Introduction
Let C be a smooth algebraic curve. The Ceresa cycle is a homologically-trivial algebraic
$1$
-cycle in the Jacobian of C that determines a canonical cycle class modulo algebraic equivalence. A celebrated result of Ceresa in [Reference Ceresa13] says that this cycle is algebraically nontrivial for very general curves of genus at least 3. While it is known to be trivial for hyperelliptic curves, understanding whether the converse holds is an active area of study (see, for example, [Reference Beauville and Schoen11, Reference Bisogno, Li, Litt and Srinivasan12]).
In [Reference Zharkov28], Zharkov defines the tropical Ceresa cycle and a notion of algebraic equivalence for tropical cycles. In an analogy to Ceresa’s original result, he proves that, for very general tropical curves with underlying graph
$K_4$
, the Ceresa cycle is algebraically nontrivial; in other words, there exists a countable union of hypersurfaces in the moduli space of tropical curves overlying
$K_4$
away from which this property holds. The main tool employed in the argument is a tropical homological invariant of tropical curves of genus 3 that vanishes whenever the Ceresa cycle is algebraically trivial. In the present paper, we extend this tool, which we call the Ceresa period
$\alpha (C)$
, to all tropical curves C. However, instead of working with a particular tropical curve C, we find it helpful to follow the approach of [Reference Corey and Li17] and consider the underlying graph G with variable edge lengths
$x_e$
. We define a corresponding ‘universal’ Ceresa period
$\alpha (G)$
that tells us how
$\alpha (C)$
behaves on the moduli space of tropical curves overlying G. Notably, if
$\alpha (G) = 0$
, then
$\alpha (C) = 0$
, and if
$\alpha (G) \neq 0$
, then
$\alpha (C) \neq 0$
for C very general.
Unlike its classical counterpart, the tropical Torelli map is not injective; in particular, there exist hyperelliptic and non-hyperelliptic tropical curves whose Jacobians coincide. Consequently, it is useful to distinguish a more general notion of hyperellipticity. A graph G is said to be of hyperelliptic type if the Jacobian of some tropical curve overlying G is isomorphic to the Jacobian of a hyperelliptic curve. Then our main result is that
Theorem A. G has trivial Ceresa period if and only if G is of hyperelliptic type.
It follows immediately from a result of Corey in [Reference Corey15] that
$\alpha (G) \neq 0$
if and only if G contains either of the graphs
$K_4$
or
$L_3$
as a minor (see Figure 4). Then Corollary 2.8 implies that a very general tropical curve whose underlying graph contains either minor has algebraically-nontrivial tropical Ceresa cycle. In analogy to the classical question, we ask whether this ‘very general’ hypothesis can be removed; that is,
Question B. Does every non-hyperelliptic-type tropical curve have algebraically-nontrivial tropical Ceresa cycle?
Zharkov remarks in [Reference Zharkov28] that the converse to this question is true, that is, every tropical curve of hyperelliptic type has algebraically-trivial Ceresa cycle, but to our knowledge, nothing is yet known about tropical Ceresa cycles on the locus of curves C where
$\alpha (C) = 0$
. Fundamentally, it seems that, as in the algebraic setting, purely homological tools do not have the specificity to see this locus.
In [Reference Corey and Li17], Corey and Li define a Ceresa–Zharkov class
$\mathbf {w}_{\tau }(G)$
using the theory of mapping class groups and the Johnson homomorphism; the definition depends on a hyperelliptic involution
$\tau $
of the genus-g surface into which they embed G. An immediate consequence of Theorem A and [Reference Corey and Li17, Theorem 5.11] is that a graph has trivial Ceresa period in our sense if and only if it has trivial Ceresa–Zharkov class. In fact, in case G is either
$K_4$
or
$L_3$
, there is a choice of
$\tau $
so that
$\alpha (G) = \mathbf {w}_{\tau }(G)$
(after the appropriate identifications); compare Example 4.1 with [Reference Corey and Li17, Proposition 5.7] and Example 4.2 with [Reference Corey and Li17, Proposition 5.9]. This generalizes an observation made in [Reference Corey, Ellenberg and Li16, Remark 3.7]. We believe that this should hold for all graphs:
Conjecture C. Let G be a graph of genus g. Then there exists a hyperelliptic involution
$\tau $
of the surface of genus g for which
$\alpha (G) = \mathbf {w}_{\tau }(G)$
.
This result, if true, would be evidence of a close link between the tropical Ceresa cycle of [Reference Zharkov28] and the tropical Ceresa class of [Reference Corey, Ellenberg and Li16].
A related invariant was recently developed in [Reference Amini, Corey and Monin5]. Also called the tropical Ceresa class, this invariant is the Abel–Jacobi image of the Ceresa cycle in a particular intermediate Jacobian. The authors expect that the Ceresa period is obtained as the image of this tropical Ceresa class under a monodromy map [Reference Amini, Corey and Monin5, §6.4]. Their invariant may be useful in studying rational triviality of the tropical Ceresa cycle and the connection to hyperellipticity; see [Reference Amini, Corey and Monin5, Theorem 3.10 and §7.2].
The Ceresa period
$\alpha (C)$
is an element of
$H^{1}(C,\mathbb {R})^{\wedge 3}$
modulo some integral lattice;
$\alpha (G)$
is defined similarly, albeit with
$\mathbb {R}$
replaced by the polynomial ring R generated by the edge length variables
$x_{e}$
for each
$e \in E(G)$
. The explicit representative
$\Theta (\Upsilon )$
for
$\alpha (G)$
that we construct in §3.1 proves to be a useful tool in demonstrating Theorem A. It has coefficients that are homogeneous of degree 2 in R. By Proposition 3.6, there is an easy combinatorial condition on pairs of edges
$(e,e^{\prime })$
and triples of cycles
$(\gamma _i,\gamma _j,\gamma _k)$
that tells us precisely when
$\Theta (\Upsilon )$
contains a monomial
$x_ex_{e^{\prime }} \epsilon _i \wedge \epsilon _j \wedge \epsilon _k$
up to scaling, where
$\epsilon _{i} \in H^{1}(C,R)$
is the dual vector to
$\gamma _{i}$
. We expect this representative of
$\alpha (G)$
to be helpful in the resolution of Conjecture C.
Outline
In §1, we discuss the necessary background for tropical curves and their Jacobians, as well as for tropical homology and algebraic cycles on real tori with integral structures. We finish the section by recalling Zharkov’s definition of the tropical Ceresa cycle and introducing our invariant
$\alpha (C)$
. In §2, we adapt these concepts so that they are well defined as graph-theoretic invariants. In particular, we define
$\alpha (G)$
and show in §2.5 that it specializes to
$\alpha (C)$
when fixing edge lengths. In §3, we construct the special representative
$\Theta (\Upsilon )$
as described above, proving in §3.3 that both
$\alpha (G)$
and
$\alpha (C)$
are independent of the choice of basepoint that goes into defining the Ceresa cycle. Finally, in §4, we use the tools that are developed in the previous sections to prove Theorem A.
1 Background
1.1 Tropical curves
Let G be a graph, by which we mean a finite, connected multigraph with vertex and edge sets
$V(G)$
and
$E(G)$
, respectively. The valence of a vertex v, denoted
$\operatorname {\mathrm {val}}(v)$
, is the number of half-edges incident to v. A leaf is an edge incident to a vertex of valence one. The genus of G is the quantity
$\# E(G) - \# V(G) + 1$
.
Fix an arbitrary orientation on the edges of G; then each edge e has a head vertex
$e^+$
and tail vertex
$e^-$
. Given the additional data of a length function
, we construct a topological space

where
$[0,\ell (e)] \subset \mathbb {R}$
is a closed interval of length
$\ell (e)$
and the gluing relations on endpoints are given by the incidence relations in G, that is, for all
$e, f \in E(G)$
not necessarily distinct,

where
$(t,e)$
denotes the point t in the closed interval
$[0,\ell (e)]$
associated to e in the disjoint union. Observe that C becomes a metric space with the ‘shortest distance’ metric induced by edge lengths. We say that a metric space C obtained via this construction is a tropical curve. We further say that the edge-weighted graph
$(G,\ell )$
is a model for C, that G underlies C, and that C overlies G.
Many graphs can underlie the same tropical curve. More precisely, consider the relation on pairs
$(G,\ell )$
generated by a single edge refinement: pick an edge e in G and subdivide it into two edges,
$e_{1}$
and
$e_{2}$
, to obtain a new graph
$G^{\prime }$
; fix the length function
$\ell ^{\prime }$
on
$G^{\prime }$
that equals
$\ell $
away from e and satisfies
$\ell ^{\prime }(e_{1}) + \ell ^{\prime }(e_{2}) = \ell (e)$
. The inverse operation is performed at vertices of valence two. If
$(G,\ell )$
is a model for C, then every other model is obtained from
$(G,\ell )$
by a sequence of refinements and inverse refinements. The genus of C is its first Betti number or, equivalently, the genus of any underlying graph. A tropical curve of genus at least two has a unique minimal model with no vertices of valence two.
We remark that our notion of tropical curve is what in the language of 1-dimensional rational polyhedral spaces should better be called a ‘smooth’ tropical curve; see, for instance, [Reference Gross and Shokrieh21, Section 2.3]. Also, our definition departs slightly from that of [Reference Mikhalkin and Zharkov26, Section 3.1], wherein leaves of a graph G are prescribed to have infinite length. This difference is of no critical importance: as we shall see in §4.2, our main object of interest, the Ceresa period, is not affected by contracting all of the leaves.
1.2 Real tori with integral structures
Given a ring R, an R-algebra S and an R-module M, we write as a shorthand and abbreviate the primitive elements
$s \otimes m$
of
$M_S$
by
$sm$
.
Fix a free abelian group N of rank g. Then we may naturally regard N as a (full-rank) lattice inside of the g-dimensional
$\mathbb {R}$
-vector space
$N_{\mathbb {R}}$
. Fix another lattice
$\Lambda \subset N_{\mathbb {R}}$
. The quotient group
with the quotient topology and smooth structure induced from
$N_{\mathbb {R}}$
is a real torus of dimension g. The tangent space at every point of X is canonically isomorphic to the universal cover
$N_{\mathbb {R}}$
, so the lattice N defines what is known as an integral structure on the real torus X. We say that a tangent vector v on X is integral if
$v \in N$
.
1.3 Smooth tropical homology
Recall that singular q-chains on a topological space X are integral formal sums of continuous maps
$\Delta ^q \to X$
, where
$\Delta ^q$
is the standard q-simplex

