Localization for logarithmic stable maps

By S. Molcho and E. Routis

Abstract

We prove a virtual localization formula for Bumsig Kim’s space of Logarithmic Stable Maps. The formula is closely related and can in fact recover the relative virtual localization formula of Graber-Vakil.

1. Introduction and background

In his papers Reference Li01 and Reference Li02, Jun Li introduced and studied the space of relative stable maps. We recall the setup: fix a pair of a smooth variety with a smooth divisor and discrete data , consisting of the arithmetic genus of a nodal curve, a vector of integers, and a homology class in . We wish to parametrize stable maps from a genus nodal curve with two sets of marked points and into the variety , whose image lies in the given homology class and with prescribed incidence conditions along the divisor, namely, .

The moduli space parametrizing is not proper: a limit of such maps may fail to exist, as in the limit that the whole curve may lie entirely in the divisor . Jun Li and, at about the same time, Li-Ruan Reference LR01 and Ionel-Parker Reference IP03,Reference IP04 from the point of view of symplectic Gromov-Witten theory, gave a beautiful solution to this issue. Jun Li’s idea can be outlined as follows. When a limit of maps tends to collapse into the divisor, the space sprouts a new component, which is isomorphic to the projective completion of the normal bundle of to contain the image, in a manner similar to a blowup. We then require that the prescribed behavior along the divisor does not happen along the original divisor , but rather the divisor at infinity in , which we denote . We call with this new component ; we then have a new pair and we may consider stable maps as above to this pair. When a family of maps to tends to collapse into , the variety sprouts a new component that replaces , as above, to create a new space with a divisor at infinity, and so forth. In general, a pair is constructed from the pair inductively. It is called the -th expansion of . Li’s moduli space parametrizes stable maps whose target is allowed to be any of the expansions above, with prescribed behavior along the divisor at infinity and with a certain compatibility requirement along the divisor , : only nodes of the source curve can map to , and when a node maps to , the two components of the curve containing the node have the same contact order with the divisor ; this is called the predeformability or “kissing” condition. This space is proper and is shown to carry a virtual fundamental class, so one can define in a standard manner a type of Gromov-Witten invariant for , called relative Gromov-Witten invariants. For details of the construction, the reader should consult Jun Li’s original paper Reference Li01.

Jun Li also considers a variant of this situation, where instead of a pair we consider a semistable nodal variety of the form . This means that is the union of two smooth varieties along a common smooth divisor in both of them that satisfies the following technical condition on the normal bundles: . Stable maps into must satisfy a similar predeformability condition as above, and the space is compactified by allowing the targets to vary as before. may deform to a target , where is replaced by , with glued along the section and along the infinity section, may deform to where the divisor connection with is replaced by another copy of , and so forth. The spaces are called the expanded degenerations of . The space of expanded degenerations also carries a virtual fundamental class, and one is thus able to extend the notion of Gromov-Witten invariants for targets , which are mildly singular. These are the correct Gromov-Witten invariants, in the sense that they satisfy deformation invarance: If is a family with smooth total space, smooth general fiber, and central fiber , the Gromov-Witten invariants of as defined by Jun Li coincide with the usual Gromov-Witten invariants of the general fiber, at least when such a comparison makes sense, i.e., for homology classes restricted from .

The relative Gromov-Witten invariants are related to the singular Gromov-Witten invariants by the degeneration formula. This was also proven by Jun Li and had also been previously considered in the symplectic category in the work of Li–Ruan Reference LR01 and Ionel–Parker Reference IP04. The degeneration formula allows one to compute Gromov-Witten invariants of expanded degenerations from the relative ones and the combinatorics of the expansions. This can be useful because it is often possible to degenerate a smooth variety into a semistable one with very simple components . Thus one can calculate Gromov-Witten invariants from relative Gromov-Witten invariants of simpler targets.

Computations of relative Gromov-Witten invariants can still be difficult, as calculations in Gromov-Witten theory often are, even if the targets are very simple. These calculations can be greatly facilitated by the use of Atiyah-Bott localization. Localization formulas for the spaces were established by Graber-Vakil in Reference GV05. The applications of such formulas are far reaching: for example, in Reference GV05, as applications of the formulas the authors recover the ELSV formula and certain striking results about the tautological ring.

Jun Li’s constructions are beautiful and geometrically transparent, but suffer from one technical drawback. The virtual fundamental classes defined are hard to work with. The reason for this is that the space of relative stable maps is not an open subset of all maps, but rather, it is locally closed. The perfect obstruction theory used to define the virtual fundamental class is thus constructed by hand and not by standard machinery. This is the main reason the paper Reference GV05 is technically difficult.

One way to avoid this issue is to use a different compactification of the space of maps to the pair or , by endowing the sources and targets of all maps with logaritmic structures and requiring that the maps between them are log maps. We will explain this more precisely in what follows, but here we would like to remark that this idea agrees with a general philosophy in the modern theory of moduli that states that instead of compactifying a moduli space of certain objects, one may try to build the moduli space of such objects with logarithmic structures. Since logarithmic structures allow mild singularities, this moduli space is often already proper. The space of logarithmic stable maps was constructed by B. Kim in his paper Reference Kim10. Kim’s space is shown to be an open substack of the space of all logarithmic maps and thus carries a natural virtual fundamental class by restriction, which is simpler than the virtual fundamental class of : its formal properties are almost identical to the fundamental class in the classical Gromov-Witten theory of smooth targets. The situation may be summarized pictorially as follows:

In this paper, we derive an analogous localization formula for the space of logarithmic stable maps. The formula is analogous to the formula of Reference GV05, but its derivation is closer in spirit with the proofs of localization in classical Gromov-Witten theory, as in Reference Kon95,Reference GP99. Specifically, we show the following.

Theorem 3.1.

is a global quotient stack and admits a equivariant immersion to a smooth Deligne-Mumford stack.

This in particular shows that admits a localization formula. Then, following the work of Graber-Vakil Reference GV05, we obtain explicitly that for suitable splittings of the discrete data into subsets , we have the following.

Theorem 5.1 (Log Virtual Localization).

The formula is essentially the same as the relative virtual localization formula of Reference GV05. The difference is that the stacks of simple relative maps and unrigidified relative stable maps of Reference GV05 are replaced by their logarithmic analogues. These are defined carefully in section 5. In fact, the log virtual localization formula can be used to recover the formula of Reference GV05; this is our Corollary 5.4.

2. Logarithmic stable maps

For completeness, we will recall here the necessary definitions and constructions that we will use. For proofs and more detailed explanations the reader should consult Kim’s paper Reference Kim10.

A family of -marked prestable curves, , carries a canonical structure of a logarithmic map, as shown in F. Kato’s paper Reference Kat00 (see also Reference Ols07). The log structures and morphisms are defined as follows. The curve corresponds to a diagram

where and is the moduli stack of -marked prestable curves and its universal family respectively. Both stacks carry natural logarithmic structures: in the log structure is given by the divisor corresponding to singular curves, and in the log structure is given by the divisors corresponding to singular curves and the markings. The log structures on and are the ones pulled back from and respectively; we denote the log structure on by and on by and refer to them as the canonical log structures. The morphism is automatically a log morphism. An explicit description of the log structures when is a geometric point can be given in terms of charts as follows: has a chart isomorphic to , ; at smooth points; at a marked point; and at a node is given by the following pushout diagram:

where the horizontal map is the diagonal and the vertical map is the inclusion corresponding to the appropriate node.

Definition 2.1.

A genus , -marked log curve is a morphism of log schemes such that is a family of genus , -marked prestable curves and the morphism is obtained from a cartesian diagram of the form

where the horizontal maps are the identities on underlying schemes. Therefore, a log curve is the same thing as the choice of a prestable curve and the choice of a homomorphism of log structures . We will reserve the notation to always refer to this homomorphism and denote it by to simplify notation if no confusion may arise.

We denote the stack parametrizing log curves by .

Definition 2.2.

A log curve is called minimal if the log structure is locally free and there is no locally free submonoid that contains the image of .

Here, we call a log structure locally free if around every point it has a chart isomorphic to for some , possibly depending on the point. For example, over , where , all surjections , , give minimal log curves but no map , .

Minimal log curves are essentially the sources of log stable maps; for a more precise statement, see Definition 2.5. Next, we discuss the possible targets, which Kim calls “extended log twisted Fulton-Macpherson type spaces”.

Fix a smooth projective variety .

Definition 2.3.

A family of schemes or algebraic spaces is called a log FM type space of if, at every point , there is étale locally an étale map

These families of spaces are required to admit canonical log structures on and on such that is given by the cocartesian diagram

and such that the morphism is in fact a log morphism . We will further require that be locally free. Its rank at equals the number of irreducible components of the singular locus of the fiber . We further require that the spaces come equipped with a map .

Definition 2.4.

An extended log twisted FM type space of is a log morphism , where

is as in Definition 2.3 above, and all relevant logarithmic data are obtained from a cartesian diagram

that is, the logarithmic data simply corresponds to a morphism of log structures . We will reserve the notation to always indicate this morphism and denote it by to simplify notation when no confusion may arise.

There is a chart for the morphism of the form

Here the top map is of the form , where is a diagonal matrix of natural numbers.

From now on we will refer to log FM type spaces of X simply as log FM spaces and to extended log twisted FM type spaces of X as extended log twisted FM spaces for brevity. Having defined both the sources and the targets of log stable maps, we can now give the definition of the central objects of study of this paper.

