Fact-checked by Grok 2 weeks ago

Maximum principle

The maximum principle is a fundamental in the of partial equations (PDEs), particularly for elliptic equations, which states that if a function u satisfies \Delta u \leq 0 (subharmonic) in a bounded \Omega, then the maximum value of u is attained on the \partial \Omega, unless u is constant throughout \Omega. This principle extends to more general elliptic operators of the form Lu = a^{ij}(x) \partial_i \partial_j u + b^i(x) \partial_i u + c(x) u \leq 0 with c(x) \leq 0, ensuring that non-constant subsolutions cannot achieve an interior maximum. A strong maximum principle refines this result: if u attains its maximum at an interior point x_0 \in \Omega and satisfies Lu \leq 0 with the conditions above, then u must be constant in \Omega. This version relies on Hopf's lemma, which asserts that at a maximum point where u(x_0) > u(x) for all x \in \Omega, the outward normal derivative satisfies \frac{\partial u}{\partial \nu}(x_0) > 0, preventing "flat" maxima and enabling proofs of for Dirichlet boundary value problems. The principle also implies comparison results: if Lu \leq 0 and Mv \leq 0 with u \leq v on \partial \Omega, then u \leq v in \Omega. Historically, the maximum principle traces its origins to properties of harmonic functions in and the Laplace equation, with key developments in the , including Hopf's contributions in the that generalized it to nonlinear elliptic PDEs. For parabolic PDEs, such as the \partial_t u - \Delta u \leq 0, the maximum principle adapts to space-time domains, stating that the maximum of u occurs on the parabolic boundary (initial time or spatial boundary), again unless u is constant. This extension applies to evolution equations like \partial_t u \leq \Delta u + \langle X, \nabla u \rangle + F(u), bounding solutions by comparison with ODEs solving \frac{d\phi}{dt} = F(\phi). The principle's robustness provides a priori bounds without explicit solutions, aiding regularity theory and results, such as those for positive solutions via moving planes methods. In applications, it underpins in boundary value problems, Harnack inequalities for controlling oscillations (e.g., \sup_U u \leq C \inf_U u for positive functions), and , including where it controls curvature evolution.

Overview and Intuition

Intuitive Explanation

The maximum principle is a fundamental property in the analysis of solutions to elliptic partial differential equations (PDEs), which model diffusion-like phenomena such as heat conduction or electrostatic potentials. At its core, the principle asserts that a non-constant solution to such an equation cannot attain its maximum value at any interior point of the domain; instead, the maximum must occur on the . This reflects the smoothing or averaging behavior inherent in these equations, preventing "peaks" or local extremes from forming inside the region without the function being constant throughout. A key intuition arises from the mean value property satisfied by solutions to , the prototypical elliptic PDE, where the value of the solution at any interior point equals the average of its values over a surrounding or . If a maximum were achieved inside, the surrounding values would all need to be less than or equal to it, making the average strictly less unless the is in that —leading to constancy everywhere by connectedness. This averaging effect ensures that interior points "inherit" their values from the boundary, much like how a calm water surface in a container takes its shape from the edges without isolated highs or lows in the middle. In one dimension, consider the simple case of solutions to -u'' = 0 on an (a, b), which are linear s u(x) = cx + d. Such a can only achieve its maximum at one unless c = 0, in which case it is constant; there are no interior maxima for non-constant solutions. This extends the intuition to higher dimensions, where non-linear but solutions behave similarly, rising or falling toward the . Physically, the principle manifests in steady-state heat distribution: in a heated object like a metal plate with fixed temperatures, the interior temperature equilibrates to a value between the hottest and coldest points, never exceeding them inside, as flows to equalize differences without creating isolated hot spots. This boundary-driven behavior underscores 's role in ensuring stable, equilibrium solutions in diffusive systems.

