<<

spaces. spaces such as the moduli of elliptic (which we discuss below) play a central role Moduli Spaces in a variety of areas that have no immediate link to the geometry being classified, in particular in alge- David D. Ben-Zvi braic number theory and algebraic . Moreover, the study of moduli spaces has benefited tremendously in recent years from interactions with Many of the most important problems in mathemat- (in particular with theory). These inter- ics concern classification. One has a class of math- actions have led to a variety of new questions and new ematical objects and a notion of when two objects techniques. should count as equivalent. It may well be that two equivalent objects look superficially very different, so one wishes to describe them in such a way that equiv- 1 Warmup: The Moduli of Lines in alent objects have the same description and inequiva- the Plane lent objects have different descriptions. Moduli spaces can be thought of as geometric solu- Let us begin with a problem that looks rather simple, tions to geometric classification problems. In this arti- but that nevertheless illustrates many of the impor- cle we shall illustrate some of the key features of mod- tant ideas of moduli spaces. uli spaces, with an emphasis on the moduli spaces Problem. Describe the collection of all lines in the of Riemann surfaces. (Readers unfamiliar with Rie- real plane R2 that pass through the origin. mann surfaces may find it helpful to begin by reading about them in Part III.) In broad terms, a moduli To save writing, we are using the word “line” to mean problem consists of three ingredients. “line that passes through the origin.” This classifica- tion problem is easily solved by assigning to each line Objects: which geometric objects would we like to L an essential parameter, or modulus, a quantity which describe, or parametrize? we can calculate for each line and which will help us Equivalences: when do we identify two of our tell different lines apart. All we have to do is take objects as being isomorphic, or “the same”? standard Cartesian coordinates x, y on the plane and Families: how do we allow our objects to vary, or measure the angle θ(L) between the line L and the x- modulate? axis, taken in counterclockwise fashion. We find that In this article we will discuss what these ingredients the possible values of θ are those for which 0  θ<π, signify, as well as what it means to solve a moduli and that for every such θ there is exactly one line problem, and we will give some indications as to why L that makes an angle of θ with the x-axis. So as a this might be a good thing to do. set, we have a complete solution to our classification Moduli spaces arise throughout algebraic geom- problem: the set of lines L, known as RP1—the real etry, and algebraic projective line—is in one-to-one correspondence with topology. (Moduli spaces in topology are often the half-open interval [0,π). referred to as classifying spaces.) The basic idea is to However, we are seeking a geometric solution to the give a geometric structure to the totality of the objects classification problem. What does this entail? We have we are trying to classify. If we can understand this geo- a natural notion of when two lines are near each other, metric structure, then we obtain powerful insights into which our solution should capture—in other words, the geometry of the objects themselves. Furthermore, the collection of lines has a natural topology.So moduli spaces themselves are rich geometric objects in far, our solution does not reflect the fact that lines their own right. They are “meaningful” spaces, in that L for which the angle θ(L) is close to π are almost any statement about their geometry has a “modular” horizontal: they are therefore close to the x-axis (for interpretation, in terms of the original classification which θ = 0) and to the lines L with θ(L) close to problem. As a result, when one investigates them one zero. We need to find some way of “wrapping round” can often reach much further than one can with other the interval [0,π) so that π becomes close to 0.

1 2 Princeton Companion to Proof

One way to do this is to take not the half-open inter- for instance, if someone tells me they have a contin- val [0,π) but the closed interval [0,π], and then to uous 37-parameter family of lines in R2, this is the “identify” the points 0 and π. (This idea can easily be same as saying that they have a continuous map from made formal by defining an appropriate equivalence R37 to RP1, which sends a point v ∈ R37 to a line relation.) If π and 0 are regarded as the same, then L(v) ∈ RP1. (More concretely, we could say that the numbers close to π are close to numbers close to 0. real function v → θ(L(v)) on R37 is continuous away This is a way of saying that if you attach the two ends from the locus where θ is close to π. Near this locus of a line segment together, then, topologically speak- we could use instead the function φ that measures the ing, you obtain a circle. angle from the y-axis.) A more natural way of achieving the same end is 1.1 Other Families suggested by the following geometric construction of RP1. Consider the unit circle S1 ⊂ R2. To each point The idea of families of lines leads to various other geo- s ∈ S1, there is an obvious way of assigning a line metric structures on the space RP1, and not just its L(s): take the line that passes through s and the ori- topological structure. For example, we have the notion gin. Thus, we have a family of lines parametrized by of a differentiable family of lines in the plane, which S1, that is, a map (or function) s → L(s) that takes is a family of lines for which the angles vary differen- points in S1 to lines in our set RP1. What is important tiably. (The same ideas apply if we replace “differen- ∞ about this is that we already know what it means for tiable” by “measurable,” “C ,” “real analytic,” etc.) two points in S1 to be close to each other, and the To parametrize such a family appropriately, we would 1 map s → L(s) is continuous. However, this map is like RP to be a differentiable , so that we a two-to-one function rather than a bijection, since s can calculate derivatives of functions on it. Such a RP1 and −s always give the same line. To remedy this, we structure on can be specified by using the angle can identify each s in the circle S1 with its antipo- functions θ and φ defined in the previous section. The dal point −s. We then have a one-to-one correspon- function θ gives us a coordinate for lines that are not dence between RP1 and the resulting quotient space too close to the x-axis, and φ gives us a coordinate for (which again is topologically a circle), and this corre- lines that are not too close to the y-axis. We can cal- RP1 spondence is continuous in both directions. culate derivatives of functions on by writing them in terms of these coordinates. One can justify this dif- The key feature of the space RP1, considered as the ferentiable structure on RP1 by checking that for any moduli space of lines in the plane, is that it captures differentiable C ⊂ R2 \ 0 the map c → L(c) the ways in which lines can modulate, or vary