Fact-checked by Grok 2 weeks ago

Maximum modulus principle

The maximum modulus principle is a central result in asserting that if a function f is holomorphic on a D \subseteq \mathbb{C} and there exists a point z_0 \in D such that |f(z_0)| \geq |f(z)| for all z \in D, then f must be constant throughout D. This principle implies that non-constant holomorphic functions cannot achieve a local maximum of their in the interior of the , as the |f| is subharmonic. A related boundary version of the theorem applies to bounded domains: if f is holomorphic in the interior of a bounded R and continuous up to the \partial R, then the maximum of |f(z)| on the closed region \overline{R} is attained on the \partial R, provided f is not . This version relies on the of the closure and the local form of the principle, ensuring that any interior maximum would force constancy by the identity for holomorphic functions. Extensions include conditions at for functions analytic outside a disk, where the modulus is bounded by its on the circle if the limit at satisfies the bound. The principle has profound implications, including proofs of the —showing every non-constant polynomial has a root by assuming otherwise and applying the principle to the entire —and Schwarz's lemma, which bounds holomorphic functions on the unit disk. It also underpins the minimum modulus principle for non-vanishing functions and the for functions, such as the real part of f, facilitating applications in asymptotics like for factorials.

Introduction

Historical Background

The maximum modulus principle traces its origins to the early , where the related for functions emerged in the context of . This version, which states that a non-constant cannot attain its maximum in the interior of a domain, was derived from the mean value property of functions. Contributions built on broader work in , including applications to and gravitation, laying essential groundwork for later developments in elliptic partial differential equations. The principle's adaptation to complex analysis occurred through the foundational efforts of in the 1820s. and the representation of analytic functions via , detailed in works such as his 1821 Cours d'analyse, laid the groundwork for the derivation of the maximum modulus principle for holomorphic functions. By the mid-19th century, the principle gained prominence through the rigorous frameworks established by and , who formalized its role in complex function theory. Riemann's 1851 dissertation introduced a geometric perspective on holomorphic functions via differential quotients, incorporating potential-theoretic ideas to emphasize conformality and boundary behavior. Weierstrass, in lectures from the 1860s onward, provided an arithmetic approach using and , solidifying the principle's role in classifying singularities and entire functions. Refinements, such as growth estimates for entire functions, were later advanced by around 1896 in his three-circle theorem, extending applications to . By the mid-20th century, the maximum modulus principle had become a cornerstone of education, prominently featured in Lars Ahlfors's influential 1953 textbook Complex Analysis, which integrated it with proofs via mean value properties and integral formulas.

Intuitive Motivation

The maximum modulus principle can be intuitively understood through an to steady-state in a bounded region. In such a physical setting, the distribution satisfies , describing a , and the hottest (or coldest) point must occur on the boundary rather than in the interior. This is because, in equilibrium with no internal heat sources or sinks, heat flows from higher to lower temperatures, preventing a local maximum inside where heat would need to flow inward against this gradient. Holomorphic functions in exhibit a similar behavior for their , as the real part of a logarithm of the function (where defined) is harmonic, linking the principle to this physical intuition. A simple example illustrates this for holomorphic functions. Consider f(z) = z on the open unit disk |z| < 1. Here, |f(z)| = |z| < 1 for all points inside the disk, while on the |z| = 1, |f(z)| = 1, achieving the supremum. This demonstrates how the grows toward the boundary without exceeding it interiorly, consistent with the principle's core idea. The underlying reason ties to subharmonicity: for a holomorphic function f and p > 0, |f(z)|^p is subharmonic on its of definition. Subharmonic functions, generalizing the mean value property of harmonic functions in a weak sense, cannot attain a local maximum in the interior of a bounded unless constant, ensuring maxima lie on the boundary. This underscores why the modulus of holomorphic functions behaves like a "smoothed" version of boundary values, avoiding interior peaks. The principle fails without holomorphy, highlighting the role of analyticity. For instance, the non-holomorphic f(z) = 1 - |z|^2 on the closed unit disk has |f(z)| = 1 - |z|^2, which attains its maximum value of 1 at the interior point z = 0 and decreases to 0 on the boundary. Such interior maxima are possible for general continuous functions, but forbidden for non-constant holomorphic ones due to their rigid structure.

