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

L2L^{2}-Extension of Adjoint bundles and Kollár’s Conjecture

Junchao Shentu stjc@ustc.edu.cn School of Mathematical Sciences, University of Science and Technology of China, Hefei, 230026, China  and  Chen Zhao czhao@ustc.edu.cn School of Mathematical Sciences, University of Science and Technology of China, Hefei, 230026, China
Abstract.

We give a new proof of Kollár’s conjecture on the pushforward of the dualizing sheaf twisted by a variation of Hodge structure. This conjecture was settled by M. Saito via mixed Hodge modules and has applications in the investigation of Albanese maps. Our technique is the L2L^{2}-method and we give a concrete construction and proofs of the conjecture. The L2L^{2} point of view allows us to generalize Kollár’s conjecture to the context of non-abelian Hodge theory.

1. Introduction

Let f:XYf:X\rightarrow Y be a surjective morphism between complex projective varieties. Assume that XX is smooth and denote by ωX\omega_{X} its dualizing sheaf. In [Kollar1986_1, Kollar1986_2], J. Kollár proves the following results which is roughly called the Kollár package in this paper.

Torsion Freeness:

RifωXR^{i}f_{\ast}\omega_{X} is torsion free for i0i\geq 0 and RifωX=0R^{i}f_{\ast}\omega_{X}=0 if i>dimXdimYi>\dim X-\dim Y.

Vanishing Theorem:

If LL is an ample line bundle on YY, then

Hj(Y,RifωXL)=0 for j>0 and i0.H^{j}(Y,R^{i}f_{\ast}\omega_{X}\otimes L)=0\textrm{ for }\forall j>0\textrm{ and }\forall i\geq 0.
Decomposition Theorem:

Rf(ωX)Rf_{\ast}(\omega_{X}) splits in D(Y)D(Y), i.e.

Rf(ωX)qRqf(ωX)[q]D(Y).Rf_{\ast}(\omega_{X})\simeq\bigoplus_{q}R^{q}f_{\ast}(\omega_{X})[-q]\in D(Y).

As a consequence, the spectral sequence

E2pq:=Hp(Y,RqfωX)Hp+q(X,ωX)E^{pq}_{2}:=H^{p}(Y,R^{q}f_{\ast}\omega_{X})\Rightarrow H^{p+q}(X,\omega_{X})

degenerates at the E2E_{2} page.

Motivated by the proofs, Kollár [Kollar1986_2, §5] conjectured that the Kollár package could be put into a more general framework which is closely related to variations of Hodge structure. More precisely, Kollár conjectured that there is a coherent sheaf SX(𝕍)S_{X}(\mathbb{V}) associated to every polarized variation of Hodge structure 𝕍=(𝒱,,,h𝕍)\mathbb{V}=(\mathcal{V},\nabla,\mathcal{F}^{\bullet},h_{\mathbb{V}}) over some dense Zariski open subset of XregX_{\rm reg}, such that the three results above hold when ωX\omega_{X} is replaced by SX(𝕍)S_{X}(\mathbb{V}). This conjecture is perfectly settled in [MSaito1991] by M. Saito’s theory of mixed Hodge modules [MSaito1988, MSaito1990] and has applications in the investigation of Albanese maps.

The purpose of this paper has two sides.

  1. (1)

    Give a concrete construction of SX(𝕍)S_{X}(\mathbb{V}) by using certain L2L^{2} holomorphic sections and reprove Kollár’s conjecture without using mixed Hodge modules. There are, in addition to the concrete construction, two other advantages of the L2L^{2} method:

    1. (a)

      It allows us to prove Kollár’s conjecture for proper Kähler morphisms. The machinery of mixed Hodge modules does not work for this more general setting because the decomposition theorem of Saito with respect to a proper Kähler morphism is still a conjecture ([MSaito1990(3), Conjecture 0.4]).

    2. (b)

      It allows us to prove Kollár’s conjecture with a twisted coefficients in a hermitian vector bundle with a Nakano semi-positive curvature. This flexibility is convenient for applications.

  2. (2)

    We observe that, rather than the structure of the variation of Hodge structure, the validity of the Kollár package for SX(𝕍)S_{X}(\mathbb{V}) is a consequence of the Nakano semipositivity of the top Hodge bundle S(𝕍):=max{k|k0}S(\mathbb{V}):=\mathcal{F}^{\max\{k|\mathcal{F}^{k}\neq 0\}} (see §4 for the abstract Kollár package). This allows us to further generalize Kollar’s conjecture to the context of non-abelian Hodge theory. For example, we show that the Kollár package holds for SX(E,h)S_{X}(E,h) (see below for definition) when EE is a subbundle of a tame harmonic bundle (H,θ,h)(H,\theta,h) such that θ¯(E)=0\overline{\theta}(E)=0 and the second fundamental form of EHE\subset H vanishes. A typical example is a system of Hodge bundles in the sense of C. Simpson [Simpson1988] on some Zariski open subset XoXregX^{o}\subset X_{\rm reg} while EE is the top nonzero Hodge summand. This is an example that supports the principle that many results for variation of Hodge structure hold for harmonic bundles. Readers may see the hard Lefschetz theorems of Simpson [Simpson1992], C. Sabbah [Sabbah2005] and T. Mochizuki [Mochizuki20072, Mochizuki20071], the vanishing theorems of D. Arapura [Arapura2019] and Y. Deng-F. Hao [Dengya2019] and Mochizuki’s degeneration theory for twistor structures [Mochizuki20072] for other examples supporting this principle. Notice that the Kollár conjecture for complex variations of Hodge structure may be proved in the patten of Saito [MSaito1991] by using the theory of complex Hodge modules (see [Popa-Schnell2017] or the MHM Project by C. Sabbah and C. Schnell on the homepage of Sabbah.).

Let XX be a complex space and XoXregX^{o}\subset X_{\rm reg} a dense Zariski open subset. Let (E,h)(E,h) be a hermitian vector bundle on XoX^{o}. Denote by j:XoXj:X^{o}\to X the open immersion. The main object of the present paper is the subsheaf SX(E,h)j(KXoE)S_{X}(E,h)\subset j_{\ast}(K_{X^{o}}\otimes E) consisting of the holomorphic forms which are locally square integrable at every point of XX. This class of objects contains Saito’s SS-sheaf and shares many properties of the dualizing sheaf. It has the following features:

  1. (1)

    SX(E,h)S_{X}(E,h) is a torsion free 𝒪X\mathscr{O}_{X}-module. It is coherent when (E,h)(E,h) is Nakano semi-positive and tame (Definition 2.8).

  2. (2)

    SX(E,h)S_{X}(E,h) has the functorial property (Proposition 2.5): Let π:XX\pi:X^{\prime}\to X be a proper bimeromorphic map such that π|π1Xo:π1XoXo\pi|_{\pi^{-1}X^{o}}:\pi^{-1}X^{o}\to X^{o} is biholomorphic, then

    SX(E,h)π(SX(πE,πh)).\displaystyle S_{X}(E,h)\simeq\pi_{\ast}(S_{X^{\prime}}(\pi^{\ast}E,\pi^{\ast}h)).
  3. (3)

    Let 𝕍:=(𝒱,,,h𝕍)\mathbb{V}:=(\mathcal{V},\nabla,\mathcal{F}^{\bullet},h_{\mathbb{V}}) be an \mathbb{R}-polarized variation of Hodge structure where h𝕍h_{\mathbb{V}} is the Hodge metric. Denote by S(𝕍):=max{k|k0}S(\mathbb{V}):=\mathcal{F}^{\max\{k|\mathcal{F}^{k}\neq 0\}} the top nonzero Hodge piece and by S(ICX(𝕍))S(IC_{X}(\mathbb{V})) the Saito’s SS-sheaf associated to the Hodge module ICX(𝕍)IC_{X}(\mathbb{V}). Then

    SX(S(𝕍),h𝕍)S(ICX(𝕍))([SC2021, Theorem 4.10]).S_{X}(S(\mathbb{V}),h_{\mathbb{V}})\simeq S(IC_{X}(\mathbb{V}))\quad\textrm{(\cite[cite]{[\@@bibref{}{SC2021}{}{}, Theorem 4.10]})}.
  4. (4)

    Consider a tame harmonic bundle (H,θ,h)(H,\theta,h) and a holomorphic subbundle EHE\subset H with vanishing second fundamental form. Assume that θ¯(E)=0\overline{\theta}(E)=0. Then the corresponding SX(E,h)S_{X}(E,h) is a coherent sheaf (Proposition 5.6) and satisfies the Kollár package (Theorem 1.1). When (E,θ,h)(E,\theta,h) is the harmonic bundle associated to an \mathbb{R}-polarized variation of Hodge structure 𝕍:=(𝒱,,,h𝕍)\mathbb{V}:=(\mathcal{V},\nabla,\mathcal{F}^{\bullet},h_{\mathbb{V}}) and E=S(𝕍)E=S(\mathbb{V}), this is reduced to Saito’s S(ICX(𝕍))S(IC_{X}(\mathbb{V})).

All complex spaces are assumed to be separated, reduced, paracompact, countable at infinity and of pure dimension throughout the present paper. We would like to point out that the complex spaces are allowed to be non-irreducible. The main result of the present paper is the following

Theorem 1.1.

Let f:XYf:X\rightarrow Y be a proper locally Kähler morphism (Definition 4.1) from a complex space XX to an irreducible complex space YY. Assume that every irreducible component of XX is mapped onto YY. Let XoXregX^{o}\subset X_{\rm reg} be a dense Zariski open subset and (H,θ,h)(H,\theta,h) a tame harmonic bundle on XoX^{o}. Let EHE\subset H be a holomorphic subbundle with vanishing second fundamental form. Let θ¯\overline{\theta} be the adjoint of θ\theta and assume that θ¯(E)=0\overline{\theta}(E)=0. Let FF be a Nakano semi-positive vector bundle on XX. Then the following statements hold.

Torsion Freeness:

Rqf(SX(E,h)F)R^{q}f_{\ast}(S_{X}(E,h)\otimes F) is torsion free for every q0q\geq 0 and vanishes if q>dimXdimYq>\dim X-\dim Y.

Injectivity:

If LL is a semi-positive holomorphic line bundle so that LlL^{l} admits a nonzero holomorphic global section ss for some l>0l>0, then the canonical morphism

Rqf(×s):Rqf(SX(E,h)FLk)Rqf(SX(E,h)FLk+l)R^{q}f_{\ast}(\times s):R^{q}f_{\ast}(S_{X}(E,h)\otimes F\otimes L^{\otimes k})\to R^{q}f_{\ast}(S_{X}(E,h)\otimes F\otimes L^{\otimes k+l})

is injective for every q0q\geq 0 and every k1k\geq 1

Vanishing:

If YY is a projective algebraic variety and LL is an ample line bundle on YY, then

Hq(Y,Rpf(SX(E,h)F)L)=0,q>0,p0.H^{q}(Y,R^{p}f_{\ast}(S_{X}(E,h)\otimes F)\otimes L)=0,\quad\forall q>0,\forall p\geq 0.
Decomposition:

Assume moreover that XX is a compact Kähler space, then Rf(SX(E,h)F)Rf_{\ast}(S_{X}(E,h)\otimes F) splits in D(Y)D(Y), i.e.

Rf(SX(E,h)F)qRqf(SX(E,h)F)[q]D(Y).Rf_{\ast}(S_{X}(E,h)\otimes F)\simeq\bigoplus_{q}R^{q}f_{\ast}(S_{X}(E,h)\otimes F)[-q]\in D(Y).

As a consequence, the spectral sequence

E2pq:Hp(Y,Rqf(SX(E,h)F))Hp+q(X,SX(E,h)F)E^{pq}_{2}:H^{p}(Y,R^{q}f_{\ast}(S_{X}(E,h)\otimes F))\Rightarrow H^{p+q}(X,S_{X}(E,h)\otimes F)

degenerates at the E2E_{2} page.

When (H,θ,h)(H,\theta,h) is a harmonic bundle associated to an \mathbb{R}-polarized variation of Hodge structure 𝕍:=(𝒱,,,h𝕍)\mathbb{V}:=(\mathcal{V},\nabla,\mathcal{F}^{\bullet},h_{\mathbb{V}}), E=S(𝕍)E=S(\mathbb{V}) and F=𝒪XF=\mathscr{O}_{X}, Theorem 1.1 implies Kollár’s conjecture. In this case, our construction of SX(𝕍)S_{X}(\mathbb{V}) coincides with Saito’s.

The present paper is organized as follows. In Section 2 we introduce the construction of the adjoint L2L^{2}-extension SX(E,h)S_{X}(E,h) of a hermitian bundle (E,h)(E,h) which generalizes Saito’s SS-sheaf. Some fundamental properties of SX(E,h)S_{X}(E,h) are proved and its L2L^{2}-Dolbeault resolution is introduced. Depending on the L2L^{2}-Dolbeault resolution, we introduce a harmonic representation of the derived pushforwards of SX(E,h)S_{X}(E,h) in Section 3 by generalizing K. Takegoshi’s work [Takegoshi1995] to complex spaces. In Section 4 we prove an abstract Kollár package and illustrate the observation that: Kollár’s conjecture is a consequence of the Nakano semi-positivity of the top Hodge piece. We prove the main theorem 1.1 in Section 5 as an application.

Notation: 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.

2. L2L^{2}-extension and its L2L^{2}-Dolbeault resolution

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

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 xUx\in U, a closed immersion ι:UΩ\iota:U\to\Omega into a complex manifold Ω\Omega and a psh (resp. strictly psh) function Λ\Lambda on Ω\Omega such that ιΛ=λ\iota^{\ast}\Lambda=\lambda. By a CC^{\infty} form on XX we mean a CC^{\infty} form α\alpha on XregX_{\rm reg} so that locally at every point xXx\in X there is an open neighborhood xUx\in U, a closed immersion ι:UΩ\iota:U\to\Omega into a holomorphic manifold Ω\Omega and βC(Ω)\beta\in C^{\infty}(\Omega) such that ιβ=α\iota^{\ast}\beta=\alpha on UXregU\cap X_{\rm reg}.

Let (Y,ds2)(Y,ds^{2}) be a hermitian manifold of dimension nn and (E,h)(E,h) a hermitian vector bundle on YY. Let L(2)p,q(Y,E;ds2,h)L^{p,q}_{(2)}(Y,E;ds^{2},h) be the space of measurable EE-valued (p,q)(p,q)-forms on YY which are square integrable 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)\}.

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

(2.1) 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 dense 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 hermitian vector bundle on XoX^{o}. 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 xVxx\in V_{x} so that

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

For each 0p,qn0\leq p,q\leq n, 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

(2.2) 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 2.1.

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 and a holomorphic closed immersion UVU\subset V into a complex manifold VV such that ds2|UdsV2|Uds^{2}|_{U}\sim ds^{2}_{V}|_{U} for some hermitian metric dsV2ds^{2}_{V} on VV. If ds02|Xregds^{2}_{0}|_{X_{\rm reg}} is moreover a Kähler metric, we say that ds02ds^{2}_{0} is a Kähler hermitian metric.

Lemma 2.2.

Let XX be a complex space and XoXregX^{o}\subset X_{\rm reg} a dense Zariski open subset. Let ds2ds^{2} be a hermitian metric on XoX^{o} and (E,h)(E,h) a hermitian vector bundle on XoX^{o}. Suppose that for every point xX\Xox\in X\backslash X^{o} there is a neighborhood xUxx\in U_{x} 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 every pp and qq.

Proof.

If 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 MM. 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 apparently 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

(2.3) 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. ∎

2.2. SX(E,h)S_{X}(E,h) and its basic properties

Let XX be a complex space of dimension nn and XoXregX^{o}\subset X_{\rm reg} a dense Zariski open subset. Let ds2ds^{2} be a hermitian metric on XoX^{o} and (E,h)(E,h) a hermitian vector bundle on XoX^{o}.

Definition 2.3.

Define

(2.4) 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}. Hence ds2ds^{2} is omitted in the notation SX(E,h)S_{X}(E,h).

Proposition 2.4.

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 X\XoX\backslash X^{o} so that X~\widetilde{X} is smooth and π\pi is biholomorphic over XoX^{o}. 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 π1{x}\pi^{-1}\{x\}. Thus it suffices to show that

(2.5) 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

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

and

(2.7) 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. By (2.6) and (2.7) we obtain

(2.8) 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 (2.5). ∎

Proposition 2.5 (Functoriality).

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.

It follows from Proposition 2.4 that

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 π1{x}\pi^{-1}\{x\}. This proves the lemma. ∎

The following simple lemma is important for applications. Roughly speaking, it says that one can shrink the domain of (E,h)(E,h) without changing SX(E,h)S_{X}(E,h). This phenomenon also appears in Saito’s S(M)S(M).

Lemma 2.6.

Let UXoU\subset X^{o} be a dense Zariski open subset. Then

SX(E,h)SX(E|U,h|U).S_{X}(E,h)\simeq S_{X}(E|_{U},h|_{U}).
Proof.

