RESEARCH STATEMENT: COMPACTIFYING RELATIVE PICARD SCHEMES

Similar documents
COMPACTIFYING THE RELATIVE PICARD FUNCTOR OVER DEGENERATIONS OF VARIETIES

ON DEGENERATIONS OF MODULI OF HITCHIN PAIRS

Two Denegeration Techniques for Maps of Curves

Geometry of the theta divisor of a compactified jacobian

Non-properness of the functor of F -trivial bundles

CANONICAL METRICS AND STABILITY OF PROJECTIVE VARIETIES

If F is a divisor class on the blowing up X of P 2 at n 8 general points p 1,..., p n P 2,

BRILL-NOETHER THEORY. This article follows the paper of Griffiths and Harris, "On the variety of special linear systems on a general algebraic curve.

Porteous s Formula for Maps between Coherent Sheaves

A TOOL FOR STABLE REDUCTION OF CURVES ON SURFACES

ON A THEOREM OF CAMPANA AND PĂUN

SESHADRI CONSTANTS ON SURFACES

The moduli stack of vector bundles on a curve

Logarithmic geometry and rational curves

Stable maps and Quot schemes

Moduli of Pointed Curves. G. Casnati, C. Fontanari

FACTORING 3-FOLD FLIPS AND DIVISORIAL CONTRACTIONS TO CURVES

MODULI SPACES OF CURVES

Mini-Course on Moduli Spaces

On a theorem of Ziv Ran

DIVISOR THEORY ON TROPICAL AND LOG SMOOTH CURVES

Compactified Jacobians of nodal curves

Parabolic sheaves, root stacks and the Kato-Nakayama space

Hodge Theory of Maps

REVISITED OSAMU FUJINO. Abstract. The main purpose of this paper is to make C n,n 1, which is the main theorem of [Ka1], more accessible.

FOUNDATIONS OF ALGEBRAIC GEOMETRY CLASS 37

HYPER-KÄHLER FOURFOLDS FIBERED BY ELLIPTIC PRODUCTS

arxiv: v1 [math.ag] 25 Jan 2013

Stable degenerations of symmetric squares of curves

Fiberwise stable bundles on elliptic threefolds with relative Picard number one

Néron Models and Compactified Picard Schemes over the Moduli Stack of Stable Curves

SERRE FINITENESS AND SERRE VANISHING FOR NON-COMMUTATIVE P 1 -BUNDLES ADAM NYMAN

Geometry of the Theta Divisor of a compactified Jacobian

AN INTRODUCTION TO MODULI SPACES OF CURVES CONTENTS

EXCLUDED HOMEOMORPHISM TYPES FOR DUAL COMPLEXES OF SURFACES

Moduli spaces of log del Pezzo pairs and K-stability

arxiv: v1 [math.ag] 16 Aug 2017

Resolution of Singularities in Algebraic Varieties

Logarithmic geometry and moduli

Cohomological Formulation (Lecture 3)

MODULI SPACES AND DEFORMATION THEORY, CLASS 1. Contents 1. Preliminaries 1 2. Motivation for moduli spaces Deeper into that example 4

PARABOLIC SHEAVES ON LOGARITHMIC SCHEMES

SPACES OF RATIONAL CURVES IN COMPLETE INTERSECTIONS

Introduction to Arithmetic Geometry

ON THE MODULI B-DIVISORS OF LC-TRIVIAL FIBRATIONS

ALGEBRAIC HYPERBOLICITY OF THE VERY GENERAL QUINTIC SURFACE IN P 3

0. Introduction 1 0. INTRODUCTION

REVISITED OSAMU FUJINO. Abstract. The main purpose of this paper is to make C n,n 1, which is the main theorem of [Ka1], more accessible.

