Fundamental polygon

Last updated

In mathematics, a fundamental polygon can be defined for every compact Riemann surface of genus greater than 0. It encodes not only the topology of the surface through its fundamental group but also determines the Riemann surface up to conformal equivalence. By the uniformization theorem, every compact Riemann surface has simply connected universal covering surface given by exactly one of the following:

Contents

In the first case of genus zero, the surface is conformally equivalent to the Riemann sphere.

In the second case of genus one, the surface is conformally equivalent to a torus C/Λ for some lattice Λ in C. The fundamental polygon of Λ, if assumed convex, may be taken to be either a period parallelogram or a centrally symmetric hexagon, a result first proved by Fedorov in 1891.

In the last case of genus g > 1, the Riemann surface is conformally equivalent to H/Γ where Γ is a Fuchsian group of Möbius transformations. A fundamental domain for Γ is given by a convex polygon for the hyperbolic metric on H. These can be defined by Dirichlet polygons and have an even number of sides. The structure of the fundamental group Γ can be read off from such a polygon. Using the theory of quasiconformal mappings and the Beltrami equation, it can be shown there is a canonical convex fundamental polygon with 4g sides, first defined by Fricke, which corresponds to the standard presentation of Γ as the group with 2g generators a1, b1, a2, b2, ..., ag, bg and the single relation [a1,b1][a2,b2] ⋅⋅⋅ [ag,bg] = 1, where [a,b] = aba−1b−1.

Any Riemannian metric on an oriented closed 2-manifold M defines a complex structure on M, making M a compact Riemann surface. Through the use of fundamental polygons, it follows that two oriented closed 2-manifolds are classified by their genus, that is half the rank of the Abelian group Γ/[Γ,Γ], where Γ = π1(M). Moreover, it also follows from the theory of quasiconformal mappings that two compact Riemann surfaces are diffeomorphic if and only if they are homeomorphic. Consequently, two closed oriented 2-manifolds are homeomorphic if and only if they are diffeomorphic. Such a result can also be proved using the methods of differential topology. [1] [2]

Fundamental polygons in genus one

Fricke-Klein-1897-hexagon-parallelogram-1.jpg
Fricke-Klein-1897-hexagon-parallelogram-2.jpg

Parallelograms and centrally symmetric hexagons

In the case of genus one, a fundamental convex polygon is sought for the action by translation of Λ = ZaZb on R2 = C where a and b are linearly independent over R. (After performing a real linear transformation on R2, it can be assumed if necessary that Λ = Z2 = Z + Zi; for a genus one Riemann surface it can be taken to have the form Λ = Z2 = Z + Z ω, with Im ω > 0.) A fundamental domain is given by the parallelogram sx + ty for 0 < s , t < 1 where x and y are generators of Λ.

If C is the interior of a fundamental convex polygon, then the translates C + x cover R2 as x runs over Λ. It follows that the boundary points of C are formed of intersections C ∩ (C + x). These are compact convex sets in ∂C and thus either vertices of C or sides of C. It follows that every closed side of C can be written this way. Translating by −x it follows that C ∩ (Cx) is also a side of C. Thus sides of C occur in parallel pairs of equal length. The end points of two such parallel segments of equal length can be joined so that they intersect and the intersection occurs at the midpoints of the line segments joining the endpoints. It follows that the intersections of al such segments occur at the same point. Translating that point to the origin, it follows that the polygon is centrally symmetric; that is, if a point z is in the polygon, so too is −z.

It is easy to see translates of a centrally symmetric convex hexagon tessellate the plane. If A is a point of the hexagon, then the lattice is generated by the displacement vectors AB and AC where B and C are the two vertices which are not neighbours of A and not opposite A. Indeed, the second picture shows how the hexagon is equivalent to the parallelogram obtained by displacing the two triangles chopped off by the segments AB and AC. Equally well the first picture shows another way of matching a tiling by parallelograms with the hexagonal tiling. If the centre of the hexagon is 0 and the vertices in order are a, b, c, −a, −b and −c, then Λ is the Abelian group with generators a + b and b + c.

Examples of Fundamental Polygons Generated by Parallelograms

There are exactly four topologies that can be created by identifying the sides of a rhombus in different ways. They are given below as directional edges A and B on a square, either as AABB or ABAB sequences.

