This may suffice to make the relation between the inversion problem and the theory of multiplicative functions and differentials clear. Abel himself proved only that part of the theorem above named for him which states that if images are the zeros and images are the poles of a meromorphic function, then the congruences (18.2) hold. But he proved this part in a sharper form, with congruence modulo G0 replaced by congruence modulo G.26 Thus with the inversion already proved, one obtains the following.

Abel’s Theorem (second version). The points images;images are the sets of zeros and poles, respectively, of a meromorphic function if and only if, for a suitable choice (independent of w) of the paths from images to the points images and images, every Abelian integral w of the first kind satisfies the equation

images

Let z be any nonconstant meromorphic function on images. It assumes, as we know, every value the same number of times, say n times. Then the surface images may be regarded as an n-sheeted covering surface images over the z-sphere, if one agrees that a point images of images lies over the point z = a of the z-sphere if the function z has the value a at images. This is the conception of a Riemann surface which Riemann himself employed in his works on algebraic functions and their integrals, and one would say today that Abel’s argument is most easily understood from this representation. Over those values a, for which the n points images at which z assumes the value a are not all distinct, there are certainly fewer than n points of the covering surface. If the function z assumes the finite value a r times at the point images, then images is a local parameter at images; the point images over a is a branch point of order r − 1. If images is any neighborhood of images, then there exists a disc images on the z-sphere such that over every point of this disc (except the center) there are exactly r points of images which lie in the neighborhood images. dz has a zero of order r − 1 at images. (Therefore there can be only finitely many such values z = a over which there are fewer than n points of images.) If z assumes the value ∞ s times at the point images, then images is a local parameter at images and we have a branch point of order s − 1; dz has a pole of order s + 1 at images. We denote the sum of the orders of all the branch points of the surface, its “branch order” by V. Then: the number of zerosthe number of poles of images where the first sum on the right is over all points of the surface except those over z = ∞, while the second sum is over precisely the points over z = ∞. The right-hand side of this equation is

images

and the left-hand side = 2p − 2. Hence

images

The branch order is always even.

Now let dw be an arbitrary differential of the first kind on images and let images in general be the points of images over the point z on the z-sphere. Then

images

is a differential dW, which is regular everywhere on the z-sphere. But such a differential does not exist, except for dW = 0. If we draw any curve on the z-sphere from the south pole z = 0 to the north pole z = ∞, then the points of images over this curve (which sometimes coalesce at branch points) form curves γ1, …, γn and it follows from integrating dW that

images

Each curve yv leads from a zero images to a pole images of z. If one joins images to images by a curve images and denotes the curve images then we obtain, as claimed, the equation (18.6).

Conversely, the congruences (18.2) follow from these equations, and hence the fact that the multiplicative function

images

has the multipliers image = 1 and is uniform on images.

The coincidence of the lattices G and G0 now comes out as follows. First, by the lemma there exist points images such that

images

provided that the ni are given integers. In this process, definite paths images and images have been chosen. From this it follows that the multiplicative function

images

is uniform on images. Hence, by the proof just carried out, paths images and images may be determined so that for these paths the left-hand sides of (18.7) all vanish. Now images and images are closed paths for which

images

The 2N closed loops images and αv give together a closed path images whose intersection numbers images with the curves αi of the integral basis, are the given integers ni.

Analytically one attacks the Jacobi inversion problem as follows. One uses an arbitrary meromorphic function f(images) on images and attempts to determine, not the points images themselves, but the values of the function f at the points images from (18.3) in their dependence on F1, F2, …, Fp. Since the values images are determined only to within a permutation, it makes better sense to replace them by their elementary symmetric functions; that is, by the coefficients of the equation

images

of degree p whose roots are the numbers

images

These coefficients Ai, expressed in terms of images are called, following Jacobi’s proposal, Abelian functions. Except for the singular systems images which form a subspace of dimension only (2p − 4) in the whole 2p-dimensional F-space, the Abelian functions are uniform and regular analytic. Furthermore, they are 2p-fold periodic. For if γI (l = 1, 2, …, 2p) is an integral basis for the closed paths on images and if

images

then, for every value of the index l, the numbers images are a period system for the Abelian function A((F)):

images

These 2p period systems are linearly independent in the following sense: the determinant whose lth row is

images

is nonzero. If one carries through explicitly the proof which has been given here of the solvability of the inversion problem, then one obtains the following result: in every bounded portion

images

of the F-space, an Abelian function A((F)) may be represented as the quotient of two functions which are regular analytic throughout this bounded portion.27 The indeterminateness for the singular systems (F) arises because for these values both the numerator and the denominator in the representation vanish. The set of the points of indeterminateness possesses no translations onto itself except for those which obviously come from the periodicity of the Abelian functions.

Furthermore, Riemann and Weierstrass showed, by a much more penetrating analysis, that the Abelian functions, without restriction to a bounded set, may be represented as quotients of transcendental entire functions, the θ-functions. They gave an explicit analytic expression in the form of a rapidly convergent infinite series for the θ-functions. These 0-functions also suffice as a basis for the general theory of the 2p-fold periodic functions of p independent complex arguments; the theory of Abelian functions is only a special case.28

The contents of this and the preceding paragraph show convincingly the ruling role played by the genus p, which is in essence a topological quantity, in the theory of functions and integrals on a closed Riemann surface. We were concerned with two trains of thought which we shall state once again with the labels

additive functions, partial fraction decomposition, Riemann-Roch theorem multiplicative functions, product representation, Abel’s theorem.

These two trains of thought permeate the theory of uniform functions on images from here on this theory is easily completed with the aid of the reciprocity laws of § 16. But the significance of this structure shows up in the proper light only when we become familiar with the system of meromorphic functions on a closed Riemann surface from a third point of view, as an algebraic function field; this will be done in the next section. And first from this aspect do those functions appear most intimately related to our other interests, the algebraic and the geometric - in so far as these relate to the theory of algebraic curves in the plane and in spaces of higher dimension.

§ 19.    The algebraic function field

The meromorphic functions on a given closed Riemann surface images form a field; this means precisely that the sum, difference, product, and quotient of two meromorphic functions (only division by 0 is ruled out) is again such a function. As was already done in the last section, one chooses a definite nonconstant meromorphic function z on images for the independent variable. Thereby images becomes an n-sheeted covering surface images over the z-sphere, and the uniform meromorphic functions f on images become n-valued algebraic functions of z. In this fashion there exists an algebraic function field belonging to images, which includes the field k(z) = k of rational functions of z. Namely, the meromorphic function f satisfies identically a definite equation

images

of degree n, in which the ri(z) are rational functions of z. More precisely, this means the following. Let images be any point of images, let t be a local parameter at images, and let z = z(t) and f = f(t) be the expansions of z and f in integral powers of t in the neighborhood of images; then the left-hand side of (19.1) becomes identically zero when one substitutes the power series z(t) and f(t) for z and f respectively. To derive the equation (19.1), we first exclude the point z = ∞ from the z-sphere as well as those points over which there are branch points, and those points over which f has poles. For every other value of z we form the number

images

where the sum on the right is over the n points over z. The sum is independent of the ordering of these points. In the neighborhood of a nonexcluded point

images

images become power series in zz0, and r1(z) is a regular analytic function at every nonexcluded point. This function cannot have an essential singularity at any excluded point; hence it has only poles. If one uses a local parameter at a point of imagesz over an excluded value z0, this may also be verified by elementary calculation. Hence r1(z) is a rational function of z. Similarly, one sees that the other elementary symmetric functions of images are rational functions of z. The equation (19.1), which is determined uniquely by f is called the field equation of f.

Furthermore, I claim that from the functions g belonging to the function field K associated with imagesz, an f may he chosen such that every g may be expressed rationally in terms of f and z. This f is called a function determining the function field. If imagesz, has the n distinct points images over z0, then it suffices to choose f such that it assumes n distinct values at these points. If k (k = 1, …, n) is a differential that has a pole only at images and has the principal part − (zz0)−2dz, then one can set, for example,

images

where one chooses any n different constants for C1, …, Cn. The field equation of f is then irreducible; that is, its left-hand side, the polynomial

images

of degree n in the variable u, cannot be factored into two polynomials images whose coefficients are also rational functions of z. For assume that this were possible; in a neighborhood of images, f may be developed in a power series in the local parameter zz0 at images. Suppose that this power series, replacing u, satisfies the equation images. Join images to any one of the points images by a curve on images whose trace on the z-sphere does not pass through any of the excluded points. Then all the function elements (z,f) along this curve must satisfy the equation images. Hence f(images) is also a root of the equation images thus it has n distinct roots and must be of degree n. Hence images is only of degree 0 in u, and a factorization of the contemplated type does not exist.

To each point images of images there belongs a function element (z, f) which satisfies the equation Fz(u) = 0. For two distinct points these function elements are always different, and the elements belonging to all points images exhaust the totality of those which satisfy that equation. In other words: the totality of those function elements which satisfy the irreducible algebraic equation Fz (u) = 0 constitutes a single analytic form in the sense of Weierstrass. This analytic form, regarded as a Riemann surface, is conformally equivalent to the given Riemann surface. The given surface is- the Riemann surface which belongs to the algebraic form defined by the equation Fz(u) = 0.

