Fact-checked by Grok 2 weeks ago

Laplace operator

The Laplace operator, commonly denoted by \Delta or \nabla^2, is a second-order linear partial differential operator applied to scalar functions, defined as the divergence of the gradient: \Delta f = \nabla \cdot (\nabla f). In three-dimensional Euclidean space with Cartesian coordinates (x, y, z), it takes the explicit form \Delta f = \frac{\partial^2 f}{\partial x^2} + \frac{\partial^2 f}{\partial y^2} + \frac{\partial^2 f}{\partial z^2}. This operator quantifies the spatial variation of f at a point, effectively measuring how much the function deviates from its local average value in the surrounding region. Geometrically, the Laplacian captures the "concavity" or "convexity" of a function relative to its neighborhood, with \Delta f > 0 indicating local minima (where f is below its average), \Delta f < 0 indicating local maxima (where f is above its average), and \Delta f = 0 signifying that f equals its local average—a property central to harmonic functions, which are solutions to Laplace's equation \Delta u = 0. These functions exhibit the mean value property, where the value at any interior point equals the average over any surrounding sphere, and they are infinitely differentiable (real analytic in \mathbb{R}^n). The operator is rotationally invariant and elliptic, ensuring well-posed boundary value problems under suitable conditions, such as the Dirichlet or Neumann problems. Named after the French mathematician and astronomer Pierre-Simon Laplace (1749–1827), who introduced it in the context of celestial mechanics and potential theory in the late 18th century, the Laplacian has become indispensable across disciplines. In physics, it governs steady-state phenomena without sources: Laplace's equation describes electrostatic potentials in charge-free regions, gravitational potentials, and irrotational fluid flows. It also appears in the heat equation \frac{\partial u}{\partial t} = \kappa \Delta u for diffusion processes, the time-independent Schrödinger equation in quantum mechanics, and the wave equation for vibrations. More generally, Poisson's equation \Delta u = -f extends it to scenarios with sources, such as mass distributions in gravity or charge densities in electromagnetism. Beyond Euclidean space, the Laplacian generalizes to the Laplace–Beltrami operator on Riemannian manifolds, defined intrinsically via the metric tensor as \Delta u = -\operatorname{div}(\nabla u), preserving key properties like self-adjointness and enabling spectral analysis through eigenvalues and eigenfunctions. Discrete analogs, such as graph Laplacians, extend its utility to computational geometry, machine learning, and network theory, where they model diffusion and connectivity. Its eigenvalues relate to manifold geometry, influencing theorems like the Chern–Gauss–Bonnet formula and heat kernel estimates.

Definition and Foundations

Formal Definition

The Laplace operator, often denoted by \Delta, is a second-order partial differential operator acting on scalar functions in Euclidean space \mathbb{R}^n. For a scalar function f: \mathbb{R}^n \to \mathbb{R}, it is defined as the divergence of the gradient of f, that is, \Delta f = \operatorname{div}(\nabla f), where \nabla f = \left( \frac{\partial f}{\partial x_1}, \dots, \frac{\partial f}{\partial x_n} \right) is the gradient vector and \operatorname{div} denotes the divergence operator. This definition captures the operator's intrinsic nature independent of coordinates. In Cartesian coordinates, the Laplace operator takes the explicit form \Delta f = \sum_{i=1}^n \frac{\partial^2 f}{\partial x_i^2}, which is the sum of the second-order pure partial derivatives of f with respect to each coordinate x_i. The operator is typically applied to sufficiently smooth functions, specifically those in the space C^2(\mathbb{R}^n) of twice continuously differentiable functions, ensuring the existence of all required second derivatives. On more general domains \Omega \subseteq \mathbb{R}^n, the domain may be restricted to C^2(\overline{\Omega}) for boundary value problems. The notation \Delta derives from the French mathematician Pierre-Simon Laplace (1749–1827), who introduced the operator in his 1785 memoir on the attractions of spheroids, applying it to problems in celestial mechanics and gravitational potentials. In physics and engineering contexts, it is alternatively denoted by \nabla^2 to highlight its composition as the del operator \nabla applied twice, though both symbols refer to the same mathematical object. Geometrically, \Delta f equals the trace of the Hessian matrix of second partial derivatives of f.

Analytic versus Geometric Laplacian

The analytic Laplacian refers to the formulation of the operator in a specific coordinate system, expressed as the sum of the second-order partial derivatives of a scalar function with respect to each coordinate variable. This coordinate-dependent expression, Δf = ∑_{i=1}^n ∂²f/∂x_i² in Cartesian coordinates on Euclidean space, arose in 18th-century mathematical analysis, notably through Pierre-Simon Laplace's applications to celestial mechanics and potential theory in works such as his Mécanique Céleste (1799–1825). In contrast, the geometric Laplacian provides an intrinsic definition independent of any chosen coordinates, formulated as the divergence of the gradient of the function on a Riemannian manifold, Δf = div(grad f). This approach, which preserves the operator's meaning under coordinate transformations, developed in 19th-century differential geometry, building on Carl Friedrich Gauss's foundational work on the intrinsic geometry of surfaces (1827) and Bernhard Riemann's generalization to higher-dimensional manifolds in his 1854 habilitation lecture. The two perspectives align precisely in flat Euclidean space, where the standard coordinate expression matches the intrinsic divergence-gradient form due to the constant metric. On curved spaces, however, the analytic expression in non-Cartesian coordinates incorporates metric-dependent terms, differing from the coordinate-free geometric version while both capture the same underlying operator.

Motivations and Interpretations

Diffusion and Heat Equation

The Laplace operator plays a central role in modeling diffusion processes, particularly in the context of heat conduction. The heat equation, which governs the evolution of temperature u(\mathbf{x}, t) in a medium, is given by \frac{\partial u}{\partial t} = \kappa \Delta u, where \kappa > 0 is the thermal diffusivity constant, and \Delta is the Laplace operator. This equation arises from fundamental physical principles and describes how heat diffuses through a material over time. The derivation of the heat equation begins with the conservation of energy in a small volume element. Consider a region in space; the rate of change of thermal energy within it equals the net heat flux across its boundary. Fourier's law of heat conduction states that the heat flux \mathbf{q} is proportional to the negative gradient of the temperature: \mathbf{q} = -k \nabla u, where k > 0 is the thermal conductivity. Applying the divergence theorem to the flux, the net heat inflow per unit volume is -\nabla \cdot \mathbf{q} = k \Delta u. Assuming constant density \rho > 0 and specific heat capacity c > 0, the thermal energy density is \rho c u, so conservation of energy yields \rho c \frac{\partial u}{\partial t} = k \Delta u, or equivalently \frac{\partial u}{\partial t} = \kappa \Delta u where \kappa = \frac{k}{\rho c}. This formulation captures the diffusive nature of heat transfer, where \Delta u represents the net rate of heat flow into the region: if \Delta u > 0, heat enters faster than it leaves, causing the local temperature to rise and smoothing out irregularities in the temperature distribution over time. In the steady-state regime, where the temperature no longer changes with time (\frac{\partial u}{\partial t} = 0), the heat equation simplifies to \Delta u = 0. Solutions to this equation are harmonic functions, which describe equilibrium temperature distributions with no net heat flow. These steady states are reached as t \to \infty, where initial temperature variations have fully diffused. A classic physical example is heat conduction along a thin metal rod of length L, insulated on the sides, with prescribed temperatures at the ends, say u(0, t) = T_1 and u(L, t) = T_2. The temperature u(x, t) satisfies the one-dimensional heat equation \frac{\partial u}{\partial t} = \kappa \frac{\partial^2 u}{\partial x^2}, with \frac{\partial^2 u}{\partial x^2} being the one-dimensional Laplacian. Initially nonuniform temperatures evolve to smooth out, approaching the linear steady-state solution u(x) = T_1 + \frac{T_2 - T_1}{L} x, which satisfies \frac{d^2 u}{dx^2} = 0. This illustrates how boundary conditions dictate the final equilibrium, with the Laplacian enforcing zero net flux in steady state.