Name Sphere Torus Projective plane Klein bottle
Orientable YesNo
Total curvature 4π02π0
Topology
ABAB
(square)
TorusAsSquare.svg
ProjectivePlaneAsSquare.svg
KleinBottleAsSquare.svg

(or )    
Topology
AABB
(square)
SphereAsSquare.svg
[3]
ProjectivePlaneAsSquare2.png

(or )   
KleinBottleAsSquare2.png
Geometry Sphere wireframe 10deg 6r.svg
Sphere
Toroidal monohedron.png
Torus
Sphere symmetry group ci.png
Hemisphere
Klein bottle translucent.png
Half torus

Fedorov's theorem

Fedorov's theorem, established by the Russian crystallographer Evgraf Fedorov in 1891, asserts that parallelograms and centrally symmetric hexagons are the only convex polygons that are fundamental domains. [4] There are several proofs of this, some of the more recent ones related to results in convexity theory, the geometry of numbers and circle packing, such as the Brunn–Minkowski inequality. [5] Two elementary proofs due to H. S. M. Coxeter and Voronoi will be presented here. [6] [7]

Coxeter's proof proceeds by assuming that there is a centrally symmetric convex polygon C with 2m sides. Then a large closed parallelogram formed from N2 fundamental parallelograms is tiled by translations of C which go beyond the edges of the large parallelogram. This induces a tiling on the torus C/NΛ. Let v, e and f be the number of vertices, edges and faces in this tiling (taking into account identifications in the quotient space). Then, because the Euler–Poincaré characteristic of a torus is zero,

On the other hand, since each vertex is on at least 3 different edges and every edge is between two vertices,

Moreover, since every edge is on exactly two faces,

Hence

so that

as required.

Voronoi's proof starts with the observation that every edge of C corresponds to an element x of Λ. In fact the edge is the orthogonal bisector of the radius from 0 to x. Hence the foot of the perpendicular from 0 to each edge lies in the interior of each edge. If y is any lattice point, then 1/2 y cannot lie in C; for if so, –1/2 y would also lie in C, contradicting C being a fundamental domain for Λ. Let ±x1, ..., ±xm be the 2m distinct points of Λ corresponding to sides of C. Fix generators a and b of Λ. Thus xi = αia + βib, where αi and βi are integers. It is not possible for both αi and βi to be even, since otherwise ± 1/2 xi would be a point of Λ on a side, which contradicts C being a fundamental domain. So there are three possibilities for the pair of integers (αi, βi) modulo 2: (0,1), (1,0) and (1,1). Consequently, if m > 3, there would be xi and xj with ij with both coordinates of xixj even, i.e. 1/2 (xi + xj) lies in Λ. But this is the midpoint of the line segment joining two interior points of edges and hence lies in C, the interior of the polygon. This again contradicts the fact that C is a fundamental domain. So reductio ad absurdumm ≤ 3, as claimed.

Dirichlet–Voronoi domains

For a lattice Λ in C = R2, a fundamental domain can be defined canonically using the conformal structure of C. Note that the group of conformal transformations of C is given by complex affine transformations g(z) = az + b with a ≠ 0. These transformations preserve Euclidean metric d(z, w) = |zw| up to a factor, as well as preserving the orientation. It is the subgroup of the Möbius group fixing the point at ∞. The metric structure can be used to define a canonical fundamental domain by C = {z: d(z, 0) < d(z, λ) for all λ ≠ 0 in Λ}. (It is obvious from the definition that it is a fundamental domain.) This is an example of a Dirichlet domain or Voronoi diagram: since complex translations form an Abelian group, so commute with the action of Λ, these concepts coincide. The canonical fundamental domain for Λ = Z + Zω with Im ω > 0 is either a symmetric convex parallelogram or hexagon with centre 0. By conformal equivalence, the period ω can be further restricted to satisfy |Re ω| ≤ 1/2 and |ω| ≥ 1. As Dirichlet showed ("Dirichlet's hexagon theorem", 1850), for almost all ω the fundamental domain is a hexagon. For Re ω > 0, the midpoints of sides are given by ±1/2, ±ω/2 and ±(ω – 1)/2; the sides bisect the corresponding radii from 0 orthogonally, which determines the vertices completely. In fact the first vertex must have the form (1 + ix)/2 and ω(1 + iy)/2 with x and y real; so if ω = a + ib, then aby = 1 and x = b + ay. Hence y = (a – 1)/b and x = (a2 + b2a)/b. The six vertices are therefore ±ω(1 – iy)/2 and ±(1 ± ix)/2. [8]

