THE EXTENDED LOCUS OF HODGE CLASSES

Similar documents
GENERIC TORELLI THEOREM FOR QUINTIC-MIRROR FAMILY. Sampei Usui

1 Moduli spaces of polarized Hodge structures.

Hodge Theory of Maps

LINKED HOM SPACES BRIAN OSSERMAN

HODGE MODULES ON COMPLEX TORI AND GENERIC VANISHING FOR COMPACT KÄHLER MANIFOLDS. A. Introduction

An overview of D-modules: holonomic D-modules, b-functions, and V -filtrations

1. Algebraic vector bundles. Affine Varieties

Algebraic Geometry. Andreas Gathmann. Class Notes TU Kaiserslautern 2014

arxiv: v1 [math.ag] 13 Mar 2019

where Σ is a finite discrete Gal(K sep /K)-set unramified along U and F s is a finite Gal(k(s) sep /k(s))-subset

ON A THEOREM OF CAMPANA AND PĂUN

APPENDIX 3: AN OVERVIEW OF CHOW GROUPS

MIXED HODGE MODULES PAVEL SAFRONOV

CHAPTER 0 PRELIMINARY MATERIAL. Paul Vojta. University of California, Berkeley. 18 February 1998

THE QUANTUM CONNECTION

The Grothendieck-Katz Conjecture for certain locally symmetric varieties

Math 249B. Nilpotence of connected solvable groups

AN OVERVIEW OF MORIHIKO SAITO S THEORY OF MIXED HODGE MODULES

Math 249B. Geometric Bruhat decomposition

Contributors. Preface

CHAPTER 1. TOPOLOGY OF ALGEBRAIC VARIETIES, HODGE DECOMPOSITION, AND APPLICATIONS. Contents

Monodromy of the Dwork family, following Shepherd-Barron X n+1. P 1 λ. ζ i = 1}/ (µ n+1 ) H.

Course 311: Michaelmas Term 2005 Part III: Topics in Commutative Algebra

Porteous s Formula for Maps between Coherent Sheaves

LECTURE 7: STABLE RATIONALITY AND DECOMPOSITION OF THE DIAGONAL

MA 206 notes: introduction to resolution of singularities

9. Birational Maps and Blowing Up

ON THE NUMBER AND BOUNDEDNESS OF LOG MINIMAL MODELS OF GENERAL TYPE

ON THE NUMBER AND BOUNDEDNESS OF LOG MINIMAL MODELS OF GENERAL TYPE

AFFINE PUSHFORWARD AND SMOOTH PULLBACK FOR PERVERSE SHEAVES

Math 248B. Applications of base change for coherent cohomology

Kähler manifolds and variations of Hodge structures

Notes on p-divisible Groups

Algebraic v.s. Analytic Point of View

THE MONODROMY-WEIGHT CONJECTURE

FOUNDATIONS OF ALGEBRAIC GEOMETRY CLASS 48

Formality of Kähler manifolds

INTEGRATION OF ONE-FORMS ON p-adic ANALYTIC SPACES

Algebraic Varieties. Notes by Mateusz Micha lek for the lecture on April 17, 2018, in the IMPRS Ringvorlesung Introduction to Nonlinear Algebra

Math 676. A compactness theorem for the idele group. and by the product formula it lies in the kernel (A K )1 of the continuous idelic norm

arxiv:alg-geom/ v1 29 Jul 1993

SPACES OF RATIONAL CURVES IN COMPLETE INTERSECTIONS

COMPLEX ALGEBRAIC SURFACES CLASS 9

Special cubic fourfolds

The generalized Hodge and Bloch conjectures are equivalent for general complete intersections

0. Introduction 1 0. INTRODUCTION

k k would be reducible. But the zero locus of f in A n+1

A SHORT PROOF OF ROST NILPOTENCE VIA REFINED CORRESPONDENCES

The geometry of Landau-Ginzburg models

Algebraic Geometry Spring 2009

5 Quiver Representations

where m is the maximal ideal of O X,p. Note that m/m 2 is a vector space. Suppose that we are given a morphism

Hyperkähler geometry lecture 3

8. Prime Factorization and Primary Decompositions

4. Images of Varieties Given a morphism f : X Y of quasi-projective varieties, a basic question might be to ask what is the image of a closed subset

Math 210B. Artin Rees and completions

ALGEBRAIC GEOMETRY I, FALL 2016.

KODAIRA-SAITO VANISHING AND APPLICATIONS

BRIAN OSSERMAN. , let t be a coordinate for the line, and take θ = d. A differential form ω may be written as g(t)dt,

LECTURE 5: SOME BASIC CONSTRUCTIONS IN SYMPLECTIC TOPOLOGY

12. Projective modules The blanket assumptions about the base ring k, the k-algebra A, and A-modules enumerated at the start of 11 continue to hold.

LINKED ALTERNATING FORMS AND LINKED SYMPLECTIC GRASSMANNIANS

SPACES OF RATIONAL CURVES ON COMPLETE INTERSECTIONS

Part V. 17 Introduction: What are measures and why measurable sets. Lebesgue Integration Theory

DELIGNE S THEOREM ON THE SEMISIMPLICITY OF A POLARIZED VHS OVER A QUASIPROJECTIVE BASE

Tamagawa Numbers in the Function Field Case (Lecture 2)

16.2. Definition. Let N be the set of all nilpotent elements in g. Define N

Proof of the Shafarevich conjecture

Compatibly split subvarieties of Hilb n (A 2 k)

Groups of Prime Power Order with Derived Subgroup of Prime Order

Construction of M B, M Dol, M DR

Introduction to Chiral Algebras

FANO MANIFOLDS AND BLOW-UPS OF LOW-DIMENSIONAL SUBVARIETIES. 1. Introduction

(1) is an invertible sheaf on X, which is generated by the global sections

SUBJECT I. TOROIDAL PARTIAL COMPACTIFICATIONS AND MODULI OF POLARIZED LOGARITHMIC HODGE STRUCTURES

( files chap2 to chap

Overview of classical mirror symmetry

9. Integral Ring Extensions

THE ENVELOPE OF LINES MEETING A FIXED LINE AND TANGENT TO TWO SPHERES

BEZOUT S THEOREM CHRISTIAN KLEVDAL

Exercises on chapter 1

Mirror Symmetry: Introduction to the B Model

THE FOURIER TRANSFORM FOR CERTAIN HYPERKÄHLER FOURFOLDS. Contents Introduction 2

Joseph Muscat Universal Algebras. 1 March 2013

Homological mirror symmetry via families of Lagrangians

HODGE THEORY, SINGULARITIES AND D-MODULES

arxiv:alg-geom/ v1 21 Mar 1996

Formal power series rings, inverse limits, and I-adic completions of rings

Generalized Tian-Todorov theorems

RIEMANN S INEQUALITY AND RIEMANN-ROCH

COMPLEX VARIETIES AND THE ANALYTIC TOPOLOGY

Math 145. Codimension

12. Hilbert Polynomials and Bézout s Theorem

LECTURE 6: J-HOLOMORPHIC CURVES AND APPLICATIONS

Chern classes à la Grothendieck

Math 396. Bijectivity vs. isomorphism

Γ 1 (N) given by the W -operator W =. It would be interesting to show

Synopsis of material from EGA Chapter II, 4. Proposition (4.1.6). The canonical homomorphism ( ) is surjective [(3.2.4)].

