Fact-checked by Grok 2 weeks ago

Geometric function theory

Geometric function theory is a branch of that studies the geometric properties of holomorphic functions, particularly focusing on conformal mappings, univalent functions, and the distortion of domains in the . It emphasizes the interplay between analytic characteristics of these functions—such as coefficients and growth rates—and the geometric features of their images, including shape preservation, boundary behavior, and extremal problems. Central to the field is the analysis of functions in the unit disk, where tools like the hyperbolic metric provide invariant measures of distance and enable bounds on mapping distortions. The foundations of geometric function theory trace back to the mid-19th century, building on Bernhard Riemann's 1851 statement of the in his dissertation, which establishes that any simply connected domain in the (other than the whole plane) is conformally equivalent to the unit disk. This theorem, along with Riemann's development of Riemann surfaces and complex differentiability, shifted complex analysis toward geometric interpretations, influencing subsequent work by figures like and in the late 19th century, who integrated and modular functions. The field gained momentum in the early with contributions such as the (first proved in a special case in 1869–70, with canonical proof in 1907, extended to the Schwarz–Pick theorem by Pick in 1915), which quantifies contractions under holomorphic self-maps of the unit disk, and Karl Weierstrass's factorization theorems from the 1870s that linked zero distributions to geometric structures. In the 1930s–1950s, and others formalized key results on quasiconformal mappings (named by Ahlfors in 1935) and extremal length (introduced by Ahlfors and Beurling in 1950), solidifying geometric function theory as a distinct discipline. Key concepts in geometric function theory include the Schwarz-Pick theorem, a generalization of the that bounds hyperbolic distances under holomorphic maps between hyperbolic domains, and the Koebe 1/4 theorem, which guarantees that univalent functions cover a significant portion of the plane near the origin. The theory also encompasses coefficient problems for classes of univalent functions, such as the Bieberbach conjecture (now de Branges' theorem, proved in 1985), which bounds the second coefficient of normalized univalent functions. Broader topics involve Loewner equations for evolving slit mappings, Carathéodory's convergence theorem for boundary extensions, and connections to and dynamical systems, with applications in areas like , , and the Schramm-Loewner evolution in probability. These elements highlight the field's role in bridging pure , , and .

Overview and Foundations

Definition and Scope

Geometric function theory is a branch of that investigates the geometric properties of holomorphic functions, particularly how these functions map domains in the while preserving or distorting geometric features such as angles, distances, and shapes. It emphasizes the spatial behavior of these mappings, including aspects like of distances, injectivity of the mapping, and the correspondence between interior points and behavior. Unlike purely algebraic or analytic treatments in classical , which often focus on series expansions or residue theorems, geometric function theory prioritizes metrics that capture intrinsic geometry, such as hyperbolic distances over ones, to quantify how mappings preserve or alter the structure of domains. The scope of geometric function theory encompasses a range of topics centered on conformal invariance, quasiconformality, univalence, and extremal metrics, all aimed at understanding the optimal or extremal behaviors of holomorphic mappings between domains. Conformal mappings, for instance, serve as angle-preserving transformations that maintain local geometry, while quasiconformal mappings allow controlled distortion to study more general homeomorphisms. Univalence addresses injectivity, ensuring one-to-one mappings that avoid overlaps, and extremal metrics, often derived from principles like extremal length, provide tools to find mappings that minimize or maximize geometric quantities such as area or length. This focus enables applications in solving value problems geometrically, where holomorphic functions model the distribution of measures on boundaries. A representative example is the role of holomorphic functions as conformal maps that preserve angles at every point, facilitating the visualization of domain transformations in the complex plane without altering infinitesimal shapes. The field emerged in the 19th century from Bernhard Riemann's foundational work on conformal mappings, which laid the groundwork for analyzing how analytic functions geometrically reshape the plane.

Historical Development

The foundations of geometric function theory were laid in the mid-19th century through Bernhard Riemann's 1851 doctoral dissertation, which introduced the concept of conformal mappings and explored the geometry of multiply connected domains in the complex plane. This work emphasized the geometric interpretation of analytic functions, shifting focus from algebraic properties to spatial configurations and paving the way for subsequent developments in mapping theory. In the late 19th and early 20th centuries, key advancements built on these ideas, with Hermann Schwarz contributing in the 1860s through his work on conformal mappings of polyhedral surfaces and symmetry principles that influenced distortion estimates. Henri Poincaré extended these concepts in the 1880s by integrating hyperbolic geometry into complex analysis, notably via his 1882 papers on the uniformization theorem, which classified Riemann surfaces and their conformal equivalences. Paul Koebe advanced the field in the early 1900s, particularly with his 1907 conjecture on the growth and covering properties of univalent functions, which highlighted geometric bounds for analytic mappings. The mid-20th century saw significant progress in extremal problems, exemplified by Ludwig Bieberbach's 1916 conjecture on the coefficients of univalent functions, which posited sharp bounds and became a central challenge in the theory. This was resolved in 1985 by Louis de Branges through his proof establishing the inequality for all coefficients, marking a major milestone after decades of effort. Concurrently, the theory of quasiconformal mappings emerged, initiated by Herbert Grötzsch in his 1928 paper introducing mappings with bounded distortion as a generalization of conformal ones, and further developed by in the 1940s with extremal quasiconformal mappings and quadratic differentials. In the modern era, Charles Loewner's 1923 theory provided a framework for representing conformal mappings via equations, enabling analysis of evolving domains and later extensions to multislit and chordal variants. Post-1980s developments incorporated computational methods, such as numerical algorithms for approximating conformal and quasiconformal maps, facilitating applications in and . Key events include the international conferences on univalent functions, beginning with the 1957 Analytic Functions conference at Princeton, which fostered collaboration and highlighted extremal problems.

Core Concepts in Mapping and Distortion

Conformal Mappings

In geometric function theory, a conformal mapping is defined as a f: D \to \Omega between domains in the where the derivative f'(z) \neq 0 at every point in D. This condition ensures that f is locally invertible and preserves oriented angles, with the angle of rotation at each point z given by \arg(f'(z)). Consequently, infinitesimal shapes and orientations are maintained up to a uniform scaling and , making conformal mappings essential for studying local geometric properties of domains. A key property of conformal mappings is their behavior as local isometries up to scaling: the infinitesimal Euclidean length ds in the domain D is transformed to |f'(z)| \, ds in the image \Omega, where |f'(z)| acts as the local scaling factor. The class of conformal mappings is closed under composition and inversion, and it includes all transformations, which are invariant under further transformations of the domains. Geometrically, this invariance facilitates the analysis of domain equivalences. Additionally, since the real part of a is , conformal mappings provide a geometric method for solving boundary value problems for by transforming complex geometries into simpler ones where solutions are known. Representative examples illustrate these properties. The exponential map f(z) = e^z conformally maps horizontal strips in the complex plane, such as \{z : -\pi < \Im(z) < \pi\}, onto the punctured plane \mathbb{C} \setminus \{0\}, preserving angles while exhibiting periodic behavior due to the imaginary part. Another classic example is the Joukowski transformation f(z) = z + \frac{1}{z}, which conformally maps the exterior of the unit disk to the exterior of a symmetric airfoil shape, aiding in the geometric modeling of fluid flow around such profiles. In geometric interpretation, Möbius transformations, as a subclass of conformal mappings, send generalized circles (circles or straight lines) to generalized circles, enabling the study of domain boundaries and symmetries. More broadly, conformal mappings serve to uniformize domains by establishing bijective correspondences that simplify global structure, as exemplified by the Riemann mapping theorem, which asserts the existence of a conformal map from any simply connected proper subdomain of the plane onto the unit disk.

