Group actions, divisors, and plane curves

By Araceli Bonifant and John Milnor

Abstract

After a general discussion of group actions, orbifolds, and weak orbifolds, this note will provide elementary introductions to two basic moduli spaces over the real or complex numbers: first the moduli space of effective divisors with finite stabilizer on the projective space , modulo the group of projective transformations of ; and then the moduli space of curves (or more generally effective algebraic -cycles) with finite stabilizer in , modulo the group of projective transformations of . It also discusses automorphisms of curves and the topological classification of smooth real curves in .

1. Introduction

A basic objective of this paper is to provide an elementary introduction to the moduli space of curves of degree in the real or complex projective plane, modulo the action of the group of projective transformations. However, in order to prepare for this, we first give an exposition of the general theory of smooth group actions and orbifolds. As a further introduction, we describe the theory of effective divisors on the real or complex projective line modulo projective transformations. We also discuss a number of related topics, including automorphism groups of curves and the topology of smooth curves in the real projective plane. The different sections are largely independent of each other.

§2. Group actions and orbifolds

We consider smooth actions of Lie groups on smooth Hausdorff manifolds. Since these actions are not always proper, we introduce the concepts of locally proper action, and weakly locally proper action. We introduce the usual concept of orbifold,⁠Footnote1 as well the concept of weak orbifold for quotient spaces which have only some of the standard orbifold properties, and we prove the following:

1

Most authors require orbifolds to be Hausdorff spaces, however we will allow orbifolds which are only locally Hausdorff.

For a proper/locally proper/or weakly proper action the quotient space is a Hausdorff orbifold/locally Hausdorff orbifold/or locally Hausdorff weak orbifold, respectively.

§3. Divisors on and the moduli space .

Working either over the real or complex numbers, the space of effective divisors of degree with finite stabilizer on the projective line , modulo projective transformations, is a space for every , but it is a Hausdorff orbifold only for . In the complex case the space is homeomorphic to , but in the real case is homeomorphic to a closed line segment in . In both the real and complex cases, there is just one point of which is “improper” in the sense that the action of the group of projective transformation at corresponding divisors is not proper. On the other hand, for , there is a unique maximal open subset which is a Hausdorff orbifold, but is not even locally Hausdorff at points outside of . This subset is compact for odd, but not for even.

The proofs make use of a Distortion Lemma for automorphisms of which describes the way in which an automorphism that lies outside a large compact subset of must distort the geometry. The proofs also make use of two familiar projective invariants associated with a 4-tuple of points in . The cross-ratio depends on the ordering of the four points, while the shape invariant (the classical -invariant times a convenient constant) is invariant under permutations of the four points.

In the complex case, a close relative of our is the classical moduli space consisting of closed Riemann surfaces of genus zero provided with an ordered list of distinct points, where two such marked Riemann surfaces are identified if there is a conformal isomorphism taking one to the other. We also discuss the compactification , a beautiful object introduced by Knudsen, based on ideas of Grothendieck, Deligne, and Mumford. (See also Keel Reference Ke and Etingof, Henriques, Kamnitzer, and Rains Reference EHKR for the analogous space .) In both the real and complex cases, there are associated embeddings

where . As an example, in the real case for and , there are ten essentially different embeddings of the circle into the surface , cutting it up into 12 hyperbolic pentagons, and thus presenting it as a hyperbolic analogue of the dodecahedron. The space is a 3-manifold which has a Jaco-Shalen-Johanssen decomposition, showing that it has some combination of hyperbolic and flat geometry. More explicitly, if we remove the ten tori corresponding to the embeddings of into , then the remainder can be given the structure of a complete hyperbolic 3-manifold with 20 infinite cusps.

§4. Curves (or -cycles) in and their moduli space

The space of all curves of degree in can be conveniently considered as a dense open subset of a projective space of dimension whose elements are formal linear combinations of curves in . These are called effective (algebraic) -cycles. For , we study the moduli space consisting of all projective equivalence classes of effective -cycles of degree with finite stabilizer. (For , there are no curves with finite stabilizer.) This moduli space has been studied by many authors. (See for example Mumford’s discussion of hypersurfaces Reference Mu, p. 79, which includes not only curves in but also divisors in .) However, our presentation is more elementary, and is addressed to nonspecialists.