In order to express the given element g of the function field K rationally in f, we apply the Lagrange interpolation formula

images

Again images denote the n points of images over the point z on the z-sphere. By the same argument as above, one finds that the coefficients of the polynomial images of degree n − 1, are rational functions of z. For all values of z for which images are different, it follows that

images

Since the polynomial Fz(u) is irreducible over the coefficient field k(z), Fz(u) and images are without common divisor over this field. Therefore the Euclidean division algorithm produces two polynomials, Hz(u) and Lz(u), with coefficients which are rational functions of z, such that

images

If we introduce the polynomial images, then the equation

images

becomes an identity on the surface. Finally, by means of the equation Fz(u) = 0, the polynomial Gz(u) may be reduced to one of degree n − 1. Hence any element g of our function field has a unique representation of the form

images

where the Ri(z) are rational functions of z. In this sense the quantities 1, f, …, fn−1 form a basis for the algebraic function field K relative to the ground field k of rational functions of z. Hence one calls n the degree of K over k.

Here one finds a purely algebraic attack on the idea of a function field. One operates with polynomials G(u) with coefficients in the field k. An irreducible polynomial F(u) of degree n is given. By identifying polynomials G(u) which are congruent modulo F(u), the “ring” of polynomials becomes a field K, of degree n over the ground field k,29

If the independent variable z is replaced by any other function, z*, on the surface, then there are infinitely many ways of choosing the function f* on the surface such that all functions may be expressed rationally in terms of z and f*. There is an irreducible algebraic equation F*z*(f*) = 0 relating z* and f*. Also, z* and f* are rational functions of the variables z and f, which are related by Fz(f) = 0; conversely, z and f are rational functions of the variables z* and f*, which are related by F*z*(f*) = 0. Through the birational transformation (z,f) image (z*, f*) the equations

images

turn into each other. The degree of this equation is, naturally, not by any means an invariant under birational transformations. But the genus p is invariant under birational transformations.

If there exists a function z on images which takes each value just once, then the associated algebraic function field is the field of rational functions of z (and p = 0). If there is no function on images which takes each value just once, but if there is a function z which takes each value exactly twice, then we may assume that the algebraic equation (which must be quadratic) determining the function field has the form

images

where the ei are all different. These l points and, if l is odd, the point ∞, are branch points of order 1, and hence the genus p is = (l/2) − 1 if l is even, and = (l − 1)/2 if l is odd. We see that l = 1 or 2 leads again to the rational field p = 0; l = 3 or l = 4 gives p = 1, which is the elliptic case; if l > 4, we get the so-called hyperelliptic function fields.

In an arbitrary algebraic function field of genus p = 1 there always exists a function with two prescribed poles, therefore a function which assumes each value only twice. In every function field of genus 2 we obtain a function of the same sort by dividing two linearly independent Abelian differentials of the first kind. But, starting with p = 3, the hyperelliptic case is no longer the general one.

In the theory of algebraic function fields, two paths are clearly indicated along which one can reach a deeper understanding of the laws governing them. One is that of abstract algebra; here the algebraic concepts of field, field extensions, the degree of a field over the ground field, the degree of transcendence, prime place, etc., are paramount; one admits coefficient fields of characteristic other than zero. Here algebraic construction must furnish that which was accomplished in our development by the Dirichlet principle and the method of integration. The other path, that trod by Riemann, is the “topological,” which we have followed. One may describe Weierstrass’ point of view as an algebraic-function-theoretic one lying between these two extremes: explicit construction reigns, but one always operates in the continuum of complex numbers. Similar remarks apply to the “Kurventheoretiker” of the German and Italian schools. The concept of an analytic form as a two-dimensional manifold lies close enough here, even if the further step, of regarding the topological properties of this manifold as more primitive than all others, is not completed. Furthermore, it is characteristic of the Riemannian type of development that it is always the Riemann surface, not the analytic form, which is regarded as the given object; the construction of an associated analytic form is a principal component of the problem to be solved. To be sure, in Riemann’s treatment itself this point of view does not appear with the complete clarity with which we can now distill it from the works of Prym, Dedekind,30 C. Neumann, and particularly Klein.31

Every closed Riemann surface of genus p may, as we saw, be represented as a multiple-sheeted covering surface over the sphere (with finitely many branch points, and without boundary). There is a great multitude of these “normal forms,” even when one normalizes the number of sheets by the condition n = p + 1 (which is always possible). A much more fundamental significance attaches to the essentially unique normal form of the Riemann surface of arbitrary genus, which is furnished by the theory of uniformization (theory of automorphic functions).

§ 20.    Uniformization

In the theory of uniformization the ideas of Weierstrass and of Riemann grow into a complete unity. With Weierstrass the analytic form (z, u) is described at each individual point by a particular representation with the aid of a parameter t (the “local parameters”): z = z(t), u = u(t). Certainly Riemann obtains a global representation z = z(Image), u = u(Image) of the whole form, but he is forced to regard the parameter Image as a point on a Riemann surface (not as a complex variable in the usual sense). Uniformization theory is concerned with obtaining a global representation z = z(t), u = u(t) with the aid of a parameter t, the uniformizing variable, which varies in a domain of the smooth complex plane. F. Klein and H. Poincaré32 must be named as the true founders of the theory of automorphic functions, to which our problem leads. Their way to the general concepts and results was prepared in the literature by important, but more specialized, investigations of Riemann, Schwarz, Fuchs, Dedekind, Klein, and Schottky. The proof of the possibility of uniformization, based on the concept of a covering surface, was furnished simultaneously in 1907 by P. Koebe and H. Poincaré.33 From then on, Koebe spent his whole scientific life in studying the problem of uniformization thoroughly from all sides, and with the most varied methods.34 To him above all we owe it that today the theory of uniformization, which certainly may claim a central role in complex function theory, stands before us as a mathematical structure of particular harmony and grandeur.

The basic idea of the following proof, to derive the existence of the uniformizing variable from the Dirichlet principle, comes from Hilbert.35

The uniformizing variable t that we seek should be such that it is suitable as a local parameter at each point of the given surface Image. Therefore it must be a uniform function, regular analytic except for poles of order one, on the universal covering surface Image. If we seek that function t which possesses the strongest uniformizing power, then we will attempt to determine t such that it assumes distinct values at any two distinct points of the surface Image; then it will map Image one-to-one and conformally onto a domain of the t-sphere. Then not only the functions on the base surface Image can be represented as uniform functions of t; but also the much larger class of functions (which are in general infinitely many-valued on Image) which arise from any function element on Image which can be continued, without branching, along all paths in Image. And since Image (in contrast to Image) is simply connected, the possibility of such a map does not contradict the analysis-situs properties of Image. We may well drop now the basic assumption of the preceding sections, that Image be closed; this assumption would not simplify anything in uniformization theory.

We obtain the desired uniformizing variable simply by applying the Dirichlet principle, not to Image, but to the universal covering surface Image. We choose a point Image on Image with the local parameter ζ; then we construct on Image, with the aid of the Dirichlet principle, the potential function U which is regular everywhere on Image except at Image, which behaves like Image(1/ζ) at Image, and with the following properties.

(1) The Dirichlet integral of U over all of Image, except for an arbitrarily small ζ-disc about Image, is finite,

(2) For every continuously differentiable function w on Image, with finite Dirichlet integral, which vanishes in a neighborhood of Image, the variation D(U, w) satisfies

Image

U generates a differential on Image; since Image is simply connected, this must be the differential of a certain function

Image

whose real part coincides with U, which is regular analytic everywhere except at Image, and which has a pole of order one at Image. Then τ is a uniformizing variable of the type we seek. The proof of this fact follows in a very elegant fashion with the aid of the following deduction, due to Koebe.36

We prove first the following fact.

If V0 is any real constant, then the points of Image at which V > V0 form a single domain; similarly for the points where V < V0.

Now 1/τ is a local parameter at Image; let K0 : | 1/τ | ≤ a0, be a (l/τ)-disc about Image. Let Image(V0) be the closed set on Image consistingofthepoints where V = V0; then certainly only two of the domains determined by Image(V0) have points in K0. If our claim were false, then among the domains determined by Image(V0) there would be one, say Image, that did not penetrate the neighborhood K0 of Image. Now let ϕ(u) and ψ(u) be any two real functions, defined and continuously differentiable for all real values u. We form the following function w on the surface Image:

Image

This function is everywhere continuously differentiable, provided that

Image

In the neighborhood K0 of Image, w vanishes identically. If Image is any point in Image and if z = x + iy is a local parameter at Image, then

Image

If ϕ, ϕ′, ψ, and ψ′ are bounded functions, then the Dirichlet integral of w over all of Image will be finite. Then, under the given conditions, D(U, w) = 0. Now

Image

If we choose ϕ and ψ such that and ϕ′ and ψ are positive for all values of their argument (except for u = V0 in the case of ψ), then we have a contradiction.37

From the fact proved, and the simple connectivity of Image, one can draw conclusions on the behavior of τ.

(1) dτ has no zeros. If at a point Image on Image, where τ = τ0, dτ = 0, then not ττ0, but (ττ0)1/r (r an integer ≥ 2) would be a local parameter at Image; let us assume r = 2; the proof for larger r is analogous. I set ττ0 = σ2 and draw in the complex σ-plane a disc K, with center σ = 0, so small that it is, via the function σ, the conformal image of a certain neighborhood of the point Image on Image. I take four points Image, Image, Image, Image in K, as indicated in Fig. 9, which are the vertices of a cross formed by two linear segments, α and β, crossing at the origin. At Image and Image, V > V0; at Image and Image, V < V0. I think of this figure transferred to the surface Image; then I can join Image and Image by a curve α′, at every point of which V > V0; likewise, I can join Image and Image by a curve β′ on which V < V0. The intersection number of the two closed curves, α + α′ and β + β′, is thus = 1. But this contradicts the fact that α + α′, as a curve on the simply connected surface Image, must be homologous to zero.