Quasiconformal Mappings

Quasiconformal mappings extend the concept of conformal mappings by permitting a controlled amount of distortion, providing a framework for studying homeomorphisms that nearly preserve angles and shapes. A homeomorphism f: \Omega \to \Omega' between domains in the complex plane is K-quasiconformal, for K \geq 1, if it satisfies the Beltrami equation \partial_{\bar{z}} f = \mu(z) \partial_z f almost everywhere, where \mu is the Beltrami coefficient with \|\mu\|_\infty \leq k = (K-1)/(K+1) < 1. This condition ensures that the mapping distorts infinitesimal circles into ellipses with bounded eccentricity, quantifying the deviation from conformality, which occurs precisely when K=1 and \mu = 0, reducing to holomorphic functions. Key properties of quasiconformal mappings include bounded analytic distortion, captured by the dilatation H(z, f) = \limsup_{r \to 0} \frac{\sup_{|h|=r} |f(z+h) - f(z)|}{\inf_{|h|=r} |f(z+h) - f(z)|} \leq K almost everywhere, where the maximum and minimum are taken over directions in the tangent space. They preserve sets of positive Lebesgue measure, as the Jacobian determinant J_f = |\partial_z f|^2 - |\partial_{\bar{z}} f|^2 \geq (1 - k^2) |\partial_z f|^2 > 0 ensures that the image of a positive-area set has positive area. The existence of such mappings is guaranteed by the measurable Riemann mapping theorem, which states that for any measurable Beltrami coefficient \mu with \|\mu\|_\infty < 1, there exists a unique (up to post-composition with a holomorphic automorphism) quasiconformal homeomorphism f solving the Beltrami equation with that \mu, and dilatation K = (1 + \|\mu\|_\infty)/(1 - \|\mu\|_\infty). Geometrically, quasiconformal mappings measure how far a deformation strays from angle-preserving behavior while maintaining topological and metric control, playing a central role in the uniformization of by enabling the comparison of complex structures through bounded distortions. A representative example is the affine stretch map f(x + iy) = x + i c y with $0 < c < 1, which stretches vertically by factor c and has dilatation K = \max(1, 1/c), illustrating simple non-conformal distortion. In applications, such mappings model earthquakes in , where quasiconformal deformations along geodesic paths simulate shear transformations on hyperbolic surfaces, connecting distinct points in via paths of bounded distortion.

Properties of Analytic Functions

Univalent Functions

In geometric function theory, a holomorphic function f defined on a domain \Omega \subset \mathbb{C} is called univalent if it is injective, meaning f(z_1) = f(z_2) implies z_1 = z_2 for all z_1, z_2 \in \Omega. This injectivity ensures that f maps \Omega bijectively onto its image, preserving the local conformal structure while providing a global one-to-one correspondence. Univalent functions form the foundation for studying conformal mappings of simply connected domains, as their images are also simply connected regions in the complex plane. The class S consists of all univalent functions analytic in the unit disk \mathbb{D} = \{z \in \mathbb{C} : |z| < 1\}, normalized by the conditions f(0) = 0 and f'(0) = 1. These functions admit a power series expansion f(z) = z + \sum_{n=2}^\infty a_n z^n for z \in \mathbb{D}. The normalization fixes the behavior at the origin, allowing for uniform estimates on growth and distortion across the class. A key property is that the image f(\mathbb{D}) under such functions is a simply connected domain containing the origin, and for certain subclasses like convex univalent functions, the image domains are convex. Another important property is the subordination principle, which relates functions within the class S. If g and f are in S with g(\mathbb{D}) \subset f(\mathbb{D}), then g(z) = f(\omega(z)) for some holomorphic \omega: \mathbb{D} \to \mathbb{D} with \omega(0) = 0, implying bounds like |g'(0)| \leq |f'(0)|. This principle facilitates comparisons between subordinate functions and extremal ones, aiding in the derivation of inequalities for coefficients and mappings. Central to the theory are geometric estimates that bound the growth and distortion of univalent functions. The Koebe growth theorem states that for f \in S and |z| = r < 1, |f(z)| \leq \frac{r}{(1 - r)^2}, with equality for rotations of the Koebe function. Complementing this, the Koebe distortion theorem provides bounds on the derivative: \frac{1 - r}{(1 + r)^3} \leq |f'(z)| \leq \frac{1 + r}{(1 - r)^3} for |z| = r < 1, again with equality attained by the Koebe function and its rotations. These estimates quantify how much univalent mappings can stretch or compress distances while remaining injective. The Koebe function k(z) = \frac{z}{(1 - z)^2} serves as the canonical example and extremal function in class S, mapping \mathbb{D} onto the complex plane minus the ray [1/4, \infty). Its coefficients are a_n = n, achieving the boundary of many inequalities, such as the Bieberbach inequality, which asserts that for f(z) = z + \sum_{n=2}^\infty a_n z^n \in S, |a_n| \leq n for all n \geq 2, with equality for rotations of the . This inequality, originally proved for n=2 and conjectured generally, underscores the role of coefficient bounds in controlling geometric properties.

Multivalent Functions

In geometric function theory, multivalent functions, often termed p-valent functions, are holomorphic mappings from a domain in the complex plane to another such that each value in the complex plane is attained at most p times, counting multiplicities. This finite multiplicity distinguishes them from univalent functions (p=1) and emphasizes the geometric structure of mappings with controlled branching. The concept originates from early studies in complex analysis. Key properties of p-valent functions include the presence of branch points, where the derivative vanishes, resulting in critical values in the image where the local multiplicity exceeds one. These branch points govern the function's behavior in covering space theory, as p-valent functions induce branched p-sheeted coverings of the target domain, with ramifications occurring at critical points. The valence p quantifies this global multiplicity, and the provides a tool to count preimages: for a closed contour enclosing the domain, the number of solutions to f(z) = w (counting multiplicity) equals \frac{1}{2\pi} \Delta \arg f(z), where \Delta \arg f(z) is the change in argument along the contour. This principle, applied to suitable contours, verifies the p-valence by ensuring no value is exceeded p times. Geometrically, p-valent functions map domains to images that form p-sheeted branched coverings, naturally represented on to resolve the branching. For a holomorphic map of degree p from a compact of genus g to one of genus g', the relates the topology via the total branching order B: $2g - 2 = p(2g' - 2) + B, where B = \sum (e_q - 1) sums the excess multiplicities e_q - 1 over all branch points q. This formula, a cornerstone for analyzing branching in multivalent mappings, highlights how local critical behavior influences global surface topology. Examples include the power function f(z) = z^p, the prototypical p-valent map from the unit disk to itself, which has a single branch point at z=0 of order p-1, and the algebraic function f(z) = \sqrt{z}, a 2-valent map with a branch point at 0, illustrating double-sheeted covering of the plane minus the origin.