Fundamental polygons in higher genus

Overview

Every compact Riemann surface X has a universal covering surface which is a simply connected Riemann surface X. The fundamental group of X acts as deck transformations of X and can be identified with a subgroup Γ of the group of biholomorphisms of X. The group Γ thus acts freely on X with compact quotient space X/Γ, which can be identified with X. Thus the classification of compact Riemann surfaces can be reduced to the study of possible groups Γ. By the uniformization theorem X is either the Riemann sphere, the complex plane or the unit disk/upper halfplane. The first important invariant of a compact Riemann surface is its genus, a topological invariant given by half the rank of the Abelian group Γ/[Γ, Γ] (which can be identified with the homology group H1(X, Z)). The genus is zero if the covering space is the Riemann sphere; one if it is the complex plane; and greater than one if it is the unit disk or upper halfplane. [9]

Bihomolomorphisms of the Riemann sphere are just complex Möbius transformations and every non-identity transformation has at least one fixed point, since the corresponding complex matrix always has at least one non-zero eigenvector. Thus if X is the Riemann sphere, then X must be simply connected and biholomorphic to the Riemann sphere, the genus zero Riemann surface. When X is the complex plane, the group of biholomorphisms is the affine group, the complex Möbius transformations fixing ∞, so the transformations g(z) = az + b with a ≠ 0. The non-identity transformations without fixed points are just those with a = 1 and b ≠ 0, i.e. the non-zero translations. The group Γ can thus be identified with a lattice Λ in C and X with a quotient C/Λ, as described in the section on fundamental polygons in genus one. In the third case when X is the unit disk or upper half plane, the group of biholomorphisms consists of the complex Möbius transformations fixing the unit circle or the real axis. In the former case, the transformations correspond to elements of the group SU(1, 1)/{±I}; in the latter case they correspond to real Möbius transformations, so elements of SL(2, R)/{±I}. [9]

The study and classification of possible groups Γ that act freely on the unit disk or upper halfplane with compact quotient—the Fuchsian groups of the first kind—can be accomplished by studying their fundamental polygons, as described below. As Poincaré observed, each such polygon has special properties, namely it is convex and has a natural pairing between its sides. These not only allow the group to be recovered but provide an explicit presentation of the group by generators and relations. Conversely Poincaré proved that any such polygon gives rise to a compact Riemann surface; in fact, Poincaré's polygon theorem applied to more general polygons, where the polygon was allowed to have ideal vertices, but his proof was complete only in the compact case, without such vertices. Without assumptions on the convexity of the polygon, complete proofs have been given by Maskit and de Rham, based on an idea of Siegel, and can be found in Beardon (1983), Iversen (1992) and Stillwell (1992). Carathéodory gave an elementary treatment of the existence of tessellations by Schwarz triangles, i.e. tilings by geodesic triangles with angles π/a, π/b, π/c with sum less than π where a, b, c are integers. When all the angles equal π/2g, this establishes the tiling by regular 4g-sided hyperbolic polygons and hence the existence of a particular compact Riemann surface of genus g as a quotient space. This special example, which has a cyclic group Z2g of bihomolomorphic symmetries, is used in the development below. [9]

The classification up to homeomorphism and diffeomorphism of compact Riemann surfaces implies the classification of closed orientable 2-manifolds up to homeomorphism and diffeomorphism: any two 2-manifolds with the same genus are diffeomorphic. In fact using a partition of unity, every closed orientable 2-manifold admits a Riemannian metric. For a compact Riemann surface a conformal metric can also be introduced which is conformal, so that in holomorphic coordinates the metric takes the form ρ(z) |dz|2. Once this metric has been chosen, locally biholomorphic mappings are precisely orientation-preserving diffeomorphisms that are conformal, i.e. scale the metric by a smooth function. The existence of isothermal coordinates—which can be proved using either local existence theorems for the Laplacian or the Beltrami equation—shows that every closed oriented Riemannian 2-manifold can be given a complex structure compatible with its metric, and hence has the structure of a compact Riemann surface. This construction shows that the classification of closed orientable 2-manifolds up to diffeomorphism or homeomorphism can be reduced to the case of compact Riemann surfaces. [10]

