Compactifying moduli spaces

By Lucia Caporaso

Abstract

The boundary of some well known algebro-geometric moduli spaces is described by highlighting the recursive combinatorial properties in connection with tropical and non-Archimedean geometry.

1. Introduction

A characteristic of algebraic geometry is the fact that the sets parametrizing equivalence classes of a certain type of objects (for example, smooth projective curves of given genus up to isomorphism or hypersurfaces of given degree in projective space up to projective equivalence) are themselves endowed with a natural algebraic structure, and are called moduli spaces. In fact, the structure of a moduli space is largely governed by the geometric properties of the parametrized objects. This phenomenon, known and used for a long time, has been established on rigorous mathematical ground during the second half of the twentieth century when moduli theory flourished; see Reference 18, Reference 23, Reference 28, Reference 19, Reference 3.

A moduli problem is thus a way of posing a classification problem. Loosely speaking, this amounts to considering a class of objects to parametrize up to some equivalence relation and a notion of a family of objects, specifying how our objects are allowed to move within . A family is, essentially, a fibration of objects in over an algebraic variety or scheme , the base of the family. The precise definition depends on the particular moduli problem; the common point is that to any point in the fiber of the family over is an object in . To say that a moduli problem admits a moduli space is to say that there is a natural algebraic structure on the set of all equivalence classes on , which makes into an algebraic variety or scheme (or a more sophisticated space, like a stack). The algebraic structure on must satisfy some basic properties. As we said, the points of are in bijective correspondence with the equivalence classes of objects in . Next, for every family of objects in over a base , the set theoretic map

sending a point in to the equivalence class of its fiber in the family, is a morphism of algebraic varieties or schemes (or stacks). One usually refers to as the moduli map of the family.

Without going into details, let us mention two subtle points. First of all, does a given moduli map determine the family uniquely? Second, is every morphism the moduli map of some family of objects in ? The answer to such questions is seldom positive. In case it is positive, we say that is a fine moduli space.

Let us look at the geometric structure of our moduli spaces. It turns out that many of them are not complete as topological spaces, or, in algebro-geometric language, they are not projective. This reflects the fact that the objects they parametrize, the elements of , are bound to degenerate to objects that are not in . Constructing completions, or compactifications, for moduli spaces has been a major area of research in algebraic geometry since at least the mid-1950s. On the one hand, compactifications of moduli spaces are projective and, hence, more tractable in computations and applications. On the other hand, they usually come with a geometric description and are moduli spaces themselves. This provides new geometric insights and has been a constant source of inspiration for new progress.

In recent years a new interesting connection has been established between the compactification of certain moduli spaces and moduli spaces of polyhedral objects. In loose words, the “skeleton” of some compactified algebro-geometric moduli space is expressed as the moduli space of the skeleta of the objects parametrized by . The connection relies on the study of the boundary of the compactification and on its recursive, combinatorial properties, some of which have long been known but are now viewed from a new perspective.

We will survey this area by focusing on the moduli spaces of stable curves and of line bundles on curves. In the first case we have a clear picture of the above-mentioned connection; in the second, partial results are known, but a complete understanding of the connection with polyhedral geometry is not yet available.

2. Compactifying moduli spaces of smooth curves

2.1. Moduli of smooth and stable curves

The moduli problem for smooth (connected, projective) curves was perhaps the first moduli problem ever studied. A curve is a projective and connected variety over an algebraically closed field, unless we specify otherwise. Classically, over the field a smooth curve is a compact, connected, orientable surface (a two-dimensional topological manifold) endowed with an algebraic structure, and our moduli problem was studied already by Riemann.

Now, is the set of all smooth curves of genus , and the equivalence is the isomorphism. A family is a flat, projective morphism such that for every point in , the fiber is a smooth curve of genus .

The moduli space of smooth algebraic curves of genus is an algebraic variety denoted by . As we said, its structure captures some of the properties of the curves it parametrizes. In particular, if , it is not a projective variety, since smooth curves of positive genus are often forced to specialize to singular ones. As we said in the introduction, it is quite important to find compactifications for , and we shall now concentrate on this.

The best known compactification is a projective variety whose boundary points (i.e., the points in ) parametrize singular curves of a very simple type, called stable curves, and were introduced for the first time by P. Deligne, A. Meyer, and D. Mumford. A remarkable feature of stable curves (whose definition will come soon) is that they can be characterized inductively using smooth curves of lower genus. In order to do this, we need to extend our range a little and consider not only smooth curves, but smooth curves together with a set of marked (and ordered) points. Our has to be chosen so that every curve of genus with marked points has only finitely many automorphisms fixing the marked points. Thus, we need if and if ; hence, we assume from now on .

We denote by the moduli space of smooth curves of genus with marked points. If , we just write .

Example 2.1.1.

The simplest cases are and , which are almost trivial. Indeed is a point since any smooth curve of genus is isomorphic to , and any two ordered triples of points in are mapped to one another by an automorphism.

If and , then has dimension , and, if we work over , it is not hard to see that can be identified with .

The variety , known to be irreducible of dimension , is not complete unless and . We shall now describe its compactification by the moduli space of stable -pointed curves. A stable curve with marked points is defined as a (connected, projective) curve having only nodes as singularities, plus (ordered) smooth points on it, and having finitely many automorphisms fixing the marked points. We shall see a more explicit description soon.

Thanks to a series of remarkable achievements (see Reference 15, Reference 21, Reference 16) we know that that is an open dense subset in , and that is a union of strata described in terms of smooth curves of genus at most with marked points. The strata parametrize curves of a fixed topological type and are described via an important combinatorial tool, the dual graph of a curve, which we shall introduce below.

The basic observation is that to give a curve with exactly one node is the same as to give the desingularization of and the two branch points of that get identified in . More generally, to give a curve having only nodes as singularities is the same as giving its desingularization (a disjoint union of smooth curves, one for each irreducible component of ) and a set of pairs of points, one for each node. To reconstruct from the data of the smooth curves with marked points, one needs gluing data, which are encoded in the dual graph of .

will be a so-called vertex-weighted graph with legs (i.e., every vertex is weighted by a nonnegative integer, which we call the genus of the vertex) and may have some leg (i.e., some half-edge) adjacent to it. For example, in Figure 1 we have a graph with three legs, labeled by . In this paper the legs are always ordered and will be fixed by any automorphism. We draw vertices of genus  by an empty circle and vertices of positive genus by a “•”, with sometimes the genus as subscript. We denote by the graph obtained from by removing all legs.

To define the dual graph, let be a nodal curve with smooth points. Its dual graph has as set of vertices, , the set of irreducible components of , and as set of edges, , the set of nodes of . An edge joins the (at most two) vertices/components on which the node lies. For every marked point, the graph has a leg attached to the vertex/component in which the marked point lies. Finally, every vertex of the graph is assigned an integer equal to the geometric genus of the corresponding component.

The (arithmetic) genus of the curve is expressed through the first Betti number of its dual graph as

with , where is the number of connected components of (here, but not always in the rest of this paper, assumed to be ). The number is defined as the genus of the graph. For example, the graphs of the above picture have genus .

The requirement that is stable translates into (and is equivalent to) a simple requirement on , namely that vertices of weight (resp., ) have degree at least (resp., ). Such graphs are called stable and form a finite set denoted by . In Figure 1 the graph on the left is stable, whereas the graph on the right is not.

Now we can exhibit a stratification of . Let be a stable graph, let be the locus of curves whose dual graph is , and it is easy to see that is never empty. We have

Example 2.1.2.

In the case where and , encountered in Example 2.1.1, it is quite easy to list all stable graphs. Together with the graph having no edges and the four legs attached to its unique vertex, we have the three cases pictured in Figure 2. To each of them there corresponds a unique curve up to isomorphism, made of two smooth components (isomorphic to ) intersecting in one node, and each component has two marked points. The three distributions of the four points give the three different isomorphism classes. It is not hard to show that .

From now on, will denote a connected, projective curve having only nodes as singularities, defined over an algebraically closed field .

2.2. The strata of

Before analyzing the stratification Equation 1, we study its strata and describe them explicitly. Let be the automorphism group of , and recall that elements of fix every leg of . For example, the three graphs in Figure 2 have no automorphisms.

Recall what we said above about reconstructing a curve from the union of its (desingularized) components plus the marked points. We shall use it now to introduce the natural morphism

where is the degree of the vertex . Let be a curve in , and consider a point in . This will correspond to the disjoint union of stable pointed (smooth) curves, which maps to birationally via a surjective normalization (or desingularization) morphism, :

with . In fact, is the desingularization of mentioned earlier. The map glues some pairs in to the nodes of . The which are not glued to anything will correspond to the marked points of . We think of Equation 3 as a presentation of . Now, acts on the gluing data, and we may have different presentations of the same curve; see Example 2.2.2.

Example 2.2.1.