Key Examples and Counterexamples

A classic example illustrating the applicability of the maximum principle is the harmonic function u(x, y) = x defined on the unit disk \Omega = \{(x, y) \in \mathbb{R}^2 : x^2 + y^2 < 1\}. Since \Delta u = 0, u is harmonic, and its maximum value of 1 is attained at the boundary point (1, 0), with no interior point exceeding this value. A related partial formulation of the maximum principle applies to subharmonic functions, where if u \in C^2(\Omega) satisfies \Delta u \geq 0 in a bounded domain \Omega \subset \mathbb{R}^n, then \max_{\overline{\Omega}} u = \max_{\partial \Omega} u. This extends the harmonic case (\Delta u = 0) and underscores the role of the Laplacian's sign in ensuring boundary maxima. In contrast, for non-elliptic partial differential equations such as the heat equation u_t - \Delta u = 0, interior maxima can occur due to time evolution, particularly on the initial time slice, distinguishing it from the elliptic setting where no such interior extrema are possible unless the solution is constant. While a space-time maximum principle holds—with the global maximum attained on the parabolic boundary (initial or lateral boundaries)—the time-dependent nature allows non-constant solutions to start with interior maxima that influence the evolution. The maximum principle fails to hold in its elliptic form for hyperbolic partial differential equations, such as the wave equation u_{tt} - \Delta u = 0. For instance, solutions can develop interior extrema through wave propagation, even if initial and boundary data are zero; this violates the boundary maximum condition, as energy conservation permits oscillations creating local maxima inside the domain. The strong maximum principle provides a sharper result, implying constancy for non-constant solutions attaining interior maxima, though its details are addressed elsewhere.

Mathematical Foundations

Elliptic Partial Differential Equations

The maximum principle is a fundamental result in the theory of partial differential equations (PDEs), particularly for elliptic equations, asserting that solutions to certain PDEs attain their maximum and minimum values on the boundary of the domain rather than in the interior. This principle has profound implications for understanding the behavior of solutions and is essential in applications ranging from physics to geometry. Originating from classical analysis, it provides bounds and regularity insights without solving the PDE explicitly.

Elliptic Partial Differential Equations

Elliptic partial differential equations form a class of second-order linear PDEs that model equilibrium or steady-state problems in various physical contexts. The general form of a second-order linear elliptic PDE in n dimensions is given by \sum_{i,j=1}^n a_{ij}(x) \frac{\partial^2 u}{\partial x_i \partial x_j} + \sum_{i=1}^n b_i(x) \frac{\partial u}{\partial x_i} + c(x) u = f(x), where u = u(x) is the unknown function, x \in \Omega \subset \mathbb{R}^n, and the coefficients a_{ij}, b_i, c, and f are sufficiently smooth functions. This equation is classified as elliptic at a point x if the symmetric matrix A(x) = (a_{ij}(x)) is positive definite, meaning that for every nonzero vector \xi \in \mathbb{R}^n, the quadratic form \sum_{i,j=1}^n a_{ij}(x) \xi_i \xi_j > 0. A stronger condition often imposed for analytical purposes is uniform ellipticity, which ensures the holds globally with quantitative bounds. Specifically, the is uniformly elliptic if there exist constants \lambda > 0 and \Lambda > 0 such that for all x \in \Omega and all \xi \in \mathbb{R}^n, \lambda |\xi|^2 \leq \sum_{i,j=1}^n a_{ij}(x) \xi_i \xi_j \leq \Lambda |\xi|^2. This condition prevents degeneracy and facilitates estimates on solutions, such as energy inequalities derived from . Canonical examples of elliptic PDEs include , \Delta u = 0, where \Delta is the \sum_{i=1}^n \frac{\partial^2}{\partial x_i^2}, and , \Delta u = f, which generalizes it to a nonhomogeneous right-hand side. These equations arise in modeling steady-state phenomena, such as electrostatic potentials where u represents the satisfying with f related to . The modern study of elliptic PDEs, particularly boundary value problems, was pioneered by in the 1890s, who developed variational methods and integral representations to address issues like the for . 's contributions laid the groundwork for existence and uniqueness theories, influencing subsequent developments in and PDE regularity.