Averaging and Harmonic Functions

A harmonic function u on a domain \Omega \subseteq \mathbb{R}^n is a twice continuously differentiable function satisfying \Delta u = 0 in \Omega. Such functions possess the mean value property, which states that for any point x \in \Omega and any radius r > 0 such that the ball B(x, r) \subset \Omega, the value of u at the center equals the average value over the ball or its boundary sphere: u(x) = \frac{1}{|B(x, r)|} \int_{B(x, r)} u(y) \, dy = \frac{1}{|\partial B(x, r)|} \int_{\partial B(x, r)} u(y) \, d\sigma(y), where |B(x, r)| and |\partial B(x, r)| denote the volume of the ball and the surface area of the sphere, respectively. To establish this property, consider the proof using Green's second identity applied to u and a suitable auxiliary function. For n \geq 3, take the fundamental solution \Phi(y) = c_n |y|^{2-n} for the Laplacian, where c_n is a dimensional constant, and apply Green's identity over the ball B(x, r): \int_{B(x, r)} (u \Delta \Phi - \Phi \Delta u) \, dy = \int_{\partial B(x, r)} \left( u \frac{\partial \Phi}{\partial n} - \Phi \frac{\partial u}{\partial n} \right) d\sigma. Since \Delta u = 0 and \Delta \Phi = 0 away from the origin, shifting coordinates so x = 0, the volume integral vanishes except at the singularity, but careful limiting as the radius approaches zero shows the boundary term simplifies to yield the spherical average equal to u(0). The ball average follows by integrating the spherical property radially. For n=2, a logarithmic fundamental solution is used analogously. The mean value property fully characterizes harmonic functions among sufficiently regular classes. Specifically, a continuous function u: \Omega \to \mathbb{R} is harmonic if and only if it satisfies the mean value property over every ball B(x, r) \subset \Omega with x \in \Omega and r > 0. The forward direction holds as shown, while the converse follows by showing that the property implies \Delta u = 0 in the sense of distributions or via differentiation under the integral sign to recover the PDE. This property underpins important qualitative behaviors of harmonic functions, notably the maximum principle. In a bounded domain \Omega with continuous boundary values, a non-constant harmonic function u cannot attain its maximum or minimum in the interior of \Omega; extrema must occur on \partial \Omega. To see this, suppose u achieves a maximum at an interior point x_0; then by the mean value property over a small ball around x_0, u(x_0) equals the average, implying u is constant on that ball, and by connectedness, throughout \Omega. A similar argument applies to minima, with the strong maximum principle ensuring the maximum is not achieved interiorly unless constant.

Potential Theory and Electrostatics

In classical potential theory, the Laplace operator governs the relationship between potentials and their sources, originating from efforts to model gravitational fields in celestial mechanics. Pierre-Simon de Laplace introduced the inhomogeneous form of the equation in his Traité de Mécanique Céleste (1799–1825), where the gravitational potential \phi due to a mass density \rho satisfies Poisson's equation \nabla^2 \phi = 4\pi G \rho, with G denoting the gravitational constant; this formulation enabled precise calculations of planetary perturbations and orbital stability. The same mathematical structure applies to electrostatics, where Poisson's equation describes the electrostatic potential \phi arising from a charge density \rho: \nabla^2 \phi = -\frac{\rho}{\epsilon_0}, with \epsilon_0 the vacuum permittivity; this equation follows from Gauss's law and the definition of the electric field as \mathbf{E} = -\nabla \phi. In regions devoid of sources where \rho = 0, the equation simplifies to Laplace's equation \nabla^2 \phi = 0, whose solutions yield the potential fields in charge-free or mass-free domains. A key tool for solving these equations is the fundamental solution, or Green's function, for the Laplace operator in \mathbb{R}^3, given by G(\mathbf{x}, \mathbf{y}) = -\frac{1}{4\pi |\mathbf{x} - \mathbf{y}|}, which satisfies \nabla^2 G(\mathbf{x}, \mathbf{y}) = \delta(\mathbf{x} - \mathbf{y}), where \delta is the Dirac delta function. This function facilitates integral representations of solutions to Poisson's equation; for instance, in electrostatics, the potential can be expressed as \phi(\mathbf{x}) = \frac{1}{\epsilon_0} \int_{\mathbb{R}^3} \frac{\rho(\mathbf{y})}{4\pi |\mathbf{x} - \mathbf{y}|} \, dV_y + \text{surface integrals}, allowing the potential at any point to be computed directly from the source distribution, subject to appropriate boundary conditions.

Energy Minimization and Variational Methods

The Dirichlet energy functional provides a variational characterization of solutions to the Laplace equation. For a domain \Omega \subset \mathbb{R}^n with boundary data g on \partial \Omega, the functional is defined as E = \frac{1}{2} \int_\Omega |\nabla f|^2 \, dV over functions f satisfying f = g on \partial \Omega. Minimizers of this functional are harmonic functions, satisfying \Delta f = 0 in \Omega in the classical sense, assuming sufficient regularity. To derive this connection, consider the first variation of E. For a variation f + \epsilon \eta with \eta = 0 on \partial \Omega, the Euler-Lagrange equation arises from setting the derivative with respect to \epsilon at \epsilon = 0 to zero: \frac{d}{d\epsilon} E[f + \epsilon \eta] \bigg|_{\epsilon=0} = \int_\Omega \nabla f \cdot \nabla \eta \, dV = 0. Integration by parts yields -\int_\Omega \eta \Delta f \, dV = 0 for all admissible \eta, implying \Delta f = 0. This variational principle, known as Dirichlet's principle, dates to the 19th century and underpins modern numerical methods like finite elements. In the weak sense, minimizers exist in Sobolev spaces H^1(\Omega), where the space of admissible functions is \{ f \in H^1(\Omega) \mid f = g \text{ on } \partial \Omega \} (in the trace sense). The existence follows from the convexity and coercivity of E, ensuring a unique minimizer via direct methods in the calculus of variations; this weak solution satisfies \Delta f = 0 in the distributional sense and coincides with the classical solution where smooth. Physically, this minimization interprets the Laplace equation as an equilibrium state. In electrostatics, the potential \phi minimizes the energy functional \frac{\epsilon_0}{2} \int_\Omega |\nabla \phi|^2 \, dV - \int_\Omega \rho \phi \, dV in charge-free regions (\rho = 0), yielding \Delta \phi = 0 via the Euler-Lagrange equation. In linear elasticity, analogous variational principles minimize strain energy functionals, leading to equilibrium equations that involve the vector Laplacian of the displacement field, such as Navier's equation -\mu \Delta \mathbf{u} - (\lambda + \mu) \nabla (\nabla \cdot \mathbf{u}) = \mathbf{f}.

