Fact-checked by Grok 2 weeks ago

Montel's theorem

Montel's theorem encompasses key results in regarding the normality of families of holomorphic or meromorphic functions on a domain in the , introduced by the French mathematician Paul Montel in his foundational work on normal families around 1912. The classical version of Montel's theorem states that a family of holomorphic functions on an \Omega \subset \mathbb{C} is —meaning every sequence in the family has a subsequence that converges uniformly on compact subsets of \Omega to a holomorphic function— the family is locally uniformly bounded. This boundedness condition ensures via the Arzelà-Ascoli theorem, providing compactness in the space of holomorphic functions. A more advanced formulation, known as the Great Montel's Theorem, applies to meromorphic functions: a family of meromorphic functions on \Omega is normal if there are three distinct points w_1, w_2, w_3 in the extended complex plane \hat{\mathbb{C}} such that each function in the family omits these values, i.e., f(\Omega) \cap \{w_1, w_2, w_3\} = \emptyset for all f in the family. This result generalizes the basic theorem by replacing boundedness with the omission of three values, leveraging the spherical metric on \hat{\mathbb{C}} to control growth. These theorems are pivotal in complex function theory, underpinning proofs of major results such as the , which establishes biholomorphic equivalence between simply connected domains and the unit disk by applying Montel's theorem to families of normalized injective mappings to extract a limit . They also facilitate extensions to Picard's theorems on value distribution and the study of iteration in , highlighting the compactness properties of normal families in understanding global behavior of analytic . Montel's original contributions, detailed in his memoir, laid the groundwork for these developments by adapting compactness ideas from to the holomorphic setting.

Background Concepts

Normal Families

A family \mathcal{F} of holomorphic functions on a domain \Omega \subset \mathbb{C} is normal if every sequence \{f_n\} \subset \mathcal{F} admits a subsequence that converges uniformly on every compact subset of \Omega either to a holomorphic function on \Omega or to infinity in the extended complex plane (meaning |f_{n_k}(z)| \to \infty uniformly on compact subsets). This notion of convergence, known as locally uniform or convergence, ensures that the limit function (or infinity) is well-behaved across the domain. The space \mathcal{H}(\Omega) of all holomorphic functions on \Omega, equipped with the , provides the natural framework for understanding normality. In this topology, a subbasis for the neighborhoods of a function f \in \mathcal{H}(\Omega) consists of sets \{g \in \mathcal{H}(\Omega) : \sup_{z \in K} |f(z) - g(z)| < \epsilon\}, where K \subset \Omega ranges over compact subsets and \epsilon > 0. A family \mathcal{F} \subset \mathcal{H}(\Omega) is normal if and only if it is relatively compact in this topology, meaning its closure is compact; this equivalence follows from the Arzelà-Ascoli theorem applied to restrictions on compact subsets. The concept of normal families originated with Paul Montel in his 1907 work on sequences of analytic functions, where he introduced ideas of for such families without initially using the term "normal." Montel's definition focused on the existence of convergent subsequences for families admitting certain representations, laying the groundwork for modern criteria; over the subsequent decades, the notion evolved to encompass broader classes of functions, including meromorphic ones, while retaining the core emphasis on sequential in the . For instance, the family of all holomorphic functions on the unit disk \mathbb{D} is not normal, as the sequence f_n(z) = n z diverges without any converging uniformly on compact subsets (e.g., on disks of radius $1/2). In contrast, families that are uniformly bounded on \mathbb{D} exhibit , highlighting how restrictive conditions can enforce the required .