Rational Curves On K3 Surfaces

CHARACTERISTIC CLASSES

Transcription:

THE EXTENDED LOCUS OF HODGE CLASSES CHRISTIAN SCHNELL Abstract. We introduce an extended locus of Hodge classes that also takes into account integral classes that become Hodge classes in the limit. More precisely, given a polarized variation of integral Hodge structure of weight zero on a Zariski-open subset of a complex manifold, we construct a canonical analytic space that parametrizes limits of integral classes; the extended locus of Hodge classes is an analytic subspace that contains the usual locus of Hodge classes, but is finite and proper over the base manifold. The construction uses Saito s theory of mixed Hodge modules and a small generalization of the main technical result of Cattani, Deligne, and Kaplan. We study the properties of the resulting analytic space in the case of the family of hyperplane sections of an odd-dimensional smooth projective variety. A. Overview 1. Hodge loci on Calabi-Yau threefolds. The purpose of this paper is to describe the construction of the extended locus of Hodge classes for polarized variations of Z-Hodge structure of weight zero. Before defining things more precisely, we shall consider a typical example that shows why this is an interesting problem, and what some of the issues are. Let X be a smooth projective Calabi-Yau threefold; this means that Ω 3 X O X, and that H 1 X, O X ) = H 2 X, O X ) = 0. We fix an embedding of X into projective space, with O X 1) the corresponding very ample line bundle, and consider the family of hyperplane sections of X. These are parametrized by the linear system B = OX 1), and we let B 0 B denote the open subset that corresponds to smooth hyperplane sections. Given a cohomology class γ H 2 S, Z) on a smooth hyperplane section S X, we can use parallel transport along paths in B 0 to move γ to other hyperplane sections; this operation is of course purely topological and does not preserve the Hodge decomposition. The Hodge locus of γ is the set { b B0 γ can be transported to a Hodge class on Sb }. Most of these loci are non-empty: in fact, Voisin [Voi92] has proved that the union of the Hodge loci of all classes γ H 2 S, Z) is a dense subset of B 0. Since Hodge classes on surfaces are algebraic, the Hodge locus is an algebraic subvariety of B 0 ; in basic terms, what we are looking at are curves or algebraic one-cycles) on X that lie on hyperplane sections. 2000 Mathematics Subject Classification. 14D07; 32G20; 14K30. Key words and phrases. Hodge class, locus of Hodge classes, variation of Hodge structure, mixed Hodge module. 1

2 CHRISTIAN SCHNELL We observe that the expected dimension of the Hodge locus is zero. Indeed, a class γ H 2 S, Z) is Hodge exactly when it pairs to zero against every holomorphic two-form on S; because X is a Calabi-Yau threefold, we have h 0 S, Ω 2 S) = h 0 X, Ω 3 XS) ) h 0 X, Ω 3 X) = h 0 X, O X 1) ) 1 = dim B. The number of conditions is the same as the dimension of the parameter space, and the Hodge locus of γ should therefore have a virtual number of points; those numbers are of interest in Donaldson-Thomas theory [KMPS10]. But there are two issues that need to be dealt with: 1) If the Hodge locus actually has finitely many points, one can of course just count them. But there may be components of positive dimension, and before one can use excess intersection theory or some other method) to assign them a number, one has to compactify such components. 2) An obvious idea is to take the closure of the Hodge locus inside the projective space B; but this is not the right thing to do because there are interesting limit phenomena that one cannot see in this way. Example. Here is a typical example. Consider a family of hyperplane sections S t X, parametrized by t, with S t smooth for t 0, and S 0 having a single ordinary double point. In this case, H 2 S t, Z) contains a vanishing cycle γ t, namely the class of an embedded two-sphere with self-intersection number γt 2 = 2. The vanishing cycle is not a Hodge class on S t, but becomes one in the limit. On the one hand, one has the limit mixed Hodge structure, which is pure of weight two in this case; γ t is a Hodge class in this Hodge structure. On the other hand, one can blow up S 0 at the node; the exceptional divisor E P 1 satisfies [E] 2 = 2, and in a sense, [E] is the limit of the γ t as t 0. 2. Statement of the problem. Abstracting from the example above, we now let H be an arbitrary polarized variation of Z-Hodge structure of weight zero, defined on a Zariski-open subset X 0 of a smooth projective variety X. The assumption about the weight is of course just for convenience: if H has even weight 2k, we can always replace it by the Tate twist Hk), which has weight zero. Let F p H denote the Hodge bundles, and let H Z denote the underlying local system. Although it is not strictly necessary for what follows, we shall assume that the polarization form Q: H Q H Q Q0) is defined over Q. Let us first recall the definition of the usual locus of Hodge classes. The local system H Z determines a not necessarily connected) covering space T Z X 0, whose sheaf of holomorphic sections is isomorphic to H Z. The points of T Z are pairs x, h), with h H Z,x a class in the fiber over the point x X 0. Definition. The locus of Hodge classes of H is the set HdgH) = { x, h) T Z h HZ,x F 0 H x is a Hodge class }. We consider the locus of Hodge classes a subset of T Z ) instead of the individual Hodge loci subsets of X 0 ) because it is useful to keep track of the Hodge classes themselves: over any given point, there may be more than one such class, and all of them may be permuted by the monodromy action. On the face of it, HdgH) is just an analytic subset of the complex manifold T Z ; the following remarkable theorem

THE EXTENDED LOCUS OF HODGE CLASSES 3 by Cattani, Deligne, and Kaplan [CDK95] shows that it is actually a countable union of algebraic varieties. Theorem 2.1 Cattani, Deligne, Kaplan). Every connected component of HdgH) is an algebraic variety, finite and proper over X 0. This theorem is one of the best results in Hodge theory. When H comes from a family of smooth projective varieties, the Hodge conjecture predicts that HdgH) should be a countable union of algebraic varieties; the point is that Cattani, Deligne, and Kaplan were able to prove this without assuming the conjecture. Theorem 2.1 shows that every connected component of the locus of Hodge classes can be extended more or less uniquely) to a projective algebraic variety that is finite and proper over X. As in the example of Calabi-Yau threefolds, this suggests that we should look for a natural compactification of HdgH) that also takes into account those integral classes that only become Hodge classes in the limit. The purpose of this paper is to solve that problem with the help of Saito s theory of mixed Hodge modules. The idea is to construct a complex analytic space T Z that extends T Z, and to use it for defining the extended locus of Hodge classes. As far as I know, it was Clemens who first suggested working directly with limits of integral classes; in any case, I learned this idea from him. 3. The case of a Hodge structure. To motivate the construction, let us first look at the case of a single Hodge structure H. We assume that H is polarized and integral of weight zero; we denote the underlying Z-module by H Z ; the polarization by Q; and the Hodge filtration by F H. Let HdgH) = H Z F 0 H be the set of Hodge classes in H. According to the bilinear relations, a class h H Z is Hodge exactly when it is perpendicular under Q) to the space F 1 H; this says that HdgH) is precisely the kernel of the linear mapping ε: H Z F 1 H), h Qh, ). At first, it may seem that ε is not good for much else, because its image is not a nice subset of F 1 H). In fact, the dimension of the vector space F 1 H can be much smaller than the rank of H Z, and so ε will typically have dense image. But it turns out that the restriction of ε to the subset H Z K) = { h H Z Qh, h) K } is well-behaved. The idea of bounding the self-intersection number of the integral classes already occurs in the paper by Cattani, Deligne, and Kaplan. To back up this claim, we have the following lemma; note that the estimate in the proof will play an important role in our analysis later on. Lemma 3.1. The mapping ε: H Z K) F 1 H) is finite and proper, and its image is a discrete subset of the vector space F 1 H). Proof. We have to show that the preimage of any bounded subset of F 1 H) is finite. It will be convenient to measure things in the Hodge norm: if h = p h p, p, with h p, p F p H F p H, is the Hodge decomposition of a vector h H, then its Hodge norm is h 2 H = h p, p 2 H = ) 1) p Q h p, p, h p, p. p p

