Fact-checked by Grok 2 weeks ago

Poisson's equation

Poisson's equation is a second-order linear of elliptic type, generally expressed in three dimensions as \nabla^2 \phi(\mathbf{r}) = f(\mathbf{r}), where \nabla^2 denotes the Laplacian operator, \phi is an unknown function, and f represents a given source term that drives the behavior of the potential. This equation serves as a cornerstone in the mathematical modeling of physical phenomena involving potentials, generalizing \nabla^2 \phi = 0, which applies in source-free regions. Named after the French mathematician and physicist , the equation was first published by him in in the Bulletin de la Société Philomatique, where he derived it in the context of electrostatic theory as a relation between and . In physics, Poisson's equation finds extensive applications across multiple domains, most notably in , where it takes the form \nabla^2 \phi = -\rho / \epsilon_0; here, \phi is the , \rho is the , and \epsilon_0 is the . Similarly, in Newtonian gravitation, the equation describes the \Phi generated by a mass density \rho via \nabla^2 \Phi = 4\pi G \rho, with G being the , linking the potential to the distribution of mass in and . Beyond these, it models steady-state heat conduction with internal heat sources (where f relates to heat generation), for incompressible flows via the pressure Poisson equation. Mathematically, Poisson's equation is well-posed under appropriate boundary conditions, such as Dirichlet (prescribed potential on the boundary) or (prescribed normal derivative), ensuring unique solutions in bounded domains, and it admits representations for explicit integral solutions in free space. Its elliptic nature implies smooth solutions away from singularities in f, and numerical methods like finite differences or finite elements are commonly employed for complex geometries due to the lack of closed-form solutions in general cases. Poisson's foundational role extends to broader , influencing developments in and the study of elliptic partial differential equations.

Mathematical foundations

General statement

Poisson's equation is a fundamental in , expressed in its general scalar form as \nabla^2 \phi = f, where \phi is the function to be determined, \nabla^2 denotes the Laplacian operator, and f is a given source term representing inhomogeneities in the domain. This equation arises in various boundary value problems over a domain \Omega \subset \mathbb{R}^n, typically supplemented by appropriate conditions on the boundary \partial \Omega. Poisson's equation reduces to when the source term vanishes (f = 0). The Laplacian operator \nabla^2 takes different explicit forms depending on the used. In Cartesian coordinates (x, y, z), it is given by \nabla^2 \phi = \frac{\partial^2 \phi}{\partial x^2} + \frac{\partial^2 \phi}{\partial y^2} + \frac{\partial^2 \phi}{\partial z^2}. In spherical coordinates (r, \theta, \phi), the expression becomes \nabla^2 \phi = \frac{1}{r^2} \frac{\partial}{\partial r} \left( r^2 \frac{\partial \phi}{\partial r} \right) + \frac{1}{r^2 \sin \theta} \frac{\partial}{\partial \theta} \left( \sin \theta \frac{\partial \phi}{\partial \theta} \right) + \frac{1}{r^2 \sin^2 \theta} \frac{\partial^2 \phi}{\partial \phi^2}. In cylindrical coordinates (\rho, \varphi, z), it is \nabla^2 \phi = \frac{1}{\rho} \frac{\partial}{\partial \rho} \left( \rho \frac{\partial \phi}{\partial \rho} \right) + \frac{1}{\rho^2} \frac{\partial^2 \phi}{\partial \varphi^2} + \frac{\partial^2 \phi}{\partial z^2}. These coordinate-specific forms facilitate solutions in domains with corresponding symmetries. To ensure well-posedness, Poisson's equation is typically paired with conditions on \partial \Omega. The specifies the value of the potential directly: \phi = g on \partial \Omega, where g is a prescribed . The , in contrast, specifies the normal derivative: \frac{\partial \phi}{\partial n} = h on \partial \Omega, where \mathbf{n} is the outward unit normal vector and h is prescribed. Mixed conditions combining both types may also be employed over different portions of the boundary.

