Fact-checked by Grok 2 weeks ago

Potential theory

Potential theory is a branch of that investigates the properties of functions and their generalizations, including subharmonic and superharmonic functions, which arise as solutions to \Delta u = 0 or related inequalities. It originated in the early from physical models of gravitation and , where potentials describe force fields generated by mass or charge distributions. Central to the field is the concept of potentials associated to measures via the Laplacian, such as the u(x) = \int \frac{1}{|x-y|} d\mu(y) in three dimensions (superharmonic for \mu \geq 0) or the logarithmic potential p_\mu(z) = \int \log |z - w| d\mu(w) in the plane (subharmonic for \mu \geq 0). Historically, potential theory developed through contributions from figures like George Green, , , and , who formalized boundary value problems such as the for functions. Green's 1828 essay laid foundational representations, while Gauss and Poisson advanced the theory in the context of flux and attraction laws. In the mid-20th century, the field was axiomatized by mathematicians including Marcel Brelot, Gustave Choquet, and Jean Deny, incorporating probabilistic interpretations via and excessive functions, as emphasized by . Key concepts include the for functions, which states that a non-constant on a bounded attains its maximum on the , and the Riesz , which decomposes a u on a compact set K as u = p_\mu + h, where h is and $2\pi \mu = \Delta u|_K. are upper semicontinuous and satisfy the sub-mean value property: u(w) \leq \frac{1}{2\pi} \int_0^{2\pi} u(w + r e^{it}) dt. These ideas extend to applications in , such as pluripotential theory on Riemann surfaces, and in probability, where functions relate to martingales and Markov processes. Potential theory also aids in computing dimensions of sets, like via Frostman's s-potentials, where \int \phi_s(x) d\mu(x) < \infty implies \dim_H(A) \geq s.

Introduction

Definition and Scope

Potential theory is a branch of mathematical analysis that studies harmonic functions and their generalizations, such as subharmonic and superharmonic functions, with a focus on their properties and applications. The term "potential theory" originated in 19th-century physics, stemming from the concept of potentials used to describe fundamental forces like gravity and electrostatic attraction, as developed in works on Newtonian gravitation and . This mathematical framework emerged to formalize these physical ideas, providing tools for analyzing fields derived from scalar potentials. The scope of potential theory encompasses solutions to elliptic partial differential equations, particularly in Euclidean spaces, where it addresses boundary value problems and integral representations of functions. In two dimensions, it maintains strong connections to complex analysis through the identification of harmonic functions with real parts of holomorphic functions, while in higher dimensions, it integrates with broader partial differential equation theory to study regularity and asymptotic behavior. Harmonic functions serve as the central objects, satisfying the mean value property and maximum principles that underpin the theory's analytical structure. A key application lies in modeling conservative force fields, such as gravitational or electrostatic fields, where the force vector is expressed as the negative gradient of a scalar potential function, ensuring path-independent work. This scalar focus distinguishes potential theory from aspects of electromagnetism involving vector potentials, which account for magnetic effects and non-conservative components not captured by scalar fields alone.

Historical Development

Potential theory originated in the physical sciences of the 18th and early 19th centuries, rooted in efforts to model gravitational and electrostatic forces. Isaac Newton's formulation of the inverse-square law of gravitation in his Philosophiæ Naturalis Principia Mathematica (1687) provided the foundational force law that later inspired the concept of gravitational potential, enabling the representation of forces as gradients of scalar fields. Similarly, Charles-Augustin de Coulomb's 1785 experiments using a torsion balance established the analogous inverse-square law for electrostatic forces between charged particles, laying groundwork for electrostatic potentials. These physical laws were mathematically unified in 1828 when George Green published An Essay on the Application of Mathematical Analysis to the Theories of Electricity and Magnetism, introducing the potential function as a central tool and Green's theorem, which relates surface integrals of potentials to boundary fluxes, thus bridging force calculations to harmonic functions. Carl Friedrich Gauss further advanced the field in 1813 with his divergence theorem, establishing key integral identities for flux through surfaces, essential for potential representations. In the early 19th century, Pierre-Simon Laplace advanced potential theory through his multi-volume Mécanique Céleste (1799–1825), where he employed potential integrals to analyze perturbations in celestial mechanics, demonstrating the stability of planetary orbits under gravitational influences. Building on this, Siméon Denis Poisson formulated in 1813 during his studies of electrostatics, expressing the relationship between charge density and the Laplacian of the potential, which generalized Laplace's equation for non-vacuum cases. Lord Kelvin (William Thomson) extended these ideas in the 1840s and 1850s, developing the method of images—a symmetry technique for solving boundary value problems in electrostatics—and the , which preserves harmonic functions under inversion, facilitating solutions for spherical and planar geometries. The mid-19th century saw potential theory evolve into a rigorous mathematical discipline through connections to complex analysis and boundary value problems. Bernhard Riemann's 1851 doctoral thesis on complex functions linked two-dimensional potentials to conformal mappings, showing how analytic functions generate harmonic potentials via real and imaginary parts, as captured by the Cauchy-Riemann equations. Peter Gustav Lejeune Dirichlet formalized boundary value problems in the 1830s and 1850s, introducing the Dirichlet principle, which posits the existence of harmonic functions minimizing energy integrals subject to boundary conditions, though its proof faced challenges until later rigorization. By the early 20th century, David Hilbert's 1904 work on integral equations provided a spectral approach to solving potential problems, treating them as Fredholm equations and establishing existence via eigenvalue expansions, which influenced operator theory in partial differential equations (PDEs). Twentieth-century developments shifted potential theory toward irregular domains and nonlinear extensions, solidifying its role in pure mathematics. Norbert Wiener's work in the 1920s introduced capacity theory, quantifying the "size" of sets with respect to harmonic measures and enabling solutions for non-smooth boundaries via Wiener's criterion for regularity. Modern extensions to nonlinear potentials, emerging in the mid-20th century, generalized classical theory to p-Laplacian equations and quasilinear PDEs, with foundational contributions addressing subharmonic functions and variational inequalities. In the mid-20th century, the field was axiomatized by mathematicians including , , and , incorporating probabilistic interpretations via and excessive functions, as emphasized by . This evolution facilitated the transition from physics-inspired methods to abstract tools in PDE analysis, influencing existence proofs, regularity theory, and stochastic processes in pure mathematics.