The classification up to homeomorphism and diffeomorphism of compact Riemann surfaces can be accomplished using the fundamental polygon. Indeed, as Poincaré observed, convex fundamental polygons for compact Riemann surfaces H/Γ can be constructed by adapting the method of Dirichlet from the Euclidean space to hyperbolic space. Then following Nevanlinna and Jost, the fundamental domain can be modified in steps to yield a non-convex polygon with vertices lying in a single orbit of Γ and piecewise geodesic sides. The pairing relation on the sides is also modified in each of these steps. Each step involves cutting the polygon by a diagonal geodesic segment in the interior of the polygon and reassembling the polygon using one of the Möbius transformations involved in the pairing. No two paired sides can have a common vertex in the final pairing relation, which satisfies similar properties to the original relation. This polygon can in turn be successively modified by reassembling the polygon after cutting it by a diagonal piecewise geodesic segment in its interior. The final polygon has 4g equivalent vertices, with sides that are piecewise geodesic. The sides are labelled by the group elements which give the Möbius transformation to the paired side. In order the labelling is

so that Γ is generated by the ai and bi subject to the single relation

Using the theory of intersection numbers, it follows that the shape obtained by joining vertices by geodesics is also a proper polygon, not necessarily convex, and is also a fundamental domain with the same group elements giving the pairing. This yields a fundamental polygon with edges given by geodesic segments and with the standard labelling. The abelianisation of Γ, the quotient group Γ/[Γ, Γ], is a free Abelian group with 2g generators. Thus the genus g is a topological invariant. It is easy to see that two Riemann surfaces with the same genus are homeomorphic since as topological space since they are obtained by identifying sides of a 4g-sided polygon—a Euclidean polygon in the Klein model—by diffeomorphisms between paired sides. [11] Applying this construction to the regular 4g-sided polygon allows the Riemann surface to be viewed topologically as a doughnut with g holes, the standard description of oriented surfaces in introductory texts on topology. [12] [13]

There are several further results:

These results are tied up with the interrelation between homeomorphisms and the fundamental group: this reflects the fact that the mapping class group of a Riemann surface—the group of quasiconformal self-homomorphisms of a Riemann surface H/Γ modulo those homotopic to the identity—can be identified with the outer automorphism group of Γ (the Dehn–Nielsen–Baer theorem). [14] To see this connection, note that if f is a quasiconformal homeomorphism of X1 = H1 onto X2 = H2, then f lifts to a quasiconformal homeomorphism f of H onto itself. This lift is unique up to pre-composition with elements of Γ1 and post-composition with elements of Γ2. If πi is the projection of H onto Xi, then fπ1 = π2f and Γi is just the group of homeomorphisms g of H such that πig = πi. If follows that fg = θ(g) f for g in Γ1 where θ is a group isomorphism of Γ1 onto Γ2. A different choice of f changes θ by composition with an inner automorphism: such isomorphisms are said to be equivalent. [15]

Two isomorphisms θ and θ′ are equivalent if and only if the corresponding homeomorphisms f and f' are homotopic. In fact it suffices to show that a quasiconformal self-homeomorphism f of a surface induces an inner automorphism of the fundamental group if and only if it is homotopic to the identity map: in other words the homomorphism of the quasiconformal self-homeomorphism group of H/Γ into Out Γ passes to the mapping class group on which it is injective. Indeed, suppose first that F(t) is a continuous path of self-homeomorphisms with F(0) = id and F(1) = f. Then there is a continuous lift F(t) with F(0) = id. Moreover, for each g in Γ, F(t) ∘ gF(t)−1 is a continuously varying element of Γ equal to g for t = 0; so discreteness of Γ forces this element to be constant and hence equal to g so that F(t) commutes with Γ, so F(1) induces the trivial automorphism. If on the other hand F is a quasiconformal lift of f inducing an inner automorphism of Γ, after composition with an element Γ if necessary it can be assumed that F commutes with Γ. Since F is quasiconformal, it extends to a quasisymmetric homeomorphism of the circle which also commutes with Γ. Each g ≠ id in Γ is hyperbolic so has two fixed points on the circle a± such that for all other points z, g±n(z) tends to a± as n tends to infinity. Hence F must fix these points; since these points are dense in the circle as g varies, it follows that F fixes the unit circle. Let μ = Fz / Fz, so that μ is a Γ-invariant Beltrami differential. Let F(t) be the solution of the Beltrami equation normalised to fix three points on the unit circle. Then F(t) commutes with Γ and so, as for F = F(1), is the identity on the unit circle. By construction F(t) is an isotopy between the identity and F. This proves injectivity. [15]

