Billiards and Teichmüller curves

By Curtis T. McMullen

Abstract

A Teichmüller curve is an isometrically immersed algebraic curve in the moduli space of Riemann surfaces. These rare, extremal objects are related to billiards in polygons, Hodge theory, algebraic geometry and surface topology. This paper presents the six known families of primitive Teichmüller curves that have been discovered over the past 30 years, and a selection of open problems.

1. Introduction

The moduli space of compact Riemann surfaces of genus is both a metric space and an algebraic variety.

The metric comes from the Teichmüller distance between , which measures the minimal conformal distortion of a map . This metric is given by a norm on each tangent space to , but for it is not Riemannian; in fact the norm balls are complicated convex sets, varying so much from point to point that is completely inhomogeneous.

The algebraic structure on comes from a projective embedding, which provides a multitude of algebraic curves . Each curve carries a natural hyperbolic metric, coming from its uniformization .

We say is a Teichmüller curve if this inclusion is an isometry. These rare and remarkable objects lie at the nexus of algebraic geometry, number theory, complex analysis, topology and automorphic forms. We focus on primitive examples, since all others are related to these by covering constructions (§2).

Teichmüller curves are elusive, but once found, they can often be viewed explicitly from many perspectives at once. For example, any primitive Teichmüller curve determines a totally real number field , with , such that:

, with ;

every Riemann surface can be assembled from triangles with vertices in , for some ; and

a factor of the Jacobian of admits real multiplication by .

The curve is rigid, so both and its map to are also defined over a number field. In particular cases one can obtain algebraic equations for , generators for , and geometric models for and for endomorphisms of its Jacobian.

Billiards.

Frequently can be chosen so there is a polygon and a finite reflection group , adapted to , such that

In this case, billiards in the polygon has optimal dynamics: every trajectory is either periodic, or uniformly distributed. Moreover, and can be reconstructed from .

The regular polygons provide the first examples of both optimal billiards and Teichmüller curves (see Figure 1.1 and §3). It is also possible that is immersed, rather than embedded in ; in this case we obtain a generalized polygon with optimal dynamics (see Figure 1.2 and §7).

The current catalog.

This paper provides a survey of the known examples of Teichmüller curves, a glimpse of their multifaceted constructions, a hint of how they were discovered, and a selection of the many open questions that remain.

The known primitive Teichmüller curves are given by 6 infinite series, and 3 sporadic examples. These are:

1.

the three Weierstrass series , in genus , and 4, §5);

2.

the sporadic examples of type , and , in genus 3 and 4 (§6);

3.

the Bouw–Möller series , providing finitely many examples in every 7); and

4.

the gothic and arabesque series, and , both in genus 4 (§8, §9).

The five horizontal series , and above each lie in a single moduli space. The index is a real quadratic discriminant, i.e. an integer , or , with .

Completeness?

In low genus, the list of primitive Teichmüller curves is almost complete.

In genus , all primitive Teichmüller curves are known: they are accounted for by the series and one other curve, associated to billiards in the regular decagon (Theorem 4.5).

In genus , there are only finitely many primitive Teichmüller curves not accounted for by the series (Theorem 5.5).

On the other hand, the vertical Bouw–Möller series gives the only known construction of primitive Teichmüller curves in genus . Thus a central open problem is to settle:

Question 1.1.

Are there infinitely many primitive Teichmüller curves in ?

The unexpected families and in genus 4 hint that similar constructions may be hidden in higher genus.

Teichmüller surfaces.

The discovery of the gothic curves also revealed an almost miraculous new phenomenon: there are primitive, totally geodesic Teichmüller surfaces in , and . This survey concludes with a description of these new surfaces from the perspective of algebraic geometry in §8, and from the perspective of quadrilaterals in §9.

Notes and references.

References and commentary are collected, section by section, in §10.

Via the action of on , Teichmüller curves are connected to the larger topic of dynamics on moduli spaces, which is itself patterned on the theory of homogeneous dynamics, Lie groups, lattices and ergodic theory. For a view of the broader setting, we recommend the many excellent surveys such as Reference D, Reference Go, Reference HS2, Reference Mas3, Reference MT, Reference Mo3, Reference Sch2, Reference Vo1, Reference Wr2, Reference Wr3, Reference Y, and Reference Z.

Outline.

In §2 and §3 we set the stage with definitions and basic examples regarding moduli spaces, polygons and billiards. The known families of primitive Teichmüller curves are described in §4 through §9.

Four appendices follow. The triangle groups are reviewed in Appendix A. There are six accidental isomorphisms between members of series of Teichmüller curves listed above; these are recorded in Appendix B. Tables of invariants of Teichmüller curves appear in Appendices C and D.

Notation.

The th Chebyshev polynomial will be denoted by ; it is characterized by

We let denote cohomology with complex coefficients. The upper half-plane in is endowed with the complete hyperbolic metric

of constant curvature . The group acts linearly on and by Möbius transformations on .

2. Moduli spaces and Teichmüller curves

This section develops background material on Riemann surfaces, polygons, and the action of on the moduli space of holomorphic 1-forms . This material will allow us to formulate the main topic we aim to address:

Problem 2.1.

Construct and classify all primitive Teichmüller curves .

Moduli space.

The moduli space parameterizes the isomorphism classes of compact Riemann surfaces of genus . It is naturally a complex orbifold, and an algebraic variety, of complex dimension when .

The Teichmüller metric on is defined by a norm on each tangent space; it can be characterized as the largest metric such that every holomorphic map

is either a contraction or an isometry. In the isometric case, we say is a complex geodesic.

Metrically, moduli space is completely inhomogeneous: the tangent spaces at are isomorphic as normed vector spaces if and only if . Nevertheless, there exists a unique complex geodesic through every point in every possible direction.

Polygons and Riemann surfaces.

How can one specify a Riemann surface ?

In the case , is a torus, thus one can write for some lattice . Alternatively, if we choose a parallelogram that is a fundamental domain for the action of , we can construct by gluing together opposite sides of .

More generally, if is any polygon, and the edges of are identified in pairs by translations, then the result is a compact Riemann surface . And in fact:

Every compact Riemann surface of genus can be presented as a polygon with its edges glued together by translations.

Note that inherits a flat metric from . At first sight this may seem paradoxical: for , admits no smooth flat metric. However the metric on has, in general, isolated singularities of negative curvature arising from the vertices of .

Example in genus 2.

Consider the polygon shown in Figure 2.1. Note that we have introduced two extra vertices, so is combinatorially an octagon. Gluing edges by vertical and horizontal translations, we obtain a Riemann surface of genus 2. The eight vertices of descend to a single point ; there, the induced flat metric has a cone angle of .

Holomorphic 1-forms.

This flat uniformization of by a polygon can be contrasted with more traditional ways of presenting a compact Riemann surface, e.g. as an algebraic curve or as a quotient of by a Fuchsian group. While a polygonal presentation of is elementary, it is not canonical; moreover, it provides with additional structure.

To explain this, recall that the space of holomorphic -forms on has dimension ; indeed, can be taken as the definition of the genus of . In local coordinates, , where is a holomorphic function. Provided , its zero set consists of points, counted with multiplicity.

The moduli space of all nonzero 1-forms of genus forms a bundle

in fact, it is a holomorphic vector bundle of rank with its zero section removed.

Strata.

The locus where the zeros of have multiplicities forms a stratum

of dimension . These strata decompose into disjoint algebraic sets, indexed by the partitions of . We sometimes use exponential notation for repeated blocks of a partition; e.g. the unique open stratum is denoted by .

From polygons to 1-forms.

Let . Since the 1-form on is invariant under translation, it descends to give a 1-form . Here is a more precise description of the relationship between Riemann surfaces and polygons.

Theorem 2.2.

Every element of can be presented in the form

for a suitable polygon .

It is often useful, as we will see below, to allow the ‘polygon’ to be disconnected. With this proviso, the proof of the result above is fairly elementary: one can construct a geodesic triangulation of the flat surface , with among its vertices, and then present as the quotient of a collection of Euclidean triangles.

Geometry of a 1-form.

A holomorphic 1-form provides with a singular flat metric . This metric has a cone angle of at each zero of order . The form determines a smooth measure on , with total mass given by

Near any point , we can choose a local flat coordinate on such that . The geodesics on are simply straight lines in these charts. Since these flat coordinates are well–defined up to translation, each geodesic has a well–defined slope . In particular, cannot cross itself; all geodesics are simple. We allow a geodesic to begin or end at a point of , but never to pass through a zero. In particular, a closed geodesic is always disjoint from .

These features are elementary to see in a polygonal model ; for example, the horizontal lines in descend to a foliation of by geodesics with slope zero. Intrinsically, this foliation is defined by the closed 1-form .

A cylinder is the closure of a maximal open set foliated by parallel closed geodesics. Every closed geodesic lies in a cylinder; in particular, is never unique its homotopy class. Most elements of are not represented by closed geodesics; rather, the loop of minimal length in a given homotopy class is a chain of geodesic segments of varying slopes, with endpoints in .

Action of .

Remarkably, upon passage to the bundle , the highly inhomogeneous space acquires a dynamical character: namely, it admits a natural action of . This action is easily described in terms of a polygonal presentation Equation 2.2: for , we have:

Here acts linearly on , and the (combinatorial) gluing instructions remains the same.

Alternatively, given one can define a harmonic form on by

and then change the complex structure on so is holomorphic on . The zeros of and have the same order, so:

leaves each stratum invariant.

Complex geodesics.

Note that if is simply a rotation, then and for some . Thus the projection of to depends only on the coset

and the map covers a unique map , making the diagram

commute.

The map is a holomorphic, isometric immersion of into moduli space, which we refer to as the complex geodesic generated by . If , then the image of simply consists of all Riemann surfaces of the form , ; see Figure 2.2.

Real geodesics.

Every 1-form also generates a distinguished real Teichmüller geodesic ray ; parameterized by arclength, it is given by

where

The Riemann surface is obtained from by shrinking its horizontal geodesics and expanding its vertical ones.

Teichmüller curves.

The stabilizer of is a discrete subgroup

It is easy to see that the complex geodesic generated by descends to give a map , where

Here the action of on is slightly twisted, since acts on the left in equation Equation 2.3; it is given by .

Now suppose has finite hyperbolic area; equivalently, suppose is a lattice in . Then the image of the map

is a Teichmüller curve in . That is, is the normalization of a totally geodesic algebraic curve.

We refer to as a generator of the Teichmüller curve . The generator of is not unique—for any and , is also generated by .

Hidden symmetries.

The pivotal group —which is large in the case of a Teichmüller curve—reflects hidden symmetries of the form itself.

More precisely, can be described as follows. Let denote the group of orientation–preserving homeomorphisms of that stabilize , and have the form

in local flat coordinates on the domain and range satisfying . Here and .

We refer to as an affine automorphism of , since it preserves the real–affine structure on determined by ; in particular, sends geodesics to geodesics. The matrix is independent of the choice of charts, and is characterized by the property that

In particular, if and only if belongs to , the group of holomorphic automorphisms of satisfying .

It is then easy to see we have an exact sequence:

For , the group is finite, so the stabilizer of in is virtually the same as its affine symmetry group.

Examples.

The square torus generates the simplest example of a Teichmüller curve. In this case, every orientation–preserving automorphism of as a Lie group is also an affine automorphism of . Thus , and the map is an isomorphism. This is the trivial Teichmüller curve.

An example in genus two is provided by the form (see Figure 2.1). Here we find

In these examples and are both triangle groups, namely and . See Appendix A for more on triangle groups, which will occur frequently in the discussions to follow.

Cylinders and parabolics.

The modulus of a cylinder of height and circumference is . In general, if is covered by a collection of horizontal cylinders with moduli , and divides for all (meaning is an integer), we can construct an affine automorphism of with

Namely we take to be a linear, right Dehn twist, iterated times. The iterate is chosen so is the matrix above for all . Since is the identity on , these twists fit together to give a map .

Conversely, it is not hard to show:

Proposition 2.3.

Suppose contains a parabolic element fixing the line of slope through the origin. Then is tiled by a family of cylinders of slope , with rational ratios of moduli.

We can now explain the appearance of the parabolic matrix in . Note that is built from three copies of the unit square. The bottom two squares define a horizontal cylinder , isometric to a Euclidean cylinder of height and circumference . Similarly the top square gives a cylinder with . Thus , and .

As for the generator , it is easy to see that has a 4-fold rotational symmetry whenever (see Figure 3.4).

Cusps of .

We note that the Teichmüller curve generated by a 1-form is properly immersed in , and hence the orbit of in is closed.

Proposition 2.4.

A Teichmüller curve has finite hyperbolic area, but it is never compact; it always has at least one cusp.

Idea of the proof.

Assume and . Construct a geodesic segment on with endpoints in . After rotating , we can assume is horizontal. Now consider the Teichmüller geodesic ray in generated by , as in equation Equation 2.5. As the length of on tends to zero. Since the orbit of is closed in its stratum, the endpoints of cannot collide, so tends to infinity in . Therefore is noncompact, and tends to a cusp of .

Combined with Proposition 2.3, we find that has many cylinder decompositions and a dense set of periodic directions.

Square–tiled surfaces.

Let us say is a square–tiled 1-form if it can be obtained by gluing together a finite number of copies of the unit square .

Generalizing the case of , one can show that has finite index in for any square–tiled 1-form. One can also check that square tiled surfaces are dense in . Consequently:

Teichmüller curves are dense in .

These Teichmüller curves, however, are simply echos in higher genus of the trivial Teichmüller curve . Every Teichmüller curve generates similar echos in higher genus, via covering constructions. For example, there is a degree 3 holomorphic map such that .

Primitivity.

For this reason we will focus our attention on primitive Teichmüller curves in : those that do not arise from lower genus.

To define these, let us say is the pullback of if there is a holomorphic map such that . A 1-form with is primitive if it is not the pullback of a form of lower genus.

Every form in , , is the pullback of a unique primitive 1-form Reference Mo1, Thm. 2.6, Reference Mc3, Thm. 2.1. We say a Teichmüller curve is primitive if it is generated by a primitive 1-form. In this case is trivial, and hence

Invariants.

We conclude this section by discussing three invariants of the Teichmüller curve generated by a 1-form .

1.

The lattice , often called the Veech group, is determined by up to conjugacy in . Indeed, it is simply the Fuchsian group uniformizing .

2.

The trace field of , defined by

is also an invariant of . It is a totally real number field, of degree at most over , satisfying

for any hyperbolic element . Moreover if and only if is the pullback of a form of genus one.

3.

All generators of lie in the same stratum , so this too is an invariant of .

The trace field and stratum are known for all the Teichmüller curves we will discuss below. The lattice , on the other hand, is often inaccessible. Nevertheless, topological invariants of , such as its Euler characteristic, can frequently be determined.

3. Billiards

We now turn to the remarkable connection between Teichmüller curves and billiards in polygons.

The first nontrivial Teichmüller curves were discovered in 1989 by Veech. They play a key role in his proof of:

Theorem 3.1.

Billiards in a regular polygon has optimal dynamics.

Here optimal dynamics means that any unit speed billiard trajectory satisfies the Veech dichotomy; it is either

(i)

periodic: meaning for some ; or

(ii)

uniformly distributed: meaning is dense, and

for any continuous function .

Which alternative holds—(i) or (ii) above—depends only on the initial slope of the trajectory. See Figure 3.1 for examples.

In this section we describe the series of Teichmüller curves associated to regular polygons, and present the proof of Theorem 3.1, following Reference V1 and Reference Mas2. We also summarize, in Theorem 3.9, the known examples of triangles with optimal billiards.

A striking feature of Theorem 3.1 is that it describes the behavior of every trajectory in , and shows that only two, radically different types of behavior are possible. For more general polygons, some trajectories may be neither periodic nor dense (see Figure 3.2), and even dense trajectories may be unevenly distributed.

Unfolding.

A polygon is rational if its angles lie in . To relate billiards to Teichmüller theory, we first explain how a rational polygon determines a holomorphic 1-form .

Suppose is a regular pentagon. The construction of is described in Figure 3.3. The idea is that, when a billiard trajectory strikes an edge of , rather than reflecting , we can reflect . The result is an adjacent polygon , and continues into along a straight line.

Now when strikes an edge of , we could add yet another polygon ; but would simply be a translate of . So instead of adding new polygons, we glue the edges of to parallel edges of . Since is combinatorially an octagon, the result is a 1-form

of genus two; and billiard trajectories in go over to geodesics on the flat surface .

A similar construction can be carried out for any rational polygon . The rationality condition means the group generated by reflections in the sides of is finite. Let be the subgroup that stabilizes up to translation. The associated 1-form is then given by

Two examples.

(i) If is a regular -gon, then is simply with opposite edges identified.

(ii) The polygon unfolds to give a Swiss cross with its sides identified; the resulting form is isomorphic to up to a factor of two (see Figure 3.4).

From billiards to Teichmüller curves.

We say is a lattice polygon if is a lattice in ; equivalently, if generates a Teichmüller curve . For brevity, we also say generates . The connection between billiards and Teichmüller curves is summed up by the following general statement.

Theorem 3.2.

The billiard flow in a lattice polygon has optimal dynamics.

With this result in hand, Theorem 3.1 follows from:

Theorem 3.3.

Every regular polygon generates a primitive Teichmüller curve. In particular, is a lattice polygon.

Corollary 3.4.

Every moduli space , , contains a primitive Teichmüller curve.

The invariants of these first examples of primitive Teichmüller curves are listed in Table 3.1. Note that is always a triangle group, that has just one or two zeros, and that every occurs.

We will sketch the proof of Theorem 3.2 at the end of this section.

Algebraic models.

One of the remarkable properties of a regular polygon is that we have an explicit algebraic model for the Teichmüller curve

it generates, to complement the flat model coming from the orbit of its unfolding. Recall that denotes the degree Chebyshev polynomial (see equation Equation 1.1).

Theorem 3.5.

For odd, the Teichmüller curve generated by a regular -gon is given by , where is the hyperelliptic curve defined by

and ranges in the space

The curves and are isomorphic, since is odd, and thus is well–defined on .

