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

L2L^{2}-Dolbeault resolution of the lowest Hodge piece of a Hodge module

Junchao Shentu [email protected] School of Mathematical Sciences, University of Science and Technology of China, Hefei, 230026, China  and  Chen Zhao [email protected] School of Mathematical Sciences, University of Science and Technology of China, Hefei, 230026, China
Abstract.

In this paper, we introduce a coherent subsheaf of Saito’s SS-sheaf, which is a combination of the SS-sheaf and the multiplier ideal sheaf. We construct its L2L^{2}-Dolbeault resolution, which generalizes MacPherson’s conjecture on the L2L^{2} resolution of the Grauert-Riemenschneider sheaf. We also prove various vanishing theorems for the SS-sheaf (Saito’s vanishing theorem, Kawamata-Viehweg vanishing theorem and some new ones like Nadel vanishing theorem) transcendentally. Finally, we discuss some applications of our results on the relative version of Fujita’s conjecture (e.g. Kawamata’s conjecture).

1. Introduction

The technique of L2L^{2}-estimates developed by Andreotti-Vesentini [AV1965] and Hörmander [Hormander1965] and Saito’s theory of Hodge modules [MSaito1988, MSaito1990] have been of great importance in the development of algebraic and complex geometry. The purpose of this paper is to resolve Saito’s SS-sheaf [MSaito1991] (a generalization of the dualizing sheaf) by locally L2L^{2}-integrable differential forms. We then prove various vanishing theorems for the SS-sheaf by a transcendental approach. Some of these theorems have previously been proved by Hodge-theoretic approach (e.g. [Suh2018, Wu2017, MSaito1991(2)]) and have applications in the investigation of Shimura varieties [Suh2018].

Let XX be a reduced, irreducible complex space of dimension nn and XoXregX^{o}\subset X_{\rm reg} a Zariski open subset. Let ds2ds^{2} be a hermitian metric on XoX^{o} and 𝕍:=(𝒱,,,h𝕍)\mathbb{V}:=(\mathcal{V},\nabla,\mathcal{F}^{\bullet},h_{\mathbb{V}}) an \mathbb{R}-polarized variation of Hodge structure on XoX^{o}, where h𝕍h_{\mathbb{V}} is the Hodge metric defined by (u,v)h𝕍:=Q(Cu,v¯)(u,v)_{h_{\mathbb{V}}}:=Q(Cu,\overline{v}) with QQ being the polarization of 𝕍\mathbb{V} and CC the Weil operator. As a generalization of the dualizing sheaf, Saito [MSaito1991] defines the SS-sheaf S(ICX(𝕍))S(IC_{X}(\mathbb{V})) associated to 𝕍\mathbb{V} as the lowest Hodge piece of the intermediate extension ICX(𝕍)IC_{X}(\mathbb{V}) and uses it to give a solution to a Kollár’s conjecture [Kollar1986]. Saito’s SS-sheaf has been of great importance in the application of the theory of Hodge modules to complex algebraic geometry (see Popa [Popa2018] for a survey). Let φ:X[,)\varphi:X\to[-\infty,\infty) be a quasi-plurisubharmonic (quasi-psh for short) function and j:XoXj:X^{o}\to X the open immersion. We introduce the multiplier SS-sheaf as

S(ICX(𝕍),φ):={αj(KXoS(𝕍))||α|ds2,h𝕍2eφvolds2< locally at every point xX},\displaystyle S(IC_{X}(\mathbb{V}),\varphi):=\left\{\alpha\in j_{\ast}\left(K_{X^{o}}\otimes S(\mathbb{V})\right)\bigg{|}\int|\alpha|^{2}_{ds^{2},h_{\mathbb{V}}}e^{-\varphi}{\rm vol}_{ds^{2}}<\infty\textrm{ locally at every point }x\in X\right\},

where S(𝕍):=max{k|k0}S(\mathbb{V}):=\mathcal{F}^{\max\{k|\mathcal{F}^{k}\neq 0\}} is the top indexed nonzero piece of the Hodge filtration \mathcal{F}^{\bullet} and KXoK_{X^{o}} is the holomorphic canonical bundle of XoX^{o}. This kind of sheaf is a combination of Saito’s SS-sheaf and the multiplier ideal sheaf (φ)\mathscr{I}(\varphi), and has the following features:

  1. (1)

    S(ICX(𝕍),φ)S(IC_{X}(\mathbb{V}),\varphi) is a coherent subsheaf of Saito’s SS-sheaf S(ICX(𝕍))S(IC_{X}(\mathbb{V})) (Proposition 4.11).

  2. (2)

    S(ICX(𝕍),φ)S(IC_{X}(\mathbb{V}),\varphi) is independent of the choice of ds2ds^{2} (Proposition 4.2).

  3. (3)

    S(ICX(𝕍),0)=S(ICX(𝕍))S(IC_{X}(\mathbb{V}),0)=S(IC_{X}(\mathbb{V})) (Theorem 4.10). S(ICX(𝕍),)=0S(IC_{X}(\mathbb{V}),-\infty)=0. This gives an alternative definition of S(ICX(𝕍))S(IC_{X}(\mathbb{V})) without using the language of Hodge modules. A similar relation regarding (0,0)(0,0)-forms is also recently observed by Schnell-Yang [SY2023].

  4. (4)

    S(ICX(𝕍),φ)S(IC_{X}(\mathbb{V}),\varphi) has the functorial property (Proposition 4.3).

  5. (5)

    Assume that XX is smooth, X\XoXX\backslash X^{o}\subset X is a (possibly empty) normal crossing divisor and 𝕍\mathbb{V} is an \mathbb{R}-polarized variation of Hodge structure with unipotent local monodromies. Let φ\varphi be a plurisubharmonic (psh for short) function with generalized analytic singularities along X\XoX\backslash X^{o} (Definition 4.7). Then, as shown in Proposition 4.9, there is an isomorphism

    (1.1) S(ICX(𝕍),φ)S(ICX(𝕍))(φ).\displaystyle S(IC_{X}(\mathbb{V}),\varphi)\simeq S(IC_{X}(\mathbb{V}))\otimes\mathscr{I}(\varphi).

    When the local monodromies are not necessarily unipotent, (1.1) only holds on X\XoX\backslash X^{o}. Nevertheless, as stated in Proposition 4.8, there is a decomposition

    (1.2) S(ICX(𝕍),φ)i=1m(φi)ωX\displaystyle S(IC_{X}(\mathbb{V}),\varphi)\simeq\bigoplus_{i=1}^{m}\mathscr{I}(\varphi_{i})\otimes\omega_{X}

    locally at every point of X\XoX\backslash X^{o}. Here φi\varphi_{i} are quasi-psh functions which depend on φ\varphi and the eigenvalues of the local monodromies of 𝕍\mathbb{V}. (1.2) plays a crucial role in the study of the relative Fujita conjecture (§1.2).

  6. (6)

    S(ICX(𝕍),φ)S(IC_{X}(\mathbb{V}),\varphi) satisfies an Ohsawa-Takegoshi extension theorem, at least when φ\varphi is generically smooth (Theorem 4.14).

Let (E,hφ)(E,h_{\varphi}) be a holomorphic vector bundle on XX with a possibly singular hermitian metric hφh_{\varphi}. Throughout this paper, by a singular hermitian metric hφ=eφh0h_{\varphi}=e^{-\varphi}h_{0} we always assume that h0h_{0} is a smooth hermitian metric and φ\varphi is a quasi-psh function. S(ICX(𝕍),φ)S(IC_{X}(\mathbb{V}),\varphi) is independent of the choice of the decomposition hφ=eφh0h_{\varphi}=e^{-\varphi}h_{0} (Lemma 4.5). Let 𝒟X,ds2n,q(S(𝕍)E,h𝕍hφ)\mathscr{D}^{n,q}_{X,ds^{2}}(S(\mathbb{V})\otimes E,h_{\mathbb{V}}\otimes h_{\varphi}) denote the sheaf of measurable S(𝕍)ES(\mathbb{V})\otimes E-valued (n,q)(n,q)-forms α\alpha such that α\alpha and ¯α\bar{\partial}\alpha are locally square integrable with respect to ds2ds^{2} and h𝕍hφh_{\mathbb{V}}\otimes h_{\varphi}. Let ω\omega be the (1,1)(1,1)-form associated to ds2ds^{2}. The main result of the present paper is

Theorem 1.1 (=Theorem 5.1).

Assume that locally at every point xXx\in X there is a neighborhood UU of xx, a strictly psh function λC2(U)\lambda\in C^{2}(U) and a bounded psh function ΦC2(UXo)\Phi\in C^{2}(U\cap X^{o}) such that 1¯λω|UXo1¯Φ\sqrt{-1}\partial\bar{\partial}\lambda\lesssim\omega|_{U\cap X^{o}}\lesssim\sqrt{-1}\partial\bar{\partial}\Phi. Then the complex of sheaves

0S(ICX(𝕍),φ)E𝒟X,ds2n,0(S(𝕍)E,h𝕍hφ)¯¯𝒟X,ds2n,n(S(𝕍)E,h𝕍hφ)0\displaystyle 0\to S(IC_{X}(\mathbb{V}),\varphi)\otimes E\to\mathscr{D}^{n,0}_{X,ds^{2}}(S(\mathbb{V})\otimes E,h_{\mathbb{V}}\otimes h_{\varphi})\stackrel{{\scriptstyle\bar{\partial}}}{{\to}}\cdots\stackrel{{\scriptstyle\bar{\partial}}}{{\to}}\mathscr{D}^{n,n}_{X,ds^{2}}(S(\mathbb{V})\otimes E,h_{\mathbb{V}}\otimes h_{\varphi})\to 0

is exact. If XX is moreover compact, then there is an isomorphism

(1.3) Hq(X,S(ICX(𝕍),φ)E)H(2),maxn,q(Xo,S(𝕍)E;ds2,h𝕍hφ),q.\displaystyle H^{q}(X,S(IC_{X}(\mathbb{V}),\varphi)\otimes E)\simeq H^{n,q}_{(2),\rm max}(X^{o},S(\mathbb{V})\otimes E;ds^{2},h_{\mathbb{V}}\otimes h_{\varphi}),\quad\forall q.

When ds2ds^{2} is the hermitian metric on XX (Definition 3.1), 𝕍=Xreg\mathbb{V}=\mathbb{C}_{X_{\rm reg}}, φ=0\varphi=0 and E=𝒪XE=\mathscr{O}_{X} is endowed with the trivial metric, (1.3) implies the results by Pardon-Stern [Pardon_Stern1991] and by Ruppenthal [Ruppenthal2014] on MacPherson’s conjecture.

1.1. Vanishing theorems

The L2L^{2}-resolution of the multiplier SS-sheaf allows us to investigate the SS-sheaf by means of analytical methods. Theorem 1.1 is used to give a transcendental prove to Kollár’s conjecture ([Kollar1986, §5]) on the derived pushforward of S(ICX(𝕍))S(IC_{X}(\mathbb{V})) in [SC2021_kollar], as well as its generalizations. In the present paper we deduce from Theorem 1.1 various vanishing theorems for Saito’s SS-sheaf.

Theorem 1.2 (Nadel type vanishing theorem, =Corollary 6.4).

Let f:XYf:X\to Y be a surjective proper Kähler holomorphic map between irreducible complex spaces. Let 𝕍\mathbb{V} be an \mathbb{R}-polarized variation of Hodge structure defined on a Zariski open subset of XX. Let (L,hφ)(L,h_{\varphi}) a holomorphic line bundle on XX with a possibly singular hermitian metric hφ:=eφhh_{\varphi}:=e^{-\varphi}h. Assume that 1Θhφ(L)\sqrt{-1}\Theta_{h_{\varphi}}(L) is ff-positive. Then

Rif(S(ICX(𝕍),φ)L)=0,i>0.\displaystyle R^{i}f_{\ast}(S(IC_{X}(\mathbb{V}),\varphi)\otimes L)=0,\quad\forall i>0.

When XX is smooth, YY is a point and 𝕍=X\mathbb{V}=\mathbb{C}_{X}, we recover the Nadel vanishing theorem [Nadel1990]. Many interesting generalizations are obtained, such as [Demailly1982, Matsumura2014, Matsumura2015, Iwai2021]. When XX is a projective variety, with a careful choice of hφh_{\varphi} we obtain

Corollary 1.3 (Demailly-Kawamata-Viehweg type vanishing theorem, =Corollary 6.2).

Let XX be a projective algebraic variety of dimension nn and 𝕍\mathbb{V} an \mathbb{R}-polarized variation of Hodge structure defined on a Zariski open subset of XX. Let LL be a line bundle such that some positive multiple mL=F+DmL=F+D where FF is a nef line bundle and DD is an effective divisor. Then

Hq(X,S(ICX(𝕍),φDm)L)=0,q>nnd(L).H^{q}(X,S(IC_{X}(\mathbb{V}),\frac{\varphi_{D}}{m})\otimes L)=0,\quad\forall q>n-{\rm nd}(L).

Here φD\varphi_{D} is the psh function associated to DD.

When LL is nef and big, this vanishing theorem has been established by Suh [Suh2018] and Wu [Wu2017] by means of Hodge theoretic methods, which generalizes Saito’s vanishing theorem for the SS-sheaf [MSaito1991(2)]. When XX is smooth and 𝕍=X\mathbb{V}=\mathbb{C}_{X} is the trivial Hodge module, it reduces to the Demailly-Kawamata-Viehweg vanishing theorem [Demailly2012, 6.25], with its roots traced back to Kawamata and Viehweg in [Kawamata1982, Viehweg1982]. Recent developments include [Cao2014, Wu2022, Inayama2022, DP2003, Demailly1991]. The Kodaira-Nakano-Kazama vanishing theorem, the relative vanishing theorem and Fujino-Enoki-Kollár injectivity theorem are generalized to coefficients in S(ICX(𝕍),φ)S(IC_{X}(\mathbb{V}),\varphi) (Theorem 6.5, Corollary 6.4, Theorem 6.6). It also implies the Esnault-Viehweg type injectivity theorem (Corollary 6.7) which has been proved in [Wu2017, Theorem 1.4] using Hodge theoretic methods. For the case that XX is smooth and 𝕍\mathbb{V} is trivial, these vanishing theorems are mainly due to the efforts of Kodaira [Kodaira1953], Nakano [Nakano1974], Kazama [Kazama1973], Takegoshi [Takegoshi1985], Kollár [Kollar1986], Enoki [Enoki1993], Fujino [Fujino2017] and Cao-Pǎun [Paun2020]. Readers may also refer to the works of Fujino and Matsumura [Matsumura2018, Matsumura20182, Matsumura20183, Matsumura20184] and the references therein.

1.2. Application to the relative version of Fujita’s conjecture

As a relative version of Fujita’s conjecture [Fujita1987], Kawamata raised the following conjecture in [Kawamata2002] with the case dimX4\dim_{\mathbb{C}}X\leq 4 settled therein.

Conjecture 1.4 (Kawamata [Kawamata2002]).

Let f:YXf:Y\to X be a proper morphism between smooth projective algebraic varieties. Assume that the degenerate loci of ff is contained in a normal crossing divisor DXD\subset X. Let LL be an ample line bundle on XX. Then RqfωYLdimX+1R^{q}f_{\ast}\omega_{Y}\otimes L^{\dim X+1} is generated by global sections for every 0qdimX0\leq q\leq\dim_{\mathbb{C}}X.

With the help of the L2L^{2}-Dolbeault resolution (Theorem 1.1) on RqfωYR^{q}f_{\ast}\omega_{Y} (an example of Saito’s SS-sheaf), we are able to investigate the separation of jets of RqfωYLdimX+1R^{q}f_{\ast}\omega_{Y}\otimes L^{\dim X+1} using the transendental method developed by Angehrn-Siu [Siu1995] and Demailly [Demailly2012, Theorem 7.4].

Corollary 1.5.

Let f:YXf:Y\to X be a proper holomorphic morphism from a Kähler manifold to a projective algebraic variety where dimX=n\dim_{\mathbb{C}}X=n. Assume that the degenerate loci of ff is contained in a normal crossing divisor DXD\subset X. Let LL be an ample line bundle on XX. Assume that there is a positive number κ>0\kappa>0 such that

LkW(12n(n+2r1)+κ)dL^{k}\cdot W\geq\left(\frac{1}{2}n(n+2r-1)+\kappa\right)^{d}

for any irreducible subvariety WW of dimension 0dn0\leq d\leq n in XX. Let 0qn0\leq q\leq n. Then the global holomorphic sections of RqfωYLR^{q}f_{\ast}\omega_{Y}\otimes L separate any set of rr distinct points x1,,xrXx_{1},\dots,x_{r}\in X, i.e. there is a surjective map

H0(X,RqfωYL)1krRqfωYL𝒪X,xk/mX,xk.H^{0}(X,R^{q}f_{\ast}\omega_{Y}\otimes L)\to\bigoplus_{1\leq k\leq r}R^{q}f_{\ast}\omega_{Y}\otimes L\otimes\mathscr{O}_{X,x_{k}}/m_{X,x_{k}}.

A similar result is obtained by Wu [Wu2017] Hodge theoretically. When f=Idf={\rm Id}, this reduces to the result in [Siu1995].

By using Demailly’s singular metric on the adjoint bundles ([Demailly2012, Theorem 7.4]) we are able obtain the relative version of [Demailly2012, Theorem 7.4].

Corollary 1.6.

Let f:YXf:Y\to X be a proper holomorphic morphism from a Kähler manifold to a projective algebraic variety where dimX=n\dim_{\mathbb{C}}X=n. Assume that the degenerate loci of ff is contained in a normal crossing divisor DXD\subset X. Let LL be an ample line bundle and GG a nef line bundle on XX. Then there is a surjective map

H0(X,RqfωYωXLmG)1krRqfωYωXLmG𝒪X,xk/mX,xksk+1H^{0}(X,R^{q}f_{\ast}\omega_{Y}\otimes\omega_{X}\otimes L^{\otimes m}\otimes G)\to\bigoplus_{1\leq k\leq r}R^{q}f_{\ast}\omega_{Y}\otimes\omega_{X}\otimes L^{\otimes m}\otimes G\otimes\mathscr{O}_{X,x_{k}}/m^{s_{k}+1}_{X,x_{k}}

at arbitrary points x1,,xrXx_{1},\dots,x_{r}\in X for every 0qdimYdimX0\leq q\leq\dim_{\mathbb{C}}Y-\dim_{\mathbb{C}}X, provided that m2+1kr(3n+2sk1n)m\geq 2+\sum_{1\leq k\leq r}\binom{3n+2s_{k}-1}{n}.

Since Theorem 1.2 holds for a general Hodge module, we actually prove the analogues of Corollary 1.5 and Corollary 1.6 for Hodge modules (Theorem 7.3 and Theorem 7.4). We also obtain the results on separating jets with an optimal bound (Theorem 7.5) when LL is ample and base point free.

Remark 1.7.

For a general Hodge module MM, the semi-simplicity of MM provides a unique decomposition M=ICZi(𝕍i)M=\bigoplus IC_{Z_{i}}(\mathbb{V}_{i}) where ICZi(𝕍i)IC_{Z_{i}}(\mathbb{V}_{i}) is a Hodge module with its strict support an irreducible Zariski closed subset ZiXZ_{i}\subset X for each ii. Then S(M,φ)S(M,\varphi) could be defined as S(ICZi(𝕍i),φ|Zi)\bigoplus S(IC_{Z_{i}}(\mathbb{V}_{i}),\varphi|_{Z_{i}}). The main results in the present paper hold for a general Hodge module as long as they are valid for Hodge modules with strict support. Therefore, we only consider Hodge modules with strict support in the present paper.

Acknowledgment: Both authors would like to thank Zhenqian Li, Ya Deng and Ruijie Yang for many helpful conversations. The first author also thanks Lei Zhang for his interest in this work.

Conventions and Notations:

  • All complex spaces are separated, irreducible, reduced, paracompact and countable at infinity. Let XX be a complex space. A Zariski closed subset (=closed analytic subset) ZXZ\subset X is a closed subset which is locally defined as the zeros of a set of holomorphic functions. A subset YXY\subset X is called Zariski open if X\ZXX\backslash Z\subset X is Zariski closed.

  • Let YY be a complex manifold and EE a hermitian vector bundle on YY. Let ΘA1,1(Y,End(E))\Theta\in A^{1,1}(Y,End(E)). Denote Θ0\Theta\geq 0 if Θ\Theta is Nakano semipositive. Let Θ1,Θ2A1,1(Y,End(E))\Theta_{1},\Theta_{2}\in A^{1,1}(Y,End(E)). Then Θ1Θ2\Theta_{1}\geq\Theta_{2} stands for Θ1Θ20\Theta_{1}-\Theta_{2}\geq 0.

  • A CC^{\infty} form on a complex space XX is a CC^{\infty} form α\alpha on XregX_{\rm reg} so that the following statement hold: Locally at every point xXx\in X there is an open neighborhood UU of xx, a holomorphic embedding ι:UN\iota:U\to\mathbb{C}^{N} and βC(N)\beta\in C^{\infty}(\mathbb{C}^{N}) such that ιβ=α\iota^{\ast}\beta=\alpha on UXregU\cap X_{\rm reg}.

  • A psh (resp. strictly psh) function on a complex space XX is a function λ:X[,)\lambda:X\to[-\infty,\infty) such that locally at every point xXx\in X there is a neighborhood UU of xx, a closed immersion ι:UΩ\iota:U\to\Omega into a holomorphic manifold Ω\Omega and a psh (resp. strictly psh) function Λ\Lambda on Ω\Omega such that ιΛ=λ\iota^{\ast}\Lambda=\lambda. A function φ\varphi on XX is called quasi-psh if it can be written locally as a sum φ=α+ψ\varphi=\alpha+\psi of a CC^{\infty} function α\alpha and a psh function ψ\psi.

  • Let φ\varphi be a quasi-psh function on a holomorphic manifold XX. (φ)𝒪X\mathscr{I}(\varphi)\subset\mathscr{O}_{X} denotes the multiplier ideal sheaf consisting of holomorphic functions ff such that |f|2eφ|f|^{2}e^{-\varphi} is locally integrable.

  • Let (Y,ds2)(Y,ds^{2}) be a hermitian manifold and EE a holomorphic vector bundle on YY. A singular hermitian metric on EE is a measurable section hEE¯h\in E^{\ast}\otimes\overline{E}^{\ast} such that h=eφh0h=e^{-\varphi}h_{0} for some smooth hermitian metric h0h_{0} and some quasi-psh function φ\varphi.

  • Let α\alpha, β\beta be functions (resp. metrics or (1,1)(1,1)-forms). We denote αβ\alpha\lesssim\beta if αCβ\alpha\leq C\beta for some constant C>0C>0. Denote αβ\alpha\sim\beta if both αβ\alpha\lesssim\beta and βα\beta\lesssim\alpha hold.

2. Preliminaries on Saito’s SS-sheaf

2.1. Saito’s SS-sheaf

Readers may see [MSaito1988, MSaito1990, MSaito1991(2), Schnell_introMHS, Peter_Steenbrink2008] for the theory of Hodge module. In the present paper we will not use the theory of Hodge module. Instead, a concrete construction of Saito’s SS-sheaf using Deligne’s extension will be used. This construction is originated by Kollár in [Kollar1986].

Let XX be a complex space and XoXregX^{o}\subset X_{\rm reg} a Zariski open subset. Let 𝕍:=(𝒱,,,h𝕍)\mathbb{V}:=(\mathcal{V},\nabla,\mathcal{F}^{\bullet},h_{\mathbb{V}}) be an \mathbb{R}-polarized variation of Hodge structure ([Cattani_Kaplan_Schmid1986, §1]) on XoX^{o}. The SS-sheaf S(ICX(𝕍))S(IC_{X}(\mathbb{V})) associated with 𝕍\mathbb{V} is defined as follows.

  1. (1)

    (Log smooth case): Assume that XX is smooth and E:=X\XoE:=X\backslash X^{o} is a simple normal crossing divisor. Let E=EiE=\cup E_{i} be the irreducible decomposition. By [Deligne1970, §II, Proposition 5.4], there is a logarithmic flat holomorphic vector bundle (𝒱~1,~)(\widetilde{\mathcal{V}}_{-1},\widetilde{\nabla}) (unique up to isomorphisms):

    ~:𝒱~1ΩX(logE)𝒱~1,\displaystyle\widetilde{\nabla}:\widetilde{\mathcal{V}}_{-1}\to\Omega_{X}(\log E)\otimes\widetilde{\mathcal{V}}_{-1},

    such that (𝒱~1,~)|Xo(\widetilde{\mathcal{V}}_{-1},\widetilde{\nabla})|_{X^{o}} is holomorphically equivalent to (𝒱,)(\mathcal{V},\nabla) and the eigenvalues of the residue operator

    ResEi~:𝒱~1|Ei𝒱~1|Ei\displaystyle{\rm Res}_{E_{i}}\widetilde{\nabla}:\widetilde{\mathcal{V}}_{-1}|_{E_{i}}\to\widetilde{\mathcal{V}}_{-1}|_{E_{i}}

    lie in (1,0](-1,0]. Let j:XoXj:X^{o}\to X denote the open immersion. By [MSaito1991, Theorem 1.1], the SS-sheaf can be described as

    S(ICX(𝕍))=R(ICX(𝕍))ωXS(IC_{X}(\mathbb{V}))=R(IC_{X}(\mathbb{V}))\otimes\omega_{X}

    where

    R(ICX(𝕍))=j(S(𝕍))𝒱~1,S(𝕍):=max{k|k0}.\displaystyle R(IC_{X}(\mathbb{V}))=j_{\ast}(S(\mathbb{V}))\cap\widetilde{\mathcal{V}}_{-1},\quad S(\mathbb{V}):=\mathcal{F}^{\max\{k|\mathcal{F}^{k}\neq 0\}}.