Mathematical Foundations

Formal Statement

The maximum modulus principle is a fundamental result in concerning the behavior of the modulus of holomorphic functions on domains in the . Let \Omega \subset \mathbb{C} be a bounded , meaning \Omega is an open and connected subset of \mathbb{C} with finite area, and let \partial \Omega denote its and \overline{\Omega} = \Omega \cup \partial \Omega its closure. Suppose f: \Omega \to \mathbb{C} is holomorphic on \Omega (i.e., complex differentiable at every point in \Omega) and continuous on \overline{\Omega}. Then, \max_{z \in \overline{\Omega}} |f(z)| = \max_{z \in \partial \Omega} |f(z)|, where |f(z)| denotes the modulus of the complex number f(z). Moreover, if equality holds at some point in the interior \Omega (i.e., |f(z_0)| = \max_{z \in \partial \Omega} |f(z)| for some z_0 \in \Omega), then f is constant on \overline{\Omega}. A stricter interior version of the principle states that if f is holomorphic on a domain \Omega and |f| attains a local maximum at any interior point z_0 \in \Omega, then f must be constant on \Omega. This version does not require boundedness or boundary continuity but applies directly to the open set. The minimum modulus principle states that if f is a holomorphic function in a bounded domain \Omega \subset \mathbb{C}, continuous up to the boundary \partial \Omega, and f \not\equiv 0 with no zeros in \Omega, then the minimum of |f(z)| on the closure \overline{\Omega} is attained on the boundary \partial \Omega. This principle follows as a counterpart to the maximum modulus principle applied to $1/f, ensuring that nonconstant holomorphic functions without interior zeros cannot achieve their smallest modulus inside the domain. A closely related result is that if f is holomorphic and nonconstant in a domain \Omega, then |f(z)| cannot attain a local maximum at any interior point of \Omega. This interior maximum principle for the modulus underscores the boundary behavior of holomorphic functions and aligns with the broader maximum modulus theorem by prohibiting interior extrema for |f| unless f is constant. The serves as a significant of the maximum modulus principle, applicable to functions on the unit disk. Specifically, if f is holomorphic in the open unit disk \mathbb{D} = \{ z \in \mathbb{C} : |z| < 1 \}, satisfies |f(z)| \leq 1 for all z \in \mathbb{D}, and f(0) = 0, then |f(z)| \leq |z| for all z \in \mathbb{D}, with |f'(0)| \leq 1. Equality holds in |f(z)| \leq |z| for some z \neq 0 or in |f'(0)| = 1 if and only if f(z) = c z for a constant c with |c| = 1. This lemma provides sharp bounds on the growth of normalized holomorphic functions and is foundational for studying automorphisms of the disk. The Phragmén–Lindelöf principle extends the maximum modulus principle to certain unbounded domains, such as sectors or strips, where standard boundedness assumptions fail. For a simply connected unbounded region G and holomorphic f: G \to \mathbb{C}, if there exists a non-vanishing bounded holomorphic function w: G \to \mathbb{C} with |w(z)| \leq 1, and boundary conditions ensure \limsup_{z \to a} |f(z)| \leq M on one part of the boundary at infinity and \limsup_{z \to b} |w(z)|^\epsilon |f(z)| \leq M on the other for every \epsilon > 0, then |f(z)| \leq M throughout G. This principle controls growth in unbounded settings by incorporating auxiliary functions to mimic bounded domain behavior.