Existence of coherent systems of rank two and dimension four

A FAMILY OF COMPACTIFIED JACOBIANS SATISFIES AUTODUALITY BECAUSE IT HAS RATIONAL SINGULARITIES

Two constructions of the moduli space of vector bundles on an algebraic curve

ON THE SLOPE_AND KODAIRA DIMENSION OF M g. FOR SMALL g

Three Descriptions of the Cohomology of Bun G (X) (Lecture 4)

Special determinants in higher-rank Brill-Noether theory

On the tautological ring of M g,n

Algebraic Geometry. Andreas Gathmann. Class Notes TU Kaiserslautern 2014

DUALITY SPECTRAL SEQUENCES FOR WEIERSTRASS FIBRATIONS AND APPLICATIONS. 1. Introduction

VARIETIES WITHOUT EXTRA AUTOMORPHISMS I: CURVES BJORN POONEN

The moduli space of curves is rigid

PROBLEMS FOR VIASM MINICOURSE: GEOMETRY OF MODULI SPACES LAST UPDATED: DEC 25, 2013

Towards Fulton s conjecture

TROPICAL METHODS IN THE MODULI THEORY OF ALGEBRAIC CURVES

Weak point property and sections of Picard bundles on a compactified Jacobian over a nodal curve

Algebraic Geometry. Question: What regular polygons can be inscribed in an ellipse?

ON SUBADDITIVITY OF THE LOGARITHMIC KODAIRA DIMENSION

h : P 2[n] P 2(n). The morphism h is birational and gives a crepant desingularization of the symmetric product P 2(n).

Birational geometry for d-critical loci and wall-crossing in Calabi-Yau 3-folds

LINKED ALTERNATING FORMS AND LINKED SYMPLECTIC GRASSMANNIANS

The Canonical Sheaf. Stefano Filipazzi. September 14, 2015

Del Pezzo surfaces and non-commutative geometry

SPACES OF RATIONAL CURVES ON COMPLETE INTERSECTIONS

Non-uniruledness results for spaces of rational curves in hypersurfaces

arxiv: v2 [math.ag] 23 Nov 2013

CANONICAL BUNDLE FORMULA AND VANISHING THEOREM

SECOND FLIP IN THE HASSETT-KEEL PROGRAM: A LOCAL DESCRIPTION

Minimal model program and moduli spaces

Quotient Stacks. Jacob Gross. Lincoln College. 8 March 2018

The Cone Theorem. Stefano Filipazzi. February 10, 2016

FOUNDATIONS OF ALGEBRAIC GEOMETRY CLASSES 43 AND 44

COMPLEX ALGEBRAIC SURFACES CLASS 9

The Affine Grassmannian

LOG MINIMAL MODEL PROGRAM FOR THE MODULI SPACE OF STABLE CURVES: THE SECOND FLIP

MODULI TOPOLOGY. 1. Grothendieck Topology

THE REPRESENTATION THEORY, GEOMETRY, AND COMBINATORICS OF BRANCHED COVERS

where m is the maximal ideal of O X,p. Note that m/m 2 is a vector space. Suppose that we are given a morphism

MODULI OF VECTOR BUNDLES ON CURVES AND GENERALIZED THETA DIVISORS

MAKSYM FEDORCHUK. n ) = z1 d 1 zn d 1.

MA 206 notes: introduction to resolution of singularities

Chern numbers and Hilbert Modular Varieties

Bjorn Poonen. MSRI Introductory Workshop on Rational and Integral Points on Higher-dimensional Varieties. January 17, 2006

The Moduli Space of Rank 2 Vector Bundles on Projective Curves

Pacific Journal of Mathematics

Math 248B. Applications of base change for coherent cohomology

Theta divisors and the Frobenius morphism

Segre classes of tautological bundles on Hilbert schemes of surfaces

SEPARABLE RATIONAL CONNECTEDNESS AND STABILITY

Vanishing theorems and holomorphic forms

David R. Morrison. String Phenomenology 2008 University of Pennsylvania 31 May 2008

Generalized Tian-Todorov theorems

Lecture Notes on Compactified Jacobians

Transcription:

