Skip to Main Content

Algebraic and Analytic Compactifications of Moduli Spaces

Patricio Gallardo
Matt Kerr

Communicated by Notices Associate Editor Steven Sam

Article cover

The Tour Ahead

The basic objects of algebraic geometry, such as subvarieties of a projective space, are defined by polynomial equations. The seemingly innocuous observation that one can vary the coefficients of these equations leads at once to unexpectedly deep questions:

When are objects with distinct coefficients equivalent?

What types of geometric objects appear if those coefficients move “towards infinity”?

Can we make sense of “equivalence classes at infinity”?

Searching for answers leads to the discovery of moduli spaces and their compactifications, parametrizing equivalence classes of said objects.

The construction of compact moduli spaces and the study of their geometry amounts nowadays to a busy and central neighborhood of algebraic geometry. Any vibrant district in an old city, of course, has too many landmarks to visit, and the first job of a tour guide is to curate a selection of sites and routes — including multiple routes to the same site for the different perspectives they afford. Our tour today has three main stops: elliptic curves, Picard curves (together with “points on a line,” their alter ego), and a brief panoramic glimpse of the general theory.

As for the routes, we first approach the elliptic curve example along the unswerving path that compactifies algebro-geometric moduli spaces with “limiting” algebro-geometric invariants. The way is straight, but entails scaling a brick wall to discover what is meant by “limits.” Our subsequent turn down 19th-century vennels will unveil a connection as old as algebraic geometry itself: associated to our algebraic varieties are analytic periods, producing a period map from our moduli space to a classifying space for periods. While lacking the ideological consistency of the former route, the “limits” of this one are more conceptually straightforward — with Calculus providing a door in the wall.

For the Picard curve example, we reverse the order of these two approaches; this example is important because it is the simplest one where there are multiple natural compactifications of both sorts. Both examples have two very nice features, besides involving objects one can draw on paper. First, the period map is close to being an isomorphism and is inverted by modular forms, an observation going back in the elliptic case to work of Weierstraß. Second, the period map extends to isomorphisms of the various algebraic and analytic compactifications.

These examples will illustrate techniques and methods in moduli theory, preparing the stage for our last stop, high above the city. From there we shall be able to see a vague outline of the modern definition of moduli spaces, as well as various algebro-geometric (GIT, KSBA, K-stable) and Hodge-theoretic (Baily-Borel, toroidal, etc.) compactifications of the same moduli space. The aim of our brief journey is to travel towards understanding their differences — and especially their spectacular coincidences.

1. Elliptic Curves

With a rich history going back to Abel, Jacobi, and Weierstraß in their guise as complex 1-tori (think of the surface of a donut), elliptic curves are central objects in many areas of mathematics, from cryptography to complex analysis. At this first stop on our tour, we’ll use their moduli space to illustrate constructions such as geometric quotients and the period map (and its inversion).

1.1. Algebraic perspective

Our starting point is the fact, first hinted at by Jacobi in 1834, that any complex 1-torus can be realized as a smooth plane cubic — that is, an algebraic curve defined as the zero-locus in of a homogeneous polynomial of degree three in 3 variables

with certain conditions on the coefficients to guarantee the smoothness of . Conversely, for an algebraic geometer, it is natural to approach the set of elliptic curves via a suitable quotient of the set of all such cubics.

Without the smoothness requirement, the coefficients of such equations comprise all ordered 10-tuples of complex numbers , not all zero, defined up to scaling (by ). Since an equation is determined uniquely by its coefficients up to scaling, we conclude that the set of all plane cubics can be identified with the set of complex points of the projective space . (Outside the open set parametrizing smooth cubics, there are different flavors of singular cubics as displayed in Figure 1.)

Figure 1.

Classification of plane cubics.

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{tikzpicture}[scale = 1.5] \draw[thick, dashed] (3.75,0) -- (3.75,-3.75); \draw[thick] [dashed] (0,-3.75) -- (3.75,-3.75); \draw[thick] plot [smooth, tension=3] coordinates { (0,0) (1,-0.5) (0,-1) (1,-1.5)}; \draw[thick] plot [smooth, tension=-5] coordinates { (3,0) (2,-0.75) (3,-1.5) }; \draw[thick] plot [smooth, tension= 0] (4.5,-0.75) circle (0.5cm); \draw[thick] (5,0)--(5,-1.5); \draw[thick] plot [smooth, tension= 0] (0.5,-2.75) circle (0.5cm); \draw[thick] (0.5,-2)--(0.5,-3.5); \draw[thick] (3.0,-2)--(3.0,-3.5); \draw[thick] (1.75,-2.25)--(3.5,-2.25); \draw[thick] (1.75,-2)--(3.5,-3.5); \draw[thick] (4.5,-2)--(4.5,-3.5); \draw[thick] (4.55,-2)--(4.55,-3.5); \draw[thick] (4.0,-2)--(5,-3.55); \draw[thick] plot [smooth, tension=3] coordinates { }; \draw[thick] (0,-4.75) to [out= 10,in= -90] (1,-4); \draw[thick] (0,-4.75) to [out= 10, in=90] (1,-5.5); \draw[thick] (2.5,-4) -- (2.5,-5.5); \draw[thick] (1.75,-4.75) -- (3.25,-4.75); \draw[thick] (1.75,-4) -- (3.25,-5.5); \draw[thick] (4.45,-4) -- (4.45,-5.5); \draw[thick] (4.5,-4) -- (4.5,-5.5); \draw[thick] (4.55,-4) -- (4.55,-5.5); \end{tikzpicture}

The fact that the set of all plane cubics is itself a complex algebraic variety is not a coincidence! Instead, it is our first encounter with one of the most important objects in moduli theory: the Hilbert scheme. Indeed, to keep track of the complex solutions of polynomial equations within projective space, we need to fix an invariant known as the Hilbert polynomial. This polynomial records geometric information about our solutions such as their dimension and degree. It was shown in 1961 by Grothendieck that there exists an algebro-geometric space (that is, a projective scheme) which parametrizes all the closed complex solutions of polynomial equations in with Hilbert polynomial . In our particular example, all plane cubics have Hilbert polynomial equal to , and the associated Hilbert scheme is . Every point in corresponds to a curve with this Hilbert polynomial and vice versa.

At this juncture the reader will point out that is certainly not the sought-for “moduli space of elliptic curves,” because it includes singular cubics. But the open subset parametrizing smooth cubics is not the solution either. The reason is that given an elliptic curve defined by the equation , we can use an invertible linear change of coordinates with , to obtain another equation . The elliptic curve defined by this second equation is isomorphic as a complex variety to the first one, and yet the Hilbert scheme tells us that they are different objects.

