1 Introduction
Suppose that C is a smooth complex projective curve of genus
$g\ge 2$
. For each point x of C, the Abel–Jacobi mapping

embeds C into its jacobian. Its image is an algebraic 1-cycle in
$\operatorname {Jac} C$
that we denote by
$C_x$
. Its image under the involution

is another algebraic 1-cycle,
$\iota _\ast C_x$
, that we denote by
$C_x^-$
. The Ceresa cycle associated to
$(C,x)$
is the algebraic 1-cycle

in
$\operatorname {Jac} C$
. Ceresa [Reference Ceresa1] proved that when C is a general curve of genus
$g\ge 3$
, the Ceresa cycle is not algebraically equivalent to 0.
A standard approach to studying the Ceresa cycle is to let C vary in the moduli space
${\mathcal M}_g$
of smooth projective curves of genus g and to study
$\nu $
, the associated normal function

Here
${\mathbb H}$
denotes the variation of Hodge structure over
${\mathcal M}_g$
whose fibre over the moduli point of C is
$H_1(C;{\mathbb Z})$
and
$J(\Lambda ^3_0 {\mathbb H})$
denotes the family of intermediate jacobians of the primitive degree three homology of
$\operatorname {Jac} C$
:

where
$\times $
denotes the Pontryagin product.
The rank of a normal function is a measure of its nontriviality. It is defined in detail in Section 3 and is closely related to Griffiths’ infinitesimal invariant of a normal function. Briefly, if one ignores the complex structure, every family of intermediate jacobians is naturally a locally constant family of real tori. This means that, locally, a normal function can be regarded as a smooth function to a real torus. Its derivative is
$\mathbb C$
-linear. The rank of
$\nu $
is the value of the complex rank of this function at a general curve. When
$g=2$
,
$\nu $
is identically 0 and thus has rank 0.
Theorem 1. For all
$g\ge 3$
, the normal function
$\nu $
of the Ceresa cycle has the maximum possible rank, namely
$\dim {\mathcal M}_g$
.
For each pointed curve
$(C,x)$
there is also the Gross–Schoen cycle [Reference Gross and Schoen11]. It is a homologically trivial algebraic 1-cycle in
$C^3$
. Its normal function is three times the normal function of the Ceresa cycle. Ziyang Gao and Shou-Wu Zhang [Reference Gao and Zhang6] have proved a stronger version of Theorem 1 which they use to prove that, for all
$g\ge 3$
, there is a nonempty Zariski open subset of
${\mathcal M}_g/{\mathbb Q}$
on which the Bloch–Beilinson height of the Gross–Schoen cycle has the Northcott property. Our proof of the theorem is very different from theirs. Whereas they use Ax–Schanuel, we proceed by studying the behaviour of the Ceresa normal function near the boundary of
${\mathcal M}_g$
using its associated monodromy representation.
Theorem 1 implies that there is a proper real analytic subvariety
$\Sigma $
of
${\mathcal M}_g$
such that
$\nu $
has maximal rank on the complement U of
$\Sigma $
. As Gao and Zhang observe via a spreading argument [Reference Gao and Zhang6, Prop. C.7], Theorem 1 implies that if C is a smooth projective curve whose moduli point lies in
$U\setminus {\mathcal M}_g(\overline {{\mathbb Q}})$
, then the Ceresa and Gross–Schoen cycles associated to C both have infinite order mod rational equivalence. This is a weaker version of their [Reference Gao and Zhang6, Thm. 1.3].
The theorem is proved by induction on the genus. The base case
$g=3$
is proved in Section 8 using a result [Reference Collino and Pirola3] of Collino and Pirola. (See Theorem 8.4.) The result is not as explicit as we had hoped. However, as a consolation prize, we obtain an explicit formula for the Green–Griffiths invariant of the genus 3 Ceresa cycle as a Teichmüller modular form, as well as a new proof of the theorem of Collino and Pirola. The inductive step is proved by studying the behaviour of the Ceresa normal function in a neighbourhood of the boundary divisor
$\Delta _0$
in the Deligne–Mumford compactification
$\overline {{\mathcal M}}_g$
of
${\mathcal M}_g$
by computing the monodromy representation of the restriction of the Ceresa normal function to this neighbourhood.
Theorem 2. In genus 3, the Green–Griffiths invariant of the Ceresa cycle is a nonzero multiple of the Teichmüller modular form
$\chi _{4,0,-1}$
. Its restriction to the hyperelliptic locus is a nonzero multiple of the restriction of the Siegel modular form
$\chi _{4,0,8}$
to the hyperelliptic locus.
For a more precise statement, see Theorem 10.2. It would be interesting to compute the multiple, which one might expect to lie in
${\mathbb Q}^\times $
.
Since the restriction of
$\chi _{4,0,8}$
to the hyperelliptic locus has no zeros [Reference van der Geer, Kouvidakis and Alexis7, Prop. 6.5], we obtain the following strengthening of the genus 3 case of a result [Reference Harris18, Thm. 6.5] of Bruno Harris.
Corollary 3. In genus 3, the rank of the Ceresa normal function is exactly 1 at every point of the hyperelliptic locus.
The proof of Theorem 2 makes essential use of the extension [Reference Cléry, Faber and van der Geer2] by Cléry, Faber and van der Geer of Ichikawa’s theory [Reference Ichikawa19] of Teichmüller modular forms in genus 3. It also gives a geometric interpretation of the meromorphic Teichmüller modular form
$\chi _{4,0,-1}$
, which has a simple pole along the hyperelliptic locus, as a multiple of the Green–Griffiths invariant of the normal function of the Ceresa cycle. Although it is a meromorphic section of an automorphic vector bundle, it is holomorphic when considered as a section of the tensor product of
$\Omega _{{\mathcal M}_3}^1$
with an automorphic vector bundle as we explain in detail in Section 10. This suggests that the definition of vector-valued Teichmüller modular forms should be enlarged to include sections of tensor products of automorphic vector bundles with the logarithmic de Rham complex of moduli spaces of curves.
As a corollary of the proof of Theorem 2, we obtain a new proof of the result of Collino and Pirola used in the proof of Theorem 1 and the strengthening of Harris’s result stated above. These results suggest that the Green–Griffiths invariant of the Ceresa normal function should be a Teichmüller modular form for all genera.
The idea of the inductive step in the proof of Theorem 1 is to study the behaviour of the Ceresa normal function
$\nu $
in a neighbourhood of the boundary divisor
$\Delta _0$
of the Deligne–Mumford moduli space
$\overline {{\mathcal M}}_g$
. For this we introduce the notion of the residual normal function associated to a normal function defined on the complement of a divisor. It is a normal function defined on the smooth locus of the divisor. In the case of the Ceresa normal function, the residual normal function is defined on the smooth locus of
$\Delta _0$
. We are able to compute it by computing its monodromy representation and show, using the inductive hypothesis, that it has rank
$3g-4$
. To complete the proof, we show that the normal rank of
$\nu $
along the smooth locus of
$\Delta _0$
is 1 using a result [Reference Hain17] about real biextensions and a monodromy computation.
1.1 Overview
The proof of Theorem 1 requires an understanding of the limiting behaviour of admissible variations of MHS (as does the proof of Gao and Zhang). It also requires some background in the topology of moduli spaces of curves, such as the fact that normal functions over them are determined by their monodromy representations when the genus is at least 3. For this reason, in an attempt to make the paper accessible to a reasonably broad audience, we have included background material on both of these topics.
Part 1 is a brief review of the Griffiths infinitesimal invariant
$\delta (\nu )$
of a normal function
$\nu $
and Green’s refinement
$\overline {\delta }(\nu )$
of it. Both vary holomorphically. In general, the derivative
$\nabla \nu _{\mathbb R}$
of the normal function varies real analytically, but not holomorphically. This is the case with the Ceresa normal function. Since
$\nabla \nu _{\mathbb R}$
is real, it is determined by its
$(1,0)$
component, which we call the canonical derivative of
$\nu $
and denote by
$\overline {\nabla } \nu $
. It determines both the rank of
$\nu $
and its Green–Griffiths invariant. It plays a fundamental role in this paper. An obstruction to obtaining an explicit formula for the rank of the Ceresa normal function at a genus 3 curve is that
$\overline {\nabla } \nu $
does not vary holomorphically.
In Part 2 we present general results about the Green–Griffiths invariant and canonical derivative of the Ceresa normal function. Part 3 focuses on the genus 3 case. One key observation is that in genus 3 the canonical derivative
$\overline {\nabla } \nu $
can be decomposed naturally

where f is real analytic and
$\overline {\delta }(\nu )$
is the Green–Griffiths invariant. Away from the hyperelliptic locus, the two components are sections of different automorphic vector bundles. At the moduli point of a nonhyperelliptic curve C, both components can be interpreted as quadratic forms on the tangent space of
${\mathcal M}_3$
at the moduli point of C. Using the fact that both components are invariant under
$\operatorname {Aut} C$
, we show that
$\overline {\nabla } f$
vanishes at the Klein quartic and, using the Collino–Pirola Theorem, that the other component has maximal rank. This establishes the base case of Theorem 1. This part concludes with the proof of Theorem 2 and related results.
Admissible normal functions are, by definition [Reference Hain12, Reference Saito25], period mappings of certain variations of mixed Hodge structure. Normal functions associated to homologically trivial algebraic cycles are admissible. This imposes strong conditions on their boundary behaviour. These restrictions play a key role in the inductive step of the proof of Theorem 1 and also in [Reference Gao and Zhang6]. Part 4 contains a review of admissible variations of MHS. We give a quick construction of the Néron model of a family of intermediate jacobians in the special case where the variation is a family of one-dimensional nilpotent orbits. We also introduce the notion of the residual normal function
$\nu _\Delta $
associated to an admissible normal function
$\nu $
defined on
$S-\Delta $
, where
$\Delta $
is a smooth divisor in the smooth variety S, and explain how its rank is related to that of
$\nu $
.
In Part 5 we apply the results in Part 4 to prove the inductive step. We use the fact that normal functions over
${\mathcal M}_{h,n}$
associated to variations corresponding to irreducible representations of
$\mathrm {Sp}_h$
are determined by their monodromy representations when
$h \ge 3$
. (See Appendix A.) This allows us to compute the monodromy of the Ceresa normal function in a neighbourhood of the smooth locus
$\Delta $
of the boundary divisor
$\Delta _0$
of
$\overline {{\mathcal M}}_g$
as well as the residual normal function
$\nu _\Delta $
. The inductive hypothesis then implies that
$\nu _\Delta $
has maximal rank
$3g-4$
. A result [Reference Hain17] about real biextensions associated to curves then implies that the ‘normal rank’ of
$\nu $
along
$\Delta $
is 1, which establishes the result.
1.2 Conventions
All algebraic varieties and stacks will be defined over the complex numbers. The moduli spaces of curves
${\mathcal M}_g$
and of principally polarized abelian varieties
$\mathcal A_g$
will be regarded as stacks.
Part 1. Preliminaries
This part is a review of some basic facts about normal functions, such as the construction of normal functions associated to families of homologically trivial cycles and the construction of the Griffiths and Green–Griffiths invariants of normal functions. The canonical derivative and the rank of a normal function are defined in Section 3.
2 Normal functions and families of algebraic cycles
In this section, we review and elaborate on work of Griffiths [Reference Griffiths9] and Green [Reference Green8] on invariants of normal functions. We also recall the definition of the canonical foliation of a family of intermediate jacobians associated to a variation of Hodge structure of weight
$-1$
.
2.1 The MHS associated to a homologically trivial cycle
Suppose that X is a smooth projective variety and that
$Z= \sum _j n_j Z_j$
is an algebraic d-cycle on X where the
$Z_j$
are distinct reduced irreducible subschemes of X of dimension d. Denote the support of Z by
$|Z|$
. When Z is homologically trivial, it determines an extension

of mixed Hodge structures. It is obtained from the long exact homology sequence of
$(X,|Z|)$
by pulling back along the map

that takes
$1$
to
$[Z] = \sum n_j[Z_j]$
and then twisting by
${\mathbb Z}(-d)$
:

The extension depends only on the rational equivalence class of Z. The extension
$(E_Z)_{\mathbb Z}$
is generated by
$H_{2d+1}(X;{\mathbb Z})$
and
$\Gamma $
, where
$\partial \Gamma = Z$
.
2.2 Extensions of mixed Hodge structure and intermediate jacobians
Suppose that V is a Hodge structure of negative weight. The group of extensions of mixed Hodge structure (MHS) of the form

forms a group
$\operatorname {Ext}^1_{\mathsf {MHS}}({\mathbb Z},V)$
that is isomorphic to the complex torus

The extension determines, and is determined by, the image of

in
$J(V)$
, where
$\mathbf {e}_{\mathbb Z} \in E_{\mathbb Z}$
and
$\mathbf {e}_F \in F^0 E$
both project to
$1\in {\mathbb Z}$
.
The torus
$J(V)$
is compact when V has weight
$-1$
. This is because, in this case,

This implies that the composite

is an isomorphism of real vector spaces which, in turn, induces an isomorphism of
$J(V)$
with the compact real torus
$V_{\mathbb R}/V_{\mathbb Z}$
.
So a homologically trivial d-cycle Z in a smooth projective variety X determines a point
$\nu _Z$
in the Griffiths intermediate jacobian


where, as above,
$\partial \Gamma = Z$
.
Remark 2.1. If a group G acts on X, then it also acts on the intermediate jacobian
$J(H_{2d+1}(X)(-d))$
. For each
$\gamma \in G$
, the point
$\gamma _\ast \nu _Z$
is the point determined by the cycle
$\gamma _\ast Z$
in X. So if
$\gamma _\ast Z$
is rationally equivalent to Z for all
$\gamma \in G$
, then
$\nu _Z$
is fixed by G.
2.3 Normal functions
Suppose
$\mathbb V$
is a variation of Hodge structure of negative weight over a smooth variety S. Set

It is a flat holomorphic vector bundle over S with connection

Denote the pth term of its Hodge filtration by
$F^p\mathcal V$
. The connection satisfies Griffiths transversality