RESEARCH STATEMENT: COMPACTIFYING RELATIVE PICARD SCHEMES ATOSHI CHOWDHURY 1. Introduction 1.1. Moduli spaces and compactification. My research is in the area of algebraic geometry concerned with moduli spaces: geometric spaces whose points classify geometric objects. Moduli spaces are fundamental in all areas of modern geometry, because their geometry encodes the relationships among the objects they classify. For many applications, the usefulness of a moduli space depends on its being compact; that is, it should contain a unique limit for each family of the objects it classifies. For example, intersection theory is generally meaningful only on compact spaces. More broadly speaking, it is valuable to understand limit behavior because limit objects often shed light on the objects of which they re the limits. But many naturally defined moduli spaces are not compact, and so it is of interest to compactify them that is, to assign limit points to families of the objects they classify in a geometrically meaningful way. The most famous example is probably that of the moduli space M g of smooth algebraic curves (or Riemann surfaces) of genus g: this space is not compact, since smooth curves can be continuously deformed to singular curves. The compactification M g of M g, due to Deligne and Mumford in 1969, supplies limits for families of smooth curves by adding to M g a boundary parametrizing certain singular curves (those which are Deligne-Mumford stable). 1.2. Noncompactness of the relative Picard scheme. The moduli spaces in which I m interested are Picard schemes (or Picard stacks), which classify line bundles on algebraic varieties. (Over a curve, the degree-0 component of the Picard scheme is also known as the Jacobian.) If X is a smooth variety, then each connected component of its Picard scheme is automatically compact. However, smooth varieties can be deformed to singular varieties, and such deformations can give rise to noncompactness in the Picard scheme. More precisely, suppose X S is a family of varieties (parametrized by a scheme S), the central fiber X X (the limit of the family) is a singular variety, and L is a line bundle on X\X (which can be viewed as a family of line bundles on the smooth varieties in the family). Then either of the following could occur to prevent the relative Picard scheme of the family (the moduli space parametrizing line bundles on the fibers) from being compact: Obstacle 1 (nonexistence of limits): No line bundle on X is a limit of L (that is, L cannot be extended to a line bundle on X). Obstacle 2 (non-uniqueness of limits, i.e. nonseparatedness): At least two (and possibly infinitely many) non-isomorphic line bundles on X are limits of L. The central goal of my work is to compactify the relative Picard scheme: to construct a modification of it in which Obstacles 1 and 2 do not occur. 1

This problem has been solved in numerous ways over families of curves (see 3.1 for a survey). This has been possible because a great deal is known about degenerations of curves and about the moduli spaces M g. The solutions for curves cannot easily be generalized to varieties of higher dimension, because moduli spaces of smooth varieties of dimension d 2 exhibit much more complicated behavior than M g (for example, unlike M g, they are highly singular (Vakil [24]). In particular, their compactifications are not nearly as well understood as M g : reasonable compactifications of moduli spaces of surfaces (d = 2) were constructed only in the 1980s (e.g. Kollár and Shepherd-Barron [12]), and the d 3 case is still a central area of research. So other methods must be found to compactify the relative Picard scheme over families of higher-dimensional varieties. My work in [4] introduces machinery (a stability condition, Definition 4.1.1) to handle line bundles on varieties of arbitrary dimension; its usefulness is indicated by Theorems 2.2.1-2.2.3 below. 2. Recent results 2.1. Nonseparatedness and reducible varieties. When attempting to compactify relative Picard schemes, there are two natural approaches one can take to overcome Obstacle 1 of 1.2 above (that is, to provide limits for families that have no limits): Approach 1: Over each singular variety X, one can add points to the moduli space to parametrize not just line bundles on X, but certain coherent sheaves (objects that generalize line bundles) on X. Approach 2: Over each singular variety X, one can add points to the moduli space to parametrize line bundles not just on X, but on certain modifications X of X. My work is aimed at a construction using Approach 2. Under this approach, the singular varieties that cause Obstacle 2 (nonseparatedness) to occur are precisely those that are reducible (which means they re formed by gluing together a finite number of irreducible components). Moreover, the nonseparatedness caused by reducibility is severe: over any one-parameter degeneration of smooth varieties to a reducible variety, every family of line bundles has infinitely many limits. Reducible varieties are unavoidable in Approach 2, even if one initially restricts one s attention to a family of varieties whose singular limit X is irreducible. This is because the modifications X of X on which Approach 2 relies are formed by adding extra irreducible components to X, making it reducible. (Specifically, Obstacle 1 occurs for families X S with singular total space. Approach 2 corrects this by replacing each such family by a desingularization of it, which has the effect of adding irreducible components to the singular limit variety while leaving the smooth members of the family unchanged. It is to the desingularized families that we aim to apply Theorem 2.2.1 below, which concerns families with nonsingular total space.) 2.2. Results in [4] and [5]. The paper [4] introduces a stability condition, Definition 4.1.1 below, for line bundles on possibly reducible varieties of arbitrary dimension d. Its purpose is to single out, for each family of line bundles, a unique limit from among the infinitely many that arise as described above from reducibility. 2

