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

Tropical curves in abelian surfaces II:
enumeration of curves in linear systems

Thomas Blomme
Abstract

In this paper, second installment in a series of three, we give a correspondence theorem to relate the count of genus gg curves in a fixed linear system in an abelian surface to a tropical count. To do this, we relate the linear system defined by a complex curve to certain integrals of 11-forms over cycles in the curve. We then give an expression for the tropical multiplicity provided by the correspondence theorem, and prove the invariance for the associated refined multiplicity, thus introducing refined invariants of Block-Göttsche type in abelian surfaces.

:
14N10, 14T90, 11G10, 14K05, 14H99
keywords:
Enumerative geometry, tropical refined invariants, abelian surfaces, floor diagrams
Data Statement: I do not have any data to point.

1 Introduction

This paper is the second in a series of three papers which study enumerative invariants of abelian surfaces from the tropical point of view. While the first paper focuses on the enumeration of genus gg curves in a fixed class passing through gg points, this paper is dedicated to the enumeration of genus gg curves in a fixed linear system that are also subject to point conditions.

1.1 Abelian surfaces and tropical tori

1.1.1 Complex abelian surfaces.

Abelian surfaces are complex tori, i.e. the quotient A=2/Λ\mathbb{C}A=\mathbb{C}^{2}/\Lambda of the complex vector space 2\mathbb{C}^{2} by some rank 44 lattice Λ\Lambda, that is subject to some condition, namely the existence of a positive line bundle called polarization on A\mathbb{C}A. Not every complex torus can be endowed with the choice of a polarization: this imposes conditions on a matrix spanning the lattice, known as Riemann bilinear relations. These are proved in [12]. Abelian surfaces have a natural structure of additive group inherited from the vector space structure of 2\mathbb{C}^{2}. Through the use of the exponential map, it is also possible to give a multiplicative description of an abelian surface as a quotient of the algebraic complex torus ()2(\mathbb{C}^{*})^{2} by some rank 22 lattice. This description is more adapted to define tropical counterparts to abelian surfaces, as it provides a logarithmic map on A=()2/Λ\mathbb{C}A=(\mathbb{C}^{*})^{2}/\Lambda.

1.1.2 Tropical abelian surfaces.

Tropicalizing the above definition, a tropical torus is obtained as a quotient of the vector space 2\mathbb{R}^{2} by some rank 22 lattice. In the rest of the paper, the vector space is N=NN_{\mathbb{R}}=N\otimes\mathbb{R}, where NN is some rank 22 lattice, and the lattice by which we quotient is still denoted by Λ\Lambda, with an inclusion S:ΛNS:\Lambda\to N_{\mathbb{R}}. As in the complex case, a tropical torus is a tropical abelian surface if it can be endowed with the choice of a polarization, i.e. a positive line bundle. We have a natural analog of the Riemann bilinear relation expressing the condition on the lattice Λ\Lambda so that 𝕋A=N/Λ\mathbb{T}A=N_{\mathbb{R}}/\Lambda.

1.1.3 Tropicalizing a family of complex tori.

The two previous definitions relate as it is possible to consider specific families of abelian surfaces N/ΛtN_{\mathbb{C}^{*}}/\Lambda_{t}, called Mumford families, that “tropicalize” to a tropical abelian surface 𝕋A=N/Λ\mathbb{T}A=N_{\mathbb{R}}/\Lambda. It is then possible to apply correspondence techniques to resolve enumerative problems inside complex abelian surfaces only by studying tropical enumerative problems.

1.1.4 Curves in abelian surfaces.

We consider a complex abelian surface A=N/Λ\mathbb{C}A=N_{\mathbb{C}^{*}}/\Lambda endowed with a polarization \mathcal{L}. It is possible to assume up to a change of basis that the pull-back of \mathcal{L} to NN_{\mathbb{C}^{*}}. Sections of \mathcal{L} can thus be viewed as certain specific holomorphic functions on NN_{\mathbb{C}^{*}} satisfying a quasi-periodic relation. These are called θ\theta-functions, and their definition is recalled in section 2.2. The zero locus of a θ\theta-function is a curve in the linear system |||\mathcal{L}|.

1.2 Enumerative geometry and Gromov-Witten invariants

1.2.1 Curves and enumerative problems.

Zero loci of sections of \mathcal{L} provide curves in the abelian surface, and it natural to try to count curves of a fixed genus that are subject to some constraints. Furthermore, using the group structure of the abelian surface, it is possible to translate curves. However, the translate of a curve does not usually belong to the same linear system anymore. It is possible to show (see [8]) that the dimension of the deformation space of a genus gg curve realizing a fixed homology class CH2(A,)C\in H_{2}(\mathbb{C}A,\mathbb{Z}) is gg. This leads to the following two enumerative problems:

  • \circ

    How many genus gg curves in the class CC pass through gg points ?

  • \circ

    How many genus gg curves in a fixed linear system pass through g2g-2 points ?

The first paper in the series focuses on the first enumerative problem, we now deal with the second.

As it happens, the answer to both problems does not depend on the choice of the points, polarization up to translation nor the abelian surface. They are denoted by 𝒩g,C\mathcal{N}_{g,C} and 𝒩g,CFLS\mathcal{N}_{g,C}^{FLS} respectively. These invariants were already studied in [8]. They coincide with Gromov-Witten invariants of abelian surfaces, as defined in [8]. Their definition differs a little from usual Gromov-Witten invariants because the choice of complex structure on an abelian surface is not generic, and a generic choice would lead to the surface having no curve at all. See [8] for more details.

More generally, Gromov-Witten invariants are obtained by integrating cohomology classes over some virtual fundamental class in the moduli space of curves inside a specific variety, here an abelian surface. The moduli space of curves with marked points is endowed with an evaluation map to the variety. Integrating pull-back of cohomology classes Poincaré dual to geometric constraints by the evaluation map amounts to count curves satisfying these geometric constraints. However, there are many other cohomology classes that it is possible to integrate, such as λ\lambda-classes, ψ\psi-classes, …The invariants from [8] were thus generalized in [9] to a wider families of Gromov-Witten invariants. The cohomology classes of an algebraic complex variety might be Poincaré dual to algebraic cycles, and thus the computation of the corresponding Gromov-Witten invariants amounts to the solving of some algebraic problem. This is the case for some toric varieties. However, this is not always the case. For abelian surfaces, there are new classes that are not dual to algebraic cycles, for instance classes in H0,1(A)H^{0,1}(\mathbb{C}A) and H1,0(A)H^{1,0}(\mathbb{C}A), providing other Gromov-Witten invariants.

1.2.2 Parametric or implicit curves.

The point of view of Gromov-Witten theory is to consider parametrized complex curves in a fixed homology class CH2(X,)C\in H_{2}(\mathbb{C}X,\mathbb{Z}), where X\mathbb{C}X is a complex variety. Counting curves in a fixed linear system uses more the implicit point of view on curves: curves are sections of a fixed line bundle, i.e. they are given with an equation. Both points of view are equivalent when the Picard group of X\mathbb{C}X is discrete in the sense that the data of the homology class CC uniquely determines the line bundle \mathcal{L} and the linear system |||\mathcal{L}|. However, when X=A\mathbb{C}X=\mathbb{C}A is for instance an abelian surface, the Picard group is not discrete as it contains a part of dimension h1,0=2h^{1,0}=2. Imposing the linear system acts as a codimension 22 condition on the space of curves. This corresponds to the fact that translates of a curve are not section of the same line bundle anymore. In [8], J. Bryan and N. Leung proved that it is possible to transform condition on the linear system into conditions on the curve. For instance, being part of a fixed linear system becomes meeting four 11-dimensional cycles whose homology classes span H1(A,)H_{1}(\mathbb{C}A,\mathbb{Z}).