Let be the stable graph in Figure 1. We have

since the action of on the product is trivial. Indeed, the only nontrivial automorphism of , interchanging the two edges on the left, acts trivially on which is a point. The stable curve of genus associated to

with and of genus , is given by identifying

so that the legs correspond to . In these identifications we made a choice which, being the same for all curves parametrized by , is irrelevant. Below is a representation of these identifications on the graph.

Example 2.2.2.

Consider the graph in the picture below, stable of genus with one leg. We marked the identifications on the edges. Now contains the automorphism , which interchanges the two edges. We shall see that acts nontrivially.

In this case we have , and the action of on swaps the two marked points of and the first two marked points of . Indeed, the following two elements in

are conjugated by and give the same point of . In other words, they are different presentations for the same curve.

Example 2.2.3.

We now consider and list the strata presented by it, with for all and . It is easy to check that . If , we leave it to the reader to check that and to find the corresponding unique (up to labeling the legs) stable graph. The remaining cases are drawn in Figures 3 and 4.

Let be a stable graph of genus with legs; recall that denotes the graph obtained from by removing all legs. To describe the inductive structure of , we need some notation.

If is a set of edges of , the graph is the subgraph obtained by removing the edges in , so that and have the same vertices. We denote by the graph obtained by replacing every edge in by a pair of legs attached to the vertices adjacent to . Hence has legs and

Consider the set of all such graphs

For any it is convenient to denote by the set of edges of such that .

We have a lattice structure (i.e., a poset structure with a unique maximum and a unique minimum) on induced by the inclusion, so that if . Of course, the maximum is and the minimum .

If is connected, is a stable graph of genus with legs. Therefore, we can consider the locus of curves in having as dual graph. More generally, if with connected, then one easily checks that is a stable graph for every , and we set

Of course, if , we recover , which we encountered earlier in Equation 2. In fact, generalizing Equation 2, for every , we have a natural surjection

which we call, again, a presentation of . The map described in Equation 2 is , and it factors for any as

The set , of all presentations of , is in a natural bijection with and hence has the lattice structure if .

If , we have a map and a factorization

In conclusion, the recursive structure of the stratum is described by the lattice .

Example 2.2.4.

In the next picture we have a stable graph and the lattice which, as we said, can be viewed as representing the recursive structure of . We denote by the graph obtained from by removing the edges with labels in .

2.3. The stratification of

We now go back to and recall that we have This decomposition has some properties reflecting how curves degenerate. Indeed, consider a family of curves in a stratum degenerating to a curve in a different stratum so that is the dual graph of . How are and related? Of course, every node of the curves in must specialize to a node of , i.e., an edge of . And the remaining nodes of (those that are not specializations of nodes of the curves in ) form a well defined set of edges of . Then the graph is obtained from by contracting to a vertex every edge of . And the converse turns out to be true: if is obtained from contracting a set of edges, we have that lies in the closure of .

We denote (weighted) edge-contractions as follows

where . Observe that the genus of a vertex of is defined as the genus of its preimage in , so that edge-contractions preserve genus and stability.

The simplest example is the contraction of every edge of . Then is the graph with one vertex of genus , no edges, and legs, so that is the dual graph of a smooth curve of genus , i.e., . Here the contraction corresponds to a family of smooth curves specializing to a curve in .

Now, we define a partial order on by setting if there exists a contraction . The poset is graded, and the rank function defined by is such that .

Example 2.3.1.

The picture of the poset is in Figure 5, with the horizontal levels marking the ranks and the minimum at the top.

So, the top vertex corresponds to . The two stata below correspond to the two codimension- strata of , parametrizing curves with exactly one node. The two bottom strata correspond to two points whose recursive presentation is in Figure 4.

The graded poset structure on corresponds to the poset structure on the strata of Equation 1 induced by inclusion of closures. More precisely, we have the following theorem.

Theorem 2.3.2.

The decomposition Equation 1 is a graded stratification of by the poset , i.e., the following properties hold.

(1)

(2)

is irreducible and locally closed.

(3)

The map sending to is a rank function.

Example 2.3.3.

We now describe . We say that two graphs have the same type if they differ only by the labeling on the legs. Then in we have three types of graphs: the graph with no edges and the two types below.

For the type on the left, we have different graphs. For the type on the right, we have different graphs (the number of partitions of shape , up to symmetry), corresponding to the top vertices of Figure 6.

2.4. Skeleta and tropical curves

Now we want to view as a category whose objects are stable graphs and whose arrows are generated by isomorphisms and edge-contractions. We shall construct a cone complex out of the category , as follows. To a graph we associate a cone

To an arrow (a composition of contractions and isomorphisms)

we associate an injective morphism of cones

If is the contraction of , then identifies with the face of , where the coordinates of the edges in are zero. If is an isomorphism, then corresponds to the bijection induced by . We can thus construct the colimit over and define it to be the skeleton of

By definition, the above space is a topological space and a so-called generalized cone complex. Let us give it an explicit geometric interpretation, so let be a point in it. Then there exists a stable graph such that lies in the interior of . Hence can be identified with , and with a set of positive real numbers , we think of as the length of the edge . Then is a so-called stable tropical curve, i.e., a stable graph whose edges have a length (a positive real number). By construction, two different points in correspond to nonisomorphic curves, and to every stable tropical curve there corresponds a point in .

Tropical curves have been studied as objects of independent interest, and, as such, they have a moduli space, , parametrizing isomorphism classes of stable tropical curves of genus with marked points, where a marked point of a tropical curve is a leg of the graph. As in the algebraic case, the legs are ordered and must be fixed by any automorphism. By its very construction, is a topological space and has the structure of a generalized cone complex; see Reference 22, Reference 8.

Our previous analysis on the geometric interpretation of enables us to conclude, at least set-theoretically, that the skeleton of can be identified with the moduli space of tropical curves of genus with legs:

It turns out that the above is not only a set-theoretic bijection, but it is an isomorphism of generalized cone complexes.

It is clear that the spaces above are not compact, but it is not hard to compactify them. We do that by extending our cones to endowed with the compact topology. This implies that we allow our tropical curves to have edges of infinite lengths. We thus obtain the compactified version of the above isomorphism (see Reference 1),

2.5. Curves over valuation fields and analytifications

The isomorphism Equation 6 is actually the reflection of a deep connection between algebraic and tropical curves. In loose words, a tropical curve encodes the local data of a family of smooth algebraic curves specializing to a singular one.

In algebraic geometry, local problems are often studied via local rings and valuation fields, whose associated algebraic schemes are the analogues of small balls in complex geometry. Let us set up the notation. We shall denote by a valuation field and by its valuation ring. The valuation of is a homomorphism, , from the multiplicative group to the additive group . We have and , so that is the unique maximal ideal of . We shall always assume that the residue field is algebraically closed. Now, the scheme associated to has a unique closed point, corresponding to the ideal , called special and denoted by . The scheme is open and dense in and is referred to as the generic point. It is useful to think of as having a limiting point built in itself: the special point of . In analogy with complex geometry, corresponds to a small ball about the origin, corresponds to the origin, and corresponds to the complement of the origin.

For the rest of this section we assume that is also complete with respect to a non-Archimedean valuation. Without going into technical details, let us mention that this has the effect of making the analysis as local as possible, enabling us to better handle limits.

Let be a smooth (connected, projective) curve over . As has a limiting point, one can ask whether has a limiting curve, i.e., if it can be completed over by adding a curve over . By the moduli properties of , we know that to there is associated a canonical moduli map mapping to the isomorphism class of . On the other hand we have a compactification of ; therefore, extends to a morphism

The image of the special point is thus a stable curve over the residue field of (which we assume is algebraically closed). In this way we associate to the stable curve . Now, is not a fine moduli space, hence there does not necessarily exist a family of curves over such that the fiber over is and the fiber over is . But, by the Stable Reduction Theorem, such a family exists for the base change of to some finite extension of non-Archimedean fields. In other words, denoting by the valuation ring of , there does exist a family over having as a generic fiber and as a special fiber.

On the one hand, of course, this involves the choice of and will therefore give different families if we vary that. But we can still ask what data do not depend on the choice of . We have already said that the curve will not change, hence its dual graph will not change either. But also, the local geometry of the family at the nodes of will be independent on the choice of , and this is precisely what will enable us to define a tropical curve, denoted by , associated to .

Our will have as underlying graph the dual graph of . To define the length of its edges, pick one of them . Then corresponds to a node of , which we shall denote by again. Consider a field extension as above, so that there is family

whose special fiber is and whose fiber over is the base change of , i.e., . Then, locally at the node , the equation of is for some in the maximal ideal of . We set the length of to be equal to where is the valuation of . We have thus defined the tropical curve , and this definition turns out to be independent of the choice of or of the local equation of at ; see Reference 29.

We shall say that the tropical curve is the skeleton, or the tropicalization, of .