Holomorphic Functions and Domains

A is a complex-valued function f: D \to \mathbb{C}, where D is an open subset of the \mathbb{C}, that is complex differentiable at every point in D. Complex differentiability at a point z_0 \in D means the limit \lim_{z \to z_0} \frac{f(z) - f(z_0)}{z - z_0} exists and is finite. satisfy key properties, such as the , which states that if f is holomorphic in a bounded D and continuous up to the boundary, then the maximum of |f(z)| on the of D is attained on the boundary. In , a is defined as a nonempty open connected subset of the \mathbb{C}. Examples include the unit disk \{z \in \mathbb{C} : |z| < 1\}, the punctured plane \mathbb{C} \setminus \{0\}, and the upper half-plane \{z \in \mathbb{C} : \operatorname{Im}(z) > 0\}. These provide the natural settings for studying holomorphic functions, as ensures path integrals and other global properties behave consistently. A sequence of holomorphic functions \{f_n\} on a domain D converges uniformly on compact subsets of D if, for every compact K \subset D, the convergence is uniform on K. Such uniform convergence on compact subsets preserves holomorphy: the limit function f = \lim_{n \to \infty} f_n is holomorphic on D. This result follows from the Arzelà-Ascoli theorem applied to holomorphic functions, which leverages their equicontinuity on compact sets due to Cauchy's estimates. Local uniform boundedness for a f on a D means that for every compact subset K \subset D, there exists a constant M_K > 0 such that |f(z)| \leq M_K for all z \in K. This condition is equivalent to f being bounded on every compact subset of D, ensuring controlled growth locally.

Statements of the Theorem

Boundedness Criterion

The boundedness criterion, a foundational result in the theory of normal families, asserts that a family of holomorphic functions on a domain is normal provided it is locally uniformly bounded. Specifically, let G \subset \mathbb{C} be a domain and let \mathcal{F} be a family of holomorphic functions on G. If for every compact subset K \subset G there exists M_K > 0 such that |f(z)| \leq M_K for all z \in K and all f \in \mathcal{F}, then \mathcal{F} is normal on G. This criterion was established by Paul Montel in his 1916 paper "Sur les familles normales de fonctions analytiques," refining his earlier investigations into sequences of analytic functions from 1907 onward. A direct is that local uniform boundedness is necessary for normality: if \mathcal{F} is not locally uniformly bounded, then \mathcal{F} is not normal. For instance, consider the family \{p_n(z) = z^n : n \in \mathbb{N}\} of holomorphic functions on the unit disk D = \{z \in \mathbb{C} : |z| < 1\}. This family is not locally bounded near any point on the boundary (approached radially), and sequences from it fail to have subsequences converging uniformly on compact subsets of D, confirming non-normality. Informally, the boundedness condition ensures that functions in \mathcal{F} cannot "escape to infinity" along any sequence of points in compact subsets of G, thereby guaranteeing equicontinuity (via Cauchy's estimates) and total boundedness in the compact-open topology. This setup enables the application of the Arzelà-Ascoli theorem to extract convergent subsequences, establishing normality.

Value Omission Criterion

The value omission criterion, also known as Montel's fundamental normality criterion, states that if F is a family of holomorphic functions on a domain G \subseteq \mathbb{C} such that each f \in F omits two distinct fixed values \alpha, \beta \in \mathbb{C} (i.e., f(G) \subseteq \mathbb{C} \setminus \{\alpha, \beta\}), then F is normal in the extended sense, meaning every sequence in F has a subsequence that converges uniformly on compact subsets of G to a holomorphic function or to the constant function \infty. This result was established by Paul Montel in 1907. The functions in such a family map the domain G into the twice-punctured complex plane \mathbb{C} \setminus \{\alpha, \beta\}, a hyperbolic Riemann surface whose restricted geometry ensures the normality of the family; in contrast, mapping into the full plane \mathbb{C} or the once-punctured plane \mathbb{C} \setminus \{\alpha\} does not guarantee normality. The universal covering space of the twice-punctured plane is the unit disk, and compositions with the covering map yield bounded lifts, which underpin the implicit local boundedness leading to normality. An illustrative example is the family of all holomorphic functions on the unit disk that map into the unit disk itself (e.g., by Schwarz lemma considerations or contractions), which omit any two fixed values outside the closed unit disk, such as 2 and 3, rendering the family normal by this criterion. The criterion is sharp in the sense that omitting only one value is insufficient to ensure normality. For instance, consider the family \{ f_n(z) = \exp(n z) \mid n = 1, 2, \dots \} of holomorphic functions on the unit disk \mathbb{D}; each f_n omits the value 0, but the family is not normal, as it fails to be locally bounded—for example, |f_n(1/2)| = \exp(n/2) \to \infty as n \to \infty.