The critical observation here is that if we are parametrizing varieties with a fixed Hilbert polynomial, then we want to account for the automorphisms of the ambient projective space. In our example the ambient space of elliptic curves is and the group associated to linear automorphisms has dimension 8. Therefore, of the 9 degrees of freedom associated with the coefficients of the equation , only one is intrinsic to the geometry of the elliptic curve, while the other eight are related to linear changes of coordinates in the ambient space.

The relation between elliptic curves and -orbits is tight; two plane cubics and represent isomorphic elliptic curves (algebraically or complex analytically) if and only if for some . Therefore, the set of elliptic curves up to isomorphism can be identified with -orbits of smooth plane cubics. Although the set of such orbits exists as a topological space, it is not at all obvious that this space is itself a complex variety. We arrive at one of the most delicate problems in algebraic geometry: Given the action of a linear group on a variety , does there exist an algebro-geometric space parametrizing the -orbits in ?

The correct framework for constructing quotients within algebraic geometry is given by Geometric Invariant Theory (GIT), initiated by Mumford in 1969 17. One of the key results from GIT is the existence of an open locus called the stable locus and a well-defined geometric quotient, which in our case is

One of the first exercises in GIT is to show that is the locus parametrizing smooth plane cubics, see 17, Example 7.12. The fact that the quotient is geometric means that every point of corresponds to a unique -orbit of a smooth cubic. Therefore, we arrive at our first (and almost correct!) example of a moduli space: is the “moduli space” of elliptic curves up to isomorphism.

We also arrive at the crux of our problem: is a non-compact variety. Is there a natural compact (projective) variety that contains it and that parametrizes a larger class of algebraic varieties? It will be tempting to consider a naive quotient of by for constructing a compactification of . However, these naive quotients are usually either of the wrong dimension or yield a non-Hausdorff topological space. Here we use a second key result from GIT: There exists a (usually larger) open set , called the semistable locus, containing and admitting a well-defined “categorical quotient.” In our case this is

a complex projective variety compactifying .

To understand the geometry of the above compactification, we recall that parametrizes all possible plane cubics. This includes our large open set parametrizing the smooth plane cubics and smaller loci parametrizing degenerations such as nodal cubics, the union of a conic and a line, etc. It is a non trivial fact that our semistable locus parametrizes all curves with at worst a singularity locally of the form ; these curves are represented at the top left section of Figure 1.

The categorical quotient is not the naive one: the points of are not in bijection with the -orbits of curves parametrized by the semistable locus . Indeed, all of the orbits within the eight-dimensional locus are identified with a single point in even though they represent non-isomorphic curves. However, there is a unique “minimal closed” -orbit associated to the point : namely, the orbit of the “triangle” cubic .

1.2. Analytic perspective

Turning to a complex-analytic perspective, we recall that an elliptic curve can be viewed as the cartesian product of two circles with a complex structure. By considering these two circles we obtain a homology basis (oriented so that ).

Up to scale, there is a unique holomorphic form , with period ratio in the upper-half plane . This is well-defined modulo the action of through fractional-linear transformations induced by changing the homology basis. We claim that the resulting (analytic) invariant captures the (algebraic) isomorphism class of .

This is closely related to the classical theory of modular forms. Here “modular” refers to the moduli of complex 1-tori , or equivalently of the lattice ; and the “forms” are essentially functions on this moduli space (i.e., of the lattice ) with certain transformation properties depending on a weight . The (biperiodic) Weierstraß -function associated to the lattice ,

satisfies , where and are modular forms of weights resp. . That is, they transform by the automorphy factor under pullback by , which makes into a well-defined function. Evidently, the image of the map sending is a Weierstraß cubic

with period ratio .

The key point is that any can be brought into this form (without changing ) through the action of on coordinates. Fix a flex point (i.e., ); since the dual curve has degree , there are 3 more tangent lines passing through . The are collinear, since otherwise one could construct a degree-1 map . So we may choose coordinates to have , , , and , which puts us in the above form 1. In fact, if , then rescaling yields a member of the family

over , whose period map composed with extends to the identity . Hence yields the claimed equivalence of analytic and algebraic “moduli,” and is an isomorphism.

While the choice of does not refine the moduli problem, keeping track of the ordered 2-torsion subgroup does. In 1, this preserves the ordering of the , which are parametrized by the weight-2 modular forms with respect to . The roles of and 2 are played by and the Legendre family , with describing the 6:1 covering . Notice that parametrizes the cross-ratio of 4 (ordered) points on .

For any we can let act on by and take the quotient to produce the universal elliptic curve with level- structure (marked -torsion) over the modular curve . To produce an algebraic realization, we can use Jacobi resp. modular forms to embed then in a suitable projective space. (Indeed, and already did this for and .) The “modular compactification” of so obtained adds points called cusps, over which the elliptic fiber degenerates to a cycle of ’s; in fact, we have . Going around a cusp subjects a basis of integral homology to a transformation conjugate to .

Now consider an algebraic realization of ; e.g., for , the Hesse pencil over has a marked subgroup as base-locus (where the curves meet the coordinate axes). The monodromy group generated by all loops in (acting on of some fiber) is tautologically . So the period ratio yields a well-defined period map inverted by modular forms (as for and ), exchanging algebraic and analytic moduli of smooth objects. The other moral here is that refining the moduli problem (e.g., level structure) produces smaller monodromy group, hence more boundary components (in this case, cusps) in the compactification.

1.3. First spectacular coincidence

In Section 1.1, we used plane cubics and GIT techniques to construct a compactification of the moduli space of smooth elliptic curves . On the other hand, by using periods and modular forms in Section 1.2 we constructed the moduli space of elliptic curves with a level -structure as well as their compactifications with . We recall that a level structure on an elliptic curve is the additional finite information arising from the choice of .

We arrive then to a natural question: Are any of the Hodge theoretic compactifications isomorphic to ? The answer turns out to be yes! From our previous discussion about the invariant it is possible to conclude that

Moreover, isomorphisms of this kind (that is, between Hodge theoretic and geometric compactifications) also exist among other geometric realizations of the elliptic curves. For instance, by keeping track of the ordered 2-torsion subgroup and the Legendre family, one can show that is isomorphic to a GIT compactification of the space of 4-tuples of points in . And with that remark, we turn the corner en route to the next stop on our tour.

2. Picard Curves and Points in a Line

As we begin to dig into this second example (did we mention that the tour includes amateur archaeological activities?), we shall unearth several ideas that are central for constructing compact moduli spaces. They include the use of finite covers to associate periods to varieties without periods, the use of limits of periods to refine compactifications, and the first example of “stable pairs.”