In this paper, we adopt a parametric point of view, so that we also need to transform the linear system condition into a more manageable data. We adopt a point of view slightly different from [8] by instead recovering the line bundle 𝒪(𝒞)\mathcal{O}(\mathscr{C}) associated to a complex curve 𝒞\mathscr{C} in terms of integrals of certain 11-forms on the curve. More precisely, the statement is as follows. We use the following facts. We refer to section 2.4 for more details.

  • -

    The Picard group of the abelian surface A=N/Λ\mathbb{C}A=N_{\mathbb{C}^{*}}/\Lambda is isomorphic to Λ/M\Lambda_{\mathbb{C}^{*}}^{*}/M, where MM is the dual lattice of NN.

  • -

    For a curve 𝒞N\mathscr{C}\subset N_{\mathbb{C}^{*}} and a circle γ\gamma inside 𝒞\mathscr{C} realizing a homology class nNn\in N, its moment is defined as exp(μγ=12iπγlogzdww)\exp\left(\mu_{\gamma}=\frac{1}{2i\pi}\int_{\gamma}\log z\frac{\mathrm{d}w}{w}\right), where zz and ww are the coordinates on NN_{\mathbb{C}^{*}} such that logz\log z is well-defined on γ\gamma. (it is the monomial in M=NM=N^{*} Poincaré dual to nn)

  • -

    An element λΛ\lambda\in\Lambda gives a 33-dimensional variety XλX_{\lambda} with boundary inside NN_{\mathbb{C}^{*}}. It is obtained as the preimage of a path in NN_{\mathbb{R}} by the logarithmic map NNN_{\mathbb{C}^{*}}\to N_{\mathbb{R}}.

The statement is as follows.

Theorem.

2.12 Given two curves 𝒞0\mathscr{C}_{0} and 𝒞1\mathscr{C}_{1} in A\mathbb{C}A lifted to periodic curves inside NN_{\mathbb{C}^{*}}. The line bundle 𝒪(𝒞1𝒞0)\mathcal{O}(\mathscr{C}_{1}-\mathscr{C}_{0}) is represented by the element of Λ\Lambda_{\mathbb{C}^{*}} that maps λΛ\lambda\in\Lambda to the product of the moments of the circles resulting from the intersection Xλ(𝒞0𝒞1)X_{\lambda}\cap(\mathscr{C}_{0}\cup\mathscr{C}_{1}).

This statement is a generalization of the 11-dimensional statement expressing the line bundle associated to a degree 0 divisor on an elliptic curve. See section 2.4 for more details.

1.2.3 Previous computations.

All the above invariants introduced above were computed in both [8] and [9] in case the class of curves CC is called primitive. This allows to prove some regularity statement. For instance, they show that certain generating series of these invariants are quasi-modular forms. To my knowledge, the computation for non-primitive classes remains open. This paper adresses the computation of 𝒩g,CFLS\mathcal{N}_{g,C}^{FLS} for non-primitive classes using the tropical method. Hopefully, generalizations of the statements from [9] should be possible with this approach.

We also mention the work of L. Halle and S. Rose [13] that also count tropical curves in abelian varieties using a completely different method: they study maps between tropical tori and maps from a curve to its Jacobian. Their method leads to the computation of some of the above invariants but generalizes in higher dimension.

1.3 Tropical geometry and correspondence theorems

1.3.1 Correspondence theorems in toric varieties.

Once defined, it remains to compute the invariants, in order to for instance study their regularity or their generating series. As the invariants do not depend on the choice of the constraints nor the choice of the abelian surface, it is now a classical method to try to compute them close to the tropical limit. This approach was first implemented by G. Mikhalkin in [16] for computing enumerative invariants of toric surfaces. The result is known as correspondence theorems. Since, other versions of correspondence theorems have been proved in various settings with different techniques, see for instance [20], [22], [23], [24] or [15].

Close to the tropical limit, a complex curve break into several pieces whose structure is encoded in a tropical curve. Tropical curves are graphs whose edges have integer slope that satisfy the balancing condition at each of their vertices. Tropical curves were first defined as graphs in 2\mathbb{R}^{2}, or n\mathbb{R}^{n}. See for instance [7]. It is possible to generalize their definition to any manifold whose tangent bundle contains a natural lattice. This is the case of tropical tori. Adhoc correspondence theorems can then relate the study of tropical curves inside such manifolds, called affine integer manifolds, to enumerative or Gromov-Witten invariants of suitable complex manifolds.

1.3.2 Correspondence in abelian surfaces.

Getting out of the toric situation, it is still possible to prove correspondence theorems by generalizing methods from the toric case. For instance, in [3], the author adapts methods from [20] to compute Gromov-Witten invariants of complex manifolds that are line bundles over an elliptic curve. However, this requires to consider families of complex varieties while correspondence for toric varieties could be seen as happening in the same variety. Concretely, in the situation of the paper, it means that we have to consider families of complex abelian surfaces At\mathbb{C}A_{t} rather than a single abelian surface.

Concerning abelian surfaces, a correspondence theorem was first proved by T. Nishinou [19]. More precisely, the main result of [19] is twofold: it consists in a realization theorem that expresses the possible deformations of a given tropical curve, and the second that uses the description of the deformations to give an expression of the deformations that satisfy gg points constraints. This way, a correspondence theorem can be seen as a recipe to get a multiplicity mΓm_{\Gamma}^{\mathbb{C}} out of a tropical curve, so that the count of tropical curves solution to a suitable enumerative problem with this multiplicity gives the desired invariant. In the first paper of the series, we proved a product expression for the complex multiplicity provided by Nishinou’s correspondence theorem.

1.3.3 Correspondence for linear systems.

In the present paper, we adapt the proof of the correspondence theorem from [19] to work for the case of curves in a fixed linear system as well. It is possible to adapt the proof using the parametric point of view thanks to the expression of the linear system constraints provided by Theorem 2.12.

Let 𝒫t\mathcal{P}_{t} be a g2g-2 point configuration in a family of abelian surfaces At\mathbb{C}A_{t} that tropicalizes to 𝒫\mathcal{P} inside 𝕋A\mathbb{T}A. See section 3.2 for more details. We have the following correspondence theorem.

Theorem.

3.3 Let h:Γ𝕋Ah:\Gamma\to\mathbb{T}A be a parametrized tropical curve passing through 𝒫\mathcal{P} and in a fixed linear system. The number of genus gg complex curves in the fixed linear system passing through 𝒫t\mathcal{P}_{t} and that tropicalize to h:Γ𝕋Ah:\Gamma\to\mathbb{T}A is

mΓ=|kerΨ|eE(Γ)we.m_{\Gamma}^{\mathbb{C}}=|\ker\Psi_{\mathbb{C}^{*}}|\prod_{e\in E(\Gamma^{\prime})}w_{e}.

In particular, we have that Ng,CFLS(𝕋A,𝒫)N_{g,C}^{FLS}(\mathbb{T}A,\mathcal{P}) does not depend on 𝒫\mathcal{P} and 𝕋A\mathbb{T}A as long as these choices are generic, and Ng,CFLS=𝒩g,CFLSN_{g,C}^{FLS}=\mathcal{N}_{g,C}^{FLS}.

The map Ψ\Psi involved in the theorem is defined in section 3.2. It plays a role similar to the evaluation map Θ\Theta from [19]. Notice that the theorem provides a new complex multiplicity with which to count tropical curves, and it may differ from the one introduced in [19] and used in the first paper of this series. Its expression using Ψ\Psi is not that important since we give below a concrete expression for mΓm_{\Gamma}^{\mathbb{C}}. The number Ng,CFLSN_{g,C}^{FLS} is the number of genus gg tropical curves in a fixed linear system passing through a generic configuration of g2g-2 points.