Great Montel's Theorem

A more general version, known as the Great Montel's Theorem, applies to families of meromorphic functions: Let \mathcal{F} be a family of meromorphic functions on a domain \Omega \subset \mathbb{C}. If there exist three distinct points w_1, w_2, w_3 \in \hat{\mathbb{C}} such that f(\Omega) \cap \{w_1, w_2, w_3\} = \emptyset for all f \in \mathcal{F}, then \mathcal{F} is normal on \Omega. This extends the value omission criterion to meromorphic functions by using the spherical metric.

Proofs of the Criteria

Proof of Boundedness Implies Normality

To prove that a locally uniformly bounded family of holomorphic functions on a domain U \subset \mathbb{C} is normal, assume \mathcal{F} is such a family, meaning that for every compact subset K \subset U, there exists M_K > 0 such that |f(z)| \leq M_K for all f \in \mathcal{F} and z \in K. The proof proceeds by showing that the spherical derivatives of functions in \mathcal{F} are locally bounded, which implies in the and hence relative compactness in the via Marty's theorem. Fix a compact K \subset U. Since \mathcal{F} is locally uniformly bounded, there exists an open neighborhood V of K contained in U such that |f(z)| \leq M for all f \in \mathcal{F} and z \in V, for some M > 0. Choose r > 0 such that the open r-neighborhood of K is contained in V. For any z \in K and f \in \mathcal{F}, gives f'(z) = \frac{1}{2\pi i} \int_{|\zeta - z| = r} \frac{f(\zeta)}{(\zeta - z)^2} \, d\zeta, so |f'(z)| \leq \frac{1}{r} \max_{|\zeta - z| = r} |f(\zeta)| \leq \frac{M}{r}. The spherical derivative is defined as f^\#(z) = \frac{|f'(z)|}{1 + |f(z)|^2}. Since |f(z)| \leq M on K, it follows that $1 + |f(z)|^2 \geq 1, yielding f^\#(z) \leq |f'(z)| \leq \frac{M}{r} for all z \in K and f \in \mathcal{F}. Thus, the spherical derivatives are uniformly bounded on K. By Marty's theorem, a family of meromorphic functions (here, holomorphic as a special case) is normal on U if and only if the family of its spherical derivatives is locally bounded on U. The local boundedness established above implies that \mathcal{F} is equicontinuous on every compact subset of U with respect to the chordal metric on the Riemann sphere. To derive equicontinuity explicitly, consider points z_1, z_2 \in K with |z_1 - z_2| < r/2. By the mean value theorem, there exists \xi between z_1 and z_2 such that |f(z_1) - f(z_2)| = |f'(\xi)| \cdot |z_1 - z_2| \leq (M/r) |z_1 - z_2|. Since |f(z)| \leq M, the chordal distance \chi(f(z_1), f(z_2)) \leq 2 |f(z_1) - f(z_2)| / (1 + |f(z_1)|^2) \leq 2 (M/r) |z_1 - z_2|, providing a uniform modulus of continuity on K. This holds for any covering of K by disks of radius r, confirming equicontinuity. The uniform boundedness and equicontinuity on compact sets imply, by the Arzelà–Ascoli theorem, that every sequence in \mathcal{F} has a subsequence converging uniformly on compact subsets of U to a holomorphic limit function (by the Weierstrass theorem on differentiation under uniform limits). Thus, \mathcal{F} is normal.