2.1. Analytic perspective

We begin this time with the complex-analytic point of view. Though ordered collections of points in do not themselves have periods, we can consider covers branched over such collections, generalizing the case of Legendre elliptic curves.

Let denote . For , compactifying

to yields a genus-3 curve with cubic automorphism and (up to scale) unique holomorphic differential with . The moduli spaces

and parametrize ordered 5-tuples in (namely and ) resp. unordered 4-tuples in (as roots of the polynomial determined by ).

Fix . To define period maps, we first describe the monodromy group through which acts on . As it sends symplectic bases to symplectic bases and must be compatible with , it should be plausible that this takes the form⁠Footnote1

1

is a lattice in a unitary group of signature , whose elements can be represented by matrices with entries in the Eisenstein integers.

and that for this is replaced by the subgroup . Now given a basis , the period vector () lies in a 2-ball (by Riemann’s bilinear relation), on which acts via

with . So we get period maps and , whose images omit 1 resp. 6 disk-quotients. Writing resp. for

and its -quotient, these maps extend to isomorphisms and .

The Picard modular forms⁠Footnote2 describing their inverses are none other than the and in 3. Indeed, the resulting “modular compactifications” of the ball quotients add only (4 resp. 1) points, extending (say) to . To understand the meaning of , notice that colliding two in 3 and normalizing yields a genus 2 curve with cubic automorphism, whose (single) period ratio is parametrized by one of the disk-quotients previously omitted. When 3 collide in 3, the normalization has genus 0 and thus no moduli, which explains the 4 boundary points in . The unnormalized scenarios are depicted in the left and middle degenerations in Fig. 2.

2

Here comprises holomorphic functions on satisfying for all ; and .

Figure 2.

Degenerating a Picard curve.

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{tikzpicture}[scale=0.6] \draw[thick] plot [tension=0.6,smooth cycle] coordinates {(1.4,4) (1,4.9) (0,5.2) (-1,4.9) (-1.4,4) (-1,3.1) (0,2.8) (1,3.1)}; \draw[thick] (-0.9,4) arc (130:50:0.4); \draw[thick] (-1,4.1) arc (-140:-40:0.5); \draw[thick] (0.1,3.4) arc (130:50:0.4); \draw[thick] (0,3.5) arc (-140:-40:0.5); \draw[thick] (0.1,4.5) arc (130:50:0.4); \draw[thick] (0,4.6) arc (-140:-40:0.5); \draw[thick] plot [tension=0.6,smooth cycle] coordinates {(-2.6,0) (-3,0.9) (-4,1.2) (-5,0.7) (-5.5,0) (-5,-0.7) (-4,-1.2) (-3,-0.9)}; \draw[thick] (-5.5,0) arc (120:54:0.8); \draw[thick] (-5.5,0) arc (-120:-60:1); \draw[thick] (-3.9,-0.6) arc (130:50:0.4); \draw[thick] (-4,-0.5) arc (-140:-40:0.5); \draw[thick] (-3.9,0.5) arc (130:50:0.4); \draw[thick] (-4,0.6) arc (-140:-40:0.5); \filldraw[red] (-5.5,0) circle (2pt); \node at (-4,-1.8) {$t_i=t_j$}; \draw[thick] plot [tension=0.6,smooth cycle] coordinates {(1.4,0) (1,0.9) (0,1.2) (-1,0.7) (-1.5,0) (-1,-0.7) (0,-1.2) (1,-0.9)}; \filldraw[red] (-1.5,0) circle (2pt); \draw[thick] (-1.5,0) arc (135:90:2); \draw[thick] (-1.5,0) arc (-135:-90:2); \draw[thick] (-1.5,0) arc (-90:-38:1.5); \draw[thick] (-1.5,0) arc (90:38:1.5); \node at (0,-1.8) {$t_i=t_j=t_k$}; \draw[blue,thick] (3.5,0) ellipse (1.3cm and 1cm); \draw[blue,thick] (3.2,0) arc (130:50:0.4); \draw[blue,thick] (3.1,0.1) arc (-140:-40:0.5); \filldraw[red] (4.3,0.8) circle (2pt); \filldraw[red] (4.3,-0.8) circle (2pt); \filldraw[red] (4.8,0) circle (2pt); \draw[thick] plot [tension=0.8,smooth] coordinates {(4.3,0.8) (5.5,1.2) (6.5,0) (5.5,-1.2) (4.3,-0.8)}; \draw[thick] plot [smooth] coordinates {(4.3,0.8) (4.9,0.7) (5.3,0.4) (5.2,0.2) (4.8,0)}; \draw[thick] plot [smooth] coordinates {(4.3,-0.8) (4.9,-0.7) (5.3,-0.4) (5.2,-0.2) (4.8,0)}; \node at (4,-1.8) {$t_i=t_j=t_k$}; \node[blue] at (4,-2.6) {with blowup}; \node[red] at (4,0.6) {\small$q_1$}; \node[red] at (4.5,0) {\small$q_2$}; \node[red] at (4,-0.6) {\small$q_3$}; \draw[thick, gray,->,decorate,decoration={snake,amplitude=3pt,pre length=2pt,post length=3pt}] (-1,3) -- (-3,1); \draw[thick, gray,->,decorate,decoration={snake,amplitude=3pt,pre length=2pt,post length=3pt}] (0,2.8) -- (0,1.2); \draw[thick, gray,->,decorate,decoration={snake,amplitude=3pt,pre length=2pt,post length=3pt}] (1,3) -- (3,1); \node[blue] at (2.7,0) {$E$}; \end{tikzpicture}

But this is not the only way to approach the collision of 3 . After a linear change in coordinates, the (2-parameter) degeneration takes the form

in a neighborhood of . Restricting to (, fixed) yields a 1-parameter family over a disk. Blowing up at produces the exceptional divisor

which is an elliptic curve with period ratio . The singular fiber of the blowup is the union of with the normalization () of , glued along (see the rightmost degeneration in Fig. 2), with acting on the lot (and cyclically permuting the ). While the modulus of is just , the ratios of the semiperiods to a period of vary in , and are related by complex multiplication by (i.e., ). In fact, as it turns out that (for some choice of ) blows up, while limits to (say) , a limit which becomes well-defined in .

The upshot is that if we replace the 4 boundary points of by copies of , then these semiperiod ratios extend to an isomorphism from to the resulting . (We have to blow up at the 4 triple-intersection points to make well-defined.) This is a first example of using limiting mixed Hodge structures (here given by the semi-period ratios) to extend period maps to a toroidal compactification (usually written ) refining the Baily-Borel compactification.