This purpose turns out to be fulfilled in a weak sense: Theorems 2.2.1 and 2.2.2 say that in generic numerical circumstances, there is exactly one semistable limit, while in special circumstances, there may be more than one but still only finitely many. In fact this is the best outcome we could have expected, because Definition 4.1.1 specializes to the stability condition given by Caporaso [3] for line bundles on nodal curves, and the latter has the same property. The existence of multiple semistable limits in [3] corresponds to the phenomenon of strict semistability in geometric invariant theory and gives rise to a space that is weakly proper in the sense of Alper [1]. The stability condition of Definition 4.1.1 is defined asymptotically, with respect to a fixed line bundle H on the base variety, so that the semistable limits of a given family of line bundles correspond to solutions of an infinite system of inequalities. (Caporaso s condition in [3] is recovered by taking d = 1 and H = K X. It turns out, however, that for curves the asymptotic aspect of the definition is redundant; hence it is absent from [3].) A priori, this system may have no solutions at all. However, for wise choices of H, it has essentially a unique solution, which yields the theorems below. Theorem 2.2.1 (Chowdhury [4]). Let X S be a one-parameter family of varieties of dimension d, with central fiber X (which may be reducible). Assume that the total space X is nonsingular, and that X has simple normal crossings. Assume furthermore that (1) the dual graph of X (the graph with vertices corresponding to the irreducible components of X, and an edge between a pair of vertices if and only if the corresponding irreducible components have nonempty intersection) is a tree; (2) the canonical bundle K X is positive (resp. negative) in a suitable sense (e.g. it suffices to have K X ample or anti-ample). Then any line bundle L on X\X has a limit on X that is K X-semistable (resp. K + X-semistable). Generically there is a unique such limit; in special cases there may be more than one, but never more than 2 n 1, where n is the number of irreducible components of X. Theorem 2.2.1 is a partial generalization of results over curves given in [3]. The combinatorial condition (1) simultaneously generalizes the conditions defining two useful classes of curves: those of compact type (i.e. having compact Jacobian), and those having only two irreducible components. (It appears, for instance, in [20] as the definition of curves of pseudo-compact type.) The condition (2) for curves says simply that the arithmetic genus of X is not 1. A more surprising theorem, in light of the curve case, is the following. For curves of genus 1 (degenerations of elliptic curves), Definition 4.1.1 (or the stability condition of [3]) breaks down completely: either all infinitely many limits of a family are semistable, or none are, regardless of the choice of the line bundle H with respect to which one defines semistability. But Theorem 2.2.2 shows that in the analogous situation in higher dimension, this does not occur! Semistability with respect to the canonical bundle is no longer useful since the canonical bundle is trivial, but semistability with respect to other line bundles can still work. Specifically, Theorem 2.2.2 concerns K3 surfaces of Type II. These, together with K3 surfaces of Type III, compactify the moduli space of smooth K3 surfaces. K3 surfaces of Type II fulfill the combinatorial condition (1) of Theorem 2.2.1, which 3