Now, the construction we just described fits into a more general picture involving the theory of analytifications of algebraic schemes (developed by V. Berkovich) and its connection to tropical geometry. By the theory developed in Reference 6, to every algebraic scheme there corresponds an analytic space, the analytification , and this correspondence is functorial and profound.

We then introduce , the analytification of . A point in corresponds, up to base change, to a stable curve over an algebraically closed field complete with respect to a non-Archimedean valuation.

The analytic space has quite a complicated structure, but, by general results, it retracts onto a simpler subspace. Indeed, to every space with a boundary and a toroidal structure, one associates the (extended) Berkovich skeleton which is a generalized, extended cone complex onto which the analytification retracts; see Reference 27. Now, does not have a proper toroidal structure associated to its boundary , but the stack associated to it, written , does. This enables us to construct the Berkovich skeleton of . Now, recall definition Equation 5 and its compactified version . How does it compare with the Berkovich skeleton of the stack ? It turns out that they are naturally isomorphic, and hence we shall identify them. This allows us to apply Berkovich’s theory to , so that we have a retraction ,

To give a geometric interpretation of the map , we recall the isomorphism introduced in Equation 6. Composing it with we have the following.

Theorem 2.5.1.

The following tropicalization map

is a continuous, surjective map that sends the class of a stable curve over the (algebraically closed, complete, non-Archimedean) field to its skeleton .

See Reference 1. Concluding in loose words: the skeleton of is the moduli space of skeleta of stable curves.

The existence of analogous “correspondence” results in other situations has been investigated, and is under investigation, as of this writing. An important case in which it has been proved to hold is the compactification of the space of admissible covers, constructed in Reference 20. For details, we refer to the original paper Reference 13, where the authors also relate their results to the ones for moduli space of stable curves described above, proving that the various correspondence results are consistent with one another.

3. Compactifying Jacobians and Néron models

3.1. Moduli of line bundles on curves

Line bundles on curves and, more generally, on abstract algebraic varieties capture the information about the projective models of the varieties, i.e., their mappings in projective spaces. Here we focus on line bundles on curves, and on their moduli spaces. Motivated by our earlier discussion, we restrict to stable curves although some of what we are going to explain holds in greater generality. For example, if we release the condition of stability by assuming our curves have at most nodes as singularities, all the results described before section 3.4 continue to hold.

The set of isomorphism classes of line bundles on a curve forms a fine moduli space, denoted by and called the Picard scheme of ; see Reference 17, Reference 23. The isomorphism class of a line bundle corresponds to the linear equivalence class of a Cartier divisor. If the line bundle corresponds to the divisor , where are smooth points of and , we write . Also, is a group under the tensor product of line bundles, and it has infinitely many connected, irreducible components. The component containing the trivial line bundle (or the zero divisor) is called the Jacobian of and is denoted by . More precisely, is the moduli space for line bundles of multidegree (i.e., of degree on every irreducible component of ) so that, with the notation of the introduction, the set consists of line bundles of multidegree on , and the equivalence is an isomorphism over . A family over a base , written

is given by a line bundle on whose restriction to every fiber of has multidegree .

If is nonsingular, then the Jacobian is projective and is thus an abelian variety, i.e., a projective algebraic group. Moreover, if has positive genus, encodes important information about itself. For example, fixing a point in we can map every to the line bundle , and this gives an injective morphism, . Our curve can thus be realized as a subvariety of its Jacobian, and in fact it generates the whole of as a group.

On the other hand, if is allowed to have nodes, easily fails to be projective; let us see how. Consider , the dual graph of . As we saw in Equation 3, the desingularization of is the disjoint union of the desingularizations of its components, and we have a birational morphism

Then we have an exact sequence of algebraic groups,

where is simply the pullback of line bundles via the map (and the first Betti number). In this sequence, the product on the right is a product of Jacobians of smooth curves, which are all abelian varieties, hence the product is also an abelian variety. The kernel of accounts for the gluing data needed to specify a line bundle on from a line bundle on . For instance, suppose is irreducible with one node , so that and the kernel of is . A line bundle on determines a line bundle on once the datum of how to identify the fibers of over the two branch points of is given. This datum is an isomorphism between two one-dimensional -vector spaces, hence it corresponds to an element of .

By the exact sequence, is complete if and only if (i.e., the dual graph of , regardless of its weights, is a tree); if that is the case, is said to be of compact type. Now, for any there exist plenty of curves that are not of compact type for example irreducible singular curves.

We are thus in the situation described in the introduction, and we want to construct compactifications for . This problem is classical and presents various aspects. As we said earlier, the noncompleteness of reflects the fact that families of line bundles on degenerate, i.e., do not admit a line bundle on as a limit. On the other hand, if is viewed, as it often happens, as the limit of a family of smooth curves, then families of line bundles on these smooth curves also degenerate. We want a good compactification of to account for both types of degenerations.

3.2. Line bundles on families of curves

As we said, we want to view our curve as a limit of smooth curves, and we do that in the following way. We denote (here and in the rest of this paper) by

a family of curves over the spectrum of a discrete valuation ring with as a special fiber and a smooth generic fiber. A discrete valuation ring can be thought of as a one-dimensional valuation ring. So, in Equation 8 we assume the generic fiber to be a nonsingular curve over the quotient field of . We shall always make the harmless (for us) assumption that has a -rational point.

The Jacobian of is a smooth projective variety over and the moduli space of line bundles of degree on . How are and related? The answer is, roughly speaking, that is a (noncomplete) limit of over . Indeed, together with the existence of the Jacobian for a fixed curve, the general theory gives the existence of a relative Jacobian,

for any family as above. The morphism Equation 9 has as fiber over and as special fiber, and it is thus a so-called model of over . It is smooth and separated, but not projective in general.

Now, as we said, a natural requirement for a good compactification of is to control not only degenerations of line bundles on but also degenerations of line bundles on as degenerates to . Therefore, we want it to appear as a (complete) limit of over , which is why we are now looking at models of over .

Of course, we also want a good compactification of to have a moduli description extending that of . From this perspective there is a fundamental model of over to consider, which is the relative degree Picard scheme,

The generic (resp., special) fiber of the above morphism is the moduli space of line bundles of degree on (resp., on ). The special fiber, written , is quite big if is not irreducible. Indeed we have a decomposition into connected components, , where and is the locus parametrizing line bundles multidegree . Notice that , and we have (noncanonical) isomorphisms ; see section 4.1.

These two models of are related by a natural inclusion , and neither of them is complete, unless is of compact type. But even when is of compact type, they present some problems, described in the next example.

Example 3.2.1.

Let be a family of curves as above, and assume that is nonsingular.

Let the special fiber have the following dual graph (such a family surely exists). So, is of compact type, with one node and two irreducible components of positive genus, and , and it is obtained by gluing the two marked curves and :

We pick a line bundle on of degree ; since is nonsingular, admits an extension to a line bundle on . We choose so that the restriction of to is the line bundle , where and are two smooth points of . Notice that has degree and multidegree . The moduli map associated to

maps the special point to the connected component of . Therefore, although the image of under lies in the Jacobian of , the image of the special point does not lie in . Hence, even if is complete, the relative Jacobian fails to parametrize degenerations of line bundles on .

Now let us show that the morphism Equation 10 is not separated; i.e., loosely speaking, a line bundle on can have more than one limit. To be more precise, let be the moduli map of ; it suffices to exhibit an extension of different from the map defined above. Since is nonsingular, is a Cartier divisor of multidegree , and is a line bundle of degree on the fibers whose restriction to coincides with the restriction of . But the restriction of to satisfies

Hence the restriction of to has multidegree . Therefore maps the special point of to , and it is obviously different from .

The problem described in this example is the existence and uniqueness of an extension for a map of the form , where is a degree line bundle on . For the relative Jacobian the existence of the extension can fail. On the other hand, for the relative Picard scheme the existence holds if is nonsingular but the uniqueness can fail.

3.3. The Néron model of the Jacobian

We shall now describe a third model for for which the problems illustrated in the previous example do not occur. The issue, as we saw, was the existence and uniqueness of an extension for maps of the form . We now approach it using Néron models, and their defining mapping property. The existence of Néron models holds in more general situations than ours; see Reference 24. For us the following special case is enough.

Theorem 3.3.1.

Let be a discrete valuation field. Let be a smooth curve over , and let be its Jacobian. Then there exists a smooth and separated model of , the Néron model

satisfying the following mapping property. For any smooth scheme and any morphism (with the fiber of over ) there exists a unique extension of to a morphism .

We just mention that the Néron model is a group scheme, and it extends the group structure of its generic fiber . Moreover, the Néron model is natural because its mapping property determines it uniquely, but it is known not to commute with ramified base change.

Now, in the setup of Example 3.2.1, we can apply the mapping property of Theorem 3.3.1 to and , obtaining that the map extends uniquely to a map from to .