Relevant Function Spaces and Domains

The maximum principle for elliptic partial differential equations is typically considered in bounded open domains \Omega \subset \mathbb{R}^n equipped with smooth boundaries \partial \Omega, often assumed to be of class C^{1,\alpha} for some \alpha > 0 to facilitate boundary regularity and analysis near \partial \Omega. Such domains ensure that the closure \overline{\Omega} is compact, which is essential for attaining maxima and controlling solution behavior. This geometric setting allows for the application of and other local estimates without complications from unboundedness or irregularities. Classical solutions to the elliptic equation are functions u \in C^2(\Omega) \cap C(\overline{\Omega}), which are twice continuously differentiable in the interior \Omega and continuous on the closed domain \overline{\Omega}. These solutions satisfy the pointwise in \Omega, enabling direct verification of differential inequalities central to the principle. The up to the is particularly important, as it permits evaluation of the solution on \partial \Omega and comparison with interior values. Although weak solutions, defined in Sobolev spaces such as H^1(\Omega) or W^{1,p}(\Omega) for appropriate p, play a key role in existence and variational formulations, the maximum principle is most straightforwardly applied to classical solutions or to weak solutions that belong to C(\overline{\Omega}) via embedding theorems. Uniform ellipticity of the operator guarantees well-posedness of boundary value problems in these spaces, ensuring unique solutions under suitable conditions. For Dirichlet boundary conditions, the solution satisfies u = g on \partial \Omega, where g is a given continuous function on the boundary. This prescription is fundamental, as the principle often relates the maximum of u in \overline{\Omega} to the values of g, emphasizing the need for continuity of u up to \partial \Omega to rigorously state and prove bounds.

Weak Maximum Principle

Formal Statement

The weak maximum principle states that subsolutions to certain elliptic partial differential equations attain their maximum value on the boundary of the domain. Specifically, let \Omega \subset \mathbb{R}^n be a bounded open domain, and consider the uniformly elliptic operator Lu = a_{ij}(x) \partial_{ij} u + b_i(x) \partial_i u + c(x) u, where the coefficients satisfy $0 < \lambda I \leq (a_{ij}(x)) \leq \Lambda I for some constants $0 < \lambda \leq \Lambda with \Lambda / \lambda < \infty, and \sup_\Omega (|b_i| / \lambda + |c| / \lambda) < \infty, with c(x) \leq 0. If u \in C^2(\Omega) \cap C(\overline{\Omega}) satisfies Lu \geq 0 in \Omega, then \sup_{\overline{\Omega}} u = \sup_{\partial \Omega} u. This result holds under the assumptions of uniform ellipticity and bounded coefficients, ensuring the operator is non-degenerate. The connectedness of \Omega is not required for the weak principle. In the case of subharmonic functions, where \Delta u \geq 0 in \Omega (corresponding to L = \Delta), the principle takes a similar form: if u \in C^2(\Omega) \cap C(\overline{\Omega}), then \sup_{\overline{\Omega}} u = \sup_{\partial \Omega} u. The uniform ellipticity assumption ensures the operator's principal part is strictly positive definite, while the boundedness of lower-order coefficients prevents degeneracy.

Proof Outline

The proof of the weak maximum principle proceeds by contradiction, assuming that a subsolution u to the elliptic equation Lu \geq 0 in a bounded domain \Omega \subset \mathbb{R}^n achieves its maximum value M at an interior point x_0 \in \Omega. At this point, the first derivative vanishes, \nabla u(x_0) = 0, and the Hessian is negative semi-definite, \mathrm{Hess}\, u(x_0) \leq 0. For a uniformly elliptic operator L u = a^{ij} \partial_{ij} u + b^i \partial_i u + c u with a^{ij} positive definite, the second-order term satisfies a^{ij} \partial_{ij} u(x_0) \leq 0, the first-order term is zero, and if c \leq 0, the zeroth-order term c u(x_0) \leq 0 (or \geq 0 if u \leq 0), yielding L u(x_0) \leq 0. This contradicts L u \geq 0 unless u is constant, in which case the maximum aligns with the boundary values. To rigorously exclude the equality case and ensure the maximum cannot occur interiorly even when L u = 0 at x_0, an auxiliary function v = u + \varepsilon |x|^2 is introduced for small \varepsilon > 0, where |x|^2 is chosen relative to a fixed origin. The maximum of v occurs near that of u, and since u is continuous up to the \partial \Omega, for sufficiently small \varepsilon, this maximum remains interior if u's was. However, L v = L u + \varepsilon \cdot 2 \mathrm{trace}(A) \geq \varepsilon \cdot 2 \lambda > 0 at the maximum of v, where \lambda > 0 is the ellipticity constant from the of the matrix A = (a^{ij}), contradicting the fact that L v \leq 0 at an interior maximum of v. Thus, the maximum of u must lie on \partial \Omega. If the coefficient c < 0 in the zeroth-order term, the simple critical-point argument may not yield a strict contradiction when u(x_0) is negative, as c u(x_0) > 0 could offset the second-order negativity. In such cases, a modified like v = u e^{\alpha |x|^2} for suitable \alpha > 0 is used, transforming the to one where the adjusted zeroth-order becomes non-positive, allowing the argument to apply similarly and force the maximum to the boundary. The of u up to \partial \Omega guarantees that the global maximum over the compact closure \overline{\Omega} is attained, and the above arguments show it cannot be interior, hence it must occur on \partial \Omega.