contin- comes out as differentiable. This means that if L(c)is uously in families. But when do families of lines arise? not close to the x-axis, then the function x → θ(L(x)) A good example is provided by the following construc- is differentiable at x = c, and similarly for φ and the tion. Whenever we have a continuous curve C ⊂ R2 \0 y-axis. The functions x → θ(L(x)) and → φ(L(x)) are in the plane, we can assign to each point c in C the called pullbacks, because they are the result of convert- line L(c) that passes through 0 and c. This gives us a ing, or “pulling back,” θ and φ from functions defined family of lines parametrized by C. Moreover, the func- on RP1 to functions defined on C. tion that takes c to L(c) is a continuous function from We now can state the fundamental property of RP1 C RP1 to —so the parametrization is a continuous one. as a differentiable space. Suppose, for example, that C is a copy of R real- R2 ized as the set of points (x, 1) at height 1. Then the A differentiable family of lines in parametrized by a map from C to RP1 gives an isomorphism between R differentiable manifold X is the same thing as a func- RP1 and the set {L : θ(L) =0 }, which is the subset of tion from X to , taking a point x to a line L(x), → → RP1 consisting of all lines apart from the x-axis. Put such that the pullbacks x θ(L(x)) and x φ(L(x)) of the functions θ, φ are differentiable functions. more abstractly, we have an intuitive notion of what it means for a collection of lines through the origin We say that RP1 (with its differentiable structure) is to depend continuously on some parameters, and this the moduli space of (differentiably varying families of) notion is captured precisely by the geometry of RP1: lines in R2. This means that RP1 carries the universal Princeton Companion to Mathematics Proof 3 differentiable family of lines. From the very definition, 1.3 Invariants of Families we have assigned to each point of RP1 a line in R2, Associated with any continuous function f from the and these lines vary differentiably as we vary the point. circle S1 to itself is an integer known as its degree. The above assertion says that any differentiable family Roughly speaking, the degree of f is the number of of lines, parametrized by a space X, is described by times f(x) goes round the circle when x goes round giving a map f : X → RP1 and assigning to x ∈ X once. (If it goes backwards n times, then we say that the line L(f(x)). the degree is −n.) Another way to think of the degree is as the number of times a typical point in S1 is passed 1.2 Reformulation: Line Bundles by f(x)asx goes round the circle, where we count this as +1 if it is passed in the counterclockwise direction It is interesting to reformulate the notion of a (contin- and −1 if it is passed in the clockwise direction. uous or differentiable) family of lines as follows. Let Earlier, we showed that the circle S1, which we X be a space and let x → L(x) be an assignment of obtained by identifying the end points of the closed lines to points in X. For each point x ∈ X, we place interval [0,π], could be used to parametrize the mod- a copy of R2 at x; in other words, we consider the uli space RP1 of lines. Combining this with the notion Cartesian product X × R2. We may now visualize the of degree, we can draw some interesting conclusions. In line L(x) as living in the copy of R2 that lies over particular, we can define its so-called “winding num- x. This gives us a continuously varying collection of ber.” Suppose that we are given a continuous function lines L(x) parametrized by x ∈ X, otherwise known γ from the circle S1 into the plane R2 and suppose that as a line bundle over X. Moreover, this line bundle it avoids 0. The image of this map will be a closed loop × R2 is embedded in the “trivial” X , C (which may cross itself). This defines for us a map which is the constant assignment that takes each x to from S1 to itself: first do γ to obtain a point c in C, R2 RP1 the plane . In the case when X is itself, we have then work out L(c), which belongs to RP1, and finally ∈ RP1 a “tautological” line bundle: to each point s , use the parametrization of RP1 to associate with L(c) R2 which we can think of as a line Ls in , it assigns apointinS1 again. The degree of the resulting com- that very same line Ls. posite map will be twice the number of times that γ, and hence C, winds around 0, so half this number is Proposition. For any topological space X there is a defined to be the winding number of γ. natural bijection between the following two sets: More generally, given a family of lines in R2 parametrized by some space X, we would like to mea- (i) the set of continuous functions f : X → RP1; sure the “manner in which X winds around the circle.” and To be precise, given a function φ from X to RP1, which (ii) the set of line bundles on X that are contained defines the parametrized family of lines, we would like × R2 in the trivial vector bundle X . to be able to say, for any map f : S1 → X, what the winding number is of the composition φf, which This bijection sends a function f to the correspond- takes a point x in S1 to its image f(x)inX and from RP1 ing pullback of the tautological line bundle on . there to the corresponding line φ(f(x)) in the fam- That is, the function f is mapped to the line bundle ily. Thus, the map φ gives us a way of assigning to → x Lf(x). (This is a pullback, because it converts L each function f : S1 → X an integer, the winding RP1 from a function defined on to a function defined number of φf. The way this assignment works does on X.) not change if φ is continuously deformed: that is, it Thus, the space RP1 carries the universal line bun- is a topological invariant of φ. What it does depend dle that sits in the trivial R2 bundle—any time we on is the class that φ belongs to in the first cohom- have a line bundle sitting in the trivial R2 bundle, we ology group of X,H1(X, Z). Equivalently, to any can obtain it by pulling back the universal (tautologi- line bundle on a space X which is contained in the cal) example on RP1.(Vector bundles are discussed trivial R2-bundle, we have associated a cohomology in more detail in Part III.) class, known as the Euler class of the bundle. This 4 Princeton Companion to Mathematics Proof is the first example of a characteristic class for vec- In order to speak of a classification problem, let tor bundles. It demonstrates that if we understand alone a moduli space, for Riemann surfaces we must the topology of moduli spaces of classes of geometric next specify when we regard two Riemann surfaces objects, then we can define topological invariants for as equivalent. (We postpone the discussion of the families of those objects. final ingredient, the notion of families of Riemann surfaces, to Section 2.2.) To do this, we must give a notion of isomorphism between Riemann surfaces: 2 The Moduli of Curves and Teichm¨uller when should two Riemann surfaces X and Y be iden- Spaces tified, or thought of as giving two equivalent realiza- tions of the same underlying object of our classifica- We now turn our attention to perhaps the most tion? This issue was hidden in our toy