Analytic Continuation

Analytic continuation is the technique of extending the domain of a holomorphic function f defined on an open connected set U \subset \mathbb{C} to a larger open connected set V \supset U, such that the extension remains holomorphic on V and agrees with f on U. This is achieved by constructing holomorphic functions on a covering of V by open sets that overlap with U, ensuring agreement on the intersections where the domains overlap. The process relies on the local uniqueness of power series representations of holomorphic functions, allowing step-by-step extension across overlapping disks. In the context of geometric function theory, analytic continuation highlights geometric constraints imposed by the domain's topology and the function's singularities. The monodromy theorem guarantees that if a holomorphic function element admits analytic continuation along every polygonal path within a simply connected domain D, then the resulting continuation is independent of the path chosen and yields a single-valued holomorphic function on D. However, in multiply connected domains, path dependence can occur, leading to monodromy—multi-valued extensions that permute function values upon encircling loops. Natural boundaries represent insurmountable geometric barriers to continuation; for instance, lacunary power series with Hadamard gaps, where the ratios of consecutive exponents exceed a fixed \lambda > 1, exhibit the unit circle |z| = 1 as a natural boundary, featuring a of singularities that prevent extension beyond the disk of . A primary method for performing involves expanding the function in at points along a desired and successively extending these series to overlapping regions, provided the path avoids singularities. The great provides insight into behavior near essential singularities, stating that in any punctured neighborhood of such a point, a assumes every complex value, with at most one possible exception, infinitely often; this underscores the dense geometric complexity near isolated singularities, limiting continuation options. Uniqueness of analytic continuation in simply connected domains follows directly from the identity theorem, which asserts that two holomorphic functions agreeing on a set with a limit point must coincide everywhere in the connected domain. Illustrative examples demonstrate these principles. The square root function \sqrt{z}, initially defined as the principal branch on \mathbb{C} \setminus (-\infty, 0], exhibits a branch point at z = 0; analytic continuation around a closed path encircling 0 results in a sign change, yielding a multi-valued extension that cannot be single-valued on a simply connected domain containing the origin. Similarly, the Gamma function \Gamma(z), originally defined by the integral \Gamma(z) = \int_0^\infty t^{z-1} e^{-t} \, dt for \operatorname{Re}(z) > 0, is extended meromorphically to \mathbb{C} \setminus \{0, -1, -2, \dots\} using the reflection formula \Gamma(z) \Gamma(1 - z) = \frac{\pi}{\sin(\pi z)}, which relates values across the positive real axis and poles, enabling continuation while preserving holomorphy in the cut plane. Riemann surfaces serve as a geometric tool to resolve the multi-valuedness arising in such continuations, representing the function as single-valued on a branched covering space.

Geometric Aspects of Special Functions

Polynomials and Their Geometric Properties

In geometric function theory, polynomials are regarded as entire functions of finite degree, meaning they are holomorphic everywhere in the with only finitely many non-zero coefficients in their expansion. This finite degree distinguishes them from more general entire functions and allows for a detailed of their properties. Geometrically, the roots of a can be visualized using Newton diagrams, which provide insight into the valuations of the roots in valued fields; for a polynomial P(x) = \sum p_k x^k, the is the lower of points (k, \ord(p_k)), where the negative slopes correspond to the valuations of the roots with multiplicity. A key property of polynomials is that they map the point at infinity to infinity, reflecting their behavior as proper maps on the . Critical points occur where the f'(z) = 0, and these points play a central role in understanding the of the . Sendov's conjecture posits that for a of degree n \geq 2 with all zeros in the closed unit disk, each zero is within distance 1 of some critical point; this remains unproven in general, though verified for degrees up to 8 and for sufficiently large n. Polynomials exhibit distinctive geometric features in the , including asymptotic rays defined by external angles that describe the approach to along specific directions. These rays are crucial in , where they parameterize the boundary behavior near . In this context, polynomial-like generalize polynomials: a f: U' \to U of degree d \geq 2, with U \Subset U' open disks, is proper and holomorphic, and by the straightening , it is hybrid equivalent to a of degree d, preserving dynamical structures like the filled K_f = \{ z \in U' \mid f^n(z) \in U' \ \forall n \geq 0 \}. A representative example is the quadratic polynomial f(z) = z^2 + c, whose J_f = \partial K_f varies dramatically with c; for c = 0, K_f is the closed unit disk, while for c = -1, it forms the connected "" set with an attracting period-2 cycle, and for c = -2, it degenerates to the interval [-2, 2]. The distribution of is further illuminated by the Gauss-Lucas theorem, which states that the roots of f' lie in the of the roots of f, implying that critical points remain geometrically confined within the root configuration. The Grace–Walsh–Szegő theorem states that if a multiaffine with real coefficients has all its zeros in a (such as the open upper half-plane), then for any univariate section obtained by setting all but one variable to the same value in that , the resulting also has all its zeros in the . Univalent polynomials, which are injective, are limited to degree 1, as higher-degree examples inevitably fold the plane.

Riemann Surfaces

A Riemann surface is defined as a one-dimensional , equipped with an atlas of charts mapping to the \mathbb{C} such that transition maps between overlapping charts are holomorphic functions. This structure allows for the natural extension of from the plane to more general domains, where functions can be defined holomorphically despite topological complexities. Compact oriented Riemann surfaces are topologically classified by their g, a non-negative that counts the number of "handles" on the surface, with g=0 corresponding to the , g=1 to the , and g \geq 2 to higher-genus surfaces. The \chi, a key topological invariant, satisfies \chi = 2 - 2g for these compact surfaces. Riemann surfaces are constructed in various ways to resolve the multi-valuedness of complex functions, particularly algebraic ones. For algebraic functions defined by polynomial equations f(z, w) = 0, the associated arises as a multi-sheeted of the , where branch points correspond to singularities and are resolved by lifting the function to separate sheets connected along branch cuts. Near these branch points, local solutions are given by Puiseux series expansions of the form w = \sum_{k \geq k_0} a_k (z - z_0)^{k/n}, which parametrize the branches and determine the ramification structure. This multi-sheeted structure eliminates artificial branch cuts in the , allowing single-valued holomorphic extensions of multi-valued functions like the inverse of z^2 + 1. Geometrically, Riemann surfaces admit canonical metrics via the uniformization theorem, which states that every simply connected Riemann surface is conformally equivalent to one of three model spaces: the hyperbolic unit disk with its Poincaré metric, the Euclidean plane, or the spherical Riemann sphere, depending on the surface's topology. For compact surfaces of genus g \geq 2, the uniformization yields a hyperbolic metric, while genus $1 admits a flat Euclidean structure, and genus &#36;0 a spherical one. The moduli space of genus g Riemann surfaces parametrizes isomorphism classes of these structures and forms a complex manifold of dimension $3g - 3 for g \geq 2, reflecting the degrees of freedom in deforming the surface while preserving its complex structure. Examples include the torus, realized as the elliptic curve \mathbb{C}/\Lambda where \Lambda is a lattice, which carries a flat metric uniformized by the plane; higher-genus surfaces can be visualized as a sphere with g handles attached, admitting hyperbolic geometry. The sheeted covering construction of Riemann surfaces facilitates analytic continuation by providing a topological framework where paths avoiding branches can be lifted unambiguously, enabling global definitions of functions that are locally analytic.