This is a consequence of the fact that if a locally L2L^{2} function on a hermitian manifold is ¯\bar{\partial}-closed away from a Zariski open subset, then it is ¯\bar{\partial}-closed over the whole manifold ([Berndtsson2010, Lemma 1.3]). ∎

Lemma 2.7.

Let (F,hF)(F,h_{F}) be a hermitian vector bundle on XX, then

SX(E,h)FSX(EF,hhF).S_{X}(E,h)\otimes F\simeq S_{X}(E\otimes F,h\otimes h_{F}).
Proof.

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

|i=1raiei|hF2i=1rai2|\sum_{i=1}^{r}a_{i}e_{i}|^{2}_{h_{F}}\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 UXoU\cap X^{o}. 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 (KXoEF)|UXo(K_{X^{o}}\otimes E\otimes F)|_{U\cap X^{o}}. Then

αds2,hhF2i=1rαids2,h2\|\alpha\|^{2}_{ds^{2},h\otimes h_{F}}\sim\sum_{i=1}^{r}\|\alpha_{i}\|^{2}_{ds^{2},h}

is finite if and only if αids2,h2\|\alpha_{i}\|^{2}_{ds^{2},h} is finite for each i=1,,ri=1,\dots,r. This proves the lemma. ∎

At the end of this section, we show that SX(E,h)S_{X}(E,h) is a coherent sheaf if the dual metric hh^{\ast} has at most polynomially growth at the boundary X\XoX\backslash X^{o} and (E,h)(E,h) is Nakano semi-positive.

Let YY be a complex manifold and EE a holomorphic vector bundle on YY. Let ΘA1,1(Y,End(E))\Theta\in A^{1,1}(Y,{\rm End}(E)). Denote Θ0\Theta\geq 0 if Θ\Theta is Nakano semi-positive, i.e.

(2.9) Θ(v,v)0,vΓ(TYE).\displaystyle\Theta(v,v)\geq 0,\quad\forall v\in\Gamma(T_{Y}\otimes E).

Let Θ1,Θ2A1,1(Y,End(E))\Theta_{1},\Theta_{2}\in A^{1,1}(Y,{\rm End}(E)), then Θ1Θ2\Theta_{1}\geq\Theta_{2} stands for Θ1Θ20\Theta_{1}-\Theta_{2}\geq 0. A hermitian vector bundle (E,h)(E,h) is Nakano semi-positive if 1Θh(E)0\sqrt{-1}\Theta_{h}(E)\geq 0.

Definition 2.8.

(E,h)(E,h) has tame singularity on XX if, for every point xXx\in X, there is an open neighborhood UU, a proper bimeromorphic morphism π:U~U\pi:\widetilde{U}\to U which is biholomorphic over UXoU\cap X^{o}, and a hermitian vector bundle (Q,hQ)(Q,h_{Q}) on U~\widetilde{U} such that

  1. (1)

    πE|π1(XoU)Q|π1(XoU)\pi^{\ast}E|_{\pi^{-1}(X^{o}\cap U)}\subset Q|_{\pi^{-1}(X^{o}\cap U)} as a subsheaf.

  2. (2)

    There is a hermitian metric hQh^{\prime}_{Q} on Q|π1(XoU)Q|_{\pi^{-1}(X^{o}\cap U)} so that hQ|πEπhh^{\prime}_{Q}|_{\pi^{\ast}E}\sim\pi^{\ast}h on π1(XoU)\pi^{-1}(X^{o}\cap U) and

    (2.10) (i=1rπfi2)chQhQ\displaystyle(\sum_{i=1}^{r}\|\pi^{\ast}f_{i}\|^{2})^{c}h_{Q}\lesssim h^{\prime}_{Q}

    for some cc\in\mathbb{R}. Here (f1,,fr)(f_{1},\dots,f_{r}) is an arbitrary local generator of the ideal sheaf defining U~\π1XoU~\widetilde{U}\backslash\pi^{-1}X^{o}\subset\widetilde{U}.

The tameness condition (2.10) is independent of the choice of the local generator. In the present paper, a tame hermitian vector bundle (E,h)(E,h) is constructed as a subsheaf of a tame harmonic bundle (see §5.2, especially Proposition 5.6) in the sense of Simpson [Simpson1990] and T. Mochizuki [Mochizuki20072, Mochizuki20071]. In this case, condition (2.10) comes from the tameness condition of the harmonic bundle. This is the origin of the name ”tame hermitian vector bundle”.

Proposition 2.9.

SX(E,h)S_{X}(E,h) is a coherent sheaf if (E,h)(E,h) is Nakano semi-positive and tame on XX.

Proof.

Since the problem is local, we assume that XX is a germ of complex space. Let π:X~X\pi:\widetilde{X}\to X be a desingularization so that π\pi is biholomorphic over XoX^{o} and D:=π1(X\Xo)D:=\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. Since (E,h)(E,h) is tame, we assume that there is a hermitian vector bundle (Q,hQ)(Q,h_{Q}) on X~\widetilde{X} such that EE is a subsheaf of Q|XoQ|_{X^{o}} and there exists mm\in\mathbb{N} such that

(2.11) z1zr2mhQ|Eπh\displaystyle\|z_{1}\cdots z_{r}\|^{2m}h_{Q}|_{E}\lesssim\pi^{\ast}h

where (z1,,zn)(z_{1},\cdots,z_{n}) is a local coordinate of X~\widetilde{X} so that D={z1zr=0}D=\{z_{1}\cdots z_{r}=0\}. There is moreover a hermitian metric hQh^{\prime}_{Q} on Q|XoQ|_{X^{o}} so that hQ|Ehh^{\prime}_{Q}|_{E}\sim h. By Proposition 2.5 there is an isomorphism

(2.12) SX(E,h)π(SX~(E,h)).\displaystyle S_{X}(E,h)\simeq\pi_{\ast}\left(S_{\widetilde{X}}(E,h)\right).

Since π\pi is a proper map, it suffices to show that SX~(E,h)S_{\widetilde{X}}(E,h) 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 D={z1zr=0}D=\{z_{1}\cdots z_{r}=0\}. Without loss of generality we assume that QQ admits a global holomorphic frame e1,,ele_{1},\dots,e_{l} which is orthonormal with respect to hQh_{Q}, i.e.

(2.13) ei,ejhQ={1,i=j0,ij.\displaystyle\langle e_{i},e_{j}\rangle_{h_{Q}}=\begin{cases}1,&i=j\\ 0,&i\neq j\end{cases}.

There is a complete Kähler metric on XoX^{o} by Lemma 2.13. Since QQ is coherent111We do not distinguish holomorphic vector bundle and its sheaf of holomorphic sections., the space Γ(X~,SX~(E,h))\Gamma(\widetilde{X},S_{\widetilde{X}}(E,h)) generates a coherent subsheaf 𝒥\mathscr{J} of QQ by strong Noetherian property of coherent sheaves. We have the inclusion 𝒥SX~(E,h)\mathscr{J}\subset S_{\widetilde{X}}(E,h) by the construction. It remains to prove the converse. By the Krull’s theorem ([Atiyah1969, Corollary 10.19]), it suffices to show that

(2.14) 𝒥x+SX~(E,h)xmX~,xk+1Q=SX~(E,h)x,k0,xX~.\displaystyle\mathscr{J}_{x}+S_{\widetilde{X}}(E,h)_{x}\cap m_{\widetilde{X},x}^{k+1}Q=S_{\widetilde{X}}(E,h)_{x},\quad\forall k\geq 0,\forall x\in\widetilde{X}.

Let αSX~(E,h)x\alpha\in S_{\widetilde{X}}(E,h)_{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 z2:=i=1nzi2\|z\|^{2}:=\sum_{i=1}^{n}\|z_{i}\|^{2}. Let

(2.15) ψk(z):=2(n+k+rm)log|zx|+z2\displaystyle\psi_{k}(z):=2(n+k+rm)\log|z-x|+\|z\|^{2}

and hψk=eψkhh_{\psi_{k}}=e^{-\psi_{k}}h. Denote ω0:=1¯z2\omega_{0}:=\sqrt{-1}\partial\bar{\partial}\|z\|^{2}, then

(2.16) 1Θhψk(E)=1¯ψk+1Θh(E)ω0.\displaystyle\sqrt{-1}\Theta_{h_{\psi_{k}}}(E)=\sqrt{-1}\partial\bar{\partial}\psi_{k}+\sqrt{-1}\Theta_{h}(E)\geq\omega_{0}.

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

(2.17) ¯(λα)ω0,hψk2¯(λα)ω0,h2¯λL2αω0,h2+λ2¯αω0,h2<\displaystyle\|\bar{\partial}(\lambda\alpha)\|^{2}_{\omega_{0},h_{\psi_{k}}}\sim\|\bar{\partial}(\lambda\alpha)\|^{2}_{\omega_{0},h}\leq\|\bar{\partial}\lambda\|^{2}_{L^{\infty}}\|\alpha\|^{2}_{\omega_{0},h}+\|\lambda\|^{2}\|\bar{\partial}\alpha\|^{2}_{\omega_{0},h}<\infty

Hence Theorem 2.10 gives a solution of the equation ¯β=¯(λα)\bar{\partial}\beta=\bar{\partial}(\lambda\alpha) so that

(2.18) βω0,h2Xo|β|ω0,h2|zx|2(n+k+rm)volω0¯(λα)ω0,hψk2<.\displaystyle\|\beta\|^{2}_{\omega_{0},h}\lesssim\int_{X^{o}}|\beta|^{2}_{\omega_{0},h}|z-x|^{-2(n+k+rm)}{\rm vol}_{\omega_{0}}\lesssim\|\bar{\partial}(\lambda\alpha)\|^{2}_{\omega_{0},h_{\psi_{k}}}<\infty.

Thus γ=βλα\gamma=\beta-\lambda\alpha is holomorphic and γΓ(X~,SX~(E,h))\gamma\in\Gamma(\widetilde{X},S_{\widetilde{X}}(E,h)).

Assume that

β=i=1lfieidz1dzn\beta=\sum_{i=1}^{l}f_{i}e_{i}dz_{1}\wedge\cdots\wedge dz_{n}

for some holormorphic functions f1,,fl𝒪X~(Xo)f_{1},\dots,f_{l}\in\mathscr{O}_{\widetilde{X}}(X^{o}). By (2.11) we obtain that

(2.19) i=1lXo|fi|2z1zr2m|zx|2(n+k+rm)volω0\displaystyle\sum_{i=1}^{l}\int_{X^{o}}|f_{i}|^{2}\|z_{1}\cdots z_{r}\|^{2m}|z-x|^{-2(n+k+rm)}{\rm vol}_{\omega_{0}}
=\displaystyle= Xo|β|ω0,hQ2z1zr2m|zx|2(n+k+rm)volω0\displaystyle\int_{X^{o}}|\beta|^{2}_{\omega_{0},h_{Q}}\|z_{1}\cdots z_{r}\|^{2m}|z-x|^{-2(n+k+rm)}{\rm vol}_{\omega_{0}}
\displaystyle\lesssim Xo|β|ω0,h2|zx|2(n+k+rm)volω0<.\displaystyle\int_{X^{o}}|\beta|^{2}_{\omega_{0},h}|z-x|^{-2(n+k+rm)}{\rm vol}_{\omega_{0}}<\infty.

This implies that z1mzrmfimX~,xk+1+rmz_{1}^{m}\cdots z_{r}^{m}f_{i}\in m_{\widetilde{X},x}^{k+1+rm} for every i=1,,li=1,\dots,l ([Demailly2012, Lemma 5.6]). Therefore βxmX~,xk+1Q\beta_{x}\in m_{\widetilde{X},x}^{k+1}Q and we prove (2.14). ∎

2.3. L2L^{2}-Dolbeault resolution of SX(E,h)S_{X}(E,h)

In this subsection we introduce an L2L^{2}-Dolbeault resolution of SX(E,h)S_{X}(E,h).

First we recall the following useful estimate.

Theorem 2.10.

[Demailly1982, Theorem 5.1] Let YY be a complex manifold of dimension nn which admits a complete Kähler metric. Let (E,h)(E,h) be a hermitian vector bundle such that

1Θh(E)ωIdE\sqrt{-1}\Theta_{h}(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,h)\alpha\in L^{n,q}_{(2)}(Y,E;\omega,h) such that ¯α=0\bar{\partial}\alpha=0, there is βL(2)n,q1(Y,E;ω,h)\beta\in L^{n,q-1}_{(2)}(Y,E;\omega,h) such that ¯β=α\bar{\partial}\beta=\alpha and βω,h2q1αω,h2\|\beta\|^{2}_{\omega,h}\leq q^{-1}\|\alpha\|^{2}_{\omega,h}.

The main theorem of this section is the following

Theorem 2.11.

Let XX be a complex space of dimension nn and ds2ds^{2} a hermitian metric on a dense Zariski open subset XoXregX^{o}\subset X_{\rm reg} with ω\omega its fundamental form. Let (E,h)(E,h) be a Nakano semi-positive hermitian vector bundle. Assume that, locally at every point xXx\in X, there is a neighborhood xUx\in U, 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

(2.20) SX(E,h)𝒟X,ds2n,(E,h)\displaystyle S_{X}(E,h)\to\mathscr{D}^{n,\bullet}_{X,ds^{2}}(E,h)

is a quasi-isomorphism. As a consequence, there is a canonical isomorphism

Hq(X,SX(E,h))H(2)n,q(Xo,E;ds2,h),q0H^{q}(X,S_{X}(E,h))\simeq H^{n,q}_{(2)}(X^{o},E;ds^{2},h),\quad\forall q\geq 0

if XX is compact.

Proof.

It suffices to show that (2.20) is exact at 𝒟X,ds2n,q(E,h)\mathscr{D}^{n,q}_{X,ds^{2}}(E,h) for q>0q>0. Since the problem is local, we consider a point xXx\in X and an open neighborhood xUx\in U which is small enough so that there is C>0C>0 and a bounded psh function ΦC2(UXo)\Phi\in C^{2}(U\cap X^{o}) such that C1¯Φω|UXoC\sqrt{-1}\partial\bar{\partial}\Phi\geq\omega|_{U\cap X^{o}}. Let h=eCΦhh^{\prime}=e^{-C\Phi}h. Since Φ\Phi is bounded and (E,h)(E,h) is Nakano semi-positive, we have hhh^{\prime}\sim h and

1Θh(E|UXo)=C1¯ΦIdE|UXo+1Θh(E|UXo)ωIdE|UXo.\sqrt{-1}\Theta_{h^{\prime}}(E|_{U\cap X^{o}})=C\sqrt{-1}\partial\bar{\partial}\Phi\otimes{\rm Id}_{E}|_{U\cap X^{o}}+\sqrt{-1}\Theta_{h}(E|_{U\cap X^{o}})\geq\omega\otimes{\rm Id}_{E}|_{U\cap X^{o}}.

By Lemma 2.13, we may assume that UXoU\cap X^{o} admits a complete Kähler metric. By Theorem 2.10, we therefore have

H(2)n,q(UXo,E|UXo;ds2,h)=H(2)n,q(UXo,E|UXo;ds2,h)=0,q>0.H^{n,q}_{(2)}(U\cap X^{o},E|_{U\cap X^{o}};ds^{2},h)=H^{n,q}_{(2)}(U\cap X^{o},E|_{U\cap X^{o}};ds^{2},h^{\prime})=0,\quad\forall q>0.

This proves the exactness of (2.20) at 𝒟X,ds2n,q(E,h)\mathscr{D}^{n,q}_{X,ds^{2}}(E,h), q>0q>0. Since ω\omega is locally bounded from below by a hermitian metric, 𝒟X,ds2n,q(E,h)\mathscr{D}^{n,q}_{X,ds^{2}}(E,h) is a fine sheaf for every qq by Lemma 2.2. The second claim therefore follows from the compactness of the space XX. ∎

In order to apply Theorem 2.11, we introduce a type of hermitian metric that is crucial for the present paper.

Definition 2.12.

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

  1. (1)

    ds2ds^{2} admits a (,1)(\infty,1) bounded potential locally on XX if, for every point xXx\in X, there is a neighborhood UU of xx, a function ΦC(U)\Phi\in C^{\infty}(U) such that |Φ|+|dΦ|ds2<|\Phi|+|d\Phi|_{ds^{2}}<\infty and ds2|U1¯Φds^{2}|_{U}\sim\sqrt{-1}\partial\bar{\partial}\Phi.

  2. (2)

    ds2ds^{2} is called locally complete on XX if, for every point xXx\in X, there is a neighborhood UU of xx such that (U¯Xo,ds2)(\overline{U}\cap X^{o},ds^{2}) is complete.

  3. (3)

    ds2ds^{2} is locally bounded from below by a hermitian metric if, for every point xXx\in X, there is a neighborhood UU of xx and a hermitian metric ds02ds^{2}_{0} on UU such that ds02|Uds2|Uds^{2}_{0}|_{U}\lesssim ds^{2}|_{U}.