Proof Techniques

Harmonic Function Approach

One approach to proving the maximum modulus principle utilizes properties of harmonic functions, specifically by considering the function u(z) = \log |f(z)| for a holomorphic function f on a bounded domain \Omega \subset \mathbb{C}, assuming f has no zeros in \Omega. Since f is holomorphic and non-vanishing, a branch of \log f(z) can be defined locally, and u(z) = \Re(\log f(z)), the real part of a holomorphic function, is therefore harmonic on \Omega. Equivalently, the Cauchy-Riemann equations for f imply that the Laplacian vanishes: \Delta u = \Delta (\log |f|) = 0 in \Omega, confirming the harmonicity of u. Harmonic functions satisfy the maximum principle: if u is harmonic and continuous on the closure of \Omega (with \Omega bounded and connected), then the maximum value of u on \overline{\Omega} is attained on the boundary \partial \Omega, unless u is constant. Applying this to u = \log |f|, the maximum of |f| on \overline{\Omega} occurs on \partial \Omega, as |f| = e^u. Suppose, for contradiction, that |f| attains its maximum at an interior point z_0 \in \Omega; then u(z_0) is a maximum for the harmonic function u, implying u is constant on \Omega. Consequently, |f(z)| is constant, and since f is holomorphic, f itself must be constant on \Omega. Note that u is actually subharmonic in general (with \Delta u \geq 0), but strict interior maxima still force constancy under these conditions. To extend the result to holomorphic functions with zeros, observe that zeros of f are isolated by the identity theorem. If the maximum of |f| is attained at an interior zero z_0, then |f(z_0)| = 0, so |f(z)| \leq 0 everywhere in \Omega, implying f \equiv 0 (a ). Otherwise, if the maximum point z_0 is not a zero, a small disk around z_0 avoids other zeros, allowing the zero-free argument to apply locally and yield constancy on \Omega. For the unit disk, zeros can alternatively be factored out using a finite B(z), which is holomorphic with |B(z)| = 1 on the ; then |f(z)/B(z)| inherits the boundary maximum from |f|, and the principle applies to the zero-free quotient.

Mean Value Theorem Approach

Gauss's mean value theorem provides a foundational tool for proving the maximum modulus principle by leveraging the averaging property of holomorphic over circles. For a f holomorphic at a point z_0 in a , the theorem states that the value at the center equals the average over any sufficiently small circle centered there: f(z_0) = \frac{1}{2\pi} \int_0^{2\pi} f(z_0 + r e^{i\theta}) \, d\theta, where r > 0 is small enough that the disk |z - z_0| < r lies within the domain of holomorphicity. This property arises as a direct consequence of Cauchy's integral formula applied to the circle. To establish the maximum modulus principle using this theorem, consider the modulus |f|. Suppose |f(z_0)| attains a local maximum at an interior point z_0, meaning |f(z)| \leq |f(z_0)| for all z in some neighborhood of z_0, with strict inequality for z \neq z_0. Taking the modulus of the mean value equation yields |f(z_0)| = \left| \frac{1}{2\pi} \int_0^{2\pi} f(z_0 + r e^{i\theta}) \, d\theta \right| \leq \frac{1}{2\pi} \int_0^{2\pi} |f(z_0 + r e^{i\theta})| \, d\theta, by the triangle inequality. The right-hand side represents the average of |f| over the circle of radius r around z_0. The modulus |f| is subharmonic in regions where f is holomorphic and non-zero, implying that its value at the center is at most the average over any circle: |f(z_0)| \leq \frac{1}{2\pi} \int_0^{2\pi} |f(z_0 + r e^{i\theta})| \, d\theta. Equality holds if and only if |f| is constant on the circle. If |f(z_0)| is a strict local maximum, then |f(z_0 + r e^{i\theta})| < |f(z_0)| for all \theta, so the average would be strictly less than |f(z_0)|, contradicting the mean value inequality. This contradiction implies that no such strict local maximum can exist unless f is constant. If equality holds throughout a neighborhood, then f must be constant there by the identity theorem for holomorphic functions. Thus, for non-constant holomorphic f, the modulus |f| attains its maximum only on the boundary of the domain, establishing the maximum modulus principle via this averaging approach.