Strong Maximum Principle

Formal Statement

The strong maximum principle asserts that non-constant solutions to certain elliptic partial differential equations cannot attain their maximum value in the interior of the . Specifically, let \Omega \subset \mathbb{R}^n be a connected open , and consider the uniformly Lu = a_{ij}(x) \partial_{ij} u + b_i(x) \partial_i u + c(x) u, where the coefficients satisfy $0 < \lambda I \leq (a_{ij}(x)) \leq \Lambda I for some constants $0 < \lambda \leq \Lambda with \Lambda / \lambda < \infty, and \sup_\Omega (|b_i| / \lambda + |c| / \lambda) < \infty. If u \in C^2(\Omega) satisfies Lu = 0 in \Omega and attains its maximum at an interior point x_0 \in \Omega, then u is constant throughout \Omega. This result, originally established by Hopf for linear elliptic equations of second order, extends the weak maximum principle by implying strict constancy rather than mere non-positivity of the interior maximum. In the case of subharmonic functions, where \Delta u \geq 0 in \Omega (with L = \Delta), the principle takes a similar form: if u \in C^2(\Omega) attains its supremum over \overline{\Omega} at an interior point x_0 \in \Omega, then u must be constant in \Omega. The uniform ellipticity assumption ensures the operator's principal part is strictly positive definite, while the boundedness of lower-order coefficients prevents degeneracy. A key corollary is Harnack's inequality, which quantifies the oscillation of positive solutions. For positive solutions u > 0 to Lu = 0 in \Omega, and for any compact subset K \subset \Omega, there exists a constant C = C(K, \Omega, L) > 0 such that \sup_K u \leq C \inf_K u. This follows directly from the strong maximum principle applied to differences of solutions and holds under the same assumptions of connectedness, uniform ellipticity, and coefficient boundedness.

Detailed Proof

The strong maximum principle for solutions to elliptic partial differential equations asserts that non-constant subsolutions cannot attain their maximum in the interior of the . We begin with the case of functions, where \Delta u = 0 in a connected bounded \Omega \subset \mathbb{R}^n, and u \in C^2(\Omega) \cap C(\overline{\Omega}). Assume, for contradiction, that u attains its maximum M = \sup_{\overline{\Omega}} u at an interior point x_0 \in \Omega. By the mean value property for functions, for any B_r(x_0) \subset \Omega, u(x_0) = \frac{1}{|B_r(x_0)|} \int_{B_r(x_0)} u(x) \, dx \leq M, with equality only if u \equiv M on B_r(x_0). Since u(x_0) = M, it follows that u \equiv M on B_r(x_0). Repeating this argument over overlapping balls covering \Omega and using the connectedness of \Omega, we conclude u \equiv M in \Omega, contradicting the assumption that u is non-constant. For the general linear elliptic case, consider the operator Lu = a_{ij}(x) \partial_{ij} u + b_i(x) \partial_i u + c(x) u \geq 0 in \Omega, where L is uniformly elliptic with bounded measurable coefficients and c \leq 0. The proof relies on reducing to the harmonic case via transformations and applying the Hopf boundary point lemma. First, a change of variables eliminates the first-order terms: locally, choose \phi such that \nabla \phi \approx -b/2, transforming L into a Schrödinger operator \tilde{\Delta} v + \tilde{c} v \geq 0 with no first-order terms and \tilde{c} \leq 0. For the zero-order term, consider perturbations or barrier functions on small balls around the assumed interior maximum point. If v attains a nonnegative interior maximum, a local barrier argument or the weak maximum principle combined with the Hopf lemma leads to a contradiction unless v (and hence u) is constant. A quantitative refinement is the strong Harnack inequality, which bounds the oscillation of positive solutions. For u > 0 solving Lu = 0 in a ball B_r(x_0) \subset \Omega, integral representations or chaining local estimates yield \sup_{B_{r/2}} u \leq C \inf_{B_{r/2}} u, with C depending on n and ellipticity constants, extending to global domains. Finally, the Hopf boundary point lemma provides a rigorous gradient estimate near the boundary, underpinning the strictness of the maximum principle. Suppose u solves Lu \geq 0 in \Omega, attains maximum M at y \in \partial \Omega with an interior ball tangent at y, and u < M in \Omega. Then, \partial u / \partial \nu (y) > 0, where \nu is the outward normal. To prove, consider auxiliary function \phi(x) = e^{-\alpha |x - z|^2} - e^{-\alpha R^2} for center z inside the tangent ball of radius R, choosing \alpha large so L\phi > 0 in \Omega. Set w = u - M + \epsilon \phi \leq 0 in \Omega, with w(y) = 0. By the weak maximum principle (applied appropriately to the sign), w < 0 in \Omega, and differentiating at y gives \partial w / \partial \nu (y) \leq 0, so \epsilon \partial \phi / \partial \nu (y) \geq \partial u / \partial \nu (y) > 0 for small \epsilon, yielding the strict inequality.