Definition 2.5.

A log stable map from a log curve to an extended log twisted FM space is a log morphism over such that, over each point :

The cokernel of the map has rank equal to the number of non-distinguished nodes.

The map is simple at the distinguished nodes.

Stability: The automorphism group is finite.

The following minimality condition holds: Either is a minimal log curve or, if not, then there exists a subsheaf of monoids of , which is a locally free log structure on , such that , minimal, and surjective.

We explain the terminology: Over each , is a nodal curve and is an extended log twisted FM space. A node of is called distinguished if it maps into the singular locus of and is non-distinguished otherwise. A morphism between locally free log structures is called simple if it is given by a diagonal matrix, as in the definition of extended log twisted FM spaces above. An automorphism of over is a cartesian diagram over

that respects the map to ; that is, on the level of underlying schemes we have

Remark 2.1.

The minimality condition slightly deviates from Kim’s definition. In Reference Kim10, the definition of a log stable map requires that the log curve be minimal. This neglects the possibility that there is no node in the curve mapping to the original divisor . The typical example of this situation is in , where a line rotates to collapse into : the limit map is the map from to that sends all of into , with the image line intersecting transversely. There is a more satisfying, intrinsic explanation of the minimality condition. Minimality is a categorical property: Minimal log schemes are precisely the log schemes one must restrict to in order to be able to consider a stack over log schemes, such as or the stack of all log maps into extended log twisted FM spaces, as a log stack. This is the content of the paper Reference Gil11. Kim’s minimal log curves are precisely the minimal objects for the stack . However, the minimal objects for the stack of all log maps into extended log twisted FM spaces that satisfy the first three properties include the case that no node of the curve maps into , and we must include this case in the definition. Here we also remark that what we call minimal here is the same thing as what is called basic in Reference Che14,Reference AC11,Reference GS11. Both notions correspond to the same categorical notion of minimality.

Let us describe the logarithmic data concretely in the case when is a geometric point. This description will be useful in what follows. The locally free log structures on are free in this case, described by a chart for certain integers ; specifically, we have

The morphisms are described on the level of characteristic monoids as follows: is given by a diagonal matrix of the form , as above. The morphism is given by a matrix of the form , where is a generalized diagonal matrix:

The integers , …, add up to . It is possible that the first row of the matrix is , in which case . This happens when no node of the curve maps to the original divisor . Otherwise, the log curve is minimal, which means that there is no common divisor between the integers , …, . For each , there is an integer such that . In other words, there is a commutative diagram

We will now fix a stack of certain log FM spaces and denote by the stack whose objects are extended log twisted FM spaces whose underlying spaces are in . We denote by and the universal family of and respectively. In other words, we consider spaces but endow them with log structures as above. We will consider the stack of log stable maps to targets in . It is proven Reference Kim10 that if the stack is algebraic, is also algebraic.

Remark 2.2.

Let us at this point explain the connection with Jun Li’s original definitions and clarify this concept geometrically. A family of expansions of a pair or, similarly, of a -semistable degeneration has canonical log structures that determine log FM spaces. The canonical log structure on a family of expansions is obtained in a manner formally identical to the way that the canonical log structure on a nodal curve is obtained. A detailed treatment of the canonical log structures on expansions can be found in Olsson’s paper Reference Ols03. Briefly, there is an algebraic stack parametrizing expansions. The stack is in fact the open substack of the stack of marked, genus prestable curves where the first two markings are on the first component of the curve and the third marking on the last; see for instance Reference GV05. In there is a normal crossings divisor corresponding to singular expansions; therefore, admits a log structure . Similarly, the universal family over admits a log structure . A family of expansions corresponds to a cartesian diagram

The pullback log structures and on and are what we denoted by and above. Therefore, expansions are examples of log FM spaces. We may thus consider log stable maps to expansions. The underlying morphism of schemes of such a log stable map is a relative stable map in the sense of Jun Li; the predeformability condition is enforced by the requirement that the map is a map of log schemes. The log structures are thus additional algebraic data on a relative stable map. The log structures encode essential geometric information very conveniently. Suppose for simplicity that is a geometric point. We have seen above the form of the log structures , and the maps between them. The rank of , which we denoted by the number above, indicates that the target is the -th expansion of . The number is the number of non-distinguished nodes. The number is the number of distinguished nodes. The matrix above indicates that of the distinguished nodes map to the first singular locus in (namely to , map to , and so forth. The contact order of the -th node mapping to the -th singular locus is . Note that once the underlying stable map is fixed, the diagram

between the characteristic monoids of the log structures is determined. This means that in order to determine the full diagram

we need to determine the elements of to which the generators of and are mapping. In fact, all generators of may be chosen to map into after automorphism, so it is enough to treat only . Matrix Equation 1 in Remark 2.1 indicates that has the following form:

Note however that the units are restricted: they must satisfy the equation , in order for the diagram

to commute. This shows that there is a finite number of ways to give to a relative stable map the structure of a log stable map. In other words, if denotes Jun Li’s space of expansions and Kim’s space of log stable maps to expansions of , which is algebraic stack since the stack of expansions is algebraic, there is a forgetful morphism

For the rigorous definition of the morphism we refer to the work of Gross and Siebert Reference GS11 and the paper Reference AMW12 of Abramovich, Marcus, and Wise. This is the left vertical arrow of diagram Equation * of the introduction. The fact that relative stable maps are a locally closed substack of the stack of all maps expresses the fact that the predeformability condition is locally closed. The fact that the stack of log stable maps is open in the stack of all log maps expresses the fact that predeformability is enforced by requiring that the map from a nodal curve to an expansion be a log map.

3. Equivariant embedding

Since Kim’s moduli space does not carry a fundamental class but rather a virtual fundamental class, in order to prove a localization formula we need to use the virtual localization formula of Graber-Pandharipande Reference GP99. Traditionally, to use the results of Reference GP99, one needs to establish the following technical condition.

Theorem 3.1.

There is a locally closed equivariant immersion of into a smooth Deligne-Mumford stack.

In fact, after the results of Chang-Kiem-Li in Reference CKL17, this is no longer necessary. We nevertheless prove the theorem, as we believe that the result and its proof are of independent interest. We will do this by proving that satisfies a slightly stronger condition, also shared by Jun Li’s space . Namely we will prove that satisfies the following property, which we will abbreviate as property SE (for strong embedding property):

is a global quotient, where is a reductive group and is a locally closed subset of a smooth projective with an action of extending that of .

There is a action on which preserves V and descends to the action on .

In Reference GV05, it is shown that satisfies SE by an explicit construction. We have seen there is a morphism ; this morphism is in fact finite, as shown in Lemma 3.2 below. Therefore, it suffices to show the following lemma; the idea of the proof is due to Vistoli.

Lemma 3.1.

Suppose is a equivariant finite morphism between Deligne-Mumford stacks, and assume that satisfies SE. Then satisfies SE as well and thus embeds equivariantly into a smooth Deligne-Mumford stack.

Proof.

Since satisfies SE, we may write with a equivariant locally closed subset and a smooth projective variety. We can find a equivariant open smooth subvariety such that:

(1)

acts on with reduced finite stabilizers,

(2)

is a closed subvariety of .

(Note that this is clearly possible since is Deligne-Mumford by the hypothesis, or, equivalently, acts on with finite reduced stabilizers). The composed morphism is then still equivariant and finite, so by replacing with we may assume is smooth. Since the morphism is equivariant, we obtain a morphism . Here denotes the stack quotient in the sense of Romagny Reference Rom05, though our notation is slightly different. Assume for a moment that has the resolution property (Reference Tot04), i.e., that any coherent sheaf on admits a surjective morphism from a locally free sheaf on . Then, the sheaf , which is coherent since is finite, is the quotient of a locally free sheaf on . In other words, embeds into a vector bundle over . That is, there is a equivariant morphism of into a equivariant vector bundle over , hence a equivariant embedding of into a smooth DM stack. In particular, admits a equivariant embedding into a stack of the form , where is a smooth vector bundle over , so it is, in fact, a quotient. To see that satisfies SE, then, the only thing that remains to be shown is that embeds equivariantly into a smooth and projective variety as a locally closed subset. We give the proof after proving the resolution property for , as it requires essentially the same argument.

We thus show that = does, indeed, have the resolution property. Equivalently, we prove that any equivariant coherent sheaf on admits a equivariant surjection from a equivariant locally free sheaf.

Let be the projection. Since we assumed above that is smooth and quasiprojective, by Reference MFK94, Corollary 1.6, we can find a -equivariant immersion . Let and consider a equivariant coherent sheaf on . Since the action on descends to the action on the quotient , the pullback is equivariant. It is also coherent. Therefore, since is quasiprojective, we may pick a large so that the twisted sheaf is generated by global sections, say , , …, . Let be their linear span in . Following the argument in the proof of Reference MFK94, pp. 25-26, Lemma , we can find

where is a equivariant finitely generated subspace. We therefore obtain a natural equivariant surjection of sheaves on

hence a natural equivariant surjection

which descends to a equivariant surjection from a equivariant locally free sheaf on to . This shows that has the resolution property. Note that this argument also suffices to complete the proof that satisfies SE: we have already seen that embeds equivariantly into a vector bundle , where U is a vector bundle over V; let for some locally free sheaf on . Then the argument just given shows that admits a equivariant surjection for some integers , . Therefore, admits a equivariant locally closed immersion to the quotient of the projectivized bundle over (i.e., a smooth projective variety) by .