The limiting curve is defined by . Indeed, the unfolding of yields the 1-form generating .

To take into account the symmetries of and , we regard as the orbifold; the unique cusp comes from . Thus is natural uniformized by as indicated in Table 3.1.

Even polygons.

When is even, a slightly different family is required. In this case is also even, and we define by the polynomial equation

for . The unfolding of gives the form on the curve defined by

In brief, these formulas arise from the close relationship between the family of curves , and the family of hyperelliptic curves branched over orbits of the dihedral group .

The hidden symmetries of the pentagon.

The fact that regular polygons generate Teichmüller curves (Theorem 3.3) can be verified by a direct calculation. To indicate the idea, we will show that

when is a regular pentagon. (This group is conjugate to .)

First, referring to Figure 3.3, observe that contains a symmetry of order that exchanges and , and a symmetry of order that rotates each. Their product gives an element of order 10 in , namely

Second, observe that decomposes into a pair of horizontal cylinders of equal modulus . One of these cylinders is obtained by gluing the top two edges of to the bottom two edges of ; it is built from 2 isosceles triangles with angles . It follows readily that . Thus, by Proposition 2.3, the affine group of contains a product of Dehn twists with

Noting that , we find

(see Appendix A). Since this triangle group is a lattice, so is . (In fact equality holds above, since is a maximal discrete subgroup of .) Its trace field is quadratic, so the corresponding Teichmüller curve in is primitive.

The golden table.

Let denote the golden ratio. It can be shown by elementary geometry that the polygon and the regular pentagon generate the same Teichmüller curve. In fact, corresponds to the orbifold point of order on , while corresponds to the point of order 5 (see Figure 3.5). As an alternative to the calculation above, one can readily check that , using the fact that decomposes into 2 horizontal cylinders of modulus . (The similar case of was discussed in §2.)

Aside: Is the Veech dichotomy effective?

Although Theorem 3.2 clearly separates the slopes in a lattice polygon into two classes, it is an open problem to distinguish between them, even in the case of a regular polygon.

To make this precise, let be a regular -sided polygon resting on the real axis, let be its set of periodic slopes, and let be the maximum number of times that a trajectory with slope hits the sides of before returning to its starting point.

It is easy to see that , where and . Equality holds when , and in fact

where denotes the height of Reference Mc10. Otherwise, Reference AS, and one can ask:

Question 3.6.

Is there an algorithm to determine, given , if a trajectory with slope in is periodic?

Equivalently:

Question 3.7.

Is there a computable function such that whenever is a periodic slope?

The first open case occurs when . In this case, an unexpected phenomenon arises: experimentally, many slopes that are not periodic are fixed by hyperbolic elements in the Veech group.

Question 3.8.

Is every point the fixed point of a parabolic or hyperbolic element of ?

A positive answer would yield an algorithm for recognizing the periodic slopes in the heptagon.

Lattice triangles.

For a more systematic enumeration of lattice polygons, it is natural to start with triangles.

To summarize what is known, let us say a triangle has type if its internal angles are proportional to the integers . The type of determines up to similarity.

Theorem 3.9.

Triangles of the following types generate Teichmüller curves:

A.

, and , ;

B.

, odd;

C.

, and ; and

D.

.

Any other lattice triangle must be scalene and obtuse, like examples B and D above.

Corollary 3.10.

All acute, right–angled, and isosceles lattice triangles are known.

Series A, discovered by Veech, produces the same Teichmüller curves as the regular polygons. Series B, discovered by Vorobets and studied by Ward, is genuinely new. These two series are special cases of the Bouw–Möller examples, to be discussed in §7. The three triangles in series C, discovered by Vorobets, Veech and Kenyon–Smillie, will be related to the exceptional Coxeter groups in §6. Example D, found by Hooper, generates a Teichmüller curve in the Weierstrass series to be discussed in §5, namely .

Question 3.11.

Is the list of lattice triangles above complete?

Dynamics on moduli space.

To conclude, we outline the proof of the Veech dichotomy, Theorem 3.2. The proof pivots on the following important result.

Theorem 3.12 (Masur).

If the Teichmüller geodesic ray generated by is recurrent in , then every horizontal geodesic in is uniformly distributed.

The ray in question, defined by equation Equation 2.5, shrinks the horizontal geodesics on . This process accelerates the horizontal geodesic flow; indeed, the special properties of lattice polygons can be traced to the role played by as a renormalization group for the dynamics of billiards in .

Sketch of the proof.

Masur’s original proof appears in Reference Mas2; see also Reference Mc7. Here we sketch a cohomological argument from Reference Mc8.

Let be the cone of closed, positive de Rham currents carried by the foliation of by horizontal geodesics. The long–term behavior of any horizontal geodesic is described by such a current. Using the Hodge norm on , one can measure the size of by the diameter of its intersection with the unit sphere.

Let be the Teichmüller ray generated by . By contraction of the complementary period mapping, which records the Hodge structure on the part of orthogonal to , every time the geodesic ray returns to a fixed compact set , one can improve the estimate on by a factor of . Thus if is recurrent, we have and reduces to the ray through . This means the horizontal foliation is uniquely ergodic, and hence every horizontal geodesic is uniformly distributed.

Proof of Theorem 3.2.

Suppose is a lattice polygon. Let

be the Teichmüller geodesic generated by .

The Veech dichotomy reflects the following alternative for the behavior of a geodesic ray on a finite–volume hyperbolic surface such as : either is recurrent, or converges to a cusp (see Figure 3.6). Recurrence implies every horizontal geodesic on is uniformly distributed, by Theorem 3.12; while convergence to a cusp implies every horizontal geodesic is periodic, by Proposition 2.3. Consequently, billiard trajectories in with initial slope zero are either periodic or uniformly distributed as well. The same reasoning applies to other slopes, by rotating .

Remark: Billiards that hit vertices.

The proof of the Veech dichotomy also shows that any infinite billiard trajectory starting at a vertex of is uniformly distributed. However, a trajectory that joins a pair of vertices need not have periodic slope. A counterexample in the lattice triangle is shown in Figure 3.7: the first trajectory joins a pair of vertices, while the second, parallel trajectory is aperiodic (cf. Reference Bo).

Question 3.13.

Must every edge of a lattice polygon have periodic slope?

A positive answer for quadratic trace fields is given in Reference Mc10, Cor. 6.3.

4. Genus 2

The regular polygons provide only three examples of Teichmüller curves in for : those generated by the regular pentagon, octagon and decagon.

It is thus natural to ask: are there infinitely many primitive Teichmüller curves ? In this section we will see the answer is yes. In fact, we will present a complete classification of such curves, following the development in Reference Mc1 and its sequels. These curves were also discovered, from another perspective, in Reference Ca.

Real multiplication.

Where should one look for Teichmüller curves ?

It turns out that membership in is reflected, not by the automorphism group of , but by the endomorphism ring of its Jacobian.

To describe this connection, recall that the Jacobian of is the compact complex torus

An endomorphism is a holomorphic group homomorphism; it is self-adjoint if its action on satisfies

with respect to the symplectic intersection pairing on –cycles. Equivalently, the action of on is self-adjoint for the inner product

Let be a totally real number field of degree over . We say admits real multiplication by if there is an inclusion

and every is self-adjoint. Then admits an orthonormal basis of eigenforms , satisfying for each .

Taking into account the integral structure, one can say more precisely that admits real multiplication by the order characterized by

The Weierstrass curves.

Now suppose . Then is a real quadratic field, and an order

is uniquely determined by its discriminant . Any integer with , or can occur, provided is not a square.

For each such , the Weierstrass curve is defined to be the locus of such that

(i)

admits real multiplication by , and

(ii)

one of its eigenforms has a double zero.

(This double zero lies at one of the six Weierstrass points of , hence the terminology.)

The first condition implies that lies on a surface in , birational to the Hilbert modular surface

(This surface describes the possibilities for .) The additional requirement that has a double zero, rather than two simple zeros, reduces to an algebraic curve.

Classification.

We can now state the classification of primitive Teichmüller curves in genus two.

Theorem 4.1.

Each component of is a primitive Teichmüller curve.

Theorem 4.2.

The curve is irreducible unless , in which case it has two components.

In the second case the components are distinguished by a spin invariant or .

Theorem 4.3.

Every Teichmüller curve in is generated by billiards in an explicit -shaped table.

For example, when is even, is generated by , where .

Corollary 4.4.

There are infinitely many primitive Teichmüller curves in .

The oasis.

When the infinite sequence of curves was first discovered, it seemed like we might be standing on the edge of a jungle, with many more curves to be found in the larger stratum . A fruitless computer search, however, began to suggest that we had stumbled upon, not a jungle, but an oasis in a vast desert. This perception is confirmed by:

Theorem 4.5.

The regular decagon generates the only other primitive Teichmüller curve in .

The proof will be sketched at the end of this section.

Classification and synthesis.

Taken together, the above results give a classification of the primitive Teichmüller curves in genus 2, in the following sense: (i) there is an explicit construction of all examples (e.g. using -shaped tables); and (ii) there are readily computable invariants that allow one to test when two constructions yield the same Teichmüller curve. For example, the regular pentagon and octagon generate and , respectively.

In contrast to the case of regular polygons, the definition of is synthetic: it is an abstractly defined algebraic curve, whose properties remain to be investigated. There is no known general algebraic formula for , although many particular cases have been treated. An understanding of the corresponding Veech groups is equally elusive.

Topology of .

Despite these mysteries, the following results provide a complete description of topology of .

Theorem 4.6 (Bouw–Möller).

If has two components, they are homeomorphic as orbifolds.

Theorem 4.7 (Bainbridge).

We have .

The Euler characteristic of the Hilbert modular surface defined by Equation 4.1 is well–studied and related to the coefficients of a modular form. Writing where is a fundamental discriminant, it is given by

(The summands involve the Jacobi symbol for and the Möbius function .)

Theorem 4.8 (Mukamel).

For , the orbifold points on all have order . Their their total number is a weighted sum of class numbers of the quadratic imaginary orders with discriminants , and .

These orbifold points can be described as octagonal pinwheels with their opposite sides identified (see Figure 4.1); they cover elliptic curves with complex multiplication, allowing one to give explicit equations for . For example, the unique orbifold point is defined by

where .

There is a simple combinatorial method to enumerate the cusps of . Combining these results, one can also compute the genus of its components. As a consequence, it is known that all components of have genus when , and that the genus grows roughly like . Appendix C gives a table of invariants of for all .

From affine maps to the Jacobian.

Let us return to the definition of , and sketch the proof of Theorem 4.1.

Where does real multiplication on come from? To answer this question, suppose contains an element with irrational trace . Here is an affine automorphism. Since has genus two, is quadratic over ; we denote its Galois conjugate by .

Typically, the action of on is only real linear. However, the transformation is always complex linear. Thus the formally defined map

preserves the holomorphic form up to a constant factor; in fact, in view of equation Equation 2.7 we have:

To make rigorous sense of , we first linearize by letting it act on . Then equation Equation 4.3 gives a well–defined endomorphism of of . Since preserves the intersection form, is self-adjoint. Moreover, immediately extends to an endomorphism of the real Lie group

We claim the action of on is holomorphic. To see this, consider the isomorphism

defined by . Extend to an orthogonal basis for ; then the splitting

corresponds, under the isomorphism Equation 4.5, to the splitting

Here is defined using the (purely topological) intersection form on .

Since preserves , the multiplicities of its eigenvalues and must be the same. By equation Equation 4.4, and is multiplication by . Since is self-adjoint, ; and hence is multiplication by . It follows that is a complex linear mapping, with eigenvectors and , and hence itself gives a holomorphic endomorphism of the Jacobian of .

Summing up, we have shown:

Theorem 4.9.

Suppose is a form of genus two and the trace field of is quadratic. Then admits real multiplication by , with as an eigenform.

The eigenform locus.

For each discriminant , define the eigenform locus in by

Theorem 4.9 shows that, if we are looking for forms that generate primitive Teichmüller curves, we can restrict attention to the 3–dimensional eigenform loci inside the 5–dimensional space . This fact motivates the definition of .

Proof of Theorem 4.1.

Using the splitting in equation Equation 4.6, one can check that is –invariant. Thus the same is true for the 2–dimensional variety

But the projection of this locus to is exactly . Thus each component of is a Teichmüller curve, generated by a form lying above it; and is primitive, because its trace field is irrational.

Using Theorem 4.9, we obtain:

Corollary 4.10.

If and the trace field of is irrational, then is a lattice.

Taking into account Proposition 2.3 and Theorem 3.2, we also find:

Corollary 4.11.

An -shaped billiard table is a lattice polygon if and only if and are both rational.

For example, is a lattice polygon if and only if for some , .

Irreducibility: Computer-assisted proof.

Next we indicate the proof of Theorem 4.2. Suppose for simplicity that is even. We wish to show is connected.

Let denote the graph whose vertices correspond to the cusps of , and whose edges join vertices that are related by a ‘butterfly move’. The graph is easily computed: its vertices correspond roughly to the -shaped polygons that generate , and its edges join cusps that are guaranteed, by an elementary argument, to lie in the same component.

To complete the proof, it suffices to show that is connected for all even . This is first verified computationally for all . The remaining cases are handled using about 11,000 different connection strategies, related to the primes and . These strategies are also generated and verified by computer, and shown to cover all cases.

The exceptional decagon.

Finally, we describe the proof of Theorem 4.5: the regular decagon generates the only remaining primitive Teichmüller curve in .

By construction, the Weierstrass curves have generators in , and they account for all the primitive Teichmüller curves with generators in this stratum. Why does the larger stratum yield only one more example? The answer is contained in the following result, which applies to all .

Theorem 4.12 (Möller).

Suppose is a lattice with trace field . Then there is a surjective map from to a compatibly polarized Abelian variety such that

(1)

corresponds to an eigenform for real multiplication by on ; and

(2)

the difference of any two zeros of is torsion in .

Corollary 4.13.

If generates a primitive Teichmüller curve in , and , then is torsion in .

Equivalently, there is a meromorphic function on whose only zeros and poles are at and .

Sketch of the proof of Theorem 4.5.

Let be a primitive Teichmüller curve generated by . By Theorem 4.9, for some .

Consider the closure of in the compactified moduli space . On the boundary, one obtains a stable 1-form with and . The real multiplication and torsion conditions persist in the limit, and give a pair of rational numbers such that . The are 15 possible values for . The pair gives rise the to regular decagon, and the remaining 14 values are ruled out by a computer–assisted calculation.

What is the shape of ?

Perhaps the main open question concerning Teichmüller curves in genus two is the structure of their Veech groups. It is remarkable that results like Corollary 4.10 allow one to certify that is a lattice, without even revealing the volume of its quotient. We conclude by stating:

Problem 4.14.

Give a direct construction of the Veech groups of the Weierstrass curves .

Remark: Square discriminant.

One can also define Weierstrass curves for square discriminant, by replacing with . The corresponding Teichmüller curves are not primitive, but many results (such as Theorem 4.2) naturally generalize to this case.

5. Genus and

In this section we will use Prym varieties, which are variants of the Jacobian, to construct infinitely many primitive Teichmüller curves in genus and , following Reference Mc3.

Eigenforms in higher genus.

One natural approach to generalizing the construction of is to consider the locus of eigenforms for real multiplication in . Unfortunately, for , this locus is generally not –invariant Reference Mc1, Thm. 7.5.

On the other hand, Theorem 4.12 shows that along a Teichmüller curve, one should only expect real multiplication on a factor of . If we arrange that this factor is 2–dimensional, then the arguments from genus two will apply; and if, moreover, we require that has only one zero, then the torsion condition in Theorem 4.12 will be vacuously satisfied.

The Weierstrass curves .

These considerations motivate the following definition. Fix , and let be a real quadratic discriminant. The Weierstrass locus consists of those such that:

(1)

the form has a single zero, of multiplicity ;

(2)

there exists a (unique) involution such that ;

(3)

the genus of is ; and

(4)

the differential is an eigenform for real multiplication by on the Prym variety

The Prym variety is the polarized Abelian subvariety of corresponding to the –eigenspace of . In the case above, it is 2–dimensional by the condition that .

The Weierstrass curve is the projection of to . It is sometimes denoted by to emphasize the fact that .

By arguments similar to those for genus 2, we find:

Theorem 5.1.

For any and , the Weierstrass curve is a finite union of primitive Teichmüller curves.

Genus 2, 3 and 4.

For , is the hyperelliptic involution on , so we recover the definition of from §4. As we will see in §6, it is straightforward to give explicit examples of forms in for and , and thereby establish:

Corollary 5.2.

There are infinitely many primitive Teichmüller curves in and .

However, is empty for . For example, when , the map must be a covering map, contradicting the fact that fixes the unique zero of .

Classification.

The analogue of Theorem 4.2 in higher genus is:

Theorem 5.3 (Lanneau–Nguyen).

For genus , is irreducible if is even; it has two components if ; and otherwise, it is empty.

For genus , is irreducible for all .

Near completeness in genus 3.

In genus three, only three additional primitive Teichmüller curves are known.

Problem 5.4.

Do the regular polygons with 7 and 14 sides, and the triangle, generate all the primitive Teichmüller curves in genus 3 outside the Weierstrass series ?

Although this problem remains open, the works of several authors, including Aulicino, Nguyen and Wright, combine to yield notable progress.

Theorem 5.5.

The Weierstrass curves account for all but finitely many primitive Teichmüller curves in genus .

The proof is sketched in §10.

As we will see in §9, the situation is surprisingly different in genus 4: two additional infinite families of primitive Teichmüller curves are now known. We remark that the triangle (from §3) generates the curve .

Topology of the Weierstrass curves.

Write where is a fundamental discriminant, and let if is odd and if is even. Let denote the number of orbifold points of order on .

The next three results, on components, Euler characteristic and orbifold points, lead to a complete description of the topology of in genus 3 and 4.

Theorem 5.6 (Zachhuber).

In genus , whenever has two components they are homeomorphic orbifolds.