Generalizations and Applications

Extensions to Other PDE Types

The maximum principle extends naturally to parabolic partial differential equations (PDEs), such as the u_t - \Delta u = 0 in a bounded \Omega \subset \mathbb{R}^n for t > 0, where solutions satisfy a weak maximum principle forward in time: the maximum value of u over the cylinder [0, T] \times \overline{\Omega} is attained on the parabolic boundary consisting of the initial time t = 0 and the lateral boundary \partial \Omega \times [0, T]. This principle arises from the smoothing and diffusive properties of parabolic operators, ensuring that interior maxima cannot exceed boundary values unless the solution is constant. For the strong maximum principle in the parabolic setting, if a subsolution attains its maximum at an interior space-time point (x_0, t_0), then the subsolution is constant throughout the backward parabolic cylinder \Omega \times [0, t_0]. In contrast, hyperbolic PDEs like the wave equation u_{tt} - \Delta u = 0 do not admit a comparable interior maximum principle due to the finite propagation speed of disturbances, which allows waves to carry maxima from the boundary into the interior without diffusion. Solutions can thus achieve local maxima inside the domain that surpass boundary values, as information propagates along characteristics rather than smoothing out irregularities, highlighting a fundamental limitation absent in elliptic or parabolic cases. For fully nonlinear elliptic PDEs of the form F(D^2 u) = 0, where F is (implying degenerate ellipticity), the maximum is established in the framework of viscosity solutions, pioneered by Crandall and Lions in the 1980s. Viscosity solutions, defined via test functions and subsolution/supersolution , ensure uniqueness and a comparison under structural conditions on F, such that if a subsolution is less than or equal to a supersolution at a maximum point, the holds globally. This extension accommodates nonclassical solutions where traditional C^2 regularity fails, relying on the degenerate elliptic nature to control oscillations. Time-dependent generalizations of the maximum principle to broader classes of equations leverage theory, where the solution operator e^{tA} generated by an A preserves positivity or bounds for appropriate initial data. For abstract semilinear parabolic equations u_t = Au + f(u) in Banach spaces, with A sectorial and f , the approach yields maximum bound principles that bound \|u(t)\|_\infty by initial and forcing terms, extending classical results to nonlinear and infinite-dimensional settings. This framework unifies handling of boundary conditions and nonlinearities, ensuring the principle holds uniformly in time under contractivity assumptions on the .