Lemma 3.2.

The morphism is finite.

Proof.

We need to show that every geometric point of , that is, every relative stable map , has a finite number of preimages and that the morphism is representable: for each preimage , the map is injective. The discussion in section 2, or Remark 6.3.1 in Reference Kim10, establishes the finiteness of the preimages. Then let be a log stable map lying over . Denote the base by , where , the source curve by , and the target by . Let be the number of distinguished and non-distinguished nodes of respectively, and let be the number of nodes of . We have , , and . An automorphism of lying over the identity automorphism of is simply an automorphism of the logarithmic structures, that is, an automorphism of that respects the maps , . In other words, we are looking for commutative diagrams

The induced matrix on the level of characteristic monoids is a generalized diagonal matrix which is the identity on the last components, and similarly is a diagonal matrix with finite cokernel on the first components. Therefore must be a diagonal matrix as well, in fact, the identity on characteristic monoids. It follows that each factor of contributes to automorphisms separately, and we may thus assume without loss of generality. In other words, the automorphism group of splits as a product of the automorphism groups contributed by each node of the target. We are thus reduced to studying three cases: (a) either there are distinguished nodes mapping to the node of ; (b) there is a node in but no node in , i.e., (cf. the minimality condition of Definition 2.5); or (c) there is one non-distinguished node and no distinguished node, i.e., , since every non-distinguished node contributes precisely one factor of in . The first case is most interesting. In this case, the two diagrams take the form

These are homomorphisms of log structures, thus lie over the map to the field , and the factor of maps identically to itself. The automorphism is determined by its action on the generator of and thus takes the form for a unit . On the other hand, we have seen in formula Equation 1’ that

and thus commutativity of the diagram implies ; that is, for all . By minimality, the greatest common divisor of the is , and thus .

In the case (b), the diagrams become

and the map is of the form . However, minimality requires that . Hence, , which has the form , must have and is thus trivial. Finally, in case (c) the diagrams become

and is the identity on the level of characteristics, so is trivial as well, by the same argument as in (b).

From the two lemmas it follows that the localization formula of Reference GP99 can be applied to .

4. The obstruction theory

In this section we analyze the obstruction theory of . We will write for to ease the notation. There is a forgetful morphism, , where

Here, the category is the category whose objects are log schemes and morphisms are strict log morphisms. The stack parametrizes, over a scheme , pairs and of an -marked log curve over and an FM space in over the same log scheme . The morphism sends a log stable map to the pair consisting of the source of the map and the target; it is the forgetful morphism forgetting the data of the map. The stack is the analogue of the Artin stack of prestable curves in ordinary Gromov-Witten theory. It is also smooth and, in fact, log smooth, as it further has the structure of a log stack.

By standard properties of the cotangent complex, the morphism induces a distinguished triangle

Consider the diagram

Here is the universal family of and the universal family over . The morphism is the evaluation map and the projection. It is proven in section 7 of Reference Kim10 that there is a canonical morphism

which is a (relative) perfect obstruction theory. The morphism is strict; therefore the log cotangent complex coincides with the ordinary cotangent complex . Furthermore, the stack is smooth; therefore is a two-term complex concentrated in degrees and (it is an Artin stack, so it has automorphisms). We therefore have a diagram

We may fill in the lower row by the cone of to obtain an (absolute) perfect obstruction theory for . We denote the tangent and obstruction sheaves on by respectively. Applying the functor , we obtain the following lemma.

Lemma 4.1.

Over a geometric point , the tangent space and obstruction space of fit into an exact sequence

The term denotes the logarithmic tangent sheaf, that is, the dual of the sheaf of relative logarithmic differentials of the logarithmic map Reference Kat, section 1.7. The term refers to the group of first order infinitesimal automorphisms of . To carry out localization calculations, we need to know the equivariant Euler classes of . Since the Euler class is a K-theory invariant, it is enough to understand the other four terms in the exact sequence. The terms , , are the cohomology groups of explicit locally free sheaves on the curve , which may be calculated by hand. In fact, more can be said:

Lemma 4.2.

Suppose denotes the canonical contraction map. Then .

Proof.

We recall the explicit construction of the log schemes . For details of the construction, the reader is referred to Reference Li01. Set ; let be the blowup of along the divisor , with a divisor defined as the proper transform of . We view as a family of log schemes over , with logarithmic structure on coming from the divisor , the product log structure on , and the natural log structure on the blowup . The log scheme over is the fiber of this family over , with the induced logarithmic structures on and the base . Next, is the blowup of along , with a divisor the proper tranform of . This is viewed as a family over , with the standard toric log structure on coming from the axes, and the log structure on again arising from the product log structure on and blowing up. The scheme is the fiber over . In general, is constructed from by blowing up ; this is viewed as a family over with the toric logarithmic structure on and the logarithmic structure on coming from and blowing up; is the proper transform of ; and is the fiber over . Then, for each , , …, , the ideal sheaf of in is actually a logarithmic ideal sheaf, so the resulting morphism in the diagram

is a sequence of logarithmic blowups in the sense of Reference Kat, 1.3 (see also, for example, Reference Niz06, section 4 for the definition of the logarithmic blowup). Since logarithmic blowups are log étale morphisms, is log étale. It follows that the log tangent bundle of over is the pullback of the log tangent bundle of over , which is simply the log tangent bundle of . Since log étale morphisms are stable under base change, it follows that the tangent bundle of over the base is pulled back from as well, as claimed.

Thus the two terms only depend on the logarithmic map from to . What we have to understand are the two terms , , that is, the infinitesimal automorphism group and tangent space of a point of . In the discussion that follows, we restrict attention to the stack when is the stack of expansions of , discussed in Remark 2.2.

We will understand the deformation group in terms of the stack of twisted stable curves of Abramovich-Vistoli Reference AV02, which is well understood. To do so, we must digress a bit. First, it will be easier for technical reasons to compare the deformation theory of with the deformation theory of the stack of log twisted curves of Olsson Reference Ols07, which we denote by . It is shown in Reference Ols07 that . Recall the definition of .

Definition 4.1.

A log twisted curve over a scheme is a log curve , where is locally free and is an injection that is locally given by a diagonal matrix.

Let us fix some notation. Denote the set of non-distinguished nodes of the curve by , the set of distinguished nodes by , and the nodes of the target by . In the notation above, we would have . Furthermore, let

denote the “universal target”: this is the moduli space that over a scheme parametrizes line bundles over , together with a section , up to isomorphism.

Consider now a family of log stable maps over a scheme , which specializes over a geometric point to a map . We must now define several morphisms.

First, consider the image of in under the natural forgetful morphism, which forgets the data of the target and the map. Étale locally around , we have a map . The morphism is defined as follows: for an étale neighborhood of in which all log structures and are actually free, the map factors as . This factorization has the following description on the level of characteristic monoids: we have seen in formula Equation 1 of section 2 that the morphism has the form , with a generalized diagonal matrix. This factors as

where is the matrix ‘made diagonal’, i.e.,

and the map is the projection that sends the first coordinates of to the first coordinate of , the next coordinates to the second coordinate of , and so forth. It remains to explain how the morphism lifts from the level of characteristic monoids to the actual log structures; this is the evident extension of formula Equation 1’ of Remark 2.2. If maps

then we now have that maps

Therefore, from the data we obtain a log twisted curve . This defines the required morphism. When we compose with the isomorphism , the twisted curve we obtain is the curve with the -th node of mapping to the -th node of twisted by .

Next, in an étale neighborhood of the image of in , we have a morphism , which can be described as follows. The nodes signify that in an étale neighborhood of the image of in , the image of is in the intersection of boundary divisors of intersecting transversally. In an étale neighborhood of the image of the divisors thus determine line bundles with sections and thus yield the desired map to . We may further obtain a morphism from an étale neighborhood of in in a similar fashion.

Putting everything together, we obtain an étale neighborhood of in and a morphism :

Let us choose a point in lying over .

Lemma 4.3.

The morphism is étale at .

Proof.

Since all stacks in question are smooth, it is sufficient to show that their tangent spaces at and respectively are isomorphic. It will be clear from the proof that we may reduce to the case where two distinguished nodes map into a single node of the target, that is, where . The argument for this is the same as the argument in the proof of Lemma 3.2. We will restrict attention to this case to simplify the notation. We are therefore given étale locally around a diagram

and we want to show that induces an isomorphism of tangent spaces. Since étale maps between smooth stacks induce isomorphisms on tangent spaces, we will work directly with the original moduli stacks and study their tangent spaces around the respective images of the point rather than their étale neighborhoods, again in order to keep the notation simple. Recall that the element consists of data of a pair and two diagrams of log structures

Here the maps and are the maps respectively. We have by formula Equation 1’ that (the here are what we would have called above, but since there is only one target node, we drop the first index to simplify notation). The map is given by mapping the generator , and the rest of the arrows send the generators of to in .

On the other hand, an element of corresponds to a triple of an element of , an element , and an isomorphism between their images in . The element corresponds to a pair of a nodal curve as above and a diagram of log structures

where the top map is an injection ; the map sends ; and the diagonal map is determined by commutativity, . An element of is a line bundle over together with a section, in other words, an element . An element of is similarly a pair . The map sends , and the map sends the data just described to the pair . Therefore the triple has as above, , and , an isomorphism of with in ; that is,

if all and are non-zero,

is arbitrary if .