§5. Cubic curves

The space is homeomorphic to the topological 2-sphere consisting of all ratios , corresponding to curves . It is isomorphic as an orbifold to the moduli space for divisors of degree four on . In the real case, the space is diffeomorphic to the unit circle. We describe the resulting circle of real cubic curves in several different ways, not only in terms of the normal form with , but also in terms of the Hesse normal form , with , and in terms of a convenient flex-slope normal form, parametrized by the slope at a flex point.

§6. Degree at least four.

When , we prove that the moduli space is not a Hausdorff space. On the other hand, we provide two different procedures for describing large open subsets of which are Hausdorff orbifolds. One of them is based on studying the distribution of virtual flex points, that is points which are either flex points or which yield flex points under perturbation. The other will be discussed in §7.

The open subset corresponding to smooth curves in is a Hausdorff space, which maps naturally into the classical moduli space consisting of all isomorphism classes of compact Riemann surfaces of genus . This map is injective for all (compare Chang Reference Ch), but is an isomorphism only for .

§7. Singularity genus and proper action

This section provides a different proof that large subsets of are Hausdorff orbifolds, based on the genus invariant for singular points. Both proofs make use of a Distortion Lemma for automorphisms of which lie outside a large compact set of group elements. These arguments apply only to points of moduli space which correspond to (possibly singular) curves, but there is a brief discussion of extending the proof to more general -cycles.

§8. Infinite automorphism groups.

Although we are primarily concerned with curves which have finite stabilizer, this section studies the opposite case of curves with infinitely many automorphisms. Following Klein and Lie, we call these -curves. We provide an explicit catalog of all such curves. (Compare Aluffi and Faber Reference AF1.)

§9. Finite automorphism groups.

The study of finite automorphism groups of curves is closely related to the study of the moduli space , since any curve with extra automorphisms gives rise to a singular point in moduli space, or at least to a ramified point. For each degree and prime , we provide an explicit criterion for deciding whether there exists a smooth curve of degree with a projective automorphism of period , and we compare this with a corresponding statement for automorphisms of arbitrary compact Riemann surfaces. We show that an arbitrary finite subgroup of is the full projective automorphism group for some smooth curve in , but show (following Chang Reference Ch) that a generic⁠Footnote2 curve of degree four has no nontrivial projective automorphism.

2

We say that a statement is true for a generic curve if it is true for all curves in some set which is dense and open in the Zariski topology. (Some authors prefer the term “general curve”, since the word “generic” has a more technical meaning in the theory of schemes.)

§10. Real curves: The Harnack-Hilbert problem

In the real case, the most studied question about smooth curves in is the Hilbert-Harnack problem concerning the number and topological arrangement of the connected components, each of which is a topological circle. We provide a report about this problem, as well as a suggested reformulation. The paper concludes with a brief Appendix describing some of the literature on the moduli spaces that we consider.

2. Group actions and orbifolds

This section will provide a general introduction to quotient spaces under a smooth group action. In the best case, with a proper action, the quotient space is Hausdorff, with an orbifold structure. Since the group actions that we consider are not always proper, we also introduce a modified requirement of weakly proper action, which suffices to prove that the quotient is locally Hausdorff, with a weak orbifold structure which includes only some of the usual orbifold properties. The section will conclude by discussing the special case of the projective general linear group and its action on projective space.

First consider the complex case. Let be a metrizable complex manifold, and let be a complex Lie group⁠Footnote3 which acts on the left by a holomorphic map ,

3

Although our Lie groups are always positive dimensional, the discussion would apply equally well to the case of a discrete group, which we might think of as a zero-dimensional Lie group.

where . (A manifold is metrizable if and only if it is paracompact and Hausdorff.) We will always assume that the action is effective in the sense that

The quotient space (or orbit space) in which is identified with if and only if for some will be denoted⁠Footnote4 by .

4

Since acts on the left, many authors would use the notation .