In general, to give a geometric interpretation to the extension and to the Néron model itself, we assume that the curve in the theorem admits a model with a stable curve as special fiber and nonsingular total space . We denote by the special fiber of the Néron model . We shall see that under these assumptions is a finite disjoint union of components isomorphic to the Jacobian of , and such union is indexed by a combinatorial invariant of . In particular, is independent of the choice of .

We need some combinatorial preliminaries. Fix an orientation (whose choice is irrelevant) on the dual graph . Let and be the usual groups of -chains, so that is the free abelian group on the vertex set and is the free abelian group on the edge set . Consider the boundary homomorphism , mapping an edge oriented from to to . Let be the coboundary, mapping a vertex to where the first sum is over all edges originating from and the second is over all edges ending at . Now we can introduce the finite group

This group is well known in graph theory. By the Kirchhoff–Trent (or Kirchhoff matrix) theorem, its cardinality is equal to the number of spanning trees of (i.e., the connected subgraphs of having the same vertices as and the first Betti number equal to ). We have (see Reference 26, Reference 25, Reference 4) the following.

Proposition 3.3.2.

Let have a stable special fiber and a nonsingular total space . Then

Using compactifications of Néron models and Jacobians, we shall give, in Equation 22, a concrete realization of the above isomorphism.

Example 3.3.3.

By what we said before the statement, the number of irreducible components of equals the number of spanning trees of . In particular, if is of compact type or is irreducible, we have .

By contrast, if is a graph consisting of two vertices joined by edges, we have spanning trees, and one can prove that .

It is not hard to find other graphs with , for example, a cycle with vertices and edges.

There is a clear relation between the Néron model and the relative Picard scheme. Indeed is the maximal separated quotient of the Picard scheme . Moreover, there is an embedding of the relative Jacobian in the Néron model; see Reference 26 and Reference 7.

3.4. Compactified Jacobians and Néron models

A compactified Jacobian for a (stable) curve is a projective variety containing not necessarily as a dense subset (which is why we use the notation rather than ) and satisfying the following requirement. For any family as in Equation 8, there is a projective morphism

having as generic fiber and as special fiber, and such that has a moduli interpretation extending that of .

Going back to the issues illustrated in Example 3.2.1, we observe that since the morphism Equation 12 is projective, any map (associated to a line bundle on ) admits a unique extension to a map .

From now on we apply the notation introduced in Proposition 3.3.2 and, for a connected nodal curve , we denote by the special fiber of the Néron model of the Jacobian associated to a family , with nonsingular.

From Proposition 3.3.2 we see that is not complete unless is complete. We thus introduce a terminology to distinguish compactified Jacobians which also compactify the Néron model. We say that a compactified Jacobian is of Néron type (or a Néron compactified Jacobian) if it contains as a dense open subset.

Compactified Jacobians of Néron type do exist, but we postpone to the next section the discussion on their existence. Assuming it, we describe how these Néron compactified Jacobians have a recursive structure in terms of Néron models.

We introduced in Equation 4 the lattice associated to the graph . We now introduce a subposet of ,

The maximum of is , and the minimal elements are the graphs such that is a spanning tree. Moreover, just as , the poset is graded by the rank function

If is the dual graph of the curve , then for every we have a set of nodes of (and edges of ) such that . We denote by the desingularization of at , so that is a connected nodal curve of genus whose dual graph is .

With our notation, is the special fiber of the Néron model of the Jacobian of a smooth curve specializing to .

The following statement (from Reference 10) describes the compactification of provided by a Néron compactified Jacobian in terms of the Néron models of all the connected partial normalizations of . This is another instance of a widespread recursive phenomenon for compactified moduli spaces. Namely, to compactify a space (e.g., ), one adds at the boundary the analogous spaces associated to simpler objects (e.g., with a connected partial normalization of ).

Theorem 3.4.1.

Let be a stable curve, and let be its dual graph. Then there exists a Néron compactified Jacobian such that

with for every . Moreover, Equation 13 is a graded stratification, i.e., the following hold.

(1)

(2)

is locally closed of pure dimension .

(3)

The map mapping to is a rank on .

Notice the similarities between this theorem and Theorem 2.3.2.

Remark 3.4.2.

The strata of minimal dimension in Equation 13 are Néron models of curves whose dual graph is a spanning tree of , hence they are irreducible and projective. By Proposition 3.3.2, the number of such strata is equal to the number of irreducible components of .

The strata of Equation 13 are, in general, not connected. Hence it is quite natural to ask whether the stratification can be refined so as to have connected strata. We will answer this question in the affirmative after Theorem 4.2.3.

4. Compactifying Jacobians of any degree

4.1. Universal Jacobians of any degree

We now concentrate on compactified Jacobians and, before continuing, we pause for a moment to recall that the Jacobian is the moduli space for line bundles of multidegree on a curve , and we ask, why not extend our consideration to all degrees and multidegrees?

We have mentioned that for any , where as usual, there are isomorphisms . How are they defined? For every vertex , pick a smooth point of lying on the component corresponding to . Then the following is an isomorphism:

which is obviously not canonical, as it depends on the choice of the points . For a smooth curve over a discrete valuation field , we have similar isomorphisms for every .

We can define compactified degree Jacobians, written , as we did in section 3.4 for . So, for any family of curves as in Equation 8, our is the special fiber of a projective morphism

having as generic fiber.

In generalizing our analysis to all degrees, we lose the group structure, but we gain a better understanding on the geometric complexity of the situation.

We shall approach the problem of compactifying the Jacobians of any degree from the point of view of the moduli theory of stable curves, described earlier. First of all, over the moduli space of stable curves, we have a universal curve, denoted by whose fiber over an automorphism-free curve is the curve itself. The requirement that be free from automorphisms is a bit annoying, but it is needed if, as is done in this paper, we work with varieties and schemes rather than stacks. On the other hand, if , the curves in admitting nontrivial automorphisms form a proper closed subset, i.e., the general stable curve has no nontrivial automorphisms. To simplify the forthcoming description, from now on we assume that curve is general in this sense.

Now, as we did for curves over valuation rings, we can consider the relative degree Jacobian associated to the universal curve over . This is often called the universal degree d Jacobian, and it is given by a morphism whose fiber over the point parametrizing the curve is . We point out that, as varies, the varieties are not always isomorphic.

Just as is not complete, is not complete, and we want to construct a compactification of parametrizing compactified degree Jacobians for all stable curves of genus . We refer to such a space as a compactified universal degree Jacobian.

The spaces can be constructed for all by imitating the Geometric Invariant Theory (GIT) construction of . Indeed, was constructed in Reference 16 as the GIT quotient of the Hilbert scheme of -canonically embedded curves (), i.e., curves in projective space embedded by the th power of their dualizing line bundle. In view of the fact that the Hilbert scheme of -canonically embedded curves has such a remarkable GIT quotient, one is led to ask about the Hilbert scheme of all curves of degree and genus in a fixed projective space. Can we construct its GIT quotient? And, if so, is this quotient a good compactification of the universal degree Jacobian?

The answer to these questions is positive, and we have, for every , a projective morphism

whose fiber over the curve is a compactified degree Jacobian, written ; see Reference 9. Now, is a connected, projective variety, all of whose irreducible components have dimension . It is singular in general, and its smooth locus, written , parametrizes line bundles on of suitable multidegree. More exactly, we have an identification between the smooth locus of and a (disjoint) union of components isomorphic to ,

where is a well-defined finite set of multidegrees of total degree . From Proposition 3.3.2 and the discussion in the earlier sections, we expect

with equality for Néron type Jacobians. This expectation does hold, and we shall discuss some examples later,

The points of the boundary, , parametrize equivalence classes of certain line bundles on nodal curves having as stable model (i.e., admitting a genus-preserving birational morphism onto ). As we shall see in the next sections, these line bundles are determined, recursively, by suitable line bundles on the partial normalizations of the given curve .

Going back to the construction of as a GIT quotient, there is a main difference with the GIT quotient defining . Namely, in the latter case the quotient is a so-called geometric quotient, which implies that every point in corresponds to exactly one orbit in the Hilbert scheme. For the quotient is geometric only for certain values of . When is not a geometric quotient, some of its points parametrize more than one orbit and its moduli description becomes more complex.

The degrees for which is a geometric quotient are precisely those satisfying the condition . In these cases, and only in these cases, all fibers of are Néron compactified Jacobians.

Now, the above numerical condition holds if , but it fails if . We shall concentrate on these two cases, which are interesting for different reasons, and give a combinatorial analysis of the compactified Jacobian, highlighting its recursive structure.

4.2. Compactified Jacobians in degree

As we said, the compactified Jacobians are of Néron type for every stable curve . We shall now describe them closely by adapting to the present case a method introduced in Reference 5 to handle the case .