Extremal and Optimization Problems

Extremal Length and Teichmüller Theory

Extremal length provides a conformal that quantifies the "connectivity" of curve families within a domain, playing a central role in geometric function theory for analyzing moduli problems and quasiconformal distortions. For a family of curves \Gamma in a domain D \subset \mathbb{C}, the extremal length \lambda(\Gamma, D) is defined as \lambda(\Gamma, D) = \sup_{\rho} \frac{\left( \inf_{\gamma \in \Gamma} \int_{\gamma} \rho \, ds \right)^2}{\int_D \rho^2 \, dA}, where the supremum is taken over all measurable functions \rho: D \to [0, \infty)&#36; such that \int_D \rho^2 , dA < \infty. This quantity measures the maximal ratio of squared minimal length (under the metric ds = \rho |dz|) to the area of the domain, achieved by an extremal density \rho$ that is constant on the domain for simple cases like annuli. The extremal length is invariant under conformal mappings of the domain, making it a robust tool for classifying up to conformal equivalence. It is closely related to the conformal modulus of curve families; for instance, the modulus M(\Gamma, D) of an annulus separating two boundaries is the reciprocal of the extremal length of the family of curves connecting those boundaries, M(\Gamma, D) = 1 / \lambda(\Gamma, D). This connection facilitates applications in quasiconformal extensions, where extremal length bounds the dilatation of mappings preserving certain curve families, ensuring controlled distortion in higher-genus surface deformations. A classic example is the Grötzsch problem, which seeks the ring domain separating two given non-overlapping compact sets E_1 and E_2 in the plane with extremal modulus. The solution minimizes the extremal length of the family of curves connecting E_1 to E_2, yielding the conformal invariant that determines the maximal quasiconformal extension constant K for mappings between such domains; specifically, for the unit disk with a radial slit, the extremal length computes to \pi / \log 2, establishing a benchmark for boundary separation problems. In Teichmüller theory, extremal length extends to study the moduli space of Riemann surfaces by incorporating quasiconformal deformations. The Teichmüller space T_g for a surface of genus g \geq 2 consists of equivalence classes of marked Riemann surfaces, where two markings are equivalent if there exists a quasiconformal map isotopic to the identity between them. Deformations are parameterized by Beltrami differentials \mu, which are -1,1-forms on the surface with \|\mu\|_\infty < 1, satisfying the Beltrami equation \bar{\partial} f = \mu \partial f for the quasiconformal map f defining the new complex structure. The geometry of T_g is governed by the Teichmüller metric, which quantifies distances between deformed structures via quasiconformal dilatations. For points in T_g represented by Beltrami differentials \mu and \nu, the distance is given by d(\mu, \nu) = \frac{1}{2} \inf \log K(\phi), where the infimum is over all quasiconformal maps \phi such that \mu(\phi) = \nu, and K(\phi) is the quasiconformal dilatation of \phi. This metric, complete and of negative curvature, integrates extremal length principles to measure minimal distortion paths in the space, with extremal geodesics often realized by Teichmüller mappings driven by holomorphic quadratic differentials. Quasiconformal mappings serve as the primary tool for defining this equivalence, linking static conformal invariants to dynamic surface moduli.

Coefficient Problems for Univalent Functions

The class \mathcal{S} consists of all analytic univalent functions f in the unit disk \mathbb{D} = \{z \in \mathbb{C} : |z| < 1\} normalized by the conditions f(0) = 0 and f'(0) = 1, so that f(z) = z + \sum_{n=2}^\infty a_n z^n. Coefficient problems in this class focus on obtaining sharp bounds for quantities such as |a_n|, \operatorname{Re}(a_n), and related functionals, which provide insights into the geometric properties of the image f(\mathbb{D}). These bounds are extremal in nature, often achieved by specific functions that maximize growth or distortion while preserving univalence. A fundamental property underlying these bounds is the area theorem, which asserts that for f \in \mathcal{S}, \sum_{n=2}^\infty \left( n |a_n|^2 - |a_{n-1}|^2 \right) \leq 1, with equality if and only if f is a rotation of the Koebe function. This inequality arises from applying Green's theorem to the inverse mapping and reflects the limited "area" covered by the omitted values outside f(\mathbb{D}). The class \mathcal{S} also exhibits rotation invariance: if f \in \mathcal{S}, then e^{-i\theta} f(e^{i\theta} z) \in \mathcal{S} for any real \theta, which implies that extremal bounds can be assumed without loss of generality for functions aligned in a particular direction. The Koebe function k(z) = \frac{z}{(1 - z)^2} = z + \sum_{n=2}^\infty n z^n serves as the canonical extremal example in \mathcal{S}, mapping \mathbb{D} onto \mathbb{C} minus the radial slit (-\infty, -1/4]. It achieves the maximum |a_n| = n among all f \in \mathcal{S}, as conjectured by Bieberbach and later proved in full generality. A key subclass relevant to coefficient maximization is the starlike functions, defined by \operatorname{Re}\left( \frac{z f'(z)}{f(z)} \right) > 0 for z \in \mathbb{D}; the Koebe function belongs to this subclass, and many coefficient bounds sharpen when restricted to it. Notable results include the Fekete–Szegő inequality, which provides the sharp bound |a_3 - \mu a_2^2| \leq 1 + \frac{2 e^{-2\mu}}{1 - \mu} for real $0 \leq \mu < 1, and |a_3 - a_2^2| \leq 1 for \mu = 1, with equality for rotations of the when \mu = 1. This inequality controls the interplay between the first few coefficients and has applications in estimating the radius of univalence for partial sums. Loewner chains offer a parametric approach to coefficient evolution: any f \in \mathcal{S} can be embedded in a family f_t(z) satisfying the driven by a continuous function \lambda(t) on the unit circle; this method facilitates deriving recursive bounds on a_n through time evolution from the identity map. The growth theorem complements these coefficient estimates by bounding the function itself: for f \in \mathcal{S} and |z| = r < 1, |f(z)| \geq \frac{r}{(1 - r)^2}, with equality for the Koebe function. This lower bound ensures that f(\mathbb{D}) contains the disk |w| < 1/4 (via the Koebe 1/4 theorem) and establishes the scale of expansion near the boundary. The Bieberbach conjecture, resolved affirmatively, posits that |a_n| \leq n for all n, representing the infinite-dimensional case of these coefficient optimizations.