The proof of surjectivity relies on comparing the hyperbolic metric on D with a word-length metric on Γ. [16] Assuming without loss of generality that 0 lies in the interior of a convex fundamental polygon C and g is an element of Γ, the ray from 0 to g(0)—the hyperbolic geodesic—passes through a succession of translates of C. Each of these is obtained from the previous one by applying a generator of Γ or a fixed product of generators (if successive translates meet in a vertex). It follows that the hyperbolic distance between 0 and g(0) is less than 4g times the word length of g plus twice diameter of the fundamental polygon. Thus the metric on Γ d1(g, h) = L(h−1g) defined by the word length L(g) satisfies

for positive constants a and b. Conversely there are positive constants c and d such that

Dirichlet polygons

Given a point in the upper half-plane H, and a discrete subgroup Γ of PSL(2, R) that acts freely discontinuously on the upper half-plane, then one can define the Dirichlet polygon as the set of points

Here, d is a hyperbolic metric on the upper half-plane. The metric fundamental polygon is more usually called the Dirichlet polygon.

Normalised polygon

In this section, starting from an arbitrary Dirichlet polygon, a description will be given of the method of Nevanlinna (1953), elaborated in Jost (2002), for modifying the polygon to a non-convex polygon with 4g equivalent vertices and a canonical pairing on the sides. This treatment is an analytic counterpart of the classical topological classification of orientable 2-dimensional polyhedra presented in Seifert & Threlfall (1934).

Fricke canonical polygon

Given a Riemann surface of genus g greater than one, Fricke described another fundamental polygon, the Fricke canonical polygon, which is a very special example of a Dirichlet polygon. The polygon is related to the standard presentation of the fundamental group of the surface. Fricke's original construction is complicated and described in Fricke & Klein (1897). Using the theory of quasiconformal mappings of Ahlfors and Bers, Keen (1965) gave a new, shorter and more precise version of Fricke's construction. The Fricke canonical polygon has the following properties:

The above construction is sufficient to guarantee that each side of the polygon is a closed (non-trivial) loop in the Riemann surface H/Γ. As such, each side can thus an element of the fundamental group . In particular, the fundamental group has 2g generators , with exactly one defining constraint,

.

The genus of the Riemann surface H/Γ is g.

Area

The area of the standard fundamental polygon is where g is the genus of the Riemann surface (equivalently, where 4g is the number of the sides of the polygon). Since the standard polygon is a representative of H/Γ, the total area of the Riemann surface is equal to the area of the standard polygon. The area formula follows from the Gauss–Bonnet theorem and is in a certain sense generalized through the Riemann–Hurwitz formula.

Explicit form for standard polygons

Explicit expressions can be given for the regular standard 4g-sided polygon, with rotational symmetry. In this case, that of a genus Riemann surface with g-fold rotational symmetry, the group may be given by generators . These generators are given by the following fractional linear transforms acting on the upper half-plane:

for . The parameters are given by

and

and

It may be verified that these generators obey the constraint

which gives the totality of the group presentation.