Fundamental Concepts

Harmonic Functions

A harmonic function u on an open domain \Omega \subset \mathbb{R}^n is a real-valued function that is twice continuously differentiable and satisfies Laplace's equation \Delta u = 0, where \Delta = \sum_{j=1}^n \frac{\partial^2}{\partial x_j^2} denotes the Laplacian operator. This condition ensures that u represents a potential without sources or sinks within \Omega. A defining property of harmonic functions is the mean value property: for any ball B(a, r) \subset \Omega centered at a with radius r > 0, the value u(a) equals the of u over the boundary \partial B(a, r), given by u(a) = \frac{1}{\sigma_{n-1}} \int_{\partial B(a, r)} u \, d\sigma, where \sigma_{n-1} is the surface area of the unit in \mathbb{R}^n. An equivalent states that u(a) is the over the entire ball B(a, r). This property implies that harmonic functions are infinitely differentiable (C^\infty) and real analytic in \Omega, meaning they can be locally represented by converging expansions. The set of harmonic functions on \Omega forms a vector space under pointwise addition and scalar multiplication, owing to the linearity of the Laplacian operator: if u and v are , then so are \alpha u + \beta v for scalars \alpha, \beta. Basic examples include constant functions u(x) = c, which trivially satisfy \Delta u = 0, and linear functions such as u(x) = a \cdot x + b (where a \in \mathbb{R}^n and b \in \mathbb{R}), as their second derivatives vanish. More generally, the fundamental solution in \mathbb{R}^n for n > 2 is \Phi(x) = -\frac{1}{(n-2) \omega_n |x|^{n-2}}, a radial function away from the , where \omega_n is the surface area of the unit . In bounded domains like balls, —homogeneous polynomials restricted to the —provide a complete for expanding functions via in spherical coordinates. In potential theory, functions serve as the foundational solutions to homogeneous boundary value problems, such as the , where they model potentials determined by boundary data without interior sources. They also connect to physical interpretations, representing equilibrium states in fields like gravitation and .

Laplace's and Poisson's Equations

In potential theory, is the central governing equilibrium states in source-free regions. It is expressed in vector form as \Delta u = \sum_{i=1}^n \frac{\partial^2 u}{\partial x_i^2} = 0, where u is the potential function in n-dimensional Euclidean space, and \Delta denotes the Laplacian operator. This equation describes the behavior of potential fields, such as electrostatic or gravitational potentials, in regions devoid of sources, representing a balance where the divergence of the field vanishes. The physical derivation of Laplace's equation arises from fundamental laws of field theory. In , states that the divergence of the \mathbf{D} equals the free \rho_v, or \nabla \cdot \mathbf{D} = \rho_v, where \mathbf{D} = \epsilon \mathbf{E} and \epsilon is the . Since the \mathbf{E} = -\nabla u, substitution yields \nabla \cdot (\epsilon \nabla u) = -\rho_v; in homogeneous media where \epsilon is and \rho_v = 0, this simplifies to \Delta u = 0. Similarly, in gravitation, for the \mathbf{g} gives \nabla \cdot \mathbf{g} = -4\pi G \rho, with \mathbf{g} = -\nabla u and G the ; in source-free regions (\rho = 0), this again leads to \Delta u = 0. Poisson's equation generalizes to include distributed sources, taking the form \Delta u = f, where f represents the source term (e.g., f = -\rho_v / \epsilon or f = 4\pi G \rho ). exist and can be constructed using the fundamental solution, which is the potential due to a unit point source; , this is \Phi(\mathbf{x}) = -\frac{1}{4\pi |\mathbf{x}|}, satisfying \Delta \Phi = \delta(\mathbf{x}), where \delta is the . For bounded domains, Green's functions provide the appropriate framework for solvability, incorporating boundary conditions to yield the general as an integral over the source f and boundary data. Boundary value problems for these equations are formulated to determine the potential within a \Omega given data on its \partial \Omega. The prescribes the potential values u = g on \partial \Omega, while the problem specifies the normal derivative \frac{\partial u}{\partial n} = h on \partial \Omega, related to the of the field. For , the Neumann formulation requires a compatibility condition \int_\Omega f \, dV = \oint_{\partial \Omega} h \, dS to ensure solvability. Uniqueness theorems guarantee that solutions, when they exist, are determined up to additive constants under suitable conditions. For the Dirichlet problem, both Laplace's and Poisson's equations have unique solutions in bounded domains with continuous boundary data, as the difference of any two solutions satisfies with zero boundary values and must vanish by the . For the Neumann problem, solutions are unique up to an additive constant, with the compatibility condition ensuring existence; this follows from integrating the equation over the domain and applying the .