If X is a smooth manifold, we may restrict our attention to smooth simplices
$\Delta ^q \to X$
; the boundary map on singular chains takes smooth q-chains to smooth
$(q-1)$
-chains, so we obtain the smooth singular chain complex
$S_{\bullet }(X)$
. We write
$H_{q}(X)$
for the q-th homology of
$S_{\bullet }(X)$
. It is well-known that the inclusion of
$S_{\bullet }(X)$
into the usual singular chain complex is a chain homotopy equivalence, and therefore that
$H_{q}(X)$
is isomorphic to the singular homology of X.
We now restrict to the case where
$X = N_{\mathbb {R}}/\Lambda $
is a real torus with integral structure N. Say that a smooth q-simplex
$\sigma $
in X is affine if it is obtained from an affine map
$\mathbb {R}^{q+1} \to N_{\mathbb {R}}$
by restricting to
$\Delta ^q \subset \mathbb {R}^{q+1}$
and pushing forward by the quotient
$N_{\mathbb {R}} \to X$
. Given points
$u_i \in N_{\mathbb {R}}$
, we write
$(u_0,\ldots ,u_q)$
for the unique affine q-simplex defined by mapping the i-th vertex of
$\Delta ^q$
to
$u_i$
. For any
$\lambda \in \Lambda $
,
$(u_0+\lambda ,\ldots ,u_q+\lambda )$
and
$(u_0,\ldots ,u_q)$
represent the same affine simplex of the torus X.
The tropical homology of rational polyhedral spaces was introduced by [Reference Itenberg, Katzarkov, Mikhalkin and Zharkov22] and has been further studied in, for example, [Reference Jell, Rau and Shaw23, Reference Jell, Shaw and Smacka24, Reference Gross and Shokrieh20]. On a real torus X with integral structure as above, a tropical
$(p,q)$
-simplex is a singular q-simplex
$\sigma $
with the additional data of a p-fold wedge product of integral tangent vectors at some point in the image of
$\sigma $
; since we have canonically identified the tangent space at each point of X with the universal cover, this extra data corresponds to an element of
$N^{\wedge p}$
, the p-th exterior power of N. However, we would like to be able to integrate over tropical chains, so we modify the definition by replacing singular chains with smooth ones:

The boundary maps on
$S_{p,\bullet }(X)$
are induced by those on
$S_{\bullet }(X)$
; we write
$H_{p,q}(X)$
for the corresponding q-th homology group. By the universal coefficient theorem and the equivalence of smooth and singular homology noted above,
$H_{p,q}(X)$
is isomorphic to the tropical homology in the sense of [Reference Itenberg, Katzarkov, Mikhalkin and Zharkov22], and in fact,

By the Künneth theorem,
$H_{q}(X) \cong \Lambda ^{\wedge q}$
. Fixing bases
$\epsilon _1,\ldots ,\epsilon _g$
for N and
$\lambda _1,\ldots ,\lambda _g$
for
$\Lambda $
, the Eilenberg–Zilber map for singular homology determines explicit generators

where
$[g] = \{1,\ldots ,g\}$
.
1.4 Tropical algebraic cycles
A (tropical) algebraic k-cycle in a rational polyhedral space X is a balanced, weighted, rational polyhedral complex of pure dimension k, defined up to refinement. For details, we refer to [Reference Allermann and Rau2, Section 2] and [Reference Allermann, Hampe and Rau1, Section 2]. In the case where
$X = N_{\mathbb {R}}/\Lambda $
is a real torus with integral structure N, a polyhedral complex in X is a stratified closed subset that locally lifts to a polyhedral complex in
$N_{\mathbb {R}}$
. We call it weighted if every facet is assigned an integer weight, and rational if the affine hull of every face has a basis that is integral in the sense of §1.2. Algebraic k-cycles form an abelian group
$\mathscr {Z}_k(X)$
, and a morphism of rational polyhedral spaces
induces a pushforward homomorphism
. In analogy with the classical situation, there is a notion of rational equivalence of cycles, which is explored in [Reference Allermann, Hampe and Rau1, Section 3]. Likewise, there is a notion of algebraic equivalence defined by [Reference Zharkov28] and explored further in [Reference Gross and Shokrieh21, Section 5.1].
There exists a group homomorphism , called the cycle class map, from algebraic k-cycles to tropical
$(k,k)$
-homology classes. We note that
$\operatorname {\mathrm {cyc}}$
commutes with pushforward maps. When X is a real torus with integral structure as above and
$Z \in \mathscr {Z}_1(X)$
, we may describe a representative
$(1,1)$
-cycle
$\operatorname {\mathrm {ch}}(Z)$
of the homology class
$\operatorname {\mathrm {cyc}}(Z)$
as follows. (For the general case, see [Reference Jell, Rau and Shaw23, Section 4.C] or [Reference Gross and Shokrieh20, Section 5].) Fix a polyhedral complex structure P for Z. Each facet
$\sigma $
of P is a line segment with rational slope, so we may lift it to a line segment in the universal cover
$N_{\mathbb {R}}$
with endpoints
$u_{\sigma }$
and
$u^{\prime }_{\sigma }$
and primitive tangent vector
$n_{\sigma } \in N$
; that is,
$n_{\sigma }$
is a positive scalar multiple of
$u^{\prime }_{\sigma } - u_{\sigma }$
such that
$a n_{\sigma } \in N$
for precisely
$a \in \mathbb {Z}$
. Then

where the summation is taken over all facets
$\sigma $
of P. The balancing condition at the vertices of P translates to
$\operatorname {\mathrm {ch}}(Z)$
having trivial boundary. We caution that the ‘chain map’
so defined is highly non-canonical; in particular, it depends on a choice of polyhedral complex structure underlying Z for each algebraic 1-cycle Z. We shall have to contend with this fact in §1.8 when we define the Ceresa period
$\alpha (C)$
.
1.5 Tropical Jacobians
The tropical Jacobian is a real torus with integral structure that we can canonically associate to any tropical curve. The original definition given by [Reference Mikhalkin and Zharkov26, Section 6.1] mimics that of the Jacobian of a complex algebraic curve; here, we present an equivalent definition that avoids mention of
$1$
-forms.
Let C be a tropical curve of genus g. Given a model
$(G,\ell )$
of C, let
$C_1(G,\mathbb {Z})$
denote the oriented simplicial 1-chains on G. The corresponding simplicial homology
$H_1(G,\mathbb {Z}) \subset C_1(G,\mathbb {Z})$
of G is isomorphic to the singular homology
$H_1(C,\mathbb {Z})$
of C; in what follows, we conflate
$H_1(G,\mathbb {Z})$
and
$H_1(C,\mathbb {Z})$
without further remark. Furthermore, by the universal coefficient theorem, we may identify
$H_1(C,\mathbb {R})^{\vee }$
with
$H^1(C,\mathbb {R})$
. Fixing an orientation on the edges of G, we define a symmetric, bilinear map
called the length pairing by

For any refinement
$G^{\prime }$
of G, this pairing descends to
$C_1(G^{\prime },\mathbb {Z}) \times C_1(G^{\prime },\mathbb {Z})$
in a way that is compatible with the restriction of either entry to
$H_1(C,\mathbb {Z})$
. Likewise,
$[\cdot ,\cdot ]$
formally extends to allow coefficients in
$\mathbb {R}$
. By the
$\otimes $
–
$\operatorname {\mathrm {Hom}}$
adjunction,
$[\cdot ,\cdot ]$
induces a homomorphism
via
$e \mapsto [e,\cdot ]$
. Consider the lattice

its dual lattice is defined by

The tropical Jacobian of C is the real torus

with integral structure given by N.
Fix a point
$v \in C$
. Following [Reference Baker and Faber8, Section 4], we define the Abel–Jacobi map
based at v by

where
$\delta $
is a path in
$C_1(G,\mathbb {Z})$
from v to w for some model
$(G,\ell )$
of C containing both v and w as vertices. We claim that
$\Phi _v(w)$
is well defined. Indeed, the definition does not depend on
$\delta $
, since any other path
$\delta ^{\prime }$
from v to w satisfies
$\delta ^{\prime } - \delta \in H_1(C,\mathbb {Z})$
, and hence,
$\pi (\delta ^{\prime }) \equiv \pi (\delta )\ \pmod \Lambda $
. Moreover, it is independent of the choice of model
$(G,\ell )$
because the length pairing is preserved under refinement.
1.6 Tropical Ceresa cycle
We identify C with its fundamental algebraic cycle in
$\mathscr {Z}_1(C)$
– that is, the unique 1-cycle with support equal to all of C and with weight one on every edge. We also write
for the inversion map. Following [Reference Zharkov28], we define the tropical Ceresa cycle based at
$v \in C$
by

where
$\Phi _{v,*}$
and
$[-1]_{*}$
are the induced pushforwards on algebraic cycles mentioned in §1.4. Given two basepoints v and w, it is not hard to see that
$\Phi _{v,*}C$
and
$\Phi _{w,*}C$
are related by translation. Then it follows from [Reference Gross and Shokrieh21, Proposition 5.5] that, modulo algebraic equivalence,
${\mathbf {Cer}}_{v}(C)$
does not depend on the choice of basepoint v.
Fix a model
$(G,\ell )$
of C for which
$v \in V(G)$
. This determines a polyhedral complex structure on
$\Phi _{v,*}C$
; with respect to this choice, we define

where
$\operatorname {\mathrm {ch}}$
is the chain map defined in §1.4. By construction, this means that
$\operatorname {\mathrm {Cer}}_{v,G}(C)$
is a
$(1,1)$
-cycle whose class in
$H_{1,1}(\operatorname {\mathrm {Jac}}(C))$
is
$\operatorname {\mathrm {cyc}}({\mathbf {Cer}}_{v}(C))$
. In fact, using the generators of
$H_{1,1}(\operatorname {\mathrm {Jac}}(C))$
given by (1), one sees that
$[-1]_*$
acts as the identity on
$H_{1,1}(\operatorname {\mathrm {Jac}}(C))$
. Then the Ceresa cycle is homologically trivial:

A straightforward computation using the expression for
$\operatorname {\mathrm {ch}}$
in (2) shows that

where
$\delta _e$
is a path in G from v to
$e^-$
. This formula will be key in defining a universal version of the Ceresa cycle in §2.3.
1.7 Integration map
Let
$\epsilon _1,\ldots ,\epsilon _g$
denote a basis for N and
$z_1,\ldots ,z_g$
the associated coordinate functions on
$N_{\mathbb {R}}$
; then the differential 1-forms
$dz_i$
on
$N_{\mathbb {R}}$
descend to
$\operatorname {\mathrm {Jac}}(C)$
. Recall from §1.3 that
$S_{1,2}(\operatorname {\mathrm {Jac}}(C))$
is generated by elements of the form
$n \otimes \sigma $
, where
$n = \sum _{i=1} n_i\epsilon _i \in N$
and
is a smooth
$2$
-simplex. Given a differential form
$\omega $
on
$\operatorname {\mathrm {Jac}}(C)$
, we define the integral of
$\omega $
over
$\sigma $
by

Let
$\binom {[g]}{3}$
denote the collection of all subsets of
$[g]$
of cardinality
$3$
. Given
$I \in \binom {[g]}{3}$
, we write
$I = \{i_1,i_2,i_3\}$
for
$i_1 < i_2 < i_3$
and adopt the multi-index notation
. Mimicking the ‘determinantal 2-form’ for
$K_4$
introduced by [Reference Zharkov28], we define the integration map
via

In the case that
$\sigma = (u_0,u_1,u_2)$
is affine (see §1.3), (5) reduces to

where M is the
$g \times 3$
matrix whose columns are the entries of n,
$u_1-u_0$
, and
$u_2-u_0$
with respect to the basis
$\epsilon _1,\ldots ,\epsilon _g$
. We see that
$\Theta $
is coordinate-independent at least on affine
$(1,2)$
-chains by rewriting this expression in terms of exterior products:

The boundary map sends
$n \otimes \sigma \mapsto n \otimes \partial \sigma $
. Then by Stokes’ theorem,
$\Theta (n \otimes \partial \sigma ) = 0$
, so
$\Theta $
descends to a map on homology
$H_{1,2}(\operatorname {\mathrm {Jac}}(C)) \to N_{\mathbb {R}}^{\wedge 3}$
. Let

Fixing a basis
$\lambda _1,\ldots ,\lambda _g$
for
$\Lambda $
, we may apply (6) to the affine generating set of
$H_{1,2}(\operatorname {\mathrm {Jac}}(C))$
given by (1) to find that

1.8 Ceresa period
We saw in §1.6 that
$\operatorname {\mathrm {Cer}}_{v,G}(C)$
is trivial in
$H_{1,1}(\operatorname {\mathrm {Jac}}(C))$
, so we may choose some
$\Sigma \in S_{1,2}(\operatorname {\mathrm {Jac}}(C))$
for which
$\partial \Sigma = \operatorname {\mathrm {Cer}}_{v,G}(C)$
. Define the Ceresa period
$\alpha (C) \in N_{\mathbb {R}}^{\wedge 3}/\mathscr {P}$
via

As the notation suggests,
$\alpha (C)$
is independent of the choice of v,
$\Sigma $
, and G. We prove independence from the basepoint v in Corollary 3.4. Given
$\Sigma ^{\prime } \in S_{1,2}(\operatorname {\mathrm {Jac}}(C))$
with the same boundary as
$\Sigma $
, we have that
$\Sigma ^{\prime } - \Sigma \in H_{1,2}(\operatorname {\mathrm {Jac}}(C))$
, and hence,
$\Theta (\Sigma ^{\prime }) \equiv \Theta (\Sigma )\ \pmod {\mathscr {P}}$
. Finally, any model
$(G^{\prime },\ell ^{\prime })$
of C is related to
$(G,\ell )$
by a series of edge refinements, so it suffices to prove independence in the case that
$G^{\prime }$
is obtained from G by subdividing a single edge e into
$e_{1}$
and
$e_{2}$
. Without loss of generality, we orient both edges in the direction of e, with
$e^{-} = e_{1}^{-}$
. Then we have both that
$\ell ^{\prime }(e_{1})^{-1}\pi (e_{1}) = \ell ^{\prime }(e_{2})^{-1}\pi (e_{2}) = \ell (e)^{-1}\pi (e)$
, and that
$\delta _{e_{1}} = \delta _{e}$
and
$\delta _{e_{2}} = \delta _{e} + e_{1}$
. Given
$\Sigma $
as above, let

Comparing the e term in
$\operatorname {\mathrm {Cer}}_{v,G}(C)$
to the
$e_{1}$
and
$e_{2}$
terms in
$\operatorname {\mathrm {Cer}}_{v,G^{\prime }}(C)$
, one sees that
$\partial \Sigma ^{\prime } = \operatorname {\mathrm {Cer}}_{v,G^{\prime }}(C)$
. Moreover, since
$\pi (\delta _{e})$
,
$\pi (\delta _{e}+e_{1})$
, and
$\pi (\delta _{e}+e)$
are colinear in
$N_{\mathbb {R}}$
, they vanish under the integration map
$\Theta $
. This forces
$\Theta (\Sigma ) = \Theta (\Sigma ^{\prime })$
, as desired.
Given the data of an algebraic equivalence between
$Z_{1}$
and
$Z_{2}$
in
$\mathscr {Z}_{1}(\operatorname {\mathrm {Jac}}(C))$
, Zharkov constructs in [Reference Zharkov28, §3.1] an affine
$(1,2)$
-chain
that satisfies
$\partial \Sigma = \operatorname {\mathrm {ch}}(Z_{1}) - \operatorname {\mathrm {ch}}(Z_{2})$
[Reference Zharkov28, Lemma 4]. As noted in [Reference Zharkov28, Lemma 5], by construction, each
$n_{i}$
lies in the affine hull of the corresponding simplex
$\sigma _{i}(\Delta ^{2})$
. The following result is a straightforward generalization of [Reference Zharkov28, Lemma 5].
Proposition 1.1. If the tropical Ceresa cycle of C is algebraically trivial, then
$\alpha (C) = 0$
.
Proof. Suppose that
$\Phi _{v,*}C$
and
$[-1]_{*}\Phi _{v,*}C$
are algebraically equivalent. Then we construct
$\Sigma $
as above so that

The fact that
$n_i$
lies in the affine hull of
$\sigma _i(\Delta ^2)$
forces
$\Theta (n_i \otimes \sigma _i) = 0$
by (6). This implies that
$\Theta (\Sigma ) = 0$
, as desired.
2 Ceresa period of a graph
We would like to work with graphs rather than particular tropical curves. To that end, we define ‘universal’ versions of the Jacobian, homology, the Ceresa cycle and the Ceresa period using edge weights in a polynomial ring. We conclude this section by showing that these universal objects specialize to their tropical counterparts when we fix specific edge lengths.
2.1 Universal Jacobian
Given a graph G of genus g with arbitrary edge orientations, we define a polynomial ring . Let
, and consider the free R-module
$N_R \cong H^1(G,R) \cong \operatorname {\mathrm {Hom}}(H_1(G,\mathbb {Z}),R)$
. We define a pairing
via

which induces a homomorphism by
$e \mapsto [e,\cdot ]$
. Then
is a free
$\mathbb {Z}$
-submodule of
$N_R$
of rank g. Formally, we define the universal Jacobian
$\operatorname {\mathrm {Jac}}(G)$
of G to be the triple
$(N_{ {R}}/\Lambda , N_{{R}}, N)$
, although we shall conflate
$\operatorname {\mathrm {Jac}}(G)$
with the quotient group
$N_{ {R}}/\Lambda $
whenever the other data are clear from context.
It should be noted that, although we use the symbol
$\operatorname {\mathrm {Jac}}(G)$
, our universal Jacobian is distinct from the Jacobian of a finite graph, also known as the abelian sandpile group or critical group, which has been studied in various contexts in physics, arithmetic geometry, and graph theory. For more on this subject, see, for instance, [Reference Bak, Tang and Wiesenfeld7, Reference Lorenzini25, Reference Dhar18, Reference Gabrielov19, Reference Bacher, de La Harpe and Nagnibeda6, Reference Nagnibeda27, Reference Baker and Norine9].
2.2 Universal homology
We define homology theories on
$\operatorname {\mathrm {Jac}}(G)$
as follows. Let
$C_q(\operatorname {\mathrm {Jac}}(G))$
denote the free abelian group generated by
$N_R^{q+1}/{\sim }$
, equivalence classes of ordered q-simplices, where we identify

for all
$\lambda \in \Lambda $
. This becomes a chain complex
$C_{\bullet }(\operatorname {\mathrm {Jac}}(G))$
via the usual boundary maps

where
$\hat {\cdot }$
means that the corresponding entry is omitted. Let
$H_q(\operatorname {\mathrm {Jac}}(G))$
denote the q-th homology of
$C_{\bullet }(\operatorname {\mathrm {Jac}}(G))$
. We further define
, with corresponding q-th homology
$H_{p,q}(\operatorname {\mathrm {Jac}}(G))$
.We call a
$(p,q)$
-chain degenerate if every q-simplex that it contains has repeated entries.
The universal coefficient theorem yields