Theorem 5.7 (Möller).

In genus , each component of satisfies

In genus , we have .

Theorem 5.8 (Torres–Zachhuber).

We have in genus , and in genus . Otherwise, only orbifold points of orders and occur, and can be calculated by elementary means for both and .

For example, in genus 3, if is an even fundamental discriminant, then

There is also a combinatorial method to enumerate the cusps of , used in the proof of Theorem 5.3, and thus the genus (of its components) can be routinely computed as well. Appendix C gives a table of these invariants of for and and .

Comparison to genus 2.

The strategy for proving the results above is similar to the case ; for example, Theorem 5.3 is established via a computer–assisted study of elementary moves connecting cusps of . However several new challenges arise in higher genus. For example, when , a new invariant is required to distinguish the components of , ; the Prym variety is not principally polarized, leading to the factor of in Theorem 5.7; and the combinatorial complexities grow substantially as increases.

6. Multicurves and Coxeter diagrams

In this section we turn to topology, to address the following question:

How can one describe a Teichmüller curve by a finite amount of combinatorial data?

We will see that one such description is provided by a weighted system of simple closed curves on a topological surface of genus .

The intersection pattern between and is recorded by a Coxeter diagram . As observed by Leininger, the spherical diagrams give rise to Teichmüller curves. In particular, the exceptional spherical diagrams and correspond to the sporadic lattice triangles from §3. We will also see that suitably chosen curve systems yield explicit examples of Weierstrass curves in for and . Finally, curve systems whose Coxeter diagrams are grid graphs will play an important role in §7.

Multicurves.

A multicurve is a union of disjoint, essential, simple closed curves, no two of which bound an annulus. A pair of multicurves bind the surface if they meet only in transverse double points, and each component of is a polygonal region with at least 4 sides (running alternately along and ).

Index the components of and so that and . We can then form the symmetric matrix

with . It is convenient to record this matrix by a Coxeter diagram with vertices and edges from vertex to vertex .

We say is orientable if the curves can be oriented so their algebraic intersection numbers satisfy for all .

Next, assign a positive integral weight to each curve (or equivalently, to each vertex of ). Let

denote the spectral radius of the matrix on the right. Let denote a right Dehn twist around , and let

We then have:

Theorem 6.1 (Thurston).

An orientable curve system canonically determines a holomorphic -form , unique up to a real scale factor, such that the multitwists and are realized by affine automorphisms satisfying

The trace field of is given by , and inherits the symmetries of the data .

Sketch of the proof.

By the Perron–Frobenius theorem, there is a positive eigenvector , unique up to scale, such that

Take one rectangle for each , , and glue to whenever and are joined by an edge of the one–complex . The result is a holomorphic 1-form such that each is represented by a horizontal or vertical cylinder of height , circumference , and modulus . It follows that the twists fit together to give a pair of globally defined affine automorphisms, and , with the indicated derivatives. To compute the trace field, observe that and apply equation Equation 2.9.

Remarks.

(1)

The stratum containing can also be read off from the topological data : the zeros of of order correspond to components of with sides.

(2)

Every Teichmüller curve can be specified by a multicurve system . To see this, let be a generator of , and let be the multicurve system coming from two different cylinder compositions of and their moduli. The corresponding 1-form is then affinely isomorphic to , so it also generates .

(3)

Conversely, if , then generates a Teichmüller curve. Indeed, the finite volume region in defined by meets every orbit of the subgroup generated by and , so these two elements already generate a lattice in .

(4)

The fact that is an eigenvalue of a real symmetric matrix shows that the trace field of any Teichmüller curve is totally real, as mentioned in §2.

Relation to Coxeter groups.

Suppose for all ; in other words, suppose that any pair of loops in the multicurve system meet at most once. Then the graph is a traditional Coxeter diagram, describing a reflection group acting on and preserving the quadratic form . This diagram is spherical if is positive–definite, or equivalently, if . In this case , where means we set all weights ; so by Remark (3) above, we have:

Theorem 6.2 (Leininger).

Whenever is a spherical Coxeter diagram, the -form associated to generates a Teichmüller curve .

In brief, we say generates .

Regular polygons revisited.

The (simply–laced) spherical Coxeter diagrams are well–known and shown in Figure 6.1, together with their corresponding curve systems (which are uniquely determined).

Note that and are infinite series, and violates our assumptions in a minor way, since it has two parallel curves. We have already seen the corresponding series of Teichmüller curves: they come from the regular polygons with sides. In fact, corresponds to , and corresponds to . (When is odd, the form is not primitive, and a double covering also intervenes.)

The three sporadic Teichmüller curves.

As Leininger shows, the exceptional diagrams are also related to billiards in the triangles listed in Theorem 3.9(C).

Theorem 6.3.

The Coxeter diagrams , and generate the same Teichmüller curves as the three sporadic triangles. All three curves are primitive.

The invariants of these sporadic Teichmüller curves are summarized in Table 6.1. The notation indicates that the Veech groups for and do not contain (see Appendix A); no other primitive Teichmüller curves are known with this property. The absence of can be traced to the asymmetry of the and diagrams.

We note that the numbers and , describing the Veech groups in Table 6.1, are simply half the Coxeter numbers and for the corresponding diagrams.

It is also easy to give algebraic generators for the sporadic Teichmüller curves. In general, the unfolding of an triangle gives, up to a complex factor, the 1-form

on the curve defined by

(assuming ). For we obtain , and similarly for and .

Weierstrass curves in genus 3 and 4.

To conclude, we will show how Coxeter diagrams can be used to construct explicit 1-forms generating infinitely many primitive Teichmüller curves for and .

Let be one of the three polygons shown in Figure 6.2. By gluing together parallel sides of , we obtain a closed surface of genus or , depending on the shape of .

The surface decomposes into horizontal and vertical cylinders, which define a multicurve system whose Coxeter diagram is shown below . The numeric labels indicate the correspondence between the cylinders of and the vertices of .

Note that the L and S shaped polygons gives diagrams isomorphic to and ; we will denote the final diagram by . Rotation of through gives an involution inducing an automorphism of . A system of weights on the vertices of is symmetric if it is invariant under . We can now state:

Theorem 6.4.

Let be a set of symmetric weights on the , or diagram, such that is quadratic over . Then the corresponding 1-form generates a Teichmüller curve

for some , with trace field .

Proof.

Due to the symmetry of , there is an involution such that . It is readily verified that and that has a unique zero. By Theorem 6.1, has trace field . A generalization of Theorem 4.9 then implies that is an eigenform for real multiplication by on , and hence for some as above.

Note that a solution to the eigenvalue equation Equation 6.1 gives explicit lengths for the edges of such that generates a component of . By varying the weights , we obtain infinitely many fields , and hence infinitely many Teichmüller curves in genus 2, 3 and 4.

Remark: Three views of the pentagon curve.

As a particular case, we now have three different constructions of the Teichmüller curve : it arises from the golden L-shaped table; from the regular pentagon; and from the Coxeter diagram.

7. Higher genus

In this section we describe the Bouw–Möller series of primitive Teichmüller curves , indexed by pairs of integers with . We refer to as a vertical series, since as . The curve depends only on the unordered pair ; however each curve has two natural generators, and .

For , the only know primitive Teichmüller curves in come from the Bouw–Möller series. The regular polygons, and the infinite series of lattice triangles of type (see Theorem 3.9), correspond to with and respectively. We will see that every is generated by billiards in a generalized polygon , and that the corresponding Veech group is commensurable to a triangle group.

Semiregular polygons.

The original approach of Bouw and Möller emphasized algebraic geometry. Here we will define using polygons, following Hooper. The two definitions coincide, as can be seen by comparing algebraic expressions for their generating 1-forms.

A semiregular polygon is a -sided polygon, with equal internal angles, whose sides alternate in length between the values and . We also allow or to equal zero, in which case becomes a regular -sided polygon. A polygon is semiregular iff all its angles are equal, and its vertices lie on a circle.

Polygonal generators for .

Let be integers with . The Teichmüller curve is generated by a form constructed by gluing together a sequence of semiregular -gons , .

To define these, let , . Let be a copy of , for . Note that and are regular polygons with sides.

We aim to glue to for each , so we rotate these polygons to make the corresponding sides of length parallel. We then define

The prime indicates that, as in the unfolding process Equation 3.1, we identify with if they are equal up to translation. Equivalently, if , we only take the union from to ; otherwise, we take the union from to . The form is well–defined up to a complex multiple.

Let denote the grid graph, whose vertices coincide with the integral points inside , and whose edges connect points that are distance one apart.

Theorem 7.1 (Hooper).

The forms and generate the same Teichmüller curve .

The curve can also specified by a multicurve system on whose Coxeter diagram is the grid graph .

Theorem 7.2 (Bouw–Möller).

The Veech group of is commensurable to , and is primitive provided .

A sketch of the proofs is deferred to the end of this section. See Theorem 7.6 for a more precise statement of Theorem 7.2.

Examples.

The polygons are shown in several cases in Figure 7.1. Figure 7.2 illustrates an example where both and are even.

The Teichmüller curves generated by -sided regular polygons coincide with . Indeed, is just the usual unfolding of a regular -gon. On the other hand, these curves are also generated by the forms , which are built out of rectangles. For example, the top row in Figure 7.1 gives the form which, as we have seen in §3, generates the same curve as the regular pentagon.

Algebraic generators.

We now turn an algebraic description of . Recall that denotes the Chebyshev polynomial of degree , characterized by equation Equation 1.1. We let denote the unique polynomial satisfying

and having positive leading coefficient. Let

Theorem 7.3.

The form is given by

on the curve defined by

Note that the automorphism shows that for all and . Observing that the zeros of lie over , we obtain:

Corollary 7.4.

The form lies in the stratum , where

if or is odd, and

if both and are even. In either case, .

Billiards.

Let us say a polygon has type if its internal angles, in order, are proportional to these values. By inspecting the algebraic formulas for , we also obtain:

Corollary 7.5.

The form is the unfolding of a generalized polygon of type

where . The billiard flow in has optimal dynamics.

Example.

The quadrilateral , of type , can be assembled from four pie-slices taken out of the semiregular polygons ; see Figure 7.3. This construction is used in the proof of both Theorem 7.3 and its Corollary 7.4 above.

Generalized polygons.

Let us now explain the statement of Corollary 7.5 for general in more detail.

The explicit formulas in Theorem 7.3 allow one to regard as a multi-valued form on the Riemann sphere with coordinate ; for example, when is odd, we have

where are the roots of . Since the roots are real, we can choose a single–valued branch of on .

Intrinsically, we can regard

as an abstract compact surface, with a flat Riemannian metric and piecewise–geodesic boundary. Topologically, is a disk; metrically, it can be constructed by gluing together finitely many Euclidean triangles.

We refer to any such metrized disk as a generalized polygon. By solving the equation on , we obtain a locally isometric developing map

Since , we can regard as the unfolding of . Geodesics reflecting off the boundary define a billiard flow in , which has optimal dynamics by a generalization of Theorem 3.2.

Now if is injective, the expression is simply an instance of the Schwarz–Christoffel formula. Thus its image is a polygon isometric to . In general, is only locally injective; but in either case, the exponents appearing in the formula for in Theorem 7.3, determine the internal angles of , yielding Corollary 7.5.

A sampler.

One can visualize as an immersed polygon by drawing the image of its boundary under the developing map; see Figure 7.4. We emphasize that and both generate . For example, the first row in Figure 7.4 generates the same Teichmüller curves as the regular polygons; the same is true of the first column.

The generalized billiard table for is shown in Figure 1.2.

Three–dimensional billiards.

One can imagine an immersed polygon as describing a traditional billiard table not in 2 dimensions, but in 3: a table that is nearly flat, but with some parts of table passing above or below other parts.

The Veech group.

The following more precise version of Theorem 7.2 completes our description of the Bouw–Möller series.

Theorem 7.6.

Provided , the Veech group of is given by

The trace field of is the same as the invariant trace field of , which is given in Appendix A. Invariants of the Bouw–Möller curves in for and are tabulated in Appendix D

Sketch of the proof of Theorems 7.1 and 7.2

We conclude by explaining why the forms constructed from biregular polygons generate Teichmüller curves.

Fix . We can assume that at least one of the polygons among has a horizontal edge, as in Figure 7.1. It is then straightforward to compute the cylinder decomposition of at angles and . The resulting multicurve system, of the form , uniquely determines the orbit of .

It turns out that the associated Coxeter diagram is a rectangular grid, and the isomorphism is reflected by an isomorphism . Thus reversing the roles of and does not change the orbit of . In particular, we obtain an isomorphism

Clearly, contains an element of order , coming from the rotational symmetry of the polygons . Similarly, the isomorphism above provides an element of order . These two elements nearly generate the triangle group. A more precise analysis leads to Theorems 7.2 and 7.6.

8. Gothic curves and the flex locus

Do the horizontal Weierstrass series and the vertical Bouw–Möller series account for all the primitive Teichmüller curves (apart from a few sporadic examples)?

In this section we will see the answer is no. To do so, we will describe the sequence of gothic curves , and their relationship to the remarkable flex locus .

The flex locus gives the first example of a primitive Teichmüller surface, i.e. a totally geodesic variety of dimension two. The curves and the surface are intertwined by the algebraic geometry of cubic curves in the plane and space curves of genus four. This section follows the development by Mukamel, Wright and the author in Reference MMW.

In the next section we will present the last known family of primitive Teichmüller curves, the arabesque series . The terminology for and is inspired by their polygonal models shown in Figure 8.1.

Curves and surfaces.

Let denote the moduli space of compact Riemann surfaces of genus with unordered marked points . We will begin by constructing the flex locus

the gothic locus

and the gothic curves . We will then sketch the proof of the following results.

Theorem 8.1.

The flex locus is a primitive, totally geodesic surface in .

The fact that is totally geodesic will follow from:

Theorem 8.2.

The gothic locus is –invariant.

Theorem 8.3.

Each component of is a primitive Teichmüller curve.

Explicit examples of gothic curves coming from billiards in quadrilaterals will be given in §9. Since these curves are generated by forms in , rather than , we obtain:

Corollary 8.4.

There exist infinitely many primitive Teichmüller curves in genus that are not accounted for by the Weierstrass series .

The gothic locus is an analogue, in genus four, of the stratum in genus two. Each of these 4–dimensional varieties provides a substrate on which one can impose the additional constraint of real multiplication, to obtain the Teichmüller curves and respectively.

Topology.

The following result describes when is nonempty, gives a lower bound on its number of components, and computes its Euler characteristic.

Theorem 8.5 (Möller–Torres-Teigell).

The gothic curve is nonempty iff is a square . It falls into or subcurves of equal Euler characteristic, and one can express in terms of and an elementary sum.

Here runs from to , which is the number of ideals of norm 6 in . Explicitly, we have or for or , respectively; otherwise . See Table 8.1 for a table of invariants of the gothic curves with . It is unknown at present if is irreducible.

Plane cubics.

We now turn to the construction of the flex locus .

First we recall some classical constructions in projective geometry. Every Riemann surface of genus 1 can be realized, in an essentially unique way, as a smooth cubic curve . Algebraically, is the zero locus of a homogeneous cubic polynomial .

Given a point , the polar conic of with respect to is defined by

Projection from to a line defines a rational map

and

Provided , has degree 3 and 6 critical points.

Let denote the tangent line to at . The polar conic picks out the six points such passes through . For each such , there is a unique cocritical point such that

Equivalently, is a fiber of . The six cocritical points of lie on the satellite conic .

The Hessian.

The cubic curve canonically determines a second cubic, its Hessian . The intersection coincides with the 9 flexes of . Classically, one picks a flex to serve as the origin for the group law on ; then the set of all flexes corresponds to the subgroup of points of order 3.

Dusk and dawn.

The Hessian can also be related to the polars of : we have

When , both the polar conic and the satellite conic degenerate to a pair of lines:

To better visualize this solar configuration, imagine that the cubic represents a (strangely shaped) planet, illuminated by rays from the sun . The sun lies on the horizon, as seen from , exactly when the tangent line to at passes through ; equivalently, when . Thus it is twilight at six points of .

In the solar configuration, we can naturally divide these six points into two groups of three, and , which we call dawn and dusk. The 3 rays of dawn also meet at three other points, namely , which we call the codawn points; see Figure 8.2.

The flex locus.

The flex locus records the set of all possible configurations of codawn points . More precisely,

The terminology comes from the fact that at . Since there are two choices for , the fiber of over is actually parameterized by a double cover , classically called the Cayleyan of . Thus is the image of an elliptic surface; however the map is not injective, and itself is birational to .

The gothic locus.

We now turn to the construction of the space from .

Recall that any Riemann surface which is not hyperelliptic admits a canonical embedding

of degree . This embedding is characterized by the property that its hyperplane sections coincide with the zero sets of holomorphic 1-forms . In the case , there exist quadric and cubic hypersurfaces and such that

To facilitate the passage between two and three dimensions, choose an involution on such that

Now consider a solar configuration as above: a cubic curve , a point , and a codawn line with . Choose a second line through , transverse to . We can then find quadric and cubic surfaces , in , such that:

and ;

; and

is the cone over with vertex .

In this way a solar configuration, together with an additional line through , naturally determines a curve of genus 4, namely

See Figure 8.3.

There is a unique hyperplane tangent to along , and hence a 1-form on such that . Because of the tangency, the zeros of have multiplicity two.

The gothic locus consists of all forms arising as above. Since is determined by the solar configuration plus a line through , and is unique up to a scale factor, we have . We also have a natural degree two map

obtained by projection from ; and a natural map , forgetting and the extra line .

The Abelian surface carrying .

We need to introduce one more key player before we can define the gothic curves .

Note that the composition has degree 6. The target of can be naturally identified with the space of lines through .

Let be the elliptic curve obtained as a 2-fold covering of branched over the three lines of dusk and the extra line determining . One can then verify that there is a unique degree three map making the diagram

commute. One also check that

By pulling back line bundles from and to , we obtain an exact sequence

with . We refer to as the hidden Abelian surface attached to . Identifying with a subspace of , equation Equation 8.2 implies:

We have for all forms in the gothic locus.