Symmetries and Transformations

Conformal Symmetries

Conformal transformations are mappings that preserve angles and locally scale distances uniformly, playing a central role in the study of functions as solutions to \Delta u = 0. In two dimensions, these transformations correspond to complex s, and the invariance of the Laplace equation under such mappings follows from the Cauchy-Riemann equations, which ensure that the real and imaginary parts of an analytic function are . Specifically, if u is in a U \subset \mathbb{R}^2 and f: V \to U is conformal (holomorphic with non-zero ), then u \circ f is in V, preserving the structure of \Delta (u \circ f) = 0. In higher dimensions, the conformal group extends to include Möbius transformations, which are compositions of inversions, translations, rotations, and scalings. These transformations preserve harmonic functions exactly, as exemplified by the Kelvin transform K[u](x) = |x|^{2-n} u(x/|x|^2) for n \geq 3, which maps harmonic functions to harmonic functions. Inversions, a key component of Möbius transformations, act conformally on \mathbb{R}^n \setminus \{0\} by mapping spheres to spheres or planes, and the resulting composition adjusts the Laplacian such that harmonic solutions remain harmonic. This preservation allows for the extension of harmonic functions across domains transformed by such symmetries. Spherical harmonics provide a concrete realization of rotational symmetries in potential theory, serving as eigenfunctions of the Laplace-Beltrami operator on the unit sphere S^{n-1} under the action of the rotation group SO(n). These functions, which are restrictions of homogeneous polynomials of d to , satisfy \Delta_{S^{n-1}} Y_d = -d(d + n - 2) Y_d, where the eigenvalue reflects the SO(n)-invariance of the spherical Laplacian. As an for L^2(S^{n-1}), decompose general functions in balls or spheres, facilitating the analysis of rotationally symmetric solutions in potential theory. Symmetries enable the generation of new from known ones, such as through principles, which extend solutions across or spheres while preserving harmonicity. For instance, reflecting a across a yields another , and combining this with inversions generates solutions in complementary domains, like extending from a to its exterior via the Kelvin transform adjusted by . This constructive approach leverages the underlying symmetries to solve boundary value problems without direct . From a group-theoretic , Laplace's equation is invariant under the full , which includes translations, rotations, dilations, and special conformal transformations (inversions composed with translations). Under the conformal group actions (translations, rotations, dilations, and special conformal transformations), the Laplacian applied to the u \circ g results in a scaled version of (\Delta u) \circ g, but since \Delta u = 0 for u, u \circ g is also harmonic. The specific scaling depends on the transformation type. This invariance underscores the conformal group's role in classifying and generating solutions in potential theory across dimensions.

Kelvin Transform and Method of Images

The Kelvin transform, introduced by in a 1845 letter to , provides a geometric inversion that preserves the harmonicity of functions in potential theory. For a function u that is in \mathbb{R}^n \setminus \{0\} with n \geq 3, the Kelvin transform is defined as v(x) = |x|^{2-n} u\left(\frac{x}{|x|^2}\right), which remains in \mathbb{R}^n \setminus \{0\}. This transformation arises from the inversion mapping x \mapsto \frac{x}{|x|^2}, which interchanges points inside and outside the unit sphere while mapping spheres to planes or vice versa, thereby facilitating solutions to boundary value problems by converting exterior domains to interior ones. In applications, the Kelvin transform is particularly useful for solving around spherical boundaries, such as mapping the potential outside a to an equivalent problem inside an inverted domain. For instance, it allows reduction of unbounded exterior problems to bounded interior ones, preserving key properties like the mean value property of functions, as detailed in classical treatments of potential theory. This technique extends naturally to under appropriate scaling, enabling analytical solutions for source distributions symmetric under inversion. The , developed by in his 1848 paper on , constructs solutions to potential problems by introducing fictitious charges (images) that enforce boundary conditions without altering the field in the region of interest. In , for a point charge q at distance d from an infinite grounded conducting plane, the image charge is -q placed symmetrically at distance d on the opposite side, yielding zero potential on the plane while matching the original field above it. This approach satisfies the (constant potential) on linear boundaries through superposition of the real and image potentials. A key example is the grounded conducting sphere of radius a with a point charge q at distance b > a from the center: the image charge is q' = -q \frac{a}{b} at distance \frac{a^2}{b} inside the , ensuring the 's surface is at zero. For a external field around an uncharged conducting , image dipoles or equivalent multipoles can be derived similarly, modeling induced surface charges. These methods extend to infinite planes for approximating parallel-plate configurations or line charges near cylindrical boundaries via two-dimensional analogs. While effective for planar and spherical geometries, the is limited to linear boundary conditions and simple symmetries, requiring infinite series of images for more complex shapes like wedges. Extensions to curved boundaries, such as circles in two dimensions, often combine the method with inversions akin to the Kelvin transform to generate valid image systems.

Dimensional Considerations

Properties in Two Dimensions

In two dimensions, the fundamental solution to , known as the logarithmic potential, takes the form \Phi(x,y) = -\frac{1}{2\pi} \ln \sqrt{x^2 + y^2}, which exhibits slower at compared to the power-law \frac{1}{|r|^{n-2}} observed for n \geq 3. This logarithmic behavior arises naturally from the for the plane and leads to distinct asymptotic properties for potentials generated by compact charge distributions, where the potential grows logarithmically rather than approaching a . Consequently, solutions to boundary value problems in unbounded domains often require careful handling of behavior at , influencing applications in and . A hallmark of two-dimensional potential theory is the central role of conformal mappings, enabled by the , which asserts that any simply connected domain in the , excluding the entire plane itself, can be conformally mapped onto the unit disk. This uniformization simplifies the solution of Dirichlet and Neumann problems by transforming irregular boundaries into circular ones, preserving harmonicity since conformal maps are angle-preserving and satisfy the Cauchy-Riemann equations. Such mappings facilitate explicit constructions of Green's functions and highlight the deep interconnection between potential theory and in this dimension. Harmonic functions in two dimensions admit a complex representation: any real-valued u(x,y) on a simply connected domain is the real part of a f(z) = u + iv, where v is the of u. This , unique up to an additive constant, leverages the Cauchy-Riemann equations and enables powerful tools from , such as the , which extends harmonic functions across straight-line boundaries by reflecting the conjugate. Similarly, variants of Morera's theorem apply to verify holomorphicity of the combined function, confirming that closed contours with vanishing integrals for both u and v imply analyticity. The governing harmonic functions in two dimensions is the full , comprising all angle-preserving transformations, which forms an infinite-dimensional generated by holomorphic and anti-holomorphic functions. This contrasts sharply with the finite-dimensional special in higher dimensions, allowing for a richer class of invariances that preserve solutions to . These symmetries underpin the extensibility of potentials across domains and the solvability of mixed problems through local adjustments. Special theorems exploit the conjugate structure, such as the argument principle for harmonic functions, which counts the of level curves around singularities by integrating the of the conjugate along the . This principle, analogous to its holomorphic counterpart, quantifies in the sense-reversing regions of harmonic mappings, providing topological insights into the global structure of solutions. Local regularity properties, including real-analyticity away from singularities, align with those in higher dimensions.