Image

FIGURE 9

(2) If I say that a point of the surface Image, at which τ has the value τ0, lies over the point τ0 of the τ-sphere, then Image becomes a covering surface Image of the τ-sphere or the τ-plane. By what we have proved under (1), this covering is unbranched. There is just one point, Image, over τ = ∞. I follow the line V = V0, starting at τ = ∞, in the τ-plane in the direction from smaller to larger values of U; on Image, the point Image, starting at Image, traces a curve over this line in the τ-plane. If I do not run into any boundary before returning to ∞, then Image describes a closed curve Image, for there is only the single point Image over ∞ then Image covers the line V = V0 in the τ-plane simply. But if I meet an obstruction, then I obtain a curve Image on Image which covers a certain piece U < U1 of the line V = V0 simply. Then I follow the line V = V0, now from larger to smaller values of U, and obtain a curve Image on Image which covers a piece U > U2 of the line simply. The sum Image + Image forms a curve Image through Image, nonclosed, and without end on Image. In either case, the curve Image separates the surface Image, since it is simply connected, into two domains, Image and Image. If, outside of Image, there were points of Image at which V = V0, say in Image, then there would be points in Image at which V < V0 and points at which V > V0. The point set Image(V0) would then determine at least three domains. Therefore Image exhausts the points at which V = V0. Thus, as a covering of the τ-sphere, Image is everywhere at most two sheeted. A value U0 + iV0 will certainly occur once and only once on Image, if V = V0 is a closed curve on Image.

Image

FIGURE 10

We still want to prove carefully that Image must always separate the surface Image. For this purpose we construct a two-sheeted, unbranched, and unlimited covering surface over Image as follows. Cut Image along Image, take two copies of this cut surface and identify the edges of the slits criss-cross. Put abstractly, this amounts to the following. To each point Image of Image we associate two points “over it,” Image and Image. If Image is a point not on Image, τ0 = τ(Image), and Image an arbitrary (ττ0)-disc which does not intersect Image, then the points Image (with upper index 1) which lie over the interior points Image of Image constitute a “neighborhood” of Image; those points Image over the same points Image constitute a “neighborhood” of Image. On the other hand, if Image is on Image, let Image denote an arbitrary (ττ0)-disc. Those points Image whose trace points Image lie inside Image and satisfy the condition VV0, together with all points Image lying over inner points Image of Image which satisfy the condition V < V0, shall constitute a “neighborhood” of Image. The neighborhood of Image is defined analogously. In the last case all the points in Image at which V = V0 certainly belong to Image; hence this definition of the concept of neighborhood is in agreement with all the demands to be made of such a definition. If Image does not separate Image, then it is clear that the manifold just defined also satisfies the condition that any two of its points can be joined by a continuous curve. But the existence of such a covering surface would contradict the fact that Image is simply connected.

The last step in the proof is the demonstration of the following theorem.

There exists at most one real number V0 such that the associated line V = V0 on Image is nonclosed.

For if there were two such lines Image and Image, say Image and Image then one makes use of a closed line U = U0 completely contained in K0 (U0 is chosen large enough). Let Image be one of the pieces Image of the curve Image; let Image be the piece Image of Image. Over the heavily drawn path (Fig. 10, the three segments) in the τ-plane there is a curve in Image, which covers the path simply, and has no ends on Image. This curve separates the simply connected Image into two domains; let Image be that one of the two domains which does not contain the point Image. Again we set

Image

For this function to be continuously differentiable on Image, we must have

Image

Let ϕ, ϕ′, ψ, ψ′ be bounded and let ϕ′ (except for u = U0) and ψ (except for u = Image or Image) be positive.38 Then we get a contradiction of the equation D(U, w) = 0. Thus we have proved the following.

    τ maps the surface Image one-to-one and conformally either onto the complete sphere (case 1)

or

    onto the sphere with one point τ0 removed (case 2)

or

    onto the sphere with a slit V = V0, U1UU2 removed (case 3).

We replace τ by a somewhat different uniformizing variable t. To be sure, t = τ in case 1. In the second case we arrange it so that the point omitted from the sphere is the point at ∞: set t = 1/(ττ0); t maps the covering surface onto the whole plane (without the point at ∞). In the third case we first apply an entire linear transformation so that the slit is given by

Image

Then, with the aid of the formula

Image

the slit τ-sphere is mapped conformally onto the interior of the unit disc | t | < 1 in the t-plane.

The construction we have described of the uniformizing variable t occurs in two clearly separate steps. First, the surface Image solves the problem in so far as it belongs to analysis situs. Then the function theoretic theorem, every simply connected Riemann surface may be mapped conformally onto a domain on the sphere, applied to Image gives the uniformizing variable. By a slight modification of the argument, one can also show that every planar surface may be mapped conformally onto a domain on the sphere. For nonplanar surfaces this is impossible, for reasons of analysis situs: there will not exist even a topological map onto a domain on the sphere. Every uniformizing variable belonging to Image (unbranched relative to Image) will map a certain unbranched, unlimited, planar covering surface Image over Image conformally onto a plane domain. One assigns two uniformizing variables which map the same covering surface Image onto subdomains of the plane to the same class {Image}. The determination of all uniformizing variables requires then the solution of two problems.

(1) The analysis-situs problem: to determine all unbranched unlimited planar covering surface Image of a given surface Image.

(2) The conformal mapping problem: to find all possible conformal maps of a planar surface Image onto a plane domain.

That the last is always possible in at least one (and hence in infinitely many) way (in other words, that every class of uniformizers {Image} conceivable under the analysis-situs condition of planarity actually exists in the function theoretic sense) is the content of Koebe’s general uniformization principle.39 As a matter of fact, this principle goes even further; it includes not only the uniformizing variables which are unbranched relative to Image, but also the great multitude of branched uniformizing variables. But without question, the uniformizing variable which we have set up and denoted by t carries more fundamental significance than any other uniformizing variable.

§ 21.    Riemann surfaces and non-Euclidean groups of motions. Fundamental regions. Poincaré Image-series

To what extent is the uniformizing variable t determined by the properties stated at the end of the last section? That is, in how many ways can one map the surface Image conformally onto the sphere, the plane, or the unit disc? This question obviously comes down to the following: in how many ways can one map the sphere, plane, or disc conformally onto one of these three domains? To begin with, it is clear that the sphere, since it is closed, can be mapped onto neither the plane nor the disc. Also, a conformal map of the plane onto the disc is impossible. If the function t*(t) transformed the t-plane conformally onto the unit disc in the t*-plane, then it would be an entire function whose absolute value would be bounded by one. By Liouville’s theorem, there is no such function (except for the constants, which make no sense here). Furthermore, the following simple theorems hold.

(CASE 1). The set of conformal maps of the sphere(represented in the usual fashion by a complex variable) onto itself is the set of linear transformations.

(CASE 2). The conformal maps of the complex plane onto itself are given precisely by the entire linear transformations.

(CASE 3). Likewise, the open unit disc can be mapped conformally onto itself only by linear transformations.

Case 1. We have to show that if the t-sphere is mapped conformally onto the t*-sphere, t* = t*(t), so that t = 0 goes into t* = 0 and t = ∞ goes into t* = ∞, then the map must be of the form t* = ct, where c is a constant. Now 1/t* is regular at the north pole (t = ∞) of the t-sphere and has a zero there; so t/t* is regular there. Since the relation tt* is one-to-one, t* vanishes only at t = 0, where it has a zero of order one; hence t/t* is regular on the complete t-sphere, and is thus a constant.

Case 2. Again we may assume that t = 0 is mapped into t* = 0. Next one has to prove that

Image

The circle | t* | = R* corresponds to a closed curve Image in the t-plane; let R0 be the maximum distance of a point on Image from the origin, and let R be an arbitrary number > R0. In the disc | t | ≤ R, | t* | assumes its maximum, which must be > R* at some boundary point, say t = t0. The image Image, of the circle | t | = R, in the t*-plane cannot meet the circle | t* | = R*; for the circle | t | = R and the curve Image, of which those two curves are the images, do not meet in the t-plane. Since the point t*(t0) on Image has an absolute value > R*, we must have | t* | > R* for all points on Image. That is, | t | > R0 implies | t* | > R*, and this is the claim (21.1). Hence l/t*(t) is regular at the north pole of the t-sphere and has a zero there. The rest of the argument is exactly the same as in case 1.

Image

FIGURE 11

We settle case 3 with the aid of the so-called Schwarz lemma.40 Again we may assume that t = 0 goes into t* = 0. If I consider the regular function t*/t in the disc | t | ≤ q (< 1), it must attain its maximum absolute value on the boundary, and hence this maximum is < 1/q ( | t* | < 1,| t | = q). Since I can choose q arbitrarily close to 1, | t*/t | ≤ 1 must hold at all points of the open unit disc. In a similar fashion, | t/t* | ≤ 1, and hence | t*/t | = 1. This is possible only if t*/t is a constant of modulus one.

With this the stated claims are demonstrated: the uniformizing parameter t is always uniquely determined to within a linear transformation. The linear transformations which map the open unit disc onto itself have the form

Image

In particular, since the cover transformations of Image are one-to-one conformal maps of Image onto itself, these cover transformations must correspond to linear transformations in the t-plane. Thus there corresponds to the group of cover transformations a certain isomorphic group Γ of linear transformations. No transformation of the group Γ (except the identity) can have a fixed point in the image domain (sphere, plane, or disc); a fixed point is a point that goes into itself under the transformation. Since every linear transformation of the sphere has a fixed point, there are no cover transformations in case I, except the identity. Then Image is the same as Image, and Image is identical, as a Riemann surface, with the sphere. Case I can arise only when the given Riemann surface Image is equivalent to the sphere.