To such a variation, we can associate a family of intermediate jacobians
$J(\mathbb V)$
over S. It is the quotient of the vector bundle
$\mathcal V/F^0\mathcal V$
by the image of
$\mathbb V_{\mathbb Z} \to \mathcal V/F^0\mathcal V$
. Its fibre over
$s \in S$
is
$J(V_s)$
, the intermediate jacobian associated to the fibre
$V_s$
of
$\mathbb V$
over s.
Each extension

of variations of mixed Hodge structure gives rise to a holomorphic section
$\nu _{{\mathbb E}}$
of the bundle

of intermediate jacobians. Its value over
$s\in S$
is the point of
$J(V_s)$
that corresponds to the extension

of MHS obtained by restricting (2) to s.
Every continuous section
$\sigma $
of
$J(\mathbb V)$
determines a class
$c(\sigma )$
in
$H^1(S,\mathbb V_{\mathbb Z})$
. A detailed description of
$c(\sigma )$
can be found in [Reference Hain16, §4.1]. Since
$H^1(S,\mathbb V_{\mathbb Z})$
is the group of congruence classes of extensions of
${\mathbb Z}_S$
by
$\mathbb V_{\mathbb Z}$
in the category of local systems over S, each section
$\sigma $
of
$J(\mathbb V)$
determines an extension of local systems

over S. The value of
$\sigma $
at
$s\in S$
determines a MHS on the fibre of
${\mathbb E}$
over s.
Definition 2.2. A section of
$J(\mathbb V) \to S$
is a normal function if the corresponding extension is a variation of MHS.
In other words, the group of normal function sections of
$J(\mathbb V)$
is isomorphic to the group of extensions of
${\mathbb Z}_S$
by
$\mathbb V$
in the category of variations of MHS over S. Normal functions have to be holomorphic. This corresponds to the condition that the Hodge filtration varies holomorphically. They also have to satisfy the Griffiths infinitesimal period relation, which is equivalent to the condition that the canonical flat connection on

satisfies Griffiths transversality,
$\nabla : F^p \mathcal E \to F^{p-1} \mathcal E \otimes \Omega ^1_S$
.
Remark 2.3. The variations of MHS that arise in algebraic geometry satisfy additional conditions. Such variations are called admissible variations of MHS. The normal functions that correspond to admissible variations of MHS are called admissible normal functions. The additional conditions on an admissible normal function
$\nu $
restrict the asymptotic behaviour of
$\nu (s)$
as s approaches the boundary of S. More precisely, the extension has to satisfy conditions which ensure that limit mixed Hodge structures exist. Admissible variations of MHS are defined in Section 11.
2.4 Families of homologically trivial cycles
Suppose now that S is a smooth variety and that
$f : X \to S$
is a family of smooth projective varieties. Suppose that Z is a relative d-cycle. That is, it is an algebraic cycle on X whose restriction
$Z_s$
to each fibre
$X_s := f^{-1}(s)$
over
$s\in S$
is a d-cycle. Suppose that each
$Z_s$
is homologically trivial in
$X_s$
. Denote the local system over S whose fibre over
$s\in S$
is
$H_{2d+1}(X_s)(-d)$
by
$\mathbb V$
. It underlies a variation of Hodge structure of weight
$-1$
. The construction in Section 2.1 gives an extension of local systems

and a mixed Hodge structure on the fibre
$E_s$
over
$s\in S$
that is an extension of
${\mathbb Z}$
by
$V_s$
. The work of Griffiths [Reference Griffiths9] implies that the corresponding section
$\nu _Z : S \to J(\mathbb V)$
is holomorphic and satisfies his infinitesimal period relation. It is therefore an extension of variations of MHS. The corresponding section
$\nu _Z$
of
$J(\mathbb V)$
is thus a normal function. The main result of [Reference Steenbrink and Zucker28] implies that it is admissible.
2.5 The foliation of
$J(\mathbb V)$
Suppose now that
$\mathbb V$
has weight
$-1$
. This implies that the map

is an isomorphism of families of tori over S. The family
$J(\mathbb V_{\mathbb R})$
is a locally constant family of tori and is thus foliated by its locally constant leaves. Consequently, the family
$J(\mathbb V) \to S$
is a foliated family of complex tori. This foliation
$\mathscr F$
is a complex foliation, but not a holomorphic foliation. This means that each leaf of
$\mathscr F$
is a complex submanifold of
$J(\mathbb V)$
but that
$\mathscr F$
is not, in general, locally biholomorphic to a product foliation.
Definition 2.4. A section of
$J(\mathbb V)$
is locally constant if its image lies in a leaf of
$\mathscr F$
or, equivalently, it lies in the image of
$H^0(S,\mathbb V_{\mathbb R}) \to H^0(S,J(\mathbb V))$
.
2.6 The Griffiths infinitesimal invariant
Griffiths’ infinitesimal invariant is an obstruction to a normal function being locally constant. Suppose that
$\nu $
is a normal function section of
$J(\mathbb V)$
. Let

be the corresponding variation of MHS over S. Suppose that
$s \in S$
and that U is an open neighbourhood of s that is biholomorphic to a complex ball. Since U is Stein, the restriction of each of the bundles
$F^p \mathcal E$
to U is holomorphically trivial, and since U is contractible, the restriction of
${\mathbb E}_{\mathbb Z}$
to U is trivial. Let

be lifts of
$1 \in {\mathbb Z}$
. Then

is a lift of the restriction of
$\nu $
to U. Since
$\mathbf {e}_F$
is a local section of
$F^0\mathcal E$
, Griffiths transversality implies that

Since
$\mathbf {e}_F$
is well defined mod
$H^0(U,F^0\mathcal V)$
, this descends to a well defined element
$\delta _U(\nu )$
of

The functor
$U \mapsto \mathcal K(U)$
is a presheaf on S.
This obstruction group has a more intrinsic description in terms of the Hodge filtration of
$\mathcal V\otimes \Omega ^{\bullet }_S$
, which is defined by

Each
$(F^p(\mathcal V\otimes \Omega _S^{\bullet }),\nabla )$
is a complex of locally free complex analytic sheaves on S. Denote their cohomology sheaves by
$\mathcal H^{\bullet }(F^p(\mathcal V\otimes \Omega _S^{\bullet }))$
. Since
$\mathcal H^1(F^0(\mathcal V\otimes \Omega _S^{\bullet }))$
is the sheafification of
$\mathcal K$
, the germs of the
$\delta _U(\nu )$
at each
$s\in S$
define a section

This is the Griffiths invariant of
$\nu $
. It vanishes if and only if
$\nu $
is locally constant on S.
2.7 Green’s refinement of the Griffiths invariant
Green’s refinement [Reference Green8, §1] of Griffiths construction is a useful tool for understanding the Griffiths invariant. A key point is that each graded quotient

of the Hodge filtration of
$\mathcal V\otimes \Omega _S^{\bullet }$
is a complex of vector bundles as the differential
$\overline {\nabla }$
induced by the connection
$\nabla $
is
${\mathcal O}_S$
-linear.
Denote the homology sheaves of the complex (4) by
$\mathcal H^{\bullet }(\operatorname {Gr}_F^p(\mathcal V\otimes \Omega _S^{\bullet }))$
. Suppose that
$\nu $
is a normal function. Denote the image of
$\delta (\nu )$
under the map

by
$\overline {\delta }(\nu )$
. In favourable situations, Green’s invariant
$\overline {\delta }(\nu )$
determines the Griffiths invariant.
Proposition 2.5 (Green).
If
$\mathcal H^1(\operatorname {Gr}_F^p(\mathcal V\otimes \Omega _S^{\bullet }))$
vanishes for all
$p>0$
, then the map (5) is injective, so that
$\delta (\nu )$
vanishes if and only if
$\overline {\delta }(\nu )$
vanishes.
Proof. Green proves this using a spectral sequence argument. Here we give an elementary direct proof. Since

is exact, so is

The vanishing of
$\mathcal H^1(\operatorname {Gr}_F^1(\mathcal V\otimes \Omega _S^{\bullet }))$
implies the injectivity of

Similarly, the vanishing of
$\mathcal H^1(\operatorname {Gr}_F^2(\mathcal V\otimes \Omega _S^{\bullet }))$
and the exactness of

implies the injectivity of

One continues in this manner to obtain injections

for all
$p>0$
. Since

for m sufficiently large, the composition of the injections (6)

is injective. The result follows by taking global sections.
3 The rank of a normal function
As above,
$\mathbb V$
is a variation of Hodge structure of weight
$-1$
over S and
$\nu : S \to J(\mathbb V)$
is normal function.
3.1 Definition of the rank
We can regard the foliation
$\mathscr F$
of
$J(\mathbb V)$
(Section 2.5) as a smooth sub-bundle of the tangent bundle of
$J(\mathbb V)$
. Since the leaves of the foliation are complex submanifolds of
$J(\mathbb V)$
, we can regard
$\mathscr F$
as a complex sub-bundle of
$TJ(\mathbb V)$
. The projection
$J(\mathbb V) \to S$
induces an isomorphism
$\mathscr F_v \to T_s S$
for all
$v\in J(V_s)$
. This gives a splitting of the sequence

Let
$\phi : TJ(\mathbb V) \to \mathcal V/F^0$
be the corresponding projection. It is
$\mathbb C$
-linear. Composing it with
$(d\nu )_s$
gives a
$\mathbb C$
-linear map
$\phi \circ (d\nu )_s : T_s S \to V_s/F^0 V_s$
:

Definition 3.1. The rank
$\operatorname {rk}_s\nu $
of
$\nu $
at
$s\in S$
is the rank of
$\phi \circ (d\nu )_s$
. Define

The rank of
$\nu $
vanishes if and only if its Griffiths invariant
$\delta (\nu )$
vanishes. In general, the relation between the Griffiths invariant and the rank is more subtle.
3.2 The canonical derivative and the Green–Griffiths invariant
In order to relate the rank of a normal function to the rank of the associated invariants defined by Griffiths and Green, we need to consider real lifts of normal functions.
We will work with complexified tangent spaces as we need to compare invariants of real analytic and holomorphic lifts of a normal function. In this subsection, and only this subsection, we denote the real tangent space of a complex manifold Y at
$y\in Y$
by
$T_y Y$
. It has a canonical almost complex structure
$I : T_y Y \to T_y Y$
. Its complexification decomposes

into the
$I=i$
and
$I=-i$
eigenspaces. The
$I=i$
eigenspace
$T_y' Y$
is the holomorphic tangent space. An
${\mathbb R}$
-linear map
$T_y Y \to V$
into a complex vector space V is complex linear if and only if the induced map
$T_y Y\otimes \mathbb C \to V$
vanishes on
$T_yY"$
. The inclusion of
$T_y Y$
into
$T_y Y\otimes \mathbb C$
induces a complex linear isomorphism

Definition 3.2. The canonical derivative
$\overline {\nabla }_s \nu $
of
$\nu $
at s is the composite

Griffiths transversality implies that its image lies in the subspace
$\operatorname {Gr}^{-1}_F V_s$
of
$V_s/F^0$
. The function
$\overline {\nabla } : s \mapsto \overline {\nabla }_s\nu $
is a real analytic section of
$\operatorname {Gr}^{-1}_F \mathcal V$
.
Denote the sheaf of real analytic functions on S by
$\mathscr E_S$
. The normal function
$\nu $
corresponds to a real analytic section
$\nu _{\mathbb R}$
of
$J(\mathbb V_{\mathbb R})$
under the isomorphism
$J(\mathbb V_{\mathbb R}) \cong J(\mathbb V)$
. Locally it lifts to a section
$\tilde {\nu }_{\mathbb R}$
of
$\mathbb V_{\mathbb R}\otimes \mathscr E_S$
. Its derivative
$(\nabla \tilde {\nu }_{\mathbb R})_s \in V_{s,{\mathbb R}}\otimes T_s^\vee S$
at
$s\in S$
does not depend on the choice of
$\tilde {\nu }_{\mathbb R}$
, so we will denote it by
$(\nabla \nu _{\mathbb R})_s$
.
The goal of the next result is to clarify the relationship between the various derivatives of
$\nu $
.
Proposition 3.3. The canonical derivative
$\overline {\nabla }_s\nu $
is related to the derivatives of
$\nu $
and
$\nu _{\mathbb R}$
at
$s\in S$
by the commutative diagram

In particular,

Proof. This is an exercise in the definitions using the fact that
$d\nu _{\mathbb R}$
is
$\mathbb C$
linear.
Corollary 3.4. The projection of the
$(1,0)$
component
$\nabla '\nu _{\mathbb R}$
of the derivative
$\nabla \nu _{\mathbb R}$
to a section of
$(\mathcal V/F^0\mathcal V)\otimes \mathscr E_S\otimes \Omega ^1_S$
equals
$\overline {\nabla } \nu \in H^0(S,\operatorname {Gr}_F^{-1}\mathcal V \otimes \Omega _S^1\otimes \mathscr E_S)$
.
It is important to note that, in general,
$\overline {\nabla } \nu $
is a real analytic, but not holomorphic, section of
$\operatorname {Gr}_F^{-1}\mathcal V \otimes \Omega _S^1$
, even though
$\overline {\delta }(\nu )$
, its reduction mod
$\overline {\nabla } \operatorname {Gr}_F^0 \mathcal V \otimes \mathscr E_S$
, is a holomorphic section of
$\operatorname {Gr}_F^{-1}\mathcal V \otimes \Omega _S^1/\overline {\nabla }\operatorname {Gr}_F^0\mathcal V$
. This will be important in Section 10 where the possibility that
$\overline {\nabla } \nu $
is not holomorphic is the obstruction to obtaining an explicit formula for
$\overline {\nabla }\nu $
in genus 3.
3.3 Symmetry
Suppose that S is a complex analytic variety and that
$\mathbb V \to S$
is a variation of Hodge structure of weight
$-1$
. Suppose that G is a finite group that acts holomorphically on S and that this action lifts to an action on
$\mathbb V \to S$
as a variation of Hodge structure. If
$s \in S$
is a fixed point of the action, then G acts on
$T_s S$
and as an automorphism group of the Hodge structure
$V_s$
and its intermediate jacobian
$J(V_s)$
. If
$\nu $
is a G-invariant normal function section of
$J(\mathbb V)$
, then
$\nu (s) \in J(V_s)$
will be a fixed point of the G action.
3.4 The rank filtration
We can filter S

by the strata
$S_r$
where
$\nu $
has rank
$\le r$
, where
$d=\dim S$
. Since
$\nu _{\mathbb R}$
is real analytic, each
$S_r$
is a real analytic subvariety of S. Gao and Zhang [Reference Gao and Zhang6, Thm. 1.4] have proved a stronger result. They define the foliation filtration
$S_{\mathscr F}$
of S to be the union of the leaves of the ‘pullback foliation’
$\nu ^\ast \mathscr F$
of codimension
$\le r$
and show that each
$S_{\mathscr F}(r)$
is a complex algebraic subvariety of S. The rank condition implies that
$S_r$
contains
$S_{\mathscr F}(r)$
for all r.
The following much weaker density result is sufficient for our purposes.
Proposition 3.6. If there is a point
$x_0 \in S$
where
$\operatorname {rk}_{x_0} \nu = \dim S$
, then

is a real analytic subvariety of S of real codimension
$\ge 1$
. In particular,
$S-\Sigma $
is dense in S.
Part 2. The Ceresa cycle
In this part we recall the construction of the normal function of the Ceresa cycle and discuss generalities related to its Green–Griffiths invariant and canonical derivative.
4 The normal function of the Ceresa cycle
Suppose that
$g\ge 3$
. The universal curve
$f : \mathcal C \to {\mathcal M}_{g,1}$
has a tautological section that takes the moduli point
$[C,x]$
of
$(C,x)$
to the point
$x \in C \subset \mathcal C$
. This allows us to embed
$\mathcal C$
into the universal jacobian
$\mathcal J := \operatorname {Jac}_{\mathcal C/{\mathcal M}_{g,1}}$
via a map
$\mu _x : \mathcal C \to \mathcal J$
whose restriction to the fibre over
$[C,x]$
is the Abel–Jacobi map (1):

The relative Ceresa cycle
$\mathcal Z_x$
in
$\mathcal J$
is

where
$\iota : u \mapsto -u$
is the involution of
$\mathcal J$
. Its restriction to the fibre over
$[C,x] \in {\mathcal M}_{g,1}$
is the Ceresa cycle
$Z_{C,x}$
in
$\operatorname {Jac} C$
.
Set

Poincaré duality implies that it has fibre
$H_1(C;{\mathbb Z})$
over
$[C,x]$
. The variation
$\Lambda ^3 {\mathbb H}$
has fibre
$H_3(\operatorname {Jac} C)$
over
$[C,x]$
. The normal function
$\nu _x$
associated to
$\mathcal Z_x$
is an admissible section of
$J(\Lambda ^3 {\mathbb H}(-1))$
over
${\mathcal M}_{g,1}$
.
Pontryagin product with the class of C defines an inclusion
$H_1(\operatorname {Jac} C) \hookrightarrow H_3(\operatorname {Jac} C)(-1)$
of Hodge structures and therefore an inclusion
$\operatorname {Jac} C \hookrightarrow J(H_3(\operatorname {Jac} C)(-1))$
. Pulte [Reference Pulte24] showed that if
$x,x' \in C$
, then

so that the image of
$\nu _x(C)$
in
$J(PH_3(\operatorname {Jac} C)(-1))$
does not depend on the choice of the base point
$x\in C$
. Set

Since
$\mathcal J = J({\mathbb H})$
, Pulte’s result implies that the normal function
$\nu _x$
descends to a normal function section of
$J(\Lambda ^3_0 {\mathbb H}(-1))$
over
${\mathcal M}_g$
:

Remark 4.1. Since
${\mathbb H}$
is defined on
${\mathcal M}_g^c$
, the normal functions
$\nu _x$
and
$\nu $
extend to normal functions over the moduli spaces
${\mathcal M}_{g,1}^c$
and
${\mathcal M}_g^c$
of curves of compact type. This follows from [Reference Hain12, Thm. 7.1].
We now take
$X = {\mathcal M}_g^c$
, the moduli space of stable genus g curves of compact type (or a suitable open subset), where
$g \ge 3$
. Let
$f : \mathcal C \to {\mathcal M}_g^c$
be the universal curve. Set

This is the weight
$-1$
polarized variation of Hodge structure (PVHS) over
${\mathcal M}_g^c$
whose fibre over the moduli point
$[C]$
of C is

