,c}}ηc\scriptstyle{\eta_{c}}fc\scriptstyle{f^{\sharp}_{c}}ψ\scriptstyle{\psi}η\scriptstyle{\eta}

such that

  1. (0)

    (Xc,DXc)(X^{\bullet}_{c},D^{X_{c}^{\bullet}}) are all klt pairs with =,1,2,\bullet=\emptyset,^{\prime}_{1},^{\prime}_{2},^{\sharp};

  2. (1)

    fcf^{\sharp}_{c} is a stable family with maximal variation;

  3. (2)

    the square 1\square_{1} is Cartesian over VV;

  4. (3)

    the square 2\square_{2} is Cartesian;

  5. (4)

    η\eta is a dominant smooth morphism, not necessarily proper;

  6. (5)

    ψ\psi is finite and flat;

  7. (6)

    for all sufficiently positive and divisible mm, we have isomorphisms

    f2,cωmX2,c/Y(mDX2,c)\displaystyle f^{\prime}_{2,c*}\omega^{m}_{X^{\prime}_{2,c}/Y^{\prime}}(mD^{X^{\prime}_{2,c}}) ηfωmXc/Y(mDXc);\displaystyle\simeq\eta^{*}f^{\sharp}_{*}\omega^{m}_{X^{\sharp}_{c}/Y^{\sharp}}(mD^{X^{\sharp}_{c}});
    detf2,cωmX2,c/Y(mDX2,c)\displaystyle\det f^{\prime}_{2,c*}\omega^{m}_{X^{\prime}_{2,c}/Y^{\prime}}(mD^{X^{\prime}_{2,c}}) ηdetfωmXc/Y(mDXc);\displaystyle\simeq\eta^{*}\det f^{\sharp}_{*}\omega^{m}_{X^{\sharp}_{c}/Y^{\sharp}}(mD^{X^{\sharp}_{c}});
  8. (7)

    for sufficiently positive and divisible NmN_{m} (depending on mm), we can find a line bundle 𝒜m\mathcal{A}_{m} on YY such that

    ψ𝒜m(detf2,cωmX2,c/Y(mDX2,c))Nm.\psi^{*}\mathcal{A}_{m}\simeq(\det f^{\prime}_{2,c*}\omega^{m}_{X^{\prime}_{2,c}/Y^{\prime}}(mD^{X^{\prime}_{2,c}}))^{N_{m}}.
Proof.

By [LMB, Theorem 16.6], 𝔐\mathfrak{M} has a finite covering S𝔐S\to\mathfrak{M} which is induced by a stable family over SS in the moduli stack of 𝔐\mathfrak{M}. We then take YY^{\prime} to be a desingularization of the main component of Y×𝔐SY\times_{\mathfrak{M}}S (the component dominant YY^{\prime}). Then we got a generic finite map ψ:YY\psi\colon Y^{\prime}\to Y. We then take the Stein factorization of the induced morphism YSY^{\prime}\to S and obtain a fibration η:YY\eta\mathrel{\mathop{\mathchar 58\relax}}Y^{\prime}\to Y^{\sharp}. Thanks to Raynaud-Gruson flattening theorem ([RG71, Théorèm 5.2.2]), after replacing YY^{\sharp} by a further resolution, we can assume that η:YY\eta\colon Y^{\prime}\to Y^{\sharp} is a flat fibration, with YY^{\sharp} being projective and smooth. Note that YY^{\prime} might have more than 1 components, and we then replace YY^{\prime} by the main component. Now we take a Kawamata covering YY{Y^{\sharp}}^{\prime}\to Y^{\sharp} so that there exists a YY^{\prime\prime}, a desingularization of the main component of Y×YYY^{\prime}\times_{Y^{\sharp}}{Y^{\sharp}}^{\prime} satisfying that YYY^{\prime\prime}\to{Y^{\sharp}}^{\prime} is semistable in codimension 1; semistable means that the morphism is flat with reduced fibres, see for instance [AK00] for a stronger result. In particular, we can remove a small closed subset of YY^{\prime\prime} to make the morphism be smooth. Replace YY^{\prime} by YY^{\prime\prime} and YY^{\sharp} by Y{Y^{\sharp}}^{\prime}. To finish the construction of the bases, we remove the largest reduced divisor FF on YY^{\prime} such that codimψ(F)2\operatorname{codim}\psi(F)\geq 2, and replace YY by the image of YY^{\prime} (with FF removal) under ψ\psi. Note that we only removed a small subset of YY.