The points of Image over one point of Image appear in the map as a system of points t equivalent under Γ. Such a system Σ has the property that any point of the system can be carried into any other by some transformation of Γ; and also, the image of any point of Σ by any transformation of Γ is another point of Σ (see p. 28). The uniform functions on Image, regular except for poles, appear, when expressed as functions of t, as automorphic functions attached to Γ. That is, functions z(t) which are invariant under the transformations of Γ:

Image

where t* = (αt + β)/(γt + δ) is any transformation in the group Γ. The group Γ must be discontinuous; that is, a system of equivalent points under Γ can never have a limit point in the image domain. If we define a Riemann surface Image by taking the “points” of Image to be the systems of equivalent points under Γ and by carrying the angular measure in the t-plane directly over to Image, then Image is equivalent, as a Riemann surface, to the given Image.

Image is the most appropriate normal form into which every Riemann surface can be brought.

In case 2, Γ must consist of entire linear transformations which have no finite fixed point. The only transformations satisfying this condition are the translations t′ = t + α. A discontinuous group of translations must be of one of the following types.41

(1) Γ contains only the identity. Then Image is identical with Image, and Image is equivalent to the plane (the “Simply punchedsphere, that is, the sphere without the north pole).

(2) Γ consists of the iterations of a single translation

Image

then clearly Image is equivalent to the infinitely long right circular cylinder, and hence, by the Mercator projection, to the doubly punched sphere (the sphere without north and south poles).

(3) Γ will consist of the transformations

Image

where α and β are translations in different directions. Then Image is closed; dt is a uniform differential, regular everywhere on Image, without any zeros, and hence Image is a closed Riemann surface of genus one. Every such surface - whose type (but not its most general form) is the torus - admits in fact a uniformization by the integral of the first kind, by means of which the universal covering surface is mapped onto the plane.

Aside from the few exceptions listed above (Image = sphere, simply or doubly punched sphere, closed surfaces of genus one), case 3, in which the image domain is the open unit disc, always occurs. Since in general the periphery of the unit disc is a natural boundary (cut) for the automorphic functions of t which represent the functions on the base surface Image, one calls t a cut-circle uniformizing variable [Grenzkreis-Uniformisierende].

The group Γ is not determined uniquely by the given surface. For t may be replaced by any variable t′ obtained from t by a linear transformation T0 which preserves the open unit disc. This transforms Γ into the group

Image

We introduce the following terminology. Any point t of the open unit disc is a “point of Image.” For the “straight lines in Image” we take the circular arcs in Image which are orthogonal to the unit circle (the diameters of the unit disc are included in these “straight lines”). Any linear transformation preserving the open unit disc is a “motion in Image”; two point sets in Image are “congruent” if one can be carried onto the other by a “motion.” The usual angular measure is retained. Then the complete Bolyai-Lobatschefsky geometry holds for these “points” and “straight lines.” That is the geometry whose axioms are the same as the axioms of Euclidean geometry, except that the parallel axiom is omitted.42 So we may call Image the non-Euclidean plane. The complex variable t (restricted by the inequality | t | < 1) is to be regarded as a coordinate representing the points in the Lobatschefskian plane, in the same way that the rectangular Cartesian coordinates x, y, or their complex fusion z = x + iy, represent the points of the Euclidean plane. Any linear transformation of t which preserves the unit disc provides another equally valid coordinate system for the Lobatschefskian plane. In this interpretation, Γ now appears as a group of motions of the non-Euclidean plane; or, more precisely, as a representation of such a group by means of a definite coordinate t. If we replace t by t′ obtained from t by a linear transformation preserving the unit disc, then the new group Γ′ thus obtained is to be regarded as a different representation of the same group of motions of the non-Euclidean plane (with the aid of another coordinate t′). There are no rotations in Γ, that is, no motions of Image which leave some point of Image fixed.

Thus to every Riemann surface (aside from the four exceptions already listed) there corresponds a single uniquely determined discontinuous group Γ of motions of the Lobatschefskian plane; and Γ contains no rotations. Two Riemann surfaces are conformally equivalent (as Riemann expressed it, belong to the same class or are realizations of one and the same ideal Riemann surface) if and only if the associated groups of non-Euclidean motions are congruent in the sense of Lobatschefskian geometry. Conversely, to every rotation-free discontinuous group of non-Euclidean motions there corresponds a definite class of Riemann surfaces. (The surface Image constructed on page 169 will serve as a representative of this class.43)

To present a discontinuous group Γ of motions of the n-E plane44 in a more visual fashion, one uses, following Klein and Poincaré, a simple gapless covering of the plane by n-E congruent regions (fundamental regions) which are permuted by the motions of Γ. A dissection of this sort, which is as simple as possible, is provided by the following considerations.

Definition of n-E distance. Let t1 and t2 be any two distinct points in the open unit disc; join them by a n-E line, that is, a circular arc in the Gaussian t-plane which intersects the unit circle orthogonally at t∞1 and t∞2. This circular arc is uniquely determined by t1 and t2. The points on it are to come in the sequence t∞1, t1, t2, t∞2 (Fig. 12). The cross-ratio

Image

has a positive real value > 1, and its real logarithm

Image

Image

Fig. 12. Non-Euclidean segment.

is positive, r (tl, t2) is invariant under all linear transformations preserving the unit disc. The n-E segment t1t2 is n-E congruent to another such segment Image if and only if Image Furthermore, if the three points tl, t2, t3 lie in that order on an n-E line, then

Image

Because of these two facts, we shall speak of the number r (tl, t2) as the non-Euclidean distance between the points t1 and t2. Now we are in a position to measure not only angles but also distances in the n-E plane.

If Image is a system of equivalent points under Γ, if t0 is a single point of Image, and t an arbitrary point of the n-E plane, then because of the discontinuity of Γ there are only a finite number of points of Image whose distance45 from t does not exceed an arbitrary given number R. Thus among the points of Image there is one or several (but certainly only a finite number) of points th such that the distance r(t, th) is the smallest distance of t from any point of Image. Then, as I shall express it briefly, t lies closest to th. If the distance r (t, th) is strictly less than the distance to all other points of Image, then I can find a neighborhood of t such that every point of this neighborhood lies closest to th. Let th be any point of Image; I collect all those t which lie closest to th into a point set Image with “center” th. The resulting sets Image with all centers in Image, have disjoint interiors and cover the whole n-E plane. The transformation of Γ which carries t0 into th carries Image into Image; hence Image is n-E congruent to Image. To every point t there corresponds a point of Image which is equivalent to t under Γ; two distinct interior points of Image are never equivalent. Thus Image has the properties of a fundamental region. Also, Image is convex; that is, if t′ and t″ are any two points of Image. then all points of the n-E segment tt″ joining t′ and t″ belong to Image. The perpendicular bisector Image of the segment t0th is the locus of points equidistant from t0 and th; to every th distinct from t0 we obtain such a line Image. The boundary of the closed set Image is composed of segments of these lines; the interior of Image consists of those points which, for every Image, lie on the same side of Image as t0. All points of the line Image have a distance Image from t0, and there are only a finite number of the centers th whose distance from t0 does not exceed an arbitrarily given 2R. Hence it follows that only a finite number of segments on the boundary of Image has a distance ≤ R from the principal center t0. Thus Image is a convex polygon, with a finite or infinite number of sides, whose vertices (that is, those points which are equidistant from three or more points of the system Image) have no limit point in the finite.

Following Fricke, Image is called a normal polygon belonging to the group Γ. The sides of the normal polygon are associated in pairs; either side of a pair can be carried onto the other by some motion in Γ. Two sides with a common vertex are never associated; for such a common vertex would have to be a fixed point of the motion carrying one side onto the other. These motions carrying one side of Image onto another (that is, the motions carrying Image onto a fundamental region Image which abuts Image along some side of Image) generate the group Γ. That is, every motion of Γ may be obtained by iterating and composing those particular motions. One proves this as follows. Join an arbitrary point th of Image to t0 by a polygonal path which avoids the vertices of the polygonal decomposition {Image}. Now observe the succession of polygons in this decomposition through which the polygonal path passes. Suppose there are in all e polygons Image equivalent to Image(including Image itself) which meet at the vertex 1 of Image. Then there are exactly e centers to which 1 lies closest, and there are e points of Image which are equivalent to 1 under Γ; or, in the terminology of Poincaré, which constitute a cycle of vertices of the polygon Image. The sum of the angles in Image at the vertices of such a cycle is the full angle 1 (= 360°).

The Riemann surface Image(p. 169) belonging to the group Γ is obviously closed if and only if there exists a positive number R such that for every point in the n-E plane there is at least one equivalent point under Γ whose n-E distance from the center t0 is ≤ R. Then every point has a distance ≤ R from the closest point in the system Image; in particular, Image is completely contained in the n-E disc of radius R about t0. Since the vertices of Image cannot have a limit point in the finite, there are, in this case, only a finite number of such vertices : Image has finitely many sides. Let the number of sides of Image, which must be even because of the pairing of the sides, be 2s; let c denote the number of distinct vertex cycles of Image, so that c is also the sum of the angles in Image. We wish to show that the genus p of this closed Riemann surface may be computed from s and c by the simple formula

Image

Since every cycle contains at least three vertices, we have

Image

Therefore 12p − 6 is an upper bound for 2s, the number of sides or vertices of a normal polygon.

We choose t0 = 0. The (n-E as well as Euclidean) linear segments joining 0 to the vertices of Image separate Image into 2s triangles. This provides a triangulation of Image into 2s triangles with 3s edges and c + 1 vertices. Then, from the most basic formula of combinatorial topology, the genus p is in fact determined by the equation

Image

In this book we have adopted the methodological attitude of avoiding boundaries as much as possible, and we have operated with coverings by overlapping neighborhoods rather than with dissections into simple pieces of surface. The appeal here to combinatorial topology would introduce a foreign element. Therefore, instead of this topological proof of the equation (21.3), I prefer a function theoretic proof. It runs as follows.