The gothic curves.

Let denote the 2–dimensional locus where is an eigenform for real multiplication by on . The gothic curve is defined, finally, as the projection of to .

Sketch of the proofs.

To conclude, we sketch the proofs of Theorems 8.1, 8.2 and 8.3. The main point is:

1. is –invariant. Note that the gothic locus is 4–dimensional, and that the gothic forms satisfy

Since as well, is locally defined by the condition above, which is given by real linear equations in period coordinates. Invariance under follows.

2. is totally geodesic. The orbits in project to complex geodesics in . Since and , there is a pencil of geodesics through every point of , and hence is totally geodesic.

3. is a finite union of Teichmüller curves. We have seen that is –invariant, and by invariance of quadratic real multiplication, so is . Thus its projection to is a totally geodesic curve .

9. Quadrilaterals

In this section we will see that the gothic curves, and the flex locus, belong to a suite of examples naturally associated to six types of quadrilaterals.

This suite yields:

(1)

six examples of –invariant 4-folds in , for various ;

(2)

three examples of primitive, totally geodesic surfaces in , for various ;

(3)

two distinct series of Teichmüller curves in ; and

(4)

two families of quadrilateral billiard tables with optimal dynamical properties.

In particular, the quadrilaterals of type will yield our last family of Teichmüller curves, the arabesque series . We follow the development in Reference EMMW.

Cyclic forms.

It is convenient to describe the type of a quadrilateral by a quadruple of integers . We require that the integers are positive and relatively prime, that

is even, and that for all .

Each quadruple determines a family of cyclic forms . Such a form can be specified by four distinct branch points in ; it is then given by on the curve defined by

The cyclic forms contain the unfoldings of every quadrilateral with internal angles , in any order; the quadrilateral’s vertices correspond to the branch points .

Let . Note that gives an automorphism of of order , satisfying . Thus the symmetric correspondence

determines a self-adjoint endomorphism of , satisfying

Action of .

The smallest closed invariant set containing the cyclic forms of type is its saturation

Our first result describes this space.

Theorem 9.1.

For each of the six values of  listed in Table 9.1, the saturation of the cyclic forms gives a primitive, irreducible, –dimensional invariant subvariety .

The remarkable feature of these six cases is that, while the action of destroys the cyclic symmetry of forms in , it merely deforms the correspondence as an algebraic cycle. The relation Equation 9.2 persists under deformation, and the original cyclic symmetries of are replaced by an action of the dihedral group

on , satisfying . The main step in the proof of Theorem 9.1 is to show the resulting variety of dihedral forms is 4–dimensional; its closure must then coincide with . The requirement that singles out the six values of in Table 9.1.

Teichmüller curves.

By equation Equation 9.2, is an eigenvector for with eigenvalue . Let denote the full eigenspace. Provided is rational, this subspace is of geometric origin: namely, it comes from a map to an Abelian surface,

Let denote the locus where is an eigenform for real multiplication by on , and let denote its projection to . As in §8, we then obtain:

Theorem 9.2.

For or , locus is a finite union of primitive Teichmüller curves for each discriminant .

These curves come from the two entries in Table 9.1 where ; for the other entries, is irrational.

The arabesque series .

The gothic locus coincides with for , and thus as well. But the curves in the arabesque series, defined by

are new; they constitute the last known series of primitive Teichmüller curves. At present, neither nor is as well–understood as the Weierstrass curves.

Determine the number of components of the curves and . Are any two components of the same curve homeomorphic? What are their topological invariants?

Billiards.

Using the relation with cyclic forms, one can readily give explicit generators for gothic and arabesque Teichmüller curves; see Figure 9.1.

Theorem 9.3.

Billiards in the quadrilateral has optimal dynamics, and the associated cyclic -form generates a gothic Teichmüller curve in , provided is irrational and

for some . The quadrilateral similarly generates an arabesque curve, provided .

Problem 9.4.

The cases and are shown at the left and right, respectively, in Figure 9.1. An unfolding of the polygon on the right appears Figure 8.1.

Teichmüller surfaces.

Finally we describe generalizations of the flex surface. We will see that the perspective of quadrilaterals leads to a unified construction of three remarkable and unexpected Teichmüller surfaces in moduli space.

First suppose that is even. We then have a natural map from to a variety in a moduli of lower dimension. The point is constructed by first forming the quadratic differential , and then marking the poles of .

The fibers of are 2–dimensional exactly when divides one of . To see why this might be the case, observe that for the curve is defined by equation Equation 9.1 with replaced by . If divides , then is only branched over the three points , . Since a configuration of three points on has no moduli, sends the entire 2–dimensional locus to a single point .

When the fibers of are 2–dimensional, the –invariance of shows is a totally geodesic surface, just as in the proof of Theorem 8.1. Summarizing, the final result is:

Theorem 9.5.

The locus is a primitive, irreducible, totally geodesic surface for the three cases indicated in Table 9.1.

The case gives the flex surface in ; the other two examples reside in and . In each case, the cyclic locus projects to the unique point in with a cyclic symmetry of order , or .

Local geodesic flatness.

Here is an indication of why a primitive, totally geodesic variety of dimension greater than one is so unusual—even more unusual than a Teichmüller curve.

For each , let

denote the space of holomorphic quadratic differentials on with at worst simple poles at . This finite–dimensional vector space is endowed with a natural –norm, given by .

Let be a totally geodesic variety of dimension . Given , the cotangent vectors that annihilate give a natural subspace of codimension . Let

The quadratic differentials in are those which generate geodesics in through . By convention, we include in .

Despite its nonlinear definition, the cone is a linear subspace of in all known examples, including those in Theorem 9.5.

The linearity of is automatic when , but for it reflects an unusual property of inside the normed vector space . Indeed, records the supporting hyperplanes for the unit ball in along its intersection with . This ball is a complicated convex body, and for most , the locus defined above is not even a real–analytic set.

It is true that isometric symmetries of may force to be linear. But these symmetries are typically ruled out by the assumption that is primitive. It is thus remarkable that exists at all. As a complement to the problem of classifying Teichmüller curves, we conclude with:

Problem 9.6.

Construct and classify all primitive totally geodesic varieties with .

It seems likely we are still in the age of discovery.

10. Notes and references

§1. Introduction.

For an algebraic perspective on , and the realization of its Deligne–Mumford compactification as a projective variety, see Reference ACG.

The dilatation of a real–linear map is given by the ratio between the major and minor axes of the image of a circle under . The dilatation of an orientation–preserving diffeomorphism between Riemann surfaces is given by , and the Teichmüller metric on is given by

This metric also comes from the natural norm on the cotangent space , which can be identified with the space of holomorphic quadratic differentials on . For background in Teichmüller theory, see, e.g. Reference Ga, Reference Nag and Reference Hub.

Figure 1.2 shows the generalized polygon discussed at the end of §7.

The five known horizontal series of Teichmüller curves all involve real multiplication by an order in a real quadratic field. According to Reference EFW, Thm. 1.5, any horizontal series must have this feature. The proof that these curves are totally geodesic relies on an argument in §4, showing that quadratic real multiplication is –invariant.

§2. Moduli spaces and Teichmüller curves.

For a proof of Theorem 2.2 with embedded and connected, see Reference Vi, Cor. 3.34.

Every complex geodesic is generated by a quadratic differential . We have concentrated on the case where arises from a 1-form, for several reasons:

(1)

Any quadratic differential becomes the square of a 1-form after passing to a 2–fold branched covering of .

(2)

Therefore, every Teichmüller curve generated by a quadratic differential is a close relative of one generated by a primitive 1-form.

(3)

Finally, a holomorphic 1-form represents a cohomology class and a differential on the Jacobian of , providing a bridge to Hodge theory.

For the stated properties of the trace field , see Reference GJ, Thm. 5.5, Reference KS, Thm. 28, Reference Ho2, Thm. 10.2 and Reference HL, Thm. 1.1.

The mild twisting of the action of , mentioned after equation Equation 2.6, arises as follows. Let . The isometric action of on gives an isomorphism sending to . The linear action of on gives a second isomorphism, , sending to . One can then readily check that the bijection , given by , induces the map from to itself. One must conjugate by this map to convert the linear action of into an action by Möbius transformations on . See Reference Mc1, Prop. 3.2.

As a complement to Proposition 2.3, we remark that the condition , , does not imply that divides for all horizontal cylinders . For example, may effect a fractional Dehn twist on some , or even permute these cylinders.

Smillie has shown that generates a Teichmüller curve if and only if is closed in Reference V2, §6.

§3. Billiards.

The fundamental result on the Veech dichotomy, Theorem 3.2, appears in Reference V1, Prop. 2.11. The proof using Reference Mas2 presented here (see also Reference HS2, §1.4) bridges a gap in the original argument, noted in Reference Mc1, §2.

The polygon in Figure 3.2 is a square with a generic rectangle attached. The trajectory shown, with starting slope , would be periodic if were absent.

Rational polygons have special properties. For example, it is unknown if every triangle has a periodic billiard trajectory; cf. Reference Sch1. On the other hand, if has angles in , then it has a dense set of periodic slopes and a full measure set of uniformly distributed directions Reference Mas1, Reference KMS. (These two papers are among the first to use Teichmüller theory to address the dynamics of billiards.)

The algebraic description of the Teichmüller curves generated by regular polygons originates in Reference Loch; see also Reference Mc4, §5 and Reference EMMW, Appendix A.

For more on relations between the golden table and the regular pentagon, see Reference Mc1, §9 and Fig. 4 and Reference DL. Periodic slopes on general Teichmüller curves are studied in Reference Mc10.

Question 3.8 is implicit in Reference HMTY, and stated explicitly in Reference Bo as Conjecture 1.4; see these works for more on periodic slopes in the regular heptagon. Winsor has shown that has more than two orbits in , contrary to a conjecture in Reference HMTY, §4.1; e.g. lies outside the previously known orbits. Experimentally, is not fixed by any hyperbolic or parabolic element in when is odd.

The discovery and classification of the lattice triangles appearing in Theorem 3.9 is contained in the work of several authors. Series A, along with the series of regular polygons, is discussed in Reference V1. For series B, see Reference Vo2, Thm. 4.4 and Reference Wa. Two of the three sporadic triangles in C also appear in Reference Vo2; for the example, see Reference KS. Example D is from Reference Ho1. The proof that this list of lattice triangles is complete, apart from the obtuse, scalene case, is given in Reference KS, Reference Pu1 and Reference Pu2.

For recent work on Question 3.11, see Reference LNZ.

The lattice polygons of genus two are classified in Reference Mc4.

Despite the Veech dichotomy, long, periodic trajectories in a lattice polygon can be unevenly distributed; see Reference DL, Reference Mc9.

§4. Genus 2.

Every form of genus two can be presented, in infinitely many ways, as the connect sum of a pair of forms of genus one. This perspective plays a central role in the proof of Theorems 4.1, 4.2 and 4.5, in Reference Mc2, Reference Mc1 and Reference Mc4, respectively; it also leads to an explicit classification of the orbit closures and the ergodic invariant measures for acting on Reference Mc5.

Veech groups for several are shown in Reference Mc1, Fig. 5 and Reference Mc4, Fig. 2. Theorem 4.3 is given in detail in Reference Mc2, Cor. 1.3. Theorem 4.5 leads to many simple examples of 1-forms such that is an infinitely generated group Reference Mc4, Thm. 1.3. For related work, see Reference HS1.

Theorem 4.7 and formula Equation 4.2, appear in Reference Ba, Thms. 1.1 and 2.12. In Reference MZ, Thm. 9.1, Möller and Zagier show is the zero locus of an explicit Hilbert modular form of weight , namely

The Euler characteristic of is directly related to the weights of this form; in brief, and give line bundles over , and

since up to boundary terms.

For Theorem 4.8 on elliptic points, and the remarks on that follow, see Reference Mu1. Algebraic models for with and are given in Reference BM1, and for all fundamental discriminants in Reference KM2.

Theorem 4.6 can be deduced from the fact that and are Galois conjugate Reference BM1, Thm. A. It also follows directly from the calculations in Reference Mc2, Reference Ba and Reference Mu1, which show the topological invariants of do not depend on .

Theorem 4.12 is contained in Reference Mo1 and Reference Mo2.

The classification of nonprimitive Teichmüller curves in is still an open problem; see Reference Du for the conjectural answer and recent progress. Integral polynomials defining , given in Reference Mu2, behave well and suggest an arithmetic theory of Teichmüller curves remains to be developed.

§5. Genus 3 and 4.

The definition of using Prym varieties, and the proof of Theorem 5.1, appear in Reference Mc3. Examples of Veech groups for Weierstrass curves in genus and are shown in Reference Mc3, Fig. 2.

The classification of Weierstrass curves given in Theorem 5.3 appears in Reference LN1 and Reference LN2.

The finiteness Theorem 5.5 appears to be stated here for first time. It follows from two known results. First, by Reference EFW, Thm. 1.5, all but finitely many primitive Teichmüller curves in are obtained by imposing quadratic real multiplication on a 4–dimensional, rank 2, –invariant subvariety , defined over in period coordinates. Second, the classification of such in genus three is now known; see Reference AN, Thm. 1.1 and the references therein. There are three examples; but two of them are contained in the hyperelliptic locus, and consist of forms pulled back from genus two, so the associated Teichmüller curves are not primitive. The one remaining possibility for gives rise to the Weierstrass curves .

For progress toward making Theorem 5.5 more effective, see Reference BaM, Thm. 1.6 and Reference LM. A related result, Reference EFW, Cor. 1.6, states that for each there are only finitely many Teichmüller curves in with trace field of degree 3 or more. For general finiteness results in hyperelliptic strata, see Reference Ap.

The formulas for in Theorem 5.7 appear in Reference Mo5, Thm. 0.2; their calculation uses theta functions and Hilbert modular forms (for the case , see equation Equation 10.1).

See Reference Za for Theorem 5.6, and Reference TZ1 and Reference TZ2 for Theorem 5.8.

§6. Multicurves and Coxeter diagrams.

Thurston’s multicurve construction appears in Reference Th. Our presentation follows Reference Mc3, §4, which also proves Theorem 6.4. The set of all multicurve systems encoding a given Teichmüller curve is described in Reference Mc9, §10.

Coxeter groups and their diagrams appear in many fields of mathematics, ranging from Lie groups and sphere packings to singularity theory; useful references include Reference Bou and Reference Hum. Theorem 6.2, with different terminology, is contained in Reference Lei, Thm. 7.1. The formulation and short proof we present here emphasize the direct connection with Coxeter groups. Theorem 6.3, relating the sporadic lattice triangles to the diagrams, was proposed in Reference Mc11 and proved in Reference Lei, §7.5.

We remark that the Coxeter element for , and the multitwists and , both act on and are related by

This explains why Coxeter numbers (which give the order of ) appear in Table 6.1 (up to a factor of two).

We note that the triangle unfolds to a Riemann surface which lies on the unique Shimura–Teichmüller curve . Indeed , initially defined by , is isomorphic to (by the change of variables ), and the latter curve represents a point on by Reference Mo4, Thm. 5.1.

The Teichmüller curve is studied in Reference CK.

§7. Higher genus.

The original construction of appears in Reference BM2. Its reformulation in terms of semiregular polygons was discovered independently by Mukamel and Hooper. We follow Reference Ho2 for much of this section; see also Reference Wr1.

Both and are separable polynomials of degree with integral coefficients. In terms of the Chebyshev polynomials of the second kind, we have

The roots of occur when and ; those of occur when , but .

For Theorem 7.3, see Reference BM2, Thm. 6.14 and Reference Ho2, Thm. 4.9. We have streamlined the formulas by using the intermediate variable and Chebyshev polynomials, to better display the integrality of the coefficients.

Theorem 7.6 is based on Reference Ho2, Thm. 4.1.

Note that gives an explicit, embedded lattice polygon for all . As can be seen in Figure 7.4, is also embedded for 8 other values of , and its interior is embedded for 2 more values. This list of lattice polygons broadens that appearing in Reference BM2, §8.

The rectilinear cousin of the -sided regular polygon is discussed in detail in Reference Mc6, §13, from the perspective of braid groups and the diagram.

§8. Gothic curves and the flex locus.

Theorems 8.1, 8.2, 8.3, and Corollary 8.4 are taken from Reference MMW. For Theorem 8.5, see Reference MT, Thm. 11.1; Table 8.1 is an excerpt of Table 1 in the same reference. We note that can also be defined by imposing real multiplication on the dual Abelian surface ; this equivalent perspective is adopted in Reference MT.

For the classical theory of plane cubics, their polars, the Hessian and related topics, see Reference Sal.

§9. Quadrilaterals.

All theorems stated in this section appear in Reference EMMW, §1. The totally geodesic surface , , is also studied in Reference KM1. Regarding Problem 9.6, Wright has shown Reference Wr4 that for each , there are only finitely many totally geodesic subvarieties with . For more on in the case of the flex locus, see Reference MMW, §4.

Appendix A. Triangle groups

Let be positive integers with . We define the triangle group by

where

and is chosen so that . Its invariant trace field is given by:

This field is a commensurability invariant.

The quotient space is an orbifold isometric to the double of the unique hyperbolic triangle with internal angles , and . We also allow , in which case and has two cusps.

The groups and are conjugate in . When and are both even, we have a natural subgroup of index two

corresponding to the orbifold covering space branched over the points of of orders and . As indicated by the notation, has two cusps. (Note that is isomorphic to .)

When is odd, we let

This group has index two in , and it does not contain . All the other groups above do contain , because .

The group is commensurable to if and only if . The remaining triangle groups are nonarithmetic. For more on triangle groups, see Reference Tak and Reference MR, Ex. 4.9).

Appendix B. Accidental isomorphisms

This supplement describes all overlaps between the series of known primitive Teichmüller curves.

Theorem B.1.

The only overlaps between the three series , the two series and , the series and the sporadic series are the following:

(1)

in genus , and ;

(2)

in genus , and ; and

(3)

in genus , and .