Coordinate Representations

In Two Dimensions

In two dimensions, the Laplace operator applied to a twice-differentiable function f(x, y) takes the form \Delta f = \frac{\partial^2 f}{\partial x^2} + \frac{\partial^2 f}{\partial y^2} in Cartesian coordinates (x, y). This expression arises as the divergence of the gradient in the Euclidean plane, providing a measure of the local variation or "curvature" of the function. To express the Laplacian in polar coordinates (r, \theta), where x = r \cos \theta and y = r \sin \theta, the chain rule is applied to transform the partial derivatives. First, the first-order partials are \frac{\partial f}{\partial r} = \cos \theta \frac{\partial f}{\partial x} + \sin \theta \frac{\partial f}{\partial y}, \quad \frac{\partial f}{\partial \theta} = -r \sin \theta \frac{\partial f}{\partial x} + r \cos \theta \frac{\partial f}{\partial y}. Differentiating again yields the second partials, such as \frac{\partial^2 f}{\partial r^2} = \cos^2 \theta \frac{\partial^2 f}{\partial x^2} + 2 \cos \theta \sin \theta \frac{\partial^2 f}{\partial x \partial y} + \sin^2 \theta \frac{\partial^2 f}{\partial y^2}, with analogous expressions for the other terms involving \theta-derivatives. Substituting into \Delta f and simplifying using trigonometric identities like \cos^2 \theta + \sin^2 \theta = 1 and collecting like terms results in the polar form \Delta f = \frac{\partial^2 f}{\partial r^2} + \frac{1}{r} \frac{\partial f}{\partial r} + \frac{1}{r^2} \frac{\partial^2 f}{\partial \theta^2}.[28] This coordinate system is particularly useful for problems with radial or angular symmetry, as it separates radial and angular dependencies more naturally. A simple class of examples illustrating the Laplacian in two dimensions consists of harmonic polynomials, which satisfy \Delta f = 0. These include the real parts of powers of the complex variable z = x + i y. For instance, \operatorname{Re}(z^n) = \operatorname{Re}((x + i y)^n) is a homogeneous harmonic polynomial of degree n. Expanding via the binomial theorem and taking the real part yields explicit forms, such as \operatorname{Re}(z^1) = x and \operatorname{Re}(z^2) = x^2 - y^2, both of which verify \Delta (\operatorname{Re}(z^n)) = 0 by direct computation. Such polynomials form a basis for harmonic functions in the plane and arise in multipole expansions or conformal mappings. Boundary value problems for the Laplacian in two dimensions, such as the Dirichlet problem on a disk, can be solved using separation of variables in polar coordinates. Consider \Delta u = 0 inside the unit disk r < 1 with boundary condition u(1, \theta) = f(\theta). Assume a product solution u(r, \theta) = R(r) T(\theta), leading to the separated equations T'' + \lambda T = 0 and r^2 R'' + r R' - \lambda R = 0. The periodic boundary in \theta implies eigenvalues \lambda = n^2 for n = 0, 1, 2, \dots, with eigenfunctions T_n(\theta) = a_n \cos(n \theta) + b_n \sin(n \theta). The radial equation then gives R_n(r) = A_n r^n (discarding the singular r^{-n} term for boundedness at r = 0). The general solution is the Fourier series u(r, \theta) = \frac{a_0}{2} + \sum_{n=1}^\infty r^n (a_n \cos(n \theta) + b_n \sin(n \theta)), where the coefficients are determined by the boundary data via a_n = \frac{1}{\pi} \int_{-\pi}^\pi f(\theta) \cos(n \theta) \, d\theta, \quad b_n = \frac{1}{\pi} \int_{-\pi}^\pi f(\theta) \sin(n \theta) \, d\theta for n \geq 1, and similarly for a_0. This approach yields the unique solution by the maximum principle for harmonic functions.