Relation to Laplace's equation

Poisson's equation, in its general form \nabla^2 \phi = f, reduces to \nabla^2 \phi = 0 in the homogeneous case where the source term f = 0, representing scenarios devoid of distributed sources or charges. This limiting case is fundamental in , where governs the behavior of functions in source-free domains. Physically, Laplace's equation describes equilibrium states in regions without internal sources, such as the electric potential inside a charge-free cavity within a conductor, whereas Poisson's equation accounts for the influence of localized sources, like charge distributions, that drive deviations from harmonicity. This distinction underscores Poisson's equation as a generalization, incorporating inhomogeneities that Laplace's equation idealizes away. Uniqueness theorems for solutions to both equations rely on boundary conditions. Under Dirichlet conditions, where the potential \phi is specified on the boundary, solutions to Poisson's equation are unique; the difference between any two solutions satisfies Laplace's equation with homogeneous Dirichlet data, which admits only the trivial solution by the maximum principle. For Neumann conditions, specifying the normal derivative \partial \phi / \partial n, uniqueness holds up to an additive constant (a harmonic function), with proofs invoking energy methods or integration by parts to show that non-trivial solutions would contradict boundary compatibility. In both cases, the homogeneous limit ensures that Laplace's solutions are a subset, uniquely determined within the same framework when f = 0. Green's identities provide a mathematical bridge between the equations, facilitating proofs of uniqueness and representation formulas. Green's second identity states that for sufficiently smooth functions \phi and \psi, \int_V (\phi \nabla^2 \psi - \psi \nabla^2 \phi) \, dV = \int_{\partial V} \left( \phi \frac{\partial \psi}{\partial n} - \psi \frac{\partial \phi}{\partial n} \right) dS, where V is a volume with boundary \partial V and \partial / \partial n denotes the outward normal derivative. When \psi satisfies (\nabla^2 \psi = 0) and \phi satisfies Poisson's (\nabla^2 \phi = f), the identity simplifies to \int_V \phi f \, dV = \int_{\partial V} \left( \phi \frac{\partial \psi}{\partial n} - \psi \frac{\partial \phi}{\partial n} \right) dS, linking source integrals to boundary data and highlighting how functions (\psi) can represent solutions to the inhomogeneous problem. This is pivotal in deriving via Green's functions, where the solution to is adjusted for the source term in Poisson's.

Derivations and theoretical context

From Gauss's law

Poisson's equation arises in physical contexts through the application of Gauss's divergence theorem, which relates the flux of a vector field through a closed surface to the divergence of that field within the enclosed volume. The theorem states that for a vector field \mathbf{F}, \int_V (\nabla \cdot \mathbf{F}) \, dV = \oint_S \mathbf{F} \cdot d\mathbf{S}, where V is the volume and S its bounding surface. This integral form allows derivation of differential equations from physical laws expressed as surface integrals. In electrostatics and gravitation, the integral forms of Gauss's law quantify the total "source" (charge or mass) enclosed by a surface, and applying the divergence theorem yields the local differential relation between the field and its source density. In , in integral form asserts that the through a closed surface equals the enclosed charge divided by the \epsilon_0, \oint_S \mathbf{E} \cdot d\mathbf{S} = \frac{Q_{\text{encl}}}{\epsilon_0}. The \epsilon_0, with value $8.854 \times 10^{-12} \, \text{F m}^{-1}, measures the electric field's strength in for a given . Applying the gives the differential form \nabla \cdot \mathbf{E} = \rho / \epsilon_0, where \rho is the . Defining the \phi such that \mathbf{E} = -\nabla \phi, yields \nabla \cdot (-\nabla \phi) = \frac{\rho}{\epsilon_0} \implies \nabla^2 \phi = -\frac{\rho}{\epsilon_0}. This is Poisson's equation for electrostatics. The gravitational analog follows similarly. Gauss's law for gravity in integral form states that the flux of the gravitational field \mathbf{g} through a closed surface equals -4\pi G times the enclosed mass, \oint_S \mathbf{g} \cdot d\mathbf{S} = -4\pi G M_{\text{encl}}, where G is the gravitational constant, with value $6.674 \times 10^{-11} \, \text{m}^3 \text{kg}^{-1} \text{s}^{-2}, quantifying the strength of gravitational attraction between masses. The divergence theorem produces the differential form \nabla \cdot \mathbf{g} = -4\pi G \rho, with \rho now the mass density. The gravitational potential \Phi is defined by \mathbf{g} = -\nabla \Phi, so \nabla \cdot (-\nabla \Phi) = -4\pi G \rho \implies \nabla^2 \Phi = 4\pi G \rho. This yields Poisson's equation for Newtonian .

In vector calculus

In , Poisson's equation arises as a fundamental expressing the relationship between a \phi and a source term f, in the coordinate-independent form \nabla^2 \phi = f, where \nabla^2 denotes the Laplacian defined abstractly as the divergence of the , \nabla^2 \phi = \nabla \cdot (\nabla \phi)./04%3A_Line_and_Surface_Integrals/4.06%3A_Gradient_Divergence_Curl_and_Laplacian) This identity holds in any where the \nabla \phi produces a from the scalar \phi, and the \nabla \cdot measures the of that field, yielding a scalar second-order independent of specific coordinate systems./04%3A_Line_and_Surface_Integrals/4.06%3A_Gradient_Divergence_Curl_and_Laplacian) Green's first identity provides a key integral formulation that connects the Laplacian to boundary behavior, stated as \int_V \left( \phi \nabla^2 \psi + \nabla \phi \cdot \nabla \psi \right) dV = \int_S \phi \frac{\partial \psi}{\partial n} dS for scalar fields \phi and \psi over a volume V with boundary S, where \partial / \partial n is the outward normal derivative. Specializing to \psi = \phi, this becomes \int_V \left( \phi \nabla^2 \phi + |\nabla \phi|^2 \right) dV = \int_S \phi \frac{\partial \phi}{\partial n} dS, which establishes variational principles for solutions to Poisson's equation by relating the volume integral of the source to energy-like functionals involving the . Poisson's equation also emerges as the time-independent limit of parabolic or equations, such as the \frac{\partial u}{\partial t} = \nabla^2 u + f, where setting \frac{\partial u}{\partial t} = 0 yields \nabla^2 u = -f. Similarly, for the wave equation \frac{\partial^2 u}{\partial t^2} = c^2 \nabla^2 u + f, the steady-state condition \frac{\partial^2 u}{\partial t^2} = 0 reduces to \nabla^2 u = -f / c^2, highlighting Poisson's role in stationary scenarios without transient dynamics. These derivations underscore the equation's abstract mathematical structure, applicable across vector fields in \mathbb{R}^n.