    Moreover, R(ICX(𝕍))R(IC_{X}(\mathbb{V})) is a holomorphic subbundle of 𝒱~1\widetilde{\mathcal{V}}_{-1} according to the nilpotent orbit theorem (Schmid [Schmid1973] and Cattani-Kaplan-Schmid [Cattani_Kaplan_Schmid1986]).

  2. (2)

    (General case): Let π:X~X\pi:\widetilde{X}\to X be a proper bimeromorphic morphism such that π\pi is biholomorphic over XoX^{o} and the exceptional loci E:=π1(X\Xo)E:=\pi^{-1}(X\backslash X^{o}) is a simple normal crossing divisor. One defines

    S(ICX(𝕍)):=π(S(ICX~(π𝕍)))Rπ(S(ICX~(π𝕍))).\displaystyle S(IC_{X}(\mathbb{V})):=\pi_{\ast}\left(S(IC_{\widetilde{X}}(\pi^{\ast}\mathbb{V}))\right)\simeq R\pi_{\ast}\left(S(IC_{\widetilde{X}}(\pi^{\ast}\mathbb{V}))\right).

    Saito [MSaito1991] shows that S(ICX(𝕍))S(IC_{X}(\mathbb{V})) is independent of the choice of the desingularization π\pi. We will provide another proof of this fact by characterizing S(ICX(𝕍))S(IC_{X}(\mathbb{V})) using L2L^{2} holomorphic sections (Corollary 4.10).

The existence of the SS-sheaf (associated to 𝕍\mathbb{V}) was conjectured by Kollár [Kollar1986], as a generalization of the dualizing sheaf, to admit a package of theorems such as Kollár’s vanishing theorem, torsion freeness and the decomposition theorem. The construction of SS-sheaf and its package of theorems are settled by Saito in [MSaito1991] through his theory of Hodge modules.

2.2. Geometric behavior of the Hodge metric

Let 𝕍=(𝒱,,,h𝕍)\mathbb{V}=(\mathcal{V},\nabla,\mathcal{F}^{\bullet},h_{\mathbb{V}}) be an \mathbb{R}-polarized variation of Hodge structure on (Δ)n×Δn(\Delta^{\ast})^{n}\times\Delta^{n^{\prime}}. Let s1,,sns_{1},\dots,s_{n} be holomorphic coordinates on (Δ)n(\Delta^{\ast})^{n} and denote Di={si=0}Δn+nD_{i}=\{s_{i}=0\}\subset\Delta^{n+n^{\prime}}. Let NiN_{i} be the unipotent part of ResDi{\rm Res}_{D_{i}}\nabla and let

p:n×Δn(Δ)n×Δn,p:\mathbb{H}^{n}\times\Delta^{n^{\prime}}\to(\Delta^{\ast})^{n}\times\Delta^{n^{\prime}},
(z1,,zn,w1,,wn)(e2π1z1,,e2π1zn,w1,,wn)(z_{1},\dots,z_{n},w_{1},\dots,w_{n^{\prime}})\mapsto(e^{2\pi\sqrt{-1}z_{1}},\dots,e^{2\pi\sqrt{-1}z_{n}},w_{1},\dots,w_{n^{\prime}})

be the universal covering. Let W(1)=W(N1),,W(n)=W(N1++Nn)W^{(1)}=W(N_{1}),\dots,W^{(n)}=W(N_{1}+\cdots+N_{n}) be the monodromy weight filtrations on V:=Γ(n×Δn,p𝒱)pV:=\Gamma(\mathbb{H}^{n}\times\Delta^{n^{\prime}},p^{\ast}\mathcal{V})^{p^{\ast}\nabla}. The following important norm estimate for flat sections is proved by Cattani-Kaplan-Schmid in [Cattani_Kaplan_Schmid1986, Theorem 5.21] for the case when 𝕍\mathbb{V} has quasi-unipotent local monodromy and by Mochizuki in [Mochizuki20072, Part 3, Chapter 13] for the general case.

Theorem 2.1.

For any 0vGrlnW(n)Grl1W(1)V0\neq v\in{\rm Gr}_{l_{n}}^{W^{(n)}}\cdots{\rm Gr}_{l_{1}}^{W^{(1)}}V, one has

|v|h𝕍(log|s1|log|s2|)l1(log|sn|)ln\displaystyle|v|_{h_{\mathbb{V}}}\sim\left(\frac{\log|s_{1}|}{\log|s_{2}|}\right)^{l_{1}}\cdots\left(-\log|s_{n}|\right)^{l_{n}}

over any region of the form

{(s1,sn,w1,,wn)(Δ)n×Δn|log|s1|log|s2|>ϵ,,log|sn|>ϵ,(w1,,wn)K}\left\{(s_{1},\dots s_{n},w_{1},\dots,w_{n^{\prime}})\in(\Delta^{\ast})^{n}\times\Delta^{n^{\prime}}\bigg{|}\frac{\log|s_{1}|}{\log|s_{2}|}>\epsilon,\dots,-\log|s_{n}|>\epsilon,(w_{1},\dots,w_{n^{\prime}})\in K\right\}

for any ϵ>0\epsilon>0 and an arbitrary compact subset KΔnK\subset\Delta^{n^{\prime}} .

Lemma 2.2.

Assume that n=1n=1. Then W1(N1)(jS(𝕍)𝒱1)|𝟎=0W_{-1}(N_{1})\cap\big{(}j_{\ast}S(\mathbb{V})\cap\mathcal{V}_{-1}\big{)}|_{\bf 0}=0.

Proof.

Assume that W1(N1)(jS(𝕍)𝒱1)|𝟎0W_{-1}(N_{1})\cap\big{(}j_{\ast}S(\mathbb{V})\cap\mathcal{V}_{-1}\big{)}|_{\bf 0}\neq 0 and let kk be the weight of 𝕍\mathbb{V}. Let l=max{l|Wl(N1)(jS(𝕍)𝒱1)|𝟎0}l=\max\{l|W_{-l}(N_{1})\cap\big{(}j_{\ast}S(\mathbb{V})\cap\mathcal{V}_{-1}\big{)}|_{\bf 0}\neq 0\}. Then l1l\geq 1. By [Schmid1973, 6.16], the filtration j𝒱1j_{\ast}\mathcal{F}^{\bullet}\cap\mathcal{V}_{-1} induces a pure Hodge structure of weight m+km+k on Wm(N1)/Wm1(N1)W_{m}(N_{1})/W_{m-1}(N_{1}). Moreover,

(2.1) Nl:Wl(N1)/Wl1(N1)Wl(N1)/Wl1(N1)\displaystyle N^{l}:W_{l}(N_{1})/W_{l-1}(N_{1})\to W_{-l}(N_{1})/W_{-l-1}(N_{1})

is an isomorphism of type (l,l)(-l,-l). Denote S(𝕍)=pS(\mathbb{V})=\mathcal{F}^{p}. By the definition of ll, any nonzero element αWl(N1)(jS(𝕍)𝒱1)|𝟎\alpha\in W_{-l}(N_{1})\cap\big{(}j_{\ast}S(\mathbb{V})\cap\mathcal{V}_{-1}\big{)}|_{\bf 0} induces a nonzero [α]Wl(N1)/Wl1(N1)[\alpha]\in W_{-l}(N_{1})/W_{-l-1}(N_{1}) of Hodge type (p,klp)(p,k-l-p). Since (2.1) is an isomorphism, there is βWl(N1)/Wl1(N1)\beta\in W_{l}(N_{1})/W_{l-1}(N_{1}) of Hodge type (p+l,kp)(p+l,k-p) such that Nl(β)=[α]N^{l}(\beta)=[\alpha]. However, β=0\beta=0 since p+l=0\mathcal{F}^{p+l}=0. This contradicts to the fact that [α]0[\alpha]\neq 0. Consequently, W1(N1)(jS(𝕍)𝒱1)|𝟎W_{-1}(N_{1})\cap\big{(}j_{\ast}S(\mathbb{V})\cap\mathcal{V}_{-1}\big{)}|_{\bf 0} must be zero. ∎

The following Nakano semi-positivity property of the curvature of S(𝕍)S(\mathbb{V}) enables us to apply Hörmander’s estimate to S(𝕍)S(\mathbb{V}).

Theorem 2.3.

[Schmid1973, Lemma 7.18] Let 𝕍=(𝒱,,,h𝕍)\mathbb{V}=(\mathcal{V},\nabla,\mathcal{F}^{\bullet},h_{\mathbb{V}}) be an \mathbb{R}-polarized variation of Hodge structure over a complex manifold. Then 1Θh𝕍(S(𝕍))0\sqrt{-1}\Theta_{h_{\mathbb{V}}}(S(\mathbb{V}))\geq 0.

2.3. L2L^{2}-adapted local frame

Let X=Δn×ΔnX=\Delta^{n}\times\Delta^{n^{\prime}}, Xo=(Δ)n×ΔnX^{o}=(\Delta^{\ast})^{n}\times\Delta^{n^{\prime}} and let j:XoXj:X^{o}\to X be the open immersion. Denote by z1,,znz_{1},\dots,z_{n} the coordinates on Δn\Delta^{n} and by w1,,wnw_{1},\dots,w_{n^{\prime}} the coordinates on Δn\Delta^{n^{\prime}}. Let Di:={zi=0}XD_{i}:=\{z_{i}=0\}\subset X, i=1,,ni=1,\dots,n. Let 𝕍=(𝒱,,,h)\mathbb{V}=(\mathcal{V},\nabla,\mathcal{F}^{\bullet},h) be an \mathbb{R}-polarized variation of Hodge structure on XoX^{o}. The aim of this subsection is to give the norm estimate of the specific frame of R(ICX(𝕍))R(IC_{X}(\mathbb{V})) at the origin 𝟎=(0,,0){\bf 0}=(0,\dots,0) which is introduced by Deligne [Deligne1970].

Let

p:n×Δn(Δ)n×Δn,p:\mathbb{H}^{n}\times\Delta^{n^{\prime}}\to(\Delta^{\ast})^{n}\times\Delta^{n^{\prime}},
(z1,,zn,w1,,wn)(e2π1z1,,e2π1zn,w1,,wn)(z_{1},\dots,z_{n},w_{1},\dots,w_{n^{\prime}})\mapsto(e^{2\pi\sqrt{-1}z_{1}},\dots,e^{2\pi\sqrt{-1}z_{n}},w_{1},\dots,w_{n^{\prime}})

be the universal covering. For each i=1,,ni=1,\dots,n, let TiT_{i} be the monodromy operators along DiD_{i} and NiN_{i} the unipotent part of ResDi(){\rm Res}_{D_{i}}(\nabla). Since T1,,TnT_{1},\dots,T_{n} are pairwise commutative, there is a finite decomposition

𝒱1|𝟎=1<α1,,αn0𝕍α1,,αn\mathcal{V}_{-1}|_{\bf 0}=\bigoplus_{-1<\alpha_{1},\dots,\alpha_{n}\leq 0}\mathbb{V}_{\alpha_{1},\dots,\alpha_{n}}

such that (Tie2π1αiId)(T_{i}-e^{2\pi\sqrt{-1}\alpha_{i}}{\rm Id}) is unipotent on 𝕍α1,,αn\mathbb{V}_{\alpha_{1},\dots,\alpha_{n}} for each i=1,,ni=1,\dots,n. Let

v1,,vN(𝒱1jS(𝕍))|𝟎1<α1,,αn0𝕍α1,,αnv_{1},\dots,v_{N}\in(\mathcal{V}_{-1}\cap j_{\ast}S(\mathbb{V}))|_{\bf 0}\cap\bigcup_{-1<\alpha_{1},\dots,\alpha_{n}\leq 0}\mathbb{V}_{\alpha_{1},\dots,\alpha_{n}}

be an orthogonal basis of (𝒱1jS(𝕍))|𝟎Γ(n×Δn,pS(𝕍))p(\mathcal{V}_{-1}\cap j_{\ast}S(\mathbb{V}))|_{\bf 0}\simeq\Gamma(\mathbb{H}^{n}\times\Delta^{n^{\prime}},p^{\ast}S(\mathbb{V}))^{p^{\ast}\nabla}. Then v1~,,vN~\widetilde{v_{1}},\dots,\widetilde{v_{N}} that are determined by

(2.2) vj~:=exp(i=1nlogzi(αiId+Ni))vj if vj𝕍α1,,αn,j=1,,N\displaystyle\widetilde{v_{j}}:={\rm exp}\left(\sum_{i=1}^{n}\log z_{i}(\alpha_{i}{\rm Id}+N_{i})\right)v_{j}\textrm{ if }v_{j}\in\mathbb{V}_{\alpha_{1},\dots,\alpha_{n}},\quad\forall j=1,\dots,N

form a frame of 𝒱1jS(𝕍)\mathcal{V}_{-1}\cap j_{\ast}S(\mathbb{V}). To be precise, we always use the notation αDi(vj~)\alpha_{D_{i}}(\widetilde{v_{j}}) instead of αi\alpha_{i} in (2.2). By (2.2) we see that

|vj~|h2\displaystyle|\widetilde{v_{j}}|^{2}_{h} |i=1nziαDi(vj~)exp(i=1nNilogzi)vj|h2\displaystyle\sim\left|\prod_{i=1}^{n}z_{i}^{\alpha_{D_{i}}(\widetilde{v_{j}})}{\rm exp}\left(\sum_{i=1}^{n}N_{i}\log z_{i}\right)v_{j}\right|^{2}_{h}
|vj|h2i=1n|zi|2αDi(vj~),j=1,,N\displaystyle\sim|v_{j}|^{2}_{h}\prod_{i=1}^{n}|z_{i}|^{2\alpha_{D_{i}}(\widetilde{v_{j}})},\quad j=1,\dots,N

where αDi(vj~)(1,0]\alpha_{D_{i}}(\widetilde{v_{j}})\in(-1,0], i=1,,n\forall i=1,\dots,n. By Theorem 2.1 and Lemma 2.2 one has

|vj|h2(log|s1|log|s2|)l1(log|sn|)ln,l1l2ln,\displaystyle|v_{j}|^{2}_{h}\sim\left(\frac{\log|s_{1}|}{\log|s_{2}|}\right)^{l_{1}}\cdots\left(-\log|s_{n}|\right)^{l_{n}},\quad l_{1}\leq l_{2}\leq\dots\leq l_{n},

over any region of the form

{(s1,,sn,w1,,wm)(Δ)n×Δm|log|s1|log|s2|>ϵ,,log|sn|>ϵ,(w1,,wm)K}\left\{(s_{1},\dots,s_{n},w_{1},\dots,w_{m})\in(\Delta^{\ast})^{n}\times\Delta^{m}\bigg{|}\frac{\log|s_{1}|}{\log|s_{2}|}>\epsilon,\dots,-\log|s_{n}|>\epsilon,(w_{1},\dots,w_{m})\in K\right\}

for any ϵ>0\epsilon>0 and an arbitrary compact subset KΔmK\subset\Delta^{m}. Hence we know that

1|vj||z1zn|ϵ,ϵ>0.\displaystyle 1\lesssim|v_{j}|\lesssim|z_{1}\cdots z_{n}|^{-\epsilon},\quad\forall\epsilon>0.

The local frame (v1~,,vN~)(\widetilde{v_{1}},\dots,\widetilde{v_{N}}) is L2L^{2}-adapted in the following sense.

Definition 2.4.

(Zucker [Zucker1979, page 433]) Let (E,h)(E,h) be a vector bundle with a possibly singular hermitian metric hh on a hermitian manifold (X,ds02)(X,ds^{2}_{0}). A holomorphic local frame (v1,,vN)(v_{1},\dots,v_{N}) of EE is called L2L^{2}-adapted if, for every set of measurable functions {f1,,fN}\{f_{1},\dots,f_{N}\}, i=1Nfivi\sum_{i=1}^{N}f_{i}v_{i} is locally square integrable if and only if fivif_{i}v_{i} is locally square integrable for each i=1,,Ni=1,\dots,N.

To see that (v1~,,vN~)(\widetilde{v_{1}},\dots,\widetilde{v_{N}}) is L2L^{2}-adapted, let us consider the measurable functions f1,,fNf_{1},\dots,f_{N}. If

j=1Nfjvj~=exp(i=1nNilogzi)(j=1Nfji=1n|zi|αDi(vj~)vj)\sum_{j=1}^{N}f_{j}\widetilde{v_{j}}={\rm exp}\left(\sum_{i=1}^{n}N_{i}\log z_{i}\right)\left(\sum_{j=1}^{N}f_{j}\prod_{i=1}^{n}|z_{i}|^{\alpha_{D_{i}}(\widetilde{v_{j}})}v_{j}\right)

is locally square integrable, then

j=1Nfji=1n|zi|αDi(vj~)vj\sum_{j=1}^{N}f_{j}\prod_{i=1}^{n}|z_{i}|^{\alpha_{D_{i}}(\widetilde{v_{j}})}v_{j}

is locally square integrable because the entries of the matrix exp(i=1nNilogzi){\rm exp}\left(-\sum_{i=1}^{n}N_{i}\log z_{i}\right) are LL^{\infty}-bounded. Since (v1,,vN)(v_{1},\dots,v_{N}) is an orthogonal basis, |fjvj~|hi=1n|zi|αDi(vj~)|fjvj|h|f_{j}\widetilde{v_{j}}|_{h}\sim\prod_{i=1}^{n}|z_{i}|^{\alpha_{D_{i}}(\widetilde{v_{j}})}|f_{j}v_{j}|_{h} is locally square integrable for each j=1,,Nj=1,\dots,N.

In conclusion, we obtain the following proposition.

Proposition 2.5.

Let (X,ds02)(X,ds^{2}_{0}) be a hermitian manifold and DD a normal crossing divisor on XX. Let 𝕍\mathbb{V} be an \mathbb{R}-polarized variation of Hodge structure on Xo:=X\DX^{o}:=X\backslash D. Then there is an L2L^{2}-adapted holomorphic local frame (v1~,,vN~)(\widetilde{v_{1}},\dots,\widetilde{v_{N}}) of 𝒱1jS(𝕍)\mathcal{V}_{-1}\cap j_{\ast}S(\mathbb{V}) at every point xDx\in D. Let z1,,znz_{1},\cdots,z_{n} be holomorphic local coordinates on XX so that D={z1zr=0}D=\{z_{1}\cdots z_{r}=0\}. Then there are constants αDi(vj~)(1,0]\alpha_{D_{i}}(\widetilde{v_{j}})\in(-1,0], i=1,,ri=1,\dots,r, j=1,,Nj=1,\dots,N and positive real functions λjC(X\D)\lambda_{j}\in C^{\infty}(X\backslash D), j=1,,Nj=1,\dots,N such that

|vj~|2λji=1r|zi|2αDi(vj~),j=1,,N\displaystyle|\widetilde{v_{j}}|^{2}\sim\lambda_{j}\prod_{i=1}^{r}|z_{i}|^{2\alpha_{D_{i}}(\widetilde{v_{j}})},\quad\forall j=1,\dots,N

and

1λj|z1zr|ϵ,ϵ>0,j=1,,N1\lesssim\lambda_{j}\lesssim|z_{1}\cdots z_{r}|^{-\epsilon},\quad\forall\epsilon>0,\quad\forall j=1,\dots,N

3. Preliminary on L2L^{2}-cohomology

3.1. L2L^{2}-Dolbeault cohomology and L2L^{2}-Dolbeault complex

Let (Y,ds2)(Y,ds^{2}) be a hermitian manifold of dimension nn and (E,h)(E,h) a holomorphic vector bundle on YY with a possibly singular hermitian metric. Let 𝒜Yp,q\mathscr{A}^{p,q}_{Y} denote the sheaf of CC^{\infty} (p,q)(p,q)-forms on YY for every 0p,qn0\leq p,q\leq n. Let L(2)p,q(Y,E;ds2,h)L^{p,q}_{(2)}(Y,E;ds^{2},h) be the space of square integrable EE-valued (p,q)(p,q)-forms on YY with respect to the metrics ds2ds^{2} and hh. Denote ¯max\bar{\partial}_{\rm max} to be the maximal extension of the ¯\bar{\partial} operator defined on the domains

Dmaxp,q(Y,E;ds2,h):=Domp,q(¯max)={ϕL(2)p,q(Y,E;ds2,h)|¯ϕL(2)p,q+1(Y,E;ds2,h)}.D^{p,q}_{\rm max}(Y,E;ds^{2},h):=\textrm{Dom}^{p,q}(\bar{\partial}_{\rm max})=\{\phi\in L_{(2)}^{p,q}(Y,E;ds^{2},h)|\bar{\partial}\phi\in L_{(2)}^{p,q+1}(Y,E;ds^{2},h)\}.

Here ¯\bar{\partial} is taken in the sense of distribution. The L2L^{2} cohomology H(2),maxp,(Y,E;ds2,h)H_{(2),\rm max}^{p,\bullet}(Y,E;ds^{2},h) is defined as the cohomology of the complex

Dmaxp,(Y,E;ds2,h):=Dmaxp,0(Y,E;ds2,h)¯max¯maxDmaxp,n(Y,E;ds2,h).\displaystyle D^{p,\bullet}_{\rm max}(Y,E;ds^{2},h):=D^{p,0}_{\rm max}(Y,E;ds^{2},h)\stackrel{{\scriptstyle\bar{\partial}_{\rm max}}}{{\to}}\cdots\stackrel{{\scriptstyle\bar{\partial}_{\rm max}}}{{\to}}D^{p,n}_{\rm max}(Y,E;ds^{2},h).

Let XX be a complex space and XoXregX^{o}\subset X_{\rm reg} a Zariski open subset of the regular locus XregX_{\rm reg}. Let ds2ds^{2} be a hermitian metric on XoX^{o} and (E,h)(E,h) a holomorphic vector bundle on XX with a possibly singular metric. Let UXU\subset X be an open subset. Define LX,ds2p,q(E,h)(U)L_{X,ds^{2}}^{p,q}(E,h)(U) to be the space of measurable EE-valued (p,q)(p,q)-forms α\alpha on UXoU\cap X^{o} such that for every point xUx\in U, there is a neighborhood VxV_{x} of xx so that

VxXo|α|ds2,h2volds2<.\int_{V_{x}\cap X^{o}}|\alpha|^{2}_{ds^{2},h}{\rm vol}_{ds^{2}}<\infty.

For each pp and qq, we define a sheaf 𝒟X,ds2p,q(E,h)\mathscr{D}_{X,ds^{2}}^{p,q}(E,h) on XX by

𝒟X,ds2p,q(E,h)(U):={ϕLX,ds2p,q(E,h)(U)|¯maxϕLX,ds2p,q+1(E,h)(U)}\mathscr{D}_{X,ds^{2}}^{p,q}(E,h)(U):=\{\phi\in L_{X,ds^{2}}^{p,q}(E,h)(U)|\bar{\partial}_{\rm max}\phi\in L_{X,ds^{2}}^{p,q+1}(E,h)(U)\}

for every open subset UXU\subset X. Define the L2L^{2}-Dolbeault complex of sheaves 𝒟X,ds2p,(E,h)\mathscr{D}_{X,ds^{2}}^{p,\bullet}(E,h) as

𝒟X,ds2p,0(E,h)¯𝒟X,ds2p,1(E,h)¯¯𝒟X,ds2p,n(E,h)\displaystyle\mathscr{D}_{X,ds^{2}}^{p,0}(E,h)\stackrel{{\scriptstyle\bar{\partial}}}{{\to}}\mathscr{D}_{X,ds^{2}}^{p,1}(E,h)\stackrel{{\scriptstyle\bar{\partial}}}{{\to}}\cdots\stackrel{{\scriptstyle\bar{\partial}}}{{\to}}\mathscr{D}_{X,ds^{2}}^{p,n}(E,h)

where ¯\bar{\partial} is defined in the sense of distribution.

Definition 3.1.