In Three Dimensions

In three-dimensional Euclidean space with Cartesian coordinates (x, y, z), the Laplace operator applied to a scalar function f is expressed as the sum of the second partial derivatives: \Delta f = \frac{\partial^2 f}{\partial x^2} + \frac{\partial^2 f}{\partial y^2} + \frac{\partial^2 f}{\partial z^2}. This form follows directly from the definition of the Laplacian as the divergence of the gradient in orthogonal curvilinear coordinates, specialized to the unit metric of Cartesian systems. For problems exhibiting rotational symmetry, such as those in spherical or cylindrical geometries, the Laplacian is transformed using the chain rule and the scale factors of the coordinate system. In cylindrical coordinates (\rho, \phi, z), where \rho = \sqrt{x^2 + y^2}, \phi = \arctan(y/x), and z unchanged, the Laplacian becomes \Delta f = \frac{1}{\rho} \frac{\partial}{\partial \rho} \left( \rho \frac{\partial f}{\partial \rho} \right) + \frac{1}{\rho^2} \frac{\partial^2 f}{\partial \phi^2} + \frac{\partial^2 f}{\partial z^2}. This expression arises from applying the general formula for the Laplacian in orthogonal curvilinear coordinates, incorporating the scale factors h_\rho = 1, h_\phi = \rho, and h_z = 1. In spherical coordinates (r, \theta, \phi), defined by r = \sqrt{x^2 + y^2 + z^2}, \theta = \arccos(z/r), and \phi = \arctan(y/x), the Laplacian takes the form \Delta f = \frac{1}{r^2} \frac{\partial}{\partial r} \left( r^2 \frac{\partial f}{\partial r} \right) + \frac{1}{r^2 \sin \theta} \frac{\partial}{\partial \theta} \left( \sin \theta \frac{\partial f}{\partial \theta} \right) + \frac{1}{r^2 \sin^2 \theta} \frac{\partial^2 f}{\partial \phi^2}. To derive this, start with the general expression for the Laplacian in orthogonal curvilinear coordinates (u^1, u^2, u^3) with scale factors h_i = |\partial \mathbf{r}/\partial u^i|: \Delta f = \frac{1}{h_1 h_2 h_3} \sum_{i=1}^3 \frac{\partial}{\partial u^i} \left( \frac{h_1 h_2 h_3}{h_i^2} \frac{\partial f}{\partial u^i} \right), where \mathbf{r}(u^1, u^2, u^3) is the position vector. For spherical coordinates, the scale factors are h_r = 1, h_\theta = r, and h_\phi = r \sin \theta, with the Jacobian volume element h_r h_\theta h_\phi = r^2 \sin \theta. Substituting these yields the radial term \frac{1}{r^2} \frac{\partial}{\partial r} (r^2 \frac{\partial f}{\partial r}), the polar term \frac{1}{r^2 \sin \theta} \frac{\partial}{\partial \theta} (\sin \theta \frac{\partial f}{\partial \theta}), and the azimuthal term \frac{1}{r^2 \sin^2 \theta} \frac{\partial^2 f}{\partial \phi^2}. This derivation relies on expressing the Cartesian Laplacian in terms of spherical variables via the chain rule, though the scale factor approach is more direct and avoids lengthy expansions of second derivatives. A classic application in three dimensions involves solving Poisson's equation \Delta \phi = 4\pi G \rho for the gravitational potential \phi inside a uniform sphere of radius R and constant density \rho, where G is the gravitational constant. Assuming spherical symmetry, \phi = \phi(r) depends only on the radial distance r, so the equation simplifies using the spherical Laplacian to \frac{1}{r^2} \frac{d}{dr} (r^2 \frac{d\phi}{dr}) = 4\pi G \rho for r < R. Integrating once gives r^2 \frac{d\phi}{dr} = \frac{4\pi G \rho}{3} r^3 + C_1; the constant C_1 = 0 by regularity at r = 0. Integrating again yields \phi(r) = \frac{2\pi G \rho}{3} (r^2 - 3R^2) + C_2 for r < R, where C_2 is determined by matching to the exterior solution \phi(r) = -GM/r (with total mass M = \frac{4\pi}{3} \rho R^3) at r = R, resulting in \phi(r) = -\frac{GM}{2R^3} (3R^2 - r^2) inside the sphere. This quadratic potential illustrates how the Laplacian captures the source distribution in symmetric geometries.

In N Dimensions

In Euclidean space \mathbb{R}^n, the Laplace operator acting on a scalar function f is defined in Cartesian coordinates as the sum of the second partial derivatives with respect to each coordinate: \Delta f = \sum_{i=1}^n \frac{\partial^2 f}{\partial x_i^2}. For radially symmetric functions f(r), where r = |x| is the Euclidean norm, the Laplacian simplifies to a one-dimensional operator: \Delta f = \frac{d^2 f}{dr^2} + \frac{n-1}{r} \frac{df}{dr}. This form arises from the divergence theorem applied to the radial gradient and highlights the influence of dimensionality on the operator's structure. In hyperspherical coordinates, which generalize spherical coordinates to n dimensions, the Laplacian is expressed using the radial coordinate r and angular variables \phi_2, \dots, \phi_n: \Delta = \frac{\partial^2}{\partial r^2} + \frac{n-1}{r} \frac{\partial}{\partial r} + \frac{1}{r^2} \Lambda_n, where \Lambda_n is the angular part, defined recursively as \Lambda_n = \frac{1}{\sin^2 \phi_n} \Lambda_{n-1} + (n-2) \cot \phi_n \frac{\partial}{\partial \phi_n} + \frac{\partial^2}{\partial \phi_n^2}, with \Lambda_2 = \frac{\partial^2}{\partial \phi_2^2}. The scale factors in these coordinates—such as h_r = 1, h_{\phi_k} = r \prod_{j=1}^{k-1} \sin \phi_j for k = 2, \dots, n—ensure the operator accounts for the geometry of the n-sphere. The fundamental solution to Laplace's equation \Delta G = -\delta(x) in n \geq 3 dimensions exhibits asymptotic behavior G(x) \sim |x|^{2-n} as |x| \to 0, specifically G(x) = \frac{1}{(n-2)\omega_n |x|^{n-2}}, where \omega_n is the surface area of the unit sphere in \mathbb{R}^n. This singularity reflects the operator's scaling properties in high dimensions, influencing solutions to boundary value problems.

Core Properties

Euclidean Invariance

The Laplace operator exhibits Euclidean invariance, meaning it remains unchanged under translations and rotations in \mathbb{R}^n. This property arises from the operator's definition as the divergence of the gradient, which respects the flat metric and affine structure of Euclidean space. Translational invariance follows directly from the constant coefficients of the Laplacian. Consider a translation T_{\mathbf{a}} f(\mathbf{x}) = f(\mathbf{x} - \mathbf{a}) for \mathbf{a} \in \mathbb{R}^n. The first partial derivative satisfies \frac{\partial}{\partial x_i} (T_{\mathbf{a}} f)(\mathbf{x}) = (T_{\mathbf{a}} \frac{\partial f}{\partial x_i})(\mathbf{x}), since differentiation commutes with the shift operator due to the independence from position. Applying this to the second derivative yields \frac{\partial^2}{\partial x_i^2} (T_{\mathbf{a}} f) = T_{\mathbf{a}} \frac{\partial^2 f}{\partial x_i^2}, and summing over i gives \Delta (T_{\mathbf{a}} f) = T_{\mathbf{a}} (\Delta f). Thus, the operator commutes with translations, preserving solutions to equations like \Delta u = 0 or \Delta u = \lambda u. Rotational invariance holds for orthogonal transformations R \in O(n), where R^T R = I. For a function g = f \circ R, the chain rule implies that the partial derivatives transform as \frac{\partial g}{\partial x_i}(\mathbf{x}) = \sum_{k=1}^n r_{k i} \frac{\partial f}{\partial y_k}(R \mathbf{x}), with \mathbf{y} = R \mathbf{x}. The second derivatives become \frac{\partial^2 g}{\partial x_i^2}(\mathbf{x}) = \sum_{k,\ell=1}^n r_{k i} r_{\ell i} \frac{\partial^2 f}{\partial y_k \partial y_\ell}(R \mathbf{x}). Summing over i, the orthogonality condition \sum_i r_{k i} r_{\ell i} = \delta_{k \ell} simplifies the expression to \Delta g(\mathbf{x}) = \Delta f (R \mathbf{x}), or \Delta (f \circ R) = (\Delta f) \circ R. This confirms that rotations preserve the form of the Laplacian. A key consequence of rotational invariance is that eigenfunctions of the Laplacian can be chosen to respect spherical symmetry, particularly in separation-of-variables solutions on rotationally symmetric domains. For instance, in spherical coordinates, the eigenfunctions separate into radial components and angular spherical harmonics, where the angular parts form irreducible representations of the rotation group SO(n), enabling the decomposition of the eigenspaces into symmetry-adapted bases. This invariance under the Euclidean group underpins the isotropy of physical laws modeled by the Laplacian, such as in the diffusion equation \partial_t u = \Delta u, where the operator ensures uniform spreading in all directions without preferred orientations, aligning with empirical observations of isotropic media.