Cauchy's Integral Formula Approach

One approach to proving the maximum modulus principle relies on Cauchy's integral formula, which expresses the value of a holomorphic function at an interior point in terms of its values on the boundary of a disk. Suppose f is holomorphic in a bounded domain D \subset \mathbb{C} and continuous up to the boundary \partial D, with |f(z)| \leq M for all z \in \overline{D}. Assume, for contradiction, that |f(z_0)| = M for some z_0 in the interior of D. Since D is open, there exists r > 0 such that the closed disk \overline{\Delta(z_0, r)} \subset D. By applied to the circle |\zeta - z_0| = r, f(z) = \frac{1}{2\pi i} \oint_{|\zeta - z_0| = r} \frac{f(\zeta)}{\zeta - z} \, d\zeta for all z with |z - z_0| < r. Taking absolute values and applying the triangle inequality yields |f(z)| \leq \frac{1}{2\pi} \int_0^{2\pi} \frac{|f(z_0 + r e^{i\theta})|}{| (z_0 + r e^{i\theta}) - z |} r \, d\theta \leq \frac{M r}{2\pi} \int_0^{2\pi} \frac{1}{|r e^{i\theta} - (z - z_0)|} \, d\theta. Since |z - z_0| = \rho < r, the denominator satisfies |r e^{i\theta} - (z - z_0)| \geq r - \rho for all \theta, so |f(z)| \leq M \cdot \frac{r}{r - \rho}. This estimate shows that |f(z)| is bounded above by a quantity exceeding M inside the disk, but it does not immediately contradict the assumption unless equality conditions are analyzed. To obtain a contradiction, evaluate at the assumed maximum point z = z_0, where the formula simplifies via the mean value property: f(z_0) = \frac{1}{2\pi} \int_0^{2\pi} f(z_0 + r e^{i\theta}) \, d\theta. Applying the triangle inequality gives |f(z_0)| \leq \frac{1}{2\pi} \int_0^{2\pi} |f(z_0 + r e^{i\theta})| \, d\theta \leq M. Since |f(z_0)| = M, equality must hold throughout: first, |f(z_0 + r e^{i\theta})| = M for almost all \theta (by continuity, for all \theta), so |f| = M constantly on the circle |\zeta - z_0| = r; second, equality in the triangle inequality requires that f(z_0 + r e^{i\theta}) has constant argument (up to a fixed phase) along the circle. This implies that f is constant on the circle |\zeta - z_0| = r, but since f is holomorphic inside and continuous up to the circle, the identity theorem extends this to f being constant in \Delta(z_0, r). Repeating this argument on smaller disks around points in the original disk shows f is constant on \Delta(z_0, r). By connectedness of D and the identity theorem for holomorphic functions, f is constant throughout D. Thus, unless f is constant, no interior maximum for |f| can exist, and the maximum must occur on \partial D. For the strict inequality inside, if f is non-constant and |f| = M on \partial \Delta(z_0, r), the equality case in the general bound requires analogous phase alignment, which again forces constancy; otherwise, |f(z)| < M \cdot \frac{r}{r - \rho}, but combined with the boundary maximum, yields |f(z)| < M.