Behavior in Higher Dimensions

In dimensions n \geq 3, the solution to \Delta \Phi = -\delta takes the form \Phi(x) = \frac{1}{|x|^{n-2}} (up to a dimensional constant), which facilitates the construction of multipole expansions for representing potentials generated by localized charge distributions. This power-law decay contrasts with the logarithmic singularity in two dimensions and enables efficient asymptotic approximations for far-field behaviors in higher-dimensional settings. The group of conformal transformations preserving the Laplace equation in \mathbb{R}^n for n \geq 3 is the finite-dimensional Möbius group, generated by inversions, isometries, translations, and dilations, which imposes limitations on generating new solutions compared to the infinite-dimensional available in two dimensions. A key tool within this framework is the transform, defined for a u in a excluding the origin as Ku(x) = |x|^{2-n} u(x/|x|^2), which preserves harmonicity throughout \mathbb{R}^n \setminus \{0\} and extends classical methods for solving boundary value problems. Regarding asymptotic behavior at infinity, harmonic functions in \mathbb{R}^n (n \geq 3) exhibit at most growth, meaning that if |u(x)| \leq C (1 + |x|)^k for some C, k > 0, then u is a of degree at most k. Liouville-type theorems further imply that bounded entire functions in \mathbb{R}^n are , providing strong rigidity results for global solutions. These properties find direct applications in modeling Newtonian gravity and in three dimensions, where the potential decays as $1/r and the field as $1/r^2, leading to faster dissipation of influences at large distances than in lower-dimensional analogs.

Analytic Properties

Local Behavior and Regularity

, defined as twice continuously differentiable solutions to \Delta u = 0 in an open domain \Omega \subset \mathbb{R}^n, exhibit exceptional smoothness properties throughout the interior of \Omega. A fundamental result in potential theory is the regularity theorem, which states that any such harmonic function u is infinitely differentiable (C^\infty) in \Omega. Furthermore, harmonic functions are real analytic in \Omega, allowing local representation by convergent expansions. The proof of this C^\infty regularity leverages the mean value property of harmonic functions: for any ball B_r(x) \subset \Omega with x \in \Omega and radius r > 0, u(x) = \frac{1}{|B_r(x)|} \int_{B_r(x)} u(y) \, dy = \frac{1}{|\partial B_r(x)|} \int_{\partial B_r(x)} u(y) \, d\sigma(y), where | \cdot | denotes the respective measure. By applying Taylor expansions to u around x and differentiating the mean value integrals with respect to parameters such as r, one inductively verifies the existence and continuity of all higher-order partial derivatives at x. This process demonstrates that the smoothness order can be arbitrarily increased, yielding C^\infty regularity without relying on the original C^2 assumption for the full result. A consequential aspect of this local regularity is the expansion of a u around any interior point x_0 \in \Omega in terms of homogeneous polynomials. Specifically, in spherical coordinates centered at x_0, u admits the series representation u(x) = \sum_{k=0}^\infty |x - x_0|^k P_k\left( \frac{x - x_0}{|x - x_0|} \right), where each P_k is a homogeneous polynomial of degree k on the unit , and the series converges uniformly on compact subsets of \Omega. This expansion underscores the analytic nature of functions and facilitates the study of their local geometry. In the broader framework of elliptic partial differential equations, the regularity of harmonic functions extends to weak solutions in Sobolev spaces W^{1,2}(\Omega). Elliptic regularity employs bootstrap arguments: starting from L^2 or L^p estimates on \Delta u = 0, one iteratively applies interior Schauder or Calderón-Zygmund estimates to upgrade the solution from W^{2,p} to C^{k,\alpha} and ultimately to C^\infty. These techniques confirm that weak harmonic functions coincide with classical ones in the interior. The local uniqueness of in follows from the , which prohibits interior local maxima or minima unless u is constant. Thus, within any B \subset \Omega, a is uniquely determined by its values on \partial B, barring the constant case.

Singularities and Bôcher's Theorem

In potential theory, an of a u at a point p in \mathbb{R}^n (with n \geq 2) is removable if u is bounded in a punctured neighborhood of p, allowing u to be extended to a on the full neighborhood. This result, analogous to Riemann's removable singularity theorem for holomorphic functions, ensures that the singularity can be "filled in" harmonically without altering the function's behavior elsewhere. Isolated singularities of harmonic functions are classified as removable, poles, or essential. A singularity at p is removable if \lim_{x \to p} |x - p|^{n-2} |u(x)| = 0 (for n > 2); it is a pole if there exists an integer M \geq 0 such that $0 < \limsup_{x \to p} |x - p|^{M + n - 2} |u(x)| < \infty; otherwise, it is essential. Near a pole, u can be expressed as h(x) + \sum_{j=1}^{m} |x - p|^{-k_j} g_j(\theta) plus higher-order terms, where the h is at p, the k_j are positive s increasing, \theta are angular coordinates, and each g_j is a spherical on the unit sphere. Essential singularities involve infinitely many such negative power terms, leading to more complex behavior. Bôcher's theorem provides a specific classification for positive harmonic functions near an isolated singularity. For a positive harmonic function u defined and twice continuously differentiable in a punctured neighborhood O \setminus \{p\} of p \in \mathbb{R}^n, there exists a harmonic function v on the full neighborhood O and a constant a \geq 0 such that, near p = 0 in the punctured unit ball B_n \setminus \{0\},
  • if n > 2, u(x) = a |x|^{2-n} + v(x);
  • if n = 2, u(x) = a \log(1/|x|) + v(x). In this case, the singular term is a non-negative multiple of the fundamental solution, and the angular dependence reduces to a constant (the degree-0 spherical harmonic). This theorem originates from Maxime Bôcher's analysis of solutions to elliptic PDEs, including , and parallels pole-like behavior but restricted to positive functions.
A representative example is the fundamental solution of , \Phi(x) = |x|^{2-n}/(n-2) for n > 2, which exhibits a pole-like of order n-2 at the , fitting the form with a = 1 and v = 0 up to constants. This underscores the theorem's role in understanding Green's functions and potentials with point sources.