(We will also regard this as a PVHS over
$\mathcal A_g$
and this as its pullback along the period map.) We will take
$\mathbb V$
to be the weight
$-1$
PVHS

where
$\theta \in \Lambda ^2 {\mathbb H}$
is the section corresponding to the polarization. The Ceresa cycle gives a normal function section
$\nu $
of
$J(\mathbb V)$
over
${\mathcal M}_g^c$
. We will give the Collino–Pirola formula for the Green–Griffiths invariant
$\overline {\delta }_C(\nu )$
of the Ceresa cycle at
$[C]$
and a partial formula for its canonical derivative
$\overline {\nabla }_C \nu $
at
$[C]$
.
5 Geometry of the period mapping
Because we are computing derivatives and so working locally, it is often more convenient to work on Torelli space and Siegel space rather than on the moduli spaces
${\mathcal M}_g$
and
$\mathcal A_g$
.
Suppose that
$g\ge 2$
. Torelli space
${\mathcal T}_g$
is the quotient of Teichmüller space by the Torelli group. It is the moduli space of pairs
$(C;\mathbf {a}_1,\dots ,\mathbf {a}_g,\mathbf {b}_1,\dots ,\mathbf {b}_g)$
consisting of ‘framed curves’ – smooth projective curves of genus g and a symplectic basis of its first integral homology group. It is a complex manifold. The symplectic group
$\mathrm {Sp}_g({\mathbb Z})$
acts on
${\mathcal T}_g$
by its action on the basis. The moduli space
${\mathcal M}_g$
of smooth projective curves of genus g is the stack quotient
$\mathrm {Sp}_g({\mathbb Z})\backslash {\mathcal T}_g$
.
The Siegel upper half plane
$\mathfrak {h}_g$
of degree g is defined by

It is the moduli space of g dimensional principally polarized abelian varieties
$(A,\theta )$
of dimension g together with a basis
$(\mathbf {a}_1,\dots ,\mathbf {a}_g,\mathbf {b}_1,\dots ,\mathbf {b}_g)$
of
$H_1(A;{\mathbb Z})$
that is symplectic with respect to the polarization
$\theta : H_1(A) \otimes H_1(A) \to {\mathbb Z}$
. The period mapping
${\mathcal T}_g \to \mathfrak {h}_g$
takes a framed curve to its framed jacobian:

It is holomorphic. The symplectic group acts on
$\mathfrak {h}_g$
via its action on the basis. The stack quotient
$\mathrm {Sp}_g({\mathbb Z})\backslash \mathfrak {h}_g$
is the moduli space
$\mathcal A_g$
of principally polarized abelian varieties. The period map is
$\mathrm {Sp}_g({\mathbb Z})$
equivariant and descends to a map
${\mathcal M}_g \to \mathcal A_g$
.
The cotangent space of
${\mathcal T}_g$
at a point corresponding to a framing of C is
$H^0(\Omega _C^{\otimes 2})$
and the cotangent space of its image in
$\mathfrak {h}_g$
is
$S^2 H^0(\Omega _C)$
. The differential is the multiplication mapping

By Noether’s theorem, it is surjective when C is not hyperelliptic. When C is hyperelliptic, it has maximal rank tangent to the hyperelliptic locus. Consequently the period mapping is a local embedding away from the hyperelliptic locus and also a local embedding when restricted to the hyperelliptic locus.
The involution
$u\mapsto -u$
of
$H_1$
induces an isomorphism

Since a genus g curve C is hyperelliptic if and only if there is an isomorphism

the period mapping
$\pi : {\mathcal T}_g \to \mathfrak {h}_g$
is an étale 2:1 covering onto its image away from the hyperelliptic locus where it is an embedding.
6 The Green–Griffiths invariant of the Ceresa cycle
Suppose that C is a smooth projective curve of genus
$g\ge 3$
. Set

The polarization induces an isomorphism
$A\cong B^\vee $
. The cotangent space
$T_{[\operatorname {Jac} C]}^\vee \mathcal A_g$
to
$\mathcal A_g$
at the moduli point
$[\operatorname {Jac} C]$
of
$\operatorname {Jac} C$
is isomorphic to
$S^2 B$
.
The fibre V of
$\mathbb V$
over C is the primitive quotient

of
$H_3(\operatorname {Jac} C)(-1)$
. The canonical polarization
$PH^3(\operatorname {Jac} C) \otimes PH^3(\operatorname {Jac} C) \to {\mathbb Q}$
induces an isomorphism

We therefore have isomorphisms

where the polarization
$\theta $
corresponds to the identity in
$\operatorname {End}(B) \cong A\otimes B$
.
The Green–Griffiths invariant
$\overline {\delta }_C(\nu )$
of
$\nu $
at C lies in the homology of the complex
$\operatorname {Gr}_F^0(V\otimes \Lambda ^{\bullet } T_{[C]}^\vee {\mathcal M}_g)$
:

It is the pullback along the period mapping
${\mathcal M}_g \to \mathcal A_g$
of the fibre

over
$[\operatorname {Jac} C] \in \mathcal A_g$
of the complex of vector bundles
$\operatorname {Gr}_F^0 (\mathcal V \otimes \Omega ^{\bullet }_{\mathcal A_g})$
. The differential is induced by the map

that is adjoint to the projection
$B^{\otimes 2} \to S^2 B$
. It is a complex in the category of
$\mathrm {GL}(B)$
modules. As such, it does not depend on the point
$[\operatorname {Jac} C]$
of
$\mathcal A_g$
.
The derivative
$\nabla \nu _{\mathbb R}$
of
$\nu $
at
$[C]$
can be regarded as the complex linear map

via the prescription in Proposition 3.3. The corresponding point in

is a 1-cocycle in the complex (10) and represents the Green-Griffiths invariant
$\overline {\delta }_C(\nu )$
of
$\nu $
at
$[C]$
in its homology.
Remark 6.1. The complex (10) and the Griffiths invariant of the Ceresa cycle have been studied by various authors. See [Reference Collino and Pirola3] and the references therein.
7 Symmetry
Symmetry can be a useful tool for computing
$\overline {\nabla }_C \nu $
. The isotropy group
$G_C$
of a point of
${\mathcal T}_g$
corresponding to a framing of C is a subgroup of
$\mathrm {Sp}_g({\mathbb Z})$
isomorphic to
$\operatorname {Aut} C$
.
Set
$\mathbb V = \Lambda ^3_0 {\mathbb H}$
, regarded as a variation of Hodge structure over
${\mathcal T}_g$
and
$\mathcal V = \mathbb V \otimes {\mathcal O}_{{\mathcal T}_g}$
.
Proposition 7.1. Suppose that
$g\ge 3$
. If C is a smooth projective curve of genus g, then
$\operatorname {Aut} C$
acts on the stalk at
$[C]$
of the complex
$\mathcal V\otimes \Omega _{{\mathcal M}_g}^{\bullet }$
and preserves its Hodge filtration. It fixes
-
(1) the Griffiths invariant
$\delta _C(\nu )$ of the normal function of the Ceresa cycle in the stalk at
$[C]$ of
$\mathcal H^1(F^0 \mathcal V\otimes \Omega _{{\mathcal M}_g}^1)$ ,
-
(2) the cocycle
$\overline {\nabla }_C \nu \in \operatorname {Gr}_F^{-1} V_C \otimes T_C^\vee {\mathcal M}_g$ ,
-
(3) the Green–Griffiths invariant
$\overline {\delta }_C(\nu )$ of
$\nu $ at C,
-
(4) the derivative
$\nabla \nu _{\mathbb R}$ at
$[C]$ of the real lift of
$\nu $ .
Proof. The group
$\mathrm {Sp}_g({\mathbb Z})$
acts on the diagram

It follows that
$\operatorname {Aut} C$
acts on the restriction of this family to an analytic neighbourhood of
$[C]$
. Now apply Lemma 3.5.
Part 3. The Ceresa cycle in genus 3
In this part we further investigate the Green–Griffiths invariant of the Ceresa normal function in genus 3 and use a result of Collino and Pirola [Reference Collino and Pirola3] to prove the base case of Theorem 1. We then give a new proof of the Collino–Pirola Theorem using Teichmüller forms and use it to prove Theorem 2.
8 The Green–Griffiths invariant in genus 3
In genus 3, the period map
${\mathcal M}_3 \to \mathcal A_3$
is a local biholomorphism away from the hyperelliptic locus. It induces an isomorphism of the complex (10) with (11), which enables us, when working at a nonhyperelliptic curve, to work locally on
$\mathcal A_3$
and exploit the action of
$\mathrm {GL}(B)$
on the complex (11).
8.1 Some representation theory
Denote the top exterior power of a representation V by
$\det V$
. Since the pairing
$B \otimes \Lambda ^2 B \to \det B$
is nonsingular in genus 3, we have a canonical isomorphism

of
$\mathrm {GL}(B)$
modules. Similarly,
$\Lambda ^2 A \cong B \otimes \det A$
. We also have a canonical isomorphism

which holds for all
$g \ge 2$
. When
$g=3$
, we thus have a canonical isomorphism

Both summands are irreducible.
Lemma 8.1. In genus 3, the are natural
$\mathrm {GL}(B)$
module isomorphisms

Proof. The first isomorphism follows from the isomorphisms

The second is proved similarly, or by taking duals.
8.1.1 Green–Griffiths cohomology
The following result combines computations of Nori [Reference Nori23, pp. 371–372] and Collino–Pirola [Reference Collino and Pirola3, Lem. 4.2.3].
Proposition 8.2. If C is a nonhyperelliptic curve of genus 3, then
-
(1) If
$p \ge 0$ , then
$H^0(\operatorname {Gr}_F^p(V\otimes \Lambda ^{\bullet } S^2 B))$ vanishes.
-
(2) If
$p> 0$ , then
$H^1(\operatorname {Gr}_F^p(V\otimes \Lambda ^{\bullet } S^2 B))$ vanishes.
-
(3) The space of 1-cocycles in the complex
$\operatorname {Gr}_F^0 (V\otimes \Lambda ^{\bullet } S^2 B)$ is naturally isomorphic to
$$ \begin{align*}S^2 S^2 B \otimes \det A \cong (S^4 B \otimes \det A) \oplus (S^2 A \otimes \det B) \end{align*} $$
$S^2 A \otimes \det B$ . Consequently,
$$ \begin{align*}H^1(\operatorname{Gr}_F^0 (V\otimes \Lambda^{\bullet} S^2 B)) \cong S^4 B \otimes \det A. \end{align*} $$
Proof. The complexes
$\operatorname {Gr}_F^p (V \otimes \Lambda ^{\bullet } S^2 B)$
are:

Each is a complex of
$\mathrm {GL}(B)$
modules. The vanishing of the homology of the first two rows in degrees 0 and 1 is sketched by Nori [Reference Nori23, §7]. Here we sketch an elementary proof. The exactness of the first row
$\operatorname {Gr}_F^2$
follows from the fact that
$S^2 B\otimes \det B$
is irreducible and the differential is nonzero and therefore injective. Exactness of the second row
$\operatorname {Gr}_F^1$
follows (using Lemma 8.1) from the fact that

which has three irreducible factors. The differential takes
$\Lambda ^3 B$
to
$\det B\otimes \operatorname {id}_{S^2B}$
. The other two factors map injectively into the degree 2 term. So we focus on the complex
$\operatorname {Gr}_F^0$
.
Lemma 8.1 implies that the degree 0 term is
$S^2 A \otimes \det B$
and that the degree 1 term is

Since the degree 0 term is irreducible and the differential is nonzero, we see that the differential is injective and that the group of 1-coboundaries is
$S^2 A\otimes \det B$
. Since the degree 2 term is irreducible and not isomorphic to either of the first two components of the degree 1 term, and since the second differential is nonzero, it follows that the group of 1-cocycles is
$(S^4 B\otimes \det A) \oplus (S^2 A \otimes \det B)$
and that
$H^1(\operatorname {Gr}_F^0)$
is
$S^4 B \otimes \det A$
.
Since each
$\operatorname {Gr}_F^p(\mathcal V \otimes \Omega _{{\mathcal M}_3}^{\bullet })$
is a complex of vector bundles over the complement of the hyperelliptic locus, the previous result implies:
Corollary 8.3. When
$g=3$
, the differential
$\nabla : F^0 \mathcal V \to \mathcal V \otimes \Omega ^1_{{\mathcal M}_3}$
is injective on the complement of the hyperelliptic locus. In addition, the homology sheaves
$\mathcal H^1(\operatorname {Gr}^p_F(\mathcal V\otimes \Omega _{{\mathcal M}_3}^{\bullet }))$
vanish on the complement of the hyperelliptic locus when
$p>0$
.
Since
$\mathbb P(H^0(\Omega ^1_C))^\vee = \mathbb P(A)$
, the Green–Griffiths invariant
$\overline {\delta }_C(\nu )$
, if it is nonzero, defines a plane quartic in
$\mathbb P(A)$
.
Theorem 8.4 (Collino–Pirola [Reference Collino and Pirola3, Thm. 4.2.4]).
If C is a nonhyperelliptic curve of genus 3, then
$\overline {\delta }_C(\nu )$
is a nonzero quartic polynomial that defines the canonical image of C in
$\mathbb P(A)$
.
We will give a new proof of it in Section 10 using Teichmüller modular forms.
8.2 Linear algebra
Suppose that C is not hyperelliptic of genus 3. We will regard
$S^2 S^2 B \otimes \det A$
as the space of symmetric bilinear forms
$S^2 A \otimes S^2 A \to \det A$
. In view of Proposition 8.2(c), we can thus regard the derivative
$\overline {\nabla }_C \nu $
as a symmetric bilinear form

It decomposes as the sum

of two symmetric forms, where
$Q_C \in S^4 B \otimes \det A$
, a natural representative of
$\overline {\delta }_C(\nu )$
, and
$R_C \in S^2 A \otimes \det B$
, which is a coboundary.
Fix a volume form
$\mathrm {vol}_A \in \det A$
and let
$\mathrm {vol}_B \in \det B$
be the dual volume form. Identify
$\det A$
and
$\det B$
with
$\mathbb C$
via the isomorphisms

The pairing
$Q_C : S^2 A \otimes S^2 A \to \mathbb C$
is (up to a nonzero multiple) given by the formula

where
$f \in S^4 B$
is a generator of the ideal of functions that vanish on the canonical image of C in
$\mathbb P(A)$
and where
$uv\in S^4 A$
is the product of
$u,v\in S^2 A$
.
The pairing
$R_C$
is the composite

where the first map is induced by the isomorphism

and the second by

We now give explicit formulas (up to multiplication by a nonzero scalar) of them. Fix a basis
$\mathbf {e}_0,\mathbf {e}_1,\mathbf {e}_2$
of A such that
$\mathrm {vol}_A = \mathbf {e}_0 \wedge \mathbf {e}_1 \wedge \mathbf {e}_2$
. Let
$x_0,x_1,x_2$
be the dual basis of B.
Proposition 8.5. If the defining equation f of C in
$\mathbb P(A)$
is

then the matrix of
$Q_C$
with respect to the basis

of
$S^2 A$
is a nonzero multiple of

Sketch of proof.
The
$\mathrm {GL}(A)$
-invariant dual pairing
$S^4 B \otimes S^4 A \to \mathbb C$
is unique up to a constant. We choose the normalization

where
$\mathbb S_4$
is the symmetric group on four letters. So, for example,

Thus

The remaining entries are obtained by permuting the indices.
Remark 8.6. For the sake of completeness, we give a formula for the bilinear form
$R_C$
. If

is the projection of
$\overline {\nabla }_C \nu $
onto
$S^2 A\otimes \det B$
, then the matrix of
$R_C$
with respect to the basis

of
$S^2 A$
is a nonzero multiple of

To prove this, one easily checks that the map
$S^2 A \to (S^2B)^{\otimes 2}$
satisfies

The formulas for the images of the other basis vectors is obtained by cyclically permuting the indices.
8.3 The Klein quartic
Here we prove Theorem 1 in genus 3 by checking that
$\overline {\nabla }\nu $
has maximal rank at the moduli point of the Klein quartic C. We do this by using the symmetries of C to show that the tensor
$R_C$
vanishes.
The Klein curve is the projective completion
$X(7)$
of the modular curve
$Y(7) = \Gamma (7)\backslash \mathfrak {h}$
, where
$\Gamma (7)$
is the full level 7 subgroup of
$\mathrm {SL}_2({\mathbb Z})$
. It has genus 3 and automorphism group
$\mathrm {PSL}_2(\mathbb F_7)$
, a simple group of order 168. Its canonical image is the plane quartic