Fundamental Theorems

Riemann Mapping Theorem

The Riemann mapping theorem asserts that if D \subset \mathbb{C} is a simply connected domain that is not the entire complex plane and z_0 \in D is a fixed point, then there exists a unique conformal bijection f: D \to \mathbb{D}, where \mathbb{D} denotes the open unit disk \{ w \in \mathbb{C} : |w| < 1 \}, such that f(z_0) = 0 and f'(z_0) > 0. This result, first formulated by in his 1851 doctoral dissertation, establishes the conformal equivalence of simply connected proper domains to the unit disk under normalized conditions. The proof proceeds by constructing a family \mathcal{F} of holomorphic functions from D to \mathbb{D} that are injective, fix z_0 at 0, and are normalized to maximize |f'(z_0)|. This family is nonempty, as one can map a square root branch to obtain an initial injective map to \mathbb{D}. By , the locally bounded family \mathcal{F} is normal, so a converges locally uniformly to a holomorphic limit function f, which is nonconstant and maps into \overline{\mathbb{D}}. Injectivity of f follows from Hurwitz's theorem applied to the sequence, ensuring that zeros of f(w) - f(z) for z \neq w imply non-injectivity only if the limit is constant, which it is not. Surjectivity onto \mathbb{D} is established by contradiction: if some point w_0 \in \mathbb{D} is omitted, a Blaschke factor adjustment yields a new map in \mathcal{F} with larger derivative at z_0, violating maximality. Uniqueness arises from the applied to compositions of such maps. Geometrically, the theorem implies that all simply connected proper domains in the complex plane are conformally equivalent to the unit disk, preserving local angles and enabling the transfer of analytic properties between domains. For domains bounded by polygons, the Schwarz-Christoffel formula provides an explicit integral representation for the inverse Riemann map from the unit disk to the polygonal domain, parametrizing the prevertices on the unit circle to match the exterior angles. Extensions of the theorem address boundary behavior, particularly through Carathéodory's work, which shows that for multiply connected domains, sequences of simply connected subdomains converging in the Carathéodory sense yield Riemann maps whose limits extend the mapping properties, though direct conformal equivalences to annuli or punctured disks require the broader . A key quantity is the conformal radius of D at z_0, defined as $1 / f'(z_0), which scales the domain locally like the unit disk and equals the radius of the largest inscribed disk centered at z_0 only if D itself is a disk.

Schwarz Lemma

The Schwarz lemma is a fundamental result in that provides sharp bounds for holomorphic functions mapping the unit disk to itself and fixing the origin. Specifically, let f: \mathbb{D} \to \mathbb{D} be holomorphic, where \mathbb{D} = \{ z \in \mathbb{C} : |z| < 1 \} is the open unit disk, and suppose f(0) = 0. Then, |f(z)| \leq |z| for all z \in \mathbb{D}, and |f'(0)| \leq 1, with equality in either inequality holding for some z_0 \in \mathbb{D} \setminus \{0\} (or at the derivative) if and only if f(z) = e^{i\theta} z for some real \theta. The proof proceeds by considering the auxiliary function g(z) = f(z)/z for z \neq 0, with g(0) = f'(0), which extends holomorphically to all of \mathbb{D} by the Riemann removable singularity theorem. Since |f(z)| < 1 on \mathbb{D}, it follows that |g(z)| \leq 1/|z| on any circle |z| = r < 1. Applying the maximum modulus principle to g on the disk |z| \leq r yields |g(z)| \leq 1 for |z| < r, and letting r \to 1 gives |g(z)| \leq 1 on \mathbb{D}. Thus, |f(z)| = |z| \cdot |g(z)| \leq |z|, and |f'(0)| = |g(0)| \leq 1. Equality in |g(z)| \leq 1 implies g is constant by the maximum principle, so f(z) = e^{i\theta} z. Geometrically, the Schwarz lemma interprets such maps as contractions with respect to the hyperbolic (Poincaré) metric on the unit disk, where the distance between points z, w \in \mathbb{D} is \rho(z, w) = \tanh^{-1} \left| \frac{z - w}{1 - \bar{z} w} \right|. The Pick-Nevanlinna generalization states that f satisfies \rho(f(z), f(w)) \leq \rho(z, w) for all z, w \in \mathbb{D}, with equality if and only if f is an automorphism of \mathbb{D}. This non-expansion property underscores the rigidity of holomorphic self-maps of the disk. Applications include the classification of the automorphism group of \mathbb{D}, consisting precisely of the Möbius transformations f(z) = e^{i\theta} \frac{z - a}{1 - \bar{a} z} for |a| < 1 and real \theta, which preserve the disk and act transitively. Another key use is Julia's lemma, which describes the boundary behavior: if f: \mathbb{D} \to \mathbb{D} is holomorphic with f(0) = 0 and \liminf_{r \to 1^-} |f(r \zeta)| / r \geq \alpha > 0 for some \zeta \in \partial \mathbb{D}, then near \zeta, f approaches the boundary in a controlled angular sector. In the context of univalent functions, the lemma yields basic bounds on coefficients, such as |a_1| \leq 1 for normalized mappings.

Maximum Modulus Principle

The maximum modulus principle states that if f is a in a bounded D \subset \mathbb{C} and continuous up to the \partial D, then the maximum of |f(z)| on the closed domain \overline{D} is attained on the \partial D. Moreover, if f is non-constant, this maximum cannot be achieved at any interior point of D. A standard proof proceeds by contradiction. Suppose |f(z_0)| = M = \max_{\overline{D}} |f(z)| for some interior point z_0 \in D. Since f is holomorphic and non-zero at z_0 (otherwise a zero would contradict the maximum by the open mapping property), consider the function g(z) = f(z) e^{\alpha z} where \alpha is chosen such that \operatorname{Re}(\alpha (z - z_0)) \leq 0 in D with strict inequality except at z_0. Then |g(z)| < M in D \setminus \{z_0\} but |g(z_0)| = M, contradicting the mean value property of the harmonic function \log |g|, which implies \log |g(z_0)| equals its average over any small disk around z_0. An alternative proof uses the subharmonicity of \log |f|: if a maximum occurs interiorly, \log |f| would be constant, hence f constant. This principle yields key geometric consequences. Non-constant holomorphic functions are open mappings: they map open sets to open sets, as an interior maximum for |f - w| with w in the image would imply constancy. Applying the principle to the entire plane \mathbb{C} shows that bounded entire functions are constant (Liouville's theorem), since otherwise |f| would achieve no maximum on compact disks but remain bounded, a contradiction. The Schwarz lemma follows as a direct application on the unit disk. The principle extends to harmonic functions: a non-constant harmonic function on a bounded domain attains neither maximum nor minimum in the interior. For unbounded domains, the Phragmén-Lindelöf principle provides a growth-controlled analog, bounding holomorphic functions in sectors or strips by their boundary values under suitable conditions at infinity. A related minimum modulus principle holds for non-vanishing holomorphic functions: if f has no zeros in D and is continuous to \partial D, then the minimum of |f(z)| on \overline{D} occurs on \partial D, unless f is constant. This follows by applying the maximum principle to $1/f.