Now we start to construct the XX^{\bullet} level. We first take a klt log smooth pair (X,DX)(X,D^{X}) with a projective morphism fY:(X,DX)Yf_{Y}\mathrel{\mathop{\mathchar 58\relax}}(X,D^{X})\to Y so that fY|V=fVf_{Y}|_{V}=f_{V}. Since general fibers of ff are of log general type, by [BCHM, Theorem 1.2] we take fcf_{c} to be the relative log canonical model of fYf_{Y}. In particular, (X,DXc)(X,D^{X_{c}}) is a klt pair. We then set X1,cX^{\prime}_{1,c}, the main component of the normalization of Xc×YYX_{c}\times_{Y}Y^{\prime}, with boundary divisor DX1,cD^{X^{\prime}_{1,c}} given by ωX1,c(DX1,c)=ψcωXc(DXc)\omega_{X^{\prime}_{1,c}}(D^{X^{\prime}_{1,c}})=\psi_{c}^{*}\omega_{X_{c}}(D^{X_{c}}). Since ψ\psi is finite, so is ψc\psi_{c}, hence (X1,c,DX1,c)(X^{\prime}_{1,c},D^{X^{\prime}_{1,c}}) is klt by [KM98, Proposition 5.20]. The induced morphism YSY^{\sharp}\to S is generic finite, so it induces a stable family fc:(Xc,DXc)Yf_{c}^{\sharp}\mathrel{\mathop{\mathchar 58\relax}}(X^{\sharp}_{c},D^{X^{\sharp}_{c}})\to Y^{\sharp}. f2,c:(X2,c,DX2,c)Yf^{\prime}_{2,c}\mathrel{\mathop{\mathchar 58\relax}}(X^{\prime}_{2,c},D^{X^{\prime}_{2,c}})\to Y^{\prime} is defined as the stable family induced by fcf^{\sharp}_{c} under the smooth base change η\eta. Note that by construction, we have that f1,cf^{\prime}_{1,c} and f2,cf^{\prime}_{2,c} coincide on ψ1V\psi^{-1}V, hence we can blowup (X1,c,DX1,c)(X^{\prime}_{1,c},D^{X^{\prime}_{1,c}}) to make ρ\rho a morphism, without changing f1,cf^{\prime}_{1,c} over ψ1V\psi^{-1}V. Now the morphisms satisfy (1) to (5), and (6) follows by the flat base change.

To prove (7), we use the argument similar to the proof of [VZ03, Corollary ix)]. Note that, over an open set V0YV_{0}\subset Y, since f2,cf^{\prime}_{2,c} can be induced by fcf_{c} using a flat base change ψ\psi, we can canonically identifying ψdetfωmXc/Y\psi^{*}\det f_{*}\omega^{m}_{X_{c}/Y} and detf2,cωmX2,c/Y(mDX2,c)\det f^{\prime}_{2,c*}\omega^{m}_{X^{\prime}_{2,c}/Y^{\prime}}(mD^{X^{\prime}_{2,c}}) over V0:=ψ1V0V^{\prime}_{0}\mathrel{\mathop{\mathchar 58\relax}}=\psi^{-1}V_{0}, and set BB the divisor support on YV0Y^{\prime}\setminus V^{\prime}_{0}, satisfying