Let be the dual graph of . We know, by the general results mentioned in the previous section, that our compactification has finitely many irreducible components all of dimension . Moreover, by Equation 15, each component contains a dense subset parametrizing line bundles on of a fixed multidegree such that ; we wrote for the set of these special multidegrees. As is of Néron type, we have , and the question is how to interpret the multidegrees lying in in a geometrically meaningful way.

At a very basic level, this amounts to distributing the integer among the vertices of . We start from the identity

The first term on the right, the summation, contains the topological data of each component (or vertex) of (or ), while the second part, , is not related to the vertices. It is thus natural to try and distribute it among the vertices in a combinatorially meaningful way. A natural approach in the graph-theoretic setting is to consider an orientation on and, for any vertex , denote by the number of edges having as target. Now we define a multidegree associated to :

By definition of orientation, every edge has exactly one target, and hence so that , which is off by from what we want (i.e., ). A way to fix this problem is to modify the orientation by allowing one edge to be bi-oriented (i.e., having both ends as targets). So, we define a -orientation on a graph as the datum of a bi-oriented edge and of an orientation on ; see Figure 7 for some examples. Now, if is any 1-orientation, we have Therefore, if we define the multidegree of as in Equation 17, we have for any 1-orientation .

In this way we have selected a special set of multidegrees of total degree , and this set is finite as there are only finitely many 1-orientations on a graph. Now, it may happen that two 1-orientations, and , have the same multidegree. If that is the case, we say that and are equivalent.

We now ask whether we got the correct set of multidegrees . It is immediately clear that the answer is negative, as this set is still too big. Indeed, from our earlier discussion we expect , but the set of equivalence classes of 1-orientations on is bigger than , as Example 4.2.1 shows.

Now, let us observe that the definition of a -orientation depends on the choice of the bi-oriented edge, which is quite arbitrary from our point of view. To eliminate this problem, we ask whether this dependence disappears when passing to equivalence classes of -orientations. The answer is no in general, but it is yes for a special type of -orientations, called rooted orientations.

A -orientation is rooted if, for every edge , the class of contains a representative having as bi-oriented edge.

A more geometric definition is the following. A 1-orientation with bi-oriented edge is rooted if for every vertex there exists a directed path from to . In the next example the orientations from 1 to 4 are rooted, the ones from 5 to 8 are not.

Example 4.2.1.

In Figure 7 we have all the 1-orientations on a 4-cycle with the same bi-oriented edge. One checks easily that they are not equivalent to one another. Since , we see that the number of multidegrees corresponding to 1-orientations is greater than .

We denote by the set of equivalence classes of rooted orientations on . We have two facts; see Reference 14, Reference 11.

Remark 4.2.2.
(a)

.

(b)

is not empty if and only if is connected.

We can go back to our main problem, the description of . Recall that we denote by the poset of connected spanning subgraphs of having legs corresponding to the removed edges. In our case, since has no legs, an element in is a graph of the form having legs. We denoted by the graph obtained by removing all legs from , hence .

We have defined neither orientations nor rooted orientations on a graph with legs. We do it now in the simplest possible way, namely by simply disregarding the legs and using exactly the same terminology. In particular, for any , we have

Now consider the set of all rooted orientations on all (connected) spanning subgraphs of :

The above set admits a natural partial order. To define it recall that for any , we denote by the set of edges such that . We write for an orientation on . Then for two classes of rooted orientations, and , we set if and if the restriction of to is equivalent to .

The forgetful map below is a (surjective) quotient of posets

The following result of Reference 14, using the terminology of Theorem 3.4.1, states that admits a recursive graded stratification governed by rooted orientations.

Theorem 4.2.3.

Let be a stable curve of genus , and let be its dual graph. Then admits the graded stratification

with a natural isomorphism for every .

4.3. Néron compactified Jacobians

Now let us recall that is of Néron type. Hence by Theorem 3.4.1 we have the following stratification of and the associated stratification map ,

such that . We know that is surjective and its fibers are not always connected.

Similarly, Theorem 4.2.3 gives us another stratification map

whose fiber over the class of is the stratum . In this case the stata are connected, so that Equation 20 is a refinement of Equation 21 with connected strata. Combining with the map in Equation 19, we have the commutative diagram

In other words, for any connected spanning subgraph of , we have

which, by Theorems 3.4.1 and 4.2.3, implies that we have an isomorphism

This is an explicit description of the isomorphism Equation 11, within a specific moduli problem.

We now show how the stratification of of Theorem 2.3.2 fits together with the stratifications of Theorems 3.4.1 and 4.2.3. We defined above two stratification maps, and . We can define the analogous stratification map for , using Theorem 2.3.2:

whose fiber over the graph is the stratum .

Now, consider the union of all posets for all stable graphs ,

We have an obvious map

mapping to . Now, has a natural poset structure which restricts to the poset structure of for every and is compatible with the poset structure of . More precisely, the map is a quotient of posets.

We can argue in a similar way for the posets of rooted orientations up to equivalence. Namely, we define

with a natural map

mapping an orientation class to the graph on which the orientation is defined.

It turns out that also has a poset structure extending that of the and is compatible with that of , so that the above map is a quotient of posets; see Reference 11.

Now we need a bit of extra care to describe the global picture over . In fact there are, as we mentioned earlier, technical problems for curves having nontrivial automorphisms. To avoid dealing with them, we here restrict to the locus in of curves free from automorphisms, denoted by ; we refer to Reference 11 for the general case. Recall that if , then is open and dense in . We write for the restriction of over .

In conclusion, we have the following commutative diagram involving the maps we described earlier.

4.4. Compactified Jacobians in degree

The case has been the object of much attention because of its connections with the Theta divisor, the Torelli and the Schottky problems, and the Prym varieties. On the other hand, the compactified Jacobian is never of Néron type, and its moduli properties are not as good.

Let us analyze by the same pattern used for . We need to describe the special set of multidegrees of total degree which correspond to the irreducible components of . The basic identity to look at is, again,

How can we partition among the components of ? As before, if we consider an orientation on and denote by the number of edges ending at the vertex , we have We define the multidegree of exactly as we did in Equation 17, by setting for every . Then, as we already noticed, we have , which now is what we need.

As for -orientations, two orientations with the same multidegree are defined to be equivalent.

How many equivalence classes of orientations do we have on a graph ? One easily sees, as is Example 4.4.2, that there are more than , which is against our expectation Equation 16.

Therefore, to compactify the degree Jacobian, we must select a special type of orientation, or we must exclude some of them. To explain which orientations to exclude, we connect to Example 3.2.1 and argue as in the next example.

Example 4.4.1.

Consider the two following orientations, and , on the same graph.

Suppose . Then and ; hence, and are not equivalent.

Now let be a curve having the above graph as a dual graph, and view it as the special fiber of a family with nonsingular, as in Example 3.2.1. So, has three nodes and two irreducible components, and . Now let be a line bundle on such that its restriction to has the multidegree

(it is not hard to prove such an exists). Now, since is nonsingular, is a Cartier divisor on of multidegree . Hence we can consider the line bundle , whose multidegree on satisfies

As the restrictions of and to coincide, we conclude that and are the multidegrees of two different limits of the same line bundle.

Now, compactified Jacobians are projective and, hence, separated. Therefore the moduli map of and from to must coincide. In fact, the image of the special point of under this moduli map turns out to lie in the boundary of and to parametrizes both and .

The conclusion we want to draw is that the two multidegrees and cannot lie in ; hence, we need to exclude the orientations and . A close look shows that the reason why and must be excluded is that they have a vertex with no incoming edge.

With this example in mind, we define an orientation to be totally cyclic if every set of vertices admits at least one incoming edge, i.e., an edge with target in and source not in . In analogy with rooted orientation, we mention that totally cyclic orientations are characterized by the property that any two vertices in the same connected component lie in a directed cycle. The set of equivalence classes of totally cyclic orientations on is denoted by

Example 4.4.2.

In Figure 8 we have the four orientations on a 2-cycle. The first two are totally cyclic and equivalent. The last two are not totally cyclic and not equivalent. In this case .

Now, as we did for , we want to consider totally cyclic orientations on all subgraphs of . For this reason we need to extend our consideration to disconnected graphs.

Remark 4.4.3.

Let be a (possibly disconnected) graph.

(a)

If is connected, then , and if , .

(b)

is empty if and only if contains some bridge (an edge whose removal disconnects the connected component in which it lies).

Since totally cyclic orientations exist only on bridgeless graphs, we introduce a new sublattice of :

The maximum of is the graph obtained from by removing all bridges, the minimum is . The poset is graded by We write

for the set of all classes of totally cyclic orientations on all spanning subgraphs of . We have a partial order on exactly as for . Now, rephrasing results from Reference 12, we are ready to exhibit a graded stratification of governed by totally cyclic orientations, and we use the same terminology as in Theorem 4.2.3.

Theorem 4.4.4.