4 CHRISTIAN SCHNELL Now suppose that h H Z satisfies Qh, h) K and εh) H R; it will be enough to prove that h H is bounded by a quantity depending only on K and R. The assumption on εh) means that Qh, v) R v H for every v F 1 H. If we apply this inequality to the vector v = p 1 1) p h p, p, we find that v 2 H = Qh, v) R v H, and hence that h p, p 2 H = v 2 H R 2. p 1 Because h is invariant under conjugation, it follows that h 2 H h0,0 2 H + 2R2. This leads to the conclusion that h 2 H K + 4R2, because Qh, h) = h 0,0 2 H + p 0 1) p h p, p 2 H K. In particular, there are only finitely many possibilities for h H Z, which means that ε is a finite mapping, and that the image of ε is a discrete subset of F 1 H). 4. The general case. Now let us return to the general case. As in [CDK95], it is not actually necessary to assume that X is projective; we shall therefore consider a polarized variation of Z-Hodge structure H of weight zero, defined on a Zariski-open subset X 0 of an arbitrary complex manifold X. By performing the construction in 3 at every point of X 0, we obtain a holomorphic mapping ε: T Z T F 1 H); here T F 1 H) = Spec Sym F 1 H ) is the holomorphic vector bundle on X 0 whose sheaf of holomorphic sections is F 1 H). The locus of Hodge classes HdgH) is then exactly the preimage of the zero section in T F 1 H). For any rational number K 0, we consider the submanifold T Z K) = { x, h) T Z Q x h, h) K }. It is a union of connected components of the covering space T Z, because the quantity Q x h, h) is obviously constant on each connected component. More or less directly by Lemma 3.1, the holomorphic mapping ε: T Z K) T F 1 H) is finite and proper, with complex-analytic image; moreover, one can show that the mapping from T Z K) to the normalization of the image is a finite covering space. For the details, please consult 18 below. To construct an extension of T Z K) to an analytic space over X, we use the theory of Hodge modules [Sai90]. Let M be the polarized Hodge module of weight dim X with strict support X, canonically associated with H. We denote the underlying filtered left D-module by the symbol M, F M). The point is that M X0 H and F k M X0 F k H; in particular, the coherent sheaf F 1 M is an extension of the Hodge bundle F 1 H to a coherent sheaf of O X -modules. Now consider the holomorphic mapping ε: T Z K) T F 1 M),

THE EXTENDED LOCUS OF HODGE CLASSES 5 where the analytic space on the right-hand side is defined as before as the spectrum of the symmetric algebra of the coherent sheaf F 1 M. We have already seen that ε T Z K) ) is an analytic subset of T F 1 H); since we are interested in limits of integral classes, we shall extend it to the larger space T F 1 M) by taking the closure. The main result of the paper is that the closure remains analytic. Theorem 4.1. The closure of ε T Z K) ) is an analytic subset of T F 1 M). The proof consists of two steps: 1) We reduce the problem to the special case where X \ X 0 is a divisor with normal crossings and H Z has unipotent local monodromy; this reduction is similar to [Sch12a]. 2) In that case, we prove the theorem by a careful local analysis, using the theory of degenerating variations of Hodge structure. In fact, we deduce the theorem from a strengthening of the main technical result of Cattani, Deligne, and Kaplan, which we prove by adapting the method introduced in [CDK95]. Rather than just indicating the necessary changes in their argument, I have chosen to write out a complete proof; I hope that this will make Chapter B useful also to those readers who are only interested in the locus of Hodge classes and the theorem of Cattani, Deligne, and Kaplan. Once Theorem 4.1 is proved, it makes sense to consider the normalization of the closure of ε T Z K) ). The mapping from T Z K) to its image in the normalization is a finite covering space; it can therefore be extended in a canonical way to a finite branched covering by appealing to the Fortsetzungssatz of Grauert and Remmert. Theorem 4.2. There is a normal analytic space T Z K) containing the complex manifold T Z K) as a dense open subset, and a finite holomorphic mapping ε: T Z K) T F 1 M), whose restriction to T Z K) agrees with ε. Moreover, T Z K) and ε are unique up to isomorphism. Since each T Z K) is a union of connected components of the covering space T Z, we can take the union over all the T Z K); this operation is well-defined because of the uniqueness statement in the theorem. In this way, we get a normal analytic space T Z, and a holomorphic mapping ε: T Z T F 1 M) with discrete fibers that extends ε. Now the preimage of the zero section in T F 1 M) gives us the desired compactification for the locus of Hodge classes. Definition 4.3. The extended locus of Hodge classes HdgH) is the closed analytic subscheme ε 1 0) T Z ; by construction, it contains the locus of Hodge classes. 5. The family of hyperplane sections. The construction above can be applied to the family of hyperplane sections of a smooth projective variety of odd dimension. In this case, one has a good description of the filtered D-module M, F ) in terms of residues [Sch12b], and it is possible to say more about the space T Z. The fact that F 1 M is the quotient of an ample vector bundle leads to the following result; it was predicted by Clemens several years ago. Theorem 5.1. The analytic space T Z K) is holomorphically convex. Every compact analytic subset of dimension 1 lies inside the extended locus of Hodge classes.