Comparing the following technical result to (1), this says that
$\operatorname {\mathrm {Jac}}(G)$
has the homology we expect of a g-dimensional real torus.
Lemma 2.1. Fix bases
$\epsilon _1,\ldots ,\epsilon _g$
of N and
$\lambda _1,\ldots ,\lambda _g$
of
$\Lambda $
. Then
-
(a)
$H_{1,1}(\operatorname {\mathrm {Jac}}(G)) = \mathbb {Z}\langle {\epsilon _i \otimes (0,\lambda _j) \; | \; i,j \in [g]}\rangle $ and
-
(b)
$H_{1,2}(\operatorname {\mathrm {Jac}}(G)) = \mathbb {Z}\langle {\epsilon _i \otimes \left ((0,\lambda _j,\lambda _j + \lambda _k) - (0,\lambda _k,\lambda _j+\lambda _k)\right ) \; | \; i,j,k \in [g],\, j < k}\rangle $ .
Proof. By restriction of scalars, the R-module
$N_R$
inherits the structure of a free
$\mathbb {Z}$
-module of infinite rank, which we denote by M. Then
$M_{\mathbb {R}}$
is an infinite-dimensional
$\mathbb {R}$
-vector space that naturally contains M as a
$\mathbb {Z}$
-submodule. We endow the g-dimensional subspace
$\Lambda _{\mathbb {R}}$
of
$M_{\mathbb {R}}$
with the Euclidean topology. Choose a subspace W complementary to
$\Lambda _{\mathbb {R}}$
and give it the Euclidean norm with respect to some basis
$w_1,w_2,\ldots $
. Give
$M_{\mathbb {R}} = \Lambda _{\mathbb {R}} \times W$
the product topology. The action of
$\Lambda $
on
$\Lambda _{\mathbb {R}}$
by translation extends to an action on
$M_{\mathbb {R}}$
; define
with the quotient topology. Equivalently, we may write
$J = \Lambda _{\mathbb {R}}/\Lambda \times W$
. It is a straightforward exercise to show that
$W \cong \varinjlim _n \mathbb {R}^n$
is contractible via a straight-line homotopy to
$0$
; therefore, the singular homology
$H_q(J)$
of J is isomorphic to that of the g-dimensional real torus
$\Lambda _{\mathbb {R}}/\Lambda $
. In particular, just as in §1.3, one can show using the Eilenberg–Zilber map that
$H_1(J)$
is freely generated by the affine simplices
$(0,\lambda _i)$
in
$M_{\mathbb {R}}$
for each i, while
$H_2(J)$
is freely generated by
$(0,\lambda _i,\lambda _i+\lambda _j) - (0,\lambda _j,\lambda _i+\lambda _j)$
for
$i < j$
.
Let
$S_{\bullet }$
denote the singular chain complex functor. There is a natural surjection
$S_{\bullet }(M_{\mathbb {R}}) \to S_{\bullet }(J)$
that identifies simplices in
$M_{\mathbb {R}}$
that are
$\Lambda $
-translates of each other, so we may refer to simplices of J by their representatives in
$M_{\mathbb {R}}$
. There is an injective chain map
sending the ordered q-simplex
$(u_0,\ldots ,u_q)$
to the affine singular simplex
$(u_0,\ldots ,u_q)$
; we naturally identify
$C_{\bullet }(\operatorname {\mathrm {Jac}}(G)) \cong \operatorname {\mathrm {im}}\Phi $
. It is clear that the given generators for
$H_q(J)$
for
$q \in \{1,2\}$
are in
$\operatorname {\mathrm {im}}\Phi $
, so the induced map
$H_q(\operatorname {\mathrm {Jac}}(G)) \to H_q(J)$
is surjective. We claim that it is in fact an isomorphism.
It suffices to show that the injection
$\Phi $
splits, since then it must induce an injection on homology. Indeed, let
$A_{\bullet }(J) \subset S_{\bullet }(J)$
denote the subcomplex generated by affine simplices; by construction,
$\operatorname {\mathrm {im}}\Phi \subset A_{\bullet }(J)$
. We define chain maps
sending a singular simplex to the affine simplex with the same vertices and
sending
$(u_0,\ldots ,u_q) \mapsto ( \left \lfloor {u_0} \right \rfloor ,\ldots , \left \lfloor {u_q} \right \rfloor )$
, where
$ \left \lfloor {\cdot } \right \rfloor $
applies the usual floor function to each coordinate of the argument with respect to the basis
$\lambda _1,\ldots ,\lambda _g,w_1,w_2,\ldots $
of
$M_{\mathbb {R}}$
.

The restriction of
$GF$
to
$\operatorname {\mathrm {im}}\Phi $
is the identity, as desired.
Since
$H_q(\operatorname {\mathrm {Jac}}(G)) \to H_q(J)$
is an isomorphism for
$q \in \{1,2\}$
, the chosen generators of
$H_q(J)$
pull back to corresponding generators for
$H_q(\operatorname {\mathrm {Jac}}(G))$
. The final result then follows from (9).
Suppose that G and
$G^{\prime }$
are graphs with universal Jacobians
$\operatorname {\mathrm {Jac}}(G) = N_R/\Lambda $
and
$\operatorname {\mathrm {Jac}}(G^{\prime }) = N^{\prime }_{R^{\prime }}/\Lambda ^{\prime }$
, respectively. Then any
$\mathbb {Z}$
-linear map
for which
$f(\Lambda ) \subset \Lambda ^{\prime }$
and
$f(N) \subset N^{\prime }$
induces a chain map
via

2.3 Universal Ceresa cycle
Fix a basepoint
$v \in V(G)$
and define the Abel–Jacobi map
by sending
$w \mapsto \pi (\delta )$
, where
$\delta $
is a path from v to w. Just as for the Abel–Jacobi map associated to a tropical curve,
$\Phi _v$
is well defined modulo
$\Lambda = \pi (H_1(G,\mathbb {Z}))$
. By an abuse of notation, we also define
by

where
$\delta _e$
is a path from v to
$e^-$
. Notice that
$\pi (\delta _e)$
and
$\pi (\delta _e + e)$
are representatives in
$N_R$
of
$\Phi _v(e^-)$
and
$\Phi _v(e^+)$
, respectively.
Taking (4) as inspiration, we define the (universal) Ceresa cycle of G based at v by

where the inversion map induces a pushforward as in (10).
Lemma 2.2.
$\operatorname {\mathrm {Cer}}_v(G)$
is independent of the orientation on the edges up to adding the boundary of a degenerate
$(1,2)$
-chain.
Proof. Let
$\bar e$
be the edge e with the opposite orientation. We observe that
$\bar e = -e$
in
$C_1(G,\mathbb {Z})$
and
$\bar {e}^- = e^+$
, so we may take
$\delta _{\bar {e}} = \delta _e + e$
. We also have
$x_{\bar {e}} = x_e$
. Then

Lemma 2.3. The
$(1,1)$
-chain
$\sum _{e\in E(G)} \Phi _v(e)$
has trivial boundary.
Proof. We compute

Then
$\partial \left (\sum _{e\in E(G)}\Phi _v(e)\right )$
is supported on the subset
$\Phi _v(V(G)) \subset \operatorname {\mathrm {Jac}}(G)$
, so it suffices to fix
$w \in V(G)$
and show that the part of the boundary supported on
$\Phi _v(w)$
vanishes. The only edges that have a boundary component at
$\Phi _v(w)$
are those that are incident to w. We further restrict our attention to the non-loop edges, since if e is a loop edge,
$\pi (e) \in \Lambda $
, and hence,
$\partial \Phi _v(e) = 0$
.
Let
$\delta $
be a path from v to w, and label the non-loop edges adjacent to w by
$e_1,\ldots ,e_n$
. By Lemma 2.2, up to adding a boundary, we may assume that each
$e_i$
is oriented away from w. We write
$\Phi _v(e_i) = x_{e_i}^{-1}\pi (e_i) \otimes (\pi (\delta ),\pi (\delta +e_i))$
. This contributes a boundary component of
$-x_{e_i}^{-1}\pi (e_i) \otimes (\pi (\delta ))$
at
$\Phi _v(w)$
, so we need only show that
$\sum _{i=1}^n x_{e_i}^{-1}\pi (e_i) = 0$
or, equivalently, that
$\sum _{i=1}^n x_{e_i}^{-1}[e_i,\gamma ] = 0$
for all
$\gamma \in H_1(G,\mathbb {Z})$
. Indeed, we may write
$\gamma = \sum _{i=1}^n c_i e_i + \gamma ^{\prime }$
, where
$\gamma ^{\prime }$
is supported away from the edges
$e_1,\ldots ,e_n$
. The fact that
$\partial \gamma = 0$
forces
$\sum _{i=1}^n c_i = 0$
, so

Immediately, we get that
$\operatorname {\mathrm {Cer}}_{v}(G)$
defines a
$(1,1)$
-cycle, and so descends to a homology class in
$H_{1,1}(\operatorname {\mathrm {Jac}}(G))$
. In fact, more is true:
Lemma 2.4.
$\operatorname {\mathrm {Cer}}_v(G)$
is trivial in
$H_{1,1}(\operatorname {\mathrm {Jac}}(G))$
.
Proof. By the construction of
$\operatorname {\mathrm {Cer}}_v(G)$
and Lemma 2.3, it suffices to show that
acts as the identity. We need only check this on the generators described in Lemma 2.1:

where the second equality follows by equivalence of simplices under translation by
$\Lambda $
.
2.4 Universal Ceresa period
We take exterior powers of
$N_R \cong R^g$
with respect to its R-module structure. We no longer have a topology, let alone a smooth structure for integration, but we can still mimic the effect of
$\Theta $
on affine simplices given by (6). Therefore, we define
by

A direct computation shows that
$\Theta \circ \partial = 0$
, so
is well defined. Let
$\Sigma \in C_{1,2}(\operatorname {\mathrm {Jac}}(G))$
be such that
$\partial \Sigma = \operatorname {\mathrm {Cer}}_v(G)$
. Then the (universal) Ceresa period of G is the image of
$\Theta (\Sigma )$
in
$N_R^{\wedge 3}/\mathscr {P}$
. Just as for
$\alpha (C)$
in §1.8,
$\alpha (G)$
depends neither on the choice of
$\Sigma $
nor on the basepoint
$v \in V(G)$
, although the latter fact will not be proved until Corollary 3.3. Although we will need the tools of §3 to compute
$\alpha (G)$
efficiently, see §4.1 for the minimal examples of nontrivial Ceresa periods.
2.5 Evaluation
Fix a tropical curve C and a model
$(G,\ell )$
. We let
$[\cdot ,\cdot ]$
continue to denote the length pairing with values in R and adopt the notation
$[\cdot ,\cdot ]_C$
for the pairing defined by (3). Likewise, we adopt the notations
$\pi _C$
,
$\Lambda _C$
,
$\Theta _C$
, and
$\mathscr {P}_C$
for the ‘real edge length’ analogues of the ‘universal’
$\pi $
,
$\Lambda $
,
$\Theta $
, and
$\mathscr {P}$
, respectively. The length function
$\ell $
induces a ring map
via
$x_e \mapsto \ell (e)$
. Then we may write
$\operatorname {\mathrm {Jac}}(G) = N_R/\Lambda $
and
$\operatorname {\mathrm {Jac}}(C) = N_{\mathbb {R}}/\Lambda _C$
, where
$N = H^1(G,\mathbb {Z})$
. The evaluation map
$\mathrm {ev}$
induces a map
$N_R \to N_{\mathbb {R}}$
that we also denote by
$\mathrm {ev}$
; this in turn induces a map
$N_R^{\wedge 3} \to N_{\mathbb {R}}^{\wedge 3}$
.
Proposition 2.5. If C is a tropical curve with underlying graph G, then
$\alpha (C) = \mathrm {ev}(\alpha (G))$
.
Proof. By definition,
$\mathrm {ev}(N) = N$
. Moreover, it follows from the definitions given in §§1.5 and 2.1 that
$[\cdot ,\cdot ]_C = \mathrm {ev} \circ [\cdot ,\cdot ]$
, and hence,
$\pi _C = \mathrm {ev} \circ \pi $
and
$\Lambda _C = \mathrm {ev}(\Lambda )$
. As a result, we obtain a chain map
sending
$(u_0,\ldots ,u_q)$
to the affine q-chain
$(\mathrm {ev}(u_0),\ldots ,\mathrm {ev}(u_q))$
, which extends to
$(p,q)$
-chains by preserving the
$N^{\wedge p}$
-component. Comparing (4) and (12), one sees that
$\operatorname {\mathrm {Cer}}_{v,G}(C) = \mathrm {ev}_{*}\operatorname {\mathrm {Cer}}_v(G)$
.
Now fix
$\Sigma \in C_{1,2}(\operatorname {\mathrm {Jac}}(G))$
for which
$\partial \Sigma = \operatorname {\mathrm {Cer}}_v(G)$
. Then the affine
$(1,2)$
-chain
satisfies
$\partial \Sigma _C = \operatorname {\mathrm {Cer}}_{v,G}(C)$
. One sees also that
$\mathrm {ev}_{*}$
sends the generators of
$H_{1,2}(\operatorname {\mathrm {Jac}}(G))$
given by Lemma 2.1 to the generators of
$H_{1,2}(\operatorname {\mathrm {Jac}}(C))$
in (1). Moreover, (6) ensures that
$\Theta _C \circ \mathrm {ev}_{*} = \mathrm {ev} \circ \Theta $
; in particular,
$\mathscr {P}_C = \mathrm {ev}(\mathscr {P})$
and
$\Theta _C(\Sigma _C) = \mathrm {ev}(\Theta (\Sigma ))$
. Then
$\alpha (C) = \mathrm {ev}(\alpha (G))$
, as desired.
Corollary 2.6. If
$\alpha (G) = 0$
, then
$\alpha (C) = 0$
.
Proof. This follows immediately from Proposition 2.5.
We naturally associate to a graph G a moduli space of isomorphism classes of tropical curves overlying G as follows (but for more general constructions, see, for example, [Reference Chan14, Section 4.1]). The group
$\operatorname {\mathrm {Aut}}(G)$
acts on
$E(G)$
; this induces an action on
$\mathbb {R}_{>0}^{E(G)}$
by permuting the coordinates. Then let