Let XX be a complex space and ds2ds^{2} a hermitian metric on XregX_{\rm reg}. We say that ds2ds^{2} is a hermitian metric on XX if, for every xXx\in X, there is a neighborhood UU of xx and a holomorphic closed immersion UVU\subset V into a complex manifold such that ds2|UdsV2|Uds^{2}|_{U}\sim ds^{2}_{V}|_{U} for some hermitian metric dsV2ds^{2}_{V} on VV. If ds2|Xregds^{2}|_{X_{\rm reg}} is moreover a Kähler metric, we say that ds2ds^{2} is a Kähler hermitian metric.

Lemma 3.2.

Let XX be a complex space and XoXregX^{o}\subset X_{\rm reg} a Zariski open subset. Let ds2ds^{2} be a hermitian metric on XoX^{o} and (E,h)(E,h) a holomorphic vector bundle on XoX^{o} with a possibly singular hermitian metric. Suppose that, for every point xX\Xox\in X\backslash X^{o}, there is a neighborhood UxU_{x} of xx and a hermitian metric ds02ds^{2}_{0} on UxU_{x} such that ds02|XoUxds2|XoUxds^{2}_{0}|_{X^{o}\cap U_{x}}\lesssim ds^{2}|_{X^{o}\cap U_{x}}. Then 𝒟X,ds2p,q(E,h)\mathscr{D}^{p,q}_{X,ds^{2}}(E,h) is a fine sheaf for each pp and qq.

Proof.

It suffices to show that for every WW¯UXW\subset\overline{W}\subset U\subset X where WW and UU are small open subsets, there is a positive continuous function ff on UU such that

  • supp(f)W¯{\rm supp}(f)\subset\overline{W},

  • ff is CC^{\infty} on UXoU\cap X^{o},

  • ¯f\bar{\partial}f has bounded fiberwise norm, with respect to the metric ds2ds^{2}.

Choose a closed embedding UMU\subset M where MM is a smooth complex manifold. Let VV¯MV\subset\overline{V}\subset M where VV is an open subset such that VU=WV\cap U=W. Let dsM2ds^{2}_{M} be a hermitian metric on MM so that ds02|UXodsM2|UXods^{2}_{0}|_{U\cap X^{o}}\sim ds^{2}_{M}|_{U\cap X^{o}}. Let gg be a positive smooth function on MM whose support lies in V¯\overline{V}. Denote f=g|Uf=g|_{U}. Then supp(f)W¯{\rm supp}(f)\subset\overline{W} and ff is CC^{\infty} on UXoU\cap X^{o}. It suffices to show the boundedness of the fiberwise norm of ¯f\bar{\partial}f. Since UXoMU\cap X^{o}\subset M is a submanifold, one has the orthogonal decomposition

TM,x=TUXo,xTUXo,x,xUXo.\displaystyle T_{M,x}=T_{U\cap X^{o},x}\oplus T_{U\cap X^{o},x}^{\bot},\quad\forall x\in U\cap X^{o}.

Therefore |¯f|ds2|¯f|ds02|¯g|dsM2<|\bar{\partial}f|_{ds^{2}}\lesssim|\bar{\partial}f|_{ds^{2}_{0}}\leq|\bar{\partial}g|_{ds^{2}_{M}}<\infty. The lemma is proved. ∎

The following estimate, which is essentially due to Hörmander in [Hormander1965] and Andreotti-Vesentini in [AV1965] and developed by Demailly in [Demailly1982], plays a crucial role in proving various types of vanishing theorems in the present paper.

Theorem 3.3.

[Demailly1982, Theorem 5.1] Let YY be a complex manifold of dimension nn which admits a complete Kähler metric. Let (E,hφ)(E,h_{\varphi}) be a hermitian vector bundle with a possibly singular hermitian metric hφ:=eφh0h_{\varphi}:=e^{-\varphi}h_{0}. Assume that

1Θhφ(E):=1¯φIdE+1Θh0(E)ωIdE\sqrt{-1}\Theta_{h_{\varphi}}(E):=\sqrt{-1}\partial\bar{\partial}\varphi\otimes{\rm Id}_{E}+\sqrt{-1}\Theta_{h_{0}}(E)\geq\omega\otimes{\rm Id}_{E}

for some (not necessarily complete) Kähler form ω\omega on YY. Then for every q>0q>0 and every αL(2)n,q(Y,E;ω,hφ)\alpha\in L^{n,q}_{(2)}(Y,E;\omega,h_{\varphi}) such that ¯α=0\bar{\partial}\alpha=0, there is βL(2)n,q1(Y,E;ω,hφ)\beta\in L^{n,q-1}_{(2)}(Y,E;\omega,h_{\varphi}) such that ¯β=α\bar{\partial}\beta=\alpha and βω,hφ2q1αω,hφ2\|\beta\|^{2}_{\omega,h_{\varphi}}\leq q^{-1}\|\alpha\|^{2}_{\omega,h_{\varphi}}.

The above theorem works effectively locally on complex analytic singularities due to the following lemma by Grauert [Grauert1956] (see also [Pardon_Stern1991, Lemma 2.4]).

Lemma 3.4.

Let xx be a point of a complex space XX and let XoXregX^{o}\subset X_{\rm reg} be a Zariski open subset. Then there is a neighborhood UU of xx and a complete Kähler metric on UXoU\cap X^{o}.

4. multiplier SS-sheaf

4.1. Adjoint L2L^{2} extension of a hermitian bundle

Let XX be a complex space of dimension nn and XoXregX^{o}\subset X_{\rm reg} a Zariski open subset. Let ds2ds^{2} be a hermitian metric on XoX^{o}.

Definition 4.1.

Let (E,h)(E,h) be a holomorphic vector bundle on XoX^{o} with a possibly singular metric. Define

SX(E,h):=Ker(𝒟X,ds2n,0(E,h)¯𝒟X,ds2n,1(E,h)).\displaystyle S_{X}(E,h):={\rm Ker}\left(\mathscr{D}_{X,ds^{2}}^{n,0}(E,h)\stackrel{{\scriptstyle\bar{\partial}}}{{\to}}\mathscr{D}_{X,ds^{2}}^{n,1}(E,h)\right).

The following proposition shows that SX(E,h)S_{X}(E,h) is independent of ds2ds^{2}. Thus ds2ds^{2} is omitted in the notation SX(E,h)S_{X}(E,h). Such property has already been discovered by Ohsawa in [Ohsawa1980].

Proposition 4.2.

Let (E,h)(E,h) be a holomorphic vector bundle on XoX^{o} with a possibly singular metric. Then SX(E,h)S_{X}(E,h) is independent of ds2ds^{2}.

Proof.

Let π:X~X\pi:\tilde{X}\to X be a desingularization of XX so that X~\widetilde{X} is smooth and π\pi is biholomorphic over XoX^{o}. We identify XoX^{o} and π1(Xo)\pi^{-1}(X^{o}) for simplicity. Let dsX~2ds^{2}_{\tilde{X}} be a hermitian metric on X~\tilde{X}. Since π\pi is a proper map, a section of KXoEK_{X^{o}}\otimes E is locally square integrable at xx if and only if it is locally square integrable near every point of π1{x}\pi^{-1}\{x\}. It is therefore sufficient to show that

(4.1) Ker(𝒟X~,πds2n,0(E,h)¯𝒟X~,πds2n,1(E,h))=Ker(𝒟X~,dsX~2n,0(E,h)¯𝒟X~,dsX~2n,1(E,h)).\displaystyle{\rm Ker}\left(\mathscr{D}_{\tilde{X},\pi^{\ast}ds^{2}}^{n,0}(E,h)\stackrel{{\scriptstyle\bar{\partial}}}{{\to}}\mathscr{D}_{\tilde{X},\pi^{\ast}ds^{2}}^{n,1}(E,h)\right)={\rm Ker}\left(\mathscr{D}_{\tilde{X},ds^{2}_{\tilde{X}}}^{n,0}(E,h)\stackrel{{\scriptstyle\bar{\partial}}}{{\to}}\mathscr{D}_{\tilde{X},ds^{2}_{\tilde{X}}}^{n,1}(E,h)\right).

Since the problem is local, we assume that there is an orthogonal frame of cotangent fields δ1,,δn\delta_{1},\dots,\delta_{n} such that

(4.2) πds2λ1δ1δ1¯++λnδnδn¯\displaystyle\pi^{\ast}ds^{2}\sim\lambda_{1}\delta_{1}\overline{\delta_{1}}+\cdots+\lambda_{n}\delta_{n}\overline{\delta_{n}}

and

(4.3) dsX~2δ1δ1¯++δnδn¯.\displaystyle ds^{2}_{\tilde{X}}\sim\delta_{1}\overline{\delta_{1}}+\cdots+\delta_{n}\overline{\delta_{n}}.

Let s=δ1δnξs=\delta_{1}\wedge\cdots\wedge\delta_{n}\otimes\xi. It follows from (4.2) and (4.3) that

(4.4) sπds2,h2\displaystyle\|s\|^{2}_{\pi^{\ast}ds^{2},h} =|δ1δnξ|πds2,h2i=1nλiδiδi¯\displaystyle=\int|\delta_{1}\wedge\cdots\wedge\delta_{n}\otimes\xi|^{2}_{\pi^{\ast}ds^{2},h}\prod_{i=1}^{n}\lambda_{i}\delta_{i}\wedge\overline{\delta_{i}}
=|ξ|h2i=1nδiδi¯\displaystyle=\int|\xi|^{2}_{h}\prod_{i=1}^{n}\delta_{i}\wedge\overline{\delta_{i}}
=sdsX~2,h2.\displaystyle=\|s\|^{2}_{ds^{2}_{\tilde{X}},h}.

Therefore sπds2,h2\|s\|^{2}_{\pi^{\ast}ds^{2},h} is locally finite if and only if sdsX~2,h2\|s\|^{2}_{ds^{2}_{\tilde{X}},h} is locally finite. This proves (4.1). ∎

Similar to Saito’s SS-sheaf [MSaito1991] and the multipler ideal sheaf [Demailly2012, Proposition 5.8], SX(E,h)S_{X}(E,h) has the functoriality property.

Proposition 4.3 (Functoriality).

Let (E,h)(E,h) be a holomorphic vector bundle on XoX^{o} with a possibly singular hermitian metric. Let π:XX\pi:X^{\prime}\to X be a proper holomorphic map between complex spaces which is biholomorphic over XoX^{o}. Then

πSX(πE,πh)=SX(E,h).\pi_{\ast}S_{X^{\prime}}(\pi^{\ast}E,\pi^{\ast}h)=S_{X}(E,h).
Proof.

By Proposition 4.2,

SX(πE,πh)=Ker(𝒟X,πds2n,0(πE,πh)¯𝒟X,πds2n,1(πE,πh))S_{X^{\prime}}(\pi^{\ast}E,\pi^{\ast}h)={\rm Ker}\left(\mathscr{D}_{X^{\prime},\pi^{\ast}ds^{2}}^{n,0}(\pi^{\ast}E,\pi^{\ast}h)\stackrel{{\scriptstyle\bar{\partial}}}{{\to}}\mathscr{D}_{X^{\prime},\pi^{\ast}ds^{2}}^{n,1}(\pi^{\ast}E,\pi^{\ast}h)\right)

and

SX(E,h)=Ker(𝒟X,ds2n,0(E,h)¯𝒟X,ds2n,1(E,h)).S_{X}(E,h)={\rm Ker}\left(\mathscr{D}_{X,ds^{2}}^{n,0}(E,h)\stackrel{{\scriptstyle\bar{\partial}}}{{\to}}\mathscr{D}_{X,ds^{2}}^{n,1}(E,h)\right).

Since π\pi is a proper map, a section of EE is locally square integrable at xx if and only if it is locally square integrable near each point of π1{x}\pi^{-1}\{x\}. The lemma is proved. ∎

4.2. Multiplier SS-sheaf

Throughout this subsection we assume that XX is a complex space, XoXregX^{o}\subset X_{\rm reg} is a Zariski open subset and 𝕍=(𝒱,,,h𝕍)\mathbb{V}=(\mathcal{V},\nabla,\mathcal{F}^{\bullet},h_{\mathbb{V}}) is an \mathbb{R}-polarized variation of Hodge structure on XoX^{o}. Let φ\varphi be a quasi-psh function on XX.

Definition 4.4.

The multiplier SS-sheaf associated to 𝕍\mathbb{V} and φ\varphi is defined by

S(ICX(𝕍),φ):=SX(S(𝕍),eφh𝕍).S(IC_{X}(\mathbb{V}),\varphi):=S_{X}(S(\mathbb{V}),e^{-\varphi}h_{\mathbb{V}}).
Lemma 4.5.

If φ1φ2\varphi_{1}-\varphi_{2} is locally bounded over XX, then S(ICX(𝕍),φ1)=S(ICX(𝕍),φ2)S(IC_{X}(\mathbb{V}),\varphi_{1})=S(IC_{X}(\mathbb{V}),\varphi_{2}).

Proof.

By assumption, we know that

|α|2eφ1vol|α|2eφ2vol\int|\alpha|^{2}e^{-\varphi_{1}}{\rm vol}\sim\int|\alpha|^{2}e^{-\varphi_{2}}{\rm vol}

for every local section α\alpha of KXoS(𝕍)K_{X^{o}}\otimes S(\mathbb{V}). Thus we prove the lemma. ∎

The proof following lemma is straightforward. Here we omit its proof.

Lemma 4.6.

Let ff be a holomorphic function on Δ12:={z||z|<12}\Delta^{\ast}_{\frac{1}{2}}:=\{z\in\mathbb{C}||z|<\frac{1}{2}\} and aa\in\mathbb{R}. Then

Δ12|f|2|z|2a𝑑z𝑑z¯<\int_{\Delta^{\ast}_{\frac{1}{2}}}|f|^{2}|z|^{2a}dzd\bar{z}<\infty

if and only if v(f)+a>1v(f)+a>-1. Here

v(f):=min{l|fl0 in the Laurent expansion f=ifizi}.v(f):=\min\{l|f_{l}\neq 0\textrm{ in the Laurent expansion }f=\sum_{i\in\mathbb{Z}}f_{i}z^{i}\}.
Definition 4.7.

A quasi-psh function φ\varphi on XX has generalized analytic singularities along a closed analytic subspace ZXZ\subset X if, for every point xZx\in Z, there is a neighborhood UU of xx, some holomorphic functions g1,,gr𝒪X(U)g_{1},\dots,g_{r}\in\mathscr{O}_{X}(U) and some real numbers b0b\geq 0, a1,,ar0a_{1},\dots,a_{r}\geq 0 such that

exp(φ|U)(i=1r|gi|2ai)b.\displaystyle\exp(\varphi|_{U})\sim\left(\sum_{i=1}^{r}|g_{i}|^{2a_{i}}\right)^{b}.

φ\varphi has analytic singularities if ai=1,i=1,,ra_{i}=1,\forall i=1,...,r. When φ\varphi has (generalized) analytic singularities along the entire XX, we briefly say that it has (generalized) analytic singularities.

The following proposition elucidates the relation between S(ICX(𝕍),φ)S(IC_{X}(\mathbb{V}),\varphi) and multiplier ideal sheaves when φ\varphi has generalized analytic singularities.

Proposition 4.8.

Assume that X=ΔnX=\Delta^{n} is the polydisc. Denote E:={z1zr=0}E:=\{z_{1}\cdots z_{r}=0\}, denote Ei:={zi=0},i=1,,rE_{i}:=\{z_{i}=0\},\forall i=1,\dots,r and denote j:Xo:=X\EXj:X^{o}:=X\backslash E\to X to be the inclusion. Let 𝕍=(𝒱,,,h𝕍)\mathbb{V}=(\mathcal{V},\nabla,\mathcal{F}^{\bullet},h_{\mathbb{V}}) be an \mathbb{R}-polarized variation of Hodge structure on XoX^{o} and let φ\varphi be a quasi-psh function on XX which has generalized analytic singularities. Let 𝟎=(0,,0)X{\bf 0}=(0,\dots,0)\in X and let v1~,,vm~\widetilde{v_{1}},\dots,\widetilde{v_{m}} be an L2L^{2}-adapted frame of R(ICX(𝕍))R(IC_{X}(\mathbb{V})) locally at 𝟎{\bf 0} as in Proposition 2.5. Let f1,,fm(j𝒪Xo)𝟎f_{1},\dots,f_{m}\in(j_{\ast}\mathscr{O}_{X^{o}})_{\bf 0}. Then

i=1mfi[vi~dz1dzn]𝟎S(ICX(𝕍),φ)𝟎\sum_{i=1}^{m}f_{i}[\widetilde{v_{i}}dz_{1}\wedge\cdots\wedge dz_{n}]_{\bf 0}\in S(IC_{X}(\mathbb{V}),\varphi)_{\bf 0}

if and only if

fi(φj=1r2αEj(vi~)log|zj|)𝟎f_{i}\in\mathscr{I}\big{(}\varphi-\sum_{j=1}^{r}2\alpha_{E_{j}}(\widetilde{v_{i}})\log|z_{j}|\big{)}_{\bf 0}

for every i=1,,mi=1,\dots,m. In conclusion, there is an isomorphism

S(ICX(𝕍),φ)𝟎ωX,𝟎i=1m(φj=1r2αEj(vi~)log|zj|)𝟎vi~.\displaystyle S(IC_{X}(\mathbb{V}),\varphi)_{\bf 0}\simeq\omega_{X,{\bf 0}}\otimes\bigoplus_{i=1}^{m}\mathscr{I}(\varphi-\sum_{j=1}^{r}2\alpha_{E_{j}}(\widetilde{v_{i}})\log|z_{j}|)_{\bf 0}\widetilde{v_{i}}.
Proof.

Denote ds02=i=1ndzidz¯ids^{2}_{0}=\sum_{i=1}^{n}dz_{i}d\bar{z}_{i}. Since v1~,,vm~\widetilde{v_{1}},\dots,\widetilde{v_{m}} is an L2L^{2}-adapted frame as in Proposition 2.5, the integral

|i=1mfivi~dz1dzn|2eφvolds02=|i=1mfivi~|2eφvolds02\int|\sum_{i=1}^{m}f_{i}\widetilde{v_{i}}dz_{1}\wedge\cdots\wedge dz_{n}|^{2}e^{-\varphi}{\rm vol}_{ds^{2}_{0}}=\int|\sum_{i=1}^{m}f_{i}\widetilde{v_{i}}|^{2}e^{-\varphi}{\rm vol}_{ds^{2}_{0}}

is finite near 𝟎{\bf 0} if and only if

|fivi~|2eφvolds02|fi|2eφj=1r|zj|2αEj(vi~)λivolds02\int|f_{i}\widetilde{v_{i}}|^{2}e^{-\varphi}{\rm}{\rm vol}_{ds^{2}_{0}}\sim\int|f_{i}|^{2}e^{-\varphi}{\rm}\prod_{j=1}^{r}|z_{j}|^{2\alpha_{E_{j}}(\widetilde{v_{i}})}\lambda_{i}{\rm vol}_{ds^{2}_{0}}

is finite near 𝟎{\bf 0} for every i=1,,mi=1,\dots,m. Here λi\lambda_{i} is a positive real function so that

(4.5) 1λi|z1zr|ϵ,ϵ>0.\displaystyle 1\lesssim\lambda_{i}\lesssim|z_{1}\cdots z_{r}|^{-\epsilon},\quad\forall\epsilon>0.

We are going to show that |fivi~|2eφvolds02|f_{i}\widetilde{v_{i}}|^{2}e^{-\varphi}{\rm}{\rm vol}_{ds^{2}_{0}} is locally integrable if and only if fi(φj=1r2αEj(vi~)log|zi|)f_{i}\in\mathscr{I}\big{(}\varphi-\sum_{j=1}^{r}2\alpha_{E_{j}}(\widetilde{v_{i}})\log|z_{i}|\big{)}.

Since φ\varphi has generalized analytic singularities, Z:={eφ=0}Z:=\{e^{\varphi}=0\} is a closed analytic subspace. Let π:X~X\pi:\widetilde{X}\to X be a desingularization so that

  1. (1)

    π\pi is biholomorphic over Xo\ZX^{o}\backslash Z^{\prime} where ZZ^{\prime} is the union of the irreducible components of ZZ that is not a component of EE.

  2. (2)

    π1(ZE)\pi^{-1}(Z\cup E) and π1(E)\pi^{-1}(E) are normal crossing divisors on X~\widetilde{X}.

Let w1,,wnw_{1},\dots,w_{n} be holomorphic local coordinates of X~\widetilde{X} such that π1(ZE)={w1ws=0}\pi^{-1}(Z\cup E)=\{w_{1}\cdots w_{s}=0\}. Then we obtain that

eπφi=1s|wi|2ai,π(j=1r|zj|2αEj(vi~))i=1s|wi|2bi\displaystyle e^{\pi^{\ast}\varphi}\sim\prod_{i=1}^{s}|w_{i}|^{2a_{i}},\quad\pi^{\ast}\left(\prod_{j=1}^{r}|z_{j}|^{2\alpha_{E_{j}}(\widetilde{v_{i}})}\right)\sim\prod_{i=1}^{s}|w_{i}|^{-2b_{i}}

for some nonnegative constants a1,,as,b1,,bsa_{1},\dots,a_{s},b_{1},\dots,b_{s}. Let ds12ds^{2}_{1} be a hermitian metric on X~\widetilde{X}. Then

πvolds02i=1s|wi|2civolds12\pi^{\ast}{\rm vol}_{ds^{2}_{0}}\sim\prod_{i=1}^{s}|w_{i}|^{2c_{i}}{\rm vol}_{ds^{2}_{1}}

for some nonnegative constants c1,,csc_{1},\dots,c_{s}. If fi(φj=1r2αEj(vi~)log|zj|)f_{i}\in\mathscr{I}\big{(}\varphi-\sum_{j=1}^{r}2\alpha_{E_{j}}(\widetilde{v_{i}})\log|z_{j}|\big{)}, then

(4.6) π(|fi|2eφ+j=1r2αEj(vi~)log|zj|volds02)|πfi|2i=1s|wi|2(aibi+ci)volds12\displaystyle\int\pi^{\ast}\left(|f_{i}|^{2}e^{-\varphi+\sum_{j=1}^{r}2\alpha_{E_{j}}(\widetilde{v_{i}})\log|z_{j}|}{\rm vol}_{ds^{2}_{0}}\right)\sim\int|\pi^{\ast}f_{i}|^{2}\prod_{i=1}^{s}|w_{i}|^{2(-a_{i}-b_{i}+c_{i})}{\rm vol}_{ds^{2}_{1}}

is locally integrable. Denote

vj(g):=min{l|gl0 in the Laurent expansion g=igiwji},j=1,,s.v_{j}(g):=\min\{l|g_{l}\neq 0\textrm{ in the Laurent expansion }g=\sum_{i\in\mathbb{Z}}g_{i}w_{j}^{i}\},\quad j=1,\dots,s.

By Lemma 4.6, the local integrability of (4.6) implies that

vj(πfi)ajbj+cj>1+ϵ,j=1,,s\displaystyle v_{j}(\pi^{\ast}f_{i})-a_{j}-b_{j}+c_{j}>-1+\epsilon,\quad j=1,\dots,s

for some ϵ>0\epsilon>0. By (4.5), we obtain that

π(|fivi~|2eφvolds02)\displaystyle\int\pi^{\ast}\left(|f_{i}\widetilde{v_{i}}|^{2}e^{-\varphi}{\rm}{\rm vol}_{ds^{2}_{0}}\right) |πfi|2j=1s|wj|2(ajbj+cj)πλivolds12\displaystyle\sim\int|\pi^{\ast}f_{i}|^{2}\prod_{j=1}^{s}|w_{j}|^{2(-a_{j}-b_{j}+c_{j})}\pi^{\ast}\lambda_{i}{\rm vol}_{ds^{2}_{1}}
|πfi|2j=1s|wj|2(ajbj+cjϵ)volds12\displaystyle\lesssim\int|\pi^{\ast}f_{i}|^{2}\prod_{j=1}^{s}|w_{j}|^{2(-a_{j}-b_{j}+c_{j}-\epsilon)}{\rm vol}_{ds^{2}_{1}}

is locally integrable. Since π\pi is a proper map, we see that |fivi~|2eφvolds02\int|f_{i}\widetilde{v_{i}}|^{2}e^{-\varphi}{\rm}{\rm vol}_{ds^{2}_{0}} is locally integrable.

Conversely, let fif_{i} be a holomorphic function on XoX^{o} such that |fivi~|2eφvolds02\int|f_{i}\widetilde{v_{i}}|^{2}e^{-\varphi}{\rm}{\rm vol}_{ds^{2}_{0}} is locally integrable. We know that

|fi|2eφ+j=1r2αEj(vi~)log|zj|volds02|fi|2eφj=1r|zj|2αEj(vi~)λivolds02|fivi~|2eφvolds02\displaystyle\int|f_{i}|^{2}e^{-\varphi+\sum_{j=1}^{r}2\alpha_{E_{j}}(\widetilde{v_{i}})\log|z_{j}|}{\rm vol}_{ds^{2}_{0}}\lesssim\int|f_{i}|^{2}e^{-\varphi}{\rm}\prod_{j=1}^{r}|z_{j}|^{2\alpha_{E_{j}}(\widetilde{v_{i}})}\lambda_{i}{\rm vol}_{ds^{2}_{0}}\sim\int|f_{i}\widetilde{v_{i}}|^{2}e^{-\varphi}{\rm}{\rm vol}_{ds^{2}_{0}}

is locally integrable by (4.5). The proof is finished. ∎

Proposition 4.9.