Solution techniques

Analytical approaches

Analytical approaches to solving Poisson's equation ∇²φ = f rely on exact methods that exploit the linearity and elliptic nature of the operator, providing closed-form expressions or series representations under suitable conditions and geometries. These techniques are particularly effective for simple s or when term f admits a convenient representation in the chosen basis. The Green's function method offers a general representation for the solution in unbounded or free space. For the three-dimensional case with the equation ∇²φ = f, the Green's function G(r, r') satisfies ∇²G = δ(r - r'), where δ is the , and the solution is given by φ(r) = ∫ G(r, r') f(r') dV' over the volume V. In three-dimensional free space, assuming the solution vanishes at infinity, the fundamental solution is G(r, r') = -1/(4π |r - r'|). This form arises from the fundamental solution of the Laplacian and ensures the correct singularity at r = r' while satisfying the homogeneous equation elsewhere. For domains with , methods provide an efficient analytical pathway. Applying the to Poisson's equation yields -|k|² φ̂(k) = f̂(k) in the transform domain, where φ̂ and f̂ are the transforms of φ and f, respectively, and k is the . Solving for φ̂(k) = -f̂(k) / |k|² (for k ≠ 0). This requires the compatibility condition that the zero-mode coefficient f̂(0) = 0, ensuring solvability up to an additive constant. Inverting the transform gives the solution φ(r) = (1/(2π)³) ∫ [-f̂(k) / |k|²] e^{i k · r} d³k. This approach is exact for periodic sources and leverages the , making it ideal for translationally invariant problems. Separation of variables is a powerful technique for bounded domains with separable geometries, such as rectangles or spheres, where boundary conditions can be imposed by . Assume Dirichlet conditions φ = 0 on the of a rectangular 0 < x < a, 0 < y < b. The source f(x,y) is expanded in a double sine series using the eigenfunctions sin(mπx/a) sin(nπy/b), leading to a solution φ(x,y) as a corresponding series ∑∑ A_{mn} sin(mπx/a) sin(nπy/b), where coefficients A_{mn} are determined by projecting f onto the basis and solving the resulting algebraic system from the eigenvalue problem ∇² (eigenfunction) = -λ (eigenfunction), with λ = (mπ/a)² + (nπ/b)². This method reduces the PDE to an infinite system of ODEs, solvable via orthogonality of the eigenfunctions. Similar expansions apply in spherical coordinates using spherical harmonics for radial symmetry. For far-field approximations, particularly in exterior problems or when sources are localized, multipole expansions provide a hierarchical series representation of the solution. In three dimensions, the potential φ(r) for large |r| is expanded as φ(r) = ∑{l=0}^∞ (1/r^{l+1}) ∑{m=-l}^l Q_{lm} Y_{lm}(θ, ϕ), where Y_{lm} are , and Q_{lm} are multipole moments computed from integrals involving f(r') and powers of r'. This series converges rapidly far from the sources, offering an asymptotic solution that captures the leading-order behavior, such as monopole, dipole, and higher terms, without solving the full equation globally.