The corresponding matrix (12) has rank 6 and determinant
$-729 = -3^6$
. To show that
$\overline {\nabla }_C(\nu )$
has maximal rank it suffices to show that
$R_C$
vanishes.
Proposition 8.7. If C is the Klein quartic, then
$R_C = 0$
. Consequently,
$\overline {\nabla } \nu $
has maximal rank at
$[C] \in {\mathcal M}_3$
.
Proof. Set
$G=\mathrm {PSL}_2(\mathbb F_7)$
. In view of Proposition 7.1, it suffices to show that
$S^2 A\otimes \det B$
does not contain the trivial representation of G. The irreducible complex representations of G have dimensions 1, 3, 3, 6, 7, 8. The two three-dimensional representations are conjugate. Since
$\mathrm {SL}_2({\mathbb Z})$
has no cusp forms of weight 2,
$B = H^0(\Omega ^1_C)$
does not contain the trivial representation. It is thus one of the two three-dimensional irreducible representations and A is its conjugate. Since the only one-dimensional G-module is the trivial representation,
$\det A$
and
$\det B$
are trivial.
Both
$S^2 A$
and
$S^2 B$
are isomorphic to the unique six-dimensional irreducible representation. It follows that
$S^2 A\otimes \det B$
is irreducible and thus contains no copy of the trivial representation.
Remark 8.8. One can easily check that
$S^4 B$
contains one copy of the trivial representation. It is spanned by a defining equation of C in
$\mathbb P(A)$
.
Since
$\overline {\nabla } \nu $
is a holomorphic section of
$\mathcal V\otimes \Omega _{{\mathcal M}_3}^1$
, we conclude that
$\overline {\nabla } \nu $
has maximal rank on an open dense subset of
${\mathcal M}_3$
.
Proposition 8.9. The derivative
$\overline {\nabla } \nu $
has maximal rank on a dense open subset of
${\mathcal M}_3$
.
9 Geometry of the period map in genus 3
Here we supplement the discussion of the geometry of the period map in Section 5 by giving a more detailed description of it in genus 3, which is needed in our proof of Theorem 8.4. This enables us to show that the Green–Griffiths invariant
$\overline {\delta }(\nu )$
is a Teichmüller modular form and to show that the derivative of
$\nu $
along the hyperelliptic locus is the restriction of the Siegel modular form
$\chi _{4,0,8}$
to the hyperelliptic locus.
Denote the closure of the locus in
$\mathfrak {h}_3$
of jacobians of hyperelliptic curves by
$\mathfrak {h}_3^{\mathrm {hyp}}$
and the locus of reducible (as a principally polarized abelian variety) abelian 3-folds by
$\mathfrak {h}_3^{\mathrm {red}}$
. Since every genus 2 curve is hyperelliptic, every reducible abelian 3-fold is the jacobian of a hyperelliptic curve of compact type. So we have a filtration

The image of the genus 3 period mapping
${\mathcal T}_3 \to \mathfrak {h}_3$
is
$\mathfrak {h}_3 - \mathfrak {h}_3^{\mathrm {red}}$
. (This is surely well known. For a proof see [Reference Hain14, Cor. 7].) Each even theta characteristic
$\alpha $
determines a theta null
$\vartheta _\alpha : \mathfrak {h}_3 \to \mathbb C$
. Their product is the Siegel modular form

of weight 18.
Proposition 9.1. The genus 3 Torelli space
${\mathcal T}_3$
is the analytic subvariety of
$(\mathfrak {h}_3-\mathfrak {h}_3^{\mathrm {red}}) \times \mathbb C$
defined by
$y^2 = \chi _{18}(\tau )$
.
Proof. Since a smooth genus 3 curve is hyperelliptic if and only if it has a (necessarily unique) vanishing even theta characteristicFootnote
1
, it follows that
$\mathfrak {h}_3^{\mathrm {hyp}}$
is the zero locus of
$\chi _{18}$
. The result follows as
${\mathcal T}_3 \to \mathfrak {h}_3-\mathfrak {h}_3^{\mathrm {red}}$
is the unique double covering of
$\mathfrak {h}_3-\mathfrak {h}_3^{\mathrm {red}}$
branched along the smooth divisor
$\mathfrak {h}_3^{\mathrm {hyp}}-\mathfrak {h}_3^{\mathrm {red}}$
.
The function
$y \in {\mathcal O}({\mathcal T}_3)$
is thus a square root of
$\chi _{18}$
. In the terminology of Ichikawa [Reference Ichikawa19] it is a Teichmüller form of genus 3 and is, in fact, the most basic of them. It has weight 9, meaning it is a section of the 9th power of the determinant of the Hodge bundle, and vanishes to order 1 along the hyperelliptic locus
${\mathcal T}_3^{\mathrm {hyp}}$
of Torelli space. It has been (re)christened
$\chi _9$
by Cléry–Faber–van der Geer in [Reference Cléry, Faber and van der Geer2].
Denote the coordinates in
$\mathfrak {h}_3$
by

The genus 3 analogue of the q-disk is the quotient of
$\mathfrak {h}_3$
by the unipotent subgroup
${\mathbb Z}^3$
of
$\mathrm {Sp}_3({\mathbb Z})$
that acts via

Set
$q_j = \exp (2\pi i \tau _{jj})$
. Denote the punctured unit disk by
$\mathbb D'$
. The map

is an open immersion. Define a partial compactification
$({\mathbb Z}^3\backslash \mathfrak {h}_3)^c$
of
${\mathbb Z}^3\backslash \mathfrak {h}_3$
by adding the three coordinate hyperplanes
$q_j=0$
, but not their intersections
$q_j=q_k=0$
,
$j\neq k$
. Set

Let
$D_j$
be the divisor
$q_j=0$
and D be their union.
The Siegel modular form
$\chi _{18}$
descends to a function
$U \to \mathbb C$
, which we will also denote by
$\chi _{18}$
. We have

where H denotes the class of the image of
$\mathfrak {h}_3^{\mathrm {hyp}}$
. (See [Reference Cléry, Faber and van der Geer2, §3].) Denote the normalization of the zero locus of

by V. It is a smooth double covering of U branched over the hyperelliptic locus
$U^{\mathrm {hyp}}$
. Since
$\chi _{18}$
vanishes to order 2 along D, the projection
$\pi : V \to U$
is not ramified over D, only over
$U^{\mathrm {hyp}}$
. Denote the inverse image of D in V by
$\Delta $
. It lies over the smooth locus of the boundary divisor
$\Delta _0$
of
$\overline {{\mathcal M}}_3$
. The divisor D lies over a boundary divisor of each toroidal compactification of
$\mathcal A_3$
.
The modular form
$\chi _9$
extends to a holomorphic function
$V \to \mathbb C$
, also denoted
$\chi _9$
. We have

The representations A and B of
$\mathrm {GL}(B)$
correspond to automorphic vector bundles
$\mathscr A$
and
$\mathscr B$
over
$\mathcal A_3$
, where
$\mathscr B$
is the Hodge bundle. Their pullbacks to
${\mathbb Z}^3\backslash \mathfrak {h}_3$
and its double covering (also denoted
$\mathscr A$
and
$\mathscr B$
) extend to U (resp. V) as graded quotients of the Hodge filtration of Deligne’s canonical extension [Reference Deligne4] of
${\mathbb H}$
to U (resp. V). When
$\mathscr A$
is regarded as a sheaf over X, where
$X=\mathcal A_3, \mathfrak {h}_3, {\mathcal T}_3, V,\dots $
, we will denote it by
$\mathscr A_X$
. Similarly with
$\mathscr B$
.
The Teichmüller modular form
$\chi _9$
is a section of
$(\det \mathscr B)^9$
over
${\mathcal M}_3$
,
${\mathcal T}_3$
and V. It vanishes to order 1 on the hyperelliptic locus
$V^{\mathrm {hyp}}$
. Denote the inclusion
$V^{\mathrm {hyp}} \to V$
by j.
Lemma 9.2. The vector bundle
$\pi ^\ast \Omega ^1_U(\log D)$
over V is isomorphic to
$S^2\mathscr B_V$
. Consequently, we have a short exact sequence

Proof. Since
$S^2\mathscr B_U$
is isomorphic to
$\Omega _U^1(\log D)$
, [Reference Faltings and Chai5, p. 117], it follows that
$S^2\mathscr B_V$
is isomorphic to
$\Omega ^1_V(\log \Delta )$
away from the hyperelliptic locus. Since the hyperelliptic locus in V has reduced defining equation
$\chi _9=0$
, it has conormal bundle
$j^\ast (\det \mathscr B_V)^9$
. The exact sequence follows as
$\pi : V \to U$
is ramified along
$V^{\mathrm {hyp}}$
.
10 Proof of Theorem 2
Theorem 2 is a consequence of Theorem 10.2 below. Set
${\mathcal T} = {\mathcal T}_3$
,
$\mathfrak {h} = \mathfrak {h}_3$
and
$\mathfrak {h}' = \mathfrak {h}_3 - \mathfrak {h}_3^{\mathrm {red}}$
. We will work over
${\mathcal T}$
,
$\mathfrak {h}'$
, U and V, as appropriate. Their hyperelliptic loci will be denoted by
${\mathcal T}^{\mathrm {hyp}}$
,
$\mathfrak {h}^{\mathrm {hyp}}$
,
$V^{\mathrm {hyp}}$
, etc.
Lemma 10.1.
-
(a) The tensor
$\overline {\nabla } \nu $ is a real analytic section of the holomorphic vector bundle
$\operatorname {Gr}_F^{-1}\mathcal V \otimes \Omega ^1_{\mathcal T}$ over
${\mathcal T}$ .
-
(b) The Green–Griffiths invariant
$\overline {\delta }(\nu )$ is a holomorphic section of
$\operatorname {Gr}_F^{-1}\mathcal V\otimes \Omega _V^1(\log \Delta )$ over V.
-
(c) The kernel of
$\overline {\nabla } : \operatorname {Gr}_F^{-1} \mathcal V \otimes \Omega ^1_V(\log \Delta ) \to \operatorname {Gr}_F^{-2} \mathcal V \otimes \Omega _V^2(\log \Delta )$ has a natural direct sum decomposition
$$ \begin{align*}\overline{\nabla} \operatorname{Gr}_F^0 \mathcal V_V \oplus (S^4\mathscr B_V \otimes \det\mathscr A_V)^\sim \end{align*} $$
$(S^4\mathscr B_V \otimes \det \mathscr A_V)^\sim $ over the open subset O of V is
$$ \begin{align*}\{\omega \in H^0(O,\operatorname{Gr}_F^{-1} \mathcal V \otimes \Omega_V^1(\log\Delta)) : \chi_9\omega \in H^0(O,S^4 \mathscr B_V \otimes \det\mathscr A_V)\}. \end{align*} $$
-
(d) There is an exact sequence
$$ \begin{align*}0 \to S^4 \mathscr B_V \otimes \det \mathscr A_V \to (S^4 \mathscr B_V \otimes \det \mathscr A_V)^\sim \to j_\ast \big[S^4 \mathscr B_V \otimes (\det\mathscr B_V)^8\big] \end{align*} $$
$j : V^{\mathrm {hyp}} \to V$ is the closed inclusion of the hyperelliptic locus.
-
(e) There is a unique real analytic section f of
$\operatorname {Gr}_F^0 \mathcal V_V$ over
$V-\Delta $ and a unique holomorphic section
$\widetilde {\delta }(\nu )$ of
$(S^4\mathscr B_V \otimes \det \mathscr A_V)^\sim $ over V such that
$$ \begin{align*}\overline{\nabla} \nu = \widetilde{\delta}(\nu) + \overline{\nabla} f. \end{align*} $$
$\widetilde {\delta }(\nu )$ is anti-invariant under the involution. Both f and
$\overline {\nabla } f$ vanish on the hyperelliptic locus.
Proof. At two points in the proof, we will appeal to general results about normal functions from Part 4. These do not depend on results in the preceding sections. Denote the sheaf of real analytic j forms on a variety X by
$\mathscr E_X^j$
and the sheaf of real analytic
$(p,q)$
forms by
$\mathscr E_X^{p,q}$
. Set
$\mathscr E_X = \mathscr E_X^0$
.
The map
$\overline {\nabla } : \operatorname {Gr}_F^0 \mathcal V \to \operatorname {Gr}_F^{-1}\mathcal V \otimes \Omega ^1_{{\mathcal T}}$
of holomorphic vector bundles extends to an
$\mathscr E_{{\mathcal T}}$
-linear map

where
$\overline {\nabla }'$
denotes the map induced by the
$(1,0)$
component of

By Corollary 3.4,
$\overline {\nabla } \nu $
is the image of
$\overline {\nabla }'\nu _{\mathbb R}$
in
$\operatorname {Gr}_F^{-1}\mathcal V \otimes _{{\mathcal O}_{\mathcal T}} \mathscr E_{\mathcal T}^{1,0}$
. It is real analytic. This proves (a). Proposition 11.3 implies that
$\nu $
is admissible. So (b) follows from Corollary 12.6.
Proposition 8.2(c) implies that the kernel of
$\overline {\nabla } : \operatorname {Gr}_F^0 \mathcal V_U \to \operatorname {Gr}_F^{-1}\mathcal V \otimes \Omega ^1_U(\log D)$
is isomorphic to
$(S^4\mathscr B_U \otimes \det \mathscr A_U) \oplus (S^2 \mathscr A_U \otimes \det \mathscr B_U)$
. Assertion (c) follows from this and the fact that, since
$\overline {\nabla }$
is
${\mathcal O}_V$
-linear,
$\overline {\nabla } \operatorname {Gr}_F^0 \mathcal V_V$
is a subsheaf of
$\pi ^\ast \operatorname {Gr}_F^{-1}\mathcal V\otimes \Omega ^1_U(\log D)$
.
Denote the hyperelliptic locus of V by H. The exactness of the sequence in (d) follows from the fact that
$(S^4\mathscr B_V\otimes \det \mathscr A_V)^\sim $
is a subsheaf of
$S^4\mathscr B_V\otimes \det \mathscr A_V(H)$
and that the conormal bundle of H in V is the restriction of
$(\det \mathscr B_V)^9$
to H. This is because H is defined by
$\chi _9$
, which vanishes to order 1 on H and is a section of
$(\det \mathscr B)^9$
.
To prove (e) first choose a local holomorphic lift
$\tilde {\nu }$
of
$\nu $
defined on an open set O of V. We can write

where h is a section of
$\operatorname {Gr}_F^0 \mathcal V_V$
. Both components are holomorphic on V. Since two lifts of
$\nu $
differ by a section of
$F^0\mathcal V_V$
, the second component does not depend on the choice of the lift
$\tilde {\nu }$
. Call it
$\widetilde {\delta }(\nu )$
. The existence of f follows from the fact that
$\overline {\nabla } : \operatorname {Gr}_F^0 \mathcal V\otimes _{{\mathcal O}_{\mathcal T}} \mathcal E_{\mathcal T} \to \operatorname {Gr}_F^{-1}\mathcal V\otimes _{{\mathcal O}_{\mathcal T}} \mathcal E^{1,0}$
is injective, so there is a unique real analytic section f of
$\operatorname {Gr}_F^0 \mathcal V$
over
$V-\Delta $
with