Assume that XX is a complex manifold and X\XoXX\backslash X^{o}\subset X is a (possibly empty) normal crossing divisor. Let 𝕍\mathbb{V} be an \mathbb{R}-polarized variation of Hodge structure on XoX^{o} which has unipotent local monodromies. Let φ\varphi be a psh function with generalized analytic singularities along X\XoX\backslash X^{o}. Then there is an isomorphism

S(ICX(𝕍),φ)S(ICX(𝕍))(φ).S(IC_{X}(\mathbb{V}),\varphi)\simeq S(IC_{X}(\mathbb{V}))\otimes\mathscr{I}(\varphi).
Proof.

Since S(ICX(𝕍))S(IC_{X}(\mathbb{V})) is locally free, it suffices to show that

(4.7) S(ICX(𝕍),φ)=(φ)S(ICX(𝕍))\displaystyle S(IC_{X}(\mathbb{V}),\varphi)=\mathscr{I}(\varphi)S(IC_{X}(\mathbb{V}))

as subsheaves of j(ωXoS(𝕍))j_{\ast}(\omega_{X^{o}}\otimes S(\mathbb{V})). Here we regard S(ICX(𝕍)):=(j(S(𝕍))𝒱1)ωXoS(IC_{X}(\mathbb{V})):=(j_{\ast}(S(\mathbb{V}))\cap\mathcal{V}_{-1})\otimes\omega_{X^{o}} as a subsheaf of j(ωXoS(𝕍))j_{\ast}(\omega_{X^{o}}\otimes S(\mathbb{V})).

Since the problem is local, we assume that XnX\subset\mathbb{C}^{n} is a germ of open subset at 𝟎=(0,,0){\bf 0}=(0,\dots,0) with the standard coordinates z1,,znz_{1},...,z_{n} such that E:=X\Xo={z1zr=0}E:=X\backslash X^{o}=\{z_{1}\cdots z_{r}=0\}. Denote Ei:={zi=0},i=1,,rE_{i}:=\{z_{i}=0\},\forall i=1,\dots,r. Let v1~,,vm~\widetilde{v_{1}},\dots,\widetilde{v_{m}} be an L2L^{2}-adapted frame of R(ICX(𝕍))R(IC_{X}(\mathbb{V})) locally at 𝟎{\bf 0} as in Proposition 2.5 and let f1,,fm(j𝒪Xo)𝟎f_{1},\dots,f_{m}\in(j_{\ast}\mathscr{O}_{X^{o}})_{\bf 0}. Proposition 4.8 tells us that

i=1mfi[vi~dz1dzn]xS(ICX(𝕍),φ)𝟎\sum_{i=1}^{m}f_{i}[\widetilde{v_{i}}dz_{1}\wedge\cdots\wedge dz_{n}]_{x}\in S(IC_{X}(\mathbb{V}),\varphi)_{\bf 0}

if and only if

fi(φj=1r2αEj(vi~)log|zj|)𝟎,i=1,,m.f_{i}\in\mathscr{I}\big{(}\varphi-\sum_{j=1}^{r}2\alpha_{E_{j}}(\widetilde{v_{i}})\log|z_{j}|\big{)}_{\bf 0},\quad i=1,\dots,m.

Since the local monodromies of 𝕍\mathbb{V} are unipotent, αEj(vi~)=0\alpha_{E_{j}}(\widetilde{v_{i}})=0, j=1,,r\forall j=1,\dots,r. Hence we prove (4.7). ∎

Theorem 4.10.

S(ICX(𝕍),0)S(ICX(𝕍))S(IC_{X}(\mathbb{V}),0)\simeq S(IC_{X}(\mathbb{V})). In particular, S(ICX(𝕍))S(IC_{X}(\mathbb{V})) is independent of the choice of the desingularization.

Proof.

Let X~X\widetilde{X}\to X be a desingularization so that π\pi is biholomorphic over XoX^{o} and π1(X\Xo)\pi^{-1}(X\backslash X^{o}) is a simple normal crossing divisor. We claim that

(4.8) S(ICX~(π𝕍),0)S(ICX~(π𝕍)).\displaystyle S(IC_{\widetilde{X}}(\pi^{\ast}\mathbb{V}),0)\simeq S(IC_{\widetilde{X}}(\pi^{\ast}\mathbb{V})).

If the claim is true, it follows from Proposition 4.3 that

S(ICX(𝕍),0)πS(ICX~(π𝕍),0)πS(ICX~(π𝕍))S(ICX(𝕍)).S(IC_{X}(\mathbb{V}),0)\simeq\pi_{\ast}S(IC_{\widetilde{X}}(\pi^{\ast}\mathbb{V}),0)\simeq\pi_{\ast}S(IC_{\widetilde{X}}(\pi^{\ast}\mathbb{V}))\simeq S(IC_{X}(\mathbb{V})).

Now it suffices to show (4.8). Since the problem is local, we assume that X~=Δn\widetilde{X}=\Delta^{n} is the polydisc with the standard holomorphic coordinates z1,,znz_{1},\dots,z_{n} such that E:=π1(X\Xo)={z1zr=0}E:=\pi^{-1}(X\backslash X^{o})=\{z_{1}\cdots z_{r}=0\}. Let j:X~\EX~j:\widetilde{X}\backslash E\to\widetilde{X} denote the inclusion and denote Ei:={zi=0},i=1,,rE_{i}:=\{z_{i}=0\},\forall i=1,\dots,r. Let 𝟎=(0,,0)X~{\bf 0}=(0,\dots,0)\in\widetilde{X} and let v1~,,vm~\widetilde{v_{1}},\dots,\widetilde{v_{m}} be an L2L^{2}-adapted frame of R(ICX~(π𝕍))R(IC_{\widetilde{X}}(\pi^{\ast}\mathbb{V})) locally at 𝟎{\bf 0} as in Proposition 2.5. Let f1,,fm(j𝒪π1(Xo))𝟎f_{1},\dots,f_{m}\in(j_{\ast}\mathscr{O}_{\pi^{-1}(X^{o})})_{\bf 0}. Proposition 4.8 shows that

i=1mfi[vi~dz1dzn]𝟎S(ICX(𝕍),0)𝟎\sum_{i=1}^{m}f_{i}[\widetilde{v_{i}}dz_{1}\wedge\cdots\wedge dz_{n}]_{\bf 0}\in S(IC_{X}(\mathbb{V}),0)_{\bf 0}

if and only if

fi(j=1r2αEj(vi~)log|zj|)𝟎f_{i}\in\mathscr{I}\big{(}-\sum_{j=1}^{r}2\alpha_{E_{j}}(\widetilde{v_{i}})\log|z_{j}|\big{)}_{\bf 0}

for every i=1,,mi=1,\dots,m. Since αEj(vi~)(1,0],j=1,,r\alpha_{E_{j}}(\widetilde{v_{i}})\in(-1,0],\forall j=1,\dots,r, Lemma 4.6 shows that exp(j=1r2αEj(vi~)log|zj|){\rm exp}\left(\sum_{j=1}^{r}2\alpha_{E_{j}}(\widetilde{v_{i}})\log|z_{j}|\right) is integrable. Hence

(j=1r2αEj(vi~)log|zj|)𝟎𝒪X~,𝟎,i=1,,m.\displaystyle\mathscr{I}\big{(}-\sum_{j=1}^{r}2\alpha_{E_{j}}(\widetilde{v_{i}})\log|z_{j}|\big{)}_{\bf 0}\simeq\mathscr{O}_{\widetilde{X},\bf 0},\quad i=1,\dots,m.

The proof of the theorem is finished. ∎

Proposition 4.11.

S(ICX(𝕍),φ)S(IC_{X}(\mathbb{V}),\varphi) is a coherent subsheaf of S(ICX(𝕍))S(IC_{X}(\mathbb{V})) for an arbitrary general quasi-psh function φ\varphi.

Proof.

First we show that S(ICX(𝕍),φ)S(ICX(𝕍))S(IC_{X}(\mathbb{V}),\varphi)\subset S(IC_{X}(\mathbb{V})). By Theorem 4.10, it is equivalent to show that

(4.9) S(ICX(𝕍),φ)S(ICX(𝕍),0).\displaystyle S(IC_{X}(\mathbb{V}),\varphi)\subset S(IC_{X}(\mathbb{V}),0).

Since φ\varphi is quasi-psh, it has an upper bound cxc_{x} locally at every point xXx\in X. Thus we have

|α|2vol|α|2ecxvol|α|2eφvol\int|\alpha|^{2}{\rm vol}\sim\int|\alpha|^{2}e^{-c_{x}}{\rm vol}\leq\int|\alpha|^{2}e^{-\varphi}{\rm vol}

locally at xx. This proves (4.9).

Next we prove its coherence. Let π:X~X\pi:\widetilde{X}\to X be a desingularization so that π\pi is biholomorphic over XoX^{o} and E:=π1(X\Xo)E:=\pi^{-1}(X\backslash X^{o}) is a simple normal crossing divisor. By abuse of notations we regard XoX~X^{o}\subset\widetilde{X} as a subset. Denote ψ=πφ\psi=\pi^{\ast}\varphi. Then ψ\psi is a quasi-psh function. There is an isomorphism

S(ICX(𝕍),φ)π(S(ICX~(𝕍),ψ))\displaystyle S(IC_{X}(\mathbb{V}),\varphi)\simeq\pi_{\ast}\left(S(IC_{\widetilde{X}}(\mathbb{V}),\psi)\right)

by Proposition 4.3. Since π\pi is proper, it suffices to show that S(ICX~(𝕍),ψ)S(IC_{\widetilde{X}}(\mathbb{V}),\psi) is a coherent sheaf on X~\widetilde{X}. Since the problem is local and X~\widetilde{X} is smooth, we may assume that X~n\widetilde{X}\subset\mathbb{C}^{n} is the unit ball so that E={z1zs=0}E=\{z_{1}\cdots z_{s}=0\} and ψ\psi has an upper bound. Denote Ei={zi=0},i=1,,sE_{i}=\{z_{i}=0\},\forall i=1,\dots,s. Notice that there is a complete Kähler metric on XoX^{o} by Lemma 3.4. Since S(ICX~(𝕍))S(IC_{\widetilde{X}}(\mathbb{V})) is coherent, the space Γ(X~,S(ICX~(𝕍),ψ))\Gamma(\widetilde{X},S(IC_{\widetilde{X}}(\mathbb{V}),\psi)) generates a coherent subsheaf 𝒥\mathscr{J} of S(ICX~(𝕍))S(IC_{\widetilde{X}}(\mathbb{V})) by the strong Noetherian property. By the construction we have the inclusion 𝒥S(ICX~(𝕍),ψ)\mathscr{J}\subset S(IC_{\widetilde{X}}(\mathbb{V}),\psi). It remains to prove the converse. By Krull’s theorem ([Atiyah1969, Corollary 10.19]), it suffices to show that

(4.10) 𝒥x+S(ICX~(𝕍),ψ)xmX~,xk+1S(ICX~(𝕍))=S(ICX~(𝕍),ψ)x,k0.\displaystyle\mathscr{J}_{x}+S(IC_{\widetilde{X}}(\mathbb{V}),\psi)_{x}\cap m_{\widetilde{X},x}^{k+1}S(IC_{\widetilde{X}}(\mathbb{V}))=S(IC_{\widetilde{X}}(\mathbb{V}),\psi)_{x},\quad\forall k\geq 0.

Let αS(ICX~(𝕍),ψ)x\alpha\in S(IC_{\widetilde{X}}(\mathbb{V}),\psi)_{x} be defined in a precompact neighborhood VV of xx. Choose a CC^{\infty} cut-off function λ\lambda so that λ1\lambda\equiv 1 near xx and suppλV{\rm supp}\lambda\subset V. Denote |z|2:=i=1n|zi|2|z|^{2}:=\sum_{i=1}^{n}|z_{i}|^{2}. Since ψ\psi is quasi-psh, there is a constant c>0c>0 such that ψ(z)+c|z|2\psi(z)+c|z|^{2} is psh. Let

ψk(z):=ψ(z)+2(n+k+1)log|zx|+(c+1)|z|2\displaystyle\psi_{k}(z):=\psi(z)+2(n+k+1)\log|z-x|+(c+1)|z|^{2}

and hψk=eψkh𝕍h_{\psi_{k}}=e^{-\psi_{k}}h_{\mathbb{V}}. Denote ω0:=1¯|z|2\omega_{0}:=\sqrt{-1}\partial\bar{\partial}|z|^{2}. By Theorem 2.3, we have

1Θhψk(S(𝕍))=1¯ψk+1Θh𝕍(S(𝕍))ω0.\displaystyle\sqrt{-1}\Theta_{h_{\psi_{k}}}(S(\mathbb{V}))=\sqrt{-1}\partial\bar{\partial}\psi_{k}+\sqrt{-1}\Theta_{h_{\mathbb{V}}}(S(\mathbb{V}))\geq\omega_{0}.

Since suppλαV{\rm supp}\lambda\alpha\subset V and ¯(λα)=0\bar{\partial}(\lambda\alpha)=0 near xx, we have

¯(λα)ω0,hψk2¯(λα)ω0,hψ2¯λL2αω0,hψ2+|λ|2¯αω0,hψ2<\displaystyle\|\bar{\partial}(\lambda\alpha)\|^{2}_{\omega_{0},h_{\psi_{k}}}\sim\|\bar{\partial}(\lambda\alpha)\|^{2}_{\omega_{0},h_{\psi}}\leq\|\bar{\partial}\lambda\|^{2}_{L^{\infty}}\|\alpha\|^{2}_{\omega_{0},h_{\psi}}+|\lambda|^{2}\|\bar{\partial}\alpha\|^{2}_{\omega_{0},h_{\psi}}<\infty

Hence, Theorem 3.3 provides a solution of the equation ¯β=¯(λα)\bar{\partial}\beta=\bar{\partial}(\lambda\alpha) so that

(4.11) βω0,hψ2Xo|β|ω0,h𝕍2eψ|zx|2(n+k+1)volω0¯(λα)ω0,hψk2<.\displaystyle\|\beta\|^{2}_{\omega_{0},h_{\psi}}\lesssim\int_{X^{o}}|\beta|^{2}_{\omega_{0},h_{\mathbb{V}}}e^{-\psi}|z-x|^{-2(n+k+1)}{\rm vol}_{\omega_{0}}\lesssim\|\bar{\partial}(\lambda\alpha)\|^{2}_{\omega_{0},h_{\psi_{k}}}<\infty.

Thus γ=λαβ\gamma=\lambda\alpha-\beta is holomorphic and γΓ(X~,S(ICX~(𝕍),ψ))\gamma\in\Gamma(\widetilde{X},S(IC_{\widetilde{X}}(\mathbb{V}),\psi)). Using the notations in §2, we have

S(ICX~(𝕍))=(j(S(𝕍))𝒱~1)ωX~\displaystyle S(IC_{\widetilde{X}}(\mathbb{V}))=\left(j_{\ast}(S(\mathbb{V}))\cap\widetilde{\mathcal{V}}_{-1}\right)\otimes\omega_{\widetilde{X}}

where j:XoX~j:X^{o}\to\widetilde{X} is the open immersion. Since ψ\psi has an upper bound, we have

βω0,h𝕍2Xo|β|ω0,h𝕍2|zx|2(n+k+1)volω0Xo|β|ω0,h𝕍2eψ|zx|2(n+k+1)volω0<.\|\beta\|^{2}_{\omega_{0},h_{\mathbb{V}}}\lesssim\int_{X^{o}}|\beta|^{2}_{\omega_{0},h_{\mathbb{V}}}|z-x|^{-2(n+k+1)}{\rm vol}_{\omega_{0}}\lesssim\int_{X^{o}}|\beta|^{2}_{\omega_{0},h_{\mathbb{V}}}e^{-\psi}|z-x|^{-2(n+k+1)}{\rm vol}_{\omega_{0}}<\infty.

Hence βS(ICX~(𝕍))𝒱~1\beta\in S(IC_{\widetilde{X}}(\mathbb{V}))\subset\widetilde{\mathcal{V}}_{-1} by Theorem 4.10.

Let 𝐯~=(v~1,,v~r)\widetilde{{\bf v}}=(\widetilde{v}_{1},\dots,\widetilde{v}_{r}) be the local frame of R(ICX~(𝕍))𝒱~1R(IC_{\widetilde{X}}(\mathbb{V}))\subset\widetilde{\mathcal{V}}_{-1} where 𝐯=ei2π1ziResEi~𝐯~{\bf v}=e^{-\sum_{i}2\pi\sqrt{-1}z_{i}{\rm Res}_{E_{i}}\widetilde{\nabla}}\widetilde{{\bf v}} is an orthogonal basis of 𝕍\mathbb{V} (c.f. §2.3). Then there are holormorphic functions f1,,fr𝒪X~(Xo)f_{1},\dots,f_{r}\in\mathscr{O}_{\widetilde{X}}(X^{o}) such that

β=i=1rfiv~idz1dzn.\beta=\sum_{i=1}^{r}f_{i}\widetilde{v}_{i}dz_{1}\wedge\cdots\wedge dz_{n}.

It follows from Theorem 2.1 that

(z1zs)ϵ|vi|,ϵ>0,i=1,,r.\displaystyle(z_{1}\cdots z_{s})^{\epsilon}\lesssim|v_{i}|,\quad\forall\epsilon>0,\forall i=1,\dots,r.

Thus one gets that

Xoi=1r|fi|2(z1zr)2ϵ|zx|2(n+k+1)volω0Xo|β|ω0,h𝕍2|zx|2(n+k+1)volω0<\displaystyle\int_{X^{o}}\sum_{i=1}^{r}|f_{i}|^{2}(z_{1}\cdots z_{r})^{2\epsilon}|z-x|^{-2(n+k+1)}{\rm vol}_{\omega_{0}}\lesssim\int_{X^{o}}|\beta|^{2}_{\omega_{0},h_{\mathbb{V}}}|z-x|^{-2(n+k+1)}{\rm vol}_{\omega_{0}}<\infty

for every ϵ>0\epsilon>0. This implies that fimX~,xk+1f_{i}\in m_{\widetilde{X},x}^{k+1} for every i=1,,ri=1,\dots,r ([Demailly2012, Lemma 5.6]). Hence

αxγx=βxmX~,xk+1S(ICX~(𝕍)).\alpha_{x}-\gamma_{x}=\beta_{x}\in m_{\widetilde{X},x}^{k+1}S(IC_{\widetilde{X}}(\mathbb{V})).

On the other hand, βS(ICX~(𝕍),ψ)\beta\in S(IC_{\widetilde{X}}(\mathbb{V}),\psi) because of (4.11). This proves (4.10). ∎

Lemma 4.12.

Let (E,hφ)(E,h_{\varphi}) be a holomorphic vector bundle on XX with a possibly singular hermitian metric hφ=eφh0h_{\varphi}=e^{-\varphi}h_{0}. Then S(ICX(𝕍),φ)ESX(S(𝕍)E,h𝕍hφ)S(IC_{X}(\mathbb{V}),\varphi)\otimes E\simeq S_{X}(S(\mathbb{V})\otimes E,h_{\mathbb{V}}\otimes h_{\varphi}).

Proof.

Let xXx\in X be a point and let UU be an open neighborhood of xx so that E|U𝒪UrE|_{U}\simeq\mathscr{O}_{U}^{\oplus r} and h0h_{0} is quasi-isometric to the trivial metric, i.e.

|i=1raiei|h02i=1r|ai|2|\sum_{i=1}^{r}a_{i}e_{i}|^{2}_{h_{0}}\sim\sum_{i=1}^{r}|a_{i}|^{2}

where {ei}\{e_{i}\} is the standard frame of 𝒪Ur\mathscr{O}_{U}^{\oplus r} and aia_{i}s are measurable functions on UU. Let ds2ds^{2} be an arbitrary hermitian metric on XoX^{o} and let α=i=1rαiei\alpha=\sum_{i=1}^{r}\alpha_{i}\otimes e_{i} be a measurable section of S(𝕍)E|UXoS(\mathbb{V})\otimes E|_{U\cap X^{o}}. Then

αds2,h𝕍hφ2i=1rαids2,eφh𝕍2\|\alpha\|^{2}_{ds^{2},h_{\mathbb{V}}\otimes h_{\varphi}}\sim\sum_{i=1}^{r}\|\alpha_{i}\|^{2}_{ds^{2},e^{-\varphi}h_{\mathbb{V}}}

is finite if and only if each αids2,eφh𝕍2\|\alpha_{i}\|^{2}_{ds^{2},e^{-\varphi}h_{\mathbb{V}}} is finite. This proves the lemma. ∎

We end this section by proving an approximation property of S(ICX(𝕍),φ)S(IC_{X}(\mathbb{V}),\varphi).

Proposition 4.13.

Let φ\varphi and ψ\psi be quasi-psh functions on XX which have generalized analytic singularities. Then

S(ICX(𝕍),φ)=S(ICX(𝕍),φ+ϵψ),0<ϵ1.S(IC_{X}(\mathbb{V}),\varphi)=S(IC_{X}(\mathbb{V}),\varphi+\epsilon\psi),\quad 0<\epsilon\ll 1.
Proof.

Since the problem is local, we assume that XX is a germ of complex space. Since φ\varphi and ψ\psi have generalized analytic singularities, Z1:={eφ=0}Z_{1}:=\{e^{\varphi}=0\} and Z2:={eψ=0}Z_{2}:=\{e^{\psi}=0\} are closed analytic subspaces. Let π:X~X\pi:\widetilde{X}\to X be a desingularization so that

  1. (1)

    π\pi is biholomorphic over Xo\(Z1Z2)X^{o}\backslash(Z_{1}\cup Z_{2});

  2. (2)

    π1(Z1Z2(X\Xo))\pi^{-1}(Z_{1}\cup Z_{2}\cup(X\backslash X^{o})) and π1(X\Xo)\pi^{-1}(X\backslash X^{o}) are normal crossing divisors of X~\widetilde{X}.

Let w1,,wnw_{1},\dots,w_{n} be holomorphic local coordinates on X~\widetilde{X} such that π1(Z1Z2(X\Xo))={w1ws=0}\pi^{-1}(Z_{1}\cup Z_{2}\cup(X\backslash X^{o}))=\{w_{1}\cdots w_{s}=0\} and π1(X\Xo)={w1wr=0}\pi^{-1}(X\backslash X^{o})=\{w_{1}\cdots w_{r}=0\} with 0rs0\leq r\leq s. Notice that r=0r=0 if we consider the problem on XoX^{o}. Denote Ei={zi=0},i=1,,rE_{i}=\{z_{i}=0\},\forall i=1,\dots,r. Then we know that

eπφi=1s|wi|2ai,eπψi=1s|wi|2bi\displaystyle e^{\pi^{\ast}\varphi}\sim\prod_{i=1}^{s}|w_{i}|^{2a_{i}},\quad e^{\pi^{\ast}\psi}\sim\prod_{i=1}^{s}|w_{i}|^{2b_{i}}

for some nonnegative constants a1,,as,b1,,bsa_{1},\dots,a_{s},b_{1},\dots,b_{s}. Denote j:π1(Xo)X~j:\pi^{-1}(X^{o})\to\widetilde{X} to be the open immersion. Let v1~,,vm~\widetilde{v_{1}},\dots,\widetilde{v_{m}} be an L2L^{2}-adapted frame of R(ICX~(π𝕍))R(IC_{\widetilde{X}}(\pi^{\ast}\mathbb{V})) locally at 𝟎=(0,,0)X~{\bf 0}=(0,\dots,0)\in\widetilde{X} as in Proposition 2.5. Let f1,,fm(j𝒪π1(Xo))𝟎f_{1},\dots,f_{m}\in(j_{\ast}\mathscr{O}_{\pi^{-1}(X^{o})})_{\bf 0} and α=i=1mfivi~dz1dzn\alpha=\sum_{i=1}^{m}f_{i}\widetilde{v_{i}}dz_{1}\wedge\cdots\wedge dz_{n}. Then Proposition 4.8 shows that

[α]xS(ICX~(π𝕍),πφ)𝟎(resp. [α]xS(ICX~(π𝕍),πφ+πψ)𝟎)[\alpha]_{x}\in S(IC_{\widetilde{X}}(\pi^{\ast}\mathbb{V}),\pi^{\ast}\varphi)_{\bf 0}\quad\left(\textrm{resp. }[\alpha]_{x}\in S(IC_{\widetilde{X}}(\pi^{\ast}\mathbb{V}),\pi^{\ast}\varphi+\pi^{\ast}\psi)_{\bf 0}\right)

if and only if

fi(πφj=1r2αEj(vi~)log|wj|)𝟎(resp. fi(πφ+πψj=1r2αEj(vi~)log|wj|)𝟎)f_{i}\in\mathscr{I}\big{(}\pi^{\ast}\varphi-\sum_{j=1}^{r}2\alpha_{E_{j}}(\widetilde{v_{i}})\log|w_{j}|\big{)}_{\bf 0}\quad\bigg{(}\textrm{resp. }f_{i}\in\mathscr{I}\big{(}\pi^{\ast}\varphi+\pi^{\ast}\psi-\sum_{j=1}^{r}2\alpha_{E_{j}}(\widetilde{v_{i}})\log|w_{j}|\big{)}_{\bf 0}\bigg{)}

for every i=1,,mi=1,\dots,m. Hence [α]xS(ICX~(π𝕍),πφ)𝟎[\alpha]_{x}\in S(IC_{\widetilde{X}}(\pi^{\ast}\mathbb{V}),\pi^{\ast}\varphi)_{\bf 0} is equivalent to that the integral

|fi|2j=1s|wj|2ajj=1r|wj|2αEj(vi~)volds02\displaystyle\int|f_{i}|^{2}\prod_{j=1}^{s}|w_{j}|^{-2a_{j}}\prod_{j=1}^{r}|w_{j}|^{2\alpha_{E_{j}}(\widetilde{v_{i}})}{\rm vol}_{ds^{2}_{0}}

is finite near 𝟎{\bf 0} for every i=1,,mi=1,\dots,m. Denote

vj(f):=min{l|fl0 in the Laurent expansion f=ifiwji}.v_{j}(f):=\min\{l|f_{l}\neq 0\textrm{ in the Laurent expansion }f=\sum_{i\in\mathbb{Z}}f_{i}w_{j}^{i}\}.