Inequalities and Principles

The strong maximum principle for functions asserts that if u is a non-constant on a connected \Omega \subseteq \mathbb{R}^n, then u cannot attain a local maximum or minimum at any interior point of \Omega. More precisely, if u attains its maximum value on \overline{\Omega} at an interior point, then u must be constant throughout \Omega. This principle follows directly from the mean value property of functions, which states that for any B(x, r) \subset \Omega, u(x) = \frac{1}{|S(x, r)|} \int_{S(x, r)} u(y) \, d\sigma(y), where S(x, r) is the sphere of radius r centered at x, and |\cdot| denotes surface measure. To prove the strong version, suppose for contradiction that u attains a local maximum at an interior point x_0 \in \Omega with u(x_0) = M \geq u(y) for all y in some neighborhood of x_0. By the mean value property applied to a small ball around x_0, the integral average equals M, implying u \equiv M on that ball, and by analytic continuation (since harmonic functions are real analytic), u is constant on the connected component containing x_0, contradicting non-constancy. The minimum principle follows analogously by considering -u. The weak maximum principle provides a boundary-focused version: if u is harmonic on a bounded domain \Omega and continuous up to the boundary \partial \Omega, then the maximum and minimum values of u on \overline{\Omega} are attained on \partial \Omega. This extends to unbounded domains via limits at infinity, where if \limsup_{|x| \to \infty} u(x) \leq M and boundary limits satisfy similar conditions, then u \leq M on \Omega. The proof again relies on the mean value property and a contradiction argument: assuming an interior maximum leads to constancy, which would violate boundary data unless the maximum is on the boundary. A key application of the is the uniqueness of solutions to the for on a bounded domain \Omega: given continuous boundary data g on \partial \Omega, there is at most one u on \Omega that extends continuously to \overline{\Omega} with u|_{\partial \Omega} = g. If two such solutions u_1 and u_2 exist, then v = u_1 - u_2 is harmonic with zero boundary values, so by the weak , v \equiv 0 on \Omega. For unbounded domains, the Phragmén–Lindelöf principle extends this idea, providing growth controls on harmonic functions; for instance, in a \Omega = \mathbb{R} \times (-1,1)^{n-1}, if a s satisfies boundary growth conditions, then s is bounded by a harmonic majorant. The maximum principle extends to subharmonic functions, defined as upper semicontinuous functions s on \Omega satisfying the submean value property: s(x) \leq \frac{1}{|S(x, r)|} \int_{S(x, r)} s(y) \, d\sigma(y) for all balls B(x, r) \subset \Omega. For non-constant subharmonic s on a connected \Omega, s attains no interior maximum, and on a compact set, the maximum is on the boundary. The proof mirrors the harmonic case, using the submean property to derive a contradiction at an assumed interior maximum, yielding local constancy and global constancy on connected sets. This extension is crucial in potential theory, where analogous minimum principles apply to superharmonic functions, such as Newtonian potentials of positive measures.

Liouville's and Harnack's Theorems

Liouville's theorem states that every bounded harmonic function on \mathbb{R}^n is constant. The proof relies on the mean value property: for a bounded harmonic function u, the average over the ball B_r(0) equals u(0), and as r \to \infty, the difference |u(x) - u(0)| tends to zero for any fixed x, implying constancy. This result extends to positive harmonic functions on \mathbb{R}^n, which must also be constant, following from the volume mean value property and the assumption of non-negativity. A related extension, known as the strong Liouville theorem, applies to real-valued functions u on \mathbb{R}^n satisfying \liminf_{|x| \to \infty} u(x)/|x| \geq 0, concluding that u is constant. These theorems provide global boundedness constraints essential in potential theory for establishing uniqueness of solutions to boundary value problems on unbounded domains. quantifies the oscillation of positive functions. For a positive u on the B(a, R) \subset \mathbb{R}^n, it holds that \frac{R - r}{R + r} u(a) \leq u(x) \leq \frac{R + r}{R - r} u(a) for all x \in B(a, r) with $0 < r < R. The proof uses the Poisson integral representation and the maximum principle to bound the function's growth within concentric balls. More generally, in a connected domain \Omega with compact K \subset \Omega, there exists C > 1 depending on \Omega and K such that \frac{1}{C} u(x) \leq u(y) \leq C u(x) for all x, y \in K and positive u on \Omega. The constants depend on the dimension n and the ratio r/R. Harnack's theorem, or Harnack's principle, addresses of of functions. If \{u_m\} is a increasing of functions on a connected \Omega that converges pointwise to a finite u on \Omega, then u is , and the is uniform on compact subsets of \Omega. The proof invokes to control the 's behavior and ensure the limit satisfies the mean value property. If the pointwise is everywhere, the diverges uniformly to on compacts. These theorems find key applications in potential theory, particularly for the of solutions to the and in value problems. For instance, Harnack's principle ensures that the Perron solution—constructed as the supremum of subharmonic functions below given data—is in the interior, yielding continuous potentials when the boundary data is continuous. This property underpins the solvability of the via the Perron method. Higher-order versions of extend to derivatives of s. For a positive u on a in \mathbb{R}^n, there exist constants C depending on n such that |\nabla u(y)| \leq C u(x) / d(x,y) for points x, y in a , providing bounds relative to the . These estimates, originally derived using maximum principles on manifolds but applicable in , control the growth of s and are crucial for regularity in potential-theoretic applications like estimates.