Numerical methods

Numerical methods are essential for solving Poisson's equation in complex geometries or with irregular boundary conditions where analytical solutions are impractical. These approaches discretize the continuous problem into a system of algebraic equations, which can then be solved iteratively or directly, balancing accuracy, computational efficiency, and scalability for large-scale problems. Common leverage structured grids, variational principles, or hierarchical structures to approximate the Laplacian operator and handle the resulting linear systems. Finite difference methods approximate the derivatives in Poisson's equation using discrete differences on a uniform grid. For a two-dimensional case, the Laplacian is approximated by central differences as \nabla^2 \phi_{i,j} \approx \frac{\phi_{i+1,j} - 2\phi_{i,j} + \phi_{i-1,j}}{h^2} + \frac{\phi_{i,j+1} - 2\phi_{i,j} + \phi_{i,j-1}}{h^2}, where h is the grid spacing and \phi_{i,j} denotes the solution at grid point (i,j). This leads to a sparse linear system that is typically solved using iterative methods due to its size. The , an iterative relaxation technique, updates each grid point sequentially by solving for \phi_{i,j} using the most recent values of neighboring points, promoting faster convergence than Jacobi iteration for elliptic problems like Poisson's equation. These methods are straightforward to implement on rectangular domains but require careful handling of boundaries to maintain second-order accuracy. Finite element methods reformulate Poisson's equation in a weak variational sense, multiplying by a test function \psi and integrating by parts to obtain \int_\Omega \nabla \phi \cdot \nabla \psi \, dV = \int_\Omega f \psi \, dV for suitable boundary conditions, where \Omega is the domain. The domain is triangulated into elements, and the solution \phi is approximated as a linear combination of basis functions (e.g., piecewise linear hat functions) over these elements, leading to a stiffness matrix that assembles element-wise contributions. This approach excels in handling irregular geometries and heterogeneous materials by conforming the mesh to the boundary, achieving optimal convergence rates of order h^k for polynomials of degree k. The resulting system is solved via direct or preconditioned iterative solvers, with the method's flexibility making it widely used in engineering simulations. Multigrid methods accelerate convergence for large discretized systems by employing a hierarchy of grids, from coarse to fine resolutions. Smoothing is applied on the fine grid to eliminate high-frequency errors, residuals are transferred to coarser grids for low-frequency correction, and the improved solution is interpolated back to the fine grid. This V-cycle or W-cycle structure reduces the condition number effectively, achieving grid-independent convergence rates near 0.1 per cycle for on structured grids. Introduced in the 1970s, these methods are particularly efficient for , enabling linear-time scaling for problems with millions of unknowns by combining geometric coarsening with robust prolongation and restriction operators. Fast Poisson solvers exploit structure in the problem to reduce complexity below O(N^2) for N degrees of freedom. For periodic boundary conditions, the fast Fourier transform (FFT) diagonalizes the discrete Laplacian in spectral space, allowing exact solution in O(N \log N) time via convolution with the inverse Green's function. This approach, dating to early FFT applications in numerical PDEs, is ideal for uniform domains like periodic boxes in simulations. For non-periodic or N-body-like problems, hierarchical matrices (H-matrices) approximate the dense potential matrix with low-rank blocks organized in a tree structure, enabling O(N \log N) or near-linear matrix-vector products and factorizations. These techniques, based on multipole expansions, are crucial for high-fidelity computations in electrostatics and gravity.

Applications in physics

Electrostatics

In electrostatics, Poisson's equation describes the relationship between the electric potential and the charge distribution in a region without time-varying magnetic fields. The equation takes the form \nabla^2 \phi = -\frac{\rho}{\epsilon_0}, where \phi is the electric potential, \rho is the charge density, and \epsilon_0 is the vacuum permittivity. This equation is derived from \nabla \cdot \mathbf{E} = \rho / \epsilon_0 combined with the definition of the electric field \mathbf{E} = -\nabla \phi. The electric field can thus be obtained from the potential as \mathbf{E} = -\nabla \phi, allowing Poisson's equation to serve as the fundamental governing equation for computing both potential and field from known charges. For a point charge q located at the origin, the solution in free space is the Coulomb potential \phi(\mathbf{r}) = \frac{1}{4\pi \epsilon_0} \frac{q}{r}, which satisfies Poisson's equation everywhere except at the origin where \rho = q \delta(\mathbf{r}), with \delta being the Dirac delta function. This form arises from the symmetry and directly follows from integrating Coulomb's law. For a general, localized charge distribution \rho(\mathbf{r}'), the solution to Poisson's equation in infinite space is given by the integral \phi(\mathbf{r}) = \frac{1}{4\pi \epsilon_0} \int \frac{\rho(\mathbf{r}')}{|\mathbf{r} - \mathbf{r}'|} \, dV', where the integration extends over all space. This Green's function approach exploits the fundamental solution to the Laplacian, \nabla^2 (1/|\mathbf{r}|) = -4\pi \delta(\mathbf{r}). In cases of spherical symmetry, such as a uniformly charged sphere, the integral simplifies using to yield piecewise potentials: constant inside the sphere and $1/r decay outside, matching the point charge form at large distances. Boundary value problems involving conductors require satisfying conditions like \phi = 0 on the conductor surface. The method of images addresses this by replacing the conductor with fictitious image charges that reproduce the correct boundary conditions in the region of interest, thereby solving Poisson's equation indirectly. For instance, a point charge q at distance d above an infinite grounded conducting plane at z=0 is equivalent to an image charge -q at z = -d, yielding \phi = 0 on the plane and the correct potential for z > 0. This technique extends to spherical conductors and other geometries, ensuring uniqueness via the properties of elliptic partial differential equations.