Riemann-Hurwitz Formula

The Riemann-Hurwitz formula provides a fundamental relation between the Euler characteristics of compact Riemann surfaces connected by a holomorphic branched covering. For a non-constant holomorphic map f: X \to Y of degree m between compact connected Riemann surfaces X and Y, the formula states that the Euler characteristic \chi(X) satisfies \chi(X) = m \cdot \chi(Y) - \sum_{p \in X} (e_p - 1), where e_p denotes the ramification index of f at the point p \in X, and the sum is over all points of ramification (with multiplicity if necessary). This can equivalently be expressed in terms of genera g_X and g_Y, using the relation \chi = 2 - 2g for a compact Riemann surface of genus g, yielding $2g_X - 2 = m(2g_Y - 2) + \sum_{p \in X} (e_p - 1).[42] A topological proof of the formula proceeds via triangulation of the surfaces. Triangulate Y compatibly with the branch values of f, ensuring that vertices include all branch values and that the triangulation lifts appropriately to X. The preimage under f of this triangulation yields a triangulation of X, where the number of vertices, edges, and faces in X is computed by summing local degrees: for non-ramified points, the local degree is 1, while at ramification points it is e_p > 1. The Euler characteristic \chi(X) = V_X - E_X + F_X then relates to \chi(Y) = V_Y - E_Y + F_Y by accounting for the total branching contribution \sum (e_p - 1), leading directly to the formula. Geometrically, the formula quantifies how branching affects the of the covering surface, with the total branching number B = \sum_{p \in X} (e_p - 1) = 2g_X - 2 - m(2g_Y - 2) measuring the "deficit" in induced by ramification points. This branching enforces constraints on possible genera for given degrees and base surfaces, with applications to classifying Riemann surfaces via their branched covers, such as in the construction of hyperelliptic curves as double covers of the . A classic example is an m-sheeted unramified cover of the (g_Y = 1, \chi(Y) = 0), which yields another (g_X = 1) since B = 0. In contrast, for a double cover (m=2) of the (g_Y = 0, \chi(Y) = 2) branched at $2g_X + 2 simple points (each with e_p = 2), the formula gives B = 2g_X + 2 - 2 \cdot 2 = 2g_X - 2, matching the total branching and confirming the genus g_X for hyperelliptic surfaces defined by equations like w^2 = f(z) of degree $2g_X + 2. Another example is a map f: \mathbb{P}^1 \to \mathbb{P}^1 of degree m, where the branching index B = 2(m-1) arises from the critical points, preserving \chi(X) = 2. The Bieberbach conjecture, now de Branges' theorem, states that if f(z) = z + \sum_{n=2}^\infty a_n z^n belongs to the class S of normalized univalent analytic functions in the unit disk, then |a_n| \leq n for every n \geq 2, with equality if and only if f is a rotation of the Koebe function k(z) = z / (1 - z)^2. This key inequality provides sharp bounds on the Taylor coefficients of such functions. Proposed by Ludwig Bieberbach in 1916, the conjecture was verified for small n (up to 6) through partial results by Bieberbach, Löwner, Garabedian-Schiffer, and others, but the general case proved elusive until Louis de Branges established it in 1985. De Branges' proof proceeds by establishing a stronger result: the Milin conjecture on the majorization of certain functionals involving the logarithmic coefficients of univalent functions, which implies both the Robertson conjecture (on bounds for bounded univalent functions) and the original Bieberbach conjecture via the Lebedev-Milin inequality. The proof relies on Loewner chains, which parametrize univalent functions (schlicht functions in German terminology) via slit mappings evolving under a differential equation, and incorporates area-preserving extensions to maintain normalization. Central to the argument are methods analyzing families of functions with positive real part in the unit disk, represented via Herglotz integrals, leading to a multiparameter inequality resolved using the Askey-Gasper inequality for hypergeometric functions. Geometrically, the theorem yields sharp bounds on the images of univalent functions, including the growth theorem |f(z)| \leq |z| / (1 - |z|)^2 for |z| < 1, which implies the Koebe one-quarter theorem: the complement of f(\mathbb{D}) omits no disk of radius $1/4 centered at the . De Branges' techniques also extend to related extremal problems, such as refinements of the for univalent functions, where the classical Bohr radius $1/3 for bounded analytic functions is improved using coefficient majorization.