ψdetfωmXc/Y(mDXc)=detf2,cωmX2,c/Y(mDX2,c)(B),\psi^{*}\det f_{*}\omega^{m}_{X_{c}/Y}(mD^{X_{c}})=\det f^{\prime}_{2,c*}\omega^{m}_{X^{\prime}_{2,c}/Y^{\prime}}(mD^{X^{\prime}_{2,c}})(B),

that is also canonically defined. Now we only need to show that BB is the pullback of some \mathbb{Q}-divisor on YY, so only need to show that, for any two components B1B_{1} and B2B_{2}, if they have a same image under ψ\psi, then they share a same coefficient. Now we only need to show that over a general point pp of ψ(B1)\psi(B_{1}). Take a local curve QQ that only intersect with ψ(B)\psi(B) at pp. Now fQf_{Q}, the restriction of fcf_{c} onto QQ, is stable over QpQ\setminus p, so we can find a cyclic cover π:Q~Q\pi\mathrel{\mathop{\mathchar 58\relax}}\tilde{Q}\to Q that is totally ramified at pp, such that the induced new family f~Q\tilde{f}_{Q} over Q~\tilde{Q} is stable and compatible with the base change of fQf_{Q} over QpQ\setminus p. Denote QQ^{\prime} any irreducible component of ψ1Q\psi^{-1}Q, and let Q~=Q~×QQ\tilde{Q}^{\prime}=\tilde{Q}\times_{Q}Q^{\prime}. Although we can induce two stable families over Q~\tilde{Q}^{\prime}, but since they are the same over general point, hence they are the same family, by the properness of the moduli functor, and we denote the family by fQ~f_{\tilde{Q}^{\prime}}. Note that fQ~,fQf_{\tilde{Q}},f_{Q^{\prime}} and fQ~f_{\tilde{Q}^{\prime}} are all compatible with flat base change, so we only need to verify the proposition on fQf_{Q} and fQ~.f_{\tilde{Q}}. This is true, since π\pi is totally ramified at pp. ∎

Since fcf^{\sharp}_{c} is a stable family with maximal variation, we have detfcωmXc/Y(mDXc)\det f^{\sharp}_{c*}\omega^{m}_{X_{c}^{\sharp}/Y^{\sharp}}(mD^{X_{c}^{\sharp}}) is big by [KP16, Theorem 7.1]. We now fix a pair of mm and NmN_{m} as in the previous proposition. To simplify notation, we set

𝒜\displaystyle\mathcal{A}^{\sharp} :=(detfcωmXc/Y(mDXc))Nm\displaystyle\mathrel{\mathop{\mathchar 58\relax}}=(\det f^{\sharp}_{c*}\omega^{m}_{X_{c}^{\sharp}/Y^{\sharp}}(mD^{X_{c}^{\sharp}}))^{N_{m}}
𝒜\displaystyle\mathcal{A}^{\prime} :=(detf2,cωmX2,c/Y(mDX2,c))Nm\displaystyle\mathrel{\mathop{\mathchar 58\relax}}=(\det f^{\prime}_{2,c*}\omega^{m}_{X^{\prime}_{2,c}/Y^{\prime}}(mD^{X^{\prime}_{2,c}}))^{N_{m}}
𝒜\displaystyle\mathcal{A} :=𝒜m.\displaystyle\mathrel{\mathop{\mathchar 58\relax}}=\mathcal{A}_{m}.

By Part (6) and (7) in Proposition 4.1.1, we have

(4.1.2) ψ𝒜=𝒜=η𝒜 and κ(𝒜)=κ(𝒜)=κ(𝒜)=Var(fV).\psi^{*}\mathcal{A}=\mathcal{A}^{\prime}=\eta^{*}\mathcal{A}^{\sharp}\textup{ and }\kappa(\mathcal{A})=\kappa(\mathcal{A}^{\prime})=\kappa(\mathcal{A}^{\sharp})=\textup{Var}(f_{V}).

4.2. Birational refinement of the stable reduction

In this subsection, we refine the stable reduction in Proposition 4.1.1, which will be used to construct Viehweg-Zuo sheaf in later section. We first show the following lemma.

Lemma 4.2.1.

