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

Kodaira-type Vanishings via Non-abelian Hodge Theory

Chuanhao Wei
Abstract.

In this paper, we use non-abelian Hodge Theory to study Kodaira type vanishings and its generalizations. In particular, we generalize Saito vanishing using mixed twistor π’Ÿ\mathscr{D}-modules. We also generalize it to a Kawamata-Viehweg type vanishing using β„š\mathbb{Q}-divisors, and we also prove a relative version for a projective morphism.

1. Introduction

Let XX be a complex projective manifold. Given a reduced normal crossing divisor D=D1+…+DnD=D_{1}+...+D_{n} on XX, and a vector 𝐝=(d1,…,dn)βˆˆβ„n,\mathbf{d}=(d_{1},...,d_{n})\in\mathbb{R}^{n}, we set 𝐝​D=d1​D1+…+dn​Dn\mathbf{d}D=d_{1}D_{1}+...+d_{n}D_{n}. We say 𝐝β‰₯𝐚\mathbf{d}\geq\mathbf{a}, if diβ‰₯aid_{i}\geq a_{i} for all ii, similarly defined for other inequalities. We use 𝟏=(1,…,1)βˆˆβ„n\mathbf{1}=(1,...,1)\in\mathbb{R}^{n}, similarly for 𝟎\mathbf{0}. We also use a​DaD to denote a​D1+…+a​DnaD_{1}+...+aD_{n}, for aβˆˆβ„a\in\mathbb{R}. A divisor denoted by a single letter, e.g. DD, is an effective integral divisor, unless there is a coefficient in front of it, like 𝐝​D,a​D,\mathbf{d}D,aD, as above, or explicitly stated otherwise. We mainly consider β„š\mathbb{Q}-divisors, and the ℝ\mathbb{R}-divisor case can be reduced to such case by an argument appeared later in Remark 1.7.

The renowned Kodaira-Akizuki-Nakano Vanishing Theorem has the following generalized form, which is due to Esnault-Viehweg [EV86], [EV92], combined with the Kawamata-Viehweg type Vanishing formulation [AMPW]:

Theorem 1.1.

Notations as above, assume that we have an ample divisor AA. Let LL be an integral divisor, such that

L≑n​u​ma​A+𝐝​D,L\equiv_{num}aA+\mathbf{d}D,

with a>0a>0, πŸŽβ‰€πβ‰€πŸ\mathbf{0}\leq\mathbf{d}\leq\mathbf{1}. Then, we have the following vanishings:

Hi​(X,Ξ©Xj​(log⁑D)​(βˆ’D)βŠ—π’ͺX​(L))=0\displaystyle H^{i}(X,\Omega^{j}_{X}(\log D)(-D)\otimes\mathcal{O}_{X}(L))=0 ,forΒ i+j>dim(X);\displaystyle,\text{for }i+j>\dim(X);
Hi​(X,Ξ©Xj​(log⁑D)βŠ—π’ͺX​(βˆ’L))=0\displaystyle H^{i}(X,\Omega^{j}_{X}(\log D)\otimes\mathcal{O}_{X}(-L))=0 ,forΒ i+j<dim(X).\displaystyle,\text{for }i+j<\dim(X).

There is another direction to generalize Kodaira-Akizuki-Nakano Vanishing using Saito’s mixed Hodge Module. In this paper, we say a pair (β„³,Fβˆ™)(\mathcal{M},F_{\bullet}) admits a mixed Hodge Module, where β„³\mathcal{M} is a right coherent π’Ÿ\mathscr{D}-module, and Fβˆ™F_{\bullet} is its increasing Hodge filtration, if we can add a weight filtration to make it an algebraic graded polarizable mixed Hodge module in the sense of [Sai90, Β§2]. We remark that, among other things, the existence of the weight compatible polarization on (β„³,Fβˆ™)(\mathcal{M},F_{\bullet}) is essential while defining mixed Hodge Module. However, for the vanishing theory we studying in this paper, we only use its formal property: the strictness of certain functors with respect to Fβˆ™F_{\bullet}. Hence we mute those information for the ease of notation.

Theorem 1.2.

Let XX be a complex projective manifold, with a reduced divisor DD and a semi-ample line bundle β„’\mathcal{L}. Assume further that ℒ​(d​D)\mathcal{L}(dD) is an ample line bundle, for some dβ‰₯0d\geq 0. Let (β„³,Fβˆ™)(\mathcal{M},F_{\bullet}) be a pair admitting a mixed Hodge Module as described above. Then, we have the following vanishings:

ℍi(DRXGrβˆ™Fβ„³[βˆ—D]βŠ—β„’)\displaystyle\mathbb{H}^{i}(\textup{DR}_{X}\textup{Gr}^{F}_{\bullet}\mathcal{M}[*D]\otimes\mathcal{L}) =0,for ​i>0;\displaystyle=0,\text{for }i>0;
ℍi(DRXGrβˆ™Fβ„³[!D]βŠ—β„’βˆ’1)\displaystyle\mathbb{H}^{i}(\textup{DR}_{X}\textup{Gr}^{F}_{\bullet}\mathcal{M}[!D]\otimes\mathcal{L}^{-1}) =0,for ​i<0.\displaystyle=0,\text{for }i<0.

In the case with the absence of DD, it is just Saito’s vanishing [Sai90, (2.g)]. Please refer [Wei20b, Theorem 20] for this general form, with the proof follows Saito’s proof of Saito’s vanishing. Please note that, in loc. cit., we use Sp to denote the de Rham functor on right modules. We use the notation DR here, to follow the convention in [Sab05] and [Moc15].

In 80s’, the non-abelian Hodge Theory is mainly developed by C. Simpson. As the starting point, he studies the harmonic bundles and uses that to give extra structures on cohomologies of a compact KΓ€hlar manifold, with the coefficient being a semisimple local system, instead of the constant local system as studied in the classical Hodge theory. Simpson also proposes the Meta theorem which states that what works for classical Hodge theory shall still work for non-abelian Hodge theory. Later on, following Saito’s idea of constructing mixed Hodge Modules, the theory of mixed twistor π’Ÿ\mathscr{D}-modules is initiated by C. Sabbah. Then, T. Mochizuki completes this spectacular theory in the past decade, which can also be viewed as a satisfying answer towards Simpson’s Meta Theorem. In the same vein, heuristically, the vanishing theorem above can be generalized to the corresponding mixed twistor π’Ÿ\mathscr{D}-modules setting. Before we state the generalization, let’s set up the notations for mixed twistor π’Ÿ\mathscr{D}-modules.

For any complex manifold XX, we denote 𝒳=XΓ—β„‚Ξ»\mathcal{X}=X\times\mathbb{C}_{\lambda}, where β„‚Ξ»\mathbb{C}_{\lambda} is the affine line with Ξ»\lambda as its coordinate, and p:𝒳→X,p:\mathcal{X}\to X, q:𝒳→ℂλq:\mathcal{X}\to\mathbb{C}_{\lambda} the natural projections. Following the notations in [Sab05] and [Moc15], in this paper, we say that a right algebraic β„›X\mathscr{R}_{X}-module β„³\mathcal{M} on 𝒳\mathcal{X} admits a mixed twistor π’Ÿ\mathscr{D}-module, if there exists an algebraic graded polarizable mixed twistor π’Ÿ\mathscr{D}-module represented by a filtered β„›\mathscr{R}-triple (𝒯,W)(\mathcal{T},W), with the triple 𝒯=(β„³β€²,β„³β€²β€²,C),\mathcal{T}=(\mathcal{M}^{\prime},\mathcal{M}^{\prime\prime},C), and β„³=β„³β€²β€²\mathcal{M}=\mathcal{M}^{\prime\prime}. We also use ΞDol:=β„³/λ​ℳ\Xi_{\textup{Dol}}:=\mathcal{M}/\lambda\mathcal{M} and ΞDR:=β„³/(Ξ»βˆ’1)​ℳ\Xi_{\textup{DR}}:=\mathcal{M}/(\lambda-1)\mathcal{M}, which are exact functors from strict coherent β„›X\mathscr{R}_{X}-modules to coherent π’ŸX\mathscr{D}_{X}-modules and coherent π’œX:=GrF​(π’ŸX)\mathscr{A}_{X}:=\textup{Gr}^{F}(\mathscr{D}_{X})-modules, respectively [Sab05, 1.1.a, Definition 1.2.1]. As we remarked in the mixed Hodge Module case, the other data while defining a graded polarizable mixed twistor π’Ÿ\mathscr{D}-module is important. However, for the vanishing theorem, we only use the holomorphic picture of a mixed twistor π’Ÿ\mathscr{D}-module, and its formal property like the strictness of certain natural functors, so we will only keep track the β„›X\mathscr{R}_{X}-module β„³=β„³β€²β€²\mathcal{M}=\mathcal{M}^{\prime\prime}.

The following vanishing result is a direct non-abelian Hodge theoretic generalization of Theorem 1.2.

Theorem 1.3.

Let XX be a complex projective manifold, with a reduced divisor DD and a semi-ample line bundle β„’\mathcal{L}. Assume further that ℒ​(d​D)\mathcal{L}(dD) is an ample line bundle, for some dβ‰₯0d\geq 0. Let β„³\mathcal{M} be a right β„›X\mathscr{R}_{X}-module admitting a mixed twistor π’Ÿ\mathscr{D}-module. Then, we have the following vanishings:

ℍi(DRXΞDolβ„³[βˆ—D]βŠ—β„’)\displaystyle\mathbb{H}^{i}(\textup{DR}_{X}\Xi_{\textup{Dol}}\mathcal{M}[*D]\otimes\mathcal{L}) =0,for ​i>0;\displaystyle=0,\text{for }i>0;
ℍi(DRXΞDolβ„³[!D]βŠ—β„’βˆ’1)\displaystyle\mathbb{H}^{i}(\textup{DR}_{X}\Xi_{\textup{Dol}}\mathcal{M}[!D]\otimes\mathcal{L}^{-1}) =0,for ​i<0.\displaystyle=0,\text{for }i<0.

In the previous theorem, β„³[βˆ—D]\mathcal{M}[*D] and β„³[!D]\mathcal{M}[!D] are prolongations of β„³\mathcal{M} as coherent β„›X\mathscr{R}_{X}-modules as in [Moc15, 3.1.2, 3.1.3]. We will use β„³(βˆ—D)\mathcal{M}(*D) to denote the canonical prolongation as a coherent β„›X⁣(βˆ—D)\mathscr{R}_{X(*D)}-module [Moc15, 3.1.1].

Naturally, we want to give the previous vanishing a Kawamata-Viehweg Vanishing type formulation, as in Theorem 1.1. The next vanishing theorem can be viewed as a such generalization, and it seems to be new even in the setting of mixed Hodge Modules. The formulation is quite technical, and we will define the notations in Β§3. At this stage, for reader’s convenience, we introduce the notation

π•πšDβ„³(Ξ±)(βˆ—D)=∩VaiDiβ„³(Ξ±)(βˆ—D),\mathbf{V}^{D}_{\mathbf{a}}\mathcal{M}^{(\alpha)}(*D)=\cap V^{D_{i}}_{a_{i}}\mathcal{M}^{(\alpha)}(*D),

where Vβˆ™Diβ„³(Ξ±)(βˆ—D)V^{D_{i}}_{\bullet}\mathcal{M}^{(\alpha)}(*D) is the KM-filtration of β„³(βˆ—D)\mathcal{M}(*D) locally around Ξ»=Ξ±\lambda=\alpha. Please note that the multi-indexed filtration π•βˆ™D\mathbf{V}^{D}_{\bullet} does not behave well in general. We need the VV-compatibility on β„³(βˆ—D)\mathcal{M}(*D), Definition 3.3.

Theorem 1.4.

Let XX be a smooth projective variety, with a reduced, normal crossing divisor DD. Let β„³\mathcal{M} be a right β„›X\mathscr{R}_{X}-module admitting a mixed twistor π’Ÿ\mathscr{D}-module, and assume that β„³(βˆ—D)\mathcal{M}(*D) is VV-compatible with respect to DD. (In particular, it is the case when β„³\mathcal{M} corresponds to a tame harmonic bundle on Xβˆ–DX\setminus D). Separate DD into two groups of components: B+C=DB+C=D, and assume that we have a semi-ample divisor AA, such that A+πžβ€‹CA+\mathbf{e}C is ample for some 𝐞β‰₯𝟎\mathbf{e}\geq\mathbf{0}. Let LL be a divisor, such that

L≑n​u​ma​A+𝐛​B+πœβ€‹C,L\equiv_{num}aA+\mathbf{b}B+\mathbf{c}C,

with a>0a>0. Then, we have the following vanishings:

ℍi(DR(X,D)(ΞDol𝐕<βˆ’π›Bπ•βˆ’πœCβ„³(0)(βˆ—D))βŠ—π’ͺX(L))=0\displaystyle\mathbb{H}^{i}(\textup{DR}_{(X,D)}(\Xi_{\textup{Dol}}\mathbf{V}^{B}_{<-\mathbf{b}}\mathbf{V}^{C}_{-\mathbf{c}}\mathcal{M}^{(0)}(*D))\otimes\mathcal{O}_{X}(L))=0 ,forΒ i>0;\displaystyle,\text{for }i>0;
ℍi(DR(X,D)(ΞDol𝐕𝐛B𝐕<𝐜Cβ„³(0)(βˆ—D))βŠ—π’ͺX(βˆ’L))=0\displaystyle\mathbb{H}^{i}(\textup{DR}_{(X,D)}(\Xi_{\textup{Dol}}\mathbf{V}^{B}_{\mathbf{b}}\mathbf{V}^{C}_{<\mathbf{c}}\mathcal{M}^{(0)}(*D))\otimes\mathcal{O}_{X}(-L))=0 ,forΒ i<0.\displaystyle,\text{for }i<0.
Remark 1.5.

Here, we consider the multi-indexed KM-filtration on the canonical prolongation β„³(βˆ—D)\mathcal{M}(*D) as a coherent β„›X(βˆ—D)\mathscr{R}_{X}(*D)-module. Due to the VV-compatibility in Defintion 3.3, for any Ξ±βˆˆβ„‚Ξ»\alpha\in\mathbb{C}_{\lambda} and π€βˆˆβ„€n\mathbf{k}\in\mathbb{Z}^{n}, it satisfies

𝐕𝐝+𝐀Dβ„³(Ξ±)(βˆ—D)=𝐕𝐝Dβ„³(Ξ±)(βˆ—D)βŠ—pβˆ—π’ͺX(𝐀D).\mathbf{V}^{D}_{\mathbf{d}+\mathbf{k}}\mathcal{M}^{(\alpha)}(*D)=\mathbf{V}^{D}_{\mathbf{d}}\mathcal{M}^{(\alpha)}(*D)\otimes p^{*}\mathcal{O}_{X}(\mathbf{k}D).

Hence we can always shift the indexes of the multi-indexed KM-filtration to β‰€πŸŽ\leq\mathbf{0}, and use the VV-filtration on β„³[βˆ—D]\mathcal{M}[*D] by [Moc15, Lemma 3.1.1]. In particular, we have that, when πβ‰€πŸŽ\mathbf{d}\leq\mathbf{0},

𝐕𝐝Dβ„³(Ξ±)(βˆ—D)=𝐕𝐝Dβ„³(Ξ±)[βˆ—D].\mathbf{V}^{D}_{\mathbf{d}}\mathcal{M}^{(\alpha)}(*D)=\mathbf{V}^{D}_{\mathbf{d}}\mathcal{M}^{(\alpha)}[*D].
Remark 1.6.

Actually, we only need to show the case that 𝐞=𝟎\mathbf{e}=\mathbf{0}, i.e. AA itself is ample, since we always have

L≑n​u​ma​A+Ο΅β€‹πžβ€‹C+𝐛​B+(πœβˆ’Ο΅β€‹πž)​C,L\equiv_{num}aA+\epsilon\mathbf{e}C+\mathbf{b}B+(\mathbf{c}-\epsilon\mathbf{e})C,

noting that a​A+Ο΅β€‹πžβ€‹CaA+\epsilon\mathbf{e}C is ample, and π•βˆ’πœC=π•βˆ’πœ+Ο΅β€‹πžC\mathbf{V}^{C}_{-\mathbf{c}}=\mathbf{V}^{C}_{-\mathbf{c}+\epsilon\mathbf{e}}, 𝐕<𝐜C=𝐕<πœβˆ’Ο΅β€‹πžC\mathbf{V}^{C}_{<\mathbf{c}}=\mathbf{V}^{C}_{<\mathbf{c}-\epsilon\mathbf{e}}, for some 0<Ο΅β‰ͺ10<\epsilon\ll 1, due to the semi-continuity of the KM-filtration.

The reason that we keep the superficially more general form as stated above, is due to that it actually carries natural geometric information that will be clarified in the proof. Furthermore, following the same proof as in Β§5, Theorem 1.4 can be further generalized to the form that assume A+πžβ€‹(C+E)A+\mathbf{e}(C+E) is ample for some effective divisor EE, and replace β„³(βˆ—D)\mathcal{M}(*D) in the first (resp. second) vanishing by β„³[βˆ—E](βˆ—D)\mathcal{M}[*E](*D) (resp. β„³[!E](βˆ—D)\mathcal{M}[!E](*D)), as in Theorem 1.3.

Remark 1.7.