Practical Applications in Analysis

The maximum principle plays a pivotal role in establishing for solutions to the associated with the Poisson equation \Delta u = f in bounded domains. Specifically, for two solutions u_1 and u_2 satisfying the same conditions on \partial \Omega, the difference w = u_1 - u_2 is and attains its on the boundary, implying w \equiv 0 and thus uniqueness. This result extends to via Perron's method, where the maximum principle ensures the convergence of the supremum of subsolutions to a matching the boundary data. In unbounded domains, the Phragmén–Lindelöf principle extends the maximum principle to control the growth of solutions at infinity, preventing unbounded oscillations for elliptic equations. For instance, in a sector of the , if a positive is bounded by an rate on the boundary rays, it remains controlled throughout the sector, yielding boundedness or constancy under suitable conditions. This principle is crucial for analyzing asymptotic behavior in exterior domains, such as in , where it implies that solutions to \Delta u = 0 with prescribed growth cannot exceed boundary estimates without violating the principle. A direct consequence is the Liouville theorem, which states that any bounded entire harmonic function on \mathbb{R}^n must be constant. The proof relies on the maximum principle applied to balls of increasing radius: since the function is bounded, its maximum on each ball occurs on the boundary, but uniformity of bounds forces constancy via the mean value property. This theorem underscores the rigidity of harmonic functions in Euclidean space, with applications in complex analysis where it parallels the constancy of bounded entire holomorphic functions. In modern analysis, the maximum principle ensures regularity in free boundary problems, such as the classical obstacle problem, where the u \geq \psi minimizes the subject to an obstacle \psi. By applying the principle to the difference between u and \psi, one verifies that u is C^{1,1} regular away from the free boundary \partial \{u > \psi\}, facilitating higher-order expansions and blow-up analysis for the interface. Similarly, in problems governed by elliptic PDEs, like minimizing a cost functional subject to \Delta y = f(u) with control u, the maximum principle provides a priori bounds on the state y, enabling uniqueness of the adjoint state and characterization of optimal controls via Pontryagin-type conditions adapted to the elliptic setting. These applications highlight the principle's utility in guaranteeing stability and in variational frameworks.