Proof.

The curves in the five horizontal series indexed by are all different, since they are generated by forms in different strata. Moreover and belong to no other series, since their Veech groups do not contain while the others do. Thus we are reduced to identifying overlaps involving , and coincidences between the vertical series and one of the 5 horizontal series.

In fact, the proof of Theorem 6.4 already shows that is the same as the Weierstrass curve in genus 3; see Figure 6.2. The value can be checked using the fact that the order 12 element in the Veech group of has trace , which generates the maximal order in .

Next we consider overlaps between and . A coincidence occurs here if and only if the curve has quadratic trace field, its stratum has the form , and . Examining Table D.1, we find the cases listed above; in all cases, the traces of elliptic elements in the Veech group show is a fundamental discriminant. This argument also shows that does not occur in the series .

Finally, considerations of strata and trace field show there is a unique remaining candidate for an accidental isomorphism: . And in fact this isomorphism can easily be seen geometrically: if we cut a symmetric quadrilateral in half, we obtain two copies of the triangle, which generates .

This shows is at least a component of . To see they are equal, one can use, for example, the fact that their Euler characteristics are both (cf. Table 8.1).

Remark on .

It is also known that the unfolding of the triangle lies in for ; see Reference EMMW, Rmk. 5 in §5.

This fact can be seen geometrically as follows. Consider the cyclic form of type on the curve defined by . The flat metric makes into a symmetric pyramid , whose base is an equilateral triangle. The cone angles of are at the vertices of its base, and at its remaining vertex .

Let be the geodesic triangle with vertices , where is a vertex of the base of , and is its barycenter (see Figure B.1). It is readily verified that is a triangle, that is tiled by six copies of , and that the unfolding of gives the form .

Appendix C. Tables of Weierstrass curves

Tables C.1, C.2 and C.3, based on Reference Mu1, Reference TZ1 and Reference TZ2, give the topological invariants of in for real quadratic discriminants and . More complete tables can be found in these references. The listed topological invariants are the genus , the number of elliptic points of order , denoted , the number of cusps , and the Euler characteristic .

A few points should be kept in mind:

(1)

In genus 2 and 3, has two components when . The given invariants are those for one of these components.

(2)

In genus 2, besides the listed orbifold points of order 2, we have .

(3)

Similarly, in genus 3, we have , and in genus 4 we have .

(4)

For any component of , the trace field is and the stratum of a generator is .

Appendix D. Table of Bouw–Möller curves

Invariants of the Bouw–Möller curves , for , are given in Table D.1.

This table is organized into groups by genus, with increasing within each group. We note that gives the trivial curve in for , and , and that is not primitive; the remaining curves are. For each form generating , there is a unique involution satisfying ; the last column gives the genus of .

The topological invariants of can be read off from its Veech group; for example, every curve has genus zero, and one or two cusps. We use the notation for triangle groups given in Appendix A.

Acknowledgements

I would like to thank M. Bainbridge, J. Boulanger, P. Hooper, P. Hubert, E. Lanneau, M. Möller, R. Mukamel and K. Winsor for many useful and informative discussions.

Table of Contents

  1. Abstract
  2. 1. Introduction
    1. Billiards.
    2. The current catalog.
    3. Completeness?
    4. Question 1.1.
    5. Teichmüller surfaces.
    6. Notes and references.
    7. Outline.
    8. Notation.
  3. 2. Moduli spaces and Teichmüller curves
    1. Problem 2.1.
    2. Moduli space.
    3. Polygons and Riemann surfaces.
    4. Example in genus 2.
    5. Holomorphic 1-forms.
    6. Strata.
    7. From polygons to 1-forms.
    8. Theorem 2.2.
    9. Geometry of a 1-form.
    10. Action of .
    11. Complex geodesics.
    12. Real geodesics.
    13. Teichmüller curves.
    14. Hidden symmetries.
    15. Examples.
    16. Cylinders and parabolics.
    17. Proposition 2.3.
    18. Cusps of .
    19. Proposition 2.4.
    20. Idea of the proof.
    21. Square–tiled surfaces.
    22. Primitivity.
    23. Invariants.
  4. 3. Billiards
    1. Theorem 3.1.
    2. Unfolding.
    3. Two examples.
    4. From billiards to Teichmüller curves.
    5. Theorem 3.2.
    6. Theorem 3.3.
    7. Corollary 3.4.
    8. Algebraic models.
    9. Theorem 3.5.
    10. Even polygons.
    11. The hidden symmetries of the pentagon.
    12. The golden table.
    13. Aside: Is the Veech dichotomy effective?
    14. Question 3.6.
    15. Question 3.7.
    16. Question 3.8.
    17. Lattice triangles.
    18. Theorem 3.9.
    19. Corollary 3.10.
    20. Question 3.11.
    21. Dynamics on moduli space.
    22. Theorem 3.12 (Masur).
    23. Sketch of the proof.
    24. Proof of Theorem 3.2.
    25. Remark: Billiards that hit vertices.
    26. Question 3.13.
  5. 4. Genus 2
    1. Real multiplication.
    2. The Weierstrass curves.
    3. Classification.
    4. Theorem 4.1.
    5. Theorem 4.2.
    6. Theorem 4.3.
    7. Corollary 4.4.
    8. The oasis.
    9. Theorem 4.5.
    10. Classification and synthesis.
    11. Topology of .
    12. Theorem 4.6 (Bouw–Möller).
    13. Theorem 4.7 (Bainbridge).
    14. Theorem 4.8 (Mukamel).
    15. From affine maps to the Jacobian.
    16. Theorem 4.9.
    17. The eigenform locus.
    18. Proof of Theorem 4.1.
    19. Corollary 4.10.
    20. Corollary 4.11.
    21. Irreducibility: Computer-assisted proof.
    22. The exceptional decagon.
    23. Theorem 4.12 (Möller).
    24. Corollary 4.13.
    25. Sketch of the proof of Theorem 4.5.
    26. What is the shape of ?
    27. Problem 4.14.
    28. Remark: Square discriminant.
  6. 5. Genus and
    1. Eigenforms in higher genus.
    2. The Weierstrass curves .
    3. Theorem 5.1.
    4. Genus 2, 3 and 4.
    5. Corollary 5.2.
    6. Classification.
    7. Theorem 5.3 (Lanneau–Nguyen).
    8. Near completeness in genus 3.
    9. Problem 5.4.
    10. Theorem 5.5.
    11. Topology of the Weierstrass curves.
    12. Theorem 5.6 (Zachhuber).
    13. Theorem 5.7 (Möller).
    14. Theorem 5.8 (Torres–Zachhuber).
    15. Comparison to genus 2.
  7. 6. Multicurves and Coxeter diagrams
    1. Multicurves.
    2. Theorem 6.1 (Thurston).
    3. Sketch of the proof.
    4. Remarks.
    5. Relation to Coxeter groups.
    6. Theorem 6.2 (Leininger).
    7. Regular polygons revisited.
    8. The three sporadic Teichmüller curves.
    9. Theorem 6.3.
    10. Weierstrass curves in genus 3 and 4.
    11. Theorem 6.4.
    12. Proof.
    13. Remark: Three views of the pentagon curve.
  8. 7. Higher genus
    1. Semiregular polygons.
    2. Polygonal generators for .
    3. Theorem 7.1 (Hooper).
    4. Theorem 7.2 (Bouw–Möller).
    5. Examples.
    6. Algebraic generators.
    7. Theorem 7.3.
    8. Corollary 7.4.
    9. Billiards.
    10. Corollary 7.5.
    11. Example.
    12. Generalized polygons.
    13. A sampler.
    14. Three–dimensional billiards.
    15. The Veech group.
    16. Theorem 7.6.
    17. Sketch of the proof of Theorems 7.1 and 7.2
  9. 8. Gothic curves and the flex locus
    1. Curves and surfaces.
    2. Theorem 8.1.
    3. Theorem 8.2.
    4. Theorem 8.3.
    5. Corollary 8.4.
    6. Topology.
    7. Theorem 8.5 (Möller–Torres-Teigell).
    8. Plane cubics.
    9. The Hessian.
    10. Dusk and dawn.
    11. The flex locus.
    12. The gothic locus.
    13. The Abelian surface carrying .
    14. The gothic curves.
    15. Sketch of the proofs.
  10. 9. Quadrilaterals
    1. Cyclic forms.
    2. Action of .
    3. Theorem 9.1.
    4. Teichmüller curves.
    5. Theorem 9.2.
    6. The arabesque series .
    7. Billiards.
    8. Theorem 9.3.
    9. Problem 9.4.
    10. Teichmüller surfaces.
    11. Theorem 9.5.
    12. Local geodesic flatness.
    13. Problem 9.6.
  11. 10. Notes and references
    1. §1. Introduction.
    2. §2. Moduli spaces and Teichmüller curves.
    3. §3. Billiards.
    4. §4. Genus 2.
    5. §5. Genus 3 and 4.
    6. §6. Multicurves and Coxeter diagrams.
    7. §7. Higher genus.
    8. §8. Gothic curves and the flex locus.
    9. §9. Quadrilaterals.
  12. Appendix A. Triangle groups
  13. Appendix B. Accidental isomorphisms
    1. Theorem B.1.
    2. Proof.
    3. Remark on .
  14. Appendix C. Tables of Weierstrass curves
  15. Appendix D. Table of Bouw–Möller curves
  16. Acknowledgements

Figures

Figure 1.1.

A periodic billiard trajectory in the regular heptagon.

Graphic without alt text
Figure 1.2.

A generalized billiard table with optimal dynamics.

Graphic without alt text
Figure 2.1.

The polygon glues up to give a Riemann surface of genus 2.

Graphic without alt text
Figure 2.2.

The –orbit of defines a complex geodesic in .

Graphic without alt text
Figure 3.1.

Three billiard trajectories in a regular pentagon.

Graphic without alt text
Figure 3.2.

A billiard trajectory that is neither periodic nor dense.

Graphic without alt text
Figure 3.3.

The double pentagon yields a surface of genus 2.

Graphic without alt text
Figure 3.4.

Unfolding an L-shaped polygon.

Graphic without alt text
Table 3.1.

Invariants of Teichmüller curves generated by -sided regular polygons.

Sides Stratum Trace field
Figure 3.5.

The Teichmüller curve generated by the golden and the regular pentagon .

Graphic without alt text
Figure 3.6.

An escaping geodesic must converge to a cusp.

Graphic without alt text
Figure 3.7.

A billiard path joining two vertices need not have periodic slope; the parallel trajectory at the left is aperiodic.

Graphic without alt text
Figure 4.1.

An orbifold point on .

Graphic without alt text
Figure 6.1.

Spherical Coxeter diagrams and their corresponding curve systems.

Graphic without alt text
Table 6.1.

The sporadic Teichmüller curves.

Diagram Triangle Stratum Trace Field
Figure 6.2.

Polygons of shapes L, S and X yields surfaces of genus 2, 3 and 4.

Graphic without alt text
Figure 7.1.

Each row shows the polygons used to construct , . The top row gives .

Graphic without alt text
Figure 7.2.

For , only the first 4 polygons are used to construct .

Graphic without alt text
Figure 7.3.

Construction of the billiard table .

Graphic without alt text
Figure 7.4.

The generalized billiard tables for .

Graphic without alt text
Figure 8.1.

Polygonal models for forms in and .

Graphic without alt text
Table 8.1.

Euler characteristics of gothic curves.

12 1
24 1
28 2
33 2
40 2
48 1
52 2
57 2
60 1
72 1
73 4
76 2
84 1
88 2
96 1
97 4
Figure 8.2.

The solar configuration.

Graphic without alt text
Figure 8.3.

The canonical curve of genus 4 is the intersection of and the cone over .

Graphic without alt text
Table 9.1.

Six types of quadrilaterals and their associated varieties.

Stratum of
5
6
6
8
8
10
Figure 9.1.

Closed billiard paths in and .

Graphic without alt text
Figure B.1.

A folded triangle that tiles a symmetric pyramid.

Graphic without alt text
Table C.1.

The Weierstrass curves in .

Table C.2.

The Weierstrass curves in .

Table C.3.

The Weierstrass curves in .

Table D.1.

The Bouw–Möller curves with .

StratumTrace field
0
0
0
0
0
0
1
0
0
0
2
1
1
1
0
0
0
1
0
0
0
3
2
2

Mathematical Fragments

Equation (1.1)
Theorem 2.2.

Every element of can be presented in the form

for a suitable polygon .

Equation (2.3)
Equation (2.5)
Equation (2.6)
Equation (2.7)
Proposition 2.3.

Suppose contains a parabolic element fixing the line of slope through the origin. Then is tiled by a family of cylinders of slope , with rational ratios of moduli.

Equation (2.9)
Theorem 3.1.

Billiards in a regular polygon has optimal dynamics.

Equation (3.1)
Theorem 3.2.

The billiard flow in a lattice polygon has optimal dynamics.

Theorem 3.3.

Every regular polygon generates a primitive Teichmüller curve. In particular, is a lattice polygon.

Question 3.8.

Is every point the fixed point of a parabolic or hyperbolic element of ?

Theorem 3.9.

Triangles of the following types generate Teichmüller curves:

A.

, and , ;

B.

, odd;

C.

, and ; and

D.

.

Any other lattice triangle must be scalene and obtuse, like examples B and D above.

Question 3.11.

Is the list of lattice triangles above complete?

Theorem 3.12 (Masur).

If the Teichmüller geodesic ray generated by is recurrent in , then every horizontal geodesic in is uniformly distributed.

Equation (4.1)
Theorem 4.1.

Each component of is a primitive Teichmüller curve.

Theorem 4.2.

The curve is irreducible unless , in which case it has two components.

Theorem 4.3.

Every Teichmüller curve in is generated by billiards in an explicit -shaped table.

Theorem 4.5.

The regular decagon generates the only other primitive Teichmüller curve in .

Theorem 4.6 (Bouw–Möller).

If has two components, they are homeomorphic as orbifolds.

Theorem 4.7 (Bainbridge).

We have .

Equation (4.2)
Theorem 4.8 (Mukamel).

For , the orbifold points on all have order . Their their total number is a weighted sum of class numbers of the quadratic imaginary orders with discriminants , and .

Equation (4.3)
Equation (4.4)
Equation (4.5)
Equation (4.6)
Theorem 4.9.

Suppose is a form of genus two and the trace field of is quadratic. Then admits real multiplication by , with as an eigenform.

Corollary 4.10.

If and the trace field of is irrational, then is a lattice.

Theorem 4.12 (Möller).

Suppose is a lattice with trace field . Then there is a surjective map from to a compatibly polarized Abelian variety such that

(1)

corresponds to an eigenform for real multiplication by on ; and

(2)

the difference of any two zeros of is torsion in .

Theorem 5.1.

For any and , the Weierstrass curve is a finite union of primitive Teichmüller curves.

Theorem 5.3 (Lanneau–Nguyen).

For genus , is irreducible if is even; it has two components if ; and otherwise, it is empty.

For genus , is irreducible for all .

Theorem 5.5.

The Weierstrass curves account for all but finitely many primitive Teichmüller curves in genus .

Theorem 5.6 (Zachhuber).

In genus , whenever has two components they are homeomorphic orbifolds.

Theorem 5.7 (Möller).

In genus , each component of satisfies

In genus , we have .

Theorem 5.8 (Torres–Zachhuber).

We have in genus , and in genus . Otherwise, only orbifold points of orders and occur, and can be calculated by elementary means for both and .

Theorem 6.1 (Thurston).

An orientable curve system canonically determines a holomorphic -form , unique up to a real scale factor, such that the multitwists and are realized by affine automorphisms satisfying

The trace field of is given by , and inherits the symmetries of the data .

Equation (6.1)
Theorem 6.2 (Leininger).

Whenever is a spherical Coxeter diagram, the -form associated to generates a Teichmüller curve .

Theorem 6.3.

The Coxeter diagrams , and generate the same Teichmüller curves as the three sporadic triangles. All three curves are primitive.

Theorem 6.4.

Let be a set of symmetric weights on the , or diagram, such that is quadratic over . Then the corresponding 1-form generates a Teichmüller curve

for some , with trace field .

Theorem 7.1 (Hooper).

The forms and generate the same Teichmüller curve .

The curve can also specified by a multicurve system on whose Coxeter diagram is the grid graph .

Theorem 7.2 (Bouw–Möller).

The Veech group of is commensurable to , and is primitive provided .

Theorem 7.3.

The form is given by

on the curve defined by

Corollary 7.4.

The form lies in the stratum , where

if or is odd, and

if both and are even. In either case, .

Corollary 7.5.

The form is the unfolding of a generalized polygon of type

where . The billiard flow in has optimal dynamics.

Theorem 7.6.

Provided , the Veech group of is given by

Theorem 8.1.

The flex locus is a primitive, totally geodesic surface in .

Theorem 8.2.

The gothic locus is –invariant.

Theorem 8.3.

Each component of is a primitive Teichmüller curve.

Corollary 8.4.

There exist infinitely many primitive Teichmüller curves in genus that are not accounted for by the Weierstrass series .

Theorem 8.5 (Möller–Torres-Teigell).

The gothic curve is nonempty iff is a square . It falls into or subcurves of equal Euler characteristic, and one can express in terms of and an elementary sum.

Equation (8.2)
Equation (9.1)
Equation (9.2)
Theorem 9.1.

For each of the six values of  listed in Table 9.1, the saturation of the cyclic forms gives a primitive, irreducible, –dimensional invariant subvariety .

Theorem 9.5.

The locus is a primitive, irreducible, totally geodesic surface for the three cases indicated in Table 9.1.

Problem 9.6.

Construct and classify all primitive totally geodesic varieties with .