Each point of
$\mathscr {M}_G^{\mathrm {tr}}$
determines a length function
$\ell $
on G and hence the tropical curve with
$(G,\ell )$
as a model. In analogy to the classical setting, we say that a property is true of a very general tropical curve overlying G if it is true of every tropical curve corresponding to a point on
$\mathscr {M}_G^{\mathrm {tr}} \setminus H$
, where H is some countable union of hypersurfaces. Then Corollary 2.6 has the following partial converse.
Proposition 2.7. If
$\alpha (G) \neq 0$
, then a very general tropical curve overlying G has nontrivial Ceresa period.
Proof. Let C be a tropical curve with
$(G,\ell )$
as a model, and adopt the notation used in the proof of Proposition 2.5. If we choose generators
$p_1,\ldots ,p_r$
of
$\mathscr {P}$
, then
$\mathrm {ev}(p_1),\ldots ,\mathrm {ev}(p_r)$
generate
$\mathscr {P}_C$
. Suppose that
$\alpha (C) = 0$
. Then by definition,
$\Theta _C(\Sigma _C) \in \mathscr {P}_C$
, so we may write

for some
$a_i \in \mathbb {Z}$
, and hence,

Observe that
$S \in N_R^{\wedge 3}$
. Since
$N_R^{\wedge 3} \cong R \otimes _{\mathbb {Z}} N^{\wedge 3}$
is a free R-module of finite rank, we may fix a basis and write S as an R-linear combination of the basis elements. Then
$S \in \ker \mathrm {ev}$
precisely when all of its finitely many R-coefficients vanish under
$\mathrm {ev}$
. The fact that
$\alpha (G) \neq 0$
implies that
$S \neq 0$
, so these coefficients cut out a subvariety of
$\mathbb {R}_{> 0}^{E(G)}$
of positive codimension. The desired statement follows by considering all possible choices of the
$a_i$
, of which there are countably many.
Corollary 2.8. If
$\alpha (G) \neq 0$
, then a very general tropical curve overlying G has nontrivial Ceresa cycle.
3 Computational tools
3.1 Explicit
$(1,2)$
-chain
$\Upsilon $
whose boundary is
$\operatorname {\mathrm {Cer}}_v(G)$
Given a graph G with basepoint
$v \in V(G)$
, we shall define a
$(1,2)$
-chain
$\Upsilon $
for which
$\partial \Upsilon = \operatorname {\mathrm {Cer}}_v(G)$
. To that end, we fix orientations on the edges of G and a spanning tree T and label the edges of
$G \setminus T$
by
$e_1,\ldots ,e_n$
. Each
$e_i$
determines a simple cycle
$\gamma _i$
that uses only edges from
$E(T) \cup \{e_i\}$
and that has positive
$e_i$
-coefficient. Order the edges of
$\gamma _i$
cyclically. Then
$H_1(G,\mathbb {Z})$
has
$\gamma _1,\ldots ,\gamma _g$
as a basis; let
$\epsilon _1,\ldots ,\epsilon _g$
be the dual basis for N, that is,

For
$e \in E(G)$
, the integer
counts the multiplicity of e in
$\gamma _i$
; by our choice of homology basis,
$f_i(e) \in \{0,\pm 1\}$
. Observe also that

From (12), we write

Fix the unique path
$\delta _i$
from v to a vertex
$p_{i,0}$
in
$\gamma _i$
so that
$\delta _i$
is contained in T and does not share any edges with
$\gamma _i$
. Following the cyclic orientation of
$\gamma _i$
, label the subsequent vertices by
$p_{i,1},p_{i,2},\ldots $
with
$p_{i,n_i} = p_{i,0}$
. Write
$e_{i,j}$
for the edge in
$\gamma _i$
with vertices
$p_{i,j}$
and
$p_{i,j+1}$
(see Figure 1). For
$0 \leq j \leq n_i$
, we define

Notice that
$\delta _i + \sum _{k=0}^{j-1} f_i(e_{i,k})e_{i,k}$
is a path from v to
$p_{i,j}$
that passes through
$p_{i,0},p_{i,1},\ldots ,p_{i,j-1}$
; in particular,
$u_{i,j}$
is a lift of
$\Phi _v(p_{i,j})$
to
$N_R$
, and (14) reduces to

the
$\epsilon _i$
-component of which is depicted in Figure 2(a). This cumbersome notation is necessary because, unlike for simplicial homology, we are not able to permute the vertices of a simplex at the cost of introducing a sign. The alternative that works in the current context is the following:

In other words, we must add the boundary of a degenerate
$(1,2)$
-chain in order to permute the vertices of each
$(1,1)$
-simplex.

Figure 1 Notation for the cycle
$\gamma _i.$

Figure 2 Notation for and
$\Upsilon$
.
Now that we have expressed
$\operatorname {\mathrm {Cer}}_v(G)$
more explicitly, we define
, where

Using the fact that
$u_{i,n_i} = u_{i,0} + \lambda _i$
, a straightforward computation yields

which by (16) equals the
$\epsilon _i$
-component of (15). This is depicted without orientations or degenerate simplices in Figure 2. We conclude that
$\partial \Upsilon = \operatorname {\mathrm {Cer}}_v(G)$
.
Remark 3.1. Notice that each
$u_{i,j}$
is independent of the orientations on the edges of G. Consequently,
$\Upsilon $
is also independent of the orientations.
3.2 Explicit representative
$\Theta (\Upsilon )$
for
$\alpha (G)$
Since
$\Theta $
kills degenerate
$(1,2)$
-simplices, we may ignore the second summation in (17) in the following computation:

Since we have bases
$\epsilon _1,\ldots ,\epsilon _g$
for N and
$\pi (\gamma _1),\ldots ,\pi (\gamma _g)$
for
$\Lambda $
, Lemma 2.1 tells us that

Applying
$\Theta $
, we find that

Since
$\pi (\gamma _j) = \sum _{m=1}^g [\gamma _j,\gamma _m] \epsilon _m$
, we may rewrite

In other words, the
$\epsilon _i \wedge \epsilon _m \wedge \epsilon _n$
-coefficient of
$2\epsilon _i \wedge \pi (\gamma _j) \wedge \pi (\gamma _k)$
is twice the
$(j,k)\times (m,n)$
-minor of the Gram matrix of the pairing
$[\cdot ,\cdot ]$
.
3.3 Basepoint independence of the Ceresa period
Proposition 3.2. Fix a graph G and let
$\Upsilon $
be the
$(1,2)$
-chain defined by (17). Then
$\Theta (\Upsilon )$
does not depend on the choice of basepoint.
Proof. Since G is connected, it suffices to show that
$\Theta (\Upsilon )$
remains unchanged whether computed using v as the basepoint or some vertex
$v^{\prime }$
separated from v by an edge a. Without loss of generality, assume that a is oriented from
$v^{\prime }$
to v.
We observe that, in the computations in §3.1, as well as in deriving (18), we do not use any properties of
$\delta _i$
other than that it is a path from v to some vertex
$p_{i,0}$
in the cycle
$\gamma _i$
. In particular, if we construct
$\Upsilon ^{\prime }$
as in §3.1 using the basepoint
$v^{\prime }$
and the paths
, then the same argument shows that
$\partial \Upsilon ^{\prime } = \operatorname {\mathrm {Cer}}_{v^{\prime }}(G)$
. Moreover, the points
$p_{i,0}$
have not changed, nor have the edge labels
$e_{i,j}$
, so it follows from (18) that

Ordering so that
$i < j$
and using the fact that
$\epsilon _j \wedge \epsilon _i = -\epsilon _i \wedge \epsilon _j$
, we obtain