Newtonian gravity

In Newtonian gravity, Poisson's equation relates the gravitational potential \phi to the mass density \rho through the form \nabla^2 \phi = 4\pi G \rho, where G is the . The \mathbf{g} is then given by \mathbf{g} = -\nabla \phi, describing the attractive force per unit mass arising from the distributed mass. This equation arises as the gravitational analog to the electrostatic case, differing primarily in the universal attractive nature of and the absence of like-charge repulsion. For a point mass M at the origin, the solution to Poisson's equation in vacuum (where \rho = 0 except at the origin) is the familiar \phi(\mathbf{r}) = -\frac{GM}{r}, with r = |\mathbf{r}|, which yields the inverse-square law for the gravitational field \mathbf{g} = -\frac{GM}{r^2} \hat{\mathbf{r}}. This potential is obtained by integrating over the Dirac delta function representation of the point mass in the source term. In cases of spherical symmetry, such as for stars or planets modeled as spherically symmetric mass distributions, the potential can be found by solving Poisson's equation radially. For the exterior region (r > R, where R is the radius of the distribution), the solution matches that of a point mass at the center, \phi(r) = -\frac{GM}{r}, with M the total mass. Inside the distribution, the potential depends on the enclosed mass up to radius r, often resulting in a quadratic form for uniform density, \phi(r) \propto - (3R^2 - r^2) (up to constants and scaling), ensuring continuity at the boundary and zero field at the center for symmetry. These solutions follow from Gauss's law applied spherically, confirming that exterior fields are unaffected by the detailed internal distribution. Poisson's equation plays a central role in galactic dynamics for modeling self-gravitating systems, where the mass distribution \rho (from stars, gas, and ) generates the potential that governs orbital motion. In such systems, the equation is solved iteratively or numerically to capture the collective gravitational effects, as in the collisionless coupled with Poisson's form, enabling studies of galactic structure and stability.

Applications in engineering and other fields

Fluid dynamics

In , Poisson's equation frequently appears in the formulation of irrotational flows via the \phi, defined such that the velocity field is \mathbf{v} = \nabla \phi. For steady, incompressible, irrotational flow, the \nabla \cdot \mathbf{v} = 0 yields \nabla^2 \phi = 0, a special case of Poisson's equation with zero source term. In compressible flows or scenarios with distributed sources—such as approximations for weak or variations—the equation generalizes to Poisson's form \nabla^2 \phi = f, where f incorporates the source; for instance, in unsteady irrotational flows with small perturbations, an approximation is \nabla^2 \phi \approx -\frac{1}{\rho} \frac{\partial \rho}{\partial t}, derived from the \frac{\partial \rho}{\partial t} + \nabla \cdot (\rho \nabla \phi) = 0 under the assumption of slowly varying \rho. This form allows modeling of flows where irrotationality holds approximately, such as in low-Mach-number or source-driven problems like point vortices or sinks. Key applications of these potential formulations include design and water wave analysis, often solved using boundary integral methods that reduce the domain to surface integrals for efficiency. In design, the Hess-Smith panel method discretizes the surface into panels with constant source and vortex distributions to satisfy the no-penetration boundary condition and , enabling computation of the and pressure distribution for prediction in flows. For water waves, the in linear theory satisfies beneath the free surface, with boundary integral methods (e.g., via ) used to evaluate the potential at control points on the domain boundaries, facilitating simulations of wave-structure interactions like those around platforms. A prominent use of Poisson's equation in (CFD) is the pressure Poisson equation within projection methods for solving the incompressible Navier-Stokes equations, which enforce mass conservation by correcting an intermediate velocity field. In Chorin's fractional-step method, the momentum equations are advanced to obtain a provisional velocity \mathbf{u}^*, followed by solving \nabla^2 p = \frac{\rho}{\Delta t} \nabla \cdot \mathbf{u}^* for the p, where the correction \mathbf{u}^{n+1} = \mathbf{u}^* - \frac{\Delta t}{\rho} \nabla p ensures \nabla \cdot \mathbf{u}^{n+1} = 0. This approach, introduced by Chorin in 1967, was extended to second-order accuracy on staggered grids by and in 1985, becoming a cornerstone for simulating viscous incompressible flows in complex geometries. Numerical solutions typically employ finite differences, multigrid, or fast transforms for the elliptic pressure solve, linking directly to broader numerical methods for PDEs.