example of famous examples of moduli spaces, the moduli spaces classifying lines in the plane: there we simply iden- of curves, and their first cousins, the Teichm¨uller tified two lines if and only if they were equal as lines spaces. These moduli spaces are the geometric solution in the plane. This naive option is not available to us to the problem of classification of compact Riemann with the more abstractly defined Riemann surfaces. surfaces, and can be thought of as the “higher theory” If we considered Riemann surfaces realized concretely of Riemann surfaces. The moduli spaces are “mean- as subsets of some larger space—for example, as solu- ingful spaces,” in that each of their points stands for tion sets to algebraic equations in complex projective a Riemann . As a result, any statement about space—we could similarly choose to identify surfaces their geometry tells us something about the geometry only if they were equal as subsets. However, this is of Riemann surfaces. too fine a classification for most applications—what We turn first to the objects. Recall that a Riemann we care about is the intrinsic geometry of is a topological surface X (connected and ori- surfaces, and not incidental features that result from ented) to which a complex structure has been given. the particular way we choose to realize them. Complex structures can be described in many ways, At the other extreme, we might choose to ignore and they enable us to do complex analysis, geom- the extra geometric structure that makes a surface etry, and algebra on the surface X. In particular, they into a Riemann surface. That is, we could identify enable us to define holomorphic (complex-analytic) two Riemann surfaces X and Y if they are topologi- and meromorphic functions on open subsets of X. cally equivalent, or homeomorphic (the “coffee mug is To be precise, X is a two-dimensional manifold, but a doughnut” perspective). The classification of com- the charts are thought of as open subsets of C rather pact Riemann surfaces up to topological equivalence than of R, and the maps that glue them together is captured by a single positive integer, the g are required to be holomorphic. An equivalent notion (“number of holes”) of the surface. Any surface of is that of a conformal structure on X, which is the genus zero is homeomorphic to the Riemann structure needed to make it possible to define angles CP1  S2, any surface of genus one is homeomorphic between curves in X. Yet another important equiva- to a torus S1 ×S1, and so on. Thus, in this case there is lent notion is that of algebraic structure on X, mak- no issue of “modulation”—the classification is solved ing X into a complex- (leading to the by giving a list of possible values of a single discrete persistent confusion in terminology: a Riemann sur- invariant. face is two dimensional, and therefore a surface, from However, if we are interested in Riemann surfaces the point of view of topology or the real numbers, as Riemann surfaces rather than simply as topological but one dimensional, and therefore a curve, from the , then this classification is too crude: it com- point of view of complex analysis and algebra). An pletely ignores the complex structure. We would now algebraic structure is what allows us to speak of poly- like to refine our classification to remedy this defect. nomial, rational, or algebraic functions on X, and is To this end, we say that two Riemann surfaces X and usually specified by realizing X as the set of solutions Y are (conformally, or holomorphically) equivalent if to equations in complex there is a topological equivalence between them that CP2 (or CPn). preserves the geometry, i.e., a homeomorphism that Princeton Companion to Mathematics Proof 5 preserves the angles between curves, or takes holo- compact Riemann surface of genus one. More gener- morphic functions to holomorphic functions, or takes ally, we can replace Z2 by any lattice L, regarding z rational functions to rational functions. (These con- and w as equivalent if z − w belongs to L.(Alattice ditions are all equivalent.) Note that we still have at L in C is an additive subgroup of C with two proper- our disposal our discrete invariant—the genus of a sur- ties. First, it is not contained in any line. Second, it is face. However, as we shall see, this invariant is not fine discrete, which means that there is a constant d>0 enough to distinguish between all inequivalent Rie- such that the distance between any two points in L is mann surfaces. In fact, it is possible to have families of at least d. Lattices are also discussed in Section ?? of inequivalent Riemann surfaces that are parametrized The General Goals of Mathematical Research by continuous parameters (but we cannot make proper in Part I. A basis for a lattice L is a pair of complex sense of this idea until we have said precisely what is numbers u and v belonging to L such that every z in meant by a family of Riemann surfaces). Thus, the L can be written in the form au + bv with a and b next step is to fix our discrete invariant and to try to integers. Such a basis will not be unique: for example, classify all the different isomorphism classes or Rie- if L = Z ⊕ Z, then the obvious basis is u = 1 and mann surfaces with the same genus by assembling v = i, but u = 1 and v = 1 + i would do just as well.) them in a natural geometric fashion. If we take a quotient of C by a lattice, then we again An important step towards this classification is the obtain a torus with complex structure. It turns out uniformization theorem. This states that any simply that any compact Riemann surface of genus one can connected Riemann surface is holomorphically iso- be produced in this way. morphic to one of the following three: the Riemann From a topological point of view, any two tori are sphere CP1, the complex plane C, or the upper half- the same, but once we consider the complex structure plane H (equivalently, the unit disc D). Since the uni- we start to find that different choices of lattice may versal covering space of any Riemann surface is lead to different Riemann surfaces. Certain changes to a simply connected Riemann surface, the uniformiza- L do not have an effect: for example, if we multiply tion theorem provides an approach to classifying arbi- a lattice L by some nonzero λ, then trary Riemann surfaces. For instance, any compact the quotient surface C/L will not be affected. That is, Riemann surface of genus zero is simply connected, C/L is naturally isomorphic to C/λL. Therefore, we and in fact homeomorphic to the , so need only worry about the difference between lattices the uniformization theorem already solves our classifi- when one is not a multiple of the other. Geometrically, 1 cation problem in genus zero: up to equivalence, CP this says that one cannot be obtained from the other is the only Riemann surface of genus zero, and so in by a combination of rotation and dilation. this case the topological and conformal classifications Notice that by taking the quotient C/L we obtain agree. not just a “naked” Riemann surface, but one equipped with an “origin”—a distinguished point e ∈ E, which 2.1 Moduli of Elliptic