References

  1. [1]
    [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. ...
  2. [2]
    [PDF] Maximum principles, Harnack inequality for classical solutions
    The strong maximum principle is typically used to prove uniqueness of solutions to elliptic Dirichlet boundary value problems. The difference u of two such ...
  3. [3]
    [PDF] Hamilton's Ricci Flow - Math (Princeton)
    The Maximum Principle. The maximum principle is the key tool in understanding many parabolic partial differential equations. It appears in many guises, but ...<|control11|><|separator|>
  4. [4]
    [PDF] Lectures on Partial Differential Equations - Applied Mathematics
    A probabilistic proof of the maximum principle will be presented after we consider the heat equation in a few weeks. We now obtain our first uniqueness theorem ...Missing: history | Show results with:history
  5. [5]
    [PDF] From Conformal Invariants to Percolation
    Apr 30, 2021 · Corollary 3.2 (Maximum Principle) A harmonic function on a connected region does not achieve its maximum unless it is constant. Proof. Let M = ...
  6. [6]
    None
    ### Intuitive Explanation of the Maximum Principle for Harmonic Functions
  7. [7]
    [PDF] Chapter 2: Laplace's equation - UC Davis Math
    Maximum principle. The maximum principle states that a non-constant harmonic function cannot attain a maximum (or minimum) at an interior point of its domain.
  8. [8]
    [PDF] 2 Heat Equation
    In this section, we prove what is known as the maximum principle for the heat equation. We will then use this principle to prove uniqueness of solutions to the ...<|separator|>
  9. [9]
    [PDF] Exam II Fall 2005 - MST.edu
    (c) Show, by exhibiting a counterexample, that solutions to the wave equation need not satisfy a maximum principle. or on a sidewall (ie, when x=0 or x=L). 8 ...
  10. [10]
    [PDF] Chapter 4: Elliptic PDEs - UC Davis Math
    next section. 4.6. General linear, second order elliptic PDEs. Consider PDEs of the form. Lu = f where L is a linear differential operator of the form. (4.16).
  11. [11]
    [PDF] 4 Classification of Second-Order Equations
    If b2 − 4ac > 0, we say the equation is hyperbolic. If b2 − 4ac = 0, we say the equation is parabolic. If b2 − 4ac < 0, we say the equation is elliptic.
  12. [12]
    [PDF] Elliptic partial differential equations
    Nov 6, 2023 · ). Applying the uniform ellipticity condition (1.8) with ξj = ∂ju and observing that. |ξ| = |∇u|, we obtain. [u, u]a = ZΩ n. X i,j=1 aij(x) ...
  13. [13]
    [PDF] Introduction and elliptic PDEs Partial differential equations
    Solutions to Laplace's equation are called harmonic functions. Examples of applications: Electrostatics, steady fluid flow, analytic functions, ... Poisson's ...
  14. [14]
    Elliptic Partial Differential Equation -- from Wolfram MathWorld
    A linear PDE is elliptic if its principal symbol, as in the theory of pseudodifferential operators, is nonzero away from the origin. For instance, (◇) has as ...
  15. [15]
    [PDF] Henri Poincar´e and partial differential equations
    Sep 3, 2012 · Henri Poincaré introduced a new approach to solve the Dirichlet prob- lem and he gave the first general solution of the initial value problem.
  16. [16]
    None
    ### Formal Statement of the Strong Maximum Principle for Elliptic PDEs
  17. [17]
    The strong maximum principle revisited - ScienceDirect.com
    Jan 1, 2004 · In this paper we first present the classical maximum principle due to E. Hopf, together with an extended commentary and discussion of Hopf's paper.Missing: original | Show results with:original
  18. [18]
    [PDF] Chapter 6: Parabolic equations - UC Davis Math
    For example, there are parabolic versions of the maximum principle and Harnack's inequality, and a Schauder theory for Hölder continuous solutions [28].
  19. [19]
    [PDF] Appetizers in Nonlinear PDEs - Mathematics - Stanford University
    May 2, 2017 · As the parabolic maximum principle underlies most of the parabolic existence and regularity theory, we first recall some basics on the ...<|separator|>
  20. [20]
    [PDF] Chapter 7: Hyperbolic equations - UC Davis Math
    The wave equation. The prototypical example of a hyperbolic PDE is the wave equation. (7.1) utt = ∆u. To begin with, consider the one-dimensional wave equation ...Missing: counterexample | Show results with:counterexample
  21. [21]
    The maximum principle for viscosity solutions of fully nonlinear ...
    We prove that viscosity solutions in W 1,∞ of the second order, fully nonlinear, equation F(D 2 u, Du, u) = 0 are unique when (i) F is degenerate ell.
  22. [22]
    Maximum Bound Principles for a Class of Semilinear Parabolic ...
    In this paper, we consider a practically desirable property for a class of semilinear parabolic equations of the abstract form 𝑢 𝑡 = L ⁢ 𝑢 + 𝑓 ⁡ [ 𝑢 ] , with L ...
  23. [23]
    Maximum bound principles for a class of semilinear parabolic ... - arXiv
    May 23, 2020 · In this paper, we consider a practically desirable property for a class of semilinear parabolic equations of the abstract form u_t=\mathcal{L}u+f[u]
  24. [24]
    [PDF] partial-differential-equations-by-evans.pdf - Math24
    I present in this book a wide-ranging survey of many important topics in the theory of partial differential equations (PDE), with particular emphasis.<|separator|>
  25. [25]
    ON TWO THEOREMS OF PHRAGMEN-LINDELOF FOR LINEAR ...
    A generalization of the classic Phragmen-Lindelof theorem to elliptic equations was given by Gilbarg [1] and Hopf [4]. A refined form of that classic theorem, ...
  26. [26]
    [PDF] 2.10 Phragmén-Lindelöf principle - HKUST Math Department
    2.10 Phragmén-Lindelöf principle. Example 2.10.1. Let f(z) = exp(γza), γ > 0 ... |F(z)| ≤ M. By the maximum modulus principle (extended version), we deduce.
  27. [27]
    [PDF] Harmonic Function Theory - Sheldon Axler
    Suppose u is a real-valued harmonic function on the closed unit disk in ... the maximum principle, if a solution exists, then it is unique. We take our ...
  28. [28]
    [PDF] The Obstacle Problem - UZH
    I., CAFFARELLI, L.A., A theorem of real analysis and its application to free boundary problems, Comm.Pure. Appl. Math. 38 (1985),. 499-502. [B-K-P] W. BECKNER ...
  29. [29]
    Maximum Principles in the Optimal Control of Semilinear Elliptic ...
    We obtain Pontryagin's maximum principle for local solutions of an optimal control problem with a monotone semilinear elliptic system and an integral cost.