The above theorem also works for ℝ\mathbb{R}-divisors by the following reduction. Due to the previous remark, we can assume AA is ample. Hence, we can find 𝟎<𝐛′′β‰ͺ𝟏\mathbf{0}<\mathbf{b}^{\prime\prime}\ll\mathbf{1}, and 𝟎<πœβ€²β€²β‰ͺ𝟏\mathbf{0}<\mathbf{c}^{\prime\prime}\ll\mathbf{1}, so that a​Aβˆ’π›β€²β€²β€‹B+πœβ€²β€²β€‹CaA-\mathbf{b}^{\prime\prime}B+\mathbf{c}^{\prime\prime}C is still ample, and both 𝐛′:=𝐛+𝐛′′\mathbf{b}^{\prime}:=\mathbf{b}+\mathbf{b}^{\prime\prime} and πœβ€²:=πœβˆ’πœβ€²β€²\mathbf{c}^{\prime}:=\mathbf{c}-\mathbf{c}^{\prime\prime} are rational. Hence, we have

L≑n​u​m(a′​Aβˆ’π›β€²β€²β€‹B+πœβ€²β€²β€‹C)+𝐛′​B+πœβ€²β€‹C,L\equiv_{num}(a^{\prime}A-\mathbf{b}^{\prime\prime}B+\mathbf{c}^{\prime\prime}C)+\mathbf{b}^{\prime}B+\mathbf{c}^{\prime}C,

with a′​Aβˆ’π›β€²β€²β€‹B+πœβ€²β€²β€‹Ca^{\prime}A-\mathbf{b}^{\prime\prime}B+\mathbf{c}^{\prime\prime}C be a β„š\mathbb{Q}-ample divisor. We also can assume 𝐕<βˆ’π›Bβ€‹π•βˆ’πœC=𝐕<βˆ’π›β€²Bβ€‹π•βˆ’πœβ€²C\mathbf{V}^{B}_{<-\mathbf{b}}\mathbf{V}^{C}_{-\mathbf{c}}=\mathbf{V}^{B}_{<-\mathbf{b}^{\prime}}\mathbf{V}^{C}_{-\mathbf{c}^{\prime}} and 𝐕𝐛B​𝐕<𝐜C=𝐕𝐛′B​𝐕<πœβ€²C\mathbf{V}^{B}_{\mathbf{b}}\mathbf{V}^{C}_{<\mathbf{c}}=\mathbf{V}^{B}_{\mathbf{b}^{\prime}}\mathbf{V}^{C}_{<\mathbf{c}^{\prime}}, due to the semi-continuity of the KM-filtration. From now on, we only consider β„š\mathbb{Q}-divisors, unless explicitly stated otherwise.

We will also establish the relative versions of the previous two vanishing theorems in Β§2. Actually, it is not straight-forward to get such generalization as e.g. [KMM87, Theorem 1-2-3], using a log-smooth compactification and Serre vanishing on coherent π’ͺ\mathcal{O}-modules. This is because, in our case, we are not working in the derived category of coherent sheaves. However, due to our proof of Theorem 1.3 being functorial, we can also get a functorial proof of its relative version, Theorem 2.1, which does not use a log-smooth compactification. We then show a Nadel-type vanishing with multiplier-ideals from β„š\mathbb{Q}-divisors, and state an effective global generalization theorem, as an application.

In Β§3, we recall the notion of multi-indexed Kashiwara-Malgrange (KM) filtration, its compatibility, and using it to get the logarithmic comparison in the setting of β„›\mathscr{R}-modules as in [Wei20a]. In Β§4, we recall the geometric construction in [EV92], which is constructing a cyclic cover from a normal crossing β„š\mathbb{Q}-divisor, and study the induced local systems and their multi-indexed KM filtration. We give the proofs of our main vanishing results in Β§5.

It is also natural to generalize the injectivity/surjectivity results in [EV92, Β§5], [Wu17], [Wu21], to the Non-abelian Hodge setting. However, the author finds such type of results have a different nature of approach, and they cannot directly imply those vanishing results above using the method summerized in [EV92, Β§1, 2. proof], with the help of Serre Vanishing, for the same reason in the relative vanishing case. Hence, the author decides to leave it for another occasion.

Let’s explain how the previous two vanishing theorems cover essentially all Kodaira-type vanishings in algebraic geometry we know so far. Nevertheless, the author has no intention to say that we have new proofs to those brilliant results. On the contrary, the vanishing theorems above shall be taken as a natural summary of currently known Kodaira-type vanishings, and the proofs are based on the proofs of those results. The proofs mainly adopt Esnault and Viehweg’s geometric construction [EV92], Saito’s proof of Saito Vanishing, and the Theory of mixed twistor π’Ÿ\mathscr{D}-modules, due to Sabbah and Mochizuki. However, we still cannot recover Nadel vanishing in the analytic setting, that is proved using L2L^{2} method.

We first restrict ourselves to the case that β„³\mathcal{M} being the Rees algebra associated to a mixed Hodge Module (𝒩,Fβˆ™)(\mathcal{N},F_{\bullet}), with a right π’Ÿ\mathscr{D}-module 𝒩\mathcal{N}, and an increasing Hodge filtration Fβˆ™F_{\bullet}. We have that ΞDol​ℳ:=β„³/λ​ℳ\Xi_{\textup{Dol}}\mathcal{M}:=\mathcal{M}/\lambda\mathcal{M} is just taking the associated graded pieces, Grβˆ™F​𝒩\textup{Gr}^{F}_{\bullet}\mathcal{N}. Then, Theorem 1.3 in this case is just Theorem 1.2.

For Theorem 1.4, if we only consider β„³\mathcal{M} admits a graded polarizable variation of mixed Hodge structures on XX, C=0C=0, and πŸŽβ‰€π›<𝟏,\mathbf{0}\leq\mathbf{b}<\mathbf{1}, we have Vβˆ’π›Bβ„³(βˆ—D)≃ℳV^{B}_{-\mathbf{b}}\mathcal{M}(*D)\simeq\mathcal{M}. In particular, if we take ℳ≃ω𝒳\mathcal{M}\simeq\omega_{\mathcal{X}}, i.e. it corresponds to the trivial variation, we get the Kawamata-Viehweg vanishing for klt pairs. Hence, via a standard argument, e.g. Corollary 2.3, it implies the big and nef vanishing for the lowest graded piece of a mixed Hodge Module, which has been proved by Suh [Suh18] and Wu [Wu17] independently. Let’s also remark that the CC part can be useful in case we want to deal with log-canonical singularities, e.g. the application on showing the zero locus of holomorphic log-one forms, [Wei20b].

If we have a reduced, possibly singular Cartier divisor DD, by taking a log resolution f:(Xβ€²,Dβ€²)β†’(X,D)f:(X^{\prime},D^{\prime})\to(X,D), with fβˆ—β€‹d​D=𝐝′​Dβ€²f^{*}dD=\mathbf{d}^{\prime}D^{\prime}. We have the multiplier ideal

(1.1) π’₯(dD)βŠ—Ο‰X≃fβˆ—ΞžDol𝐕<βˆ’πβ€²D′ω𝒳′(βˆ—Dβ€²).\mathcal{J}(dD)\otimes\omega_{X}\simeq f_{*}\Xi_{\textup{Dol}}\mathbf{V}^{D^{\prime}}_{<-\mathbf{d}^{\prime}}\omega_{\mathcal{X}^{\prime}}(*D^{\prime}).

Hence we get the Nadel vanishing with the multiplier ideal of β„š\mathbb{Q}-divisors. Such relation between multiplier ideals and Hodge modules has also been studied in [BS05] and [MP20]. See also Corollary 2.3 in the next section.

In the case that β„³\mathcal{M} is tame and underlies a variation of twistor structures on Xβˆ–DX\setminus D, this is equivalent to that ΞDol​ℳ\Xi_{\textup{Dol}}\mathcal{M} gives us a slope parabolic polystable higgs bundle on Xβˆ–DX\setminus D, due to [Moc06, Theorem 1.4]. Since a parabolic semistable higgs bundle can be realized as a sequentially extension of stable higgs bundles, as argued in [AHL19, Lemma 7.1], then the two vanishings above generalize those vanishing results proved by Arapura-Hao-Li [AHL19], and by Deng-Hao [DH22]. In [DH22], their vanishing is more refined in the sense that they build a relation between the range of the vanishing degrees and the number of positive eigenvalues of the curvature form of β„’\mathcal{L}. In the case that β„’\mathcal{L} being kk-ample as in [Som78, Β§1], (which is a stronger condition, see also [Tot13],) we have a projective dominate morphism f:Xβ†’Yf:X\to Y, with β„’n=fβˆ—β€‹π’œ\mathcal{L}^{n}=f^{*}\mathcal{A} for some ample line bundle π’œ\mathcal{A} on YY, and nβˆˆβ„€+n\in\mathbb{Z}^{+}. Hence we can achieve the vanishing with the expected range as in [DH22], due to the compatibility of the direct image, DR and ΞDol\Xi_{\textup{Dol}} functor. We leave the details to the interested readers.

In this paper, we use βŠ—\otimes (without sub-index) to denote the tensor product over the corresponding structure sheaf π’ͺ\mathcal{O}, unless specified otherwise. We use f+f_{+} and f†f_{\dagger} (resp. fβˆ—f_{*} and f!f_{!}) to denote the derived direct image and direct image with proper support in the derived category of β„›\mathscr{R}-modules (resp. quasi-coherent sheaves or constructable sheaves). We use β„βˆ™\mathbb{H}^{\bullet} (resp. ℍcβˆ™\mathbb{H}^{\bullet}_{c}) to denote taking the hypercohomoloy (resp. hypercohomoloy with proper support), which is the same as the (derived) functor of taking derived direct image aβˆ—a_{*} (resp. a!a_{!}), where aa is the canonical map to Spec​(β„‚)\text{Spec}(\mathbb{C}). We consider right π’Ÿ\mathscr{D}-modules or β„›\mathscr{R}-modules by default, unless specified otherwise. Although we only consider the vanishing theory for algebraic mixed twistor π’Ÿ\mathscr{D}-modules on a quasi-projective variety, the constructions in Β§3 and Β§4 work for the analytic setting.

2. Relative Vanishings and applications

In this section, we state the relative vanishings, a generalized nef and big vanishing, and an effective global generalization theorem, as an application. Most notations shall be standard, following [Sab05] and [Moc15]. Some notations will be carefully defined in the next section.

Let’s first state a relative version of Theorem 1.3.

Theorem 2.1.

Let f:Xβ†’Sf:X\to S be a projective morphism between smooth quasi-projective varieties. On XX, assume that we have a reduced divisor DD and an ff-semi-ample line bundle β„’\mathcal{L}. Assume further that ℒ​(d​D)\mathcal{L}(dD) is an ff-ample line bundle, for some dβ‰₯0d\geq 0. Let β„³\mathcal{M} be a right β„›X\mathscr{R}_{X}-module admitting a graded polarizable mixed twistor π’Ÿ\mathscr{D}-module on XX. Then, we have the following vanishings:

β„›ifβˆ—(DRXΞDolβ„³[βˆ—D]βŠ—β„’)\displaystyle\mathcal{R}^{i}f_{*}(\textup{DR}_{X}\Xi_{\textup{Dol}}\mathcal{M}[*D]\otimes\mathcal{L}) =0,for ​i>0;\displaystyle=0,\text{for }i>0;
β„›ifβˆ—(DRXΞDolβ„³[!D]βŠ—β„’βˆ’1)\displaystyle\mathcal{R}^{i}f_{*}(\textup{DR}_{X}\Xi_{\textup{Dol}}\mathcal{M}[!D]\otimes\mathcal{L}^{-1}) =0,for ​i<βˆ’d,\displaystyle=0,\text{for }i<-d,

where d=dimSd=\dim S.

We can also show a relative version of Theorem 1.4:

Theorem 2.2.

Fix a projective morphism between smooth quasi-projective varieties f:Xβ†’Sf:X\to S. DD is a reduced, normal crossing divisor on XX. Let β„³\mathcal{M} be a right β„›X\mathscr{R}_{X}-module admitting a graded polarizable mixed twistor π’Ÿ\mathscr{D}-module on XX, and assume that β„³(βˆ—D)\mathcal{M}(*D) is VV-compatible with respect to DD. Separate DD into two groups of components: B+C=DB+C=D, and assume that we have an ff-semi-ample divisor AA, such that A+πžβ€‹CA+\mathbf{e}C is ff-ample for some 𝐞β‰₯𝟎\mathbf{e}\geq\mathbf{0}. Let LL be a divisor, such that

L≑l​i​na​A+𝐛​B+πœβ€‹C,L\equiv_{lin}aA+\mathbf{b}B+\mathbf{c}C,

with a>0a>0. Then, we have the following vanishings:

β„›ifβˆ—(DR(X,D)(ΞDol𝐕<βˆ’π›Bπ•βˆ’πœCβ„³(0)(βˆ—D))βŠ—π’ͺX(L))=0\displaystyle\mathcal{R}^{i}f_{*}(\textup{DR}_{(X,D)}(\Xi_{\textup{Dol}}\mathbf{V}^{B}_{<-\mathbf{b}}\mathbf{V}^{C}_{-\mathbf{c}}\mathcal{M}^{(0)}(*D))\otimes\mathcal{O}_{X}(L))=0 ,forΒ i>0;\displaystyle,\text{for }i>0;
β„›ifβˆ—(DR(X,D)(ΞDol𝐕𝐛B𝐕<𝐜Cβ„³(0)(βˆ—D))βŠ—π’ͺX(βˆ’L))=0\displaystyle\mathcal{R}^{i}f_{*}(\textup{DR}_{(X,D)}(\Xi_{\textup{Dol}}\mathbf{V}^{B}_{\mathbf{b}}\mathbf{V}^{C}_{<\mathbf{c}}\mathcal{M}^{(0)}(*D))\otimes\mathcal{O}_{X}(-L))=0 ,forΒ i<βˆ’d,\displaystyle,\text{for }i<-d,

where d=dimSd=\dim S.

Recall that, due to Hitchin-Kobayashi correspondence, we have that, for a stable vector bundle β„°\mathcal{E} on a smooth projective variety XX with vanishing Chern classes, we can view it as a Higgs bundle with the trivial Higgs map, that can be lift as a harmonic bundle corresponding to a unitary representation. A semistable vector bundle 𝒩\mathcal{N} with vanishing Chern classes can be realized as extensions of stable vector bundle β„°\mathcal{E} with vanishing Chern classes, [Sim91, Theorem 2]. Hence 𝒩\mathcal{N}, equipped with the trivial higgs map, can be lift as a smooth variation of mixed twistor structures.

Then, we can show the following Nadel vanishing on a semistable vector bundle with vanishing Chern classes.

Corollary 2.3.

Let XX be a smooth projective variety, 𝒩\mathcal{N} a semistable vector bundle on XX with vanishing Chern classes. Given a divisor LL, such that

L≑n​u​ma​A+b​B,a>0,bβ‰₯0L\equiv_{num}aA+bB,a>0,b\geq 0

with AA a nef and big divisor, and BB an effective divisor. Then we have

ℍi​(π’©βŠ—Ο‰β€‹(L)βŠ—π’₯​(b​B))=0,for ​i>0.\mathbb{H}^{i}(\mathcal{N}\otimes\omega(L)\otimes\mathcal{J}(bB))=0,\text{for }i>0.

where π’₯​(b​B)\mathcal{J}(bB) stands for the multiplier ideal of the β„š\mathbb{Q}-divisor b​BbB.

π’₯​(b​B)\mathcal{J}(bB) can also be replaced by π’₯​(b​|B|)\mathcal{J}(b|B|) the multiplier ideal of linear series, or π’₯​(β€–b​Bβ€–)\mathcal{J}(||bB||) the asymptotic multiplier ideal. Let’s refer [Laz04b] for more details.

Proof.

Take an embedded log-resolution of XX and the support of BB, getting Ο€:Xβ€²β†’X.\pi:X^{\prime}\to X. Set Ο€βˆ—β€‹a​A=a′​Aβ€²\pi^{*}aA=a^{\prime}A^{\prime} and Ο€βˆ—β€‹b​B=𝐛′​Bβ€²\pi^{*}bB=\mathbf{b}^{\prime}B^{\prime}. Since AA is nef and big, so is a′​Aβ€²a^{\prime}A^{\prime}. Up to a further log-resolution, we have a′​A′≑l​i​naβˆ˜β€‹A∘+πžβ€‹E,a^{\prime}A^{\prime}\equiv_{lin}a^{\circ}A^{\circ}+\mathbf{e}E, with aβˆ˜β€‹A∘a^{\circ}A^{\circ} ample and 𝟎<𝐞β‰ͺ𝟏\mathbf{0}<\mathbf{e}\ll\mathbf{1}. (Since we can apply

a′​A′≑l​i​n(a∘m​A∘+a′​(mβˆ’1)m​Aβ€²)+1mβ€‹πžβ€‹E,a^{\prime}A^{\prime}\equiv_{lin}(\frac{a^{\circ}}{m}A^{\circ}+\frac{a^{\prime}(m-1)}{m}A^{\prime})+\frac{1}{m}\mathbf{e}E,

with the β„š\mathbb{Q}-divisor in the parentheses being ample.) To summarize, we have

Ο€βˆ—β€‹L≑n​u​maβˆ˜β€‹A∘+𝐛′​Bβ€²+πžβ€‹E.\pi^{*}L\equiv_{num}a^{\circ}A^{\circ}+\mathbf{b}^{\prime}B^{\prime}+\mathbf{e}E.

Set 𝒩′=Ο€βˆ—β€‹π’©\mathcal{N}^{\prime}=\pi^{*}\mathcal{N}, and it, equipped with the trivial higgs map, still can be lifted as a variation of mixed twistor structures, and we use β„³β€²\mathcal{M}^{\prime} to denote its corresponding β„›\mathscr{R}-module. In particular, it is non-characteristic with respect to any smooth divisor, hence the multi-indexed KM-filtration of β„³β€²(βˆ—(Bβ€²+E))\mathcal{M}^{\prime}(*(B^{\prime}+E)) only jumps at the integers, along each component of Bβ€²B^{\prime} and EE, e.g. (3.11). In the case that Bβ€²B^{\prime} and EE share no common component, we have

(2.1) 𝐕<βˆ’π›β€²B′𝐕<βˆ’πžEβ„³β€²(βˆ—(Bβ€²+E))≃𝐕<βˆ’π›β€²B′𝐕<𝟎Eβ„³β€²(βˆ—(Bβ€²+E))≃𝐕<βˆ’π›β€²Bβ€²β„³β€²(βˆ—Bβ€²),\mathbf{V}^{B^{\prime}}_{<-\mathbf{b}^{\prime}}\mathbf{V}^{E}_{<-\mathbf{e}}\mathcal{M}^{\prime}(*(B^{\prime}+E))\simeq\mathbf{V}^{B^{\prime}}_{<-\mathbf{b}^{\prime}}\mathbf{V}^{E}_{<\mathbf{0}}\mathcal{M}^{\prime}(*(B^{\prime}+E))\simeq\mathbf{V}^{B^{\prime}}_{<-\mathbf{b}^{\prime}}\mathcal{M}^{\prime}(*B^{\prime}),

where the second identity can be checked directly. In the case that Bβ€²B^{\prime} and EE share common components, e.g. B0β€²=E0B^{\prime}_{0}=E_{0}, we change the first term by omitting e0​E0e_{0}E_{0} in EE, and use b0β€²+e0b^{\prime}_{0}+e_{0} to replace b0β€²b^{\prime}_{0}, both as index of KM-filtration and coefficient of B0β€²B^{\prime}_{0}. Since e0e_{0} is small, it is still isomorphic to the other two terms with index 𝐛′\mathbf{b}^{\prime} the original coefficients of Bβ€²B^{\prime}. See also the remark below. Now, due to Theorem 1.4, we have

ℍi(DR(Xβ€²,Bβ€²+E)(ΞDol𝐕<βˆ’π›β€²B′𝐕<βˆ’πžEβ„³β€²(βˆ—(Bβ€²+E)))βŠ—Ο€βˆ—π’ͺ(L))=0,i>0.\mathbb{H}^{i}(\textup{DR}_{(X^{\prime},B^{\prime}+E)}(\Xi_{\textup{Dol}}\mathbf{V}^{B^{\prime}}_{<-\mathbf{b}^{\prime}}\mathbf{V}^{E}_{<-\mathbf{e}}\mathcal{M}^{\prime}(*(B^{\prime}+E)))\otimes\pi^{*}\mathcal{O}(L))=0,i>0.