Thermodynamics

In thermodynamics, Poisson's equation governs the steady-state distribution of temperature in systems with internal heat generation. The steady-state heat conduction equation, derived from the conservation of energy under constant thermal properties, takes the form \nabla^2 T = -Q / k, where T is the field, Q represents the volumetric generation rate (such as from chemical reactions or ), and k is the material's thermal conductivity. This balances diffusive with localized sources, ensuring no net accumulation of over time. The assumption of implies \partial T / \partial t = 0, simplifying the transient to this Poisson form, which is fundamental for analyzing temperature profiles in bounded domains with specified conditions. This formulation finds critical application in heat conduction through solids featuring distributed internal sources, notably in design. In reactor elements, processes produce volumetric heat Q that varies spatially due to distributions, necessitating solutions to \nabla^2 T = -Q / k to predict gradients and prevent hotspots that could compromise structural or coolant efficiency. Analytical solutions are limited to simple geometries, so numerical methods like finite element analysis are employed to resolve the equation across complex core configurations, informing safety margins and operational limits. For instance, in cylindrical rods, radial allows separation into ordinary differential equations, but full three-dimensional modeling is required for heterogeneous assemblies. In the realm of thermodynamic potentials, Poisson's equation emerges within descriptions of inhomogeneous systems, particularly those involving charged particles or density variations. The Poisson-Nernst-Planck () equations couple Poisson's equation for the electrostatic potential \phi, \nabla^2 \phi = -\rho / \epsilon (where \rho is and \epsilon is ), with transport equations for species densities driven by gradients in \mu. Here, the excess chemical potential often follows a Boltzmann form \mu_\text{ex} = k_B T \ln \rho, linking the Laplacian source term directly to density-dependent functions f(\rho), which capture local inhomogeneities in fluids or electrolytes. This framework models in non-uniform environments, such as near interfaces or under external fields, where spatial variations in \mu influence phase stability and diffusion. Seminal developments in PNP theory emphasize steady-state solutions that resolve these coupled effects, providing insights into thermodynamic consistency across scales. Poisson's equation also connects to in , where it simulates dissipative processes far from . In models of transport networks or branching structures, solving the Poisson equation with source terms representing energy dissipation maximizes local rates, aligning with Prigogine's principle of minimum near steady states or maximum production in far-from- regimes. For example, in optimizing fluid or flow paths, the equation's solutions yield configurations that enhance irreversible generation while minimizing total dissipation, as seen in biological or engineered systems with tree-like architectures. This relation underscores Poisson's role in quantifying thermodynamic irreversibility, where the source term f embodies fluxes and affinities driving non-equilibrium evolution. In source-free cases, the equation simplifies to \nabla^2 T = 0, modeling reversible, uniform conduction without increase.

Surface reconstruction

Poisson surface reconstruction is a technique that formulates the problem of generating a smooth, watertight surface from an oriented as solving Poisson's equation over a . Given a set of points with estimated surface normals, the method constructs an indicator function \chi whose level set approximates the underlying surface. The core idea is to solve the Poisson equation \nabla^2 \chi = \nabla \cdot \mathbf{V}, where \mathbf{V} is a smoothed vector field derived from the input point normals, ensuring that the gradient of \chi aligns with the surface orientation. This approach treats reconstruction globally, avoiding local partitioning or blending issues common in other methods. The algorithm begins by estimating oriented normals at each input point, often using on local neighborhoods, followed by orienting them consistently via . To solve the equation efficiently, the volume is discretized using an adaptive structure, which refines resolution near the points to handle non-uniform sampling while keeping computational costs manageable. The resulting sparse is solved iteratively with a multigrid solver, producing values of \chi at nodes. Finally, the at \chi = 0.5 is extracted using an adaptive algorithm, yielding a that is watertight and manifold. This method finds applications in , where it reconstructs detailed models from laser or structured light scans of objects, producing smooth surfaces despite sparse or irregular point distributions. In medical imaging, Poisson reconstruction processes point clouds derived from MRI or data to generate accurate models of anatomical structures, such as liver surfaces for surgical planning and navigation. For instance, screened variants of the algorithm have been used to create patient-specific models from sampled point data in orthopedic and reconstruction. Compared to explicit reconstruction methods like or alpha shapes, Poisson surface reconstruction offers superior robustness to noise in point positions and normals, as well as non-uniform sampling densities, by implicitly fitting a smooth function that minimizes deviation from the input orientations. It produces high-quality, watertight meshes without requiring post-processing for hole filling or seam alignment, making it particularly effective for real-world scanned data with imperfections.