makes them amenable to the kind of analysis used for Theorem 2.2.1; K3 surfaces of Type III are more complicated to work with (see 5.3). Theorem 2.2.2 (Chowdhury [5]). Let X be a K3 surface of Type II. Then there is a line bundle H of degree 0 on X such that the following holds: whenever X is the limit of a semistable degeneration X S of K3 surfaces, any line bundle L on X\X has an H + -semistable limit on X. Generically there is a unique such limit; in special cases there may be up to 2 n 1, where n is the number of irreducible components of X. As a final observation, when the line bundles H behave well in families, so does the notion of asymptotic semistability with respect to H. Hence the set of all semistable line bundles forms an algebraic moduli space. The following statement makes this precise when H is the canonical bundle: Theorem 2.2.3 (Chowdhury [4]). The functor that associates to any scheme T the set of families of varieties X T together with line bundles L on X whose restriction to each fiber X t is K X t -semistable defines an Artin stack. 3. History 3.1. Curves. The problem of compactifying Picard schemes over curves, considered as early as the 1950s (e.g. by Igusa [10]), has been quite thoroughly studied. For certain fixed nodal curves, compactified Picard schemes (or Jacobians) were constructed in the 1970s (D Souza [6], Oda and Seshadri [18]); these were followed by compactifications over certain special families of curves (e.g. Ishida [11]). In the 1990s results were established that encompassed far more general types of families: the universal construction of Caporaso [3] already mentioned in 2.2, as well as the fine moduli spaces of Esteves [7]. The more general problem of compactifying moduli spaces of higher-rank vector bundles over curves has been studied with similar success, with compactifications over a fixed nodal curve or specific degenerations to nodal curves (e.g. Gieseker [8], Nagaraj and Seshadri [17]), as well as universal constructions over M g (Pandharipande [21], Schmitt [22]). Many of these constructions rely heavily on geometric invariant theory, although they can be recast in terms of stacks (e.g. see Melo [15]). 3.2. Higher dimension. Over higher-dimensional varieties, the picture is far less complete. Altman and Kleiman [2] constructed compactified Picard schemes for families of higher-dimensional varieties that are irreducible. Degenerations of higherdimensional varieties to reducible varieties, on the other hand, have not been considered. There is, however, a related problem worth mentioning, namely that of compactifying moduli spaces of higher-rank vector bundles on fixed smooth varieties (which, unlike their rank-1 counterparts, are not compact). It was pointed out as early as [18] that the number of irreducible components of a variety and the rank of a vector bundle play analogous roles with respect to non-compactness. The analogy is not directly visible in the the original construction of Gieseker [9] and Maruyama [14], which compactifies the moduli space of stable vector bundles of arbitrary rank on a fixed smooth variety of arbitrary dimension. More suggestive are the recent 4

alternative compactifications given by Markushevich, Tikhomirov, and Trautmann [13] (based on ideas from differential geometry) and by Timofeeva [23] for moduli spaces of vector bundles over a fixed smooth surface. 4. Details of recent work 4.1. Stability condition. The notion of stability used in Theorems 2.2.1-2.2.3 is defined as follows. Definition 4.1.1 (Chowdhury [4]). Let X be a variety of dimension d, and let L and H be line bundles on X. For each union Y X of irreducible components of X, let D = Y X\Y, and let 1 d+1 ( ) d + 1 e Y (L) = d! χ(x, L) [L d+1 j D j 1 Y ] [L d X] χ(y, L). d + 1 j j=1 We say L is H -semistable (resp. H + -semistable) if for every sufficiently large positive integer m, e Y (L + mh) 0 (resp. e Y (L + mh) 0) for every union of irreducible components Y X. (Here, products in brackets are intersection products, while sums of line bundles denote their tensor product, or the sum of the corresponding divisors; see [4] for specifics.) The quantity e Y (L) in Definition 4.1.1 is a polynomial of degree d in L. Its definition is motivated by geometric invariant theory. In particular, for d = 1 (the curve case), Definition 4.1.1 is precisely a GIT stability condition. The same is not clear when d 2, because in that case the GIT analysis is much more difficult to carry out than it is for curves. The virtue of Definition 4.1.1 is that it lets us avoid these difficulties and study the higher-dimensional case directly. The essential property of e Y (L) is the following (stated with some abuse of notation; see [4] for details). Lemma 4.1.2. Let X be the central fiber in a nonsingular one-parameter family of varieties, Y a union of irreducible components of X, Z = X\Y, and D = Y Z. Then e Y (L) = e Z (L + Z). 4.2. Remarks on Theorems 2.2.1 and 2.2.2. Let us point out a few aspects of the proofs of Theorems 2.2.1 and 2.2.2. The condition (1) of Theorem 2.2.1 and the Type II assumption of Theorem 2.2.2 allow one to study varieties with arbitrarily many irreducible components inductively, dealing with two components at a time. The heart of the proof of each theorem is therefore an analysis of the two-component case. For a variety X of dimension d with two irreducible components, Lemma 4.1.2 implies that the semistable limits of a family of line bundles can be read off from the roots of a certain infinite sequence (p m ) of polynomials of degree at most d in a single variable. In particular, there will be a unique semistable limit roughly when there is an essentially unique convergent sequence of the form (r m ), where each r m is a root of p m. In general, one has little control over whether this happens; furthermore, the correspondence between roots and semistable limits is muddied by the possibility 5