See also

Notes

  1. See:
  2. See:
  3. Example of sphere construction from fundamental polygon.
  4. E. Fedorov (1891) "Симметрія на плоскости" (Simmetriya na ploskosti, Symmetry in the plane), Записки Императорского С.-Петербургского минералогического общества (Zapiski Imperatorskova Sankt-Petersburgskova Mineralogicheskova Obshchestva, Proceedings of the Imperial St. Petersburg Mineralogical Society), 2nd series, 28 : 345–390 (in Russian).
  5. See:
  6. Voronoi's proof has the advantage of generalising to n dimensions: it shows that if translates of a centrally symmetric convex polyhedron tessallate Rn, then the polyhedron has at most 2(2n − 1) faces.
  7. See:
  8. See:
  9. 1 2 3 Beardon 1984
  10. Imayoshi & Taniguchi 1992
  11. Note that a simple polygon in the plane with n ≥ 4 vertices is homeomorphic to one, and hence any, convex n-gon by a piecewise linear homeomorphism, linear on the edges: this follows by induction on n from the observation of Max Dehn that any simple polygon possesses a diagonal, i.e. an interior chord between vertices, so can be broken up into smaller polygons; see Guggenheimer (1977). For a regular 4g-gon, the pairing between sides can be made linear, by reparametrising triangles made up of the centre and one side from each pair of sides.
  12. Jost 2002 , pp. 47–57
  13. Shastri 2011
  14. Farb & Margalit 2012
  15. 1 2 Ahlfors 2006 , pp. 67–68
  16. Farb & Margalit 2012 , pp. 230–236

Related Research Articles

In the mathematical field of algebraic topology, the fundamental group of a topological space is the group of the equivalence classes under homotopy of the loops contained in the space. It records information about the basic shape, or holes, of the topological space. The fundamental group is the first and simplest homotopy group. The fundamental group is a homotopy invariant—topological spaces that are homotopy equivalent have isomorphic fundamental groups. The fundamental group of a topological space is denoted by .

<span class="mw-page-title-main">Riemann mapping theorem</span>

In complex analysis, the Riemann mapping theorem states that if is a non-empty simply connected open subset of the complex number plane which is not all of , then there exists a biholomorphic mapping from onto the open unit disk

<span class="mw-page-title-main">Covering space</span> Type of continuous map in topology

In topology, a covering or covering projection is a map between topological spaces that, intuitively, locally acts like a projection of multiple copies of a space onto itself. In particular, coverings are special types of local homeomorphisms. If is a covering, is said to be a covering space or cover of , and is said to be the base of the covering, or simply the base. By abuse of terminology, and may sometimes be called covering spaces as well. Since coverings are local homeomorphisms, a covering space is a special kind of étale space.

In mathematics, a modular form is a (complex) analytic function on the upper half-plane, , that satisfies:

<span class="mw-page-title-main">Klein quartic</span> Compact Riemann surface of genus 3

In hyperbolic geometry, the Klein quartic, named after Felix Klein, is a compact Riemann surface of genus 3 with the highest possible order automorphism group for this genus, namely order 168 orientation-preserving automorphisms, and 168 × 2 = 336 automorphisms if orientation may be reversed. As such, the Klein quartic is the Hurwitz surface of lowest possible genus; see Hurwitz's automorphisms theorem. Its (orientation-preserving) automorphism group is isomorphic to PSL(2, 7), the second-smallest non-abelian simple group after the alternating group A5. The quartic was first described in (Klein 1878b).

In mathematics, the Schwarzian derivative is an operator similar to the derivative which is invariant under Möbius transformations. Thus, it occurs in the theory of the complex projective line, and in particular, in the theory of modular forms and hypergeometric functions. It plays an important role in the theory of univalent functions, conformal mapping and Teichmüller spaces. It is named after the German mathematician Hermann Schwarz.

In number theory and algebraic geometry, a modular curveY(Γ) is a Riemann surface, or the corresponding algebraic curve, constructed as a quotient of the complex upper half-plane H by the action of a congruence subgroup Γ of the modular group of integral 2×2 matrices SL(2, Z). The term modular curve can also be used to refer to the compactified modular curvesX(Γ) which are compactifications obtained by adding finitely many points to this quotient. The points of a modular curve parametrize isomorphism classes of elliptic curves, together with some additional structure depending on the group Γ. This interpretation allows one to give a purely algebraic definition of modular curves, without reference to complex numbers, and, moreover, prove that modular curves are defined either over the field of rational numbers Q or a cyclotomic field Qn). The latter fact and its generalizations are of fundamental importance in number theory.

In mathematics, a Fuchsian model is a representation of a hyperbolic Riemann surface R as a quotient of the upper half-plane H by a Fuchsian group. Every hyperbolic Riemann surface admits such a representation. The concept is named after Lazarus Fuchs.