Applications

In Complex Analysis

In complex analysis, the maximum modulus principle underpins several foundational theorems regarding the global properties of holomorphic functions, particularly those defined on the entire complex plane. A primary application is Liouville's theorem, which asserts that any bounded entire function is constant. Specifically, if f: \mathbb{C} \to \mathbb{C} is holomorphic and satisfies |f(z)| \leq M for some constant M > 0 and all z \in \mathbb{C}, then the maximum modulus principle applied to the closed disk |z| \leq R for arbitrary R > 0 shows that the supremum of |f| on this disk equals M and is attained on the boundary |z| = R. As R \to \infty, this boundary behavior, combined with the boundedness, forces |f| to be constant throughout \mathbb{C}, implying f is constant. The principle also facilitates Picard's little theorem, which states that a non-constant omits at most one value in the . This result arises from growth controls imposed by the maximum modulus principle on entire functions; if a non-constant entire f omitted two distinct values a and b, a transformation mapping \mathbb{C} \setminus \{a, b\} to the unit disk would yield a bounded entire function (by , constant), leading to a . For instance, the e^z omits only 0, illustrating the theorem's bound. Another key consequence is the open mapping theorem, affirming that a non-constant maps open sets to open sets. The maximum modulus principle contributes by implying that non-constant s have no local maxima in their within the , ensuring that the cannot have isolated boundary points and thus must be open.) The principle further elucidates the growth behavior of , providing precise asymptotic control at . For a p(z) = a_n z^n + a_{n-1} z^{n-1} + \cdots + a_0 of degree n \geq 1 with a_n \neq 0, the maximum modulus principle implies that |p(z)| grows without bound as |z| \to \infty, specifically satisfying |p(z)| \sim |a_n| |z|^n for large |z|. This follows from applying the principle to large disks, where the leading term dominates on the boundary, bounding lower-order contributions and establishing the precise rate of growth.

In Boundary Value Problems

The maximum modulus principle plays a crucial role in establishing for solutions to the in boundary value problems for functions. Consider a bounded domain \Omega \subset \mathbb{R}^n with smooth \partial \Omega, and let u be a on \Omega that is continuous up to the boundary and attains prescribed continuous boundary values \phi on \partial \Omega. If v is another such with the same boundary values, then w = u - v is on \Omega, continuous up to the boundary, and vanishes on \partial \Omega. By the principles for functions—analogous to the modulus principle in the complex case—w attains its on \partial \Omega, where it is zero, implying w \equiv 0 on \Omega. Thus, u = v, proving of the . In the context of the Laplace equation \Delta u = 0, solutions in simply connected planar domains can be represented as the real part of a f, i.e., u = \operatorname{Re}(f), where f is analytic on the domain. The maximum modulus principle applied to |f| then bounds |u| \leq |f|, providing estimates for approximate s in problems. For instance, if data is approximated, the principle ensures that deviations in the holomorphic extension propagate in a controlled manner, limiting interior s to those on the . This representation facilitates solving Dirichlet problems by extending conditions holomorphically while leveraging modulus bounds to verify accuracy. A concrete example arises in the unit disk D = \{ z \in \mathbb{C} : |z| < 1 \}, where the seeks a u on D matching continuous data \phi on the unit \partial D. The solution, given explicitly via the Poisson integral, satisfies the : the values of u in D lie strictly between the minimum and maximum of \phi unless \phi is constant, ensuring the solution remains within the range of the data without interior extrema. This property not only confirms the solution's boundedness but also underscores the principle's utility in verifying that perturbations in conditions do not cause unbounded interior growth. For unbounded domains, such as infinite strips, the standard maximum modulus principle does not directly apply due to lack of compactness, but the Phragmén–Lindelöf principle provides a modified version that controls growth. In a strip S = \{ z \in \mathbb{C} : a < \operatorname{Im}(z) < b \}, if f is holomorphic and bounded on the boundaries with |f| \leq 1 there, and satisfies a growth condition like |f(z)| \leq \exp(\exp(c |\operatorname{Re}(z)|)) for some c < \pi/(b-a), then |f(z)| \leq 1 throughout S. This extension ensures growth control in boundary value problems on unbounded regions, preventing exponential blow-up while adapting the modulus principle to strip geometry.