In fact, we also have a product expression provided by the following theorem. Let h:Γ𝕋Ah:\Gamma\to\mathbb{T}A be a tropical curve passing through a generic point configuration 𝒫\mathcal{P} of g2g-2 points and that belongs to a fixed linear system. The complement of 𝒫\mathcal{P} inside Γ\Gamma is connected and has genus 22. It retracts onto a genus 22 subgraph ΣΓ\Sigma\subset\Gamma. Let ΛΓΣ\Lambda_{\Gamma}^{\Sigma} be the index of H1(Σ,)H_{1}(\Sigma,\mathbb{Z}) inside H1(𝕋A,)ΛH_{1}(\mathbb{T}A,\mathbb{Z})\simeq\Lambda. Let also δΓ\delta_{\Gamma} be the gcd of the weights of the edges of the curve.

Theorem.

3.4. One has mΓ=δΓΛΓΣmΓm_{\Gamma}^{\mathbb{C}}=\delta_{\Gamma}\Lambda_{\Gamma}^{\Sigma}m_{\Gamma}, where mΓ=VmVm_{\Gamma}=\prod_{V}m_{V} is the usual multiplicity, and mVm_{V} is the usual vertex multiplicity.

The new complex multiplicity possesses a new term ΛΓΣ\Lambda_{\Gamma}^{\Sigma}. Tropically, its presence can be explained by the appearance of new kind of walls. These new walls are as follows. Usually, when moving the constraints, one of the following events, called wall, can happen:

  • \circ

    An edge is contracted leading to the appearance of a quadrivalent vertex. In this case, two solutions may become a unique solution on the other side of the wall.

  • \circ

    A cycle is contracted to a segment, leading to a pair of quadrivalent vertices linked by a pair of parallel edges. In this case, one curve is replaced by another.

  • \circ

    Last, one of the marked points meets a vertex of the curve.

For the latter kind of wall, in the usual case of toric surfaces, the fact that the complement of the marked point in the curve is a forest allows one to prove that the marked point can only go from on edge adjacent to the vertex to another, so that there are only two out of the three adjacent combinatorial types that can provide a solution. Here, if a marked point merges with a vertex of Σ\Sigma, the marked point may move onto any of the adjacent edges, and we can have two solutions becoming one on the other side of the wall. The local invariance is ensured by the new term ΛΓΣ\Lambda_{\Gamma}^{\Sigma}.

1.4 Refined invariants for curves in linear systems

The usual complex multiplicity provided by the correspondence theorem from [16] expresses as a product over the vertices of the tropical curve. In [2], F. Block and L. Göttsche proposed to refine this multiplicity into a Laurent polynomial one by replacing the vertex multiplicities by their quantum analog. In the case of toric surfaces, I. Itenberg and G. Mikhalkin proved in [14] that the count of tropical curves with fixed genus and degree passing through the right number of points with refined multiplicity is invariant. These results were generalized in various settings, see for instance [21], [1], [10], [3], [4].

In our case, despite the appearance of the new term ΛΓΣ\Lambda_{\Gamma}^{\Sigma}, the complex multiplicity provided by Theorem 3.3 and Theorem 3.4 still expresses as a product over the vertices. Thus, we can define the refined multiplicity mΓq=VqmV/2qmV/2q1/2q1/2m_{\Gamma}^{q}=\prod_{V}\frac{q^{m_{V}/2}-q^{-m_{V}/2}}{q^{1/2}-q^{-1/2}}. We then introduce the following enumerative counts:

Ng,C,kFLS(𝕋A,𝒫)\displaystyle N_{g,C,k}^{FLS}(\mathbb{T}A,\mathcal{P}) =h(Γ)𝒫δ(Γ)=kΛΓΣmΓ,\displaystyle=\sum_{\begin{subarray}{c}h(\Gamma)\supset\mathcal{P}\\ \delta(\Gamma)=k\end{subarray}}\Lambda_{\Gamma}^{\Sigma}m_{\Gamma}\in\mathbb{N},
BGg,C,kFLS(𝕋A,𝒫)\displaystyle BG_{g,C,k}^{FLS}(\mathbb{T}A,\mathcal{P}) =h(Γ)𝒫δ(Γ)=kΛΓΣmΓq[q±1/2].\displaystyle=\sum_{\begin{subarray}{c}h(\Gamma)\supset\mathcal{P}\\ \delta(\Gamma)=k\end{subarray}}\Lambda_{\Gamma}^{\Sigma}m^{q}_{\Gamma}\in\mathbb{Z}[q^{\pm 1/2}].

counting curves with a fixed gcd, and the following counts for curves without a gcd condition:

Mg,CFLS(𝕋A,𝒫)\displaystyle M_{g,C}^{FLS}(\mathbb{T}A,\mathcal{P}) =h(Γ)𝒫ΛΓΣmΓ=k|δ(C)Ng,C,kFLS(𝕋A,𝒫),\displaystyle=\sum_{h(\Gamma)\supset\mathcal{P}}\Lambda_{\Gamma}^{\Sigma}m_{\Gamma}=\sum_{k|\delta(C)}N_{g,C,k}^{FLS}(\mathbb{T}A,\mathcal{P})\in\mathbb{N},
Ng,CFLS(𝕋A,𝒫)\displaystyle N_{g,C}^{FLS}(\mathbb{T}A,\mathcal{P}) =h(Γ)𝒫δΓΛΓΣmΓ=k|δ(C)kNg,C,kFLS(𝕋A,𝒫),\displaystyle=\sum_{h(\Gamma)\supset\mathcal{P}}\delta_{\Gamma}\Lambda_{\Gamma}^{\Sigma}m_{\Gamma}=\sum_{k|\delta(C)}kN_{g,C,k}^{FLS}(\mathbb{T}A,\mathcal{P})\in\mathbb{N},
Rg,CFLS(𝕋A,𝒫)\displaystyle R_{g,C}^{FLS}(\mathbb{T}A,\mathcal{P}) =h(Γ)𝒫δΓΛΓΣmΓq=k|δ(C)kBGg,C,kFLS(𝕋A,𝒫)[q±1/2],\displaystyle=\sum_{h(\Gamma)\supset\mathcal{P}}\delta_{\Gamma}\Lambda_{\Gamma}^{\Sigma}m^{q}_{\Gamma}=\sum_{k|\delta(C)}kBG_{g,C,k}^{FLS}(\mathbb{T}A,\mathcal{P})\in\mathbb{Z}[q^{\pm 1/2}],
BGg,CFLS(𝕋A,𝒫)\displaystyle BG_{g,C}^{FLS}(\mathbb{T}A,\mathcal{P}) =h(Γ)𝒫ΛΓΣmΓq=k|δ(C)BGg,C,kFLS(𝕋A,𝒫)[q±1/2].\displaystyle=\sum_{h(\Gamma)\supset\mathcal{P}}\Lambda_{\Gamma}^{\Sigma}m^{q}_{\Gamma}=\sum_{k|\delta(C)}BG_{g,C,k}^{FLS}(\mathbb{T}A,\mathcal{P})\in\mathbb{Z}[q^{\pm 1/2}].

We already know that the count Ng,CFLS(𝕋A,𝒫)N_{g,C}^{FLS}(\mathbb{T}A,\mathcal{P}) does not depend on the choice of 𝒫\mathcal{P} as long as it is generic, nor the choice of 𝕋A\mathbb{T}A as long as it is generic. This invariance is provided by the correspondence theorem. The invariance for other counts cannot be deduced from any complex invariants, and has thus to be studied on its own, looking at the walls depicted above.

Theorem.