We take a meromorphic differential dz on the closed surface Image. As we know, its order is 2p − 2. We shall assume for the moment that none of the zeros or poles of dz lie on the sides of Image. Consider the analytic function dz/dt of t in the region Image. Being a convex polygon, is simply connected, and one of the simplest consequences of the Cauchy integral theorem in function theory says that the increase in the azimuth of dz/dt, in tracing the boundary of Image, is equal to the number of zeros less the number of poles of dz/dt in the interior of Image; hence it is = 2p − 2. To determine the increase of the azimuth of dz/dt in tracing the boundary, consider a side σ = t1 t2 = AB of Image and its equivalent side Image where the second comes from

Image

Fig. 13. Determination of the genus from the normal polygon.

the first by the substitution t* = tS of the group Γ. This substitution S carries the polygon Image into an equivalent one which abuts on Image along the side Image. If, in tracing Image in the positive sense, the segment σ is traced from t1 to t2, then the segment σ* will be traced in the sense Image Since dz is invariant under the substitution S, the sum of the contributions of these two segments to the increase of the azimuth is equal to the difference

Image

Here Image denotes the negative of the angle (measured in Euclidean fashion) through which the tangent to σ = AB turns when this side is traced in the positive sense from A to B. This angle is positive and = Image minus the sum of the angles of the n-E triangle 0AB. In this way we get, for the desired increase in the azimuth of dz/dt when the boundary of Image is traced in the positive sense, the value

Image

and the equation claimed, 2p − 2 = sc + 1, follows.

It is inherent in the situation that one cannot completely avoid bordered regions in a dissection into equivalent fundamental regions, but the boundaries here are simple enough to be grasped easily.

If dz has zeros or poles on the boundary of Image, then one can proceed in one of two ways. Either one cuts out disc neighborhoods of these zeros and poles and later lets the radii tend to zero, or one avoids this unpleasant eventuality by a suitable choice of the center t0. If t = a is a zero or pole of dz which lies on the mid-line of Image, that is, if t0 and Image have the same distance from a, then t0 has the same distance from a as from aS− 1. So one need only choose t0 such that it does not lie on any of the mid-lines associated with a pair of distinct but equivalent zeros or poles of dz; then it is certain that no point a lies on the boundary of the normal polygon with center t0. Since only a finite number of these mid-lines come within an n-E distance of the origin ≤ a given R, this is certainly possible. Finally, one can bring the point t = t0 into the point t = 0 by a suitable n-E motion and thus make t = 0 the center of Image.

The dissection of the n-E plane into normal polygons furnishes not only a visualization of the n-E group of motions, but also a means of constructing such groups directly.46

In the group Γ of motions of the non-Euclidean plane (or in the associated manifold Image) we meet the purest embodiment of the concept of a Riemann surface, freed of all accidental properties. To crown the whole development of this part of Riemannian function theory, the solution of the following problem would be desired. Given a Riemann surface in its normal form (that is, the associated group of motions of the non-Euclidean plane), to express each of the uniform or many-valued unbranched functions on the surface by means of a closed analytic formula in terms of the coordinate t of the points in the Lobatschefskian plane. The Image-series introduced by Poincaré47 constitute an important attack on the solution of this problem. If the given group Γ which preserves the open unit disc consists of the substitutions

Image

then, for example,

Image

are Poincaré Image-series. To establish the uniform absolute convergence of these series, take an ordinary disc κ of radius a: |tt0 | ≤ a about an arbitrary point t0 ( | t0 | < 1) so small that the equivalent discs κSi are mutually disjoint. The surface integral, Ji, of

Image

over κ is the (Euclidean) area of κSi; consequently, Image is convergent. If the Euclidean distance from t to t0 satisfies | tt0 | ≤ a/2, then by (12.16)

Image

Therefore

Image

is uniformly convergent in the neighborhood of t = t0. Also,

Image

holds uniformly for | t | ≤ q < 1, for the equivalent points Image condense only on the boundary of the unit disc; the uniform absolute convergence of the two series, Image, and Image′ follows. At t = 0, and at all equivalent points under Γ, Image has a pole of order 1; otherwise Image is regular. At the same points Image′ has poles of order 3 and is regular everywhere else. From the equation

Image

it follows that Image(t)(dt)2 goes into itself under an arbitrary transformation Si of the group

Image

Similarly,

Image

Hence Image(t)/Image′(t) is an automorphic function relative to the group Γ; it is not identically zero, but is has a zero of order 2 at t = 0 (and at every equivalent point). In short, it is a uniform nonconstant function, without essential singularities, on the ground surface Image.

The examples of Image-series given here are of dimension − 4. Poincaré succeeded (by using n-E instead of Euclidean area) in proving48 the absolute convergence of Image-series of dimension − r, for r > 2. Once a Riemann surface is given in its normal form, one might hope, on the basis of Poincaré’s attack, that one could succeed in deriving all the theorems on the existence of functions and integrals on the surface with the aid of analytic formulae. This would be the analog of what has long been possible in the case p = 1 with the theory of elliptic functions. Perhaps one would prefer then to construct Riemannian function theory as follows : with the aid of the Dirichlet principle or a competing method derive, not the existence of functions and integrals, but only the existence of the cut-circle uniformizing variable; then construct the functions and differentials by explicit formulae with this uniformizing variable as argument. The recent extensive investigations of H. Petersson are aimed in this direction. The difficulty lies in the limit passage from r > 2 to r = 2. The structural form of the Poincaré series does not appear to furnish an adequate foundation for this process; Petersson reaches his goal only through a new device of regarding the Image-series as eigenfunctions of certain functionals.49 In spite of the tremendous advance attained by Petersson, it seemed advisable to me to retain, in this book, the independent construction of all functions and differentials by means of the Dirichlet principle.

§ 22.    The conformal mapping of a Riemann surface onto itself

We say that a group of motions of the n-E plane contains infinitesimal transformations if there exist motions in the group which differ arbitrarily little from the identity. A discontinuous group of motions, say Γ, certainly contains no infinitesimal transformations. Also the converse of this theorem is true.

An n-E group of motions is discontinuous if and only if it contains no infinitesimal motions.

To prove this we use, as before, the complex coordinate t. In general, a linear transformation T : tt* has two fixed points on the t-sphere, τ′ and τ″; then the transformation may be written50

images

the constant μ is called the multiplier of T. If T preserves the open unit disc, then there are two possibilities: (1) either τ′ and τ″ both lie on the circumference of the unit disc, and μ is positive (hyperbolic transformation); or (2) τ′ and τ″ are inverse points relative to the unit circle, and τ′ is interior to the same; then | μ | = 1 (elliptic transformation). In the n-E plane, T is a rotation about τ′, and the real number ϕ, determined to within an additive integer by the equation μ = ex(ϕ), is the angle of rotation of T.

The parabolic transformations insert themselves as the trasnitional case between the hyperbolic and elliptic transformations. A parabolic transformation has only one fixed point, τ′. If it preserves the open unit disc, then it has the form

images

We base our proof on the following lemma.

Lemma. Let

images

be two sequences of points, in the n-E plane, which converge to the same point t0 (| t0 | < 1). Let Tn be an n-E motion which carries tn into images. If there exists a neighborhood of t0 which contains no fixed point of any of the motions Tn(n = 1,2,3,…), then Tn converges to the identity.

Let images (| images | ≤ 1) and images be the two fixed points of Tn (they may coincide); then there is a positive number l such that

images

holds for all n. If I set

images

then I can assume that

images

for all n. Then every one of the four differences

images

is > l. If Tn is not parabolic, then it has the form

images

If I set t = tn, and thus t* = images, in here, I get

images

This yields the estimate

images

In place of (22.3) I can write

images

for every linear substitution with the fixed point images must be of this form. Also if the transformation Tn is parabolic, I can put it in the form (22.5); in this case μn = 1 and (22.4) certainly holds. We get the limit equation

images

If again we set t = tn and t* = images now in (22.5), we find

images

Finally, if we bring (22.5) into its natural form

images

we get the values

images

which differ from the corresponding coefficients in the identity substitution

images

by less than

images

This completes the proof of the lemma.

Now let Γ* be an n-E group of motions, without infinitesimal transformations, and let τ be a fixed point in the n-E plane of some rotation belonging to Γ*. By Sτ I denote all rotations in Γ* which have the fixed point τ; they form a group in themselves. If I normalize the angles of rotation ϕ of the Sτ by the condition 0 image ϕ < 1, then, because of the lack of infinitesimal transformations, there exists one of the Sτ, say images, with the smallest angle of rotation ϕ0 > 0. The angle of rotation of every Sτ is an integral multiple of ϕ0; for otherwise, by appropriate choice of the integer k, one could generate a transformation Sτ (images)k with a smaller angle of rotation than images. Thus images is a “primitive” transformation in the group of the Sτ; all the others are obtained as powers of images. If one determines the integer h by the condition

images

then clearly (images)h+1 has a smaller angle of rotation, (h + 1)ϕ0 − 1, than images, unless 0 = 1. Therefore ϕ0 = 1/h must hold, and h is the order of the finite cyclic group (Sτ) which consists of the rotations

images

The fixed points of the rotations in Γ* can have no limit point in the finite. If τ (| τ | < 1) were such a limit point and if

images

were a sequence of primitive rotations in Γ*, with distinct fixed points τn which converged to τ, then there would be two a priori possibilities.