By Lemma 4.6, we observe that

f(πφj=1r2αEj(vi~)log|wj|)f\in\mathscr{I}\big{(}\pi^{\ast}\varphi-\sum_{j=1}^{r}2\alpha_{E_{j}}(\widetilde{v_{i}})\log|w_{j}|\big{)}

if and only if

(4.12) vj(f)aj+αEj(vi~)>1,j=1,,s.\displaystyle v_{j}(f)-a_{j}+\alpha_{E_{j}}(\widetilde{v_{i}})>-1,\quad\forall j=1,\dots,s.

Here we set αEj(vi~)=0,j=r+1,,s\alpha_{E_{j}}(\widetilde{v_{i}})=0,\forall j=r+1,\dots,s.

Similar arguments show that

f(π(φ+ϵψ)j=1r2αEj(vi~)log|wj|)f\in\mathscr{I}\big{(}\pi^{\ast}(\varphi+\epsilon\psi)-\sum_{j=1}^{r}2\alpha_{E_{j}}(\widetilde{v_{i}})\log|w_{j}|\big{)}

if and only if

(4.13) vj(f)ajϵbj+αEj(vi~)>1,j=1,,s.\displaystyle v_{j}(f)-a_{j}-\epsilon b_{j}+\alpha_{E_{j}}(\widetilde{v_{i}})>-1,\quad\forall j=1,\dots,s.

Conditions (4.12) and (4.13) are equivalent when ϵ>0\epsilon>0 is small enough. We obtain therefore that

S(ICX~(π𝕍),πφ)S(ICX~(π𝕍),π(φ+ϵψ)),0<ϵ1.S(IC_{\widetilde{X}}(\pi^{\ast}\mathbb{V}),\pi^{\ast}\varphi)\simeq S(IC_{\widetilde{X}}(\pi^{\ast}\mathbb{V}),\pi^{\ast}(\varphi+\epsilon\psi)),\quad 0<\epsilon\ll 1.

By Proposition 4.3,

S(ICX(𝕍),φ)π(S(ICX~(π𝕍),πφ))π(S(ICX~(π𝕍),π(φ+ϵψ)))S(ICX(𝕍),φ+ϵψ)S(IC_{X}(\mathbb{V}),\varphi)\simeq\pi_{\ast}(S(IC_{\widetilde{X}}(\pi^{\ast}\mathbb{V}),\pi^{\ast}\varphi))\simeq\pi_{\ast}(S(IC_{\widetilde{X}}(\pi^{\ast}\mathbb{V}),\pi^{\ast}(\varphi+\epsilon\psi)))\simeq S(IC_{X}(\mathbb{V}),\varphi+\epsilon\psi)

when ϵ>0\epsilon>0 is small enough. ∎

4.3. Extension and adjunction

In this section we consider the extension and adjunction properties for S(ICX(𝕍),φ)S(IC_{X}(\mathbb{V}),\varphi). These results will be used in proving the Demailly-Kawamata-Viehweg vanishing theorem for Saito’s SS-sheaf S(ICX(𝕍))S(IC_{X}(\mathbb{V})) (Corollary 6.2).

Theorem 4.14 (Ohsawa-Takegoshi extension theorem for S(ICX(𝕍),φ)S(IC_{X}(\mathbb{V}),\varphi)).

Let XX be a complex space and ΩXreg\Omega\subset X_{\rm reg} a Zariski open subset which is a Stein manifold. Let 𝕍\mathbb{V} be an \mathbb{R}-polarized variation of Hodge structure on some Zariski open subset XoXregX^{o}\subset X_{\rm reg}. Let HXH\subset X be a Cartier divisor such that 𝒪X(H)𝒪X\mathscr{O}_{X}(H)\simeq\mathscr{O}_{X} and HregΩXoH_{\rm reg}\cap\Omega\cap X^{o} is dense in HregH_{\rm reg}. Let (E,hφ)(E,h_{\varphi}) be a vector bundle on XX with a singular hermitian metric hφ=eφh0h_{\varphi}=e^{-\varphi}h_{0}. Assume that φ\varphi is smooth over some Zariski open subset of XX and 1Θhφ(E)0\sqrt{-1}\Theta_{h_{\varphi}}(E)\geq 0 as a current. Let hHh_{H} be a smooth hermitian metric on 𝒪X(H)\mathscr{O}_{X}(H) with zero curvature. Then there is a constant C>0C>0 such that, for every αΓ(H,S(ICH(𝕍|HXo),φ|H)E|H𝒪X(H)|H)\alpha\in\Gamma(H,S(IC_{H}(\mathbb{V}|_{H\cap X^{o}}),\varphi|_{H})\otimes E|_{H}\otimes\mathscr{O}_{X}(H)|_{H}) satisfying that αh𝕍hφhH<\|\alpha\|_{h_{\mathbb{V}}h_{\varphi}h_{H}}<\infty, there is βΓ(X,S(ICX(𝕍),φ)E)\beta\in\Gamma(X,S(IC_{X}(\mathbb{V}),\varphi)\otimes E) which satisfies that β|H=α\beta|_{H}=\alpha and βh𝕍hφCαh𝕍hφhH\|\beta\|_{h_{\mathbb{V}}h_{\varphi}}\leq C\|\alpha\|_{h_{\mathbb{V}}h_{\varphi}h_{H}}.

Proof.

Assume that φ\varphi is smooth over a Zariski open subset UXU\subset X. Let HH^{\prime} be a Cartier divisor of Ω\Omega so that Y:=Ω\HXoUY:=\Omega\backslash H^{\prime}\subset X^{o}\cap U and YY is a Stein manifold. Let

αΓ(H,S(ICH(𝕍|HXo),φ|H)E|H𝒪X(H)|H)\alpha\in\Gamma(H,S(IC_{H}(\mathbb{V}|_{H\cap X^{o}}),\varphi|_{H})\otimes E|_{H}\otimes\mathscr{O}_{X}(H)|_{H})

such that αh𝕍hφhH<\|\alpha\|_{h_{\mathbb{V}}h_{\varphi}h_{H}}<\infty. By the Ohsawa-Takegoshi extension theorem [OT1988] (see also [Guan-Zhou2015, Theorem 2.2]), there is βΓ(Y,KYS(𝕍|Y)E|Y)\beta\in\Gamma(Y,K_{Y}\otimes S(\mathbb{V}|_{Y})\otimes E|_{Y}) such that β|YH=α|YH\beta|_{Y\cap H}=\alpha|_{Y\cap H} and βY,h𝕍hφCα|YHh𝕍hφhH\|\beta\|_{Y,h_{\mathbb{V}}h_{\varphi}}\leq C\|\alpha|_{Y\cap H}\|_{h_{\mathbb{V}}h_{\varphi}h_{H}} for some constant C>0C>0. It follows from Lemma 4.12 that

βΓ(X,S(ICX(𝕍|Y),φ)E)=Γ(X,S(ICX(𝕍),φ)E).\beta\in\Gamma(X,S(IC_{X}(\mathbb{V}|_{Y}),\varphi)\otimes E)=\Gamma(X,S(IC_{X}(\mathbb{V}),\varphi)\otimes E).

Since α\alpha and β\beta are both holomorphic sections, we get β|H=α\beta|_{H}=\alpha. ∎

An immediate consequence of the extension theorem is the following

Corollary 4.15 (Restriction Formula).

Let XX be a complex space and XoXregX^{o}\subset X_{\rm reg} a Zariski open subset. Let 𝕍\mathbb{V} be an \mathbb{R}-polarized variation of Hodge structure on some Zariski open subset XoXregX^{o}\subset X_{\rm reg}. Let φ\varphi be a quasi-psh function on XX which is smooth over some Zariski open subset of XX. Let HH be a reduced Cartier divisor of XX such that HregXoH_{\rm reg}\cap X^{o} is dense in HregH_{\rm reg}. Then

(4.14) S(ICH(𝕍|HXo),φ|H)S(ICX(𝕍),φ)|H𝒪X(H)|H\displaystyle S(IC_{H}(\mathbb{V}|_{H\cap X^{o}}),\varphi|_{H})\subset S(IC_{X}(\mathbb{V}),\varphi)|_{H}\otimes\mathscr{O}_{X}(H)|_{H}

as subsheaves of S(ICX(𝕍),0)|H𝒪X(H)|HS(IC_{X}(\mathbb{V}),0)|_{H}\otimes\mathscr{O}_{X}(H)|_{H}.

Proof.

By Theorem 4.14, there is an inclusion

S(ICH(𝕍|HXo),φ|H)SX(S(𝕍)𝒪X(H),h𝕍eφh0)|HS(IC_{H}(\mathbb{V}|_{H\cap X^{o}}),\varphi|_{H})\subset S_{X}(S(\mathbb{V})\otimes\mathscr{O}_{X}(H),h_{\mathbb{V}}\otimes e^{-\varphi}h_{0})|_{H}

where h0h_{0} is an arbitrary smooth hermitian metric on 𝒪X(H)\mathscr{O}_{X}(H). By Lemma 4.12 there is an isomorphism

SX(S(𝕍)𝒪X(H),h𝕍eφh0)S(ICX(𝕍),φ)𝒪X(H).S_{X}(S(\mathbb{V})\otimes\mathscr{O}_{X}(H),h_{\mathbb{V}}\otimes e^{-\varphi}h_{0})\simeq S(IC_{X}(\mathbb{V}),\varphi)\otimes\mathscr{O}_{X}(H).

The corollary is proved. ∎

Generally, the inclusion (4.14) is strict even if φ=0\varphi=0. From the perspective of Hodge modules, ICX(𝕍)|HIC_{X}(\mathbb{V})|_{H} could be a mixed Hodge module while ICH(𝕍|HXo)IC_{H}(\mathbb{V}|_{H\cap X^{o}}) is pure. When φ\varphi has generalized analytic singularities, (4.14) is an equality if HH is in a general position.

Proposition 4.16.

Let XX be a projective algebraic variety and let 𝕍\mathbb{V} be an \mathbb{R}-polarized variation of Hodge structure on some Zariski open subset XoXregX^{o}\subset X_{\rm reg}. Let φ\varphi be a quasi-psh function on XX which has generalized analytic singularities. Let Λ\Lambda be a free linear system on XX. Then there is a canonical isomorphism

S(ICX(𝕍),φ)|H𝒪X(H)|H=S(ICH(𝕍|HXo),φ|H)S(IC_{X}(\mathbb{V}),\varphi)|_{H}\otimes\mathscr{O}_{X}(H)|_{H}=S(IC_{H}(\mathbb{V}|_{H\cap X^{o}}),\varphi|_{H})

for a general HΛH\in\Lambda.

Proof.

Let π:X~X\pi:\widetilde{X}\to X be a desingularization so that E:=π1(X\Xo)E:=\pi^{-1}(X\backslash X^{o}) is a simple normal crossing divisor and π\pi is biholomorphic over XoX^{o}. Let HH be in a general position so that H~:=πH\widetilde{H}:=\pi^{\ast}H is smooth and intersects transversally with every stratum of EE. By Proposition 4.3, it suffices to show that

(4.15) S(ICX~(π𝕍),πφ)|H~𝒪X~(H~)|H~=S(ICH~(π𝕍|π(HXo)),πφ|H~)\displaystyle S(IC_{\widetilde{X}}(\pi^{\ast}\mathbb{V}),\pi^{\ast}\varphi)|_{\widetilde{H}}\otimes\mathscr{O}_{\widetilde{X}}(\widetilde{H})|_{\widetilde{H}}=S(IC_{\widetilde{H}}(\pi^{\ast}\mathbb{V}|_{\pi^{\ast}(H\cap X^{o})}),\pi^{\ast}\varphi|_{\widetilde{H}})

for a general HΛH\in\Lambda.

Since the problem is local, we assume that X~=Δn\widetilde{X}=\Delta^{n} is a polydisc with the standard holomorphic coordinates z1,,znz_{1},\dots,z_{n} such that E={z1zr=0}E=\{z_{1}\cdots z_{r}=0\}. Denote Ei={zi=0}E_{i}=\{z_{i}=0\}, i=1,,r\forall i=1,\dots,r. Let (v1~,,vm~)(\widetilde{v_{1}},\dots,\widetilde{v_{m}}) be an L2L^{2}-adapted frame of R(ICX(𝕍))R(IC_{X}(\mathbb{V})) as in Proposition 2.5. Denote

ψi:=j=1rαEj(vi~)log|zj|2.\psi_{i}:=-\sum_{j=1}^{r}\alpha_{E_{j}}(\widetilde{v_{i}})\log|z_{j}|^{2}.

By Proposition 4.8, we see that

S(ICX~(π𝕍),πφ)ωX~i=1m(πφ+ψi)vi~,\displaystyle S(IC_{\widetilde{X}}(\pi^{\ast}\mathbb{V}),\pi^{\ast}\varphi)\simeq\omega_{\widetilde{X}}\otimes\bigoplus_{i=1}^{m}\mathscr{I}(\pi^{\ast}\varphi+\psi_{i})\widetilde{v_{i}},
S(ICH~(π𝕍|π(HXo))),πφ|H~)ωH~i=1m(πφ|H~+ψi|H~)vi~|H~\displaystyle S(IC_{\widetilde{H}}(\pi^{\ast}\mathbb{V}|_{\pi^{\ast}(H\cap X^{o})})),\pi^{\ast}\varphi|_{\widetilde{H}})\simeq\omega_{{\widetilde{H}}}\otimes\bigoplus_{i=1}^{m}\mathscr{I}(\pi^{\ast}\varphi|_{\widetilde{H}}+\psi_{i}|_{\widetilde{H}})\widetilde{v_{i}}|_{\widetilde{H}}

and

(πφ+ψi)|H~(πφ|H~+ψi|H~)\mathscr{I}(\pi^{\ast}\varphi+\psi_{i})|_{\widetilde{H}}\simeq\mathscr{I}(\pi^{\ast}\varphi|_{\widetilde{H}}+\psi_{i}|_{\widetilde{H}})

since HH is in a general position. Consequently, (4.15) is obtained. ∎

5. L2L^{2}-Dolbeault resolution of multiplier SS-sheaf

The purpose of this section is to prove Theorem 1.1.

Theorem 5.1.

Let XX be a complex space of dimension nn and ds2ds^{2} a hermitian metric on a Zariski open subset XoXregX^{o}\subset X_{\rm reg} with ω\omega its fundamental form. Let 𝕍=(𝒱,,,h𝕍)\mathbb{V}=(\mathcal{V},\nabla,\mathcal{F}^{\bullet},h_{\mathbb{V}}) be an \mathbb{R}-polarized variation of Hodge structure on XoX^{o}. Let (E,hφ)(E,h_{\varphi}) be a holomorphic vector bundle on XX with a (possibly) singular hermitian metric hφ:=eφh0h_{\varphi}:=e^{-\varphi}h_{0}. Assume that, locally at every point xXx\in X, there is a neighborhood UU of xx, a strictly psh function λC2(U)\lambda\in C^{2}(U) and a bounded psh function ΦC2(UXo)\Phi\in C^{2}(U\cap X^{o}) such that 1¯λ|UXoω|UXo1¯Φ\sqrt{-1}\partial\bar{\partial}\lambda|_{U\cap X^{o}}\lesssim\omega|_{U\cap X^{o}}\lesssim\sqrt{-1}\partial\bar{\partial}\Phi. Then the canonical map

(5.1) S(ICX(𝕍),φ)E𝒟X,ds2n,(S(𝕍)E,h𝕍hφ)\displaystyle S(IC_{X}(\mathbb{V}),\varphi)\otimes E\to\mathscr{D}^{n,\bullet}_{X,ds^{2}}(S(\mathbb{V})\otimes E,h_{\mathbb{V}}\otimes h_{\varphi})

is a quasi-isomorphism. If XX is moreover compact, then there is an isomorphism

Hq(S(ICX(𝕍),φ)E)H(2),maxn,q(Xo,S(𝕍)E|Xo;ds2,h𝕍hφ),q.\displaystyle H^{q}(S(IC_{X}(\mathbb{V}),\varphi)\otimes E)\simeq H^{n,q}_{(2),\rm max}(X^{o},S(\mathbb{V})\otimes E|_{X^{o}};ds^{2},h_{\mathbb{V}}\otimes h_{\varphi}),\quad\forall q.
Proof.

By Lemma 4.12 we have

S(ICX(𝕍),φ)EKer(𝒟X,ds2n,0(S(𝕍)E,h𝕍hφ)¯𝒟X,ds2n,1(S(𝕍)E,h𝕍hφ)).S(IC_{X}(\mathbb{V}),\varphi)\otimes E\simeq{\rm Ker}\left(\mathscr{D}^{n,0}_{X,ds^{2}}(S(\mathbb{V})\otimes E,h_{\mathbb{V}}\otimes h_{\varphi})\stackrel{{\scriptstyle\bar{\partial}}}{{\to}}\mathscr{D}^{n,1}_{X,ds^{2}}(S(\mathbb{V})\otimes E,h_{\mathbb{V}}\otimes h_{\varphi})\right).

It remains to show that (5.1) is exact at 𝒟X,ds2n,q(S(𝕍)E,h𝕍hφ)\mathscr{D}^{n,q}_{X,ds^{2}}(S(\mathbb{V})\otimes E,h_{\mathbb{V}}\otimes h_{\varphi}) for q>0q>0. Since the problem is local, we consider a point xXx\in X and an open neighborhood UU of xx such that E|U𝒪UrkEE|_{U}\simeq\mathscr{O}_{U}^{\oplus{\rm rk}E} and h0h_{0} is quasi-isometric to the trivial metric h1h_{1} on 𝒪UrkE\mathscr{O}_{U}^{\oplus{\rm rk}E}. Since

1Θeφh1(E|UXo)=1¯φIdEc1¯λIdEC1¯ΦIdE\sqrt{-1}\Theta_{e^{-\varphi}h_{1}}(E|_{U\cap X^{o}})=\sqrt{-1}\partial\bar{\partial}\varphi\otimes{\rm Id}_{E}\geq c\sqrt{-1}\partial\bar{\partial}\lambda\otimes{\rm Id}_{E}\geq C\sqrt{-1}\partial\bar{\partial}\Phi\otimes{\rm Id}_{E}

for some negative constants c,Cc,C. By assumptions, there is a constant C>0C^{\prime}>0 such that

(C+C)1¯Φω|UXo.(C^{\prime}+C)\sqrt{-1}\partial\bar{\partial}\Phi\geq\omega|_{U\cap X^{o}}.

Let h=eφCΦh1h^{\prime}=e^{-\varphi-C^{\prime}\Phi}h_{1}. The boundedness of Φ\Phi implies that hhφh^{\prime}\sim h_{\varphi}. It follows from Theorem 2.3 that

1Θh𝕍h(S(𝕍)E|UXo)C1¯ΦIdS(𝕍)E+1Θeφh1(E|UXo)IdS(𝕍)ωIdS(𝕍)E\sqrt{-1}\Theta_{h_{\mathbb{V}}\otimes h^{\prime}}(S(\mathbb{V})\otimes E|_{U\cap X^{o}})\geq C^{\prime}\sqrt{-1}\partial\bar{\partial}\Phi\otimes{\rm Id}_{S(\mathbb{V})\otimes E}+\sqrt{-1}\Theta_{e^{-\varphi}h_{1}}(E|_{U\cap X^{o}})\otimes{\rm Id}_{S(\mathbb{V})}\geq\omega\otimes{\rm Id}_{S(\mathbb{V})\otimes E}

holds on UXoU\cap X^{o}. By Lemma 3.4 we may assume that UXoU\cap X^{o} admits a complete Kähler metric. Consequently, we have

H(2)n,q(UXo,S(𝕍)E|UXo;ds2,h𝕍hφ)=H(2)n,q(UXo,S(𝕍)E|UXo;ds2,h𝕍h)=0,q>0H^{n,q}_{(2)}(U\cap X^{o},S(\mathbb{V})\otimes E|_{U\cap X^{o}};ds^{2},h_{\mathbb{V}}h_{\varphi})=H^{n,q}_{(2)}(U\cap X^{o},S(\mathbb{V})\otimes E|_{U\cap X^{o}};ds^{2},h_{\mathbb{V}}h^{\prime})=0,\quad\forall q>0

by Theorem 3.3. This proves the exactness of (5.1) at 𝒟X,ds2n,q(S(𝕍)E,h𝕍hφ)\mathscr{D}^{n,q}_{X,ds^{2}}(S(\mathbb{V})\otimes E,h_{\mathbb{V}}\otimes h_{\varphi}) for all q>0q>0. Since 𝒟X,ds02n,(E,hφ)\mathscr{D}^{n,\bullet}_{X,ds^{2}_{0}}(E,h_{\varphi}) is a complex of fine sheaves (Lemma 3.2), we obtain the second claim. ∎

A typical example of the metric ds2ds^{2} that satisfies the conditions in Theorem 5.1 is the hermitian metric on XX, which always exists by partition of unity. For applications we require ds2ds^{2} to be a complete Kähler metric. Such kind of metric exists if XX is the truncation of a weakly pseudoconvex Kähler space. Recall that a complex space is a weakly pseudoconvex (Kähler) space if it is a (Kähler) complex space that admits a smooth psh exhausted function.

Lemma 5.2.

Let XX be a weakly pseudoconvex Kähler space with ψ\psi a smooth psh exhausted function on XX. Denote Xc:={xX|ψ(x)<c}X_{c}:=\{x\in X|\psi(x)<c\}. Let XoXregX^{o}\subset X_{\rm reg} be a Zariski open subset. Then, for every cc\in\mathbb{R}, there exists a complete Kähler metric ds2ds^{2} on XcXoX_{c}\cap X^{o} satisfying that for every point xXcx\in X_{c}, there is a neighborhood UU of xx, a bounded function ΦC(U)\Phi\in C^{\infty}(U) and a smooth strictly psh function λ\lambda on UU such that 1¯λds21¯Φ\sqrt{-1}\partial\bar{\partial}\lambda\lesssim ds^{2}\sim\sqrt{-1}\partial\bar{\partial}\Phi.

Proof.

The construction of the metric is the motivated by Ohsawa [Ohsawa2018, Lemma 2.6]. Let UU be a neighborhood of a point xX\Xox\in X\backslash X^{o}. Assume that U\XoUU\backslash X^{o}\subset U is defined by f1,,fr𝒪U(U)f_{1},\dots,f_{r}\in\mathscr{O}_{U}(U). Let

φU:=1log(logi=1r|fi|2)+ϕU,\varphi_{U}:=\frac{1}{\log(-\log\sum_{i=1}^{r}|f_{i}|^{2})}+\phi_{U},

where ϕU\phi_{U} is a strictly CC^{\infty} psh function on UU so that φU\varphi_{U} is strictly psh. Then the quasi-isometric class of 1¯φU\sqrt{-1}\partial\bar{\partial}\varphi_{U} is independent of the choice of {f1,,fr}\{f_{1},\dots,f_{r}\} and ϕU\phi_{U}. By partition of unity, the potential functions φU\varphi_{U} can be glued to a global function φ\varphi (not necessarily psh) on XX so that

1¯φ|U1¯φU\displaystyle\sqrt{-1}\partial\bar{\partial}\varphi|_{U}\sim\sqrt{-1}\partial\bar{\partial}\varphi_{U}

near every point xX\Xox\in X\backslash X^{o} and φ0\varphi\equiv 0 away from a neighborhood of X\XoX\backslash X^{o}.