3.5 The refined count BGg,C,kFLS(𝕋A,𝒫)BG_{g,C,k}^{FLS}(\mathbb{T}A,\mathcal{P}) of genus gg curves in the class CC with fixed gcd passing through 𝒫\mathcal{P} in the fixed linear system (and thus all the others) does not depend on the choice of 𝒫\mathcal{P} and the line bundle as long as it is generic.

We remove 𝒫\mathcal{P} from the notation to denote the associated invariant. Then, we have an invariance statement regarding the choice of the abelian surface 𝕋A\mathbb{T}A.

Theorem.

3.7 The refined invariant BGg,C,kFLS(𝕋A)BG_{g,C,k}^{FLS}(\mathbb{T}A) (and thus all the others) does not depend on the choice of 𝕋A\mathbb{T}A as long as it is chosen generically among the surfaces that contain curves in the class CC.

Interpretations of refined invariants in toric varieties have already been proved: they correspond both to refined signed count of real curves according to the value of some quantum index, see [17], [5], or to generating series of Gromov-Witten invariants with insertions of λ\lambda-classes, as proven by P. Bousseau in [6]. It is also conjectured that they correspond to the refinement of the Euler characteristic of some relative Hilbert scheme by the Hirzebruch genus, see [11]. Although it should be possible to adapt some methods of the above citations, the meaning of the refined invariants in abelian surfaces remains open. Computations from [9] and results from [6] suggest that refined invariants enable the computation of Gromov-Witten invariants with insertions of λ\lambda-classes in this situation as well.

The refinement of the complex multiplicity by the refined multiplicity proposed in [2] adapts in our situation due to the appearance of the product of vertex multiplicities. The term δΓ\delta_{\Gamma} could be replaced by anything else since we have invariance for the count of tropical curves with a fixed gcd. It would be interesting to find a way to refine the new term ΛΓΣ\Lambda_{\Gamma}^{\Sigma}.

1.5 Plan of the paper

The paper is organized as follows. The second section deals with line bundles on abelian surfaces, their sections called θ\theta-functions, and how to relate the parametric and implicit points of view for curves both in the complex and tropical case. The third section studies the enumerative problem considered in this paper. It states the correspondence theorem, and results concerning tropical multiplicities and the invariants. The fourth section is devoted to the proof of the results stated in the third. The last section provides small examples. We refer to the last paper for more examples of computations using the pearl diagram algorithm.

Acknowledgments. Research is supported in part by the SNSF grant 204125.

2 Line bundles and curves in linear systems

In the following section, we consider a tropical torus 𝕋A=N/Λ\mathbb{T}A=N_{\mathbb{R}}/\Lambda, where NN and Λ\Lambda are two lattices of rank 22, and NN_{\mathbb{R}} is the real vector space NN\otimes\mathbb{R}. The matrix of the inclusion ΛN\Lambda\hookrightarrow N_{\mathbb{R}} is denoted by SS.

2.1 Line bundles on abelian surfaces

Line bundles on tropical tori are defined analogously to the classical case. See [12] for the classical setting, and [13] or [18] for a more complete reference in the tropical setting. In the classical case, the group of isomorphism classes of line bundles on a variety YY is the cohomology group PicY=H1(Y,𝒪Y×)\mathrm{Pic}\ Y=H^{1}(Y,\mathcal{O}^{\times}_{Y}). In the tropical world, the sheaf of affine functions with integer slope Aff\mathrm{Aff}_{\mathbb{Z}} plays that role. Locally, an affine function is of the form xUNm,x+lx\in U\subset N_{\mathbb{R}}\mapsto\langle m,x\rangle+l\in\mathbb{R}.

Definition 2.1.

A isomorphism class of line bundles on 𝕋A\mathbb{T}A is an element of the Picard group H1(𝕋A,Aff)H^{1}(\mathbb{T}A,\mathrm{Aff}_{\mathbb{Z}}), where Aff\mathrm{Aff}_{\mathbb{Z}} is the sheaf of affine functions with integer slope, i.e. slope in MM.

In tropical geometry, the following exact sequence plays the role of the exponential sequence:

0AffM0.0\rightarrow\mathbb{R}\rightarrow\mathrm{Aff}_{\mathbb{Z}}\rightarrow M\rightarrow 0.

The second arrow maps a function to its slope mMm\in M, which is a section of the cotangent bundle. We then have the long exact sequence associated to the short exact sequence:

H0(𝕋A,M)H1(𝕋A,)H1(𝕋A,Aff)c1H1(𝕋A,M)H2(𝕋A,).\cdots\to H^{0}(\mathbb{T}A,M)\to H^{1}(\mathbb{T}A,\mathbb{R})\to H^{1}(\mathbb{T}A,\mathrm{Aff}_{\mathbb{Z}})\xrightarrow{c_{1}}H^{1}(\mathbb{T}A,M)\to H^{2}(\mathbb{T}A,\mathbb{R})\to\cdots.

Morever, we have the following isomorphisms:

  • -

    The group H0(𝕋A,M)H^{0}(\mathbb{T}A,M) of global 11-forms with integer slope is just MM,

  • -

    As H1(𝕋A,)ΛH_{1}(\mathbb{T}A,\mathbb{Z})\simeq\Lambda, we have H1(𝕋A,)ΛH^{1}(\mathbb{T}A,\mathbb{R})\simeq\Lambda^{*}_{\mathbb{R}},

  • -

    We similarly have H1(𝕋A,M)ΛM=hom(Λ,M)H^{1}(\mathbb{T}A,M)\simeq\Lambda^{*}\otimes M=\hom(\Lambda,M).

On the left side of the sequence, the cobordism map MΛM\to\Lambda^{*}_{\mathbb{R}} is just the dual map to the inclusion S:ΛNS:\Lambda\to N_{\mathbb{R}}. By assumption, it is an injection. Thus, we have an injection of the dual torus into the Picard group:

𝕋A=H1(𝕋A,)/H0(𝕋A,M)=Λ/MH1(𝕋A,Aff).\mathbb{T}A^{\vee}=H^{1}(\mathbb{T}A,\mathbb{R})/H^{0}(\mathbb{T}A,M)=\Lambda^{*}_{\mathbb{R}}/M\hookrightarrow H^{1}(\mathbb{T}A,\mathrm{Aff}_{\mathbb{Z}}).

Meanwhile, on the right side of the sequence, the map c1:H1(𝕋A,Aff)H1(𝕋A,M)hom(Λ,M)c_{1}:H^{1}(\mathbb{T}A,\mathrm{Aff}_{\mathbb{Z}})\to H^{1}(\mathbb{T}A,M)\simeq\hom(\Lambda,M) is called the Chern class map. Let \mathcal{L} be a line bundle and c=c1()c=c_{1}(\mathcal{L}) denotes its Chern class. By definition, we have c:ΛMc:\Lambda\to M. Then, using the inclusion S:ΛNS:\Lambda\hookrightarrow N_{\mathbb{R}}, we set the following bilinear form:

B(λ1,λ2)=c(λ1)(Sλ2),B(\lambda_{1},\lambda_{2})=c(\lambda_{1})(S\lambda_{2}),

that we also write c(λ1)(λ2)c(\lambda_{1})(\lambda_{2}) or c(λ1),λ2\langle c(\lambda_{1}),\lambda_{2}\rangle if the inclusion S:ΛNS:\Lambda\to N_{\mathbb{R}} is implicit. We have the following proposition, already noticed in [18].

Proposition 2.2.

For any H1(𝕋A,Aff)\mathcal{L}\in H^{1}(\mathbb{T}A,\mathrm{Aff}_{\mathbb{Z}}), the bilinear form BB induced by its Chern class c:ΛMc:\Lambda\to M is symmetric.

Proof.