Finally,
$[\cdot ,\cdot ]$
is symmetric, so the summation vanishes and we are left with
$\Theta (\Upsilon ^{\prime }) = \Theta (\Upsilon )$
, as desired.
Corollary 3.3. The Ceresa period
$\alpha (G)$
does not depend on the choice of basepoint.
Proof. This follows immediately from Proposition 3.2 and the fact that
$\Theta (\Upsilon )$
is a representative of
$\alpha (G)$
.
Corollary 3.4. Let C be a tropical curve overlying G. Then
$\alpha (C)$
does not depend on the choice of basepoint.
Proof. Fix
$v, v^{\prime } \in C$
. After refining G if necessary, we may assume that
$v, v^{\prime } \in V(G)$
. We recall from Proposition 2.5 that
$\alpha (C) = \mathrm {ev}(\alpha (G))$
. But Corollary 3.3 implies that
$\alpha (G)$
is the same whether computed using v or
$v^{\prime }$
, so the same must be true for
$\alpha (C)$
.
3.4 Rewriting
$\Theta (\Upsilon )$
in terms of edge pairs
Lemma 3.5. The variable
$x_{e_i}$
does not appear in any coefficient of
$\Theta (\Upsilon )$
for any
$i \in [g]$
.
Proof. For each edge e,
$\pi (e) = x_e\sum _{i=1}^g f_i(e)\epsilon _i \in x_eN$
. In particular, the variable
$x_{e_i}$
appears in
$\pi (e)$
only for
$e = e_i$
. Since no
$\delta _j$
contains
$e_i$
, and
$\gamma _j$
contains
$e_i$
only for
$j = i$
, the only computation
$\Theta (\epsilon _j \otimes \Upsilon _j)$
in which
$\pi (e_i)$
appears is for
$j = i$
(see (18)). But every term of
$\Theta (\epsilon _i \otimes \Upsilon _i)$
with
$\pi (e_i) = x_{e_i}\epsilon _i$
as a factor also has
$\epsilon _i$
as a factor, so the alternating property implies that these terms all vanish.
It is not hard to see from (18) that
$\Theta (\Upsilon )$
is homogeneous of degree 2 in the variables
$x_e$
. We now develop a characterization for when a pair of edges e and
$e^{\prime }$
contributes a monomial
$x_ex_{e^{\prime }}\, \epsilon _i \wedge \epsilon _j \wedge \epsilon _k$
to
$\Theta (\Upsilon )$
. Recall from §3.1 that

Likewise, we define

Given distinct edges e and
$e^{\prime }$
in
$\gamma _i$
, say that
$e <_i e^{\prime }$
if e occurs before
$e^{\prime }$
in the ordering determined by the endpoint of
$\delta _i$
and the cyclic orientation of
$\gamma _i$
. Equivalently, in the notation of Figure 1, we declare
$e_{i,j} <_i e_{i,j+1}$
for
$0 \leq j \leq n_i - 2$
. Then let

Fix a subset
$S \subset E(G) \times E(G)$
that contains exactly one element from
$\{(e,e^{\prime }),(e^{\prime },e)\}$
for each pair of distinct edges e and
$e^{\prime }$
. Likewise, in lieu of requiring that
$i < j < k$
, we instead fix a subset
$S^{\prime } \subset [g]^3$
that contains exactly one permutation of each tuple
$(i,j,k) \in [g]^3$
of all distinct indices.
Proposition 3.6. Fix a graph G and let
$\Upsilon $
be the
$(1,2)$
-chain defined by (17). Then we may write

where
$a_{i,j,k}(e,e^{\prime })$
takes values in
$\{0,\pm 2\}$
and is nonzero precisely when, up to relabeling
$(i,j,k)$
,
$\gamma _i$
contains both e and
$e^{\prime }$
,
$\gamma _j$
contains e but not
$e^{\prime }$
, and
$\gamma _k$
contains
$e^{\prime }$
but not e. In particular,

and reversing the cyclic orientation of any one of the cycles
$\gamma _i$
,
$\gamma _j$
, or
$\gamma _k$
flips the sign.
Proof. From (18), we may write

We may write
$\gamma _i = \sum _{e \in E(G)} f_i(e)e$
and
$\delta _i = \sum _{e \in E(G)} g_i(e)e$
, so that

Expanding
$\pi (e) = x_e\sum _{j=1}^gf_j(e)\epsilon _j$
and
$\pi (e^{\prime }) = x_{e^{\prime }}\sum _{k=1}^gf_k(e^{\prime })\epsilon _k$
, we obtain

We reindex by
$S^{\prime }$
to get

where

As one might expect, reordering the tuple
$(e,e^{\prime })$
does not affect
$a_{i,j,k}(e,e^{\prime })$
, while permuting
$(i,j,k)$
changes it by the sign of the permutation. We record here the possible combinations of values of
${{\lvert {f_{\cdot }(\cdot )} \rvert }}$
up to such reordering:

A
$*$
indicates that the corresponding entry can have value either
$0$
or
$1$
. While these cases are not all pairwise disjoint, they do cover all the possibilities. Indeed, up to permuting the rows and columns, if zero or one of the six values is
$0$
, then we are in case (a). If exactly two values are
$0$
, then we fall into one of (b), (c) or (d). If three are
$0$
, then we are in cases (c) or (d). Finally, if at least four values are
$0$
, then case (d) applies.
We claim that the
$a_{i,j,k}(e,e^{\prime }) \neq 0$
only in case (b). As a shortcut, we may consider only edges e and
$e^{\prime }$
in T; by Lemma 3.5, the remaining edges do not contribute terms. Then
$T \setminus \{e,e^{\prime }\}$
has three connected components; let
$G^{\prime }$
be the graph obtained from G by contracting each of these components to a vertex and removing the additional edges
$e_t$
for
$t \not \in \{i,j,k\}$
. We depict in Figure 3 the resulting graph for each of the cases (a), (b) and (c) with the edges
$e_i,e_j,e_k$
drawn as necessary. The basepoint v descends to one of the three vertices of
$G^{\prime }$
, but by Proposition 3.2, we may fix it arbitrarily. We remark that the values of
$f_t$
,
$g_t$
and
$h_t$
on e and
$e^{\prime }$
remain unchanged for
$t \in \{i,j,k\}$
when passing from G to
$G^{\prime }$
.
The following observations will be helpful in further narrowing down the cases. By Remark 3.1,
$\Theta (\Upsilon )$
is independent of the orientations on the edges of G. Then without loss of generality, we may orient the edges of
$G^{\prime }$
as shown in Figure 3. Meanwhile, replacing
$\gamma _i$
with the cycle
having the opposite cyclic orientation of edges (i.e., the reverse of the partial order
$<_i$
) changes the sign of
$f_i$
and
$h_i$
. Hence, only
$A_i$
,
$B_{i,k}$
, and
$B_{i,j}$
change sign, thereby causing
$a_{i,j,k}(e,e^{\prime })$
also to change sign. The same applies to the indices j and k. Notably, this means that the choice for each t of
$\gamma _t$
versus
$\bar \gamma _t$
does not affect whether or not
$a_{i,j,k}(e,e^{\prime }) = 0$
, so in Figure 3, we orient
$\gamma _t$
(not depicted) in the same direction as
$e_t$
. Finally, recall from §3.1 that the path
$\delta _t$
is defined uniquely by the fact that it lies in T and has disjoint support from the cycle
$\gamma _t$
.

Figure 3 Cases for the contracted graph
$G'$
.
To reiterate, all of the choices that went into drawing Figure 3 were made without loss of generality up to a sign. Therefore, one may read off the values of
$f_t$
,
$g_t$
, and
$h_t$
on e and
$e^{\prime }$
for each of the relevant indices of t from
$\{i,j,k\}$
; one finds that
$a_{i,j,k}(e,e^{\prime })$
vanishes in cases (a) and (c) and equals
$-2$
in case (b). That
$a_{i,j,k}(e,e^{\prime })$
vanishes in case (d) is not hard to see, since
$f_k(e) = f_k(e^{\prime }) = 0$
forces
$B_{j,k} = B_{i,k} = A_k = 0$
. The particular expression for the coefficient given by (21) in the statement of the result follows from case (b) by plugging into
$a_{i,j,k}(e,e^{\prime })$
only the values of the indicator functions that vanish.
4 Forbidden minor characterization of nontriviality of the Ceresa period
With the powerful tools of §3 in hand, we now endeavor to prove Theorem A by following the approach of [Reference Corey and Li17]. In §4.1, we show that
$K_4$
and
$L_3$
, the graphs depicted in Figure 4, have nontrivial Ceresa period. In §§4.2 and 4.3, we show that having nontrivial Ceresa period is preserved under certain contraction and deletion operations, respectively. Finally, in §4.4, we show that graphs of hyperelliptic type have trivial Ceresa period and that there are no other cases left to consider.

Figure 4 Minimal graphs with nontrivial Ceresa period.
Throughout this section, whenever we are working with two graphs G and
$G^{\prime }$
, for each object O associated to G that was defined in §§2 and 3, we let
$O^{\prime }$
denote the corresponding object for
$G^{\prime }$
.
4.1 Base cases
Example 4.1. Let be the graph with basepoint v and oriented edges
$e_i$
as depicted in Figure 4(a), with
$T = \{e_4,e_5,e_6\}$
, paths
$\delta _i = 0$
for all i, and the cyclic orientation of
$\gamma _i$
chosen to match the orientation of
$e_i$
. Let
. Then
$[\cdot ,\cdot ]$
has Gram matrix

One can show using either (18) or Proposition 3.6 that

Meanwhile, (19) and (20) imply that
$\mathscr {P}$
is generated by the elements

Simplifying, we are left with six generators:

Suppose that we could obtain
$\Theta (\Upsilon )$
as some
$\mathbb {Z}$
-linear combination of these generators. Since
$x_1x_4$
,
$x_1x_2$
,
$x_1x_3$
, and
$x_2x_3$
each appear in only one generator and not in
$\Theta (\Upsilon )$
, those generators cannot contribute. This leaves the second and the sixth. Continuing along the same lines, the second generator uniquely contains the term
$x_2x_5$
, so it must be trivial. Now the sixth generator uniquely contains
$x_3x_6$
, so it also vanishes, a contradiction. In other words,
$\alpha (K_4) \neq 0$
, with

Example 4.2. Consider the graph with basepoint v and oriented edges
$e_i$
as shown in Figure 4(b), with
$T = \{e_5,e_6\}$
, paths
$\delta _i = 0$
for all i, and the cyclic orientation of
$\gamma _i$
matching
$e_i$
. We again abbreviate
. Then
$[\cdot ,\cdot ]$
has Gram matrix

One can show that

By Corollary 2.6, to prove that
$\alpha (L_3) \neq 0$
, it suffices to show that some tropical curve overlying
$L_3$
has nontrivial Ceresa period. Indeed, let C be the tropical curve obtained by evaluating every edge length
$x_i$
to
$1$
, with the corresponding map
. Then
$\mathscr {P}_C$
is generated by the four elements