Let be a stable curve of genus , and let be its dual graph. Then we have a graded stratification

and a natural isomorphism for every ,

The isomorphism Equation 25 exhibits a recursive behavior which we have already encountered in our earlier statements. Indeed, and is the multidegree associated to a totally cyclic orientation on . Hence the boundary of the compactified degree Jacobian of is stratified by Jacobians of degree of partial normalizations of .

Now, having diagram Equation 23 in mind, we ask whether the stratifications of Theorem 4.4.4 glue together over consistently with the stratification of of Theorem 2.3.2. The answer is positive, and, reasoning as we did for with some obvious modification, we have the commutative diagram below, where the posets and are defined exactly as and (in the previous section). We refer to Reference 11 for details.

From Remark 4.4.3a, it is clear that is not of Néron type unless is irreducible or of compact type. Nevertheless as been used in various applications.

An example of these applications concerns the Theta divisors and the generalized Torelli map. Recall that on a smooth curve a general line bundle of degree has no nontrivial sections, whereas as soon as the degree is at least , Riemann–Roch theory predicts the existence of nontrivial sections for every line bundle of that degree. The subset in of all line bundles admitting some nontrivial global section is a proper closed subset of codimension , called the Theta divisor, denoted by ,

It is well known that is a prime divisor giving a principal polarization on . By using the isomorphisms , we obtain Theta divisors on the Jacobian (or on any other ), but since these isomorphisms are not natural, the Theta divisor of is canonically given only in the case where (in the other cases, its class is canonically given).

A famous theorem involving the Theta divisor is the Torelli theorem, stating that a smooth curve is uniquely determined, up to isomorphism, by the pair given by its Jacobian and its Theta divisor.

The definition of given in Equation 27, with small modifications, makes sense also for our singular curve , and enables us to define the Theta divisor in as the closure of the locus of line bundles on parametrized by admitting nontrivial global sections. Moreover, is a Cartier divisor and a principal polarization, and the pair is endowed with a natural group action of . Using the language of moduli theory for abelian varieties, is a so-called principally polarized stable semi-abelic pair. Such pairs appear as the boundary points in the compactification of the moduli space of principally polarized abelian varieties constructed in Reference 2. Moreover, they form the image of the compactified Torelli morphism

mapping a curve to the pair . By the Torelli theorem, the restriction of to is injective, but is easily seen not to be injective. The combinatorial structure of described in Theorem 4.4.4 is a key tool in Reference 12 to describe the fibers of in detail.

We conclude by observing that, by the moduli properties of , the compactified Torelli map can be viewed as the moduli map associated to the family .

About the author

Lucia Caporaso wrote her thesis in algebraic geometry in 1993 at Harvard University under the supervision of Joe Harris. She held positions at Harvard and the Massachusetts Institute of Technology before moving to Italy in 2001 as professor of mathematics at Roma Tre University.

Figures

Figure 1.

On the left is a (weighted) graph with three legs. On the right is the underlying leg-free graph.

Figure 2.

The three genus , stable graphs with four legs.

Figure 3.

Presentation of and in by

Figure 4.

Presentation of and in by

Figure 5.

The poset , or the dual graph of

Figure 6.

The poset of strata of

Figure 7.

The eight 1-orientations on a 4-cycle

Figure 8.

The four orientations on a 2-cycle

Mathematical Fragments

Example 2.1.1.

The simplest cases are and , which are almost trivial. Indeed is a point since any smooth curve of genus is isomorphic to , and any two ordered triples of points in are mapped to one another by an automorphism.

If and , then has dimension , and, if we work over , it is not hard to see that can be identified with .

Equation (1)
Equation (2)
Equation (3)
Example 2.2.2.

Consider the graph in the picture below, stable of genus with one leg. We marked the identifications on the edges. Now contains the automorphism , which interchanges the two edges. We shall see that acts nontrivially.

In this case we have , and the action of on swaps the two marked points of and the first two marked points of . Indeed, the following two elements in

are conjugated by and give the same point of . In other words, they are different presentations for the same curve.

Equation (4)
Theorem 2.3.2.

The decomposition Equation 1 is a graded stratification of by the poset , i.e., the following properties hold.

(1)

(2)

is irreducible and locally closed.

(3)

The map sending to is a rank function.

Equation (5)
Equation (6)
Equation (8)
Equation (9)
Equation (10)
Example 3.2.1.

Let be a family of curves as above, and assume that is nonsingular.

Let the special fiber have the following dual graph (such a family surely exists). So, is of compact type, with one node and two irreducible components of positive genus, and , and it is obtained by gluing the two marked curves and :

We pick a line bundle on of degree ; since is nonsingular, admits an extension to a line bundle on . We choose so that the restriction of to is the line bundle , where and are two smooth points of . Notice that has degree and multidegree . The moduli map associated to

maps the special point to the connected component of . Therefore, although the image of under lies in the Jacobian of , the image of the special point does not lie in . Hence, even if is complete, the relative Jacobian fails to parametrize degenerations of line bundles on .

Now let us show that the morphism Equation 10 is not separated; i.e., loosely speaking, a line bundle on can have more than one limit. To be more precise, let be the moduli map of ; it suffices to exhibit an extension of different from the map defined above. Since is nonsingular, is a Cartier divisor of multidegree , and is a line bundle of degree on the fibers whose restriction to coincides with the restriction of . But the restriction of to satisfies

Hence the restriction of to has multidegree . Therefore maps the special point of to , and it is obviously different from .

Theorem 3.3.1.

Let be a discrete valuation field. Let be a smooth curve over , and let be its Jacobian. Then there exists a smooth and separated model of , the Néron model

satisfying the following mapping property. For any smooth scheme and any morphism (with the fiber of over ) there exists a unique extension of to a morphism .

Proposition 3.3.2.

Let have a stable special fiber and a nonsingular total space . Then

Equation (12)
Theorem 3.4.1.

Let be a stable curve, and let be its dual graph. Then there exists a Néron compactified Jacobian such that

with for every . Moreover, 13 is a graded stratification, i.e., the following hold.

(1)

(2)

is locally closed of pure dimension .

(3)

The map mapping to is a rank on .

Equation (15)
Equation (16)
Equation (17)
Example 4.2.1.

In Figure 7 we have all the 1-orientations on a 4-cycle with the same bi-oriented edge. One checks easily that they are not equivalent to one another. Since , we see that the number of multidegrees corresponding to 1-orientations is greater than .

Equation (19)
Theorem 4.2.3.

Let be a stable curve of genus , and let be its dual graph. Then admits the graded stratification

with a natural isomorphism for every .

Equation (21)
Equation (22)
Equation (23)
Example 4.4.2.

In Figure 8 we have the four orientations on a 2-cycle. The first two are totally cyclic and equivalent. The last two are not totally cyclic and not equivalent. In this case .

Remark 4.4.3.

Let be a (possibly disconnected) graph.

(a)

If is connected, then , and if , .

(b)

is empty if and only if contains some bridge (an edge whose removal disconnects the connected component in which it lies).

Theorem 4.4.4.

Let be a stable curve of genus , and let be its dual graph. Then we have a graded stratification

and a natural isomorphism for every ,

Equation (27)