Equation (10.1)

References

Reference [Ap]
P. Apisa, orbit closures in hyperelliptic components of strata, Duke Math. J. 167 (2018), no. 4, 679–742, DOI 10.1215/00127094-2017-0043. MR3769676,
Show rawAMSref \bib{Apisa:hype}{article}{ label={Ap}, author={Apisa, Paul}, title={$\mathrm {GL}_2\mathbb {R}$ orbit closures in hyperelliptic components of strata}, journal={Duke Math. J.}, volume={167}, date={2018}, number={4}, pages={679--742}, issn={0012-7094}, review={\MR {3769676}}, doi={10.1215/00127094-2017-0043}, }
Reference [ACG]
E. Arbarello, M. Cornalba, and P. A. Griffiths, Geometry of algebraic curves. Volume II, Grundlehren der mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], vol. 268, Springer, Heidelberg, 2011. With a contribution by Joseph Daniel Harris, DOI 10.1007/978-3-540-69392-5. MR2807457,
Show rawAMSref \bib{Arbarello:Cornalba:Griffiths:book:II}{book}{ label={ACG}, author={Arbarello, Enrico}, author={Cornalba, Maurizio}, author={Griffiths, Phillip A.}, title={Geometry of algebraic curves. Volume II}, series={Grundlehren der mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]}, volume={268}, note={With a contribution by Joseph Daniel Harris}, publisher={Springer, Heidelberg}, date={2011}, pages={xxx+963}, isbn={978-3-540-42688-2}, review={\MR {2807457}}, doi={10.1007/978-3-540-69392-5}, }
Reference [AS]
P. Arnoux and T. A. Schmidt, Veech surfaces with nonperiodic directions in the trace field, J. Mod. Dyn. 3 (2009), no. 4, 611–629, DOI 10.3934/jmd.2009.3.611. MR2587089,
Show rawAMSref \bib{Arnoux:Schmidt:SAF}{article}{ label={AS}, author={Arnoux, Pierre}, author={Schmidt, Thomas A.}, title={Veech surfaces with nonperiodic directions in the trace field}, journal={J. Mod. Dyn.}, volume={3}, date={2009}, number={4}, pages={611--629}, issn={1930-5311}, review={\MR {2587089}}, doi={10.3934/jmd.2009.3.611}, }
Reference [AN]
D. Aulicino and D.-M. Nguyen, Rank 2 affine manifolds in genus 3, J. Differential Geom. 116 (2020), no. 2, 205–280, DOI 10.4310/jdg/1603936812. MR4168204,
Show rawAMSref \bib{Aulicino:Nguyen:genus3}{article}{ label={AN}, author={Aulicino, David}, author={Nguyen, Duc-Manh}, title={Rank 2 affine manifolds in genus 3}, journal={J. Differential Geom.}, volume={116}, date={2020}, number={2}, pages={205--280}, issn={0022-040X}, review={\MR {4168204}}, doi={10.4310/jdg/1603936812}, }
Reference [Ba]
M. Bainbridge, Euler characteristics of Teichmüller curves in genus two, Geom. Topol. 11 (2007), 1887–2073, DOI 10.2140/gt.2007.11.1887. MR2350471,
Show rawAMSref \bib{Bainbridge:thesis}{article}{ label={Ba}, author={Bainbridge, Matt}, title={Euler characteristics of Teichm\"{u}ller curves in genus two}, journal={Geom. Topol.}, volume={11}, date={2007}, pages={1887--2073}, issn={1465-3060}, review={\MR {2350471}}, doi={10.2140/gt.2007.11.1887}, }
Reference [BaM]
M. Bainbridge and M. Möller, The Deligne-Mumford compactification of the real multiplication locus and Teichmüller curves in genus 3, Acta Math. 208 (2012), no. 1, 1–92, DOI 10.1007/s11511-012-0074-6. MR2910796,
Show rawAMSref \bib{Bainbridge:Moeller:g3}{article}{ label={BaM}, author={Bainbridge, Matt}, author={M\"{o}ller, Martin}, title={The Deligne-Mumford compactification of the real multiplication locus and Teichm\"{u}ller curves in genus 3}, journal={Acta Math.}, volume={208}, date={2012}, number={1}, pages={1--92}, issn={0001-5962}, review={\MR {2910796}}, doi={10.1007/s11511-012-0074-6}, }
Reference [Bo]
J. Boulanger, Central points of the double heptagon translation surface are not connexion points, arXiv:2009.01748, 2020.
Reference [Bou]
N. Bourbaki, Groupes et algèbres de Lie, Ch. IV–VI, Actualites Scientifiques et Industrielles, 1337, Hermann, Paris, 1968; Masson, Paris, 1981.
Reference [BM1]
I. I. Bouw and M. Möller, Differential equations associated with nonarithmetic Fuchsian groups, J. Lond. Math. Soc. (2) 81 (2010), no. 1, 65–90, DOI 10.1112/jlms/jdp059. MR2580454,
Show rawAMSref \bib{Bouw:Moeller:genus2}{article}{ label={BM1}, author={Bouw, Irene I.}, author={M\"{o}ller, Martin}, title={Differential equations associated with nonarithmetic Fuchsian groups}, journal={J. Lond. Math. Soc. (2)}, volume={81}, date={2010}, number={1}, pages={65--90}, issn={0024-6107}, review={\MR {2580454}}, doi={10.1112/jlms/jdp059}, }
Reference [BM2]
I. I. Bouw and M. Möller, Teichmüller curves, triangle groups, and Lyapunov exponents, Ann. of Math. (2) 172 (2010), no. 1, 139–185, DOI 10.4007/annals.2010.172.139. MR2680418,
Show rawAMSref \bib{Bouw:Moeller:tris}{article}{ label={BM2}, author={Bouw, Irene I.}, author={M\"{o}ller, Martin}, title={Teichm\"{u}ller curves, triangle groups, and Lyapunov exponents}, journal={Ann. of Math. (2)}, volume={172}, date={2010}, number={1}, pages={139--185}, issn={0003-486X}, review={\MR {2680418}}, doi={10.4007/annals.2010.172.139}, }
Reference [Ca]
K. Calta, Veech surfaces and complete periodicity in genus two, J. Amer. Math. Soc. 17 (2004), no. 4, 871–908, DOI 10.1090/S0894-0347-04-00461-8. MR2083470,
Show rawAMSref \bib{Calta:genus2}{article}{ label={Ca}, author={Calta, Kariane}, title={Veech surfaces and complete periodicity in genus two}, journal={J. Amer. Math. Soc.}, volume={17}, date={2004}, number={4}, pages={871--908}, issn={0894-0347}, review={\MR {2083470}}, doi={10.1090/S0894-0347-04-00461-8}, }
Reference [CK]
M. Costantini and A. Kappes, The equation of the Kenyon-Smillie -Teichmüller curve, J. Mod. Dyn. 11 (2017), 17–41, DOI 10.3934/jmd.2017002. MR3588522,
Show rawAMSref \bib{Costantini:Kappes:E7}{article}{ label={CK}, author={Costantini, Matteo}, author={Kappes, Andr\'{e}}, title={The equation of the Kenyon-Smillie $(2,3,4)$-Teichm\"{u}ller curve}, journal={J. Mod. Dyn.}, volume={11}, date={2017}, pages={17--41}, issn={1930-5311}, review={\MR {3588522}}, doi={10.3934/jmd.2017002}, }
Reference [DL]
D. Davis and S. Lelièvre, Periodic paths on the pentagon, double pentagon and golden L, arXiv:1810.11310, 2018.
Reference [D]
L. DeMarco, The conformal geometry of billiards, Bull. Amer. Math. Soc. (N.S.) 48 (2011), no. 1, 33–52, DOI 10.1090/S0273-0979-2010-01322-7. MR2738905,
Show rawAMSref \bib{DeMarco:billiards}{article}{ label={D}, author={DeMarco, Laura}, title={The conformal geometry of billiards}, journal={Bull. Amer. Math. Soc. (N.S.)}, volume={48}, date={2011}, number={1}, pages={33--52}, issn={0273-0979}, review={\MR {2738905}}, doi={10.1090/S0273-0979-2010-01322-7}, }
Reference [Du]
E. Duryev, Teichmuller Curves in Genus Two: Square-tiled Surfaces and Modular Curves, ProQuest LLC, Ann Arbor, MI, 2018. Thesis (Ph.D.)–Harvard University. MR4187609,
Show rawAMSref \bib{Duryev:square}{book}{ label={Du}, author={Duryev, Eduard}, title={Teichmuller Curves in Genus Two: Square-tiled Surfaces and Modular Curves}, note={Thesis (Ph.D.)--Harvard University}, publisher={ProQuest LLC, Ann Arbor, MI}, date={2018}, pages={116}, isbn={979-8678-13604-6}, review={\MR {4187609}}, }
Reference [EFW]
A. Eskin, S. Filip, and A. Wright, The algebraic hull of the Kontsevich-Zorich cocycle, Ann. of Math. (2) 188 (2018), no. 1, 281–313, DOI 10.4007/annals.2018.188.1.5. MR3815463,
Show rawAMSref \bib{Eskin:Filip:Wright:hull}{article}{ label={EFW}, author={Eskin, Alex}, author={Filip, Simion}, author={Wright, Alex}, title={The algebraic hull of the Kontsevich-Zorich cocycle}, journal={Ann. of Math. (2)}, volume={188}, date={2018}, number={1}, pages={281--313}, issn={0003-486X}, review={\MR {3815463}}, doi={10.4007/annals.2018.188.1.5}, }
Reference [EMMW]
A. Eskin, C. T. McMullen, R. E. Mukamel, and A. Wright, Billiards, quadrilaterals and moduli spaces, J. Amer. Math. Soc. 33 (2020), no. 4, 1039–1086, DOI 10.1090/jams/950. MR4155219,
Show rawAMSref \bib{Eskin:McMullen:Mukamel:Wright:dm}{article}{ label={EMMW}, author={Eskin, Alex}, author={McMullen, Curtis T.}, author={Mukamel, Ronen E.}, author={Wright, Alex}, title={Billiards, quadrilaterals and moduli spaces}, journal={J. Amer. Math. Soc.}, volume={33}, date={2020}, number={4}, pages={1039--1086}, issn={0894-0347}, review={\MR {4155219}}, doi={10.1090/jams/950}, }
Reference [Ga]
F. P. Gardiner, Teichmüller theory and quadratic differentials, Pure and Applied Mathematics (New York), John Wiley & Sons, Inc., New York, 1987. A Wiley-Interscience Publication. MR903027,
Show rawAMSref \bib{Gardiner:book}{book}{ label={Ga}, author={Gardiner, Frederick P.}, title={Teichm\"{u}ller theory and quadratic differentials}, series={Pure and Applied Mathematics (New York)}, note={A Wiley-Interscience Publication}, publisher={John Wiley \& Sons, Inc., New York}, date={1987}, pages={xviii+236}, isbn={0-471-84539-6}, review={\MR {903027}}, }
Reference [Go]
E. Goujard, Sous–variétés totalement géodésiques des espaces de modules de Riemann, Astérisque 430 (2021), 407–424.
Reference [GJ]
E. Gutkin and C. Judge, Affine mappings of translation surfaces: geometry and arithmetic, Duke Math. J. 103 (2000), no. 2, 191–213, DOI 10.1215/S0012-7094-00-10321-3. MR1760625,
Show rawAMSref \bib{Gutkin:Judge:billiards}{article}{ label={GJ}, author={Gutkin, Eugene}, author={Judge, Chris}, title={Affine mappings of translation surfaces: geometry and arithmetic}, journal={Duke Math. J.}, volume={103}, date={2000}, number={2}, pages={191--213}, issn={0012-7094}, review={\MR {1760625}}, doi={10.1215/S0012-7094-00-10321-3}, }
Reference [HMTY]
E. Hanson, A. Merberg, C. Towse, and E. Yudovina, Generalized continued fractions and orbits under the action of Hecke triangle groups, Acta Arith. 134 (2008), no. 4, 337–348, DOI 10.4064/aa134-4-4. MR2449157,
Show rawAMSref \bib{Hanson:Merberg:Towse:Yudovina}{article}{ label={HMTY}, author={Hanson, Elise}, author={Merberg, Adam}, author={Towse, Christopher}, author={Yudovina, Elena}, title={Generalized continued fractions and orbits under the action of Hecke triangle groups}, journal={Acta Arith.}, volume={134}, date={2008}, number={4}, pages={337--348}, issn={0065-1036}, review={\MR {2449157}}, doi={10.4064/aa134-4-4}, }
Reference [Ho1]
W. P. Hooper, Another Veech triangle, Proc. Amer. Math. Soc. 141 (2013), no. 3, 857–865, DOI 10.1090/S0002-9939-2012-11379-6. MR3003678,
Show rawAMSref \bib{Hooper:Veech}{article}{ label={Ho1}, author={Hooper, W. Patrick}, title={Another Veech triangle}, journal={Proc. Amer. Math. Soc.}, volume={141}, date={2013}, number={3}, pages={857--865}, issn={0002-9939}, review={\MR {3003678}}, doi={10.1090/S0002-9939-2012-11379-6}, }
Reference [Ho2]
W. P. Hooper, Grid graphs and lattice surfaces, Int. Math. Res. Not. IMRN 12 (2013), 2657–2698, DOI 10.1093/imrn/rns124. MR3071661,
Show rawAMSref \bib{Hooper:grids}{article}{ label={Ho2}, author={Hooper, W. Patrick}, title={Grid graphs and lattice surfaces}, journal={Int. Math. Res. Not. IMRN}, date={2013}, number={12}, pages={2657--2698}, issn={1073-7928}, review={\MR {3071661}}, doi={10.1093/imrn/rns124}, }
Reference [Hub]
J. H. Hubbard, Teichmüller theory and applications to geometry, topology, and dynamics. Vol. 2: Surface homeomorphisms and rational functions, Matrix Editions, Ithaca, NY, 2016. MR3675959,
Show rawAMSref \bib{Hubbard:book:T1}{book}{ label={Hub}, author={Hubbard, John Hamal}, title={Teichm\"{u}ller theory and applications to geometry, topology, and dynamics. Vol. 2}, subtitle={Surface homeomorphisms and rational functions}, publisher={Matrix Editions, Ithaca, NY}, date={2016}, pages={x+262}, isbn={978-1-943863-00-6}, review={\MR {3675959}}, }
Reference [HL]
P. Hubert and E. Lanneau, Veech groups without parabolic elements, Duke Math. J. 133 (2006), no. 2, 335–346, DOI 10.1215/S0012-7094-06-13326-4. MR2225696,
Show rawAMSref \bib{Hubert:Lanneau:twists}{article}{ label={HL}, author={Hubert, Pascal}, author={Lanneau, Erwan}, title={Veech groups without parabolic elements}, journal={Duke Math. J.}, volume={133}, date={2006}, number={2}, pages={335--346}, issn={0012-7094}, review={\MR {2225696}}, doi={10.1215/S0012-7094-06-13326-4}, }
Reference [HS1]
P. Hubert and T. A. Schmidt, Infinitely generated Veech groups, Duke Math. J. 123 (2004), no. 1, 49–69, DOI 10.1215/S0012-7094-04-12312-8. MR2060022,
Show rawAMSref \bib{Hubert:Schmidt:infgen}{article}{ label={HS1}, author={Hubert, Pascal}, author={Schmidt, Thomas A.}, title={Infinitely generated Veech groups}, journal={Duke Math. J.}, volume={123}, date={2004}, number={1}, pages={49--69}, issn={0012-7094}, review={\MR {2060022}}, doi={10.1215/S0012-7094-04-12312-8}, }
Reference [HS2]
P. Hubert and T. A. Schmidt, An introduction to Veech surfaces, Handbook of dynamical systems. Vol. 1B, Elsevier B. V., Amsterdam, 2006, pp. 501–526, DOI 10.1016/S1874-575X(06)80031-7. MR2186246,
Show rawAMSref \bib{Hubert:Schmidt:survey}{article}{ label={HS2}, author={Hubert, Pascal}, author={Schmidt, Thomas A.}, title={An introduction to Veech surfaces}, conference={ title={Handbook of dynamical systems. Vol. 1B}, }, book={ publisher={Elsevier B. V., Amsterdam}, }, date={2006}, pages={501--526}, review={\MR {2186246}}, doi={10.1016/S1874-575X(06)80031-7}, }
Reference [Hum]
J. E. Humphreys, Reflection groups and Coxeter groups, Cambridge Studies in Advanced Mathematics, vol. 29, Cambridge University Press, Cambridge, 1990, DOI 10.1017/CBO9780511623646. MR1066460,
Show rawAMSref \bib{Humphreys:book:refl}{book}{ label={Hum}, author={Humphreys, James E.}, title={Reflection groups and Coxeter groups}, series={Cambridge Studies in Advanced Mathematics}, volume={29}, publisher={Cambridge University Press, Cambridge}, date={1990}, pages={xii+204}, isbn={0-521-37510-X}, review={\MR {1066460}}, doi={10.1017/CBO9780511623646}, }
Reference [KS]
R. Kenyon and J. Smillie, Billiards on rational-angled triangles, Comment. Math. Helv. 75 (2000), no. 1, 65–108, DOI 10.1007/s000140050113. MR1760496,
Show rawAMSref \bib{Kenyon:Smillie:billiards}{article}{ label={KS}, author={Kenyon, Richard}, author={Smillie, John}, title={Billiards on rational-angled triangles}, journal={Comment. Math. Helv.}, volume={75}, date={2000}, number={1}, pages={65--108}, issn={0010-2571}, review={\MR {1760496}}, doi={10.1007/s000140050113}, }
Reference [KMS]
S. Kerckhoff, H. Masur, and J. Smillie, Ergodicity of billiard flows and quadratic differentials, Ann. of Math. (2) 124 (1986), no. 2, 293–311, DOI 10.2307/1971280. MR855297,
Show rawAMSref \bib{KMS:billiards}{article}{ label={KMS}, author={Kerckhoff, Steven}, author={Masur, Howard}, author={Smillie, John}, title={Ergodicity of billiard flows and quadratic differentials}, journal={Ann. of Math. (2)}, volume={124}, date={1986}, number={2}, pages={293--311}, issn={0003-486X}, review={\MR {855297}}, doi={10.2307/1971280}, }
Reference [KM1]
A. Kumar and R. E. Mukamel, Real multiplication through explicit correspondences, LMS J. Comput. Math. 19 (2016), no. suppl. A, 29–42, DOI 10.1112/S1461157016000188. MR3540944,
Show rawAMSref \bib{Kumar:Mukamel:real}{article}{ label={KM1}, author={Kumar, Abhinav}, author={Mukamel, Ronen E.}, title={Real multiplication through explicit correspondences}, journal={LMS J. Comput. Math.}, volume={19}, date={2016}, number={suppl. A}, pages={29--42}, review={\MR {3540944}}, doi={10.1112/S1461157016000188}, }
Reference [KM2]
A. Kumar and R. E. Mukamel, Algebraic models and arithmetic geometry of Teichmüller curves in genus two, Int. Math. Res. Not. IMRN 22 (2017), 6894–6942, DOI 10.1093/imrn/rnw193. MR3737325,
Show rawAMSref \bib{Kumar:Mukamel:genus2}{article}{ label={KM2}, author={Kumar, Abhinav}, author={Mukamel, Ronen E.}, title={Algebraic models and arithmetic geometry of Teichm\"{u}ller curves in genus two}, journal={Int. Math. Res. Not. IMRN}, date={2017}, number={22}, pages={6894--6942}, issn={1073-7928}, review={\MR {3737325}}, doi={10.1093/imrn/rnw193}, }
Reference [LM]
E. Lanneau and M. Möller. Non–existence and finiteness results for Teichmüller curves in Prym loci, Exp. Math., to appear.
Reference [LN1]
E. Lanneau and D.-M. Nguyen, Teichmüller curves generated by Weierstrass Prym eigenforms in genus 3 and genus 4, J. Topol. 7 (2014), no. 2, 475–522, DOI 10.1112/jtopol/jtt036. MR3217628,
Show rawAMSref \bib{Lanneau:Nguyen:wgenus3}{article}{ label={LN1}, author={Lanneau, Erwan}, author={Nguyen, Duc-Manh}, title={Teichm\"{u}ller curves generated by Weierstrass Prym eigenforms in genus 3 and genus 4}, journal={J. Topol.}, volume={7}, date={2014}, number={2}, pages={475--522}, issn={1753-8416}, review={\MR {3217628}}, doi={10.1112/jtopol/jtt036}, }
Reference [LN2]
E. Lanneau and D.-M. Nguyen, Weierstrass Prym eigenforms in genus four, J. Inst. Math. Jussieu 19 (2020), no. 6, 2045–2085, DOI 10.1017/s1474748019000057. MR4167002,
Show rawAMSref \bib{Lanneau:Nguyen:wgenus4}{article}{ label={LN2}, author={Lanneau, Erwan}, author={Nguyen, Duc-Manh}, title={Weierstrass Prym eigenforms in genus four}, journal={J. Inst. Math. Jussieu}, volume={19}, date={2020}, number={6}, pages={2045--2085}, issn={1474-7480}, review={\MR {4167002}}, doi={10.1017/s1474748019000057}, }
Reference [LNZ]
A. Larsen, C. Norton, and B. Zykoski, Strongly obtuse rational lattice triangles, Trans. Amer. Math. Soc. 374 (2021), no. 10, 7119–7142, DOI 10.1090/tran/8415. MR4315599,
Show rawAMSref \bib{Larsen:Norton:Zykoski:triangles}{article}{ label={LNZ}, author={Larsen, Anne}, author={Norton, Chaya}, author={Zykoski, Bradley}, title={Strongly obtuse rational lattice triangles}, journal={Trans. Amer. Math. Soc.}, volume={374}, date={2021}, number={10}, pages={7119--7142}, issn={0002-9947}, review={\MR {4315599}}, doi={10.1090/tran/8415}, }
Reference [Lei]
C. J. Leininger, On groups generated by two positive multi-twists: Teichmüller curves and Lehmer’s number, Geom. Topol. 8 (2004), 1301–1359, DOI 10.2140/gt.2004.8.1301. MR2119298,
Show rawAMSref \bib{Leininger:Dehn}{article}{ label={Lei}, author={Leininger, Christopher J.}, title={On groups generated by two positive multi-twists: Teichm\"{u}ller curves and Lehmer's number}, journal={Geom. Topol.}, volume={8}, date={2004}, pages={1301--1359}, issn={1465-3060}, review={\MR {2119298}}, doi={10.2140/gt.2004.8.1301}, }
Reference [Loch]
P. Lochak, On arithmetic curves in the moduli spaces of curves, J. Inst. Math. Jussieu 4 (2005), no. 3, 443–508, DOI 10.1017/S1474748005000101. MR2197065,
Show rawAMSref \bib{Lochak:rm}{article}{ label={Loch}, author={Lochak, Pierre}, title={On arithmetic curves in the moduli spaces of curves}, journal={J. Inst. Math. Jussieu}, volume={4}, date={2005}, number={3}, pages={443--508}, issn={1474-7480}, review={\MR {2197065}}, doi={10.1017/S1474748005000101}, }
Reference [MR]
C. Maclachlan and A. W. Reid, The arithmetic of hyperbolic 3-manifolds, Graduate Texts in Mathematics, vol. 219, Springer-Verlag, New York, 2003, DOI 10.1007/978-1-4757-6720-9. MR1937957,
Show rawAMSref \bib{Maclachlan:Reid:book}{book}{ label={MR}, author={Maclachlan, Colin}, author={Reid, Alan W.}, title={The arithmetic of hyperbolic 3-manifolds}, series={Graduate Texts in Mathematics}, volume={219}, publisher={Springer-Verlag, New York}, date={2003}, pages={xiv+463}, isbn={0-387-98386-4}, review={\MR {1937957}}, doi={10.1007/978-1-4757-6720-9}, }
Reference [Mas1]
H. Masur, Closed trajectories for quadratic differentials with an application to billiards, Duke Math. J. 53 (1986), no. 2, 307–314, DOI 10.1215/S0012-7094-86-05319-6. MR850537,
Show rawAMSref \bib{Masur:closed}{article}{ label={Mas1}, author={Masur, Howard}, title={Closed trajectories for quadratic differentials with an application to billiards}, journal={Duke Math. J.}, volume={53}, date={1986}, number={2}, pages={307--314}, issn={0012-7094}, review={\MR {850537}}, doi={10.1215/S0012-7094-86-05319-6}, }
Reference [Mas2]
H. Masur, Hausdorff dimension of the set of nonergodic foliations of a quadratic differential, Duke Math. J. 66 (1992), no. 3, 387–442, DOI 10.1215/S0012-7094-92-06613-0. MR1167101,
Show rawAMSref \bib{Masur:criterion}{article}{ label={Mas2}, author={Masur, Howard}, title={Hausdorff dimension of the set of nonergodic foliations of a quadratic differential}, journal={Duke Math. J.}, volume={66}, date={1992}, number={3}, pages={387--442}, issn={0012-7094}, review={\MR {1167101}}, doi={10.1215/S0012-7094-92-06613-0}, }
Reference [Mas3]
H. Masur, Ergodic theory of translation surfaces, Handbook of dynamical systems. Vol. 1B, Elsevier B. V., Amsterdam, 2006, pp. 527–547, DOI 10.1016/S1874-575X(06)80032-9. MR2186247,
Show rawAMSref \bib{Masur:survey}{article}{ label={Mas3}, author={Masur, Howard}, title={Ergodic theory of translation surfaces}, conference={ title={Handbook of dynamical systems. Vol. 1B}, }, book={ publisher={Elsevier B. V., Amsterdam}, }, date={2006}, pages={527--547}, review={\MR {2186247}}, doi={10.1016/S1874-575X(06)80032-9}, }
Reference [MT]
H. Masur and S. Tabachnikov, Rational billiards and flat structures, Handbook of dynamical systems, Vol. 1A, North-Holland, Amsterdam, 2002, pp. 1015–1089, DOI 10.1016/S1874-575X(02)80015-7. MR1928530,
Show rawAMSref \bib{Masur:Tabachnikov:survey}{article}{ label={MT}, author={Masur, Howard}, author={Tabachnikov, Serge}, title={Rational billiards and flat structures}, conference={ title={Handbook of dynamical systems, Vol. 1A}, }, book={ publisher={North-Holland, Amsterdam}, }, date={2002}, pages={1015--1089}, review={\MR {1928530}}, doi={10.1016/S1874-575X(02)80015-7}, }