6 CHRISTIAN SCHNELL 6. Acknowledgements. In writing this paper, I have benefited a lot from a new survey article by Cattani and Kaplan [CK14] that explains the results of [CDK95] in the case n 2. I thank Eduardo Cattani for letting me read a draft version, and for answering some questions. Several years ago, Davesh Maulik asked me about the case of hyperplane sections of a Calabi-Yau threefold; I thank him for many useful conversations, and for his general interest in the problem. Most of all, I thank my former thesis adviser, Herb Clemens, for suggesting that one should study limits of integral classes with the help of residues; as in many other cases, his idea contained the seed for the solution of the general problem. B. Local analysis in the normal crossing case 7. Main result. The purpose of this chapter is to prove the following special case of Theorem 4.1. We shall see later how the general case can be reduced to this one. Theorem 7.1. Let X be a complex manifold, and let X 0 = X \ D be the complement of a divisor D X with normal crossing singularities. Let H be a polarized variation of Z-Hodge structure of weight zero on X 0 whose local monodromy at each point of D is unipotent. Then the closure of the image of the holomorphic mapping ε: T Z K) T F 1 M) is a complex-analytic subspace of T F 1 M). Here is a brief outline of the proof. The assertion is local on X, and unaffected by enlarging the divisor D. We may therefore assume that X = n, with coordinates s 1,..., s n, and that D is the divisor defined by s 1 s n = 0. Denote by H Z the generic fiber of the local system H Z, by Q: H Q H Q Q the pairing on H Q = H Z Z Q giving the polarization, and by N 1,..., N n the logarithms of the unipotent monodromy transformations. Define H Z K) = { h H Z Qh, h) K }. After pulling back to the universal covering space H n, with coordinates z 1,..., z n related by s j = e 2πizj to the coordinates on n ), we get a holomorphic mapping ε: H n H Z K) T F 1 M). The main point is to show the following: Suppose that zm), hm) ) H n H Z K) is a sequence such that s 1 m),..., s n m) are going to zero along a bounded sector, and ε zm), hm) ) remains bounded. Then after passing to a subsequence, hm) is constant and partially monodromy invariant. Roughly speaking, this means that if we let Z n denote the smallest analytic subvariety containing all the points s1 m),..., s n m) ), then hm) is monodromy invariant on Z. As explained in Theorem 9.1 below, such a result quickly leads to a proof of Theorem 7.1. For technical reasons, we prove a slightly more general result. Let Φ: H n H C denote the period mapping associated with the variation of Hodge structure H on ) n. Since εz, h) = 0 if and only if h Φ 0 z), it is reasonable to expect that a bound on εz, h) should control the distance between h and Φ 0 z). In fact, we show that if the sequence ε zm), hm) ) remains bounded in T F 1 M), then hm) bm) mod Φ 0 zm) ). Here bm) H C is a bounded sequence with property that, for some α > 0, every N j bm) is in Oe α Im zjm) ). This uses the description of the D-module M in terms

THE EXTENDED LOCUS OF HODGE CLASSES 7 of Deligne s canonical extension H, and the fact that F 1 M, compared to F 1 H, contains additional sections with poles along the divisors s j = 0. In Theorem 10.4 below, we prove that even under this weaker assumption, a subsequence of hm) is constant and partially monodromy invariant. In the special case when bm) is in Oe α maxj Im zjm) ), this result is due to Cattani, Deligne, and Kaplan [CDK95, Theorem 2.16]. Their proof is an application of the theory of degenerating variations of Hodge structure, especially the multivariable SL2)-orbit theorem [CKS86]. We prove Theorem 10.4 by adapting their method; there are several difficulties, caused by the fact that the sequence bm) is not necessarily going to zero, but these difficulties can be overcome. As in the original, we argue by induction on the dimension of Z; the description of period mappings in [CKS86] lends itself very well to such an approach. A subtle point is that the assumption hm) H Z is needed in many places: it ensures that certain terms that would only be going to zero when hm) H R are actually equal to zero after passing to a subsequence. Rather than giving an abstract description of the proof, I have decided to include in 12) a careful discussion of the special case n = 1. All the interesting features of the general case are present here, but without the added complications of having several nilpotent operators N 1,..., N n and several variables z 1 m),..., z n m). Hopefully, this will help the reader understand the proof of the general case. 8. Local description of the problem. Since Theorem 7.1 is evidently a local statement, we shall begin by reviewing the local description of polarized variations of Hodge structure [Sch73, Kas85, CKS86]. Fortunately, Cattani and Kaplan have written a beautiful survey article, where they describe all the major results [CK89]. Rather than citing the original sources, I will only quote from this article. Let n, with coordinates s = s 1,..., s n ), be the product of n copies of the unit disk; then ) n is the complement of the divisor defined by s 1 s n = 0. Let H be a polarized variation of Z-Hodge structure of weight zero on ) n ; we assume that the underlying local system of free Z-modules H Z has unipotent monodromy around each of the divisors s j = 0. Let H n, with coordinates z = z 1,..., z n ), be the product of n copies of the upper half-plane; the holomorphic mapping H n ) n, z e 2πiz1,..., e 2πizn) makes it into the universal covering space of ) n. If we pull back the local system H Z, it becomes trivial; let H Z denote the free Z-module of its global sections, and Q: H Q H Q Q0) the symmetric bilinear form coming from the polarization on H. By assumption, the monodromy transformation around s j = 0 is of the form e Nj, where N j is a nilpotent endomorphism of H Q = H Z Z Q that satisfies QN j h 1, h 2 ) + Qh 1, N j h 2 ) = 0. It is clear that N 1,..., N n commute. We now review the description of H that results from the work of Cattani, Kaplan, and Schmid. Let Ď denote the parameter space for filtrations F = F H C that satisfy QF p, F q ) = 0 whenever p+q > 0; let D Ď denote the subset of those F that define a polarized Hodge structure on H C = H Z Z C with polarization Q. Recall that Ď is a closed subvariety of a flag variety, and that the so-called period domain D is an open subset of Ď. The variation of Hodge structure H can be lifted to a period mapping Φ: H n D

8 CHRISTIAN SCHNELL which is holomorphic and horizontal. It is known that every element of the cone CN 1,..., N n ) = { a 1 N 1 + + a n N n a1,..., a n > 0 } defines the same monodromy weight filtration [CK89, Theorem 2.3]; we denote the common filtration by W = W N 1,..., N n ). In the limit, H determines another filtration F Ď for which the pair W, F ) is a mixed Hodge structure on H C, polarized by Q and every element of CN 1,..., N n ). According to the nilpotent orbit theorem [CK89, Theorem 2.1], the period mapping is approximated with good bounds on the degree of approximation) by the associated nilpotent orbit Φ nil : H n D, Φ nil z) = e z jn j F. One can use the mixed Hodge structure W, F ) to express Φz) in terms of the nilpotent orbit and additional holomorphic data on n. Denote by g = { X EndH C ) QXh1, h 2 ) + Qh 1, Xh 2 ) = 0 } the Lie algebra of infinitesimal isometries of Q. The mixed Hodge structure W, F ) determines a decomposition of H C with the following properties: H C = I p,q, W w = I p,q, F k = I p,q, p,q p+q w p k A formula for the subspaces I p,q can be found in [CK89, 1.12)]. The decomposition leads to a corresponding decomposition of the Lie algebra g = p,q g p,q, with g p,q consisting of those X that satisfy XI a,b ) I a+p,b+q. In this notation, we have N 1,..., N n g 1, 1 ; moreover, the restriction of Q to the subspace I p,q I p,q is nondegenerate if p = p and q = q, and zero otherwise. The more precise version of the nilpotent orbit theorem [CK89, Theorem 2.8] is that the period mapping of H can be put into the normal form 8.1) Φ: H n D, Φz) = e z jn j e Γs) F, for a unique holomorphic mapping Γ: n p 1 with Γ0) = 0. When we write Γs), it is of course understood that s j = e 2πizj for every j = 1,..., n. The horizontality of the period mapping has the following very useful consequence [CK89, Proposition 2.6]. Proposition 8.2. Let Φz) = e z jn j e Γs) F be the normal form of a period mapping on H n. Then for every j = 1,..., n, the commutator [ Nj, e Γs)] = N j e Γs) e Γs) N j vanishes along the divisor s j = 0. The presentation of the period mapping in 8.1) is also convenient for describing the polarizable Hodge module M that we obtain by taking the intermediate extension of H to n. Here is a brief explanation of how this works. g p,q