In the real case, the definitions are completely analogous, although we could equally well work in either the category or in the real analytic category. To fix ideas, let us choose the real analytic category. Thus in the real case, we will assume that is a real Lie group, that is a metrizable real analytic manifold, and that is a real analytic map. It will often be convenient to use the word analytic, by itself, to mean “real analytic” in the real case, or “complex analytic” in the complex case.

Definition 2.1.

For each the set consisting of all images with is called the -orbit of . We will also use the notation

if we are thinking of as a fiber of the projection map .

Remark 2.2.

In other words, each fiber is an equivalence class, where two points of are equivalent if and only if they belong to the same orbit under the action of . More generally, given any equivalence relation on , we can form the quotient space . Such a quotient space always has a well-defined quotient topology, defined by the condition that a set is open if and only if the preimage is open as a subset of . (We will always assume that is a Hausdorff space, but it does not necessarily follow that is Hausdorff.)

Remark 2.3 (Closed orbits and the condition).

By definition, a topological space is a -space if every point of is closed as a subset of . Evidently, a quotient space (or more generally ) is a -space if and only if each orbit (or each equivalence class) is closed as a subset of .

Orbifolds and weak orbifolds.

The concept and basic properties of orbifolds are due to I. Satake, who called them V-manifolds (see Reference Sa1Reference Sa2). They were later studied by W. Thurston (see Reference Th), who introduced the term “orbifold”.

Let stand for either the real or the complex numbers.

Definition 2.4.

By a -dimensional -orbifold chart around a point of a topological space we will mean the following:

(1)

a finite group acting linearly on ;

(2)

an -invariant open neighborhood of the point ; and

(3)

a homeomorphism from the quotient space onto an open neighborhood of in such that the zero vector in maps to .

The group will be called the ramification group at , and its order will be called the ramification index. A point is ramified if and unramified if .

The space together with an integer valued function will be called a -dimensional weak orbifold over if there exists such an orbifold chart around every point , such that the associated ramification function from to the positive integers coincides with the specified function restricted to .

Lemma 2.5.

Every weak orbifold is a locally Hausdorff space; and the function from to the set of positive integers is always upper semicontinuous, taking the value on a dense open set. More precisely, for any orbifold chart , we have for every , with on a dense open subset of .

Of course in good cases will be a Hausdorff space; but even a locally Hausdorff space can be quite useful.⁠Footnote5

5

Perhaps the most startling application of locally Hausdorff spaces in science would be to the “Many Worlds” interpretation of quantum mechanics, in which the space-time universe continually splits into two or more alternate universes. (See, for example, Reference Bec.) The resulting object is possibly best described as a space which is locally Hausdorff, but wildly non-Hausdorff. It can be constructed mathematically out of infinitely many copies of the Minkowski space by gluing together corresponding open subsets. (Of course it does not make any objective sense to ask whether these alternate universes “really exist”. The only legitimate question is whether a mathematical model including such alternate universes can provide a convenient and testable model for the observable universe.)

Proof of Lemma 2.5.

For any chart around and any nonidentity element , the set of elements of fixed by must be a linear subspace of of dimension at most . The complement of this finite union of linear subspaces within is a dense open subset . If is its image, then the associated mapping is locally bijective, and it is precisely -to-one. Since this map is one-to-one in a small neighborhood of any point of , it follows that all points of are unramified. Each -orbit in is compact and nonempty, so we can use the Hausdorff metric for compact subsets of to make into a metric space. In particular, it follows that is locally Hausdorff.

Definition 2.6.

An orbifold chart around gives rise to a smaller orbifold chart around any point which is called the restriction of this chart to a neighborhood of . In fact, choosing a representative point over , let be the stabilizer, consisting of all for which . Evidently acts linearly on , fixing the point . Choose an -invariant neighborhood of which is small enough so that the various images with are all disjoint from . Then the projection from to is the required restriction to an orbifold chart around .

Definition 2.7.

An orbifold atlas on is a collection of orbifold chart homeomorphisms

where the are open sets covering , which satisfy the following.