Curves is the image of the origin 0 ∈ C—in other words, an Next, we consider Riemann surfaces whose universal : cover is C, which is the same as saying that they are Definition. An elliptic curve (over C) is a Riemann quotients of C. For example, we can look at a quotient surface E of genus one, equipped with a marked point of C by Z, which means that we regard two complex e ∈ E. Elliptic curves, up to isomorphism, are in bijec- numbers z and w as equivalent if z − w is an integer. tion with lattices L ⊂ C up to rotation. This has the effect of “wrapping C round” into a cylin- der. Cylinders are not compact, but to get a compact Remark. In fact, since L ⊂ C is a subgroup of the surface we could take a quotient by Z2 instead: that Abelian group C, the elliptic curve E = C/L is nat- is, we could regard z and w as equivalent if their dif- urally an Abelian group, with e as its identity ele- ference is of the form a + bi, where a and b are both ment. This is an important motivation for keeping e integers. Now C is wrapped round in two directions as part of the data that defines an elliptic curve. A and the result is a torus with a complex (or, equiv- more subtle reason for remembering the location of e alently, conformal or algebraic) structure. This is a when we speak of E is that it helps us to define E more 6 Princeton Companion to Mathematics Proof

uniquely. This is useful, because any surface E of genus multiply L by a complex number so as to turn ω1 into one has lots of symmetries, or automorphisms: there is 1 and hence ω2 into ω = ω2/ω1. The orientation condi- always a holomorphic automorphism of E taking any tion now says that ω is in the upper half-plane H, i.e., point x to any other given point y. (If we think of E as its imaginary part is positive, Im ω>0. Conversely, a group, these are achieved by translations.) Thus, if any complex number ω ∈ H in the upper half-plane someone hands us another genus-one surface E, there determines a unique oriented lattice L = Z1 ⊕ Zω— may be no way to identify E and E, or there may be that is, the set of all integer combinations a + bω of 1 infinitely many ways: we can always compose a given and ω—and no two of these lattices are related by a isomorphism between them with a self-symmetry of E. rotation. As we will discuss later, automorphisms haunt almost What does this tell us about elliptic curves? We saw every moduli problem, and are crucial when we con- earlier that an elliptic curve is defined by a lattice L sider the behavior of families. It is usually convenient and an identity e. Now we have seen that if we give L to “rigidify” the situation somewhat, so that the pos- some extra structure, namely an oriented basis, then sible isomorphisms between different objects are less we can parametrize it by a complex number ω ∈ H. “floppy” and more uniquely determined. In the case This makes precise for us the “additional structure” of elliptic curves, distinguishing the point e achieves that we want to place on elliptic curves. We say that a this by reducing the symmetry of E. Once we do that, marked elliptic curve is an elliptic curve E,e together there is usually at most one way to identify two elliptic with the choice of an oriented basis ω1,ω2 for the asso- curves (one way, that is, that takes origin to origin). ciated lattice (fundamental group) L of E. The point is that any lattice has infinitely many different bases, We see that Riemann surfaces of genus one (with which lead to many automorphisms of E. By “mark- the choice of a marked point) can be described by con- ing” one of these bases, we stop them being automor- crete “linear algebra data”: a lattice L ⊂ C, or rather phisms. the equivalence class consisting of all nonzero scalar multiples λL of L. This is the ideal setting to study 2.2 Families and Teichm¨uller Space a classification, or moduli, problem. The next step is to find an explicit parametrization of the collection With our new definition, we can summarize the earlier of all lattices, up to multiplication, and to decide in discussion by saying that marked elliptic curves are in what sense we have obtained a geometric solution to bijection with points ω ∈ H of the upper half-plane. the classification problem. The upper half-plane is, however, much more than just In order to parametrize the collection of lattices, we a set of points: it carries a host of geometric structures, follow a procedure used for all moduli problems: first in particular a topology and a complex structure. In parametrize lattices together with the choice of some what sense do these structures reflect geometric prop- additional structure, and then see what happens when erties of marked elliptic curves? In other words, in we forget this choice. For every lattice L we choose a what sense is the H, known in this basis ω1,ω2 ∈ L: that is, we represent L as the set context as the Teichm¨uller space T1,1 of genus-one of all integer combinations aω1 + bω2. We do this in Riemann surfaces with one marked point, a geometric an oriented fashion: we require that the fundamental solution to the problem of classifying marked elliptic parallelogram spanned by ω1 and ω2 is positively ori- curves? ented. (That is, the numbers 0, ω1, ω1 +ω2, and ω2 list In order to answer this question, we need the notion the vertices of the parallelogram in an counterclock- of a continuous family of Riemann surfaces, and also wise order. From the geometric point of view of the the notion of a complex-analytic family. A continuous elliptic curve E, L is the fundamental group of E, family of Riemann surfaces parametrized by a topo- and the orientation condition says that we generate L logical space S—for example, by the circle S1—is the by two loops, or “meridians,” A = ω1, B = ω2, which assignment of a Riemann surface Xs to every point s are oriented, in that their oriented intersection num- of S, which “varies continuously.” In our example of ber A∩B is equal to +1 rather than −1.) Since we are the moduli of lines in the plane, a continuous family of interested in lattices only up to multiplication, we can lines was characterized by the property that the angles Princeton Companion to Mathematics Proof 7 between the lines and the x-ory-axes defined continu- a single point s = S is indeed just a single Riemann ous functions of the parameters. Geometrically defined surface Xs. Just as in this simple case we wish to collections of lines, such as those produced by a curve consider Riemann surfaces only up to equivalence, so C in the plane, then gave rise to continuous families. there is a notion of equivalence or isomorphism of two  More abstractly, a continuous family of lines defined a analytic families {Xs} and {Xs} parametrized by the line bundle over the parameter space. A good criterion same space S. We simply regard the families as equiva-  for a family of Riemann surfaces is likewise that any lent if the surfaces Xs and Xs are isomorphic for every “reasonably defined” geometric quantity that we can s, and if the isomorphism depends analytically on s. calculate for every Riemann surface should vary con- Armed with the notion of family, we can now formu- tinuously in the