Denote u=logi=1r|fi|2u=-\log\sum_{i=1}^{r}|f_{i}|^{2}. Then

1¯φ|U\displaystyle\sqrt{-1}\partial\bar{\partial}\varphi|_{U} 12+loguu2log3uu¯u+1¯uulog2u+1¯ϕU\displaystyle\sim\sqrt{-1}\frac{2+\log u}{u^{2}\log^{3}u}\partial u\wedge\bar{\partial}u+\sqrt{-1}\frac{-\partial\bar{\partial}u}{u\log^{2}u}+\sqrt{-1}\partial\bar{\partial}\phi_{U}
1u¯uu2log2u+1¯uulog2u+1¯ϕU.\displaystyle\sim\sqrt{-1}\frac{\partial u\wedge\bar{\partial}u}{u^{2}\log^{2}u}+\sqrt{-1}\frac{-\partial\bar{\partial}u}{u\log^{2}u}+\sqrt{-1}\partial\bar{\partial}\phi_{U}.

Hence loglogu\log\log u is a smooth psh exhausted function near X\XoX\backslash X^{o} such that

|dloglogu|1¯φ2.|d\log\log u|_{\sqrt{-1}\partial\bar{\partial}\varphi}\lesssim 2.

By the Hopf-Rinow theorem, 1¯φ\sqrt{-1}\partial\bar{\partial}\varphi is locally complete near X\XoX\backslash X^{o}.

Let cc\in\mathbb{R} and let ω0\omega_{0} be a Kähler hermitian metric on XX. By adding a constant to ψ\psi we assume that ψ0\psi\geq 0. Then ψc:=ψ+1cψ\psi_{c}:=\psi+\frac{1}{c-\psi} is a smooth psh exhausted function on Xc={xX|ψ(x)<c}X_{c}=\{x\in X|\psi(x)<c\}. Hence, ω0+1¯ψc2\omega_{0}+\sqrt{-1}\partial\bar{\partial}\psi^{2}_{c} is a complete Kähler metric on XcX_{c} ([Demailly1982, Theorem 1.3]). Since Xc¯\overline{X_{c}} is compact,

1¯φ+K(ω0+1¯ψc2),K0\sqrt{-1}\partial\bar{\partial}\varphi+K(\omega_{0}+\sqrt{-1}\partial\bar{\partial}\psi^{2}_{c}),\quad K\gg 0

is positive definite and it provides the desired complete Kähler metric on XcXoX_{c}\cap X^{o}. ∎

6. Vanishing theorems for SS-sheaf

Vanishing theorems and injectivity theorems for the SS-sheaf are deduced from Theorem 5.1 in this section.

6.1. Nadel vanishing theorem

Theorem 6.1.

Let XX be a weakly pseudoconvex Kähler space and ω\omega a Kähler hermitian metric on XX. Let 𝕍\mathbb{V} be an \mathbb{R}-polarized variation of Hodge structure on some Zariski open subset XoXregX^{o}\subset X_{\rm reg} and (L,hφ)(L,h_{\varphi}) a holomorphic line bundle with a possibly singular hermitian metric hφ:=eφhh_{\varphi}:=e^{-\varphi}h. If 1Θhφ(L)ϵω\sqrt{-1}\Theta_{h_{\varphi}}(L)\geq\epsilon\omega as currents for some ϵ>0\epsilon>0, then

Hq(X,S(ICX(𝕍),φ)L)=0,q>0.H^{q}(X,S(IC_{X}(\mathbb{V}),\varphi)\otimes L)=0,\quad\forall q>0.
Proof.

By Theorem 5.1 there is a quasi-isomorphism

S(ICX(𝕍),φ)Lq.i.s.𝒟X,ωn,(S(𝕍)L,h𝕍hφ).S(IC_{X}(\mathbb{V}),\varphi)\otimes L\simeq_{\rm q.i.s.}\mathscr{D}^{n,\bullet}_{X,\omega}(S(\mathbb{V})\otimes L,h_{\mathbb{V}}\otimes h_{\varphi}).

Lemma 3.2 implies that 𝒟X,ωn,(S(𝕍)L,h𝕍hφ)\mathscr{D}^{n,\bullet}_{X,\omega}(S(\mathbb{V})\otimes L,h_{\mathbb{V}}\otimes h_{\varphi}) is a complex of fine sheaves. Thus

Hq(X,S(ICX(𝕍),φ)L)Hq(Γ(X,𝒟X,ωn,(S(𝕍)L,h𝕍hφ))).\displaystyle H^{q}(X,S(IC_{X}(\mathbb{V}),\varphi)\otimes L)\simeq H^{q}\left(\Gamma(X,\mathscr{D}^{n,\bullet}_{X,\omega}(S(\mathbb{V})\otimes L,h_{\mathbb{V}}\otimes h_{\varphi}))\right).

Now let q>0q>0 and let αΓ(X,𝒟X,ωn,q(S(𝕍)L,h𝕍hφ))\alpha\in\Gamma(X,\mathscr{D}^{n,q}_{X,\omega}(S(\mathbb{V})\otimes L,h_{\mathbb{V}}\otimes h_{\varphi})) be a locally L2L^{2} form such that ¯α=0\bar{\partial}\alpha=0. We would like to show that there exists βΓ(X,𝒟X,ωn,q1(S(𝕍)L,h𝕍hφ))\beta\in\Gamma(X,\mathscr{D}^{n,q-1}_{X,\omega}(S(\mathbb{V})\otimes L,h_{\mathbb{V}}\otimes h_{\varphi})) satisfying that ¯β=α\bar{\partial}\beta=\alpha.

Let ψ\psi be a smooth psh exhausted function on XX and let χ\chi be a convex increasing function which is of fast growth at infinity so that

Xo|α|ω,h𝕍hφ2eχψvolω<.\displaystyle\int_{X^{o}}|\alpha|_{\omega,h_{\mathbb{V}}\otimes h_{\varphi}}^{2}e^{-\chi\circ\psi}{\rm vol_{\omega}}<\infty.

Let h:=eχψhφh^{\prime}:=e^{-\chi\circ\psi}h_{\varphi}. Consequently, we have

1Θh𝕍h(S(𝕍)L)\displaystyle\sqrt{-1}\Theta_{h_{\mathbb{V}}\otimes h^{\prime}}(S(\mathbb{V})\otimes L) =1Θh𝕍(S(𝕍))+1¯(χψ)IdS(𝕍)+1Θhφ(L)IdS(𝕍)\displaystyle=\sqrt{-1}\Theta_{h_{\mathbb{V}}}(S(\mathbb{V}))+\sqrt{-1}\partial\bar{\partial}(\chi\circ\psi)\otimes{\rm Id}_{S(\mathbb{V})}+\sqrt{-1}\Theta_{h_{\varphi}}(L)\otimes{\rm Id}_{S(\mathbb{V})}
ϵωIdS(𝕍)\displaystyle\geq\epsilon\omega\otimes{\rm Id}_{S(\mathbb{V})}

by Theorem 2.3. Moreover, Lemma 5.2 implies that XcXo:={xXo|ψ(x)<c}X_{c}\cap X^{o}:=\{x\in X^{o}|\psi(x)<c\} admits a complete Kähler metric for every cc\in\mathbb{R}. It follows from Theorem 3.3 that for every cc\in\mathbb{R} there is

βcL(2)n,q1(XcXo,S(𝕍)L;ω,h𝕍h)Γ(Xc,𝒟X,ωn,q1(S(𝕍)L,h𝕍hφ))\beta_{c}\in L^{n,q-1}_{(2)}(X_{c}\cap X^{o},S(\mathbb{V})\otimes L;\omega,h_{\mathbb{V}}\otimes h^{\prime})\subset\Gamma(X_{c},\mathscr{D}^{n,q-1}_{X,\omega}(S(\mathbb{V})\otimes L,h_{\mathbb{V}}\otimes h_{\varphi}))

such that ¯βc=α|Xc\bar{\partial}\beta_{c}=\alpha|_{X_{c}} and βcϵω,h𝕍h1qαϵω,h𝕍h\|\beta_{c}\|_{\epsilon\omega,h_{\mathbb{V}}\otimes h^{\prime}}\leq\frac{1}{q}\|\alpha\|_{\epsilon\omega,h_{\mathbb{V}}\otimes h^{\prime}}. By taking a weak limit of βc\beta_{c} we obtain a solution of the equation α=¯β\alpha=\bar{\partial}\beta such that βΓ(X,𝒟X,ωn,q1(S(𝕍)L,h𝕍hφ))\beta\in\Gamma(X,\mathscr{D}^{n,q-1}_{X,\omega}(S(\mathbb{V})\otimes L,h_{\mathbb{V}}\otimes h_{\varphi})). ∎

Corollary 6.2.

Let XX be a projective algebraic variety of dimension nn and let 𝕍\mathbb{V} be an \mathbb{R}-polarized variation of Hodge structure on some Zariski open subset XoXregX^{o}\subset X_{\rm reg}. Let LL be a line bundle such that some positive multiple mL=F+DmL=F+D where FF is a nef line bundle and DD is an effective divisor. Then

Hq(X,S(ICX(𝕍),φDm)L)=0,q>nnd(F).H^{q}(X,S(IC_{X}(\mathbb{V}),\frac{\varphi_{D}}{m})\otimes L)=0,\quad\forall q>n-{\rm nd}(F).

Here φD\varphi_{D} is the psh function associated to DD and

nd(F)=max{k=1,,n|c1(F)k0H2k(X,)}.{\rm nd}(F)=\max\{k=1,\dots,n|c_{1}(F)^{k}\neq 0\in H^{2k}(X,\mathbb{R})\}.
Remark 6.3.

Let sDΓ(X,𝒪X(D))s_{D}\in\Gamma(X,\mathscr{O}_{X}(D)) and hDh^{\prime}_{D} a hermitian metric on 𝒪X(D)\mathscr{O}_{X}(D). Let φD:=log|sD|hD2\varphi_{D}:=\log|s_{D}|^{2}_{h^{\prime}_{D}}. Then the singular metric hD:=eφDhDh_{D}:=e^{-\varphi_{D}}h^{\prime}_{D} is independent of the choice of hDh^{\prime}_{D} and 1ΘhD(𝒪X(D))0\sqrt{-1}\Theta_{h_{D}}(\mathscr{O}_{X}(D))\geq 0 as a current. Since φD\varphi_{D} is differed by a smooth function for different choices of sDs_{D} and hDh^{\prime}_{D}, Lemma 4.5 shows that S(ICX(𝕍),φDm)S(IC_{X}(\mathbb{V}),\frac{\varphi_{D}}{m}) is independent of the choice of sDs_{D} and hDh^{\prime}_{D}.

Proof.

Case I: nd(F)=n{\rm nd}(F)=n. In this case FF is big and nef. Then bF=A+EbF=A+E for some constant b>0b>0, an ample line bundle AA and an effective Cartier divisor EE. Let hAh_{A} be a metric with positive curvature on AA and hEh^{\prime}_{E} a metric on 𝒪X(E)\mathscr{O}_{X}(E). Let hE:=eφEhEh_{E}:=e^{-\varphi_{E}}h^{\prime}_{E} (Remark 6.3). Then hF,0=(hAhE)1bh_{F,0}=(h_{A}\otimes h_{E})^{\frac{1}{b}} is a singular metric such that

1ΘhF,0(F)1bω.\displaystyle\sqrt{-1}\Theta_{h_{F,0}}(F)\geq\frac{1}{b}\omega.

Here ω=1ΘhA(A)\omega=\sqrt{-1}\Theta_{h_{A}}(A) is a hermitian metric on XX. Since FF is nef, there is a metric hF,ϵh_{F,\epsilon} such that 1ΘhF,ϵ(F)ϵω\sqrt{-1}\Theta_{h_{F,\epsilon}}(F)\geq-\epsilon\omega for every ϵ>0\epsilon>0. Let hD=eφDhDh_{D}=e^{-\varphi_{D}}h^{\prime}_{D} be the singular metric on 𝒪X(D)\mathscr{O}_{X}(D) as in Remark 6.3. Define

hL:=(hF,ϵ1δhF,0δhD)1m=eδφEbmφDm(hF,ϵ1δhAδ/bhEδ/bhD)1m,0<bϵδ1.\displaystyle h_{L}:=(h_{F,\epsilon}^{1-\delta}\otimes h_{F,0}^{\delta}\otimes h_{D})^{\frac{1}{m}}=e^{-\frac{\delta\varphi_{E}}{bm}-\frac{\varphi_{D}}{m}}(h_{F,\epsilon}^{1-\delta}\otimes h_{A}^{\delta/b}\otimes h^{\prime\delta/b}_{E}\otimes h^{\prime}_{D})^{\frac{1}{m}},\quad 0<b\epsilon\ll\delta\ll 1.

Then

1ΘhL(L)\displaystyle\sqrt{-1}\Theta_{h_{L}}(L) =1m((1δ)1ΘhF,ϵ(F)+δ1ΘhF,0(F)+1ΘhD(D))\displaystyle=\frac{1}{m}\left((1-\delta)\sqrt{-1}\Theta_{h_{F,\epsilon}}(F)+\delta\sqrt{-1}\Theta_{h_{F,0}}(F)+\sqrt{-1}\Theta_{h_{D}}(D)\right)
1m((1δ)ϵω+δbω)δϵmω.\displaystyle\geq\frac{1}{m}\left(-(1-\delta)\epsilon\omega+\frac{\delta}{b}\omega\right)\geq\frac{\delta\epsilon}{m}\omega.

Consequently, Theorem 6.1 yields

Hq(X,S(ICX(𝕍),φL)L)=0,q>0,H^{q}(X,S(IC_{X}(\mathbb{V}),\varphi_{L})\otimes L)=0,\quad\forall q>0,

where φL=δφEbm+φDm\varphi_{L}=\frac{\delta\varphi_{E}}{bm}+\frac{\varphi_{D}}{m}. Moreover, Proposition 4.13 implies that S(ICX(𝕍),φL)=S(ICX(𝕍),1mφD)S(IC_{X}(\mathbb{V}),\varphi_{L})=S(IC_{X}(\mathbb{V}),\frac{1}{m}\varphi_{D}) when δ>0\delta>0 is small enough. This proves the theorem under the condition that nd(F)=n{\rm nd}(F)=n.

Case II: nd(F)<n{\rm nd}(F)<n. Let π:X~X\pi:\widetilde{X}\to X be a desingularization so that π\pi is biholomorphic over Xo\DX^{o}\backslash D and the exceptional loci E:=π1((X\Xo)D)E:=\pi^{-1}((X\backslash X^{o})\cup D) is a simple normal crossing divisor. Let HH be a reduced ample hypersurface in a general position so that π1H\pi^{-1}H is smooth and has normal crossings with EE. By taking HH sufficiently ample we assume that F+mHF+mH is ample.

By Proposition 4.16 we have

(6.1) S(ICX(𝕍),1mφD)|H𝒪X(H)|HS(ICH(𝕍|XoH),1mφDH)\displaystyle S(IC_{X}(\mathbb{V}),\frac{1}{m}\varphi_{D})|_{H}\otimes\mathscr{O}_{X}(H)|_{H}\simeq S(IC_{H}(\mathbb{V}|_{X^{o}\cap H}),\frac{1}{m}\varphi_{D\cap H})

when HH is in a general position. We further assume that HH contains no associated points of S(ICX(𝕍),1mφD)S(IC_{X}(\mathbb{V}),\frac{1}{m}\varphi_{D}). This implies that the sequence

0S(ICX(𝕍),1mφD)LS(ICX(𝕍),1mφD)L𝒪X(H)S(ICX(𝕍),1mφD)L𝒪X(H)𝒪H00\to S(IC_{X}(\mathbb{V}),\frac{1}{m}\varphi_{D})\otimes L\to S(IC_{X}(\mathbb{V}),\frac{1}{m}\varphi_{D})\otimes L\otimes\mathscr{O}_{X}(H)\to S(IC_{X}(\mathbb{V}),\frac{1}{m}\varphi_{D})\otimes L\otimes\mathscr{O}_{X}(H)\otimes\mathscr{O}_{H}\to 0

is exact. There is therefore a long exact sequence

(6.2) \displaystyle\cdots Hq(X,S(ICX(𝕍),1mφD)L)Hq(X,S(ICX(𝕍),1mφD)L𝒪X(H))\displaystyle\to H^{q}\left(X,S(IC_{X}(\mathbb{V}),\frac{1}{m}\varphi_{D})\otimes L\right)\to H^{q}\left(X,S(IC_{X}(\mathbb{V}),\frac{1}{m}\varphi_{D})\otimes L\otimes\mathscr{O}_{X}(H)\right)
Hq(H,S(ICX(𝕍),1mφD)|HL|H𝒪X(H)|H).\displaystyle\to H^{q}\left(H,S(IC_{X}(\mathbb{V}),\frac{1}{m}\varphi_{D})|_{H}\otimes L|_{H}\otimes\mathscr{O}_{X}(H)|_{H}\right)\to\cdots.

Since F+mHF+mH is ample, nd(F+mH)=n{\rm nd}(F+mH)=n. Then we have

Hq(X,S(ICX(𝕍),1mφD)L𝒪X(H))=0,q>0\displaystyle H^{q}\left(X,S(IC_{X}(\mathbb{V}),\frac{1}{m}\varphi_{D})\otimes L\otimes\mathscr{O}_{X}(H)\right)=0,\quad\forall q>0

by Case I. It follows from (6.1) and (6.2)that there are isomorphisms

(6.3) Hq(X,S(ICX(𝕍),1mφD)L)Hq1(H,S(ICH(𝕍|XoH),1mφDH)L|H),q>1.\displaystyle H^{q}\left(X,S(IC_{X}(\mathbb{V}),\frac{1}{m}\varphi_{D})\otimes L\right)\simeq H^{q-1}\left(H,S(IC_{H}(\mathbb{V}|_{X^{o}\cap H}),\frac{1}{m}\varphi_{D\cap H})\otimes L|_{H}\right),\quad\forall q>1.

Subsequently, since nd(L|H)=nd(L){\rm nd}(L|_{H})={\rm nd}(L), we have

Hq(H,S(ICH(𝕍|XoH),1mφDH)L|H)=0,q>n1nd(L|H)\displaystyle H^{q}\left(H,S(IC_{H}(\mathbb{V}|_{X^{o}\cap H}),\frac{1}{m}\varphi_{D\cap H})\otimes L|_{H}\right)=0,\quad\forall q>n-1-{\rm nd}(L|_{H})

by induction on dimX\dim X. Consequently, (6.3) implies that

Hq(X,S(ICX(𝕍),1mφD)L)=0,q>nnd(L).\displaystyle H^{q}\left(X,S(IC_{X}(\mathbb{V}),\frac{1}{m}\varphi_{D})\otimes L\right)=0,\quad\forall q>n-{\rm nd}(L).

6.2. Relative vanishing theorem

Let f:XYf:X\to Y be a proper holomorphic map between complex spaces. A (1,1)(1,1)-current α\alpha on XX is ff-positive if, for every point yYy\in Y there is a neighborhood UU of yy, a hermitian metric ωU\omega_{U} on UU and a hermitian metric ω\omega^{\prime} on f1(U)f^{-1}(U) such that α|f1U+fωUω\alpha|_{f^{-1}U}+f^{\ast}\omega_{U}\geq\omega^{\prime}.

Corollary 6.4.

Let f:XYf:X\to Y be a surjective proper Kähler holomorphic map between complex spaces. Let 𝕍\mathbb{V} be an \mathbb{R}-polarized variation of Hodge structure on some Zariski open subset XoXregX^{o}\subset X_{\rm reg}. Let (L,hφ)(L,h_{\varphi}) be a holomorphic line bundle on XX with a possibly singular hermitian metric hφ:=eφhh_{\varphi}:=e^{-\varphi}h. Assume that 1Θhφ(L)\sqrt{-1}\Theta_{h_{\varphi}}(L) is ff-positive. Then

Rif(S(ICX(𝕍),φ)L)=0,i>0.\displaystyle R^{i}f_{\ast}(S(IC_{X}(\mathbb{V}),\varphi)\otimes L)=0,\quad\forall i>0.
Proof.

Since the problem is local, we may assume that YY admits a non-negative smooth strictly psh exhausted function ψ\psi so that ψ1{0}={y}Y\psi^{-1}\{0\}=\{y\}\subset Y is a point. To achieve this, one may embed YY into N\mathbb{C}^{N} as a closed Stein analytic subspace and take ψ=|z|2\psi=|z|^{2}. Moreover, since ff is proper, fψf^{\ast}\psi is a smooth psh exhausted function on XX.

By Lemma 5.2, for every cc\in\mathbb{R} there is a complete Kähler metric ωc\omega_{c} on Xof1(Yc)X^{o}\cap f^{-1}(Y_{c}) where Yc={yY|ψ(y)<c}Y_{c}=\{y\in Y|\psi(y)<c\}. Since 1Θhφ(L)\sqrt{-1}\Theta_{h_{\varphi}}(L) is ff-positive, we assume that

1ΘeCψhφ(L)=1Θhφ(L)+Cf(1¯ψ)ω\displaystyle\sqrt{-1}\Theta_{e^{-C\psi}h_{\varphi}}(L)=\sqrt{-1}\Theta_{h_{\varphi}}(L)+Cf^{\ast}\left(\sqrt{-1}\partial\bar{\partial}\psi\right)\geq\omega^{\prime}

for some hermitian metric ω\omega^{\prime} on f1(Yc)f^{-1}(Y_{c}) and some constant C>0C>0. It follows from Theorem 6.1 that

Hi(f1(Yc),S(ICX(𝕍),φ)L)Hi(f1(Yc),S(ICX(𝕍),φ+Cψ)L)=0,i>0.\displaystyle H^{i}(f^{-1}(Y_{c}),S(IC_{X}(\mathbb{V}),\varphi)\otimes L)\simeq H^{i}(f^{-1}(Y_{c}),S(IC_{X}(\mathbb{V}),\varphi+C\psi)\otimes L)=0,\quad\forall i>0.

Here the isomorphism follows from the boundedness of ψ\psi (Lemma 4.5). Taking the limit c0c\to 0 we see that

Rif(S(ICX(𝕍),φ)L)y=0,i>0.\displaystyle R^{i}f_{\ast}(S(IC_{X}(\mathbb{V}),\varphi)\otimes L)_{y}=0,\quad\forall i>0.

This proves the corollary. ∎

6.3. Kodaira-Nakano-Kazama type vanishing theorem

Theorem 6.5.

Let XX be a weakly pseudoconvex Kähler space of dimension nn. Let 𝕍\mathbb{V} be an \mathbb{R}-polarized variation of Hodge structure on some Zariski open subset XoXregX^{o}\subset X_{\rm reg}. Let (E,h)(E,h) be a hermitian vector bundle on XX. Assume that (S(𝕍)E,h𝕍h)(S(\mathbb{V})\otimes E,h_{\mathbb{V}}\otimes h) is mm-Nakano positive. Then

Hq(X,S(ICX(𝕍))E)=0H^{q}(X,S(IC_{X}(\mathbb{V}))\otimes E)=0

whenever q1q\geq 1 and mmin{nq+1,rkS(𝕍)+rkE}m\geq{\rm min}\{n-q+1,{\rm rk}S(\mathbb{V})+{\rm rk}E\}.

Proof.

By Theorem 5.1 there is a quasi-isomorphism

S(ICX(𝕍))Eq.i.s.𝒟X,ωn,(S(𝕍)E,h𝕍h)S(IC_{X}(\mathbb{V}))\otimes E\simeq_{\rm q.i.s.}\mathscr{D}^{n,\bullet}_{X,\omega}(S(\mathbb{V})\otimes E,h_{\mathbb{V}}\otimes h)

where ω\omega is a Kähler hermitian metric on XX. Notice that 𝒟X,ωn,(S(𝕍)E,h𝕍h)\mathscr{D}^{n,\bullet}_{X,\omega}(S(\mathbb{V})\otimes E,h_{\mathbb{V}}\otimes h) is a complex of fine sheaves by Lemma 3.2. Hence

Hq(X,S(ICX(𝕍))E)Hq(Γ(X,𝒟X,ωn,(S(𝕍)E,h𝕍h))).\displaystyle H^{q}(X,S(IC_{X}(\mathbb{V}))\otimes E)\simeq H^{q}\left(\Gamma(X,\mathscr{D}^{n,\bullet}_{X,\omega}(S(\mathbb{V})\otimes E,h_{\mathbb{V}}\otimes h))\right).