Assume that we have two log smooth projective morphisms

fi:(Xi,DXi)(Y,DY),with i=1,2,f_{i}\mathrel{\mathop{\mathchar 58\relax}}(X_{i},D^{X_{i}})\to(Y,D^{Y}),\text{with }i=1,2,

whose generic fibers share the same birational equivalent model, i.e. we have log resolutions over ξ\xi, ϕi,ξ:(Xξ,DXξ)(Xi,ξ,DXi,ξ),\phi_{i,\xi}\mathrel{\mathop{\mathchar 58\relax}}(X_{\xi},D^{X_{\xi}})\to(X_{i,\xi},D^{X_{i,\xi}}), satisfying

ϕi,ξωXξ(DXi,ξ)ωXξ(DXξ),\phi^{*}_{i,\xi}\omega_{X_{\xi}}(D^{X_{i,\xi}})\simeq\omega_{X_{\xi}}(D^{X_{\xi}}),

where ξ\xi is the generic point of YY. After adding components to DYD^{Y} if necessary (cf. §3.4), forgetting the information over a small closed subset of YY, and replacing DXivD^{X_{i}}_{v} by f1(DY)f^{-1}(D^{Y}), we have f1ωf1m=f2ωf2mf_{1*}\omega_{f_{1}}^{m}=f_{2*}\omega_{f_{2}}^{m}.

Proof.

According to the assumption, we can have a common log resolution, hence the following commutative diagram, with all pairs are log smooth,

(X1,DX1){(X_{1},D^{X_{1}})}(X,DX){(X,D^{X})}(X2,DX2){(X_{2},D^{X_{2}})}(Y,DY),{(Y,D^{Y}),}f1\scriptstyle{f_{1}}ϕ1\scriptstyle{\phi_{1}}ϕ2\scriptstyle{\phi_{2}}f\scriptstyle{f}f2\scriptstyle{f_{2}}

so that DX|Xξ=DXξD^{X}|_{X_{\xi}}=D^{X_{\xi}}. We can assume that every irreducible component of DXD^{X} is dominant over YY, i.e. the vertical part of DXD^{X} over YY is zero. After we extend DYD^{Y} and forget a small closed subset of YY, we can assume that ff is smooth away from DYD^{Y} without changing log smoothness of f1f_{1} and f2f_{2}. We then replace the vertical parts of DX1D^{X_{1}}, DXD^{X} and DX2D^{X_{2}} by f11(DY)f_{1}^{-1}(D^{Y}), f1(DY)f^{-1}(D^{Y}) and f21(DY)f_{2}^{-1}(D^{Y}) respectively. Using Lemma 3.1.4, we can assume that all vertical morphisms are projective log smooth morphisms.

To prove the statement, by symmetry we only need to show that ϕ1(ωX(DX))m=(ωX1(DX1))m\phi_{1*}(\omega_{X}(D^{X}))^{m}=(\omega_{X_{1}}(D^{X_{1}}))^{m}. Consider a Cartier divisor BB with

𝒪(B)=(ωX(DX))mϕ1(ωX1(DX1))m,\mathscr{O}(B)=(\omega_{X}(D^{X}))^{m}\otimes\phi_{1}^{*}(\omega_{X_{1}}(D^{X_{1}}))^{-m},

which is an ϕ1\phi_{1}-exceptional divisor. Since ff is log smooth, any irreducible ϕ1\phi_{1}-exceptional divisor EE on XX maps to DYD^{Y} or dominates YY. By the assumption of fif_{i} over the generic point of YY, EE cannot dominant YY. Then EE maps to DYD^{Y} and EE is ff-exceptional. Hence the coefficient of EE in DXD^{X} is 11 by construction (see (Assumption. 0) in §3.1). Since (X1,DX1)(X_{1},D^{X_{1}}) is log canonical (by (Assumption. 0) in §3.1), the coefficient of EE in BB is positive, which concludes the proof. ∎

Proposition 4.2.2.

Assume that we have a commutative diagram (4.1.1) as in Proposition 4.1.1, we can always find a log resolution π:(X,DX)(Xc,DXc)\pi\mathrel{\mathop{\mathchar 58\relax}}(X,D^{X})\to(X_{c},D^{X_{c}}), so that we can have the following commutative diagram, allowing forgetting a small closed subset of YY, YY^{\prime} and YY^{\sharp}.