Lemma 2.13.

Let XX be a weakly pseudoconvex Kähler space with ψ\psi a smooth exhausted psh function on XX. Denote Xc:={xX|ψ(x)<c}X_{c}:=\{x\in X|\psi(x)<c\}. Let XoXX^{o}\subset X be a dense Zariski open subset. Then, for every cc\in\mathbb{R}, there exists a complete Kähler metric ds2ds^{2} on XcXoX_{c}\cap X^{o} such that

  1. (1)

    ds02ds2ds^{2}_{0}\lesssim ds^{2} for some hermitian metric ds02ds^{2}_{0} on XcX_{c};

  2. (2)

    ds2ds^{2} admits a (,1)(\infty,1) bounded potential locally on XcX_{c}.

Proof.

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=1rfi2)+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 φ\varphi on XX so that

(2.21) 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 VV of X\XoX\backslash X^{o}.

Denoting u=logi=1rfi2u=-\log\sum_{i=1}^{r}\|f_{i}\|^{2}, we assume that u>eu>e on VV by a possible shrinking of VV. Then

(2.22) 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

|φ|+|dφ|1¯φ1logu<1.|\varphi|+|d\varphi|_{\sqrt{-1}\partial\bar{\partial}\varphi}\lesssim\frac{1}{\log u}<1.

Since loglogu\log\log u is a smooth exhausted function such that

|dloglogu|1¯φ2.|d\log\log u|_{\sqrt{-1}\partial\bar{\partial}\varphi}\leq 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 ψ0\psi\geq 0. Then ψc:=ψ+1cψ\psi_{c}:=\psi+\frac{1}{c-\psi} is a smooth exhausted psh 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 hermitian 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 gives a desired complete Kähler metric on XcXoX_{c}\cap X^{o} which we need. ∎

3. Harmonic Representation of the derived pushforwards

Let f:XYf:X\rightarrow Y be a proper locally Kähler morphism between complex spaces and XoXregX^{o}\subset X_{\rm reg} a dense Zariski open subset. Let (E,h)(E,h) be a hermitian vector bundle on XoX^{o} with a Nakano semi-positive curvature. The aim of this section is to introduce a harmonic representation of RqfSX(E,h)R^{q}f_{\ast}S_{X}(E,h). The results of this section are generalizations of Section 4 of [Takegoshi1995] to L2L^{2}-Dolbeault complexes on complex spaces.

3.1. Harmonic forms on weakly 1-complete spaces

Let ZZ be a complex space of pure dimension nn and ZoZregZ^{o}\subset Z_{\rm reg} a dense Zariski open subset. Let ds2ds^{2} be Kähler metric on ZoZ^{o} which is locally complete and admits a (,1)(\infty,1) bounded potential locally on XX (Definition 2.12). Denote ω\omega to be the associated Kähler form. Let (E,h)(E,h) be a hermitian vector bundle on ZoZ^{o}. Let Ap,q(Zo,E)A^{p,q}(Z^{o},E) (resp. Ak(Zo,E)A^{k}(Z^{o},E)) be the space of CC^{\infty} EE-valued (p,q)(p,q) (resp. kk) forms on ZoZ^{o} and let Acptp,q(Zo,E)Ap,q(Zo,E)A^{p,q}_{\rm cpt}(Z^{o},E)\subset A^{p,q}(Z^{o},E) be the subspace of forms with compact support in ZoZ^{o}. Let :Ap,q(Zo,E)Anq,np(Zo,E)\ast:A^{p,q}(Z^{o},E)\to A^{n-q,n-p}(Z^{o},E) be the Hodge star operator relative to ds2ds^{2} and let E:Ap,q(Zo,E)Aq,p(Zo,E)\sharp_{E}:A^{p,q}(Z^{o},E)\to A^{q,p}(Z^{o},E^{\ast}) be the anti-isomorphism defined by hh. Denote by ,h\langle-,-\rangle_{h} the pointwise inner product on Ap,q(Zo,E)A^{p,q}(Z^{o},E). These operators are related by

(3.1) α,βhvolds2=αEβ.\displaystyle\langle\alpha,\beta\rangle_{h}{\rm vol}_{ds^{2}}=\alpha\wedge\ast\sharp_{E}\beta.

Denote

(3.2) (α,β)h:=Zoα,βhvolds2\displaystyle(\alpha,\beta)_{h}:=\int_{Z^{o}}\langle\alpha,\beta\rangle_{h}{\rm vol}_{ds^{2}}

and αh:=(α,α)h\|\alpha\|_{h}:=\sqrt{(\alpha,\alpha)_{h}}. Let =D+¯\nabla=D^{\prime}+\bar{\partial} be the Chern connection associated to hh. Let ¯h=D\bar{\partial}^{\ast}_{h}=-\ast D^{\prime}\ast and Dh=¯D^{\prime\ast}_{h}=-\ast\bar{\partial}\ast be the formal adjoints of ¯\bar{\partial} and DD^{\prime} respectively.

For two operators SS and TT acting on Ar(Z0,E)A^{r}(Z^{0},E) with degree aa and bb respectively, we define the graded Lie bracket [S,T]:=ST(1)abTS[S,T]:=S\circ T-(-1)^{ab}T\circ S.

Denote by LL the Lefschetz operator with respect to ds2ds^{2} and by Λ\Lambda the formal adjoint of LL. Then we have the following Kähler identities ([Wells1980], Chapter V):

(3.3) Dh=1[¯,Λ],\displaystyle D_{h}^{{}^{\prime}\ast}=-\sqrt{-1}[\bar{\partial},\Lambda],
(3.4) ¯h=1[D,Λ]and\displaystyle\bar{\partial}_{h}^{\ast}=\sqrt{-1}[D^{\prime},\Lambda]\quad and
(3.5) e(θ)=1[e(θ¯),Λ]\displaystyle e(\theta)^{\ast}=\sqrt{-1}[e(\bar{\theta}),\Lambda]

for θA1,0(Zo,E)\theta\in A^{1,0}(Z^{o},E).

Denote Δ¯=¯¯h+¯h¯\Delta_{\bar{\partial}}=\bar{\partial}\bar{\partial}^{\ast}_{h}+\bar{\partial}^{\ast}_{h}\bar{\partial} and ΔD=DDh+DhD\Delta_{D^{\prime}}=D^{\prime}D^{\prime\ast}_{h}+D^{\prime\ast}_{h}D^{\prime}. Since e(Θh)=[¯,D]e(\Theta_{h})=[\bar{\partial},D^{\prime}], by (3.3), (3.4) and Jacobi’s identity

[[S,T],U]+(1)a(b+c)[[T,U],S]+(1)c(a+b)[[U,S],T]0[[S,T],U]+(-1)^{a(b+c)}[[T,U],S]+(-1)^{c(a+b)}[[U,S],T]\equiv 0

where a=degSa=\deg S, b=degTb=\deg T and c=degUc=\deg U respectively, we obtain that the formula

(3.6) Δ¯=ΔD+1[e(Θh),Λ]\displaystyle\Delta_{\bar{\partial}}=\Delta_{D^{\prime}}+\sqrt{-1}[e(\Theta_{h}),\Lambda]

holds on ZoZ^{o} where e(α)(β):=αβe(\alpha)(\beta):=\alpha\wedge\beta.

Let φ:Zo\varphi:Z^{o}\to\mathbb{R} be a CC^{\infty} function and let hφ:=eφhh_{\varphi}:=e^{-\varphi}h. Since Dhφ=De(φ)D^{\prime}_{h_{\varphi}}=D^{\prime}-e(\partial\varphi), we obtain the formulae:

(3.7) ¯hφ=¯h+e(¯φ)\displaystyle\bar{\partial}_{h_{\varphi}}^{\ast}=\bar{\partial}_{h}^{\ast}+e(\bar{\partial}\varphi)^{\ast}

and

(3.8) Δ¯hφ=ΔDhφ+1[e(Θ(h)+¯φ),Λ]\displaystyle\Delta_{\bar{\partial}_{h_{\varphi}}}=\Delta_{D^{\prime}_{h_{\varphi}}}+\sqrt{-1}[e(\Theta(h)+\partial\bar{\partial}\varphi),\Lambda]

where Δ¯hφ:=[¯,¯hφ]\Delta_{\bar{\partial}_{h_{\varphi}}}:=[\bar{\partial},\bar{\partial}_{h_{\varphi}}^{\ast}] and ΔDhφ:=[Dhφ,Dhφ]\Delta_{D^{\prime}_{h_{\varphi}}}:=[D^{\prime}_{h_{\varphi}},D^{\prime\ast}_{h_{\varphi}}].

By (3.3), (3.4), (3.5) and Jacobi’s identity, the Donnelly and Xavier’s formula ([Takegoshi1995, (1.9), (1.10)]) can be stated as follows.

(3.9) [¯,e(¯φ)]+[Dh,e(φ)]=1[e(¯φ),Λ]\displaystyle[\bar{\partial},e(\bar{\partial}\varphi)^{\ast}]+[D^{\prime\ast}_{h},e(\partial\varphi)]=\sqrt{-1}[e(\partial\bar{\partial}\varphi),\Lambda]

and

(3.10) [D,e(φ)]+[¯h,e(¯φ)]=1[e(¯φ),Λ)].\displaystyle[D^{\prime},e(\partial\varphi)^{\ast}]+[\bar{\partial}^{\ast}_{h},e(\bar{\partial}\varphi)]=\sqrt{-1}[e(\partial\bar{\partial}\varphi),\Lambda)].

Fix a Whitney stratification of ZZ so that ZoZ^{o} is the union of open strata. By Sard’s theorem there is a subset Σ\Sigma\subset\mathbb{R} of measure zero so that for every c\Σc\in\mathbb{R}\backslash\Sigma and every stratum SS of ZZ, cc is a regular value of φ|S\varphi|_{S}. Any c\Σc\in\mathbb{R}\backslash\Sigma is called a regular value of φ\varphi. For any regular value cc of φ\varphi, {φ=c}\{\varphi=c\} is a piecewise smooth submanifold of ZZ. In particular, {φc}Zo\{\varphi\leq c\}\cap Z^{o} is a submanifold of ZoZ^{o} with a smooth boundary {φ=c}Zo\{\varphi=c\}\cap Z^{o}.

Denote

(α,β)c,h={φc}Zoα,βhvolds2,[α,β]c,h:={φ=c}Zoα,βhvol{φ=c}(\alpha,\beta)_{c,h}=\int_{\{\varphi\leq c\}\cap Z^{o}}\langle\alpha,\beta\rangle_{h}{\rm vol}_{ds^{2}},\quad[\alpha,\beta]_{c,h}:=\int_{\{\varphi=c\}\cap Z^{o}}\langle\alpha,\beta\rangle_{h}{\rm vol}_{\{\varphi=c\}}

and αc,h:=(α,α)c,h\|\alpha\|_{c,h}:=\sqrt{(\alpha,\alpha)_{c,h}} for any regular value cc of φ\varphi. By Stokes’s theorem we acquire that ([Takegoshi1995, (1.1)])

(3.11) (¯α,β)c,h=(α,¯hβ)c,h+[α,e(¯φ)β]c,h,\displaystyle(\bar{\partial}\alpha,\beta)_{c,h}=(\alpha,\bar{\partial}^{\ast}_{h}\beta)_{c,h}+[\alpha,e(\bar{\partial}\varphi)^{\ast}\beta]_{c,h},

and

(3.12) (Dα,β)c,h=(α,Dhβ)c,h+[α,e(φ)β]c,h\displaystyle(D^{\prime}\alpha,\beta)_{c,h}=(\alpha,D^{\prime\ast}_{h}\beta)_{c,h}+[\alpha,e(\partial\varphi)^{\ast}\beta]_{c,h}

hold for every CC^{\infty} forms α\alpha and β\beta on {φc}Zo\{\varphi\leq c\}\cap Z^{o} such that either α\alpha or β\beta has compact support in {φc}Zo\{\varphi\leq c\}\cap Z^{o}.

Proposition 3.1.

Let (M,ωM)(M,\omega_{M}) be a complete Kähler manifold of dimension nn and let (E,h)(E,h) be a Nakano semi-positive holomorphic vector bundle on MM. If q1q\geq 1 and α𝒟n,q(M,E;ωM,h)KerΔ¯\alpha\in\mathscr{D}^{n,q}(M,E;\omega_{M},h)\cap{\rm Ker}\Delta_{\bar{\partial}}, then α\alpha satisfies the following equations:

  1. (1)

    ¯α=0\bar{\partial}\alpha=0, ¯hα=0\bar{\partial}^{\ast}_{h}\alpha=0, Dhα=0D^{\prime\ast}_{h}\alpha=0 and e(Θh)Λα,αh=0\langle e(\Theta_{h})\Lambda\alpha,\alpha\rangle_{h}=0 on MM. In particular ¯(α)=0\bar{\partial}(\ast\alpha)=0, i.e. αΓ(M,ΩMnq(E))\ast\alpha\in\Gamma(M,\Omega^{n-q}_{M}(E)).

  2. (2)

    e(¯φ)α=0e(\bar{\partial}\varphi)^{\ast}\alpha=0 and e(¯φ)Λα,αh=0\langle e(\partial\bar{\partial}\varphi)\Lambda\alpha,\alpha\rangle_{h}=0 on MM for any CC^{\infty} psh function φ\varphi on MM with

    supxM{|φ(x)|+|dφ(x)|ωM}<.{\rm sup}_{x\in M}\{|\varphi(x)|+|d\varphi(x)|_{\omega_{M}}\}<\infty.
Proof.

See [Takegoshi1995, Theorem 3.4]. ∎

Denote by 𝒫e(Z)\mathscr{P}_{e}(Z) the set of CC^{\infty} psh functions φ:Z(,c)\varphi:Z\to(-\infty,c_{\ast}) for some c(,]c_{\ast}\in(-\infty,\infty] such that Xc:={zZ|φ(z)<c}X_{c}:=\{z\in Z|\varphi(z)<c\} is precompact in ZZ for every c<cc<c_{\ast}. ZZ is called a weakly 1-complete space if 𝒫e(Z)\mathscr{P}_{e}(Z)\neq\emptyset. For every φ𝒫e(Z)\varphi\in\mathscr{P}_{e}(Z), denote

(3.13) p,q(Z,E,h,φ):={α𝒟Z,ds2p,q(E,h)(Z)|¯α=¯hα=0,e(¯φ)α=0}.\displaystyle\mathscr{H}^{p,q}(Z,E,h,\varphi):=\left\{\alpha\in\mathscr{D}^{p,q}_{Z,ds^{2}}(E,h)(Z)\big{|}\bar{\partial}\alpha=\bar{\partial}^{\ast}_{h}\alpha=0,e(\bar{\partial}\varphi)^{\ast}\alpha=0\right\}.

By the regularity theorem for elliptic operators of second order, every element of p,q(Z,E,h,φ)\mathscr{H}^{p,q}(Z,E,h,\varphi) is CC^{\infty} on ZoZ^{o}.

Proposition 3.2.

Let ZZ be a weakly 1-complete space of pure dimension nn and ZoZregZ^{o}\subset Z_{\rm reg} a dense Zariski open subset. Let ds2ds^{2} be a Kähler metric on ZoZ^{o} which is locally complete on ZZ and let (E,h)(E,h) be a hermitian vector bundle on ZoZ^{o} with a Nakano semi-positive curvature. Then the following assertions hold.

  1. (1)

    Let φ𝒫e(Z)\varphi\in\mathscr{P}_{e}(Z). Assume that α𝒟Z,ds2n,q(E,h)(Z)\alpha\in\mathscr{D}^{n,q}_{Z,ds^{2}}(E,h)(Z) satisfies e(¯φ)α=0e(\bar{\partial}\varphi)^{\ast}\alpha=0. Then ¯α=¯hα=0\bar{\partial}\alpha=\bar{\partial}^{\ast}_{h}\alpha=0 if and only if Dhα=1e(Θh+¯φ)Λα,αh=0D_{h}^{\prime\ast}\alpha=\langle\sqrt{-1}e(\Theta_{h}+\partial\bar{\partial}\varphi)\Lambda\alpha,\alpha\rangle_{h}=0. Here 1e(Θh+¯φ)Λα,αh=0\langle\sqrt{-1}e(\Theta_{h}+\partial\bar{\partial}\varphi)\Lambda\alpha,\alpha\rangle_{h}=0 is equivalent to 1(¯φ)Λα,αh=1(Θh)Λα,αh=0\langle\sqrt{-1}(\partial\bar{\partial}\varphi)\Lambda\alpha,\alpha\rangle_{h}=\langle\sqrt{-1}(\Theta_{h})\Lambda\alpha,\alpha\rangle_{h}=0.

  2. (2)

    Let φ,ψ𝒫e(Z)\varphi,\psi\in\mathscr{P}_{e}(Z). Then n,q(Z,E,h,φ)=n,q(Z,E,h,ψ)\mathscr{H}^{n,q}(Z,E,h,\varphi)=\mathscr{H}^{n,q}(Z,E,h,\psi) for every q0q\geq 0.