Reference [Mc1]
C. T. McMullen, Billiards and Teichmüller curves on Hilbert modular surfaces, J. Amer. Math. Soc. 16 (2003), no. 4, 857–885, DOI 10.1090/S0894-0347-03-00432-6. MR1992827,
Show rawAMSref \bib{McMullen:bild}{article}{ label={Mc1}, author={McMullen, Curtis T.}, title={Billiards and Teichm\"{u}ller curves on Hilbert modular surfaces}, journal={J. Amer. Math. Soc.}, volume={16}, date={2003}, number={4}, pages={857--885}, issn={0894-0347}, review={\MR {1992827}}, doi={10.1090/S0894-0347-03-00432-6}, }
Reference [Mc2]
C. T. McMullen, Teichmüller curves in genus two: discriminant and spin, Math. Ann. 333 (2005), no. 1, 87–130, DOI 10.1007/s00208-005-0666-y. MR2169830,
Show rawAMSref \bib{McMullen:spin}{article}{ label={Mc2}, author={McMullen, Curtis T.}, title={Teichm\"{u}ller curves in genus two: discriminant and spin}, journal={Math. Ann.}, volume={333}, date={2005}, number={1}, pages={87--130}, issn={0025-5831}, review={\MR {2169830}}, doi={10.1007/s00208-005-0666-y}, }
Reference [Mc3]
C. T. McMullen, Prym varieties and Teichmüller curves, Duke Math. J. 133 (2006), no. 3, 569–590, DOI 10.1215/S0012-7094-06-13335-5. MR2228463,
Show rawAMSref \bib{McMullen:Prym}{article}{ label={Mc3}, author={McMullen, Curtis T.}, title={Prym varieties and Teichm\"{u}ller curves}, journal={Duke Math. J.}, volume={133}, date={2006}, number={3}, pages={569--590}, issn={0012-7094}, review={\MR {2228463}}, doi={10.1215/S0012-7094-06-13335-5}, }
Reference [Mc4]
C. T. McMullen, Teichmüller curves in genus two: torsion divisors and ratios of sines, Invent. Math. 165 (2006), no. 3, 651–672, DOI 10.1007/s00222-006-0511-2. MR2242630,
Show rawAMSref \bib{McMullen:tor}{article}{ label={Mc4}, author={McMullen, Curtis T.}, title={Teichm\"{u}ller curves in genus two: torsion divisors and ratios of sines}, journal={Invent. Math.}, volume={165}, date={2006}, number={3}, pages={651--672}, issn={0020-9910}, review={\MR {2242630}}, doi={10.1007/s00222-006-0511-2}, }
Reference [Mc5]
C. T. McMullen, Dynamics of over moduli space in genus two, Ann. of Math. (2) 165 (2007), no. 2, 397–456, DOI 10.4007/annals.2007.165.397. MR2299738,
Show rawAMSref \bib{McMullen:abel}{article}{ label={Mc5}, author={McMullen, Curtis T.}, title={Dynamics of $\mathrm {SL}_2(\mathbb {R})$ over moduli space in genus two}, journal={Ann. of Math. (2)}, volume={165}, date={2007}, number={2}, pages={397--456}, issn={0003-486X}, review={\MR {2299738}}, doi={10.4007/annals.2007.165.397}, }
Reference [Mc6]
C. T. McMullen, Braid groups and Hodge theory, Math. Ann. 355 (2013), no. 3, 893–946, DOI 10.1007/s00208-012-0804-2. MR3020148,
Show rawAMSref \bib{McMullen:bn}{article}{ label={Mc6}, author={McMullen, Curtis T.}, title={Braid groups and Hodge theory}, journal={Math. Ann.}, volume={355}, date={2013}, number={3}, pages={893--946}, issn={0025-5831}, review={\MR {3020148}}, doi={10.1007/s00208-012-0804-2}, }
Reference [Mc7]
C. T. McMullen, Diophantine and ergodic foliations on surfaces, J. Topol. 6 (2013), no. 2, 349–360, DOI 10.1112/jtopol/jts033. MR3065179,
Show rawAMSref \bib{McMullen:dioph}{article}{ label={Mc7}, author={McMullen, Curtis T.}, title={Diophantine and ergodic foliations on surfaces}, journal={J. Topol.}, volume={6}, date={2013}, number={2}, pages={349--360}, issn={1753-8416}, review={\MR {3065179}}, doi={10.1112/jtopol/jts033}, }
Reference [Mc8]
C. T. McMullen, Teichmüller dynamics and unique ergodicity via currents and Hodge theory, J. Reine Angew. Math. 768 (2020), 39–54, DOI 10.1515/crelle-2019-0037. MR4168686,
Show rawAMSref \bib{McMullen:ue}{article}{ label={Mc8}, author={McMullen, Curtis T.}, title={Teichm\"{u}ller dynamics and unique ergodicity via currents and Hodge theory}, journal={J. Reine Angew. Math.}, volume={768}, date={2020}, pages={39--54}, issn={0075-4102}, review={\MR {4168686}}, doi={10.1515/crelle-2019-0037}, }
Reference [Mc9]
C. T. McMullen, Modular symbols for Teichmüller curves, J. Reine Angew. Math. 777 (2021), 89–125, DOI 10.1515/crelle-2021-0019. MR4292865,
Show rawAMSref \bib{McMullen:mod}{article}{ label={Mc9}, author={McMullen, Curtis T.}, title={Modular symbols for Teichm\"{u}ller curves}, journal={J. Reine Angew. Math.}, volume={777}, date={2021}, pages={89--125}, issn={0075-4102}, review={\MR {4292865}}, doi={10.1515/crelle-2021-0019}, }
Reference [Mc10]
C. T. McMullen, Billiards, heights, and the arithmetic of non-arithmetic groups, Invent. Math. 228 (2022), no. 3, 1309–1351, DOI 10.1007/s00222-022-01101-4. MR4419633,
Show rawAMSref \bib{McMullen:gold}{article}{ label={Mc10}, author={McMullen, Curtis T.}, title={Billiards, heights, and the arithmetic of non-arithmetic groups}, journal={Invent. Math.}, volume={228}, date={2022}, number={3}, pages={1309--1351}, issn={0020-9910}, review={\MR {4419633}}, doi={10.1007/s00222-022-01101-4}, }
Reference [Mc11]
C. McMullen, Letter to Leininger, et al., 13 July 2003.
Reference [MMW]
C. T. McMullen, R. E. Mukamel, and A. Wright, Cubic curves and totally geodesic subvarieties of moduli space, Ann. of Math. (2) 185 (2017), no. 3, 957–990, DOI 10.4007/annals.2017.185.3.6. MR3664815,
Show rawAMSref \bib{McMullen:Mukamel:Wright:gothic}{article}{ label={MMW}, author={McMullen, Curtis T.}, author={Mukamel, Ronen E.}, author={Wright, Alex}, title={Cubic curves and totally geodesic subvarieties of moduli space}, journal={Ann. of Math. (2)}, volume={185}, date={2017}, number={3}, pages={957--990}, issn={0003-486X}, review={\MR {3664815}}, doi={10.4007/annals.2017.185.3.6}, }
Reference [Mo1]
M. Möller, Periodic points on Veech surfaces and the Mordell-Weil group over a Teichmüller curve, Invent. Math. 165 (2006), no. 3, 633–649, DOI 10.1007/s00222-006-0510-3. MR2242629,
Show rawAMSref \bib{Moeller:tor}{article}{ label={Mo1}, author={M\"{o}ller, Martin}, title={Periodic points on Veech surfaces and the Mordell-Weil group over a Teichm\"{u}ller curve}, journal={Invent. Math.}, volume={165}, date={2006}, number={3}, pages={633--649}, issn={0020-9910}, review={\MR {2242629}}, doi={10.1007/s00222-006-0510-3}, }
Reference [Mo2]
M. Möller, Variations of Hodge structures of a Teichmüller curve, J. Amer. Math. Soc. 19 (2006), no. 2, 327–344, DOI 10.1090/S0894-0347-05-00512-6. MR2188128,
Show rawAMSref \bib{Moeller:Hodge}{article}{ label={Mo2}, author={M\"{o}ller, Martin}, title={Variations of Hodge structures of a Teichm\"{u}ller curve}, journal={J. Amer. Math. Soc.}, volume={19}, date={2006}, number={2}, pages={327--344}, issn={0894-0347}, review={\MR {2188128}}, doi={10.1090/S0894-0347-05-00512-6}, }
Reference [Mo3]
M. Möller, Affine groups of flat surfaces, Handbook of Teichmüller theory. Vol. II, IRMA Lect. Math. Theor. Phys., vol. 13, Eur. Math. Soc., Zürich, 2009, pp. 369–387, DOI 10.4171/055-1/11. MR2497782,
Show rawAMSref \bib{Moeller:survey:affine}{article}{ label={Mo3}, author={M\"{o}ller, Martin}, title={Affine groups of flat surfaces}, conference={ title={Handbook of Teichm\"{u}ller theory. Vol. II}, }, book={ series={IRMA Lect. Math. Theor. Phys.}, volume={13}, publisher={Eur. Math. Soc., Z\"{u}rich}, }, date={2009}, pages={369--387}, review={\MR {2497782}}, doi={10.4171/055-1/11}, }
Reference [Mo4]
M. Möller, Shimura and Teichmüller curves, J. Mod. Dyn. 5 (2011), no. 1, 1–32, DOI 10.3934/jmd.2011.5.1. MR2787595,
Show rawAMSref \bib{Moeller:Shimura}{article}{ label={Mo4}, author={M\"{o}ller, Martin}, title={Shimura and Teichm\"{u}ller curves}, journal={J. Mod. Dyn.}, volume={5}, date={2011}, number={1}, pages={1--32}, issn={1930-5311}, review={\MR {2787595}}, doi={10.3934/jmd.2011.5.1}, }
Reference [Mo5]
M. Möller, Prym covers, theta functions and Kobayashi curves in Hilbert modular surfaces, Amer. J. Math. 136 (2014), no. 4, 995–1021, DOI 10.1353/ajm.2014.0026. MR3245185,
Show rawAMSref \bib{Moeller:Prym}{article}{ label={Mo5}, author={M\"{o}ller, Martin}, title={Prym covers, theta functions and Kobayashi curves in Hilbert modular surfaces}, journal={Amer. J. Math.}, volume={136}, date={2014}, number={4}, pages={995--1021}, issn={0002-9327}, review={\MR {3245185}}, doi={10.1353/ajm.2014.0026}, }
Reference [MT]
M. Möller and D. Torres-Teigell, Euler characteristics of Gothic Teichmüller curves, Geom. Topol. 24 (2020), no. 3, 1149–1210, DOI 10.2140/gt.2020.24.1149. MR4157552,
Show rawAMSref \bib{Moeller:Torres:gothic}{article}{ label={MT}, author={M\"{o}ller, Martin}, author={Torres-Teigell, David}, title={Euler characteristics of Gothic Teichm\"{u}ller curves}, journal={Geom. Topol.}, volume={24}, date={2020}, number={3}, pages={1149--1210}, issn={1465-3060}, review={\MR {4157552}}, doi={10.2140/gt.2020.24.1149}, }
Reference [MZ]
M. Möller and D. Zagier, Modular embeddings of Teichmüller curves, Compos. Math. 152 (2016), no. 11, 2269–2349, DOI 10.1112/S0010437X16007636. MR3577896,
Show rawAMSref \bib{Moeller:Zagier:modular}{article}{ label={MZ}, author={M\"{o}ller, Martin}, author={Zagier, Don}, title={Modular embeddings of Teichm\"{u}ller curves}, journal={Compos. Math.}, volume={152}, date={2016}, number={11}, pages={2269--2349}, issn={0010-437X}, review={\MR {3577896}}, doi={10.1112/S0010437X16007636}, }
Reference [Mu1]
R. E. Mukamel, Orbifold points on Teichmüller curves and Jacobians with complex multiplication, Geom. Topol. 18 (2014), no. 2, 779–829, DOI 10.2140/gt.2014.18.779. MR3180485,
Show rawAMSref \bib{Mukamel:thesis}{article}{ label={Mu1}, author={Mukamel, Ronen E.}, title={Orbifold points on Teichm\"{u}ller curves and Jacobians with complex multiplication}, journal={Geom. Topol.}, volume={18}, date={2014}, number={2}, pages={779--829}, issn={1465-3060}, review={\MR {3180485}}, doi={10.2140/gt.2014.18.779}, }
Reference [Mu2]
R. E. Mukamel, Polynomials defining Teichmüller curves and their factorizations , Exp. Math. 30 (2021), no. 1, 19–31, DOI 10.1080/10586458.2018.1488156. MR4223280,
Show rawAMSref \bib{Mukamel:modp}{article}{ label={Mu2}, author={Mukamel, Ronen E.}, title={Polynomials defining Teichm\"{u}ller curves and their factorizations $\mod \,p$}, journal={Exp. Math.}, volume={30}, date={2021}, number={1}, pages={19--31}, issn={1058-6458}, review={\MR {4223280}}, doi={10.1080/10586458.2018.1488156}, }
Reference [Nag]
S. Nag, The complex analytic theory of Teichmüller spaces, Canadian Mathematical Society Series of Monographs and Advanced Texts, John Wiley & Sons, Inc., New York, 1988. A Wiley-Interscience Publication. MR927291,
Show rawAMSref \bib{Nag:book}{book}{ label={Nag}, author={Nag, Subhashis}, title={The complex analytic theory of Teichm\"{u}ller spaces}, series={Canadian Mathematical Society Series of Monographs and Advanced Texts}, note={A Wiley-Interscience Publication}, publisher={John Wiley \& Sons, Inc., New York}, date={1988}, pages={xiv+427}, isbn={0-471-62773-9}, review={\MR {927291}}, }
Reference [Pu1]
J.-C. Puchta, On triangular billiards, Comment. Math. Helv. 76 (2001), no. 3, 501–505, DOI 10.1007/PL00013215. MR1854695,
Show rawAMSref \bib{Puchta:billiards}{article}{ label={Pu1}, author={Puchta, Jan-Christoph}, title={On triangular billiards}, journal={Comment. Math. Helv.}, volume={76}, date={2001}, number={3}, pages={501--505}, issn={0010-2571}, review={\MR {1854695}}, doi={10.1007/PL00013215}, }
Reference [Pu2]
J.-C. Puchta, Addendum to “On triangular billiards”, Preprint, 2021.
Reference [Sal]
G. Salmon, Higher Plane Curves, Hodges, Foster and Figgis, Dublin, 1879.
Reference [Sch1]
R. E. Schwartz, Obtuse triangular billiards. II. One hundred degrees worth of periodic trajectories, Experiment. Math. 18 (2009), no. 2, 137–171. MR2549685,
Show rawAMSref \bib{Schwartz:100}{article}{ label={Sch1}, author={Schwartz, Richard Evan}, title={Obtuse triangular billiards. II. One hundred degrees worth of periodic trajectories}, journal={Experiment. Math.}, volume={18}, date={2009}, number={2}, pages={137--171}, issn={1058-6458}, review={\MR {2549685}}, }
Reference [Sch2]
R. Schwartz, Billiards from the square to the stadium, ICM Proceedings, 2022, to appear.
Reference [Tak]
K. Takeuchi, Arithmetic triangle groups, J. Math. Soc. Japan 29 (1977), no. 1, 91–106, DOI 10.2969/jmsj/02910091. MR429744,
Show rawAMSref \bib{Takeuchi:triangles}{article}{ label={Tak}, author={Takeuchi, Kisao}, title={Arithmetic triangle groups}, journal={J. Math. Soc. Japan}, volume={29}, date={1977}, number={1}, pages={91--106}, issn={0025-5645}, review={\MR {429744}}, doi={10.2969/jmsj/02910091}, }
Reference [Th]
W. P. Thurston, On the geometry and dynamics of diffeomorphisms of surfaces, Bull. Amer. Math. Soc. (N.S.) 19 (1988), no. 2, 417–431, DOI 10.1090/S0273-0979-1988-15685-6. MR956596,
Show rawAMSref \bib{Thurston:surfaces}{article}{ label={Th}, author={Thurston, William P.}, title={On the geometry and dynamics of diffeomorphisms of surfaces}, journal={Bull. Amer. Math. Soc. (N.S.)}, volume={19}, date={1988}, number={2}, pages={417--431}, issn={0273-0979}, review={\MR {956596}}, doi={10.1090/S0273-0979-1988-15685-6}, }
Reference [TZ1]
D. Torres-Teigell and J. Zachhuber, Orbifold points on Prym-Teichmüller curves in genus 3, Int. Math. Res. Not. IMRN 4 (2018), 1228–1280, DOI 10.1093/imrn/rnw277. MR3801461,
Show rawAMSref \bib{Torres:Zachhuber:g3}{article}{ label={TZ1}, author={Torres-Teigell, David}, author={Zachhuber, Jonathan}, title={Orbifold points on Prym-Teichm\"{u}ller curves in genus 3}, journal={Int. Math. Res. Not. IMRN}, date={2018}, number={4}, pages={1228--1280}, issn={1073-7928}, review={\MR {3801461}}, doi={10.1093/imrn/rnw277}, }
Reference [TZ2]
D. Torres-Teigell and J. Zachhuber, Orbifold points on Prym-Teichmüller curves in genus 4, J. Inst. Math. Jussieu 18 (2019), no. 4, 673–706, DOI 10.1017/s1474748017000196. MR3963516,
Show rawAMSref \bib{Torres:Zachhuber:g4}{article}{ label={TZ2}, author={Torres-Teigell, David}, author={Zachhuber, Jonathan}, title={Orbifold points on Prym-Teichm\"{u}ller curves in genus 4}, journal={J. Inst. Math. Jussieu}, volume={18}, date={2019}, number={4}, pages={673--706}, issn={1474-7480}, review={\MR {3963516}}, doi={10.1017/s1474748017000196}, }
Reference [V1]
W. A. Veech, Teichmüller curves in moduli space, Eisenstein series and an application to triangular billiards, Invent. Math. 97 (1989), no. 3, 553–583, DOI 10.1007/BF01388890. MR1005006,
Show rawAMSref \bib{Veech:triangles}{article}{ label={V1}, author={Veech, W. A.}, title={Teichm\"{u}ller curves in moduli space, Eisenstein series and an application to triangular billiards}, journal={Invent. Math.}, volume={97}, date={1989}, number={3}, pages={553--583}, issn={0020-9910}, review={\MR {1005006}}, doi={10.1007/BF01388890}, }
Reference [V2]
W. A. Veech, Geometric realizations of hyperelliptic curves, Algorithms, fractals, and dynamics (Okayama/Kyoto, 1992), Plenum, New York, 1995, pp. 217–226. MR1402493,
Show rawAMSref \bib{Veech:hyperelliptic}{article}{ label={V2}, author={Veech, William A.}, title={Geometric realizations of hyperelliptic curves}, conference={ title={Algorithms, fractals, and dynamics}, address={Okayama/Kyoto}, date={1992}, }, book={ publisher={Plenum, New York}, }, date={1995}, pages={217--226}, review={\MR {1402493}}, }
Reference [Vi]
M. Viana, Dynamics of interval exchange transformations and Teichmüller flows, Preprint, 2008.
Reference [Vo1]
Ya. B. Vorobets, Plane structures and billiards in rational polygons: the Veech alternative (Russian), Uspekhi Mat. Nauk 51 (1996), no. 5(311), 3–42, DOI 10.1070/RM1996v051n05ABEH002993; English transl., Russian Math. Surveys 51 (1996), no. 5, 779–817. MR1436653,
Show rawAMSref \bib{Vorobets:billiards}{article}{ label={Vo1}, author={Vorobets, Ya. B.}, title={Plane structures and billiards in rational polygons: the Veech alternative}, language={Russian}, journal={Uspekhi Mat. Nauk}, volume={51}, date={1996}, number={5(311)}, pages={3--42}, issn={0042-1316}, translation={ journal={Russian Math. Surveys}, volume={51}, date={1996}, number={5}, pages={779--817}, issn={0036-0279}, }, review={\MR {1436653}}, doi={10.1070/RM1996v051n05ABEH002993}, }
Reference [Vo2]
Ya. B. Vorobets, Plane structures and billiards in rational polyhedra (Russian), Uspekhi Mat. Nauk 51 (1996), no. 1(307), 145–146, DOI 10.1070/RM1996v051n01ABEH002769; English transl., Russian Math. Surveys 51 (1996), no. 1, 177–178. MR1392678,
Show rawAMSref \bib{Vorobets:ward}{article}{ label={Vo2}, author={Vorobets, Ya. B.}, title={Plane structures and billiards in rational polyhedra}, language={Russian}, journal={Uspekhi Mat. Nauk}, volume={51}, date={1996}, number={1(307)}, pages={145--146}, issn={0042-1316}, translation={ journal={Russian Math. Surveys}, volume={51}, date={1996}, number={1}, pages={177--178}, issn={0036-0279}, }, review={\MR {1392678}}, doi={10.1070/RM1996v051n01ABEH002769}, }
Reference [Wa]
C. C. Ward, Calculation of Fuchsian groups associated to billiards in a rational triangle, Ergodic Theory Dynam. Systems 18 (1998), no. 4, 1019–1042, DOI 10.1017/S0143385798117479. MR1645350,
Show rawAMSref \bib{Ward:billiards}{article}{ label={Wa}, author={Ward, Clayton C.}, title={Calculation of Fuchsian groups associated to billiards in a rational triangle}, journal={Ergodic Theory Dynam. Systems}, volume={18}, date={1998}, number={4}, pages={1019--1042}, issn={0143-3857}, review={\MR {1645350}}, doi={10.1017/S0143385798117479}, }
Reference [Wr1]
A. Wright, Schwarz triangle mappings and Teichmüller curves: the Veech-Ward-Bouw-Möller curves, Geom. Funct. Anal. 23 (2013), no. 2, 776–809, DOI 10.1007/s00039-013-0221-z. MR3053761,
Show rawAMSref \bib{Wright:BM}{article}{ label={Wr1}, author={Wright, Alex}, title={Schwarz triangle mappings and Teichm\"{u}ller curves: the Veech-Ward-Bouw-M\"{o}ller curves}, journal={Geom. Funct. Anal.}, volume={23}, date={2013}, number={2}, pages={776--809}, issn={1016-443X}, review={\MR {3053761}}, doi={10.1007/s00039-013-0221-z}, }
Reference [Wr2]
A. Wright, Translation surfaces and their orbit closures: an introduction for a broad audience, EMS Surv. Math. Sci. 2 (2015), no. 1, 63–108, DOI 10.4171/EMSS/9. MR3354955,
Show rawAMSref \bib{Wright:survey:EMS}{article}{ label={Wr2}, author={Wright, Alex}, title={Translation surfaces and their orbit closures: an introduction for a broad audience}, journal={EMS Surv. Math. Sci.}, volume={2}, date={2015}, number={1}, pages={63--108}, issn={2308-2151}, review={\MR {3354955}}, doi={10.4171/EMSS/9}, }
Reference [Wr3]
A. Wright, From rational billiards to dynamics on moduli spaces, Bull. Amer. Math. Soc. (N.S.) 53 (2016), no. 1, 41–56, DOI 10.1090/bull/1513. MR3403080,
Show rawAMSref \bib{Wright:survey:BAMS}{article}{ label={Wr3}, author={Wright, Alex}, title={From rational billiards to dynamics on moduli spaces}, journal={Bull. Amer. Math. Soc. (N.S.)}, volume={53}, date={2016}, number={1}, pages={41--56}, issn={0273-0979}, review={\MR {3403080}}, doi={10.1090/bull/1513}, }
Reference [Wr4]
A. Wright, Totally geodesic submanifolds of Teichmüller space, J. Differential Geom. 115 (2020), no. 3, 565–575, DOI 10.4310/jdg/1594260019. MR4120819,
Show rawAMSref \bib{Wright:tot_geod}{article}{ label={Wr4}, author={Wright, Alex}, title={Totally geodesic submanifolds of Teichm\"{u}ller space}, journal={J. Differential Geom.}, volume={115}, date={2020}, number={3}, pages={565--575}, issn={0022-040X}, review={\MR {4120819}}, doi={10.4310/jdg/1594260019}, }
Reference [Y]
J.-C. Yoccoz, Interval exchange maps and translation surfaces, Homogeneous flows, moduli spaces and arithmetic, Clay Math. Proc., vol. 10, Amer. Math. Soc., Providence, RI, 2010, pp. 1–69. MR2648692,
Show rawAMSref \bib{Yoccoz:translation:survey}{article}{ label={Y}, author={Yoccoz, Jean-Christophe}, title={Interval exchange maps and translation surfaces}, conference={ title={Homogeneous flows, moduli spaces and arithmetic}, }, book={ series={Clay Math. Proc.}, volume={10}, publisher={Amer. Math. Soc., Providence, RI}, }, date={2010}, pages={1--69}, review={\MR {2648692}}, }
Reference [Za]
J. Zachhuber, The Galois action and a spin invariant for Prym-Teichmüller curves in genus 3 (English, with English and French summaries), Bull. Soc. Math. France 146 (2018), no. 3, 427–439, DOI 10.24033/bsmf.2766. MR3936530,
Show rawAMSref \bib{Zachhuber:spin}{article}{ label={Za}, author={Zachhuber, Jonathan}, title={The Galois action and a spin invariant for Prym-Teichm\"{u}ller curves in genus 3}, language={English, with English and French summaries}, journal={Bull. Soc. Math. France}, volume={146}, date={2018}, number={3}, pages={427--439}, issn={0037-9484}, review={\MR {3936530}}, doi={10.24033/bsmf.2766}, }
Reference [Z]
A. Zorich, Flat surfaces, Frontiers in number theory, physics, and geometry. I, Springer, Berlin, 2006, pp. 437–583, DOI 10.1007/978-3-540-31347-2_13. MR2261104,
Show rawAMSref \bib{Zorich:survey}{article}{ label={Z}, author={Zorich, Anton}, title={Flat surfaces}, conference={ title={Frontiers in number theory, physics, and geometry. I}, }, book={ publisher={Springer, Berlin}, }, date={2006}, pages={437--583}, review={\MR {2261104}}, doi={10.1007/978-3-540-31347-2\_13}, }