(4.2.1) (X,DX){(X,D^{X})}(X1,DX1){(X^{\prime}_{1},D^{X^{\prime}_{1}})}(X,DX){(X^{\prime},D^{X^{\prime}})}(X2,DX2){(X^{\prime}_{2},D^{X^{\prime}_{2}})}(X,DX){(X^{\sharp},D^{X^{\sharp}})}(Y,DY){(Y,D^{Y})}(Y,DY){(Y^{\prime},D^{Y^{\prime}})}(Y,DY){(Y^{\sharp},D^{Y^{\sharp}})}f\scriptstyle{f}1{\square_{1}}f1\scriptstyle{f^{\prime}_{1}}ψ¯\scriptstyle{\bar{\psi}}ϕ1\scriptstyle{\phi_{1}}ϕ2\scriptstyle{\phi_{2}}f\scriptstyle{f^{\prime}}2{\square_{2}}f2\scriptstyle{f^{\prime}_{2}}η¯\scriptstyle{\bar{\eta}}f\scriptstyle{f^{\sharp}}ψ\scriptstyle{\psi}η\scriptstyle{\eta}

such that

  1. (1)

    all the log pairs are log smooth;

  2. (2)

    ff is a projective compactification of fVf_{V};

  3. (3)

    all the downwards morphisms are log smooth, with vertical boundaries are reduced and the coefficients of horizontal boundaries are in (0,1)(0,1);

  4. (4)

    the square 1\square_{1} is Cartesian over YDYY\setminus D^{Y} and ψ\psi is étale over YDYY\setminus D^{Y};

  5. (5)

    the square 2\square_{2} is Cartesian;

  6. (6)

    all ϕ\phi_{\bullet} are birational, with =1,2\bullet=1,2;

  7. (7)

    ψ¯\bar{\psi} factors through the natural morphism ψ0:X0X\psi_{0}\mathrel{\mathop{\mathchar 58\relax}}X^{\prime}_{0}\to X, where X0X^{\prime}_{0} is the normalization of the main component of X×YY,X\times_{Y}Y^{\prime}, and ϕ0:X1X0\phi_{0}\mathrel{\mathop{\mathchar 58\relax}}X^{\prime}_{1}\to X^{\prime}_{0} is birational, and denote f0:X0Yf^{\prime}_{0}\mathrel{\mathop{\mathchar 58\relax}}X^{\prime}_{0}\to Y^{\prime} the induced morphism;

  8. (8)

    for all mm sufficiently positive and divisible, we have isomorphisms

    f0ψ0ωmff1ψ¯ωmff1ωmf1fωmff2ωmf2ηfωmf;f^{\prime}_{0*}\psi_{0}^{*}\omega^{m}_{f}\simeq f^{\prime}_{1*}\bar{\psi}^{*}\omega^{m}_{f}\simeq f^{\prime}_{1*}\omega^{m}_{f^{\prime}_{1}}\simeq f^{\prime}_{*}\omega^{m}_{f^{\prime}}\simeq f^{\prime}_{2*}\omega^{m}_{f^{\prime}_{2}}\simeq\eta^{*}f^{\sharp}_{*}\omega^{m}_{f^{\sharp}};
  9. (9)

    denoting f:(X,DX)(Y,DY)\lceil f^{\bullet}\rceil\mathrel{\mathop{\mathchar 58\relax}}(X^{\bullet},\lceil D^{X^{\bullet}}\rceil)\to(Y^{\bullet},D^{Y^{\bullet}}), with =,1,2,\bullet=\emptyset,^{\prime}_{1},^{\prime}_{2},^{\sharp}, i.e. only considering the boundary with reduced structure, for all mm sufficiently positive and divisible, we have isomorphisms

    f0ψ0ωmff1ψ¯ωmff1ωmf1fωmff2ωmf2ηfωmf.f^{\prime}_{0*}\psi_{0}^{*}\omega^{m}_{\lceil f\rceil}\simeq f^{\prime}_{1*}\bar{\psi}^{*}\omega^{m}_{\lceil f\rceil}\simeq f^{\prime}_{1*}\omega^{m}_{\lceil f^{\prime}_{1}\rceil}\simeq f^{\prime}_{*}\omega^{m}_{\lceil f^{\prime}\rceil}\simeq f^{\prime}_{2*}\omega^{m}_{\lceil f^{\prime}_{2}\rceil}\simeq\eta^{*}f^{\sharp}_{*}\omega^{m}_{\lceil f^{\sharp}\rceil}.
Proof.