family. For example, a classical con- late the characteristic property that the upper half- struction of Riemann surfaces of genus g comes from plane possesses when we think of it as the moduli taking 4g-gons and gluing opposite sides together. The space of marked elliptic curves. We define a contin- resulting Riemann surface is fully determined by the uous or analytic family of marked elliptic curves to edge-lengths and angles of the polygon. Therefore, a be a family where the underlying genus-one surfaces continuous family of Riemann surfaces described in vary continuously or analytically, while the choice of this fashion should be precisely a family such that the basepoint e ∈ E and the basis of the lattice L vary edge-lengths and angles give continuous functions of s s s continuously. the parameter set. H In more abstract topological terms, if we have a col- The upper half-plane plays a role for marked ellip- tic curves that is similar to the role played by RP1 for lection {Xs,s∈ S} of Riemann surfaces depending on lines in the plane. The following theorem makes this points in a space S and we wish to make it into a con- statement precise. tinuous family, then we should give the union s∈S Xs itself the structure of a topological space X , which Theorem. For any topological space S, there is a one- should simultaneously extend the topology on each to-one correspondence between continuous maps from individual X . The result is called a Riemann surface s H bundle. Associated with X is the map that takes each S to and isomorphism classes of continuous families of marked elliptic curves parametrized by S. Similarly, point x to the particular s for which x belongs to Xs. We should demand that this map is continuous, and there is a one-to-one correspondence between analytic H perhaps more (it could be a fibration, or fiber bundle). maps from any complex manifold S to and isomor- This definition has the advantage of great flexibility. phism classes of analytic families of marked elliptic For example, if S is a complex manifold, then in just curves parametrized by S. the same way we can speak of a complex-analytic fam- If we apply the theorem in the case where S is a sin- ily of Riemann surfaces {Xs,s∈ S} parametrized by gle point, it simply tells us that the points of H are in S: now we ask for the union of the Xs to carry not just a topology but a complex structure (i.e., it should form bijection with the isomorphism classes of marked ellip- a complex manifold), extending the complex struc- tic curves, as we already knew. However, it contains H ture on the fibers and mapping holomorphically to more information: it says that , with its topology and the parameter set. The same holds with “complex- complex structure, embodies the structure of marked analytic” replaced by “algebraic.” These abstract def- elliptic curves and the ways in which they can modu- initions have the property that if our Riemann surfaces late. At the other extreme, we could take S = H itself, are described in a concrete way—cut out by equa- mapping S to H by the identity map. This expresses tions, glued from coordinate patches, etc.—then the the fact that H itself carries a family of marked elliptic coefficients of our equations or gluing data will vary curves, i.e., the collection of Riemann surfaces defined as complex-analytic functions in our family precisely by ω ∈ H fit together into a complex manifold fibering when the family is complex analytic (and likewise for over H with elliptic curve fibers. This family is called continuous or algebraic families). the universal family, since by the theorem any family As a reality check, note that a (continuous, analytic, is “deduced” (or pulled back) from this one universal or other) family of Riemann surfaces parametrized by example. 8 Princeton Companion to Mathematics Proof

2.3 From Teichm¨uller Space to Moduli be given a complex-analytic structure, which, it turns Space out, identifies it with the complex plane C itself. To see this one uses the classical modular function j(z), We have arrived at a complete and satisfying picture a complex-analytic function on H which is invariant for the classification of elliptic curves when we choose under the modular group PSL Z and which therefore in addition a marking (that is, an oriented basis of 2 defines a natural coordinate H/ PSL Z → C. the associated lattice L = π (E)). What can we say 2 1 It appears that we have solved the moduli problem about elliptic curves themselves, without the choice of for elliptic curves: we have a topological, and even marking? We somehow need to “forget” the marking, complex-analytic, space M = H/ PSL Z whose by regarding two points of H as equivalent if they cor- 1,1 2 points are in one-to-one correspondence with isomor- respond to two different markings of the same elliptic phism classes of elliptic curves. This already qualifies curve. M as the coarse moduli space for elliptic curves, Now, given any two bases of the group (or lattice) 1,1 which means it is as good a moduli space as we can Z ⊕ Z, there is an invertible 2 × 2 matrix with integer hope for. However, M fails an important test for entries that takes one basis to the other. If the two 1,1 a moduli space that T passed (as we saw in Sec- bases are oriented, then this matrix will have determi- 1,1 tion 2.2): it is not true, even for the circle S = S1, nant 1, which means that it is an element   that every continuous family of elliptic curves over S ab corresponds to a map from S to M1,1. A = ∈ SL2 Z cd The reason for this failure is the problem of auto- morphisms. These are equivalences from E to itself, of the group of invertible unimodular matrices over Z. that is, complex-analytic maps from E to E that pre- Similarly, given any two oriented bases (ω ,ω ) and 1 2 serve the basepoint e. Equivalently, they are given by (ω ,ω ) of a lattice L, which can be thought of as 1 2 complex-analytic self-maps of C that preserve 0 and oriented identifications of L with Z ⊕ Z, there is a the lattice L. Such a map must be a rotation: that is, matrix A ∈ SL Z such that ω = aω + bω and ω = 2 1 1 2 2 multiplication by some complex number λ of modu- cω + dω . If we now consider the normalized bases 1 2 lus 1. It is easy to check that for most lattices L in (1,ω) and (1,ω), where ω = ω /ω and ω = ω /ω , 1 2 1 2 the plane, the only rotation that sends L to itself is then we obtain a transformation of the upper half- multiplication by λ = −1. Note that this is the same plane. It is given by the formula −1 that we quotiented out by to pass from SL2 Z to  aω + b Z ω = . PSL2 . However, there are two special lattices that cω + d have greater symmetry. These are the square lattice