Proof of Omission Implies Normality

To prove that a family \mathcal{F} of meromorphic functions on a domain U \subset \mathbb{C}, each omitting three distinct fixed values a, b, c \in \hat{\mathbb{C}}, is normal, the argument leverages the Zalcman rescaling lemma and Picard's great theorem. Without loss of generality, by a suitable (which preserves meromorphicity and normality), assume the omitted values are $0, 1, \infty. The thrice-punctured \hat{\mathbb{C}} \setminus \{0, 1, \infty\} is hyperbolic, but the proof proceeds via contradiction using rescaling. Suppose \mathcal{F} is not normal. Then implies the existence of a subsequence \{f_n\} \subset \mathcal{F}, points z_n \in U, radii \rho_n \to 0, such that the rescaled functions g_n(\zeta) = f_n(z_n + \rho_n \zeta) converge uniformly on compact subsets of \mathbb{C} to a meromorphic function g: \mathbb{C} \to \hat{\mathbb{C}} \setminus \{0, 1, \infty\}, with g nonconstant (since if constant, the original family would be normal by the boundedness criterion). Moreover, the spherical derivative satisfies g^\#(\zeta) \leq 1 for all \zeta \in \mathbb{C}. However, Picard's great theorem states that any nonconstant meromorphic function on \mathbb{C} omits at most two values in \hat{\mathbb{C}}. Thus, no such nonconstant g omitting three values exists, yielding a contradiction. Therefore, \mathcal{F} must be normal.

Necessity and Equivalence

Local Boundedness as Necessary Condition

To establish the necessity of local uniform boundedness in Montel's boundedness criterion, suppose a family \mathcal{F} of holomorphic functions on a domain \Omega \subset \mathbb{C} is normal but not locally uniformly bounded. Then there exists a compact set K \subset \Omega such that \mathcal{F} is not uniformly bounded on K, so there exists a sequence \{f_n\} \subset \mathcal{F} with \sup_{z \in K} |f_n(z)| \geq n for each n. By normality, there is a subsequence \{f_{n_k}\} converging uniformly on K to a holomorphic function f on \Omega. Since f is continuous on the compact set K, it is bounded: \sup_{z \in K} |f(z)| < \infty. Uniform convergence implies that for sufficiently large k, \sup_{z \in K} |f_{n_k}(z)| is also bounded, contradicting the choice of the sequence. Thus, normality implies local uniform boundedness. A concrete example illustrating non-normality due to lack of local boundedness is the family \{f_n(z) = n z : n \in \mathbb{N}\} on the unit disk \mathbb{D}. Here, |f_n(1)| = n \to \infty, so \mathcal{F} is not locally bounded on the compact set \{1/2 \leq |z| \leq 1\} \subset \mathbb{D}. Moreover, \mathcal{F} is not normal on \mathbb{D}, as any subsequence diverges to \infty uniformly on compact subsets avoiding 0 but converges to 0 near 0, precluding uniform convergence on the whole \mathbb{D}.

Equivalence in the Bounded Case

In the bounded case, Montel's theorem establishes a precise equivalence between and local uniform boundedness for families of holomorphic functions on a domain \Omega \subseteq \mathbb{C}. Specifically, a family \mathcal{F} of holomorphic functions on \Omega is if and only if it is locally uniformly bounded, meaning that for every compact subset K \subset \Omega, there exists a constant M_K > 0 such that |f(z)| \leq M_K for all z \in K and all f \in \mathcal{F}. This bidirectional formulation captures the full sharpness of the criterion, where local uniform boundedness ensures the existence of convergent subsequences uniformly on compact subsets, and conversely, normality implies such boundedness via properties of uniform limits of holomorphic functions. This equivalence was comprehensively developed in Paul Montel's 1927 treatise, where he synthesized earlier results on normal families and demonstrated the if-and-only-if relationship in the context of boundedness. Montel's work highlighted how this criterion provides a complete for families that remain "well-behaved" under in the space of holomorphic functions, influencing subsequent developments in . In contrast, the value omission criterion—stating that a family omitting two fixed values in \mathbb{C} is —is not equivalent to normality, as it represents a stricter sufficient condition rather than a necessary one. While omitting two values implies local uniform boundedness (and thus normality) through growth estimates on the functions, the converse fails: there exist normal families that do not omit any two fixed values collectively. For instance, consider the family \mathcal{F} = \{f_n(z) = z/n : n = 1, 2, \dots \} of holomorphic functions on the entire \mathbb{C}. This family is locally uniformly bounded, hence normal, with subsequences converging uniformly on compact sets to the zero function; however, it omits no values commonly, as f_1(z) = z attains every . Such examples underscore that boundedness provides the fundamental equivalence, while value omission serves as a specialized tool for ensuring normality without requiring explicit bounds.