From Proposition 4.1.1, we first replace (Xc,DXc)(X^{\bullet}_{c},D^{X^{\bullet}_{c}}) by their log resolutions (X,DX)(X^{\bullet},D^{X^{\bullet}}), with =,1,2,\bullet=\emptyset,^{\prime}_{1},^{\prime}_{2},^{\sharp}, and the horizontal part of the boundary divisor DXD^{X^{\bullet}} is defined by taking the positive and horizontal part of ω1XπωXc(DXc)\omega^{-1}_{X^{\bullet}}\otimes\pi^{\bullet*}\omega_{X^{\bullet}_{c}}(D^{X^{\bullet}_{c}}), where π\pi^{\bullet} are the corresponding log resolutions. Since (Xc,DXc)(X^{\bullet}_{c},D^{X^{\bullet}_{c}}) are klt, so are (X,DX)(X^{\bullet},D^{X^{\bullet}}). Moreover, by the construction of fcf_{c} in Proposition 4.1.1, we can assume f:(X,DX)Yf\mathrel{\mathop{\mathchar 58\relax}}(X,D^{X})\to Y is a projective compactification of the initial log smooth family fV:(V,DU)Vf_{V}\mathrel{\mathop{\mathchar 58\relax}}(V,D^{U})\to V. In particular, we can keep track the initial boundary divisor EY=YVE^{Y}=Y\setminus V because ultimately we are interested in the log Kodaira dimension of VV.

We do not consider the vertical parts of DXD^{X^{\bullet}} at this stage, but they will be specified after we fixed the boundary divisors on YY^{\bullet}. Since base changing by an étale or smooth morphism will keep the log smoothness, we can keep 1\square_{1} and 2\square_{2} are Cartesian over the assigned locus, and make ψ¯\bar{\psi} and η¯\bar{\eta} are morphisms of log smooth pairs. By the construction of f1,cf^{\prime}_{1,c} in Proposition 4.1.1, we have (7).

Set DYD^{Y} and DYD^{Y^{\sharp}} be divisors on YY and YY^{\sharp} respectively, so that ff and ff^{\sharp} are log smooth over the base outside of the boundary (cf. Lemma 3.1.2). By expanding DYD^{Y}, we can assume that ψ\psi is étale over YDYY\setminus D^{Y}. Set DYD^{Y^{\prime}} a divisor on YY^{\prime}, so that it contains both ψ1DY\psi^{-1}D^{Y} and η1DY\eta^{-1}D^{Y^{\sharp}}, and both f1f^{\prime}_{1} and f2f^{\prime}_{2} are log smooth over YDYY\setminus D^{Y^{\prime}}. By Lemma 3.1.4, also Remark 3.1.3, we can assume that f1f^{\prime}_{1} and f2f^{\prime}_{2} are log smooth, by removing a small subset of YY^{\prime}. Hence, we can apply the previous lemma, by further expending DYD^{Y^{\prime}}, constructing ff^{\prime} as in the proof of the lemma, and replacing all vertical boundaries by the inverse image DYD^{Y^{\prime}}, we have the identities of the third term to the fifth term of (8). Then, we replace DYD^{Y} and DYD^{Y^{\sharp}}, by ψDY\psi_{*}D^{Y^{\prime}} and ηDY.\eta_{*}D^{Y^{\prime}}. Now we expanding DXD^{X^{\bullet}} by setting DXv=f1DYD^{X^{\bullet}_{v}}=f^{\bullet-1}D^{Y^{\bullet}}. Meanwhile we take further log resolutions of (X,DX)(X^{\bullet},D^{X^{\bullet}}) if needed, to keep them being log smooth, but do not change the locus that are already log smooth, so that we can keep the Cartesian condition. At last, remove some small subsets of YY, YY^{\prime}, and YY^{\sharp}, we have all the downwards morphisms are log smooth.

(X,DX)(X0,DX0)(X1,DX1)(Y,DY)(Y,DY).ff0ψ¯0ϕ0f1ψ¯ψ