The vanishing of f and
$\overline {\nabla } f$
on the hyperelliptic locus follows as
$\overline {\nabla }\nu $
is anti-invariant under the involution of
$\sigma : V\to U$
, which forces f to vanish on the hyperelliptic locus, and as
$\overline {\nabla }$
is a
$\sigma $
-invariant map of vector bundles.
We now assume familiarity with [Reference Cléry, Faber and van der Geer2] and use some of the notation in it, especially the notation for the weight
$(j,k,\ell )$
of Siegel and Teichmüller modular forms.Footnote
2
Theorem 10.2. The Green–Griffiths invariant
$\overline {\delta }(\nu )$
of the Ceresa cycle is a section of the subsheaf
$(S^4\mathscr B_V \otimes \det \mathscr A_V)^\sim $
of
$S^4\mathscr B_V \otimes \det \mathscr A_V(H)$
. It is a nonzero multiple of the Teichmüller modular form
$\chi _{4,0,-1}$
. Its restriction to the hyperelliptic locus of V projects, under the projection in Lemma 10.1(d), to a nonzero multiple of the section
$\chi _{4,0,8}$
of
$S^4\mathscr B_V\otimes (\det \mathscr B_V)^8$
.
Proof. We can identify
$\overline {\delta }(\nu )$
with its canonical representative
$\widetilde {\delta }(\nu )$
as both are sections of
$S^4 \mathscr B \otimes \det \mathscr A$
over V. Lemma 10.1(e) implies that
$\chi _9 \widetilde {\delta }(\nu )$
is a section of
$S^4 \mathscr B_V \otimes \det \mathscr A_V$
that is invariant under the involution of V. It is therefore a Siegel modular form of weight
$(4,0,8)$
. Since
$\chi _9$
vanishes on
$\Delta $
, it is a cusp form. Taïbi [Reference Taïbi29] has shown that the space of Siegel cusp forms of weight
$(4,0,8)$
is one dimensional and spanned by the cusp form
$\chi _{4,0,8}$
. It follows that
$\overline {\delta }(\nu )$
is a multiple of the meromorphic Teichmüller modular form

and that it is a section of
$(S^4 \mathscr B_V \otimes \det \mathscr A_V)^\sim $
.
This multiple is nonzero: The derivative
$\nabla \nu _{\mathbb R}$
of the Ceresa normal function is a de Rham representative of the class of
$\nu $
in
$H^1({\mathcal M}_3,\mathbb V)$
. This class is nonzero as its restriction to the Torelli group
$T_3$
is twice the Johnson homomorphism. (See [Reference Hain12, Thm. 11.1]. There are many other proofs of nontriviality, such as Ceresa’s original proof [Reference Ceresa1].) This implies that the Griffiths invariant
$\delta (\nu )$
is nonzero. Propositions 2.5 and 8.2 then imply that
$\overline {\delta }(\nu )$
is also nonzero.
Next we compute the restriction of
$\widetilde {\delta }(\nu )$
to the hyperelliptic locus. We already know that
$\chi _9 \widetilde {\delta }(\nu )$
descends to a nonzero multiple of the section
$\chi _{4,0,8}$
of the vector bundle
$S^4\mathscr B_U\otimes (\det \mathscr B_U)^8$
. Since
$U^{\mathrm {hyp}}$
is the divisor defined by
$\chi _{18} = 0$
and since the restriction of
$\nu $
to
$V^{\mathrm {hyp}}$
is zero, we can write

after identifying
$S^4\mathscr B_U\otimes \det \mathscr A_U$
with
$\operatorname {Gr}^{-1}_F\mathcal V\otimes \Omega ^1_U(\log D)$
. Here h is a section of the restriction of
$S^4\mathscr B\otimes (\det \mathscr B_U)^8$
to
$U^{\mathrm {hyp}}$
as the conormal bundle of
$U^{\mathrm {hyp}}$
is
$(\det \mathscr B_U)^{18}$
, which is trivialized by
$d\chi _{18}$
. It follows that the restriction of
$\widetilde {\delta }(\nu )$
to
$V^{\mathrm {hyp}}$
is

Since the conormal bundle of
$V^{\mathrm {hyp}}$
is isomorphic to
$(\det \mathscr B_V)^9$
, which is trivialized by
$d\chi _9$
, we see that the restriction of
$\widetilde {\delta }(\nu )$
to
$V^{\mathrm {hyp}}$
can be identified with a nonzero multiple of the restriction of
$\chi _{4,0,8}$
to
$V^{\mathrm {hyp}}$
.
Remark 10.3. The result of Collino and Pirola (Thm. 8.4) now follows from [Reference Cléry, Faber and van der Geer2, Prop. 10.1].
10.1 Proof of Corollary 3
Since the Ceresa normal function vanishes on the hyperelliptic locus, a smooth divisor, the rank of
$\nu $
at each point of the hyperelliptic locus is at most 1. By Lemma 10.1(e),
$\overline {\nabla } f$
vanishes on the hyperelliptic locus. So, to prove the result, we need to show that
$\overline {\delta }(\nu )$
has no zeros on the hyperelliptic locus. By Theorem 10.2, the restriction of
$\overline {\delta }(\nu )$
to the hyperelliptic locus can be identified with
$\chi _{4,0,8}$
. So it suffices to show that the restriction of
$\chi _{4,0,8}$
to the hyperelliptic locus has no zeros.
In genus 3, the restriction of
$\mathscr B$
to the hyperelliptic locus (regarded as a stack) is
$S^2\mathscr W$
, where
$\mathscr W$
is the rank 2 vector bundle whose fibre over the moduli point of a hyperelliptic curve is the linear system associated to its hyperelliptic series. By the representation theory of
$\mathrm {GL}_2$
, the restriction of
$S^4\mathscr B$
to the hyperelliptic locus in genus 3 decomposes

and
$\det \mathscr B$
restricts to
$(\det \mathscr W)^3$
. So the restriction of
$S^4\mathscr B\otimes (\det \mathscr B)^8$
to the hyperelliptic locus projects to

By [Reference van der Geer, Kouvidakis and Alexis7, §6], the projection of the restriction of
$\chi _{4,0,8}$
to the hyperelliptic locus onto this factor is the product
$f_{8,-2}$
of the universal octic times the discriminant
$\mathfrak {d}$
of the binary octic, which is a section of
$(\det \mathscr W)^{28}$
. Since neither vanish at the moduli point of a smooth hyperelliptic curve,
$\chi _{4,0,8}$
has no zeros there.
This strengthens the genus 3 case of the result [Reference Harris18, Thm. 6.5] of Bruno Harris. He proved that the derivative of the normal function of the Ceresa cycle has rank 1 at the general hyperelliptic curve of genus
$g\ge 3$
.
Part 4. Admissible normal functions
The inductive proof of the general case requires technical Hodge theory. In particular, it requires an understanding of the asymptotic behaviour of the variations of mixed Hodge structure that correspond to normal functions.
After recalling the definition and basic properties of admissible variations of mixed Hodge structure, we apply them to describe the boundary behaviour of normal functions. We review the construction of the Néron model of a family of intermediate jacobians by Green, Griffiths and Kerr [Reference Green, Griffiths and Kerr10] in the case where the variation is a nilpotent orbit. This is applied to study the asymptotic behaviour of admissible normal functions in codimension 1. In particular, it is used in the definition of residual normal functions, which plays a key role in the inductive proof of Theorem 1.
11 Admissible variations of mixed Hodge structure
This is a terse review of variations of mixed Hodge structure. Basic references include the papers of Steenbrink–Zucker [Reference Steenbrink and Zucker28] and Kashiwara [Reference Kashiwara22].
Suppose that S is a smooth quasi-projective variety. We write
$S= {\overline {S}} - D$
, where
${\overline {S}}$
is a smooth projective variety and D is a normal crossing divisor. An admissible variation of
${\mathbb Z}$
-MHS
${\mathbb E}$
over S consists of the following data:
-
(1) A local system
${\mathbb E}_{\mathbb Z}$ over S of finitely generated abelian groups which, for simplicity, we assume to have unipotent local monodromy.Footnote 3
-
(2) An increasing filtration (the weight filtration)
$$ \begin{align*}0 = W_m {\mathbb E}_{{\mathbb Q}} \subseteq W_{m+1} {\mathbb E}_{{\mathbb Q}} \subseteq \cdots \subseteq W_n {\mathbb E}_{{\mathbb Q}} = {\mathbb E}_{{\mathbb Q}} \end{align*} $$
${\mathbb E}_{{\mathbb Q}} := {\mathbb E}_{\mathbb Z}\otimes {\mathbb Q}$ by sub local systems.
-
(3) A decreasing filtration (the Hodge filtration)
$$ \begin{align*}\mathcal E = F^a \mathcal E \supseteq F^{a+1} \mathcal E \supseteq \cdots \supseteq F^b \mathcal E = 0 \end{align*} $$
$\mathcal E$ of
${\mathbb E}_{\mathbb Z}\otimes {\mathcal O}_S$ to
${\overline {S}}$ by holomorphic sub-bundles. That is,
$F^p\mathcal E$ and
$\mathcal E/F^p\mathcal E$ are locally free for each p.
The flat connection on
${\mathbb E}_{\mathbb Z}\otimes {\mathcal O}_S$
extends to a connection

Under the assumption of unipotent local monodromy, the canonical extension is characterized by the property that the residue of the connection at each smooth point of D is nilpotent.
These data are required to satisfy the following conditions:
-
(a) The connection on
$\mathcal E$ satisfies Griffiths transversality:
(15)for each p.$$ \begin{align} \nabla : F^p \mathcal E \to F^{p-1}\mathcal E \otimes \Omega^1_{\overline{S}}(\log D) \end{align} $$
-
(b) For each
$s\in S$ , the restriction of the Hodge and weight filtrations to the fibre of
${\mathbb E}_{\mathbb Z}$ over s is a mixed Hodge structure.
-
(c) Each weight graded quotient
$\operatorname {Gr}_W^m {\mathbb E}$ admits a polarization in the sense of [Reference Schmid26, §2].
-
(d) For each holomorphically embedded arc
$\alpha : \mathbb D \to {\overline {S}}$ satisfying
$\alpha ^{-1}(D) = \{0\}$ , there is a relative weight filtration
$$ \begin{align*}0 = M_r E_0 \subseteq M_{r+1} E_0 \subseteq \cdots \subseteq M_s = E_0 \end{align*} $$
$E_0$ of
$\mathcal E$ over
$\alpha (0)$ . Set
$$ \begin{align*}N = -\operatorname{Res}_0 \nabla|_{\mathbb D} \in \operatorname{End} E_0. \end{align*} $$
-
(i)
$N (M_j E_0) \subseteq M_{j-2} (E_0)$ and
$N(W_r E_0) \subseteq W_r E_0$ for each
$j,r \in {\mathbb Z}$ .
-
(ii) for each
$r,k \in {\mathbb Z}$ ,
$N^k : M_{r+k} W_r E_0 \to M_{r-k} W_m E_0$ induces an isomorphism
(16)See [Reference Steenbrink and Zucker28] for more details and also [Reference Hain15, §7] for an exposition with some relevant examples. Note that the filtration$$ \begin{align} \operatorname{Gr}^M_{r+k} \operatorname{Gr}^W_r E_0 \overset{\simeq}{\longrightarrow} \operatorname{Gr}^M_{r-k} \operatorname{Gr}^W_r E_0. \end{align} $$
$M_{\bullet }$ is called the monodromy weight filtration when
${\mathbb E}$ is a variation of Hodge structure.
-
These data determine, for each nonzero tangent vector
${\vec v} \in T_0 \mathbb D$
, a canonical limit mixed Hodge structure which we denote by
$E_{\vec v}$
. It is defined as follows. Choose a holomorphic parameter t in
$\mathbb D$
such that
${\vec v} = \partial /\partial t$
. First observe that each weight graded quotient
$\operatorname {Gr}^W_m {\mathbb E}$
is a polarized variation of Hodge structure in the sense of Schmid [Reference Schmid26].
Standard results in the theory of ordinary differential equations imply that there is a unique trivialization

of the restriction of
$\mathcal E$
to
$\mathbb D$
in which the connection is

(This trivialization depends on the choice of parameter t.) The flat sections are of the form

Identify the fibre
$E_t$
of
$\mathcal E$
over
$\alpha (t) \in \mathbb D'$
with
$E_0$
using the trivialization (17) above. This also allows us to regard N as an endomorphism of
$E_t$
. With these identifications, we have

where
$h_t : E_t \to E_t$
is the local monodromy operator.
The identification of the fibre
$E_\lambda $
over
$\lambda \in \mathbb D'$
with
$E_0$
defines a
${\mathbb Z}$
structure on
$E_0$
. Denote it by
$E_{\lambda {\vec v},{\mathbb Z}}$
. The limit mixed Hodge structure
$E_{\lambda {\vec v}}$
is the trifiltered vector space

endowed with the lattice
$E_{\lambda {\vec v},{\mathbb Z}}$
. Schmid’s result implies that each of its
$W_{\bullet }$
graded quotients comprises an MHS with weight filtration
$M_{\bullet }$
.Footnote
4
It then follows that for each
$\lambda \in \mathbb D'$
,
$E_{\lambda {\vec v}}$
is a MHS with the filtration

by sub MHS. Although the identification of
$E_\lambda $
with
$E_0$
depends on the choice of parameter t, the
${\mathbb Z}$
-structure (and thus the MHS) depends only on the tangent vector
$\lambda {\vec v} = \lambda \partial /\partial t \in T_0 \mathbb D$
.
In fact, the
${\mathbb E}_{\lambda {\vec v}}$
are defined for all
$\lambda \neq 0$
and form a nilpotent orbit of MHS over the punctured tangent space
$T_0'\mathbb D := T_0 \mathbb D - \{0\}$
.
Griffiths transversality (15) implies that
$N (F^p E_0) \subseteq F^{p-1}E_0$
. Since
$N(M_rE_0) \subset M_{r-2}E_0$
, we have:
Proposition 11.1. For each
${\vec v} \in T_0'\mathbb D$
, the monodromy logarithm
$N : E_0 \to E_0$
induces a morphism
$N : E_{\vec v} \to E_{\vec v}(-1)$
of MHS.
11.1 Nilpotent orbits
The 1-variable nilpotent orbit above is the restriction of a several variable nilpotent orbit. Such nilpotent orbits of limit MHS approximate admissible variations of MHS near a point of the boundary divisor D and are useful for understanding the boundary behaviour of admissible variations.
Suppose that
$U = \mathbb D^n$
is a polydisk neighbourhood in
${\overline {S}}$
of a point of D. Suppose that the intersection of D with U is defined by
$t_1t_2\dots t_k = 0$
, where
$t_j$
is the coordinate in the jth disk. Let
$D_j$
be the component of
$D\cap U$
defined by
$t_j = 0$
. Denote the fibre of
$\mathcal E$
over the origin of U by
$E_0$
. Set

Then
$N_1,\dots ,N_k$
are commuting nilpotent endomorphisms of
$E_0$
. There is a trivialization

of the restriction of
$\mathcal E$
to U in which
-
(1)
$\nabla $ is given by
$$ \begin{align*}\nabla = d - \sum_{j=1}^k N_j \frac{dt_j}{t_j} \end{align*} $$
-
(2) The isomorphism (18) restricts to an isomorphism of
$W_m \mathcal E|_U$ with
$U\times W_m E_0$ .
Define a Hodge filtration on
$U\times E_0$
by
$F^p(U\times E_0) = U \times F^p E_0$
. The fibre of
$F^p \mathcal E$
over
$t\in U$
corresponds to a subspace
$F^p_t E_0$
of
$E_0$
via the isomorphism (18). Since both Hodge filtrations of the bundle
$U\times E_0$
are holomorphic and agree at the origin, we have:
Proposition 11.2. There is a holomorphic mapping
$\Phi _{{\mathbb E}} : U \to \operatorname {Aut} E_0$
with the property that:
-
(a)
$\Phi _{{\mathbb E}}(0) = \operatorname {id}_{E_0}$ ,
-
(b)
$\Phi _{{\mathbb E}}(t)$ preserves
$W_{\bullet }$ and the induced automorphism of
$\operatorname {Gr}^W_0 E_0$ is the identity,
-
(c) the original Hodge filtration on
$\mathcal E|_U$ is related to the Hodge filtration on
$U\times E_0$ by
$$ \begin{align*}F^p_t E_0 = \Phi_{{\mathbb E}}(t)(F^p E_0). \end{align*} $$
Suppose that
$x=(0,\dots ,0,t_{k+1},\dots ,t_n) \in \cap D_j$
. For each tangent vector

that is not tangent to any component of D — equivalently,
$\lambda _j \neq 0$
for
$1\le j \le k$
— we can restrict
${\mathbb E}$
to the arc
$t\mapsto (\lambda _1 t_1,\dots , \lambda _k t_k,t_{k+1},\dots ,t_n)$
to obtain a limit MHS
$E_{\vec v}$
whose weight filtration is the relative weight filtration of