Applications and Relations

Connections to Entire Function Theorems

Montel's theorem establishes deep connections to classical results in the theory of s by specializing its normality criteria to the case of functions holomorphic on the entire \mathbb{C}. The boundedness criterion, which states that a locally bounded of holomorphic functions on a domain is , directly generalizes when applied to entire functions. Specifically, a bounded of entire functions is normal, meaning every sequence has a converging uniformly on compact subsets of \mathbb{C} to a holomorphic limit function. Since the family is globally bounded, the limit is a bounded entire function and thus constant by . Consequently, all functions in the family must be constant, underscoring that non-constant entire functions cannot belong to a non-trivial bounded . The value omission criterion of Montel's theorem similarly links to Picard's little theorem. A family of holomorphic functions on a that collectively omit two fixed values in \mathbb{C} is . When restricted to entire functions, such a family admits a converging uniformly on compacta to an entire that also omits the two values. By Picard's little theorem, this must be , implying that the original family consists solely of functions. This interplay demonstrates how Montel's theorem bridges local normality on domains to global rigidity for entire functions, where omitting two values enforces . Bloch's principle offers a unifying for these connections, suggesting that a property compelling entire functions to be typically induces in families of holomorphic functions on more general domains. Boundedness exemplifies this: it yields for entire functions via and for families via Montel's boundedness criterion. Omitting two values aligns similarly, producing for entire functions by Picard's little theorem while ensuring for families under Montel's omission criterion. The principle highlights the global-to-local transition inherent in entire function theory, where properties like on a disk can imply broader structural constraints. A key illustration of the one-value omission limit arises with the family of scaled exponentials \{f_\lambda(z) = \lambda e^z : \lambda \in \mathbb{C}, |\lambda| \leq 1\}, consisting of entire functions that omit 0. This family is not , as the subsequence f_n(z) = e^{z - n} converges to 0 on \{z : \operatorname{Re} z < 0\} but diverges to \infty on \{z : \operatorname{Re} z > 0\}, precluding a uniformly convergent limit on \mathbb{C}. This example emphasizes that omitting a single value permits non-constant entire functions, such as the itself, and fails to guarantee , contrasting sharply with the two-value case.

Role in Picard's Theorem and Bloch's Principle

Montel's theorem plays a pivotal role in establishing Picard's theorems by leveraging the normality of families of holomorphic functions to derive restrictions on value omission for entire or meromorphic functions. In the proof of Picard's little theorem, which states that a non-constant entire function omits at most one complex value, the argument proceeds by contradiction: assume a non-constant entire function f omits two values, say 0 and 1 (after a suitable Möbius transformation). Consider the family of rescaled and translated functions \{f(2^k \phi_a) \mid \phi_a \in \Aut(\Delta), k = 1, 2, \dots \}, where \Delta is the unit disk and \Aut(\Delta) denotes its automorphisms; this family consists of holomorphic functions on \Delta that omit 0 and 1. By Montel's theorem, since the family omits two fixed values, it is normal on \Delta. Applying Marty's theorem to bound the spherical derivatives and extracting a subsequence converging uniformly on compact subsets leads to f'(z) = 0 everywhere, implying f is constant, a contradiction. Thus, Montel's normality criterion forces the value omission bound. For Picard's great theorem, which asserts that near an essential singularity at z = a, a holomorphic function takes every complex value, with at most one exception, infinitely often, Montel's theorem similarly ensures normality in punctured neighborhoods. Assume f has an essential singularity at 0 and omits three distinct values a_1, a_2, a_3; by the Casorati-Weierstrass theorem, there exists a sequence c_n \to 0 such that the spherical derivative f^\#(c_n) \to \infty. Define the rescaled family f_n(z) = f(c_n + c_n z), holomorphic on the unit disk with f_n^\#(0) \to \infty, implying non-normality by Marty's criterion. However, each f_n omits a_1, a_2, a_3, so Montel's theorem declares the family normal, yielding a contradiction unless the singularity is not essential. This application highlights how local normality via Montel controls global behavior near singularities. Bloch's principle, a linking local to global restrictions, is formalized through Montel's theorem to imply Picard-type behavior on the . Attributed to André Bloch, the principle posits that a family of s on a is if no non-constant entire shares the same restrictive property (e.g., omitting values). Zalcman's formalization defines a "Bloch property" P such that if a single function on the satisfies P, it is constant, and families satisfying P locally are ; Montel's value-omission criterion exemplifies this, as omitting three values forces constancy globally ('s little theorem) and locally. This duality ensures that Montel's local criterion extends to global Picard assertions, providing a unified framework for value distribution. Modern extensions of Montel's theorem to several complex variables preserve the normality criterion: a family of holomorphic maps from a domain in \mathbb{C}^m to \mathbb{P}^n(\mathbb{C}) omitting $2n+1 hyperplanes in is normal, generalizing the one-variable case (where n=1 corresponds to omitting 3 values in \hat{\mathbb{C}}) and facilitating analogs of Picard's theorems in higher dimensions. In non-archimedean fields, such as p-adic numbers, a version of Montel's theorem holds for analytic functions over complete valued fields, where families avoiding certain disks are , enabling studies of dynamics and value distribution in this setting; recent works further refine these for rational maps and periodic points.