The morphism then sends the data corresponding to to the triple , where is the curve and the diagram (3) has determined by and , with as in the definition of .

To show that induces an isomorphism of tangent spaces we consider isomorphism classes of morphisms from to all stacks in question extending the given data over . A morphism corresponds to a pair of an infinitesimal deformation of and a diagram

lying over (3). Therefore, we must have under ; maps . The diagonal arrow is determined by commutativity .

Morphisms and lying over the given elements are again line bundles over , which are thus trivial, together with a section restricting to 0 over ; hence they correspond to respectively. Under the morphism maps the extension of to the pair , where the are the ones appearing in diagram (4). Isomorphisms between and restricting to over are of the form . Note that can be arbitrary; however, in order for an isomorphism to exist, the condition is forced. Therefore, the choices involved in extending are the choices of the deformation of and the numbers . Notice however that the choice of either the or the can be eliminated via an isomorphism, for consider two pairs and . There is an isomorphism

with the vertical arrow being the isomorphism where and . In other words, if we denote by the extension of determined by choosing as the extension of (that is, with the choice of the unit ) and by the one with as the extension of (with the unit ) we have an isomorphism between the triple and .

To summarize, the choice of the extension of the image of in corresponds to the data of a choice of a deformation of and the choices of the numbers .

On the other hand, a choice of an extension of in corresponds to the data of a deformation of , a deformation of in , which is necessarily trivial, and diagrams

lying over (2). The extension maps , maps , maps , and the diagonal arrows are determined by commutativity. Notice that up to isomorphism, there is only one choice for the right diagram, the choice of the number . Again, this is because from the diagram

with the isomorphism , and we get an isomorphism between the extension determined by the extensions of with the extension determined by . Since by varying the expression varies through all elements of , the choice of is eliminated up to isomorphism, as claimed. Once the isomorphism is fixed, though, the diagrams on the left with different choices of remain distinct. In other words, the choices involved in extending are up to isomorphism the extension of and the numbers . These are precisely the same choices as involved in extending . This concludes the proof of the lemma.

Remark 4.1.

The geometric meaning of the number in the map is that the -th node is smoothed with speed in the moduli space of twisted curves. The geometric significance of Lemma 4.3 then is that in , all nodes mapping to the same node of the target must be smoothed simultaneously, with the same speed: the speed with which the node of the target is being smoothed.

The lemma in particular implies that we may calculate as follows: Let us write for the divisor of marked points and the twisted curve obtained from as explained. The tangent space to the stack of twisted curves is given by the ext group

where and are the structure sheaf and sheaf of Kahler differentials of the twisted curve respectively. The “local-to-global” spectral sequence for says that the tangent space fits into the short exact sequence

Here Hom and Ext are underlined to indicate that we are taking the sheaf Hom and Ext respectively. The rightmost group in the exact sequence has a canonical description as follows. Let denote the set of non-distinguished nodes of the curve, the set of distinguished nodes of the curve, and the set of nodes of the target, as above. Furthermore, given a node in the curve , let , denote the two components of at . Then we have

There is a diagonal map which simply sends the coordinate corresponding to a node to the sum of the coordinates corresponding to the nodes in mapping to , that is, . This is in fact the map of tangent spaces of the map described above. Just as the tensor product describes intrinsically the part of the deformations of the curve that smooth the nodes, the group describes the part of the deformations of that smooth the nodes of . Here, is obtained from by twisting along the divisor by the integer via the map , just as is obtained from via the map . Then, the fiber diagram of Lemma 4.3 implies the following.

Corollary 4.1.

The tangent space to is the fiber product

The proof of Lemma 4.3 also allows us to understand the automorphism group .

Corollary 4.2.

An infinitesimal automorphism of in consists of an infinitesimal automorphism of the source curve fixing the marked points and an infinitesimal automorphism of in :

Proof.

We give the details for the case when two nodes of map into a node of , as in Lemma 4.3. The general case reduces to this as in the proof of Lemma 3.2. The group of infinitesimal automorphisms of consists of the group of automorphisms of the trivial extension of over lying over the identity automorphism of . This is an infinitesimal automorphism of , an infinitesimal automorphism of in , which consists of a copy of for each expanded component of , and an automorphism of the log structure which lies over the identity automorphism over and which is compatible with both diagrams

Here , as always. An automorphism of must map the generator of for some if it is to reduce to the identity over . In order for the automorphism to be compatible with the second diagram, that is, in order for

to commute, we must have , which implies that . So the logarithmic structures contribute no infinitesimal automorphisms.

5. The virtual localization formula

We are now in a position to derive the virtual localization formula for in the case when the pair carries a -action leaving pointwise fixed. The ideas of this section can essentialy be found in the paper of Graber-Vakil Reference GV05. For the convenience of the reader, we recall the form of relative virtual localization formulas and refer the reader to the paper of Graber-Pandharipande Reference GP99 for details.

5.1. Graber-Pandharipande virtual localization

Suppose is a DM stack with a -action equipped with a -equivariant perfect obstruction theory . Let denote the connected components of the fixed locus of , which we refer to as the fixed loci for brevity, and denote the natural inclusion by . The virtual localization formula reads

Here is a class in , the equivariant Chow ring of (or in equivariant cohomology). The integral is the proper pushforward map from . The term denotes the equivariant Euler class of a vector bundle, in this case, of the virtual normal bundle of in . The qualification that the normal bundle is virtual means that we are not only taking the ordinary normal bundle in the tangent space but also keeping track of the obstruction bundle. More precisely, is defined as , the moving part of the tangent space minus the moving part of the obstruction space (the moving part of a representation is the subrepresentation where acts non-trivially). The Euler class of a sum of vector bundles is by definition the product of the Euler classes; the Euler class of the difference is thus the quotient. Finally, the virtual fundamental class of a fixed locus is by definition the virtual fundamental class arising from the fixed part of the tangent/obstruction theory. Therefore, in order to give a localization formula we must identify the fixed loci and calculate the classes for each of them.

5.2. Types of fixed loci

Consider a pair as above. This defines a stack of expanded targets , as explained in Remark 2.2. We fix discrete data , consisting of the genus of a curve, marked points, and the contact orders of the last marked points at the divisor at infinity of the target . In this section we study the fixed loci of . In the localization formula we will distinguish between two different types of fixed loci; the first type consists of morphisms to itself rather than an expansion of . Such a locus is much simpler to understand than a general locus. The other type of fixed loci consist of those with targets with . The idea of the localization formula is to express the virtual fundamental classes of these loci recursively in terms of the simple loci and moduli spaces of log stable maps to the expanded part of only. We formalize this below.

Definition 5.1.

A fixed locus is called simple if the general, hence every, element has target . We denote a simple fixed locus with discrete data by .

Similarly we define composite loci.

Definition 5.2.

A fixed locus is called composite if the elements map to targets with .

The simple loci are open substacks of ; therefore, all results of section 4 apply to the simple loci without change.

To understand composite loci, we need to understand log stable maps to the expanded part of , that is, to the scheme theoretic closure of the complement of the first component in . The formal definition of such a log stable map is identical to the one given in Definition 2.5. The only differences in the theory of such maps arise from the fact that the rigidifying map to is much simpler, contracting all components to . In the case of , which was the first case to be studied in the literature, the rigidifying map to becomes trivial. We thus call these maps unrigidified log stable maps. We denote the stack parametrizing the expanded part of expansions of by and its universal family by . Similarly to , the stack is isomorphic to the open substack of where the first marking is on the first component and the second marking is on the last component; see Remark 2.2. We will abusively denote the space of log stable maps to targets in by . The analogous stack of unrigidified relative stable maps is introduced and studied in Reference GV05.

Remark 5.1.

The stack is very similar and often simpler than . For instance, observe that

The minimality condition in Definition 2.5 is actually simpler: it simply requires that the log curve be minimal, since it is not possible to have a node of the target with no distinguished node mapping to it anymore.

The analogous map of Lemma 3.2 is also finite, as the map to is never required in the proof of Lemma 3.2.

Theorem 3.1 applies without change. This is because it is shown in Reference GV05 that satisfies the property SE, so Lemma 3.1 applies.

For the purposes of localization, we need to understand the deformation theory of carefully. This is essentially the same as that of . Specifically, we introduce the analogue of the stack ,

and its canonical morphism,

The analogues of Lemma 4.1 and of Corollaries 4.1 and 4.2 are then as follows.

Lemma 5.1.

Over a geometric point , the tangent space and obstruction space of fit into an exact sequence

Proof.

The deformation theory of is the same as in section 4, induced by the relative perfect obstruction theory

where has been replaced by and by . This is true since Kim’s proof works for any stack of log FM spaces, thus expansions and unrigidified expansions alike. Therefore, the discussion of section 4 applies as well, which results in the six-term exact sequence.

Furthermore, with the notation of Corollary 4.1, we have the following.

Corollary 5.1.

The tangent space to is the fiber product

Proof.

The corollary follows from Lemma 4.3; the lemma applies verbatim, as the map to is required nowhere in the proof.

Finally,

Corollary 5.2.

An infinitesimal automorphism of in consists of an infinitesimal automorphism of the source curve fixing the marked points and an automorphism of in :

The two main differences between and are the following: First, the action of on is trivial. This is because dilation of each component of by is an automorphism of , and so the map is isomorphic to . Second, the log tangent bundle is trivial.