2.2. Algebraic perspective

Going back to ordered -tuples of points in , and adopting a geometric viewpoint, we should phrase the moduli problem in terms of objects up to an equivalence relation. An “object” here is an -pointed curve , which is equivalent to if () for some . By considering the open set associated to distinct points, we obtain the quotient

This moduli space is an -dimensional non-compact variety, and every point of it parametrizes a unique configuration of distinct labelled points in up to isomorphism. For , it is the same as above.

Now we describe a new approach to compactify : we expand the set of objects in consideration, so it contains more than just -pointed curves . Indeed, it was discovered in the late 1960s by Grothendieck and later by Knudsen that we can define pairs of a more general sort, called stable -pointed curves of genus . The set of all such stable pairs corresponds to the points of a smooth, compact algebraic variety known as . Moreover, it is the first example of a so-called “fine moduli space” which will be described in §3.1.

This new type of “stable pair,” parametrized by the boundary of this compactification, is a connected but possibly reducible complex curve together with smooth distinct labelled points , …, in , satisfying the following conditions:

has only ordinary double points and every irreducible component of is isomorphic to the projective line .

has arithmetic genus , or equivalently . (Think of a “tree” of ’s.)

On each component of there are at least three points which are either one of the marked points or a double point, i.e., the intersection of two components of .

is a well-behaved compactification. For example, the boundary is a normal crossing divisor with smooth irreducible components.

Let’s consider the case of closely. The moduli space is two-dimensional and isomorphic to the blow-up of at four points in general position. The boundary is equal to the union of irreducible divisors and they are labelled by subsets with . We can explicitly identify these divisors from our blow up construction. Indeed, they correspond to the four exceptional divisors obtained from the points we are blowing up, and the (strict transform of the) lines passing through pairs of such points. Each divisor parametrizes a different type of stable curve; e.g., the divisor generically parametrizes the union of two s with the points distributed as in Fig. 3.

Figure 3.

Generic limit parametrized by .

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{tikzpicture}[scale=0.6] \draw[thick] plot [tension=0.6,smooth cycle] coordinates {(1.4+-3.5,4) (1+-3.5,4.9) (0+-3.5,5.2) (-1+-3.5,4.9) (-1.4+-3.5,4) (-1+-3.5,3.1) (0+-3.5,2.8) (1+-3.5,3.1)}; \node[red] at (0.5+-3.5,4.8) {\small$1$}; \filldraw[red] (0.5+-3.5,4.4) circle (2pt); \node[red] at (-0.6+-3.5,4.7) {\small$3$}; \filldraw[red] (-0.3+-3.5,4.7) circle (2pt); \node[red] at (0.7+-3.5,3.8) {\small$2$}; \filldraw[red] (0.7+-3.5,3.4) circle (2pt); \node[red] at (-0.8 +-3.5,3.9) {\small$4$}; \filldraw[red] (-0.5+-3.5,3.9) circle (2pt); \node[red] at (-0.2 +-3.5,3.2) {\small$5$}; \filldraw[red] (-0.5+-3.5,3.2) circle (2pt); \draw[thick] plot [tension=0.6,smooth cycle] coordinates {(1.4+3.2,4) (1+3.2,4.9) (0+3.2,5.2) (-1+3.2,4.9) (-1.4+3.2,4) (-1+3.2,3.1) (0+3.2,2.8) (1+3.2,3.1)}; \draw[thick, gray,->,decorate,decoration={snake,amplitude=3pt,pre length=2pt,post length=3pt}] (-1.8,4) -- (1.5,4); \node[gray] at (0,3) { }; \draw[thick] plot [tension=0.6,smooth cycle] coordinates {(1.4+6.0,4) (1+6.0,4.9) (0+6.0,5.2) (-1+6.0,4.9) (-1.4+6.0,4) (-1+6.0,3.1) (0+6.0,2.8) (1+6.0,3.1)}; \node[red] at (0.5+6.0,4.8) {\small$1$}; \filldraw[red] (0.5+6.0,4.4) circle (2pt); \node[red] at (-0.3+3.2,4.7) {\small$3$}; \filldraw[red] (-0.0+3.2,4.7) circle (2pt); \node[red] at (0.7+6.0,3.8) {\small$2$}; \filldraw[red] (0.7+6.0,3.4) circle (2pt); \node[red] at (-0.8 +3.2,3.9) {\small$4$}; \filldraw[red] (-0.5+3.2,3.9) circle (2pt); \node[red] at (-0.3 +3.2,3.2) {\small$5$}; \filldraw[red] (-0.0+3.2,3.2) circle (2pt); \filldraw[blue] (1.4+3.2,4) circle (3pt); \end{tikzpicture}

By now our tourists must all be ready to shout: “But we already have (from §1.1) a technique for compactiying moduli spaces! Couldn’t we compactify these spaces by going down the same route as for elliptic curves?” The answer is yes — there are indeed GIT compactifications — but with a new twist. We determined already that is a quotient of an open locus within by . We also mentioned that Geometric Invariant Theory and subsequent developments imply that there is a semistable open locus whose quotient yields a projective variety. However, this semistable locus is not unique! In our particular case, there are finitely many open loci , depending on a collection of rational numbers with and , such that

Furthermore, the categorical quotient

is a projective variety compactifying . The choice of the numbers reflects a more general fact: GIT uses line bundles on the space, here , and characters of the group to construct different semistable loci. A framework known as “variation of GIT” (VGIT), developed by Dolgachev, Hu, and Thaddeus, shows that there is only a finite number of non-isomorphic GIT quotients, and that changing the values of induces birational transformations among them. For example, for each there are choices of that yield and as quotients. In general, it is difficult to determine all possible GIT compactifications. In our particular example (of ), depending on the choice of “weights” , the quotients can be either , , or a blow-up of at points in general position with . Two of these cases, of course, match the compactifications of §2.1.

This plethora of distinct geometric compactifications is a feature of moduli theory. Moreover, the above GIT quotients are philosophically different from . Remember that allows itself to degenerate, so as to keep the points distinct (as in the figure 3). On the other hand, any of the GIT quotients enables the points to collide amongst themselves in a controlled manner, and does not degenerate. This scenario is depicted in Fig. 4.

Figure 4.

GIT limit when .