(1) The orders hn are bounded for all n. Then there would be infinitely many of the Sn which have the same order h; let them be the transformations images . Then images is infinitesimal for infinitely large n. Therefore this case is ruled out.

(2) The orders hn are not bounded. Then one can choose a subsequence images of the Sn such that the associated orders tend to ∞ and the angles of rotation tend to zero. But that also would contradict the absence of infinitesimal operations in Γ*.

Having settled this, we can now show that a system of equivalent points, under Γ*, can have no limit point in the finite. Suppose that t0 (| t0 | < 1) were the limit of a sequence of equivalent points tn (n = 1, 2, 3, …):

images

Let Sn be the motion of Γ* which carries tn into tn+1. I pick a neighborhood images of t0 which, with the possible exception of t0 itself, contains no fixed point of a rotation in Γ*. By the lemma, the fixed point images(| images | ≤ 1) of Sn can lie outside of images for only finitely many n. Therefore we must have images = t0 for nn0; it is no restriction to assume that n0 = 1. Then all tn result from t1 by rotations about t0, which are contained in Γ*; but by such rotations I can carry t1 into only finitely many different positions. Thus the possibility of a limit point is contradicted, and the proof of the theorem stated at the beginning of § 22 is complete.

With this preparation out of the way, we arrive at the real object of this final section of our exposition. We are concerned with the one-to-one conformal maps of a Riemann surface images onto itself. These maps form a group; the significance of this group for the theory of the Riemann surface images is obviously analogous to that, for example, of the group of motions for metric geometry. Therefore when we formulate the following theorem, due in essence to Klein, we are stating a fact of fundamental importance.

The group of conformal maps of a Riemann surface onto itself is always discontinuous, except for the following seven exceptional cases: images = complete sphere, simply or doubly punched sphere, a disc on the sphere [literally, “skull cap on the sphere”], a punched disc (i.e., a disc without its center), a zone on the sphere (between two circles of latitude51), a closed Riemann surface of genus one.

Proof.52 Let images denote the universal covering surface over images, let t be the cut-circle uniformizing variable which maps images conformally onto the open unit disc, and let Γ be the group of n-E motions belonging to the Riemann surface images. The cases in which the t-sphere or the infinite t-plane, instead of the unit disc, arises may be ignored; our theorem is obviously invalid in those cases (exceptional cases 1, 2, 3, 7). Let C be a conformal map of images onto itself which carries the point images into images. Let images be a point of images over images, images a point of images over images. Because of the simple connectivity of images, there exists a single topological map images of images onto itself such that:

(1) images goes into images, and

(2) every point images on images goes into a point images, whose trace point images is the image under C of the trace images of images.

imagesis also conformal, and therefore appears in the t-plane as a linear transformation T, which preserves the open unit disc; also T carries any system of equivalent points under Γ into a system of the same sort. This last fact is expressed by the equation

images

which states that T and Γ commute; or, for every substitution S in Γ, the “transform” TST−1 also belongs to Γ. Conversely, it is also clear that every linear substitution T that commutes with Γ, and which carries |t| < 1 onto itself, furnishes a conformal map of images onto itself. Obviously, the set of motions of the n-E plane which commute with Γ form a group, Γc, which contains Γ as a subgroup. The discontinuity of this group Γc is to be proved; by the theorem at the beginning of this section, it suffices to demonstrate the absence of infinitesimal operations in Γc.

If S and T are any two linear transformations preserving the open unit disc, then the two points which T throws into the fixed points of S are the fixed points of the transformation

images

If S′ has the same fixed points as S, then either T must carry each of the fixed points of S into itself (that is, if S is not parabolic, T must have the same fixed points as S; if S is parabolic, at least one of the fixed points of T must coincide with the fixed point of S); or T must interchange the two fixed points σ′ and σ″ of the (nonparabolic) transformation S, that is, T must carry σ′ into σ″ and σ″ into σ′. As is clear from (22.2), a parabolic transformation cannot accomplish this last; a nonparabolic T(22.1) can accomplish it only if μ2 = 1, μ = − 1, and T is an elliptic transformation with rotation angle images. Thus it is possible for S and T to commute

images

only if

I. both transformations S and T are nonparabolic and have the same fixed points, or

II. both transformations S and T are parabolic and have the same fixed point, or

III. S and T are elliptic transformations with angles of rotation = images, or

IV. one of the two transformations, S and T, is the identity.

If Γ consists of only the identity, then images is conformally equivalent to the open unit disc, or a disc on the sphere (exceptional case 4).

Now let S be an arbitrary substitution in Γ, distinct from the identity. Assume that, contrary to our claim, there is a sequence of operations Tn ≠ 1 in in Γc, such that Tn converges to the identity. Then both images and images would be elements of Γ. Along with Tn, the last of these operations converges to the identity as n tends to infinity. But since Γ contains no infinitesimal transformations, it must be that

images

holds for nn0; that is, S and Tn commute. We have listed four cases in which this is possible, but since S is not elliptic, only cases I and II remain in question. If we apply this result to all possible operations S ≠ 1 of the group Γ, we find that only two possibilities remain open:

I. All operations in Γ are hyperbolic and have the same two fixed points, which one may assume are + i and − i.

II. All operations in Γ are parabolic and have the same fixed point, which we may take to be + i.

(CASE I). The map

images

carries the open unit disc in the t-plane onto the parallel strip − π/2 < y < + π/2 in the z = (x + iy)-plane; Γ goes into a group of translations parallel to the x-axis ; that is, Γ takes the form

images

Then

images

maps images one-to-one onto the annulus

images

and by stereographic projection this may be turned into a zone on the sphere with the equator for its mid-line (exceptional case 6).

(CASE II.) The map

images

carries the open unit disc in the t-plane into the upper half-plane y > 0, and Γ into a group of translations parallel to the x-axis. So again Γ has the form (22.6) and w = eiaz carries images onto the punched unit disc in the w-plane:

images

With this our principal theorem is completely proved; at the same time we have had the opportunity to observe the applicability of the normal form images of the Riemann surface through an important example. It is trivial that in each of the seven excluded cases there is a continuous group of conformal maps of the surface onto itself. It is also easy to describe these groups completely.

A closed Riemann surface of genus p > 1 admits only a finite number of conformal maps onto itself.53 For a system of equivalent points under this group can have no limit point on the closed surface, and therefore contains only a finite number of points.


1)See Weierstrass, Werke, 2, 49–54, ϋber das sog. Dirichletsche Prinzip (1870).

2) H. A. Schwarz: ϋber einen Grenzübergang durch alternierendes Verfahren, Gesammelte Abh., Vol. II, 133–143; ϋber die Integration der partiellen Differential-gleichung Δu = 0 unter vorgeschriebenen Grenz-und Unstetigkeits-Bedingungen, ibid., 144–171; Zur Integration der partiellen Differentialgleichung Δu = 0, ibid., 175–210. C. Neumann: Ber. Sächs. Akad. d. Wiss. Leipzig, 1870; Vorlesungen über Riemann’s Theorie der Abelschen Integrate, 2nd edition, Leipzig 1884, 388–471. Among the more recent literature I mention in particular the following two works of C. Carathéodory: Elementarer Beweis für den Fundamentalsatz der konformen Abbildung, in the Festschrift zum 50jährigen Doktorjubiläum von H. A. Schwarz, Berlin 1914, 19–41; Conformal Representation, Cambridge Tracts in Math., 28, 1932.

3)See the two papers, ϋber das Dirichletsche Prinzip, in Hilbert’s Gesammelte Abh., 3, Berlin 1935,10–14 and 15–37; and also his paper, Zur Theorie der konformen Abbildung, ibid., 73–80. The second was originally published in the Festschrift zur Feier des 150jährigen Bestehens der Gesellschaft der Wissenschaftenzu Göttingen, 1901, and was reprinted in Math. Ann., 59 (1904). The other two originated in 1905 and 1909.

4)I cite here particularly the following early works which appeared before the first edition of this book: B. Levi, Sulprincipio di Dirichlet, Rend. Circ. Mat. Palermo,22 (1906) 293–360; G. Fubini, II principio di minimo e i teoremi di esistenza per i problemi al contorno relativi alle equazione alle derivate parziali di ordini pari, ibid.,23 (1907) 58–84; H. Lebesgue, Sur le probleme de Dirichlet, ibid., 24 (1907) 371–402; S. Zaremba, Sur le principe du minimum, Bull, de l’Ac. des sciences de Cracovie, July 1909, 206fF; R. Courant, ϋber die Methode des Dirichletschen Prinzips, Math. Ann., 72 (1912) 517–550. What I give here, essentially unchanged, is the proof that I developed in 1913 on the basis of the work of Hilbert, B. Levi, and Zaremba; the attack is simpler than that of Riemann and Hilbert. The proof incorporates the modifications which are the natural consequence of Dieudonne’s improved method of surface integration. Since that time a rich literature on the subject has accumulated. Through Courant, among others, the “direct methods of the calculus of variations” has come to be a familiar term in analysis. For a comprehensive presentation we refer to Courant’s book on the Dirichlet principle, already mentioned in the preface.

5)Gesammelte Abhandlungen, 2, 186–198.

6)J. Hadamard, Sur le principe de Dirichlet, Bull. Soc. math. France, 34 (1906) 135–139. S. Zaremba, loc. cit., 206ff.

7) The Schwarz inequality gives

image

    where image. The rest follows from the Bessel inequality corresponding to the annulus (q, 1):

image

8)The presentation in Klein’s monograph of 1882, already cited, is based on these and similar physical considerations. The diagrams of streams in that monograph are very instructive.