5.3. Description of the fixed loci in terms of known stacks

Consider now an element in a composite fixed locus, mapping to a target . Define with as above, and let denote the restriction maps. The discrete data then splits into two sets of discrete data: , consisting of the genus of , a partition describing the behavior of over , the subset of the marked points on not mapping to , and the homology class ; and , consisting of the genus of , the subset of the marked points on not mapping to , the same partition , the original partition describing the behavior along the divisor at infinity, and the homology class determined by . The data are locally constant on the fixed locus. Furthermore, and are naturally logarithmic stable maps, with logarithmic data determined only by . Then, the map belongs to and belongs to . In what follows we will describe the substacks of the fixed locus of , obtained by all possible splittings of the discrete data into , in terms of known stacks involving and . To this end, we need to be able to glue any pair of log stable maps in into a log stable map in . In order to do so, we need first of all to be able to glue the underlying maps, so we work right away with and only need to glue the logarithmic data.

It is convenient to think about log structures in the Borne-Vistoli setting Reference BV12; thus, a log structure on a scheme is a sheaf of sharp monoids and a functor . Concretely, to every element of we assign a line bundle with a section on in a functorial manner. In our case everything is particularly simple, as the characteristic monoids will be just of the form for some integer , as in fact the characteristic monoids of the relevant sublog structures will be constant, so all this data will amount to choosing line bundles with sections on . A morphism between two log structures on is a map such that is naturally isomorphic to .

A. The canonical log structures on the glued source and target.

Consider a family of nodal curves , for which nodes persist. We label the nodes (connected components of the relative singular locus) by , …, , and by and the two components of the partial normalization of at the nodes. For notational convenience, we denote the images of the sections of and which map to the node by the same symbol. The canonical log structure on (see section 2) has a locally free sublog structure corresponding to the nodes. In the language of Borne-Vistoli the latter is the log structure ; that is, the line bundle that parametrizes deformations of the node in and with the section, as corresponds to a map to that maps into the boundary divisor determined by the node, and the log structure on is pulled back by the divisorial log structure on under this map. Similarly, for a family of singular expansions in for which the singularity persists, we will write . Here is the simple part of , its expanded part, and denotes the Cartier divisor in and , canonically identified with (see Reference Li01, section 4.1), along which we glue them to obtain . Since the fibers of are trivial, a standard cohomology and base change argument shows that the pushforward of via is locally free and that the adjunction is an isomorphism (this can be checked on the local models of described in Reference Li01). Then the canonical log structure of on (see section 2) has a locally free sublog structure corresponding to given by , where , since again the divisor persists.

B. Gluing the logarithmic maps.

Suppose now that we have two families and over in and respectively. On the level of schemes, the problem has a unique gluing to a family

The map however is not a log map at the moment. To promote it to one, we must correct the log structure on . Any log structure on that makes simultaneously and log smooth over must receive maps from both canonical log structures and . Furthermore, if is the contact order of at , the map at must, at the level of log structures, have the form

Here, in the two pushouts, the map is the map determined from the maps and as the image of the generator corresponding to and respectively. The map is the diagonal in both cases. Consequently, the image of the part of the log structure corresponding to in must be identified with times the image of the part of the log structure corresponding to in . We claim there is a log structure on with a unique map , i.e., an initial log structure that makes into a logarithmic stable map or, in other words, a minimal one (Definition 2.2). To see this, note that the maps restricted to give isomorphisms and . Hence, pushing forward to , we get isomorphisms

Since and restricted to are isomorphisms, the notation can be simplified to

On the other hand, recall that (section 5.3.A). Since and restricted to are equal to and respectively we get simply that

Therefore, if is defined by

we see that there is a unique map which sends to and to .

This does not solve the problem still, as to produce a family in , the log structure on the base must have characteristic . In fact, the issue is simply that is not a sheaf of saturated monoids. Saturating indeed gives and corresponds to adding a generator such that , , with , . On the level of log structures (not just characteristics), consider the diagram

Here is viewed as the substack of parametrizing pairs of line bundles with the zero section and is the morphism which, over a scheme , sends a -tuple of pairs of line bundles with their zero sections to and the diagonal morphism. We then have a morphism to the fiber product

which sends the data of the two maps to the collection

and isomorphism Equation 8. Then, saturating is equivalent to taking the fiber product

Lemma 5.2.

The morphism is finite étale and surjective of pure degree .

Proof.

The lower arrow of the above cartesian square factors as

where sends a pair of a line bundle with its zero section to copies of . The second morphism is base change from the -th power morphism , hence a gerbe banded by . The first one is a -torsor; indeed an -point of the fiber product in the middle corresponds to line bundles , , …,  on with isomorphisms . Then the pullback of via is equivalent to the fibered category associated to the functor on -schemes that sends to -tuples of isomorphisms whose -th power is .

C. Gluing the forgetful morphisms.

Recall that there exist forgetful morphisms and . There exists an analogous forgetful morphism from to a stack , to be defined below, whose objects are obtained by appropriately gluing the objects of and . To be more precise, we define a stack via the fiber product

where is given by diagram Equation 9 above. To define the right vertical arrow we glue the underlying scheme data of curves and targets of a pair of elements in and over canonically along their sections and common distinguished divisor respectively. With notation as above, let the resulting glued pair have underlying scheme data and ; write for the glued family of divisors. Then we map the glued pair to

Then , over a scheme , parametrizes pairs of elements of and , together with isomorphisms

Now the morphism factors through . Indeed, this follows from isomorphism Equation 8, which was obtained as soon as the underlying maps were glued.

All in all, there is a stack , which, over a scheme , parametrizes

a pair of elements of and , with isomorphisms

a line bundle on such that , and isomorphisms such that .

Concretely, we have a diagram

Here, the map is precisely as in the previous section. The map sends a pair of elements of and over , together with the isomorphisms defined above, to

Consequently, the composed map is the one considered in the previous section as well. From the fact that the bottom square in the diagram is cartesian, we conclude that is algebraic and the map is étale of pure degree . Furthermore, since the big square is cartesian, the top square is also cartesian.

Now let be the group of automorphisms of the partition , that is, the number of bijections with , and consider its natural action on . We then have a canonical identification

of the fixed locus with the quotient stack of by the (finite) group in the sense of Reference Rom05. Finally, also acts naturally on , and, what is more, the image of the fixed locus in can be canonically identified with . Let be the quotient map. We then have a cartesian diagram

We thus have a description of the fixed loci in terms of known stacks. What is not a priori clear is the relation between the natural obstruction theory induced on by the obstruction theory of and those of the two factors .

D. Comparison of obstruction theories.

Let us denote by and the tangent and obstruction space of at a point , as in Lemma 4.1, and by the tangent and obstruction spaces of the two component maps .

Lemma 5.3.

At the point we have the equality

in -theory.

Proof.

Denote by , …, the nodes connecting with , that is, the nodes over . Then , …,  become marked points in . We denote each of the two markings on and respectively over the node by the same symbol for ease of notation. Let be the composition of with the contraction , and let . By Lemmas 4.1 and 4.2 combined we have that the fibers of fit into the six-term exact sequence

We have two similar six-term exact sequences for with replaced by and by and replaced by for . Write for the vector of all marked points on , as in Corollary 4.2 above, and for the vector of marked points in . Observe that

-vector fields that vanish on the nodes and marked points of are simply vector fields that vanish on the nodes and marked points of and and the nodes connecting the two. Furthermore,

and thus

Similarly, from the local-to-global sequence Equation 6 we have an equality in -theory

On the other hand, we have by Corollary 4.1 that differs in -theory from by replacing with . All nodes of persist as nodes over some target node in and , except precisely the nodes over , as is not a target node for either or . Therefore,

It remains to analyze the relative deformations and obstructions of given by the cohomology groups . We have the normalization sequence

By Reference Kat96, Example 10.2 and Reference Gro11, Example 3.31, the log tangent bundle fits into the short exact sequence

and thus coincides at a point of with the tangent space at that point in . Therefore, twisting the normalization sequence by and taking cohomology we get

Therefore, the difference between and is precisely . Putting everything together yields the lemma.

From the above lemma it follows that in -theory the pullbacks of the (virtual) sheaves and to differ by two bundles: the first is the bundle with fiber , which may be identified with the pullback of the tangent bundle under the evaluation map ; the second one is the bundle with fiber . This is the line bundle that parametrizes deformations of the node ; it may be identified with the pullback of to , where and are the projections of to the two factors, and are the respective similar bundles. Note that is a trivial bundle with non-trivial action, while is a non-trivial bundle with trivial action. All the above sheaves on descend to ; we will denote the sheaves descended from them by the same symbol. To keep consistent with existing literature, we write . We then obtain:

Corollary 5.3.

If and , we have

Proof.

Lemma 5.3 implies that differs from the sum of the only by the bundle , since has trivial action. Furthermore, since the torus action on is trivial, the bundles have no moving part.

The perfect obstruction theory of the fixed locus is by definition obtained by restriction from the torus fixed part of the perfect obstruction theory of relative to discussed in section 4. Recall that the image of in was identified with the étale quotient of the stack by the group (section 5.3.C). Now note that the relative cotangent complex of in has trivial torus-fixed part: the torus action on the deformation space of the nodes and the divisor obtained from gluing the objects in and only has trivial fixed locus. Therefore has a perfect obstruction theory relative to its image, which yields the same virtual fundamental class . Consequently, in view of the cartesian diagram Equation 10, we may pull back the above obstruction theory via the étale quotient map to obtain a perfect obstruction theory for relative to and thus a virtual fundamental class . Now, by Reference Cos06, Theorem 5.0.1, we have

We have the following cartesian diagram

Now let be the morphism defined right before Lemma 5.2. Putting everything together, we obtain

Theorem 5.1 (Log Virtual Localization).