Since we also have ΞDol𝐕<βˆ’π›β€²Bβ€²β„³β€²(βˆ—Bβ€²)β‰ƒπ’©β€²βŠ—ΞžDol𝐕<βˆ’π›β€²B′ω𝒳′(βˆ—Bβ€²)\Xi_{\textup{Dol}}\mathbf{V}^{B^{\prime}}_{<-\mathbf{b}^{\prime}}\mathcal{M}^{\prime}(*B^{\prime})\simeq\mathcal{N}^{\prime}\otimes\Xi_{\textup{Dol}}\mathbf{V}^{B^{\prime}}_{<-\mathbf{b}^{\prime}}\omega_{\mathcal{X}^{\prime}}(*B^{\prime}), and the Higgs connection has set to be trivial, combining (2.1),

DR(Xβ€²,Bβ€²+E)(ΞDol𝐕<βˆ’π›β€²B′𝐕<βˆ’πžEβ„³β€²(βˆ—(Bβ€²+E)))\textup{DR}_{(X^{\prime},B^{\prime}+E)}(\Xi_{\textup{Dol}}\mathbf{V}^{B^{\prime}}_{<-\mathbf{b}^{\prime}}\mathbf{V}^{E}_{<-\mathbf{e}}\mathcal{M}^{\prime}(*(B^{\prime}+E)))

decomposes as ⨁(π’©β€²βŠ—βˆ§i𝒯(X,Bβ€²+E)βŠ—ΞžDol𝐕<βˆ’π›β€²B′ω𝒳′(βˆ—Bβ€²))[i]\bigoplus(\mathcal{N}^{\prime}\otimes\wedge^{i}\mathcal{T}_{(X,B^{\prime}+E)}\otimes\Xi_{\textup{Dol}}\mathbf{V}^{B^{\prime}}_{<-\mathbf{b}^{\prime}}\omega_{\mathcal{X}^{\prime}}(*B^{\prime}))[i]

Due to local vanishing, we have

β„›fβˆ—ΞžDol𝐕<βˆ’π›β€²B′ω𝒳′(βˆ—Bβ€²)≃ℛfβˆ—Ο‰X(βˆ’βŒŠπ›β€²βŒ‹Bβ€²)≃ℛ0fβˆ—Ο‰X(βˆ’βŒŠπ›β€²βŒ‹Bβ€²)≃ωXβŠ—π’₯(bB),\mathcal{R}f_{*}\Xi_{\textup{Dol}}\mathbf{V}^{B^{\prime}}_{<-\mathbf{b}^{\prime}}\omega_{\mathcal{X}^{\prime}}(*B^{\prime})\simeq\mathcal{R}f_{*}\omega_{X}(-\lfloor\mathbf{b}^{\prime}\rfloor B^{\prime})\simeq\mathcal{R}^{0}f_{*}\omega_{X}(-\lfloor\mathbf{b}^{\prime}\rfloor B^{\prime})\simeq\omega_{X}\otimes\mathcal{J}(bB),

See also [BS05]. (Of course, we can directly apply Theorem 2.1, but it is not necessary.) Combining the vanishing above and the projection formula, we get the vanishing we need. ∎

Remark 2.4.

Note that, for the nef and big vanishing, we cannot add the CC part in Theorem 1.4 in a naive way. The reason is that we cannot identify π•πŸŽE\mathbf{V}^{E}_{\mathbf{0}} and π•βˆ’πžE\mathbf{V}^{E}_{-\mathbf{e}} as in (2.1), even when 𝐞\mathbf{e} is very small.

Remark 2.5.

The essential point of this version of Nadel vanishing is Hitchin-Kobayashi correspondence. Once we assume that, we can view the semistable vector bundle 𝒩\mathcal{N} as extensions of variation of Hodge structures coming from unitary representations, i.e. with trivial Hodge filtration. Hence itself is the lowest filtered piece, and we can apply the vanishing in the setting of Hodge module, e.g. [Wu17]. Another way to achieve the vanishing is to use the fact that 𝒩\mathcal{N} is Nakano semi-positive, and we can apply the usual L2L^{2} type Nadel vanishing.

It is natural to apply the previous Nadel vanishing to get the following effective global generation result, and we follow the exposition in [Laz04b, 10.4]. Please also refer [dC98b], [dC98a] for a much more comprehensive study on this topic.

Theorem 2.6 (Theorem of Angehrn and Siu).

Let X be a smooth projective variety of dimension nn, with AA an ample divisor on it. Fix a point x∈Xx\in X, and assume that, for every irreducible subvariety ZβŠ‚XZ\subset X passing through xx, (including XX itself,) setting mm as its dimension, and assume that

(Amβ‹…Z)>(12​n​(n+1))m.(A^{m}\cdot Z)>(\frac{1}{2}n(n+1))^{m}.

Then, for any semistable vector bundle 𝒩\mathcal{N} with vanishing Chern classes, π’©βŠ—Ο‰X​(A)\mathcal{N}\otimes\omega_{X}(A) is free at xx, i.e. π’©βŠ—Ο‰X​(A)\mathcal{N}\otimes\omega_{X}(A) has a global section that does not vanish at x.

In particular, if A≑n​u​mk​L,A\equiv_{num}kL, for some kβ‰₯12​n​(n+1)k\geq\frac{1}{2}n(n+1) and an ample divisor LL, then π’©βŠ—Ο‰X​(A)\mathcal{N}\otimes\omega_{X}(A) is free, i.e. globally generated.

Proof.

According to [Laz04a, 10.4.C], we are able to find an effective β„š\mathbb{Q}-divisor d​DdD on XX, with l​c​t​(D;x)=dlct(D;x)=d, and xx being d​DdD’s isolated LC locus, and d​D≑n​u​mλ​AdD\equiv_{num}\lambda A, for some Ξ»<1\lambda<1. Due to the vanishing in Corollary 2.3, we have

H1​(X,π’©βŠ—Ο‰X​(A)βŠ—π’₯​(d​D))=0.H^{1}(X,\mathcal{N}\otimes\omega_{X}(A)\otimes\mathcal{J}(dD))=0.

Locally around xx, π’₯​(d​D)\mathcal{J}(dD) is just xx’s ideal sheaf, so we have the surjection of the following natural restriction

H0​(X,π’©βŠ—Ο‰X​(A))β†’H0​(X,π’©βŠ—Ο‰X​(A)βŠ—C​(x)),H^{0}(X,\mathcal{N}\otimes\omega_{X}(A))\to H^{0}(X,\mathcal{N}\otimes\omega_{X}(A)\otimes C(x)),

where C​(x)C(x) denoting the one-dimensional sky-scraper sheaf, supported at xx, and this is what we need. ∎

3. Multi-indexed KM-filtration and Logarithmic comparison

In this section, we would love to generalize some results in [Wei20a] about logarithmic comparison in mixed Hodge modules setting, to the mixed twistor π’Ÿ\mathscr{D}-modules case.

Let’s first recall the definition of the Kashiwara-Malgrange (KM-)filtration on a coherent β„›X⁣(βˆ—D)\mathscr{R}_{X(*D)}-module, with respect to a smooth component HH of the reduced normal crossing divisor DD on a complex manifold XX. Since in this paper, given a β„›\mathscr{R}-module β„³\mathcal{M} on XX, we will find ourselves only use its information on the open part Xβˆ–DX\setminus D, so it is more natural and easier to consider the KM-filtration on the coherent β„›X⁣(βˆ—D)\mathscr{R}_{X(*D)}-module β„³(βˆ—D):=β„³βŠ—π’ͺ𝒳(βˆ—(pβˆ—D))\mathcal{M}(*D):=\mathcal{M}\otimes\mathcal{O}_{\mathcal{X}}(*(p^{*}D)) than to consider the KM-filtration on the coherent β„›X\mathscr{R}_{X}-modules like β„³[βˆ—D]\mathcal{M}[*D] or β„³[!D]\mathcal{M}[!D], see also Remark 1.5. Actually, β„³[βˆ—D]\mathcal{M}[*D] and β„³[!D]\mathcal{M}[!D] themselves are built from the KM-filtration on β„³(βˆ—D)\mathcal{M}(*D), as in [Moc15, 3.1].

There is a canonical β„€\mathbb{Z}-indexed increasing filtration on β„›X\mathscr{R}_{X}, stalk-wise defined by, for any (x,Ξ±)βˆˆπ’³(x,\alpha)\in\mathcal{X},

VkHβ„›X⁣(βˆ—D),(x,Ξ±)={Pβˆˆβ„›X⁣(βˆ—D),(x,Ξ±)|\displaystyle V^{H}_{k}\mathscr{R}_{X(*D),(x,\alpha)}=\{P\in\mathscr{R}_{X(*D),(x,\alpha)}|
Pβ‹…(pβˆ—β€‹π’ͺX​(i​H))\displaystyle P\cdot(p^{*}\mathcal{O}_{X}(iH)) βŠ‚(x,Ξ±)pβˆ—(π’ͺX(i+kH))(x,Ξ±),βˆ€iβˆˆβ„€}.{}_{(x,\alpha)}\subset p^{*}(\mathcal{O}_{X}(i+kH))_{(x,\alpha)},\forall i\in\mathbb{Z}\}.

We note that V0H​ℛX⁣(βˆ—D)V^{H}_{0}\mathscr{R}_{X(*D)} is a coherent ring. It can be directly checked that VkH​ℛX⁣(βˆ—D)=V0H​ℛX⁣(βˆ—D)βŠ—π’ͺX​(k​H)V^{H}_{k}\mathscr{R}_{X(*D)}=V^{H}_{0}\mathscr{R}_{X(*D)}\otimes\mathcal{O}_{X}(kH), and Vβˆ™H​ℛX⁣(βˆ—D)V^{H}_{\bullet}\mathscr{R}_{X(*D)} is a good filtration, [Bjo93, Appendix III], as a V0H​ℛX⁣(βˆ—D)V^{H}_{0}\mathscr{R}_{X(*D)}-module.

Now, we use 𝒩\mathcal{N} to denote a strict coherent β„›X⁣(βˆ—D)\mathscr{R}_{X(*D)}-module. For βˆ€Ξ±βˆˆβ„‚\forall\alpha\in\mathbb{C}, we use 𝒩(Ξ±)\mathcal{N}^{(\alpha)} to denote restricting 𝒩\mathcal{N} onto an open subset X×Δ​(Ξ±,Ο΅)βŠ‚π’³X\times\Delta(\alpha,\epsilon)\subset\mathcal{X}, for some small Ο΅>0\epsilon>0, where Δ​(Ξ±,Ο΅)\Delta(\alpha,\epsilon) is the open disk in β„‚Ξ»\mathbb{C}_{\lambda}, centered at Ξ»=Ξ±\lambda=\alpha and with radius Ο΅\epsilon.

Definition 3.1.

Let 𝒩\mathcal{N} be a strict coherent β„›X⁣(βˆ—D)\mathscr{R}_{X(*D)}-module. We say that it is strictly specializable along HH, a smooth component of DD, [Moc15, 2.1.2.2], if, for βˆ€Ξ±βˆˆβ„‚\forall\alpha\in\mathbb{C}, there exists a Kashiwara-Malgrange (KM-)filtration Vβˆ™HV^{H}_{\bullet} on 𝒩(Ξ±)\mathcal{N}^{(\alpha)}, (for some Ο΅>0\epsilon>0,) which is an exhaustive ℝ\mathbb{R}-indexed increasing filtration by coherent V0H​ℛX⁣(βˆ—D)V^{H}_{0}\mathscr{R}_{X(*D)}-modules, satisfying the following conditions:

  1. (1)

    for all aβˆˆβ„a\in\mathbb{R}, locally around any point Pβˆˆπ’³(Ξ±)P\in\mathcal{X}^{(\alpha)}, there exists some Ο΅>0\epsilon>0 VaH​𝒩(Ξ±)=Va+Ο΅H​𝒩(Ξ±)V^{H}_{a}\mathcal{N}^{(\alpha)}=V^{H}_{a+\epsilon}\mathcal{N}^{(\alpha)};

  2. (2)

    each β„›H⁣(βˆ—Dβ€²|H)\mathscr{R}_{H(*D^{\prime}|_{H})}-module GraVH​𝒩(Ξ±):=VaH/V<aH​𝒩(Ξ±)\textup{Gr}^{V^{H}}_{a}\mathcal{N}^{(\alpha)}:=V^{H}_{a}/V^{H}_{<a}\mathcal{N}^{(\alpha)} is strict , where Dβ€²=Dβˆ’HD^{\prime}=D-H;

  3. (3)

    locally around any point Pβˆˆπ’³(Ξ±)P\in\mathcal{X}^{(\alpha)},

    VaH​𝒩(Ξ±)β‹…t=Vaβˆ’1H​𝒩(Ξ±),for all ​aβˆˆβ„,V^{H}_{a}\mathcal{N}^{(\alpha)}\cdot t=V^{H}_{a-1}\mathcal{N}^{(\alpha)},\text{for all }a\in\mathbb{R},

    where tt is any local holomorphic function on XX that defines HβŠ‚XH\subset X.

  4. (4)

    For any aβˆˆβ„a\in\mathbb{R} and Pβˆˆπ’³(Ξ±)P\in\mathcal{X}^{(\alpha)}, there exists a finite set

    𝒦​(a,Ξ±,P)βŠ‚{uβˆˆβ„Γ—β„‚|𝔭​(Ξ±,u)=a},\mathcal{K}(a,\alpha,P)\subset\{u\in\mathbb{R}\times\mathbb{C}|\mathfrak{p}(\alpha,u)=a\},

    such that

    ∏uβˆˆπ’¦β€‹(a,Ξ±,P)(t​ðt+𝔒​(Ξ»,u))\prod_{u\in\mathcal{K}(a,\alpha,P)}(t\eth_{t}+\mathfrak{e}(\lambda,u))

    is nilpotent on GraVH​𝒩(Ξ±)\textup{Gr}^{V^{H}}_{a}\mathcal{N}^{(\alpha)}, where 𝔭​(Ξ±,βˆ™):ℝ×ℂ→ℝ\mathfrak{p}(\alpha,\bullet):\mathbb{R}\times\mathbb{C}\to\mathbb{R} and 𝔒​(Ξ±,βˆ™):ℝ×ℂ→ℂ\mathfrak{e}(\alpha,\bullet):\mathbb{R}\times\mathbb{C}\to\mathbb{C} are functions defined in [Moc07a, Β§2.1]. (They build a relation between the KMS-spectrum at Ξ»=0\lambda=0 and Ξ»=Ξ±\lambda=\alpha. See [Moc07a, Corollary 7.71] for the case on a tame harmonic bundle on the puncture disk.)

Remark 3.2.

The KM-filtration Vβˆ™HV^{H}_{\bullet} on 𝒩(Ξ±)\mathcal{N}^{(\alpha)} is actually unique if exists, [Moc11, Lemma 22.3.4].

Then, we consider the multi-indexed KM-filtration. Let XX be a complex manifold, with a reduced normal crossing divisor D=D1+…+DnD=D_{1}+...+D_{n}, with irreducible components DiD_{i}. For any

𝐚=(a1,…,an)βˆˆβ„n,\mathbf{a}=(a_{1},...,a_{n})\in\mathbb{R}^{n},

we denote

(3.1) π•πšD​ℛX⁣(βˆ—D)=∩iVaiDi​ℛX⁣(βˆ—D).\mathbf{V}_{\mathbf{a}}^{D}\mathscr{R}_{X(*D)}=\cap_{i}V_{a_{i}}^{D_{i}}\mathscr{R}_{X(*D)}.

For 𝐚=𝟎:=(0,…,0)\mathbf{a}=\mathbf{0}:=(0,...,0), π•πŸŽD​ℛX⁣(βˆ—D)\mathbf{V}_{\mathbf{0}}^{D}\mathscr{R}_{X(*D)} is a coherent sub-ring of β„›X\mathscr{R}_{X}, denoted by β„›(X,D)\mathscr{R}_{(X,D)}.

Let 𝒩\mathcal{N} be a coherent right β„›X⁣(βˆ—D)\mathscr{R}_{X(*D)}-module, which is strictly specializable with respect to all DiD_{i}. In particular, it is the case when β„³\mathcal{M}, a coherent β„›X\mathscr{R}_{X}-module, admits a graded polarizable mixed twistor π’Ÿ\mathscr{D}-module, and 𝒩=β„³(βˆ—D)\mathcal{N}=\mathcal{M}(*D).

Define a multi-indexed Kashiwara-Malgrange filtration with respect to DD by

π•πšD​𝒩(Ξ±)=∩VaiDi​𝒩(Ξ±),\mathbf{V}^{D}_{\mathbf{a}}\mathcal{N}^{(\alpha)}=\cap V^{D_{i}}_{a_{i}}\mathcal{N}^{(\alpha)},

for any 𝐚=(a1,…,an)βˆˆβ„n.\mathbf{a}=(a_{1},...,a_{n})\in\mathbb{R}^{n}. It is not hard to see that π•βˆ™D​𝒩(Ξ±)\mathbf{V}^{D}_{\bullet}\mathcal{N}^{(\alpha)} is a multi-indexed filtered module over the filtered ring π•βˆ™D​ℛX\mathbf{V}_{\bullet}^{D}\mathscr{R}_{X}, and their filtrations are compatible in the sense that

(3.2) π•πšD𝒩(Ξ±)⋅𝐕𝐛Dβ„›X(βˆ—D)βŠ‚π•πš+𝐛D𝒩(Ξ±),\mathbf{V}^{D}_{\mathbf{a}}\mathcal{N}^{(\alpha)}\cdot\mathbf{V}_{\mathbf{b}}^{D}\mathscr{R}_{X}(*D)\subset\mathbf{V}^{D}_{\mathbf{a}+\mathbf{b}}\mathcal{N}^{(\alpha)},

for any 𝐚,π›βˆˆβ„n.\mathbf{a},\mathbf{b}\in\mathbb{R}^{n}.