That is, the group SL2 Z is acting on the upper half- L = Z · 1 ⊕ Z · i, corresponding to the fourth root of plane by linear fractional (or M¨obius) transformations unity i, and the hexagonal lattice L = Z · 1 ⊕ Z · e2πi/6, with integer coefficients, and two points in the upper corresponding to a sixth root of unity. (Note that half-plane correspond to the same elliptic curve if one the hexagonal lattice is also represented by the point can be turned into the other by means of such a trans- ω =e2πi/3.) The square lattice, which corresponds to formation. If this is the case, then we should regard the elliptic curve formed by gluing the opposite sides the two points as equivalent: that is how we formalize of a square, has as its symmetries the group Z/4Z of the idea of “forgetting” the marking. Note also that rotational symmetries of the square. The hexagonal the scalar matrix − Id in SL2 Z, which negates both ω1 lattice, which corresponds to the elliptic curve formed and ω2, acts trivially on the upper half-plane, so that by gluing the opposite sides of a regular hexagon, has we in fact get an action of PSL2 Z =SL2 Z/{± Id} on as its symmetries the group Z/6Z of rotational sym- H. metries of a hexagon. So we come to the conclusion that elliptic curves We see that the number of automorphisms of an (up to isomorphism) are in bijection with orbits of elliptic curve jumps discontinuously at the special 2πi/6 PSL2 Z on the upper half-plane, or equivalently with points ω = i and ω =e . This already suggests points of the quotient space H/ PSL2 Z. This quotient that something might be wrong with M1,1 asamod- space has a natural quotient topology, and in fact can uli space. Note that we avoided this problem with Princeton Companion to Mathematics Proof 9

1 1 the moduli T1,1 of marked elliptic curves, since there the trivial family S ×Ei of elliptic curves over S , that are no automorphisms of an elliptic curve that also is, the family where every curve is equal to E(i) but preserve the marking. Another place we might have the curves do not twist as we go around! Thus, there observed this problem with M1,1 is when we passed are more families of elliptic curves than there are maps to the quotient H/ PSL2 Z. We avoided the automor- to M1,1—the quotient space H/ PSL2 Z cannot handle phism λ = −1 by quotienting by PSL2 Z rather than the complications caused by automorphisms. A vari- 2πi/6 SL2 Z. However, the two special points i and e ant of this construction applies to complex-analytic are preserved by integer M¨obius transformations of H families with S1 replaced by C×. This is a very general other than the identity, and they are the only points phenomenon in moduli problems: whenever objects with that property. This means that the quotient have nontrivial automorphisms, we can imitate the

H/ PSL2 Z naturally comes with conical singularities construction above to get nontrivial families over an at the points corresponding to these two orbits: one interesting parameter set, all of whose members are looks like a cone with angle π, and the other like a cone the same. As a result, they cannot be classified by a 2 with angle 3 π. (To see why this is plausible, imag- map to the set of all isomorphism classes. ine the following simpler instance of the same phe- What do we do about this problem? One approach nomenon. If for every complex number z you identify is to resign ourselves to having coarse moduli spaces, z with −z, then the result is to wrap the complex plane which have the right points and right geometry but do round into a cone with a singularity at 0. The reason not quite classify arbitrary families. Another approach

0 is singled out is that it is preserved by the transfor- is the one that leads to T1,1: we can fix markings of mation z →−z. Here the angle would be π because one kind or another, which “kill” all automorphisms. the identification of points is two-to-one away from In other words, we choose enough extra structure on the singularity and π is half of 2π.) It is possible to our objects so that there do not remain any (nontriv- massage these singularities away using the j-function, ial) automorphisms that preserve all this decoration. but they are indicating a basic difficulty. In fact, one can be far more economical than picking So why do automorphisms form an obstacle to the a basis of the lattice L and obtaining the infinite cov- existence of “good” moduli spaces? We can demon- ering T1,1 of M1,1: one can fix a basis of L only up to strate the difficulty by considering an interesting con- some congruence (for example, of L/2L). Finally, we tinuous family of marked elliptic curves parametrized can simply learn to come to terms with the automor- by the circle S = S1. Let E(i) be the “square” elliptic phisms, keeping them as part of the data, resulting curve that we considered earlier, based on the lattice in “spaces” where points have internal symmetries. of integer combinations of 1 and i. Next, for every t This is the notion of an ,orstack, which between 0 and 1, let Et beacopyofE(i). Thus, we is flexible enough to deal with essentially all moduli have taken the constant, or “trivial,” family of elliptic problems. curves over the closed unit interval [0, 1], where every curve in the family is E(i). Now we identify the ellip- tic curves at the two ends of this family, not in the 3 Higher-Genus Moduli Spaces and obvious way, but by using the automorphism given by Teichm¨uller Spaces a90◦ rotation, or multiplication by i. This means that we are looking at the family of elliptic curves over the We would now like to generalize as much as possi- circle where each member of the family is a copy of ble of the picture of elliptic curves and their mod- the elliptic curve E(i), but these copies twist by 90◦ uli to higher-genus Riemann surfaces. For each g we as we go around the circle. would like to define a space Mg, called the mod- It is easy to see that there is no way to capture this uli space of curves of genus g, that classifies com- family of elliptic curves by means of a map from S1 to pact Riemann surfaces of genus g and tells us how the space M1,1. Since all of the members of the family they modulate. Thus, the points of Mg should cor- are isomorphic, each point of the circle should map to respond to our objects, compact Riemann surfaces of the same point in M1,1 (the equivalence class of i in genus g, or, to be more accurate, equivalence classes 1 H). But the constant map S →{i}∈M1,1 classifies of such surfaces, where two surfaces are considered 10 Princeton Companion to Mathematics Proof to be equivalent if there is a complex-analytic iso- ing of their fundamental group. This means that our morphism between them. In addition, we would like objects are genus-g surfaces X plus a set of generators