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{tikzpicture}[scale=0.7] \draw[thick] plot [tension=0.6,smooth cycle] coordinates {(1.4+-3.5,4) (1+-3.5,4.9) (0+-3.5,5.2) (-1+-3.5,4.9) (-1.4+-3.5,4) (-1+-3.5,3.1) (0+-3.5,2.8) (1+-3.5,3.1)}; \node[red] at (0.5+-3.5,4.8) {\small$1$}; \filldraw[red] (0.5+-3.5,4.5) circle (2pt); \node[red] at (-0.6+-3.5,4.7) {\small$3$}; \filldraw[red] (-0.3+-3.5,4.7) circle (2pt); \node[red] at (0.7+-3.5,3.8) {\small$2$}; \filldraw[red] (0.7+-3.5,3.5) circle (2pt); \node[red] at (-0.8 +-3.5,3.9) {\small$4$}; \filldraw[red] (-0.5+-3.5,3.9) circle (2pt); \node[red] at (-0.2 +-3.5,3.2) {\small$5$}; \filldraw[red] (-0.5+-3.5,3.2) circle (2pt); \draw[thick] plot [tension=0.6,smooth cycle] coordinates {(1.4+3.2,4) (1+3.2,4.9) (0+3.2,5.2) (-1+3.2,4.9) (-1.4+3.2,4) (-1+3.2,3.1) (0+3.2,2.8) (1+3.2,3.1)}; \draw[gray,thick,->,decorate,decoration={snake,amplitude=3pt,pre length=2pt,post length=3pt}] (-1.8,4) -- (1.5,4); \node[gray] at (0,3) { }; \node[red] at (-0.3+3.2,4.7) {\small$3$}; \filldraw[red] (-0.0+3.2,4.7) circle (2pt); \node[red] at (0.6+3.2,4.1) {\small$1=2$}; \filldraw[red] (0.7+3.2,3.8) circle (2pt); \node[red] at (-0.8 +3.2,3.9) {\small$4$}; \filldraw[red] (-0.5+3.2,3.9) circle (2pt); \node[red] at (-0.3 +3.2,3.2) {\small$5$}; \filldraw[red] (-0.0+3.2,3.2) circle (2pt); \end{tikzpicture}

And so we arrive at one of the main questions within moduli theory: How are a priori different compactifications of a moduli space related to each other? In our case, it is a theorem of Kapranov that for every as above there is a morphism

whose restriction to is an isomorphism.

2.3. More spectacular coincidences

Let’s stop and look at the road traveled thus far. We learned that given a collection of five ordered points in , it is possible to construct a genus three curve (known as a Picard curve) associated with them. By using a certain eigenspace of its homology we can define a period map that embeds as a dense open subset of . Moreover, this embedding extends to isomorphisms of algebraic and analytic compactifications in two different ways. It would be remarkable if such phenomena persist for other moduli of points .

It turns out that they do. [insert gasp here] Deligne and Mostow showed in 1986, with additional contributions by Doran in 2004, that for with , there exist certain ball quotients such that

where is a certain symmetric group, and is a well-chosen arithmetic group depending on some weights .

Furthermore, the same weights induce GIT compactifications of which are isomorphic to Baily-Borel compactifications of -dimensional ball quotients: that is,

where the adds finitely many points. The isomorphisms 5 compactify period maps 4 associated to cyclic covers of branched in a manner dictated by the configuration of weighted points. In each case, the cyclic automorphism of the covering curve has an eigenspace in with Hodge numbers .

Moreover, it is possible to enrich the above picture. Indeed, by using a slight generalization of the stable pairs described before, we obtain the Hassett moduli space of -pointed rational curves with weights . This compactification of is a smooth projective variety, and it admits morphisms

The Hassett compactifications allow both collisions of points and degenerations of but in a controlled manner depending on the weight .

From the analytic perspective, we also have a unique “toroidal” compactification , discussed at the end of §2.1, that refines the Baily-Borel compactification. (Instead of points, the boundary components are -dimensional, meaning that more information about asymptotic behavior of periods is retained.) Recent work of the authors with L. Schaffler 7 found that there is an isomorphism between and the toroidal compactification. Thus we arrive at the following commutative diagram for :

For a list of these cases see Tables 2 and 3 in 7.

3. A Panoramic View of the Theory

Our tour has arrived at the base of the funicular, on which we now ascend for a theoretical overview.

3.1. What is a moduli space?

If there was really a Temple of Moduli off in the distance, emblazoned on its facade would be some version of: We desire more than a space parametrizing objects, such as smooth elliptic curves up to isomorphism. We seek to understand all well-behaved families of them. A delicate question arises from this ideal. What is meant by adjectives such as “all” and “well-behaved” for a family of algebraic objects? To answer it, we need to reformulate the moduli problem.

Let’s start with a family where is an algebraic variety such as . As a first guess, we might ask that key invariants such as the dimension and the degree — or more precisely, the Hilbert polynomial alluded to in §1.1 — be the same for all fibers over (geometric) points . This intuition turns out to be correct, but does not yield a complete answer. For one thing, we need to consider families over a base which is not an algebraic variety but rather a scheme (which is a more general object). The right well-behavedness condition for families is due to Serre, and it is called flatness. We won’t define it here, but only say that if is an algebraic variety, then flatness is equivalent to the fibers having constant Hilbert polynomial. Part of the reformulated moduli problem, then, is to understand all flat families where is any scheme.

With the above remarks in mind, let be a “reasonable” class of objects — for example, either the stable -pointed curves of genus described in §2 or the smooth elliptic curves of §1. Let be the category of schemes, and let be the category of sets. For every scheme , we consider the set of all flat families . This yields the moduli functor

defined as

Well, that escalated quickly, didn’t it? We started with points in and now we have a functor from the category of all schemes. Here another insight is required: to any variety or scheme , we can associate a functor from to by defining

The functor determines completely. This result (based on Yoneda’s lemma) tells us roughly that to determine a variety or scheme, we just need to understand all maps from other objects to it.

With such ideas in mind, we can return to our moduli problem. We say that a moduli functor is represented by a scheme if there is a natural isomorphism from to . In that event, is called a fine moduli space for , and constructing a family of objects over a base is equivalent to defining a morphism .

We already have seen two examples of a fine moduli space. The smooth algebraic variety presented in §2 represents the moduli functor associated to stable -pointed curves of genus up to isomorphism. Our second example is the Hilbert scheme , which represents the functor

The fact that the Hilbert scheme represents a moduli functor has profound applications in algebraic geometry: Many moduli spaces are constructed by taking GIT quotients of an appropriate Hilbert scheme, as in §1.1.

3.2. A Faustian bargain

Unfortunately, most moduli functors are not represented by a variety or even a scheme. For example, the moduli functor associated with isomorphism classes of smooth elliptic curves fails to be represented by . So we are left with two options.