  3. (3)

    For every q0q\geq 0 and every φ𝒫e(Z)\varphi\in\mathscr{P}_{e}(Z), the Hodge star operator gives a well defined map

    (3.14) :n,q(Z,E,h,φ)Ker¯(𝒟Z,ds2nq,0(E,h)(Z)𝒟Z,ds2nq,1(E,h)(Z)).\displaystyle\ast:\mathscr{H}^{n,q}(Z,E,h,\varphi)\to{\rm Ker}\bar{\partial}\left(\mathscr{D}^{n-q,0}_{Z,ds^{2}}(E,h)(Z)\to\mathscr{D}^{n-q,1}_{Z,ds^{2}}(E,h)(Z)\right).
Proof.

To show (1), we suppose that e(¯φ)α=0e(\bar{\partial}\varphi)^{\ast}\alpha=0 and ¯α=¯hα=0\bar{\partial}\alpha=\bar{\partial}^{\ast}_{h}\alpha=0. Take any regular value cc of φ\varphi. Since ({φc}Zo,ds2)(\{\varphi\leq c\}\cap Z^{o},ds^{2}) is complete, the Hopf-Rinow theorem implies that there exists an exhaustive sequence (Kν)(K_{\nu}) of compact sets of Zo{φc}Z^{o}\cap\{\varphi\leq c\} and functions θνC({φc}Zo,)\theta_{\nu}\in C^{\infty}(\{\varphi\leq c\}\cap Z^{o},\mathbb{R}) such that

θν=1on a neighbourhood of Kν,Supp θνKν+1,\theta_{\nu}=1\quad\textrm{on a neighbourhood of $K_{\nu}$,}\quad\textrm{Supp $\theta_{\nu}\subset K^{\circ}_{\nu+1}$},
0θνθν+11and|dθν|g2ν,ν.0\leq\theta_{\nu}\leq\theta_{\nu+1}\leq 1\quad\textrm{and}\quad|d\theta_{\nu}|_{g}\leq 2^{-\nu},\quad\forall\nu.

Then θνα\theta_{\nu}\alpha converges to α\alpha under the norm +¯+D+¯h+Dh\|-\|+\|\bar{\partial}-\|+\|D^{\prime}-\|+\|\bar{\partial}^{\ast}_{h}-\|+\|D^{\prime\ast}_{h}-\|.

Let hφ=eφhh_{\varphi}=e^{-\varphi}h. Since ¯α=¯hα=0\bar{\partial}\alpha=\bar{\partial}^{\ast}_{h}\alpha=0, by (3.6) we obtain that

(3.15) (DDhφα,θνα)c,hφ+(1e(Θhφ)Λα,θνα)c,hφ=0.\displaystyle(D^{\prime}D^{\prime\ast}_{h_{\varphi}}\alpha,\theta_{\nu}\alpha)_{c,h_{\varphi}}+(\sqrt{-1}e(\Theta_{h_{\varphi}})\Lambda\alpha,\theta_{\nu}\alpha)_{c,h_{\varphi}}=0.

By (3.12), this is equivalent to

(3.16) (Dhφα,Dhφ(θνα))c,hφ+(1e(Θhφ)Λα,θνα)c,hφ+[e(φ)Dhφα,θνα]c,hφ=0.\displaystyle(D^{\prime\ast}_{h_{\varphi}}\alpha,D^{\prime\ast}_{h_{\varphi}}(\theta_{\nu}\alpha))_{c,h_{\varphi}}+(\sqrt{-1}e(\Theta_{h_{\varphi}})\Lambda\alpha,\theta_{\nu}\alpha)_{c,h_{\varphi}}+[e(\partial\varphi)D^{\prime\ast}_{h_{\varphi}}\alpha,\theta_{\nu}\alpha]_{c,h_{\varphi}}=0.

By the assumptions and (3.9) we have

(3.17) e(φ)Dhφα=1e(¯φ)Λα.\displaystyle e(\partial\varphi)D^{\prime\ast}_{h_{\varphi}}\alpha=\sqrt{-1}e(\partial\bar{\partial}\varphi)\Lambda\alpha.

Substituting (3.17) into (3.16), we get that

(3.18) (Dhφα,Dhφ(θνα))c,hφ+(1e(Θhφ)Λα,θνα)c,hφ+[1e(¯φ)Λα,θνα]c,hφ=0.\displaystyle(D^{\prime\ast}_{h_{\varphi}}\alpha,D^{\prime\ast}_{h_{\varphi}}(\theta_{\nu}\alpha))_{c,h_{\varphi}}+(\sqrt{-1}e(\Theta_{h_{\varphi}})\Lambda\alpha,\theta_{\nu}\alpha)_{c,h_{\varphi}}+[\sqrt{-1}e(\partial\bar{\partial}\varphi)\Lambda\alpha,\theta_{\nu}\alpha]_{c,h_{\varphi}}=0.

Letting ν\nu\to\infty, we obtain that

(3.19) Dhφαc,hφ2+(1e(Θhφ)Λα,α)c,hφ+[1e(¯φ)Λα,α]c,hφ=0.\displaystyle\|D^{\prime\ast}_{h_{\varphi}}\alpha\|^{2}_{c,h_{\varphi}}+(\sqrt{-1}e(\Theta_{h_{\varphi}})\Lambda\alpha,\alpha)_{c,h_{\varphi}}+[\sqrt{-1}e(\partial\bar{\partial}\varphi)\Lambda\alpha,\alpha]_{c,h_{\varphi}}=0.

Since φ\varphi is a psh function and 1Θh\sqrt{-1}\Theta_{h} is Nakano semi-positive, 1Θhφ=1Θh+1¯φ\sqrt{-1}\Theta_{h_{\varphi}}=\sqrt{-1}\Theta_{h}+\sqrt{-1}\partial\bar{\partial}\varphi is also Nakano semi-positive. Then all the three terms in (3.19) are semi-positive. Hence they are all zero for every regular value cc of φ\varphi. This proves the necessity of (1).

To prove the sufficiency of (1), we assume that Dhα=1e(Θh+¯φ)Λα,αh=0D_{h}^{\prime\ast}\alpha=\sqrt{-1}\langle e(\Theta_{h}+\partial\bar{\partial}\varphi)\Lambda\alpha,\alpha\rangle_{h}=0 and e(¯φ)α=0e(\bar{\partial}\varphi)^{\ast}\alpha=0. By (3.6) we have Δ¯α=0\Delta_{\bar{\partial}}\alpha=0. By (3.11) we obtain

(3.20) (Δ¯α,θνα)c,h\displaystyle(\Delta_{\bar{\partial}}\alpha,\theta_{\nu}\alpha)_{c,h}
=\displaystyle= (¯hα,¯h(θνα))c,h+(¯α,¯(θνα))c,h+[¯hα,e(¯φ)(θνα)]c,h[e(¯φ)¯α,θνα]c,h\displaystyle(\bar{\partial}^{\ast}_{h}\alpha,\bar{\partial}^{\ast}_{h}(\theta_{\nu}\alpha))_{c,h}+(\bar{\partial}\alpha,\bar{\partial}(\theta_{\nu}\alpha))_{c,h}+[\bar{\partial}^{\ast}_{h}\alpha,e(\bar{\partial}\varphi)^{\ast}(\theta_{\nu}\alpha)]_{c,h}-[e(\bar{\partial}\varphi)^{\ast}\bar{\partial}\alpha,\theta_{\nu}\alpha]_{c,h}
=\displaystyle= 0.\displaystyle 0.

By (3.9) we acquire that

(3.21) 1e(¯φ)¯α,θναh=1e(¯φ)Λα,θναh=θν1e(¯φ)Λα,αh=0.\displaystyle\langle\sqrt{-1}e(\bar{\partial}\varphi)\bar{\partial}\alpha,\theta_{\nu}\alpha\rangle_{h}=\langle\sqrt{-1}e(\partial\bar{\partial}\varphi)\Lambda\alpha,\theta_{\nu}\alpha\rangle_{h}=\theta_{\nu}\sqrt{-1}\langle e(\partial\bar{\partial}\varphi)\Lambda\alpha,\alpha\rangle_{h}=0.

Combining (3.20), (3.21) and e(¯φ)(θνα)=θνe(¯φ)α=0e(\bar{\partial}\varphi)^{\ast}(\theta_{\nu}\alpha)=\theta_{\nu}e(\bar{\partial}\varphi)^{\ast}\alpha=0, we obtain that

(3.22) (¯hα,¯h(θνα))c,h+(¯α,¯(θνα))c,h=0.\displaystyle(\bar{\partial}^{\ast}_{h}\alpha,\bar{\partial}^{\ast}_{h}(\theta_{\nu}\alpha))_{c,h}+(\bar{\partial}\alpha,\bar{\partial}(\theta_{\nu}\alpha))_{c,h}=0.

Taking its limit we know that

(3.23) ¯hαc,h2+¯αc,h2=0\displaystyle\|\bar{\partial}^{\ast}_{h}\alpha\|^{2}_{c,h}+\|\bar{\partial}\alpha\|^{2}_{c,h}=0

for every regular value cc of φ\varphi. This implies that ¯α=¯hα=0\bar{\partial}\alpha=\bar{\partial}^{\ast}_{h}\alpha=0.

To prove (2) we set hψ=eψhh_{-\psi}=e^{\psi}h. By (3.9) we obtain that

(3.24) ¯e(¯ψ)(θνα)+e(¯ψ)¯(θνα)=1e(¯ψ)Λ(θνα)\displaystyle\bar{\partial}e(\bar{\partial}\psi)^{\ast}(\theta_{\nu}\alpha)+e(\bar{\partial}\psi)^{\ast}\bar{\partial}(\theta_{\nu}\alpha)=\sqrt{-1}e(\partial\bar{\partial}\psi)\Lambda(\theta_{\nu}\alpha)

if αn,q(Z,E,h,φ)\alpha\in\mathscr{H}^{n,q}(Z,E,h,\varphi). By (3.7) and (3.11), we get that

(3.25) (¯e(¯ψ)(θνα),α)c,hψ\displaystyle(\bar{\partial}e(\bar{\partial}\psi)^{\ast}(\theta_{\nu}\alpha),\alpha)_{c,h_{-\psi}} =(e(¯ψ)(θνα),¯hψα)c,hψ+[e(¯ψ)(θνα),e(¯φ)α]c,hψ\displaystyle=(e(\bar{\partial}\psi)^{\ast}(\theta_{\nu}\alpha),\bar{\partial}^{\ast}_{h_{-\psi}}\alpha)_{c,h_{-\psi}}+[e(\bar{\partial}\psi)^{\ast}(\theta_{\nu}\alpha),e(\bar{\partial}\varphi)^{\ast}\alpha]_{c,h_{-\psi}}
=(e(¯ψ)(θνα),(¯he(¯ψ))α)c,hψ\displaystyle=(e(\bar{\partial}\psi)^{\ast}(\theta_{\nu}\alpha),(\bar{\partial}^{\ast}_{h}-e(\bar{\partial}\psi)^{\ast})\alpha)_{c,h_{-\psi}}
=(e(¯ψ)(θνα),e(¯ψ)α)c,hψ.\displaystyle=-(e(\bar{\partial}\psi)^{\ast}(\theta_{\nu}\alpha),e(\bar{\partial}\psi)^{\ast}\alpha)_{c,h_{-\psi}}.

Taking ν\nu\to\infty on (3.24), we know that

(3.26) (1e(¯ψ)Λ(α),α)c,hψ+e(¯ψ)αc,hψ2=0\displaystyle(\sqrt{-1}e(\partial\bar{\partial}\psi)\Lambda(\alpha),\alpha)_{c,h_{-\psi}}+\|e(\bar{\partial}\psi)^{\ast}\alpha\|^{2}_{c,h_{-\psi}}=0

for every regular value cc of φ\varphi. Since both terms are semi-positive, we show that e(¯ψ)α=0e(\bar{\partial}\psi)^{\ast}\alpha=0. Hence αn,q(Z,E,h,ψ)\alpha\in\mathscr{H}^{n,q}(Z,E,h,\psi).

It remains to show (3). Since \ast is a bounded operator, it suffices to show that ¯α=0\bar{\partial}\ast\alpha=0 for every αn,q(Z,E,h,φ)\alpha\in\mathscr{H}^{n,q}(Z,E,h,\varphi). This follows from ¯α=Dhα=0-\ast\bar{\partial}\ast\alpha=D^{\prime\ast}_{h}\alpha=0 which is proved in (1). ∎

By Proposition 3.2-(2), n,q(Z,E,h,φ)\mathscr{H}^{n,q}(Z,E,h,\varphi) is independent of the choice of φ𝒫e(Z)\varphi\in\mathscr{P}_{e}(Z). Hence we simply denote n,q(Z,E,h):=n,q(Z,E,h,φ)\mathscr{H}^{n,q}(Z,E,h):=\mathscr{H}^{n,q}(Z,E,h,\varphi) for every ZZ with 𝒫e(Z)\mathscr{P}_{e}(Z)\neq\emptyset.

3.2. Harmonic Representation

Let us return to the relative setting. Let f:XYf:X\rightarrow Y be a proper morphism from a complex space XX to an irreducible complex space YY. Denote n:=dimXn:={\rm dim}X and m:=dimYm:={\rm dim}Y respectively. Assume that every irreducible component of XX is mapped onto YY. Let XoXregX^{o}\subset X_{\rm reg} be a dense Zariski open subset and (E,h)(E,h) a hermitian vector bundle on XoX^{o} with a Nakano semi-positive curvature. Assume that there is a Kähler metric ds2ds^{2} on XoX^{o} such that

  1. (1)

    ds2ds^{2} admits a (,1)(\infty,1) bounded potential locally on XX;

  2. (2)

    ds2ds^{2} is locally complete on XX and is locally bounded from below by a hermitian metric.

By Lemma 2.13, such kind of metric exists locally near every fiber of ff and globally on XoX^{o} when XX is a compact Kähler space. By Theorem 2.11 and Lemma 2.2, there is a resolution by fine sheaves

(3.27) SX(E,h)𝒟X,ds2n,(E,h).\displaystyle S_{X}(E,h)\to\mathscr{D}^{n,\bullet}_{X,ds^{2}}(E,h).

Denote X(T):=f1TX(T):=f^{-1}T for every subset TYT\subset Y. Denote by LL the Lefschetz operator with respect to ds2ds^{2} and denote by Λ\Lambda the formal adjoint of LL.

Proposition 3.3.

Notations as above. Let SYS\subset Y be a Stein open subset. Let SoSregS^{o}\subset S_{\rm reg} and X(S)oX(S)regX(So)X(S)^{o}\subset X(S)_{\rm reg}\cap X(S^{o}) be dense Zariski open subsets so that f:X(S)oSof:X(S)^{o}\to S^{o} is a submersion. Then the Hodge star operator

(3.28) :n,q(X(S),E,h){α𝒟X(S),ds2nq,0(E,h)(X(S))|¯α=0,α|X(S)oΓ(X(S)o,ΩXonmqfΩSom)}\displaystyle\ast:\mathscr{H}^{n,q}(X(S),E,h)\to\left\{\alpha\in\mathscr{D}^{n-q,0}_{X(S),ds^{2}}(E,h)(X(S))\bigg{|}\bar{\partial}\alpha=0,\alpha|_{X(S)^{o}}\in\Gamma(X(S)^{o},\Omega_{X^{o}}^{n-m-q}\otimes f^{\ast}\Omega^{m}_{S^{o}})\right\}

is well defined and injective for every qq. As a consequence,

  1. (1)

    n,q(X(S),E,h)=0\mathscr{H}^{n,q}(X(S),E,h)=0 for every q>nmq>n-m;

  2. (2)

    For every open subset SSS^{\prime}\subset S such that 𝒫e(S)\mathscr{P}_{e}(S^{\prime})\neq\emptyset, the restriction map

    (3.29) n,q(X(S),E,h)n,q(X(S),E,h)\displaystyle\mathscr{H}^{n,q}(X(S),E,h)\to\mathscr{H}^{n,q}(X(S^{\prime}),E,h)

    is well defined.

Proof.

Let αn,q(X(S),E,h)\alpha\in\mathscr{H}^{n,q}(X(S),E,h). By Proposition 3.2-(1), Dhα=0D^{\prime\ast}_{h}\alpha=0, i.e. ¯α=0\bar{\partial}\ast\alpha=0. It remains to show that α|X(S)oΓ(X(S)o,ΩXonmqfΩSom)\ast\alpha|_{X(S)^{o}}\in\Gamma(X(S)^{o},\Omega_{X^{o}}^{n-m-q}\otimes f^{\ast}\Omega^{m}_{S^{o}}).