If we separate DD into two groups of components D=B+CD=B+C, we will also use 𝐕𝐛Bβ€‹π•πœC​𝒩(Ξ±):=𝐕𝐛B​𝒩(Ξ±)βˆ©π•πœC​𝒩(Ξ±).\mathbf{V}^{B}_{\mathbf{b}}\mathbf{V}^{C}_{\mathbf{c}}\mathcal{N}^{(\alpha)}:=\mathbf{V}^{B}_{\mathbf{b}}\mathcal{N}^{(\alpha)}\cap\mathbf{V}^{C}_{\mathbf{c}}\mathcal{N}^{(\alpha)}.

Recall that we say 𝐛<𝐚\mathbf{b}<\mathbf{a}, if bi<aib_{i}<a_{i} for all ii. We denote

(3.3) 𝐕<𝐚D​𝒩(Ξ±):=βˆͺ𝐛<πšπ•π›D​𝒩(Ξ±).\mathbf{V}^{D}_{<\mathbf{a}}\mathcal{N}^{(\alpha)}:=\cup_{\mathbf{b}<\mathbf{a}}\mathbf{V}^{D}_{\mathbf{b}}\mathcal{N}^{(\alpha)}.

In general, the nn filtrations Vβˆ™DiV^{D_{i}}_{\bullet} do not behave well between each other. This motivates us to make the following definition, which will be essentially used in the logarithmic comparison, Proposition 3.7.

Definition 3.3.

Notations as above, we say that π•βˆ™D\mathbf{V}^{D}_{\bullet}, the multi-indexed Kashiwara-Malgrange filtration with respect to DD on 𝒩(Ξ±)\mathcal{N}^{(\alpha)} is V-compatible, if we have the following strictness relation

(3.4) π•πšD​𝒩(Ξ±)β†’β‹…tiπ•πšβˆ’πŸiD​𝒩(Ξ±)\mathbf{V}^{D}_{\mathbf{a}}\mathcal{N}^{(\alpha)}\xrightarrow{\cdot t_{i}}\mathbf{V}^{D}_{\mathbf{a}-\mathbf{1}^{i}}\mathcal{N}^{(\alpha)}

are isomorphisms, for all πšβˆˆβ„n\mathbf{a}\in\mathbb{R}^{n}, where 𝟏i:=[0,…,0,1,0,…,0],\mathbf{1}^{i}:=[0,...,0,1,0,...,0], with the only 11 at the ii-th position. If locally for any Ξ±βˆˆβ„‚\alpha\in\mathbb{C}, π•βˆ™D​𝒩(Ξ±)\mathbf{V}^{D}_{\bullet}\mathcal{N}^{(\alpha)} is VV-compatible, we say that such β„›(βˆ—D)\mathscr{R}(*D)-module 𝒩\mathcal{N} is VV-compatible with respect to DD. If we further have 𝒩=β„³(βˆ—D)\mathcal{N}=\mathcal{M}(*D) for some coherent β„›X\mathscr{R}_{X}-module β„³\mathcal{M}, then we also say that β„³\mathcal{M} is VV-compatible with respect to DD.

Remark 3.4.

Let HH be a smooth component of DD, and denote Dβ€²=Dβˆ’HD^{\prime}=D-H. If we fix the Dβ€²D^{\prime} part of the multi-indexed KM-filtration, the HH part induces a filtration Vβˆ™HV^{H}_{\bullet} on π•πšβ€²D′​𝒩(Ξ±)\mathbf{V}^{D^{\prime}}_{\mathbf{a}^{\prime}}\mathcal{N}^{(\alpha)}, for a fixed πšβ€²βˆˆβ„nβˆ’1\mathbf{a}^{\prime}\in\mathbb{R}^{n-1}. Assuming the VV-compatibility, we can check that such an induced filtration satisfies all conditions in Definition 3.1, replacing β„›X⁣(βˆ—D)\mathscr{R}_{X(*D)} by π•πŸŽDβ€²β„›X(βˆ—D)\mathbf{V}^{D^{\prime}}_{\mathbf{0}}\mathscr{R}_{X}(*D) and it is unique if exits, following the same argument. We may still call it the KM-filtration on a π•πŸŽDβ€²β„›X(βˆ—D)\mathbf{V}^{D^{\prime}}_{\mathbf{0}}\mathscr{R}_{X}(*D)-module with respect to HH.

As in the filtered π’Ÿ\mathscr{D}-module case in [Wei20a, Lemma 12], we have the following

Lemma 3.5.

Let HH be a smooth component of DD and denote Dβ€²=Dβˆ’HD^{\prime}=D-H. For any right coherent β„›(X,D)\mathscr{R}_{(X,D)}-module 𝒱\mathcal{V}, such that it is tt-torsion free, for any local holomorphic function tt on XX that locally defines HH, we have

β„‹i​(π’±βŠ—β„›(X,D)𝐋ℛ(X,Dβ€²))=0,Β for all ​iβ‰ 0.\mathcal{H}^{i}(\mathcal{V}\otimes^{\mathbf{L}}_{\mathscr{R}_{(X,D)}}\mathscr{R}_{(X,D^{\prime})})=0,\text{ for all }i\neq 0.
Proof.

Working locally on XX, we can assume X=YΓ—β„‚tX=Y\times\mathbb{C}_{t}, with H=YΓ—0H=Y\times{0}, and Dβ€²=pYβˆ—β€‹DYD^{\prime}=p_{Y}^{*}D^{Y}, for some normal crossing DYβŠ‚YD^{Y}\subset Y. Then, locally we have

β„›(X,D)\displaystyle\mathscr{R}_{(X,D)} =β„›(Y,DY)β€‹βŸ¨t,t​ðt⟩,\displaystyle=\mathscr{R}_{(Y,D^{Y})}\left<t,t\eth_{t}\right>,
β„›(X,Dβ€²)\displaystyle\mathscr{R}_{(X,D^{\prime})} =β„›(Y,DY)β€‹βŸ¨t,Γ°t⟩.\displaystyle=\mathscr{R}_{(Y,D^{Y})}\left<t,\eth_{t}\right>.

We also consider β„›(X,D)β€‹βŸ¨ΞΎβŸ©\mathscr{R}_{(X,D)}\left<\xi\right>, which is, as a left free β„›(X,D)\mathscr{R}_{(X,D)}-module, isomorphic to the polynomial ring β„›(X,D)​[ΞΎ]\mathscr{R}_{(X,D)}[\xi]. It also possesses a right β„›(X,D)\mathscr{R}_{(X,D)}-module structure with the non-commutative relations

[ΞΎ,t]=Ξ»,[ΞΎ,t​ðt]=λ​ξ.[\xi,t]=\lambda,\ [\xi,t\eth_{t}]=\lambda\xi.

We have the following left β„›(X,D)\mathscr{R}_{(X,D)}-linear complex

(3.5) β„›(X,D)β€‹βŸ¨ΞΎβŸ©β†’β‹…(tβ€‹ΞΎβˆ’t​ðt)β„›(X,D)β€‹βŸ¨ΞΎβŸ©\displaystyle\mathscr{R}_{(X,D)}\left<\xi\right>\xrightarrow{\cdot(t\xi-t\eth_{t})}\mathscr{R}_{(X,D)}\left<\xi\right> β†’β„›(X,Dβ€²).\displaystyle\to\mathscr{R}_{(X,D^{\prime})}.
ΞΎ\displaystyle\xi ↦ðt\displaystyle\mapsto\eth_{t}

Let’s show that it actually is a short exact sequence. We give both ΞΎ,t​ðt\xi,t\eth_{t} and Γ°t\eth_{t} of grading 11, which induces an increasing filtration Fβˆ™F_{\bullet} on β„›(X,D)β€‹βŸ¨ΞΎβŸ©\mathscr{R}_{(X,D)}\left<\xi\right> and β„›(X,Dβ€²)\mathscr{R}_{(X,D^{\prime})}. Both maps in the complex are strict with respect to Fβˆ™F_{\bullet}, so to show the complex is exact, we only need to consider the induced complex of their associated graded pieces:

β„›(Y,DY)​[t,t​ðt,ΞΎ]β†’β‹…(tβ€‹ΞΎβˆ’t​ðt)β„›(Y,DY)​[t,t​ðt,ΞΎ]\displaystyle\mathscr{R}_{(Y,D^{Y})}[t,t\eth_{t},\xi]\xrightarrow{\cdot(t\xi-t\eth_{t})}\mathscr{R}_{(Y,D^{Y})}[t,t\eth_{t},\xi] β†’β„›(Y,DY)​[t,Γ°t],\displaystyle\to\mathscr{R}_{(Y,D^{Y})}[t,\eth_{t}],
ΞΎ\displaystyle\xi ↦ðt\displaystyle\mapsto\eth_{t}

which is exact by induction on the grading. Hence, (3.5) gives a free resolution of β„›(X,Dβ€²)\mathscr{R}_{(X,D^{\prime})} as a left β„›(X,D)\mathscr{R}_{(X,D)} module.

Now, we only need to argue that π’±β€‹βŸ¨ΞΎβŸ©β†’β‹…(tβ€‹ΞΎβˆ’t​ðt)π’±β€‹βŸ¨ΞΎβŸ©\mathcal{V}\left<\xi\right>\xrightarrow{\cdot(t\xi-t\eth_{t})}\mathcal{V}\left<\xi\right> is injective, which can be checked by considering the degree of ΞΎ\xi combined with the tt-torsion freeness of 𝒱\mathcal{V}. ∎

Proposition 3.6.

Let β„³\mathcal{M} be a strict coherent β„›X\mathscr{R}_{X}-module. Assume β„³(βˆ—D)\mathcal{M}(*D) is VV-compatible with respect to DD, and HH is a smooth component of DD. Denote Dβ€²=Dβˆ’HD^{\prime}=D-H. For any Ξ±βˆˆβ„‚Ξ»\alpha\in\mathbb{C}_{\lambda} and πšβ€²βˆˆβ„nβˆ’1\mathbf{a}^{\prime}\in\mathbb{R}^{n-1}, we have

π•πšβ€²Dβ€²V0Hβ„³(Ξ±)(βˆ—D)βŠ—β„›(X,D)𝐋ℛ(X,Dβ€²)\displaystyle\mathbf{V}^{D^{\prime}}_{\mathbf{a}^{\prime}}V^{H}_{0}\mathcal{M}^{(\alpha)}(*D)\otimes^{\mathbf{L}}_{\mathscr{R}_{(X,D)}}\mathscr{R}_{(X,D^{\prime})} β‰ƒπ•πšβ€²Dβ€²β„³(Ξ±)[βˆ—H](βˆ—Dβ€²);\displaystyle\simeq\mathbf{V}^{D^{\prime}}_{\mathbf{a}^{\prime}}\mathcal{M}^{(\alpha)}[*H](*D^{\prime});
π•πšβ€²Dβ€²V<0Hβ„³(Ξ±)(βˆ—D)βŠ—β„›(X,D)𝐋ℛ(X,Dβ€²)\displaystyle\mathbf{V}^{D^{\prime}}_{\mathbf{a}^{\prime}}V^{H}_{<0}\mathcal{M}^{(\alpha)}(*D)\otimes^{\mathbf{L}}_{\mathscr{R}_{(X,D)}}\mathscr{R}_{(X,D^{\prime})} β‰ƒπ•πšβ€²Dβ€²β„³(Ξ±)[!H](βˆ—Dβ€²).\displaystyle\simeq\mathbf{V}^{D^{\prime}}_{\mathbf{a}^{\prime}}\mathcal{M}^{(\alpha)}[!H](*D^{\prime}).
Proof.

Let’s first show that,

V0Hβ„³(Ξ±)(βˆ—D)βŠ—V0H​ℛX⁣(βˆ—D)𝐋ℛX⁣(βˆ—Dβ€²)\displaystyle V^{H}_{0}\mathcal{M}^{(\alpha)}(*D)\otimes^{\mathbf{L}}_{V^{H}_{0}\mathscr{R}_{X(*D)}}\mathscr{R}_{X(*D^{\prime})} ≃ℳ(Ξ±)[βˆ—H](βˆ—Dβ€²);\displaystyle\simeq\mathcal{M}^{(\alpha)}[*H](*D^{\prime});
V<0Hβ„³(Ξ±)(βˆ—D)βŠ—V0H​ℛX⁣(βˆ—D)𝐋ℛX⁣(βˆ—Dβ€²)\displaystyle V^{H}_{<0}\mathcal{M}^{(\alpha)}(*D)\otimes^{\mathbf{L}}_{V^{H}_{0}\mathscr{R}_{X(*D)}}\mathscr{R}_{X(*D^{\prime})} ≃ℳ(Ξ±)[!H](βˆ—Dβ€²).\displaystyle\simeq\mathcal{M}^{(\alpha)}[!H](*D^{\prime}).

These two identities are due to [Moc15, Lemma 3.1.2, Lemma 3.1.10], combining Lemma 3.5 above. From now on, we focus on the first identity in the statement of the proposition, since the second one follows similarly. Lemma 3.5 will be used repeatedly to show various tensor functors are exact, without being mentioned explicitly.

We want to show the following naturally induced map

(3.6) π•πšβ€²Dβ€²V0Hβ„³(Ξ±)(βˆ—D)βŠ—β„›(X,D)β„›(X,Dβ€²)β†’β„³(Ξ±)[βˆ—H](βˆ—Dβ€²)\mathbf{V}^{D^{\prime}}_{\mathbf{a}^{\prime}}V^{H}_{0}\mathcal{M}^{(\alpha)}(*D)\otimes_{\mathscr{R}_{(X,D)}}\mathscr{R}_{(X,D^{\prime})}\to\mathcal{M}^{(\alpha)}[*H](*D^{\prime})

is injective, for any πšβ€²βˆˆβ„nβˆ’1\mathbf{a}^{\prime}\in\mathbb{R}^{n-1}. Note that, using a similar resolution as (3.5), we have

β„³(Ξ±)[βˆ—H](βˆ—Dβ€²)\displaystyle\mathcal{M}^{(\alpha)}[*H](*D^{\prime}) ≃V0Hβ„³(Ξ±)(βˆ—D)βŠ—V0H​ℛX⁣(βˆ—D)β„›X⁣(βˆ—Dβ€²)\displaystyle\simeq V^{H}_{0}\mathcal{M}^{(\alpha)}(*D)\otimes_{V^{H}_{0}\mathscr{R}_{X(*D)}}\mathscr{R}_{X(*D^{\prime})}
≃V0Hβ„³(Ξ±)(βˆ—D)βŠ—β„›(X,D)β„›(X,Dβ€²).\displaystyle\simeq V^{H}_{0}\mathcal{M}^{(\alpha)}(*D)\otimes_{\mathscr{R}_{(X,D)}}\mathscr{R}_{(X,D^{\prime})}.

Now the injectivity of (3.6) can be deduced from the fact that the cokernel of the natural inclusion

π•πšβ€²Dβ€²V0Hβ„³(Ξ±)(βˆ—D)β†’V0Hβ„³(Ξ±)(βˆ—D)\mathbf{V}^{D^{\prime}}_{\mathbf{a}^{\prime}}V^{H}_{0}\mathcal{M}^{(\alpha)}(*D)\to V^{H}_{0}\mathcal{M}^{(\alpha)}(*D)

is tt-torsion free, due to the VV-compatibility of β„³\mathcal{M} in the assumption. In particular, we have the following injection

VaEV0Hβ„³(Ξ±)(βˆ—D)βŠ—β„›(X,D)β„›(X,Dβ€²)β†’β„³(Ξ±)[βˆ—H](βˆ—Dβ€²),V^{E}_{a}V^{H}_{0}\mathcal{M}^{(\alpha)}(*D)\otimes_{\mathscr{R}_{(X,D)}}\mathscr{R}_{(X,D^{\prime})}\to\mathcal{M}^{(\alpha)}[*H](*D^{\prime}),

for any component Eβ‰ HE\neq H of DD. It is straightforward to check that the filtration on β„³(Ξ±)[βˆ—H](βˆ—Dβ€²)\mathcal{M}^{(\alpha)}[*H](*D^{\prime}) induced by the image above satisfies all of those conditions in Definition 3.1, which means it gives the KM-filtration with respect to EE. This implies that the multi-indexed filtration induced by the images of (3.6) is indeed the multi-indexed KM-filtration of β„³(Ξ±)[βˆ—H]\mathcal{M}^{(\alpha)}[*H], with respect to Dβ€²D^{\prime}. It can also be argued directly by using Remark 3.4. ∎

Apply the previous proposition inductively, we get the following

Proposition 3.7 (Logarithmic Comparison).

With the same assumptions in the previous proposition, separate DD into two groups of components D=B+CD=B+C. For any Ξ±βˆˆβ„‚Ξ»\alpha\in\mathbb{C}_{\lambda}, we have

(3.7) 𝐕<𝟎Bπ•πŸŽCβ„³(Ξ±)(βˆ—D)βŠ—β„›(X,D)𝐋ℛX≃ℳ(Ξ±)[!B+βˆ—C].\mathbf{V}^{B}_{<\mathbf{0}}\mathbf{V}^{C}_{\mathbf{0}}\mathcal{M}^{(\alpha)}(*D)\otimes^{\mathbf{L}}_{\mathscr{R}_{(X,D)}}\mathscr{R}_{X}\simeq\mathcal{M}^{(\alpha)}[!B+*C].

In particular, we have

(3.8) DR(X,D)𝐕<𝟎Bπ•πŸŽCβ„³(Ξ±)(βˆ—D)≃DRXβ„³(Ξ±)[!B+βˆ—C].\textup{DR}_{(X,D)}\mathbf{V}^{B}_{<\mathbf{0}}\mathbf{V}^{C}_{\mathbf{0}}\mathcal{M}^{(\alpha)}(*D)\simeq\textup{DR}_{X}\mathcal{M}^{(\alpha)}[!B+*C].