Our first option is to weaken our expectations. In this case, we look for a scheme that best approximates our moduli functor. still parametrizes all our objects, but maps into it will not parametrize all of their families. This scheme is known as a coarse moduli space. If a moduli functor has a coarse moduli space, the latter is unique (up to canonical isomorphism). For example, the GIT quotient is the coarse moduli space for smooth elliptic curves up to isomorphism.

Our second choice is the Faustian one. We greatly generalize the idea of “geometric space” via categorical tools. Indeed, to keep track of all the families, we need new geometric objects known as stacks. Introduced by Deligne, Mumford, and Artin in the 1970s, they are (loosely speaking) enrichments of schemes obtained by attaching an automorphism group to every point. In our particular context, the stack of objects is a category whose objects are families of our “reasonable” objects, and the morphisms are maps among such families, for details see 18.

Both choices, coarse moduli spaces and stacks are available for a well-behaved moduli problem. For example, when automorphisms of all parametrized objects are finite and the stack is “Hausdorff,” Keel and Mori showed in 1997 that besides the stack representing our moduli functor there is also a coarse moduli space . Recent results by Alper, Halpern-Leistner, and Heinloth have generalized this result to a larger class of stacks, whose parametrized objects can have positive-dimensional automorphism groups.

3.3. Algebraic compactifications

Suppose we are given a moduli problem for which the corresponding (coarse or fine) moduli space is noncompact. To go further, we are faced with another delicate question: Can we define a moduli problem with a larger “reasonable” set of objects such that its associated (coarse or fine) moduli space is compact and contains ? The answer depends on the types of varieties we are parametrizing.

If the varieties parametrized by are “positive” enough (that is, their canonical bundles are ample), then we add degenerations of which have ample canonical bundles and “well-behaved” singularities known as semi-log-canonical (slc) singularities, see 1, Def 1.3.1. Unfortunately, this case does not include many varieties of interest, such as plane cubics.

As a result, it is common to “enrich” our objects to pairs where is a codimension one subvariety (that is, a divisor) such that is ample. Examples of pairs include -pointed curves with . In the presence of , the “correct” new objects for compactifying our moduli problem are called stable pairs: They are degenerations of such that is ample and certain invariants are the same, but we allow for singularities that are at worst slc.

In fact, this approach is already familiar from §2. We began with a moduli problem parametrizing -pointed curves , which was represented by a (noncompact, fine) moduli space . Then, we allowed for more general curves as in Figure 3. The resulting moduli functor was then represented by a compact variety.

The above theory fails when is not ample, so a new perspective is necessary. Here, the concept of -(semi)stability is central. Introduced in 1997 by Tian, it became a leading conjecture — now theorem — that the existence of a (Kähler-Einstein) KE metric on a smooth Fano variety is equivalent to satisfying a K-stability condition defined via the so-called Donaldson-Futaki invariant. There is a local-to-global interplay that restricts the geometry of -semistable varieties. For example, by work of Odaka, a reasonable -semistable Fano surface is irreducible. The construction and explicit description of the moduli space of -semistable varieties constitutes the ongoing work of many people, including Alper, Blum, Halpern-Leistner, Li, Liu, Wang, Xu, and Zhuang, among others. A particular case involving pairs is described by the first author’s work with Martinez-Garcia and Spotti 8.

The reader who wants to go beyond mathematical tourism is referred (in the ample case) to 4 and 10 for curves, 1 and 15 for higher-dimensional cases, and 13 (and references therein) for technical details. For more details about K-stability see 20.

3.4. Analytic compactifications

Our tour has reached the crenellations in the walls above our city — no one said getting into (or out of) Algebraic Geometry was easy! — from which distant vantage §§1.22.1 suggest the outline of a completely different idea for compactifying .

Suppose (i) we have a period map where is a “period domain” like or and a monodromy group. Next, say that (ii) is injective (which is called a “Torelli theorem”), and (iii) is a dense open subset of , as in the examples we saw. Finally, we need that (iv) has a “natural” compactification by “asymptotic period data” — which can mean different things as in §2.1. Then taking the closure of in gives a compactification of ; and if we are lucky and make the right choices then (v) extends to a map from an algebraic compactification to (as in §2.3).

Here we briefly address (i) and (iv). Given a family of smooth projective varieties, we can identify the cohomologies with a fixed -vector space up to the action of monodromy. By the Hodge theorem, each decomposes into a direct sum of subspaces represented by differential forms of type with . This amounts to a decomposition which varies over . More precisely, the flag varies holomorphically over , satisfying the differential condition , and yields what is called a variation of Hodge structure, as first defined by Griffiths.

It also yields a holomorphic map — this is our — into a Hodge domain , modulo the action of monodromy. This domain is an analytic open subset of a generalized flag variety which depends on the Hodge numbers , the (orthogonal or symplectic) intersection form on , and possible additional “symmetries” of the variation. In §1.2 was , while in §2.1 it was . These are both instances of what Hodge theorists call the classical case, where the above differential condition is vacuous and is an algebraic variety.

In the classical case, we can use generalizations of the modular forms encountered above to embed in a projective space. The resulting compactification is called the Baily-Borel compactification. Given a normal-crossing compactification , there is an extension recording the limits of the flag as degenerates. More refined limiting invariants for Hodge flags (limiting mixed Hodge structures), together with a choice of fan, lead to toroidal compactifications , which are typically resolutions of singularities of .

When is not algebraic, the obvious question is “what about the image of the period map?” Using algebraization results in -minimal geometry, Bakker, Brunebarbe, and Tsimerman proved in 2018 that a projective compactification of always exists 3. The construction of Hodge-theoretic completions of , or partial compactifications of that complete , remain areas of active research; cf. the book by Kato and Usui 11 and ongoing work of Green, Griffiths, Laza, and Robles.

3.5. Some final spectacular coincidences

Cubic hypersurfaces in have captivated the imaginations of algebraic geometers since the discovery by Cayley and Salmon (circa 1850) that each smooth one contains exactly 27 lines. They provide another case in which the entire programme (i)–(v) from §3.4 can be worked out.⁠Footnote3 Since the Hodge decomposition on of a smooth cubic surface is trivial, we pass to 3-to-1 cyclic covers of branched along . The associated period map sends the moduli space of cubic surfaces to a four-dimensional ball quotient .

3

An expanded version of this article at arXiv:2107.08316 provides several additional examples and further technical details of the general theory.

Figure 5.

A cubic surface with three singularities. (Image created by Oliver Labs.)