References

  1. [1]
    Poisson's equation - Richard Fitzpatrick
    Poisson's equation is written where $u({\bf r})$ is some scalar potential which is to be determined, and $ v({\bf r})$ is a known ``source function.''
  2. [2]
  3. [3]
    A Poisson Solver Based on Iterations on a Sylvester System
    Poisson's equation ∇2u=f , an elliptic partial differential equation [1] , was first published in 1813 in the Bulletin de la Société Philomatique by Siméon ...<|control11|><|separator|>
  4. [4]
    [PDF] Poisson's Equation in Electrostatics
    Mar 8, 2011 · Poisson's equation is derived from Coulomb's law and Gauss's theorem. In math- ematics, Poisson's equation is a partial differential ...
  5. [5]
    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.
  6. [6]
    [PDF] Fast Solvers - Computer Science | UC Davis Engineering
    Poisson's equation is a differential equation of elliptic type with broad utility in physical models that include gravitation, temperature distribution, ...
  7. [7]
    [PDF] 7 Laplace and Poisson equations - NYU Courant
    On a more purely mathematical line, Laplace's equation appears prominently in the theory of complex variables and in connection with issues of analyticity.
  8. [8]
    Laplacian -- from Wolfram MathWorld
    The Laplacian for a scalar function phi is a scalar differential operator defined by (1) where the h_i are the scale factors of the coordinate system ...
  9. [9]
    [PDF] 1 Poisson Equation
    The short-hand notation for the Poisson equation (2) is. −∇2u = f(x) in Ω, u = ub on ∂ΩD, ∇u · n = g on ∂ΩN , which applies in any number of space ...
  10. [10]
    [PDF] A Crash Course in Linear Elasticity
    Two forms of the Navier equation exist, depending on which constitutive relation, (26) or (28), is used: G∇2ui +. G. (1 − 2ν) uk,ki = α. ∂p. ∂xi. − Fi. (41).
  11. [11]
    [PDF] Chapter 2: Laplace's equation - UC Davis Math
    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 ...
  12. [12]
    [PDF] Notes on Partial Differential Equations John K. Hunter - UC Davis Math
    A physical explanation of this property is that Laplace's equation ... the source as a continuous distribution of point sources and superposing the corre-.
  13. [13]
    [PDF] Lectures on Electromagnetic Field Theory
    Oct 31, 2022 · ... Equation, and Static Green's Function. 33. 3.1 Simple Constitutive ... Poisson's Equation ...
  14. [14]
    [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. Finally, for the ...
  15. [15]
    4.2 The Divergence Theorem
    Solving Poisson's Equation. ·. In this section we shall use the divergence theorem to find a formula for the solution of Poisson's equation.
  16. [16]
    [PDF] PHY481 - Lecture 8
    the differential form of Gauss's law then becomes,. ∇2V = −ρ/ǫ0. (1). This is Poisson's equation and is the most commonly solved form of Gauss's law. The.
  17. [17]
    vacuum electric permittivity - CODATA Value
    Numerical value, 8.854 187 8188 x 10-12 F m ; Standard uncertainty, 0.000 000 0014 x 10-12 F m ; Relative standard uncertainty, 1.6 x 10 ...Missing: ε₀ | Show results with:ε₀<|separator|>
  18. [18]
    [PDF] Part A : Gravity
    Introducing the gravitational potential Φ such that g = ∇Φ, Gauss's Law can be written as. ∇2Φ=4πGρ,. (8) a form known as Poisson's equation. In class, we ...
  19. [19]
    Newtonian constant of gravitation - CODATA Value
    Newtonian constant of gravitation $ G $. Numerical value, 6.674 30 x 10-11 m3 kg-1 s-2. Standard uncertainty, 0.000 15 x 10-11 m3 kg-1 s-2.
  20. [20]
    [PDF] Vector Calculus - DAMTP - University of Cambridge
    Note that the gradient of a scalar field gave a vector field. Now the divergence of a vector field gives a scalar field. The divergence isn't the only way ...
  21. [21]
    [PDF] Green's Functions, the Helmholtz Theorem and Green's Identities 1 ...
    Green's identities are 3 identities that can be derived from vector calculus, with particular cases when Laplace's equation holds. They lead to theorems and to ...
  22. [22]
    [PDF] Green's Identities, Uniqueness, Dirichlet and Neumann Green's ...
    For Dirichlet boundary conditions, since U = 0 on S, then the above constant must vanish and U = 0 everywhere in V . Thus φ1 = φ2 and the solution is unique. ...
  23. [23]
    [PDF] 12 Green's first identity Having studied Laplace's equation in regions ...
    Previously we saw that the maximum principle implies uniqueness of solutions to the Dirichlet problem for Laplace's, and more generally Poisson's equations.
  24. [24]
    [PDF] Method of Green's Functions
    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 . The ...
  25. [25]
    [PDF] solving the poisson equation - Department of Computer Science
    The first method we describe involves Fourier transforms in each variable. We extend the sequences u = {u} and f = {f} to be odd doubly periodic sequences of.
  26. [26]
    [PDF] The Dirichlet Problem on a Rectangle - Trinity University