Mg to do the best it can to embody the structure Ai, Bi of π1(X), which give an isomorphism between 1 of continuous families of genus-g surfaces. Likewise, π1(X) and Γg, up to conjugation. Our equivalences there are spaces Mg,n parametrizing “n-punctured” are complex-analytic maps that preserve the mark- Riemann surfaces of genus g. This means we consider ings. Finally, our continuous (respectively, complex- not “bare” Riemann surfaces, but Riemann surfaces analytic) families are continuous (complex-analytic) together with a “decoration” or “marking” by n dis- families of Riemann surfaces with continuously vary- tinct labeled points (punctures). Two of these are con- ing markings of the fundamental group. In other sidered to be equivalent if there is a complex-analytic words, we are asserting the existence of a topological isomorphism between them that takes punctures to space/complex manifold Tg, with a complex-analytic punctures and preserves labels. Since there are Rie- family of marked Riemann surfaces over it, and the mann surfaces with automorphisms, we do not expect following strong property. M to be able to classify all families of Riemann sur- g The characteristic property of T . For any topo- faces: that is, we will expect examples similar to the g logical space (respectively, complex manifold) S, there twisted square-lattice construction discussed earlier. is a bijection between continuous maps (respectively, However, if we consider Riemann surfaces with enough holomorphic maps) S →T and isomorphism classes extra markings, then we will be able to obtain a mod- g of continuous (respectively, complex-analytic) families uli space in the strongest sense. One way to choose of marked genus-g surfaces parametrized by S. such markings is to consider Mg,n with n large enough (for fixed g). Another approach will be to mark gen- 3.1 Digression: “Abstract Nonsense” erators of the fundamental group, leading to the It is interesting to note that, while we have yet to see Teichm¨uller spaces Tg and Tg,n. We now outline this process. why such a space exists, it follows from general, non- geometric principles—category theory or “abstract To construct the space Mg, we return to the uni- formization theorem. Any compact surface X of genus nonsense”—that it is completely and uniquely deter- g>1 has as its universal cover the upper half-plane H, mined, both as a topological space and as a com- so it is represented as a quotient X = H/Γ , where Γ plex manifold, by this characteristic property. In a is a representation of the fundamental group of X as a very abstract way, every topological space M can be subgroup of conformal self-maps of H. The group of all uniquely reconstructed from its set of points, the set of paths between these points, the set of surfaces span- conformal automorphisms of H is PSL2 R, the group of linear fractional transformations with real coeffi- ning these paths, and so on. To put it differently, we cients. The fundamental groups of all compact genus- can think of M as a “machine” that assigns to any g Riemann surfaces are isomorphic to a fixed abstract topological space S the set of continuous maps from S to M. This machine is known as the “functor of points group Γg, with 2g generators Ai, Bi (i =1,...,g) and one relation: that the product of all commutators of M.” Similarly, a complex manifold M provides a −1 −1 ⊂ R machine that assigns to any other complex manifold S AiBiAi Bi is the identity. A subgroup Γ PSL2 that acts on H in such a way that the quotient H/Γ the set of complex-analytic maps from S to M. A curi- is a Riemann surface (technically, the action should ous discovery of category theory (the Yoneda lemma) have no fixed points and should be properly discon- is that for very general reasons (having nothing to do tinuous) is known as a Fuchsian group. Thus, the with geometry), these machines (or functors) uniquely analog of the representation of elliptic curves by lat- determine M as a space, or a complex manifold. tices L  Z ⊕ Z in the plane is the representation of H higher-genus Riemann surfaces as /Γ , where Γ is a 1. Note that while the fundamental group of X depends on the Fuchsian group. choice of a basepoint, π1(X, x) and π1(X, y) may be identified by choosing a path from x to y, and the different choices are related The Teichm¨uller space Tg of genus-g Riemann sur- by conjugation by a loop. Thus, if we are willing to identify sets of faces is the space that solves the moduli problem generators Ai, Bi when they differ only by a conjugation, then we for genus-g surfaces, when they come with a mark- can ignore the choice of a basepoint. Princeton Companion to Mathematics Proof 11

Any moduli problem in the sense we have described achieved by the action of a group, the genus-g mapping

(giving objects, equivalences, and families) also gives class group MCGg or Teichm¨uller modular group,on such a machine, where to S we assign the set of all Tg, which generalizes the modular group PSL2 Z that families over S, up to isomorphism. So just by set- acts on H = T1,1. (The is defined ting up the moduli problem we have already uniquely as the group of all self-diffeomorphisms of a genus-g determined the topology and complex structure on surface—remember that all such surfaces are topolog- Teichm¨uller space. The interesting part then is to ically the same—modulo those diffeomorphisms that know whether or not there actually exists a space giv- act trivially on the fundamental group.) As in the ing rise to the same machine we have constructed, case of elliptic curves, Riemann surfaces with auto- whether we can construct it explicitly, and whether we morphisms correspond to points in Tg fixed by some can use its geometry to learn interesting facts about subgroup of MCGg, and give rise to singular points in

Riemann surfaces. the quotient Mg = Tg/ MCGg. Representation varieties, or moduli spaces of rep- 3.2 Moduli Spaces and Representations resentations, are an important and concrete class of moduli spaces that arise throughout geometry, topol- Coming back to earth, we discover that we have a ogy, and number theory. Given any (discrete) group fairly concrete model of Teichm¨uller space at our dis- Γ , we ask (for example) for a space that parametrizes posal. Once we have fixed the marking π1(X)  Γg, homomorphisms of Γ into the group of n × n matrices. we are simply looking at all ways to represent Γg as The notion of equivalence is given by conjugation by a Fuchsian subgroup of PSL2 R. Ignoring the Fuch- GL , and that of families by continuous (or analytic, sian condition for a moment, this means finding 2g n or algebraic, etc.) families of matrices. This problem real matrices (up to ± Id) Ai,Bi ∈ PSL2 R satisfying is interesting even when the group Γ is Z. Then we the commutator relation of Γg. This gives an explicit are simply considering invertible n × n matrices (the set of (algebraic!) equations for the entries of the 2g image of 1 ∈ Z) up to conjugacy. It turns out that matrices, which determine the space of all represen- there is no moduli space for this problem, even in the tations Γg → PSL2 R. We must now quotient out by coarse sense, unless we consider only “nice enough” the action of PSL2 R that simultaneously conjugates matrices: for example, matrices that consist of only a all 2g matrices to obtain the representation variety single Jordan block. This is a good example of a ubiq- Rep(Γg, PSL2 R). This is analogous to considering lat- uitous phenomenon in moduli problems: one is often tices in C up to rotation, and is motivated by the fact forced to throw out some “bad” (unstable) objects in that the quotients of H by two conjugate subgroups of order to have any chance of obtaining a moduli space. PSL2 R will be isomorphic. (See the paper by Mumford and Suominen (1972) for Once we have described the space of all represen- a detailed discussion.) tations of Γg into PSL2 R, we can then single out Teichm¨uller space as the subset of the representation 3.3 Moduli Spaces and