Graphic without alt text

In 2000, Allcock, Toledo, and Carlson showed that extends to an isomorphism between the GIT compactification of and the Baily-Borel compactification of , which adds only one point. The associated cubic surface, defined by , is depicted in Figure 5. A related isomorphism, involving stable pairs and a toroidal compactification, was recently discovered by L. Schaffler and the authors 7.

Acknowledgment

The authors thank the editors and referees for their comments.

References

[1]
Valery Alexeev, Moduli of weighted hyperplane arrangements, Advanced Courses in Mathematics. CRM Barcelona, Birkhäuser/Springer, Basel, 2015. Edited by Gilberto Bini, Martí Lahoz, Emanuele Macrìand Paolo Stellari, DOI 10.1007/978-3-0348-0915-3. MR3380944Show rawAMSref\bib{alexeev2013moduli}{book}{ author={Alexeev, Valery}, title={Moduli of weighted hyperplane arrangements}, series={Advanced Courses in Mathematics. CRM Barcelona}, note={Edited by Gilberto Bini, Mart\'{\i } Lahoz, Emanuele Macr\`\i and Paolo Stellari}, publisher={Birkh\"{a}user/Springer, Basel}, date={2015}, pages={vii+104}, isbn={978-3-0348-0914-6}, isbn={978-3-0348-0915-3}, review={\MR {3380944}}, doi={10.1007/978-3-0348-0915-3}, } Close amsref.
[2]
Avner Ash, David Mumford, Michael Rapoport, and Yung-Sheng Tai, Smooth compactifications of locally symmetric varieties, 2nd ed., Cambridge Mathematical Library, Cambridge University Press, Cambridge, 2010. With the collaboration of Peter Scholze, DOI 10.1017/CBO9780511674693. MR2590897Show rawAMSref\bib{AMRT}{book}{ author={Ash, Avner}, author={Mumford, David}, author={Rapoport, Michael}, author={Tai, Yung-Sheng}, title={Smooth compactifications of locally symmetric varieties}, series={Cambridge Mathematical Library}, edition={2}, note={With the collaboration of Peter Scholze}, publisher={Cambridge University Press, Cambridge}, date={2010}, pages={x+230}, isbn={978-0-521-73955-9}, review={\MR {2590897}}, doi={10.1017/CBO9780511674693}, } Close amsref.
[3]
B. Bakker, Y. Brunebarbe, and J. Tsimerman, o-minimal GAGA and a conjecture of Griffiths, preprint, arXiv:1811.12230.
[4]
Lucia Caporaso, Compactifying moduli spaces, Bull. Amer. Math. Soc. (N.S.) 57 (2020), no. 3, 455–482, DOI 10.1090/bull/1662. MR4108092Show rawAMSref\bib{caporaso}{article}{ author={Caporaso, Lucia}, title={Compactifying moduli spaces}, journal={Bull. Amer. Math. Soc. (N.S.)}, volume={57}, date={2020}, number={3}, pages={455--482}, issn={0273-0979}, review={\MR {4108092}}, doi={10.1090/bull/1662}, } Close amsref.
[5]
James Carlson, Stefan Müller-Stach, and Chris Peters, Period mappings and period domains, Cambridge Studies in Advanced Mathematics, vol. 168, Cambridge University Press, Cambridge, 2017. Second edition of [ MR2012297]. MR3727160Show rawAMSref\bib{CMSP}{book}{ author={Carlson, James}, author={M\"{u}ller-Stach, Stefan}, author={Peters, Chris}, title={Period mappings and period domains}, series={Cambridge Studies in Advanced Mathematics}, volume={168}, note={Second edition of [ MR2012297]}, publisher={Cambridge University Press, Cambridge}, date={2017}, pages={xiv+562}, isbn={978-1-316-63956-6}, isbn={978-1-108-42262-8}, review={\MR {3727160}}, } Close amsref.
[6]
Igor Dolgachev, Lectures on invariant theory, London Mathematical Society Lecture Note Series, vol. 296, Cambridge University Press, Cambridge, 2003, DOI 10.1017/CBO9780511615436. MR2004511Show rawAMSref\bib{GITDolgachev}{book}{ author={Dolgachev, Igor}, title={Lectures on invariant theory}, series={London Mathematical Society Lecture Note Series}, volume={296}, publisher={Cambridge University Press, Cambridge}, date={2003}, pages={xvi+220}, isbn={0-521-52548-9}, review={\MR {2004511}}, doi={10.1017/CBO9780511615436}, } Close amsref.
[7]
Patricio Gallardo, Matt Kerr, and Luca Schaffler, Geometric interpretation of toroidal compactifications of moduli of points in the line and cubic surfaces, Adv. Math. 381 (2021), Paper No. 107632, 48, DOI 10.1016/j.aim.2021.107632. MR4214398Show rawAMSref\bib{GKS}{article}{ author={Gallardo, Patricio}, author={Kerr, Matt}, author={Schaffler, Luca}, title={Geometric interpretation of toroidal compactifications of moduli of points in the line and cubic surfaces}, journal={Adv. Math.}, volume={381}, date={2021}, pages={Paper No. 107632, 48}, issn={0001-8708}, review={\MR {4214398}}, doi={10.1016/j.aim.2021.107632}, } Close amsref.
[8]
Patricio Gallardo, Jesus Martinez-Garcia, and Cristiano Spotti, Applications of the moduli continuity method to log K-stable pairs, J. Lond. Math. Soc. (2) 103 (2021), no. 2, 729–759, DOI 10.1112/jlms.12390. MR4230917Show rawAMSref\bib{GMGS}{article}{ author={Gallardo, Patricio}, author={Martinez-Garcia, Jesus}, author={Spotti, Cristiano}, title={Applications of the moduli continuity method to log K-stable pairs}, journal={J. Lond. Math. Soc. (2)}, volume={103}, date={2021}, number={2}, pages={729--759}, issn={0024-6107}, review={\MR {4230917}}, doi={10.1112/jlms.12390}, } Close amsref.
[9]
Joe Harris and Ian Morrison, Moduli of curves, Graduate Texts in Mathematics, vol. 187, Springer-Verlag, New York, 1998. MR1631825Show rawAMSref\bib{moduliHarris}{book}{ author={Harris, Joe}, author={Morrison, Ian}, title={Moduli of curves}, series={Graduate Texts in Mathematics}, volume={187}, publisher={Springer-Verlag, New York}, date={1998}, pages={xiv+366}, isbn={0-387-98438-0}, isbn={0-387-98429-1}, review={\MR {1631825}}, } Close amsref.
[10]