Now let q>0q>0 such that mmin{nq+1,rkS(𝕍)+rkE}m\geq{\rm min}\{n-q+1,{\rm rk}S(\mathbb{V})+{\rm rk}E\}. Let αΓ(X,𝒟X,ωn,q(S(𝕍)E,h𝕍h))\alpha\in\Gamma(X,\mathscr{D}^{n,q}_{X,\omega}(S(\mathbb{V})\otimes E,h_{\mathbb{V}}\otimes h)) be a locally L2L^{2} form such that ¯α=0\bar{\partial}\alpha=0. It suffices to find βΓ(X,𝒟X,ωn,q1(S(𝕍)E,h𝕍h))\beta\in\Gamma(X,\mathscr{D}^{n,q-1}_{X,\omega}(S(\mathbb{V})\otimes E,h_{\mathbb{V}}\otimes h)) so that ¯β=α\bar{\partial}\beta=\alpha.

Denote A=[1Θh𝕍h(S(𝕍)E),Λω]A=[\sqrt{-1}\Theta_{h_{\mathbb{V}}\otimes h}(S(\mathbb{V})\otimes E),\Lambda_{\omega}]. Let ψ\psi be a smooth psh exhausted function on XX and χ\chi a convex increasing function. Denote hχ:=eχψhh_{\chi}:=e^{-\chi\circ\psi}h. Then

1Θh𝕍hχ(S(𝕍)E)\displaystyle\sqrt{-1}\Theta_{h_{\mathbb{V}}\otimes h_{\chi}}(S(\mathbb{V})\otimes E) =1Θh𝕍h(S(𝕍)E)+1¯(χψ)IdS(𝕍)E\displaystyle=\sqrt{-1}\Theta_{h_{\mathbb{V}}\otimes h}(S(\mathbb{V})\otimes E)+\sqrt{-1}\partial\bar{\partial}(\chi\circ\psi)\otimes{\rm Id}_{S(\mathbb{V})\otimes E}
1Θh𝕍h(S(𝕍)E).\displaystyle\geq\sqrt{-1}\Theta_{h_{\mathbb{V}}\otimes h}(S(\mathbb{V})\otimes E).

Denote Aχ=[1Θh𝕍hχ(S(𝕍)E),Λω]A_{\chi}=[\sqrt{-1}\Theta_{h_{\mathbb{V}}\otimes h_{\chi}}(S(\mathbb{V})\otimes E),\Lambda_{\omega}]. Then AχA>0A_{\chi}\geq A>0 in bidegree (n,q)(n,q). Thus the integrals

Xo|α|ω,hχ2volω=Xo|α|ω,h2eχψvolω\int_{X^{o}}|\alpha|_{\omega,h_{\chi}}^{2}{\rm vol_{\omega}}=\int_{X^{o}}|\alpha|_{\omega,h}^{2}e^{-\chi\circ\psi}{\rm vol_{\omega}}

and

Xo(Aχ1α,α)ω,hχ2volωXo(A1α,α)ω,h2eχψvolω\int_{X^{o}}(A^{-1}_{\chi}\alpha,\alpha)_{\omega,h_{\chi}}^{2}{\rm vol_{\omega}}\leq\int_{X^{o}}(A^{-1}\alpha,\alpha)_{\omega,h}^{2}e^{-\chi\circ\psi}{\rm vol_{\omega}}

are finite if χ\chi grows fast enough at infinity. By Lemma 5.2,

XcXo:={xXo|χψ(x)<c}X_{c}\cap X^{o}:=\{x\in X^{o}|\chi\circ\psi(x)<c\}

admits a complete Kähler metric for every cc\in\mathbb{R}. Subsequently, [Demailly2012, Theorem 5.1] implies that there exists

βcL(2)n,q1(XcXo,S(𝕍)E;ω,h𝕍hχ)Γ(Xc,𝒟X,ωn,q1(S(𝕍)E,h𝕍h))\beta_{c}\in L^{n,q-1}_{(2)}(X_{c}\cap X^{o},S(\mathbb{V})\otimes E;\omega,h_{\mathbb{V}}\otimes h_{\chi})\subset\Gamma(X_{c},\mathscr{D}^{n,q-1}_{X,\omega}(S(\mathbb{V})\otimes E,h_{\mathbb{V}}\otimes h))

such that ¯βc=α|Xc\bar{\partial}\beta_{c}=\alpha|_{X_{c}} and

βcω,h𝕍hχ1qXo(A1α,α)ω,h2eχψvolω.\|\beta_{c}\|_{\omega,h_{\mathbb{V}}\otimes h_{\chi}}\leq\frac{1}{q}\int_{X^{o}}(A^{-1}\alpha,\alpha)_{\omega,h}^{2}e^{-\chi\circ\psi}{\rm vol_{\omega}}.

By taking a weak limit of a certain subsequence of {βc}\{\beta_{c}\} we obtain the solution of the equation α=¯β\alpha=\bar{\partial}\beta such that βΓ(X,𝒟X,ωn,q1(S(𝕍)E,h𝕍h))\beta\in\Gamma(X,\mathscr{D}^{n,q-1}_{X,\omega}(S(\mathbb{V})\otimes E,h_{\mathbb{V}}\otimes h)). This proves the theorem. ∎

6.4. Enoki-Kollár type injectivity theorem

Theorem 6.6.

Let XX be a compact Kähler space of dimension nn. Let 𝕍\mathbb{V} be an \mathbb{R}-polarized variation of Hodge structure on some Zariski open subset XoXregX^{o}\subset X_{\rm reg}. Let (L,hφL)(L,h_{\varphi_{L}}) and (F,hφF)(F,h_{\varphi_{F}}) be holomorphic line bundles on XX with singular hermitian metrics hφL:=eφLhh_{\varphi_{L}}:=e^{-\varphi_{L}}h and hφF:=eφFhh_{\varphi_{F}}:=e^{-\varphi_{F}}h^{\prime} which are smooth on a Zariski open subset UXU\subset X. Assume the following conditions.

  1. (1)

    1ΘhφF(F)0\sqrt{-1}\Theta_{h_{\varphi_{F}}}(F)\geq 0 on UU.

  2. (2)

    1(ΘhφF(F)ϵΘhφL(L))0\sqrt{-1}(\Theta_{h_{\varphi_{F}}}(F)-\epsilon\Theta_{h_{\varphi_{L}}}(L))\geq 0 on UU for some positive constant ϵ\epsilon.

Let sΓ(X,L)s\in\Gamma(X,L) be a nonzero holomorphic section such that sup|s|hφL<\sup|s|_{h_{\varphi_{L}}}<\infty. Then the multiplication homomorphism

×s:Hq(X,S(ICX(𝕍),φF)F)Hq(X,S(ICX(𝕍),φF+φL)FL)\times s:H^{q}(X,S(IC_{X}(\mathbb{V}),\varphi_{F})\otimes F)\to H^{q}(X,S(IC_{X}(\mathbb{V}),\varphi_{F}+\varphi_{L})\otimes F\otimes L)

is injective for every integer q0q\geq 0.

Proof.

The proof is parallel to the arguments in [Fujino2012]. By virtue of Lemma 5.2 and Theorem 5.1, there exists a complete Kähler metric ω\omega on V:=XoUV:=X^{o}\cap U such that the canonical maps

Hq(X,S(ICX(𝕍),φF)F)H(2)n,q(V,S(𝕍)F;ω,h𝕍hφF)\displaystyle H^{q}(X,S(IC_{X}(\mathbb{V}),\varphi_{F})\otimes F)\to H^{n,q}_{(2)}(V,S(\mathbb{V})\otimes F;\omega,h_{\mathbb{V}}\otimes h_{\varphi_{F}})

and

Hq(X,S(ICX(𝕍),φF+φL)FL)H(2)n,q(V,S(𝕍)FL;ω,h𝕍hφFhφL)\displaystyle H^{q}(X,S(IC_{X}(\mathbb{V}),\varphi_{F}+\varphi_{L})\otimes F\otimes L)\to H^{n,q}_{(2)}(V,S(\mathbb{V})\otimes F\otimes L;\omega,h_{\mathbb{V}}\otimes h_{\varphi_{F}}\otimes h_{\varphi_{L}})

are isomorphisms for every q0q\geq 0. Since S(ICX(𝕍),φF)S(IC_{X}(\mathbb{V}),\varphi_{F}) and S(ICX(𝕍),φF+φL)S(IC_{X}(\mathbb{V}),\varphi_{F}+\varphi_{L}) are coherent, H(2)n,q(V,S(𝕍)F;ω,h𝕍hφF)H^{n,q}_{(2)}(V,S(\mathbb{V})\otimes F;\omega,h_{\mathbb{V}}\otimes h_{\varphi_{F}}) and H(2)n,q(V,S(𝕍)FL;ω,h𝕍hφFhφL)H^{n,q}_{(2)}(V,S(\mathbb{V})\otimes F\otimes L;\omega,h_{\mathbb{V}}\otimes h_{\varphi_{F}}\otimes h_{\varphi_{L}}) are finite dimensional for each q0q\geq 0. Thus there are isomorphisms

H(2)n,q(V,S(𝕍)F;ω,h𝕍hφF)(2)n,q(S(𝕍)F)\displaystyle H^{n,q}_{(2)}(V,S(\mathbb{V})\otimes F;\omega,h_{\mathbb{V}}\otimes h_{\varphi_{F}})\simeq\mathscr{H}^{n,q}_{(2)}(S(\mathbb{V})\otimes F)

and

H(2)n,q(V,S(𝕍)FL;ω,h𝕍hφFhφL)(2)n,q(S(𝕍)FL),\displaystyle H^{n,q}_{(2)}(V,S(\mathbb{V})\otimes F\otimes L;\omega,h_{\mathbb{V}}\otimes h_{\varphi_{F}}\otimes h_{\varphi_{L}})\simeq\mathscr{H}^{n,q}_{(2)}(S(\mathbb{V})\otimes F\otimes L),

where

(2)n,q(S(𝕍)F)={αDmaxn,q(V,S(𝕍)F;ω,h𝕍hφF)|¯α=0,¯α=0}\displaystyle\mathscr{H}^{n,q}_{(2)}(S(\mathbb{V})\otimes F)=\{\alpha\in D^{n,q}_{\rm max}(V,S(\mathbb{V})\otimes F;\omega,h_{\mathbb{V}}\otimes h_{\varphi_{F}})|\bar{\partial}\alpha=0,\bar{\partial}^{\ast}\alpha=0\}

and

(2)n,q(S(𝕍)FL)={αDmaxn,q(V,S(𝕍)FL;ω,h𝕍hφFhφL)|¯α=0,¯α=0}.\displaystyle\mathscr{H}^{n,q}_{(2)}(S(\mathbb{V})\otimes F\otimes L)=\{\alpha\in D^{n,q}_{\rm max}(V,S(\mathbb{V})\otimes F\otimes L;\omega,h_{\mathbb{V}}\otimes h_{\varphi_{F}}\otimes h_{\varphi_{L}})|\bar{\partial}\alpha=0,\bar{\partial}^{\ast}\alpha=0\}.

We claim that the multiplication map

×s:(2)n,q(S(𝕍)F)(2)n,q(S(𝕍)FL)\times s:\mathscr{H}^{n,q}_{(2)}(S(\mathbb{V})\otimes F)\longrightarrow\mathscr{H}^{n,q}_{(2)}(S(\mathbb{V})\otimes F\otimes L)

is well-defined. If the claim is true, the theorem is obvious. Assume that su=0su=0 in (2)n,q(S(𝕍)FL)\mathscr{H}^{n,q}_{(2)}(S(\mathbb{V})\otimes F\otimes L). Since ss is holomorphic over XX, the locus {s0}\{s\neq 0\} is dense in VV. Hence u=0u=0 for u(2)n,q(S(𝕍)F)u\in\mathscr{H}^{n,q}_{(2)}(S(\mathbb{V})\otimes F) since uu is smooth over VV. This implies the desired injectivity. Thus it is sufficient to prove the above claim.

Take an arbitrary u(2)n,q(S(𝕍)F)u\in\mathscr{H}^{n,q}_{(2)}(S(\mathbb{V})\otimes F). Since sup|s|hφL<\sup|s|_{h_{\varphi_{L}}}<\infty, u<\|u\|<\infty and ¯u=0\bar{\partial}u=0, we obtain that

susup|s|hφLu<,¯(su)=0.\|su\|\leq\sup|s|_{h_{\varphi_{L}}}\|u\|<\infty,\quad\bar{\partial}(su)=0.

By the Bochner-Kodaria-Nakano identity,

Δ¯u=ΔDu+[1Θh𝕍hφF(S(𝕍)F),Λ]u,\displaystyle\Delta_{\bar{\partial}}u=\Delta_{D^{\prime}}u+[\sqrt{-1}\Theta_{h_{\mathbb{V}}\otimes h_{\varphi_{F}}}(S(\mathbb{V})\otimes F),\Lambda]u,

where Λ\Lambda is the adjoint of ω\omega\wedge\cdot and =D+¯\nabla=D^{\prime}+\bar{\partial} is the bidegree decomposition of the Chern connection \nabla associated to h𝕍hφFh_{\mathbb{V}}\otimes h_{\varphi_{F}}. Since u(2)n,q(S(𝕍)F)u\in\mathscr{H}^{n,q}_{(2)}(S(\mathbb{V})\otimes F), we have Δ¯u=0\Delta_{\bar{\partial}}u=0. So

(6.4) ΔDu+[1Θh𝕍hφF(S(𝕍)F),Λ]u=0.\displaystyle\Delta_{D^{\prime}}u+[\sqrt{-1}\Theta_{h_{\mathbb{V}}\otimes h_{\varphi_{F}}}(S(\mathbb{V})\otimes F),\Lambda]u=0.

By Assumption (1) and Theorem 2.3, we obtain that

[1Θh𝕍hφF(S(𝕍)F),Λ]u,uh𝕍hφF0,\displaystyle\langle[\sqrt{-1}\Theta_{h_{\mathbb{V}}\otimes h_{\varphi_{F}}}(S(\mathbb{V})\otimes F),\Lambda]u,u\rangle_{h_{\mathbb{V}}\otimes h_{\varphi_{F}}}\geq 0,

where ,h𝕍hφF\langle,\rangle_{h_{\mathbb{V}}\otimes h_{\varphi_{F}}} is the pointwise inner product with respect to h𝕍hφFh_{\mathbb{V}}\otimes h_{\varphi_{F}} and ω\omega. Since ω\omega is complete, we obtain that

(6.5) ΔDu,u=Du2+Du20.\displaystyle\langle\Delta_{D^{\prime}}u,u\rangle=\|D^{\prime}u\|^{2}+\|D^{\prime*}u\|^{2}\geq 0.

By combining (6.4) with (6.5), we obtain that

Du2=Du2=1Θh𝕍hφF(S(𝕍)F)Λu,uh𝕍hφF=0,\|D^{\prime}u\|^{2}=\|D^{\prime*}u\|^{2}=\langle\sqrt{-1}\Theta_{h_{\mathbb{V}}\otimes h_{\varphi_{F}}}(S(\mathbb{V})\otimes F)\Lambda u,u\rangle_{h_{\mathbb{V}}\otimes h_{\varphi_{F}}}=0,

which implies that

(6.6) Du=0 and 1Θh𝕍hφF(S(𝕍)F)Λu,uh𝕍hφF=0.\displaystyle D^{\prime*}u=0\textrm{ and }\langle\sqrt{-1}\Theta_{h_{\mathbb{V}}\otimes h_{\varphi_{F}}}(S(\mathbb{V})\otimes F)\Lambda u,u\rangle_{h_{\mathbb{V}}\otimes h_{\varphi_{F}}}=0.

Since u(2)n,q(S(𝕍)F)u\in\mathscr{H}^{n,q}_{(2)}(S(\mathbb{V})\otimes F) and ss is holomorphic over XX, (6.6) yields

D(su)=¯(su)=sDu=0\displaystyle D^{\prime*}(su)=-\ast\bar{\partial}\ast(su)=sD^{\prime*}u=0

where \ast is the Hodge star operator with respect to ω\omega. Moreover, due to the degree, we have D(su)=0D^{\prime}(su)=0. As a result, ΔD(su)=0\Delta_{D^{\prime}}(su)=0. Applying the Bochner-Kodaria-Nakano identity to susu, we get that

0¯(su)2Δ¯(su),su=1Θh𝕍hφFhφL(S(𝕍)FL)Λsu,suh𝕍hφFhφL.0\leq\|\bar{\partial}^{*}(su)\|^{2}\leq\langle\Delta_{\bar{\partial}}(su),su\rangle=\langle\sqrt{-1}\Theta_{h_{\mathbb{V}}\otimes h_{\varphi_{F}}\otimes h_{\varphi_{L}}}(S(\mathbb{V})\otimes F\otimes L)\Lambda su,su\rangle_{h_{\mathbb{V}}\otimes h_{\varphi_{F}}\otimes h_{\varphi_{L}}}.

By the assumptions and Theorem 2.3 we obtain that

0\displaystyle 0\leq 1Θh𝕍hφFhφL(S(𝕍)FL)Λsu,suh𝕍hφFhφL\displaystyle\langle\sqrt{-1}\Theta_{h_{\mathbb{V}}\otimes h_{\varphi_{F}}\otimes h_{\varphi_{L}}}(S(\mathbb{V})\otimes F\otimes L)\Lambda su,su\rangle_{h_{\mathbb{V}}\otimes h_{\varphi_{F}}\otimes h_{\varphi_{L}}}
\displaystyle\leq (1+1ϵ)|s|hφL21Θh𝕍hφF(S(𝕍)F)Λu,uh𝕍hφF=0.\displaystyle(1+\frac{1}{\epsilon})|s|^{2}_{h_{\varphi_{L}}}\langle\sqrt{-1}\Theta_{h_{\mathbb{V}}\otimes h_{\varphi_{F}}}(S(\mathbb{V})\otimes F)\Lambda u,u\rangle_{h_{\mathbb{V}}\otimes h_{\varphi_{F}}}=0.

Thus ¯(su)=0\bar{\partial}^{*}(su)=0. Hence Δ¯(su)=0\Delta_{\bar{\partial}}(su)=0, equivalently, su(2)n,q(S(𝕍)FL)su\in\mathscr{H}^{n,q}_{(2)}(S(\mathbb{V})\otimes F\otimes L). The proof of the claim is finished. ∎

Assume that FF is a semi-ample holomorphic line bundle such that H0(X,FkL1)0H^{0}(X,F^{\otimes k}\otimes L^{-1})\neq 0 for some k>0k>0. Denote B={s=0}B=\{s=0\} and take a nonzero divisor D|FkL1|D^{\prime}\in|F^{\otimes k}\otimes L^{-1}|. Since FF is semi-ample, there is a smooth reduced divisor DD so that B+D+D|Fm|B+D+D^{\prime}\in|F^{\otimes m}| where mm is large enough so that S(ICX(𝕍),φD+Dm)=S(ICX(𝕍))S(IC_{X}(\mathbb{V}),\frac{\varphi_{D+D^{\prime}}}{m})=S(IC_{X}(\mathbb{V})) (Proposition 4.13). Let hBh_{B} be the singular hermitian metric on LL associated to BB and hD+Dh_{D+D^{\prime}} the singular hermitian metric on 𝒪X(D+D)\mathscr{O}_{X}(D+D^{\prime}) associated to the divisor D+DD+D^{\prime} (Remark 6.3). Then U=Xreg\(BDD)U=X_{\rm reg}\backslash(B\cup D\cup D^{\prime}), hF:=(hBhD+D)1mh_{F}:=(h_{B}h_{D+D^{\prime}})^{\frac{1}{m}} and hBh_{B} satisfy the conditions in Theorem 6.6.

Corollary 6.7 (Esnault-Viehweg type injectivity theorem).

Let XX be a compact Kähler space and let 𝕍\mathbb{V} be an \mathbb{R}-polarized variation of Hodge structure on some Zariski open subset XoXregX^{o}\subset X_{\rm reg}. Let (L,hL)(L,h_{L}) be a holomorphic line bundle on XX and FF a semi-ample holomorphic line bundle such that H0(X,FkL1)0H^{0}(X,F^{\otimes k}\otimes L^{-1})\neq 0 for some k>0k>0. Let sΓ(X,L)s\in\Gamma(X,L) be a nonzero holomorphic section. Then the multiplication homomorphism

×s:Hq(X,S(ICX(𝕍))F)Hq(X,S(ICX(𝕍))FL)\times s:H^{q}(X,S(IC_{X}(\mathbb{V}))\otimes F)\to H^{q}(X,S(IC_{X}(\mathbb{V}))\otimes F\otimes L)

is injective for every integer q0q\geq 0.

7. Applications in the relative Fujita conjecture

Some applications of the Nadel type vanishing theorem (Theorem 6.1) are given on the separation of jets of adjoint bundles in the relative case. Before that we recall the singular metrics constructed by Angehrn-Siu [Siu1995] and Demailly [Demailly2012].

Proposition 7.1 (Angehrn-Siu [Siu1995]).

Let XX be a smooth projective algebraic variety and LL an ample line bundle on XX. Assume that there is a rational number κ>0\kappa>0 such that

LkW(12n(n+2r1)+κ)dL^{k}\cdot W\geq\left(\frac{1}{2}n(n+2r-1)+\kappa\right)^{d}

for any irreducible subvariety WW of dimension 0dn0\leq d\leq n in XX. Let x1,,xrXx_{1},\dots,x_{r}\in X and denote m0=12n(n+2r1)m_{0}=\frac{1}{2}n(n+2r-1). Then there is a rational number 0<ϵ<κ0<\epsilon<\kappa and a singular hermitian metric hh^{\prime} on Lm0+ϵm0+κL^{\otimes\frac{m_{0}+\epsilon}{m_{0}+\kappa}} with analytic singularities such that

  1. (1)

    1Θh(Lm0+ϵm0+κ)0\sqrt{-1}\Theta_{h^{\prime}}(L^{\otimes\frac{m_{0}+\epsilon}{m_{0}+\kappa}})\geq 0;

  2. (2)

    x1,,xrsupp𝒪X/(h)x_{1},\dots,x_{r}\in{\rm supp}\mathscr{O}_{X}/\mathscr{I}(h^{\prime}) and x1x_{1} is an isolated point of supp𝒪X/(h){\rm supp}\mathscr{O}_{X}/\mathscr{I}(h^{\prime}). Moreover there is a neighborhood UU of x1x_{1} so that hh^{\prime} is smooth on U\{x1}U\backslash\{x_{1}\}.

Proof.

This is statement ()0(*)_{0} on page 299 in [Siu1995] which is proved in Lemma 9.2 in loc. cit. ∎

Proposition 7.2.

(Demailly [Demailly2012, Theorem 7.4]) Let XX be a smooth projective algebraic variety and LL an ample line bundle on XX. Let ω\omega be a Kähler form on XX. Let x1,,xrXx_{1},\dots,x_{r}\in X and let s1,,srs_{1},\dots,s_{r}\in\mathbb{N} be non-negative integers. Denote m0=2+1jr(3n+2sj1n)m_{0}=2+\sum_{1\leq j\leq r}\binom{3n+2s_{j}-1}{n}. Then there is a singular hermitian metric hh^{\prime} on ωXLm0\omega_{X}\otimes L^{\otimes m_{0}} with analytic singularities such that

  1. (1)

    1Θh(ωXLm0)ϵω\sqrt{-1}\Theta_{h^{\prime}}(\omega_{X}\otimes L^{\otimes m_{0}})\geq\epsilon\omega for some ϵ>0\epsilon>0.

  2. (2)

    The singular loci of hh^{\prime} is isolated and the weight φ\varphi of hh^{\prime} (h=eφh0h^{\prime}=e^{-\varphi}h_{0} for some smooth hermitian metric h0h_{0}) satisfies that

    ν(φ,xj)n+sj,j=1,,r.\nu(\varphi,x_{j})\geq n+s_{j},\quad j=1,\dots,r.
Theorem 7.3.

Let XX be a projective nn-fold and DD a (possibly empty) normal crossing divisor on XX. Assume that 𝕍\mathbb{V} is an \mathbb{R}-polarized variation of Hodge structure on Xo:=X\DX^{o}:=X\backslash D. Let LL be an ample line bundle on XX. Assume that there is a positive number κ>0\kappa>0 such that

LkW(12n(n+2r1)+κ)dL^{k}\cdot W\geq\left(\frac{1}{2}n(n+2r-1)+\kappa\right)^{d}

for any irreducible subvariety WW of dimension 0dn0\leq d\leq n in XX. Then the global holomorphic sections of S(ICX(𝕍))LS(IC_{X}(\mathbb{V}))\otimes L separate any set of rr distinct points x1,,xrXx_{1},\dots,x_{r}\in X, i.e. there is a surjective map

H0(X,S(ICX(𝕍))L)1krS(ICX(𝕍))L𝒪X,xk/mX,xk.H^{0}(X,S(IC_{X}(\mathbb{V}))\otimes L)\to\bigoplus_{1\leq k\leq r}S(IC_{X}(\mathbb{V}))\otimes L\otimes\mathscr{O}_{X,x_{k}}/m_{X,x_{k}}.
Proof.