Fix a closed immersion SNS\subset\mathbb{C}^{N} where z1,,zNz_{1},\dots,z_{N} is the coordinate of N\mathbb{C}^{N}. Let φ=i=1Nzi2𝒫e(S)\varphi=\sum_{i=1}^{N}\|z_{i}\|^{2}\in\mathscr{P}_{e}(S). By Proposition 3.2-(1) and (3.5), one has

(3.30) i=1N|e(f(zi)¯)α|h2=\displaystyle\sum_{i=1}^{N}|e(\overline{\partial f^{\ast}(z_{i})})^{\ast}\alpha|^{2}_{h}= i=1Ne(f(zi)¯)α,1e(f(zi))Λα(3.5)\displaystyle\sum_{i=1}^{N}\langle e(\overline{\partial f^{\ast}(z_{i})})^{\ast}\alpha,\sqrt{-1}e(\partial f^{\ast}(z_{i}))\Lambda\alpha\rangle\quad(\ref{align_theta})
=\displaystyle= i=1N1e(f(zi)¯)e(f(zi))Λα,α\displaystyle\sum_{i=1}^{N}\langle\sqrt{-1}e(\overline{\partial f^{\ast}(z_{i})})e(\partial f^{\ast}(z_{i}))\Lambda\alpha,\alpha\rangle
=\displaystyle= 1e(¯fφ)Λα,αh=0.\displaystyle\langle\sqrt{-1}e(\partial\bar{\partial}f^{\ast}\varphi)\Lambda\alpha,\alpha\rangle_{h}=0.

Hence f(dzi)α=0f^{\ast}(dz_{i})\wedge\ast\alpha=0, i=1,,N\forall i=1,\dots,N. This implies that α|X(S)o=0\alpha|_{X(S)^{o}}=0 if q>nmq>n-m and α|X(V)X(S)o\ast\alpha|_{X(V)\cap X(S)^{o}} can be divided by fθf^{\ast}\theta for any open subset VSoV\subset S^{o} which admits a non-vanishing holomorphic mm-form θ\theta. Therefore

α|X(S)oΓ(X(S)o,ΩXonmqfΩSom).\ast\alpha|_{X(S)^{o}}\in\Gamma(X(S)^{o},\Omega_{X^{o}}^{n-m-q}\otimes f^{\ast}\Omega^{m}_{S^{o}}).

This proves that (3.28) is well defined and n,q(X(S),E,h)=0\mathscr{H}^{n,q}(X(S),E,h)=0 for every q>nmq>n-m. (3.28) is injective because

(3.31) Lq=c(n,q)Id,c(n,q)=1(nq)(nq+3)q!\displaystyle L^{q}\circ\ast=c(n,q){\rm Id},\quad c(n,q)=\sqrt{-1}^{(n-q)(n-q+3)}q!

holds on (n,q)(n,q)-forms ([Wells1980, Theorem 3.16]).

It remains to show that α|X(S)n,q(X(S),E,h)\alpha|_{X(S^{\prime})}\in\mathscr{H}^{n,q}(X(S^{\prime}),E,h), i.e. e(¯fψ)α|X(S)=0e(\bar{\partial}f^{\ast}\psi)^{\ast}\alpha|_{X(S^{\prime})}=0 for some ψ𝒫e(S)\psi\in\mathscr{P}_{e}(S^{\prime}). By (3.31), α=Lqβ\alpha=L^{q}\beta for β=c(n,q)1α\beta=c(n,q)^{-1}\ast\alpha. Choose an arbitrary ψ𝒫e(S)\psi\in\mathscr{P}_{e}(S^{\prime}). Since β|X(S)oΓ(X(S)o,ΩXonmqfΩSom)\beta|_{X(S)^{o}}\in\Gamma(X(S)^{o},\Omega_{X^{o}}^{n-m-q}\otimes f^{\ast}\Omega^{m}_{S^{o}}), e((fψ))Lkβ|X(S)X(S)o=0e(\partial(f^{\ast}\psi))L^{k}\beta|_{X(S^{\prime})\cap X(S)^{o}}=0 for every k=0,,q1k=0,\dots,q-1. Because

(3.32) 1[e(¯(fψ)),L]=e((fψ)), ([Wells1980, Chapter V, (3.22)]),\displaystyle\sqrt{-1}[e(\bar{\partial}(f^{\ast}\psi))^{\ast},L]=e(\partial(f^{\ast}\psi)),\textrm{\quad(\cite[cite]{[\@@bibref{}{Wells1980}{}{}, Chapter V, (3.22)]})},

we obtain that

(3.33) e(¯(fψ))Lqβ|X(S)X(S)o=Lqe(¯(fψ))β|X(S)X(S)o=0.\displaystyle e(\bar{\partial}(f^{\ast}\psi))^{\ast}L^{q}\beta|_{X(S^{\prime})\cap X(S)^{o}}=L^{q}e(\bar{\partial}(f^{\ast}\psi))^{\ast}\beta|_{X(S^{\prime})\cap X(S)^{o}}=0.

Thus e(¯(fψ))α|X(S)=0e(\bar{\partial}(f^{\ast}\psi))^{\ast}\alpha|_{X(S^{\prime})}=0 by continuity. This proves the proposition. ∎

By Proposition 3.3, the restriction map

n,q(X(V),E,h)n,q(X(U),E,h)\mathscr{H}^{n,q}(X(V),E,h)\to\mathscr{H}^{n,q}(X(U),E,h)

is well defined for any pair of Stein open subsets UVYU\subset V\subset Y. Hence the data

Un,q(X(U),E,h),UY is a Stein open subset\displaystyle U\mapsto\mathscr{H}^{n,q}(X(U),E,h),\quad U\subset Y\textrm{ is a Stein open subset}

determines a sheaf fn,q(E,h)\mathscr{H}^{n,q}_{f}(E,h) on YY (after a sheafification).

By (3.27) and Lemma 2.2, there is a natural morphism

(3.34) fn,(E,h)f(𝒟X,ds2n,(E,h))Rf(SX(E,h)).\displaystyle\mathscr{H}^{n,\bullet}_{f}(E,h)\to f_{\ast}(\mathscr{D}_{X,ds^{2}}^{n,\bullet}(E,h))\simeq Rf_{\ast}(S_{X}(E,h)).

This induces a canonical morphism

fn,q(E,h)RqfSX(E,h)\mathscr{H}^{n,q}_{f}(E,h)\to R^{q}f_{\ast}S_{X}(E,h)

for every 0qn0\leq q\leq n. The main result of this section is

Theorem 3.4.

fn,q(E,h)\mathscr{H}^{n,q}_{f}(E,h) is a sheaf of 𝒪Y\mathscr{O}_{Y}-modules for every q0q\geq 0. Assume that SX(E,h)S_{X}(E,h) is a coherent sheaf, then the canonical morphism

fn,q(E,h)RqfSX(E,h)\mathscr{H}^{n,q}_{f}(E,h)\to R^{q}f_{\ast}S_{X}(E,h)

is an isomorphism of 𝒪Y\mathscr{O}_{Y}-modules for every q0q\geq 0. Moreover,

fn,q(E,h)(U)=n,q(f1(U),E,h)\mathscr{H}^{n,q}_{f}(E,h)(U)=\mathscr{H}^{n,q}(f^{-1}(U),E,h)

for every Stein open subset UYU\subset Y.

Proof.

Let SYS\subset Y be a Stein open subset and φ𝒫e(X(S))\varphi\in\mathscr{P}_{e}(X(S)). For every αn,q(X(S),E,h)\alpha\in\mathscr{H}^{n,q}(X(S),E,h) and every g𝒪Y(S)g\in\mathscr{O}_{Y}(S), denote g:=fgg^{\prime}:=f^{\ast}g. Then Proposition 3.2-(1) implies that

Dh(gα)=¯(gα)=gDhα=0D^{\prime\ast}_{h}(g^{\prime}\alpha)=-\ast\bar{\partial}\ast(g^{\prime}\alpha)=g^{\prime}D^{\prime\ast}_{h}\alpha=0

and

1e(Θh+¯φ)Λ(gα),gαh=g21e(Θh+¯φ)Λ(α),αh=0.\langle\sqrt{-1}e(\Theta_{h}+\partial\bar{\partial}\varphi)\Lambda(g^{\prime}\alpha),g^{\prime}\alpha\rangle_{h}=\|g^{\prime}\|^{2}\langle\sqrt{-1}e(\Theta_{h}+\partial\bar{\partial}\varphi)\Lambda(\alpha),\alpha\rangle_{h}=0.

Hence gαn,q(X(S),E,h,φ)g^{\prime}\alpha\in\mathscr{H}^{n,q}(X(S),E,h,\varphi) by Proposition 3.2-(1). This shows that fn,q(E,h)\mathscr{H}^{n,q}_{f}(E,h) is a sheaf of 𝒪Y\mathscr{O}_{Y}-modules for every 0qn0\leq q\leq n.

Since SX(E,h)S_{X}(E,h) is a coherent sheaf, to prove the remaining claims, it suffices to show that the natural morphism

(3.35) τUq:n,q(X(U),E,h)HqΓ(X(U),𝒟X,ds2n,(E,h))\displaystyle\tau^{q}_{U}:\mathscr{H}^{n,q}(X(U),E,h)\to H^{q}\Gamma(X(U),\mathscr{D}^{n,\bullet}_{X,ds^{2}}(E,h))

is an isomorphism for every Stein open subset UYU\subset Y and every q0q\geq 0. Fix a CC^{\infty} exhausted strictly psh function φU\varphi_{U} on UU. Denote Uc:={φU<c}U_{c}:=\{\varphi_{U}<c\} and φ:=fφU\varphi:=f^{\ast}\varphi_{U}.

Claim 1: τUq\tau^{q}_{U} is injective. Assume that αn,q(X(U),E,h)\alpha\in\mathscr{H}^{n,q}(X(U),E,h) and α=¯β\alpha=\bar{\partial}\beta for some β𝒟X,ds2n,q1(E,h)(X(U))\beta\in\mathscr{D}^{n,q-1}_{X,ds^{2}}(E,h)(X(U)). By (3.11), we obtain that

(α,α)c,h=(β,¯hα)c,h=0(\alpha,\alpha)_{c,h}=(\beta,\bar{\partial}^{\ast}_{h}\alpha)_{c,h}=0

for every regular value cc of φ\varphi. Hence α=0\alpha=0.

Claim 2: τUq\tau^{q}_{U} is surjective. Let αΓ(X(U),𝒟X,ds2n,q(E,h))\alpha\in\Gamma(X(U),\mathscr{D}^{n,q}_{X,ds^{2}}(E,h)) be ¯\bar{\partial}-closed.

Step 1: In this step we show that for every cc\in\mathbb{R}, α|X(Uc)=uc+¯βc\alpha|_{X(U_{c})}=u_{c}+\bar{\partial}\beta_{c} for some ucn,q(X(Uc),E,h)u_{c}\in\mathscr{H}^{n,q}(X(U_{c}),E,h) and βcΓ(X(Uc),𝒟X,ds2n,q1(E,h))\beta_{c}\in\Gamma(X(U_{c}),\mathscr{D}^{n,q-1}_{X,ds^{2}}(E,h)).

Fix 0<c0<c10<c_{0}<c_{1}. Denote ωds2\omega_{ds^{2}} to be the Kähler form associated to ds2ds^{2} and let ωλ:=ωds2+1¯λ(φc)\omega_{\lambda}:=\omega_{ds^{2}}+\sqrt{-1}\partial\bar{\partial}\lambda(\varphi-c) for some smooth convex function λ:0\lambda:\mathbb{R}\to\mathbb{R}_{\geq 0} such that λ(t)=0\lambda(t)=0 if tc0t\leq c_{0}, λ(t)>0\lambda(t)>0, λ(t)>0\lambda^{\prime}(t)>0, λ′′(t)>0\lambda^{\prime\prime}(t)>0 if t>c0t>c_{0} and 0c1λ′′(t)𝑑t=+\int_{0}^{c_{1}}\sqrt{\lambda^{\prime\prime}(t)}dt=+\infty. Let hλ:=eλ(φ)hh_{\lambda}:=e^{-\lambda(\varphi)}h. Then ωλωds2\omega_{\lambda}\geq\omega_{ds^{2}} is a complete Kähler metric on X(Uc+c1)X(U_{c+c_{1}}) and (E,hλ)(E,h_{\lambda}) is Nakano semi-positive. By choosing λ\lambda sufficiently large we assume that αL(2)n,q(X(Uc+c1),E;ωλ,hλ)\alpha\in L^{n,q}_{(2)}(X(U_{c+c_{1}}),E;\omega_{\lambda},h_{\lambda}). Noting that ¯α=0\bar{\partial}\alpha=0, there is a unique decomposition α=vc+γc\alpha=v_{c}+\gamma_{c} so that vcv_{c} is a harmonic form on Uc+c1U_{c+c_{1}} with respect to ωλ\omega_{\lambda} and hλh_{\lambda} while γc\gamma_{c} lies in the closure of the range of ¯\bar{\partial} in the Hilbert space L(2)n,q(X(Uc+c1),E;ωλ,hλ)L^{n,q}_{(2)}(X(U_{c+c_{1}}),E;\omega_{\lambda},h_{\lambda}). Since ωλ|Uc=ωds2|Uc\omega_{\lambda}|_{U_{c}}=\omega_{ds^{2}}|_{U_{c}} and hλ|Uc=h|Uch_{\lambda}|_{U_{c}}=h|_{U_{c}}, by Proposition 3.1 and Proposition 3.5 below, we see that vc|Ucn,q(X(Uc),E,h)v_{c}|_{U_{c}}\in\mathscr{H}^{n,q}(X(U_{c}),E,h) and γc|Uc=¯βc\gamma_{c}|_{U_{c}}=\bar{\partial}\beta_{c} for some βcΓ(X(Uc),𝒟X,ds2n,q1(E,h))\beta_{c}\in\Gamma(X(U_{c}),\mathscr{D}^{n,q-1}_{X,ds^{2}}(E,h)).

Step 2: By step 1, there is a decomposition

α|X(Uk)=uk+¯βk\alpha|_{X(U_{k})}=u_{k}+\bar{\partial}\beta_{k}

where ukn,q(X(Uk),E,h)u_{k}\in\mathscr{H}^{n,q}(X(U_{k}),E,h) and βkΓ(X(Uk),𝒟X,ds2n,q1(E,h))\beta_{k}\in\Gamma(X(U_{k}),\mathscr{D}^{n,q-1}_{X,ds^{2}}(E,h)) for every kk\in\mathbb{N}. Since uk+1u_{k+1} and uku_{k} are cohomologous on X(Uk)X(U_{k}), we have uk+1|X(Uk)=uku_{k+1}|_{X(U_{k})}=u_{k} by Claim 1. Hence there is a unique un,q(X(U),E,h)u\in\mathscr{H}^{n,q}(X(U),E,h) such that u|X(Uk)=uku|_{X(U_{k})}=u_{k}. Hence the cohomology classes [α][\alpha] and [u][u] are equal over every X(Uk)X(U_{k}), kk\in\mathbb{N}. Since UkU_{k} is a Stein space for each k{+}k\in\mathbb{N}\cup\{+\infty\} and SX(E,h)S_{X}(E,h) is a coherent sheaf, there is a canonical isomorphism

Γ(X(Uk),𝒟X,ds2n,q(E,h))Hq(X(Uk),SX(E,h))Γ(Uk,Rqf(SX(E,h)))\Gamma(X(U_{k}),\mathscr{D}^{n,q}_{X,ds^{2}}(E,h))\simeq H^{q}(X(U_{k}),S_{X}(E,h))\simeq\Gamma(U_{k},R^{q}f_{\ast}(S_{X}(E,h)))

for every k{+}k\in\mathbb{N}\cup\{+\infty\}. By regarding [α][\alpha] and [u][u] as sections of Rqf(SX(E,h))R^{q}f_{\ast}(S_{X}(E,h)) (which are equal over every UkU_{k}, kk\in\mathbb{N}) we obtain that [α]=[u][\alpha]=[u]. This proves the surjectivity of τUq\tau^{q}_{U}. ∎

To complete the proof of Theorem 3.4, we go on proving the following proposition.

Proposition 3.5.

Assume that SX(E,h)S_{X}(E,h) is a coherent sheaf on XX. Let VXV\subset X be an open subset and denote Vo:=VXoV^{o}:=V\cap X^{o}. Let

¯:L(2)n,q1(Vo,E;ds2,h)L(2)n,q(Vo,E;ds2,h)\bar{\partial}:L^{n,q-1}_{(2)}(V^{o},E;ds^{2},h)\to L^{n,q}_{(2)}(V^{o},E;ds^{2},h)

be the unbounded operator in the sense of distribution where q1q\geq 1. Suppose that αIm¯¯\alpha\in\overline{{\rm Im}\bar{\partial}}. Then there is an (n,q1)(n,q-1)-form βLV,ds2n,q1(E,h)(V)\beta\in L^{n,q-1}_{V,ds^{2}}(E,h)(V) such that α=¯β\alpha=\bar{\partial}\beta.

Proof.

To prove the proposition, we need the following lemma which is a modified version of Theorem 2.10 and we leave its proof to the end of this section.

Lemma 3.6.