In the statement of the theorem the symbols , , and denote the proper pushforward, flat, and Gysin pullback operations, respectively, on the Chow groups of the corresponding Deligne-Mumford stacks (see Reference Vis89).

Proof.

It is enough to show that

Then the theorem will follow immediately from equality Equation 11 and Lemma 5.3.

First, we claim that has a perfect obstruction theory relative to , which yields the same virtual fundamental class as its perfect obstruction theory relative to discussed above. The reason is that the relative cotangent complex of over has trivial torus-fixed part. Indeed, since is étale over , it suffices to consider the action of on the cotangent bundle of the fibers of . In the notation of section 5.3.C, consider an element of with underlying scheme data and and isomorphisms , , …, . Then an element acts on by multiplying them by , so its -action on the cotangent bundle of is non-trivial.

We have a commutative diagram

where is the composite morphism , , the forgetful morphisms, and the morphism defined in diagram Equation 12. By the previous paragraph and the proof of Lemma 5.3, the deformation and obstruction spaces of the map and its component maps , over the same point in , differ by the term . In the formalism of Manolache Reference Man12, Definition 4.5, this is equivalent to the statement that the perfect obstruction theories of and together with the trivial obstruction theory of , given by its relative cotangent complex , form a compatible triple. Clearly is of Deligne-Mumford type, and the same argument as in Reference AMW12, Lemma 4.2.1 shows that is also of Deligne-Mumford type. Then, using Reference Man12, Lemma 4.9, we deduce that is the virtual pullback of (Reference Man12, Definition 3.7]) with respect to , which, in our case, is simply its Gysin pullback via followed by the flat pullback (see Remarks 3.9 and 3.10 in Reference Man12).

In section 2, we discussed the finite morphism from the moduli space of log stable maps to the moduli space of relative stable maps. In the paper Reference AMW12 it is shown that the pushforward of the virtual fundamental class of under coincides with the virtual fundamental class of Jun Li’s space. We may modify these results to include the maps and , with the appropriate modifications of the spaces in the setting of relative stable maps as targets. Then, applying to both sides of the equation in Theorem 5.1 yields the relative virtual localization theorem of Graber-Vakil.

Corollary 5.4.

The log virtual localization formula becomes the relative virtual localization formula under the functor .

Proof.

First, note that the fixed locus of Li’s space corresponding to the splitting data is identified as the quotient stack of by Reference GV05, page 13; let be the quotient map. Then we have a commutative diagram

where is the finite map induced by and . Therefore, the effect of on the numerators appearing on the right-hand side of the formula is the same as the effect of on . By the projection formula and Lemma 5.2, we have

Now, we have the cartesian diagram

over . Therefore, by Reference Vis89, Theorem 3.12 and the results of Reference AMW12, we get

which, in turn, after applying gives by Reference GV05, Lemma 3.2.

So what remains is to analyze the Euler classes appearing in the denominators of the formula. The term does not change under , as it is the Euler class of the virtual normal bundle of a fixed locus inside the simple locus, and is an isomorphism over the simple locus. On the other hand, let be the line bundle in parametrizing deformations of the node; its fiber at a point is . We denote the Euler class of by , as in Reference GV05. Note that over a fixed locus , the pullback , where is the line bundle of with fiber parametrizing deformations of the node , which is the -th root of . We thus have , which justifies the choice of notation for .

Summing over all gives precisely the relative virtual localization formula of Reference GV05.

Acknowledgments

We are grateful to our advisor, Dan Abramovich, for suggesting this problem to us and for numerous valuable discussions. We also thank Angelo Vistoli for providing the idea of the proof of Lemma 3.1. Finally, we would like to thank the anonymous referee for useful comments.

Mathematical Fragments

Equation (*)
Definition 2.2.

A log curve is called minimal if the log structure is locally free and there is no locally free submonoid that contains the image of .

Definition 2.5.

A log stable map from a log curve to an extended log twisted FM space is a log morphism over such that, over each point :

The cokernel of the map has rank equal to the number of non-distinguished nodes.

The map is simple at the distinguished nodes.

Stability: The automorphism group is finite.

The following minimality condition holds: Either is a minimal log curve or, if not, then there exists a subsheaf of monoids of , which is a locally free log structure on , such that , minimal, and surjective.

Equation (1)
Remark 2.2.

Let us at this point explain the connection with Jun Li’s original definitions and clarify this concept geometrically. A family of expansions of a pair or, similarly, of a -semistable degeneration has canonical log structures that determine log FM spaces. The canonical log structure on a family of expansions is obtained in a manner formally identical to the way that the canonical log structure on a nodal curve is obtained. A detailed treatment of the canonical log structures on expansions can be found in Olsson’s paper Reference Ols03. Briefly, there is an algebraic stack parametrizing expansions. The stack is in fact the open substack of the stack of marked, genus prestable curves where the first two markings are on the first component of the curve and the third marking on the last; see for instance Reference GV05. In there is a normal crossings divisor corresponding to singular expansions; therefore, admits a log structure . Similarly, the universal family over admits a log structure . A family of expansions corresponds to a cartesian diagram

The pullback log structures and on and are what we denoted by and above. Therefore, expansions are examples of log FM spaces. We may thus consider log stable maps to expansions. The underlying morphism of schemes of such a log stable map is a relative stable map in the sense of Jun Li; the predeformability condition is enforced by the requirement that the map is a map of log schemes. The log structures are thus additional algebraic data on a relative stable map. The log structures encode essential geometric information very conveniently. Suppose for simplicity that is a geometric point. We have seen above the form of the log structures , and the maps between them. The rank of , which we denoted by the number above, indicates that the target is the -th expansion of . The number is the number of non-distinguished nodes. The number is the number of distinguished nodes. The matrix above indicates that of the distinguished nodes map to the first singular locus in (namely to , map to , and so forth. The contact order of the -th node mapping to the -th singular locus is . Note that once the underlying stable map is fixed, the diagram

between the characteristic monoids of the log structures is determined. This means that in order to determine the full diagram

we need to determine the elements of to which the generators of and are mapping. In fact, all generators of may be chosen to map into after automorphism, so it is enough to treat only . Matrix Equation 1 in Remark 2.1 indicates that has the following form:

Note however that the units are restricted: they must satisfy the equation , in order for the diagram

to commute. This shows that there is a finite number of ways to give to a relative stable map the structure of a log stable map. In other words, if denotes Jun Li’s space of expansions and Kim’s space of log stable maps to expansions of , which is algebraic stack since the stack of expansions is algebraic, there is a forgetful morphism

For the rigorous definition of the morphism we refer to the work of Gross and Siebert Reference GS11 and the paper Reference AMW12 of Abramovich, Marcus, and Wise. This is the left vertical arrow of diagram Equation * of the introduction. The fact that relative stable maps are a locally closed substack of the stack of all maps expresses the fact that the predeformability condition is locally closed. The fact that the stack of log stable maps is open in the stack of all log maps expresses the fact that predeformability is enforced by requiring that the map from a nodal curve to an expansion be a log map.

Theorem 3.1.

There is a locally closed equivariant immersion of into a smooth Deligne-Mumford stack.

Lemma 3.1.

Suppose is a equivariant finite morphism between Deligne-Mumford stacks, and assume that satisfies SE. Then satisfies SE as well and thus embeds equivariantly into a smooth Deligne-Mumford stack.

Lemma 3.2.

The morphism is finite.

Lemma 4.1.

Over a geometric point , the tangent space and obstruction space of fit into an exact sequence

Lemma 4.2.

Suppose denotes the canonical contraction map. Then .

Lemma 4.3.

The morphism is étale at .

Equation (6)
Corollary 4.1.

The tangent space to is the fiber product

Corollary 4.2.

An infinitesimal automorphism of in consists of an infinitesimal automorphism of the source curve fixing the marked points and an infinitesimal automorphism of in :

Equation (8)
Equation (9)
Lemma 5.2.

The morphism is finite étale and surjective of pure degree .

Equation (10)
Lemma 5.3.

At the point we have the equality

in -theory.

Corollary 5.3.

If and , we have

Equation (11)
Equation (12)
Theorem 5.1 (Log Virtual Localization).
Corollary 5.4.

The log virtual localization formula becomes the relative virtual localization formula under the functor .