Function Spaces

Classical Spaces of Harmonic Functions

Harmonic functions on an open domain \Omega \subset \mathbb{R}^n form a real , as the sum of two functions and any scalar multiple of a function satisfy \Delta u = 0 due to its linearity. This space is infinite-dimensional when \Omega is unbounded, such as \mathbb{R}^n for n \geq 2, because it contains homogeneous polynomials of arbitrarily high degree, including of all orders. The collection of harmonic functions belonging to L^p(\Omega) for $1 \leq p \leq \infty constitutes a of L^p(\Omega), often denoted h^p(\Omega), where membership requires \Delta u = 0 in the distributional sense alongside the integrability condition. functions exhibit enhanced regularity compared to general L^p functions; specifically, they lie in all Sobolev spaces W^{k,p}(\Omega) for k \geq 0 due to their C^\infty away from singularities, enabling Sobolev theorems to imply continuous or Hölder continuity on bounded subdomains when p > n/k. In two dimensions, harmonic Hardy spaces H^p consist of harmonic functions u on the unit disk such that the L^p-means over circles of radius r < 1 remain bounded as r \to 1^-, i.e., \sup_{0 < r < 1} \left( \frac{1}{2\pi} \int_0^{2\pi} |u(re^{i\theta})|^p \, d\theta \right)^{1/p} < \infty. These spaces connect closely to , as every in H^p on the disk is the real part of a in the analytic H^p, with the imaginary part recoverable via the . Spaces of harmonic functions with controlled growth at infinity, such as those satisfying |u(x)| = O(|x|^k) as |x| \to \infty for some fixed k \in \mathbb{N}, form finite-dimensional subspaces on complete Riemannian manifolds with non-negative . On , such functions are precisely the harmonic polynomials of degree at most k, with the dimension of this space being O(k^{n-1}), highlighting the role of growth conditions in classifying entire functions. Spherical harmonics provide an for expansions of harmonic functions; specifically, the functions Y_\ell^m(\theta, \phi) for \ell = 0, 1, 2, \dots and m = -\ell, \dots, \ell form a complete for L^2(S^{n-1}) with respect to the surface measure, allowing in in spherical coordinates. This basis is complete in the sense that any on the sphere admits a unique convergent expansion \sum_{\ell=0}^\infty \sum_{m=-\ell}^\ell a_{\ell m} Y_\ell^m, with coefficients given by inner products.

Modern Analytic Spaces

Modern analytic spaces in potential theory extend classical function spaces by incorporating specialized norms that capture the analytic and growth properties of harmonic functions, often forming Hilbert or Banach structures suited to operator-theoretic investigations and boundary behavior analysis. These spaces build upon precursors like the classical L^p spaces but emphasize semi-norms tailored to harmonicity, such as those controlling growth or energy integrals, enabling deeper insights into approximation and . The Bloch space consists of harmonic functions u on the unit disk \mathbb{D} \subset \mathbb{C} for which the satisfies \sup_{z \in \mathbb{D}} (1 - |z|^2) |\nabla u(z)| < \infty, measuring controlled "block-like" growth that prevents excessive expansion while allowing linear-type behavior at . This semi-norm, denoted \|u\|_B = \sup_{z \in \mathbb{D}} (1 - |z|^2) |\nabla u(z)|, endows the space with a Banach structure, where functions exhibit in a sense, crucial for studying radial limits and maximal functions in unbounded domains. The Bergman space comprises L^2 functions on a \Omega \subset \mathbb{R}^n, defined as the of polynomials under the L^2(\Omega) , forming a with a reproducing that facilitates point evaluations and integral representations via the Bergman . For the unit , the Bergman explicitly computes inner products and projections, enabling characterizations of boundedness for Toeplitz operators and decompositions essential for L^p estimates. The Dirichlet space includes harmonic functions u on a domain with finite Dirichlet integral \int_\Omega |\nabla u|^2 \, dA < \infty, representing minimal energy configurations in variational problems and forming a Hilbert space under the norm \|u\|_D^2 = |u(0)|^2 + \int_\Omega |\nabla u|^2 \, dA. This integral quantifies the "energy" of the function, linking directly to solutions of the Dirichlet problem and providing a framework for embedding theorems in potential theory. These spaces find applications in , where composition and multiplication operators on Bloch or Bergman spaces reveal spectral properties and compactness criteria through Carleson measure characterizations. In , Littlewood-Paley decompositions adapted to functions decompose them into dyadic blocks via Poisson integrals, yielding square-function estimates that bound L^p norms and facilitate multiplier theorems for . Connections to probabilistic models arise through Dirichlet forms, where the energy integral \int |\nabla u|^2 \, dA defines quadratic forms associated with symmetric Markov processes, enabling potential-theoretic representations of transition densities and hitting probabilities. In electrical networks, these forms correspond to effective resistances via Thomson's principle, interpreting functions as voltage potentials across graphs modeled by Dirichlet spaces.