that some of the p m have non-real roots, multiple roots, or roots close together. The remarkable thing is that when H is well chosen, these difficulties vanish: it turns out that the sequence (p m ) converges to a linear polynomial, which guarantees the existence and uniqueness of the desired convergent sequence of roots. As a point of comparison, in the curve case (d = 1) as seen in [3], the asymptotic part of Definition 4.1.1 ends up being irrelevant, and the unique semistable limit of a family simply corresponds to the unique root of a certain linear polynomial. Thus the behavior of the curve case reappears in the asymptotic behavior of the higher-dimensional case. 5. Current and future projects 5.1. Proper moduli spaces. As indicated in 2.1, the usefulness of the stability condition, Definition 4.1.1, is in inducing separatedness (uniqueness of limits). To create a proper (i.e. compact) moduli space, it must be deployed together with Approach 2 of 2.1 to guarantee existence of limits. More precisely, a proper moduli space must parametrize, over each variety X, something between the following two sets: (1) the set of all semistable line bundles on X; (2) the set of all semistable line bundles on all modifications (cf. Approach 2) X of X. In the space defined by (1), existence of limits may fail: if X is singular, there will be families of varieties with limit X over which Obstacle 1 of 1.2 occurs. The space defined by (2) may fail to be separated or to be of finite type: the prevention of Obstacle 1 could require the space to include line bundles on infinitely many distinct modifications of X. (The stability condition guarantees that no family of line bundles will have infinitely many limits on any fixed modification of X. But a family could still have a limit on each of infinitely many different modifications of X.) The solution is to understand how limits over different modifications of X are related, and to collapse those that are redundant. This is essentially what happens over curves in [3]: it turns out to suffice to include a finite collection of modifications of each nodal curve X. Question 5.1.1. Does a finite collection of modifications suffice for varieties of higher dimension? This question can be addressed by studying the behavior of the stability condition on the exceptional components in a variety. More precisely, the necessary modifications of a variety X come from blowing up (thus desingularizing) families of varieties degenerating to X. This process adds irreducible components to X which are projective cones over varieties of lower dimension. The special form of these components places restrictions on the values of the quantities e Y in Definition 4.1.1, which in turn point towards an affirmative answer to Question 5.1.1. 5.2. K3 surfaces. Olsson in [19] constructs a compactification of the moduli space of polarized K3 surfaces via a quotient in which nonseparatedness is corrected by the identification of all multiple limits. Theorem 2.2.2 suggests that the stability 6

condition of Definition 4.1.1 could be used to define a section of this quotient; that is, to assign polarizations in a canonical way to degenerations of K3 surfaces. Such a construction would require answers to the following: Question 5.2.1. Does the line bundle H of Theorem 2.2.2 extend to families of K3 surfaces? Question 5.2.2. Does a result like Theorem 2.2.2 hold for K3 surfaces of Type III, or for appropriate modifications thereof? 5.3. Combinatorial improvements. The proofs of Theorems 2.2.1 and 2.2.2 rely heavily on the hypothesis that the underlying variety X is combinatorially simple: there are no cycles of irreducible components (the dual graph is a tree). This means that the theorems do not apply, for example, to K3 surfaces of Type III, or, more generally, to any variety in which three or more components intersect. But varieties like these cannot be avoided altogether in degeneration problems, even after modifications (blowups) such as those used in Approach 2 of 2.1. Question 5.3.1. Can Theorem 2.2.1 be extended to varieties that are more combinatorially complicated, such as those with triple intersections? 5.4. Comparison with other stability conditions. When d = 1 and H = K X, Definition 4.1.1 reduces to the stability condition of Caporaso [3] for line bundles on curves. This in turn determines a moduli space which is isomorphic to the rank-1 version of Pandharipande s space in [21], which parametrizes slope-semistable vector bundles on curves. (As shown in Schmitt [22], the two interpretations diverge for vector bundles of higher rank.) This suggests the following question: Question 5.4.1. Can Definition 4.1.1 be interpreted in terms of slopes of coherent sheaves? Such an interpretation may also make it feasible to address the higher-rank case: Question 5.4.2. Can Definition 4.1.1 be generalized to vector bundles of higher rank? Finally, the proofs of Theorems 2.2.1 and 2.2.2 depend only on a few formal properties possessed by Definition 4.1.1 (notably Lemma 4.1.2). It would be interesting to understand to what extent these properties determine the stability condition: Question 5.4.3. Can one formulate other stability conditions with the same formal properties as Definition 4.1.1? References [1] J. Alper. Good moduli spaces for Artin stacks. Ann. Inst. Fourier 63 (2013), no. 6, 2349-2402. [2] A. Altman and S. Kleiman. Compactifying the Picard scheme. Adv. in Math. 35 (1980), no. 1, 50-112. [3] L. 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. [4] A. Chowdhury. Compactifying the relative Picard functor over degenerations of varieties. Preprint, available at http://math.berkeley.edu/~atoshi, 2014. [5] A. Chowdhury. Stability of line bundles on K3 surfaces of Type II. In preparation. 7