References

Reference [AC11]
Dan Abramovich and Qile Chen, Stable logarithmic maps to Deligne-Faltings pairs II, Asian J. Math. 18 (2014), no. 3, 465–488, DOI 10.4310/AJM.2014.v18.n3.a5. MR3257836,
Show rawAMSref \bib{AC}{article}{ label={AC11}, author={Abramovich, Dan}, author={Chen, Qile}, title={Stable logarithmic maps to Deligne-Faltings pairs II}, journal={Asian J. Math.}, volume={18}, date={2014}, number={3}, pages={465--488}, issn={1093-6106}, review={\MR {3257836}}, doi={10.4310/AJM.2014.v18.n3.a5}, }
Reference [AMW12]
Dan Abramovich, Steffen Marcus, and Jonathan Wise, Comparison theorems for Gromov-Witten invariants of smooth pairs and of degenerations (English, with English and French summaries), Ann. Inst. Fourier (Grenoble) 64 (2014), no. 4, 1611–1667. MR3329675,
Show rawAMSref \bib{AMW}{article}{ label={AMW12}, author={Abramovich, Dan}, author={Marcus, Steffen}, author={Wise, Jonathan}, title={Comparison theorems for Gromov-Witten invariants of smooth pairs and of degenerations}, language={English, with English and French summaries}, journal={Ann. Inst. Fourier (Grenoble)}, volume={64}, date={2014}, number={4}, pages={1611--1667}, issn={0373-0956}, review={\MR {3329675}}, }
Reference [AV02]
Dan Abramovich and Angelo Vistoli, Compactifying the space of stable maps, J. Amer. Math. Soc. 15 (2002), no. 1, 27–75, DOI 10.1090/S0894-0347-01-00380-0. MR1862797,
Show rawAMSref \bib{AV}{article}{ label={AV02}, author={Abramovich, Dan}, author={Vistoli, Angelo}, title={Compactifying the space of stable maps}, journal={J. Amer. Math. Soc.}, volume={15}, date={2002}, number={1}, pages={27--75}, issn={0894-0347}, review={\MR {1862797}}, doi={10.1090/S0894-0347-01-00380-0}, }
Reference [BV12]
Niels Borne and Angelo Vistoli, Parabolic sheaves on logarithmic schemes, Adv. Math. 231 (2012), no. 3-4, 1327–1363, DOI 10.1016/j.aim.2012.06.015. MR2964607,
Show rawAMSref \bib{bv}{article}{ label={BV12}, author={Borne, Niels}, author={Vistoli, Angelo}, title={Parabolic sheaves on logarithmic schemes}, journal={Adv. Math.}, volume={231}, date={2012}, number={3-4}, pages={1327--1363}, issn={0001-8708}, review={\MR {2964607}}, doi={10.1016/j.aim.2012.06.015}, }
Reference [Che14]
Qile Chen, Stable logarithmic maps to Deligne-Faltings pairs I, Ann. of Math. (2) 180 (2014), no. 2, 455–521, DOI 10.4007/annals.2014.180.2.2. MR3224717,
Show rawAMSref \bib{Cann}{article}{ label={Che14}, author={Chen, Qile}, title={Stable logarithmic maps to Deligne-Faltings pairs I}, journal={Ann. of Math. (2)}, volume={180}, date={2014}, number={2}, pages={455--521}, issn={0003-486X}, review={\MR {3224717}}, doi={10.4007/annals.2014.180.2.2}, }
Reference [CKL17]
Huai-Liang Chang, Young-Hoon Kiem, and Jun Li, Torus localization and wall crossing for cosection localized virtual cycles, Adv. Math. 308 (2017), 964–986, DOI 10.1016/j.aim.2016.12.019. MR3600080,
Show rawAMSref \bib{CKL}{article}{ label={CKL17}, author={Chang, Huai-Liang}, author={Kiem, Young-Hoon}, author={Li, Jun}, title={Torus localization and wall crossing for cosection localized virtual cycles}, journal={Adv. Math.}, volume={308}, date={2017}, pages={964--986}, issn={0001-8708}, review={\MR {3600080}}, doi={10.1016/j.aim.2016.12.019}, }
Reference [Cos06]
Kevin Costello, Higher genus Gromov-Witten invariants as genus zero invariants of symmetric products, Ann. of Math. (2) 164 (2006), no. 2, 561–601, DOI 10.4007/annals.2006.164.561. MR2247968,
Show rawAMSref \bib{cos}{article}{ label={Cos06}, author={Costello, Kevin}, title={Higher genus Gromov-Witten invariants as genus zero invariants of symmetric products}, journal={Ann. of Math. (2)}, volume={164}, date={2006}, number={2}, pages={561--601}, issn={0003-486X}, review={\MR {2247968}}, doi={10.4007/annals.2006.164.561}, }
Reference [Gil11]
W. D. Gillam, Logarithmic stacks and minimality, Internat. J. Math. 23 (2012), no. 7, 1250069, 38, DOI 10.1142/S0129167X12500693. MR2945649,
Show rawAMSref \bib{GMinimality}{article}{ label={{Gil}11}, author={Gillam, W. D.}, title={Logarithmic stacks and minimality}, journal={Internat. J. Math.}, volume={23}, date={2012}, number={7}, pages={1250069, 38}, issn={0129-167X}, review={\MR {2945649}}, doi={10.1142/S0129167X12500693}, }
Reference [GP99]
T. Graber and R. Pandharipande, Localization of virtual classes, Invent. Math. 135 (1999), no. 2, 487–518, DOI 10.1007/s002220050293. MR1666787,
Show rawAMSref \bib{GP}{article}{ label={GP99}, author={Graber, T.}, author={Pandharipande, R.}, title={Localization of virtual classes}, journal={Invent. Math.}, volume={135}, date={1999}, number={2}, pages={487--518}, issn={0020-9910}, review={\MR {1666787}}, doi={10.1007/s002220050293}, }
Reference [Gro11]
Mark Gross, Tropical geometry and mirror symmetry, CBMS Regional Conference Series in Mathematics, vol. 114, Published for the Conference Board of the Mathematical Sciences, Washington, DC; by the American Mathematical Society, Providence, RI, 2011, DOI 10.1090/cbms/114. MR2722115,
Show rawAMSref \bib{gross}{book}{ label={Gro11}, author={Gross, Mark}, title={Tropical geometry and mirror symmetry}, series={CBMS Regional Conference Series in Mathematics}, volume={114}, publisher={Published for the Conference Board of the Mathematical Sciences, Washington, DC; by the American Mathematical Society, Providence, RI}, date={2011}, pages={xvi+317}, isbn={978-0-8218-5232-3}, review={\MR {2722115}}, doi={10.1090/cbms/114}, }
Reference [GS11]
Mark Gross and Bernd Siebert, Logarithmic Gromov-Witten invariants, J. Amer. Math. Soc. 26 (2013), no. 2, 451–510, DOI 10.1090/S0894-0347-2012-00757-7. MR3011419,
Show rawAMSref \bib{GS}{article}{ label={GS11}, author={Gross, Mark}, author={Siebert, Bernd}, title={Logarithmic Gromov-Witten invariants}, journal={J. Amer. Math. Soc.}, volume={26}, date={2013}, number={2}, pages={451--510}, issn={0894-0347}, review={\MR {3011419}}, doi={10.1090/S0894-0347-2012-00757-7}, }
Reference [GV05]
Tom Graber and Ravi Vakil, Relative virtual localization and vanishing of tautological classes on moduli spaces of curves, Duke Math. J. 130 (2005), no. 1, 1–37, DOI 10.1215/S0012-7094-05-13011-3. MR2176546,
Show rawAMSref \bib{GV}{article}{ label={GV05}, author={Graber, Tom}, author={Vakil, Ravi}, title={Relative virtual localization and vanishing of tautological classes on moduli spaces of curves}, journal={Duke Math. J.}, volume={130}, date={2005}, number={1}, pages={1--37}, issn={0012-7094}, review={\MR {2176546}}, doi={10.1215/S0012-7094-05-13011-3}, }
Reference [IP03]
Eleny-Nicoleta Ionel and Thomas H. Parker, Relative Gromov-Witten invariants, Ann. of Math. (2) 157 (2003), no. 1, 45–96, DOI 10.4007/annals.2003.157.45. MR1954264,
Show rawAMSref \bib{ionrel}{article}{ label={IP03}, author={Ionel, Eleny-Nicoleta}, author={Parker, Thomas H.}, title={Relative Gromov-Witten invariants}, journal={Ann. of Math. (2)}, volume={157}, date={2003}, number={1}, pages={45--96}, issn={0003-486X}, review={\MR {1954264}}, doi={10.4007/annals.2003.157.45}, }
Reference [IP04]
Eleny-Nicoleta Ionel and Thomas H. Parker, The symplectic sum formula for Gromov-Witten invariants, Ann. of Math. (2) 159 (2004), no. 3, 935–1025, DOI 10.4007/annals.2004.159.935. MR2113018,
Show rawAMSref \bib{IP}{article}{ label={IP04}, author={Ionel, Eleny-Nicoleta}, author={Parker, Thomas H.}, title={The symplectic sum formula for Gromov-Witten invariants}, journal={Ann. of Math. (2)}, volume={159}, date={2004}, number={3}, pages={935--1025}, issn={0003-486X}, review={\MR {2113018}}, doi={10.4007/annals.2004.159.935}, }
Reference [Kat]
K. Kato, Logarithmic degeneration and Dieudonne theory, preprint.
Reference [Kat96]
Fumiharu Kato, Log smooth deformation theory, Tohoku Math. J. (2) 48 (1996), no. 3, 317–354, DOI 10.2748/tmj/1178225336. MR1404507,
Show rawAMSref \bib{fkd}{article}{ label={Kat96}, author={Kato, Fumiharu}, title={Log smooth deformation theory}, journal={Tohoku Math. J. (2)}, volume={48}, date={1996}, number={3}, pages={317--354}, issn={0040-8735}, review={\MR {1404507}}, doi={10.2748/tmj/1178225336}, }
Reference [Kat00]
Fumiharu Kato, Log smooth deformation and moduli of log smooth curves, Internat. J. Math. 11 (2000), no. 2, 215–232, DOI 10.1142/S0129167X0000012X. MR1754621,
Show rawAMSref \bib{fK}{article}{ label={Kat00}, author={Kato, Fumiharu}, title={Log smooth deformation and moduli of log smooth curves}, journal={Internat. J. Math.}, volume={11}, date={2000}, number={2}, pages={215--232}, issn={0129-167X}, review={\MR {1754621}}, doi={10.1142/S0129167X0000012X}, }
Reference [Kim10]
Bumsig Kim, Logarithmic stable maps, New developments in algebraic geometry, integrable systems and mirror symmetry (RIMS, Kyoto, 2008), Adv. Stud. Pure Math., vol. 59, Math. Soc. Japan, Tokyo, 2010, pp. 167–200. MR2683209,
Show rawAMSref \bib{Kim}{article}{ label={Kim10}, author={Kim, Bumsig}, title={Logarithmic stable maps}, conference={ title={New developments in algebraic geometry, integrable systems and mirror symmetry}, address={RIMS, Kyoto}, date={2008}, }, book={ series={Adv. Stud. Pure Math.}, volume={59}, publisher={Math. Soc. Japan, Tokyo}, }, date={2010}, pages={167--200}, review={\MR {2683209}}, }
Reference [Kon95]
Maxim Kontsevich, Enumeration of rational curves via torus actions, The moduli space of curves (Texel Island, 1994), Progr. Math., vol. 129, Birkhäuser Boston, Boston, MA, 1995, pp. 335–368, DOI 10.1007/978-1-4612-4264-2_12. MR1363062,
Show rawAMSref \bib{KTorus}{article}{ label={Kon95}, author={Kontsevich, Maxim}, title={Enumeration of rational curves via torus actions}, conference={ title={The moduli space of curves}, address={Texel Island}, date={1994}, }, book={ series={Progr. Math.}, volume={129}, publisher={Birkh\"{a}user Boston, Boston, MA}, }, date={1995}, pages={335--368}, review={\MR {1363062}}, doi={10.1007/978-1-4612-4264-2\_12}, }
Reference [Li01]
Jun Li, Stable morphisms to singular schemes and relative stable morphisms, J. Differential Geom. 57 (2001), no. 3, 509–578. MR1882667,
Show rawAMSref \bib{Li}{article}{ label={Li01}, author={Li, Jun}, title={Stable morphisms to singular schemes and relative stable morphisms}, journal={J. Differential Geom.}, volume={57}, date={2001}, number={3}, pages={509--578}, issn={0022-040X}, review={\MR {1882667}}, }
Reference [Li02]
Jun Li, A degeneration formula of GW-invariants, J. Differential Geom. 60 (2002), no. 2, 199–293. MR1938113,
Show rawAMSref \bib{Lideg}{article}{ label={Li02}, author={Li, Jun}, title={A degeneration formula of GW-invariants}, journal={J. Differential Geom.}, volume={60}, date={2002}, number={2}, pages={199--293}, issn={0022-040X}, review={\MR {1938113}}, }
Reference [LR01]
An-Min Li and Yongbin Ruan, Symplectic surgery and Gromov-Witten invariants of Calabi-Yau 3-folds, Invent. Math. 145 (2001), no. 1, 151–218, DOI 10.1007/s002220100146. MR1839289,
Show rawAMSref \bib{LR}{article}{ label={LR01}, author={Li, An-Min}, author={Ruan, Yongbin}, title={Symplectic surgery and Gromov-Witten invariants of Calabi-Yau 3-folds}, journal={Invent. Math.}, volume={145}, date={2001}, number={1}, pages={151--218}, issn={0020-9910}, review={\MR {1839289}}, doi={10.1007/s002220100146}, }
Reference [Man12]
Cristina Manolache, Virtual pull-backs, J. Algebraic Geom. 21 (2012), no. 2, 201–245, DOI 10.1090/S1056-3911-2011-00606-1. MR2877433,
Show rawAMSref \bib{manol}{article}{ label={Man12}, author={Manolache, Cristina}, title={Virtual pull-backs}, journal={J. Algebraic Geom.}, volume={21}, date={2012}, number={2}, pages={201--245}, issn={1056-3911}, review={\MR {2877433}}, doi={10.1090/S1056-3911-2011-00606-1}, }
Reference [MFK94]
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}{ label={MFK94}, 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}, }
Reference [Niz06]
Wiesława Nizioł, Toric singularities: log-blow-ups and global resolutions, J. Algebraic Geom. 15 (2006), no. 1, 1–29, DOI 10.1090/S1056-3911-05-00409-1. MR2177194,
Show rawAMSref \bib{Niz}{article}{ label={Niz06}, author={Nizio\l , Wies\l awa}, title={Toric singularities: log-blow-ups and global resolutions}, journal={J. Algebraic Geom.}, volume={15}, date={2006}, number={1}, pages={1--29}, issn={1056-3911}, review={\MR {2177194}}, doi={10.1090/S1056-3911-05-00409-1}, }
Reference [Ols03]
Martin C. Olsson, Universal log structures on semi-stable varieties, Tohoku Math. J. (2) 55 (2003), no. 3, 397–438. MR1993863,
Show rawAMSref \bib{OlssonSemistable}{article}{ label={Ols03}, author={Olsson, Martin C.}, title={Universal log structures on semi-stable varieties}, journal={Tohoku Math. J. (2)}, volume={55}, date={2003}, number={3}, pages={397--438}, issn={0040-8735}, review={\MR {1993863}}, }
Reference [Ols07]
Martin C. Olsson, (Log) twisted curves, Compos. Math. 143 (2007), no. 2, 476–494, DOI 10.1112/S0010437X06002442. MR2309994,
Show rawAMSref \bib{Otw}{article}{ label={Ols07}, author={Olsson, Martin C.}, title={(Log) twisted curves}, journal={Compos. Math.}, volume={143}, date={2007}, number={2}, pages={476--494}, issn={0010-437X}, review={\MR {2309994}}, doi={10.1112/S0010437X06002442}, }
Reference [Rom05]
Matthieu Romagny, Group actions on stacks and applications, Michigan Math. J. 53 (2005), no. 1, 209–236, DOI 10.1307/mmj/1114021093. MR2125542,
Show rawAMSref \bib{R}{article}{ label={Rom05}, author={Romagny, Matthieu}, title={Group actions on stacks and applications}, journal={Michigan Math. J.}, volume={53}, date={2005}, number={1}, pages={209--236}, issn={0026-2285}, review={\MR {2125542}}, doi={10.1307/mmj/1114021093}, }
Reference [Tot04]
Burt Totaro, The resolution property for schemes and stacks, J. Reine Angew. Math. 577 (2004), 1–22, DOI 10.1515/crll.2004.2004.577.1. MR2108211,
Show rawAMSref \bib{Totaro}{article}{ label={Tot04}, author={Totaro, Burt}, title={The resolution property for schemes and stacks}, journal={J. Reine Angew. Math.}, volume={577}, date={2004}, pages={1--22}, issn={0075-4102}, review={\MR {2108211}}, doi={10.1515/crll.2004.2004.577.1}, }
Reference [Vis89]
Angelo Vistoli, Intersection theory on algebraic stacks and on their moduli spaces, Invent. Math. 97 (1989), no. 3, 613–670, DOI 10.1007/BF01388892. MR1005008,
Show rawAMSref \bib{intersection}{article}{ label={Vis89}, author={Vistoli, Angelo}, title={Intersection theory on algebraic stacks and on their moduli spaces}, journal={Invent. Math.}, volume={97}, date={1989}, number={3}, pages={613--670}, issn={0020-9910}, review={\MR {1005008}}, doi={10.1007/BF01388892}, }