In mathematics, the Teichmüller space of a (real) topological surface is a space that parametrizes complex structures on up to the action of homeomorphisms that are isotopic to the identity homeomorphism. Teichmüller spaces are named after Oswald Teichmüller.

<span class="mw-page-title-main">Schwarz triangle</span> Spherical triangle that can be used to tile a sphere

In geometry, a Schwarz triangle, named after Hermann Schwarz, is a spherical triangle that can be used to tile a sphere, possibly overlapping, through reflections in its edges. They were classified in Schwarz (1873).

In mathematical complex analysis, a quasiconformal mapping, introduced by Grötzsch (1928) and named by Ahlfors (1935), is a homeomorphism between plane domains which to first order takes small circles to small ellipses of bounded eccentricity.

In geometric topology, Busemann functions are used to study the large-scale geometry of geodesics in Hadamard spaces and in particular Hadamard manifolds. They are named after Herbert Busemann, who introduced them; he gave an extensive treatment of the topic in his 1955 book "The geometry of geodesics".

Geometric function theory is the study of geometric properties of analytic functions. A fundamental result in the theory is the Riemann mapping theorem.

In mathematics, the Schoenflies problem or Schoenflies theorem, of geometric topology is a sharpening of the Jordan curve theorem by Arthur Schoenflies. For Jordan curves in the plane it is often referred to as the Jordan–Schoenflies theorem.

In mathematics, and more precisely in topology, the mapping class group of a surface, sometimes called the modular group or Teichmüller modular group, is the group of homeomorphisms of the surface viewed up to continuous deformation. It is of fundamental importance for the study of 3-manifolds via their embedded surfaces and is also studied in algebraic geometry in relation to moduli problems for curves.

In mathematics, the Beltrami equation, named after Eugenio Beltrami, is the partial differential equation

In mathematics a translation surface is a surface obtained from identifying the sides of a polygon in the Euclidean plane by translations. An equivalent definition is a Riemann surface together with a holomorphic 1-form.

In mathematics, a quasicircle is a Jordan curve in the complex plane that is the image of a circle under a quasiconformal mapping of the plane onto itself. Originally introduced independently by Pfluger (1961) and Tienari (1962), in the older literature they were referred to as quasiconformal curves, a terminology which also applied to arcs. In complex analysis and geometric function theory, quasicircles play a fundamental role in the description of the universal Teichmüller space, through quasisymmetric homeomorphisms of the circle. Quasicircles also play an important role in complex dynamical systems.

In mathematics, the Douady–Earle extension, named after Adrien Douady and Clifford Earle, is a way of extending homeomorphisms of the unit circle in the complex plane to homeomorphisms of the closed unit disk, such that the extension is a diffeomorphism of the open disk. The extension is analytic on the open disk. The extension has an important equivariance property: if the homeomorphism is composed on either side with a Möbius transformation preserving the unit circle the extension is also obtained by composition with the same Möbius transformation. If the homeomorphism is quasisymmetric, the diffeomorphism is quasiconformal. An extension for quasisymmetric homeomorphisms had previously been given by Lars Ahlfors and Arne Beurling; a different equivariant construction had been given in 1985 by Pekka Tukia. Equivariant extensions have important applications in Teichmüller theory; for example, they lead to a quick proof of the contractibility of the Teichmüller space of a Fuchsian group.

In mathematics, differential forms on a Riemann surface are an important special case of the general theory of differential forms on smooth manifolds, distinguished by the fact that the conformal structure on the Riemann surface intrinsically defines a Hodge star operator on 1-forms without specifying a Riemannian metric. This allows the use of Hilbert space techniques for studying function theory on the Riemann surface and in particular for the construction of harmonic and holomorphic differentials with prescribed singularities. These methods were first used by Hilbert (1909) in his variational approach to the Dirichlet principle, making rigorous the arguments proposed by Riemann. Later Weyl (1940) found a direct approach using his method of orthogonal projection, a precursor of the modern theory of elliptic differential operators and Sobolev spaces. These techniques were originally applied to prove the uniformization theorem and its generalization to planar Riemann surfaces. Later they supplied the analytic foundations for the harmonic integrals of Hodge (1941). This article covers general results on differential forms on a Riemann surface that do not rely on any choice of Riemannian structure.

References