The
$E_{\vec v}$
form a nilpotent orbit of MHS over

that is constant on the cosets of
$T_x(\bigcap _{j=1}^k D_j)$
. It therefore descends to a nilpotent orbit of MHS on the normal bundle of the intersection of the
$D_j$
with the normal bundles of the
$D_j$
removed. We shall denote this nilpotent orbit by
${\mathbb E}^{\mathrm {nil}}$
. It is an admissible variation of MHS. The case
$k=1$
plays an important role in the proof of Theorem 1.
Nilpotent orbits are themselves admissible VMHS. They behave well under subquotients of admissible variations of MHS. In particular, the nilpotent orbit associated to the subquotient
$\operatorname {Gr}^W_r {\mathbb E}$
is
$\operatorname {Gr}^W_r{\mathbb E}^{\mathrm {nil}}$
.
11.2 Admissible normal functions
Suppose that
$\mathbb V$
is a polarizable variation of Hodge structure of negative weight over a smooth variety S. An admissible normal function is a section
$\nu $
of
$J(\mathbb V)$
that corresponds to an admissible variation of MHS
${\mathbb E}$
over S that is an extension of
${\mathbb Z}_S(0)$
by
$\mathbb V$
.
Proposition 11.3. Normal functions associated to families of homologically trivial cycles are admissible.
Proof. Suppose that we are in the setting of Section 2.4. Let
${\mathbb E}$
be the extension
${\mathbb E}_Z$
associated to Z. There is a Zariski open subset U of S on which the map
$(X,|Z|) \to S$
is topologically locally trivial. To prove
${\mathbb E}$
is an admissible variation of MHS over U it suffices, by [Reference Kashiwara22], to show that its restriction to each curve in S is admissible. This follows from the main result of [Reference Steenbrink and Zucker28] applied to
$(X-|Z|)|_U \to U$
. It also follows by appealing to Saito’s theory of mixed Hodge modules [Reference Saito25]. The result of [Reference Hain12, §7] implies that
${\mathbb E}$
is an admissible variation over S.
12 Codimension 1 degenerations of normal functions
We now study codimension 1 degenerations of admissible normal functions in the case where
$\mathbb V$
is an admissible variation of Hodge structure of weight
$-1$
. In this case,
$J(\mathbb V)$
is a holomorphic family of compact complex tori.
12.1 Setup
Suppose that
${\overline {S}}$
is a smooth variety (not necessarily compact) and that
$\Delta $
is a connected smooth divisor in
${\overline {S}}$
. Set
$S={\overline {S}}-\Delta $
. Suppose that
$\mathbb V$
is a polarizable variation of Hodge structure of weight
$-1$
over S and that
$\nu : S \to J(\mathbb V)$
is an admissible normal function. Denote the corresponding variation of MHS by
${\mathbb E}$
. We shall assume that
$\mathbb V_{\mathbb Z}$
is torsion free, a condition that is satisfied by all variations in Part 5.
Denote the normal bundle of
$\Delta $
in
${\overline {S}}$
by L and the complement of its 0-section, which we identify with
$\Delta $
, by
$L'$
. Let
$\pi : L \to \Delta $
be the projection and
$\pi '$
its restriction to
$L'$
. By the discussion in Section 11.1, there are nilpotent orbits of MHS
${\mathbb E}^{\mathrm {nil}}$
and
$\mathbb V^{\mathrm {nil}}$
over
$L'$
and an extension

of variations over
$L'$
. Both
$\mathbb V^{\mathrm {nil}}$
and
${\mathbb E}^{\mathrm {nil}}$
are filtered by subvariations
$M_j\mathbb V^{\mathrm {nil}}$
and
$M_j{\mathbb E}^{\mathrm {nil}}$
.
Even though
$\mathbb V^{\mathrm {nil}}$
is not a variation of Hodge structure of pure weight
$-1$
(except when
$N=0$
), we can still define
$J(\mathbb V^{\mathrm {nil}})$
to be the (not necessarily topologically locally trivial) family of complex tori over
$L'$
whose fibre over
${\vec v} \in L_x'$
is

Schmid’s result [Reference Schmid26, Thm. 4.9] implies that there is a neighbourhood U of the zero section of L such that
$\mathbb V^{\mathrm {nil}}$
is a polarizable variation of HS of weight
$-1$
over
$U' := U - \Delta $
. Consequently,
$J(\mathbb V^{\mathrm {nil}})$
is a family of compact complex tori over
$U'$
. Denote by
$\nu ^{\mathrm {nil}} : U' \to J(\mathbb V^{\mathrm {nil}})$
the normal function section of it that corresponds to the restriction of (19) to
$U'$
.
Since the monodromy logarithm
$N : \mathbb V^{\mathrm {nil}} \to \mathbb V^{\mathrm {nil}}(-1)$
is a morphism,
$\ker N$
is a subvariation of
$M_{-1}\mathbb V^{\mathrm {nil}}$
. Since it has trivial monodromy, it is constant on each
$L_x'$
and thus extends to a variation over L, which we denote by
${\mathbb K}$
. Since it is constant on each
$L_x$
, it is the pullback along
$\pi $
of its restriction
${\mathbb K}_0$
to
$\Delta $
.
12.2 The Néron model associated to a nilpotent orbit
Here we give the construction of the Néron model for
$\mathbb V^{\mathrm {nil}}$
on the punctured neighbourhood
$U'$
of the zero section of L. It is more general than the construction given in [Reference Green, Griffiths and Kerr10] as we allow the base to have dimension
$>1$
, but it is less general in that we construct Néron models only for nilpotent orbits. Working with nilpotent orbits is sufficient for our purposes and simplifies the construction as nilpotent orbits are filtered by their relative weight filtrations.
Regard
$J(\mathbb V_{\mathbb R})$
as a local system of abelian groups over
$L'$
.
Lemma 12.1. For each
$x\in \Delta $
, there is an exact sequence

where
$G_x$
is the finite abelian group

In more concrete terms

where
${\vec v} \in L_x'$
and
$h : V_{{\vec v},{\mathbb Z}} \to V_{{\vec v},{\mathbb Z}}$
is the local monodromy operator.
Proof. Denote
$\mathbb V\otimes _{\mathbb Z} {\Bbbk }$
by
$\mathbb V_{\Bbbk }$
, where
${\Bbbk }$
is a subring of
${\mathbb R}$
. Since
$V_{\mathbb Z}$
is torsion free, the long exact sequence of cohomology associated to the exact sequence

of local systems over
$L_x'$

Since
$H^1(L_x',\mathbb V_{\mathbb R})$
is torsion free and since
$H^0(L_x',\mathbb V_{\Bbbk }) = K_{\Bbbk }$
, the sequence

is exact, where
$K_{\Bbbk }$
denotes the fibre of
${\mathbb K}_{\Bbbk }$
over
$x\in \Delta $
. Since
$L_x'$
has the homotopy type of a circle, we have the Poincaré duality isomorphism
$H^1(L_x',\mathbb V_{\mathbb Z}) \cong H_0(L_x',\mathbb V_{\mathbb Z})$
. It restricts to an isomorphism on torsion subgroups. The results now follows as
$H^0(L_x',J({\mathbb K})) = K_{\mathbb R}/K_{\mathbb Z}$
.
The groups
$G_x$
form a local system over L. Denote it by
$\mathbb G$
. Denote the local system over L whose restriction to
$L_x$
is
$H^0(L_x',\mathbb V)$
by
$\mathbb T$
. We have a short exact sequence of local systems

All terms in this sequence are admissible variations of MHS over L. The quotient variation
$\mathbb G$
is Tate of weight 0. Define
$J(\mathbb T)$
by defining its fibre over
$x\in \Delta $
to be

There is an exact sequence

over L. The inclusion
$\mathbb T \to \mathbb V^{\mathrm {nil}}$
induces a morphism

over
$L'$
. The Néron model
$\widehat {J}(\mathbb V^{\mathrm {nil}})$
of
$J(\mathbb V^{\mathrm {nil}})$
is the family of abelian complex Lie groups over U obtained by glueing
$J(\mathbb V^{\mathrm {nil}})|_{U'}$
to
$J(\mathbb T)|_U$
by identifying
$J(\mathbb T)|_{U'}$
. In other words

is a pushout square. Its restriction to
$U'$
is
$J(\mathbb V^{\mathrm {nil}})|_{U'}$
and its fibre over the zero section is
$J(\mathbb T)|_\Delta $
. It is a separated slit analytic space. (See [Reference Green, Griffiths and Kerr10, p. 308] for the definition.) If
$N^2 \neq 0$
,
$\widehat {J}(\mathbb V^{\mathrm {nil}})$
is not a manifold.
Remark 12.2. If S is a modular curve and if
$\mathbb V$
is the local system
$R^1 f_\ast {\mathbb Z}(1)$
associated to the universal elliptic curve
$f : \mathcal E \to S$
, then
$J(\mathbb V)$
is the universal elliptic curve. The restriction of
$\mathbb V$
to a punctured neighbourhood
$U'$
of a cusp
$\Delta $
is a nilpotent orbit. So
$J(\mathbb V) = J(\mathbb V^{\mathrm {nil}})$
. The Néron model
$\widehat {J}(\mathbb V)$
over U is the usual Néron model.
This example illustrates another feature of the construction. The family
$J({\mathbb K})$
over U is the trivial
$\mathbb C^\times $
bundle and, with the correct choice of coordinates, the fibre of
$J(\mathbb V)$
over
$q\in U'$
is
$\mathbb C^\times /q^{\mathbb Z}$
. In particular, the map
$J({\mathbb K}) \to J(\mathbb V)$
has infinite degree. In this case it is surjective as
$N^2 = 0$
.
The following result is closely related to the main results of [Reference Green, Griffiths and Kerr10] and [Reference Schnell27].
Theorem 12.3. The normal function
$\nu ^{\mathrm {nil}}$
lifts to a holomorphic section
$\tilde {\nu }^{\mathrm {nil}}$
of
$J(\mathbb T)$
over U and this section projects to a holomorphic section
$\widehat {\nu }^{\mathrm {nil}}$
of
$\widehat {J}(\mathbb V^{\mathrm {nil}})$
whose restriction to
$U'$
is
$\nu ^{\mathrm {nil}}$
.

Proof. We work locally in L. Let X be a contractible open Stein subset of
$\Delta $
. Set

where
$\pi : L \to \Delta $
denotes the projection, and
$Z' = Z\cap U'$
. Recall that the restriction of
$\mathcal E$
(and thus
$\mathcal V$
) to each
$L_x$
is trivial, so we can (and will) identify their fibres
$E_z$
and
$V_z$
over
$z \in Z$
with
$E_x$
and
$V_x$
, respectively, where
$z\in L_x'\cap U$
.
Exactness of
$F^0$
and
$M_0$
implies that there is a holomorphic section
$\mathbf {e}_F$
of
$F^0 M_0 \mathcal E$
over Z that is constant (with respect to the trivialization) on each
$L_x'$
and projects to the section
$1$
of
$\operatorname {Gr}^W_0{\mathbb E}_{\mathbb Z} \cong {\mathbb Z}$
. There is also a multivalued section
$\mathbf {e}_{\mathbb Z}$
of
${\mathbb E}_{\mathbb Z}$
over
$Z'$
that projects to
$1\in H^0(Z',\operatorname {Gr}^W_0 {\mathbb E}^{\mathrm {nil}}$
). Then

is a multivalued holomorphic section of
$\mathcal V$
over
$Z'$
which descends to
$\nu ^{\mathrm {nil}} \in H^0(Z',J(\mathbb V^{\mathrm {nil}}))$
. To prove the theorem, we will show that it lifts to a section of
$J(\mathbb T)$
over
$Z'$
.
Denote the local monodromy operator by h and its logarithm by N. Since h is unipotent, the endomorphisms
$(h-1)$
and N of
$E_{z,{\Bbbk }}$
satisfy

when
${\Bbbk }$
is a subfield of
$\mathbb C$
. The property (16) of the relative weight filtration implies that

Since N is a morphism of type
$(-1,-1)$
,

So
$N \mathbf {e}_F$
is a section of
$F^{-1}M_{-2}\mathcal V$
. The properties of the relative weight filtration imply that there is a holomorphic section
$\mathbf {e}'$
of
$F^0 M_0 \mathcal V$
over Z such that
$N\mathbf {e}_F = N\mathbf {e}'$
. It is constant (with respect to the trivialization) on each
$L_x'\cap U$
. By replacing
$\mathbf {e}_F$
by
$\mathbf {e}_F-\mathbf {e}'$
, we may assume that
$N\mathbf {e}_F = 0$
. That is, the restriction of
$\mathbf {e}_F$
to each
$L_x$
is also a horizontal section of
$\mathcal K$
.
The class of
$\nu ^{\mathrm {nil}}$
in
$H^0(X,\mathbb G)$
is represented by
$(h-1)\mathbf {e}_{\mathbb Z}(z)$
in

If it vanishes at one (and hence all)
$x\in X$
, there is a multivalued section
$\mathbf {v}_{\mathbb Z}$
of
$\mathbb V_{\mathbb Z}$
over
$Z'$
such that
$(h-1)\mathbf {v}_{\mathbb Z} = (h-1)\mathbf {e}_{\mathbb Z}$
. By replacing
$\mathbf {e}_{\mathbb Z}$
by
$\mathbf {e}_{\mathbb Z} - \mathbf {v}_{\mathbb Z}$
, we see that if the class of
$\nu ^{\mathrm {nil}}$
vanishes in
$H^0(Z',\mathbb G)$
, then we may choose
$\mathbf {e}_{\mathbb Z}$
so that
$(h-1)\mathbf {e}_{\mathbb Z} = 0$
. That is,
$\mathbf {e}_{\mathbb Z}$
extends to a section of
${\mathbb K}_{\mathbb Z}$
over Z. Since
$\mathbf {e}_F$
extends to a holomorphic section of
$\mathcal K$
over Z,
$\mathbf {v} = \mathbf {e}_{\mathbb Z} - \mathbf {e}_F$
also extends to a holomorphic section of
$\mathcal K$
over Z. It is constant on each slice
$L_x'\cap U$
and descends to a holomorphic section of
$J({\mathbb K})$
over Z whose restriction to
$Z'$
is a lift of
$\nu ^{\mathrm {nil}}$
. This completes the proof when the obstruction in
$H^0(X,\mathbb G)$
vanishes. We now prove the general case.
Every element of
$H^0(U',J(\mathbb V))$
extends to U. Since
$J(\mathbb V) \cong J(\mathbb V_{\mathbb R})$
, every element of
$G_x$
is the class of a constant section of
$J(\mathbb V)$
over
$U\cap L_x'$
. All such sections lift to sections of
$\widehat {J}(\mathbb T)$
over
$U\cap L_x$
. So, if the class of
$\nu ^{\mathrm {nil}}$
in
$H^0(X,\mathbb G)$
is nonzero, we can find a constant section
$\overline {\nu }$
of
$J(\mathbb V)$
with the same class in
$G_x$
. Since the class of
$\nu ^{\mathrm {nil}}-\overline {\nu }$
vanishes, by the previous paragraph, it lifts to a holomorphic section of
$J({\mathbb K})$
over
$U\cap L_x$
for all
$x \in X$
. Since
$\overline {\nu }$
lifts to a section of
$\widehat {J}(\mathbb T)$
, it follows that
$\nu ^{\mathrm {nil}}$
lifts to a section
$\widehat {\nu }^{\mathrm {nil}}$
of
$\widehat {J}(\mathbb T)$
over Z whose restriction to
$Z'$
projects to
$\nu ^{\mathrm {nil}}$
.
Remark 12.4. The Néron model
$\widehat {J}(\mathbb V)$
of
$J(\mathbb V)$
can be constructed from
$\widehat {J}(\mathbb V^{\mathrm {nil}})$
using the Nilpotent Orbit Theorem in the guise of Proposition 11.2. Just use the function
$\Phi $
to perturb the Hodge filtration of
$\mathbb V^{\mathrm {nil}}$
to obtain the Hodge filtration of
$\mathbb V$
to construct
$J(\mathbb V)$
from
$J(\mathbb V^{\mathrm {nil}})$
. The restrictions of
$\widehat {J}(\mathbb V)$
and
$\widehat {J}(\mathbb V^{\mathrm {nil}})$
to
$\Delta $
are equal. Similarly, the Nilpotent Orbit Theorem implies that the normal function section
$\nu $
of
$J(\mathbb V)$
extends to a section
$\widehat {\nu }$
of
$\widehat {J}(\mathbb V)$
and that the restrictions of
$\widehat {\nu }$
and
$\widehat {\nu }^{\mathrm {nil}}$
to
$\Delta $
are equal.
Since
${\overline {S}}-S$
is a smooth connected divisor
$\Delta $
, the local system
$\mathbb V$
extends to a local system on
${\overline {S}}$
if and only if
$N=0$
. In this case, the variation
$\mathbb V$
extends to a polarized variation of Hodge structure over
${\overline {S}}$
. The fibre over
$x\in \Delta $
is the limit MHS associated to any tangent vector
${\vec v}$
of
${\overline {S}}$
at x that is not tangent to
$\Delta $
. The limit does not depend on the choice of
${\vec v}$
.
Corollary 12.5 [Reference Hain12, Thm. 7.1].
If
$N=0$
, then
$\nu $
extends to a holomorphic section of
$J(\mathbb V)$
over
${\overline {S}}$
.
The derivative of the lift
$\tilde {\nu }^{\mathrm {nil}}$
of
$\nu ^{\mathrm {nil}}$
constructed in Theorem 12.3 is a section of
$F^{-1}\mathcal V\otimes \Omega ^1_{\overline {S}}(\log \Delta )$
. Combined with the Nilpotent Orbit theorem (Proposition 11.2), this implies that the Green–Griffiths invariant of an admissible normal function is logarithmic along the smooth boundary divisor
$\Delta $
.
Corollary 12.6. The Green–Griffiths invariant
$\overline {\delta }(\nu )$
of
$\nu $
extends to a holomorphic section of

over
${\overline {S}}$
.
13 Residual normal functions and the normal rank
The nilpotent Orbit Theorem (Proposition 11.2) implies that the rank of
$\nu $
on S is bounded below by the rank of
$\nu ^{\mathrm {nil}}$
on
$L'$
. In this section we introduce several tools for estimating the rank of
$\nu ^{\mathrm {nil}}$
. The first is the normal rank.
Definition 13.1. The normal rank
$\operatorname {rk}^\perp _x\nu $
of
$\nu $
at
$x\in \Delta $
is defined to be the rank of the restriction of
$\nu ^{\mathrm {nil}}$
to
$L_x'$
. The normal rank
$\operatorname {rk}_\Delta ^\perp \nu $
of
$\nu $
along
$\Delta $
is defined by