[6] C. D Souza. Compactification of generalized Jacobians. Proc. Indian Acad. Sci. Sect., Math. Sci. 88 (1979), no. 5, 419-457. [7] E. Esteves. Compactifying the relative Jacobian over families of reduced curves. Trans. Amer. Math. Soc. 353 (2001), 3045-3095. [8] D. Gieseker. A degeneration of the moduli space of stable bundles. J. Diff. Geom. 191 (1984), 173-206. [9] D. Gieseker. On the moduli of vector bundles on an algebraic surface. Ann. of Math. (2) 106 (1977), no. 1, 45-60. [10] J. Igusa. Fiber systems of Jacobian varieties. Amer. J. Math. 78 (1956), 171-199. [11] M. Ishida. Compactifications of a family of generalized Jacobian varieties. Proc. Internat. Sympos. Algebraic Geometry, Kyoto, 1977 (M. Nagata, ed.). Kinokuniya, Tokyo (1978), 503-524. [12] J. Kollár, N. Shepherd-Barron. Threefolds and deformations of surface singularities. Invent. Math. 91 (1988), 299-338. [13] D. Markushevich, A.S. Tikhomirov, G. Trautmann. Bubble tree compactification of moduli spaces of vector bundles on surfaces. Cent. Eur. J. Math. 10 (2012), no. 4, 1331-1355. [14] M. Maruyama. Moduli of stable sheaves II. J. Math. Kyoto Univ. 18 (1978), no. 3, 557-614. [15] M. Melo. Compactified Picard stacks over M g. Math. Z. 263 (2008), no. 4, 939-957. [16] D. Mumford. Stability of projective varieties. Enseign. Math. (2) 23 (1977), 39-110. [17] D. Nagaraj, C. Seshadri. Degenerations of the moduli spaces of vector bundles on curves I. Proc. Indian Acad. Sci. Math. Sci. 107 (1997), 101-137. [18] T. Oda, C. Seshadri. Compactifications of the generalized Jacobian variety. Trans. Amer. Math. Soc. 253 (1979), 1-90. [19] M. Olsson. Semistable degenerations and period spaces for polarized K3 surfaces. Duke Math. J. 125 (2004), no. 1, 121-203. [20] B. Osserman. Limit linear series for curves not of compact type. Available at arxiv:1406.6699, 2014. [21] R. Pandharipande. A compactification over M g of the universal moduli space of slopesemistable vector bundles. J. Amer. Math. Soc. 9 (1996), no. 2, 425-471. [22] A. Schmitt. The Hilbert compactification of the universal moduli space of semistable vector bundles over smooth curves. J. Differential Geom. 66 (2004), no. 2, 169-209. [23] N. Timofeeva. On a new compactification of the moduli of vector bundles on a surface. Sbornik: Mathematics 199 (2008), no. 7, 1051-1070. [24] R. Vakil. Murphy s law in algebraic geometry: badly-behaved deformation spaces. Invent. Math. 164 (2006), no. 3, 569-590. 8