Spectral Theory

The spectral theory of the Laplace operator centers on the analysis of its eigenvalues and eigenfunctions, particularly for the Dirichlet Laplacian on a bounded domain \Omega \subset \mathbb{R}^n with smooth boundary. The Dirichlet Laplacian is defined as the self-adjoint operator -\Delta on L^2(\Omega) with domain H^2(\Omega) \cap H_0^1(\Omega), where H_0^1(\Omega) consists of functions vanishing on \partial \Omega. Its spectrum is purely discrete, consisting of eigenvalues $0 < \lambda_1 \leq \lambda_2 \leq \cdots \to \infty with finite multiplicity, and a complete orthonormal basis of corresponding L^2-eigenfunctions \{\phi_k\}_{k=1}^\infty satisfying -\Delta \phi_k = \lambda_k \phi_k in \Omega and \phi_k = 0 on \partial \Omega. These eigenvalues admit a variational characterization through the Rayleigh quotient R(u) = \frac{\int_\Omega |\nabla u|^2 \, dx}{\int_\Omega u^2 \, dx} for u \in H_0^1(\Omega) \setminus \{0\}. The smallest eigenvalue is given by \lambda_1 = \min R(u), achieved at the ground state eigenfunction \phi_1 > 0 in \Omega, which minimizes the Dirichlet energy subject to L^2-normalization. More generally, the Courant-Fischer min-max theorem provides \lambda_k = \min_{\dim V = k} \max_{u \in V \setminus \{0\}} R(u), where the minimum is over k-dimensional subspaces V \subset H_0^1(\Omega), yielding a complete ordering of the spectrum without relying on explicit solutions. The eigenfunctions form an orthonormal basis for L^2(\Omega), enabling the spectral decomposition of the heat semigroup generated by \Delta. For t > 0 and f \in L^2(\Omega), the solution to the heat equation \partial_t u = \Delta u with u(0) = f is u(t) = e^{t \Delta} f = \sum_{k=1}^\infty e^{-\lambda_k t} \langle f, \phi_k \rangle_{L^2} \phi_k, where the series converges in L^2. This expansion reflects the diffusive decay, with higher modes (larger \lambda_k) attenuating faster as t \to \infty. A key asymptotic result is Weyl's law, which describes the distribution of eigenvalues for large k. The counting function N(\lambda) = \# \{ k : \lambda_k \leq \lambda \} satisfies N(\lambda) \sim \frac{\omega_n}{(2\pi)^n} |\Omega| \lambda^{n/2} as \lambda \to \infty, where \omega_n is the volume of the unit ball in \mathbb{R}^n and |\Omega| is the measure of \Omega. Equivalently, the eigenvalues grow as \lambda_k \sim c_n k^{2/n} |\Omega|^{-2/n}, with c_n = (2\pi)^2 / \omega_n^{2/n}, providing a geometric link between the spectrum and the domain's size.

Additional Mathematical Properties

The Laplacian operator \Delta is self-adjoint on the Hilbert space L^2(\Omega) equipped with suitable boundary conditions, such as Dirichlet or Neumann conditions on a bounded domain \Omega \subset \mathbb{R}^n. This property follows from integration by parts, yielding \int_\Omega u \Delta v \, dx = \int_\Omega v \Delta u \, dx for sufficiently smooth functions u, v \in C_c^\infty(\Omega) or their extensions in the appropriate Sobolev spaces, with the boundary terms vanishing under the specified conditions. Green's first identity relates the Laplacian to the gradient: for smooth functions u \in C^2(\overline{\Omega}) and v \in C^1(\overline{\Omega}), it states \int_\Omega (u \Delta v + \nabla u \cdot \nabla v) \, dx = \int_{\partial \Omega} u \frac{\partial v}{\partial n} \, dS, where \frac{\partial}{\partial n} denotes the outward normal derivative. Green's second identity, a direct consequence, is \int_\Omega (u \Delta v - v \Delta u) \, dx = \int_{\partial \Omega} \left( u \frac{\partial v}{\partial n} - v \frac{\partial u}{\partial n} \right) dS, which underscores the symmetry and boundary behavior essential for boundary value problems. The operator exponential e^{t\Delta} generates a contraction semigroup on L^p(\mathbb{R}^n) for $1 \leq p \leq \infty, solving the heat equation \partial_t w = \Delta w with initial data w(0) = f. This semigroup satisfies \|e^{t\Delta} f\|_{L^p} \leq \|f\|_{L^p} for all t \geq 0, reflecting the diffusive smoothing and non-expansive nature of heat flow. The negative Laplacian -\Delta is positive definite on L^2(\Omega) under Dirichlet boundary conditions, as \langle -\Delta u, u \rangle_{L^2} = \int_\Omega |\nabla u|^2 \, dx \geq 0 with equality only for constant functions (or zero in the Dirichlet case), establishing a quadratic form that underpins stability in elliptic problems. This positivity implies the maximum principle: for a solution u to \Delta u = 0 in \Omega, the maximum of u occurs on \partial \Omega, preventing interior maxima and ensuring uniqueness in Dirichlet problems.