Let . Then
$\Theta (\Upsilon _C) = 2\epsilon _1 \wedge \epsilon _2 \wedge \epsilon _4 + 2\epsilon _1 \wedge \epsilon _2 \wedge \epsilon _3$
, which is clearly not in
$\mathscr {P}_C$
, as desired. Therefore,
$\alpha (L_3) \neq 0$
, with

4.2 Contraction
Given a graph G with a non-loop edge a, we let
$G / a$
denote the graph obtained from G by contracting a. We identify
$E(G) = E(G / a) \sqcup \{a\}$
in the usual way.
Proposition 4.3. Let G be a graph with a non-loop edge a. If
$\alpha (G) = 0$
, then
$\alpha (G/a) = 0$
. Moreover, if a is a separating edge, then the converse also holds.
Although the explicit
$(1,2)$
-chain
$\Upsilon $
is not strictly necessary for proving the first part of this result, it is nonetheless helpful in simplifying the argument.
Proof. Let . We shall declare
to be the basepoint of both G and
$G^{\prime }$
. Fix a spanning tree T of G containing a and let
be the corresponding spanning tree in
$G^{\prime }$
. Let
$\gamma _1,\ldots ,\gamma _g$
be the basis of
$H_1(G,\mathbb {Z}) \cong H_1(G^{\prime },\mathbb {Z})$
determined by T in G and
$T^{\prime }$
in
$G^{\prime }$
. Finally, choose orientations of the edges and cycles in G; these descend to
$G^{\prime }$
, allowing us to define both
$\Upsilon $
and
$\Upsilon ^{\prime }$
as in §3.1.
Let be the natural inclusion, which misses
$x_a$
, and define a projection map
via
$x_a \mapsto 0$
and
$x_e \mapsto x_e$
for
$e \neq a$
. Notice that
$\rho \circ \iota = \mathrm {id}$
. Then the induced maps
and
making the natural identification
$N \cong N^{\prime }$
also satisfy
$\rho \circ \iota = \mathrm {id}$
. We write
for the homomorphism sending
$a \mapsto 0$
and
$e \mapsto e$
for
$e \neq a$
.
It is straightforward to check that
$\rho ([e,e^{\prime }]) = [c(e),c(e^{\prime })]^{\prime }$
for all
$e,e^{\prime } \in E(G)$
, and hence,
$\rho \circ \pi = \pi ^{\prime } \circ c$
. In particular,

by (19),
$\rho $
induces a group isomorphism
$\mathscr {P} \cong \mathscr {P}^{\prime }$
. Moreover, it is clear from the edge pair characterization in Proposition 3.6 that, for edges
$e,e^{\prime } \in E(G^{\prime })$
, the indicator functions
$f_t$
,
$g_t$
and
$h_t$
are the same in
$G^{\prime }$
as they are in G. In other words,

Then
$\rho (\Theta (\Upsilon )) = \Theta ^{\prime }(\Upsilon ^{\prime })$
, proving the first part of the statement. If a is a separating edge, then a is not part of any cycle. This implies that
$x_a$
does not appear in
$\pi (\gamma _t)$
for any t, so
$\iota $
induces
$\mathscr {P}^{\prime } \cong \mathscr {P}$
. Likewise,
$a_{i,j,k}(e,a) = 0$
for all
$e \in E(G)$
, so
$\iota (\Theta ^{\prime }(\Upsilon ^{\prime })) = \Theta (\Upsilon )$
. This proves the remaining part of the statement.
4.3 Deletion
Fix a graph G and let
$\Upsilon $
be the explicit
$(1,2)$
-chain defined in §3.1. If
$\alpha (G) = 0$
, then by definition, we can write
$\Theta (\Upsilon )$
as some
$\mathbb {Z}$
-linear combination of the generators in (19). In fact, not all of the generators are necessary.
Lemma 4.4. If
$\alpha (G) = 0$
, then
$\Theta (\Upsilon )$
is a
$\mathbb {Z}$
-linear combination of the generators

Proof. Fix indices i, j, and k all distinct with
$j < k$
. By (20), the generator
$2\epsilon _i \wedge \pi (\gamma _j) \wedge \pi (\gamma _k)$
contributes a term

which contains
$2x_{e_j}x_{e_k}\, \epsilon _i \wedge \epsilon _j \wedge \epsilon _k$
. Since
$e_t$
appears only in
$\gamma _t$
, it is straightforward to show that this is in fact the only generator that contains a nonzero multiple of
$2x_{e_j}x_{e_k} \epsilon _i \wedge \epsilon _j \wedge \epsilon _k$
. Moreover, by Lemma 3.5, such a term also does not appear in
$\Theta (\Upsilon )$
; as there is no way to cancel this term out with a different generator, we conclude that
$\epsilon _i \wedge \pi (\gamma _j) \wedge \pi (\gamma _k)$
cannot contribute to a combination equaling
$\Theta (\Upsilon )$
. In other words, we must have either
$i = j$
or
$i = k$
.
Proposition 4.5. Let G be a graph with an edge a. If either
-
(a) a is a loop edge or
-
(b) a is parallel to an edge
$a^{\prime }$ ,
then
$\alpha (G) = 0$
implies that
$\alpha (G \setminus a) = 0$
.
Proof. Let . In either case (a) or (b), a is part of some cycle. Without loss of generality, we may choose the spanning tree T and the labeling on the edges of
$G \setminus T$
so that
$e_g = a$
. We also fix a basepoint v and orientations on the edges and cycles. These choices descend to
$G^{\prime }$
. Identifying edges of
$G^{\prime }$
with the corresponding edges of G, there is a natural inclusion
that misses
$x_{e_g}$
. Likewise, identifying the cycles of
$G^{\prime }$
with the corresponding cycles in G, we let
denote the induced inclusion on homology and
the projection that kills
$\gamma _g$
. Define

In coordinates, we may identify
$N^{\prime }_{R^{\prime }} \cong R^{\prime }\langle {\epsilon _1,\ldots ,\epsilon _{g-1}}\rangle $
and
$N_R \cong R\langle {\epsilon _1,\ldots ,\epsilon _g}\rangle $
; then
$\nu $
is the inclusion of
$R^{\prime }$
-modules sending
$\epsilon _i \mapsto \epsilon _i$
, while
$\xi $
is the projection of R-modules sending
$\epsilon _i \mapsto \epsilon _i$
for
$i < g$
and
$\epsilon _g \mapsto 0$
.
Because
$x_{e_g}$
appears in
$[e,\gamma _i]$
only for
$e = e_g$
and
$i = g$
, it is straightforward to check that

for all edges
$e \in E(G^{\prime }) \subset E(G)$
. We claim that
$\nu _{*}\Upsilon ^{\prime } = \xi _{*}\Upsilon $
. Indeed, observe first that
$\Upsilon ^{\prime }$
does not have an
$\epsilon _g$
-component, and that the
$\epsilon _g$
-component of
$\Upsilon $
is killed by
$\xi _{*}$
. Fixing
$i \neq g$
, we recall that the paths
$\delta _i$
defined in §3.1 remain in
$T = T^{\prime }$
, so
$\delta _i^{\prime } = \delta _i$
. The definition of
$\Upsilon ^{\prime }$
given by (17) depends only on the points
$u_{i,j}^{\prime } = \pi ^{\prime }\left (\delta _i + \sum _{k=0}^{j-1} f_i(e_{i,k})e_{i,k} \right ) \in N^{\prime }_{R^{\prime }}$
. Then
$\nu (u_{i,j}^{\prime }) = \xi (u_{i,j})$
for all j, so the definition of the pushforward in (10) implies that
$\nu _{*}(\epsilon _i \otimes \Upsilon _i^{\prime }) = \xi _{*}(\epsilon _i \otimes \Upsilon _i)$
. The claim follows. Consequently,

A similar computation on generators of
$H_{1,2}(\operatorname {\mathrm {Jac}}(G))$
shows that
$\nu (\mathscr {P}^{\prime }) \subset \xi (\mathscr {P})$
, but we may not have equality in general. Explicitly,

leaving open the possibility that
$\xi $
fails to map generators of the form
$2\epsilon _i \wedge \pi (\gamma _j) \wedge \pi (\gamma _g)$
into
$\nu (\mathscr {P}^{\prime })$
. Here, we have restricted our attention to the case where
$i < g$
, since
$\xi (\epsilon _g) = 0$
. Our assumption that
$\alpha (G) = 0$
means that we may write
$\Theta (\Upsilon )$
explicitly as a
$\mathbb {Z}$
-linear combination of the generators in Lemma 4.4. Therefore, if we can show that
$2\epsilon _i \wedge \xi (\pi (\gamma _i)) \wedge \xi (\pi (\gamma _g))$
lies in
$\nu (\mathscr {P}^{\prime })$
for any such generator appearing in the linear combination, then we will have shown that
$\nu (\Theta ^{\prime }(\Upsilon ^{\prime })) \in \nu (\mathscr {P}^{\prime })$
; by injectivity of
$\nu $
, this would imply in turn that
$\Theta ^{\prime }(\Upsilon ^{\prime }) \in \mathscr {P}^{\prime }$
, as desired.
For case (a), this is straightforward because
$\pi (\gamma _g) = \pi (e_g) = x_{e_g}\epsilon _g$
, so
$\xi (\pi (\gamma _g)) = 0$
. In case (b), recall that
$e_g$
is parallel to an edge
$a^{\prime }$
. Without loss of generality, assume that
$e_g$
and
$a^{\prime }$
have the same orientation. If
$a^{\prime }$
is a separating edge in
$G^{\prime }$
, then
$\gamma _g = -a^{\prime } + e_g$
and
$a^{\prime }$
is not part of any other cycle
$\gamma _i$
. In particular,
$\pi (\gamma _g) = (x_{a^{\prime }} + x_{e_g})\epsilon _g$
, so we again have that
$\xi (\pi (\gamma _g)) = 0$
. Otherwise,
$a^{\prime }$
is not separating in
$G^{\prime }$
, so we may choose T so that
$e_{g-1} = a^{\prime }$
. We write explicitly

for
$a_i \in \mathbb {Z}$
, where we have omitted from the linear combination the generators of
$H_{1,2}(\operatorname {\mathrm {Jac}}(G))$
that we already know map to
$\nu (\mathscr {P}^{\prime })$
under
$\xi $
– that is, those that do not contain
$\pi (\gamma _g)$
as a factor or that have
$\epsilon _g$
as the first factor.
The fact that
$\gamma _g = \gamma _{g-1} - e_{g-1} + e_g$
allows us to rewrite