THE EXTENDED LOCUS OF HODGE CLASSES 9 Let H, ) be the flat vector bundle on ) n underlying the variation of Hodge structure. The monodromy being unipotent, this bundle admits a canonical extension to a vector bundle H on n, on which the connection has a logarithmic pole along each of the divisors s j = 0 with nilpotent residue [Del70, Proposition 5.2]. Explicitly, for each v H C, the holomorphic mapping 8.3) σ v : H n H C, σ v z) = e z jn j e Γs) v descends to a holomorphic section of H on ) n, and H is the locally free subsheaf of j H generated by all such sections [CK89, 2.2)]. The Hodge bundles F p H extend uniquely to holomorphic subbundles F p H of the canonical extension; concretely, F p H is generated by those sections in 8.3) with v F p. Now let M, F ) denote the filtered D-module underlying M. Then M is simply the D-submodule of j H generated by H. Moreover, the Hodge filtration on M is given by F k M = j 0 F j D n F j k H. It satisfies F j D n F k M F j+k M, and each F k M is a coherent sheaf on n whose restriction to ) n agrees with F k H. This is a translation of Saito s results in [Sai90, 3.10]; note that Saito uses right D-modules. For the purposes of our construction, the important point is that F 1 M has more sections than F 1 H; the following lemma exhibits the ones that we will use. Lemma 8.4. For any vector v F 2, and any index 1 k n, the formula σ v,k z) = e z jn j e Γs) N kv s k defines a holomorphic section of the coherent sheaf F 1 M on n. Proof. It is clear from the description above that σ v z) is a holomorphic section of F 2 M. By [CK89, 2.7)], the horizontality of the period mapping is equivalent to d e z jn j e Γs)) = e n z jn j e Γs) d Γ 1 s) + N j dz j ). Using this identity and the fact that s k = e 2πiz k, we compute that σ v z) = e z jn j e Γs) Γ 1 s) + N ) k v s k s k 2πis k = e z jn j e Γs) Γ 1s) v + 1 s k 2πi σ v,kz). This section belongs to F 1 M by the definition of the filtration; we now obtain the result by noting that Γ 1 s) v is a holomorphic mapping from n into F 1. We close this section by describing the mapping ε: T Z T F 1 M) in coordinates. With the conventions in [CK89, 1.8)], the étalé space T Z of the local system H Z can be obtained as the quotient of H n H Z by the following Z n -action: 8.5) a z, h) = z + a, e a jn j h ) As in the general problem, we define, for any integer K 0, a set j=1 H Z K) = { h H Z Qh, h) K }.

10 CHRISTIAN SCHNELL Then T Z K) is the quotient of H n H Z K) by the action in 8.5). Now F 1 H is a subsheaf of F 1 M, and so we have a commutative diagram 8.6) H n H Z K) T Z K) T F 1 M) ε ε p T F 1 H) The concrete description of F 1 H shows that T F 1 H) n HomF 1, C). We therefore obtain a holomorphic mapping H n H Z n HomF 1, C), which, in coordinates, is given by the formula z, h) s, v Q h, e z jn j e Γs) v )). As usual, the relation s j = e 2πizj is implicit in the notation. 9. Reformulation of the problem. We continue to use the notation introduced in the previous section. Our goal is to deduce Theorem 7.1 from the following more precise local statement. Theorem 9.1. Suppose we are given a sequence of points zm), hm) ) H n H Z K) with x j m) = Re z j m) bounded and y j m) = Im z j m) going to infinity for every j = 1,..., n. If ε zm), hm) ) remains bounded inside T F 1 M), then there is a subsequence with the following properties: a) The sequence hm) is constant and equal to h H Z K). b) One has a 1 N 1 + + a n N n )h = 0 for certain positive integers a 1,..., a n ; in particular, h W 0. c) There is a vector w C n such that lim m e zjm)n j hm) = e w jn j h. d) Each N k h is a rational Hodge class in the mixed Hodge structure W 2, e w jn j F ) 1). Let us prove that Theorem 9.1 implies Theorem 7.1. It suffices to show that the closure of the image of ε: T Z K) T F 1 M) is analytic in a neighborhood of any given point in T F 1 M). After choosing local coordinates, we may therefore assume without loss of generality that X = n and X 0 = ) n, and consider the behavior of the closure over the origin. For every h H Z, we have a holomorphic mapping ε h = ε, h): H n T F 1 M). We only need to prove that the closure of the image of ε h is analytic; this is because, by assertion a) in Theorem 9.1, any bounded subset of T F 1 M) can intersect only finitely many of the sets ε h H n ). Furthermore, we may assume that there is a vector w C n such that each N j h is a rational Hodge class in the mixed Hodge structure W, e w jn j F ) 1), and that a 1 N 1 h+ +a n N n h = 0 for certain

THE EXTENDED LOCUS OF HODGE CLASSES 11 positive integers a 1,..., a n Z; otherwise, the closure of ε h H n ) does not actually contain any points over the origin, according to Theorem 9.1. Under these assumptions on h, we can prove the stronger result that the image of p ε h has an analytic closure; here p: T F 1 M) T F 1 H) denotes the holomorphic mapping induced by F 1 H F 1 M; see 8.6). This suffices to conclude the proof, because the image of ε h is then contained in the closed analytic subset p 1 ) p ε h )H n ). As p is an isomorphism over ) n, it follows that the closure of ε h H n ) is also analytic in fact, it is a connected component of the above set. Proposition 9.2. Let h H Z be an element with N j h W 2 e w jn j F 1 for all j. Then the image of the holomorphic mapping H n n HomF 1, C), z s, v Q h, e z jn j e Γs) v )) has an analytic closure where s j = e 2πizj as usual). Proof. Let H C = p,q Ip,q be Deligne s decomposition of the mixed Hodge structure W, e w jn j F ). Since e z jn j e Γs) = e z j w j)n j e w jn j e Γs) e w jn j ) e w jn j, we may replace F by e w jn j F and e Γs) by the expression in parentheses, and assume without essential loss of generality that w = 0. We then have N j h I 1, 1 for every j = 1,..., n. Under the isomorphism Q: I p,q HomF 1, C) p 1 induced by Q, the linear functional v Q h, e z jn j e Γs) v ) corresponds to e Γs) e ) p,q z jn j h = e h) p,q Γs) + e Γs) e ) z jn j id h. p 1 p 1 Here we have used the fact that N j h I 1, 1 and Γs) p 1 gp,q. Since Γs) is holomorphic on n, it is therefore enough to prove that the image of 9.3) C n C n I p,p, z e 2πiz1,..., e 2πizn, e z jn j id ) ) h p 1 has an analytic closure. This is what we are going to do next. We denote by V the direct sum of the I p,p with p 1; given a vector v V, we write v p,p for its component in the summand I p,p. Let Sh) = { a Z n a j N j h = 0 } ; note that Z n /Sh) embeds into H Q, and is therefore a free Z-module, say of rank r. We can thus find a matrix A SL n Z) whose last n r columns give a basis for the submodule Sh) Z n. If we now introduce new coordinates z 1,..., z n) C n by defining n n z j = a j,k z k and N k = a j,k N j, k=1 j=1