9)Sul principio di Dirichlet, Rend. Circ. Mat. Palermo, 22 (1908) 293–360, §7. For the method under discussion see H. Weyl, The method of orthogonal projection in potential theory, Duke Math. Jour., 7 (1940) 411–444.

10)Let it be permitted here that we use the bar to mean something other than the passage from a complex number to its conjugate.

11)Our proof is closely related to that given by Zaremba in his paper cited above.

12) The “shrunken block” terminology was introduced in §10: Bθ(image): |x | ≤ θb, | y | ≤ θb.

13) It would have been quite possible, and might have even meant a simplification, to consider only Riemann surfaces and harmonic and Abelian differentials. Then h would have been introduced as the number of linearly independent everywhere regular harmonic differentials. Instead, we subsumed Riemann surfaces under smooth oriented surfaces. I make this remark with regard to the tendency in the literature, sometimes explicit, to construct the theory of Riemann surfaces “without topology.” For example, see L. Bieberbach, Über die Einordnung des Hauptsatzes der Uniformisierung in die Weierstrasssche Funktionentheorie, Math. Ann., 78 (1918) 312–331. I can see neither difficulties nor any particular profit in this.

14) Earlier it was customary to derive these symmetry laws (following Riemann and C. Neumann) by integrating around the boundary of the canonically cut surface. Our normalization of the elementary integrals renders this dissection superfluous. I regard this as an essential improvement of the method.

15) See Riemann, Theorie der Abelschen Funktionen, Werke (2nd edition), p. 140, and especially Appell, Journal de mathematiques, 3. ser., 9 (1883), and Acta Mathematica, 13 (1890).

16) This terminology comes from the arithmetic theory of algebraic functions of Hensel and Landsberg; see Hensel and Landsberg, Theorie der algebraischen Funktionen einer Variablen, Leipzig, 1902.

17) Riemann considered only the so-called “general” case (R = p in our notation). The complete result for integral divisors images is due to G. Roch, Jour. f. Math., 64 (1865) 372–376. Fractional divisors were considered by Klein (Riemannsche Flächen I, autograph lectures, Göttingen 1892, pp. 110–111), E. Ritter, Math. Ann., 44 (1894) 314, and Hensel and Landsberg loc. cit., pp. 362–364.

18) Compare the references to Klein and Hensel-Landsberg in the preceding footnote. E. Ritter, loc. cit., argues as follows. If f0 = images′/images is one function which is a multiple of 1/images (images′ integral), then I obtain all such functions f as follows: f = f0f′, f′ a multiple of 1/images′, and thus I come back to the case of an integral divisor images′. But here the proof that [if B + (m + 1 − p) > 0] an f0 exists is lacking.

19) The last method is also available on open Riemann surfaces and yields here immediately the proof that there exist functions on the surface which are not constant.

20) A. Brill and M. Noether, Math. Ann., 7 (1874) 283.

21) K. Weierstrass, Mathematische Werke, Vol. 4, Vorlesungen über die Theorie der Abelschen Transzendenten, Berlin, 1902, pp. 224–225.

22) Riemann, Theorie der Abelschen Funktionen, Jour. f. Math., 54 (1857) = Werke, 2nd ed., pp. 88–142; Über das Verschwinden der Theta-Funktionen, Jour. f. Math., 65 (1865) = Werke, pp. 212–224. Weierstrass, Vorlesungen über die Theorie der Abelschen Transzendenten, Werke Vol. 4. Stahl, Theorie der Abelschen Funktionen Leipzig 1896. H. F. Baker, Abel's theorem and the allied theory, including the theory of theta functions, Cambridge 1897. Prym and Rost, Theorie der Prymschen Funktionen erster Ordnung, Leipzig 1911, 2. Teil, 7. Abschnitt. Krazer, Lehrbuch der Thetafunktionen, Leipzig 1903.

    The significance principal of the inversion problem to us today lies primarily, not in its intrinsic value, but in the splendid developments created by Riemann and Weierstrass in their efforts to solve the problem.

23) Jour. f. Math., 70 (1869) 354–362; 71 (1870) 223–236, and 305–315. See the work of Appell cited on page 133, and also the following. Prym and Rost, Theorie der Prymschen Funktionen erster Ordnung, Leipzig 1911; R. König, Zur arithmethischen Theorie der auf einem algebraischen Gebilde existierenden Funktionen, Ber. d. Verh. Sächs. Ges. Wiss. Leipzig, math.-phys. Kl., 63 (1911) 348–368; O. Haupt, Zur Theorie der Prymschen Funktionen, Math. Ann., 77 (1915) 24–64.

24) First formulated by E. Ritter, Math. Ann., 44, p. 314. (See also the footnote on page. 137). Instead of the symbolism of divisors, Ritter uses the theory of forms on Riemann surfaces, founded by Klein, which makes possible a real representation of divisors with the aid of multiplicative forms. In Math. Ann., 47 (1896) 157–221, aided by the basic theorems introduced into the theory of linear differential equations by Riemann, he generalizes his investigations to systems of n forms which transform in a homogeneous linear way under the influence of a cover transformation S.

25) In case p = 1, X is not present (p. 143); in case p = 2, X consists of a single character system images0.

26) Abel’s theorem [developed in splendid simplicity in the short note, Démonstration d’une propriété générate d’une certaine classe de fonctions transcendentes, Jour. f. Math., 4 (1829) 200–201 = Œuvres complètes, nouvelle édition (1881), 1, pp. 515–517] is more general, in that it concerns not only integrals of the first kind. A paper of Abel, Mémoire sur une propriété générate d’une classe très-étendue de fonctions transcendentes, on this subject, which he submitted to the Paris Academy in 1826, was lost for a long time because of Cauchy’s carelessness; it was first published after Abel’s death: Mémoires présentés par divers savants, 7 (1841) = Abel, Œuvres completes, nouvelle édition (1881), 1. pp. 145–241. Furthermore, the theorem is contained in the manuscript left by Abel, Sur la comparaison des fonctions transcendentes, Œuvres complètes, 2, pp. 55–66. The converse of Abel’s theorem for integrals of the first kind may be read between the lines in Riemann; it was first stated explicitly (without completely adequate proof) by Clebsch, Jour. f. Math., 63 (1864) 198; Clebsch used it for all it was worth in the theory of algebraic curves.

27) See Weierstrass, Werke, 4, pp. 451–456.

28) See the fourth chapter in Krazer, Lehrbuch der Thetafunktionen, Leipzig 1903, where references to the other literature may be found.

29) The z-plane with finitely many points deleted becomes a simply connected surface if one draws linear cuts from these points to infinity. Riemann’s monodromy problem is concerned with a far-reaching generalization of algebraic functions (the meromorphic functions on an n-sheeted Riemann surface over the z-sphere). It asks for a system of n functions analytic in the slit plane, and the functions are to be related across the slits by given linear substitutions. (A side condition ruling out essential singularities must be added.) Hilbert, Grundzüge einer allgemeinen Theorie der Integralgleichungen, Leipzig 1912, 81–108 and S. J. Plemelj, Riemannsche Funktionenscharen mit gegebener Monodromiegruppe, Monatsh. f. Math. u. Physik, Jahrg., 19, 211–245, demonstrated the solvability of the monodromy problem. Then Robert König developed a complete theory of these “Riemann transcendentals,” including the exact analog of the reciprocity laws of § 16. See in particular the following works of König: Die Reduktions- und Reziprozitäts-Theoreme bei den Riemannschen Transzendenten, Math. Ann., 79 (1918) 76–135 (where one also finds his earlier work developed); Die Integrate der Riemannschen Transzendenten, Math. Ann., 80 (1919) 1–28; Die Elementartheoreme bei den Riemannschen Transzendenten, Math. Zeit., 15 (1922) 26–65. See also, R. König and M. Kraft, Über Reihenentwicklung analytischer Funktionen, Jour. f. Math., 154 (1935) 154–173; ibid., Über Primfunktionen, Jour. f. Math., 165 (1931) 96–107. On this foundation is based the work of H. Schmidt, Über multiplikative Funktionen und die daraus entspringenden Differentialsysteme, Math. Ann., 105 (1931) 325–380, and, in most recent times, H. Röhrl, Differentialsysteme, welche aus multiplikativen Klassen mit exponentiellen Singularitaten entspringen, I, Math. Ann., 123 (1951) 53–75; II, ibid., 124 (1952) 187–218; III, ibid., 125 (1953) 448–466. Further: Funktionenklassen auf geschlossenen Riemannschen Flächen, Math. Nachr., 6 (1952) 355–384; Die Elementartheoreme der Funktionenklassen auf geschlossenen Riemannschen Flächen, ibid., 7 (1952) 65–84; Über gewisse Verallgemeinerungen der Abelschen Integrate, ibid., 9 (1953) 23–44.

30) Prym espoused this point of view in his work from 1869 on; here the work of Dedekind, cited on page 33, on elliptic modular functions, from the year 1877, is pertinent.

31) See the author’s article, Topologie und abstrakte Algebra als zwei Wege mathematischen Verständnisses, in Unterrichtsblätter f. Math. u. Naturwiss., 38 (1932) 177–188.