Since
$x_{e_g}$
appears as a coefficient in
$\pi (\gamma _j)$
only for
$j = g$
, the only place where it appears in (23) is where we have explicitly written it before the second summation. In particular,
$x_{e_g}$
does not appear in any of the omitted generators of the form
$2\epsilon _g \wedge \pi (\gamma _g) \wedge \pi (\gamma _j)$
because the leading factor of
$\epsilon _g$
kills the
$\epsilon _g$
-term of
$\pi (\gamma _g)$
, nor does it appear in
$\Theta (\Upsilon )$
by Lemma 3.5. We conclude that

and hence,
$[\gamma _i,\gamma _j](a_i - a_j) = 0$
for all
$i,j \in [g-1]$
. Applying
$\xi $
to (23) kills terms that contain
$\epsilon _g$
as a factor, so we obtain

completing the proof.
4.4 Graphs of hyperelliptic type
Given a graph G, we write
$G^2$
to mean the 2-edge-connectivization of G, obtained by contracting each of the separating edges of G.
Corollary 4.6.
$\alpha (G) = 0$
if and only if
$\alpha (G^2) = 0$
.
Proof. This follows immediately from repeated application of Proposition 4.3.
Lemma 4.7. Let
$G_1$
and
$G_2$
be graphs, with
the wedge sum. If
$G_1$
and
$G_2$
are both period-trivial, then so is G. In particular, a graph has trivial Ceresa period if each of its maximal 2-connected components is.
Proof. The second statement follows trivially from the first. To prove the first statement, let
$v_1 \in V(G_1)$
and
$v_2 \in V(G_2)$
be the two vertices identified in G. For
$t \in \{1,2\}$
, fix
$v_t$
as the basepoint of
$G_t$
with spanning tree
$T_t$
. Notice that
is a spanning tree of G; let
$v_1 = v_2$
be the basepoint for G. Let
$\gamma _{t,1},\ldots ,\gamma _{t,g_t}$
be the simple cycles determined by
$T_t$
with arbitrary orientation; then
$\gamma _{1,1},\ldots ,\gamma _{1,g_1},\gamma _{2,1},\ldots ,\gamma _{2,g_2}$
are the simple cycles of G. Consequently, we identify
$H_1(G,\mathbb {Z}) \cong H_1(G_1,\mathbb {Z}) \oplus H_1(G_2,\mathbb {Z})$
. The paths
$\delta _{t,i}$
in
$T_t$
descend to the corresponding paths in T, allowing us to define
$(1,2)$
-chains
$\Upsilon _1$
,
$\Upsilon _2$
, and
$\Upsilon $
using the explicit construction given in §3.1.
We may regard R as the coproduct of
$\mathbb {Z}$
-algebras
$R \cong R_1 \oplus _{\mathbb {Z}} R_2$
. Then
$N_R \cong N_{1,R} \oplus N_{2,R}$
as R-modules. It is straightforward to see using Proposition 3.6 that
$\Theta (\Upsilon )$
may be obtained as
$\Theta (\Upsilon _1) + \Theta (\Upsilon _2)$
. Indeed, no pair of edges
$(e_1,e_2)$
with
$e_t \in E(G_t)$
share a common cycle, so they do not contribute to
$\Theta (\Upsilon )$
. Meanwhile, any pairs
$(e,e^{\prime })$
with both edges coming from the same subgraph
$G_1$
have the same
$f_{1,i}$
,
$g_{1,i}$
and
$h_{1,i}$
values in G as they do in
$G_1$
, with
$f_{2,i}$
,
$g_{2,i}$
and
$h_{2,i}$
values all zero (and vice versa for edge pairs coming from
$G_2$
). Likewise, one can see from the generators given by (19) that
$\mathscr {P}_1 \oplus \mathscr {P}_2 \subset \mathscr {P}$
; it follows that if
$\Theta (\Upsilon _t) \in \mathscr {P}_t$
for both t, then
$\Theta (\Upsilon ) \in \mathscr {P}$
, as desired.
Let G and
$G^{\prime }$
be graphs. We say that
$G^{\prime }$
is a permissible minor of G if we may obtain
$G^{\prime }$
from G by deleting only loops or parallel edges and contracting only non-loop edges. Then Propositions 4.3 and 4.5 immediately imply that
$\alpha (G) = 0$
only if
$\alpha (G^{\prime }) = 0$
.
A tropical curve C is hyperelliptic if it admits an involution
$\iota $
for which the quotient
$C/\iota $
is a tree. More generally, it is of hyperelliptic type if its Jacobian is isomorphic to that of a hyperelliptic tropical curve. We say that a graph G is of hyperelliptic type if some choice of edge lengths makes it into a hyperelliptic-type tropical curve. For more details on these notions, we refer to [Reference Baker and Norine10, Reference Amini, Baker, Brugallé and Rabinoff3, Reference Amini, Baker, Brugallé and Rabinoff4, Reference Chan14]. By [Reference Corey15, Proposition 3.3], the property of being of hyperelliptic type does not in fact depend on the choice of edge lengths. Following [Reference Corey15], we say that G is strongly of hyperelliptic type if some choice of edge lengths yields a hyperelliptic tropical curve.

Figure 5 The graph
$L(T)$
for the tree T marked with bold edges.
Let T be a tree of maximal valence 3 and fix a disjoint copy
$T^{\prime }$
of T. Let
be the involution that identifies each vertex of T with the corresponding vertex of
$T^{\prime }$
. Following [Reference Chan14, Definition 4.7], we define the ladder over T to be the graph
$L(T)$
obtained by adding
$3 - \operatorname {\mathrm {val}}(v)$
parallel edges between v and
$\iota (v)$
for each
$v \in V(T)$
. We call the edges added in this way vertical edges.
Lemma 4.8. Let G be a
$2$
-edge-connected graph that is strongly of hyperelliptic type. Then
$\alpha (G) = 0$
.
Proof. We claim that G is a permissible minor of
$L(T)$
for some tree T of maximal valence
$3$
. Indeed, this follows almost immediately from [Reference Chan14, Theorem 4.9], which states that ladders are precisely the maximal cells of the moduli space of
$2$
-edge-connected hyperelliptic tropical curves. However, since we do not allow weighted vertices, we must delete loops rather than contracting them. Then without loss of generality, we may assume that
$G = L(T)$
.
We show that
$\Theta (\Upsilon )$
is identically zero for
$L(T)$
. Fix a 1-valent vertex
$v_0$
of T and an edge
$e_0$
of
$L(T)$
from
$v_0$
to
$\iota (v_0)$
. Let S be the spanning tree of
$L(T)$
with edges
$E(T) \cup E(\iota (T)) \cup \{e_0\}$
. Each remaining vertical edge determines a unique cycle in
$L(T)$
; label these edges arbitrarily and orient them from T to
$\iota (T)$
. Orient each
$e \in E(T)$
away from
$v_0$
; there is a partial order on
$E(T)$
given by declaring that
$e < e^{\prime }$
whenever the unique path in T from v to the tail of
$e^{\prime }$
contains e. Rephrasing Proposition 3.6, nonzero terms of
$\Theta (\Upsilon )$
correspond to pairs of distinct edges e and
$e^{\prime }$
satisfying:
-
(*) e and
$e^{\prime }$ share a cycle and each is part of another cycle that the other is not in.
We claim that (
$*$
) is not satisfied for any pair of edges in
$L(T)$
. Indeed, neither edge can be any of the vertical edges:
$e_0$
is part of every cycle, while each other vertical edge is part of only one cycle. If both e and
$e^{\prime }$
lie in T, then either
$e < e^{\prime }$
or the two edges are incomparable. In the first case, every cycle containing
$e^{\prime }$
also contains e. In the second case, there are no common cycles between e and
$e^{\prime }$
. By symmetry, (
$*$
) also fails if both e and
$e^{\prime }$
lie in
$T^{\prime }$
. If
$e^{\prime } = \iota (e)$
, then
$e^{\prime }$
and e form a separating pair; in particular, they are contained in precisely the same cycles. If
$e \in T$
and
$e^{\prime } \in T^{\prime }$
with
$e^{\prime } \neq \iota (e)$
, then the fact that (
$*$
) fails for e and
$\iota (e^{\prime })$
implies that it also fails for e and
$e^{\prime }$
. This proves the claim, so
$\Theta (\Upsilon ) = 0$
, as desired.
Proof of Theorem A.
Suppose first that G is not of hyperelliptic type. By [Reference Corey15, Theorem 1.1], G contains
$G^{\prime } \in \{K_4,L_3\}$
as a minor. In fact, [Reference Corey and Li17, Lemma 5.10] implies that, in this special case,
$G^{\prime }$
must be a permissible minor of G. That
$\alpha (G) \neq 0$
follows from our computations in Examples 4.1 and 4.2 showing that
$\alpha (G^{\prime }) \neq 0$
in either case.
Conversely, suppose that G is of hyperelliptic type. By [Reference Corey15, Theorem 1.1], G does not contain
$K_4$
or
$L_3$
as a minor. Then neither do the maximal 2-connected components of the 2-edge-connectivization
$G^2$
, so any such component is still of hyperelliptic type. In particular, Corollary 4.6 and Lemma 4.7 imply that we may assume that G itself is 2-connected. By definition, there exists a 2-connected tropical curve C of hyperelliptic type with underlying graph G. Then by [Reference Corey15, Theorem 4.5], there exists a hyperelliptic tropical curve
$C^{\prime }$
of which C is a permissible minor. Let
$G^{\prime }$
denote the underlying graph of
$C^{\prime }$
. Contracting any separating edges, the resulting tropical curve
${C^{\prime }}^2$
is hyperelliptic by [Reference Chan14, Corollary 3.11]. Then
${G^{\prime }}^2$
is 2-edge-connected and strongly of hyperelliptic type, so Lemma 4.8 implies that
$\alpha ({G^{\prime }}^2) = 0$
. By Corollary 4.6,
$\alpha (G^{\prime }) = 0$
; since G is a permissible minor of
$G^{\prime }$
, we have that
$\alpha (G) = 0$
.
Acknowledgements
The author thanks Farbod Shokrieh for his invaluable guidance throughout this project; Alexander Waugh for homology advice; Samouil Molcho, Thibault Poiret, Felix Röhrle and Jonathan Wise for various insightful conversations at the BIRS workshop on ‘Curves: Algebraic, Tropical, and Logarithmic’; and the anonymous referee for their helpful comments and suggestions.
Competing interest
The author has no competing interests to declare.
Funding statement
The author was partially supported by NSF CAREER grant DMS–2044564.