References

  1. [1]
    Geometric Function Theory 2018 - staff.math.su.se
    The course aims to introduce students to geometric function theory in a broad sense, and to define concepts and present techniques required in modern ...
  2. [2]
    [PDF] Janowski type functions related to $(j, k) - AIMS Press
    Mar 25, 2025 · Geometric Function Theory (GFT) is a field within complex analysis that focuses on the geometric and analytic properties of holomorphic ...
  3. [3]
    [PDF] The hyperbolic metric and geometric function theory
    As a consequence we obtain a number of results in geometric function theory. ... This paper is intended to be both an introduction to the hyperbolic metric and a ...
  4. [4]
    [PDF] ON SOME HISTORICAL ASPECTS OF THE THEORY OF RIEMANN ...
    the theory of complex functions came about in the period from 1880 to around 1910. Geometric function theory is the use of geometric, chie y. Riemannian ...<|control11|><|separator|>
  5. [5]
    [PDF] geometric function theory - JKU
    GEOMETRIC FUNCTION THEORY. WS 1995/96. J.B. Cooper. I. Basic ... A useful and beautiful introduction to function theory is the classical text of Ahlfors.
  6. [6]
    [PDF] Conformal invariants: topics in geometric function theory
    This book covers topics in geometric function theory, including conformal invariants, functions of complex variables, and Riemann surfaces.
  7. [7]
    Geometric Function Theory - an overview | ScienceDirect Topics
    Geometric function theory is defined as the study of schlicht mappings and their applications, primarily focusing on conformal mappings within the broader ...
  8. [8]
    Hermann Schwarz (1843 - 1921) - Biography - MacTutor
    Hermann Schwarz worked on the conformal mapping of polyhedral surfaces onto the spherical surface and on a problem of the calculus of variation, namely surfaces ...
  9. [9]
    [PDF] Poincaré's Path to Uniformization | Connemara Doran
    In the three long Acta papers which followed in 1882 and 1883, Poincaré developed the theory from this starting point, and the main lines of his exposition have ...
  10. [10]
    Paul Koebe (1882 - 1945) - Biography - MacTutor
    Koebe was awarded his secondary school certificate in 1900. In his lectures at the University of Berlin on the theory of elliptic functions, Weierstrass took ...Missing: early | Show results with:early
  11. [11]
    Untersuchungen über schlichte konforme Abbildungen des ... - EuDML
    Untersuchungen über schlichte konforme Abbildungen des Einheitskreises. I. K. Löwner · Mathematische Annalen (1923). Volume: 89, page 103-121; ISSN: 0025 ...Missing: paper | Show results with:paper
  12. [12]
    A proof of the Bieberbach conjecture - Project Euclid
    1985 A proof of the Bieberbach conjecture. Louis Branges. Author Affiliations +. Louis Branges1 1Purdue University. DOWNLOAD PDF + SAVE TO MY LIBRARY. Acta ...Missing: URL | Show results with:URL
  13. [13]
    [PDF] On five papers by Herbert Grötzsch - HAL
    Dec 16, 2019 · Abstract. Herbert Grötzsch is the main founder of the theory of quasicon- formal mappings. We review five of his papers, written between ...Missing: 1920s 1940s
  14. [14]
    A commentary on Teichmüller's paper "Extremale quasikonforme ...
    Nov 4, 2015 · Abstract:We provide a commentary on Teichm{ü}ller's paper "Extremale quasikonforme Abbildungen und quadratische Differentiale" (Extremal ...
  15. [15]
    [PDF] Coefficient Problems in Geometric Function Theory Iason Efraimidis
    Since the 1980's researches in univalent functions tried to extend this theory to harmonic mappings, especially after an influential paper of Clu- nie and Sheil ...
  16. [16]
    THE BIEBERBACH CONJECTURE Proceedings of the Symposium ...
    42. , "On certain coefficients of univalent functions," in Analytic functions,. Princeton. Univ. Press, Princeton, N. J., 1960, pp. 159-194. 43. , On certain ...
  17. [17]
    [PDF] arXiv:1803.09113v3 [math.MG] 11 Dec 2019
    Dec 11, 2019 · In fact, conformal mappings in the plane correspond to the holomorphic functions on. C with non-zero derivative on their respective domain, and ...<|control11|><|separator|>
  18. [18]
    [PDF] Introductory Complex Analysis in 30 Hours ——–186 problems in 36 ...
    Jan 13, 2024 · with a pure rotation and thus preserves oriented angles. To see ... Some conformal mapping building blocks to consider are ez, Logz, zn ...<|control11|><|separator|>
  19. [19]
    From i to ζ: the basics of complex analysis
    A function f ∈ C1(Ω) is called conformal if df =0inΩ and df preserves the angle and orientation at each point. Thus, the holomorphic functions are precisely ...
  20. [20]
    Conformal Mappings in Euclidean Space
    Rigidity will prevent a wide and flexible class of conformal transformations in dimensions n ≥ 3, and as we have said, this is a prime motivation for the ...
  21. [21]
    [PDF] Numerical conformal mapping - arXiv
    Jul 22, 2025 · Conformal mapping may be the best-known topic in complex analysis. Any simply connected nonempty domain Ω in the complex plane C (assum-.Missing: holomorphic | Show results with:holomorphic
  22. [22]
    [PDF] arXiv:math/0311480v6 [math.DS] 28 Oct 2007
    Oct 28, 2007 · This article is one step in a program to describe the dynamics of the family of exponential maps Eκ : z 7→ exp(z) + κ and the structure of its ...<|separator|>
  23. [23]
    [PDF] arXiv:1712.09730v2 [physics.flu-dyn] 5 Dec 2022
    Dec 5, 2022 · One of the well-known analytic methods for calculating the equivalent of airflow over a 2-dimensional airfoil shape has involved conformal ...
  24. [24]
    [PDF] Quasiconformal Mappings Christopher J. Bishop Stony Brook ...
    The purpose of these notes is to introduce readers to the basic results about quasiconformal maps of planar domains and their application to vari-.
  25. [25]
    [PDF] Complex earthquakes and Teichmüller theory
    Mar 4, 1996 · Abstract. It is known that any two points in Teichmüller space are joined by an earthquake path. In this paper we show any earthquake path.
  26. [26]
    [PDF] On Certain Properties of P-valent Functions Defined by a ... - Hikari Ltd
    Definition 1.1. A function )( zf is called a p-valent function if it does not take any value more than p times. Definition 1.2. A function. )( pAf. ∈ is said ...
  27. [27]
    Analytic Continuation -- from Wolfram MathWorld
    Analytic continuation (sometimes called simply "continuation") provides a way of extending the domain over which a complex function is defined.
  28. [28]
    Monodromy Theorem -- from Wolfram MathWorld
    If a complex function f is analytic in a disk contained in a simply connected domain D and f can be analytically continued along every polygonal arc in D,
  29. [29]
    ON LACUNARY SERIES TAKAFUMI MURAI § 1. Introduction We ...
    An x- lacunary series is called Hadamard lacunary. Proposition 4 shows that an L-lacunary series is Hadamard lacunary. For an L-lacunary sequence E = {nk)^x.
  30. [30]
    Picard's Great Theorem -- from Wolfram MathWorld
    Every analytic function assumes every complex value, with possibly one exception, infinitely often in any neighborhood of an essential singularity.
  31. [31]
    [PDF] Branch Points and Branch Cuts (18.04, MIT). - MIT Mathematics
    Oct 11, 1999 · Thus log(z) is analytic in region R, as long as a branch of log(z) is chosen. Obviously, we may enlarge the region R, and log(z) in this ...
  32. [32]
    Reflection Relation -- from Wolfram MathWorld
    Perhaps the best known example of a reflection formula is the gamma function identity Gamma(z)Gamma(1-z)=pi/(sin(piz)), (1) originally discovered by Euler ...Missing: original | Show results with:original<|control11|><|separator|>
  33. [33]
    [PDF] lectures on the theory of entire functions - Uni DUE
    May 26, 2014 · Clearly, each polynomial is an entire function. In this case only a finite number of the coefficients an in (1.1) differs from zero. If an ...Missing: geometric | Show results with:geometric
  34. [34]