References

  1. [1]
    [PDF] Potential theory - Department of Mathematics
    May 8, 2009 · Mathemat- ical potential theory originated in the work of Green, Gauss and others in the early 19th century, and has since that time served as a ...
  2. [2]
  3. [3]
    Potential Theory - an overview | ScienceDirect Topics
    Potential theory is defined as a mathematical framework that studies the properties of potentials associated with mass distributions, particularly through the s ...
  4. [4]
    Potential Theory in the Complex Plane
    Potential Theory in the Complex Plane. Search within full text. Access. Thomas Ransford, Université Laval, Québec.<|control11|><|separator|>
  5. [5]
    Foundations of Potential Theory - SpringerLink
    Book Title: Foundations of Potential Theory · Authors: Oliver Dimon Kellogg · Series Title: Grundlehren der mathematischen Wissenschaften · Publisher: Springer ...Missing: definition | Show results with:definition
  6. [6]
    Newton's Philosophiae Naturalis Principia Mathematica
    Dec 20, 2007 · In the first segment of the “Copernican Scholium” Newton identifies the center of gravity of the planetary system as the appropriate point to ...
  7. [7]
    June 1785: Coulomb Measures the Electric Force
    Jun 1, 2016 · Charles Augustin Coulomb (top) used a calibrated torsion balance (bottom) to measure the force between electric charges.Missing: potential | Show results with:potential
  8. [8]
    The Green of Green Functions - Physics Today
    George Green's essay introducing Green's theorem and Green functions was published at the author's expense in 1828. View larger. The customary route to ...
  9. [9]
    Siméon-Denis Poisson (1781 - 1840) - Biography - MacTutor
    Siméon-Denis Poisson worked on definite integrals and Fourier series. This was the foundation of later work in this area by Dirichlet and Riemann. Thumbnail of ...
  10. [10]
    8. The Image Method in Electrostatics - Galileo and Einstein
    Thomson realized its importance, and created the image technique as an easy way to construct Green's functions for some simple geometries, in particular a plane ...
  11. [11]
    STUDIES IN THE HISTORY OF COMPLEX FUNCTION THEORY II
    Riemann establishes that u and v are potential functions and that a conformai mapping of the x, j>-plane to the u, t>plane is effected via the analytic function ...
  12. [12]
    [PDF] Dirichlet's Principle - Penn Math
    Oct 16, 2014 · By 1840 it was known that if S ⊂ R is a closed and bounded set and f : S → R is a continuous function, then there are points p and q in S where ...
  13. [13]
    [PDF] FREDHOLM, HILBERT, SCHMIDT Three Fundamental Papers on ...
    Dec 15, 2011 · Hilbert's contribution to the theory of integral equations was not just a matter of introducing eigenvalues and eigenfunctions. After all ...
  14. [14]
    (PDF) Developments and perspectives in Nonlinear Potential Theory
    Feb 8, 2019 · Nonlinear Potential theory aims at replicating the classical linear potential theory when nonlinear equations are considered.
  15. [15]
    [PDF] Notes on Partial Differential Equations John K. Hunter - UC Davis Math
    (1) Potential theory e.g. in the Newtonian theory of gravity, electrostatics, heat flow, and potential flows in fluid mechanics. (2) Riemannian geometry ...
  16. [16]
    [PDF] Harmonic Function Theory - Sheldon Axler
    Harmonic functions also have a mean-value property with respect to volume measure. The polar coordinates formula for integration on Rn is indispensable here ...
  17. [17]
    [PDF] Chapter 2: Laplace's equation - UC Davis Math
    A solution of. Laplace's equation is called a harmonic function. Laplace's equation is a linear, scalar equation. It is the prototype of an elliptic partial ...
  18. [18]
    [PDF] 3 Laplace's Equation
    We say a function u satisfying Laplace's equation is a harmonic function. 3.1 The Fundamental Solution. Consider Laplace's equation in Rn,. ∆u = 0 x ∈ Rn ...
  19. [19]
    5.15: Poisson's and Laplace's Equations - Engineering LibreTexts
    Sep 12, 2022 · Laplace's Equation (Equation. ) states that the Laplacian of the electric potential field is zero in a source-free region. Like Poisson's ...Missing: seminal papers
  20. [20]
    Gravitational Potential
    where the integral is over all space. , is known as Poisson's equation. Of course, Equation (1.272) is the integral form of Poisson's equation.
  21. [21]
    Green's functions - Richard Fitzpatrick
    The Green's function $G({\bf r}, {\bf r}')$ is the potential, which satisfies the appropriate boundary conditions, generated by a unit amplitude point source.
  22. [22]
    Green's Function--Poisson's Equation -- from Wolfram MathWorld
    Poisson's equation is del ^2phi=4pirho, where phi is often called a potential function and rho a density function, so the differential operator in this case is ...
  23. [23]
    [PDF] Uniqueness of solutions to the Laplace and Poisson equations
    For the case of Dirichlet boundary conditions or mixed boundary conditions, the solution to Poisson's equation always exists and is unique.
  24. [24]
  25. [25]
    [PDF] Harmonic analysis on spheres, I
    Feb 28, 2011 · of rotation groups SO(n). Spheres themselves are rarely groups, but are acted-upon transitively by rotation groups. An essential simplicity is ...
  26. [26]
    [PDF] Lecture 17: Conformal Invariance - MIT Mathematics
    Nov 7, 2006 · The Laplace's equation is conformally invariant. So if we need to solve it in a domain with complicated geometry, we can apply a conformal ...
  27. [27]
    The Kelvin transformation in potential theory and Stokes flow - ADS
    This property, which was the content of a letter sent by Kelvin to Liouville in 1845, provides a powerful machinery for solving particular potential problems ...Missing: inversion historical origin
  28. [28]
    Kelvin transformation - Encyclopedia of Mathematics
    Jul 25, 2021 · Because of this property, the Kelvin transformation enables one to reduce exterior problems in potential theory to interior ones and vice versa ...Missing: historical | Show results with:historical
  29. [29]
    m - American Mathematical Society
    Foundations of Potential Theory. By O. D. Kellogg. Berlin, Springer, 1929 ... in order to satisfy the requirements of the definition of regular surface.
  30. [30]
    [PDF] On Kelvin Transformation
    Jun 7, 2006 · We employ standard theory of resolvents ... (to regular harmonic functions) of the fact that Kelvin transform preserves harmonic functions.
  31. [31]
    The Kelvin Transform | SpringerLink
    The Kelvin transform, similar to f(z) ↦ f(1/z), transforms a function harmonic inside the unit sphere into a function harmonic outside the sphere.Missing: reference | Show results with:reference
  32. [32]
    [PDF] Kelvin transformation and inverse multipoles in electrostatics - arXiv
    also known as inversion in the sphere — which, among other things ...