The map to H2(𝕋A,)=Λ2H1(𝕋A,)H^{2}(\mathbb{T}A,\mathbb{R})=\Lambda^{2}H^{1}(\mathbb{T}A,\mathbb{R}) corresponds to the skew-symmetrization. Hence the symmetry. ∎

These considerations allow us to find a decomposition of the Picard group H1(𝕋A,Aff)H^{1}(\mathbb{T}A,\mathrm{Aff}_{\mathbb{Z}}): it contains the dual torus 𝕋A\mathbb{T}A^{\vee}, and surjects to a discrete lattice: the elements of hom(Λ,M)\hom(\Lambda,M) that induce a symmetric pairing on Λ\Lambda_{\mathbb{R}}. Let us describe this lattice in coordinates. Choosing basis of Λ\Lambda and MM, we have an identification of hom(Λ,M)\hom(\Lambda,M) with 2()\mathcal{M}_{2}(\mathbb{Z}), and let SS be the matrix of the inclusion ΛN\Lambda\hookrightarrow N_{\mathbb{R}}.

Proposition 2.3.

A matrix cc belongs to Imc1\mathrm{Im}\ c_{1} if and only if STcS^{T}c is symmetric: STc𝒮2()S^{T}c\in\mathcal{S}_{2}(\mathbb{R}).

Remark 1.

Notice that this condition is equivalent to Sc1Sc^{-1} being symmetric when cc is invertible.

Proof 2.1.

The dual map MΛM\to\Lambda^{*}_{\mathbb{R}} is given by the matrix STS^{T}. Meanwhile, cc is the matrix of a map ΛM\Lambda\to M. Hence, the matrix of the bilinear map BB is given by STcS^{T}c, which is by composition a map ΛΛ\Lambda\to\Lambda_{\mathbb{R}}^{*}.

Thus, the image of the Chern class map is the intersection of the lattice hom(Λ,M)hom(Λ,M)\hom(\Lambda,M)\subset\hom(\Lambda,M)_{\mathbb{R}} with some hyperplane depending on the inclusion SS. In particular, if it is chosen generically, it is empty. One other way to view it is that a given class in hom(Λ,M)\hom(\Lambda,M) being realized as a Chern class imposes a condition on the inclusion SS.

Example 2.2.

Assume we have S=(αβγδ)S=\begin{pmatrix}\alpha&\beta\\ \gamma&\delta\\ \end{pmatrix} and c=(n001)c=\begin{pmatrix}n&0\\ 0&1\\ \end{pmatrix}. Then cc is realized as a Chern class in 𝕋A\mathbb{T}A if and only if nβ=γn\beta=\gamma. \lozenge

We finish by giving the definition of a polarization.

Definition 2.4.

The choice of a polarization on 𝕋A\mathbb{T}A is the data of a cc such that BB is symmetric and positive definite bilinear form on Λ\Lambda.

Concretely, a line bundle with chern class cc can be seen as the quotient of the trivial line bundle N×N_{\mathbb{R}}\times\mathbb{R} by the following action of Λ\Lambda, extending the action by translation on NN_{\mathbb{R}}:

λ(x,ξ)=(x+λ,ξ+c(λ)(x)+βλ).\lambda\cdot(x,\xi)=(x+\lambda,\xi+c(\lambda)(x)+\beta_{\lambda}).

The βλ\beta_{\lambda} are some real numbers. For this formula to define an action, we need to have the following relation:

βλ+μ=βλ+βμ+c(λ)(μ).\beta_{\lambda+\mu}=\beta_{\lambda}+\beta_{\mu}+c(\lambda)(\mu).

This is another way to see that cc defines a symmetric form. It implies that βλ=12c(λ)(λ)\beta_{\lambda}=\frac{1}{2}c(\lambda)(\lambda) up to a linear map in λ\lambda. This linear map corresponds to the Λ/M\Lambda^{*}_{\mathbb{R}}/M part in the Picard group and is defined by translations of the line bundle.

2.2 complex and tropical θ\theta-functions

The goal of this section is to define tropical θ\theta-function in a setting more general than the one of [18], recall the definition of θ\theta-functions in the complex setting, as in [12] but with multiplicative notations, and relate both when considering a Mumford family of tori At\mathbb{C}A_{t}.

2.2.1 Tropical θ\theta-functions.

Let us consider the line bundle \mathcal{L} with fixed Chern class cc, that is the quotient of N×N_{\mathbb{R}}\times\mathbb{R} by the action λ(x,ξ)=(x+λ,ξ+c(λ)(x)+βλ)\lambda\cdot(x,\xi)=(x+\lambda,\xi+c(\lambda)(x)+\beta_{\lambda}) for βλ=12c(λ)(λ)\beta_{\lambda}=\frac{1}{2}c(\lambda)(\lambda). All line bundles with Chern class cc are translate of this specific line bundle.

Definition 2.5.

A θ\theta-function for \mathcal{L} is a global convex piecewise affine function Θ\Theta on NN_{\mathbb{R}} with integer slope and satisfying the quasi-periodicity condition

Θ(x+λ)=Θ(x)+c(λ)(x)+βλ.\Theta(x+\lambda)=\Theta(x)+c(\lambda)(x)+\beta_{\lambda}.

We have the following proposition that expresses all the θ\theta-functions.

Proposition 2.

Let {m0}\{m_{0}\} be a system of representatives of MM modulo c(Λ)c(\Lambda). Any θ\theta-function is of the form θ(x)=maxm0(am0+θm0(x))\theta(x)=\max_{m_{0}}\left(a_{m_{0}}+\theta_{m_{0}}(x)\right) for some choice of coefficients am0a_{m_{0}}, where

Θm0(x)=m0,x+maxlΛ(c(l),x12c(l),lm0,l),\Theta_{m_{0}}(x)=\langle m_{0},x\rangle+\max_{l\in\Lambda}\left(\langle c(l),x\rangle-\frac{1}{2}\langle c(l),l\rangle-\langle m_{0},l\rangle\right),
Proof 2.3.

As a convex function on NN_{\mathbb{R}}, a θ\theta-function Θ\Theta is determined by its Legendre transform:

Θ^(m)=maxxN(m,xΘ(x)).\widehat{\Theta}(m)=\max_{x\in N_{\mathbb{R}}}\left(\langle m,x\rangle-\Theta(x)\right).

The quasi-periodicity condition on Θ\Theta translates to the following quasi-periodicity condition on Θ^\widehat{\Theta}:

Θ^(m+c(λ))\displaystyle\widehat{\Theta}(m+c(\lambda)) =maxxN(m+c(λ),xΘ(x))\displaystyle=\max_{x\in N_{\mathbb{R}}}\left(\langle m+c(\lambda),x\rangle-\Theta(x)\right)
=maxxN(m,x+c(λ),xΘ(x)βλ)+βλ\displaystyle=\max_{x\in N_{\mathbb{R}}}\left(\langle m,x\rangle+\langle c(\lambda),x\rangle-\Theta(x)-\beta_{-\lambda}\right)+\beta_{-\lambda}
=maxxN(m,xΘ(xλ))+βλ\displaystyle=\max_{x\in N_{\mathbb{R}}}\left(\langle m,x\rangle-\Theta(x-\lambda)\right)+\beta_{-\lambda}
=Θ^(m)m,λ+βλ.\displaystyle=\widehat{\Theta}(m)-\langle m,\lambda\rangle+\beta_{-\lambda}.

It means that the values of the function Θ^\widehat{\Theta} are determined by a system of representatives modulo c(Λ)Mc(\Lambda)\subset M. We check that the functions Θm0\Theta_{m_{0}} satisfy the relations, as in [18]. They are well-defined since the function lc(l),ll\mapsto\langle c(l),l\rangle is positive definite. The result follows.

Remark 3.