Compatibility condition. For each point in an overlap and each sufficiently small neighborhood of , the restriction of the th and th orbifold charts to are isomorphic in the following sense. Let

be the two restrictions. Then we require that there should be an analytic isomorphism so that . (Compare Figure 1.) Furthermore, there should be a group isomorphism so that

for every and every .

Two such atlases are equivalent if their union also satisfies this compatibility condition. The space together with an equivalence class of such atlases is called an orbifold.⁠Footnote6 (Compare Reference Th, Reference BMP.)

6

Caution. Most authors require orbifolds to be Hausdorff spaces, but as noted in the Introduction, we will allow orbifolds which are only locally Hausdorff.

There may well be examples of weak orbifolds which are not compatible with any orbifold structure, but we do not know any such examples.

Example 2.8.

On any Riemann surface, we can choose any function which takes the value except at finitely many points. A corresponding collection of complex orbifold charts is easily constructed, and the compatibility condition is easily verified.

Here is a pair of more interesting examples.

Example 2.9.

Let be the plane consisting of all with , and let be the group of permutations of the three coordinates. Then under the action of , each point of can be put into a unique normal form with . If we set and , then we can solve easily for as linear functions of . Thus the quotient space can be identified with the positive quadrant in the -plane. The ramification index is throughout the interior of the quadrant, with along the two edges, and with at the origin. Note that the compatibility is automatically satisfied, since we have specified only one coordinate chart.

The analogous example with looks quite different. In this case, the quotient space is best described as the complex -plane, where

are elementary symmetric functions, with . Again the ramification index of a point in the quotient space is if are all distinct, if only two are equal to each other, and if . Any triple with takes the form , up to permutation of the coordinates, with and . Thus if the expression⁠Footnote7 is nonzero; with for most points where this expression is zero; but with if .

7

Up to sign, this expression is just the discriminant of the polynomial .

The main object of this section will be to describe conditions on the group action which guarantee that the quotient will be an orbifold or weak orbifold.

Proper and weakly proper actions.

Definition 2.10.

A continuous action of on a locally compact space is called proper if the associated map

from to is a proper map (in the usual sense that the preimage of any compact set is compact). A completely equivalent requirement is the following.

For any pair of compact subsets the set of all with is compact.

Here is another completely equivalent condition.

For every pair of points and in , there exist neighborhoods of and of which are small enough so that the set of all with has compact closure.

The proof that these three forms of the definition are equivalent is straightforward and will be left to the reader.⁠Footnote8

8

In the special case of a discrete group, such an action is called properly discontinuous. (For more about proper actions, see the discussion beginning with Lemma 2.25.)

The action is locally proper at if this condition is satisfied for the special case where . It then follows that the action is proper throughout some -invariant open neighborhood of . In fact, if each which maps some point of into is contained in a compact set , then it follows that each which maps a point of to a point of is contained in the compact set .

The action will be called weakly proper at if the following still weaker local condition is satisfied:

There should be a neighborhood of and a compact set such that, whenever two points and of belong to the same -orbit, there exists at least one element with .

It will be convenient to call a point in either proper or improper according as the action of on corresponding points of is or is not locally proper. Similarly, an improper point in will be called weakly proper if the action of on corresponding points of is weakly proper. (Caution. Even when a point is improper, the quotient space may have a perfectly good orbifold structure.)

Definition 2.11.

Given an action of on , the stabilizer of a point is the closed subgroup of consisting of all for which . Note that points on the same fiber have isomorphic stabilizers, since

If the stabilizer is finite, then it follows easily that the fiber through (consisting of all images with ) is a smoothly embedded submanifold which is locally diffeomorphic to .

Under the hypothesis that all stabilizers are finite, we will prove the following in Theorem 2.18 together with Lemma 2.13 and Corollary 2.28:

For a proper action the quotient space is a Hausdorff orbifold.

For a locally proper action the quotient is a locally Hausdorff orbifold.

For a weakly proper action the quotient is a locally Hausdorff weak orbifold.

(See Figure 2 for an example of a smooth locally proper action with trivial stabilizers where the quotient is not a Hausdorff space.)

Remark 2.12.