The second tool is the residual normal function
$\nu _\Delta $
defined below.
13.1 The residual normal function
$\nu _\Delta $
By Theorem 12.3,
$\nu ^{\mathrm {nil}}$
has a natural lift
$\tilde {\nu }^{\mathrm {nil}}$
to a section of
$J(\mathbb T)$
over a neighbourhood U of
$\Delta $
in L. The order of the class of
$\nu ^{\mathrm {nil}}$
in
$H^0(U,\mathbb G)$
is the least positive integer such that
$k\tilde {\nu }^{\mathrm {nil}}$
is a section of
$J({\mathbb K})$
. Since
${\mathbb K} \subseteq M_{-1}\mathbb V^{\mathrm {nil}}$
,
$k\tilde {\nu }^{\mathrm {nil}}$
descends to a section of
$J(\operatorname {Gr}^M_{-1}\mathbb V^{\mathrm {nil}})$
over U. Set

This is a polarizable variation of Hodge structure over
$\Delta $
of weight
$-1$
.
Definition 13.2. The residual normal function
$\nu _\Delta $
is the restriction to
$\Delta $
of the section of
$J(\mathbb V_\Delta )$
determined by
$k\tilde {\nu }^{\mathrm {nil}}$
.
Remark 13.3. The section
$k\tilde {\nu }^{\mathrm {nil}}$
of
$J({\mathbb K})$
over
$\Delta $
corresponds to an extension

of admissible variations of MHS over
$\Delta $
, which we call the residual variation. This extension, after tensoring with
${\mathbb Q}$
, is a subquotient of
$M_0\mathbb V^{\mathrm {nil}}$
. The variation that corresponds to the residual normal function
$\nu _\Delta $
is
$\mathbb L/M_{-2}$
.
13.2 The real Néron model
The real local system
$\mathbb T_{\mathbb R}$
underlying
$\mathbb T$
has constant fibre
$H^0(L_x',\mathbb V_{\mathbb R})$
over
$L_x$
for each
$x\in \Delta $
. Define

This is a family of compact real tori over L which fits in the exact sequence

The inclusion
$J(\mathbb T_{\mathbb R}) \hookrightarrow J(\mathbb T)$
induces an isomorphism on fundamental groups.
Even though
$J(\mathbb T)|_{U'} \to J(\mathbb V^{\mathrm {nil}})$
has infinite degree when
$N\neq 0$
, the map

is an inclusion.
Remark 13.4. The variation
${\mathbb K}$
is a constant subvariation of
$\mathbb V^{\mathrm {nil}}$
with strictly negative weights. Denote its fibre by K. Since
$K_{\mathbb R} \cap F^0 K = 0$
, the torus
$J(K_{\mathbb R}) := K_{\mathbb R}/K_{\mathbb Z}$
is the kernel of the map

Thus
$J(K_{\mathbb R})$
consists of those extensions of
${\mathbb Z}$
by K that split as an
${\mathbb R}$
-MHS.
Remark 13.5. This is a continuation of Remark 12.2. In the situation described there,
$J({\mathbb K}_{\mathbb R})$
is the trivial
$S^1$
bundle over
$\mathbb D$
. Its image in
$J({\mathbb K})$
, the trivial
$\mathbb C^\times $
bundle, is the unit circle. It is the kernel of the map

The group
$G_\Delta $
is the group of torsion sections of
$J(\mathbb V)$
modulo those that lie in the circle bundle
$J({\mathbb K}_{\mathbb R})$
. The bundle
$J(\mathbb T_{\mathbb R})$
is
$G_x \times J({\mathbb K}_{\mathbb R})$
. Its image in a fibre of
$J(\mathbb V)$
is the real subgroup whose identity component is the image of the unit circle in
$\mathbb C^\times $
and its translates by torsion sections that correspond to the nontrivial elements of
$G_x$
.
13.3 The normal rank
To study the normal rank of
$\nu $
at
$x\in \Delta $
, we restrict the variations to the fibre
$L_x'$
of
$L'$
over x. The open neighbourhood U of
$\Delta $
in L will be as in Section 12.1. Recall that
$U'=U-\Delta $
.
Lemma 13.6. Let
$k>0$
be the order of the class of
$\nu ^{\mathrm {nil}}$
in
$H^0(U\cap L_x',\mathbb G)$
. The restriction of the normal function
$\nu ^{\mathrm {nil}}$
to
$U\cap L_x'$
has rank 0 if and only if the restriction of
$k\widehat {\nu }^{\mathrm {nil}}$
to
$U\cap L_x'$
factors through the inclusion
$\widehat {J}({\mathbb K}_{\mathbb R})|_{U\cap L_x'} \hookrightarrow \widehat {J}(\mathbb V^{\mathrm {nil}})|_{U\cap L_x'}$
.
Proof. First note that multiplication of a normal function by a positive integer does not change the rank. By replacing
$\nu ^{\mathrm {nil}}$
by
$k\nu ^{\mathrm {nil}}$
, we may assume that
$\tilde {\nu }^{\mathrm {nil}}$
is a section of
$J({\mathbb K})$
. Then
$\nu ^{\mathrm {nil}}$
has rank 0 over
$U\cap L_x'$
if and only if it lifts to a constant section of
$\mathbb V_{\mathbb R}$
over
$U\cap L_x'$
. In other words, it lifts to an element of
$H^0(L_x',{\mathbb K}_{\mathbb R})$
and thus descends to a section of
$\widehat {J}({\mathbb K}_{\mathbb R})$
over
$L_x'$
.
Suppose that
$\mathbb D$
is an analytic disk in
${\overline {S}}$
that intersects
$\Delta $
transversally at x. Set
$\mathbb D'= \mathbb D\cap L'$
. Observe that the nilpotent orbit associated to
$\mathbb V|_{\mathbb D'}$
is
$\mathbb V^{\mathrm {nil}}|_{L_x'}$
.
Corollary 13.7. We have
$0 \le \operatorname {rk} \nu ^{\mathrm {nil}}_{U\cap L_x'} \le \operatorname {rk} \nu |_{\mathbb D'} \le 1$
.
Proof. By multiplying
$\nu |_{\mathbb D'}$
and
$\nu ^{\mathrm {nil}}|_{L_x'}$
by a positive integer, we may assume that
$\widehat {\nu }^{\mathrm {nil}}|_{U\cap L_x'}$
lifts to a section of
$J({\mathbb K})$
. It suffices to show that if
$\nu |_{\mathbb D'}$
has rank 0, then so does
$\nu ^{\mathrm {nil}}|_{U\cap L_x'}$
. If
$\nu |_{\mathbb D'}$
has rank 0, then
$\nu (0) \in J(K_{\mathbb R})$
, where K is the fibre of
${\mathbb K}$
over x. Since
$\nu (0) = \widehat {\nu }^{\mathrm {nil}}(0)$
, this implies that
$\nu ^{\mathrm {nil}}|_{U\cap L_x'}$
is a section of
$J({\mathbb K})$
. Lemma 13.6 now implies that
$\nu ^{\mathrm {nil}}|_{U\cap L_x'}$
also has rank 0.
13.4 The residual normal function
$\nu _\Delta $
and the rank of
$\nu $
Proposition 13.8. With the notation above, we have
-
(1)
$\operatorname {rk} \nu \ge \operatorname {rk} \nu ^{\mathrm {nil}}$ ,
-
(2)
$\operatorname {rk} \nu ^{\mathrm {nil}} \ge \operatorname {rk} \nu _\Delta + \operatorname {rk}_\Delta ^\perp \nu $ ,
-
(3)
$\operatorname {rk}_\Delta ^\perp \nu = 0$ if and only if there is a positive integer k such that
$k\nu ^{\mathrm {nil}}$ is a section of
$J({\mathbb K}_{\mathbb R})$ that is constant on each fibre of
$L' \to \Delta $ .
Note that equality may not hold in (2) when
$\pi _1(\Delta ,x)$
acts nontrivially on the kernel of the restriction mapping
$H_0(L_x,W_0\operatorname {End} V_{\vec v}) \to \operatorname {End} K_x$
.
Proof. As previously remarked, the first assertion follows from the Nilpotent Orbit Theorem (Prop. 11.2). The second assertion follows from the commutativity of the diagram

where
$x\in \Delta $
and
${\vec v} \in L_x'$
. The third assertion follows from Lemma 13.6.
Corollary 13.9. If
$\operatorname {rk} \nu ^{\mathrm {nil}} = \operatorname {rk} \nu _\Delta $
, then there is a normal function section
$\overline {\nu }$
of
$J({\mathbb K}_{\mathbb R})$
over
$\Delta $
and a positive integer k such that the diagram

commutes. Consequently, each fibre of the extension of
${\mathbb Z}$
by
${\mathbb K}$
(as variations of MHS) determined by
$k\nu ^{\mathrm {nil}}$
splits after tensoring with
${\mathbb R}$
.
Proof. If
$k\nu ^{\mathrm {nil}}$
is a section of the subtorus
$J({\mathbb K}_{\mathbb R})$
of
$J(\mathbb V^{\mathrm {nil}})$
, it is the pullback along
$L' \to \Delta $
of a section of
$J({\mathbb K}_{\mathbb R})$
over
$\Delta $
as it is constant over each
$L_x'$
. The second assertion follows from Remark 13.4.
Part 5. Higher genus
In this part we use the results of Part 4 and induction to prove that the Ceresa normal function has rank
$3g-3$
when
$g\ge 4$
. The basic idea behind the proof is to approximate the genus g Ceresa normal function by its nilpotent orbit on the normal bundle of the smooth points
$\Delta $
of the divisor
$\Delta _0$
in
$\overline {{\mathcal M}}_g$
. A global monodromy computation allows us to compute the residual normal function. This and the inductive hypothesis imply that the residual normal function
$\nu _\Delta $
has rank
$3g-4$
. The proof is completed by using Corollary 13.9 to show that the normal rank is 1, which will establish the result.
14 Behaviour of the Ceresa normal function near
$\Delta _0$
Suppose that
$g> 3$
. We will apply the results and constructions of Section 12 with

where
$\Delta _+$
is the union of the boundary divisors
$\Delta _j$
where
$j>0$
. Recall that L is the normal bundle of
$\Delta $
in
${\mathcal M}_g$
and that
$L'$
is L with its 0-section removed.
14.1 A moduli description of
$L'$
A point of
$\Delta $
corresponds to a smooth projective curve C of genus
$g-1$
together with an unordered pair
$\{p,q\}$
of distinct points of C. We will identify
$\Delta $
with the quotient of the moduli space
${\mathcal M}_{g-1,2}$
that parameterizes smooth projective curves of genus
$g-1$
with two distinct marked points. The involution
$\sigma $
of
${\mathcal M}_{g-1,2}$
swaps the two points.
The moduli space
${\mathcal M}_{g-1,\vec {2}}$
is the
$\mathrm {\mathbb {G}_m}\times \mathrm {\mathbb {G}_m}$
torsor over
${\mathcal M}_{g-1,2}$
whose fibre over
$(C;p,q)$
is
$T_p' C \times T_q' C$
— ordered pairs
$({\vec v}_p,{\vec v}_q)$
of nonzero tangent vectors of C at p and q. The group
$\mathrm {\mathbb {G}_m}$
acts on
${\mathcal M}_{g-1,\vec {2}}$
by

A point of
${\mathcal M}_{g-1,\vec {2}}/\mathrm {\mathbb {G}_m}$
corresponds to a 4-tuple
$(C;p,q,{\vec v})$
, where C is smooth and projective of genus
$g-1$
,
$(p,q)$
is an ordered pair of distinct points of C, and
${\vec v}$
is a nonzero element of
$T_p C\otimes T_q C$
.
By deformation theory, there is an étale double covering (in the sense of stacks)

The automorphism group of the covering is generated by the involution
$\sigma $
that swaps the two points p and q.
The pullback of the normal function of the genus
$g-1$
Ceresa cycle along
${\mathcal M}_{g-1,\vec {2}} \to {\mathcal M}_{g-1}$
descends to a normal function section of
$J(\Lambda ^3_0 {\mathbb H}_\Delta )$
over
$\Delta $
. Denote it by
$\nu _0$
. We also have the normal function sections
$\kappa _p$
and
$\kappa _q$
of
$J({\mathbb H}_\Delta )$
defined by

where
$K_C$
denotes the canonical divisor class of C.
Proposition 14.1. If
$g\ge 4$
, then

where
$\kappa := (\kappa _p + \kappa _q)/2$
. That is,

Proof. Since
$g\ge 4$
,
$\Delta $
and
$L'$
are quotients of moduli spaces of curves of genera
$\ge 3$
. So we can apply the results of Appendix A. More precisely, we see that (after tensoring with
${\mathbb Q}$
) the normal function sections of
$J(\Lambda ^3 {\mathbb H}_\Delta )$
over
$\Delta $
(resp.
$L'$
) are the normal function sections of its pullback that are invariant under the involution that swaps p and q. Since, after tensoring with
${\mathbb Q}$
,
$\Lambda ^3 {\mathbb H}_\Delta (-1)$
is isomorphic to the direct sum of
${\mathbb H}_\Delta $
and
$\Lambda ^3_0 {\mathbb H}_\Delta (-1)$
, the result follows from the classification result in the appendix.
14.2 The family of nilpotent orbits over
$L'$
Denote by
$\mathbb V^{\mathrm {nil}}$
the family of nilpotent orbits over
$L'$
associated to the
${\mathbb Z}$
variation
$\mathbb V := (\Lambda ^3 {\mathbb H}/\theta \cdot {\mathbb H})(-1)$
over
${\mathcal M}_g$
. Denote the fibrewise nilpotent monodromy operator
$\mathbb V^{\mathrm {nil}} \to \mathbb V^{\mathrm {nil}}$
by N. It satisfies
$N^2 = 0$
. Denote the associated relative weight filtration by

Suppose that
$(C;p,q,{\vec v})$
is in
${\mathcal M}_{g-1,\vec {2}}/\mathrm {\mathbb {G}_m}$
. Set
$C'= C-\{p,q\}$
. Denote the homology class of a small positive loop about p in
$C'$
by
$\mathbf {a}$
(this is the ‘vanishing cycle’). Choose a path in C from q to p and denote its class in
$H_1(C,\{p,q\})$
by
$\mathbf {b}$
. It is well defined mod
$H_1(C)$
.
Denote the first order smoothing associated to
${\vec v}$
of the nodal curve obtained from C by identifying p and q by
$C_{\vec v}$
. This has a limit MHS; it is the fibre of
${\mathbb H}^{\mathrm {nil}}$
over
${\vec v} \in L'$
. The monodromy logarithm N takes
$\mathbf {b}$
to
$\mathbf {a}$
. The graded quotients of the relative weight filtration are:

The involution
$\sigma $
that swaps p and q acts on
$H_1(C_{\vec v})$
by

Denote the weight
$-1$
polarized variation of HS over
$L'$
with fibre
$H_1(C)$
over
$(C;\{p,q\},{\vec v})$
by
${\mathbb H}_\Delta $
. It is pulled back from
$\Delta $
. The
$M_{\bullet }$
graded quotients of
${\mathbb H}^{\mathrm {nil}}$
over
${\mathcal M}_{g-1,\vec {2}}/\mathrm {\mathbb {G}_m}$
are