Article Information

MSC 2010
Primary: 14N35 (Gromov-Witten invariants, quantum cohomology, Gopakumar-Vafa invariants, Donaldson-Thomas invariants)
Secondary: 14D23 (Stacks and moduli problems)
Author Information
S. Molcho
Department of Mathematics, University of Colorado Boulder, Campus Box 395, Boulder, Colorado 80309-0395; and Mathematical Research Center, Scuola Normale Superiore di Pisa 56126 Pisa, Italy
samouil.molcho@sns.it
MathSciNet
E. Routis
Department of Mathematics, Brown University, Box 1917, 151 Thayer Street, Providence, Rhode Island 02912; and Kavli IPMU (WPI), UTIAS, The University of Tokyo, Kashiwa, Chiba 277-8583, Japan
Address at time of publication: Max Planck Institute for Mathematics, Vivatsgasse 7, 53111 Bonn, Germany
evangelos.routis@ipmu.jp, routis@mpim-bonn.mpg.de
MathSciNet
Additional Notes

The second author was supported by the World Premier International Research Center Initiative (WPI), MEXT, Japan.

Journal Information
Transactions of the American Mathematical Society, Series B, Volume 6, Issue 3, ISSN 2330-0000, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on , revised on , , and published on .
Copyright Information
Copyright 2019 by the authors under Creative Commons Attribution-Noncommercial 3.0 License (CC BY NC 3.0)
Article References
  • Permalink
  • Permalink (PDF)
  • DOI 10.1090/btran/31
  • MathSciNet Review: 3905962
  • Show rawAMSref \bib{3905962}{article}{ author={Molcho, S.}, author={Routis, E.}, title={Localization for logarithmic stable maps}, journal={Trans. Amer. Math. Soc. Ser. B}, volume={6}, number={3}, date={2019}, pages={80-113}, issn={2330-0000}, review={3905962}, doi={10.1090/btran/31}, }

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.