The tropical θ\theta-functions are functions on NN_{\mathbb{R}}, but they can be seen as section of line bundles on 𝕋A=N/Λ\mathbb{T}A=N_{\mathbb{R}}/\Lambda.

2.2.2 Complex θ\theta-functions.

Concerning the complex line bundles on a complex abelian surface, we refer to the corresponding section of [12] for a detailed version of the following. Consider the abelian surface A=N/Λ\mathbb{C}A=N_{\mathbb{C}^{*}}/\Lambda. This is a multiplicative presentation of an abelian surface. In [12], an abelian surface is presented as a quotient N/ΩN_{\mathbb{C}}/\Omega for a matrix Ω=(I2Z~)\Omega=\begin{pmatrix}I_{2}&\tilde{Z}\end{pmatrix}, where Z~:ΛN\tilde{Z}:\Lambda\to N_{\mathbb{C}} is some matrix. The relation between both notations is provided by the exponential map

z~Nz=e2iπz~N.\tilde{z}\in N_{\mathbb{C}}\longmapsto z=e^{2i\pi\tilde{z}}\in N_{\mathbb{C}^{*}}.

In order to relate the complex setting to the tropical setting, we need to adopt the multiplicative notation. However, the reader might be more accustomed to the additive notation. Thus, we try to specify both notations as often as possible. Notation with a \sim are for the additive setting, and without for the multiplicative one.

Consider a line bundle on A\mathbb{C}A whose pull-back to NN_{\mathbb{C}^{*}} is the trivial bundle N×N_{\mathbb{C}^{*}}\times\mathbb{C}. Its Chern class is a skew-symmetric map on the lattice NΛN\oplus\Lambda. Assume it is of the form Q=(0ccT0)Q=\begin{pmatrix}0&-c\\ c^{T}&0\\ \end{pmatrix}, where c:ΛMc:\Lambda\to M, and satisfies the Riemann bilinear relation from [12]: Z~c1\tilde{Z}c^{-1} is symmetric and 𝔪(Z~c1)\mathfrak{Im}(-\tilde{Z}c^{-1}) is positive. Moreover, there exists a function α:Λ\alpha:\Lambda\to\mathbb{C}^{*} such that the line bundle is obtained quotienting N×N_{\mathbb{C}^{*}}\times\mathbb{C} by the action of Λ\Lambda:

λ(z,ξ)=(λz,αλzc(λ)ξ).\lambda\cdot(z,\xi)=(\lambda\cdot z,\alpha_{\lambda}z^{c(\lambda)}\xi).

For this relation to define an action, we need α\alpha to satisfy the relation αλ+μ=αλαμλc(μ)\alpha_{\lambda+\mu}=\alpha_{\lambda}\alpha_{\mu}\lambda^{c(\mu)}. In particular, λc(μ)=μc(λ)\lambda^{c(\mu)}=\mu^{c(\lambda)}, which amounts to the Riemann bilinear relation Z~c1\tilde{Z}c^{-1} is symmetric. For what follows, we fix αλ=eiπc(λ),Zλ\alpha_{\lambda}=e^{i\pi\langle c(\lambda),Z\lambda\rangle}, which satisfies the relation. This fixes the line bundle which was previously defined up to translation.

Definition 2.6.

A θ\theta-function is a holomorphic function on NN_{\mathbb{C}} (resp. NN_{\mathbb{C}^{*}}) that satisfies the following relations:

{θ(z~+n)=θ(z~) for nNN,θ(z~+Z~λ)=αλe2iπc(λ),z~θ(z~) or multiplicatively θ(λz)=αλzc(λ)θ(z).\left\{\begin{array}[]{l}\theta(\tilde{z}+n)=\theta(\tilde{z})\text{ for }n\in N\subset N_{\mathbb{C}},\\ \theta(\tilde{z}+\tilde{Z}\lambda)=\alpha_{\lambda}e^{2i\pi\langle c(\lambda),\tilde{z}\rangle}\theta(\tilde{z})\\ \end{array}\right.\text{ or multiplicatively }\theta(\lambda\cdot z)=\alpha_{\lambda}z^{c(\lambda)}\theta(z).

These relations need only to check it on a basis of the lattice. If (e1,e2)(e_{1},e_{2}) is a basis of NN and (λ1,λ2)(\lambda_{1},\lambda_{2}) a basis of Λ\Lambda,

{θ(z~+ej)=θ(z~)θ(z~+Z~λj)=e2iπc(λj),z~+iπc(λj),Z~λjθ(z~) or multiplicatively θ(λjz)=eiπc(λj),Z~λjzc(λj)θ(z).\left\{\begin{array}[]{l}\theta(\tilde{z}+e_{j})=\theta(\tilde{z})\\ \theta(\tilde{z}+\tilde{Z}\lambda_{j})=e^{2i\pi\langle c(\lambda_{j}),\tilde{z}\rangle+i\pi\langle c(\lambda_{j}),\tilde{Z}\lambda_{j}\rangle}\theta(\tilde{z})\\ \end{array}\right.\text{ or multiplicatively }\theta(\lambda_{j}\cdot z)=e^{i\pi\langle c(\lambda_{j}),\tilde{Z}\lambda_{j}\rangle}z^{c(\lambda_{j})}\theta(z).
Proposition 4.

Any θ\theta-function is a linear combination of the following θ\theta-functions:

θm0(z~)=e2iπm0,z~lΛeiπc(l),Zl2iπm0,Zle2iπc(l),z~\displaystyle\theta_{m_{0}}(\tilde{z})=e^{2i\pi\langle m_{0},\tilde{z}\rangle}\sum_{l\in\Lambda}e^{-i\pi\langle c(l),Zl\rangle-2i\pi\langle m_{0},Zl\rangle}e^{2i\pi\langle c(l),\tilde{z}\rangle}
or multiplicatively θm0(z)=zm0lΛeiπc(l),Zl2iπm0,Zlzc(l).\displaystyle\theta_{m_{0}}(z)=z^{m_{0}}\sum_{l\in\Lambda}e^{-i\pi\langle c(l),Zl\rangle-2i\pi\langle m_{0},Zl\rangle}z^{c(l)}.
Proof 2.4.

Following [12], any θ\theta-function can be developped in power series as follows

θ(z~)=mMame2iπm,z~ or multiplicatively θ(z)=mMamzm.\theta(\tilde{z})=\sum_{m\in M}a_{m}e^{2i\pi\langle m,\tilde{z}\rangle}\text{ or multiplicatively }\theta(z)=\sum_{m\in M}a_{m}z^{m}.

As in the tropical case, the relation imposes the following relations on the coefficients

amc(λ)=e2iπm,Z~λeiπc(λ),Z~λam.a_{m-c(\lambda)}=e^{2i\pi\langle m,\tilde{Z}\lambda\rangle}e^{-i\pi\langle c(\lambda),\tilde{Z}\lambda\rangle}a_{m}.

Thus, all the coefficients are determined by the coefficients am0a_{m_{0}} for {m0}\{m_{0}\} a system of representatives of the quotient M/c(Λ)M/c(\Lambda). We check that the functions θm0\theta_{m_{0}} satisfy the relations. These are well-defined since lc(l),Zll\mapsto\langle c(l),Zl\rangle has negative imaginary part, so that the series converge. The result follows.

2.2.3 Varying the surface.

We now consider a family of abelian surfaces At=N/eAtS\mathbb{C}A_{t}=N_{\mathbb{C}^{*}}/\langle e^{A}t^{S}\rangle with a line bundle having Chern class c:ΛMc:\Lambda\to M. This corresponds to the choice Zt=12iπ(A+Slogt)Z_{t}=\frac{1}{2i\pi}(A+S\log t), where AA is a complex matrix, and SS is an integer matrix chosen such that Sc1Sc^{-1} is symmetric and positive definite. The study of the complex setting tells us that we can take as a basis of sections the θ\theta-functions having the following multiplicative expressions:

θm0(z)=zm0lΛe12c(l),Alm0,Alt12c(l),Slm0,Slzc(l).\theta_{m_{0}}(z)=z^{m_{0}}\sum_{l\in\Lambda}e^{-\frac{1}{2}\langle c(l),Al\rangle-\langle m_{0},Al\rangle}t^{-\frac{1}{2}\langle c(l),Sl\rangle-\langle m_{0},Sl\rangle}z^{c(l)}.