References

[1]
Dan Abramovich, Lucia Caporaso, and Sam Payne, The tropicalization of the moduli space of curves (English, with English and French summaries), Ann. Sci. Éc. Norm. Supér. (4) 48 (2015), no. 4, 765–809, DOI 10.24033/asens.2258. MR3377065, Show rawAMSref\bib{ACP}{article}{ author={Abramovich, Dan}, author={Caporaso, Lucia}, author={Payne, Sam}, title={The tropicalization of the moduli space of curves}, language={English, with English and French summaries}, journal={Ann. Sci. \'{E}c. Norm. Sup\'{e}r. (4)}, volume={48}, date={2015}, number={4}, pages={765--809}, issn={0012-9593}, review={\MR {3377065}}, doi={10.24033/asens.2258}, } Close amsref.
[2]
Valery Alexeev, Complete moduli in the presence of semiabelian group action, Ann. of Math. (2) 155 (2002), no. 3, 611–708, DOI 10.2307/3062130. MR1923963, Show rawAMSref\bib{Alexeev}{article}{ author={Alexeev, Valery}, title={Complete moduli in the presence of semiabelian group action}, journal={Ann. of Math. (2)}, volume={155}, date={2002}, number={3}, pages={611--708}, issn={0003-486X}, review={\MR {1923963}}, doi={10.2307/3062130}, } Close amsref.
[3]
Enrico Arbarello, Maurizio Cornalba, and Pillip A. Griffiths, Geometry of algebraic curves. Volume II, Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], vol. 268, Springer, Heidelberg, 2011. With a contribution by Joseph Daniel Harris, DOI 10.1007/978-3-540-69392-5. MR2807457, Show rawAMSref\bib{GAC}{book}{ author={Arbarello, Enrico}, author={Cornalba, Maurizio}, author={Griffiths, Pillip A.}, title={Geometry of algebraic curves. Volume II}, series={Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]}, volume={268}, note={With a contribution by Joseph Daniel Harris}, publisher={Springer, Heidelberg}, date={2011}, pages={xxx+963}, isbn={978-3-540-42688-2}, review={\MR {2807457}}, doi={10.1007/978-3-540-69392-5}, } Close amsref.
[4]
M. Artin, Néron models, Arithmetic geometry (Storrs, Conn., 1984), Springer, New York, 1986, pp. 213–230. MR861977, Show rawAMSref\bib{artin}{article}{ author={Artin, M.}, title={N\'{e}ron models}, conference={ title={Arithmetic geometry}, address={Storrs, Conn.}, date={1984}, }, book={ publisher={Springer, New York}, }, date={1986}, pages={213--230}, review={\MR {861977}}, } Close amsref.
[5]
Arnaud Beauville, Prym varieties and the Schottky problem, Invent. Math. 41 (1977), no. 2, 149–196, DOI 10.1007/BF01418373. MR0572974, Show rawAMSref\bib{beauville}{article}{ author={Beauville, Arnaud}, title={Prym varieties and the Schottky problem}, journal={Invent. Math.}, volume={41}, date={1977}, number={2}, pages={149--196}, issn={0020-9910}, review={\MR {0572974}}, doi={10.1007/BF01418373}, } Close amsref.
[6]
Vladimir G. Berkovich, Spectral theory and analytic geometry over non-Archimedean fields, Mathematical Surveys and Monographs, vol. 33, American Mathematical Society, Providence, RI, 1990. MR1070709, Show rawAMSref\bib{Berkovich}{book}{ author={Berkovich, Vladimir G.}, title={Spectral theory and analytic geometry over non-Archimedean fields}, series={Mathematical Surveys and Monographs}, volume={33}, publisher={American Mathematical Society, Providence, RI}, date={1990}, pages={x+169}, isbn={0-8218-1534-2}, review={\MR {1070709}}, } Close amsref.
[7]
Siegfried Bosch, Werner Lütkebohmert, and Michel Raynaud, Néron models, Ergebnisse der Mathematik und ihrer Grenzgebiete (3) [Results in Mathematics and Related Areas (3)], vol. 21, Springer-Verlag, Berlin, 1990, DOI 10.1007/978-3-642-51438-8. MR1045822, Show rawAMSref\bib{BLR}{book}{ author={Bosch, Siegfried}, author={L\"{u}tkebohmert, Werner}, author={Raynaud, Michel}, title={N\'{e}ron models}, series={Ergebnisse der Mathematik und ihrer Grenzgebiete (3) [Results in Mathematics and Related Areas (3)]}, volume={21}, publisher={Springer-Verlag, Berlin}, date={1990}, pages={x+325}, isbn={3-540-50587-3}, review={\MR {1045822}}, doi={10.1007/978-3-642-51438-8}, } Close amsref.
[8]
Silvia Brannetti, Margarida Melo, and Filippo Viviani, On the tropical Torelli map, Adv. Math. 226 (2011), no. 3, 2546–2586, DOI 10.1016/j.aim.2010.09.011. MR2739784, Show rawAMSref\bib{BMV}{article}{ author={Brannetti, Silvia}, author={Melo, Margarida}, author={Viviani, Filippo}, title={On the tropical Torelli map}, journal={Adv. Math.}, volume={226}, date={2011}, number={3}, pages={2546--2586}, issn={0001-8708}, review={\MR {2739784}}, doi={10.1016/j.aim.2010.09.011}, } Close amsref.
[9]
Lucia Caporaso, A compactification of the universal Picard variety over the moduli space of stable curves, J. Amer. Math. Soc. 7 (1994), no. 3, 589–660, DOI 10.2307/2152786. MR1254134, Show rawAMSref\bib{Cthesis}{article}{ author={Caporaso, Lucia}, title={A compactification of the universal Picard variety over the moduli space of stable curves}, journal={J. Amer. Math. Soc.}, volume={7}, date={1994}, number={3}, pages={589--660}, issn={0894-0347}, review={\MR {1254134}}, doi={10.2307/2152786}, } Close amsref.
[10]
Lucia Caporaso, Néron models and compactified Picard schemes over the moduli stack of stable curves, Amer. J. Math. 130 (2008), no. 1, 1–47, DOI 10.1353/ajm.2008.0000. MR2382140, Show rawAMSref\bib{Cneron}{article}{ author={Caporaso, Lucia}, title={N\'{e}ron models and compactified Picard schemes over the moduli stack of stable curves}, journal={Amer. J. Math.}, volume={130}, date={2008}, number={1}, pages={1--47}, issn={0002-9327}, review={\MR {2382140}}, doi={10.1353/ajm.2008.0000}, } Close amsref.
[11]
Lucia Caporaso and Karl Christ, Combinatorics of compactified universal Jacobians. arXiv:1801.04098 (2018).
[12]
Lucia Caporaso and Filippo Viviani, Torelli theorem for stable curves, J. Eur. Math. Soc. (JEMS) 13 (2011), no. 5, 1289–1329, DOI 10.4171/JEMS/281. MR2825165, Show rawAMSref\bib{CV2}{article}{ author={Caporaso, Lucia}, author={Viviani, Filippo}, title={Torelli theorem for stable curves}, journal={J. Eur. Math. Soc. (JEMS)}, volume={13}, date={2011}, number={5}, pages={1289--1329}, issn={1435-9855}, review={\MR {2825165}}, doi={10.4171/JEMS/281}, } Close amsref.
[13]
Renzo Cavalieri, Hannah Markwig, and Dhruv Ranganathan, Tropicalizing the space of admissible covers, Math. Ann. 364 (2016), no. 3-4, 1275–1313, DOI 10.1007/s00208-015-1250-8. MR3466867, Show rawAMSref\bib{CMR}{article}{ author={Cavalieri, Renzo}, author={Markwig, Hannah}, author={Ranganathan, Dhruv}, title={Tropicalizing the space of admissible covers}, journal={Math. Ann.}, volume={364}, date={2016}, number={3-4}, pages={1275--1313}, issn={0025-5831}, review={\MR {3466867}}, doi={10.1007/s00208-015-1250-8}, } Close amsref.
[14]
Karl Christ, Orientations, break Divisors and compactified Jacobians, PhD thesis in Mathematics, Roma Tre University (2017).
[15]
P. Deligne and D. Mumford, The irreducibility of the space of curves of given genus, Inst. Hautes Études Sci. Publ. Math. 36 (1969), 75–109. MR0262240, Show rawAMSref\bib{DM}{article}{ author={Deligne, P.}, author={Mumford, D.}, title={The irreducibility of the space of curves of given genus}, journal={Inst. Hautes \'{E}tudes Sci. Publ. Math.}, number={36}, date={1969}, pages={75--109}, issn={0073-8301}, review={\MR {0262240}}, } Close amsref.
[16]
D. Gieseker, Lectures on moduli of curves, Tata Institute of Fundamental Research Lectures on Mathematics and Physics, vol. 69, Published for the Tata Institute of Fundamental Research, Bombay; Springer-Verlag, Berlin-New York, 1982. MR691308, Show rawAMSref\bib{gieseker}{book}{ author={Gieseker, D.}, title={Lectures on moduli of curves}, series={Tata Institute of Fundamental Research Lectures on Mathematics and Physics}, volume={69}, publisher={Published for the Tata Institute of Fundamental Research, Bombay; Springer-Verlag, Berlin-New York}, date={1982}, pages={iii+99}, isbn={3-540-11953-1}, review={\MR {691308}}, } Close amsref.
[17]