Denote the symplectic forms of
${\mathbb H}_\Delta $
and
$\operatorname {Gr}^M_{\bullet }{\mathbb H}^{\mathrm {nil}}$
by
$\theta _\Delta \in \Lambda ^2 {\mathbb H}_\Delta $
and
$\theta \in \operatorname {Gr}^M_{-2}\Lambda ^2 {\mathbb H}$
, respectively. We have

Set
$\Lambda ^2_0 {\mathbb H}_\Delta = \Lambda ^2 {\mathbb H}_\Delta /\theta _\Delta $
. It is an irreducible local system over
$\Delta $
and
$L'$
.
Proposition 14.2. The
$M_{\bullet }$
graded quotients of the pullback of
$\mathbb V^{\mathrm {nil}}$
to the double cover
${\mathcal M}_{g-1,\vec {2}}/\mathrm {\mathbb {G}_m}$
of
$L'$
are:

Here we are regarding
${\mathbb H}_\Delta $
as a variation over
$L'$
by pulling it back along the projection
$L' \to \Delta $
.
Proof. Since

we have

The inclusion
$\theta \cdot \operatorname {Gr}^M_{-1}{\mathbb H}^{\mathrm {nil}} \hookrightarrow \Lambda ^3 \operatorname {Gr}^M_{-1}{\mathbb H}^{\mathrm {nil}}$
takes
$\theta \cdot {\mathbb H}_\Delta $
diagonally into the two copies
$\theta _\Delta \cdot {\mathbb H}_\Delta $
and
$(\mathbf {a}\wedge \mathbf {b})\cdot {\mathbb H}_\Delta $
of
${\mathbb H}_\Delta $
. The result follows.
14.3 Computation of
$\nu ^{\mathrm {nil}}$
Here we assume that
$g\ge 4$
. Since
$N^2=0$
,
${\mathbb K} = M_{-1}\mathbb V^{\mathrm {nil}}$
. As explained in Section 12,
$\nu ^{\mathrm {nil}}$
lifts to a section
$\tilde {\nu }^{\mathrm {nil}}$
of
$J(M_{-1}\mathbb V^{\mathrm {nil}})$
and projects to a normal function
$\nu _\Delta $
, which is a section of

over
$\Delta $
.
The first step in understanding
$\nu ^{\mathrm {nil}}$
is to compute the residual normal function
$\nu _\Delta $
. Since
$g-1 \ge 3$
and since

Proposition 14.1 implies that
$\nu _\Delta $
is a linear combination of the normal function
$\nu _0$
and
$\kappa $
defined there.
Proposition 14.3. The residual normal function is a linear combination

where
$c\neq 0$
.
Sketch of proof.
The exact linear combination is easy to compute once one fixes an
$\mathrm {Sp}(H_\Delta )$
splitting of
$\Lambda ^3 {\mathbb H}_\Delta $
. All that matters to us here, though, is that c is nonzero. For this reason, we will not specify a splitting. To establish the result, it is sufficient to work with variations of
${\mathbb Q}$
-MHS.
I will assume familiarity with relative completion of mapping class groups [Reference Hain13]. Fix
${\vec v} \in L_x'$
. We will use it as a base point of
${\mathcal M}_g$
and
$L'$
and x as a base point of
$\Delta $
. Denote the fibres of
${\mathbb E}$
and
${\mathbb H}$
over
${\vec v}$
by E and H, and the fibres of
${\mathbb E}_\Delta $
and
${\mathbb H}_\Delta $
over x by
$E_\Delta $
and
$H_\Delta $
.
Recall that in this section
$S={\mathcal M}_g$
. Denote the Lie algebra of the completion of
$\pi _1({\mathcal M}_g,{\vec v})$
with respect to the standard homomorphism
$\pi _1({\mathcal M}_g,{\vec v}) \to \mathrm {Sp}(H)$
by
$\mathfrak {g}_S$
. Since the MHS on it is a limit MHS, it has two weight filtrations,
$W_{\bullet }$
and
$M_{\bullet }$
. The MHS on
$(\mathfrak {g}_S,M_{\bullet })$
is filtered by
$W_{\bullet }$
. Denote the Lie algebra of the relative completion of
$\pi _1(L',{\vec v})$
with respect to the homomorphism
$\pi _1(L',{\vec v}) \to \mathrm {Sp}(H_\Delta )$
by
$\mathfrak {g}_{L'}$
. We will denote its weight filtration by
$M_{\bullet }$
instead of
$W_{\bullet }$
. We do this because, with this weight filtration,
$\mathfrak {g}_{L'} \to \mathfrak {g}_S$
is a morphism of MHS. Denote the Lie algebra of the relative completion of
$\pi _1(\Delta ,x)$
with respect
$\pi _1(\Delta ,x) \to \mathrm {Sp}(H_\Delta )$
by
$\mathfrak {g}_\Delta $
. We will denote its weight filtration by
$M_{\bullet }$
. (You can think of it as having two weight filtrations
$M_{\bullet }$
and
$W_{\bullet }$
which are equal.)
The homomorphisms

induce Lie algebra homomorphisms

which are morphisms of MHS with respect to the weight filtration
$M_{\bullet }$
.
The monodromy representations of the normal functions
$\nu $
,
$\nu ^{\mathrm {nil}}$
and
$\nu _\Delta $
are related by the commutative diagram

where A is the subgroup of
$\operatorname {Aut} E$
whose Lie algebra isFootnote
5

These induce Lie algebra homomorphisms

The Lie algebra
$\mathfrak {a}$
has a natural MHS induced by that of
$\operatorname {End} E$
. The homomorphisms in the diagram are morphisms of MHS with respect to the weight filtration
$M_{\bullet }$
. Exactness of
$\operatorname {Gr}^M_{\bullet }$
implies that these are determined by the induced maps on the
$M_{\bullet }$
graded quotients. Each
$\operatorname {Gr}^M_{\bullet } \mathfrak {g}_X$
is an
$\mathfrak {sp}(H_\Delta )$
module and all morphisms between them, in this proof, are
$\mathfrak {sp}(H_\Delta )$
equivariant.
The monodromy of the Ceresa cycle over
${\mathcal M}_g$
induces (and is determined by) the
$\mathrm {Sp}(H)$
equivariant isomorphism

(This is the ‘Johnson homomorphism’.) It corresponds to the
$\mathfrak {sp}(H)$
invariant isomorphism

which is a morphism of MHS with respect to
$M_{\bullet }$
. We have to compute its restriction to
$\mathfrak {g}_{L'}$
and its projection to
$\operatorname {End} E_\Delta $
.
Proposition 14.2 implies that the
$\operatorname {Gr}^M_{\bullet }\operatorname {Gr}^W_{\bullet }$
quotients of
$\mathfrak {g}_S/W_{-2}$
are:

The bigraded quotients of E are

and of
$\mathfrak {a}$
are

The action of
$\mathfrak {g}_S/W_{-2}$
on E is determined by the bigraded action

The isomorphism (21) implies that the action of
$\mathfrak {g}_S/W_{-2}$
on E and its associated bigraded version are both faithful. The
$\mathfrak {sp}(H_\Delta )$
invariance of the action and Schur’s Lemma determine the bigraded action up to scaling on each bigraded summand. This will enable us to compute the action of
$\operatorname {Gr}^M_{\bullet }\mathfrak {g}_L$
on
$\operatorname {Gr}^M_{\bullet } E$
.
Before doing this, note that the summand
${\mathbb Q}\mathbf {a}^2$
is spanned by the logarithm of the Dehn twist that corresponds to a small loop that encircles
$\Delta $
. It is the image of the logarithm of the loop in
$L_x'$
that encircles the origin.
The
$\operatorname {Gr}^M_{\bullet }$
quotients of the image of
$\mathfrak {g}_{L'}/M_{-3}$
in
$\operatorname {Gr}^M_{\bullet } \mathfrak {g}_S/W_{-3}$
, and thus in
$\operatorname {End} \operatorname {Gr}^M_{\bullet } E$
as well, are:

Since the homomorphism
$\mathfrak {g}_{L'} \to \mathfrak {g}_\Delta $
is surjective, the diagram (20) and the computation (22) imply that the image of
$\mathfrak {g}_\Delta \to \operatorname {End} E_\Delta $
is an extension of
$\mathfrak {sp}(H_\Delta )$
by
$H_\Delta \oplus \Lambda ^3_0 H_\Delta $
. The result now follows from Proposition 14.1. The copy of
$\Lambda ^3_0 H_\Delta $
in the image corresponds to
$\nu _0$
and the copy of
$H_\Delta $
to
$\kappa $
.
Remark 14.4. A consequence of the proof is that the image of
$\mathfrak {g}_\Delta $
in
$\operatorname {End} K$
is an extension of
$\mathfrak {sp}(H_\Delta )$
by the two step nilpotent Lie algebra whose associated graded is generated by
$H_\Delta \oplus \Lambda ^3_0 H_\Delta $
in weight
$-1$
and whose weight
$-2$
graded quotient is
$\Lambda ^2_0 H_\Delta $
. The Lie subalgebra generated by
$H_\Delta $
is nonabelian and also has weight
$-2$
graded quotient
$\Lambda ^2_0 H_\Delta $
.
Corollary 14.5. We have
$\operatorname {rk} \nu _\Delta \ge \operatorname {rk} \nu _0 + 2$
.
Proof. Suppose that
$x\in \Delta $
is the moduli point of
$(C;\{p,q\})$
. Since the diagram

commutes, where c is the constant appearing in Proposition 14.3, we see that
$\operatorname {rk} \nu _\Delta \ge \operatorname {rk} \nu _0 + \operatorname {rk} \kappa $
. The result follows as, for generic
$(p,q) \in C^2$
,
$d\kappa |_{C^2}$
has rank 2 at
$(p,q)$
.
15 Proof of Theorem 1
The first step is to show
$\nu ^{\mathrm {nil}}$
cannot be a section of
$J({\mathbb K}_{\mathbb R})$
. We do this by restriction to a curve in
$\Delta $
. Let C be a smooth curve of genus
$g-1$
and p a point of C. Set
$C_p' = C-\{p\}$
. There is a morphism
$C_p' \to \Delta $
that takes
$q \in C$
to the moduli point of the nodal curve
$\overline {C}_q$
obtained from C by identifying p with q. The inclusion
$C_p' \hookrightarrow \Delta $
induces a group homomorphism
$\pi _1(C_p',q) \to \pi _1(\Delta ,x)$
, where x is the point of
$\Delta $
that corresponds to
$\overline {C}_q$
. This induces a Lie algebra homomorphism
$\mathfrak {p}(C_p') \to \mathfrak {g}_\Delta $
from the Lie algebra of the unipotent completion of
$\pi _1(C_p',q)$
to
$\mathfrak {g}_\Delta $
. This homomorphism is an injective morphism of MHS, [Reference Hain13]. In particular,
$\mathfrak {p}(C_p')/W_{-3} \to \mathfrak {g}_\Delta /W_{-3}$
is injective. The weight graded quotients of
$\mathfrak {p}(C_p')/W_{-3}$
are
$H_\Delta $
in weight
$-1$
and
$\Lambda ^2 H_\Delta $
in weight
$-2$
. The bracket
$H_\Delta \otimes H_\Delta \to \Lambda ^2 H_\Delta $
is surjective.
As mentioned in Remark 13.3, the section
$\tilde {\nu }^{\mathrm {nil}}$
of
$J({\mathbb K})$
over
$\Delta $
corresponds to an admissible variation of MHS
$\mathbb L$
over
$\Delta $
. The restriction of
$\mathbb L_{{\mathbb Q}}$
to
$C_p'$
is a unipotent variation of MHS with weight graded quotients

Denote it by
$\mathbb L_C$
. In the terminology of [Reference Hain17], it is a unipotent biextension. Since its monodromy representation has nonabelian image, the main result of [Reference Hain17] implies that there is a dense open subset of C over which the fibre of
$\mathbb L_C$
does not split as a real biextension. That is, no positive multiple of the restriction of
$\nu ^{\mathrm {nil}}$
to C is a section of
$J({\mathbb K}_{\mathbb R})$
and so no positive multiple of
$\nu ^{\mathrm {nil}}$
is a section of
$J({\mathbb K}_{\mathbb R})$
. Corollary 13.9 implies that
$\operatorname {rk} \nu ^{\mathrm {nil}}> \operatorname {rk} \nu _\Delta $
.
The inductive hypothesis and Corollary 14.5 imply that

Since
$\operatorname {rk} \nu \ge \operatorname {rk} \nu ^{\mathrm {nil}}$
, we must have
$\operatorname {rk}\nu = 3g-3$
.
Appendix A Normal functions over
${\mathcal M}_{h,m+\vec {r}}$
Denote the category of admissible variations of MHS over a smooth variety X by
$\mathsf {MHS}(X)$
. When
$\mathbb V$
is a polarized variation of Hodge structure over X of weight
$-1$
, the space of normal function sections of
$J(\mathbb V)$
is, by definition,
$\operatorname {Ext}^1_{\mathsf {MHS}(X)}({\mathbb Z}(0),\mathbb V)$
.
Suppose that
$2h-2+m+r> 0$
. Denote by
${\mathcal M}_{h,m+\vec {r}}$
the moduli space (more accurately, stack) that parameterizes isomorphism classes of
$(C;x_1,\dots ,x_m,{\vec v}_1,\dots ,{\vec v}_r)$
where C is smooth, projective of genus h, each
${\vec v}_j \in T_{y_j} C$
is nonzero, and where
$x_1,\dots ,x_m,y_1,\dots ,y_r$
are distinct points of C.
The following is a special case of [Reference Hain16, Thm. A.1] when
$r=0$
. (See also [Reference Hain12, §8].) The
$r>0$
case follows from the same arguments combined with well-known facts about mapping class groups. Here
${\mathbb H}$
denotes the local system over
${\mathcal M}_{h,m+\vec {r}}$
whose fibre over
$(C;x_1,\dots ,x_m,{\vec v}_1,\dots ,{\vec v}_r)$
is
$H_1(C)$
. For
$0 < k \le h$
, we set

It is irreducible and corresponds to the kth fundamental representation of
$\mathrm {Sp}_h$
.
Theorem A.1. Suppose that
$h\ge 3$
and that
$\mathbb V$
is a polarized variation of Hodge structure whose underlying local system corresponds to a nontrivial irreducible
$\mathrm {Sp}_h$
-module. If
$\mathbb V$
is not isomorphic to variations
${\mathbb H}$
or
$\Lambda ^3_0 {\mathbb H}(-1)$
of weight
$-1$
, then

Otherwise,

and

Here
$\nu $
corresponds to the normal function of the Ceresa cycle and
$\kappa _j$
corresponds to the section
$(C;x_1,\dots ,x_m,{\vec v}_1,\dots ,{\vec v}_r) \mapsto (2h-2)x_j - K_C$
of the universal jacobian
$J({\mathbb H}) \to {\mathcal M}_{g,n+\vec {r}}$
, where
$x_j = y_{j-m}$
when
$j> m$
.
Remark A.2. Denote the genus g Torelli group by
$T_g$
. The Johnson homomorphism [Reference Johnson20] induces a homomorphism

where C is a curve that corresponds to the base point of Torelli space. Johnson [Reference Johnson21] proved that it is an isomorphism mod 2-torsion. This implies that there are isomorphisms

One can show that the generator of
$H^1({\mathcal M}_g,\Lambda ^3_0{\mathbb H}_{\mathbb Z})$
corresponds to
$2\tau _g$
and that this is the class of the Ceresa cycle. Moreover, the class map

is an isomorphism. This implies that when
$g\ge 3$
, the normal functions (over
${\mathcal M}_g$
) of the Gross–Schoen cycle and all variants of both the Ceresa cycle and the Gross–Schoen cycle are, mod torsion, integral multiples of the normal function of the Ceresa cycle. The class of the Gross Schoen cycle is easily seen to be

which implies that the normal function of the Gross–Schoen cycle is three times that of the Ceresa cycle.
Acknowledgments
I am especially grateful to Shou-Wu Zhang for asking me about the rank of the Ceresa normal function. Without his interest, this paper would never have seen the light of day. I am also grateful to Gerard van der Geer for bringing his work [Reference van der Geer, Kouvidakis and Alexis7] with Kouvidakis to my attention. This yielded Corollary 3. I am also indebted to Haohua Deng for pointing out several errors, now fixed, in Part 5 of an early draft of this paper and to the referees whose numerous helpful comments and corrections resulted in significant improvements to the paper. Finally, I would like to thank the Simons Foundation for travel support.
Competing interest
The author has no competing interests to declare.
Financial support
Simons travel grant MP-TSM-00002403.