Taking logt\log_{t} and the limit as tt goes to \infty, we have the following propoition.

Proposition 5.

The tropical limit of the θ\theta-function θm0\theta_{m_{0}} is the tropical θ\theta-function Θm0\Theta_{m_{0}}.

2.3 Curves in abelian surfaces

2.3.1 Parametrized curves.

For sake of completeness of the present paper, we include a quick reminder of tropical curves in abelian surfaces from the parametric point of view and refer the reader to the first paper for more details.

Definition 2.7.

An abstract tropical curve is a finite metric graph Γ\Gamma. A parametrized tropical curve is a map h:Γ𝕋Ah:\Gamma\to\mathbb{T}A such that:

  • -

    If ee is an edge, hh is affine with integer slope on the edges of Γ\Gamma. The slope is of the form weuew_{e}u_{e}, where wew_{e} is an integer called weight of the edge, and ueNu_{e}\in N is a primitive vector. We set δΓ=gcdewe\delta_{\Gamma}=\mathrm{gcd}_{e}w_{e} to be the gcd of the weight of the edges.

  • -

    For each vertex VV of Γ\Gamma, one has the balancing condition: eVweue=0\sum_{e\ni V}w_{e}u_{e}=0, where weuew_{e}u_{e} denotes the slope of hh on ee oriented outside VV.

Definition 2.8.

Let h:Γ𝕋Ah:\Gamma\to\mathbb{T}A be a parametrized tropical curve.

  • -

    The curve is irreducible if Γ\Gamma is connected.

  • -

    If the curve is irreducible, its genus is the first Betti number b1(Γ)b_{1}(\Gamma).

  • -

    Its degree is the class that it realizes inside H1,1(𝕋A)ΛNH_{1,1}(\mathbb{T}A)\simeq\Lambda\otimes N.

  • -

    The curve is said to be simple if Γ\Gamma is trivalent en hh is an immersion.

We also have the following statement proved in the first paper that enables a concrete computation of the degree CC of a tropical curve h:Γ𝕋Ah:\Gamma\to\mathbb{T}A.

Proposition 6.

Let CΛNC\in\Lambda\otimes N be a class and S:ΛNS:\Lambda\hookrightarrow N_{\mathbb{R}} be the inclusion defining the abelian surface 𝕋A=N/Λ\mathbb{T}A=N_{\mathbb{R}}/\Lambda. Then,

  • -

    The class CC is realized by a tropical curve if and only if the product CST𝒮2++()CS^{T}\in\mathcal{S}_{2}^{++}(\mathcal{R}).

  • -

    Given a parametrized tropical curve h:ΓNh:\Gamma\to N_{\mathbb{R}}, the class CΛNC\in\Lambda\otimes N that it realizes is obtained by adding the slopes of the edges intersected by two loops realizing a basis of H1(𝕋A,)H_{1}(\mathbb{T}A,\mathbb{Z}).

2222
(a)(a) (b)(b) (c)(c)
Figure 1: Three examples of tropical curves in tropical tori.
Example 2.5.

On Figure 1, we can see three different examples of tropical curves in different tropical tori 𝕋A\mathbb{T}A. Each tropical torus is represented either by a parallelogram or by an hexagon whose pairs of opposite edges have been glued together. The curve (a)(a) has genus 22 and degree (1001)\begin{pmatrix}1&0\\ 0&1\\ \end{pmatrix}, curve (b)(b) genus 55 and degree (2003)\begin{pmatrix}2&0\\ 0&3\\ \end{pmatrix}, and curve (c)(c) genus 22 and degree (2101)\begin{pmatrix}2&1\\ 0&1\\ \end{pmatrix}. \lozenge

2.3.2 Implicit curves.

In the planar setting, tropical curve inside NN_{\mathbb{R}} have two possible descriptions: they are either the image of a parametrized tropical curve, or the corner locus of a tropical polynomial. For an abelian surface with a line bundle, the sections play the role of the tropical polynomials, and the distinguished tropical θ\theta-functions Θm0\Theta_{m_{0}} defined in the previous section play the role of the monomials in NN_{\mathbb{R}}.

Definition 2.9.

Let cc be the Chern class of a line bundle \mathcal{L}. Let Θ(x)=maxm0(am0+Θm0(x))\Theta(x)=\max_{m_{0}}\left(a_{m_{0}}+\Theta_{m_{0}}(x)\right) be a tropical θ\theta-function, defining a section of \mathcal{L}. Then the corner locus of θ\theta is called a planar tropical curve in 𝕋A\mathbb{T}A.

As for tropical curves inside NN_{\mathbb{R}}, any planar tropical curve in 𝕋A\mathbb{T}A admits a parametrization h:Γ𝕋Ah:\Gamma\to\mathbb{T}A. A planar tropical curve is irreducible if it cannot be written as the union of two planar tropical curves. The genus of an irreducible planar tropical curve is the minimal genus among the possible parametrizations of the curve.

The following proposition relates the parametric point of view and the implicit point of view by giving the relation between the class CC realized by a tropical curve Γ\Gamma, and the Chern class cc of the line bundle O(Γ)O(\Gamma).

Proposition 7.

Let h:Γ𝕋Ah:\Gamma\to\mathbb{T}A be a parametrization of a tropical curve that is a section of a line bundle \mathcal{L} with Chern class c:ΛMc:\Lambda\to M. Let C:ΛNC:\Lambda^{*}\to N be the degree of the parametrized tropical curve. Then CC and cc are Poincaré dual to each other: cc is the comatrix of CC.

Proof 2.6.

The statement comes from Poincaré duality inside 𝕋A\mathbb{T}A: the integration of cc over a cycle is obtained by intersecting the cycle with CC.

Up to a multiplicative constant, it means that C=(c1)TC=(c^{-1})^{T}. We can then check that the conditions Sc1𝒮2++()Sc^{-1}\in\mathcal{S}^{++}_{2}(\mathbb{R}) and CST𝒮2++()CS^{T}\in\mathcal{S}^{++}_{2}(\mathbb{R}) are equivalent. By abuse of notation, by tropical curve we mean a parametrized tropical curve or a planar tropical curve.

Remark 8.

Notice that as in the case of tropical curves inside NN_{\mathbb{R}}, tropical curves are dual to some particular subdivisions of the torus M/c(Λ)M_{\mathbb{R}}/c(\Lambda).

Example 2.7.

The curve on Figure 1 (a)(a) is the corner locus of the unique θ\theta-function associated to the principal polarization of 𝕋A\mathbb{T}A. The curve on (c)(c) is the corner locus of one pf the two θ\theta-functions given by the polarization. To get the curve on (b)(b), one needs several θ\theta-functions. \lozenge

2.4 From curves to linear systems

We have seen that there are two points of views for curves inside abelian surfaces: either as parametrized curve, or as zero-locus of a section of some line bundle \mathcal{L}. Tautologically, a curve 𝒞\mathscr{C} is always the zero-locus of a section of the line bundle 𝒪(𝒞)\mathcal{O}(\mathscr{C}). The goal of this section is to relate the two points of views: how to recover the line bundle 𝒪(𝒞)\mathcal{O}(\mathscr{C}) from a parametrization of the curve. By that, we mean that given two curves 𝒞0\mathscr{C}_{0} and 𝒞1\mathscr{C}_{1} in the same homology class, how to recover the element 𝒪(𝒞1𝒞0)\mathcal{O}(\mathscr{C}_{1}-\mathscr{C}_{0}) of the dual torus Λ/M\Lambda^{*}_{\mathbb{C}^{*}}/M.