Recall the de Rham functor on a right coherent β„›X\mathscr{R}_{X}-module β„³\mathcal{M}

DRX​ℳ=β„³βŠ—β„›X𝐋π’ͺ𝒳,\textup{DR}_{X}\mathcal{M}=\mathcal{M}\otimes^{\mathbf{L}}_{\mathscr{R}_{X}}\mathcal{O}_{\mathcal{X}},

which can be explicitly expressed using the Spencer complex Spβˆ™β€‹(π’ͺ𝒳)\textup{Sp}^{\bullet}(\mathcal{O}_{\mathcal{X}}), [Sab05, Β§0.6], as a resolution of π’ͺ𝒳\mathcal{O}_{\mathcal{X}}, by locally free left β„›X\mathscr{R}_{X}-modules. In the case of right β„›(X,D)\mathscr{R}_{(X,D)}-modules, we have the log-de Rham functor on a right coherent β„›(X,D)\mathscr{R}_{(X,D)}-module 𝒱\mathcal{V}:

DR(X,D)​𝒱=π’±βŠ—β„›(X,D)𝐋π’ͺ𝒳.\textup{DR}_{(X,D)}\mathcal{V}=\mathcal{V}\otimes^{\mathbf{L}}_{\mathscr{R}_{(X,D)}}\mathcal{O}_{\mathcal{X}}.

We can use the logarithmic Spencer complex in [Wei20a, Β§2], to get the explicit expression. Please note that the Sp functor in loc. cit. is the same as the de Rham functor here.

Let’s also show the following compatibility of the multi-indexed KM-filtration with respect to certain pushforward functor, which has essentially been proved in [Sab05, Theorem 3.1.8]. It will be used in the proof of Theorem 1.4.

Proposition 3.8.

Assume that f:Xβ†’Yf:X\to Y is a proper map between complex manifolds, and F=fΓ—Id:XΓ—β„‚nβ†’YΓ—β„‚nF=f\times\text{Id}:X\times\mathbb{C}^{n}\to Y\times\mathbb{C}^{n}. Let DXD^{X} (resp. DYD^{Y}) be a normal crossing divisor on XΓ—β„‚nX\times\mathbb{C}^{n} (resp. YΓ—β„‚nY\times\mathbb{C}^{n}), defined by those coordinates of β„‚n\mathbb{C}^{n}, and β„³\mathcal{M} a strict coherent β„›\mathscr{R}-module on XΓ—β„‚nX\times\mathbb{C}^{n}, and β„³(βˆ—DX)\mathcal{M}(*D^{X}) is VV-compatible with respect to DXD^{X}. We have the following compatibility of the direct image functor with multi-indexed KM-filtration:

(3.9) β„‹iF†(𝐕𝐝DXβ„³(Ξ±)(βˆ—DX))≃𝐕𝐝DYβ„‹iF†ℳ(Ξ±)(βˆ—DX),Β for anyΒ πβˆˆβ„n\mathcal{H}^{i}F_{\dagger}(\mathbf{V}^{D^{X}}_{\mathbf{d}}\mathcal{M}^{(\alpha)}(*D^{X}))\simeq\mathbf{V}^{D^{Y}}_{\mathbf{d}}\mathcal{H}^{i}F_{\dagger}\mathcal{M}^{(\alpha)}(*D^{X}),\text{ for any }\mathbf{d}\in\mathbb{R}^{n}

where the F†F_{\dagger} on the left shall be read as the direct image functor on a β„›(XΓ—β„‚n,DX)\mathscr{R}_{(X\times\mathbb{C}^{n},D^{X})}-module, [Wei20a, Β§2], see also [Sab05, Remark 1.4.3(2)].

Proof.

Although Sabbah’s theorem only states the case that DD has only one component, we can apply it inductively on the number of components of DD. To be more precise, the inductive assumption implies that

β„‹iF†(𝐕𝐝′D′⁣Xβ„³(Ξ±)(βˆ—DX))≃𝐕𝐝′D′⁣Yβ„‹iF†ℳ(Ξ±)(βˆ—DX),Β for anyΒ πβ€²βˆˆβ„nβˆ’1\mathcal{H}^{i}F_{\dagger}(\mathbf{V}^{D^{\prime X}}_{\mathbf{d}^{\prime}}\mathcal{M}^{(\alpha)}(*D^{X}))\simeq\mathbf{V}^{D^{\prime Y}}_{\mathbf{d}^{\prime}}\mathcal{H}^{i}F_{\dagger}\mathcal{M}^{(\alpha)}(*D^{X}),\text{ for any }\mathbf{d}^{\prime}\in\mathbb{R}^{n-1}

where D′⁣X=DXβˆ’DnXD^{\prime X}=D^{X}-D^{X}_{n}, D′⁣Y=DYβˆ’DnYD^{\prime Y}=D^{Y}-D^{Y}_{n}. Now, as in Remark 3.4, Vβˆ™DnV^{D_{n}}_{\bullet} gives the KM-filtration on 𝐕𝐝′D′⁣X​ℳ(Ξ±)\mathbf{V}^{D^{\prime X}}_{\mathbf{d}^{\prime}}\mathcal{M}^{(\alpha)}, hence we can apply Sabbah’s argument to get that β„‹i​F†​(Vβˆ™Dn​𝐕𝐝′D′⁣X​ℳ)\mathcal{H}^{i}F_{\dagger}(V^{D_{n}}_{\bullet}\mathbf{V}^{D^{\prime X}}_{\mathbf{d}^{\prime}}\mathcal{M}) gives the KM-filtration on 𝐕𝐝′D′⁣Y​ℋi​F†​ℳ(Ξ±)\mathbf{V}^{D^{\prime Y}}_{\mathbf{d}^{\prime}}\mathcal{H}^{i}F_{\dagger}\mathcal{M}^{(\alpha)} with respect to DnYD^{Y}_{n}. ∎

Please also refer to [Wei20a, Theorem 3], and the Remark after its proof. In that case, the map is more general, and we cannot reduce to the form of the map FF as above. The other results in [Wei20a] about the direct image and dual functors on log-representations shall still work in the setting of β„›\mathscr{R}-modules admitting graded polarizable mixed twistor π’Ÿ\mathscr{D}-modules, by essentially the same arguments. Since we are not using them in this paper, we do not copy them here.

Before we end this section, we consider an easy special case. Let XX be a complex manifold with a reduced normal crossing divisor DD. According to [Sab05, Β§3.7], assume a holonomic β„›X\mathscr{R}_{X}-module β„³\mathcal{M} is strictly non-characteristic with respect to all components of DD. Then, β„³(βˆ—D)\mathcal{M}(*D) is strictly specializable and VV-compatible, and the multi-indexed KM-filtration π•βˆ™D\mathbf{V}^{D}_{\bullet} on β„³(βˆ—D)\mathcal{M}(*D) is globally defined, i.e. does not depend on Ξ±\alpha, and π•βˆ™D\mathbf{V}^{D}_{\bullet} only jumps at β„€n\mathbb{Z}^{n}, with

π•πšDβ„³(βˆ—D)=β„³βŠ—π’ͺ𝒳((𝐚+𝟏)D),for anyΒ πšβˆˆβ„€n.\mathbf{V}^{D}_{\mathbf{a}}\mathcal{M}(*D)=\mathcal{M}\otimes\mathcal{O}_{\mathcal{X}}((\mathbf{a}+\mathbf{1})D),\text{for any }\mathbf{a}\in\mathbb{Z}^{n}.

Let HH be one component of DD and Dβ€²:=Dβˆ’HD^{\prime}:=D-H. Following [Moc15, Lemma 3.1.1], let’s first compute,

Vβˆ™Hβ„³(βˆ—Dβ€²)[βˆ—H]:=Vβˆ™Hβ„³(βˆ—D)βˆ©β„³(βˆ—Dβ€²)[βˆ—H].V^{H}_{\bullet}\mathcal{M}(*D^{\prime})[*H]:=V^{H}_{\bullet}\mathcal{M}(*D)\cap\mathcal{M}(*D^{\prime})[*H].

We claim that

(3.10) VkHβ„³(βˆ—Dβ€²)[βˆ—H]={β„³(βˆ—Dβ€²)βŠ—π’ͺ𝒳((k+1)H),if ​k≀0;β„³(βˆ—Dβ€²)βŠ—π’ͺ𝒳(H)+…+β„³(βˆ—Dβ€²)βŠ—π’ͺ𝒳((k+1)H)β‹…Ξ»k,if ​kβ‰₯1.V^{H}_{k}\mathcal{M}(*D^{\prime})[*H]=\begin{cases}\mathcal{M}(*D^{\prime})\otimes\mathcal{O}_{\mathcal{X}}((k+1)H),&\text{if }k\leq 0;\\ \mathcal{M}(*D^{\prime})\otimes\mathcal{O}_{\mathcal{X}}(H)+...\\ \ +\mathcal{M}(*D^{\prime})\otimes\mathcal{O}_{\mathcal{X}}((k+1)H)\cdot\lambda^{k},&\text{if }k\geq 1.\end{cases}

The k≀0k\leq 0 part is due to [Sab05, Lemma 3.7.4]. The kβ‰₯1k\geq 1 part is due to that

mtβˆ’kβ‹…Γ°t=(mΓ°t)tβˆ’k+m(βˆ’k)tβˆ’kβˆ’1Ξ»,βˆ€mβˆˆβ„³(βˆ—Dβ€²),mt^{-k}\cdot\eth_{t}=(m\eth_{t})t^{-k}+m(-k)t^{-k-1}\lambda,\forall m\in\mathcal{M}(*D^{\prime}),

which implies VkHβ„³(βˆ—Dβ€²)[βˆ—H]+VkHβ„³(βˆ—Dβ€²)[βˆ—H]β‹…Γ°t=Vk+1Hβ„³(βˆ—Dβ€²)[βˆ—H]V^{H}_{k}\mathcal{M}(*D^{\prime})[*H]+V^{H}_{k}\mathcal{M}(*D^{\prime})[*H]\cdot\eth_{t}=V^{H}_{k+1}\mathcal{M}(*D^{\prime})[*H], which is what we need. Set

π•πšDβ„³[βˆ—D]:=π•πšDβ„³(βˆ—D)βˆ©β„³[βˆ—D].\mathbf{V}^{D}_{\mathbf{a}}\mathcal{M}[*D]:=\mathbf{V}^{D}_{\mathbf{a}}\mathcal{M}(*D)\cap\mathcal{M}[*D].

By induction, we can get that it only jumps at β„€n\mathbb{Z}^{n}, and for any πšβˆˆβ„€n,\mathbf{a}\in\mathbb{Z}^{n},

(3.11) π•πšDβ„³[βˆ—D]=βˆ‘π€βˆˆβ„€n,π€β‰€πšβ„³βŠ—π’ͺ𝒳((𝐀+𝟏)D)β‹…Ξ»s𝐀,\mathbf{V}^{D}_{\mathbf{a}}\mathcal{M}[*D]=\sum_{\mathbf{k}\in\mathbb{Z}^{n},\mathbf{k}\leq\mathbf{a}}\mathcal{M}\otimes\mathcal{O}_{\mathcal{X}}((\mathbf{k}+\mathbf{1})D)\cdot\lambda^{s_{\mathbf{k}}},

where

s𝐀:=βˆ‘{i|kiβ‰₯1}ki.s_{\mathbf{k}}:=\sum_{\{i|k_{i}\geq 1\}}k_{i}.
Remark 3.9.

When β„³\mathcal{M} is algebraic and holonormic, it is non-characteristic with respect to a general hypersurface, due to Bertini’s theorem, e.g. [Moc15, Β§14.3.1.3]. If we further assume that it is strictly specializable, then it is strictly non-characteristic.

4. Esnault-Viehweg’s covering construction

On a smooth variety XX, assume we have a line bundle β„’\mathcal{L}, with a section s∈H0​(X,β„’βŠ—N)s\in H^{0}(X,\mathcal{L}^{\otimes N}) defining an integral divisor N​(𝐝​D)N(\mathbf{d}D), with D=D1+…+DnD=D_{1}+...+D_{n} reduced and normal crossing, πβˆˆβ„šn\mathbf{d}\in\mathbb{Q}^{n}, 0<di≀10<d_{i}\leq 1, and

gcd⁑(N​d1,…,N​dn,N)=1.\gcd(Nd_{1},...,Nd_{n},N)=1.

Applying the construction in [EV92, Β§3], the section ss gives a cyclic covering g:X^β†’X,g:\hat{X}\to X, that only ramifies along DD, with X^\hat{X} being smooth and irreducible, and the ramification number along DiD_{i} being Ngcd⁑(N​di,N)\frac{N}{\gcd(Nd_{i},N)}. If we denote di=diβ€²diβ€²β€²d_{i}=\frac{d^{\prime}_{i}}{d^{\prime\prime}_{i}} as reduced fractional, then the ramification number is just diβ€²β€²d^{\prime\prime}_{i}. Let’s use h:X^0β†’X0h:\hat{X}^{0}\to X^{0} to denote gg restricted over X0=Xβˆ–DX^{0}=X\setminus D, i.e. the Γ©tale part of gg. Fix an β„›X\mathscr{R}_{X}-module β„³\mathcal{M} that admits a graded polarizable mixed twistor π’Ÿ\mathscr{D}-module, assuming that β„³(βˆ—D)\mathcal{M}(*D) is VV-compatible with respect to DD, and denote β„³[βˆ—D]\mathcal{M}[*D], its prolongation along DD. Let β„³0=β„³|X0\mathcal{M}^{0}=\mathcal{M}|_{X^{0}}, and denote β„³^0=hβˆ—β€‹β„³0\hat{\mathcal{M}}^{0}=h^{*}\mathcal{M}^{0}. Denote β„³^[βˆ—D^]\hat{\mathcal{M}}[*\hat{D}], the prolongation of β„³^0\hat{\mathcal{M}}^{0} along D^\hat{D}, which also underlies a mixed twistor π’Ÿ\mathscr{D}-module.

Due to the projection formula for β„›\mathscr{R}-modules, [HTT08, Corollary 1.7.5] for the π’Ÿ\mathscr{D}-module case, and [Moc11, Lemma 22.7.1], on X0X^{0} we have

(4.1) h+​ℳ^0=h+​h+​ℳ0≃ℳ0βŠ—hβˆ—β€‹π’ͺ𝒳^0≃⨁0≀i<N(β„³0βŠ—pβˆ—β€‹β„’βˆ’(i,𝐝​D)|X0),h_{+}\hat{\mathcal{M}}^{0}=h_{+}h^{+}\mathcal{M}^{0}\simeq\mathcal{M}^{0}\otimes h_{*}\mathcal{O}_{\hat{\mathscr{X}}^{0}}\simeq\bigoplus_{0\leq i<N}(\mathcal{M}^{0}\otimes p^{*}\mathcal{L}^{-(i,\mathbf{d}D)}|_{X^{0}}),

where

β„’βˆ’(i,𝐝​D)=(β„’(i,𝐝​D))βˆ’1=(β„’i​(βˆ’βŒŠi​𝐝​DβŒ‹))βˆ’1,\mathcal{L}^{-(i,\mathbf{d}D)}=(\mathcal{L}^{(i,\mathbf{d}D)})^{-1}=(\mathcal{L}^{i}(-\lfloor i\mathbf{d}D\rfloor))^{-1},

as in [EV92, 3.1 Notation]. We will just use (i)(i) to replace (i,𝐝​D)(i,\mathbf{d}D), if 𝐝​D\mathbf{d}D is clear from the context. We will also use the notation

⟨x⟩=xβˆ’βŒŠxβŒ‹,\left<x\right>=x-\lfloor x\rfloor,

for the fractional part of xx.

Note that

hβˆ—β€‹π’ͺ𝒳^0≃⨁0≀i<aβ€²β€²pβˆ—β€‹β„’βˆ’(i)|X0h_{*}\mathcal{O}_{\hat{\mathscr{X}}^{0}}\simeq\bigoplus_{0\leq i<a^{\prime\prime}}p^{*}\mathcal{L}^{-(i)}|_{X^{0}}

shall also be viewed as a decomposition of left β„›\mathscr{R}-modules, that underlies mixed twistor π’Ÿ\mathscr{D}-modules, corresponds to the cyclic decomposition of hβˆ—β€‹β„‚X^0h_{*}\mathbb{C}_{\hat{X}^{0}}. We also note that the tensor products in (4.1) are between a right β„›\mathscr{R}-module and a left β„›\mathscr{R}-module, and such gives us a right β„›\mathscr{R}-module, as the π’Ÿ\mathscr{D}-module case in [HTT08, Proposition 1.2.9. (ii)].

Fix Ξ΄\delta such that 0≀δ≀N0\leq\delta\leq N. Using the notations in [Moc07a, 6.1.5], locally around a general point of DiD_{i}, pβˆ—β€‹β„’βˆ’(Ξ΄)|X0p^{*}\mathcal{L}^{-(\delta)}|_{X^{0}} underlies the tame harmonic bundle q1βˆ—β€‹L​(βˆ’βŸ¨Ξ΄β€‹di⟩,0)q_{1}^{*}L(-\left<\delta d_{i}\right>,0). See also [EV92, 3.16 Lemma c)]. In particular, after we change it into its corresponding right β„›\mathscr{R}-module, [Moc07b, 14.1.2], we have

VaiDi(Ο‰π’³βŠ—pβˆ—β„’βˆ’(Ξ΄)(βˆ—Di))=Ο‰π’³βŠ—pβˆ—β„’βˆ’(Ξ΄)(nDi),V^{D_{i}}_{a_{i}}(\omega_{\mathscr{X}}\otimes p^{*}\mathcal{L}^{-(\delta)}(*D_{i}))=\omega_{\mathscr{X}}\otimes p^{*}\mathcal{L}^{-(\delta)}(nD_{i}),

if βˆ’βŸ¨Ξ΄β€‹di⟩+nβˆ’1≀ai<βˆ’βŸ¨Ξ΄β€‹di⟩+n.-\left<\delta d_{i}\right>+n-1\leq a_{i}<-\left<\delta d_{i}\right>+n. Recall that ω𝒳:=Ξ»βˆ’nβ‹…pβˆ—β€‹Ο‰X,\omega_{\mathscr{X}}:=\lambda^{-n}\cdot p^{*}\omega_{X}, e.g. [Moc07b, Example 14.4].