Jacobians variety that consists of Fuchsian representations of Γg into PSL2 R. Luckily this subset is open in the repre- The upper half-plane H = T1,1, together with the sentation variety, which gives a nice realization of Tg action of PSL2 Z, gives an appealingly complete pic- as a topological space—in fact, Tg is homeomorphic ture of the moduli problem for elliptic curves and its to R6g−6. (This can be seen very explicitly in terms of geometry. The same cannot be said, unfortunately, for the Fenchel–Nielsen coordinates, which parametrize a the picture of Tg as an open subset of the representa- surface in Tg via a cut-and-paste procedure involving tion variety. In particular, the representation variety 3g − 3 lengths and 3g − 3 angles.) We may now try to does not even carry a natural complex structure, so ∼ “forget” the marking π1(X) = Γg, to obtain the mod- we cannot see from this description the geometry of uli space Mg of unmarked Riemann surfaces. In other Tg as a complex manifold. This failure reflects some words, we would like to take Tg and identify any two of the ways in which the study of moduli spaces is points that represent the same underlying Riemann more complicated for genus greater than one. In par- surface with different markings. This identification is ticular, the moduli spaces of higher-genus surfaces are 12 Princeton Companion to Mathematics Proof not described purely by linear algebra plus data about Invariants from moduli spaces. An important orientation, as is the case in genus one. application of moduli spaces in geometry and topology Part of the blame for this complexity lies with the is inspired by quantum field theory, where a particle, fact that the fundamental group Γg  π1(X)(g>1) rather than follow the “best” classical path between is no longer Abelian, and in particular it is no longer two points, follows all paths with varying probabilities. equal to the first homology group H1(X, Z). A related Classically, one calculates many topological invariants problem is that X is no longer a group. A beauti- by picking a geometric structure (such as a metric) on ful solution to this problem is given by the construc- a space, calculating some quantity using this struc- tion of the Jacobian Jac(X), which shares with ellip- ture, and finally proving that the result of the calcu- tic curves the properties of being a torus (homeomor- lation did not depend on the structure we chose. The phic to (S1)2g), an Abelian group, and a complex new alternative is to look at all such geometric struc- (in fact complex-algebraic) manifold. (The Jacobian tures, and integrate some quantity over the space of all of an elliptic curve is the elliptic curve itself.) The choices. The result, if we can show convergence, will Jacobian captures the “Abelian” or “linear” aspects manifestly not depend on any choices. of the geometry of X. There is a moduli space Ag has given rise to many important applications of this for such complex-algebraic tori (known as Abelian idea, in particular by giving a rich structure to the varieties), which does share all of the nice proper- collection of integrals obtained in this way. Donaldson ties and linear algebraic description of the moduli and Seiberg–Witten theories use this philosophy to of elliptic curves M1,1 = A1. The good news—the give topological invariants of four-manifolds. Gromov– —is that by assigning to each Riemann Witten theory applies it to the topology of symplec- surface X its Jacobian we embed Mg as a closed, tic manifolds, and to counting problems in algebraic complex-analytic subset of Ag. The interesting news— geometry, such as, How many rational plane curves of the —is that the image is quite com- degree 5 pass through 14 points in general position? plicated to characterize intrinsically. In fact, solutions (Answer: 87 304.) to this problem have come from as far afield as the Modular forms. One of the most profound ideas in study of nonlinear partial differential equations! mathematics, the Langlands program, relates num- ber theory to function theory (harmonic analysis) on 3.4 Further Directions very special moduli spaces, generalizing the moduli In this section we give hints at some interesting ques- space of elliptic curves. These moduli spaces (Shimura tions about, and applications of, moduli spaces. varieties) are expressible as quotients of symmetric spaces (such as H) by arithmetic groups (such as Deformations and degenerations. Two of the PSL2 Z). Modular and automorphic forms are main topics in moduli spaces ask which objects are special functions on these moduli spaces, described by very near to a given one, and what lies very far their interaction with the large symmetry groups of away. Deformation theory is the calculus of mod- the spaces. This is an extremely exciting and active uli spaces: it describes their infinitesimal structure. area of mathematics, which counts among its recent In other words, given an object, deformation theory triumphs the proof of Fermat’s last theorem and the is concerned with describing all its small perturba- Shimura–Taniyama–Weil conjecture (Wiles, Taylor– tions (see Mazur (2004) for a beautiful discussion of Wiles, Breuil–Conrad–Diamond–Taylor). this). At the other extreme, we can ask what happens when our objects degenerate? Most moduli spaces, for example the moduli of curves, are not compact, so Further Reading there are families “going off to infinity.” It is impor- tant to find “meaningful” compactifications of moduli For further reading, historical accounts and bibliogra- spaces, which classify the possible degenerations of our phies on moduli spaces, the following articles are objects. Another advantage of compactifying moduli highly recommended. spaces is that we can then calculate integrals over the A beautiful and accessible overview of moduli completed space. This is crucial for the next item. spaces, with an emphasis on the notion of deforma- Princeton Companion to Mathematics Proof 13 tions, is given by Mazur (2004). The articles by Hain (2000) and Looijenga (2000) give excellent introduc- tions to the study of the moduli spaces of curves, per- haps the oldest and most important of all moduli prob- lems. The article by Mumford and Suominen (1972) introduces the key ideas underlying the study of mod- uli spaces in .

Hain, R. 2000. Moduli of Riemann surfaces, transcendental aspects. In School on Algebraic Geometry, Trieste, 1999, pp. 293–353. ICTP Lecture Notes Series, No. 1. Trieste: The Abdus Salam International Centre for Theoretical Physics. Looijenga, E. 2000. A minicourse on moduli of curves. In School on Algebraic Geometry, Trieste, 1999, pp. 267– 291. ICTP Lecture Notes Series, No. 1. Trieste: The Abdus Salam International Centre for Theoretical Physics. Mazur, B. 2004. Perturbations, deformations and varia- tions (and “near–misses”) in geometry. Physics and num- ber theory. Bulletin of the American Mathematical Soci- ety 41(3):307–336. Mumford, D. and K. Suominen. 1972. Introduction to the theory of moduli. In Algebraic Geometry, Oslo, 1970: Proceedings of the Fifth Nordic Summer School in Math- ematics (ed. F. Oort), pp. 171–222. Groningen: Wolters- Noordhoff.