Physical and Interpretive Aspects

Potential Theory Interpretation

In classical potential theory, harmonic functions serve as models for potentials in charge-free regions, satisfying Laplace's equation \Delta u = 0. These functions describe electrostatic potentials, where the electric field is the negative gradient of the potential, and gravitational potentials outside mass distributions. The maximum principle for such harmonic functions asserts that on a bounded domain, a non-constant harmonic function attains its maximum and minimum values on the boundary, precluding interior local extrema. This property implies that electrostatic fields exhibit no interior equilibrium maxima for the potential, ensuring that field lines converge toward charges without spurious stable points inside the domain unless the potential is constant. In the context of complex potentials, where a f represents the complex potential \Phi + i\Psi, the modulus |f| relates to via \log |f| being subharmonic on the . satisfy a analogous to that for harmonics: a non-constant subharmonic function on a bounded achieves its maximum on the . Consequently, |f| models field strengths, such as magnitude, that are bounded above by values, preventing interior peaks in the absence of sources. A representative example arises in gravitational potential theory: in a bounded region devoid of masses, the gravitational potential \Phi, being harmonic, attains its extreme values—maximum (least negative) and minimum (most negative)—on the boundary, reflecting the influence of external mass distributions without interior singularities or equilibria. The maximum principle further connects to Green's functions in potential theory, where representation formulas express solutions to Dirichlet boundary value problems as integrals involving the Green's function G(x,y). This principle provides bounds on these solutions, estimating their magnitudes relative to boundary data and ensuring uniqueness in charge-free regions.

Fluid Dynamics and Steady States

In the context of steady-state heat conduction, the temperature distribution u in a region without internal heat sources satisfies \nabla^2 u = 0, making u a . The for functions implies that the values of u occur on the of the domain, preventing the formation of interior hotspots or cold spots in equilibrium. This property ensures thermal stability, as any interior extremum would require generation or , which is absent in the . For irrotational incompressible fluid flows, the velocity field can be derived from a \phi that satisfies \nabla^2 \phi = 0, rendering \phi throughout the flow domain. The fluid speed is given by |\mathbf{v}| = |\nabla \phi|, and since |\nabla \phi|^2 is subharmonic when \phi is , the applies to |\nabla \phi|^2, implying that the maximum fluid speed occurs on the domain boundary rather than in the interior. This subharmonicity follows from the \nabla^2 (|\nabla \phi|^2) = 2 \sum_{i,j=1}^n \left( \frac{\partial^2 \phi}{\partial x_i \partial x_j} \right)^2 \geq 0, ensuring no interior velocity maxima in inviscid potential flows. A representative example is the steady irrotational flow of an incompressible fluid around a fixed obstacle, such as a circular cylinder or sphere. In the exterior domain, the potential \phi is harmonic, and boundary conditions enforce no normal flow on the obstacle surface. The maximum principle guarantees that the highest speeds arise on the obstacle boundary, as seen in uniform flow past a cylinder where the speed reaches $2U_\infty (twice the far-field speed U_\infty) at the equatorial points, with speeds decreasing toward the far field. This prevents unphysical interior accelerations and aligns with observed flow patterns. The maximum modulus principle ties into Bernoulli's equation for steady irrotational flows, where \frac{p}{\rho} + \frac{1}{2} |\mathbf{v}|^2 + gz = \text{[constant](/page/Constant)} holds everywhere in the domain due to the absence of . With maximum speeds confined to boundaries, minimum pressures also occur there, bounded by surface conditions like no-slip approximations or prescribed pressures, ensuring the flow remains physically consistent without unbounded interior variations.