Let YY be a complex manifold of dimension nn which admits a complete Kähler metric. Let ω=1¯φ\omega=\sqrt{-1}\partial\bar{\partial}\varphi be a Kähler metric of YY with sup|φ|<\sup|\varphi|<\infty. Let (E,h)(E,h) be a Nakano semi-positive hermitian vector bundle on YY. Then for every q>0q>0 and every αL(2)n,q(Y,E;ω,h)\alpha\in L^{n,q}_{(2)}(Y,E;\omega,h) such that ¯α=0\bar{\partial}\alpha=0, there is βL(2)n,q1(Y,E;ω,h)\beta\in L^{n,q-1}_{(2)}(Y,E;\omega,h) such that ¯β=α\bar{\partial}\beta=\alpha and βω,h2Cαω,h2\|\beta\|^{2}_{\omega,h}\leq C\|\alpha\|^{2}_{\omega,h}. Here CC is a constant depending on qq and φ\varphi, but not depending on α\alpha.

Proof.

Let h:=heφh^{\prime}:=he^{-\varphi}, then hhh^{\prime}\sim h. Since (E,h)(E,h) is Nakano semi-positive, we know that

1Θh(E)=1¯φIdE+1Θh(E)ωIdE.\sqrt{-1}\Theta_{h^{\prime}}(E)=\sqrt{-1}\partial\bar{\partial}\varphi\otimes{\rm Id}_{E}+\sqrt{-1}\Theta_{h}(E)\geq\omega\otimes{\rm Id}_{E}.

By Theorem 2.10, there exists βL(2)n,q1(Y,E;ω,heφ)\beta\in L_{(2)}^{n,q-1}(Y,E;\omega,he^{-\varphi}) such that ¯β=α\bar{\partial}\beta=\alpha and βω,heφ21qαω,hφ\|\beta\|_{\omega,he^{-\varphi}}^{2}\leq\frac{1}{q}\|\alpha\|_{\omega,h^{-\varphi}}. Since φ\varphi is a bounded function, we have βω,h2Cαω,h2\|\beta\|^{2}_{\omega,h}\leq C\|\alpha\|^{2}_{\omega,h} where CC is a constant depending on qq and φ\varphi and not depending on α\alpha. The lemma is proved. ∎

Let us return to the proof of the proposition. First we take a locally finite Stein open cover {Vj}\{V_{j}\} of VV. Denote Vj0,j1,,jp:=Vj1VjpV_{j_{0},j_{1},\dots,j_{p}}:=V_{j_{1}}\cap\cdots\cap V_{j_{p}}. By Lemma 2.13, after a possible refinement we assume that there is a complete Kähler metric on Vj0,j1,,jpo:=Vj0,j1,,jpXoV_{j_{0},j_{1},\dots,j_{p}}^{o}:=V_{j_{0},j_{1},\dots,j_{p}}\cap X^{o} which has a bounded potential for every j0,j1,,jpj_{0},j_{1},\dots,j_{p}. Since αIm¯¯L(2)n,q(Vo,E;ds2,h)\alpha\in\overline{{\rm Im}\bar{\partial}}\subset L^{n,q}_{(2)}(V^{o},E;ds^{2},h) and Ker¯\rm Ker\bar{\partial} is closed in L(2)n,q(Vo,E;ds2,h)L^{n,q}_{(2)}(V^{o},E;ds^{2},h), we know that ¯α=0\bar{\partial}\alpha=0.

Since SX(E,h)S_{X}(E,h) is a coherent sheaf on XX, by Theorem 2.11 and Lemma 2.2 there are isomorphisms of cohomologies

(3.36) Hk({Vj},SX(E,h))Hk(V,SX(E,h))Hk(Γ(V,𝒟X,E;ds2,h)),\displaystyle H^{k}(\{V_{j}\},S_{X}(E,h))\simeq H^{k}(V,S_{X}(E,h))\simeq H^{k}(\Gamma(V,\mathscr{D}^{\bullet}_{X,E;ds^{2},h})),

where Hk({Vj},SX(E,h))H^{k}(\{V_{j}\},S_{X}(E,h)) is the Čech cohomology with respect to the covering {Vj}\{V_{j}\}.

Let us recall the corresponding Čech cocycle of α\alpha. Let αj=α|Vjo\alpha_{j}=\alpha|_{V_{j}^{o}}. By Lemma 3.6, there exists bjL(2)n,q1(Vjo,E;ds2,h)b_{j}\in L_{(2)}^{n,q-1}(V_{j}^{o},E;ds^{2},h) such that αj=¯bj\alpha_{j}=\bar{\partial}b_{j} on VjoV_{j}^{o} for each jj.

Suppose that {αj0,,jr}\{\alpha_{j_{0},\dots,j_{r}}\} and {bj0,,jr}\{b_{j_{0},\dots,j_{r}}\} are determined in the way as:

(3.37) bj0,,jrL(2)n,qr1(Vj0,j1,,jpo,E;ds2,h),αj0,,jrL(2)n,qr(Vj0,j1,,jpo,E;ds2,h),\displaystyle b_{j_{0},\dots,j_{r}}\in L_{(2)}^{n,q-r-1}(V_{j_{0},j_{1},\dots,j_{p}}^{o},E;ds^{2},h),\quad\alpha_{j_{0},\dots,j_{r}}\in L_{(2)}^{n,q-r}(V_{j_{0},j_{1},\dots,j_{p}}^{o},E;ds^{2},h),
αj0,,jr=¯bj0,,jronVj0,,jroand(δα)j0,,jr+1=0.\displaystyle\alpha_{j_{0},\dots,j_{r}}=\bar{\partial}b_{j_{0},\dots,j_{r}}\quad{\rm on}\quad V_{j_{0},\dots,j_{r}}^{o}\quad{\rm and}\quad(\delta\alpha)_{j_{0},\dots,j_{r+1}}=0.

Set

αj0,,jr+1:=(δb)j0,,jr+1,\alpha_{j_{0},\dots,j_{r+1}}:=(\delta b)_{j_{0},\dots,j_{r+1}},

which is a ¯\bar{\partial}-closed form. It follows from Lemma 3.6 that the same statements in (3.37) also hold for αj0,,jr+1\alpha_{j_{0},\dots,j_{r+1}}. Repeating the above steps, we can obtain a qq-cocycle {αj0,,jq}Zq({Vj},SX(E,h))\{\alpha_{j_{0},\dots,j_{q}}\}\in Z^{q}(\{V_{j}\},S_{X}(E,h)) which corresponds to α\alpha by (3.36).

By the hypothesis there exits a sequence {γm}L(2)n,q1(Vo,E;ds2,h)\{\gamma_{m}\}\in L_{(2)}^{n,q-1}(V^{o},E;ds^{2},h) such that α¯γm0\|\alpha-\bar{\partial}\gamma_{m}\|\rightarrow 0 as mm\rightarrow\infty. Let νm:=α¯γm\nu_{m}:=\alpha-\bar{\partial}\gamma_{m}. Since ¯νm=0\bar{\partial}\nu_{m}=0, by Lemma 3.6 there exists {μj,m}L(2)n,q1(Vjo,E;ds2,h)\{\mu_{j,m}\}\in L_{(2)}^{n,q-1}(V_{j}^{o},E;ds^{2},h) such that νm=¯μj,m\nu_{m}=\bar{\partial}\mu_{j,m} and μj,mCjνm\|\mu_{j,m}\|\leq C_{j}\|\nu_{m}\| for CjC_{j} independent of mm but depending on jj. Set ρj,m:=bjμj,mγm|Vjo\rho_{j,m}:=b_{j}-\mu_{j,m}-\gamma_{m}|_{V_{j}^{o}}. Then we have ¯ρj,m=0\bar{\partial}\rho_{j,m}=0, ¯(αijδ{ρj,m})=0\bar{\partial}(\alpha_{ij}-\delta\{\rho_{j,m}\})=0 and αijδ{ρj,m}Cijνm\|\alpha_{ij}-\delta\{\rho_{j,m}\}\|\leq C_{ij}\|\nu_{m}\|.

Suppose that {ρj0,,jr,m}\{\rho_{j_{0},\dots,j_{r},m}\} are already determined as:

(3.38) ρj0,,jr,mL(2)n,qr1(Vj0,,jro,E;ds2,h),¯(αj0,,jr+1(δρ)j0,.,jr+1,m)=0,\displaystyle\rho_{j_{0},\dots,j_{r},m}\in L_{(2)}^{n,q-r-1}(V_{j_{0},\dots,j_{r}}^{o},E;ds^{2},h),\bar{\partial}(\alpha_{j_{0},\dots,j_{r+1}}-(\delta\rho)_{j_{0},\dots.,j_{r+1},m})=0,
¯ρj0,,jr,m=0,αj0,,jr+1(δρ)j0,.,jr+1,mCj0,,jr+1νm.\displaystyle\bar{\partial}\rho_{j_{0},\dots,j_{r},m}=0,\|\alpha_{j_{0},\dots,j_{r+1}}-(\delta\rho)_{j_{0},\dots.,j_{r+1},m}\|\leq C_{j_{0},\dots,j_{r+1}}\|\nu_{m}\|.

We construct {ρj0,,jr+1,m}\{\rho_{j_{0},\dots,j_{r+1},m}\} as follows. (3.38) and Lemma 3.6 imply that there exists γj0,,jr,mL(2)n,qr2(Vj0,,jro,E;ds2,h)\gamma_{j_{0},\dots,j_{r},m}\in L_{(2)}^{n,q-r-2}(V_{j_{0},\dots,j_{r}}^{o},E;ds^{2},h) and μj0,,jr+1,mL(2)n,qr2(Vj0,,jr+1o,E;ds2,h)\mu_{j_{0},\dots,j_{r+1},m}\in L_{(2)}^{n,q-r-2}(V_{j_{0},\dots,j_{r+1}}^{o},E;ds^{2},h) such that ρj0,,jr,m=¯γj0,,jr,m\rho_{j_{0},\dots,j_{r},m}=\bar{\partial}\gamma_{j_{0},\dots,j_{r},m}, αj0,,jr+1(δρm)j0,,jr+1=¯μj0,,jr+1,m\alpha_{j_{0},\dots,j_{r+1}}-(\delta\rho_{m})_{j_{0},\dots,j_{r+1}}=\bar{\partial}\mu_{j_{0},\dots,j_{r+1},m} and μj0,,jr+1,mCj0,,jr+1νm\|\mu_{j_{0},\dots,j_{r+1},m}\|\leq C_{j_{0},\dots,j_{r+1}}\|\nu_{m}\|. Then we set

ρj0,,jr+1,m=bj0,,jr+1μj0,,jr+1,m(δγ)j0,,jr+1,m,\rho_{j_{0},\dots,j_{r+1},m}=b_{j_{0},\dots,j_{r+1}}-\mu_{j_{0},\dots,j_{r+1},m}-(\delta\gamma)_{j_{0},\dots,j_{r+1},m},

which satisfies the statements in (3.38). Repeating the above steps, we obtain a q1q-1 cochain ρm={ρj0,,jq1,m}Cq1({Vj},SX(E,h))\rho_{m}=\{\rho_{j_{0},\dots,j_{q-1},m}\}\in C^{q-1}(\{V_{j}\},S_{X}(E,h)) such that αδρm0\|\alpha-\delta\rho_{m}\|\rightarrow 0 as mm\rightarrow\infty. By [Hormander1990, Theorem 2.2.3], δρm\delta\rho_{m} tends to α\alpha as mm\rightarrow\infty uniformly on compact subsets of XX. Since SX(E,h)S_{X}(E,h) is coherent, δCq1({Vj},SX(E,h))\delta C^{q-1}(\{V_{j}\},S_{X}(E,h)) is a closed subspace of Cq({Vj},SX(E,h))C^{q}(\{V_{j}\},S_{X}(E,h)) with respect to the topology of uniform convergence on compact subsets. Hence there exists ρCq1({Vj},SX(E,h))\rho\in C^{q-1}(\{V_{j}\},S_{X}(E,h)) such that δρ=α\delta\rho=\alpha. By (3.36), [α]=0Hq(Γ(V,𝒟X,E;ds2,h))[\alpha]=0\in H^{q}(\Gamma(V,\mathscr{D}^{\bullet}_{X,E;ds^{2},h})). Thus we prove the proposition. ∎

4. An abstract Kollár package

Definition 4.1.

[Takegoshi1995, Definition 6.1] A morphism f:XYf:X\rightarrow Y between complex spaces is locally Kähler if f1Uf^{-1}U is a Kähler space for any relatively compact open subset UYU\subset Y.

Throughout this section, f:XYf:X\rightarrow Y is a proper locally Kähler morphism from a complex space XX to an irreducible complex space YY. Assume that every irreducible component of XX is mapped onto YY. XoXregX^{o}\subset X_{\rm reg} is a dense Zariski open subset and (E,h)(E,h) is a hermitian vector bundle on XoX^{o} with a Nakano semi-positive curvature. In this section we establish the Kollár package of SX(E,h)S_{X}(E,h) under the coherence assumptions.

Theorem 4.2 (Torsion Freeness).

Assume that SX(E,h)S_{X}(E,h) is a coherent sheaf on XX, then RqfSX(E,h)R^{q}f_{\ast}S_{X}(E,h) is torsion free for every q0q\geq 0 and vanishes if q>dimXdimYq>\dim X-\dim Y.

Proof.

Since the problem is local, we assume that YY is Stein and there is a Kähler metric on XoX^{o} which is locally complete and locally bounded from below by a hermitian metric and which admits a (,1)(\infty,1) potential locally on XX (Lemma 2.13). Define the sheaf ΩX,ds2,(2)p(E,h)\Omega^{p}_{X,ds^{2},(2)}(E,h) as

ΩX,ds2,(2)p(E,h)(U)={α𝒟X,ds2p,0(E,h)(U)|¯α=0}\Omega^{p}_{X,ds^{2},(2)}(E,h)(U)=\left\{\alpha\in\mathscr{D}^{p,0}_{X,ds^{2}}(E,h)(U)\bigg{|}\bar{\partial}\alpha=0\right\}

for every open subset UXU\subset X.

By Proposition 3.2-(3), the Hodge star operator induces a well defined map

:fn,q(E,h)fΩX,ds2,(2)nq(E,h).\ast:\mathscr{H}^{n,q}_{f}(E,h)\to f_{\ast}\Omega^{n-q}_{X,ds^{2},(2)}(E,h).

Since the Lefschetz operator LL with respect to ds2ds^{2} is bounded and [L,¯]=0[L,\bar{\partial}]=0, taking the finesss of 𝒟X,ds2n,(E,h)\mathscr{D}^{n,\bullet}_{X,ds^{2}}(E,h) (Lemma 2.2) into account we get a well defined map

Lq:fΩX,ds2,(2)nq(E,h)Rqf(𝒟X,ds2n,(E,h)).L^{q}:f_{\ast}\Omega^{n-q}_{X,ds^{2},(2)}(E,h)\to R^{q}f_{\ast}(\mathscr{D}^{n,\bullet}_{X,ds^{2}}(E,h)).

By Theorem 2.11 and Theorem 3.4, we get the morphisms

RqfSX(E,h)fΩX,ds2,(2)nq(E,h)LqRqfSX(E,h)R^{q}f_{\ast}S_{X}(E,h)\stackrel{{\scriptstyle\ast}}{{\to}}f_{\ast}\Omega^{n-q}_{X,ds^{2},(2)}(E,h)\stackrel{{\scriptstyle L^{q}}}{{\to}}R^{q}f_{\ast}S_{X}(E,h)

such that

(4.1) Lq=c(n,q)Id,c(n,q)=1(nq)(nq+3)q!(3.31).\displaystyle L^{q}\circ\ast=c(n,q){\rm Id},\quad c(n,q)=\sqrt{-1}^{(n-q)(n-q+3)}q!\quad(\ref{align_Last=1}).

This proves the first claim since fΩX,ds2,(2)nq(E,h)f_{\ast}\Omega^{n-q}_{X,ds^{2},(2)}(E,h) is torsion free for every q0q\geq 0. The vanishing result follows from Proposition 3.3-(1). ∎

Theorem 4.3 (Injectivity Theorem).

Assume that SX(E,h)S_{X}(E,h) is a coherent sheaf on XX. If LL is a semi-positive holomorphic line bundle on XX so that LlL^{l} admits a nonzero holomorphic global section ss for some l>0l>0, then the canonical morphism

Rqf(s):Rqf(SX(E,h)Lk)Rqf(SX(E,h)Lk+l)R^{q}f_{\ast}(\otimes s):R^{q}f_{\ast}(S_{X}(E,h)\otimes L^{\otimes k})\to R^{q}f_{\ast}(S_{X}(E,h)\otimes L^{\otimes k+l})

is injective for every q0q\geq 0 and every k1k\geq 1.

Proof.