Denote i:X0β†’X,i:X^{0}\to X, the natural embedding, and denote

β„³Ξ΄=i+​(β„³0βŠ—pβˆ—β€‹β„’βˆ’(Ξ΄)|X0),\mathcal{M}_{\delta}=i_{+}(\mathcal{M}^{0}\otimes p^{*}\mathcal{L}^{-(\delta)}|_{X^{0}}),

which is an β„›\mathscr{R}-module that underlies a mixed twistor π’Ÿ\mathscr{D}-module on XX, satisfying β„³Ξ΄=β„³Ξ΄[βˆ—D].\mathcal{M}_{\delta}=\mathcal{M}_{\delta}[*D]. Take prolongation of (4.1) along DD, we have

g+β„³^[βˆ—D^]≃⨁0≀i<Nβ„³i.g_{+}\hat{\mathcal{M}}[*\hat{D}]\simeq\bigoplus_{0\leq i<N}\mathcal{M}_{i}.

By using the functoriality for the tensor product in [Moc07b, 7.2.6], we have

VaiDiβ„³Ξ΄(Ξ±)(βˆ—D)=\displaystyle V^{D_{i}}_{a_{i}}\mathcal{M}^{(\alpha)}_{\delta}(*D)= VβŸ¨Ξ΄β€‹di⟩+aiDiβ„³(Ξ±)(βˆ—D)βŠ—Vβˆ’βŸ¨Ξ΄β€‹diβŸ©βˆ’1Di(Ο‰π’³βŠ—pβˆ—β„’βˆ’(Ξ΄)(βˆ—D))βŠ—Ο‰π’³βˆ’1\displaystyle V^{D_{i}}_{\left<\delta d_{i}\right>+a_{i}}\mathcal{M}^{(\alpha)}(*D)\otimes V^{D_{i}}_{-\left<\delta d_{i}\right>-1}(\omega_{\mathscr{X}}\otimes p^{*}\mathcal{L}^{-(\delta)}(*D))\otimes\omega^{-1}_{\mathscr{X}}
≃\displaystyle\simeq VβŸ¨Ξ΄β€‹di⟩+aiDiβ„³(Ξ±)(βˆ—D)βŠ—pβˆ—β„’βˆ’(Ξ΄)(βˆ—(Dβˆ’Di)).\displaystyle V^{D_{i}}_{\left<\delta d_{i}\right>+a_{i}}\mathcal{M}^{(\alpha)}(*D)\otimes p^{*}\mathcal{L}^{-(\delta)}(*(D-D_{i})).

The shifting of degree 11 in the KM-filtration is due to the convention of the shifting from the parabolic structure to the KM-filtration as in [Moc07b, 15.1.2.]. Twisting Ο‰π’³βˆ’1\omega^{-1}_{\mathscr{X}} at the end is to make the tensor product works between right and left modules β„›\mathscr{R}-modules. By taking intersections of those KM-filtrations respect to all components of DD, we get

(4.2) π•πšDβ„³Ξ΄(Ξ±)(βˆ—D)=\displaystyle\mathbf{V}^{D}_{\mathbf{a}}\mathcal{M}^{(\alpha)}_{\delta}(*D)= π•βŸ¨Ξ΄β€‹πβŸ©+𝐚Dβ„³(Ξ±)(βˆ—D)βŠ—π•βˆ’βŸ¨Ξ΄β€‹πβŸ©βˆ’πŸD(Ο‰π’³βŠ—pβˆ—β„’βˆ’(Ξ΄)(βˆ—D))βŠ—Ο‰π’³βˆ’1\displaystyle\mathbf{V}^{D}_{\left<\delta\mathbf{d}\right>+\mathbf{a}}\mathcal{M}^{(\alpha)}(*D)\otimes\mathbf{V}^{D}_{-\left<\delta\mathbf{d}\right>-\mathbf{1}}(\omega_{\mathscr{X}}\otimes p^{*}\mathcal{L}^{-(\delta)}(*D))\otimes\omega^{-1}_{\mathscr{X}}
≃\displaystyle\simeq π•βŸ¨Ξ΄β€‹πβŸ©+𝐚Dβ„³(Ξ±)(βˆ—D)βŠ—pβˆ—β„’βˆ’(Ξ΄).\displaystyle\mathbf{V}^{D}_{\left<\delta\mathbf{d}\right>+\mathbf{a}}\mathcal{M}^{(\alpha)}(*D)\otimes p^{*}\mathcal{L}^{-(\delta)}.

In particular, we have the following

Lemma 4.1.

In the above setting, β„³Ξ΄=β„³Ξ΄[βˆ—D]\mathcal{M}_{\delta}=\mathcal{M}_{\delta}[*D], and β„³Ξ΄(βˆ—D)\mathcal{M}_{\delta}(*D) is VV-compatible with respect to DD. For any Ξ±βˆˆβ„‚Ξ»,\alpha\in\mathbb{C}_{\lambda},

𝐕δ​𝐝Dβ„³(Ξ±)(βˆ—D)βŠ—pβˆ—β„’βˆ’Ξ΄β‰ƒπ•βŸ¨Ξ΄β€‹πβŸ©Dβ„³(Ξ±)(βˆ—D)βŠ—pβˆ—β„’βˆ’(Ξ΄)β‰ƒπ•πŸŽDβ„³Ξ΄(Ξ±)(βˆ—D),\mathbf{V}^{D}_{\delta\mathbf{d}}\mathcal{M}^{(\alpha)}(*D)\otimes p^{*}\mathcal{L}^{-\delta}\simeq\mathbf{V}^{D}_{\left<\delta\mathbf{d}\right>}\mathcal{M}^{(\alpha)}(*D)\otimes p^{*}\mathcal{L}^{-(\delta)}\simeq\mathbf{V}^{D}_{\mathbf{0}}\mathcal{M}^{(\alpha)}_{\delta}(*D),

and

𝐕<δ​𝐝Dβ„³(Ξ±)(βˆ—D)βŠ—pβˆ—β„’βˆ’Ξ΄β‰ƒπ•<βŸ¨Ξ΄β€‹πβŸ©Dβ„³(Ξ±)(βˆ—D)βŠ—pβˆ—β„’βˆ’(Ξ΄)≃𝐕<𝟎Dβ„³Ξ΄(Ξ±)(βˆ—D),\mathbf{V}^{D}_{<\delta\mathbf{d}}\mathcal{M}^{(\alpha)}(*D)\otimes p^{*}\mathcal{L}^{-\delta}\simeq\mathbf{V}^{D}_{<\left<\delta\mathbf{d}\right>}\mathcal{M}^{(\alpha)}(*D)\otimes p^{*}\mathcal{L}^{-(\delta)}\simeq\mathbf{V}^{D}_{<\mathbf{0}}\mathcal{M}^{(\alpha)}_{\delta}(*D),

as β„›(X,D)\mathscr{R}_{(X,D)}-modules.

We can also generalize an intermediate result in Saito’s proof of Saito vanishing [Sai90, (2.33.3)], which will be used in the proof of Theorem 1.3.

Lemma 4.2.

In the above setting, if we further assume that the β„›X\mathscr{R}_{X}-module β„³\mathcal{M} is non-characteristic with respect to all components of DD, then we have that, for any 1≀δ≀N1\leq\delta\leq N, β„³Ξ΄=β„³Ξ΄[βˆ—D]=β„³Ξ΄[!D]\mathcal{M}_{\delta}=\mathcal{M}_{\delta}[*D]=\mathcal{M}_{\delta}[!D]. Furthermore, we have

ΞDolβ„³Ξ΄β‰ƒΞžDolβ„³[βˆ—D]βŠ—β„’βˆ’(Ξ΄)β‰ƒΞžDolβ„³[!D]βŠ—β„’βˆ’(Ξ΄)(D),\Xi_{\textup{Dol}}\mathcal{M}_{\delta}\simeq\Xi_{\textup{Dol}}\mathcal{M}[*D]\otimes\mathcal{L}^{-(\delta)}\simeq\Xi_{\textup{Dol}}\mathcal{M}[!D]\otimes\mathcal{L}^{-(\delta)}(D),

as π’œX\mathcal{A}_{X}-modules, where β„’βˆ’(Ξ΄)\mathcal{L}^{-(\delta)} and β„’βˆ’(Ξ΄)​(D)\mathcal{L}^{-(\delta)}(D) carry trivial higgs structure, i.e. differential operators act trivially on them.

Proof.

We have

(4.3) π•πŸŽDβ„³(βˆ—D)βŠ—pβˆ—β„’βˆ’(Ξ΄)=π•πŸŽ+βŸ¨Ξ΄β€‹πβŸ©Dβ„³(βˆ—D)βŠ—pβˆ—β„’βˆ’(Ξ΄)β‰ƒπ•πŸŽDβ„³Ξ΄(βˆ—D),\mathbf{V}^{D}_{\mathbf{0}}\mathcal{M}(*D)\otimes p^{*}\mathcal{L}^{-(\delta)}=\mathbf{V}^{D}_{\mathbf{0}+\left<\delta\mathbf{d}\right>}\mathcal{M}(*D)\otimes p^{*}\mathcal{L}^{-(\delta)}\simeq\mathbf{V}^{D}_{\mathbf{0}}\mathcal{M}_{\delta}(*D),

where the first identity is due to β„³\mathcal{M} being non-characteristic with respect to DD, and the second one is due to the previous lemma. We also have

π•πŸŽDβ„³(βˆ—D)βŠ—pβˆ—β„’βˆ’(Ξ΄)=𝐕<𝟎+βŸ¨Ξ΄β€‹πβŸ©Dβ„³(βˆ—D)βŠ—pβˆ—β„’βˆ’(Ξ΄)≃𝐕<𝟎Dβ„³Ξ΄(βˆ—D),\mathbf{V}^{D}_{\mathbf{0}}\mathcal{M}(*D)\otimes p^{*}\mathcal{L}^{-(\delta)}=\mathbf{V}^{D}_{<\mathbf{0}+\left<\delta\mathbf{d}\right>}\mathcal{M}(*D)\otimes p^{*}\mathcal{L}^{-(\delta)}\simeq\mathbf{V}^{D}_{<\mathbf{0}}\mathcal{M}_{\delta}(*D),

which implies π•πŸŽDβ„³Ξ΄(βˆ—D)≃𝐕<𝟎Dβ„³Ξ΄(βˆ—D).\mathbf{V}^{D}_{\mathbf{0}}\mathcal{M}_{\delta}(*D)\simeq\mathbf{V}^{D}_{<\mathbf{0}}\mathcal{M}_{\delta}(*D). Hence, the first statement holds due to the logarithmic comparison Proposition 3.7, since we only need to compare the π•πŸŽD\mathbf{V}^{D}_{\mathbf{0}} and 𝐕<𝟎D\mathbf{V}^{D}_{<\mathbf{0}} part.

For the second statement, according to the previous computation, we note that

ΞDolπ•πŸŽDβ„³Ξ΄=ΞDolπ•πŸŽDβ„³(βˆ—D)βŠ—β„’βˆ’(Ξ΄)=ΞDol𝐕<𝟎Dβ„³(βˆ—D)βŠ—β„’βˆ’(Ξ΄)(D).\Xi_{\textup{Dol}}\mathbf{V}^{D}_{\mathbf{0}}\mathcal{M}_{\delta}=\Xi_{\textup{Dol}}\mathbf{V}^{D}_{\mathbf{0}}\mathcal{M}(*D)\otimes\mathcal{L}^{-(\delta)}=\Xi_{\textup{Dol}}\mathbf{V}^{D}_{<\mathbf{0}}\mathcal{M}(*D)\otimes\mathcal{L}^{-(\delta)}(D).

Due to the logarithmic comparison again and use the explicit resolution (3.5), ΞDolβ„³Ξ΄,ΞDolβ„³[βˆ—D]βŠ—β„’βˆ’(Ξ΄)\Xi_{\textup{Dol}}\mathcal{M}_{\delta},\Xi_{\textup{Dol}}\mathcal{M}[*D]\otimes\mathcal{L}^{-(\delta)} and ΞDolβ„³[!D]βŠ—β„’βˆ’(Ξ΄)(D),\Xi_{\textup{Dol}}\mathcal{M}[!D]\otimes\mathcal{L}^{-(\delta)}(D), can be computed from the terms in the previous identities respectively, using a same functor. ∎

Remark 4.3.

Actually, we have

ℳδ≃ℳ[βˆ—D]βŠ—pβˆ—β„’βˆ’(Ξ΄).\mathcal{M}_{\delta}\simeq\mathcal{M}[*D]\otimes p^{*}\mathcal{L}^{-(\delta)}.

Note that β„³[βˆ—D]βŠ—pβˆ—β„’βˆ’(Ξ΄)\mathcal{M}[*D]\otimes p^{*}\mathcal{L}^{-(\delta)} is a priori just a sub-β„›(X,D)\mathscr{R}_{(X,D)}-module of β„³Ξ΄(βˆ—D)\mathcal{M}_{\delta}(*D), but we will see from the computation below that it is indeed closed under the action of β„›X\mathscr{R}_{X}. However, we cannot identify them to β„³[!D]βŠ—pβˆ—β„’βˆ’(Ξ΄)(D)\mathcal{M}[!D]\otimes p^{*}\mathcal{L}^{-(\delta)}(D) naively, since it does not carry a natural β„›X\mathscr{R}_{X}-module structure.

Let’s compute π•βˆ™Dβ„³Ξ΄:=π•βˆ™Dβ„³Ξ΄(βˆ—D)βˆ©β„³Ξ΄\mathbf{V}^{D}_{\bullet}\mathcal{M}_{\delta}:=\mathbf{V}^{D}_{\bullet}\mathcal{M}_{\delta}(*D)\cap\mathcal{M}_{\delta}. We first note that, due to (4.2), π•βˆ™D​ℳδ\mathbf{V}^{D}_{\bullet}\mathcal{M}_{\delta} only jumps at β„€n+βŸ¨Ξ΄β€‹πβŸ©\mathbb{Z}^{n}+\left<\delta\mathbf{d}\right>. Recall (3.11) is just the special case when Ξ΄=0\delta=0. We claim that

(4.4) π•πš+βŸ¨Ξ΄β€‹πβŸ©Dβ„³[βˆ—D]=βˆ‘π€β‰€πšβ„³βŠ—pβˆ—β„’βˆ’(Ξ΄)βŠ—π’ͺ𝒳((𝐀+𝟏)D)β‹…Ξ»s𝐀.\mathbf{V}^{D}_{\mathbf{a}+\left<\delta\mathbf{d}\right>}\mathcal{M}[*D]=\sum_{\mathbf{k}\leq\mathbf{a}}\mathcal{M}\otimes p^{*}\mathcal{L}^{-(\delta)}\otimes\mathcal{O}_{\mathcal{X}}((\mathbf{k}+\mathbf{1})D)\cdot\lambda^{s_{\mathbf{k}}}.

Similarly, we just need to compute it component by component, which is a more general version of (3.10). Let H=D1H=D_{1} a component of DD, Vβˆ™Hβ„³Ξ΄:=Vβˆ™Hβ„³Ξ΄(βˆ—D)βˆ©β„³Ξ΄V^{H}_{\bullet}\mathcal{M}_{\delta}:=V^{H}_{\bullet}\mathcal{M}_{\delta}(*D)\cap\mathcal{M}_{\delta}. We already know that it only jumps at β„€+βŸ¨Ξ΄β€‹d1⟩\mathbb{Z}+\left<\delta d_{1}\right>. We want to show

(4.5) Vk+βŸ¨Ξ΄β€‹d1⟩H​ℳδ={β„³(βˆ—Dβ€²)βŠ—pβˆ—β„’βˆ’(Ξ΄)βŠ—π’ͺ𝒳((k+1)H),if ​k≀0;β„³(βˆ—Dβ€²)βŠ—pβˆ—β„’βˆ’(Ξ΄)βŠ—π’ͺ𝒳(H)+…+β„³(βˆ—Dβ€²)βŠ—pβˆ—β„’βˆ’(Ξ΄)βŠ—π’ͺ𝒳((k+1)H)β‹…Ξ»k,if ​kβ‰₯1.V^{H}_{k+\left<\delta d_{1}\right>}\mathcal{M}_{\delta}=\begin{cases}\mathcal{M}(*D^{\prime})\otimes p^{*}\mathcal{L}^{-(\delta)}\otimes\mathcal{O}_{\mathcal{X}}((k+1)H),&\text{if }k\leq 0;\\ \mathcal{M}(*D^{\prime})\otimes p^{*}\mathcal{L}^{-(\delta)}\otimes\mathcal{O}_{\mathcal{X}}(H)+...\\ \ +\mathcal{M}(*D^{\prime})\otimes p^{*}\mathcal{L}^{-(\delta)}\otimes\mathcal{O}_{\mathcal{X}}((k+1)H)\cdot\lambda^{k},&\text{if }k\geq 1.\end{cases}

The i≀0i\leq 0 part is due to [Sab05, Lemma 3.7.4] and (4.2). The iβ‰₯1i\geq 1 part is due to that, for any local sections mβˆˆβ„³(βˆ—Dβ€²),l∈pβˆ—β„’βˆ’(Ξ΄)m\in\mathcal{M}(*D^{\prime}),l\in p^{*}\mathcal{L}^{-(\delta)},

(mβŠ—l)​tβˆ’kβ‹…Γ°t\displaystyle(m\otimes l)t^{-k}\cdot\eth_{t} =(m​ðtβŠ—l)​tβˆ’kβˆ’(mβŠ—(t​ðt​l))​tβˆ’kβˆ’1+mβŠ—l​(βˆ’k)​tβˆ’kβˆ’1​λ,\displaystyle=(m\eth_{t}\otimes l)t^{-k}-(m\otimes(t\eth_{t}l))t^{-k-1}+m\otimes l(-k)t^{-k-1}\lambda,
=(m​ðtβŠ—l)​tβˆ’kβˆ’(mβŠ—(βŸ¨Ξ΄β€‹d1βŸ©β€‹Ξ»β€‹l))​tβˆ’kβˆ’1+mβŠ—l​(βˆ’k)​tβˆ’kβˆ’1​λ,\displaystyle=(m\eth_{t}\otimes l)t^{-k}-(m\otimes(\left<\delta d_{1}\right>\lambda l))t^{-k-1}+m\otimes l(-k)t^{-k-1}\lambda,
=(m​ðtβŠ—l)​tβˆ’kβˆ’mβŠ—l​(βŸ¨Ξ΄β€‹d1⟩+k)​tβˆ’kβˆ’1​λ.\displaystyle=(m\eth_{t}\otimes l)t^{-k}-m\otimes l(\left<\delta d_{1}\right>+k)t^{-k-1}\lambda.