    Mar 18, 2014 · (x) sin (nπxa ) dx. Separation of variables to shows that the solution to (A) is u1(x,y) = ∞. ∑ n=1. An sin(nπxa ) sinh. (nπ(b − y) a. ) ...
  27. [27]
    [PDF] Today in Physics 217: multipole expansion
    Oct 21, 2002 · The guaranteed uniqueness of solutions has spawned several creative ways to solve the Laplace and Poisson equations for the electric potential.
  28. [28]
    Electric Potential | My Notes
    This is known as Poisson's equation. In regions where there is no charge, so ρ = 0 \rho = 0 ρ=0, Poisson's equation reduces to Laplace's equation,. ∇ 2 V = 0 ...
  29. [29]
    Poisson's Equation - Richard Fitzpatrick
    Poisson's equation is written Here, the function $ v({\bf r})$ is conventionally referred to as a source. Suppose that we have to solve Equation (137) over all ...<|control11|><|separator|>
  30. [30]
    4.7 Method of Images - MIT
    Given a charge distribution throughout all of space, the superposition integral can be used to determine the potential that satisfies Poisson's equation.
  31. [31]
    Method of Images
    Thus, we need to solve Poisson's equation, (2.99), in the region $z>0$ , with a single point charge $q$ located at coordinates (0, 0, $d$ ), subject to the ...
  32. [32]
    [PDF] Lecture 2: Potential Theory I
    Poisson's equation. Substituting −∇Φ for F we have Poisson's equation: Hoorray!! ∇. 2. Φ=4πGρ. (11). The derivation of the Poisson's equation stems from the ...
  33. [33]
    4. Gravitation - Lecture Notes on General Relativity - S. Carroll
    We have now justified Einstein's equations in two different ways: as the natural covariant generalization of Poisson's equation for the Newtonian gravitational ...
  34. [34]
    [PDF] Gravitational Potentials
    Gauss's theorem (8.6) can be used to find the required surface density ... the potential in Poisson's equation. As Fig. 2.8 of BT87 shows, this density ...
  35. [35]
    [PDF] Poisson Equation & Virial Theorem Gravity in Astrophysical Fluids
    Poisson Equation: It can be shown that the gravitational potential obeys the Poisson equation: ∇2Φ=4πGρ. For a derivation, see Section 3.2 of Astrophysical ...
  36. [36]
    Dynamics of Elliptical Galaxies and Other Spheroidal Components
    Equation 10, (10). If the system is self-gravitating, one obtains on solving Poisson's equation with rho replaced by (10),. Equation 11, (11). where. Equation ...
  37. [37]
    Poisson's equation - Steady-state Heat Transfer - Nuclear Power
    Poisson's equation - Steady-state Heat Transfer. Under steady-state conditions, there can be no change in the amount of energy storage (∂T/∂t = 0).
  38. [38]
    [PDF] A Condensed Review of Nuclear Reactor - DSpace@MIT
    Poisson's equation (heat conduction equation). p c. -. = ' (K VT) + Q"'. [26] where T,t,p and cp represent the temperature, time, material density and specific ...
  39. [39]
    [PDF] Numerical Solution of 3D Poisson-Nernst-Planck Equations ...
    We have developed efficient numerical algorithms for solving 3D steady- state Poisson-Nernst-Planck (PNP) equations with excess chemical potentials described by ...
  40. [40]
    Tree Network Formation in Poisson Equation Models and the ...
    We prepare two simulation models: one is the Poisson equation model that simulates the state far from equilibrium, and the other is the Laplace equation model ...
  41. [41]
    Nonequilibrium Thermodynamics and Entropy Production in ... - MDPI
    Nov 17, 2019 · The Poisson equation for the effect of the special charge of the i-th species is: ∂ φ ∂ x = 4 π n i e i. (7). which gives a value of 107V cm−2 ...
  42. [42]
    [PDF] Poisson surface reconstruction - Hugues Hoppe
    We show that surface reconstruction from oriented points can be cast as a spatial Poisson problem. This Poisson formulation considers all the points at once ...
  43. [43]
    CGAL 6.1 - Poisson Surface Reconstruction: User Manual
    The output scalar function, represented in an adaptive octree, is then iso-contoured using an adaptive marching cubes. CGAL implements a variant of this ...
  44. [44]
    Construction of low-cost 3D scanner using triangulation and ...
    Screened Poisson Surface Reconstruction techniques was applied to improve the uncomplete and uneven surface mesh. In order to test the setup, 3D scanning was ...
  45. [45]
    Surface reconstruction for planning and navigation of liver resections
    In this work, we propose the application of Poisson surface reconstruction (PSR) to obtain 3D models of the liver surface with applications to planning and ...
  46. [46]
    Anatomic Surface Reconstruction from Sampled Point Cloud Data ...
    Sep 21, 2015 · In this paper, we propose an approach for reconstruction of an anatomic surface model from point cloud data using the Screened Poisson Surface Reconstruction ...
  47. [47]
    [PDF] Poisson Surface Reconstruction - People @EECS
    We show that surface reconstruction from oriented points can be cast as a spatial Poisson problem. This Poisson formulation considers all the points at once ...