References

  1. [1]
    [PDF] NORMAL FAMILIES AND PICARD'S GREAT THEOREM
    The beginning of a modern complex analysis dates back to 1907 when. P. Montel introduced concepts of compactness into complex analysis. [8]. In 1912, Montel ...
  2. [2]
    [PDF] Normal families and Montel's theorem
    The theorem below says that it takes very little for a sequence of analytic functions to be normal: the functions only have to be uniformly bounded on any ...Missing: Paul paper
  3. [3]
    [PDF] A Heuristic Principle in Complex Function Theory
    Perhaps the most celebrated criterion for normality is the following theorem, due to Paul Montel. MONTEL'S THEOREM. Let F be a family of functions meromorphic ...Missing: primary | Show results with:primary<|control11|><|separator|>
  4. [4]
    [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 ...
  5. [5]
    [PDF] Sur les familles normales de fonctions analytiques - Numdam
    N.S.. PAUL MONTEL. Sur les familles normales de fonctions analytiques. Annales scientifiques de l'É.N.S. 3e série, tome 33 (1916), p. 223-302. <http://www ...
  6. [6]
    [PDF] arXiv:math/0511048v1 [math.CV] 2 Nov 2005
    Nov 2, 2005 · We first note how. Zalcman's lemma can be used to deduce Montel's Theorem from Picard's Theo- rem. In fact, suppose that F is a non-normal ...
  7. [7]
    [PDF] Lecture Note for Math 220B Complex Analysis of One Variable
    Use the Open Mapping Theorem to prove the Maximum Modulus The- orem of holomorphic function: Let f be holomorphic in a domain D so that |f(z)|≤|f(z0)| for all z ...
  8. [8]
    2.2 Domains
    A domain is a set that is non-empty, open, and path connected. Examples include open balls, annuli, and half-planes.
  9. [9]
    [PDF] Complex Analysis I, Christopher Bishop 2024 - Stony Brook University
    Compactness for families of functions is often given by Arzela-Ascoli theorem. ... a subsequence which converges uniformly on compact subsets of Ω. • The ...
  10. [10]
    [PDF] 12. Normality and holomorphic functions - UCSD Math
    Remark 12.2. Note that to say that a function is uniformly bounded on compact subsets is equivalent to the statement that it is locally bounded.
  11. [11]
    [PDF] Normal families of holomorphic functions - Harold P. Boas
    Nov 10, 2003 · Montel's theorem says that a family F of holomorphic functions is normal if and only if it is locally bounded. Paul Montel 1876–1975 functions ...
  12. [12]
    Sur les familles normales de fonctions analytiques - EuDML
    Sur les familles normales de fonctions analytiques. Paul Montel · Annales scientifiques de l'École Normale Supérieure (1916). Volume: 33, page 223-302; ISSN ...
  13. [13]
    Normal Families - Joel L. Schiff - Google Books
    This is the first book devoted solely to the subject of normal families of analytic and meromorphic functions since the 1927 treatise of Paul Montel.
  14. [14]
    [PDF] Joel L. Schiff Normal Families
    Paul Montel (1876–1975). ' Photo reprinted with permission of Birkhäuser ... A book on the subject of normal families more than sixty years after the.
  15. [15]
    [PDF] Montel's theorem - Harold P. Boas