Since the problem is local, we assume that YY is Stein and there is a Kähler metric on XoX^{o} which is locally complete and locally bounded from below by a hermitian metric and which admits a (,1)(\infty,1) potential locally on XX (Lemma 2.13). Let hLh_{L} be the hermitian metric on LL with a semi-positive curvature, then we have

SX(E,h)LkSX(ELk,hhLk),k1S_{X}(E,h)\otimes L^{\otimes k}\simeq S_{X}(E\otimes L^{\otimes k},h\otimes h_{L}^{k}),\quad k\geq 1

by Lemma 2.7. By Theorem 3.4, it is therefore sufficient to show that the canonical map

(4.2) s:n,q(X(U),ELk,hhLk)n,q(X(U),ELk+l,hhLk+l)\displaystyle\otimes s:\mathscr{H}^{n,q}(X(U),E\otimes L^{\otimes k},h\otimes h_{L}^{k})\to\mathscr{H}^{n,q}(X(U),E\otimes L^{\otimes k+l},h\otimes h_{L}^{k+l})

is well defined for every Stein open subset UU of YY. Let αn,q(X,ELk,hhLk,φ)\alpha\in\mathscr{H}^{n,q}(X,E\otimes L^{\otimes k},h\otimes h_{L}^{k},\varphi) for some φ𝒫e(X)\varphi\in\mathscr{P}_{e}(X) (𝒫e(X)\mathscr{P}_{e}(X)\neq\emptyset since YY is Stein and ff is proper). It follows from Proposition 3.2-(1) that

Dh(αs)=¯(αs)=Dhαs=0D^{\prime\ast}_{h}(\alpha\otimes s)=-\ast\bar{\partial}\ast(\alpha\otimes s)=D^{\prime\ast}_{h}\alpha\otimes s=0

and

0\displaystyle 0 1e(ΘhhLk+l(ELk+l)+¯φ)Λ(αs),αshhLk+l\displaystyle\leq\langle\sqrt{-1}e(\Theta_{h\otimes h_{L}^{k+l}}(E\otimes L^{k+l})+\partial\bar{\partial}\varphi)\Lambda(\alpha\otimes s),\alpha\otimes s\rangle_{h\otimes h_{L}^{k+l}}
k+lk|s|hLl21e(ΘhhLk(ELk)+¯φ)Λ(α),αhhLk=0.\displaystyle\leq\frac{k+l}{k}|s|_{h_{L}^{l}}^{2}\langle\sqrt{-1}e(\Theta_{h\otimes h_{L}^{k}}(E\otimes L^{k})+\partial\bar{\partial}\varphi)\Lambda(\alpha),\alpha\rangle_{h\otimes h_{L}^{k}}=0.

Hence αsn,q(X,ELk+l,hhLk+l,φ)\alpha\otimes s\in\mathscr{H}^{n,q}(X,E\otimes L^{\otimes k+l},h\otimes h_{L}^{k+l},\varphi) by Proposition 3.2-(1). Hence s\otimes s is well defined and injective. The proof is finished. ∎

Theorem 4.4 (Decomposition).

Assume that there exists a Kähler metric ds2ds^{2} on XoX^{o} such that

  1. (1)

    ds2ds^{2} admits a (,1)(\infty,1) bounded potential locally on XX;

  2. (2)

    ds2ds^{2} is locally complete on XX and is locally bounded from below by a hermitian metric.

Assume that SX(E,h)S_{X}(E,h) is a coherent sheaf on XX. Then RfSX(E,h)Rf_{\ast}S_{X}(E,h) splits in D(Y)D(Y), i.e.

RfSX(E,h)qRqfSX(E,h)[q]D(Y).Rf_{\ast}S_{X}(E,h)\simeq\bigoplus_{q}R^{q}f_{\ast}S_{X}(E,h)[-q]\in D(Y).

As a consequence, the spectral sequence

(4.3) E2pq:=Hp(Y,RqfSX(E,h))Hp+q(X,SX(E,h))\displaystyle E^{pq}_{2}:=H^{p}(Y,R^{q}f_{\ast}S_{X}(E,h))\Rightarrow H^{p+q}(X,S_{X}(E,h))

degenerates at the E2E_{2} page if YY is compact.

Remark 4.5.

By Lemma 2.13, such kind of metric exists if XX is compact.

Proof.

Let

(4.4) α:qfn,q(E,h)[q]f𝒟X,E;ds2,hRf𝒟X,E;ds2,h\displaystyle\alpha:\bigoplus_{q}\mathscr{H}^{n,q}_{f}(E,h)[-q]\to f_{\ast}\mathscr{D}^{\bullet}_{X,E;ds^{2},h}\simeq Rf_{\ast}\mathscr{D}^{\bullet}_{X,E;ds^{2},h}

be the inclusion map. By Theorem 3.4, α\alpha is a quasi-isomorphism under the hypothesis that SX(E,h)S_{X}(E,h) is coherent. Hence by Theorem 2.11 and Theorem 3.4 we obtain that

RfSX(E,h)Rf𝒟X,E;ds2,hqfn,q(E,h)[q]qRqfSX(E,h)[q]Rf_{\ast}S_{X}(E,h)\simeq Rf_{\ast}\mathscr{D}^{\bullet}_{X,E;ds^{2},h}\simeq\bigoplus_{q}\mathscr{H}^{n,q}_{f}(E,h)[-q]\simeq\bigoplus_{q}R^{q}f_{\ast}S_{X}(E,h)[-q]

in D(Y)D(Y). The degeneration of the spectral sequence follows from standard arguments. ∎

Theorem 4.6 (Vanishing Theorem).

Assume that SX(E,h)S_{X}(E,h) is a coherent sheaf on XX. If YY is a projective algebraic variety and LL is an ample line bundle on YY, then

Hq(Y,RpfSX(E,h)L)=0,q>0,p0.H^{q}(Y,R^{p}f_{\ast}S_{X}(E,h)\otimes L)=0,\quad\forall q>0,\forall p\geq 0.
Proof.

Since SX(E,h)S_{X}(E,h) is coherent, so is RpfSX(E,h)R^{p}f_{\ast}S_{X}(E,h). Then there is kk large enough so that H0(Y,Lk)0H^{0}(Y,L^{\otimes k})\neq 0 and

(4.5) Hq(Y,RpfSX(E,h)Lk+1)=0,q>0,p0.\displaystyle H^{q}(Y,R^{p}f_{\ast}S_{X}(E,h)\otimes L^{k+1})=0,\quad\forall q>0,\forall p\geq 0.

By Lemma 2.7, we acquire that

(4.6) SX(E,h)fLlSX(ELl,hfhLl),l>0\displaystyle S_{X}(E,h)\otimes f^{\ast}L^{\otimes l}\simeq S_{X}(E\otimes L^{\otimes l},h\otimes f^{\ast}h_{L}^{l}),\quad\forall l>0

where hLh_{L} is a hermitian metric on LL with positive curvature. By Theorem 4.3, the canonical map

fs:Hq(X,SX(EL,hfhL))Hq(X,SX(ELk+1,hfhLk+1)),q0\otimes f^{\ast}s:H^{q}(X,S_{X}(E\otimes L,h\otimes f^{\ast}h_{L}))\to H^{q}(X,S_{X}(E\otimes L^{\otimes k+1},h\otimes f^{\ast}h_{L}^{k+1})),\quad\forall q\geq 0

is injective. By Theorem 4.4, we know that the canonical map

s:Hq(Y,RpfSX(EL,hfhL))Hq(Y,RpfSX(ELk+1,hfhLk+1)),p,q0\otimes s:H^{q}(Y,R^{p}f_{\ast}S_{X}(E\otimes L,h\otimes f^{\ast}h_{L}))\to H^{q}(Y,R^{p}f_{\ast}S_{X}(E\otimes L^{\otimes k+1},h\otimes f^{\ast}h_{L}^{k+1})),\quad\forall p,q\geq 0

is injective. Combining this with (4.5) and (4.6), we prove the theorem. ∎

5. Non-abelian Hodge theory and Kollár package

5.1. Harmonic bundle and variation of Hodge structure

The notion of harmonic bundle is used by Simpson [Simpson1992] to establish a correspondence between local systems and semistable higgs bundles with vanishing Chern classes over a compact Kähler manifold. A typical example of harmonic bundles comes from a polarized variation of Hodge structure (loc. cit.). A harmonic bundle produces a λ\lambda-connection structure which gives a 𝒟\mathscr{D}-module on λ=1\lambda=1 and a higgs bundle on λ=0\lambda=0. This is the main subject of non-abelian Hodge theory. We only review the necessary knowledge of this topic that is used in the present paper. Readers may consult [Simpson1988, Simpson1990, Simpson1992, Sabbah2005, Mochizuki20072, Mochizuki20071] for more details.

Let (M,ds2)(M,ds^{2}) be a Kähler manifold and (𝒱,)(\mathcal{V},\nabla) a holomorphic vector bundle with a flat connection on MM. Let hh be a hermitian metric on 𝒱\mathcal{V} which is not necessarily compatible with \nabla.

Let =1,0+0,1\nabla=\nabla^{1,0}+\nabla^{0,1} be the bi-degree decomposition. There are unique operators δh\delta^{\prime}_{h} and δh′′\delta^{\prime\prime}_{h} so that 1,0+δh′′\nabla^{1,0}+\delta^{\prime\prime}_{h} and 0,1+δh\nabla^{0,1}+\delta^{\prime}_{h} are connections compatible with hh. Denote hc=δh′′δh\nabla^{c}_{h}=\delta^{\prime\prime}_{h}-\delta^{\prime}_{h} and Θh()=hc+hc\Theta_{h}(\nabla)=\nabla\nabla^{c}_{h}+\nabla^{c}_{h}\nabla.

Definition 5.1.

(𝒱,,h)(\mathcal{V},\nabla,h) is called a harmonic bundle if Θh()=0\Theta_{h}(\nabla)=0. In this case, hh is called a harmonic metric.

Let (𝒱,,h)(\mathcal{V},\nabla,h) be a harmonic bundle. Denote θ=12(1,0δh)\theta=\frac{1}{2}(\nabla^{1,0}-\delta^{\prime}_{h}) and ¯=12(0,1+δh′′)\bar{\partial}=\frac{1}{2}(\nabla^{0,1}+\delta^{\prime\prime}_{h}). Then ¯2=0\bar{\partial}^{2}=0. Denote by 𝒱~\widetilde{\mathcal{V}} the underlying complex CC^{\infty}-vector bundle of 𝒱\mathcal{V}, then H=(𝒱~,¯)H=(\widetilde{\mathcal{V}},\bar{\partial}) is a holomorphic vector bundle and θ\theta is a higgs field on HH (i.e. θ\theta is 𝒪M\mathscr{O}_{M}-linear and θ2=0\theta^{2}=0).

Let h\nabla_{h} be the Chern connection on HH with respect to hh and let Θh(H)=h2\Theta_{h}(H)=\nabla_{h}^{2} be its curvature form. Then we have the self-dual equation

(5.1) Θh(H)+θθ¯+θ¯θ=0\displaystyle\Theta_{h}(H)+\theta\wedge\overline{\theta}+\overline{\theta}\wedge\theta=0

where θ¯=12(0,1δh′′)\overline{\theta}=\frac{1}{2}(\nabla^{0,1}-\delta^{\prime\prime}_{h}) is the adjoint of θ\theta with respect to the metric hh.

Conversely, let (H,θ,h)(H,\theta,h) be a hermitian higgs bundle . Let θ¯\overline{\theta} be the adjoint of θ\theta and let \partial be the unique (1,0)(1,0)-connection such that +¯\partial+\bar{\partial} is compatible with hh. hh is called harmonic if Θh(θ):=(+¯+θ+θ¯)2=0\Theta_{h}(\theta):=(\partial+\bar{\partial}+\theta+\overline{\theta})^{2}=0. There is a 1:11:1 correspondence (c.f. [Simpson1988, Simpson1992])

(5.2) {(𝒱,,h):Θh()=0}1:1{(H,θ,h):Θh(θ)=0}.\displaystyle\left\{(\mathcal{V},\nabla,h):\Theta_{h}(\nabla)=0\right\}\stackrel{{\scriptstyle 1:1}}{{\leftrightarrow}}\left\{(H,\theta,h):\Theta_{h}(\theta)=0\right\}.

Therefore by a harmonic bundle we mean an object on either side of (5.2).

For the purpose of the present paper, we are interested in tame harmonic bundles in the sense of Simpson [Simpson1990] and Mochizuki [Mochizuki20072, Mochizuki20071].

Definition 5.2.

Let XX be a complex space and XoXregX^{o}\subset X_{\rm reg} a dense Zariski open subset. A harmonic bundle (𝒱,,h)(\mathcal{V},\nabla,h) on XoX^{o} is called a tame harmonic bundle on (X,Xo)(X,X^{o}) if, locally at every point xXx\in X, there exists some c>0c>0 such that

(5.3) (i=1rfi2)c|v|h(i=1rfi2)c\displaystyle(\sum_{i=1}^{r}\|f_{i}\|^{2})^{c}\lesssim|v|_{h}\lesssim(\sum_{i=1}^{r}\|f_{i}\|^{2})^{-c}

holds for every (multivalued) flat section vv. Here {f1,,fr}\{f_{1},\dots,f_{r}\} is a local generator of the ideal sheaf defining X\XoXX\backslash X^{o}\subset X and c>0c>0 is a constant independent of vv.

The tameness is independent of the choice of the local generator {f1,,fr}\{f_{1},\dots,f_{r}\}. Let π:X~X\pi:\widetilde{X}\to X be a desingularization of X\XoXX\backslash X^{o}\subset X so that the preimage of X\XoX\backslash X^{o} is supported on a simple normal crossing divisor EE. Then (5.3) is equivalent to the estimate

(5.4) z1zrc|v|hz1zrc, for some c>0\displaystyle\|z_{1}\cdots z_{r}\|^{c^{\prime}}\lesssim|v|_{h}\lesssim\|z_{1}\cdots z_{r}\|^{-c^{\prime}},\quad\textrm{ for some }c^{\prime}>0

where (z1,,zn)(z_{1},\dots,z_{n}) is an arbitrary holomorphic local coordinate of X~\widetilde{X} such that E={z1zr=0}E=\{z_{1}\cdots z_{r}=0\}.

Remark 5.3.

There is an equivalent definition of tameness, using the corresponding higgs bundle. By desingularization, we assume that XX is smooth and X\XoXX\backslash X^{o}\subset X is a normal crossing divisor. Then the higgs field θ\theta of the higgs bundle (H,θ)(H,\theta) associated to (𝒱,,h)(\mathcal{V},\nabla,h) can be described as:

(5.5) θ=i=1rfidzizi+j=r+1ngjdzj\displaystyle\theta=\sum_{i=1}^{r}f_{i}\frac{dz_{i}}{z_{i}}+\sum_{j=r+1}^{n}g_{j}dz_{j}

under a holomorphic coordinate UXU\subset X such that U(X\Xo)={z1zr=0}U\cap(X\backslash X^{o})=\{z_{1}\cdots z_{r}=0\}. The harmonic bundle (𝒱,,h)(\mathcal{V},\nabla,h) is tame if and only if the coefficients of the characteristic polynomials det(tfi)\det(t-f_{i}), i=1,,ri=1,\dots,r and det(tgj)\det(t-g_{j}), j=r+1,,nj=r+1,\dots,n can be extended to the holomorphic functions on UU. Readers may see [Mochizuki2002] for more details.

A typical type of tame harmonic bundles is the variation of Hodge structure.

Definition 5.4.

[Simpson1988, §8] Let XX be a complex space and XoXregX^{o}\subset X_{\rm reg} a dense Zariski open subset. Denote by 𝒜Xo0\mathscr{A}^{0}_{X^{o}} the sheaf of CC^{\infty} functions on XoX^{o}. A polarized complex variation of Hodge structure on XoX^{o} of weight kk is a harmonic bundle (𝒱,,h)(\mathcal{V},\nabla,h) on XoX^{o} together with a decomposition 𝒱𝒪X𝒜Xo0=p+q=k𝒱p,q\mathcal{V}\otimes_{\mathscr{O}_{X}}\mathscr{A}^{0}_{X^{o}}=\bigoplus_{p+q=k}\mathcal{V}^{p,q} of CC^{\infty}-bundles such that

  1. (1)

    The Griffiths transversality condition

    (5.6) (𝒱p,q)𝒜0,1(𝒱p+1,q1)𝒜1(𝒱p,q)𝒜1,0(Vp1,q+1)\displaystyle\nabla(\mathcal{V}^{p,q})\subset\mathscr{A}^{0,1}(\mathcal{V}^{p+1,q-1})\oplus\mathscr{A}^{1}(\mathcal{V}^{p,q})\oplus\mathscr{A}^{1,0}(V^{p-1,q+1})

    holds for every pp and qq. Here 𝒜p,q(𝒱i,j)\mathscr{A}^{p,q}(\mathcal{V}^{i,j}) (resp. 𝒜k(𝒱i,j)\mathscr{A}^{k}(\mathcal{V}^{i,j})) denotes the sheaf of CC^{\infty} (p,q)(p,q)-forms (resp. kk-forms) with values in 𝒱i,j\mathcal{V}^{i,j}.