By induction on rr, the canonical morphism

H0(X,S(ICX(𝕍))L)2krS(ICX(𝕍))L𝒪X,xk/mX,xkH^{0}(X,S(IC_{X}(\mathbb{V}))\otimes L)\to\bigoplus_{2\leq k\leq r}S(IC_{X}(\mathbb{V}))\otimes L\otimes\mathscr{O}_{X,x_{k}}/m_{X,x_{k}}

is surjective. It is therefore sufficient to show that, for every vS(ICX(𝕍))L𝒪X,x1/mX,x1v\in S(IC_{X}(\mathbb{V}))\otimes L\otimes\mathscr{O}_{X,x_{1}}/m_{X,x_{1}}, there is a section s1H0(X,S(ICX(𝕍))L)s_{1}\in H^{0}(X,S(IC_{X}(\mathbb{V}))\otimes L) such that s1(x1)=vs_{1}(x_{1})=v and s1(xk)=0s_{1}(x_{k})=0, k=2,,rk=2,\dots,r.

By Proposition 7.1, there is a singular hermitian metric hh^{\prime} on Lm0+ϵm0+κL^{\otimes\frac{m_{0}+\epsilon}{m_{0}+\kappa}} so that

  1. (1)

    hh^{\prime} has analytic singularities;

  2. (2)

    1Θh(Lm0+ϵm0+κ)0\sqrt{-1}\Theta_{h^{\prime}}(L^{\otimes\frac{m_{0}+\epsilon}{m_{0}+\kappa}})\geq 0;

  3. (3)

    x1,,xrsupp𝒪X/(h)x_{1},\dots,x_{r}\in{\rm supp}\mathscr{O}_{X}/\mathscr{I}(h^{\prime}) and x1x_{1} is an isolated point of supp𝒪X/(h){\rm supp}\mathscr{O}_{X}/\mathscr{I}(h^{\prime}).

Let h′′h^{\prime\prime} be a smooth hermitian metric on Lκϵm0+κL^{\otimes\frac{\kappa-\epsilon}{m_{0}+\kappa}} such that 1Θh′′(Lκϵm0+κ)>0\sqrt{-1}\Theta_{h^{\prime\prime}}(L^{\otimes\frac{\kappa-\epsilon}{m_{0}+\kappa}})>0. Let h=hh′′h=h^{\prime}h^{\prime\prime}. Then

(7.1) 1Θh(L)1Θh′′(Lκϵm0+κ)>0.\displaystyle\sqrt{-1}\Theta_{h}(L)\geq\sqrt{-1}\Theta_{h^{\prime\prime}}(L^{\otimes\frac{\kappa-\epsilon}{m_{0}+\kappa}})>0.

Let φ\varphi be the weight of hh, i.e. h=eφh0h=e^{-\varphi}h_{0} for some smooth hermitian metric h0h_{0}. Let φ\varphi^{\prime} be a quasi-psh function so that

  1. (1)

    φ\varphi^{\prime} is smooth on X\{x2,,xr}X\backslash\{x_{2},\dots,x_{r}\}.

  2. (2)

    eφe^{-\varphi^{\prime}} is not locally integrable at any of x2,,xrx_{2},\cdots,x_{r}.

  3. (3)

    φφ+C\varphi^{\prime}\geq\varphi+C for some CC\in\mathbb{R}.

Such φ\varphi^{\prime} can be constructed as follows. Since φ\varphi has analytic singularities, we assume that there is an open subset U1U_{1} such that U1¯{x1,,xr}={x1}\overline{U_{1}}\cap\{x_{1},\dots,x_{r}\}=\{x_{1}\} and φ=alog(j=1s|gj|2)+λ\varphi=a\log(\sum_{j=1}^{s}|g_{j}|^{2})+\lambda for some holomorphic functions g1,,gsg_{1},\dots,g_{s} on U1U_{1}, a constant a>0a>0 and a bounded function λ\lambda. Since x1,,xrsupp𝒪X/(h)x_{1},\dots,x_{r}\in{\rm supp}\mathscr{O}_{X}/\mathscr{I}(h), {x1}{g1==gs=0}\{x_{1}\}\subset\{g_{1}=\cdots=g_{s}=0\}. Let φ1=alog(j=1l|gj|2)+λ\varphi^{\prime}_{1}=a\log(\sum_{j=1}^{l}|g_{j}|^{2})+\lambda where gs+1,,glg_{s+1},\dots,g_{l} are holomorphic functions such that {g1==gl=0}=\{g_{1}=\cdots=g_{l}=0\}=\emptyset. Let UU^{\prime} be an open neighborhood of {x2,,xr}\{x_{2},\dots,x_{r}\} such that U1¯U¯=\overline{U_{1}}\cap\overline{U^{\prime}}=\emptyset. Choose a smooth extension φC(X\{x2,,xr})\varphi^{\prime}\in C^{\infty}(X\backslash\{x_{2},\dots,x_{r}\}) of φ1C(U1)\varphi^{\prime}_{1}\in C^{\infty}(U_{1}) and φC(U\{x2,,xr})\varphi\in C^{\infty}(U^{\prime}\backslash\{x_{2},\dots,x_{r}\}). Then φ\varphi^{\prime} is the desired function.

Consider the short exact sequence of sheaves

0S(ICX(𝕍),φ)LS(ICX(𝕍),φ)LS(ICX(𝕍),φ)/S(ICX(𝕍),φ)L0.\displaystyle 0\to S(IC_{X}(\mathbb{V}),\varphi)\otimes L\to S(IC_{X}(\mathbb{V}),\varphi^{\prime})\otimes L\to S(IC_{X}(\mathbb{V}),\varphi^{\prime})/S(IC_{X}(\mathbb{V}),\varphi)\otimes L\to 0.

Taking the cohomologies we obtain the exact sequence

H0(X,S(ICX(𝕍),φ)L)H0(X,S(ICX(𝕍),φ)/S(ICX(𝕍),φ)L)H1(X,S(ICX(𝕍),φ)L).\displaystyle H^{0}(X,S(IC_{X}(\mathbb{V}),\varphi^{\prime})\otimes L)\to H^{0}(X,S(IC_{X}(\mathbb{V}),\varphi^{\prime})/S(IC_{X}(\mathbb{V}),\varphi)\otimes L)\to H^{1}(X,S(IC_{X}(\mathbb{V}),\varphi)\otimes L).

By (7.1) and Theorem 6.1 we have

H1(X,S(ICX(𝕍),φ)L)=0.\displaystyle H^{1}(X,S(IC_{X}(\mathbb{V}),\varphi)\otimes L)=0.

So the canonical map

(7.2) H0(X,S(ICX(𝕍),φ)L)H0(X,S(ICX(𝕍),φ)/S(ICX(𝕍),φ)L)\displaystyle H^{0}(X,S(IC_{X}(\mathbb{V}),\varphi^{\prime})\otimes L)\to H^{0}(X,S(IC_{X}(\mathbb{V}),\varphi^{\prime})/S(IC_{X}(\mathbb{V}),\varphi)\otimes L)

is surjective.

Now let us investigate the structure of S(ICX(𝕍),φ)/S(ICX(𝕍),φ)S(IC_{X}(\mathbb{V}),\varphi^{\prime})/S(IC_{X}(\mathbb{V}),\varphi) at x1x_{1}. Let z1,,znz_{1},\dots,z_{n} be holomorphic coordinates centered at x1x_{1} such that D={z1zs=0}D=\{z_{1}\cdots z_{s}=0\}. Let UXU\subset X denote the domain of the coordinate and denote Di={zi=0}D_{i}=\{z_{i}=0\}, i=1,,s\forall i=1,\dots,s. Let v1~,,vm~\widetilde{v_{1}},\dots,\widetilde{v_{m}} be an L2L^{2}-adapted frame of R(ICX(𝕍))R(IC_{X}(\mathbb{V})) as in Proposition 2.5. Denote

ψi:=j=1sαEj(vi~)log|zj|2.\psi_{i}:=-\sum_{j=1}^{s}\alpha_{E_{j}}(\widetilde{v_{i}})\log|z_{j}|^{2}.

By Proposition 4.8, we see that

S(ICX(𝕍),φ)|UωUi=1m(φ+ψi)|Uvi~\displaystyle S(IC_{X}(\mathbb{V}),\varphi)|_{U}\simeq\omega_{U}\otimes\bigoplus_{i=1}^{m}\mathscr{I}(\varphi+\psi_{i})|_{U}\widetilde{v_{i}}

and

S(ICX(𝕍),φ)|UωUi=1m(φ+ψi)|Uvi~.\displaystyle S(IC_{X}(\mathbb{V}),\varphi^{\prime})|_{U}\simeq\omega_{U}\otimes\bigoplus_{i=1}^{m}\mathscr{I}(\varphi^{\prime}+\psi_{i})|_{U}\widetilde{v_{i}}.

Thus

(7.3) S(ICX(𝕍),φ)/S(ICX(𝕍),φ)|UωUi=1m(φ+ψi)/(φ+ψi)|Uvi~.\displaystyle S(IC_{X}(\mathbb{V}),\varphi^{\prime})/S(IC_{X}(\mathbb{V}),\varphi)|_{U}\simeq\omega_{U}\otimes\bigoplus_{i=1}^{m}\mathscr{I}(\varphi^{\prime}+\psi_{i})/\mathscr{I}(\varphi+\psi_{i})|_{U}\widetilde{v_{i}}.

After a possible shrinking of UU we assume that φ\varphi is smooth on U\{x1}U\backslash\{x_{1}\}. Recall that αEj(vi~)(1,0]\alpha_{E_{j}}(\widetilde{v_{i}})\in(-1,0] for every j=1,,sj=1,\cdots,s in Proposition 2.5. Then eψ1,,eψme^{-\psi_{1}},\dots,e^{-\psi_{m}} are locally integrable on UU. This implies that

(φ+ψi)x𝒪U,x,i=1,,m,xU\{x1}.\mathscr{I}(\varphi+\psi_{i})_{x}\simeq\mathscr{O}_{U,x},\quad i=1,\dots,m,\quad x\in U\backslash\{x_{1}\}.

As a consequence

(7.4) Usupp(φ+ψi)/(φ+ψi){x1}.\displaystyle U\cap{\rm supp}\mathscr{I}(\varphi^{\prime}+\psi_{i})/\mathscr{I}(\varphi+\psi_{i})\subset\{x_{1}\}.

Since φ\varphi^{\prime} is smooth at x1x_{1} and eψ1,,eψme^{-\psi_{1}},\dots,e^{-\psi_{m}} are locally integrable on UU, one gets that

(φ+ψi)x1𝒪U,x1,i=1,,m.\displaystyle\mathscr{I}(\varphi^{\prime}+\psi_{i})_{x_{1}}\simeq\mathscr{O}_{U,x_{1}},\quad i=1,\dots,m.

Since x1supp𝒪U/(φ)x_{1}\in{\rm supp}\mathscr{O}_{U}/\mathscr{I}(\varphi), one sees that

(7.5) (φ+ψi)x1(φ)x1mU,x1,i=1,,m.\displaystyle\mathscr{I}(\varphi+\psi_{i})_{x_{1}}\subset\mathscr{I}(\varphi)_{x_{1}}\subset m_{U,x_{1}},\quad i=1,\dots,m.

By combining (7.3) with (7.4) and (7.5), we obtain a canonical surjective map

H0(X,S(ICX(𝕍),φ)/S(ICX(𝕍),φ)L)S(ICX(𝕍))L𝒪X,x1/mX,x1.H^{0}(X,S(IC_{X}(\mathbb{V}),\varphi^{\prime})/S(IC_{X}(\mathbb{V}),\varphi)\otimes L)\to S(IC_{X}(\mathbb{V}))\otimes L\otimes\mathscr{O}_{X,x_{1}}/m_{X,x_{1}}.

Since (7.2) is surjective, we get a surjective map

H0(X,S(ICX(𝕍),φ)L)S(ICX(𝕍))L𝒪X,x1/mX,x1.H^{0}(X,S(IC_{X}(\mathbb{V}),\varphi^{\prime})\otimes L)\to S(IC_{X}(\mathbb{V}))\otimes L\otimes\mathscr{O}_{X,x_{1}}/m_{X,x_{1}}.

For every vS(ICX(𝕍))L𝒪X,x1/mX,x1v\in S(IC_{X}(\mathbb{V}))\otimes L\otimes\mathscr{O}_{X,x_{1}}/m_{X,x_{1}}, there is a section

s1H0(X,S(ICX(𝕍),φ)L)s_{1}\in H^{0}(X,S(IC_{X}(\mathbb{V}),\varphi^{\prime})\otimes L)

such that s1(x1)=vs_{1}(x_{1})=v. Since φ\varphi^{\prime} is not locally integrable at any of x2,,xrx_{2},\dots,x_{r}, we know that s1(xi)=0s_{1}(x_{i})=0, i=2,,ri=2,\dots,r. This proves the theorem. ∎

Theorem 7.4.

Let XX be a smooth projective algebraic variety and DD a (possibly empty) normal crossing divisor of XX. Let 𝕍\mathbb{V} be an \mathbb{R}-polarized variation of Hodge structure on Xo:=X\DX^{o}:=X\backslash D. Let LL be an ample line bundle and GG a nef line bundle on XX. Then S(ICX(𝕍))ωXLmGS(IC_{X}(\mathbb{V}))\otimes\omega_{X}\otimes L^{\otimes m}\otimes G generates simultaneous jets of order s1,,srs_{1},\dots,s_{r}\in\mathbb{N} at arbitrary points x1,,xrXx_{1},\dots,x_{r}\in X, i.e. there is a surjective map

H0(X,S(ICX(𝕍))ωXLmG)1jrS(ICX(𝕍))ωXLmG𝒪X,xk/mX,xksk+1,H^{0}(X,S(IC_{X}(\mathbb{V}))\otimes\omega_{X}\otimes L^{\otimes m}\otimes G)\to\bigoplus_{1\leq j\leq r}S(IC_{X}(\mathbb{V}))\otimes\omega_{X}\otimes L^{\otimes m}\otimes G\otimes\mathscr{O}_{X,x_{k}}/m^{s_{k}+1}_{X,x_{k}},

provided that m2+1kr(3n+2sk1n)m\geq 2+\sum_{1\leq k\leq r}\binom{3n+2s_{k}-1}{n}. In particular, S(ICX(𝕍))ωXLmGS(IC_{X}(\mathbb{V}))\otimes\omega_{X}\otimes L^{\otimes m}\otimes G is globally generated for m2+(3n1n)m\geq 2+\binom{3n-1}{n}.

Proof.

Denote m0=2+1kp(3n+2sk1n)m_{0}=2+\sum_{1\leq k\leq p}\binom{3n+2s_{k}-1}{n}. Let hh^{\prime} be the singular hermitian metric on ωXLm0\omega_{X}\otimes L^{\otimes m_{0}} as constructed in Proposition 7.2. Since LL and GG are nef, there is a smooth hermitian metric h′′h^{\prime\prime} on L(mm0)GL^{\otimes(m-m_{0})}\otimes G such that

1Θh′′(L(mm0)G)ϵ2ω\sqrt{-1}\Theta_{h^{\prime\prime}}(L^{\otimes(m-m_{0})}\otimes G)\geq-\frac{\epsilon}{2}\omega

where ω\omega is a Kähler form on XX.

Let h:=hh′′h:=h^{\prime}h^{\prime\prime} be the singular hermitian metric on A:=ωXLmGA:=\omega_{X}\otimes L^{\otimes m}\otimes G. Then

  1. (1)

    1Θh(A)ϵ2ω\sqrt{-1}\Theta_{h}(A)\geq\frac{\epsilon}{2}\omega,

  2. (2)

    supp(𝒪X/(h)){\rm supp}(\mathscr{O}_{X}/\mathscr{I}(h)) is 0-dimensional and the weight φ\varphi of hh satisfies

    ν(φ,xk)n+sk,k=1,,r.\nu(\varphi,x_{k})\geq n+s_{k},\quad k=1,\dots,r.

Consider the short exact sequence of sheaves

0S(ICX(𝕍),φ)AS(ICX(𝕍))AS(ICX(𝕍))/S(ICX(𝕍),φ)A0.\displaystyle 0\to S(IC_{X}(\mathbb{V}),\varphi)\otimes A\to S(IC_{X}(\mathbb{V}))\otimes A\to S(IC_{X}(\mathbb{V}))/S(IC_{X}(\mathbb{V}),\varphi)\otimes A\to 0.

Taking the cohomology of this sequence, we obtain the exact sequence

H0(X,S(ICX(𝕍))A)H0(X,S(ICX(𝕍))/S(ICX(𝕍),φ)A)H1(X,S(ICX(𝕍),φ)A).\displaystyle H^{0}(X,S(IC_{X}(\mathbb{V}))\otimes A)\to H^{0}(X,S(IC_{X}(\mathbb{V}))/S(IC_{X}(\mathbb{V}),\varphi)\otimes A)\to H^{1}(X,S(IC_{X}(\mathbb{V}),\varphi)\otimes A).

Since 1Θh(A)ϵ2ω\sqrt{-1}\Theta_{h}(A)\geq\frac{\epsilon}{2}\omega, we have

H1(X,S(ICX(𝕍),φ)A)=0\displaystyle H^{1}(X,S(IC_{X}(\mathbb{V}),\varphi)\otimes A)=0

by Theorem 6.1. So the canonical map

(7.6) H0(X,S(ICX(𝕍))A)H0(X,S(ICX(𝕍))/S(ICX(𝕍),φ)A)\displaystyle H^{0}(X,S(IC_{X}(\mathbb{V}))\otimes A)\to H^{0}(X,S(IC_{X}(\mathbb{V}))/S(IC_{X}(\mathbb{V}),\varphi)\otimes A)

is surjective.

Now let us investigate the structure of S(ICX(𝕍))/S(ICX(𝕍),φ)S(IC_{X}(\mathbb{V}))/S(IC_{X}(\mathbb{V}),\varphi) at xkx_{k}, k=1,,rk=1,\dots,r.

Let z1,,znz_{1},\dots,z_{n} be holomorphic coordinates on a neighborhood UkU_{k} of xkx_{k}, centered at xkx_{k}, such that D={z1zs=0}D=\{z_{1}\cdots z_{s}=0\}. Denote Di={zi=0},i=1,,sD_{i}=\{z_{i}=0\},\forall i=1,\dots,s. Let v1~,,vm~\widetilde{v_{1}},\dots,\widetilde{v_{m}} be an L2L^{2}-adapted frame of R(ICX(𝕍))R(IC_{X}(\mathbb{V})) at xkx_{k} as in Proposition 2.5. Denote

ψi:=j=1sαEj(vi~)log|zj|2.\psi_{i}:=-\sum_{j=1}^{s}\alpha_{E_{j}}(\widetilde{v_{i}})\log|z_{j}|^{2}.

It follows from Proposition 4.8 that

S(ICX(𝕍),φ)|UkωUki=1m(φ+ψi)|Ukvi~.\displaystyle S(IC_{X}(\mathbb{V}),\varphi)|_{U_{k}}\simeq\omega_{U_{k}}\otimes\bigoplus_{i=1}^{m}\mathscr{I}(\varphi+\psi_{i})|_{U_{k}}\widetilde{v_{i}}.

Hence

(7.7) S(ICX(𝕍))/S(ICX(𝕍),φ)|UkωUki=1m𝒪X/(φ+ψi)|Ukvi~.\displaystyle S(IC_{X}(\mathbb{V}))/S(IC_{X}(\mathbb{V}),\varphi)|_{U_{k}}\simeq\omega_{U_{k}}\otimes\bigoplus_{i=1}^{m}\mathscr{O}_{X}/\mathscr{I}(\varphi+\psi_{i})|_{U_{k}}\widetilde{v_{i}}.

After a possible shrinking of UkU_{k} we assume that φ\varphi is smooth on Uk\{xk}U_{k}\backslash\{x_{k}\}. Since αEj(vi~)(1,0]\alpha_{E_{j}}(\widetilde{v_{i}})\in(-1,0] for every j=1,,sj=1,\cdots,s, eψ1,,eψme^{-\psi_{1}},\dots,e^{-\psi_{m}} are locally integrable on UkU_{k}. This implies that

(φ+ψi)x𝒪U,x,i=1,,m,xUk\{xk}.\mathscr{I}(\varphi+\psi_{i})_{x}\simeq\mathscr{O}_{U,x},\quad i=1,\dots,m,\quad\forall x\in U_{k}\backslash\{x_{k}\}.

As a consequence,

(7.8) Usupp𝒪X/(φ+ψi){xk},i=1,,m.\displaystyle U\cap{\rm supp}\mathscr{O}_{X}/\mathscr{I}(\varphi+\psi_{i})\subset\{x_{k}\},\quad i=1,\dots,m.

It is evident that ν(φ,xk)n+sk\nu(\varphi,x_{k})\geq n+s_{k}, which implies that

(7.9) (φ+ψi)xk(φ)xkmX,xksk+1,i=1,,m.\displaystyle\mathscr{I}(\varphi+\psi_{i})_{x_{k}}\subset\mathscr{I}(\varphi)_{x_{k}}\subset m^{s_{k}+1}_{X,x_{k}},\quad i=1,\dots,m.

Combining (7.7), (7.8) and (7.9) we obtain a canonical surjective map

H0(X,S(ICX(𝕍))/S(ICX(𝕍),φ)A)1krS(ICX(𝕍))A𝒪X,xk/mX,xksk+1.H^{0}(X,S(IC_{X}(\mathbb{V}))/S(IC_{X}(\mathbb{V}),\varphi)\otimes A)\to\bigoplus_{1\leq k\leq r}S(IC_{X}(\mathbb{V}))\otimes A\otimes\mathscr{O}_{X,x_{k}}/m^{s_{k}+1}_{X,x_{k}}.

Since (7.6) is surjective, we get the desired surjective map

H0(X,S(ICX(𝕍))A)1krS(ICX(𝕍))A𝒪X,xk/mX,xksk+1.H^{0}(X,S(IC_{X}(\mathbb{V}))\otimes A)\to\bigoplus_{1\leq k\leq r}S(IC_{X}(\mathbb{V}))\otimes A\otimes\mathscr{O}_{X,x_{k}}/m^{s_{k}+1}_{X,x_{k}}.

Let f:YXf:Y\to X be a proper holomorphic morphism from a Kähler manifold to a projective algebraic variety. Assume that the degenerate loci of ff is contained in a normal crossing divisor DXD\subset X. Denote by 𝕍q:=Rqf(f1(X\D))\mathbb{V}^{q}:=R^{q}f_{\ast}(\mathbb{C}_{f^{-1}(X\backslash D)}) the variation of Hodge structure on X\DX\backslash D. Then RqfωYS(ICX(𝕍q))R^{q}f_{\ast}\omega_{Y}\simeq S(IC_{X}(\mathbb{V}^{q})) is a locally free 𝒪X\mathscr{O}_{X}-module for every q0q\geq 0 ([Kollar1986, Theorem 2.6] or [Takegoshi1995, Theorem V]). In this case, Corollary 1.5 and Corollary 1.6 in §1.2 can be deduced.

When LL is ample and base point free, we could obtain the optimal bound.

Theorem 7.5.

Let XX be a projective algebraic variety and let 𝕍\mathbb{V} be an \mathbb{R}-polarized variation of Hodge structure on some Zariski open subset XoXregX^{o}\subset X_{\rm reg}. Let LL be an ample line bundle on XX which is generated by global sections and GG is a nef line bundle on XX. Then S(ICX(𝕍))LmGS(IC_{X}(\mathbb{V}))\otimes L^{\otimes m}\otimes G generates simultaneous jets of order s1,,srs_{1},\dots,s_{r}\in\mathbb{N} at arbitrary points x1,,xrXx_{1},\dots,x_{r}\in X, i.e. there is a surjective map

H0(X,S(ICX(𝕍))LmG)1krS(ICX(𝕍))LmG𝒪X,xk/mX,xksk+1,H^{0}(X,S(IC_{X}(\mathbb{V}))\otimes L^{\otimes m}\otimes G)\to\bigoplus_{1\leq k\leq r}S(IC_{X}(\mathbb{V}))\otimes L^{\otimes m}\otimes G\otimes\mathscr{O}_{X,x_{k}}/m^{s_{k}+1}_{X,x_{k}},

provided that mdimX+1kr(sk+1)m\geq\dim_{\mathbb{C}}X+\sum_{1\leq k\leq r}(s_{k}+1).

Proof.

By Theorem 6.1, it follows that S(ICX(𝕍))L(dimX+1)GS(IC_{X}(\mathbb{V}))\otimes L^{\otimes(\dim_{\mathbb{C}}X+1)}\otimes G is 0-regular in the sense of Castelnuovo-Mumford [Mumford1966]. Consequently, the theorem follows from [Shentu2020, Theorem 1.1]. ∎

References