12 CHRISTIAN SCHNELL we have z 1 N 1 + +z n N n = z 1N 1+ +z nn n. The vectors N 1h,..., N rh I 1, 1 are linearly independent, while N r+1h = = N nh = 0. The mapping in 9.3) therefore has the same image as n ) n C n C n V, z e 2πia 1,kz k,..., e 2πia n,kz k, e z k N ) k id h. k=1 Because N 1h,..., N rh are linearly independent, we can find linear functionals ϕ 1,..., ϕ r : I 1, 1 C with the property that k=1 z j = ϕ j z 1N 1h + + z rn rh). Every point s, v) C n V in the image therefore satisfies the polynomial equations ph 9.4) v p, p = 1)p ϕ 1 v 1, 1 )N 1 + + ϕ r v 1, 1 )N p! r) for every p 1. Moreover, one has s j = n r e 2πia j,kz k = e 2πia j,kϕ k v 1, 1) k=1 k=1 n k=r+1 ) e 2πiz aj,k k. Now observe that the closure of the image of the monomial mapping n n C ) n r C n, t r+1,..., t n ) t a 1,k k,..., k=r+1 k=r+1 t a n,k k is a closed algebraic subvariety Z C n, defined by finitely many polynomial equations. In fact, each of these polynomials can be taken as a difference of two monomials, and Z is a possibly not normal) toric variety. For every polynomial fs) in the ideal of Z, we therefore obtain a holomorphic equation ) r r 9.5) f s 1 e 2πia 1,kϕ k v 1, 1),..., s n e 2πia n,kϕ k v 1, 1 ) = 0 k=1 that is satisfied by every point s, v) C n V in the image. Together, 9.4) and 9.5) give a system of holomorphic equations for the closure of the image of the mapping in 9.3), proving that it is indeed an analytic subvariety. 10. Another reformulation of the problem. The purpose of this section is to reduce the proof of Theorem 9.1 to a statement that only involves the variation of Hodge structure H. Suppose then that we are given a sequence of points zm), hm) ) H n H Z K) with the properties listed in Theorem 9.1. Observe that h H Z is a Hodge class in the Hodge structure defined by Φz) if and only if h Φ 0 z) if and only if εz, h) = 0. This suggests that the boundedness of ε zm), hm) ) should allow us to control the distance from hm) to the subspace Φ 0 zm) ). To quantify this idea, we fix an inner product on H C, and let denote the corresponding norm. Definition 10.1. A sequence of vectors bm) H C is called harmless with respect to ym) if there is a positive real number α > 0 such that the quantity n bm) + e αykm) N k bm) remains bounded as m. k=1 k=1 )

THE EXTENDED LOCUS OF HODGE CLASSES 13 In other words, a sequence bm) H C is harmless if and only if it is bounded and N k bm) is in Oe αy km) ) for every k = 1,..., n. We can now turn the idea from above into a precise statement. Proposition 10.2. If the sequence ε zm), hm) ) T F 1 M) is bounded, then 10.3) hm) bm) mod Φ 0 zm) ), for a sequence of vectors bm) H C that is harmless with respect to Im zm). Proof. We are going to use the collection of holomorphic sections σ v z) = e z jn j e Γs) v for v F 1 ) σ v,k z) = e z jn j e Γs) N kv s k for v F 2 ) of the coherent sheaf F 1 M. Define the auxiliary sequence of vectors h m) = e Γsm)) e z jm)n j hm) H C. Using Deligne s decomposition H C = p,q Ip,q of the mixed Hodge structure W, F ), we also define h m) 1 = h m) p,q I p,q. p 1 p 1 We have h m) h m) 1 modulo F 0 = p 0 Ip,q, and therefore 10.3) holds with bm) = e z jm)n j e Γsm)) h m) 1 H C. It remains to show that bm) is harmless with respect to the sequence of imaginary parts ym) = Im zm). By assumption, the sequence of complex numbers ) ) Q hm), σ v zm) = Q hm), e ) z jm)n j e Γsm)) v = Q h m), v ) is bounded for every v F 1. Since the pairing Q is nondegenerate and compatible with Deligne s decomposition, we conclude that h m) 1 is bounded. Likewise, the boundedness of the sequence ) ) Nk h ) m) Q hm), σ v,k zm) = Q s k m), v for every v F 2 implies that N k h m) 1 is in Oe 2πykm) ). Combining both observations, we find that the sequence h m) 1 is harmless with α = 2π). But then bm) is also harmless with α < 2π) by Lemma 11.3 below. To summarize, we have reduced Theorem 9.1 to the following slightly stronger statement. It has the advantage of being expressed purely in terms of H. Theorem 10.4. Suppose we are given a sequence of points zm), hm) ) H n H Z K) with x j m) = Re z j m) bounded and y j m) = Im z j m) going to infinity for every j = 1,..., n. Also suppose that hm) bm) mod Φ 0 zm) ) for a sequence of vectors bm) H C that is harmless with respect to ym). Then there exists a subsequence with the following properties: a) The sequence hm) is constant and equal to h H Z K).

14 CHRISTIAN SCHNELL b) There are positive integers a 1, a 2,..., a n such that a 1 N 1 + + a n N n )h = 0. c) There is a vector w C n such that zjm)n j hm) = e w jn j h. lim m e Here is why Theorem 10.4 implies Theorem 9.1. Let zm), hm) ) be a sequence with the properties described in Theorem 9.1. According to Proposition 10.2, we have hm) bm) mod Φ 0 zm) ) for a harmless sequence bm) H C ; the first three assertions in Theorem 9.1 therefore follow immediately from Theorem 10.4. To prove the fourth one, note that h W 0. Since we also know that Q hm), e z jm)n j e Γsm)) N ) kv s k m) is bounded for every choice of v F 2, we can pass to the limit and obtain 0 = lim Q e ) z jm)n j hm), e Γsm)) N k v = Q e ) w jn j h, N k v. m Now the properties of Q imply that N k h e w jn j F 1. 11. Properties of harmless sequences. This section contains a few elementary results about harmless sequences that will be useful later. First, we prove the following structure theorem; to simplify the notation, we define y n+1 m) = 0. Proposition 11.1. A harmless sequence can always be written in the form bm) = b 0 m) + b 1 m) + + b n m), where b k m) ker N 1 ker N k and b k m) is in Oe αy k+1m) ). In other words, b 0 m) is of size e αy1m) ; b 1 m) is in the kernel of N 1 and of size e αy2m) ; and so on, down to b n m), which is in the kernel of all the N j and bounded. The proof is based on the following simple result from linear algebra. Lemma 11.2. Let T : V V be a linear operator on a finite-dimensional vector space. Then every v V can be written in the form v = v 0 + v 1, where T v 1 = 0 and v 0 C T v, for a constant C that depends only on V, T, and. Proof. Recall that comes from an inner product on V. By projecting to ker T, we get v = v 0 + v 1, with T v 1 = 0 and v 0 ker T. In particular, T v = T v 0. Now T : ker T ) im T is an isomorphism, and therefore has an inverse S. We then get v 0 = ST v) C T v, for a constant C that depends only on V, T, and the choice of norm. Proof of Proposition 11.1. Since N 1,..., N n commute with each other, we can use the lemma and induction. First, we apply the lemma for V = H C and T = N 1 ; this gives bm) = b 0 m) + b m), with N 1 b m) = 0 and b 0 m) in Oe αy1m) ). In the next step, we apply the lemma for V = ker N 1 and T = N 2 to decompose b m), etc. We also need to know that harmless sequences are preserved when we apply certain operators; this fact has already been used during the proof of Proposition 10.2.