Extensions to Vectors and Beyond

Vector Laplacian

The vector Laplacian extends the scalar Laplacian operator to vector fields in Euclidean space \mathbb{R}^3. For a sufficiently smooth vector field \mathbf{V}, it is defined by the identity \Delta \mathbf{V} = \nabla (\nabla \cdot \mathbf{V}) - \nabla \times (\nabla \times \mathbf{V}), where \nabla denotes the del operator, \nabla \cdot \mathbf{V} is the divergence, and \nabla \times \mathbf{V} is the curl. This expression provides a coordinate-free formulation that relates the vector Laplacian to fundamental vector calculus operations. In Cartesian coordinates, the vector Laplacian simplifies to the componentwise application of the scalar Laplacian: (\Delta \mathbf{V})_i = \Delta V_i = \sum_{j=1}^3 \frac{\partial^2 V_i}{\partial x_j^2} for each component i = 1, 2, 3. This equivalence underscores its role as a second-order differential operator measuring the "diffusion" or variation of the vector field across space. The identity \Delta \mathbf{V} = \nabla (\nabla \cdot \mathbf{V}) - \nabla \times (\nabla \times \mathbf{V}) derives from rearranging the standard vector calculus identity \nabla \times (\nabla \times \mathbf{V}) = \nabla (\nabla \cdot \mathbf{V}) - \Delta \mathbf{V}, where the componentwise Laplacian appears on the right. This form is particularly useful in domains without holes or singularities, ensuring the curl and divergence operators behave predictably under integration theorems like Stokes' theorem. In practice, the componentwise definition is often employed for computational simplicity in rectangular grids, while the identity form highlights physical interpretations, such as decomposing the field into irrotational and solenoidal parts via Helmholtz decomposition. In electromagnetism, the vector Laplacian governs the vector potential \mathbf{A} in the Coulomb gauge, where the gauge condition \nabla \cdot \mathbf{A} = 0 simplifies Maxwell's equations. For magnetostatics, this yields the Poisson-like equation \Delta \mathbf{A} = -\mu_0 \mathbf{J}, with \mu_0 the vacuum permeability and \mathbf{J} the current density, allowing \mathbf{A} to be solved directly via Green's functions similar to the scalar case. This equation arises from substituting the gauge into Ampère's law, \nabla \times \mathbf{B} = \mu_0 \mathbf{J}, and using \mathbf{B} = \nabla \times \mathbf{A}, resulting in \nabla \times (\nabla \times \mathbf{A}) = \mu_0 \mathbf{J}; the vector Laplacian identity then isolates \Delta \mathbf{A}. In time-dependent scenarios under the low-frequency approximation, the same form persists, facilitating solutions for magnetic fields in conductors or antennas. In fluid dynamics, the vector Laplacian appears in the Navier-Stokes equations as the viscous diffusion term for incompressible Newtonian fluids. The momentum equation includes \nu \Delta \mathbf{u}, where \nu is the kinematic viscosity and \mathbf{u} is the velocity field, representing the net effect of shear stresses that smooth velocity gradients and dampen turbulence. This term originates from the divergence of the viscous stress tensor, \nabla \cdot (2\nu \mathbf{D}) with \mathbf{D} the rate-of-strain tensor, which simplifies to \nu \Delta \mathbf{u} under incompressibility (\nabla \cdot \mathbf{u} = 0) and constant viscosity. For example, in laminar pipe flow, \Delta \mathbf{u} drives the parabolic velocity profile by balancing pressure gradients with viscous forces.

Laplace–Beltrami Operator

The Laplace–Beltrami operator extends the classical Laplace operator to functions on Riemannian manifolds, providing an intrinsic differential operator that respects the geometry defined by the metric tensor. For a smooth function f on a Riemannian manifold (M, g), it is defined as the divergence of the gradient: \Delta_g f = \operatorname{div}_g (\operatorname{grad}_g f). In local coordinates \{x^i\}, this expression becomes \Delta_g f = \frac{1}{\sqrt{|g|}} \partial_i \left( \sqrt{|g|} \, g^{ij} \partial_j f \right), where g = \det(g_{kl}) is the determinant of the metric tensor g_{kl}, g^{ij} are the components of its inverse, and summation over repeated indices i, j from 1 to \dim M is implied. This form ensures the operator is second-order, elliptic, and self-adjoint with respect to the L^2 inner product induced by the volume measure \sqrt{|g|} \, dx^1 \wedge \cdots \wedge dx^n. The operator's intrinsic nature arises because it is defined solely in terms of the Riemannian metric g, independent of any isometric embedding of M into a Euclidean space. Equivalently, \Delta_g f can be expressed as the trace of the Hessian tensor of f with respect to g: \Delta_g f = \operatorname{tr}_g (\nabla^2 f) = g^{ij} \nabla_i \nabla_j f, where \nabla denotes the Levi-Civita connection. This perspective highlights its role as a contraction that averages second derivatives along geodesics, preserving the manifold's geometry without reference to ambient coordinates. On the unit 2-sphere S^2 with the standard round metric of constant sectional curvature 1, the Laplace–Beltrami operator admits spherical harmonics Y_{lm} as eigenfunctions, satisfying \Delta_{S^2} Y_{lm} = -l(l+1) Y_{lm} for nonnegative integers l and m = -l, \dots, l. These eigenvalues -l(l+1) are negative with multiplicity $2l+1, reflecting the operator's negative semi-definiteness and enabling spectral decompositions for solving boundary value problems on the sphere. Applications of the Laplace–Beltrami operator span analysis and geometry on manifolds. In the heat method for geodesic distance computation, one solves the heat equation \partial_t u = \Delta_g u with an initial Dirac delta source; the short-time solution approximates the geodesic distance function via normalization and gradient flow, offering an efficient alternative to traditional eikonal methods on curved domains. Heat kernels, as fundamental solutions p_g(x,y,t) = \sum_k e^{-\lambda_k t} \phi_k(x) \phi_k(y) to the same equation (where \{\lambda_k, \phi_k\} are eigenvalues and eigenfunctions of -\Delta_g), encode manifold invariants like volume growth and injectivity radius through asymptotic expansions. Additionally, in Ricci flow—a metric evolution equation \partial_t g = -2 \operatorname{Ric}_g—Perelman's entropy functionals from the early 2000s incorporate the Laplace–Beltrami operator in their monotonicity proofs, linking spectral properties to the resolution of singularities and the proof of the Poincaré conjecture.

D'Alembertian and Other Relativistic Variants

In special relativity, the d'Alembertian operator, often denoted by □, serves as the wave operator in four-dimensional Minkowski spacetime, defined as □ = ∂²/∂t² - Δ, where Δ is the spatial Laplacian and units are chosen such that the speed of light c = 1. This operator is the natural hyperbolic analogue of the elliptic Laplacian in Euclidean space, arising from the invariant wave equation for massless fields propagating at the speed of light. The d'Alembertian plays a central role in the Klein-Gordon equation, which describes the dynamics of massive scalar fields in relativistic quantum field theory: (□ + m²)φ = 0, where m is the mass parameter and φ is the scalar field. For m = 0, this reduces to the massless wave equation □φ = 0, governing phenomena like electromagnetic waves in vacuum. In the broader context of pseudo-Riemannian manifolds with Lorentzian metrics, the d'Alembertian generalizes the Laplace-Beltrami operator to account for the indefinite metric signature, typically (-,+,+,+), enabling the study of wave propagation on curved spacetimes. This extension preserves the divergence form but incorporates the metric's hyperbolic structure, contrasting with the positive-definite Riemannian case. Other variants include the biharmonic operator Δ², a fourth-order elliptic operator that arises in the theory of thin plate bending under transverse loading, where it models the deflection of elastic plates in static equilibrium. For nonlinear extensions, the p-Laplacian, defined as div(|∇u|^{p-2} ∇u) for p > 1, appears in models of nonlinear diffusion processes, such as non-Newtonian fluid flow or anomalous transport in porous media.