    Paul Montel, 1876–1975. He wrote Leçons sur les familles normales de fonctions analytiques et leurs applications, Paris,. 1927. Math 618. May 4, 2004 ...
  16. [16]
    [PDF] Class Notes Math 618: Complex Variables II Spring 2016
    The same example shows that the punctured plane ℂ ⧵ {0} is still too big for the range of the functions. Remarkably, the twice-punctured plane is small enough.
  17. [17]
  18. [18]
    [PDF] A non-archimedean Montel's theorem
    Trucco. Charles Favre. A non-archimedean Montel's theorem. Page 2. The complex case. Basics in non-archimedean analysis. Berkovich point of view. Ideas of proof.Missing: history | Show results with:history
  19. [19]
    [PDF] Introduction to Complex Analysis Michael Taylor
    Chapter 2 starts with two major theoretical results, the Cauchy integral theorem, and its corollary, the Cauchy integral formula. These theorems have a major ...
  20. [20]
    [PDF] Lemma 5.16 then tells us that this definition agrees with Definition ...
    Theorem 5.24 is a stronger result than both of them. Proof. Assume f omits at least two values in C. We will show that f must be a constant function on ˆB1 ...
  21. [21]
    [PDF] An Introduction to Holomorphic Dynamics
    Zalcman's rescaling lemma. Zalcman's lemma has revolutionized the study of normal families. It can not only be used to prove the equivalence of results for ...
  22. [22]
    [PDF] phic if there exists a holomorphic bijection - IISc Math
    Theorem 0.3 (Montel's theorem). A family F of holomorphic functions on Ω is normal if and only if it is locally uniformly bounded. To prove this, we first ...
  23. [23]
    [PDF] Theorem VII.1.2 (continued 1)
    Jul 8, 2017 · Theorem VII.2.1. If {f} is a sequence in H(G) and f belongs to. C(G, C) such that lim∞ f f, then f is analytic and the derivatives.
  24. [24]
    [PDF] math 566 lecture notes 6: normal families and the theorems of picard
    Dec 7, 2010 · A family F ⊂ O(Ω,M) is called normal if every sequence in F has a subsequence that converges locally uniformly on Ω. Theorem 9 (Marty). Assume ...
  25. [25]
    [PDF] The Picard Theorems via Geometry - Dartmouth Mathematics
    Aug 25, 2015 · Theorem 1.15 (The Little Picard Theorem). Let f be an entire function whose range omits at least two points. Then f is constant. Ok, time to ...
  26. [26]
    [PDF] Bloch's principle
    Abstract. A heuristic principle attributed to André Bloch says that a family of holomorphic functions is likely to be normal if there is no nonconstant.
  27. [27]
    [PDF] Radially distributed values and normal families - Purdue Math
    A major guideline in the theory of normal families is the heuristic Bloch principle which says that the family of all holomorphic functions having a certain ...
  28. [28]
  29. [29]
  30. [30]
    [1105.0746] A non-archimedean Montel's theorem - arXiv
    May 4, 2011 · Abstract:We prove a version of Montel's theorem for analytic functions over a non-archimedean complete valued field.
  31. [31]
    [PDF] Lectures on Complex Analysis - IISc Math
    Feb 1, 2025 · entire function is a function that is holomorphic on the entire complex plane C. ... A holomorphic definition would have primitive d dz log.<|control11|><|separator|>