This implies Vk+βŸ¨Ξ΄β€‹d1⟩H​ℳδ+Vk+βŸ¨Ξ΄β€‹d1⟩H​ℳδ⋅ðt=Vk+1+βŸ¨Ξ΄β€‹d1⟩H​ℳδV^{H}_{k+\left<\delta d_{1}\right>}\mathcal{M}_{\delta}+V^{H}_{k+\left<\delta d_{1}\right>}\mathcal{M}_{\delta}\cdot\eth_{t}=V^{H}_{k+1+\left<\delta d_{1}\right>}\mathcal{M}_{\delta}, which is what we need.

Comparing (4.4) with (3.11), we get 𝐕𝐀+βŸ¨Ξ΄β€‹πβŸ©Dℳδ≃𝐕𝐀Dβ„³[βˆ—D]βŠ—pβˆ—β„’βˆ’(Ξ΄)\mathbf{V}^{D}_{\mathbf{k}+\left<\delta\mathbf{d}\right>}\mathcal{M}_{\delta}\simeq\mathbf{V}^{D}_{\mathbf{k}}\mathcal{M}[*D]\otimes p^{*}\mathcal{L}^{-(\delta)}.

5. Proofs of main theorems

Although the proof of Theorem 1.3 is mainly motivated from Saito’s proof of Saito-Kodaira vanishing, we give it a more geometric and functorial treatment and hopefully provide a clearer picture. Later, we will find that it is also helpful to prove the relative version.

Fix XX, a smooth algebraic variety of dimension nn, with a normal crossing reduced divisor DD, and a line bundle π’œ\mathcal{A}. We also fix an algebraic coherent β„›X\mathscr{R}_{X}-module β„³\mathcal{M}, admitting a mixed twistor π’Ÿ\mathscr{D}-module, that is VV-compatible with respect to DD. We say a sequence of smooth varieties X=Y0βŠƒY1βŠƒY2βŠƒβ€¦βŠƒYnX=Y_{0}\supset Y_{1}\supset Y_{2}\supset...\supset Y_{n}, with YjY_{j} of codimension jj in XX or Yj=βˆ…Y_{j}=\emptyset, is a filtration induced by π’œ\mathcal{A}, if π’ͺYjβˆ’1​(Yj)β‰ƒπ’œ|Yjβˆ’1\mathcal{O}_{Y_{j-1}}(Y_{j})\simeq\mathcal{A}|_{Y_{j-1}}. Note that, once π’œ|Yj\mathcal{A}|_{Y_{j}} is a trivial line bundle, then Yj+1=βˆ…,Y_{j+1}=\emptyset, so are the successive terms.

Such a filtration is called normal crossing, (with respect to DD,) if D|YjD|_{Y_{j}} is still a normal crossing reduced divisor on YjY_{j}, for all jj. Such a filtration is called non-characteristic, (with respect to β„³\mathcal{M},) if YjY_{j} is non-characteristic with respect to β„³|Yjβˆ’1\mathcal{M}|_{Y_{j-1}}, for all jj. Note that strict specializablity is part of the assumption of mixed twistor π’Ÿ\mathscr{D}-module, hence being non-characteristic automatically implies being strictly non-characteristic. If π’œ\mathcal{A} is base point free, then due to Remark 3.9, we know that, for a generic filtration induced by π’œ\mathcal{A}, it is both normal crossing and non-characteristic (NCNC). If we fix a NCNC filtration, it induces an exact sequence of β„›X\mathscr{R}_{X}-modules

0β†’β„³β†’β„³[βˆ—Y1]β†’i+(β„³|Y1[βˆ—Y2])→…→i+(β„³Ynβˆ’1[βˆ—Yn])β†’i+β„³|Ynβ†’0.0\to\mathcal{M}\to\mathcal{M}[*Y_{1}]\to i_{+}(\mathcal{M}|_{Y_{1}}[*Y_{2}])\to...\to i_{+}(\mathcal{M}_{Y_{n-1}}[*Y_{n}])\to i_{+}\mathcal{M}|_{Y_{n}}\to 0.

Once Yj+1=βˆ…,Y_{j+1}=\emptyset, all the successive terms are just 0 by default. This can be checked directly using [Moc15, Lemma 3.1.23], and the explicit computation (3.11).

Dually, using [Moc15, Lemma 3.1.24], we can similar consider the exact sequence of β„›X\mathscr{R}_{X}-modules

0β†’i+β„³|Ynβ†’i+(β„³Ynβˆ’1[!Yn])→…→i+(β„³|Y1[!Y2])β†’β„³[!Y1]β†’β„³β†’0.0\to i_{+}\mathcal{M}|_{Y_{n}}\to i_{+}(\mathcal{M}_{Y_{n-1}}[!Y_{n}])\to...\to i_{+}(\mathcal{M}|_{Y_{1}}[!Y_{2}])\to\mathcal{M}[!Y_{1}]\to\mathcal{M}\to 0.
Proof for Theorem 1.3.

We only prove the second vanishing here. Saito’s proof also focus on this case, and readers can compare these two approaches. The first vanishing can be derived using the dual construction with little extra effort. Please note that we cannot directly apply Grothendieck-Serre duality here, since we are not working in the derived category of coherent sheaves.

We first set π’œ=β„’m\mathcal{A}=\mathcal{L}^{m}, for some integer mm such that it is base point free. According to the previous construction, we get an exact sequence of β„›X\mathscr{R}_{X}-modules

(5.1) 0β†’β„³[!D]β†’\displaystyle 0\to\mathcal{M}[!D]\to β„³[!D+βˆ—Y1]β†’i+(β„³[!D]|Y1[βˆ—Y2])→…\displaystyle\mathcal{M}[!D+*Y_{1}]\to i_{+}(\mathcal{M}[!D]|_{Y_{1}}[*Y_{2}])\to...
β†’i+(β„³[!D]|Ynβˆ’1[βˆ—Yn])β†’i+β„³[!D]|Ynβ†’0.\displaystyle\to i_{+}(\mathcal{M}[!D]|_{Y_{n-1}}[*Y_{n}])\to i_{+}\mathcal{M}[!D]|_{Y_{n}}\to 0.

Apply ΞDol\Xi_{\textup{Dol}} on each of them, we get an exact sequence of Higgs sheaves. In particular, the Higgs complex is π’ͺX\mathcal{O}_{X}-linear. Hence, we can twist the complex by β„’βˆ’1\mathcal{L}^{-1} and keep the exactness. If we view β„’βˆ’1\mathcal{L}^{-1} as a Higgs bundle with trivial Higgs connection, then the next complex we get is actually an exact sequence of Higgs sheaves:

(5.2) 0β†’ΞžDolβ„³[!D]βŠ—β„’βˆ’1β†’β„°0β†’β„°1→…→ℰnβ†’0,0\to\Xi_{\textup{Dol}}\mathcal{M}[!D]\otimes\mathcal{L}^{-1}\to\mathcal{E}^{0}\to\mathcal{E}^{1}\to...\to\mathcal{E}^{n}\to 0,

with

β„°j=ΞDol(i+β„³[!D]|Yj[βˆ—Yj+1])βŠ—β„’βˆ’1=ΞDoli+(β„³[!D]|Yj[βˆ—Yj+1]βŠ—pβˆ—β„’βˆ’1|Yj).\mathcal{E}^{j}=\Xi_{\textup{Dol}}(i_{+}\mathcal{M}[!D]|_{Y_{j}}[*Y_{j+1}])\otimes\mathcal{L}^{-1}=\Xi_{\textup{Dol}}i_{+}(\mathcal{M}[!D]|_{Y_{j}}[*Y_{j+1}]\otimes p^{*}\mathcal{L}^{-1}|_{Y_{j}}).

Due to Lemma 4.2, we get that β„°j\mathcal{E}^{j} is isomorphic to ΞDoli+𝒩Yj[!D|Yj]\Xi_{\textup{Dol}}i_{+}\mathcal{N}_{Y_{j}}[!D|_{Y_{j}}], where 𝒩Yj\mathcal{N}_{Y_{j}} a β„›Yj\mathscr{R}_{Y_{j}}-module admitting a mixed twistor π’Ÿ\mathscr{D}-module, satisfying

𝒩Yj[!D|Yj]=𝒩Yj[!D|Yj][!Yj+1]=𝒩Yj[!D|Yj][βˆ—Yj+1].\mathcal{N}_{Y_{j}}[!D|_{Y_{j}}]=\mathcal{N}_{Y_{j}}[!D|_{Y_{j}}][!Y_{j+1}]=\mathcal{N}_{Y_{j}}[!D|_{Y_{j}}][*Y_{j+1}].

Due to the assumption that ℒ​(d​D)\mathcal{L}(dD) is ample, hence so is π’œ|Yj​(d​D|Yj)=π’ͺYj​(d​D|Yj+Yj+1)\mathcal{A}|_{Y_{j}}(dD|_{Y_{j}})=\mathcal{O}_{Y_{j}}(dD|_{Y_{j}}+Y_{j+1}), which implies that Zj:=Yjβˆ–(d​D|Yj+Yj+1)r​e​dZ_{j}:=Y_{j}\setminus(dD|_{Y_{j}}+Y_{j+1})_{red} is affine.

By Artin’s vanishing, we have

ℍk(iβˆ—DRYjΞDR𝒩Yj[!D|Yj])=ℍck(DRZjΞD​R𝒩Yj[!D|Yj]|Zj)=0,Β forΒ k<0.\mathbb{H}^{k}(i_{*}\textup{DR}_{Y_{j}}\Xi_{\textup{DR}}\mathcal{N}_{Y_{j}}[!D|_{Y_{j}}])=\mathbb{H}^{k}_{c}(\textup{DR}_{Z_{j}}\Xi_{DR}\mathcal{N}_{Y_{j}}[!D|_{Y_{j}}]\big{|}_{Z_{j}})=0,\text{ for }k<0.

Further, recall that qq is the projection 𝒳→ℂλ\mathcal{X}\to\mathbb{C}_{\lambda}. We have that q+𝒩Yj[!D|Yj]q_{+}\mathcal{N}_{Y_{j}}[!D|_{Y_{j}}] is a free module on β„‚Ξ»\mathbb{C}_{\lambda}, since it shall admit a mixed twistor structure, [Moc15, Proposition 7.2.7]. In particular, ℍk(DRXΞDol𝒩Yj[!D|Yj])\mathbb{H}^{k}(\textup{DR}_{X}\Xi_{\textup{Dol}}\mathcal{N}_{Y_{j}}[!D|_{Y_{j}}]) has the same dimension as of ℍk(DRXΞDR𝒩Yj[!D|Yj])\mathbb{H}^{k}(\textup{DR}_{X}\Xi_{\textup{DR}}\mathcal{N}_{Y_{j}}[!D|_{Y_{j}}]). Note that this argument is used to replace the classical Hodge-de Rham complex degeneration for proving Kodaira vanishing. So we have

(5.3) ℍk(DRXΞDol𝒩Yj[!D|Yj])=ℍk(DRXβ„°j)=0,Β forΒ k<0.\mathbb{H}^{k}(\textup{DR}_{X}\Xi_{\textup{Dol}}\mathcal{N}_{Y_{j}}[!D|_{Y_{j}}])=\mathbb{H}^{k}(\textup{DR}_{X}\mathcal{E}^{j})=0,\text{ for }k<0.

By considering (5.2), now we can conclude the proof by a standard argument of the degeneration of the Hodge to de Rham spectral sequence, e.g. [EV92, Appendix 25] ∎

Then, we prove the relative version.

Proof of Theorem 2.1.

We only prove the second vanishing as in the previous proof. The first one follows using the dual construction. Since we can add the pullback of a sufficiently ample line bundle on SS, we can assume that β„’\mathcal{L} is semi-ample on XX. Hence, as in the proof of Theorem 1.3, we have an NCNC sequence of varieties

X=Y0βŠƒY1βŠƒβ€¦βŠƒYn,X=Y_{0}\supset Y_{1}\supset...\supset Y_{n},

with respect to β„³[!D]\mathcal{M}[!D], induced by π’œ=β„’m\mathcal{A}=\mathcal{L}^{m}. Then, we have the following exact sequence of Higgs sheaves:

(5.4) 0β†’ΞžDolβ„³[!D]βŠ—β„’βˆ’1β†’β„°0β†’β„°1→…→ℰnβ†’0,0\to\Xi_{\textup{Dol}}\mathcal{M}[!D]\otimes\mathcal{L}^{-1}\to\mathcal{E}^{0}\to\mathcal{E}^{1}\to...\to\mathcal{E}^{n}\to 0,

satisfying

β„°jβ‰ƒΞžDoli+𝒩Yj[!D|Yj],\mathcal{E}^{j}\simeq\Xi_{\textup{Dol}}i_{+}\mathcal{N}_{Y_{j}}[!D|_{Y_{j}}],

with 𝒩Yi\mathcal{N}_{Y_{i}} are β„›Yi\mathscr{R}_{Y_{i}}-modules underlying mixed twistor π’Ÿ\mathscr{D}-modules, satisfying

𝒩Yj[!D|Yj]=𝒩Yj[!D|Yj+!Yj+1]=𝒩Yj[(!D|Yj+βˆ—Yj+1)].\mathcal{N}_{Y_{j}}[!D|_{Y_{j}}]=\mathcal{N}_{Y_{j}}[!D|_{Y_{j}}+!Y_{j+1}]=\mathcal{N}_{Y_{j}}[(!D|_{Y_{j}}+*Y_{j+1})].

Due to the assumption that ℒ​(d​D)\mathcal{L}(dD) is ff-ample, so is π’œ|Yj​(d​D|Yj)=π’ͺYj​(d​D|Yj+Yj+1)\mathcal{A}|_{Y_{j}}(dD|_{Y_{j}})=\mathcal{O}_{Y_{j}}(dD|_{Y_{j}}+Y_{j+1}). Set Zj:=Yjβˆ–(d​D|Yj+Yj+1)r​e​dZ_{j}:=Y_{j}\setminus(dD|_{Y_{j}}+Y_{j+1})_{red} is affine, being relative ample implies that the induced gj:Zjβ†’Sg_{j}:Z_{j}\to S are affine morphisms. Using Artin Vanishing, we get

β„‹pkgj,!(DRZjΞDR(𝒩Yi[!D|Yj]|Zj))=0,forΒ k<0,\mathcal{H}_{p}^{k}g_{j,!}(\textup{DR}_{Z_{j}}\Xi_{\textup{DR}}(\mathcal{N}_{Y_{i}}[!D|_{Y_{j}}]|_{Z_{j}}))=0,\text{for }k<0,

where β„‹pk\mathcal{H}_{p}^{k} stands for the kk-th cohomology with respect to the perverse tt-structure on constructable sheaves. This implies that, using Riemann-Hilbert correspondence, we have

(5.5) β„›kf+i+(𝒩Yj[!D|Yj])\displaystyle\mathcal{R}^{k}f_{+}i_{+}(\mathcal{N}_{Y_{j}}[!D|_{Y_{j}}]) ≃ℛkf†i†(𝒩Yj[!D|Yj])\displaystyle\simeq\mathcal{R}^{k}f_{\dagger}i_{\dagger}(\mathcal{N}_{Y_{j}}[!D|_{Y_{j}}])
≃ℛkgj,†(𝒩Yi[!D|Yj]|Zj)=0,Β forΒ k<0,\displaystyle\simeq\mathcal{R}^{k}g_{j,\dagger}(\mathcal{N}_{Y_{i}}[!D|_{Y_{j}}]|_{Z_{j}})=0,\text{ for }k<0,

in the category of coherent β„›\mathscr{R}-modules. Since we work locally on SS, we can also assume that SS is affine. In the case that kβ‰₯0k\geq 0, we can apply Artin Vanishing again, (also the discussion before (5.3),) to get

(5.6) ΞDolβ„›la†ℛkf†i†(𝒩Yj[!D|Yj])\displaystyle\Xi_{\textup{Dol}}\mathcal{R}^{l}a_{\dagger}\mathcal{R}^{k}f_{\dagger}i_{\dagger}(\mathcal{N}_{Y_{j}}[!D|_{Y_{j}}]) β‰ƒΞžDRβ„›la†ℛkf†i†(𝒩Yj[!D|Yj])\displaystyle\simeq\Xi_{\textup{DR}}\mathcal{R}^{l}a_{\dagger}\mathcal{R}^{k}f_{\dagger}i_{\dagger}(\mathcal{N}_{Y_{j}}[!D|_{Y_{j}}])
≃ℍclDRSΞDRβ„›kgj,†(𝒩Yi[!D|Yj]|Zj)=0,forΒ l<0,\displaystyle\simeq\mathbb{H}_{c}^{l}\textup{DR}_{S}\Xi_{\textup{DR}}\mathcal{R}^{k}g_{j,\dagger}(\mathcal{N}_{Y_{i}}[!D|_{Y_{j}}]|_{Z_{j}})=0,\text{for }l<0,

where aa is the universal map Sβ†’Spec​(β„‚)S\to\text{Spec}(\mathbb{C}).

We have the Grothendieck-Leray spectral sequence:

E2p,q=ΞDolβ„›pa†ℛqf†i†(𝒩Yj[!D|Yj])=ℍcpDRSΞDolβ„›qgj,†𝒩Yj[!D|Yj],E^{p,q}_{2}=\Xi_{\textup{Dol}}\mathcal{R}^{p}a_{\dagger}\mathcal{R}^{q}f_{\dagger}i_{\dagger}(\mathcal{N}_{Y_{j}}[!D|_{Y_{j}}])=\mathbb{H}_{c}^{p}\textup{DR}_{S}\Xi_{\textup{Dol}}\mathcal{R}^{q}g_{j,\dagger}\mathcal{N}_{Y_{j}}[!D|_{Y_{j}}],

which converges to

Ep+q=ℍcp+qDRXΞDoli†𝒩Yj[!D|Yj].E^{p+q}=\mathbb{H}_{c}^{p+q}\textup{DR}_{X}\Xi_{\textup{Dol}}i_{\dagger}\mathcal{N}_{Y_{j}}[!D|_{Y_{j}}].