    The sketch of the theory of algebraic functions, which we could give here, is very incomplete. From the older literature, which preceded the development of modem abstract algebra, I mention here, besides the works of Riemann, Weierstrass, F. Klein, C. Neumann, H. F. Baker, and Hensel-Landsberg already cited, the following presentations: Clebsch and Gordan, Theorie der Abelschen Funktionen, Leipzig 1866; A Brill and M. Noether, Über die algebraischen Funktionen und ihre Anwendung in der Geometrie, Math. Ann., 7 (1874) 269–310; Brill and Noether, Die Entwicklung der Theorie der algebraischen Funktionen, Bericht der DMV, 3, Berlin 1894; F. Severi, Lezione di Geometria algebrica, Padova 1908 (as representative of the very fruitful Italian school of algebraic geometers). In all of these works the curve-theoretic point of view predominates. A trail-blazing and classical work for the algebraic-arithmetic foundations of the theory is: R. Dedekind and H. Weber, Theorie der algebraischen Funktionen einer Veränderlichen, Jour. f. Math., 92 (1882) 181–290, (also contained in H. Weber, Algebra, v. Ill, 2nd edition, Braunschweig 1908, p. 623ff.). Also, F. Klein, Riemannsche Flächen I, II, autograph lectures, Göttingen 1892/93. Appell et Goursat, Theorie des fonctions algébriques, Paris 1895. H. F. Baker, Analytic principles of the theory of curves, Cambridge 1933. The first steps in the domain of algebraic functions with arbitrary fields of constants (including those with prime characteristic) were due to H. Hasse, Theorie der Differentiate in algebraischen Funktionenkörpern mit vollkommenem Konstantenkörper, Jour. f. Math., 172 (1943) 55–64, F. K. Schmidt, Zur arith-metischen Theorie der algebraischen Funktionen, I, Math. Zeit., 41 (1936) 415, and André Weil, Zur algebraischen Theorie der algebraischen Funktionen, Jour. f. Math., 179 (1938) 129–133. A modern overall presentation is that of Claude Chevalley, Introduction to the theory of algebraic functions of one variable, Math. Surveys No. 6, Am. Math. Soc., New York 1951.

32) For Poincaré, see, besides the numerous notes in the Comptes Rendus for the years 1881/82, the papers in the Acta Mathematica, Vols. 1, 3, 4, 5 (1882/84). For Klein, see the papers in Math. Ann., Vols. 19, 20, 21 (1882/83), and also the inclusive presentation: Fricke and Klein, Vorlesungen über die Theorie der automorphen Funktionen, Leipzig 1897 to 1912.

33) H. Poincaré, Sur l’uniformisation des fonctions analytiques, Acta Math., 31 (1907) 1-63. P. Koebe, ϋber die Uniformisierung beliebiger analytischer Kurven. Nachr. d. Ges. Wiss. Göttingen, from 1907 on.

34) I give here the titles and places of publication of the larger series of his works; these were frequently announced in shorter notes in the Nachrichten der Gesellschaft der Wissenschaften zu Göttingen and the Sitzungsberichte der Sächsischen Akademie der Wissenschaften zu Leipzig.

    (a) ϋber die Uniformisierung der algebraischen Kurven, I, II, III (first proof of the general fundamental theorem of Klein, the iteration procedure), IV (second existence proof for the general canonical uniformizing variable: method of continuity), Math. Ann., 67 (1909) 145–224 ; 69 (1910) 1–81; 72 (1912) 437–516; 75 (1914) 42-129.

    (b) ϋber die Uniformisierung beliebiger analytischer Kurven, I. Teil: Das allgemeine Uniformisierungsprinzip; II Teil: Die zentralen Uniformisierungsprobleme, Jour. f. Math., 138 (1910) 192–253, and 139 (1911) 251–292.

    (c) Abhandlungen zur Theorie der konformen Abbildung. I. Die Kreisabbildung des allgemeinsten einfach und zweifach zusammenhängenden schlichten Bereichs und die Ränder Zuordnung bei konformer Abbildung, Jour. f. Math., 145 (1915) 177–223. II. Die Fundamentalabbildung beliebiger mehrfach zusammenhängender schlichter Bereiche nebst einer Anwendung auf die Bestimmung algebraischer Funktionen zu gegebener Riemannscher Fläche, Acta Math., 40 (1916) 251-290. III. Der allgemeine Fundamentalsatz der konformen Abbildung nebst einer Anwendung auf die konforme Abbildung der Oberfläche einer körperlichen Ecke, Jour. f. Math., 147 (1917) 67-104. IV. Abbildung mehrfach zusammenhängender schlichter Bereiche auf Schlitzbereiche, Acta Math., 41 (1918) 305-344. V. Fortsetzung, Math. Zeit., 2 (1918) 198-236. VI. Abbildung mehrfach zusammenhängender Bereiche auf Kreisbereiche. Uniformisierung hyperelliptischer Kurven, Math. Zeit., 7 (1920) 235-301.

    (d) Riemannsche Mannigfaltigkeiten und nichteuklidische Raumformen, acht Mitteilungen; Sitzber. Akad. Berlin, 1927, 164–196; 1928, 345–384 and 385–442; 1929, 414–457; 1930, 304–364 and 505–541; 1931, 506–534; 1932, 249–284. Also the prize winning monograph of 1920, first published in Acta Math., 50 (1927) 27–157, Allgemeine Theorie der Riemannschen Mannigfaltigkeiten.

    See also Koebe’s survey lectures: ϋber ein allgemeines Uniformisierungsprinzip, Int. Math. Cong. Rome, Atti (1909) 25–30; Referat über automorphe Funktionen und Uniformisierung, Jahresber. DMV 21 (1912) 157-163; Methodender konformen Abbildung und Uniformisierung, Math. Congr. Bologna, Atti, 3 (1928) 195–203.

    Of the more recent literature, the following should be noted especially: B. L. van der Waerden, Topologie und Uniformisierung der Riemannschen Flächen, Ber. Sächs. Ak. Wiss. Leipzig, math.-phys. Kl., 93 (1941) 147–160; and R. Nevanlinna’s book, Uniformisierung, Berlin 1953.

35) Zur Theorie der konformen Abbildung, Nachr. Ges. Wiss. Göttingen, 1909, 314–323.

36) ϋber die Hilbertsche Uniformisierungsmethode, Nachr. Ges. Wiss. Göttingen, 1910, 61–65.

37) All the required conditions are satisfied if we use the function

Image

38) All the demands are satisfied by [see footnote 37] :

Image

39) See particularly P. Koebe, ϋber die Uniformisierung beliebiger analytischer Kurven, erster Teil; Das Allgemeine Uniformisierungsprinzip, Jour. f. Math., 138 (1910) 192–253.

40) H. A. Schwarz, Gesammelte Abhandlungen, Vol. II, pp. 109-111. The application of this lemma to our problem goes back to Poincare, Acta Math., 4 (1884) 23–232.

41)That the three cases listed exhaust the possibilities follows from the process of the adaptation of a (two-dimensional) number lattice relative to a sublattice. See page 89 of this book or G. Kowalewski, Die komplexen Veränderlichen und ihre Funktionen, Leipzig 1911, pp. 57 f.

42)The model of the plane non-Euclidean geometry which presents itself here is intimately related to the model, discovered by Klein in the year 1871, which is based on Cayley’s metric. Klein, Über die sogenannte Nicht-Euklidische Geometrie, Math. Ann., 4 (1871) 573-625. See also Fricke and Klein, Vorlesungen über die Theorie der automorphen Funktionen, Vol. I, Leipzig 1897, pp. 3-59. The book of Bonola gives a good orientation on non-Euclidean geometry: Non-Euclidean geometry: a critical and historical study of its developments, English translation by H. S. Carslaw, Dover Pubs., 1955 (originally published 1911)

43)See the final remarks in Koebe’s first communication Über die Uniformisierung beliebiger analytischer Kurven, Göttinger Nachrichten, 1907, 209–210.

44) n-E = non-Euclidean

45) The word “distance,” and all other geometric expressions, are to be understood, until further notice, in the n-E sense.

46) Besides the normal polygons, also the “canonical polygons,” whose theory was developed by Fricke (Fricke-Klein, Vorlesungen über die Theorie der automorphen Funktionen, Leipzig, 1897–1912) are of great importance. With their aid, Fricke succeeded in formulating and proving rigorously the theorem which had already been stated by Riemann: the Riemann surfaces of genus p ( >1) form a (6p ‒ 6)-dimensional manifold. These matters are intimately connected with the “method of continuity,” by means of which Klein and Poincaré attempted to prove, simultaneously at the beginning of the 1880’s, the uniformizibility of algebraic forms. See in particular P. Koebe, Math. Ann., 75 (1914) 42–129.

47) Poincaré, Mémoire sur les fonctions fuchsiennes, Acta Math., 1 (1882) 207.

48) For a brief formulation of this proof see H. Petersson: Abh. Math. Sem. Univ. Hamburg, 16 (1948) 127–130; and, in more detail, Über den Bereich absoluter Konvergenz der Poincaréschen Reihen, Acta Math., 80 (1948) 23–63. See also: Über die Transformationsfaktoren der relativen Invarianten linearer Substitutionsgruppen, Monatsh. f. Math., 53 (1949) 17’41.

49) Automorphe Formen als metrische Invarianten, Vols. I and II, Math. Nachr., 1 (1948) 158–212 and 218–257; Über Weierstrass-Punkte und die expliziten Darstellungen der automorphen Formen von reeller Dimension, Math. Zeit., 52 (1949) 32–59.

50) If τ″ = ∞, this becomes

images

51) These constitute according to the breadth of the zone, infinitely many essentially differenr Riemann surfaces.

52) This is Poincaré’s proof: Acta Math. 7 (1885) 16–19. To be sure, Poincaré considers only closed Riemann surfaces; but his proof, word for word, remains valid for open surfaces.

53) For this more special theorem, first stated by H. A. Schwarz, there exist other more algebraic proofs. One of Weierstrass (1875), which was first published in 1895 (Werke, Vol. II, pp. 235–244) ; one of M. Noether (Math. Ann., 20, pp. 59–62 and 21, pp. 138–140; 1882); and one of A. Hurwitz, Math. Ann., 41 (1893) 403–411.