References

  1. [1]
    Maximum Modulus Principle -- from Wolfram MathWorld
    If an analytic function f has a point z_0 where |f(z_0)| >= |f(z)| for all z in U, then f is constant. Also, if |f| has a local max at z_0, then f is constant.
  2. [2]
    [PDF] Section 4.54. Maximum Modulus Principle
    Apr 26, 2020 · The Maximum Modulus Theorem states that if f is analytic and not constant in a domain, then |f(z)| has no maximum value in that domain.
  3. [3]
    [PDF] m2pm3 handout: the maximum modulus theorem
    The Maximum Modulus Theorem states that if f is holomorphic in N(a, R) and |f(z)|≤|f(a)| for all z in N(a, R), then f is constant.
  4. [4]
  5. [5]
    [PDF] Theory of Complex Functions
    Theory of complex functions / Reinhold Remmert ; translated by. Robert B. Burckel. p. cm. - (Graduate texts in mathematics ; 122. Readings in mathematics).
  6. [6]
  7. [7]
    [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-.
  8. [8]
    [PDF] Complex Varables Lecture Notes for Math 122A
    Jul 27, 2011 · We would like to determine the steady-state temperature within the region when the temperature is given on the boundary. Let u(x, y) be the ...
  9. [9]
    [PDF] Subharmonic functions
    Nov 2, 2020 · Suppose f (z) is a holomorphic function on Ω and B(z,r) is a disk ... |f (z)|p = exp p log |f (z)| is subharmonic for any 0 < p < ∞. Oleg ...
  10. [10]
    Continuous, Non-analytic function with maximum on the interior of a ...
    Jan 17, 2021 · I'm looking for a non-analytic function where the maximum is not on the boundary, but the function is continuous. −|z| on the closed unit disk. ...Demonstrating that a non-constant holomorphic function attains its ...Prove the open mapping theorem by using maximum modulus ...More results from math.stackexchange.com
  11. [11]
    [PDF] 15 - The Maximum Modulus Principle and the Mean Value Property
    Oct 7, 2025 · Theorem 2 (The maximum modulus principle). Let 𝑈 ⊂ ℂ be a domain and 𝑓 ∶ 𝑈 → ℂ be holomorphic/analytic. If |𝑓| has a local maximum on ...
  12. [12]
    None
    ### Formal Statement of the Maximum Modulus Principle
  13. [13]
    [PDF] 1 Lecture 1 - Penn Math
    Theorem 6.2 (Maximum Modulus Principle). If f is holomorphic in Ω a con- nected open set and |f(z)| has an interior local maximum, then f is constant on. Ω.<|control11|><|separator|>
  14. [14]
    [PDF] 2.10 Phragmén-Lindelöf principle - HKUST Math Department
    Phragmén (1863-1937) observed that this example is the best possible in 1904. Theorem 2.10.2 (Phragmén). Let G = ( z : |arg z| <.Missing: source | Show results with:source
  15. [15]
    [PDF] Lecture 28: The maximum modulus theorem - UW Math Department
    Maximum principle for harmonic functions. Assume u is harmonic on E, and continuous on E, where E is a bounded, connected, open set. Then the maximum of u on E.
  16. [16]
    [PDF] The Maximum Principle - Trinity University
    The mean value property has the following important consequence. Theorem 2 (Strict Maximum Principle for Harmonic Functions) ... Gauss proved the Fundamental ...
  17. [17]
    [PDF] 18.04 S18 Topic 4: Cauchy's integral formula - MIT OpenCourseWare
    In order to prove the maximum modulus principle we will first prove the mean value property. This will give you a good feel for the maximum modulus principle.Missing: authoritative sources
  18. [18]
  19. [19]
    proof of maximal modulus principle - PlanetMath
    Mar 22, 2013 · By definition there will exist r>0 r > 0 such that the set Sr={z∈C:|z−z0|2≤r2}⊂K S r = { z ∈ ℂ : | z - z 0 | 2 ≤ r 2 } ⊂ K .
  20. [20]
    [PDF] Introduction to Complex Analysis Michael Taylor
    ... Cauchy integral theorem and consequences. 5. The Cauchy integral theorem and the Cauchy integral formula. 6. The maximum principle ... Hadamard's factorization ...
  21. [21]
    [PDF] XII.2. The Little Picard Theorem.
    Dec 15, 2017 · Note. Recall that Theorem XI.3.6 states that an entire function of finite order assumes every complex number with one possible exception.
  22. [22]
    [PDF] Lecture Note for Math 220B Complex Analysis of One Variable