Now, using (5.5) and (5.6), we can get

Ek=ℍckDRXΞDoli†𝒩Yj[!D|Yj]≃ℍckDRXβ„°j=0,forΒ k<0.E^{k}=\mathbb{H}_{c}^{k}\textup{DR}_{X}\Xi_{\textup{Dol}}i_{\dagger}\mathcal{N}_{Y_{j}}[!D|_{Y_{j}}]\simeq\mathbb{H}_{c}^{k}\textup{DR}_{X}\mathcal{E}^{j}=0,\text{for }k<0.

Considering (5.4), we can use a standard argument of the degeneration of the Hodge to de Rham spectral sequence as in the proof Theorem 1.3, to get

(5.7) ℍckDRXΞDolβ„³[!D]βŠ—β„’βˆ’1=0,forΒ k<0.\mathbb{H}_{c}^{k}\textup{DR}_{X}\Xi_{\textup{Dol}}\mathcal{M}[!D]\otimes\mathcal{L}^{-1}=0,\text{for }k<0.

Note that aβˆ—=a!​[d]a_{*}=a_{!}[d] is an exact functor on the category of quasi-coherent sheaves, and actually it is fully faithful from the category of π’ͺS\mathcal{O}_{S}-modules to the category of aβˆ—β€‹π’ͺSa_{*}\mathcal{O}_{S}-modules, [Har77, II.Exercise.5.3]. Decomposing ℍc\mathbb{H}_{c} into a!​fβˆ—a_{!}f_{*}, and combining (5.7), we can conclude

β„›kfβˆ—DRXΞDolβ„³[!D]βŠ—β„’βˆ’1=0,forΒ k<βˆ’d.\mathcal{R}^{k}f_{*}\textup{DR}_{X}\Xi_{\textup{Dol}}\mathcal{M}[!D]\otimes\mathcal{L}^{-1}=0,\text{for }k<-d.

∎

Let’s prove the generalized Kawamata-Viehweg vanishing with β„š\mathbb{Q}-divisors.

Proof for Theorem 1.4.

According to Remark 1.5, we can always assume that πŸŽβ‰€π›<𝟏,πŸŽβ‰€πœ<𝟏\mathbf{0}\leq\mathbf{b}<\mathbf{1},\mathbf{0}\leq\mathbf{c}<\mathbf{1}. We will focus on the second vanishing, the other one is similar, by just shifting those indexes to πŸŽβ‰€βˆ’π›<𝟏,πŸŽβ‰€βˆ’πœ<𝟏\mathbf{0}\leq-\mathbf{b}<\mathbf{1},\mathbf{0}\leq-\mathbf{c}<\mathbf{1}, instead. We work locally around Ξ»=0\lambda=0, hence we omit the super-index (0)(0) when taking the KM-filtration.

We first try to reduce numerical equivalence to linear equivalence. Note that

L≑n​u​ma​A+𝐛​B+πœβ€‹C,L\equiv_{num}aA+\mathbf{b}B+\mathbf{c}C,

is equivalent to that (see e.g. [Laz04a, Remark 1.1.20]) there exists a topologically trivial line bundle 𝒫\mathcal{P}, and nβˆˆβ„€+n\in\mathbb{Z}^{+}, such that

π’ͺ​(n​L)βŠ—π’«β‰ƒπ’ͺ​(n​(a​A+𝐛​B+πœβ€‹C)).\mathcal{O}(nL)\otimes\mathcal{P}\simeq\mathcal{O}(n(aA+\mathbf{b}B+\mathbf{c}C)).

Apply Bloch–Gieseker coverings on 𝒫\mathcal{P} and (a​A+𝐛​B+πœβ€‹C)(aA+\mathbf{b}B+\mathbf{c}C) [Laz04a, 4.1.10]. That means there exists a covering map Ο€:X~β†’X\pi:\tilde{X}\to X, satisfying Ο€βˆ—β€‹π’«β‰ƒπ’«~n,\pi^{*}\mathcal{P}\simeq\tilde{\mathcal{P}}^{n}, for some topologically trivial line bundle 𝒫~\tilde{\mathcal{P}} on X~\tilde{X}, and L~=Ο€βˆ—β€‹(a​A+𝐛​B+πœβ€‹C)\tilde{L}=\pi^{*}(aA+\mathbf{b}B+\mathbf{c}C) is a integral divisor. Actually, Ο€\pi can be constructed as a cyclic cover ramified along smooth divisors of sufficient general location. Hence, we can require that Ο€\pi is non-characteristic with respect to β„³(βˆ—D)\mathcal{M}(*D). In particular, Ο€\pi does not ramify along DD, and A~(:=Ο€βˆ—β€‹A),B~,C~\tilde{A}(:=\pi^{*}A),\tilde{B},\tilde{C} are reduced and normal crossing, so we can set

β„³~(βˆ—D~):=π†ℳ(βˆ—D)β‰ƒΟ€βˆ—β„³(βˆ—D).\tilde{\mathcal{M}}(*\tilde{D}):=\pi^{\dagger}\mathcal{M}(*D)\simeq\pi^{*}\mathcal{M}(*D).

By a local computation, the KM-filtrations on β„³~(βˆ—D~)\tilde{\mathcal{M}}(*\tilde{D}) along the components of DD are also just the pull back the corresponding filtrations on β„³[βˆ—D]\mathcal{M}[*D].

V𝐛B~V<𝐜C~β„³~(βˆ—D~)β‰ƒΟ€βˆ—V𝐛BV<𝐜Cβ„³(βˆ—D)V^{\tilde{B}}_{\mathbf{b}}V^{\tilde{C}}_{<\mathbf{c}}\tilde{\mathcal{M}}(*\tilde{D})\simeq\pi^{*}V^{B}_{\mathbf{b}}V^{C}_{<\mathbf{c}}\mathcal{M}(*D)

In particular, β„³~[βˆ—D~]\tilde{\mathcal{M}}[*\tilde{D}] is still VV-compatible with respect to D~\tilde{D}.

Due to Proposition 3.8, we have

π†(V𝐛B~V<𝐜C~β„³~(βˆ—D~))≃V𝐛BV<𝐜Cπ†ℳ(βˆ—D).\pi_{\dagger}(V^{\tilde{B}}_{\mathbf{b}}V^{\tilde{C}}_{<\mathbf{c}}\tilde{\mathcal{M}}(*\tilde{D}))\simeq V^{B}_{\mathbf{b}}V^{C}_{<\mathbf{c}}\pi_{\dagger}\mathcal{M}(*D).

This implies that V𝐛BV<𝐜Cβ„³(βˆ—D)V^{B}_{\mathbf{b}}V^{C}_{<\mathbf{c}}\mathcal{M}(*D) is a direct summand of π†(V𝐛B~V<𝐜C~β„³~(βˆ—D~))\pi_{\dagger}(V^{\tilde{B}}_{\mathbf{b}}V^{\tilde{C}}_{<\mathbf{c}}\tilde{\mathcal{M}}(*\tilde{D})). Due to the compatibility of DR, ΞDol\Xi_{\textup{Dol}} and the direct image functor, and by projection formula, we get

DR(X,D)(ΞDolV𝐛BV<𝐜Cβ„³(βˆ—D))βŠ—π’ͺX(βˆ’L)\textup{DR}_{(X,D)}(\Xi_{\textup{Dol}}V^{B}_{\mathbf{b}}V^{C}_{<\mathbf{c}}\mathcal{M}(*D))\otimes\mathcal{O}_{X}(-L)

is a direct summand of

Ο€βˆ—(DR(X~,D~)(ΞDolV𝐛B~V<𝐜C~β„³~(βˆ—D~))βŠ—π’«~βŠ—π’ͺX~(βˆ’L~)),\pi_{*}(\textup{DR}_{(\tilde{X},\tilde{D})}(\Xi_{\textup{Dol}}V^{\tilde{B}}_{\mathbf{b}}V^{\tilde{C}}_{<\mathbf{c}}\tilde{\mathcal{M}}(*\tilde{D}))\otimes\tilde{\mathcal{P}}\otimes\mathcal{O}_{\tilde{X}}(-\tilde{L})),

where L~=a​A~+𝐛​B~+πœβ€‹C~\tilde{L}=a\tilde{A}+\mathbf{b}\tilde{B}+\mathbf{c}\tilde{C}, and 𝒫~βŠ—π’ͺX~​(βˆ’L~)β‰ƒΟ€βˆ—β€‹π’ͺX​(βˆ’L)\tilde{\mathcal{P}}\otimes\mathcal{O}_{\tilde{X}}(-\tilde{L})\simeq\pi^{*}\mathcal{O}_{X}(-L). Note that 𝒫~\tilde{\mathcal{P}}, as a topologically trivial line bundle, can be viewed as a stable higgs bundle that is associated to a rank one unitary representation, i.e. a rank one harmonic bundle with trivial higgs connection, so we can use β„³~βŠ—pβˆ—β€‹π’«~\tilde{\mathcal{M}}\otimes p^{*}\tilde{\mathcal{P}} to replace β„³~\tilde{\mathcal{M}}, still getting a mixed twisor π’Ÿ\mathscr{D}-module. Hence we reduce the problem to the case that we replace ≑n​u​m\equiv_{num} in the initial statement by ≑l​i​n\equiv_{lin}.

In this setting, we can further apply Bloch–Gieseker coverings with respect to AA, and we reduce to the case that a​AaA is linear equivalent to some β„€\mathbb{Z}-divisor Aβ€²A^{\prime}, hence 𝐛​B+πœβ€‹C\mathbf{b}B+\mathbf{c}C is also linear equivalent to some β„€\mathbb{Z}-divisor Lβ€²L^{\prime}.

Now, we are ready to apply Esnault-Viehweg covering construction. Denote π’œ=π’ͺ​(Aβ€²)\mathcal{A}=\mathcal{O}(A^{\prime}), 𝐛​B+πœβ€‹C=𝐝​D\mathbf{b}B+\mathbf{c}C=\mathbf{d}D and β„’=π’ͺ​(𝐝​D)\mathcal{L}=\mathcal{O}(\mathbf{d}D). Let NN be the least positive integer such that Nβ€‹πβˆˆβ„€nN\mathbf{d}\in\mathbb{Z}^{n}, following the construction in Β§4, and using Lemma 4.1, we get

V𝐛BV<𝐜Cβ„³(βˆ—D)βŠ—pβˆ—β„’βˆ’1≃V𝟎BV<𝟎Cβ„³1(βˆ—D)V^{B}_{\mathbf{b}}V^{C}_{<\mathbf{c}}\mathcal{M}(*D)\otimes p^{*}\mathcal{L}^{-1}\simeq V^{B}_{\mathbf{0}}V^{C}_{<\mathbf{0}}\mathcal{M}_{1}(*D)

This implies

DR(X,D)(ΞDolV𝐛BV<𝐜Cβ„³(βˆ—D))βŠ—π’ͺX(βˆ’L)\displaystyle\textup{DR}_{(X,D)}(\Xi_{\textup{Dol}}V^{B}_{\mathbf{b}}V^{C}_{<\mathbf{c}}\mathcal{M}(*D))\otimes\mathcal{O}_{X}(-L)
≃\displaystyle\simeq DR(X,D)(ΞDolV𝟎BV<𝟎Cβ„³1(βˆ—D))βŠ—π’œβˆ’1\displaystyle\textup{DR}_{(X,D)}(\Xi_{\textup{Dol}}V^{B}_{\mathbf{0}}V^{C}_{<\mathbf{0}}\mathcal{M}_{1}(*D))\otimes\mathcal{A}^{-1}
(5.8) ≃\displaystyle\simeq DRX(ΞDolβ„³1[βˆ—B][!C])βŠ—π’œβˆ’1.\displaystyle\textup{DR}_{X}(\Xi_{\textup{Dol}}\mathcal{M}_{1}[*B][!C])\otimes\mathcal{A}^{-1}.

We use the logarithmic comparison (Proposition 3.7) to get the second isomorphism. Now, we can conclude the proof using Theorem 1.3. ∎

Proof of Theorem 2.2.

Once we get the identity (5.8), we can apply Theorem 2.1 to conclude. ∎

References

  • [AHL19] Donu Arapura, Feng Hao, and Hongshan Li. Vanishing theorems for parabolic Higgs bundles. Math. Res. Lett., 26(5):1251–1279, 2019.
  • [AMPW] Donu Arapura, Kenji Matsuki, Deepam Patel, and JarosΕ‚aw WΕ‚odarczyk. A kawamata-viehweg type formulation of the logarithmic akizuki-nakano vanishing theorem. preprint, arXiv:1806.01137.
  • [Bjo93] Jan-Erik Bjork. Analytic π’Ÿ{\mathscr{D}}-modules and applications, volume 247 of Mathematics and its Applications. Kluwer Academic Publishers Group, Dordrecht, 1993.
  • [BS05] Nero Budur and Morihiko Saito. Multiplier ideals, VV-filtration, and spectrum. J. Algebraic Geom., 14(2):269–282, 2005.
  • [dC98a] Mark AndreaΒ A. deΒ Cataldo. Effective nonvanishing, effective global generation. Ann. Inst. Fourier (Grenoble), 48(5):1359–1378, 1998.
  • [dC98b] Mark AndreaΒ A. deΒ Cataldo. Singular Hermitian metrics on vector bundles. J. Reine Angew. Math., 502:93–122, 1998.
  • [DH22] YaΒ Deng and Feng Hao. Vanishing theorem for tame harmonic bundles via l2l^{2}-cohomology. 2022.
  • [EV86] HΓ©lΓ¨ne Esnault and Eckart Viehweg. Logarithmic de Rham complexes and vanishing theorems. Invent. Math., 86(1):161–194, 1986.
  • [EV92] HΓ©lΓ¨ne Esnault and Eckart Viehweg. Lectures on vanishing theorems, volumeΒ 20 of DMV Seminar. BirkhΓ€user Verlag, Basel, 1992.
  • [Har77] Robin Hartshorne. Algebraic geometry. Springer-Verlag, New York-Heidelberg, 1977. Graduate Texts in Mathematics, No. 52.
  • [HTT08] Ryoshi Hotta, Kiyoshi Takeuchi, and Toshiyuki Tanisaki. DD-modules, perverse sheaves, and representation theory, volume 236 of Progress in Mathematics. BirkhΓ€user Boston, Inc., Boston, MA, 2008. Translated from the 1995 Japanese edition by Takeuchi.
  • [KMM87] Yujiro Kawamata, Katsumi Matsuda, and Kenji Matsuki. Introduction to the minimal model problem. In Algebraic geometry, Sendai, 1985, volumeΒ 10 of Adv. Stud. Pure Math., pages 283–360. North-Holland, Amsterdam, 1987.
  • [Laz04a] Robert Lazarsfeld. Positivity in algebraic geometry. I, volumeΒ 48 of Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge. A Series of Modern Surveys in Mathematics [Results in Mathematics and Related Areas. 3rd Series. A Series of Modern Surveys in Mathematics]. Springer-Verlag, Berlin, 2004. Classical setting: line bundles and linear series.
  • [Laz04b] Robert Lazarsfeld. Positivity in algebraic geometry. II, volumeΒ 49 of Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge. A Series of Modern Surveys in Mathematics [Results in Mathematics and Related Areas. 3rd Series. A Series of Modern Surveys in Mathematics]. Springer-Verlag, Berlin, 2004. Positivity for vector bundles, and multiplier ideals.
  • [Moc06] Takuro Mochizuki. Kobayashi-Hitchin correspondence for tame harmonic bundles and an application. AstΓ©risque, (309):viii+117, 2006.
  • [Moc07a] Takuro Mochizuki. Asymptotic behaviour of tame harmonic bundles and an application to pure twistor DD-modules. I. Mem. Amer. Math. Soc., 185(869):xii+324, 2007.
  • [Moc07b] Takuro Mochizuki. Asymptotic behaviour of tame harmonic bundles and an application to pure twistor DD-modules. II. Mem. Amer. Math. Soc., 185(870):xii+565, 2007.
  • [Moc11] Takuro Mochizuki. Wild harmonic bundles and wild pure twistor DD-modules. AstΓ©risque, (340):x+607, 2011.
  • [Moc15] Takuro Mochizuki. Mixed twistor DD-modules, volume 2125 of Lecture Notes in Mathematics. Springer, Cham, 2015.
  • [MP20] Mircea MustaΕ£Δƒ and Mihnea Popa. Hodge ideals for β„š\mathbb{Q}-divisors, VV-filtration, and minimal exponent. Forum Math. Sigma, 8:Paper No. e19, 41, 2020.
  • [Sab05] Claude Sabbah. Polarizable twistor D{D}-modules. AstΓ©risque, (300):vi+208, 2005.
  • [Sai90] Morihiko Saito. Mixed Hodge modules. Publ. Res. Inst. Math. Sci., 26(2):221–333, 1990.
  • [Sim91] CarlosΒ T. Simpson. Nonabelian Hodge theory. In Proceedings of the International Congress of Mathematicians, Vol. I, II (Kyoto, 1990), pages 747–756. Math. Soc. Japan, Tokyo, 1991.
  • [Som78] AndrewΒ John Sommese. Submanifolds of Abelian varieties. Math. Ann., 233(3):229–256, 1978.
  • [Suh18] Junecue Suh. Vanishing theorems for mixed Hodge modules and applications. J. Eur. Math. Soc. (JEMS), 20(11):2589–2606, 2018.
  • [Tot13] Burt Totaro. Line bundles with partially vanishing cohomology. J. Eur. Math. Soc. (JEMS), 15(3):731–754, 2013.
  • [Wei20a] Chuanhao Wei. Logarithmic comparison theorems in mixed hodge modules. Michigan Math. J., 69(1):201–223, 2020.
  • [Wei20b] Chuanhao Wei. Logarithmic Kodaira dimension and zeros of holomorphic log-one-forms. Math. Ann., 378(1-2):485–512, 2020.
  • [Wu17] Lei Wu. Vanishing and injectivity theorems for Hodge modules. Trans. Amer. Math. Soc., 369(11):7719–7736, 2017.
  • [Wu21] Lei Wu. Vanishing and injectivity for R-Hodge modules and R-divisors. Michigan Math. J., 1(1):1–27, 2021.