Alexander Grothendieck, Technique de descente et théorèmes d’existence en géométrie algébrique. V. Les schémas de Picard: théorèmes d’existence (French), Séminaire Bourbaki, Vol. 7, Soc. Math. France, Paris, 1995, pp. Exp. No. 232, 143–161. MR1611170, Show rawAMSref\bib{GrothendieckP}{article}{ author={Grothendieck, Alexander}, title={Technique de descente et th\'{e}or\`emes d'existence en g\'{e}om\'{e}trie alg\'{e}brique. V. Les sch\'{e}mas de Picard: th\'{e}or\`emes d'existence}, language={French}, conference={ title={S\'{e}minaire Bourbaki, Vol. 7}, }, book={ publisher={Soc. Math. France, Paris}, }, date={1995}, pages={Exp. No. 232, 143--161}, review={\MR {1611170}}, } Close amsref.
[18]
Alexander Grothendieck, Techniques de construction et théorèmes d’existence en géométrie algébrique. IV. Les schémas de Hilbert (French), Séminaire Bourbaki, Vol. 6, Soc. Math. France, Paris, 1995, pp. Exp. No. 221, 249–276. MR1611822, Show rawAMSref\bib{GrothendieckH}{article}{ author={Grothendieck, Alexander}, title={Techniques de construction et th\'{e}or\`emes d'existence en g\'{e}om\'{e}trie alg\'{e}brique. IV. Les sch\'{e}mas de Hilbert}, language={French}, conference={ title={S\'{e}minaire Bourbaki, Vol. 6}, }, book={ publisher={Soc. Math. France, Paris}, }, date={1995}, pages={Exp. No. 221, 249--276}, review={\MR {1611822}}, } Close amsref.
[19]
Joe Harris and Ian Morrison, Moduli of curves, Graduate Texts in Mathematics, vol. 187, Springer-Verlag, New York, 1998. MR1631825, Show rawAMSref\bib{HM}{book}{ author={Harris, Joe}, author={Morrison, Ian}, title={Moduli of curves}, series={Graduate Texts in Mathematics}, volume={187}, publisher={Springer-Verlag, New York}, date={1998}, pages={xiv+366}, isbn={0-387-98438-0}, isbn={0-387-98429-1}, review={\MR {1631825}}, } Close amsref.
[20]
Joe Harris and David Mumford, On the Kodaira dimension of the moduli space of curves, Invent. Math. 67 (1982), no. 1, 23–88, DOI 10.1007/BF01393371. With an appendix by William Fulton. MR664324, Show rawAMSref\bib{HarMum}{article}{ author={Harris, Joe}, author={Mumford, David}, title={On the Kodaira dimension of the moduli space of curves}, note={With an appendix by William Fulton}, journal={Invent. Math.}, volume={67}, date={1982}, number={1}, pages={23--88}, issn={0020-9910}, review={\MR {664324}}, doi={10.1007/BF01393371}, } Close amsref.
[21]
Finn F. Knudsen, The projectivity of the moduli space of stable curves. II. The stacks , Math. Scand. 52 (1983), no. 2, 161–199, DOI 10.7146/math.scand.a-12001. MR702953, Show rawAMSref\bib{Knudsen}{article}{ author={Knudsen, Finn F.}, title={The projectivity of the moduli space of stable curves. II. The stacks $M_{g,n}$}, journal={Math. Scand.}, volume={52}, date={1983}, number={2}, pages={161--199}, issn={0025-5521}, review={\MR {702953}}, doi={10.7146/math.scand.a-12001}, } Close amsref.
[22]
Grigory Mikhalkin, Moduli spaces of rational tropical curves, Proceedings of Gökova Geometry-Topology Conference 2006, Gökova Geometry/Topology Conference (GGT), Gökova, 2007, pp. 39–51. MR2404949, Show rawAMSref\bib{MIK5}{article}{ author={Mikhalkin, Grigory}, title={Moduli spaces of rational tropical curves}, conference={ title={Proceedings of G\"{o}kova Geometry-Topology Conference 2006}, }, book={ publisher={G\"{o}kova Geometry/Topology Conference (GGT), G\"{o}kova}, }, date={2007}, pages={39--51}, review={\MR {2404949}}, } Close amsref.
[23]
D. Mumford, J. Fogarty, and F. Kirwan, Geometric invariant theory, 3rd ed., Ergebnisse der Mathematik und ihrer Grenzgebiete (2) [Results in Mathematics and Related Areas (2)], vol. 34, Springer-Verlag, Berlin, 1994, DOI 10.1007/978-3-642-57916-5. MR1304906, Show rawAMSref\bib{GIT}{book}{ author={Mumford, D.}, author={Fogarty, J.}, author={Kirwan, F.}, title={Geometric invariant theory}, series={Ergebnisse der Mathematik und ihrer Grenzgebiete (2) [Results in Mathematics and Related Areas (2)]}, volume={34}, edition={3}, publisher={Springer-Verlag, Berlin}, date={1994}, pages={xiv+292}, isbn={3-540-56963-4}, review={\MR {1304906}}, doi={10.1007/978-3-642-57916-5}, } Close amsref.
[24]
André Néron, Modèles minimaux des variétés abéliennes sur les corps locaux et globaux (French), Inst. Hautes Études Sci. Publ.Math. No. 21 (1964), 128. MR0179172, Show rawAMSref\bib{Neron}{article}{ author={N\'{e}ron, Andr\'{e}}, title={Mod\`eles minimaux des vari\'{e}t\'{e}s ab\'{e}liennes sur les corps locaux et globaux}, language={French}, journal={Inst. Hautes \'{E}tudes Sci. Publ.Math. No.}, volume={21}, date={1964}, pages={128}, issn={0073-8301}, review={\MR {0179172}}, } Close amsref.
[25]
Tadao Oda and C. S. Seshadri, Compactifications of the generalized Jacobian variety, Trans. Amer. Math. Soc. 253 (1979), 1–90, DOI 10.2307/1998186. MR536936, Show rawAMSref\bib{os}{article}{ author={Oda, Tadao}, author={Seshadri, C. S.}, title={Compactifications of the generalized Jacobian variety}, journal={Trans. Amer. Math. Soc.}, volume={253}, date={1979}, pages={1--90}, issn={0002-9947}, review={\MR {536936}}, doi={10.2307/1998186}, } Close amsref.
[26]
M. Raynaud, Spécialisation du foncteur de Picard (French), Inst. Hautes Études Sci. Publ. Math. 38 (1970), 27–76. MR0282993, Show rawAMSref\bib{Raynaud}{article}{ author={Raynaud, M.}, title={Sp\'{e}cialisation du foncteur de Picard}, language={French}, journal={Inst. Hautes \'{E}tudes Sci. Publ. Math.}, number={38}, date={1970}, pages={27--76}, issn={0073-8301}, review={\MR {0282993}}, } Close amsref.
[27]
Amaury Thuillier, Géométrie toroïdale et géométrie analytique non archimédienne. Application au type d’homotopie de certains schémas formels (French, with English summary), Manuscripta Math. 123 (2007), no. 4, 381–451, DOI 10.1007/s00229-007-0094-2. MR2320738, Show rawAMSref\bib{thuillier}{article}{ author={Thuillier, Amaury}, title={G\'{e}om\'{e}trie toro\"{i}dale et g\'{e}om\'{e}trie analytique non archim\'{e}dienne. Application au type d'homotopie de certains sch\'{e}mas formels}, language={French, with English summary}, journal={Manuscripta Math.}, volume={123}, date={2007}, number={4}, pages={381--451}, issn={0025-2611}, review={\MR {2320738}}, doi={10.1007/s00229-007-0094-2}, } Close amsref.
[28]
Eckart Viehweg, Quasi-projective moduli for polarized manifolds, Ergebnisse der Mathematik und ihrer Grenzgebiete (3) [Results in Mathematics and Related Areas (3)], vol. 30, Springer-Verlag, Berlin, 1995, DOI 10.1007/978-3-642-79745-3. MR1368632, Show rawAMSref\bib{Viehweg}{book}{ author={Viehweg, Eckart}, title={Quasi-projective moduli for polarized manifolds}, series={Ergebnisse der Mathematik und ihrer Grenzgebiete (3) [Results in Mathematics and Related Areas (3)]}, volume={30}, publisher={Springer-Verlag, Berlin}, date={1995}, pages={viii+320}, isbn={3-540-59255-5}, review={\MR {1368632}}, doi={10.1007/978-3-642-79745-3}, } Close amsref.
[29]
Filippo Viviani, Tropicalizing vs. compactifying the Torelli morphism, Tropical and non-Archimedean geometry, Contemp. Math., vol. 605, Amer. Math. Soc., Providence, RI, 2013, pp. 181–210, DOI 10.1090/conm/605/12116. MR3204272, Show rawAMSref\bib{viviani}{article}{ author={Viviani, Filippo}, title={Tropicalizing vs. compactifying the Torelli morphism}, conference={ title={Tropical and non-Archimedean geometry}, }, book={ series={Contemp. Math.}, volume={605}, publisher={Amer. Math. Soc., Providence, RI}, }, date={2013}, pages={181--210}, review={\MR {3204272}}, doi={10.1090/conm/605/12116}, } Close amsref.

Article Information

MSC 2010
Primary: 14H10 (Families, moduli), 14D06 (Fibrations, degenerations), 14D20 (Algebraic moduli problems, moduli of vector bundles)
Author Information
Lucia Caporaso
Dipartimento di Matematica e Fisica, Università Roma Tre, Largo San Leonardo Murialdo, I-00146 Roma, Italy
caporaso@mat.uniroma3.it
MathSciNet
Journal Information
Bulletin of the American Mathematical Society, Volume 57, Issue 3, ISSN 1088-9485, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on and published on .
Copyright Information
Copyright 2018 American Mathematical Society
Article References

Settings

Change font size
Resize article panel
Enable equation enrichment

Note. To explore an equation, focus it (e.g., by clicking on it) and use the arrow keys to navigate its structure. Screenreader users should be advised that enabling speech synthesis will lead to duplicate aural rendering.

For more information please visit the AMS MathViewer documentation.