    By an induction argument, f(z) = f(zn) = f(z0). Corollary 3.2 (Minimum Modulus Theorem) Let D be a domain in C and f holomorphic in D with f(z) 6= ...Missing: source | Show results with:source
  23. [23]
    11.5 Maximum principle and the fundamental theorem of algebra
    In this section we study the local behavior of polynomials, and analytic functions in general, and the growth of polynomials as goes to infinity.
  24. [24]
    [PDF] Harmonic functions - Purdue Math
    Mar 18, 2024 · By the Maxi- mum/Minimum Principle, they must be attained on ∂D, but if the functions are equal on ∂D, these maximum and minimum are both 0. ...
  25. [25]
    [PDF] Chapter 2: Laplace's equation - UC Davis Math
    The maximum principle gives a uniqueness result for the Dirichlet problem for the Poisson equation. Theorem 2.18. Suppose that Ω is a bounded, connected ...<|control11|><|separator|>
  26. [26]
    [PDF] Harmonic Function Theory - Sheldon Axler
    lowing maximum principle for harmonic functions. 1.8. Maximum Principle: Suppose Ω is connected, u is real valued and harmonic on Ω, and u has a maximum or a ...<|control11|><|separator|>
  27. [27]
    [PDF] Something about Poisson and Dirichlet - Purdue Math
    The Dirichlet problem on the unit disc is: given a continuous real valued function ϕ on the unit circle, find a real valued function u that is continuous on ...Missing: disk | Show results with:disk
  28. [28]
    [PDF] 19. Dirichlet problem and the Poisson kernel - UCSD Math
    We first show that if h is identically zero then u is identically zero. By the maximum principle applied to u we have that the maximum of u on U ∪ ∂U is ...
  29. [29]
    [PDF] 12a. Phragmén-Lindelöf Theorems
    Feb 7, 2021 · The maximum modulus principle [1] does not apply to unbounded regions. That is, holomorphic functions on an unbounded region may be bounded by 1 ...
  30. [30]
    [PDF] Chapter 2: Laplace's equation - UC Davis Mathematics
    The maximum principle states that a non-constant harmonic function cannot attain a maximum (or minimum) at an interior point of its domain. This result implies ...
  31. [31]
    [PDF] Lesson 35. Potential theory, Electrostatic fields - Purdue Math
    Potential theory is the theory of harmonic functions, that is, solutions to Laplace's equation ∇2 Φ = 0. In applications, electrostatic and gravitational ...<|control11|><|separator|>
  32. [32]
    [PDF] 1 Harmonic Functions
    Apr 6, 2012 · Applying now the Maximum Principle to the function −h we conclude that h = 0 on D, hence h1 = h2 there. By evoking the Identity Principle (Thm 6) ...
  33. [33]
    [PDF] Potential Theory in the Complex Plane
    Hence, for every holomorphic function f : Ω → C, the Laplacian of f vanishes on the whole of Ω and thus Re(f) and Im(f) are both harmonic functions on. Ω.
  34. [34]
    [PDF] 5 Potential Theory
    Notice that Green's function gives us a solution to the Interior Dirichlet Problem which is similar to a double layer potential. We will see that for an ...
  35. [35]
    [PDF] Chapter 1 Maximum principle and the symmetry of solutions of ...
    Apr 1, 2017 · The maximum principle, for equations like ∆u + F(x, u)=0, states that if a function attains a maximum inside a domain, it must be a constant. ...
  36. [36]
    [PDF] 3 IRROTATIONAL FLOWS, aka POTENTIAL FLOWS - DAMTP
    ∇2φ = 0 ,. i.e., φ is a harmonic function, in the sense of satisfying Laplace's equation. The boundary condition for impermeable boundaries, also called the ...
  37. [37]
    [PDF] HARMONIC FUNCTIONS Contents 1. Introduction 1 2. Green's ...
    Sep 21, 2018 · If we think of ϕ as the electrostatic potential, this leads to the integral form of the Gauss law expressed in terms of ϕ. Similarly, letting u ...
  38. [38]
    Potential Flow Theory – Introduction to Aerospace Flight Vehicles
    ... flow field remains invariant about the {x} -axis. The maximum velocity occurs at the sphere's equator ( \theta = \pm \pi/2 ), is 1.5 {V_{\infty}} , with a ...
  39. [39]
    Bernoulli's Equation In Irrotational Flow - NPTEL Archive
    Therefore, the total mechanical energy remains constant everywhere in an inviscid and irrotational flow, while it is constant only along a streamline for an ...