Brendan Hassett, Classical and minimal models of the moduli space of curves of genus two, Geometric methods in algebra and number theory, Progr. Math., vol. 235, Birkhäuser Boston, Boston, MA, 2005, pp. 169–192, DOI 10.1007/0-8176-4417-2_8. MR2166084Show rawAMSref\bib{hassett}{article}{ author={Hassett, Brendan}, title={Classical and minimal models of the moduli space of curves of genus two}, conference={ title={Geometric methods in algebra and number theory}, }, book={ series={Progr. Math.}, volume={235}, publisher={Birkh\"{a}user Boston, Boston, MA}, }, date={2005}, pages={169--192}, review={\MR {2166084}}, doi={10.1007/0-8176-4417-2\_8}, } Close amsref.
[11]
Kazuya Kato and Sampei Usui, Classifying spaces of degenerating polarized Hodge structures, Annals of Mathematics Studies, vol. 169, Princeton University Press, Princeton, NJ, 2009, DOI 10.1515/9781400837113. MR2465224Show rawAMSref\bib{KU}{book}{ author={Kato, Kazuya}, author={Usui, Sampei}, title={Classifying spaces of degenerating polarized Hodge structures}, series={Annals of Mathematics Studies}, volume={169}, publisher={Princeton University Press, Princeton, NJ}, date={2009}, pages={xii+336}, isbn={978-0-691-13822-0}, review={\MR {2465224}}, doi={10.1515/9781400837113}, } Close amsref.
[12]
Matt Kerr and Gregory Pearlstein, Boundary components of Mumford-Tate domains, Duke Math. J. 165 (2016), no. 4, 661–721, DOI 10.1215/00127094-3167059. MR3474815Show rawAMSref\bib{KP}{article}{ author={Kerr, Matt}, author={Pearlstein, Gregory}, title={Boundary components of Mumford-Tate domains}, journal={Duke Math. J.}, volume={165}, date={2016}, number={4}, pages={661--721}, issn={0012-7094}, review={\MR {3474815}}, doi={10.1215/00127094-3167059}, } Close amsref.
[13]
J. Kollár, Families of varieties of general type, 2010, https://web.math.princeton.edu/~kollar/book/modbook20170720-hyper.pdf.
[14]
János Kollár, Singularities of the minimal model program, Cambridge Tracts in Mathematics, vol. 200, Cambridge University Press, Cambridge, 2013. With a collaboration of Sándor Kovács, DOI 10.1017/CBO9781139547895. MR3057950Show rawAMSref\bib{singularitiesMMP}{book}{ author={Koll\'{a}r, J\'{a}nos}, title={Singularities of the minimal model program}, series={Cambridge Tracts in Mathematics}, volume={200}, note={With a collaboration of S\'{a}ndor Kov\'{a}cs}, publisher={Cambridge University Press, Cambridge}, date={2013}, pages={x+370}, isbn={978-1-107-03534-8}, review={\MR {3057950}}, doi={10.1017/CBO9781139547895}, } Close amsref.
[15]
S. Kovács, Young person’s guide to moduli of higher dimensional varieties, Algebraic geometry—Seattle 2005. Part 2 (2009): 711–743. MR2483953
[16]
Sándor J. Kovács and Zsolt Patakfalvi, Projectivity of the moduli space of stable log-varieties and subadditivity of log-Kodaira dimension, J. Amer. Math. Soc. 30 (2017), no. 4, 959–1021, DOI 10.1090/jams/871. MR3671934Show rawAMSref\bib{KZ}{article}{ author={Kov\'{a}cs, S\'{a}ndor J.}, author={Patakfalvi, Zsolt}, title={Projectivity of the moduli space of stable log-varieties and subadditivity of log-Kodaira dimension}, journal={J. Amer. Math. Soc.}, volume={30}, date={2017}, number={4}, pages={959--1021}, issn={0894-0347}, review={\MR {3671934}}, doi={10.1090/jams/871}, } Close amsref.
[17]
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. MR1304906Show rawAMSref\bib{mumford1994geometric}{book}{ author={Mumford, D.}, author={Fogarty, J.}, author={Kirwan, F.}, title={Geometric invariant theory}, series={Ergebnisse der Mathematik und ihrer Grenzgebiete (2) [Results in Mathematics and Related Areas (2)]}, volume={34}, edition={3}, publisher={Springer-Verlag, Berlin}, date={1994}, pages={xiv+292}, isbn={3-540-56963-4}, review={\MR {1304906}}, doi={10.1007/978-3-642-57916-5}, } Close amsref.
[18]
Martin Olsson, Algebraic spaces and stacks, American Mathematical Society Colloquium Publications, vol. 62, American Mathematical Society, Providence, RI, 2016, DOI 10.1090/coll/062. MR3495343Show rawAMSref\bib{Staks}{book}{ author={Olsson, Martin}, title={Algebraic spaces and stacks}, series={American Mathematical Society Colloquium Publications}, volume={62}, publisher={American Mathematical Society, Providence, RI}, date={2016}, pages={xi+298}, isbn={978-1-4704-2798-6}, review={\MR {3495343}}, doi={10.1090/coll/062}, } Close amsref.
[19]
Orsola Tommasi, The geometry of the moduli space of curves and abelian varieties, Surveys on recent developments in algebraic geometry, Proc. Sympos. Pure Math., vol. 95, Amer. Math. Soc., Providence, RI, 2017, pp. 81–100. MR3727497Show rawAMSref\bib{abelian}{article}{ author={Tommasi, Orsola}, title={The geometry of the moduli space of curves and abelian varieties}, conference={ title={Surveys on recent developments in algebraic geometry}, }, book={ series={Proc. Sympos. Pure Math.}, volume={95}, publisher={Amer. Math. Soc., Providence, RI}, }, date={2017}, pages={81--100}, review={\MR {3727497}}, } Close amsref.
[20]
Chenyang Xu, K-stability of Fano varieties: an algebro-geometric approach, EMS Surv. Math. Sci. 8 (2021), no. 1-2, 265–354, DOI 10.4171/emss/51. MR4307210Show rawAMSref\bib{xu2020k}{article}{ author={Xu, Chenyang}, title={K-stability of Fano varieties: an algebro-geometric approach}, journal={EMS Surv. Math. Sci.}, volume={8}, date={2021}, number={1-2}, pages={265--354}, issn={2308-2151}, review={\MR {4307210}}, doi={10.4171/emss/51}, } Close amsref.

Credits

The opening graphic is courtesy of Marco Rosario Venturini Autieri via Getty.

Figures 14 are courtesy of the authors.

Figure 5 is courtesy of Oliver Labs.

Photo of Patricio Gallardo is courtesy of Branka Vidic.

Photo of Matt Kerr is courtesy of Margaret Milligan Kerr.