Let A\mathbb{C}A be an abelian surface and 𝒞0\mathscr{C}_{0} and 𝒞1\mathscr{C}_{1} be two curves inside A\mathbb{C}A realizing the same homology class. The divisor 𝒞1𝒞0\mathscr{C}_{1}-\mathscr{C}_{0} defines a line bundle 𝒪(𝒞1𝒞0)\mathcal{O}(\mathscr{C}_{1}-\mathscr{C}_{0}) of degree 0. It belongs to the zero-component of the Picard group of A\mathbb{C}A. If 𝒞0\mathscr{C}_{0} and 𝒞1\mathscr{C}_{1} were given as zero locus of sections of their respective line bundles represented by two θ\theta-functions θ0,θ1:N\theta_{0},\theta_{1}:N_{\mathbb{C}^{*}}\to\mathbb{C}, then the quotient satisfies the relation

θ1θ0(λz)=Kλθ1θ0(z),\frac{\theta_{1}}{\theta_{0}}(\lambda\cdot z)=K_{\lambda}\frac{\theta_{1}}{\theta_{0}}(z),

for some KλK_{\lambda} where λΛKλ\lambda\in\Lambda\mapsto K_{\lambda}\in\mathbb{C}^{*} is a morphism. As θ1\theta_{1} and θ0\theta_{0} are defined up to monomials, KK is only well-defined up to the action of MM, which is not surprising since it is an element in the dual torus. If Kλ=1K_{\lambda}=1, it would mean that 𝒞0\mathscr{C}_{0} and 𝒞1\mathscr{C}_{1} belong to the same linear system: they are equivalent since θ1θ0\frac{\theta_{1}}{\theta_{0}} descends to a meromorphic function on A\mathbb{C}A. Finding the line bundle 𝒪(𝒞1𝒞0)\mathcal{O}(\mathscr{C}_{1}-\mathscr{C}_{0}) amounts to find KK. This is more or less straightforward if θ1\theta_{1} and θ0\theta_{0} are given. However, this is not the case if the curves are given by parametrizations φ0:C0A\varphi_{0}:\mathbb{C}C_{0}\to\mathbb{C}A and φ1:C1A\varphi_{1}:\mathbb{C}C_{1}\to\mathbb{C}A.

First, we explain how to recover KK from the curves in the tropical case and close to the tropical limit before getting a general statement for complex tori.

2.4.1 Tropically and close to the tropical limit

Let h0:Γ0𝕋Ah_{0}:\Gamma_{0}\to\mathbb{T}A and h1:Γ1𝕋Ah_{1}:\Gamma_{1}\to\mathbb{T}A be two parametrized tropical curves realizing the same homology class CC in a tropical abelian variety 𝕋A=N/Λ\mathbb{T}A=N_{\mathbb{R}}/\Lambda. Both curves lifts to periodic curves inside NN_{\mathbb{R}} that we denote by the same letters, and that are the corner locus of a tropical θ\theta-function Θ0\Theta_{0} and Θ1\Theta_{1}, defined up to addition of an affine function with slope in MM. The function f(x)=Θ1(x)Θ0(x)f(x)=\Theta_{1}(x)-\Theta_{0}(x) satisfies the quasi-periodicity relation

f(x+λ)=f(x)+Kλ,f(x+\lambda)=f(x)+K_{\lambda},

where λKλ\lambda\mapsto K_{\lambda} is linear. Adding a monomial to one of the θ\theta functions changes ff and thus KK. Thus, only the class of KK in Λ/M\Lambda^{*}_{\mathbb{R}}/M is well-defined. We intend to recover the element KΛ/MK\in\Lambda^{*}_{\mathbb{R}}/M.

For the statement as well as the proof, recall the notion of moment of an edge for a tropical curve. Let h:ΓNh:\Gamma\to N_{\mathbb{R}} be a tropical curve, ee be an oriented edge where hh has slope weuew_{e}u_{e}. The moment of the oriented edge is det(weue,p)\det(w_{e}u_{e},p), where pp is any point on the edge. It corresponds to the position of the edge along a transversal axis.

Theorem 2.10.

Let h0:Γ0𝕋Ah_{0}:\Gamma_{0}\to\mathbb{T}A and h1:Γ1𝕋Ah_{1}:\Gamma_{1}\to\mathbb{T}A be two parametrized tropical curves realizing the same homology class CC, lifting to periodized curves inside NN_{\mathbb{R}}. Choose some point p0p_{0} in the complement of the curves inside NN_{\mathbb{R}}. The degree 0 line bundle 𝒪(Γ1Γ0)\mathcal{O}(\Gamma_{1}-\Gamma_{0}) in the dual torus Λ/M\Lambda^{*}_{\mathbb{R}}/M is represented by the element KK of Λ\Lambda^{*}_{\mathbb{R}} that maps a class λ\lambda to the following: choose a path γ\gamma between p0p_{0} and p0+λp_{0}+\lambda, and add the moments of the oriented edges of Γ0\Gamma_{0} and Γ1\Gamma_{1} met by the path, oriented so that the intersection sign between γ\gamma and the oriented edge is - for Γ1\Gamma_{1} and ++ for Γ0\Gamma_{0}.

Remark 9.

It is important to take the same base-point for lifting all the loops: during the computation, we assume that the slope of ff at p0p_{0} is 0, and this determines uniquely ff, not up to addition of a monomial anymore.

Proof 2.8.

Up to addition of a monomial, we can assume that ff has slope 0 at p0p_{0}. Let λ\lambda be an element of Λ\Lambda, lifted to a path between p0p_{0} and p0+λp_{0}+\lambda that intersects Γ0\Gamma_{0} and Γ1\Gamma_{1} transversally. Let x1,,xn1x_{1},\dots,x_{n-1} be the intersection points between the path and the lifts of Γ0\Gamma_{0} and Γ1\Gamma_{1}. Let also x0=p0x_{0}=p_{0} and xn=p0+λx_{n}=p_{0}+\lambda. Let mim_{i} is the slope of ff in the region where the path between xix_{i} and xi+1x_{i+1} lies. By assumption, m0=mn=0m_{0}=m_{n}=0. We have that

Kλ=f(x0+λ)f(x0)\displaystyle K_{\lambda}=f(x_{0}+\lambda)-f(x_{0}) =i=0n1f(xi+1)f(xi)\displaystyle=\sum_{i=0}^{n-1}f(x_{i+1})-f(x_{i})
=i=0n1mi,xi+1xi,\displaystyle=\sum_{i=0}^{n-1}\langle m_{i},x_{i+1}-x_{i}\rangle,
=i=1nmi1,xii=0n1mi,xi\displaystyle=\sum_{i=1}^{n}\langle m_{i-1},x_{i}\rangle-\sum_{i=0}^{n-1}\langle m_{i},x_{i}\rangle
=i=1n1mi1mi,xi.\displaystyle=\sum_{i=1}^{n-1}\langle m_{i-1}-m_{i},x_{i}\rangle.

Furthermore, xix_{i} belongs to an edge ee of either Γ0\Gamma_{0} or Γ1\Gamma_{1} directed by weuew_{e}u_{e}. The monomial mimi1m_{i}-m_{i-1} is equal to det(weue,)\det(w_{e}u_{e},-) when ee is oriented so that the intersection index eγe\cdot\gamma at xix_{i} is positive. The result follows since Γ1\Gamma_{1} appears positively in the divisor and