THE EXTENDED LOCUS OF HODGE CLASSES 15 Lemma 11.3. Let Φz) = e z jn j e Γs) F be the normal form of a period mapping on H n. If bm) H C is harmless with respect to Im zm), then so are e z jm)n j bm) and e Γsm)) bm), provided that Im z 1 m),..., Im z n m) are going to infinity. Proof. Since the operator e z jn j is polynomial in z 1,..., z n, whereas N j bm) is in Oe α Im zjm) ), it is clear that e z jm)n j bm) is again harmless for a slightly smaller value of α). On the other hand, the operator e Γs) is holomorphic on n, and therefore bounded; moreover, Proposition 8.2 shows that the norm of N j e Γsm)) bm) e Γsm)) N j bm) is bounded by a constant multiple of s j m) = e 2π Im zjm). This is clearly enough to conclude that e Γsm)) bm) is a harmless sequence, too. Next, we consider the case when the sequence hm) belongs to certain subspaces. For any subset J {1,..., n}, we let W J) denote the weight filtration of the cone { } CJ) = a j N j a j > 0 for every j J. j J We would like to know that when hm) W w J), we can also take bm) W w J). This requires the following assumption on the period mapping. Definition 11.4. Let J {1,..., n} be a subset of the index set. We say that Φz) = e z jn j e Γs) F is a nilpotent orbit in the variables {s j } j J if Γs) s j = 0 for every j J; in other words, if Γs) does not depend on the variables {s j } j J. The point is that Γs) then commutes with N j for j J by Proposition 8.2), and therefore preserves the weight filtration W J). Note that nilpotent orbits in the usual sense are the special case when J = {1,..., n}. Lemma 11.5. Suppose that hm) bm) mod Φ 0 zm) ), with bm) H C harmless. If hm) W w J), and if Φz) is a nilpotent orbit in the variables {s j } j J, then one can arrange that bm) W w J) as well. Proof. This follows from the proof of Proposition 10.2. Because of the assumption on Φz), the function Γs) commutes with N j for j J by Proposition 8.2), and therefore preserves the weight filtration W J). Consequently, h m) = e Γsm)) e z jm)n j hm) W w J). Because each N j is a 1, 1)-morphism, W w J) is a sub-mixed Hodge structure of W, F ); it is therefore compatible with Deligne s decomposition, and so h m) 1 = h m) p,q W w J). p 1 At the same time, we clearly have h m) b m) mod F 0, which means that h m) 1 = b m) 1 is a harmless sequence. We conclude as before from Proposition 10.2 that e z jn j e Γs) h m) 1 W w J) is harmless.

16 CHRISTIAN SCHNELL We end this section with a word of caution. During the proof of Theorem 10.4, the fact that bm) H R causes some trouble. If being harmless was preserved by the Hodge decomposition for the Hodge structure Φ zm) ), we could easily arrange that bm) H R. Unfortunately, this is not the case. Example 11.6. Let n = 1, and consider the special case H C = I 1,1 I 1, 1. If b I 1, 1, then Nb = 0, and so b is harmless. Now we decompose with respect to A short calculation gives e iyn F 0 e iyn F 1. b = e iyn N + b 2iy e iyn N + b 2iy = b 2 + N + ) b b + 2iy 2 N + ) b 2iy and neither of the two components is harmless. The best one can say is that, after applying yn, they are bounded; this is consistent with [Sch12a, Proposition 24.3]. 12. Proof in the one-dimensional case. In this section, we prove Theorem 10.4 in the special case n = 1. This case is technically easier, because it avoids the complications coming from the presence of several variables and several nilpotent operators. Because many key features of the proof are the same as in the general case, it may be useful to understand them first in this special case. Suppose then that zm), hm) ) H H Z K) is a sequence of the type considered in Theorem 10.4. Fix an inner product on the space H C, and denote by the corresponding norm. By Proposition 11.1, we can arrange that hm) bm) = b 0 m) + b 1 m) mod Φ 0 zm) ), with b 0 m) in Oe αym) ), and b 1 m) ker N bounded. Our goal is to prove that, after taking a subsequence, hm) is constant and Nhm) = 0. We first introduce some notation. Let W, ˆF ) denote the R-split mixed Hodge structure canonically associated with W, F ) by the SL2)-orbit theorem [CK89, Corollary 3.15]. If Y g R denotes the corresponding splitting, the eigenspaces E l Y ) define a real grading of the weight filtration W, meaning that W k = l k E l Y ). To simplify some of the arguments below, we shall choose the inner product on H C in such a way that this decomposition is orthogonal. The most important tool in the proof will be the following sequence of real operators: ) 1 em) = exp log ym) Y EndH R ) 2 Note that em) acts as multiplication by ym) l/2 on the subspace E l Y ), and preserves the filtration ˆF. Because [Y, N] = 2N, we have 12.1) em)n = 1 ym) Nem). Since the sequence of real parts xm) is bounded, [CK89, Theorem 4.8] shows that 12.2) F = def e in ˆF = lim m em)φ zm) ) D.

THE EXTENDED LOCUS OF HODGE CLASSES 17 The filtration F has two important properties: on the one hand, it belongs to D, and therefore defines a polarized Hodge structure of weight zero on H C ; on the other hand, the pair W, F ) is a mixed Hodge structure. We divide the proof of the theorem in the case n = 1) into six steps; each of the six steps will appear again in a similar form during the proof of the general case. Step 1. We prove that bm) and hm) are bounded with respect to the Hodge norm. This will also show that hm) grows at most polynomially in ym). Lemma 12.3. The two sequences bm) Φzm)) and hm) Φzm)) are bounded. Proof. Recall that the Hodge norm of a vector h H C with respect to the polarized Hodge structure Φz) D is defined as h 2 Φz) = p Z h p, p 2 Φz) = p Z 1) p Q h p, p, h p, p), where h = h p, p is the Hodge decomposition of h in Φz). We begin the proof by observing that the sequence em)bm) = em)b 0 m) + em)b 1 m) is bounded, for the following reason. On the one hand, b 1 m) ker N W 0 implies that b 1 m) = l 0 b 1 m) l l 0 E l Y ); consequently, the boundedness of b 1 m) implies the boundedness of em)b 1 m) = l 0 ym) l/2 b 1 m) l. On the other hand, the term em)b 0 m) is going to zero, because b 0 m) is in Oe αym) ), whereas em) grows at most polynomially in ym). Because em) is a real operator, we have bm) Φzm)) = em)bm) em)φzm)), which is bounded by virtue of 12.2). Now let hm) = p Z hm) p, p denote the Hodge decomposition of hm) in the Hodge structure Φzm)). The difference hm) bm) is an element of Φ 0 zm)), and for p 1, hm) p, p Φzm)) = bm) p, p Φzm)) is bounded. Recalling that hm) H Z K), we now have hm) 2 Φzm)) = Q hm), hm) ) + p 0 1 1) p ) hm) p, p 2 Φzm)) K + 4 bm) 2 Φzm)), and so the Hodge norm of hm) is bounded, too.