  2. (2)

    The hermitian form QQ which equals (1)ph(-1)^{p}h on 𝕍p,q\mathbb{V}^{p,q} is parallel with respect to \nabla.

Denote S(𝕍):=pmaxS(\mathbb{V}):=\mathcal{F}^{p_{\rm max}} where pmax=max{p|p0}p_{\rm max}=\max\{p|\mathcal{F}^{p}\neq 0\}.

The following proposition is known to experts. We recall the proof in sketch for the convenience of readers.

Proposition 5.5.

If a harmonic bundle (𝒱,,h)(\mathcal{V},\nabla,h) admits a complex variation of Hodge structure, then (𝒱,,h)(\mathcal{V},\nabla,h) is tame.

Proof.

Let 𝕍:=(𝒱,,{𝒱p,q},h𝕍)\mathbb{V}:=(\mathcal{V},\nabla,\{\mathcal{V}^{p,q}\},h_{\mathbb{V}}) be a complex variation of Hodge structure of weight kk. Take the decomposition

=θ¯++¯+θ\nabla=\overline{\theta}+\partial+\bar{\partial}+\theta

according to (5.6). By [Simpson1988, §8], the corresponding higgs bundle of (𝒱,,h𝕍)(\mathcal{V},\nabla,h_{\mathbb{V}}) is (H=Ker¯,θ,h𝕍)(H={\rm Ker}\bar{\partial},\theta,h_{\mathbb{V}}). There is moreover an orthogonal decomposition of holomorphic subbundles H=p+q=kHp,qH=\oplus_{p+q=k}H^{p,q} where Hp,q=H𝒱p,qH^{p,q}=H\cap\mathcal{V}^{p,q} and

θ(Hp,q)Hp1,q+1ΩXo.\theta(H^{p,q})\subset H^{p-1,q+1}\otimes\Omega_{X^{o}}.

Thus the higgs field is nilpotent. Therefore fif_{i}, i=1,,ri=1,\dots,r and gjg_{j}, j=r+1,,nj=r+1,\dots,n in the local expression (5.5) are nilpotent matrix. Hence det(tfi)=tn\det(t-f_{i})=t^{n} and det(tgj)=tn\det(t-g_{j})=t^{n}. So the harmonic bundle is tame due to the alternative definition of tameness (Remark 5.3). ∎

5.2. Harmonic bundle and Kollár package

Proposition 5.6.

Let XX be a complex space and XoXregX^{o}\subset X_{\rm reg} a dense Zariski open subset. Let (H,θ,h)(H,\theta,h) be a tame harmonic bundle on XoX^{o}. Let EHE\subset H be a holomorphic subbundle with vanishing second fundamental form. Assume that θ¯(E)=0\overline{\theta}(E)=0. Then (E,h)(E,h) is a tame hermitian vector bundle with a Nakano semi-positive curvature. As a consequence, SX(E,h)S_{X}(E,h) is a coherent sheaf on XX.

Proof.

To prove the tameness (Definition 2.8), we construct QQ by Mochizuki’s prolongation construction. Since the problem is local, we assume that there is a desingularization π:X~X\pi:\widetilde{X}\to X such that π\pi is biholomorphic over XoX^{o} and D:=π1(X\Xo)D:=\pi^{-1}(X\backslash X^{o}) is a simple normal crossing divisor. By abuse of notations we identify XoX^{o} and π1Xo\pi^{-1}X^{o}. Since (H,θ,h)(H,\theta,h) is tame, by [Mochizuki20072, Theroem 8.58] there is a logarithmic higgs bundle (H~,θ~)(\widetilde{H},\widetilde{\theta}):

θ~:H~ΩX~(logD)H~,\displaystyle\widetilde{\theta}:\widetilde{H}\to\Omega_{\widetilde{X}}(\log D)\otimes\widetilde{H},

such that (H~,θ~)|Xo(\widetilde{H},\widetilde{\theta})|_{X^{o}} is holomorphically equivalent to (H,θ)(H,\theta). Let D=i=1kDiD=\cup_{i=1}^{k}D_{i} be the irreducible decomposition and let (z1,,zn)(z_{1},\dots,z_{n}) be a local coordinate such that Di={zi=0}D_{i}=\{z_{i}=0\}, i=1,,ki=1,\dots,k. By [Mochizuki20072, Part 3, Chapter 13], the tameness of (H,θ,h)(H,\theta,h) forces a norm estimate

(5.7) |z1zk|2b|s|h\displaystyle|z_{1}\cdots z_{k}|^{2b}\lesssim|s|_{h}

for any local holomorphic section ss of H~\widetilde{H} and a constant b>0b>0 which is independent of ss.

Let Q:=H~Q:=\widetilde{H}, then EE is a holomorphic subbundle of Q|XoQ|_{X^{o}}. By (5.7), we see that (E,h)(E,h) is tame.

To see that (E,h)(E,h) is Nakano semi-positive, we take the decomposition

=θ¯++¯+θ\nabla=\overline{\theta}+\partial+\bar{\partial}+\theta

according to (5.6). Since EEE\oplus E^{\bot} is an orthogonal holomorphic direct sum of HH and θ¯(E)=0\overline{\theta}(E)=0, we get from (5.1) that

1Θh(E)=1θθ¯0.\sqrt{-1}\Theta_{h}(E)=\sqrt{-1}\theta\wedge\overline{\theta}\geq 0.

Hence (E,h)(E,h) is Nakano semi-positive. By Proposition 2.9, SX(E,h)S_{X}(E,h) is a coherent sheaf on XX. ∎

Let 𝕍:=(𝒱,,{𝒱p,q},h𝕍)\mathbb{V}:=(\mathcal{V},\nabla,\{\mathcal{V}^{p,q}\},h_{\mathbb{V}}) be a complex variation of Hodge structure of weight kk. Let

=θ¯++¯+θ\nabla=\overline{\theta}+\partial+\bar{\partial}+\theta

be the decomposition according to (5.6). The corresponding higgs bundle of (𝒱,,h𝕍)(\mathcal{V},\nabla,h_{\mathbb{V}}) is (H=Ker¯,θ,h𝕍)(H={\rm Ker}\bar{\partial},\theta,h_{\mathbb{V}}) by [Simpson1988, §8]. There is moreover an orthogonal decomposition of holomorphic subbundles H=p+q=kHp,qH=\oplus_{p+q=k}H^{p,q} where Hp,q=H𝒱p,qH^{p,q}=H\cap\mathcal{V}^{p,q} and

θ(Hp,q)Hp1,q+1ΩXo.\theta(H^{p,q})\subset H^{p-1,q+1}\otimes\Omega_{X^{o}}.

For the reason of degrees, we have θ¯(S(𝕍))=0\overline{\theta}(S(\mathbb{V}))=0. By Proposition 5.5, we acquire that (H=Ker¯,θ,h𝕍)(H={\rm Ker}\bar{\partial},\theta,h_{\mathbb{V}}) and S(𝕍)S(\mathbb{V}) satisfy the conditions in Proposition 5.6. Hence SX(S(𝕍),h𝕍)S_{X}(S(\mathbb{V}),h_{\mathbb{V}}) is a coherent sheaf.

The following proposition shows that SX(S(𝕍),h𝕍)S_{X}(S(\mathbb{V}),h_{\mathbb{V}}) coincides with Saito’s SS-sheaf associated to ICX(𝕍)IC_{X}(\mathbb{V}).

Proposition 5.7.

Let 𝕍:=(𝕍,,,h𝕍)\mathbb{V}:=(\mathbb{V},\nabla,\mathcal{F}^{\bullet},h_{\mathbb{V}}) be an \mathbb{R}-polarized variation of Hodge structure. Denote by ICX(𝕍)IC_{X}(\mathbb{V}) the intermediate extension of 𝕍\mathbb{V} on XX as a pure Hodge module and by S(ICX(𝕍))S(IC_{X}(\mathbb{V})) the Saito’s SS-sheaf associated to ICX(𝕍)IC_{X}(\mathbb{V}) ([MSaito1991]). Then

SX(S(𝕍),h𝕍)S(ICX(𝕍)).S_{X}(S(\mathbb{V}),h_{\mathbb{V}})\simeq S(IC_{X}(\mathbb{V})).
Proof.

See [Schnell2020] or [SC2021, Theorem 4.10]. ∎

Now we are ready to show the Kollár package with respect to a tame harmonic bundle.

Theorem 5.8.

Let f:XYf:X\rightarrow Y be a proper locally Kähler morphism from a complex space XX to an irreducible complex space YY. Assume that every irreducible component of XX is mapped onto YY. Let XoXregX^{o}\subset X_{\rm reg} be a dense Zariski open subset and (H,θ,h)(H,\theta,h) a tame harmonic bundle on XoX^{o}. Let EHE\subset H be a holomorphic subbundle with vanishing second fundamental form. Assume that θ¯(E)=0\overline{\theta}(E)=0. Let FF be a Nakano semi-positive vector bundle on XX. Then the following statements hold.

Torsion Freeness:

Rqf(SX(E,h)F)R^{q}f_{\ast}(S_{X}(E,h)\otimes F) is torsion free for every q0q\geq 0 and vanishes if q>dimXdimYq>\dim X-\dim Y.

Injectivity:

If LL is a semi-positive holomorphic line bundle so that LlL^{\otimes l} admits a nonzero holomorphic global section ss for some l>0l>0, then the canonical morphism

Rqf(×s):Rqf(SX(E,h)FLk)Rqf(SX(E,h)FLk+l)R^{q}f_{\ast}(\times s):R^{q}f_{\ast}(S_{X}(E,h)\otimes F\otimes L^{\otimes k})\to R^{q}f_{\ast}(S_{X}(E,h)\otimes F\otimes L^{\otimes k+l})

is injective for every q0q\geq 0 and every k1k\geq 1.

Vanishing:

If YY is a projective algebraic variety and LL is an ample line bundle on YY, then

Hq(Y,Rpf(SX(E,h)F)L)=0,q>0,p0.H^{q}(Y,R^{p}f_{\ast}(S_{X}(E,h)\otimes F)\otimes L)=0,\quad\forall q>0,p\geq 0.
Decomposition:

Assume moreover that XX is a compact Kähler space, then Rf(SX(E,h)F)Rf_{\ast}(S_{X}(E,h)\otimes F) splits in D(Y)D(Y), i.e.

Rf(SX(E,h)F)qRqf(SX(E,h)F)[q]D(Y).Rf_{\ast}(S_{X}(E,h)\otimes F)\simeq\bigoplus_{q}R^{q}f_{\ast}(S_{X}(E,h)\otimes F)[-q]\in D(Y).

As a consequence, the spectral sequence

E2pq:Hp(Y,Rqf(SX(E,h)F))Hp+q(X,SX(E,h)F)E^{pq}_{2}:H^{p}(Y,R^{q}f_{\ast}(S_{X}(E,h)\otimes F))\Rightarrow H^{p+q}(X,S_{X}(E,h)\otimes F)

degenerates at the E2E_{2} page.

Proof.

By Proposition 5.6, SX(E,h)S_{X}(E,h) is a coherent sheaf. Let hFh_{F} be a hermitian metric on FF with Nakano semi-positive curvature. By Lemma 2.7, we see that

SX(E,h)FSX(EF,hhF)S_{X}(E,h)\otimes F\simeq S_{X}(E\otimes F,h\otimes h_{F})

is a coherent sheaf on XX. Note that (EF,hhF)(E\otimes F,h\otimes h_{F}) is Nakano semi-positive by Proposition 5.6. It follows from the abstract Kollár package in §4 that the claims of the theorem are valid. ∎

Taking SX(S(𝕍),h𝕍)S(ICX(𝕍))S_{X}(S(\mathbb{V}),h_{\mathbb{V}})\simeq S(IC_{X}(\mathbb{V})) for a variation of Hodge structure 𝕍\mathbb{V} and F=𝒪XF=\mathscr{O}_{X}, we obtain Kollár’s conjecture.

We end this section with remarks on two other packages of Kollár’s conjecture.

Remark 5.9 (Remarks on the Intersection cohomology package).

In [Kollar1986_2, §5.8], Kollár also predicts that S(ICX(𝕍))S(IC_{X}(\mathbb{V})) is related to the intersection complex ICX(𝕍)IC_{X}(\mathbb{V}) when 𝕍=(𝒱,,,h)\mathbb{V}=(\mathcal{V},\nabla,\mathcal{F}^{\bullet},h) is a polarized variation of Hodge structure. This involves the L2L^{2}-representation of the intersection complex.

Theorem 5.10.

Let XX be a compact Kähler space of pure dimension nn and XoXregX^{o}\subset X_{\rm reg} a dense Zariski open subset. Let 𝕍=(𝒱,,,h)\mathbb{V}=(\mathcal{V},\nabla,\mathcal{F}^{\bullet},h) be an \mathbb{R}-polarized variation of Hodge structure of weight rr on XoX^{o}. Then IHk(X,𝕍)IH^{k}(X,\mathbb{V}) admits a pure Hodge structure of weight kk:

IHk(X,𝕍)=p+q=k+rp,q0IHp,q(X,𝕍),IHp,q(X,𝕍)=IHq,p(X,𝕍)¯IH^{k}(X,\mathbb{V})=\bigoplus_{\stackrel{{\scriptstyle p,q\geq 0}}{{p+q=k+r}}}IH^{p,q}(X,\mathbb{V}),\quad IH^{p,q}(X,\mathbb{V})=\overline{IH^{q,p}(X,\mathbb{V})}

for every 0k2dimX0\leq k\leq 2\dim X. There is moreover a morphism

ICX(𝕍)SX(𝕍)IC_{X}(\mathbb{V})\to S_{X}(\mathbb{V})

in the derived category of sheaves of \mathbb{C}-vector spaces which induces an isomorphism

IHn+r,q(X,𝕍)Hq(X,SX(𝕍)),q0.IH^{n+r,q}(X,\mathbb{V})\to H^{q}(X,S_{X}(\mathbb{V})),\quad\forall q\geq 0.
Proof.

The first statement is a consequence of [SC2021_CGM, Theorem 1.4]. Roughly speaking, there is a complete Kähler metric ds2ds^{2} on XoX^{o} whose L2L^{2}-de Rham complex 𝒟X,𝕍;ds2,h\mathscr{D}^{\bullet}_{X,\mathbb{V};ds^{2},h} is quasi-isomorphic to ICX(𝕍)IC_{X}(\mathbb{V}). As a consequence, there is a canonical isomorphism

IHk(X,𝕍)H(2)k(Xo,𝕍;ds2,h),k.IH^{k}(X,\mathbb{V})\simeq H^{k}_{(2)}(X^{o},\mathbb{V};ds^{2},h),\quad\forall k.

The (p,q)(p,q)-decomposition of forms in 𝒟X,𝕍;ds2,h\mathscr{D}^{\bullet}_{X,\mathbb{V};ds^{2},h} provides the Hodge structure on IHk(X,𝕍)IH^{k}(X,\mathbb{V}). See [SC2021_CGM, §8.3] for details. The second claim follows from the diagram

𝒟X,𝕍;ds2,h\textstyle{\mathscr{D}^{\bullet}_{X,\mathbb{V};ds^{2},h}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}τ\scriptstyle{\tau}\scriptstyle{\simeq}𝒟X,ds2n,(S(𝕍),h)\textstyle{\mathscr{D}^{n,\bullet}_{X,ds^{2}}(S(\mathbb{V}),h)\ignorespaces\ignorespaces\ignorespaces\ignorespaces} (Theorem 2.11, Proposition 5.7)\scriptstyle{\simeq\textrm{ (Theorem \ref{thm_main_local1}, Proposition \ref{prop_coincide_Saito_S})}}ICX(𝕍)\textstyle{IC_{X}(\mathbb{V})}SX(𝕍)\textstyle{S_{X}(\mathbb{V})}

where τ\tau is taking the projection to the S(𝕍)S(\mathbb{V})-valued (n,)(n,\bullet)-component. ∎

Remark 5.11 (Remarks on the direct image package).

In [Kollar1986_2, §5.8], Kollár also predicts that

RqfS(ICX(𝕍))SY(RqfICX(𝕍)|Yo)R^{q}f_{\ast}S(IC_{X}(\mathbb{V}))\simeq S_{Y}(R^{q}f_{\ast}IC_{X}(\mathbb{V})|_{Y^{o}})

where YoY^{o} is the Zariski open subset of YY so that RqfICX(𝕍)|YoR^{q}f_{\ast}IC_{X}(\mathbb{V})|_{Y^{o}} is a local system whose fiber at yYoy\in Y^{o} is canonically isomorphic to q(Xy,ICXy(𝕍|XyXo))\mathbb{H}^{q}(X_{y},IC_{X_{y}}(\mathbb{V}|_{X_{y}\cap X^{o}})). This is also a consequence of the L2L^{2}-representation of ICX(𝕍)IC_{X}(\mathbb{V}) ([SC2021_CGM, Theorem 1.4]).

References