Article Information

MSC 2020
Primary: 32G15 (Moduli of Riemann surfaces, Teichmüller theory (complex-analytic aspects in several variables))
Author Information
Curtis T. McMullen
Mathematics Department, Harvard University, 1 Oxford St, Cambridge, Massachussetts 02138-2901
MathSciNet
Additional Notes

The author’s research was supported in part by the NSF.

Journal Information
Bulletin of the American Mathematical Society, Volume 60, Issue 2, ISSN 1088-9485, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on and published on .
Copyright Information
Copyright 2022 Curtis T. McMullen
Article References
  • Permalink
  • Permalink (PDF)
  • DOI 10.1090/bull/1782
  • MathSciNet Review: 4557380
  • Show rawAMSref \bib{4557380}{article}{ author={McMullen, Curtis}, title={Billiards and Teichm\"uller curves}, journal={Bull. Amer. Math. Soc.}, volume={60}, number={2}, date={2023-04}, pages={195-250}, issn={0273-0979}, review={4557380}, doi={10.1090/bull/1782}, }

Settings

Change font size
Resize article panel
Enable equation enrichment

Note. To explore an equation, focus it (e.g., by clicking on it) and use the arrow keys to navigate its structure. Screenreader users should be advised that enabling speech synthesis will lead to duplicate aural rendering.

For more information please visit the AMS MathViewer documentation.