References

  1. [1]
    [PDF] Curl, Divergence and Laplacian - Purdue Math
    Remark: The Laplace operator takes a scalar function f as input and ∆f is a scalar function. It appears in the mathematical description of problems like ...
  2. [2]
    Laplacian (operator) - MIT
    The Laplacian is defined as. ∇ 2 = ∇ ⋅ ∇ . \nabla^2 = \nabla \cdot \nabla. ∇2=∇⋅∇. In cartesian coordinates, ∇ 2 F = ∂ 2 F x ∂ x 2 + ∂ 2 F y ∂ y 2 + ∂ 2 F z ...
  3. [3]
    [PDF] The geometrical significance of the Laplacian Abstract
    Jul 20, 2015 · The Laplacian operator can be defined, not only as a differential operator, but also through its averaging properties.
  4. [4]
    [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.
  5. [5]
    [PDF] 8 Laplace's equation: properties - UCSB Mathematics Department
    So Laplace's operator is indeed invariant under rotations. The rotation invariance also implies that Laplace's equation allows rotationally invariant solutions,.
  6. [6]
    [PDF] Introducing the Laplacian - Fan Chung Graham
    The Laplacian is named after Pierre-Simon Laplace, the famous French mathematician who studied what has come to be known as the Laplace-Beltrami operator. The ...
  7. [7]
    [PDF] Analysis on Manifolds via the Laplacian - Mathematics and Statistics
    ... define our star operator known as the Laplacian, or Laplace operator, or Laplace-Beltrami operator. Definition 14. The Laplacian on (M,g) is the operator.<|control11|><|separator|>
  8. [8]
    [PDF] Discrete Laplace Operators
    1. Introduction. The Laplacian is perhaps the prototypical differential operator for various. physical phenomena. It describes, for example, heat diffusion, ...
  9. [9]
    [PDF] Partial Differential Equations (based on L.C. Evans's textbook) by ...
    The Laplace operator on Rn is defined by. (2.1). ∆u = n. X k=1 uxkxk = div(∇u) ... We consider the second-order linear differential operator ∂t + L defined by.
  10. [10]
    Pierre-Simon Laplace (1749 - 1827) - Biography - MacTutor
    Laplace's first paper which was to appear in print was one on the integral calculus which he translated into Latin and published at Leipzig in the Nova acta ...
  11. [11]
    [PDF] 16.5 Curl and Divergence
    The operator. ∇2 = ∇·∇ is called the Laplace operator, or Laplacian, because of its relation to Laplace's equation.
  12. [12]
    [PDF] A Visual Introduction to Partial Differential Equations - Math N54
    Aug 13, 2019 · Let's first start with some basic definitions. Definition. The Laplacian is a second order partial differential operator, given in Rn by. ∆ = ∂
  13. [13]
    Laplacian -- from Wolfram MathWorld
    The Laplacian is extremely important in mechanics, electromagnetics, wave theory, and quantum mechanics, and appears in Laplace's equation.
  14. [14]
    June 10, 1854: Riemann's classic lecture on curved space
    Gauss described Riemann as having “a gloriously fertile originality” in his report on the thesis, and two years later, when Riemann was required to give a ...
  15. [15]
    [PDF] Analytic and geometric aspects of Laplace operator on Riemannian ...
    The relationship between geometric structure of manifolds and spectrum of differential operators created a new concept which is spectral geometry. In the case ...
  16. [16]
    The Heat Equation - Pauls Online Math Notes
    Sep 5, 2025 · In this section we will do a partial derivation of the heat equation ... With Fourier's law we can easily remove the heat flux from this equation.
  17. [17]
    [PDF] 2 Heat Equation
    Below we provide two derivations of the heat equation, ut − kuxx = 0 ... cρut(x, t)dx. Fourier's Law says that heat flows from hot to cold regions at ...
  18. [18]
    [PDF] Théorie analytique de la chaleur - University of Notre Dame
    Cette theorie formera desormais nne, des branches les plus' importantes de la .physique ge- nerale. Les cODnaiesances que les· plus anciens peuples avaient pu ...
  19. [19]
    [PDF] The heat equation
    steady-state solution (with ∂u/∂t = 0). This turns our differential equation into Laplace's equation, being. ∂2u. ∂x2. +. ∂2u. ∂y2. = 0. (5.10). 7. Page 8. To ...
  20. [20]
    [PDF] The 1-D Heat Equation
    Sep 8, 2006 · Fourier's law of heat transfer: rate of heat transfer proportional to negative temperature gradient,. Rate of heat transfer. ∂u. = (1). −K0.
  21. [21]
    [PDF] Section 11.1
    The heat equation is more commonly expressed in terms of the Laplace operator. ... The Steady State Heat Equation. An interesting and important special case of ...
  22. [22]
    [PDF] Harmonic Function Theory - Sheldon Axler
    . Green's identity is the key to the proof of the mean-value property. Before stating the mean-value property, we introduce some notation: B(a, r ) = {x ...
  23. [23]
    [PDF] Griffiths D.J. Introduction to electrodynamics (3ed., PH, 1999)(T ...
    Poisson's Equation and Laplace's Equation. The Potential of a Localized Charge Distribution. Summary; Electrostatic Boundary Conditions. 79. 83. 83. 87. 2.4.
  24. [24]
    [PDF] Method of Green's Functions - DSpace@MIT
    For 3D domains, the fundamental solution for the Green's function of the. Laplacian is −1/(4πr), where r = (x − ξ). 2. + (y − η). 2. + (z − ζ). 2. (see ...
  25. [25]
    [PDF] 7 Calculus of Variations
    In this section, we show that the solution of Laplace's equation can be rewritten as a mini- mization problem. Let. A≡{w ∈ C2(Ω),w = g for x ∈ ∂Ω}. Let. I( ...
  26. [26]
    [PDF] ************************************* Introduction to Variational ...
    Example 1 – Dirichlet's Principle. The starting example of variational method for. PDE is the Dirichlet principle for Laplace's equation: ∆u = 0, u|∂Ω = f ...
  27. [27]
    [PDF] Sect. 1.12 Variational Approach to the Solution of the Laplace and ...
    With → and g → p/eo, the minimization of the functional yields the "equation of motion" of the electrostatic potential in the presence of a charge density and ...
  28. [28]
    [PDF] Laplace's equation in the Polar Coordinate System - UC Davis Math
    In this note, I would like to derive. Laplace's equation in the polar coordinate system in details. Recall that Laplace's equation in R2 in terms of the usual ( ...
  29. [29]
    Inequalities for Harmonic Polynomials In Two and Three Dimensions
    The equality holds if and only if U(r, 4))=rn cos n(4)-4)o), g)a real. GENERALIZED GRADIENT THEOREM. If U(r, 4)) is a harmonic polynomial of degree n satisfying ...
  30. [30]
    [PDF] Laplace's Equation on a Disk - MATH 467 Partial Differential ...