  33. [33]
  34. [34]
    Method of Images
    As a second example of the method of images, consider a grounded conducting sphere of radius $a$ centered on the origin. Suppose that a point electric ...
  35. [35]
    [PDF] METHOD OF IMAGES
    We make the voltage 0 on the plane and on the half sphere at ∞. This is called “grounding.” Physically “grounding” means connecting the plane to a very large ...
  36. [36]
    [PDF] LECTURE NOTES 6 - High Energy Physics
    a. Example #3: Point charge +q near a charged conducting sphere of radius R. (variation on image charge Example #2). → Use the superposition principle for image ...
  37. [37]
    [PDF] Today in Physics 217: the method of images
    Oct 9, 2002 · method of images. ❑ Caveats. ❑ Example: a point charge and a grounded conducting sphere. ❑ Multiple images y x b b b b a a q q. -q. -q a a ...
  38. [38]
    [PDF] Dielectric (and Magnetic) Image Methods - Kirk T. McDonald
    The method of images is most often employed in electrostatic examples with point or line charges in vacuum outside conducting planes, cylinders or spheres ...<|separator|>
  39. [39]
    [PDF] 5 Potential Theory
    As usual, let Φ(x) denote the fundamental solution of Laplace's equation. That is, let. Φ(x) ≡. ½. − 1. 2π ln|x| n = 2. 1 n(n−2)α(n). · 1. |x|n−2 n ≥ 3. 1. Page ...<|control11|><|separator|>
  40. [40]
    None
    Summary of each segment:
  41. [41]
    [PDF] Logarithmic Potential Theory with Applications to Approximation ...
    Oct 12, 2010 · Abstract. We provide an introduction to logarithmic potential theory in the complex plane that par- ticularly emphasizes its usefulness in ...
  42. [42]
    [PDF] A concise course in complex analysis and Riemann surfaces ...
    THE POTENTIAL THEORY PROOF OF THE RIEMANN MAPPING THEOREM. 147. 2. The potential theory proof of the Riemann mapping theorem. This discussion raises the ...
  43. [43]
    [PDF] Introduction to Potential Theory via Applications - Christian Kuehn
    We introduce the basic concepts related to subharmonic functions and potentials, mainly for the case of the complex plane and prove the Riesz decomposition ...
  44. [44]
    [PDF] Harmonic functions - Purdue Math
    Mar 18, 2024 · In any simply connected region in the plane, every harmonic function is the real part of an analytic function f. This f is defined up to.Missing: theory | Show results with:theory
  45. [45]
    [PDF] 18.04 S18 Topic 5: Introduction to harmonic functions
    To complete the tight connection between analytic and harmonic functions we show that any har- monic function is the real part of an analytic function. Theorem ...
  46. [46]
    [PDF] 4. Introducing Conformal Field Theory - DAMTP
    The purpose of this section is to get comfortable with the basic language of two dimen- sional conformal field theory4. This is a topic which has many ...
  47. [47]
    [PDF] Lectures on Conformal Field Theories - DAMTP
    In two dimensions there is the infinite dimensional Virasoro algebra whereas in higher dimensions the conformal group is finite dimensional.
  48. [48]
    [PDF] Preliminaries - Assets - Cambridge University Press
    The argument principle for harmonic functions can now be formulated as a direct generalization of the classical result for analytic functions. Theorem. Let f be ...
  49. [49]
    Harmonic Mappings of Multiply Connected Domains - MSP
    The second step is then to apply the argument principle for harmonic functions to conclude that f is globally univalent. The third step is to show that f cannot ...
  50. [50]
    Classical Potential Theory - Book - SpringerLink
    This book provides a comprehensive treatment of classical potential theory: it covers harmonic and subharmonic functions, maximum principles, polynomial
  51. [51]
    [PDF] A TUTORIAL ON CONFORMAL GROUPS - Biblioteka Nauki
    Moreover all such Möbius maps are conformal. This is the Theorem of Liouville. The obvious analogue of this theorem then holds for positive-definite quadratic ...
  52. [52]
    Elliptic Partial Differential Equations of Second Order - SpringerLink
    This is a book of interest to any having to work with differential equations, either as a reference or as a book to learn from.
  53. [53]
    [PDF] Bôcher's Theorem Michael Taylor Contents 1. Introduction 2. Proof ...
    Bôcher's theorem states that if u satisfies (1.1), then u(x) = AV(x) + h(x), where h is harmonic on O and A is a constant.
  54. [54]
    [PDF] Bocher's Theorem - Sheldon Axler
    Sep 6, 2015 · INTRODUCTION. Bocher's Theorem characterizes the behavior of positive har- monic functions in the neighborhood of an isolated singularity.
  55. [55]
    [PDF] Classical Potential Theory
    Armitage,DavidH. Classical potential theory. - (Springf:r monographs in mathematics). 1. Potential theory (Mathematics) 2.
  56. [56]
    ap.analysis of pdes - Gradient estimates - MathOverflow
    Feb 19, 2021 · Gradient estimates (and especially the differential Harnack) for harmonic functions on Riemannian manifolds were proved by Cheng and Yau in 1975.Missing: source | Show results with:source
  57. [57]
    [PDF] Stability of the elliptic Harnack inequality - Annals of Mathematics
    If the EHI holds, then iterating condition (1.6) gives a.e. Hölder continuity of harmonic functions, and it follows that any harmonic function has a continuous.Missing: source | Show results with:source
  58. [58]
    DIMENSION ESTIMATE OF POLYNOMIAL GROWTH HARMONIC ...
    In particular, this implies that the space of harmonic forms of fixed growth order on the Euclidean space with any periodic metric must be finite dimensional.
  59. [59]
    [PDF] Radial Variation of Positive Harmonic Functions and Bloch Functions
    The first one is the variation of positive harmonic functions and the second one the radial variation of Bloch functions. A theorem on the variation of positive ...<|control11|><|separator|>
  60. [60]
    [PDF] Harmonic Bergman Spaces - The Library at SLMath
    We present a simple derivation of the explicit formula for the harmonic Bergman reproducing kernel on the ball in euclidean space and give an elementary proof ...Missing: potential | Show results with:potential
  61. [61]
    [PDF] The classical Dirichlet space - University of Richmond
    If u ∈ Dh, the harmonic Dirichlet space, then u = P(k ∗ g) for some g ∈ L2 and moreover, the radial boundary function for u is k ∗ g almost everywhere. This.
  62. [62]
    The Littlewood-Paley inequalities for Hardy-Orlicz spaces of ...
    Littlewood-Paley inequalities are also known to be valid for harmonic functions in the unit ball in JRn. In fact Stevie [7] h¥ recently proved that for n 2: 3, ...
  63. [63]
    [PDF] Random walks and electric networks - Dartmouth Mathematics
    We shall show, in fact, that the Dirichlet problem has a natural generalization in the context of absorbing Markov chains and can be solved by Markov chain ...<|separator|>