    [PDF] Newton polygons - UBC Math
    Oct 21, 2018 · Newton polygons are associated to polynomials with coefficients in a discrete valuation ring, and they give information about the valuations ...
  35. [35]
    Sendov's conjecture for sufficiently-high-degree polynomials
    Abstract. Sendov's conjecture asserts that if a complex polynomial f of degree n⩾2 n ⩾ 2 has all of its zeros in closed unit disk {z:|z|⩽1} { z : | z | ⩽ 1 } , ...
  36. [36]
    [PDF] On the dynamics of polynomial-like mappings - Cornell Mathematics
    The Straightening Theorem (1,1) allows one to deduce various properties of polynomial- like mappings from the analogous properties of polynomials. In particular ...
  37. [37]
    [PDF] Julia Sets and the Mandelbrot Set
    Several more filled Julia sets for quadratic polynomials are shown in Figure 4. All of these are for polynomials of the form f(z) = z2 + c for some constant c, ...
  38. [38]
    [PDF] The Gauss–Lucas theorem - UCR Math Department
    The Gauss–Lucas theorem says that for any complex polynomial P, the roots of the derivative P0 lie in the convex hull of the roots of P. In other words, ...Missing: source | Show results with:source
  39. [39]
    None
    ### Summary of the Grace-Walsh-Szegő Theorem and Its Application to Interlacing Zeros of Orthogonal Polynomials
  40. [40]
    [PDF] Riemann Surfaces
    Dec 2, 2004 · complex manifold. So a Riemann surface is the same thing as a 1 ... a complex torus C/Λ (genus 1). Notice that this shows that some ...
  41. [41]
    [PDF] Riemann Surfaces - Berkeley Math
    Then, ˜U is homeomorphic to U. Proof of the Monodromy Theorem: Let (R, r, π, F) be the Riemann surface of f, π(r) = a. Let ˜U be the component of π−1(U) ...<|separator|>
  42. [42]
    [PDF] arXiv:1706.04386v1 [math.AG] 14 Jun 2017
    Jun 14, 2017 · The Riemann surface of an algebraic function on the plane represents a complex ... Puiseux series are called Puiseux. 1When talking of a (p ...<|control11|><|separator|>
  43. [43]
    [PDF] Simple proofs of uniformization theorems - arXiv
    Oct 4, 2005 · Each simply-connected Riemann surface is conformally equivalent to either unit disc, or C, or the Riemann sphere. In the proofs of the ...
  44. [44]
    [PDF] Universal moduli spaces of Riemann surfaces - arXiv
    Nov 26, 2016 · We construct a moduli space for Riemann surfaces that is universal in the sense that it represents compact Riemann surfaces of any finite genus.
  45. [45]
    Conformal invariants and function-theoretic null-sets
    Aug 28, 2006 · Conformal invariants and function-theoretic null-sets ... Article PDF. Download to read the full article text. Use our pre-submission checklist.
  46. [46]
    [PDF] Introduction to Teichmüller Theory - UC Davis Math
    Aug 31, 2008 · This set of notes contains basic material on Riemann surfaces, Teichmüller spaces and. Kleinian groups. It is based on a course I taught at ...Missing: seminal | Show results with:seminal
  47. [47]
    [PDF] Introduction to Teichmüller theory - IMJ-PRG
    The Teichmüller space of a surface S is the deformation space of complex structures on S and can also be seen as a space of hyperbolic metrics on S. The aim ...
  48. [48]
  49. [49]
    Univalent Functions | SpringerLink
    Univalent Functions ; Overview. Authors: P. L. Duren ; Softcover Book USD 119.99. Price excludes VAT (USA) ; Hardcover Book USD 119.99 ; Explore related subjects.
  50. [50]
    [PDF] The Riemann Mapping Theorem - UChicago Math
    The Riemann mapping theorem is a major result in the study of complex functions because it states conditions which are sufficient for biholomorphic equivalence.
  51. [51]
    [PDF] History of Riemann Mapping Theorem - Stony Brook University
    Poincaré and uniformisation. The theory of conformal mappings took a novel turn with the arrival of Poincaré and problems to do with the theory of algebraic.Missing: uniformization | Show results with:uniformization<|separator|>
  52. [52]
    [PDF] Riemann Mapping Theorem
    Abstract The purpose of this report is to reproduce the proof of Riemann Mapping. Theorem. We assume the readers are familiar with the basic results in real ...
  53. [53]
    [PDF] The Riemann Mapping Theorem Christopher J. Bishop
    The celebrated Riemann mapping theorem says that any two simply connected planar domains (other that the whole plane) can be mapped to each other by a conformal ...
  54. [54]
    conformal radius - PlanetMath.org
    Mar 22, 2013 · By uniqueness of the map arising from the Riemann mapping theorem we can see that γ is also unique. Thus we could define the conformal radius ...
  55. [55]
    [PDF] Geometric Complex Analysis - Imperial College London
    May 9, 2017 · A pair of harmonic functions u and v defined on the same domain Ω ⊆ C are called harmonic conjugates, if they satisfy the Cauchy-Riemann ...
  56. [56]
    [PDF] Chapter 1.1 Schwarz's lemma and Riemann surfaces
    Jan 1, 2011 · The main connection between analytical and geometrical aspects of the theory is the invariant version of Schwarz's lemma proved by Pick, stating ...
  57. [57]
    [PDF] Complex Analysis review notes for weeks 1-6
    One significant application of Schwarz's Lemma is the classification of all conformal self-maps of the unit disk. Suppose f : D → D is a bijective analytic ...
  58. [58]
    [PDF] The Schwarz Lemma at the Boundary - Mathematics
    Jan 13, 2010 · Abstract: The most classical version of the Schwarz lemma in- volves the behavior at the origin of a bounded, holomorphic function.
  59. [59]
    [PDF] Ahlfors, Complex Analysis
    Complex Analysis has successfully maintained its place as the standard elementary text on functions of one complex variable. There is, never-.
  60. [60]
    Ueber Riemann'sche Flächen mit gegebenen Verzweigungspunkten
    Hurwitz, A. Ueber Riemann'sche Flächen mit gegebenen Verzweigungspunkten. Math. Ann. 39, 1–60 (1891).
  61. [61]
    [PDF] The Riemann-Hurwitz Formula - webspace.science.uu.nl
    This paper is meant to indicate the enormous influence the “Riemann theorem” had and still has in mathematics. Basically no new results are contained in this.
  62. [62]
    Bieberbach Conjecture -- from Wolfram MathWorld
    The nth coefficient in the power series of a univalent function should be no greater than n. In other words, if f(z)=a_0+a_1z+a_2z^2+.Missing: original | Show results with:original
  63. [63]
    A proof of the Bieberbach conjecture | Acta Mathematica
    Oct 26, 1984 · De Branges, L.,A proof of the Bieberbach conjecture. Preprint E-5-84, Leningrad Branch of the V. A. Steklov Mathematical Institute, 1984.Missing: URL | Show results with:URL
  64. [64]
    [PDF] A proof of the Bieberbach conjecture
    An elementary argument shows that the Robertson conjecture implies the Bieberbach conjecture [17]. The logarithmic coefficients of a power series f(z) with ...Missing: resolution | Show results with:resolution
  65. [65]
    [PDF] The Bieberbach Conjecture - Mathematics Department
    De Branges' version was published in 1985 [5]. (See also [15], [10], and. [9] for a slightly different version of the proof and for more detailed commentary ...Missing: resolution | Show results with:resolution
  66. [66]
    (PDF) Bieberbach's conjecture, the de Branges and Weinstein ...
    Aug 10, 2025 · The Bieberbach conjecture about the coefficients of univalent functions of the unit disk was formulated by Ludwig Bieberbach in 1916 [4].Missing: resolution | Show results with:resolution
  67. [67]
    [PDF] Bieberbach's Conjecture, the de Branges and Weinstein Functions ...
    Abstract. The Bieberbach conjecture about the coefficients of univalent functions of the unit disk was formulated by Ludwig Bieberbach in 1916 ...Missing: resolution | Show results with:resolution
  68. [68]
    [PDF] Bohr's Phenomenon for Some Univalent Harmonic Functions
    Bohr's Phenomenon for Some Univalent Harmonic Functions. 253. Since h is univalent in D, so by De Brange's Theorem, we have |am| ≤ m and hence we get. (3.7).