    Solve Laplace's equation on the unit disk with the following Dirichlet boundary condition. ∆u = 0 for x2 + y2 < 1 u(1,θ) = π - θ for -π<θ< ...
  31. [31]
    [PDF] 3 Laplace's Equation
    The Laplace equation is one of the most fundamental differential equations in all of mathematics, pure as well as applied. A function ψ : M → R obeying ∇2ψ = 0 ...
  32. [32]
    [PDF] Solution to Laplace's Equation in Cylindrical Coordinates
    In cylindrical coordinates apply the divergence of the gradient on the potential to get Laplace's equation. ∇2V (ρ, φ, z) = ρ∂. 2V. ∂ρ2 +. ∂V.<|separator|>
  33. [33]
    An Instructional Derivation of the Laplacian Operator in Spherical ...
    We present an instructional derivation of the Laplacian operator in spherical coordinates. Our derivation is self-contained and employs well-known mathematical ...
  34. [34]
    [PDF] 5 The Poisson and Laplace Equations
    Here we have derived them in the context of gravity and electrostatics, but their applications spread much further. To give just one further example, in ...Missing: operator celestial
  35. [35]
    [PDF] The Laplace operator in polar coordinates in several dimensions
    In case n = 2, we can write y = x1, x = x2. The polar coordinates (r, θ) are defined by r2 = x2 + y2,. (2) x = r cos θ and y = r sin θ,.
  36. [36]
    [PDF] Laplace's Equation The Fundamental Solution and Green's Function
    Mar 24, 2021 · n(n - 2)ωn. 1. |x|n−2 is the fundamental solution for Laplace's equation for n ≥ 3. In all cases, Φ ∈ L1 loc(Rn), so given a function f ...
  37. [37]
    [PDF] Chapter 2: Diffusion
    Then, substitution of ∂c/∂t and ∂2c/∂x2 in the diffusion equation yields: ... We here assume isotropy in diffusion (single D value) but allow for anisotropy in ...
  38. [38]
    [PDF] Spectral Theory of Partial Differential Equations - Publish
    This book presents highlights of spectral theory for selfadjoint partial differential operators, emphasizing problems with discrete spectrum. Style of the ...
  39. [39]
    [PDF] 6 Eigenvalues of the Laplacian
    Example 1. ... That is, m = λ1 and u is a corresponding eigen- function. Proof. Suppose u is the minimizer of the Rayleigh quotient and m is the Rayleigh quotient.
  40. [40]
    [PDF] Laplacian Eigenfunctions - UC Davis Mathematics
    May 23, 2007 · Laplacian eigenfunctions are used to analyze spatial frequency in domains of complicated shapes, and are tailored to the domain. They form a ...
  41. [41]
    Das asymptotische Verteilungsgesetz der Eigenwerte linearer ...
    Das asymptotische Verteilungsgesetz der Eigenwerte linearer partieller Differentialgleichungen (mit einer Anwendung auf die Theorie der Hohlraumstrahlung).
  42. [42]
    [PDF] Notes on Partial Differential Equations John K. Hunter
    These are notes from a two-quarter class on PDEs that are heavily based on the book Partial Differential Equations by L. C. ... self-adjoint operator on L2(Ω) for.<|separator|>
  43. [43]
    [PDF] Chapter 2: Laplace's equation - UC Davis Mathematics
    Green's first identity provides a proof of the uniqueness of solutions of the. Dirichlet problem based on estimates of L2-norms of derivatives instead of maxi-.
  44. [44]
    [PDF] The Heat Equation - UC Davis Mathematics
    The heat equation semigroup on X = L2(Rn) is an example of a contraction semigroup. The term 'contraction' is not used here in a strict sense.
  45. [45]
    The meaning of the vector Laplacian - ScienceDirect
    The meaning of the vector Laplacian. Author links open overlay panelParry Moon, Domina Eberle Spencer.
  46. [46]
    Vector Laplacian -- from Wolfram MathWorld
    A vector Laplacian can be defined for a vector A by del ^2A=del (del ·A)-del x(del xA), (1) where the notation ✡ is sometimes used to distinguish the vector ...
  47. [47]
    [PDF] superposition integral and boundary value points of view
    Specification of the potential in this way is sometimes called setting the gauge, and with (2) we have established the Coulomb gauge. We turn now to the ...
  48. [48]
    [PDF] 3 The Navier-Stokes Equation
    This follows from a uniqueness theorem that is proven in the same way as the uniqueness of solutions to the Laplace equation (see the lectures on Vector ...
  49. [49]
    None
    Summary of each segment:
  50. [50]
    [PDF] I. INTRODUCTION TO MEAN CURVATURE FLOW 1. The Hessian ...
    If ∇ is the Levi-Civita connection of a metric g on M, we obtain the Laplace-. Beltrami operator ∆g by taking the trace: ∆gf = trg∇2f. If M is a ...
  51. [51]
    [1204.6216] Geodesics in Heat - arXiv
    Apr 24, 2012 · We introduce the heat method for computing the shortest geodesic distance to a specified subset (eg, point or curve) of a given domain.
  52. [52]
    [PDF] Generalized Causal Set d'Alembertians - arXiv
    Mar 6, 2014 · Abstract: We introduce a family of generalized d'Alembertian operators in D-dimensional. Minkowski spacetimes MD which are manifestly ...
  53. [53]
    [PDF] arXiv:2309.10848v1 [hep-th] 19 Sep 2023
    Sep 19, 2023 · Unless otherwise specified, we work in n spacetime dimensions, assume h = c = 1 and use metric signature (−,+,...,+). The D'Alembertian operator ...<|separator|>
  54. [54]
    [PDF] Natural Higher-Derivatives Generalization for the Klein-Gordon ...
    Nov 4, 2020 · A general polynomial in the d'Alembertian operator wave equation is briefly discussed in reference [18], followed by a carefull treatment of a ...
  55. [55]
    [PDF] From D'Alembert to Klein-Gordon and Schrödinger - arXiv
    Jun 18, 2020 · The procedure is illustrated by reducing the D'Alembert theory on a five-dimensional Minkowski space-time to a massive Klein-Gordon theory in ...
  56. [56]
    [PDF] arXiv:2406.19147v2 [gr-qc] 18 Dec 2024
    Dec 18, 2024 · The Laplace-Beltrami operator formally coincides with the d'Alembertian (7), but throughout this paper we will re- serve ∆ for the (Riemannian) ...Missing: relativistic | Show results with:relativistic
  57. [57]
    [1410.3488] A nonlocal biharmonic operator and its connection with ...
    Oct 13, 2014 · Abstract:We introduce here a nonlocal operator as a natural generalization to the biharmonic operator that appears in plate theory.
  58. [58]
    Evolution of Interfaces for the Nonlinear Parabolic p-Laplacian Type ...
    May 24, 2016 · Abstract page for arXiv paper 1605.07279: Evolution of Interfaces for the Nonlinear Parabolic p-Laplacian Type Reaction-Diffusion Equations.<|control11|><|separator|>