Although there are examples which are weakly proper but not locally proper, they seem to be hard to find. Lemma 3.11 will show that divisors of degree four with only three distinct points give rise to such examples; and the proof of Lemma 5.4 will show that curves of degree three with a simple double point provide closely related examples. However, these are the only examples we know.

The following is well known.

Lemma 2.13.

If the action is proper, then the quotient is a Hausdorff space.

It follows as an immediate corollary that a locally proper action yields a quotient space which is locally Hausdorff. However the quotient under a locally proper action need not be Hausdorff. (Compare Figure 2.) If stabilizers are finite, then we will see in Theorem 2.18 that even a weakly proper action yields a quotient space which is locally Hausdorff.

Remark 2.14.

Note that every locally Hausdorff space is . In fact if one point belongs to the closure of a different point , then no neighborhood of is Hausdorff.

Proof of Lemma 2.13.

It will be convenient to choose a metric on . Given and there are two possibilities. If we can choose neighborhoods and so that no translate intersects , then the images and in the quotient space are disjoint open sets.

On the other hand, taking and to be a sequence of neighborhoods of and , respectively, of radius , if we can choose a group element for each with , then by compactness we can pass to an infinite subsequence so that the converge to a limit . It follows easily that , so that and map to the same point in the quotient space.

Remark 2.15.

The converse to Lemma 2.13 is false: A quotient space may be Hausdorff even when the action is not proper. Compare the discussions of and in Lemmas 3.4, 3.5, and 3.11, and of the quotient spaces and in Section 5. These quotients are Hausdorff orbifolds, even though the associated group action fails to be proper everywhere. (See Example 2.8, as well as Lemma 5.4.) However, for larger we will have to deal with moduli spaces which are definitely not Hausdorff.  (See Theorem 3.2 for and Theorem 6.1 for .)

Weak orbifold structures.

The passage from weakly proper actions to weak orbifold structures will be based on the following. Given any fiber , and given any point , we will refer to the quotient of tangent vector spaces

as the transverse vector space to at . (If is provided with a Riemannian metric, then can be identified with the normal vector space at .)

In order to describe a weak orbifold structure on the quotient, we must first construct the associated ramification groups. Note that the stabilizer acts linearly on both and , and hence acts linearly on the -dimensional quotient space , where is the codimension of in . In practice, we will always assume that the stabilizer is finite, so that is equal to the difference . However, this action is not always effective: the group may act nontrivially on , while leaving the transverse vector space pointwise fixed.

Definition 2.16.

Let be the normal subgroup of consisting of all group elements which act as the identity map on (that is, all such that for all ). The quotient group

will be called the ramification group at . Note that by its very definition, comes with a linear action on the vector space , which is isomorphic to or . It is not hard to check that different points on the same fiber have isomorphic ramification groups. Let be the image of the fiber in . As in Definition 2.4, the order of this finite group will be called the ramification index , and will be called unramified if .

We will need the following.

Lemma 2.17 (Invariant metrics).

In the real case, given any finite subgroup there exists a smooth -invariant Riemannian metric on the space Similarly, in the complex case has a smooth -invariant Hermitian metric.

Proof.

Starting with an arbitrary smooth Riemannian or Hermitian metric, average over its transforms⁠Footnote9 by elements of . Then each element of will represent an isometry for the averaged metric.

9

A Riemannian metric can be described as a smooth function which assigns to each a symmetric positive definite inner product on the vector space of tangent vectors at . Given any diffeomorphism , and given a Riemannian metric on , we can use the first derivative map to pull back the metric, setting

In particular, given any finite group consisting of diffeomorphisms from to itself, we can form the average The construction in the complex case is similar, using Hermitian inner products.

Now let be any fiber with finite stabilizers, and let be an arbitrary base point. Since the stabilizer is a finite group acting on , we can choose a -invariant metric on . Using this metric, the transverse vector space can be identified with the normal vector space consisting of all tangent vectors to at which are orthogonal to the fiber at . Given , we can consider geodesics of length starting at which are orthogonal to at . If is small enough, these geodesics will sweep out a smooth -dimensional disk which meets