18 CHRISTIAN SCHNELL Step 2. Next, we reduce the problem to the case of a nilpotent orbit. Lemma 12.3 gives the boundedness of the sequence em)hm). Since em) 1 is polynomial in ym), it follows that hm) grows at most like a fixed power of ym). We have e zm)n e Γsm)) e zm)n hm) bm) ) e zm)n F 0, and by using Lemma 11.3 and the bounds on hm) and bm), we see that hm) is congruent to a harmless sequence modulo e zm)n F 0. We may therefore assume without loss of generality that Φz) = e zn F is a nilpotent orbit. Note that we only have Φz) D when the imaginary part of z H is sufficiently large; this does not cause any problems because Im zm) is going to infinity anyway. Step 3. We exploit the boundedness of em)hm) to prove that hm) W 0. By passing to a subsequence, we can arrange that there is a limit v = lim m em)hm) H R. With respect to the eigenspace decomposition of Y, we have em)hm) = l Z ym) l/2 hm) l, and so hm) l is going to zero when l 1, and is bounded when l = 0. Let l Z be the largest index such that hm) l 0 along a subsequence. The projection from E l Y ) to gr W l is an isomorphism, and because hm) H Z, it follows that hm) l lies in a discrete subset of E l Y ). This is only possible if l 0, and hence hm) W 0 ; moreover, the component hm) 0 takes values in a finite set. After passing to a subsequence, we may therefore assume that where h 0 E 0 Y ) is constant. hm) h 0 mod W 1, Step 4. We prove that Nh 0 = 0. Consider again the decomposition b 1 m) = l 0 b 1 m) l, where b 1 m) l E l Y ) ker N; note that all summands are bounded, and that only terms with l 0 appear because b 1 m) ker N W 0. Consequently, em)b 1 m) = b 1 m) 0 + ym) l/2 b 1 m) l l 1 has the same limit as b 1 m) 0 E 0 Y ) ker N. If we now look back at em)hm) em)b 0 m) + em)b 1 m) mod em)φ 0 zm) ), we find that all terms converge individually, and hence that v = lim em)hm) lim b 1m) 0 mod F 0. m m To show that Nv = 0, we apply the following version of [CDK95, Proposition 3.10] to the R-split mixed Hodge structure W, ˆF ), recalling that F = e in ˆF. This result is, in a sense, an asymptotic form of Theorem 10.4. Lemma 12.4. Let W, F ) be an R-split mixed Hodge structure on H C, and let N be a real 1, 1)-morphism of W, F ). If a vector h W 2l H R satisfies h b mod e in F l

THE EXTENDED LOCUS OF HODGE CLASSES 19 for some b E 2l Y ) ker N, then h E 2l Y ) ker N. Proof. If we apply e in to both sides and use the fact that Nb = 0, we get e in h E 2l Y ) + W 2l F l. Thus Nh W 2l 2 F l H R, which can only happen if Nh = 0, because W, F ) is a mixed Hodge structure. But then Y h 2lh W 2l 1 F l H R, and so Y h = 2lh for the same reason. Alternatively, one can use the decomposition into the subspaces I p,q = W p+q F p F q, which is preserved by Y. Consequently, v E 0 Y ) ker N. We can now project the congruence em)hm) h 0 mod W 1 to the subspace E 0 Y ) to conclude that v = h 0, and hence that Nh 0 = 0. Step 5. Next, we shall argue that Nhm) = 0. Since Nh 0 = 0, we already know that Nhm) W 3. In addition, we have Nb 1 m) = 0, and so em)nhm) em)nb 0 m) mod Nem)Φ 0 zm) ) ; here we have used 12.1) to interchange N and em). We claim that em)nhm) is bounded by a constant multiple of em)nb 0 m), and therefore in Oe αym) ). If not, then the ratios em)nb 0 m) em)n hm) would be going to zero; after passing to a subsequence, the unit vectors em)n hm) em)nhm) W 3 H R would then converge to a unit vector in W 3 NF 0 H R W 3 F 1 H R ; but this is not possible because W, F ) is a mixed Hodge structure. Consequently, em)nhm) is in Oe αym) ); because em) 1 only grows like a power of ym), the same is true for the norm of Nhm). But these vectors lie in a discrete set, and so Nhm) = 0 after passing to a subsequence. Step 6. To finish the proof, we have to show that hm) is bounded. Choose a point w H with sufficiently large imaginary part to ensure that Φw) = e wn F D; this filtration defines a polarized Hodge structure of weight zero on H C. Since Φz) is a nilpotent orbit, we then have hm) = e wn zm)n hm) e wn zm)n b 0 m) + b 1 m) mod Φ 0 w); note that both terms on the right-hand side are bounded. This relation shows that, with respect to the Hodge structure Φw), all the Hodge components hm) p, p with p 1 are bounded. Because hm) H Z K), we conclude as in Lemma 12.3 that hm) is bounded in the Hodge norm for Φw), and therefore bounded. After passing to a subsequence, the sequence hm) becomes constant. This completes the proof of Theorem 10.4 in the case n = 1.

20 CHRISTIAN SCHNELL 13. Setup in the general case. In the remaining sections, we shall prove Theorem 10.4 in general, by adapting the method in [CDK95, Section 4] to our setting. The proof uses many results from the theory of degenerating variations of Hodge structures; these will be introduced in the appropriate places. Thoughout the discussion, we fix an inner product on H C, and denote by the corresponding norm; we shall make a more specific choice later on. To explain the idea of the proof, let us first consider a sequence zm) H n, with the property that the real parts x j m) = Re z j m) are bounded, and the imaginary parts y j m) = Im z j m) are going to infinity. In the course of the argument, it will often be necessary to pass to a subsequence; to reduce clutter, we shall use the same notation for the subsequence. A new feature of the general case is that we no longer have a unique scale on which we can measure the rate of growth of a sequence; the reason is that y 1 m),..., y n m) may be going to infinity at different rates. The most efficient way to deal with this problem is as follows. Following [CDK95, 4.1.3)], we expand the sequence zm) according to the rate of growth of its imaginary parts. After passing to a subsequence, we can find an integer 1 d n and an n d-matrix A with nonnegative real entries, such that 13.1) zm) = iatm) + wm). Here wm) C n is a convergent sequence, with the property that Φ wm) ) D for every m N; and tm) R d has the property that all the ratios t 1 m) t 2 m), t 2m) t 3 m),..., t d m) t d+1 m) are going to infinity. To avoid having to deal with special cases, we always define t d+1 m) = 1.) Moreover, we can partition the index set {1, 2,..., n} = J 1 J 2 J d in such a way that a j,k 0 if and only if j J 1 J k. By construction, s j m) is in Oe αtkm) ) for every j J k. We define new operators n T k = a j,k N j CJ 1 J k ) j=1 and have the identity n z j m)n j = j=1 d it k m)t k + k=1 n w j m)n j, Now suppose that bm) H C is a harmless sequence with respect to Im zm). By definition, there is some α > 0 such that n bm) + e αyjm) N j bm) j=1 is bounded; it is easy to see that the same is true with a different α > 0) for d bm) + e αtjm) T j bm). k=1 From now on, we shall use the expression harmless or harmless with respect to tm) to refer to sequences with this property. j=1